You are on page 1of 749

Algebraic Geometry

Andrew Kobin
Contents

I Commutative Algebra vii


1 Preliminaries 1
1.1 Radicals . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 2
1.2 Nakayama’s Lemma and Consequences . . . . . . . . . . . . . . . . . . . . . 6
1.3 Localization . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 7
1.4 Transcendence Degree . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 12

2 Integral Dependence 16
2.1 Integral Extensions of Rings . . . . . . . . . . . . . . . . . . . . . . . . . . . 17
2.2 Integrality and Field Extensions . . . . . . . . . . . . . . . . . . . . . . . . . 21
2.3 Integrality, Ideals and Localization . . . . . . . . . . . . . . . . . . . . . . . 25
2.4 Normalization . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 32
2.5 Valuation Rings . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 37
2.6 Dimension and Transcendence Degree . . . . . . . . . . . . . . . . . . . . . . 38

3 Noetherian and Artinian Rings 42


3.1 Ascending and Descending Chains . . . . . . . . . . . . . . . . . . . . . . . . 43
3.2 Composition Series . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 46
3.3 Noetherian Rings . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 49
3.4 Primary Decomposition . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 54
3.5 Artinian Rings . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 61
3.6 Associated Primes . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 65

4 Discrete Valuations and Dedekind Domains 68


4.1 Discrete Valuation Rings . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 69
4.2 Dedekind Domains . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 73
4.3 Fractional, Invertible Ideals . . . . . . . . . . . . . . . . . . . . . . . . . . . 75
4.4 The Class Group . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 81
4.5 Dedekind Extensions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 83

5 Completion and Filtration 87


5.1 Topological Abelian Groups and Completion . . . . . . . . . . . . . . . . . . 88
5.2 Inverse Limits . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 91
5.3 Topological Rings and Module Filtrations . . . . . . . . . . . . . . . . . . . 95
5.4 Graded Rings and Modules . . . . . . . . . . . . . . . . . . . . . . . . . . . 97

i
Contents Contents

6 Dimension Theory 102


6.1 Hilbert Functions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 103
6.2 Local Noetherian Rings . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 109
6.3 Complete Local Rings . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 113

II Homological Algebra 118


7 Category Theory 119
7.1 Categories and Functors . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 120
7.2 Exactness of Sequences and Functors . . . . . . . . . . . . . . . . . . . . . . 127

8 Special Modules 136


8.1 Projective Modules . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 137
8.2 Injective Modules . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 142
8.3 Flat Modules . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 149

9 Categorical Constructions 160


9.1 Products and Coproducts . . . . . . . . . . . . . . . . . . . . . . . . . . . . 161
9.2 Limits and Colimits . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 164
9.3 Abelian Categories . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 170
9.4 Projective and Injective Resolutions . . . . . . . . . . . . . . . . . . . . . . . 172

10 Homology 176
10.1 Chain Complexes and Homology . . . . . . . . . . . . . . . . . . . . . . . . . 177
10.2 Derived Functors . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 183
10.3 Tor and Ext . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 189
10.4 Universal Coefficient Theorems . . . . . . . . . . . . . . . . . . . . . . . . . 202

11 Ring Homology 204


11.1 Dimensions of Rings . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 205
11.2 Hilbert’s Syzygy Theorem . . . . . . . . . . . . . . . . . . . . . . . . . . . . 208
11.3 Regular Sequences and the Koszul Complex . . . . . . . . . . . . . . . . . . 212
11.4 Projective Dimension . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 217
11.5 Depth and Cohen-Macauley Rings . . . . . . . . . . . . . . . . . . . . . . . . 222
11.6 Gorenstein Rings . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 231

III Algebraic Geometry 237


12 Algebraic Varieties 238
12.1 Affine Algebraic Varieties . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 239
12.2 Morphisms of Affine Varieties . . . . . . . . . . . . . . . . . . . . . . . . . . 249
12.3 Sheaves of Functions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 254
12.4 Finite Morphisms and the Fibre Theorem . . . . . . . . . . . . . . . . . . . 256
12.5 Projective Varieties . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 259

ii
Contents Contents

12.6 Nonsingular Varieties . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 265


12.7 Abstract Curves . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 269
12.8 Products of Varieties . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 272
12.9 Blowing Up . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 274
12.10Prevarieties and Varieties . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 276
12.11Complete Varieties . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 278

13 Sheaf Theory 281


13.1 Sheaves and Sections . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 282
13.2 The Category of Sheaves . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 289
13.3 Sheaf Cohomology . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 295
13.4 Čech Cohomology . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 299

14 Scheme Theory 308


14.1 The Spectrum of a Ring . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 309
14.2 Schemes . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 316
14.3 Properties of Schemes . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 318
14.4 Sheaves of Modules . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 324
14.5 Group Schemes . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 330
14.6 The Proj Construction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 333

15 Morphisms of Schemes 336


15.1 Flat Morphisms . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 337
15.2 Kähler Differentials . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 341
15.3 Sheaves of Relative Differentials . . . . . . . . . . . . . . . . . . . . . . . . . 347
15.4 Smooth Morphisms . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 350
15.5 Unramified Morphisms . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 357
15.6 Étale Morphisms . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 361
15.7 Henselian Rings . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 365

16 Cohomology 373
16.1 Direct and Inverse Image . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 374
16.2 Acyclic Sheaves . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 383
16.3 Noetherian Scheme Cohomology . . . . . . . . . . . . . . . . . . . . . . . . . 388
16.4 Cohomology of Projective Space . . . . . . . . . . . . . . . . . . . . . . . . . 391

17 Curves 394
17.1 Divisors . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 395
17.2 Morphisms Between Curves . . . . . . . . . . . . . . . . . . . . . . . . . . . 398
17.3 Linear Equivalence . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 401
17.4 Répartitions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 405
17.5 Differentials . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 407
17.6 Serre Duality and the Riemann-Roch Theorem . . . . . . . . . . . . . . . . . 411
17.7 The Riemann-Hurwitz Formula . . . . . . . . . . . . . . . . . . . . . . . . . 414
17.8 The Canonical Map . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 416

iii
Contents Contents

17.9 Bézout’s Theorem . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 417


17.10Rational Points of Conics . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 419

18 Complex Varieties 424


18.1 Complex Manifolds . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 425
18.2 Dolbeault Cohomology . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 430
18.3 Kähler Manifolds . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 433
18.4 Analytic Varieties . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 438
18.5 Divisors, Line Bundles and Sections . . . . . . . . . . . . . . . . . . . . . . . 441
18.6 Cohomology and Chern Classes . . . . . . . . . . . . . . . . . . . . . . . . . 447

IV Algebraic Groups 456


19 Introduction 457
19.1 Chevalley’s Theorem . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 458

20 Algebraic Groups 459


20.1 Hopf Algebras . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 460
20.2 Topological Properties . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 462
20.3 G-spaces . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 466
20.4 Jordan Decomposition . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 471
20.5 Lie Algebra of an Algebraic Group . . . . . . . . . . . . . . . . . . . . . . . 478

21 Classifying Algebraic Groups 483


21.1 Quotients of Algebraic Groups . . . . . . . . . . . . . . . . . . . . . . . . . . 484
21.2 Parabolic and Borel Subgroups . . . . . . . . . . . . . . . . . . . . . . . . . 486
21.3 Diagonalizable Groups . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 492
21.4 Connected Solvable Algebraic Groups . . . . . . . . . . . . . . . . . . . . . . 496
21.5 Semisimple Algebraic Groups . . . . . . . . . . . . . . . . . . . . . . . . . . 502
21.6 Root Data . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 507
21.7 Representations of Algebraic Groups . . . . . . . . . . . . . . . . . . . . . . 517

22 Finite Groups of Lie Type 519


22.1 Jordan’s Theorem . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 521
22.2 Restricted Lie Algebras . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 525
22.3 The Frobenius . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 531
22.4 Chevalley Groups . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 535

V Generalized Jacobians 538


23 Classical Jacobians 539
23.1 Elliptic Functions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 540
23.2 Elliptic Curves . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 547
23.3 The Classical Jacobian . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 553

iv
Contents Contents

23.4 Jacobians of Higher Genus Curves . . . . . . . . . . . . . . . . . . . . . . . . 559

24 Maps From Curves 561


24.1 Moduli and Local Symbols . . . . . . . . . . . . . . . . . . . . . . . . . . . . 562
24.2 The Main Theorem . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 567
24.3 A Word About Symmetric Products . . . . . . . . . . . . . . . . . . . . . . . 574

25 Curves with Singularities 575


25.1 Normalization . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 576
25.2 Singular Riemann-Roch . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 579
25.3 Differentials on Singular Curves . . . . . . . . . . . . . . . . . . . . . . . . . 581

26 Constructing Jacobians 583


26.1 The Jacobian of a Curve . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 584
26.2 Generalized Jacobians . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 587
26.3 The Canonical Map . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 590
26.4 Invariant Differential Forms . . . . . . . . . . . . . . . . . . . . . . . . . . . 594
26.5 The Structure of Generalized Jacobians . . . . . . . . . . . . . . . . . . . . . 596

27 Class Field Theory for Curves 602


27.1 Coverings and Isogenies . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 604
27.2 Class Field Theory . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 606

VI Algebraic Fundamental Groups 609


28 Introduction 610
28.1 Topology Review . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 611
28.2 Finite Étale Algebras . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 614
28.3 Étale Morphisms . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 616

29 Fundamental Groups of Algebraic Curves 618


29.1 Proper Normal Curves . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 620
29.2 Finite Branched Covers . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 622
29.3 Fundamental Group of Curves . . . . . . . . . . . . . . . . . . . . . . . . . . 628
29.4 The Outer Galois Action . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 632
29.5 The Inverse Galois Problem . . . . . . . . . . . . . . . . . . . . . . . . . . . 637

30 Riemann’s Existence Theorem 643


30.1 Riemann Surfaces . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 644
30.2 The Existence Theorem over P1C . . . . . . . . . . . . . . . . . . . . . . . . . 651
30.3 The General Case . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 658
30.4 The Analytic Existence Theorem . . . . . . . . . . . . . . . . . . . . . . . . 660

v
Contents Contents

31 Fundamental Groups of Schemes 664


31.1 Galois Theory for Schemes . . . . . . . . . . . . . . . . . . . . . . . . . . . . 666
31.2 The Étale Fundamental Group . . . . . . . . . . . . . . . . . . . . . . . . . . 669
31.3 Properties of the Fundamental Group . . . . . . . . . . . . . . . . . . . . . . 672
31.4 Structure Theorems . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 676
31.5 Results in Characteristic p . . . . . . . . . . . . . . . . . . . . . . . . . . . . 680

VII Étale Cohomology 684


32 Introduction 685

33 Sites 686
33.1 Grothendieck Topologies and Sites . . . . . . . . . . . . . . . . . . . . . . . 687
33.2 Sheaves on Sites . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 690
33.3 The Étale Site . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 694

34 Cohomology 701
34.1 Direct and Inverse Image Functors . . . . . . . . . . . . . . . . . . . . . . . 702
34.2 Étale Cohomology . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 705

VIII Algebraic Spaces and Stacks 709


35 Introduction 710

36 Categories Fibred in Groupoids 711


36.1 Fibred Categories . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 714
36.2 Fibre Products . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 718

37 Descent 721
37.1 Galois Descent . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 723
37.2 Fields of Definition . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 734
37.3 Galois Descent for Varieties and Schemes . . . . . . . . . . . . . . . . . . . . 736

vi
Part I

Commutative Algebra

vii
Chapter 1

Preliminaries

The contents of Part I come from an Algebra IV course taught by Dr. Peter Abramenko
at the University of Virginia in Spring 2016. The main theory of commutative rings are
covered, with an eye toward applications to algebraic geometry. Topics include:

ˆ Integral dependence

ˆ Noetherian, Artinian and Dedekind rings

ˆ Valuations and discrete valuation rings

ˆ Dimension theory

ˆ Affine algebras

ˆ Hilbert’s Nullstellensatz

The commutative algebra may be found in Atiyah and MacDonald’s Introduction to Com-
mutative Algebra.
We will adhere to several conventions throughout these notes:

ˆ All rings are commutative with unity 1. A will always denote such a ring.

ˆ All subrings contain 1. In particular, proper ideals are never subrings.

ˆ For any ring homomorphism f : A → B, in addition to the homomorphism axioms, we


stipulate that f (1A ) = 1B .

ˆ In any integral domain, we assume 1 6= 0.

Several topics also come from a course on commutative algebra taught by Dr. Craig
Huneke at UVA in Spring 2019. The main topics I’ve included from this course are: associ-
ated primes, some dimension theory and notions of singularity in rings.

1
1.1. Radicals Chapter 1. Preliminaries

1.1 Radicals
There are three important types of radicals that we may define in a commutative ring A.
The first two are defined below.

Definition. The nil radical of A is defined as the intersection of all prime ideals of A,
written \
N (A) = P.
prime ideals
P ⊂A

We say A is reduced if N (A) = 0.

Definition. The Jacobson radical of A is the intersection of all maximal ideals of A,


written \
J(A) = M.
maximal ideals
M ⊂A

Remark. Clearly N (A) ⊆ J(A).

Lemma 1.1.1. For any x ∈ A, x ∈ J(A) if and only if 1 − xy is a unit in A for all y ∈ A.

Proof. ( =⇒ ) Suppose A(1 − xy) 6= A for some y ∈ A. Then A(1 − xy) ⊂ M for a maximal
ideal M . By definition, x ∈ J(A) ⊆ M and since J(A) is an ideal, xy ∈ M as well. This
means 1 = xy + (1 − xy) ∈ M , a contradiction. Therefore A(1 − xy) = A so there is some
a ∈ A such that a(1 − xy) = 1; that is, 1 − xy is a unit.
( ⇒= ) Conversely, suppose x 6∈ M for some maximal ideal M ⊂ A. Then M + Ax = A
by maximality of M , so b + xy = 1 for some b ∈ M , or 1 − xy = b ∈ M , meaning 1 − xy
cannot be a unit in A.

Proposition 1.1.2. N (A) = {x ∈ A | xn = 0 for some n ≥ 1}. That is, the nil radical
consists of all nilpotent elements in A.

Proof. (⊇) Suppose xn = 0. Then for any prime ideal P , xn = 0 ∈ P which implies x ∈ P .
Thus x ∈ N (A).
(⊆) If x ∈ A is not nilpotent, we will produce a prime ideal not containing x. Let Σ
denote the set of ideals I ⊂ A such that xn 6∈ I for any n ≥ 1. Notice that 0 is an element
of Σ since we are assuming x is not nilpotent; thus Σ is nonempty. Therefore we may apply
Zorn’s Lemma to choose an ideal P ∈ Σ which is maximal among this collection of ideals.
We claim that P is prime. Clearly P 6= A since A is not an element of Σ. Suppose y, z ∈ A
but y, z 6∈ P . Since P is maximal in Σ, the ideals P + Ay and P + Az are both strictly larger
ideals than P . Thus P + Ay, P + Az 6∈ Σ so there exist m, n ∈ N such that xm ∈ P + Ay and
xn ∈ P + Az. Then xm+n ∈ (P + Ay)(P + Az) ⊆ P + Ayz, but xm+n 6∈ P by assumption,
so we must have yz 6∈ P . This proves P is prime.

Definition. For an ideal I ⊂ A, the radical of I √is r(I) = {x ∈ A | xn ∈ I for some n ≥ 1}.
Alternate notations for the radical ideal include I and rad(I).

2
1.1. Radicals Chapter 1. Preliminaries

Corollary 1.1.3. For any ideal I ⊂ A, r(I) equals the intersection of all prime ideals of A
containing I. In particular, r(I) is an ideal.

Proof. Consider the quotient ring Ā = A/I. For x ∈ A we have the following equivalent
statements:

x ∈ r(I) ⇐⇒ there exists an n ∈ N such that xn ∈ I


⇐⇒ there exists an n ∈ N such that (x + I)n = 0 ∈ Ā
⇐⇒ x + I is nilpotent, i.e. x + I ⊂ N (Ā)
⇐⇒ x + I is contained in every prime of Ā
⇐⇒ x lies in every prime ideal of A containing I.
\
Therefore r(I) = P as desired.
prime ideals
P ⊇I

Remark. By Corollary 1.1.3, N (A) = r(0).

Lemma 1.1.4. Suppose I, J ⊂ A are ideals. Then

(a) I ⊆ r(I).

(b) I = r(I) if and only if A/I is reduced. In particular, every prime ideal is radical.

(c) r(r(I)) = r(I).

(d) r(I m ) = r(I) for all m ∈ N.

(e) r(IJ) ⊇ r(I)r(J).

(f ) r(I + J) = r(r(I) + r(J)).

(g) r(I) = A if and only if I = A.

Proof. (a) is clear from the definition.


(b) By Corollary 1.1.3, r(I) equals the intersection of all primes lying over I. Let π : A →
A/I be the quotient map. Then N (A/I) = 0 ⇐⇒ π(r(I)) = 0 ⇐⇒ r(I) ⊆ ker π = I. By
part (a), we always have I ⊆ r(I) so this shows that N (A/I) = 0 ⇐⇒ r(I) = I.
(c) By part (a), it suffices to show r(r(I)) ⊆ r(I). If x ∈ r(r(I)) then xn ∈ r(I) for some
n ≥ 1, but then (xn )m = xnm ∈ I for some m ≥ 1. Thus x ∈ r(I).
(d) If x ∈ r(I m ) then xn ∈ I m for some n ≥ 1, but I m ⊆ I so this shows x ∈ r(I). On
the other hand, if x ∈ r(I) then xn ∈ I for some n ≥ 1. Letting k = max{n, m}, we see that
xk ∈ I m . Hence r(I) = r(I m ).
(e) It suffices to prove this for elements of the form xy ∈ r(I)r(J), where x ∈ r(I) and
y ∈ r(J). Then xn ∈ I and y m ∈ J for some n, m ≥ 1. Without loss of generality, assume
n ≥ m. Then (xy)n = xn y n ∈ IJ so xy ∈ r(IJ).
(f) First take z ∈ r(I + J), so that z n ∈ I + J for some n ≥ 1. Then z n = x + y for
x ∈ I, y ∈ J. Since I ⊆ r(I) and J ⊆ r(J) by part (a), we have z n = x + y ∈ I + J ⊆

3
1.1. Radicals Chapter 1. Preliminaries

r(I) + r(J). Therefore z ∈ r(r(I) + r(J)). On the other hand, if w ∈ r(r(I) + r(J)) then
wm ∈ r(I) + r(J) for some m ≥ 1. This means wm = x + y for some x ∈ r(I), y ∈ r(J). In
turn this says that xk ∈ I and y ` ∈ J for some k, ` ≥ 1. Using the binomial formula, write
 
m(k+`) k+` k+` k+`−1 k + ` k+`−q q
w = (x + y) = x + (k + `)x y + ... + x y + . . . + y k+` .
q

Note that q ranges from 0 to k + `. When 0 ≤ q ≤ `, the qth term (the first term in the sum
being the 0th term) is divisible by xk and so it lies in I. Otherwise when 1 + ` ≤ q ≤ k + `,
the qth term is divisible by y ` and so it lies in J. Therefore wm(k+`) is of the form a + b
where a ∈ I and b ∈ J. Hence w ∈ r(I + J).
(g) Suppose r(I) = A. Then in particular 1 ∈ r(I) so for some n ≥ 1, 1 = 1n ∈ I.
Thus I = A. On the other hand, if I = A then part (a) says A ⊆ r(A), which implies
r(A) = A.

Example 1.1.5. Let A = Z and I = J = 2Z. Then IJ = 4Z so r(IJ) = 2Z, but I and J
are each prime and thus radical, so r(I)r(J) = IJ = 4Z. Hence r(IJ) ) r(I)r(J), showing
that the containment in (e) may be strict.

Example 1.1.6. Let k be an algebraically closed field and consider A = k[x, y]. Take
I = (y) and J = (y − x2 ). Then I = r(I) since A/I = k[x] which is a PID and hence
reduced. Likewise, k[x, y]/(y − x2 ) ∼ = k[x] by the first isomorphism theorem applied to the
map k[x, y] → k[x], f (x, y) 7→ f (x, x2 ). Thus J = r(J), so r(I) + r(J) = I + J. However,
I + J = (y) + (y − x2 ) = (y, x2 ) and the radical of this ideal is r((y, x2 )) = (y, x), a strictly
larger ideal than I + J. This shows the containment r(I + J) ⊂ r(I) + r(J) is sometimes
strict, so the second radical on the right side of property (f) is necessary for equality.

Lemma 1.1.7. Let J1 , . . . , Jn be ideals of a ring A and J = ni=1 Ji . Then r(J) = ni=1 r(Ji ).
T T

Proof. It suffices to prove the case when n = 2. Suppose x ∈ r(J1 ∩J2 ). Then xn ∈ J1 ∩J2 for
some n ∈ N, so in particular xn ∈ J1 , implying x ∈ r(J1 ), and xn ∈ J2 , implying x ∈ r(J1 ).
Hence x ∈ r(J1 ) ∩ r(J2 ).
On the other hand, if y ∈ r(J1 ) ∩ r(J2 ), then y ∈ r(J1 ) so y k ∈ J1 for some k ∈ N.
Similarly, y ∈ r(J2 ) implies y ` ∈ J2 for some ` ∈ N. Set m = max(k, `). Then y m =
y m−k y k ∈ J1 and y m = y m−` y ` ∈ J2 by the setup, so y m ∈ J1 ∩ J2 . Hence y ∈ r(J1 ∩ J2 ).
This proves r(J1 ∩ J2 ) = r(J1 ) ∩ r(J2 ).

Example 1.1.8. The conclusion of Lemma 1.1.7 does not hold for arbitrary intersections
of ideals. Let kTbe a field and A = k[t] and consider the family of ideals Jn = (tn ) for
`
n ∈T N. Then n∈N Jn = 0 since if t | f (t) for all ` ≥ 1 then necessarily f = 0. So
rT n∈N Jn = r(0) = N (A) = 0 since A is a PID. However, for each n ∈ N, r((tn )) = (t) so
n∈N r(Jn ) = (t) 6= 0.

Lemma 1.1.9. If f : A → B is a ring homomorphism and J is an ideal of B, then


f −1 (r(J)) = r(f −1 (J)).

4
1.1. Radicals Chapter 1. Preliminaries

Proof. Let x ∈ A. Then

x ∈ f −1 (r(J)) ⇐⇒ f (x) ∈ r(J)


⇐⇒ f (x)n ∈ J for some n ∈ N
⇐⇒ f (xn ) ∈ J for some n ∈ N since f is a homomorphism
⇐⇒ xn ∈ f −1 (J) for some n ∈ N
⇐⇒ x ∈ r(f −1 (J)).

Hence f −1 (r(J)) = r(f −1 (J)).

5
1.2. Nakayama’s Lemma and Consequences Chapter 1. Preliminaries

1.2 Nakayama’s Lemma and Consequences


A classic result in ring theory is Nakayama’s Lemma, which will be useful later on.
Theorem 1.2.1 (Nakayama’s Lemma). If I ⊂ A is an ideal such that I ⊆ J(A), and M is
a finitely generated A-module such that IM = M , then M = 0.
Proof. Assume IM = M but M 6= 0. Since M is finitely generated, pick some minimal finite
set of generators {m1 , . . . , mn } of M . Then mn ∈ M = IM = Im1 + . . . + Imn so there exist
elements a1 , . . . , an ∈ I such that mn = a1 m1 + . . . + an mn . This can be written
(1 − an )mn = a1 m1 + . . . + an−1 mn−1 .
Since an ∈ I ⊆ J(A), Lemma 1.1.1 shows 1−an is a unit in A. Then mn ∈ Am1 +. . .+Amn−1 .
This shows that {m1 , . . . , mn−1 } generates M as an A-module, contradicting minimality.
Therefore we must have M = 0.
One can find examples of ideals not lying in the Jacobson radical for which IM = M for
nontrivial modules M .
Example 1.2.2. Let A be a local ring which is an integral domain a but not a field.
As an
example, for any prime p, the ring of p-adic integers Z(p) = b : a, b ∈ Z, p - b is a local
subring of Q which is not a field. Let M be the unique maximal ideal of A and take F to be
its field of fractions (see below for details). Since A is not a field, M 6= 0. In fact, in the local
ring situation, M = J(A). Then we have M F = F for the A-module F , but F 6= 0. This
provides a counterexample to Nakayama’s Lemma when the module is not finitely generated,
as is the case with the residue field of a local ring.
The following is another version of Nakayama’s Lemma which may be useful in some
contexts.
Corollary 1.2.3. Let M be a finitely generated A-module and take N ⊆ M to be a submod-
ule. If I ⊂ A is an ideal such that I ⊆ J(A) and M = IM + N , then M = N .
Proof. Consider the finitely generated A-module M = M/N . Then
IM = (IM + N )/N = M/N = M .
Applying Nakayama’s Lemma shows that M = 0, that is, M = N .
Let A be a local ring with maximal ideal M = J(A). We call the quotient field k =
A/M the residue field of A. If N is a finitely generated A-module then N/M N is a finite
dimensional k-vector space.
Proposition 1.2.4. If x1 , . . . , xn are elements of N such that x̄1 , . . . , x̄n ∈ N/M N span
N/M N as a k-vector space, then N is generated by the xi as an A-module.
Proof. Set N 0 = ni=1 Axi , which is a submodule of N . Consider the canonical projection
P
π : N → N/M N . By assumption, π(N 0 ) = N/M N . Thus N = N 0 + M N but since
M = J(A) and N is finitely generated as an A[x]-module, the conditions of Corollary 1.2.3
are satisfied. Hence N 0 = N .

6
1.3. Localization Chapter 1. Preliminaries

1.3 Localization
One of the most important constructions in commutative algebra is the localization of a ring
at a multiplicatively closed subset.

Definition. A subset S ⊆ A is said to be multiplicatively closed provided that 1A ∈ S


and for all x, y ∈ S, xy ∈ S as well.

Examples. Two of the most important examples of multiplicatively closed subsets are:

1 For any prime ideal P ⊂ A, the set S = A r P is multiplicatively closed.

2 For a fixed element f ∈ A (usually a function in a ring of functions), the set S = {f n |


n ≥ 0} is multiplicatively closed, where by convention f 0 = 1A .

Definition. Let S ⊆ A be multiplicatively closed and suppose M is a finitely generated A-


module. The module of fractions S −1 M is defined as S −1 M = M × S/ ∼ where ∼ is the
equivalence relation (m, s) ∼ (n, t) iff u(tm − sn) = 0 for some u ∈ S. It is common to write
elements of S −1 M as ms instead of (m, s). Addition and the A-module action of S −1 M are
given by
m n tm + sn m am
+ = and a = .
s t st s s
Lemma 1.3.1. The above operations make S −1 M into an A-module.

Proposition 1.3.2. For the A-module A, S −1 A is a ring under the multiplication law as · bt =
ab
st
, with unity 1S −1 A = 11AA .

Proof. We will verify that multiplication is well-defined. The other ring axioms are easily
checked.
Example.

3 If A is an integral domain then S = A r {0} is multiplicatively closed. The ring S −1 A


is called the field of fractions of A.

Lemma 1.3.3. There is a homomorphism j : A → S −1 A defined by a 7→ a


1
which is one-to-
one if and only if S contains no zero divisors.

Proposition 1.3.4 (Universal Property of Localization). Let S ⊆ A be a multiplicatively


closed subset. For any ring homomorphism ϕ : A → B such that ϕ(S) ⊆ B × , there exists a
unique ring homomorphism ϕ0 : S −1 A → B making the following diagram commute:
ϕ
A B

j ϕ0

S −1 A

7
1.3. Localization Chapter 1. Preliminaries

Proof. First, if as ∈ S −1 A and such a ϕ0 is defined, we must have ϕ0 as  = ϕ0 a1 ϕ0 1s =


  

ϕ(a)ϕ(s)−1 . This implies uniqueness. Now define ϕ0 in this way: ϕ0 as = ϕ(a)ϕ(s)−1 for
any a ∈ A, s ∈ S. To see that ϕ0 is well-defined, suppose as = bt in S −1 A. Then there is an
element u ∈ S such that u(at − bs) = 0. Applying ϕ, we have ϕ(u)ϕ(at − bs) = 0. Since
ϕ(u) is a unit in B, it follows that

ϕ(at − bs) = 0 =⇒ ϕ(a)ϕ(t) = ϕ(b)ϕ(s) =⇒ ϕ(a)ϕ(s)−1 = ϕ(b)ϕ(t)−1 .

Hence ϕ0 is well-defined. The rest follows easily from this definition of ϕ0 .


Examples.

4 If S = ArP for a prime ideal P ⊂ A, the localization of A at P is the ring AP := S −1 A.

5 If S = {f n | n ≥ 0} for some element f ∈ A then the localization of A at f is


Af := S −1 A.

Proposition 1.3.5. If f, g ∈ A are elements such that r((f )) = r((g)) then Af ∼


= Ag .
Proof. The assumption r((f )) = r((g)) means that f n = xg and g m = yf for some n, m ≥ 1
and x, y ∈ A. Let jf : A → Af and jg : A → Ag be the canonical homomorphisms sending
a 7→ a1 in each localization. Then jf (g) = g1 is invertible in Af , since f n − xg = 0 and so
g
· x = 11 . Symmetrically, jg (f ) = f1 is invertible in Ag . Therefore by the universal property
1 fn
of localization, there are unique homomorphisms ϕ1 : Ag → Af and ϕ2 : Af → Ag such that
ϕ1 ◦ jg = jf and ϕ2 ◦ jf = jg :
A
jf jg
ϕ2
Af Ag
ϕ1

Then it is clear that ϕ1 and ϕ2 are inverses of each other, so Af ∼


= Ag .
Proposition 1.3.6. If M is an A-module and S ⊆ A is a multiplicatively closed subset then
there exists a well-defined isomorphism of S −1 A-modules (or of A-modules)

f : S −1 A ⊗A M −→ S −1 M
a
s
⊗ m 7−→ am
s
.

Proof. Define a map

S −1 A × M −→ S −1 M
a
, m 7−→ am

s s
.

It is easy to check that is A-bilinear, so by the universal property of tensor products, we get a
well-defined homomorphism of A-modules f : S −1 A ⊗A M → S −1 M . Clearly f is surjective.

8
1.3. Localization Chapter 1. Preliminaries

For an arbitrary element i asii ⊗ mi ∈ S −1 A ⊗ M , set s = i si ∈ S and ti = j6=i sj ∈ S.


P Q Q
Then X X X X
ai ai ti 1 1
si
⊗ mi = s
⊗ m i = s
⊗ a t m
i i i = s
⊗ ai ti mi .
i i i i
−1 1
Therefore every element of S A ⊗ M has the form ⊗ m for s ∈ S and m ∈ M . Now
s
suppose f s ⊗ m = 0. Then ms = 0 by definition of the map, so tm = 0 for some t ∈ S.
1


Thus 1s ⊗ m = stt ⊗ m = st1 ⊗ tm = st1 ⊗ 0 = 0. This proves f is injective and thus we have
an isomorphism S −1 A ⊗A M ∼ = S −1 M as A-modules. Finally, it’s easy to see that the same
map is also an isomorphism of S −1 A-modules.
Given an ideal I ⊂ A, we can define the subset S −1 I ⊆ S −1 A by viewing I as an
−1 a

A-module. Specifically, S I = s : a ∈ I, s ∈ S .

Lemma 1.3.7. Let I ⊂ A be an ideal and let S ⊂ A be a multiplicatively closed subset.


Then

(a) S −1 I ⊆ S −1 A is an ideal.

(b) S −1 I = S −1 A if and only if S ∩ I 6= ∅.

Proof. (a) Let as ∈ S −1 A and bt ∈ S −1 I, meaning s, t ∈ S, a ∈ A and b ∈ I. Then as · bt = ab


st
lies in S −1 I since ab ∈ I by absorption and st ∈ S by the multiplicatively closed property.
Hence S −1 I is an ideal of S −1 A.
(b) On one hand, if s ∈ S ∩ I then ss = 11 ∈ S −1 I so S −1 I = S −1 A. On the other hand,
S −1 I = S −1 A implies 11 ∈ S −1 I so there exist elements a ∈ I and s ∈ S such that as = 11 .
Then there is a t ∈ S such that t(a − s) = 0, that is, ts = ta ∈ S ∩ I. Thus S ∩ I is
nonempty.

Proposition 1.3.8. Let S be a multiplicatively closed subset of A. Then

(a) Every proper ideal J ⊂ S −1 A is of the form S −1 I for some ideal I not intersecting S.

(b) For any prime ideal P ⊂ A not intersecting S, j −1 (S −1 P ) = P .

(c) The map

{P ⊂ A | P is a prime ideal, S ∩ P = ∅} −→ {prime ideals of S −1 A}


P 7−→ S −1 P

is an inclusion-preserving bijection.

Proof. (a) Let J ⊂ S −1 A be an ideal. Then I = j −1 (J) is an ideal of A. If as ∈ S −1 I for


a ∈ I, s ∈ S, then a1 ∈ J so as = a1 · 1s ∈ J. On the other hand, if as ∈ J then a1 = as · 1s ∈ J.
Then a ∈ I since I = j −1 (J). Hence as ∈ S −1 I, which proves that S −1 I = J. Also, S ∩ I = ∅
since S −1 I = J ( S −1 A.
(b) By definition of the map j, P ⊆ j −1 (S −1 P ). For the other containment, suppose
a ∈ j −1 (S −1 P ). Then a1 ∈ S −1 P so there exist b ∈ P and s ∈ S such that a1 = sb . Thus for
some t ∈ S we have t(sa − b) = 0, or tsa = tb which lies in P since b ∈ P . Since s, t ∈ S and

9
1.3. Localization Chapter 1. Preliminaries

S ∩ P = ∅, s, t 6∈ P but then tsa ∈ P implies a ∈ P since P is prime. Thus P = j −1 (S −1 P )


as claimed.
(c) Suppose P ⊂ A is a prime ideal with S ∩ P = ∅. Let xs , yt ∈ S −1 A such that
x
s
· yt = xy
st
∈ S −1 P . Then there are elements a ∈ P and u ∈ S such that xy st
= ua . So for
some v ∈ S, v(uxy − sta) = 0; that is, vuxy = vsta which lies in P since a ∈ P . Now
v, u ∈ S so v, u 6∈ P , meaning x ∈ P or y ∈ P . Thus xs ∈ S −1 P or yt ∈ S −1 P . This proves
that S −1 P is a prime ideal of S −1 A. Finally, denote the map P 7→ S −1 P by Φ. By (b), Φ is
injective. For surjectivity, assume J ⊂ S −1 A is prime. Then P = j −1 (J) is also prime. By
(a), J = S −1 P = Φ(P ). Also, S ∩ P = ∅ since J is prime and therefore J 6= S −1 A. Hence
Φ is surjective, so it is a bijection.

Corollary 1.3.9. If P ⊂ A is prime then AP is a local ring with unique maximal ideal
S −1 P , where S = A r P . Moreover, there is an inclusion-preserving bijection

{Q ⊂ A | Q is a prime ideal, Q ⊆ P } ←→ {prime ideals of AP }.

Proof. We know that S −1 P ⊂ S −1 A is a prime ideal. Observe that

S −1 A r S −1 P = as : a ∈ A r P, s ∈ S = as : a ∈ S, s ∈ S = A×
 
P,

the set of units in AP . This implies S −1 P is the unique maximal ideal of S −1 A. The bijection
follows from (c) of Proposition 1.3.8, using the fact that S ∩ Q = ∅ if and only if Q ⊆ P .

Lemma 1.3.10. Let S be a multiplicatively closed subset and I, J ideals of the ring A. Then

(a) S −1 (IJ) = (S −1 I)(S −1 J).

(b) S −1 (I ∩ J) = (S −1 I) ∩ (S −1 J).

(c) S −1 r(I) = r(S −1 I).


Pn
Proof. (a) Let xs ∈ S −1 (IJ), with x ∈ IJ and s ∈ S. Then x = k=1 ak bk for ak ∈ I and
bk ∈ J. Moreover, we assume 1 ∈ S by convention so
n n
x X ak b k X ak b k
= = · .
s k=1
s k=1
s 1

Then each summand is a product of an element in S −1 I and an element in S −1 J, so xs ∈


(S −1 I)(S −1 J). Thus S −1 (IJ) ⊆ (S −1 I)(S −1 J). On the other hand, if yt ∈ (S −1 I)(S −1 J)
then there are elements ak ∈ I, bk ∈ J, sk , tk ∈ S such that
n Pn
y X ak bk s1 t1 · · · ŝk t̂k · · · sn tn ak bk
= = k=1 .
t k=1
sk tk s1 t1 · · · sn tn

The numerator lies in IJ since ak ∈ I, bk ∈ J so each summand is a product of an element of


I and an element of J. The denominator is an element of S since this set is multiplicative.
Therefore yt ∈ S −1 (IJ) by definition. Hence (S −1 I)(S −1 J) ⊆ S −1 (IJ).

10
1.3. Localization Chapter 1. Preliminaries

(b) Observe that for any x


s
∈ S −1 A,
x
∈ S −1 (I ∩ J) ⇐⇒ x ∈ I ∩ J and s ∈ S
s
x x
⇐⇒ ∈ S −1 I and ∈ S −1 J
s s
x
⇐⇒ ∈ (S I) ∩ (S −1 J).
−1
s
So S −1 (I ∩ J) = (S −1 I) ∩ (S −1 J).
(c) If I ∩ S 6= ∅, I ⊆ r(I) implies r(I) ∩ S 6= ∅ as well so we have

S −1 r(I) = S −1 A = r(S −1 A) = r(S −1 I).

Now assume I ∩ S = ∅. Since S is multiplicative, if s were in r(I) for any s ∈ S then


sn ∈ I ∩TS for some n ∈ N, which is impossible. Therefore r(I) ∩ S = ∅. By Corollary 1.1.3,
r(I) = p where p runs over all prime ideals of A that contain I. Then by Proposition 1.3.8,
we have !
\ \ \
S −1 r(I) = S −1 p = S −1 p = q = r(S −1 I)
p⊇I p⊇I q⊇S −1 I

where p runs over the primes of A containing I and q runs over the primes of S −1 A containing
S −1 I. Hence we have S −1 r(I) = r(S −1 I) in all cases.

Lemma 1.3.11. For a prime ideal P ⊂ A, the quotient field AP /S −1 P is isomorphic to the
field of fractions of A/P .
π i
Proof. Let ϕ : A → − A/P → − Frac(A/P ) be the canonical quotient map composed with

i : ā 7→ 1 . Then P ⊆ ker ϕ, so S = ArP gets mapped to nonzero elements in Frac(A/P ), but
these are all units since Frac(A/P ) is a field. Therefore the universal property of localization
gives us a unique homomorphism f : AP → Frac(A/P) such that f ◦ j = ϕ. Notice that if
a
s
∈ S −1 P then a ∈ P so ϕ(a) = 0 and therefore f as = 0. So S −1 P ⊆ ker f . Clearly f is
onto, so by the first isomorphism theorem, we have AP /S −1 P ∼ = Frac(A/P ).
Example.

6 Let A = Z. Since Z is an integral domain, (0) is a prime ideal and its localization is
A(0) = Q, the field of fractions of Z. For any nonzero prime p ∈ Z, the localization
Z(p) = as : a, s ∈ Z, p - s is called the ring of p-adic integers. Each Z(p) contains only

the prime ideals (0) and pZ(p) := S −1 (p).

11
1.4. Transcendence Degree Chapter 1. Preliminaries

1.4 Transcendence Degree


Let K/k be an extension of fields and let E ⊆ K be any subset.

Definition. We say E is algebraically independent over k if for any finite subset {x1 , . . . , xn } ⊆
E and any function F ∈ k[t1 , . . . , tn ] for which F (x1 , . . . , xn ) = 0, it must be that F = 0.
When E = {x}, we say the element x is transcendental (or transcendent) over k.

Clearly x ∈ K is transcendental over k if and only if it is not algebraic over k.

Definition. A set E ⊆ K is a transcendence basis for K/k if E is a maximal algebraically


independent subset of K. In the case that E is a transcendence basis for K/k and K = k(E),
we say the extension is purely transcendental.

Note that the empty set E = ∅ is a transcendence basis if and only if K/k is an algebraic
extension.

Example 1.4.1. Let K = k(t1 , . . . , tn ) be the field of fractions of the polynomial ring
k[t1 , . . . , tn ]. Then K is a purely transcendental extension of K, with transcendence basis
E = {t1 , . . . , tn }.

Lemma 1.4.2. If {x1 , . . . , xn } ⊂ K is algebraically independent then k(x1 , . . . , xn ) ∼


= k(t1 , . . . , tn ).
Proof. Consider the surjection

ϕ : k[t1 , . . . , tn ] −→ k[x1 , . . . , xn ]
ti 7−→ xi .

Then ϕ is injective because {x1 , . . . , xn } is algebraically independent. Moreover, ϕ is a k-


algebra homomorphism so this extends to an isomorphism k(t1 , . . . , tn ) → k(x1 , . . . , xn ).

Lemma 1.4.3. Let K/k be a field extension and suppose E ⊆ K with |E| = n. Then

(a) If E is algebraically independent over k then there exists a transcendence basis E 0 of


K/k such that E 0 ⊇ E.

(b) If E is finite and K = k(E) then the elements of E may be numbered such that
k(x1 , . . . , xr )/k is purely transcendental for some 0 ≤ r ≤ n and K/k(x1 , . . . , xr ) is
algebraic.

(c) E is a transcendence basis for K/k if and only if K/k(E) is algebraic and K/k(E1 )
is not algebraic for any proper subset E1 ( E.

Proof. (a) Apply Zorn’s Lemma (exercise).


(b) Let E be numbered so that {x1 , . . . , xr } is a maximal algebraically independent subset
of E. Clearly {x1 , . . . , xr } is a transcendence basis for k(x1 , . . . , xr ) so k(x1 , . . . , xr )/k is
purely transcendental. We will show that for each r + 1 ≤ j ≤ n, xj is algebraic over

12
1.4. Transcendence Degree Chapter 1. Preliminaries

k(x1 , . . . , xr ). We know for each j, {x1 , . . . , xr , xj } is algebraically dependent over k, so there


exists a nonzero polynomial F ∈ k[t1 , . . . , tr+1 ] such that F (x1 , . . . , xr , xj ) = 0. Write

F = Fd (t1 , . . . , tr )tdr+1 + . . . + F1 (t1 , . . . , tr )tr+1 + F0 (t1 , . . . , tr )

where Fi ∈ k[t1 , . . . , tr ] and Fd 6= 0. Then

0 = Fd (x1 , . . . , xr )xdj + . . . + F1 (x1 , . . . , xr )xj + F0 (x1 , . . . , xr )

and Fd (x1 , . . . , xr ) 6= 0 since {x1 , . . . , xr } is algebraically independent. We must therefore


have d ≥ 1. This shows xj is algebraic over k(x1 , . . . , xr ) for each r + 1 ≤ j ≤ n, as we had
claimed. Finally, it follows that K/k(x1 , . . . , xr ) is algebraic.
(c) Pick any x ∈ K r E. Then since E is a maximal algebraically independent subset
of K, E ∪ {x} is algebraically dependent over k. This means there exist x1 , . . . , xr ∈ E
such that {x1 , . . . , xr , x} is algebraically dependent over k. Hence x must be algebraic over
k(x1 , . . . , xr ) as in (b). In particular, x is algebraic over k(E). To show the second property,
we may suppose without loss of generality that E1 = {x1 , . . . , xr−1 } so that K/k(x1 , . . . , xr−1 )
is algebraic. Consider the tower K ⊃ k(E) ⊃ k(x1 , . . . , xr−1 ). Then k(E) = k(x1 , . . . , xr ) is
algebraic over k(x1 , . . . , xr−1 ), which implies xr is algebraic over k. This clearly contradicts
the algebraic independence of {x1 , . . . , xr }. Hence K/k(E1 ) cannot be algebraic for any
proper subset E1 ( E.
Going the other direction, suppose E satisfies the properties that K/k(E) is algebraic and
K/k(E1 ) is not algebraic for any E1 ( E. By (b), we may number E = {x1 , . . . , xr , . . . , xn }
so that k(x1 , . . . , xr )/k is purely transcendental and K/k(x1 , . . . , xr ) is algebraic. By as-
sumption {x1 , . . . , xr } cannot be a proper subset of E, so we get r = n and k(E)/k is purely
transcendental. This implies E is algebraically independent over k. Finally, since K/k(E) is
algebraic, for any x ∈ K r k(E), x is algebraic. So E ∪ {x} is not algebraically independent.
This proves maximality of E, hence E is a transcendence basis.

Proposition 1.4.4. Assume E1 , E2 ⊆ K are finite subsets with K/k(E1 ) algebraic and E2
algebraically independent over k. Then |E1 | ≥ |E2 |.

Proof. Let E1 = {x1 , . . . , xm } and E2 = {y1 , . . . , yn }, so that |E1 | = m and |E2 | = n.


We claim that the elements of E1 can be numbered so that K/k(y1 , . . . , y` , x`+1 , . . . , xm )
is algebraic for all 0 ≤ ` ≤ min{m, n}. We prove this by induction on `. For ` = 0,
K/k(E1 ) is already algebraic by assumption. Now suppose we can number E1 so that
K/k(y1 , . . . , y`−1 , x` , . . . , xm ) is algebraic; set K`−1 = k(y1 , . . . , y`−1 , x` , . . . , xm ). Now y` is
algebraic over K`−1 since y` ∈ K. This implies there exist polynomials G0 , . . . , Ge in m
variables such that

Ge (y1 , . . . , y`−1 , x` , . . . , xm )y`e + . . . + G0 (y1 , . . . , y`−1 , x` , . . . , xm ) = 0

and Ge (y1 , . . . , y`−1 , x` , . . . , xm ) 6= 0. In particular, Ge must be nonzero. It follows that there


exists a nonzero polynomial F in m + 1 variables such that

F (y1 , . . . , y` , x` , . . . , xm ) = 0 (∗)

13
1.4. Transcendence Degree Chapter 1. Preliminaries

We may assume F is such a polynomial of minimal degree. Now {y1 , . . . , y` } is algebraically


independent over k so one of x` , . . . , xm must occur in (∗). After renumbering, we may
assume x` occurs in (∗), which then becomes

Fd (y1 , . . . , y` , x`+1 , . . . , xm )xd` + . . . + F0 (y1 , . . . , y` , x`+1 , . . . , xm ) = 0

for polynomials Fi in m variables over k with Fd 6= 0. Then Fd (y1 , . . . , y` , x`+1 , . . . , xm ) 6= 0


due to the minimality of deg F . This also implies d ≥ 1. Then deg F ≥ deg Fd + d.
The above equation shows that x` is algebraic over K` := k(y1 , . . . , y` , x`+1 , . . . , xm ). By
induction, K/K`−1 is algebraic but since K`−1 ⊆ K` , we have now shown that K/K` (x` )
and K` (x` )/K` are both algebraic extensions. By the tower property of algebraic extensions,
K/K` is algebraic, proving the claim.
If min{m, n} = n then n ≤ m and we’re done. On the other hand, if min{m, n} = m
then we can use the above induction to conclude that K/k(y1 , . . . , ym ) is algebraic. But then
we must have n = m since if n > m, ym+1 will be algebraic over k(y1 , . . . , ym ), in which case
there is a nontrivial polynomial equation H(y1 , . . . , ym+1 ) = 0. This however contradicts
{y1 , . . . , ym+1 } being algebraically independent. In every case we have n ≤ m as desired.
Corollary 1.4.5. If E1 and E2 are transcendence bases for K/k then |E1 | = |E2 | or they
are both infinite.
Proof. Apply Proposition 1.4.4 and (c) of Lemma 1.4.3.
This allows us to define the following:
Definition. The transcendence degree of a field extension K/k is defined as tr degk K =
|E| if K/k admits a finite transcendence basis E, and tr degk K = ∞ otherwise.
Remark. tr degk K = 0 if and only if K/k is algebraic.
Corollary 1.4.6. The number r in Lemma 1.4.3(b) such that k(x1 , . . . , xr )/k is purely tran-
scendental and K/k(x1 , . . . , xr ) is algebraic is equal to tr degk K.
Proof. We proved that {x1 , . . . , xr } is algebraically independent over k and {x1 , . . . , xr , x}
is algebraically dependent for any x ∈ K r {x1 , . . . , xr }. Hence {x1 , . . . , xr } is a maximal
algebraically independent set in K, and thus a transcendence basis for K/k. Therefore
r = #{x1 , . . . , xr } = tr degk K.
Example 1.4.7. For the purely transcendental extension k(t1 , . . . , tn ), its transcendence
degree is tr degk k(t1 , . . . , tn ) = n.
Proposition 1.4.8. For a tower of fields L ⊃ K ⊃ k in which each extension is finitely
generated, we have
tr degk L = tr degK L + tr degk K.
Proof. Suppose E1 = {x1 , . . . , xn } is a transcendence basis for K/k and E2 = {y1 , . . . , ym }
is a transcendence basis for L/K. Set E = E1 ∪ E2 = {x1 , . . . , xn , y1 , . . . , ym }. We first show
E is algebraically independent over k. Suppose f ∈ k[t1 , . . . , tn+m ] such that

f (x1 , . . . , xn , y1 , . . . , ym ) = 0.

14
1.4. Transcendence Degree Chapter 1. Preliminaries

Then we may write X a n+1 a n+m


f= fi1 ,...,im (t1 , . . . , tn )tn+1 · · · tn+m
i1 ,...,im

where each fi1 ,...,im ∈ k[t1 , . . . , tn ], aj ≥ 0 and the sum is over all tuples (i1 , . . . , im ) ∈
Nm . Notice that f (E1 , tn+1 , . . . , tn+m ) ∈ K[tn+1 , . . . , tn+m ] so because E2 is algebraically
independent over K and f (E1 , E2 ) = 0, we must have f (E1 , tn+1 , . . . , tn+m ) ≡ 0. Therefore
each fi1 ,...,im (E1 ) = 0 but because E1 is algebraically independent over k, each fi1 ,...,im ≡ 0.
Hence f ≡ 0, showing E is algebraically independent over k.
To finish the proof, note that K/k(E1 ) and L/K(E2 ) are algebraic by Lemma 1.4.3(c),
and we have a tower k(E) ⊂ K(E2 ) ⊂ L in which each intermediate extension is algebraic,
so L/k(E) is algebraic by transitivity. Hence E is a transcendence basis for L/k. It follows
that tr degk L = |E| = |E1 | + |E2 | = tr degk K + tr degK L.

Proposition 1.4.9. If k is a perfect field, K = k(E) and tr degk K = r, then the elements
of E can be numbered such that k(x1 , . . . , xr )/k is purely transcendental and K/k(x1 , . . . , xr )
is a separable algebraic extension.

Proof. If char k = 0 then this follows from Corollary 1.4.6 and (b) of Lemma 1.4.3. If
char k = p > 0 then k = k p and the result follows from Proposition 4.1 in Lang.

15
Chapter 2

Integral Dependence

A critical concept in algebraic number theory and algebraic geometry is the idea of integrality:
the property that an element (or a collection of elements) satisfies a polynomial identity with
coefficients in a ring. In this chapter, we define integrality and explore the basic properties
of integral extensions of rings, eventually proving the ‘going up’ and ‘going down’ theorems
(Section 2.3) and Noether’s Normalization Lemma (Section 2.4). These results have vital
consequences in later chapters.

16
2.1. Integral Extensions of Rings Chapter 2. Integral Dependence

2.1 Integral Extensions of Rings


In this section we define and outline the main results about integrality over a ring A. Let
A ⊆ B be a ring extension, i.e. A is a subring of B.

Lemma 2.1.1. If M is a B-module which is finitely generated as an A-module, then for any
x ∈ B, there exists a monic polynomial f (t) ∈ A[t] of degree at least 1 such that f (x)M = 0.
Pn
Proof. Let M = i=1 Ami . Since M is a B-module, there exist elements aij ∈ A for
1 ≤ i, j ≤ n, such that x satisfies the equations
n
X
xmi = aij mj for each 1 ≤ i ≤ n.
j=1

Pn
On the other hand, x also satisfies xmi = j=1 xδij mj for each i, so subtracting these
equations gives us a system
n
X
(xδij − aij )mj = 0 for each 1 ≤ i ≤ n.
j=1

Set cij = xδij − aij ∈ B and consider the matrix C = (cij ) ∈ Mn (B). Then the system of
equations above can be written
   
m1 0
 ..   .. 
C  .  = . (∗)
mn 0

From linear algebra, there exists an adjoint matrix adj(C) ∈ Mn (B) such that adj(C)C =
(det C)In . Multiplying through (∗) on the left by adj(C) gives us
   
m1 0
 ..   .. 
(det C)  .  =  . 
mn 0

Since the mi generated M , this proves that (det C)M = 0. Now, expanding det C using the
Leibniz formula for a determinant yields
X
det C = sgn(σ)cσ(1)1 · · · cσ(n)n
σ∈Sn
Yn
= (x − aii ) + a0n−2 xn−2 + . . . + a00 for some a0i ∈ A
i=1
= x + an−1 xn−1 + . . . + a0 for some ai ∈ A,
n

after combining coefficients. Set f (t) = tn + an−1 tn−1 + . . . + a0 . Then f (x)M = 0.

17
2.1. Integral Extensions of Rings Chapter 2. Integral Dependence

Definition. An element x ∈ B is integral over A if there exists a monic polynomial


f (t) ∈ A[t] of degree at least 1 such that f (x) = 0. If every element in B is integral over A
then we say B is integral over A, or B is an integral extension of A.

Lemma 2.1.2. For x ∈ B the following are equivalent:

(i) x is integral over A.

(ii) A[x] is a finitely generated A-module.

(iii) There exists a subring C ⊆ B containing A[x] such that C is finitely generated as an
A-module.

(iv) There exists a faithful A[x]-module M which is finitely generated as an A-module.

Proof. (i) =⇒ (ii) If xn +an−1 xn−1 +. . .+a1 x+a0 = 0 for ai ∈ A then xn ∈ M := n−1 i
P
i=0 Ax
which is a finitely generated A-module. By induction, for all m ≥ n, xm ∈ M which proves
A[x] = M . Hence A[x] is finitely generated.
(ii) =⇒ (iii) Let C = A[x]. Then C is finitely generated as an A-module by assumption.
(iii) =⇒ (iv) Let M = C. Then since C is a subring, 1A ∈ C and hence C is a faithful
A-module.
(iv) =⇒ (i) Lemma 2.1.1 implies that f (x)M = 0 for some monic polynomial f (t) ∈ A[t],
but if M is faithful, f (x) = 0. Hence x is integral over A.

Corollary 2.1.3. Let B/A be a ring extension and suppose x1 , . . . , xn ∈ B are integral over
A. Then A[x1 , . . . , xn ] is a finitely generated A-module.

Proof. We prove this by induction on n. The base case for n = 1 follows from (ii) of
Lemma 2.1.2. For n ≥ 2, suppose C := A[x1 , . . . , xn ] is finitely generated over A. Then
there exist elements y1 , . . . , y` ∈ C such that
`
X
C = A[x1 , . . . , xn−1 ] = Ayi .
i=1
Pm−1
By hypothesis, xn is integral over A and therefore over C as well. Then C[xn ] = j=0 Cxjn
for some m ∈ N. Thus
m−1
XX `
A[x1 , . . . , xn ] = C[xn ] = Ayi xjn ,
j=0 i=1

so in particular, A[x1 , . . . , xn ] is a finitely generated A-module.

Definition. For a ring extension B/A, the set A0 of elements x ∈ B that are integral over
A is called the integral closure of A in B. If A0 = A, we say A is integrally closed in
B. Further, if A0 = B then B/A is called an integral extension.

Example 2.1.4. By Corollary 2.1.3, A[x1 , . . . , xn ] is an integral extension of A for any set
of algebraic integers x1 , . . . , xn over A.

18
2.1. Integral Extensions of Rings Chapter 2. Integral Dependence

Lemma 2.1.5. Let B/A be a ring extension. Then A0 , the integral closure of A in B, is a
subring of B which contains A.
Proof. It is trivial that A ⊆ A0 , for every a ∈ A is a root of t − a ∈ A[t]. Suppose x, y ∈ A0 .
Then by Corollary 2.1.3, A[x, y] is a finitely generated A-module which is a subring of B.
Clearly x + y, xy ∈ A[x, y] so applying Lemma 2.1.2(iii) shows that x + y and xy are integral
over A.
Proposition 2.1.6. If B/A and C/B are integral extensions of rings, then C/A is also an
integral extension.
Proof. Take x ∈ C. We must show that x is integral over A. First, x is integral over B so
there exist elements b0 , . . . , bn−1 ∈ B such that
xn + bn−1 xn−1 + . . . + b0 = 0.
Then since B/A is integral, x is integral over B 0 := A[b0 , . . . , bn−1 ]. By Corollary 2.1.3, B 0
is finitely generated as an A-module since all the bi are integral over A. Then there are
elements y1 , . . . , y` ∈ B 0 which generate B 0 as an A-module. Also, B 0 [x] is a B 0 -module
which is finitely generated by xj , 0 ≤ j ≤ m − 1. Then we have
m−1
X m−1
XX `
0 0 j
A[x] ⊆ B [x] = Bx = Ayi xj .
j=0 j=0 i=1

Thus by Lemma 2.1.2(iii), x is integral over A. This proves C/A is an integral extension.
Corollary 2.1.7. If A0 is the integral closure of A in B then A0 is integrally closed in B.
Proof. Take x ∈ B which is integral over A0 . Then A0 [x]/A0 is integral by Corollary 2.1.3 and
A0 /A is integral by definition of A0 . By Proposition 2.1.6, this implies A0 [x]/A is integral, so
in particular x is integral over A. Thus x ∈ A0 .
Definition. An integral domain A is called integrally closed (or alternatively, normal)
if A is integrally closed in its field of fractions.
Proposition 2.1.8. Every UFD is integrally closed.
Proof. Let A be a UFD and denote the field of fractions of A by K. Take x ∈ K and write
x = ab for a, b ∈ A and b 6= 0; we may assume a and b are relatively prime, that is, there
is no prime p ∈ A that divides both a and b. Suppose x is integral over A. Then there are
elements a0 , . . . , an−1 ∈ A such that
 a n  a n−1
+ an−1 + . . . + a0 = 0.
b b
Multiplying through by bn gives us
an + an−1 b + . . . + a0 bn = 0.
Subtracting an to the other side shows that b | an in A, but since a and b are relatively
prime, and therefore an and b are also relatively prime, this is only possible when b is a unit.
Hence b−1 exists in A, so x = ab = ab−1 ∈ A, proving A is integrally closed.

19
2.1. Integral Extensions of Rings Chapter 2. Integral Dependence

Example 2.1.9. Both the integers Z and the Gaussian integers Z[i] are Euclidean domains,
so in particular they are UFDs and therefore integrally closed by Proposition 2.1.8.

Example 2.1.10. Let K be √ a quadratic extension of Q, i.e. a field extension such that
[K : Q] = 2. Then K = Q( d) for some squarefree integer d ∈ Z r {0, 1}, meaning d has
prime factorization d = ±p1 p2 · · · pr for primes pi . We want to compute the integral closure
of Z in K. Since K/Q is Galois, there is one nontrivial Q-automorphism σ ∈ Gal(K/Q):

σ : K −→ K
√ √
a + b d 7−→ a − b d.

Using the norm and trace maps for K/Q, we have that

α ∈ K is integral over Z ⇐⇒ N (α) = ασ(α) and Tr(α) = α + σ(α) ∈ Z.

To verify this, suppose f (α) = 0 for a monic polynomial f (t) ∈ Z[t]. Then 0 = σ(f (α)) =
f (σ(α)), so σ(α) is integral over Z. Since the integral elements form a ring, N (α) = ασ(α)
and Tr(α) = α + σ(α) are also integral over Z. But the norm and trace maps take values in
Q, so N (α), Tr(α) ∈ Q and they’re integral, so both lie in Z.
Conversely, if ασ(α), α + σ(α) ∈ Z then α is a root of the monic polynomial

(t − α)(t − σ(α)) = t2 − (α + σ(α))t + ασ(α) ∈ Z[t].

Therefore α is integral.
For any number field K/Q, we let OK denote the integral closure of Z in K, called the
ring of integers of K. In the quadratic case, our work above shows that

OK = {a + b d | a, b ∈ Q and a2 − b2 d, 2a ∈ Z}
( √
Z[h d] i if d ≡ 2, 3 (mod 4)
= √
Z 1+2 d if d ≡ 1 (mod 4).

1+ d
To see how the second case arises, suppose d ≡ 1 (mod 4). Then ω = 2
satisfies the
integral expression ω 2 − ω + 1−d
4
= 0.

20
2.2. Integrality and Field Extensions Chapter 2. Integral Dependence

2.2 Integrality and Field Extensions


In this section, let A and B be integral domains and let K be the field of fractions of A.

Lemma 2.2.1. If B/A is an integral extension then A is a field if and only if B is a field.

Proof. ( =⇒ ) Assume A is a field and choose x ∈ B r {0}. Then x is integral over A so


there exist elements a0 , . . . , an−1 ∈ A such that xn + an−1 xn−1 + . . . + a0 = 0. This can be
rewritten as
x(xn−1 + an−1 xn−2 + . . . + a1 ) = −a0 (∗)
Choose such an expression for x with n minimal. By assumption, x 6= 0 and since n was
minimal, xn−1 + an−1 xn−2 + . . . + a1 6= 0. Then since B is an integral domain, (∗) implies
a0 6= 0, and since A is a field, a−1
0 ∈ A exists. Divide through (∗) by −a0 to obtain

x−1 := −a−1
0 (x
n−1
+ . . . + a1 ).

This shows x is invertible in B.


( ⇒= ) Conversely, suppose B is a field and take a ∈ A r {0}. Then a−1 ∈ B exists.
Since B/A is integral, there exist elements ai ∈ A such that

a−n + an−1 a−(n−1) + . . . + a0 = 0 ⇐⇒ a−1 = −(an−1 + . . . + a0 an−1 ) ∈ A.

This shows that a is invertible in A.

Lemma 2.2.2. Let A be integrally closed in its fraction field K, let L/K be a finite extension
and suppose α ∈ L is algebraic over K. Then α is integral over A if and only if α has a
minimal polynomial f (t) ∈ A[t].

Proof. ( ⇒= ) is the definition of integrality.


( =⇒ ) Take a splitting field E of f over K. Then f splits into linear factors:
n
Y
f (t) = (t − αi ) for αi ∈ E, α1 = α.
i=1

Since α is integral, there are K-isomorphisms

σi : K(α) −→ K(αi ), α 7→ αi ,

using the fact that K(αi ) ∼


= K[t]/(f (t)) for each root αi . By assumption, there exists a
monic polynomial g(t) ∈ A[t] such that g(α) = 0. Applying σi to this equation gives

0 = σi (g(α)) = g(σi (α)) = g(αi ).

Thus αi is integral over A for all 1 ≤ i ≤ n. It follows that f (t) = ni=1 (t − αi ) ∈ K[t]
Q
has coefficients which are integral over A. Finally, since A is integrally closed, this proves
f ∈ A[t].

21
2.2. Integrality and Field Extensions Chapter 2. Integral Dependence

Remark. The most important rings in algebraic number theory are rings of integers: for a
finite extension K/Q, OK denotes the integral closure of Z in K. By Lemma 2.2.2, we can
write OK = {α ∈ K | fα ∈ Z[t]} where fα denotes the minimal polynomial of α over Q.
Definition. For any finite field extension L/K of degree n = [L : K], there is a K-algebra
homomorphism

ϕ : L ,→ EndK (L) ∼= Mn (K)


α 7→ (ϕα : x 7→ αx).

The trace form of the extension L/K is defined as

Tr = TrL/K : L −→ K
α 7−→ tr(ϕα ).

By properties of the matrix trace tr, the trace form TrL/K is K-linear.
Lemma 2.2.3. A finite extension L/K is separable if and only if TrL/K 6= 0.
Proof. We will only show the forward direction. Suppose L/K is separable. Then L = K(α)
for some α ∈ L by the primitive element theorem. We claim that for some 0 ≤ i ≤ n − 1,
Tr(αi ) 6= 0. Let
Yn
f (t) = (t − αj )
j=1

be the minimal polynomial of α over K. By separability, f has distinct roots α = α1 , . . . , αn


in some splitting field E of f over K. By the Cayley-Hamilton theorem, the characteristic
polynomial χ of ϕα satisfies 0 = χ(ϕ(α)) = ϕ(χ(α)). Thus χ(α) = 0 by injectivity of ϕ.
Hence f (t) | χ(t) in K[t] but deg f = n = deg χ and both polynomials are monic, so we have
f = χ. This implies that α1 , . . . , αn are the distinct eigenvalues
P of ϕα in E. Let Mα be the
matrix representing ϕα in Mn (K). Then Tr(α) = tr(Mα ) = nj=1 αj . Taking powers of Mα ,
we have n
X
i i
Tr(α ) = tr(Mα ) = αji .
j=1

Consider the n × n matrix C = (αji )


where 0 ≤ i ≤ n − 1 and 1 ≤ j ≤ n. Then C ∈ Mn (E)
and its determinant is a Vandermonde determinant:
Y
det C = (αj − α` )2
`<j

which is nonzero because the αj are distinct. In particular,


       
0 1 1 + ... + 1 Tr(1)
 ..  .  .. ..
 .  6= C  ..  =  = .
  
. .
n−1 n−1 n−1
0 1 α1 + . . . + αn Tr(α )

So Tr(αi ) 6= 0 for some i. Therefore the trace form on L/K is nondegenerate.

22
2.2. Integrality and Field Extensions Chapter 2. Integral Dependence

When Tr 6= 0, one may define a nondegenerate, symmetric, K-bilinear form b : L×L → K


by b(x, y) = TrL/K (xy).
Lemma 2.2.4. For any β ∈ L, the characteristic polynomial χ of ϕβ satisfies
χ(t) = (f (t))[L:K(β)]
where f (t) is the minimal polynomial of β over K. Therefore TrL/K (β) = [L : K(β)] TrK(β)/K (β).
Proof. Let ϕβ be the map L → L given by ϕβ (x) = βx. Set m = [L : K(β)] and n = [K(β) :
K]. If m = 1, then L = K(β) so deg f (t) = [L : K] = n. By Cayley-Hamilton, f (t) divides
χ(t), but both are by definition monic polynomials and deg χ(t) = n, so we have χ(t) = f (t).
In the general case, let {x1 , . . . , xr } be a basis of K(β)/K and let {y1 , . . . , ym } be a basis
of L/K(β). We know the set {xi yj : 1 ≤ i ≤ r, 1 ≤ j ≤ m} is a basis forPL/K. Let B = (bk` )
be the standard matrix for ϕβ in the extension K(β)/K. Then βxi = rk=1 bki xk and thus
r
X
β(xi yj ) = bki (xk yj ).
k=1

If we write the basis {xi yj } as {x1 y1 , x2 y1 , . . . , xr y1 , x1 y2 , . . . , xr ym } then ϕβ has the following


standard matrix in L/K:  
B
 B
0 
 
 . . 

0 . 
B
There are m blocks, each of which is B, so χ(t) = [det(B − tI)]m but by the m = 1 case,
det(B − tI) = f (t). Therefore χ(t) = (f (t))m as claimed.
Proposition 2.2.5. If A is an integrally closed integral domain with K its fraction field,
and L/K is a finite separable extension, then for Pthe integral closure B of A in L, there
exists a K-basis {v1 , . . . , vn } of L such that B ⊆ ni=1 Avi .
Proof. If x ∈ L then we claim there is some a ∈ A r {0} such that ax ∈ B. Since x is
algebraic, there exist q0 , . . . , qn−1 ∈ K such that
xn + qn−1 xn−1 + . . . + q0 = 0.
Then a may be chosen to be the common denominator of the qi , for which we have
(ax)n + aqn−1 (ax)n−1 + . . . + an q0 = 0.
Each coefficient of this expression now lies in A, so ax is integral over A and thus lies in B.
Now this implies that there is a K-basis {u1 , . . . , un } of L with ui ∈ B for each 1 ≤ i ≤ n.
Since L/K is separable, Lemma 2.2.3 says that TrL/K 6= 0 and so b(·, ·) is nondegenerate.
Pn a dual basis {v1 , . . . , vn } of L with respect to b(·, ·). Let x ∈ B. Then we
Then there exists
can write x = j=1 cj vj for cj ∈ K. We finish by showing each cj ∈ A. Consider
n
! n
X X
Tr(ui x) = b(ui , x) = b ui , cj vj = cj b(ui , vj ) = ci .
j=1 j=1

23
2.2. Integrality and Field Extensions Chapter 2. Integral Dependence

Now each ui , x ∈ B so ui x ∈ B and thus ui x is integral over A. Up to sign, Tr(ui x) is a


coefficient of the characteristic polynomial χ of ϕ(ui x). By Lemma 2.2.4, χ is a power of the
minimal polynomial of ui x, which according to Lemma Pn 2.2.2 has coefficients in A. Hence
χ ∈ A[t] so Tr(ui x) = ci ∈ A. This shows that x ∈ i=1 Avi as desired.

Corollary 2.2.6. If A is a PID and L/K is a finite separable extension, where K is the
fraction field of A, then the integral closure B of A in L is a finitely generated free A-module
of rank n = [L : K].

Proof. First, by Proposition 2.1.8, every PID is integrally closed so we may apply Proposi-
tion 2.2.5 and its proof to produce K-bases {u1 , . . . , un } and {v1 , . . . , vn } of L such that
n
X n
X
Aui ⊆ B ⊆ Avi .
i=1 i=1

The
Pn vi form a K-basis, so in particular they are linearly independent over A and thus
i=1 Avi is a free A-module of rank n. By the theory of modules over a PID, we have that
B is free of rank at most n over A, but reversing the roles of the ui and vi , we see that the
rank of B equals n.

Example 2.2.7. If K/Q is finite, then there exists a Q-basis of K, say {α1 , . . . , αn }, which
is a Z-basis of OK . Such a basis is called an integral basis of K (or of OK ). In particular,
Corollary 2.2.6 shows that OK is a free abelian group of rank
√ n = [K : Q].
In the special case of a quadratic extension, K = Q( d), the ring of integers is OK =
Z ⊕ Zω, with (√
d if d ≡ 2, 3 (mod 4)
ω = 1+√d
2
if d ≡ 1 (mod 4).
Thus an integral basis of OK is {1, ω}.

Remark. Proposition 2.2.5 and Corollary 2.2.6 are not true in general if L/K is not a
separable extension, as we will show in Section 4.5.

24
2.3. Integrality, Ideals and Localization Chapter 2. Integral Dependence

2.3 Integrality, Ideals and Localization


In this section we develop the theory of ideals in integral extensions and explore in particular
how these notions interact with localization. We will prove the famous ‘going up’ and ‘going
down’ theorems.

Lemma 2.3.1. Assume B/A is a ring extension with I ⊂ A and J ⊂ B ideals such that
J ∩ A = I. Then

(a) B/J is integral over A/I.

(b) If I and J are prime ideals then I is maximal in A if and only if J is maximal in B.

Proof. (a) We have a sequence A ,→ B  B/J where the kernel of A → B/J is A ∩ J = I.


Therefore A/I injects into B/J so the ring extension (B/J)/(A/I) makes sense. Take
x̄ ∈ B/J with x̄ = x + J for some x ∈ B. Then there exist ai ∈ A such that

xn + an−1 xn−1 + . . . + a0 = 0 mod J


=⇒ x̄n + ān−1 x̄n−1 + . . . + ā0 = 0.

The coefficients of the second equation are in A/I, so we see that x̄ is integral over A/I.
(b) If I and J are prime ideals then A/I and B/J are integral domains. By (a), B/J is
integral over A/I, and by Lemma 2.2.1, B/J is a field exactly when A/I is a field. Hence J
is maximal in B if and only if I is maximal in A.

Lemma 2.3.2. Let B/A be a ring extension and S ⊆ A be a multiplicatively closed subset.

(a) If B/A is integral then S −1 B/S −1 A is integral.

(b) If C is the integral closure of A in B then S −1 C is the integral closure of S −1 A in


S −1 B.

Proof. First, there is a canonical map S −1 A → S −1 B, as 7→ as which is a well-defined, one-


to-one ring homomorphism. As a consequence, we can regard S −1 A as a subring of S −1 B.
(a) Let xs ∈ S −1 B where x ∈ B and s ∈ S. Given that x is integral over A, there exist
a0 , . . . , an−1 ∈ A such that

xn + an−1 xn−1 + . . . + a0 = 0 in B.

Applying the map j : B → S −1 B to this expression and dividing by sn gives us


 x n a  x n−1 a0
n−1
+ + . . . + n = 0 in S −1 B.
s s s s
Since each coefficient lies in S −1 A, this shows that xs is integral over S −1 A and thus the
extension S −1 B/S −1 A is integral.
(b) We have A ⊆ C ⊆ B so by the preliminary remark, we may consider the tower of
ring extensions S −1 A ⊆ S −1 C ⊆ S −1 B. Since C/A is integral, part (a) implies S −1 C/S −1 A

25
2.3. Integrality, Ideals and Localization Chapter 2. Integral Dependence

is integral. If x
s
∈ S −1 B is integral over S −1 A, there exist fractions a0
s0
, . . . , asn−1
n−1
∈ S −1 A such
that  x n a  x n−1
n−1 a0
+ + ... + = 0.
s sn−1 s s0
Multiplying through by sn and the common denominator, we can write

t0 xn + a0n−1 xn−1 + . . . + a00 = 0 in S −1 B

where t0 ∈ S and a0i ∈ A. Then there is some t00 ∈ S such that

t00 (t0 xn + a0n−1 xn−1 + . . . + a00 ) = 0 in B.

Setting t = t0 t00 ∈ S, we see that

txn + a00n−1 xn−1 + . . . + a000 = 0 in B, where a00i ∈ A.

Multiplying through by tn−1 gives us

(tx)n + a00n−1 (tx)n−1 + . . . + a000 tn−1 = 0.

Therefore tx is integral over A, so tx ∈ C. This means x


s
= tx
ts
∈ S −1 C, so we conclude that
S −1 C is the integral closure of S −1 A in S −1 B.
Note that Lemma 2.3.2(b) does not imply that if B/A is an integral extension of rings
with Q ⊂ B prime and Q ∩ B = p, then BQ /Ap is integral.

Lemma 2.3.3. Suppose B/A is an integral extension of rings. If Q, Q0 ⊂ B are prime ideals
with Q ⊆ Q0 and Q ∩ A = Q0 ∩ A, then Q = Q0 .

Proof. Set p = Q ∩ A = Q0 ∩ A. Then p is a prime ideal of A. If S = A r p then by


Lemma 2.3.2(a), S −1 B/Ap is an integral extension. By Corollary 1.3.9, S −1 p is the unique
maximal ideal of the local ring Ap . Notice that S ∩ Q0 = S ∩ A ∩ Q0 = S ∩ p = ∅, and
likewise S ∩Q = ∅. Therefore S −1 Q ⊆ S −1 Q0 are prime ideals of S −1 B by Proposition 1.3.8.
Also, 11 6∈ S −1 Q0 so S −1 Q0 ∩ S −1 A 6= S −1 A. Thus S −1 Q0 ∩ S −1 A = S −1 p, and likewise
S −1 Q ∩ S −1 A = S −1 p. Since S −1 p is maximal in S −1 A, Lemma 2.3.1(b) implies S −1 Q and
S −1 Q0 are both maximal ideals in S −1 B. Therefore S −1 Q = S −1 Q0 so Proposition 1.3.8(c)
shows that we must have Q = Q0 .

Lemma 2.3.4. Suppose B/A is an integral extension of rings and p ⊂ A is a prime ideal.
Then there exists a prime ideal Q ⊂ B with Q ∩ A = p.

Proof. Let S = A r p and consider the commutative diagram


A B

i j

Ap = S −1 A S −1 B

26
2.3. Integrality, Ideals and Localization Chapter 2. Integral Dependence

Choose any maximal ideal m ⊂ S −1 B and set Q = j −1 (m) which is a prime ideal of B by
Proposition 1.3.8. We claim that Q ∩ A = p. Since B is integral over A, Lemma 2.3.2(a)
implies S −1 B is integral over S −1 A = Ap . Then by Lemma 2.3.1(b), m maximal in S −1 B
implies m ∩ Ap is maximal in Ap . But Ap is a local ring, so m ∩ Ap = S −1 p. Since the first
diagram commutes, we have another commutative diagram:
Q∩A Q

i j

S −1 p = m ∩ Ap m

Therefore Q ∩ A = j −1 (m) ∩ A = i−1 (m ∩ Ap ) = i−1 (S −1 p). Finally, Proposition 1.3.8(b)


shows that Q ∩ A = p as desired.
Lemmas 2.3.3 and 2.3.4 together show that in an integral extension B/A, any prime ideal
p ⊂ A lifted to B is contained in a prime ideal of B which is unique in any chain of ideals
of B. This generalizes in an important way, in what is known as the going up theorem.

Theorem 2.3.5 (Going Up). Suppose B/A is an integral extension of rings and

p0 ( p1 ( · · · ( pn

is a chain of prime ideals in A, in which each inclusion is strict. Then for any chain of
prime ideals
Q0 ( Q1 ( · · · ( Qm−1
in B such that 0 ≤ m ≤ n and Qi ∩ A = pi for each 0 ≤ i ≤ m − 1, there exists a prime
ideal Qm ⊂ B such that Qm−1 ( Qm and Qm ∩ A = pm .

Proof. When m = 0, Lemma 2.3.4 says there exists a prime ideal Q0 ⊂ B with Q0 ∩ A = p0 .
Now to induct, suppose the theorem holds for some m, 1 ≤ m ≤ n. Then Qm−1 ∩ A = pm−1
so consider the extension of quotient rings

Ā = A/pm−1 ,−→ B̄ = B/Qm−1 .

By Lemma 2.3.1(a), B̄/Ā is an integral extension. By the correspondence theorem, p̄m :=


pm /pm−1 is a prime ideal of Ā. Then by Lemma 2.3.4, there exists a prime ideal Qm ⊂ B̄
such that Qm ∩ Ā = p̄m . Let Qm be the preimage π −1 (Qm ) where π : B → B̄ is the canonical
projection. Clearly Qm ) Qm−1 since p̄m 6= 0. Finally, the above work implies Qm ∩ A = pm .
Hence the statement holds by induction.

Definition. If A is a ring and p ⊂ A is prime, the height of p is defined as

ht(p) = sup{` ≥ 0 | there is a chain of prime ideals p0 ( p1 ( · · · ( p`−1 ( p}.

The Krull dimension of A is then defined by

dim A = sup{ht(p) | p ⊂ A is prime}.

27
2.3. Integrality, Ideals and Localization Chapter 2. Integral Dependence

Remark. If p is not maximal then p ( m for some maximal ideal m ⊂ A, so we can


reformulate the Krull dimension definition in one of the following ways:

dim A = sup{ht(m) | m ⊂ A is maximal}


= sup{` ≥ 0 | p0 ( p1 ( · · · ( p` is a chain of prime ideals}.

Examples.

1 If K is a field then dim K = 0 since (0) is the only prime ideal.

2 If A is an integral domain, then dim A = 0 if and only if A is a field. However, there


exist non-domains which have dimension 0.

3 If A is a PID but not a field, then all nonzero prime ideals are of the form (π) for a
prime element π ∈ A. This shows that dim A = 1.

Lemma 2.3.6. Suppose p ⊂ A is a prime ideal. Then

(a) ht(p) = dim Ap .

(b) ht(p) + dim A/p ≤ dim A.

Proof. (a) By Proposition 1.3.8(c), prime ideals of Ap are in bijective, inclusion-preserving


correspondence with primes Q ⊂ A intersecting S = A r p trivially, that is, primes Q ⊆ p.
Therefore any chain p0 ( p1 ( · · · ( pn−1 ( p of prime ideals in A corresponds to a
chain S −1 p0 ( S −1 p1 ( · · · ( S −1 pn−1 ( S −1 p of prime ideals in Ap = S −1 p; this proves
dim Ap ≥ ht(p). Conversely, any chain Q0 ( Q1 ( · · · ( Q`−1 ( S −1 p of prime ideals in
Ap corresponds to a chain j −1 (Q0 ) ( j −1 (Q1 ) ( · · · ( j −1 (Q`−1 ) ( p in A; this establishes
ht(p) ≥ dim Ap . Hence ht(p) = dim Ap .
(b) Suppose dim A/p = ` and let Q0 ( Q1 ( · · · ( Q` be a maximal chain of prime
ideals in A/p. By the correspondence theorem, this lifts to a chain of prime ideals

Q0 ( Q1 ( · · · ( Q` (∗)

in A such that each Qi ⊇ p. Of course we can always add p to the bottom of such a chain,
so if ` is maximal, we must have Q0 = p. Now for any chain of prime ideals p0 ( p1 ( · · · (
pn−1 ( p in A, we can extend (∗) by this new chain, giving

p0 ( p1 ( · · · ( pn−1 ( p = Q0 ( Q1 ( · · · ( Q` ,

a chain of prime ideals in A. This shows dim A ≥ n + `. In particular, this holds for any
such n ≤ ht(p), so we have

dim A ≥ ht(p) + ` = ht(p) + dim A/p.

Theorem 2.3.7. If B is an integral extension of A then dim A = dim B.

28
2.3. Integrality, Ideals and Localization Chapter 2. Integral Dependence

Proof. If Q0 ( Q1 ( · · · ( Q` is a chain of prime ideals in B, then

Q0 ∩ A ( Q1 ∩ A ( · · · ( Q` ∩ A

is a chain of prime ideals in A – the strictness of each containment comes from Lemma 2.3.3.
This implies dim B ≤ dim A. On the other hand, if p0 ( p1 ( · · · ( p` is a chain of prime
ideals in A, then by the going up theorem, we get a chain of prime ideals

Q0 ( Q1 ( · · · ( Q`

in B, with Qi ∩ A = pi for each i. Hence dim A ≤ dim B so we have dim A = dim B.


There is an analog of the going up theorem for descending chains of prime ideals, but it
requires additional conditions. First we prove two important lemmas.

Lemma 2.3.8. For a ring extension B/A and any prime ideal p ⊂ A, there exists a prime
ideal Q ⊂ B with the property that Q ∩ A = p if and only if pB ∩ A ⊆ p.

Proof. ( =⇒ ) If Q ⊂ B exists and satisfies Q ∩ A = p, then pB ∩ A = (Q ∩ A)B ∩ A ⊆


Q ∩ A = p.
( ⇒= ) Set T = A r p. By assumption, pB ∩ A ⊆ p implies (pB ∩ A) ∩ T = ∅. Therefore
T −1 (pB) 6= T −1 B so there is a maximal ideal m ⊂ T −1 B such that T −1 (pB) ⊆ m. By
Proposition 1.3.8, there is a prime ideal Q ⊂ B such that Q ∩ T = ∅ and m = T −1 Q. Then
p ⊆ pB ∩ A ⊆ Q ∩ A = Q ∩ p ⊆ p so it follows that p = Q ∩ A as required.

Lemma 2.3.9. Suppose A and B are integral domains with B/A an integral extension of
rings, K the field of fractions of A and A integrally closed (in K). If p ⊂ A is a prime ideal
of A and α ∈ pB is any element, then the nonleading coefficients of the minimal polynomial
f of α over K lie in p.

Proof. By Lemma 2.2.2, the coefficients of f must lie in A. Let α = b1 p1 + . . . + bk pk for


bi ∈ B, pi ∈ p. Replacing B with A[b1 , . . . , bk ], we may assume BPis a finitely generated
n
Pn generated by x1 , . . . , xn . For each 1 ≤ i ≤ n, write αxi = j=1 aij xj with aij ∈ p.
A-module
Then j=1 (δij αxi −aij ) = 0. Multiplying by the classical adjoint of the matrix (δij αxi −aij ),
we get
det(δij αxi − aij )xj = 0 for all 1 ≤ j ≤ n.
Therefore αxi satisfies a monic polynomial with coefficients in p. Since this holds for every
generator xi , it follows that α itself satisfies such a polynomial, say g ∈ p[t]. If f is the
minimal polynomial of α over K, f must divide g, say g = f h for h ∈ A[t]. Reducing
modulo p, we get
tdeg g = ḡ = f¯ h̄ ∈ (A/p)[t].
Since p is prime, A/p is a domain, so f¯ and h̄ must also be powers of t. Thus all nonleading
terms of f lie in p.

Theorem 2.3.10 (Going Down). Suppose B/A is an integral extension of integral domains
with A integrally closed. If
p0 ) p1 ) · · · ) pn

29
2.3. Integrality, Ideals and Localization Chapter 2. Integral Dependence

is a chain of prime ideals in A and

Q0 ) Q1 ) · · · ) Qm−1

is a chain of prime ideals in B such that 0 ≤ m ≤ n and Qi ∩ A = p for each i ≤ m − 1,


then there is a prime ideal Qm ⊂ B such that Qm ( Qm−1 and Qm ∩ A = pm .

Proof. If m = 0, this is given to us by Lemma 2.3.4. Otherwise assume 1 ≤ m ≤ n. Then


since B is an integral domain, the map B → BQm−1 is injective. Set B 0 = BQm−1 = S −1 B,
where S = B r Qm−1 . Then we get a tower of ring extensions A ⊆ B ⊆ B 0 . It suffices to
show there exists a nonzero prime ideal Q0 ⊂ B 0 such that Q0 ∩ A = pm , for then we will have
pm = Q0 ∩ A = (Q0 ∩ B) ∩ A = Qm ∩ A for some prime ideal Qm ⊂ B, by Proposition 1.3.8.
Set p = pm . By Lemma 2.3.8, showing the existence of Q0 ⊂ B 0 satisfying Q0 ∩ A = p is
equivalent to showing pB 0 ∩ A ⊆ p. If pB 0 ∩ A = 0, we are done. Otherwise, take a nonzero
element a ∈ pB 0 and write a = αs for some α ∈ pB and s ∈ S. Also let f be the minimal
polynomial of α over K. Then Lemma 2.3.9 implies f = tn + cn−1 tn−1 + . . . + c0 for ci ∈ p.
Since a ∈ A, we may write
1 cn−1 n−1 c0
n
f (at) = tn + t + . . . + n = tn + c0n−1 tn−1 + . . . + c00 ,
a a a
ci
with c0i = an−i
. Then 1
an
f (at) is also irreducible, and plugging in t = s gives

1 1
n
f (as) = n f (α) = 0.
a a
It follows that a1n f (at) is the minimal polynomial of s over K. Thus each c0i lies in A for
0 ≤ i ≤ n − 1. Suppose that a 6∈ p. Then since p is prime, c0n−i ai = cn−i ∈ p implies c0n−i ∈ p
for each 0 ≤ i ≤ n − 1. Since s ∈ B, we get

sn = −c0n−1 sn−1 − . . . − c00 ∈ pB 0 ⊆ pm−1 B ⊆ Qm−1

by hypothesis. Then s ∈ Qm−1 since Qm−1 is a prime ideal, but this contradicts s ∈ S =
B r Qm−1 . Hence a ∈ p, implying pB 0 ∩ A ⊆ p after all.
For the remainder of the section, let A be an integral domain with field of fractions K.
Whenever S ⊆ Ar{0}, we can consider the local ring S −1 A as a subring of K. In particular,
Ap ⊆ K for every prime p ⊂ A.
\
Lemma 2.3.11. Any integral domain A can be written A = Am .
maximal ideals
m⊂A

Proof. Let D be the intersection of Am over all maximal ideals m ⊂ A. Then D ⊆ K. For
a fixed x ∈ D, consider the ideal I = {a ∈ A | ax ∈ A} ⊆ A. If m is a fixed maximal ideal
of A, by assumption x ∈ Am so there exist elements a, b ∈ A with a 6∈ m such that x = ab .
It follows that ax = b ∈ A, so a ∈ I. Since we took a 6∈ m, this implies I 6⊂ m. Now m was
arbitrary, so we must have I = A. Thus 1 ∈ I, implying x = 1 · x ∈ A and therefore A = D
as required.

30
2.3. Integrality, Ideals and Localization Chapter 2. Integral Dependence

Proposition 2.3.12. For an integral domain A, the following are equivalent:

(i) A is integrally closed.

(ii) Ap is integrally closed for every prime ideal p ⊂ A.

(iii) Am is integrally closed for every maximal ideal m ⊂ A.

Proof. (i) =⇒ (ii) Take p ⊂ A to be a prime ideal and set S = A r p. By assumption, A


is the integral closure of A in K, and by Lemma 2.3.2(b), S −1 A = Ap is integrally closed in
K, which is also the field of fractions of Ap .
(ii) =⇒ (iii) is trivial.
(iii) =⇒ (i) If m ⊂ A is a maximal ideal, we have A ⊆ Am ⊆ K. Then the assumption
that Am is integrally closed implies that the integral closure of A in K lies in Am for any
such m. Therefore by Lemma 2.3.11, A is integrally closed.

31
2.4. Normalization Chapter 2. Integral Dependence

2.4 Normalization
Let k be a field and A a finitely generated k-algebra, but not necessarily an integral domain.
Definition. Elements a1 , . . . , an ∈ A are said to be algebraically dependent over k if
there exists a nonzero polynomial f ∈ k[t1 , . . . , tn ] such that f (a1 , . . . , an ) = 0. Otherwise
the ai are said to be algebraically independent over k.
If a1 , . . . , an ∈ A are algebraically independent, then the k-algebra homomorphism
ϕ : k[t1 , . . . , tn ] −→ A
ti 7−→ ai
is injective. As a consequence, k[a1 , . . . , an ] ∼
= k[t1 , . . . , tn ] as k-algebras.
Theorem 2.4.1 (Noether’s Normalization Lemma). If A is a finitely generated k-algebra
with A = k[x1 , . . . , xn ] for xi ∈ A, then there exist algebraically independent elements
y1 , . . . , yr ∈ A, 0 ≤ r ≤ n, such that A/k[y
Pn 1 , . . . , yr ] is an integral extension. Moreover,
if k is infinite, we may choose the yi in j=1 kxj .
Proof. We prove the theorem by induction on n. Starting with n = 1, there are two possi-
bilities:
ˆ x1 is transcendental over k, in which case we may take r = 1 and y1 = x1 .
ˆ x1 is algebraic over k, in which case we may take r = 0, so that k[x1 ]/k is integral.
Now suppose n ≥ 2. If x1 , . . . , xn are already algebraically independent, take r = n and
xi for each 1 ≤ i ≤ n. If not, we claim that there exist elements z1 , . . . , zn−1 ∈ A (and
yi = P
zi ∈ nj=1 kxj if k is infinite) such that the following hold:
(i) k[z1 , . . . , zn−1 , xn ] = A.
(ii) xn is integral over k[z1 , . . . , zn−1 ].
Once we prove the claim, the proof finishes as follows. Set B = k[z1 , . . . , zn−1 ] ⊆ A. Then
A = B[xn ] and A/B by (i) and (ii). By induction, there exist y1 , . . . , yr ∈ B (and
Pis integral P
if k is infinite, yi ∈ n−1 j=1 kzj ⊆ n
j=1 kxj ) such that the yi are algebraically independent over
k and B is integral over k[y1 , . . . , yr ]. It will then follow that A/k[y1 , . . . , yr ] is an integral
extension by Proposition 2.1.6.
To prove the existence of z1 , . . . , zn−1 satisfying (i) and (ii), assume k is infinite. By
assumptionP there is a nonzero polynomial f ∈ k[t1 , . . . , tn ] such that f (x1 , . . . , xn ) = 0.
Write f = di=0 fi (t1 , . . . , tn ) where each fi is homogeneous of (total) degree i and fd 6= 0.
(Since f0 ∈ k, we know d ≥ 1.) We can write fd in two ways:
X
fd (t1 , . . . , tn ) = cJ tj11 · · · tjnn where cJ ∈ k
J=(j1 ,...,jn )∈Nn
0
j1 +...+jn =d
d
X
fd (t1 , . . . , tn ) = Gj (t1 , . . . , tn−1 )tnd−j where each Gj is homogeneous of degree i.
j=0

32
2.4. Normalization Chapter 2. Integral Dependence

Since fd 6= 0, not all of the Gj are 0. This implies fd (t1 , . . . , tn−1 , 1) 6= 0. Since k is infinite,
there exist scalars λ1 , . . . , λn−1 ∈ k such that fd (λ1 , . . . , λn−1 , 1) 6= 0. Plugging these into
the top expression for fd , we get
X jn−1 j1 +...+jn
fd (λ1 xn , . . . , λn−1 xn , xn ) = cJ λj11 · · · λn−1 xn
J
X j
= cJ λj11 · · · λn−1
n−1 d
xn = fd (λ1 , . . . , λn−1 , 1)xdn .
J

Now set zi = xi − λi xn for each 1 ≤ i ≤ n − 1. Then k[z1 , . . . , zn−1 , xn ] = k[x1 , . . . , xn ] and


d
X
0 = f (x1 , . . . , xn ) = fj (x1 , . . . , xn )
j=0
d
X
= fj (z1 + λ1 xn , . . . , zn−1 + λn−1 xn , xn )
j=0

= fd (z1 + λ1 xn , . . . , zn−1 + λn−1 xn , xn ) + lower degree terms


X
= cJ (z1 + λ1 xn )j1 · · · (zn−1 + λn−1 xn )jn−1 xjnn + lower degree terms
J
= fd (λ1 xn , . . . , λn−1 xn , xn ) + lower terms in xn
d−1
X
= fd (λ1 , . . . , λn−1 , 1)xdn + Hj (z1 , . . . , zn−1 )xjn
j=1

where fd (λ1 , . . . , λn−1 , 1)xdn ∈ k × and each Hj ∈ k[t1 , . . . , tn−1 ]. This shows xn satisfies
a nontrivial monic polynomial equation with coefficients in B = k[z1 , . . . , zn−1 ], since in
particular we can divide out by fd (λ1 , . . . , λn−1 , 1) 6= 0. Thus xn is integral over B.
If k is finite, we need a slightly different argument. By the assumption on x1 , . . . , xn ,
there is a nonzero polynomial F ∈ k[t1 , . . . , tn ] such that F (x1 , . . . , xn ) = 0. Write
X j
F (t1 , . . . , tn ) = cj t11 · · · tjnn
j∈Nn
0

for cj ∈ k for all j = (j1 , . . . , jn ) ∈ Nn0 . Set J = {j | cj 6= 0}, which is a finite set since F is a
polynomial. We claim there exists a tuple m = (m1 , . . . , mn ) ∈ Nn0 such that for any distinct
j, j 0 ∈ J, jm 6= j 0 m. To justify such a claim, let N ∈ N be greater than any component
ji of any tuple j = (j1 , . . . , jn ) ∈ J. Then m = (N n−1 , N n−2 , . . . , N, 1) satisfies the desired
conclusion, since for N large enough, every base N expansion of a tuple j ∈ J is unique.
For each 1 ≤ i ≤ n − 1, set zi = xi − xm n . Then k[z1 , . . . , zn−1 , xn ] = k[x1 , . . . , xn ] so (i)
i

33
2.4. Normalization Chapter 2. Integral Dependence

is satisfied. Moreover, we have


X j
0 = F (x1 , . . . , xn ) = cj xj11 · · · xn−1
n−1 jn
xn
j∈J
X
= cj (z1 + xm 1 j1 mn−1 jn−1 jn
n ) · · · (zn−1 + xn ) xn
j∈J
X
= cj x m
n
1 j1
· · · xm
n
n−1 jn−1 jn
xn + lower terms in xn
j∈J
X
= cj x m
n
1 j1 +...+mn−1 jn−1 +mn jn
+ lower terms in xn
j∈J
X
= cj xjm
n + lower terms,
j∈J

where the lower terms at the end have total degree strictly bounded by m. By our choice
of m, we have that the powers xjm n over all j ∈ J are distinct, so in particular there is a
maximum exponent j m, j ∈ J and for this j ∗ , cj ∗ 6= 0. This proves xn satisfies a nonzero
∗ ∗

polynomial in k[z1 , . . . , zn−1 ][t] which may be made monic by dividing out by cj ∗ 6= 0. Hence
xn is integral over k[z1 , . . . , zn−1 ].
Remark. If A is an integral domain with field of fractions K = k(x1 , . . . , xn ), the r in the
normalization lemma equals tr degk K.
We use Noether’s normalization lemma to prove the fundamental theorem in algebraic ge-
ometry, Hilbert’s Nullstellensatz. This first version is stated in terms of algebraic extensions
of fields; we will prove other versions later.
Theorem 2.4.2 (Hilbert’s Nullstellensatz, Algebraic Version). If K/k is a field extension
such that K is a finitely generated k-algebra, then K/k is a finite algebraic extension.
Proof. Noether’s normalization lemma says that there are yi ∈ K such that K/k[y1 , . . . , yr ]
is integral. Then by Lemma 2.2.1, K a field implies k[y1 , . . . , yr ] is also a field. However,
k[y1 , . . . , yr ] ∼
= k[t1 , . . . , tr ] which is not a field unless r = 0. Thus no such yi exist, so K/k
is an integral extension, which is obviously equivalent to K/k being algebraic. Finally, the
extension is of finite degree because K is finitely generated as a k-algebra.
Definition. For k a field and a = (a1 , . . . , an ) ∈ k n , the ideal associated to a is

ma := (t1 − a1 , . . . , tn − an ) ⊂ k[t1 , . . . , tn ].

Remark. Notice that for each 1 ≤ i ≤ n, ti ≡ ai (mod ma ), so for any polynomial F ∈


k[t1 , . . . , tn ],
F (t1 , . . . , tn ) ≡ F (a1 , . . . , an ) (mod ma ).
Therefore k[t1 , . . . , tn ] = k + ma . This defines an evaluation homomorphism

ϕa : k[t1 , . . . , tn ] −→ k
F (t1 , . . . , tn ) 7−→ F (a1 , . . . , an ).

34
2.4. Normalization Chapter 2. Integral Dependence

One can check that ϕa is a k-algebra homomorphism. Clearly ma ⊆ ker ϕa , and by k[t1 , . . . , tn ] =
k + ma , we get ma = ker ϕa and this is a maximal ideal of k[t1 , . . . , tn ]. Also, if a 6= b ∈ k n
then ma 6= mb . This defines an injection
k n ,−→ MaxSpec(k[t1 , . . . , tn ])
a 7−→ ma
where MaxSpec(A) = {maximal ideals of A} for any ring A. This mapping is not surjective
in general, but in the case that k is algebraically closed, the Nullstellensatz (2.4.2) implies
that it is surjective.
Corollary 2.4.3. If k is algebraically closed, then the mapping
k n −→ MaxSpec(k[t1 , . . . , tn ])
is bijective.
Proof. We saw that if a 6= b then ma 6= mb , so it remains to show surjectivity. Let m be
a maximal ideal of A := k[t1 , . . . , tn ]. Then the quotient ring K = A/m is a field which is
finitely generated as a k-algebra by images of the monomials t̄i = ti + m in A/m. By the
Nullstellensatz (2.4.2), K/k is algebraic, but since k was assumed to be algebraically closed,
this implies K = k. Therefore for all 1 ≤ i ≤ n, there exists an ai ∈ k such that t̄i = ai = āi
in K. Equivalently, ti ≡ ai (mod m) so ti − ai ∈ m for all i. This proves ma ⊆ m for
a = (a1 , . . . , an ), but we saw that ma is always maximal. Thus ma = m. We conclude that
k n → A is both surjective and injective.
By Corollary 2.4.3, if k is algebraically closed, every proper ideal I ⊂ k[t1 , . . . , tn ] has
a common zero in k n , called a Nullstelle in German. Explicitly, I ⊆ ma for some maximal
ideal ma defined by a point a ∈ k n .
Lemma 2.4.4. If A is a ring in which every prime ideal P which is not maximal is an
intersection of prime ideals properly containing P , then every prime ideal is an intersection
of maximal ideals.
Proof. Suppose to the contrary that p ⊂ A is a prime ideal that is not an intersection of
maximal ideals. In particular, p is not maximal so by hypothesis,
\
p = {Q ⊂ A | Q prime, Q ) p}.
Consider the integral domain B = A/p. Then J(B) 6= 0, where J(B) denotes the Jacobson
radical of B, so there is a nonzero element x ∈ J(B). Let S = {xn | n ≥ 0} be the
multiplicatively closed subset of powers of x in B, and consider the localization Bx = S −1 B.
Since p is prime, A/p is reduced, i.e. N (A/p) = 0, but this means x 6∈ N (B). In particular,
x is not nilpotent, so Bx is nonzero. Let m ⊂ Bx be a maximal ideal and let Q be its
corresponding prime ideal in B. If Q were itself maximal, we would have x ∈ Q but
then Q ∩ S 6= ∅. Therefore Q is not maximal, but for any prime ideal Q0 ) Q, there is no
corresponding prime in Bx , so Q0 ∩S must be nonempty. In particular, x ∈ Q0 for every prime
ideal Q0 ) Q. However, the assumption on p implies every non-maximal
T 0 prime in B =0 A/p is
an intersection of prime ideals properly containing it. Thus Q = {Q ⊂ B prime | Q ) Q},
but then x ∈ Q, a contradiction. Hence no such p exists in A, so every prime ideal in A is
an intersection of maximal ideals.

35
2.4. Normalization Chapter 2. Integral Dependence

Corollary 2.4.5. If A is a finitely generated k-algebra, for k a field, every prime ideal of A
is an intersection of maximal ideals.

Proof. By Lemma T 2.4.4, it suffices to show that for every prime ideal p ⊂ A which is not
maximal, p = {Q T ⊂ A | Q prime, Q ) p}. Suppose to the contrary that there exists such
a prime p with p ( {Q ⊂ A | Q prime, Q ) p}. Let B = A/p. Then B is not a field, and
by assumption the intersection of all nonzero prime ideals in B is nonzero. In particular this
intersection contains a nonzero element f , which
n as in Lemma o 2.4.4 is not nilpotent.
b
Consider the (nonzero) localization Bf = f n : b ∈ B, n ≥ 0 . Then Bf is finitely gener-
ated as a k-algebra since A and B are. Suppose Bf is a field. Then by Nullstellensatz (2.4.2),
Bf is an algebraic extension of k. In particular, Bf /k is integral, so Bf is integral over B
as well. Then Lemma 2.2.1 implies Bf and B are fields simultaneously, but this contradicts
the fact that B is not a field. Therefore Bf cannot be a field, so it has a nonzero prime ideal
m ⊂ Bf . This pulls back along j : B → Bf to a prime ideal m0 ⊂ B. Then f 6∈ m0 but
clearly this is impossible, since f was chosen to lie in the intersection of all prime ideals of
B. Thus no such prime p ⊂ A can exist, so every non-maximal prime ideal of A is equal to
the intersection of all prime ideals properly containing it.

36
2.5. Valuation Rings Chapter 2. Integral Dependence

2.5 Valuation Rings


Definition. An integral domain A with field of fractions K is called a valuation ring if
x ∈ A or x−1 ∈ A for every x ∈ K × .

Example 2.5.1. If A is a UFD and p ∈ A is a prime element, then the localization A(p) =
a
: a, b ∈ A, p - b is a valuation ring. To see this, take dc ∈ K × with c, d ∈ A r {0}. By

b
unique factorization, we may assume dc is in lowest terms, i.e. gcd(c, d) = 1. Then if p - d,
−1 d
we have dc ∈ A(p) . On the other hand, if p | d, we must have p - c so dc = c ∈ A(p) .

Theorem 2.5.2. Every valuation ring A is a local ring and is integrally closed.

Proof. We first prove A is a local ring. Let A× be the set of units in A and let m = A r A× .
One can prove that m is an ideal of A, and since every element outside of m is a unit, m
must be the unique maximal ideal of A.
To prove A is integrally closed, assume x ∈ K × is integral over A. Then there exist
a0 , . . . , an−1 ∈ A such that xn + an−1 xn−1 + . . . + a0 = 0. If x−1 ∈ A, we can multiply the
polynomial expression through by x−(n−1) to obtain

x = −(an−1 + an−2 x−1 + . . . + a0 x−(n−1) ).

Since x−1 ∈ A, the right side lies in A and thus x ∈ A. This shows A is integrally closed.

Theorem 2.5.3. For K a field and A ⊆ K a nonzero subring, the integral closure of A in
K is equal to \
A= B.
valuation rings B
A⊆B⊆K

We will see in Section 4.1 how valuation rings arise naturally as subrings of a field K.

37
2.6. Dimension and Transcendence Degree Chapter 2. Integral Dependence

2.6 Dimension and Transcendence Degree


Question. How does one determine the dimension of a finitely generated k-algebra?
Example 2.6.1. Consider A = k[t1 , . . . , tn ]. We would like to conclude that dim A = n, but
this is not obvious from the definition of Krull dimension. Clearly dim A ≥ n, since
0 ( (t1 ) ( (t1 , t2 ) ( · · · ( (t1 , . . . , tn )
is a chain of prime ideals in A. We will show the opposite inequality shortly.
Assume C and D are integral domains, with D/C a ring extension and K = Frac(C)
and L = Frac(D) their fields of fractions. Then L/K is a field extension. Define the
transcendence degree of D over C to be tr degC D := tr degK L. If C = k is a field, we have
tr degk D = tr degk L = sup{|E| : E ⊂ D, E is algebraically independent over k}.
Lemma 2.6.2. Assume A and B are finitely generated k-algebras which are integral do-
mains, and there is a surjective k-algebra homomorphism π : B → A with ker π 6= 0. Then
tr degk A < tr degk B.
Proof. Set r = tr degk B and s = tr degk A. Since A and B are finitely generated, r and s are
each finite. If a1 , . . . , as ∈ A are algebraically independent over k, choose b1 , . . . , bs ∈ B such
that π(bi ) = ai for each i. Then for any polynomial F ∈ k[t1 , . . . , ts ] vanishing at (b1 , . . . , bs ),
we have
0 = π(F (b1 , . . . , bs )) = F (π(b1 ), . . . , π(bs )) = F (a1 , . . . , as ).
Since the ai are algebraically independent over k, we must have F = 0, so s ≤ r.
To show s < r strictly, we demonstrate that any set of r elements in A is algebraically
dependent over k. Let a1 , . . . , ar ∈ A be given. Again choose bi ∈ B such that π(bi ) = ai
for 1 ≤ i ≤ r and take b0 ∈ ker π r {0}. Since tr degk B = r, there is a nonzero polynomial
F ∈ k[t0 , t1 , . . . , tr ] such that F (b0 , b1 , . . . , br ) = 0. Choose F of minimal degree in t0 . Then
d
X
F (t0 , t1 , . . . , tr ) = Fi (t1 , . . . , tr )ti0 for Fi ∈ k[t1 , . . . , tr ].
i=1
Pd
We claim F0 6= 0; otherwise, F = t0 i=1 Fi (t1 , . . . , tr )ti−1
0 and evaluating in the integral
domain B implies
d
X d
X
0 = b0 Fi (b1 , . . . , br )bi−1
0 =⇒ Fi (b1 . . . , br )bi−1
0 = 0.
i=1 i=1

This is a nonzero polynomial of strictly smaller degree in t0 that vanishes on (b0 , . . . , br ), so


we in fact must have F0 6= 0. Now plug in the bj and apply π to obtain
d
! d
X X
i
0=π Fi (b1 , . . . , br )b0 = Fi (a1 , . . . , ar )π(b0 )i = F0 (a1 , . . . , ar ).
i=1 i=1

Since F0 6= 0, we see that a1 , . . . , ar are algebraically dependent over k.

38
2.6. Dimension and Transcendence Degree Chapter 2. Integral Dependence

Corollary 2.6.3. For any n ≥ 1, dim k[t1 , . . . , tn ] = n.

Proof. Set A = k[t1 , . . . , tn ]. We showed in Example 2.6.1 that dim A ≥ n. Let

p0 ( p1 ( · · · ( p`

be a chain of prime ideals in A. Then we have a well-defined sequence of surjections

A/p0  A/p1  · · ·  A/p`

with all kernels nonzero since pi ( pi+1 for all i. By Lemma 2.6.2,

n ≥ tr degk A/p0 > tr degk A/p1 > · · · > tr degk A/p` ≥ 0.

In this sequence, there are ` strict inequalities, so ` ≤ n. Further, ` was arbitrary so


dim A = n.

Theorem 2.6.4. If A is a finitely generated k-algebra which is an integral domain, then


dim A = tr degk A.

Proof. By Noether’s normalization lemma, we can pick y1 , . . . , yr ∈ A which are algebraically


independent over k such that A/k[y1 , . . . , yr ] is an integral extension of rings. Then dim A =
dim k[y1 , . . . , yr ] = r by Theorem 2.3.7. Now for all x ∈ A, x is integral over k[y1 , . . . , yr ].
If K is the field of fractions of A, then every z ∈ K is algebraic over the field k(y1 , . . . , yr ),
but then Lemma 1.4.3(c) shows that {y1 , . . . , yr } is a transcendence basis of K/k. Hence
tr degk A = tr degk K = r = dim A.

Definition. We say a ring A of finite Krull dimension has the strong dimension property
(SDP) if whenever p0 ( · · · ( p` is a maximal (non-refinable) chain of prime ideals in A,
we have ` = dim A.

Lemma 2.6.5. Assume A has SDP and n = dim A. Then

(1) ht(p) + dim A/p = dim A for any prime ideal p ⊂ A.

(2) ht(m) = n for all maximal ideals m ⊂ A.

(3) If p0 ( p1 ( · · · ( p` is a chain of prime ideals in A which cannot be refined, then


` = ht(p` ) − ht(p0 ).

(4) If p ⊂ A is a prime ideal then A/p has SDP.

Proof. (1) Given a prime ideal p ⊂ A with ht(p) = m, choose a chain

p0 ( · · · ( pm−1 ( p

of prime ideals in A. If dim A/p = d then there is a maximal chain

0 = Q0 ( Q1 ( · · · ( Qd

39
2.6. Dimension and Transcendence Degree Chapter 2. Integral Dependence

of prime ideals in A/p. Then we may form a chain

p0 ( · · · ( pm−1 ( p = Q0 ( Q1 ( · · · ( Qd

where Qi ⊂ A are the prime ideals such that Qi = Qi /p. This is a maximal, non-refinable
chain of primes in A, so by SDP, dim A = m + d = ht(p) + dim A/p.
(2) follows immediately from (1) since A/m is a field and dim F = 0 for any field F .
(3) follows from a similar argument to the proof of (1). Given such a chain of primes

p0 ( · · · ( p`

we can extend this to a maximal chain of primes in A:

Q0 ( · · · ( Qj = p0 ( · · · ( P` ( p`+1 ( · · · ( pm .

Then SDP implies n = j + m. When such a chain is chosen so that j = ht(p0 ) and
m − ` = dim A/p` , we must have

n = j + m = j + ` + (m − `) = ht(p0 ) + ` + dim A/p` .

However by (1), n = ht(p` ) + dim A/p` so we see that ht(p` ) = ht(p0 ) + `.


(4) Any maximal prime ideal chain in A/p is of the form

0 ( p1 /p ( · · · ( p` /p

where p ( p1 ( · · · ( p` is a non-refinable prime ideal chain in A. By (3), ` = ht(p` ) − ht(p)


but p` must be maximal, so by (2), ht(p` ) = n. Finally, applying (1) gives ` = n − ht(p) =
dim A/p. Hence SDP holds for A/p.
An important example of a ring with the strong dimension property is the polynomial
algebra k[t1 , . . . , tn ]. Before proving this has SDP, we need a preliminary result which relies
on the going down theorem.

Proposition 2.6.6. Let A be a finitely generated k-algebra that is an integral domain. Then
for any nonzero prime ideal p ⊂ A such that ht(p) = 1, we have dim A/p = dim A − 1.

Proof. By Noether’s normalization lemma, there exist elements y1 , . . . , yn ∈ A which are


algebraically independent over k and have the property that A/k[y1 , . . . , yn ] is an integral
extension. Then by Theorems 2.3.7 and 2.6.4, dim A = dim k[y1 , . . . , yn ] = n. Since ht(p) =
1, there is no nonzero prime ideal of A properly contained in p. Set q = p ∩ k[y1 , . . . , yn ].
If q = 0, then we would have 0 = p ∩ k[y1 , . . . , yn ] = 0 ∩ k[y1 , . . . , yn ], which would imply
p = 0 by Lemma 2.3.3, but this contradicts ht(p) = 1. Suppose there exists a nonzero
prime ideal q0 ⊂ k[y1 , . . . , yn ] such that q0 ( q. Then the going down theorem would give
us a nonzero prime ideal p0 ( p, contradicting ht(p) = 1. This shows that ht(q) = 1
also. Now by Lemma 2.3.1(a), A/p is integral over k[y1 , . . . , yn ]/q, so it suffices to prove
dim k[y1 , . . . , yn ]/q = n − 1.
Set B := k[y1 , . . . , yn ]. In light of Lemma 1.4.2 and Corollary 2.6.3, we have that dim B =
n. We first show that q is a principal ideal. Since q 6= 0, there exists a nonzero element

40
2.6. Dimension and Transcendence Degree Chapter 2. Integral Dependence

f 0 ∈ q. Then because B is a UFD, f 0 is divisible by some irreducible element f ∈ q, but


then f is prime and 0 6= (f ) ⊆ (f 0 ) ⊆ q. Since q is a minimal prime, we must have (f ) = q.
Finally, B/q = B/(f ) is an integral extension of k[z2 , . . . , zn ] where {1, z2 , . . . , zn } is a change
of basis of {y1 , . . . , yn } such that f (1, z2 , . . . , zn ) is monic. It follows from Theorem 2.3.7 and
Corollary 2.6.3 that dim B/q = dim k[z2 , . . . , zn ] = n − 1.

Theorem 2.6.7. If A is a finitely generated k-algebra which is an integral domain, then A


has the strong dimension property.

Proof. We induct on n = dim A. If n = 0, A is a field so the property holds trivially. Now


assume that every k-algebra which is a domain and has dimension less than n has the strong
dimension property. Take a maximal chain of prime ideals p0 ( · · · ( p` in A. Since A is a
domain, we must have p0 = 0 and p` maximal. Consider the integral domain B = A/p1 . By
the correspondence theorem,

0 = p1 /p1 ( p2 /p1 ( · · · ( p` /p1

is a maximal chain of prime ideals in B of length ` − 1. Moreover, ht(p1 ) = 1 so by


Proposition 2.6.6, dim B = dim A/p1 = dim A − 1 = n − 1. Thus by induction, n − 1 = ` − 1
and so we get n = ` as required.

Corollary 2.6.8. For any n ≥ 1, k[t1 , . . . , tn ] has the strong dimension property. In partic-
ular, dim k[t1 , . . . , tn ]/p = n − ht(p) for any prime ideal p ⊂ k[t1 , . . . , tn ].

41
Chapter 3

Noetherian and Artinian Rings

42
3.1. Ascending and Descending Chains Chapter 3. Noetherian and Artinian Rings

3.1 Ascending and Descending Chains


Definition. A partially ordered set (Σ, ≤) satisfies the ascending chain condition (ACC)
if any ascending chain x1 ≤ x2 ≤ · · · in Σ is stationary, i.e. there is some N ∈ N such that
xn = xN for all n ≥ N .
Definition. A partially ordered set (Σ, ≤) satisfies the descending chain condition (DCC)
if any descending chain x1 ≥ x2 ≥ · · · in Σ is stationary, i.e. there is some N ∈ N such
that xn = xN for all n ≥ N .
In ring theory, we study chain conditions on the collection ΣM of submodules of a fixed
A-module M , partially ordered by inclusion. The modules M that satisfy the above chain
conditions have special names.
Definition. An A-module M is noetherian if the collection of submodules of M satisfies the
ascending chain condition. On the other hand, M is artinian if the collection of submodules
of M satisfies the descending chain condition.
Lemma 3.1.1. Let (Σ, ≤) be a poset. Then
(i) Σ satisfies ACC if and only if every nonempty subset of Σ has a maximal element.
(ii) Σ satisfies DCC if and only if every nonempty subset of Σ has a minimal element.
In particular, M is noetherian iff every nonempty collection of submodules has a maximal
element, and M is artinian iff every nonempty collection of submodules has a minimal
element.
Proposition 3.1.2. An A-module M is noetherian if and only if every submodule of M is
finitely generated (as an A-module).
f g
Lemma 3.1.3. Let 0 → M 0 → − M 00 → 0 be a short exact sequence of A-modules. Then
− M→
(a) M is noetherian if and only if M 0 and M 00 are noetherian.
(b) M is artinian if and only if M 0 and M 00 are artinian.
Proof. ( =⇒ ) Consider the maps induced by f and g:

f ∗ : ΣM 0 −→ ΣM g ∗ : ΣM 00 −→ ΣM
N 0 7−→ f (N 0 ) N 00 7−→ g −1 (N 00 ).

Then f ∗ and g ∗ are both order-preserving and injective maps, so the posets ΣM 0 and ΣM 00
inherit the chain conditions from ΣM .
( ⇒= ) We prove this for the artinian property; the proof for noetherian modules is
similar. Let C : M1 ⊇ M2 ⊇ M3 ⊇ · · · be a descending chain of submodules Mi ⊂ M . These
correspond to chains of submodules

f −1 (C) :f −1 (M1 ) ⊇ f −1 (M2 ) ⊇ f −1 (M3 ) ⊇ · · · in M 0


g(C) :g(M1 ) ⊆ g(M2 ) ⊇ g(M3 ) ⊇ · · · in M 00 .

43
3.1. Ascending and Descending Chains Chapter 3. Noetherian and Artinian Rings

By the DCC on ΣM 0 and ΣM 00 , there exists a single N ∈ N such that f −1 (Mn ) = f −1 (MN )
and g(Mn ) = g(MN ) for all n ≥ N . For all n ≥ N , we already have Mn ⊇ MN . On the
other hand, given x ∈ MN , there exists a y ∈ Mn such that g(x) = g(y). This implies
g(x) − g(y) = g(x − y) = 0 by A-linearity. So x − y ∈ ker g which equals im f by exactness.
Thus there is some z ∈ M 0 such that f (z) = x − y ∈ MN . Then z ∈ f −1 (MN ) = f −1 (Mn )
so x − y = f (z) ∈ Mn . Hence x = y + f (z) ∈ Mn so MN ⊆ Mn . This proves the DCC on
ΣM , so M is artinian.
Ln
Corollary 3.1.4. If M1 , . . . , Mn are noetherian (resp. artinian) A-modules then i=1 Mi
is also noetherian (resp. artinian).

Proof. This is proved by induction on n, considering short exact sequences


n
M n−1
M
0 → Mn → Mi → Mi → 0
i=1 i=1

and applying Lemma 3.1.3.


We distinguish the special case of A as a module over itself.

Definition. A (commutative) ring A is a noetherian ring if it is noetherian as a module


over itself. Similarly, A is an artinian ring if it is artinian as a module over itself.

Let A be a ring. Then the A-submodules of A are exactly the ideals of A. Then by
Proposition 3.1.2, A is noetherian if and only if all ideals of A are finitely generated.

Corollary 3.1.5. If A is noetherian (resp. artinian), then any finitely generated A-module
M is also noetherian (resp. artinian).

Proof. Write M = Am1 + . . . + Amn for mi ∈ M . Consider the projection

An −→ M
n
X
(a1 , . . . , an ) 7−→ ai mi .
i=1

Then An is noetherian (resp. artinian) by Corollary 3.1.4 so M is noetherian (resp. artinian)


by Lemma 3.1.3.

Lemma 3.1.6. For a field k and a k-vector space V , the following are equivalent:

(i) dimk V < ∞.

(ii) V is a noetherian k-module.

(iii) V is an artinian k-module.

44
3.1. Ascending and Descending Chains Chapter 3. Noetherian and Artinian Rings

Proof. (i) =⇒ (ii), (iii) follow from the linear algebra fact that every subspace of a finite
dimensional vector space is finite dimensional, and thus has a finite basis.
(ii), (iii) =⇒ (i) If dimk V = ∞ then there is a countable linear independent subset
{vj }∞
j=1 ⊂ V . Let Vi = Span{vj | j ≤ i} for each i ∈ N. Then V1 ( V2 ( · · · is an
ascending chain in which each subspace is finitely generated, so it cannot stabilize. Likewise,
let Wi = Span{vj | j ≥ i} for each i ∈ N. Then W1 ) W2 ) · · · is a descending chain which
doesn’t stabilize. Hence dimk V = ∞ implies is V is neither noetherian nor artinian.

Proposition 3.1.7. Suppose A has maximal ideals m1 , . . . , mn , not necessarily distinct, with
m1 · · · mn = 0. Then A is noetherian if and only if A is artinian.

Proof. Say N is an A-module and m ⊂ A is a maximal ideal. Then N/mN becomes a k-


vector space for k = A/m via (a + m)(n + mN ) = an + mN . Then A-submodules of N/mN
are the same as k-subspaces of N/mN . Thus N/mN is noetherian (resp. artinian) as an
A-module if and only if N/mN is noetherian (resp. artinian) as a k-vector space; this in
turn is equivalent to dimk N/mN < ∞ by Lemma 3.1.6. Hence N/mN is noetherian if and
only if N/mN is artinian.
Now let m1 , . . . , mn ⊂ A be maximal ideals satisfying m1 · · · mn = 0. Set Ai = A/m1 · · · mi
for each 1 ≤ i ≤ n. In particular, An = A/{0} = A by hypothesis. If A is noetherian (resp.
artinian) then each Ai is noetherian (resp. artinian) by Lemma 3.1.3. Note that A1 = A/m1
is a field, so A1 is both noetherian and artinian. We claim that Ai is noetherian for each
1 ≤ i ≤ n if and only if Ai is artinian for each 1 ≤ i ≤ n. Indeed, A1 is the base case so for
n ≥ 2, consider the short exact sequence

0 → m1 · · · mi−1 /m1 · · · mi → Ai → Ai−1 → 0.

Set Ni = m1 · · · mi−1 /m1 · · · mi . Then by Lemma 3.1.3,

Ai is noetherian ⇐⇒ Ni and Ai−1 are noetherian


⇐⇒ Ni is noetherian and Ai−1 is artinian (by induction)
⇐⇒ Ni is artinian and Ai−1 is artinian (by preliminary remarks)
⇐⇒ Ai is artinian.

Finally, since A = An , this implies that A is noetherian if and only if A is artinian.

45
3.2. Composition Series Chapter 3. Noetherian and Artinian Rings

3.2 Composition Series


Let M be an A-module.
Definition. A chain in M is a strictly descending chain of submodules of M of the form

M = M0 ) M1 ) · · · ) Mn = 0.

The integer n is called the length of the chain.


Definition. A composition series of M is a chain in M which cannot be refined, or
equivalently, the quotient Mi /Mi+1 is simple for all 0 ≤ i ≤ n − 1.
Definition. The length of a module M , denoted `(M ), is by convention `(M ) = ∞ if M
does not have a composition series. Otherwise,

`(M ) = min{n ∈ N | M has a composition series of length n}.

One can view the length function `(·) as a generalization of the dimension of a vector
space.
Proposition 3.2.1. Assume M is an A-module that has a composition series. Then
(a) For every submodule N ⊂ M , `(N ) ≤ `(M ) and if N is a proper submodule, `(N ) <
`(M ).

(b) Any chain in M has length at most `(M ).

(c) Every composition series of M has length `(M ).

(d) Any chain in M can be extended to a composition series.


Proof. (a) Choose a composition series of minimal length,

M = M0 ) M1 ) · · · ) M` = 0.

Then ` = `(M ). Intersecting with N , setting Ni = N ∩ Mi , gives a sequence of containments


that are not necessarily strict:

N = N0 ⊇ N1 ⊇ · · · ⊇ N` = 0.

However for each i, we have an inclusion Ni /Ni+1 ,→ Mi /Mi+1 since

Ni+1 = N ∩ Mi+1 = N ∩ Mi ∩ Mi+1 = Ni ∩ Mi+1 .

By assumption, each Mi /Mi+1 is simple, so either Ni /Ni+1 = 0 or Ni /Ni+1 ∼


= Mi /Mi+1 , in
which case Ni /Ni+1 is also simple. After removing repetitions, i.e. where Ni = Ni+1 , we
obtain a composition series of N of length at most `. Hence `(N ) ≤ `(M ). Now suppose
N ( M but `(N ) = `(M ). Then Ni ) Ni+1 for all 0 ≤ i ≤ ` − 1 in the above chain. By
construction, M`−1 is a simple A-module and 0 6= N`−1 ⊆ M`−1 , so we have N`−1 = M`−1 .

46
3.2. Composition Series Chapter 3. Noetherian and Artinian Rings

By induction, Ni = Mi for all 0 ≤ i ≤ ` − 1. Hence N = N0 = M0 = M , a contradiction.


Therefore `(N ) < `(M ) for every proper submodule N .
(b) Consider any chain M = M0 ) M1 ) · · · ) Mn = 0. Then by (a),

` = `(M0 ) > `(M1 ) > · · · > `(Mn ) > 0.

In this string we have n strict inequalities so it must be that n ≤ `.


(c) By definition of `(M ), any compositions series of M has length at least `(M ), but by
(b), every composition series of M has length at most `(M ).
(d) Given any chain, refine it as many times as possible. By (b), we cannot get a chain of
length strictly greater than `(M ), so the process will eventually terminate when the length
of the chain is `(M ). At this point we will have a composition series for M .
Corollary 3.2.2. An A-module M has a composition series if and only if M is both noethe-
rian and artinian.
Proof. ( =⇒ ) Assume `(M ) < ∞. Then by Proposition 3.2.1, all chains in M have length
at most `(M ). In particular, any ascending or descending chain must be finite, which implies
M has satisfies ACC and DCC.
( ⇒= ) Assume M 6= 0. Then the set of all proper submodules of M has a maximal
element by Lemma 3.1.1; call it M1 . If M1 = 0, then M ) M1 = 0 is a composition series
for M . Otherwise, the set of proper submodules of M1 has a maximal element M2 , by
Lemma 3.1.1 again, using the fact that submodules of noetherian modules are noetherian.
Continue in this fashion to construct a chain

M = M0 ) M1 ) M2 ) · · ·

By maximality in each step, Mi /Mi+1 is simple for all i ≥ 0. Since M is artinian, this process
must terminate, else the DCC is violated. Then at some point we have

M = M0 ) M1 ) M2 ) · · · ) Mn ) 0

which is a composition series for M .


Lemma 3.2.3. If M is an A-module that admits a composition series, then for any short
exact sequence 0 → M 0 → M → M 00 → 0 of A-modules, `(M ) = `(M 0 ) + `(M 00 ). That is,
length is additive on short exact sequences.
f g
Proof. Label the morphisms 0 → M 0 → − M → − M 00 → 0. First, M is both noetherian
and artinian by Corollary 3.2.2. Note that M 0 and M 00 are also noetherian and artinian
by Lemma 3.1.3, so each admits a composition series. Consequently, `(M 0 ) and `(M 00 ) are
defined. Let

M 0 = M00 ) M10 ) · · · ) M`0 = 0 and M 00 = M000 ) M100 ) · · · ) Mn00 = 0

be composition series of M 0 and M 00 . Then by Proposition 3.2.1(c), `(M 0 ) = ` and `(M 00 ) =


n. By exactness, we get a sequence of inclusions

M = g −1 (M000 ) ⊃ · · · ⊃ g −1 (Mn00 ) = ker g = im f = f (M00 ) ⊃ f (M10 ) ⊃ · · · ⊃ f (M`0 ) = 0. (∗)

47
3.2. Composition Series Chapter 3. Noetherian and Artinian Rings

Now since g is surjective, g(g −1 (Mi00 )) = Mi00 so the inclusions are strict between the g −1 (Mi00 ),
0 ≤ i ≤ n. Similarly, since f is injective, f −1 (f (Mi0 )) = Mi0 so the inclusions between the
f (Mj0 ), 0 ≤ j ≤ `, are strict. Therefore (∗) is a strictly descending chain of submodules of
M of total length ` + n. Consider each quotient in the first half; since g is surjective, we get
an isomorphism
'
g −1 (Mi00 )/g −1 (Mi+1
00
→ Mi00 /Mi+1
)− 00
for 0 ≤ i ≤ n.
Then each g −1 (Mi00 )/g −1 (Mi+1
00
) is simple. Similarly, injectivity of f gives us an isomorphism
'
0
Mi0 /Mi+1 → f (Mi0 )/f (Mi+1
− 0
) for 0 ≤ j ≤ `,
0
which shows the f (Mi0 )/f (Mi+1 ) are all simple. Hence (∗) is a composition series for M , so
by Proposition 3.2.1(c), `(M ) = ` + n = `(M 0 ) + `(M 00 ).

48
3.3. Noetherian Rings Chapter 3. Noetherian and Artinian Rings

3.3 Noetherian Rings


In this section we present the key details regarding noetherian rings and their modules in
commutative algebra.
Lemma 3.3.1. If A is a noetherian ring, then
(a) For any ideal I ⊂ A, A/I is a noetherian ring.

(b) For any multiplicatively closed subset S ⊆ A, S −1 A is a noetherian ring. In particular,


Ap is a local noetherian ring for any prime ideal p ⊂ A.
Proof. (a) By Lemma 3.1.3, we know A/I is noetherian as an A-module. It follows from
the isomorphism theorems that the A-submodules of A/I are exactly the ideals of A/I, so
it follows that A/I is noetherian as a ring.
(b) If J1 ⊆ J2 ⊆ · · · is a chain of ideals in S −1 A then applying j −1 gives a chain of
ideals j −1 (J1 ) ⊆ j −1 (J2 ) ⊆ · · · in A. Since A is noetherian, the chain stabilizes, i.e. there
is some n ∈ N such that j −1 (Ji ) = j −1 (Jn ) for all i ≥ n. Applying the localization functor
S −1 to these modules returns the original Ji , since S −1 (j −1 (Ji )) = Ji for each i ≥ 1 by
Proposition 1.3.8:
J1 ⊆ J2 ⊆ · · · ⊆ Jn = Jn+1 = · · · .
Hence the ACC holds for S −1 A, so S −1 A is noetherian. The final remark about Ap follows
from Corollary 1.3.9.
Remark. A similar argument as in Lemma 3.3.1(b) shows that for every noetherian A-
module M , any localization S −1 M is also a noetherian A-module.
One of the fundamental results in the theory of noetherian rings is known as Hilbert’s ba-
sis theorem. We will see that it implies every finitely generated ring extension of a noetherian
ring is also noetherian.
Theorem 3.3.2 (Hilbert’s Basis Theorem). If A is a noetherian ring, then A[t] is also
noetherian.
Proof. Let I ⊂ A[t] be an ideal and let L denote the set of all leading coefficients of elements
of I. That is,
( n
)
X
i
L = α ∈ A : there is some ai t ∈ I with an = α .
i=0

We claim L is an ideal. Clearly 0 ∈ L since I contains the zero polynomial. Suppose f, g ∈ I


such that f = an tn + . . . + a0 and g = bm tm + . . . + b0 . We must show an + bm ∈ L. If n = m
then the leading coefficient of f + g is an + bm , and since f + g ∈ I, we have an + bn ∈ L. If
n 6= m, without loss of generality assume n < m. Then

tm−n f + g = (an tm + . . . + a0 tm−n ) + (bm tm + . . . + b0 )


= (an + bm )tm + . . . + (a0 + bm−n )tm−n + . . . + b0

49
3.3. Noetherian Rings Chapter 3. Noetherian and Artinian Rings

which lies in I, showing an +bm ∈ L. Finally, if c ∈ A then can is the leading term of cf ∈ I, so
L is an ideal of A. Now, since A is noetherian, L is finitely generated, say by a1 , . . . , an ∈ A.
For each 1 ≤ i ≤ n, let fi ∈ I be a polynomial having ai as its leading coefficient. For each
of these, write fi = ai tei + . . . so that deg fi = ei . Set E = max{e1 , . . . , en }.
For each 0 ≤ d ≤ E − 1, let Ld be the set of all leading coefficients of polynomials in I
of degree d, together with 0:
( d
)
X
Ld = α ∈ A : there is some ai ti ∈ I with ad = α .
i=0

The proof that each Ld is an ideal of A is similar to the proof above for L. Then each
Ld is finitely generated, say by bd1 , . . . , bdnd ∈ A. Let fd1 , . . . , fdnd ∈ I be polynomials
such that the leading coefficient of fdr is bdr , 1 ≤ r ≤ nd . We claim that f1 , . . . , fn and
fdr , 0 ≤ d ≤ E − 1, 1 ≤ r ≤ nd , are a generating set for I. Let Q denote the ideal generated
by these polynomials. By construction, Q ⊆ I since the generators of Q are elements of I.
If Q ( I, choose f ∈ I r Q of minimum degree d with leading coefficient a.
First suppose d ≥ E. Then a ∈ L so there exist elements c1 , . . . , cn ∈ A such that
a = c1 a1 + . . . + cn an . Observe that g = c1 td−e1 f1 + . . . + cn td−en fn is a polynomial in Q
of degree d and leading coefficient a, so f − g ∈ I has strictly smaller degree than f . By
minimality, we must have f − g = 0, that is, f = g ∈ Q, a contradiction. On the other
hand, if d < E then f ∈ Ld so we have f = c1 bd1 + . . . + cnd bdnd for some cr ∈ A. Then
g = c1 fd1 + . . . + cnd fnd is a polynomial in Q of degree d and leading coefficient a, producing
the same contradiction. It follows that Q = I, so I is finitely generated. This proves that
A[t] is noetherian.

Remark. The converse of Hilbert’s basis theorem is trivially true: A[t] noetherian implies
A noetherian since A ∼
= A[t]/(t) as rings.
Corollary 3.3.3. If A is noetherian, then

(a) A[t1 , . . . , tn ] is noetherian for any n ≥ 1.

(b) B is noetherian for any finitely generated ring extension B/A.

Proof. (a) Induct on the number of generators. The base case is Hilbert’s basis theorem.
(b) We can write B = A[b1 , . . . , bn ] for some bi ∈ B. Then there is a (unique) surjective
ring homomorphism

ϕ : A[t1 , . . . , tn ] −→ B
ti 7−→ bi .

So B ∼
= A[t1 , . . . , tn ]/ ker ϕ and we can apply (a) and Lemma 3.3.1.
Corollary 3.3.4. For a field k, any finitely generated k-algebra is noetherian.

Proof. Every field is noetherian (it admits a trivial composition series) so Lemma 3.3.3(b)
may be applied.

50
3.3. Noetherian Rings Chapter 3. Noetherian and Artinian Rings

Corollary 3.3.5. Every finitely generated commutative ring is noetherian.


Proof. Let A be such a ring. There is a unique ring homomorphism f : Z → A given by
f (n) = n · 1A for all n. Set A0 = f (Z). Then A0 is noetherian by Lemma 3.3.1(a) and the
hypothesis that A is finitely generated as a ring means that A/A0 is a finitely generated ring
extension. Then the statement follows from Corollary 3.3.3(b).
Proposition 3.3.6. Suppose A is noetherian and C ⊃ B ⊃ A are extensions of rings such
that C/A is a finitely generated extension and C is a finitely generated B-module. Then B/A
is also finitely generated.
Proof. Choose x1 , . . . , xm ∈ C and y1 , . . . , yn ∈ C with y1 = 1 such that
n
X
C = A[x1 , . . . , xm ] = Byj .
j=1

Then there exist bij` ∈ B such that for all 1 ≤ i ≤ m, 1 ≤ j ≤ n,


n
X
xi yj = bij` y` . (∗)
`=1

Define the subring B0 = A[bij` | 1 ≤ i ≤ m, 1 ≤ j, ` ≤ n]. Then B0 is finitely generated


as an A-module by construction, so by Corollary 3.3.3, B0 is noetherian. Consider the
B0 -submodule
Xn
M= B0 y` ⊆ C.
`=1
Pn
By (∗), xi yj ∈ M for all i, j. Thus xi j=1 B0 yj ⊆ M , or in other words xi M ⊆ M for each
i. In particular, C = A[x1 , . . . , xm ] ⊆ M ⊆ C, so M = C. Now we have
A ⊆ B0 ⊆ B ⊆ C
with B0 noetherian and C a finitely generated B0 -module. Thus by Corollary 3.1.5, C is a
noetherian B0 -module, and Proposition 3.1.2 implies that B ⊆ C is finitely generated as a
B0 -module. Hence B/B0 is a finitely generated ring extension, and by construction B0 /A is
a finitely generated ring extension, so we conclude that B/A is finitely generated as well.
As a result, we obtain the following interesting fact.
Theorem 3.3.7. If a field K is finitely generated as a ring then K is a finite field.
Proof. Consider the map f : Z → K. Since K is a field, f (Z) is an integral domain. We have
two cases: (1) when char K = p > 0, f (Z) ∼ = Fp ; and (2) when char K = 0, f (Z) ∼ = Z. In
case (1), K is a finitely generated Fp -algebra. By Hilbert’s Nullstellensatz (2.4.2), K/Fp is an
algebraic extension, so in fact it is an extension of finite fields. Therefore K is finite. On the
other hand, in (2), K is finitely generated over Z and K is a field, so we have Z ⊂ Q ⊂ K.
This means K is a finitely generated Q-algebra, so once again the Nullstellensatz says that
K/Q is a finite extension. However, one sees that the conditions of Proposition 3.1.2 are
satisfied, so Q is a finitely generated ring extension of Z, which is false. Therefore the only
possibility is that K is a finite field.

51
3.3. Noetherian Rings Chapter 3. Noetherian and Artinian Rings

Corollary 3.3.8. Any matrix ring GLn (K), n ≥ 2, over an infinite field K cannot be finitely
generated.
The proof of Hilbert’s basis theorem can be generalized to prove that power series rings
over a noetherian ring are also noetherian.
Theorem 3.3.9. If A is noetherian, then the power series ring A[[t]] is also noetherian.
Proof. For a power series f ∈ A[[t]], where f = ∞ i
P
i=0 ai t , let j be the smallest index such
that aj 6= 0. We say aj is the coefficient of lowest degree of f . If ai = 0 for all i ≥ 0, i.e. f
is the zero power series, we will say its coefficient of lowest degree is 0.
Let I ⊂ A[[t]] be an ideal and define L ⊂ A to be the set of all coefficients of lowest
degree of elements in I:
( ∞
)
X
L = aj : ai ti ∈ I for some j ∈ N0 and ai , i > j .
i=j

We first show L is an ideal.


P∞ Since 0 ∈ I andP0 has lowest degree coefficient 0, we have 0 ∈ L.
Let f, g ∈ I, with f = i=j ai t and g = ∞
i k
k=` bk t . We must show aj + b` ∈ L. If j = `,
the lowest degree coefficient of f + g is aj + bj , and since f + g ∈ I, we have aj + bj ∈ L. If
j 6= `, without loss of generality assume j < `. Then

X ∞
X ∞
X ∞
X ∞
X
`−j `−j i k i+`−j k
t f +g =t ai t + bk t = ai t + bk t = (ak+j−` + bk )tk
i=j k=` i=j k=` k=`

so we see that the lowest degree coefficient is aj + b` . Therefore aj + b` ∈ L. Now, if α ∈ A,


the lowest degree coefficient of αf is αai , and αf ∈ I since I is an ideal, so αai ∈ L. This
proves L is an ideal of A. By the noetherian hypothesis, L is finitely generated by some
α1 , . . . , αn ∈ A.
Now for each k = 1, . . . , n, take fk ∈ I such that the lowest degree termPof fk is αk ; let

ek be the exponent of t in fk having αk as its coefficient, i.e. fk = αk t + i=ek +1 ai ti . Set
ek

E = max{ek | 1 ≤ k ≤ n}. For all 0 ≤ d ≤ E − 1, define Jd ⊂ A by


( ∞
)
X
Jd = ad : ai ti ∈ I for some ai , i > d ∪ {0}.
i=d

The proof that each Jd is an ideal of A is similar to the proof for L. In particular, since A is
noetherian we get that Jd is finitely generated for each 0 ≤ d ≤ E − 1. Say Jd is generated
by βd1 , . . . , βdnd ∈ A. ForP each 0 ≤ d ≤ E − 1 and for each 1 ≤ r ≤ nd , let fdr ∈ A[[t]] be
such that fdr = βdr td + ∞ a
i=d+1 i t i
∈ I. Let Q be the ideal of A[[t]] finitely generated by all
fk , 1 ≤ k ≤ n and fdr , 0 ≤ d ≤ E − 1, 1 ≤ r ≤ nd . We claim that Q = I. Since all of the
generators of Q lie in I, it is clear that Q ⊆ I. On the other hand, take f ∈ I and suppose
the lowest degree term of f is atdP , for some a ∈ L and d ≥ 0. If d ≤ E − 1, then a ∈ Jd so
nd
there exist cr ∈ A such that a = r=1 cr βdr , which implies

nd nd
! nd ∞ X nd
X X X X X
d i d
f− cr fdr = f − cr βdr t + ai t = f − cr βdr t − cr ai ti
r=1 r=1 i=d+1 r=1 i=d+1 r=1

52
3.3. Noetherian Rings Chapter 3. Noetherian and Artinian Rings

is a power series in A[[t]] of lowest degree


Pnd at least d + 1. (We can switch the sums in the last
step since one is finite.) Since f − r=1 cr fdr ∈ I, we may assume that the lowest degree of
f is at least E. If the lowest degree of f is d > E, then
nE
X
f− cr fdr td−E
r=1

which is a power series in A[[t]] of lowest degree at least d + 1. Notice that the ideals Jd
are ascending: J0 ⊆ J1 ⊆ J2 ⊆ · · · . Since A is noetherian, this chain stabilizes, or in other
words, we can always find power series

X
gr = c`r t`−E for 1 ≤ r ≤ nE
`=d

such that f can be written


nE
X
f= gr fdr .
r=1

Hence f ∈ Q so Q = I and it follows that I is finitely generated.

53
3.4. Primary Decomposition Chapter 3. Noetherian and Artinian Rings

3.4 Primary Decomposition


For this section, we assume A is a noetherian ring. Our goal is to prove an analog of primary
decomposition (of modules over a PID) for modules over A.
Lemma 3.4.1. For any ideal I ⊂ A, there exists n ∈ N such that r(I)n ⊆ I.
Proof. Since A is noetherian, r(I) is finitely generated; say x1 , . . . , xm ∈ r(I) such that
(x1 , . . . , xm ) = r(I). Then there exist ei ∈ N such that xei i ∈ I for each 1 ≤ i ≤ m. Set
n = e1 + . . . + em . Then xj11 · · · xjmm ∈ I whenever j1 + . . . + jm ≥ n, but the products
xj11 · · · xjmm generate r(I)n , so we see that for this choice of n, r(I)n ⊆ I.
We obtain the following fact about the nil radical of a noetherian ring.
Corollary 3.4.2. There exists an n ∈ N such that N (A)n = 0.
Proof. For the ideal I = (0), r(0) = N (A). The statement follows from Lemma 3.4.1.
Definition. A proper ideal I ⊂ A is irreducible if for any ideals J, K ⊂ A with J ∩ K = I,
we have J = I or K = I.
Example 3.4.3. Prime ideals are always irreducible. Indeed, if J ∩K = p is prime but p ( J
and p ( K, then there are some elements x ∈ J rp and y ∈ K rp. But xy ∈ JK ⊆ J ∩K = p
which is a contradiction since p is prime. Thus p is irreducible.
Lemma 3.4.4. Any proper ideal I ⊂ A is a finite intersection of irreducible ideals.
Proof. Consider the collection

Σ = {proper ideals I ⊂ A | I is not an intersection of irreducible ideals}.

If we assume Σ is nonempty, then since A is noetherian, Lemma 3.1.1 guarantees that Σ has
a maximal element I0 . In particular, I0 is not irreducible. Then there exist ideals J, K ⊂ A
such that I0 = J ∩ K but I0 ( J and I0 ( K. Since I0 is maximal in Σ, neither J nor K lie
in Σ. Thus J and K are each a finite intersection of irreducible ideals, but then I0 = J ∩ K is
also a finite intersection of irreducible ideals, a contradiction. Hence Σ must be empty.
Definition. A proper ideal I ⊂ A is primary if for every x, y ∈ A, xy ∈ I implies x ∈ I
or y ∈ r(I). Equivalently, I is primary if xy ∈ I implies that x ∈ I, y ∈ I or x, y ∈ r(I).
Lemma 3.4.5. If a proper ideal I ⊂ A is irreducible, then I is primary.
Proof. Consider B = A/I. Then if I is irreducible, (0) is irreducible in B, i.e. there do not
exist two nonzero ideals J, K ⊂ B with J ∩ K = 0. Assume x, y ∈ B such that xy = 0 and
x 6= 0. We will show that y n = 0 for some n ∈ N, i.e. y is nilpotent. Consider the ascending
chain
Ann(y) ⊆ Ann(y 2 ) ⊆ Ann(y 3 ) ⊆ · · ·
By the ACC, there is some n ∈ N such that Ann(y i ) = Ann(y n ) for all i ≥ n. Now we show
(x) ∩ (y n ) = 0. If b ∈ (x) ∩ (y n ) then b = rx for some r ∈ B, so by = rxy = 0. Also, b = cy n

54
3.4. Primary Decomposition Chapter 3. Noetherian and Artinian Rings

for some c ∈ B, so cy n+1 = cy n y = by = 0. This shows c ∈ Ann(y n+1 ) = Ann(y n ), and thus
0 = cy n = b. Hence (x) ∩ (y n ) = 0 but since (0) is irreducible and x 6= 0, we must have
(y n ) = 0.
Finally, this implies I ⊂ A is primary, since if x0 , y 0 ∈ A such that x0 y 0 ∈ I but x0 6∈ I,
we have x0 + I 6= I and (x0 + I)(y 0 + I) = x0 y 0 + I = I. Then by the first part of the proof,
x0 + I 6= I means (y 0 + I)n = (y 0 )n + I = I for some n ∈ N, and thus (y 0 )n ∈ I.

Theorem 3.4.6 (Primary Decomposition). For any proper ideal I ⊂ A, there exist finitely
many primary ideals Q1 , . . . , Qn such that I = Q1 ∩ · · · ∩ Qn .

Proof. Apply Lemmas 3.4.4 and 3.4.5.

Lemma 3.4.7. If Q ⊂ A is a primary ideal, then r(Q) is the smallest prime ideal of A
containing Q.

Proof. Assume x, y ∈ A with xy ∈ r(Q). Then (xy)n = xn y n ∈ Q for some n ∈ N, so xn ∈ Q


or y n ∈ r(Q). In the first case, we see immediately that x ∈ r(Q). Otherwise, y n ∈ r(Q)
implies that y ∈ r(r(Q))T= r(Q) by Lemma 1.1.4(c). Hence r(Q) is a prime ideal. Now, by
Corollary 1.1.3, r(Q) = {p0 prime | Q ⊆ p0 ⊂ A} but since r(Q) is prime, it is the smallest
prime containing Q.

Definition. For a prime ideal p ⊂ A, any proper ideal Q ⊂ A is called p-primary if


p = r(Q).

Examples.

1 If A is a UFD and p ∈ A is a prime element, then (pn ) is p-primary (the usual notation
for (p)-primary) for any n ∈ N.

2 In general, a p-primary ideal need not be a power of p. Let A = k[x, y] and Q = (x, y 2 ).
Then A/Q ∼ = k[y](y 2 ) and in this quotient, the zero divisors are the multiplies of
y, hence nilpotent. This implies Q is primary, and clearly its radical is p = (x, y).
However, p2 ( Q ( p are strict inclusions, so Q is not a power of a prime ideal.

3 If p is a prime ideal, pn need not always be primary. Let A = k[x, y, z]/(xy − z 2 ) and
let x̄, ȳ, z̄ ∈ A be the respective images of x, y, z in the quotient. Set p = (x̄, z̄). Then
p is a prime ideal of A since A/p ∼ = k[y] is an integral domain. However, x̄ȳ = z̄ 2 ∈ p2 ,
but x̄ 6∈ p2 and ȳ 6∈ r(p2 ) = p. So p2 is not primary.

The next result shows that the counterexamples in 2 and 3 do not occur when we
consider maximal ideals.

Lemma 3.4.8. Let A be a noetherian ring.

(a) If I ⊂ A is a proper ideal such that r(I) is maximal, then I is primary.

(b) If m ⊂ A is a maximal ideal, then mn is primary for all n ∈ N.

55
3.4. Primary Decomposition Chapter 3. Noetherian and Artinian Rings

Proof. (a) Assume r(I) is maximal. Then by Corollary 1.1.3, the set of all prime ideals in
A containing I just consists of r(I). Thus B = A/I is a local ring with unique maximal
ideal r(I)/I. If x, y ∈ B such that xy = 0 but y 6∈ r(I)/I, then since r(I)/I is the unique
maximal ideal, y ∈ B × . So xy = 0 =⇒ xyy −1 = 0 =⇒ x = 0. Now if x0 , y 0 ∈ A such that
x0 y 0 ∈ I but y 0 6∈ r(I), then by the previous statement,

(x0 + I)(y 0 + I) = I =⇒ x0 + I = I =⇒ x0 ∈ I.

This shows I is primary.


(b) By Lemma 1.1.4, r(mn ) = r(m) = m. By (a), this shows that mn is m-primary.
Corollary 3.4.9. If m ⊂ A is a maximal ideal and I ⊂ A is any proper ideal, then I is
m-primary if and only if there is some n ∈ N such that mn ⊆ I ⊆ m.
Proof. ( =⇒ ) By Lemma 1.1.4, I ⊆ r(I) = m. Then by Lemma 3.4.1, r(I)n = mn ⊆ I for
some n.
( ⇒= ) Conversely, by Lemma 3.4.8(a), m = r(mn ) ⊆ r(I) ⊆ r(m) = m so we have
r(I) = m. By definition, this means I is m-primary.
Proposition 3.4.10. Assume I ⊂ A is a proper ideal and I = ni=1 Qi for primary ideals
T
Qi . If p1 , . . . , pn are (not necessarily distinct) prime ideals such that Qi is pi -primary for
each 1 ≤ i ≤ n, then
(a) For every prime ideal p ⊂ A such that p ⊇ I, p ⊇ pi for some 1 ≤ i ≤ n.
(b) The minimal elements of the collection {p ⊂ A prime | p ⊇ I} are exactly the minimal
elements of {pi }ni=1 .
(c) If A is noetherian, it only has finitely many minimal prime ideals.
Proof. (a) If p ⊇ I = ni=1 Qi then by Lemma 1.1.4,
T

n
\ n
\
p = r(p) ⊇ r(I) ⊇ r(Qi ) = pi .
i=1 i=1

Thus p contains one of the pi .


(b) If p ⊂ A is a minimal prime ideal containing I, then by (a), p ⊇ pi for some pi . But
by minimality, p = pi .
(c) Apply (b) and primary decomposition (3.4.6) to the ideal I = (0).
Lemma
Tn 3.4.11. If Q1 , . . . , Qn are p-primary ideals for a fixed prime ideal p ⊂ A, then
Q
i=1 i is also p-primary.
Tn Tn Tn
Tn r(Qi ) = p for all 1 ≤ i ≤ n, so r ( i=1 Qi ) = Tni=1 r(Qi ) = i=1
Proof. By assumption Tnp = p.
It remains to show i=1 Qi is primary. Assume x, y ∈ A with xy ∈ i=1 Qi but x 6∈ i=1 Qi .
Then for each i, xy ∈ Qi but T Tn that x 6∈ Qj . Since Qj is p-primary, we
there is some j such
n
must have y ∈ r(Qi ) = p = r ( i=1 ). Therefore i=1 Qi is primary.
Definition. A primaryTdecomposition I = ni=1 Qi is called minimal (or reduced) if for
T
every 1 ≤ i ≤ n, Qi 6⊃ j6=i Qj and r(Qi ) 6= r(Qj ) whenever i 6= j.

56
3.4. Primary Decomposition Chapter 3. Noetherian and Artinian Rings

By Theorem 3.4.6 and Lemma 3.4.11, minimal primary decompositions exist for all ideals
I ⊂ A.
Example 3.4.12. Let A = k[x, y, z] and consider the ideals p1 = (x, y), p2 = (x, z) and
m = (x, y, z). Then A/p1 ∼ = k[z] and A/p2 ∼ = k[y] which are integral domains, so p1 and p2

are prime. Also, A/m = k which is a field, so m is a maximal ideal of A. We claim that
p1 ∩ p2 = (x, yz). Clearly x and yz both lie in (x, y) ∩ (x, z), so (x, yz) ⊆ p1 ∩ p2 . On the
other hand, for any f ∈ p1 ∩ p2 , f ∈ (x, y) implies f = gx + hy for g, h ∈ A. But f ∈ (x, z) is
possible if and only if h = h0 z for some h0 ∈ A, in which case we have f = gx+h0 yz ∈ (x, yz).
Hence p1 ∩ p2 = (x, yz).
Now we show that p1 p2 = p1 ∩ p2 ∩ m2 and this is a minimal primary decomposition
of p1 p2 . The left can be written p1 p2 = (x2 , xy, xz, yz). Notice that x2 , xy, xz and yz all
lie in m2 , and since p1 p2 ⊆ p1 ∩ p2 , this shows that x2 , xy, xz, yz ∈ p1 ∩ p2 ∩ m2 . Hence
p1 p2 ⊆ p1 ∩ p2 ∩ m2 . Going the other way, suppose F ∈ p1 ∩ p2 ∩ m2 . Then F ∈ m2 so
F = ax2 + bxy + cxz + dy 2 + eyz + f z 2 for some a, b, c, d, e, f ∈ A. Notice that by the
paragraph above, ax2 + bxy + cxz + eyz ∈ p1 p2 so to show that F ∈ p1 p2 , it will suffice to
show that dy 2 + f z 2 ∈ p1 p2 . By the previous statement, ax2 + bxy + cxz + eyz ∈ p1 ∩ p2
so because F ∈ p1 ∩ p2 , we must have dy z + f z 2 ∈ p1 ∩ p2 . Observe that f z 2 ∈ p2 so it
follows that dy 2 ∈ p2 . But y 2 6∈ p2 and this ideal is prime, so we must have d ∈ p2 . A similar
argument shows that f ∈ p1 . Thus we have dy 2 ∈ p2 p1 and f z 2 ∈ p1 p2 , so dy 2 + f z 2 ∈ p1 p2 .
This proves F ∈ p1 p2 , so p1 ∩ p2 ∩ m ⊆ p1 p2 .
Since p1 and p2 are prime, they are primary (3.4.5). Also, since m is maximal, m2 is
primary (3.4.8(b)). Each term in the primary decomposition has a distinct radical, so we
need only check that none of them contains the intersection of the remaining two. Notice:

x ∈ p1 ∩ p2 but x 6∈ m2 ,
y 2 ∈ p1 ∩ m2 but y 2 6∈ p2 ,
and z 2 ∈ p2 ∩ m2 but z 2 ∈
6 p1 .

This shows p1 p2 = p1 ∩ p2 ∩ m2 is a minimal primary decomposition.


Definition. Let I ⊂ A be an ideal and take x ∈ A. The colon ideal, or ideal quotient of
I by x is defined as
(I : x) = {y ∈ A | xy ∈ I}.
If J ⊂ A is another ideal, we can define the ideal quotient of I by J as

(I : J) = {y ∈ A | yJ ⊆ I}.

The following lemma collects some basic facts about ideal quotients.
Lemma 3.4.13. Let I, J ⊂ A be ideals and x ∈ A. Then
(a) (I : J) is an ideal of A and I ⊆ (I : J).

(b) x ∈ I if and only if (I : x) = A.


T  T
(c) For a collection of ideals {Ij } of A, j Ij : x = j (Ij : x).

57
3.4. Primary Decomposition Chapter 3. Noetherian and Artinian Rings

(d) AnnA (x) = (0 : x).


(e) Let D = {z ∈ A | yz = 0 for some y ∈ A r {0}} be the set of zero divisors of A. Then
[ [ [
D= (0 : x) = AnnA (x) = r(AnnA (x)).
x6=0 x6=0 x6=0

Proof. (a) Since IJ ⊆ I, I ⊆ (I : J) is clear. To prove (I : J) is an ideal, let y, z ∈ (I : J)


and r ∈ A. Then for any j ∈ J, yj, zj ∈ I and so (ry − z)j = ryj − zj = r(yj) − zj ∈ I.
Hence (I : J) is an ideal.
(b) This is the definition of an ideal I.
(c) Observe that
!
\ \
y∈ Ij : x ⇐⇒ xy ∈ Ij
j j

⇐⇒ xy ∈ Ij for each Ij
⇐⇒ y ∈ (Ij : x) for each j
\
⇐⇒ y ∈ (Ij : x).
j

(d) This is the definition of the annihilator.


(e) The first equality is trivial and the second is immediate from (d). To prove the
third, we have AnnA (x) ⊆ r(AnnA (x)) by Lemma 1.1.4(a). On the other hand, for any
y ∈ r(AnnA (x)), there is some n ≥ 1 such that y n ∈ AnnA (x), i.e. xy n = 0. Then y ∈ D so
all equalities hold.
Lemma 3.4.14. If Q ⊂ A is a p-primary ideal and x ∈ ArQ then (Q : x) is also p-primary.
Proof. Fix x ∈ A r Q. Then for any y ∈ (Q : x), xy ∈ Q and since Q is primary, this implies
y ∈ r(Q) = p. So (Q : x) ⊆ p. By Lemma 3.4.13(a), we have
p = r(Q) ⊆ r((Q : x)) ⊆ r(p) = p.
Therefore r((Q : x)) = p. It remains to show (Q : x) is primary. Suppose y, z ∈ A such that
yz ∈ (Q : x) but z 6∈ r((Q : x)) = p. Then xyz ∈ Q but z 6∈ r(Q) = p, so xy ∈ Q since Q is
primary. By definition this implies y ∈ (Q : x). Hence (Q : x) is p-primary.
Proposition 3.4.15. If ni=1 Qi = 0 is a minimal primary
T
Sn decomposition of the zero ideal,
with Qi a pi -primary ideal for each 1 ≤ i ≤ n, then i=1 pi ⊆ D.
T
Proof. By minimality, for each 1 ≤ i ≤ n there is some xi ∈ j6=i Qj such that xi 6∈ Qi .
Then by Lemma 3.4.13(c) and (d),
n
! n
\ \
AnnA (xi ) = (0 : xi ) = Qj : xi = (Qj : xi ) = (Qi : xi ).
j=1 j=1

S each 1 ≤ i ≤ n. Moreover, pi = r((Qi : xi )) =


By Lemma 3.4.14, (Qi : xi ) is pi -primary for
r(AnnA (xi )) ⊆ D by Lemma 3.4.13(e), so ni=1 pi ⊆ D as claimed.

58
3.4. Primary Decomposition Chapter 3. Noetherian and Artinian Rings

Corollary 3.4.16. If A is noetherian and p ⊂ A is a minimal prime ideal, then p ⊆ D.

Proof. By Theorem 3.4.6, (0)Tnhas a primary decomposition and we may assume it is minimal
by Lemma 3.4.11. Set 0 = i=1 Qi where Qi ⊂ A is a pi -primary ideal for some prime ideal
pi . By Proposition 3.4.10, the minimal prime ideals of A are elements of {pi | 1 ≤ i ≤ n}
and by Proposition 3.4.15, each pi ⊆ D. Thus all minimal prime ideals of A are contained
in D.

Lemma 3.4.17. If Qi is a pi -primary ideal for 1 ≤ i ≤ n and pi + pj = A whenever i 6= j,


then n n
Y \
Qi = Qi .
i=1 i=1
Qn Tn
Proof. We always have i=1 Qi ⊆ i=1 Qi so it suffices to show the reverse inclusion. First
note that by Lemma 1.1.4(f), whenever j 6= i, we have

r(Qi + Qj ) = r(r(Qi ) + r(Qj )) = r(pi + pj ) = r(A) = A.

So 1 ∈ r(Qi + Qj ) which implies Qi + Qj = A. Now if n = 2 and x ∈ Q1 ∩ Q2 then the


previous statement tells us that there exist q1 ∈ Q1 , q2 ∈ Q2 such that

x = x · 1 = x(q1 + q2 ) = xq1 + xq2 ∈ Q1 Q2 .

Therefore the statement holds when n = 2. To induct, we need to show that Q1 · · · Qn−1 +
Qn = A. By hypothesis, for each 1 ≤ i ≤ n − 1 there are elements xi ∈ Qi , yi ∈ Qn−1 such
that xi + yi = 1. Then
n−1
Y
1= (xi + yi ) ∈ x1 · · · xn−1 + Qn ⊆ Q1 · · · Qn−1 + Qn .
i=1

Therefore Q1 · · · Qn−1 + Qn = A. Finally, using the base case and inductive hypothesis, we
have ! ! !
\n n−1
\ n−1
\ n−1
Y
Qi = Qi ∩ Qn = Qi ∩ Qn = Qi Qn .
i=1 i=1 i=1 i=1

Hence the statement holds for any n.


In the next proposition, we generalize the correspondence of Proposition 1.3.8 to a bijec-
tion on primary ideals.

Proposition 3.4.18. Let S ⊆ A be a multiplicatively closed subset and suppose Q ⊂ A is a


p-primary ideal. Then

(a) If S ∩ p 6= ∅ then S −1 Q = S −1 p.

(b) If S ∩ p = ∅ then j −1 (S −1 Q) = Q and S −1 Q is S −1 p-primary.

59
3.4. Primary Decomposition Chapter 3. Noetherian and Artinian Rings

(c) The correspondence


 0 
Q ⊂ A is primary and
Φ: ←→ {primary ideals of S −1 A}
r(Q0 ) ∩ S = ∅

given by Q 7→ S −1 Q is bijective and inclusion-preserving.


n
Proof. (a) Take s ∈ S ∩ p. Then for some n ∈ N, sn ∈ Q ∩ S which implies that s1 ∈ S −1 Q,
n
but s1 is a unit in S −1 A, so we must have S −1 Q = S −1 A.
(b) The inclusion Q ⊆ j −1 (S −1 Q) is trivial: if q ∈ Q then 1q ∈ S −1 Q, so q ∈ j −1 (S −1 Q) by
definition of the map j. For the reverse inclusion, suppose x ∈ j −1 (S −1 Q). Then x1 ∈ S −1 Q
so there exist elements q ∈ Q, s ∈ S such that x1 = qs . Hence for some t ∈ S, t(xs − q) = 0,
i.e. stx = tq ∈ Q. Now stx ∈ Q but st ∈ S, so since S ∩ p = ∅, we must have st 6∈ p. Since
Q is p-primary, this implies x ∈ Q. This proves Q = j −1 (S −1 Q).
Next, r(Q) = p by assumption and r(S −1 Q) = S −1 (r(Q)) = S −1 p, so it remains to
show S −1 Q is a primary ideal. Suppose xs , yt ∈ S −1 A such that xy st
∈ S −1 Q. Then there are
xy q
elements u ∈ S, q ∈ Q such that st = u . Thus for some v ∈ S, v(uxy − stq) = 0, that is,
vuxy = stq ∈ Q. Since vu ∈ S by multiplicative closure, vu 6∈ p as before, so xy ∈ Q because
Q is primary. Thus x ∈ Q or y ∈ r(Q), which implies xs ∈ S −1 Q or yt ∈ S −1 (r(Q)) = S −1 p.
Hence S −1 Q is S −1 p-primary.
(c) By part (b), Φ is injective and for any primary ideal Q0 ⊂ A, S −1 Q0 is primary in
S −1 A. Inclusion-preserving is as in Proposition 1.3.8. Suppose J ⊂ S −1 A is a primary ideal.
To show Φ is surjective, it’s enough to show I = j −1 (J) is a primary ideal of A, since then
Proposition 1.3.8(a) says that S −1 I = S −1 (j −1 (J)) = J. Take x, y ∈ A such that xy ∈ I
but x 6∈ I. Then xy 1
∈ S −1 I = J so since J is primary, x1 ∈ J or y1 ∈ r(J). The former
is impossible, since x1 ∈ J = S −1 I would imply x ∈ I. So we must have y1 ∈ r(J). Then
n
for some n ≥ 1, y1 ∈ J, and thus y n ∈ I. Thus I is primary. We conclude that Φ is an
inclusion-preserving bijection.

60
3.5. Artinian Rings Chapter 3. Noetherian and Artinian Rings

3.5 Artinian Rings


Recall from Section 3.1 that a (commutative) ring A is artinian if the descending chain
condition holds on the set of ideals of A. We have seen many examples of noetherian rings,
but it turns out that the artinian condition is much more restrictive. Assume that A 6= 0.
Example 3.5.1. If k is a field and A is a finite dimensional k-algebra, then A is artinian.
Proposition 3.5.2. Let A be artinian. Then
(a) If A is an integral domain then A is a field.
(b) dim A = 0.
(c) J(A) = N (A).
Proof. (a) Let x ∈ A, x 6= 0, be given. Then the ideals generated by powers of x form a
descending chain in A:
(x) ⊇ (x2 ) ⊇ (x3 ) ⊇ · · ·
By the descending chain condition, there is some n ∈ N such that (xn ) = (xn+1 ). So there is
a y ∈ A such that xn = xn+1 y, that is, xn (1 − xy) = 0. Since A is a domain, xn 6= 0 so we
must have 1 − xy = 0. Hence x is a unit.
(b) If A is any artinian ring with a prime ideal p ⊂ A, then A/p is also artinian by
Lemma 3.1.3. Since A/p is a domain, by (a) it is a field, so p is maximal. Thus all prime
ideals in A are maximal, and thus also minimal, so dim A = 0.
(c) follows directly from the definitions of J(A) and N (A) as intersections of maximal
and prime ideals.
Lemma 3.5.3. If A is artinian then Spec(A) is finite.
Proof. Consider the collection Σ = {p1 ∩ · · · ∩ pn | pi ⊂ is a prime ideal and n ≥ 1}. Since
A 6= 0, Σ is nonempty. By Lemma 3.1.1, Σ has a minimal element, say I = p1 ∩ · · · pn . For
any prime ideal p ∈ Spec(A), I ∩ p ∈ Σ but by minimality, I = I ∩ p, so p ⊇ I. Since p
is prime, we must have p ⊇ pi for some 1 ≤ i ≤ n. Now all prime ideals are maximal by
Proposition 3.5.2(b), so p = pi . Hence Spec(A) = {p1 , . . . , pn }.
Proposition 3.5.4. If A is artinian, then N (A)n = 0 for some n ∈ N.
Proof. Consider the descending chain
N (A) ⊇ N (A)2 ⊇ N (A)3 ⊇ · · ·
By DCC, there is an n ∈ N such that N (A)m = N (A)n for all m ≥ n. Set I = N (A)n
and assume I 6= 0. Then the collection Σ = {ideals J ⊂ A | IJ 6= 0} is nonempty. By
Lemma 3.1.1, Σ contains a minimal element J0 . Then IJ0 6= 0 so there exists an element
x ∈ J0 so that xI 6= 0. This means (x) ∈ Σ, but by minimality of J0 , we must have (x) = J0 .
Now I 2 = N (A)2n = N (A) = I so we have 0 6= xI = xI 2 = (xI)I. Thus xI ∈ Σ and
xI ⊆ (x) = J0 . By minimality, this implies xI = J0 = (x), and thus there is some y ∈ I so
that x = xy. By induction, x = xy m for all m ∈ N, but y ∈ I ⊆ N (A) so y is nilpotent.
Thus for some m, y m = 0, giving us x = xy m = 0, contradicting the initial assumption that
x 6= 0. Hence I = N (A)n = 0.

61
3.5. Artinian Rings Chapter 3. Noetherian and Artinian Rings

Theorem 3.5.5 (Hopkins). For a (commutative) ring A, the following are equivalent:

(i) A is artinian.

(ii) A is noetherian and dim A = 0.

Proof. By Proposition 3.5.2(b), A artinian directly implies dim A = 0. Assuming (i),


Lemma 3.5.3 implies that A has finitely many maximal (and prime) ideals m1 , . . . , mn . Then

N (A) = J(A) = m1 ∩ · · · mn ⊇ m1 · · · mn .

By Proposition 3.5.4, there is some k ∈ N such that N (A)k = 0, but (m1 · · · mn )k ⊆ N (A)k =
0 so by Proposition 3.1.7, A is noetherian. On the other hand, assuming (ii) implies that
A has finitely many minimal prime ideals, by Proposition 3.4.10. But if dim A = 0, each
minimal prime ideal is maximal as well. Thus by Corollary 3.4.2, N (A)k = 0 for some k ∈ N.
Then once again the hypothesis of Proposition 3.1.7 is satisfied, so A is artinian.

Theorem 3.5.6. Let A be a local noetherian ring with unique maximal ideal m. Then one
of the following conditions holds:

(i) mn 6= mn+1 for all n ∈ N, and thus A is not artinian.

(ii) mn = 0 for some n ∈ N, in which case A is artinian.

Proof. Consider the descending chain

m ⊇ m2 ⊇ m3 ⊇

If mn 6= mn+1 for all n ∈ N, the chain is not stationary so A is not artinian. Clearly in this
case mn 6= 0 for any n ∈ N, or else the chain would stabilize. Alternatively, if mn = mn+1
for some n, the locality condition implies mn+1 = J(A)mn . Since A is noetherian, mn is a
finitely generated submodule of A, so Nakayama’s Lemma (1.2.1) implies mn = 0. Now if
p ⊂ A is a prime ideal, 0 = mn ⊆ p so we must have m ⊆ p; otherwise, if x ∈ m r p then
xn ∈ p, which is impossible. By maximality, m = p so m is in fact the unique prime ideal of
A. Hence dim A = 0, so by Theorem 3.5.5, A is artinian.
Examples.

1 If k is a field, A = k[[t]] is noetherian by Theorem 3.3.9. Also, A is local with maximal


ideal m = (t). In this ring, (0) is prime so dim A = 1. Therefore A cannot be artinian,
so by Theorem 3.5.6, mn = (tn ) 6= 0 for all n ∈ N.

2 Choose a prime p ∈ Z, an integer n ∈ N and set A = Z/pn Z. Since A is finite, it is


local artinian with maximal ideal pZ/pn Z.

3 Let k be a field and consider the polynomial ring B = k[t]. If f ∈ B is a prime


element, i.e. an irreducible polynomial in t, then A = B/(f n ) is a local artinian ring
with maximal ideal (f )/(f n ).

62
3.5. Artinian Rings Chapter 3. Noetherian and Artinian Rings

Proposition 3.5.7. Let A be a local artinian ring with maximal ideal m and residue field
k = A/m. Then the following are equivalent:
(a) Every ideal of A is principal.

(b) m is principal.

(c) dimk m/m2 ≤ 1, where m/m2 is viewed as a k-vector space.


Proof. (a) =⇒ (b) is trivial and (b) =⇒ (c) follows from the fact that if m = Ax, then
m/m2 = k(x + m2 ) which has dimension at most 1.
To prove (c) =⇒ (a), note that if dimk m/m2 ≤ 1 then there is some x ∈ m such that
m/m2 = k(x+m2 ). By Theorem 3.5.5, A is noetherian so in particular m is finitely generated
as an A-module. Thus by Nakayama’s Lemma (1.2.1), with M = N = J(A) = m, we get
N/M N = m/m2 =⇒ m = (x). Now Proposition 3.5.4 implies there exists an n ∈ N such
that mn = N (A)n = 0. Take a nonzero ideal I ⊂ A. Then I ⊆ m and there is an ` ∈ N such
that I ⊆ m` = (x` ) but I 6⊆ m`+1 = (x`+1 ). So there is a y ∈ I such that y = ax` for some
a ∈ A r m. Since A is local, this a must be a unit in A. Thus we have

(y) ⊆ I ⊆ m` = (x` ) = (y),

so I = (y) and I = m` as claimed.


Proposition 3.5.8. Let k be a field and A a finitely generated k-algebra which is also a local
artinian ring. Then dimk A is finite.
Proof. For any local ring A which is finitely generated as a k-algebra, we have surjections
k[t1 , . . . , tr ] → A → A/m so that A/m is also a finitely generated k-algebra. By the algebraic
Nullstellensatz (2.4.2), A/m is a finite (algebraic) extension.
Since A is artinian, A is also noetherian by Theorem 3.5.5. Then Theorem 3.5.6 shows
that mn = 0 for some n ∈ N, in which case the chain of submodules m` ⊇ m`+1 eventually
stabilizes:
A ⊇ m ⊇ m2 ⊇ · · · ⊇ mn−1 ⊇ mn = 0.
For each 0 ≤ ` ≤ n − 1, m` is a finitely generated A-module (since A is noetherian), so
each quotient m` /m`+1 is a finite dimensional A/m-vector space. By the first paragraph and
transitivity of dimension, we then have that

dimk m` /m`+1 = (dimk A/m)(dimA/m m` /m`+1 ) < ∞

for each 0 ≤ ` ≤ n − 1. In particular, taking ` = n − 1 shows that mn−1 /mn = mn−1 is


finite dimensional over A/m. From linear algebra, (∗) if W ⊆ V is a vector subspace over a
field F such that W and V /W are both finite dimensional F -vector spaces, then V is finite
dimensional. Taking V = mn−2 and W = mn−1 , we now have that mn−2 is a finite dimensional
A/m-vector space. By induction, m` is finite dimensional for all 1 ≤ ` ≤ n − 1. In particular,
m1 = m is finite dimensional over A/m, and of course A/m is finite dimensional over itself,
so by (∗), A is a finite dimensional A/m-vector space. This proves the statement.
The structure of commutative artinian rings is completely described in the next theorem.

63
3.5. Artinian Rings Chapter 3. Noetherian and Artinian Rings

Theorem 3.5.9. Any artinian ring A is the finite direct product of local artinian rings, and
the factors are uniquely determined by A up to isomorphism.

Corollary 3.5.10. If A is a finitely generated k-algebra which is artinian, then dimk A is


finite.

Proof. By Theorem 3.5.9, A is the direct product of finitely many local artinian rings, and
each of these has finite dimension by Proposition 3.5.8.

64
3.6. Associated Primes Chapter 3. Noetherian and Artinian Rings

3.6 Associated Primes


Definition. A prime filtration of an A-module M is a sequence of submodules

0 = M0 ⊆ M1 ⊆ · · · ⊆ Mn = M

such that for each 0 ≤ i ≤ n − 1, Mi+1 /Mi ∼


= R/pi+1 for some prime ideals p1 , . . . , pn ⊂ A.
Definition. Let M be an A-module. A prime ideal p ⊂ A is an associated prime of M if
there is an embedding of A-modules A/p ,→ M . Let Ass(M ) denote the set of all associated
primes of M .
Lemma 3.6.1. For any A-module M ,

Ass(M ) = {p ⊂ A prime | there is some nonzero x ∈ M with AnnA (x) = p}.

Proof. If p ∈ Ass(M ) then x = 1 + p ∈ A/p ,→ M is a nonzero element of M and by


definition AnnA (x) = p. Conversely, if p = AnnA (x) for some x ∈ M , then the natural map
ϕx : A → M, a 7→ ax has kernel p so it factors through A/p ∼= im ϕx ⊆ M .
We will prove that every A-module has a prime filtration, but first we need a lemma.
Lemma 3.6.2. For all nonzero A-modules M , Ass(M ) is nonempty.
Proof. Consider the set

Λ = {I = AnnA (x) | x ∈ M, x 6= 0}.

Then since M 6= 0, Λ is nonempty. Let I = AnnA (x) be a maximal element of Λ; we claim


I is a prime ideal of A. Suppose a, b ∈ A such that ab ∈ I, so that abx = 0. If a, b 6∈ I then
ax 6= 0, but then b ∈ AnnA (ax) which shows I ( AnnA (ax). This contradicts maximality of
I, so either a or b lies in I, showing I is prime.
Theorem 3.6.3. If A is noetherian, then every finitely generated A-module has a prime
filtration.
Proof. Let M be a finitely generated A-module and consider the collection of submodules

Γ = {N ⊆ M | N has a prime filtration}.

By Lemma 3.6.2, there exists a prime p ∈ Ass(M ) so the submodule A/p ,→ M has trivial
prime filtration 0 ⊆ A/p. Thus Γ is nonempty. Let N be a maximal element of Γ, say with
prime filtration
0 ⊆ N1 ⊆ · · · ⊆ Nk = N.
If N 6= M , use Lemma 3.6.2 again to choose p ∈ Ass(M/N ) with p = AnnA (x + N ) for some
x ∈ M r N . Then N + Ax has a prime filtration:

0 ⊆ N1 ⊆ · · · ⊆ Nk = N ( N + Ax

with (N + Ax)/N ∼
= AnnA (x + N ) = p, contradicting maximality of N . Hence M = N .

65
3.6. Associated Primes Chapter 3. Noetherian and Artinian Rings

Lemma 3.6.4. For every short exact sequence of A-modules 0 → K → M → N → 0,


Ass(M ) ⊆ Ass(N ) ∪ Ass(K).
Proof. Suppose p ∈ Ass(M ) so that by Lemma 3.6.1, p = AnnA (y) for some nonzero y ∈ M .
If y ∈ N then p ∈ Ass(N ) already. Otherwise, u = y + N ∈ M/N is a nonzero element of K.
Suppose p 6= AnnA (u). Then there is some s ∈ AnnA (u) such that s 6∈ p and 0 = su = sy +N
in K, i.e. sy ∈ N . Then p = AnnA (y) ⊆ AnnA (sy). On the other hand, if t ∈ AnnA (sy)
then sty = 0, so st ∈ p but since s 6∈ p and p is prime, this means t ∈ p. Hence p = AnnA (y)
or AnnA (sy), so p lies in either Ass(K) or Ass(N ).
Corollary 3.6.5. For any A-module M , Ass(M n ) = Ass(M ) for all n ≥ 1.
Example 3.6.6. If A is an integral domain, then Ass(A) = {0} (by Lemma 3.6.1 for
example).
Theorem 3.6.7. Let A be a noetherian ring and M a finitely generated A-module. Then
(1) Let p ⊂ A be a prime ideal. Then p ∈ Ass(M ) if and only if pAp ∈ Ass(Mp ).
(2) | Ass(M )| < ∞.
(3) If p is minimal among prime ideals containing AnnA (M ) then p ∈ Ass(M ).
Proof. (1) On one hand, if p ∈ Ass(M ) then A/p ,→ M by definition, but since localization
is exact (see Example 8.3.8), (A/p)p ,→ Mp as well so pAp ∈ Ass(Mp ). Conversely, suppose
pAp ∈ Ass(Mp ). By Lemma 3.6.1, pAp = AnnAp (α) for some nonzero α ∈ Mp . Write
p = (x1 , . . . , xn ) so that xi α = 0 for all 1 ≤ i ≤ n. Then there exist elements si ∈ A r p
such that si xi α = 0. Set s = s1 · · · sn . Since p is prime, s 6∈ p and we have sα ∈ M and
xi (sα) = sxi α = 0 for each i. Thus p ⊆ AnnA (sα). If t ∈ AnnA (sα) r p then we would
have (ts)α = 0 but this implies α = 0 in Mp , a contradiction. Hence p = AnnA (sα) so
p ∈ Ass(M ).
(2) By Theorem 3.6.3, M has a prime filtration
0 = M0 ⊆ M1 ⊆ · · · ⊆ Mn = M
with Mi+1 /Mi ∼
= A/pi+1 for some primes p1 , . . . , pn ⊂ A. If n = 1, notice that M1 ∼
= A/p1 is
a domain so Ass(M1 ) = {p1 } by Example 3.6.6. For n ≥ 2, we have a short exact sequence
0 → Mn−1 → M → M/Mn−1 → 0
so Lemma 3.6.4 says that Ass(M ) ⊆ Ass(Mn−1 ) ∪ Ass(M/Mn−1 ). By induction, Ass(Mn−1 )
is finite, and Ass(M/Mn−1 ) = Ass(A/pn ) = {pn } so Ass(M ) is finite as well.
(3) Suppose p is minimal among all prime ideals of A containing AnnA (M ). Then Mp 6= 0
and Mp is an (A/ AnnA (M ))p -module, so we may assume AnnA (M ) = 0. Now p is minimal,
so Spec(Ap ) = {pAp }. By (1), we may replace A by Ap and M by Mp . Consider the filtration
M ⊇ pM ⊇ p2 M ⊇ · · ·
Since A is noetherian and Spec(A) = {p}, Theorem 3.5.5 shows that A is artinian and thus
pn+1 M = pn M for some smallest n ≥ 1. Then p(pn M ) = pn M so Nakayama’s lemma (1.2.1)
says that pn M = 0. Take y ∈ pn−1 M, y 6= 0. Then p ⊆ AnnA (y) but since Spec(A) = {p}, p
is a maximal ideal and hence p = AnnA (y). So by Lemma 3.6.1, p ∈ Ass(M ).

66
3.6. Associated Primes Chapter 3. Noetherian and Artinian Rings

Remark. The proof of (2) shows that if p1 , . . . , pn are the primes corresponding to a prime
filtration of M , then Ass(M ) ⊆ {p1 , . . . , pn }.

Example 3.6.8. Let A = k[x, y](x2 − y 3 ) and define an A-module M by the free resolution
 

2x 
−3y 2
0 → A −−−−−−→ A2 → M → 0.

(M is called the module of Kähler differentials of A; see Section 15.2 and specifically Exam-
ple 15.4.3 for more details). We claim the only associated primes of M are 0 and (x, y). Let
p be a prime ideal of A and consider the localized sequence

0 → Ap → A2p → Mp → 0

which remains exact  since localization is exact.


  If p 6=  y) then x 6= 0 or y 6= 0, so
 (x,
2x 1 0
the matrix can be put into the form or after a change of basis. Thus
−3y 2 0 1
Mp ∼ = Ap so Ass(Mp ) = {0} and thus by Theorem 3.6.7(1), p 6∈ Ass(M ). On the other hand,
if p = (x, y) then consider the element
 3 
2x
m= ∈ M.
−3y 5

Notice that
     2 
2x4 2x4 2 2x
xm = 5 = 2 2 =x =0
−3xy −3x y −3y 2
 3     
2x y 2x3 y 2 2x
while ym = = =x y = 0.
−3y 6 −3x2 y 3 −3y 2

Hence AnnA (m) = (x, y) so (x, y) is an associated prime of M .

Proposition 3.6.9. For a noetherian ring A and a finitely generated A-module M ,


[
p = {a ∈ A | ax = 0 for some nonzero x ∈ M }.
p∈Ass(M )

Proof. Let Z be the set of all a ∈ A for which ax = 0 for some x 6= 0 in M . It is clear that
any p ∈ Ass(M ) is contained in Z, so one inclusion is true. On the other hand, suppose
a ∈ Z with ax = 0 for some nonzero x ∈ M . The set

Λ = {AnnA (bx) | b ∈ A, b 6= 0}

is nonempty since AnnA (ax) ∈ Λ, so it has a maximal element, say q = AnnA (bx) for some
b 6= 0. If we can show q is prime, then we are done by Lemma 3.6.1. Suppose uv ∈ q
but u 6∈ q. Then ubx 6= 0 and q ⊆ AnnA (ubx), so by maximality, q = AnnA (ubx). Hence
v ∈ AnnA (ubx) = q which shows q is prime as desired.

67
Chapter 4

Discrete Valuations and Dedekind


Domains

68
4.1. Discrete Valuation Rings Chapter 4. Discrete Valuations and Dedekind Domains

4.1 Discrete Valuation Rings


Definition. A discrete valuation on a field K is a surjective group homomorphism v :
K × → Z such that v(x + y) ≥ min(v(x), v(y)) for all x, y ∈ K × . We often extend v to all of
K by setting v(0) = ∞.
Lemma 4.1.1. Let A be an integral domain with field of fractions K. If v 0 : A r {0} → N0
is a function satisfying
(i) v 0 (xy) = v 0 (x) + v 0 (y)
(ii) v 0 (x + y) ≥ min(v 0 (x), v 0 (y))
for all x, y ∈ A r {0} and v 0 (0) = ∞ when necessary, then there exists a unique discrete
valuation v : K × → Z which extends v 0 , i.e. v|A = v 0 .
Proof. Define v : K × → Z by v ab = v 0 (a)−v 0 (b). Then one can check that v is well-defined,


a surjective homomorphism and satisfies v(x + y) ≥ min(v(x), v(y)).


Corollary 4.1.2. If A is a UFD with field of fractions K and p ∈ A is a prime element,
there is a discrete valuation vp : K × → Z such that for any a ∈ A, vp (a) = n if pn | a and
pn+1 - a.
Proof. It’s enough to show that vp satisfies conditions (i) and (ii) of Lemma 4.1.1. Take
x, y ∈ A r {0} and write x = pn x0 and y = pm y 0 where n, m ≥ 0 and p - x0 , y 0 . Then
xy = pn+m x0 y 0 and since p is prime, p - x0 y 0 . Thus vp (xy) = n + m, so (i) is satisfied. On
the other hand, suppose without loss of generality that n ≥ m, so that min(v(x), v(y)) = m.
Then x + y = pn x0 + pm y 0 = pm (pn−m x0 + y 0 ) but since p | pn−m x0 and p - y 0 , we have
p - (pn−m x0 + y 0 ). Thus vp (x + y) = m and condition (ii) is satisfied.
Definition. For a prime element p in a UFD A, the valuation vp : K × → Z is called the
p-adic valuation on A (or on K).
Examples.
1 For A =Z and K = Q, and p a prime integer, the classical p-adic valuation is defined
by vp ab = vp (a) − vp (b), where vp (a) = n if a = pn m, p - m.

2 Let k be a field and consider the field of rational polynomials K = k(t). The so-
v∞: K × → Z defined by the degree
called “infinite place” of K is a discrete valuation 
function: v∞ (f ) = − deg f for f ∈ k[t] and v∞ fg = deg g − deg f for fg ∈ K × .
For A = k[t−1 ], the prime element t−1 induces the same valuation on Frac(A) = k(t):
vt−1 = v∞ .
3 For k a field, consider the field of rational power series
K = k((t)) = {t−n f | n ∈ N0 , f ∈ k[[t]]}.
Define a function v on K by v(x) = m ∈ Z if x = ∞ i
P
i=m ai t for ai ∈ k and am 6= 0.
Formally let v(0) = ∞ so that v : K → Z is a discrete valuation. As in 2 , the prime
element t in the UFD A[[t]] induces this valuation.

69
4.1. Discrete Valuation Rings Chapter 4. Discrete Valuations and Dedekind Domains

Definition. If v : K × → Z is a discrete valuation, then O(v) = {v ∈ K | v(x) ≥ 0} is called


the valuation ring of v.

Lemma 4.1.3. For any discrete valuation v : K × → Z, O(v) is a subring of K.

Proof. By convention, v(0) = ∞ so 0 ∈ O(v). If x, y ∈ O(v) then by definition v(xy) =


v(x)+v(y) ≥ 0 so xy ∈ O(v). Likewise, v(x+y) ≥ min(v(x), v(y)) ≥ 0 shows that xy ∈ O(v).
Hence O(v) ⊆ K is a subring.

Proposition 4.1.4. An integral domain A is a DVR if there exists a discrete valuation v


on K = Frac(A) such that O(v) = A.

Proof. Suppose v : K × → Z is a discrete valuation with A = O(v). By Lemma 4.1.3, A is a


ring so it is enough to show that A is a DVR. If x ∈ K × such that x−1 6∈ A, then v(x−1 ) < 0.
Note that the definition of a valuation forces v(1) = 0 since v(1) = v(1 · 1) = v(1) + v(1), so
for such an x we have

0 = v(1) = v(xx−1 ) = v(x) + v(x−1 ) < v(x) + 0

and so v(x) > 0. Thus x ∈ A, so A is a DVR.


Examples, continued.

For a prime p and the p-adic valuation vp : Q → Z, the valuation ring O(vp ) =
1 
a
b
: a, b ∈ Z, b 6= 0, p - b coincides with Z(p) , the p-adic integers described in Exam-
ple 1.2.2.

2 For v∞ on k(t), the valuation ring is similar to the p-adic integers:


 
f
O(v∞ ) = : f, g ∈ k[t], g 6= 0, deg g ≥ deg f .
g

3 For the valuation v on k((t)), O(v) = k[[t]].

Proposition 4.1.5. Let v : K × → Z be a discrete valuation with valuation ring O. Then

(1) O× = {x ∈ K | v(x) = 0}.

(2) O is a local ring with unique maximal ideal m = O r O× = {x ∈ K | v(x) > 0}.

(3) O is a DVR.

(4) If π ∈ O satisfies v(π) = 1, then m = (π).

(5) Every proper ideal I ⊂ O is of the form I = (π m ) = {x ∈ K | v(x) ≥ m}, where


m = min(v(x) | x ∈ I).

(6) O is a PID with unique prime elements of the form uπ, where u is a unit.

(7) Spec(O) = {(0), m} and hence dim O = 1.

70
4.1. Discrete Valuation Rings Chapter 4. Discrete Valuations and Dedekind Domains

Proof. Exercise.

Definition. For a valuation ring O = O(v), an element π ∈ O such that m = (π) is the
unique maximal ideal of O is called a uniformizer of O.

Remark. Suppose A is a local noetherian domain.

(1) If dim A = 1 then the fact that A is local and a domain implies Spec(A) = {0, m}, where
m is the maximal ideal of A.

(2) Further, if I ⊂ A is a nonzero proper ideal then Spec(A/I) = {m/I} by the corre-
spondence theorem. Thus dim A/I = 0. By Lemma 3.1.3, A noetherian implies A/I is
noetherian as well. Thus by Theorem 3.5.5, A/I is also artinian.

(3) Now m/I = N (A/I) so by Proposition 3.5.4, (m/I)` = 0 for some ` ∈ N, and hence
m` ⊆ I ⊆ m. This shows that r(I) = m, i.e. that I is m-primary by Lemma 3.4.8.

(4) Finally, since dim A = 1, Theorem 3.5.5 says that A is not artinian, so by Theorem 3.5.6,
m` 6= m`+1 for any ` ∈ N0 .

The next result completely characterizes local noetherian DVRs.

Theorem 4.1.6. Let A be a local noetherian domain with maximal ideal m and residue field
k = A/m, such that dim A = 1. Then the following are equivalent:

(i) A is a DVR.

(ii) A is integrally closed.

(iii) m is a principal ideal.

(iv) dimk m/m2 = 1.

(v) For all nonzero proper ideals I ⊂ A, there exists an n ∈ N such that I = mn .

(vi) There exists an x ∈ A such that for any nonzero proper ideal I ⊂ A, I = (xn ) for
some n ∈ N.

Proof. (i) =⇒ (ii) If A is a DVR then A is a PID by (6) of Proposition 4.1.5, so A is a


UFD. By Proposition 2.1.8, this proves A is integrally closed.
(ii) =⇒ (iii) Choose a nonzero element a ∈ m. By Remark (3), there is some ` ∈ N
with m` ⊆ (a) and m`−1 6⊂ (a). Choose b ∈ m`−1 with b 6∈ (a) and set x = ab ∈ k. Then
x−1 = ab 6∈ A since b 6∈ (a). Since A is integrally closed, this means x−1 is not integral over A.
Suppose x−1 m ⊂ m. Then m is an A[x−1 ]-module, m is faithful since m 6= 0 and A[x−1 ] ⊆ k,
and m is finitely generated as an A-module because A is noetherian. Therefore the conditions
of Lemma 2.1.2(iv) are satisfied, showing x−1 is integral over A, a contradiction. Thus we
must have x−1 m 6⊂ m. Now
b 1 1 1
x−1 m = m = (bm) ⊆ m` ⊆ (a) = A,
a a a a

71
4.1. Discrete Valuation Rings Chapter 4. Discrete Valuations and Dedekind Domains

showing x−1 m is an ideal of A. But x−1 m 6⊂ m so by maximality we must have x−1 m = A,


i.e. m = (x).
(iii) =⇒ (iv) If m = (x) then m/m2 = k(x + m2 ) as a k-vector space, so dimk m/m2 . By
(4) of the Remark, m 6= m2 , so we must have dimk m/m2 = 1.
(iv) =⇒ (v) By (3) of the Remark, there exists an ` ∈ N such that m` ⊆ I ⊆ m.
Without loss of generality, assume ` ≥ 2 and consider the quotient A/m` . Then by (2) of
the Remark, A/m` is local artinian. Thus

dimk (m/m` )/(m2 /m` ) = dimk m/m2 = 1

by the isomorphism theorems. By Proposition 3.5.7, this implies I/m` = (m/m` )n = mn /m`
for some n ≤ `. However, I ⊇ m` and mn ⊇ m` so we must have I = mn .
(v) =⇒ (vi) By Remark (4), m 6= m2 so we can find some element x ∈ m r m2 . By
assumption, there is some n ∈ N such that (x) = mn . Since x 6∈ m2 , n has to equal 1, so
m = (x). Hence for any nonzero ideal I ⊂ A, I = m` = (x` ) for some ` ∈ N.
(vi) =⇒ (i) Suppose m = (xn ) for some x ∈ A and n ∈ N. Since m is maximal, x ∈ m
so we have (x) ⊆ m = (xn ) ⊆ (x) and thus m = (x), that is, m is principal. By Remark (4),
(x` ) 6= (x`+1 ) for any ` ∈ N0 . This means for any nonzero element a ∈ A, there is a unique
` ∈ N0 such that (a) = (x` ) by assumption. Define v : A r {0} → N0 by v(a) = ` where ` is
chosen as described. This map is surjective because x` 7→ ` for any ` ∈ N0 . If a, b ∈ A such
that (a) = (x` ) and (b) = (xm ), then

(ab) = (x` )(xm ) = (x`+m ) =⇒ v(ab) = v(a) + v(b).

Moreover, suppose v(a) = ` and v(b) = m. Then (a) = (x` ) and (b) = (xm ) so there exist
u, w ∈ A× such that a = ux` and b = wxm . Without loss of generality assume ` ≤ m. Then
a + b = ux` + wxm = x` (u + vxm−` ) and u + vxm−` ∈ A, so this shows a + b ∈ (x` ) and
hence (a + b) ⊆ (x` ). If (a + b) = (xr ), we have (xr ) ⊆ (x` ) so in particular r ≥ ` and thus
v(a + b) ≥ ` = v(a).
Now by Lemma 4.1.1, v extends uniquely to a discrete valuation v : K × → Z. It
remains to verify that O(v) = A and we will be finished. By construction A ⊆ O(v) since
v|Ar{0} : A r {0} → N0 . On the other hand, if y ∈ O(v) is nonzero then y = ab for some
a, b ∈ A r {0}. Then 0 ≤ v(y) = v(a) − v(b), that is, v(a) ≥ v(b). If v(a) = ` and v(b) = m
as above, then (a) = (x` ) and (b) = (xm ), and since ` ≥ m, we have (a) ⊆ (b). Hence a ∈ (b)
so y = ab ∈ A. This completes the proof that A = O(v), i.e. A is a valuation ring.

72
4.2. Dedekind Domains Chapter 4. Discrete Valuations and Dedekind Domains

4.2 Dedekind Domains


Definition. An integral domain A is called a Dedekind domain if
(i) A is noetherian.
(ii) dim A = 1, i.e. every prime ideal is maximal.
(iii) A is integrally closed.
Notice that the Krull dimension requirement means that a field is not a Dedekind domain.
Example 4.2.1. Every PID that is not a field is a Dedekind domain.
Dedekind domains are the central object in algebraic number theory, as one can construct
a canonical Dedekind domain OK , called the ring of integers, for any number field K ⊇ Q.
In this way the axioms and properties of Dedekind domains were devised in an attempt to
generalize the algebraic properties of Z.
Proposition 4.2.2. If A is a noetherian domain of Krull dimension 1, the following are
equivalent:
(i) A is integrally closed, and hence a Dedekind domain.
(ii) Ap is a DVR for every nonzero prime ideal p ⊂ A.
(iii) Every primary ideal in A is a prime power.
Proof. (i) =⇒ (ii) It follows from Proposition 2.3.12 that Ap is integrally closed for every
prime ideal p in A. Also, by Lemma 2.3.6, dim Ap = ht(p) = 1 and by Lemma 3.3.1, Ap is
noetherian, so Ap is a DVR by Lemma 3.3.1.
(ii) =⇒ (i) If Ap is a DVR then it is a PID, hence a UFD, hence integrally closed by
Proposition 2.1.8. Since this holds for all nonzero prime ideals p, Proposition 2.3.12 implies
that A is also integrally closed.
(ii) =⇒ (iii) Take a p-primary ideal 0 6= Q ⊂ A so that p = r(Q). Then p is maximal
since dim A = 1. For S = Arp, S −1 Q is a proper ideal in S −1 A = Ap . By Proposition 3.4.18,
S −1 Q is S −1 p-primary in Ap . Now Ap is a DVR, so S −1 Q = (S −1 p)n for some n ∈ N
according to Theorem 4.1.6. Since p is maximal, Lemma 3.4.8 says that pn is p-primary, and
by Lemma 1.3.10, (S −1 p)n = S −1 pn , so Proposition 3.4.18 implies Q = pn .
(iii) =⇒ (ii) We are assuming A is a noetherian domain, so Ap is a local noetherian
domain for all nonzero prime ideals p ∈ Spec(A) = MaxSpec(A) by Lemma 3.3.1. Then
dim A = 1 implies dim Ap = 1. We want to show Ap is a DVR, so by Theorem 4.1.6, it’s
enough to show that all nonzero ideals in Ap are powers of S −1 p, where S = A r p. Take
some nonzero ideal I ⊂ Ap . By Remark (3) preceding Theorem 4.1.6, I is S −1 p-primary so
Proposition 3.4.18 says there exists a primary ideal Q ⊂ A, Q ∩ S = ∅, with S −1 Q = I.
Since Q ∩ S = ∅, Q ⊆ p and thus r(Q) = p since p is the only prime ideal of A containing
Q. By assumption, Q = pn for some n ∈ N, and thus Lemma 1.3.10 gives us
I = S −1 Q = S −1 pn = (S −1 p)n .
Hence I is a power of S −1 p and we conclude that Ap is a DVR.

73
4.2. Dedekind Domains Chapter 4. Discrete Valuations and Dedekind Domains

Corollary 4.2.3. If A is a Dedekind domain, then any nonzero ideal of A is a product of


prime ideals in A.

Proof. Suppose I ⊂ A is a nonzero ideal. Since A is noetherian, I = `i=1 Qi for primary


T
ideals Qi ⊂ A with r(Qi ) = pi prime, by Theorem 3.4.6. By Lemma 3.4.11, we may take
this to be a minimal primary decomposition of I, so pi 6= pj for i 6= j. Since dim A = 1, each
pi is maximal so pi + pj = A for all i 6= j. Hence I = `i=1 Qi = `i=1 Qi by Lemma 3.4.17.
T Q
Now by Proposition 4.2.2(iii), each primary ideal Qi is a powerQ of a prime ideal, namely its
radical r(Qi ) = pi . So there exist n1 , . . . , n` ∈ N such that I = `i=1 pni i .
We will prove in the next section that nonzero ideals in a Dedekind domain factor into
a unique product of prime ideals. This is the most important feature of Dedekind domains,
as it allows us to generalize the unique factorization properties of Z to similar objects in
number theory which turn out to be Dedekind domains. For example, taking the integral
closure of Z in a finite separable extension of Q preserves the Dedekind property:

Theorem 4.2.4. If A is a Dedekind domain with field of fractions K and L/K is a finite
separable field extension, then the integral closure B of A in L is a Dedekind domain.

Proof. First, B is integrally closed by Corollary 2.1.7. Moreover, since B/A is an integral
extension, dim B = dim A = 1 by Theorem 2.3.7. Finally, since L/K isPseparable, Proposi-
tion 2.2.5 says that there is a K-basis {v1 , . . . , vn } of L such that B ⊆ ni=1 Avi . Since A is
noetherian, we get that B is noetherian as well. Hence B is a Dedekind domain.

74
4.3. Fractional, Invertible Ideals Chapter 4. Discrete Valuations and Dedekind Domains

4.3 Fractional, Invertible Ideals


In this section we prove that every nonzero ideal in a Dedekind domain factors uniquely into
a product of prime ideals. In order to prove this, we need to expand the notion of an ideal
to include other submodules of the fraction field. Let A be an integral domain with field of
fractions K. Then A-ideals are also A-submodules of K.
Definition. For two A-submodules I, J ⊆ K, the generalized ideal quotient of I by J
with respect to K is
(I :K J) = {x ∈ K | xJ ⊆ I}.
Lemma 4.3.1. For every pair of submodules I, J ⊆ K, (I :K J) is an A-submodule of K.
Lemma 4.3.2. Let I, J ⊆ K be submodules. If S is any multiplicatively closed subset of A,
then S −1 (I :K J) = (S −1 I :K S −1 J) if J is a finitely generated A-module.
Definition. A nonzero A-submodule I ⊆ K is called a fractional ideal of A if there is
some x ∈ K × such that xI ⊆ A. Equivalently, I is fractional if there is an ideal J ⊂ A such
that I = yJ for some y ∈ K × .
Definition. A nonzero A-submodule I ⊆ K is an invertible ideal of A if there is an
A-submodule J ⊆ K such that IJ = A.
Proposition 4.3.3. Let I ⊆ K be a nonzero A-submodule. Then
(a) If I is finitely generated then I is a fractional ideal.
(b) If A is noetherian and I is fractional, then I is finitely generated.
(c) If I is invertible with IJ = A, then J = (A :K I) and J is unique.
(d) If I is invertible then I is finitely generated and therefore fractional.
Proof. (a) Let I = ni=1 Axi where xi = abii for ai , bi ∈ A r {0}. Then b1 · · · bn I ⊆ A so I is
P
fractional.
(b) If A is noetherian, then any A-ideal J is finitely generated, so if I = yJ for J ⊂ A
and y ∈ K × then I is clearly finitely generated as an A-module.
(c) Suppose IJ = A for some submodule J ⊆ K. Then J ⊆ (A :K I) by definition of the
generalized ideal quotient. On the other hand,
(A :K I) = (A :K I)A = (A :K I)IJ ⊆ AJ = J.
Therefore we have J = (A :K I) and this implies uniqueness since J was arbitrary.
Pnsome submodules J ⊆ A. Since 1 ∈ A = IJ, there are xi ∈ I
(d) Suppose IJ = A for
and yi ∈ J such that 1 = i=1 xi yi . For each x ∈ I, we have
n
X n
X
x=x·1=x xi y i = xi (xyi ).
i=1 i=1

each xyi ∈ IJ = A, we have ni=1 xi (xyi ) ∈


P Pn
SinceP i=1 Axi ⊆ I. Therefore x ∈ I, so
n
I = i=1 Axi , meaning I is finitely generated. Finally, (a) implies that I is a fractional
A-ideal.

75
4.3. Fractional, Invertible Ideals Chapter 4. Discrete Valuations and Dedekind Domains

If I is an invertible ideal, we will write I −1 = (A :K I). When I ⊂ A is an ideal, we will


call it an integral ideal to distinguish from fractional ideals that are not submodules of A.
Take a nonzero prime ideal p ∈ Spec(A) and consider the local ring Ap = S −1 A, where
S = A r p. If I ⊆ K is an A-submodule, we will write Ip = S −1 I.
Proposition 4.3.4. For an A-submodule I ⊆ K, the following are equivalent:
(i) I is an invertible ideal of A.
(ii) I is finitely generated as an A-module and Ip is an invertible Ap -ideal for all prime
ideals p ∈ Spec(A).
(iii) I is a finitely generated A-module and Im is an invertible Am -ideal for all maximal
ideals m ∈ MaxSpec(A).
Proof. (i) =⇒ (ii) By Proposition 4.3.3(d), I is finitely generated as an A-module. If
IJ = A for a submodule J ⊆ K, then by Lemma 1.3.10(a), we have Ap = Ip Jp so in
particular Ip is an invertible ideal of Ap .
(ii) =⇒ (iii) is trivial.
(iii) =⇒ (i)) Set L = I(A :K I). Then L ⊆ A is an ideal. For any maximal ideal m ⊂ A,
set S = A r m. Then

Lm = Im S −1 (A :K I) = Im (S −1 A :K S −1 I) by Lemma 4.3.2
= Im (Am :K Im ) = Im Im−1 = Am .

In particular Lm is not contained in the maximal ideal S −1 m, so either L ∩ S 6= ∅ or L 6⊂ m.


However, if Lm ⊂ Am is an ideal, then L ∩ S = ∅, so we must have L 6⊂ m. Since m is
maximal, it follows that L = A, i.e. I is invertible.
Proposition 4.3.5. For a local domain A which is not a field, A is a DVR if and only if
every fractional ideal of A is invertible.
Proof. (i) =⇒ (ii) Suppose A is a DVR. By Proposition 4.1.5, A is a PID so every fractional
ideal of A is of the form Ax for some x ∈ K × . The inverse of Ax is Ax−1 , so all fractional
ideals of A are also invertible.
(ii) =⇒ (i) Every nonzero (integral) ideal of A is fractional, so by assumption they are
all invertible. Then Proposition 4.3.3(d) implies every ideal of A is finitely generated, so by
Proposition 3.1.2, A is noetherian. Let m be the unique maximal ideal of A. Then there
exists a fractional ideal m ⊆ K such that mm−1 = A. We claim that every ideal of A is
of the form mn for some n ∈ N. Let J ⊂ A be a nonzero integral ideal. Then J ⊆ m so
Jm−1 ⊆ mm−1 = A, which shows that Jm−1 is an integral ideal of A. Moreover, A ( m
implies J ( Jm−1 , for if J = Jm−1 then we would have Jm = JA = J, and thus m = 0 by
Nakayama’s Lemma (1.2.1). Now starting with any ideal J ⊂ A, we get a strictly ascending
chain of ideals
J ( Jm−1 ( Jm−2 (
Since A is noetherian, the chain stabilizes, i.e. Im−n = A for some n ∈ N. Therefore I = mn
for this n. This implies Spec(A) = {0, m} so dim A = 1. Finally, Theorem 4.1.6 shows that
A is a DVR.

76
4.3. Fractional, Invertible Ideals Chapter 4. Discrete Valuations and Dedekind Domains

Our results in the previous two sections combine to give us the following characterization
of Dedekind domains.

Theorem 4.3.6. For an integral domain A which is not a field, the following are equivalent:

(i) A is a Dedekind domain.

(ii) Every fractional ideal of A is invertible.

Proof. (i) =⇒ (ii) Suppose A is Dedekind. Then by Proposition 4.2.2, Ap is a DVR


for every prime ideal p ∈ Spec(A) = MaxSpec(A). Thus Proposition 4.3.5 implies every
fractional ideal Ip is invertible (as an Ap -module) for any such prime ideal p. Finally, since A
is noetherian, every ideal I is finitely generated (3.1.2) and hence Proposition 4.3.4 implies
I is an invertible ideal.
(ii) =⇒ (i) If every fractional ideal of A is invertible, then each is finitely generated
by Proposition 4.3.3(d), so A is noetherian. We want to show Ap is a DVR for all maximal
ideals m ∈ MaxSpec(A). By Proposition 4.3.5, it’s enough to show that every fractional
ideal of Am is invertible. Let I 0 ⊂ Am be a nonzero ideal. Then I 0 = S −1 I = Im for some
nonzero ideal I ⊂ A. By assumption I is invertible, so Im is invertible. It follows that every
fractional ideal of Am is invertible, so we have that ht(m) = dim Am = 1 by Lemma 2.3.6(a)
and this holds for any maximal ideal m of A. Hence dim A = 1 so Proposition 4.2.2 tells us
that A is Dedekind.
Dedekind domains have the important property that every nonzero ideal factors uniquely
into a product of prime ideals.

Theorem 4.3.7. If A is a Dedekind Qdomain then every nonzero ideal I ⊂ A factors uniquely
m
into a product of prime ideals I = i=1 pi , with repetitions allowed.

Proof. By Corollary 4.2.3, I factors as I = m


Q
i=1 pi for nonzero prime ideals piQ⊂ A that
need not be distinct. Thus it remains to check uniqueness. Suppose that I = nj=1 qj for
prime ideals qj ⊂ A. Then q1 · · · qn ⊆ p1 so there exists a 1 ≤ j ≤ n such that qj ⊆ p1 . Since
primes are maximal in a Dedekind domain, qj = p1 . We may relabel the qj so that q1 = p1 .
By Theorem 4.3.6, p−11 exists so we have

p1 p2· · · pm = p1 q2 · · · qn
−1
=⇒ p1 p1 p2 · · · pm = p−1
1 q2 · · · qn
=⇒ p2 · · · pm = q2 · · · qn .

By induction, m = n and after relabelling, qi = pi for all 1 ≤ i ≤ m. Hence prime


factorizations are unique up to reordering.

Proposition 4.3.8. Take nonzero ideals I, J ⊂ A with I = m


Q ei Q m fi
i=1 pi and J = i=1 pi for
distinct primes p1 , . . . , pm ⊂ A and exponents ei , fi ≥ 0. Prove the following:

(a) I ⊆ J if and only if ei ≥ fi for all i.


max(ei ,fi )
(b) I ∩ J = m
Q
i=1 pi .

77
4.3. Fractional, Invertible Ideals Chapter 4. Discrete Valuations and Dedekind Domains

Qm min(ei ,fi ) f
(c) I + J = i=1 pi . In particular, pei i + pj j = A for all ei and fj when i 6= j.
Proof. (a) One direction is clear: if ei ≥ fi for all iQthen pei i ⊆ pfi i for all i, and hence I ⊆ J.
On the other hand, suppose I ⊆ J. Then I ⊆ m fi f1
i=1 pi ⊆ p1 . By Theorem 4.3.6, p1 is
invertible, so we may multiply by p−11 to obtain:

I ⊆ pf11 =⇒ pe11 pe22 · · · pemm ⊆ pf11


=⇒ p−e 1 e1 e2 em −e1 f1
1 p1 p2 · · · pm ⊆ p1 p1

=⇒ pe22 · · · pemm ⊆ pf11 −e1 .


If f1 − e1 > 0 then pf11 −e1 ⊆ p1 . Since p1 is prime, the above implies p1 ⊇ pj for some
2 ≤ j ≤ m. But prime ideals are maximal in a Dedekind domain, so p1 = pj , contradicting
the fact that p1 , . . . , pm are distinct primes. Therefore f1 − e1 ≤ 0. Repeating the proof for
each pi , 2 ≤ i ≤ m shows that ei ≥ fi for all i.
(b) Using unique factorization of ideals, write
n
Y n
Y n
Y n
Y
I= pei i , J= pfi i , I ∩J = pgi i , I +J = phi i
i=1 i=1 i=1 i=1

for distinct prime ideals p1 , . . . , pm , pm+1 , . . . , pn , n ≥ m, and unique exponents ei , fi , gi , hi ≥


0, with ei = fi = 0 for m < i ≤ n. Then by (a), we get the following implications:
I ∩ J ⊆ I =⇒ gi ≥ ei for all 1 ≤ i ≤ n and
I ∩ J ⊆ J =⇒ gi ≥ fi for all 1 ≤ i ≤ n.
So gi ≥ max(ei , fi ) for all 1 ≤ i ≤ n. This shows gi ≥ max(ei , fi ) ≥ ei and gi ≥ max(ei , fi ) ≥
fi for all 1 ≤ i ≤ n, so because I ∩ J is the largest ideal of A which is contained in both I
and J, part (a) implies
n
Y n
Y
max(ei ,fi ) max(ei ,fi )
I ∩J ⊆ pi ⊆ I ∩ J =⇒ I ∩ J = pi .
i=1 i=1

(c) As in part (b), we have:


I ⊆ I + J =⇒ ei ≥ hi for all 1 ≤ i ≤ n and
J ⊆ I + J =⇒ fi ≥ hi for all 1 ≤ i ≤ n.
So min(ei , fi ) ≥ hi for all 1 ≤ i ≤ n, meaning ei ≥ min(ei , fi ) ≥ hi and fi ≥ min(ei , fi ) ≥ hi .
Since I + J is the smallest ideal of A containing both I and J, part (a) means that
n
Y n
Y
min(ei ,fi ) min(ei ,fi )
I⊆ pi ⊆I +J and J ⊆ pi ⊆I +J
i=1 i=1
n
Y min(ei ,fi )
imply I + J ⊆ pi ⊆I +J
i=1
n
Y min(ei ,fi )
so I + J = pi .
i=1

78
4.3. Fractional, Invertible Ideals Chapter 4. Discrete Valuations and Dedekind Domains

Theorem 4.3.9 (Chinese Remainder Theorem). If I = m ei


Q
i=1 pi for distinct prime ideals
pi ⊂ A and exponents ei ≥ 0, then there is a canonical isomorphism
m
Y
A/I −→ A/pei i .
i=1

Remark. Let A be a Dedekind domain and take a nonzero prime (maximal) ideal Q p ∈
Spec(A), so that Ap is a DVR. For 0 6= x ∈ A, we have a prime factorization (x) = m ei
i=1 pi
with p1 = p and ei ≥ 0. The ei are determined by (x), so define a function

vp : A r {0} −→ N0
x 7−→ e1 .

Then vp extends to a valuation on K = Frac(A) such that O(vp ) = Ap . As this generalizes


the construction of the p-adic valuation on Q, the function vp is called the p-adic valuation
on K.

Proposition 4.3.10. If A is a Dedekind domain with Spec(A) finite, then A is a PID.

Proof. Let p1 , . . . , pn be the prime ideals of A. By the Chinese remainder theorem,

A/p21 p2 · · · pn ∼
= A/p21 × A/p2 × · · · × A/pn .

In particular, for any fixed x1 ∈ p1 r p21 , there exists y ∈ A that satisfies y + p21 = x1 + p21
and y + pi = 1 + pi for all 2 ≤ i ≤ n. Thus y − x1 ∈ p21 ⊂ p1 so it follows that y ∈ p1 . We
have (y) ⊆ p1 but (y) 6⊂ p21 and (y) 6⊂ pi for any 2 ≤ i ≤ n. Since A is a Dedekind domain,
(y) has a prime factorization
Y n
(y) = pei i for ei ≥ 0.
i=1

By Proposition 4.3.8(a) and the above work, we conclude that 1 ≤ e1 < 2 and 0 ≤ ei < 1
for each 2 ≤ i ≤ n. In other words, e1 = 1 and e2 = . . . = en = 0, so (y) = p1 . Repeating
the proof for each pi , 1 ≤ i ≤ n shows that every prime ideal of A is principal. Suppose
pi = (yi ). Then any ideal I ⊂ A has a prime factorization
n
Y n
Y
I= pai i = (yiai ) = (y1a1 · · · ynan ).
i=1 i=1

Therefore I is principal.

Corollary 4.3.11. Let I be a nonzero ideal in a Dedekind domain A. Then for any nonzero
a ∈ I, there is some b ∈ I such that I = (a, b).

Proof. We first prove A/I is a PID. By the correspondence theorem, every prime ideal of A/I
along the quotient map A → A/I to a unique prime ideal of A which contains I.
pulls back Q
a
Write I = rj=1 pj j for prime ideals pj ⊂ A and integers aj ≥ 1. Then Proposition 3.4.10(a)
and the fact that dim A = 1 imply that the only prime ideals containing I are p1 , . . . , pr . It

79
4.3. Fractional, Invertible Ideals Chapter 4. Discrete Valuations and Dedekind Domains

follows that A/I has finitely many prime ideals. Therefore by Proposition 4.3.10, A/I is a
PID.
Now take 0 6= a ∈ I. Then by the first paragraph, A/(a) is principal and therefore the
ideal I/(a) can be generated by a single element b+(a), b ∈ I. (If b ∈ (a), this means I = (a)
so I was principal to begin with.) On one hand, it’s immediate that (a, b) ⊆ I. On the other
hand, every y ∈ I r(a) has reduction y+(a) ⊆ I/(a) in A/(a), so y+(a) = r(b+(a)) = rb+(a)
for some r ∈ A, and thus y − rb ∈ (a), so y = rb + sa ∈ (a, b). This shows I = (a, b).

80
4.4. The Class Group Chapter 4. Discrete Valuations and Dedekind Domains

4.4 The Class Group


Definition. For a Dedekind domain A with field of fractions K, the ideal group of A is
the free abelian group

IA = {I ⊂ K | I is a fractional ideal of A} = {I ⊆ K | I is an invertible ideal}.

The subgroup PA = {Ax | x ∈ K × } ≤ IA is called the subgroup of principal fractional


ideals of A. The class group of A is then defined as the quotient group

CA = IA /PA .

The order of this group, denoted h(A), is called the class number of A.

In general, class numbers can be arbitrarily large, but a highly important fact for number
fields is that their rings of integers have finite class number:

Theorem 4.4.1. For an algebraic extension K/Q, with ring of integers OK , the class number
hK := h(OK ) is finite.

Lemma 4.4.2. For a Dedekind domain A, the following are equivalent:

(i) h(A) = 1.

(ii) A is a PID.

(iii) A is a UFD.

Proof. (i) ⇐⇒ (ii) Suppose h(A) = 1. Then for any ideal I ⊂ A, I is a fractional ideal
(e.g. by Proposition 4.3.3), but since the class group of A is trivial, I must be principal.
Hence A is a PID. Conversely, suppose every ideal I ⊂ A is principal.
Qr Any fractional ideal
ei
J ∈ IA may be written as a formal product of prime ideals J = i=1 pi with ei ∈ Z for each
i. Then each prime ideal is principal: pi = (xi ) for an element xi ∈ A, 1 ≤ i ≤ r, so we have
r
Y r
Y
J= pei i = (xi )ei = Axe11 · · · xerr .
i=1 i=1

Therefore J is a principal fractional ideal, and it follows that PA = IA , that is, h(A) =
|IA /PA | = 1.
(ii) ⇐⇒ (iii) It is well-known that every PID is a UFD. For the converse, suppose p ⊂ A
is a (proper) nonzero prime ideal. Choose 0 6= x ∈ p (which is not a unit since p 6= A) and
write x = pe11 · · · perr for distinct irreducible elements pi ∈ A and exponents ei ∈ N. Since p
is a prime ideal, pi ∈ p for some 1 ≤ i ≤ r. Then (pi ) ⊆ p. Since pi is irreducible, (pi ) is a
prime ideal, so the fact that dim A = 1 implies (pi ) = p. Thus every nonzero prime ideal of
A is principal. This implies the result by unique factorization of ideals.

81
4.4. The Class Group Chapter 4. Discrete Valuations and Dedekind Domains


Example√ 4.4.3. Let K = Q( −5) be the quadratic number field with ring of integers
OK = Z[ −5]. Then OK is a Dedekind domain, but it is not a UFD, since for example
√ √
6 = 2 · 3 = (1 + −5)(1 − −5)

and each of the elements in the two factorizations is irreducible but not prime. However,
since OK is Dedekind, the ideal (6) factors uniquely into a product of prime ideals:

(6) = p21 p2 p3 ,
√ √
where p1 = (2, 1 + −5) = (2, 1 − −5)

p2 = (3, 1 + −5)

p3 = (3, 1 − −5).
√ √
(2) = p21 , (3) √
In particular, √ = p2 p3 , (1 +
√ −5) = p1 p2 and
√ (1 − −5) = p1 p3 . We have
(2) ( p1 ( Z[ −5] and [Z[ −5] : 2Z[ −5]] = 4, so [Z[ −5] : p1 ] = 2 and hence p1 is a
maximal ideal. Similar arguments show that p2 and p3 are also √ maximal (thus prime) ideals.
Moreover, notice that each of p1 , p2 , p3 is not principal, so h(Z[ −5]) > 1 (in fact, the class
number is 2 in this case).

82
4.5. Dedekind Extensions Chapter 4. Discrete Valuations and Dedekind Domains

4.5 Dedekind Extensions


Theorem 4.2.4 says that extending a Dedekind domain in a finite separable extension of
its fraction field (i.e. taking its integral closure) again yields a Dedekind domain. The
critical step in the proof of this result is to prove the integral closure is noetherian, which
requires Proposition 2.2.5. The issue in generalizing Theorem 4.2.4 is that Proposition 2.2.5
does not hold if L/K is not separable – in fact, we will demonstrate in a moment several
counterexamples to 2.2.5 in the inseparable case, even when A and B are DVRs.
Recall the following definition:

Definition. A field extension L/K is purely inseparable if for each α ∈ L, the minimal
polynomial of α over K has a single distinct root.

Lemma 4.5.1. For a finite extension L/K, the following are equivalent:

(i) L/K is purely inseparable.

(ii) If α ∈ L is separable over K then α ∈ K.


n
(iii) There is some n ∈ N such that for all α ∈ L, αp ∈ K.

Lemma 4.5.2. Let L/K be any algebraic extension of fields. Then there is a unique subfield
K ⊆ K sep ⊆ L such that K sep /K is separable and L/K sep is purely inseparable.

We now prove the general version of 4.2.4.

Theorem 4.5.3. Let A be a Dedekind domain with field of fractions K, L/K a finite ex-
tension and B the integral closure of A in L. Then B is a Dedekind domain.

Proof. It remains to prove the theorem when L/K is inseparable. Assume char K = p > 0.
By Lemma 4.5.2, there is a unique subextension K ⊂ L0 ⊂ L such that L0 /K is separable
and L/L0 is purely inseparable. Let B 0 denote the integral closure of A in L0 . Since L0 /K is
separable, Theorem 4.2.4 shows B 0 is Dedekind. Thus it suffices to consider the case when
K = L0 and A = B 0 .
After this reduction, L/K is purely inseparable, so by Lemma 4.5.1, there is some n ∈ N
n
so that for every α ∈ L, αp ∈ K. Set q = pn . Let M be the splitting field of all polynomials
of the form tq − α, with α ∈ K. Alternatively,

M = {α1/q ∈ K | α ∈ K}

where K is an algebraic closure of K. Let C be the integral closure of A in M .


purely
insep.
K L M

A B C

83
4.5. Dedekind Extensions Chapter 4. Discrete Valuations and Dedekind Domains

n
Define the Frobenius map ϕ : M → K by ϕ(x) = xq = xp . Notice that ϕ is well-defined
by Lemma 2; injective since char K = p; and surjective by the construction of M . It is
clear that ϕ is a homomorphism, so it is an isomorphism (of rings) M → K. Moreover, we
claim A = ϕ(C). Indeed, if y ∈ A then any root of tq − y lies in C and maps to y via ϕ.
Conversely, if x ∈ C then xq − α = 0 for some α ∈ K, i.e. xq ∈ K, but xq ∈ C as well, so
ϕ(x) = xq ∈ C ∩ K = A. We have thus established that A ∼ = C as rings, so C is in particular
a Dedekind domain.
By Theorem 4.3.6, B is a Dedekind domain if and only if every ideal of B is invertible.
For an ideal I ⊂ B, letP J = IC be the corresponding ideal of C. Then J is invertible
m −1
since C is Dedekind, Pm soq q i=1 ai bi = 1q for some elements ai ∈ I, bi ∈ J . Since we are in
q−1 q
characteristic
Pm p, a b
i=1 i i = 1 and b i ∈ K. For each 1 ≤ i ≤ m, set c i = a i b i so that
q−1 −q q −q q −q
i=1 ai ci = 1 and ci ∈ I J ⊆ L. Now ci I ⊆ I J ⊆ J J = C but the previous
statementPshows that ci I ⊆ L. So ci I ⊆ C ∩ L = B, and thus ci ∈ (B : I). In summary,
we have m i=1 ai ci = 1 for elements ai ∈ I and ci ∈ (B : I), so it follows that I(B : I) = B.
Hence every ideal of B is invertible, so B is a Dedekind domain.
In fact, Theorem 4.5.3 can be obtained as a corollary to the much more powerful Krull-
Akizuki Theorem, which we don’t quite have the tools to prove yet.

Theorem (Krull-Akizuki). Let A be a noetherian integral domain and suppose dim A =


1, K = Frac(A), L/K is a finite algebraic extension and B ⊂ L is any subring containing
A. Then B is a noetherian domain with dim B ≤ 1.

For the remainder of the section, we discuss counterexamples to Proposition 2.2.5 and
discuss a generalization of this result for affine algebras.

Example 4.5.4. (Hochster) Consider a perfect field k of characteristic p > 0 (e.g. k = Fp


if youSlike) and let K = k(tpn | n ∈ N). Also, for each n ∈ N, set SKn = k(t1 , . . . , tn ) and
L = n=1 Kn = k(tn | n ∈ N) so that L = K. Define a ring A = ∞
∞ p
n=1 Kn [[x]] where x is
an indeterminate. It is a fact that A is a local noetherian domain with maximal ideal (x),
and any ideal of A is of the form I = (xm ) for some m ∈ N. In particular, Theorem 4.1.6
shows that A is a DVR. Observe that K[[x]] ⊆ A ⊆ L[[x]]. Set u = ∞ i
P
i=1 i x ∈ L[[x]] so
t
that up ∈ A. If F = Frac(A), the extension F (u)/F is purely inseparable, but the integral
closure of A in F (u) is not a finitely generated A-module.

Example 4.5.5. (Kaplansky) Let k be a field of characteristic 2 and set C = k[[t]]. Take
some u ∈ C r k and let K = k(t, u2 ) and L = k(t, u). If u is chosen so that u and t are
algebraically independent over k, for example, then we will have [L : K] = 2. Consider the
rings A = C ∩ K, B = C ∩ L and the field of fractions M = Frac(C) = k((t)).

k K L M

A B C

84
4.5. Dedekind Extensions Chapter 4. Discrete Valuations and Dedekind Domains

Since C is a DVR (Example 3 in Section 4.1), A and B are also DVRs with Frac(A) = K
and Frac(B) = L. In particular, B is the integral closure of A in L – if B 0 is this integral
closure, then B ⊆ B 0 because B is integral over A, but on the other hand B is integrally
closed (using the fact that char k = 2), so it must be that B = B 0 . Suppose B were a finitely
generated A-module. Then we would have B ⊆ Aα1 + . . . + Aαr + Aβ1 u + . . . + Aβs u for
some αi , βj ∈ K. By clearing denominators, there is some m ∈ N such that tm αi ∈ A and
tm βj ∈ A for all 1 ≤ i ≤ r, 1 ≤ j ≤ s. So tm B ⊆ A[u]. Write u = ∞ `
P
`=0 a` t with a` ∈ k. Set


X
m −m−1 −m−1
v = (u − a0 − a1 t − . . . − am t )t =t a` t` .
`=m+1

By the first description, v ∈ L and by the second, v ∈ C, so v ∈ C ∩L = B ⊆ A[u]. However,


the second expression for v shows the coefficient of u in the expansion of tm v must be t−1 ,
contradicting the fact that v ∈ A[u]. Thus B cannot be finitely generated over A.

To close, we generalize Proposition 2.2.5 to the context of affine algebras.

Theorem 4.5.6 (Noether). Let k be a field and suppose A is a finitely generated k-algebra
that is also an integral domain with field of fractions K. If L/K is any finite extension and
B is the integral closure of A in L, then B is a finitely generated A-module.

Proof. (Eisenbud) By Noether normalization (2.4.1), there exist y1 , . . . , yr ∈ A such that


A is integral over the polynomial ring k[y1 , . . . , yr ]. Then the result will follow if we prove
B/k[y1 , . . . , yr ] is a finite extension of rings, so we may replace A with k[y1 , . . . , yr ], and K
with k(y1 , . . . , yr ), to begin with. Moreover, if L00 is the normal closure of L over K and B 00
is the integral closure of A in L00 , then demonstrating B 00 /A is finite also implies the result.
So we make the reduction L = L00 , so that L/K is a normal extension. Set G = Gal(L/K)
and let L0 = LG be the subfield fixed by G. It is a general fact from Galois theory that L/L0
is Galois and L0 /K is a purely inseparable extension.
purely
insep. Galois
K L0 L

A B0 B

Since L/L0 is finite and separable, Proposition 2.2.5 gives us that B is a finitely generated
B 0 -module, so it’s enough to prove B 0 is a finitely generated A-module to obtain the result.
If L0 = K, we’re done so assume L0 6= K. This a nontrivial purely inseparable extension, so
char K = p > 0 and by Lemma 4.5.1, there is some n ∈ N with q = pn and αq ∈ K for every
α ∈ L0 . In particular, if L0 = K(z1 , . . . , zn ) for z1 , . . . , zn ∈ L0 , then for each 1 ≤ i ≤ n,
ziq ∈ K, so there exist fi , gi ∈ A such that ziq = fgii . Since A = k[y1 , . . . , yr ], we have
X X
fi = cJ,i y1j1 · · · yrjr and gi = dJ,i y1j1 · · · yrjr
J=(j1 ,...,jr )∈Nr0 J=(j1 ,...,jr )∈Nr0

85
4.5. Dedekind Extensions Chapter 4. Discrete Valuations and Dedekind Domains

for cJ,i , dJ,i ∈ k, finitely many of which are nonzero. Let S be the finite set of all nonzero
cJ,i and dJ,i , where J ranges over all tuples (j1 , . . . , jr ) ∈ Nr0 and 1 ≤ i ≤ n. Then the field
k 0 := k(s1/q | s ∈ S) is a finite extension of k, and consequently L e := k 0 (y11/q , . . . , yr1/q ) is
a finite extension of L0 . If B e is the integral closure of A in L, e then Be contains B 0 so it’s
0 0
enough to prove the result when L = L e and B = B.e
0 1/q 1/q 0
Now consider C := k [y1 , . . . , yr ] ⊆ L . Then C is integrally closed (e.g. it’s a UFD)
1/q 1/q 1/q
and Frac(C) = L0 , so B 0 ⊆ C. But each yj is integral over A, so C = k 0 [y1 , . . . , yr ] ⊆ B 0
since B 0 is the integral closure of A in L0 . Hence B 0 = C, which is finitely generated over A
by definition. This finishes the proof.

86
Chapter 5

Completion and Filtration

87
5.1. Topological Abelian Groups and Completion Chapter 5. Completion and Filtration

5.1 Topological Abelian Groups and Completion


Definition. A topological abelian group is a topological space G with an abelian group
structure such that the addition and inversion maps
α : G × G −→ G η : G −→ G
(x, y) 7−→ x + y x 7−→ −x
are continuous, where G × G is endowed with the product topology.
Lemma 5.1.1. The set {0} is closed in G if and only if G is Hausdorff as a topological
space.
Proof. Since addition is continuous, {0} is closed in G if and only if the diagonal ∆ =
{(x, x) | x ∈ G} is closed in G × G. It is well known that the latter condition is equivalent
to G being Hausdorff.
Lemma 5.1.2. For a fixed a ∈ G, the translation map ta : G → G, x 7→ x + a, is continuous
with inverse t−a , and hence a homeomorphism.
Lemma 5.1.3. For any subgroup N ≤ G of a topological abelian group G, N is a topological
abelian group with the subspace topology and G/N is a topological abelian group with the
quotient topology.
If U = U (0) is a neighborhood of 0, then U + a = ta (U ) is also open, and thus a
neighborhood of a. Since ta is a homeomorphism, every neighborhood of any point of G is
of the form U + a for some U = U (0). Thus the topology of G is uniquely determined by
the collection {U (0)} of neighborhoods of the identity.
T
Lemma 5.1.4. Let H = U (0)30 U (0) be the intersection of all neighborhoods of 0. Then
(i) H is a subgroup of G.
(ii) H = {0}.
(iii) G/H is Hausdorff.
(iv) G is Hausdorff if and only if H = {0}.
Proof. (i) Take x, y ∈ H. Then x, y ∈ U (0) for every neighborhood U (0) of the identity. In
particular, 0 ∈ −x + U (0) and this is an open set since t−x is a homeomorphism. Hence
H ⊆ −x + U (0) so y ∈ −x + U (0). Thus x + y ∈ U (0). Since U (0) was arbitrary, we have
that x + y ∈ H. This argument also shows that −x ∈ U (0) for any neighborhood U (0), so
−x ∈ H. Hence H is a subgroup.
(ii) For any x ∈ G, x ∈ H ⇐⇒ −x ∈ H ⇐⇒ −x ∈ U (0) for all neighborhoods
U (0) ⇐⇒ 0 ∈ x + U (0) for all U (0) ⇐⇒ x is a limit point of {0}. Therefore H = {0}.
(iii) For a fixed a ∈ G, a + H = ta (H) = ta ({0}) is closed in G/H since ta is continuous.
Thus cosets, aka the ‘points’ in G/H, are closed, so G/H is Hausdorff.
(iv) If H = {0} then (iii) implies G is Hausdorff. On the other hand, for all x 6= 0, there
is a neighborhood Ux of x such that 0 6∈ Ux . Thus x 6∈ H, which implies H = {0}.

88
5.1. Topological Abelian Groups and Completion Chapter 5. Completion and Filtration

Assume 0 ∈ G has a countable basis of subgroups U (0) ⊆ G.


Definition. A Cauchy sequence in G is a sequence of elements (xn )∞ n=0 , where xn ∈ G
for all n ∈ N0 , such that for all neighborhoods U (0), there is some N = N (U (0)) ∈ N such
that xn − xm ∈ U (0) for all n, m ≥ N .
Definition. We say two Cauchy sequences (xn ) and (yn ) are equivalent, written (xn ) ∼
(yn ), if for any neighborhood U (0), there is some N = N (U (0)) ∈ N such that xn −yn ∈ U (0)
for all n ≥ N .
Lemma 5.1.5. ∼ is an equivalence relation on the set of Cauchy sequences in G.
Proof. Let (xn ), (yn ) and (zn ) be Cauchy sequences in G. Then for any U (0), xn − xn =
0 ∈ U (0) for every n ∈ N, so (xn ) ∼ (xn ). If (xn ) ∼ (yn ) then there is an N ∈ N
such that for all n ≥ N , xn − yn ∈ U (0). Since U (0) is assumed to be a subgroup of G,
yn − xn = −(xn − yn ) ∈ U (0) as well, so the same N shows that (yn ) ∼ (xn ). Finally, if
(xn ) ∼ (yn ) and (yn ∼ zn ) and U (0) is a neighborhood of 0, there are N1 , N2 ∈ N such
that for all n ≥ N1 , xn − yn ∈ U (0) and for all n ≥ N2 , yn − zn ∈ U (0). Thus for all
n ≥ max{N1 , N2 }, we have xn − zn = xn − yn + yn − zn ∈ U (0) since U (0) is a subgroup.
Therefore (xn ) ∼ (zn ), so ∼ is an equivalence relation.

Definition. The completion of a topological group G is the quotient space G


b of all equiva-
lence classes under ∼ of Cauchy sequences in G. For a Cauchy sequence (xn ), we let [(xn )]
denote its equivalence class in G.
b

Proposition 5.1.6. For any topological group G, G b is a topological group via pointwise
addition and inversion: [(xn )] + [(yn )] = [(xn + yn )] and −[(xn )] = [(−xn )].
Proof. Consider Cauchy sequences (xn ) and (yn ) in G. We first prove pointwise addition
and inversion of Cauchy sequences are invariant on equivalences classes. If (x0n ) ∼ (xn ) and
(yn0 ) ∼ (yn ) then for all U (0), there are numbers N1 , N2 ∈ N such that x0n − xn ∈ U (0) for
all n ≥ N1 and yn0 − yn ∈ U (0) for all n ≥ N2 . Taking N = max{N1 , N2 }, we have that for
all n ≥ N , (x0n + yn0 ) − (xn + yn ) = (x0n − xn ) + (yn0 − yn ) ∈ U (0) since U (0) is a subgroup.
Thus (x0n + yn0 ) ∼ (xn + yn ), so pointwise addition is well-defined. Similarly, (−x0n ) ∼ (−xn )
so inversion is well-defined.
To prove that G b is a group under these actions, consider two Cauchy sequences (xn ) and
(yn ) in G. Then for a neighborhood U (0), there are N1 , N2 ∈ N such that xm − xn ∈ U (0)
for all m, n ≥ N1 and ym − yn ∈ U (0) for all m, n ≥ N2 . Then for all n ≥ N = max{N1 , N2 },
(xm + ym ) − (xn − yn ) = (xm − xn ) + (ym − yn ) ∈ U (0) since U (0) is a subgroup. Hence
(xn + yn ) is a Cauchy sequence. The zero sequence (0) represents the identity class in G b and
it’s easy to check that [(xn )] + [(−xn )] = [(0)] for all Cauchy sequences (xn ), so we have that
Gb is an abelian group.
Finally, for each neighborhood U (0) in G, define a subgroup U b (0) in G
b by

b (0) = {[(xn )] : there exists an N ∈ N such that xn ∈ U (0) for all n ≥ N }.


U

Then {U
b (0) : U (0) is a neighborhood of 0 in G} forms a basis for a topology on G
b which is
compatible with the group structure on G.b Hence G b is a topological group.

89
5.1. Topological Abelian Groups and Completion Chapter 5. Completion and Filtration

Proposition 5.1.7. There is a homomorphism ϕ : G → G b which is injective precisely when


G is Hausdorff.
T
Proof. Define ϕ(x) = [(x, x, x, . . .)] = [(x)]. Then ker ϕ = H where H = U (0) U (0) so ϕ is
injective if and only if H = 0, which by Lemma 5.1.4 is equivalent to G being Hausdorff.

Definition. A homomorphism of topological abelian groups is a continuous homo-


morphism f : G → H. Such a map f is an isomorphism of topological abelian groups
if f is a homeomorphism and an isomorphism, that is, f is a continuous homomorphism
with a continuous inverse f −1 : H → G which is also a homomorphism.

Lemma 5.1.8. Given a homomorphism of topological abelian groups f : G → H, if (xn ) is


a Cauchy sequence in G then (f (xn )) is a Cauchy sequence in H.

Proof. Suppose V (0) is a neighborhood of 0 ∈ H. Since f is continuous, U (0) := f −1 (V (0))


is open in G, and since f is a homomorphism, f (0) = 0 and hence U (0) is a neighborhood
of 0 ∈ G. Thus there is some N ∈ N such that for all m, n ≥ N , xm − xn ∈ U (0). Applying
f and using the fact that it is a homomorphism, we get f (xm ) − f (xn ) = f (xm − fn ) ∈
f (U (0)) = V (0) for all m, n ≥ N . Therefore (f (xn )) is Cauchy.

Corollary 5.1.9. Every homomorphism of topological abelian groups f : G → H induces a


homomorphism fˆ : G
b→H b such that fˆ[(xn )] = [f (xn )].

In particular, this says that the completion G


b a solution to a universal mapping property
for G. In addition, the completion operation (·) c is a covariant functor on the category
TopAbGps of topological abelian groups:
f g
Proposition 5.1.10. For every sequence G → − H →
− K of homomorphisms of topological
\
abelian groups, we have (g ◦ f ) = ĝ ◦ fˆ.

90
5.2. Inverse Limits Chapter 5. Completion and Filtration

5.2 Inverse Limits


In this section we formulate an alternative construction of the completion G
b from Section 5.1
using inverse limits.

Definition. An inverse system is a family of groups (Am )m∈I , for some directed index set
I, together with homomorphisms θm` : Am → A` for every m ≥ ` such that when m ≥ ` ≥ k,
the following diagram commutes:

θm`
Am A`

θmk θ`k

Ak

Definition. Given an inverse system Q (Am , θm` ) over a directed set I, we define a coherent
sequence to be a sequence (ym ) ∈ m∈I Am such that θm` (ym ) = y` for all m ≥ `. The set
of all coherent sequences of (Am , θm` ) is called the inverse limit of the inverse system:
( )
Y
lim Am = (ym ) ∈ Am : θm` (ym ) = y` for all m ≥ ` .
←−
m∈I

Now let G be a topological group possessing a countable basis of open subgroups Gn =


Gn (0) such that G = G0 ⊇ G1 ⊇ G2 ⊇ · · ·

Lemma 5.2.1. Each Gn is both open and closed.

Proof. For all a ∈ G, a + Gn = ta (Gn ) is open since translation is a homeomorphism.


S Since
the a + Gn partition the complement of Gn in G, we have G r Gn = a6∈Gn (a + Gn ) which
is a union of open sets and therefore open. Hence Gn is closed.
For each n ≥ 1, Gn ⊆ Gn−1 so we have projections θn : G/Gn → G/Gn−1 .

Lemma 5.2.2. The system (G/Gn , θn )n∈N is an inverse system.

Lemma 5.2.3. lim G/Gn is a topological group.


←−
Q
Proof. Consider n∈N G/Gn with the product topology. Then lim G/Gn has the subspace
Q Q ←−
topology from n≥0 G/Gn . Indeed, if pm : n∈N G/Gn → G/Gm is the mth projection map,
then the sets S(x, m) := p−1
m (x + Gm ) = {(yn ) | ym = x + Gm } over all m form a subbasis
Q e m) := S(x, m) ∩ lim G/Gn form a subbasis
for the topology on n∈N G/Gn , so the sets S(x,
←−
for the (subspace) topology on lim G/Gn . In fact, S(0,
e m) form a basis of this topology
←−
because the collection of S(0,
e m) are closed under intersections to begin with. The axioms
of a topological abelian group are then easily checked for lim G/Gn .
←−

91
5.2. Inverse Limits Chapter 5. Completion and Filtration

When (G/Gn , θn )n∈N is the inverse system of quotients in G, the inverse limit coincides
with the completion of G from Section 5.1.
Proposition 5.2.4. For a topological abelian group G with inverse system (G/Gn , θn )n∈N ,
b∼
there is a canonical isomorphism of topological abelian groups G = lim G/Gn .
←−

Proof. For each n ∈ N, define a map


b −→ G/Gn
ζn : G
[(xk )] 7−→ xkn + Gn

where kn is the smallest N (or any N ) such that xk + Gn = xN + Gn for all k ≥ N . Such a
choice is possible because (xk ) is Cauchy. Clearly the ζn are homomorphisms, and one can
verify that they are well-defined on equivalence classes of Cauchy sequences. Now define a
map into the inverse limit by
b −→ lim G/Gn
ζ:G
←−
[(xk )] 7−→ (ζn ([(xk )]))n∈N .

The image of ζ consists of coherent sequences: for all n ∈ N, we have

θn+1 (xkn +1 + Gn+1 ) = xkn +1 + Gn = xkn + Gn

since kn+1 ≥ kn , and therefore θn+1 (ζn+1 [(xk )]) = ζn [(xk )]. Now given (yn ) ∈ lim G/Gn ,
←−
construct a Cauchy sequence mapping to (yn ) by taking a coset representative xn for each
n ∈ N such that yn = xn + Gn . Since (yn ) is a coherent sequence, for each n ∈ N we have

θn+1 (yn+1 ) = yn =⇒ θn+1 (xn+1 + Gn+1 ) = xn+1 + Gn = xn + Gn .

So xn+1 − xn ∈ Gn and hence (xn ) is a Cauchy sequence. Clearly ζ([(xn )]) = (yn ) so ζ is
surjective.
To see that ζ is continuous, suppose [(xk )] ∈ G.
b Then using the notation from the proof
of Lemma 5.2.3, for each m ∈ N,

[(xk )] ∈ ζ −1 (S(0,
e m)) ⇐⇒ xkm + Gm = 0 + Gm in G/Gm ⇐⇒ [(xk )] ∈ G
bm .

Since G
bm is open for all m, we have that ζ is continuous.
Finally, to show ζ is an isomorphism, we define an inverse function by

η : lim G/Gn −→ G
b
←−
xn + Gn 7−→ [(xn )].

We showed above that this map is well-defined and a homomorphism of topological abelian
groups, and clearly ζ ◦ η = idlim G/G by surjectivity of ζ. Finally, for any [(xk )] ∈ G,
b we
n
←−
have η ◦ ζ([(xk )]) = η(xkn + Gn ) = [(xkn )] = [(xk )] since subsequences of Cauchy sequences
are equivalent to their parent Cauchy sequence. This proves that ζ is an isomorphism of
topological abelian groups.

92
5.2. Inverse Limits Chapter 5. Completion and Filtration

Definition. When (Gn )n∈N is the system of all subgroups of finite index in G, the inverse
limit G
b = lim G/Gn is called the profinite completion of G.
←−

Definition. A surjective inverse system is one in which every map θm` is surjective.
Definition. Let (Am , αm` ) and (Bm , βm` ) be two inverse systems indexed over the same set
I. A morphism of inverse systems is a set of maps (ϕ• ) such that ϕm : Am → Bm is a
homomorphism for each m ∈ I that is compatible with the αm` and βm` , i.e. the following
diagrams commute for all m ≥ ` ∈ I:
ϕm
Am Bm
αm` βm`
A` B`
ϕ`

Definition. Let (Am , αm` ), (Bm , βm` ) and (Cm , γm` ) be inverse systems with the same index
set I. A short exact sequence of inverse systems is a sequence of maps of inverse
(im ) (pm ) m i pm
systems 0 → (Am ) −−→ (Bm ) −−→ (Cm ) → 0 such that 0 → Am −→ Bm −→ Cm → 0 is an
exact sequence for all m ∈ I.
Proposition 5.2.5. For any short exact sequence of inverse systems 0 → (Am ) → (Bm ) →
(Cm ) → 0, there is an exact sequence of topological abelian groups

0 → lim Am → lim Bm → lim Cm .


←− ←− ←−

In other words, inverse limit is a left exact functor from the category of inverse systems to
the category of topological abelian groups. Furthermore, if (Am ) is a surjective inverse system
then 0 may be added on the right, i.e. lim(−) is an exact functor.
←−

Proof. Let A = Am and define a map dA : A → A by (am ) 7→ (am − θm+1,m (am+1 )). Define
Q
B, dB , C and dC similarly for (Bm ) and (Cm ). Then ker dA = lim Am , ker dB = lim Bm and
←− ←−
kerdC = lim Cm . We have a commutative diagram with exact rows:
←−

0 A B C 0

dA dB dC

0 A B C 0

so by the Snake Lemma, we get an exact sequence

0 → ker dA → ker dB → ker dC → coker dA → coker dB → coker dC → 0.

The first three terms of this are 0 → lim Am → lim Bm → lim Cm so we have our exact
←− ←− ←−
sequence. For the final statement, suppose (Am , αm` ) is a surjective system. We need to prove
coker dA = 0, i.e. for each (am ) ∈ A, we want to find (xm ) ∈ A such that dA (xm ) = (am ).

93
5.2. Inverse Limits Chapter 5. Completion and Filtration

Let x0 = 0. Then a0 = x0 − θ1 (x1 ) so θ1 (x1 ) = −a0 and since θ1 is surjective, this allows us
to choose some x1 ∈ A1 so that this holds. Proceed to inductively define (xm ). This shows
dA is surjective, so we are done.
p
Corollary 5.2.6. Let 0 → G0 → G → − G00 → 0 be a short exact sequence of topological
abelian groups having systems of open subgroups {Gn }, {G0n } and {G00n }, where G0n = G0 ∩ Gn
and G00n = p(Gn ) for each n ∈ N. Then the sequence of completions

b0 → G
0→G b00 → 0
b→G

is exact.

Proof. This is done by considering the short exact sequence of surjective inverse systems
0 → (G0 /G0n ) → (G/Gn ) → (G00 /G00n ) → 0 and applying Proposition 5.2.5.

Corollary 5.2.7. For a topological abelian group G with system of open subgroups Gn and
b for each n ∈ N, G
completion G, bn is a topological subgroup of G
b and G/
b Gbn ∼
= G/Gn .
Proof. Apply Corollary 5.4.9 to the sequence 0 → Gn → G → G/Gn → 0. Notice that
G/Gn ∼ \n ∼
= G/G = G/
b G bn because each has the discrete topology.

bb ∼ b
Corollary 5.2.8. For any topological abelian group G, G = G.
Definition. A topological abelian group G is complete if the canonical map ϕ : G → G,
b x 7→
[(x)], is an isomorphism.

In this language, Corollary 5.2.8 says that completions are complete.


b is an isomorphism, so T∞ Gn = ker ϕ =
Remark. Notice that if G is complete, ϕ : G → G n=1
0 and hence by Lemma 5.1.4, G is Hausdorff.

94
5.3. Topological Rings and Module Filtrations Chapter 5. Completion and Filtration

5.3 Topological Rings and Module Filtrations


One of the most important concepts in commutative algebra is the completion of a ring with
respect to a certain topology induced by one of its ideals. We first define the notion of a
topological ring.

Definition. A topological ring is a ring A which is a topological abelian group such that
the ring multiplication A × A → A, (a, b) 7→ ab, is a continuous map. Topological fields are
defined analagously.

Given an ideal J ⊂ A, one can always give A the structure of a topological ring:

Definition. Let A be a ring and J ⊂ A an ideal. Then the chain

A = J0 ⊇ J1 ⊇ J2 ⊇ J3 ⊇ · · ·

forms a basis of additive subgroups of (A, +) as a topological abelian group. This topology,
called the J-adic topology, makes A into a topological ring.

Lemma 5.3.1. The completion A


b of a topological ring A is a topological ring.

Examples.

1 Let A = k[t] be the polynomial ring in a single variable and take J = (t). Then
A
b = k[[t]], the power series ring.

2 When A = Z and J = (p) for a prime integer p, the completion Zb is called the p-adic
completion of Z, denoted Zp . Alternatively, Zp = lim Z/pn Z.
←−

Definition. Let A be a topological ring. An A-module M is called a topological A-module


if M is a topological abeliain group and the A-action map A × M → M, (a, m) 7→ am, is a
continuous map.

If J ⊂ A is an ideal of A, so that A is a topological ring with the J-adic topology,


for an A-module M one can define open sets Mn = J n M that make M into a topological
A-module. Then the topological completion M c of M with respect to this topology is a
topological A-module.
b
Rather than using J n M to define the J-adic topology on M , we instead use systems of
additive subgroups of M called filtrations.

Definition. A filtration of an A-module M is an infinite chain of submodules

M = M0 ⊇ M1 ⊇ M2 ⊇ · · ·

If J ⊂ A is an ideal, a J-filtration of M is a filtration such that JMn ⊆ Mn+1 for all n ≥ 0.


If in addition JMn = Mn+1 for all n larger than some n0 ∈ N, then the filtration is called a
stable J-filtration.

95
5.3. Topological Rings and Module Filtrations Chapter 5. Completion and Filtration

Lemma 5.3.2. If (Mn ) and (Mn0 ) are stable J-filtrations of M , then there exists an n0 ∈ N
0
such that Mn+n0 ⊆ Mn0 and Mn+n 0
⊆ Mn for all n ≥ 0.
Proof. Let n0 ∈ N such that JMn = Mn+1 and JMn0 = Mn+1 0
for all n ≥ n0 . Then
JMn0 = Mn0 +1 and JMn0 = Mn0 +1 . By induction, J Mn0 = Mn0 +n and J n Mn0 0 = Mn0 0 +n .
0 0 n

Then we have
J n M ⊆ J n−1 M ⊆ · · · ⊆ JM ⊆ M
and J n M 0 ⊆ J n−1 M 0 ⊆ · · · ⊆ JM 0 ⊆ M 0 for all n ≥ n0 .
From this we obtain Mn+n0 = J n Mn0 ⊆ J n M ⊆ Mn0 . Reversing the roles of the Mn and Mn0
gives the other containment.
Remark. T∞The ncompletion A along an ideal J can also be described in terms of a metric, as
b
long as n=1 J = 0: if x, y ∈ A, define
1
d(x, y) = where v(x, y) = sup{k ≥ 1 | x − y ∈ J k }.
2v(x,y)
Formally, we set d(x, y) = 0 if v(x, y) = ∞. Completion of the induced metric topology
on A, that is, by formally adding in limits of Cauchy sequences in this topology, yields a
topological ring isomorphic to A.b Moreover, this description makes it easier to see that the
ring operations + and · on A are continuous with respect to the metric topology and extend
to continuous operations on A, b and that the canonical embedding A ,→ A b has dense image.
Proposition 5.3.3. Suppose A is a ring and J ⊂ A is an ideal such that ∞ n
T
n=1 J = 0. Let
Ab be the completion of A along J. Then J A b ⊆ J(A),
b the Jacobson radical of A. b
Proof. By Lemma 1.1.1 it suffices to show 1 − x is a unit in A b for all x ∈ J. Consider the
formal expansion
1
= 1 + x + x2 + . . .
1−x
Set sn = 1 + x + . . . + xn ∈ A. Then for each n ≥ 1, sn − sn+1 = −xn+1 ∈ J n+1 so (sn ) is a
1
Cauchy sequence with respect to the J-adic metric and hence its limit 1−x exists in A.
b

Lemma 5.3.4. If J ⊂ A is an ideal with ∞


T n b nb∼ n
n=1 J = 0, then A/J A = A/J for all n ≥ 1.
In particular, if m ⊂ A is a maximal ideal, then Ab is a local ring with maximal ideal m b = mA.
b
Example 5.3.5. The completion of a polynomial ring A = k[t1 , . . . , tn ] along the maximal
ideal m = (t1 , . . . , tn ) is a power series ring A b = k[[t1 , . . . , tn ]], which is local with maximal
ideal (t1 , . . . , tn ). More generally:
Proposition 5.3.6. If J = (a1 , . . . , an ) is a finitely generated ideal of A with ∞ n
T
n=1 J = 0,
then
Ab = A[[t1 , . . . , tn ]]/(t1 − a1 , . . . , tn − an ).
Corollary 5.3.7. If A is noetherian then any completion A b is also noetherian.
Example 5.3.8. Be aware that localization and completion do not commute. For example,
consider the ideal J = (x) in A = k[x, y]. Then A(x) = (k(y)[x])(x) and the completion of this
ring along (x) is k(y)[[x]]. Meanwhile, the completion A b of A along (x) is k[[x]][y], whose
localization at (x)Ab is the proper subring (k[[x]][y])(x) ( k(y)[[x]].

96
5.4. Graded Rings and Modules Chapter 5. Completion and Filtration

5.4 Graded Rings and Modules


L∞
Definition. A ring A is said to be graded if as an abelian group, A = n=0 An for sub-
groups An ≤ A with the property that Am An ≤ Am+n for all m, n ∈ N0 .

Notice
L∞ that A0 is always a subring of A, so each An is an A0 -module. We also define
A+ = n=1 An , which is an ideal of A.

Definition. If A is aL
graded ring, an abelian group M is a graded A-module if M is an

A-module and M = n=0 Mn for subgroups Mn ≤ M such that Am Mn ⊆ Mm+n for all
m, n ∈ N0 .

As above, each Mn is not just an A-module but also an A0 -module.


L∞
Definition. Given a graded A-module M = n=0 Mn , an element x ∈ Mn is said to be
homogeneous of degree n in M .

Definition. Let A and B be graded rings. A homomorphism of graded rings is a ring


map f : A → B such that f (An ) ⊆ Bn for all n ∈ N0 . Likewise, if M and N are two graded
A-modules, a homomorphism of graded modules is an A-module map f : M → N such
that f (Mn ) ⊆ Nn for all n ∈ N0 .

Example 5.4.1. If A = k[t1 , . . . , tn ] then A is a graded ring with respect to total degree:

M X
A= An where A0 = k and An = kti11 · · · tirr
n=0 (i1 ,...,ir )∈Nr0
i1 +...+ir =n

This is called the standard grading on the polynomial ring A. The ideal A+ here is (t1 , . . . , tn ).

Lemma 5.4.2. For a graded ring A, the following are equivalent:

(i) A is noetherian.

(ii) A0 is noetherian and A/A0 is a finitely generated ring extension.

Proof. (ii) =⇒ (i) follows directly from Corollary 3.3.3.


(i) =⇒ (ii) If A is noetherian, then A0 ∼ = A/A+ is noetherian and A+ is finitely generated.
Let A+ = (x1 , . . . , xs ) for xi ∈ A. We may assume each xi ∈ Aki is homogeneous of degree
ki . We will show An ⊆ A0 [x1 , . . . , xs ] for all n ∈ N. The n = 0 case is trivial so Pnassume
n ≥ 1. Then An ⊂ A+ so for any x ∈ An there are elements ci ∈ A such that x = i=1 ci xi .
Since x is homogeneous of degree n, we may assume ci ∈ An−ki where by convention we treat
A` = 0 when ` < 0. Since all ki > 0, then ci ∈ A0 [x1 , . . . , xs ] by induction, and therefore
x ∈ A0 [x1 , . . . , xs ]. It follows that A ⊆ A0 [x1 , . . . , xs ], so A = A0 [x1 , . . . , xs ] and A/A0 is a
finitely generated extension of rings.

97
5.4. Graded Rings and Modules Chapter 5. Completion and Filtration

Remark. If A is a ring and I ⊂ A is an ideal, there are two canonical graded rings we can
associate to I: ∞ ∞
M M
∗ n
A := I and GI (A) := I n /I n+1 .
n=0 n=0

By convention, we let I 0 = A. Notice that I 0 = A ,→ A∗ , so that A∗0 = A. If A is noetherian,


then I is finitely generated as a submodule, that is, I = (x1 , . . . , xr ) for some xi ∈ I. Then
powers of I are generated by homogeneous products of the generators of I:
r
!
Y
In = xei i : ei ∈ N0 , e1 + . . . + er = n .
i=1

In particular, this shows that A∗ = A[x1 , . . . , xs ], so A∗ is a noetherian ring by Corol-


lary 3.3.3(b).

Suppose I is an ideal of a ring A. Let M be an A-module with I-filtration (Mn )n∈N0 ;


recall that this means there is a chain of submodules

M = M0 ⊇ M1 ⊇ M2 ⊇ · · ·

such that IMn ⊆ Mn+1 for all n ∈ N0 . The standard I-filtration on M is when Mn = I n M ,
which defines the I-topology (or I-adic topology) on M . For any filtration (Mn )n∈N0 , we
define a graded module from M by

M

M = Mn .
n=0

Lemma 5.4.3. M ∗ is a graded A∗ -module.

Lemma 5.4.4. Let A be a noetherian ring, I ⊂ A and ideal and M a finitely generated
A-module. Then for an I-filtration (Mn )n∈N0 of M , the following are equivalent:

(i) M ∗ = ∞ ∗
L
n=0 Mn is a finitely generated A -module.

(ii) (Mn )n∈N0 is a stable I-filtration.

Proof. Note that by Corollary 3.1.5, the assumptions that A is noetherian and M is finitely
generated imply that M is a noetherian A-module. Thus each submodule Mn ⊆ M is finitely
generated, so the modules
Mn
Qn := Mi
n=0

are finitely generated for each n ∈ N, by Corollary 3.1.4. For each n, Qn generates the
following A∗ -submodule of M ∗ :

M
Q∗n := Qn ⊕ I m Mn .
m=1

98
5.4. Graded Rings and Modules Chapter 5. Completion and Filtration

(That Q∗n is an A∗ -submodule follows from the fact that I m Mn ⊆ Mm+n for all m ∈ N0 .)
Moreover, Q∗n is a finitely generated A∗ -module generated by the A-module generators of
Q
Sn∞. By∗ construction, Q∗n ⊆ Q∗n+1 for all n, since I m Mn ⊆ I m−1 Mn+1 for all m ≥ 1. Also,

n=0 Qn = M . With these observations, we now proceed to the main proof.
(i) =⇒ (ii) If M ∗ is a finitely generated A∗ -module then it follows (3.1.5) that M ∗ is
noetherian as an A∗ -module, i.e. M ∗ satisfies the ACC on A∗ -submodules. In particular,
the ascending chain Q∗n ⊆ Q∗n+1 stabilizes,
S∞ so there is some n0 ∈ N0 such that Q∗n = Q∗n0 for
all n ≥ n0 . By the above, M = n=0 Qn = Q∗n0 so Q∗n0 is a finitely generated A∗ -module.
∗ ∗

Then ∞ n0 ∞
M M M
Mn = Mn ⊕ I m Mn0 =⇒ Mn0 +m = I m Mn0 for all m ∈ N0 .
n=0 n=0 m=n0

This is equivalent by Lemma 5.3.2 to (Mn ) ∈n∈N0 being a stable I-filtration.


(ii) =⇒ (i) If (Mn ) is a stable filtration, there is some n0 ∈ N0 such that Mn0 +m = I m Mn0
for all m ∈ N0 . This implies M ∗ = Q∗n0 so by the preliminary remarks, M ∗ is a finitely
generated A∗ -module.
Proposition 5.4.5. Assume A is a noetherian ring, I ⊂ A is an ideal, M is a finitely
generated A-module with stable I-filtration (Mn )n∈N0 and M 0 ⊆ M is a submodule. Then the
filtration Mn0 = M 0 ∩ Mn induced on M 0 is a stable I-filtration.
Proof. First, (Mn0 ) is an I filtration, since for each n,

IMn0 = I(M 0 ∩ Mn ) ⊆ IM 0 ∩ IMn ⊆ M 0 ∩ Mn+1 ⊆ M 0 ∩ Mn .

If (Mn ) is stable then by Lemma 5.4.4, M ∗ = ∞ ∗


L
n=0 Mn is a finitely generated A -module, and
thus by Lemma 5.4.2 a noetherian A∗ -module. Then (M 0 )∗ exists and is an A∗ -submodule
of M ∗ . Since M ∗ is noetherian, (M 0 )∗ is finitely generated as an A∗ -submodule. Finally,
Lemma 5.4.4 implies that (Mn0 ) is a stable I-filtration.
We now obtain the useful Artin-Rees Lemma, which has a number of important conse-
quences that follow.
Corollary 5.4.6 (Artin-Rees Lemma). Assume A is a noetherian ring, I ⊂ A is an ideal, M
is a finitely generated A-module and M 0 ⊆ M is a submodule. Then there exists an n0 ∈ N0
such that M 0 ∩ I n+n0 M = I n (M 0 ∩ I n0 M ) for all n ∈ N0 .
Proof. Apply Proposition 5.4.5 to the stable I-filtration Mn = I n M of M .
Corollary 5.4.7. Assume A is a noetherian ring, I ⊂ A is an ideal, M is a finitely generated
A-module and M 0 ⊆ M is a submodule. Then the subspace topology induced on M 0 by the
I-topology on M is precisely the I-topology on M 0 .
Corollary 5.4.8 (Krull’s Intersection Theorem). Suppose A is a noetherian ring. Then
If I ⊂ A is an ideal such that I ⊆ J(A) and M is a finitely generated A-module, then
(a) T
∞ n
n=0 I M = 0. In particular, the I-topology on M is Hausdorff.

(b) If A is an integral domain and I is a proper ideal of A, then ∞ n


T
n=0 I = 0.

99
5.4. Graded Rings and Modules Chapter 5. Completion and Filtration

Proof. (a) Set M 0 = ∞ n 0


T
n=0 I M so that M ⊆ M is a finitely generated A-submodule of M (A
is noetherian and M is finitely generated). By the Artin-Rees lemma, there exists an n0 ∈ N0
so that for all n ∈ N0 , M 0 ∩ I n+n0 M = I n (M 0 ∩ I n0 M ). But notice that M 0 ∩ I n+n0 M = M 0
and M 0 ∩ I n0 M = M 0 , so we have M 0 = I n M 0 . Since I ⊆ J(A), Nakayama’s lemma (1.2.1)
implies M 0 = 0.
(b) Now suppose A is an integral domain. Set J = ∞ n
T
n=0 I and suppose x ∈ J. Then
x ∈ I k for all k ∈ N0 , so in particular (x) ⊆ I k for every k. By the Artin-Rees lemma, there
is an n0 ∈ N0 such that for all n ∈ N0 , (x) ∩ I n+n0 = I n ((x) ∩ I n0 ). Since (x) ⊆ I k for all
k, this becomes (x) = I n (x) for all n, even (x) = I(x). Thus there is some a ∈ I such that
x = ax, which can be written x(1 − a) = 0. If a were 1, I would be the unit ideal, but this
contradicts the assumption that I is proper. So a 6= 1, and since A is a domain, it follows
that x = 0.

Corollary 5.4.9. Suppose A is a noetherian ring and 0 → M 0 → M → M 00 → 0 is a


short exact sequence of A-modules. For a fixed ideal I ⊂ A, consider the I-topologies on
M 0 , M, M 00 and the corresponding completions M
c0 , M c00 . Then the sequence
c, M

c0 → M
0→M c00 → 0
c→M

is also exact.

Proof. This follows from Corollary 5.4.7 and Lemma 5.3.2.


Now recall the other canonical graded ring associated to an ideal I ⊂ A, namely

M
GI (A) = I n /I n+1 .
n=0

The multiplication law is given by

I m /I m+1 × I n /I n+1 −→ I m+n /I m+n+1


(x + I m+1 , y + I n+1 ) 7−→ xy + I m+n+1 .

This is well-defined since I n ⊇ I n+1 is a chain. If M is an A-module with I-filtration


(Mn )n∈N0 , we can define a GI (A)-module by

M
GI (M ) = Mn /Mn+1 .
n=0

Then GI (M ) is a graded GI (A)-module via the multiplication law

I m /I m+1 × Mn /Mn+1 −→ Mm+n /Mm+n+1


(x + I m+1 , z + Mn+1 ) 7−→ xz + Mm+n+1

which is well-defined since xz ∈ I m Mn ⊆ Mm+n .

Lemma 5.4.10. If A is a noetherian ring, then for every ideal I ⊂ A,

100
5.4. Graded Rings and Modules Chapter 5. Completion and Filtration

(a) GI (A) is noetherian.

(b) For any finitely generated A-module M with stable I-filtration (Mn )n∈N0 , GI (M ) is
finitely generated as a GI (A)-module.

Proof. (a) Since A is noetherian, I is finitely generated, say by x1 , . . . , xr ∈ I. Set x̄i =


xi + I 2 ∈ I/I 2 . Then as before, it can be checked that each I n /I n+1 is generated by
products of powers of the x̄i . By this, GI (A) = A/I[x̄1 , . . . , x̄r ] which is noetherian by
Corollary 3.3.3(b).
(b) If (Mn ) is a stable I-filtration of M , there is some n0 ∈ N0 such that Mn0 +n = I n Mn0
for all n0 ∈ N0 . This implies Mn0 +n /Mn0 +n+1 = (I n /I n+1 )Mn0 /Mn0 +1 for all n ∈ N0 , and
thus n0
M
GI = G(A) Mi /Mi+1 .
i=0

Now each Mi /Mi+1 is a finitely generated A-module and even an A/I-module, and A/I ⊆
GI (A). It follows that GI (M ) is a finitely generated GI (A)-module.

101
Chapter 6

Dimension Theory

102
6.1. Hilbert Functions Chapter 6. Dimension Theory

6.1 Hilbert Functions


Let A = ∞
L
n=0 An be a graded noetherian ring. Recall from Lemma 5.4.2 that A 0 is noethe-
rian and A = A0 [x1 , . . . , xs ] for some xi ∈ Aki , ki ≥ 1, 1 ≤ i ≤ s. Also let M = ∞
L
n=0 Mn be
a finitely generated graded A-module, generated by homogeneous elements m1 , . . . , mt such
that mj ∈ Mrj for each 1 ≤ j ≤ t.

Lemma 6.1.1. Each Mn is a finitely generated A0 -module.

Proof. It is clear that each Mn is an A0 -module by the definitionP of a graded A-module.


Take z ∈ Mn . Then there exist aj ∈ A for 1 ≤ j ≤ t such that z = tj=1 aj mj . Further, we
can choose polynomials fj ∈ A0 [t1 , . . . , ts ] such that aj = fj (x1 , . . . , xs ) for each 1 ≤ j ≤ t.
Note that monomials in the xi are homogeneous: xe11 · · · xess ∈ Ak1 e1 +...+ks es . Since z ∈ Mn ,
we can write
X t
z= gj (x1 , . . . , xs )mj
j=1

where for each j, gj (x1 , . . . , xs ) is an A0 -linear combination of monomials xe11 · · · xess such
that k1 e1 + . . . + ks es = n − rj . This shows Mn is finitely generated as an A0 -module: the
explicit generators are all elements of the form xe11 · · · xess mj for 1 ≤ j ≤ t and ei ≥ 0 such
that k1 e1 + . . . + ks es = n − rj .
Let C be the class of finitely generated A0 -modules.

Definition. A function λ : C → Z is said to be additive if for any short exact sequence


0 → M 0 → M → M 00 → 0 in C, we have

λ(M ) = λ(M 0 ) + λ(M 00 ).

Examples.

1 If A0 is noetherian, the length function `(M ) is additive by Lemma 3.2.3.

2 In the special case that A0 = k is a field, C is the class of all finite dimensional k-vector
spaces. Here, the dimension function λ(M ) = dimk M is an additive function.
f1 f2 fn fn+1
Lemma 6.1.2. Suppose 0 → M0 − → M1 − → · · · −→ Mn −−→ 0 is an exact sequence in C.
Then for any additive function λ : C → Z,
n
X
(−1)i λ(Mi ) = 0.
i=0

Proof. For each 1 ≤ i ≤ n + 1, set Ni = im fi which also equals ker fi+1 by exactness. Then
we have short exact sequences
fi+1
0 → Ni → Mi −−→ Ni+1 → 0

103
6.1. Hilbert Functions Chapter 6. Dimension Theory

for every 1 ≤ i ≤ n + 1. By additivity, λ(Ni ) − λ(Mi ) + λ(Ni+1 ) = 0 for all i, so we get a


telescoping sum:
n
X n
X
(−1)i λ(Mi ) = (−1)i (λ(Ni ) + λ(Ni+1 )) = λ(N0 ) ± λ(Nn+1 ).
i=0 i=0

Finally, additivity implies λ(0) = 0 so we have λ(N0 ) = λ(Nn+1 ) = 0. The result then
follows.

Definition. Let A be a graded ring, C the class of finitely generated A0 -modules and λ an
additive function on C. Then for any module M ∈ C, we define the Poincaré series of M
with respect to λ as

X
PM (t) = λ(Mn )tn ∈ Z[[t]].
n=0

Formally, the Poincaré series of any module is a power series with coefficients in Z. A
surprising result is that for finitely generated A0 -module M , PM (t) is in fact a rational
function.

Theorem 6.1.3 (Hilbert, Serre). For any M ∈ C and additive function λ, there is some
polynomial f (t) ∈ Z[t] such that

f (t)
PM (t) = Qs ki
∈ Q(t).
i=1 (1 − t )

Proof. We prove this by inducting on s, the number of generators of A0 -generators of A. If


s = 0, A = A0 and M is a finitely generated A-module, so Mn = 0 for all n > max{rj | 1 ≤
j ≤ t}. Thus PM (t) is a polynomial in this case, so f (t) = PM (T ) ∈ Z[t] works.
Now assume the result holds for all modules in C generated as an A0 -module by s − 1
elements or fewer. Consider the A-module homomorphism f : M → M, z 7→ xs z. This gives
us A-modules K = ker f and L = coker f = M/xs M . Since A is noetherian, K and L are
finitely generated (by Corollary 3.1.5 and Proposition 3.1.2). Note that xs K = f (K) = 0 and
xs L =L xs M/xs M = 0, so L
K and L are in fact graded modules over the quotient Ā = A/xs A:
K = ∞ n=0 K n and L = ∞
n=0 Ln . Also notice that Lr = Mr whenever r < ks . Moreover,
by Lemma 3.1.3, Ā is noetherian so K and L are finitely generated as Ā-modules. We can
view Ā as a graded noetherian ring:

M
Ā = Ān = Ā0 [x̄1 , . . . , x̄s ]
n=0

where Ān = An /xs An−ks and x̄i = xi + xs Aki −ks for all n, i. Since ki ≥ 1 for all i, we actually
have Ā0 ∼
= A0 , so the finitely generated modules over Ā0 are precisely the elements of C. In
a moment, this will allow us to apply the induction hypothesis to K and L.
For n ∈ N0 , let fn denote the restriction of f to Mn . We get the following exact sequences:
fn
0 → Kn → Mn −→ Mn+ks → Ln+ks → 0 for n ∈ N0 .

104
6.1. Hilbert Functions Chapter 6. Dimension Theory

By Lemma 6.1.2, λ(Kn ) − λ(Mn ) + λ(Mn+ks ) − λ(Ln+ks ) = 0. Multiplying through by tn+ks


and rearranging terms yields

tks (tn λ(Kn ) − tn λ(Mn )) = tn+ks (λ(Ln+ks ) − λ(Mn+ks ))


=⇒ tks (PK (t) − PM (t))n+ks = (PL (t) − PM (t))n+ks for all n ∈ N0
ks
=⇒ t (PK (t) − PM (t)) = PL (t) − PM (t)

since both power series have 0 coefficients for indices less than ks , and the second line implies
that the power series agree in every remaining term. This can be rearranged to read

(1 − tks )PM (t) = PL (t) − tks PK (t).

By the induction hypothesis, there exist polynomials g(t), h(t) ∈ Z[t] such that
s−1
Y s−1
Y
(1 − tki )PL (t) = g(t) and (1 − tki )PK (t) = h(t).
i=1 i=1

It follows that s
Y
(1 − tki )PM (t) = g(t) − tks h(t) ∈ Z[t].
i=1

Thus the result holds for PM (t) for all M ∈ C by induction.


Definition. The dimension of a graded A-module M is defined to be the order of the pole
of PM (t) at t = 1, denoted d(M ). Formally, we set d(M ) = 0 if there is no pole at t = 1.
For a graded A-module M of dimension d = d(M ), Theorem 6.1.3 allows us to write the
Poincaré series M as
g(t)
PM (t) =
h(t)(1 − td )
where h(1) 6= 0, and if d ≥ 1, g(1) 6= 0.
Corollary 6.1.4. If d(M ) ≥ 1 and x ∈ Ak such that xz 6= 0 for all z ∈ M r {0}, then
d(M/xM ) = d(M ) − 1.
Proof. Consider the map f : M → M, z 7→ xz. As in Theorem 6.1.3, we have modules
K = ker f and L = coker f = M/xM . Since xz 6= 0 for any z ∈ M r {0}, K = 0. Then the
proof of Theorem 6.1.3 shows that (1−tk )PM (t) = PL (t). It follows that d(L) = d(M )−1.
Lemma 6.1.5. For all n, s ∈ N0 with n ≥ s,
 
n+s−1
#{(i1 , . . . , is ) ∈ Ns0 : i1 + . . . + is = n} = .
s−1
Example 6.1.6. Let k be a field and consider the polynomial ring A = k[t1 , . . . , ts ]. Then
by Example 5.4.1, A is a graded ring with
M
A0 = k and An = kti11 · · · tiss for each n ≥ 1.
ij ∈N0
i1 +...+is =n

105
6.1. Hilbert Functions Chapter 6. Dimension Theory

Let λ(M ) = dimk M be the additive dimension function for k-vector spaces M . Notice that
n+s−1
by Lemma 6.1.5, dimk An = s−1 for each n ≥ 1, so the Poincaré series for A as a module
over itself is ∞ ∞  
X
n
X n+s−1 n 1
PA (t) = (dimk An )t = t = s
.
n=0 n=0
s − 1 (1 − t)
Hence d(A) = s so the notion of vector space dimension coincides with the dimension of a
graded ring. Notice that in this example, k1 = k2 = · · · = ks , i.e. the generators t1 , . . . , ts
are all homogeneous of degree 1. The following proposition generalizes this and introduces
the Hilbert polynomial of an A-module.

Proposition 6.1.7. Suppose A = A0 [x1 , . . . , xs ] is a graded noetherian ring such that xi ∈


A1 for all 1 ≤ i ≤ s. Then for any additive function λ and graded A-module M , there exists
n0 ∈ N0 and a polynomial g(t) ∈ Q[t] with deg g = d(M ) − 1 such that λ(Mn ) = g(n) for all
n ≥ n0 .

Proof. For fixed k, ` ∈ N, n+k


 1
`
= `! (n + k)(n + k − 1) · · · (n + k − ` + 1) is a polynomial in
n of degree ` with coefficients in Q. The assumption that each xi ∈ A1 means k1 = . . . ks ,
so Theorem 6.1.3 gives us a function f (t) ∈ Z[t] such that
∞  
f (t) X n+s−1 n
PM (t) = = f (t) t
(1 − t)s n=0
s − 1
Pn0
by Lemma 6.1.5. Write f (t) = k=0 ak tk , with ak ∈ Z and n0 ∈ N. If d(M ) = 0, then
PM (t) ∈ Z[t] and so λ(Mn ) = 0 for all n > deg f . In this case, g = 0 works; formally we
let deg 0 =P−1. Now if d(M ) 6= 0, we may assume s = d = d(M ) and f (1) 6= 0. Then in
particular nk=0
0
ak = f (1) 6= 0 so we have
n0
! ∞  ! ∞
X X m + d − 1 X
k m
PM (t) = ak t t = λ(Mn )tn .
k=0 m=0
d − 1 n=0

This shows that for all n ≥ n0 ,


n0  
X n−k+d−1
λ(Mn ) = ak .
k=0
d−1

Set g(t) = nk=0 ak t−k+d−1


P 0 
d−1
. Then by the preliminary comment, g is a polynomial of degree
at most d − 1 in Q[t], and g(n) = λ(Mn ) for all n ≥ n0 . Finally, notice that
n0 n0
!
nd−1
 
X n−k+d−1 X
λ(Mn ) = ak = ak + lower terms
k=0
d − 1 k=0
(d − 1)!
Pn0
and k=0 ak 6= 0, so the degree of g is exactly d − 1.

Definition. The function g is called the Hilbert polynomial of M with respect to the
additive function λ.

106
6.1. Hilbert Functions Chapter 6. Dimension Theory

Proposition 6.1.8. Given a noetherian ring A with maximal ideal m ∈ MaxSpec(A), an m-


primary ideal Q ⊂ A and a finitely generated A-module M with stable Q-filtration (Mn )∞
n=0 ,
the following hold:

(a) `(M/Mn ) < ∞ for all n ∈ N0 .

(b) If Q is generated by s elements, there exist n0 ∈ N and a polynomial g(t) ∈ Q[t] of


degree at most s with the property that `(M/Mn ) = g(n) for all n ≥ n0 .

(c) The degree and leading coefficient of g(t) are independent of the choice of Q-filtration.

Proof. (a) Recall the graded noetherian ring



M
G(A) = Gn (A), where G0 (A) = A/Q and Gn (A) = Qn /Qn+1 .
n=0

Notice that G0 (A) is noetherian and dim A/Q = 0 since r(Q) = m is the only prime ideal
of A containing Q. L∞Thus by Theorem 3.5.5, G0 (A) is artinian. The associated graded
module G(M ) = n=0 Mn /Mn+1 is a graded G(A)-module. By Lemma 5.4.10(b), G(M )
is finitely generated as a G(A)-module. Moreover, for each j ∈ N0 , Gj (M ) = Mj /Mj+1
is a finitely generated G0 (A)-module, hence artinian and noetherian by Corollary 3.1.5.
Hence
Pn−1 by Corollary 3.2.2, each quotient Mj /Mj+1 has a composition series, so `(M/Mn ) =
j=0 `(Mj /Mj+1 ) is finite.
(b) We have that Q = (x1 , . . . , xs ) as an ideal of A. For each xi , define x̄i = xi + Q2 ∈
Q/Q2 = G1 (A). Then G(A) = G0 (A)[x̄1 , . . . , x̄s ]. Note that k1 = . . . = ks = 1 in the
notation of Theorem 6.1.3. Consider the Poincaré series

X
PG(M ) (t) = `(Mn /Mn+1 )tn+1 .
n=0

By Proposition 6.1.7, there exist n0 ∈ N and a polynomial f (t) ∈ Q[t] of degree d−1 ≤ s−1,
where d = d(G(M )), such that `(Gj (M )) = f (j) for each j ≥ n0 . Then by (a), for all n ≥ n0
we have
n−1
X 0 −1
nX n−1
X
`(M/Mn ) = `(Mj /Mj+1 ) = `(Mj /Mj+1 ) + f (j).
j=0 j=0 j=n0
Pn−1
Now we show that j=n0 f (j) is a polynomial in n of degree d ≤ s with rational coefficients.
It is a factP that if f ∈ Q[t] with deg f = d − 1, then there exists g̃ ∈ Q[t] P with deg g̃ = d
such that nj=1 f (j) = g̃n for all n. To see this explicitly, write f (t) = `r=0 ar tr so that
Pn P` Pn r Pn r
j=1 f (j) = r=0 a r j=1 j . For all r ∈ N0 , there exists an hr ∈ Q[t] satisfying j=1 j =
hr (n); indeed, these arise as certain Bernoulli polynomials:

ˆ If r = 0, h0 (t) = t.

ˆ If r = 1, h1 (t) = 21 t + 12 t2 .

ˆ If r = 2, h2 (t) = 61 t + 12 t2 + 13 t3 , etc.

107
6.1. Hilbert Functions Chapter 6. Dimension Theory

In any case, we have nj=1 f (j) = `r=0 ar hr (n) which is now a polynomial in n as claimed.
P P
This proves (b).
(c) If (Mn ) and (Mn0 ) are two stable Q-filtrations of M , then by Lemma 5.3.2, there is
some m0 ∈ N0 such that Mn+m0 ⊆ Mn0 and Mn+m 0
0
⊆ Mn for all n ∈ N0 . By (b), we get
integers n0 , n0 ∈ N0 and polynomials g, g ∈ Q[t] such that for all n ≥ max{n0 , n00 },
0 0

g(n + m0 ) = `(M/Mn+m0 ) ≥ `(M/Mn0 ) = g 0 (n)


and g 0 (n + m0 ) = `(M/Mn+m
0
0
) ≥ `(M/Mn ) = g(n).

g(n)
It follows that lim = 1 so g and g 0 must have the same degree and leading coefficients.
n→∞ g 0 (n)

In the next section, we will prove that when A is local noetherian with maximal ideal m,
the dimension of A may be computed by

dim A = min{s ∈ N0 | there exists an m-primary ideal Q ⊂ A generated by s elements}.

With A, m, M, Q and (Mn )∞ n M


n=0 , where Mn = Q M , as in Proposition 6.1.8, let χQ (t) ∈
M
Q[t] denote the polynomial (g in the proposition) such that `(M/Mn ) = χQ (n) for all large
enough n.

Definition. When M = A, the polynomial χQ (t) := χA


Q (t) is called the characteristic
polynomial of Q.

Corollary 6.1.9. For a maximal ideal m ⊂ A and an m-primary ideal Q ⊂ A, χQ (t) is a


polynomial of degree at most min{s ∈ N0 | Q is generated by s elements}.

Proposition 6.1.10. For a maximal ideal m ⊂ A and any m-primary ideal Q ⊂ A,


deg χQ (t) = deg χm (t).

Proof. Since A is assumed to be noetherian, mr ⊆ Q ⊆ m for some r ∈ N by Lemma 3.4.1.


Then mnr ⊆ Qn ⊆ mn for all n ∈ N, so `(A/mn ) ≤ `(A/Qn ) ≤ `(A/mnr ). Thus for
sufficiently large n, χm (n) ≤ χQ (n) ≤ χm (nr), so taking the limit as n → ∞ implies the
statement.

Remark. Let A be a graded noetherian ring and m ⊂ A a maximal ideal. By Proposi-


tions 6.1.7 and 6.1.8, deg χm (t) = d(Gm (A)) where Gm (A) is considered as a graded Gm (A)-
module. By definition, d(Gm (A)) equals the order of the pole at t = 1 of the Poincaré series
PGm (A) (t), which equals deg g(Gm (A)) +1 when g is the Hilbert function of Gm (A). Therefore
we can write d(A) = deg χm (t) = d(Gm (A)).

108
6.2. Local Noetherian Rings Chapter 6. Dimension Theory

6.2 Local Noetherian Rings


For the entirety of this section, assume A is a local noetherian ring with maximal ideal m
and set

δ(A) = min{s ∈ N0 | there exists an m-primary ideal Q ⊂ A generated by s elements}.

Our goal is to show that δ(A) = d(A) = deg χm (t) = dim A.

Lemma 6.2.1. δ(A) ≥ d(A).

Proof. This follows immediately from Corollary 6.1.9 and Proposition 6.1.10.

Lemma 6.2.2. If Q is an m-primary ideal and M is a finitely generated A-module, then for
any x ∈ A satisfying xz =
6 0 for all z ∈ m r {0},
0
deg χM M
Q (t) ≤ deg χQ (t) − 1

where M 0 = M/xM .

Proof. Set N = xM ⊆ M . We have exact sequences


p
0→N →M →
− M →0 (E)
pn
0 → N/N ∩ Qm → M/Qn M −→ M 0 /Qn M 0 → 0 (En )

for each n ∈ N. Set Nn = N ∩ Qn M . Now (Qn M )∞ n 0 ∞ ∞


n=0 , (Q M )n=0 and (Nn )n=0 are
stable Q-filtrations of M , M 0 and N , respectively (for (Nn ), use Proposition 5.4.5). The
length function ` is additive, so by the short exact sequence (En ), `(N/Nn ) − `(M/Qn M ) +
`(M 0 /Qn M 0 ) = 0. Thus for some n0 ∈ N0 ,
0
χN M M
Q (n) − χQ (n) + χQ (n) = 0

for all n ≥ n0 . By assumption, M → N, z 7→ xz is an isomorphism of A-modules. So the sta-


ble Q-filtration (Nn ) of N corresponds to a stable Q-filtration of M . By Proposition 6.1.8(c),
χN M
Q (t) and χQ (t) have the same degree and leading coefficient. Taking n → ∞, this implies
0
that det χM M N M
Q (t) = deg(χQ (t) − χQ (t)) ≤ deg χQ (t) − 1.

Corollary 6.2.3. If x ∈ A is not a zero divisor or a unit then d(A/(x)) ≤ d(A) − 1.


A/(x)
Proof. Let M = A, M 0 = A/(x) and apply Lemma 6.2.2 to get deg χm (t) ≤ deg χm (t)−1 =
A/(x)
d(A) − 1. By definition, d(A/(x)) = deg χm/(x) (t) = deg χm (t) so the result follows.

Proposition 6.2.4. d(A) ≥ dim A.

Proof. We prove this by induction on d = d(A). If d = 0, `(A/mn ) is stationary for


large enough n, so there exists an n ∈ N0 for which `(A/mn ) = `(A/mn+1 ). This im-
plies mn = mn+1 = J(A)mn , so by Nakayama’s Lemma (1.2.1), mn = 0. Thus A is artinian
by Theorem 3.5.6. In particular, dim A = 0 by Theorem 3.5.5.

109
6.2. Local Noetherian Rings Chapter 6. Dimension Theory

Now suppose d = d(A) ≥ 1. Let p0 ( · · · ( pr be a chain of prime ideals in A. We


must show r ≤ d for all such r. Assuming r ≥ 1, pick x ∈ p1 r p0 and set A0 = A/p0
and x0 = x + p0 ∈ A0 . Since A0 is an integral domain, x0 is not a zero divisor and x0 6= 0
because x 6∈ p0 . Thus by Corollary 6.2.3, d(A0 /(x0 )) ≤ d(A0 ) − 1. Notice that m0 = m/p0
is the maximal ideal of A0 . Since ` is additive, the projections A/mn → A0 /(m0 )n imply
A0
`(A/mn ) ≥ `(A0 /(m0 )n ) for all n ∈ N0 . So deg χA 0
m (t) ≥ deg χm0 (t) which means d(A) ≥ d(A ).
This gives us d(A0 /(x0 )) ≤ d(A) − 1. Set p0i = pi /p0 for 1 ≤ i ≤ r, each of which is a prime
ideal of A0 . Then x0 ∈ p01 so

p01 /(x0 ) ( p02 /(x0 ) ( · · · ( p0r /(x0 )

is a chain of prime ideals in A0 /(x0 ). Since d(A0 /(x0 )) ≤ d(A) − 1, and dim A0 /(x0 ) ≤
d(A0 /(x0 )) ≤ d−1 by the inductive hypothesis, we get r −1 ≤ dim A0 /(x0 ) ≤ d−1. Therefore
r ≤ d. Taking the supremum over all such chains of prime ideals, we get dim A ≤ d.
Corollary 6.2.5. Let A be any noetherian ring. Then
(a) If A is local, dim A < ∞.

(b) For any prime ideal p ∈ Spec(A), ht(p) < ∞.


Proof. (a) follows immediately from Proposition 6.2.4. For (b), note that Ap is local noethe-
rian and ht(p) = dim Ap by Lemma 2.3.6(a).
Corollary 6.2.6. For a local noetherian ring, δ(A) ≥ d(A) ≥ dim A.
Example 6.2.7. Let (A, m, k) be a local noetherian ring. Then dimk m/m2 ≥ dim A. Indeed,
if x1 , . . . , xs ∈ m such that x̄1 , . . . , x̄s ∈ m/m2 form a k-basis of m/m2 , Proposition 1.2.4 says
that m is generated as an ideal by x1 , . . . , xs . Thus s ≥ δ(A) ≥ dim A.
For any m-primary ideal Q ⊂ A, set δ(Q) = min{s ∈ N0 | Q is generated by s elements},
so that δ(A) = min{δ(Q) | Q is m-primary}. Clearly δ(m) ≥ δ(A). Further, for any
x1 , . . . , xs ∈ m,
s
X
2
m = (x1 , . . . , xs ) ⇐⇒ m/m = Ax̄i where x̄i = xi + m2 , by Proposition 1.2.4
i=1
s
X
⇐⇒ m/m2 = kx̄i .
i=1

Taking the minimum over s on both sides, it follows that δ(m) = dimk m/m2 .
Definition. A local noetherian ring A is said to be a regular ring if dim A = dimk m/m2 .
By the above, A is regular if and only if m = (x1 , . . . , xd ), where d = dim A.
Theorem 6.2.8. Let (A, m, k) be a local noetherian ring of dimension d = dim A. Then the
following are equivalent:
(a) Gm (A) ∼
= k[t1 , . . . , td ] for indeterminates ti .

110
6.2. Local Noetherian Rings Chapter 6. Dimension Theory

(b) A is regular, i.e. dimk m/m2 = d.

(c) m = (x1 , . . . , xd ) for some x1 , . . . , xd ∈ A.

Lemma 6.2.9. Let A be a regular ring. Then A is an integral domain which is integrally
closed.

Proof. We prove T A is an integral domain. Take x, y ∈ A r {0}. By Krull’s intersection


theorem (5.4.8), ∞ n r
n=1 m = 0 so there exist r, s ∈ N0 such that x ∈ m , x 6∈ m
r+1
, y ∈ ms
s+1
and y 6∈ m . Let x̄ be the image of x in Gr (A) and ȳ the image of y in Gs (A). Then x̄ 6= 0
and ȳ 6= 0, so x̄ · ȳ 6= 0 in Gm (A), since by Theorem 6.2.8, Gm (A) ∼
= k[t1 , . . . , td ] which is an
integral domain. Thus xy 6= 0 in A.

Theorem 6.2.10 (Krull’s Height Theorem). Suppose A is noetherian and x1 , . . . , xs ∈ A.


Then for any prime ideal p ⊂ A which is minimal among prime ideals containing (x1 , . . . , xs ),
ht(p) ≤ s.

Proof. Assume I = (x1 , . . . , xs ) ⊆ p and p is minimal among primes containing I. Consider


the localization Ap = S −1 A, where S = A r p. Then Ap is local noetherian with maximal
ideal m = S −1 p. If p0 ⊂ A is any other prime ideal containing I and p0 ⊆ p, then by
minimality, p0 = p. Therefore m is the unique prime ideal −1
 of Ap containing S A, so by
−1 −1 x1 xs
Lemma 3.4.8(a), S is m-primary. Now S I = 1 , . . . , 1 so by definition, Corollary 6.2.6
and Lemma 2.3.6(a), s ≥ δ(Ap ) ≥ dim Ap = ht(p).

Theorem 6.2.11 (Krull’s Principal Ideal Theorem). Assume A is noetherian, x ∈ A is a


nonunit and p ∈ Spec(A) is a prime ideal minimal among prime ideals containing (x). Then

(a) ht(p) ≤ 1.

(b) If x is not a zero divisor of A, then ht(p) = 1.

Proof. (a) is immediate from Krull’s height theorem. For (b), suppose (x) ⊆ p and ht(p) = 0.
Then all elements of p are zero divisors by Corollary 3.4.16, so it must be that ht(p) = 1.

Theorem 6.2.12. Let A be a local noetherian ring. Then dim A = δ(A).

Proof. We have already proved dim A ≤ d := δ(A) in Corollary 6.2.6. For the other inequal-
ity, we construct a sequence of elements x1 , . . . , xd ∈ m satisfying the following condition:
For each 1 ≤ j ≤ d, every prime p ∈ Spec(A)
(∗)
containing (x1 , . . . , xj ) has height ht(p) ≥ j.
Given such x1 , . . . , xd , let Ij = (x1 , . . . , xj ) for each 1 ≤ j ≤ d. Then if p ⊇ Id , ht(p) ≥ d =
ht(m) but this is only possible if p = m. Hence Id is m-primary by Lemma 3.4.8(a), so we
get
dim A = d ≥ δ(Id ) ≥ δ(A).
Hence the result follows once we prove (∗).
We verify (∗) inductively. For 1 ≤ j ≤ d, let x1 , . . . , xj−1 be constructed. If p ∈ Spec(A)
such that p ⊇ Ij−1 , then ht(p) ≥ j − 1 by the inductive hypothesis. By Proposition 3.4.10,

111
6.2. Local Noetherian Rings Chapter 6. Dimension Theory

there are finitely many minimal elements of the set C = {p ∈ Spec(A) | p ⊇ Ij−1 }. Suppose
S1n, . . . , pn are the elements of C S
p with height j − 1 < d. Then pi 6= m for all 1 ≤ i ≤ n, so
n
i=1 i p ( m. Choose x j ∈ m r i=1 pi . It now follows that (∗) holds for these x1 , . . . , xd ,
and the theorem is proven.

Corollary 6.2.13. If A is a local noetherian ring, then δ(A) = d(A) = dim A.

We finish the section by proving a partial converse to Theorem 6.2.10. First we need a
technical lemma.

Lemma 6.2.14 (Prime Avoidance). Suppose S I1 , . . . , In ⊂ A are ideals of which at least n−2
are prime. Then for any ideal J ⊂ A, if J ⊆ nk=1 Ik then J ⊆ Ik for some 1 ≤ k ≤ n.

Proof. Suppose J 6⊂ Ik for any 1 ≤ k ≤ n. If n = 2, choose x1 ∈ J such that x1 6∈ I1 and


x2 ∈ J such that x2 6∈ I2 .SThen x1 + x2 ∈ J but x1 + x2 6∈ I1 ∪ I2 , so J 6⊂ I1 ∪ I2 . Inducting
on n > 2, we have J 6⊂ k6=` Ik for each 1 ≤ ` ≤ n. Without S loss of generality, assume I1
is prime. For each 1 ≤ ` ≤ n, choose x` ∈ J not lying in k6=` Ik , so in particular x1 ∈ I1 ,
and set x = x1 + x2 x3 · · · xn . If x ∈ I` for some ` ≥ 2, then x1 = x − x2 x3 · · · xn ∈ I` ,
contradicting the base case. On the other hand, if x ∈ I1 then x2 x3 · · · xn ∈ IS 1 , so x` ∈ I1
for some ` ≥ 2 by primality, also a contradiction. Hence J is not contained in nk=1 Ik .

Theorem 6.2.15. Let A be a local noetherian ring of dimension d, with maximal ideal m.
Then there exist d elements x1 , . . . , xd ∈ m such that r(x1 , . . . , xd ) = m.

Proof. If d = 0, we take (x1 , . . . , xd ) = 0. In this case, Theorems 3.5.5 and 3.5.6 say that
A is artinian and mn = 0 for some n ≥ 1. Thus r(0) = m by definition. Now induct on
d > 0. Since A is noetherian, the set of minimal primes of A is finite (Proposition 3.4.10),
say {p1 , . . . , pk }. Clearly m 6⊂ pi for any 1 ≤ i ≤ k, or else we would be in the base case. By
the prime avoidance lemma, m 6⊂ p1 ∪ · · · pk . Choose x1 ∈ m such that x1 6∈ p1 ∪ · · · ∪ pk .
Then dim(A/(x1 )) ≤ d − 1 so by induction, there exist d − 1 elements x̄2 , . . . , x̄d ∈ A/(x1 )
such that r(x̄2 , . . . , x̄d ) = m̄ in A/(x1 ). Lifting back to A, we see that r(x1 , . . . , xd ) = m as
desired.

Corollary 6.2.16. If A is local noetherian of dimension d then for any nonzero x ∈ m,


dim(A/(x)) = d − 1.

Proof. If dim(A/(x)) < d − 1 then we could choose a smaller number of elements to generate
m up to radical, but this would contradict Krull’s height theorem.

112
6.3. Complete Local Rings Chapter 6. Dimension Theory

6.3 Complete Local Rings


We say a ring A is complete if A is a completion of some ring along an ideal. The structure
of complete local rings is fully described in this section.
Theorem 6.3.1 (Hensel’s Lemma). Let (A, m, k) be a complete local ring with ∞ n
T
n=1 m = 0.
Suppose F (t) ∈ A[t] is a monic polynomial which factors as
F (t) ≡ g(t)h(t) mod m
for some monic, relatively prime polynomials g, h ∈ k[t]. Then there exist polynomials monic
G, H ∈ A[t] of the same degrees as g and h, respectively, such that F (t) = G(t)H(t) in A[t]
and G(t) = g(t) and H(t) = h(t) in k[t].
Proof. We inductively construct polynomials Gn , Hn ∈ A[t] satisfying
(1) Gn = g, H n = h and Gn − Gn+1 , Hn − Hn+1 ∈ mn A[t] for all n ≥ 1.
(2) F − Gn Hn ∈ mn A[t] for all n ≥ 1.
For n = 1, choose any monic polynomials G1 , H1 ∈ A[t] lifting g and h, i.e. so that G1 = g
and H 1 = h. Then
F − G1 H1 = F − gh ≡ 0 mod m
so F − G1 H1 ∈ mA[t]. Now assume Gn , Hn ∈ A[t] have been constructed and satisfy (1) and
(2). Then by (2), there are some y1 , . . . , y` ∈ mn and L1 , . . . , L` ∈ A[t] such that
`
X
F − Gn Hn = yi Li .
i=1

Since g and h are relatively prime, for each 1 ≤ i ≤ ` we can write


L i = ai g + b i h
for some ai , bi ∈ k[t] with deg ai < deg h by the division algorithm; note that this implies
deg bi < deg g as well. Lift each ai to Ai ∈ A[t] and bi to Bi ∈ A[t], so that Li −Ai Gn −Bi Hn ∈
mA[t] for each i. Then
`
X
F ≡ Gn Hn + yi (Ai Gn + Bi Hn ) mod mn+1 .
i=1

Set Gn+1 = Gn + i=1 yi Bi and Hn+1 = Hn + `i=1 yi Ai . Then since each yi ∈ mn , we see
P` P
that Gn − Gn+1 , Hn − Hn+1 ∈ mn A[t]. Moreover,
`
X X
F − Gn+1 Hn+1 = F − Gn Hn − yi (Ai Gn + Bi Hn ) − yi yj Bi Aj
i=1 i,j
X
2n
=− yi yj Bi Aj ∈ m A[t]
i,j

which implies F − Gn+1 Hn+1 ∈ mn+1 A[t] since 2n ≥ n + 1. Hence (Gn ) and (Hn ) are Cauchy
sequences in A[t] so they converge to some G, H ∈ A[t]. By construction, F − GH ∈ mn A[t]
for all n ≥ 1, so by hypothesis, F = GH.

113
6.3. Complete Local Rings Chapter 6. Dimension Theory

Corollary 6.3.2. If (A, m, k) is a complete local ring with ∞


T
n=1 m = 0 and f (t) is a monic
¯
polynomial such that f (t) has a simple root α ∈ k, then there exists an element a ∈ A with
ā = α and f (a) = 0.
Proof. This is the case where g(t) = t − α.
Definition. A coefficient field for a local ring (A, m, k) is a field L ⊆ A such that the
composition L ,→ A → A/m = k is an isomorphism.
There are three possibilities for the situation of a local ring having a coefficient field:
ˆ char A = 0 and Q ⊆ A, in which case A is called equicharacteristic of characteristic 0;

ˆ char A = p > 0 and Fp ⊆ A, in which case A is called equicharacteristic of characteristic


p;
ˆ A does not contain a field, which is only possible if char A = 0 and char k = p > 0,
and in this case A is called mixed characteristic.
Example 6.3.3. The power series rings C[[t]] and R[[t]] are examples of equicharacteristic
complete local rings in characteristic 0.
Example 6.3.4. For a prime p, Fp [[t]] is a complete local ring that is equicharacteristic of
characteristic p.
Example 6.3.5. The p-adic integers Zp are the prototypical example of a mixed character-
istic ring; they do not contain a field and have residue field Zp /pZp ∼
= Fp .
Theorem 6.3.6 (Cohen’sT∞ First Structure Theorem). If (A, m, k) is a complete local ring
n
containing a field and n=1 m = 0, then A has a coefficient field.
Proof. First assume char A = 0, so Q ⊆ A. Use Zorn’s Lemma to choose a maximal subfield
π
L ⊆ A. If L ,→ A → − k is already an isomorphism, we’re done. Otherwise, L still embeds
into k since both are fields, so choose some α ∈ k r π(L). If α is transcendental over π(L),
lift α to a ∈ A. Then for all nonzero f ∈ L[t], π(f (a)) = f (α) 6= 0 since α is transcendental,
so f (a) 6∈ m. Thus L(a) ⊆ A, contradicting maximality of L. So α must be algebraic, say
with minimal polynomial g(t) ∈ π(L)[t]. Since char L = 0, α is a simple root of g so by
Corollary 6.3.2, α lifts to a root a ∈ A of g ∈ π(L)[t] = L[t]. Then a is algebraic, so L[a]
is a field and it is contained in A, again contradicting maximality of L. So we must have
π(L) = k.
Now assume char A = p > 0. By Proposition 5.2.4, A b ∼= lim A/mn so to prove the
←−
theorem, we will exhibit a coefficient for each A/mn that are compatible with the maps
πn+1 : A/mn+1 → A/mn of the inverse system. We induct on n; for n = 1, k = A/m itself
is a coefficient field. For n = 2, set R = A/m2 , which is a local ring with maximal ideal
mR = m/m2 , and consider the ring

Rp = {xp | x ∈ R}

of pth powers in R – note that it is a ring by the ‘freshman’s dream’: (x + y)p = xp + y p .


Suppose xp ∈ Rp is nonzero. Then since m2R = 0, we must have x 6∈ mR and thus x is a unit

114
6.3. Complete Local Rings Chapter 6. Dimension Theory

in R. If x−1 ∈ R is an inverse of x, then (x−1 )p ∈ Rp is an inverse of xp so we see that Rp is


a subfield of R. Let L be a maximal subfield of R containing Rp and consider the projection
π : R → R/mR = k. If π(L) 6= k, choose some α ∈ k r π(L) and lift it to x ∈ R. Then
xp ∈ Rp ⊆ L is nonzero, so by the above it is a unit. This shows L[x] ) L ⊇ Rp which
contradicts maximality of L, so we must have π(L) = k. Thus R has a coefficient field.
To induct, suppose we have constructed a coefficient field Ln for A/mn . Set B =
−1
πn+1 (Ln ) ⊆ A/mn+1 . Then B is a subring of A/mn+1 containing mn /mn+1 . If b ∈ B is
not a unit, then b ∈ A/mn+1 and πn+1 (b) ∈ Ln , so πn+1 (b) = 0 since Ln is a field. Thus
b ∈ mn /mn+1 . This shows B is a local ring with maximal ideal mB = mn /mn+1 . Moreover,
m2n ⊆ mn+1 implies m2B = 0 so we are in the n = 2 case and there exists a coefficient field
Ln+1 ⊆ B ⊆ A/mn+1 . By construction, πn+1 (Ln+1 ) = Ln so taking L = lim Ln , we obtain a
←−
coefficient field of A.
b

Corollary
T∞ 6.3.7. Let (A, m, k) be a complete local ring of characteristic p > 0 such that
n
n=1 m = 0 and k is perfect. Then

e
\
L= Ap
e=1

is the unique coefficient field for A.


e
Proof. We first prove L is a field. Let a ∈ L be nonzero. For each e ≥ 1, we canTwrite a = xp
for some x ∈ A. If a ∈ m, then x ∈ m since m is prime. This shows a ∈ ∞ e=1 m
pe
= 0,
contradicting a 6= 0. Therefore a 6∈ m, so a is a unit and L is a field. Now let F be a
coefficient field for A. Then F ∼
e
= k so F p = F for all e ≥ 1. This implies
∞ ∞
pe e
\ \
F = F ⊆ Ap = L
e=1 e=1

so F ⊆ L. Reversing F and L gives an equality, so L is in fact unique.

Lemma 6.3.8. Let A T∞= An be the completion along an ideal I ⊆ A and suppose M is an
b
A-module such that n=1 I M = 0. If x1 , . . . , xr ∈ M such that M/IM is generated by
x̄i = xi + IM, 1 ≤ i ≤ r, then M is generated by the xi .

Proof. Set N = ri=1 Axi ⊆ M . For each m0 ∈ M , we can write


P

r
X
m0 = ai0 xi + m1
i=1

for some ai0 ∈ A and m1 ∈ IM . By hypothesis M = N + IM , so IM = IN + I 2 M and by


extension, I n M = I n N + I n+1 M for all n ≥ 1. Therefore we can write
r
X
m1 = ai1 xi + m2
i=1

115
6.3. Complete Local Rings Chapter 6. Dimension Theory

for ai1 ∈ A and m2 ∈ I 2 M . Repeating this argument, we can write


r n
!
X X
m0 = aij xi + mn+1
i=1 j=0
Pn
for any n ≥ 1, where aij ∈ I j and mn+1 ∈ I n+1 M . But for each i, j=0 aij is a Cauchy
sequence in A, so it converges to some ai ∈ A and
r
X
m0 − ai x i ∈ I n M
i=1

for all n. Hence m0 ∈ N .


Theorem 6.3.9 (Cohen’s T Second Structure Theorem). Let (A, m, k) be a complete local ring
containing a field with ∞ n=1 m n
= 0 and suppose m is finitely generated, say by x1 , . . . , xr .
Then there is a surjective ring map

k[[t1 , . . . , tr ]] −→ A

sending ti 7→ xi .
, . . . , tr ]] → A by sending ti 7→ xi as suggested. This means for an
Proof. Define π : k[[t1P
arbitrary power series αn1 ,...,nr tn1 1 · · · tnr r in k[[t1 , . . . , tr ]],
X  X
π αn1 ,...,nr tn1 1 · · · tnr r = αn1 ,...,nr xn1 1 · · · xnr r

and the partial sums of this series form a Cauchy sequence in A, so they converge. Hence π
is well-defined and is a homomorphism by construction. It remains to show π is surjective.
But A is a module over k[[t1 , . . . , tr ]] and A/(t1 , . . . , tr )A = A/m is generated by 1, so by
Lemma 6.3.8, A is generated by 1 as a k[[t1 , . . . , tr ]]-module. This implies π is surjective.
Corollary 6.3.10. If A is a complete local ring containing a field with ∞ n
T
n=1 m = 0 and m
is finitely generated, then A is noetherian.
Proof. Apply Theorem 3.3.9 and Lemma 3.3.1(a).
Definition. For a local ring A with maximal ideal m, a system of parameters for A is a
set of elements x1 , . . . , xn ∈ m such that r((x1 , . . . , xn )) = m.
Recall (Theorem 6.2.12) that if x1 , . . . , xn is a system of parameters for A, then n =
dim A.
Theorem 6.3.11 (Cohen’s Third Structure Theorem). Let (A, m, k) be a complete local
noetherian ring of dimension d containing a field and suppose x1 , . . . , xd ∈ m is a sytem of
parameters for A. Then the map

ϕ : k[[t1 , . . . , td ]] −→ A
ti 7−→ xi

identifies k[[t1 , . . . , td ]] ∼
= im ϕ ⊆ A and the resulting ring extension A/k[[t1 , . . . , td ]] is finite.

116
6.3. Complete Local Rings Chapter 6. Dimension Theory

Proof. As in the proof of the second structure theorem, ϕ is a well-defined homomorphism


since A is complete. Set I = (x1 , . . . , xd ), so that r(I) = m. Notice that A/I is a local
noetherian ring with maximal ideal m/I, and by Lemma 3.4.1, mn ⊆ I for some n ≥ 1. Then
Theorem 3.5.6 implies A/I is artinian. In particular, dimk A/I is finite by Proposition 3.5.8.
Choose a k-basis y1 , . . . , ym of A/I. Setting S = k[[t1 , . . . , td ]], we have k = S/(t1 , . . . , td ) so

A = Sy1 + . . . + Sym + (t1 , . . . , td )A.

Now Lemma 6.3.8 implies A = Sy1 +. . . Sym as an S-module. Let B = im ϕ ⊆ A. Then A is a


finite extension of B ∼
= S/ ker ϕ so d = dim A = dim B by Theorem 2.3.7. On the other hand,
by Krull’s height theorem (6.2.10), dim S ≤ d. If ker ϕ 6= 0, then Corollary 6.2.16 implies
dim B = dim(S/ ker ϕ) ≤ d − 1, a contradiction. Hence ker ϕ = 0 so k[[t1 , . . . , td ]] ∼
= B.
Remark. One may regard Cohen’s third structure theorem as a complete version of Noether’s
normalization lemma (Theorem 2.4.1).

Example 6.3.12. Consider the 1-dimensional local ring A = R[x](x2 +1) with maximal ideal
m = (x2 + 1)A. Then the residue field of A is C but C does not embed into A. By Cohen’s
first structure theorem (6.3.6), C does embed into the completion A
b but it’s not immediately
obvious how this embedding is defined. Note that understanding C ,→ A b comes down to
2 2 2
finding an element u ∈ A b such that u = −1. Since x + 1 ∈ m, 1 − (x + 1) is a unit and
we have
x2
= −1.
1 − (x2 + 1)
Applying Hensel’s lemma (6.3.1) produces a square root of 1 − (x2 + 1), so set
x
u= p .
1 − (x2 + 1)

Explicitly, u can be identified with the Cauchy sequence


1 3 5
u = x + x(x2 + 1) + x(x2 + 1)2 + x(x2 + 1)3 + . . .
2 8 16

117
Part II

Homological Algebra

118
Chapter 7

Category Theory

119
7.1. Categories and Functors Chapter 7. Category Theory

7.1 Categories and Functors


Category theory is a way of describing ‘classes’ of things in mathematics and the relations
between them in a general way.
Definition. A category C is a class of objects obj(C) and for every pair of objects A, B
in obj(C) a set of morphisms HomC (A, B) with a composition law:

HomC (B, C) × HomC (A, B) −→ HomC (A, C)


(g, f ) 7−→ g ◦ f = gf.

For every A ∈ obj(C) there is an identity morphism 1A ∈ HomC (A, A) such that for all
f ∈ HomC (A, B), f ◦ 1A = f and likewise for HomC (B, A).
Examples.
1 Veck is a category consisting of vector spaces over a field k as the objects and linear
transformations as the morphisms. This is the same category as k-Mod (see below).

2 For a ring R with identity, R-Mod is the category of (left) R-modules with R-linear
maps as the morphisms. In this case we will denote HomR-Mod (A, B) as HomR (A, B).

3 Groups (or Grps) is the category consisting of groups together with group homomor-
phisms. Even more generally, we can define the category Sets: objects are sets and
morphisms are just regular functions. Other algebraic categories arise in a similar fash-
ion: AbGrps or Ab is the category of abelian groups with the usual homomorphisms;
Rings is the category of rings with ring homomorphisms, where we assume all rings
have unity; likewise, ComRings are commutative rings.

4 For a topological space X, we can form a category by taking the open subsets of X as
objects and for any open U, V ⊂ X,
(
∅ if U 6⊆ V
HomX (U, V ) =
{iU,V } if U ⊆ V

where iU,V is the inclusion U ,→ V . This category can be generalized to any space X
which has a partial ordering .

5 As another topological example, Top is the category of topological spaces together with
continuous functions between them.
How do we compare two categories?
Definition. A functor F from C to D is an assignment

F : obj(C) −→ obj(D) HomC (A, B) −→ HomD (F (A), F (B))


A 7−→ F (A) f :A→B F (f ) : F (A) → F (B)

such that F (gf ) = F (g)F (f ).

120
7.1. Categories and Functors Chapter 7. Category Theory

Remark. The functor defined above is sometimes called a covariant functor: the image of
a morphism f : A → B is a morphism F (A) → F (B). On the other hand, a contravariant
functor takes f : A → B to F (f ) : F (B) → F (A), and F (gf ) = F (f )F (g).

Examples.

1 The forgetful functor is useful in many contexts. For example,

Forget : Veck −→ Ab
V 7−→ (V, +)
f 7−→ f

takes a vector space V and “forgets” its vector space structure. We can go further and
forget a group structure, giving a functor Ab → Sets.

2 For any category C, the identity functor 1C : C → C takes A 7→ A and f 7→ f .

3 In topology, the fundamental group is a functor:

π1 : Top −→ Grps

X 7−→ π1 (X)

f :X→Y f ∗ : π1 (X) → π1 (Y ).

3 The nth homology group of a topological space is a functor Hn (·) : Top → Ab.

4 Let C be the category of field extensions K/k with morphisms ϕ : K → L such that
ϕ|k = idk , i.e. the set of k-homomorphisms. Then Aut is the functor defined by

Aut : C −→ Grps
K/k 7−→ Aut(K/k).

Another category can be formed by taking obj(C) to be all Galois extensions K/k.
Then all automorphism groups are called Galois groups, and the functor is

Gal : C −→ Grps

K/k 7−→ Gal(K/k)

ϕ:K→L Gal(ϕ) : Gal(L/k) → Gal(K/k)

where the last map is the quotient map. Notice that both of these are contravariant
functors.

121
7.1. Categories and Functors Chapter 7. Category Theory

5 For the category Veck we have the contravariant functor

( )∗ : Veck −→ Veck

V 7−→ Homk (V, k)

f :V →W f∗ : W∗ → V ∗

g 7→ gf

6 Abelianization is a functor:

(·)ab : Groups −→ Ab

G 7−→ G/G0 where G0 is the commutator subgroup


f ¯
f.

The main functors of interest in this course will be for the categories R-Mod:

ˆ HomR (−, N ), a contravariant functor

ˆ HomR (M, −), a covariant functor

ˆ − ⊗R N , a covariant functor.

It is usually better to think of everything in terms of covariant functors. To explain why,


we have the following generalization of a concept from ring theory.

Definition. Given a category C, its opposite category C op consists of obj(C op ) = obj(C)


and HomC op (A, B) = HomC (B, A).

Then a contravariant functor from C is simply a covariant functor from C op . For example,
we see that the category of left R-modules is isomorphic (to be defined) to the category of
right Rop -modules, where Rop = R as sets and a ·Rop b = b ·R a.

Proposition 7.1.1. If T : A → B is a covariant functor, define T op : Aop → B by T op (A) =


T (A) for all A ∈ obj(A) and T op (f op ) = T (f ) for all morphisms in A. Then T op is a
contravariant functor.

Proof. First, if C ∈ obj(Aop ), then T op (C) = T (C) ∈ obj(B).


Next, T op (f op ) : T op (A) → T op (A0 ) since T op (f op ) = T (f ) : T (A) → T (A0 ), and these
objects are equal in both maps.
Now to show that T op composes, take f : A → A0 and g : A0 → A00 so that f op : A0 → A
and g op : A00 → A0 . Then

T op (g op ) : T op (A00 ) → T op (A0 ) and T op (f op ) : T op (A) → T op (A0 )

122
7.1. Categories and Functors Chapter 7. Category Theory

so we see that
T op (g op f op ) = T op ((f g)op ) = T (f g) = T (f )T (g)
since T is covariant. But this last part is just T op (f op )T op (g op ) so we have shown that T op is
contravariant.
Finally T op (1A ) = T (1A ) = 1T (A) = 1T op (A) so in particular T op is a contravariant functor.

Just as homomorphisms compare algebraic objects and functors compare categories, there
is a way to compare functors themselves.

Definition. For covariant functors S, T : A → B, a natural transformation τ : S → T


is a family of morphisms in B:

τ = {τA : S(A) → T (A)}A∈obj(A)

such that the following diagram commutes for all morphisms f : A → A0 .


τA
S(A) T (A)

S(f ) T (f )

S(A0 ) T (A0 )
τA0

We can define natural transformations between contravariant functors in a similar way,


or just treat them as covariant functors from Aop .

Definition. A natural isomorphism is a natural transformation such that every τA is an


isomorphism.

Proposition 7.1.2. If σ : S → T and τ : T → U are natural transformations of functors


S, T, U : A → B, then τ σ : S → U defined by (τ σ)A = τA σA is a natural transformation.

Proof. Easy to check.

Example 7.1.3. For any functor T : A → B, the identity natural transformation


ωT : T → T is defined by (ωT )A = 1T (A) : T (A) → T (A).

Proposition 7.1.4. A natural transformation τ : S → T is a natural isomorphism if and


only if there is a natural transformation σ : T → S such that στ = ωS and τ σ = ωT .

Proof omitted.

Definition. For two (covariant) functors F, G : A → B, define

Nat(F, G) = {natural transformations τ : F → G}.

123
7.1. Categories and Functors Chapter 7. Category Theory

Theorem 7.1.5. If A ∼ = B then HomC (A, −) and HomC (B, −) are naturally isomorphic
functors. Similarly, HomC (−, A) ∼
= HomC (−, B).
Proof. Let A ∼= B via the isomorphism α : A → B and let α−1 be its inverse. Define a
transformation τ : HomC (A, −) → HomC (B, −) by

τC = (α−1 )∗ : HomC (A, C) −→ HomC (B, C)


f 7−→ f α−1

for each C ∈ obj(C). Note that f α−1 : B → C so this map is well-defined.


Next we verify that for any morphism f : C → C 0 , the following diagram commutes:
τC
Hom(A, C) Hom(B, C)

f f

Hom(A, C 0 ) Hom(B, C 0 )
τC 0

If h ∈ Hom(A, C) then it maps h 7→ hα−1 7→ f hα−1 via the top map, and h 7→ f h 7→ f hα−1
via the bottom map, so indeed the diagram commutes. Thus τ is a natural transformation.
Finally we need to check that each τC is an isomorphism. To do this, construct a trans-
formation σC : HomC (B, C) → HomC (A, C) taking g 7→ gα. Then for any f ∈ HomC (A, C)
and g ∈ HomC (B, C) we have

τC (σC (g)) = τC (gα) = gαα−1 = g


and
σC (τC (f )) = σC (f α−1 ) = f α−1 α = f.

Hence τC and σC are inverse morphisms, so τC is an isomorphism. This proves that τ is a


natural isomorphism. The proof that HomC (−, A) ∼
= HomC (−, B) is similar.
We will see that the converse of this statement is also true. That is, HomC (−, A) com-
pletely determines the object A itself (as does HomC (A, −)). But what’s the meaning be-
hind statements like this? For an object A ∈ obj(C), the functor HomC (−, A) : C op → Sets
gives us a lot of information about A. For example, if C is the category of groups and
HomC (Z/2Z, A) contains a nontrivial element, this tells us that A has at least one element of
order 2. Likewise, in the category Top of topological spaces, elements of Hom(S 1 , A) encode
the information of closed loops in A. So “probing” A by different objects B, i.e. considering
HomC (B, A), tells us different sorts of information about A. By probing A with all objects
in C, would it be possible to know all information about A? The answer is yes, and the
formal statement of this is called Yoneda’s Lemma.
Lemma 7.1.6 (Yoneda’s Lemma). Let C be a category, A ∈ obj(C) and G : C → Sets a
covariant functor. Then there is a bijection

Nat(HomC (−, A), G) ←→ G(A).

124
7.1. Categories and Functors Chapter 7. Category Theory

Proof. Let τ : HomC (−, A) → G be a natural transformation. Then for each B ∈ obj(C) and
morphism ϕ : A → B, the following diagram commutes:
τA
Hom(A, A) G(A)

ϕ∗ G(ϕ)

Hom(B, A) G(B)
τB

where ϕ∗ : f 7→ ϕ ◦ f . In particular, 1A ∈ HomC (A, A) maps to τA (1A ) ∈ G(A). Define


the map Nat(HomC (−, A), G) → G(A) by sending τ 7→ τA (1A ). If σ : HomC (−, A) → G is
another natural transformation such that σA (1A ) = τA (1A ), then the above diagram shows
that for all ϕ ∈ HomC (B, A),

σB (ϕ) = σB ϕ∗ (1A ) = G(ϕ)τA (1A ) = G(ϕ)σA (1A ) = τB ϕ∗ (1A ) = τB (ϕ).

Thus σ = τ , so the assignment is one-to-one.


We next show every element x ∈ G(A) is induced by such a natural transformation
τ : HomC (−, A) → G. For each B ∈ obj(C) and f ∈ HomC (B, A), define τB (f ) = G(f )(x).
Then for any B → C, the following diagram commutes:
τB
Hom(A, B) G(B)

Hom(A, C) G(C)
τC

Hence τ is a natural transformation and by construction, τA (1A ) = G(1A )(x) = 1G(A) (x) = x.
This proves the bijection.
Corollary 7.1.7. If HomC (−, A) and HomC (−, B) are naturally isomorphic, then A ∼
= B.
Proof. Suppose τ : HomC (−, A) → HomC (−, B) is a natural isomorphism. Applying Yoneda’s
Lemma to G = HomC (−, B), we get a unique element x ∈ G(A) = HomC (A, B). Applying
Yoneda to the inverse of τ yields a unique inverse to x, proving A ∼
= B.
For any category C, there is a functor C → Fun(C op , Sets) given by A 7→ HomC (−, A).
The Yoneda Lemma shows that this functor is fully faithful, i.e. it embeds C as a full
subcategory of Fun(C op , Sets). More precisely:
Corollary 7.1.8 (Yoneda Embedding). There is a functor C → Fun(C op , Sets), A 7→ HomC (−, A),
which is an isomorphism on HomC (B, C) for all B, C ∈ C.
Lemma 7.1.9 (Yoneda’s Lemma, Covariant Version). Let C be a category, A ∈ obj(C) and
G : C → Sets a covariant functor. Then there is a bijection

Nat(HomC (A, −), G) ←→ G(A)


τ 7−→ τA (1A ).

125
7.1. Categories and Functors Chapter 7. Category Theory

Proof. Similar to the proof of Lemma 7.1.6.

Corollary 7.1.10 (Yoneda Embedding, Covariant Version). There is a fully faithful functor
C → Fun(C, Sets), A 7→ HomC (A, −).

126
7.2. Exactness of Sequences and Functors Chapter 7. Category Theory

7.2 Exactness of Sequences and Functors


f g
Definition. A sequence A →
− B→
− C in a category C is exact at B if ker g = im f .

Definition. An exact sequence of the form


f g
0→A→
− B→
− C→0

is called a short exact sequence. In particular, this means that f is injective, g is surjective
and ker g = im f . This is sometimes called an “extension of A by C”.
f g
If 0 → A →
− B→ − C → 0 is a short exact sequence of R-modules, the first isomorphism
theorem for modules tells us that A ∼
= im f and B/ im f ∼
= C.
Example 7.2.1. For a commutative ring R, take the set S of all R-module short exact
sequences:
S = {0 → A → B → C → 0 exact | A, B, C ∈ R-Mod}
Then there is an equivalence relation ∼ such that

S/ ∼∼
= Ext1R (C, A).

Definition. Let 0 → A → B → C → 0 be an exact sequence in a category C and let


F : C → D be a covariant functor. F is left exact if

0 → F (A) → F (B) → F (C)

is exact, and F is right exact if

F (A) → F (B) → F (C) → 0

is exact. If it is both left and right exact, then F is simply called exact.

Note that if F is a contravariant functor, left exactness looks like

0 → F (C) → F (B) → F (A)

and right exactness is analogous.

Proposition 7.2.2. HomR (X, −) is a left exact functor on the category R-Mod.
f g
Proof. Let 0 → A → − B → − C → 0 be an exact sequence of left R-modules and take
X ∈ obj(R-Mod). If we apply HomR (X, −), the sequence becomes
f∗ g∗
0 → HomR (X, A) −
→ HomR (X, B) −
→ HomR (X, C).

We need to show that f∗ is one-to-one and ker g∗ = im f∗ .

127
7.2. Exactness of Sequences and Functors Chapter 7. Category Theory

First, take h ∈ HomR (X, A) such that f h = 0. Note that f is one-to-one ⇐⇒ there
exists a morphism f˜ : im f → A such that f˜f = idA . In this case
h = idA h = f˜f h = f˜0 = 0
which shows that h = 0, i.e. f∗ is one-to-one.
Now let j ∈ HomR (X, B) such that gj = 0, i.e. j ∈ ker g∗ . We need to construct a
function k : X → A such that f k = j. Let x ∈ X. Then g(j(x)) = 0 so j(x) ∈ ker g. But
ker g = im f so there exists a (unique) element a ∈ A such that f (a) = j(x). Now we can
define k : X → A by mapping x to this unique a. Then j(x) = f (a) = f (k(x)) so j = f k as
desired. Hence HomR (X, −) is left exact.
Note that HomR (X, −) may not always be right exact. In fact, even in the category of
abelian groups, HomZ (X, −) is not right exact, as the following example shows.
Example 7.2.3. Consider the exact sequence
0 → 2Z → Z → Z/2Z → 0.
Let X = Z/2Z. Then Proposition 7.2.2 tells us that
0 → Hom(Z/2Z, 2Z) → Hom(Z/2Z, Z) → Hom(Z/2Z, Z/2Z)
is exact. But up to isomorphism, this sequence is
0 → 0 → 0 → Z/2Z.
Clearly adding a zero on the right makes the sequence not exact, since the kernel of Z/2Z → 0
is necessarily Z/2Z.
Two important examples in the theory of modules and their homological properties are
Definition. An R-module P is projective if HomR (P, −) is exact.
Definition. An R-module E is injective if HomR (−, E) is exact.
The proof of Proposition 7.2.2 may be adapted to show that the contravariant functor
HomR (−, Y ) is always left exact, so proving that a module is projective or injective comes
down to proving that the associated Hom is right exact. Example 7.2.3 shows that Z/2Z is
not a projective Z-module. We will explore these special modules in Chapter 8.
Proposition 7.2.4. P is a projective module if and only if for every surjection f : A → B,
f∗ : HomR (P, A) → HomR (P, B) is surjective. This in turn is equivalent to saying for every
diagram of the form shown below, where the bottom row is exact, there exists a g̃ that makes
the diagram commute.
P

g

A B 0
f

128
7.2. Exactness of Sequences and Functors Chapter 7. Category Theory

Theorem 7.2.5 (Hom-Tensor Adjointness). Given rings R and S and modules AR ,R BS and
CS , there is an isomorphism

τA,B,C : HomS (A ⊗R B, C) −→ HomR (A, HomS (B, C))


f 7−→ [a 7→ (b 7→ f (a ⊗ b))]

i.e. given any f : A ⊗R B → C and an element a ∈ A, we can construct a linear map


fa : B → C sending b 7→ f (a ⊗ b).

The term adjointness comes from the idea that Hom may be viewed as an inner product
on R-Mod:
h·, ·i = HomR (−, −) :R Mod ×R Mod −→ Ab
which has ⊗R as its adjoint functor. In particular, we have functors

FB := − ⊗R B : ModR −→ ModS
GB := HomS (B, −) : ModS −→ ModR .

Then Theorem 7.2.5 says that FB and GB are adjoint functors. FB is sometimes called the
left adjoint and GB the right adjoint, and together they are called an adjoint pair.
Notice that the above setup uses right actions on A ⊗R B and HomS (B, C). There is a
similar setup involving left actions, which is not discussed here.

Proposition 7.2.6. − ⊗R B : ModR → Ab is a right exact functor.

Proof. Given f : A → A0 a homomorphism of right R-modules, let ϕ : A × B → A0 ⊗ B be


the map sending (a, b) 7→ f (a) ⊗ b. This map is biadditive:

ϕ(a + a0 , b) = f (a + a0 ) ⊗ b = (f (a) + f (a0 )) ⊗ b = f (a) ⊗ b + f (a0 ) ⊗ b = ϕ(a, b) + ϕ(a0 , b).

(The other properties are proven similarly.) By the universal mapping property, this induces
a linear map

f ⊗ 1 : A ⊗R B 7−→ A0 ⊗R B
a ⊗ b 7−→ f (a) ⊗ b.

Therefore − ⊗R B : ModR → Ab is a functor. The proof that A ⊗R − is a functor is similar.


To prove that − ⊗R B is right exact, take an exact sequence
i p
B0 → − B 00 → 0
− B→

of left R-modules. We need to show that the induced sequence is exact:


1⊗i 1⊗p
A ⊗ B 0 −−→ A ⊗ B −−→ A ⊗ B 00 → 0.

It is clear that im(1 ⊗ i) ⊆ ker(1 ⊗ p) because pi = 0, but it is not as obvious in the other
direction. We will prove exactness at A ⊗ B another way in a moment.

129
7.2. Exactness of Sequences and Functors Chapter 7. Category Theory

X
To show 1 ⊗ p is onto, let ai ⊗ b00i ∈ A ⊗ B 00 . By exactness of the original sequence at
B 00 , for each b00i ∈ B 00 there exists some bi ∈ B such that p(bi ) = b00i . Then
X  X X
(1 ⊗ p) ai ⊗ b i = ai ⊗ p(bi ) = ai ⊗ b00i .

Hence 1 ⊗ p is onto.
Finally, notice that if we prove that (A ⊗ B)/ im(1 ⊗ i) is a solution to the universal
mapping property in the definition of A ⊗ B 00 then this will show that (A ⊗ B)/ im(1 ⊗ i) ∼
=
00
A ⊗ B which implies exactness at A ⊗ B. Given a biadditive map ϕ : A × B → G, we need
to produce a biadditive map ψ and a linear map ϕ̄ making the following commute:
A × B 00
ϕ
ψ

(A ⊗ B)/ im(1 ⊗ i) G
ϕ̄

Define ψ(a, b00 ) = (a ⊗ b) + im(1 ⊗ i) where p(b) = b00 . We need to be carefull about defining
maps out of A × B 00 . In particular, if p(b1 ) = p(b2 ), we must show that (a ⊗ b1 ) + im(1 ⊗ i) =
(a⊗b2 )+im(1⊗i). But p(b1 ) = p(b2 ) implies there exists some b0 ∈ B 0 such that i(b0 ) = b1 −b2 .
Tensoring with a yields

(a ⊗ b1 ) − (a ⊗ b2 ) = a ⊗ (b1 − b2 ) = a ⊗ i(b0 )

which lies in im(1 ⊗ i) by exactness. Hence ψ is well-defined. To check that ψ is biadditive,

ψ(a + a0 , b00 ) = (a + a0 ) ⊗ b + im(1 ⊗ i)


= (a ⊗ b) + (a0 ⊗ b) + im(1 ⊗ i)
= ((a ⊗ b) + im(1 ⊗ i)) + ((a0 ⊗ b) + im(1 ⊗ i))
= ψ(a, b00 ) + ψ(a0 , b00 ).

The rest is proven similarly.


Now we need to produce ϕ̄ : (A ⊗ B)/ im(1 ⊗ i) → G. We will first define a map
A × B → G by (a, b) 7→ ϕ(a, p(b)); it is easy to see that this is biadditive. This induces a
linear map Φ : A ⊗ B → G. Moreover, for any b0 ∈ B 0 ,

Φ(a ⊗ i(b0 )) = ϕ(a, pi (b0 )) = ϕ(a, 0) = 0

which shows that ker Φ = im(1 ⊗ i). Therefore we may define the quotient map out of A ⊗ B:

ϕ̄ : (A ⊗ B)/ im(1 ⊗ i) −→ G

which inherits linearity from Φ. This completes the proof of right exactness.
Example 7.2.7. For an integer n > 0,
n
0→Z→
− Z → Z/nZ → 0

130
7.2. Exactness of Sequences and Functors Chapter 7. Category Theory

is a short exact sequence. Tensoring with the Z-module Z/nZ gives a sequence
n
Z/nZ →
− Z/nZ → Z/nZ ⊗ Z/nZ → 0.

However, multiplication by n is the zero map on Z/nZ, so this shows that − ⊗ Z/nZ is not
left exact. As a bonus, the tensored exact sequence splits into

0 → Z/nZ → Z/nZ ⊗ Z/nZ → 0.

So Z/nZ ⊗ Z/nZ ∼
= Z/nZ as abelian groups.
Definition. A short exact sequence of R-modules
f g
0→A→
− B→
− C→0

is split if there exists a map h : C → B with gh = 1C .

Proposition 7.2.8. If a short exact sequence of R-modules


f g
0→A→
− B→
− C→0

is split, then B ∼
= A ⊕ C.
Proof. We will show that B = im f ⊕ im h where h : C → B satisfies gh = 1C . If b ∈ B,
then g(b) ∈ C and b − h(g(b)) ∈ ker g. To see this, note that

g(b − h(g(b))) = g(b) − g(h(g(b))) = g(b) − g(b) = 0

since gh = 1C . By exactness of the sequence, there exists an element a ∈ A such that


f (a) = b − h(g(b)). Then B = im f + im h. To show the product is direct, it suffices to
show that im f ∩ im h = {0}. Suppose x = f (a) = h(c). Then g(x) = g(f (a)) = 0 because
gf = 0. On the other hand, g(x) = g(h(c)) = c because gh = 1C . Hence x = 0, and so we
have proven B ∼= A ⊕ C.

Lemma 7.2.9 (3×3 Lemma). Consider the following commutative diagram in R-Mod having
exact columns.

131
7.2. Exactness of Sequences and Functors Chapter 7. Category Theory

0 0 0

0 A0 A A00 0

0 B0 B B 00 0

0 C0 C C 00 0

0 0 0

If the bottom two rows are exact, the top row is exact, and if the top two rows are exact, the
bottom row is exact.

Proof. Picking a00 ∈ A00 , let b00 be its image in B 00 . Then b00 must map to zero by exactness
of the third column. Thus there is a b ∈ B mapping to b00 since the map is surjective; let c
be the image of b in C. By commutativity, c 7→ 0. Therefore there is some c0 ∈ C 0 mapping
to c, using exactness of the third row.
Now by surjectivity of B 0 → C 0 , there is some b0 ∈ B 0 mapping to c0 ; denote its image in
B by b̄. By exactness, this b̄ maps to 0 in B 00 . This further implies that b − b̄ 7→ b00 . Since
the lower left square commutes, b̄ 7→ c =⇒ b − b̄ 7→ 0 =⇒ for some a ∈ A, a 7→ b − b̄. Let
a map to ā00 ∈ A00 . Then since A00 ,→ B 00 is injective, ā00 = a00 so it is guaranteed that there
will be an a mapping to a00 . Hence A → A00 is surjective.
Next, take some a0 ∈ A0 which maps to 0 in A. Then a0 7→ b0 ∈ B 0 and by commutativity
around the square, b0 7→ 0 ∈ B. But A0 ,→ B 0 is injective, so a0 = 0 and A0 → A is seen to
be injective.
Now pick another a0 ∈ A0 with image a ∈ A and subsequently a 7→ x ∈ A00 . We need to
show x = 0. Well a0 maps to some b0 ∈ B 0 and b0 7→ b ∈ B, so by commutativity a 7→ b. Since
b is in the image of B 0 → B and the second row is exact, b 7→ 0 ∈ B 00 . By commutativity of
the upper right square, x 7→ 0 ∈ B 00 . But since this map is injective, this forces x = 0.
We have thus shown that the top row is exact. The second statement is proven similarly.

The proof above is an example of a variety of proofs which are common to homological
algebra, called diagram chasing. Another example of this is

Lemma 7.2.10. Consider the following commutative diagram in R-Mod having exact rows
and columns.

132
7.2. Exactness of Sequences and Functors Chapter 7. Category Theory

A0 A A00 0

B0 B B 00 0

C0 C C 00 0

0 0 0

If A00 → B 00 and B 0 → B are injections, then C 0 → C is an injection. Similarly, if C 0 → C


and A → B are injections, then A00 → B 00 is an injection. In other words, if the last column
and the second row are short exact sequences, then the third row is a short exact sequence,
and similarly, if the bottom row and the second column are short exact sequences, then the
third column is a short exact sequence.

Proof. First suppose A00 ,→ B 00 and B 0 ,→ B are injections. The following diagram serves as
a useful blueprint for the diagram chase to follow.

A0 (7) A A00 0
0 a 00
a a =0 (6)

(8)
(4)
0
β
B0 B B 00 0
=

0 0
(9) b (3)
b (5)

(2)

C0 (1) C C 00 0
0 0
c

0 0 0

Take c0 ∈ C 0 such that c0 7→ 0 in C – we must show c0 = 0. By exactness of the first


column, there is a b0 ∈ B 0 such that b0 7→ c0 . Let b be the image of b0 under B 0 ,→ B; the
diagram commutes, so b0 7→ c0 7→ 0 ∈ C implies b0 7→ b 7→ 0 ∈ C. Therefore b is in the
kernel of B → C, and by exactness of the middle column, there exists some a ∈ A such that
a 7→ b in B. Denote the image of a along A → A00 by a00 . We also have b0 7→ b 7→ 0 since the
middle row is exact, and commutativity gives us a 7→ a00 7→ 0 along the top. But A00 ,→ B 00
is injective, so a00 = 0. Thus a is in the kernel of A → A00 so by exactness of the top row,
there is some a0 ∈ A0 such that a0 7→ a ∈ A. Let the image of a0 under A0 → B 0 be denoted
β 0 . Then since a0 7→ a 7→ b ∈ B, β 0 is also mapped to b. But B 0 ,→ B is injective, so we must

133
7.2. Exactness of Sequences and Functors Chapter 7. Category Theory

have β 0 = b0 . This shows that b0 is in the image of A0 → B 0 , so it’s in the kernel of B 0 → C 0


and hence b0 7→ c0 = 0. This completes the proof that C 0 → C is injective.
The second statement follows from an identical proof if we reflect the diagram across
the line running through A0 , B and C 00 , since this transformation preserves the homological
properties of the arrows.
Note that the statements about short exact sequences follow immediately from the proofs
of injectivity.
One of the most famous diagram chasing proofs is the Snake Lemma, which will become
useful when we define complexes in Chapter 10. The version below involves a morphism in
the category of short exact sequences.

Lemma 7.2.11 (Snake Lemma). If the commutative diagram

0
α0 α
0 A A A00 0

f g h
0
β β
0 B0 B B 00 0

has exact rows, then there is an exact sequence

0 → ker f → ker g → ker h → coker f → coker g → coker h → 0.

Proof. To begin, note that the left square commutes so α0 restricts to α0 : ker f → ker g.
Likewise α restricts to α : ker g → ker h and these both inherit injectivity. Consider the
β0
projection B 0 − → B → B/ im g = coker g. For any b0 ∈ im f , there is some a0 ∈ A0 such that
f (a0 ) = b0 , but also gα0 (a0 ) = β 0 (b0 ) by commutativity. So β 0 (b0 ) ∈ im g which induces a map
β̄ 0 : coker f → coker g – in other words β 0 factors through f . Likewise, β induces a map β̄
and both β̄ 0 and β̄ inherit surjectivity. We pause to label the sequence:
α0 α ð β̄ 0 β̄
0 → ker f −
→ ker g −
→ ker h →
− coker f −
→ coker g →
− coker h → 0.

The interesting part of the sequence is at ker h → coker f ; we have labelled this with the
character ð. How can we define ð? Start with a00 ∈ ker h, so that h(a00 ) = 0. By exactness,
there is some a ∈ A such that α(a) = a00 . If b is the image of a under g, b 7→ 0 in B 00 and so
there is some b0 ∈ B 0 such that b0 7→ b 7→ 0.
With this notation, we define

ð : ker h −→ coker f
a00 7−→ b0 + im f.

Since all maps in this diagram are linear, it suffices to check well-definedness. Once we do
so, we will have all maps defined and can then proceed to check exactness.
Suppose we have α(a1 ) = α(a2 ) = a00 . Then there is some a0 ∈ A0 such that α0 (a0 ) = a1 −a2
since this difference lies in ker α. Let b01 and b02 be the unique (by injectivity) elements of B 0

134
7.2. Exactness of Sequences and Functors Chapter 7. Category Theory

such that β 0 (b01 ) = g(a1 ) and β 0 (b02 ) = g(a2 ). We want to show that b01 − b02 = f (a0 ). But this
is easily seen, since commutativity of the left square gives us a0 7→ a1 − a2 7→ g(a1 ) − g(a2 ) :=
b1 − b2 around the top, and a0 7→ f (a0 ) 7→ b1 − b2 around the bottom. Then by injectivity,
f (a0 ) must equal b01 − b02 as desired. This shows that ð maps into B 0 / im f = coker f , i.e. it
is well-defined.
We now have all our maps, so we can proceed to show exactness. First α0 α = 0 implies
im(α0 |ker f ) ⊆ ker(α |ker g ). For the other containment, take a ∈ ker α that maps to 0 under
g. Then by exactness, there exists an a0 7→ a but the left square commutes, so a0 7→ b0 7→ 0
and injectivity of β 0 shows b0 = 0.
Next suppose b ∈ coker g maps to b00 ∈ coker h = B/ im h. Then there is some a00 mapping
to b00 under h, and α is onto to there also exists a ∈ A mapping to a00 . Let g(a) = b̃ ∈ B;
by commutativity b̃ 7→ b00 . Then b̃ − b 7→ 0 ∈ B 00 implies there exists b0 ∈ B 0 such that
β 0 (b0 ) = b̃ − b = g(a) − b. So b = β 0 (b0 ) − g(a) =⇒ β̄ 0 (b0 + im g) = β 0 (b0 ) + im g = b + im g.
The hard part is showing exactness around the connecting map ð. For a00 ∈ ker h, we can
lift back to some a ∈ A which maps to a00 ∈ A00 and has image g(a) = b. But b 7→ 0 under
β so there is some b0 ∈ B 0 such that b0 7→ b. Recall that we defined ð(a00 ) = b0 + im f . Then
im(α |ker g ) ⊂ ker ð because g(a) = 0. On the other hand, suppose ð(a00 ) = 0. Then in the
diagram chase for the definition of ð, choose a0 ∈ A0 such that f (a0 ) = b0 . Set ã = α0 (a0 ). By
commutativity, g(ã − a) = 0 and by exactness, α(ã − a) = α(a) = a00 .
Finally let b0 ∈ B 0 such that β̄ 0 (b0 + im f ) = im g, so β 0 (b0 ) ∈ im g. The fact that
im ð ⊂ ker β̄ 0 is just by definition of the connecting map. Conversely, suppose b0 7→ b such
that there is some a ∈ A with g(a) = b. If we set α(a) = a00 , we have already constructed
the connecting map for a00 . Moreover we know b0 7→ b 7→ 0 so a00 must map to 0 along h, i.e.
a00 ∈ ker h. This shows exactness at every part of the sequence.

Lemma 7.2.12 (Five Lemma). Suppose the following diagram is commutative with exact
rows.
A1 A2 A3 A4 A5

f1 f2 f3 f4 f5

B1 B2 B3 B4 B5

(i) If f2 and f4 are surjective and f5 is injective, then f3 is surjective.

(ii) If f2 and f4 are injective and f1 is surjective, then f3 is injective.

(iii) If f1 , f2 , f4 and f5 are isomorphisms, then so is f3 .

Proof. Commence chase.

135
Chapter 8

Special Modules

136
8.1. Projective Modules Chapter 8. Special Modules

8.1 Projective Modules


We mentioned projective modules in Section 7.2 as they relate to the covariant Hom functor,
but here we will define them in a more module-oriented context. We will see shortly that
the two notions coincide.

Definition. A left R-module P is projective if for every diagram with exact rows and solid
arrows, shown below, there is a dashed arrow making the diagram commute.

M N 0

In other words, projectives allow you to lift along surjections. We will prove

Theorem 8.1.1. Over a ring R, a left R-module P is projective if and only if it is a direct
summand of a free R-module.

Example 8.1.2. As a Z/6Z-module, Z/6Z is isomorphic to Z/2Z ⊕ Z/3Z which shows


that Z/2Z and Z/3Z are projective Z/6Z-modules. However, Z/2Z and Z/3Z are not free
(6 kills everything in both submodules) and therefore are not projective in the category of
Z/6Z-modules.

A long-standing open question, proposed by Serre and proven by Quillen and Suslin, is:
If P is a projective module over k[x1 , . . . , xn ] then must P be free?
We now show that the earlier definition of projectives agrees with our definition in this
chapter.

Proposition 8.1.3. P is projective ⇐⇒ HomR (P, −) is exact.

Proof. We proved HomR (P, −) is always left exact. For a sequence


α0 α
0 → M0 − → M 00 → 0
→M −

applying HomR (P, −) induces an exact sequence with solid arrows:

0 → Hom(P, M 0 ) → Hom(P, M ) → Hom(P, M 00 ) 99K 0.

To get the 0 on the right, note that by definition of projective, every f : P → M 00 factors
through α : M → M 00 :
P

f
α
M 00 M 0

137
8.1. Projective Modules Chapter 8. Special Modules

Notice that is equivalent to showing Hom(P, M ) → Hom(P, M 00 ) is surjective. The converse


follows by reversing the entire argument.
Perhaps the most important example of a projective module is a free module. This is
highly useful since in the future when we wish to build modules out of projectives, we may
instead use free modules which are easier to work with.
Proposition 8.1.4. Every free module is projective.
Proof. Suppose F is a free module with basis {fi }i∈I . Let α : M → N be a surjection, i.e.
the row of the following diagram with solid arrows is exact.
F
α̃ ϕ
α
M N 0

If ϕ : F → N is an R-linear map, denote ϕ(fi ) = ni ∈ N . Since α is surjective, there exists


an mi ∈ M such that α(mi ) = ni . Then we will define α̃(fi ) = mi and extend by linearity
to all of F . It is clear that αα̃ and ϕ agree on {fi } which is a basis for F ; therefore the
diagram commutes.
An interesting example of projective modules arises from localization in rings:
Proposition 8.1.5. For a commutative ring R and P a finitely generated R-module, P is
projective if and only if for all prime ideals p ⊂ R, the localization Pp is free.
A useful property of projectives is
Proposition 8.1.6. P is projective ⇐⇒ every short exact sequence ending in P splits.
Proof. ( =⇒ ) Suppose we have a short exact sequence

0 −→ M −→ N −→ P −→ 0.

where P is projective. The identity on P induces a surjection f :


P
f
id
0 M N P 0

Hence the sequence splits.


( ⇒= ) Conversely, take a finite presentation of P :

0 −→ K −→ F −→ P −→ 0

where F is free and the map F → R takes a generating set {fi } of F to a generating set
{pi } of P . By hypothesis the sequence splits, so F ∼
= P ⊕ K, and by the earlier (unproven)
characterization of projectives, P is projective.

138
8.1. Projective Modules Chapter 8. Special Modules

In the above proof we used the fact that direct summands of free modules are projective.
We will now prove an even stronger result which implies this.
Theorem 8.1.7. In the category of R-modules for a ring R,
(1) Direct summands of projective modules are projective.

(2) Direct sums of projective modules are projective.


Proof. (1) Suppose a projective module P can be written P = P 0 ⊕ P 00 . Consider the
following diagram with the bottom row exact.
P

π i

f˜ P0

g

M N 0
f

Given g surjective, we want to construct a map g̃ making the smaller triangle commute.
Note that P = P 0 ⊕ P 00 induces π : P → P 0 the natural projection and i : P 0 ,→ P the
inclusion map. Since P is projective, there exists a map f˜ : P → M making the larger
triangle commute. It suffices to show that g̃ = f˜i is what we are looking for, but this is
obvious since for any p0 ∈ P 0 ,

f f˜i(p0 ) = f f˜(p0 ) = g(p0 ).

Hence P 0 is projective.
(2) If P 0 and P 00 are projective then HomR (P 0 , −) and HomR (P 00 , −) are exact functors.
Thus we have a natural isomorphism of functors

Hom(P 0 ⊕ P 00 , −) ∼
= Hom(P 0 , −) ⊕ Hom(P 00 , −).

Since direct sums of exact functors are exact, this shows Hom(P 0 ⊕ P 00 , −) is exact and
therefore P 0 ⊕ P 00 is projective. In the infinite case, note that
!
Ai , B ∼
M Y
Hom = Hom(Ai , B).
i∈I i∈I

Then the result follows from the fact that direct products of exact functors are exact.
Corollary 8.1.8.
(a) Every module over a field is projective.

(b) If R is a PID, then every projective R-module is also free.

139
8.1. Projective Modules Chapter 8. Special Modules

Proof. (a) A module over a field is a vector space which is free, hence projective by Propo-
sition 8.1.4.
(b) For modules over a PID, every submodule of a free module is free.
Projective modules have many desirable properties.

Proposition 8.1.9. Every projective left R-module P has a free complement, that is, a free
left R-module F such that P ⊕ F is free.

Proof. Since P is projective, it’s the direct summand of a free module, i.e. there exists a Q
so that P ⊕ Q is free. The trick is to consider

F = Q ⊕ P ⊕ Q ⊕ ···

which is free since each pair in the sum is free. But notice that P ⊕ F = P ⊕ Q ⊕ P ⊕ Q ⊕ · · ·
is free for the same reasons; hence P has a free complement.

Proposition 8.1.10. For a commutative ring R and projective R-modules P and Q, P ⊗R Q


is a projective R-module.

Proof. Since P, Q are projective, there exist free modules F1 , F2 and modules G1 , G2 such
that
F1 = P ⊕ G1 and F2 = Q ⊕ G2 .
Tensoring these two expressions gives us

F1 ⊗ F2 = (P ⊕ G1 ) ⊗ (Q ⊕ G2 ) = (P ⊗ Q) ⊕ (P ⊗ G2 ) ⊕ (G1 ⊗ Q) ⊕ (G1 ⊕ G2 ).

Hence it suffices to prove F1 ⊗ F2 is free which will imply P ⊗ Q is projective. This is easily
seen from the fact that tensor distributes across direct sums (we already used this above):
! !
M M M
F1 ⊗ F2 = R ⊗ R = (R ⊗ R).
i j i,j

Recall that a left R-module M is finitely generated if M is the quotient of a free left
R-module F ∼
= Rn for n < ∞. We may then write M ∼ = F/K where K is a submodule of F .
Definition. A left R-module M is finitely presented if it is finitely generated and M ∼
=
F/K where K is finitely generated.

This can be summarized with an exact sequence

0 −→ K −→ Rn −→ M −→ 0.

If M is finitely presented we will instead write

Rm −→ Rn −→ M −→ 0.

140
8.1. Projective Modules Chapter 8. Special Modules

Proposition 8.1.11. Every finitely generated projective left R-module P is finitely presented.

Proof. Since P is finitely generated, we have an exact sequence

0 −→ K −→ F −→ M −→ 0

where F is free. P projective implies the sequence splits, so F ∼


= P ⊕ K. Then we see that
K is finitely generated since it is a direct summand, and hence an image, of F a finitely
generated module. Therefore P is finitely presented.
Finitely generated modules are the most important type of module in any module cate-
gory. In the next section we will see that finitely generated modules are intimately related
to conditions on chains of submodules.

141
8.2. Injective Modules Chapter 8. Special Modules

8.2 Injective Modules


There is another special type of module that shows up frequently in homological algebra.

Definition. A left R-module E is injective if for every diagram with exact rows and solid
arrows, shown below, there is a dashed arrow making the diagram commute.

0 A B

Notice the difference between the diagrams for projectives and injectives. We remarked
that projectives allow you to lift along surjections. Similarly, injectives allow you to extend
injections.
In Section 7.2 we defined injectives in terms of exactness of the contravariant Hom functor.
We now prove that these two definitions coincide.

Proposition 8.2.1. A left R-module E is injective ⇐⇒ HomR (−, E) is exact.

Proof. Suppose we have an exact sequence


i p
0→A→
− B→
− C → 0.

We must show that the sequence


p∗ i∗
0 → HomR (C, E) −
→ HomR (B, E) −
→ HomR (A, E) → 0

is exact. We proved in Proposition 7.2.2 that HomR (−, E) is always left exact, so it remains
to show exactness at HomR (A, E). In other words we must prove i∗ is surjective if and
only if i is injective. On one hand, if f ∈ Hom(A, E) there exists g ∈ Hom(B, E) with
f = i∗ (g) = gi; that is,
E

f g

0 A B
i

commutes, showing E is injective. Conversely, if E is injective then for any f : A → E there


exists g : B → E making the above diagram commute. Then we see that f = gi = i∗ (g) ∈
im i∗ so i∗ is surjective. This proves Hom(−, E) is an exact functor.
Similar to projectives, we have

Proposition 8.2.2. E is injective ⇐⇒ every short exact sequence beginning with E splits.

142
8.2. Injective Modules Chapter 8. Special Modules

Proof. ( =⇒ ) Consider the diagram


E

id g

0 E B
i

Since E is injective, the identity on E induces g : B → E such that gi = idE . Hence the
sequence splits.
( ⇒= ) Suppose E → M → M 00 → 0 is exact, so that by hypothesis M ∼ = E ⊕ M 00 .
We will see shortly that direct summands of injectives are injective, so it follows that E is
injective.
There is a useful characterization of injectives that allows us to restrict our attention to
ideals of R, rather than all module injections A ,→ B. Before stating and proving Baer’s
Criterion, we need

Zorn’s Lemma. Let X be a poset. If every chain in X has an upper bound in X then X
has maximal elements.

Zorn’s Lemma is of immense consequence in set theory, for it is logically equivalent to


the Axiom of Choice. Zorn’s Lemma is often invoked to prove the existence of maximal
ideals in a ring, and in particular a non-noetherian ring since the ascending chain condition
guarantees the existence of maximal ideals when R is noetherian. We will sometimes say a
set is Zornable (not an official term) if it satisfies the conditions of Zorn’s Lemma.

Theorem 8.2.3 (Baer’s Criterion). Let E be a left R-module. E is left injective ⇐⇒ every
R-linear map f : I → E, where I is an ideal of R, can be extended to R:

f g

0 I R
i

Proof. ( =⇒ ) is obvious by the definition of injectives. To prove ( ⇒= ) suppose


E

0 A B
i

143
8.2. Injective Modules Chapter 8. Special Modules

has an exact row. Let X = {(A0 , g 0 ) | A ⊆ A0 ⊆ B and g 0 extends f }. We define a partial


ordering on X by (A0 , g 0 )  (A00 , g 00 ) if A0 ⊆ A00 and g 00 extends g 0 . By construction X is
0 0 0 0
bounded above, so this makes [ X into a Zornable set. Suppose (A1 , g1 )  (A2 , g2 )  · · · is
a chain in X. Define A0 = A0i and g 0 (a0 ) = gi0 (a0 ) where a0 ∈ A0i . This defines an upper
i≥1
bound on the chain, so by Zorn’s Lemma there exists a maximal element in the chain. Let
(A0 , g0 ) be such a maximal element. We claim A0 = B. Suppose not; then there is some
b ∈ B r A0 . Define a left ideal of R by I = {r ∈ R | rb ∈ A0 } – this is sometimes called a
colon ideal, denoted (bR : A0 ). We construct a map h : I → E by defining h(r) = g0 (rb).
This map may be lifted by
E

h h∗

0 I R
i

Let A1 = A0 + Rb and define g1 (a0 + rb) = g0 (a0 ) + rh∗ (1). Note that g1 may not be well-
defined since a0 + rb is not necessarily the unique way to write an element of A1 . However,
if g1 is well-defined, we will have shown (A0 , g0 ) ≺ (A1 , g1 ) contradicting maximality of
(A0 , g0 ). Thus to finish the proof we show well-definedness. Suppose a0 + rb = a00 + r0 b.
Then (r − r0 )b = a00 − a0 ∈ A =⇒ r − r0 ∈ I. This gives us

g0 (a00 − a0 ) = g0 ((r − r0 )b) = h(r − r0 ).

Now we can’t take the constant r − r0 out of h, since this would leave 1 behind, and 1 6∈ I.
But we do have h(r − r0 ) = h∗ (r − r0 ) and h∗ is defined on R, so h∗ (r − r0 ) = (r − r0 )h∗ (1).
It follows easily that g1 is well-defined.
We also have

Proposition 8.2.4. Direct sums and direct summands of injectives are injective.

Proof. Similar to Theorem 8.1.7.

Definition. A module M over a domain R is divisible if for all m ∈ M and nonzero r ∈ R,


there is some m0 ∈ M such that m = rm0 .

Informally, this says that in a divisible R-module, you can ‘divide’ by R.

Example 8.2.5. Q is a divisible Z-module. In fact, this holds for the field of fractions of
any domain.

Theorem 8.2.6. Every injective left R-module is divisible.

Proof. Consider the inclusion nZ ,→ Z. Then divisibility of an R-module A is equivalent to


completing the diagram

144
8.2. Injective Modules Chapter 8. Special Modules

A y
x -
↑ 1
n
0 nZ Z

Of course this is immediate when A is injective.


The converse holds when R is a PID:

Theorem 8.2.7. Let R be a PID. Then

(1) Every divisible R-module is injective.

(2) Quotients of injectives are injective.

Proof. (2) follows from (1), while the proof of (1) is similar to the first part of the proof of
Baer’s Criterion (8.2.3).

Example 8.2.8. By Example 8.2.5 we see that Q is an injective Z-module. Moreover, by


(2) of Theorem 8.2.7, Q/Z is also an injective Z-module.

Our next goal is to show that every left R-module can be realized as a submodule of an
injective left R-module. We begin by proving this for Z-modules (abelian groups). Let M
be a Z-module. Then M ∼ = F/S where F is some free abelian group
M and S is L the module
of relations. By the fundamental theorem of abelian groups, M ∼= Z/S and Z can be
L i∈I
embedded in Q so we have a composition

M∼
M M
= Z/S ,→ Q/S.
i∈I i∈I
L L
Now Q is divisible so Q is also divisible. Then by Theorem 8.2.7, Q/S is injective.
Before proving the property for R-modules in general, we will need

Proposition 8.2.9. Let ϕ : R → S be a ring homomorphism and E an injective left R-


module. Then HomR (S, E) is an injective left S-module.

Proof. First let’s convince ourselves that HomR (S, E) is a left S-module at all. The map
ϕ : R → S induces a left R-action on S given by r · s = ϕ(r)s. S is also a right module over
itself, so there is an available action for S to act on HomR (S, E) on the left.
Now, note that HomR (S, E) is an injective left S-module ⇐⇒ HomS (−, HomR (S, E))
is an exact functor. By Hom-Tensor Adjointness,

HomS (−, HomR (S, E)) ∼


= HomR (S ⊗S −, E) ∼
= HomR (−, E)

and HomR (−, E) is exact precisely when E is injective. In fact we have proven that the
converse of the proposition holds as well.

145
8.2. Injective Modules Chapter 8. Special Modules

Corollary 8.2.10. For any divisible abelian group D, HomZ (R, D) is an injective left R-
module.
Definition. For an abelian group A, the group A∗ := HomZ (A, Q/Z) is called the dual of
A.
Theorem 8.2.11. Let A be an abelian group and A∗ its dual. Then
(1) If A is a left (resp. right) R-module, then A∗ is a right (resp. left) R-module.
(2) Given a surjection (resp. injection) of R-modules A → B, the induced map B ∗ → A∗
is injective (resp. surjective).
(3) If P is projective, P ∗ is injective. Likewise, if E is injective then E ∗ is projective.
One might hope for a “duality theorem” that says A ∼ = A∗ for any abelian group A.
Unfortunately, this is false in general. We will however prove that A embeds into A∗∗ .
Example 8.2.12. If A is finite, then A ∼ = A∗ . Indeed, since Hom splits over finite direct
sums, it suffices to consider A = Z/nZ, in which case the map

Z/nZ −→ HomZ (Z/nZ, Q/Z)


1 7−→ 1 7→ n1


is an isomorphism. Hence there is a canonical isomorphism between A and its double dual:

A −→ A∗∗ = HomZ (A∗ , Q/Z)


a 7−→ (ϕ 7→ ϕ(a)).

On the other hand, if A = Z, we have Z∗ = Q/Z so that Z∗∗ = HomZ (Q/Z, Q/Z). One
can show that HomZ (Q/Z, Q/Z) ∼ b the profinite completion of Z. There is a natural
= Z,
embedding Z ,→ Z.
b

Lemma 8.2.13. For any abelian group A, there is an embedding A ,→ A∗∗ .


Proof. Define Φ : A → A∗∗ by a 7→ ϕa , where ϕa (f ) = f (a) for any f : A → Q/Z. Take
any nonzero a ∈ A. We must show that there exists some f : A → Q/Z such that f (a) 6= 0.
Let B = Za. Then there is a map ψ : B → Q/Z given by a 7→ ψ(a) 6= 0. Now since
Q/Z is injective, ψ extends along the inclusion B ,→ A to a map f : A → Q/Z which by
construction has f (a) 6= 0. Hence Φ is injective.
Theorem 8.2.14. For any left R-module M , there is an embedding M ,→ E where E is an
injective left R-module.
Proof. By Proposition 8.2.9, M ∗∗ = HomZ (M ∗ , Q/Z) is injective. Therefore the embedding
M ,→ M ∗∗ realizes M as a submodule of an injective module.

This shows we can embed a module into some injective module, but HomZ (R, D) is often
way bigger than necessary. The natural follow-up question is: What is the smallest injective
E such that M is a submodule of E? To answer this, first we need

146
8.2. Injective Modules Chapter 8. Special Modules

Definition. E is an essential extension of M if there exists an injective map α : M ,→ E


such that for all nonzero submodules S ⊆ E, α(M ) ∩ S 6= 0.
Example 8.2.15. If M = M 0 ⊕ M 00 then M is not an essential extension of M 0 or M 00 .
The most important essential extensions are the maximal ones. Essential extensions are
related to injectives by
Proposition 8.2.16. An R-module M is injective ⇐⇒ M has no proper essential exten-
sions.
Proof. ( =⇒ ) Suppose α : M ,→ E is an essential extension. Since M is injective, the
sequence
α
0→M − →E
splits, so E ∼
= M ⊕ N for some N . If N 6= 0, α : M ,→ E is not essential so we must have
N = 0 from which it follows that E ∼ = M.
α
( ⇒= ) Embed M − → E where E is injective; this is possible by Theorem 8.2.14. Note that
α is not an essential extension, so there exists a submodule S ⊂ E such that α(M ) ∩ S = 0.
The set of such S is Zornable so pick N to be a maximal submodule of E with this property.
α π
We will show M − →E→ − E/N is essential.
The kernel of the projection π is exactly N , so since α(M ) ∩ N = 0, the composite
π ◦ α is one-to-one. Let S/N be any nonzero submodule of E/N , i.e. N ( S ⊆ E. By
maximality of N , α(M ) ∩ S 6= 0, but this implies (S/N ) ∩ (α(M )/N ) 6= 0 and thus π ◦ α
is an essential extension. By hypothesis, M has no essential extensions so E = α(M ) + N .
Further, α(M ) ∩ N = 0 so E = α(M ) ⊕ N . Then α(M ) is injective since it is the direct
summand of an injective. It follows that M is injective.
Theorem 8.2.17. The following are equivalent for M ⊆ E.
(i) E is a maximal essential extension M .
(ii) E is an essential extension of M and E is injective.
(iii) E is injective and there does not exist an injective module E 0 such that M ⊆ E 0 ( E.
An extension E satisfying (i) – (iii) is called the injective hull of M and it is unique up
to isomorphism.
Proof omitted.

Examples.
1 If R is a domain, the field of fractions Q of R is the injective hull of R.

2 Let R = k[x1 , . . . , xn ] and consider the ideal m = (x1 , . . . , xn ). Let E(R/m) denote the
injective hull of R/m. E(R/m) has a k-basis given by monomials x−k 1
1
· · · x−k
n
n
where
ki ≥ 0. Further, R acts on E(R/m) via
(
rx1−k1 · · · x−k n
if for all i, powers of xi in rx−k1
· · · x−k n
are ≤ 0
r · (x−k
1
1
· · · x −kn
n ) = n 1 n
0 otherwise.

147
8.2. Injective Modules Chapter 8. Special Modules

For example, x−2 −1


2 · x2 = 0 and xi · 1 = 0 for all i. The embedding

R/m ,→ E(R/m)
1̄ 7→ 1

shows that E(R/m) is an essential extension of R/m.

148
8.3. Flat Modules Chapter 8. Special Modules

8.3 Flat Modules


As projectives and injectives arose from the Hom functor, there is a third type of module
that arises from the tensor product.
Definition. For a ring R, a right R-module A is flat if A⊗R − is an exact functor. Flatness
of left R-modules is defined similarly for the covariant functor − ⊗R A.
Equivalently, a right R-module A is flat if and only if for every injection of left R-modules
i : B → C, 1A ⊗ i : A ⊗R B → A ⊗R C is also an injection.
Example 8.3.1. For any ring, R is a flat R-module. To see this, recall that R ⊗R A ∼
=A

and R ⊗R B = B and so we have a commutative diagram
i
A B


= ∼
=

R ⊗R A R ⊗R B
1A ⊗ i

showing 1A ⊗ i is injective.
Proposition 8.3.2. For a ring R,
(i) Every projective right (or left) R-module is flat.
(ii) A direct sum of right (or left) R-modules is flat if and only if each summand is flat.
Proof. (i) follows from (ii) since a free right R-module is the direct sum of copies of R, which
is flat. Moreover, P is projective if and only if it is the direct summand of a free module,
which, combined with the first statement Msays that projectives are always flat.
To prove (ii), consider a direct sum Mk of right R-modules Mk . For any family of R-
k M M
linear maps {fk : Mk → Nk } there is a map f : Mk → Nk taking (mk ) 7→ (fk (mk )),
k k
and clearly f is injective if and only if each fk is injective. So if i : A → B is a left R-module
injection, there is a commutative diagram
L 1⊗i L
( Mk ) ⊗R A ( Mk ) ⊗R B


= ∼
=
L L
(Mk ⊗R A) (Mk ⊗R B)
f

where ϕ is defined by (mk ⊗ a) M


7→ (mk ⊗ i(a)). By the above, 1 ⊗ i is an injection ⇐⇒ each
1Mk ⊗ i is an injection. Hence Mk is flat ⇐⇒ each Mk is flat.
k

149
8.3. Flat Modules Chapter 8. Special Modules

Lemma 8.3.3. For a left R-module M , the following are equivalent:

(a) M is flat.

(b) Every inclusion 0 → N 0 → N of R-modules induces an inclusion 0 → N 0 ⊗R M →


N ⊗R M .

(c) Every inclusion 0 → N 0 → N of finitely generated R-modules induces an inclusion


0 → N 0 ⊗R M → N ⊗R M .

Corollary 8.3.4. If every finitely generated submodule of a right R-module M is flat, then
M is flat.

However, the converse to Corollary 8.3.4 is false in general:

Example 8.3.5. For a field k, let R = k[x, y]. Then R is flat as a module over itself (it is
free) and the ideal M = (x, y) is a finitely generated submodule of R, but M is not flat.

Proposition 8.3.6. If R is an integral domain and A is a flat R-module, then A is torsion-


free.

Proof. Let Q be the field of fractions of R. Since A is flat, the functor − ⊗R A is exact, so
the exact sequence 0 → R → Q induces an exact sequence

0 → R ⊗R A → Q ⊗R A.

We know R ⊗R A ∼ = A, and Q ⊗R A is a vector space over Q, so it is torsion-free and it


follows that A is torsion-free.
The converse holds over a PID:

Proposition 8.3.7. If R is a PID, then every torsion-free R-module B is flat.

Proof. The theory of modules over a PID says that every finitely generated R-module M
that is torsion-free is also free. Thus every finitely generated submodule M ⊆ B is free,
hence projective, hence flat. By Corollary 8.3.4, this implies that B is also flat.

Example 8.3.8. For any multiplicative set S ⊆ R, the localization S −1 R is a flat module.

Similar to a property for injectives, we have

Theorem 8.3.9. A right R-module A is flat if and only if the sequence


A1 ⊗i
0 → A ⊗R I −−−→ A ⊗R R

is exact for every left ideal I ⊂ R, where i : I → R is the inclusion map.

Definition. A left R-module M is faithfully flat provided that for all left R-modules A
and B, the sequence 0 → A → B is exact if and only if 0 → A ⊗R M → B ⊗R M is exact.
One can define faithful flatness similarly for right modules.

150
8.3. Flat Modules Chapter 8. Special Modules

Lemma 8.3.10. A left R-module M is faithfully flat if and only if M is flat and for all left
R-modules A, A ⊗R M = 0 implies A = 0.

Example 8.3.11. If F is free, say F ∼ ∼ L A which is


L L
= i∈I R, then A ⊗R i∈I R = i∈I
nonzero precisely when A is nonzero. Hence free modules are faithfully flat.

Example 8.3.12. Any localization S −1 R of a ring R is flat by Example 8.3.8, but is not
faithfully flat in general. For example, if R is an integral domain with field of fractions K –
the localization of R at the multiplicative set R r {0} – then for any nonzero torsion module
M , M ⊗R K = 0.

Example 8.3.13. In general, projective modules are not faithfully flat. For instance, sup-
pose R has a nontrivial idempotent element e ∈ R, that is, e2 = e. Then e(1 − e) = 0 and
1 − e is also an idempotent, and we have R ∼ = eR ⊗ (1 − e)R. So eR is a projective R-module,
but it is not faithfully flat since (1 − e)R ⊗R eR = (1 − e)R ⊗R e2 R = e(1 − e)R ⊗R eR = 0.

Theorem 8.3.14. Let R be a commutative ring and M an R-module. Then the following
are equivalent:

(a) M is flat.

(b) For all prime ideals p ⊂ R, the localization Mp is flat as an R-module (equivalently,
as an Rp -module).

(c) For all maximal ideals m ⊂ R, the localization Mm is flat as an R-module (equivalently,
as an Rm -module.

Recall that an R-module M is finitely presented if there exists a short exact sequence
0 → K → Rn → M → 0 with n finite.

Proposition 8.3.15. Let M be a finitely generated R-module. Then

(a) If M is projective, then it is finitely presented.

(b) If M is finitely presented and 0 → L → Rm → M → 0 is any short exact sequence


with m finite, then L is finitely generated.

Proof. (1) Let 0 → K → Rn → M → 0 be any presentation of M . Then since M is


projective, the sequence splits, so K is a direct summand of Rn , thus a quotient of Rn and
therefore finitely generated.
(2) Assuming M is finitely presented, there is some short exact sequence 0 → K → Rn →
M → 0 with n finite and K finitely generated. Consider the diagram
0 K Rn M 0

f g id

0 L Rm M 0

151
8.3. Flat Modules Chapter 8. Special Modules

Then since Rm is free, we get a map g : Rn → Rm making the right square commute. Call
f the restriction of g to K, which gives a commutative square on the left. By the Snake

Lemma, 0 = ker id → coker f − → coker g → coker id = 0 is exact, so the middle arrow is
an isomorphism. Now coker g is finitely generated as it is a quotient of Rm , so this implies
coker f = L/f (K) is finitely generated. Since K was finitely generated, f (K) is also finitely
generated and it now follows that L is finitely generated.
We now relate finite presentability to flatness. First, we need a lemma.

Lemma 8.3.16. Suppose 0 → N → F → M → 0 is a short exact sequence of left R-modules,


where M and F are flat. Then

(a) For any ideal I ⊆ R, N ∩ IF = IN .


P
(b) If F is free with R-basis {yi }i∈IP
, then for any n = i∈I ri yi ∈ N ⊆ F , there exist
elements ni ∈ N such that n = i∈I ri ni .

(c) Suppose F is free. For any finite set of elements {n1 , . . . , nr } ⊆ N , there exists a
morphism f : F → N such that f (ni ) = ni for all 1 ≤ i ≤ n.

Proof. (a) Tensoring with I gives an exact sequence

I ⊗R N → I ⊗R F → I ⊗R M → 0.

Notice that since M, F are flat, I ⊗R F ∼


= IF and I ⊗R M ∼= IM , and the kernel of the
resulting map IF → IM is precisely N ∩ IF . Therefore we have a commutative diagram
with exact rows
I ⊗R N I ⊗R F I ⊗R M 0

ψ ∼
= ∼
=

0 N ∩ IF IF IM 0

This induces a map ψ : I ⊗R N → N ∩IF which by a quick diagram chase is an isomorphism.


This identifies N ∩ IF with im ψ = IN ⊆ IF .
(b) Consider the ideal I = (ri )i∈I ⊆ R. Then by (a),
X
n ∈ N ∩ IF = IN = ri N
i∈I

and the statement follows. P


(c) We induct on r. When r = 1, let n = n1 = i∈I0 ri yi for some finite I0 ⊆ I and some
ri ∈ R. Applying (b), we get n = i∈I0 ri ni for some n0i ∈ N . For each yi in a basis of F ,
0
P
set (
n0i , if i ∈ I0
f (yi ) =
0, if i 6∈ I0 .

152
8.3. Flat Modules Chapter 8. Special Modules

Then this extends uniquely to an R-linear map f : F → N such that


!
X X X
f (n) = f ri yi = ri f (yi ) = ri n0i = n.
i∈I0 i∈I0 i∈I0

To induct, assume the statement holds for all finite subsets of N consisting of r − 1 elements.
For a given {n1 , . . . , nr } ⊆ N , the inductive hypothesis constructs for us two maps f1 , f2 :
F → N satisfying
f1 (n1 ) = n1 and f2 (ni − f1 (ni )) = ni − f1 (ni ) for all 2 ≤ i ≤ r.
Set f = f1 + f2 − f2 ◦ f1 , where f2 ◦ f1 is defined by changing targets, i.e. considering
f1 : F → N ⊆ F . Then f (n1 ) = f1 (n1 ) = n1 and for each 2 ≤ i ≤ r,
f (ni ) = f1 (ni ) + f2 (ni − f1 (ni )) = f1 (ni ) + ni − f1 (ni ) = ni .
Therefore the statement holds for all r by induction.
Theorem 8.3.17. For a left R-module M , the following are equivalent:
(1) M is finitely generated and projective.
(2) M is finitely presented and flat.
(3) M is finitely presented and for all maximal ideals m ⊂ R, the localization Mm is free.
Proof. (1) =⇒ (2) is Propositions 8.3.15 and 8.3.2.
(2) =⇒ (1) If M is finitely presented, then there exists a short exact sequence 0 →
N → F → M → 0 where F is finitely generated and free and N is finitely generated, say by
n1 , . . . , nr ∈ N . Then since M is flat, Lemma 8.3.16 says there is a morphism f : F → N
with f (ni ) = ni for all i, but since the ni generate N , this means f |N = idN . Hence the short
exact sequence splits, so in particular M is a direct summand of F and thus projective.
(1) =⇒ (3) By the above, M is finitely presented (and flat), so there exists a presentation
Rm → Rn → M → 0 with m, n finite. Flatness implies that tensoring with Rm for any
maximal ideal m yields an exact sequence of Rm -modules
m n
Rm → Rm → Mm → 0.
Therefore Mm is finitely presented. Theorem 8.3.14 also shows Mm is flat. By finite presenta-
tion, Mm /m is a finite dimensional vector space over the residue field κ(m) = Rm /m. Choose
m /m which are the images of x1 , . . . , xr ∈ Mm . Then by Nakayama’s
a basis x̄1 , . . . , x̄r of MP
Lemma (1.2.1), Mm = ri=1 Rm xi . Thus Mm has a finite presentation of Rm -modules
N → F → Mm → 0
where F is free with basis y1 , . . . , yr whose P
images in Mm are x1 , . . . , xP
r . Suppose n =
P r r r
i=1 ri yi ∈ N . By exactness, n maps to i=1 ri xi = 0 in Mm , so i=1 r̄i x̄i = 0 in
Mm /m. Since the x̄i are k(m)-linearly independent, this implies each Prr̄i = 0, that is, ri
lies in ker(Rm → κ(m)), and so ri ∈ m. Hence we see that n = i=1 ri yi ∈ mF . By
Lemma 8.3.16(a), N = N ∩ mF = mF , so by Nakayama’s Lemma (1.2.1), N = 0. This
shows Mm ∼ = F.
(3) =⇒ (1) will be proven later.

153
8.3. Flat Modules Chapter 8. Special Modules

We now turn our attention back to characterizing faithfully flat R-modules.


Lemma 8.3.18. For an R-module M , the following are equivalent:
(a) M is faithfully flat.
(b) M is flat and for every maximal ideal m ⊂ R, mM 6= M .
Proof. (a) =⇒ (b) For any maximal ideal m ⊂ R, R/m 6= 0, so if M is faithfully flat,
M/mM ∼ = R/m ⊗R M 6= 0. Hence mM 6= M .
(b) =⇒ (a) Take an R-module N 6= 0. By Lemma 8.3.10, it suffices to show N ⊗R M 6= 0.
Choose a nonzero element x ∈ N and consider the module homomorphism ϕ : R → N defined
by setting ϕ(1) = x. Let I = ker ϕ ⊂ R. Then R/I is a submodule of N and I 6= R (since
x 6= 0), so there is some maximal ideal m ⊂ R containing I. Thus IM ⊆ mM 6= M so
R/I ⊗R M ∼ = M/IM 6= 0. Now by flatness, R/I ⊗R M is a submodule of N ⊗R M , so this
shows N ⊗R M 6= 0 as desired.
Example 8.3.19. For any ring R, the module
M
M= Rm
m∈MaxSpec(R)

is faithfully flat. Indeed, M is flat since it is a direct sum of flat R-modules (use Proposi-
tion 8.3.2(ii) and Theorem 8.3.14). On the other hand, suppose N ⊗R M = 0. Then
!
M M M
0 = N ⊗R Rm = (N ⊗R Rm ) = Nm .
m m m

Therefore Nm = 0 for all maximal ideals m ⊂ R, so N = 0.


This gives
L examples of modules which are faithfully flat but not projective, such as the
direct sum p Z(p) over all prime integers p ∈ Z.
Lemma 8.3.20. If M is a finitely presented R-module and N is any R-module, then for all
multiplicative sets S ⊆ R, there is an isomorphism
S −1 HomR (M, N ) ∼
= HomS −1 R (S −1 M, S −1 N ).
Proof. The assumption that M is finitely presented means there is an exact sequence Rm →
Rn → M → 0. Localization is an exact functor, so S −1 Rm → S −1 Rn → S −1 M → 0 is still
exact, meaning S −1 M is finitely presented as an S −1 R-module, since S −1 Rn ∼
= (S −1 R)n . On
the other hand, since HomR (−, N ) is left exact we get an exact sequence
0 → HomR (M, N ) → HomR (Rn , N ) = N n → HomR (Rm , N ) = N m .
Localizing this sequence gives the top row in the following diagram:
0 S −1 HomR (M, N ) S −1 N n S −1 N m

ϕ ∼
= ∼
=

0 HomS −1 R (S −1 M, S −1 N ) (S −1 N )n (S −1 N )m

154
8.3. Flat Modules Chapter 8. Special Modules

This induces a morphism ϕ : S −1 HomR (M, N ) → HomS −1 R (S −1 M, S −1 N ) and a quick


diagram chase shows that it is an isomorphism.
Let f : A → B be a ring homomorphism. This induces two functors on the corresponding
categories of modules:

ModB −→ ModA
N 7−→ N, with a · n = f (a)n,
and ModA −→ ModB
M 7−→ MB := B ⊗A M.

Lemma 8.3.21. If M is a (faithfully) flat A-module, then MB is a (faithfully) flat B-module.

Definition. A ring homomorphism f : A → B is called (faithfully) flat if B is (faithfully)


flat as an A-module.

Corollary 8.3.22. If f : A → B is a (faithfully) flat morphism and S ⊆ A is a multiplicative


set with T = f (S), then the induced morphism S −1 f : S −1 A → T −1 B is (faithfully) flat.

Proof. This follows from the fact that there is a canonical isomorphism of S −1 A-modules
T −1 B ∼
= S −1 A ⊗A B.
On the other hand, we have:

Lemma 8.3.23. Suppose f : A → B is faithfully flat and M is a faithfully flat B-module.


Then M is also (faithfully) flat as an A-module.
f g
Corollary 8.3.24. If A →
− B →
− C are faithfully flat ring homomorphisms, then so is
g ◦ f : A → C.

Proposition 8.3.25. Let f : A → B be a faithfully flat homomorphism. Then

(a) For every A-module M , the map M → M ⊗A B is injective. In particular, f itself is


injective.

(b) If I ⊂ A is an ideal, then IB ∩ A = I.

(c) The induced morphism f ∗ : Spec B → Spec A is surjective.

Proof. (a) Suppose x ∈ M r {0}. Then Ax is a nonzero submodule of M , so because B is a


faithfully flat A-module, 0 6= Ax⊗A B ,→ M ⊗A B is an inclusion of a nonzero submodule. On
the other hand, Ax⊗A B = (x⊗10b, so x⊗1 6= 0 in Ax⊗A B and hence x⊗1 6= 0 in M ⊗A B.
This shows that x is not an element of ker(M → M ⊗A B), so ker(M → M ⊗A B) = 0.
(b) Apply (a) to the A-module M = A/I to see that

A/I ,→ A/I ⊗A B ∼
= B/IB.

In general, the kernel of this map is (IB∩A)/I but because the map is injective, (IB∩A)/I =
0, that is, IB ∩ A = I.

155
8.3. Flat Modules Chapter 8. Special Modules

(c) Take a prime ideal p ⊂ A. We want to find a prime ideal q ⊂ B with p = q ∩ A. Let
Ap be the localization at p, which is a local ring with maximal ideal pAp . Since A → B is
faithfully flat, Corollary 8.3.22 shows that Ap → Bp is also faithfully flat, and hence injective
by (a). Applying (b) to the ideal I = pBp ⊂ Bp , we get pBp ∩ Ap = pAp 6= Ap , so pBp 6= Bp .
Thus there is some maximal ideal m ⊂ Bp containing pBp . Now m ∩ Ap ⊇ pAp but 1 6∈ m, so
m ∩ Ap is a proper ideal of Ap containing its unique maximal ideal pAp ; hence m ∩ Ap = pAp .
By Proposition 1.3.8, there is a prime ideal q ⊂ B extending to m ⊂ Bp , and by construction
q ∩ A = pAp ∩ A = p.
Remark. It turns out that f : A → B is faithfully flat if and only if f is flat and f ∗ :
Spec B → Spec A is surjective.
Theorem 8.3.26 (Generic Faithful Flatness). Let A be an integral domain with field of
fractions K, B/A a finitely generated ring extension and assume B ,→ K ⊗A B is injective.
Then there exist nonzero elements a ∈ A and b ∈ B such that the induced morphism Aa → Bb
is faithfully flat.
Proof. Suppose B = A[b1 , . . . , bn ] with bi ∈ B. By Noether’s normalization lemma (Theo-
rem 2.4.1), there exist elements x1 , . . . , xm ∈ BK = K ⊗A B such that BK /K[x1 , . . . , xm ] is
an integral extension. After multiplying through by a common denominator, we may assume
x1 , . . . , xm ∈ B. In particular, b1 , . . . , bn are integral over K[x1 , . . . , xm ], so they are integral
over Aa0 [x1 , . . . , xm ] for some nonzero element a0 ∈ A. This implies Ba0 /Aa0 [x1 , . . . , xm ] is a
finite ring extension. Thus we may replace A with Aa0 and B with Bb0 to have a finite ring
extension B/A[x1 , . . . , xm ].
Set C = A[x1 , . . . , xm ], E = K(x1 , . . . , xm ) and consider the diagram
B BK E ⊗C B

C K[x1 , . . . , xm ] E

After inverting an element of C, we may assume BK → E ⊗C B is injective. Note that


the columns in the diagram are all finite ring extensions; also, the right column is injective
0 0
Ps {z1 , 0. . . , zr } of E ⊗0C B and let z1 , . . . , zs be
because E is a flat C-module. Choose an E-basis
generators of B as a C-module, so that B = i=1 Czi . Then each Pr zi is a linear combination
of the zj , so there is a nonzero polynomial q ∈ C with Bq = j=1 Cq zj . Observe that the
map

ϕ : Cqr −→ Bq
r
X
(c1 , . . . , cr ) 7−→ cj zj
j=1

b : E ⊗Cq Cqr → E ⊗Cq Bq =


is surjective. Then since E ⊗Cq − is right exact, the base change ϕ
E ⊗C B is also surjective. But both E ⊗Cq Cqr and E ⊗C B are E-vector spaces of dimension
b is also injective. Thus E ⊗Cq ker ϕ = 0, which implies ker ϕ = 0 since E is faithfully
r, so ϕ
flat. Hence Cqr ∼= Bq as Cq -modules.

156
8.3. Flat Modules Chapter 8. Special Modules

Choose a nonzero coefficient a in the polynomial q ∈ C = A[x1 , . . . , xm ] and set b = aq.


Then Aa ,→ Bb is injective. We claim that it is faithfully flat. Consider the tower of ring
extensions
Aa ,−→ Aa [x1 , . . . , xm ] ,→ Aa [x1 , . . . , xm ]q ,→ Bb .
The above paragraph shows that Aa [x1 , . . . , xm ] is a free Aa -module, so in particular it is
faithfully flat by Example 8.3.11. Likewise, Bb is a free module over Aa [x1 , . . . , xm ]q , so
the third map is also faithfully flat. Finally, by Corollary 8.3.24 we need only show the
middle map is faithfully flat to complete the proof. Note that Aa [x1 , . . . , xm ]q is flat over
Aa [x1 , . . . , xm ] by Example 8.3.8, so by Lemma 8.3.18 we need only check that if m ⊂ Aa
is a maximal ideal, mAa [x1 , . . . , xm ]q 6= Aa [x1 , . . . , xm ]q . Since a is a coefficient of q, clearly
mAa [x1 , . . . , xm ]q cannot contain q. Also, mAa [x1 , . . . , xm ] 6= Aa [x1 , . . . , xm ] since Aa ,→
Aa [x1 , . . . , xm ]q is faithfully flat (this follows from Lemma 8.3.18). Together, these imply that
mAa [x1 , . . . , xm ]q 6= Aa [x1 , . . . , xm ]q , since otherwise q k ∈ mAa [x1 , . . . , xm ]q for some k ≥ 1,
but m is radical, so we would have q ∈ mAa [x1 , . . . , xm ]q . Hence Aa ,→ Aa [x1 , . . . , xm ]q is
flat.
We next prove that the completion functor A 7→ A b defined in Section 5.1 is flat.
Theorem 8.3.27. Suppose A is a noetherian ring, I ⊂ A is an ideal with ∞ n
T
n=1 I = 0 and
M is a finitely generated A-module. Then M c∼ = M ⊗A A.
b
bn ∼
Proof. Notice that for a free module An , we have A = An ⊗A A.
b Then taking a presentation
of M , say
Am → An → M → 0,
and applying Corollary 5.4.9, we get a diagram

Am ⊗A A
b An ⊗A A
b M ⊗A A
b 0

∼ ∼ α

bm
A bn
A M
c 0

with exact rows. Then by the five lemma, there exists a map α which fills in the diagram
and is an isomorphism.
Corollary 8.3.28. Completion is flat, i.e. for any local noetherian ring A, the natural map
A→A b is flat.
The following is an analogue of the going down theorem (2.3.10) for flat morphisms.
Theorem 8.3.29. Let ϕ : A → B be a flat morphism of noetherian rings. Suppose q ⊂ B
is a prime ideal, p = ϕ−1 (q) and there is a chain of prime ideals
p ) p1 ) p2 ) · · · ) pn
in A. Then there exists a chain of prime ideals
q ) q1 ) q2 ) · · · ) qn
in B such that ϕ−1 (qi ) = pi for all 1 ≤ i ≤ n.

157
8.3. Flat Modules Chapter 8. Special Modules

Proof. If n = 1, choose q1 ⊆ q which is minimal among primes containing p1 B. Suppose


ϕ−1 (q1 ) ) p1 . Then there exists some x ∈ ϕ−1 (q1 ) r p1 . Consider the short exact sequence
x
0 → A/p1 →
− A/p1 →
− A/(p1 , x) → 0.
(The left map is injective since p1 is prime and x 6∈ p1 .) Since B is a flat A-module, tensoring
with B yields another short exact sequence
x
0 → B/p1 B →
− B/p1 B →
− B/(p1 , x)B → 0.
Since q1 is minimal over p1 B, it is an associated prime of B/p1 B (Theorem 3.6.7), so Propo-
sition 3.6.9 implies x is a zero divisor in B/p1 B which contradicts exactness of the second
sequence on the left. Hence we must have ϕ−1 (q1 ) = p1 . The rest of the chain is constructed
inductively using the same argument.
Theorem 8.3.30. Let ϕ : (A, m) → (B, n) be a local morphism of local noetherian rings.
Then
(1) dim A + dim B/mB ≥ dim B.
(2) If ϕ is flat, then dim A + dim B/mB = dim B.
Proof. (1) Set d = dim A, e = dim B/mB and choose a system of parameters x1 , . . . , xd for A
and elements y1 , . . . , ye ∈ B whose images ȳ1 , . . . , ȳe form a system of parameters in B/mB.
Set I = (ϕ(x1 ), . . . , ϕ(xd ), y1 , . . . , ye ). We claim r(I) = n. It will then follow by Krull’s height
theorem (6.2.10) that dim B ≤ d + e. To prove the claim, set J = (ϕ(x1 ), . . . , ϕ(xd )) ⊆ I
and write
r(I) r(r(J), y1 , . . . , ye )
=
r(J) r(J)
r(mB, y1 , . . . , ye )
=
r(mB)
r(r(mB), ȳ1 , . . . , ȳe ) n
= = .
r(mB) r(mB)
This proves r(I) = n by the correspondence theorem for prime ideals.
(2) We must show dim B ≥ d + e. Choose a chain of prime ideals
q̄0 ( q̄1 ( · · · ( q̄e = n/mB
in B/mB and use the correspondence theorem to lift it to a chain
q0 ( q1 ( · · · ( qe = n
in B. Then q0 must contract to m, so if
p0 ( p1 ( · · · ( pn = m
is a chain of primes in A, since ϕ is flat, Theorem 8.3.29 provides a chain of prime ideals
q00 ( q01 · · · ( q0d−1 ( q0
lying over the pi . This proves dim B ≥ d + e as required.

158
8.3. Flat Modules Chapter 8. Special Modules

Corollary 8.3.31. If A is a local noetherian ring, then dim A = dim A.


b

Proof. This follows from the fact that A → A


b is flat and A/m
b A b∼
= A/m which is a field.
We also obtain a generalization of Corollary 2.6.3.

Corollary 8.3.32. If A is a noetherian ring, then dim A[x1 , . . . , xn ] = dim A + n for every
n ≥ 1.

159
Chapter 9

Categorical Constructions

160
9.1. Products and Coproducts Chapter 9. Categorical Constructions

9.1 Products and Coproducts


Category theory requires us to think outside the box when it comes to our preconceived
notions about set theory. Since a category is determined in terms of objects (which may be
sets) and morphisms, often ignoring elements completely, it will be advantageous for us to
better understand certain set theoretic constructions in the language of categories.
We begin by redefining the disjoint union. In set theory, two subsets A and B of a larger,
ambient set can be forced to be disjoint by taking the product (A ∪ B) × {0, 1} and defining
subsets A0 = A × {0} and B 0 = B × {1}. Clearly A0 ∩ B 0 = ∅.

Definition. The union A0 ∪ B 0 is called the formal disjoint union of A and B. We denote
this by A t B.

There are two functions associated with this construction, α : A → A0 ∪ B 0 which takes
a 7→ (a, 0), and β : B → A0 ∪ B 0 which maps b 7→ (b, 1). In fact, given functions f : A → X
and g : B → X, there is a unique function θ : A t B → X that extends both f and g:
(
f (a) if u = (a, 0) ∈ A0
θ(u) =
f (b) if u = (b, 1) ∈ B 0 .

This is a well-defined map since A0 ∩ B 0 = ∅. This describes disjoint unions in terms that
are easily compatible with categories; notice that we didn’t use an element at all.

Definition. If A, B ∈ obj(C), their coproduct is a triple (A t B, α, β) for some A t B ∈


obj(C) and morphims α : A → A t B and β : B → A t B, called injections, such that
for every X ∈ obj(C) and morphisms f : A → X and g : B → X, there exists a unique
morphism θ : A t B → X making the diagram commute.

A
f
α
θ
AtB X
β
g
B

This shows that the coproduct is a solution to a universal mapping problem. Below we
see that some familiar objects together with their most closely-associated morphisms can be
realized as coproducts.
Examples.

1 We showed above that in the category Sets, the coproduct of two sets A and B is their
disjoint union.

2 In the category R-Mod of left R-modules, two modules A and B have a coproduct: their
direct sum A ⊕ B together with the natural injections α : a 7→ (a, 0) and β : b 7→ (0, b).

161
9.1. Products and Coproducts Chapter 9. Categorical Constructions

3 Let k be a commutative ring and suppose A and B are commutative k-algebras. Then
the tensor product A⊗k B is their coproduct in the category of commutative k-algebras.

4 Coproducts may not exist in every category. For example, let X be a set and define C
to be the category of objects as subsets of X and morphisms as inclusion maps between
them. For an object A ( X, its complement is Ac = X r A. Then A and Ac do not
have a coproduct.

Definition. An object A in a category C is an initial object if for every object X in C,


there is a unique morphism A → X.

Examples.

5 The empty set ∅ is an initial object in the category Sets.

6 The zero module {0} is an initial object in the category R-Mod.

7 Not all categories have initial objects. For example, the the partially-ordered set of
integers Z does not have an initial object.

Proposition 9.1.1. If A and A0 are initial objects in a category C, then they are isomorphic
and hence initial objects, if they exist, are unique.

Proof. By definition there is only one morphism f : A → A0 and only one in the other
direction, g : A0 → A. It’s easy to see that gf = idA and f g = idB , so f and g are
isomorphisms.

Corollary 9.1.2. For objects A, B ∈ obj(C), if the coproduct A t B exists it is unique up to


isomorphism.

Proof. Show A t B is an initial object. See Rotman for details.


The dual notion of a coproduct is a product, which we construct next.

Definition. Let A, B ∈ obj(C). Their product is a triple (A u B, p, q) for some object


A u B in C and morphisms p : A u B → A and q : A u B → B, called projections, such that
for every object X ∈ C and every pair of morphisms f : X → A and g : X → B, there exists
a unique morphism θ : X → A u B making the diagram commute.

A
f
p
θ
X AtB
β
g
B

Examples.

162
9.1. Products and Coproducts Chapter 9. Categorical Constructions

8 For two sets A, B ∈ Sets, their product is the regular cartesian product A×B, together
with the usual coordinate projections.

9 If R is a ring and A and B are left R-modules, their product is the direct product
A ⊕ B. Thus the product and coproduct coincide in R-Mod.

10 In Groups, the coproduct of two groups G and H is called the free product, denoted
G ∗ H, which is related to the construction of free groups. In any case, G ∗ H is distinct
from G × H, which is the product in the category Groups.

11 Recall that in topology, the fundamental group is a functor π1 : Top → Groups. On one
side, the coproduct of two topological spaces X and Y is their wedge, X ∨ Y . On the
other side, the Seifert-van Kampen Theorem tells us that π1 (X ∨ Y ) ∼= π1 (X) ∗ π1 (Y ).
So this functor preserves coproducts.

As we did with initial objects and coproducts, we use terminal products to prove that
products are unique (up to isomorphism).

Definition. An object B in a category C is a terminal object if for every object X in C,


there is a unique morphism X → B.

Proposition 9.1.3. If B and B 0 are terminal objects in a category C, then they are isomor-
phic and hence terminal objects, if they exist, are unique.

Proof. Similar to the proof for initial objects; just reverse the arrows!

Corollary 9.1.4. For objects A, B ∈ obj(C), if the product A u B exists it is unique up to


isomorphism.

Proof. Again, see Rotman.

163
9.2. Limits and Colimits Chapter 9. Categorical Constructions

9.2 Limits and Colimits


Products and coproducts are special cases of a more general notion in category theory. Let
C be a diagram category, i.e. a category whose objects are the vertices of a directed graph
and whose morphisms are in bijection with the directed edges of this graph (there may be
multiple loops on a given vertex, but every vertex is assumed to possess a distinguished
‘identity’ loop).

Definition. A direct (resp. inverse) system in a category A is a covariant (resp. con-


travariant) functor C → A where C is a diagram category.

Definition. Let F : C → A be an inverse system in A. The limit (also called the projective
or inverse limit) of F is an object lim F ∈ obj(A) such that for all objects C ∈ C, there
are morphisms ϕC : lim F → F(C) making the diagrams

lim F

ϕC ϕC 0

F(C) F(C 0 )

commute whenever HomA (F(C), F(C 0 )) 6= ∅, and such that lim F is universal among all
such objects in A.

Limits are sometimes also written lim F. By the universal property, limits are unique up
←−
to unique isomorphism.

Example 9.2.1. Let C be the category consisting of two objects {1, 2} and only identity
morphisms. An inverse system F : C → A is just defined by specifying two objects, F(1) = a1
and F(2) = a2 . Then lim F is the product of these elements, a1 u a2 .

Definition. Let F : C → A be a direct system in A. The colimit (also called the injective
or direct limit) of F is an object colim F ∈ obj(A) such that for all C ∈ C, there are
morphisms ψC : F(C) → colim F making the diagrams

F(C) F(C 0 )

ψC ψC 0

colim F

commute whenever HomA (F(C), F(C 0 )) 6= ∅, and such that colim F is universal among all
such objects in A.

Example 9.2.2. Let C be the same category as in Example 9.2.1. Then a direct system
F : C → A is once again defined by specifying F(1) = a1 and F(2) = a2 , but colim F is the
coproduct of these elements, a1 t a2 .

164
9.2. Limits and Colimits Chapter 9. Categorical Constructions

There are plenty of other important examples of limits and colimits in category theory.
Definition. Let C be the diagram category
ˆ
ˆ
ˆ

The colimit of a direct system F : C → A is called the pushout of F and the limit of an
inverse system F : C → A is called the pullback of F.
B
Explicitly, if a direct system F has image A then we write the pushout Q =
C
colim F as a square diagram
A B

C Q

B
Similarly, if F is an inverse system with image A then we write the pullback P = lim F
C
as a square diagram
P B

C A

Definition. Suppose A is a category with an object 0 ∈ obj(A) that is both initial and
f B
terminal. For any morphism f : A → B, the pushout of the diagram A is called
0
the cokernel of f , written coker f . Similarly, for such a morphism f , the pullback of the
0
diagram B is called the kernel of f , written ker f .
A f
Example 9.2.3. The category AbGps of abelian groups has the trivial group 0 as a zero
object, i.e. an object that is both initial and terminal. Then for a homomorphism of abelian
groups f : A → B, the kernel and cokernel of f coincide exactly with these notions from
abstract algebra:
ker f = {a ∈ A | f (a) = 0}
and coker f = {[b] | b ∈ B and [b] = [b0 ] if b0 = b + f (a)}.

165
9.2. Limits and Colimits Chapter 9. Categorical Constructions

This also holds in ModR for any ring R, and indeed in any abelian category as we shall see.
In the category ModR , notice that the functor HomR (C, −) does not preserve colimits since
it doesn’t preserve cokernels (this functor is not right exact; see Example 7.2.3). However,
by Proposition 7.2.2, HomR (C, −) does preserve kernels. We will show in general that
this condition is nearly equivalent to the property of preserving all limits. Similarly, recall
(Proposition 7.2.6) that C ⊗R − is a right exact functor, so it preserves cokernels, but does not
preserve kernels. We will also show that the property of right exactness is nearly equivalent
to preserving all colimits.

Definition. Let F : A → B and G : B → A be functors. Then (F, G) is called an adjoint


pair if there is a natural isomorphism

HomB (F−, −) ∼
= HomA (−, G−).

In this case F is called a left adjoint of G and G is a right adjoint of F.

Example 9.2.4. Let R and S be rings and fix a left R-module M , a left S-module N and
an (S, R)-bimodule L. Then by Theorem 7.2.5, there is a natural isomorphism

HomS (L ⊗R M, N ) ∼
= HomR (M, HomS (L, N )).

This shows (L ⊗R −, HomS (L, −)) is an adjoint pair of functors.

Example 9.2.5. In representation theory, Frobenius reciprocity is a special version of Hom-


tensor adjointness used to study adjoint pairs of representations.

Lemma 9.2.6. Let F : A → B and G : B → A be a pair of functors. Then (F, G) is an


adjoint pair if and only if there exist natural transformations η : idA → GF and ε : FG → idB
such that the compositions Gε ◦ ηG and εF ◦ Fη are both the identity natural transformation.

Proof. ( =⇒ ) Suppose (F, G) is an adjoint pair. Then for any objects X ∈ A and Y ∈ B,
there is an isomorphism

ΦX,Y : Hom(FX, Y ) −→ Hom(X, GY ).

Applying this to Y = FX, we have ΦF X,F X : Hom(FX, FX) − → Hom(X, GFX) which takes
the identity idF X to some morphism ηX = ΦF X,F X (idF X ) : X → GFX. This defines the
natural transformation η : idA → GF. On the other hand, applying the natural isomorphism

Φ to X = GY gives an isomorphism ΦF GY,Y : Hom(FGY, Y ) − → Hom(GY, GY ) under which
the identity idGY is mapped to by some εY = Φ−1
F GY,Y (idGY ) : FGY → Y . This defines the
natural transformation ε : FG → idB . To check the identity conditions, apply Φ to the
idF GY εY
sequence εY : FGY −−−→ FGY −→ Y to get:
ηGY Gε
GY −−→ GFGY −→ GY.

Thus Gε ◦ ηGY = ΦF GY,Y (εY ) which by definition equals idGY . The proof of the other identity
is similar.

166
9.2. Limits and Colimits Chapter 9. Categorical Constructions

( ⇒= ) Given natural transformations η : idA → GF and ε : FG → idB , we define


a natural transformation ΦX,Y : Hom(FX, Y ) → Hom(X, GY ) by sending α : FX → Y
to G(α) ◦ η : X → GFX → GY . Similarly, define its inverse Φ−1 X,Y : Hom(X, GY ) →
Hom(FX, Y ) by sending β : X → GY to ε ◦ F(β) : FX → FGY → Y . To see that these
are natural inverses, fix X ∈ A, Y ∈ B and α : FX → Y and consider

Φ−1 −1
X,Y ΦX,Y (α) = ΦX,Y (G(α) ◦ η) = ε ◦ F(G(α) ◦ η)
= ε ◦ FG(α) ◦ Fη since F is a functor
= εF(G(α))Fη = εF ◦ Fη(α) = α

by the identity εF ◦ Fη = id. Similarly, for β : X → GY , we have

ΦX,Y Φ−1
X,Y (β) = ΦX,Y (ε ◦ F(β)) = G(ε ◦ F(β)) ◦ η
= Gε ◦ GF(β) ◦ η since G is a functor
= Gε ◦ ηG(β) = β

by the identity Gε ◦ ηG = id. Hence ΦX,Y and Φ−1


X,Y form a natural isomorphism.

Definition. For an adjoint pair (F, G), the natural transformations η : idA → GF and
ε : FG → idB are called the unit and counit of the adjunction, respectively. The conditions
Gε ◦ ηG = id and εF ◦ Fη = id are called the triangle identities.

Theorem 9.2.7. If (F, G) is an adjoint pair of functors, then F preserves colimits and G
preserves limits.

Proof. Assume F : A → B and G : B → A. Let X : C → A be a direct system and


for each C ∈ C, let ϕC : X(C) → colim X be the induced morphism making the relevant
diagrams commute. We must prove that F(colim X) is the colimit of the direct system
F ◦ X : C → A → B. For any C ∈ C and B ∈ B, there is an isomorphism

HomB (F(X(C)), B) ∼
= HomA (X(C), G(B)).

Thus for any diagram


F(X(C)) B

colim(F ◦ X)

the corresponding diagram can be completed:


X(C) G(B)

colim X

167
9.2. Limits and Colimits Chapter 9. Categorical Constructions

On the other hand, applying F to colim X gives a diagram


F(colim X) B

F ◦ ϕC

F(X(C))

In other words, every map F(X(C)) → B which is compatible with the F ◦ ϕC will factor
through F(colim X). Hence by the universal property of colimits, F(colim X) = colim(F ◦
X). The proof that G preserves limits is dual.

Corollary 9.2.8. For any (S, R)-bimodule L, L ⊗R − preserves colimits and HomS (L, −)
preserves limits.

Theorem 9.2.9. Let F : ModR → AbGps be an additive functor. Then the following are
equivalent:

(1) F ∼
= L ⊗R − for some R-module L.
(2) F preserves colimits.

(3) F is right exact and preserves direct sums.

(4) F has a right adjoint.

Proof. (1) =⇒ (4) is Corollary 9.2.8.


(4) =⇒ (2) is Theorem 9.2.7.
(2) =⇒ (3) is immediate since a direct sum is a colimit.
(3) =⇒ (1) Consider L = F(R), which is a right R-module under right multiplication by
elements of R. Then for any R-module M , we have a presentation of M by free R-modules:
ϕ
RY −
→ RX → M → 0,

where X and Y are some sets. Note that coker ϕ = M . Since F preserves direct sums,
F(RY ) = (F(R))Y = LY ∼ = L ⊗R RY and likewise F(RX ) ∼ = L ⊗R RX . Then applying F to
the exact sequence above gives the top row in the following diagram with exact rows:

LY LX F(M ) 0


= ∼
=

L ⊗R R Y L ⊗R RX L ⊗R M 0

Then there is a map F(M ) → L⊗R M which by the Snake Lemma (7.2.11) is an isomorphism.
Naturality is easy to check.

Theorem 9.2.10. Let F : ModR → AbGps be an additive functor. Then the following are
equivalent:

168
9.2. Limits and Colimits Chapter 9. Categorical Constructions

(1) F ∼
= HomR (L, −) for some R-module L.
(2) F preserves limits.

(3) F is left exact and preserves direct products.

(4) F has a left adjoint.

Proof. Similar.

169
9.3. Abelian Categories Chapter 9. Categorical Constructions

9.3 Abelian Categories


Abelian categories are the preferred setting for working with derived functors, which are the
main tools used in homological algebra.

Definition. C is an additive category if

ˆ HomC (A, B) is an abelian group for all A, B ∈ obj(C). Note that we normally only
require Hom to be a set.

ˆ For all a ∈ A, b ∈ B, f ∈ Hom(A, B) and g ∈ Hom(B, A),

b(f + g) = bf + bf and (f + g)a = f a + ga.

ˆ There exists a 0 object, which is both initial and terminal, meaning for all A ∈ obj(C),
there exist maps 0 → A and A → 0.

ˆ C has finite products ( ) and coproducts ( ) and they agree on finite-index sets.
Q L

Q L
The last condition, that and agree on finite sets, can be proven using the other
axioms. We can also define additive functors between additive categories, which are
functors that preserve the additive structure of the category.
To fully understand abelian categories, it requires us to redefine our concept of kernels
and cokernels. These are best understood in terms of monics and epics.

Definition. A monomorphism, or monic, in a category C is a morphism u : B → C so


that if there are maps f, g : A → B such that uf = ug, then f = g.

f
u
A B C
g

Definition. An epimorphism, or epic, is the dual notion, that is π is an epic if for every
f, g : B → C, f π = gπ implies f = g.

f
π
A B C
g

In general, if u is one-to-one then u is monic, and if π is onto then π is epic, but the
converse does not hold in a general category. The converse does hold however for abelian
categories, which we take to be our definition.

Definition. A category C is abelian if every one-to-one morphism is monic and every onto
morphism is epic.

170
9.3. Abelian Categories Chapter 9. Categorical Constructions

Example 9.3.1. Let Comm be the category of commutative rings, with ring homomorphisms
as morphisms. Consider the map i : Z → Q. This is an epimorphism in C but it is clearly not
f
onto. To see that it is epic, suppose we have Q R so that f i = gi, i.e. f (n) = g(n) for
g
all n ∈ Z. Clearly if f and g agree on Z they agree on Q, so i is an epimorphism. However,
i cannot be onto because Q is much bigger than Z. This shows that Comm is not an abelian
category.

Next we relate kernels and cokernels to monics and epics.

Proposition 9.3.2. Let u : A → B be a morphism in an additive category A.

(i) If ker u exists, then u is monic ⇐⇒ ker u = 0.

(ii) Likewise, if coker u exists then u is epic ⇐⇒ coker u = 0.

Proof omitted.
From this we define

Definition. An additive category A is abelian if

(1) Every morphism has a kernel and a cokernel.

0
(2) If f : M → N is monic, then M is the kernel of the diagram coker f
N

M
(3) If f is epic, then N is the cokernel of the diagram ker f
0
Note that condition (2) implies that if f is monic, then there exists a short exact sequence
f
0→M →
− N→
− coker f → 0.

Likewise, condition (3) implies that if f is epic, there exists a short exact sequence
f
0 → ker f →
− M→
− N → 0.

Example 9.3.3. For any ring R, the category ModR is an abelian category. If R is (left)
noetherian, then the subcategory modR of finitely generated (left) R-modules is also an
abelian category.

Theorem 9.3.4. Any abelian category is naturally isomorphic to a full subcategory of ModR
for some ring R.

171
9.4. Projective and Injective Resolutions Chapter 9. Categorical Constructions

9.4 Projective and Injective Resolutions


Free resolutions are easy to motivate, and give rise naturally to the idea of projective resolu-
tions (since free and projective are closely related). For an R-module A, a free resolution
of A is an exact sequence
· · · → Fn → Fn−1 → · · · → F1 → F0 → A → 0
where each Fi is a free R-module. This can be thought of in terms of generators and relations:
in the first step, we let F0 be a free module with submodule K0 = ker(F0 → A) such that A
is the quotient of F0 by K0 . Then K0 is the module of relations that describe A and F0 /K0
is a presentation of A. Next we map F1 → K0 and let K1 be the module of relations for this
presentation, and so forth. Free resolutions give a natural way of thinking about equations
in A in terms of sequences of equations in free modules.
Proposition 9.4.1. If the short exact sequence 0 → M 0 → M → M 00 → 0 is split, and N is
another R-module, then the sequences
0 → M 0 ⊗ N → M ⊗ N → M 00 ⊗ N → 0 (9.1)
0 → Hom(M 00 , N ) → Hom(M, N ) → Hom(M 0 , N ) → 0 (9.2)
0 → Hom(N, M 0 ) → Hom(N, M ) → Hom(N, M 00 ) → 0 (9.3)
are all exact.
Proof. (1) Tensor is right exact so it suffices to show i ⊗ 1 : M 0 ⊗ N → M ⊗ N is injective.
Notice that (π ⊗ 1)(i ⊗ 1) = πi ⊗ 1 = 0 ⊗ 1 = 0 so i ⊗ 1 has a left inverse and is therefore
injective.
(2) By Proposition 7.2.2, we only need to check that i∗ : Hom(M, N ) → Hom(M 0 , N ) is
π
surjective. But the splitting M 0 ← − M gives i∗ π ∗ = (πi)∗ = (idM 0 )∗ = 1 so i∗ has a right
inverse and is therefore surjective.
(3) Similarly, covariant Hom is always left exact so it’s enough to show j∗ : Hom(N, M ) →
f
Hom(N, M 00 ) is surjective. The other splitting M ←
− M 00 gives us j∗ f∗ = (jf )∗ = (idM 00 )∗ = 1,
so j∗ has a right inverse and is surjective.
In Proposition 8.1.4 we showed that every free module was projective. This gives rise to
a more general notion of projective resolutions.
Definition. A projective resolution of an object A in an abelian category A is an exact
sequence
P = · · · → P2 → P1 → P0 → A → 0
where each Pi is projective.
Definition. For a projective resolution P = · · · → P2 → P1 → P0 → A → 0 of A, we define
the nth syzygy of P to be Kn = ker(Pn → Pn−1 ).
Definition. We say an abelian category A has enough projectives if projective resolutions
exist for every object A ∈ obj(A); equivalently there exists a projective P ∈ obj(A) and an
epic P → A.

172
9.4. Projective and Injective Resolutions Chapter 9. Categorical Constructions

Lemma 9.4.2. (a) Every module over a field has a projective resolution of length 0.

(b) Every module over a PID has a projective resolution of length at most 1.

Proof. (a) If V is a module over a field, it is a vector space and therefore free and projective.
id
Hence 0 → V − → V → 0 is a projective resolution.
(b) Let P be a projective module with a surjection P → M , so that P is also free by
Corollary 8.1.8. Then the kernel K of this map is a submodule of P , so it’s free and therefore
projective by Corollary 8.1.8 again. Hence 0 → K → P → M → 0 is a projective resolution
of M .
Before proving some results for projective resolutions, we need

Proposition 9.4.3 (Schanuel’s Lemma). Given exact sequences


i p
0→K→
− P →
− M →0
i0 p0
0 → K0 −
→ P0 −
→M →0

where P and P 0 are projective, there is an isomorphism K ⊕ P 0 ∼


= K0 ⊕ P .

Proof. See Rotman.


This generalizes nicely to long exact sequences of modules.

Proposition 9.4.4. Let A be a left R-module and suppose we have two exact sequences
i pn pn−1 p2 p1 p0
− Pn −→ Pn−1 −−−→ · · · −
0→K→ → P1 −
→ P0 −
→A→0
and
i0 qn qn−1 q2 q1 q0
0 → K0 −
→ Qn −→ Qn−1 −−→ · · · −
→ Q1 −
→ Q0 −
→ A → 0,

where the Pi and Qi are projective. Then

K ⊕ Qn ⊕ Pn−1 ⊕ · · · ∼
= K 0 ⊕ Pn ⊕ Qn−1 ⊕ · · ·

Proof. For all j = 0, . . . , n we will show that there are short exact sequences

0 → ker pj → Pj ⊕ Qj−1 · · · → ker pj−1 ⊕ Qj−1 ⊕ · · · → 0


and
0 → kerqj → Qj ⊕ Pj−1 · · · → ker qj−1 ⊕ Pj−1 ⊕ · · · → 0.

We prove this using induction. For the base case,


i p0
0 → ker p0 →
− P0 −
→B→0
and
i0 q0
0 → ker q0 −
→ Q0 −
→B→0

173
9.4. Projective and Injective Resolutions Chapter 9. Categorical Constructions

By Schanuel’s Lemma (Proposition 9.4.3), we have ker p0 ⊕ Q0 ∼ = ker q0 ⊕ P0 , proving the


base case. Next suppose the sequences above are exact for k < n, i.e. ker pk ⊕Qk ⊕Pk−1 · · · ∼
=
ker qk ⊕ Pk ⊕ Qk−1 · · · . Consider

0 → ker pk+1 → Pk+1 ⊕ Qk · · · → ker pk ⊕ Qk ⊕ · · · → 0


and
0 → kerqk+1 → Qk+1 ⊕ Pk · · · → ker qk ⊕ Pk ⊕ · · · → 0.

In each case the kernel of the first map is contained in Pk+1 ⊕ Qk (flip these for the bottom
sequence) so the inductive step follows easily.
Now since the original sequence is exact at K, K embeds into Pn via i and therefore
K = ker pn . Hence
K ⊕ Qn ⊕ Pn−1 ⊕ · · · ∼ = K 0 ⊕ Pn ⊕ Qn−1 · · ·
as desired.
Definition. If A is a left R-module and · · · → P2 → P1 → P0 → A → 0 is a projective
resolution of A, we say A has projective dimension n if there is a smallest n such that Kn
is projective. Otherwise A has infinite projective dimension (or no projective dimension).
If A has projective dimension n, we write pd(A) = n. The following result tells us that
the integer n does not depend on the projective resolution chosen for A.
Proposition 9.4.5. If one projective resolution of a module A has a projective nth syzygy,
then the nth syzygy of every projective resolution of A is projective.
Proof. Consider the projective resolutions

P = · · · → Pn → Pn−1 → · · · → P1 → P0 → A → 0
and
P = · · · → Pn0 → Pn−1
0 0
→ · · · → P10 → P00 → A → 0.

Make a soft truncation at the nth location of each to produce exact sequences

0 → Kn → Pn → Pn−1 → · · · → P1 → P0 → A → 0

0 → Kn0 → Pn0 → Pn−1


0
→ · · · → P10 → P00 → A → 0.

Then by Schanuel’s Lemma (Proposition 9.4.3), Kn ⊕Pn0 ⊕Pn−1 ⊕· · · ∼ = Kn0 ⊕Pn ⊕Pn−10
⊕· · · .
0 0 0
Let Qn = Pn ⊕ Pn−1 ⊕ · · · and Qn = Pn ⊕ Pn−1 ⊕ · · · . These are both projective since they
are direct sums of projectives, and so we see that Kn ⊕ Q0n ∼ = Kn0 ⊕ Qn . Notice that we
haven’t yet used the hypothesis that some Kn is projective, so this proof works for all n.
Now if Kn is projective, the isomorphism Kn ⊕ Q0n ∼ = Kn0 ⊕ Qn shows that Kn0 is a direct
summand of a projective, so it too is projective. Hence if the nth syzygy in a projective
resolution of A is projective, the nth syzygy is projective in all projective resolutions of A,
reassuring us that the term ‘projective dimension’ is well-defined.
There is a dual notion for injectives.

174
9.4. Projective and Injective Resolutions Chapter 9. Categorical Constructions

Definition. An injective resolution of an object A in an abelian category A is an exact


sequence
E = 0 → A → E0 → E1 → E2 → · · ·
such that each E n is injective.

Notice that when the indices increases along the arrows, we write them as superscripts,
and when they decrease along arrows we write them as subscripts. This notational convention
will be seen throughout homological algebra.

Definition. As a dual to the definition of syzygies, in an injective resolution E = 0 → A →


E0 → E1 → E2 → · · · of A we call V n = coker(En−1 → En ) the nth cosyzygy of E.

Definition. An abelian category A is said to have enough injectives if injective resolutions


exist for every object A ∈ obj(A); equivalently there exists an injective E ∈ obj(A) and a
monic A → E.

175
Chapter 10

Homology

176
10.1. Chain Complexes and Homology Chapter 10. Homology

10.1 Chain Complexes and Homology


The environment in which homological algebra is performed is abelian categories. In general,
abelian categories may be realized as the category of left R-modules for some ring R, so we
will often reduce proof to the case where C = R-Mod.

Definition. A chain complex, usually shortened to complex, in an abelian category A is


a chain of differentials ∂n ,
∂n+1 n ∂
A• = · · · → An+1 −−−→ An −→ An−1 → · · ·

such that for all n, ∂n ∂n+1 = 0.

Definition. If A is an abelian category, the category of complexes over A is denoted


Comp(A).

Proposition 10.1.1. If A is an abelian category then Comp(A) is also abelian.

Proof omitted.

Definition. If A• and B• are complexes in Comp(A), a chain map f• : A• → B• is a


sequence of morphisms (fn : An → Bn ) making the following commute

∂n+1 ∂n
··· An+1 An An−1 ···

fn+1 fn fn−1

··· Bn+1 0
Bn Bn−1 ···
∂n+1 ∂n

Note that the condition ∂n ∂n+1 = 0 (sometimes denoted ∂ 2 = 0) is equivalent to im ∂n+1 ⊆


ker ∂n . We will usually write ker ∂n = Zn (A• ), called the n-cycles of A• . Likewise, we denote
im ∂n+1 = Bn (A• ), called the n-boundaries of A• . Then we have subcomplexes Z(A• ) and
B(A• ) of A• given by
0 0 0
Z(A• ) = · · · →
− Zn (A• ) →
− Zn−1 (A• ) →
− ···
0 0 0
B(A• ) = · · · →
− Bn (A• ) →
− Bn−1 (A• ) →
− ···

Note that B(A• ) ⊆ Z(A• ). This allows us to define the central object in homological algebra:

Definition. Let A• be a complex in an abelian category A. The nth homology of A• is


the quotient Hn (A• ) := Zn (A• )/Bn (A• ), which is an object in A.

Intuitively, the nth homology measures the failure of exactness of a complex at the nth
location in the chain. Since A is abelian, we also have H(A• ) = Z(A• )/B(A• ). Notice that
A• is exact at An ⇐⇒ Hn (A• ) = 0, and the entire complex A• is exact ⇐⇒ H(A• ) = 0.

177
10.1. Chain Complexes and Homology Chapter 10. Homology

Proposition 10.1.2. For every abelian category A, the nth homology Hn : Comp(A) → A is
an additive functor for each n ∈ Z.

Proof. We prove this for A = Ab, the category of abelian groups. The definition above
defines the functor Hn : Comp(Ab) → Ab on objects by A• 7→ Hn (A• ) so it remains to define
homology on morphisms. Let f : C → C 0 be a chain map and define

Hn (f ) : Hn (C) −→ Hn (C 0 )
zn + Bn (C) 7−→ fn (zn ) + Bn (C 0 ).

We must show that fn (zn ) is a cycle and that Hn (f ) is well-defined, i.e. independent of the
choice of cycle zn . Since f is a chain map, we have the following commutative diagram
∂n+1 ∂n
··· Cn+1 Cn Cn−1 ···

fn+1 fn fn−1

0 0
··· Cn+1
0
Cn0 Cn−1 ···
∂n+1 ∂n

showing (by a quick diagram chase) that indeed fn (zn ) ∈ Zn (C 0 ), and if zn + Bn (C) =
wn + Bn (C) then fn (zn − wn ) must land in Bn (C 0 ).
Next we prove Hn is a functor. Obviously Hn (1C ) must be the identity on Hn (C). If f
and g are chain maps such that gf is defined in some Hom set, then for all z ∈ Zn (C) we
have

Hn (gf )(z) = (gf )n (z) + Bn (C 0 ) = gn (fn (z) + Bn (C 0 ))


= Hn (g)(fn (z) + Bn (C 0 )) = (Hn (g)Hn (f ))(z).

Finally, if f, g : C → C 0 are chain maps, then

Hn (f + g)(z) = (fn + gn )(z) = (fn (z) + gn (z)) + Bn (C 0 ) = (Hn (f ) + Hn (g))(z).

Hence Hn is a functor.
The first big result we will prove is a statement for short exact sequences of complexes.
We will make extensive use of the Snake Lemma (7.2.11). Note that by the equivalence of
abelian categories and module categories R-Mod for rings R, we are able to state things in
terms of R-modules.
i p
Theorem 10.1.3 (The Long Exact Sequence). Let 0 → C 0 → − C → − C 00 → 0 be an exact
sequence in Comp(R), the category of complexes of left R-modules. Then there is a long exact
sequence in R-Mod

· · · → Hn (C 0 ) → Hn (C) → Hn (C 00 ) → Hn−1 (C 0 ) → Hn−1 (C) → Hn−1 (C 00 ) → · · ·

Proof. First we expand the short exact sequence:

178
10.1. Chain Complexes and Homology Chapter 10. Homology

.. .. ..
. . .

0
in+1 pn+1 00
0 Cn+1 Cn+1 Cn+1 0
0 00
∂n+1 ∂n+1 ∂n+1
in pn
0 Cn0 Cn Cn00 0

∂n0 ∂n ∂n00
0
in−1 pn−1 00
0 Cn−1 Cn−1 Cn−1 0

.. .. ..
. . .

How do we define Hn (i) : Hn (C 0 ) → Hn (C)? In the proof of the Snake Lemma (7.2.11), we
in
proved that Zn (C 0 ) −
→ Zn (C) induces
i
Zn (C 0 ) −
→n
Zn (C) → Hn (C)
so we just need to check that if z ∈ Bn (C 0 ) then in (z) ∈ Bn (C). This follows from the
commutativity of the left squares in the diagram above. The dual argument works for
pn
Hn (C) → Zn (C) −→ Zn (C 00 ). Hence we have well-defined maps Hn (i) : Hn (C 0 ) → Hn (C)
and Hn (p) : Hn (C) → Hn (C 00 ).
The Snake Lemma (7.2.11) is also a good place to consult when defining a connecting
map ð : Hn (C 00 ) → Hn−1 (C 0 ). In that proof, we saw that there is a map
Zn (C 00 ) −→ Cn−1
0
/Bn−1 (C 0 ) = coker ∂n0
z 00 7−→ c0 + im ∂n0 .
0
By the Correspondence Theorem, the submodules of Cn−1 /Bn−1 (C 0 ) are in correspondence
with submodules of Cn−1 containing Bn−1 , and one of these if Zn−1 (C 0 ). Thus we want to
0 0

prove c0 + im ∂n0 = z 0 + im ∂n0 for some z 0 ∈ ker ∂n−1 , which is equivalent to c0 − z 0 ∈ im ∂n0 .
A diagram chase actually reveals that c0 itself lies in ker ∂n−1
0
. So by the lifting property of
kernels, we have a map
Zn (C 00 ) −→ Zn−1 (C 0 )/Bn−1 (C 0 ) = Hn−1 (C 0 )
z 00 7−→ c0 + im ∂n−1
0
.
Another diagram chase shows that if z 00 ∈ Bn (C 00 ) then z 00 7→ c0 ∈ Bn−1 (C 0 ) in the above
map. This induces a map ð : Hn (C 00 ) → Hn−1 (C 0 ).
Now that we have all the maps in hand, it remains to show that the long sequence is
exact at each homology. The proof of this is very similar to the latter part of the proof of
the Snake Lemma (7.2.11).

179
10.1. Chain Complexes and Homology Chapter 10. Homology

Proposition 10.1.4. Let R and A be rings and suppose T : R Mod → A Mod is an exact
additive functor. Then T commutes with homology, that is, for every complex C ∈ Comp(R)
and for every n, there is an isomorphism

Hn (T (C)) ∼
= T (Hn (C)).

Proof. The nth homology Hn (C) defines two exact sequences in R Mod:

0 → Bn (C) → Zn (C) → Hn (C) → 0 (10.1)


0 → Zn (C) → Cn → Bn−1 (C) → 0. (10.2)

Since T is an additive exact functor, applying it to the sequences above gives two new exact
sequences

0 → T (Bn (C)) → T (Zn (C)) → T (Hn (C)) → 0 (10.3)


0 → T (Zn (C)) → T (Cn ) → T (Bn−1 (C)) → 0. (10.4)

n ∂ ∂n−1
On the other side of things, expand the complex C = · · · → Cn −→ Cn−1 −−−→ · · · and apply
the exact functor T , giving
T (∂n ) T (∂n−1 )
T (C) = · · · → T (Cn ) −−−→ T (Cn−1 ) −−−−→ · · ·

Since T is additive, T (∂n )T (∂n−1 ) = T (∂n ∂n−1 ) = T (0) = 0 (the zero map) so we see that
T (C) is a complex in Comp(A). Therefore we can define its nth homology Hn (T (C)) =
Zn (T (C))/Bn (T (C)). As above, this defines two sequences in A Mod:

0 → Bn (T (C)) → Zn (T (C)) → Hn (T (C)) → 0 (10.5)


0 → Zn (T (C)) → T (Cn ) → Bn−1 (T (C)) → 0 (10.6)

Now consider the diagram whose rows are exact sequences (6) and (4).
0 Zn (T (C)) T (Cn ) Bn−1 (T (C)) 0

f id g

0 T (Zn (C)) T (Cn ) T (Bn−1 (C)) 0

Since the middle arrow is the identity (i.e. a boring isomorphism), if we can show that there
exist maps f and g where either one is an isomorphism, then the Five Lemma (7.2.12) proves
the other is an isomorphism. This will nearly finish the problem.
To prove that there is an isomorphism f , note that any exact additive functor T commutes
ϕ
with kernels, since applying T to an exact sequence 0 → K → A − → B produces an exact
T (ϕ)
sequence 0 → T (K) → T (A) −−→ T (B). Then T (K) is the kernel of T (ϕ), and kernels
are unique up to isomorphism so T (ker ϕ) ∼ = ker T (ϕ). In our context this proves that
Zn (T (C)) = ker T (∂n ) ∼
= T (ker ∂n ) = T (Zn (C)). Now we can fill one arrow:

180
10.1. Chain Complexes and Homology Chapter 10. Homology

0 Zn (T (C)) T (Cn ) Bn−1 (T (C)) 0


= id g

0 T (Zn (C)) T (Cn ) T (Bn−1 (C)) 0

Clearly the left square commutes since both vertical arrows are isomorphisms.
Now we turn our attention to constructing g. If b ∈ Bn−1 (T (C)) then exactness of the
top row says there is an element c ∈ T (Cn ) mapping to b. Since the middle vertical arrow is
the identity, c 7→ c ∈ T (Cn ) in the bottom row, which has image b0 ∈ T (Bn−1 (C)). We use
this to define g : Bn−1 (T (C)) → T (Bn−1 (C)), b 7→ b0 .
The only concern is that g may not be well defined, so we take a moment to verify this
with a diagram chase:
c̃ (1)
(2) c
z c − c̃ b
0 Zn (T (C)) T (Cn ) Bn−1 (T (C)) 0

(3) ∼
= id g

0 T (Zn (C)) T (Cn ) T (Bn−1 (C)) 0


z0 (4)
c − c̃ (5)
0

Suppose b pulls back to c and c̃ in T (Cn ). Then their difference c − c̃ is in the kernel of
T (Cn ) → Bn−1 (T (C)) so c−c̃ pulls back to some z ∈ Zn (T (C)). The left vertical arrow takes
z 7→ z 0 isomorphically, so z 0 7→ c − c̃ along the bottom left by commutativity. By exactness
of the bottom row, pushing z 0 along both maps on the bottom takes z 0 7→ c − c̃ 7→ 0, which
shows that g is well defined after all.
So g exists and by construction it makes the right square commute. We can apply the
Five Lemma (7.2.12) to see that g must be an isomorphism, and this holds for any n so if
we consider the diagram with (5) and (3) as its rows we can put in isomorphisms on the left
and middle arrows.
0 Bn (T (C)) Zn (T (C))) Hn (T (C)) 0


= ∼
=

0 T (Bn (C)) T (Zn (C)) T (Hn (C)) 0

The proof above generalizes to any diagram of this form, showing there exists an isomorphism
(the dashed arrow) Hn (T (C)) ∼= T (Hn (C)) as desired.

181
10.1. Chain Complexes and Homology Chapter 10. Homology

The dual notion to homology is cohomology, which as with many ‘dual’ concepts in
homological algebra, is formed by reversing the arrows. We define things explicitly below.

Definition. A co-chain complex in an abelian category A is a chain


∂n+1 ∂
A• = · · · ← An+1 ←−−− An ←−
n
An−1 ← · · ·

where ∂n−1 ∂n = 0 for all n.

Note that the subscripts of homology become superscripts in cohomology, and most
objects acquire a “co-” prefix. For example, ker ∂n = Z n (A• ) are called the n-cocycles of
A• , im ∂n+1 = B n (A• ) are the n-coboundaries of A• and these allow us to define

Definition. The nth cohomology of a cochain complex A• is the quotient

H n (A• ) = Z n (A• )/B n (A• )

which is an object in A. Therefore H n is a functor Comp(A) → A.

182
10.2. Derived Functors Chapter 10. Homology

10.2 Derived Functors


Definition. Suppose T : A → C is a left exact functor between abelian categories. Then a
cohomological δ-functor extending T is a family of additive functors (T n : A → C)n∈N
such that for any short exact sequence in A,
0 → A → B → C → 0,
there exists a long exact sequence in C:
δ δ
− T 1A → T 1B → T 1C →
0 → TA → TB → TC → − T 2A → · · ·
Likewise, for a right exact functor T : A → C, a homological δ-functor extending T is
a family (Tn )n∈N such that for every short exact sequence as above, there is a long exact
sequence
δ δ
· · · T2 C →
− T1 A → T1 B → T1 C →
− T A → T B → T C → 0.
To make notation consistent, we will put T 0 = T .
Example 10.2.1. We will see below that there is a cohomological δ-functor ExtnR (M, −)
extending the left exact functor HomR (M, −) : ModR → AbGps. Also, there is a cohomolog-
ical δ-functor ExtnR (−, N ) extending HomR (−, N ) and a homological δ-functor TorR
n (M, −)
extending M ⊗R −.
Example 10.2.2. Let p be prime and consider the functor T : AbGps → AbGps, A 7→ A/pA.
Then there is a δ-functor (T n ) extending T given by definng T 1 A = p A, the p-torsion of
A, and T n A = 0 for all n ≥ 2. Dually, the p-torsion functor S : A 7→ p A is extended
by the δ-functor (S n ) with S 1 A = A/pA and S n A = 0 for n ≥ 2. It turns out that
T n = Extn (Z/pZ, −) and S n = Extn (−, Z/pZ) from the previous example.
Theorem 10.2.3. Let T : A → C be a left (resp. right) exact functor on an abelian category
A with enough injectives (resp. projectives). Then there exists a universal cohomological
(resp. homological) δ-functor (T n ) (resp. (Tn )) extending T and such that T n (E) = 0 for all
injectives E ∈ A (resp. Tn (P ) = 0 for all projectives P ∈ A).
Definition. A cohomological δ-functor (T n ) is effaceable if for any n ≥ 0 and any object
M , there is a monomorphism M ,→ N such that the induced map T n (M ) → T n (N ) is 0.
Proposition 10.2.4. If A is an abelian category with enough injectives, then a δ-functor is
effaceable if and only if it vanishes on all injectives.
Proof. ( =⇒ ) Let (T n ) be a δ-functor and E ∈ A an injective. Then given a monomorphism
E ,→ N , by Proposition 8.2.2 there exists a splitting N → E. Hence N ∼ = E ⊕ N 0 for some
object N 0 , where ⊕ denotes the coproduct. Now by left exactness and Theorem 9.2.10,
T n (N ) ∼
= T n (E) ⊕ T n (N 0 ), and the map T n (E) → T n (N ) ∼= T n (E) ⊕ T n (N 0 ) is 0 by
n
hypothesis, so T (E) = 0.
( ⇒= ) Conversely, suppose T n (E) = 0 for all injectives and take an arbitrary M ∈ A.
Then by Theorem 8.2.14, M ,→ E for some injective E and therefore T n (M ) → T n (E) = 0
must be the zero map.

183
10.2. Derived Functors Chapter 10. Homology

Corollary 10.2.5. Effaceable δ-functors are unique up to unique isomorphism.


Proof. Assume (T n ) is an effaceable δ-functor. Then by Proposition 10.2.4, for any monomor-
phism M ,→ N , the map T 1 (M ) → T 1 (N ) is 0. Thus in the long exact sequence for T n , we
have
0
0 → T 0 (M ) → T 0 (N ) → T 0 (N/M ) → T 1 (M ) → − T 1 (N ) → · · ·
so T 1 (M ) ∼
= T 0 (N/M )/ im T 0 (N ). More generally, for k ≥ 0, T k+1 (M ) → T k+1 (N ) is
0 and this implies T k+1 (M ) ∼= T k (N/M )/ im T k (N ) so the objects T n (M ) are completely
determined up to isomorphism by the objects T 0 (M ) and the choices of monomorphisms
M ,→ N in Proposition 10.2.4.
Corollary 10.2.6. If (T n ) is a cohomological δ-functor on an abelian category A with enough
injectives, then there is a natural isomorphism T n ∼
= H n (T 0 (−)) for all n ≥ 0.
Proof. Take an object M ∈ A. Since A has enough injectives, we may take an injective
resolution of M :
M E1 E2 E3 ···

K1 K2

Then T n (M ) ∼
= T n−1 (K1 ) ∼
= ··· ∼
= T 1 (Kn−1 ) and this last term is
T 0 (Kn ) ∼ ker(T 0 (En ) → T 0 (En+1 ))
T 1 (Kn−1 ) ∼
= = = H n (T 0 (E• )).
im T 0 (En−1 ) im(T 0 (En−1 ) → T 0 (En ))

This proves Theorem 10.2.3 for left exact functors. Dually, we have:
Definition. A homological δ-functor (Tn ) is coeffaceable if for any n ≥ 0 and any object
M , there is an epimorphism N → M such that the induced map Tn (N ) → Tn (M ) is 0.
Proposition 10.2.7. If A is an abelian category with enough projectives, then a δ-functor
is coeffaceable if and only if it vanishes on all projectives.
Corollary 10.2.8. Coeffaceable δ-functors are unique up to unique isomorphism.
Corollary 10.2.9. If (Tn ) is a homological δ-functor on an abelian category A with enough
projectives, then there is a natural isomorphism Tn ∼
= Hn (T0 (−)) for all n ≥ 0.
This proves Theorem 10.2.3 for right exact functors.
Definition. Suppose T : A → C is a covariant, additive, right exact functor between abelian
categories, where A has enough projectives. The nth left derived functor of T is the
universal homological δ-functor of T :

Ln T : A −→ C
A 7−→ Hn (T (P• ))

where P• is a fixed projective resolution of A.

184
10.2. Derived Functors Chapter 10. Homology

Derived functors in some sense measure the failure of exactness in the nth homological
dimension of the functor T .
Definition. In the category R-Mod of left R-modules, if TM = − ⊗R M then its left derived
functors are called Tor:

TorR
n (M, N ) := (Ln TM )(N ) = Hn (P•,N ⊗R N ).

0
Likewise for a right R-module M , the left derived functors of TM = M ⊗R − are called tor:
0
torR
n (N, M ) := (Ln TM )(N ) = Hn (M ⊗R P•,N ).

The notation tor is only temporary, as we will show in Corollary 10.3.11.


The dual of left derived functors is right derived functors, which we define now.
Definition. For a covariant, additive, left exact functor S : A → C between abelian cate-
gories, where A has enough injectives, the nth right derived functor of S is the universal
cohomological δ-functor of S:

Rn S : A −→ C
A 7−→ H−n (S(E• ))

where E• is a fixed injective resolution of A.


The counterpart to Tor is a right derived functor called Ext.
Definition. In R-Mod, the right derived functors of SM = HomR (M, −) are called Ext:

ExtnR (M, N ) := (Rn SM )(N ) = H−n (HomR (M, E•,N )).

We summarize the different variances and types of derived functors in the table below.
Variance of T Category requirements Derived functor
covariant enough projectives Ln T
covariant enough injectives Rn T
contravariant enough projectives Rn T
contravariant enough injectives Ln T

Example 10.2.10. For all right R-modules A and left R-modules B, TorR ∼
0 (A, B) = A ⊗R B.
Likewise, if A and B are both left R-modules then Ext0R (A, B) ∼
= HomR (A, B).
In defining left and right derived functors, we have neglected the fact that we are choosing
a particular projective (or injective) resolution of A. The Comparison Theorem says that
unique chain maps exist between projective resolutions of M and N when M → N is a
module homomorphism, so we need not worry when defining Ln T and Rn T .
Theorem 10.2.11 (Comparison). Let P• : · · · → P2 → P1 → P0 be a projective chain
complex and suppose C• : · · · C2 → C1 → C0 is an acyclic chain complex. Then for any
homomorphism ϕ : H0 (P• ) → H0 (C• ), there is a chain map f : P• → C• whose induced map
on H0 is ϕ, and ϕ is unique up to chain homotopy.

185
10.2. Derived Functors Chapter 10. Homology

Proof. Consider the diagram


∂2 ∂1 ∂0
P2 P1 P0 H0 (P• ) 0

f2 f1 f0 ϕ
∂20 ∂10 ∂00
C2 C1 C0 H0 (C• ) 0

Since P0 is projective, there exists an f0 lifting ϕ to P0 → C0 . Inductively, given fn−1 we


have a diagram
∂n
Pn Pn−1

fn fn−1
∂n0
Cn Cn−1
0
Note that fn−1 ∂ has image lying in ker ∂n0 ⊆ Cn−1 , but C• is acyclic, so ∂n0 (Cn ) = ker ∂n−1 .
Since Pn is projective, we can lift fn−1 ∂n to the desired map fn : Pn → Cn . By construction,
f = {fn }∞
n=0 satisfies the desired properties.
For uniqueness, suppose g : P• → C• is another chain map restricting to ϕ on H0 . Since
f0 − g0 = 0 on H0 , it must be that (f0 − g0 )(P0 ) ⊆ ker ∂00 = im ∂10 , so by projectivity of P0
there exists s0 : P0 → C1 making the following diagram commute:
P0

s0 f0 − g0

C1 im ∂10 0
0
Inductively, given s0 , . . . , sn−1 satisfying fk − gk = ∂k+1 sk + sk+1 ∂k for all 0 ≤ k ≤ n − 1, we
have

∂n0 (fn − gn − sn−1 ∂n ) = (fn−1 − gn−1 )∂n − ∂n0 sn−1 ∂n since f, g are chain maps
= (∂n0 sn−1 − sn−2 ∂n−1 )∂n − ∂n0 sn−1 ∂n = 0.

Hence there is a commutative diagram


Pn

sn fn − gn − sn−1 ∂n

Cn+1 0
ker ∂n0 0
∂n+1
0
This establishes the chain homotopy s : P• → C• such that fn − gn = ∂n+1 sn + sn+1 ∂n for
all n ≥ 0. Hence f is unique up to chain homotopy.

186
10.2. Derived Functors Chapter 10. Homology

Corollary 10.2.12. Let g : M → N be R-linear and pick projective resolutions P• and Q•


of M and N , respectively. Then there exists a chain map f : P• → Q• such that H0 (f ) = g
and f is unique up to chain homotopy.

Proof. Given projective resolutions P• , Q• → M , we have M = H0 (P• ) = H0 (Q• ) so let


ϕ = idM . Since projective resolutions are acyclic, the comparison theorem gives us a chain
map f : P• → Q• . Reversing the roles of P• and Q• gives a chain map in the opposite
direction, and uniqueness forces the composition of these maps to be the identity in either
direction.

Corollary 10.2.13. For R-modules M and N , the assignments (M, N ) 7→ Hn (P• ⊗ N ) and
(M, N ) 7→ H n (Hom(P• , N )) are independent of the deleted projective resolution P• → M
and are functorial in both M and N . In particular, Ext and Tor are well-defined functors.

Proof. If P• , Q• → M are two projective resolutions, then by Corollary 10.2.12, P• and Q•


are chain homotopy equivalent, and it easily follows that P• ⊗ N and Q• ⊗ N , as well as
Hom(P• , N ) and Hom(Q• , N ), are chain homotopy equivalent as well.
Next, a map N → N 0 induces a map P• ⊗N → P• ⊗N 0 which is functorial since ⊗ and H•
are functors to begin with. On the other hand, suppose P• → M and P•0 → M 0 are projective
resolutions. Then given a map M → M 0 , the uniqueness portion of Corollary 10.2.12 says
that the map is functorial.
The proof for Hom is analagous.
Some of the basic facts about derived functors are described in detail in section 6.2 of
Rotman. The following properties of left derived functors have analogs that hold for right
derived functors.

Theorem 10.2.14. If T : A → C is an additive covariant functor between abelian categories,


and A has enough projectives, then each Ln T : A → C is an additive covariant functor.

Proof sketch. Use the fact that Hn is an additive covariant functor Comp(A) → A. Be careful
checking that Ln T is well-defined on chain maps.

Proposition 10.2.15. If T : A → C is an additive covariant functor of abelian categories,


then (Ln T )A = 0 for all A ∈ obj(A) whenever n is negative.

Proof. This is immediate from the definition since the nth term of a resolution P of A is 0
whenever n is negative.
Notice that since TorR R R
• (L, −) is an effaceable δ-functor, Tor1 (L, M ) = 0 implies Torn (L, M ) =
0 for n ≥ 2 and for all modules M . Similarly, Ext1R (L, M ) = 0 implies ExtnR (L, M ) = 0 for
n ≥ 2 and all M .

Definition. The depth of an R-module M , denoted depth(M ), is the smallest n for which
ExtnR (R, M ) is nonzero.

The Krull dimension (see Section 2.3) of an R-module M is defined to be the supremum
of the lengths of all chains of prime ideals. An important result, which we won’t prove, is

187
10.2. Derived Functors Chapter 10. Homology

Theorem 10.2.16. The depth of M is bounded above by the Krull dimension of M .

There is a deep connection between projective dimension and depth.

Theorem 10.2.17 (Auslander-Buchsbaum). For an R-module M with finite projective di-


mension n, n + depth(M ) = depth(R), where R is viewed as a module over itself.

In particular, if the projective dimension of M is finite, we see that

n ≤ depth(R) ≤ dim R.

188
10.3. Tor and Ext Chapter 10. Homology

10.3 Tor and Ext


In this section, we explore the properties of Tor and Ext. Recall the definition in Section 10.2:
TorR
n (M, N ) = Hn (P•,N ⊗R N ).

Theorem 10.3.1. Let R be a ring, A a right R-module and B a left R-module. Then
(i) For all n ≥ 0, TorR ∼ R op
n (A, B) = Torn (B, A).

(ii) If R is commutative, then for all n ≥ 0, TorR ∼ R


n (A, B) = Torn (B, A).

Proof. It suffices to prove (i) since (ii) will follow immediately. We know that every left
R-module is a right Rop -module and likewise every right R-module is a left Rop -module. Let
P• be a projective resolution of A. Then t : P• ⊗R B −→ B ⊗Rop P• is a chain map of
Z-complexes with tn : Pn ⊗R B 7−→ B ⊗Rop Pn defined by pn ⊗ b 7→ b ⊗ pn . It is easy to
see that each tn is an isomorphism, so the complexes are isomorphic and therefore have the
same homology:
TorR ∼ Rop
n (A, B) = Hn (P• ⊗R B) = Hn (B ⊗Rop P• ) = Torn (B, A).

The key property is that Torn vanishes on projective modules of homological degree n ≥ 1,
since a projective resolution P• of a projective module P is simply → 0 → 0 → P → 0.
Proposition 10.3.2. Let R be a ring and let M and N be R-modules.
(a) If M is a free R-module, then TorR n
n (M, N ) = ExtR (M, N ) = 0 for n > 0.

(b) If R is a PID, then TorR n


n (M, N ) = ExtR (M, N ) = 0 for n > 1.

Proof. Apply Corollary 8.1.8.


Theorem 10.3.3 (Long Exact Sequence for Tor). For every short exact sequence
0 → A0 → A → A00 → 0
in ModR and R-module B, TorR
n (−, B) induces an exact sequence
00 0 00 0
· · · → TorR R R R R
n+1 (A , B) → Torn (A , B) → Torn (A, B) → Torn (A , B) → Torn−1 (A , B) → · · ·

ending in · · · → A0 ⊗R B → A ⊗R B → A00 ⊗R B → 0.
Proof. Use the definition of Tor via homology and the long exact sequence from Theo-
rem 10.1.3.
This theorem is true more generally for any left derived functor, and we will see a similar
sequence with Ext and right derived functors. The long exact sequence for Tor is usually
summarized in the literature by saying that the Tor sequence repairs loss of exactness on
the left of tensoring a short exact sequence.
The next result says that TorR R
n (A, −) and Torn (−, B) can be computed using projective
resolutions (or alternatively, flat resolutions) in either slot.

189
10.3. Tor and Ext Chapter 10. Homology

Theorem 10.3.4. For all projective (or flat) resolutions P• of A and Q• of B and for all
n ≥ 0,
Hn (P• ⊗R B) ∼ = TorR ∼
n (A, B) = Hn (A ⊗R Q• ).

Proof. See Rotman, section 7.1.


Proposition 10.3.5. If A and some family {Bi }i∈I are R-modules then for all n ≥ 0 there
are natural isomorphisms
!
Bi ∼
M M
TorR A,
n = TorR (A, Bi ).
n
i∈I i∈I

The same is true for direct sums in the first slot.


Proof. The base case is the natural equivalence between TorR
0 (A, −) and A ⊗R − and the
full theorem can be proven using induction on n.
Example 10.3.6. Let B be an abelian group and consider TorZ1 (Z/nZ, B). For every n
there is an exact sequence
fn
0 → Z −→ Z → Z/nZ → 0
where fn is multiplication by n. Applying − ⊗Z B gives an exact sequence
1⊗fn
Tor1 (Z, B) → Tor1 (Z/nZ, B) → Z ⊗ B −−−→ Z ⊗ B.
Since Z is projective, Tor1 (Z, B) = 0. Moreover 1 ⊗ fn is also just multiplication by n. We
also know that Z ⊗ B = B since in general R ⊗R − is naturally isomorphic to the identity
functor on R-Mod. This shows that Tor1 (Z/nZ, B) = {b ∈ B | nb = 0}.
In the case when A and B are finitely generated abelian groups, we can now compute
Tor1 (A, B). Decompose A and B by the fundamental theorem of finitely generated abelian
groups: M M
A = Zr ⊕ Ai and B = Zs ⊕ Bj
where Ai and Bj are cyclic. Since Tor1 commutes with direct sums,

Tor1 (A, B) ∼ r s
M M
= (Tor1 (Z, B)) ⊕ (Tor1 (A, Z)) ⊕ Tor1 (Ai , B) ⊕ Tor1 (A, Bj ).
i j

Since Z is projective, Tor vanishes on the free parts in the first two summands. Then we’re
left with Tor of cyclic groups. By the work above, each Tor1 (Ai , B) = {b ∈ B | ni b = 0}
where ni = |Ai |. Thus Tor1 (Ai , B) is a cyclic group of order di = gcd(ni , |B|) and the same
holds for the Tor1 (A, Bj ).
Example 10.3.7. If R is an integral domain with field of fractions Q and residue field
K = Q/R, then Tor captures the idea of torsion submodules. In fact, if A is an R-module
and At is its torsion submodule, then TorR ∼
1 (K, A) = At . Moreover, there is an exact sequence

0 → At → A → Q ⊗R A → K ⊗R A → 0
which shows that A is torsion ⇐⇒ Q ⊗R A = 0. This explains the name Tor for in general,
TorR
n is a torsion R-module.

190
10.3. Tor and Ext Chapter 10. Homology

Lemma 10.3.8 (Horseshoe Lemma). Given a short exact sequence of R-modules 0 → A →


B → C → 0 and deleted projective resolutions P• → A and R• → C, there is a deleted
projective resolution Q• → B and a short exact sequence of chain complexes 0 → P• →
Q• → R• → 0 inducing the maps of the original short exact sequence on H0 .

Proof. We must have Q• = P• ⊕R• so that the short exact sequence 0 → P• → Q• → R• → 0


splits. Consider the following diagram with exact columns and exact top and bottom row:
0

ε
··· Pn ··· P1 P0 A 0

i
δ0
Qn ··· Q1 Q0 B 0

Φ j

··· Rn ··· R1 R0 C 0

Our goal is to complete the middle row such that the diagram commutes. We begin by
constructing δ0 : Q0 → B. Since R0 is projective, there exists a map Φ : R0 → B lifting j,
given by the dashed arrow above. Then define δ0 on Q0 = P0 ⊕ R0 by δ0 = (i ◦ ε) ⊕ (−Φ).
Having constructed δ0 , . . . , δn−1 , we have the following portion of the above diagram:
··· Pn Pn−1 Pn−2 ···

δn−1
··· Qn Qn−1 Qn−2 ···

··· Rn Rn−1 Rn−2 ···

Setting K = ker(Pn−1 → Pn−2 ), L = ker(Qn−1 → Qn−2 ) and M = ker(Rn−1 → Rn−2 ), the


above diagram induces a smaller diagram of the same form as the base case:
··· Pn K 0

δn
Qn L

··· Rn M 0

Complete the diagram as in the base case to construct δn .

191
10.3. Tor and Ext Chapter 10. Homology

Corollary 10.3.9. If 0 → N 0 → N → N 00 → 0 is a short exact sequence of R-modules and


M is any R-module, then there is a long exact sequence

→ Torn (M, N 0 ) → Torn (M, N ) → Torn (M, N 00 ) → Torn−1 (M, N 0 ) →

Proof. Take projective resolutions P•0 → N 0 and P•00 → N 00 ; then Lemma 10.3.8 provides a
projective resolution P• for N and a short exact sequence of complexes

0 → P•0 → P• → P•00 → 0.

Since each term in the sequence is a complex of projective modules, applying M ⊗ − yields
a short exact sequence again. Then the desired long exact sequence is merely the long exact
sequence in homology for this induced short exact sequence.

Theorem 10.3.10. For each n ≥ 0, there exists a functor TorR


n : R−Mod×R−Mod → R−Mod
which satisfies

(1) TorR
0 (M, N ) = M ⊗ N .

(2) For any short exact sequence 0 → M 0 → M → M 00 → 0 and any R-module N , there
is a long exact sequence

→ Torn (M 0 , N ) → Torn (M, N ) → Torn (M 00 , N ) → Torn−1 (M 0 , N ) →

which is natural in N .

(3) For any free module F , TorR


n (F, N ) = 0 for all n > 0.

Moreover, any functor satisfying these three properties is naturally isomorphic to TorR
n.

Proof. We have proven that Torn satisfies the stated properties so it remains to show that
these in fact characterize Torn . We prove this inductively on n. For n = 0, uniqueness
follows from the universal property of the tensor product. For n ≥ 1, take modules M and
N and a free module F such that there is an exact sequence 0 → K → F → M → 0. Then
by (2), there is a long exact sequence

0 = Torn (F, N ) → Torn (M, N ) → Torn−1 (K, N ) → Torn−1 (F, N ) → · · ·

When n > 1, Torn−1 (F, N ) = 0 as well so Torn (M, N ) ∼ = Torn−1 (K, N ). When n = 1,
Tor1 (M, N ) ∼
= ker(K ⊗ N → F ⊗ N ). In all cases, induction implies that Torn is determined
as a functor by Torn−1 so uniqueness holds.

Corollary 10.3.11. For any n ≥ 0 and any modules M and N , there is a natural isomor-
phism
Torn (M, N ) ∼
= Torn (N, M ).
Proof. Consider the assignment (M, N ) 7→ torn (M, N ) := Torn (N, M ). Then

(1) tor0 (M, N ) = Tor0 (N, M ) = N ⊗ M ∼


= M ⊗ N.

192
10.3. Tor and Ext Chapter 10. Homology

(2) torn has a long exact sequence in the first variable because Torn has a long exact sequence
in the second variable by Corollary 10.3.9.
(3) For a free module F , torn (F, N ) = Torn (N, F ) = 0 for each n > 0 since − ⊗ F preserves
exactness.
Hence by Theorem 10.3.10, Torn and torn are naturally isomorphic.
Next we shift our focus to Ext. This is in some ways the more interesting of the two
derived functors, since in any module category it naturally admits an R-algebra structure.
As we saw for Tor, Ext defines a long exact sequence which repairs the loss of exactness of
applying the Hom functor to a short exact sequence.
Theorem 10.3.12 (Long Exact Sequence for Ext). If 0 → B 0 → B → B 00 → 0 is a short
exact sequence in R Mod then for every left R-module A, ExtnR (A, −) induces a long exact
sequence

0 → HomR (A, B 0 ) → HomR (A, B) → HomR (A, B 00 )


→ Ext1R (A, B 0 ) → Ext1R (A, B) → Ext1R (A, B 00 )
→ Ext2R (A, B 0 ) → Ext2R (A, B) → Ext2R (A, B 00 ) → · · ·

Proof. Same as above, except reverse the arrows.


Corollary 10.3.13. If 0 → M 0 → M → M 00 → 0 is a short exact sequence of R-modules
and N is any R-module, then there is a long exact sequence

→ Extn (M 00 , N ) → Extn (M, N ) → Extn (M 0 , N ) → Extn+1 (M 00 , N ) →

Proof. Similar to the proof of Corollary 10.3.9.


Theorem 10.3.14. For each n ≥ 0, there exists a functor ExtnR : R−Mod×R−Mod → R−Mod
which satisfies
(1) Ext0R (M, N ) = HomR (M, N ).
(2) For any short exact sequence 0 → M 0 → M → M 00 → 0 and any R-module N , there
is a long exact sequence

→ Extn (M 00 , N ) → Extn (M, N ) → Extn (M 0 , N ) → Extn+1 (M 00 , N ) →

which is natural in N .
(3) For any free module F , ExtnR (F, N ) = 0 for all n > 0.
Moreover, any functor satisfying these three properties is naturally isomorphic to ExtnR .
Proof. Reverse the arrows in the proof of Theorem 10.3.10.
Ext is characterized by the property that Extn vanishes on all injective modules whenever
n ≥ 1, for the same reason that Torn vanishes on projectives. The next proposition shows
that Ext interacts with direct sums and products in the same way that Hom does.

193
10.3. Tor and Ext Chapter 10. Homology

Proposition 10.3.15. For R-modules B and {Ai }i∈I and for all n ≥ 0, there are natural
isomorphisms !
Ai , B ∼
M Y
ExtnR = ExtnR (Ai , B).
i∈I i∈I

As with Hom, things are slightly different in the second slot.

Proposition 10.3.16. For R-modules A and {Bi }i∈I and for all n ≥ 0, there are natural
isomorphisms !
Bi ∼
Y Y
ExtnR A, = ExtnR (A, Bi ).
i∈I i∈I

Example 10.3.17. Fix n ≥ 0, m ≥ 1 and consider ExtnZ (Z/mZ, Z). A simple projective
resolution of Z/mZ is the short exact sequence
m
0→Z−
→ Z → Z/mZ → 0.

Applying Hom(−, Z) and deleting Z/mZ, we obtain a complex


m
0 → Hom(Z, Z) = Z −
→ Hom(Z, Z) = Z → 0.

The homology of this sequence is now easy to calculate:

Ext0 (Z/mZ, Z) = 0 and Ext1 (Z/mZ, Z) = Z/mZ.

In general, for an abelian group B and for any n ≥ 2,

Ext1Z (Z/nZ, B) ∼
= B/nB.

(Compare to Example 10.3.6.) As before, one can use this to show that for any finitely
generated abelian groups A, B, Ext1 (A, B) is the direct sum of Ext1 (Ai , Bj ) where Ai and
Bj are finite cyclic groups.

Definition. For R-modules A, C, an extension of A by C is a short exact sequence


i p
0→A→
− B→
− C → 0.

An extension splits if there exists a map s : C → B such that ps = 1C .

Lemma 10.3.18. If Ext1R (C, A) = 0 then every extension of A by C splits.


i p
Proof. Starting with an extension 0 → A →
− B→− C → 0, apply HomR (C, −) to obtain an
exact sequence
p∗
− Ext1 (C, A) = 0.
ð
Hom(C, B) − → Hom(C, C) →
So p∗ is surjective, meaning there is some s ∈ Hom(C, B) such that 1C = p∗ (s), or 1C = ps.
Hence the extension splits.

194
10.3. Tor and Ext Chapter 10. Homology

The elements in Ext1R (C, A) can be thought of as ‘obstructions’ to the splitting property
of extensions of A by C. This explains where the name Ext comes from. An important
property of Ext is
Proposition 10.3.19. A left R-module P is projective ⇐⇒ Ext1R (P, B) = 0 for every
R-module B. Similarly, a left R-module E is injective ⇐⇒ Ext1R (A, E) = 0 for every
R-module A.
Proof. If P is projective, then Ext1R (P, B) = 0 for all B by Proposition 8.1.3. On the other
hand, if Ext1R (P, B) = 0 for all B then Lemma 10.3.18 shows that every exact sequence
ending in P splits and hence P is projective.
The second statement is proven similarly.
Likewise, we have a homological characterization of flat modules.
Proposition 10.3.20. For an R-module M , the following are equivalent:
(a) M is flat.
(b) TorR
n (N, M ) = 0 for all n ≥ 1 and R-modules N .

(c) TorR
1 (N, M ) = 0 for all R-modules N .

(d) TorR
1 (N, M ) = 0 for all finitely generated R-modules N .

(e) TorR
1 (R/p, M ) = 0 for all prime ideals p ⊂ R.

(f ) For all prime ideals p ⊂ R, the multiplication map p ⊗R M → pM is an isomorphism.


rij x0j for
P P
(g) If ri xi = 0 for some elements r i ∈ R and x i ∈ M , then each x i =
rij ∈ R and x0j ∈ M such that
P
rij ri = 0.
Proof. (a) ⇐⇒ (b) =⇒ (c) =⇒ (d) =⇒ (e) ⇐⇒ (f) and (c) ⇐⇒ (g) are all
immediate from the definitions and the long exact sequence in Tor. In addition, (d) =⇒
(a) is straightforward. Therefore it suffices to prove (e) =⇒ (d). Use Theorem 3.6.3 to get
a prime filtration of N :
0 ⊆ N1 ⊆ N2 ⊆ · · · ⊆ Nk = N
with Ni+1 /Ni ∼ = R/pi for primes pi ⊂ R. If k = 1, then N ∼ = R/p1 and (3) implies
R
Tor1 (N, M ) = 0. Inducting on k, the short exact sequence

0 → Nk−1 → N → R/pk → 0

induces a long exact sequence in TorR


i , part of which reads:

0 = TorR R R
1 (Nk−1 , M ) → Tor1 (N, M ) → Tor1 (R/pk , M ) = 0.

So TorR
1 (N, M ) = 0.

Theorem 10.3.21 (Local Criterion for Flatness). Suppose ϕ : (R, m) → (S, n) is a local
homomorphism of local noetherian rings and N is a finitely generated S-module. Then N is
a flat R-module if and only if TorR
1 (R/m, N ) = 0.

195
10.3. Tor and Ext Chapter 10. Homology

Proof. ( =⇒ ) follows from Proposition 10.3.20.


( ⇒= ) By Proposition 10.3.20, it’s enough to show that for all finitely generated R-
modules, Tor1 (N, M ) = 0. We induct on d = dim(R/ AnnR (N )). If d = 0, let

N0 ⊆ · · · ⊆ N` = N

be a prime filtration of N . Then since r(Ann(N )) = m, the only prime appearing as a


quotient is m. There is a short exact sequence

0 → k = R/m → N → N 0 → 0

so by induction on `, Tor1 (N 0 , M ) = 0 and by hypothesis, Tor1 (k, M ) = 0. Therefore


Tor1 (N, M ) = 0. Now let d > 0; we will show Tor1 (R/p, M ) = 0 for all primes p in a prime
filtration of N (keep the same notation as above). Notice that dim(R/p) ≤ d for any such p.
If p = m, then by hypothesis Tor1 (R/m, M ) = 0. Otherwise, choose x ∈ m r p and consider
the short exact sequence
x
0 → R/p →
− R/p →
− R/(p, x) → 0.

Then dim(R/(p, x)) < dim(R/p) ≤ d so by induction, Tor1 (R/(p, x), M ) = 0. Thus the long
exact sequence in Tor for the above sequence contains
x
Tor1 (R/p, M ) →
− Tor1 (R/p, M ) → Tor1 (R/(p, x), M ) = 0.

That is, Tor1 (R/p, M ) = x Tor1 (R/p, M ) but since this is a finitely generated R-module,
Nakayama’s lemma (Theorem 1.2.1) implies Tor1 (R/p, M ) = 0.

Proposition 10.3.22. Let R be a commutative ring, a ∈ R a non-zero-divisor and M an


R-module. Write a M = {m ∈ M | am = 0} for the a-torsion part of M . Then

(a) R/aR ⊗ M ∼
= M/aM .
(b) Tor1 (R/aR, M ) ∼
= aM .
(c) Hom(R/aR, M ) ∼
= aM .
(d) Ext1 (R/aR, M ) ∼
= M/aM .
Proof. (a) Define a map φ : R/aR × M → M/aM by (r, m) 7→ rm. If r − r0 ∈ aR, then
r − r0 = as for some s ∈ R. Then (r − r0 )m = (as)m = a(sm) ∈ aM so the map is
well-defined on the quotient R/aR. It is also clearly bilinear. By the universal property of
tensor products, this determines a linear map Φ : R/aR ⊗ M → M/aM . If Φ(r ⊗ m) = 0 in
M/aM then Φ(r ⊗ m) = rm ∈ aM . Thus rm = am0 for some m0 ∈ M , and since the tensor
is over R, we can pass elements of R across the tensor:

r ⊗ m = 1 ⊗ rm = 1 ⊗ am0 = a ⊗ m0 = 0 ⊗ m0 = 0 in M/aM.

So Φ is one-to-one. Clearly Φ is also surjective: 1 ⊗ m 7→ m for any m ∈ M . Hence we have


the desired isomorphism.

196
10.3. Tor and Ext Chapter 10. Homology

(b) Consider the exact sequence


a
0→R→
− R→
− R/aR → 0.
Tensoring with M gives a long exact sequence in Tor:
0 = Tor1 (R, M ) → Tor1 (R/aR, M ) → R ⊗ M → R ⊗ M → R/aR ⊗ M → 0.
(By Proposition 10.3.2, since R is a free R-module, Tor1 (R, M ) = 0.) Using the facts that
R⊗M ∼ = M and, from part (a), R/aR ⊗ M ∼ = M/aM , we get an isomorphic exact sequence:
a
0 → Tor1 (R/aR, M ) → M →
− M → M/aM → 0.
Since the map out of Tor1 (R/aR, M ) is injective and the sequence is exact, we see that
a
Tor1 (R/aR, M ) is isomorphic to the kernel of M → − M , which is clearly a M . Hence
Tor1 (R/aR, M ) ∼
= aM .
(c) For each m ∈ M , there is an R-map fm : R → M, r 7→ rm. Moreover, if m ∈ a M , then
fm factors through the quotient R/aR, giving an element f¯m of Hom(R/aR, M ). This de-
termines a map ϕ : a M → Hom(R/aR, M ), m 7→ f¯m . Conversely, each g ∈ Hom(R/aR, M )
determines an element g(1) ∈ a M , since ag(1) = g(a1) = g(a) = 0. This gives ψ :
Hom(R/aR, M ) → a M , and we have
ϕψ(g)(r) = ϕ(g(1)) = f¯g(1) (r) = rg(1) = g(r)
and ψϕ(m) = ψ(f¯m ) = f¯m (1) = 1m = m.
So we see that ϕ and ψ are inverses. Hence Hom(R/aR, M ) ∼
= aM .
(d) Using the same exact sequence
a
0→R→
− R→
− R/aR → 0
and applying Hom(−, M ), we get a long exact sequence in Ext:
0 → Hom(R/aR, M ) → Hom(R, M ) → Hom(R, M ) → Ext1 (R/aR, M ) → Ext1 (R, M ) = 0.
(Again since R is free, Ext1 (R, M ) = 0.) Replacing Hom(R/aR, M ) with a M and Hom(R, M )
with M , we get
a
0 → aM → M → − M → Ext1 (R/aR, M ) → 0.
a
By exactness, Ext1 (R/aR, M ) is isomorphic to the cokernel of the map M →
− M , which is
M/aM by definition.
Example 10.3.23. Let G be an abelian group, with torsion part T (G). For another abelian
group B, applying − ⊗ B to the short exact sequence 0 → T (G) → G → G/T (G) → 0 gives
a long exact sequence in Tor (coefficients in Z):
0 → Tor1 (T (G), B) → Tor1 (G, B) → Tor1 (G/T (G), B) = 0,
with the final 0 coming from Proposition 10.3.2 (using that G/T (G) is torsion-free). Thus
one may replace G by T (G) when computing Tor1 . Further, consider the short exact sequence
0 → Z → Q → Q/Z → 0. Applying T (G) ⊗ − gives another long exact sequence
Tor1 (T (G), Z) → Tor1 (T (G), Q) → Tor1 (T (G), Q/Z) → T (G) ⊗ Z → T (G) ⊗ Q = 0.
= Tor1 (Q/Z, T (G)) ∼
Further, Tor1 (T (G), Q) = 0, so we get Tor1 (Q/Z, G) ∼ = T (G).

197
10.3. Tor and Ext Chapter 10. Homology

Example 10.3.24. Let m ≥ 2 and n ≥ 0 be integers. Then using the above results, we can
compute
(
Z/mZ, n = 0
Torn (Z, Z/mZ) = Torn (Z/mZ, Z) =
0, n>0

0,
 n=0
n
Ext (Z/mZ, Z) = Z/mZ, n = 1

0, n>1

(
Z/mZ, n = 0
Extn (Z, Z/mZ) =
0, n > 0.

Example 10.3.25. Let R be a commutative ring and m ⊂ R a maximal ideal. Let k = R/m
be the residue field. An interesting Ext group in commutative algebra is Ext1R (k, k), as we
will see in more detail in Section 11.1. To compute this group, consider the short exact
sequence
0 → m → R → k → 0.
Since R is a free module over itself, by Theorem 10.3.14, the long exact sequence in ExtR (−, k)
becomes

=
→ HomR (R, k) → HomR (m, k) → Ext1R (k, k) → 0.
0 → HomR (k, k) −

Moreover, since HomR (k, k) ∼


= k and HomR (R, k) ∼ = k, the first arrow is an isomorphism.
By exactness, the second arrow must be 0, so we get

Ext1R (k, k) ∼
= HomR (m, k).

For any R-module N , every map N → R/m factors through N/mN , so we see that HomR (N, R/m) ∼
=
HomR (N/mN, R/m). Setting N = m, we then get HomR (m, k) ∼ = HomR (m/m2 , k) ∼ =
(m/m2 )∗ , the dual space of m/m2 . This is more commonly known as the tangent space to
Spec R at m, Tm Spec R = (m/m2 )∗ . Thus tangent spaces have a homological interpretation.
Example 10.3.26. For an R-module M , we can similarly consider Ext1R (M, M ), which
consists of R-modules M
f for which there is a short exact sequence

0→M →M
f → M → 0.

There is an endomorphism t : M f → M f making M f into a module over the ring R[ε] =


R[t]/(t2 ); in fact, M
f is free over R[ε]. There is a bijective correspondence between elements
1
of ExtR (M, M ) and equivalence classes of free R[ε]-modules M f such that M
f/tMf∼ = M , where
f f 0 f ∼ f0
M and M are equivalent if there exists an isomorphism M = M inducing the identity on
M . One can then view M as a ‘point’ in the ‘space’ of all R-modules and M f as a ‘tangent
vector’ to M in this space. (Officially, M is called a deformation of M .)
f
The deformation M f = M [ε] has an explicit R-module structure given by

α : R −→ End(M [ε]), α(r) · m1 + εm2 = rm1 + εrm2 .

198
10.3. Tor and Ext Chapter 10. Homology

Then other deformations of M come from maps α̃ : R → End(M [ε]) such that α ≡ α̃ mod
ε. That is, α̃ = α + εµ for µ : R → End(M ). For any r ∈ R, m1 , m2 ∈ M ,
α̃(r)(m1 + εm2 ) = rm1 + ε(rm2 + µ(r)µ1 ).
Thus for α̃ to define an R-module structure on M [ε], we must have α̃(r1 )α̃(r2 ) = α̃(r1 r2 ),
i.e. for any m ∈ M ,
α̃(r1 )(r2 m + εµ(r2 )m) = r1 r2 m + εµ(r1 )(r2 m) + εr1 µ(r2 )m
= r1 r2 m + εµ(r1 r2 )m = α̃(r1 r2 )m.
Thus the necessary condition on µ is
µ(r1 )r2 + r1 µ(r2 ) − µ(r1 r2 ) = 0.
This is obviously a cocycle condition, so it can be measured with homology.
On the other hand, when are α̃ and α̃0 equivalent? i.e. when do they determine equivalent
deformations? Let α̃0 = α + εµ0 . For equivalence to be true, there must be some β =
1 + εγ ∈ End(M [ε]) such that β α̃ = α̃0 β. (Notice that β −1 = 1 − εγ since ε2 = 0.) Then
α̃ = β −1 α̃0 β = (1 − εγ)α̃0 (1 + εγ), so the necessary condition on µ and µ0 is
µ = µ0 + ε(αγ − γα).
This proves there is a bijection
{µ : R → End(M ) | µ(r1 )r2 + r1 µ(r2 ) − µ(r1 r2 ) = 0}
Ext1R (M, M ) ←→ .
hαγ − γα | γ ∈ End(M )i
One can also check that this correspondence is compatible with the group operations, where
addition on the right is pointwise addition of the maps µ : R → End(M ).
Now assume R is a finitely presented ring with generators x1 , . . . , xn , and M is both an
R-module and a finite dimensional vector space over some field k (e.g. k is the residue field
of a local ring as in the previous example). A potential R-action on M is specified by a map
Span{x1 , . . . , xn } −→ Endk (M ) ∼
= Mn (k), xi 7−→ Xi
such that the relations on the xi are mapped to 0, i.e. there are polynomials pj for which we
have pj (X1 , . . . , Xn ) = 0. This defines an algebraic set Y , the space of all potential R-actions
on M , which has an action of GL(M ). The quotient space ModR (M ) = Y /GL(M ) is called
the moduli space of R-actions on M . Under the quotient map Y → Y /GL(M ), a tangent
vector to M maps to a tangent vector modulo a tangent vector to GL(M ), which is just a
matrix. These tangent vectors on ModR (M ) are precisely given by certain cocycle conditions
like the one above.
Example 10.3.27. Let Γ be a quiver, that is, an oriented graph.
ˆ 3

ˆ ˆ
1 2
ˆ 4

199
10.3. Tor and Ext Chapter 10. Homology

The graph algebra of Γ is defined to be the span of all oriented paths in Γ over some commu-
tative ring R (usually a field), with multiplication defined by concatenation where possible
(and 0 otherwise) and addition given by formal sums over R. This algebra is written Π(Γ).
The representation theory of Π(Γ) is quite important in modern P representation theory. If 1v
denotes the trivial path at a vertex v ∈ V (Γ), notice that 1 = v∈V (Γ) 1v and 1v 1v0 = δvv0 1v ,
that is, the elements 1v are orthogonal idempotents in Π(Γ). Therefore for any module M
over the path algebra, there is a decomposition

M∼
M
= 1v M.
v∈V (Γ)

Set Mv = 1v M . We can label the vertices of the graph with these submodules:

ˆ M3

ˆ ˆ
M1 M2
ˆ M4

This allows us to study the representations of Π(Γ) in more detail by making choices at each
vertex. For example, let R = k be a field and consider the following module:
ˆ 0

ˆ ˆ
0 k
ˆ 0

This depicts a simple Π(Γ)-module since it is 1-dimensional. (In fact, if Γ has no oriented
cycles, then these are the only simple modules.) Let Sv denote the simple module with k at
vertex v, e.g. S1 is depicted above. We are interested in computing the group ExtΠ(Γ) (Sv , Sv0 )
for different v, v 0 ∈ V (Γ). Consider the short exact sequence

0 → K → Π(Γ) → Sv → 0.

Here, 1w 7→ δvw for any idempotent 1w ∈ Π(Γ). The kernel K is spanned by all paths of
length at least one, together with all 1w for w 6= v. Decompose the path algebra as:
M
Π(Γ) = Pw
w∈V (Γ)

where Pw := Π(Γ)1w . Then each summand Pw is a projective module. Note that HomΠ(Γ) (Pw , M ) ∼
=

Mw for any module M – in general, this is the fact that HomR (Re, M ) = eM for an idem-
potent e. Thus we have a short exact sequence

0 → K → Pv → Sv → 0

200
10.3. Tor and Ext Chapter 10. Homology

which is easier to understand. Applying ExtΠ(Γ) (−, Sv0 ) yields an exact sequence

0 → Hom(Sv , Sv0 ) → Hom(Pv , Sv0 ) → Hom(K, Sv0 ) → Ext(Sv , Sv0 ) → 0

since Pv is projective. Since im Hom(Pv , Sv0 ) = 0, we get an isomorphism

Ext(Sv , Sv0 ) ∼
= Hom(K, Sv0 ).

Further, one can see that K is in canonical bijection with the set of paths in Γ starting at
any w adjacent to v, counted with multiplicity given by the number of edges v → w. That

is, K = v→w Pwjw , where jw is this multiplicity of paths. Thus we can write
L

!
Ext(Sv , Sv0 ) ∼
= Hom(K, Sv0 ) ∼ P jw , Sv0 ∼
M
= Hom w = k `vv0 ,
v→w

where `vv0 represents the number of paths v → v 0 . In practice, these Ext groups contain
enough information (number of paths) to reconstruct the quiver from just a set of vertices.

201
10.4. Universal Coefficient Theorems Chapter 10. Homology

10.4 Universal Coefficient Theorems


Now that we have sufficiently described the functors ⊗ and Hom and their derived functors,
it is natural to ask: what is the relation between H• (C• ⊗ M ) and H• (C• ) ⊗ M for a chain
complex C• and module M ? Likewise, what is the relation between H• (Hom(C• , M )) and
Hom(H• (C• ), M )? The universal coefficient theorems give answers to these questions in
terms of Tor and Ext.
Theorem 10.4.1 (Universal Coefficient Theorem for Homology with Coefficients). For a
free chain complex C• over a PID and a coefficient module M , for each n ≥ 0 there is a
short exact sequence

0 → Hn (C• ) ⊗ M → Hn (C• ⊗ M ) → Tor1 (Hn−1 (C• ), M ) → 0.

Moreover, this short exact sequence splits and is natural in C• and M .


Proof. The map g : Hn (C• ) ⊗ M → Hn (C• ⊗ M ) may be defined by [a] ⊗ m 7→ [a ⊗ m].
Write Zn = ker(∂ : Cn → Cn−1 ), Bn = im(∂ : Cn+1 → Cn ) and Hn = Hn (C• ). Then for each
n ≥ 0, there are exact sequences
i q
0 → Bn →
− Zn →
− Hn → 0 (1)
j ∂
0 → Zn →
− Cn →
− Bn−1 → 0 (2)

where (2) is split. Tensoring with M , we get a diagram with exact rows and columns:
0 0
0

i q
Bn ⊗ M Zn ⊗ M Hn ⊗ M 0
g
j
∂ ∂
Cn+1 ⊗ M Cn ⊗ M Cn−1 ⊗ M
f

0 Tor1 (Hn−1 , M ) Bn−1 ⊗ M Zn−1 ⊗ M

0
0 0

The bottom row is the long exact sequence in Tor coming from the short exact sequence (1),
but it terminates after Tor1 because Zn−1 is free. Also, since (2) is split we get the zeroes at
the top of the middle column and bottom of the right column. By diagram chasing, one can
define f and g and show that the sequence is exact. Moreover, the sequence is split due to
the fact that sequence (2) is split. Finally, naturality follows from naturality of tensor, Hn
and Tor1 .

202
10.4. Universal Coefficient Theorems Chapter 10. Homology

Theorem 10.4.2 (Universal Coefficient Theorem for Cohomology). For a free chain complex
C• over a PID and a coefficient module M , for each n ≥ 0 there is a short exact sequence

0 → Ext1 (Hn−1 (C• ), M ) → H n (Hom(C• , M )) → Hom(Hn (C• ), M ) → 0

which splits and is natural in C• and M .

Proof. Replacing − ⊗ M with Hom(−, M ), Tor1 with Ext1 and reversing the arrows in the
proof of Theorem 10.4.1 gives a diagram:
0 0
0

i q
Hom(Bn , M ) Hom(Zn , M ) Hom(Hn , M ) 0
g
j
∂ ∂
Hom(Cn+1 , M ) Hom(Cn , M ) Hom(Cn−1 , M )
f

0 Ext1 (Hn−1 , M ) Hom(Bn−1 , M ) Hom(Zn−1 , M )

0
0 0

The proof is dual to the proof of Theorem 10.4.1.

Lemma 10.4.3. Let G = F (G) ⊕ T (G) be a finitely generated abelian group, where F (G)
is the free part of G and T (G) is the torsion part of G. Then Hom(G, Z) ∼= F (G) and
Ext1 (G, Z) ∼
= T (G).

Proof. Use the universal coefficient theorem, Example 10.3.24 and additivity.

Corollary 10.4.4. Suppose C• is a chain complex and M is a module. If Hn (C• ; M ) and


Hn−1 (C• , M ) are finitely generated, then so is H n (C• , M ) and there is an isomorphism

H n (C• , M ) ∼
= Fn ⊕ Tn−1 ,

where Fn = F (Hn (C• , M )) and Tn−1 = T (Hn−1 (C• , M )).

Note that the isomorphism in Corollary 10.4.4 is not canonical.

203
Chapter 11

Ring Homology

In this chapter we use homological algebra to study several types of singularities in commu-
tative algebra. The strongest type of nonsingularity is regularity: as defined in Section 6.2, a
local ring (A, m, k) is regular if its Krull dimension equals the dimension of the k-vector space
m/m2 . By Theorem 6.2.8, this is equivalent to m being generated by dim A elements. The
dimension theory of Section 6.2 can be weakened in several ways to yield different notions
of singularity, including the Cohen-Macauley and Gorenstein conditions.
Note to the reader: in the first two sections, I use R and S to denote rings and A and B
for modules, in contrast to later sections. Sorry for the confusion!

204
11.1. Dimensions of Rings Chapter 11. Ring Homology

11.1 Dimensions of Rings


Homology, especially derived functors like Ext and Tor, is used to define various dimensions
of a ring R, which measure how far R is from having some nice characteristic. We have
already introduced projective dimension, which was defined by
Definition. If A is a left R-module and · · · → P2 → P1 → P0 → A → 0 is a projective
resolution of A, we say A has projective dimension n if there is a smallest n such that Kn
is projective. This is written pd(A) = n. If no such finite n exists, A has infinite projective
dimension.
In commutative algebra we are often interested in proving a theorem of the same flavor
as the one below.
Theorem 11.1.1. The following are equivalent for a left R-module A.
(1) pd(A) ≤ n.

(2) ExtiR (A, B) = 0 for all left R-modules B and all i ≥ n + 1.

(3) Extn+1
R (A, B) = 0 for all left R-modules B.

(4) There exists a projective resolution P• of A such that the (n − 1)st syzygy Kn−1 is
projective.

(5) In every projective resolution of A the (n − 1)st syzygy is projective.


Proof. (1) =⇒ (2) By assumption there is a projective resolution Pn → · · · → P0 → A → 0
with Pi = 0 for all i ≥ n + 1. Therefore the maps d∗i : HomR (Pi−1 , B) → HomR (Pi , B)
induced by HomR (−, B) are all 0 for i ≥ n + 1. Hence ExtiR (A, B) = 0 for all i ≥ n + 1.
(2) =⇒ (3) is immediate.
(3) =⇒ (4) Let Kn−1 be the (n − 1)st syzygy of a projective resolution of A. Since
Extn+1
R (A, B) = 0 for all B, by properties of right derived functors we have

Extn+1 ∼ 1
R (A, B) = ExtR (Kn−1 , B) = 0.

Hence Kn−1 is projective.


(4) =⇒ (5) follows from Proposition 9.4.5.
(5) =⇒ (1) Let P• = · · · → P1 → P0 → A → 0 be a projective resolution of A. If Kn−1
is the (n − 1)st syzygy of P• then

0 → Kn−1 → Pn−1 → · · · → P1 → P0 → A → 0

is exact. By assumption Kn−1 is projective, so this is a finite projective resolution of A and


hence pd(A) ≤ n.

Proposition 11.1.2. Suppose 0 → M 0 → M → M 00 → 0 is an exact sequence of left


R-modules. If two of the modules have finite projective dimension, so does the third.

205
11.1. Dimensions of Rings Chapter 11. Ring Homology

Proof omitted.
The proof of the next proposition shows that the long exact Ext sequence can be useful
when dealing with projective dimension.
Proposition 11.1.3. Let 0 → M 0 → M → M 00 → 0 be an exact sequence of left R-modules.
(i) If pd(M 0 ) < pd(M ) then pd(M 00 ) = pd(M ).
(ii) If pd(M 0 ) > pd(M ) then pd(M 00 ) = pd(M 0 ) + 1.
(iii) If pd(M 0 ) = pd(M ) then pd(M 00 ) ≤ pd(M 0 ) + 1.
Proof. Let 0 → M 0 → M → M 00 → 0 be exact, take A to be a left R-module and consider
the long exact sequence for Ext:
0 → Hom(M 00 , A) → Hom(M, A) → Hom(M 0 , A)
→ Ext1 (M 00 , A) → Ext1 (M, A) → Ext1 (M 0 , A) → · · ·
· · · → Extn (M 00 , A) → Extn (M, A) → Extn (M 0 , A)
→ Extn+1 (M 00 , A) → Extn+1 (M, A) → Extn+1 (M 0 , A) → · · ·
(i) Let pd(M ) = n and suppose pd(M 0 ) < n. Then Extn (M 0 , A) = 0 and Extn+1 (M, A) =
0 so exactness implies Extn+1 (M 00 , A) = 0. Thus pd(M 00 ) ≤ n. Notice that since n is the
projective dimension of M , Extn (M, A) 6= 0. Then we cannot have Extn (M 00 , A) = 0 since
the right arrow of then nth row
· · · → Extn (M 00 , A) → Extn (M, A) → 0 → · · ·
is a surjection and hence nonzero. Therefore pd(M 00 ) = n.
(ii) Now suppose pd(M 0 ) > n. Let pd(M 0 ) = N so that ExtN +1 (M 0 , A) = 0. Since
N > n, ExtN (M, A) and ExtN +1 (M, A) are both 0 so the N th and (N + 1)st rows of the
long sequence look like
· · · ExtN (M 00 , A) → 0 → ExtN (M 0 , A) → ExtN +1 (M 00 , A) → 0 → 0 → · · ·
By exactness, this shows that ExtN (M 0 , A) ∼ = ExtN +1 (M 00 , A) and in fact this will hold for
Extk (M 0 , A) and Extk+1 (M 00 , A) for all k ≥ N (all of the Exts on M in between will be
0). Hence ExtN +1 (M 00 , A) 6= 0 and Extk (M 00 , A) = 0 for all k > N + 1 so by definition
pd(M 00 ) = N + 1.
(iii) Finally, suppose pd(M 0 ) = n. Then Extk (M 0 , A) = 0 = Extk (M, A) for all k ≥ n + 1.
Consider
· · · → Extn (M 00 , A) → Extn (M, A) → Extn (M 0 , A)
→ Extn+1 (M 00 , A) → 0 → 0
→ Extn+2 (M 00 , A) → 0 → 0 → · · ·
As before, the 0’s on either side show that Extn+2 (M 00 , A) = 0 which is enough to prove that
pd(M 00 ) ≤ n + 1.
The proofs when any of the projective dimensions are infinite follows from Proposi-
tion 11.1.2.

206
11.1. Dimensions of Rings Chapter 11. Ring Homology

Example 11.1.4. For (R, p, k) a local ring, pick a projective resolution P• of k over R.
Since − ⊗R Rp is exact, P• ⊗R Rp is a projective resolution of k ⊗R Rp . Then we can compute

ExtiRp (kp , kp ) = H−i HomRp (P• ⊗R Rp , kp )



= H−i (HomR (P• , k))p
= H−i (HomR (P• , k) ⊗R Rp )

= (H−i HomR (P• , k)) ⊗R Rp since ⊗R commutes with homology
= ExtiR (k, k) ⊗ Rp
= ExtiR (k, k) p .


Thus Ext commutes with localization.


(aij )
Example 11.1.5. Let F −−→ G be a map of left modules over (R, m, k) a local noetherian
ring. Tensoring with k yields
(āij )
F ⊗R k −−→ G ⊗R k.
Thus for any finitely generated left R-module M , TorR ∼ rank Fi where Fi is the ith
i (M, k) = k
free module in a minimal free resolution of M . Likewise, applying HomR (−, R/m) gives us
(āji )
HomR (G, k) −−→ HomR (F, k)

so ExtiR (M, k) ∼
= k rank Fi as well. This proves (2) of Theorem ??.
In practice, we apply this to an exact sequence 0 → M 0 → M → M 00 → 0 to produce

0 → Tor0 (M 0 , k) → Tor0 (M, k) → Tor0 (M 00 , k)


→ Tor1 (M 0 , k) → Tor1 (M, k) → Tor1 (M 00 , k) → · · ·
· · · → Torn (M 0 , k) → Torn (M, k) → Torn (M 00 , k) → 0

where n = max{pd(M ), pd(M 0 ), pd(M 00 )}. The above shows that the Tor are all vector
spaces, and rank is additive along exact sequences, so
n
X n
X
i
(−1) βi (M ) = (−1)i (βi (M 0 ) + βi (M 00 ))
i=0 i=0
X
where βi denotes the ith Betti number of the module in its parentheses. The sum (−1)i βi (M )
is called the Euler characteristic of M , denoted χ(M ). We have therefore shown that
χ(M ) = χ(M 0 ) + χ(M 00 ), i.e. Euler characteristic is additive along short exact sequences.

207
11.2. Hilbert’s Syzygy Theorem Chapter 11. Ring Homology

11.2 Hilbert’s Syzygy Theorem


M
Recall that for a graded ring R, M is a graded R-module if M = Mi for submodules Mi
i∈Z
that satisfy Ri Mj ⊆ Mi+j for all i, j. In proving his Syzygy Theorem, Hilbert was trying
to answer the following question: If M is a finitely generated graded R-module, is there a
recurrence formula for the dimensions of the Mi ?

Definition. Given a graded R-module M , its Hilbert series is the formal power series
X
HM (t) = (dim Mi )ti .
i∈Z

Example 11.2.1. If R = k[x1 , . . . , xn ] as above, then


1
HR (t) = 1 + 3t + 6t2 + 10t3 + . . . = .
(1 − t)3

Hilbert’s question can be restated by asking if there is a recurrence relation among the
coefficients of HM (t). This in turn is equivalent to asking if HM (t) is rational.

Proposition 11.2.2. Suppose we have an exact sequence of graded modules

0 → M 0 → M → M 00 → 0

where the arrows are graded homomorphisms. Then HM (t) = HM 0 (t) + HM 00 (t).

Proof. For every i, we get a short exact sequence of vector spaces

0 → Mi0 → Mi → Mi00 → 0

and dimension is additive along short exact sequences of vector spaces.


More generally, one can show

Proposition 11.2.3. Suppose C• = 0 → M1 → · · · → Mk → 0 is a complex of graded


R-modules. Then X X
(−1)i HMi (t) = (−1)i HH i (C• ) (t).
i∈Z i∈Z

Suppose M is a graded left R-module generated by the elements {m1 , . . . , m` } which


have degrees b01 , . . . , b0` . Consider the following projective (free) resolution of M .
L L
0 M j R(−b 0j ) j R(−b1j ) ···

K1

0 0

208
11.2. Hilbert’s Syzygy Theorem Chapter 11. Ring Homology

The ring R(−bij ) is called a shifting of the graded ring R, which is used to line up the
degrees of the mi with the graded pieces of the ring. One can prove

Lemma 11.2.4. If k ∈ Z then HR(−k) (t) = tk HR (t).

tbij
In particular, HR(−bij ) (t) = tbij HR (t) = for all i, j. Hilbert’s proof of the Syzygy
(1 − t)n
Theorem boiled down to showing
X (−1)i+1 tbij
HM (t) =
i,j
(1 − t)n

whenever pd(M ) ≤ n. In modern language, this can be stated in terms of global dimension.

Definition. For a ring R, its global dimension is

gldim(R) = sup{pd(A) | A is an R-module}.

Theorem 11.2.5. Let R be a ring. Then

(i) For any R-module A, pd(A) = max{n ≥ 0 | ExtnR (A, M ) 6= 0 for some M }.

(ii) gldim(R) = max{n ≥ 0 | ExtnR (A, M ) 6= 0 for some A, M }.

Proposition 11.2.6. For n ∈ N,

(a) If k is a field of characteristic p | n then k[Z/nZ] has infinite global dimension.

(b) Z[Z/nZ] has infinite global dimension.

One defines injective and flat dimension in analogy with projective dimension:

Definition. If A is an R-module and 0 → A → E0 → E1 → E2 → · · · is an injective


resolution of A, then A has injective dimension id(A) = n if there is a smallest n such
that Kn is injective. Likewise, if · · · → F2 → F1 → F0 → A → 0 is a flat resolution of A,
then A has flat dimension fd(A) = n if there is a smallest n such that Kn is flat. If no
such n exists, we say A has infinite injective/flat dimension.

Theorem 11.2.7. The following are equivalent for a left R-module A:

(1) id(A) ≤ n.

(2) ExtiR (B, A) = 0 for all left R-modules B and all i ≥ n + 1.

(3) Extn+1
R (B, A) for all left R-modules B.

(4) There exist an injective resolution E• of A such that the (n − 1)st cosyzygy Kn−1 is
injective.

(5) In every injective resolution of A the (n − 1)st cosyzygy is injective.

209
11.2. Hilbert’s Syzygy Theorem Chapter 11. Ring Homology

Proof. Similar to Theorem 11.1.1.


Theorem 11.2.8. For any ring R,

max{id(M ) | M is an R-module} = max{pd(M ) | M is an R-module}


= max{pd(R/I) | I ⊆ R is a left ideal}
= max{n ≥ 0 | ExtnR (A, B) 6= 0 for some R-modules A, B}.

Proof. By Theorems 11.1.1 and 11.2.7,

max{pd(M ) | M ∈ R-Mod} ≤ n ⇐⇒ max{n | ExtnR (A, B) 6= 0} ≤ n


⇐⇒ max{id(M ) | M ∈ R-Mod} ≤ n.

Further, max{pd(R/I) | I ⊆ R} ≤ max{pd(M ) | M ∈ R-Mod} is trival. To complete the


circle, assume e = max{pd(R/I) | I ⊆ R} < id(N ) for some R-module N . If

0 → N → E0 → E1 → · · · → Ee−1 → C

is an injective resolution of N of length e − 1 with (e − 1)st cosyzygy C, then C must not


be injective since ExtnR (C, N ) 6= 0 for some n > e. By Baer’s criterion (Theorem 8.2.3),
Ext1R (R/I, C) 6= 0 implies Exte+1
R (R/I, N ) 6= 0, which implies pd(R/I) > e, a contradiction.
Hence e ≥ id(N ) for all R-modules N , so e ≥ max{id(N ) | N ∈ R-Mod} = max{pd(M ) |
M ∈ R-Mod}, completing the proof.
Example 11.2.9. R has global dimension 0 if and only if R is a semisimple ring.
Example 11.2.10. R has global dimension 1 if and only if every submodule of a projective
R-module is projective. Such a ring is called hereditary. An example of a hereditary ring is
a Dedeking domain, e.g. the ring of integers in a number field or the coordinate ring of a
smooth algebraic curve.
Theorem 11.2.11 (Hilbert’s Syzygy Theorem). For any ring R,

gldim k[x1 , . . . , xn ] = gldim k + n.

In particular, if k is a field then gldim k[x1 , . . . , xn ] = n.


Example 11.2.12. For a graded module M over S = R/I where R = k[x1 , . . . , xn ], define
βi to be the rank of the ith free S-module in a minimal free resolution of M . These are called
the Betti numbers of M . One can prove that βi = dimk ExtiS (M, k). In other words, if I
contains anything at all, there exists an S-module M which has infinite projective dimension
by Theorem ??. The question that arises from this is
Question (Serre-Kaplansky). Is the Poincaré-Betti series
X
PSM (t) = βi ti
i≥0

rational? That is, is there some recurrence on the size of the free modules in a minimal
resolution of M ?

210
11.2. Hilbert’s Syzygy Theorem Chapter 11. Ring Homology

Anick disproved this in 1982 by finding a counterexample in algebraic topology whose


Poincaré-Betti series is irrational. However, there are still interesting cases when PSM (t) is
rational.

1 If S is what is called a complete intersection, then PSM (t) is rational.

2 (Berglund) If S = R/I with I a monomial ideal, then PSk (t) is rational.

3 Let S = k[x, y]/(x2 , xy) and consider


x y −y 0 0
 

(x y) 2
( x0 y0 −y
x )
0 0 x x 0
0 0 0 0 x
0 ← k ← S ←−−− S ←−−−− −− S 3 ←−−−−−−−− S 5 ← S 8 ← · · ·

This is a minimal resolution of k and we can see that the ith map relates the (i − 1)st
map in a recursive way. Interestingly, the ranks of the free modules are the Fibonacci
numbers (starting at the second 1). Then the Poincaré-Betti series is rational:
1
PSk (t) = 1 + 2t + 3t2 + 5t3 + 8t4 + . . . = .
1 − t + t2

Amazingly, PSk (t) is an analytic function with radius of convergence (lim sup n βn)−1 .
This gives a nice characterization of the growth of the Betti numbers βi .

Definition. The curvature of an R-module M is curv(M ) = lim sup n βn , where βi are
the Betti numbers for M .

A powerful characterization theorem using curvature is

Theorem 11.2.13. Let S = k[x1 , . . . , xn ]/I. Then

(1) curv k = 0 ⇐⇒ S is a regular ring.

(2) curv k = 1 ⇐⇒ S is a complete intersection.

211
11.3. Regular Sequences and the Koszul Complex Chapter 11. Ring Homology

11.3 Regular Sequences and the Koszul Complex


Definition. An element x ∈ A is a nonzero divisor for a module M if xm = 0 implies
x
m = 0 for all m ∈ M , or equivalently, if 0 → M → − M is exact. A regular sequence on
M is a sequence of elements (x1 , . . . , xn ) in A satisfying:

(1) x1 is a nonzero divisor for M and for each 2 ≤ i ≤ n, xi is a nonzero divisor for
M/(x1 , . . . , xi−1 )M .

(2) (x1 , . . . , xn )M 6= M .

Remark. If M is finitely generated and each xi lies in the Jacobson radical J(A), then
condition (2) is guaranteed.

Definition. A local noetherian ring A is Cohen-Macauley (abbreviated CM) if there is a


system of parameters x1 , . . . , xd ∈ A which forms a regular sequence on A.

Example 11.3.1. If (A, m) is a 0-dimensional local ring then A is trivially Cohen-Macauley.

Example 11.3.2. If (A, m) is a 1-dimensional local integral domain, then every nonzero
element is a nonzero divisor, so A is Cohen-Macauley.

Example 11.3.3. Let A = C[x4 , y 4 , x3 y, xy 3 ]m where m = (x4 , y 4 , x3 y, xy 3 ). Then A is a


local subring of C[x, y](x,y) , so dim A = 2. Note that x4 , y 4 is a system of parameters for A,
since
(x3 y)4 = (x4 )3 y 4 ∈ (x4 , y 4 ) and likewise (xy 3 )4 = x4 (y 4 )3 ∈ (x4 , y 4 ).
However (x4 , y 4 ) is not a regular sequence on A because y 4 is a zero divisor for A/(x4 ):

y 4 (x3 y)2 = x4 (xy 3 )2 = 0

but x3 y 6= 0 in A/(x4 ).
x
Lemma 11.3.4. An element x ∈ A is a nonzero divisor if and only if 0 → A →
− A is exact,
and in this case the free resolution
x
0→A→
− A→
− A/(x) → 0

is exact.

Proof. Easy.
More generally, we will show that (x1 , . . . , xn ) is a regular sequence on A if and only if
the induced free resolution
(x1 ··· xn )
0 → Amn → · · · → Am1 −−−−−−→ A → A/(x1 , . . . , xn ) → 0

is exact.

212
11.3. Regular Sequences and the Koszul Complex Chapter 11. Ring Homology

Example 11.3.5. For a regular sequence (x, y) on A, this free resolution is


 
−y 
  
x x y
0 −→ A −−−−→ A2 −−−−−→ A −→ A/(x, y) −→ 0
which is exact by direct calculation.
For an element x ∈ A, let K• (x; A) denote the chain complex
x
0→A→
− A→0
and for any chain complex C• of A-modules, write C• (x) = C• ⊗ K• (x; A), where ⊗ here
means the tensor product of chain complexes. Then for each n ∈ Z, Cn (x) ∼
= Cn ⊕ Cn−1 and
the differential on C• (x) is given by
d : Cn (x) −→ Cn−1 (x)
(cn , cn−1 ) 7−→ (dcn + (−1)n−1 xcn−1 , dcn−1 ).
Proposition 11.3.6. For any x ∈ A and any chain complex of A-modules C• , there is a
long exact sequence
x
· · · → Hn (C• ) → Hn (C• (x)) → Hn−1 (C• ) →
− Hn−1 (C• ) → Hn−1 (C• (x)) → · · ·
Proof. Apply homology to the short exact sequence of complexes
0 → C• → C• (x) → C• (−1) → 0
and use Hn (C• (−1)) ∼
= Hn−1 (C• ).
Corollary 11.3.7. For any x ∈ A and any chain complex of A-modules C• , there is a short
exact sequence
0 → Hn (C• )/xHn (C• ) → Hn (C• (x)) → AnnHn−1 (C• ) (x) → 0.
The complex K• (x; A) is an example of the Koszul complex, which we define next.
Definition. The Koszul complex for a sequence of elements (x1 , . . . , xn ) in A is the chain
complex K• (x1 , . . . , xn ; A) defined inductively by
K• (x1 , . . . , xn ; A) = K• (x1 , . . . , xn−1 ; A)(xn ).
If M is an A-module, write K• (x1 , . . . , xn ; M ) = K• (x1 , . . . , xn ; A) ⊗A M .
L
Remark. In general K• (x1 , . . . , xn ; M ) and n (Kn (x1 , . . . , xn ; A) ⊗A M ) are not the same
chain complex. Keep in mind that K• ⊗A M is the tensor product of K• with the complex
0 → M → 0.
Lemma 11.3.8. For any sequence V of elements (x1 , . . . , xn ) in A, K• (x1 , . . . , xn ; A) is iso-
morphic to the exterior algebra An . In particular, the graded pieces of the Koszul complex
for (x1 , . . . , xn ) are
n
Ki (x1 , . . . , xn ; A) = A( i ) .

213
11.3. Regular Sequences and the Koszul Complex Chapter 11. Ring Homology

Corollary 11.3.9. For any sequence of elements (x1 , . . . , xn ) in A, K• (x1 , . . . , xn ; A) is a


graded A-algebra as well as a finitely generated free resolution of K0 (x1 , . . . , xn ; A) = A.

Theorem 11.3.10. Let x1 , . . . , xn ∈ A and let M be a finitely generated A-module. Then

(1) H0 (K• (x1 , . . . , xn ; M )) ∼


= M/(x1 , . . . , xn )M .
(2) Hn (K• (x1 , . . . , xn ; M )) ∼
= AnnM (x1 , . . . , xn ) = {m ∈ M | xi m = 0 for all 1 ≤ i ≤ n}.
(3) If (x1 , . . . , xn ) is a regular sequence on M , then Hi (K• (x1 , . . . , xn ; M )) = 0 for all
i > 0.

(4) If A is noetherian and x1 , . . . , xn ∈ J(A) then the converse of (3) is true: if K• (x1 , . . . , xn ; M )
is acyclic then (x1 , . . . , xn ) is a regular sequence on M .

Proof. Write Hi (x1 , . . . , xn ; M ) := Hi (K• (x1 , . . . , xn ; M )). If n = 1, then (1) – (3) are
immediate for the Koszul complex
1 x
0 → M −→ M → 0.

Further, if H1 (x1 ; M ) = 0 then x1 is a nonzero divisor for M . Since A is noetherian and


x1 ∈ J(A), Nakayama’s lemma (Theorem 1.2.1) implies M 6= x1 M , assuming M 6= 0. Now
assume n > 1 and induct.
(1) Let C• (M ) = K• (x1 , . . . , xn−1 ; M ) for any finitely generated A-module M and set
C• = C• (A). Notice that K• (x1 , . . . , xN −1 ; M ) = C• ⊗A M as chain complexes, so

C• (M )(xn ) = K• (x1 , . . . , xn−1 ; M )(xn ) = K• (x1 , . . . , xn ; A) ⊗A M = K• (x1 , . . . , xn ; M ).

Applying Corollary 11.3.7 with n = 0, we get

0 → H0 (C• (M ))/xn H0 (C• (M )) → H0 (x1 , . . . , xn ; M ) → AnnH−1 (C• ) (xn ) = 0

and by induction, H0 (C• (M )) ∼


= M/(x1 , . . . , xn−1 )M so we get

H0 (x1 , . . . , xn ; M ) ∼
= M/(x1 , . . . , xn ; M ).

(2) follows from the same argument.


(3) For each i ≥ 1, induction on the short exact sequence in Corollary 11.3.7 implies
Hi (x1 , . . . , xn ; M ) = 0. Then H1 (x1 , . . . , xn ; M ) ∼ = AnnH0 (C• ) (xn ) so by (1), H0 (C• ) =
M/(x1 , . . . , xn−1 )M . Finally, since (x1 , . . . , xn ) is a regular sequence, we have

AnnH0 (C• ) (xn ) = 0

as desired.
(4) Again, the short exact sequence in Corollary 11.3.7 implies Hi (C• )/xn Hi (C• ) = 0 for
all i ≥ 1 and Nakayama’s lemma then implies Hi (C• ) = 0 for all i ≥ 1. By the inductive
hypothesis, (x1 , . . . , xn−1 ) is a regular sequence and we are assuming H1 (x1 , . . . , xn ; M ) =
AnnH0 (C• ) (xn ) = 0, so it follows that xn is a nonzero divisor on M/(x1 , . . . , xn−1 )M and we
are done.

214
11.3. Regular Sequences and the Koszul Complex Chapter 11. Ring Homology

Corollary 11.3.11. If (x1 , . . . , xn ) is a regular sequence on A then the Koszul complex


K• (x1 , . . . , xn ; A) is a free resolution of A/(x1 , . . . , xn ).
Example 11.3.12. Let (A, m, k) be a complete local noetherian ring containing a coefficient
field, so that k ,→ A, and suppose A is Cohen-Macauley of dimension d. Fix a system of
parameters x1 , . . . , xd ∈ A that also forms a regular sequence. We know by Cohen’s third
structure theorem (6.3.11) that B = k[[x1 , . . . , xd ]] → A is a finite ring extension. We claim
that A is flat as a B-module. Notice that (x1 , . . . , xd ) is a regular sequence on B, so that
K• (x1 , . . . , xd ; B) is a free resolution of B/(x1 , . . . , xd ) ∼
= k. Then

TorB ∼ ∼
1 (k, A) = H1 (K• (x1 , . . . , xd ; B) ⊗B A) = H1 (K• (x1 , . . . , xd ; A)) = 0

since (x1 , . . . , xd ) is a regular sequence on A. By Theorem ??, this shows A is a flat B-module.
Conversely, if A is flat over B, then Tori (B)(k, A) = 0 for all i ≥ 1 which means
Hi (x1 , . . . , xd ; A) = 0 for all i ≥ 1. Thus by Theorem 11.3.10(4), (x1 , . . . , xd ) is a regu-
lar sequence. This shows that the Cohen-Macauley property (in this context) is equivalent
to being a flat B-module, where B is a power series ring. Even better, a finitely generated
flat module over any local noetherian ring is free, so the Cohen-Macauley property can be
detected locally by a certain freeness condition.
Example 11.3.13. Let B = k[[x4 , y 4 ]] and A = k[[x4 , y 4 , x3 y, xy 3 ]] be the completions of
the polynomial rings in Example 11.3.3. Then

A = B + (x3 y)B + (xy 3 )B + (x6 y 2 )B + (x2 y 6 )B

but A is not free as a B-module, since for example y 4 (x6 y 2 ) − x4 (x2 y 6 ) = 0. Hence A is not
Cohen-Macauley.
Definition. A minimal free resolution of an A-module M is a free resolution
ϕ2 ϕ1 ϕ0
· · · → F2 −→ F1 −→ F0 −→ M → 0

in which ϕi (Fi ) ⊆ mFi−1 for each i ≥ 1.


Let (A, m) be a local noetherian ring and M a finitely generated A-module. We construct
a minimal free resolution of M as follows. Let b0 = dimk M/mM . If {m̄1 , . . . , m̄b0 } is a k-
basis of M/mM , lift it to a generating set m1 , . . . , mb0 (by Proposition 1.2.4) and define
ϕ0 : Ab0 → M by ei 7→ mi . Set M1 = ker ϕ0 and repeat this procedure inductively to
construct
ϕ2 ϕ1 ϕ0
· · · → Ab2 −→ Ab1 −→ Ab0 −→ M → 0.
This sequence is a free resolution by observation and we have built it so that ϕi (Abi ) ⊆
m(Abi−1 ) at each step.
Example 11.3.14. If (x1 , . . . , xn ) is a regular sequence on A, then K• (x1 , . . . , xn ; A) is a
minimal free resolution of A/(x1 , . . . , xn ).
ϕ2 ϕ1 ϕ0
Definition. If · · · → Ab2 −→ Ab1 −→ Ab0 −→ M → 0 is a minimal free resolution of M ,
then bi is called the ith Betti number of M , sometimes written βi (M ).

215
11.3. Regular Sequences and the Koszul Complex Chapter 11. Ring Homology

The Betti numbers of a module have a homological characterization: note that TorA
i (k, M )
is the ith homology of the chain complex
ϕi+1 ϕ
· · · → k bi+1 −−→ k bi −
→i
k bi−1 → · · ·

but by minimality, each ϕi is the zero map. Hence TorA ∼ bi


i (k, m) = k .

216
11.4. Projective Dimension Chapter 11. Ring Homology

11.4 Projective Dimension


ϕ2 ϕ1 ϕ0
Definition. If M is an A-module and · · · → P2 −→ P1 −→ P0 −→ M → 0 is a projective
resolution of M , we say M has projective dimension pdA (M ) = n if there is a smallest n
such that ker ϕn is projective. Otherwise M has infinite projective dimension (or no projective
dimension).
Example 11.4.1. pdA (M ) = 0 if and only if M is itself projective.
Proposition 11.4.2. For a local noetherian ring (A, m, k) and a finitely generated A-module
M , pdA (M ) = sup{i | TorA
i (k, M ) 6= 0}.

Proof. Since projective modules are flat and k ⊗A − is exact, we have automatically that
pdA (M ) ≤ sup{i | TorA i (k, M ) 6= 0}. Conversely, set n = pdA (M ) and induct on n; the base
ϕ2 ϕ1 ϕ0
case follows from Nakayama’s lemma. When n ≥ 1, let · · · → P2 −→ P1 −→ P0 −→ M → 0 be
a projective resolution of M and set Ki = ker ϕi . Then since P• is also a flat resolution of M ,
we have TorA ∼ A A
i+2 (k, M ) = Tor1 (k, Ki ). The base case shows that Ki is free, so Tori (k, M ) = 0
for all i > n, proving the opposite inequality.
Corollary 11.4.3. For a local noetherian ring (A, m, k) and a finitely generated A-module
M , pdA (M ) = n < ∞ if and only if M has a minimal free resolution of length n.
We will need the following lemma to establish Serre’s inequality for dimk TorA
i (k, k).

Lemma 11.4.4. Let (A, m, k) be a local noetherian ring and ϕ : F → F 0 a morphism of


finitely generated free A-modules. If the base change ϕ : F ⊗A k → F 0 ⊗A k is injective, then
ϕ splits.
Proof. Write F = An , F 0 = Am and let e1 , . . . , en be the standard basis of An . By hypothesis,
the images ϕ(e1 ), . . . , ϕ(en ) are all linearly independent in F 0 ⊗A k, so we may complete
this to a basis {ϕ(e1 ), . . . , ϕ(en ), f¯n+1 , . . . , f¯m }. By Proposition 1.2.4, F 0 is generated as a
free module by ϕ(e1 ), . . . , ϕ(en ), fn+1 , . . . , fm so we may define a splitting ψ : F 0 → F by
ψ(ϕ(ei )) = ei for each 1 ≤ i ≤ m and ψ(fi ) = 0 for n + 1 ≤ i ≤ m.
Lemma 11.4.5 (Serre). If (A, m, k) is a local noetherian ring and m = δ(A) is the minimal
number of generators of m, then for all i ≥ 0,
 
A m
dimk Tori (k, k) ≥ .
i
In particular, pdA (k) ≥ m.
Proof. Let F• → k be a minimal free resolution (as A-modules). Then since the image of Fi
lies in mFi−1 , we can easily see that TorA ∼
i (k, k) = Fi ⊗ k for each i. Let m = (x1 , . . . , xm ) and
consider the diagram below whose bottom row is the Koszul complex K• (x1 , . . . , xm ; A):
··· F2 F1 A k 0

ϕ2 ϕ1 id id
m
··· A( 2 ) Am A k 0

217
11.4. Projective Dimension Chapter 11. Ring Homology

By the comparison theorem (Corollary 10.2.12), there is a chain map ϕ• : K• := K• (x1 , . . . , xm ; A) →


F• inducing the identity on both A and k. Notice that each ϕi splits, then rank Fi ≥
rank Ki = mi , proving the lemma. Consider the square
δ
Fi Fi−1

ψi ψi−1
d
Ki Ki−1

Assume inductively that ψi−1 : Fi−1 → Ki−1 has been constructed so that ψi−1 ◦ ϕi−1 = 1.
To show ϕi splits, by Lemma 11.4.4 it suffices to show that each ϕi : K• ⊗A k → F• ⊗A k
is injective. Take z ∈ Ki such that ϕi (z) ∈ mFi and note that δ(Fi ) ⊆ mFi−1 implies
δ ◦ ϕi (z) ∈ m2 Fi−1 . Thus ψi−1 ◦ δ ◦ ϕi (z) ∈ m2 Ki−1 , so we see that

ψi−1 ◦ ϕi−1 ◦ d(z) = d(z) ∈ m2 Ki−1 .


xi
Explicitly, the graded pieces of K• = m
N
i=1 (0 → A −
→ A → 0) are
m
= A( i ) =
Ki ∼
M
AeI
I⊆{1,...,m}
|I|=i

where, if I = {j1 , . . . , ji }, then eI is the generator of A ⊗ Aej1 ⊗ · · · ⊗ Aeji ⊗ A with standard


ej` generators in degree 1. The differential on K• is then given by

d : Ki −→ Ki−1
X
eI 7−→ εj xj eIr{j}
j∈I
P
where εj = ±1. Write z = |I|=i zI eI ∈ Ki . Then
!
X X X X
d(z) = εj xj zI eIr{j} = εj xj zJ∪{j} eJ ∈ m2 Ki−1 .
I⊆{1,...,m} j∈I J⊆{1,...,m} j6∈J
|I|=i |J|=i−1

Since Ki−1 is a free A-module, this means each term j6∈J ±xj zJ∪{j} lies in m2 . However, in
P

m/m2 the xj form a basis so we must have zI ∈ m for all |I| = i. Hence d(z) ∈ mKi−1 , so ϕi
is one-to-one as required.
Remark. The following is an open problem independently posed by Buchsbaum-Eisenbud
and Harack. If (A, m) is a local noetherian ring of dimension d and M is a finitely generated,
artinian A-module with pdA (M ) < ∞, it is not known whether the Betti numbers bi =
d
dimk TorAi (k, M ) satisfy Serre’s lemma: bi ≥ i . However, Walker proved the following
weak form of the problem:
X d
bi ≥ 2d .
i=0

218
11.4. Projective Dimension Chapter 11. Ring Homology

Auslander, Buchsbaum and Serre gave a fundamental description of regular local rings
using projective dimension.
Theorem 11.4.6 (Auslander-Buchsbaum-Serre). For a local noetherian ring (A, m, k) of
dimension d, the following are equivalent:
(1) A is a regular local ring.
(2) m is generated by a regular sequence.
(3) pdA (k) < ∞.
(4) For all finitely generated A-modules M , pdA (M ) < ∞.
Proof. (1) =⇒ (2) Let m = (x1 , . . . , xd ). By Lemma 6.2.9, A is a domain and by induction
so is A/(x1 , . . . , xi ) for each 1 ≤ i ≤ d. Hence (x1 , . . . , xd ) is a regular sequence.
(2) =⇒ (3) Suppose m = (x1 , . . . , xt ) where (x1 , . . . , xt ) is a regular sequence on A.
Then K• (x1 , . . . , xt ; A) is a minimal free resolution of H0 (x1 , . . . , xt ; A) = k, so we have
pdA (k) = t < ∞.
(3) =⇒ (4) If n = pdA (k) then TorA i (k, k) = 0 for all i > n, so there is a minimal free
resolution of k with length n:
0 → Abn → · · · → Ab1 → A → k → 0.
Tensoring with M shows TorA i (k, M ) = 0 for all i > n, so pdA (M ) ≤ pdA (k).
(4) =⇒ (1) Assume pdA (k) = n < ∞. We claim n ≤ d = dim A. Let t be the length of
the longest regular sequence in A, say (y1 , . . . , yt ); we will see that t is equal to the depth of A.
By Proposition 3.6.9, m ∈ Ass(A/(y1 , . . . , yt )). That is, R/m = k sits inside of A/(y1 , . . . , yt )
and in fact we have an exact sequence
0 → k → A/(y1 , . . . , yt ) → M → 0.
If n > t, then tensoring with k gives a long exact sequence
· · · → TorA A A
n+1 (M, k) → Torn (k, k) → Torn (A/(y1 , . . . , yt ), k) → · · ·

but (4) implies TorA n+1 (M, k) = 0. On the other hand, since K• (y1 , . . . , yt ; A) has length t, we
A
get Torn (A/(y1 , . . . , yt ), k) = 0. Therefore TorA n (k, k) = 0, contradicting Proposition 11.4.2.
This Sproves n ≤ t. Next, we claim t ≤ d. If t = 1, A contains a nonzero divisor y1 such that
y1 6∈ p∈Ass(A) p. In particular y1 does not lie in any minimal prime of A by Theorem 3.6.7(3).
Thus d ≥ 1. Now suppose t ≥ 2 and note by Corollary 6.2.3 that dim(A/(y1 )) = d − 1.
Since (ȳ2 , . . . , ȳt ) is a regular sequence in A/(y1 ), so induction on t proves t − 1 ≤ d − 1 and
thus t ≤ d. Finally, let δ = δ(A) be the minimal number of generators of m. We know from
Corollary 6.2.6 that d ≤ δ so to complete the proof it is enough to show δ ≤ n. But this just
follows from Lemma 11.4.5.
Remark. In the course of the proof of Auslander-Buchsbaum-Serre, we established that if
(A, m, k) is a local noetherian ring with pdA (k) < ∞, then
δ(A) = pdA (k) = depth(A) = dim A
which strengthens the conclusion in Section 6.2. Further, this gives a new criterion for a
local noetherian ring to be regular: dimk m/m2 ≤ pdA (k).

219
11.4. Projective Dimension Chapter 11. Ring Homology

Corollary 11.4.7. Every regular local ring is Cohen-Macauley.


Corollary 11.4.8. If A is a regular local ring and p ⊂ A is a prime ideal, then Ap is also
regular.
Proof. By Theorem 11.4.6(2), the A-module M = A/p has a free resolution of finite length,

0 → Abr → · · · → Ab1 → A → M → 0,

where r = pdA (M ). Tensoring with Ap yields

0 → Apbr → · · · → Abp1 → Ap → k(p) → 0

so Corollary 11.4.3 implies pdAp (k(p)) < ∞. Hence (3) of Theorem 11.4.6 says that Ap is
regular.
Corollary 11.4.9. If A is the completion of a regular local ring, then Ap is regular for all
prime ideals p ⊂ A.
The Auslander-Buchsbaum-Serre theorem also allows us to prove the ‘fibrewise criterion
for regularity’.
Theorem 11.4.10. Let ϕ : (A, m) → (B, n) be a flat, local homomorphism of local rings. If
A and B/mB are regular, then so is A.
Proof. Set d = dim A and e = dim B/mB, so that m = (x1 , . . . , xd ) and n/mB = (ȳ1 , . . . , ȳe )
for some yi ∈ B having images ȳi ∈ B/mB. Then n = (ϕ(x1 ), . . . , ϕ(xd ), y1 , . . . , ye ) so we
are done by Theorem 8.3.30.
Corollary 11.4.11. If A is a regular local ring then A[[x1 , . . . , xn ]] is regular for any n ≥ 1.
Proof. The map A → A[[x1 , . . . , xn ]] is flat since it is the direct limit of a system of maps
to free modules. Moreover, the fibre of this map is k[[x1 , . . . , xn ]] which is regular by Theo-
rem 6.2.8(c). Hence Theorem 11.4.10 applies.
Corollary 11.4.12. If A is a regular local ring then A[x1 , . . . , xn ] is also regular for any
n ≥ 1, meaning for all prime ideals q ⊂ A[x1 , . . . , xn ], the local ring A[x1 , . . . , xn ]q is regular.
Proof. By induction, it’s enough to prove the A[x] case. Choose a prime ideal q ⊂ A[x]
and set p = q ∩ A. Then the map Ap → A[x]q factors through Ap → Ap [x] and thus is
flat. Meanwhile, the fibre is (Ap /pAp [x])q = k(p)[x]q which is regular of dimension 1 since
k(p)[x] is a PID, for example, or using Noether normalization (Theorem 2.4.1). Therefore
Theorem 11.4.10 implies A[x]q is regular.
Remark. Let A = k[x1 , . . . , xn ]. Since Ap is regular by the above, for any finitely generated
A-module M and any prime p ⊂ A, pdAp (Mp ) ≤ dim Ap ≤ n. So if

0 → Q → Fn−1 → · · · → F1 → F0 → M → 0

is a resolution with Fi free, Qp will be a free Ap -module. Hence Q is projective, so any


pdA (M ) ≤ n. (In fact, the Quillen-Suslin theorem implies Q is even free as an A-module.)

220
11.4. Projective Dimension Chapter 11. Ring Homology

Corollary 11.4.13. For a regular local ring (A, m) and any nonzero x ∈ m, the quotient
A/(x) is regular if and only if x 6∈ m2 .

Proof. This follows from Theorem 11.4.10 and Corollary 6.2.16.

Theorem 11.4.14 (Jacobian Criterion). Let k be a perfect field and set A = k[x1 , . . . , xn ].
For a prime p = (F1 , . . . , Fm ) ⊂ A of height h = ht(p), let B = A/p, choose a prime q ⊂ B,
set ξi = x̄i ∈ B/q and define the Jacobian
 
∂Fi
J = J(ξ1 , . . . , ξn ) := (ξ1 , . . . , ξn ) .
∂xj

Then rankk(q) J ≤ h with equality if and only if Bq is a regular local ring.

Example 11.4.15. Let A = k[x, y, s, t] and p = (xy − st), which has height 1. Then for the
dimension 3 ring B = A/p,  
y
x
J = 
 −t 
−s
so clearly (1) is satisfied modulo any prime q. Moreover, rank J = 1 if and only if J 6= 0
mod q, so we see that there is only an isolated singularity at q = (x, y, s, t).

Remark. The Jacobian matrix J determines a presentation of the module of Kähler differ-
entials of B from Example 3.6.8:
J
Bn −
→ B m → Ω1A/k → 0.

221
11.5. Depth and Cohen-Macauley Rings Chapter 11. Ring Homology

11.5 Depth and Cohen-Macauley Rings


Recall from Section 11.3 that a ring A is Cohen-Macauley if there is a system of parameters
x1 , . . . , xd ∈ A which is a regular sequence. As projective dimension was used in the last
section to characterize regularity, in this section we will give a homological characterization
of the Cohen-Macauley condition using depth. As one of many consequences, we will see
that if the CM condition holds for some system of parameters for A, then it holds for all
systems of parameters.

Proposition 11.5.1. Let A be a noetherian ring, I ⊂ A an ideal and M a finitely generated


A-module such that IM 6= M . Let (y1 , . . . , yt ) be any regular sequence on M which lies in
I. Then
t = min{i ≥ 0 | ExtiA (A/I, M ) 6= 0}.
In particular, the length of any regular sequence on M in I is constant.

Proof. Suppose HomA (A/I, M ) = Ext0A (A/I, M ) 6= 0. Choose a nonzero map f : A/I → M
and set x = f (1) ∈ M . Then for any y ∈ I, 0 = f (y) = yf (1) = yx so Ix = 0. Thus
every element of I is a nonzero divisor for M and therefore every regular sequence on M
has length
S 0. Conversely, if every regular sequence has length 0, then by Proposition 3.6.9,
I ⊆ p∈Ass(M ) p. By prime avoidance (Lemma 6.2.14), I ⊆ p for some p ∈ Ass(M ). This
means A/I → A/p ,→ M determines a nonzero element of HomA (A/I, M ). So the equality
holds when either side is zero.
Now induct on t; let (y1 , . . . , yt ) be a regular sequence in I for M of maximal length.
Applying HomA (A/I, −) to the short exact sequence
y1
0→M −
→M →
− M1 → 0

yields a long exact sequence


y1
· · · → Exti−1 i
→ ExtiA (A/I, M ) → ExtiA (A/I, M1 ) → · · ·
A (A/I, M1 ) → ExtA (A/I, M ) −

Note that (ȳ2 , . . . , ȳt ) is a maximal length regular sequence on M1 = M/y1 M , so by induction
ExtiA (A/I, M1 ) = 0 for i < t − 1 and ExttA (A/I, M1 ) 6= 0. Therefore the long exact sequence
implies ExtiA (A/I, M ) = 0 for i < t − 1 and the map
t−1 y1
ExtA → Extt−1
(A/I, M ) − A (A/I, M )

is injective. However, y1 ∈ AnnA (A/I) since y1 ∈ I, so the multiplication-by-y1 map is 0 and


hence ExtA t−1
(A/I, M ) = 0. Finally, ExttA (A/I, M ) ∼
= Extt−1
A (A/I, M1 ) 6= 0 by induction.

Definition. For an ideal I ⊂ A and a finitely generated A-module M with IM 6= M , the


I-depth of M is the minimal length of any regular sequence in I on M , written depthI (M ).

Definition. If (A, m) is a local noetherian ring and M is a finitely generated A-module, the
depth of M is defined to be the m-depth, depth(M ) = depthm (M ).

222
11.5. Depth and Cohen-Macauley Rings Chapter 11. Ring Homology

Remark. For any regular sequence (y1 , . . . , yt ) in I on M , iterating the above argument
yields an isomorphism

ExttA (A/I, M ) ∼
= HomA (A/I, M/(y1 , . . . , yt )M ).

Corollary 11.5.2. For any ideal I ⊂ A and any collection of finitely generated A-modules
{Mj }j∈J , with IMj 6= Mj for any j ∈ J, we have
!
M
depthI Mj = min{depthI (Mj ) | j ∈ J}.
j∈J

Theorem 11.5.3 (Auslander-Buchsbaum). Let (A, m, k) be a local noetherian ring and M a


finitely generated A-module such that pdA (M ) < ∞. Then depth(M )+pdA (M ) = depth(A).

Proof. First assume depth(A) = 0. We claim M is free – by Corollary 11.5.2, this would
then imply depth(M ) = depth(A) = 0. Suppose that n = pdA (M ) > 0. Given a minimal
free resolution of M , say
ϕn ϕn−1 ϕ1 ϕ0
0 → Fn −→ Fn−1 −−−→ · · · −→ F0 −→ M → 0,

we know ϕi (Fi ) ⊆ mFi−1 for all i. Now depth(A) = 0 if and only if HomA (k, A) = 0, the
latter of which is only possible if there is some nonzero z ∈ A with mz = 0. Consider the
element  
z
0
u =  ..  ∈ Fn .
 
.
0
Then ϕn (u) = zϕn (e1 ) ⊆ zmFn−1 = (mz)Fn−1 = 0, but since n > 0 this contradicts
injectivity of ϕn . Therefore n must be 0 in the first place, but this of course implies M ∼ = F0
is free and as a consequence depth(M ) = 0 as explained above.
Now assume depth(M ) = 0, so that HomA (k, M ) 6= 0. Let t = depth(A) and choose a reg-
ular sequence (y1 , . . . , yt ) in m for A. Notice that since n = pdA (M ), TorA
i (A/(y1 , . . . , yt ), M ) =
A
0 for all i > n and Torn (A/(y1 , . . . , yt ), M ) 6= 0. For a minimal free resolution F• → M as
above, TorA n (A/(y1 , . . . , yt ), M ) can be computed as the nth homology of

0 → A/(y1 , . . . , yt ) ⊗A Fn → A/(y1 , . . . , yt ) ⊗A Fn−1 → · · ·

But m ∈ Ass(A/(y1 , . . . , yt )) so by Proposition 3.6.9, there is some nonzero z ∈ A/(y1 , . . . , yt )


such that mz = 0. Then the same argument as in the paragraph above shows that

TorA
n (A/(y1 , . . . , yt ), M ) 6= 0.

On the other hand, consider the Koszul complex K• (y1 , . . . , yt ; A):


t
0 → A → At → · · · → A(2) → A → A/(y1 , . . . , yt ) → 0.

223
11.5. Depth and Cohen-Macauley Rings Chapter 11. Ring Homology

Since (y1 , . . . , yt ) is a regular sequence, this is acyclic by Theorem 11.3.10. That is,

TorA
i (A/(y1 , . . . , yt ), M ) = 0 for all i > t.

Since depth(M ) = 0, there is some nonzero z ∈ M with mz = 0, so the same trick shows
TorA
t (A/(y1 , . . . , yt ), M ) 6= 0. Hence t = n.
The remaining cases are when d = depth(M ) > 0 and t = depth(A) > 0. Note that
HomA (k, M ) = 0 implies m 6∈ Ass(M ). By prime avoidance (Lemma 6.2.14), there is
some y ∈ M lying outside the union Ass(M ) ∪ Ass(A) – note that these sets are finite
by Theorem 3.6.7. Then Proposition 3.6.9 implies y is a nonzero divisor for M and for A.
Applying − ⊗A M to the short exact sequence
y
0→A→ − A→0
− A→

yields another short exact sequence


y
0→M →
− M→
− A ⊗A M → 0

and the corresponding long exact sequence in Tor shows that TorA
i (A, M ) = 0 for all i > 0.
On the other hand, let F• → M be a minimal free resolution and apply − ⊗A A:

0 → F n → · · · → F 1 → F 0 → M ⊗A A → 0

where F i = Fi ⊗A A is a free A-module. Note that the Tor vanishing condition on A means
this sequence is exact. We have thus shown that pdA (M ) = pdA (A ⊗A M ). By induction,

depth(M ⊗A A) + pdA (M ⊗A A) = depth(A)

but since A = A/(y) for a nonzero divisor y, depth(A) = depth(A)−1 and depth(M ⊗A A) =
depth(M ) − 1. This proves the formula.

Corollary 11.5.4. A local noetherian ring A is Cohen-Macauley if and only if depth(A) =


dim A.

Corollary 11.5.5. Let A be a regular local ring and I ⊂ A an ideal. Then B = A/I is
Cohen-Macauley if and only if pdA (B) = dim A − dim B.

Proof. By Theorem 11.5.3, depth(B)+pdA (B) = depth(A) = dim A, which can be rewritten
pdA (B) = dim A − depth(B). Therefore B is CM if and only if depth(B) = dim B if and
only if pdA (B) = dim A − dim B
The following is an analogue of Theorem 8.3.30 for depth, and also provides a version of
Theorem 11.4.10 for the Cohen-Macauley condition. In this way it justifies the inclusion of
depth on our list of types of dimension in ring theory.

Theorem 11.5.6. If (A, m) → (B, n) is a flat, local homomorphism of local noetherian


rings, then depth(A) + depth(B/mB) = depth(B). In particular, B is Cohen-Macauley if
and only if A and B/mB are Cohen-Macauley.

224
11.5. Depth and Cohen-Macauley Rings Chapter 11. Ring Homology

Proof. Set k = A/m, t = depth(A) and suppose t > 0. Choose a maximal regular sequence
(x1 , . . . , xt ) in A. Consider the short exact sequence
1x
0 → A −→ A→
− A1 → 0.

Since A → B is flat, the sequence


1x
0 → B −→ B→
− B/x1 B → 0

is also exact, so x1 is a nonzero divisor on B. Moreover, A1 → B1 = B/x1 B is also flat


and if m1 = m/(x1 ) is the maximal ideal of A1 , then B1 /m1 B1 ∼
= B/mB. At the same time,
depth(A1 ) = depth(A) − 1 and depth(B1 ) = depth(B) − 1 so by induction,

depth(A)+depth(B/mB) = (depth(A1 )+1)+depth(B1 /m1 B1 ) = depth(B1 )+1 = depth(B).

It remains to show the base case, when t = depth(A) = 0. Fix some y ∈ B whose image
ȳ ∈ B/mB is a nonzero divisor. We’ll be done by induction, since dim B/mB < dim B, if
we can show A → B/yB is flat and y is a nonzero divisor in B. We begin with the second
statement. We claim that y is a nonzero divisor on B/mn B for all n ≥ 1. Indeed, the base
case is by hypothesis and in general, applying − ⊗A B to the exact sequence

0 → k` ∼
= mn /mn+1 → A/mn+1 → A/mn → 0

(for some ` < ∞) yields the exact sequence in both rows of the following commutative
diagram:
0 (B/mB)` B/mn+1 B B/mn B 0

ȳ y y

0 (B/mB)` B/mn+1 B B/mn B 0

Then by the Snake Lemma and induction, the kernel of the middle column is 0, so y is a
nonzero divisor on B/mn+1 B. Suppose yz = 0 for some z ∈ B. T Then ȳz̄ = 0 in B/mn B
for all n ≥ 1 so by the claim, z̄ = 0 in every B/mn B. Thus z ∈ ∞ n
n=1 m B but by Krull’s
intersection theorem (Corollary 5.4.8), this is 0 so y is a nonzero divisor in B. Finally, to
show flatness, by Theorem ?? it suffices to prove TorA 1 (k, B/yB) = 0. Since y is a nonzero
divisor in B, the sequence
y
0→B→ − B→ − B/yB → 0
is exact, so applying − ⊗A k and taking Tor yields an exact sequence

0 = TorA A
1 (k, B) → Tor1 (k, B/yB) → B/mB →
− B/mB.

(The 0 on the left follows from Theorem ??.) By assumption, multiplication by ȳ is injective
on B/mB, so exactness implies TorA 1 (k, B/yB) = 0.

Corollary 11.5.7. For any local noetherian ring A, depth(A) = depth(A).


b

225
11.5. Depth and Cohen-Macauley Rings Chapter 11. Ring Homology

Proof. Same as the analogous statement for dimension: A → A


b is flat, A/
b mb∼
= A/m = k and
all fields have depth 0.
Corollary 11.5.8. If A is Cohen-Macauley then for any n ≥ 1, A[x1 , . . . , xn ] is also Cohen-
Macauley.
Proof. It suffices to prove the case n = 1, in which case A → A[x] is flat and has regular fibres
which are thus Cohen-Macauley by Corollary 11.4.7. Hence Theorem 11.5.6 applies.
Definition. Let A be a noetherian ring and M a finitely generated A-module. The dimen-
sion of the module M is

dim M = sup{ht(p) | p ∈ Ass(M )}.

By Theorem 3.6.7(2), dim M < ∞ for any finitely generated A-module M .


Theorem 11.5.9. Let (A, m, k) be a local noetherian ring and suppose M and N are finitely
generated A-modules. Then ExtiA (M, N ) = 0 for all i < depth(N ) − dim M .
Proof. If dim M = 0, Ass(M ) = {m} so the quotients in any prime filtration

0 = M` ⊆ M`−1 ⊆ · · · ⊆ M0 = M

of M are all isomorphic to A/m = k. By definition of depth, ExtiA (k, N ) = 0 for all
i < depth(A), so for each short exact sequence

0 → Mj → Mj+1 → Mj+1 /Mj ∼


=k→0

the long exact sequence in Ext implies ExtiA (Mj , N ) = 0. In particular, ExtiA (M, N ) = 0.
Now induct; if dim M > 0 then by the same argument it’s enough to show ExtiA (A/p, N ) = 0
for all i < depth(N ) − dim M and all primes p in a prime filtration of M . Fix such a p.
Then by definition, dim A/p ≤ dim M . Choose x ∈ m which does not lie in p and consider
the short exact sequence
x
0 → A/p →
− A/p →
− A/(p, x) → 0.

By Corollary 6.2.16, dim A/(p, x) = dim A/p − 1 so by induction, ExtiA (A/(p, x), N ) = 0 for
all i < depth(N ) − dim A/p + 1. Applying HomA (−, N ) to the short exact sequence above
yields
x
0 = ExtiA (A/(p, x), N ) → ExtiA (A/p, N ) →
− ExtiA (A/p, N ) → ExtiA (A/(p, x), N ) = 0

whenever i < depth(N ) − dim A/p. Since x ∈ m r p, Nakayama’s lemma (1.2.1) implies
ExtiA (A/p, N ) = 0 for these i. Since p was any prime in a prime filtration of M , we are
done.
Corollary 11.5.10. If (A, m, k) is a local noetherian ring and M is a finitely generated
A-module, then for any p ∈ Ass(M ), depth(M ) ≤ dim A/p.
Corollary 11.5.11. If A is Cohen-Macauley, then

226
11.5. Depth and Cohen-Macauley Rings Chapter 11. Ring Homology

(a) The minimal primes of A are exactly the associated primes of A.

(b) For any minimal prime p ⊂ A, dim A/p = dim A.


Proof. Applying Corollary 11.5.10 with M = A says that for all p ∈ Ass(A), dim A =
depth(A) ≤ dim A/p ≤ dim A, so they must all be associated and the equality in (2) holds.

Example 11.5.12. By Corollary 11.5.11, the ring k[x, y, z]/(x)∩(y, z) is not Cohen-Macauley
since it has two minmal primes, (x) and (y, z), of different codimensions.
Example 11.5.13. Similarly, Corollary 11.5.11 implies that k[x, y]/(x2 , xy) is not Cohen-
Macauley because there is an associated prime, (x, y), which is not minimal – (x) is the only
minimal prime in this ring.
Proposition 11.5.14. Let (A, m) → (B, n) be a local homomorphism of local noetherian
rings such that A is regular and B is a finite A-module. Then B is Cohen-Macauley if and
only if B is a free A-module.
Proof. Since A is regular, Theorem 11.4.6 shows pdA (B) < ∞. Then by the Auslander-
Buchsbaum formula (Theorem 11.5.3), pdA (B) + depth(B) = depth(A). Next, since A → B
is finite, Theorem 2.3.7 implies dim A = dim B. But A is also Cohen-Macauley (Corol-
lary 11.4.7) so depth(A) = dim A by Corollary 11.5.4. Thus

B is Cohen-Macauley ⇐⇒ depth(B) = dim B


⇐⇒ pdA (B) = dim A − dim B = 0 by Theorem 11.5.3
⇐⇒ B is a free A-module.

We can now show that the definition of Cohen-Macauley is independent of the choice of
regular sequence on M .
Proposition 11.5.15. Let (A, m) be a d-dimensional Cohen-Macauley ring containing a field
and suppose x1 , . . . , xd is a system of parameters. Then (x1 , . . . , xd ) is a regular sequence.

Proof. To begin, note that A is CM if and only if A b is CM – this follows from Theorem 11.5.6.
Also note that (x1 , . . . , xd ) is a system of parameters if and only if dim A/(x1 , . . . , xd ) = 0.
Then A/(x b∼
b 1 , . . . , xd )A = A/(x1 , . . . , xd ) so because dim A
b = dim A, x1 , . . . , xd is a system of
parameters for A. Applying Theorem 11.3.10, we have:
b

b ⇐⇒ K(x1 , . . . , xd ; A)
(x1 , . . . , xd ) is a regular sequence on A b is acyclic
⇐⇒ K(x1 , . . . , xd ; A) ⊗A A b is acyclic (A → A
b is flat)
⇐⇒ K(x1 , . . . , xd ; A) is acyclic
⇐⇒ (x1 , . . . , xd ) is a regular sequence on A.

Thus we may reduce to the case when A is a complete CM ring. By Cohen’s third structure
theorem (6.3.11), there is a finite ring extension B = k[[x1 , . . . , xd ]] → A, where k ⊆ A is a

227
11.5. Depth and Cohen-Macauley Rings Chapter 11. Ring Homology

coefficient field. By Corollary 11.4.11, B is regular and because A is CM, Proposition 11.5.14
implies A is a free B-module (beware that A and B have switched roles here compared
to that proposition). On the other hand, K(x1 , . . . , xd ; B) is acyclic by Theorem 11.3.10,
but since B → A is flat, K(x1 , . . . , xd ; A) = K(x1 , . . . , xd ; B) ⊗B A is still acyclic. Hence
Theorem 11.3.10 again implies that (x1 , . . . , xd ) is a regular sequence on A.
The last result holds in general but requires a more flexible version of Cohen’s structure
theorems when A does not contain a field.
Proposition 11.5.16. Let A be a local Cohen-Macauley ring. Then every non-refinable
chain of prime ideals
p0 ( p1 ( · · · ( pk = m
has length k = dim A.
Proof. We induct on d = dim A. If d = 0, there’s only one prime ideal, m, so every chain of
primes is trivial. If d > 0, suppose p0 ( · · · ( pk = m is a chain of primes. Since p1 is not a
minimal prime, the prime avoidance lemma (6.2.14) says that
[
p1 6⊂ q.
q⊂A
q minimal

Pick x ∈ p1 that does not lie in any minimal prime q. By Corollary 11.5.11, the minimal
primes of A are exactly the associated primes, so Proposition 3.6.9 says x is not a zero divisor
on A. Thus A/(x) is also CM and dim A/(x) = d − 1. Consider the chain of primes

p1 ( p2 ( · · · ( pk = m

in A/(x), where pi = pi /(x) for each 1 ≤ i ≤ k. By induction, this chain must have length
k − 1 = dim A/(x) = d − 1 so k = d as desired.
For an ideal I in a noetherian ring A, we extend the definition of height to I by:

ht(I) := min{ht(p) | I ⊆ p is prime}.

The proof of Lemma 2.3.6(b) goes through for any ideal I, so that ht(I) + dim A/I ≤ dim A.
Moreover, Proposition 11.5.16 gives us:
Corollary 11.5.17. If A is a Cohen-Macauley ring, then for any ideal I ⊂ A, ht(I) +
dim A/I = dim A.
Example 11.5.18. Let B = k[x, y1 , . . . , yn ](x,y1 ,...,yn ) , let J = (x) ∩ (y1 , . . . , yn ) and set
A = B/I. Then the minimal primes of A are (x̄) and I = (ȳ1 , . . . , ȳn ) and ht(I) = 0, but
dim A/I = dim k[x̄] = 1 so ht(I) + dim A/I = 1 < dim A = n as long as n ≥ 2. Therefore A
is not Cohen-Macauley.
Theorem 11.5.19 (Unmixedness). Let A be a local Cohen-Macauley ring and suppose I =
(x1 , . . . , xn ) ⊆ m such that ht(I) = n. Then (x1 , . . . , xn ) is a regular sequence and A/I is
Cohen-Macauley.

228
11.5. Depth and Cohen-Macauley Rings Chapter 11. Ring Homology

Proof. First note that in general, Krull’s height theorem (6.2.10) implies ht(I) ≤ n. As-
suming for the moment that (x1 , . . . , xn ) is a regular sequence, Proposition 11.5.15 says
that every maximal regular sequence in A has length depth A = dim A, so it follows that
A/I is also CM. Thus it suffices to show (x1 , . . . , xn ) is a regular sequence. If n = 1, then
ht((x1 )) = 1 is equivalent to saying x1 is not contained in any minimal prime of A, but since
A is CM, Corollary 11.5.11 and Proposition 3.6.9 show x1 is not a zero divisor. For n > 1,
we induct. Suppose x1 lies in some minimal prime p ⊂ A. Then for any prime pk which is
minimal over I, we have the following non-refinable chain of primes in A/(x1 ):
p/(x1 ) ( p1 /(x1 ) ( · · · ( pk /(x1 ).
Then k = ht(pk ) = n, but pk /(x1 ) is minimal with respect to primes of A/(x1 ) containing
(x2 , . . . , xn )/(x1 ), contradicting Krull’s height theorem (6.2.10). Hence x1 cannot be con-
tained in any minimal prime of A, but since A is CM, the argument in the base case shows
x1 is a nonzero divisor on A. That is, A/(x1 ) is CM and ht((x̄2 , . . . , x̄n )) = n − 1. Induction
finishes the proof.
Corollary 11.5.20. If A is a local Cohen-Macauley ring then Ap is Cohen-Macauley for all
primes p ⊂ A.
Proof. By Lemma 2.3.6(a), dim Ap = ht(p). Set n = ht(p). We claim there are x1 , . . . , xn ∈ p
such that ht((x1 , . . . , xn )) = n. Assuming this is possible, the unmixedness theorem then
implies (x1 , . . . , xn ) is a regular sequence in A, and because A → Ap is flat, (x1 , . . . , xn )
will also be a regular sequence in Ap . Hence dim Ap ≥ depth(Ap ) ≥ n = dim Ap so by
Corollary 11.5.4, Ap is CM. Finally, we can find such elements x1 , . . . , xn ∈ p as follows. By
the prime avoidance lemma (6.2.14), there exists an x1 ∈ p that does not lie in any minimal
prime of A. Moreover, there are a finite number of minimal primes containing (x1 ) and by
Krull’s height theorem (6.2.10), each has height 1. So choose x2 ∈ p avoiding these primes
and repeat the argument to construct the rest of the sequence (x1 , . . . , xn ).
Definition. A ring A is catenary if for all primes p ⊂ q in A, every non-refinable chain
of prime ideals between them,
p ( p1 ( · · · ( p` ( q,
has the same length.
Theorem 11.5.21. Suppose A = B/I for some ideal I ⊂ B. Then if B is Cohen-Macauley,
A is catenary.
Proof. For any prime ideal p = q/I ⊂ A, Corollary 11.5.20 shows that Bq is still CM, but
by Proposition 11.5.16, every maximal chain of primes in Bq has the same length. Therefore
the same holds in Ap = Bq /Iq , so A is catenary.
The next result says that the property of being Cohen-Macauley is an ‘open condition’
on the set of primes of a regular ring. This will be made precise once we define Spec(A) and
the Zariski topology in Section 14.1, but for now, define
V (I) := {p ⊂ A | I ⊆ p is prime}
for any ideal I ⊂ A.

229
11.5. Depth and Cohen-Macauley Rings Chapter 11. Ring Homology

Proposition 11.5.22. Suppose B is a regular ring and A = B/q for a prime ideal q ⊂ B.
Then the set
C = {p ⊂ A | p is prime and Ap is not CM}
is equal to V (I) for some ideal I ⊂ A.

Proof. Consider a minimal free resolution


ϕ
· · · → B rc+1 −
→ B rc → · · · → B r1 → B → A → 0

where c = ht(q) = dim B −dim A by Corollary 11.5.17. If ϕ is the zero map then pdB (A) = c
so the Auslander-Buchsbaum formula (Theorem 11.5.3) and Corollary 11.5.4 show A is
CM. Otherwise, it’s enough to show that after localizing at an arbitrary prime p ⊂ B, the
r
corresponding map ϕp : Bp c+1 → Bprc splits. After choosing a basis of these free Bp -modules,
ϕp is split if and only if the ideal of k × k minors, where k is the rank of ϕp , is equal to the
unit ideal. Let I be this minor ideal. Then we have shown that the set of primes p ⊂ A for
which Ap is not CM is precisely V (I).
To finish, we list some nontrivial examples and properties of Cohen-Macauley rings.
Details can be found in papers by the cited authors.

Example 11.5.23. (Hochster-Eagan) Let X = (xij ) be an n × m matrix of indeterminates


and let It (X) be the ideal in the polynomial ring k[xij ] generated by the t × t minors of
X. Then k[xij ]/It (X) is Cohen-Macauley for all t. More generally, Hochster-Roberts show
that if G ⊆ GLn (k) is a linearly reductive group, then the ring of invariants k[V ]G is Cohen-
Macauley, where V = k n .

Proposition 11.5.24 (Hochster-Huneke). Suppose A and B are local noetherian ring such
that A contains a field and B is regular. If there exists a split local homomorphism A ,→ B,
then A is Cohen-Macauley.

Proposition 11.5.25. Let A = k[x1 , . . . , xn ]/I for a homogeneous ideal I ⊂ k[x1 , . . . , xn ].


Then A(x1 ,...,xn ) is Cohen-Macauley if and only if Ap is Cohen-Macauley for all primes p ⊂ A.

230
11.6. Gorenstein Rings Chapter 11. Ring Homology

11.6 Gorenstein Rings


Recall that an ideal I ⊂ A is irreducible if it cannot be written as the intersection of two
proper ideals of A.

Definition. A local noetherian ring (A, m) is Gorenstein if A is Cohen-Macauley of di-


mension d and there exists a system of parameters x1 , . . . , xd ∈ A such that (x1 , . . . , xd ) is
an irreducible ideal.

We will show that this definition is equivalent to the more common definition of A having
finite injective dimension.

Proposition 11.6.1. A regular local ring is Gorenstein.

Proof. By Theorem 6.2.8, A is regular if and only if m = (x1 , . . . , xd ) for a system of param-
eters x1 , . . . , xd , but in general a maximal ideal is irreducible, so A is Gorenstein.
Suppose (A, m) is a local noetherian ring and x1 , . . . , xd is a system of parameters. Then A
is Gorenstein if and only if 0 is an irreducible ideal in the 0-dimensional ring A/(x1 , . . . , xd ).
So to begin our study of Gorenstein rings, let’s characterize when a 0-dimensional local
noetherian ring is Gorenstein. Recall that a local noetherian ring has dimension 0 if and
only if it’s artinian (this is Hopkins’ theorem, 3.5.5).

Example 11.6.2. Let k be a field. Then A = k[x, y]/(x, y)2 is 0-dimensional but not
Gorenstein since (x, y)2 = (x, y 2 ) ∩ (x2 , y) is not irreducible in k[x, y].

Example 11.6.3. If n ≥ 2, the ring A = k[x]/(xn ) is Gorenstein. Indeed, A is a k-vector


space with basis 1, x̄, . . . , x̄n−1 and every ideal in A is of the form I = (x̄k ) for some k. This
means any intersection of proper ideals looks like (x̄j ) ∩ (x̄k ) = (x̄max{j,k} ) 6= 0. Thus 0 is not
an irreducible ideal.

Example 11.6.4. For a nontrivial example, we will show later that

k[x, y, z]/(x2 − y 2 , y 2 − z 2 , xy, xz, yz)

is Gorenstein.

Remark. One might hope to classify all Gorenstein rings of any dimension by the tech-
niques which we will describe for local artinian rings. However, it turns out that there is no
classification of Gorenstein rings of the form k[x1 , . . . , xn ]/I for n ≥ 4.

Recall the following definitions and basic properties from ring theory.

Definition. For an A-module M , a submodule N ⊆ M is called essential, or M is an


essential extension of N , if for all nonzero submodules S ⊆ M , N ∩ S 6= 0.

Definition. The socle of a local ring (A, m) is the set soc(A) = {x ∈ A | xm = 0}.

Lemma 11.6.5. For any local ring (A, m, k),

231
11.6. Gorenstein Rings Chapter 11. Ring Homology

(a) soc(A) is an ideal of A.

(b) m soc(A) = 0, so in particular soc(A) is a k-vector space.

(c) If A is artinian, then A is an essential extension of soc(A).

Proof. (a) and (b) are standard. For (c), take a nonzero ideal I ⊂ A. Then for a nonzero
element x ∈ I, either xm = 0, so x ∈ soc(A) ∩ I, or by Theorem 3.5.6, xmn = 0 for some
minimal n ≥ 2. In the latter case, xmn−1 6= 0 but xmn−1 ⊆ soc(A) ∩ I. Thus soc(A) is an
essential ideal of A.

Theorem 11.6.6. A local artinian ring (A, m, k) is Gorenstein if and only if dimk soc(A) =
1.

Proof. ( =⇒ ) If dimk soc(A) ≥ 2, choose nonzero, linearly independent vectors x, y ∈


soc(A). Then (x) ∩ (y) = 0 so A is not Gorenstein.
( ⇒= ) Assume I ∩ J = 0 for some ideals I, J 6= 0. Then I ∩ J ∩ soc(A) = 0 but by
Lemma 11.6.5(c), we must have I ∩soc(A) 6= 0 and J ∩soc(A) 6= 0. However, these I ∩soc(A)
and J ∩ soc(A) are distinct k-subspaces of soc(A), so we must have dimk soc(A) ≥ 2.
Let (B, m, k) be a regular local ring of dimension d and suppose m = (x1 , . . . , xd ). For a
local artinian B-algebra of the form A = B/I, a minimal free resolution of A,

0 → B bd → · · · → B b1 → B → A → 0,

is necessarily of length d since pdB (A) = dim B − dim A by Theorem 11.5.19 and Corol-
lary 11.5.5. Then tensoring with k produces a resolution with all zero maps, so TorB ∼
d (k, A) =
bd
k . On the other hand, the Koszul complex for A as a B-module is
ϕ
→ Ad → · · · → Ad → A → k ⊗B A → 0
0→A−

so by definition,

TorB
d (k, A) = ker ϕ = ker(a 7→ (±xi )a)
= {a ∈ A | axi = 0 for all 1 ≤ i ≤ d}
= soc(A).

This proves:

Proposition 11.6.7. If A is a local artinian ring that is a quotient of a regular local ring
of dimension d with residue field k, then dimk soc(A) = bd , the dth Betti number of A.

Corollary 11.6.8. A local artinian ring A that is a quotient of a regular local ring of
dimension d is Gorenstein if and only if bd = 1.

Corollary 11.6.9. Suppose f1 , . . . , fd ∈ k[x1 , . . . , xd ] form a system of parameters in the


local ring k[x1 , . . . , xd ](x1 ,...,xd ) . Then the quotient k[x1 , . . . , xd ](x1 ,...,xd ) /(f1 , . . . , fd ) is Goren-
stein.

232
11.6. Gorenstein Rings Chapter 11. Ring Homology

Proof. By Corollaries 11.5.8 and 11.5.20, k[x1 , . . . , xd ](x1 ,...,xd ) is Cohen-Macauley, so by


Proposition 11.5.15, (f1 , . . . , fd ) is a regular sequence. Now the conclusion follows from
analyzing the Koszul complex for (f1 , . . . , fd ).

Theorem 11.6.10. Let (B, n, k) be a regular local ring of dimension n, I ⊂ B an ideal and
suppose A = B/I is Cohen-Macauley of dimension d. Set c = ht(I) = n − d. Then A is
Gorenstein if and only if dimk TorB
c (k, A) = 1.

Proof. By the Auslander-Buchsbaum formula (Theorem 11.5.3) and Corollary 11.5.4,

pdB (A) = depth(B) − depth(A) = dim B − dim A = c

so bc = dimk TorB
c (k, A) is the last Betti number in a minimal free resolution of A:

0 → B bc → B bc−1 → · · · → B b1 → B → A → 0.

Set Fi = B bi for each 1 ≤ i ≤ c. Suppose y1 , . . . , yd ∈ B reduce to a system of parameters


ȳ1 , . . . , ȳd in A. Since A is Cohen-Macauley, Proposition 11.5.15 shows (ȳ1 , . . . , ȳd ) is also
a regular sequence; by prime avoidance (Lemma 6.2.14), we may assume (y1 , . . . , yd ) is a
regular sequence in B. To prove the theorem, we will compute TorB i (B/(y1 , . . . , yd ), A) in
two ways. First, since (y1 , . . . , yd ) is a regular sequence, the Koszul complex K• (y1 , . . . , yd ; B)
is acyclic by Theorem 11.3.10. But the images of the yi form a regular sequence in A, so
because K• (ȳ1 , . . . , ȳd ; A) = K• (y1 , . . . , yd ; B) ⊗B A, this reduced Koszul complex is also
acyclic. That is, TorB i (B/(y1 , . . . , yd ), A) = 0 for i > 0.
On the other hand, TorB ∼
i (B/(y1 , . . . , yd ), A) = Hi (K• (y1 , . . . , yd ; B) ⊗ F• ) which implies
K• (y1 , . . . , yd ; B) ⊗ F• is a free resolution of TorB 0 (B/(y1 , . . . , yd ), A) = A/(ȳ1 , . . . , ȳd ). Now
dim A/(ȳ1 , . . . , ȳd ) = 0, by Krull’s height theorem (6.2.10) e.g., so Proposition 11.6.7 says

dimk soc(A/(ȳ1 , . . . , ȳd )) = dimk TorB


n (k, A/(ȳ1 , . . . , ȳd ))

equals the last Betti number in K• (y1 , . . . , yd ; B) ⊗ F• . The resolutions K• (y1 , . . . , yd ; B) and
F• have lengths d and c, respectively, and the last free module in the tensor complex, in
dimension n = d + c, is exactly Kd (y1 , . . . , yd ; B) ⊗B Fc = B ⊗B Fc ∼ = Fc . Therefore we have
TorB (k, A) ∼
= Tor B
(k, A/(ȳ 1 , . . . , ȳ d )) so we conclude b c = 1 if and only if A/(ȳ1 , . . . , ȳd ) is
c n
Gorenstein, by Theorem 11.6.6. However, the same argument shows that bc = 1 if and only
if dimk TorB c (k, A/(z1 , . . . , zd )) = 1 for any system of parameters z1 , . . . , zd in A, which is
equivalent to (z1 , . . . , zd ) being an irreducible ideal in A for any such zi .

Corollary 11.6.11. Suppose A is Gorenstein and x1 , . . . , xd ∈ A is any system of parame-


ters. Then (x1 , . . . , xd ) is an irreducible ideal.

Example 11.6.12. (Complete intersections) A local noetherian ring A is called a complete


intersection if A ∼
= B/I for a regular local ring B and an ideal I = (f1 , . . . , fc ) generated
by a regular sequence in B. In this case we know the Koszul complex K• (f1 , . . . , fc ; B) is
a minimal free resolution of A, so by Corollaries 11.4.3 and 11.5.5, A is Cohen-Macauley.
Further, ht(I) = c implies TorB ∼
c (k, A) = k so we have inclusions

regular ⊆ complete intersection ⊆ Gorenstein ⊆ Cohen-Macauley.

233
11.6. Gorenstein Rings Chapter 11. Ring Homology

Example 11.6.13. Consider a graph

b
a

with vertices labeled with variables a, . . . , e. Let I = (ab, bc, cd, de, ea) be the edge ideal in
B = k[a, b, c, d, e](a,b,c,d,e) . Then ht(I) = 3 so A = B/I is a local ring of dimension 2 with
free B-resolution
0 → B → B 5 → B 5 → B → A → 0.
Hence by the Auslander-Buchsbaum formula (Theorem 11.5.3), A is Cohen-Macauley. Fur-
ther, the last Betti number of A is 1 so Theorem 11.6.10 shows A is Gorenstein. However,
A is not a complete intersection since I is 5-generated.

Example 11.6.14. Let k = C and choose a polynomial f ∈ C[y1 , . . . , yn ] such that f (0) = 0.
The polynomial ring C[X1 , . . . , Xn ] acts on such f by partial differentiation: Xi · f = ∂y∂ i f .
Set
(f )⊥ := {G ∈ C[X1 , . . . , Xn ] | G · f = 0}.
Then A = C[X1 , . . . , Xn ]/(f )⊥ is a 0-dimensional Gorenstein ring and one can prove that
every 0-dimensional Gorenstein ring is isomorphic to such a quotient. For example, if f =
y12 + y22 + y32 then (f )⊥ = (X1 X2 , X1 X3 , X2 X3 , X12 − X22 , X12 − X32 ). On the other hand, for
g = y1 y2 y3 the corresponding ideal is (g)⊥ = (X12 , X22 , X32 ) and A = C[x1 , x2 , x3 ]/(g)⊥ is a
complete intersection in this case. It is an open question which f give quotients that are
complete intersections.

Theorem 11.6.15. Let (A, m, k) → (B, n, `) be a flat, local homomorphism of local rings.
Then B is Gorenstein if and only if A and B/mB are Gorenstein.

Proof. If dim A or dim B > 0, one can induct using Theorem 11.6.10 so we will reduce to
the case that dim A = dim B = 0. By Theorem 11.6.6, we must show dim soc(B) = 1 if and
only if dim soc(A) = 1 and dim soc(B/mB) = 1. First assume A and B/mB are Gorenstein.
Choose a ∈ A such that soc(A) = (a) ∼= k and b ∈ B such that soc(B/mB) = (b̄) ∼= k. We
claim soc(B) = (ab). On one hand, we have

abn = a(bn) ⊆ a(mB) = (am)B = 0.

If ab = 0, then b ∈ AnnB (a) = AnnA (a)B = mB, so we would have b̄ = 0 in B/mB, a


contradiction. Hence ab is a nonzero element of soc(B). On the other hand, let x ∈ soc(B).
Then xn = 0, so x(mB) = 0 which is equivalent to x ∈ AnnB (mB) = AnnA (m)B = aB.

234
11.6. Gorenstein Rings Chapter 11. Ring Homology

Thus x ∈ aB, say x = ay for some y ∈ B. Then ayn = 0 implies yn ⊆ AnnB (a) =
AnnA (a)B = mB. Hence y ∈ soc(B/mB), so by hypothesis, y = zb + π for some z ∈ B and
π ∈ mB. This implies

x = ay = a(zb + π) = azb + aπ = zab ∈ (ab).

Thus soc(B) ⊆ (ab) so we conclude soc(B) = (ab).


Conversely, assume B is Gorenstein, so soc(B) = (x) for some x ∈ B. By the same
argument as above, if a ∈ soc(A) and b ∈ B such that b̄ ∈ soc(B/mB) r {0}, then
ab ∈ soc(B). The flatness hypothesis guarantees ab 6= 0 for such a, b. If soc(A) has di-
mension greater than 1, then we can find linearly independent elements a1 , a2 ∈ soc(A) with
(a1 b), (a2 b) ⊆ soc(A). So because dim soc(B) = 1, there is some unit c ∈ B × such that
a2 b = ca1 b, or (a2 − ca1 )b = 0. Thus a2 = ca1 , a contradiction. A similar contradiction arises
from assuming dim soc(B/mB) > 1, so we conclude dim(A) = dim(B/mB) = 1.
Let A be a ring and M an A-module. Denote by EA (M ) the injective hull of M , which
is the unique smallest injective A-module containing M (there are several equivalent charac-
terizations, namely EA (M ) is the maximal essential extension of M ). If (A, m, k) is a local
ring, we will write E = EA (k) and for any M , M̌ = HomA (M, E).

Proposition 11.6.16. If (A, m, k) is a local artinian ring with E = EA (k),

(1) The functor M 7→ M̌ is exact.

(2) ǩ ∼
= k.
(3) If λ(M ) denotes the length of a prime filtration of M , then λ(M ) = λ(M̌ ).
ˇ ∼
(4) For any A-module M , we have M̌ = M.
Proof. (1) E is injective.
(2) Note that HomA (k, E) ∼ = soc(E) as k-vector spaces. Moreover, if dim soc(E) ≥ 2,
consider the essential extension

k ,−→ E
1 7−→ x ∈ soc(E).

Then there is a y ∈ soc(E) linearly independent from x so (y) ∩ (x) = 0, contradicting the
fact that E is an essential extension of k. Hence dim soc(E) = 1.
(3) Inducting on λ(M ), note that AssA (M ) = {m}, so there exists an embedding k ,→ M .
Let
0 → k → M → M1 → 0
be the corresponding short exact sequence. Since λ is additive, λ(M ) = λ(k) + λ(M1 ) =
1 + λ(M1 ) and by (1),
0 → ǩ → M̌ → M̌ 1 → 0
is also exact and thus λ(M̌ ) = λ(ǩ) = λ(M̌ 1 ). So by induction, λ(M̌ 1 ) = λ(M1 ). Moreover,
(2) implies λ(ǩ) = λ(k) so it follows that λ(M̌ ) = λ(M ).

235
11.6. Gorenstein Rings Chapter 11. Ring Homology

ˇ = Hom (M̌ , E) = Hom (Hom (M, E), E) determines a map i : M ,→


(4) Identifying M̌ A A A
ˇ
M̌ which is necessarily injective. Then by (3), λ(M ) = λ(M̌ ) so coker i = 0. Hence i is an
isomorphism.
ϕ0 ϕ1 ϕ2
Definition. If M is an A-module and 0 → M −→ E0 −→ E1 −→ · · · is an injective resolution
of M , we say M has injective dimension idA (M ) = n if there is a smallest n such that
coker ϕn is injective. If no such n exists, M has infinite injective dimension.
The next theorem presents some more common definitions of the Gorenstein property,
namely (d).
Theorem 11.6.17. Let (A, m, k) be a local artinian ring and set E = EA (k). Then the
following are equivalent:
(a) A is Gorenstein.
(b) A ∼
= E.
(c) A is injective as an A-module.
(d) idA (A) < ∞.
Proof. (a) =⇒ (b) By Theorem 11.6.6, soc(A) is 1-dimensional. But A is an essential
extension of soc(A), so there are embeddings k ,→ A ,→ E since E is the maximal essential
extension of k. Applying λ to the short exact sequence

0 → A → E → E/A → 0

yields λ(E) = λ(A) = λ(E/A), but Proposition 11.6.16(4) says λ(A) = λ(Ǎ) = λ(HomA (A, E)) =
λ(E) since E is injective. Thus λ(E/A) = 0, which implies E/A = 0.
(b) =⇒ (c) =⇒ (d) are trivial.
(d) =⇒ (a) It is well-known that over a local noetherian ring, every injective A-module
is isomorphic to a direct sum of copies of E. In particular, a finite injective resolution of A
over itself may be taken to be of the form

0 → A → E b0 → E b1 → · · · → E bn → 0.

Applying ˇ(·) and Proposition 11.6.16 gives a free resolution

0 → Abn → · · · → Ab1 → Ab0 → E → 0

so in particular pdA (E) < ∞. Further, the Auslander-Buchsbaum formula (Theorem 11.5.3)
implies pdA (E)+depth(E) = depth(A), but depth(E) and depth(A) are both 0 so pdA (E) =
0 as well. In particular, E ∼= Ab for some b ≥ 1. On the other hand, λ(E) = λ(A) by
additivity of λ, so we must have b = 1. Hence A is Gorenstein.
Theorem 11.6.18. Let f1 , . . . , fd be a system of parameters in the power series ring k[[x1 , . . . , xd ]]
and set A = k[[x1P
, . . . , xd ]]/(f1 , . . . , fd ). Then A is a 0-dimensional complete intersection.
d
Further, if fi = j=1 aij xj for some aij ∈ k, then soc(A) is generated by the image of
det(aij ) in k.

236
Part III

Algebraic Geometry

237
Chapter 12

Algebraic Varieties

One of the principal mathematical offshoots of commutative algebra is the study of algebraic
geometry. Many of the definitions and results in Part I have geometric interpretations which
give further insight into their meaning and importance. However, the beauty of algebraic
geometry lies in how it transports ideas in both directions: the algebraic to the geometric
and the geometric to the algebraic.
In this chapter, we begin the study of algebraic geometry by exploring algebraic varieties.
For awhile, this will illustrate enough of the spirit of algebraic geometry for our purposes,
but eventually we will want to generalize to schemes; this is done in Chapter 14.

238
12.1. Affine Algebraic Varieties Chapter 12. Algebraic Varieties

12.1 Affine Algebraic Varieties


In this section we introduce the fundamental object in algebraic geometry: an affine variety.
Let k be a field. We will often need the assumption that k is algebraically closed in order to
prove nontrivial results.

Definition. For each n ∈ N, we define affine n-space over k to be

An = Ank = {(x1 , . . . , xn ) | xi ∈ k}.

As sets, An = k n , but the new notation carries with it the implication that An is viewed
geometrically.

Remark. Alternatively, for any field k ⊆ K ⊆ k̄, one can define Ank (K) to be the fixed points
of Ank under the action of the Galois group Gal(k̄/K). In particular, Ank = Ank (k) = (k̄ n )Gk
where Gk = Gal(k̄/k) is the absolute Galois group of the field k.

We will let A denote the polynomial ring k[t1 , . . . , tn ].

Definition. For a polynomial f ∈ A, define its zero set (or zero locus) to be

Z(f ) = {P ∈ An | f (P ) = 0}.

We extend the definition of zero set to sets of polynomials f1 , . . . , fr ∈ A by


r
\
Z(f1 , . . . , fr ) = Z(fi ).
i=1

Remark. The definition of zero set can be extended to arbitrary subsets F ⊆ A by


\
Z(F) = Z(f ).
f ∈F

Notice that if I = (F) is the ideal of A generated by F, then Z(F) = Z(I). By Hilbert’s basis
theorem (3.3.2), there exists a finite subset {f1 , . . . , fr } ⊆ F such that Z(F) = Z(f1 , . . . , fr ).

Definition. A subset X ⊆ An is called an algebraic set if X = Z(F) for a set F ⊆ A,


that is, X is algebraic if it is the zero set of some collection of polynomials in k[t1 , . . . , tn ].

By the remark, it is equivalent to say X is a zero set if X = Z(I) for some ideal I ⊂ A.
Thus the operation Z(·) takes a subset of a ring and assigns to it a geometric space. There
is a dual notion:

Definition. For any subset X ⊆ An , we define the vanishing ideal of X to be

J(X) = {f ∈ A | f (P ) = 0 for all P ∈ X}.

Lemma 12.1.1. For all X ⊆ An , J(X) is a radical ideal of A.

239
12.1. Affine Algebraic Varieties Chapter 12. Algebraic Varieties

Proof. Take f, g ∈ J(X) and r ∈ A. Then for any P ∈ X, (f − g)(P ) = f (P ) − g(P ) = 0


and (rf )(P ) = r(P )f (P ) = 0 so f + g, rf ∈ J(X) and thus J(X) is an ideal. Moreover, for
any m ∈ N, f m (P ) = 0 if and only if f (P ) = 0 so we see that r(J(X)) = J(X).
Examples.
1 ∅ = Z(A) and An = Z(0) are both algebraic sets.

2 If U ⊆ An is an affine subspace, i.e. U = P0 + V for a point P0 ∈ An and a linear


subspace V ⊆ k n , then U = Z(L1 , . . . , Ln−d ) where d = dimk V and L1 , . . . , Ln−d are
linear polynomials in A.
3 For any point P = (a1 , . . . , an ) ∈ An , {P } = Z(t1 − a1 , . . . , tn − an ). Consider the max-
imal ideal mP = (t1 − a1 , . . . , tn − an ) ⊂ A (we saw in Section 2.4 that mP is maximal).
Then {P } = Z(mP ). Recall that when k is algebraically closed, Corollary 2.4.3 shows
that points of An are in one-to-one correspondence with the maximal ideals of A via
the association P ↔ mP .
4 In A2 , an example of an algebraic curve is C = {(T 2 − 1, T (T 2 − 1))} = Z(t21 + t31 − t22 ):

5 The algebraic set Z(y, y − x2 ) = Z(x, y) consists of just the point (0, 0) in A2k :

Z(y − x2 )

Z(y)

Z(y, y − x2 )

Definition. If X = Z(S) ⊆ Ank (k̄) is an algebraic set and K is a field such that k ⊆ K ⊆ k̄,
define the K-points of X by X(K) := X ∩ Ank (K) = X GK , where GK = Gal(k̄/K).
Moreover, we say X is defined over K if J(X) has a generating set consisting of elements
of K[t1 , . . . , tn ].

240
12.1. Affine Algebraic Varieties Chapter 12. Algebraic Varieties

Lemma 12.1.2. Let X, Y ⊆ An be sets and I, I1 , I2 and I` ⊂ A be ideals, with ` ∈ L some


indexing set. Then
(a) If Y ⊆ X then J(Y ) ⊇ J(X).

(b) If I2 ⊆ I1 then Z(I2 ) ⊇ Z(I1 ).

(c) Z(J(X)) ⊇ X.

(d) J(Z(I)) ⊇ I.

(e) Z(J(Z(I))) = Z(I).

(f ) J(Z(J(X))) = J(X).

(g) Z(I1 ) ∪ Z(I2 ) = Z(I1 ∩ I2 ) = Z(I1 I2 ).


! !
\ [ X
(h) Z(I` ) = Z I` = Z I` .
`∈L `∈L `∈L

Proof. (a) – (d) are obvious from the definitions of Z and J.


(e) By (c), Z(I) ⊆ Z(J(Z(I))) so it remains to prove the reverse containment. However,
we have that I ⊆ J(Z(I)) by (d) so then applying Z gives Z(I) ⊇ Z(J(Z(I))) by (b).
(f) is similar to (e). Here, (d) gives us J(X) ⊆ J(Z(J(X))). On the other hand, we have
X ⊆ Z(J(X)) by (c), so applying J yields J(X) ⊇ J(Z(J(X))) by (a).
(g) We get Z(I1 ) ∪ Z(I2 ) ⊆ Z(I1 ∩ I2 ) ⊆ Z(I1 I2 ) immediately from the containments
I1 ⊇ I1 ∩ I2 ⊇ I1 I2 and I2 ⊇ I1 ∩ I2 ⊇ I1 I2 , using (b). Suppose P ∈ Z(I1 I2 ) and P 6∈ Z(I1 ).
Then there is some f ∈ I1 such that f (P ) 6= 0, but for any g ∈ I2 , we have (f g)(P ) =
f (P )g(P ) = 0. Since A = k[t1 , . . . , tn ] is a domain and f (P ) 6= 0, we must have g(P ) = 0.
This shows P ∈ Z(I2 ). Hence Z(I1 ) ∪ Z(I2 ) ⊇ Z(I1 I2 ) so we have established all three
equalities. S P S  P 
(h) The containments I` ⊆ `∈L I` ⊆T `∈L I` give us Z(I S ` ) ⊇Z PI` ⊇ Z
`∈L  `∈L I`

T ` ∈ L, by (b), and therefore `∈L Z(I` ) ⊇ Z


for each `∈L I` ⊇ Z `∈L I` .PSuppose
P
P ∈ `∈L Z(I ` ). Then for every f ` ∈ I` , f ` (P ) = 0. In particular,
P  for any fT= `∈L f` ∈
P I` , f` (P ) = 0 for each ` so f (P ) = 0. Thus P ∈ Z
`∈L `∈L I` . This shows `∈L Z(I` ) ⊆
Z I
`∈L ` , so we have all three equalities.
In particular, these properties demonstrate that the algebraic subsets of An form the
closed sets of a topology on An .
Definition. The topology on An having as its closed sets all algebraic subsets of An is called
the Zariski topology on An .
In (c) and (d), we see that Z and J are not quite inverse operations, but that they may
enlarge the given object (set or ideal). The next two important results shows when Z and
J are inverses: namely, ZJ is the identity on closed sets and when k is algebraically closed,
JZ is the identity on radical ideals.
Lemma 12.1.3. If X ⊆ An is any subset and X is the Zariski-closure of X in An , then

241
12.1. Affine Algebraic Varieties Chapter 12. Algebraic Varieties

(a) J(X) = J(X).

(b) Z(J(X)) = X.
Proof. (a) Since X ⊆ X, we immediately get J(X) ⊇ J(X) by Lemma 12.1.2(a). On the
other hand, if f ∈ J(X), f (P ) = 0 for all P ∈ X. In other words, X ⊆ Z(f ) but Z(f ) is
closed by definition, so Z(f ) ⊇ X. Thus f ∈ J(X).
(b) X is algebraic by definition so there exists some ideal I ⊂ A such that Z(I) = X. Now
by (a), Z(J(X)) = Z(J(X)) = Z(J(Z(I))) which by Lemma 12.1.2(e) equals Z(I) = X. So
Z(J(X)) = X as required.
Example 12.1.4. Let k = R and consider f = t12e1 + . . . + tn2en + 1 ∈ R[t1 , . . . , tn ] for some
ei ∈ N. Then Z(f ) = ∅ so J(Z(f )) = A and this holds for all choices of ei ∈ N. However,
f does not generate the whole ring A, so this shows JZ is not always the identity on ideals
of A. Hilbert’s Nullstellensatz, proven below, says that this correspondence is bijective on
radical ideals when k is an algebraically closed field.
Theorem 12.1.5 (Hilbert’s Nullstellensatz, Geometric Version). If k is algebraically closed,
then J(Z(I)) = r(I) for every ideal I ⊂ A.
Proof. By Lemma 12.1.2(d), J(Z(I)) ⊇ I and by Lemma 12.1.1, J(Z(I)) is radical. To-
gether, these imply J(Z(I)) ⊇ r(I). For the reverse containment, take f ∈ J(Z(I)). Since

r
P∈
A is noetherian, I = (f1 , . . . , fm ) for a finite set of polynomials f1 , . . . , fm
m
I. We must
show that there exist g1 , . . . , gm ∈ A and r ∈ N so that we can write f = i=1 fi gi .
Consider the ideal I 0 = (f1 , . . . , fm , 1 − tn+1 f ) ⊂ A[tn+1 ] = k[t1 , . . . , tn=1 ]. If P 0 =
(P, an+1 ) is a point in An+1 , with P ∈ An , assume P 0 ∈ Z(I 0 ). Then fi (P 0 ) = fi (P ) = 0 for
all 1 ≤ i ≤ m. Since I = (f1 , . . . , fm ), this means P ∈ Z(I). Then f ∈ J(Z(I)) implies
f (P ) = 0, but 1 − tn+1 f (P ) = 1 6= 0, contradicting P 0 ∈ Z(I 0 ). Therefore it must be that
Z(I 0 ) = ∅, so I 0 6⊂ mP 0 for any P 0 ∈ An+1 , where mP 0 is the maximal ideal of A[tn+1 ] given
by
mP 0 = (t1 − a1 , . . . , tn+1 − an+1 ).
By Corollary 2.4.3, the algebraic version of Hilbert’s Nullstellensatz for algebraically closed
fields, {mP 0 | P 0 ∈ An+1 } = MaxSpec k[t1 , . . . , tn+1 ]. We have therefore shown that I 0
is not contained in any maximalPideal of k[t1 , . . . , tn+1 ], but this is only possible if I 0 =
k[t1 , . . . , tn+1 ]. In particular, 1 = m
i=1 fi hi +(1−tn+1 f )hm+1 for some h1 , . . . , hm+1 ∈ A[tn+1 ].
1
Substituting tn+1 = f , we obtain
m
X  
1= fi (t1 , . . . , tn )hi t1 , . . . , tn , f1 .
i=1

Then multiplying by f r for sufficiently large r clears all powers of f1 , yielding


m
X
r
f = fi gi ∈ I for some gi ∈ k[t1 , . . . , tn ].
i=1

Hence f ∈ r(I).

242
12.1. Affine Algebraic Varieties Chapter 12. Algebraic Varieties

We have seen that J and Z establish a correspondence, though not always bijective,
between the ideals of A and the closed subsets of An . We next explore how concepts like the
noetherian property and prime ideals translate into geometry.

Definition. A topological space X is noetherian if it satisfies the descending chain condi-


tion on closed subsets, or equivalently, if it satisfies the ascending chain condition on open
subsets.

Example 12.1.6. Affine n-space An equipped with the Zariski topology is a noetherian
space. Indeed, if X1 ⊇ X2 ⊇ X3 ⊇ · · · is a descending chain of Zariski-closed subsets
of An , applying the vanishing ideal operation J to each set gives an ascending chain of
ideals J(X1 ) ⊆ J(X2 ) ⊆ J(X3 ) ⊆ · · · in A = k[t1 , . . . , tn ]. By Hilbert’s basis theorem
(3.3.2), A is noetherian so there is some n ∈ N for which we have J(Xm ) = J(Xn ) for
all m ≥ n. Now by Lemma 12.1.3, Z(J(Xi )) = Xi since each Xi is closed, so we get
Xm = Z(J(Xm )) = Z(J(Xn )) = Xn for all m ≥ n. Hence the chain of subsets stabilizes.

Lemma 12.1.7. Let X be a noetherian space. Then

(a) X is quasi-compact.

(b) Every subspace Y ⊆ X is noetherian, with the subspace topology.


S
Proof. (a) Assume X = i∈I Ui is an open cover of X for some index set I, but there is no
finite subcover. Then for any sequence (ik ) ⊆ I, we can construct an ascending chain

Ui1 ( Ui1 ∪ Ui2 ( Ui1 ∪ Ui2 ∪ Ui3 ( · · ·

which is strictly ascending because no finite union of the Ui can equal all of X. Therefore
X is not noetherian.
(b) Let Y1 ⊇ Y2 ⊇ Y3 ⊇ · · · be a descending chain of closed subsets of Y . TThen there
exists a closed set Xn0 ⊆ X such that Yn = Xn0 ∩ X for each n ∈ N. Set Xn = nj=1 Xj0 for
each n. Then each Xn is closed in X and
n
\ n
\
Xn ∩ Y = Xj0 ∩Y = Yj = Yn .
j=1 j=1

Now X1 ⊇ X2 ⊇ X3 ⊇ · · · is a descending chain of closed subsets in X by construction, so


because X is noetherian, there exists an n ∈ N such that for all m ≥ n, Xm = Xn . Thus for
all m ≥ n, Ym = Xm ∩ Y = Xn ∩ Y = Yn so the chain in Y stabilizes. Hence Y is noetherian
as a subspace of X.

Definition. A nonempty topological space X is said to be irreducible if for any two closed
subsets X1 , X2 ⊆ X such that X1 ∪ X2 = X, we have X = X1 or X = X2 .

Lemma 12.1.8. Let Y be a subspace of a topological space X. Then Y is irreducible if and


only if for any closed sets X1 , X2 ⊆ X such that Y ⊆ X1 ∪ X2 , we have Y ⊆ X1 or Y ⊆ X2 .

Proof. Obvious.

243
12.1. Affine Algebraic Varieties Chapter 12. Algebraic Varieties

Lemma 12.1.9. A set X ⊆ An is irreducible if and only if J(X) is a prime ideal of A.

Proof. ( =⇒ ) Assume f, g ∈ A such that f g ∈ J(X). Then X ⊆ X(f g) = Z(f ) ∪ Z(g) by


Lemma 12.1.2(g), so we can write X = (Z(f ) ∩ X) ∪ (Z(g) ∩ X) – note that each of these
sets is closed in X. If X is irreducible, then we must have X = Z(f ) ∩ X or X = Z(g) ∩ X.
In particular, X ⊆ Z(f ) or X ⊆ Z(g), so f ∈ J(X) or g ∈ J(X). Hence J(X) is prime.
( ⇒= ) Given that J(X) is prime, suppose X ⊆ X1 ∪ X2 for two closed sets X1 , X2 ⊆
An . Then there exist ideals I1 , I2 ⊂ A such that Z(I1 ) = X1 and Z(I2 ) = X2 . By
Lemma 12.1.2(g), X ⊆ Z(I1 ) ∪ Z(I2 ) = Z(I1 I2 ) so applying J, we get J(X) ⊇ J(Z(I1 I2 )) ⊇
I1 I2 by Lemma 12.1.2(a) and (d). Since J(X) is prime, we must have J(X) ⊇ I1 or
J(X) ⊇ I2 , but then X ⊆ Z(J(X)) ⊆ Z(I1 ) = X1 or X ⊆ Z(J(X)) ⊆ Z(I2 ) = X2 .
By Lemma 12.1.8 we are done.

Proposition 12.1.10. Let X be a topological space and Y ⊆ X a subspace. Then

(a) X is irreducible if and only if for any nonempty open subset U ⊆ X, U = X.

(b) Y is irreducible if and only if Y is irreducible.

(c) If X is irreducible then X is connected.

(d) If X is irreducible and f : X → Z is a continuous function, then f (X) is irreducible.

Proof. (a) For a nonempty open set U ⊆ X, X r U is a proper closed subspace and X =
(X r U ) ∪ U is a decomposition into closed subsets. If X is irreducible, then X 6= X r U
implies X = U . Conversely, suppose X = X1 ∪X2 where X1 , X2 ⊆ X are closed and X1 6= X.
Then X r X1 is a nonempty open subset of X, so by assumption X r X1 = X. Then

X = X1 ∪ X2 =⇒ X r X1 ⊆ X2 =⇒ X r X1 ⊆ X2 =⇒ X ⊆ X2 =⇒ X = X2 .

Therefore X is irreducible.
(b) By Lemma 12.1.8, Y is irreducible if and only if whenever Y ⊆ X1 ∪ X2 for closed
X1 , X2 ⊆ X, we have Y ⊆ X1 or Y ⊆ X2 . Notice that this is equivalent to Y ⊆ X1 ∪ X2
implying Y ⊆ X1 or Y ⊆ X2 . So Y is irreducible precisely when Y is irreducible.
(c) Suppose X = U ∪ V for disjoint, nonempty open sets U, V ⊂ X. Then U = X r V
and V = X r U are both closed, but neither U nor V is equal to the whole space X since
they are disjoint and nonempty.
(d) Suppose f (X) = Y1 ∪ Y2 for closed subsets Y1 , Y2 ⊆ f (X). By continuity, X1 :=
f −1 (Y1 ) and X2 := f −1 (Y2 ) are both closed subsets of X. Then X1 ∪ X2 = X so by
irreduciblity of X, X = X1 or X = X2 . Suppose without loss of generality that X = X1 .
Then f (X) = f (X1 ) = Y1 , so it follows that f (X) is irreducible.

Definition. A subset Y of a noetherian space X is called an irreducible component of


X if Y is a maximal irreducible subspace of X.

Example 12.1.11. Consider the affine plane A2 . Take f = t1 t2 in k[t1 , t2 ]. Then V(f ) is
the union of the t1 and t2 axes, each of which is an irreducible subspace of A2 :

244
12.1. Affine Algebraic Varieties Chapter 12. Algebraic Varieties

t2 A2

t1

Example 12.1.12. Take an irreducible polynomial f ∈ k[t1 , t2 ]. Since k[t1 , t2 ] is a UFD, (f )


is a prime ideal so C := Z(f ) is irreducible by Lemma 12.1.9. C is called the (affine) algebraic
curve defined by f , sometimes written f (x, y) = 0. In general, an irreducible polynomial
in k[t1 , . . . , tn ] corresponds to an affine variety Y = Z(f ) ⊆ An , called an (affine) algebraic
hypersurface.
Proposition 12.1.13. If X is a nonempty noetherian space, then it has finitely many irre-
ducible components X1 , . . . , Xm such that X = m
S
i=1 X i .
Proof. First we show X can be written as a finite union of some closed irreducible subsets.
Define collections
( `
)
[
M = ∅ 6= Y ⊆ X : Y = Yi for closed irreducible subsets Yi ⊆ X
i=1
and N = {∅ 6= Y ⊆ X : Y is closed and Y 6∈ M}.

Observe that M is closed under finite unions. We claim that N is empty. If Y ∈ N then Y
is not irreducible (since every irreducible set is an element of M), so there exist Y 0 , Y 00 ⊆ Y
closed such that Y 0 ∪ Y 00 = Y and Y 0 , Y 00 6= Y . Also, at least one of Y 0 , Y 00 is in N in
light of the observation that M is closed under unions. Without loss of generality, assume
Y 0 ∈ N . Continuing this construction inductively, with Y1 = Y and Y2 = Y 0 , we get a
strictly descending chain
Y = Y1 ) Y2 ) Y3 ) · · ·
contradicting the fact that X is noetherian. Therefore N must be empty.
Sm X ∈ M, so there exist closed, irreducible subsets X1 , . . . , Xm ⊆ X such
In particular,
that X = i=1 Xi . We may assume Xi 6⊂ Xj for any i 6= j. Now any irreducible subset
Y ⊆ X is contained in at least one Xi , hence only X1 , . . . , Xm can be irreducible components
of X. In fact, each Xi is an irreducible component since if Y is another irreducible subset
of X such that Xi ⊆ Y , there exists some 1 ≤ j ≤ m for which Xi ⊆ Y ⊆ Xj , but
by assumption this is only possible when i = j. In this case, Xi = Y , proving Xi is an
irreducible component of X for each 1 ≤ i ≤ m.
Definition. Assume X is a noetherian space. If X is irreducible, the dimension of X is
defined by
 
` ∈ N0 | there exists a chain of closed, irreducible subsets
dim X = sup .
Y0 ( Y1 ( · · · ( Y` with Yi ⊆ X

245
12.1. Affine Algebraic Varieties Chapter 12. Algebraic Varieties

By Proposition 12.1.13, we may write X = m


S
i=1 Xi for closed, irreducible subsets Xi ⊆ X.
Then we extend the definition of dimension to X by

dim X = max{dim Xi | 1 ≤ i ≤ m}.

Recall that all algebraic subsets of An are noetherian spaces (with the Zariski topology). We
now define one of the fundamental objects of algebraic geometry.
Definition. An affine algebraic variety over k is an irreducible algebraic subset of An .
Definition. An quasi-affine variety if a nonempty, open subset of an affine variety.
By Proposition 12.1.10, we have that if X is an affine variety, then any nonempty, open
subset U ⊆ X is dense in X. In particular, the Zariski topology on X is not Hausdorff in
general.
Remark. Suppose k is an algebraically closed field. Set A = k[t1 , . . . , tn ]. Then we have
the following correspondence:

{algebraic subsets of An } ←→ {radical ideals of A}


X 7−→ J(X)
Z(I)→−I.7

By Hilbert’s Nullstellensatz, this correspondence is bijective and inclusion-reversing. Fur-


thermore, Lemma 12.1.9 refines this correspondence to

{affine varieties of An } ←→ Spec(A)


and {points P ∈ An } ←→ MaxSpec(A).

This correspondence is preserved when restricting our attention to an algebraic subset of An .


Definition. For an algebraic subset X ⊆ An , define the coordinate ring of X by k[X] :=
A/J(X).
If k is algebraically closed, there are bijective, inclusion-reversing correspondences

{algebraic subsets of X} ←→ {radical ideals of k[X]}


{affine subvarieties of X} ←→ Spec(k[X])
{points of X} ←→ MaxSpec(k[X]).

Lemma 12.1.14. Let X ⊆ An be an algebraic subset. Then


(a) k[X] is a reduced ring.

(b) k[X] is an integral domain if and only if X is irreducible.


Proof. (a) By Lemma 12.1.1, J(X) = r(J(X)). Apply Lemma 1.1.4(b).
(b) By Lemma 12.1.9, X being irreducible is equivalent to J(X) being a prime ideal of
A = k[t1 , . . . , tn ].

246
12.1. Affine Algebraic Varieties Chapter 12. Algebraic Varieties

Definition. For an affine variety X ⊆ An , we define its function field over k to be the field
of fractions Frac k[X], denoted k(X). An element of k(X) is called a rational function on
X. If X is defined over some k ⊆ K ⊆ k̄, then the field K(X) := Frac K[X] is called the
field of K-rational functions on X.

Lemma 12.1.15. For any tower k ⊆ K ⊆ k̄ over which X is defined, K[X] = k̄[X]GK and
K(X) = k̄(X)GK , where GK = Gal(k̄/K) is the absolute Galois group of K.

Proposition 12.1.16. Let X ⊆ An be an affine variety. Then

(a) dim X ≤ dim k[X].

(b) If k is algebraically closed, then dim X = dim k[X].

Proof. (a) If Y0 ( Y1 ( · · · ( Y` is a chain of closed, irreducible subsets of X, then J(X0 ) )


J(Y1 ) ) · · · ) J(Y` ) is a chain of prime ideals in k[X], by Lemma 12.1.9. (The inclusions are
strict since Z(J(Yi )) = Yi for each 0 ≤ i ≤ `, by Lemma 12.1.3(b).) Thus dim X ≤ dim k[X].
(b) Assume k is algebraically closed and let p0 ( p1 ( · · · ( pm be a strictly ascending
chain of prime ideals in k[X]. For each i, pi = p0i /J(X) for some prime ideal p0i ⊂ A =
k[t1 , . . . , tn ] containing J(X). Thus by Hilbert’s Nullstellensatz, J(Z(p0i )) = p0i for each i,
which gives us
Z(p00 ) ) Z(p01 ) ) · · · ) Z(p0m ),
a strictly descending chain of affine subsets of An . Since each p0i contains J(X), this is a
chain of closed, irreducible subsets of X. Hence dim k[X] ≤ dim X so we have equality.

Corollary 12.1.17. Suppose k is algebraically closed and X ⊆ An is an affine variety. Then

(a) dim X = tr degk k(X).

(b) If Y ( X is an affine subvariety, dim Y < dim X.

Proof. (a) Apply Proposition 12.1.16 and Theorem 2.6.4.


(b) This follows immediately from (a) and Lemma 2.6.2.

Theorem 12.1.18. Suppose k is algebraically closed. Then

(a) If I = (f1 , . . . , fs ) is an ideal of A = k[t1 , . . . , tn ], then dim Z(I) ≥ n − s.

(b) If X ⊆ An is any algebraic subset with irreducible components X1 , . . . , Xm ⊆ X, then


dim Xi = n − 1 for all 1 ≤ i ≤ m if and only if there exists an f ∈ A r k with
X = Z(f ).

Proof. (a) By Proposition 3.4.10(b), the set {p ∈ Spec(A) | p ⊇ I} has only finitely many
minimal elements, p1 , . . . , pm . ThenTCorollary 1.1.3 allows us to write r(I) = m
T
i=1 i . By
p
m
Hilbert’s Nullstellensatz, J(Z(I)) = i=1 pi and so by Lemma 12.1.2,
m
! m
\ [
Z(I) = Z(J(Z(I))) = Z pi = Z(pi ).
i=1 i=1

247
12.1. Affine Algebraic Varieties Chapter 12. Algebraic Varieties

Lemma 12.1.9 then says that since J(Z(pi )) = pi is prime, Xi := Z(pi ) is an irreducible
set. We have Xi 6⊂ Xj for any i 6= j since J(Xi ) = pi 6⊃ pj = J(Xj ). Hence by the proof
of Proposition 12.1.13, X1 , . . . , Xm are the irreducible components of J(I) = X. Now by
Krull’s height theorem (6.2.10), ht(pi ) ≤ s for each 1 ≤ i ≤ m. We now exploit the fact that
the polynomial ring has the strong dimension property. One has for each i,

dim Xi = dim k[Xi ] = dim A/pi = dim A − ht(pi ) ≥ n − s

by Corollary 2.6.8. Taking the maximum over all Xi , we get dim X ≥ n − s.


(b) Starting with X = Z(f ), we know A is a UFD so there exist a unit u ∈ A× , prime
elements
Sm p1 , . . . , pm and exponents e1 , . . . , em ∈ N such that f = upe11 · · · pemm . Then Z(f ) =
i=1 Z(pi ) so it follows that Z(pi ) are the irreducible components of Z(f ). Then dim Xi =
dim Z/(pi ) = dim A/(pi ) = n − 1 by Proposition 2.6.6. Conversely, suppose dim Xi = n − 1
for every irreducible component Xi of X. Set pi = J(Xi ) to be the corresponding prime
ideal in A. Then dim A/pi = dim k[Xi ] = dim Xi = n − 1 which implies ht(pi ) = 1 by
Lemma 2.6.5(a). Since A is a UFD, there is some prime element pi ∈ A with pi = (pi ). Then
Xi = Z(pi ) = Z(pi ) so it follows that f = p1 · · · pm is the element we seek.

Example 12.1.19. Let X ⊆ An be an affine variety. Then for any r ≥ 1, X × Ar is an


affine variety with coordinate ring k[X × Ar ] ∼
= k[X][t1 , . . . , tr ]. In particular, dim X × Ar =
dim X + r.

Lemma 12.1.20. If k is algebraically closed and U ⊆ X is a nonempty open subset of


X ⊆ An an algebraic variety, then U is also a noetherian space with the subspace topology,
U is irreducible and dim U = dim X.

248
12.2. Morphisms of Affine Varieties Chapter 12. Algebraic Varieties

12.2 Morphisms of Affine Varieties


In this section we define the notion of a morphism, or map, between algebraic varieties. In
a basic sense, these can be thought of as polynomial maps; between affine spaces An and
Am , this is well-defined but on subsets of affine space, the definition will be slightly different.
Consider a map ϕ = (ϕ1 , . . . , ϕm ) : An → Am . We say ϕ is a morphism of varieties if
there exist polynomials f1 , . . . , fm ∈ A = k[t1 , . . . , tn ] such that fi (P ) = ϕi (P ) for all points
P ∈ An . If k is infinite, notice that for all f, g ∈ A, f = g if and only if f (P ) = g(P ) for all
P ∈ An .
Now suppose X ⊆ An is an affine variety. A morphism from X into affine m-space is a
map ϕ = (ϕ1 , . . . , ϕm ) : X → Am such that there exist f1 , . . . , fm ∈ A with ϕi (P ) = fi (P ) =
f¯i (P ) for all P ∈ X, where f¯i = fi + J(X) ∈ k[X]. Formally, we identify ϕi = f¯i : X → k.
We generalize this construction to maps between two affine varieties below.
Definition. If X ⊆ An and Y ⊆ Am are affine varieties, a morphism of affine varieties
between them is a map ϕ = (ϕ1 , . . . , ϕm ) : X → Y such that ϕ1 , . . . , ϕm ∈ k[X] and for all
P ∈ X, ϕ(P ) ∈ Y .
Let ϕ : X → Y be a morphism of affine varieties. Since Y is closed, Y = Z(J(Y ))
by Lemma 12.1.3(b) so ϕ(P ) ∈ Y is equivalent to saying g(ϕ1 (P ), . . . , ϕm (P )) = 0 for all
g ∈ J(Y ), or alternatively, g ◦ ϕ = 0 for all g ∈ J(Y ).
Definition. For affine varieties X ⊆ An and Y ⊆ Am , we define the ring of morphisms
from X to Y as the set

M(X, Y ) = {ϕ : X → Y | ϕ is a morphism}

with ring operations given by pointwise addition and composition. A morphism ϕ ∈ M(X, Y )
is an isomorphism if ϕ is bijective and ϕ−1 is a morphism in the ring M(Y, X).
Example 12.2.1. Let k be a perfect field of characteristic p > 0 and consider the p-power
map

ϕ : A1 −→ A1
x 7−→ xp .

Then ϕ is a bijective morphism (it is explicitly given by a polynomial map and is known to
be an automorphism of any field of characteristic p) but it is not an isomorphism of varieties,
since taking the pth root is not a polynomial map.
Example 12.2.2. For any affine variety X over k, k[X] = M(X, A1 ).
Proposition 12.2.3. For an affine variety X over a field k, the points of X are in bijective
correspondence with Homk (k[X], k).
Proof. The correspondence is given by sending a point P ∈ X to the evaluation homomor-
phism evP : f 7→ f (P ), and sending a k-algebra homomorphism f : k[X] → k to the point
(f (t̄1 ), . . . , f (t̄n )) ∈ k[X], where k[X] = k[t̄1 , . . . , t̄n ].

249
12.2. Morphisms of Affine Varieties Chapter 12. Algebraic Varieties

Lemma 12.2.4. Let X and Y be affine varieties. Then every morphism ϕ ∈ M(X, Y ) is
continuous with respect to the Zariski topologies on X and Y .

Proof. Let ϕ : X → Y be a morphism. Take a closed set Z = Z(f1 , . . . , fr ) ⊆ Y , with


f1 , . . . , fr ∈ k[t1 , . . . , tm ]. Then

ϕ−1 (Z) = {P ∈ X | ϕ(P ) ∈ Z} = {P ∈ X | fi ◦ ϕ(P ) = 0}


= Z(f1 ◦ ϕ, . . . , fr ◦ ϕ).

Since each fi ◦ ϕ ∈ k[t1 , . . . , tn ], Z(f1 ◦ ϕ, . . . , fr ◦ ϕ) is a closed set in X. Hence ϕ is


continuous.

Lemma 12.2.5. For affine varieties X, Y and W and morphisms ϕ : X → Y and ψ : Y →


W , the composition ψ ◦ ϕ : X → W is also a morphism.

Definition. For a morphism of varieties ϕ : X → Y , we define a comorphism ϕ∗ ∈


Homk (k[Y ], k[X]) of the k-algebras k[Y ] and k[X] by

ϕ∗ : k[Y ] −→ k[X]
g 7−→ (g ◦ ϕ : X → A1 ).

Conversely, given a k-algebra homomorphism α : k[Y ] → k[X], we define its corresponding


comorphism of varieties by

α∗ : X −→ Y
(t̄1 , . . . , t̄n ) 7−→ (α(t̄1 ), . . . , α(t̄n )).

Lemma 12.2.6. Let X, Y and W be affine varieties. Then

(a) (idX )∗ = idk[X] and (idk[X] )∗ = idX .


ϕ ψ
(b) For morphisms of varieties X − → W , we have (ψ ◦ ϕ)∗ = ϕ∗ ◦ ψ ∗ . That is,
→ Y −
comorphisms are natural.
α β
(c) For k-algebra homomorphisms k[W ] − − k[X], we have (β ◦ α)∗ = α∗ ◦ β ∗ .
→ k[Y ] →

(d) For any ϕ ∈ M(X, Y ), ϕ∗∗ = ϕ. Likewise, for any α ∈ Homk (k[Y ], k[X]), α∗∗ = α.

Definition. A morphism ϕ : X → Y is said to be dominant if ϕ(X) = Y , that is, if it


has dense image in its target. A dominant morphism is called finite if the ring extension
k[X]/ϕ∗ (k[Y ]) is an integral extension.

Remark. Let ϕ : X → Y be a finite morphism. Then the injection ϕ∗ : k[Y ] ,→ k[X] induces
a field extension k(Y ) ,→ K(X) which is finite and algebraic. Further, if k is algebraically
closed, Proposition 12.1.16 and Theorem 2.3.7 imply that dim X = dim Y .

Example 12.2.7. Let ϕ : A1 → A1 be the map sending x 7→ xn for some n ≥ 2. Then for
all infinite fields k, ϕ is a finite morphism but usually not surjective.

250
12.2. Morphisms of Affine Varieties Chapter 12. Algebraic Varieties

Proposition 12.2.8. If k is algebraically closed and ϕ ∈ M(X, Y ) is a finite morphism,


then ϕ is surjective.

Proof. Take a point x = (x1 , . . . , xn ) ∈ X ⊆ An and associate to it the maximal ideal


mx = (t1 − x1 , . . . , tn − xn ) ⊂ A. Recall that J({x}) = mx . The ideal mx = mx /J(X) is
maximal in k[X]. Similarly, for any point y ∈ Y we have my = J({y}) ∈ k[t1 , . . . , tm ] and
an associated maximal ideal my ∈ k[Y ]. Now

ϕ(x) = y ⇐⇒ ϕ(x) ∈ {y} = Z(my )


⇐⇒ ḡ(ϕ(x)) = 0 for all ḡ ∈ my
⇐⇒ (ḡ ◦ ϕ)(x) = 0 for all ḡ ∈ my
⇐⇒ ϕ∗ (ḡ)(x) = 0 for all ḡ ∈ my
⇐⇒ ϕ∗ (ḡ) ∈ mx for all ḡ ∈ my
⇐⇒ ϕ∗ (my ) ⊆ mx .

Assuming ϕ is finite, the extension k[X]/ϕ∗ (k[Y ]) is integral and in particular ϕ∗ : k[Y ] →
ϕ∗ (k[Y ]) is an isomorphism of k-algebras. For any given y ∈ Y , Lemma 2.3.4 shows there
exists a prime ideal m ⊂ k[X] with m ∩ ϕ∗ (k[Y ]) = ϕ∗ (my ). Since ϕ∗ (my ) is maximal, this m
must be maximal in k[X] by Lemma 2.3.1(b). Then by Hilbert’s Nullstellensatz (algebraic
version, Theorem 2.4.2), there exists x ∈ X such that m = mx . Thus ϕ∗ (my ) ⊆ m = mx . By
the above equivalences, this implies ϕ(x) = y so we have proven ϕ is a surjection.

Remark. If ϕ : X → Y is a finite morphism, the fibre ϕ−1 (y) is a finite set for all y ∈ Y .
In particular, dim ϕ−1 (y) = 0 as a variety.

Lemma 12.2.9. Let ϕ : X → Y be a morphism of affine varieties. Then

(a) ϕ∗ is injective if and only if ϕ is dominant.

(b) ϕ∗ is surjective if and only if ϕ(X) is closed in Y and ϕ0 : X → ϕ(X) is an isomor-


phism.

(c) ϕ∗ is an isomorphism of k-algebras if and only if ϕ is an isomorphism of affine vari-


eties.

Proof. (a) Notice that

ϕ∗ is injective ⇐⇒ ϕ∗ (g) 6= 0 for all g ∈ k[Y ] r {0}


⇐⇒ g ◦ ϕ 6= 0 for all g ∈ k[Y ] r {0}
⇐⇒ g(ϕ(X)) 6= {0} for all g ∈ k[Y ] r {0}
⇐⇒ J(ϕ(X)) = J(Y ) ⊆ k[t1 , . . . , tn ]
⇐⇒ ϕ(X) = Y by Lemma 12.1.3.

Therefore the injectivity of ϕ∗ is equivalent to ϕ being dominant.


(b) Suppose ϕ∗ is surjective. We claim ϕ(X) = Z(ker ϕ∗ ). On one hand, if y ∈ ϕ(X), let
x ∈ X such that ϕ(x) = y. Then for any g ∈ ker ϕ∗ , g(y) = g(ϕ(x)) = (g◦ϕ)(x) = ϕ∗ (g)(x) =

251
12.2. Morphisms of Affine Varieties Chapter 12. Algebraic Varieties

0, so ϕ(X) ⊆ Z(ker ϕ∗ ). This gives us a map ϕ0 : X → Z(ker ϕ∗ ) which corresponds on the


level of k-algebras to (ϕ0 )∗ : k[Y ]/ ker ϕ∗ → k[X]. Since (ϕ0 )∗ is surjective, the isomorphism
theorems give us k[Y ]/ ker ϕ∗ ∼ = k[X] which implies, so ϕ0 : X → ϕ(X) is an isomorphism.
In particular, ϕ(X) = Z(ker ϕ∗ ) so the image of ϕ is closed. The converse follows from
Lemma 12.2.6.
(c) follows from (a) and (b).
Corollary 12.2.10. For any affine varieties X and Y , there is an isomorphism

HomAffk (X, Y ) ∼
= Homk-alg (k̄[Y ], k̄[X]).

In particular, there is an equivalence of categories between Affk , the affine varieties over k
together with variety morphisms, and (k-alg)op , the opposite category of finitely generated
k-algebras together with k-algebra homomorphisms.
Remark. Suppose k ⊆ K ⊆ k̄ and X and Y are defined over K. If ϕ = (ϕ1 , . . . , ϕm ) : X →
Y is a morphism such that each ϕi ∈ K[t1 , . . . , tn ], we say the morphism is defined over K. In
particular, any ϕ : X → Y induces a morphism of K-rational points, ϕK : X(K) → Y (K),
that is defined over K.
Example 12.2.11. (Interpreting Noether normalization) Assume k is algebraically closed.
Take an affine variety X ⊆ An with associated affine algebra k[X] = k[t1 , . . . , tn ]/J(X) =
k[t̄1 , . . . , t̄n ]. By Corollary 12.1.17 and Theorem 2.6.4, d := dim X = tr degk k[X] =
Pn k[X]. Then Noether normalization (Theorem 2.4.1) prescribes elements∼s1 , . . . , sd ∈
dim
i=1 k t̄i such that k[X]/k[s1 , . . . , sd ] is an integral extension and k[s1 , . . . , sd ] = k[t1 , . . . , td ]
as k-algebras. Consider the embedding α : k[s1 , . . . , sd ] ,→ k[X]. Viewing k[s1 , . . . , sd ] =
k[t1 , . . . , td ] = k[Ad ], we get a comorphism ϕ = α∗ : X → Ad . Since k[X]/k[t1 , . . . , td ] is
integral, ϕ is a finite morphism and hence surjective by Proposition 12.2.8. This says that
any affine variety can be viewed as a finite-sheeted cover of Ad for some d.
Suppose U ⊆ X is a nonempty, open subset of an affine variety X ⊆ An . Then Y = X rU
is a closed set in X, so Y = Z(I) for some ideal I ⊂ k[X]. Conversely, for every ideal I ⊂ k[X]
we associate to it an open set D(I) = X r Z(I).
Definition. A principal open set in X is D(f ) := D((f )) for some f ∈ k[X] r {0},
sometimes also denoted Xf . Explicitly, D(f ) = {x ∈ X | f (x) 6= 0}.
If I = (f1 ,S. . . , fm ) for f1 , . . . , fm ∈ k[X], then by Lemma 12.1.2(h), Z(I) = m
T
i=1 Z(fi ) so
m
that D(I) = i=1 D(fi ). This shows that every open subset in X can be covered by finitely
many principal open subsets.
For f ∈ k[X] r {0}, choose a lift F ∈ k[t1 , . . . , tn ] such that f = F + J(X) ∈ k[X]. Set
A = k[t1 , . . . , tn ] and B = k[t1 , . . . , tn+1 ].
Lemma 12.2.12. Define W ⊆ An+1 by W = Z(L)where L  is the ideal (J(X), tn+1 F − 1) ⊂
B. Let ϕ : D(f ) → W be the map given by x 7→ x, f (x) with inverse ϕ−1 : W → D(f ).
1

Then
(a) ϕ is a Zariski-homeomorphism.

252
12.2. Morphisms of Affine Varieties Chapter 12. Algebraic Varieties

(b) ϕ induces a k-algebra isomorphism k[W ] ∼


= k[X]f , the localization of k[X] at the
powers of f .

As a consequence, the principal open subsets of X can be identified with affine subvarieties
of An+1 , giving rise to the quasi-affine varieties mentioned previously.

Remark. If k is an arbitrary field, k[X]f does not only depend on the principal open set
Xf , but also on the choice of f . However, if k is algebraically closed, we have

Xf = Xg ⇐⇒ Z(f ) = Z(g)
⇐⇒ r((f )) = r((g)) by Hilbert’s Nullstellensatz
=⇒ k[X]f = k[X]g by Proposition 1.3.5.

In this context, we denote k[X]f = k[Xf ].

253
12.3. Sheaves of Functions Chapter 12. Algebraic Varieties

12.3 Sheaves of Functions


Definition. Let X ⊆ An be an irreducible affine variety. A sheaf of functions on X
consists of the following data:

(a) For x ∈ X, the stalk at x is defined by


n o
OX,x = fg ∈ k(X) : f, g ∈ k[X], g(x) 6= 0 .

Alternatively, if mx = J({x}) is the maximal ideal associated to the point x, we may


identify OX,x = k[X]mx .

(b) For each nonempty, open subset U ⊆ X, we define a ring of functions on U by


\ \n o
OX (U ) = OX,x = h = fgxx ∈ k(X) : fx , gx ∈ k[X] such that gx (x) 6= 0 .
x∈U x∈U

Each h ∈ OX (U ) gives a well-defined, k-valued function

h : U −→ k
fx (x)
x 7−→
gx (x)

called a regular function on U . Formally, we set OX (∅) = 0.

Lemma 12.3.1. Let X ⊆ An be a variety and U ⊆ X a nonempty, open subset. Then

(a) For each h ∈ OX (U ), h : U → k is continuous with respect to the Zariski topology.

(b) Suppose h1 , h2 ∈ OX (U ) and V ⊆ U is a nonempty, open subset such that h1 |V = h2 |V .


Then h1 = h2 on U .

Proof. (a) Clear from the definitions.


f1
(b) Fix x ∈ V and f1 , g1 , f2 , g2 ∈ k[X] with g1 (x), g2 (x) 6= 0 and h1 = g1
and h2 = fg22 .
Without loss of generality we may assume V ⊆ Xg1 ∩ Xg2 . By assumption, fg11 (y)
(y)
= fg22 (y)
(y)
for
all y ∈ V . Thus f1 g2 − f2 g1 = 0 on V , so V ⊆ Z(f1 g2 − f2 g1 ), but Z(f1 g2 − f2 g1 ) is closed
and V is dense, so it follows that Z(f1 g2 − f2 g1 ) = X. Thus f1 g2 − f2 g1 = 0 in k[X], meaning
f1 f2
h1 = = = h2
g1 g2
in k(X).

Proposition 12.3.2. Suppose k is algebraically closed and X is an affine variety over k.


Then

(a) OX (X) = k[X], that is, the coordinate ring of X consists of regular k-valued functions
X → k.

254
12.3. Sheaves of Functions Chapter 12. Algebraic Varieties

(b) For any f ∈ k[X] r {0}, OX (Xf ) = k[X]f .

Proof. (a) By the Nullstellensatz, the maximal ideals of k[X] are precisely mx for x ∈ X.
Since k[X] is a domain, Lemma 2.3.11 allows us to write
\ \ \
k[X] = k[X]m = k[X]mx = OX,x = OX (X).
maximal ideals x∈X x∈X
m⊂k[X]

(b) This follows from a similar proof to (a), using the fact that the maximal ideals of
k[X]f are precisely those of the form S −1 mx , where x ∈ Xf and S = {f m | m ∈ N0 }.

Definition. Assume U is a (nonempty) quasi-affine variety. The function field of U is


defined to be k(U ) = Frac(OX (U )), where X = U is the Zariski-closure of U . Viewing
OX (U ) as a subring of k[X], we can identify k(U ) = k(X) whenever X = U .

Definition. Two quasi-affine varieties U, V ⊆ X are said to be birationally equivalent if


k(U ) ∼
= k(V ) as k-algebras.
Proposition 12.3.3. Suppose X and Y are affine varieties over k. Then the following are
equivalent:

(i) X is birationally equivalent to Y .

(ii) There exist nonzero functions f ∈ k[X] and g ∈ k[Y ] such that Xf ∼
= Yg as affine
varieties.

(iii) There exist nonzero functions f ∈ k[X] and g ∈ k[Y ] such that k[X]f ∼
= k[Y ]g as
k-algebras.

Proof. (i) ⇐⇒ (ii) By Proposition 12.1.10(a), the principal open sets Xf and Yg are dense
in X and Y , respectively, so k(X) = k(Xf ) ∼
= k(Yg ) = k(Y ). Thus X and Y are birationally
equivalent. The converse comes from the fact that the principal open sets form bases of the
Zariski topologies on X and Y .
(ii) ⇐⇒ (iii) follows from the identification k[Xf ] = k[X]f .
A major area of interest in algebraic geometry is the classification of varieties up to
birational equivalence. For curves, there is a canonical invariant called the genus which
completely classifices curves up to birational equivalence over the algebraic closure k̄ of a
field k.

Definition. A rational variety is a variety X over k which is birationally equivalent to


An for some n.

255
12.4. Finite Morphisms and the Fibre Theorem Chapter 12. Algebraic Varieties

12.4 Finite Morphisms and the Fibre Theorem


Assume k is an algebraically closed field and let X and Y be affine varieties over k. Recall
from Section 12.2 that a morphism ϕ : X → Y is finite if it is dominant – i.e. ϕ∗ : k[Y ] ,→
k[X] is injective – and k[X]/k[Y ] is integral. By Proposition 12.2.8, ϕ must also be surjective.

Proposition 12.4.1. Let ϕ : X → Y be a finite morphism. Then

(a) For any closed set E ⊆ X, ϕ(E) ⊆ Y is closed and the restriction ϕ|E : E → ϕ(E) is
also a finite morphism.

(b) Suppose W ⊆ Y is a closed, irreducible subset and ϕ−1 (W ) ⊆ X has irreducible


components E1 , . . . , E` , with Ei ⊆ X closed for each 1 ≤ i ≤ `. Then ϕ(Ei ) = W if
and only if dim Ei = dim ϕ−1 (W ). Furthermore, dim ϕ−1 (W ) = dim W .

Proof. (a) Set A = k[Y ] and B = k[X], so that B/A is an integral extension. Given a closed
set E ⊆ X, set I = J(E) to be its vanishing ideal in B. By Lemma 2.3.1(a), (B/I)/(A/A∩I)
is also an integral extension. From Lemma 12.1.3, we have Z(I) = Z(J(E)) = E = E.
Let I 0 = A ∩ I and set E 0 = Z(I 0 ). Then ϕ(E) = E 0 so in particular ϕ(E) is closed.
Define ψ : E → E 0 by x 7→ ϕ(x). We claim ψ is finite. First, since I = J(E) is a
radical ideal by Lemma 12.1.1, I 0 = A ∩ I must be radical, too, e.g. by Lemma 1.1.9.
Thus by the Nullstellensatz, I 0 = r(I 0 ) = J(Z(I)) = J(E 0 ). Viewing J(E) ⊆ B, we have
B/I = k[X]/J(E), but by isomorphism theorems, this can be written

k[X]/J(E) = (k[t1 , . . . , tn ]/J(X))/(J(E)/J(X)) ∼


= k[t1 , . . . , tn ]/J(E) = k[E].

Similarly, A/I 0 ∼
= k[E 0 ], so we get a comorphism ψ ∗ : k[E 0 ] → k[E] which is injective by the
above comments. Therefore Lemma 12.2.9 shows that ψ is dominant. Moreover, k[E]/k[E 0 ]
is integral by the above, so ψ is finite.
(b) Since each component Ei is closed in X, (a) says that each ϕ(Ei ) is closed in Y .
Further, irreducibility of Ei implies ϕ(Ei ) is irreducible by Proposition 12.1.10(d). By (a),
ϕ|Ei : Ei → ϕ(Ei ) is a finite morphism, so Theorem 2.6.4 implies dim Ei = dim ϕ(Ei ). In
particular, Corollary 12.1.17 implies that since ϕ(Ei ) is a subvariety of W , ϕ(Ei ) = W if
and only if dim W = dim ϕ(Ei ) = dim Ei .
To finish, we show dim ϕ−1 (W ) = dim W . Corollary 12.1.17 and the previous paragraph
give us dim Ei = dim ϕ−1 (Ei ) ≤ dim W , so it follows from the definition of dimension for
ϕ−1 (W ) that dim ϕ−1 (W ) = max{dim Ei | 1 ≤ i ≤ `} ≤ dim W . Conversely, since ϕ is
surjective we get that

W = ϕ(ϕ−1 (W )) = ϕ(E1 ) ∪ · · · ∪ ϕ(E` ).

Since W is irreducible, we must have W = ϕ(Ej ) for some 1 ≤ j ≤ `. Then

dim ϕ−1 (W ) ≥ dim Ej = dim ϕ(Ej ) = dim W.

Hence the dimensions of W and ϕ−1 (W ) are equal.

256
12.4. Finite Morphisms and the Fibre Theorem Chapter 12. Algebraic Varieties

We now prove the powerful Fibre Theorem for dominant morphisms. Recall from Exam-
ple 12.1.19 that if X is an affine variety, so is X × Ar , and dim X × Ar = dim X + r.
Theorem 12.4.2 (Fibre Theorem). Suppose ϕ : X → Y is a dominant morphism of affine
varieties. Set r = dim X − dim Y . Then there exists a nonempty, open subset U ⊆ Y such
that
(a) U ⊆ ϕ(X).
(b) If W ⊆ Y is closed and irreducible and W ∩ U 6= ∅, then dim(ϕ−1 (W ) ∩ ϕ−1 (U )) =
dim W + r. In particular, dim ϕ−1 (y) = r for all y ∈ U .
Proof. Suppose X ⊆ Ap and Y ⊆ Aq for p, q ≥ 1. Let B = k[X] and A = k[Y ] so that
A ,→ B. Set K = k(Y ) and L = k(X).
K L

A B

For the multiplicative subset D = A r {0}, we have that K = D−1 A and L = Frac(D−1 B).
Write B = k[t̄1 , . . . , t̄p ] = A[t̄1 , . . . , t̄p ] and note that B is a finitely generated A-module.
Then K[t̄1 , . . . , t̄p ] = D−1 A[t̄1 , . . . , t̄p ] = D−1 B so D−1 B is a finitely generated K-algebra.
By Noether’s normalization lemma (Theorem 2.4.1), there exist elements h1 , . . . , h` ∈ D−1 B
which are algebraically independent over K such that D−1 B/K[h1 , . . . , h` ] is an integral
extension. Without loss of generality, we may assume h1 , . . . , h` ∈ B. Then we have the
following diagram:
K K(h1 , . . . , h` ) L

A K[h1 , . . . , h` ] D−1 B

By Proposition 1.4.8 and Corollary 12.1.17(a), we get


` = tr degK L = tr degk L − tr degk K = dim X − dim Y = r.
Note that while B/A[h1 , . . . , hr ] may not necessarily be integral, we can find an element f ∈ A
such that Bf /Af [h1 , . . . , hr ] is integral. Then h1 , . . . , hr are algebraically independent over
Af , so Af [h1 , . . . , hr ] is isomorphic to a polynomial ring in r variables, say Af [h1 , . . . , hr ] ∼
=
Af [u1 , . . . , ur ]. It follows that we have maps
ψ p
→ Yf × A r →
Xf − − Yf
where p is the coordinate projection and ψ is the finite morphism induced as a comorphism
by the k-algebra maps
k[Yf × Ar ] ∼
= k[Yf ][u1 , . . . , ur ] ∼
= k[Yf ][h1 , . . . , hr ] = Af [h1 , . . . , hr ] ,→ Bf .

257
12.4. Finite Morphisms and the Fibre Theorem Chapter 12. Algebraic Varieties

Set U = Yf , the principal open subset of Y generated by f . By construction, U ⊆ ϕ(X).


Take W ⊆ Y to be a closed, irreducible subset with W ∩ U 6= ∅. Set W 0 = W ∩ U ; then
W is a closed, irreducible subset of U = Yf so W 0 × Ar is closed and irreducible in U × Ar .
0

Hence W 0 × Ar is an affine variety. Moreover, we have

dim(ϕ−1 (W ) ∩ ϕ−1 (U )) = dim ϕ−1 (W ∩ U ) = dim ϕ−1 (W 0 )


= dim ψ −1 (W 0 × Ar ) since ψ is a finite morphism
= dim W 0 × Ar by Proposition 12.4.1(b)
= dim W 0 + r by Example 12.1.19
= dim W + r by Lemma 12.1.20.

The final statement follows by applying this to the irreducible subset W = {y} ⊆ U .

258
12.5. Projective Varieties Chapter 12. Algebraic Varieties

12.5 Projective Varieties


In this section we construct a projective analog to the notion of an affine variety. Let k be
a field; as in previous sections, we will often make the assumption that k is algebraically
closed.
Definition. For n ∈ N, we define projective n-space over k to be the quotient space
Pn = Pnk = An+1 r{0}/ ∼ where (a0 , . . . , an ) ∼ (b0 , . . . , bn ) if and only if there is some λ ∈ k ×
such that (b0 , . . . , bn ) = (λa0 , . . . , λan ). The coordinates of Pn are written [a0 , . . . , an ], called
homogeneous coordinates.
As in the affine case, for k ⊆ K ⊆ k̄ we can define Pnk (K) = {[a0 , . . . , an ] : ai ∈ K}.
Lemma 12.5.1. For any k ⊆ K ⊆ k̄, Pnk (K) = (Pnk (k̄))GK , where GK = Gal(k̄/K).
Proof. Apply Hilbert’s Theorem 90.
Definition. For a point P = [a0 , . . . , an] ∈ Pnk (k̄), the minimal field of definition for P
over k is the field k(P ) = k aa0i , . . . , aani where ai 6= 0. Alternatively, k(P ) = k̄ G(P ) where
G(P ) = {σ ∈ Gk | σ(P ) = P } ≤ Gk .
√ √ √
Example 12.5.2. The point P = ( 2, 2, 2) ∈ P3Q (Q) has minimal field of definition
Q(P ) = Q since scaling by √12 gives (1, 1, 1) ∈ A3Q .

Let S = k[t0 , . . . , tn ] be the polynomial ring in n + 1 indeterminates. Recall that S is a


graded ring with graded pieces given by total degree:

M
S= Sd where Sd = {f ∈ S | deg f = d}.
d=0

An arbitrary polynomial in S does not have a well-defined vanishing set in Pn . However,


homogeneous polynomials do have vanishing sets:
Definition. For f ∈ Sd , define the zero set of f to be

Z(f ) = {P ∈ Pn | f (P ) = 0},

where f (P ) = f (p0 , . . . , pn ) if P = [p0 , . . . , pn ]. This set is well-defined, since f ∈ Sd implies


f (λa0 , . . . , λan ) = λd f (a0 , . . . , an ) for all λ ∈ k × .
Set S h = ∞ of homogeneous polynomials F ⊆ S h , define the zero
S
d=0 Sd . For a collection
T
set of this collection by Z(F) = f ∈F Z(f ).
Definition. Let S = k[t0 , . . . , tn ] and suppose I ⊂ S is an ideal. Then S is homogeneous
if I = ∞
L
d=0 d where Id = I ∩ Sd for each d ∈ N0 .
I
Lemma 12.5.3. Let I ⊂ S be an ideal. Then
P∞
(a) I is homogeneous if and only if for all f = d=0 fd ∈ I, each fd ∈ I.

259
12.5. Projective Varieties Chapter 12. Algebraic Varieties

P T
(b) For any family (Iα )α∈A of homogeneous ideals of S, α∈A Iα and α∈A Iα are also
homogeneous ideals.

(c) For an element f ∈ Sd , f S is homogeneous.

(d) I is a homogeneous ideal if and only if I is generated by finitely many homogeneous


elements.

(e) If I is homogeneous then r(I) is homogeneous.

(f ) A homogeneous ideal p ⊂ S is prime if and only if f g ∈ p implies f ∈ p or g ∈ p for


all homogeneous elements f, g ∈ S h .

Proof. (a) is routine and (b) follows from applying (a) and the definitions of sums and
intersections of ideals.
(c) Write S = ∞
L L∞
j=d Sj−d . Then f S = j=d f Sj−d and note that (f S)j = f Sj−d makes
f S into a homogeneous ideal. L
(d) If I is homogeneous, I = ∞
S∞
d=0 (I ∩Sd ) so I is generated by the set d=0 (I ∩Sd ) which
consists of homogeneous elements. Finite generation
S∞ follows from Hilbert’s basis theorem
(3.3.2). Conversely, if I = (B) where B = P d=0 Bd and Bd ⊆ Sd for each d ∈ N0 , then for
any a ∈ I, there exist rb ∈ S such that a = b∈B rb b and finitely many of the rb are nonzero.
Then ∞ X ∞
X X M
a= rb b = rb b ∈ (I ∩ Sd ).
b∈B d=0 b∈Bd d=0
L∞ L∞
Therefore we have d=0 (I ∩ Sd ) ⊆ I ⊆ d=0 (I ∩ Sd ) and hence I is homogeneous.
(e) If f = m
P
f
d=0 d ∈ r(I) then by induction on 0 ≤ d ≤ m, it follows that fd ∈ r(I)
as well. For the base case, f ∈ r(I) implies there is some r ∈ N such that f r ∈ I. Then
(f r )0 = f0r ∈ I0r so f0 ∈ r(I). Then f − f0 ∈ r(I) which implies (f − f0 )` ∈ I for some ` ∈ N,
so f1` ∈ I` . Inductively, we get fd ∈ r(I).
(f) One direction is trivial. For the other, assume the statement holds for homogeneous
elements
P` in S and suppose Pm f g ∈ p and g 6∈ p for arbitrary elements f, g ∈ S. Write
f = d=0 fd and g = d=j gd with gj 6∈ p and fd , gd ∈ Sd . Then f0 gj ∈ pj = p ∩ Sj and so
f0 ∈ p. As in (e), we can continue inductively to get fd ∈ p for all d ≥ 0, so f ∈ p. Hence p
is prime.

Definition. Let X ⊆ Pn be any subset. The (homogeneous) vanishing ideal of X is


defined to be
J(X) = {f ∈ S h | f (P ) = 0 for all P ∈ X}.
X is called an algebraic subset if X = Z(I) for some homogeneous ideal I ⊂ S.

Lemma 12.5.4. Let I be an ideal of S and X ⊆ Pn a subset. Then

(a) If I = (f1 , . . . , fm ) then Z(I) = m


T
i=1 Z(fi ).

(b) J(X) is a homogeneous, radical ideal of S.

Proof. Similar to the proof of Lemma 12.1.1.

260
12.5. Projective Varieties Chapter 12. Algebraic Varieties

As in the affine case, the sets Z(I) form the closed sets in the Zariski topology on Pn .

Definition. For an algebraic subset X ⊆ Pn , the coordinate ring of X is S[X] = S/J(X).

Lemma 12.5.5. An algebraic set X ⊆ Pn is irreducible if and only if J(X) is prime in S.

Proof. See the proof of Lemma 12.1.9.

Remark. By Hilbert’s basis theorem (3.3.2), S = k[t0 , . . . , tn ] is noetherian and so Pn is a


noetherian space, as are its subspaces equipped with the Zariski topology.

We next define the projective version of an algebraic variety.

Definition. A projective variety is an irreducible closed subset of Pn . A quasi-projective


variety is a nonempty, open subset of a projective variety.

Theorem 12.5.6 (Hilbert’s Nullstellensatz, Projective Version). Let k be an algebraically


closed field and set S = k[t0 , . . . , tn ]. Then for any homogeneous ideal I ⊂ S,

(a) J(Z(I)) = r(I) if Z(I) 6= ∅.

(b) Z(I) = ∅ if and only if I = S or r(I) = (t0 , . . . , tn ).

Proof. (a) By Lemma 12.1.9(b) and Corollary 1.1.3, r(I) is a homogeneous ideal and clearly
r(I) ⊆ J(Z(I)). For the other direction, suppose f ∈ J(Z(I)) ∩ Sd and f 6= 0. Since
Z(I) 6= ∅, d must be at least 1, so f ∈ (t0 , . . . , tn ). In particular, f (0) = 0, i.e. f vanishes
at the origin in An+1 . Set

Za (I) = {P ∈ An+1 | f (P ) = 0 for all f ∈ I}.

Then the above shows f ∈ J(Za (I)). Notice that [a0 , . . . , an ] ∈ Z(I) ⊆ Pn if and only
if (a0 , . . . , an ) ∈ Za (I) r {0} ⊆ An+1 . By the affine Nullstellensatz (Theorem 12.1.5),
J(Za (I)) = r(I) so by the above, we must have f ∈ r(I). Therefore J(Z(I)) = r(I).
(b) If r(I) ⊇ (t0 , . . . , tn ) then Z(r(I)) ⊆ Z(t0 , . . . , tn ) = ∅, implying Z(I) = ∅. Con-
versely, Z(I) = ∅ implies Za (I) ⊆ {0}, where Za (I) is as defined in part (a). But Za (I) ⊆ {0}
impies J(Za (I)) ⊇ J({0}) = (t0 , . . . , tn ) by Lemma 12.1.2(a), so we have r(I) ⊇ (t0 , . . . , tn )
by the affine Nullstellensatz (Theorem 12.1.5). This completes the proof.

Corollary 12.5.7. For an algebraically closed field k, the algebraic subsets of Pn are in
bijective correspondence with the homogeneous radical ideals I ⊂ S, I =
6 (t0 , . . . , tn ).

As in the affine case, when k is algebraically closed we have correspondences


   
algebraic subsets homogeneous radical ideals
←→
X ⊆ Pn I ⊆ S, I 6= (t0 , . . . , tn )
   
projective varieties homogeneous prime ideals
←→ .
X ⊆ Pn p ∈ Spec(S), p 6= (t0 , . . . , tn )

261
12.5. Projective Varieties Chapter 12. Algebraic Varieties

Definition. Given a polynomial f ∈ k[t1 , . . . , tn ] of degree d, we obtain a homogeneous


form fh ∈ k[t0 , . . . , tn ] by defining
 
d x1 xn
fh (x0 , . . . , xn ) = x0 f ,..., ,
x0 x0

called the homogenization of f .


Conversely, a homogeneous polynomial F ∈ k[t0 , . . . , tn ] determines a polynomial F(i) ∈
k[t1 , . . . , tn ] for each 0 ≤ i ≤ n given by

F(i) (x1 , . . . , xn ) = F (x1 , . . . , xi−1 , 1, xi , . . . , xn ),

called the ith dehomogenization of F .

Definition. For an ideal I ⊆ k[t1 , . . . , tn ], define the homogenization of I by

Ih = {fh | f ∈ I} ⊆ k[t0 , . . . , tn ].

Likewise, for an ideal J ⊆ k[t0 , . . . , tn ], the ith dehomogenization of J is

J(i) = {F(i) | F ∈ J} ⊆ k[t1 , . . . , tn ].

Define the ith projective hyperplane by Hi = Z(ti ) ⊂ Pn for 0 ≤ i ≤ n. Set Ui = Pn r Hi ,


n
an open set in Pn . Then Pn = i=0 Ui , that is, the complements of the coordinate hyperplanes
S
are an open cover of Pn .

Proposition 12.5.8. Each Ui is homeomorphic to An . That is, Pn is locally affine.


 
Proof. Define ϕi : Ui → An by ϕi [a0 , . . . , an ] = aa0i , . . . , ai−1
ai
, ai+1
ai
, . . . , an
ai
. This is Zariski-
continuous and has a continuous inverse given by ψi (b1 , . . . , bn ) = [b1 , . . . , bi , 1, bi+1 , . . . , bn ].
Therefore ϕi is a Zariski-homeomorphism for each 0 ≤ i ≤ n.

Corollary 12.5.9. If Y ⊆ Pn is a projective variety, then Y = ni=0 (Y ∩ Ui ). In particu-


S
lar, every projective variety may be covered by open sets which are homeomorphic to affine
varieties in An .

For a projective algebraic set Y ⊂ Pnk , where Y = Z(J) for an ideal J ⊆ k[t0 , . . . , tn ],
we get n + 1 affine algebraic sets Yi = ϕ−1i (Y ∩ Ui ) = Z(J(i) ). These are called the deho-
mogenizations of Y . Conversely, for an affine algebraic set X ⊆ Akn , with X = Z(I), the
projective closure of X in Pnk is the Zariski closure in Pnk of ϕ0 (X), denoted X. Note that
X = Z(I(ϕ0 (X))) = Z(Ih ).

Lemma 12.5.10. The map ϕ0 |X : X → X ∩ ϕ0 (X) is a homeomorphism.

Using the affine patches Ui as charts on Pnk , we can define regular functions and morphisms
on a projective variety X as follows.

Definition. A function on X ⊆ Pnk is regular if it pulls back along Ui ,→ Pnk (i.e. restricts)
to a regular function on each affine patch Xi = Ui ∩ X.

262
12.5. Projective Varieties Chapter 12. Algebraic Varieties

Definition. Let X ⊆ Pnk be a projective variety. A rational function on X is an equiva-


lence class of quotients of homogeneous forms of the same degree,

F (x0 , . . . , xn )
f=
G(x0 , . . . , xn )
F1 F2
for F, G ∈ k[t0 , . . . , tn ], G 6∈ J(X), where we say f = G1
and g = G2
are equivalent if
F1 G2 − F2 G1 ∈ J(X).

Definition. The function field of X ⊆ Pnk is the set of rational functions on X, denoted
k(X).

Lemma 12.5.11. For each affine patch Xi = X ∩ Ui , k(X) ∼


= k(Xi ) as k-algebras.
In particular, if Y ⊆ Ank is an affine variety, then k(Y ) ∼
= k(Y ), where Y is the projective
closure of Y .
F
Definition. A function f ∈ k(X) is regular at a point P ∈ X if f can be written f = G
for homogeneous forms F, G ∈ k[t0 , . . . , tn ] such that G(P ) 6= 0.

Proposition 12.5.12. A projective variety X ⊆ Pnk is a ringed space with structure sheaf
OX : U 7→ OX (U ) defined on open sets U ⊆ X by

OX (U ) = {f ∈ k(X) | f is regular at P for all P ∈ U }.

We can now define morphisms between projective varieties using this ringed space struc-
ture.

Definition. A morphism of (quasi-)projective varieties is a map ϕ : X → Y that is a


morphism of the ringed spaces.

The definition of rational maps between affine varieties extends to projective varieties in
the following way.

Definition. For projective varieties X ⊆ Pnk and Y ⊆ Pm k , a rational map ϕ : X 99K Y is


a pair of open sets U ⊆ X and V ⊆ Y and a morphism ϕ = (ϕ0 , . . . , ϕm ) : U → V , such
that each ϕi ∈ k[t0 , . . . , tm ] is a homogeneous polynomial, ϕ(P ) ∈ Y for each P ∈ X and
some ϕi 6∈ J(X).

Definition. A map ϕ : X → Y is regular at a point P ∈ X if at least one ϕi (P ) 6= 0.


We say ϕ is a regular map if it is regular at every P ∈ X.

Lemma 12.5.13. A map ϕ : X → Y is regular if and only if it is a morphism of varieties.

Note that a quasi-projective set with the Zariski topology is not Hausdorff in general.
Indeed, if X is irreducible, then any nonempty open set is dense. Thus we need a notion to
replace the Hausdorff condition for algebraic sets.

Proposition 12.5.14. A quasi-projective set is T1.

263
12.5. Projective Varieties Chapter 12. Algebraic Varieties

Proof. If P = [α0 , . . . , αn ] ∈ X is a point then P = Z((αi tj − αj ti )i,j ) = Z(mP ). Thus points


are closed in the Zariski topology.

Corollary 12.5.15. If U ⊆ X is open, P, Q ∈ X and f (P ) = f (Q) for all f ∈ OX (U ), then


P = Q.

Corollary 12.5.16. Let X and Y be quasi-projective sets and ϕ, ψ : X → Y two morphisms.


Then if the set Uϕ,ψ := {P ∈ X | ϕ(P ) = ψ(P )} contains an open dense set, we have ϕ = ψ.

Definition. For a function f ∈ k[X], define the principal open subset of f by D(f ) :=
{P ∈ X | f (P ) 6= 0}.

Lemma 12.5.17. If X is a quasi-projective variety, then the collection {D(f ) | f ∈ k[X]}


is a basis for the Zariski topology on X.

264
12.6. Nonsingular Varieties Chapter 12. Algebraic Varieties

12.6 Nonsingular Varieties


Assume k is an algebraically closed field. Let X ⊆ An be a d-dimensional affine variety.

Definition. If J(X) = (f1 , . . . , fr ) is the vanishing ideal of X in A = k[t1 , . . . , tn ] and


x ∈ X, then the Jacobian of X at x is the r × n matrix
 
∂fi
Jx = (x) ,
∂tj

where derivatives of polynomials are just the formal power rule tkj 7→ ktjk−1 . We say X is
nonsingular, or smooth, at x ∈ X if rank Jx = n−d. Otherwise, X is said to be singular
at x.

Remark. X is nonsingular at a point x if and only if every open set U ⊆ X containing x is


nonsingular at x, which is even equivalent to there existing some open U ⊆ X containing x
such that U is nonsingular at x.

Lemma 12.6.1. Let A be a ring, M ⊂ A a maximal ideal and m ⊂ AM the corresponding


maximal ideal in the localization at M . Then

(a) A/M ∼
= AM /m.
(b) A/M 2 ∼
= AM /m2 .
(c) M/M 2 ∼
= m/m2 as A/M -vector spaces.
The following asserts that the geometric condition of smoothness is equivalent to the
algebraic condition of a ring being regular local. This is a fundamental connection in algebraic
geometry.

Theorem 12.6.2. An affine variety X ⊆ An is nonsingular at a point x ∈ X if and only if


the stalk OX,x is a regular local ring.

Proof. Let x = (x1 , . . . , xn ) ∈ X ⊆ An and define the corresponding maximal ideal mx =


(t1 − x1 , . . . , tn − xn ) ⊂ A = k[t1 , . . . , tn ]. Then there is a map

θ : A −→ k n
 
∂f ∂f
f 7−→ (x), . . . , (x) .
∂t1 ∂tn

Notice that θ is linear, and θ(ti − xi ) = (0, . . . , 1, . . . , 0) (with 1 in the ith coordinate), so θ
is surjective. In fact, this shows that θ|mx is surjective. We claim that ker θ|mx = m2x . On
one hand, for any 1 ≤ i, j, k ≤ n, we have

∂((ti − xi )(tj − xj ))
= δik (tj − xj ) + δjk (ti − xi )
∂tk

265
12.6. Nonsingular Varieties Chapter 12. Algebraic Varieties

which is zero on m2x . Take


X
g= gij (ti − xi )(tj − xj ) ∈ m2x ,
i,j

where gij ∈ A. Then for each 1 ≤ k ≤ n,

∂g X ∂((ti − xi )(tj − xj )) ∂gij


= gij + (ti − xi )(tj − xj )
∂tk i,j
∂tk ∂tk

2
P at x. Hence mx ⊆ ker θ|mx . On the other
which is zero hand, suppose f ∈ ker θ|mx r m2x .
Then f = i gi (ti − xi ) for some gi ∈ A. Since f 6∈ m2x , we must have gj 6∈ mx for some j.
Thus
∂f ∂ X ∂gi
= (gj (tj − xj )) + (ti − xi )
∂tj ∂tj i6=j
∂t j

which vanishes on x. Thus gj (x) = 0, implying gj ∈ J(x) = mx , a contradiction. Therefore


ker θ|mx = m2x as claimed.
We now get a k-linear isomorphism θ0 : mx /m2x → k n . Then J(X) = (f1 , . . . , fr ) ⊆ mx
is a k-linear subspace, so θ(J(X)) is a k-subspace of k n . This implies (θ0 )−1 (θ(J(X))) =
J(X) + mx /m2x and so the Jacobian at x is
 
θ(f1 )
Jx =  ...  .
 
θ(fr )

Thus rank Jx = dimk θ(J(X)) = dimk (J(X) + mx /m2x ). From Section 12.3, we know that
OX,x = k[X]mx and this is a local ring because mx is a maximal ideal. Let m be the unique
maximal ideal of OX,x and let m be the corresponding ideal in k[X]. By Lemma 12.6.1,
m/m2 ∼ = mx /m2x as k-vector spaces, so m/m2 ∼ = mx /m2x + J(X). As a result, we get
2
dimk m/m + rank Jx = n. Since k is algebraically closed, dim X = dim k[X] by Propo-
sition 12.1.16, so dim OX,x = dim k[X]mx = ht(mx ) by Lemma 2.3.6(a). Since X is affine,
k[X] is both a finitely generated k-algebra and an integral domain (Lemma 12.1.14), and
k[X]/mx ∼ = k as k-algebras, so Lemma 2.6.5 and Theorem 2.6.7 imply that ht(mx ) = d.
Putting everything together, we get

OX,x is regular ⇐⇒ dimk m/m2 = d ⇐⇒ rank Jx = n − d ⇐⇒ X is smooth at x.

One can define analogs of the sheaf of regular functions on an arbitrary variety X (includ-
ing affine, projective, quasi-affine and quasi-projective). The notion of smoothness generalize
to these cases:

Definition. An arbitrary variety X is nonsingular at a point x ∈ X if the stalk OX,x is a


regular ring. We say X is a nonsingular variety, or a smooth variety, if OX,x is regular
for all x ∈ X.

266
12.6. Nonsingular Varieties Chapter 12. Algebraic Varieties

One of the most important types of varieties is a hypersurface. As in ordinary geometry,


hypersurfaces are those objects with codimension 1.

Definition. A subset Y ⊆ An is a hypersurface if Y = Z(f ) for some irreducible, non-


constant polynomial f ∈ A.

Notice that if Y = Z(f ) is a hypersurface in An , with k[Y ] = k[t1 , . . . , tn ]/J(f ), Propo-


sition 2.6.6 implies that ht(f ) = 1 and dim Y = dim k[Y ] = dim k[t1 , . . . , tn ] − 1 = n − 1. So
indeed hypersurfaces are the varieties with codimension 1. The next proposition illustrates
why hypersurfaces are ubiquitous in modern algebraic geometry. First, we need:

Lemma 12.6.3. If A is a UFD with field of fractions K and f ∈ A[t] is an irreducible


polynomial, then K[t]/(f ) ∼
= Frac(A[t]/(f )).
Proposition 12.6.4. Let X be a d-dimensional affine variety. Then X is birationally equiv-
alent to a hypersurface in Ad+1 .

Proof. Let K = k(X) be the fractional field of X. Since k[X]/k is a finitely gener-
ated algebra, K/k is a finitely generated field extension. Moreover, assuming k is alge-
braically closed and therefore perfect, Proposition 1.4.9 says that K/k can be written as a
tower K ⊃ k(t1 , . . . , td ) ⊃ k such that K/k(t1 , . . . , td ) is finite and separable (using that
tr degk K = dim X = d from Corollary 12.1.17) and k(t1 , . . . , td ) is purely transcendental.
By the primitive element theorem, we may write K = k(t1 , . . . , td )(g) for some g ∈ K. In
particular, there exists an irreducible polynomial f ∈ k(t1 , . . . , td )[t] such that f (g) = 0 and
K∼ = k(t1 , . . . , td )[t]/(f ). Clearing denominators (e.g. using Gauss’s lemma), we get an irre-
ducible polynomial f˜ = λf ∈ k[t1 , . . . , td ][t] for some λ ∈ k[t1 , . . . , td ]. Let Y = Z(f˜) ⊆ Ad+1 .
Then k[Y ] = k[t1 , . . . , td ][t]/(f˜) so applying Lemma 12.6.3 with A = k[t1 , . . . , td ] and its field
of fractions k(t1 , . . . , td ), we get

k(X) = K ∼
= k(t1 , . . . , td )[t]/(f ) ∼
= Frac(k[t1 , . . . , td ][t]/(f˜)) = Frac(k[Y ]) = k(Y ).

Hence X and Y are birationally equivalent by definition.


For a d-dimensional affine variety X ⊆ An , let Sing X denote the set of singular points
of X, that is,
Sing X = {x ∈ X | rank Jx < n − d}.
(We will prove in a moment that rank Jx ≤ n − d always holds.) The following result
characterizes this set of singular points when X is affine.

Theorem 12.6.5. Suppose X ⊆ An is a d-dimensional affine variety. Then Sing X is a


proper closed subset of X.

Proof. We first prove Sing X is closed. In the proof of Theorem 12.6.2, we saw that for any
x ∈ X, rank Jx = n − dimk m/m2 where m is the maximal ideal of the local ring OX,x . By
Example 6.2.7, dimk m/m2 ≥ dim OX,x = d and so rank Jx ≤ n − d (justifying the above
description of Sing X). By linear algebra,

rank Jx = max{m ≥ 1 | there is some m × m minor C of Jx with det C 6= 0}.

267
12.6. Nonsingular Varieties Chapter 12. Algebraic Varieties

 
∂fi
Let J(X) = (f1 , . . . , fr ) and consider the operator J = . Then any determinant of a
∂tj
square minor of J is an element of A = k[t1 , . . . , td ]. Thus the set

D = {det C | C is an (n − d) × (n − d) minor of J }

is an ideal of A, and Sing X = Z(D) ∩ X. This shows Sing X is closed in X.


It remains to show Sing X 6= X. By Proposition 12.6.4, X is birationally equivalent to
a hypersurface Y ⊆ Ad+1 , say Y = Z(f ) for an irreducible polynomial f ∈ k[t1 , . . . , td+1 ].
Then  
∂f
Sing Y = y ∈ Y : (y) = 0 for all 1 ≤ i ≤ d + 1 .
∂ti
∂f ∂f
For each i, we have that deg ∂t i
≤ deg f − 1. If Sing Y = Y , we would have ∂t i
∈ J(Y ) = (f )
∂f
for all 1 ≤ i ≤ d + 1, and thus ∂ti = 0 for all i. If char k = 0, this implies f is a constant,
∂f
which is a contradiction. In the case that char k = p > 0, ∂t i
= 0 for all i implies that
f is a polynomial in the terms tpi . Taking pth roots of the coefficients, which is possible
since k is algebraically closed, we get that f = g p for some g ∈ k[t1 , . . . , td+1 ], contradicting
the irreducibility of f . Therefore Sing Y is a proper subset of Y . Now since X and Y
are birationally equivalent, there exist nonzero elements g ∈ k[X] and h ∈ k[Y ] such that
Xg ∼= Yh , by Proposition 12.3.3. Suppose Sing Yh = Yh . Then Sing Yh = Yh = Yh but
this implies Y ⊆ Sing Y , a contradiction. Thus Sing Yh is a proper subset of Yh . Now by
Proposition 12.3.3, we get

Xg ∼
= Yh =⇒ k[Xg ] ∼
= k[Yh ] =⇒ OX,x ∼
= OXg ,x ∼
= OYh ,y ∼
= OY,y

for all x ∈ Xg and y ∈ Yg . Thus Sing Yh ( Yh implies Sing Xg ( Xg , so it follows that


Sing X ( X.

Corollary 12.6.6. Let X be any variety. Then Sing X is a proper closed subset of X.

Proof. If X is quasi-affine, with X ⊆ Y ⊆ An , then X may be covered by finitely many


principal open subsets of Y , and Sing X is the intersection with the sets of singularities of
each of these. Apply Theorem 12.6.5 to get that Sing X is closed and Sing X ( X. If X
is projective, Corollary 12.5.9 says that X has a covering by finitely many affine varieties,
so Theorem 12.6.5 gives the result. Finally, if X is quasi-projective, then X has a finite
covering by quasi-affine varieties, so apply the above.

268
12.7. Abstract Curves Chapter 12. Algebraic Varieties

12.7 Abstract Curves


Assume k is algebraically closed.
Definition. An affine curve is an affine variety X over k of dimension 1.
By Corollary 12.1.17(a), an affine curve is a variety whose function field k(X) has tran-
scendence degree 1 over k. If X is a nonsingular affine curve, as in Section 12.6, for each
P ∈ X, the ring OP := OX,P = k[X]mP ⊆ k(X) is a regular local ring (Theorem 12.6.2).
T OP is a DVR for each P ∈ X. Also, for any open subset U ⊆ X, let
In particular,
OX (U ) = P ∈U OP be the corresponding ring of k-valued functions on U .
We seek to reproduce this notion of an algebraic curve in a more general setting. Let K
be a finitely generated field extension of k with tr degk K = 1, that is, K is a function field
of degree 1 over k. Suppose v : K × → Z is a discrete valuation, with discrete valuation ring
R = O(v) = {x ∈ K | v(x) ≥ 0} and maximal ideal mv = {x ∈ K | v(x) > 0}. We say R is
a DVR over k if v(k × ) = {0}. Let CK be the set of all DVRs over k contained in K.
Lemma 12.7.1. Let Y be a quasi-projective variety and P, Q ∈ Y be two points. Then
OP ⊆ OQ implies P = Q.
Proof. In general, OP only depends on open sets around P so we may replace Y with
its Zariski-closure Y and assume Y is projective. Further, supposing P and Q don’t lie
in a coordinate hyperplane, we may in fact assume Y is affine. Then OP = k[Y ]mP and
OQ = k[Y ]mQ , so OP ⊆ OQ implies that mQ ⊆ mP . Since these are maximal ideals, we get
mQ = mP , but then Corollary 2.4.3 gives us P = Q.
Lemma 12.7.2. If K is a function field of degree 1 over k and x ∈ K, then {R ∈ CK | x 6∈
R} is a finite set.
Proof. For any DVR R over k, x 6∈ R if and only if x1 ∈ mR so it suffices to show that for
any nonzero y ∈ K × , the set {R ∈ CK | y ∈ mR } is finite. If y ∈ k × , then y 6∈ mR for
any R ∈ CK since mR consists of the elements of K with positive valuation. Now assume
y 6∈ k. Then k[y] ⊆ K, k[y] is isomorphic as a k-algebra to some polynomial ring over k, and
by Lemma 1.4.3(b), K/k(y) is a finitely generated, algebraic extension of fields. Let B be
the integral closure of k[y] in K. Then k[y] is a PID, and in particular a Dedekind domain,
so by Theorem 4.5.3, B is also a Dedekind domain. In particular, B is a finitely generated
k-algebra with field of fractions K.
Suppose y ∈ R for some DVR R ∈ CK . Then k[y] ⊆ R implies B ⊆ R since R is
integrally closed in K. Let n = mR ∩ B; we claim this is a maximal ideal in B. To see
this, note that since n = mR ∩ B, this is a prime ideal, and since B is Dedekind, it suffices
to prove n 6= 0. Let v be the valuation for R. If n = 0, then v(B r {0}) = {0} implying
v(K × ) = {0}, contradicting the fact that R is a discrete valuation ring. So n must in fact
be maximal in B. Now B ⊆ R and B r n ⊆ R× so localizing at n gives us Bn ⊆ R. Since B
is Dedekind and n is nonzero, Bn must in fact be a DVR by Proposition 4.2.2.
We proceed to show that Bn = R. To do this, let v 0 be the valuation for Bn and let
S = B rn, i.e. so Bn = S −1 B. If x ∈ R r B 0
n then we would have v (x) < 0, and
thus v 0 x1 > 0, so x1 ∈ S −1 n ⊆ mR . Thus v x1 > 0 which is equivalent to v(x) < 0,


269
12.7. Abstract Curves Chapter 12. Algebraic Varieties

contradicting x ∈ R. Hence Bn = R. Now suppose y ∈ mR = S −1 n. Since B is a finitely


generated k-algebra, it is the coordinate ring for some affine variety Y over k. Moreover,
dim B = 1 because it is a Dedekind domain, so dim Y = 1 and the localizations of B at all
maximal ideals are all DVRs (Proposition 4.2.2), so it follows from Theorem 12.6.2 that Y is
nonsingular. Then y ∈ k[Y ] = B so y may viewed as a regular function on Y . In particular,
y = y1 ∈ S −1 n = mR is equivalent to y vanishing at the point in affine space corresponding
to mR by Corollary 2.4.3. However, Z(y) is a nonempty, proper subset of Y and dim Y = 1,
so by Corollary 12.1.17(b), Z(y) must be a finite set of points. This says that y vanishes at
only finitely many points in affine k-space, so by the above, y ∈ mR for only finitely many
R ∈ CK .

Corollary 12.7.3. Any DVR of the field extension K/k is the local ring of a variety Y at
some point P ∈ Y .

Proof. Let Y and y be as in the proof of Lemma 12.7.2.

Corollary 12.7.4. If R ∈ CK , then R/mR ∼


= k.
Proof. By Corollary 12.7.3, R = OP = k[Y ]mP for some point P ∈ K. Further, Lemma 1.3.11
gives us k[Y ]mP /S −1 mP ∼
= k[Y ]/mP = k, where S = k[Y ] r mP .
Let K/k be a function field of degree 1 and as above, let CK be the set of all DVRs of
K/k. For any affine variety X over k with k(X) = K, dim X = 1 and in particular X is
infinite (as a set). By Corollary 12.7.3, CK contains the local ring OP for each point P ∈ X,
but by Lemma 12.7.1, these OP are all distinct, so it follows that CK itself is an infinite set.
Then the finite sets of CK along with CK itself define the closed sets of a topology on CK .
A ‘point’ in CK is a discrete valuation ring RP ; we will conflate the notation
T P and RP .
For an open set U ⊆ CK , define its sheaf of regular functions O(U ) = P ∈U RP , where an
element f ∈ O(U ) is represented by a regular function f : U → k such that f (P ) is the
residue of f mod mP . Notice that if f, g ∈ O(U ) define the same function, then f − g ∈ mR
for infinitely many R ∈ CK . By Lemma 12.7.2, f = g so elements of O(U ) are well-defined.

Definition. An abstract nonsingular curve over a field k is an open subset U ⊆ CK for


some function field K/k of degree 1, with the induced topology and sheaf of regular functions
as described above.

Definition. A morphism of nonsingular curves is a continuous map ϕ : X → Y , where


X and Y are nonsingular curves, such that for all open subsets V ⊆ Y and all regular
functions f : V → k, f ◦ ϕ : ϕ−1 (V ) → k is regular.

The next result shows that abstract nonsingular curves properly generalize nonsingular
quasi-projective curves as defined in Section 12.6.

Theorem 12.7.5. If X is a nonsingular quasi-projective curve over k then X is isomorphic


to an abstract nonsingular curve.

Proof. Let K = k(X) and recall that Corollary 12.1.17(a) implies tr degk K = 1. Define a
map ϕ : X → CK that sends P 7→ OP . If we set U = ϕ(X), changing targets gives us a

270
12.7. Abstract Curves Chapter 12. Algebraic Varieties

surjection ϕ : X → U . Moreover, Lemma 12.7.1 T says that ϕ isTinjective. If we show U is an


open set in CK , we will be done since O(U ) = P ∈U OP ∼ = P ∈ϕ(U ) RP = O(ϕ(U )) shows
that ϕ is an isomorphism. Now to show U is an open set, it is enough to show U contains
an open set of CK since open sets are by definition the complements of finite sets.
By Corollary 12.5.9, X may be covered by affine varieties X1 , . . . , X` . Then ϕ(X1 ), . . . , ϕ(X` )
cover U = ϕ(X) so it suffices to show each ϕ(Xj ) is open in CK . Thus we may assume X = Xj
is affine. We claim that U = {R ∈ CK | k[X] ⊆ R}. Note that the statement k[X] ⊆ R is
logical because K = k(X) and so k[X] ⊆ K. On one hand, if R ∈ U is a DVR over k, then
R = OP for some point P ∈ X, and in this case k[X] ⊆ R. On the other hand, suppose
k[X] ⊆ R for R ∈ CK . Then k[X] is a Dedekind domain and R is a DVR, so as in the proof
of Lemma 12.7.2, n = mR ∩ k[X] is a maximal ideal of k[X] and k[X]n = R. Thus R is the
localization of k[X] at a maximal ideal corresponding to a point in X, i.e. R ∈ ϕ(X). Hence
U has the desired description.
Now use the fact that k[X] is a finitely generated k-algebra to write k[X] = k[x1 , . . . , xn ]
for elements x1 , . . . , xn ∈ k[X]. Then for any R ∈ CK , k[X] ⊆ R if and only if xi ∈ R for
each 1 ≤ i ≤ n. Thus U = ni=1 Ui , where Ui = {R ∈ CK | xi ∈ R}. By Lemma 12.7.2, each
T
Ui is the complement of a finite set in CK , so U is open as required.

Theorem 12.7.6 (Extension). If X is an abstract nonsingular curve, P ∈ X is any point


and ϕ : X r {P } → Y is a morphism into a projective variety Y , then there exists a unique
morphism Φ : X → Y extending ϕ to all of X.

Proof. (Sketch) Let Y ⊆ Pn for n ≥ 1. If the map ψ : X r {P } → Pn (obtained from ϕ by


changing targets) can be extended to all of X, then ψ −1 (Y ) is a closed set containing X r{P }.
In particular, X r {P } is nonempty and open, so this would imply ψ −1 (Y ) = X. Thus
ϕ(X) = ψ(X) ⊆ Y so we may reduce to the case where T Y = Pn . Let U0 , . . . , Un be the affine
coordinate charts of Pn as S in Corollary 12.5.9. Set U = ni=0 Ui . If ϕ(X r {P }) ∩ U = ∅, we
would have ϕ(X r{P }) ⊆ ni=0 Hi , where Hi = Pn rUi is the ith projective hyperplane. Note
that X r {P } is irreducible, so ϕ(X r {P }) is irreducible and thus ϕ(X r {P }) ⊆ Hi ∼ = Pn−1
for some 0 ≤ i ≤ n. Restricting the target of ϕ to projective n − 1 space, we may assume by
induction that ϕ(X r {P }) ∩ U 6= ∅.
On U , each xxji for 0 ≤ i, j ≤ n is a regular function. Let fij = xxji ◦ ϕ : ϕ−1 (U ) → k for
0 ≤ i, j ≤ n; each fij is also a regular function, so view fij ∈ K. Then we can choose ` ≥ 1
such that Φ(P ) = [f0` (P ), . . . , fn` (P )]. This defines an extension Φ : X → Pn ; one may now
check that Φ is a morphism.
One can prove:

Theorem 12.7.7. If K is a function field over k of degree 1 then CK is isomorphic to a


nonsingular projective curve over k.

Corollary 12.7.8. Every abstract nonsingular curve is isomorphic to some quasi-projective


nonsingular curve over a field k.

Thus the abstract theory of curves presented in this section captures the more concrete
notions of affine, projective and quasi-affine/projective varieties detailed in earlier sections.

271
12.8. Products of Varieties Chapter 12. Algebraic Varieties

12.8 Products of Varieties


Consider two ringed spaces X and Y . We may take their set-theoretic product X × Y and,
if each space is a topological space, endow X × Y with the product topology. Unfortunately,
in the category of algebraic varieties, this operation does not preserve the structure of two
varieties X and Y ; that is, the product topology arising from the spaces’ Zariski topologies
does not suffice to do algebraic geometry.
Instead, consider the projections πX : X × Y → X and πY : X × Y → Y . For any ringed
space Z, we must have a bijection

Hom(Z, X × Y ) ←→ Hom(Z, X) × Hom(Z, Y )


ϕ 7−→ (ϕ ◦ πX , ϕ ◦ πY ).

We thus make X × Y into a ringed space with OX×Y (U × V ) defined for all open sets
U ⊆ X, V ⊆ Y by stipulating that anything of the form
X

f= (πX gi )(πY∗ hi ), for gi ∈ OX (U ) and hi ∈ OY (V ),

is regular on U × V . If g ∈ OX (U ), we must have πX g ∈ OX×Y (U × V ) and likewise, if

h ∈ OY (V ), then πY h ∈ OX×Y (U × V ). Thus for such an f as above, D(f ) is an open subset
of X × Y that would not be open in the usual product topology.
Example 12.8.1. Under the above description of products of affine varieties, An × Am ∼
=
n+m 2
A for any n, m ∈ N. Note that even for n = m = 1, the Zariski topology on A is not
equivalent to the product topology on A1 × A1 .
Lemma 12.8.2. If X and Y are affine varieties, then
(a) X × Y is an affine variety.

(b) k[X × Y ] = k[X] ⊗k k[Y ].


To define products of projective varieties requires a little more care.
Proposition 12.8.3 (Segre Embedding). For any n, m ∈ N, there is an embedding

σn,m : Pn × Pm −→ P(n+1)(m+1)−1
([x0 , . . . , xn ], [y0 , . . . , ym ]) 7−→ [xi yj ]i,j

such that the image Σn,m := σn,m (Pn × Pm ) has the structure of an algebraic subset that
coincides with the Zariski topology of the product Pn × Pm .
Proof. (Sketch) Viewing P(n+1)(m+1)−1 as a space of (n + 1) × (m + 1) matrices, we have that

Σn,m = {[zij ]i,j | all 2 × 2 minors of (zij ) vanish}.

Then clearly Σn,m = Z((zij zk` − zkj zi` )i,j,k,` ), so Σn,m is an algebraic set. The fact that σn,m
is a bijection is obvious. One can now verify that the induced topology corresponds to the
topology on Pn × Pm .

272
12.8. Products of Varieties Chapter 12. Algebraic Varieties

Definition. Let V be a vector space over k. The set of lines, i.e. 1-dimensional subspaces,
of V is called the projective space over V , denoted P(V ).

Example 12.8.4. If V = k n is finite dimensional, then P(V ) can be identified with Pnk .

Lemma 12.8.5. If V and W are k-vector spaces, then P(V ) × P(W ) ∼


= P(V ⊗k W ).
Now if X ⊆ Pn and Y ⊆ Pm are projective algebraic sets, we can realize X × Y as a
subset of P(n+1)(m+1)−1 by identifying it with the embedded image σn,m (X × Y ). Setting

OX×Y (U × V ) := OP(n+1)(m+1)−1 (σn,m (U × V ))

gives X × Y the structure of a ringed space which coincides with the previous description of
the product of two varieties.

Proposition 12.8.6. Subvarieties of Pn × Pm are zero sets of polynomials of the form

G(x0 , . . . , xn , y0 , . . . , ym ),

where G is homogeneous in the xi of degree d and homogeneous in the yj of degree e.

Proof. Without loss of generality, suppose d ≥ e. Then G(x0 , . . . , ym ) = 0 if and only if


yid−e G(x0 , . . . , ym ) = 0 and the latter polynomial is homogeneous of a single degree. Viewing
Pn × Pm as the embedded image Σn,m ⊆ P(n+1)(m+1)−1 gives the result.

Example 12.8.7. Consider the Segre embedding P1k × P1k ,→ P3k and set Q = Σ1,1 =
Z(z00 z11 − z01 z10 ). The polynomial z00 z11 − z01 z10 is called a quadric and the embedded
image Q is called a quadric surface. For each α, β ∈ P1k , one gets lines on the quadric surface
realized by {α} × P1k ,→ Q and P1k × {β} ,→ Q. Note that lines of these forms cover Q, for
which reason Q is called a ruled surface.

273
12.9. Blowing Up Chapter 12. Algebraic Varieties

12.9 Blowing Up
We now have a working notion of products of varieties, so consider the space An × Pn−1 .
Coordinates in this space are (P, [`]), where P ∈ An is a point and [`] ∈ Pn−1 is the class of
some line through the origin ` in An . Consider the set B ⊆ An × Pn−1 defined by

B = {(P, [`]) | P ∈ `}.

Then B is an algebraic subset: B = Z((xi yj − xj yi )i,j ) if An = {(x1 , . . . , xn )} and Pn−1 =


{[Y1 , . . . , Yn ]}.
In dimension n = 2, notice that for any point P = (u, v) and line [`] = [α, β], we have
u α
P ∈ ` ⇐⇒ = ⇐⇒ uβ − vα = 0.
v β

This explains why we can write B = Z(x1 Y2 − x2 Y1 ) ⊆ A2 × P1 .


Now let π : An × Pn−1 → An be the canonical projection. If P 6= 0 in An , then π −1 (P ) =
(P, [`P ]) is defined, where `P is the unique line through the origin containing P . Therefore
π is an isomorphism on an open subset of B:

π : B r {(P, [`]) | P = 0} −→ An r {0}.

On the other hand, if P = 0, the set π −1 (0) = {(0, [`]) ∈ An × Pn−1 } is isomorphic to Pn−1 .
In the dimension 2 case, B is covered by the following affine patches:

U1 = {((x, y), [Y1 , Y2 ]) | Y1 6= 0} ∩ B and U2 = {((x, y), [Y1 , Y2 ]) | Y2 6= 0} ∩ B

On U1 , set t = YY12 so that in local coordinates (x, y, t), U1 = Z(xt − y) ∼


= A2 . Likewise, for
U2 , set s = YY21 so that in the coordinates (x, y, s), U2 = Z(x − ys) ∼
= A2 . Thus we see that
each affine patch Ui is a quadric surface. Effectively, we have replace a point (0, 0) in A2 with
a copy of P1 so that every line through the origin in A2 , all of which are indistinguishable in
P1 to begin with, now corresponds to a unique line on one of the affine quadric surfaces.

Definition. The set B is called the blowup of An at the point 0, denoted B = Bl0 An . The
set E0 An := π −1 (0) ∼
= Pn−1 is called the exceptional divisor of the blowup.
Definition. Let X ⊆ An be an affine variety and π : An × Pn−1 → An the canonical
projection. The pullback π −1 (X) is called the total transform of X, while the proper (or
strict) transform of X is defined as

Bl0 X := π −1 (X r {0}).

As the notation suggests, this set is also called the blowup of X at 0. The set

E0 X := Bl0 X ∩ E0 An

is called the exceptional divisor of the blowup of X.

274
12.9. Blowing Up Chapter 12. Algebraic Varieties

Remark. More generally, for any subvariety Z ⊆ X, one can define the blowup of X along
Z, a variety BlZ X that is birationally equivalent to X, such that Z is a codimension 1
subvariety of BlZ X.

Example 12.9.1. Consider the plane curve X = Z(y 2 − x2 (x + 1)) ⊆ A2 .

Note that this variety has a singularity at the point (0, 0). Using the blowup of A2 defined
above, Bl0 A2 , we can blowup X to ‘remove the singularity’ at 0. Let U1 be the first affine
patch and ϕ : U1 → A2 the standard isomorphism. We make the substitution y = xt, so
that ϕ(π −1 (X)) = Z(x2 (t2 − x − 1)). The x2 factor of this polynomial corresponds to the
exceptional divisor E0 X under this blowup, so the proper transform of X at 0 looks like

ϕ(Bl0 (X)) = Z(t2 − x − 1)

on the affine patch U1 . Note also that

E0 X = Bl0 X ∩ E0 A2 = Z(t2 − x − 1) ∩ Z(x) = Z(t2 − 1) = {±1},

so the exceptional set of X consists of two points.

Lemma 12.9.2. The projection π : Bl0 X 99K X is a birational equivalence.

Blowing up allows us to replace singular curves (or more generally, varieties) with non-
singular curves by a sequence of blowups, such that in each step the birational equivalence
class of the curve is preserved. The problem of finding such a nonsingular blowup is known
as resolution of singularities. Much progress has been made on this problem (e.g. Hironaka’s
theorem says that nonsingular blowups exist for any finite dimensional variety over a field
of characteristic zero), but there is still much to be done (e.g. in finite characteristic cases).

275
12.10. Prevarieties and Varieties Chapter 12. Algebraic Varieties

12.10 Prevarieties and Varieties


In this section we give a more abstract notion of an algebraic variety in the vein of the
generalization of curves described in Section 12.7.
Informally, a prevariety is a space that locally looks like an affine variety. The rigorous
definition is given below.
Definition. For an algebraically closed field k, a prevariety over k is a pair (X, OX )
consisting of a topological space X and a sheaf OX of k-valued functions on X Snsuch that
n
there exists a collection {Ui }i=1 of nonempty open subsets of X for which X = i=1 Ui and
(Ui , OX (Ui )) is an affine variety for each i.
One can see that prevarieties closely resemble manifolds, which are ‘locally euclidean’.
Prevarieties over k form a category pShk with products (X × Y again). The name is often
shortened to just ‘variety’ even though there is a formal definition of varieties, given here.
Definition. A variety is a prevariety in which the diagonal ∆X := {(x, x) | x ∈ X} is
closed in X × X.
The first thing to check is that this new notion of variety captures the affine case.
Proposition 12.10.1. If (X, OX ) is an affine variety then it is a variety.
Proof. Let A = k[X]. Then k[X × X] = A ⊗k A which is an affine k-algebra. Let I be the
ideal in A ⊗k A generated by {a ⊗ 1 − 1 ⊗ a : a ∈ A}. Then ∆X = V(I). For any x, y ∈ X,
(a ⊗ 1 ⊗ a)(x, y) = a(x) − a(y) so if x = y, this evaluates to 0 but if x 6= y, there’s always an
a ∈ A such that a(x) 6= a(y). Therefore ∆X is Zariski closed.
Proposition 12.10.2. Let X and Y be varieties. Then the product X × Y is a variety and
is unique up to isomorphism.
Proof. Let X = m
S Sn
i=1 Ui and Y = j=1 Vj be covers of X, Y by affine open sets. Then X × Y
is covered by {Ui × Vj }i,j and each of these is affine. Suppose f is a function defined on an
open neighborhood U ⊂ Ui × Vj with x ∈ U . Then f is regular if f |U ∩(Ui ×Vj ) is regular for
the affine structure on Ui × Vj . The space of such regular functions is a sheaf on X × Y . It is
easy to verify that X × Y is a variety with these features defined, and that it is unique.
Proposition 12.10.3. Let Y be a prevariety and let X be a variety, and suppose ϕ ∈
HompShk (Y, X) is a morphism of prevarieties. Then
(a) The graph Γϕ := {(y, ϕ(y)) | y ∈ Y } is closed in Y × X.
(b) Given a morphism ψ : Y → X such that ψ = ϕ on a dense subset of Y , ψ = ϕ
everywhere on Y .
Proof. (a) Define the continuous map α : Y × X → X × X defined by (y, x) 7→ (ϕ(y), x).
Then Γϕ = α−1 (∆X ) and since X is a variety, the diagonal is closed. Hence Γϕ is closed.
(b) Suppose ϕ and ψ agree on a dense subset D ⊂ Y . Consider the map ϕ × ψ : Y →
X × X given by y 7→ (ϕ(y), ψ(y)). Then (ϕ × ψ)−1 (∆X ) = {y ∈ Y | ϕ(y) = ψ(y)} =: B,
but clearly D is contained in B. Since X is a variety, B is closed, but the only closed set
containing D is the whole space Y . Therefore {y ∈ Y | ϕ(y) = ψ(y)} = Y .

276
12.10. Prevarieties and Varieties Chapter 12. Algebraic Varieties

Example 12.10.4. Part (b) of Proposition 12.10.3 does not hold in the category of preva-
rieties. For instance, take the affine line with a double point, i.e. A e 1 := A1 ∪ {00 } where
e 1 r {0} ∼
A =A e 1 r {00 } ∼
= A1 . Then A
e 1 is a prevariety but not a variety, and there are maps
ϕ, ψ : A1 → A e 1 with ϕ(0) = 0 and ψ(0) = 00 . The maps agree on a dense subset, namely
e 1 r {0, 00 }, but they are not the same map on the whole space.
A

Proposition 12.10.5. Let X be a variety. If U and V are nonempty, open affine subvarieties
of X then U ∩ V is an open affine variety.

Proof. First, ∆X ∩ (U × V is closed in U × V . Then the inclusion i : X → ∆X induces an


isomorphism U ∩ V ∼ = ∆X (U × V ). Let OX (U ∩ V ) denote the restriction of all functions in
OX (U ) and OX (V ) to U ∩ V . Then OX (U ∩ V ) is a sheaf of k-valued functions on U ∩ V .
In other words, OX (U ∩ V ) = k[U ] ⊗ k[V ], which is generated by k[U ] ⊗ 1 and 1 ⊗ k[V ]. It
follows that U ∩ V is an affine variety.

Remark. If U and V are nonempty, open affine subvarieties of a variety X then the ring of
k-valued functions on U ∩ V , OX (U ∩ V ), is generated by the restrictions of all functions in
OX (U ) and OX (V ) to U ∩ V .

277
12.11. Complete Varieties Chapter 12. Algebraic Varieties

12.11 Complete Varieties


Definition. A variety X is complete if for any variety Y , the projection map X × Y →
Y, (x, y) 7→ y, is a closed map.

Proposition 12.11.1. Let X be a complete variety. Then

(1) Any closed subvariety of X is complete.

(2) If Y is complete then X × Y is also complete.

(3) For every morphism ϕ : X → Y , ϕ(X) is closed in Y and complete.

(4) If X ⊆ Y as a subvariety, then X is closed.

(5) If X is irreducible, then OX (X) consists of constant functions.

(6) If X is affine then |X| < ∞.

Proof. (1) Let X 0 ⊆ X be a closed subvariety and Y any variety, and consider π 0 : X 0 × Y →
Y . Suppose Z ⊆ X 0 × Y is a closed subset. In general X 0 × Y is closed in X × Y so the
diagram
i
Z ⊆ X0 × Y X ×Y

π0 π

commutes and thus the image of Z is closed.


(2) Assume Y is complete and let Z be an arbitrary variety. We can factor the map
X × Y × Z → Z as
X ×Y ×Z →Y ×Z →Z
but both of these maps are closed since X and Y are each complete. The composition of
closed maps is closed, so X × Y is complete.
(3) Let Γ = {(x, ϕ(x)) | x ∈ X} ⊆ X × Y be the graph of ϕ.. Then Γ is closed, so ϕ(X)
is the projection of Γ = X × ϕ(X) onto Y , and since X is complete, ϕ(X) is closed. For
completeness, use (1).
(4) follows from applying (3) to the inclusion i : X → Y .
(5) Let f : X → k be a regular function on X. This induces a morphism ϕ : X → A1 ⊆ P1
where P1 is the projective line. We know ϕ(X) is closed in A1 but since the only proper
closed subsets of A1 are finite, we see that ϕ is surjective onto A1 . Moreover, ϕ(X) is also
closed in P1 , but this is impossible since A1 = ϕ(X) is open in P1 and therefore cannot equal
P1 . Therefore ϕ(X) must be empty, so f is a constant function, i.e. OX (X) = k.
(6) follows from (5).
Examples.

278
12.11. Complete Varieties Chapter 12. Algebraic Varieties

1 Finite varieties (collections of points) are complete.

2 If V is a finite dimensional vector space over k, then the projective space P(V ) consist-
ing of one-dimensional subspaces of V is complete. By (1) of Proposition 12.11.1 below,
every closed subvariety of P(V ) is complete, so all projective varieties are complete.
We now prove an important property of projective varieties.
Theorem 12.11.2. Every projective variety is complete.
Proof. We proved that every closed subvariety of a complete variety is complete, so it suffices
to prove Pn is complete for all n ≥ 1. In other words we will show that if π : Pn × Y → Y
is the projection map and C ⊂ Pn × Y is closed then π(C) ⊆ Y is closed. Set A = k[Y ] and
B = A[T0 , T1 , . . . , Tn ]. Then B is a ring of k-valued functions on k n+1 × Y . For every proper
homogeneous ideal I ⊂ B, define
V ∗ (I) = {(x∗ , y) | f (x, y) = 0 for all f ∈ I} ⊆ Pn × Y.
Then the V ∗ (I) are the closed subsets of Pn × Y so it suffices to prove π(V ∗ (I)) is closed for
all proper homogeneous ideals I ⊂ B. We may assume V ∗ (I) is irreducible, i.e. I is prime.
We may also assume π|V ∗ (I) is dominant (changing the target to π(V ∗ (I)) if necessary). Then
we must show for every y ∈ Y , there exists x∗ ∈ Pn so that (x∗ , y) ∈ V ∗ (I), since then we
will have π(V ∗ (I)) = π(V ∗ (I)).
Take M ⊂ A to be the maximal ideal that vanishes at y. Then J = M B + I is a
homogeneous ideal so V ∗ (J) is defined, and if we show V ∗ (J) is nonempty, we’ll be done.
Assume to the contrary that V ∗ (J) = ∅. Then there is a k > 0 such that Tik ∈ J for each Ti .
Equivalently, there is an m > 0 so that Bm , the set of all degree m homogeneous polynomials
in B, is contained in J. Set N = Bm /(Bm ∩ I). This is a finitely generated A-module in the
obvious way. Moreover, notice that M N = N . Then by Nakayama’s Lemma (1.2.1), this
implies N = 0. But then Bm = Bm ∩ I so it follows that V ∗ (I) = ∅, which is impossible for
a proper ideal I ⊂ B. Hence V ∗ (J) 6= ∅ so the theorem is proved.
There are examples of complete, non-projective varieties but they are difficult to come
up with. However, we have:
Theorem 12.11.3. (Chow’s Lemma) For any complete irreducible variety X, there exists a
projective variety X and a surjective birational morphism f : X → X.
Recall from general topology that a topological space X is compact if and only if for
all spaces Y , X × Y → Y is a closed map. In this way complete varieties are an algebraic
analogue of compact spaces. This analogy runs quite deep:
Proposition 12.11.4. Let X be a connected complete variety. Then the sheaf OX (X) =
k[X] equals k. That is, every regular k-valued function on X is constant.
Proof. Take f ∈ OX (X). Then f is a map f : X → k = A1 . Extend this to a map
g : X → A1 ,→ P1 ,
so g is not surjective onto P1 . By completeness of X, g(X) is closed in P1 , but the only
proper closed subsets of P1 are point-sets. Therefore g(X) = {x0 }, or in other words, g is
constant. This implies f is constant.

279
12.11. Complete Varieties Chapter 12. Algebraic Varieties

This fact is analagous to the theorem in complex analysis that every holomorphic function
on a connected compact domain is constant.

Corollary 12.11.5. Nontrivial affine varieties are not projective.

Proof. Let X be an affine variety of dimension at least 1. View X as a proper subset of affine
n-space An , which has coordinate algebra k[T1 , . . . , Tn ]. Then some coordinate function Ti
does not vanish on X, so Ti ∈ OX (X) is a nonconstant regular function on X.

280
Chapter 13

Sheaf Theory

In this chapter, we describe the theory of sheaves on topological spaces. The basic intuition
is that for a topological space X, one wishes to study a given property defined on open sets
U ⊆ X that behaves well under union and intersection of open sets. Sheaves will allow us
in the following chapter to construct a generalization of a variety: a scheme.

281
13.1. Sheaves and Sections Chapter 13. Sheaf Theory

13.1 Sheaves and Sections


Definition. Let p : E → X be a morphism in a category C. A section of p is a morphism
σ : X → E such that p ◦ σ = idX .
Definition. When p : E → X is a continuous map of topological spaces, a local section
of p over an open set U ⊆ X is a map σ : U → E such that p ◦ σ = idX |U . The set of all
sections of p over U is denoted Γp (U ). A global section of p is a section over the open set
U = X.
The following observations, which are easy to prove, are a natural starting point for the
study of sheaves.
Lemma 13.1.1. Let p : E → X be a continuous map of topological spaces. Then
(a) If V ⊆ U ⊆ X are open sets, then there is a restriction map

Γp (U ) −→ Γp (V )
σ 7−→ σ|V .

That is, Γp (−) is a contravariant functor on the category TopX of open sets on X.
S
(b) Let {Ui } be a collection of open sets in X, with U = Ui , and suppose there exist
local sections σi ∈ Γp (Ui ) such that σi |Ui ∩Uj = σj |Ui ∩Uj for all i, j. Then there exists a
unique section σ ∈ Γp (U ) for which σ|Ui = σi holds for all i.
This generalizes in the following way.
Definition. For a topological space X, a presheaf on X is a contravariant functor F :
TopX → C, where C is any category. For each inclusion of open sets V ,→ U , the induced
morphism FU V : F(U ) → F(V ) is called restriction, written FU V (σ) = σ|V .
Definition. Let C be a set category. A presheaf F : TopX → C is called a sheaf on X
provided it satisfies the following ‘descent conditions’:
(1) For any open covering U = Ui , if there exist sections σ, σ 0 ∈ F(U ) such that σ|Ui =
S
σ 0 |Ui for all i, then σ = σ 0 .
S
(2) For any open covering U = Ui admitting sections σi ∈ F(Ui ) such that σi |Ui ∩Uj =
σj |Ui ∩Uj for all i, j, then there exists a unique section σ ∈ F(U ) such that σ|Ui = σi
for all i.
Notice that the uniqueness in (2) is guaranteed by (1) so it is not strictly necessary to
include it in the definition.
Definition. A morphism of sheaves is simply a natural transformation of functors F →
G. Likewise, an isomorphism of sheaves is a natural isomorphism of the underlying functors.
Example 13.1.2. For a (pre)sheaf F on X and an open set U ⊆ X, the functor F|U : V 7→
F(V ) is a (pre)sheaf. The operation F 7→ F|U is called restriction of a sheaf.

282
13.1. Sheaves and Sections Chapter 13. Sheaf Theory

Remark. The descent conditions in the definition of a sheaf are stated for set categories,
S category C by saying that for every open set
but they can be reinterpreted in an arbitrary
U ∈ TopX and every open covering U = Ui , the object F(U ) is a limit of the following
diagram (also called an equalizer):
Y Y
F(U ) F(Ui ) F(Ui ∩ Uj )
i i,j

Example 13.1.3. The classic example of a sheaf on a topological space X assigns to each
open U ⊆ X the set F(U ) of all continuous (or differentiable, smooth, holomorphic, etc.)
functions on U , with restriction
S maps given by restriction of functions. To see that this is
a sheaf, suppose U = Ui is an open cover in X and f, g : U → Y are continuous maps
such that f |Ui = g|Ui . This means f (x) = g(x) for all x ∈ Ui . Since the Ui cover U , this
means f (x) = g(x) for all x ∈ U and hence f = g. If instead we have continuous functions
fi : Ui → Y for each i with fi |Ui ∩Uj = fj |Ui ∩Uj , then fi (x) = fj (x) for all x ∈ Ui ∩ Uj . For
x ∈ U , define f : U → Y by f (x) = fi (x) if x ∈ Ui . Then the previous statement implies f
is well-defined and continuous on U . Hence f ∈ F(U ) and f |Ui = fi for all i.

Example 13.1.4. Let F(U ) be the set of bounded functions U → R. Then F : U 7→ F(U )
is a presheaf on X, with the usual restriction maps, but F is not a sheaf. For example,
1 4

when X = R, consider the open cover of X given by Un = n − 3 , n + 3 . Then fn (x) = x
defines a sequence of bounded functions on each Un which agree on intersections, but the
only possible lift of {fn } to X is f (x) = x, an unbounded function. Hence axiom (2) fails.

Example 13.1.5. Let S be a set and define a functor F : TopX → Set by F(U ) = S for all
U ⊆ X, with the restriction maps given by σ|V = σ ∈ S for any inclusion V ,→ U . Then F
is a presheaf, called the constant presheaf on S, but if X has two disjoint open sets, then F
is not a sheaf.

Example 13.1.6. The right notion of a constant sheaf is constructed as follows. Let S be a
set considered with the discrete topology. For all open sets U ⊆ X, let FS (U ) = C(U, S) be
the set of all continuous (i.e. locally constant) functions U → S. Then FS defines a sheaf
on X. Notice that if U is a connected open set, then FS (U ) ∼= S, hence the name constant
sheaf.

Example 13.1.7. For a continuous map p : E → X, the assignment U 7→ Γp (U ) defines


a sheaf on X by Lemma 13.1.1. Conversely, every sheaf F on X determines a topological
space EF , called the éspace étale (or étale space) of F, together with a map pF : EF → X,
called the étale cover associated to F. This is constructed below.

Definition. For a sheaf F on X and a point x ∈ X, the stalk of F at x is the direct limit

Fx := lim F(U )
−→

taken over all open sets U ⊆ X containing x.

Example 13.1.8. For the constant sheaf FS , the stalks are all the same: FS,x = S.

283
13.1. Sheaves and Sections Chapter 13. Sheaf Theory

Example 13.1.9. Suppose X is a T1 topological space and A is an abelian group. The


skyscraper sheaf at x ∈ X with coefficients in A is the presheaf x∗ A defined by
(
A, x ∈ U
x∗ A(U ) =
0, x 6∈ U.
S
To see that it is a sheaf, suppose U = Ui is an open cover and s, t ∈ x∗ A(U ) such that
s|Ui = t|Ui for all Ui . If x 6∈ U , then x∗ A(U ) = 0 so s = t is immediate. If x ∈ U , then
x∗ A(U ) = A and x ∈ Ui for some i, but s|Ui = s ∈ A and t|Ui = t ∈ A, so s = t. On the
other hand, if si ∈ x∗ A(Ui ) such that si |Ui ∩Uj = sj |Ui ∩Uj for all Ui ∩ Uj 6= ∅, then there are
two cases again. If x 6∈ U , there is a trivial lift of the si to s ∈ x∗ A(U ) = 0. If x ∈ U , x
lies in some Ui so we may take s = si ∈ x∗ A(Ui ) = A = x∗ A(U ). If x also lies in Uj , then
Ui ∩ Uj 6= ∅ so si |Ui ∩Uj = sj |Ui ∩Uj and in particular si = sj . Therefore s is well-defined, so
x∗ A is a sheaf.
For a point y ∈ X, the stalk (x∗ A)y is A if y = x and 0 otherwise. Therefore we can
think of a skyscraper sheaf as a sort of constant sheaf concentrated at the point x. When X
is not T1 and x is not a closed point of X, then the stalks are a little ‘fuzzier’: for y ∈ X,
(
A, y ∈ {x}
(x∗ A)y =
0, y 6∈ {x}.

Example 13.1.10. Let X = C be the space of complex numbers and for each open U ⊂ C,
let O(U ) denote the set of all holomorphic functions U → C. Then O is a sheaf of rings
with stalks given by Oz = C[t](t−z) , the ring of Taylor series at z with positive radius of
convergence.

Example 13.1.11. For an affine algebraic variety V ⊆ An , the structure presheaf OV (see
Section 12.3) is a sheaf. The proof is similar to the argument in Example 13.1.3. In fact, we
already proved axiom (2) in Lemma 12.3.1. The stalk of the structure sheaf OV at a point
x ∈ V is
OV,x = K[V ]mx ,
the localization of K[V ] at the maximal ideal mx = {f ∈ K[V ] | f (x) = 0}. Equivalently,
OV,x consists of all rational functions f ∈ K(V ) which are defined at x, meaning on some
neighborhood U of x, f = hgxx for some gx , hx ∈ OV (U ) with hx (x) 6= 0.

Proposition 13.1.12. Let p : E → X be a local homeomorphism. Then

(1) The functor Γp (−) : TopX → Set, U 7→ Γp (U ), is a sheaf on X with restriction maps
Γp (U ) → Γp (V ) for V ⊆ U given by restriction of sections s 7→ s|V .

(2) Sets of the form s(U ), where U ⊆ X is open and s ∈ Γp (U ), form a basis for the
topology on E.

(3) The stalk of Γp (−) at x ∈ X is the fibre Ex .

284
13.1. Sheaves and Sections Chapter 13. Sheaf Theory

Proof. (1) follows from Lemma 13.1.1 and (2) is routine. Here’s a proof of (3):
For each neighborhood U ⊆ X of x, define a map

νU : Γp (U ) −→ Ex
s 7−→ s(x).

This is compatible with restriction, i.e. if V ⊆ U is a smaller neighborhood of x, the diagram


Γp (U )
νU

Ex

νV
Γp (V )

commutes. Therefore the νU induce a morphism ν : Γp (−)x = lim Γp (U ) → Ex . For e ∈ Ex ,


−→
take a neighborhood Ve ⊆ E such that p|Ve : Ve → π(Ve ) = U is a homeomorphism. Then
the map s = p|−1
Ve is a section of p over U and s(x) = s(p(e)) = e, so ν is onto. Finally, if
s and t are sections of p over some neighborhoods U1 and U2 of x, respectively, such that

s(x) = t(x) in Ex , then we may choose neighborhoods V1 , V2 ⊆ E such that p|V1 : V1 − → U1
and p|V2 : V2 → U2 are homeomorphisms. Set V = V1 ∩ V2 and U = π(V ). Then p is a
homeomorphism on V , so p ◦ s|U = p ◦ t|U implies s|U = t|U so s = t in Γp (−)x . Thus ν is a
bijection.
S
Definition. The éspace étale of a sheaf F on X is the space EF = x∈X Fx with the
topology induced by the open sets

VU,f := {(x, f |Fx ) | x ∈ U, f ∈ F(U )}

where U runs over all open sets in X. The étale cover pF : EF → X is naturally defined
as (x, f |Fx ) 7→ x.

Example 13.1.13. For the sheaf of holomorphic functions O on C, the open sets VU,f are
given by

VU,f = {(z0 , p(z)) | p(z) is the Taylor series expansion of f (z) at z0 }.

Theorem 13.1.14. There is an equivalence of categories between maps p : E → X and


sheaves F on X given by (p : E → X) 7→ Γp and F 7→ (pF : EF → X).

Proof. We have seen, courtesy of Proposition 13.1.12, that Γp is a sheaf on X. On the other
hand, it follows from the definition of the topology on EF that the map pF : EF → X is
continuous. Therefore it remains to show that the assignments p 7→ Γp and F 7→ pF define
an equivalence of categories. That is, we must show that for a sheaf F on X, there is an
isomorphism ΓpF ∼ = F, and for a map p : E → X, there is a homeomorphism EΓp ∼ = E
making the diagram

285
13.1. Sheaves and Sections Chapter 13. Sheaf Theory


EΓp E

commute. First let F be a sheaf, let U ⊆ X be open, take σ ∈ F(U ) and define s ∈ ΓpF (U )
by s(x) = (x, σ|x ). Consider s−1 (VU,f ) ⊆ U for some f ∈ F(U ). Then

x ∈ s−1 (VU,f ) ⇐⇒ s(x) = f (x)


⇐⇒ s|W = f |W for some neighborhood W of x.

For such an x we have W ⊆ s−1 (VU,f ), so the morphism F → ΓpF sending σ 7→ s is well-
defined and continuous. SupposeSs0 ∈ ΓpF (U ) such that s(x) = s0 (x) for all x ∈ U 0 . Then
there exists an open cover U = Ui with s|Ui = s0 |Ui for all Ui , so by the sheaf axioms,
s = s0 . Therefore F → ΓpF is one-to-one.
On the other hand, suppose s : U → EΓp is a section of pF . Then

s−1 (VU 0 ,s0 ) = {x0 ∈ U 0 | s(x0 ) = s0 (u0 )}.

On the open set s−1 (VU 0 ,s0 ) ⊆ U , s comes from s0 uniquely. Therefore if U 0 is a neighborhood
of x, then s−1 (VUS 0
0 ,s0 ), where s (x) = s(x), is a neighborhood of x. This means there is an

open cover U = Ui with σi ∈ F(Ui ) such that σi (x) = s(x) for all x ∈ U . Moreover,
since F → ΓpF is one-to-one, we have σi |Ui ∩Uj = σj |Ui ∩Uj for all i, j. Therefore by the sheaf
axioms, there is a unique section σ ∈ F(U ) with σ|Ui = σi for all i, which then satisfies
σ(x) = s(x) for all x ∈ U . Thus F → ΓpF is onto, hence an isomorphism.
Now let p : E → X be a continuous map. Define E → EΓp as follows. For y ∈ E, set
x = p(y) ∈ X. Then for all open sets U ⊆ X containing X, y ∈ p−1 (U ), so y in fact lies
in the stalk Γp,x = lim Γp (U ). Thus the assignment E → EΓp , y 7→ y is well-defined. It is
−→
clearly one-to-one and onto, and continuity follows easily from the definition of the topology
on EΓp .
Definition. A sheaf F on X is locally constant if for all x ∈ X, there is a neighborhood
U ⊆ X of x such that F|U is isomorphic to a constant sheaf on U .
Example 13.1.15. If p : E → X is a topological covering, then Γp is a locally constant
sheaf on X. Indeed, for each x ∈ X, choose a connected, locally trivial neighborhood U
of X such that p−1 (U ) ∼= U × p−1 (x). Then for a local section σ : U → E, p maps σ(U )
homeomorphically onto U , so σ(U ) must be a connected component of p−1 (U ). It follows
that Γp (U ) ∼
= p−1 (x) and therefore Γp |U is isomorphic to the constant sheaf Fp−1 (x) .
Conversely, for every locally constant sheaf F, the étale cover EF → X is a topological
cover. To see this, note that for each locally constant neighborhood U ⊆ X, there is a fixed
discrete set S such that Fx = S for all x ∈ U . Thus p−1 (U ) ∼= U × S, so EF is a cover of X.
Thus we have proven:
Corollary 13.1.16. There is an equivalence of categories between covering spaces of X and
locally constant sheaves on X.

286
13.1. Sheaves and Sections Chapter 13. Sheaf Theory

Remark. For a presheaf F on X, there is still a notion of éspace étale pF : EF → X, but


in general the natural maps F(U ) → ΓpF (U ) need not be isomorphisms.

Theorem 13.1.17 (Sheafification). For every presheaf (of sets) F on X, there exists a
presheaf F sh together with a morphism of presheaves θ : F → F sh such that for every x ∈ X,
θx : Fx → Fxsh is a bijection. Moreover, for any morphism ϕ : F → G where G is a sheaf,
there is a unique morphism of sheaves ψ : F sh → G such that the diagram

θ
F F sh

ϕ ψ

commutes.

Corollary 13.1.18. If F is a sheaf, then F sh = F .

Proof. This will from Proposition 13.2.8(2) and the fact that Fx → Fxsh is a bijection for all
x ∈ X.
π
Proof of 13.1.17. Let E = EF →− X be the étale space of F and define F = Γπ (−), the sheaf
of sections of π. Then the morphism θ : F → F sh is defined by

θU : F (U ) −→ F sh (U )
s 7−→ (s̃ : U → E, x 7→ s|x ).

Suppose V ⊆ U is an inclusion of open sets. It is immediate that the diagram

θU
F (U ) F sh (U )

F (V ) F sh (V )
θV

commutes, so the θU indeed give a morphism θ : F → F sh . On stalks, we have that


Fxsh = Γ(−, π)x = Ex = Fx by Proposition 13.1.12(3), and it is clear that θx : Fx → Fx is
just the identity morphism sx 7→ sx .
To prove the universal property, note that Corollary 13.1.18 already follows from the
paragraph above, so for any sheaf G, θ : G → Gsh is the identity on G. If ϕ : F → G is a
morphism (of presheaves), then by Theorem 13.1.14 there is a continuous map Eϕ : EF → EG
and thus a morphism ψ : F sh = ΓπF → ΓπG = Gsh = G. Now for all open U ⊆ X and
sections s ∈ F (U ),
ψU ◦ θU (s) = ψU (s̃) = Eϕ ◦ se = ϕU (s)
by construction, so the diagram

287
13.1. Sheaves and Sections Chapter 13. Sheaf Theory

θU
F (U ) F sh (U )

ϕU ψU

G(U )

commutes, and once again this is compatible with restriction of sections along V ⊆ U . The
fact that such a ψ is unique follows from the bijection Fx → Fxsh and Proposition 13.2.8.

Definition. For a presheaf F on X, we call F sh the sheafification of F , or the sheaf


associated to F .

Remark. Abstractly, we have an adjoint pair of functors

sh
Presh(C) Sh(C)
forget

which means sheafification is right exact and the forgetful functor is left exact. In a more
general setting, such as when we consider the categories of presheaves and sheaves of abelian
groups, rings, algebras, etc., we can take this as our definition of the sheafification functor:
it is the left adjoint to the (left exact) forgetful functor ShX ,→ PreshX .
Alternatively, we can equip an associated sheaf F sh with the structure of a sheaf of
π
abelian groups (or rings, algebras, etc.) as follows. Let E → − X be the étale cover for the
presheaf F and consider the fibre product

E ×X E := {(e1 , e2 ) ∈ E × E | π(e1 ) = π(e2 )}.

Then E ×X E → E, (e1 , e2 ) 7→ e1 + e2 is a well-defined, continuous map of topological spaces


since the fibres of E (i.e. the stalks of F ) are abelian groups. The induced map

F sh (U ) × F sh (U ) = Γπ (U ) × Γπ (U ) ∼
= Γπ×X π (U ) −→ Γπ (U ) = F sh (U )

gives an additive structure on each F sh (U ), making F sh into a sheaf of abelian groups.

288
13.2. The Category of Sheaves Chapter 13. Sheaf Theory

13.2 The Category of Sheaves


In this section, we study the category ShX of sheaves on a topological space X. We start
by introducing several constructions in the category PreshX of presheaves on X. Many
categorical constructions in PreshX can be constructed locally, i.e. over open sets U ⊆ X.
For example:
Definition. For two presheaves F and G on X, the presheaf product `
F × G is defined
` × G)(U ) = F(U
by (F ` ) × G(U ). Likewise, the presheaf coproduct F G is defined by
(F G)(U ) = F(U ) G(U ).
Definition. Let ϕ : F → G be a morphism of presheaves.
ˆ The presheaf image im ϕ is defined by (im ϕ)(U ) = im(F(U ) → G(U )).

ˆ The presheaf kernel ker ϕ is defined by (ker ϕ)(U ) = ker(F(U ) → G(U )).

ˆ The presheaf cokernel coker ϕ is defined by (coker ϕ)(U ) = coker(F(U ) → G(U )).
Lemma 13.2.1. If ϕ : F → G is a morphism of sheaves, then ker ϕ is a sheaf.
S
Proof. Suppose U = i Ui is an open cover in X and s, t ∈ (ker ϕ)(U ) such that s|Ui = t|Ui
for all Ui . Then since (ker ϕ)(U ) ⊆ F(U ), the first sheaf axiom for F implies directly that
s = t. Now suppose there are sections si ∈ (ker ϕ)(Ui ) such that si |Ui ∩Uj = sj |Ui ∩Uj for all
overlapping Ui , Uj . Since (ker ϕ)(Ui ) ⊆ F(Ui ), the second sheaf axiom for F implies there
exists a section s ∈ F(U ) such that s|Ui = si for all i. We must show s ∈ (ker ϕ)(U ) ⊆ F(U ).
Note that ϕU (s)|Ui = ϕUi (s|Ui ) = ϕUi (si ) = 0 for each Ui , so by the first sheaf axiom for G,
we must have ϕU (s) = 0.
Theorem 13.2.2. If C is an abelian category, then the category PreshX (C) of presheaves
TopX → C is an abelian category.
Example 13.2.3. For a morphism ϕ of sheaves, coker ϕ is not a sheaf in general. For exam-
ple, take X = C and let F = O be the sheaf of holomorphic functions from Example 13.1.10.
If O× denotes the sheaf of nonvanishing holomorphic functions on C, then there is a mor-
phism of sheaves ϕ : O → O× given locally by ϕU : O(U ) → O(U )× , f 7→ e2πif . In fact,
one can see that ϕ is an isomorphism on stalks (using the complex logarithm). However, on
U = C× , the map O(C× ) → O(C× )× is not surjective, since f (z) = z is not in the image.
Thus on the level of stalks we have exact sequences
ψx ϕx
Zx −→ Ox −→ Ox×

where Z denotes the constant sheaf, but the cokernel of the map ψ : Z → O is not O× , but
rather (O/Z)sh .
Definition. The sheaf cokernel of a morphism of sheaves ϕ : F → G is the sheafification
of the presheaf cokernel:

U 7−→ (coker ψ)(U ) := (G(U )/ϕ(F(U )))sh .

289
13.2. The Category of Sheaves Chapter 13. Sheaf Theory

Theorem 13.2.4. For any space X, ShX is an abelian category.

Let ϕ : F → G be a morphism of presheaves on X. Then for any x ∈ X, since the stalks


Fx and Gx are defined via direct limits, there is an induced morphism ϕx : Fx → Gx .

Lemma 13.2.5. Let F be a sheaf on X and U ⊆ X an open set with sections s, t ∈ F(U ).
Then s =Qt if and only if for all x ∈ U , s|x = t|x in Fx . In other words, the morphism
F(U ) → x∈U Fx is injective.

S neighborhood Ux ⊆ U of x such that s|Ux = t|Ux . Since


Proof. If s|x = t|x , then there is some
this holds for all x ∈ U and U = x∈U Ux , we have s|Ux = s|Ux for all Ux , but by the sheaf
axioms this implies s = t.

Corollary 13.2.6. Let F be a presheaf, G a sheaf and ϕ, ψ : F → G two morphisms of


presheaves. Then ϕ = ψ if and only if ϕx = ψx : Fx → Gx for all x ∈ X.

Proof. The ( =⇒ ) implication is clear, so suppose ϕx = ψx for all x. Let U ⊆ X be an open


set and take s ∈ F(U ). Then the elements ϕU (s), ψU (s) ∈ G(U ) restrict on stalks to:

ϕU (s)|x = ϕx (s|x ) since ϕ is a morphism


= ψx (s|x ) by hypothesis
= ψU (s)|x since ψ is a morphism.

Therefore by Lemma 13.2.5, ϕU (s) = ψU (s). Since this holds for all U , we get ϕ = ψ.

Definition. Let F be a sheaf of abelian groups on X and s ∈ Γ(X, F) a global section. Then
the support of s is the set

supp(s) = {x ∈ X | s|x 6= 0 in Fx }.

Lemma 13.2.7. For any s ∈ Γ(X, F), supp(s) is a closed set in X.

Proposition 13.2.8. Let ϕ : F → G be a morphism of presheaves, with F a sheaf. Then

(1) The morphisms ϕx : Fx → Gx are injective for all x ∈ X if and only if the morphisms
ϕU : F(U ) → G(U ) are injective for all open sets U ⊆ X.

(2) The morphisms ϕx : Fx → Gx are bijective for all x ∈ X if and only if the morphisms
ϕU : F(U ) → G(U ) are bijective for all open sets U ⊆ X.

Proof. (1) Assume ϕx are injective for all x ∈ X and suppose s, t ∈ F(U ) are sections such
that ϕU (s) = ϕU (t) in G(U ). Then for all x ∈ X,

ϕx (s|x ) = ϕU (s)|x = ϕU (t)|x = ϕx (t|x )

since ϕ is a morphism, so injectivity implies s|x = t|x . Finally, Lemma 13.2.5 shows that
s = t in F(U ).
Conversely, assume that each ϕU is injective. For a fixed x ∈ X, take sx , tx ∈ Fx such
that ϕx (sx ) = ϕx (tx ) in Gx . Then there is a neighborhood U of x such that sx = s|x and

290
13.2. The Category of Sheaves Chapter 13. Sheaf Theory

tx = t|x for some s, t ∈ F(U ). Since Gx = lim G(U ) over all neighborhoods U of x, the
−→
condition that ϕx (sx ) = ϕx (tx ) in Gx means there is some smaller neighborhood V ⊆ U
containing x on which ϕU (s)|V = ϕU (t)|V . Since ϕ is a morphism, this is equivalent to
ϕV (s|V ) = ϕV (t|V ), but by injectivity of ϕV , we get s|V = t|V . Passing to the direct limit,
we have sx = tx in Fx . Hence ϕx is injective.
(2) The ( ⇒= ) implication is trivial. For ( =⇒ ), suppose each ϕx is bijective. Fix
U ⊆ X and t ∈ G(U ); we must find s ∈ F(U ) so that ϕU (s) = t. For any x ∈ U ,
consider the image tx of t in the stalk Gx . By hypothesis, there is some sx ∈ Fx such that
ϕx (sx ) = tx . Then there is a neighborhood Ux ⊆ U of x on which sx = sUx |x for some
sUx ∈ F(Ux ). By construction, ϕUx (sUx ) = t in Gx , so there is a neighborhood Vx ⊆ Ux of x
with ϕVx (sUx |Vx ) = ϕUxS(sUx )|Vx = tVx . Set sVx = sUx |Vx . Letting x range over the points of
U , we get a cover U = x∈U Vx and moreover, if x, y ∈ U , then

ϕVx ∩Vy (sVx |Vx ∩Vy ) = ϕVx (sVx )|Vx ∩Vy since ϕ is a morphism
= t|Vx ∩Vy
= ϕVy (sVy )|Vx ∩Vy
= ϕVx ∩Vy (sVy |Vx ∩Vy ) for the same reason.

We are assuming each ϕx is bijective, so in particular (1) implies each ϕVx ∩Vy is injective, and
thus sVx |Vx ∩Vy = sVy |Vx ∩Vy . Now the second sheaf axiom guarantees that there is a section
s ∈ F(U ) such that s|Vx = sVx for all x ∈ U . Finally, observe that

ϕU (s)|Vx = ϕVx (s|Vx ) = ϕVx (sVx ) = t|Vx

for all x ∈ U . Thus the first sheaf axiom implies ϕU (s) = t.

Remark. Notice that the proof of surjectivity in (2) crucially relies on the fact that the ϕx
and ϕU are already injective. It is not true in general that the ϕx are all surjective if and
only if the ϕU are all surjective.

Proposition 13.2.9. For any x ∈ X, the stalk functor PreshX → Ab, F 7→ Fx , is exact.
ϕ ψ
Proof. Suppose F 0 − →F − → F 00 is an exact sequence of presheaves on X. We must show the
ϕx ψx
induced sequence of stalks Fx0 −→ Fx −→ Fx00 is exact. First, since ψ ◦ ϕ = 0 globally, on
each stalk we get ψx ◦ ϕx = 0 so im ϕx ⊆ ker ψx . On the other hand if sx ∈ ker ψx , then for
some neighborhood U of x, there is a section s ∈ F(U ) such that s|x = sx and ψU (s)|x =
ψx (sx ) = 0. Thus on some smaller neighborhood V ⊆ U of x, ψV (s|V ) = ψU (s)|V = 0, so
s|V ∈ ker ψV . By exactness, this means s|V = ϕV (s0 ) for some s0 ∈ F 0 (V ). Then we have
sx = ϕV (s0 )x = ϕx (s0x ) for s0x = s0 |x in Fx . Hence ker ψx ⊆ im ϕx so the sequence is exact.
This gives us a good definition for what it means for a sequence of sheaves to be exact.

Definition. A sequence of sheaves F 0 → F → F 00 is exact if for all x ∈ X, the sequence of


stalks Fx0 → Fx → Fx00 is exact.

291
13.2. The Category of Sheaves Chapter 13. Sheaf Theory

ϕ ψ
Theorem 13.2.10. Let 0 → F 0 − →F − → F 00 → 0 be a short exact sequence of sheaves on
X. Then for every open set U ⊆ X, the sequence

0 → F 0 (U ) → F(U ) → F 00 (U )

is exact. That is, the forgetful functor ShX → PreshX is left exact.
ϕx ψx
Proof. By definition, 0 → Fx0 −→ Fx −→ Fx00 → 0 is exact for all x ∈ X, so Propo-
sition 13.2.8(1) implies 0 → F 0 (U ) → F(U ) is exact for all open sets U ⊆ X. Set
K(U ) = ker(F(U ) → F 00 (U )). Then by Lemma 13.2.1, K : U 7→ K(U ) is a sheaf and it is
easy to see that the stalks are Kx = ker(Fx → Fx00 ). Now we have (ψ ◦ ϕ)x = ψx ◦ ϕx = 0 for
all x ∈ X by exactness of the sequence of stalks, so (ψ ◦ϕ)U = ψU ◦ϕU = 0 for all U , and thus
im ϕU ⊆ K(U ). Finally, ϕ can be written as a morphism F → K, but on the level of stalks

this is a bijection ϕx : Fx −
→ Kx = im ϕx , so by Proposition 13.2.8(2), ϕU : F(U ) → K(U )
is also bijective. Hence F ∼ = im ϕ = K so the sequence 0 → F 0 (U ) → F(U ) → F 00 (U ) is
exact.
In general, if F → F 00 is surjective, it need not be true that F(U ) → F 00 (U ) is surjective.
By definition, surjectivity of sheaves means surjectivity of stalks, so an element s00x ∈ Fx00 is
the image of some sx ∈ Fx but this sx need only be defined on some neighborhood of x, not
necessarily on arbitrary U . Example 13.2.3 gives a counterexample.

Definition. A sheaf F on X is called flasque (or flabby) if for every open inclusion V ⊆ U ,
the restriction map F(U ) → F(V ) is surjective.
ϕ ψ
Theorem 13.2.11. If 0 → F 0 − → F 00 → 0 is a short exact sequence of sheaves with F 0
→F −
flasque, then the sequence

0 → F 0 (U ) → F(U ) → F 00 (U ) → 0

is exact for any open U ⊆ X. If in addition F is flasque, then so is F 00 .

Proof. By Theorem 13.2.10, it’s enough to show ψU : F(U ) → F 00 (U ) is surjective. Further,


after restriction of sheaves, it’s enough to consider the case U = X. Let t ∈ F 00 (X). Then
for any x ∈ X, there is a neighborhood Ux ⊆ X of x and a section sUx ∈ F(Ux ) such that
ψUx (sUx ) = t|Ux , by surjectivity of stalks. Consider the set of pairs

C = {(U, s) | U ⊆ X, s ∈ F(U ), ψU (s) = t|U }

of open sets which admit an extension of t. Then C is a partially ordered set via (U1 , s1 ) 
(U2 , s2S) if and only if U1 ⊆ U2 and s2 |U1 = s1 . For S
a linearly ordered subset (Uα , sα ) in C, the
∗ ∗
pair ( Uα , s ) is an upper bound, where s ∈ F ( Uα ) is defined using the sheaf conditions
on F. Thus by Zorn’s Lemma, C contains a maximal element, say (U, s). Assume x ∈ X rU .
Then there is a neighborhood V of x and a section sV,0 ∈ F(U ) with ψV (sV,0 ) = t|V by
surjectivity on stalks. Note that

ψU ∩V (s|U ∩V − sV,0 |U ∩V ) = ψU ∩V (s|U ∩V ) − ψU ∩V (sV,0 |U ∩V ) = t|U ∩V − t|U ∩V = 0

292
13.2. The Category of Sheaves Chapter 13. Sheaf Theory

so s|U ∩V −sV,0 |U ∩V ∈ ψU ∩V = im ϕU ∩V by exactness at F(U ∩V ). That is, s|U ∩V −sV,0 |U ∩V =


ϕU ∩V (s00 ) for some s00 ∈ F 0 (U ∩ V ). Now since F 0 is flasque, s00 = s0 |U ∩V for some s0 ∈ F 0 (V ).
Set sV = sV,0 + ϕV (s0 ). By construction, s|U ∩V = sV |U ∩V so by the sheaf axioms on F over
U ∪ V , there is a section sU ∪V ∈ F(U ∪ V ) with sU ∪V |U = s and sU ∪V |V = sV . This shows
that (U, s) ≺ (U ∪ V, sU ∪V ), contradicting the fact that (U, s) is a maximal element in C.
Hence X r U = ∅, or U = X, and s ∈ F(X) satisfies ψX (s) = t by definition.
For the last statement, consider the following diagram with exact rows:
0 F 0 (U ) F(U ) F 00 (U ) 0

0 F 0 (V ) F(V ) F 00 (V ) 0

If F is flasque, then the middle column is surjective, so it follows that the right column is
also surjective. Hence F 00 is also flasque.

Proposition 13.2.12. Let ϕ : F → G be a morphism of presheaves on X. Then

(1) ϕ is surjective over all open sets U ⊆ X if and only if ϕ is an epimorphism in the
category PreshX .

(2) If F and G are sheaves, then ϕ is surjective on stalks if and only if ϕ is an epimorphism
in the category ShX .

Proof. (1) Let coker ϕ be the presheaf cokernel of ϕ. Then ϕ is an epimorphism if and only
if coker ϕ = 0. For any open U ⊆ X, (coker ϕ)(U ) is the cokernel of ϕU : F (U ) → G(U )
in the category of abelian groups, so it is 0 if and only if ϕU is surjective. On the level of
sheaves, coker ϕ = 0 if and only if (coker ϕ)(U ) = 0 for all U , thus ϕ is surjective on open
sets if and only if it’s an epimorphism.
(2) First suppose ϕ is surjective on stalks. If

ϕ f
F G H
g

is a diagram of sheaves such that f ◦ ϕ = g ◦ ϕ, then for all x ∈ X, we have

fx ◦ ϕx = (f ◦ ϕ)x = (g ◦ ϕ)x = gx ◦ ϕx

but since ϕx is surjective, this implies fx = gx . Applying Corollary 13.2.6 gives us f = g, so


ϕ is an epimorphism.
Conversely, if ϕ is an epimorphism, let C = coker ϕ be the presheaf cokernel of ϕ. If
p : G → G/ϕ(F ) ∼ = C is the natural projection of presheaves, consider the diagram
ϕ p
F G C
0

293
13.2. The Category of Sheaves Chapter 13. Sheaf Theory

Then p ◦ ϕ = 0 = 0 ◦ ϕ so by assumption, p = 0. That is, C = 0, which means Cx = 0 for all


x ∈ X. But Cx = coker(ϕx : Fx → Gx ), so Cx = 0 implies ϕx is surjective for all x ∈ X.
Definition. Let X be a topological space and ShX (C) the category of sheaves Topop
X → C with
values in a category C. Then the section functor is defined by
Γ : ShX (C) −→ C
F 7−→ Γ(X, F) = F(X).
For any open set U ⊆ X, the sections over U are simply the restricted sections Γ(U, F) =
F(U ).
Proposition 13.2.13. For any X, the section functor Γ is left exact.
Proof. Follows from Theorem 13.2.10.
Example 13.2.14. When C = Ab is the category of abelian groups, the section functor is
representable:
Γ(X, F) ∼= HomShX (ZX , F)
where ZX is the constant sheaf on X with stalks Z. Recall that Hom(ZX , −) is left exact;
this gives an easier proof of Proposition 13.2.13 in this case.
Theorem 13.2.15. Suppose C has products and enough injectives. Then ShX (C) has enough
injectives.
Proof. Given x ∈ X and an object A ∈ A, let x∗ A denote the skyscraper sheaf at x with
coefficients A: (
A, x ∈ U
x∗ A(U ) =
0, x 6∈ U.
Then Hom(G, x∗ A) consists of collections of morphisms G(U ) → A in C for each U containing
x which are compatible with the restrictions G(U ) → G(V ) for V ,→ U . By the universal
property of stalks (they are limits), this is the same as a map Gx → A. Thus
HomShX (C) (G, x∗ A) ∼
= HomC (Gx , A).
If A is an injective of C, this shows that x∗ A is an injective of ShX (C). Moreover, since Hom
commutes with products, we have
!
x∗ Ax ∼ HomSh (C) (G, x∗ Ax ) ∼
Y Y Y
HomSh (C) G,
X = X = HomC (Gx , Ax )
x∈X x∈X x∈X

for any family of objects (Ax )x∈X in C. Since C has enough injectives, for each x ∈ X we
may choose a monic ϕx : Gx ,→ Ax with Ax injective, so that the induced morphism
Y
G −→ x∗ Ax
x∈X
Q
is monic and x∈X x∗ Ax is injective.
Thus for a category C with products and enough injectives, the right derived functors of
the section functor Γ : ShX (C) → C are defined.

294
13.3. Sheaf Cohomology Chapter 13. Sheaf Theory

13.3 Sheaf Cohomology


Theorem 13.2.15 allows us to construct the right derived functors of the global section func-
tor.
Definition. Let X be a topological space and C a category with products and enough in-
jectives. Then the sheaf cohomology of X is the sequence of right derived functors of
Γ : ShX (C) → C, denoted
H i (X, F) := Ri Γ(X, F).
One of the main phenomena to exploit is that ShX has many acyclic objects which are
not injective. This is important because of the following result.
Theorem 13.3.1. Let F be a sheaf on X and suppose L• is an acyclic resolution of F , i.e.
a complex of sheaves for which each Ln is acyclic. Then for all i ≥ 0,

H i (X, F ) ∼
= H i (Γ(X, L• )).

Proof. The statement is trivial for i = 0. Let


f0 f1
0 → F → L0 −
→ L1 −
→ L2 → · · ·

be our acyclic resolution of F . Set K 0 = F and for each i ≥ 1, K i = im fi−1 ⊆ Li . Then


there are short exact sequences of sheaves
e gi
0 → Ki −
→i
Li −
→ K i+1 → 0

where ei is the inclusion of a subsheaf and the gi satisfy ei+1 ◦ gi = fi . Since Γ is left exact
(Theorem 13.2.10), each sequence
Γe Γgi
0 → Γ(X, K i ) −−→
i
Γ(X, Li ) −−→ Γ(X, K i+1 )

is exact. Note that Γei identifies Γ(X, K i ) with ker Γgi , which is equal to ker Γfi since Γei is
injective. Likewise, im Γgi−1 = im Γfi−1 . Therefore Γ(X, K i )/ im Γgi−1 ∼ = ker Γfi / im Γfi−1 =
H i (Γ(X, L• )). On the other hand, there is a long exact sequence
Γgi
0 → Γ(X, K i ) → Γ(X, L• ) −−→ Γ(X, K i+1 ) → H 1 (X, K i ) → H 1 (X, Li ) = 0

with a zero on the end because Li is acyclic. Hence

H 1 (X, K i ) ∼
= Γ(X, K i+1 )/ im Γgi = H i+1 (Γ(X, L• )).

The same argument shows that there is an isomorphism H j−1 (X, K i+1 ) ∼
= H j (X, K i ) for each
j ≥ 2. In particular, H (X, K ) = H (X, K ) = H (K, F ) so we obtain H i+1 (X, F ) ∼
1 i ∼ i+1 0 i+1
=
H i+1 (Γ(X, L• )) for all i ≥ 0.
In fact, Theorem 13.3.1 holds in much greater generality, namely for the derived functors
of a left exact functor out of an abelian category with enough injectives, but we will only
need it for sheaf cohomology computations in these notes.

295
13.3. Sheaf Cohomology Chapter 13. Sheaf Theory

Example 13.3.2. Let X be a smooth, paracompact manifold and define the sheaf of rings
U 7−→ OX (U ) = {f : U → R | f is smooth}.
The pair (X, OX ) is called a ringed space. A sheaf of abelian groups F on X is called an
OX -module if for all U ⊆ X, F(U ) is an OX (U )-module such that the module structure
OX (U ) × F(U ) → F(U ) is compatible with restrictions F(U ) → F(V ). Examples of OX -
modules include the sheaves ΩkX (differential k-forms), VF X (smooth vector fields), Γ(E)
(sections of any vector bundle E), etc. (See Section 14.4 for more details.) Using this
language, we have our first of several ‘vanishing theorems’:
Theorem 13.3.3. For a smooth, paracompact manifold X and any OX -module F,
H i (X, F) = 0
for all i > 0.

Q We induct on i > 0. For an OX -module F, the proof of Theorem 13.2.15 shows that
Proof.
I = x∈X Fx is an injective sheaf and F ,→ I is injective. It’s easy to check that I is also
an OX -module. Thus we obtain a short exact sequence of OX -modules
0 → F → I → I/F → 0.
By definition H 1 (X, F) = coker(Γ(X, I) → Γ(X, I/F)). Take a section s ∈ Γ(X, I/F).S By
the sheaf axioms and paracompactness, we may choose a locally finite open cover Ui = X
and sections ti ∈ I(Ui ) such that ti 7→ s|Ui . Fix a partition of unity fi subordinate to the
Ui , i.e. functions fi ∈ OX (X) such that for each Ui ,
X
fi |XrUi = 0 and fj ≡ 1.
j

Then fi ti extendsPto a global section ofPI vanishing away from Ui , so by the sheaf axioms,
the section t = f t
j j j maps to s = j fj s. This shows that Γ(X, I) → Γ(X, I/F) is
1
surjective, so H (X, F) = 0. The inductive step follows from analyzing the same short exact
sequence, which gives H k (X, F) ∼= H k−1 (X, I/F).
In fact, the proof of Theorem 13.3.3 shows:
Corollary 13.3.4. If F is a flasque sheaf on a smooth, paracompact manifold X, then
H i (X, F) = 0 for all i > 0.
This shows that flasque sheaves are acyclic, so we may compute H i (X, G) using a flasque
resolution of G. (We will generalize this in Section 16.2.)
In general, cohomology is hard to compute from the derived functors definition. However,
we may begin by recovering ordinary (singular) cohomology with coefficients:
Theorem 13.3.5. Let X be a topological space, A an abelian group and AX the constant
sheaf on X with stalks A. Then
H i (X, AX ) ∼
= H i (X; A)
where H i (X; A) denotes the ordinary cohomology of X with coefficients in A.

296
13.3. Sheaf Cohomology Chapter 13. Sheaf Theory

Proof. For each open set U ⊆ X, let Cn (U ; A) denote the A-module spanned by singular
simplices on U and let ∂ : Cn (U ; A) → Cn−1 (U ; A) denote the simplicial boundary map.
Dualizing gives us the modules of singular cosimplices C n (U ; A) := HomA (Cn (U ; A), A) with
coboundary map given by the adjoint, d = ∂ ∗ : C n (U ; A) → C n+1 (U ; A). The assignment
U 7→ C n (U ; A) defines a presheaf C n on X for each n ≥ 0. In fact by construction, for any
inclusion V ,→ U , the restriction map C n (U ; A) → C n (V ; A) is surjective, so C n is a flasque
presheaf on X. Note that C n satisfies the gluing axiom of a sheaf, but it is not separated.
However, sheafifying a flasque presheaf which satisfies the gluing axiom yields a flasque sheaf
in general, so setting C n := (C n )sh , we get a complex of flasque sheaves C • . The differential
d : C n → C n+1 is induced by the coboundary maps dU : C n (U ; A) → C n+1 (U ; A). In fact,
one can show that C n is isomorphic to the quotient sheaf C n /C0n , where

C0n (U ) = {ϕ ∈ C n (U ) | there exists an open cover U = {Ui } of U such that ϕ|Ui ≡ 0}.

To verify that C • is an exact complex, it suffices to compute exactness on stalks. On


a neighborhood of each point, however, this sequence becomes C • (U ) which is exact by
construction in algebraic topology. Note that ker(d : C 0 → C 1 ) is precisely the constant sheaf
AX , so C • is a flasque resolution of X. Since flasque sheaves are acyclic (more generally, we
will show this in Theorem 16.2.3), Theorem 13.3.1 implies that

H i (X, AX ) ∼
= H i (Γ(X, C • ))

for all i ≥ 0.
Finally, singular cohomology is computed by the complex of abelian groups C • (X) but
we have a short exact sequence of complexes of presheaves

0 → C0• → C • → C • → 0

so to prove H i (X; A) = H i (Γ(X, C • )) ∼


= H i (Γ(X, C • )) = H i (X, AX ), it remains to show that
C0• is acyclic. Fix an open cover U of X and let CnU (U ; A) denote the submodule of Cn (U ; A)
spanned by simplices lying in an element of U. Set CUn (U ; A) = HomA (CnU (U ; A), A) and let
CU• be the resulting complex of presheaves on X. There is a natural morphism of presheaves
C n → CUn which is surjective; call its kernel CUn,0 . Then we can see that

C0n = lim CUn,0


−→

where the limit is taken over all open covers U of X. This identification is compatible with
differentials, so we get C0• = lim CU• ,0 . By a result in algebraic topology, there exists a cover
−→
U for which CU• ,0 is acyclic, which finishes the proof.
Now the strategy for computing general sheaf cohomology is clear: we may take an in-
jective resolution of AX and compute the cohomology of the resulting complex. Of course,
different ways of resolving a constant sheaf can give different ways of computing cohomol-
ogy, and these observations will produce beautiful comparisons between different flavors of
cohomology theory.

297
13.3. Sheaf Cohomology Chapter 13. Sheaf Theory

Let ΩkX be the sheaf of differential k-forms on a smooth


L∞ manifold X. This defines a sheaf
• k
of graded rings (differential graded algebras) ΩX = k=0 ΩX . Explicitly, there is a complex
d d d
OX = Ω0X →
− Ω1X →
− Ω2X →
− ···
where d is the exterior derivative.
Definition. The sheaf cohomology of Ω•X is called the de Rham cohomology of X, written
i
HdR (X) := H i (Γ(X, Ω•X )).
Theorem 13.3.6. For a smooth, paracompact manifold X, the complex of sheaves
i d d d
− Ω0X →
0 → RX → − Ω1X →
− Ω2X →
− ···
is acyclic, where RX denotes the constant sheaf on X with stalks R, and i is the inclusion of
constant functions.
Proof. It is enough to compute this on stalks, where we have for each x ∈ X a coordinate
chart U ∼= Rn . By Poincaré’s Lemma in algebraic topology, HdR i
(Rn ) = 0 for all i > 0, so
i
HdR (U ) = 0 for all i > 0 as well. Taking direct limits commutes with cohomology, so
H i (X; Ω•X,x ) = lim H i (U ; Ω•X (U )) = lim HdR
i
(U ) = 0.
−→ −→

Corollary 13.3.7. For a smooth manifold X, there is an isomorphism


H • (X) ∼= H • (X, RX )
dR

where RX is the constant sheaf.


Proof. Theorem 13.3.6 shows that Ω0X → Ω1X → Ω2X → · · · is a deleted acyclic resolution of
RX , so this follows from the definition of sheaf cohomology as the right derived functors of
Γ(X, RX ).
Corollary 13.3.8 (De Rham’s Theorem). For a smooth manifold X, there is an isomor-
phism between de Rham cohomology and singular cohomology:
H • (X) ∼= H • (X; R).
dR

Proof. Apply Corollary 13.3.7 and Theorem 13.3.5.


i
Corollary 13.3.9. For a smooth, compact manifold X, HdR (X) is finite dimensional for all
i ≥ 0.
Proof. Endow X with a Riemannian metric. Then every point in X admits a neighborhood
which is geodesically convex, meaning any two points in the neighborhood lie on a geodesic
which is contained in the neighborhood. Since X is compact, we may choose a finite cover
U = {U1 , . . . , Um } consisting of these geodesically convex open sets. Note that for any
i1 , . . . , ir ∈ {1, . . . , m}, the intersection Ui1 ,...,ir = Ui1 ∩ · · · ∩ Uir is also geodesically convex.
A geodesically convex set has vanishing de Rham cohomology, so de Rham’s theorem says
that H i (Ui1 ,...,ir , RX |Ui1 ,...,ir ) ∼ i
= HdR (Ui1 ,...,ir ) = 0 for all i > 0. Therefore RX is acyclic for the
i
cover U, and we will prove Leray’s theorem (16.2.6) which says that H i (X, RX ) = Ȟ (U, RX ).
Now observe that each C i (U, RX ) in the Čech complex is a finite product of one-dimensional
i
vector spaces, hence itself finite dimensional, so Ȟ (U, RX ) is finite dimensional.

298
13.4. Čech Cohomology Chapter 13. Sheaf Theory

13.4 Čech Cohomology


For a general space X, the sheaf cohomology groups may be quite difficult to compute – how
does one produce a flasque or even injective resolution resolution in general? Fortunately,
there is another construction of sheaf cohomology which, though cumbersome to define, is
much more amenable to computation.
We motivate the definition of Čech cohomology with a classical problem originally studied
by Mittag-Leffler. Let X be a Riemann surface, that is, a 2-manifold admitting a complex
structure with holomorphic transition functions. We do not assume X to be closed. Suppose
E is a closed, discrete subset of X and for each point a ∈ E, we are given a function
za : Ua → C on some neighborhood Ua ⊆ X of a such that za (a) = 0. Consider a function

pa (za ) = α−m za−m + α−m+1 za−m+1 + . . . + α−1 za−1 , αj ∈ C.

That is, pa is a polynomial in za−1 , or a Laurent polynomial on Ua centered at a. The Mittag-


Leffler problem is to find a meromorphic function f : X → C such that f is holomorphic on
X r E and for all a ∈ E, the function f − pa has a removable discontinuity at a. Then pa
will be the principal part of f on Ua .
In short, the Mittag-Leffler problem asks us to extend meromorphic functions defined on
open sets in X to a meromorphic function on the whole Riemann surface. The following
restatement will be useful for later generalization. Let U = {Ui }i∈I be an open cover of X
and suppose {fi : Ui → C} is a collection of meromorphic functions such that each fi is
either holomorphic on Ui or has a single pole ai ∈ Ui , with ai 6∈ Uj for any j 6= i. The
Mittag-Leffler problem is then to find a meromorphic function f : X → C such that for each
i ∈ I, f |Ui − fi is holomorphic.
First notice that if the fi agree on all overlaps Ui ∩ Uj , then the sheaf condition on M
(the sheaf of meromorphic functions on X) guarantees that there is a global meromorphic
function f ∈ M(X) so that f |Ui − fi = 0 for all i, a much stronger conclusion than the
Mittag-Leffler problem asks for. Thus in some sense, the problem is to study how far one
can lift local meromorphic functions that do not glue together on overlaps. In any case,
one way to find such a function f : X → C in the previous paragraph is to find a family
{hi : Ui → C} of holomorphic functions on each Ui such that (fi + hi )|Ui ∩Uj = (fj + hj )|Ui ∩Uj
for all i, j. This can be rewritten as

fi |Ui ∩Uj − fj |Ui ∩Uj = hj |Ui ∩Uj − hi |Ui ∩Uj .

Set tij = fi |Ui ∩Uj − fj |Ui ∩Uj . Then if the above equation is satisfied, we have tij ∈ O(Ui ∩ Uj ),
where O is the sheaf of holomorphic functions on X and

tjk − tik + tij = 0.

Thus we want to find holomorphic functions hi ∈ O(Ui ) satisfying:

(1) tij = hj − hi on Ui ∩ Uj (the boundary condition); and

(2) tjk − tik + tij = 0 on Ui ∩ Uj ∩ Uk (the cycle condition).

299
13.4. Čech Cohomology Chapter 13. Sheaf Theory

This generalizes to any space X with open cover U = {Ui }i∈I .


Q
Definition. Let F be a presheaf on X. A family of sections (tij ) ∈ i,j F (Ui ∩ Uj ) is called
a Čech 1-cocycle if for all i, j, k, tjk − tik + tij = 0 on Ui ∩ Uj ∩ Uk . The group of Čech
1
1-cocycles is written Ž (U, F ).
Q
Definition. AQfamily (tij ) ∈ i,j F (Ui ∩ Uj ) is called a Čech 1-coboundary if there is a
family (hi ) ∈ i F (Ui ) such that tij = hj − hi on each Ui ∩ Uj . This forms a subgroup of
1 1
Ž (U, F ) which is denoted B̌ (U, F ).

Definition. The first Čech cohomology group of the cover U with coefficients in F is
the quotient group
1 1 1
Ȟ (U, F ) := Ž (U, F )/B̌ (U, F ).

Now to solve the Mittag-Leffler problem, it’s enough to show that for a Riemann surface
1
X with cover U, Ȟ (U, O) = 0 where O is the sheaf of holomorphic functions on X. Note
that these Čech cohomology groups depend on the cover U. To package together all of
the information about holomorphic functions on covers of X, we introduce the notion of a
refinement of covers.

Definition. Let U = {Ui }i∈I and V = {Vj }j∈J be open covers of X. We say V is a refine-
ment of U if there exists a function τ : J → I such that for all j ∈ J, Vj ⊆ Uτ (j) .

Notice that such a refinement τ induces a map


1 1
τ 1 : Ž (U, F ) −→ Ž (V, F )

(tik )i,k∈I tτ (j)τ (`) |Vj ∩V` j,`∈J .

1 1
Lemma 13.4.1. If τ is a refinement from U to V, then τ 1 (B̌ (U, F )) ⊆ B̌ (V, F ).
1 1
Thus there is an induced map on Čech cohomology groups, Ȟ (U, F ) → Ȟ (V, F ) and
one can check that this does not depend on the choice of map τ : J → I.

Definition. The first Čech cohomology of X with coefficients in F is the direct limit
1 1
Ȟ (X, F ) = lim Ȟ (U, F )
−→

over all open covers U of X, ordered by refinement.

Lemma 13.4.2. For any sheaf F on a Riemann surface X, the maps


1 1
Ȟ (U, F ) −→ Ȟ (X, F )

are injective for every open cover U of X.

300
13.4. Čech Cohomology Chapter 13. Sheaf Theory

1
Proof. Let U = {Ui }i∈I be an open cover of X. If g = (gik )i,k∈I ∈ Ȟ (U, F ) becomes 0 in
1
Ȟ (X, F ), this means there is some open cover V = {Vj }j∈J which is a refinement of U,
1
say by τ : J → I, such that g 0 = (gj`
0
)j,` = gτ (j)τ (`) |Vj ∩V` j,` lies in B̌ (V, F ). Since g 0 is a

0
Q
1-coboundary, there is a family (hj ) ∈ j F (Vj ) such that gj` = hj − h` on Vj ∩ V` . For a
fixed index i ∈ I, observe that
0
gj` = gτ (j)τ (`)
= gτ (j)i + giτ (`) by the cocycle condition
= giτ (`) − giτ (j) by the cocycle condition again
= hj − h` on Vj ∩ V` .

This can be rewritten as

giτ (j) + hj = giτ (`) + h` on Ui ∩ Vj ∩ V` .

Consider the cover {Ui ∩ Vj }j∈J of Ui . By the sheaf axioms for F , there exists a section
ti ∈ F (Ui ) such that for each Vj , ti |Ui ∩Vj = (giτ (j) + hj )|Ui ∩Vj . Now let i, k ∈ I be arbitrary
and fix some j ∈ J. Then

gik = giτ (j) + gτ (j)k = giτ (j) − gkτ (j) by the cocycle condition
= (ti − hj ) − (tk − hj ) by the above
= ti − tk on Ui ∩ Uk ∩ Vj .

Since j was arbitrary and {Ui ∩ Uk ∩ Vj }j∈J covers Ui ∩ Uk , the sheaf axioms imply gik − ti − tk
1 1
on Ui ∩ Uk . Hence gik ∈ B̌ (U, F ) so g = 0 in Ȟ (U, F ).

Theorem 13.4.3 (Mittag-Leffler). For a non-compact Riemann surface X with sheaf of


1
holomorphic functions O, Ȟ (U, O) = 0 for every open cover U of X.
1
Proof. We will show that Ȟ (X, O) = 0 in this situation, so Lemma 13.4.2 implies that
1
Ȟ (U, O) = 0.

Definition. Let U = {Ui }i∈I be an open cover of X and let F be a sheaf of abelian groups
on X. The Čech complex of F with respect to U is the cochain complex C • (U, F ) defined
by Y
C p (U, F ) = F (Ui0 ∩ · · · ∩ Uip )
i0 ,...,ip ∈I

with differential

d : C p (U, F ) −→ C p+1 (U, F )


p+1
!
X
α 7−→ (−1)k αi0 ,...,ik−1 ,ik+1 ,...,ip+1 |Ui0 ,...,ip+1
k=0

where Ui0 ,...,ip+1 = Ui0 ∩ · · · ∩ Uip+1 .

301
13.4. Čech Cohomology Chapter 13. Sheaf Theory

Lemma 13.4.4. d2 = 0; that is, C • (U, F ) is a cochain complex.


Definition. The pth Čech cohomology with respect to an open cover U of a space
X with coefficients in a sheaf F is the pth cohomology of the Čech complex:
p
Ȟ (U, F) := H p (C • (U, F)).
0
Lemma 13.4.5. For any sheaf F and cover U of X, Ȟ (U, F) = H 0 (X, F) = Γ(X, F).
0
Proof. By definition, Ȟ (U, F) = ker(d : C 0 (U, F) → C 1 (U, F)). For α = (αi ) ∈ C 0 (U, F),
we have dα = (αi − αj )i,j which is zero if and only if αi = αj on Ui ∩ Uj for all i, j. Thus
ker d = Γ(X, F).
Suppose U 0 is a refinement of U, that is, U 0 = {Uj0 }j∈J is a cover of X and there is
a function λ : J → I such that for all j ∈ J, Uj0 ⊆ Uλ(j) . Then there is a chain map
C • (U 0 , F) → C • (U, F) given by
C p (U 0 , F) −→ C p (U, F), (αj0 ,...,jp ) 7−→ (αλ(j0 ),...,λ(jp ) |Uj0 ∩···∩Ujp ).

This in turn induces maps on Čech cohomology:


p p
Ȟ (U 0 , F) −→ Ȟ (U, F)
for all p.
Definition. The pth Čech cohomology of a space X with coefficients in a sheaf F is the
direct limit
p p
Ȟ (X, F ) := lim Ȟ (U, F )
−→
taken over all covers U ordered with respect to refinement.
Theorem 13.4.6. If X is paracompact, then there is a natural isomorphism
p
Ȟ (X, F) ∼
= H p (X, F)
for all p and F.
Fix a space X, an open cover U = {Ui }i∈I and a sheaf F on X. We construct a complex
of sheaves of abelian groups C p (U, F ) on X as follows. For an open set V ⊆ X, define
Y
C p (U, F ) : V 7−→ Γ(V, C p (U, F )) := F (V ∩ Ui0 ,...,ip )
i0 ,...,ip ∈I

where as usual Ui0 ,...,ip = Ui0 ∩ · · · ∩ Uip .


Proposition 13.4.7. If F is a sheaf on X, then C p (U, F ) is a sheaf for all p ≥ 0.
We give C • (U, F ) the structure of a complex of sheaves by letting
dV : Γ(V, C p (U, F )) −→ Γ(V, C p+1 (U, F ))
be the map induced from the Čech differential d : C p (U, F ) → C p+1 (U, F ) restricted to each
V ∩ Ui0 ,...,ip .

302
13.4. Čech Cohomology Chapter 13. Sheaf Theory

Proposition 13.4.8. For an open cover U and a sheaf F , there is an exact sequence of
sheaves
0 → F → C 0 (U, F ) → C 1 (U, F ) → · · ·
Proof. The sheaf axioms on F imply F → C 0 (U, F ) is injective. To prove exactness at
C p (U, F ), we need to check that the sequence of stalks
α β
C p−1 (U, F )x −
→ C p (U, F )x →
− C p+1 (U, F )x

is exact for all x ∈ X. Since U covers X, we know x ∈ Uj for some Uj . Take fx ∈ C p (U, F )x
and let V be a neighborhood of x and f ∈ Γ(V, C p (U, F )) with f |x = fx . We may assume
V ⊆ Uj . Then for each p ≥ 1,

θp : C p (U, F )x −→ C p−1 (U, F )x


fx 7−→ fj,i0 ,...,ip−1 |x

is well-defined and independent of the choice of V and Uj . Now with the same fx and f as
above, we have

θp+1 (df )x = (df )j,i0 ,...,ip |x


p
!
X
= fi0 ,...,ip − (−1)k fj,i0 ,...,ik−1 ,ik+1 ,...,ip


k=0 x
p
= fx − d(θ fx ).

Hence θp+1 d + dθp is the identity on C p (U, F )x , which shows the identity map on this stalk
is chain homotopic to the zero map and thus ker β = im α, proving exactness.
Definition. For a sheaf F on X, the exact sequence

0 → F → C 0 (U, F ) → C 1 (U, F ) → · · ·

is called the Čech resolution of F with respect to the cover U.


Theorem 13.4.9. For any sheaf F on X and any open cover U of X, there is a natural
map
p
Ȟ (U, F ) −→ H p (X, F )
for all p ≥ 0 which is compatible with refinement. In particular, there is a natural map
p
Ȟ (X, F ) −→ H p (X, F )

for all p ≥ 0.
Proof. Applying the global sections functor to the Čech resolution of F yields
0 1
0 → F (X) → Č (U, F ) → Č (U, F ) → · · ·

Clearly the cohomology groups of this sequence are precisely the Čech cohomology groups
p
Ȟ (U, F ). Therefore there is a chain map

303
13.4. Čech Cohomology Chapter 13. Sheaf Theory

0 F C 0 (U, F ) C 1 (U, F ) ···

id

0 F E0 E1 ···
p
for any injective resolution E • of F . This induces the maps on cohomology, Ȟ (U, F ) →
H p (X, F ), and one can check that they commute with the maps coming from refinements of
open covers. The second statement follows.
The sheafification functor PreshX → ShX (Theorem 13.1.17) can alternatively be con-
structed using the Čech resolution. In fact, this construction works in a much more general
context known as Grothendieck topology. Let F be a presheaf on a space X, U = {Ui } an
open cover of X and
0
n Y o
Ȟ (U, F ) = f = (fi ) ∈ F (Ui ) : fi |Ui ∩Uj = fj |Ui ∩Uj for all i, j .

For an open set U ⊆ X, let


0 0
Ȟ (U, F |U ) = lim Ȟ (U 0 , F |U )
−→

0
where the limit is over all open covers U 0 of U . Define the set F # (U ) = Ȟ (U, F |U ). For
0
an inclusion of open sets V ,→ U in X, there is a natural restriction map Ȟ (U, F |U ) →
0
Ȟ (V, F |V ) coming from restriction of an open cover. This makes U 7→ F # (U ) into a presheaf
on X. Moreover, Proposition 13.4.8 shows that F → F # is injective.
Definition. We call a presheaf
S F on X a separated presheaf if it satisfies the first sheaf
axiom; that is, if U = Ui is an open cover and there are sections s, t ∈ F (U ) such that
s|Ui = t|Ui for all Ui , then s = t.
Theorem 13.4.10. Let F be a presheaf on X. Then
(1) F # is a separated presheaf on X.

(2) If F is a separated presheaf, then F # is a sheaf. In particular, F ## is always a sheaf


on X.

(3) F ## is a sheafification of F .
Proof. (1) Suppose s, t ∈ F # (U ) and there is an open cover {Vj } of U such that s|Vj = t|Vj
for all Vj . We must show s = t in F # (U ). Since they are elements of a direct limit, s and t
have representatives
0
sU = (si ), tU = (ti ) ∈ Ȟ (U, F |U )
on some open cover U of U – a priori there may be two different covers supporting s and t,
but after a common refinement of the covers we may assume this takes place on just a single
cover. By hypothesis, there is an open cover Wj = {Wjk }k of each Vj that is a refinement of
{Vj ∩ Ui }i such that si |Wjk = ti |Wjk for all k. Then W = {Wjk }j,k is an open cover of U on

304
13.4. Čech Cohomology Chapter 13. Sheaf Theory

0
which si |Wjk = ti |Wjk on every element. Therefore the images of sU and tU in lim Ȟ (U, F |U )
−→
are equal, so s = t as required.
(2) Now assume F itself is separated; we must show that F # satisfies the second sheaf
axiom. As an intermediate step, we claim that for any open cover U = {Ui }i∈I of U , the
map
0 0
Ȟ (U, F |U ) −→ Ȟ (U, F |U )
0
is injective (this is conditioned on F being separated). Notice that if s, t ∈ Ȟ (U, F |U )
0
become equal in the direct limit Ȟ (U, F |U ), then there exists a cover V = {Vj }j∈J of U
such that s|Vj = t|Vj for all j ∈ J. We may assume V is a refinement of U, say by a map
τ : J → I. Then sτ (j) |Vj = tτ (j) |Vj for all j ∈ J. Since s and t are Čech 0-cocycles, we have

si |Ui ∩Uτ (j) = sτ (j) |Ui ∩Uτ (j) and ti |Ui ∩Uτ (j) = tτ (j) |Ui ∩Uτ (j)

for all i, j. Since Vj ⊆ Uτ (j) for all j, this means

si |Ui ∩Vj = ti |Ui ∩Vj

for all i, j. Fixing i ∈ I, the collection {Ui ∩Vj }j is an open cover of Ui and si |Ui ∩Vj = ti |Ui ∩Vj ,
0
so by the separated condition, si = ti on Ui . This implies s = t in Ȟ (U, F |U ), proving the
claim.
Now take si ∈ F # (Ui ) for some cover U = {Ui }i∈I of U such that si |Ui ∩Uj = sj |Ui ∩Uj for
all i, j ∈ I. We must construct a section s ∈ F # (U ) restricting to si on each Ui . Each si is
0
represented by some sVi = (sik ) ∈ Ȟ (Vi , F |Ui ) for an open cover Vi = {Vik }k of Ui . Then
V = {Vik }i,k is an open cover of U . Fixing i, j ∈ I, consider the cover W = {Vik ∩ Vj` }k,` of
Ui ∩ Uj . The sections
0
s̃i = (sik |Vik ∩Vj` ), s̃j = (sj` |Vik ∩Vj` ) ∈ Ȟ (W, F |Ui ∩Uj )
0
become equal in Ȟ (Ui ∩ Uj , F |Ui ∩Uj ) by assumption, so because F is separated, the previous
0
paragraph shows s̃i = s̃j in Ȟ (W, F |Ui ∩Uj ). Therefore the s̃i give a well-defined element
0
s ∈ Ȟ (U, F |U ) restricting to si on each Ui . Hence F # is a sheaf.
(3) Suppose F → G is a morphism of presheaves, where G is a sheaf. This determines a
chain map
C • (U, F ) −→ C • (U, G)
for any open cover U of X. Hence there is a map on Čech cohomology
p p
Ȟ (U, F ) −→ Ȟ (U, G)

for all p ≥ 0, which is compatible with refinement. In particular, there is a map


0 0
Ȟ (X, F ) −→ Ȟ (X, G).
0 0
Let U ⊆ X be an open set. Then there is a map Ȟ (U, F |U ) → Ȟ (U, G|U ) which fits into
a diagram

305
13.4. Čech Cohomology Chapter 13. Sheaf Theory

F (U ) G(U )


=
0 0
Ȟ (U, F |U ) Ȟ (U, G|U )

The right column is an isomorphism since G is a sheaf. Thus we may defined the dashed
0
arrow Ȟ (U, F |U ) → G(U ) which induces a morphism of presheaves F # → G through which
F → G factors. Repeating the argument shows that F → G further factors through a
morphism F ## → G. Uniqueness follows from a similar argument as the one in the proof
of Theorem 13.1.17.
Changing topics, we next describe some useful applications of sheaf cohomology (using
Čech complexes for computation).
Theorem 13.4.11. For any paracompact space X and sheaf F, there exists an injective
sheaf I on X such that
H n (X, F) ∼
= H n−1 (X, I/F)
for all n.
Let us illustrate this for n = 1. Note that H 1 (X, F) = Ext1ShX (Ab) (ZX , F) where ZX is a
constant sheaf. Explicitly, elements of Ext1 (ZX , F) are isomorphism classes of short exact
sequences of sheaves
0 → F → G → ZX → 0.
Assume X is connected, so that Γ(X, ZX ) ∼ = Z. For any short exact sequence 0 → F →
G → ZX → 0, we have a long exact sequence
δ
− H 1 (X, F) → · · ·
0 → Γ(X, F) → Γ(X, G) → Z →

Set ε = δ(1). Then ε = 0 if and only if there is a splitting Z → Γ(X, G) of δ. There are open
sets {Ui } in X such that Γ(Ui , G) → Γ(Ui , ZX ) is surjective for all i (although this need not
be true for global sections). So on Ui , the short exact sequence

0 → F|Ui → G|Ui → ZUi → 0

splits, i.e. there is some σi ∈ G(Ui ) mapping to 1 ∈ Z. For each i, j, set τij = σi − σj ∈
G(Ui ∩ Uj ). Then τij maps to 1 − 1 = 0 in Γ(Ui ∩ Uj , ZX ), so by exactness τij ∈ F(Ui ∩ Uj ).
Also note that each σi is unique up to adding any bi ∈ F(Ui ), but if σi0 = σi + bi , we get
τij0 = τij + bi − bj . Viewing (τij ) and (τij0 ) as Čech 1-cocycles, the previous statement shows
1
that [(τij )] = [(τij0 )] in Ȟ (X, F). Q
Conversely, given a 1-cocycle (τij ) in F(Ui ∩ Uj ), we show how to reconstruct the
extension 0 → F → G → ZX → 0. We define G by its éspace étale:

G ét = {(x, f, n, Ui ) | x ∈ Ui , f ∈ Fx , n ∈ Z}/ ∼

where (x, f, n, Ui ) ∼ (x, f +nτij , n, Uj ) for every x ∈ Ui ∩Uj . Thus locally, G ét |Ui ∼
= F ét |Ui ×Z.
Over Ui , the section σi : x 7→ (x, 0, 1, Ui ) gives a splitting of δ : Γ(X, G) → Z, since

306
13.4. Čech Cohomology Chapter 13. Sheaf Theory

(x, 0, 1, Ui ) ∼ (x, τij , 1, Uj ) whenever x ∈ Ui ∩ Uj . This defines G as a sheaf of sets on X, but


with a corresponding action of F given locally by F(U ) × (F(U ) × Z) → F(U ) × Z. In other
words, G is a torsor for F on X. In general, torsors are in correspondence with cohomology
classes in H 1 (X, F), as we have seen in this case.

Example 13.4.12. Suppose X is a manifold (possibly with extra structure), G is a topo-


logical group and P → X is a principal G-bundle – this means that over certain open sets
U ⊆ X, there is a G-equivariant isomorphism

=
P |U U ×G

making the diagram commute. The data of a principal bundle is equivalent to specifying
an open cover {Ui } of X and “transition functions” τij : Ui ∩ Uj → G (which preserve the
extra structure on the manifold). Explicitly, if x ∈ Ui ∩ Uj then (x, g) 7→ (x, gτij (x)) is an
isomorphism on (Ui ∩ Uj ) × G as a trivial bundle over Ui ∩ Uj . It follows from this description
that τji = τij−1 and τij τjk = τik for all i, j, k. So (τij ) is a Čech cocycle for some sheaf, but
which one? Let ΓG be the sheaf on X given locally by

ΓG (U ) = {σ : U → U × G | σpσ = idU }.

Then the above description shows that principal G-bundles up to isomorphism are in one-
to-one correspondence with elements of H 1 (X, ΓG ).

Example 13.4.13. Let F be a field and G = GLn (F ). Then principal G-bundles are in
one-to-one correspondence with rank n vector bundles over X, so the isomorphism classes
of rank n vector bundles are determined by classes in H 1 (X, ΓGLn (F ) ). If F = R and n = 1,
so that GL1 (R) = R× , we have a short exact sequence of abelian groups
exp sgn
0 → R −−→ R× −−→ Z/2Z → 0.

By de Rham’s theorem (13.3.8), H 1 (X, ΓR ) = H 2 (X, ΓR ) = 0, so the long exact sequence in


sheaf cohomology implies H 1 (X, ΓR× ) ∼= H 1 (X; Z/2Z). That is, isomorphism classes of line
bundles (rank 1 vector bundles) on X are in one-to-one correspondence with cohomology
classes H 1 (X; Z/2Z). The class ω(L) ∈ H 1 (X; Z/2Z) determined by a line bundle L → X
is called the Stiefel-Whitney class of L.
A similar calculation for F = C and the short exact sequence
exp
− C −−→ C× → 0
0→Z→

shows that complex line bundles up to isomorphism are determined by classes in H 1 (X, ΓC× ) ∼
=
H 2 (X; Z), called Chern classes, denoted c(L).

307
Chapter 14

Scheme Theory

308
14.1. The Spectrum of a Ring Chapter 14. Scheme Theory

14.1 The Spectrum of a Ring


In this section we discuss another generalization of affine varieties in the form of the prime
spectrum Spec(A) for a ring A. This forms the groundwork for scheme theory, which in many
ways is the main language of modern algebraic geometry. Let Spec(A) denote the set of prime
ideals of A and let MaxSpec(A) denote the set of maximal ideals. As we have seen with
Corollary 2.4.3 (the algebraic Nullstellensatz), when A is the coordinate ring of some affine
variety X ⊆ An , the points of An correspond bijectively with elements (ideals) of MaxSpec(A)
via P ↔ mP . This correspondence transfers the Zariski topology of X to A. Further, if
ϕ : X → Y is a morphism of varieties, with A = k[X] and B = k[Y ], then identifications
X ↔ MaxSpec(A) and Y ↔ MaxSpec(B) induce a map ϕ∗ : MaxSpec(A) → MaxSpec(B),
given by ϕ∗ (m) = (ϕ∗ )−1 (m), where ϕ∗ : B → A is the induced comorphism.
We can mimic this in a general setting; however, it is not true in general that if ϕ : A → B
is a ring homomorphism then a maximal ideal in B pulls back to a maximal ideal in A. To
remedy this, note the following fact.
Lemma 14.1.1. Suppose f : A → B is a ring homomorphism and q is a prime ideal of B.
Then f −1 (q) is a prime ideal of A.
Proof. Suppose x, y ∈ A such that xy ∈ f −1 (q) but x 6∈ f −1 (q). Then f (x)f (y) = f (xy) ∈
f (f −1 (q)) = q but f (x) 6∈ q, so since q is prime, f (y) ∈ q. This implies y ∈ f −1 (q) so we
have that f −1 (q) is prime.
This suggests a natural replacement for MaxSpec A, the prime spectrum:

Spec(A) = {p ⊂ A | p is a prime ideal}.

For a subset E ⊂ A, let V (E) be the set of prime ideals of A containing E:

V (E) = {p ∈ Spec(A) | p ⊇ E}.

Lemma 14.1.2. Let E be a subset of A and set a = (E), the ideal generated by E. Then
(i) If a is the ideal generated by E then V (E) = V (a) = V (r(a)).

(ii) V (0) = X, V (1) = ∅.

(iii) If (Ei )i∈I is any family of subsets of A, then


!
[ \
V Ei = V (Ei ).
i∈I i∈I

(iv) V (a ∩ b) = V (ab) = V (a) ∪ V (b) for any ideals a, b of A.


Proof. As a preliminary remark, if E, F ⊆ Spec(A) such that N ⊆ M , then any prime ideal
containing M also contains N . Thus V (M ) ⊆ V (N ).
(i) E ⊆ a so by the preliminary remark, V (E) ⊇ V (a). On the other hand, if p is a
prime ideal containing E then p ⊇ a since by definition a is the smallest ideal containing

309
14.1. The Spectrum of a Ring Chapter 14. Scheme Theory

E. Thus p ∈TV (a) so we have V (E) = V (a). For the second inequality, Corollary 1.1.3 says
that r(a) = {P primes | P ⊇ a}, which directly implies V (a) = V (r(a)).
(ii) Every prime ideal p ⊂ A contains 0, so V (0) = X. On the other hand, if p is a prime
ideal containing 1 then p = A, which is impossible. Therefore V (1) = ∅.
(iii) For each Ei , let ai = (E
S Si ) be the
 ideal generated by Ei , and let a be the ideal generated
by i∈I Ei . Then by (i), V T each i. Clearly ai ⊆ a
i∈I i = V (a) and V (Ei ) = V (ai ) for
E
for each i, soTV (a) ⊆ V (ai ), and since this holds for all i, V S(a) ⊆ i∈I V (ai ). On the other
hand, if p ∈ i∈I V (ai ) then p ⊇ aiT⊇ Ei for each i, so p ⊇ i∈I Ei . By definition of a, this
implies p ⊇ a, so p ∈ V (a). Hence i∈I V (ai ) ⊆ V (a) so we have equality.
(iv) By Lemma 1.1.4, r(a)r(b) ⊆ r(ab) so (i) gives us

V (ab) = V (r(ab)) ⊆ V (r(a)r(b)) ⊆ V (r(a)) ∪ V (r(b)) = V (a) ∪ V (b).

Next, a ⊇ a ∩ b and b ⊇ a ∩ b so by the preliminary remark, V (a) ⊆ V (a ∩ b) and V (b) ⊆


V (a ∩ b). Thus V (a) ∪ V (b) ⊆ V (a ∩ b). Finally, ab ⊆ a ∩ b implies V (a ∩ b) ⊆ V (ab). So
we have V (ab) ⊆ V (a ∪ V (a) ∪ V (b) ⊆ V (a ∩ b) ⊆ V (ab) and hence equality all the way
through.
The above properties show that the sets V (E) for E ⊆ A form the closed sets of a
topology on Spec(A), called the Zariski topology on Spec(A).
Definition. For a ring A, the space Spec(A) equipped with the Zariski topology is called an
affine scheme.
Proposition 14.1.3. Let f : A → B be a ring homomorphism. Then
(a) The induced map f ∗ : Spec(B) → Spec(A), q → f −1 (q), is continuous with respect to
the Zariski topology.

(b) If f is surjective, f ∗ is a closed map.

(c) More generally, if B/f (A) is an integral ring extension, f ∗ is a closed map.
Proof. (a) By Lemma 14.1.2(i), it suffices to show that for any ideal I ⊂ A, (f ∗ )−1 (V (I)) =
V (f (I)). Note that for a prime p ∈ Spec(B),

p ∈ (f ∗ )−1 (V (I)) ⇐⇒ f ∗ (p) = f −1 (p) ∈ V (I)


⇐⇒ f −1 (p) is a prime containing I, by (a)
⇐⇒ p ⊇ f (I)
⇐⇒ p ∈ V (f (I)).

Therefore (f ∗ )−1 (V (I)) = V (f (I)) so closed sets pull back to closed sets. This implies f ∗ is
continuous.
(b) Let J ⊂ B be any ideal. We claim that f ∗ (V (J)) = V (f −1 (J)). By Lemma 14.1.2(i),
this is enough to conclude that f ∗ is closed. Set I = f −1 (J) ⊂ A. Since f is surjective,
the map B/J → A/I is also surjective. Then V (J) = Spec(B/J) and V (I) = Spec(A/I)
so replacing A with A/I and B with B/J, it suffices to show that f ∗ (Spec(B)) = Spec(A).
If p ∈ f ∗ (Spec(B)) then p = f −1 (Q) for some prime ideal q ⊂ B. By Lemma 14.1.1,

310
14.1. The Spectrum of a Ring Chapter 14. Scheme Theory

p ∈ Spec(A). The other inclusion follows from surjectivity of f , so we have f ∗ (Spec(B)) =


Spec(A) and f ∗ is a closed map.
(c) Viewing f as the composition f : A  f (A) ,→ B, (b) shows that the surjection is
always closed so it suffices to consider the case when A ⊆ B is a subring and f : A ,→ B is
the inclusion map. Let J ⊂ B be an ideal, so that V (J) is closed in Spec(B). Set I = J ∩ A.
We claim that f ∗ (V (J)) = V (I). Consider the inclusion A/I ,→ B/J. By Lemma 2.3.1(a),
B/J is integral over A/I, so replacing B with B/J and A with A/I, it is enough to show
that f ∗ (Spec(B)) is closed. If p ∈ f ∗ (Spec(B)) then p = f −1 (Q) = Q ∩ A for some prime
ideal Q ⊂ B, but Q ∩ A is prime in A, so this means p ∈ Spec(A). On the other hand, if
p ∈ Spec(A) then by Lemma 2.3.4, there is a prime ideal Q ⊂ B such that Q ∩ A = p. In
particular p = f −1 (Q) so p ∈ f ∗ (Spec(B)). This shows f ∗ (Spec(B)) = Spec(A) so f is a
closed map.
The operator V takes sets (and in particular, ideals) of A to closed sets in Spec(A) and
reverses inclusions. We define
T an operator going in the opposite direction: for a subset
Y ⊆ Spec(A), let J(Y ) = p∈Y p. Then J(Y ) is an ideal of A and clearly the operator J is
inclusion-reversing.

Proposition 14.1.4. Let A be a ring, I ⊂ A an ideal and Y ⊆ Spec(A) a subset. Then

(a) V (J(Y )) = Y , the Zariski closure of Y in Spec(A).

(b) J(V (I)) = r(I).

(c) V (I) is irreducible if and only if r(I) is a prime ideal.

Proof. (a) On one hand,

V (J(Y )) = {p ∈ Spec(A) | p ⊇ J(Y )}


( )
\
= p ∈ Spec(A) | p ⊇ q
q∈Y

⊇ {p ∈ Spec(A) | p ∈ Y } = Y

so V (J(Y )) ⊇ Y . Since V (J(Y )) is a closed set in Spec(A), we have V (J(Y )) ⊇ Y . Con-


versely, we may write Y = V (E) for some subset E ⊂ A. If E 6⊂ J(Y ) then Y = V (E) 6⊃
V (J(Y )) ⊇ Y , a contradiction. Hence E ⊆ J(Y ), implying Y = V (E) ⊇ V (J(Y )). This
gives us the desired equality. T T
(b) By definition, J(V (I)) = p∈V (I) p = p⊇I p where the latter intersection is over all
prime ideals p containing I. Thus by Corollary 1.1.3, J(V (I)) = r(I).
(c) ( =⇒ ) Suppose V (I) is irreducible and let f, g ∈ A with f g ∈ J(V (I)) = r(I). Then
by (a), we have V (I) ⊆ V (J(V (I))) ⊆ V (f g) but by Lemma 14.1.2(iv), V (f g) = V ((f )(g)) =
V ((f )) ∪ V ((g)) = V (f ) ∪ V (g), so in particular we can write V (I) = (V (f ) ∩ V (I)) ∪
(V (g) ∩ V (I)). Since V (I) is assumed to be irreducible, it follows that V (I) = V (f ) ∩ V (I)
or V (I) = V (g) ∩ V (I). In other words, V (I) ⊆ V (f ) or V (I) ⊆ V (g), so f ∈ r(I) or
g ∈ r(I). Hence r(I) is prime.

311
14.1. The Spectrum of a Ring Chapter 14. Scheme Theory

( ⇒= ) Conversely, assume r(I) is prime and V (I) = Y1 ∪Y2 for closed sets Y1 , Y2 ⊆ V (I).
Then there exist closed sets X1 , X2 ⊆ Spec(A) with Y1 = X1 ∩ V (I) and Y2 = X2 ∩ V (I).
For some ideals a1 , a2 ⊂ A, we have X1 = V (a1 ) and X2 = V (a2 ), so V (I) ⊆ X1 ∪ X2 =
V (a1 ) ∪ V (a2 ) = V (a1 a2 ) by Lemma 14.1.2(iv). Applying J, we get r(I) ⊇ J(V (a1 a2 )) =
r(a1 a2 ) ⊇ a1 a2 . Since r(I) is prime, a1 ⊆ r(I) or a2 ⊆ r(I); without loss of generality, assume
a1 ⊆ r(I). Then by (a), V (I) ⊆ V (r(I)) ⊆ V (a1 ) = X1 . This implies Y1 = X1 ∩V (I) = V (I),
so V (I) is irreducible.
Proposition 14.1.5. Let A be a noetherian ring and X = Spec(A) its prime spectrum.
Then
(a) X is noetherian.

(b) dim X = dim A.


Proof. (a) Suppose X1 ⊇ X2 ⊇ X3 ⊇ · · · is a descending chain of Zariski-closed subsets of
X. We may assume each Xj = V (aj ) for ideals aj ⊂ A. Applying J, we get an ascending
chain of radical ideals in A:

r(a1 ) ⊆ r(a2 ) ⊆ r(a3 ) ⊆ · · ·

Since A is noetherian, there is an n ∈ N such that r(an ) = r(an+1 ) = · · · . Then since each
Xj was taken to be closed, Proposition 14.1.4(a) implies V (J(Xj )) = V (r(aj )) = Xj and so
we get

V (r(a1 )) ⊇ V (r(a2 )) ⊇ V (r(a3 )) ⊇ · · · ⊇ V (r(an )) = V (r(an+1 )) = · · ·


=⇒ X1 ⊇ X2 ⊇ X3 ⊇ · · · ⊇ Xn = Xn+1 = · · ·

Thus every descending chain of closed subsets stabilizes, meaning X is noetherian.


(b) Suppose Y0 ) Y1 ) · · · ) Y` is a chain of closed, irreducible subsets of X. Write
Yk = V (Ik ) for ideals Ik ⊂ A. By Proposition 14.1.4(c), each J(Yk ) = r(Ik ) is prime, so we
get a proper chain
r(I0 ) ( r(I1 ) ( · · · ( r(I` )
of prime ideals in A. (The inclusions must be strict, else r(Ij ) = r(Ij+1 ) for some 0 ≤ j ≤
` − 1, in which case we would have Yj = V (Ij ) = V (r(Ij )) = V (r(Ij+1 )) = V (Ij+1 ) = Yj+1
by Lemma 14.1.2(i).) By definition, this means ` ≤ dim A, so taking the supremum over
all such chains of closed, irreducible subsets in X, we obtain dim X ≤ dim A. Conversely,
if p0 ( p1 ( · · · ( pm is a strict chain of prime ideals in A, Xk = V (pk ) is irreducible for
each 0 ≤ k ≤ m, by Proposition 14.1.4(c). This determines a chain of closed, irreducible
subsets in X: X0 ) X1 ) · · · ) Xm . (Again, if Xj = Xj+1 for any 0 ≤ j ≤ m − 1,
then pj = r(pj ) = J(V (pj )) = J(Xj ) = J(Xj+1 ) = J(V (pj+1 )) = r(pj+1 ) = pj+1 by
Proposition 14.1.4(b), a contradiction.) Therefore m ≤ dim X, so taking the supremum over
all such chains of primes in A gives us dim A ≤ dim X.
Example 14.1.6. The converse of Proposition 14.1.5(a) is not true: if Spec(A) is a noethe-
rian, A need not be noetherian. For an example, consider the polynomial ring k[tn | n ∈ N]
in countably many variables. We have an ideal I = (t2n | n ∈ N) in this ring, so one can take

312
14.1. The Spectrum of a Ring Chapter 14. Scheme Theory

the quotient to get A = k[tn | n ∈ N]/I. The ideal J = (tn | n ∈ N) strictly contains I,
so J = J/I is an ideal of A which is clearly not finitely generated. So A is not noetherian.
However, k[tn | n ∈ N]/J ∼= k is a field, so J is maximal and hence J is maximal in A. On
the other hand, since every generator of J is nilpotent, every element of J is nilpotent, so
J must also be a minimal prime ideal. This implies J is the unique prime ideal of A, so
Spec(A) is finite and in particular a noetherian space.
Next, for any prime ideal p ⊂ A, let Ap denote the localization at p. For any open set
U ⊆ Spec A, we define
( )
a f
O(U ) = s : U → Ap s(p) ∈ Ap , ∃ q ∈ V ⊆ U such that s(q) = for all q ∈ V, f, g ∈ A .

p∈U
g

Theorem 14.1.7. (Spec A, O) is a ringed space. Moreover,


(1) For any p ∈ Spec A, Op ∼
= Ap as rings.
(2) Γ(Spec A, O) ∼
= A as rings.
(3) For any f ∈ A, define the open set D(f ) = {p ∈ Spec A | f ∈ 6 p}. Then the D(f )

form a basis for the topology on Spec A and O(D(f )) = Af as rings.
Proof. (1) Define ϕp : Op → Ap by s 7→ sU (p) where sU ∈ OU for some neighborhood U of p
and sU (p) is the image of p in Ap . Take fa ∈ Ap . Then a ∈ A and f ∈ A r p, so fa = sU (p)
where U = D(f ) and s = fa itself. On the other hand, suppose U is a open set containing p
and s, t ∈ O(U ) such that ϕp (s) = ϕp (t) in Ap . We may assume that on U , s = fa and t = gb
for a, b ∈ A and f, g ∈ A r p. Then there is some u ∈ A r p such that u(ga − f b) = 0 in
A. This shows s = fa = gb = t in Aq for every q ∈ D(f ) ∩ D(g) ∩ D(u), so s = t on an open
neighborhood of p and hence s = t in Op .
(2) follows from (3) by setting f = 1, in which case A1 = A.
(3) Define π : Af → O(D(f )) by fan 7→ s where s : p 7→ fan ∈ Ap for p ∈ O(D(f )).
   
Suppose π fan = π fbm for some a, b ∈ A and m, n ≥ 0. Then fan = fbm in Ap for every
p ∈ D(f ). For such a p, we have u(f m a − f n b) = 0 in A. Setting I = Ann(f m a − f n b) ⊆ A,
we see that I 6⊂ p. Since this holds for any p ∈ D(f ), this implies V (I) ∩ D(f ) = ∅, so
f ∈ rad(I), that is, f k ∈ I for some k ≥ 1 which means f k (f m a − f n b) = 0, i.e. fan = fbm in
Af . Thus π is injective.
To show π is surjective, take s ∈ O(D(f )). Cover D(f ) with open sets Vi on which s = agii
S
for ai ∈ A and gi ∈ A r p∈Vi p. In fact, we may arrange so that Vi = D(gi ) on the nose.
Furthermore, using Lemma 14.1.2 one can show that for some n ≥ 1, f n is a finite sum
r
f n =P i=1 bi gi , so D(f ) is even covered by finitely many open sets D(g1 ), . . . , D(gr ). Taking
P
a = ri=1 ai bi , we have that on D(gj ),
r
X
n
gj a − f aj = (gj ai bi − bi gi aj ) = 0.
i=1
 
a aj a
So fn
= gj
= s on D(gj ), which shows π fn
= s. Hence π is an isomorphism.

313
14.1. The Spectrum of a Ring Chapter 14. Scheme Theory

Example 14.1.8. For any field k, Spec k is a single point ∗ corresponding to the zero ideal,
with sheaf O(∗) ∼
= k.

Example 14.1.9. Let A = k[t1 , . . . , tn ] be the polynomial ring in n variables over k. Then
Spec A = Ank , the affine n-space over k. For example, when A = k[t] is the polynomial ring
in a single variable, Spec k[t] = A1k , the affine line.
When k = C, Hilbert’s Nullstellensatz (Corollary 2.4.3) tells us that all the closed points
of A1k correspond to maximal ideals of the form (t − α) for α ∈ C. But there is also a
non-closed, ‘generic point’ corresponding to the zero ideal which was not detected before.

closed points
−2 0 1+i generic point
Spec C[t]
(t + 2) (t) (t − (1 + i)) (0)

On the other hand, if k = Q or another non-algebraically closed field, the same closed
points corresponding to linear ideals (t − α) show up, as well as the generic point cor-
responding to (0), but there are also points corrresponding to ideals generated by higher
degree irreducible polynomials like t2 + 1. Thus the structure of Spec Q[t] is much different
than the algebraically closed case.

closed points
−2 0 ?? generic point
Spec Q[t]
(t + 2) (t) (t2 + 1) (0)

Example 14.1.10. Let X be an algebraic variety over a field k, x ∈ X a point and consider
the affine scheme Y = Spec(k[ε]/(ε2 )). We can think of Y as a “big point” with underlying
space ∗ corresponding to the zero ideal, along with a “tangent vector” extending infinites-
imally in every direction around ∗. Then any map Y → X determines a unique tangent
vector in Tx X, the tangent space of X at x. This idea is useful in intersection theory. For
example, consider the tangency of the x-axis and the parabola y = x2 in A2k :

y − x2

y
(0, 0)

314
14.1. The Spectrum of a Ring Chapter 14. Scheme Theory

As a variety, this point (0, 0) corresponds to the quotient of k-algebras k[x, y]/r(y, y − x2 ) =
k[x]/r(x2 ) = k[x]/(x) = k. Thus the information of tangency is lost. However, as an affine
scheme, (0, 0) corresponds to Spec(k[x, y]/(y, y − x2 )) = Spec(k[x]/(x2 )) so the intersection
information is preserved.

Example 14.1.11. Every affine algebraic variety X may be viewed as an affine scheme. In-
deed, let I(X) ⊆ k[t1 , . . . , tn ] be the vanishing ideal of X and set A = k[X] = k[t1 , . . . , tn ]/I(X).
Then (X, OX ) determines an affine scheme (Spec A, O). Moreover, the assignment

(X, OX ) 7→ (Spec k[X], OSpec k[X] )

is a fully faithful functor from affine k-varieties to schemes.

315
14.2. Schemes Chapter 14. Scheme Theory

14.2 Schemes
In this section we define a scheme and prove some basic properties resulting from this defi-
nition. Recall that a ringed space is a pair (X, F) where X is a topological space and F is
a sheaf of rings on X.
Definition. A locally ringed space is a ringed space (X, F) such that for all P ∈ X,
there is a ring A such that FP ∼
= Ap for some prime ideal p ⊂ A.
Example 14.2.1. Any affine scheme Spec A is a locally ringed space by (1) of Theo-
rem 14.1.7. We will sometimes denote the structure sheaf O by OA .
Definition. The category of locally ringed spaces is the category whose objects are
locally ringed spaces (X, F) and whose morphisms are morphisms of ringed spaces (X, F) →
(Y, G) such that for each P ∈ X, the induced map fP# : OY,f (P ) → OX,P is a morphism of
local rings, i.e. (fP# )−1 (mP ) = mf (P ) where mP (resp. mf (P ) ) is the maximal ideal of the local
ring OX,P (resp. OY,f (P ) ).
We are now able to define a scheme.
Definition. A scheme is a locally ringed space (X, OX ) that admits an open covering {Ui }
such that each Ui is affine, i.e. there are rings Ai such that (Ui , OX |Ui ) ∼
= (Spec Ai , OAi ) as
locally ringed spaces.
The category of schemes Sch is defined to be the full subcategory of schemes in the
category of locally ringed spaces. Denote the subcategory of affine schemes by AffSch. Also
let CommRings denote the category of commutative rings with unity.
Proposition 14.2.2. There is an isomorphism of categories

AffSch −→ CommRingsop
(X, OX ) 7−→ OX (X)
(Spec A, O) →−7 A.
Proof. (Sketch) First suppose we have a homomorphism of rings f : A → B. By Lemma 14.1.1
this induces a morphism f ∗ : Spec B → Spec A, p 7→ f −1 (p) which is continuous since
f −1 (V (a)) = V (f (a)) for any ideal a ⊂ A. Now for each p ∈ Spec B, define the localiza-
tion fp : Af ∗ p → Bp using the universal property of localization. Then for any open set
V ⊆ Spec A, we get a map
f # : OA (V ) −→ OB ((f ∗ )−1 (V )).
One checks that each is a homomorphism of rings and commutes with the restriction maps.
Thus f # : OA → OB is defined. Moreover, the induced map on stalks is just each fp , so the
pair (f ∗ , f # ) gives a morphism (Spec B, OB ) → (Spec A, OA ) of locally ringed spaces, hence
of schemes.
Conversely, take a morphism of schemes (ϕ, ϕ# ) : (Spec B, OB ) → (Spec A, OA ). This
induces a ring homomorphism Γ(Spec A, OA ) → Γ(Spec B, OB ) but by (2) of Theorem 14.1.7,
Γ(Spec A, OA ) ∼ = A and Γ(Spec B, OB ) ∼
= B so we get a homomorphism A → B. It’s easy
to see that the two functors described give the required isomorphism of categories.

316
14.2. Schemes Chapter 14. Scheme Theory

Example 14.2.3. We saw in Example 14.1.8 that for any field k, Spec k = ∗ is a point
` O(∗)
with structure sheaf ` = k. If A = L1 × · · · × Lr is a finite étale k-algebra, then
Spec A = Spec L1 · · · Spec Lr is (schematically) a disjoint union of points.

Example 14.2.4. Let A be a DVR with residue field k. Then Spec A = {0, mA }, a closed
point for the maximal ideal m and a generic point for the zero ideal. There are two open
subsets here, {0} and Spec A, and we have OA ({0}) = k and OA (Spec A) = A.

Example 14.2.5. If k is a field and A is a finitely generated k-algebra, then the closed
points of X = Spec A are in bijection with the closed points of an affine variety over k with
coordinate ring A (see Example 14.1.11).

Example 14.2.6. Let A = Z (or any Dedekind domain). Then dim A = 1 and it turns out
that dim Spec A = 1 for some appropriate notion of dimension (see Section 14.3). Explicitly,
Spec Z has a closed point for every prime p ∈ Z and a generic point for (0):

closed points
generic point
Spec Z
2 3 5 7 11 (0)

Example 14.2.7. Let k be a field, X1 = X2 = A1k two copies of the affine line and U1 =
U2 = A1k r {0}, where 0 is the closed point of A1k corresponding to (x) in k[x]. Then we can
glue together X1 and X2 along the identity map U1 → U2 to get a scheme X which looks
like the affine line with the origin “doubled”. Note that X is not affine!

A1k r {0}

317
14.3. Properties of Schemes Chapter 14. Scheme Theory

14.3 Properties of Schemes


Many definitions in ring theory can be rephrased for schemes. For example:

Definition. A scheme X is reduced if for all open U ⊆ X, OX (U ) has no nilpotent


elements.

Definition. A scheme X is integral if for all open U ⊆ X, OX (U ) has no zero divisors.

Lemma 14.3.1. X is integral if and only if X is reduced and irreducible as a topological


space.

Proof. ( =⇒ ) Clearly integral implies reduced, so we just need to prove X is irreducible.


Suppose X = U ∪ V for open subsets U, V ⊆ X. Then OX (U ∪ V ) = OX (U ) × OX (V ) which
is not a domain unless one of OX (U ), OX (V ) is 0. In that case, U or V is empty, so this
shows X is irreducible.
( ⇒= ) Suppose X is reduced and irreducible, but there exists an open set U ⊆ X and
f, g ∈ OX (U ) with f g = 0. Define closed sets

C = {P ∈ U | fP ∈ mP ⊂ OX,P }
D = {P ∈ U | gP ∈ mP ⊂ OX,P }.

Then by definition of OX , we must have C ∪ D = U . By irreducibility, C = U without loss of


generality. Thus for any affine open set U 0 ⊆ U with U 0 = Spec A, we have (OX |U 0 )(D(f )) =
0 but by (3) of Theorem 14.1.7, OU 0 (D(f )) ∼
= Af , the localization of A at powers of f . When
Af = 0, f is nilpotent but by assumption this means f = 0. Hence X is integral.

Definition. The dimension of a scheme X (or any topological space) is

dim X = sup{n ∈ N0 | there exists a chain of irreducible, closed sets X0 ( X1 ( · · · ( Xn ⊆ X}.

Proposition 14.3.2. Let A be a noetherian ring. Then dim Spec A = dim A, the Krull
dimension of A.

Be warned that the converse to Proposition 14.3.2 is false in general.

Definition. Let X be any scheme. For a point P ∈ X, we define the codimension of P to


be the Krull dimension of the local ring at P , that is codim P = dim OX,P .

Note that by commutative algebra, the codimension of P is equal to the height of the
prime ideal p ⊂ A associated to P for any choice of affine open neighborhood P ∈ U =
Spec A.

Definition. Let X be a scheme. Then

ˆ X is locally noetherian if each stalk OX,P is a local noetherian ring.

ˆ X is noetherian if X is integral and locally noetherian.

318
14.3. Properties of Schemes Chapter 14. Scheme Theory

ˆ An integral scheme X is normal if each stalk OX,P is integrally closed in its field of
fractions.
ˆ X is regular if each OX,P is regular as a local ring, that is, dim OX,P = dim mP /m2P
as OX,P /mP -vector spaces.
Definition. Let U ⊆ X be an open subset. Then (U, OX |U ) is a scheme which we call an
open subscheme of X. The natural morphism j : U ,→ X, j # : OX → j∗ OX |U is called
an open immersion.
Example 14.3.3. For X = Spec A, let f ∈ A and recall the open set D(f ) defined in
Theorem 14.1.7. Then D(f ) is an open subscheme of X and the open immersion D(f ) ,→ X
corresponds to the natural inclusion of prime ideals Spec Af ,→ Spec A (this is a property of
any localization).
Definition. Let A → A/I be a quotient homomorphism of rings. Then the induced mor-
phism Spec(A/I) → Spec A is called an affine closed immersion. For a general morphism
of schemes f : X → Y , f is called a closed immersion if f is injective, f (X) ⊆ Y is closed
and there exists a covering of X by affine open sets {Ui } such that each f |Ui : Ui → f (Ui ) is
an affine closed immersion. The set f (X) is called a closed subscheme of Y .
Definition. Let X be a scheme. The category of schemes over X, denoted SchX , consists
p
of objects Y →
− X, where Y is a scheme and p is a morphism, and morphisms Y → Z making
the following diagram commute:
Y Z

Example 14.3.4. Every scheme Y is a scheme over Spec Z. Write Y = Ui where Ui ∼


S
=
Spec Ai for rings Ai . Then for each of these there is a canonical homomorphism ϕi : Z → Ai
which induces ϕ∗i : Spec ∼
S Ai → Spec Z. Composing these with the isomorphisms Ui = Spec Ai ,
we get a map Y = Ui → Spec Z.
The fibre of a topological cover p : Y → X can be interpreted as a fibre product:

p−1 (x) := {x} ×X Y Y

{x} X

We next construct fibre products in the category SchX and use these to construct the alge-
braic analogue of a fibre.
Definition. Let X be a scheme and Y, Z schemes over X. A fibre product of Y and Z
over X, denoted Y ×X Z, is a scheme over both Y and Z such that the diagram

319
14.3. Properties of Schemes Chapter 14. Scheme Theory

Y ×X Z

Y Z

commutes and Y ×X Z is universal with respect to such diagrams, i.e. for any scheme W
over both Y and Z, the following diagram can be completed uniquely:

∃!

Y ×X Z

Y Z

Given f : Y → X and any scheme Z over X, the induced map fZ : Y ×X Z → Z is called


the base change of f over Z.

Theorem 14.3.5. For any schemes Y, Z over X, there exists a fibre product Y ×X Z which
is unique up to unique isomorphism.

Proof. (Sketch) First suppose X, Y and Z are all affine; write X = Spec A, Y = Spec B
and C = Spec Z. Then Spec(B ⊗A C) is a natural candidate for the fibre product in this
case. Indeed, the tensor product satisfies the universal property conveyed by the following
diagrams:

320
14.3. Properties of Schemes Chapter 14. Scheme Theory

∃!

B ⊗A C B ⊗A C

B C B C

A A

Applying the functor Spec yields the right diagrams with arrows reversed, by Proposi-
tion 14.2.2, so the fibre product exists in the affine case.
In general, note that once we construct any fibre product, it will be unique up to unique
isomorphism by the universal property, just as in every proof of the solution to a universal
mapping problem. Now suppose S X and Z are affine and Y is arbitrary. Write Y as a union
of affine open subschemes Y = Yi . Then by the affine case, Yi ×X Z exists for each Yi .
For each pair of overlapping open sets Yi ∩ Yj , set Uij = p−1
Yi (Yi ∩ Yj ) ⊆ Yi ×X Z, where pYi
is the morphism Yi ×X Z → Yi . Then it’s easy to verify that Uij = (Yi ∩ Yj ) ×X Z (that
is, Uij satisfies the definition of the fibre product for Yi ∩ Yj and Z over X), and by the
universal property, there are unique isomorphisms ϕij : Uij → Uji for each overlapping pair,
commuting with all projections. Therefore we may glue together the fibre products Yi ×X Z
along the isomorphisms ϕij to get a scheme Y ×X Z which then satisfies the definition of the
fibre product for Y and Z over X. Now, covering Z by affine open subschemes and repeating
this process will construct Y ×X Z for any schemes Y, ZSover an affine scheme X.
Finally, let X be an arbitrary scheme and write X = Xi for affine open subschemes Xi .
Let q : Y → X and r : Z → X be the given morphisms and for each Xi , set Yi = q −1 (Xi )
and Zi = r−1 (Xi ). By the affine case, each Yi ×Xi Zi exists, but any morphisms f : W → Yi
and g : W → Z making the diagram
W

Yi Z

commute must satisfy g(W ) ⊆ Zi . It follows that Yi ×Xi Z = Yi ×X Zi and the gluing
procedure from above allows us to construct Y ×X Z from these. Hence the fibre product
exists in every case.

321
14.3. Properties of Schemes Chapter 14. Scheme Theory

Definition. Let p : Y → X be a morphism of schemes, x ∈ X a point and k(x) = OX,x /mx


the residue field at x, with natural map Spec k(x) ,→ X. Then the fibre of p at x is the fibre
product Yx := Y ×X Spec k(x).
Lemma 14.3.6. Let p : Y → X be a morphism of schemes and x ∈ X any point. Then
(a) The fibre Yx = Y ×X Spec k(x) is a scheme over the point Spec k(x).

(b) The underlying topological space of Yx is homeomorphic to the set p−1 (x) of preimages
of x.
p
(c) The assignment (Y →
− X) 7→ Yx is functorial.
Example 14.3.7. Let A be a DVR and consider the affine scheme X = Spec A. We saw
in Example 14.2.4 that X has a closed point m = mA and a generic point (0). For any
morphism p : Y → X, there are two fibres:
ˆ The generic fibre Y(0) , which is an open subscheme of Y

ˆ The special fibre Ym , which is a closed subscheme of Y .


Let Y be a scheme over X and define the diagonal map ∆ : Y → Y ×X Y coming from
the universal property applied to the diagram


id id
Y ×X Y

Y Y

Definition. A morphism Y → X is called separated if the diagonal ∆ : Y → Y ×X Y


is a closed immersion of schemes. We will say a scheme Y over X is separated if the
corresponding morphism Y → X is separated, and a scheme is simply separated if it is
separated as a scheme over Spec Z.
Example 14.3.8. Let X = Spec A and Y = Spec B be affine schemes, with Y → X a
morphism between them. This corresponds to a ring homomorphism A → B which makes
B into an A-module. The diagonal ∆ : Y → Y ×X Y corresponds to the multiplication map
B ⊗A B → B, b ⊗ b0 7→ bb0 , which is a homomorphism of B-modules. This map is clearly
surjective, so ∆ is a closed immersion. Hence every affine scheme (and morphism of affine
schemes) is separated.

322
14.3. Properties of Schemes Chapter 14. Scheme Theory

Example 14.3.9. One can show that the affine line with the origin doubled (Example 14.2.7)
is not separated as a scheme over Spec k.
One perspective on separatedness is that it is a suitable replacement for the Hausdorff
condition in algebraic geometry. In the Zariski topology on any scheme, there are always
proper open subsets that are dense, so the Hausdorff property usually fails to hold.
Definition. A morphism f : Y → X is of finite type if there exists an affine covering
Ui , with Ui = Spec Ai , such that each f −1 (Ui ) has an open covering f −1 (Ui ) =
S
X =
S ni
j=1 Spec Bij for ni < ∞ and Bij a finitely generated Ai -algebra. Further, we say f is
a finite morphism if each ni = 1, i.e. f −1 (Ui ) = Spec Bi for some finitely generated
Ai -algebra Bi .
Definition. A separated morphism f : Y → X is proper if it is of finite type and for every
morphism Z → X, the base change morphism Y ×X Z → Z is closed.
Lemma 14.3.10. Let X, Y, Z be noetherian schemes. Then for any morphism f : Y → X,
(a) If f is an open immersion, then f is separated.
(b) If f is a closed immersion, then f is separated and proper.
(c) If g : Z → Y is separated (resp. proper) then the composition f ◦ g : Z → X is
separated (resp. proper).
(d) If Z is a scheme over X, the base change Y ×X Z → Z is separated and proper.
(e) If f is finite, then f is proper.
Example 14.3.11. (Projective line over a scheme) Let X = Spec A be an affine scheme.
Then the “affine line” A1X = Spec A[t] is an affine scheme over X. Set X1 = A1X = Spec A[t]
and X2 = A[t−1 ]. Then each contains an open subscheme isomorphic to U = Spec A[t, t−1 ],
coming from applying Spec to the diagram of A-algebras

A[t, t−1 ]

A[t] A[t−1 ]

Gluing along these isomorphic open subschemes gives us a scheme P1X = X1 ∪U X2 , called
the projective line over X. In the affine case, we will write P1X = P1A .
S When X is an arbitrary scheme, X has a covering by 1open affine subschemes X =
Ui and a gluing construction defines the projective line PX . We will generalize this in
Section 14.6.

323
14.4. Sheaves of Modules Chapter 14. Scheme Theory

14.4 Sheaves of Modules


Through Proposition 14.2.2, we are able to transfer commutative ring theory to the language
of affine schemes. In this section, we define a suitable setting for transferring module theory
to the language of sheaves and schemes.

Definition. Let (X, OX ) be a ringed space. A sheaf of OX -modules, or an OX -module


for short, is a sheaf of abelian groups F on X such that each F(U ) is an OX (U )-module and
for each inclusion of open sets V ⊆ U , the following diagram commutes:

OX (U ) × F(U ) F(U )

OX (V ) × F(V ) F(V )

If F(U ) ⊆ OX (U ) is an ideal for each open set U , then we call F a sheaf of ideals on X.

Example 14.4.1. Let f : Y → X be a morphism of ringed spaces. Then the pushforward


sheaf f∗ OY is naturally an OX -module on X via f # : OX → f∗ OY . Additionally, the kernel
sheaf of f # , defined on open sets by (ker f # )(U ) = ker(OX (U ) → f∗ OY (U )), is a sheaf of
ideals on X.

Most module terminology extends to sheaves of OX -modules. For example,

ˆ A morphism of OX -modules is a morphism of sheaves F → G such that each F(U ) →


G(U ) is an OX (U )-module map. We write HomX (F, G) = HomOX (F, G) for the set of
morphisms F → G as OX -modules. This defines the category of OX -modules, written
OX -Mod.

ˆ Taking kernels, cokernels and images of morphisms of OX -modules again give OX -


modules.

ˆ Taking quotients of OX -modules by OX -submodules again give OX -modules.

ˆ An exact sequence of OX -modules is a sequence F 0 → F → F 00 such that each F 0 (U ) →


F(U ) → F 00 (U ) is an exact sequence of OX (U )-modules.

ˆ Basically any functor on modules over a ring generalizes to an operation on OX -


Vn HomOX (F, G); direct sum F ⊕ G; tensor product
modules, including Hom, written
F ⊗OX G; and exterior powers F.

The most important of these constructions for our purposes will be the direct sum oper-
ation.

Definition. An OX -module F Sis free (of rank r) if F ∼ ⊕r


= OX as OX -modules. F is locally
free if X has a covering X = Ui such that each F|Ui is free as an OX |Ui -module.

324
14.4. Sheaves of Modules Chapter 14. Scheme Theory

Remark. The rank of a locally free sheaf of OX -modules is constant on connected com-
ponents. In particular, the rank of a locally free OX -module is well-defined whenever X is
connected.
Definition. A locally free OX -module of rank 1 is called an invertible sheaf.
Let A be a ring, M an A-module and set X = Spec A. To extend module theory to
the language of schemes, we want to define an OX -module M f on X. To start, for each
p ∈ Spec A, let Mp = M ⊗A Ap be the localization of the module M at p. Then Mp is an
Ap -module consisting of ‘formal fractions’ ms where m ∈ M and s ∈ S = A r p. For each
open set U ⊆ X, define
( )
f(U ) = h : U →
a m
M Mp s(p) ∈ Mp , ∃ p ∈ V ⊆ U, m ∈ M, s ∈ A with s(q) = for all q ∈ V .

p∈U
s

(Compare this to the construction of the structure sheaf OA on Spec A in Section 14.1. Also,
note that necessarily the s ∈ A in the definition above must lie outside of all q ∈ V .)

Proposition 14.4.2. Let M be an A-module and X = Spec A. Then M


f is a sheaf of
OX -modules on X, and moreover,
fp ∼
(1) For any p ∈ Spec A, M = Mp as rings.
f) ∼
(2) Γ(X, M = M as A-modules.
f(D(f )) ∼
(3) For any f ∈ A, M = Mf = M ⊗A Af as A-modules.
The proof is similar to the proof of Theorem 14.1.7; both can be found in Hartshorne.
Proposition 14.4.3. Let X = Spec A. Then the association

A-Mod −→ OX -Mod
M 7−→ M
f

defines an exact, fully faithful functor.


Proof. Similar to the proof of Proposition 14.2.2.
These Mf will be our affine model for modules over a scheme X. We next define the
general notion, along with an analogue of finitely generated modules over a ring.
Definition. Let (X, OX ) be a scheme. An OX -module F is quasi-coherent if there is an
affine covering X = Xi , with Xi = Spec Ai , and Ai -modules Mi such that F|Xi ∼
S
=Mfi as
OX |Xi -modules. Further, we say F is coherent if each Mi is a finitely generated Ai -module.
Example 14.4.4. For any scheme X, the structure sheaf OX is obviously a coherent sheaf
on X.
Let QCohX (resp. CohX ) be the category of quasi-coherent (resp. coherent) sheaves of
OX -modules on X.

325
14.4. Sheaves of Modules Chapter 14. Scheme Theory

Theorem 14.4.5. QCohX and CohX are abelian categories.


Example 14.4.6. Let X = Spec A, I ⊆ A an ideal and Y = Spec(A/I). Then the natural
inclusion i : Y ,→ X is a closed immersion by definition, and it turns out that i∗ OY ∼
= A/I
g
as OX -modules, so i∗ OY is a quasi-coherent, even coherent, sheaf on X.
We next identify the image of the functor M 7→ M
f from Proposition 14.4.3.

Theorem 14.4.7. Let X = Spec A. Then there is an equivalence of categories



A-Mod −
→ QCohX .
Moreover, if A is noetherian, this restricts to an equivalence

A-mod −
→ CohX
where A-mod denotes the subcategory of finitely generated A-modules.
Proof. (Sketch) The association M 7→ M f sends an A-module to a quasi-coherent sheaf on
X = Spec A by definition of quasi-coherence. Further, one can prove that a sheaf F on X is
a quasi-coherent OX -module if and only if F ∼
=M f for an A-module M . The inverse functor
QCohX → A-Mod is given by F 7→ Γ(X, F).
When A is noetherian, the above extends to say that F is coherent if and only if F ∼
=Mf
for a finitely generated A-module M . The rest of the proof is identical.
The following lemma generalizes Example 14.4.6.
Lemma 14.4.8. Let f : Y → X be a morphism of schemes and let G be a quasi-coherent
sheaf on Y . Then f∗ G is a quasi-coherent sheaf on X. Further, if G is coherent and f is a
finite morphism, then f∗ G is also coherent.
Note that the second statement is false in general.
Next, we construct an important example of a quasi-coherent sheaf on a scheme. As
always, we begin with a construction on rings. All of the following will be defined again and
proven, as well as significantly expanded, in Chapter 15.
Definition. Let A → B be a ring homomorphism. The module of relative differentials
for B/A is defined to be
Ω1B/A := Zhdb | b ∈ Bi/N,
the quotient of the free B-module generated by formal symbols db for all b ∈ B by the
submodule N = hda, d(b + b0 ) − db − db0 , d(bb0 ) − b(db0 ) − (db)b0 i. This is the universal B-
module for these three relations.
Example 14.4.9. If A = k is a field and B is a finitely generated k-algebra, write B =
k[t1 , . . . , tn ]/(f1 , . . . , fr ). Then B is the coordinate ring of the variety in Ank cut out by the
fi and * n +
X ∂fj
Ω1B/k = khdti i/ dti
i=1
∂ti
is the module of total derivatives on this variety.

326
14.4. Sheaves of Modules Chapter 14. Scheme Theory

Lemma 14.4.10. Let A → B be a ring homomorphism. Then

(a) For any A-algebra C, Ω1B⊗A C/C ∼


= Ω1B/A ⊗A C.

(b) For any multiplicative set S ⊆ B, Ω1S −1 B/A ∼


= S −1 Ω1B/A = Ω1B/A ⊗B S −1 B.
Proof. See Proposition 15.2.6.
That is, the functor B 7→ Ω1B/A commutes with base change and localization. We now
give the analogous construction for OX -modules, starting in the affine case.

Definition. Let A → B be a ring homomorphism. The sheaf of relative differentials is


e B/A on Spec B defined by the module Ω1 .
the OB -module Ω B/A

Lemma 14.4.11. Let A → B be a ring homomorphism. Then

(a) Ω
e B/A is a quasi-coherent sheaf on Spec B.

e B/A (D(f )) ∼
(b) For any element f ∈ B, Ω = ΩBf /A where Bf is the localization of B at
powers of f .

Now consider the map m : B ⊗A B → B, m(b1 ⊗ b2 ) = b1 b2 from Example 14.3.8. Let I


be the kernel of m. Since m is surjective, this means B ⊗A B/I ∼= B. Since I acts trivially
2 2
on I/I , there is an induced module action of B ⊗A B/I on I/I , and thus a corresponding
B-module structure on I/I 2 . The proof of the following fact can be found in Eisenbud,
among other places.

Lemma 14.4.12. ΩB/A ∼


= I/I 2 .
Example 14.4.13. In Example 14.4.9, the isomorphism ΩB/k ∼
= I/I 2 is induced by the map

B −→ ΩB/k
ti 7−→ dti .

Let Y → X be a separated morphism of schemes and let ∆ : Y → Y ×X Y be the


corresponding diagonal. This induces a morphism of sheaves ∆# : OY ×X Y → ∆∗ OX which
has kernel sheaf I (a sheaf on Y ×X Y ). This I in fact defines the closed subscheme
∆(Y ) ⊆ Y ×X Y .

Lemma 14.4.14. For Y → X, ∆ and I as above,

(a) O∆(Y ) ∼
= OY ×X Y /I as sheaves on ∆(Y ).
(b) I/I 2 is an O∆(Y ) -module.

Identifying Y with its image ∆(Y ) in the fibre product Y ×X Y allows us to define a
sheaf analogue of the module of differentials by pulling back I/I 2 .

Definition. For a separated morphism Y → X, the sheaf of relative differentials ΩY /X


is the pullback:

327
14.4. Sheaves of Modules Chapter 14. Scheme Theory

ΩY /X I/I 2

Y ∆(Y )

Remark. ΩY /X is a sheaf of OY -modules on Y . Moreover, on an affine patch Spec B ⊆ Y , the


e B/A ∼
sheaf of relative differentials restricts to Ω g2 for some rings A → B. In particular,
= I/I
ΩY /X is quasi-coherent.
We finish the section by discussing some applications of relative differentials. Again,
differentials and smoothness will be studied in much greater detail in Chapter 15. Let k be
a field and X a connected scheme over Spec k of finite type and dimension d.
Definition. A k-scheme X is smooth over k if the sheaf of relative differentials ΩX/ Spec k
is locally free of rank d.
Theorem 14.4.15. Assume k is algebraically closed and dim X = d. Then the following
are equivalent:
(1) X is smooth over k.
(2) For every affine open subset U ∼
= Spec(k[t1 , . . . , tn ]/(f1 , . . . , fm )), the Jacobian matrix
 
∂fi
JP = (P )
∂tj
has rank n − d at all closed points P ∈ X.
(3) For every closed point P ∈ X, the stalk OX,P is a regular local ring.
This amazing result unites geometry (ΩX/ Spec k being locally free), algebra (OX,P being
regular) and analysis (the vanishing of partial derivatives in JP ) into one concept of smooth-
ness. Unfortunately, the theorem fails when k is not algebraically closed, but it still hints at
a deep intersection between all three areas of math.
Theorem 14.4.16. Let X be any variety over an algebraically closed field. Then there is a
dense open subset which is smooth.
Recall that a finitely generated k-algebra A is finite étale if A = L1 × · · · × Lr for finite,
separable extensions Li /k.
Proposition 14.4.17. A finitely generated k-algebra A is finite étale if and only if Ω1A/k = 0.
Finally, we give an important construction relating the geometry and algebra of a smooth
scheme. Let X be smooth over a field k and let n = dim X.
Definition. The canonical sheaf of X is the nth exterior power sheaf
^
ωX := n ΩX/ Spec k .

328
14.4. Sheaves of Modules Chapter 14. Scheme Theory

Here are some interesting facts about the canonical sheaf:

ˆ ωX is an invertible sheaf on X.

ˆ One can define the geometric genus of X by g(X) := dimk Γ(X, ωX ). Then when X
is a curve (a smooth scheme of dimension 1), g(X) is equal to the arithmetic genus of
X, another important algebraic invariant. These genera are not equal in general.

ˆ If X is a curve over C, then the genus of the corresponding Riemann surface X(C)
is precisely g(X), so the canonical bundle carries important topological information
about X(C).

329
14.5. Group Schemes Chapter 14. Scheme Theory

14.5 Group Schemes


Recall that a group is a set G together with three maps,

µ : G × G −→ G (multiplication)
e : {e} ,−→ G (identity)
i : G −→ G (inverse)

satisfying associativity, identity and and inversion axioms. This generalizes to the notion of
a group object in an arbitrary category C. We state the definition for scheme categories here.
p
Definition. A group scheme over a scheme X is a scheme G →
− X with morphisms

µ : G × G −→ G
e : {e} ,−→ G
i : G −→ G

satisfying the following axioms:

(1) (Associativity) µ ◦ (id × µ) = µ ◦ (µ × id).

(2) (Identity) µ ◦ (id × e) = id = µ ◦ (e × id).

(3) (Inversion) µ ◦ (id × i) = e ◦ p = µ ◦ (i × id).

Definition. A group scheme G over X is finite if p : G → X is a finite morphism, and


flat if p : G → X is a flat morphism, i.e. p∗ OG is a sheaf of flat OX -modules.

Remark. When G is a finite group scheme over X, flat is equivalent to locally free.

The following describes an equivalent, and equally important, perspective on group


schemes using the language of functors.

Proposition 14.5.1. Let G be a scheme over X. Then a choice of group scheme structure
on G is equivalent to a compatible choice of group structure on the sets HomX (Y, G) for all
schemes Y over X. That is, a group scheme structure is a functor SchX → Groups such
that the composition with the forgetful functor Groups → Sets is representable.

Proof. This is a basic application of Yoneda’s Lemma.

Example 14.5.2. Let G be a finite group of order Qn n and let X be any scheme. The constant
group scheme on G over X is defined as GX := i=1 X, with projection map induced by the
identity on each disjoint copy of X. Multiplication QnGX ×X GX → GX is given by sending
(P, Q), where P = x ∈ X in the gi th component of i=1 X and Q = x in the gj th component
(gi , gj ∈ G) to the corresponding point P Q = x in the gi gj th component of the disjoint union.
(Note that P and Q must correspond to the same point x ∈ X by definition of the fibre
product – draw the diagram!) Similarly, the identity is the morphism Qtaking X onto the
copy of X indexed by the identity element eG ∈ G, e : X → XeG ⊆ ni=1 X. Finally, the

330
14.5. Group Schemes Chapter 14. Scheme Theory

inversion morphism i : GX → GX takes P in the gi th component to the corresponding point


P −1 in the gi−1 th component.
This definition can be extended to an arbitrary group G. Note that when G is finite, GX
is a finite (étale) group scheme over X. A special case of this is the trivial group scheme
{1}X = X. Thus every scheme is a group scheme.

Example 14.5.3. Let X = Spec A be affine and recall the affine line A1X = Spec A[t]
constructed in Example 14.3.11. Then A1X is an affine group scheme over X, denoted Ga ,
called the additive group scheme over X, with morphisms induced by the following ring
homomorphisms:

µ∗ : A[t] −→ A[t] ⊗ A[t]


t 7−→ t ⊗ 1 + 1 ⊗ t

e : A[t] −→ A
t 7−→ 0

i : A[t] −→ A[t]
t 7−→ −t.

Notice that these are just the axioms of a Hopf algebra! This construction generalizes to the
affine line over a non-affine scheme as well.

Example 14.5.4. For X = Spec A, the multiplicative group scheme over X is Gm :=


Spec A[t, t−1 ] with morphisms induced by

µ∗ : A[t, t−1 ] −→ A[t, t−1 ] ⊗ A[t, t−1 ]


t 7−→ t ⊗ t
t 7−→ t−1 ⊗ t−1
−1

e∗ : A[t, t−1 ] −→ A
t, t−1 7−→ 1
i∗ : A[t, t−1 ] −→ A[t, t−1 ]
t 7−→ t−1
t−1 7−→ t.

Note that when A = k is a field, these are just the schematic versions of the algebraic groups
Ga,k and Gm,k . This shows that group schemes are a direct generalization of algebraic groups.

Example 14.5.5. For X = Spec A, the nth roots of unity form a group scheme defined by
µn = Spec(A[t, t−1 ]/(tn − 1)). This is a finite group subscheme of Gm .

Example 14.5.6. If char A = p > 0, then αp = Spec(A[t]/(tp )) defines a group scheme over
Spec A which is isomorphic as a scheme to Spec A, but not as a group scheme!

Example 14.5.7. The Jacobian of a curve is a group scheme. In particular, an elliptic curve
(a dimension 1 scheme with a specified point O) is a group scheme with identity O.

331
14.5. Group Schemes Chapter 14. Scheme Theory

p q
Definition. Let G → − X be a finite, flat group scheme. A left G-torsor is a scheme Y → − X
with q finite, locally free and surjective, together with a group action ρ : G ×X Y → Y , which
satisfies:

(1) ρ ◦ (e × idY ) is equal to the projection map X ×X Y → Y .

(2) ρ ◦ (idG × ρ) = ρ ◦ (µ × idY ) : G ×X G ×X Y → G ×X Y → Y .

(3) ρ × idY : G ×X Y → Y ×X Y is an isomorphism of X-schemes.

Right G-torsors are defined similarly.

Remark. The idea is that a G-torsor is exactly the same as G, except we have forgotten
the “identity point” e (which is a morphism, not a point).

Example 14.5.8. Left multiplication defines a G-torsor structure on G itself.

Example 14.5.9. Let k be a field and m an integer such that char k - m. Let µm =
Spec(k[t, t−1 ]/(tm − 1)) be the group scheme of mth roots of unity √ over Spec k. Take a ∈
× × m
k /(k ) (that is, a is not an mth power in k) and set L = k( a), which is a finite field
m

extension of k. We claim Y = Spec L is a µm -torsor over Spec k.


Let ζ be a primitive mth root of unity. Up to finite base change, we may assume ζ ∈ k.
Define
√ √
ρ∗ : k( m a) −→ k[t, t−1 ]/(tm − 1) ⊗k k( m a)
√ √
m
a 7−→ ζ ⊗ m a.

This defines a morphism ρ : µm ×Spec k Y → Y and one can prove that it satisfies axioms (1)
and (2) of a torsor by checking the corresponding properties for ρ∗ . When char k - m, µm
is a reduced scheme over Spec k and it’s easy to see that µm ∼ = (Z/mZ)Spec k , the constant
group scheme on Z/mZ over Spec k. Moreover, L/k is a Galois extension with Gal(L/k) ∼ =
∼ Qm
Z/mZ, L ⊗k L = i=1 L and this has a corresponding Galois action which induces the

isomorphism L ⊗k L − → k[t, t−1 ]/(tm − 1) ⊗k L. Applying Spec again, we get the isomorphism

µm ×Spec k Y −
→ Y ×Spec k Y so Y is indeed a µm -torsor.
Using Artin-Schreier
√ theory, one can show that every µm -torsor arises in this way, i.e. as
Spec L for L = k( a).
m

332
14.6. The Proj Construction Chapter 14. Scheme Theory

14.6 The Proj Construction


The construction of projective varieties in Section 12.5 generalizes in the same way affine
varieties generalized to affine schemes in Section 14.1.
Let S = d=0 Sd be a graded ring and set S+ = ∞
L∞ L
d=1 Sd . Define the projective spectrum
of S:
Proj(S) = {p ⊂ S | p is a homogeneous prime ideal, p 6⊃ S+ }.
For each homogeneous ideal a ⊆ S, let V (a) denote the set of homogeneous prime ideals
containing a:
V (a) = {p ∈ Proj(S) | p ⊇ a}.

Lemma 14.6.1. Let S be a graded ring. Then

(i) V (0) = Proj(S), V (1) = ∅ and V (a) = ∅ for any homogeneous ideal a ⊇ S+ .

(ii) If (ai )i∈I is any family of homogeneous ideals of S, then


!
X \
V ai = V (ai ).
i∈I i∈I

(iii) V (a ∩ b) = V (ab) = V (a) ∪ V (b) for any homogeneous ideals a, b of A.

Proof. Same as the proof of Lemma 14.1.2, using the properties in Lemma 12.5.3.
As in the affine case, this shows that the sets V (a form the closed sets of a Zariski
topology on Proj(S). Next, we define a sheaf of rings O on Proj(S). If p is a homogeneous
prime ideal of S, let S(p) be the subring of elements of degree 0 in the localization T −1 S,
where T = (S r p) ∩ S h and S h is the set of homogeneous elements of S. Define S(f ) similarly
for f ∈ S h . For an open set U ⊆ Proj(S), define
( )
a f h
O(U ) = s : U → S(p) s(p) ∈ S(p) , ∃ q ∈ V ⊆ U such that s(q) = for all q ∈ V, f, g ∈ S .

p∈U
g

Theorem 14.6.2. (Proj S, O) is a scheme. Moreover,

(1) For any p ∈ Proj S, Op ∼


= S(p) .
(2) For any f ∈ S h ∩ S+ , define the open set D+ (f ) = {p ∈ Proj S | f ∈ 6 p}. Then the

D+ (f ) form a basis for the topology on Proj S and O(D+ (f )) = S(f ) as rings.

(3) Further, for any f ∈ S h ∩ S+ , D+ (f ) ∼


= Spec S(f ) as affine schemes.
Proof. Similar to the proof of Theorem 14.1.7. The fact that (Proj S, O) is a locally ringed
space follows from (1), and then (3) shows that it is covered by affine schemes, so it is a
scheme.

Definition. For a graded ring S, the scheme (Proj S, O) is called a projective scheme.

333
14.6. The Proj Construction Chapter 14. Scheme Theory

Example 14.6.3. Let A be a commutative ring and S = A[x0 , . . . , xn ] be the graded poly-
nomial ring in n + 1 variables over A. The projective n-space over A is the projective scheme

PnA := Proj S = Proj A[x0 , . . . , xn ].

If k is an algebraically closed field, then the closed points of Pnk are in bijection with the
points of the projective variety defined in Section 12.5.
Proposition 14.6.4. Suppose ϕ : S → T is a morphism of graded rings and set

U = {p ∈ Proj T | p 6⊃ ϕ(S+ )}.

Then
(a) U is an open subset of Proj T .
(b) ϕ induces a morphism of schemes ϕ∗ : U → Proj S.
(c) If ϕ|Sd : Sd → Td is an isomorphism for all d, then U = Proj T . Moreover, ϕ∗ :
Proj T → Proj S is an isomorphism.
Example 14.6.5. More generally, if ϕ : S → T is merely surjective, then U = Proj T and
ϕ : Proj T ,→ Proj S is a closed immersion. One can show that every closed subscheme of
Proj S is of the form Proj(S/I) for some homogeneous ideal I ⊆ S.
Example 14.6.6. Suppose f : A → B is a homomorphism of rings and f ∗ : Spec B →
Spec A is the corresponding morphism of affine schemes, then

PnB ∼
= PnA ×Spec A Spec B.
Example 14.6.7. One defines projective n-space over an arbitrary scheme X, written PnX ,
by
PnX := PnZ ×Spec Z Y.
Alternatively, PnX may be constructed by gluing together n + 1 copies of affine n-space
A1X = Spec A[t1 , . . . , tn ] along the isomorphic open subsets Xi = {ti 6= 0} (when X is affine;
in the general case, glue affine subschemes together as in the previous example). The natural
morphism PnX → X generalizes in the following way.
Definition. A morphism of schemes Y → X is projective if it factors through a closed
immersion Y → PnX for some n ≥ 1.
Example 14.6.8. For any ring A and any graded ring S such that S0 = A, the morphism
Proj S → Spec A is a projective morphism.
Theorem 14.6.9. If f : Y → X is a projective morphism of noetherian schemes, then f is
proper.
The idea behind the proof of Theorem 14.6.9 is to first prove Pnk → Spec k is proper
for any n, which is a straightforward adaptation of the proof when Pnk is considered as a
projective algebraic variety. One can then modify this proof for PnZ → Spec Z and then use
the properties of proper morphisms in Lemma 14.3.10 to obtain the general result.

334
14.6. The Proj Construction Chapter 14. Scheme Theory

Definition. A morphism Y → X is quasi-projective if it factors into an open immersion


Y → X 0 followed by a projective morphism X 0 → X.

Theorem 14.6.10. Over an algebraically closed field k, there is an equivalence of categories


between algebraic k-varieties and quasi-projective, integral schemes over k. Under this equiv-
alence, projective varieties correspond to projective, integral schemes over k.

Let S be a graded ring and X = Proj S. As with affine schemes, there is a functor
GrModS → ShX which sends a graded S-module M to a sheaf of OX -modules M
f defined in
the same way as in Proposition 14.4.2.
f is a sheaf of OX -module on
Proposition 14.6.11. Let M be a graded S-module. Then M
X = Proj S and

(1) For any p ∈ X, Mfp ∼ = M(p) , where the latter denotes the submodule of degree 0
−1
elements in T M , as above.
f|D+ (f ) ∼
(2) For any f ∈ S h ∩ S+ , M =M(f ) as O(D+ (f ))-modules.
g

(3) M
f is quasi-coherent, and if S is noetherian and M is finitely generated, then M
f is
coherent.

Now let A be a noetherian ring and S = A[x0 , . . . , xn ] so that X = Proj S = PnA is the
projective n-space over A defined above. Recall the “shifted” graded ring S(d) for any d ∈ Z,
defined by S(d)k = Sd+k , which is naturally a graded S-module.

Definition. For d ∈ Z, the dth twisting sheaf of X = PnA is the OX -module OX (d) = S(d).
g
For any OX -module F , write F (d) = F ⊗ OX (d).

Proposition 14.6.12. Suppose S is generated by S1 as an S0 -algebra. Then

(1) For each d ∈ Z, OX (d) is an invertible sheaf on X.

(2) For each j, k ∈ Z, OX (j) ⊗ OX (k) ∼


= OX (j + k).

(3) Let ϕ : S → T be a homomorphism of graded rings, where T is generated by T1


as a T0 -algebra, set Y = Proj T and let U ⊆ Y be the open set and ψ = ϕ∗ :
U → X the morphism defined in Proposition 14.6.4. Then ψ ∗ OX (d) ∼
= OY (d)|U
and ψ∗ (OY (d)|U ) ∼
= (ψ∗ OU )(d).
Proof. Routine – see Hartshorne.

335
Chapter 15

Morphisms of Schemes

336
15.1. Flat Morphisms Chapter 15. Morphisms of Schemes

15.1 Flat Morphisms


Recall from Section 8.3 that a module M over a ring A is flat if the functor M ⊗A − is exact,
and M is faithfully flat if it is flat and M ⊗A N = 0 implies N = 0. We will several different
characterizations of flatness and faithful flatness from previous sections.
Definition. A morphism of schemes f : X → Y is said to be flat at x ∈ X if OX,x is a flat
OY,f (x) -module. We say f is a flat morphism if it is flat at every point x ∈ X.
Definition. A morphism f : X → Y is faithfully flat if it is flat and surjective.
Remark. Note that a faithfully flat morphism is necessarily faithfully flat on stalks, meaning
each OX,x is faithfully flat as an OY,f (x) -module. However, faithful flatness on stalks does
not guarantee surjectivity, so our definition is strictly stronger than the stalkwise definition.
Example 15.1.1. An open immersion is flat but not faithfully flat (unless it is a isomor-
phism). Indeed, open immersions are locally of the form Spec Af ,→ Spec A for a ring A and
an element f ∈ A, and Af is a flat A-module by Example 8.3.8.
Lemma 15.1.2. If f : X → Y and g : Y → Z are (faithfully) flat morphisms, then so is
g ◦ f : X → Z.
Proposition 15.1.3. Suppose A and B are noetherian rings and g : A → B is a ring
homomorphism making B a faithfully flat A-algebra. If b ∈ B is an element such that
b̄ ∈ B/mB is a nonzero divisor for all maximal ideals m ⊂ A, then B/(b) is a flat A-algebra.
Proof. Let m be a maximal ideal of A. Since g : A → B is faithfully flat, there exists a
maximal ideal n ⊂ B retracting to m along g. Then since localization preserves flatness
(Corollary 8.3.22), we may replace A with Am and B with Bn so that g is a morphism of
local rings. Denote the resulting maximal ideals by m ⊂ A and n ⊂ B.
Suppose c ∈ B such that bc = 0. Since b̄ is not a zero divisor in B/mB, we must have
c̄ = 0, i.e. c ∈ mB. We claim c ∈ mn B for all n ≥ 1. To induct, Pr assume c ∈ m B. Let
n
n
{a1 , . . . , ar } be a minimal generating set for m and write c = i=1 ai bi for bi ∈ B. Then
r
X r
X
0 = bc = b ai b i = ai (bbi ).
i=1 i=1
Ps 0 0
It follows P
from Proposition 10.3.20 that bbi = j=1 aij bj for some aij ∈ A and bj ∈ B
satisfying aij ai = 0. Now since {a1 , . . . , ar } is a minimal generating set for mn , we must
have aij ∈ m for each i, j. This shows that bbi ∈ mB, but since b̄ is not T a zero divisor in
B/mB, that means each bi ∈ mB, so c ∈ mn+1 B. By induction, c ∈ ∞ n=1 mn
B but since
g : A → B is a morphism of local rings,

\ ∞
\
mn B = nn = 0
n=1 n=1

by Krull’s intersection theorem. Therefore c = 0, so b is not a zero divisor of B. Repeating


the argument for the ring homomorphism A/I → B/IB where I ⊂ A is any ideal, we see
that b + IB is not a zero divisor in any B/IB. This implies by Proposition 10.3.22 that
TorA
1 (A/I, B/(b)) = 0 for all ideals I ⊂ A, so by Proposition 10.3.20, B/(b) is flat.

337
15.1. Flat Morphisms Chapter 15. Morphisms of Schemes

Definition. Let x, x0 ∈ X be points. We say x0 is a generalization of x, and x is a


specialization of x0 , if x ∈ {x0 }. A point η ∈ X is called a generic point if {η} = X.

Example 15.1.4. For an affine scheme X = Spec A, a prime p is a generalization of q if


and only if p ⊆ q. Thus p is generic if and only if it’s the unique maximal ideal of A.

Proposition 15.1.5. Suppose f : X → Y is a flat morphism of affine schemes and take


x ∈ X with y = f (x) ∈ Y . Then for any generalization y 0 of y, there exists a point
x0 ∈ f −1 (y 0 ) which is a generalization of x.

Proof. The flat condition implies that OX,x is a flat OY,y -module, and by Lemma 8.3.18 it is
even faithfully flat, so by Proposition 8.3.25(c), fx : Spec OX,x → Spec OY,y is surjective.

Definition. Suppose X is a noetherian space. A subset U ⊆ X is locally closed if it is an


intersection of open and closed sets. A set is constructible if it is a union of finitely many
locally closed subsets.

Proposition 15.1.6. Let X be a noetherian space. Then a subset U ⊆ X is constructible


if and only if for every irreducible closed subset Z for which Z ∩ U is dense in Z, the set
Z ∩ U contains an open subset of Z.

Theorem 15.1.7 (Chevalley). If f : X → Y is a finite morphism of noetherian schemes,


then f (X) is constructible.

Proof. Cover Y with affine open sets V1 , . . . , Vr and for each 1 ≤ j ≤ r, cover f −1 (Vj ) with
affine open sets U1j , . . . , Usj j . Then it’s enough to prove each f (Uij ) is constructible in Vj , so
we may reduce to the case of a finite morphism of affine schemes f : Spec B → Spec A. Now
any irreducible closed subset of Spec A is of the form Z = V (p) for some prime ideal p ⊂ A.
Suppose f (Spec B) ∩ V (p) is dense in V (p). Then f (Spec B) ∩ V (p) may be identified with
the image of the morphism

ϕ : Spec(B/pB) −→ Spec(A/p).

Thus we may reduce to the case when A is a domain, so that B is a finitely generated
A-algebra. Then B = A[x1 , . . . , xn , xn+1 , . . . , xm ] where x1 , . . . , xn are algebraically indepen-
dent over A and xn+1 , . . . , xm are algebraic over A[x1 , . . . , xn ]. Thus for each n + 1 ≤ i ≤ m,
there exist polynomials pij ∈ A[x1 , . . . , xn ] such that
d d −1
p0j xj j + p1j xj j + . . . + pdj j = 0

(and p0j 6= 0). Consider the following polynomial with coefficients in A:


m
Y
Q= p0j .
j=n+1

Then for any prime ideal p ⊂ A, the ideal P = p[x1 , . . . , xn ] is prime in A[x1 , . . . , xn ] and
Q 6∈ P. Hence BP is integral over A[x1 , . . . , xn ]P and the result follows from the going up
theorem.

338
15.1. Flat Morphisms Chapter 15. Morphisms of Schemes

Corollary 15.1.8. If f : X → Y is a finite morphism of noetherian schemes and Z is


constructible in X, then f (Z) is constructible in Y .
Proposition 15.1.9. Suppose X is a noetherian space such that every irreducible closed set
in X has a generic point. Let U ⊆ X be a constructible set and x ∈ X. Then U contains
an open neighborhood of x if and only if every generalization of x lies in U .
Theorem 15.1.10. Let X be a noetherian space such that every irreducible closed set in X
has a generic point. Then U ⊆ X is open if and only if the following conditions hold:
(a) For all x ∈ U , every generalization of x lies in U .

(b) For all x ∈ U , U ∩ {x} contains a nonempty open subset of {x}.


Proof. ( =⇒ ) is clear from Propositions 15.1.6 and 15.1.9.
( ⇒= ) We claim U is constructible. Let Z be an irreducible closed set such that Z ∩ U
is dense in Z. Since X is noetherian, Z contains a generic point η, so Z = {η}. Now by
condition (b), Z ∩ U contains a nonempty open subset of Z so Proposition 15.1.6 implies U
is constructible. Hence (a) and Proposition 15.1.9 imply that U is open.
Proposition 15.1.11. If f : X → Y is a finite morphism of noetherian schemes, x ∈ X
and y = f (x) ∈ Y , then the following are equivalent:
(1) f takes every open neighborhood of x to an open neighborhood of y.

(2) For any generalization y 0 of y, there exists x0 ∈ f −1 (y 0 ) which is a generalization of x.


Proof. (1) =⇒ (2) Suppose y ∈ {y 0 }. Take Z to be the union of all irreducible components
of f −1 ({y 0 }) not containing x. Then X r Z is an open neighborhood of x, so f (X r Z) is an
open neighborhood of y. Thus by Proposition 15.1.9, y 0 ∈ f (X r Z). Take x1 ∈ X r Z with
f (x1 ) = y 0 . Then x1 lies in some irreducible component of f −1 ({y 0 }) which also contains x,
say C. Since X is noetherian, C contains a generic point, i.e. C = {x0 } for some x0 ∈ C.
Thus x0 is a generalization of x, so x1 ∈ {x0 }, which implies y 0 = f (x1 ) ∈ {f (x0 )}. But
x0 ∈ C ⊆ f −1 ({y 0 }) implies that f (x0 ) ∈ {y 0 }. Hence y 0 = f (x0 ).
(2) =⇒ (1) Take an open neighborhood U of x. Then U is constructible, so by Corol-
lary 15.1.8, f (U ) is constructible in Y . By Proposition 15.1.9, f (U ) is an open neighborhood
of y = f (x).
Theorem 15.1.12. Let f : X → Y be a finite, flat morphism of noetherian schemes. Then
f is an open map.
Proof. This follows immediately from Propositions 15.1.11 and 15.1.5.
Our main goal in this section is to prove that every finite morphism of noetherian schemes
is generically flat, meaning there exists an open set of the source over which the morphism
is flat. To prove this, we need the following two results in commutative algebra.
Lemma 15.1.13. Let f : A → B be a homomorphism of noetherian rings such that B is
a finitely generated A-algebra, let q ∈ Spec B and p = f −1 (q) ∈ Spec A be prime ideals and
suppose M is a finitely generated B-module such that Mq is a flat Ap -module. Then

339
15.1. Flat Morphisms Chapter 15. Morphisms of Schemes

(a) There exists an element g ∈ B r q such that (M/pM )g is a flat A/p-module and
TorA
1 (M, A/p)g = 0.

(b) For any q0 ∈ D(g) ∩ {q}, Mq0 is a flat A-module.

Theorem 15.1.14 (Generic Flatness). If f : X → Y is a finite morphism of noetherian


schemes, then the set

U = {x ∈ X | OX,x is a flat OY,f (x) -module}

is open in X.

Proof. It is enough to prove the theorem when X and Y are affine, say X = Spec B and
Y = Spec A. Note that Corollary 8.3.22 implies the set U contains the generalizations of all
of its points. Also, it follows from Lemma 15.1.13(b) that for all x ∈ U , the set U ∩ {x}
contains an open set of {x}. Hence the conditions in Theorem 15.1.10 are satisfied, so U is
open.

340
15.2. Kähler Differentials Chapter 15. Morphisms of Schemes

15.2 Kähler Differentials


In many areas of math, especially complex and differential geometry, smooth maps are
essential tools of study. To define smoothness for schemes, we first introduce a module
construction developed in commutative algebra to parallel the classical theory of differential
forms. (Note that some of the definitions and results in this and the following section were
first introduced in Section 14.4. Here, we aim to give a full account of relative differentials
and smooth morphisms.)
Definition. For a commutative ring A, an A-algebra B and a B-module M , an A-derivation
of B into M is an A-linear homomorphism d : B → M that satisfies the Leibniz rule: for
all b, b0 ∈ B, d(bb0 ) = b(db0 ) + (db)b0 . Let DerA (B, M ) denote the set of all A-derivations
B → M.
Lemma 15.2.1. If d : B → M is an A-derivation, then for all a ∈ A, da = 0.
Proof. Here, da denotes d(a1B ) so the Leibniz rule for this expression is:
da = d(a1B ) = (da)1B + a(d1B ) = da + da
since d is A-linear. Thus da = 2da, or da = 0.
Definition. For an A-algebra B, a module of relative Kähler differentials (usually just
relative differentials) of B is a B-module Ω1B/A together with an A-derivation d : B → Ω1B/A
which are universal with respect to A-derivations. That is, for any A-derivation e : B → M ,
there exists a unique B-module homomorphism g : Ω1B/A → M making the following diagram
commute:
e
B M

d
g
Ω1B/A

As with any universal object, if it exists, Ω1B/A is unique up to isomorphism. For its
existence, we have:
Proposition 15.2.2. For any A-algebra B, the module of relative differentials of B may be
defined by
Ω1B/A = Zhdb | b ∈ Bi/N
where N is the submodule generated by elements of the form da for a ∈ A and d(b+b0 )−db−db0
and d(bb0 ) − b(db0 ) − (db)b0 for b, b0 ∈ B.
Proof. The map d : B → Ω1B/A is simply b 7→ db + N , which is a derivation by definition of
N . Suppose e : B → M is an A-derivation. Then there is a map g : Ω1B/A → M obtained
P P
by setting g ( rb db + N ) = rb e(b). Note that g(N ) = 0 since e is a derivation, so g is
well-defined. Moreover, the diagram for the universal property of differentials commutes by
construction of g, and uniqueness is proven as usual.

341
15.2. Kähler Differentials Chapter 15. Morphisms of Schemes

The universal property of Ω1B/A implies the following nice fact.


Lemma 15.2.3. For any A-algebra B and B-module M , there is an isomorphism

HomB (Ω1B/A , M ) −→ DerA (B, M )
ϕ 7−→ ϕ ◦ d

which is natural in M .
Example 15.2.4. If B = A[x1 , . . . , xn ], then Ω1B/A is a free B-module generated by {dx1 , . . . , dxn }
and the derivation d : B → Ω1B/A is given by
n
X ∂f
df = dxi
i=1
∂x i

∂f
where ∂xi
denote the formal partial derivatives of the polynomial f .
Lemma 15.2.5. Suppose A is a commutative ring. Then
(a) If I ⊂ A is an ideal, then Ω1(A/I)/A = 0.

(b) For any multiplicative subset S ⊂ A, Ω1S −1 A/A = 0.

Proof. (a) For any a + I ∈ A/I, d(a + I) = ad(1 + I) since d is A-linear, and as we saw in
the proof of Lemma 15.2.1, d(1 + I) = 0.
(b) For any b ∈ S −1 A, there is some s ∈ S so that sb ∈ A. Then by Lemma 15.2.1,
0 = d(sb) = s(db) + (ds)b = s(db) since s ∈ S ⊂ A. But since s 6= 0, we must have
db = 0.
Now let ϕ : B → C be an A-algebra homomorphism. This induces a homomorphism of
C-modules

dϕ : Ω1B/A ⊗B C −→ Ω1C/A
db ⊗ c 7−→ cd(ϕ(b)).

On the other hand, by Lemma 15.2.3, the canonical derivation C → Ω1C/A induces a C-
module map βϕ : Ω1C/A → Ω1C/B which is just dc 7→ dc. (One can define βϕ directly this
way and then check it is well-defined and preserves the C-module structure, but there is a
universal property for a reason.)
Proposition 15.2.6. Let B be an A-algebra. Then
(a) For any other A-algebra A0 , set B 0 = B ⊗A A0 . Then there is a natural isomorphism
of B 0 -modules Ω1B 0 /A0 ∼
= Ω1B/A ⊗B B 0 .

(b) For any A-algebra homomorphism ϕ : B → C, there is an exact sequence of C-modules


dϕ βϕ
Ω1B/A ⊗B C −→ Ω1C/A −→ Ω1C/B → 0.

342
15.2. Kähler Differentials Chapter 15. Morphisms of Schemes

(c) For any multiplicative subset S ⊂ B, there is an isomorphism of S −1 B-modules


S −1 Ω1B/A ∼
= Ω1S −1 B/A .

(d) Suppose C = B/I for an ideal I ⊂ B. Then there is an exact sequence of C-modules
δ dϕ
I/I 2 →
− Ω1B/A ⊗B C −→ Ω1C/A → 0

where ϕ is the quotient map B → C and δ is defined by x + I 2 7→ dx ⊗ 1.


Proof. (a) The derivation d : B → Ω1B/A induces a map d0 = d ⊗ idA0 : B 0 → Ω1B/A ⊗A A0 ∼
=
1 0
ΩB/A ⊗B B and one can check that it satisfies the same universal property as the canonical
derivation B 0 → Ω1B 0 /A0 . Therefore Ω1B 0 /A0 ∼
= Ω1B/A ⊗B B 0 .
(b) By Yoneda’s Lemma (7.1.6), it’s equivalent to show

0 → HomC (Ω1C/B , N ) → HomC (Ω1C/A , N ) → HomC (Ω1B/A ⊗B C, N )

is exact for an arbitrary C-module N . By Hom-tensor adjointness, HomC (Ω1B/A ⊗B C, N ) ∼


=
1
HomB (ΩB/A , N ) so Lemma 15.2.3 shows that the sequence in question is really

0 → DerB (C, N ) → DerA (C, N ) → DerA (B, N )

which is exact by the definition of derivations.


(c) By Lemma 15.2.5(b), Ω1S −1 B/B = 0, so the sequence in (b) becomes a surjection

S −1 Ω1B/A → Ω1S −1 B/A → 0.

The map here sends dbs ∈ S −1 Ω1B/A to 1s ·d 1b which is clearly injective, hence an isomorphism.


(d) Again by Yoneda’s Lemma and Lemma 15.2.3, it’s enough to show

0 → DerA (B/I, N ) → DerA (B, N ) → HomB/I (I/I 2 , N )

is exact for any B/I-module N . Note that HomB/I (I/I 2 , N ) ∼


= HomB (I, N ) by the universal
property of quotients, and the map from DerA (B, N ) sends d : B → N to d|I : I → N . It is
clear that the quotient of this map is DerA (B/I, N ), so the sequence is exact.
Corollary 15.2.7. If B is a finitely generated A-algebra or a localization of such an algebra,
then Ω1B/A is finitely generated as a B-module.
Proof. Given a presentation
n
ϕ
M
0→I→
− A−
→B→0
i=1
Ln
set F = i=1 A and apply Proposition 15.2.6(d) to the quotient map ϕ : F → B to get an
exact sequence of B-modules

I/I 2 → Ω1F/A ⊗F B −→ Ω1B/A → 0.

Then by definition Ω1F/A is finitely generated as an A-module, so Ω1F/A ⊗F B is finitely


generated as a B-module and this implies Ω1B/A is as well.

343
15.2. Kähler Differentials Chapter 15. Morphisms of Schemes

Example 15.2.8. Suppose B = A[x1 , . . . , xn ] is a polynomial ring over A, take f ∈ B and


set C = B/(f ). Then by Proposition 15.2.6(d) there is an exact sequence

(f )/(f 2 ) → Ω1B/A ⊗B C → Ω1C/A → 0

so we can see that !


n
M
Ω1C/A = Chdxi i /Chdf i
i=1

where df denotes the formal differential


n
X ∂f
df = dxi .
i=1
∂x i

More generally, if C = B/(f1 , . . . , fm ) then


Ln
Chdxi i
Ω1C/A = i=1
.
Chdf1 , . . . , dfm i
Lemma 15.2.9. Let R and S be A-algebras. Then
(a) If B = R ⊗A S then there is a canonical isomorphism of B-modules

(Ω1R/A ⊗R B) ⊕ (Ω1S/A ⊗S B) −→ Ω1B/A
(dr ⊗ b1 , ds ⊗ b2 ) 7−→ b1 d(r ⊗ 1) + b2 d(1 ⊗ s).

(b) If S = R[x1 , . . . , xn ]/I for an ideal I, let ρ : R → S be the restriction of the quotient
map to R ⊆ R[x1 , . . . , xn ] and let α = dρ : Ω1R/A ⊗R S → Ω1S/A and δ : I/I 2 →
Ω1R[x1 ,...,xn ]/A ⊗R[x1 ,...,xn ] S be the maps defined above. Then there is a surjective B-
module homomorphism ker δ → ker α.
= Ω1S/A ⊗S B and Ω1B/S ∼
Proof. (a) From Proposition 15.2.6(a), Ω1B/R ∼ = Ω1R/A ⊗R B, so (b) of
the same proposition gives us exact sequences
ϕ
Ω1R/A ⊗R B −
→ Ω1B/A → Ω1S/A ⊗S B (15.1)
ψ
and Ω1S/A ⊗S B −
→ Ω1B/A → Ω1R/A ⊗R B. (15.2)

Then the desired map is ϕ ⊕ ψ which is of the desired form by construction and is an
isomorphism because the last maps in (1) and (2) are sections of ψ and ϕ, respectively.
(b) Set A0 = A[x1 , . . . , xn ] and R0 = R[x1 , . . . , xn ]. Then the fact that R0 = A0 ⊗A R
implies by (a) that

Ω1R0 /A ∼
= (Ω1A0 /A ⊗A0 R0 ) ⊕ (Ω1R/A ⊗R R0 ) = Ω1R0 /R ⊕ (Ω1R/A ⊗R R0 ).

Applying − ⊗R0 S yields an isomorphism

Ω1R0 /A ⊗R0 S ∼
= (Ω1R0 /R ⊗R0 S) ⊕ (Ω1R/A ⊗R S) (15.3)

which fits into a commutative diagram

344
15.2. Kähler Differentials Chapter 15. Morphisms of Schemes

δ0
I/I 2 Ω1R0 /A ⊗R0 S Ω1S/A 0

δ ∼
= α
p i
0 Ω1R0 /R ⊗R0 S (Ω1R0 /R ⊗R0 S) ⊕ (Ω1R/A ⊗R S) Ω1R/A ⊗R S 0

where the top row is the exact sequence from Proposition 15.2.6(d) and the bottom row is
the canonical split sequence for the direct sum. A quick diagram chase reveals there is a
surjection ker δ → ker α.
We next study relative differentials for field extensions. Let k be a field.
Lemma 15.2.10. Let E/k be an arbitrary field extension and K = E[x]/(p(x)) a simple
algebraic extension of E. Then
(a) If K/E is separable, then Ω1K/E = 0 and there is an isomorphism of K-vector spaces
Ω1E/k ⊗E K ∼
= Ω1K/k .

(b) If K/E is inseparable, then Ω1K/E ∼


= K and dimE Ω1E/k ≤ dimK Ω1K/k ≤ dimE Ω1E/k + 1.

Proof. (a) By Example 15.2.8,

Ω1K/E = Khdxi/(p0 (x) dx) ∼


= K/(p0 (x))

so if the extension is separable, p0 (x) is relatively prime to p(x). Thus K/(p0 (x)) = 0. The
isomorphism Ω1E/k ⊗E K ∼ = Ω1K/k then follows from Proposition 15.2.6(d).
(b) When K/E is inseparable, p0 (x) = 0 so Ω1K/E = Khdxi/(p0 (x) dx) ∼ = K. From the
isomorphism (3) in the proof of Lemma 15.2.9, we obtain

dimK (Ω1E[x]/k ⊗E[x] K) = dimK Khdxi + dimK (Ω1E/k ⊗E K) = 1 + dimE Ω1E/k

from which the inequalities in (b) are easily obtained.


Corollary 15.2.11. Suppose K/k is a finite extension. Then K/k is separable if and only
if Ω1K/k = 0.

Proof. ( =⇒ ) is immediate from Lemma 15.2.10(a) since a separable extension is always


simple.
( ⇒= ) Suppose K/k is inseparable. Then there is a tower of field extensions K ⊇ E ⊇ k
where K/E is simple and inseparable so by Lemma 15.2.10(b), Ω1K/E ∼ = K. However, by
1 1 1
Proposition 15.2.6 there is a surjection ΩK/k → ΩE/k so ΩK/k 6= 0.

Definition. A function field K/k is said to be separable if it has a purely transcendental


subextension E/k such that K/E is finite and separable.
Proposition 15.2.12. Suppose K is a function field over k of transcendence degree n. Then
Ω1K/k is a finite dimensional K-vector space of dimension dimK Ω1K/k ≥ n, with equality if
and only if K/k is separable.

345
15.2. Kähler Differentials Chapter 15. Morphisms of Schemes

Proof. By definition K is a finite extension of L = k(x1 , . . . , xn ). Moreover, since L =


Frac(k[x1 , . . . , kn ]), Example 15.2.4 and Proposition 15.2.6(c) imply dimL Ω1L/k = n. Thus
Lemma 15.2.10 gives n ≤ dimK Ω1K/k < ∞. If K/k is separable, then K/L is separable and
a simple extension, and Lemma 15.2.10(a) gives dimK Ω1K/k = dimL Ω1L/k = n. Conversely if
dimK Ω1K/k = n, we may take α1 , . . . , αn ∈ K such that Ω1K/k = SpanK {dα1 , . . . , dαn }. Then
E = k(α1 , . . . , αn ) defines a subextension of K/k and the exact sequence

Ω1E/k ⊗E K → Ω1K/k → Ω1K/E → 0

from Proposition 15.2.6(b) implies Ω1K/E = 0 since the first arrow is an isomorphism. By
Lemma 15.2.10, K/E must be separable and hence so is K/k.

346
15.3. Sheaves of Relative Differentials Chapter 15. Morphisms of Schemes

15.3 Sheaves of Relative Differentials


Sheafifying the construction of Ω1B/A gives a scheme-theoretic version of differential forms
that serves to define smooth morphisms in the category of schemes. We begin with the affine
construction.

Definition. Let A → B be a ring homomorphism with corresponding morphism of affine


schemes f : Spec B → Spec A. Then the sheaf of relative differentials for f is the sheaf
Ω1Spec B/ Spec A = Ω
e 1 on Spec B.
B/A

Lemma 15.3.1. Let f : X = Spec B → Y = Spec A be a morphism of affine schemes. Then

(a) Ω1X/Y is a sheaf of OX -modules on X.

(b) Ω1X/Y is quasi-coherent.

(c) For any x ∈ B, Ω1X/Y (D(x)) ∼


= Ω1Bx /A .

Proof. (a) and (b) follow from the construction M 7→ M


f in Section 14.4, while (c) is an
application of Proposition 15.2.6(c).
To extend this construction to all schemes, we first provide an alternate definition of
Ω1B/A . Let m : B ⊗A B → B be the multiplication map m(b ⊗ b0 ) = bb0 . Set I = ker m.

Proposition 15.3.2. The map

d : B −→ I/I 2
b 7−→ 1 ⊗ b − b ⊗ 1

is an A-derivation inducing an isomorphism of B-modules I/I 2 ∼


= Ω1B/A .

Proof. Suppose e : B → M is an A-derivation. Then the map ϕ : B⊗A B → M, b⊗b0 7→ be(b0 )


is a homomorphism of B-modules and we have

ϕ((1 ⊗ b − b ⊗ 1)(1 ⊗ b0 − b0 ⊗ 1)) = ϕ(1 ⊗ bb0 − b0 ⊗ b − b ⊗ b0 − bb0 ⊗ 1)


= ϕ(1 ⊗ bb0 ) − ϕ(b0 ⊗ b) − ϕ(b ⊗ b0 ) − ϕ(bb0 ⊗ 1)
= e(bb0 ) − b0 e(b) − be(b0 ) − bb0 e(1) = 0

since e is a derivation. One can show that I 2 is generated as a B-module by elements of


the form (1 ⊗ b − b ⊗ 1)(1 ⊗ b0 − b0 ⊗ 1) for b, b0 ∈ B, so this shows ϕ induces a B-module
homomorphism g : I/I 2 → M making the diagram
e
B M

d
g
2
I/I

347
15.3. Sheaves of Relative Differentials Chapter 15. Morphisms of Schemes

commute. Uniqueness follows from a similar proof using the explicit description of I/I 2 .
Hence I/I 2 is universal with respect to A-derivations, so it is isomorphic to Ω1B/A .
Now let f : X → Y be a morphism of schemes. Then the diagonal morphism ∆ : X →
X ×Y X maps X to a closed subscheme of an open set U ⊆ X ×Y X. Let I be the ideal
sheaf of ∆(X) in OU .

Definition. The sheaf of relative differentials for f : X → Y is the pullback sheaf

Ω1X/Y := ∆∗ (I/I 2 ).

Example 15.3.3. Let A be a ring, B an A-algebra, X = Spec B, Y = Spec A and consider


the structure morphism f : X → Y . Then it follows from Proposition 15.3.2 that Ω1X/Y =
e 1 , in agreement with our affine definition of Ω1
Ω above. More generally:
B/A X/Y

Proposition 15.3.4. Suppose f : X → Y is a morphism of schemes. Then for any open


sets V ⊆ Y and U ⊆ f −1 (V ) ⊆ X, there is an isomorphism

Ω1X/Y |U ∼
=Ωe1
OX (U )/OY (V )

of sheaves on U . In particular, for all x ∈ X, there is an isomorphism of stalks

Ω1X/Y,x ∼
= Ω1OX,x /OY,f (x) .

Corollary 15.3.5. For all schemes X → Y , Ω1X/Y is a quasi-coherent OX -module.

Example 15.3.6. Let A be a ring, put Y = Spec A and let X = AnY be affine n-space over
n
A. Then by Example 15.2.4, we have Ω1X/A = OX .

Example 15.3.7. With Y = Spec A again, let X = PnY be projective n-space over A. Let us
compute Ω1X/Y carefully. Recall that X has an affine cover {U0 , U1 } where U0 = Spec A[t] and
U1 = Spec A[t−1 ]. Then Ω1U0 /A is the sheafification of the A-module A[t]dt; likewise, Ω1U1 /A
is the sheafification of the A-module A[t−1 ]d 1t = A[t−1 ] −1 dt. We claim that Ω1X/Y ∼

t2 =
O(−2), the twisting sheaf of degree −2 – recall that this is defined as O(−2) = M f where
M = A[t](−2) is the −2 shift of the graded ring A[t]. On the open sets U0 and U1 , we have
isomorphisms

ϕ0 : Ω1U0 /A −→ M
f|U0
dt 7−→ 1
and ϕ1 : Ω1U1 /A −→ M f|U1
 
1 1
d 7−→ − 2
t t

So Ω1X/A is a line bundle on X (a locally free sheaf of rank 1). Moreover, on the overlap
U0 ∩ U1 = Spec A[t, t−1 ], we have d 1t = − t12 dt, so we see that ϕ0 and ϕ1 determine a global


isomorphism Ω1X/A ∼ = O(−2) as claimed.

348
15.3. Sheaves of Relative Differentials Chapter 15. Morphisms of Schemes

Example 15.3.8. Let X be an affine curve defined over a field k by an equation F (x, y) = 0
for F ∈ k[x, y] such that at any point on X, the formal partial derivatives ∂F ∂x
and ∂F
∂y
do not simultaneously vanish. Such a curve is said to be smooth in the classical sense,
though we will soon define smoothness for general schemes and recover this very definition.
Since X = Spec(k[x, y]/(F )) is affine, Ω1X/k is the sheafification of the module Ω1B/k where
B = k[x, y]/(F ), and Example 15.2.8 shows that

Ω1B/k = (Bhdxi ⊕ Bhdyi)/hdF i.

In particular, this means that ∂F


∂x
dx = − ∂F
∂y
dy but since they do not vanish simultaneously,
∂F ∂F
∂x
and ∂y generate the unit ideal in B. Thus

Ω1B/k = (Bhdxi ⊕ Bhdyi)/hdF i ∼


=B

so Ω1X/k is isomorphic to OX = B.
e That is, Ω1 is a free OX -module of rank 1 with basis
X/k
n o n o
1 1
∂F/∂x
dy or ∂F/∂y dx .

Proposition 15.3.9. Let f : X → Y be a morphism of schemes. Then

(a) For any Y -scheme Y 0 , set X 0 = X ×Y Y 0 . Then there is an isomorphism Ω1X 0 /Y 0 ∼


=
∗ 1 0 0
p ΩX/Y of sheaves on X , where p : X → X is the canonical projection.

(b) For any morphism of schemes Y → Z, there is an exact sequence of sheaves on X,

f ∗ Ω1Y /Z → Ω1X/Z → Ω1X/Y → 0.

(c) For any open set U ⊆ X, there is an isomorphism Ω1X/Y |U ∼ = Ω1U/Y of sheaves on U .
In particular, for any x ∈ X, Ω1X/Y,x ∼
= Ω1OX,x /OY,f (x) .

(d) Let Z be a closed subscheme of X with ideal sheaf I. Then there is an exact sequence
of sheaves on Z,
I/I 2 → Ω1X/Y ⊗OX OZ → Ω1Z/Y → 0.

Proof. This is the global version of Proposition 15.2.6.

Example 15.3.10. Let E be an elliptic curve over a field k, i.e. a nonsingular projective
curve in P2k with distinguished point O ∈ E(k) and affine Weierstrass equation

y 2 + a1 xy + a3 y = x3 + a2 x2 + a4 x + a6 for ai ∈ k.

Then ω = 2y+adx1 x+a3


is a differential form on E that generates Ω1k(E)/k , so Ω1E/k is a free
OE -module of rank 1 generated by ω.

349
15.4. Smooth Morphisms Chapter 15. Morphisms of Schemes

15.4 Smooth Morphisms


Let k be a field. We can use ring theory to adapt the notion of smoothness from differential
geometry to the setting of k-varieties, and ultimately arbitrary schemes. Although our first
definition of smoothness won’t be linked to Kähler differentials, we will later see that the
property of a morphism X → Y being smooth can be phrased in terms of the sheaf Ω1X/Y .
This will agree with the perspective that Kähler differentials are an abstraction of differential
forms to ring theory.
Recall (Section 6.2) that a local noetherian ring (A, m, k) is regular if dim A = dimk m/m2 .
In general, we have dimk m/m2 by Example 6.2.7, and A is regular of dimension n if and
only if m can be generated by n elements (Theorem 6.2.8). Moreover, by Lemma 6.2.9 a
regular local ring is an integrally closed integral domain.

Definition. A locally noetherian scheme X is regular at a point x ∈ X if OX,x is a


regular local ring, i.e. if dim OX,x = dimk(x) Tx X, where Tx X = (mx /m2x )∗ is the tangent
space of X at x. We say X is regular if it is regular at every x ∈ X.

Note that a locally noetherian scheme is regular if and only if it is regular at all of its
closed points.

Example 15.4.1. Let X = Ank = Spec k[x1 , . . . , xn ] be affine n-space over a field k. Since X
is an affine variety, for any closed point x ∈ X we have dim OX,x = dim X = n (Lemma 2.6.5,
Corollary 2.6.3 and Proposition 14.1.5), and if x = (α1 , . . . , αn ) then the maximal ideal in
OX,x is mx = (x1 − α1 , . . . , xn − αn ). Thus mx is generated by n elements, so OX,x is regular.

Example 15.4.2. Likewise, Pnk is regular at all of its (closed) points since each one is
contained in an affine open subscheme isomorphic to Ank .

Example 15.4.3. Let char k 6= 2, 3 and consider the affine curve X ⊆ A2k defined by

X = Spec B = Spec(k[x, y]/(x2 − y 3 )).

P = (0, 0)

Then B ∼ = k[t3 , t2 ] and it’s easy to compute the dimensions of tangent spaces by localization
and see that X is regular at all P 6= (0, 0) and singular at P = (0, 0), but here’s a way of
seeing it using the results of the preceding sections. The local ring at any P = (x0 , y0 ) is

OX,P = BmP where mP = (x − x0 , y − y0 ). Then TX,P = mP /m2P has dimension equal to

350
15.4. Smooth Morphisms Chapter 15. Morphisms of Schemes

the rank of Ω1X/k,P ∼


= Ω1B/k,mP as an OX,P -module by Proposition 15.3.4, and as we saw in
Example 15.2.8,
Bdx ⊕ Bdy
Ω1B/k = .
hd(x2 − y 3 )i
Now d(x2 − y 3 ) = 2x dx − 3y 2 dy is 0 when P = (0, 0), so dim TX,P = 2 6= 1 = dim OX,P in
this case, so X is singular at P = (0, 0). On the other hand, if P = (x0 , y0 ) 6= (0, 0), Ω1B/k,mP
dx dy
is a free BmP -module of rank 1 generated by either 3y 2 if y0 6= 0 or by 2x if x0 6= 0. Thus X

is regular on the locus {P 6= (0, 0)}.

The last example is also easy to verify using the Jacobian condition for regularity, which
we recall now. Let X ⊆ Ank be an (irreducible) affine variety with vanishing ideal I ⊆
k[x1 , . . . , xn ]. Then I = (f1 , . . . , fm ) for some polynomials fj ∈ k[x1 , . . . , xn ]. For each point
x ∈ X, the Jacobian of X at x is the m × n matrix
 
∂fi
Jx = (x) ∈ Mm×n (k).
∂xj

Theorem 15.4.4 (Jacobian Condition, Theorem 12.6.2). An affine variety X ⊆ Ank is


regular at x ∈ X if and only if rank Jx = n − dim X.

Corollary 15.4.5. The rank of the Jacobian matrix of X at x is independent of the choice
of generators f1 , . . . , fm for the vanishing ideal of X.

Definition. Let Y be a locally noetherian scheme. A morphism f : X → Y is smooth at


a point x ∈ X if the following conditions hold:

(i) f is of finite type at x.

(ii) f is flat at x.

(iii) If y = f (x), then the fibre Xy := X ×Y k(y) is regular at x.

Otherwise f is singular at x. We say f is a smooth morphism if it is smooth at ev-


ery point x ∈ X, and call X a smooth Y -scheme. Finally, f is smooth of relative
dimension n if f is smooth and for each y in the image of f , dim Xy = n.

Denote by Xsm the set of smooth points of a Y -scheme X → Y and by Xsing the set of
singular points of X. Note that in general, the fibres of a morphism need not have the same
dimension.

Remark. To define smoothness for an arbitrary morphism of schemes, one replaces condition
(i) with “f is locally of finite presentation at x”. When Y is locally noetherian, this condition
is equivalent to being locally of finite type. We will assume for the rest of the section that
Y is always locally noetherian.

Theorem 15.4.6. If f : X → Y is a smooth morphism and Y is regular, then so is X.

351
15.4. Smooth Morphisms Chapter 15. Morphisms of Schemes

Proof. Take x ∈ X and set y = f (x) ∈ Y , m = dim OX,x and n = dim OY,y . Since f is flat,
Theorem 8.3.30 says that

dim OXy ,x = dim OX,x − dim OY,y = m − n,

so by smoothness, the maximal ideal of OXy ,x is generated by r = m − n elements, say


α1 , . . . , αr . On the other hand, OXy ,x = OX,x /my OX,x where my is the maximal ideal of
OY,y . So each αi = āi for some ai ∈ OX,x . Let fx# : OY,y → OX,x be the morphism of local
rings induced by f at x. Then by hypothesis, my is generated by some b1 , . . . , bn , and the
definition of the αi shows that the set

{a1 , . . . , ar , fx# (b1 ), . . . , fx# (bn )}

generates mx . Therefore OX,x is regular as desired.

Proposition 15.4.7. Suppose f : X → Y is a smooth morphism. Then

(a) If Y 0 → Y is any morphism, then the base change X 0 := X ×Y Y 0 → Y 0 is also smooth.

(b) If g : Y → Z is smooth, then so is the composition g ◦ f : X → Z.

Proof. These are mostly straightforward from the definitions. For (a), note that finite gener-
ation and flatness for rings are preserved under tensor product (the latter by Lemma 8.3.21).
Moreover, the fibre of a base change is precisely the original fibre, so all parts of the definition
of smoothness hold for X 0 → Y 0 .
For (b), finite type is a property preserved in towers of rings. Moreover, flatness is
preserved under composition by Lemma 15.1.2. To prove the regularity condition for g ◦ f ,
set y = f (x) and z = g(y). Then

Xz = X ×Z k(z) = (X ×Y Y ) ×Z k(z) = X ×Y (Y ×Z k(z)) = X ×Y Yz .

Since g is smooth, Yz is regular but then by Theorem 15.4.6, Xz is also regular.

Example 15.4.8. For a scheme Y , one can define affine n-space over Y to be the affine
scheme
AnY := Spec(OY (Y )[x1 , . . . , xn ]).
Then the ring map OY (Y ) → OY (Y )[x1 , . . . , xn ] induces a natural projection morphism
π : AnY → Y . We claim π is always smooth of relative dimension n. By Hilbert’s Basis
Theorem, π is of finite type. Flatness is trivial. Finally, each fibre of π is Ank(y) where k(y)
is the residue field at y ∈ Y , and these are regular by Example 15.4.1 and have dimension
n. Thus π is smooth of relative dimension n.

Definition. A smooth morphism of relative dimension 0 is called an étale morphism.

The next result says that a smooth Y -scheme is étale-locally a subscheme in AnY . The
notion of ‘étale-locally’ can be made precise using Grothendieck topologies (namely, the étale
topology on Y ), which will be introduced in Section 33.1.

352
15.4. Smooth Morphisms Chapter 15. Morphisms of Schemes

Proposition 15.4.9. Let f : X → Y be a flat morphism of finite type. Then f is smooth at


x ∈ X if and only if there exists an open neighborhood U ⊆ X of x and an étale morphism
g : U → AnY for some n such that f |U = π ◦ g : U → AnY → Y , where π is the canonical
projection.
Proof. ( ⇒= ) is immediate, since g is smooth by definition, π is smooth by Example 15.4.1
and thus their composition is smooth by Proposition 15.4.7(b).
( =⇒ ) Assume f is smooth at x ∈ X and set y = f (x). By definition, Xy is a regular
k(y)-scheme at x, say of dimension n, so after passing to an affine neighborhood of x in
X, we may assume there is an étale morphism Xy → Ank(y) of k(y)-schemes. Indeed, since
Xy is regular, the maximal ideal of OXy ,x = OX,x /my OX,x is generated by n elements, say
t1 , . . . , tn . Write Ank(y) = Spec(k(y)[x1 , . . . , xn ]). Then there is a map

k(y)[x1 , . . . , xn ] = Γ(Ank(y) , OAnk(y) ) −→ Γ(Xy , OXy )


xi 7−→ ti .

The corresponding morphism Xy → Ank(y) is smooth of relative dimension 0 by construction.


Now Γ(X, OX ) → Γ(Xy , OXy ) need not be surjective, but we can find a small enough open
neighborhood U ⊆ X of x so that t1 , . . . , tn lie in the image of Γ(U, OU ) → Γ(Xy , OXy ), say
they are the images of t01 , . . . , t0n . Then there is a map of OY -algebras

OY [x1 , . . . , xn ] −→ f∗ OU

induced by sending xi 7→ f∗ (t0i ) which in turn determines a morphism

g : U −→ Spec(OY (Y )[x1 , . . . , xn ]) = AnY .

By construction, f |U = π ◦ g : U → AnY → Y . Now π is smooth of relative dimension n


by Example 15.4.8 and fibre dimension adds along compositions, so it remains to show g is
smooth at x. It is clear that g is of finite type, and since each Xy → Ank(y) is flat, so is g.
For regularity, use the fact that Xy → Ank(y) is smooth together with Theorem 15.4.6.
Remark. The proof above shows that the n in g : U → AnY may be chosen to be the
dimension of the fibre Xy . In the language of complex geometry, Proposition 15.4.9 says
that (étale-locally) every smooth morphism is a local submersion.
Corollary 15.4.10. Let f : X → Y be a smooth morphism of relative dimension n. Then
(a) Ω1X/Y is locally free of rank n.

(b) For any morphism Y → Z, there is a short exact sequence of sheaves

0 → f ∗ Ω1Y /Z → Ω1X/Z → Ω1X/Y → 0.

Proof. By Proposition 15.4.9 we may assume f = π ◦ g where g : X → AnY is étale and


π : AnY → Y is the projection. From Proposition 15.3.9(b), we have an exact sequence

g ∗ Ω1AnY /Y → Ω1X/Y → Ω1X/AnY → 0.

353
15.4. Smooth Morphisms Chapter 15. Morphisms of Schemes

If (b) holds for the map g, we could extend this with a zero on the left and, since rank
is additive along short exact sequences, this would imply rank Ω1X/Y = rank g ∗ Ω1An /Y +
Y
rank Ω1X/An . Further, Lemma 15.4.11(a) below shows that rank Ω1An /Y = n and, since g ∗ is
Y Y
exact (Corollary 16.1.13), rank g ∗ Ω1An /Y = n as well. Then if (a) holds for g, rank Ω1X/An = 0
Y Y
so we would get rank Ω1X/Y = n. Moreover, consider the commutative diagram

0 f ∗ Ω1Y /Z Ω1X/Z Ω1X/Y 0

0 g ∗ π ∗ Ω1Y /Z g ∗ Ω1An /Y g ∗ Ω1An /Y 0


Y Y

f
The top row is our desired sequence for X →− Y → Z. The bottom row comes from applying
g ∗ to the sequence
0 → π ∗ Ω1Y /Z → Ω1AnY /Y → Ω1AnY /Y → 0.
We claim that it is enough to prove that (a) and (b) hold for π and g. Indeed, if they hold
for g then the short exact sequences of relative differentials associated to the compositions
g g
X → − AnY → Z and X → − AnY → Y become isomorphisms g ∗ Ω1An /Z ∼ = Ω1X/Z and g ∗ Ω1AnY /Y ∼
=
Y
1
ΩX/Y so the middle and right columns of the diagram are isomorphisms. Then (b) for π
along with exactness of g ∗ imply the bottom row is exact, so we can deduce exactness of
the top row. To summarize, the Corollary follows from establishing (a) and (b) for g and π.
This is done in the following lemma.
Lemma 15.4.11. Let Y → Z be a morphism of locally noetherian schemes. Then
(a) For every n ≥ 1, let π : AnY → Y be the canonical projection. Then Ω1An /Y is a free
Y
OY -module of rank n and there is a short exact sequence
0 → π ∗ Ω1Y /Z → Ω1AnY /Z → Ω1AnY /Y → 0.

(b) If g : X → Y is étale then g ∗ Ω1Y /Z ∼


= Ω1X/Z .
Proof. (a) Example 15.2.4 implies that Ω1An /Y is free of rank n. The second statement can
Y
be deduced from the local property that the sequence
0 → Ω1B/A ⊗B B[x1 , . . . , xn ] → Ω1B[x1 ,...,xn ]/A → Ω1B[x1 ,...,xn ]/B → 0
from Proposition 15.2.6(b) is in fact split exact when C = B[x1 , . . . , xn ]. This is obvious
from the definition of the symbols df .
(b) Consider the commutative diagram
∆X q
X X ×Y X X ×Z X

g×g
g
∆Y
Y Y ×Z Y

354
15.4. Smooth Morphisms Chapter 15. Morphisms of Schemes

where ∆X and ∆Y are the diagonal maps. Let I be the OY ×Z Y -ideal corresponding to
the image of ∆Y , so that Ω1Y /Z = ∆∗Y (I/I 2 ). Then since g and therefore g × g is flat,
J := (g × g)∗ I is the OX×Z X -ideal corresponding to the image of q. We will see that since
g is étale, ∆X is an open immersion. Assuming this, Ω1X/Z ∼
= (q ◦ ∆X )∗ (J /J 2 ) and we have

Ω1X/Z ∼
= (q ◦ ∆X )∗ (J /J 2 ) = ∆∗X ◦ q ∗ (J /J 2 )
= ∆∗X ◦ q ∗ ◦ (g × g)∗ (I/I 2 )

= g ∗ ◦ ∆∗ (I/I 2 ) by commutativity
Y
∗ 1
=g ΩY /Z by definition.

Therefore g ∗ Ω1Y /Z ∼
= Ω1X/Z .
In fact, the converse of (a) is also true. We omit the proof.
Theorem 15.4.12. Let f : X → Y be a flat morphism of finite type with fibres of dimension
n. Then f is smooth if and only if Ω1X/Y is locally free of rank n.
There are two more important properties of smoothness to discuss. The first says that
every morphism is generically smooth, generalizing Theorem 12.6.5.
Theorem 15.4.13 (Generic Smoothness). For a morphism of finite type f : X → Y between
locally noetherian schemes, the smooth locus Xsm is a nonempty open subset of X.
g π
Proof. By Proposition 15.4.9, f |Xsm factors locally as Xsm → − AnY →− Y for some étale g.
n
Since the smooth locus of π is all of AY (Example 15.4.8), it’s enough to prove the theorem
for étale morphisms. This will be done in the next section.
Let Y = Spec A be an affine scheme and Y0 ⊆ Y a closed subscheme defined by an ideal
I ⊂ A. Then Y0 is called an infinitesimal subscheme of Y if I 2 = 0 (or equivalently, if I is
nilpotent).
Theorem 15.4.14 (Infinitesimal Lifting). Let f : X → S be a morphism of finite type over
a locally noetherian scheme S. Then f is smooth if and only if for every affine S-scheme Y
and every infinitesimal subscheme Y0 ⊆ Y ,

HomS (Y, X) −→ HomS (Y0 , X)

is surjective.
The second condition in the theorem is called formal smoothness. Informally, this result
says that smoothness is detected by the property of being able to lift tangent vectors. Another
way to view formal smoothness is as an abstraction of Hensel’s Lemma. In the case of an
arbitrary scheme, one has:
Theorem 15.4.15. A morphism of schemes f : X → S is smooth if and only if f is formally
smooth and locally of finite presentation.
One can use the infinitesimal lifting criterion to prove the following generalization of the
Jacobian condition.

355
15.4. Smooth Morphisms Chapter 15. Morphisms of Schemes

Theorem 15.4.16 (Jacobian Condition for Schemes). Let f : X → Y be a morphism of


locally noetherian schemes and suppose j : U ,→ W is a closed immersion, where W is an
open subset of AnY . Then f is smooth of relative dimension n at x ∈ U if and only if the
sheaf of ideals I defining U ,→ W is generated on a neighborhood of j(x) by m = n − d local
sections f1 , . . . , fm such that  
∂fi
rank = n − d.
∂xj

356
15.5. Unramified Morphisms Chapter 15. Morphisms of Schemes

15.5 Unramified Morphisms


In this section we defined unramified morphisms and give several different characterizations
of this condition. In the next section this will be used to define étale morphisms.
Definition. An algebra A over a field k is called a finite étale k-algebra if A ∼
Qn
= i=1 Li
as k-algebras, where each Li is a finite, separable field extension of k.
Lemma 15.5.1. Let k be a field Qmwith algebraic closure k̄. Then a k-algebra A is finite étale
if and only if Ak̄ = A ⊗k k̄ ∼
= i=1 k̄ as k̄-algebras for some finite integer m.
Proof. ( =⇒ ) Suppose A ∼
Qn
= i=1 Li for finite, separable extensions Li /k. For each Li ,
there is a primitive element αi ∈ Li such that Li = k(αi ). Fixing i, let p(t) be the minimal
polynomial of αi over k, which factors over k̄ as
r
Y
p(t) = (t − aj ) for some aj ∈ k̄.
j=1

Since Li /k is separable, the ideals (t − a1 ), . . . , (t − ar ) are coprime in k̄[t], so the Chinese


remainder theorem allows us to write
r r
= k̄[t]/(p(t)) ∼
Li ⊗k k̄ ∼ k̄[t]/(t − aj ) ∼
Y Y
= = k̄
j=1 j=1

as k̄-algebras. This implies A ⊗k k̄ is also a product of copies of k̄.


( ⇒= ) is similar, using integrality and invariance of domain under integral extensions.
For details, see Bosch 8.4.6.
Definition. A morphism f : X → Y of schemes which is locally of finite type is said to
be unramified at a point x ∈ X if my OX,x = mx , where y = f (x), and the extension of
residue fields k(x)/k(y) is finite and separable. Otherwise f is ramified at x and the set
of all points at which f is ramified is called the ramification locus of f . We say f is an
unramified morphism if it is unramified at all x ∈ X.
Theorem 15.5.2. Let f : X → Y which is locally of finite type and fix x ∈ X. Then the
following are equivalent:
(a) f is unramified at x.

(b) Ω1X/Y,x = 0.

(c) There is an open neighborhood U ⊆ X of x such that the diagonal morphism ∆X :


X → X ×Y X restricts to an open immersion ∆X |U : U ,→ X ×Y X.
Proof. (a) =⇒ (b) The problem is local, so let us assume X = Spec B, Y = Spec A and
x = q ⊂ B and y = p ⊂ A are prime ideals. By assumption, pBq = qBq so

Bq ⊗Ap k(p) ∼
= Bq /pBq = k(q).

357
15.5. Unramified Morphisms Chapter 15. Morphisms of Schemes

By Proposition 15.3.4, Ω1X/Y,x ∼


= Ω1Bq /Ap and Proposition 15.2.6(a) gives

Ω1Bq /Ap ⊗Bq k(p) ∼


= Ω1(Bq ⊗Ap k(p))/k(p) = Ω1k(q)/k(p)

but this vanishes by Corollary 15.2.11. Further, since f is of finite type (locally, but we may
assume globally on Spec B → Spec A), Corollary 15.2.7 says Ω1Bq /Ap is a finitely generated
Bq -module so Nakayama’s lemma together with the above calculation imply Ω1Bq /Ap = 0.
(b) =⇒ (c) We may again assume X = Spec B, Y = Spec A and f is of finite type. In
the affine case, ∆ = ∆X is a closed immersion, say with ideal sheaf I. Then

0 = Ω1X/Y,x ∼
= (I/I 2 )∆(x) = I∆(x) /I∆(x)
2
.

It is routine to check that I∆(x) is a coherent sheaf, i.e. corresponds to a finitely generated
B ⊗A B-module via Theorem 14.4.7. Then by Nakayama’s lemma, I∆(x) = 0 and since it is
coherent, I then vanishes in some neighborhood V ⊆ X ×Y X of ∆(x). Thus ∆ restricts to
an isomorphism on U = ∆−1 (V ) ⊆ X.
(c) =⇒ (a) Since the question is again local, we may assume U = X = Spec B,
Y = Spec A and ∆ = ∆X is an open immersion. Consider the pullback diagram
Spec(Bp /pBp ) Spec B

Spec k(p) Spec A

Let q̄ be the image of qBp in Bp /pBp . The localization of Bp /pBp at q̄ is Bq /pBq and the
residue field of this local ring is (Bq /pBq )/q̄ ∼
= k(q). To prove f is unramified at x = q,
we must show qBq /pBq = 0 and k(q)/k(p) is a finite, separable extension. Since statement
(c) is preserved under base change, we may in fact assume Y = Spec k for a field k. Then
Lemma 15.5.1 shows that we can assume k is algebraically closed. Let π1 , π2 : X ×Y X → X
be the two coordinate projections and consider the morphism

h : X −→ X ×Y X
r 7−→ (r, p).

Then π1 ◦ h = idX and π2 ◦ h is the constant morphism X → Spec k ,→ X at p. Thus


h−1 (∆(X)) = {p} but since we are assuming ∆ is open, this actually implies the topology
on X is discrete. Furthermore, since B/k is finite type, this means B is artinian and hence
Spec B is finite. But OX,x ∼ = k for all k ∈ X so by the structure theory of local artinian
rings, B ∼
= k × · · · × k for a finite number of copies of k.
Remark. There are good geometric interpretations of each of the conditions in Theo-
rem 15.5.2:
(a) The definition of unramified is a technical way of saying that there are a ‘full’ number
of preimage points above x. This is analogous to the fact that a degree n cover of
Riemann surfaces is locally n-to-1.

358
15.5. Unramified Morphisms Chapter 15. Morphisms of Schemes

(b) The Kähler differential condition in (b) says that at every point in the fibre over x,
there are the same number of tangent directions as at x.

(c) Condition (c) says that ramification points are precisely the ‘limit points’ of ∆X (X)
which are not in the image of the diagonal.

Corollary 15.5.3. Let X be locally noetherian and suppose f : X → Y is a morphism


locally of finite type. Then the ramification locus of f is a proper closed subset of X.

Proof. This follows from Theorem 15.5.2(b) since the support of a coherent sheaf on a locally
noetherian scheme is closed.

Corollary 15.5.4. A morphism f : X → Y locally of finite type is unramified at x ∈ X if


and only if the morphism Xy = X ×Y Spec k(y) → Spec k(y) of fibres is unramified.

The unramified condition is preserved under composition and base change, as we saw for
flat and smooth morphisms in Lemmas 8.3.21/15.1.2 and Proposition 15.4.7, respectively.

Proposition 15.5.5. Suppose f : X → Y is unramified. Then

(a) If Y 0 → Y is any morphism, then the base change X 0 = X ×Y Y 0 → Y 0 is unramified.

(b) If g : Y → Z is unramified, then the composition g ◦ f : X → Z is also unramified.

Proof. (a) follows from Theorem 15.5.2(b) and Proposition 15.3.9(a).


(b) Fix x ∈ X. Set y = f (x), z = g(y) and suppose f is unramified at x and g is
unramified at y. Then mz OX,x = (mz OY,y )OX,x = my OX,x = mx and a tower finite, separable
extensions is again finite and separable, so k(x)/k(z) is finite and separable. Therefore g ◦ f
is unramified at x.

Example 15.5.6. Let Y be the cuspidal curve from Example 15.4.3 defined by y 2 − x3 = 0
over a field k of characteristic different from 2, 3. There is a normalization map f : A1k → Y
corresponding to the ring map A → B where A = k[t2 , t3 ] → k[t]. Then Ω1A1 /Y is the
sheafification of Ω1B/A , but notice that B = A[s]/(s2 − t2 ) so by Example 15.2.8,

Ω1B/A = B dx/2B dx ∼
= k[t]/(2t) ∼
= k[t]/(t).

This shows that the support of Ω1A1 /Y is the point corresponding to (t), so f is unramified over
A1 r {0}. This illustrates the ‘number of tangent directions’ interpretation of ramification
in the remark above.

Example 15.5.7. Let us show that the definition of ramification in this section agrees
with the notion in algebraic number theory. Let B/A be an integral extension of Dedekind
domains. Then for any nonzero prime p ⊂ A, the extension pB factors as a product of prime
ideals g
Y
pB = Pegg
i=1

359
15.5. Unramified Morphisms Chapter 15. Morphisms of Schemes

where each Pi is a prime of B lying over p and ei ≥ 1. In algebraic number theory, one says
that p is ramified if ei > 1 for some i and unramified if e1 = · · · = eg = 1. Fix 1 ≤ j ≤ g
and consider the morphism of affine schemes f : Spec B → Spec A. Then
g
Y
pBPj = (pB)Pj = (Pi BPj )ei = (Pj BPj )ej .
i=1

So f is unramified at Pj ∈ Spec B if and only if ej = 1. For example, the morphism


Spec Z[i] → Spec Z corresponding to the extension of Dedekind domains Z[i]/Z in the
quadratic extension Q(i)/Q is ramified at the point P = (1 + i) lying over p = (2) ∈ Spec Z
and unramified everywhere else. More generally, f : Spec B → Spec A is ramified at
P ∈ Spec B precisely when P divides the different ideal of B/A, DB/A . This ideal may
be define using fractional ideals, but a more enlightening definition here is that

DB/A = AnnB (Ω1B/A )

where Ω1B/A is the module of Kähler differentials of B/A. Note that this reinforces the
connection to ramification in number theory, since unramified is equivalent to Ω1 vanishing
after localization (Theorem 15.5.2).

360
15.6. Étale Morphisms Chapter 15. Morphisms of Schemes

15.6 Étale Morphisms


We have already defined a morphism f : X → Y to be étale if it is smooth of relative
dimension 0. The following result relates this to the more common definition of étale.
Theorem 15.6.1. Let f : X → Y be a morphism of schemes. Then the following are
equivalent:
(a) f is étale.

(b) f is smooth and unramified.

(c) f is flat and unramified.


Proof. (a) =⇒ (b) follows from Theorems 15.5.2 and 15.4.12.
(b) =⇒ (c) is trivial since a smooth morphism is by definition flat.
(c) =⇒ (a) Suppose f is flat and unramified. We need only check that the fibres Xy are
regular of dimension 0 over k(y) for each y ∈ Y . For such a y, fix x ∈ Xy . The unramified
condition means k(x)/k(y) is finite separable, so Ω1k(x)/k(y) = 0 by Corollary 15.2.11. There-
fore Spec k(x) → Spec k(y) is étale by Theorem 15.4.12 and as a consequence, Theorem 15.4.6
implies Xy is regular over k(y). On the other hand, my OX,x = mx which implies

OXy ,x = OX,x /my OX,x = OX,x /mx = k(x).

This shows Xy has dimension 0.


Corollary 15.6.2. Open immersions are étale.
Proof. An open immersion is locally of the form Spec Af ,→ Spec A for some ring A and
nonzero element f ∈ A. This is flat by Example 8.3.8 and unramified by Lemma 15.2.5(b),
along with Theorem 15.5.2.
Corollary 15.6.3 (Generic Étaleness). For a morphism of finite type f : X → Y between
locally noetherian schemes, there is a nonempty open set Xét ⊆ X over which f is étale.
Proof. This now follows from the same results for flatness (Theorem 15.1.14) and unrami-
fiedness (Corollary 15.5.3).
Note that this also proves generic smoothness (Theorem 15.4.13).
Proposition 15.6.4. Suppose f : X → Y is étale. Then
(a) If Y 0 → Y is any morphism, then the base change X 0 = X ×Y Y 0 → Y 0 is étale.

(b) If g : Y → Z is étale, then the composition g ◦ f : X → Z is also étale.


Proof. This already follows from Proposition 15.4.7 since étale morphisms are smooth and
fibre dimension is preserved in each case, but now we obtain a second proof by applying
Theorem 15.6.1, Lemmas 8.3.21 and 15.1.2 (for flat) and Proposition 15.5.5 (for unramified).

361
15.6. Étale Morphisms Chapter 15. Morphisms of Schemes

g f
Proposition 15.6.5. Suppose a composition X →
− Y →
− Z is étale and f is unramified.
Then g is also étale.

Proof. Let Γg : X → X ×Z Y be the graph of g, which sends x 7→ (x, g(x)). Then g = p2 ◦ Γg ,


where p2 : X ×Z Y → Y is the second coordinate projection. So by Proposition 15.6.4(b),
it’s enough to show p2 and Γg are each étale. For Γg , consider the pullback diagram
g
X Y

Γg ∆

X ×Z Y Y ×Z Y
g×1

where ∆ = ∆Y is the diagonal morphism. Since f is unramified, Theorem 15.5.2 says that
∆ is an open immersion, hence étale by Corollary 15.6.2. Then Proposition 15.6.4(a) implies
Γg is étale, too.
On the other hand, p2 is also a pullback:
p1
X ×Z Y X

p2 f ◦g
g
Y Z

So the hypothesis that f ◦ g is étale implies p2 is étale, again by Proposition 15.6.4(a).

Proposition 15.6.6. A closed immersion of noetherian schemes which is flat is also an


open immersion.

Proof. Let f : X → Y be such a morphism. Since f is flat, it is open by Theorem 15.1.12


so we may replace Y by f (X) and assume f is surjective. Then since f is finite, f∗ OX is a
locally free OY -module by Lemma 14.4.8. Finally the closed immersion assumption means
f∗ OX ∼= OY as OY -modules, so we conclude that f is an isomorphism.

Example 15.6.7. Let L/k be a finite extension of fields. The morphism f : Spec L → Spec k
is always flat, so f is étale if and only if it is unramified. On the other hand, Theorem 15.5.2
and Corollary 15.2.11 say that f is unramified if and only if L/k is separable. Note that L
is a trivial example of a finite étale k-algebra. For any k-algebra A, the structure morphism
f : Spec A → Spec k is étale if and only if A is a finite étale k-algebra. If A is of the
form A = k[t]/(p(t)), where p(t) is a monic polynomial over k, then the étale condition is
equivalent to p(t) being separable for the same reason as the case of L/k.

Example 15.6.8. Let p(t) be a monic polynomial over an arbitrary commutative ring A
and consider B = A[t]/(p(t)). Then B is a free A-module of rank n = deg p and for any
prime p ∈ Spec B,
B ⊗A k(p) ∼
= k(p)[t]/(p̄(t))

362
15.6. Étale Morphisms Chapter 15. Morphisms of Schemes

where p̄(t) is the image of p(t) in k(p)[t]. Then Corollary 15.5.4 shows that B is an étale
A-algebra (or equivalently Spec B → Spec A is étale) if and only if (p(t), p0 (t)) = 1, where p0
is the formal derivative of p. The same statement holds when we replace B by a localization
Bb at a nonzero element b ∈ B; namely, Bb is étale over A if and only if p0 (t) is a unit in
Bb [t].
Definition. Let B = A[t]/(p(t)) for a monic polynomial p(t) ∈ A[t] and let b ∈ B be nonzero
such that p0 (t) is a unit in Bb [t]. Then Bb is called a standard étale A-algebra.
In a moment we will prove that every étale morphism is locally standard étale, meaning
isomorphic to Spec Bb → Spec A for a standard étale A-algebra Bb . We will need Zariski’s
Main Theorem, which we state but do not prove below.
Definition. A morphism f : X → Y is called quasi-finite if f −1 (y) is discrete for all
y ∈Y.
Theorem 15.6.9 (Zariski’s Main Theorem). Let Y be a noetherian scheme and f : X → Y
a separated, quasi-finite morphism. Then f factors as
j g
− X0 →
f :X→ − Y

where j is an open immersion and g is a finite morphism.


Theorem 15.6.10. Let f : X → Y be a morphism over a locally noetherian scheme Y .
If f is étale over an open neighborhood of x ∈ X, then there are affine open neighborhoods
V ⊆ X of x and U ⊆ Y of y = f (x) such that f |V : V → U is a standard étale morphism.
Proof. We may assume X = Spec C, Y = Spec A and f is separated. The proof of Theo-
rem 15.6.1 showed that unramified morphisms are quasi-finite, so by Zariski’s main theorem,
we may assume C is a finite A-algebra. Let x = q ∈ Spec C. We will construct a standard
étale A-algebra Bb and show that Bb ∼ = Cc for some c ∈ C r q. Since C/A is finite, we
may localize at y = p ∈ Spec A, so assume A itself is a local ring with maximal ideal mA
and we have q ∩ A = mA . Choose an element γ ∈ C whose image γ̄ in C/pC generates
the residue field extension k(q)/k(p), which is possible since this field extension is separa-
ble. Set q0 = q ∩ A[γ]. Since f is unramified, q is the unique prime lying over q0 . This
means C ⊗A[γ] A[γ]q0 ∼= Cq , but since A[γ] → C is finite and injective, so is the localization
A[γ]q0 → Cq . On the other hand, this map is also surjective: k(q0 ) → k(q) is surjective, so
Nakayama’s lemma (1.2.1) implies the cokernel of A[γ]q0 → Cq is 0.
Next, A[γ]q0 ∼
= Cq induces an isomorphism A[γ]c0 ∼ = Cc for some c ∈ C r q and c0 ∈
A[γ] r q . Since A[γ] is finite as an A-algebra, we may replace C by A[γ], so that q = q0 .
0

Set n = [k(q) : k(p)]. Then {1, γ̄, . . . , γ̄ n−1 } is a basis for the field extension k(q)/k(p),
so Nakayama’s lemma (version 1.2.4) implies {1, γ, . . . , γ n−1 } generates C as an A-algebra.
Thus there is a monic polynomial p(t) ∈ A[t] such that

B := A[t]/(p(t)) −→ A[γ] = C

is surjective. Note that the image p̄(t) of p(t) in k(q)[t] is exactly the characteristic polynomial
of γ̄ over k(p), so p̄(t) is separable. Therefore for some b ∈ B not landing in q ⊂ C,

363
15.6. Étale Morphisms Chapter 15. Morphisms of Schemes

Example 15.6.8 shows that Bb is a standard étale algebra over A. For the corresponding
choice of c ∈ C, we get a surjection

h : Bb −→ Cc .

We finish by showing h is an isomorphism. Notice that we have a composition


h∗ r
Spec Cc −→ Spec Bb →
− Spec A

where r and h∗ ◦ r are both étale. Therefore Proposition 15.6.5 shows that h∗ is also étale.
But since h is surjective, h∗ is an étale closed immersion and hence an open immersion by
Proposition 15.6.6. This proves h is an isomorphism.

364
15.7. Henselian Rings Chapter 15. Morphisms of Schemes

15.7 Henselian Rings


Definition. A local ring (A, m, k) is henselian if it satisfies Hensel’s Lemma, i.e. if when-
ever f (x) ∈ A[x] is a monic polynomial and α ∈ A such that f (α) ∈ m but f 0 (α) 6∈ m, then
there exists an element a ∈ A for which a ≡ α mod m and f (a) = 0.
The condition that f (α) ∈ m but f 0 (α) 6∈ m is the same as saying ᾱ is a simple root of
f¯(x) over k = A/m. As with Hensel’s Lemma, there are multiple equivalent characterizations
of henselian rings. We need to set up a few things first before stating the main theorem.
Definition. A local homomorphism of local rings ϕ : (A, m) → (B, n) is called a pointed
étale neighborhood if B is an étale A-algebra and A/m ∼
= B/n as A-modules.
Note that by Theorem 15.6.10, any pointed étale neighborhood A → B is of the form
B = (A[x]/(f (x)))q where f (x) ∈ A[x] is monic and q = (x − a, mB) for some a ∈ A such
that f (a) ∈ m and f 0 (a) 6∈ m.
Definition. A ring B is decomposable if B ∼
Q`
= i=1 Bi for local rings (Bi , mi ).
Proposition 15.7.1. Let (A, m, k) be a local ring and B an A-module. Then
= `i=1 Bi is decomposable, then MaxSpec(B) = {M1 , . . . , M` } where Mi =
Q
(1) If B Q
mi × j6=i Bj . Moreover, BMi ∼
= Bi for each 1 ≤ i ≤ `.
Let ei = (0, . . . , 0, 1, 0, . . . , 0) be the ith orthogonal idempotent in B. Then B ∼
(2) Q =
`
Be i and Be ∼
i = B ei for each 1 ≤ i ≤ `.
i=1

(3) For any finite ring map A → B, B ⊗A k is decomposable.


(4) If MaxSpec(B) = {M1 , . . . , M` }, then the following are equivalent:
(i) B is decomposable.
(ii) B ∼
Q`
= Mi .
i=1
(iii) For any finite ring map A → B, the decomposition of B ⊗A k lifts to a decom-
position of B.
Proof. (1) and (2) are routine. For (3), use the Chinese remainder theorem along with the
fact that B/mB is artinian. In (4), (ii) =⇒ (iii) =⇒ (i) are trivial and (i) =⇒ (ii) follows
directly from (1).
As above, let (A, m, k) be a local ring and A → B a finite ring map. Set B = B ⊗A k.
Then every idempotent of B determines an idempotent of B, say e 7→ ē.
Lemma 15.7.2. The assignment {idempotents in B} → {idempotents in B} is one-to-one.
Moreover, it is onto if either of the following equivalent conditions hold:
(a) B is decomposable.
(b) The orthogonal idempotents ē1 , . . . , ē` ∈ B = B 1 × · · · × B ` lift to a set of idempotents
of B.

365
15.7. Henselian Rings Chapter 15. Morphisms of Schemes

Proof. Let e and f be idempotents in B, set y = e − f and assume ȳ = ē − f¯ = 0 in B.


Notice that

y 3 = (e − f )3 = e3 + 3e2 f − 3ef 2 − f 3 = e + 3ef − 3ef − f = e − f = y,

or in other words, y 3 − y = 0, so y(y 2 − 1) = 0. However, y ∈ J(B) so y 2 − 1 is a unit and


therefore y = 0. This shows e 7→ ē is one-to-one in general.
Now suppose ei are idempotent lifts of the ēi to B. Then for each 1 ≤ i, j ≤ `, ei ej is an
idempotent lift of ēi ēj but ēi ēj = 0 whenever i 6= j, so by the first paragraph, ei ej = 0 for
all i 6= j. This then implies e1 + . . . + e` is an idempotent, and it is a lift of ē1 + . . . + ē` = 1
from B̄ to B, so again by the one-to-one property, e1 + . . . + e` = 1 in B. This proves
B∼ = Be1 × · · · × Be` which is a lifting of B = B 1 × · · · × B ` .
We can now state and prove the main theorem characterizing henselian rings.

Theorem 15.7.3. For a local ring (A, m, k), the following are equivalent:

(1) A is henselian.

(2) Every pointed étale neighborhood (A, m) → (B, n) is an isomorphism of local rings.

(3) For all F1 , . . . , Fn ∈ A[x1 , . . . , xn ], if there exist elements a1 , . . . , an ∈ A such that


Fi (a1 , . . . , an ) ∈ m for all 1 ≤ i ≤ n and the Jacobian determinant
 
∂Fi
J(a1 , . . . , an ) := det (a1 , . . . , an )
∂xj

does not lie in m, then there exist elements b1 , . . . , bn ∈ A with bi ≡ ai mod m and
Fi (b1 , . . . , bn ) = 0 for all 1 ≤ i ≤ n.

(4) If F (x) ∈ A[x] is monic and its reduction mod m factors as F = gh for some relatively
prime (necessarily monic) polynomials g, h ∈ k[x], then F = GH for some monic
polynomials G, H ∈ A[x] with G = g and H = h.

(5) Every module-finite extension of rings A → B is decomposable.

Proof. (1) =⇒ (2) Write B = (A[x]/(f (x)))q as above, with q = (x − α, mB) for α ∈ A
such that f (α) ∈ m but f 0 (α) 6∈ m. By assumption, there exists a lift a ∈ A such that a ≡ α
mod m and f (a) = 0. Then f (x) = (x − a)g(x) for some g ∈ A[x]; since f 0 (a) 6= 0, g is not
divisible by x − a. Thus g 6∈ q, so it follows that q = (x − a, mB). Localizing at q then gives

B = (A[x]/(f (x)))q = (A[x]/(x − a))q ∼


= A.

(2) =⇒ (3) Set B = (A[x1 , . . . , xn ]/(F1 , . . . , Fn ))q where q = (x1 − a1 , . . . , xn − an , mB).


Then B/q ∼ = k = A/m, showing B is a pointed étale neighborhood of A, so by hypothesis

B = A. For 1 ≤ i ≤ n, denote the image of xi under this isomorphism by bi . Then
Fj (b1 , . . . , bn ) = 0 and by construction bi ≡ ai mod m for all j.

366
15.7. Henselian Rings Chapter 15. Morphisms of Schemes

(3) =⇒ (4) Let F, g and h be as in the statement of (4) and write


n−1
X
n
F (x) = x + ai x i ,
i=0
n−1
X
f (x) = F (x) = xn + λ i xi where λi = āi ,
i=0
s−1
X
g(x) = xs + α i xi
i=0
t−1
X
and h(x) = xt + β i xi , where s + t = n.
i=0

Then the condition f = gh is equivalent to the existence of a solution to the system of


equations
α0 β0 − λ0 = 0
α1 β0 + α0 β1 − λ1 = 0
..
.
αs−1 + βt−1 − λn−1 = 0.
Likewise, F = GH corresponds to a solution in A to the system
F1 : z0 y0 − a0 = 0
F2 : z1 y0 + z0 y1 − a1 = 0
..
.
Fn : zs−1 + yt−1 − an−1 = 0.
By (3), such a solution exists as long as the Jacobian condition
 
∂Fi
det (α0 , . . . , αs−1 , β0 , . . . , βt−1 ) 6= 0
∂zi , yi
is satisfied. Explicitly, this matrix is the determinant of the resultant of f and g:
· · · · · · yt−1 1 0 ···
 
y0 y1 0
 0 y0
. . · · · · · · yt−2 yt−1 1 · · · 0 
. . .. .. .. .. .. .. 
. . . . . . . . 

Res(f, g) = 
0 0 ··· ··· y0 y1 y2 · · · yt−1 

z0 z1 · · · zs−1 · · · · · · · · · · · · 0 
 
.. .. ... .. ... ... ... ... ..
. . . .
Set J = det(Res(f, g)). Then J(α, β) := J(α0 , . . . , αs−1 , β0 , . . . , βt−1 ) = 0 if and only if there
is a nonzero solution to
(z0 , . . . , zs−1 , w0 , . . . , wt−1 )J(α, β) = 0

367
15.7. Henselian Rings Chapter 15. Morphisms of Schemes

for some zi , wj . But this equation is really

(z0 + z1 x + . . . zs−1 xs−1 )h(x) + (w0 + . . . + wt−1 xt−1 )g(x) = 0

so J(α, β) = 0 precisely when g and h are not relatively prime. Hence (3) guarantees the
existence of such a solution and consequently a factorization F = GH lifting the equation
f = gh to A[x].
(4) =⇒ (5) Notice that condition (4) implies that every A-algebra of the form A[x]/(f (x)),
for f (x) monic, is decomposable. Given a finite ring extension A → B, write B = B ⊗A k =
B 1 × · · · × B ` using Proposition 15.7.1(3). By Lemma 15.7.2, it’s enough to lift the orthog-
onal idempotents ē1 , . . . , ē` ∈ B; since the argument is identical for each one, we will just
construct the lift for ē = ē1 . For any lift b ∈ B of ē, let f (x) ∈ A[x] be the minimal monic
polynomial having b as a root. Set R = A[x]/(f (x)) and consider the diagram

A A[b]

ϕ
R B

As in the statement of Proposition 15.7.1, let M1 , . . . , M` be the maximal ideals of B 1 , . . . , B ` ,


respectively. Then b ∈ M2 ∩· · ·∩M` and bM1 = {1}. Set n = ϕ∗ M1 ∈ Spec(R). We claim M1
is the unique maximal ideal of R lying over n – this follows from the fact that (x, M1 ) = R
and ϕ(x) = b ∈ M2 ∩ · · · ∩ M` . Further, we know R is decomposable and Rn is one of its
factors, so the idempotent corresponding to this factor lifts ē.
(5) =⇒ (1) Assume f (x) ∈ A[x] is monic and α ∈ A satisfies f (α) ∈ m and f 0 (α) 6∈ m.
Set B = A[x]/(f (x)) – since f is monic, B is finite over A. Then the hypothesis says that B
is decomposable, say B = B1 × · · · × B` , but by Proposition 15.7.1(4), this decomposition is
a lift of B = B ⊗A k = B 1 × · · · × B ` . We know f¯(x) = (x − α)g(x) for some g ∈ k[x] not
divisible by x − α), so without loss of generality, assume B 1 ∼ = k[x]/(x − α) ∼
= k. Then the
decomposition of B lifts to a decomposition B = B1 × · · · × B` with B1 ∼ = A[x]/(x − a) for
some a ≡ α mod m for which f (a) = 0.
Remark. Hensel’s Lemma is often stated in the form of condition (4), from which (1) is
easily derived.
Example 15.7.4. The original version of Hensel’s Lemma states that Tthe lifting criterion for

simple roots holds in any complete local ring which is separated, i.e. n=1 mn = 0. Therefore
separated complete local rings are henselian.
Definition. For a local ring (A, m, k), a henselization of A is a local ring (Ah , mh ) satis-
fying:
(1) Ah is a henselian ring admitting a local homomorphism i : A → Ah .

(2) Ah is universal with respect to such local homomorphisms: if f : A → B is any local


homomorphism to a henselian ring B, then there is a unique map g : Ah → B such
that f = g ◦ i.

368
15.7. Henselian Rings Chapter 15. Morphisms of Schemes

We will show that every local ring has a henselization. To construct Ah , we need the
following facts about pointed étale neighborhoods.
Proposition 15.7.5. Suppose B and C are pointed étale neighborhoods of a local ring
(A, m, k). Then
(1) If Q is the kernel of B ⊗A C → k ⊗A k = k, then Q is a prime ideal and (B ⊗A C)Q
is a pointed étale neighborhood of A.

(2) There is at most one local homomorphism B → C and if one exists, C is a pointed
étale neighborhood of B.

(3) If there exist local homomorphisms B → C and C → B, then they are both isomor-
phisms.
Proof. (1) By Theorem 15.6.10, we may write

B = (A[x]/(f (x)))q1 and C = (A[y]/(g(y)))q2

for monic polynomials f (x) and g(y) and prime ideals q1 and q2 . Then B⊗A C is a localization
of A[x, y]/(f, g) and f 0 (x), g 0 (y) 6∈ Q, so (B ⊗A C)Q is a localization of a finite étale A-algebra.
Moreover, k(Q) = k, so (B ⊗A C)Q is a pointed étale neighborhood.
(2) If there exists a map α : B → C then Proposition 15.6.5 implies α is also étale.
Since the residue fields of A, B and C are all k, this implies B → C is a pointed étale
neighborhood so the second statement holds. For uniqueness, suppose β : B → C is another
homomorphism. Then the map B ⊗A B → C given by b ⊗ b0 7→ α(b)β(b0 ) takes Q from
(1) to the maximal ideal of C and hence induces a map (B ⊗A B)Q → C, with b⊗1 1
7→ α(b)
and 1⊗b 1

7 β(b). On the other hand, the natural map B ⊗ A B → B takes Q to the
maximal ideal of B, so induces a map (B ⊗A B)Q → B. If I is the kernel of B ⊗A B → B, by
Proposition 15.3.2 we know I/I 2 ∼ 1
= ωB/A = 0 since A → B is étale. Therefore IQ ⊂ (B⊗A B)Q
is finitely generated, lies in the maximal ideal and satisfies IQ2 = IQ so it is 0 by Nakayama’s
lemma (Theorem 1.2.1). That is, (B ⊗A B)Q → B is an isomorphism, but this means that
under (B ⊗A B)Q → C, both b⊗1 1
and 1⊗b
1
map to the same element, so α(b) = β(b) and we
conclude that α = β.
(3) It follows from (2) that the only local homomorphism B → B is the identity, so any
compositions B → C → B and C → B → C must be the identity.
Corollary 15.7.6. The isomorphism classes of pointed étale neighborhoods of A form a
directed system.
Theorem 15.7.7. Let (A, m, k) be a local ring. Then
(1) There exists a henselization of A, namely Ah = lim B where the limit is taken over
−→
some choice of representatives of the isomorphism classes of pointed étale neighbor-
hoods A → B.

(2) If mh is the maximal ideal of Ah , then mAh = mh and Ah /mh ∼


= A/m = k.
(3) The natural map A → Ah is étale.

369
15.7. Henselian Rings Chapter 15. Morphisms of Schemes

(4) If A → C is a pointed étale neighborhood then C h ∼


= Ah .

(5) (Ah )h = Ah and (A


dh ) = A.
b

(6) If A is noetherian, then so is Ah .


Proof. (1) Set A0 = lim B with maximal ideal m0 , let f (x) ∈ A0 [x] be monic and suppose
−→
a ∈ A0 such that f (a) ∈ m0 and f 0 (a) 6∈ m0 . Then there exists a pointed étale neighborhood
B of A such that a ∈ B and f (x) ∈ B[x]. Let mB be the maximal ideal of B and define
B1 = (B[x]/(f (x)))(x−a,mB B[x])
which is a pointed étale neighborhood of B by construction. Then Proposition 15.7.5(2) says
B1 is also a pointed étale neighborhood of A. Thus there is a map ϕ : B1 → A0 compatible
with the directed system defining A0 such that ϕ(x̄) is a lifting of a which is a root of f (x).
That is, A0 is henselian. Now we show A0 is universal with respect to maps from A to
henselian rings. Indeed, given such a map A → C with C henselian, take a pointed étale
neighborhood A → B. Since C is henselian, Theorem 15.7.3(2) says C is isomorphic to
any pointed étale neighborhood (C[x]/(f (x)))q , so with the right choice of q we get a local
homomorphism B → C. These assemble to give a map A0 = lim B → C, so A0 = Ah .
−→
(2) Similar to the analogous property for the completion (A,
b m).
b
h
(3) By Theorem 15.6.1 it’s enough to show A → A is flat and unramified, but this
follows from the fact that flatness and Ω1−/A are preserved under direct limits, along with
(1).
(4) follows easily from the description in (1).
(5) The first statement is obvious from the universal property of henselization. As with
completion, it’s easy to show that for all n ≥ 1,
(A/mn )h ∼
= Ah /mn Ah ∼
= Ah /(mh )n .

Then taking inverse limits gives A bh∼


b = (A) = (A
d h ).

(6) can be found in Milne, 2.8.8.


Proposition 15.7.8. Suppose (A, m) → (B, n) is a local homomorphism of local rings and
B is henselian. Let A be the integral closure of A in B and set C = An∩A , which is a local
ring with maximal ideal mC = n ∩ A. Then (C, mC ) is a henselian ring.
Proof. Let f (x) be a monic polynomial in C[x] and suppose α ∈ C is a simple root mod mC ,
that is, f (α) ∈ mC but f 0 (α) 6∈ mC . Write
a1 an−1 an
f (x) = xn + xn−1 + . . . + x+
s s s
for ai ∈ A and s ∈ A r (n ∩ A). Then since B is henselian, there exists an element a ∈ B
with f (a) = 0 and a ≡ α mod n. The condition f (a) = 0 implies
(sa)n + a1 (sa)n−1 + . . . + an−1 (sa) + sn−1 an = 0
so sa ∈ A since A is integral over A. On the other hand, s 6∈ n ∩ A forces a ∈ An∩A = C.
Hence C is henselian.

370
15.7. Henselian Rings Chapter 15. Morphisms of Schemes

Definition. A local ring (A, m) is an approximation ring if whenever F1 , . . . , F` ∈ A[x1 , . . . , xn ]


have a common zero (â1 , . . . , ân ) over the completion A, b there exists a solution over A, i.e.
there exist elements a1 , . . . , an ∈ A with Fj (a1 , . . . , an ) = 0 for all 1 ≤ j ≤ `.
It is immediate from the definitions that:
complete local rings ⊆ approximation rings ⊆ henselian rings.
Proposition 15.7.9. Suppose (A, m) is an approximation ring. Then for any polynomials
F1 , . . . , F` in A[x1 , . . . , xn ] admitting “approximate roots” â1k , . . . , ânk ∈ A
b for all k ≥ 1 such
k
that Fj (âjk ) ∈ m , there exist actual roots a1 , . . . , an ∈ A with Fj (a1 , . . . , an ) = 0 for all
1 ≤ j ≤ ` and aj ≡ âjk mod mk for all k ≥ 1.
Conversely:
Theorem 15.7.10 (Popescu). Let (A, m) be a henselian local ring. If the formal fibres of A
are geometrically regular, then A is an approximation ring.
Remark. Recall that the formal fibres of A are k(p) ⊗A A b for p ∈ Spec(A), and these are
geometrically regular if k(p) ⊗A A
b are all regular, where k(p) denotes an algebraic closure.

Corollary 15.7.11 (Artin Approximation). Let (A, m, k) be a henselian local ring, fix N ≥ 1
and suppose there are polynomials F1 , . . . , Fn ∈ A[x] and elements â1 , . . . , ân ∈ Ab such
that Fj (â1 , . . . , ân ) = 0 for all 1 ≤ j ≤ n. Then there exist a1 , . . . , an ∈ A such that
Fj (a1 , . . . , an ) = 0 for all 1 ≤ j ≤ n and ai ≡ âi mod mN A
b for all 1 ≤ i ≤ n.

Proof. Let mN = (u1 , . . . , ur ) for elements u` ∈ A. For each i, choose any bi ∈ A such that
bi ≡ âi mod mN A
b and write
n
X
bi − âi = âi` u`
`=1

for some âi` ∈ A.


b Consider the following system of equations over A:

F1 (x1 , . . . , xn ) = 0
..
.
Fn (x1 , . . . , xn ) = 0
X n
x 1 − b1 = y1` u`
`=1
..
.
n
X
x n − bn = yn` u` .
`=1

Since the system has a solution over A,


b namely xi = âi and yi` = âi` , Popescu’s theorem says
there’s a solution over A, say xi = ai and yi` = ai` . Then Fj (a1 , . . . , an ) = 0, ai − bi ∈ mN
and âi ≡ bi ≡ ai mod mN A b as required.

371
15.7. Henselian Rings Chapter 15. Morphisms of Schemes

Theorem 15.7.12. If (A, m, k) is a local noetherian ring which is excellent and normal,
then
Ah = Amb ∩A
where A is the integral closure of A in its completion A.
b

Proof. By Theorem 15.7.7 and Proposition 15.7.8, Ah and B := Amb ∩A are henselian and Ah
is the smallest henselian extension of A, so it suffices to show B ⊆ Ah . In fact, we just need
to show A ⊆ Ah since then its localization at m b ∩ A will also be contained in Ah . Take
a∈A b integral over A. Then since A is normal and excellent, there is a monic polynomial
f (x) ∈ A[x] with f (a) = 0 and fintely many roots in A b which can be distinguished mod
m A for any sufficiently large N . Applying Artin approximation to Ah , along with the
N b

fact (Theorem 15.7.7) that (A bh ) = A,


b we get an element b ∈ Ah with f (b) = 0 and b ≡ a
mod mN A.b But this implies a = b, so a ∈ Ah .

Definition. A local ring (A, m, k) is strictly henselian if it is henselian and k is separably


closed, i.e. k = k sep .

Qn henselian, then for every finite étale morphism A → B,


Proposition 15.7.13. If A is strictly

B is ‘strictly decomposable’: B = i=1 A.
Proof. By Theorem 15.7.3(5), B is decomposable but the étale hypothesis guarantees that
the residue fields of the factors are all finite separable extensions of k = A/m. Hence
Theorem 15.7.3(2) implies each factor is isomorphic to A.
Definition. For a local ring (A, m, k), a strict henselization of A is a local ring (Ash , msh )
satisfying:
(1) Ash is a strictly henselian ring admitting a local homomorphism j : A → Ash .

(2) Ash is universal with respect to such local homomorphisms: if f : A → B is any local
homomorphism to a strictly henselian ring B, then there is a unique map g : Ash → B
such that f = g ◦ j.
Similar to the properties of Ah , one can prove:
Theorem 15.7.14. Let (A, m, k) be a local ring. Then
(1) There exists a strict henselization Ash , namely Ash = lim B where the limit is over all
−→
local rings (B, mB ) admitting local étale maps A → B such that B/mB ∼ = k sep .
(2) If msh is the maximal ideal of Ash , then mAsh = msh and Ash /msh ∼
= k sep .
(3) The natural maps A → Ash and Ah → Ash are étale.

(4) If A → C is a local étale map such that C/mC ∼


= k sep , then C sh ∼
= Ash .
(5) (Ash )sh = Ash and (Ah )sh = Ash .

(6) If A is noetherian, then so is Ash .

372
Chapter 16

Cohomology

373
16.1. Direct and Inverse Image Chapter 16. Cohomology

16.1 Direct and Inverse Image


Let ShX denote the category of sheaves (of abelian groups) on a scheme X. For a morphism
f : X → Y , there are various functors between ShX and ShY which may illuminate the
structure of these categories. The easiest to define is the direct image.
Definition. For a morphism f : X → Y and a presheaf F on X, define the direct
image (or pushforward) of F along f to be the presheaf f∗ F : Topop
Y → Ab sending
V 7→ (f∗ F )(V ) := F (f −1 (V )) for all open V ⊆ Y .
Proposition 16.1.1. For any morphism f : X → Y , f∗ is a functor ShX → ShY .
Proof. Suppose ϕ : F → G is a morphism of presheaves on X. Then for all open V ⊆ Y ,
−1
define (f∗ ϕ)V : (f∗ F )(V ) → (f∗ G)(V ) to be the map ϕf −1 (VS ) : F (f (V )) → G(f −1 (V )).
This shows that f∗ is a functor.S To finish, take −1 a covering V = Vi in Y , set U = f −1 (V ) and
consider the covering U = Ui where Ui = f (Vi ). Suppose s, t ∈ (f∗ F )(V ) = F (U ) such
that s|Vi = t|Vi for all Vi . This means s|Vi = t|Vi in (f∗ F )(Vi ) = F (Ui ), so by the first sheaf
axiom for F , s = t in F (U ) = (f∗ F )(V ). On the other hand, suppose si ∈ (f∗ F )(Vi ) = F (Ui )
for all i such that si |Vi ∩Vj = sj |Vi ∩Vj for all Vi , Vj . This means si and sj are equal in each
(f∗ F )(Vi ∩ Vj ) = F (f −1 (Vi ∩ Vj )) = F (Ui ∩ Uj ) so by the second sheaf axiom on F , there is
a section s ∈ F (U ) restricting to si on each Ui . That is, s ∈ (f∗ F )(V ) such that s|Vi = si in
each (f∗ F )(Vi ). Hence f∗ F is a sheaf.
Example 16.1.2. If i : {x} → X is a geometric point of X, then any sheaf on {x} is an
abelian group, say A, and i∗ A = Ax is the skyscraper sheaf on X supported at x with stalk
A. This explains the usual notation x∗ A for skyscraper sheaves.
Example 16.1.3. Let p : X → Y be a finite morphism of degree n > 1 and let F = FS
`nfunctions on Y with values in S, |S| > 1. Then for V ⊆ Y
be the sheaf of locally constant
−1
sufficiently small, f (V ) = i=1 Ui for disjoint neighborhoods U1 , . . . , Un which are each
isomorphic to V . Thus (p∗ F )(V ) = S ⊕n . This shows that for each y ∈ Y , the stalk at y of
F is
(p∗ F )y = S ⊕n .
Another sheaf on Y with stalk S ⊕n is the locally constant sheaf G = FS ⊕n . However, p∗ F is
not isomorphic to G, as we have
(p∗ F )(Y ) = S 6= S ×n = G(Y ).
Proposition 16.1.4. For any morphism f : X → Y , the functor f∗ : ShX → ShY is left
exact.
Proof. Suppose 0 → F → G → H → 0 is a short exact sequence in ShX . By Theo-
rem 13.2.10, the sequence
0 → F (U ) → G(U ) → H(U )
is exact for all open sets U ⊆ X. In particular, if V ⊆ Y is open, then the above applied to
U = f −1 (V ) shows that
0 → (f∗ F )(V ) → (f∗ G)(V ) → (f∗ H)(V )
is exact. Passing to stalks proves the statement.

374
16.1. Direct and Inverse Image Chapter 16. Cohomology

This allows us to define the right derived functors of f∗ .

Definition. Let E • be an injective resolution of a sheaf F ∈ ShX and suppose f : X → Y


is a morphism. The higher direct images of f are the right derived functors of the direct
image functor:
Rn f∗ F := Hn (f∗ E • )
where for a complex of sheaves G• on Y and n ≥ 0, Hn (G• ) denotes the quotient sheaf
ker(Gn → Gn+1 )/ im(Gn−1 → Gn ).
f g
Proposition 16.1.5. Suppose X →
− Y →
− Z are morphisms of schemes. Then (g ◦ f )∗ =
g∗ ◦ f∗ .

Lemma 16.1.6. Let f : X → Y be a morphism, V = {Vj } an open cover of Y and


U = {f −1 (Vj )} the induced open cover of X. Then for all sheaves F on X and all p ≥ 0,
p p
Ȟ (U, F ) = Ȟ (V, f∗ F ).

Proof. Set Vj0 ,...,jp = Vj0 ∩ · · · ∩ Vjp and Uj0 ,...,jp = f −1 (Vj0 ,...,jp ). Then (f∗ F )(Vj0 ,...,jp ) =
F (Uj0 ,...,jp ) so it follows that C p (U, F ) = C p (V, f∗ F ) for all p ≥ 0. Hence the cohomologies
of these complexes are the same.

Example 16.1.7. Let i : X ,→ Y be a closed immersion and let F be a sheaf on X. Then


for any y ∈ Y , the stalk of i∗ F at y is
(
Fx , y = i(x) ∈ i(X)
(i∗ F )y =
0, y 6∈ i(X).

In particular, this proves:

Lemma 16.1.8. If i : X ,→ Y is a closed immersion, then i∗ : ShX → ShY is an exact


functor.

Therefore, if E • is an injective resolution of a sheaf F on X, then for a closed immersion


i : X ,→ Y the cohomology of i∗ F may be computed via the resolution i∗ E • :

H p (X, F ) = H p (Y, i∗ F ) = H p (Γ(Y, i∗ E • )).

Let f : X → Y be any continuous map. There is also an inverse image functor f ∗ :


ShY → ShX which is a little harder to construct, since f (U ) need not be open in Y for all
open U ⊆ X.

Definition. Let f : X → Y be a morphism, F a sheaf on Y and define a presheaf f −1 F by

(f −1 F )(U ) = lim F (V )
−→

where the direct limit is taken over all open sets V ⊆ Y containing f (U ). The inverse
image (or pullback) of F along f , denoted f ∗ F , is the sheafification of f −1 F .

375
16.1. Direct and Inverse Image Chapter 16. Cohomology

Remark. The inverse image sheaf may also be defined topologically on the level of étale
spaces as follows. Let EF → Y be the étale space of F . Then f ∗ F is the sheaf of sections of
the map E = Ef ∗ F → X whose total space is the pullback in the diagram

E X

EF Y

Example 16.1.9. If f : X → Y is an open morphism, then for all U ⊆ X, (f −1 F )(U ) =


F (f (U )) and this already defines a sheaf, so f −1 F = f ∗ F .
Example 16.1.10. Let i : {x} ,→ X be the inclusion of a closed point in X and let F be a
sheaf on X. Then i∗ F = Fx , the (constant sheaf on {x} given by the) stalk of F at x. More
generally:
Proposition 16.1.11. For any morphism f : X → Y , sheaf F on Y and point x ∈ X, there
is a natural isomorphism (f ∗ F )x ∼
= Ff (x) .
Proof. Let f −1 F be the presheaf U 7→ lim F (V ) as in the definition of inverse image, set
−→
y = f (x) and suppose U ⊆ X is an open neighborhood of x and V ⊆ Y is an open set in
Y containing f (U ). Then y lies in every such V so we get morphisms F (V ) → Fy which
assemble into a morphism out of the direct limit,

λU : (f −1 F )(U ) −→ Fy .

Taking the limit over all such neighborhoods U of x gives a morphism λ : (f −1 F )x → Fy .


We claim this is an isomorphism. For sy ∈ Fy , choose a representative sV ∈ F (V ) on some
open neighborhood V of y. Then U = f −1 (V ) is an open neighborhood of x in X such that
f (U ) ⊆ V . Hence there are maps
λ
F (V ) → (f −1 F )(U ) −→
U
Fy

whose composition is the stalk map F (V ) → Fy . Thus λU (f −1 sV ) maps to sy in the stalk


at y, so λ is surjective.
On the other hand, if t1 , t2 ∈ (f −1 F )x with λ(t1 ) = λ(t2 ), then there is some neighborhood
U of x such that t1 = tU,1 |x and t2 = tU,2 |x for sections tU,1 , tU,2 ∈ (f −1 F )(U ). (A priori,
these happen on different open neighborhoods of x but we may pass to the intersection.)
In turn these tU,1 , tU,2 are represented by sections sV,1 , sV,2 ∈ F (V ) for some open set V
containing f (U ) and the fact that λ(t1 ) = λ(t2 ) ensures that sV,1 |y = sV,2 |y in Fy . Thus by
shrinking V , we have sV,1 |V 0 = sV,2 |V 0 for some open neighborhood V 0 ⊆ V containing y. Let
U 0 = f −1 (V 0 ) ⊆ U . Then x ∈ U 0 and tU,1 |U 0 = tU,2 |U 0 so t1 = t2 in (f −1 F )x . Thus λ is an
isomorphism.
Finally, the sheafification map f −1 F → f ∗ F induces a map on stalks (f −1 F )x → (f ∗ F )x
such that the composition (f −1 F )x → (f ∗ F )x ∼ = Fy is precisely the isomorphism λ. This
shows (f ∗ F )x ∼ F
= y as desired.

376
16.1. Direct and Inverse Image Chapter 16. Cohomology

Proposition 16.1.12. For any continuous map f : X → Y , f ∗ is a functor ShY → ShX .


Proof. By definition, f ∗ F is a sheaf on X for all sheaves F on Y , so we need only prove
naturality. Let ϕ : F → G be a morphism of sheaves on Y and fix an open set U ⊆ X. Then
the maps ϕV : F (V ) → G(V ) over all open V ⊆ Y containing U assemble into a map

(f −1 F )(U ) = lim F (V ) −→ lim G(V ) = (f −1 G)(U ).


−→ −→

Passing to the sheafification, we get a map f ∗ ϕ : f ∗ F → f ∗ G.


Corollary 16.1.13. For any f : X → Y , the pullback functor f ∗ : ShY → ShX is exact.
Proof. Let 0 → F → G → H → 0 be an exact sequence of sheaves on Y . Then for all
y ∈ Y , the sequence on stalks 0 → Fy → Gy → Hy → 0 is exact. Thus Proposition 16.1.11
says that 0 → (f ∗ F )x → (f ∗ G)x → (f ∗ H)x → 0 is exact for all x ∈ X, with f (x) = y. By
definition this means 0 → f ∗ F → f ∗ G → f ∗ H → 0 is exact.
Proposition 16.1.14. For any f : X → Y and any sheaf F on Y , f ∗ F is the sheaf of
sections of the map EF ×Y X → X, where EF → Y is the étale space of F .
Proof. Set E = EF ×Y X and let Ef ∗ F be the étale space of f ∗ F = (f −1 F )sh . Then
Ef ∗ F = Ef −1 F so it’s enough to show that Ef −1 F = E. The maps λ : (f −1 F )x → Ff (x) for
all x ∈ X in the proof of Proposition 16.1.11 induce a map of sets
a a
Φ : Ef −1 F = (f −1 F )x −→ Fy = EF .
x∈X y∈Y

It’s easy to check that Φ is continuous with respect to the topology on the étale spaces.
Moreover, the diagram
Ef −1 F EF

f
X Y

commutes by construction, so by the universal property of pullbacks, we get a map Ψ :


Ef −1 F → E = EF ×Y X. Notice that the fibres of Ef −1 F and E are the same and Ψ induces
bijections on these fibres. Hence Ψ is a continuous bijection. Finally, the fact that Ef −1 F →
X and E → X are both local homeomorphisms ensures that Ψ is a homeomorphism.
Example 16.1.15. Let f : X → Y be continuous and F = FS the sheaf of locally constant
functions on Y with values in S. Then EF = S × Y → Y where S has the discrete topology.
By Proposition 16.1.11, Ef ∗ F = EF ×Y X = (S×Y )×Y X = S×X, which by Theorem 13.1.14
has sheaf of sections FS → X, the sheaf of locally constant functions on X. So pullback of
a locally constant sheaf is again locally constant.
Example 16.1.16. Let i : X ,→ Y be the inclusion of a subspace. Then i∗ is just the
restriction functor F →
7 F |X .

377
16.1. Direct and Inverse Image Chapter 16. Cohomology

Theorem 16.1.17. For a continuous map f : X → Y , (f ∗ , f∗ ) is an adjoint pair of functors.


Proof. Let F be a sheaf on X and G be a sheaf on Y . Then since f ∗ G = (f −1 G)sh ,
Theorem 13.1.17 gives a natural isomorphism
HomShX (f ∗ G, F ) ∼
= HomPreshX (f −1 G, F )
so it suffices to construct a natural isomorphism
HomPreshX (f −1 G, F ) ∼
= HomShY (G, f∗ F ).
Notice that for any open V ⊆ Y ,
(f∗ f −1 G)(V ) = (f −1 G)(f −1 (V )) = lim G(V 0 )
−→

over all open sets V 0 ⊆ Y containing f (f −1 (V )). But since V itself is an open set containing
f (f −1 (V )), we get a map G(V ) → lim G(V 0 ) which is compatible with restriction. This
−→
defines a natural transformation
η : idShY −→ f∗ f −1
which will be the unit of our adjunction. Next, we construct a counit of the adjunction. For
any open U ⊆ X, we have
(f −1 f∗ F )(U ) = lim(f∗ F )(V ) = lim F (f −1 (V ))
−→ −→

where the limit is over all open V ⊆ Y containing f (U ). Well for every such V , f −1 (V ) ⊇ U
so the restriction maps F (f −1 (V )) → F (U ) assemble to give a map lim F (f −1 (V )) → F (U ).
−→
This defines the counit
ε : f −1 f∗ −→ idPreshX .
Finally, by Lemma 9.2.6 it remains to show that η and ε satisfy the triangle identities. For
an open set V ⊆ Y , we have
f∗ εF ◦ ηf∗ F (V ) : f∗ F (V ) → f∗ f −1 f∗ F (V ) → f∗ F (V )
which is F (f −1 (V )) → lim F (f −1 (V 0 )) → F (f −1 (V ))
−→

where the limit in the middle is over all open V 0 ⊆ Y containing f (f −1 (V )). Since V ⊇
f (f −1 (V )), this composition is clearly the identity on F (f −1 (V )). Therefore f∗ εF ◦ ηf∗ F is
the identity on f∗ F . Likewise, for U ⊆ X open,
εf −1 G ◦ f −1 ηG (U ) : f −1 G(U ) → f −1 f∗ f −1 G(U ) → f −1 G(U )
which is lim G(V ) → lim G(V 0 ) → lim G(V )
−→ −→ −→

where the outer limits are over all open V ⊆ Y containing f (U ) and the middle limit is
over all open V 0 ⊆ Y containing f (f −1 (V )) for any V containing f (U ). Any such V in
the first limit also appears in the second limit, so the composition is given by the trivial
restrictions G(V ) → G(V ) for all V . That is, the composition is the identity on lim G(V ) =
−→
f −1 G(U ).

378
16.1. Direct and Inverse Image Chapter 16. Cohomology

f g
Corollary 16.1.18. For any morphisms X → − Z, (g ◦ f )∗ = f ∗ ◦ g ∗ .
− Y →

Proof. Each of (g ◦ f )∗ and f ∗ ◦ g ∗ is a left adjoint of (g ◦ f )∗ = g∗ ◦ f∗ .

Corollary 16.1.19. Let f : X → Y be a continuous map and E an injective sheaf on X.


Then f∗ E is an injective sheaf on Y .

Proof. Suppose 0 → F → G is an exact sequence in ShY . Then since f ∗ is exact, 0 → f ∗ F →


f ∗ G is an exact sequence in ShX and by adjointness of (f ∗ , f∗ ), we have a commutative
diagram
Hom(G, f∗ E) Hom(F, f∗ E) 0


= ∼
=

Hom(f ∗ G, E) Hom(f ∗ F, E) 0

with isomorphisms on the columns. Since E is injective, the bottom row is exact. Therefore
the top row is also exact, which proves f∗ E is injective.

Remark. More generally, any functor with an exact left adjoint preserves injectives.

Corollary 16.1.20. For any continuous map f : X → Y and sheaf F on X, there is a


homomorphism
H p (X, F ) −→ H p (Y, f∗ F )
for all p ≥ 0.

Proof. Take an injective resolution E • of F in ShX . Then by Corollary 16.1.19, f∗ E • is an


injective resolution of f∗ F in ShY , so we can compute cohomology using this resolution:

H p (Y, f∗ F ) = H p (Γ(Y, f∗ E • )).

Since each f∗ E n is defined locally by a direct limit lim E n (f −1 (V )), there is an induced map
−→
of global sections Γ(X, E n ) → Γ(Y, f∗ E n ) for each n ≥ 0. This induces the desired maps on
cohomology: H p (X, F ) −→ H p (Y, f∗ F ).

Remark. To compute H p (X, f ∗ G) for a sheaf G on Y is harder in general, since if E • is an


injective resolution of G in ShY , the inverse images f ∗ E • need not form an injective resolution
of f ∗ G on X. However, by the theory of derived categories, there exists an injective complex
of sheaves J • on X and an embedding f ∗ E • → J • which is a quasi-isomorphism, i.e. it
induces isomorphisms on all cohomology groups. Therefore the cohomology of f ∗ G can be
computed with this resolution:

H p (X, f ∗ G) = H p (Γ(X, J • )).

Further, there are chain maps

Γ(Y, E • ) → Γ(X, f ∗ E • ) → Γ(X, J • )

379
16.1. Direct and Inverse Image Chapter 16. Cohomology

which induce a homomorphism on cohomology

H p (Y, G) −→ H p (X, f ∗ G)

for all p ≥ 0. In principle these can be defined directly from the definitions of sheaf coho-
mology and f ∗ , but the derived categorical perspective makes this straightforward.

One thing to note about the direct image functor is that it does not preserve stalks.
Indeed, if f : X → Y is an arbitrary morphism and y ∈ Y , then for any sheaf F on X,
(f∗ F )y = Fx if y = f (x) ∈ f (X), but if y 6∈ f (X), then (f∗ F )y need not be 0 (the value of F
over the empty set). One way to remedy this is to define a new type of pushforward functor
that respects stalks. We start by giving the definition for open immersions.

Definition. Let j : U ,→ X be an open immersion. For a sheaf F on U , define the ex-


tension by zero of F along j to be the sheafification j! F of the presheaf P on X defined
by (
F (V ), V ⊆ U
P (V ) =
0, V 6⊆ U.

Lemma 16.1.21. For any open immersion j : U ,→ X, the assignment F 7→ j! F is a functor


ShU → ShX .

Proposition 16.1.22. For any open immersion j : U ,→ X, (j! , j ∗ ) is an adjoint pair of


functors.

Proof. Let F be a sheaf on U and G a sheaf on X. First recall that the inverse image functor
j ∗ is simply restriction: j ∗ G = G|U . Then for the presheaf P defined above, we have

HomShX (j! F, G) ∼
= HomPreshX (P, G) by the universal property of sheafification

= HomShU (F, G|U ) by definition of P
= HomShU (F, j ∗ G).

These isomorphisms are all natural in F and G, so (j! , j ∗ ) is an adjoint pair.

Corollary 16.1.23. For any open immersion j : U ,→ X, j! is exact and j ∗ preserves


injectives.

Proof. Exactness of j! can be checked on stalks, and the second statement is a consequence
of the remark following Corollary 16.1.19.
If j : U ,→ X is an open immersion, the above gives us a triple of functors (j! , j ∗ , j∗ ) in
which each consecutive pair is an adjunction and j ∗ is exact. In a similar fashion, we can
extend the adjunction (i∗ , i∗ ) for a closed embedding i : Z ,→ X.

Definition. Let i : Z ,→ X be a closed immersion. For a sheaf F on X, define the inverse


image of F supported on Z to be the sheaf i! F whose value on an open set V ⊆ Z is

(i! F )(V ) = ΓZ (V, F ) := {s ∈ Γ(V, F ) | supp(s) ⊆ V }.

380
16.1. Direct and Inverse Image Chapter 16. Cohomology

Lemma 16.1.24. For any closed immersion i : Z ,→ X, F 7→ i! F is a functor ShX → ShZ .


Proposition 16.1.25. For any closed immersion i : Z ,→ X, (i∗ , i! ) is an adjoint pair of
functors.
Corollary 16.1.26. For any closed immersion i : Z ,→ X, i∗ is exact and i! preserves
injectives.
Therefore we have a triple of functors (i∗ , i∗ , i! ) in which each consecutive pair is an
adjunction and i∗ is exact. We remark that the functor f ! does not exist in general –
although a derived analogue does exist and agrees with the derived functor of i! for a closed
embedding.
Let i : Z ,→ X be a closed immersion, set U = X r Z and let j : U ,→ X be the
corresponding open immersion. Define a category C whose objects are triples (G, H, ϕ)
where G is a sheaf on Z, H is a sheaf on U and ϕ : G → i∗ j∗ H is a morphism of sheaves, and
whose morphisms are pairs of morphisms (ξ : G → G0 , χ : H → H 0 ) making the diagram
ϕ
G i∗ j∗ H

ξ i∗ j∗ (χ)

G0 0
i∗ j∗ H 0
ϕ

commute. Let η : F → j∗ j ∗ F be the unit of the adjunction (j ∗ , j∗ ). Then this induces a


morphism of sheaves i∗ (η) : i∗ F → i∗ j∗ j ∗ F .
Theorem 16.1.27 (Recollment of Sheaves). For i : Z ,→ X and j : U = X r Z ,→ X as
above,
ShX −→ C
F 7−→ (i∗ F, j ∗ F, i∗ (η))
is an equivalence of categories.
When i : Z ,→ X is a closed immersion and G is a sheaf on Z, note that the stalks of
the direct image sheaf F = i∗ G are precisely
(
Gx , x ∈ Z
Fx =
0, x 6∈ Z.
Since i∗ preserves stalks, this shows that the counit ε : i∗ i∗ G → G of the adjunction (i∗ , i∗ ) is
in fact an isomorphism. Hence i∗ G = F = i! G. On the other hand, for the open embedding
j : U = X r Z ,→ X, we have
(
Gx , x ∈ U
(j∗ G)x =
0, x 6∈ U .
Therefore j∗ G 6= j! G in general. This presents an obstacle to constructing the adjunctions
(f! , f ∗ , f∗ , f ! ) for more general morphisms f .

381
16.1. Direct and Inverse Image Chapter 16. Cohomology

Definition. A subset W ⊆ X is locally closed if

(1) W is open in W .

(2) W = U ∩ C where U ⊆ X is open and C ⊆ X is closed.

(3) For all x ∈ W , there is a neighborhood V ⊆ X of x such that V ∩ W is closed in V .

Proposition 16.1.28. The adjunctions (i! , i∗ , i∗ , i! ) exist for any locally closed immersion
i : W ,→ X.

Proof. We have an open immersion W ,→ W and a closed immersion W ,→ X, so apply


Theorem 16.1.27.
Conversely:

Proposition 16.1.29. Let i : W ,→ X be the inclusion of an arbitrary subset. If every sheaf


on X is of the form i! G for some G ∈ ShW , then W is locally closed in X.

Theorem 16.1.30. Let i : Z ,→ X be a closed immersion, set U = X rZ and let j : U ,→ X


be the corresponding open immersion. Then for any sheaf F on X, the sequence

0 → j! j ∗ F → F → i∗ i∗ F → 0

is exact.

Proof. By definition, it suffices to check this on stalks: for x ∈ X, the sequence is


( id
0 → Fx − → Fx → − 0 → 0, x ∈ U
id
0→0→
− Fx −
→ Fx → 0, x ∈ Z.

Remark. Note that when j : Z ,→ X is a closed immersion, the morphism of sheaves


j! G → j∗ G is injective for all G ∈ ShZ , so for every open set V ⊆ X, we may view Γ(V, j! G)
as a subset of Γ(V, j∗ G) = Γ(V ∩ U, G), where U = X r Z. Explicitly,

Γ(V, j! G) = {s ∈ Γ(V ∩ U, G) | supp(s) is closed in U }.

This explains another common name, ‘sections with compact support’, for j! in the topolog-
ical setting. For a morphism of schemes f : X → Y , this suggests defining the pushforward
with proper support f! : ShZ → ShX by

(f! F )(V ) = {s ∈ Γ(f −1 (V ), F ) | f |supp(s) : supp(s) → V is proper}.

(Recall that a morphism is proper if it is separated, of finite type and is a closed morphism
under all base changes. This is in contrast to the topological definition of properness: the
preimage of every compact set is compact.)

382
16.2. Acyclic Sheaves Chapter 16. Cohomology

16.2 Acyclic Sheaves


Injective resolutions are hard to come by in sheaf theory, but Theorem 13.3.1 shows that
sheaf cohomology can be computed via acyclic resolutions. This gives us the following result,
which will be useful in reducing cohomology computations to simpler subspaces.
Proposition 16.2.1. Let i : Z ,→ X be a closed immersion. Then for any sheaf of abelian
groups F on Z,
H i (Z, F ) ∼
= H i (X, i∗ F )
for all i ≥ 0.
Proof. Suppose E • is an injective resolution of F on Z. Then by exactness of i∗ , i∗ E • is
an injective resolution of i∗ F on X and Γ(Z, E • ) = Γ(X, i∗ E • ) so the cohomology groups
agree.
Recall (Section 13.2) that a sheaf F on X is flasque if for all open sets V ⊆ U , the
restriction F (U ) → F (V ) is surjective.
Lemma 16.2.2. Let (X, OX ) be any ringed space. Then any injective OX -module is flasque.
Proof. For any open set U , let j : U ,→ X be the inclusion map and set OU = j! (OX |U ).
Then OU is an OX -module. Suppose F is an injective OX -module and V ⊆ U are open sets
in X. Then there is an induced map OV ,→ OU which is an inclusion since j! is exact, so
by the injective property for F , HomOX (OU , F ) → HomOX (OV , F ) is surjective. Finally, the
adjunction (j! , j ∗ ) gives us an isomorphism

HomOX (OU , F ) ∼
= HomOX |U (OX |U , j ∗ F ) = F (U ).

Hence F (U ) → F (V ) is surjective, so F is flasque.


Theorem 16.2.3. Let F be a flasque sheaf on X. Then F is acyclic.
Proof. Since ShX has enough injectives by Theorem 13.2.10, there is an exact sequence of
sheaves
0→F →G→H→0
where G is injective. By Theorem 13.2.11,

0 → F (X) → G(X) → H(X) → 0

is also exact, so in the long exact sequence in sheaf cohomology, we have

0 → F (X) → G(X)  H(X) → H 1 (X, F ) → H 1 (X, G) = 0

where the last term is zero since G is injective, and thus acyclic. Therefore exactness implies
H 1 (X, F ) = 0. Now for i ≥ 2, inductively assume that H i−1 (X, Q) = 0 for all flasque sheaves
Q on X. Part of the long exact sequence looks like

0 = H i−1 (X, G) → H i−1 (X, H) → H i (X, F ) → H i (X, G) = 0,

383
16.2. Acyclic Sheaves Chapter 16. Cohomology

once again using that G is injective. Thus H i (X, F ) ∼


= H i−1 (X, H). Meanwhile, Lemma 16.2.2
shows that G is flasque, so the second statement in Theorem 13.2.11 implies H is also flasque
and hence H i−1 (X, H) = 0 by the inductive hypothesis. This proves H i (X, F ) = 0 for all
i ≥ 2.
Remark. If (X, OX ) is any ringed space, Theorem 16.2.3 shows that flasque sheaves are
acyclic for the functor Γ : ModOX → Ab, so the derived functors of Γ agree with the sheaf
cohomology H i (X, −) for sheaves of abelian groups.
Theorem 16.2.4 (Grothendieck Vanishing). Let X be a noetherian topological space of
dimension n. Then for any sheaf of abelian groups F on X,

H i (X, F ) = 0

for all i > n.


Proof. First suppose X is reducible; let Z be one of its irreducible components and set
U = X r Z. Following the notation above, let j : U ,→ X be the open inclusion and set
FU = j! (F |U ). Likewise, let i : Z ,→ X be the closed inclusion and set FZ = i∗ (F |Z ). Then
by Theorem 16.1.30, there is an exact sequence

0 → FU → F → FZ → 0.

By the long exact sequence in sheaf cohomology, it’s enough to prove H i (X, FU ) = 0 and
H i (X, FZ ) = 0 for all i > n. Note that FU extends to a sheaf on U , which is a closed
subset with one less irreducible component than X. Applying Proposition 16.2.1 to U ,→ X
and noting that Z is closed and irreducible, we may reduce to the case where X is itself
irreducible.
We induct on n = dim X. If n = 0, Γ : ShX → Ab is an equivalence of categories (X is
irreducible) and in particular exact, so Ri Γ is trivial for i > 0. For n > 0, note that we have

F = lim FB
−→
S
where the limit is over all finite subsets B ⊆ U ⊆X F (U ), directed by inclusion, and FB is
the subsheaf of F generated by the sections in B. Since cohomology commutes with direct
limits, it suffices to prove H i (X, FB ) = 0 for all B. Suppose B 0 ⊆ B and set q = |B r B 0 |.
Then there is an exact sequence

0 → FB 0 → FB → G → 0

where G is a sheaf generated by q sections over some open subsets of X. Using the long
exact sequence coming from this sequence, we may reduce to the case where q = 1. In this
case we have a short exact sequence

0 → K → ZU → F → 0

for some open set U ⊆ X, where Z denotes the (locally) constant sheaf on X with values
in Z. Thus we may prove the theorem by showing that H i (X, ZU ) = 0 for any open U and

384
16.2. Acyclic Sheaves Chapter 16. Cohomology

H i (X, K) = 0 for any nonzero subsheaf K ⊆ ZU . For such a subsheaf K, we have that Kx
is a subgroup of ZU,x for all x ∈ U , so there is some open set V ⊆ U for which K|V ∼
= dZ|V .

Abstractly, dZV = ZV so there is a short exact sequence

0 → ZV → K → Q → 0.

The quotient sheaf Q is supported on U r V which has dimension < n since X is irreducible.
Therefore H i (X, Q) = 0 for all i ≥ n by the inductive hypothesis. Thus it remains to show
H i (X, ZU ) = 0 for i > n and all U (which implies vanishing for ZV if V ⊆ U ). Set Z = X rU .
Since X is irreducible, dim Z < n so by induction, H i (X, ZZ ) = 0 for i ≥ n. On the other
hand, Z is obviously flasque so Theorem 16.2.3 shows that H i (X, Z) = 0 for i > 0. Thus the
long exact sequence induced by the short exact sequence

0 → ZU → Z → ZZ → 0

gives us H i (X, ZU ) ∼
= H i−1 (X, ZZ ) which is 0 for all i − 1 ≥ n. This finishes the proof.
Suppose U = {Ui } is an open covering of a space X and F is a sheaf of abelian groups
p
on X. By Theorem 13.4.9, there is a natural map Ȟ (U, F ) → H p (X, F ) for every p ≥ 0.
p
Proposition 16.2.5. If F is a flasque sheaf on X, then Ȟ (U, F ) = 0 for all open covers
U of X and all p > 0.

Proof. Let C • (U, F ) be the Čech resolution of F with respect to U. Recall that C p (U, F ) is
defined for any open V ⊆ X by
Y
Γ(V, C p (U, F )) = F (V ∩ Ui0 ,...,ip )
i0 ,...,ip

where Ui0 ,...,ip = Ui0 ∩ · · · ∩ Uip . For each of these Ui0 ,...,ip , the sheaf V 7→ F (V ∩ Ui0 ,...,ip ) is
flasque and since products of flasque sheaves are flasque, the entire sheaf C p (U, F ) is flasque.
Therefore C p (U, F ) is acyclic by Theorem 16.2.3, so the Čech resolution can used to compute
sheaf cohomology:
p
Ȟ (U, F ) = H p (C • (U, F )) = H p (X, F ) = 0.

Definition. A sheaf F on X is acyclic for an open cover U = {Ui } if for all p > 0,
H p (Ui0 ,...,ip , F |Ui0 ,...,ip ) = 0.

Such covers exist in many common situations, e.g. for manifolds with a Riemannian
metric. The following result allows for relatively easy computations of sheaf cohomology
when an acyclic cover is available.

Theorem 16.2.6 (Leray). If F is a sheaf on X which is acyclic for an open cover U, then
the maps
p
Ȟ (U, F ) −→ H p (X, F )
are isomorphisms for all p ≥ 0.

385
16.2. Acyclic Sheaves Chapter 16. Cohomology

0
Proof. We prove this by induction on p. For p = 0, this is clear since both Ȟ (U, F ) and
j
H 0 (X, F ) are Γ(X, F ). Now assume Ȟ (U, G) → H j (X, G) is an isomorphism for all j ≤ p
and all acyclic sheaves G. For our given F , take a short exact sequence of sheaves
0→F →E→Q→0
where E is injective. Set U = Ui0 ,...,ip . Then by the hypothesis on F , H i (U, F |U ) = 0 for all
i > 0. Since E is injective, and thus flasque by Lemma 16.2.2, we have H i (U, E|U ) = 0 as
well. Thus in the long exact sequence induced by the above sequence, we can see
0 = H i (U, E|U ) → H i (U, Q) → H i+1 (U, F ) = 0.
This shows H i (U, Q) = 0 for all i ≥ 1, so by taking the product over all such Ui0 ,...,ip , we
conclude that Q is acyclic for U. Therefore, the sequence of complexes of sheaves
0 → C • (U, F ) → C • (U, E) → C • (U, Q) → 0
is exact and induces a long exact sequence
p−1 p−1 p p
· · · → Ȟ (U, E) → Ȟ (U, Q) → Ȟ (U, F ) → Ȟ (U, E) → · · ·
p−1 p
Now since E is flasque, Proposition 16.2.5 shows that Ȟ (U, E) = Ȟ (U, E) = 0. Therefore
we have a diagram with exact rows
p−1 p
0 Ȟ (U, Q) Ȟ (U, F ) 0

0 H p−1 (X, Q) H p (X, F ) 0


p
By induction, the left column is an isomorphism, so it follows that Ȟ (U, F ) → H p (X, F ) is
an isomorphism.
1
Corollary 16.2.7. For any sheaf F on X, the map Ȟ (X, F ) → H 1 (X, F ) is an isomor-
phism.
Proof. Let E be an injective sheaf and 0 → F → E → Q → 0 a short exact sequence of
sheaves on X. Fix an open cover U of X. Then the morphism of complexes of sheaves
C • (U, F ) → C • (U, E) is injective; let K• (U) be its cokernel, so that we have a short exact
sequence of complexes
0 → C • (U, F ) → C • (U, E) → K• (U) → 0.
1
Then by Proposition 16.2.5, Ȟ (U, E) = 0 and there is a diagram with exact rows:
1
0 F (X) E(X) H 0 (K• (U)) Ȟ (U, F ) 0

id id

0 F (X) E(X) Q(X) H 1 (X, F ) 0

386
16.2. Acyclic Sheaves Chapter 16. Cohomology

After passing to the direct limit over all U, it’s enough to show that

α : lim H 0 (K• (U)) → H 0 (X, Q)


−→

0
is an isomorphism. Injectivity is easy. For surjectivity, take s ∈ H 0Q (X, Q) = Ȟ (X, Q).
Then there exists an open cover U = {Ui } of X such that s = (si ) ∈ E(Ui ) and these si
satisfy si − sj ∈ ker(F (Ui ∩ Uj ) → E(Ui ∩ Uj )) = F (Ui ∩ Uj ) for all overlaps Ui ∩ Uj . By
0
definition this (si ) defines a Čech 0-cocycle in Ȟ (U, E) whose image in H 0 (K• (U)) maps to
s along α. Hence α is an isomorphism.
Definition. The support of a morphism of sheaves ϕ : F → G is the closed set

supp(ϕ) = {x ∈ X | ϕx 6= 0}

where ϕx : Fx → Gx is the map on stalks.


Definition. An open cover U of X is locally finite if every x ∈ X has a neighborhood
which has nonempty intersection with only finitely many elements of U. The space X is
paracompact if every open cover has a locally finite refinement.
Example 16.2.8. All metric spaces are paracompact.
Definition. Let U = {Ui }i∈I be a locally finite cover of X and F a sheaf on X. A partition
of unity for F subordinate to U is a collection of morphisms P of sheaves {ηi : F → F }i∈I
such that for each i ∈ I, supp(ηi ) ⊆ Ui , and for all x ∈ X, i∈I ηi,x = idFx .
Note that when U is locally finite, only
P finitely many of the maps ηi,x : Fx → Fx are
nonzero for any given x ∈ X, so the sum i∈I ηi,x is always defined.
Definition. A sheaf F on X is fine if it admits partitions of unity subordinate to any locally
finite cover of X.
Proposition 16.2.9. Let 0 → F → G → H → 0 be a short exact sequence of sheaves on
a paracompact space X. If F is fine, then 0 → F (X) → G(X) → H(X) → 0 is exact.
Moreover, if every open set U ⊆ X is paracompact and G is flasque, then H is flasque as
well.
Proof. Similar to the proof of Theorem 13.2.11.
Corollary 16.2.10. Let F be a fine sheaf on X. Then F is acyclic.
Proof. Similar to the proof of Theorem 16.2.3.
Example 16.2.11. Let (X, OX ) be a ringed space. Assume U = {Ui }i∈I is a locally finite
open cover and OX admits a partition of unity subordinate to U; equivalently,
Passume there
exist sections {si ∈ OX (X)}i∈I for which supp(si ) ⊆ Ui for all i ∈ I and i∈I si = 1 in
OX (X). Let F be an OX -module. Then the maps ηi,U : F (U ) → F (U ), t 7→ si |U · t induce
a partition of unity {ηi : F → F } for F subordinate to U. Hence sheaves of modules over a
fine sheaf of rings are also fine.

387
16.3. Noetherian Scheme Cohomology Chapter 16. Cohomology

16.3 Noetherian Scheme Cohomology


Let A be a noetherian ring, X = Spec A the corresponding affine scheme and let I ⊂ A be
an ideal. For any A-module M , set ΓI (M ) = {m ∈ M | I n m = 0 for some n > 0} which is
a submodule of M .

Lemma 16.3.1. If E is an injective A-module, then ΓI (E) is also injective.

Proof. We prove this using Baer’s criterion (Theorem 8.2.3); let J ⊂ A be an ideal and
ϕ : J → ΓI (E) a homomorphism. Since J is finitely generated, there is some N ∈ N such
that I N ϕ(J) = 0, or ϕ(I N J) = 0 since ϕ is a homomorphism. By Corollary 5.4.7, there is
some n ≥ N such that I N J ⊇ J ∩ I n , so ϕ(J ∩ I n ) = 0. Consider the diagram

A A/I n
ψ0

J J/(J ∩ I n ) ΓI (E) E
ϕ

Since E is injective, we get an induced map ψ 0 as in the diagram, but then I n im ψ 0 = 0 so


ψ 0 really lands in ΓI (E) and thus we get the dashed arrow A/I n → ΓI (E). Composing with
the quotient map along the top row gives the lift ψ : A → ΓI (E) of ϕ needed for Baer’s
criterion.
For any A-module M , let M
f be the associated sheaf on X, which is coherent by Theo-
rem 14.4.7.

Proposition 16.3.2. If E is an injective A-module, then E


e is flasque.

Proof. Apply Proposition 14.4.3 and Lemma 16.2.2.

Theorem 16.3.3. Let A be a noetherian ring, X = Spec A and suppose F is a quasi-coherent


sheaf on X. Then for all i > 0, H i (X, F ) = 0.

Proof. Consider the A-module M = Γ(X, F ) and take an injective resolution E • of M . Then

0→M e0 → E
f→E e1 → · · ·

is an exact sequence of sheaves on X. Since F = M f by Proposition 14.4.2 and each E i is


flasque by Proposition 16.3.2, the cohomology of the deleted complex 0 → E e • is H i (X, F ).
But applying Γ to 0 → E e • yields 0 → E • which has cohomology H 0 (E • ) = M and H i (E • ) =
0 for i ≥ 1 since the E i are injective.

Corollary 16.3.4. Let X be a noetherian scheme. Then every quasi-coherent sheaf on X


can be embedded in a flasque quasi-coherent sheaf on X.

388
16.3. Noetherian Scheme Cohomology Chapter 16. Cohomology

S
Proof. Write X = Ui where Ui = Spec Ai for a noetherian ring Ai . For a quasi-coherent
sheaf F , there is an Ai -module Mi such that F |Ui = M fi for each I. Choose an injective
L
Ai -module Ei for which Mi ,→ Ei . Set E = (fi )∗ E ei where fi : Ui ,→ X is the canonical
inclusion. The morphisms F |Ui ,→ Ei glue to give a morphism F → (fi )∗ E
e ei , so by taking
the direct sum, there is a morphism F → E which is injective. For each i, Corollary 16.1.19
shows that (fi )∗ E
ei is an injective sheaf on X, hence flasque (Lemma 16.2.2). Moreover,
Lemma 14.4.8 shows each (fi )∗ E ei is quasi-coherent, so taking the direct sum gives a flasque,
quasi-coherent sheaf E into which F embeds.

Corollary 16.3.5. Let X be a noetherian scheme, U = {Ui } an affine open cover of X and
F a quasi-coherent sheaf on X. Then for all p ≥ 0, the natural map
p
Ȟ (U, F ) −→ H p (X, F )

is an isomorphism.

Proof. The p = 0 case was Lemma 13.4.5. For p ≥ 1, use Corollary 16.3.4 to embed F in
a flasque quasi-coherent sheaf G and let H = G/F be the quotient sheaf. Since F and G
are quasi-coherent and each Ui0 ,...,ip = Ui0 ∩ · · · ∩ Uip is affine, Theorem 16.3.3 implies that
H i (Ui0 ,...,ip , F ), H i (Ui0 ,...,ip , G) and H i (Ui0 ,...,ip , H) are all zero for i ≥ 1. Therefore the short
exact sequence 0 → F → G → H → 0 stays exact over each overlap:

0 → F (Ui0 ,...,ip ) → G(Ui0 ,...,ip ) → H(Ui0 ,...,ip ) → 0.

Taking the product over all such overlaps, we see that the sequence of Čech complexes

0 → C • (U, F ) → C • (U, G) → C • (U, H) → 0

is exact. This induces a long exact sequence


0 0 0 1 1
0 → Ȟ (U, F ) → Ȟ (U, G) → Ȟ (U, H) → Ȟ (U, F ) → Ȟ (U, G) = 0

with the final term vanishing since G is flasque (Proposition 16.2.5). Also note that for the
i i+1
same reason there will be isomorphisms Ȟ (U, H) ∼ = Ȟ (U, F ) for all i ≥ 1. Since H is
1 ∼
quasi-coherent and Ȟ (U, F ) −→ H 1 (X, F ) is an isomorphism by Corollary 16.2.7, the result
for all i ≥ 1 follows by induction.

Theorem 16.3.6 (Serre). For a noetherian scheme X, the following are equivalent:

(a) X is affine.

(b) For any quasi-coherent sheaf F on X, H i (X, F ) = 0 for all i > 0.

(c) For any coherent sheaf of ideals I, H 1 (X, I) = 0.

(d) There exist elements fi ∈ Γ(X, OX ) such that each D(fi ) = {x ∈ X | fi (x) 6= 0} is
affine and the ideal generated by the fi is Γ(X, OX ).

389
16.3. Noetherian Scheme Cohomology Chapter 16. Cohomology

Proof. (a) =⇒ (b) is Theorem 16.3.3.


(b) =⇒ (c) is trivial.
(c) =⇒ (d) Let x ∈ X be a closed point and U = Spec A an affine open subscheme
containing x. Then Z = X r U is a closed subscheme defined by a (necessarily coherent)
sheaf of ideals I, and likewise Z ∪ {x} is defined by a coherent sheaf of ideals J. Then J ⊆ I
and I/J is supported on {x}. By hypothesis, H 1 (X, J) = 0 so the long exact sequence in
sheaf cohomology for 0 → J → I → I/J → 0 is

0 → Γ(X, J) → Γ(X, I) → Γ(X, I/J) → 0

and thus Γ(X, I) → Γ(X, I/J) is surjective. This means there is some f ∈ Γ(X, I) such that
f (x) 6= 0, so x ∈ D(f ) ⊆ U . Note that D(f ) = DU (f |U ) is indeed affine. Repeating this
for all closed pointsLof X, we get a family (fi )i∈S ⊂ Γ(X, OL X ). These define a morphism
of OX -modules ϕ : i∈S OX → OX such that each ϕD(fi ) : i∈S OX (D(fi )) → OX (D(fi ))
is surjective, L
but since the D(fi ) cover X, ϕ itself is surjective. The short exact sequence
0 → ker ϕ → i∈S OX → OX → 0 induces a long exact sequence
!
Γϕ
M
0 → Γ(X, ker ϕ) → Γ X, OX −→ Γ(X, OX ) → H 1 (X, ker ϕ) = 0
i∈S

by hypothesis, so Γϕ is surjective and thus the fi generate Γ(X, OX ).


(d)S =⇒ (a) Since (fi ) generates the unit ideal in A := Γ(X, OX ), we must have
X = D(fi ). There is a natural morphism of schemes g : X → Spec A associated to
idA under the correspondence HomSch (X, Spec A) ∼ = HomRing (Γ(X, OX ), A). Let gi be the
restriction g|D(fi ) : D(fi ) → DA (fi ) where DA (f ) is the principal open subset of Spec A
defined in Theorem 14.1.7. By the same theorem, Γ(D(fi ), OX ) ∼ = Afi so gi is an isomorphism
on global sections and is therefore an isomorphism by Proposition 14.2.2 since D(fi ) and
DA (fi ) are both affine. This implies g is an isomorphism, so X itself is affine.

390
16.4. Cohomology of Projective Space Chapter 16. Cohomology

16.4 Cohomology of Projective Space


Let A be a noetherian ring, S = A[x0 , . . . , xn ] and X = Proj S = PnA the projective n-space
over A defined in Section 14.6. For each d ∈ Z, set O(d) = OX (d).
Theorem 16.4.1. For a noetherian ring A and X = PnA ,
(a) There is an isomorphism of graded S-modules

S∼
M
= H 0 (X, O(d)).
d∈Z

In particular, H 0 (X, O(d)) = 0 for d < 0.

(b) For all d ∈ Z and 0 < p < n, H p (X, O(d)) = 0.

(c) H n (X, O(−n − 1)) ∼


= A.
(d) There is a perfect pairing of free A-modules

H 0 (X, O(d)) × H n (X, O(−d − n − 1)) −→ H n (X, O(−n − 1))

for all d ∈ Z.
L
Proof. By (3) of Proposition 14.6.11, each O(d) is quasi-coherent, so the sheaf F = d∈Z O(d)
is quasi-coherent as well. Moreover, since cohomology comutes with direct sums, we have
M
H p (X, O(d)) = H p (X, F )
d∈Z

for all p ≥ 0.
(a) For 0 ≤ i ≤ n, set Ui = D+ (xi ), which is an affine open subset of X and X = ni=0 Ui .
S
p
Therefore by Corollary 16.3.5, H p (X, F ) ∼ = Ȟ (U, F ) for all p ≥ 0. For each i0 , . . . , ip ,
Ui0 ,...,ip = D+ (xi0 · · · xip ) so by (3) of Theorem 14.6.2, F (Ui0 ,...,ip ) ∼
= S(xi0 ···xip ) . This shows
that the Čech complex for F with respect to U is
Y Y
C • (U, F ) : 0 → S(xi ) → S(xi xj ) → · · · → S(x0 ···xn )
i i,j

and this is compatible with the grading on F . The kernel of the first map is on one hand
0
Ȟ (U, F ) but on the other hand one can easily see that this kernel is precisely S.
(b) We may assume n > 1 and induct on n. Localizing the above Čech complex C • (U, F )
at the element xn (as a graded module) yields
Y Y
0 → S(xn ) × S(xi xn ) → S(xn ) × S(xi xj xn ) → · · · → S(x0 ···xn )
i<n i,j<n

which is precisely the Čech complex C • (U 0 , F |Un ) where U 0 = {Ui ∩ Un } is an open cover of
p
Un . Taking cohomology of this yields Ȟ (U 0 , F |Un ) ∼
= H p (Un , F |Un ) by Corollary 16.3.5. But

391
16.4. Cohomology of Projective Space Chapter 16. Cohomology

since Un is affine and F is quasi-coherent, H p (Un , F |Un ) = 0 for p > 0 by Theorem 16.3.3,
and since localization is exact, H p (X, F )xn = 0 for all 0 < p < n. To finish, consider the
short exact sequence of graded S-modules
n x
0 → S(−1) −→ S→
− S/(xn ) → 0.

Applying the sheaf construction in Proposition 14.6.11 gives a short exact sequence of OX -
modules
0 → O(−1) → OX → OH → 0
where H ⊂ X is the hyperplane defined by xn = 0. Shifting the entire sequence by d yields
a short exact sequence
0 → O(d − 1) → O(d) → OH (n) → 0
and taking the direct sum over all d ∈ Z gives

0 → F (−1) → F → FH → 0

where FH = d∈Z OH (d). In the long exact sequence for this sequence, the map H p (X, F (−1)) →
L
H p (X, F ) is just multiplication by xn and, up to one degree of grading, H p (X, F (−1)) =
H p (X, F ). Since H ∼= Pn−1 p
A , induction gives H (X, FH ) = 0 for 0 < p < n − 1. The end of
the long exact sequence can be written as
δ x
0 → H n−1 (X, FH ) →
− H n (X, F (−1)) −→
n
H n (X, F ) → 0

but this may not be exact. However, it is clear from the description below of H n (X, F )
as a free A-module that the second map is surjective and δ is injective, so we have that

multiplication by xn is an isomorphism H p (X, F (−1)) − → H p (X, F ) for all 0 < p < n. Since
we showed H p (X, F )xn = 0, this implies H p (X, F ) = 0 for the same values of p.
(c) The last map in the Čech complex,
Y
α: S(x0 ···x̂k ···xn ) −→ S(x0 ···xn ) ,
k

n
has kernel Ȟ (U, F ) which is isomorphic to H n (X, F ) by Corollary 16.3.5. The localization
Sx0 ···xn is a free A-module with basis {xj00 · · · xjnn | ji ∈ Z}, and

im α = {xj00 · · · xjnn ∈ Sx0 ···xn | ji ≥ 0 for some 0 ≤ i ≤ n}.

Thus ker α is the free A-module generated by the basis {xj00 · · · xjnn | ji < 0 for all i}. In
degree −n−1, the only monomial with this condition is x−1 −1
0 · · · xn , so (ker α)−n−1 is a free A-
module of rank 1. But the (−n−1)st graded piece of H (X, F ) is precisely H n (X, O(−n−1)).
n

(d) If d < 0, then H 0 (X, O(d)) = 0 by (a). To see that H n (X, O(−d − n − 1)) = 0, note
that there are no negatively graded monomials of degree > −n − 1 so a similar argument to
that in (c) works. So the statement holds trivially for d < 0. If d ≥ 0,
( n
)
H 0 (X, O(d)) ∼ j
X
= A x 0 · · · xjn | ji ≥ 0 and
0 n ji = n
i=0

392
16.4. Cohomology of Projective Space Chapter 16. Cohomology

as a free A-module and the pairing is given by

(xj00 · · · xjnn ) · (xk00 · · · xknn ) = xj00 +k0 · · · xjnn +kn .

The kernel of the pairing is any monomials for which ji + ki ≥ 0 for any i, so we can see that
the dual bases {xj00 · · · xjnn } and {x−j
0
0 −1
· · · x−j
n
n −1
} induce a perfect pairing.
The remaining results in this section are stated without proof but can be found in
Hartshorne.

Definition. An invertible sheaf L on a scheme Y is called very ample (or very ample
relative to a base X) if there exists an n ∈ N and an immersion f : Y → Pn (or to PnX )
such that f ∗ O(1) ∼
= L. We say L is ample if there is some m ∈ N such that L⊗m is very
ample.

Proposition 16.4.2. An X-scheme Y is projective if and only if Y → X is proper and


there exists a very ample sheaf on Y relative to X.

Theorem 16.4.3 (Serre). Let A be a noetherian ring and X a projective A-scheme. Suppose
O(1) is a very ample sheaf on X and F is any coherent sheaf on X. Then there exists an
N ∈ N such that for all n ≥ N , F (n) is generated by finitely many global sections, i.e.
there exists a collection of sections {s1 , . . . , sr } ⊂ Γ(X, F (n)) such that for every x ∈ X, the
elements si,x ∈ F (n)x generate F (n)x as an OX,x -module.

Corollary 16.4.4. If X is a projective scheme over a noetherian ring,


Lrthen any coherent
sheaf on X can be written as a quotient of a free sheaf of the form i=1 O(di ) for some
d1 , . . . , dr ∈ Z.

Theorem 16.4.5 (Serre’s Theorem A). Let A be a noetherian ring and X a projective A-
scheme, with very ample sheaf O(1) relative to Spec A. Then for any coherent sheaf F on
X,

(a) For every i ≥ 0, H i (X, F ) is a finitely generated A-module.

(b) There exists an N ∈ N such that for each n ≥ N , H i (X, F (n)) = 0 for all i > 0.

393
Chapter 17

Curves

In this chapter we further study the geometry of algebraic varieties of dimension 1.


Definition. An irreducible, projective algebraic variety X of dimension dim X = 1 is called
an algebraic curve.
For the rest of the chapter, X will denote an algebraic curve. The first important result
is that the local rings OX,P of a nonsingular curve are discrete valuation rings.
Theorem 17.0.1. Let X be an algebraic curve and P ∈ X a nonsingular point. Then OX,P
is a DVR.
Proof. Fix P ∈ X and let OP = OX,P be the local ring at P , with maximal ideal mP and
residue field κ(P ) = OP /mP . Then by Theorem 12.6.2, OP is a regular local ring. Thus
dimκ(P ) (mP /m2P ) = dim X = 1. Let t ∈ mP such that dP t 6= 0. Then by Theorem 6.2.8, t
generates mP so for f ∈ k̄(X) with f (P ) = 0, we have f = tr u in OP , for some u ∈ OP× .
Define a map
ordP : OP −→ Z
f 7−→ ordP (f ) = max{d ∈ Z | f ∈ mdP }.
Explicitly, if f = tr u where u is a unit, then ordP (f ) = r. Formally, we also set ordP (f ) = 0
if f (P ) 6= 0, to get a map on all of k̄(X). One then shows that ordP is a discrete valuation
with OP as its valuation ring.
Corollary 17.0.2. For any nonsingular point P ∈ X, OP is a PID and therefore a UFD.
Proof. By the above, every ideal of OP is of the form (tr ) where t ∈ mP is a generator.
Definition. An element t ∈ mP such that t̄ generates mP /m2P is called a uniformizer at P .
Definition. Fix a rational function f ∈ k(X) and an integer r > 0. We say f has a pole
of order r at P if ordP (f ) = −r, and a zero of order r at P if ordP (f ) = r.
Remark. A rational function f ∈ k(X) is regular at P if and only if ordP (f ) ≥ 0.
Proposition 17.0.3. Every nonconstant, rational function f ∈ k̄(X) has at least one pole.
Proof. A rational function f ∈ k̄(X) with no poles is regular everywhere on X, and therefore
constant by Proposition 12.11.4, since X is projective.
Remark. Each f ∈ k̄(X) has only finitely many zeroes and poles, or none at all.

394
17.1. Divisors Chapter 17. Curves

17.1 Divisors
Definition. Let X be a variety. An irreducible divisor on X is a closed, irreducible
k-subvariety x of X of codimension 1.
When X is a curve over k, an irreducible divisor is a closed point of MaxSpec k[X ∩ Ui ]
for some affine patch Ui , or alternatively, a Gk -orbit of points in X(k̄).
Definition. The degree of an irreducible divisor x on X is the size of the Gk -orbit in X(k̄)
corresponding to x, i.e. deg(x) = [κ(P ) : k] for any P ∈ x.
Example 17.1.1. Let X = P1 . On an affine patch A1 ,→ Ui ⊆ P1 , the irreducible divisors
correspond to irreducible polynomials in k[A1 ] = k[t].
Definition. Let X be a curve over k. The divisor group on X, Div(X), is the free abelian
group on the set of irreducible divisors on X:
( )
X
Div(X) = D = nx x : nx ∈ Z, nx 6= 0 for finitely many x .
x∈X
P
The elements of Div(X) are called
P divisors on X. For a divisor D = x∈X nx x ∈ Div(X),
the degree of D is deg(D) = x∈X nx deg(x).
Example 17.1.2. If k is algebraically closed, then the irreducibleP
divisors are the points of
X, so each D ∈ Div(X) is a weighted sum of points of P X: D = x∈X nx x. The degree of
such a divisor is just the sum of the weights: deg(D) = x∈X nx .

P Now assume X is a nonsingular curve. For f ∈ k(X) , we can define a divisor D(f ) =
x∈X ordx (f )x, called the principal divisor of f . This defines a map

D : k(X)∗ −→ Div(X)
whose image is denoted PDiv(X), the group of principal divisors on X.
Definition. The Picard group, or divisor class group, of X is the quotient group
Pic(X) = Div(X)/ PDiv(X).
This defines an equivalence relation on divisors: D1 ∼ D2 if D1 = D2 + D(f ) for some
f ∈ k(X)∗ .
Example 17.1.3. Consider the variety E = Z(y 2 − x3 − 3x2 − 2x). This is the elliptic curve
defined by y 2 = f (x) where f = x3 + 3x2 + 2x = x(x + 1)(x + 2). The projective closure of
E is E = Z(fh ), where
fh = ZY 2 − X 3 − 3X 2 Z − 2XZ 2 .
Y
Setting y = Z
, we can compute its divisor on E:
X
D(y) = ordP (y)P.
P ∈X

On the affine part, there are only zeroes of y, and they occur precisely at P = (−2, 0), (−1, 0)
and (0, 0).

395
17.1. Divisors Chapter 17. Curves

Note that t ∈ OE,P is a uniformizer whenever dP t 6= 0. Viewing t ∈ k[x, y], i.e. as a lift of
[t] ∈ OE,P , we have that
∂f
ˆ t = x is a uniformizer as long as dP x = x|TP E 6= 0, which is equivalent to (P ) 6= 0.
∂y
∂f
ˆ t = y is a uniformizer as long as dP y = y|TP E 6= 0, that is, (P ) 6= 0.
∂x
In particular, we can always find a uniformizer! For P = (−2, 0), (−1, 0) and (0, 0), t = y is
a uniformizer. It follows that ordP (y) = 1 at each of these points, and ordQ (y) = 0 for any
other point Q ∈ E. Thus the divisor for y is

(y) = (−2, 0) + (−1, 0) + (0, 0) + ord∞ (y)∞.

The point at infinity is where Z = 0, so by the defining equation for E, X = 0 and, in


projective space, Y = 1. Set P = ∞ = [0, 1, 0]. On a different affine patch
  containing P , we
have coordinates ζ = Y and ξ = Y . Then y = ζ so ordP (Y ) = ordP ζ1 = − ordP (ζ). In
Z X 1

these coordinates, the defining equation for E becomes

g = ζ − (ξ 3 + 3ξ 2 ζ + 2ξζ 2 ).
∂g
Notice that ∂ζ
(0, 0) = 1, so ξ is a uniformizer on this patch. Now

ordP (ζ) = ordP (ξ 3 + 3ξ 2 ζ + 2ξζ 2 ) ≥ min{ordP (ξ 3 ), ordP (3ξ 2 ζ), ordP (2ξζ 2 )}.

396
17.1. Divisors Chapter 17. Curves

We have ordP (ξ 3 ) = 3 and ordP (3ξ 2 ζ), ordP (2ξζ 2 ) ≥ 3. If all three orders are equal to 3, then
by the ultrametric inequality ordP (ζ) must be strictly greater than the minimum, which is
3 in this case. But then ordP (3ξ 2 ζ) = ordP (ξ 2 ) + ordP (ζ) > 2 + 3 > 3, so in fact we cannot
have all three orders equal to 3. Hence ordP (ζ) = 3. We have thus calculated the divisor of
y on the elliptic curve E:

(y) = (−2, 0) + (−1, 0) + (0, 0) − 3∞.

397
17.2. Morphisms Between Curves Chapter 17. Curves

17.2 Morphisms Between Curves


Proposition 17.2.1. Let C be an algebraic curve and X a projective variety, and suppose
ϕ : C 99K X is a rational map. If P ∈ C is a nonsingular point then ϕ is regular at P .

Proof. A more general result is that if Y is a normal variety, i.e. the local rings OY,P are
normal rings, then the locus of nondeterminacy of such a rational map ϕ : Y 99K X is a
subvariety of codimension at least 2. For Y = C a curve, this means there are no points
where ϕ fails to be regular.
A nonconstant rational map ϕ : C1 99K C2 between curves induces a field extension
k(C2 ) ,→ k(C1 ). Since both function fields have transcendence degree 1, this is in fact a
finite field extension.

Definition. For curves C1 and C2 and a rational map ϕ : C1 99K C2 , define the degree
of ϕ by deg ϕ = [k(C1 ) : k(C2 )]; the separable degree of ϕ by degs ϕ = [k(C1 ) : k(C2 )]s ;
and the inseparable degree of ϕ by degi ϕ = [k(C1 ) : k(C2 )]i . We say ϕ is separable if
k(C1 ) ⊇ k(C2 ) is a separable extension.

Definition. Any finitely generated field extension of k with transcendence degree 1 over k
is called a function field of degree 1 over k.

Proposition 17.2.2. There is an equivalence of categories


   
nonsingular curves over k function fields of deg. 1 over k
←→ .
with nonconstant, rational maps with k-homomorphisms

Proof. (Sketch) The assignment X 7→ k(X) determines one direction: we have seen that
k(X) is indeed a function field over k. Conversely, for a function field K/k, we associate
an abstract algebraic curve XK to K by putting a Zariski topology on theTmaximal ideals
of the valuation rings O ⊂ K. The structure sheaf is given by OXK (U ) = P ∈U OP where
U ⊆ XK is open and OP is the valuation ring corresponding to P . This determines the
reverse assignment K 7→ XK . One now checks that these assignments are inverse and
preserve categorical structure.
Now fix nonsingular curves X and Y over k and a morphism ϕ : X → Y defined over k.
Then an irreducible divisor y ∈ Div(Y ) corresponds to a maximal ideal mY (on some affine
patch) with uniformizer ty ∈ k(Y ).

Definition. The pullback of ϕ is a map ϕ∗ : Div(Y ) → Div(X) defined on irreducible


divisors by X
ϕ∗ y = ordX (ϕ∗ ty )x,
x∈X

where ty is a uniformizer at y, and extended linearly.

Example 17.2.3. Let X be the plane curve defined by y 2 − x and Y = P1 the projective
line, and let ϕ : X → Y be the x-coordinate projection.

398
17.2. Morphisms Between Curves Chapter 17. Curves

x2
X

x0

x1

Y
y0 y1

Then ϕ∗ y0 = 2x0 + ord∞ (ϕ∗ ty0 )∞ and ϕ∗ y1 = x1 + x2 + ord∞ (ϕ∗ ty1 )∞.

Definition. Let ϕ : X → Y be a morphism, x ∈ X and y = ϕ(x) ∈ Y . The number


eϕ (x) = ordx (ϕ∗ ty ) is called the ramification index of ϕ at x. If eϕ (x) = 1 and the residue
field extension κ(x)/κ(y) is separable, we say ϕ is unramified at x. Otherwise, we say ϕ
is ramified at x, and y is called a branch point of ϕ.

Proposition 17.2.4. Fix a morphism ϕ : X → Y , x ∈ X and y = ϕ(x) ∈ Y . Then

(1) eϕ (x) does not depend on the choice of uniformizer ty .


P
(2) For any Q ∈ Y , P ∈ϕ−1 (Q) eϕ (P ) = deg ϕ.

(3) All but finitely many Q ∈ Y have #ϕ−1 (Q) = degs ϕ.

(4) If ψ : Y → Z is a morphism then eψϕ (x) = eϕ (x)eψ (y).

Definition. Given a morphism ϕ : X → Y , the pushforward of ϕ is a map ϕ∗ : Div(X) →


Div(Y ) defined on irreducible divisors x ∈ X by

ϕ∗ x = [κ(x) : κ(ϕ(x))]ϕ(x)

and extended linearly.

Proposition 17.2.5. Let ϕ : X → Y be a morphism and D ∈ Div(Y ) and D0 ∈ Div(X)


divisors. Then

(1) deg(ϕ∗ D) = (deg ϕ)(deg D).

(2) ϕ∗ (f ) = (ϕ∗ f ) for any function f ∈ k(Y ).

(3) deg(ϕ∗ D0 ) = deg(D0 ).

399
17.2. Morphisms Between Curves Chapter 17. Curves

(4) ϕ∗ ϕ∗ D = (deg ϕ)D.

Corollary 17.2.6. For any function f ∈ k(X) on a curve X, deg(f ) = 0.

Proof. First, if f is constant then (f ) = 0 since X is complete. Otherwise, we may view f


as a function f : X → P1k . Then (f ) = f −1 (0) − f −1 (∞) for the points 0, ∞ ∈ P1k , since
X
f −1 (P ) = vQ (f ∗ tP )Q.
Q∈X

But on the other hand, this means deg(f ) = deg(f −1 (0) − f −1 (∞)) = [k(X) : k(f )] − [k(X) :
k(f )] = 0.
Let Div0 (X) be the subgroup of Div(X) consisting of divisors of degree zero. Then
Corollary 17.2.6 shows that PDiv(X) ⊆ Div0 (X). Set

Pic0 (X) := Div0 (X)/ PDiv(X).

Then the degree map determines an exact sequence


D
0 → k × → k(X)× −
→ Div0 (X) → Pic0 (X) → 0.

If X is defined over the algebraic closure k̄, write Pic(X/k̄) for Pic(X(k̄)). Consider Div(X/k̄)Gk .
Then we have an embedding

Pic0 (X/k) ,→ Pic0 (X/k̄)Gk .

Unfortunately, this map is not surjective in general.

400
17.3. Linear Equivalence Chapter 17. Curves

17.3 Linear Equivalence


Definition. The classes [D] = {D + (f ) : f ∈ k(X)× } in the Picard group of X determines
a linear equivalence: D ∼ D0 if there exists an f ∈ k(X)× such that D + (f ) = D0 .
Lemma 17.3.1. For two divisors D, D0 ∈ Div(X), D ∼ D0 if and only if deg(D) = deg(D0 ).
Therefore the degree map descends to a map on the Picard group,

deg : Pic(X) −→ Z.
P
Definition. A divisor D = nx x on X is called effective if nx ≥ 0 for all x ∈ X. In this
case we will write D ≥ 0. Also, if D1 , D2 ∈ Div(X) and D1 − D2 is an effective divisor, we
write D1 ≥ D2 . This defines an ordering on Div(X).
Definition. Let D be an effective divisor on X. Then the Riemann-Roch space associated
to D is the k-vector space

L(D) = {f ∈ k(X)× | D + (f ) ≥ 0} ∪ {0}.

We denote its dimension by `(D) = dimk L(D).


P
The condition that D + (f ) ≥ 0 can be restated as (f ) ≥ −D, or if D = nx x then
ordx f ≥ −nx for all x ∈ X.
Example 17.3.2. Let x ∈ X and n > 0. For the divisor D = nx, the space L(D) consists
of all f ∈ k(X)× with no poles except possibly at x of order at most n.
Definition. Fix a divisor D ∈ Div(X). The projective space

|D| := {D0 ∈ Div(X) : [D0 ] = [D] and D0 ≥ 0} ∼


= P(L(D))

is called the complete linear system of D on X. Any projective subspace of |D| is called
a linear system of D on X.
Note that D is linearly equivalent to an effective divisor if and only if L(D) 6= 0.
Theorem 17.3.3. For any D ∈ Div(X), L(D) is finite dimensional.
Lemma 17.3.4. If D1 , D2 ∈ Div(X) are linearly equivalent, say D1 − D2 = (g) for some
g ∈ k(X)× , then there is an isomorphism

L(D1 ) −→ L(D2 )
f 7−→ gf.

In particular, `(D) is a well-defined invariant of each class [D] ∈ Pic(X).


Remark. If X is defined over an extension k ⊆ K ⊆ k̄, write LK (D) and `K (D) for the
Riemann-Roch space of D on X(K) and its dimension. Then Lk̄ (D) has a basis consisting
of functions f ∈ k(X)× , so `k̄ (D) = `k (D). Thus we are justified in writing `(D) for any of
these.

401
17.3. Linear Equivalence Chapter 17. Curves

Proposition 17.3.5. Let D, D1 , D2 ∈ Div(X). Then

(1) `(D) ≤ deg(D) + 1 if D ≥ 0.

(2) If D1 ≤ D2 then L(D1 ) ⊆ L(D2 ).

Example 17.3.6. For X = P1 , any divisor D is linearly equivalent to d∞ for some d ∈ Z.


Then L(D) ∼ = L(d∞) = {f ∈ k[t] : deg f ≤ d} which has dimension exactly d + 1. Thus the
equality `(D) = deg(D) + 1 holds for any divisor on P1 .

Example 17.3.7. If X 6= P1 and D is an effective divisor, then `(D) ≤ deg(D). In partic-


ular, if deg(D) ≤ 0 then `(D) = 0.

Next, we explore how much less than deg(D) + 1 the dimension `(D) can be. This
culminates with the Riemann-Roch theorem in Section 17.6. Set γ(D) = deg(D) + 1 − `(D).

Theorem 17.3.8 (Riemann Inequality). For an nonsingular algebraic curve X, there is a


bound γX such that γ(D) ≤ γX and 1 + deg(D) − γX < `(D) for all divisors D ∈ Div(X).

The Riemann-Roch spaces are useful for constructing maps X → PN and in particular
embeddings into projective space. Given a rational map ϕ = (ϕ0 , . . . , ϕN ) : X 99K PN with
ϕi ∈ k(X), define the divisor of ϕ to be

Dϕ = gcd{(ϕ0 ), . . . , (ϕN )}.

Then for each ϕi , (ϕi ) − Dϕ ≥ 0 so ϕi ∈ L(−Dϕ ). Set D = Dϕ . Let M be the subspace of


L(−D) spanned by (ϕ0 ), . . . , (ϕN ). We may assume that these (ϕi ) are linearly independent,
lowering N if necessary. Then dim M = N + 1. Next, δ = {(g) − D | g ∈ M } is a linear
system of dimension N , i.e. a subspace of | − D|. Thus every rational map X 99K PN
determines a linear system of D, and it turns out the converse is also true.
Given δ ≤ |D| a linear subspace of |D| ⊆ PN of dimension N , define the base locus of δ
by n X o
0
B(δ) = P ∈ X : nP 6= 0 for all D = nP P ∈ δ .

Choose a basis f0 , . . . , fN of functions for L(D) corresponding to δ. Then ϕδ = (f0 , . . . , fN ) :


X 99K PN is a rational map that restricts to a morphism on X r B(δ). This is in fact unique
up to automorphism of PN – corresponding to a choice of basis.

Definition. A linear system δ ≤ |D| is called basepoint-free if B(δ) = ∅.

A basepoint-free linear system δ determines a regular map ϕδ : X → PN .

Definition. If the complete linear system |D| is basepoint-free and the morphism ϕ|D| : X →
PN is an embedding, we say |D| is very ample. If for some m > 0, the complete linear
system |mD| is very ample, then we say |D| is ample.
P
Theorem 17.3.9. Let X be a curve and D = nx x an effective divisor on X. Then

(1) D is basepoint-free if and only if for all x ∈ X such that nx 6= 0, `(D − x) < `(D).

402
17.3. Linear Equivalence Chapter 17. Curves

(2) |D| is very ample if and only if `(D − P − Q) < `(D − P ) < `(D) for all P, Q ∈ X.

For an algebraic curve X, the assignment

U 7→ OX (U )

defines a sheaf OX on X, called the structure sheaf of X. Note that for each P ∈ X,
OX,P = OP is the local ring at P , and for any open set U ⊆ X,
\
OX (U ) = OP .
P ∈U

For any sheaf F on X, let H i (X, F) be the ith cohomology of X with coefficients in F.
Theorem 16.2.4 implies the following:

Proposition 17.3.10. Let X be an algebraic curve and F a sheaf on X. Then H i (X, F) = 0


for all i ≥ 2.
P
We now define a sheaf on X associated to a divisor D = nP P ∈ Div(X). For each
open set U ⊆ X, set

L(D)(U ) = {f ∈ k(X)× : vP (f ) ≥ −nP for all P ∈ U }.

Then U 7→ L(D)(U ) defines a sheaf of k-vector spaces on X, denoted L(D). In fact, L(D)
is a subsheaf of the constant sheaf on X associated to k(X) (which by abuse of notation we
will also write as k(X)). Note that H 0 (X, L(D)) = L(D), the Riemann-Roch space for D.

Proposition 17.3.11. For any divisor D on X, H i (X, L(D)) = 0 for i ≥ 2 and is a finite
dimensional k-vector space for i = 0, 1.

Proof. First suppose D = 0. Then L(0) = OX is the sheaf of regular functions on X and
since X is complete,

H 0 (X, L(0)) = H 0 (X, OX ) = OX (X) = k.

Hence h0 (L(0)) = 1. We will prove dimk H 1 (X, L(0)) < ∞ later. For the rest, it’s enough
to prove that the statements hold simultaneously for D and D0 = D + P , for any divisor
D and any point P ∈ X. For any point Q ∈ X, we have L(D)Q ⊆ L(D + P )Q so we may
regard L(D) as a subsheaf of L(D + P ). Consider the exact sequence of sheaves

0 → L(D) → L(D + P ) → F → 0

where F is the quotient sheaf. Then FQ = 0 for all Q 6= P , and for Q = P we get
L(D + P )P
dimk FP = dimk = 1, so H 1 (X, F) = 0 and H 0 (X, F) = k. Thus the long exact
L(D)P
sequence in sheaf cohomology is

0 → L(D) → L(D0 ) → k → H 1 (X, L(D)) → H 1 (X, L(D0 )) → 0.

403
17.3. Linear Equivalence Chapter 17. Curves

This yields two short exact sequences:


0 → L(D) → L(D0 ) → A → 0
0 → B → H 1 (X, L(D)) → H 1 (X, L(D0 )) → 0.
Applying the additive function dimk gives
dimk L(D) + dimk A = dimk L(D0 )
dimk B + dimk H 1 (X, L(D0 )) = dimk H 1 (X, L(D))
so H i (X, L(D)) and H i (X, L(D0 )) are finite dimensional simultaneously, for i = 0, 1.
Definition. The arithmetic genus of X is the dimension
g = dimk H 1 (X, L(0)) = dimk H 1 (X, OX ).
The following is a preliminary version of the Riemann-Roch theorem, which we will prove
in full generality in Section 17.6.
Theorem 17.3.12 (Riemann-Roch, First Version). For an algebraic curve X and a divisor
D ∈ Div(X),
`(D) − h1 (L(D)) = 1 − g + deg(D).
Proof. First suppose D = 0. Then `(0) = dimk L(0) = 1 since (f ) ≥ 0 is equivalent to
f being regular on X, which is equivalent to f being constant since X is complete. Also,
deg(0) = 0 so we get 1 − g on both sides of the equation. Now let P ∈ X. To prove
the theorem in general, it will be enough to show the formula holds for any D and D + P
simultaneously. On the left, we have `(D)−h1 (L(D)) = h0 (L(D))−h1 (L(D)) = χ(X, L(D)),
the Euler characteristic of the sheaf L(D) on X. Set χ(D) := χ(X, L(D)). Then for any P ∈
X, we have deg(D+P ) = deg(D)+1, so in particular 1−g +deg(D+P ) = 1−g +deg(D)+1.
Thus it’s enough to show χ(D + P ) = χ(D) + 1.
As in the proof of Proposition 17.3.11, consider the exact sequence of sheaves
0 → L(D) → L(D + P ) → F → 0
where F is the quotient sheaf. Then as before, h1 (X, F) = 0 and h0 (X, F) = 1. Consider
the long exact sequence in sheaf cohomology:
0 → L(D) → L(D + P ) → H 0 (X, F) → H 1 (X, L(D)) → H 1 (X, L(D + P )) → 0.
Applying dimk gives
`(D) − `(D + P ) + h0 (F) − h1 (L(D)) + h1 (L(D + P )) = 0
=⇒ `(D) − `(D + P ) + 1 − h1 (L(D)) + h1 (L(D + P )) = 0,
which is equivalent to χ(D) + 1 = χ(D + P ). Hence the Riemann-Roch equality holds for
all D.
This gives us a reasonable tool for calculations, but to obtain the most general form of
Riemann-Roch, we will need to interpret h1 (L(D)). It turns out that there is a ‘canonical
divisor’ K on X for which H 1 (X, L(D)) ∼= L(K − D) for all divisors D.

404
17.4. Répartitions Chapter 17. Curves

17.4 Répartitions
To interpret h1 (L(D)) in Theorem 17.3.12, we want to find a ‘canonical divisor’ for which
h1 (L(D)) = `(K − D) for all divisors D on X. We first interpret H 1 (X, L(D)) using
répartitions and then relate the dual of this space to the space of meromorphic differentials
on our curve. Finally, this will allow us to define the canonical divisor.

Definition. A répartition on X is a family {rP }P ∈X with rP ∈ k(X) and such that rP 6∈


OP for finitely many P . The set of all répartitions on X is denoted R. Note that R is a
k-algebra.

Remark. R may also be identified as the ring of adèles of k, as in adèlic number theory.
P
For a divisor D = nP P ∈ Div(X), define the subspace

R(D) = {{rP } : vP (rP ) ≥ nP } ⊆ R.

As with the Riemann-Roch spaces, we have R(D0 ) ⊇ R(D) when D0 ≥ D. Also, note that
[
R= R(D).
D∈Div(X)

As with divisors, there is a natural embedding k(X) ,→ R, f 7→ {f }.

Proposition 17.4.1. For any D ∈ Div(X), H 1 (X, L(D)) ∼


= R/(R(D) + k(X)).
Proof. Regard L(D) as a subsheaf of the constant sheaf k(X) on X. This determines a short
exact sequence of sheaves
0 → L(D) → k(X) → Q → 0
where Q is the sheafification of the presheaf k(X)/L(D). The long exact sequence in coho-
mology is

H 0 (X, k(X)) → H 0 (X, Q) → H 1 (X, L(D)) → H 1 (X, k(X)) → · · ·

Then we have H 0 (X, k(X)) ∼ = k(X) since k(X) is a constant sheaf, and H 1 (X, k(X)) = 0,
either by computing using Čech cohomology or using the fact that k(X) is a flasque sheaf.
In any case, the long exact sequence becomes:

k(X) → H 0 (X, Q) → H 1 (X, L(D)) → 0.

Moreover, Q has finite support (it vanishes off of a finite set, namely the support of D), so
it is a skyscraper sheaf and thus

H 0 (X, Q) ∼ (k(X)/L(D))P ∼
M
= = R/R(D)
P ∈X

Now since k(X) is a field, the natural k-algebra map k(X) → R/R(D), f 7→ {f } is injective,
so by exactness we get H 1 (X, L(D)) ∼= R/(R(D) + k(X)) as desired.

405
17.4. Répartitions Chapter 17. Curves

Set I(D) = R/(R(D) + k(X)) and let J(D) = I(D)∗ denote the algebraic dual. By
duality, we have J(D0 ) ⊆ J(D) if D0 ≥ D. Set
[
J= J(D).
D∈Div(X)

For f ∈ k(X) and α ∈ J, define a map

f α : R −→ k
r 7−→ α(f r).

Since α vanishes on k(X), f α vanishes on k(X) as well. Moreover, if α ∈ J(D) and f ∈


L(D0 ), then f α vanishes on R(D − D0 ) and hence f α ∈ J(D − D0 ) ⊆ J. This shows that J
has the structure of a k(X)-vector space.

Proposition 17.4.2. dimk(X) J ≤ 1.

Proof. Suppose α, α0 ∈ J are k(X)-linearly independent. We may find a divisor D ∈ Div(X)


such that α, α0 ∈ J(D). Set d = deg(D). For each n ≥ 0, pick some divisor ∆n of degree
n. If f ∈ L(∆n ) then by the above, f α, f α0 ∈ J(D − ∆n ). Moreover, if f α + gα0 = 0 for
g ∈ L(∆n ), then we must have f = g = 0. This defines an injection

L(∆n ) × L(∆n ) −→ J(D − ∆n )


(f, g) 7−→ f α + gα0 .

Thus dim J(D − ∆n ) ≥ 2`(∆n ), but by the first form of Riemann-Roch (Theorem 17.3.12),

dim J(D − ∆n ) = dim I(D − ∆n )


= − deg(D − ∆n ) + g − 1 + `(D − ∆n )
= −(d − n) + g − 1 + `(D − ∆n ).

In particular, when n > d, we have deg(D − ∆n ) < 0 so `(D − ∆n ) = 0. Hence the


Riemann-Roch formula becomes

dim J(D − ∆n ) = −(d − n) + g − 1 = n + (g − 1 − d).

On the other hand, by Riemann-Roch,

2`(∆n ) ≥ 2(1 − g + deg(∆n )) = 2(1 − g + n) = 2n + (2 − 2g),

so dim J(D − ∆n ) ≥ 2`(∆n ) implies that n + (g − 1 − d) ≥ 2n + (2 − 2g). Taking n large


enough gives a contradiction, so α, α0 must be linearly independent.
In fact, it will follow from our work in the next section that dimk(X) J = 1.

406
17.5. Differentials Chapter 17. Curves

17.5 Differentials
Definition. For a curve X, the space of meromorphic differentials on X is the k(X)-
vector space ΩX consisting of formal differentials df for each f ∈ k(X)× satisfying
ˆ d(f + g) = df + dg,

ˆ dα = 0 if α ∈ k,

ˆ d(f g) = f dg + g df .
That is, ΩX is the k-algebra

Ω(X) = hdf | f ∈ k(X)i/(df g − gdf − f dg).

If ϕ : X → Y is a morphism of curves, we get a map of fields ϕ∗ : k(Y ) → k(Y ). Define


the induced map on meromorphic differentials by

ϕ∗ : ΩY −→ ΩX
X  X

ϕ fi dti 7−→ ϕ∗ fi d(ϕ∗ ti ).

Lemma 17.5.1. For any algebraic curve X, dimk(X) ΩX = dim X = 1.


Proposition 17.5.2. For any f ∈ k(X), the following are equivalent:
(i) df 6= 0.

(ii) df is a basis for ΩX .

(iii) k(X)/k(f ) is finite and separable.

(iv) f 6∈ k if char k = 0, or f 6∈ k(X)p if char k = p > 0.


Lemma 17.5.3. A nonconstant morphism ϕ : X → Y is separable if and only if the induced
map ϕ∗ : ΩY → ΩX is nonzero.
For a point P ∈ X, choose a uniformizer t = tP in OX,P . Then ΩX is generated by dt.
Hence for any ω ∈ ΩX , there exists g ∈ k(X) such that ω = g dt.
Definition. Define the order of ω at P ∈ X to be ordP (ω) = ordP (g), where ω = g dt. The
principal divisor associated to ω is then defined to be
X
(ω) = ordP (ω)P.
P ∈X

Proposition 17.5.4. Let X be a curve, P ∈ X, f ∈ k(X) and ω ∈ ΩX . Then


(1) If f is regular at P then df = f dt for t = tP a local uniformizer.

(2) For any s ∈ k(X) such that s(P ) = 0, ordP (f ds) = ordP (f ) + ordP (s) − 1 if p -
ordP (s), and ordP (f ds) ≥ ordP (f ) + ordP (s) if p | ordP (s).

407
17.5. Differentials Chapter 17. Curves

(3) ordP (ω) = 0 for all but finitely many P ∈ X.

Definition. The canonical class on a curve X is the class KX = [(ω)] in Pic(X) for any
nonzero differential ω ∈ ΩX .

Lemma 17.5.5. The canonical class is well-defined, i.e. does not depend on the choice of
ω ∈ ΩX .

Proof. For nonzero ω1 , ω2 ∈ ΩX , write ω1 = f ω2 for some f ∈ k(X)× . Then (ω1 ) = (f ω2 ) =


(f ) + (ω2 ). Thus [(ω1 )] = [(ω2 )].

Definition. We say ω ∈ ΩX is a holomorphic (or regular) differential on X if ordP (ω) ≥


0 for all P ∈ X. We denote the space of holomorphic differentials on X by Ω[X].

Note that Ω[X] is a k-vector space but need not be a k(X)-vector space.

Definition. The geometric genus of X is defined as g(X) := `(KX ), the dimension of the
Riemann-Roch space L(KX ) of the canonical class.

Lemma 17.5.6. There is an isomorphism L(KX ) → Ω[X].

Proof. The map is f 7→ f ω for any fixed ω ∈ Ω[X] defining the canonical class.

Corollary 17.5.7. For any curve X, g(X) = dimk Ω[X].

Remark. For any divisor D ∈ Div(X), `k (D) = `k̄ (D) implies g(X(k)) = g(X(k̄)), so the
geometric genus is unchanged when passing to the algebraic closure k̄. Moreover, g(X) is a
birational invariant of X.

Example 17.5.8. Let X = P1 and let t be a coordinate function on some affine patch U of
P1 . We claim that (dt) = −2∞. Indeed, for any α ∈ U ∼ = A1 , t − α is a local uniformizer at
α. Thus ordα(dt) = ordα (d(t − α)) = 0. At infinity, 1t is a local uniformizer so we can write
dt = −t2 d 1t . Hence

ord∞ (dt) = ord∞ −t2 d 1t = ord∞ − t−2 1


+ ord∞ d 1t = −2 + 0 = −2.
  

So (dt) = −2∞ as claimed. Now for any ω ∈ ΩP1 , deg(ω) = −2 so we see that `(KP1 ) =
`(−2∞) = 0. Hence the genus of the projective line is g(P1 ) = 0.

Corollary 17.5.9. There are no holomorphic differentials on P1 .

Proof. By Corollary 17.5.7, g(P1 ) = dimk Ω[P1 ] but by the calculations above, the genus of
P1 is zero.

Definition. For a divisor D ∈ Div(X), define the meromorphic differentials at D by

Ω(D) = {ω ∈ ΩX : (ω) ≥ D}.

408
17.5. Differentials Chapter 17. Curves

There is another useful local invariant of a meromorphic differential, which we define now.
For P ∈ X, let ObP = k[[T ]] be the topological completion of the local ring OP , which comes
equipped with a map OP → O bP , t 7→ T for a local uniformizer t; this map is an embedding.
Then the fraction field of this ring is FbP := Frac(O bP ) = k((T )) so via the embedding just
defined, we can view functions f ∈ OP as formal Laurent series in T ,

X
f= ai T i .
i=−n

In analogy with complex analysis, we define:

Definition.
P∞ The residue of ω ∈ ΩX at P ∈ X is resP (ω) := a−1 if ω = f dt and f =
i
i=−n ai T .

Fix P ∈ X and let (O, m, v) be the DVR of the local field FbP and let Dk (FbP ) be the ring
of k-differentials on FbP . Set
\∞
Q= mn dO ⊆ Dk (FbP )
n=1

and define D0 = Dk (FbP )/Q.


P∞ i
Lemma P∞ 17.5.10. Let f = i=−m ai t ∈ k((T )), where t is a uniformizer at P , and set
ft = i=−m iai t , the formal derivative of f with respect to t. Then df = ft dt in D0 and
i−1

dt is a basis for D0 .

Proof. It is enough to show df − ft dt ∈ mN dO for all N . Write f = f0 + tN +1 f1 for some


f0 ∈ k((T )) and f1 ∈ O. Then ft = (f0 )t + (N + 1)tN g for some g ∈ O, so

df − ft dt = (N + 1)tN f1 dt + tN +1 df1 − tN +1 g dt ∈ mN dO.

Hence df − ft dt = 0 in D0 . Finally, note that t 7→ ft is a nontrivial derivation on FbP which


vanishes on Q, so D0 6= 0. This proves that D0 has dimension 1 and is generated by dt.

Proposition 17.5.11. Let P ∈ X and let t be a uniformizer at P . Then

(1) resP is k-linear.

(2) resP (ω) = 0 for any ω ∈ Odt.

(3) resP (dg) = 0 for any g ∈ FbP .


 
(4) resP dgg
= vP (g) for any g ∈ FbP .

Proposition 17.5.12. The residue resP (ω) is well-defined, i.e. does not depend on the
choice of uniformizer t.

409
17.5. Differentials Chapter 17. Curves

Proof. By Lemma 17.5.10, write ω = ω0 + N −n


P
n=−N an t dt for some ω0 ∈ ΩX with vP (ω0 ) ≥
0. Let the residue of ω with respect to t be rest (ω) P = a−1 . If u is another uniformizer
N −n
at P , define resu (ω) the same way. Then resu (ω) = n=−N an resu (t dt) but by (4) of
dt −n

Proposition 17.5.11, resu t = vP (t) = 1 so it suffices to show resu (t dt) = 0 for n ≥ 2.
Let g = − n−1 t , so that for n ≥ 2, t−n dt = dg. Then by (3) of Proposition 17.5.11,
1 n−1

resu (dg) = 0, so resu (t−n dt) = 0 as required.


P
Theorem 17.5.13 (Residue Formula). For any ω ∈ ΩX , P ∈X resP (ω) = 0.

Proof. We first establish this for X = P1 . In this case, k(X) = k(t) so for any differential
ω ∈ ΩP1 , ω = f dt for some f ∈ k(t). Since resP is k-linear by Proposition 17.5.11, we may
assume that f = tn or f = (t − a)−n for some n > 0 and a ∈ k. If f = tn , the only pole of
ω = tn dt is at P = ∞. At ∞, u = 1t is a uniformizer, so we get

ω = f dt = −u−(n+2) du =⇒ res∞ (f dt) = 0


P
by Proposition 17.5.11. Hence P ∈P1 resP (ω) = res∞ (ω) = 0. The other case is similar.
Now let X be any algebraic curve and take a rational function ϕ ∈ k(X). This may
be viewed as a morphism of curves ϕ : X → P1 , corresponding to the field extension
k(P1 ) = k(t) ,→ k(X). Set E = k(t) and F = k(X). Then the trace TrF/E gives a map on
meromorphic differentials

Tr : ΩX −→ ΩP1
ω 7−→ TrF/E (f ) dϕ

if ω = f dϕ for f ∈ k(X). Take P ∈ P1 and Q ∈ X such that ϕ(Q) = P , and let E bP


and FbQ be the complete local fields at P and Q, respectively. Then the valuations at these
points are related by vQ = evP where e is the ramification index of F/E. Further, there is
an isomorphism
bP ∼
Y
F ⊗E E = FbQ
Q→P

where the product isPover all Q ∈ X for which ϕ(Q) = P . Under this isomorphism, for any
f ∈ F , TrF/E (f ) = Q→P TrQ (f ), which implies
X
resP (Tr(f ) dϕ) = resP (TrQ (f )) dϕ,
Q→P

but resQ (f dϕ) = resP (Tr(f ) dϕ), so we get


X X X X
resQ (ω) = resQ (ω) = resP (Tr(ω))
Q∈X P ∈P1 Q→P P ∈P1

which is 0 by the P1 case proven above. Therefore the residue formula holds for X.

410
17.6. Serre Duality and the Riemann-Roch Theorem Chapter 17. Curves

17.6 Serre Duality and the Riemann-Roch Theorem


Residues allow us to define a pairing

Ω(X) × R −→ k
X
(ω, r) 7−→ hω, ri := resP (rP ω)
P ∈X

if r = {rP }. Note that if r ∈ k(X) ⊆ R, then by the residue theorem (17.5.13), hω, ri = 0 for
all ω ∈ Ω(X). Further, if D is a divisor, r = {rP } ∈ R(D) and ω ∈ Ω(D), then rP ω ∈ Ω(X)P
for all P ∈ X, since vP (rP ) ≥ −vP (D) and vP (ω) ≥ vP (D). Hence the pairing passes to the
quotient I(D) = R/(R(D) + k(X)):

h·, ·i : Ω(D) × I(D) −→ k.

Consider the map θ : Ω(D) → J(D) defined by ω 7→ hω, ·i.


Lemma 17.6.1. Let D be a divisor on X and ω ∈ Ω(X). If θ(ω) ∈ J(D), then ω ∈ Ω(D).
Proof. If ω 6∈ Ω(D), there is a point P ∈ X for which vP (ω) < vP (D). Set n = vP (ω) + 1
and construct a répartition r = {rQ } by
(
t−n , if Q = P
rQ =
0, if Q 6= P.

Then vP (rP ω) = −1, so resP (rP ω) 6= 0. Thus hω, ri 6= 0 since the pairing is 0 everywhere
else. However, r ∈ R(D) implies hω, ri = 0, a contradiction.
Theorem 17.6.2 (Serre Duality). For any divisor D ∈ Div(X), θ : Ω(D) → J(D) is an
isomorphism.
Proof. Suppose θ(ω) = 0. Then θ(ω) ∈ J(D) for every D, so by Lemma 17.6.1, ω ∈ Ω(D)
for every D. However, \
Ω(D) = 0
D∈Div(X)

since X is complete, so ω = 0. This proves θ is injective. On the other hand, dimk(X) Ω(X) =
1 and dimk(X) J ≤ 1 by Proposition 17.4.2, so the map Ω(X) → J must be surjective.
Suppose α ∈ J(D). Then there is some ω ∈ Ω(X) with θ(ω) = α, but by Lemma 17.6.1,
this means ω ∈ Ω(D). Therefore θ is surjective and hence an isomorphism.
Let D be a divisor on X. By Theorem 17.3.12, we already have

`(D) − i(D) = deg(D) − g + 1

where `(D) = dimk L(D) = dimk H 0 (X, L(D)) and i(D) = dimk Ω(D) = dimk H 1 (X, L(D)).
Let ω, ω 0 ∈ Ω(X). Since Ω(X) has dimension 1, we may write ω = f ω 0 for f ∈ k(X). Then
(ω) = (ω 0 ) + (f ), so in particular every meromorphic differential form on X belongs to the
same nontrivial linear equivalence class in Pic(X), call it K.

411
17.6. Serre Duality and the Riemann-Roch Theorem Chapter 17. Curves

Definition. The class K ∈ Pic(X) such that [(ω)] = K for any ω ∈ Ω(X) is called the
canonical divisor class of X.
Fix a differential form ω0 ∈ Ω(X), so that K = [(ω0 )]. For any D ∈ Div(X) and
ω ∈ Ω(X), (ω) ≥ D is equivalent to (f ) ≥ D − (ω0 ), ω = f ω0 for f ∈ k(X). This proves:
Lemma 17.6.3. ω ∈ Ω(D) if and only if f ∈ L(K − D), where ω = f ω0 .
Theorem 17.6.4 (Riemann-Roch). For an algebraic curve X with canonical divisor K and
any divisor D ∈ Div(X),

`(D) − `(K − D) = 1 − g + deg(D).

Proof. Lemma 17.6.3 shows that i(D) = `(K − D) so this is just a reinterpretation of
Theorem 17.3.12.
Corollary 17.6.5. If KX is the canonical divisor on X, then deg(KX ) = 2g − 2.
Proof. Set D = K = KX . Then the Riemann-Roch theorem says that

`(K) − `(0) = deg(K) + 1 − g

but `(K) = g by definition and `(0) = 1. Solving for deg(K) we get deg(K) = 2g − 2.
Corollary 17.6.6. Suppose deg(D) > 2g − 2 for some divisor D ∈ Div(X). Then `(D) =
deg(D) + 1 − g.
The genus is a discrete invariant of nonsingular curves. There are two natural questions
that arise:
(1) What are the curves with genus g for a particular g ∈ N0 ?

(2) How do we describe the structure of the collection of all genus g curves?
We will see that one can put the structure of a variety on the collection of genus g curves.
Lemma 17.6.7. Let X be an algebraic curve. Then X ∼ = P1 if and only if there is some
divisor D ∈ Div(X) such that deg(D) = 1 and `(D) ≥ 2.
Proof. ( =⇒ ) If X ∼
= P1 then g(X) = g(P1 ) = 0 by Example 17.5.8. Take a point P ∈ X
and set D = P ∈ Div(X); of course deg(D) = 1. Then by the Riemann-Roch theorem,

`(D) = 1 − g + deg(D) + `(K − D) = 1 − 0 + 1 + `(K − D) = 2 + `(K − D) ≥ 2.

( ⇒= ) Since `(D) ≥ 2, there exists a nonconstant function g ∈ L(D). Then D ∼ D +


(g) ≥ 0 so we may assume D is effective. The only way for deg(D) = 1 is for D = P for some
point P ∈ X(k). Now g determines a map g : X → P1 , under which g ∗ ∞ = ord∞ (g) = P ,
so we must have deg(g) = 1. Hence g is an isomorphism of curves.
Proposition 17.6.8. For an algebraic curve X with genus g = g(X), the following are
equivalent:

412
17.6. Serre Duality and the Riemann-Roch Theorem Chapter 17. Curves

(1) X ∼
= P1 .
(2) g = 0 and there exists a divisor D ∈ Div(X) with deg(D) = 1.

(3) g = 0 and X(k) 6= ∅.

Proof. (1) =⇒ (2) follows immediately from Lemma 17.6.7.


(2) =⇒ (1) Since the genus is 0, deg(D) > 2g − 2 = −2 is certainly true. By Corol-
lary 17.6.6, `(D) = deg(D) + 1 − g = 1 + 1 − 0 = 2, so Lemma 17.6.7 once again applies.
(2) =⇒ (3) follows from the proof of Lemma 17.6.7.
(3) =⇒ (2) Any rational point P ∈ X(k) is a divisor on X of degree 1.
This shows that the main interest for curves of genus 0 is in finding rational points
P ∈ X(k). Moreover, when g(X) = 0, the complete linear system |KX | is very ample by
Theorem 17.3.9 and the Riemann-Roch theorem, and the embedding ϕ|KX | : X ,→ P2 realizes
X as a plane conic.

Remark. If ϕ : P1 → X is a morphism, Corollary 17.7.3 says that g(X) = 0. Further, when


k is algebraically closed or we consider the k̄-points X(k̄), one has X ∼
= P1 .

413
17.7. The Riemann-Hurwitz Formula Chapter 17. Curves

17.7 The Riemann-Hurwitz Formula


Let ϕ : X → Y be a nonconstant morphism of curves and fix P ∈ X. Then eϕ (P ) =
ordP (ϕ∗ tϕ(P ) ) where tϕ(P ) is a local uniformizer. We would like to see what happens to the
canonical class KX under a morphism. Take t to be a uniformizer at Q = ϕ(P ) and set
eϕ (P ) = e. Then ϕ∗ (dt) = d(ϕ∗ t). Moreover, if s is a uniformizer on X at P , then ϕ∗ t = use
for some unit u ∈ OP× . Now d(ϕ∗ t) = d(use ) = se du + uese−1 ds. Write du = g ds for a
regular function g ∈ OP ; this is possible by (1) of Proposition 17.5.4. Then

d(ϕ∗ t) = se g ds + euse−1 ds
=⇒ ordP (d(ϕ∗ t)) = ordP (se g + euse−1 )
= min{ordP (se g), ordP (euse−1 )}.

If char k - e, then this minimum is e − 1; otherwise, when char k | e the minimum is at least
e.
Definition. If ϕ is ramified and char k - eϕ (P ) for all P ∈ X, we say ϕ is tamely ramified.
Otherwise ϕ is wildly ramified.
Remark. If ϕ is tamely ramified, then ordP (d(ϕ∗ t)) = eϕ(P ) − 1 for each P . If ϕ is wildly
ramified at P , then ordP (d(ϕ∗ t)) ≥ eϕ (P ).
Definition. For a morphism ϕ : X → Y , define the ramification divisor
X
Rϕ = ordP (d(ϕ∗ t))P.
P ∈X

Now for ω ∈ ΩY , the canonical classes on X and Y can be defined by KY = [(ω)] and
KX = [(ϕ∗ ω)]. On the other hand, the pullback defines a divisor ϕ∗ KY ∈ Div(X). We want
to determine the relation between these three divisors.
Lemma 17.7.1. If ϕ : X → Y is a morphism of curves, then KX = ϕ∗ KY + [Rϕ ], where
Rϕ is the ramification divisor of ϕ.
Proof. If ω = f dt ∈ ΩY , then

ordP (ϕ∗ ω) = ordP (ϕ∗ f d(ϕ∗ t)) = ordP (ϕ∗ f ) + ordP (d(ϕ∗ t)),

so we see that ordP (ϕ∗ ω) gives the coefficient in KX , ordP (ϕ∗ f ) gives the coefficient in
ϕ∗ KY and ordP (d(ϕ∗ t)) gives the coefficient in Rϕ . Summing over P ∈ X gives the desired
equality.
P
Taking ϕ to be tamely ramified, Rϕ = P ∈X (eϕ (P ) − 1)P so the degree function applied
to the equation in Lemma 17.7.1 gives
X
deg(KX ) = deg(ϕ∗ KY ) + (eϕ (P ) − 1).
P ∈X

We will show in Section 17.6 that deg(KX ) = 2g(X) − 2. This proves:

414
17.7. The Riemann-Hurwitz Formula Chapter 17. Curves

Theorem 17.7.2 (Riemann-Hurwitz Formula). For any morphism ϕ : X → Y ,


X
2g(X) − 2 = (deg ϕ)(2g(Y ) − 2) + (eϕ (P ) − 1).
P ∈X

Corollary 17.7.3. For any morphism ϕ : X → Y , g(X) ≥ g(Y ).

415
17.8. The Canonical Map Chapter 17. Curves

17.8 The Canonical Map


We saw at the end of Section 17.6 that the theory of genus 0 curves for the most part reduces
to studying whether X has rational points and describing the embedding ϕ|KX | : K ,→ P2 .
What about higher genus curves?

Proposition 17.8.1. Let X be a nonsingular algebraic curve over k of genus g ≥ 1. If KX


is the canonical divisor of X then the complete linear system |KX | is basepoint-free.

Proof. This follows from the Riemann-Roch theorem (17.6.4) and Theorem 17.3.9, taking
D = KX .
Thus |KX | determines a regular map into projective space.

Definition. The canonical map of a genus g ≥ 1 curve X is the map ϕ|KX | : X → Pg−1 .

Definition. A hyperelliptic curve is a smooth curve X together with a separable, degree


2 map X → P1 .

Example 17.8.2. When char k 6= 2, a hyperelliptic curve is of the form X = Z(y 2 − f (x))
for a polynomial f ∈ k[x]. More generally, the minimal degree of a nonconstant morphism
X → P1 is called the gonality of X. Thus, a hyperelliptic curve is a curve of gonality 2.

Proposition 17.8.3. If X is not hyperelliptic and g ≥ 2, the canonical map ϕ|KX | : X →


Pg−1 is an embedding.

Proposition 17.8.4. If X is a nonsingular algebraic curve of genus g and D ∈ Div(X),


then

(1) If deg(D) ≥ 2g then |D| is basepoint-free.

(2) If deg(D) ≥ 2g + 1 then |D| is very ample.

Corollary 17.8.5. If g ≥ 2 then ϕ|3KX | : X → P5g−6 is an embedding.

Definition. The map ϕ|3KX | is called the tricanonical map of a curve X.

Theorem 17.8.6 (Faltings). If X is a curve of genus g ≥ 2 then #X(Q) is finite.

We have for the most part dealt completely with the cases of curves of genus g = 0 and
g ≥ 2, so the most interesting work remains to be done for curves of genus g = 1.

416
17.9. Bézout’s Theorem Chapter 17. Curves

17.9 Bézout’s Theorem


For this section let k be algebraically closed, fix X ⊆ PN a projective curve and Y ⊆ PN
a hypersurface defined by Y = Z(F ) for some F ∈ k[X0 , . . . , XN ]. Further suppose that
X 6⊂ Y , i.e. that F 6∈ J(X). Then by counting codimensions, X ∩Y must be some dimension
0 variety in PN , i.e. X and Y intersect in some discrete set of points. We want to count
these points, including some notion of multiplicity, in a rigorous way.

Definition. The intersection multiplicity of X and Y = Z(F ) at a point P ∈ X ∩ Y ,


denoted (X · F )P , is defined as follows. Let G ∈ k[X0 , . . . , XN ] be any form of the same
degree as F such that G(P ) 6= 0. Then F/G ∈ k(X) so the intersection multiplicity at P is
defined: (X · F )P := ordP (F/G). Further, the intersection divisor of F on X is
X
divX (F ) = (X · F )P P,
P ∈X∩Y
P
and its order (X · F ) := P ∈X∩Y (X · F )P is called the intersection number of X and Y .

Q
P

If L is the linear form representing the line in the figure, then (X · L)P = 1, (X · L)Q = 2 and the
intersection number is (X · L) = 1 + 2 = 3.

Proposition 17.9.1. If F1 ∈ k[X0 , . . . , XN ] r J(X) is another form with deg F1 = deg F ,


then (X · F ) = (X · F1 ).

Proof. Set f = F/F1 ∈ k(X). Then divX (F ) ∼ divX (F1 ), so deg(divX (F )) = deg(divX (F1 )),
and thus the intersection number is well-defined.

Corollary 17.9.2. If deg F = m and L is any linear form such that L 6∈ J(X), then
(X · F ) = m(X · L).

Proof. Since intersection multiplicity at a point is multiplicative, this formula is clear.

Lemma 17.9.3. For any form F 6∈ J(X) and any point P ∈ X ∩ Z(F ), (X · F )P = 1 if
and only if F (P ) = 0 and TP X 6⊂ TP Z(F ).

Stated another way, Lemma 17.9.3 says that the intersection multiplicity at P is 1 if and
only if X and Z(F ) meet transversely.

Lemma 17.9.4. For any smooth curve X, there exists a linear form L such that (X ·L)P ≤ 1
for all P ∈ X ∩ Z(L).

417
17.9. Bézout’s Theorem Chapter 17. Curves

Definition. The degree of a projective curve X ⊆ PN is defined to be

degPN X := max{#(X ∩ H) : H is a hyperplane and X 6⊂ H}.

Corollary 17.9.5. Let X be a projective curve in PN . Then degPN X = (X · L) for any


linear form L.

Theorem 17.9.6 (Bézout). Let X ⊂ PN be a projective curve and F ∈ k[X0 , . . . , XN ] a


form such that F 6∈ J(X). Then (X · F ) = (degPN X)(deg F ).

Example 17.9.7. If X ⊂ P2 is a planar curve given by a form G = 0, then degP2 X = deg G


so we can count intersection multiplicities in the plane by:

(X · F ) = (deg G)(deg F )

for any F ∈ k[X0 , X1 , X2 ] r J(X).

418
17.10. Rational Points of Conics Chapter 17. Curves

17.10 Rational Points of Conics


Given a plane conic C over a field of characteristic char k 6= 2, say

C : ax2 + 2bxy + 2cx + dy 2 + 2ey + f = 0

in A2k , we can homogenize to get a curve in P2k :

C : F (X, Y, Z) = aX 2 + 2bXY + 2cXZ + dY 2 + 2cY Z + f Z 2 = 0.

Then F is a quadratic form on the vector space V = k 3 .

Definition. For a k-vector space V , a function q : V → k is a quadratic form if

(a) q(λv) = λ2 v for all λ ∈ k and v ∈ V .

(b) The pairing bq (v, w) = 12 (q(v + w) − q(v) − q(w)) is symmetric and k-bilinear.

A quadratic form q is said to be nondegenerate if bq induces an isomorphism V ∼


= V ∗.
Otherwise q is degenerate.

If F (X, Y, Z) is a quadratic form on V = k 3 , then there is a matrix


 
a b c
MF = d e f 
g h i

such that F (X, Y, Z) = (X Y Z)MF (X Y Z)t . The determinant deg MF is called the
discriminant of F .

Lemma 17.10.1. A quadratic form F (X, Y, Z) is nondegenerate if and only if deg MF 6= 0.

Since MF is symmetric when F is quadratic, we may transform it by some invertible


matrix T ∈ GL3 (k) to a diagonal form DF = T t MF T . In these coordinates of k 3 , we have
3
X
F = ai Xi2 .
i=1

Further, if k = Q, we may assume the ai ∈ Z are squarefree and relatively prime.

Definition. A quadratic form F represented by a diagonal matrix M with squarefree, co-


prime integer entries is called a primitive quadratic form.

The crucial Hasse-Minkowski theorem says that a plane conic having a Q-rational point
is equivalent to the conic having a rational point over every completion of Q.

Theorem 17.10.2 (Hasse-Minkowski). Let F ∈ Q[X0 , . . . , Xn ] be a primitive quadratic form


and let X = Z(F ) ⊆ PnQ . Then X(Q) 6= ∅ if and only if X(Qv ) 6= ∅ for all places v of Q.

419
17.10. Rational Points of Conics Chapter 17. Curves

This theorem is the classic example of Hasse’s “local-to-global principle”: points over the
local fields Qv determine points over Q. Note that the Hasse-Minkowski theorem does not
hold for general varieties X, nor for general fields k.

Example 17.10.3. For a conic X, X(R) 6= ∅ if and only if there is a change of sign among
the coefficients ai in the form F defining X. This condition is easily checked as long as one
can diagonalize MF .

Thus to find rational points of a conic, we need only ask if there is an algorithm for
checking whether X has points over each p-adic field Qp .

Example 17.10.4. Let X = Pn . Then Pn (Q) = Pn (Z) and for any prime p, Pn (Qp ) =
Pn (Zp ), so it’s enough to look for integer solutions. If P = [α0 , . . . , αn ] ∈ Pn (Qp ), then
we can clear denominators so that P = [β0 , . . . , βn ] for βi ∈ Zp and some βj ∈ Z× p . The
n
reduction mod p of P is then given by Pe = [β̄0 , . . . , β̄n ] ∈ P (Fp ).

It turns out that quadratic forms always have points over finite fields. To prove this, we
will need the following counting lemma.

Lemma 17.10.5. For a sum s = α∈Fqn α1k1 · · · αnkn , where α = (α1 , . . . , αn ) and ki ∈ Z≥0 ,
P

if at least one ki is not a positive integer multiple of q − 1, then s = 0.

Proof. Write  
X n
Y X
s= α1k1 · · · αnkn =  ak i  .
α∈Fqn i=1 a∈Fq

If any ki = 0 then a∈Fq aki = a∈Fq 1 = q ≡ 0 so we may assume all ki 6= 0. Let φ be a


P P

generator of the cyclic group Fq× and write ψ = φki . If ki is not a positive multiple of q − 1,
then ψ 6= 1. Now we have
q−2
X X X
ki ki
a = a = (φm )ki
a∈Fq a∈Fq× m=0

q−2
X 1 − ψ q−1 1−1
= ψm = ≡ = 0.
m=0
1−ψ 1−ψ

Therefore s = 0.

Theorem 17.10.6 (Chevalley-Warning). Let Fq be a finite P field of characteristic p and let


f0 , . . . , fr ∈ Fq [X1 , . . . , Xn ] be polynomials satisfying n > rj=1 deg fj . Set X = Z(f0 , . . . , fr ) ⊆
AnFq . Then

(a) #X(Fq ) ≡ 0 (mod p).

(b) If (0, . . . , 0) ∈ X(Fq ) is a point on the curve then #X(Fq ) ≥ p.

420
17.10. Rational Points of Conics Chapter 17. Curves

Proof. Define the indicator function for X(Fq ):


r
Y
P (X1 , . . . , Xn ) = (1 − fj (X1 , . . . , Xn )q−1 ).
j=1

Notice that P (α) = 1 if α ∈ X(Fq ) and 0 otherwise. Then


X
#X(Fq ) = P (α) mod p.
α∈Fq

Now we have r n
X X
deg P = deg fi (q − 1) < n(q − 1)
i=1 i=1

by hypothesis. So for each monomial term X1k1 · · · Xnkn in P (X1 , . . . , Pn ), k1 + . . . + kn <


n(q − 1), so at least one ki must be less than q − 1. Hence by Lemma 17.10.5,
X
#X(Fq ) = P (α) = 0 mod p.
α∈Fqn

This proves (a), and (b) follows trivially.

Corollary 17.10.7. Every quadratic form in at least three variables has a point over each
finite field.

The theory of Hasse-Minkowski extends more generally to number fields K/Q, with
similar local-global principles at work.
We next determine when solutions to quadratic equations F = 0 over finite fields lift to
solutions in Zp , similar to Hensel’s Lemma. To do so, we introduce the notion of an integral
model for a variety over Q.

Definition. For a projective variety X ⊆ PN Q , an integral model for X is a choice of


homogenous forms F1 , . . . , Fm ∈ Z[X0 , . . . , Xn ] such that X = Z(F1 , . . . , Fm ). Denote these
forms {F1 , . . . , Fm } by X .

Note that we may assume the set of all coefficients of an integral model X = {F1 , . . . , Fm }
is coprime.

Definition. Let X = {F1 , . . . , Fm } be an integral model of X over Q. For a prime p, the


reduction of X mod p is the variety

XFp = Z(F 1 , . . . , F m ) ⊆ PN
Fp ,

where F i = Fi mod p. We say XFp is geometrically reduced if the ideal (F 1 , . . . , F m ) is


radical in Fp [X0 , . . . , XN ].

Notice that XFp depends on the integral model X chosen for X.

421
17.10. Rational Points of Conics Chapter 17. Curves

Definition. We say an integral model X has good reduction mod p if XFp is geometrically
reduced and nonsingular, and bad reduction mod p otherwise.
Lemma 17.10.8. An integral model X = {F1 , . . . , Fm } has good reduction mod p if and only
if Z[X0 , . . . , XN ]/(F1 , . . . , Fm ) ⊗ Fp is a regular ring.
Example 17.10.9. If X ⊆ P2Q is a plane conic and X is an integral model of X over Q given
by a primitive quadratic form F ∈ Z[X0 , X1 , X2 ], then X has bad reduction at a prime p if
and only if p divides the discriminant ∆(F ).
Corollary 17.10.10. A primitive quadratic form F ∈ Z[X0 , X1 , X2 ] has bad reduction at
only finitely many primes.
The following is a version of Hensel’s Lemma that we will need for lifting solutions of
quadratic forms.
Theorem 17.10.11. Let (R, v) be a complete DVR, f ∈ R[x1 , . . . , xN ] and suppose (a1 , . . . , aN ) ∈
RN such that  
∂f
v(f (a1 , . . . , aN )) > 2v (a1 , . . . , aN )
∂xi
for some 1 ≤ i ≤ N . Then f has a root in RN .
Corollary 17.10.12. If X = Z(F ) is an integral model over Zp and P is a smooth point of
X (Fp ), then P lifts to a point of X (Zp ).
Corollary 17.10.13. Let β ∈ Z× 2 2
p . Then x = β has a solution in Zp if and only if x ≡ β
mod pε has a solution, where ε = 3 when p = 2 and ε = 1 otherwise.
Proof. Let f (x) = x2 − β ∈ Z[x] so that f 0 (x) = 2x. Suppose α0 ∈ Zp is a solution to
f (x) ≡ 0 mod pε , i.e. v(f (α0 )) ≥ ε. Then α02 ≡ β 6= 0 mod p so since Zp is a DVR, α0
must be a unit, i.e. v(α0 ) = 0. Now we have
2vp (f 0 (α0 )) = 2vp (2a0 ) = 2(vp (a0 ) + vp (2))
(
2, p = 2
= 2vp (2) =
0, p 6= 2
< v(f (α0 )) in all cases.
Therefore Hensel’s Lemma applies.
This leaves the question of lifting singular points.
Pn 2
Theorem 17.10.14. Let F = i=0 ai Xi be a primitive quadratic form over Zp and set
X = Z(F ) ⊆ P2Qp . Suppose β0 , . . . , βn ∈ Zp such that ordp (βj ) = 0 for some 0 ≤ j ≤ n, with
F (β0 , . . . , βn ) = 0 mod pε+1 , where
(
1, p 6= 2
ε=
3, p = 2.

Then there exists a nontrivial root of F in Zp , that is, α = (α0 , . . . , αn ) ∈ Znp , with α` 6= 0
for some 0 ≤ ` ≤ n, and F (α0 , . . . , αn ) = 0.

422
17.10. Rational Points of Conics Chapter 17. Curves

Proof. Since F is primitive, ai , βj ∈ Z× p for some 0 ≤ i, j ≤ n. If i = j, then the point


P = (β̄0 , . . . , β̄n ) is a smooth point of XFp . By Theorem 17.10.11, P lifts to a solution in
Zp . On the other hand, assume without loss of generality that β0 ∈ Z× ×
p and a0 6∈ Zp . Then
0 0 × 0 2
a0 = pa0 for some a0 ∈ Zp . Set c = a1 β1 + . . . + an βn . Then pa0 β + c ≡ 0 (mod pε+1 )
2 2

so p | c; write c = pc0 . Then pa00 β02 + c0 ≡ 0 (mod pε ). This implies c0 ∈ Z× p – in fact,


0 ε c0 × c0
c ∈ 1 + p Op – so a0 ∈ Zp . In particular, − a0 is a square in Zp by Corollary 17.10.13.
0 0
0
Write − ac0 = θ2 for θ ∈ Zp . Then α = (θ, β1 , . . . , βn ) is a solution to F (α) = 0 over Zp as
0
required.
We have proven the following theorem characterizing rational points of quadratic forms
(conics) over Q.

Theorem 17.10.15. Let F be a nondegenerate, primitive quadratic form over Z and let
X = Z(F ) be the corresponding conic over Q. Then X(Q) 6= ∅ if and only if

(1) There is a sign change in the coefficients – i.e. X(R) 6= ∅.

(2) F = 0 has a primitive solution mod 16 – i.e. X(Q2 ) 6= ∅.

(3) F = 0 has a primitive solution mod p2 for all primes p > 2 – i.e. X(Qp ) 6= ∅.

In practice, one need only check (2) and (3) for primes at which X has bad reduction,
and by Corollary 17.10.10 there are only finitely many of these.

423
Chapter 18

Complex Varieties

This chapter is a survey of the theory of complex manifolds and varieties. This part of
the notes comes from a course on 4-manifolds taught by Dr. Tom Mark at the University
of Virginia in Spring 2019. I stuck around for the beginning of the course to take notes
on complex manifolds, complex surfaces and surface classification since these topics are
developed in parallel to the algebro-geometric theory of surfaces. The main reference for
this course is Griffiths-Harris, Principles of Algebraic Geometry.

424
18.1. Complex Manifolds Chapter 18. Complex Varieties

18.1 Complex Manifolds


Recall that a complex function f : C → C is holomorphic if it satisfies the Cauchy-Riemann
equations:
∂u ∂v ∂u ∂v
= and =−
∂x ∂y ∂y ∂x
where f = u + iv. To express this for functions of multiple complex variables, we can
reinterpret this definition as follows. Set z = x + iy and define differentials dz = dx + i dy,
dz̄ = dx − i dy and
∂f ∂f
df = dx + dy.
∂x ∂y
Then dx = 21 (dz + dz̄), dy = − 2i (dz − dz̄) and
   
1 ∂f ∂f 1 ∂f ∂f
df = −i dz + +i dz̄
2 ∂x ∂y 2 ∂x ∂y

which suggests defining the following complex derivatives:


   
∂f 1 ∂f ∂f ∂f 1 ∂f ∂f
= −i and = +i .
∂z 2 ∂x ∂y ∂ z̄ 2 ∂x ∂y
∂f ∂f
Then df = ∂z
dz + ∂ z̄
dz̄.

Lemma 18.1.1. A complex function in one variable f = u + iv : C → C is holomorphic if


and only if ∂f
∂ z̄
= 0.

For a multivariable function f : Cn → C, write zj = xj + iyj , 1 ≤ j ≤ n, for the


coordinates on Cn . Then
n n n n
X ∂f X ∂f X ∂f X ∂f
df = dxj + dyj = dzj + dz̄j
j=1
∂x j j=1
∂y j j=1
∂z j j=1
∂ z̄j

∂f ∂f
where ∂zj
and ∂ z̄j
are defined in the same way as their single-variable counterparts above.

Definition. Let f : Cn → C be a complex function in the variables zj = xj + iyj , 1 ≤ j ≤ n,


and define
n n
X ∂f ¯
X ∂f
∂f = dzj and ∂f = dz̄j .
j=1
∂z j j=1
∂ z̄j

¯ = 0.
Then f is holomorphic if ∂f

Lemma 18.1.2. A function f : Cn → C is holomorphic if and only if it satisfies the Cauchy-


Riemann equations in each variable zj .

Definition. More generally, f = (f1 , . . . , fm ) : Cn → Cm is holomorphic if f1 , . . . , fm are


all holomorphic.

425
18.1. Complex Manifolds Chapter 18. Complex Varieties

Definition. A complex n-manifold is a Hausdorff topological space that is locally homeo-


morphic to (an open subset of ) Cn and has holomorphic transition functions. Explicitly, M
is a complex manifold if it admits a complex atlas: a choice of open covering {Ui } of X
together with homeomorphisms ϕi : Ui → ϕi (Ui ) ⊆ Cn such that for each pair of overlapping
charts Ui , Uj , the transition map

ϕij := ϕ−1
j ◦ ϕi : ϕi (Ui ∩ Uj ) −→ ϕj (Ui ∩ Uj )

and its inverse are holomorphic. A complex structure on M is the choice of such a
complex atlas, up to holomorphic equivalence of charts, defined by a similar condition to the
above.

Let M be a complex n-manifold and fix a point p ∈ M . Then in a chart U ⊆ M around


p, the complex structure determines local parameters dx1 , . . . , dxn , dy1 , . . . , dyn . As in the
previous section, there are also complex parameters dzj = dxj + i dyj and dz̄j = dxj − i dyj
for 1 ≤ j ≤ n. Each of the dxj , dyj , dzj , dz̄j are R-linear functionals Tp M → R. Writing
dzj = dxj + i dyj extends these to C-linear functionals  Tp M →  C (and likewisefor dz̄j ), so 
each dzj , dz̄j ∈ Tp∗ M ⊗R C. Similarly, each ∂z∂ j = 12 ∂x∂ j − i ∂y∂ j and ∂∂z̄j = 21 ∂x∂ j + i ∂y∂ j
are elements of the complexified tangent space Tp M ⊗R C. Note that Tp M ⊗ C and Tp∗ ⊗ C
are both 2n-dimensional complex vector spaces (so 4n-dimensional real vector spaces).
The above functionals are cotangent vectors and can be evaluated on the above tangent
vectors as follows: for 1 ≤ j, k ≤ n,
    
∂ 1 ∂ ∂
dzj = (dxj + i dyj ) −i
∂zk 2 ∂xk ∂yk
         
1 ∂ ∂ ∂ ∂
= dxj + dyj + i dyj − dxj
2 ∂xk ∂yk ∂xk ∂yk
1
= [(δij + δij ) + i(0 − 0)] = δij .
2
      n o
Likewise, dz̄j ∂∂z̄k = δij and dzj ∂∂z̄k = dz̄j ∂z∂k = 0. Thus ∂z∂ 1 , . . . , ∂z∂n , ∂∂z̄1 , . . . , ∂∂z̄n
and {dz1 , . . . , dzn , dz̄1 , . . . , dz̄n } are dual bases for the vector spaces Tp M ⊗ C and Tp∗ M ⊗ C,
respectively. As a consequence, there are several different notions of “tangent space” for a
complex manifold:

ˆ Real tangent space: the underlying real vector space of Tp M , which has real dimen-
sion 2n;

ˆ Complex tangent space: the complex vector space TpC M := Tp M ⊗ C, which has
complex dimension 2n;
n on
ˆ Holomorphic tangent space: the subspace Tp0 M := SpanC ∂z∂ j of TpC M ;
j=1
n on
ˆ Antiholomorphic tangent space: the subspace Tp00 M := SpanC ∂
∂ z̄j
of TpC M .
j=1

426
18.1. Complex Manifolds Chapter 18. Complex Varieties

We also define the real and complex cotangent spaces Tp∗ M and TpC,∗ M , as well as their
holomorphic and antiholomorphic counterparts Tp1,0 M := (Tp0 M )∗ and Tp0,1 M := (Tp00 M )∗ ,
respectively.

Proposition 18.1.3. Let M be a complex manifold of dimension n and p ∈ M . Then

(1) TpC M = Tp0 M ⊕ Tp00 M .

(2) Complex conjugation acts C-linearly on TpC M and exchanges Tp0 M and Tp00 M .

(3) TpC,∗ M = Tp1,0 M ⊕ Tp0,1 M , and specifically we have

Tp1,0 M = {α ∈ TpC,∗ M | α ≡ 0 on Tp00 M } = SpanC {dzj }nj=1


Tp0,1 M = {α ∈ TpC,∗ M | α ≡ 0 on Tp0 M } = SpanC {dz̄j }nj=1 .

(4) Complex conjugation acts C-linearly on TpC,∗ M and exchanges Tp1,0 M and Tp0,1 M .

Suppose M m and N n are complex manifolds with open subsets U ⊆ M, V ⊆ N , and


f : U → V is an arbitrary function. For 1 ≤ j ≤ m, write the coordinates on U as
zj = xj + iyj and similarly, for 1 ≤ k ≤ n, write the components of f on V as wk = uk + ivk .
The real Jacobian of f is an m × n matrix
 
∂uk ∂uk
 ∂xj ∂yj 
JR f = 
 ∂vk ∂vk 

∂xj ∂yj

which defines a linear map JR f : Tp M → Tf (p) N for any p ∈ U . Extending scalars allows
us to define the complex Jacobian JC f : TpC M → TfC(p) N given by the same matrix JR f with
n o n o
∂ ∂ ∂ ∂ ∂ ∂ ∂ ∂
respect to the real bases ∂xj , ∂ x̄j , ∂yj , ∂ ȳj and ∂uk , ∂ ūk , ∂vk , ∂v̄k of these tangent spaces.
n o n o
With respect to the complex bases ∂z∂ j , ∂∂z̄j and ∂w∂ k , ∂ w̄∂ k , the matrix is
 
∂wk ∂wk
 ∂zj ∂ z̄j 
JC f = 
 ∂ w̄k ∂ w̄k  .

∂zj ∂ z̄j

Proposition 18.1.4. Let U ⊆ M, V ⊆ N and f : U → V be as above. Then the following


are equivalent:

(1) f is holomorphic.
∂wk ∂ w̄k
(2) For all 1 ≤ j ≤ m and 1 ≤ k ≤ n, ∂ z̄j
= ∂zj
= 0.

(3) For all p ∈ U , JC f takes Tp1,0 M to Tf1,0 0,1 0,1


(p) N and Tp M to Tf (p) N .

427
18.1. Complex Manifolds Chapter 18. Complex Varieties

Corollary 18.1.5. Suppose M and N are complex manifolds of the same dimension and
f : M → N is holomorphic. Then
  2
∂w k
≥ 0.
det(JR f ) = det(JC f ) = det
∂zj

Corollary 18.1.6. A holomorphic map f : M → N between complex manifolds of the same


dimension is orientation-preserving.

In particular, transition maps are orientation-preserving, so every complex manifold has


a naturally defined orientation.

Example 18.1.7. Note that the operators ∂z∂ j and ∂∂z̄j act as expected on the coordinates
{zj , z̄j }. For example, consider the map f : C2 → C, f (z, w) = z 2 + w3 . Then

JC f = 2z 3w

which is surjective whenever (z, w) 6= (0, 0) so for any constant c ∈ C r {0}, the equation
z 2 + w3 = c defines a smooth submanifold of C2 .

More generally, there are holomorphic versions of the inverse and implicit function theo-
rems from differential geometry:

Theorem 18.1.8 (Inverse Function Theorem). If f : M → N is holomorphic and JC f is


an isomorphism at p ∈ M , then there is a neighborhood V ⊆ N of f (p) and an inverse
f −1 : V → f −1 (V ) to f around p that is holomorphic on V .

Theorem 18.1.9 (Implicit Function Theorem). Let f : M1 × M2 → N be a holomorphic


function such that JC f is invertible at (p, q) ∈ M1 × M2 and set w = f (p, q). Then there is
a unique holomorphic map g : M → N such that f (p, g(p)) = w.

To prove these, one need only prove that f −1 (in the inverse function theorem) and g
(in the implicit function theorem) are holomorphic. The rest follows from real differential
geometry.

Remark. In the above example, the complex submanifold defined by the equation z 2 +w3 = c
is not compact. In fact, a complex submanifold of Cn is never compact. To see this, fix an
embedding M ,→ Cn of a complex submanifold. Then the composition f : M ,→ Cn → C
with projection onto the ith component is a holomorphic function, so by the maximum
principle, f must be constant. A version of this argument can be used to prove the following
important fact.

Theorem 18.1.10. If M is a closed complex manifold, then there are no nonconstant holo-
morphic functions M → C.

Example 18.1.11. Let Pn be complex projective n-space, defined as the quotient Cn+1 r
{0}/ ∼ where (z0 , . . . , zn ) ∼ (w0 , . . . , wn ) if wj = λzj for some λ ∈ C r {0}. Alternatively,
Pn is the space of complex lines through the origin in Cn+1 . Write the coordinates on Pn as

428
18.1. Complex Manifolds Chapter 18. Complex Varieties

[z0 , . . . , zn ]. There is a natural complex structure on Pn via the open sets Uj = {[z0 , . . . , zn ] ∈
Pn | zj 6= 0} and charts

ϕj : Uj −→ Cn
 
z0 zj−1 zj+1 zn
[z0 , . . . , zn ] 7−→ ,..., , ,..., .
zj zj zj zj

One can check that the transition maps are holomorphic, so this gives Pn the structure of an
n-dimensional complex manifold. For any homogeneous polynomial f ∈ C[z  0 , . . . , zn ], the
n ∂f ∂f
equation f (z0 , . . . , zn ) = 0 defines an algebraic set V (f ) in P . If JC f = ∂z 0
· · · ∂z n
is
nonzero at every point in V (f ) except for the point corresponding to 0 ∈ Cn+1 , then V (f )
is a complex submanifold of Pn .

429
18.2. Dolbeault Cohomology Chapter 18. Complex Varieties

18.2 Dolbeault Cohomology


Let M be a complex manifold of dimension n and for k ≥ 1, let k T C,∗ M be the vector
V
bundle of differential k-forms on M . Then the splitting T C,∗ M = T 1,0 M ⊕ T 0,1 M from
Vk C,∗
Proposition 18.1.3 induces a splitting of T M:
^ M ^ ^ ^ ^ 
k C,∗ p,q C,∗ p,q C,∗ p 1,0
T M= T M where T M := T M ⊗ q T 0,1 M .
p+q=k

Each graded piece p,q T C,∗ M has basis {dzi1 ∧· · ·∧dzip ∧dz̄j1 ∧· · ·∧dz̄jq | 1 ≤ i1 , . . . , ip , j1 , . . . , jq ≤
V
n} which we can write more compactly as {dzI ∧ dz̄J : |I| = p, |J| = q} where I, J ⊂ N and
dzI = dzi1 ∧ · · · ∧ dzip for I = {i1 , . . . , ip } (likewise for dz̄J ). Passing to global sections of
this vector bundle gives a type decomposition on the vector space of differential k-forms:
 ^  M  ^  M
Ωk M := Γ M, k T C,∗ M = Γ M, p,q T C,∗ M =: Ωp,q M.
p+q=k p+q=k

We say a differential k-form ω ∈ Ωp,q M ⊆ Ωk M has type (p, q). Write


X
ω= fIJ dzI ∧ dz̄J
|I|=p,|J|=q

using the same indexing notation as above, we have


X X X
dω = dfIJ ∧ dzI ∧ dz̄J = ∂fIJ ∧ dzI ∧ dz̄J + ¯ IJ ∧ dzI ∧ dz̄J .
∂f
|I|=p,|J|=q |I|=p,|J|=q |I|=p,|J|=q

In the last expression, the first term is a differential form of type (p + 1, q) and the second
term is of type (p, q + 1). Put
X X
∂ω = ∂fIJ ∧ dzI ∧ dz̄J and ∂ω ¯ = ¯ IJ ∧ dzI ∧ dz̄J .
∂f
|I|=p,|J|=q |I|=p,|J|=q

Then dω = ∂ω + ∂ω ¯ with ∂ω ∈ Ωp+1,q M and ∂ω ¯ ∈ Ωp,q+1 M . Observe that 0 = d2 =


∂ + ∂ ∂¯ + ∂∂
2 ¯ + ∂¯ and each piece lands in a different direct summand of Ωk+1 M , so we must
2

have ∂ 2 = ∂¯2 = 0 and ∂ ∂¯ = −∂∂.


¯ This proves:

Lemma 18.2.1. For any complex manifold M , there are C-linear operators

∂ : Ωp,q M → Ωp+1,q M and ∂¯ : Ωp,q M → Ωp,q+1 M

which satisfy ∂ 2 = ∂¯2 = 0 and ∂ ∂¯ = −∂∂.


¯

Definition. The Dolbeault cohomology of a complex manifold M is the bigraded vector


space M
H •,• (M ) = H p,q (M ) ¯
where H p,q (M ) := H q (Ωp,• (M ), ∂).
p,q≥0

430
18.2. Dolbeault Cohomology Chapter 18. Complex Varieties

Remark. Taking the differential ∂ on Ω•,q (M ) yields an isomorphic version of H •,• (M ).

Definition. A differential form on M of type (p, 0) is called a holomorphic p-form.

A holomorphic differential form on M is locally of the form


X
ω= fI dzI
|I|=p

where the fI are holomorphic functions.

Remark. The grading Ωk (M ) =


L p,q
Ω (M ) together with the decomposition of the differ-
¯ •
ential d = ∂ + ∂ gives Ω (M ) the structure of a double complex of vector spaces. There is
an associated spectral sequence, called the Hodge-de Rham spectral sequence, whose E 2 -page

is given by H p,q (M ) and which converges to HdR (M ). In particular, if hp,q = dimC H p,q (M )
k
and bk = dimC HdR (M ) is the kth Betti number of M , then
X
hp,q ≥ bk
p+q=k

for all k ≥ 0. In the case when M is a Kähler manifold (see Section 18.3),
Pthen the Hodge-de
Rham spectral sequence collapses on the second page, in which case p+q=k hp,q = bk for
all k. Moreover, when M embeds holomorphically into Cn for some n ≥ 1, then M has
the cohomology of an affine complex variety, so H p,q (M ) = 0 for all q > 0 and there is an
isomorphism HdR p
(M ) ∼
= H p (H •,0 (M ), ∂) for all p ≥ 0.
Example 18.2.2. Let E be a complex torus given by C/Λ for a lattice Λ ⊆ C. We may
write Λ = Z ⊕ Zτ for some τ = a + bi ∈ C such that b > 0. Then E carries the structure of
a complex elliptic curve (see Section 23.2). Using the above techniques, one may show that

H 1,0 (E) = Chdzi ∼


= SpanC {1, τ } ⊂ C2 and H 0,1 (E) = Chdz̄i ∼
= SpanC {1, τ̄ } ⊂ C2 .

So different choices of τ determine different subspaces of H 2 (E; C) ∼ = C2 and therefore


different complex structures on E. In fact, the set of all elliptic curves up to isomorphism
may be identified with the set of all τ ∈ C with b > 0.

Example 18.2.3. For n ≥ 1, complex projective n-space Pn has a distinguished differential


form known as the Fubini-Study form, which can be defined as follows. Recall that Pn has
the following cohomology groups:
(
C, k = 0, 2, . . . , 2n
H k (Pn ; C) =
0, otherwise.

Let z ∈ Pn be a point with a neighborhood U having holomorphic map Z : U → Cn+1 r {0}


which is a section of π : Cn+1 r {0} → Pn . Define ωF S ∈ Ω2 (Pn ) locally about z by
i ¯
ωF S = ∂ ∂ log |Z|2

431
18.2. Dolbeault Cohomology Chapter 18. Complex Varieties

where |·| is the usual Euclidean norm on Cn+1 . Then ωF S is a (1, 1)-form and is real, meaning
ω̄F S = ωF S . Moreover, dωF S = ∂ωF S = ∂ω ¯ F S = 0, so ωF S is a closed 2-form and defines a
class [ωF S ] ∈ H 1,1 (Pn ). On the other hand, one can show that ωF S is Poincaré dual to the
class of a hyperplane H ⊆ Pn representing Pn−1 ,→ Pn , so ωF S is not exact and therefore
[ωF S ] generates H 1,1 (Pn ). Since H 2 (Pn ; C) = H 2,0 (Pn ) ⊕ H 1,1 (Pn ) ⊕ H 0,2 (Pn ), it follows that

H 2 (Pn ; C) = H 1,1 (Pn ) ∼


= C and H 2,0 (Pn ) = H 0,2 (Pn ) = 0.

Recall that as a ring, H • (Pn ; C) ∼


= C[x]/(xn+1 ) where x is a generator in degree 2. The above
shows we can take x = [ωF S ], so that [ωFk S ] generates H k (Pn ; C) for each 1 ≤ k ≤ n. However,
since ωFk S is a (k, k)-form, this means the type decomposition of H • (Pn ; C) is precisely
(
C, p = q
H p,q (Pn ) =
0, p 6= q.

432
18.3. Kähler Manifolds Chapter 18. Complex Varieties

18.3 Kähler Manifolds


Recall that a Riemannian metric on a smooth manifold M is a smooth assignment g of an
inner product on Tx M for each x ∈ M . This can alternatively be characterized as a tensor
field g ∈ Γ(M, T ∗ M ⊗ T ∗ M ) that restricts to a symmetric, positive definite form at each
x ∈ M . If M is a complex manifold, then there is an R-linear endomorphism

J : T M −→ T M

satisfying J 2 = −id. Explicitly, on Tx M this can be written Jv = iv.


Definition. A Riemannian metric g on a complex manifold M is compatible with the
complex structure if g(Jv, Jw) = g(v, w) for all x ∈ M and v, w ∈ Tx M .
Proposition 18.3.1. Every complex manifold admits a compatible metric.
Fix a complex Riemannian manifold (M, g) with g a compatible metric. Extending J to
T M defines a C-bilinear form
C

gC : TxC M ⊗ TxC M −→ C

for each x ∈ M such that if v, w ∈ Tx0 M , then gC (v, w) = 0. This is because Tx0 M may be
written Tx0 M = {X − iJX | X ∈ Tx M }.
Definition. For a complex Riemannian manifold M with compatible metric g, the associ-
ated Hermitian form on M is the assignment h of an R-bilinear map h : Tx0 M ⊗Tx0 M → C
to each x ∈ M , defined by h(v, w) = gC (v, w̄) for all v, w ∈ Tx0 M .
Lemma 18.3.2. Let (M, g) be a complex manifold with compatible metric and associated
Hermitian form h. Then
(a) For all v, w ∈ Tx0 M , h(v, w) = h(w, v). In particular, h(v, v) ∈ R.

(b) For all v ∈ Tx0 M , h(v, v) ≥ 0 with equality if and only if v = 0.

(c) For all v, w ∈ Tx0 M and λ ∈ C, h(λv, w) = λh(v, w) = h(v, λ̄w).


Definition. A (1, 1)-form ω ∈ Ω1,1 (M ) is a positive form if ω̄ = ω and for all nonzero
v ∈ Tx0 M , −iω(v, v̄) > 0.
P
The reason for this terminology
P is that any real (1, 1)-form ω = fjk dzj ∧dz̄k determines
a Hermitian form h = −i fjk dzj ⊗ dz̄k and h is positive definite (and thus defines a metric
on M ) if and only if ω is a positive form.
Definition. To a complex manifold M with compatible metric g, we associated a (1, 1)-form
ω defined by
ω(X, Y ) = −g(X, JY ) for X, Y ∈ Tx M.
Then M is a Kähler manifold if ω is a closed form, and in this case g is called a Kähler
metric on M .

433
18.3. Kähler Manifolds Chapter 18. Complex Varieties

Example 18.3.3. On Pn , the Fubini-Study form ωF S is closed and of the form ω(X, Y ) =
−g(X, JY ) where g is the standard metric inherited from Pn+1 . Thus Pn is a Kähler manifold.

Example 18.3.4. On Cn , the usual inner product may be expressed as a tensor field
n n
X 1X
g= (dxj ⊗ dxj + dyj ⊗ dyj ) = (dzj ⊗ dz̄j + dz̄j ⊗ dzj )
j=1
2 j=1

where zj = xj + iyj . The corresponding Hermitian form on T 0 Cn is given by


   
∂ ∂ ∂ ∂ 1
h , =g , = δjk
∂zj ∂zk ∂zj ∂ z̄k 2

so the ∂z∂ j are mutually Hermitian-orthogonal of length √12 . Meanwhile, the associated 2-form
for this metric is determined by
   
∂ ∂ ∂ ∂
ω , = −g ,i =0
∂zj ∂zk ∂zj ∂zk
   
∂ ∂ ∂ ∂
ω , = −g ,i =0
∂ z̄j ∂ z̄k ∂ z̄j ∂ z̄k
   
∂ ∂ ∂ ∂ i
ω , = −g ,i = δjk .
∂zj ∂ z̄k ∂zj ∂ z̄k 2

So ω has the form n n


iX X
ω= dzj ∧ dz̄j = dxj ∧ dyj
2 j=1 j=1

as expected.

Example 18.3.5. Any Riemann surface (complex 1-manifold) is automatically a Kähler


manifold because any 2-form is closed.

Example 18.3.6. Let M be a Kähler manifold of complex dimension n, x ∈ M and choose


an orthonormal basis {ε1 , . . . , ε2n } of Tx∗ M with respect to a basis {e1 , . . . , e2n } of Tx M . We
may assume J : Tx M → Tx M acts by Je2j−1 = e2j for each 1 ≤ j ≤ n. Set ϕj = ε2j−1 + iε2j .
Then {ϕ1 , . . . , ϕn } is a basis for Tx1,0 M . The calculation on Cn in Example 18.3.4 then shows
that the associated (1, 1)-form ω on M locally has the form
n n
X iX
ω= ε2j−1 ∧ ε2j = ϕj ∧ ϕ̄j .
j=1
2 j=1

A short calculation shows that ω n = n! ε1 ∧ · · · ∧ ω2n = n! volM , where volM is the volume
form on M with respect to g. In particular this is easy to see from the calculation at the
end of Example 18.3.4 for Cn .

Proposition 18.3.7. Let (M, g) be a Kähler manifold and N ⊆ M any submanifold. Then
restricting g to N makes N a Kähler manifold.

434
18.3. Kähler Manifolds Chapter 18. Complex Varieties

Proof. Let i : N ,→ M be the subspace inclusion and write ωM and ωN for the associated
(1, 1)-forms on M and N , respectively. Then ωN = i∗ ωM and since d commutes with pullback,
we get dωN = 0. Therefore N is Kähler.
Corollary 18.3.8. On a complex manifold M of dimension n with Hermitian inner product
h and associated (1, 1)-form ω, the volume form on M is given by volM = n!1 ω n . In particular,
for any complex submanifold N ⊆ M of dimension d,
Z
1
volM (N ) = ωd.
d! N
Corollary 18.3.9. If (M, g) is a Kähler manifold, then
(1) The volume of any closed complex submanifold N ⊆ M depends only on the homology
class [N ] ∈ H • (M ).

(2) A closed complex submanifold N ⊆ M has [N ] 6= 0.


Corollary 18.3.10. If (M, g) is a compact Kähler manifold then for each 0 ≤ k ≤ n,
[ω k ] 6= 0 and in particular H 2k (M ) 6= 0.
Proof. If [ω k ] = 0 then ω k = dα for some α ∈ Ω2k−1 (M ). In this case, we have
Z Z
n
n! vol(M ) = ω = ω k ∧ ω n−k
ZM M Z
n−k
= dα ∧ ω = (dα ∧ ω n−k + α ∧ dω n−k ) since ω n−k is closed
ZM M

= d(α ∧ ω n−k ) by the Leibniz rule


ZM
= α ∧ ω n−k = 0 by Stokes’ theorem.
∂M

But this implies vol(M ) = 0 which is impossible.


Theorem 18.3.11 (Hodge Decomposition). If M is a compact Kähler manifold, then for
each k ≥ 0 there is an isomorphism

(M ) ∼
M
k
HdR = H p,q (M ).
p+q=k

Proof. (Sketch) Let ∆ be the Laplace-Beltrami operator on M , defined by

∆ : Ω• (M ) −→ Ω• (M )
ω 7−→ dd∗ ω + d∗ dω

where d∗ : Ωk (M ) → Ωk−1 (M ) is the adjoint of the exterior derivative on M . Explicitly, d∗


is uniquely determined by hdα, βi = hα, d∗ βi for all α, β ∈ Ω• (M ). A differential k-form ω
on M is called harmonic if ∆ω = 0; this is equivalent to dω = d∗ ω = 0, which is easy to
prove using the inner product definition of d∗ . Thus a harmonic form ω ∈ Ωk (M ) defines a

435
18.3. Kähler Manifolds Chapter 18. Complex Varieties

k
cohomology class [ω] ∈ HdR (M ). Let Hk (M ) be the vector space of all harmonic k-forms on
M . The Hodge theorem says that the natural inclusion

Hk (M ) ,−→ HdR
k
(M )

is in fact an isomorphism for all k ≥ 0, so every de Rham cohomology class has a unique
harmonic representative.
On the other hand, since M is Kähler, its metric is compatible with the complex structure
on M in aLway that ensures ∆ preserves the type decomposition of differential k-forms
Ωk (M ) = p,q
p+q=k Ω (M ), and therefore this type decomposition is also compatible with
the ‘anti-holomorphic Laplace operator’

∆∂¯ = ∂¯∂¯∗ + ∂¯∗ ∂¯ : Ω• (M ) → Ω• (M ).

Hence the Hodge theorem, suitably applied to ∆∂¯, says that every Dolbeault cohomology
class has a unique ∆∂¯-harmonic representative, so there are isomorphisms

(M ) ∼ Hp,q (M ) ∼
M M
k
HdR = Hk (M ) = = H p,q (M ).
p+q=k p+q=k

Example 18.3.12. Let Σ = Σg be a Riemann surface of genus g, which is Kähler by


Example 18.3.5. Note that for any p or q > 1, H p,q (Σ) = 0 since the complex dimension of
Σ is 1. Then the Hodge decomposition for Σ is:

C∼ 2
= HdR (Σ) = H 1,1 (Σ)
C2g ∼ 1
= HdR (Σ) = H 1,0 (Σ) ⊕ H 0,1 (Σ) ∼
= Cg ⊕ Cg since H 1,0 (Σ) ∼
= H 0,1 (Σ)
C∼ = H 0 (Σ) = H 0,0 (Σ).
dR

This information can be encoded in a Hodge diamond:


1
g g
1

For an arbitrary Kähler manifold M , H p,q (M ) ∼


= H q,p (M ) so if we set hp,q = dimC H p,q (M ),
p,q q,p
then h = h for all p, q ≥ 0. Thus the Hodge diamond for any Kähler manifold is always
symmetric about the vertical axis. In fact, Poincaré duality ensures that the diamond is also
symmetric about the horizontal line k = dim M/2.
Example 18.3.13. Let Σg and Σh be Riemann surfaces of genera g and h, respectively, and
set M = Σg × Σh . It is easy to check M is also Kähler. By the Künneth theorem, we have
2
HdR (M ) = H 2,0 (M ) ⊕ H 1,1 (M ) ⊕ H 0,2 (M )
= (H 1,0 (Σg ) ⊗ H 1,0 (Σh )) ⊕ (H 1,1 (Σg ) ⊗ H 0,0 (Σh )) ⊕ (H 1,0 (Σg ) ⊗ H 0,1 (Σh ))
⊕ (H 0,1 (Σg ) ⊗ H 1,0 (Σh )) ⊕ (H 0,0 (Σg ) ⊗ H 1,1 (Σh )) ⊕ (H 0,1 (Σg ) ⊗ H 0,1 (Σh ))

= (Cg ⊗ Ch ) ⊕ C ⊕ (Cg ⊗ Ch ) ⊕ (Cg ⊗ C) ⊕ C ⊕ (Cg ⊗ Ch ).

436
18.3. Kähler Manifolds Chapter 18. Complex Varieties

Similarly,
1
HdR (M ) = H 1,0 (M ) ⊕ H 0,1 (M )
= (H 1,0 (Σg ) ⊗ H 0,0 (Σh )) ⊕ (H 0,0 (Σg ) ⊗ H 1,0 (Σh ))
⊕ (H 0,1 (Σg ) ⊗ H 0,0 (Σh )) ⊕ (H 0,0 (Σh ) ⊗ H 0,1 (Σh ))

= Cg ⊕ Ch ⊕ Cg ⊕ Ch .

Applying the symmetry described above, we get the following Hodge diamond for M :

1
g+h g+h
gh 2(gh + 1) gh
g+h g+h
1

Remark. The duality


H p,q (M ) ∼
= H n−p,n−q (M )
inducing the symmetry of the Hodge diamond for a Kähler manifold is an instance of Serre
duality.

437
18.4. Analytic Varieties Chapter 18. Complex Varieties

18.4 Analytic Varieties


Definition. Let M be a complex manifold. An analytic variety in M is a subset V ⊆ M
given locally on an open subset U ⊆ M by the vanishing locus

V ∩ U = V (f1 , . . . , fk )

of some holomorphic functions f1 , . . . , fk : U → C.


Example 18.4.1. An analytic hypersurface in M is an analytic variety which is locally de-
fined by a single holomorphic function f : U → C. Such varieties have complex codimension
1 in M .
Example 18.4.2. Any complex submanifold N (say of dimension n) of M (of dimension
m) is an analytic subvariety of M , as N is locally defined by the vanishing of the last m − n
coordinates in a complex chart.
From complex analysis, we know that if f : C → C is a holomorphic function, then it is
analytic, meaning in an open disk D about z0 ∈ C, f is equal to its Taylor series expansion:
Z
1 f (w)
f (z) = dw.
2πi ∂D w − z
In several complex variables, a similar argument says that if U ⊆ Cn is open then f (z1 , . . . , zn ) :
U → C is holomorphic if and only if f is locally of the form
X
f= ak1 ,...,kn (z1 − u1 )k1 · · · (zn − un )kn

in a neighborhood of (u1 , . . . , un ) and this power series converges on that neighborhood. One
useful consequence of this fact is that if f, g : U → C are holomorphic, U is connected and
f and g agree on some open subset of U , then f ≡ g on all of U . (This is sometimes called
the principle of analytic continuation.)
Definition. A Weierstrass polynomial on an open set U ⊆ Cn is a holomorphic function
g : U → C of the form

g(z1 , . . . , zn−1 , w) = wd + g1 (z1 , . . . , zn−1 )wd−1 + . . . + gd (z1 , . . . , zn−1 )

for some holomorphic functions g1 , . . . , gd : Cn−1 → C. More generally, a Weierstrass


polynomial on a complex manifold M is one that is locally given by a Weierstrass polynomial
on the complex charts of M .
We have the following analogue of the fundamental theorem of calculus for complex
functions of several variables.
Theorem 18.4.3 (Weierstrass Preparation). Let f be a holomorphic function on a neigh-
borhood of 0 in Cn = Cn−1
z × Cw , where z = (z1 , . . . , zn−1 ), and suppose f (0, . . . , 0, w) 6≡ 0.
Then on a (possibly smaller) neighborhood of 0, one can write f = gh for a Weierstrass
polynomial g and a holomorphic function h such that h(0) 6= 0.

438
18.4. Analytic Varieties Chapter 18. Complex Varieties

Example 18.4.4. Let V = V (f ) be a hypersurface in Cn . After a coordinate change, we may


assume f satisfies f (0, . . . , 0, w) 6≡ 0. Then by the Weierstrass preparation theorem, f = gh
in a neighborhood of 0, for g a Weierstrass polynomial and h(0) 6= 0, so V = V (g). Thus
we may always assume hypersurfaces are (locally) cut out by a Weierstrass polynomial. For
each z = (z1 , . . . , zn−1 ), the zero set of g(z, w) as a function of w is the collection of discrete
roots of a polynomial, so other than where there are multiple roots, each of these roots are in
fact holomorphic functions in the multivariable z. Thus the projection (z, w) 7→ z exhibits
V as a branched cover of (an open set of) Cn−1 with branch locus precisely equal to

{z ∈ Cn−1 | g(z, w) has a multiple root} = V (disc(g))

where disc(g) is the discriminant of g. Therefore the branch locus of a hypersurface is itself
an analytic variety. For example, take n = 2 and g(z, w) = w2 − z. Then the projection
V → C, (z, w) 7→ z 2 gives the classic degree 2 branched cover of C with branch locus {0}.

Real coordinates of V = V (w2 − z)

Theorem 18.4.5 (Extension). Let D be a disk around 0 ∈ Cn and f : D → C be a


holomorphic function such that f (0) = 0. Suppose h : D r V (f ) → C is holomorphic and
bounded. Then h extends uniquely to a holomorphic function h̃ : D → C.
For a complex manifold M and a point P ∈ M , let OP = OM,P be the local ring of the
sheaf of holomorphic functions at P (i.e. the germs of holomorphic functions defined at P ).
Then the Extension Theorem implies that each OP is an integral domain. What’s more, OP
is a UFD:
Proposition 18.4.6. For each P ∈ M , the local ring OM,P is a UFD.
Proof. (Sketch) Since the question is local, we may assume M = Cn and P = 0. We induct
on n. For f ∈ OCn ,0 , we can use the Weierstrass preparation theorem (18.4.3) to write
f = gh for some Weierstrass polynomial g ∈ OCn−1 ,0 [w] and some holomorphic h ∈ OCn ,0
with h(0) 6= 0, i.e. h ∈ OC×n ,0 . By induction, OCn−1 ,0 is a UFD and by Gauss’s Lemma, so
is OCn−1 ,0 [w], so g is a product of irreducibles uniquely up to a unit. Since h is a unit, this
gives an irreducible factorization of f uniquely up to a unit as well.
Theorem 18.4.7 (Division Algorithm). Let f, g ∈ OM,P where g is a Weierstrass polynomial
of degree k. Then f = gh + r for some h, r ∈ OM,P with r a Weierstrass polynomial of degree
strictly less than k.

439
18.4. Analytic Varieties Chapter 18. Complex Varieties

Proof. Follows a similar proof as for the polynomial division algorithm.

Corollary 18.4.8. If g ∈ OM,P is irreducible and h ∈ OM,P such that h ≡ 0 on V (g), then
g | h.

These results show that for any P ∈ M , OM,P more or less behaves like a polynomial
ring. This makes the connection to algebraic geometry even more explicit.

440
18.5. Divisors, Line Bundles and Sections Chapter 18. Complex Varieties

18.5 Divisors, Line Bundles and Sections


The notion of a divisor from Section 17.1 can be defined in the context of analytic varieties.
Consider the set of all hypersurfaces V ⊆ M , i.e. those analytic varieties defined locally
by V (f ) for some holomorphic function f . (The following definitions are the same for any
analytic subvariety of M , but we will not need them here.)
Definition. For a hypersurface V ⊆ M , a point P ∈ V is called a smooth point (or
nonsingular point) of V if on a neighborhood U ⊆ M of P intersecting V nontrivially, the
holomorphic differential map
∂P f : TP0 M −→ C
is nonzero. Otherwise, P is called a singular point of V . Write Vsm and Vsing for the set
of smooth and singular points, respectively, of V .
In analogy with Theorem 12.6.5, we have:
Proposition 18.5.1. For any hypersurface V ⊆ M , Vsing is a proper analytic subvariety of
V.
Proof. Locally, if V is defined by a holomorphic function f then Vsing is defined by the
∂f ∂f
vanishing of the partial derivatives ∂z 1
, . . . , ∂z n
.
Definition. A hypersurface in M is irreducible if it cannot be written as the union of two
hypersurfaces in M .
Proposition 18.5.2. Let V be a hypersurface in M . Then
(1) V is irreducible if and only if Vsm is connected.

(2) If V1 , . . . , Vm are the connected components of Vsm , then their closures V 1 , . . . , V m are
the irreducible components of V .
Corollary 18.5.3. Every hypersurface V ⊆ M has a unique decomposition

V = V1 ∪ · · · ∪ Vm

where Vj are irreducible hypersurfaces.


Definition. A divisor on a complex manifold M is a locally finite linear combination of
irreducible hypersurfaces in M ,
X
D= nV V, nV ∈ Z.
V ⊆M

Let Div(M ) be the abelian group of divisors on M under addition of coefficients:


X X X
mV V + nV V = (mV + nV )V.
V ⊆M V ⊆M V ⊆M

A divisor D is called effective if nV ≥ 0 for all irreducible hypersurfaces V ⊆ M .

441
18.5. Divisors, Line Bundles and Sections Chapter 18. Complex Varieties

Definition. Let g be a holomorphic function on a neighborhood U ⊆ M of a point P ∈ V .


The order of vanishing of g at P is

ordP (g) = max{k ≥ 0 | g = f k h for h ∈ OM,P }

where V (f ) = V ∩ U .

Lemma 18.5.4. For a hypersurface V ⊆ M and g ∈ OM (U ), the order of vanishing ordP (g)
is independent of the point P ∈ V ∩ U , that is, P →
7 ordP (g) is locally constant.

As a result, the order of vanishing ordV (g) of g along any hypersurface V ⊆ M is well-
defined.

Example 18.5.5. Let Σ be a Riemann surface. Then a divisor on Σ is just a formal linear
combination of points of Σ: X
D= nP P.
P ∈X

Thus Div(Σ) is the free abelian group on the points of Σ.

Example 18.5.6. The above situation is atypical, as many complex manifolds have no
divisors at all. However, projective manifolds, i.e. those admitting complex embeddings
M ,→ Pn , contain many hypersurfaces. For example, if V ⊆ Pn is a hypersurface then
M ∩ V is a hypersurface in M . Examples of such hypersurfaces in Pn can be found by taking
the vanishing set of any homogeneous polynomial f ∈ C[z0 , . . . , zn ].

Recall that a meromorphic function on M is an almost-everywhere defined function f


which is locally of the form f = hg where g and h are holomorphic and h is not identically
the zero function. Such an f is specificed by an open cover {Uα } of M and holomorphic
functions gα , hα ∈ OM (Uα ) such that hα 6≡ 0 and on Uα ∩ Uβ ,

gα hβ |Uα ∩Uβ = gβ hα |Uα ∩Uβ .

By Theorem 18.4.7 we may choose gα , hα relatively prime in OM (Uα ).

Example 18.5.7. Let M = P2 and consider the meromorphic function f : P2 → C defined


by
−z02 + z12 + z22
f (z0 , z1 , z2 ) = .
z1 z2
Then f defines a divisor (f ):

(f ) = V (−z02 + z12 + z22 ) − V (z1 z2 ).

Here, V (−z02 + z12 + z22 ) is called the divisor of zeroes of f sometimes written (f )0 , while
V (z1 z2 ) is called the divisor of poles and is denoted (f )∞ . Further, in this example V (z1 z2 )
is reducible (it’s the union of the z0 z2 - and z0 z1 -planes in P2 ) so we have

(f ) = V (−z02 + z12 + z22 ) − V (z1 ) − V (z2 ).

442
18.5. Divisors, Line Bundles and Sections Chapter 18. Complex Varieties

Definition. The principal divisor of a meromorphic function f : M → C is the divisor


X
(f ) = ordV (f )V
V ⊆M

where ordV (f ) := ordV (gα ) − ordV (hα ) for a local expression f = hgαα on some open Uα ⊆ M
with Uα ∩ V 6= ∅. Write
X X
(f )0 = ordV (gα ) and (f )∞ = ordV (hα )
V ⊆M V ⊆M

for the divisor of zeroes and divisor of poles of f , so that (f ) = (f )0 − (f )∞ .


There is an important connection between divisors andPline bundles which we exhibit
now for complex manifolds. Observe that any divisor D = nV V can be described locally
as a principal divisor of a meromorphic function: if V is locally defined by V (gα,v ) for each
Uα , then on Uα , D can be represented by (fα ) where
Y nV
fα = gα,V .
V ∩Uα 6=∅

These local expressions must satisfy the following condition on Uα ∩ Uβ :



(fα ) = (fβ ) ⇐⇒ ∈ OM (Uα ∩ Uβ )× .

(Here, OM (U )× denotes the nonvanishing holomorphic functions on U .) Set gαβ = ffαβ . Notice
that if gαβ = 1 on all overlaps, then the fα glue together to define a global meromorphic
function f such that D = (f ). However, in general we have gαβ gβγ gγα = 1 on all triple
overlaps Uα ∩ Uβ ∩ Uγ . Then the maps gαβ : Uα ∩ Uβ → C× are the transition maps of a line
bundle on M , denoted OM (D) or just O(D) if the underlying complex manifold is clear.
Definition. A holomorphic vector bundle on a complex manifold M is a holomorphic
map π : E → M which is a complex vector bundle such that the transition functions are
holomorphic.
Lemma 18.5.8. For any divisor D ∈ Div(M ), O(D) is a holomorphic line bundle.
Definition. The set of isomorphism classes of holomorphic line bundles on M , which is a
group under ⊗, is called the Picard group of M , written Pic(M ).
Theorem 18.5.9. For any complex manifold M , the assignment

Φ : Div(M ) −→ Pic(M )
D 7−→ O(D)

is a group homomorphism.
Proof. From the above description of O(D) using transition functions, it is clear that O(D1 +
D2 ) ∼
= O(D1 ) ⊗ O(D2 ) for any divisors D1 , D2 ∈ Div(M ).

443
18.5. Divisors, Line Bundles and Sections Chapter 18. Complex Varieties

Definition. Two divisors D1 , D2 ∈ Div(M ) are called linearly equivalent, denoted D1 ∼


D2 , if there exists a nonzero meromorphic function f on M such that D1 = D2 + (f ).
Lemma 18.5.10. Let D1 , D2 ∈ Div(M ). Then O(D1 ) ∼
= O(D2 ) if and only if D1 ∼ D2 .
Proof. The line bundles O(D1 ) and O(D2 ) are isomorphic if and only if there is an open
cover {Uα } of M and a collection of holomorphic functions hα ∈ O(Uα )× such that gαβ 2

1 i
hβ = hα ◦ gαβ on Uα ∩ Uβ , where gαβ are the transition functions defining O(Di ), i = 1, 2.
i
i
By the above, gαβ = ffαi where fαi are the meromorphic functions locally defining Di . So
β
O(D1 ) ∼
= O(D2 ) if and only if

fα2 fβ2
= 1 on Uα ∩ Uβ .
fα1 ◦ hα fβ ◦ hβ
n 2 o
This latter condition says that the collection f 1f◦h α
α
defines a global meromorphic function
α
f on M such that
X
D2 − D1 = (ordV (fα2 ) − ordV (fα1 ))V
V ⊆M

fα2
X  
= ordV V
V ⊆M
fα1
fα2
X  
= ordV V since hα is holomorphic
V ⊆M
fα1 ◦ hα
= (f ).

This completes the proof.


Let M(M ) denote the abelian group of meromorphic functions on M , under pointwise
addition.
Corollary 18.5.11. There is an exact sequence of abelian groups
Φ
M(M ) → Div(M ) −
→ Pic(M ).

Moreover, if M is compact then this extends to an exact sequence


Φ
0 → C× → M(M ) → Div(M ) −
→ Pic(M )

where C× → M(M ) is the map α 7→ cα , the constant function on α.


Example 18.5.12. Let M = Pn and consider the hyperplane H = V (f ) where f is a
homogeneous linear polynomial. Notice that different choices of f give hyperplanes with
isomorphic line bundles (since the ratio of any two homogeneous linear polynomials is a
global meromorphic on Pn ), so we might call L = O(H) the hyperplane line bundle on Pn .
For now, assume f = z0 . Consider the open cover {Uj } of Pn , where

Uj = {[z0 , . . . , zn ] ∈ Pn | zj 6= 0}.

444
18.5. Divisors, Line Bundles and Sections Chapter 18. Complex Varieties

On this cover, H is defined by a collection of functions fj : Uj → C given by


(
1, j = 0
fj ([z0 , . . . , zn ]) = z0
zj
, j 6= 0.

These determine the following transition maps of O(H):


fj zk
gjk = = on Uj ∩ Uk .
fk zj
On the other hand, there is a distinguished line bundle E → Pn defined by

E = {(`, v) | ` ∈ Pn , v ∈ `} ⊂ Pn × Cn+1 ,

called the tautological line bundle. Over the same open cover {Uj }, E is trivialized by

ϕ−1j : Uj −→ E|Uj
  
z0 zn
([z0 , . . . , zn ], λ) 7−→ [z0 , . . . , zn ], λ , . . . , 1, . . . ,
zj zj
 
λ
= [z0 , . . . , zn ], (z0 , . . . , zn ) .
zj
Then the transition maps for E are:
 
λ
(ϕj ◦ ϕ−1
k )([z0 , . . . , zn ], λ) = ϕj [z0 , . . . , zn ], (z0 , . . . , zn )
zk
 
λzj
= [z0 , . . . , zn ],
zk
E z
i.e. gjk = zkj , the inverse of the transition map on Uj ∩ Uk for O(H). This proves that
E∼ = O(H)∗ = O(−H).
Definition. For a holomorphic vector bundle π : E → M , a holomorphic section (resp.
meromorphic section) of π is a holomorphic (resp. meromorphic) function s : M → E
such that π ◦ s = idM . The group of all holomorphic (resp. meromorphic) sections of π is
denoted Γh (M, π) or Γh (M, E) (resp. Γ(M, π) or Γ(M, E)).
Locally, a holomorphic section s ∈ Γh (M, π) is determined by sections sα : Uα → E|Uα
such that sα = gαβ sβ on each overlap Uα ∩ Uβ .
P
Example 18.5.13. Suppose D = nV V ∈ Div(M ) is a divisor with local defining functions
fα on Uα . By definition, the transition maps of O(D) are gαβ = ffαβ on Uα ∩ Uβ so the fα glue
together to give a global section fD of O(D). If D is an effective divisor, fD is a holomorphic
section of O(D), but in general fD need only be meromorphic. Conversely, if L → M is a
line bundle and s : M → L is a nonvanishing meromorphic section, there is a natural divisor
associated to the pair (L, s), namely
X
(s) = ordV (s)V.
V ⊆M

445
18.5. Divisors, Line Bundles and Sections Chapter 18. Complex Varieties

Theorem 18.5.14. For any complex manifold M , there is a one-to-one correspondence

Div(M ) ←→ {(L, s) | L ∈ Pic(M ), s ∈ Γ(M, L) is nonvanishing}


D 7−→ (O(D), fD )
(s) →−7 (L, s).

This restricts to a one-to-one correspondence

Diveff (M ) ←→ {(L, s) | L ∈ Pic(M ), s ∈ Γh (M, L) is nonvanishing}

where Diveff (M ) denotes the subset of effective divisors on M .

Corollary 18.5.15. The image of Φ : Div(M ) → Pic(M ) is precisely the isomorphism


classes of line bundles on M admitting nonzero meromorphic sections.

For a divisor D ∈ Div(M ), define a vector space

L(D) = {f ∈ M(M ) | D + (f ) is effective},

as in Section 17.3.

Proposition 18.5.16. Let L = O(D) be the line bundle defined by a divisor D ∈ Div(M )
and let sD be a meromorphic section of L defined by a set of local defining functions fα for
D. Then there is a one-to-one correspondence

L(D) ←→ Γh (M, L)
f 7−→ f sD .

Example 18.5.17. For M = Pn and L = O(H) the hyperplane bundle, this correspondence
gives

Γ(Pn , O(H)) ∼
= L(H) = {f ∈ M(Pn ) | f has a pole of order ≤ 1 along H}.

Suppose H = V (z0 ). Then these functions all have the form

p(z0 , . . . , zn )
f (z0 , . . . , zn ) =
z0

where p is a homogeneous linear polynomial on Cn+1 , so Γ(Pn , O(H)) ∼


= Cn+1 . Similarly, for
n ⊗d n
any d ≥ 2, Γ(P , O(H) ) = Γ(P , O(dH)) can be identified with the space of homogeneous
polynomials p(z0 , . . . , zn ) of degree d, so

Γ(Pn , O(dH)) ∼
= Symd (Cn+1 )∗ .

446
18.6. Cohomology and Chern Classes Chapter 18. Complex Varieties

18.6 Cohomology and Chern Classes


p
Recall (Section 13.4) that for a space X, the Čech cohomology groups Ȟ (U, F ) for a sheaf
F on X, with respect to an open cover U = {Ui }, are defined as the cohomology of the
complex Y
C p (U, F ) := F (Ui0 ,...,ip )
i0 ,...,ip

with differential

d : C p (U, F ) −→ C p+1 (U, F )


p+1
!
X
α 7−→ (−1)k αi0 ,...,ik−1 ,ik+1 ,...,ip+1 |Ui0 ,...,ip+1
k=0

where Ui0 ,...,ip denotes the intersection Ui0 ∩ · · · ∩ Uip . We will use the fact that for any sheaf
0
F on X and any open cover U of X, Ȟ (U, F ) = F (X) = H 0 (X, F ) (Lemma 13.4.5) and
1
Ȟ (X, F ) → H 1 (X, F ) is always an isomorphism (Corollary 16.2.7).

Example 18.6.1. If M is a complex manifold, for each open U ⊆ M let M(U ) be the
vector space of meromorphic functions on U . Then U 7→ M(U ) defines a sheaf on M and
H 0 (M, M) = M(M ) by the above. Moreover, there is an injective morphism of sheaves
×
OM ,→ M and we have

Div(M ) ∼
= H 0 (M, M/O× ) and Pic(M ) ∼ ×
= H 1 (M, OM ).

Explicitly, the isomorphism Pic(M ) ∼ ×


= H 1 (M, OM ) is given by L 7→ [(gαβ )] where (gαβ ) is
the 1-cocycle given by the transition functions gαβ defining L (see Section 18.5). Recall from
Corollary 18.5.11 that for any complex manifold M , there is an exact sequence

M(M ) → Div(M ) → Pic(M ).

This in fact coincides with the long exact sequence in sheaf cohomology coming from the
short exact sequence of sheaves
× ×
0 → OM → M → M/OM → 0.

Recall that by de Rham’s theorem (Corollary 13.3.8), the singular cohomology of any real
manifold can be computed by de Rham cohomology. Dolbeault’s theorem is an analogous
statement for complex manifolds.

Theorem 18.6.2 (Dolbeault). For any complex manifold M and every p, q ≥ 0, there is an
isomorphism
H p,q (M ) ∼
= H q (M, Ωp ).
Consider the short exact sequence of sheaves
× exp
0→Z→
− OM −−→ OM →0

447
18.6. Cohomology and Chern Classes Chapter 18. Complex Varieties

where Z is the constant sheaf and exp denotes the exponential map f 7→ e2πif . This induces
a long exact sequence in sheaf cohomology:
× δ
· · · → H 1 (M, OM ) → H 1 (M, OM − H 2 (M, Z) → · · ·
)→
×
Again note that H 1 (M, OM ) = Pic(M ).

Definition. The (first) Chern class of a holomorphic line bundle L on M is the cohomology
×
class c1 (L) := δ([L]) ∈ H 2 (M, Z), where [L] denotes the class in H 1 (M, OM ) corresponding
to L.

The fact that the connecting map δ in a long exact sequence is a homomorphism implies
the following basic facts about Chern classes.

Lemma 18.6.3. Let L be a holomorphic line bundle on a complex manifold M . Then

(a) If L is trivial, then c1 (L) = 0.

(b) For any other holomorphic line bundle L0 on M , c1 (L ⊗ L0 ) = c1 (L) + c1 (L0 ). In


particular, c1 (L∗ ) = −c1 (L).

(c) For any holomorphic map f : N → M between complex manifolds, c1 (f ∗ L) = f ∗ c1 (L).

The same arguments define the first Chern class of any smooth (i.e. topological) line
bundle on M . Explicitly, replacing OM with the sheaf A of smooth functions on M , there
is also a short exact sequence of sheaves
exp
− A −−→ A× → 0.
0→Z→

The corresponding long exact sequence is


δ
· · · → H 1 (M, A) → H 1 (M, A× ) →
− H 2 (M, Z) → · · ·

Then H 1 (M, A× ) may be identified with the set of isomorphism classes of smooth line bundles
on M and for any such bundle L, we take c1 (L) to be the class δ([L]) ∈ H 2 (M, Z).

Proposition 18.6.4. For any holomorphic line bundle L → M , c1 (L) only depends on the
underlying smooth line bundle of L and this is classified by its first Chern class.

Proof. The morphism of sheaves OM → A fits into a commutative diagram


exp ×
0 Z OM OM 0

id
exp
0 Z A A× 0

Taking cohomology yields another commutative diagram

448
18.6. Cohomology and Chern Classes Chapter 18. Complex Varieties

× δ
H 1 (M, OM ) H 2 (M, Z)

id

0 = H 1 (M, A) H 1 (M, A× ) H 2 (M, Z) H 2 (M, A)

Then since A is acyclic, H 1 (M, A) = H 2 (M, A) = 0 so the connecting homomorphism in


the bottom row is an isomorphism. Both statements follow immediately.
Example 18.6.5. Let Σ = Σg be a Riemann surface of genus g with structure sheaf O = OΣ .
By Dolbeault’s theorem and Example 18.3.12, H 1 (Σ, O) ∼ = H 0,1 (Σ) ∼
= Cg and H 2 (Σ, O) ∼ =
0,2 1 × 1 ∼ 2g 2
H (Σ) = 0. Moreover, H (Σ, O ) = Pic(Σ), H (Σ, Z) = Z and H (Σ, Z) = Z since Σ ∼
is a real 2-manifold, so the long exact sequence coming from the exponential sequence on Σ
becomes
0 → Z2g → Cg → Pic(Σ) → Z → 0.
Hence Pic(Σ) ∼ = Z × (Cg /Z2g ). The second factor, H 1 (Σ, O)/H 1 (Σ, Z) ∼= Cg /Z2g , is called
the Jacobian of Σ, written J(Σ). Topologically, J(Σ) is homeomorphic to a torus of genus
g. (For more on Jacobian, see Part V). For a line bundle L over Σ, pairing c1 (L) with the
fundamental class [Σ] ∈ H2 (Σ; Z) defines an integer called the degree of L:

deg(L) = hc1 (L), [Σ]i.


P
Definition. Let Σ be a Riemann
P surface and D = nP P a divisor on Σ. The degree of
D is the integer deg(D) = nP .
Lemma 18.6.6. For any divisor D ∈ Div(Σ), deg(D) = deg(O(D)).
Proposition 18.6.7. If L ∈ Pic(Σ) such that deg(L) < 0 then L admits no nonzero holo-
morphic sections.
Proof. We saw in Example 18.5.13 that if L is a line bundle over Σ which admits a holomor-
phic section s, then L = O(D) where D = (s) is the effective divisor defined by s. Hence
deg(L) = deg(s) ≥ 0 since s is holomorphic.
Let M be a complex manifold and recall from Proposition 18.5.16 that there is an iso-
morphism L(D) ∼ = H 0 (M, O(D)) given by multiplication by sD , where sD is a meromorphic
section of L defined by a collection of local defining functions for D. More generally, let
E → M be a vector bundle and write

E(D) = {meromorphic sections s : M → E | D + (s) ≥ 0}.

Then the same argument proves:


Proposition 18.6.8. If D ∈ Div(M ) is a divisor and E → M is a vector bundle, then there
is an isomorphism

E(D) −→ H 0 (M, E ⊗ O(D))
f 7−→ f ⊗ sD .

449
18.6. Cohomology and Chern Classes Chapter 18. Complex Varieties

Definition. The holomorphic Euler characteristic of a vector bundle E → M is the


integer

X
χh (M, E) = (−1)i hi (E)
i=0
i i
where h (E) := dim H (M, E). The holomorphic Euler characteristic of a complex
manifold M is the holomorphic Euler characteristic of its structure sheaf,

χh (M ) = χh (M, OM ).

Proposition 18.6.9. For any complex manifold M ,



X
χh (M ) = (−1)i h0,i (M )
i=0

where hp,q (M ) = dim H p,q (M ).


Proof. Apply Dolbeault’s theorem.
Suppose D = V is a smooth, irreducible hypersurface in M and consider the restriction
map E → E|V .
Lemma 18.6.10. For an irreducible hypersurface V ⊆ M and any vector bundle E → M ,
the map E → E|V is a surjective morphism of sheaves fitting into a short exact sequence

0 → E ⊗ O(−V ) → E → E|V → 0.

Therefore there is a long exact sequence

0 → H 0 (M, E ⊗ O(−V )) → H 0 (M, E) → H 0 (V, E|V ) → H 1 (M, E ⊗ O(−V )) → · · ·

Corollary 18.6.11. If V ⊆ M is an irreducible hypersurface, then

χh (M, E) = χh (M, E ⊗ O(−V )) + χh (V, E|V ).

Corollary 18.6.12. Let Σ = Σg be a Riemann surface of genus g and L → Σ a line bundle.


Then for any point P ∈ Σ, χh (Σ, L) = χh (Σ, L ⊗ O(−P )) + 1.
Proof. For any P ∈ Σ, χh (P, L|P ) = h0 (L|P ) is the dimension of the fibre of L over P , which
is 1.
Corollary 18.6.13. For any Riemann surface Σ of genus g, χh (Σ) = 1 − g.
Proof. Applying Example 18.3.12, we get χh (Σ) = h0,0 (Σ) − h0,1 (Σ) = 1 − g.
From this we obtain a version of the Riemann-Roch theorem (17.3.12) for Riemann
surfaces.
Corollary 18.6.14 (Riemann-Roch Theorem). Suppose Σ is a Riemann surface of genus g
and D is a divisor on Σ. Then

χh (Σ, O(D)) = deg(D) − g + 1.

450
18.6. Cohomology and Chern Classes Chapter 18. Complex Varieties

Alternatively, this may be written χh (Σ, L) = deg(L) − g + 1 for any line bundle L ∈
Pic(Σ). We might want to know more about the dimensions h0 (L) and h1 (L) for a given
line bundle, but it turns out that these are not determined solely by the topology of L → Σ.
When M is compact, a version of Serre duality (Theorem 17.6.2) tells us that h1 (L) may be
computed as h0 of a different line bundle.

Theorem 18.6.15 (Serre Duality). Suppose M is a compact complex manifold of dimension


n and E → M is a vector bunde. Then there is an isomorphism of C-vector spaces

H p (M, E) ∼
= H n−p (M, E ∗ ⊗ O(KM ))∗

where (−)∗ denotes vector space dual and KM is a distinguished divisor on M which is defined
up to linear equivalence.

Here’s a sketch of the proof; along the way we will define KM ∈ Div(M ). Compare this
to its counterpart in Section 17.6. For any vector bundle E → M , define the vector spaces
 ^   ^ 
k k ∗ p,q p,q ∗
Ω (M, E) := Γ M, T M ⊗E and Ω (M, E) := Γ M, T M ⊗E .

Then we have a decomposition

Ωk (M, E) ∼
M
= Ωp,q (M, E).
p+q=k

An element of Ωp,q (M, E) can be thought of V


as a ‘vector-valued differential form of type (p, q),
with values in E’. Equivalently, sections of k T C,∗ M ⊗ E may be identified with alternating
multilinear maps T M ×T M → E and such maps have a natural type decomposition agreeing
with the above. Locally, ω ∈ Ωp,q (M, E) is of the form
X
ω= ωIJ dz I ∧ dz̄ J ⊗ σ
|I|=p,|J|=q

for some σ ∈ Γ(M, E).


We define a ‘star operator’ ? : Ωp,q (M, E) → Ωn−p,n−q (M, E ∗ )∗ as follows. Assume E and
M are endowed with Hermitian metrics which are compatible with the map E → M . When
E∼ = Cr is a trivial bundle, ? is uniquely defined by the formula

α ∧ ?β = hα, βi · d volM

for any α, β ∈ Ωp,q (M ) and volM a Hermitian volume form on M . Locally, given a Hermitian
n-frame ε1 , . . . , εn for T 1,0 M with dual basis ε̄1 , . . . , ε̄n for T 0,1 M , the operator ? acts by

?(εi1 ∧ · · · ∧ εip ∧ ε̄j1 ∧ · · · ∧ ε̄jq ) = ±εi01 ∧ · · · ∧ εi0n−p ∧ ε̄j1 ∧ · · · ∧ ε̄jn−q


0

where {i1 , . . . , ip , i01 , . . . , i0n−q } = {1, . . . , n}, {j1 , . . . , jq , j10 , . . . , jn−q


0
} = {1, . . . , n} and ± is
the product of signs of these permutations of (1, . . . , n). For example, when n = 3,

?(ε1 ∧ ε2 ) = ε3 ∧ ε̄1 ∧ ε̄2 ∧ ε̄3 , ?(ε2 ∧ ε̄1 ) = −ε1 ∧ ε3 ∧ ε̄2 ∧ ε̄3 , and so on.

451
18.6. Cohomology and Chern Classes Chapter 18. Complex Varieties

For nontrivial E, this local definition extends to the whole bundle using the Hermitian metric
on E.
Now by the theory of harmonic functions (see Section 18.3 and the proof of Hodge
decomposition), we may identify H p,q (M, E) with the space Hp,q (M, E) of harmonic functions
with values in E. Explicitly, α ∈ Hp,q (M, E) if and only if ∆E ∂¯
α = 0, where

∆E ¯ ¯∗ ¯∗ ¯ • •
∂¯ = ∂ ∂ + ∂ ∂ : Ω (M, E) −→ Ω (M, E)

as in the proof of Hodge decomposition. Let ∆∂¯ be the Laplace operator on Ωbullet (M ).

Lemma 18.6.16. For any vector bundle E → M , ∆E


∂¯
? = ?∆∂¯.

Therefore ? descends to an operator ? : Hp,q (M, E) → Hn−p,n−q (M, E ∗ )∗ .

Lemma 18.6.17. For all p, q ≥ 0, ? : Hp,q (M, E) → Hp,q (M, E ∗ ) is an isomorphism.

This gives us a string of isomorphisms

H p (M, E) ∼
= H 0,p (M, E) by Dolbeault’s theorem

= H0,p (M, E) by the proof of Hodge decomposition

= Hn,n−p (M, E ∗ )∗ by the Lemma
 ^ ∗

=H n,n−p ∗ ∗ ∼
(M, E ) = H n−p
M, n ∗
T M ⊗E ∗
.

It remains to interpret n T ∗ M ⊗ E ∗ . Note that the bundle n T ∗ M is a line bundle, so it is


V V
isomorphic to O(KM ) for some divisor KM . This completes the sketch of Serre duality.

Definition.
Vn ∗ ∼ For a compact complex manifold M of dimension n, any divisor KM such that
T M = O(KM ) is called a canonical divisor on M .

Example 18.6.18. When Σ is a Riemann surface and L → Σ is a line bundle, Serre duality
says that
H 1 (Σ, L) ∼
= H 0 (Σ, L∗ ⊗ O(KΣ )).

Corollary 18.6.19 (Riemann-Roch Theorem, Second Version). Suppose Σ is a Riemann


surface of genus g and L is a line bundle on Σ. Then

h0 (L) − h0 (L∗ ⊗ O(KΣ )) = deg(L) − g + 1.

Alternatively, if D is a divisor on Σ then

h0 (D) − h0 (KΣ − D) = deg(D) − g + 1.

Corollary 18.6.20. If KΣ is a canonical divisor on a Riemann surface Σ of genus g, then


deg(KΣ ) = 2g − 2.

Proof. Apply Riemann-Roch with D = KΣ .

452
18.6. Cohomology and Chern Classes Chapter 18. Complex Varieties

Note that deg(KΣ ) = 2g − 2 is precisely −χ(Σ) where χ denotes the topological Euler
characteristic. To see this directly, let e(Σ) = e(T Σ) be the Euler class of Σ and observe
that O(KΣ ) = T ∗ Σ, so we have

−χ(Σ) = −he(T Σ), [Σ]i = h−c1 (T Σ), [Σ]i = hc1 (T ∗ Σ), [Σ]i = hc1 (KΣ ), [Σ]i = deg(KΣ ).

Corollary 18.6.21. If D is a divisor on a Riemann surface Σ of genus g and deg(D) >


2g−2, then h0 (D) = deg(D)−g+1. Likewise, if L is a line bundle on Σ with deg(L) > 2g−2
then h0 (Σ, L) = deg(L) − g + 1.
Proof. If deg(D) > 2g − 2 then deg(KΣ − D) = 2g − 2 − deg(D) < 0 so Proposition 18.6.7
shows that h0 (KΣ − D) = 0.
This shows that if L is a line bundle on a Riemann surface of genus g (the same statement
will hold for divisors) such that deg(L) < 0 or deg(L) > 2g − 2, then h0 (L) is topologically
determined. However for line bundles with 0 ≤ deg(L) ≤ 2g − 2, this is not the case in
general.
What are holomorphic sections good for?, you might ask. Well, suppose L is a holomor-
phic line bundle on a complex manifold M and dim H 0 (M, L) = N + 1 for some N ≥ 0. Let
{s0 , . . . , sN } be a basis for H 0 (M, L). Then there is a point x ∈ M for which not all sj (x)
are zero. In a local trivialization of L near x, the tuple (s0 (x), . . . , sN (x)) defines a point in
CN +1 r {0} and on a different trivialization, this tuple is well-defined up to gαβ (x) where
gαβ is a transition function defining L. In other words, the point [s0 (x), . . . , sN (x)] ∈ PN is
well-defined.
Definition. A complete linear system for a line bundle L → M is the collection |L| of
effective divisors corresponding to H 0 (M, L), i.e. if L ∼
= O(D) for a divisor D ∈ Div(M ),
then
|L| = {D0 ∈ Div(M ) | D0 ∼ D} ∼ = P(H 0 (M, L)) ∼
= P(L(D)).
A linear system for L is a collection |W | of effective divisors corresponding to a linear sub-
space W ⊆ H 0 (M, L). The dimension of a linear system is the dimension of the projective
subspace of P(L(D)) to which it corresponds.
In particular, note that dim |W | = dim W − 1. We call a linear system of dimension 1,
i.e. when h0 (L) = 2, a pencil (of divisors) for L. The following discussion closely parallels
the discussion of the canonical embedding in Section 17.3.
Definition. The base locus of a linear system |W | for L → M is the analytic subvariety

B|W | = {x ∈ M | s(x) = 0 for all s ∈ W } ⊆ M.

If B|W | = ∅, we say |W | is basepoint free.


The important point here is that if B|L| = ∅, then choosing a basis for H 0 (M, L) produces
a well-defined holomorphic map ϕ|L| : M → PN . Moreover, a different choice of basis changes
ϕ|L| by a projective transformation.
Let’s consider the case of a Riemann surface Σ of genus g. Let L ∈ Pic(Σ) be a line
bundle. Then a natural question to ask is: when does ϕ|L| give an embedding Σ ,→ PN ?

453
18.6. Cohomology and Chern Classes Chapter 18. Complex Varieties

First note that B|L| = ∅ if and only if for all P ∈ Σ, there is a section s ∈ H 0 (Σ, L) such
that s(P ) 6= 0. Equivalently, we are asking if the natural map H 0 (Σ, L) → H 0 (P, L|P ) is
surjective for all P . Consider the short exact sequence of sheaves

0 → IP → L → L|P → 0.

Here, IP is the sheaf of sections of L which vanish at P . Explicitly, IP = L ⊗ O(−P ) and


the long exact sequence in cohomology from this reads:

H 0 (Σ, L) → H 0 (P, L|P ) → H 1 (Σ, L ⊗ O(−P )).

By Serre duality, H 1 (Σ, L ⊗ O(−P )) ∼= H 0 (Σ, L∗ ⊗ O(KΣ − P ))∗ , so surjectivity of the


restriction map is equivalent to having H 0 (Σ, L∗ ⊗ O(KΣ − P )) = 0. If deg(L) > 2g − 1,
then
deg(L∗ ⊗ O(KΣ − P )) = − deg(L) + (2g − 2) − 1 < 0
and this cohomology group will vanish by Proposition 18.6.7. Therefore if deg(L) > 2g − 1,
|L| is basepoint free and ϕ|L| defines a holomorphic map Σ → PN . The next question is
when this map is an embedding. This happens precisely when, for all P 6= Q in Σ, the
vectors (s0 (P ), . . . , sN (P )) and (s0 (Q), . . . , sN (Q)) are linearly independent in CN +1 . Let
AP := H 0 (Σ, IP ) ⊂ H 0 (Σ, L) be the subspace of sections of L vanishing at P . Note that
AP has codimension 1 in H 0 (Σ, L). Then P 6= Q map to distinct points in PN if and only if
there is a section s ∈ AP such that s(Q) 6= 0. Consider the sequence of sheaves

0 → IP,Q → L → L|{P,Q} → 0

where IP,Q now denotes the sheaf of sections of L vanishing at P and Q simultaneously.
Also note that H 0 ({P, Q}, L|{P,Q} ) ∼= H 0 (P, L|P ) ⊕ H 0 (Q, L|Q ) = L(P ) ⊕ L(Q). Then ϕ|L|
is one-to-one if and only if H 0 (Σ, L) → L(P ) ⊕ L(Q) is onto and to show this condition, it
suffices to prove H 1 (Σ, IP,Q ) = 0. By Serre duality,

H 1 (Σ, IP,Q ) = H 1 (Σ, L ⊗ O(−(P + Q))) ∼


= H 0 (Σ, L∗ ⊗ O(KΣ + P + Q)).

Thus we see that if deg(L) > 2g, deg(KΣ + P + Q) − deg(L) = 2g − deg(L) < 0 which implies
H 1 (Σ, IP,Q ) = 0.
Finally, we might ask when ϕ|L| is an immersion, i.e. when dP ϕ|L| : TP Σ → Tϕ|L| (P ) PN is
an injective linear map for all P ∈ Σ. Dualizing, this is the same as dP ϕ∗|L| : Tϕ∗|L| (P ) PN →
TP∗ Σ being surjective. For a given P ∈ Σ, we may choose a basis {s0 , . . . , sN } for H 0 (Σ, L)
with s0 (P ) 6= 0 and s1 (P ) = · · · = sN (P ) = 0. Then in a neighborhood of P , ϕ|L| is given
by
sj
ϕ|L| (Q) = (t1 (Q), . . . , tN (Q)) ∈ CN ⊂ PN where tj = .
s0
Then the differential is given by (dtj ) so the adjoint is the row vector dP ϕ∗|L| = dt1 · · · dtN .


Thus dP ϕ∗|L| is surjective precisely when for any α ∈ TP∗ Σ, there exist a section s ∈ H 0 (Σ, IP )
such that ds = α. Consider the exact sequence of sheaves
d
0 → L ⊗ O(−2P ) → − (T ∗ Σ ⊗ L)|P → 0
− L→

454
18.6. Cohomology and Chern Classes Chapter 18. Complex Varieties

where d sends a section s to ds. In the long exact sequence, we have

· · · → H 0 (Σ, L) → H 0 (P, (T ∗ Σ ⊗ L)|P ) → H 1 (Σ, L ⊗ O(−2P )) → · · ·

so we are interested in when H 1 (Σ, L ⊗ O(−2P )) = 0. Again by Serre’s duality and a


dimension argument, this vanishing occurs when deg(L) > 2g. Therefore we have proven:

Theorem 18.6.22. Every compact Riemann surface Σ admits a holomorphic embedding


Σ ,→ PN for some N ≥ 1.

Proof. Let g be the genus of Σ and take L → Σ to be a line bundle with degree deg(L) > 2g.
Then the above argument shows that the complete linear system |L| is basepoint free and
hence ϕ|L| : Σ → PN is defined, where h0 (L) = N + 1. Moreover, since deg(L) > 2g, the
above also shwos ϕ|L| is an embedding.

Remark. It is a deep theorem that every analytic submanifold of PN is algebraic, so as a


consequence we see that every Riemann surface may be realized as a complex subvariety of
some projective space.

We have seen that for a line bundle L on a Riemann surface Σ, the conditions that ϕ|L|
is well-defined and an embedding are related to the vanishing of H 1 (Σ, −) for various line
bundles built from L. Since dim Σ = 1, there is a nice topological criterion for the vanishing of
such cohomology groups: the degree of those line bundles need only be sufficiently positive
(this is essentially because irreducible divisors are merely points). This nice situation is
special to the dimension 1 case.
If M is a complex manifold of dimension greater than 1, there still exists a criterion for
1
H (M, L) to vanish for a particular line bundle L → M , but it is geometric and not merely
topological in nature. To introduce this criterion, called positivity of a line bundle, we first
discuss the theory of connections on topological vector bundles.

455
Part IV

Algebraic Groups

456
Chapter 19

Introduction

The notes in Part IV come from two courses on Lie groups taught by Dr. Brian Parshall
at the University of Virginia in the fall of 2015 (Lie groups and algebraic groups) and the
spring of 2018 (finite groups of Lie type). The topics covered are:

ˆ Basic algebraic geometry

ˆ Algebraic groups and their properties

ˆ The classical theory of Lie algebras and Lie groups

ˆ Complete varieties, parabolic subgroups and Borel subgroups

ˆ Diagonalizable, connected solvable and semisimple groups

ˆ Finite groups of Lie type and their representation theory.

The primary texts for this course are Procesi’s Lie Groups: An Approach through Invariants
and Representations, T.A. Springer’s Linear Algebraic Groups, Malle-Testerman’s Linear Al-
gebraic Groups and Finite Groups of Lie Type, Geck’s An Introduction to Algebraic Geometry
and Algebraic Groups and Jacobson’s Basic Algebra I.
Given a group G, we have many ways of assigning further structure to study its connection
to different areas of mathematics:

(1) As a set – this is abstract group theory;

(2) As a topological space – this gives rise to topological groups;

(3) As a smooth manifold – these are Lie groups;

(4) As an algebraic variety – these are algebraic groups, which are covered in Chapters 20
and 21;

(5) As a scheme – this fits in with modern algebraic geometry;

(6) As a quantum group – this has nice applications in representation theory.

457
19.1. Chevalley’s Theorem Chapter 19. Introduction

19.1 Chevalley’s Theorem


The following theorem will be useful.

Theorem 19.1.1 (Chevalley). Let ϕ : X → Y be a morphism of varieties. Then ϕ(X)


contains a nonempty open subset of ϕ(X).

To prove the theorem, we first need the following lemma, which we won’t prove.

Lemma 19.1.2. Let A ⊆ B be finitely generated integral domains over an algebraically


closed field k. Then for every b ∈ B, b 6= 0, there exists an a ∈ A, a 6= 0 such that any
homomorphism ϕ ∈ Homk (A, k) for which ϕ(a) 6= 0 extends to a k-algebra homomorphism
Φ : B → k with Φ(b) 6= 0.

Now we prove Chevalley’s theorem:


Proof. We will reduce to the case where X and Y are irreducible and affine, and ϕ(X) = Y .
Let A = k[Y ] and B = k[X], so that ϕ induces an algebra homomorphism:
ϕ
X−
→Y
ϕ∗
B ←− A.

Since ϕ(X) = Y , ϕ∗ is an injection. To apply Lemma 19.1.2, let b = 1. Then there is a


nonzero a ∈ A satisfying Lemma 19.1.2. We claim that D(a) is an open subset of Y which
is contained in ϕ(X). First, every principal open set is open of course. Take f ∈ D(a), so
f (a) 6= 0. By Lemma 19.1.2, there exists a k-algebra homomorphism F : B → k such that
F (1) 6= 0. Note that the following diagram commutes:
ϕ∗
B A

f
F
k

Therefore ϕ(F ) = f , so D(a) ⊂ ϕ(X) as claimed.

Definition. A morphism ϕ : X → Y between irreducible algebraic varieties is dominant if


ϕ(X) = Y , that is, if the image of ϕ is dense in the Zariski topology on Y .

Chevalley’s Theorem says that the image of any dominant morphism of varieties contains
an open subset of the codomain.

458
Chapter 20

Algebraic Groups

Definition. An algebraic group is a group G which is a variety over an algebraically


closed field k, whose group operations are variety morphisms: G × G → G, (g, h) 7→ gh and
(·)−1 : G → G, g 7→ g −1 .
Definition. G is an affine algebraic group if G is affine as a variety. Such a group is
also sometimes called a linear algebraic group.
Examples.
1 The base field k is an affine algebraic group under addition. This is often denoted Ga .

2 The nonzero elements of the field, k × , form an affine algebraic group under multiplica-
tion, denoted Gm . In this case, the coordinate ring is k[G] = k[t, t−1 ], a one-dimensional
torus. Note that dim Ga = dim Gm = 1.
3 Given affine groups G and H, G×H is an affine group so products exist in the category
AlgGroups. We could take Ga × · · · × Ga = Gna which is an affine group constructed
| {z }
n
from the underlying space An . Similarly, Gm × · · · × Gm = Gnm is an affine group called
| {z }
n
the n-dimensional torus.
4 The general linear group GLn (k) is an affine group. Here
A = k[GLn (k)] = k[Tij | 1 ≤ i, j ≤ n][T ]/(1 − T D)
where D is the determinant function and (1 − T D) is the ideal generated by 1 − T D.

5 The special linear group SLn (k) is an affine group with coordinate algebra A =
k[Tij ]/(D − 1). We can continue defining algebraic groups for the orthogonal linear
group, projective general/special linear groups, etc.
Later we will prove:
Theorem 20.0.1. Every irreducible one-dimensional affine algebraic group G is isomorphic
to either Ga or Gm .

459
20.1. Hopf Algebras Chapter 20. Algebraic Groups

20.1 Hopf Algebras


Let G be an affine algebraic group and set A = k[G], so that G = Homk (A, k). Then G × G
is affine with coordinate algebra k[G × G] = A ⊗k A. The comorphism of multiplication
G × G → G is the algebra
Pn homomorphism ∆ : A → A ⊗k A, called comultiplication.
Pn Take
f ∈ A; then ∆(f ) = i=1 fi ⊗ fi for fi , fi ∈ A. For all x, y ∈ G, f (xy) = i=1 fi (x)fi0 (y).
0 0

Since multiplication commutes, we have a commutative diagram

m × 1G
G×G×G G×G

1G × m m

G×G m G

On the algebra side, there is a corresponding commutative diagram:

∆⊗1
A ⊗k A ⊗k A A ⊗k A

1⊗∆ ∆

A ⊗k A A

In other words (1 ⊗ ∆)∆ = (∆ ⊗ 1)∆. This is called the law of coassociativity.


Similarly, the inverse morphism G → G, g 7→ g −1 induces a comorphism γ : A → A. The
rule is given by γ(f )(x) = f (x−1 ) for all f ∈ A. The comorphism γ is called an antipode.
Notice that m(1 ⊗ γ)∆ = m(γ ⊗ ∆)∆ = ε where ε(f ) = f (e) for all f ∈ A (e ∈ G is the
identity element). The element ε is called a counit and it satisfies the counit law:

(ε ⊗ 1)∆ = (1 ⊗ ε)∆ = idA .

This may also be viewed as the comorphism ε : A → k of the injection {e} ,→ G.


Any algebra satisfying coassociativity and having antipodes and counits may be endowed
with the structure of a coordinate ring for some algebraic group.

Definition. A Hopf algebra over k is a 4-tuple (A, ∆, γ, ε) consisting of a k-algebra A


together with algebra homomorphisms ∆ : A → A ⊗k A, γ : A → A and ε : A → k satisfying
the laws of coassociativity, antipodes and counits.

Examples.

1 Recall that Ga is the additive group k as an algebraic group. Its coordinate algebra is

460
20.1. Hopf Algebras Chapter 20. Algebraic Groups

A = k[T ], which is a Hopf algebra whose morphisms are

∆ : A −→ A ⊗k A
T 7−→ T ⊗ 1 + 1 ⊗ T ;
γ : A −→ A
T 7−→ −T ;
and ε : A −→ k
T 7−→ 0.

2 Similarly, Gm is an algebraic group whose coordinate ring is A = k[T, T −1 ]. A can be


made into a Hopf algebra by defining the morphisms

∆ : A −→ A ⊗k A
T 7−→ T ⊗ T
T 7−→ T −1 ⊗ T −1 ;
−1

γ : A −→ A
T 7−→ T −1 ;
and ε : A −→ k
T 7−→ 1.

3 For GLn (k), the Hopf algebra structure consists of ∆ : A → A ⊗k A which maps
Tij 7→ n`=1 Ti` ⊗ T`j and T 7→ D−1 ⊗ D−1 ; and γ : Tij 7→ D−1 adjTji by Cramer’s Rule,
P

where adjT denotes the transpose of the adjoint.

461
20.2. Topological Properties Chapter 20. Algebraic Groups

20.2 Topological Properties


Recall that Chevalley’s Theorem (19.1.1) says for a variety morphism f : X → Y , f (X)
contains a nonempty, open subset of f (X).

Proposition 20.2.1. Let G be an affine algebraic group.

(1) G has a unique irreducible component which is closed and contains e, the identity.
This component is denoted G0 .

(2) G0 is a normal subgroup of finite index in G and the other irreducible components of
G are the cosets aG0 .

(3) Given a homomorphism ϕ : G → H of affine algebraic groups, ϕ(G0 ) ⊆ H 0 .

Proof. (1) By Proposition 12.1.13, G is a union of a finite number of closed, irreducible


components X1 , . . . , Xn . Each Xi has the induced topology of G on it. Since G = X1 ∪ · · · ∪
Xn , consider only the irreducible components Y1 , . . . , Ym which contain the identity. Then
Y1 × · · · × Ym is an affine algebraic variety which is irreducible. We can consider the diagram
Y1 × · · · × Ym Y1 · · · Ym

G×m G

The map along the diagonal is continuous, so it takes irreducible components to irreducible
components. Thus Y1 · · · Ym is irreducible in G and so is its closure Y1 · · · Ym . However
Yi ⊆ Y1 · · · Ym for each Yi , but Yi is irreducible. This forces m = 1, so there is a single
irreducible component containing the identity.
(2) For G0 ≤ G, note that the multiplication and inverse maps are homeomorphisms, so
they take a 7→ e and G0 → G0 ; therefore G0 is closed under these operations. Since inner
automorphisms are also homeomorphisms, we can see that for any x ∈ G, xG0 x−1 = G0
so G0 is a normal subgroup. Moreover, since the components of G are mutually disjoint, it
follows that these must be the cosets xG0 .
(3) follows from the uniqueness part of (1), and the fact that morphisms map irreducible
components into irreducible components.
Examples.

1 GLn (k) is connected (irreducible) because GLn (k) = GLn (k)0 . This can be seen from
the fact that the ideal (1 − T D) is irreducible in k[Tij , T ], and since the coordinate al-
gebra for GLn (k) is A = k[Tij , T ]/(1 − T D), irreducibility follows. The most important
subgroup, SLn (k), is also irreducible.

2 Recall the orthogonal group On (k) = {g ∈ GLn (k) | g T g = 1}. An important subgroup
is the special orthogonal group SOn (k) = {g ∈ On (k) | det g = 1}. It turns out that
SOn (k) = On (k)0 so the special orthogonal group is a connected component of On (k).

462
20.2. Topological Properties Chapter 20. Algebraic Groups

3 Set n = 2m, m ≥ 1 and define


 
0 Im
J= .
−Im 0

From this we define the nth symplectic group Spn (k) = {g ∈ GLn (k) | g T Jg = J}.

The algebraic groups in 1 through 3 comprise what are known as the classical groups.

Let G be an affine group and take x ∈ G. The left multiplication map g 7→ xg is in fact
a homeomorphism of topological spaces G → G. In general, how might we show that a map
between algebraic groups is continuous?

Definition. The map ϕ is continuous if ϕ−1 (V(J)) is closed in X for every ideal J ⊂ k[Y ].

Notice that in the case of left multiplication ϕ(g) = xg, we have a commutative diagram

ϕ∗
k[G] k[G]

g ϕ(g)

where g and ϕ(g) are viewed as maps k[G] → k. Then for an ideal J ⊂ k[G], ϕ(g) ∈ V(J) if
and only if g vanishes on the ideal of k[G] generated by ϕ∗ (J). Hence ϕ−1 (J) is closed in G
and we see that left multiplication is continuous. Similar proofs show the continuity of right
multiplication, conjugation, etc.

Lemma 20.2.2. If U and V are dense, open subsets of an affine group G then G = U V .

Proof. Let x ∈ G and consider U −1 x. This is still open and dense, so (U −1 x) ∩ V 6= ∅.


Thus there is some u ∈ U and v ∈ V such that u−1 x = v, that is, x = uv ∈ U V . Hence
G = UV .

Lemma 20.2.3. Let H be a subgroup of G, an affine algebraic group.

(a) H is also a subgroup of G. In particular, H is an affine algebraic group.

(b) If H contains a nonempty open subset of H then H = H.

Proof. (a) Let x ∈ H. Then H = xH ⊆ xH. This shows that H ⊆ xH, or x−1 H ⊆ H.
Therefore HH ⊆ H. So if h̄ ∈ H then H h̄ ⊆ H. Now right multiplication is continuous so
it maps the closure of H h̄ into H, meaning H h̄ ⊆ H. Since h̄ ∈ H was arbitrary, HH ⊆ H.
Hence H is closed under multiplication. Inverses are proven similarly. Finally, one can check
that multiplication and inverses on H are the restrictions of their counterparts on G, thus
they are variety morphisms. Hence H ≤ G and H is an affine algebraic group.

463
20.2. Topological Properties Chapter 20. Algebraic Groups

(b) Now suppose H contains a nonempty open subset Ω ⊆ H. S Given x ∈ H, xΩ is also


open in H. We may assume Ω contains the identity. Then H = x∈H xΩ and the arbitrary
S
union of open sets is open, so H is open in H. On the other hand, H r H = g∈HrH gH
is also open for similar reasons. So its complement, H, is closed in H. Finally, H open and
closed in H implies that H = H.
Theorem 20.2.4. Let ϕ : G → G0 be a morphism of affine algebraic groups. Then
(1) ker ϕ is a closed subgroup of G.

(2) im ϕ is a closed subgroup of G0 .

(3) ϕ(G0 ) = ϕ(G)0 .


Proof. (1) The Zariski topology is T1, so point-sets are closed. Since ϕ is continuous, ker ϕ =
ϕ−1 (eG0 ) is closed.
(2) By Chevalley’s Theorem (19.1.1), ϕ(G) contains a nonempty, open subset of ϕ(G).
Then by Lemma 20.2.3(a), ϕ(G) = ϕ(G) so im ϕ = ϕ(G) is closed.
(3) follows from Proposition 20.2.1.
The property that every kernel and image is closed implies that the kernels and images
of algebraic group homomorphisms are themselves algebraic groups. In other words, the
category AffGroupsk has kernels and images.
Examples.
4 Let G = GLn (k) and for 1 ≤ i < j ≤ n, set Xij = A1 and ϕij (λ) as the matrix with 1’s
along the diagonal, λ in the ijth position and 0’s elsewhere. Also set Xii = A1 r {0}
and ϕii (λ) as the matrix with λ in the ith position of the diagonal, 1’s along the rest
of the diagonal and 0’s elsewhere. One can see that H = GLn (k) in this case.

5 Let G = Spn (k), the symplectic group. The matrix J gives us a k-vector space V of
dimension n = 2m. Given a nondegenerate, alternating bilinear form b on V ,

b : V × V −→ k,

for every nonzero vector v ∈ V , there exists a u ∈ V such that b(u, v) 6= 0 and
b(u, u) = 0. Whenever char k 6= 2, b(u, v) = −b(v, u) for all u, v ∈ V . Thus

Spn (k) = {g : V → V | g is invertible and b(gu, gv) = b(u, v) for all u, v ∈ V }.

For each u ∈ V and c ∈ k, define τu,c ∈ Spn (k) by τu,c = x − cb(x, u)u; these are called
symplectic transvection. For each u, this corresponds to a morphism

A1 −→ Spn (k), c 7−→ τu,c .

Theorem 20.2.5. Spn (k) is generated by the symplectic transformations {τu,c | u ∈


V, c ∈ k}.
Corollary 20.2.6. Spn (k) = Spn (k)0 so the symplectic groups are connected.

464
20.2. Topological Properties Chapter 20. Algebraic Groups

Theorem 20.2.7. Let G be an affine algebraic group and suppose {Xi , ϕi }i∈I is a collection
of irreducible varieties Xi and morphisms ϕi : Xi → G such that e ∈ Yi := ϕi (Xi ) for every
i. Let H be the closed subgroup generated by the Yi , i.e. H = hYi ii∈I . Then there exist
elements a(1), a(2), . . . , a(m) ∈ I and a choice ε(h) = ±1 for each h = 1, 2, . . . , m such that
ε(1) ε(2) ε(m)
H = Ya(1) Ya(2) · · · Ya(m) =: Yaε .

Proof. First, for each i define ϕ−1


i : Xi → G by ϕ−1 i (x) = ϕi (x)
−1
for all x ∈ Xi . Enlarging
−1 −1
the index set I if necessary, we may assume ϕi (Xi ) = Yi for all i. Then we can drop the
signs ε in Yaε and focus on Ya .
Observe that Y a is closed and irreducible for any sequence a = (a(1), . . . , a(m)). Then
Y a is a subvariety of G containing the identity e. If Y a ( Y b for some other sequence b, then
dim Y a < dim Y b ≤ dim G < ∞. So among the choices of sequences a, select one so that Y a
has maximum dimension. For any other sequence b = (b(1), . . . , b(m)), Y a ⊆ Ya Yb =: Y (a,b)
but by maximality of dim Y a , we must have Y a = Y (a,b) . But e ∈ Ya so Yb ⊆ Y (a,b) = Y a
and thus Y b ⊆ Y a . Since H is generated by these Yb , H ⊆ Y a but of course Y a ⊆ H as well,
so we have equality: Y a = H.
Notice that Ya is the image of the morphism

τ : Xa(1) × · · · × Xa(m) −→ G
(xa(1) , . . . , xa(m) ) 7−→ ϕa(1) (xa(1) ) · · · ϕa(m) (xa(m) ).

Then by Chevalley’s Theorem, Ya contains a nonempty, open subset U of Y a = H. Therefore


by Lemma 20.2.2, H = U U ⊆ Ya Ya = Y(a,a) and we have proven the theorem.
Examples.

4 (continued) Let G = GLn (k) again. Recall the maps ϕij (λ) for i, j = 1, . . . , n that are
homomorphisms A1 → G. Notice that when i < j, ϕji (λ) = ϕij (λ)T . The ϕij (λ) gen-
erate GLn (k) and from this we can deduce that H = GLn (k) so GLn (k) is irreducible
(connected).

Theorem 20.2.7 has some useful corollaries which we mention here.

Corollary 20.2.8. Let {Gi }i∈I be a family of closed, connected subgroups of an affine group
G. Then the subgroup generated by the Gi , hGi ii∈I is closed and connected.

Corollary 20.2.9. Let H, K ≤ G with H closed and connected. Then the commutator
subgroup [H, K] := hhkh−1 k −1 | h ∈ H, k ∈ Ki is closed and connected.

Proof. Let I = K and consider the collection {Hx , ϕx }x∈K where Hx = xHx−1 and ϕx :
Hx → G, xhx−1 7→ hxh−1 x−1 . It is clear that the images im ϕx over all x ∈ K generate
[H, K], so apply Theorem 20.2.7 to deduce that [H, K] is closed and irreducible.

465
20.3. G-spaces Chapter 20. Algebraic Groups

20.3 G-spaces
Let G be an affine algebraic group over k.

Definition. A variety X over k is called a G-space if there is a morphism G × X →


X, (g, x) 7→ g · x, such that for all g1 , g2 ∈ G and x ∈ X, the following properties hold:

(1) (g1 g2 ) · x = g1 · (g2 · x).

(2) e · x = x.

This is the algebraic group version of G-actions.

Examples.

1 Take X = G. Then G acts on itself in several ways: (g, x) 7→ gx (the left action);
(g, x) 7→ xg −1 (the right action); (g, x) 7→ gxg −1 (conjugation). These all make G into
a G-space.

2 Let V be a finite dimensional vector space over k and suppose G acts on V by linear
transformations, (g, v) 7→ g · v, so that the G-space axioms hold. Then there exists a
homomorphism
ϕ : G −→ GL(V ) ∼ = GLn (k).
(Here the isomorphism depends on choice of basis.) We say the mapping G → GLn (k)
is rational (formally, a rational representation) if it is a morphism of algebraic groups.
In this case, V is said to be a rational G-module and this structure makes V into a
G-space. However, not every G-space is a rational module since the G-action might
not be linear.
To expand this notion a bit, suppose V is a (possibly infinite dimensional) k[G]-module.
We say V is a rational G-module if it is the union of finite dimensional rational G-
submodules as defined above. Equivalently, V is rational if every v ∈ V lies in a finite
dimensional rational submodule.
Every G-module V is a comodule over the coordinate ring A = k[G]: there is a linear
map
∆V : V −→ V ⊗k A
such that (1 ⊗ ∆)∆V = (∆V ⊗ 1)∆V and (1 ⊗ ε)∆V = idV for ∆ the comultiplication of
A and ε the counit. For example, A = k[G] is a rational G-module from this viewpoint.
Pn∆A = ∆, 0we see that coassociativity holds, and for f ∈ A, g ∈ G such that
By setting
∆(f ) = i=1 (fi ⊗ fi ), we set
n
X
g·f = fi fi0 (g).
i=1

This action makes A into a rational G-module.

466
20.3. G-spaces Chapter 20. Algebraic Groups

3 Take a finite dimensional vector space V . A flag F in V is a chain of subspaces

F : 0 = V0 ⊆ V1 ⊆ · · · ⊆ Vm = V.

If in addition dim Vi = i for each i = 0, 1, . . . , m then we say F is a complete flag in


V . The collection Flag(V ) = {F | F is a complete flag in V } is an algebraic variety,
but it is usually not affine. What’s more, given any variety Y over k, the projection
map Flag(V ) × Y → Y is a closed map (varieties satisfying this property are called
complete varieties).
Observe that if g ∈ G and F : V0 ⊆ V1 ⊆ · · · ⊆ Vm is a flag in V , then

g · F : gV0 ⊆ gV1 ⊆ · · · ⊆ gVm

is another flag in V . This action makes Flag(V ) into a G-space. An important example
of this object is when G = GLn (k). Let V be the n-dimensional vector space over k
with standard basis {e1 , . . . , en } and set Vi = Span{e1 , . . . , ei } for each i = 1, . . . , n.
Then F : 0 = V0 ⊆ V1 ⊆ · · · ⊆ Vn = V is a complete flag in V and StabG (F) = B, the
subspace of upper triangular matrices in GLn (k). Hence we can view the flag variety
as a quotient: Flag(V ) ∼= G/B.
The following is perhaps the most important theorem in the study of G-spaces. We will
prove this in Section 12.11.

Theorem 20.3.1 (Borel, 1955). Suppose X is a complete variety and G is a connected,


solvable algebraic group that acts on X. Then there exists a fixed point x0 ∈ X, i.e. g·x0 = x0
for all g ∈ G.

Definition. An orbit of a G-space X is a subset O ⊂ X of the form O = {g · x0 | g ∈ G}


for a fixed x0 ∈ X.

Theorem 20.3.2. Suppose X is a G-space and take an orbit O ⊂ X. Then O is open in


O in the Zariski topology and every irreducible component of O r O has dimension strictly
less than dim O.

Proof. Notice that O is the image of a morphism G → X, namely g 7→ g·x0 , so by Chevalley’s


Theorem, O contains a nonempty open subset U ⊂ O. If g ∈ G then g · U is also open in O
and gS· U ⊂ O. Letting g range over all elements of G, we get an open cover of O; explicitly,
O = g∈G Ug which is a union of open sets. Therefore O is open in O.
One can define a partial order ≤ on the collection of orbits of the G-space X by O1 ≤ O2
if O1 ⊆ O2 .

Examples.

4 Let G = GLn (k) and X = Mn (k) ∼


2
= An . Then G acts on X by conjugation. Orbits
are described by Jordan canonical forms. Now let

Xn = {x ∈ Mn (k) | x is nilpotent} = {x ∈ Mn (k) | xn = 0}.

467
20.3. G-spaces Chapter 20. Algebraic Groups

Then G also acts on Xn by conjugation, but what are the orbits this time? An n × n
matrix x is called unipotent if 1−x is nilpotent. Let Yn = {x ∈ Mn (k) | x is unipotent}.
There is an obvious bijection Xn ↔ Yn . Observe that Xn is closed in Mn (k) since the
points are given by a polynomial expression on their components. So Xn is an affine
algebraic variety and consequently so is Yn . Now x ∈ Xn if and only if x is similar to
an upper diagonal matrix with 0’s along the diagonal.
Next define   
1


 1
 * 




Un =   ∈ Mn (k) ≤ GLn (k).
 
...




0  


 1 

Theorem 20.3.3. Let H be a closed, irreducible subgroup of GLn (k) consisting entirely
of unipotent matrices. Then there exists a g ∈ GLn (k) such that gHg −1 ≤ Un .

Proof. Let’s induct on n. If n = 1, the only 1×1 unipotent matrix is [1] so the base case
is trivial. Suppose there is some n ∈ N where for all m < n, there exists a g ∈ GLm (k)
such that gHg −1 ≤ Um . View GLn (k) as GL(V ) where V is the vector space of n × 1
column vectors. We may assume the action of H on V is irreducible; otherwise there
is a subspace 0 6= W ( V such that W is stabilized by H, and thus for some g ∈ G,
" #
gHg −1 =
W *
0 V /W

In this case dim W and dim V /W are both strictly less than n so we’re done by induc-
tion.
Now, assuming V is an irreducible H-space, consider the subalgebra E of Endk (V )
spanned by the elements of H. Then by Wedderburn-Artin theory, E = Endk (V ). For
g, h ∈ H, tr(g) = tr(h) = n = dim V . Hence tr((1 − g)h) = tr(h) − tr(gh) = n − n = 0.
Fixing g ∈ H and letting h vary, we can write all elements of E as linear combinations
of (1 − g)h. Then tr((1 − g)r) = 0 for all r ∈ Endk (V ). This means 1 − g = 0, so
H = {e}.

Returning to the GLn (k)-space Xn of all nilpotent n × n matrices, we know Xn is a


variety and in fact its dimension is
n(n − 1)
dim Xn = n2 − .
2
Every element in Xn can be put in Jordan canonical form. For a positive integer a,
define the standard a × a Jordan block
 
0 1 0
0 0 1
 0 
0 0 0
Ja = 


 . . 

0 . 1
0 0

468
20.3. G-spaces Chapter 20. Algebraic Groups

Then for every x ∈ Xn , there is some g ∈ GLn (k) such that


 
Ja1
 Ja2
0 
−1
gxg = 
 
. . .


0 
Jat
Pt
where a1 > a2 > · · · > at ≥ 0 and r=1 ar = n. Such a choice of a1 , . . . , at corresponds
to a partition of N , denoted λ = (a1 , a2 , . . . , at ) ` n. Jordan canonical forms are
unique orbit representations, so the orbits O of Xn are in one-to-one correspondence
with partitions λ ` n. Let Oλ denote the orbit corresponding to λ.
Now we’d like to describe the closures of orbits, Oλ (this wasn’t solved until the 1960s).
Let Λ(n) be the set of partitions of n. Define the dominance ordering on Λ(n) in
the followingPway: for λP= (λ1 , . . . , λr ) and µ = (µ1 , . . . , µs ), write λ E µ if for all
i = 1, . . . , r, ik=1 λk ≤ ik=1 µk . Then we have the following characterization:
Theorem 20.3.4. Oλ ⊆ Oµ if and only if λ E µ.

5 For an algebraic group G with coordinate algebra A = k[G], we can regard A as a


G-space in several ways:
ˆ For g ∈ PG and f ∈ A, define ρ(g)f ∈ A Pg · f = (ρ(g)f )(x) = f (xg). Then if
∆(f ) = ni=1 fi ⊗ fi0 , we have ρ(g)f = ni=1 fi0 (g)fi ∈ A and ρ(e) = idA . This
action corresponds to the left regular representation of G.
ˆ We could similarly define the right action of G on A by f · g = (λ(g)f )(x) =
f (g −1 x) for any g ∈ G, f ∈ A. This is the right regular representation.
ˆ Finally, we can define the conjugation action by g · f = (µ(g)f )(x) = f (g −1 xg).
Proposition 20.3.5. For every f ∈ A, there exists a finite dimensional subspace V ⊂ A
such that f ∈ V and ρ(G)V = V ; that is, the ρ-action of G on A is locally finite.
Proof. Let {gi }i∈I be a basis for A over k. Apply comultiplication to f : ∆(f ) = ni=1 fi ⊗ gi
P
for some fi ∈ A and gi in the basis. Notice
n
X n
X
f = (1 ⊗ ε)∆(f ) = ε(gi )fi = gi (e)fi .
i=1 i=1

Therefore f ∈ Span{f1 , . . . , fn }. Set V = Span{f1 , . . . , fn }. We claim V satisfies the


conditions of the proposition. Consider
n
X n
X
∆(fj ) ⊗ gj = fj ⊗ ∆(gj ) by coassociativity
j=1 j=1
n
!
X X
= fj ⊗ gji ⊗ gi
j=1 i
" n
! #
X X
= fj ⊗ gji ⊗ gi .
i j=1

469
20.3. G-spaces Chapter 20. Algebraic Groups

Pn Pn
Thus ∆(fi ) = j=1 fj ⊗ gji so we see that ρ(g)fi = j=1 gji (g)fj ∈ V . This implies that
ρ(G)V = V .
The following result is critical in the general theory of affine algebraic groups.

Theorem 20.3.6. Let G be an affine algebraic group. Then G is isomorphic to a closed


subgroup of GLn (k) for some integer n ≥ 1.

Proof. Given elements a1 , . . . , am ∈ A, there exists a finite dimensional G-submodule V ⊂ A


such that a1 , . . . , am ∈ V and ρ(G)V = V (applying Proposition 20.3.5 m times and taking
the span of all the resulting vector spaces). We also know A is a finitely generated k-
algebra since it’s noetherian. Then there exists a finite dimensional G-submodule of A
which generates A as an algebra (e.g. V may be Pnchosen to correspond with f1 , . . . , fn , the
generators of A over k). For g ∈ G, ρ(g)fj = i=1 aij (g)fi and applying comultiplication
yields
Xn
∆(fj ) = fi ⊗ aij .
i=1

It follows that the map ϕ : G → GLn (k), g 7→ (aij (g)), is a group homomorphism and hence
a morphism of algebraic groups. The comorphism for ϕ is

ϕ∗ : k[GLn (k)] −→ A
p 7−→ p ◦ ϕ.

Pn xij ∈ k[GLn (k)] be the ijth entry of p. Then ϕ (xij ) = aij ∈ A, but fj (g) = (ρ(g)fj )(e) =
Let
i=1 aij (g)fi (e). Thus
Xn n
X
fj = fi (e)aij = fi (e)ϕ∗ (xij ).
i=1 i=1

Therefore ϕ∗ is onto, so A is a homomorphic image of k[GLn (k)]. In other words, A ∼ =


k[GLn (k)]/I for some radical ideal I ⊂ k[GLn (k)]. By the Nullstellensatz, this identifes G
with a closed subgroup of GLn (k).
Instead of affine algebraic groups, it is common to call these ‘linear algebraic groups’
instead, a term which is explained by Theorem 20.3.6.

470
20.4. Jordan Decomposition Chapter 20. Algebraic Groups

20.4 Jordan Decomposition


Let V be a finite dimensional vector space over k and take a linear operator a : V → V ,
i.e. a ∈ Endk (V ). Then a is semisimple if V has a basis of eigenvectors of a; that is if there
is an ordered basis such that the matrix of a is diagonal. Let cha (T ) be the characteristic
polynomial of a:
m
Y
cha (T ) = det(T I − a) = (T − λi )ni
i=1

where λi are the distinct eigenvalues of a. We will need the following lemmas.
Lemma 20.4.1. If a, b ∈ Endk (V ) are semisimple and commute, then a + b and ab are
semisimple, and a and b can be simultaneously diagonalized.
Lemma 20.4.2. If a and b are semisimple, commuting, nilpotent (resp. unipotent) operators
then a + b and ab are also semisimple, commuting and nilpotent.
By the Chinese remainder theorem, there exists a polynomial p(T ) satisfying p(T ) ≡ 0
mod T and p(T ) ≡ ai mod (T − λi )ni for all i =L1, . . . , m. Define as = p(a) and an = a − as .
m
We can decompose V as a direct sum V = i=1 Vi where a(Vi ) ⊆ Vi and cha|Vi (T ) =
(T − λi )ni . (In other words, Vi is a generalized eigenspace for λi .) We claim that as is
semisimple and an is nilpotent, as an = an as , and a = as + an is a unique decomposition of
a having these properties.
Suppose a = a0s + a0n for a0s , a0n satisfying the above conditions. Then as + an = a0s + a0n .
Note that a commutes with a0s and a0n , so a0s commutes with p(a) = as ; likewise a0n commutes
with a − p(a) = an . Now subtract: as − a0s = a0n − an . By Lemma 20.4.1, any two commuting
semisimple operators can be simultaneously diagonalized, but this diagonalization turns
a0n − an into 0. Hence as − a0s = 0 = a0n − an so a0s = as and a0n = an .
Definition. The unique choice of semisimple as and nilpotent an such that a = as + an is
called the Jordan decomposition of a.
We have thus proved the following theorem.
Theorem 20.4.3 (Jordan Decomposition). Every linear operator a ∈ Endk (V ) on a finite
dimensional vector space V has a unique Jordan decomposition a = as + an where as is
semisimple and an is nilpotent.
Corollary 20.4.4. Let a ∈ GL(V ) for a finite dimensional vector space V . Then there is a
unique factorization a = as au = au as , where as is semisimple and au is unipotent.
Proof. Write a = as + an by the Jordan decomposition theorem. Since as an = an as , a−1 s an
−1 −1
is also nilpotent so we can write a = as + an = as (1 + as an ) = as au where au = 1 + as an
is unipotent.
We next develop some of the properties of the Jordan decomposition.
Proposition 20.4.5. Let a ∈ Endk (V ) for a finite dimensional vector space V and suppose
W ⊂ V is a subspace which is a-invariant.

471
20.4. Jordan Decomposition Chapter 20. Algebraic Groups

(1) W is as -invariant and an -invariant. Moreover, (a|W )s = as |W and (a|W )n = an |W .

(2) Given the operator aV /W : V /W → V /W induced by a, (aV /W )s = (as )V /W and


(aV /W )n = (an )V /W .

Proof. Clear from the definitions.

Proposition 20.4.6. Let V, W be finite dimensional vector spaces over k and suppose a ∈
Endk (V ), b ∈ Endk (W ) and ϕ ∈ Homk (V, W ) such that the following diagram commutes:
ϕ
V W

a b

V ϕ W

Then the semisimple and nilpotent parts of a and b also make the diagram commute:
ϕ ϕ
V W V W

as bs an bn

V ϕ W V ϕ W

Proof. Proposition 20.4.5 implies that the property holds in the special case that ϕ is an
injection or a surjection. In general, every morphism ϕ : V → W can be factored through
the direct sum:
σ π
V −−−→ V ⊕ W −−−→ W
v 7−→ (v, ϕ(v))
(v, w) 7−→ w.

Then the special case may be applied.

Corollary 20.4.7. Let a ∈ Endk (V ) and b ∈ Endk (W ) be linear operators and consider the
linear operators

a ⊕ b : V ⊕ W −→ V ⊕ W
(v, w) 7−→ (a(v), b(w));
and a ⊗ b : V ⊗k W −→ V ⊗k W
v ⊗ w 7−→ a(v) ⊗ b(w).

Then (a ⊕ b)s = as ⊕ bs , (a ⊕ b)n = an ⊕ bn , (a ⊗ b)s = as ⊗ bs and (a ⊗ b)n = an ⊗ bn .

472
20.4. Jordan Decomposition Chapter 20. Algebraic Groups

We next describe what happens when dim V = ∞.


Definition. An operator a ∈ Endk (V ) is called locally finite if for every v ∈ V , there
exists a finite dimensional subspace W ⊂ V such that v ∈ W and a(W ) ⊆ W .
Proposition 20.4.8. Every locally finite operator on V has a Jordan decomposition.
Let V be a rational G-module. Recall from Section 20.3 that this means V is a comodule
for A = k[G] with structure map ∆V : V → V ⊗k A and counit ε : A → k satisfying
(1 ⊗ ∆)∆V = (∆V ⊗ 1)∆V and (1 ⊗ ε)∆V = idV .
Proposition 20.4.9. Every rational G-module V is locally finite; that is, for every v ∈ V
there is a finite dimensional subspace W ⊂ V such that v ∈ W and G(W ) = W .
Proof. Look at ∆V (v) = ri=1 vi ⊗ gi for vi ∈ V and gi are some elements of a basis {gi }i∈I
P

Pr Then W = Span{v1 , . . . , vr } is G-stable and the second comodule axiom gives us


for A.
v = i=1 gi (e)vi ∈ W .
Example 20.4.10. The coordinate algebra A itself P is a rational G-module with structure
map ∆A = ∆ : A → A ⊗k A and G-action x · f = ri=1 gi (x)fi = ρ(x)(f ) for all x ∈ G.
Recall that this corresponds to the left regular representation of G.
Remark. If V, W are rational G-modules then V ⊕ W and V ⊗k W are also rational G-
modules. For V ⊗k W , the structure map ∆V ⊗k W is a result of the following commuting
diagram (and there is an analagous one for V ⊕ W ):
∆V ⊗ ∆W
V ⊗k W (V ⊗k A) ⊗k (W ⊗k A)

(2, 3)

∆V ⊗k W (V ⊗k W ) ⊗k (A ⊗k A)

idV ⊗ m

(V ⊗k W ) ⊗k A

where m is the multiplication map A ⊗k A → A.


Lemma 20.4.11. Let V be a finite dimensional rational G-module, set V0 = V as a vector
space and view V0 as a G-module by the trivial action of G. Then V is a G-submodule of
V0 ⊗k A.
Proof sketch. Show that the map ∆V : V → V0 ⊗k A is an injection.
Notice that V0 ⊗k A = A ⊕ A ⊕ · · · ⊕ A = A⊕n where n = dim V . Then every finite
dimensional G-space can be viewed as a submodule of some copies of A. In other words, A
acts like an injective hull for finite dimensional rational G-modules. S
Now let a : V → V be a locally finite operator on a vector space V = i∈I Vi where each
Vi is finite dimensional and a(Vi ) ⊆ Vi . Then a|Vi is defined and has Jordan decomposition
a|Vi = (a|Vi )s + (a|Vi )n . Therefore a can be decomposed uniquely into a = as + an where
as , an commute, as is locally semisimple and an is locally nilpotent. In addition, if

473
20.4. Jordan Decomposition Chapter 20. Algebraic Groups

ϕ
V W

a b

V ϕ W

commutes, then as in the finite dimensional case, the following diagrams all commute:
ϕ
V W

as , an , au bs , bn , bu

V ϕ W

Theorem 20.4.12 (Abstract Jordan Decomposition). Let G be an affine algebraic group.

(1) There exist unique elements gs , gu ∈ G that commute such that ρ(g) has Jordan de-
composition ρ(g) = ρ(g)s ρ(g)u = ρ(gs )ρ(gu ).

(2) If ϕ : G → G0 is a morphism of affine algebraic groups then ϕ(gs ) = ϕ(g)s and


ϕ(gu ) = ϕ(g)u .

(3) If G = GLn (k) then for every g ∈ G, gs and gu are the semisimple and unipotent parts
of g.

Proof. (1) Let m denote the multiplication map A ⊗k A → A, where A = k[G]. Then the
following diagram commutes:
m
A⊗k A

ρ(g) ⊗ ρ(g) ρ(g)

A⊗k A
m

since ρ(g) is an algebra automorphism of A. Therefore by the previous statement, the


semisimple and unipotent parts of ρ(g) induce commuting squares as well, so ρ(g)s and
ρ(g)u are also algebra automorphisms of A. The counit of A gives an algebra homomorphism
ε ◦ ρ(g)s : A → k but these are all identified with the elements of G. Thus there is a unique
gs ∈ G corresponding to ε ◦ ρ(g)s ; similar logic shows there is a unique element gu ∈ G for
ε ◦ ρ(g)u . (Explicitly, gs satisfies (ρ(g)s f )(e) = f (gs ) for all f ∈ A.) Then for all x ∈ A,

(ρ(g)s f )(x) = (λ(x−1 )ρ(g)s f )(e) = (ρ(g)s λ(x−1 )f )(e) since λ, ρ commute
= (λ(x−1 )f )(gs ) by the defining property of gs
= f (xgs ) = (ρ(gs )f )(x).

474
20.4. Jordan Decomposition Chapter 20. Algebraic Groups

Since they agree at every x ∈ A, ρ(g)s and ρ(gs ) must be the same operator. The proof is
the same for ρ(gu ). Then we have ρ(g) = ρ(g)s ρ(g)u = ρ(gs )ρ(gu ) = ρ(gs gu ) = ρ(gu gs ) so the
elements gs , gu must commute.
To see that gs and gu are unique, suppose gs0 , gu0 ∈ G such that g = gs0 gu0 and ρ(gs0 ) =
ρ(g)s , ρ(gu0 ) = ρ(g)u . Let V be a G-space. Then for any nonzero f ∈ V ∗ = Homk (V, k),
the map f˜ : V → A sending v to f˜(v)(g) = f (gv) is an injective G-module homomorphism.
Indeed, if f˜ = 0 then f (gv) = 0 for all g ∈ G, v ∈ V which implies v = 0, so f˜ is injective.
On the other hand, for g, h ∈ G, v ∈ V , f˜(hv)(g) = f (ghv) = f˜(v)(gh) = (ρ(h)f˜(v))(g) so
f˜(hv) = ρ(h)f˜(v) and thus f˜ is a map of G-modules. For all g ∈ G, the diagram


V A

g ρ(g)

V A

commutes, so by the work above, gs0 and ρ(gs0 ) also make the diagram commute, as do gu0
and ρ(gu0 ). Since ρ(gs0 ) = ρ(g)s = ρ(gs ) and ρ(gu0 ) = ρ(g)u = ρ(gu ) and f ∈ V ∗ was arbitrary,
this identifies gs = gs0 and gu = gu0 .
(2) Given ϕ : G → G0 , ϕ is surjective onto im ϕ, which is a closed subgroup of G0 by (2)
of Theorem 20.2.4. Then we can factor

ϕ : G  im ϕ ,→ G0 .

We can reduce to two cases: either G is isomorphic to a closed subgroup of G0 , in which


case ϕ is injective, or ϕ is surjective. In the former case, A = k[G] = A0 /I for I an ideal of
A0 = k[G0 ]. Then ϕ(G) = {x ∈ G0 | ρ(x)I = I} so ρ(gs )I ⊆ I and ρ(gu )I ⊆ I. Therefore
gs , gu ∈ ϕ(G). On the other hand if ϕ is surjective, then A may be viewed as a subspace of
A0 which is stable under ρ(g) for all g ∈ G. In both cases, (2) follows by Proposition 20.3.5.
(3) is straightforward.
Example 20.4.13. Let G = Ga = (k, +). Recall that this has the coordinate algebra
A = k[G] = k[t]. Since G is affine, it is isomorphic to a linear algebraic group. In fact,
  
1 a
Ga ∼ = U2 = : a ∈ k ≤ GL2 (k).
0 1

Every element of U2 is unipotent, so Ga is a unipotent group. Consider a representation


ϕ : Ga → GL(V ). Then ϕ(Ga ) consists of unipotent operators. Recall that V is a submodule
of V0 ⊗ A = A ⊕ A ⊕ · · · ⊕ A = A⊕ dim V . Then A is a locally finite Ga -module so it is a union
of finite dimensional, Ga -stable submodules. Set An = {f (t) ∈ A | deg f ≤ n}. If b ∈ k is a
scalar, then b acts on monomials by

ρ(b)(tn ) = (t + b)n = tn + nbtn−1 + . . .

Viewed in the basis {1, t, . . . , tn } of An , ρ(b) is represented by a unipotent matrix.

475
20.4. Jordan Decomposition Chapter 20. Algebraic Groups

Theorem 20.4.14. Let G be a commutative affine algebraic group and let Gs (resp. Gu )
be the subset of semisimple (resp. unipotent) elements of G. Then Gs and Gu are closed
subgroups of G and Gs × Gu ∼ = G. Further, if G is irreducible then so are Gs and Gu .
Proof. Assume G is a closed subgroup of some GLn (k). View GLn (k) as GL(V ) where
V is the vector space k n . Then x ∈ GL(V )u ⇐⇒ (x − 1)m = 0 for some m ≥ 1.
Therefore GL(V )u is closed since the elements satisfy polynomial equations, so in particular
Gu = G ∩ GL(V )u is closed in G. If x, y ∈ Gu then x − 1, y − 1 are both nilpotent and

xy − 1 = xy − y + y − 1 = (x − 1)y + y − 1

is nilpotent since the elements of G commute. Hence xy ∈ Gu so Gu is a closed subgroup.


Now, Gs consists of commuting semisimple matrices, so they can be simultaneously diag-
onalized, i.e. there is a basis of V consisting of eigenvectors g ∈ Gs . Thus there
L are distinct
homomorphisms ϕi : Gs → k × and a vector space decomposition V = i Vi into finite
dimensional nonzero subspaces Vi such that given s ∈ Gs , s acts as scalar multiplication by
ϕi (s) on Vi for each i. In other words, Vi = {v ∈ V | sv = ϕi (s)v for all s ∈ Gs }. Since
elements of Gu commute with elements of Gs , the Vi are all Gu -stable. Thus we may assume
that the elements of Gu act as upper triangular on each Vi and therefore on V itself. Let Tn
be the subgroup of diagonal matrices in GLn (k). Then Gs = G ∩ Tn so in particular Gs is a
closed subgroup.
Lastly, Gs × Gu → G, (s, u) 7→ su is an isomorphism of affine algebraic groups by the
Jordan decomposition theorem. If G is irreducible, then since it surjects onto Gs and Gu , it
follows that Gs and Gu are also irreducible.
Remark. In general, Gs and Gu need not be subgroups, though the proof above shows that
Gu as always closed.
Proposition 20.4.15. Suppose G is a one-dimensional, connected, affine algebraic group.
Then
(1) G is abelian.

(2) G is either semisimple or unipotent.

(3) If G = Gu then any element x ∈ G has finite order dividing char k. (When char k = 0,
this is interpreted to mean that G has no elements of finite order.)
Proof. (1) Let g ∈ G and define the morphism

ϕ : G −→ G
x 7−→ xgx−1 .

Then im ϕ is connected so im ϕ is also connected. Thus either im ϕ = {g} or im ϕ = G.


Suppose the latter is true. Then by Chevalley’s Theorem, im ϕ contains a nonempty open
subset U ⊂ G. Since G is one-dimensional, G r im ϕ is finite. Assume G is a closed subgroup
of some GLn (k). Then the elements of im ϕ are all similar so they have the same characteristic
polynomial. Therefore the morphism χ : G → k[t] taking g to its characteristic polynomial

476
20.4. Jordan Decomposition Chapter 20. Algebraic Groups

over k has finite image consisting of the characteristic polynomials of g and the finitely many
characteristic polynomials of the elements of G r im ϕ. Since G is connected (irreducible),
the image of χ is a finite, irreducible subset of k[t], hence one point, so it must be that χ(G)
is a single element, which must be (x − 1)n since this is the characteristic polynomial of the
identity. This shows every element of G is unipotent, so by Theorem 20.3.3, G is similar to
a subgroup of GLn (k) consisting of upper triangular unipotent matrices. In particular, G
is solvable, so the derived subgroup G0 = [G, G] is a proper subgroup of G. We know from
Corollary 20.2.9 that G0 is closed and connected, but since G is one-dimensional, this means
G0 = {e}. Thus g −1 ϕ(x) = g −1 xgx−1 ∈ G0 = {e} for all x ∈ X, contradicting the hypothesis
that im ϕ = G. Hence im ϕ = {g} which implies that G is commutative.
(2) Theorem 20.4.14 shows that G ∼ = Gs × Gu but since G is connected, each of these
components is connected. Moreover, 1 = dim G = dim Gs + dim Gu so it must be that either
G = Gu or G = Gs .
(3) Assume G = Gu and char k = p > 0. Let x ∈ G be a nonidentity element. For
h h
some h ∈ N, define G(p ) ≤ G to be the image of the map G → G, y 7→ y p , which is a
h
morphism of affine algebraic groups since G is abelian. Then G(p ) is a closed subgroup of G
h h
which is connected. Since G is one-dimensional, either G(p ) = {e} or G(p ) = G. Suppose
ph > n. Since G = Gu , each x ∈ G may be written as x = 1 + λ for a nilpotent element λ,
h h h h
so xp = (1 + λ)p = 1 + λp = 1. Thus when ph > n, G(p ) = {e}. If G(p) = G then for all
h h−1 h
h > 0, G(p ) = (G(p ) )(p) but this contradicts the fact that G(p ) = {e} for large enough h.
Therefore G(p) = {e} so every element of G has order dividing the characteristic of k.
Example 20.4.16. Let G = GLn (k). Define the following subgroups of G:
     

 1 ∗ 
 
 ∗ ∗ 
 
 ∗ 0 

 ..  ..  ..
Un =  , = , = .  .
  
.  B n  .  Tn 
     
 0 1   0 ∗   0 ∗ 
Then Un is a nilpotent, solvable subgroup of GLn (k). Bn (called a Borel subgroup; see
Chapter 21) is also nilpotent but it is not solvable, since [Bn , Bn ] = Un . The subgroup Tn is
known as an n-dimensional torus; it is abelian, solvable and nilpotent. It’s also not hard to
see that Tn ∼
= G×n
m = Gm × · · · × Gm . We can regard Tn and Un as closed subgroups of Bn .
We will see in Chapter 21 that Bn ∼ = Tn n Un .
We will prove the following characterization in Section 21.4, which says that every con-
nected, solvable, affine algebraic group behaves in precisely the same way as in the example
above.
Theorem 20.4.17. Let G be a connected, solvable, affine algebraic group. Then
(1) For any s ∈ Gs , ZG (s) is connected.
(2) Gs lies in a maximal torus in G, i.e. a maximal closed subgroup isomorphic to a
product of copies of Gm .
(3) Any two maximal tori in G are conjugate.
(4) Gu is a closed, connected subgroup of G and if T is a maximal torus, then T ×Gu → G
is an isomorphism of varieties (not necessarily of algebraic groups).

477
20.5. Lie Algebra of an Algebraic Group Chapter 20. Algebraic Groups

20.5 Lie Algebra of an Algebraic Group


In this section we define a Lie algebra g for an affine algebraic group G. In the early days, the
theory was developed for algebraic groups over fields of characteristic zero. In this setting,
g and G are closely related. In characteristic p > 0, things are not so nice – it’s better in
this case to work with affine group schemes.
Let G be an affine algebraic group over a field k with char k = 0.

Definition. The Lie algebra of G is the tangent space to G at the identity, g = Te G.


That is, g consists of the point derivations at e, i.e. the k-linear maps δ : A → k satisfying
δ(f g) = f (e)δ(g) + δ(f )g(e).

We can also view this tangent space as Te G = (Me /M2e )∗ where Me = {f ∈ k[G] | f (e) =
0}. Note that since G is smooth, dim g = dim G = tr degk k(G) (using Theorem 12.1.17 for
the second equality). To prove g is indeed a Lie algebra, consider

Der(A) := Derk (A, A) = {D : A → A | D is k-linear and D(f g) = f D(g) + D(f )g}.

Then for any X, Y ∈ Der(A), [X, Y ] := XY − Y X ∈ Der(A) which defines a Lie algebra
structure on Der(A). However, dimk Der(A) = ∞ so this space is too big.

Example 20.5.1. Leibnitz’s derivation formula for D ∈ Der(A) may be expanded:


n  
n
X n
D (f g) = Di (f )Dn−i (g) for any n > 0.
i=0
i

When char k = p > 0, we have Dp (f g) = Dp (f )g + f Dp (g) so Dp is also a derivation! (This


is not true in characteristic zero.) This additional structure on Der(A) leads to the notion
of a p-restricted Lie algebra, which comes equipped with a [p]-operator:

g −→ g
x 7−→ x[p] .

This satisfies

ˆ (λx)[p] = λp x[p] for all λ ∈ k;

ˆ ad(x[p] ) = (ad x)p ;


Pp−1 fi (x,y)
ˆ (The Jacobson formula): (x + y)[p] = x[p] + y [p] + i=1 i
where fi are polynomials
in x, y taking values in g.

There is a finite dimensional algebra U(g)− , called the restricted enveloping algebra of g,
that contains g as a Lie subalgebra and which satisfies x[p] = xp ∈ U(g)− for all x ∈ g. The
dimension of U(g)− is pdim g . This algebra U(g)− is actually what’s known as a Frobenius
algebra, meaning it’s similar in nature to the group algebra of a finite group. Further, U(g)−
is also a Hopf algebra.

478
20.5. Lie Algebra of an Algebraic Group Chapter 20. Algebraic Groups

In our setting, Der(A) is a restricted enveloping algebra, but it’s not finite dimensional so
we will relate Te G to a finite dimensional subalgebra of Der(A) that satisfies the properties
of the Lie algebra we’re looking for.

Lemma 20.5.2. Let δ ∈ Te G be a point derivation. Then D = (1 ⊗ δ)∆ is a derivation of


A.

Proof. For f, g ∈ A, we have

D(f g) = (1 ⊗ δ)∆(f g) = (1 ⊗ δ)∆(f )∆(g)


! !
X X
= (1 ⊗ δ) fi ⊗ fi0 gj ⊗ gj0
i j
X
= (1 ⊗ δ) f i gj ⊗ fi0 gj0
i,j
X
= fi gj (fi0 (e)δ(gj0 ) + δ(fi0 )gj0 (e))
i,j
! !
X X X X
= fi fi0 (e) gj δ(gj0 ) + gj gj0 (e) fi δ(fi0 )
j i i j
X X
= f gj δ(gj0 ) + gfi δ(fi0 ) by the counit axiom
j i
X X
=f gj δ(gj0 ) + g fi δ(fi0 )
j i

= f (1 ⊗ δ)∆(g) + g(1 ⊗ δ)∆(f )


= f D(g) + gD(f ).

Therefore D is a derivation.

Lemma 20.5.3. Let D ∈ Der(A) and define δ = ε ◦ D. Then δ ∈ Te G.

Proof. Let f, g ∈ A. Then we have

δ(f g) = εD(f g) = ε(f D(g) + gD(f ))


= f (e)D(g)(e) + g(e)D(f )(e)
= f (e)δ(g) + g(e)δ(f ).

Hence δ is a point derivation of A at e.


Lemmas 20.5.2 and 20.5.3 give us maps
α
Te G Der(A)
β

479
20.5. Lie Algebra of an Algebraic Group Chapter 20. Algebraic Groups

We claim that β ◦ α = id. To see this, let δ ∈ Te G be a point derivation of A. Then

α(δ)(f ) = (1 ⊗ δ)∆(f ) = D(f ) ∈ Der(A)


!
X X
and β(D)(f ) = (1 ⊗ δ) fi ⊗ fi0 (e) = fi (e)δ(fi0 )
i i
!
X
=δ fi (e)fi0 = δ(f ).
i

Therefore β ◦ α is the identity on Te G so α is injective. This tells us that there is a copy of


Te G sitting inside Der(A). We next identity its image.
Recall the left and right actions of G on A:

ρg : A → A,f →
7 g·f
λg : A → A,f →7 f · g −1 ,

where (g · f )(x) = f (xg) and (f · g −1 )(x) = f (g −1 x) for all g, x ∈ G and f ∈ A.


Theorem 20.5.4. The image of Te G under α : Te G → Der(A) is the set

{D ∈ Der(A) | D(f · x) = D(f ) · x for all f ∈ A, x ∈ G}.

Proof. Assume D ∈ im α, i.e. D = (1 ⊗ δ)∆ for some point derivation δ ∈ Te G. Observe


that f · x = (εx ⊗ 1)∆(f ) where εx : A → k, εx (h) = h(x). By coassociativity,

(1 ⊗ δ)∆(εx ⊗ 1)∆ = (1 ⊗ 1 ⊗ δ)(εx ⊗ 1 ⊗ 1)(1 ⊗ ∆)∆


= (εx ⊗ 1 ⊗ 1)(1 ⊗ 1 ⊗ δ)(∆ ⊗ 1)∆
= (εx ⊗ 1 ⊗ δ)(∆ ⊗ 1)∆;
and (εx ⊗ 1)∆(1 ⊗ δ)∆ = (εx ⊗ 1 ⊗ 1)(1 ⊗ 1 ⊗ δ)(∆ ⊗ 1)∆
= (εx ⊗ 1 ⊗ δ)(∆ ⊗ 1)∆.

So we see that (1 ⊗ δ)∆ and (εx ⊗ 1)∆ commute. Hence D(f · x) = D(f ) · x.
This space α(Te G) has a Lie algebra structure with Lie bracket
X
[σ, τ ] = (σ(fi )τ (fi0 ) − τ (fi )σ(fi0 )).
i

Corollary 20.5.5. If ϕ : G → G0 is a morphism of algebraic groups, there is a corresponding


Lie algebra homomorphism

ϕ∗ : g −→ g0
δ 7−→ δ ◦ ϕ∗ ,

where ϕ∗ : k[G0 ] → k[G] is the induced algebra homomorphism.


Corollary 20.5.6. The assignment g(−) of an algebraic group over k to its Lie algebra is
a covariant functor on the category AffGpsk of affine algebraic groups over k.

480
20.5. Lie Algebra of an Algebraic Group Chapter 20. Algebraic Groups

Examples.

1 Let G = Ga with coordinate algebra A = k[t]. To compute g, note that any derivation
is determined by its value on t. For all x, y ∈ Ga , we have (λx t)(y) = y − x so every
left translation of t may be written λx t = t − x. Suppose D is a derivation on k[t]
commuting with all λx for x ∈ Ga . Then D(t) = ∞ i
P
a
i=0 i t for some ai ∈ k, all but
finitely many of which are zero, and we have

X ∞
X
i
λx D(t) = λx ai t = ai (t − x)i
i=0 i=0

X
while Dλx (t) = D(t − x) = D(t) − D(x) = D(t) = ai ti .
i=0

all i ≥ 1, so that D(t) = a = a0 is a


If λx D = Dλx , this is only possible if ai = 0 for
scalar. Thus D = a dtd . Hence g has a basis dtd and for any x ∈ Ga = k, we have

d d
(t + x)n = (tn ) · x.
dt dt
Furthermore, when char k = p > 0,
 p
d
tn = n(n − 1) · · · (n − p + 1)tn−p = 0.
dt

Thus g is a p-restricted Lie algebra in this case.

2 Let G = Gm with coordinate algebra A = k[t, t−1 ]. Then if D is a derivation on k[t, t−1 ]
commuting with all λx , x ∈ Gm , we may write D(t) as a Laurent series,

X
D(t) = ai ti
i=−∞

with ai ∈ k × and ai = 0 for all but finitely many i. Consider



X ∞
X
λx D(t) = λx i
ai t = ai x−i ti
i=−∞ i=−∞

X
−1 −1 −1
while Dλx (t) = D(x t) = x D(t) = x ai ti .
i=−∞

This implies x−1 ai = ai x−i for all i, which is only possible if ai = 0 for all i 6= 1 and
D(t) = at for a nonzero scalar a = a1 ∈ k × . Thus D = at dtd , which shows that the
Lie algebra g is spanned by the basis t dtd . Note that t dtd (tn ) = ntn . In characteristic
p > 0,  p
d
t tn = tn
dt

481
20.5. Lie Algebra of an Algebraic Group Chapter 20. Algebraic Groups

p
so t dtd = t dtd , in which case the [p]-operator is just the identity. Once again, g is a
p-restricted Lie algebra.
The previous two examples show that the Lie algebras of Ga and Gm are abstractly
isomorphic (they are one-dimensional k-algebras in every characteristic), but not as
restricted Lie algebras. Interestingly, we may view the map

Ga −→ Gm
x 7−→ xn

as a covering space in characteristic zero, and thus the tangent spaces are preserved.
However, in characteristic p > 0, x 7→ xp is not a covering space (in particular, it is
not étale; see Part VI).

3 Let G = GLn (k). The coordinate algebra may be written k[xij ] ∇1 where ∇ = det
 

as an operator. Then g is spanned by dxdij and the Lie bracket is simply [X, Y ] =

x=e
XY − Y X as one should expect. In characteristic p > 0, g is a restricted Lie algebra
under the operator (aij )[p] = (apij ).

Theorem 20.5.7 (Chevalley). Let H be a closed subgroup of an algebraic group G, with Lie
algebras h ⊆ g. Then there exists a representation ρ : G → GL(V ) and a nonzero vector
v ∈ V such that

H = {x ∈ G | ϕ(x)(v) ∈ kv} and h = {a ∈ g | dϕ(a)(v) ∈ kv}.

482
Chapter 21

Classifying Algebraic Groups

483
21.1. Quotients of Algebraic Groups Chapter 21. Classifying Algebraic Groups

21.1 Quotients of Algebraic Groups


Let G be an algebraic group and H a closed subgroup of G. A natural question is: how do
we put a variety structure on the quotient G/H = {aH | a ∈ G}? First we must topologize
G/H. Given the quotient map π : G → G/H, we stipulate that U ⊆ G/H is open if and
only if π −1 (U ) is open in G. Next we define a sheaf of functions on G/H. Given the sheaf
OG for G, set O(U ) = OG (π −1 (U )) for each open U ⊆ G/H.

Proposition 21.1.1. The space OG/H = {O(U ) = OG (π −1 (U )) | U ⊆ G/H open} is a sheaf


of k-valued functions on G/H.

We have thus defined a topological space G/H, which inherits the Zariski topology from
the quotient map π : G → G/H, and a sheaf of functions OG/H .

Theorem 21.1.2. Let G be an algebraic group and let H be a closed subgroup.

(1) G/H is an algebraic variety with sheaf of functions OG/H as defined above.

(2) If G is affine then G/H is affine.

(3) If H is a normal subgroup, then G/H is an algebraic group.

(4) G/H is universal with respect to G-spaces which are H-trivial. That is, for any G-
space Y and any y ∈ Y such that H ⊆ Gy , there is a unique G-equivariant morphism
ϕ : G/H → Y making the diagram

αy

G/H Y
ϕ

commute, where αy : g 7→ g · y.

Given G and H as above, let A = k[G] be the coordinate algebra of G. Recall that A is a
rational G-module (via translation). We saw that every finite dimensional subspace W ⊂ A
is contained in a finite dimensional G-stable subspace V ⊂ A.

Lemma 21.1.3. For any closed subgroup H ≤ G, there exists a finite dimensional subspace
V ⊂ k[G] and a subspace W ⊂ V such that

(1) V is G-stable; that is, ρ(g)V = V for all g ∈ G.

(2) H is the stabilizer of W ; that is, H = {g ∈ G | ρ(g)W = W }.

(3) h = {X ∈ g | dρ(X)W ⊆ W } is the Lie algebra of H as an algebraic group and is a


Lie subalgebra of g.

484
21.1. Quotients of Algebraic Groups Chapter 21. Classifying Algebraic Groups

Proof. Let I ⊂ A be the ideal corresponding to H under the Nullstellensatz. Then I =


(f1 , . . . , fr ) for functions fi ∈ A, so there exists a finite dimensional, G-stable subspace
B ⊂ A containing these generators. Take W = I ∩ V . Note that if h ∈ H and f ∈ I, then
h · f ∈ I. This shows that W is H-stable, since V itself is G-stable and hence H-stable.
Conversely, if g ∈ G and ρ(g)W = W then ρ(g)fi ∈ W ⊂ I for each fi . So (ρ(g)fi )(e) = 0
which implies fi (g) = 0 for all i. Thus g ∈ H. This proves (1) and (3), and (3) is just an
exercise in the definitions.
For a finite dimensional vector space VVwith a d-dimensional subspace W ⊆ V , every
x ∈ GL(V ) acts on the dth exterior power d V via the action

g · (v1 ∧ · · · ∧ vd ) = gv1 ∧ · · · ∧ gvd .


V 
This induces a representation ϕ : GL(V ) → GL d V .
Vd
Lemma 21.1.4. Let {v1 , . . . , vd } be a basis of W and set L = k(v1 ∧ · · · ∧ vd ) ⊆ V . Then

(1) For each x ∈ GL(V ), xW = W if and only if ϕ(x)L = L.

(2) For each X ∈ gl(V ), XW ⊆ W if and only if dϕ(L) ⊆ L.

Proof. (1) Choose a basis {v1 , . . . , vd } of W and a basis {v`+1 , . . . , v`+d } of xW . These
may overlap since xW ∩ W is a subspace of V . For any x ∈ GL(V ), x · (v1 ∧ · · · ∧ vd ) ∈
k(v`+1 , . . . , v`+d ). If ϕ(x)L = L then v1 ∧· · ·∧vd and v`+1 ∧· · ·∧v`+d are linearly independent,
but this can only happen if ` = 0. This means v1 , . . . , vd also span xW , but dim xW = d =
dim W so we must have xW = w.
(2) is proven similarly.

Theorem 21.1.5. Given G an algebraic group and H a closed subgroup, there exists a
quasi-projective variety X such that

(1) X is homogeneous and connected as a G-space.

(2) There exists a point x ∈ X such that Gx = H.

(3) The fibres of G → X, g 7→ g · x are exactly the left cosets of H in G.

(4) The map G → X, g 7→ g · x is separable.

(5) X ∼
= G/H as varieties.
Proof. (Sketch) By Lemma 21.1.3, there exists a finite dimensional rational G-module V
containing an element v for which Gv = H. Form the projective space P(V ) consisting of
all one-dimensional subspaces of V . Then G acts on P(V ), and v corresponds to a point
x ∈ P(V ). Let X = G · x be the orbit of x under this action. One may check that X satisfies
(1) through (5).

485
21.2. Parabolic and Borel Subgroups Chapter 21. Classifying Algebraic Groups

21.2 Parabolic and Borel Subgroups


Lemma 21.2.1. Let ϕ : X → Y be a bijective morphism of homogeneous G-spaces. Then
X is complete if and only if Y is complete.

Proof. ( =⇒ ) The image of any complete variety is complete.


( ⇒= ) For any variety Z, there is a homeomorphism X × Z → Y × Z which is an open
bijection. Then the following diagram commutes:
X ×Z Y ×Z

π π

This implies X is complete.

Definition. A subgroup P of an algebraic group G is parabolic provided that G/P is com-


plete.

We will prove the following characterization of connected solvable groups.

Theorem 21.2.2. Let G be connected. Then G has a proper parabolic subgroup if and only
if G is not solvable.

Example 21.2.3. If P C G is a normal subgroup then G/P is affine, so P is parabolic if


and only if G0 ⊆ P .

Example 21.2.4. If P ≤ G is any parabolic subgroup then G/P is homeomorphic to a


closed subvariety of a projective space P(V ) for some homogeneous G-space V .

Remark. Let P ≤ G be a parabolic subgroup. Then for every variety X there is a morphism

πX : G × X −→ G/P × X
(g, x) 7−→ (gP, x).
−1
Let A ⊆ G/P ×X and set A0 = πX (A) ⊆ G×X. Observe that if (g, x) ∈ A0 then (gp, x) ∈ A0
for all p ∈ P . If some closed subset B ⊆ G × X satisfies (gP, x) ⊆ B then πX (B) is closed.
This is another way of phrasing the completeness condition.

Lemma 21.2.5. The parabolic property is transitive. Specificallly, let Q ≤ P ≤ G be a chain


of subgroups in an algebraic group G and assume P is parabolic in G and Q is parabolic in
P . Then Q is also parabolic in G.

Proof. Let X be a variety, take A ⊆ G/Q × X and set A0 = p−1 (A), where p : G × X →
G/Q × X. Consider the following diagram:

486
21.2. Parabolic and Borel Subgroups Chapter 21. Classifying Algebraic Groups

α
P ×G×X G × X ⊇ A0

σ p

P/Q × G × X G/Q × X ⊇ A

πG×X πX

e0 ⊆ G × X
A X
πX
e0 = πG×X (σα−1 (A0 )) ⊆ G × X. Then following the diagram around counterclockwise
Let A
e0 ) = πX (A), which is closed by completeness of P/Q. Therefore G/Q is also
shows that πX (A
complete.
Lemma 21.2.6. Let Q ≤ P ≤ G be closed subgroups of an algebraic group G. Then P is
parabolic if Q is parabolic.
Proof. Suppose G/Q is complete. There is a surjective morphism of varieties G/Q → G/P ,
so by (3) of Proposition 12.11.1, G/P is complete.
We will now prove the characterization of connected solvable groups in terms of parabolic
subgroups.
Theorem 21.2.7. A connected algebraic group G has a proper parabolic subgroup if and
only if G is not solvable.
Proof. We will prove the theorem when G is affine. We may then take G to be a closed
subgroup of GL(V ) for some finite dimensional k-vector space V . There is a mapping

V r {0} −→ P(V )
v 7−→ {cv | c ∈ k}.

The action of G on V induces an action on P(V ) which makes P(V ) into a homogeneous
G-space. Take an orbit O = G · x under this action and consider its closure O ⊆ P(V ). We
may assume O has minimal dimension, so that O = O. Since P(V ) is complete, Proposi-
tion 12.11.1(1) implies O is also complete. Thus there is a bijective morphism G/Gx ↔ O.
Let P = Gx = StabG (x). By Lemma 21.2.1, G/P is complete so P is parabolic. If P 6= G
then we’re done. Otherwise suppose P = G. Then we can replace V with the quotient space
V /x and repeat the argument to find x0 ∈ P(V /x) such that Gx0 is parabolic. Continue
this process, inductively defining x0 = x, x1 = x0 , . . . and V0 = V, V1 = V /x0 , . . . Either this
process terminates with a proper parabolic subgroup Pk = Gxk 6= G for some xk , in which
case we’re done, or G stabilizes a flag of subspaces,

0 = Vn ⊆ Vn−1 ⊆ · · · ⊆ V1 ⊆ V0 = V,

such that dim Vi /Vi+1 for i = 0, . . . , n − 1. Choose an ordered basis {v0 , . . . , vn } such that
the subgroup hv0 , . . . , vi i is G-stable for each i = 0, . . . , n − 1. Then relative to this basis, G
is isomorphic to a group of upper triangular matrices which is solvable.

487
21.2. Parabolic and Borel Subgroups Chapter 21. Classifying Algebraic Groups

If G does have a proper parabolic subgroup P , assume G is of minimal dimension among


all connected affine algebraic groups with this property. We may further assume dim P is
maximal among all proper parabolic subgroups of G. Take the derived subgroup G0 = [G, G].
If G is solvable then G0 C G is a proper, closed, normal subgroup. Let Q = P G0 which is
a closed subgroup of G. Certainly P ≤ Q ≤ G so by Lemma 21.2.6, Q is parabolic.
By maximality of dim P , either Q = P or Q = G. If Q = G then by the isomorphism
theorems, G/P = P G0 /P ∼ = G0 /(G0 ∩ P ), so G0 ∩ P is parabolic in G0 . But G0 has strictly
smaller dimension than G, so by minimality of dim G, we must have Q = P G0 ≤ G0 6= G,
contradicting Q = G. On the other hand, if Q = P , then G0 ≤ P so P must be a normal
subgroup of G. Hence G/P is affine by Theorem 21.1.2, which implies by connectedness
and Proposition 12.11.1(6) that G/P is a point. This can only happen if P = G but this
contradicts the assumption that P is a proper subgroup of G. Hence in the solvable case G
does not contain a proper parabolic subgroup.
The theory of complete varieties and parabolic subgroups is closely related to Borel
subgroups.

Definition. A maximal, closed, connected, solvable subgroup B of an affine algebraic group


G is called a Borel subgroup.

Example 21.2.8. The most important example of a Borel subgroup is the group Bn ≤
GLn (k) of upper triangular matrices. In this case GLn (k)/Bn ∼
= Pn , projective n-space.
n
By Theorem 12.11.2, P is complete, which implies that B is also a parabolic subgroup of
GLn (k).

Theorem 21.2.9 (Borel’s Fixed Point Theorem). Let G be a connected solvable group acting
on a complete variety X. Then there exists an element x ∈ X that is fixed by the action of
G.

Proof. Let O = G · x be a closed orbit of the action of G on X, which exists by . Then by


Proposition 12.11.1(1), O is complete and O ∼ = G/Gx so Gx is a parabolic subgroup of G.
Since G is solvable, Theorem 21.2.7 implies Gx = G, and hence x is fixed by the action of
G.

Proposition 21.2.10. Let G be a connected algebraic group, B ≤ G a Borel subgroup and


suppose that σ ∈ Aut(G) fixes B pointwise. Then σ = 1.

Proof. Suppose σ(b) = b for all b ∈ B. This defines an action of G on itself by

G × G −→ G
(g, x) 7−→ σ(g)xg −1 .

If G1 is the stabilizer of 1 ∈ G, then for any b ∈ B, σ(b)1b−1 = bb−1 = 1 so b ∈ G1 .


Thus B ⊆ G1 . By the universal property of quotients (Theorem 21.1.2), we have a map
f : G/B → G making the diagram

488
21.2. Parabolic and Borel Subgroups Chapter 21. Classifying Algebraic Groups

G/B G
f

commute, where α(g) = σ(g)g −1 . Now since G/B is complete and G is affine, by Propo-
sition 12.11.1 f must be constant, so for all g ∈ G, σ(g)g −1 = 1. That is, σ(g) = g as
required.

Theorem 21.2.11. Let G be an affine algebraic group. Then

(1) A closed subgroup P ≤ G is parabolic if and only if P contains a Borel subgroup of G.


In particular, every Borel subgroup of G is parabolic.

(2) Any two Borel subgroups of G are conjugate.

Proof. (1) Suppose P is parabolic and take a Borel subgroup B ≤ G. Consider the action
of B on G/P via b · xP = bxP . Then by Borel’s fixed point theorem, the action has a fixed
point x ∈ G/P , i.e.

B · xP = xP =⇒ x−1 bx ∈ P for all b ∈ B =⇒ x−1 Bx ⊆ P.

Clearly x−1 Bx is also a Borel subgroup, so the forward implication holds. For the converse,
we prove directly that every Borel subgroup B ≤ G is parabolic and use Lemma 21.2.6. If G
itself is connected and solvable, then we’re done. Otherwise, by Theorem 21.2.7, G contains
a proper parabolic subgroup P . Take a Borel subgroup B ≤ G. By the argument above, we
may assume B ≤ P . Then dim G − dim B > dim P − dim B. By induction, we can show that
B is parabolic in P . Thus by Lemma 21.2.6, B is parabolic in G so every Borel subgroup is
parabolic.
(2) Let B, B 0 ≤ G be two Borel subgroups. Then by (1), B 0 is parabolic so by the
fixed point theorem, B is conjugate to a subgroup of B 0 , i.e. there is an x ∈ G such that
xBx−1 ⊆ B 0 . But B is maximal among connected, solvable subgroups of G so we must have
xBx−1 = B 0 . Hence every pair of Borel subgroups is conjugate.

Corollary 21.2.12. Suppose ϕ : G → G0 is a surjective morphism of algebraic groups and


let B (resp. P ) be a Borel (resp. parabolic) subgroup of G. Then ϕ(B) (resp. ϕ(P )) is a
Borel (resp. parabolic) subgroup of G0 .

Proof. Let B ≤ G be a Borel subgroup. Then ϕ(B) is also connected and solvable. Consider
the induced map ϕ : G/B → G0 /ϕ(B). Since B is also parabolic, G/B is complete so
ϕ(G/B) = G0 /ϕ(B) is complete. Thus ϕ(B) is parabolic in G0 , so it is Borel. The proof for
parabolic subgroups follows this proof closely.

Corollary 21.2.13. If G is connected and B ≤ G is a Borel subgroup, then Z(G)0 ⊆


Z(B) ⊆ Z(G).

489
21.2. Parabolic and Borel Subgroups Chapter 21. Classifying Algebraic Groups

Proof. Z(G)0 is closed and connected, and solvable since it’s abelian. So Z(G)0 ⊆ B 0 for
some Borel subgroup B 0 ≤ G. Since B 0 is conjugate to B and Z(G)0 ⊆ Z(G), we may
assume Z(G)0 ⊆ B. This implies Z(G)0 ⊆ Z(B). Now take g ∈ Z(B) and consider the
commutator map
f : G −→ G
x 7−→ gxg −1 x−1 .
Since g ∈ Z(B), this map is constant on left cosets of B:
f (xb) = gxbg −1 b−1 x−1 = gxg −1 bb−1 x−1 = gxg −1 x−1 = f (x).
By the universal property of quotient varieties, this gives us a map
τ : G/B −→ G
xB 7−→ f (x) = gxg −1 x−1 .
Sicne G/B is complete and G is affine, τ is a constant map, which forces gxg −1 x−1 = 1 for
all x ∈ G. Hence g ∈ Z(G).
Corollary 21.2.14. Let G be connected and B ≤ G be a nilpotent Borel subgroup. Then
B = G. That is, G is nilpotent.
Proof. Since B is nilpotent, B contains a closed abelian subgroup Z ≤ B with Z ⊆ Z(B)
and Z 6= {1}. By Corollary 21.2.13, Z ⊆ Z(G). Also, the surjection G → G/Z shows
that G/Z has B/Z as a Borel subgroup. Since B/Z is nilpotent and dim B/Z < dim G,
by induction we obtain B/Z = G/Z. Therefore G/Z is nilpotent, so G is nilpotent. This
implies B = G by maximality of B.
Examples.
3 Let G = GLn (k) and consider the Borel subgroup B of upper triangular matrices.
Then B is in fact the point-stabilizer in the projective variety Flagn (k) of complete
flags
0 = V0 ( V1 ( · · · ( Vn = V
where V = k n . Using the fact that Flagn (k) is complete, this proves that G/B is
complete so B is parabolic.
4 Consider B − ≤ GLn (k), the subgroup of lower triangular matrices. Then B − is a Borel
subgroup of GLn (k) since it is conjugate to B:
 
0 1
wBw−1 = B − where w = 
.

 
..


1 0
The set B − B is a dense open subvariety of G, called the big cell. Let W be the subgroup
of permutation matrices in GLn (k), so that W ∼ = Sn . One can prove that
[
G= BwB
w∈W

490
21.2. Parabolic and Borel Subgroups Chapter 21. Classifying Algebraic Groups

is a disjoint union. This is called the Bruhat decomposition of G, mimicking the decom-
position of the same name for Lie algebras. We will discuss the Bruhat decomposition
further in Section 21.5.

5 Take V = k n and assume char k 6= 2. For x, y ∈ V with x = (x1 , . . . , xn and y =


(y1 , . . . , yn ), set
n
X
hx, yi = xi y i .
i=1

Then GLn (k) acts on the group On (k) = {g ∈ G | hgx, gyi = hx, yi} of orthogonal
linear transformations of V . Recall that a subspace W ⊆ V is isotropic if for all
x, y ∈ W , hx, yi = 0. The Borel subgroups of On (k) are all conjugate to the point-
stabilizer of the variety of isotropic flags of V , i.e. flats

0 = V0 ( V1 ( · · · ( Vn = V

where dim Vi+1 /Vi = 1 and each subspace Vi is isotropic.

We next mention a characterization of algebraic groups which are complete as varieties.


To do so, we need the following lemma.

Lemma 21.2.15. Let X, Y, Z be irreducible varieties with X complete, and suppose there is
a morphism ϕ : X × Y → Z. If for some a ∈ Y , ϕ(·, a) : X → Z is a constant map then
ϕ(·, y) is constant for all y ∈ Y .

Proof. Set Γ = {(x, y, ϕ(x, y)) | x ∈ X, y ∈ Y } ⊆ X × Y × Z, the graph of ϕ, which


is closed. Then the set A = {(y, ϕ(x, y)) | x ∈ X, y ∈ Y } is the image of Γ under the
projection X × Y × Z → Y × Z, so by completeness of X, A is closed. Also, since X, Y, Z
are irreducible, A is irreducible. Let π : A → Y be projection onto Y . If ϕ(·, a) is constant
for some a ∈ Y , then π −1 (a) is a single point. It is well-known that π −1 (a) finite implies that
dim A = dim Y . Now let x ∈ X and set Ax = {(y, ϕ(x, y)) | y ∈ Y } ⊆ A. This is a closed
subset of A and Ax ∼ = Y in the obvious way, so Ax is irreducible. But A was irreducible, so
Ax = A and therefore A ∼ = Y . This implies that ϕ(·, y) is constant for any y ∈ Y .
Theorem 21.2.16. Every connected algebraic group which is complete as a variety is abelian.

Proof. Consider the map ϕ : G × G → G defined by ϕ(x, y) = xyx−1 . Conjugation is a mor-


phism of algebraic groups, and clearly ϕ(·, e) is constant on G. Therefore by Lemma 21.2.15,
ϕ(·, y) is constant for every y ∈ G. But then we must have xyx−1 = y.
Abelian varieties are important tools in modern number theory. For every algebraic curve
C of genus ≥ 1, we can associate to it an abelian variety called the Jacobian variety, J(C)
– this will be done in Part V. These are in some ways analagous to compact complex Lie
groups in Lie theory.

491
21.3. Diagonalizable Groups Chapter 21. Classifying Algebraic Groups

21.3 Diagonalizable Groups


Take an affine algebraic group G and let X ∗ (G) = Hom(G, Gm ) be the set of morphisms
from G to Gm , a one-dimensional torus with coordinate algebra k[Gm ] = k[T, T −1 ]. Recall
that this is a Hopf algebra with comultiplication, counit and antipode maps defined on the
basis elements of k[T, T −1 ] by

∆(T a ) = T a ⊗ T a , ε(T a ) = 1 and γ(T a ) = T −a for all a ∈ Z.

Definition. The set X ∗ (G) = Hom(G, Gm ) is called the character group of G.


Notice that if θ, µ ∈ X ∗ (G) then θµ ∈ X ∗ (G) as well, where (θµ)(x) = θ(x)µ(x). That
is, X ∗ (G) is an abelian group under pointwise multiplication.
Definition. The set X∗ (G) = Hom(Gm , G) is called the set of cocharacters of G.
If G is abelian then X∗ (G) is also an abelian group, but in the general case X∗ (G) may
not be a group. In the abelian case, there is a pairing

X ∗ (G) × X∗ (G) −→ Hom(Gm , Gm ) ∼


=Z
(χ, α) 7−→ χ ◦ α.

The composition χ◦α ∈ Hom(Gm , Gm ) is a Hopf algebra homomorphism, so it takes T 7→ T a


for some a ∈ Z, and it is this integer that identifies Hom(Gm , Gm ) ∼
= Z.
Theorem 21.3.1 (Artin). If χ1 , . . . , χm ∈ X ∗ (G) are characters of G then they are linearly
independent over k.
Definition. An affine algebraic group G is diagonalizable if G is isomorphic to a closed
subgroup of Tn , a maximal torus in GLn (k).
Example 21.3.2. G = Ga has character group X ∗ (Ga ) = Hom(Ga , Gm ) = {1}.
Example 21.3.3. When G = Gm , X ∗ (Gm ) = Hom(Gm , Gm ) ∼ = Z and every character is of
n
the form t 7→ t for some n ∈ Z (on the level of coordinate algebras). More generally, an
n-dimensional torus G = Tn itself is diagonalizable by definition. The term ‘diagonalizable’
comes from the fact that, under a choice of basis in GLn (k), the subgroup Tn consists of
diagonal matrices. Write Tn = Gm × · · · × Gm = G×n m . Its coordinate algebra is the algebra
of Laurent polynomials in n variables: k[Tn ] = k[T1 , T1−1 , . . . , Tn , Tn−1 ]. This has a basis
consisting of monomials T1a1 · · · Tnan , ai ∈ Z. Moreover, the monomials are characters, so
by Artin’s theorem they are linearly independent. This proves X ∗ (Tn ) ∼ = Zn and therefore
k[Tn ] ∼
= k[Zn ] = k[X ∗ (Tn )], the group algebra on X ∗ (Tn ). This generalizes as follows.
Theorem 21.3.4. For an affine algebraic group G, the following are equivalent:
(1) G is diagonalizable.

(2) X ∗ (G) is an abelian group of finite type, and therefore isomorphic to Zn ⊕ F where F
is finite, and the elements of X ∗ (G) form a basis for k[G].

492
21.3. Diagonalizable Groups Chapter 21. Classifying Algebraic Groups

(3) Any finite dimensional rational G-module V is a direct sum of one-dimensional G-


submodules.
Proof. (1) =⇒ (2) Assume G is a closed subgroup of Tn . Then the induced morphism
k[Tn ] → k[G] is a surjection. Thus the distinct restrictions χ|G for χ ∈ X ∗ (Tn ) form a basis
for k[G]. It also follows that the restriction map X ∗ (Tn ) → X ∗ (G) is surjective. Hence we
have a surjection of groups Zn  X ∗ (G) so X ∗ (G) is abelian of finite type.
(2) =⇒ (3) Let ϕ : G → GL(V ) be a finite dimensional rational representation of G.
For each x ∈ G, ϕ(x) has the form ϕ(x) = (aij (x)) ∈ GLn (k), where aij ∈ k[G]. Write
X
ϕ(x) = χ(x)Aχ (∗)
χ∈X ∗ (G)

for linear operators Aχ : V → V . (This is possible since by hypothesis, the characters form
a basis of k[G].) Since ϕ is a homomorphism, for any x, y ∈ G we have
  
X X X X
ϕ(xy) = ψ(xy)Aψ =  θ(x)Aθ   µ(y)Aµ  = θ(x)µ(y)Aθ Aµ .
ψ∈X ∗ (G) θ∈X ∗ (G) µ∈X ∗ (G) θ,µ

Now ψ × ψ is a character of G × G, so we can rewrite the above expression as


X X
ϕ(xy) = (ψ × ψ)(x, y)Aψ = (θ × µ)(x, y)Aθ Aµ .
ψ∈X ∗ (G) θ,µ

By linear independence, we must have Aθ Aµ = δθ,µ Aθ where δθ,µ is the Kronecker delta. Now
if x = 1, ϕ(1) = idV and the expression (∗) becomes
X
idV = Aχ (∗∗)
χ∈X ∗ (G)

Thus the nonzero Aχ form a complete system of orthogonal idempotent linear operators on
V . If Aχ 6= 0, set Vχ = Aχ (V ). Then (∗∗) implies that we can decompose V into these
subspaces: M
V = Vχ .
χ∈X ∗ (G)

It remains to show the Vχ are one-dimensional. Notice that G acts on each Vχ via
X
x · v = ϕ(x)v = θ(x)Aθ (v) for x ∈ G, v ∈ Vχ
θ∈X ∗ (G)
X
= θ(x)Aθ Aχ (v 0 )
θ∈X ∗ (G)

= χ(x)Aχ (v 0 ) = χ(x)v.
Therefore G acts by the character χ on Vχ , so it follows that each Vχ is one-dimensional.
(3) =⇒ (1) Finally, by the work above we have a basis of eigenvectors for the image
of G in GL(V ), so by linear algebra G is diagonalizable in the linear sense. Then G ⊆ Tn
under this basis.

493
21.3. Diagonalizable Groups Chapter 21. Classifying Algebraic Groups

Definition. The character of a rational G-module V is


n
X
ch V = (dim Vi )e(χi )
i=1
Ln
where V = i=1 Vi and χi are as above.

Corollary 21.3.5. Let G be a diagonalizable group over a field k of characteristic p > 0.


Then X ∗ (G) is an abelian group of finite type with no p-torsion.

Proof. Suppose χ ∈ X ∗ (G) satisfies χp = 1. Then for any x ∈ G, 1 = χp (x) = χ(x)p which
implies χ(x) = 1, so χ is identically the operator 1. Hence the p-torsion part of X ∗ (G) is
trivial.
We can push this characterization further. Let M be an abelian group of finite type with
no p-torsion, where char k = p > 0. Define the k-algebra k[M ] having basis elements e(m)
for all m ∈ M ; these are formal exponentiation symbols, which distinguish the operation
e(m)e(m0 ) = e(m+m0 ) in k[M ] from the addition operation in M itself. Define multiplication
in this way, e(m)e(m0 ) = e(m+m0 ) and extend linearly. One can put a Hopf algebra structure
on k[M ] by defining

∆ : k[M ] −→ k[M ] ⊗ k[M ]


e(m) 7−→ e(m) ⊗ e(m);
γ : k[M ] −→ k[M ]
e(m) 7−→ e(−m);
and ε : k[M ] −→ k
e(m) 7−→ 1,

and extending by linearity.

Theorem 21.3.6. k[M ] is an affine algebra which is the coordinate algebra for a diagonal-
izable algebraic group G(M ) for which k[G(M )] = k[M ] as algebras and M ∼
= X ∗ (G(M )) as
abelian groups.

Corollary 21.3.7. There is an equivalence of categories between diagonalizable algebraic


groups and abelian groups of finite type with no p-torsion.

Proposition 21.3.8. G(M ) is a torus if and only if M is a free abelian group.

Proposition 21.3.9. Let M be a finitely generated abelian group with submodules {Mi }ni=1 .
Then the following are equivalent:

(1) M = ni=1 Mi .
L

(2) G(M ) = ni=1 G(Mi ).


Q

(3) k[M ] = nI=1 k[Mi ].


N

494
21.3. Diagonalizable Groups Chapter 21. Classifying Algebraic Groups

Lemma 21.3.10. Suppose G is an affine alglebraic group acting on a variety X. If G is


connected and X is a finite set then gx = x for all g ∈ G, x ∈ X. That is, G fixes X
pointwise.
Proof. Write X = {x1 , . . . , xn } and fix one of these elements, call it x. Consider the mor-
phism of varieties ϕ : G → X, g 7→ gx. Then ϕ(G) is irreducible, but since X is discrete,
this means ϕ(G) = {x}. Hence the action of G fixes X pointwise.
Theorem 21.3.11 (Rigidity). Let α : V × H → H 0 be a morphism of varieties such that:
(i) H 0 is an algebraic group and for each m > 0, there are finitely many elements of H 0
of order m.
(ii) H is an algebraic group whose set of finite order elements is dense.
(iii) V is a connected variety and for every x ∈ V , the map αx : H → H 0 , h 7→ α(x, h) is
a morphism of algebraic groups.
Then the map V → Hom(H, H 0 ), x 7→ αv is constant.
Proof. For each h ∈ H, let βh : V → H 0 be defined by βh (x) = α(x, h). For each h ∈ H of
finite order and each x ∈ V , α(x, h) = αx (h) is an element of H 0 with finite order dividing
|h|. Thus βh (x) = α(x, h) has finite order, so for fixed h, βh has finite image, but since
V is connected, βh must be constant. Now for x, y ∈ V , let γ : H → H 0 be the map
h 7→ αx (h)αy (h)−1 . Then
γ(h) = α(x, h)α(y, h)−1 = βh (x)βy−1 = e
since βh is constant. It follows that γ is trivial on a dense subset of H, but since γ is
continuous, this extends to a trivial map on all of H. Hence αx = αy , so the map x 7→ αx is
constant.
Corollary 21.3.12. Let G be a connected affine algebraic group acting on a diagonalizable
group D. Then the action is trivial.
Proof. For any integer n ≥ 1, set Dn = {x ∈ D | xn = e} (not to be confused with the
dihedral group). Since D is diagonalizable, D ∼ = G×`
m × F for a finite group F . Then every
x ∈ D can be written x = (x1 , . . . , x` , f ) for xi ∈ Gm , f ∈ F . In particular, if x ∈ Dn
then xni = 1 and f n =S1. Thus Dn is a finite set which is closed in D, and thus a discrete
affine subvariety, and ∞
n=1 Dn is a dense subset of D so by Theorem 21.3.11, the morphism
D × G → D corresponds to a trivial action of G.
Corollary 21.3.13. Suppose H ≤ G is a closed subgroup of a diagonalizable group G. Then
NG (H)0 = ZG (H)0 , where these denote, respectively, the connected component of the identity
in the normalizer of H and the centralizer of H.
Proof. Apply Theorem 21.3.11 to the morphism
NG (H)0 × H −→ H
(x, h) 7−→ xhx−1
to see that the map x 7→ (h 7→ xhx−1 ) is constant for every h ∈ H.

495
21.4. Connected Solvable Algebraic Groups Chapter 21. Classifying Algebraic Groups

21.4 Connected Solvable Algebraic Groups


Let G be an affine algebraic group with coordinate algebra A and Lie algebra g. Recall that
g is defined as

g = {D ∈ Derk (A) | D(f · g) = D(f ) · g for all f ∈ A, g ∈ G}.

Every g ∈ G has a Jordan decomposition g = gs gu = gu gs . Moreover, any embedding


ϕ
G− → GL(V ) = GLn (k) preserves this decomposition: ϕ(g) = ϕ(g)s ϕ(g)u = ϕ(gs )ϕ(gu ). It
turns out that there is a Jordan decomposition for g as well, which allows us to generalize
the results of Section 20.4.
Proposition 21.4.1. Let G be an affine algebraic group and let g be its Lie algebra. Given
X ∈ g, there exist elements Xs , Xn ∈ g such that

(1) X = Xs + Xn .

(2) [Xs , Xn ] = 0, i.e. Xs and Xn commute.

(3) If ϕ : G → GLn (k) is an embedding with differential dϕ : g → gln (k), then dϕ(Xs ) =
dϕ(X)s and dϕ(Xn ) = dϕ(X)n .

(4) If G ≤ GLn (k) and g ⊆ gln (k) then X = Xs + Xn is the usual Jordan decomposition
from linear algebra.
Lemma 21.4.2. Let T = G×n m be an n-dimensional torus. Then T consists entirely of
semisimple elements, that is Ts = T . If H ≥ T is a closed, connected subgroup then H = G×r
m
for some r ≤ n, that is, every closed, connected subgroup of a torus is also a torus.

The following is an analogue of Lie’s Theorem for solvable Lie algebras.


Theorem 21.4.3 (Lie-Kolchin). Let G ≤ GLn (k) be a closed, connected, solvable linear
algebraic group. Then there exists x ∈ GLn (k) such that xGx−1 ⊆ Bn , the Borel subgroup
of upper triangular matrices in GLn (k). In other words, every connected solvable linear
algebraic group may be simultaneously upper triangularized.
Proof. It suffices to show that there exists a nonzero vector v ∈ V = k n such that v is an
eigenvector for all x ∈ G. Let d be the length of the derived series of G. We induct on n + d,
noting that in the case n = 1 or d = 1, the proof is trival. For n, d ≥ 2, suppose there is a
nontrivial, proper, G-invariant subspace W ⊂ V . Then G is contained in the set of matrices
of the form  
W ∗
0 V /W
Therefore by induction on dimension, there exists a common eigenvector for the G-action on
W and on V /W , which is thus a common eigenvector for the entire G-action on V . Hence
we may assume V is irreducible. Since d ≥ 2, [G, G] 6= {1} is a subgroup of G with strictly
smaller derived series length and it will follows from Corollary 21.4.5 that [G, G] is closed,
connected and solvable. Therefore by induction, [G, G] has a common eigenvector v ∈ V . Let

496
21.4. Connected Solvable Algebraic Groups Chapter 21. Classifying Algebraic Groups

W be the span of all common eigenvectors for [G, G]. Then for all g ∈ G, x ∈ [G, G], v ∈ W ,
say with xv = λv for some λ ∈ k, we have

x(gv) = g(g −1 xgv) = g(λv) = λgv

since g −1 xg ∈ [G, G]. Hence W is a G-submodule of V , so by irreducibility, V = W . This


proves V has a basis of common eigenvectors for [G, G], i.e. every x ∈ [G, G] is a diagonal
matrix with respect to this basis, so [G, G] is abelian. Further, for any x ∈ [G, G], the
conjugates gxg −1 over all g ∈ G act diagonally on V relative to this basis and have the
same characteristic polynomial as x. Thus there can be only finitely many conjugates of x,
but since the map G → {gxg −1 | g ∈ G} sending g 7→ gxg −1 is a surjective morphism of
varieties and G is connected, there is only one conjugate, namely x itself. Thus x ∈ Z(G),
but at the same time x ∈ [G, G] so viewed as a matrix, det x = 1. That is, x is a diagonal
matrix with roots of unity along the diagonal, so it commutes with the G-action. By Schur’s
Lemma, [G, G] then consists of scalar matrices of determinant 1 and the scalars are roots of
unity, so there are only finitely many elements in [G, G]. But since [G, G] is connected, this
contradicts [G, G] 6= {1}.
For an alternate proof, consider the action of G on the projective variety Pn−1 (k) = P(V ).
By Theorem 12.11.2, P(V ) is complete, so by Borel’s fixed point theorem (21.2.9), there exists
a point L ∈ P(V ) such that xL = L for all x ∈ G. Then L is a one-dimensional subspace of
V spanned by some vector v 6= 0, and Gv = v so v is an eigenvector for every element in G.
This implies the result.
The next theorem generalizes the theory of Jordan decomposition and Theorem 20.4.14
for commutative algebraic groups.

Corollary 21.4.4 (Structure Theorem for Connected Nilpotent Groups). Assume G is con-
nected and nilpotent. Let Gs be the subset of all semisimple elements in G and let Gu be the
closed subvariety of all unipotent elements of G. Then

(1) Gs and Gu are closed subgroups of G.

(2) Gs ∼
= Gm × · · · × Gm , that is, Gs is a torus.

(3) There is an isomorphism Gs × Gu → G, (s, u) 7→ su.

Proof. We first show that Gs ⊆ Z(G). Take s ∈ Gs and define a map

χ : G −→ G
x 7−→ sxs−1 x−1 .

Let σ : G → G, x 7→ sxs−1 , so that χ(x) = σ(x)x−1 . Since G is nilpotent, there is an integer


n ≥ 1 such that χn (x) = e = 1G for all x ∈ G. Moreover, χ is a morphism of algebraic
groups so it has a differential at the identity:

de χ : g −→ g
X 7−→ ad(s)(X) − X.

497
21.4. Connected Solvable Algebraic Groups Chapter 21. Classifying Algebraic Groups

Since χn = 1, we see that ad(s) − I is nilpotent, but s is semisimple so ad(s) − I is also


semisimple. This can only happen if ad(s) − I = 0, or ad(s) = I, i.e. conjugation by s is
trivial g = Te G. This implies s ∈ Z(G), so Gs ⊆ Z(G). In particular, Gs is a subgroup of G.
Now, since G is nilpotent, it is solvable so we can apply the Lie-Kolchin Theorem (21.4.3).
Let V = k n and view G ≤ GLn (k). Since Gs consists of commuting semisimple elements,
the elements can be simultaneously diagonalized. Explicitly, there exist homomorphisms
χ1 , . . . , χr : Gs → G×
m such that

r
M
V = Vχi where Vχi = {v ∈ V | sv = χi (s)v for all s ∈ Gs }.
i=1

(One can think of the Vχi as generalized eigenspaces with respect to the χi .) Now each Vχi
is a G-submodule of V , so by the Lie-Kolchin Theorem (21.4.3), we may simultaneously
triangularize G so that xGx−1 ⊆ Bn , the subgroup of upper triangular matrices in GLn (k).
Then with respect to this choice of basis, Gs consists of diagonal matrices in GLn (k). In
particular, Gs is a closed subgroup of G. The proof is similar for Gu , so we have proven (1).
Now by Jordan decomposition, the map

Gs × Gu −→ G
(s, u) 7−→ su

is a bijection. Since Gs , Gu are closed subgroups and varieties, this is in fact an isomorphism
of groups and varieties, so (2) holds.
Finally, since G is connected, Gs and Gu are connected. We proved in part (1) that Gs
consists of diagonal matrices in GLn (k) so by Lemma 21.4.2, Gs is isomorphic to a torus.
With the next result, we complete our description of solvable algebraic groups.

Corollary 21.4.5. Assume G is connected and solvable. Then

(1) The commutator subgroup G0 = [G, G] is a closed, connected and unipotent subgroup
of G.

(2) Gu is a closed, connected, normal subgroup of G.

(3) G/Gu is isomorphic to a torus.

Proof. (1) For each x ∈ G, set [G, x] = {gxg −1 x−1 | g ∈ G}. This is a closed, irreducible
subvariety of G containing the identity. The commutator G0 = [G, G] is generated by these
sets [G, x] as a subgroup of G, so G0 is also closed and connected.
(2) View G ≤ Bn by the Lie-Kolchin Theorem (21.4.3). By matrix multiplication, G0 ⊆
Un , the subgroup of Bn with 1’s along the diagonal, so in particular G0 consists of unipotent
matrices. This means G0 ⊆ Gu so Gu is normal. Moreover, Gu = G ∩ Un so Gu is closed and
connected.
(3) By (3) of Corollary 21.4.4, G/Gu ∼= Gs so G/Gu consists of semisimple elements, that
0
is, (G/Gu )s = G/Gu . Since G ⊆ Gu by part (2), G/Gu is abelian. This means G/Gu is
solvable so by (2) of Corollary 21.4.4, G/Gu is isomorphic to a torus.

498
21.4. Connected Solvable Algebraic Groups Chapter 21. Classifying Algebraic Groups

Example 21.4.6. If G is a connected algebraic group of dimension 1, then G is solvable


and G = Gm or G = Ga .

Proposition 21.4.7. If G is a connected algebraic group of dimension dim G ≤ 2, then G


is solvable.

Proof. Let B ≤ G be a Borel subgroup. If B = G, we are done, so assume B 6= G. If


B = {1}, then G = G/B is complete so by Proposition 12.11.1, G must be finite, hence
trivial. The remaining case is when dim B = 1. Since B is solvable, [B, B] 6= B so [B, B] =
{1} and thus B is abelian. Fix any nontrivial g ∈ ZG (B) and consider the map ϕ : G →
G, x 7→ gxg −1 x−1 . Then by the argument in the proof of Corollary 21.2.13, ϕ is constant, so
g ∈ Z(G). This shows that {1} ( ZG (B) ⊆ Z(G), so G/Z(G) has dimension at most 1 and
is thus solvable. Hence G too is solvable.
Given s ∈ G consider the morphism

σs : G −→ G
x 7−→ sxs−1 .

Then σs has a differential, dσs = ad(s) : g → g. For example, if G = GLn (k) then
ad(s) = sXs−1 , the adjoint action of G on g. Let ZG (s) = {g ∈ G | gs = sg} be the
centralizer of s. This is a closed subgroup, so we may ask what the Lie algebra is for this
algebraic group.

Lemma 21.4.8. Let s ∈ G.

(1) If char k = 0 then the Lie algebra of ZG (s) is Zg (ad(s)).

(2) If char k = p > 0 and s is semisimple, then the same holds. In this case, we simply
write the Lie algebra of ZG (s) as Zg (s).

As in the Lie algebra case, we have

Definition. If T ≤ G is a maximal torus in an algebraic group G, then ZG (T )0 is called a


Cartan subgroup of G.

It turns out that every Cartan subgroup of G is nilpotent. Let T ≤ G be a maximal


torus and set C = ZG (T )0 . If G is semisimple, then by Corollary 21.3.13, C = T ( NG (T )
and NG (T )0 = T . In particular, NG (T )/T is a finite group. We will prove that NG (T )/T is
isomorphic to the Weyl group of G, to be defined in Section 21.5.

Definition. For an algebraic group G, we define its central series to be {Cn }∞


n=0 , where
C0 = G and for each n ≥ 1, Cn+1 = [Cn , G].

Definition. The derived series of an algebraic group G is {Dn }∞


n=0 , where D0 = G and
for each n ≥ 1, Dn+1 = [Dn , Dn ].

that when G is connected, all the subgroups Cn and Dn are closed and connected.
Recall T
Set C∞ = ∞ n=0 Cn . Since Cn ⊇ Cn+1 for all n, there is some n ≥ 0 such that C∞ = Cn .

499
21.4. Connected Solvable Algebraic Groups Chapter 21. Classifying Algebraic Groups

Lemma 21.4.9. G is nilpotent if and only if C∞ = {1}.


Lemma 21.4.10. G is nilpotent if and only if G = T × Gu , where T is a torus and Gu is
the closed subgroup of unipotent elements of G.
Theorem 21.4.11. Let G be a connected solvable group. Then
(1) If s ∈ G is semisimple, there exists a maximal torus T ≤ G such that s ∈ T .
(2) If T and T 0 are maximal tori of G, there exists an element g ∈ C∞ such that gT g −1 =
T 0.
Corollary 21.4.12. In any connected affine algebraic group G, any two maximal tori are
conjugate.
Proof. Let T and T 0 be two maximal tori in G. Then there is a Borel subgroup B ≤ G such
that T and T 0 act on G/B. Since G/B is complete, Borel’s fixed point theorem says that
T fixes some coset gB. In other words, g −1 T g ⊆ B for some g ∈ G. Likewise, h−1 T 0 h ⊆ B
for some h ∈ G, so by (2) of Theorem 21.4.11, g −1 T g and h−1 T 0 h are conjugate in B. This
implies the result.
Theorem 21.4.13. Let B be a Borel subgroup in a connected algebraic group G. Then
NG (B) = B.
Proof. We induct on the dimension of G. Let H = NG (B), x ∈ H and T ≤ B a maximal
torus. Since all maximal tori are conjugate by Corollary 21.4.12, we may assume xT x−1 = T .
Define a map
ψ : T −→ T
t 7−→ [x, t] = xtx−1 t−1 .
Since xtx−1 ∈ T for all t ∈ T , ψ is a group homomorphism, so its image is a closed,
connected subgroup of T by Theorem 20.2.4(2). In particular, im ψ is a torus. Now if
dim(im ψ) < dim T , then ker ψ 6= {1} so S = (ker ψ)0 is also a torus. Then x ∈ ZG (S), so x
normalizes ZG (S) ∩ B which is a Borel subgroup of ZG (S). If ZG (S) ( G, then by induction
x ∈ ZG (S) ∩ B and in particular x ∈ B. Otherwise, we have ZG (S) = G and S ⊆ Z(G),
so B/S is a Borel subgroup of G/S which is normalized by xS. Once again by induction,
xS ∈ B/S so x ∈ B.
It remains to analyze what happens when dim(im ψ) = dim T . In this case, im ψ = T
since T is connected. By Theorem 20.5.7, there is a rational G-representation ϕ : G →
GL(V ) and a nonzero vector v ∈ V such that
H = {x ∈ G | ϕ(x)v ∈ kv} and h = {X ∈ g | dϕ(X)v ∈ kv}.
Now ϕ(Bu ) fixes v, but T = im ψ is contained in [H, H], which implies im ψ fixes v as well.
By Corollary 21.4.5, B ∼
= T × Bu so we have shown that B fixes v. This induces a morphism
G/B −→ V
gB 7−→ gv.
Since G/B is complete and V is affine, this map must be constant, i.e. gv = v for all g ∈ G.
Hence G = B, so H = NG (G) = G = B.

500
21.4. Connected Solvable Algebraic Groups Chapter 21. Classifying Algebraic Groups

Corollary 21.4.14. Let P be a parabolic subgroup in a connected algebraic group G. Then


P is connected and NG (P ) = P .

Proof. Let B ≤ P be a Borel subgroup. Then B is connected so B ⊆ P 0 C P and B 0 is


a Borel subgroup of P 0 . Also, for all y ∈ NG (P ), yBy −1 ⊆ P 0 . Since B and yBy −1 are
both Borel subgoups of P 0 , then there is some x ∈ P 0 such that xyBy −1 x−1 = B. Therefore
xy ∈ NG (B) = B ⊆ P 0 by Theorem 21.4.13, which implies NG (P ) ⊆ P 0 ⊆ P ⊆ NG (P ).
Hence P 0 = P and NG (P ) = P .

Corollary 21.4.15. Let P, Q ≤ G be parabolic subgroups which are conjugate in G and


suppose B ⊆ P ∩ Q for some Borel subgroup B ≤ G. Then P = Q.

Proof. Let P = yQy −1 for some y ∈ G. Then B and yBy −1 are both Borel subgroups of
P , so there exists x ∈ P such that B = xyBy −1 x−1 . Thus xy ∈ NG (B) = B ⊆ P by
Theorem 21.4.13, so y ∈ P and hence P = Q.

Corollary 21.4.16. Let G be an affine algebraic group and let B be the set of all Borel
subgroups of G. Then B is a projective G-space via the map

ϕ : G/B −→ B
xB 7−→ xBx−1

for any fixed B ∈ B.

Proposition 21.4.17. Let G be a connected algebraic group of dimension dim G ≤ 2. Then


G is solvable.

Proof. Let B ≤ G be a Borel subgroup. If B = G, we are done, so assume B 6= G. Also,


if B = {1}, then G = G/B is complete, but the only complete, connected, affine group is
trivial, in which case G is trivially solvable. The remaining case is when dim B = 1. Since
B itself is solvable, [B, B] is a proper subgroup of B, so [B, B] = {1}. Thus B is abelian
and ZG (B) 6= {1}. By the same argument as in the proof of Theorem 21.4.13, for any fixed
g ∈ ZG (B), the map

ϕ : G −→ G
x 7−→ gxg −1 x−1

is constant, so g ∈ Z(G). Thus {1} ( ZG (B) ⊆ Z(G) which implies dim G/Z(G) ≤ 1, so
G/Z(G) is solvable and hence so is G.

Proposition 21.4.18. If T ≤ G is a torus in a connected algebraic group G, then ZG (T ) is


connected.

501
21.5. Semisimple Algebraic Groups Chapter 21. Classifying Algebraic Groups

21.5 Semisimple Algebraic Groups


The goal of this section is to adapt the program of classifying complex semisimple Lie algebras
to semisimple algebraic groups. Recall the standard classification:

ˆ g is a complex semisimple Lie algebra

ˆ h is its Cartan subalgebra, which is a maximal torus

ˆ The Cartan decomposition: g = h ⊕ α∈Φ gα


L

ˆ The simple roots R ⊂ h∗

ˆ The sl2 subgroups hgα , g−α i ∼ = sl2 (C), each spanned by eα , hα , fα for α ∈ X, which
satisfy [hα , eα ] = 2eα , [hα , fα ] = −2fα , [eα , fα ] = hα

The classification of semisimple algebraic groups is modelled after this setup. Suppose
S ≤ G is a torus, ϕ : S → GL(V ) is a finite dimensional representation and χ ∈ X ∗ (S) is a
character. Then we define the submodule

Vχ = {v ∈ V | ϕ(s)v = χ(s)v for all s ∈ S}.

Then V decomposes into these eigenspaces:


M
V = Vχ .
χ∈X ∗ (S)

Definition. For a character χ such that Vχ 6= 0, we say χ is a weight of S in V and Vχ is


a weight space.

Definition. The character of the representation V is


X
ch V = dim Vχ e(χ) ∈ ZX ∗ (S)
χ∈X ∗ (S)

where e(χ) is the formal exponential of χ, satisfying e(χ)e(ψ) = e(χ + ψ).

Remark. In characteristic 0, ch V is determined by Weyl’s character formula. In positive


characteristic however, not much is known.

Definition. Let T be a maximal torus in a connected, affine algebraic group G. Then dim T
is called the rank of G.

Proposition 21.5.1. Rank is well-defined and an algebraic invariant of G.

Proof. For well-definedness, use the fact that any two tori in G are conjugate. Invariance is
obvious.

502
21.5. Semisimple Algebraic Groups Chapter 21. Classifying Algebraic Groups

Fix g ∈ G and consider the morphism

ϕg : G −→ G
x 7−→ gxg −1 .

This induces a differential on the Lie algebra of G:

dϕg = Ad(g) : g −→ g.

In this way g is a rational G-module. Moreover, for any maximal torus T ≤ G, this action
restricts to a T -action on g.

Lemma 21.5.2. There exists a unique maximum connected, normal, solvable subgroup of
every affine algebraic group G.

Proof. Uniqueness follows from the fact that two connected, normal, solvable subgroups
generate a larger connected, normal, solvable subgroup.

Definition. The unique maximum connected, normal, solvable subgroup of G is called the
solvable radical of G, denoted Rs (G). An algebraic group is semisimple if Rs (G) = {1}.

There is an analagous definition for unipotent subgroups; existence and uniqueness are
proven in a similar fashion.

Definition. The unique maximum, connected, normal, unipotent subgroup of G is called the
unipotent radical of G, denoted Ru (G). An algebraic group is reductive if Ru (G) = {1}.

Example 21.5.3. GLn (k) is not semisimple, since Rs (GLn (k)) consists of the upper trian-
gular scalar matrices:   
t
∗ 

 

 t
 

×
Rs (G) =  :t∈k .
 
..

 
0 .  

 
 t 

One can check that Ru (GLn (k)) = {I}, so GLn (k) is reductive.
On the other hand, SLn (k) is semisimple. To see this, suppose H C SLn (k) is a con-
nected, normal, solvable subgroup. By the Lie-Kolchin Theorem (21.4.3), H can be put in
upper triangular form, so assume H ⊆ Bn . Conjugating H by the matrix
 
0 1
.

X=
 
..


±1 0

makes H lower triangular, where the sign in the lower left is chosen to make det X = 1.
Then XHX −1 = H since H C SLn (k), and hence H ⊆ Tn , the maximal torus of SLn (k).
So H consists of scalar matrices, so |H| < ∞. Since H is connected, this can only occur if
H = {1}. Therefore SLn (k) is semisimple.

503
21.5. Semisimple Algebraic Groups Chapter 21. Classifying Algebraic Groups

Example 21.5.4. Consider G the group of block matrices


 
∗ ∗ ∗ ∗
 ∗ ∗ ∗ ∗ 
  ∈ GL4 (k).
0 ∗0 ∗ 


Then the solvable and unipotent radicals of G are seen to be
     

 t 0 ∗ ∗ 
 
 1 0 ∗ ∗ 

0 t ∗ ∗ 0 1 ∗ ∗ 
   
 : t ∈ k×  : t ∈ k× .

Rs (G) =  Ru (G) = 
0 10 ∗1 0 1 ∗ 
      
   
0 1
   

Remarks.
1 If G is reductive then G0 = [G, G] is semisimple.

2 If S = Z(G) is the center of a reductive group G, then S is a torus and there is an


exact sequence of algebraic groups

1 → K → G0 × S → G → 1

with K a finite group. In other words, G is an almost direct product. By this token,
every reductive group is an almost direct product of semisimple groups so it’s enough
to study semisimple groups to capture all of the theory.
Example 21.5.5. Let k be a field of characteristic p > 0 and suppose (n, p) = 1. If
G = GLn (k) with S = {tI : t ∈ k × } and G0 = SLn (k) then
  
 ξ
 0 

 ..
. :ξ =1 ∼
 n
SLn (k) ∩ S =  = Z/nZ.

 0 
ξ 

We will need a good definition of the projective special linear group P SL2 (k) in the next
section. If char k 6= 2, then we define P SL2 (k) = SL2 (k)/h−Ii. However, if char k = 2, then
the subgroup h−Ii is trivial, and therefore this construction does not work. Instead, recall
the coordinate algebra of SL2 (k):

k[SL2 (k)] = k[T11 , T12 , T21 , T22 ]/hT11 T22 − T12 T21 − 1i = k[tij ].

Lemma 21.5.6. k[SL2 (k)] is reduced.


Set A = k[SL2 (k)]. Then A is a Hopf algebra whose comultiplication and antipode maps
are defined on the tij in the following way:
2
X
∆(tij ) = tik ⊗ tkj
k=1
γ(t11 ) = t22 , γ(t12 ) = −t12 , γ(t21 ) = −t21 , γ(t22 ) = t11 .

504
21.5. Semisimple Algebraic Groups Chapter 21. Classifying Algebraic Groups

Define a subalgebra B of A by

B = k[tij tk` | 1 ≤ i, j, k, ` ≤ 2] ⊆ A.

Then ∆(B) ⊆ B ⊗ B ⊆ A ⊗ A, γ(B) ⊆ A and ε(B) ⊆ k, so B is clearly a Hopf subalgebra


of A.
Definition. The projective special linear group P SL2 (k) is the algebraic subgroup of
SL2 (k) whose coordinate algebra is the Hopf algebra B.
Note that the inclusion B ,→ A induces a surjective morphism π : SL2 (k) → P SL2 (k). If
char k 6= 2, it’s easy to see that SL2 (k)/P SL2 (k) ∼
= Z/2Z. In characteristic 2, π : SL2 (k) →
P SL2 (k) is an isomorphism of groups, but not of algebraic groups.
Another way to view P SL2 (k) is via the diagram

i π
A B SL2 (k) P SL2 (k)

X X ◦ i = π(X)

k k

Here SL2 (k) is the space of k-maps A → k and for each X ∈ SL2 (k), π(X) = X ◦ i : B → k.
Suppose ϕ : B → k is a k-homomorphism, i.e. ϕ ∈ P SL2 (k). Then ϕ(tij tk` ) ∈ k, so for
example ϕ(t211 ) = ϕ(t11 )2 = ξ11 ∈ k. In characteristic 2, the equation ϕ(t11 )2 − ξ11 has a
unique solution which allows us to define Φ(tij ) for all tij such that ϕ = Φ ◦ i. This is how
to exhibit the isomorphism π : SL2 (k) → P SL2 (k) in the characteristic 2 case.
Definition. An automorphism σ ∈ Aut(G) is called semisimple if the comorphism σ ∗ :
k[G] → k[G] is semisimple as an operator.
Note that any σ ∈ Aut(G) induces an automorphism of the Lie algebra g of G by taking
the differential.
Lemma 21.5.7. An automorphism σ ∈ Aut(G) is semisimple if and only if there is an
isomorphism of algebraic groups ϕ : G → G ≤ GLn (k) and a semisimple matrix s ∈ GLn (k)
such that for all x ∈ G, ϕ(σ(x)) = sϕ(x)s−1 .
Theorem 21.5.8. Let G be an affine algebraic group over a field k. If σ ∈ Aut(G) is
semisimple, then the Lie algebra of Gσ = {x ∈ G | σx = x} is gσ = {X ∈ g | dσ(X) = X}.
Moreover, if char k = 0 then the same holds for any σ ∈ Aut(G).
Example 21.5.9. Suppose char k = 2 and G = SL2 (k). Then the Lie algebra of G is

g = sl2 (k) = {x ∈ M2 (k) | tr(x) = 0}

as in the complex case. Since char k = 2, sl2 (k) has a basis


     
0 1 1 0 0 0
, , .
0 0 0 1 1 0

505
21.5. Semisimple Algebraic Groups Chapter 21. Classifying Algebraic Groups

 
1 1
Let σ be the automorphism of SL2 (k) given by conjugation by the matrix . Then
0 1
one can check that
      
1 b 1 0 0 1
Gσ = :b∈k , gσ = ,
0 1 0 1 0 0

and dim Gσ = 1 and dimk gσ = 2. So the Lie algebra of Gσ cannot be gσ . However, this does
not present a counterexample to Theorem 21.5.8 since σ is not a semisimple operator.

Corollary 21.5.10. Let D be a diagonalizable group acting on an affine group G and define

ZG (D) = {x ∈ G | dx = x for all d ∈ D} and Zg (D) = {X ∈ g | dX = X for all d ∈ D}.

Then the Lie algebra of ZG (D) is Zg (D).

Proof. Use induction on dim g and apply Theorem 21.5.8 to the automorphisms G → G, x 7→
dx for each fixed d ∈ D.

506
21.6. Root Data Chapter 21. Classifying Algebraic Groups

21.6 Root Data


Theorem 21.6.1. Let G be an affine algebraic group and fix a maximal torus T ≤ G. Regard
g as a T -module. Then g has the following decomposition:
M
g=h⊕ gα
α∈X ∗ (T )

where h is the Lie algebra of T and for each α ∈ X ∗ (T ),

gα = {X ∈ g | Ad(t)(X) = α(t)(X) for all t ∈ T }.

Definition. If gα 6= 0, α is called a root of T in G. The set of roots is denoted R.


Theorem 21.6.2. If gα 6= 0, i.e. α is a root of T , then dim gα = 1.
Lemma 21.6.3. Let S ≤ T be a torus. Then ZG (S) = ZG (T ) if and only if S is not
contained in ZG ((ker α)0 ) for any root α ∈ R.
Proof. Note that ZG (S) is connected by Proposition 21.4.18 and ZG (T ) ⊆ ZG (S). By the
above, we can decompose g as M
g = Zg (T ) ⊕ gα ,
α∈R

so ZG (T ) ( ZG (S) if and only if Ad(S) fixes a nonzero element of gα for some α ∈ R, or


equivalently, α(S) = 1. In other words, ZG (T ) ( ZG (S) if and only if S ⊆ (ker α)0 for some
α ∈ R.
For α ∈ R, let S = (ker α)0 ⊆ T . Then T /S ∼
= Gm , a one-dimensional torus. Define
Gα = ZG (S) = ZG (S)0 .
Lemma 21.6.4. For α ∈ R, the Borel subgroups of Gα are all of the form B ∩Gα for B ≤ G
a Borel subgroup of G containing S.
Lemma 21.6.5. Let G be an affine algebraic group. Then
(1) G is generated by the subgroups Gα , i.e. G = hGα iα∈R .
(2) If every Gα is solvable then G is solvable.
Proof. (1) If T is a maximal torus in G, then for any S = (ker α)0 , ZG (S) ⊇ T . Then the
Lie algebra of G decomposes as M
g = Zg (T ) ⊕ gα
α∈R
0
equal to the Lie algebra of Gα = ZG ((ker α)0 ).
and by Corollary 21.5.10, gα = Zg ((ker α) ) is L
Thus the subgroup H = hGα i has Lie algebra α∈R gα = g, so dim H = dim G and we must
have H = G.
(2) Assume each Gα is solvable, so that it is its own (unique) Borel subgroup. For any
Borel subgroup B ≤ G containing T , we know S = (ker α)0 ⊆ T , so ZG (S) ⊆ B. Since this
holds for all α ∈ R, we get hGα i ⊆ B. Then by (1), B = G so G is solvable.

507
21.6. Root Data Chapter 21. Classifying Algebraic Groups

Let R0 be the set of roots α ∈ R such that Gα is not solvable. Also set Wα =
NG (S)/ZG (S). This will be an analogue of the Weyl group. Note that Wα ⊆ Aut(Gm ) ∼
=
Z/2Z, so |W | ≤ 2.
Theorem 21.6.6. If α ∈ R0 then Wα ∼ = Z/2Z.
To prove Theorem 21.6.6, assume T ≤ G is a torus and T ⊆ GL(V ) for a rational
G-module V . Choose a basis {v1 , . . . , vn } for V consisting of T -eigenvectors, so that T ⊆
diag(v1 , . . . , vn ) ⊆ GL(V ). Then for each t ∈ T and vi , we have tvi = γi (t)vi for some
γi (t) ∈ Gm . Thus γi ∈ X ∗ (T ), and in fact, X ∗ (T ) = hγ1 , . . . , γn i. For any cocharacter
λ : Gm → T , set mi = hλ, γi i ∈ Z for each 1 ≤ i ≤ n. The composition
λ γi
Gm → − T −
→ Gm
is given by γi ◦ λ(z) = z mi for all z ∈ Gm . Now if v = ni=1 ai vi is a nonzero vector in V , we
P
have n
X
λ(z)v = ai z mi vi .
i=1

On the other hand, Gm also acts on P(V ) via z[v] = [zv], so we can extend this to
" n #
X
λ(z)[v] = [λ(z)v] = ai z m i v i .
i=1
0
Let m0 and m be respectively the minimum and maximum of m1 , . . . , mn . Define
I = {1 ≤ i ≤ n | ai 6= 0}, I0 = {1 ≤ i ≤ n | mi = m0 }, I 0 = {1 ≤ i ≤ n | mi = m0 }.
Consider the map ϕ : Gm → P(V ) given by z 7→ λ(z)[v]. We extend ϕ to a map on P1k
as follows. Let P1k = U0 ∪ U1 where
U0 = {[x, y] : y 6= 0} and U1 = {[x, y] : x 6= 0}.
Viewing Gm = U0 ∩ U1 ⊂ P1k , we need only define ϕ for 0 = [0, 1] and ∞ = [1, 0]. In fact,
there are two ways to extend ϕ:
" #
X
ϕ0 (z) = ai z mi −m0 vi
i∈I
" #
−m0
X
and ϕ1 (z) = ai z m i vi .
i∈I

Then ϕ0 extends ϕ to U0 since each mi − m0 ≥ 0. Likewise ϕ1 extends ϕ to U1 since each


mi − m0 ≤ 0. Moreover, ϕ0 = ϕ1 on U0 ∩ U1 since
" # " # " #
X X X
ai z mi −m0 vi = z −m0 ai z m i v i = ai z m i v i
i∈I i∈I i∈I
" # " #
0 0
X X
= z −m ai z m i v i = ai z mi −m vi .
i∈I i∈I

508
21.6. Root Data Chapter 21. Classifying Algebraic Groups

P 
Call ϕ the extension by ϕ0 and ϕ1 to all of P1k . Notice that ϕ(0)[v] = i∈I 0
a i v i and
P 
ϕ(∞)[v] = i∈I 0 ai vi are both fixed poitns for the Gm -action on P(V ). Also notice that
ϕ(0)[v] = ϕ(∞)[v] is equivalent to m0 = m0 , in which case all the mi are equal and hence [v]
is itself a fixed point of the Gm -action. So we may choose λ ∈ X∗ (T ) in the first place so that
mi 6= mj for all i 6= j, so that ϕ(0) 6= ϕ(∞). Since h·, ·i : X ∗ (T ) × X∗ (T ) → Z is a perfect
pairing, this shows that T and λ(Gm ) ⊆ T have the same eigenspaces in V . Theorem 21.6.6
follows from the next three results.

Lemma 21.6.7. Let W ⊆ V be a vector subspace of codimension 1 and let X be a closed,


irreducible subvariety of P(V ) which is not contained in P(W ). If dim X ≥ 1, then X ∩
P(W ) 6= ∅ and each irreducible component of X ∩ P(W ) has codimension 1 in X.

Proof. Suppose X ∩ P(W ) = ∅. If dim V = n, then P(V ) r P(W ) ∼ = An−1


k and X embeds
into P(V ) r P(W ), but X is projective, hence complete, so this embedding is a constant
map. Thus dim X = 0. The last statement of the lemma follows from dimension theory.

Theorem 21.6.8. Let T be a torus in GL(V ) acting on P(V ) and let X ⊆ P(V ) be a closed,
irreducible subvariety stabilized by the action of T . If X does not lie in a hyperplane in P(V ),
then

(1) If dim X ≥ 1, then T fixes at least two points of X.

(2) If dim X ≥ 2, then T fixes at least three points of X.

Proof. (1) Assume dim X ≥ 1. If T fixes all of X, the statement holds trivially. Otherwise,
there is some [v] ∈ X which is not fixed by T . But by the above, for a suitable choice of
λ ∈ X∗ (T ), λ(0)[v] and λ(∞)[v] are distinct and fixed by the action of T .
(2) Apply Lemma 21.6.7.

Corollary 21.6.9. Let P ≤ G be a parabolic subgroup and T a maximal torus in G. Then


T fixes at least 2 points in G/P . In particular, T lies in at least 2 distinct Borel subgroups
of G.

Proof. By Theorem 20.3.6, we may view G ⊆ GL(V ) so that G/P is an orbit of a point in
P(V ). Apply Theorem 21.6.8 to this orbit to get at least 2 fixed points.
This completes the proof of Theorem 21.6.6, since if α ∈ R0 , Gα has a proper Borel
subgroup B, so dim Gα /B ≥ 1 implies |Wα | = |NGα (T )/ZGα (T )| ≥ 2.
To classify semisimple algebraic groups, we next generalize the notion of a root system
for a complex Lie algebra.

Definition. A root datum is a 4-tuple Ψ = (X, R, X ∨ , R∨ ) consisting of free abelian groups


X and X ∨ of finite rank with a pairing

X × X ∨ −→ Z, (χ, λ) 7−→ hχ, λi

and sets R and R∨ , called the roots and coroots of Ψ, respectively, which satisfy

(a) R ⊆ X, R∨ ⊆ X ∨ and there is a bijection R ↔ R∨ , α 7→ α∨ .

509
21.6. Root Data Chapter 21. Classifying Algebraic Groups

(b) For each α ∈ R, there are maps sα : X → X and s∨α : X ∨ → X ∨ such that sα (x) =
x − hx, α∨ iα and s∨α (y) = y − hα, yiα∨ for all x ∈ X, y ∈ X ∨ .

(c) For all α ∈ R, hα, α∨ i = 2.

(d) For all α ∈ R, sα (R) = R and s∨α (R∨ ) = R∨ .


Lemma 21.6.10. For any α ∈ R, s2α = 1 and (s∨α )2 = 1.
Proof. Let x ∈ X. Then

s2α (x) = sα (x − hx, α∨ iα)


= x − hx, α∨ iα − hx − hx, α∨ iα, α∨ iα
= x − hx, α∨ iα − hx, α∨ iα + hx, α∨ ihα, α∨ iα
= x − 2hx, α∨ iα + 2hx, α∨ iα = x.

The other identity is proven similarly.


Definition. The set W = hsα | α ∈ Ri is a subgroup of the automorphism group of X, called
the Weyl group of Ψ. We also define a subset W ∨ = hs∨α | α ∈ Ri ⊆ Aut(X ∨ ). Finally,
Q = ZR ⊆ X is called the root lattice of Ψ.
As in the Lie algebra setting, there is a set R+ ⊂ R consisting of the positive roots of X,
and R− ⊂ R consisting of the negative roots. One can choose x ∈ X ∨ such that hα, xi = 6 0
for all α ∈ R. Then R+ may be defined as the set of α ∈ R such that hα, xi > 0. Likewise,
R− is the set of roots α ∈ R such that hα, xi < 0. Then R = R+ ∪ R− .
Definition. S = {indecomposable roots α ∈ R+ } is called the set of simple roots of R.
Theorem 21.6.11. Let Ψ be a root datum with R its root system and S the set of simple
roots in R. For α, β ∈ S with sα =
6 sβ , let m(α, β) denote the order of sα sβ in W , which is
finite. Then
(1) If α ∈ S then s2α = 1.

(2) For all α, β ∈ S such that sα 6= sβ , (sα sβ )m(α,β) = 1.

(3) These two relations give a presentation of W . In particular, W is a Coxeter group.


This theorem is the starting place for an axiomatic formulation of the theory of semisimple
algebraic groups. We next show that this captures our description of semisimple groups in
Section 21.5. Let G be a semisimple algebraic group, fix a maximal torus T ≤ G and let B
be a Borel subgroup of G containing T . Let g, b and h be the Lie algebras of G, B and T ,
respectively. We can decompose g as a direct sum
M
g=h⊕ gα
α∈X ∗ (T )

where gα = {X ∈ g | Ad(t)(X) = α(t)(X) for all t ∈ T }. Let Φ be the set of roots of T ,


i.e. those α ∈ X = X ∗ (T ) for which gα 6= 0. Then Φ ⊆ X. Further, we can associate

510
21.6. Root Data Chapter 21. Classifying Algebraic Groups

X = Hom(T, Gm ) ∼ = Zr , where r = dim T , so Φ is a sublattice of X. Let Q ≤ X be the


subgroup generated by Φ. Since ZG (T ) = T , it’s easy to see that Q ∼ = Zr , that is, Φ is full
rank in X.
Now let X ∨ = X∗ (T ) = Hom(Gm , T ). If T = Grm then for any χ ∈ X ∨ , χ(t) =
(ta1 , ta2 , . . . , tar ) for some a1 , a2 , . . . , ar ∈ Z. This establishes an isomorphism X ∨ ∼ = Zr , so
the characters and cocharacters of T are dual lattices. This means there is a bilinear pairing

X × X ∨ −→ Z
(α, χ) 7−→ hα, χi := m where α ◦ χ(t) = tm .

Explicitly, X ∨ = Hom(X, Z).

Definition. X = X ∗ (T ) is called the weight lattice of G and X ∨ is called the coweight


lattice of G.

Given a root α ∈ Φ, there is a homomorphism ϕ : SL2 (k) → G such that ϕ(SL2 (k)) = Gα
is isomorphic to either SL2 (k) or P SL2 (k). Also, T normalizes Gα so the Lie algebra of Gα
may be written
h̄ ⊕ gα ⊕ g−α
for h̄ a one-dimensional subalgebra of h. This mimics the classical decomposition of sl2 (k)
in the Lie algebra setting. Consider the map
 
t 0
ν : Gm −→ SL2 (k), t 7−→ .
0 t−1

Then α∨ = ϕ ◦ ν : Gm → T is a cocharacter satisfying hα, α∨ i = 2. Define Q∨ to be the


subgroup of X ∨ generated by all α∨ for α ∈ Φ. Then Q∨ ⊆ X ∨ is a sublattice.

Definition. We say G is simply connected if Q∨ = X ∨ .

Definition. The Weyl group of G is the group W = NG (T )/ZG (T ), where T is a maximal


torus in G.

Note that by Proposition 21.4.18, ZG (T ) = ZG (T )0 and by Corollary 21.3.13, ZG (T )0 =


NG (T )0 , so W = NG (T )/NG (T )0 is in fact a finite group. As in Corollary 21.4.16, let B be
the set of all Borel subgroups in G. Then for a fixed Borel subgroup B ≤ G containing T ,
the map

W = NG (T )/ZG (T ) −→ B T := {C ∈ B | T ⊆ C}
x̄ 7−→ xBx−1

is a bijection. This allows us to study the Weyl group in terms of the algebraic geometry of
the projective variety B.

Theorem 21.6.12. For a connected algebraic group G over an algebraically closed field k,
with maximal torus T , Borel subgroup B and Weyl group W = NG (T )/ZG (T ), the following
are equivalent:

511
21.6. Root Data Chapter 21. Classifying Algebraic Groups

(a) The rank of G/Rs (G) is 1.


(b) |W | = 2.
(c) |B T | = 2.
(d) dim G/B = 1.
(e) G/B ∼
= P1k as projective varieties.
(f ) There is a surjective morphism of algebraic groups ϕ : G → P GL2 (k) with finite
kernel.
Proof. For simplicity, assume G is semisimple, so that Rs (G) = {1} and T ∼ = Gm . The
reduction to this case is done in Malle-Testerman. Also, note that the equality |W | = |B T |
follows from the paragraph above, so (b) ⇐⇒ (c) is done.
(a) =⇒ (b) Assume G is not solvable, so it has a proper Borel subgroup {1} ( B ( G.
Then dim G/B ≥ 1 so Corollary 21.6.9 implies |W | = |B T | ≥ 2. In general, we saw that
|W | ≤ 2 so we have |W | = 2.
(c) =⇒ (d) If dim G/B ≥ 2, then Theorem 21.6.8 implies |B T | ≥ 3.
(d) =⇒ (e) In the proof of Theorem 21.6.6, we constructed a surjective morphism
ϕ : P1 → G/B by defining ϕ0 and ϕ1 on the affine patches U0 , U1 and extending over 0
and ∞. Moreover, a consequence of Corollary 17.7.3 is that the existence of a dominant
morphism P1 → X for any X implies X ∼ = P1 .
(e) =⇒ (f) The natural action of G on G/B induces a morphism
ϕ : G −→ Aut(G/B) ∼
= Aut(P1 ).
Note that ker ϕ = B 0 ∈B B 0 is finite, so im ϕ = G/ ker ϕ has dimension at least 3. We will
T
show (Theorem 22.0.1) that Aut(P1 ) ∼ = P GL2 (k) which has dimension equal to 3, so in fact
ϕ must be surjective.
(f) =⇒ (a) Since T is a torus in G, ϕ(T ) is a torus in P GL2 (k), so ϕ(T ) has rank 1,
i.e. ϕ(T ) ∼
= Gm . Since ker ϕ is finite, ϕ|T : T → Gm is an isomorphism and we are done.
Corollary 21.6.13. Let G be a connected, semisimple algebraic group of rank 1. Then G is
isomorphic to either SL2 (k) or P SL2 (k).
As in the classical case, for each α ∈ Φ we can define an element sα ∈ W by sα (χ) =
χ − hχ, α∨ iα. Then a presentation of the Weyl group is W = hsα | α ∈ Φi.
For α, β ∈ Φ, notice that hα, β ∨ i ∈ Z. Define E = X ⊗Z R ∼ = Rr . Then W acts on E as
follows. For w ∈ W , choose a coset representative nw ∈ NG (T ). Then for χ ∈ X ∗ (T ), we
define w(χ) ∈ X ∗ (T ) by w(χ)(t) = χ(n−1 w tnw ) for all t ∈ T . This is well-defined on W by
construction, so we will just write w tw for n−1
−1
w tnw in the future. This defines an action of
W on X (T ) = Z , so extend by scalars to get an action on E = X ∗ (T ) ⊗Z R ∼
∗ r
= Rr . If (·, ·)
is a positive definite symmetric bilinear form on E, there is an associated bilinear form on E
defined by X
hx, yi = (w(x), w(y)).
w∈W

It follows immediately that h·, ·i is also positive definite and W -invariant.

512
21.6. Root Data Chapter 21. Classifying Algebraic Groups

Lemma 21.6.14. Let W = W (G, T ) = NG (T )/ZG (T ) be the Weyl group of G. Then

(a) W acts transitively on R and R0 .

(b) If S ≤ T is any torus, then W (ZG (S), T ) is a subgroup of W (G, T ).

(c) When S = (ker α)0 for a root α ∈ R, Wα = W (Gα , T ) = W (Gα /S, T /S).

Theorem 21.6.15. Let G be a connected semisimple algebraic group with maximal torus T ,
Borel subgroup B, Weyl group W = NG (T )/ZG (T ) and root system Φ. If E = X ⊗Z R, then

(1) E admits an inner product with respect to which W acts by orthogonal linear trans-
formations. In particular, W is a finite Coxeter group.

(2) Φ is a classical root system in the Euclidean space E and there is an embedding X ∨ ,→
E given by

α∨ 7−→ .
hα, αi

(3) For each α ∈ Φ, there is an isomorphism



u α : Ga −
→ Uα ⊆ G

onto a closed subgroup Uα ≤ G which is stable under the action of T . In addition,


Lie(Uα ) = gα .

(4) For any ordered subset Γ ⊆ Φ+ which is closed under the inner product, multiplication
defines an isomorphism

Y
A|Γ| = Uα −→ UΓ := hUα | α ∈ Γi.
α∈Γ

Further, Ru (B) = UΦ+ ∼


Q
= α∈Φ+ Uα .
Definition. The subgroups Uα for α ∈ Φ are called one-parameter subgroups of G.

Since Φ is a root system, we may define a set T = {α1 , . . . , αr } of simple roots, and
subsets Φ+ , Φ− ⊂ Φ of positive and negative roots, with respect to the bilinear form h·, ·i.

Definition. A set of fundamental dominant weights of G consists of elements $1 , . . . , $r ∈


X such that for each 1 ≤ i, j ≤ r, h$i , αj∨ i = δij .

Theorem 21.6.16. G is simply connected if and only if its weight lattice X contains a set
of fundamental dominant weights.

Example 21.6.17. For G = SLn (k), there are characters


 
t1 0
εi : T −→ Gm , 
 ...  7−→ ti .

0 tn

513
21.6. Root Data Chapter 21. Classifying Algebraic Groups

Then ε1 , . . . , εn form an orthonormal basis of E. A root system for G is Φ = {εi − εj | i 6= j}


and a set of simple roots is Π = {ε1 − ε2 , ε2 − ε3 , . . . , εn−1 − εn }. Note that for each root
α = εi − εj , gα = heij i, where eij is the standard ijth matrix unit. A set of fundamental
dominant weights for SLn (k) is

$1 = ε1 , $2 = ε1 ε2 , . . . , $n = ε1 · · · εn .

Therefore by Theorem 21.6.16, SLn (k) is simply connected.

The following is the fundamental theorem in the classification of semisimple algebraic


groups.

Theorem 21.6.18. Let G and G0 be two simply connected, semisimple algebraic groups over
k with the same root system. Then G ∼ = G0 as algebraic groups. Conversely, given a classical
root system Φ, there exists a simply connected, semisimple algebraic group with Φ as its root
system.

This is an analogue to Theorem 21.6.18 in the Lie algebra case. For an arbitrary (meaning
not necessarily simply connected) semisimple algebraic group, we have the following.

Corollary 21.6.19. A semisimple algebraic group G is uniquely determined by its root datum
(X, Φ, X ∨ , Φ∨ ). Moreover, if G
b is a simply connected, semisimple algebraic group with root
system Φ then G is a homomorphic image of G b with finite kernel.

Definition. Let Φ be an irreducible root system and set Q = ZΦ and Ω = Hom(Q∨ , Z) ⊆


Hom(X ∨ , Z) = X. Then the quotient group Λ(Φ) = Ω/Q is called the fundamental group
of Φ.

Theorem 21.6.20. For every irreducible root system Φ, Λ(Φ) is a finite group. Moreover,
there is a bijective correspondence

{root data with root system Φ} ←→ {subgroups H ≤ Γ(Φ)}.

Example 21.6.21. The irreducible root systems are classified by their Dynkin diagrams,
which are labelled by their type: infinite families An , Bn , Cn , Dn (which splits into the even
and odd cases) and exceptional types G2 , F4 , E6 , E7 , E8 . For an irreducible root system
Φ of one of these types, there is a semisimple algebraic group Gsc corresponding to the
trivial subgroup {1} ⊆ Γ(Φ) – the notation is suggestive, as Gsc is simply connected in the
sense of our definition above. Likewise, there is a group GAd corresponding to the whole
fundamental group Γ(Φ). Below are listed some data about the irreducible root systems and
their corresponding semisimple algebraic groups.

514
21.6. Root Data Chapter 21. Classifying Algebraic Groups

Φ W Γ(Φ) Gsc GAd intermediates


An−1 Sn Z/nZ SLn (k) P SLn (k) SLn (k)/(Z/d), d | n
Bn Sn o (Z/2Z) Z/2Z Spin(2n + 1) SO(2n + 1) n/a
Cn Sn o (Z/2Z) Z/2Z SP2n (k) P CSp2n (k) n/a
D2n+1 Sn o (Z/2Z)n−1 Z/4Z Spin(4n + 2) P CO0 (4n + 2) SO(4n + 2)
D2n Sn o (Z/2Z)n−1 Z/2Z × Z/2Z Spin(4n) P CO0 (4n) SO(4n), two more
G2 D6 {1} G2 G2 n/a
F4 order 27 · 32 {1} F4 F4 n/a
E6 51840 Z/3Z E6,sc E6,Ad n/a
E7 2903040 Z/2Z E7,sc E7,Ad n/a
E8 696729600 {1} E8 E8 n/a

Let Φ be an irreducible root system and take Π = {α1 , . . . , αr } to be a set of simple roots.
Note that there are many choices of Π, but if Π0 is another set of simple roots, there is a unique
element w ∈ W such that w(Π) = Π0 . We have the positive roots Φ+ = (Nα1 ⊕· · ·⊕Nαr )∩Φ
and the negative roots Φ− = −Φ+ , so that Φ = Φ+ ∪ Φ− . Notice that −Π = {−α1 , . . . , −αr }
is also a set of simple roots, so by the above, there is a unique element w0 ∈ W such that
w0 (Π) = −Π.

Proposition 21.6.22. Given the length function ` : W → Z, w0 is the unique element of


maximum length, i.e. `(w0 ) > `(w) for any w ∈ W, w 6= w0 . Moreover, for each w ∈ W ,
`(w) = #{α ∈ Φ+ | w−1 (α) ∈ Φ− } and `(w0 ) = |Φ+ |.

Definition. Let I ⊆ Π be a subset of simple roots and define the subgroup LI = hT, Uα iα∈±I
of G, where T is a maximal torus and the Uα are one-parameter subgroups. Then LI is called
a Levi subgroup, or Levi factor, corresponding to I.

Theorem 21.6.23. For a semisimple algebraic group G with maximal torus T , Borel sub-
group B and irreducible root system Φ containing simple roots Π,

(1) The subgroups PI := hLI , Bi, where LI is the Levi factor corresponding to some subset
I ⊆ Π, are parabolic subgroups of G. Moreover, the correspondence I ↔ PI is bijective.

(2) Any parabolic subgroup of G is conjugate to a unique PI .

(3) For each I ⊆ Φ, PI ∼


= LI o Ru (PI ).
There is an analogue of the Bruhat decomposition for semisimple algebraic groups. Let
W be the Weyl group of G. If w ∈ W and B ⊇ T is a Borel subgroup of G, then let
nw ∈ NG (T ) be a representative of w ∈ W . Observe that nw B and Bnw are independent of
the choice of nw , so we can write nw B = wB and Bnw = Bw.

Theorem 21.6.24 (Bruhat). Any semisimple algebraic group G may be written as a disjoint
union of double cosets [
G= BwB.
w∈W

515
21.6. Root Data Chapter 21. Classifying Algebraic Groups

Set Xw = BwB/B ⊆ G/B. Then X w is called a Schubert variety. It turns out that
each Xw is an affine subvariety of the complete variety G/B, with projective closure X w .
Specifically, Xw ∼
= A`(w) where `(w) is the length of w in the Weyl group.
We can also write G as a non-disjoint union
[
G= BwB.
w∈W

Since G is closed, BwB = G for some w ∈ W . Write B = U T , so that BwB = U T wB =


U ww−1 T wB = U wB. For each w ∈ W , set
Y Y
Uw0 = Uα and Uw = Uα .
α∈Φ+ α∈Φ+
w−1 (α)>0 w−1 (α)<0

Then U = Uw Uw0 so by the above work,

BwB = U wB = Uw Uw0 wB = Uw wB.

This proves the following result.

Theorem 21.6.25. For each w ∈ W , BwB ∼ = Uw × {w} × B as algebraic groups and


+
dim Uw = `(w), so dim BwB = `(w) + r + |Φ |.

Corollary 21.6.26. For the unique element w0 ∈ W of maximum length, Bw0 B = U w0 B


is open in G.

Definition. B − = w0 Bw0−1 is called the opposite Borel subgroup to B.

Corollary 21.6.27. Set U − = w0 U − w0−1 . Then U − B ∼


= U − × B is open and dense in G.

516
21.7. Representations of Algebraic Groups Chapter 21. Classifying Algebraic Groups

21.7 Representations of Algebraic Groups


In this section we sketch the representation theory of algebraic groups which mirrors the
representation theory of Lie algebras. Let k be an algebraically closed field and take an
affine algebraic group G over k.
Definition. A representation of G is a morphism of algebraic varieties ρ : G → GL(V )
where V is a rational G-module.
Unlike in the theory of finite group representations, much is unknown in the representa-
tion theory of algebraic groups. In particular, it is not known in general how to compute
the dimensions and characters of the irreducible G-representations.
Let G be semisimple and simply connected with maximal torus T , Borel subgroup B,
root system Φ with positive roots Φ+ ⊂ Φ, coroots Φ∨ and simple roots and coroots Π and
Π∨ . Recall from the Lie-Kolchin theorem (21.4.3) that if L is an irreducible G-module, then
B stabilizes a one-dimensional subspace kv ⊂ L, with b · v = ξ(b)v for some ξ ∈ X ∗ (B) =
Hom(B, Gm ). We write kv = Lξ and call ξ a dominant weight of L.
Definition. If G is an algebraic group having root system Φ, let gC be the complex simple
Lie algebra with root system Φ and define the distribution algebra
Dist(G) = U(gC )Z ⊗ k
where U(gC )Z is the Kostant Z-form of gC .
Theorem 21.7.1. For an algebraic group G with root system Φ, the category of G-representations
is equivalent to the category of locally finite Dist(G)-modules.
Theorem 21.7.2. The irreducible G-modules are indexed by the set of dominant weights
X+ := Z≥0 $1 ⊕ · · · ⊕ Z≥0 $r .
Proof. (Sketch) Let L be an irreducible G-module. Then for a maximal torus T ≤ G, L is
a T -module and the character of L as a T -module can be written
X
ch L = dim Lξ e(ξ),
ξ∈X ∗ (T )

where Lξ = {v ∈ L | t · v = ξ(t)v for all t ∈ T }. It is known that X ∗ (T ) has a partial


ordering: X
λ ≥ µ if λ − µ = nα α.
α∈Φ+
nα ≥0

There is a unique element λ ∈ X ∗ (T ) which is maximal with respect to this ordering, and
such that Lλ 6= 0. One can prove dim Lλ = 1 and there is a nonzero eigenvector vλ ∈ Lλ such
that L = kG · vλ . Moreover, λ is dominant, i.e. hλ, α∨ i ∈ Z≥0 . Following the Lie algebra
case, we call λ the highest weight of L. This determines an assignment
τ : Irr(G) −→ X+
L 7−→ λ.

517
21.7. Representations of Algebraic Groups Chapter 21. Classifying Algebraic Groups

Chevalley proved that the mapping is also onto, i.e. for every λ ∈ X+ there is an (irreducible)
G-module L(λ). Borel later gave the following elementary proof.
It’s enough to consider λ = $i a fundamental dominant weight. Let Qi be the subgroup
nj αj ∈ Φ+ , where ni = 0. Then each Qi is closed
P
of G generated by B and U−α for α =
and parabolic. By Theorem 21.1.5, there exists a finite dimensional G-module V containing
a one-dimensional subspace ` ⊆ V such that Qi = StabG (`). Take a basis {v1 , . . . , vn } of V
such that kv1 = `. Then each x ∈ Qi corresponds to a matrix of the form
 
∗ ∗ ··· ∗
0 ∗ · · · ∗
 
 .. .. . . .. 
. . . .
0 ∗ ··· ∗

Let ϕ : G → GL(V ) = GLn (k) be the representation of G corresponding to V . Write


ϕ(g) = (aij ) where aij = ϕ(g)ij ∈ A = k[G]. Then since Qi = StabG (`), a11 has weight
m$i for some constant m. Restrict a11 ∈ A to U − B. Then there exists f ∈ k(G) of weight
$i such that f m = a11 . But f m ∈ A and A is integrally closed in its fraction field, so f
is integral over A. Thus f ∈ A. Consider the G-module generated by f and factor out its
largest submodule. This gives an irreducible G-module L of highest weight $i . Hence τ is
onto and the theorem is proven.

518
Chapter 22

Finite Groups of Lie Type

In this chapter we study an important class of finite groups arising from linear algebraic
groups over a finite field F = Fq , where q = pm is a power of a prime p. First, we prove
an analogy of the fact from complex analysis that P GL2 (C) is equal to the automorphism
group of the Riemann sphere C ∪ {∞}.

Theorem 22.0.1. For all fields k, Aut(P1k ) ∼


= P GL2 (k).
Proof. Let V = A2 = k 2 and consider the projection defining P1 = P1k :

V r {0} −→ P1
(x, y) 7−→ [x, y].
   
1 x x
Viewing points in V and P as column vectors and , respectively, GL2 (k) acts on
y y
V r {0} by left multiplication and this action descends to an action on P1 , written
    
a b x ax + by
= .
c d y cx + dy

Notice that Gm ∼
= Z(GL2 (k)) acts trivially on P1 , so we get an induced action of P GL2 (k) =
GL2 (k)/Z(GL2 (k)) on P1 , which we write as
    
a b x ax + by
= .
c d y cx + dy
 
1 a b
Consider the points 0 = [0, 1], 1 = [1, 1] and ∞ = [1, 0] in P . For each g = , define
c d
the triple
ϕ(g) := (g(0), g(1), g(∞)) = ([b, d], [a + b, c + d], [a, c]).
This defines a map ϕ : P GL2 (k) → P1 ×P1 ×P1 which is clearly injective (look at the first and
third components in particular). In fact, P GL2 (k) acts doubly transitively on P1 ×P1 ×P1 via
this map, with trivial isotropy subgroup. Therefore, for any (x, y, z) ∈ P1 ×P1 ×P1 with x, y, z

519
Chapter 22. Finite Groups of Lie Type

all distinct, there exists an element g ∈ P GL2 (k) such that (g(x), g(y), g(z)) = (0, [u, v], ∞)
with [u, v] 6= 0, ∞. Thus neither of u, v is 0, so we can write
 −1    
u 0 u
0, , ∞ = (0, 1, ∞).
0 v −1 v

This shows that P GL2 (k) acts triply transitively on the middle factor of P1 . Thus the map

P GL2 (k) −→ Aut(P1 )


g 7−→ ([u, v] 7→ g · [u, v])

is injective. On the other hand, for σ ∈ Aut(P1 ), there exists an element g ∈ P GL2 (k) such
that (σ(0), σ(1), σ(∞)) = (g(0), g(1), g(∞)) by triple-transitivity, so τ = σ −1 g fixes ∞ and
restricts to an automorphism of A1 = P1 r {∞}. Since k[A1 ] = k[t], the only invertible
endomorphisms are linear, so τ |A1 has the form x 7→ ax + b for some a, b ∈ k. But we also
know τ (0) = 0 and τ (1) = 1, which imply b = 0 and a = 1, so τ = 1 after all. Hence σ = g,
so P GL2 (k) ∼= Aut(P1 ).

520
22.1. Jordan’s Theorem Chapter 22. Finite Groups of Lie Type

22.1 Jordan’s Theorem


In this section, we prove Jordan’s theorem, which shows the existence of an infinite family
of finite simple groups of Lie type coming from the classical groups GLn (Fq ) and SLn (Fq ).
Example 22.1.1. Let GLn (q) = GLn (Fq ). This is a finite group whose order is equal to
the number of ordered bases of an n-dimensional vector space over Fq . Explicitly,
|GLn (q)| = (q n − 1)(q n − q) · · · (q n − q n−1 )
= q 1+2+...+n−1 (q n − 1)(q n−1 − 1) · · · (q − 1)
n−1
Y
n(n−1)/2 n
=q (q − 1) (q i + q i−1 + . . . + q + 1).
i=1

This tells us that q n(n−1)/2 is the order of a Sylow p-subgroup of GLn (q), namely the nilpotent
subgroup  

 1 ∗ 

 ..
Un =  .  .


 0 
1 
On the other hand, (q − 1)n is the order of the maximal torus of GLn (q),
 

 ∗ 0 

 ..
Tn =  .  .

 
 0 ∗ 
The remaining term has an interpretation in terms of topology, as we will see later.
Example 22.1.2. For SLn (q) = SLn (Fq ), consider the short exact sequence of groups
1 → SLn (q) → GLn (q) → F× q → 1.

Then |SLn (q)| = |GLn (q)|/(q − 1) = q n(n−1)/2 (q − 1)n−1 n−1 i i−1


Q
i=1 (q + q + . . . + q + 1). Further,
if C denotes the center of SLn (q), then
  

 ζ 0 

 ..  n
C=  .  : ζ = 1 ,

 0 
ζ 

i.e. C consists of diagonal matrices diag(ζ, . . . , ζ) where ζ is an nth root of unity. There-
fore |C| = (n, q − 1), the greatest common divisor, so the projective special linear group
P SLn (q) = P SLn (Fq ) = SLn (Fq )/C has order
|SLn (q)| |GLn (q)|
|P SLn (q)| = = .
(n, q − 1) (q − 1)(n, q − 1)
For example, P SL2 (2) = SL2 (2) is a finite group of order 6 which turns out to be isomorphic
to S3 (since it is nonabelian). Also, P SL2 (3) is a nonsimple group of order 12 which is
isomorphic to A4 . In what can be regarded as the origin of the classification of finite simple
groups, the following important result by Jordan gives an infinite class of finite simple groups
different from the alternating groups.

521
22.1. Jordan’s Theorem Chapter 22. Finite Groups of Lie Type

Theorem 22.1.3 (Jordan). Let F be any field. For n > 2, or for n = 2 and |F | =
6 2, 3, the
group |P SLn (F )| is simple.
Consider the set of n × n matrices Mn (F ) with basis {eij }1≤i,j≤n consisting of elementary
matrices. For b ∈ F and i 6= j, define
 
1
Tij (b) = 1 + beij =  . . . 
 
b 1

where b appears in the ith row and the jth column. Then Tij (b)Tij (c) = Tij (b + c), so
Tij : b 7→ Tij (b) defines a group homomorphism F → Mn (F ). Left (resp. right) multiplication
by each Tij (b) defines a map Mn (F ) → Mn (F ) which adds b times the jth row (resp. column)
to the ith row (resp. column). Also, for i 6= j, define

Pij = 1 + eij + eji − eii − ejj = (1 + eij )(1 − eji )(1 + eij )(1 − 2eii ).

Then left (resp. right) multiplication by Pij defines a map Mn (F ) → Mn (F ) which switches
the ith and jth rows (resp. columns). Note that
 
1
 ... 
 
1 − 2eii =  −1
 


 . .. 

(with −1 in the ith entry on the diagonal) is a diagonal matrix.


Lemma 22.1.4. For any n ≥ 2, SLn (F ) is generated by {Tij (b) | 1 ≤ i, j ≤ n, i 6= j, b ∈ F }.
Proof. Note that each Tij (b) has determinant 1, so hTij (b)i ⊆ SLn (F ). If A ∈ SLn (F ), then
we may write P AQ = diag(d1 , . . . , dn ) for some P, Q which are products of various Tij (b)
and Pij . Notice that Pij = Tij (1)Tji (1)Tij (1)(1 − 2eii ) so we can assume P and Q are each a
product of some Tij (b) and 1 − 2eii . Now consider the following commutator relations:

Tk` (b)(1 − 2eii ) = (1 − 2eii )Tk` when k, ` 6= i


Tij (b)(1 − 2eii ) = (1 − 2eii )Tij (−b) for any i 6= j
Tji (b)(1 − 2eii ) = (1 − 2eii )Tji (−b) for any i 6= j.

Then P may be written with all the 1 − 2eii on the left and factors of Tij (b) otherwise (and
likewise for Q, with the 1 − 2eii on the right). Multiplying these factors across maintains
a diagonal matrix on the right, so we may assume P AQ = diag(d1 , . . . , dn ) where P and Q
are each a product of some Tij (b). Thus det(P ) = det(A) = det(Q) = 1 so d1 · · · dn = 1.
For d ∈ F × , there is a sequence of elementary matrix operations taking
 −1   
d 0 1 0
S(d) = to = T21 (d−1 ).
0 d d−1 1

522
22.1. Jordan’s Theorem Chapter 22. Finite Groups of Lie Type

In particular, S(d) is a product of the Tij (b). The same argument applies to
 
d−1 0 0 ··· 0
 0 d
 0 · · · 0
diag(d−1 , d, 1, . . . , 1) =  0 0
 1 · · · 0
 .. .. .. . . .. 
 . . . . .
0 0 0 ··· 1

for any number of 1’s on the diagonal. Next, our diagonal matrix diag(d1 , . . . , dn ) satisfies

diag(d−1
1 , d1 , 1, . . . , 1) diag(d1 , d2 , . . . , dn ) = diag(1, d1 d2 , d3 , . . . , dn ).

Repeating this, we get

diag(1, (d1 d2 )−1 , d1 d2 , 1, . . . , 1) diag(1, d1 d2 , d3 , . . . , dn ) = diag(1, 1, d1 d2 d3 , d4 , . . . , dn )

and eventually this reduces to diag(1, 1, 1, . . . , d1 d2 · · · dn ) on the right. Since d1 · · · dn = 1


and diag(1, 1, . . . , 1) = T12 (0), we are done because everything is now written as a product
of the Tij (b).

Lemma 22.1.5. For n > 2, or for n = 2 and |F | =


6 2, 3, SLn (F ) is equal to its commutator
0
subgroup SLn (F ) .

Proof. By Lemma 22.1.4, it suffices to show Tij (b) ∈ SLn (F )0 for every i 6= j and b ∈ F . For
n > 2, let 1 ≤ k ≤ n, k 6= i, j. Then for any b ∈ F , we can write

Tij (b) = Tik (b)Tkj (1)Tik (−b)Tkj (−1)

which is a commutator. Thus the lemma holds in this case. When n = 2, consider the
commutator      
d 0 1 c 1 c(d2 − 1)
, = .
0 d−1 0 1 0 1
Since F 6= F2 , F3 , there exists a d ∈ F with d2 6= 1. Taking c = (d2 − 1)−1 b, the above
identity gives us

1 (d2 − 1)−1 b
   
d 0
T12 (b) = , ∈ SLn (F )0 .
0 d−1 0 1

This completes the proof.


To prove Jordan’s theorem, we need the following group theory result due to Iwasawa in
the 1940s.

Theorem 22.1.6 (Iwasawa). Suppose G acts on a set S with kernel K = ker(G → Aut(S)).
Then G/K is simple provided the following conditions hold:

(1) G acts primitively on S, i.e. S cannot be written as a disjoint union of subsets which
are permuted transitively by G.

523
22.1. Jordan’s Theorem Chapter 22. Finite Groups of Lie Type

(2) G is equal to its commutator G0 .

(3) There exists an element x ∈ S such that the stabilizer Gx has a normal abelian sub-
group Ax and G is generated by its conjugates {gAx g −1 | g ∈ G}.

Example 22.1.7. Every 2-transitive group action is primitive. (Recall that a group action
is n-transitive if any subsetset of n elements can be taken to any other subset of n elements.)

Example 22.1.8. The natural action of G on itself by left (or right) multiplication is not
primitive since for any subgroup H ≤ G, the cosets gH are permuted transitively by G.

Proof of Jordan’s Theorem (22.1.3). Take G = SLn (F ) and let S = Pn−1 be projective
(n − 1)-space over F , i.e. the set of all lines through the origin in F n . Then G acts on S
with kernel C, the center of SLn (F ) so to prove that SLn (F ) is simple, we need only check
the conditions of Theorem 22.1.6. It is a standard fact that the action of SLn (F ) on lines
through the origin is 2-transitive, so by Example 22.1.7, the action is primitive. Further,
Lemma 22.1.5 showed that SLn (F ) = SLn (F )0 holds under our conditions on n and F , so
it remains to show condition (3) in Iwasawa’s theorem. Fix an ordered basis (e1 , . . . , en ) of
F n and let ` be the line spanned by e1 , so ` ∈ Pn−1 . The stabilizer of ` in SLn (F ) is
  

 ∗ ∗ · · · ∗ 

0 ∗ · · · ∗
 

G` =  .. .. . . ..  : ∗ ∈ F
 
 . .
 . . 


 0 ∗ ··· ∗ 

Consider the subgroup A` ≤ G` defined by


 
 1 b 0 
A` =  0 1  .

0 I 

 
1 b
Then A` is a normal abelian subgroup of the stabilizer and since T12 (b) = , the
0 1
conjugates of A` generate SLn (F ). Hence SLn (F )/C = P SLn (F ) is simple.

524
22.2. Restricted Lie Algebras Chapter 22. Finite Groups of Lie Type

22.2 Restricted Lie Algebras


This section is taken from a lecture on Lie algebras in characteristic p > 0 that I gave in Dr.
Parshall’s class in spring 2018.
The theory of semisimple Lie algebras over a field k of characteristic 0 is known. When
char k = p > 0 however, the theory is much richer and is still open, though much is under-
stood. For starters, an algebra over a field of finite characteristic can be given the structure of
a restricted algebra, which gives rise to new types of representation theory and classification
results. We first elaborate on our observations in Example 20.5.1.
Let k be a field of characterstic p > 0, let A be an associative k-algebra and consider the
vector space of k-derivations

Derk (A) = {D : A → A | D is k-linear and D(xy) = D(x)y + xD(y) for all f, g ∈ A}.

Then [x, y] = xy − yx endows Derk (A) with the structure of a Lie algebra. The Leibniz rule
D(xy) = D(x)y + xD(y) can be extended inductively to the following formula for all n ≥ 1:
n  
n
X n
D (xy) = Di (x)Dn−i (y).
i=0
i

When n = p, this formula becomes

Dp (xy) = Dp (x)y + xDp (y)

which shows that Dp ∈ Derk (A) as well. In other words, the Lie algebra Derk (A) is closed
under the p-mapping D 7→ Dp . The notion of a restricted Lie algebra generalizes this.

Definition. For a prime p and a field k of characteristic p, a p-restricted Lie algebra


(or just restricted Lie algebra is a Lie algebra L over k which admits a map [p] : L →
L, x 7→ x[p] , such that

(i) (λx)[p] = λp x[p] for all λ ∈ k, x ∈ L.

(ii) ad(x[p] ) = (ad(x))p for all x ∈ L, where ad(x)(−) = [x, −].

(iii) (Jacobson Formula) For all x, y ∈ L,


p−1
[p] [p] [p]
X fi (x, y)
(x + y) =x +y +
i=1
i

where fi (x, y) is the coefficient of ti−1 in the formal expression (ad(tx + y))p−1 (x).

Proposition 22.2.1. Let A be an associative k-algebra and L = A− the corresponding Lie


algebra, i.e. A equipped with bracket [x, y] = xy − yx. Then L is a restricted Lie algebra with
x[p] = xp .

525
22.2. Restricted Lie Algebras Chapter 22. Finite Groups of Lie Type

Proof. (i) is obvious.


(ii) We must show that ad(xp )(y) = (ad(x))p (y). One can prove by induction that in any
associative algebra and for any n ≥ 1,
n  
i n
X
n
(ad(x)) (y) = (−1) xn−i yxi . (*)
i=0
i
p

So when n = p, p divides i
for all 1 ≤ i ≤ p − 1, so we get

[x, y]p = xp y − yxp = [xp , y].

(iii) Consider the polynomial expression


p−1
X
p p p p
f (t) = (tx + y) = t x + y + si (x, y)ti ∈ L[t],
i=1

where si (x, y) is a polynomial in x, y of total degree p. Then


p−1
X
p p p
(x + y) = f (1) = x + y + si (x, y)
i=1

so it suffices to show isi (x, y) = fi (x, y). Taking the formal derivative of the expressions for
f (t) with respect to t, we get
p−1 p−1
X X
0 i p−i−1
f (t) = (tx + y) x(tx + y) = isi (x, y)ti−1 .
i=0 i=1

On the other hand, (∗) shows that f 0 (t) = (ad(tx + y))p−1 (x), so we have isi (x, y) = fi (x, y)
as desired.
Definition. Let L and L0 be restricted Lie algebras over k.
ˆ A [p]-restricted homomorphism is a Lie algebra homomorphism ϕ : L → L0 such
that ϕ(x[p] ) = ϕ(x)[p] for all x ∈ L.

ˆ A [p]-restricted ideal of L is a Lie ideal I ≤ L such that x[p] ∈ I for all x ∈ I. Note
that [p]-restricted ideals are precisely kernels of [p]-restricted homomorphisms.

ˆ Similarly, a [p]-restricted subalgebra is a Lie subalgebra M ≤ L that is closed under


x 7→ x[p] .

ˆ A [p]-restricted representation of L is a restricted Lie algebra homomorphism ρ :


L → gl(V ), where V is a k-vector space and gl(V ) is considered as an associative
algebra with the [p]-restricted structure from Proposition 22.2.1.
From now on, we will drop the “[p]-” and refer to everything as “restricted”.
Question. We will consider the following:

526
22.2. Restricted Lie Algebras Chapter 22. Finite Groups of Lie Type

(1) Does every restricted Lie algebra admit a finite dimensional, faithful, restricted repre-
sentation?
(2) When does an ordinary Lie algebra admit the structure of a restricted Lie algebra?
Starting with (2), notice that for L to be a restricted Lie algebra, we must have
(ad(x))p = ad(x[p] ) for all x ∈ L,
i.e. each (ad(x))p is an inner derivation on L. We will show that in the right circumstances
this condition is also sufficient for L to admit a restricted structure.
For an ordinary Lie alebra L with k-basis {ui }, recall the Poincaré-Birkhoff-Witt theorem:
Theorem 22.2.2 (PBW). The universal enveloping algebra U = U(L) has basis
{uki11 · · · ukirr | r ≥ 1, i1 < . . . < ir , kj ≥ 0}.
We will adapt this slightly to construct an analogue of the universal enveloping algebra
for restricted Lie algebras.
Theorem 22.2.3. Let L be a Lie algebra with basis {ui } such that every (ad(ui ))p is an
[p] [p]
inner derivation, say (ad(ui ))p = ad(ui ) for some ui ∈ L. Then there exists a unique
[p]-mapping L → L, x 7→ x[p] extending this definition and defining a p-restricted structure
on L.
[p]
Proof. Let U = U(L) be the universal enveloping algebra of L. Note that each zi := upi − ui
[p]
lies in the center of U since ad(zi ) = ad(ui )p − ad(ui ) = 0. Slightly adapting the proof of
the PBW theorem, one can prove that U has basis
{zih11 · · · zihrr uki11 · · · ukirr | r ≥ 1, i1 < . . . < ir , hj ≥ 0, 0 ≤ kj ≤ p − 1}.

Define the ideal I = hzi i ⊆ U and the quotient algebra U = U/I. Then the images of the
monomials uki11 · · · ukirr form a basis for U. Moreover, x 7→ x̄ := x + I ∈ U/I restricts to a
homomorphism L → L := (L + I)/I, but since the ūi = ui + I are linearly independent, this
mapping is an isomorphism of Lie algebras L ∼ = L.
Next, the associative algebra U is a restricted Lie algebra by Proposition 22.2.1 and by
construction L is a restricted subalgebra of U:
[p]
ūpi = (ui + I)p = upi + I p = upi + I = ui + I.
0
For uniqueness, suppose L admits another mapping x 7→ x[p] satisfying the axioms of a
0
restricted Lie algebra. Then x[p] − x[p] lies in the center of L for all x ∈ L. In other words,
0
the map f : x 7→ x[p] − x[p] is a p-semilinear mapping. Then ker f is a vector subspace of L,
[p] [p]0
so we need only observe that f is trivial on a basis of L: f (ui ) = ui − ui = upi − upi = 0.
Hence [p] = [p]0 .
Corollary 22.2.4. If L is a finite dimensional Lie algebra with nondegenerate Killing form,
then there is a unique p-restricted structure on L. Moreover, there is a restricted universal
enveloping algebra U as above which has dimension pdim L and is universal with respect to all
maps

527
22.2. Restricted Lie Algebras Chapter 22. Finite Groups of Lie Type

ψ
ϕ
L A

where A is any associative k-algebra, ϕ : L → A− is a restricted homomorphism with respect


to a[p] = ap on A, and ψ is a unique homomorphism of associative k-algebras.
Proof. Same argument as for the ordinary universal enveloping algebra.
Corollary 22.2.5. Every restricted Lie algebra admits a finite dimensional, faithful, re-
stricted representation.
Proof. A restricted representation is a Lie algebra homomorphism ρ : L → gl(V ) which is a
[p]-mapping, or equivalently, such that ρ(L) is a restricted subalgebra of gl(V ). Since U has
a finite dimensiona, faithful representation, so does L by Corollary 22.2.4.
Corollary 22.2.6. For any k-algebra A, Derk (A) is a p-restricted subalgebra of Endk (A)
with respect to D 7→ Dp . In particular, the Lie algebra of any algebraic group G over k is
restricted.
In light of this, we examine the [p]-restricted structure on the Lie algebras of two impor-
tant algebraic groups. Let k be an algebraically closed field of characterstic p.
Example 22.2.7. Let G = Ga be the additive group of k with Lie algebra ga = Lie(Ga ) =
k dtd . Then for all n ≥ 0,

 p
d
tn = n(n − 1) · · · (n − p + 1)tn−p = 0
dt

so D[p] = Dp = 0 for all D ∈ g. In other words, ga is a trivial restricted Lie algebra.


Example 22.2.8.

d Now let G = Gm , the multiplicative group of k, with Lie algebra gm =
Lie(Gm ) = k t dt . Then for all n ≥ 0,
 p
d
t tn = tn
dt

so in this case D[p] = Dp = D, i.e. D 7→ Dp is the identity on gm . This shows that gm and ga
are not isomorphic as restricted Lie algebras, even though they are isomorphic as abstract
Lie algebras over k.
Remark. Geometrically, the (restriction to k r {0} of the) map

Ga → Gm , x 7→ xn

may be viewed as a covering space (i.e. a local homeomorphism, or a locally trivial surjec-
tion) as long as char k does not divide n. In particular, this is always a covering space in

528
22.2. Restricted Lie Algebras Chapter 22. Finite Groups of Lie Type

characterstic 0. However, when char k = p | n, the above examples show that the tangent
space structure of Ga does not push forward to Gm under x 7→ xp . So this map cannot be a
local homeomorphism. It turns out that the correct analogue of the cyclic cover x 7→ xn in
characteristic p is the Artin-Schreier cover

Ga → Ga , x 7→ xp − x.

(See Part VI for the full theory of algebraic fundamental groups.) The general case of a
cover of degree n = pr m, (m, p) = 1 is given by Artin-Schreier-Witt theory.

The theory of algebraic groups in characteristic p is closely related to restricted Lie


algebras, as the following results show.

Theorem 22.2.9 (Demazure-Gabriel). For a field k of characteristic p > 0, there is a


one-to-one correspondence

{restricted Lie algebras over k} ←→ {finite group schemes over k of height 1}


g = Te G →−7 G
L 7−→ group scheme corresponding to U = U(L)/hxp − x[p] i.

(Here, a group scheme has height 1 if the Frobenius F : G → G is zero.)

Theorem 22.2.10 (Jacobson). Suppose k is a field of characteristic p > 0 such that [k :


k p ] < ∞. Then there is a one-to-one correspondence

{restricted Lie algebras over k} ←→ {inseparable subfields k ⊇ E ⊇ k p of exponent 1}


DerE (k) →−7 E
L 7−→ {x ∈ k | D(x) = 0 for all D ∈ L}.

The classical simple Lie algebras are completely classified by their Dynkin diagram: An ,
Bn , Cn , Dn , E6 , E7 , E8 , F4 , G2 . Each of these admits a model over Z (given by the
theory of Chevalley bases, which are related to Chevalley groups; see Section 22.4) which
therefore descends to a Lie algebra over Fp for any prime p. The reduction mod p of one
of the classical simple Lie algebras is semisimple in general, but by taking quotients, one
obtains all “classical” simple restricted Lie algebras. However, there are simple restricted
Lie algebras not arising from reduction mod p:

ˆ (Witt) W1 = Lie(k[x]/(xp )) for all p ≥ 5.

ˆ (Witt-Jacobson) More generally, Wn = Lie(k[x1 , . . . , xn ]/(xp1 , . . . , xpn )) for n ≥ 1.

ˆ (Special) Sn , defined as the derived algebra of the subalgebra of Wn which fixes a


certain volume form.

ˆ (Hamiltonian) Hn defined similarly, but for a Hamiltonian form.

ˆ (Contact) Kn , defined similarly, but for a contact form.

529
22.2. Restricted Lie Algebras Chapter 22. Finite Groups of Lie Type

Theorem 22.2.11 (Block-Wilson, formerly Kostrikin-Shafarevich Conjecture). Every re-


stricted simple Lie algebra over a field of characteristic p > 5 is classical or one of the
four types (Witt-Jacobson, special, Hamiltonian, contact) listed above. These are collectively
referred to as the Cartan algebras.

In small characteristic, there are known exceptions to the Block-Wilson classification:

ˆ In characteristic 5, there is only one exception (the Melikian algebra).

ˆ In characteristics 2 and 3, there are many exceptions.

530
22.3. The Frobenius Chapter 22. Finite Groups of Lie Type

22.3 The Frobenius


Let p be a prime, q = pr for some r ≥ 1 and k = Fq the algebraic closure of Fq , the finite field
with q elements. Take G to be an algebraic group over k and let A = k[G] = k[x1 , . . . , xn ]/I
be the coordinate algebra of G (so that I is a radical ideal in this polynomial ring).
Definition. For a field extension L/Fq contained in F q , we say G is defined over L if the
radical ideal I can be generated by polynomials in L[x1 , . . . , xn ].
Definition. An Fq -structure on G is an Fq -algebra A0 such that A ∼
= A0 ⊗Fq k as Hopf
algebras.
Note that G is defined over Fq if and only if G admits an Fq -structure A0 . If G is any
algebraic group over k, then by Corollary 22.2.6, its Lie algebra is a [p]-restricted Lie algebra.
Definition. For an algebraic group G over k, the Frobenius is the morphism F : G → G
[p]
induced by the comorphism F ∗ : A → A, xi 7→ xi , where [p] is the restricted structure on
the Lie algebra of G. If A0 is an Fq -structure on G, then the Frobenius can also be defined
by F ∗ : a0 ⊗ λ 7→ aq0 ⊗ λ for a0 ∈ A0 , λ ∈ k.
Lemma 22.3.1. The Frobenius F : G → G is a morphism of algebraic groups which is an
isomorphism of abstract groups but not of algebraic varieties.
This observation, while perhaps surprising, gives us two new objects to study: ker F
and GF = {g ∈ G | F (g) = g}. Notice that GF is a finite group (which will give us many
examples of finite simple groups of Lie type), while ker F is a group scheme, but not a variety.
Example 22.3.2. Let G = GLn (k). Then the Frobenius is F : (aij ) 7→ (aqij ) and

GLn (k)F = {(aij ) ∈ GLn (k) | (aqij ) = (aij ) for all 1 ≤ i, j ≤ n} = GLn (Fq ).

Thus we recover our first finite (though non-simple) group of Lie type, GLn (q).
Let A = k[GLn (k)] be the coordinate algebra of GLn (k) and let B be any commutative
k-algebra. Then setting GLn (B) := Homk (A, B) defines a functor

GLn : Algk −→ Group


B 7−→ GLn (B).

Likewise, ker F defines a functor

G1 : Algk −→ Group
B 7−→ ker(F : GLn (B) → GLn (B)),

called the first infinitesimal subgroup of GLn . Note that k[G1 ] is a finite dimensional, commu-
tative Hopf algebra over k. In fact, its dual k[G1 ]∗ turns out to be the restricted enveloping
algebra U = U(gln ) of the Lie algebra gln = Lie(GLn (k)), in the sense of Section 22.2.
For any algebraic group G over k, Theorem 20.3.6 says that G is isomorphic to a closed
subgroup of GLn (k) for some n ≥ 1.

531
22.3. The Frobenius Chapter 22. Finite Groups of Lie Type

Lemma 22.3.3. If G ≤ GLn (k) is a closed subgroup, then the Frobenius F : G → G is


equal to the restriction to G of the standard Frobenius F : GLn (k) → GLn (k) given by
(aij ) 7→ (aqij ).
Definition. For m ≥ 1, the mth infinitesimal subgroup of G is Gm = ker(F m ). These
subgroups form a filtration of group schemes: G1 ⊆ G2 ⊆ G3 ⊆ · · · ⊆ G.
To obtain other finite groups of Lie type, we have the following generalization of the
Frobenius.
Definition. Let G be an algebraic group over k = Fq . An endomorphism σ : G → G is
called a Steinberg endomorphism if σ r = F is the Frobenius for some r ≥ 1.
Note that for any Steinberg endomorphism σ of G, the fixed points form a finite group:

Gσ ⊆ GF ⊆ GLn (k)F = GLn (q d ) (for some n, d ≥ 1).

Example 22.3.4. When G = GLn (k), the map σ : (aij ) 7→ ((aqij )−1 )T satisfies σ 2 (aij ) =
2
(aqij ), so σ 2 = F : GLn (q 2 ) → GLn (q 2 ). Thus σ is Steinberg. The group GLn (k)σ ≤ GLn (q 2 )
is called the general unitary group, written GUn (q). Explicitly,

GUn (q) = {A ∈ GLn (q 2 ) | A fixes h·, ·i}

where h·, ·i is the inner product defined by

h·, ·i : Fnq2 × Fnq2 −→ Fq2


n
X
(x, y) = (x1 , . . . , xn , y1 , . . . , yn ) 7−→ hx, yi := xi yiq .
i=1

Also, the subgroup SUn (q) := GUn (q) ∩ SLn (k) is called the special unitary group.
Example 22.3.5. When G = Gm is the multiplicative group, all endomorphisms are of the
form

σm : Gm −→ Gm
γ 7−→ γ m , for some m ∈ Z.

When m = q, σq = F is the Frobenius on F× q , so we see that the Steinberg endomorphisms


σ
are precisely the Frobenius maps for the powers of p. Here, Gmq = F× q is a finite group of
order q − 1.
Example 22.3.6. When G = Ga is the additive group, a map

σh : Ga −→ Ga
x 7−→ h(x)

is an endomorphism precisely when h(x) = f (xp ) for some f ∈ k[x]. Then for σhr = F , we
i
must have h(x) = xp for some i ≥ 1. Setting q = pi , σxq is a Steinberg endomorphism, but
the fixed points are just Gσaxq = Fq , so we don’t obtain any new groups this way.

532
22.3. The Frobenius Chapter 22. Finite Groups of Lie Type

Theorem 22.3.7 (Steinberg). Let G be a simple algebraic group over k = Fq . For any
endomorphism σ : G → G, exactly one of the following holds:

(1) σ is an automorphism.

(2) σ is a Steinberg endomorphism and thus Gσ is a finite group.

One can construct Steinberg endomorphisms for all of the classical semisimple groups
over Fq (in the sense of Chevalley’s classification; see Theorem 21.6.18) except for types
B2 , C2 , G2 and F4 in small characteristic. This realizes “most” of the finite simple groups,
but misses types B2 = C2 in characteristic p = 2 (the Ree group), G2 in characteristic p = 3
(the Suzuki group) and F4 in characteristic p = 2 (the Tits group).

Theorem 22.3.8 (Lang-Steinberg). Let G be a connected algebraic group over k = Fq and


let σ : G → G be a Steinberg endomorphism. Then the map L : G → G, g 7→ σ(g)g −1 is
surjective.

Proof. We give the proof when σ = F is the Frobenius, which is due to Lang. The general-
ization by Steinberg isn’t too much harder.
Let g be the Lie algebra of G. For L : G → G, g 7→ F (g)g −1 , the differential acts on g is
dL = d(F × (·)−1 ) = dF × (·)−1 − 1 = −1, so dL is a Lie algebra isomorphism. This means
L(G) is dense in G. Similarly, for x ∈ G, we may define Lx : G → G by L(g) = F (g)xg −1 ,
and since Tx G ∼
= g as k-vector spaces, dLx : g → Tx G is also an isomorphism. Thus Lx (G) is
dense in G, so by Chevalley’s theorem (19.1.1), Lx (G) ∩ L(G) 6= ∅. Take y ∈ Lx (G) ∩ L(G).
Then F (g1 )xg1−1 = y = F (g2 )g2−1 for some g1 , g2 ∈ G, so

x = F (g1−1 g2 )g2−1 g1 = F (w)w−1

for w = g1−1 g2 . Thus L is surjective.

Corollary 22.3.9. Let G be a connected reductive group and σ : G → G a Steinberg endo-


morphism. Then there exists a Borel subgroup B ≤ G such that F (B) ⊆ B and a maximal
torus T ≤ B such that σ(T ) ⊆ T . Moreover, all such pairs (B, T ) are Gσ -conjugate.

Proof. Let V = {(B, T ) | B ≤ G is a Borel subgroup and T ≤ B is a maximal torus}.


Then G acts transitively on V by g(B, T ) = (gBg −1 , gT g −1 ). Further, note that the Frobe-
nius commutes with this action: F (g(B, T )) = F (g)F (B, T ). Take v = (B, T ) ∈ V and
choose g ∈ G such that gv = F (v). By the Lang-Steinberg theorem (22.3.8), g = σ(x)x−1
for some x ∈ G, so σ(x)x−1 v = F (v). Set x−1 v = (B 0 , T 0 ) ∈ V . Then

(B 0 , T 0 ) = x−1 v = F (x−1 v) = (F (B 0 ), F (T 0 )).

This shows that F (B 0 ) = B 0 and F (T 0 ) = T 0 , so B 0 and T 0 work for the first statement of
the corollary. The second statement also follows from the above computation.
This allows one to construct Steinberg endomorphisms, and thus finite groups of Lie type,
and descend them to simple subgroups. To study representations of such groups, we have
the following construction, which we won’t discuss in detail.

533
22.3. The Frobenius Chapter 22. Finite Groups of Lie Type

Definition. Let ρ : G → GL(V ) be a representation of an algebraic group G and let F :


G → G be the Frobenius of G. Then the Frobenius twist of ρ is the representation

ρ ◦ F : G −→ GL(V ).

We denote the corresponding G-module by V [p] .

This allows one to extend the representation theory of Section 21.7 to characteristic p.
For example, recall that for a semisimple algebraic group G, the irreducible G-modules are
in one-to-one correspondence with the set of dominant weights λ ∈ X ∗ (T ) = Hom(T, Gm ),
where T is a maximal torus in G.

Definition. Let q = pr . A dominant weight λ is called q-restricted if the element ρ =


1
P
2 α∈Φ+ α satisfies
0 ≤ hλ + ρ, αi∨ i ≤ q
for all αi∨ ∈ Π∨ .

Note that any weight ξ ∈ X ∗ (T ) can be written ξ = ξ0 + pξ1 + p2 ξ2 + . . . + pr ξr where ξi


is p-restricted.

Theorem 22.3.10 (Steinberg). Let ξ ∈ X ∗ (T ) and write ξ = ξ0 + pξ1 + p2 ξ2 + . . . + pr ξr as


above. Then
Lξ ∼
[p] [p] [p]
= Lξ0 ⊗ Lξ1 ⊗ Lξ2 ⊗ · · · ⊗ Lξr .
Moreover, if ξ itself is p-restricted, then Lξ is infinitesimally simple as a G-module.

Example 22.3.11. For G = SLn (k) with highest weight module L = Lλ , if λ is q-restricted
then Lλ |SLn (Fq ) gives an irreducible SLn (Fq )-module.

More generally, if Lξ is simple when restricted to the fixed points of the Frobenius, GF ,
then ξ is q-restricted. The converse also holds when q = p but not in general.

534
22.4. Chevalley Groups Chapter 22. Finite Groups of Lie Type

22.4 Chevalley Groups


In this section, which follows a presentation for Dr. Parshall’s spring 2018 class given by
Chris Leonard, we discuss how to associate to any complex semisimple Lie algebra g a family
of algebraic groups G for various fields k having g as their Lie algebras. For example, when
k = R or C, G will be a Lie group; for algebraically closed k, G will be a semisimple,
connected algebraic group (and in fact all such algebraic groups arise this way); and for k a
finite field, G will be a finite simple group of Lie type (and most such finite groups arise this
way).
Let g be a complex semisimple Lie algebra. Fix a finite dimensional, faithful representa-
tion ρ : g → gl(V ). Notice that for any X ∈ g, the object

ρ(X)2 ρ(X)3 X ρ(X)n
exp(ρ(X)) = 1 + ρ(X) + + + ... =
2 3! n=0
n!

makes sense in GL(V ). So taking G = hexp(ρ(X)) | X ∈ gi gives us a subgroup of GL(V ).


However, to make this construction work over any field k, we must first make sense of g, ρ
and V over any k.
Let U = U(g) be the universal enveloping algebra of g and for each root α ∈ Φ, fix a
nonzero element Xα ∈ gα . The Kostant Z-form of g is defined by

UZ = ZhXα(n) | α ∈ Φ, n ≥ 0i
(n) n
where Xα = Xn!α . Then UZ is a free abelian group and UZ ⊗ C = U. Thus for any finite
dimensional, faithful g-module V , there exists a free abelian Z-submodule VZ ⊆ V satisfying:

ˆ VZ ⊗ C = V

ˆ UZ · VZ ⊆ VZ .

Since UZ and VZ are defined over the integers, we can now define

U k = UZ ⊗Z k and V k = VZ ⊗Z k

for any field k.

Definition. The algebra U k is called the hyperalgebra of g over k.

For a Lie algebra g with g-module V , let

ρ : U k −→ End(V k )

be the corresponding representation over k. For each t ∈ k and α ∈ Φ, set



X
Xα (t) = exp(tXα ) = tn ρ(Xα(n) ).
n=0

535
22.4. Chevalley Groups Chapter 22. Finite Groups of Lie Type

Definition. The root group for a root α ∈ Φ is the subgroup

uα = hXα (t) | t ∈ ki ⊆ End(V k ).

For the triple (g, V, k), the Chevalley group is then defined as

GkV = huα | α ∈ Φi.

Example 22.4.1. For g = sln (C) with root system Φ = {εi − εj | i 6= j}, the root groups
are generated by Xεi −εj = Eij , the ijth matrix unit in SLn (k). Taking V = Cn to be the
defining representation of sln (C), we have that V k = k n for any field k. Then

Xij (t) := Xεi −εj (t) = 1 + tEij ∈ GLn (k)

so by Lemma 22.1.4, we see that GkV = SLn (k) for any field k. This construction even works
for any commutative ring R in place of k.

Let Λr = X ∨ = ni=1 Zαi be the root lattice


L
Ln of g, where αi are the simple roots. This is a
sublattice of the weight lattice Λw = X = i=1 Z$i , where $i are the fundamental dominant
weights. Now if V is any finite dimensional, faithful G-module, there is an intermediate
lattice Λr ⊆ ΛV ⊆ Λw defined by

ΛV = SpanZ {µ ∈ Λr | Vµ 6= 0}.

Proposition 22.4.2. The quotient Λw /Λr is finite, and for any intermediate lattice Λr ⊆
Λ ⊆ Λw , there is a finite dimensional, faithful g-module V for which ΛV = Λ.

Fix a Lie algebra g, a finite dimensional, faithful g-module V and a Chevalley group
GV = GkV for some field k.

Theorem 22.4.3. If V, V 0 are finite dimensional, faithful g-modules such that ΛV ⊆ ΛV 0 ,


then there exists a surjective homomorphism ϕ : GV 0 → GV such that ker ϕ ⊆ Z(GV 0 ) is a
finite subgroup. In particular, the Chevalley group GV only depends on the lattice ΛV .

Example 22.4.4. The weight lattice Λw corresponds to the universal group Gsc , also known
as the simply connected group (in the sense of Example 21.6.21). Likewise, the Cheval-
ley group GAd corresponding to the root lattice Λr is GAd , the adjoint group from Ex-
ample 21.6.21. Thus, given any g-module V with Chevalley group GV , there is a pair of
α β
surjections Gsc −
→ GV →
− GAd .
α β
Proposition 22.4.5. Let V be a finite dimensional, faithful g-module and Gsc −
→ GV →
− GAd
the corresponding morphisms of Chevalley groups. Then

(a) ker β = Z(GV ). In particular, Z(GAd ) = {1}.

(b) When k = C, GV is a Lie group, ker β ∼ = ΛV /Λr and ker α ∼ = Λw /ΛV ∼= π1 (GV ), the
topological fundamental group of GV . In particular, π1 (Gsc ) = {1} so Gsc is simply
connected as a manifold.

536
22.4. Chevalley Groups Chapter 22. Finite Groups of Lie Type

Example 22.4.6. For the semisimple Lie algebra g = sln (C), the universal group is Gsc =
SLn (k), while the adjoint group is GAd = Gsc /Z(Gsc ) = SLn (k)/Z(SLn (k)) = P SLn (k).

Theorem 22.4.7. Assume Φ is an irreducible root lattice and k is a field such that Φ is not
of type A1 , B2 or G2 if k = F2 , and Φ is not of type A1 if k = F3 . Then for any g-module V
with root lattice Φ, GAd = GV /Z(GV ) is simple as an abstract group.

Theorem 22.4.8. Let k be an algebraically closed field and V a g-module with Chevalley
group GV . Then GV is a semisimple, connected algebraic group over k. Furthermore, every
semisimple, connected algebraic group over k arises in this way.

It turns out that if k is any field of characteristic p > 0, then every Chevalley group can
be defined over the prime subfield Fp . This describes a large portion of the theory of finite
simple groups, although as noted at the end of Section 22.2, there are exceptional cases.

537
Part V

Generalized Jacobians

538
Chapter 23

Classical Jacobians

The notes in Part V come from the 2016-2017 Galois-Grothendieck Seminar at the University
of Virginia. The general theme of the year’s seminar was “generalized Jacobians”: their
construction and application for algebraic curves over an arbitrary field.
Let C be a curve over a field k. We will almost always assume C is irreducible, projective
and smooth. For the first chapter, k will be the complex numbers C. Our goal is to construct
a new object J, the Jacobian of C, which is a projective algebraic variety over k of dimension
equal to the genus of C. We will see that J is also a group, and in fact is an algebraic group.
This means that since J is a projective (hence complete) algebraic group, it is abelian. Such
an algebraic group is called an abelian variety. When the genus of C is greater than 0, there
will be a canonical embedding C ,→ J.
The class field theory of k(C) = K will also be of interest. In general, for a field K,
class field theory describes the abelian extensions L/K, i.e. finite Galois extensions with
Gal(L/K) abelian, in terms of C itself. In the case of K = k(C) the function field of a
curve, such an extension will be of the form L = k(C 0 ) for a curve C 0 . Then the field
embedding K ⊆ L corresponds to a ramified cover C 0 → C. It turns out that if C 0 → C is
unramified, then it corresponds to a finite index subgroup of the surface group
g
* +
Y
Σg = x1 , y1 , . . . , xg , yg : [xi , yi ] = 1
i=1

where g is the genus of C. Even better, these covers of curves can be realized by covers of
algebraic varieties. When C 0 → C is a ramified cover, the usual notion of Jacobian will be
replaced by a generalized Jacobian, which is isomorphic to a direct product of the Jacobian
of C and another algebraic group.
We will see in the first chapter that when k = C, the theory has powerful connections to
the theory of complex manifolds. In particular, the Jacobian J can be realized as a complex
torus Cg /Λ where Λ is a lattice, and we will have π1 (C)ab ∼
= π1 (J) = Z2g .

539
23.1. Elliptic Functions Chapter 23. Classical Jacobians

23.1 Elliptic Functions


Let Λ ⊆ C be a lattice, i.e. a free abelian subgroup of rank 2. Then Λ can be written
ω1
Λ = Zω1 + Zω2 for some ω1 , ω2 ∈ C such that 6∈ R.
ω2
Definition. A function f : C → C ∪ {∞} is doubly periodic with lattice of periods Λ
if f (z + `) = f (z) for all ` ∈ Λ and z ∈ C.
Definition. An elliptic function is a function f : C → C ∪ {∞} that is meromorphic and
doubly periodic.
It is not obvious that doubly periodic functions even exist! We will prove this shortly.
Definition. Let Λ ⊆ C be a lattice. The set

Π = Π(ω1 , ω2 ) = {t1 ω1 + t2 ω2 | 0 ≤ ti < 1}

is called the fundamental parallelogram, or fundamental domain, of Λ. We say a


subset Φ ⊆ C is fundamental for Λ if the quotient map C → C/Λ restricts to a bijection
on Φ.

ω1

ω2
Π

Lemma 23.1.1. For any choice of basis [ω1 , ω2 ] of Λ, Π(ω1 , ω2 ) is fundamental for Λ.
Lemma 23.1.2. Let Λ be a lattice. Then
(a) If Π is the fundamental domain of Λ, then for any α ∈ C, Πα := Π + α is fundamental
for Λ.
[
(b) If Φ is fundamental for Λ, then C = Φ + `.
`∈Λ

Corollary 23.1.3. Suppose f is an elliptic function with lattice of periods Λ and Φ funda-
mental for Λ. Then f (C) = f (Φ).

540
23.1. Elliptic Functions Chapter 23. Classical Jacobians

Proposition 23.1.4. A holomorphic elliptic function is constant.


Proof. Let f be such an elliptic function and let Φ be the fundamental domain for its lattice
of periods. Then Π is compact and hence f (Π) is as well. In particular, f (C) = f (Π) ⊆ f (Π)
is bounded, so by Liouville’s theorem, f is constant.
The prominence of tools from complex analysis (e.g. Liouville’s theorem in the above
proof) is obvious in the study of elliptic functions. Another important result for computations
is the residue theorem:
Theorem (Residue Theorem). For any meromorphic function f on a region R ⊆ C, with
isolated singularities z1 , . . . , zk ∈ R. Then if ∆ = ∂R,
Z k
X
f (z) dz = 2πi Res(f ; zi ).
∆ i=1

Proposition 23.1.5. Let f be an elliptic function. If α ∈ C is a complex number such that


∂Πα does not contain any of the poles of f , then the sum of the residues of f inside ∂Πα
equals 0.
Proof.R Fix a basis [ω1 , ω2 ] of Λ and set ∆ = ∂Πα . By the residue theorem, it’s enough to
show ∆ f (z) dz = 0. We parametrize the boundary of Π as follows:
γ1 = α + tω1
γ2 = α + ω1 + tω2
γ3 = α + (1 − t)ω1 + ω2
γ4 = α + (1 − t)ω2 .
γ3

γ4 Πα γ2

γ1
α

R R R R
We show that γ1 f (z) dz+ γ3 f (z) dz = 0 and leave the proof that γ2 f (z) dz+ γ2 f (z) dz = 0
for exercise. Consider
Z Z Z 1 Z 1
f (z) dz + f (z) dz = f (α + tω1 )(ω1 dt) + f (α + (1 − t)ω1 + ω2 )(−ω1 dt)
γ1 γ3 0 0
Z 1 Z 0
= ω1 f (α + tω1 ) dt + ω1 f (α + sω1 ) ds since f is elliptic
0 1
Z 1 Z 1 
= ω1 f (α + tω1 ) dt − f (α + sω1 ) ds = 0.
0 0

Hence the sum of the residues equals 0.

541
23.1. Elliptic Functions Chapter 23. Classical Jacobians

Corollary 23.1.6. Any elliptic function has either a pole of order at least 2 or two poles on
the fundamental domain of its lattice of periods.

Proposition 23.1.7. Suppose f is an elliptic function with fundamental domain Π and


n
α ∈ C such that ∆ = ∂Πα does not contain any zeroes or poles of f . LetPn {aj }j=1 be a finite
set of zeroes and poles in Πα , with mj the order of the pole aj . Then j=1 mj = 0.

Proof. For a pole z0 , we can write f (z) = (z − z0 )m g(z) for some holomorphic function g(z),
with g(z0 ) 6= 0. Then

f 0 (z) g 0 (z)
 
−1
= (z − z0 ) m + (z − z0 ) .
f (z) g(z)
 0 
Hence Res ff ; z0 = m. Then the statement follows from Proposition 23.1.5.

Proposition 23.1.7 has an analogue in algebraic


P geometry: if f is a rational function on
an algebraic curve C, the formal sum (f ) = mj aj , where the aj and mj arePdefined as
above, is called the principal divisor associated to f and its degree is deg(f ) = mj . Then
one can prove that deg(f ) = 0.
Continuing in the complex setting, let f be an elliptic function and let a1 , . . . , ar be the
poles and zeroes of f in the fundamental domain of Λ. Write ordai fP= mi if ai is a pole
of order −mi or if ai is a zero of multiplicity mi . The sum ord(f ) = ri=1 mi is called the
order of f . Then Corollary 23.1.6 says that there are no elliptic functions of order 1. We
will show that the field of elliptic functions with period lattice Λ is generated by an order 2
and an order 3 function.
Let f be elliptic and z0 ∈ C with ordz0 f = m. Then for any ` ∈ Λ, ordz0 +` f = m as
well. Indeed, if z0 is a zero then

0 = f (z0 ) = f (z0 ) = . . . = f (m−1) (z0 )

but f (k) (z) is also elliptic for all k ≥ 1. If z0 is a pole of f , the same result can be obtained
using f1 instead of f .
If Φ1 and Φ2 are any two fundamental domains for Λ, then for all a1 ∈ Φ1 , there is a
unique a2 ∈ Φ2 such that a2 = a1 + ` for some ` ∈ Λ. Thus Propositions 23.1.5 and 23.1.7
hold for any fundamental domain of Λ, so it follows that ord(f ) is well-defined on the quotient
C/Λ.
Now given any meromorphic function f (z) on C, we would like to construct an elliptic
function F (z) with lattice Λ. Put
X
F (z) = f (z + `).
`∈Λ

There are obvious problems of convergence and (in a related sense) the order of summation.
It turns out we can do this construction with f (z) = z1m , m ≥ 3 though. First, we need the
following result from complex analysis, which can be proven using Cauchy’s integral formula
and Morera’s theorem.

542
23.1. Elliptic Functions Chapter 23. Classical Jacobians

Lemma 23.1.8. Let U ⊆ C be an open set and suppose (fn ) is a sequence of holomorphic
functions on U such that fn → f uniformly on every compact subset of U . Then f is
holomorphic on U and fn0 → f 0 uniformly on every compact subset of U .

Proposition 23.1.9. Let Λ be a lattice with basis [ω1 , ω2 ]. Then the sum
X 1
|ω|s
ω∈Λr{0}

converges for all s > 2.

Proof. Extend the fundamental domain by translation by the vectors ω1 , ω2 and ω1 + ω2 ,


and call the boundary of the resulting region ∆:

Λ Λ

Λ Λ

Then ∆ is compact, so there exists c > 0 such that |z| ≥ c for all z ∈ ∆. We claim that for
all m, n ∈ Z,
|mω1 + nω2 | ≥ c · max{|m|, |n|}.
The cases when m = 0 or n = 0 are trivial, so without loss of generality assume m ≥ n > 0.
Then n
|mω1 + nω2 | = |m| ω1 + ω2 ≥ |m|c.

m
Hence the claim holds. Set M = max{|m|, |n|} and arrange the sum in question so that the
1
|ω|s
are added in order of increasing M values. Then the sum can be estimated by
∞ ∞
X 1 X 8M X 1
≤ ∼ .
|ω|s M =1 cs M s M =1 M s−1
ω∈Λr{0}

This converges for s > 2 by p-series.

Proposition 23.1.10. Let n ≥ 3 and define


X 1
Fn (z) = n
.
ω∈Λ
(z − ω)

Then Fn (z) is holomorphic on C r Λ and has poles of order n at the points of Λ. Moreover,
Fn is doubly periodic and hence elliptic.

543
23.1. Elliptic Functions Chapter 23. Classical Jacobians

Proof. Fix r > 0 and let Br = Br (0) be the open complex r-ball centered at the origin in C.
Let Λr = Λ ∩ B r be the lattice points contained in the closed r-ball. Then the function
X 1
Fn,r (z) =
ω∈ΛrΛr
(z − ω)n

1 C
is holomorphic on Br . To see this, one has |z−ω| n ≤ |ω|n for some constant C and for all

z ∈ Br , ω ∈ Λ r Λr . Then |ω|Cn converges by Proposition 23.1.9, so by the Weierstrass M -test,


1
|z−ω|n
converges uniformly and hence Fn,r (z) is holomorphic. It follows from the definition
that Fn has a pole of order n at each ω ∈ Λ. Finally, for ` ∈ Λ, we have
X 1 X 1
Fn (z + `) = n
= = Fn (z)
ω∈Λ
(z + ` − ω) η∈Λ
(z − η)n

since the series is absolutely convergent and we can rearrange the terms.
This shows that elliptic functions exist and more specifically that for each n ≥ 3, there
is at least one elliptic function of order n. Unfortunately the previous proof won’t work
to construct an elliptic function of order 3. However, Weierstrass discovered the following
elliptic function.

Definition. The Weierstrass ℘-function for a lattice Λ is defined by

1 X  1 1

℘(z) = 2 + − .
z (z − w)2 ω 2
ω∈Λr{0}

Theorem 23.1.11. For any lattice Λ, ℘(z) is an elliptic function with poles of order 2 at
the points of Λ and no other poles. Moreover, ℘(−z) = ℘(z) and ℘0 (z) = −2F3 (z).

Proof. (Sketch) To show ℘(z) is meromorphic, one estimates the summands by



1 1 D
(z − ω)2 − ω 2 ≤ |ω|3

for some constant D and all z ∈ Br , ω ∈ Λ r Λr as in the previous proof.


Next, ℘(z) can be differentiated term-by-term to obtain the expression ℘0 (z) = −2F3 (z).
And proving that ℘(z) is odd is straightforward:

1 X  1 1

℘(−z) = + −
(−z)2 (−z − ω)2 ω 2
ω∈Λr{0}

1 X  1 1

= 2+ − = ℘(z)
z (z − (−ω))2 (−ω)2
−ω∈Λr{0}

after switching the order of summation.

544
23.1. Elliptic Functions Chapter 23. Classical Jacobians

Finally, proving ℘(z) is doubly periodic is difficult since we don’t necessarily have absolute
convergence. However, one can reduce to proving ℘(z + ω1 ) = ℘(z) = ℘(z + ω2 ). Then using
the formula for ℘0 (z), we have

d
[℘(z + ω1 ) − ℘(z)] = −2F3 (z + ω1 ) + 2F3 (z)
dz
= −2F3 (z) + 2F3 (z) = 0

since F3 (z) is elliptic by Proposition 23.1.10. Hence ℘(z+ ω1 ) − ℘(z) = c is constant.


Evaluating at z = − ω21 , we see that c = ℘ ω21 − ℘ − ω21 = 0 since ℘(z) is odd. Hence
c = 0, so it follows that ℘(z) is doubly periodic and therefore elliptic.

Lemma 23.1.12. Let ℘(z) be the Weierstrass ℘-function for a lattice Λ ⊆ C and let Π be
the fundamental domain of Λ. Then

(1) For any u ∈ C, the function ℘(z) − u has either two simple roots or one double root
in Π.

(2) The zeroes of ℘0 (z) in Π are simple and they only occur at ω21 , ω22 and ω1 +ω
2
2
.

(3) The numbers u1 = ℘ ω21 , u2 = ℘ ω22 and u3 = ℘ ω1 +ω


  
2
2
are precisely those u for
which ℘(z) − u has a double root.

Proof. (1) follows from Corollary 23.1.6.


(2) By Theorem 23.1.11, deg ℘0 (z) = 3 so it suffices to show that ω21 , ω22 and ω1 +ω2
2
are all
roots. For z = ω21 , we have
ω   ω  ω  ω 
1 1 1 1
℘0 = −℘0 − = −℘0 − ω1 = −℘0
2 2 2 2
since ℘0 (z) is elliptic. Thus ℘0 ω21 = 0. The others are similar.


(3) The double roots occur exactly when ℘0 (u) = 0, so use (2).
We now prove that any elliptic function can be written in terms of ℘(z) and ℘0 (z).

Theorem 23.1.13. Fix a lattice Λ ⊆ C and let E(Λ) be the field of all elliptic functions
with lattice of periods Λ. Then E(Λ) = C(℘, ℘0 ).

Proof. Take f (z) ∈ E(Λ). Then f (−z) ∈ E(Λ) as well and thus we can write f (z) as the
sum of an even and an odd elliptic function:

f (z) + f (−z) f (z) − f (−z)


f (z) = feven (z) + fodd (z) = + .
2 2
We will prove that every even elliptic function is rational in ℘(z), but this will imply the
theorem, since then feven (z) = ϕ(℘(z)) and f℘odd (z)
0 (z) = ψ(℘(z)) for some ϕ, ψ ∈ C(℘(z)) and
0
we can then write f (z) = ϕ(℘(z)) + ℘ (z)ψ(℘(z)).
Assume f (z) is an even elliptic function. It’s enough to construct ϕ(℘(z)) such that
f (z)
ϕ(℘(z))
only has (potential) zeroes and poles at z = 0 in the fundamental parallelogram for

545
23.1. Elliptic Functions Chapter 23. Classical Jacobians

f (z)
Λ, since then by Corollary 23.1.6, ϕ(℘(z)) is holomorphic and then by Proposition 23.1.4
it is constant. Suppose f (a)
 = 0 for a some zero of order m. Consider ℘(z) = u. If
ω1 ω2 ω1 +ω2
 
u 6= ℘ 2 , ℘ 2 , ℘ 2
then ℘(z) = u has precisely two solutions in the fundamental
parallelogram, z = a and z = a∗ where

ω1 + ω2 − a if a ∈ Int(Π)


a = ω1 − a if a is parallel to ω1

ω2 − a if a is parallel to ω2 .

(Notice that since f is even, f (a) = 0 implies f (a∗ ) = 0 as well.) Moreover, if orda f = 0 then
orda∗ f = m. Note that a = a∗ holds precisely when a is in the set Θ := 0, ω21 , ω22 , ω1 +ω

2
2
.
Let Z (resp. P ) be the set of zeroes (resp. poles) of f (z) in Π. Then the assignment
a 7→ a∗ is in fact an involution on Z and P , so we can write

Z = Z10 ∪ · · · ∪ Zr0 ∪ Z100 ∪ · · · ∪ Zs00


P = P10 ∪ · · · ∪ Pu0 ∪ P100 ∪ · · · ∪ Pv00

where the Zi0 and Pi0 are the 2-element orbits of the involution and the Zj00 and Pj00 are the
1-element orbits. Of course then s, v ≤ 3. For a0i ∈ Zi0 , set orda0i f = m0i and for a00j ∈ Zj00 ,
set orda00i f = m00i , which is even. Likewise, for b0i ∈ Pi0 , set ordb0i f = n0i and for b00j ∈ Pj00 , set
ordb00i f = n00i which is even. Then we define ϕ(℘(z)) by

0 m0i 00 m00
Qr Qs
j /2
i=1 (℘(z) − ℘(ai )) j=1 (℘(z) − ℘(aj ))
ϕ(℘(z)) = Qu 0 n0i
Qv 00 nj
.
i=1 (℘(z) − ℘(bi )) j=1 (℘(z) − ℘(bj ))

Then ϕ(℘(z)) has only potential zeroes/poles at z = 0 in the fundamental parallelogram, so


we are done.

546
23.2. Elliptic Curves Chapter 23. Classical Jacobians

23.2 Elliptic Curves


Let Λ ⊆ C be a lattice. There is a canonical way to associate to the complex torus C/Λ
an elliptic curve E such that C/Λ ∼ = E(C). We would also like to reverse this process, i.e.
given an elliptic curve E, define a lattice Λ ⊆ C such that C/Λ ∼ = E(C). This procedure
generalizes for a curve C of genus g > 1 and produces its Jacobian, C ,→ Cg /Λ = J(C).
We need the following lemma from complex analysis.

Lemma 23.2.1. Suppose f0 , f1 , f2 , . . . is a sequence of analytic functions on the ball Br (z0 )


with Taylor expansions
X∞
(n)
fn (z) = ak (z − z0 )k .
k=0
P∞
Then if F (z) = n=0 fn (z) converges uniformly on Bρ (z0 ) for all ρ < r, each series Ak =
P∞ (n)
n=0 ak converges and F (z) has Taylor expansion

X
F (z) = Ak (z − z0k ).
k=0

Let ℘(z) be the Weierstrass ℘-function for Λ. Then ℘0 (z)2 is an even elliptic function, so
by Theorem 23.1.13, ℘0 (z)2 ∈ C(℘). On a small enough neighborhood around z0 = 0,

1 X  1 1

℘(z) − 2 = −
z (z − ω)2 ω 2
ω∈Λr{0}

is analytic. Moreover, for each ω ∈ Λ r {0} we have

1 1 2z 3z 2
= + + 4 + ...
(z − ω)2 ω2 ω3 ω
2
1 1 2z 3z
=⇒ 2
− 2 = 2 + 4 + ...
(z − ω) ω ω ω

which is uniformly convergent. Hence Lemma 23.2.1 shows that


∞ ∞
1 X X k+1 X
℘(z) − = zk = (k + 1)Gk+2 z k
z2 ω k+2
ω∈Λr{0} k=1 k=1

1
P
where Gm = Gm (Λ) := ω∈Λr{0} ω m .
These Gm are examples of modular forms.
1
P
Definition. The series Gm (Λ) = ω∈Λr{0} ω m is called the Eisenstein series for Λ of
weight m.

547
23.2. Elliptic Curves Chapter 23. Classical Jacobians

From the above work, we obtain the following formulas:


1
℘(z) = + 3G4 z 2 + 5G6 z 4 + 7G8 z 6 + . . .
z2
1
℘(z)2 = 4 + 6G4 + . . .
z
1 9G4
℘(z)3 = 6 + 2 + 15G6 + . . .
z z
2
℘0 (z) = − 3 + 6G4 z + . . .
z
4 24G4
℘0 (z)2 = 6 − 2 − 80G6 − . . .
z z
This implies:
Proposition 23.2.2. The functions ℘ and ℘0 satisfy the following relation:
℘0 (z)2 = 4℘(z)3 − g2 ℘(z) − g3
where g2 = 60G4 and g3 = 140G6 .
Consider the polynomial p(x) = 4x3 − g2 x − g3 , where the gn are defined for the lattice
Λ ⊆ C.
Proposition 23.2.3. p(x) = 4(x − u1 )(x − u2 )(x − u3 ) where u1 = ℘ ω21 , u2 = ℘ ω22 and
 

u3 = ℘ ω1 +ω
2
2
are distinct roots.
Thus (x, y) = (℘(z), ℘0 (z)) determine an equation y 2 = 4x3 −g2 x−g3 which is the defining
equation for an elliptic curve E0 over C. Let E = E0 ∪ {[0, 1, 0]} ⊆ P2 be the projective
closure of E0 . The point [0, 1, 0] is sometimes denoted ∞.
Theorem 23.2.4. The map
ϕ : C/Λ −→ E(C)
(
[℘(z), ℘0 (z), 1], z ∈
6 Λ
z + Λ 7−→ ϕ(z + Λ) =
[0, 1, 0], z∈Λ
is a bijective, biholomorphic map.
Proof. Assume z1 , z2 ∈ C are such that z1 + Λ 6= z2 + Λ. Without loss of generality we may
assume z1 , z2 ∈ Π, the fundamental domain of Λ (otherwise, translate). If ℘(z1 ) = ℘(z2 ) and
℘0 (z1 ) = ℘0 (z2 ), then
 with the notation of Theorem 23.1.13, we must have z2 = z1∗ 6= z1 and
ω1 ω2 ω1 +ω2
. Since ℘0 (z) is odd, we get ℘0 (z1 ) = ℘0 (z2 ) = −℘0 (−z2 ) =

thus z1 , z2 6∈ Θ = 0, 2 , 2 , 2
−℘0 (z1 ), but this implies ℘(z1 ) = 0, contradicting z1 6∈ Θ. Therefore ϕ is one-to-one.
Next, we must show that for any (x0 , y0 ) ∈ E(C), x0 = ℘(z) and y0 = ℘0 (z) for some
z ∈ C. If ℘(z1 ) = x0 , then it’s clear that ℘0 (z1 ) = y0 or −y0 . Now one shows as in the
previous paragraph that we must have ℘0 (z1 ) = y0 .
Now consider F (x, y) = y 2 − p(x), where p(x) = 4x3 − g2 x − g3 . If (x0 , y0 ) satisfies
F (x0 , y0 ) = 0 and y0 6= 0, then ∂F∂y
(x0 , y0 ) 6= 0 and thus the assignment (x, y) 7→ x is a local
chart about (x0 , y0 ). Likewise, (x, y) 7→ y defines a local chart about (x0 , y0 ) when x0 6= 0.
Finally, we conclude by observing that a locally biholomorphic map is biholomorphic.

548
23.2. Elliptic Curves Chapter 23. Classical Jacobians

In general, an elliptic curve can be defined by a Weierstrass equation

E : y 2 = f (x) = ax3 + bx2 + cx + d.


X Y
This embeds into projective space via (x, y) 7→ [x, y, 1]. Setting x = Z
and y = Z
, we also
obtain a homogeneous equation for the curve:

E : ZY 2 = aX 3 + bX 2 Z + cXZ 2 + dZ 3 .

The single point at infinity, [0, 1, 0], can be studied by dehomogenizing via the coordinates
z̃ = YZ and x̃ = X
Y
, which yield

E : z̃ = ax̃3 + bx̃2 z̃ + ax̃z̃ 2 + dz̃ 3 .

We have shown that a lattice Λ ⊆ C determines elliptic functions ℘(z) and ℘0 (z) that satisfy
℘0 (z)2 = 4℘(z)3 − g2 ℘(z) − g3 and that this polynomial expression has no multiple roots.
Therefore the mapping z 7→ (℘(z), ℘0 (z)) determines a bijective correspondence C/Λr{0} →
E(C) r {∞} which can be extended to all of C/Λ → E(C) (this is Theorem 23.2.4). There
is a natural group structure on C/Λ induced from C, but what is not so obvious is that E(C)
also possesses a group structure, the so-called “chord-and-tangent method”.
Let E be an elliptic curve over an arbitrary field k, let O ∈ E(k) be the point at infinity
and fix two points P, Q ∈ E(k). In the plane P2k , there is a unique line containing P and Q;
call it L. (If P = Q, then take L to be the tangent line to E at P .) By Bézout’s theorem,
E ∩ L = {P, Q, R} for some third point R ∈ E(k), which may not be distinct from P and Q
if multiplicity is counted. Let L0 be the line through R and O and call its third point R0 .

Q
P

P +Q

Addition of two points P, Q ∈ E(k) is defined by P + Q = R0 , where R0 is the unique


point lying on the line through R and O. If R = O, we set R0 = O.

Proposition 23.2.5. Let E be an elliptic curve with O ∈ E(k). Then

(a) If L is a line in P2 such that E ∩ L = {P, Q, R}, then (P + Q) + R = O.

(b) For all P ∈ E(k), P + O = P .

549
23.2. Elliptic Curves Chapter 23. Classical Jacobians

(c) For all P, Q ∈ E(k), P + Q = Q + P .


(d) For all P ∈ E(k), there exists a point −P ∈ E(k) satisfying P + (−P ) = O.
(e) For all P, Q, R ∈ E(k), (P + Q) + R = P + (Q + R).
Together, (b) – (e) say that chord-and-tangent addition of points defines an associative,
commutative group law on E(k). The proofs of (a) – (d) are rather routine using the
definition of this addition law, whereas verifying associativity is notoriously difficult. There
are formulas for the coordinates of P + Q that make this possible though (see Silverman).
Theorem 23.2.6. The map ϕ : C/Λ → E(C) is an isomorphism of abelian groups.
Proof. Consider the diagram
ϕ×ϕ
C/Λ × C/Λ E(C) × E(C)

α β

C/Λ E(C)
ϕ

where α and β are the respective group operations. Since C/Λ × C/Λ is a topological group,
it’s enough to show the diagram commutes on a dense subset of C/Λ × C/Λ. Consider
e = {(u1 , u2 ) ∈ C2 | u1 , u2 , u1 ± u2 , 2u1 + u2 , u1 + 2u2 6∈ Λ}.
X

Then X e ∼
= C2 so X = X e mod Λ × Λ is dense in C/Λ × C/Λ. Take (u1 + Λ, u2 + Λ) ∈ X
and set u3 = −(u1 + u2 ). Then u1 + u2 + u3 = 0 in C/Λ. Set P = ϕ(u1 ), Q = ϕ(u2 ) and
R = ϕ(u3 ) ∈ E(C). By the assumptions on X, the points P, Q, R are distinct. We want to
show ϕ(u1 + u2 ) = ϕ(u1 ) + ϕ(u2 ) = P + Q. Since ℘(z) is even and ℘0 (z) is odd, we see that
ϕ(−z) = −ϕ(z) for all z ∈ C/Λ. Thus ϕ(u1 + u2 ) = −ϕ(−(u1 + u2 )) = −R so we need to
show P + Q + R = O, i.e. P, Q, R are colinear. Since u1 6= u2 , the line P Q is not vertical,
so there exist a, b such that ℘0 (ui ) = a℘(ui ) + b for i = 1, 2. Consider the elliptic function

f (z) = ℘0 (z) − (a℘(z) + b).

Then on the fundamental domain Π, f only has a pole at 0, so ord0 f = −3. Also, u1 and u2
are distinct zeroes of f , so there is a third point ω ∈ Π such that deg(f ) = u1 +u2 +ω−3·0 = 0,
i.e. u1 + u2 + ω = 0. Solving for ω, we get ω = −(u1 + u2 ) = u3 . It follows that R = ϕ(u3 )
is on the same line as P and Q, so we are done.
The compatibility of the group operations of C/Λ and E(C) is highly useful. For example,
fix N ∈ N and let
E[N ] = {P ∈ E(C) | [N ]P = O},
where [N ]P = P
| + .{z
. . + P}. The points of E[N ] are called the N -torsion points of E. For
N
N = 2, the points P such that P = −P are exactly the intersection points of E with the
x-axis along with O = [0, 1, 0]:

550
23.2. Elliptic Curves Chapter 23. Classical Jacobians

In general, one can show that #E[N ] = N 2 . This is hard to see from the geometric
picture, but working with the isomorphism E(C) ∼ = C/Λ from Theorem 23.2.6, we see
that since C/Λ = R/Z × R/Z as an abelian group, the N -torsion is given by (C/Λ)[N ] =
1
N
Z/Z × N1 Z/Z. This is a group of order N 2 .
A morphism in the category of elliptic curves is called an isogeny. Explicitly, ϕ : E1 → E2
is an isogeny between two elliptic curves if it is a (nonconstant) morphism of schemes that
takes the basepoint O1 ∈ E1 to the basepoint O2 ∈ E2 .
Proposition 23.2.7. Suppose Λ1 , Λ2 ⊆ C are lattices and f : C/Λ1 → C/Λ2 is a holomor-
phic map. Then there exist a, b ∈ C such that aΛ1 ⊆ Λ2 and

f (z mod Λ1 ) = az + b mod Λ2 .

Proof. As topological spaces, C/Λ1 and C/Λ2 are complex tori with the same universal
covering space C, so any f : C/Λ1 → C/Λ2 lifts to F : C → C making the diagram
commute:
F
C C

π1 π2

C/Λ1 C/Λ2
f

Since covers are local homeomorphisms, it follows that F is holomorphic as well. Thus for
any z ∈ C, ` ∈ Λ1 ,

π2 (F (z + `) − F (z)) = f (π1 (z + `) − π1 (z)) = f (π1 (z) − π1 (z)) = f (0) = 0.

So F (z + `) − F (z) ∈ Λ1 for any ` ∈ Λ1 and the function L(z) = F (z + `) − F (z) is constant.


It follows that F 0 (z + `) = F 0 (z), so F 0 is holomorphic and elliptic, but this means by
Proposition 23.1.4 that F 0 (z) = a for some constant a. Hence F (z) = az + b as claimed.

551
23.2. Elliptic Curves Chapter 23. Classical Jacobians

Corollary 23.2.8. Any holomorphic map f : C/Λ1 → C/Λ2 is, up to translation, a group
homomorphism. In particular, if f (0) = 0 then f is a homomorphism.

Corollary 23.2.9. For any elliptic curve E, the group of endomorphisms End(E) has rank
at most 2.

Proof. Viewing E(C) = C/Λ for some Λ = Z + Zτ , we get

End(E) = {f : E → E | f is an isogeny}
= {f : C/Λ → C/Λ | f is holomorphic and f (0) = 0} by Corollary 23.2.8
= {z ∈ C | zΛ ⊆ Λ}
= {z ∈ C | z(Z + Zτ ) ⊆ (Z + Zτ )}
⊆ Z + Zτ.

Hence rank End(E) ≤ 2.


It turns out that there are two possible cases for the rank of End(E), breaking down as
follows:

ˆ End(E) = Z.

ˆ End(E) is an order O in some imaginary quadratic number field K/Q. In this case, E
is said to have complex multiplication.

552
23.3. The Classical Jacobian Chapter 23. Classical Jacobians

23.3 The Classical Jacobian


For the isomorphism ϕ : C/Λ → E(C) in Theorem 23.2.6, let ψ = ϕ−1 : E(C) → C/Λ be
the inverse map. To understand this map explicitly, we will show how to construct a torus
for every elliptic curve, i.e. find a lattice Λ ⊆ C such that C/Λ ∼
= E(C).
Lemma 23.3.1. Any lattice Λ ⊆ C can be written
Z P 
Λ= dz : P ∈ Λ .
0

Notice that each differential form dz on C satisfies d(z + `) = dz for all ` ∈ Λ by


Lemma 23.3.1. Thus dz descends to a differential form on C/Λ, which by abuse of notation
we will also denote by dz. Formally, this is the pushforward of dz along the quotient π :
C → C/Λ. This implies:

Lemma 23.3.2. Any lattice Λ ⊆ C can be written


Z 
Λ= dz : γ is a closed curve in C/Λ passing through 0 .
γ

For an elliptic curve E defined by the equation y 2 = f (x), fix a holomorphic differential
form ω on E(C). (In general, the space of holomorphic differential forms on a curve has
dimension equal to the genus of the curve, so in the elliptic curve case, there is exactly one
such ω, up to scaling.)

Definition. The lattice of periods for an elliptic curve E is


Z 
Λ= ω : γ is a closed curve in E passing through P
γ

where P ∈ E(C) is fixed.

Example 23.3.3. Under the map ϕ : C/Λ → E(C), z 7→ (x, y) = (℘(z), ℘0 (z)), we see that

dx = ℘0 (z) dz = y dz

so ω = dx
y
is a differential form on E(C). In fact, ω = f dx 2
0 (x) , where E is defined by y = f (x),

is holomorphic because f 0 (x) 6≡ 0. This differential form is also holomorphic at O = [0, 1, 0],
so up to scaling, this is the unique holomorphic form on E.

Historically, mathematicians were interested in studying solutions to elliptic integrals, or


integrals of the form Z
dx
√ .
3
ax + bx + c
When f (x) = ax3 + bx + c, the expression ω = √ax3dx +bx+c
is precisely the holomorphic
differential form defining the lattice of periods of the elliptic curve E : y 2 = f (x).

553
23.3. The Classical Jacobian Chapter 23. Classical Jacobians

For a more functorial description, let VE = Γ(E, ΩE ) be the space of all holomorphic
differential forms on E. If γ is a curve in E(C), there is an associated linear functional
ϕγ ∈ VE∗ defined by

ϕγ : VE −→ C
Z
ω 7−→ ω.
γ

Fixing the basepoint O ∈ E(C), the lattice of periods for E can be written

Λ = {ϕγ : γ ∈ π1 (E(C), O)}.

In other words, this defines a map π1 (E(C), O) → VE∗ , γ 7→ ϕγ .

Definition. The Jacobian of an elliptic curve E is the quotient J(E) = VE∗ /Λ.

For each point P ∈ E(C), the coset ϕγ + Λ is an element of the Jacobian, where γ is a
path from O to P . This defines an injective map i : E ,→ J(E).

Proposition 23.3.4. Suppose σ : E1 → E2 is an isogeny between elliptic curves, so that


σ(O1 ) = O2 . Then there is a map τ : J(E1 ) → J(E2 ) making the following diagram commute:
σ
E1 E2

i1 i2

J(E1 ) J(E2 )
τ

Proof. The pullback gives a contravariant map σ ∗ : VE2 → VE1 , ω 7→ σ ∗ ω = ω ◦ σ. Taking


the dual of this gives a linear map σ ∗∗ : VE∗1 → VE∗2 defined by (σ ∗∗ ρ)(ω) = ρ(σ ∗ ω) for any
ρ ∈ VE∗1 and ω ∈ VE2 . Taking ρ = ϕγ1 for a path γ1 in E1 gives
Z Z
∗ ∗ ∗
ρ(σ ω) = ϕγ1 (σ ω) = σ ω= ω = ϕσ(γ1 ) ω.
γ1 σ(γ1 )

Thus σ ∗∗ ϕγ1 = ϕσ(γ1 ) . If γ1 is a closed curve through O1 , then σ(γ1 ) is a closed curve passing
through O2 = σ(O1 ). Hence if ΛE1 , ΛE2 are the lattices of periods for E1 , E2 , respectively,
we have σ ∗∗ (λE1 ) ⊆ ΛE2 . So σ ∗∗ factors through the quotients, defining τ :

τ = σ ∗∗ : VE∗1 /ΛE1 −→ VE∗2 /ΛE2 .

It is immediate the diagram commutes.

Lemma 23.3.5. For any elliptic curve E, the inclusion i : E ,→ J(E) induces an isomor-
phism
i∗ : π1 (E, O) −→ π1 (J(E), i(O)).

554
23.3. The Classical Jacobian Chapter 23. Classical Jacobians

Unfortunately, the construction of the Jacobian given so far is not algebraic so it would
be hard to carry over to curves over an arbitrary ground field. To construct Jacobians
algebraically, we will prove Abel’s theorem:

Theorem 23.3.6 (Abel). Suppose Λ ⊆ C is a lattice with fundamentalP domainPΠ and take
any set {ai } ⊂ Π such that there are integers mi ∈ Z satisfying mi = 0 and mi ai ∈ Λ.
Then there exists an elliptic function f (z) whose set of zeroes and poles is {ai } and whose
orders of vanishing/poles are ordai f = mi .

Given a lattice Λ ⊆ C, we may assume Λ = Z + Zτ for some τ ∈ C with im τ > 0.

Definition. The theta function for a lattice Λ is



2 τ +2nz)
X
θ(z, τ ) = eπi(n .
n=−∞

2 2
One has |eπi(n τ +2nz) | = e−π(n im τ +2n im z)
for any z ∈ C, which implies that the above
series converges absolutely.

Proposition 23.3.7. Fix a theta function θ(z) = θ(z, τ ). Then

(1) θ(z) = θ(−z).

(2) θ(z + 1) = θ(z).

(3) θ(z + τ ) = e−πi(τ +2z) θ(z).

Properties (2) and (3) together say that θ(z) is what’s known as a semielliptic function.
For our purposes, this will be good enough. Notice that for z = 1+τ 2
, we have
   
1+τ 1+τ
θ =θ − + (1 + τ )
2 2
 
πi(τ +2(− 1+τ )) 1 + τ
=e 2 θ −
2
   
πi 1+τ 1+τ
=e θ − = −θ .
2 2
1+τ
Thus z = 2
is a zero of θ(z).

Lemma 23.3.8. All zeroes of θ(z, t) are simple and are of the form 1+τ
2
+ ` for ` ∈ Λ.

Lemma 23.3.9. For x ∈ C, set θ(x) (z, τ ) = θ z − 1+τ



2
− x . Then θ(x) (z) = θ(x) (z, τ )
satisfies:

(1) θ(x) (z + 1) = θ(x) (z).

(2) θ(x) (z + τ ) = e−πi(2(z−x)−1) θ(x) (z).

We now prove Abel’s theorem (23.3.6).

555
23.3. The Classical Jacobian Chapter 23. Classical Jacobians

Proof. Given such a set {ai } ⊂ Π, let x1 , . . . , xn be the list of all ai with mi > 0, listed with
repetitions corresponding to the number mi . For example, if m1 = 2 then x1 = x2 = a1 .
Likewise, letPy1 , . . . , yn be the list of all ai with mi < 0, once again with repetitions. By the
hypothesis mi = 0, there are indeed an equal number of each. Set
Qn (xi )
θ (z)
f (z) = Qi=1
n (yi ) (z)
.
i=1 θ

Then by Lemma 23.3.9, f (z + 1) = f (z). On the other hand, the lemma also gives
Qn (xi )
i=1 θ (z + τ )
f (z + τ ) = Q n (y i ) (z)
i=1 θ
Pn Pn
= e2πi( i=1 xi − i=1 yi ) f (z)
P
= e2πi mi ai
f (z)
X
= f (z) since mi ai = 0.

Therefore f (z) is elliptic.


Note that θ(z) is a meromorphic function, so by complex analysis, the integral

θ0 (z)
Z
1
dz
2πi ∂Π θ(z)

counts the number of zeroes of θ(z) in the fundamental domain Π, up to multiplicity. To


ensure no zeroes lying on ∂Π are missed, we may shift Π → Πα for an appropriate α ∈ C.
Parametrize ∂Π as in Proposition 23.1.5. Then once again the integrals along γ2 and γ4
cancel since θ(z + 1) = θ(z). On the other hand,

θ(z + τ ) = e−πi(τ +2z) θ(z)


=⇒ θ0 (z + τ ) = e−πi(τ +2z) (−2πiθ(z) + θ0 (z))
θ0 (z + τ ) θ0 (z)
=⇒ = −2πi + .
θ(z + τ ) θ(z)

This implies

θ0 (z) θ0 (z) θ0 (z) θ0 (z) θ0 (z)


Z Z Z Z Z
dz = dz + dz + dz + dz
∂Π θ(z) γ1 θ(z) γ2 θ(z) γ3 θ(z) γ4 θ(z)
θ0 (z) θ0 (z) θ0 (z) θ0 (z)
Z Z  Z Z 
= dz + dz + dz + dz
γ1 θ(z) γ3 θ(z) γ2 θ(z) γ4 θ(z)
θ0 (z) θ0 (z)
Z Z 
= dz − dz + 2πi + 0
γ1 θ(z) γ1 θ(z)
= 2πi.
1+τ
It follows that θ(z) has exactly one zero in Π, and it must be z = 2
.

556
23.3. The Classical Jacobian Chapter 23. Classical Jacobians

Definition. For a curve E (need not be elliptic), define:


ˆ A divisor on E is a formal sum D =
P
nP P over the points P ∈ E, with nP ∈ Z.
The abelian group of all divisors is denoted Div(E).
ˆ The degree of a divisor D =
P P
nP P ∈ Div(E) is deg(D) = nP . The set of all
0
degree 0 divisors is denoted Div (E).
ˆ For a meromorphic
P function f on E(C) = C/Λ, the principal divisor associated to
f is (f ) = degP P where nP = ordP f . The group of all principal divisors is denoted
PDiv(E).
ˆ The Picard group of E is the quotient group Pic(E) = Div(E)/ PDiv(E). The degree
zero part of the Picard group is written Pic0 (E) = Div0 (E)/ PDiv(E).
The inverse map ψ : E → C/Λ extends to the group of divisors on E:
Ψ : Div(E) −→ C/Λ
X X
nP P 7−→ nP ψ(P ).

Definition. The map Ψ : Div(E) → C/Λ is called the Abel-Jacobi map.


R
Recall that ψ : P 7→ γP ω + Λ ∈ C/Λ where ω is a fixed holomorphic differential form
on E and γP is a path connecting O ∈ E(C) to P . If O0 is another basepoint and ψ 0 is the
corresponding map, we have ψ(P ) = ψ(O0 ) + ψ 0 (P ) for all P ∈ E. So it appears that
P Ψ is
0
not well-defined. However, this issue vanishes when we restrict Ψ to Div (E): if D = nP P
is a degree 0 divisor, then
X
Ψ(D) = nP ψ(P )
X
= nP (ψ(O0 ) + ψ 0 (P ))
X X
= ψ(O0 ) nP + nP ψ 0 (P )
X
=0+ nP ψ 0 (P ) = Ψ0 (D).

Corollary 23.3.10. The map Ψ : Div0 (E) → C/Λ induces an isomorphism Pic0 (E) ∼
= C/Λ.
Proof. One can prove that Ψ is a surjective group homomorphism. Moreover, Abel’s theorem
(23.3.6) implies that ker Ψ = PDiv(E).
Consider the map iO : E → Div0 (E) that sends P 7→ P −O. This fits into a commutative
diagram:

Div0 (E)
Ψ

iO C/Λ

ψO
E

557
23.3. The Classical Jacobian Chapter 23. Classical Jacobians

On the level of the Picard group, this diagram looks like

Pic0 (E)
Ψ

iO C/Λ

ψO
E

and every arrow is a bijection.

558
23.4. Jacobians of Higher Genus Curves Chapter 23. Classical Jacobians

23.4 Jacobians of Higher Genus Curves


Let C be a complex curve of genus g ≥ 2 and let V = Γ(C, ΩC ) be the vector space of
holomorphic differential forms on C. Then dimC V =R g, so V ∗ ∼ = Cg . As in the previous
section, for any path ω in C the assignment ϕγ : ω 7→ γ ω defines a functional ϕγ ∈ V ∗ . As
for elliptic curves, we define:
Definition. The lattice of periods for C is

Λ = {ϕγ ∈ V ∗ | γ is a closed curve in C}.

Lemma 23.4.1. Λ is a lattice in V ∗ .


Definition. The Jacobian of C is the quotient space J(C) = V ∗ /Λ.
As with elliptic curves, we have a map ψ : C → J(C) called the Abel-Jacobi map, which
sends P 7→ ϕγP + Λ, where γP is a curve through P . Also, ψ extends to the divisor group of
C as a map
Ψ : Div(C) −→ J(C)
which is canonical when restricted to Div0 (C). The Abel-Jacobi theorem generalizes Theo-
rem 23.3.6 and Corollary 23.3.10.
Theorem 23.4.2. Let C be a curve of genus g > 0 and let Ψ : Div0 (C) → J(C) be the
Abel-Jacobi map. Then
(1) (Abel) ker Ψ = PDiv(C).

(2) (Jacobi) Ψ is surjective.


Therefore Ψ induces an isomorphism Pic0 (C) ∼
= J(C).
Just as with elliptic curves, if we fix a basepoint O ∈ C, the map iO : C → Div0 (C), P 7→
P − O determines a commutative diagram

Pic0 (C)
Ψ

iO J(C)

ψO
C

However, this time not every map is a bijection. In particular, dim C = 1 < g = dim J(C).
To remedy this, let C g be the g-fold product of C and consider the map

ψ g : C g −→ J(C)
(P1 , . . . , Pg ) 7−→ ψ(P1 ) + . . . + ψ(Pg )

where + denotes the group law on J(C).

559
23.4. Jacobians of Higher Genus Curves Chapter 23. Classical Jacobians

Theorem 23.4.3 (Jacobi). ψ g : C g −→ J(C) is surjective.

There is still work to do to show that the natural map C g → Pic0 (C) is surjective.
It turns out that J(C) is birationally equivalent to the symmetric power C (g) = C g / ∼,
where (P1 , . . . , Pg ) ∼ (Pσ(1) , . . . , Pσ(g) ) for any permutation σ ∈ Sg . Jacobi proved that this
birational equivalence is enough to endow Pic0 (C) ∼ = J(C) with the structure of an algebraic
group.

Theorem 23.4.4. J(C) is an abelian variety.

560
Chapter 24

Maps From Curves

561
24.1. Moduli and Local Symbols Chapter 24. Maps From Curves

24.1 Moduli and Local Symbols


In this chapter, let k be an algebraically closed field and X a curve over k. Fix a finite subset
of points S ⊆ X.
Definition. A modulus on X supported on S, written m, is an assignment of a positive
integer nP > 0 toPeach point P ∈ S. Equivalently, a modulus m may be identified with an
effective divisor P ∈X nP P supported on S.
Definition. For a function g ∈ k(X) and a modulus m supported on S, saying (g) is prime
to S, denoted g ≡ 1 mod m, will mean vP (1 − g) ≥ nP for all P ∈ S. In addition, saying
g ≡ 1 mod m at P will mean vP (1 − g) ≥ nP . Let Div(X − S) denote the set of divisors of
X that are prime to S.
The goal in this chapter is to study maps f : X r S → G where G is a commutative
algebraic group. Any such map extends to a map f : Div(X − S) → G by
X X
D= nP P 7−→ f (D) := nP f (P ).
P ∈X P 6∈S

For instance, this map is given on principal divisors by


X
(g) 7→ f ((g)) = vP (g)f (P ).
P 6∈S

Definition. A modulus for a map f : X r S → G is a modulus m supported on S such


that f ((g)) = 0 for all functions g ≡ 1 mod m.
Definition. Let f : X rS → G be a map to a commutative algebraic group. A local symbol
is the assignment of an element (f, g)P ∈ G for each g ∈ k(X)× and P ∈ X satisfying:
(a) (f, gg 0 )P = (f, g)P + (f, g 0 )P .
(b) If P ∈ S, then (f, g)P = 0 for any g ≡ 1 (mod m) at P .
(c) (f, g)P = vP (g)f (P ) for all P 6∈ S.
×
P
(d) P ∈X (f, g)P = 0 for all g ∈ k(X) .

Proposition 24.1.1. Let f : X r S → G be a map and m any effective divisor supported


on S. Then m is a modulus for f if and only if there exists a local symbol associated to f
and m. Further, if such a local symbol exists, it is unique.
Proof. First suppose a local symbol exists for f, m. If g ≡ 1 mod m, we have
X X
f ((g)) = vP (g)f (P ) = (f, g)P by (c)
P 6∈S P 6∈S
X
=− (f, g)P by (d)
P ∈S
= 0 by (b).

562
24.1. Moduli and Local Symbols Chapter 24. Maps From Curves

Hence m is a modulus for f by definition. Conversely,


P suppose m is a modulus for f , which
may be viewed as an effective divisor m = nP P . We will define a local symbol (f, g)P
×
for all g ∈ k(X) . For any P 6∈ S, condition (c) forces us to define (f, g)P = vP (g)f (P ).
For P ∈ S, there is a function gP ∈ k(X)× such that gP ≡ 1 mod m at any Q ∈ S r {P }
and ggP ≡ 1 mod m at P , by the weak approximation theorem (see ANT). Define the local
symbol at P ∈ S by X
(f, g)P = − vQ (gP )f (Q).
Q6∈S

Now if g, g ∈ k(X) with associated functions gP , gP0 as above, the product gP gP0 satisfies
0 ×

the weak approximation theorem for gg 0 . Thus


X
(f, gg 0 )P = − vP (gP gP0 )f (Q)
Q6∈S
X X
=− vP (gP )f (Q) − vP (gP0 )f (Q)
Q6∈S Q6∈S
0
= (f, g)P + (f, g )P .
So (a) is satisfied. Next, when g ≡ 1 mod m at P ∈ S, the function
P gP = 1 satisfies the
weak approximation theorem for g so by definition (f, g)P = − Q6∈S vP (1)f (Q) = 0. Thus
(b) holds. Finally, note that
X XX
(f, g)P = − vQ (gP )f (Q)
P ∈X P ∈X Q6∈S
XX
=− vQ (gP )f (Q) since G is associative
Q6∈S P ∈X
!
X Y
=− vQ gP f (Q).
Q6∈S P ∈X

gP and k = hg . Then k ≡ 1 mod m so


Q P
Set h = P ∈X Q6∈S vQ (k)f (Q) = 0 by (b). This
gives us
X X
0= (f, g)Q = − vQ (g)f (Q) by (c)
Q6∈S Q6∈S
X X
=− vQ (g)f (Q) + vQ (k)f (Q)
Q6∈S Q6∈S
X g  X
=− vQ f (Q) = − vQ (h)f (Q)
Q6∈S
k Q6∈S
X
= (f, g)P
P ∈X

from above. Hence (d) is satisfied.


Remark. If m, m0 are two moduli supported on S and m ≥ m0 (in the sense of effective
divisors), then g ≡ 1 mod m0 implies g ≡ 1 mod m, which means a modulus for f is not
unique. However, local symbols are unique by Proposition 24.1.1, so they must not depend
on m.

563
24.1. Moduli and Local Symbols Chapter 24. Maps From Curves

Suppose θ : G → G0 is a homomorphism of commutative algebraic groups.


Lemma 24.1.2. Local symbols are natural, i.e. θ(f, g)P = (θ ◦ f, g)P for all g ∈ k(X)× and
P ∈ X.
Proposition 24.1.3. If m is a modulus of f : X r S → G, then m is also a modulus of
θ ◦ f : X r S → G0 .
Proof. Follows from Lemma 24.1.2.
For a morphism of curves π : X → X 0 , set π(S) = S 0 . Then π : X r S → X 0 r S 0 is a
cover with pullback divisors X
π ∗ (P 0 ) = eP P
P →P 0
0 0 0 ∗ 0
P
defined for every P ∈ X r S . Thus f (π (P )) = P →P 0 eP f (P ) is well-defined.
Definition. The trace of a map f : X r S → G along π : X → X 0 is the map
Trπ f : X 0 r S 0 −→ G
P 0 7−→ f (π ∗ (P )).
Proposition 24.1.4. If m is a modulus for f : X r S → G, then there is a modulus m0 for
Trπ f and X
(Trπ f, g 0 )P 0 = (f, g 0 ◦ π)P
P →P 0
for any P 0 ∈ X 0 and g 0 ∈ k(X 0 )× .
Proof. It suffices to show that (Trπ f, g 0 )P 0 satisfies the conditions of a local symbol, since
then Proposition 24.1.1 provides a modulus m0 for Trπ f .
(a) We have
X
(Trπ f, gg 0 )P 0 = (f, (gg 0 ) ◦ π)P
P →P 0
X X
= (f, g ◦ π)P + (f, g 0 ◦ π)P
P →P 0 P →P 0
= (Trπ f, g)P 0 + (Trπ f, g 0 )P 0 .
(b) Suppose g 0 ≡ 1 mod m0 at P 0 . Then for any P → P 0 , vP 0 (1 − g 0 ) ≥ nP 0 and
vP (1 − g 0 ◦ π) = eP vP 0 (1 − g 0 ) ≥ eP n0P ≥ nP , so g 0 ◦ π ≡ 1 mod m at P . Hence (f, g 0 ◦ π)P = 0
for each P ∈ S and if P 6∈ S, vP (g 0 ◦ π) = 0 so (f, g 0 ◦ π)P = vP (g 0 ◦ π)f (P ) = 0. This implies
X
(Trπ f, g 0 )P 0 = (f, g 0 ◦ π)P = 0.
P →P 0

(c) Let P 0 6∈ S 0 . Then


X X
(Trπ f, g 0 )P 0 = (f, g 0 ◦ π)P = vP (g 0 ◦ π)f (P ) since P 6∈ S = π −1 (S 0 )
P →P 0 P →P 0
X X
= eP vP 0 (g 0 )f (P ) = vP 0 (g 0 ) eP f (P )
P →P 0 P →P 0
= vP 0 (g 0 ) Trπ f (P 0 )

564
24.1. Moduli and Local Symbols Chapter 24. Maps From Curves

by definition of the trace.


(d) Finally, for any g 0 ∈ k(X 0 )× ,
X X X X
(Trπ f, g 0 )P 0 = (f, g 0 ◦ π)P = (f, g 0 ◦ π)P = 0
P 0 ∈X 0 P →P 0 P 0 ∈X 0 P ∈X

by (d) for the local symbol (f, −)P . Hence (Trπ f, −)P 0 is a local symbol on X 0 r S 0 .
Let π : X → X 0 be a ramified covering of curves and suppose f 0 : X 0 r S 0 → G0 is a
morphism to a commutative algebraic group G0 .

Definition. For g ∈ k(X), the norm of g is defined to be its norm in the field extension
k(X)/π(k(X 0 )):
Nπ g := Nk(X)/π(k(X 0 )) (g).

Proposition 24.1.5. Suppose f 0 : X 0 r S 0 → G0 has modulus m0 . Then f 0 ◦ π has a modulus


m with local symbol X
(f 0 , Nπ g)P 0 = (f 0 ◦ π, g)P
P →P 0

for all P 0 ∈ X 0 and g ∈ k(X)× .

Proof. By Proposition 24.1.1, it’s enough to show the formula above defines a local Q symbol.
Note that the norm map is continuous and satisfies a local-global principle N = P ∈X NP ,
where NP is the norm of the extension of complete local fields at P → P 0 . Thus if a
modulus m supported on S exists P and g ≡ 1 mod m then P Nπ g ≡ 1 mod m0 . Moreover,
for any g ≡P1 mod m, if (g) = vP (g)P then π(g) = vP (g)π(P ). We must show
vP 0 (Nπ g) = P →P 0 vP (g) for all g ≡ 1 mod m. If w1 , . . . , wd are the distinct extensions of
vP from k(X 0 )vP to k(X) ⊗k(X 0 ) k(X 0 )vP , then we have
d
! d
Y X
vP 0 (Nπ g) = vP 0 Ni (g) = vP 0 (Ni (g))
i=1 i=1
d
X
= wi (g) by algebraic number theory
i=1
= π(vP (g)).

So f 0 ◦ π((g)) = f 0 (Nπ g) = 0 since m0 is a modulus for f 0 and Nπ g ≡ 1 mod m0 . This shows


that if a modulus m supported on S exists, it is a modulus for f 0 ◦ π.
Now assume P 0 6∈ S 0 and P → P 0 , so that P 6∈ S. Then

(f 0 , Nπ g)P 0 = vP 0 (Nπ g)f 0 (P 0 )


X
= vP (g)(f 0 ◦ π)(P )
P →P 0
X
= (f 0 ◦ π, g)P
P →P 0

565
24.1. Moduli and Local Symbols Chapter 24. Maps From Curves

as desired. On the other hand, if P 0 ∈ S 0 , choose h ∈ k(X)× such that hg ≡ 1 mod m at


all P → P 0 and h ≡ 1 mod m at all P ∈ S, P 6→ P 0 , by weak approximation. Then by the
Nπ g
same argument as above, N πh
≡ 1 mod m0 at every P 0 ∈ S 0 and Nπ h ≡ 1 mod m0 at every
0 0
P 6∈ S . Hence

(f 0 , Nπ g)P 0 = (f 0 , Nπ h)P 0
X
=− (f, Nπ h)Q0 by condition (d) for the local symbol
Q0 6∈S 0
X
=− (f 0 ◦ π, h)P by the above
Q0 6∈S 0
X
= (f 0 ◦ π, h)P by condition (d) again
P →P 0
X
= (f ◦ π, g)P .
P →P 0

Therefore the local symbol formula holds for all P 0 ∈ X 0 .

566
24.2. The Main Theorem Chapter 24. Maps From Curves

24.2 The Main Theorem


Let G be a commutative, connected algebraic group and suppose f : X → G is a rational
map, i.e. regular away from a finite set S ⊆ X. We will prove:

Theorem 24.2.1. The map f has a modulus m supported on S.

Example 24.2.2. Let G = Ga = k, the additive group of the field. We can show Theo-
rem 24.2.1 holds for any curve X over k and any map f : X → Ga . For this, it suffices
to prove the existence
 of a local symbol (by Proposition 24.1.1), which we claim is given
by (f, g)P = resP f gdg where resP denotes the residue at P , as in Section 17.5. For each
 
P ∈ S, set nP = 1−vP (f ) and define m = P ∈S nP P . Let’s verify that (f, g)P = resP f gdg
P

defines a local symbol for this map.


(a) follows from the definition of resP .
(b) If g ≡ 1 mod m at P , then vP (1 − g) ≥ nP and

vP (dg) = vP (d(1 − g)) = vP (1 − g) − 1 ≥ nP − 1 = −vP (f ).


   
Thus, since vP (g) = 0, we have vP f gdg ≥ 0 so resP f gdg = 0 by Proposition 17.5.11.
(c) Let P 6∈ S. Then f is regular at P and dg g
has a simple pole at P , so f dg
g
has a simple
pole at P as well. By (4) of Proposition 17.5.11,
   
f dg dg
resP = f (P ) resP = f (P )vP (g).
g g

(d)
 follows
 immediately from the residue formula (Theorem 17.5.13). Hence (f, g)P =
f dg
resP g is a local symbol and m is a modulus for f .

One
 consequence
 of this construction on G = Ga is that the local symbol (f, g)P =
f dg
resP g commutes with the Frobenius:

Proposition 24.2.3. If char k = p > 0, then for any map f : X r S → k, g ∈ k(X)× and
P ∈ X,  p    p
f dg f dg
resP = resP .
g g
Proof. Let θ : Ga → Ga , x 7→ xp be the Frobenius map. Then f factors through θ:
f
X rS Ga

Ga

Apply Lemma 24.1.2.

567
24.2. The Main Theorem Chapter 24. Maps From Curves

Example 24.2.4. Let G = Gm = k × be the multiplicative group of the field and suppose
f : X 99K Gm is a rational map and S is the finite set ofPpoints of X at which f is not
regular, i.e. the set of zeroes and poles of f . We claim m = P ∈S P is a modulus for f , that
is, nP = 1 for all P ∈ S works. To show this, we show the assignment
fn
(f, g)P = (−1)mn (P ) where n = vP (g), m = vP (f )
gm
 n
is a local symbol for f . Note that vP gfm = mn − mn = 0 so this symbol is well-defined.
Also, (f, g)P satisfies the following properties:

(i) (f, g)P (g, f )P = 1

(ii) (−f, f )P = 1

(iii) (1 − f, f )P = 1.
P
To prove this is a local symbol corresponding to the modulus m = P ∈S P , keep in mind
that the abelian group structure on Gm is multiplicative with identity 1.
(a) becomes (f, gg 0 )P = (f, g)P (f, g 0 )P in multiplicative notation, which is clear from the
definition of the symbol.
(b) Suppose vP (1−g) ≥ nP = 1. Then vP (g) = 0 and g(P ) = 1, so (f, g)P = g(P )−m = 1,
the multiplicative identity in Gm .
(c) Take P 6∈ S. Then by the definition of S, vP (f ) = 0 and we have (f, g)P = f (P )n =
f (P )vP (g) , the correct equality in multiplicative notation.
(d) View g ∈ k(X)× as a rational function g : X 99K P1 . If g ≡ a 6= 0 is constant, then
for all P ∈ X, (f, g)P = g(P )−vP (f ) = a−vP (f ) so we have
Y Y P
(f, g)P = a−vP (f ) = a P ∈X −vP (f ) = a− deg((f ))
P ∈X P ∈X

which is a0 = 1 by Corollary 17.2.6. So the formula holds for constant functions. More
generally, any g ∈ k(X)× may be viewed as a branched cover g : X → P1 . Set E =
k(P1 ) = k(t) and F = k(X) so that g induces a field extension F/E with norm map
N = NF/E : F × → E × .

Lemma 24.2.5. Let t denote the identity map on E. Then for all P ∈ P1 ,
Y
(f, g)Q = (N f, t)P .
Q→P

×
Proof. Extend the definition of (f, g)P to the completions E d and Fb× = k(X)
bP = k(t) [ .
P Q Q
It is immediate that Q
(f, g)P is still multiplicative in both arguments. By algebraic number
theory, N = NF/E = Q→P NQ where NQ = NFbQ /EbP . Thus we have
Y
(N f, t)P = (NQ f, t)P
Q→P

568
24.2. The Main Theorem Chapter 24. Maps From Curves

so it suffices to show (f, g)Q = (NQ f, t)P for all Q → P . Fix such a Q ∈ g −1 (P ). By
multiplicativity of the local norm NQ , we may assume f and g are uniformizers of FbQ and
EbP , respectively, i.e. that vQ (f ) = vP (g) = 1, FbQ = k((f )) and E
bP = k((g)). Then

(NQ f )n NQ f
(NQ f, g)P = (−1)nm m
(P ) = − (P )
g g

since vP (NQ f ) = vQ (f ) = 1. On the other side, vQ (g) = e = [FbQ : E


bP ], the ramification
index of Q/P , so viewing E bP ⊂ FbQ , we have

fe ef
e
(f, g)Q = (−1)em (Q) = (−1) (Q).
gm g

bP , we see that NQef (P ) = (−1)e−1 so we get


Considering the minimal polynomial of f over E f
(f, g)Q = (NQ f, t)P as desired.
Y
Lemma 24.2.6. For every function f 0 ∈ k(P1 )× , (f 0 , t)P = 1.
P ∈P1

Proof. Write f 0 = µ − λ)nλ for µ ∈ k × . Then


Q
λ∈k (t
!nλ
Y
(f 0 , t)P = (µ, t)P (t − λ, t)P ,
λ∈k

(µ, t)P = (µ(P ))vP (g) = µn and


Q
P ∈P1 (µ, t)P
= 1. For the remaining pieces, if λ = 0 we have
(
2 2 −1, P = 0, ∞
(t, t)P = (−1)n 1(P ) = (−1)n =
1, P 6= 0, ∞.
Q
Thus P ∈P1 (t, t)P = 1. On the other hand, for λ 6= 0, we have


 −λ, P =0
n 1

(t − λ)  , P =λ
(t − λ, t)P = (−1)nm m
(P ) = λ
t 
 −1, P =∞

6= 0, ∞, λ.

1, P
0
Q Q
Thus P ∈P1 (t − λ, t)P = 1 so we get P ∈P1 (f , t)P = 1.
These lemmas give us condition (d) and therefore show that Theorem 24.2.1 holds for
Gm , since Y Y Y Y
(f, g)Q = (f, g)Q = (N f, g)P = 1.
Q∈X Q→P P ∈P1 P ∈P1

Proposition 24.2.7. Suppose f, g ∈ k(X)× and Sf , Sg are, respectively, their sets of zeroes
and poles. If Sf ∩ Sg = ∅, then f ((g)) = g((f )).

569
24.2. The Main Theorem Chapter 24. Maps From Curves

n
Proof. Let (f, g)P = (−1)nm gfm (P ) be the local symbol from Example 24.2.4. Then
Y Y
f ((g)) = f (P )vP (g) = (f, g)P
P 6∈Sf P 6∈Sf
Y Y Y 1
and g((f )) = g(P )vP (f ) = (g, f )P = P 6∈ Sg .
P 6∈Sg P 6∈Sg
(f, g)P

Also, for all P outside Sf ∪ Sg , (f, g)P = 1 so putting these together, we have
Y Y Y 1
1= (f, g)P = (f, g)P (f, g)P = g((f )).
P ∈X P ∈Sf P ∈Sg
f ((g))

Hence f ((g)) = g((f )).


Proposition 24.2.8. If G is any commutative algebraic group and f : X r S → G is a
regular map, then for all ramified covers π : X → X 0 , Trπ f is a regular map X 0 r S 0 → G.
Proof. Let n = deg π and consider the n-fold symmetric product X (n) (see Section 24.3 for
more details). We will see that X (n) is irreducible, normal and nonsingular if X is. Consider
the pullback map

π ∗ : X 0 −→ X (n)
X
P 0 7−→ eP P.
P →P 0

Set Y = X r S, Y 0 = X 0 r S 0 and define

F : Y n −→ G
n
X
(P1 , . . . , Pn ) 7−→ f (Pi ).
i=1

Clearly the sum on the right is invariant under the symmetric group action, so F descends
to a map F 0 : Y (n) → G. Consider the diagram
Y Yn
F
π
π∗ F0
Y0 Y (n) G

Trπ f

Since F 0 is regular by definition, it suffices to show π ∗ is regular which we will do in Sec-


tion 24.3.
Let g ∈ k(X)× , m = P ∈S nP P and suppose vP (g) ≥ nP for all P ∈ S (this condition will
P
be denoted g ≡ 0 mod m). Then by Proposition 24.2.8, Trg f is a regular map P1 r{0} → G.

570
24.2. The Main Theorem Chapter 24. Maps From Curves

P
Proposition 24.2.9. Let m = P ∈S nP P be a modulus and f : X r S → G a regular map.
Then m is a modulus for f if and only if for all g ∈ k(X)× such that g ≡ 0 mod m, Trg f
is constant.

Proof. For a 6= ∞, let (g)a be the divisor of zeroes of g − a and let (g)∞ be the divisor of
poles of g. Then for all a ∈ P1 r {0},
!
X
Trg f (a) = f eP P = f ((g)a ).
P →a

We now proceed to the proof.


( =⇒ ) Suppose m is a modulus for f . For a ∈ P1 r {0, ∞}, write g − a = −a 1 − ag


and notice that since −a ∈ k × , we have


  g  g 
vP 1 − 1 − = vP = vP (g) ≥ nP
a a
g
so in particular 1 − a
≡ 1 mod m. Applying f to the divisor (g − a), we get
 g 
0 = f ((g − a)) = f 1 − = f ((g)a − (g)∞ ) = f ((g)a ) − f ((g)∞ ),
a
so f ((g)∞ ) = f ((g)a ) = Trg f (a). As this holds for any a ∈ P1 r {0}, Trg f is constant.
( ⇒= ) Suppose Trg f is constant for every g ≡ 0 mod m. Setting h = 1 − g, we have
h ≡ 1 mod m and (h) = (g)1 − (g)∞ , so

f ((h)) = f ((g)1 − (g)∞ ) = f ((g)1 ) − f ((g)∞ ) = Trg f (1) − Trg f (∞) = 0.

Thus m is a modulus for f .


Let us now move towards a proof of Theorem 24.2.1. We will prove it in the case that
char k = 0; although it also holds when char k > 0, the proof is quite different and can be
found in Serre. Let f : X r S → G be a regular map, set r = dim G and take ω1 , . . . , ωr
to be a basis of degree 1 invariant differential forms on G, that is, a basis of the Lie algebra
associated to G. Set αi = f ∗ (ωi ) for each 1 P
≤ i ≤ r. For P ∈ S, choose nP ≥ 0 such that
vP (αi ) ≥ −nP for all 1 ≤ i ≤ r and set m = P ∈S nP P .

Theorem 24.2.10. m is a modulus for f .

Proof. Fix g ≡ 0 mod m. By Proposition 24.2.9 it is enough to show that the function
f 0 = Trg f is constant. To do this, we will show (f 0 )∗ ωi = 0 for all 1 ≤ i ≤ r. To see why this
implies the trace function is constant, notice that 0 = (f 0 )∗ ωi (P ) = ωi (f 0 (P )) so the map of
tangent spaces induced by f 0 is the zero map and thus f 0 is constant.
Our first step is to show (f 0 )∗ ωi = Trg ωi for all 1 ≤ i ≤ r. Let π : Y → X be a finite
Galois cover and consider the Galois, separable map g ◦ π : Y → P1 . Since it is separable, the
induced map ω 7→ (g ◦ π)∗ ω is injective. Thus if we can show that π ∗ g ∗ (f 0 )∗ ωi = π ∗ g ∗ Trg ωi ,
it will follow that (f 0 )∗ ωi = Trg ωi . To do this, we observe that all lifts along π are Galois
conjugate, or more generally:

571
24.2. The Main Theorem Chapter 24. Maps From Curves

π g
Proposition 24.2.11. Suppose Y → − X → − X 0 are Galois covers of curves with G =
Gal(k(Y )/k(X 0 )) and H = Gal(k(Y )/k(X)) ≤ G. Then for all P ∈ X,
X
g −1 (g(P )) = [k(X) : k(X 0 )]i π ◦ σj (Q)
j

where [k(X) : k(X 0 )]i denotes the inseparable degree, π(Q) = P and {σj } is a transversal of
G/H.
Proof. Set h = g ◦ π and note that
X X
h−1 (g(P )) = nσ σ(Q) = nσ(Q)
σ∈G σ∈G

for a fixed n ∈ N, since π is Galois. On the other hand, h(h−1 (P 0 )) = [k(Y ) : k(X 0 )]P 0 for
all P 0 ∈ X 0 , so we must have h−1 (g(P )) = σ∈G σ(Q). Now to get g −1 (g(P )), apply π:
P

X
π(h−1 (g(P ))) = π ◦ σ(Q)
σ∈G
XX
= π ◦ α ◦ σj (Q)
α∈H j
XX
= π ◦ σj (Q) since α ∈ H
α∈H j
X
= [k(Y ) : k(X)] π ◦ σj (Q).
j

On the other hand, π(h−1 (g(P ))) = [k(Y ) : k(X)]g −1 (g(P )) so dividing by [k(Y ) : k(X)],
we get the result.
Corollary 24.2.12. For any f ∈ k(X) and g ∈ k(X)× , (Trg f ) ◦ g ◦ π = j f ◦ π ◦ σj .
P

π g
− P1 , we have f 0 ◦ g ◦ π = j f ◦ π ◦ σj , so
P
Applying this to our maps Y →− X→
X
π ∗ g ∗ Trg αi = σj∗ π ∗ αi .
j

On the other hand,


X X
π ∗ g ∗ (f 0 )∗ ωi = σj∗ π ∗ f ∗ (ωi ) = σj∗ π ∗ (αi )
j j

by definition of the αi . Thus π ∗ g ∗ (f 0 )∗ ωi = π ∗ g ∗ Trg ωi as claimed.


Finally, we use this to show each (f 0 )∗ ωi = 0. We start by showing (f 0 )∗ ωi has at most
a simple pole at 0. Let t : P1 → P1 denote the identity map once again and suppose
v0 ((f 0 )∗ ωi ) = −m − 1 for m ∈ Z. Then v0 (tm (f 0 )∗ ωi ) = m − m − 1 = −1 so by definition
res0 (tm (f 0 )∗ ωi ) 6= 0. The function g : X → P1 induces a field extension k(P1 ) ,→ k(X), t 7→ g,
in which
Tr(g m αi ) = Tr(tm αi ) = tm Tr(αi ) = tm (f 0 )∗ αi .

572
24.2. The Main Theorem Chapter 24. Maps From Curves

By the residue formula (Theorem 17.5.13),


X
resQ (g m αi ) = res0 (Tr(g m αi )) = res0 (tm (f 0 )∗ ωi ) 6= 0.
Q→0

For any Q → 0, if Q 6∈ S, g and αi are regular at Q, so g m αi is regular at Q. If Q ∈ S, we


have
vQ (g m αi ) = mvQ (g) + vQ (αi ) ≥ mnQ − nQ = (m − 1)nQ
since g ≡ 0 mod m. Since the residue formula holds, we must have m − 1 ≤ −1, which
means m ≤ 0, that is, (f 0 )∗ ωi has at most a simple pole at 0.
By Proposition 24.2.8, f 0 = Trg f is regular on P1 r {0}. It follows that (f 0 )∗ ωi is regular
on P1 r {0}. Moreover, since v0 ((f 0 )∗ ωi ) ≥ −1 and the residue theorem tells us that
X
res0 ((f 0 )∗ ωi ) = resP ((f 0 )∗ ωi ) = 0,
P ∈P1

(f 0 )∗ ω must be regular at 0 as well. Hence (f 0 )∗ ωi = 0, which finishes the proof of Theo-


rem 24.2.10.

573
24.3. A Word About Symmetric Products Chapter 24. Maps From Curves

24.3 A Word About Symmetric Products


Let X be an algebraic variety such that every finite subset of X is contained in an affine
open subset (for example, this is true when X is projective). Let X (n) = X n /Sn denote the
n-fold symmetric product of X. This can be given the structure of a variety. More generally:

Proposition 24.3.1. If G is a finite group acting on a variety V such that every orbit of G
is contained in any affine open subset of V , then the quotient V /G is a variety.

Proof. Let S1 , . . . , Sk denote the orbits. Since each orbit is invariant under the G-action, we
can cover V by a collection of affine open sets {Ui0 } with Si ⊆ Ui0 . Set Ui = Ui0 /G. Then Ui
is affine: if Ui0 = Spec Ai for a ring Ai , then Ui ∼= Spec AG i , the affine scheme on the ring of
G-invariants of Ai . It is easy to see that V /G is covered by these affine open sets Ui . 
T
Finally, we must show the diagonal map ∆V /G is closed. It suffices to show ∆V /G U
i,j i × Uj
is closed in Uk × U` for all k, `. Consider the commutative diagram
π
V V /G

∆V ∆V /G

V ×V V /G × V /G
π×π
T 
0 0
Since V is a variety, ∆V U
i,j i × U j is closed in Uk0 × U`0 for all k, `. Since π : V → V /G
T 
0 0
is a finite map, (π × π) ◦ ∆V U i × Uj is closed in Uk × U` , but by commutativity, this
T  i,j
is precisely ∆V /G i,j Ui × Uj .

Remark. If dim V > 1, V /G may be a singular variety. For example, when X is a curve
and dim X > 1, the symmetric product X (n) is always singular.

Proposition 24.3.2. Suppose X is a variety and n ≥ 2. Then

(a) If X is irreducible, so is X (n) .

(b) If X is normal, so is X (n) .

Define the map δn : X → X (n) which sends P 7→ nP := [P, P, . . . , P ]. This factors as


δn = π ◦ ∆ : X → X n → X (n) . An important result for us in Chapter 26 will be:

Proposition 24.3.3. When char k = 0, the map δn : X → δn (X) ⊆ X (n) is a birational


homeomorphism.

574
Chapter 25

Curves with Singularities

575
25.1. Normalization Chapter 25. Curves with Singularities

25.1 Normalization
To discuss singular curves, we introduce a process of “normalization” to get rid of singularities
in some way. The goal will be to prove a Riemann-Roch theorem and describe a canonical
divisor by way of meromorphic differentials on a singular curve.
As a motivating example, consider the irreducible algebraic curve X 0 defined by the
equation y 2 − x3 = 0 on an affine patch and compactified by adding a point O at infinity.
Let O0 be the structure sheaf of X 0 , which has local rings
(
k[x2 , x3 ], P = O
OP0 =
k[x]P , P 6= O.

Then k[x2 , x3 ] is not a regular local ring (the dimension of the tangent space at P = O is 2;
see Section 6.2), so X 0 is singular at O.
In general, let X 0 be an irreducible algebraic curve with singular locus S 0 and field of
rational functions K = k(X 0 ). For each Q ∈ X 0 , let OQ denote the integral closure of OQ
0
in
K.

Lemma 25.1.1. For any irreducible algebraic curve X 0 ,

(a) OQ is a finitely generated k-algebra for all Q ∈ X 0 .


0 0
(b) [OQ : OQ ] < ∞ with OQ = OQ if and only if Q is a nonsingular point of X 0 .

(c) The assignment Q 7→ OP defines a sheaf of k-algebras on X 0 .

Definition. A normalization of X 0 is an algebraic curve X and a finite rational morphism


ϕ : X → X 0 such that

(a) X is nonsingular and normal.

(b) ϕ is a bijection X r S → X 0 r S 0 , where S = ϕ−1 (S 0 ).

(c) ϕ is a birational equivalence.

If U 0 ⊆ X 0 is an open affine subset, let A0 = k[U 0 ] be the corresponding coordinate


ring and denote by A the integral closure of A0 in K, which by Lemma 2.1.2 is also finitely
generated over k. This allows us to explicitly normalize X 0 on affine patches: replace each
U 0 with an affine curve U corresponding to AS and glue these U together using the same
gluing information as an affine cover X 0 = U 0 to obtain X. For each U 0 , there is a
0 0
morphism U → U induced by the inclusion A ,→ A which glue together to give a morphism
ϕ : X → X 0 . Note that for an open V ⊆ U with p(V )∩S = ∅, the projection p|V : V → p(V )
is a birational isomorphism.

Lemma 25.1.2. The X just constructed is an algebraic curve. Moreover, if X 0 is a projective


curve then so is X.

For any Q ∈ S 0 , note that OQ = P ∈ϕ−1 (Q) OP where OP is the local ring of X at P .
T

576
25.1. Normalization Chapter 25. Curves with Singularities

Theorem 25.1.3. For every algebraic curve X 0 , there exists a normalization ϕ : X → X 0 .

Definition. The conductor of the normalization X → X 0 is the sheaf c defined as the


annihilator of O/O0 as O0 -modules, or equivalently, as the assignment of local rings
0 0 0
Q 7→ cQ := {f ∈ OQ | f OQ ⊆ OQ } = AnnOQ0 (OQ /OQ ).

Given the singular locus S 0 ⊆ X 0 , we define a relation R on S = ϕ−1 (S 0 ) by aRb if


ϕ(a) = ϕ(b) in S 0 . Then normalization may be viewed as the assignment

{algebraic curves over k} −→ {nonsingular curves /k} × {finite sets} × {relations}


X 0 7−→ (X, S, R).

For any Q ∈ S 0 , define the radical of OQ :

rQ = {f ∈ OQ | f (P ) = 0 for all P ∈ ϕ−1 (Q)}.


0
Then we have k + cQ ⊆ OQ ⊆ k + rQ ⊆ OQ and each quotient is finite dimensional over k.
n
Hence for some n ≥ 0, k + rQ ⊆ k + cQ .

Example 25.1.4. For X 0 defined by y 2 − x3 = 0 and the point Q = O, the chain above
(with n = 2) is
0
k + r2Q ⊆ k + cQ ⊆ OQ ⊆ k + rQ ⊆ OQ
2 2 3 2 3
k + hxi ⊆ k[x , x ] = k[x , x ] ⊆ k + hxi ⊆ k[x].

For every singular curve X 0 , we have constructed a nonsingular normalization X → X 0 .


Conversely, every nonsingular curve can be made into a singular curve by taking the quotient
with respect to an equivalence relation at a finite set of points, giving an assignment in the
opposite direction. Let (X, S, R) be a triple consisting of a nonsingular curve X, a finite set
S ⊆ X and a relation R on S. Define X 0 = (X r S) ∪ S/R. To make X 0 into a singular
curve, we define a sheaf of rings O0 on X 0 as follows. For Q ∈ X 0 , set OQ 0
= OQ if Q 6∈ S
and OQ to be any ring such that k + rQ ⊆ OQ ⊆ k + rQT⊆ OQ . Next, we declare U ⊆ X 0 to
0 n 0

be open if its complement in X 0 is finite. Set O0 (U ) = Q∈U OQ .

Lemma 25.1.5. (X 0 , O0 ) is an algebraic curve which is singular at every Q ∈ S/R.

Proof. The sheaf axioms for O0 are easily checked. To prove X 0 is a curve, it suffices to
prove the case when X is affine, say X = Spec A. In this case let A0 = Q∈X 0 OQ 0
T
and set
0 0 0
Y = Spec A . It is clear that A ⊆ A. Moreover, by the chain condition on OQ , there exists
an f ∈ A such that f A ⊆ A0 ⊆ A so in particular A0 is finitely generated as a k-algebra.
Since A is an A0 -algebra and is finitely generated, it follows from Lemma 2.1.2 that A/A0 is
an integral extension. Hence dim A0 = dim A = 1 by Theorem 2.3.7 so Y is a curve. In fact,
since A is integrally closed (X is nonsingular), A is precisely the integral closure of A0 in
its fraction field, which shows X = Spec A is the normalization of Y . Finally, one can show
that X 0 ∼= Y so X 0 is also an affine curve, finishing the proof.

577
25.1. Normalization Chapter 25. Curves with Singularities

Theorem 25.1.6. The assignment

{algebraic curves over k} −→ {nonsingular curves /k} × {finite sets} × {relations}


X 0 7−→ (X, S, R)

is a bijection.

Therefore we may treat singular curves formally as nonsingular curves equipped with
‘singular data’ in the form of S ⊆ X with a relation R on S.

Example
P 25.1.7. Let X be a complete, irreducible, nonsingular curve over k. For a modulus
m= nP P on X with support S = {P ∈ X | nP 6= 0} (we will assume |S| ≥ 2), define
the singular curve X 0 associated to the data (X, S, R) where R identifies the entire set S to
a point Q ∈ X 0 . Let cQ = {f ∈ K | f ≡ 0 mod m} and OQ 0
= k + cQ , so that c is the
0
conductor of O in O. We denote the resulting singular curve by Xm , which has a unique
singular point Q.
In our example X 0 : y 2 − x3 = 0, the singular curve X 0 is isomorphic to Xm where X is
the normalization of X 0 and m = 2O.

578
25.2. Singular Riemann-Roch Chapter 25. Curves with Singularities

25.2 Singular Riemann-Roch


Let X 0 be a complete, irreducible algebraic curve over k with normalization
P X and let g be
0
the genus of X. For each Q ∈ X, let δQ = dim(OQ /OQ ) and set δ = Q∈S δQ . Also put
π = g + δ. For a divisor D on X which is prime to S, let L(D) be the sheaf defined by D
as in Section 17.3. On X 0 , setting
(
0 L(D)Q , Q 6∈ S 0
L (D)Q = 0
OQ , Q ∈ S0

defines a sheaf L0 (D) since X r S → X 0 r S 0 is a birational isomorphism. In analogy with


the nonsingular Riemann-Roch theory, define

L0 (D) = H 0 (X 0 , L0 (D))
I 0 (D) = H 1 (X 0 , L0 (D))
`0 (D) = h0 (X 0 , L0 (D)) = dim L0 (D)
i0 (D) = h1 (X 0 , L0 (D)) = dim I 0 (D).

When X 0 = Xm for a modulus m, we will also write these as Lm (D), Im (D), `m (D) and im (D),
respectively.

Proposition 25.2.1. For any divisor D on X, `0 (D) and i0 (D) are finite.
0
Proof. Both L(D) and OQ are coherent sheaves, so this follows from the proof of Proposi-
tion 17.3.11.

Proposition 25.2.2. For all q ≥ 0, H q (X, O) = H q (X 0 , O).

Proof. See FAC.

Theorem 25.2.3 (Riemann-Roch for Singular Curves). Let X 0 be a singular curve with
normalization X. Then for any divisor D on X prime to S,

`0 (D) − i0 (D) = deg(D) + 1 − π.

Proof. For a divisor D prime to S and any point P 6∈ S 0 , let Q0 = L0 (D + P )/L0 (D) be the
cokernel sheaf in the exact sequence

0 → L0 (D) → L0 (D + P ) → Q0 → 0.

Then as in the proof of Theorem 17.3.12, Q0 is 0 away from P and has dimension 1 at
P . Thus the same argument as in that proof shows that χ0 (D + P ) = χ0 (D) + 1, where
χ0 (D) = `0 (D) − i0 (D). Thus it suffices to prove the theorem for D = 0. In this case, we have
L0 (D) = O0 and since O0 is the pushforward of O along the normalization map X → X 0 ,
there is a short exact sequence

0 → O0 → O → O/O0 → 0.

579
25.2. Singular Riemann-Roch Chapter 25. Curves with Singularities

Thus by additivity, χ0 (O0 ) + χ0 (O/O0 ) = χ0 (O). Further, as in the proof of Theorem 17.3.12,
O/O0 is a skyscraper sheaf concentrated on S 0 , so
X X
dim H 0 (X 0 , O/O0 ) = dim(OQ /OQ0
)= δQ = δ.
Q∈S Q∈S

This shows that χ0 (O0 ) + δ = χ0 (O), but by the standard Riemann-Roch theorem, χ(O) =
1 − g so to finish, it suffices to observe χ0 (O) = χ(O) by Proposition 25.2.2.

Corollary 25.2.4. π = i0 (0) = h1 (X 0 , O0 ).

580
25.3. Differentials on Singular Curves Chapter 25. Curves with Singularities

25.3 Differentials on Singular Curves


Once again, let X 0 be a complete, irreducible curve over k with normalization X. For Q 6∈ S 0 ,
let Ω0Q = ΩQ = Dk (OX,Q ), the ring of regular differential forms on X at Q. For Q ∈ S 0 , set

Ω0Q = {ω ∈ Dk (OX,Q
0 0
) | hf, ωiQ = 0 for all f ∈ OQ },
P
where hf, ωiQ = P ∈ϕ−1 (Q) resP (f ω) is the residue pairing at Q.

Lemma 25.3.1. For any Q ∈ X 0 ,

(a) ΩQ ⊆ Ω0Q .
0
(b) hOQ , Ω0Q iQ = 0.

(c) hOQ , ΩQ iQ = 0.
0
Thus the vector spaces OQ /OQ and Ω0Q /ΩQ are dual to each other with respect to h·, ·iQ .

Example 25.3.2. When X 0 = Xm for a modulus m =


P
nP P , with unique singular point
0
Q, the condition hf, ωi = 0 for all f ∈ OQ is equivalent to h1, ωi = 0 and vP (ω) ≥ −nP for
all P ∈ S.

Let ϕ : X → X 0 be the normalization map and let g : X r S → X 0 r S 0 be the


corresponding birational equivalence. For an algebraic group G and a map f : X r S → G,
define the trace of f along g (as in Section 24.1) to be
X
Trg f (Q) = f (P ).
P →Q

Extend this to a meromorphic differential ω on X by putting Trg ω(Q) = f where ω = f dtQ .

Proposition 25.3.3. Let ω be a meromorphic differential on X. Then ω ∈ Ω0Q for all


Q ∈ X 0 if and only if Trg ω = 0 for all g ∈ OQ
0
and Q ∈ S 0 .

Now for a divisor D on X prime to S, set


(
Ω(D)Q , Q 6∈ S
Ω0 (D)Q =
Ω0Q , Q ∈ S.

This defines a sheaf of differentials Ω0 (D) on X 0 for which Ω0 (D) = H 0 (X 0 , Ω0 (D)). As with
Serre duality, we have:

Theorem 25.3.4. For any divisor D on X prime to S, Ω0 (D) is isomorphic to the dual of
I 0 (D) = H 1 (X 0 , L0 (D)).

581
25.3. Differentials on Singular Curves Chapter 25. Curves with Singularities

Proof. Let R denote the répartitions of X and define the subsets

R0 = {r ∈ R | rP1 = rP2 whenever ϕ(P1 ) = ϕ(P2 )} ⊆ R


R0 (D) = {r ∈ R0 | vP (rP ) + vP (D) ≥ 0 for P 6∈ S and rP ∈ OP0 for P ∈ S} ⊆ R0 .

As in Section 17.5, Dk (K) is isomorphic to the dual of R/K, so for any ω ∈ Dk (K), ω ∈ Ω0 (D)
if and only if ω|R0 (D) = 0. Hence Ω0 (D) ∼
= (R/(k(X) + R0 (D)))∗ . Consider the short exact
sequence of sheaves
0 → L0 (D) → k(X) → Q → 0
where k(X) is the constant sheaf and Q is the sheafification of k(X)/L0 (D). Note that
since X 0 is irreducible, H 0 (X 0 , k(X)) = k(X) and H 1 (X 0 , k(X)) = 0. Taking the long exact
sequence in sheaf cohomology, we get

k(X) → H 0 (X 0 , Q) → I 0 (D) → 0.

As in the nonsingular case, Q is a skyscraper sheaf which allows us to write

H 0 (X 0 , Q) ∼ (k(X)/L0 (D))Q ∼
M
= = R0 /R0 (D).
Q∈X 0

Further, k(X) → R0 /R0 (D) is injective since k(X) is a field, so the second exact sequence
gives us I 0 (D) ∼
= R0 /(k(X) + R0 (D)). One finishes by showing R/(k(X) + R0 (D)) and
R /(k(X) + R (D)) have the same dual, from which it follows that Ω0 (D) ∼
0 0
= I 0 (D)∗ .

Corollary 25.3.5. i0 (D) = dimk Ω0 (D).

Corollary 25.3.6. If Ω0 is the vector space of regular differential forms on X 0 , then dimk Ω0 =
π = g + δ.

Assume Ω0 is a locally free sheaf on X 0 . Set K 0 = K + c where c is the conductor of the


map X → X 0 . Then K 0 is a canonical divisor on X 0 and we obtain the full Riemann-Roch
theorem for singular curves:

Theorem 25.3.7. Suppose Ω0 is locally free. For any divisor D on X prime to S,

`0 (D) − `0 (K 0 − D) = deg(D) + 1 − π.

582
Chapter 26

Constructing Jacobians

Recall that every algebraic curve X over the complex numbers has a Jacobian J(X) = Cg /Λ
where Λ is a full lattice in Cg . In this chapter, we construct Jacobians for more general
algebraic curves, including so-called generalized Jacobians for singular curves.

583
26.1. The Jacobian of a Curve Chapter 26. Constructing Jacobians

26.1 The Jacobian of a Curve


Let X be a complete, nonsingular algebraic curve of genus g over a field k. We will prove
the following result:

Theorem. There exists an abelian variety J(X) of dimension g over k and a rational map
ϕ : X → J(X) such that the group law of J(X) corresponds to a ‘rational group law’ on the
gth symmetric power of X.

We adopt Weil’s conventions in Foundations of Algebraic Geometry (FAC):

ˆ Assume X is a projective algebraic curve which is irreducible and nonsingular over k.

ˆ Assume for the moment k is algebraically closed, although the results will hold in
general.

ˆ Fix a universal domain Ω, which is an algebraically closed field extension Ω/k of infinite
transcendence degree within which all other extensions we consider will be contained.

ˆ A point of X will be treated as a point with coordinates in Ω.

ˆ The field of definition k(P ) for a point P ∈ X is the finite subextension k(P )/k of Ω/k
generated by the coordinates of P .

ˆ For a subfield Ω ⊇ L ⊇ k and a point P ∈ X, we will say P is L-rational if k(P ) ⊆ L.


P
Let D = nP P be a divisor on X and choose an affine open set U ⊆ X containing the
support of D. Then U = Spec A for some Dedekind ring A/k. Set AΩ = A ⊗k Ω so that the
Ω-points
P U (Ω) Q correspond to the ring AΩ . Then D corresponds to a fractional ideal a of AΩ
via nP P ↔ P nP .

Lemma 26.1.1. There exists a smallest field extension K/k such that a is generated by
polynomials with coefficients in K.

Definition. Such a field K is called the field of rationality of D, written K = k(D). For
any field L/k, we say D is rational over L if k(D) ⊆ L.

Lemma 26.1.2. If D1 , D2 ∈ Div(X) are rational over L/k, then D1 − D2 is also rational
over L.

Proof. This follows from the fact that k(D1 − D2 ) ⊆ k(D1 ) ∩ k(D2 ).

Lemma 26.1.3. Let P1 , . . . , Pn be independent P


generic points of X, i.e. k(P1 , . . . , Pn )/k has
transcendence degree n. Then the divisor D = Pi has field of rationality

k(D) = k(P1 , . . . , Pn )s ,

the subfield of k(P1 , . . . , Pn ) fixed by all permutations of {P1 , . . . , Pn }.

584
26.1. The Jacobian of a Curve Chapter 26. Constructing Jacobians

Proof. Set K = k(P1 , . . . , Pn )s . If σ is an automorphism of Ω/K, then σ(D) = D so D is


rational over some purely inseparable extension of K. Since k(P1 , . . . , Pn )/K is separable (it
is Galois), D must be rational over K, i.e. k(D) ⊆ K. On the other hand, any automorphism
σ ∈ Aut(Ω/k(D)) fixes D and hence fixes K, so K is purely inseparable over k(D). However,
k(D)(P1 , . . . , Pn )/k(D) is separable, so K/k(D) is also separable. Hence K = k(D).
Take D ∈ Div(X). Following the conventions of Chapter 17, let L(D) = {f ∈ k(X) |
(f ) + D ≥ 0} be the Riemann-Roch space of D and set `(D) = dimk L(D). By the Riemann-
Roch theorem (17.6.4),
`(D) = deg(D) + 1 − g + `(C − D)
where C is any divisor in the canonical class of X.

Lemma 26.1.4. Suppose `(C − D) 6= 0 and D is rational over some L/k. Then for any
generic point P ∈ X which is L-rational, we have `(C − (D + P )) = `(C − D) − 1.

Proof. If this were not the case, Riemann-Roch would imply `(C − (D + P )) = `(C − D)
but then every f ∈ L(C − D) vanishes at P , which is impossible since P is generic.

Lemma 26.1.5. Fix a degree 0 divisor D0 ∈ Div0 (X) and let P1 , . . . , Pg be P independent
generic points of X defined over some L/k for which P D0 is rational. Set D = Pi . Then
there is a unique effective divisor of the form E = Qi , where Q1 , . . . , Qg are independent
generic points, such that E ∼ D0 + D. Moreover,

L(P1 , . . . , Pg )s = L(Q1 , . . . , Qg )s .

Proof. By the Riemann inequality (17.3.8), deg(D0 ) = 0 implies `(D0 ) = 0 or 1. Using


Riemann-Roch and Lemma 26.1.4, if `(D0 ) = 0, then `(C −D0 ) = g−1 and `(C −(D0 +D)) =
0. On the other hand, if `(D0 ) = 1, then `(C − D0 ) = g and `(C − (D0 + D)) = 0 once
again. This means there is a unique effective divisor E in the complete linear system for
D0 + D. Moreover, since D0 and D are rational over L(P1 , . . . , Pg )s , Lemma 26.1.2 shows
that D0 + D is as well. Thus E is rational over L. It follows that D and E have the same
field of rationality
P over L, and since the transcendence degree of this extension is g, we must
have E = Qi for independent generic points Q1 , . . . , Qn . The final statement about fixed
fields is immediate from Lemma 26.1.3.
Assume X has an effective divisor D with deg(D) = g which is rational over k. To
construct the Jacobian of X, we define what Weil calls a normal composition law (see FAC)
on the field k(D) = k(P1 , . . . , Pg )s . Since k(D)/k has transcendence degree g, we may choose
a model V /k for k(D), i.e. an open subset of X g with function field k(D) and dim V = g.
There is a rational map ϕ : X (g) → V on the symmetric product (see Section 24.3) such P that
for any generic point q ∈ V , ϕ−1 (q) can be uniquely identified with P a divisor D 0
= Pi for
P1 , . . . , Pg independent generic points of X over k such that k ( Pi ) = k(q). Thus we can
write q = ϕ ( Pi ) = ϕ(D0 ).
P
P
Now P if P1 , . . . , Pg , Q1 , . . . , Qg are independent generic points of X/k, set D1 = Pi and
D2 = Qi . Then by Lemma 26.1.5, `(D1 + D2 − D) = 1 and D1 + D2 − D ∼ E for some

585
26.1. The Jacobian of a Curve Chapter 26. Constructing Jacobians

P
positive divisor E = Mi which is unique for D1 , D2 . Set q1 = ϕ(D1 ), q2 = ϕ(D2 ) and
q3 = ϕ(E). We define the additive composition law on V by

V × V −→ V
(q1 , . . . , q2 ) 7−→ q1 + q2 := q3 .

Proposition 26.1.6. Let q0 = ϕ(D0 ) for a divisor D0 ∈ Div(X) and let q1 , q2 , q3 as above.
Then

(1) V × V → V is a rational map which is rationally surjective.

(2) q1 + q2 = q2 + q1 .

(3) q0 + (q1 + q2 ) = (q0 + q1 ) + q2 .

Proof. (1) and (2) are straightforward. For (3), consider the divisor class [D0 +D1 +D2 −2D].
Then associativity follows easy.

Theorem 26.1.7 (Weil, FAC). There is an algebraic group G/k and a birational isomor-
phism V → G such that

(1) The group law on G is obtained from V by this map.

(2) G is unique up to birational isomorphism.

Thus for an effective divisor D ∈ Div(X) of degree deg(D) = g, we may choose a model
over k for k(D) which is an algebraic group.

Definition. The algebraic group thus defined is called the Jacobian of X, denoted J =
J(X).

Theorem 26.1.8. For a complete, nonsingular curve X/k of genus g ≥ 1, the Jacobian J =
J(X) is an abelian variety of dimension g equipped with a rational map f : X → J such that
P f is defined, and for independent generic points P1 , . . . , Pg ∈ X
for any field L/k over which
over L, the point Q = f (Pi ) is a generic point of J such that L(Q) = L(P1 , . . . , Pg )s .
Moreover, every generic point of J is of this form and the Pi are uniquely determined up to
permutation.

Remark. The morphism f : X → J factors through ϕ : X (g) → J, (P1 , . . . , Pg ) 7→


P
f (Pi ).
Also, J satisfies a universal property

f
X J(X)

for abelian varieties A. We will show this in Section 26.3 in the case of generalized Jacobians.

586
26.2. Generalized Jacobians Chapter 26. Constructing Jacobians

26.2 Generalized Jacobians


In this section, we generalize the construction of the ordinary Jacobian of a curve X to an
abelian variety Jm , for any modulus m on X, which is universal among algebraic groups G
admitting a morphism f : X → G for which m is a modulus.
Fix a curve X, a modulus m with support S and assume k is algebraically closed and the
points of S are k-rational.

Definition. We say two divisors D, D0 ∈ Div(X) prime to S are m-equivalent, written


D ∼m D0 , if D0 = D + (g) for some function g ≡ 1 mod m. For m = 0, we take ∼0 to be
the ordinary linear equivalence relation on Div(X).

Lemma 26.2.1. For any modulus m, ∼m is an equivalence relation on Div(X) which forms
a finer partition than ∼.

Let Cm be the group of m-equivalence classes of divisors prime to S and let Cm0 be the
subgroup of degree 0 divisor classes in Cm . Following the construction in Section 25.2, for a
divisor D ∈ Div(X) prime to S, let Lm (D) be the k-vector space Lm (D) = H 0 (Xm , Lm (D))
for the sheaf (
0
OQ , Q∈S
Lm (D)Q =
L(D)Q , Q 6∈ S.

Proposition 26.2.2. Fix D ∈ Div(X) prime to S. The set {D0 ∈ Div(X) : D0 ≥


0 and D0 ∼m D} is in one-to-one correspondence with P(Lm (D)) r P(L(D − m)).

Proof. Such a divisor D0 = D + (g) corresponds to a function g ∈ k(X) satisfying g ≡ 1


0
mod m and (g) ≥ −D outside S. This means g ∈ OQ for the unique singular point Q ∈ Xm .
On the other hand, suppose g ∈ Lm (D) and assume g 6= 0 on S. Then D + (g) ∼m D.
Finally, g ∈ Lm (D) is nonzero on S if and only if g ∈ L(D − m) so we have the desired
correspondence.

Lemma 26.2.3. Let D ∈ Div(X) be prime to S and suppose L/k is a field containing k(D).
Then

(1) Lm (D) has a basis of functions which are rational over L.

(2) Im (D)∗ = Ωm (D) has a basis of differential forms which are rational over L.

Proof. (1) Let f ∈ k(X). Then f ∈ Lm (D) is equivalent to f ≡ 1 mod m and (f ) ≥ −D by


definition, but each of these conditions is L-linear since D is rational over L. Hence a basis
of L-rational functions can be chosen.
(2) For m = 0, Serre duality (17.6.2) gives us I(D) ∼ = L(C − D) for a divisor C in the
canonical class. Since C can be chosen to be L-rational, the result follows. The case for
general m is similar, using the results of Chapter 25.

Corollary 26.2.4. Fix D ∈ Div(X) prime to S. Suppose there is a unique positive divisor
D0 ∼m D. Then D0 is rational over k(D).

587
26.2. Generalized Jacobians Chapter 26. Constructing Jacobians

Proof. Since D0 is unique, dim Lm (D) = 1 and dim L(D − m) = 0 so by Lemma 26.2.3,
we can choose g ∈ Lm (D) which is rational over k(D). Write D0 = D + (g) and apply
Lemma 26.1.2.
Fix D ∈ Div(X) prime to S and set

`m (D) = dimk Lm (D) = h0 (Xm , Lm (D))


im (D) = dimk Im (D) = h1 (Xm , Lm (D))
(
g, m=0
π=
g + deg(m) − 1, m 6= 0.

Then by singular Riemann-Roch (Theorem 25.2.3), `m (D) − im (D) = deg(D) + 1 − π.

Lemma 26.2.5. Let K be a field over which D is rational and let P ∈ X be a generic point
over K. Then im (D + P ) = im (D) − 1.

Proof. Basically the same as Lemma 26.1.4, as that proof only needed L(C −D) ∼
= I(D).
Lemma 26.2.6. Let D0 ∈ Div0 (X) be rational over K and suppose P1 , . . . , Pπ are inde-
pendent generic
Pπ points over a field K/k. Then there exists a unique effective divisor of the
form E = i=1 Qi , wherePQ1 , . . . , Qπ are independent generic points over K, such that
E ∼m D0 + D, where D = πi=1 Pi . Moreover, K(D) = K(E).

Proof. This is similar to the proof of Lemma 26.1.5, which is the m = 0 case. For general m,
note that since deg(D) = 0, `m (D) ≤ 1 so by singular Riemann-Roch, im (D) ≤ π. Applying
Lemma 26.2.5 multiple times, we see that im (D0 + D) = 0 implies `m (D0 + D) = 1. Thus
there is a unique (up to scaling) function g ∈ Lm (D0 + D). When m 6= 0, we must also
check L(D0 + D − m) = 0 so that g 6∈ L(D0 + D − m). Suppose `(D0 + D − m) ≥ 1. Since
deg(D0 +D−m) = g−1, ordinary Riemann-Roch implies i(D0 +D−m) = `(D0 +D−m) ≥ 1.
On the other hand, i(D0 +D −m) = i(D0 −m)−π by Lemma 26.2.5, so we get i(D0 −m) > π.
However, deg(D0 − m) < 0 implies `(D0 − m) = 0 so by Riemann-Roch, i(D0 − m) = π, a
contradiction. Thus E can be chosen to be the unique effective divisor in the complete linear
system for D0 + D and the above shows that E ∼m D0 + D.
Finally, Lemma 26.2.3 allows us to choose g rational over K(D0 + D) = K(P1 , . . . , Pπ )s .
By Corollary 26.2.4, E is rational over K(P1 , . . . , Pπ )s . On the other hand, the above para-
graph implies `m (D) = 1, so D is also the unique effective divisor that is m-equivalent to
E − D0 . Thus by the same argument, K(D) ⊆ K(E − D0 ) = K(E). This completes the
proof.
Let X (π) be the symmetric product of π copies of X. We now construct a birational
composition on an open subset of X (π) as we did in the nonsingular case. Fix a field K/k
and a point P0 ∈ X which is rational over K.

Lemma 26.2.7. Let P, Q ∈ X (π) be independent generic points over K. Then there exists
a unique divisor R ∈ Div(X) such that R ∼m P + Q − πP0 as divisors and K(P, Q) =
K(R, P ) = K(R, Q).

588
26.2. Generalized Jacobians Chapter 26. Constructing Jacobians

Proof. Every generic point of X (π) is identified with aP


divisor of X (consisting of independent
generic points) and conversely, so we can write P = πi=1 Pi , where P1 , . . . , Pπ are indepen-
dent generic points of X over K(Q) = K(Q − πP0 ). By Lemma 26.2.6, R = E satisfies the
desired properties.

Proposition 26.2.8. The map X (π) × X (π) → X (π) given by (P, Q) 7→ P + Q := R is a


unique birational composition law making X (π) into a birational group.

Proof. One need only check that π is defined over k – for this, look at the Gal(Ω/k)-action
– and that π is associative – as in Proposition 26.1.6, this can be done by considering the
divisor P + P 0 + P 00 − 2πP0 .
Thus by Weil’s theorem (26.1.7), there exists a unique (up to birational equivalence)
algebraic group G over k birationally equivalent to X (π) with group law induced from the
one constructed above on X (π) .

Definition. The algebraic group Jm = Jm (X) associated to the described birational group
law on X (π) is called the generalized Jacobian of X for the modulus m.

The birational isomorphism ϕm : X (π) → Jm guaranteed by Weil’s theorem satisfies


ϕ(P · Q) = ϕ(P ) + ϕ(Q) where · is the composition law on X (π) .

589
26.3. The Canonical Map Chapter 26. Constructing Jacobians

26.3 The Canonical Map


Let X be an algebraic curve, m a modulus on X with support S and fix a point P0 ∈ X r S.
Let Jm = Jm (X) be the generalized Jacobian of X for m and let ϕ = ϕm : X (π) → Jm
be the canonical map given by Weil’s theorem. We would like to extend this to a map
θ : Div0 (X) → Jm , in analogy with the Abel-Jacobi theorem (see PSection 23.4).
For any divisor D ∈ Div(X) rational over a field K/k, let M = πi=1 Mi be a generic point
of X (π) over K. By Lemma 26.2.6,
P there exists a unique
(π)
P point N ∈ X such that as a divisor
on X, N ∼m D − deg(D)P0 + Mi and K(N ) = K ( Mi ). Define θM (D) = ϕ(N ) − ϕ(M ).

Lemma 26.3.1. Let K/k be a field, suppose D, D0 ∈ Div(X) are rational over K and set
D00 = D + D0 . Then if M, M 0 , M 00 are independent generic points of X (π) defined over K,
then
θM 00 (D00 ) = θM (D) + θM 0 (D0 ).

Proof. Lemma 26.2.6 allows us to write down three unique points N, N 0 , N 00 ∈ X (π) for
D, D0 , D00 , respectively, satisfying

N ∼m D − deg(D)P0 + M, N 0 ∼m D0 − deg(D0 )P0 + M 0 , N 00 ∼m D00 − deg(D00 )P0 + M 00 .

Then ϕ(N 00 ) − ϕ(M 00 ) = ϕ(N ) − ϕ(M ) + ϕ(N 0 ) − ϕ(M 0 ), and so ϕ(N ) + ϕ(N 0 ) + ϕ(M 00 ) =
ϕ(N 00 ) + ϕ(M ) + ϕ(M 0 ). Since k(N ) = k(M ), k(N 0 ) = k(M 0 ) and k(N 00 ) = k(M 00 ), we see
that N, N 0 , N 00 are independent over k. Therefore N + N 0 + M 00 ∼m N 00 + M + M 0 , which
implies θM 00 (D00 ) = θM (D) + θM 0 (D0 ).

Lemma 26.3.2. For D ∈ Div(X), the definition of θM (D) is independent of the choice of
M ∈ X (π) .

Proof. When D0 = 0, N ∼m M implies N = M , so θM (D0 ) = 0. Thus by Lemma 26.3.1,


θM (D) = θM 00 (D) for any independent generic points M, M 00 ∈ X (π) . In general, if M, M 0 ∈
X (π) are generic points over K, we can find a generic point M 00 such that θM 0 (D) = θM (D) =
θM 00 (D).
In light of this, we will write θ(D) = θM (D) = ϕ(N ) − ϕ(M ) for any generic point
M ∈ X (π) .

Lemma 26.3.3. If P = πi=1 Pi is a generic point of X (π) , then θ(P ) = ϕ(P ).


P

Proof. If M is a generic point of X (π) which is independent of P , then by Lemma 26.2.7


there exists a unique point N ∈ X (π) satisfying N ∼m P + M − πP0 . Thus N = P · M using
the birational group law on X (π) , and we have

θ(P ) = ϕ(N ) − ϕ(M ) = ϕ(P )

by definition of ϕ.

Lemma 26.3.4. If D ∈ Div(X) is rational over K, then θ(D) is rational over K.

590
26.3. The Canonical Map Chapter 26. Constructing Jacobians

Proof. By Lemma 26.2.6, θ(D) = ϕ(M ) − ϕ(N ) for N ∈ X (π) satisfying N ∼m D −


deg(D)P0 + M and K(M ) = K(N ). Thus θ(D) is rational over K(N ). By Lemma 26.3.2,
we may vary M , so this holds for any generic point N ∈ X (π) , so θ(D) is rational over
T
N ∈X (π) K(N ) = K.

Proposition 26.3.5. Let DivS (X) be the divisors on X which are prime to S. Then θ :
DivS (X) → Jm is a surjective group homomorphism with kernel

ker θ = {D ∈ DivS (X) | D ∼m nP0 for some n ∈ Z}.

Proof. Lemmas 26.3.1 and 26.3.3 prove that θ is a homomorphism. For surjectivity, note
that Lemma 26.3.3 shows that the image of θ is open and hence generates Jm . Since θ is a
homomorphism, we have im θ = Jm . Lastly, θ(D) = 0 if and only if ϕ(N ) = ϕ(M ) for all
(equivalently, for any) generic points M ∈ X (π) , which is equivalent to N ∼m D−deg(D)P0 +
M . Hence θ(D) = 0 if and only if D ∼m deg(D)P0 .
We next describe the universal property of ϕ and θ.

Lemma 26.3.6. For any separable cover g : X → X 0 of nonsingular, irreducible, projective


curves of degree deg(g) = n + 1 and for any P ∈ X, the divisor corresponding to g −1 (g(P ))
is of the form P + HP for an effective divisor HP ∈ X (n) . Therefore P 7→ HP is a regular
map X → X (n) .
π g
Proof. Let Y be the Galois closure of X → X 0 , which is a Galois covering Y → − X 0.
− X →
0
Set g = Gal(Y → X ) and h = Gal(Y → X) ≤ g. Then the coset space g/h has coset
representatives {σi }n+1
i=1 where [σi+1 ] = h and we can write

n+1
X
g −1 (g(P )) = π ◦ σi (Q)
i=1
Pn
for a fixed Q ∈ π −1 (P ). Set HP = i=1 π ◦ σi (Q), which is an effective divisor on X (n) .
Next, define a map

h : Y −→ X n
Q 7−→ (π ◦ σ1 (Q), . . . , π ◦ σn (Q)).

Composing this with X n → X (n) , we get a map h0 : Y → X (n) which is regular. But
explicitly, X n → X (n) takes (P1 , . . . , Pn ) to the point of X (n) that identifies with the divisor
P n
i=1 Pi , so passing to the quotient Y /h = X, we get precisely the map P 7→ HP .

The map θ : DivS (X) → Jm defines a rational map θ : X r S → Jm . We will prove that
θ is in fact regular.

Lemma 26.3.7. For any generic point M of X (π) over K, there exists a rational map
θ0 : X → Jm defined over K(M ) and regular over X r S which is rational over K, and such
that θ0 = θ on X r S over K.

591
26.3. The Canonical Map Chapter 26. Constructing Jacobians

Proof. For π = 0, this is trivial. Otherwise, choose P1 ∈ X r S such that P1 6= P0 . By


Lemma 26.2.7, there is a unique point N ∈ X (π) satisfying N ∼m −P1 + P0 + M and
K(P1 ) = K(N ). Take g ∈ k(X) such that (g) = N + P1 − P0 − M and g ≡ 1 mod m. By
Lemma 26.2.7, (g) is unique, so `m (−P0 + P1 + M ) = 0 and thus v∞ (g) = P0 + M . This
shows g : X → P1 is a covering of degree π + 1. Moreover, P0 is a simple pole and g is not
a pth power, so g is separable (of degree π + 1). By Lemma 26.3.6, g −1 (g(P )) = P + HP
for an effective divisor HP ∈ X (n) and the map s : X → X (π) , P 7→ HP is regular. Set
θ0 (P ) = ϕ(N ) − ϕ(HP ), which is rational over K(M ) by Lemma 26.2.7. Then we must show
θ0 agrees with θ on X r S.
Suppose P 6∈ S is rational over K; set a = g(P ). When P 6= P0 , a 6= ∞ since v∞ (g) =
P0 + M by the above. Thus (g − a) ≥ P − P0 − M , so by Lemma 26.2.7, g − a is unique.
Also note that g − a 6= 0 on S, so the function g 0 = g−a
1−a
satisfies g 0 ≡ 1 mod m and

(g 0 ) = g −1 (a) − v∞ (g) = P + HP − (P0 + M ).

Thus HP ∼m −P + P0 + M , but by uniqueness, HP is also generic over K. Thus ϕ(HP ) is


defined. Finally, θ(P ) = ϕ(M ) − ϕ(HP ) follows by definition, so θ = θ0 on X r S.
Lemma 26.3.8. The extension of θ to X (π) coincides with the birational map ϕ : X (π) → Jm .
Proof. Define Sθ : (X − S)(π) → Jm by πi=1 Mi 7→ πi=1 θ(Mi ). Then Sθ is regular, so it
P P
extends to a rational map
Sθ : X (π) 99K Jm .
We must show Sθ(M ) = ϕ(M ) for any generic point M = πi=1 Mi of X (π) . By P
P
definition and
Lemma 26.3.3, and viewing M as a divisor on X prime to S, we have Sθ(M ) = πi=1 θ(Mi ) =
ϕ(M ) = θ(M ).
We have proven:
Theorem 26.3.9. Let X be a nonsingular curve over k, m a modulus on X with support S
and ϕm : X → Jm the canonical map. Then
(1) ϕm is a rational map over k which is regular over X r S.

(2) The extension of ϕm to DivS (X) descends to an isomorphism Cm0 −
→ Jm .

(3) The extension of ϕm to X (π) defines a birational map X (π) → Jm .


Theorem 26.3.10. Let X be a nonsingular curve over k, P0 ∈ X and G a commutative
algebraic group. For any map f : X → G with modulus m and such that f (P0 ) = 1G , there
exists a unique morphism of algebraic groups F : Jm → G such that f = F ϕm + 1G :

Jm
ϕm
F

X G
f

592
26.3. The Canonical Map Chapter 26. Constructing Jacobians

Proof. Assume 1G = 0. Then for D ∈ Div0 (X) which is prime to S, Theorem 26.3.10(2)
says Cm0 ∼
= Jm , so the natural map

Cm0 −→ G
D 7−→ f (D)

determines a homomorphism of algebraic groups F : Jm ∼ = Cm0 → G. It is clear that


f = F ◦ ϕm by construction. Consider the following extension of f :
π
X π
X
(π)
Sf : X −→ G Pi 7−→ f (Pi ).
i=1 i=1

Then Sf is rational on X (π) and regular on (X r S)(π) , so Sf = F ◦ ϕ where ϕ : X (π) → Jm


is the extension of ϕm to the symmetric product. By Theorem 26.3.10(3), ϕ is biregular on
a nonempty open set U ⊆ Jm . Thus F coincides with Sf ◦ ϕ−1 on U , and Sf ◦ ϕ−1 is regular
on U , so F is as well. Translating U to cover all of Jm , we get that F is regular on Jm .

Corollary 26.3.11. The construction of (Jm , ϕm ) is independent of the basepoint P0 ∈ X,


up to translating ϕm .

Proof. Suppose P00 ∈ X is another basepoint with associated canonical map ϕ0m : X → Jm .
(By Weil’s theorem, Jm is unique already.) Then by Theorem 26.3.10, there are unique maps
F, F 0 : Jm → Jm making the following diagram commute:

ϕm Jm

X F F0
ϕ0m Jm

Then F ◦ F 0 = idJm = F 0 ◦ F , so ϕm = ϕm + ϕ0m (P0 ).

Corollary 26.3.12. Every rational map f : X → G into a commutative algebraic group G


factors through ϕm : X → Jm for some modulus m on X.

Proof. By Theorem 24.2.10, such an f has a modulus m. Now apply Theorem 26.3.10.

593
26.4. Invariant Differential Forms Chapter 26. Constructing Jacobians

26.4 Invariant Differential Forms


Let G be any algebraic group.
Definition. An invariant differential form on G is a differential form ω ∈ Ωk (G) which
is invariant under translation by elements of G.
Lemma 26.4.1. If dim G = n, then the invariant differential forms on G form a vector
space of dimension n.
In our case, we take G = Jm to be the generalized Jacobian of a curve with modulus
m, which is an algebraic group of dimension π = deg m + g − 1. Let ϕm : X → Jm be the
canonical map. Then for any invariant differential form ω ∈ Ω(Jm ), there is an induced
invariant differential form ϕ∗m (ω) ∈ Ω(X).
Proposition 26.4.2. The map ω 7→ ϕ∗m (ω) is a bijection between the invariant differential
forms on Jm and the differential forms α ∈ Ω(X) such that (α) ≥ −m.
Proof. First suppose ϕ∗m (ω) = 0. Consider the maps
g : X π −→ Jm and hi : X π −→ Jm
π
X
(x1 , . . . , xπ ) 7−→ ϕ(xi ) (x1 , . . . , xπ ) 7−→ ϕ(xi ) for each i.
i=1
Pπ ∗
Pπ ∗
Then g = i=1 hi , so g (ω) = i=1 hi (ω) = 0 by hypothesis. Note that g factors as
X π → X (π) → Jm , where the first map is surjective and separable and the second map is the
birational isomorphism from Weil’s theorem (26.1.7). It follows that g is generically surjective
and separable, so g ∗ is injective and thus ω = 0 to begin with. Therefore ω 7→ ϕ∗m (ω) is
injective.
Next, note that Ω(−m) = {α ∈ Ω(X) | (α) ≥ −m}, using the notation of Section 17.5.
By Riemann-Roch (17.6.4),
`(−m) − dim Ω(−m) = deg(−m) + 1 − g
=⇒ 0 − dim Ω(−m) = − deg(m) + 1 − g
=⇒ dim Ω(−m) = deg(m) + g − 1 = π.
Thus it’s enough to show that for each ω ∈ Ω(Jm ), ϕm (ω) lands in Ω(−m) and count dimen-
sions. Notice
P that any element of Ω(−m) has no poles outside S, so by the residue theorem

(17.5.13), P ∈S resP (α) = 0. Thus to show α = ϕ (ω) ∈ Ω(−m), it suffices by Proposi-
tion 24.2.9 to show that Trg (α) = 0 for all rational maps g : X → k such that g ≡ 0 mod m
(which are not pth powers). Let h = Trg (ϕm ) : P1 → Jm , where ϕm is the map coming from
the diagram
ϕm
X Jm

g Trg (ϕ)

P1

594
26.4. Invariant Differential Forms Chapter 26. Constructing Jacobians

By the proof of Theorem 24.2.10, Trg (α) = h∗ (ω) where α = ϕ∗ (ω). On the other hand,
since m is a modulus for ϕm , Proposition 24.2.9 implies h is constant. Thus h∗ (ω) = 0 as
required.

Corollary 26.4.3. If deg(m) ≥ 2, then the map ϕm is not regular at any point P ∈ S.

Proof. Given any P ∈ S, we have deg(m − P ) > 0, so by singular Riemann-Roch (Theo-


rem 25.2.3), dim Ω(−m+P ) = π −1. On the other hand, from the proof of Proposition P26.4.2
we know that dim Ω(−m) = π, so there is some α ∈ Ω(−m) r Ω(−m + P ). If m = nP P ,

then this α has a pole of order nP at P . On the other hand, α = ϕm (ω) for some ω ∈ Ωk (Jm ),
so ϕm cannot be regular at P .

595
26.5. The Structure of Generalized Jacobians Chapter 26. Constructing Jacobians

26.5 The Structure of Generalized Jacobians


In this section, we explore the structure of generalized Jacobians and in addition give some
important examples.

Proposition 26.5.1. Let m and m0 be moduli on X such that m ≥ m0 and fix a point P0 ∈ X.
Then there exists a unique surjective map F : Jm → Jm0 with connected kernel Hm/m0 such
that ϕm0 = F ◦ ϕm .

Proof. Since m0 is a modulus for ϕm0 , m ≥ m0 implies m is also a modulus for ϕm0 . Thus by
Theorem 26.3.10, there exists a unique map F : Jm → Jm0 such that F ϕm + ϕm0 (P0 ) = ϕm0 .
0
To show F is onto, we construct a birational section of F . Consider the map X π → X π given
0
by (M1 , . . . , Mπ0 ) 7→ (M1 , . . . , Mπ0 , P0 , . . . , P0 ). This descends to a map S : X (π ) → X (π) ,
which in turn determines S 0 : Jm0 → Jm via the birational isomorphisms X (π) → Jm and
0
X (π ) → Jm0 . By construction, F ◦ S 0 = 1, so F is surjective.
As a result, anytime m ≥ m0 , the generalized Jacobian for m0 can be decomposed as
Jm0 ∼
= Jm × Hm/m0 . For the special case m0 = 0, J0 = J is the ordinary Jacobian and we write
Lm = Hm/0 . A point d ∈ Jm corresponds to a linear equivalence class of divisors D prime to
S. When D = (g) for some g ∈ k(X), vP (g) = 0 for all P ∈ S, so the image of d in J via the
map F is 0. On the other hand, the class of D determines the origin 0 ∈ Jm when D = (h)
for some h ≡ 1 mod m, so g = λh for some λ ∈ k × .
For a fixed point P ∈ X, define the groups
(n)
UP = {f ∈ k(X)× | vP (f ) = 0} and UP = {f ∈ UP | vP (1 − f ) ≥ n}.
(n)
For each P ∈ S and g ∈ k(X), we get a function gP ∈ UP /UP for some n ≥ 0, and
Q (n )
conversely, every collection of cosets (gP )P ∈S ∈ P ∈S UP /UP P is the image of such a g ∈
(n )
k(X)× . Set Rm = P ∈S UP /UP P , let ∆ be the image of k × in Rm via λ 7→ (λ, . . . , λ) and
Q
define
Hm = Rm /∆.
We have proven the following.

Proposition 26.5.2. The map θ : Hm → Lm , (g) 7→ (gP ) is an isomorphism of groups.

We next describe an algebraic group structure on Lm . To begin, let k((t)) be the field
of formal (Laurent) power series over k, which is a local field when equipped with the usual
power series valuation v. Let

U = {f ∈ k((t)) | v(f ) = 0} and U (n) = {f ∈ U | v(1 − f ) ≥ n}.

Then any f ∈ U (n) can be written f = 1 + an tn + an+1 tn+1 + . . . so that the elements of the
quotient U/U (n) are truncated power series: f = a0 + a1 t + . . . + an−1 tn−1 , a0 6= 0. Therefore
U/U (n) has the structure of a variety, and pointwise multiplication, inversion and change of
uniformizer t are all regular functions on U/U (n) . Thus:

596
26.5. The Structure of Generalized Jacobians Chapter 26. Constructing Jacobians

Proposition 26.5.3. U/U (n) is an algebraic group under pointwise multiplication and in-
version.
Now let V(n) = U (1) /U (n) ⊆ U/U (n) be the subgroup of truncated power series of the form
1 + a1 t + . . . + an−1 tn−1 . This too has the structure of an algebraic group.
Lemma 26.5.4. There is an isomorphism of algebraic groups U/U (n) ∼
= Gm × V(n) .
Proof. Every function f ∈ U/U (n) is of the form f = a0 f 0 for some f 0 ∈ U (1) /U (n) and
a0 ∈ k × , so we have the abstract isomorphism U/U (n) ∼
= Gm × V(n) . Since multiplication by
−1
a0 and a0 are regular, the isomorphism preserves the algebraic group structure.
Lemma 26.5.5. For each 1 ≤ i ≤ n − 1, let gi ∈ k((t)) be a formal power series with
v(gi ) = i. Then every element g ∈ V(n) can be written uniquely as

g = (1 + a1 g1 )(1 + a2 g2 ) · · · (1 + an−1 gn−1 )

for a1 , . . . , an−1 ∈ k.
b1
Proof. Write g = 1 + b1 t + . . . + bn−1 tn−1 and g1 = c1 t + c2 t2 + . . .. Then setting a1 = c1
and
h1 = 1+ag1 g1 , we see that v(1 − h1 ) ≥ 2. The subsequent ai are defined similarly.

Corollary 26.5.6. For each n ≥ 1, V(n) ∼


= Akn−1 as varieties.
For the rest of the section, we will assume char k = 0. The structure theory in charac-
teristic p > 0 can be found in Serre.
Definition. For a formal power series g ∈ k((t)), define the exponential of g to be the formal
expansion
g2 g3
exp(g) = 1 + g + + + . . . ∈ k((t)).
2 6
Lemma 26.5.7. For all g1 , g2 ∈ k((t)), exp(g1 + g2 ) = exp(g1 ) exp(g2 ).
Lemma 26.5.8. For gi ∈ k((t)) with v(gi ) = i as above, we may write every g ∈ V(n)
uniquely as
g = exp(a1 g1 ) exp(a2 g2 ) · · · exp(an−1 gn−1 )
for ai ∈ k.
Proof. Same proof as for Lemma 26.5.5.
By Lemma 26.5.4 and Corollary 26.5.6, we have U/U (n) ∼ = Gm × Gan−1 as an algebraic
(n )
group. The preceding analysis applies to each of the ‘local groups’ UP /UP P in the decom-
position of Rm , so we have

Rm ∼ UP /UP P ∼
Y (n )
Y
= = G|S|
m × Gna P −1 .
P ∈S P ∈S

Then since ∆ ∼
= Gm , we may write

Hm = Rm /∆ ∼
Y
= G|S|−1
m × Gna P −1 .
P ∈S

597
26.5. The Structure of Generalized Jacobians Chapter 26. Constructing Jacobians

Let g, h ∈ k(X) and for every z ∈ P1k , let Dz be the principal divisor defined by Dz = (g +zh)
for z 6= ∞ and D∞ = (h). Let

T = {z ∈ P1k | Dz ∩ S 6= ∅}.

Lemma 26.5.9. The assignment z 7→ ϕ(Dz ) defines a regular map P1k r T → Lm .

Proof. Notice that ϕ(Dz ) ∈ Lm since Dz is the divisor of a rational function. To prove
regularity, we may assume Dz = (g − z). Let ψ = Trg (ϕ), which is regular on P1k r T by
Proposition 24.2.8. Then ϕ(Dz ) = ψ(z) − ψ(∞), so z 7→ ϕ(Dz ) is regular off of T .

Lemma 26.5.10. θ : Hm → Lm is a regular map.

Proof. Write Hm ∼
|S|−1 Q
= Gm × P ∈S Gna P −1 . Then it is enough to check regularity on each com-
ponent of the product. For each P ∈ S, let Gm,P be the multiplicative group corresponding
to P in the first product. Let v be a function with v ≡ 1 mod m at P and v ≡ 0 mod m
on S r {P }. Then for λ ∈ Gm,P , uλ = (λ − 1)v + 1 satisfies uλ ≡ λ mod m at P and uλ ≡ 1
mod m on S r {P }. Moreover, θ(λ) = ϕ((uλ )) for (uλ ) the principal divisor associated to uλ ,
so by Lemma 26.5.9, θ is regular on Gm,P . A similar proof works for each factor Gna P −1 .

Theorem 26.5.11. The Q map θ : Hm → Lm is an isomorphism of algebraic groups. In


∼ |S|−1
particular, Lm = Gm × P ∈S Gna P −1 .

Proof. By Proposition 26.5.2, θ is a bijection and by Lemma 26.5.10, it’s regular, so we at


least know that Hm and Lm have the same dimension. Thus [k(Hm ) : k(Lm )] < ∞. This
means θ is biregular on an open subset, so it is a bijective, birational isomorphism, hence an
isomorphism of algebraic groups.

Remark. The proof in the case char k = p > 0 requires a bit more care, as a bijective,
birational isomorphism is not necessarily a global isomorphism (it may correspond to an
inseparable extension). However, an argument using the trace map and residue formula
show the result; see Serre for the details.

Corollary 26.5.12. For any curve X with modulus m, the generalized Jacobian Jm is a
group extension of the ordinary Jacobian J of X, i.e. there is a short exact sequence of
groups
1 → Lm → Jm → J → 1.

Example 26.5.13. For X = P1k and the modulus m = 0 + ∞, J(P1k ) = 0 so the generalized
Jacobian is Jm = P1k r {0, ∞} ∼
= Gm . More generally, if m = (P ) + (Q) for distinct points
1 ∼
P, Q ∈ Pk , then Jm = Gm . Let us explicitly describe the multiplication on Jm . First, we
have
Div0m (X) = {D = (g) | g : P1 → k, ordP (g) = ordQ (g) = 0}.

598
26.5. The Structure of Generalized Jacobians Chapter 26. Constructing Jacobians

Suppose D1 = (f1 ) and D2 = (f2 ). Then

D1 ∼m D2 ⇐⇒ there is some f ∈ k(P1 )× with (f ) = D1 − D2 and f ≡ 1 mod m


 
f1
⇐⇒ there is some f with (f ) = and f (P ) = f (Q) = 1
f2
f1 (P ) f1 (Q)
⇐⇒ =
f2 (P ) f2 (Q)
f1 (P ) f2 (P )
⇐⇒ = .
f1 (Q) f2 (Q)

Thus the map ϕm : Cm0 → Gm = P1 r {P, Q} is given by [(g)] 7→ [g(P ), g(Q)]. It is clear that
this ϕm is both well-defined and injective. To see that it is surjective, take [α, β] ∈ P1 r{0, ∞}.
Then the function α
β
(t − Q) t−P
g(t) = −
P −Q Q−P
defines an m-equivalence class [(g)] ∈ Cm0 and ϕ([(g)]) = [α, β] by construction.

Example 26.5.14. Let X = E be an elliptic curve with distinguished point O ∈ E. Then


Pic0 (E) = E by the Abel-Jacobi map (as in Chapter 23):

E −→ Pic0 (E)
P 7−→ [(P ) − (O)].

For D = nP P ∈ Div0 (E), there is some function f ∈ k(E)× such that D = D − (O) + (f ).
P
Consider the modulus m = (M ) + (N ) for distinct points M, N ∈ E r {O}. If D ∈ Div0m (E),
then either

ˆ D = (M ) or D = (N ); or

ˆ D 6= (M ), (N ).

In the first case, D = (M ), so that ordM (f ) = −1.P


In the second case, ordM (f ) = ordN (f ) =
0, so D ∼ S − (O) ∼ (S + T ) − (T ) for S = nP P and any T ∈ E. Choosing T 6∈
{O, M, N, M − N, N − M }, we may assume (M + T ) − (T ) and (N + T ) − (T ) have support
which is disjoint from the support of m. Set
(
O, if S 6= M, N
R=
T, if S = M or N.

Then there is a function f ∈ k(E)× such that D = (S + R) − (R) + (f ) and ordM (f ) =


ordN (f ) = 0.

599
26.5. The Structure of Generalized Jacobians Chapter 26. Constructing Jacobians

Repeating this for two divisors D1 , D2 ∈ Div0m (E), we can write Di = (Si +Ri )−(Ri )+(fi )
for i = 1, 2. Then

D1 ∼m D2 ⇐⇒ there is an f ∈ k(E)× with (f ) = D1 − D2 , f ≡ 1 mod m


 
f1
⇐⇒ (f ) = (S1 + R1 ) − (S2 + R2 ) + (R2 ) − (R1 ) +
f2
f1
for some f ∈ k(E)× with f = and f (M ) = f (N ) = 1
f2
⇐⇒ S1 + R1 − (S2 + R2 ) + R2 − R1 = O in E
f1 (M ) f2 (M )
⇐⇒ S1 = S2 in E and = .
f1 (N ) f2 (N )

Thus the canonical map is given by

ϕm : Cm0 (E) −→ Jm = Gm × E
 
f (M )
[D] 7−→ ,S if D = (S + R) − (R) + (f ).
f (N )

We need the following result to describe the group structure of Jm in this case.

Theorem 26.5.15. Let (G, +) and (G, ⊕) be two groups and


i p
− G→
1→A→ − G→0

a group extension by an abelian group A. Then

(1) There is a section s : G → G such that G = A × G as sets.

(2) G acts on A via the map

A × G −→ A
(a, σ) 7−→ aσ := x ⊕ a x

where x ∈ G is any element such that p(x) = σ.

(3) There is a 2-cocycle c : G × G → A such that

c(σ, τ ) · c(σ + τ, ρ) = c(τ, ρ)σ · c(σ, τ + ρ)


and (a, σ) ⊕ (b, τ ) = (a · b · c(σ, τ ), σ + τ )

for all σ, τ, ρ ∈ G and a, b ∈ A. Explicitly, c(σ, τ ) = s(σ) ⊕ s(τ ) s(σ + τ ), where s


is the section from (1).

In our situation, we have a group extension

1 → Gm → Jm → E → 0

600
26.5. The Structure of Generalized Jacobians Chapter 26. Constructing Jacobians

and Jm = Gm × E as sets, with (b1 , P ) ⊕ (b2 , Q) = (b1 b2 · cm (P, Q), P + Q) for all b1 , b2 ∈
Gm , P, Q ∈ E and cm : E × E → Gm a 2-cocycle. To determine this cocycle, let D1 =
(P1 ) − (O) + (f1 ) and D2 = (P2 ) − (O) + (f2 ) be two divisors on E prime to the support
of m. Then under the canonical map ϕm : Cm0 (E) → Jm , we have ϕm ([Di ]) = (bi , Pi )
for i = 1, 2 and (b1 , P1 ) ⊕ (b2 , P2 ) = (b3 , P3 ) for some b3 ∈ Gm , P3 ∈ E. Notice that
if D3 = (P1 ) + (P2 ) − 2(O) + (f1 f2 ), then it will be enough to determine f3 such that
D3 = (P3 ) − (O) + (f3 ). By the Abel-Jacobi isomorphism, we have

(P1 ) + (P2 ) − 2(O) = (P1 + P2 ) − (O) + (`P1 ,P2 )

for some `P1 ,P2 ∈ k(E)× . Set f3 = f1 f2 `P1 ,P2 ,

f3 (M ) `P ,P (M ) `P1 ,P2 (M )
b3 = = b 1 b2 1 2 and cm (P1 , P2 ) = .
f3 (N ) `P1 ,P2 (N ) `P1 ,P2 (N )

We claim (`P1 ,P2 ) = (P1 ) + (P2 ) − (P1 + P2 ) − (O) will work. Let LP1 ,P2 be the linear form
defining the line in E ⊆ P2 containing P1 , P2 and −(P1 + P2 ).

−(P1 + P2 )

P2
P1

LP1 ,P2
P1 + P2

LP1 +P2 ,O

Then dividing out by z gives a function with the following zeroes and poles:
 
LP1 ,P2 (z)
= (P1 ) + (P2 ) + (−(P1 + P2 )) − 3(O).
z

On the other hand, the linear form LP1 +P2 ,O defining the line containing P1 + P2 , −(P1 + P2 )
and O yields  
LP1 +P2 ,O (z)
= (P1 + P2 ) + (−(P1 + P2 )) − 2(O).
z
LP1 ,P2 `P1 ,P2 (M )
So taking `P1 ,P2 = LP1 +P2 ,O
, we get cm (P1 , P2 ) = `P1 ,P2 (N )
as the desired 2-cocycle.

601
Chapter 27

Class Field Theory for Curves

The goal of class field theory is to classify all abelian extensions L/K in terms of the
arithmetic of K itself. To this end, let I(K) be the set of fractional ideals of K and set
GL/K = Gal(L/K) for any abelian Galois extension.

Definition. The Artin map of an unramified abelian extension L/K is the map on frac-
tional ideals given by

θL/K : I(K) −→ GL/K


Y
a 7−→ Frobvordv (a)
v

where the product runs over all places v of K, Frobv is the Frobenius element at v and ordv (a)
is the valuation of a at v.

Theorem 27.0.1. For any principal ideal (α) ∈ I(K) and any unramified abelian ex-
tension L/K, θL/K ((α)) = 1. Therefore θL/K descends to a map on the class group:
CK = I(K)/P (K) → GL/K .

Class field theory for number fields is summarized in the following theorem.

Theorem 27.0.2. Let L/K be a finite abelian extension with Galois group GL/K . Then

(1) (Artin Reciprocity) The Artin map θL/K : CK → GL/K is surjective with kernel
NL/K (CL ), where NL/K is the norm map.

(2) For any normal subgroup N ≤ CK , there exists a finite abelian extension L/K for
which N = ker θL/K . In this case, CK /N ∼
= GL/K .

(3) The Artin map is natural, i.e. for all finite unramified abelian extensions M ⊇ L ⊇ K,
there is a commutative diagram

602
Chapter 27. Class Field Theory for Curves

θM/L
CL GM/L

NL/K

CK GM/K
θM/K

There is an analogous theory for covers of curves, constructed as follows. Let X be a


smooth, geometrically connected curve over a perfect field k and specify a set of closed points
S ⊂ X.

Theorem 27.0.3. Let H any connected commutative algebraic group with principal homo-
geneous space H 0 admitting a regular map f : X r S → H 0 . Then

(1) There exists a modulus m on X for f supported on S and a commutative diagram


defined over k:

ϕm
X rS Jm0

fm0
f
H0

where Jm0 is a principal homogeneous space for the generalized Jacobian Jm and ϕm :
X r S → Jm0 is the restriction of the canonical map.

(2) There exists a map fm : Jm → H with respect to which fm0 : Jm0 → H 0 is an equivariant
map of principal homogeneous spaces.

(3) For every modulus m of f , the map fm0 is unique and thus there is a unique modulus
f dividing all such m.

(4) Any geometrically connected abelian cover of H 0 pulls back along f to give a geomet-
rically connected cover of X which is unramified outside S.

Definition. The modulus f is called the conductor of the cover f : X r S → H 0 .

603
27.1. Coverings and Isogenies Chapter 27. Class Field Theory for Curves

27.1 Coverings and Isogenies


Let k = Fq be a finite field and π : X 0 → X a finite, unramified cover of curves defined
over k. This corresponds to a field extension K 0 /k(X) of transcendence degree 1 fields. If
Gal(K 0 /k(X)) is abelian, we say the cover π is abelian.

Theorem 27.1.1. Let π : X 0 → X be an abelian cover of curves corresponding to a field


extension K 0 /k(X), with ramification locus S. Then there exists a modulus m supported on
S and a finite connected abelian cover G1 → Jm1 such that the diagram

X 0 r S0 G1

X Jm1
ϕm

commutes, where S 0 = π −1 (S) and Jm1 is the principal homogeneous space for Jm given by
Theorem 27.0.3. Moreover, there exists a unique modulus f dividing all moduli m for π, and
for this f, the cover G1 → Jm1 is geometrically connected if and only if X 0 is geometrically
connected.

Suppose π : X 0 → X is an abelian cover such that X 0 (k) 6= ∅. Then for any y0 ∈ X 0 (k),
with image x0 = π(y0 ) ∈ X(k), all elements of π −1 (x0 ) are k-points of X 0 since π is abelian.
Let

ϕ : X −→ J = J(X)
x 7−→ x − x0

be the ordinary Abel map. Then by Theorem 27.1.1, there is a finite, connected abelian
cover G → J such that the diagram

X0 G

X ϕ J

commutes. Then J = G/N for some normal subgroup N ⊆ G. In other words, G → J is an


isogeny.
In the case when G = J, the cover J → J is an Artin-Schreier cover, that is, it is
of the form x 7→ xq − x. The map Fq : x 7→ xq is called the qth Frobenius on J (or on
any algebraic group G over k). Then the map G = J → J can be written as Fq − 1.
Take a point ξ ∈ X with field of definition Fqr , r ≥ 1. Then ξ determines a Galois orbit
(ξ = ξ1 , . . . , ξr ) with ξi ∈ X(k̄), Fq (ξi ) = ξi+1 for 1 ≤ i ≤ r − 1 and Fq (ξr ) = ξ1 . Take
y ∈ X 0 = {y ∈ J | Fq (y) − y ∈ X} lying above ξ. Note that J acts on X 0 by z · y = y + z,

604
27.1. Coverings and Isogenies Chapter 27. Class Field Theory for Curves

and through this action, we may lift Fq to a Frobenius Frobξ : X 0 → X 0 . Explicitly,


Frobξ (y) = Fqr (y), but notice that z = Fqr (y) − y ∈ J:
r
X n
X n
X r
X
Fqr (y) − y = (Fqi (y) − Fqi−1 (y)) = Fqi−1 (Fq (y) − y) = Fqi−1 (ξ) = ξi .
i=1 i=1 i=1 i=1
Pn
Thus Frobξ may be identified with the element z = i=1 ξi .

Definition. The Artin map of X is defined by

θ : Div0 (X) −→ J
X X
D= nξ ξ 7−→ nξ ξ.

Note that if D is a principal divisor, then θ(D) = 0 in J. This is a completely trivial


consequence of the geometry of J, whereas the proof of the analogous Theorem 27.0.1 is one
of the most difficult parts of global class field theory. As a consequence, θ descends to a map
on the Picard group: Pic0 (X) = Div0 (X)/ PDiv0 (X) → J.

605
27.2. Class Field Theory Chapter 27. Class Field Theory for Curves

27.2 Class Field Theory


We next give a little more detail on covers of curves before stating the analogue of class
field theory for curves over a finite field. Let k be any field and V a normal irreducible
variety over k with function field K = k(V ). Given a finite separable extension L/K, let
W be the normalization of V in L. This defines a covering π : W → V corresponding to
L/K. If L/K is Galois with G = Gal(L/K), then we say pi : W → V is Galois, write
Gal(W/V ) = Gal(L/K), and we have V ∼ = W/G. For a point P ∈ V , we say P is unramified
in π if it has exactly n = [L : K] preimages.
Now suppose f : X → V is a rational map. Then there is an induced map of varieties

f W → W given by pullback:

f ∗W X

f
π
W V

In fact, we have the following amazing result, which can be found in Serre.
Theorem 27.2.1. Every abelian cover W → V is the pullback of an isogeny f : G0 → G of
smooth, abelian varieties with kernel ker f ∼
= Gal(W/V ).
Given a finite group N , let A(N ) be the group algebra of N (over some universal domain
Ω, e.g.) and let G(N ) be the subgroup of invertible elements:

A(N ) = {(as )s∈N | as ∈ Ω}


G(N ) = {(as )s∈N ∈ A(N ) | det(ast ) 6= 0}.

Then N embeds as a subgroup of G(N ) via s 7→ (at ) where at = δst for all t ∈ N .
Proposition 27.2.2. If π : W → V is a Galois cover with Galois group N , then there exist
a rational map f : V → G(N )/N such that W ∼
= f ∗ G(N ). Moreover, if W and π are defined
over k, then both f and the isomorphism W ∼
= f ∗ G(N ) are also defined over k.
Proof. Define g : W → A(N ) by x 7→ (ϕs (x)), where ϕ ∈ k̄(W ) is any fixed rational function.
Then since det(ϕst ) 6= 0, g has image in G(N ). Consider the diagram
g
W G(N )

π
f
V G(N )/N

Then by the normal basis theorem from Galois theory, there exists an element ψ ∈ k(W )
such that {ϕs }s∈N is a basis of the field extension k(W )/k(V ). It follows that f is equal to
g after taking quotients, so the diagram commutes.

606
27.2. Class Field Theory Chapter 27. Class Field Theory for Curves

Proposition 27.2.3. If P ∈ V is unramified in π : W → V , then f : V → G(N )/N from


above may be chosen to be regular on π −1 (P ).

Proof. To prove this, we must find a normal basis {ψ s }s∈N where ψ is regular on the set
π −1 (P ). Let OP be the local ring at P in K = k(V ) with maximal ideal mP and as in
Chapter 25, let OP0 be the integral closure of OP in L = k(W ), so that
\
OP0 = OQ .
Q∈π −1 (P )

Set k 0 (P ) = OP0 /mP OP0 and k(P ) = OP /mP . Since P is unramified, k 0 (P ) is an étale algebra
of degree n = [L : K] over k(P ). Thus by the normal basis theorem, we can choose λ ∈ k 0 (P )
such that {λs }s∈N is a normal basis of k 0 (P )/k(P ). Now λ may be lifted to OP0 for all P ,
giving a rational function ψ on W which is regular on π −1 (P ).
This proves Theorem 27.2.1.

Example 27.2.4. Let N = Z/nZ where (n, p) = 1. Then A(N ) ∼ = k[t]/(tn − 1) and if k
contains a primitive nth root of unity, then G(N ) ∼
= G×n
m . Let θn : Gm → Gm be the isogeny
n
x 7→ x (called the nth Kummer cover). Then G(N ) → G(N )/N is the pullback of this
isogeny:

G(N ) Gm

θn

G(N )/N Gm

Example 27.2.5. Let N = Z/pZ. Then A(N ) ∼ = k[t]/(tp − 1), G(N ) ∼= Ga and G(N ) →
G(N )/N is the pullback of the Artin-Schreier cover ℘ : Ga → Ga , x 7→ xp − x:

G(N ) Ga

G(N )/N Ga

The general case is given by Artin-Schreier-Witt theory.

Now let X be a complete, irreducible, nonsingular curve over k with field of rational
functions K = k(X).

Corollary 27.2.6. If π : Y → X is an abelian cover of curves with Galois group N =


Gal(Y /X), then there exists a separable isogeny G → H of abelian varieties with kernel N
such that Y ∼
= f ∗ G for some f : X → H:

607
27.2. Class Field Theory Chapter 27. Class Field Theory for Curves

Y G

π
f
X H

Proof. Follows from Theorem 27.2.1.


Let S be the ramification locus of π : Y → X. By Theorems 24.2.1 and 26.3.10, f : X →
H has a modulus m supported on S such that f factors through the generalized Jacobian
Jm = Jm (X):
ϕm f¯
f : X −→ Jm →
− H.
Thus we can pull back the isogeny G → H to an isogeny J 0 = f¯∗ G → Jm such that Y ∼
= ϕ∗m J 0 .
This proves:

Corollary 27.2.7. Every abelian cover Y → X is the pullback of a separable isogeny J 0 → Jm


for some modulus m on X. Moreover, there exists a modulus f such that f ≤ m for all such
moduli m.

Definition. The modulus f is called the conductor of the cover Y → X.

Corollary 27.2.8. There is a one-to-one correspondence

{unramified abelian covers Y → X} ←→ {isogenies J 0 → J(X)}.

608
Part VI

Algebraic Fundamental Groups

609
Chapter 28

Introduction

The following notes are taken from a reading course on étale fundamental groups led by Dr.
Lloyd West at the University of Virginia in Spring 2017. The contents were presented by stu-
dents throughout the course and mostly follow Szamuely’s Galois Groups and Fundamental
Groups. Main topics include:

ˆ A review of covering space theory in topology (universal covers, monodromy)

ˆ Covers and ramified covers of normal curves

ˆ The algebraic fundamental group for curves

ˆ Finite étale covers of schemes and the étale fundamental group

ˆ Grothendieck’s main theorems for the fundamental group of a scheme

ˆ Applications.

610
28.1. Topology Review Chapter 28. Introduction

28.1 Topology Review


There are two key concepts in algebraic topology that, for various reasons, one might want
to consider in an algebraic setting. These are covering spaces and fundamental groups, and
they are intimately connected. The more familiar concept might be that of the fundamental
group, which at the beginning is usually defined in terms of homotopy classes of based loops
in a given topological space. To define an algebraic analogue, we will need an alternative
perspective on the fundamental group.
Let X be a connected, locally simply connected topological space.
Definition. A cover of X is a space Y and a map p : Y → X that is a local homeomorphism.
That is, for every x ∈ X there is a neighborhood U ⊆ X of x such that p−1 (U ) is a disjoint
union of open sets in Y and p restricts to a homeomorphism on each open set.
One consequence of this definition is that for all x, y ∈ X, p−1 (x) is a discrete space and
p
p−1 (x) ∼
= p−1 (y). A primary goal in topology is to study and classify all such covers Y → − X.
p p0
− X and Y 0 −
Definition. A morphism of covers between Y → → X is a map f : Y → Y 0
making the following diagram commute:

f
Y1 Y2

p1 p2

This defines a category CovX of covers over X. In this category, we will abbreviate
HomCovX (Y, Z) by HomX (Y, Z).
Example 28.1.1. The unit interval [0, 1] ⊆ R has no nontrivial covers. However, S 1 ⊆ C
does: for each n ∈ Z, the map

pn : S 1 −→ S 1 , z 7−→ z n

is a cover. Further, there is a special cover

π : R −→ S 1 , t 7−→ e2πit

such that for every n ∈ Z, the following diagram commutes:

R S1

π pn

S1

All covers of the circle arise in this way.

611
28.1. Topology Review Chapter 28. Introduction

The special property of R → S 1 leads to the notion of a universal cover.


Definition. A covering space π : X e → X is a universal cover for X if for every other
cover p : Y → X, there is a unique map f : X
e → Y making the diagram commute:

f
X
e Y

π p

It is equivalent to say that a universal cover is any simply connected cover of X, and
one shows easily that universal covers are unique up to equivalence of covers. An important
result is that a universal cover exists, under certain mild conditions on X.
In topology, the topological fundamental group is defined using homotopy:
 
top homotopy classes of loops
π1 (X, x) := .
in X based at x

The universal cover has important connections to this fundamental group. In particular,
consider an automorphism α ∈ AutX (X)e = HomX (X, e Fix x ∈ X and a lift x̃ ∈ X
e X). e of x.
e is simply connected, any path x̃ → αx̃ is unique up to
Then πα(x̃) = x. Moreover, since X
homotopy. This determines a map AutX (X) e → π1top (X, x).

e → π1top (X, x) is an isomorphism.


Theorem 28.1.2. For any x ∈ X, AutX (X)

Unfortunately, such a space X e does not exist in the algebraic world. Thus we describe a
slightly different interpretation of the fundamental group.
Definition. The fibre functor over x ∈ X is the assignment

Fibx : CovX −→ Sets


p
− X) 7−→ p−1 (x).
(Y →

By the universal property of a universal cover X e ∈ CovX , to give a morphism of covers


f :Xe → Y is the same as to choose a point y = f (x̃) ∈ p−1 (x). In other words, Fibx is a
representable functor, i.e. for x ∈ X, there is a natural isomorphism

Fibx (−) ∼ ex̃ , −),


= HomX (X

where the Hom set consists of morphisms based at x̃. This fibre functor is constructible in
algebraic categories, though it fails to be representable.
Going further, there is a natural left action of AutX (X)
e on X;
e however, it will be more
convenient to view this as a right action of AutX (X)e op on X.
e This induces a left action of
top
AutX (X)
e = π1 (X, x) on HomX (X, e Y ):

α · f = f ◦ α for any α ∈ AutX (X), e → Y.


e f :X

612
28.1. Topology Review Chapter 28. Introduction

This action is called the monodromy action. Often, one views this as an action of π1top (X, x)
on the fibre Fibx (Y ) given by lifting paths. In any case, we get a map

π1top (X, x) −→ Aut(Fibx ),

where Aut(Fibx ) is the automorphism group of the fibre functor in the following sense. For
any functor F : C → D, an automorphism of F is a natural transformation of F that
has a two-sided inverse. The set Aut(F ) of all automorphisms of F is then a group under
composition. Moreover, Aut(F ) has a natural action on F (C) for any object C ∈ C.

Theorem 28.1.3. For all x ∈ X, π1top (X, x) → Aut(Fibx ) is an isomorphism.

Theorem 28.1.4. Let X be a connected, locally simply connected space and fix x ∈ X. Then
the fibre functor Fibx defines an equivalence of categories

CovX −
→ {left π1 (X, x)-sets}

with connected covers corresponding to transitive π1 (X, x)-sets and Galois covers to coset
spaces of X
ex by normal subgroups.

Proof. For a transitive π1 (X, x)-set S, define YS = X/H


e where H = Stabπ1 (X,x) (s) for any
point s ∈ S. This defines a Galois cover YS → X and one can extend this to arbitrary
π1 (X, x)-sets orbitwise for the full correspondence.
The picture gets more interesting if we restrict ourselves to finite covers. Given such a
cover p : Y → X, there is an exact sequence of groups

1 → N → π1top (X, x) → AutX (p−1 (x)) → 1

where N is some finite index kernel. This shows that the monodromy action factors through
a finite quotient. As a result, this action can be defined on the level of a profinite group,
namely the profinite completion of π1top (X, x):

π1top
\ (X, x) := lim π1top (X, x)/N,
←−

where the inverse limit is over all finite index subgroups N ≤ π1top (X, x).

Corollary 28.1.5. The fibre functor Fibx defines an equivalence of categories



→ {finite, continuous π1top
{finite covers of X} − \ (X, x)-sets}.

Moreover, the correspondence restricts to



→ {finite π1top
{connected covers} − \ (X, x)-sets with transitive action}

and → {π1top
{Galois covers} − \ (X, x)/N | N an open normal subgroup}.

613
28.2. Finite Étale Algebras Chapter 28. Introduction

28.2 Finite Étale Algebras


Fix a field k and an algebraic closure k̄, which comes equipped with a separable closure
ks ⊆ k̄. Set Gk = Gal(ks /s) and let L/k be any finite, separable extension.

Lemma 28.2.1. Homk (L, ks ) is a finite, continuous, transitive Gk -set.

Proof. By Galois theory, # Homk (L, ks ) = [L : k], so this is finite when the extension is
assumed to be finite. The Gk -action is defined by σ·f = σ◦f for σ ∈ Gk and f ∈ Homk (L, ks );
it is routine to verify that this is indeed a group action. Now to show the action is continuous,
since Homk (L, ks ) is discrete, this is equivalent to showing the stabilizer StabGk (f ) is open
for each f ∈ Homk (L, ks ). Notice that

StabGk (f ) = {σ ∈ Gk | σ ◦ f = f } = {σ ∈ Gk | σ fixes f (L)}

and this is open by Galois theory / the topology of the profinite group Gk . Finally, since
L/k is separable, we may pick a minimal polynomial h(t) for a primitive element of L/k.
Then Gk permutes the roots of h(t) transitively, so it follows that Gk acts transitively on
Homk (L, ks ).

Corollary 28.2.2. There exists an open subgroup H ≤ Gk such that Homk (L, ks ) ∼ = Gk /H
as Gk -sets. Further, if L/k is Galois, H may be chosen to be an open normal subgroup.

Proof. We may pick H = StabGk (h), the stabilizer of the minimal polynomial of a primitive
element of L/k. The Galois case follows from the fundamental theorem of Galois theory.

Theorem 28.2.3. The assignment L/k 7→ Homk (L, ks ) is a contravariant functor

{finite separable extensions of k} −→ {finite, continuous, transitive Gk -sets}

which is an anti-equivalence of categories. Moreover, Galois extensions L/k correspond to


finite quotients of Gk .

Proof. For any finite, continuous, transitive Gk -set S, define a finite separable extension
LS /k by taking the subfield of ks /k fixed by the stabilizer StabGk (s) of any point s ∈ S.
One now checks that this is an inverse functor to the Homk (−, ks ) functor.
To study all finite continuous Gk -sets, we replace separable field extensions with a slightly
more general object. Recall (Section 15.5) that a k-algebra A is a finite étale algebra if
A∼
= L1 × · · · × Lr for finite separable extensions Li /k.
Corollary 28.2.4 (Grothendieck). The assignment A/k 7→ Homk (A, ks ) is an anti-equivalence
of categories

{finite étale k-algebras} −
→ {finite continuous Gk -sets}
which reduces to the above case when A = L is a finite separable extension of k.

614
28.2. Finite Étale Algebras Chapter 28. Introduction

Topologically, we may view k as a point space covering the ‘smaller’ point space ks , and
any intermediate extension L/k as an intermediate cover. In this setting, Gk plays the role
of the deck transformations of the ‘universal cover’ k → ks and Homk (L, ks ) plays the role
of the fibre of a cover. Also, under this analogy, a finite separable extension L represents
a connected cover while a finite étale algebra A may be viewed as a disconnected cover.
Finally, the choice of a algebraic (and separable) closure of k is analagous to the choice of a
basepoint of a topological space, which also determines a universal cover. We will see that
this subtlety conceals a wealth of information about the algebraic fundamental group.

615
28.3. Étale Morphisms Chapter 28. Introduction

28.3 Étale Morphisms


Let us now shift focus to the algebraic setting, meaning a category of algebro-geometric
objects such as nonsingular varieties over a field or schemes. The Zariski topology makes it
tricky to define a cover in terms of a local homeomorphism – in fact, by Zariski’s theorem,
such a map is only ever a trivial inclusion of an open set, so nothing terribly interesting.
However, in the geometric setting, we have the notion of the differential on tangent spaces.

Definition. For a morphism p : Y → X of nonsingular varieties X and Y over an alge-


braically closed field k, we say p is étale at y ∈ Y if the differential dpy : Ty Y → Tp(y) X is
an isomorphism. The map is said to be étale if it is étale at every y ∈ Y .
If k is not algebraically closed, consider the base change pk̄ : Y (k̄) → X(k̄). Then we say
p is étale at y ∈ Y if for all geometric points ȳ → y, pk̄ is étale at ȳ. As above, p is étale if
it is étale at every y ∈ Y .

A consequence of this definition is that the fibres of an étale morphism p : Y → X are


finite and their cardinality is constant (at least, on connected components). In Section 15.6,
we defined a morphism of schemes p : X → Y to be étale at y ∈ Y if the induced morphism
on local rings p# : OX,p(y) → OY,y was an étale morphism of local rings, that is, if p# was
flat and unramified. One can show that if p : Y → X is a morphism of nonsingular varieties
whose differential dpy is an isomorphism for all y ∈ Y , then p is flat and unramified, and
conversely. Thus our definition for schemes properly encapsulates the differential definition.
Let FétX be the category consisting of finite étale covers of X, together with morphisms
of covers defined as in the topological case (i.e. commuting with covering maps over X). For
a geometric point x̄ : Spec Ω → X in a scheme X, define the fibre functor

Fibx̄ : FétX −→ Sets


p
(Y →
− X) 7−→ Spec Ω ×X Y.

Definition. The étale fundamental group of a scheme X at a geometric point x̄ :


Spec Ω → X is the automorphism group of the fibre functor over x̄,

π1ét (X, x̄) = Aut(Fibx̄ ).

As in the topological case, we will prove:

Theorem 28.3.1 (Grothendieck). For X a connected scheme and x̄ : Spec Ω → X a geo-


metric point,

(1) π1ét (X, x̄) is a profinite group which acts continuously on each fibre Fibx̄ (Y ) for Y → X
a cover.

(2) Fibx̄ : FétX → {continuous, finite π1ét (X, x̄)-sets} is an equivalence of categories.

Further, we will see that π1ét (X, x̄) = lim AutX (Fibx̄ (Y )), where the inverse limit may be
←−
taken over all Galois covers p : Y → X.

616
28.3. Étale Morphisms Chapter 28. Introduction

Example 28.3.2. Let k be a field and consider the scheme X = Spec k which is a point. A
geometric point x̄ : Spec Ω → X is a choice of algebraic closure Ω ⊇ k; such a choice also
defines a separable closure ksep ⊇ k. In this case, the correspondence of Theorem 28.3.1 is:
FétX ←→ {Spec A | A is a finite étale k-algebra}.
Here, the functor Fibx̄ is representable by Homk (A, Ω) ∼ = Homk (A, ksep ). Moreover, we
have a Galois action of Gk := Gal(ksep /k) on Homk (A, ksep ) given by σ · f = σ ◦ f for any
f : A → ksep and σ ∈ Gk . In particular, π1 (X, x̄) = Gk and this action is the monodromy
action as described in Section 28.1.
Theorem 28.3.3. For a field k, Homk (−, ksep ) induces an equivalence of categories

{finite étale k-algebras} −
→ {finite continuous Gk -sets}.
Example 28.3.4. Let (A, m, k) be a complete DVR and set X = Spec A. There is an
equivalence between étale covers
FétA ←→ Fétk .
(One direction is obvious; the other is an application of Hensel’s Lemma.) Therefore
π1ét (Spec A, x̄) ∼
= π1ét (Spec k, x̄) for any geometric point x̄ of A.
Theorem 28.3.5. Suppose X is connected and normal, K = k(X) is the function field of X
and L/K is a finite, separable field extension. Let Y be the normalization of X in L. Then
p : Y → X is a ramified cover and for some set S ⊂ Y , p : Y r S → X is étale. Further,
every étale cover of X arises in this way.
Corollary 28.3.6. If X is a connected, normal scheme with function field K and K ur is
the compositum of all extensions of K in which normalization of X is unramified, then
π1ét (X) ∼
= Gal(K ur /K).
Example 28.3.7. A famous result in number theory says that Q has no nontrivial unramified
extensions. Therefore π1ét (Spec Z) = 1. However, ‘removing some points’
 1  from Spec Z, e.g.
ét
localizing, yields nontrivial fundamental groups, such as π1 Spec Z n .
Example 28.3.8. We will show that π1ét (AC ) = 1 (compare to the topological case!), but
for char k > 0, we may have π1ét (Ak ) 6= 1.
There is a classic and important connection between curves and Riemann surfaces which
we will explored further in Chapter 30. The arithmetic version of this correspondence is
borne out by the following theorem.
Theorem 28.3.9. Let X be a normal curve of genus g over k. If char k = p > 0, then
g
* +
π ét (X)(p ) ∼
0
Y
1 = x1 , y1 , . . . , xg , yg : [xi , yi ] = 1 [p]
i=1
0
where (·)(p ) denotes the prime-to-p part of a profinite group. Further, if k = C, then
π ét (X) ∼
1 1
top
= π\ (X).

617
Chapter 29

Fundamental Groups of Algebraic


Curves

Let A be a finitely generated k-algebra which is an integral domain with field of fractions K,
and suppose tr degk K = 1. Consider the space X = Spec A = {p ⊂ A | p is a prime ideal}.
Taking closed sets to be of the form V (I) = {p ∈ Spec A | I ⊇ p} for ideals I ⊂ A defines
a Zariski topology on X. The zero ideal 0 defines a generic point of X: for any nonempty
open subset U ⊆ X, 0 ∈ U . It follows that A0 = K. On the other hand, any nonzero prime
p is maximal (since dim A = tr degk K = 1), so V (p) = {p}. Thus prime ideals correspond
to closed points in X. In particular, all proper closed subsets of X are finite.

Definition. For P ∈ X, let OX,P := AP be the local ring at P . For any open U ⊆ X, set
\
OX (U ) := OX,P .
P ∈U

Proposition 29.0.1. U 7→ OX (U ) defines a sheaf of functions on X.

Definition. We call the ringed space (X, OX ) an integral affine curve.

A morphism of integral affine curves is a morphism of the underlying schemes.

Proposition 29.0.2. There is an anti-isomorphism of categories


 
finitely generated integral domains
{integral affine curves over k} −→
A/k with tr degk K = 1
(X, OX ) 7−→ O(X)

(ϕ : Y → X) (ϕ# : O(X) → O(Y ))

Spec A →−7 A

(f ∗ : Spec B → Spec A) (f : A → B).

618
Chapter 29. Fundamental Groups of Algebraic Curves

Let X = Spec A be an integral affine curve. Suppose L/k is a field extension such that
A ⊗k L is still a domain. Then XL := Spec(A ⊗k L) is an integral affine curve over L, called
the base change of X to L. The natural map A → A ⊗k L, a 7→ a ⊗ 1 induces a morphism
of ringed spaces XL → X.
Definition. If A⊗k k̄ is an integral domain, we say X = Spec A is geometrically integral.
In this case, XL is integral for all intermediate extensions k̄ ⊇ L ⊇ k. For any of these
L, there is a functor
   
geometrically integral integral affine
FL : −→ .
affine curves over k curves over k

Example 29.0.3. Let X = Spec(R[x, y]/(x2 + y 2 + 1)). Then closed points of X are
in correspondence with maximal ideals of the ring R[x, y]/(x2 + y 2 + 1). These in turn
correspond with Galois orbits of XC . Explicitly, each closed point (x2 , y 2 + 1) of X is covered
by exactly two points (x, y − i) and (x, y + i) via XC → X.
Recall that a curve X is normal at P ∈ X if O(X)P is integrally closed and X is normal
if it is normal at every P ∈ X.
Lemma 29.0.4. Let ϕ : Y → X be a finite morphism. Then
(1) O(Y ) is integral over O(X).

(2) ϕ is surjective.

(3) ϕ# is injective.

(4) k(Y )/k(X) is a finite extension.

(5) ϕ has finite fibres.


Theorem 29.0.5. Let X be an integral normal affine curve. Then the assignments Y 7→
k(Y ) and ϕ 7→ ϕ∗ induce an anti-equivalence of categories
   
normal affine curves over k with embeddings of fields
−→ .
finite maps Y → X k(X) ,→ L

Proof. Given an embedding i : k(X) ,→ L, let B be the integral closure of O(X) in L. Then
by commutative algebra, B is a finitely generated k-algebra which is integrally closed, and
since L/k(X) is a finite extension, tr degk L = 1. Hence by Proposition 29.0.2, Y = Spec B
is a normal affine curve with a morphism ϕ = i∗ : Y → X. One now checks that these
assignments are natural inverses.
Definition. For an integral affine curve X and a field extension L/k(X), the curve Y
constructed above is called the normalization of X in L.
Remark. If X is geometrically integral, then XL is equal to the normalization of X in L
for any L ⊇ k(X).

619
29.1. Proper Normal Curves Chapter 29. Fundamental Groups of Algebraic Curves

29.1 Proper Normal Curves


To fill out the analogy with the topological case (Section 28.1), we want to consider some
version of compact curves. Let X be an affine curve with function field K = k(X).

Lemma 29.1.1. The set {OX,P | P 6= 0} is equal to the set of DVRs of K containing O(X).

Proof. On one hand, (⊆) is obvious. For the other containment, let R be such a DVR and
call its maximal ideal m. Then p := m ∩ O(X) is a nonzero maximal ideal in O(X). This
means O(X)p ⊆ R but each of these is a DVR with the same function field, so we must have
OX,P = R.

Example 29.1.2. Let X = A1k so that O(A1k ) = k[t], the polynomial ring in a single
indeterminate. If Y is also A1k , we may take O(Y ) = k[t−1 ]. Let R be a DVR with Frac(R) =
k(X) ∼= k(t) ∼
= k(t−1 ). Then either R ⊃ O(X) or R ⊃ O(Y ). In fact, the only DVR not
containing O(X) is R = k[t−1 ](t−1 ) , while for O(Y ) the exception is R0 = k[t](t) . We want
some way of “gluing” these two DVRs together so that all DVRs of the function field over k
are considered.

Now assume X is a normal affine curve. By Noether normalization, there exists a function

f ∈ O(X) such that O(X) is a finitely generated
h i k[f ]-module. Define X to be the normal
curve associated to the integral closure of k f1 in K. Then every DVR R with fraction field
K is a local ring for X + = X or for X − , corresponding to whether f ∈ R or f1 ∈ R.
More abstractly, for a finitely generated field extension K/k of transcendence degree 1,
let XK be the set of DVRs of K/k. Define a topology on XK by declaring the complements
T
of finite subsets to be open. Define a sheaf of rings on XK by U 7→ OXK (U ) := R∈U R.

Theorem 29.1.3. (XK , OXK ) is a ringed space which is “locally affine” in the sense that
+ − + −
there is a decomposition XK = XK ∪ XK where XK and XK are affine curves.

Definition. The ringed space (XK , OXK ) is called an integral proper normal curve.

Definition. A morphism of (integral) proper normal curves ϕ : YL → XK is a


morphism of the underlying ringed spaces in which the morphism of sheaves ϕ# : OXK → OYL
is a morphism of local rings, i.e. sends maximal ideals to maximal ideals.
ϕ
Proposition 29.1.4. The functor that sends a morphism of proper normal curves YL − → XK
to the induced injection of fields ϕ∗ : K ,→ L is an anti-equivalence of categories
   
proper normal curves with finite ∼ finite field extensions

→ .
surjections YL → XK K ,→ L over k

We will now drop the subscript on XK denoting the field unless it becomes convenient
to have it. Take a morphism of proper normal curves ϕ : Y → X with affine covers
X = X + ∪ X − and Y = Y + ∪ Y − . Then there are induced morphisms of affine curves
ϕ+ : Y + → X + and ϕ− : Y − → X − .

620
29.1. Proper Normal Curves Chapter 29. Fundamental Groups of Algebraic Curves

Definition. We say an open set U ⊆ X in a proper normal curve is affine if OX (U ) is a


finitely generated k-algebra.

Lemma 29.1.5. Let X be a proper normal curve. Then a subset U ⊆ X is open if and only
if U ∼
= Spec A for some finitely generated k-algebra A.

Proposition 29.1.6. There is an equivalence of categories


 
∼ affine open subsets of
{normal affine curves over k} −
→ .
proper normal curves over k

Proof. (Sketch) Given a normal affine curve X, set X + = X and embed X into X 0 =
X + ∪ X − as shown above. This yields a proper normal curve, and to show the bijection is an
equivalence of categories, one need only check that a morphism of affine curves ϕ : Y → X
extends uniquely to ϕe : Y 0 → X 0.

Definition. We say a morphism of proper normal curves ϕ : Y → X is finite if for any


affine U ⊆ X,

(1) ϕ−1 (U ) is affine in Y .

(2) ϕ∗ O(U ) is a finitely generated O(U )-module.

Notice that for any affine set U ⊆ X and finite morphism ϕ : Y → X, the restriction
ϕ|ϕ−1 (U ) : ϕ−1 (U ) → U is a finite morphism of affine curves. Thus Lemma 29.0.4 yields:

Corollary 29.1.7. Any finite morphism is surjective.

Conversely, we have:

Proposition 29.1.8. Let ϕ : YL → XK be a surjective morphism of proper normal curves.


Then ϕ is finite.

Proof. Let U ⊆ X be affine. Then ϕ−1 (U ) = {R | R is a DVR of L containing O(U )}. For
any R ∈ ϕ−1 (U ), the integral closure B of O(U ) is contained in R since R is itself integrally
closed. But by Theorem 29.0.5, this B is precisely the integral closure of the normalization
V of U in L. Hence V = ϕ−1 (U ) so (1) is satisfied. It also follows from the identification of
V as the normalization of U in L that ϕ∗ O(U ) = O(ϕ−1 (U )) = O(V ) is finitely generated
over O(U ).

Corollary 29.1.9. For any proper normal curve X, the assignment Y 7→ k(Y ) induces an
anti-equivalence of categories
   
proper normal curves with finite ∼ finitely generated field extensions k(X) ,→ L

→ .
morphisms Y → X with tr degk L = 1

621
29.2. Finite Branched Covers Chapter 29. Fundamental Groups of Algebraic Curves

29.2 Finite Branched Covers


In this section we describe the analogue of branched covers in the algebraic setting. We begin
with the case of affine curves and later generalize to proper normal curves. Fix a morphism
ϕ : Y → X of integral affine curves. If ϕ is finite and separable, then it is étale at a closed
point P ∈ X if O(Y )/P O(Y ) is a finite étale algebra over the residue field κ(P ) = O(X)/P .
Further suppose X and Y are normal affine curves. Then O(X) and O(Y ) are Dedekind
domains, so P O(Y ) can be written
r
Y
P O(Y ) = Qei i
i=1

for distinct closed points Qi ∈ Y and integers ei . By the Chinese remainder theorem,
r
O(Y )/P O(Y ) ∼
Y
= (O(Y )/Qei i ).
i=1

Definition. The integer ei is called the ramification index of ϕ at Qi (over P ). If ei > 1


for any Qi , we say ϕ is ramified at P (or P is a branch point of ϕ).
The following is a summary of the discussion in Example 15.5.7.
Proposition 29.2.1. For aQmorphism of normal affine curves ϕ : Y → X and a point
P ∈ X such that P O(Y ) = ri=1 Qei i , the following are equivalent:
(1) ϕ is étale at P .
(2) ei = 1 for each 1 ≤ i ≤ r and each residue field κ(Qi ) = O(Y )/Qi is separable over
κ(P ).
(3) For each Qi over P , κ(Qi )/κ(P ) is separable and mP OY,Qi = mQi , where mP ⊂ OX,P
and mQi ⊂ OY,Qi are the maximal ideals in the given local rings.
Note that O(Y )/P O(Y ) ∼ = (ϕ∗ OY )P ⊗OX,P κ(P ). We can view the set {Qi } as the
geometric fibre ϕ−1 (P ), and under this identification O(Y )/P O(Y ) acts as the space of
regular functions on ϕ−1 (P ).
Example 29.2.2. Let k = C and X = A1C . Then the space of regular functions at the points
0 and 2 is given by
C[t]/t(t − 2) ∼
= C × C.
Example 29.2.3. Let k = C and define the map
ϕn : AnC → AnC , x 7→ xn .
This induces the ring extension C[tn ] ,→ C[t]. Here, the maximal ideals are (tn − a) for
a ∈ C, and we have the following ramification behavior:
(Q √
n ∼
n−1
i=0 (C[t]/(t − ζn
i n
a)) ∼
= Cn , a 6= 0
C[t]/(t − a) =
C[t]/(tn ), a = 0.
Thus ϕn is étale at each a 6= 0 but not at a = 0.

622
29.2. Finite Branched Covers Chapter 29. Fundamental Groups of Algebraic Curves

The next two results were proven in a more general context in Section 15.6; however,
concrete proofs are given here to illustrate some of the techniques we will need in future
discussions of étale and branched covers.
ψ ϕ
Lemma 29.2.4. If Z − →Y −
→ X are finite separable morphisms of affine curves and P ∈ X
is a closed point, then

(1) If ϕ is étale at P and ψ is étale over ϕ−1 (P ), then ψ ◦ ϕ is étale at P .

(2) If X, Y and Z are normal affine curves, then the converse holds.

Proof. Let X = Spec A, Y = Spec B and Z = Spec Z for k-algebras A, B and C. Then we
ϕ∗ ψ∗
have ring extensions A −→ B −→ C as κ(P )-algebras. By the above comments,

C/P C ∼
= C ⊗A κ(P )

= C ⊗B (B ⊗A κ(P ))
!

Y
= C ⊗B κ(Q)
Q→P


Y
= κ(R).
R→P

Thus the composition is étale at P if each morphism


Pr is. Now suppose the curves are all
normal. Then the ramification indices satisfy i=1 ei fi = [Frac(B) : Frac(A)], where fi =
[κ(Qi ) : κ(P )] and likewise for the ring extensions C/B and C/A. It follows easily that if
one of ϕ, ψ is ramified, then so is the composition.

Proposition 29.2.5. Let ϕ : Y → X be a finite separable morphism of affine curves. Then


there is a nonempty open set U ⊆ X such that ϕ is étale over U . In particular, ϕ has finitely
many branch points.

Proof. Let X = Spec A and Y = Spec B. By hypothesis, B is a finitely generated A-module,


so write B = A[f1 , . . . , fr ] for integral elements f1 , . . . , fr . Then there is a tower of ring
extensions
A ⊆ A[f1 ] ⊆ A[f1 , f2 ] ⊆ · · · ⊆ A[f1 , . . . , fr−1 ] ⊆ B.
This determines a sequence of morphisms

Y → Xr−1 → · · · → X2 → X1 → X

whose composition is ϕ, where Xi = Spec A[f1 , . . . , fi ] for 1 ≤ i ≤ r − 1. By induction and


using Lemma 29.2.4, we may reduce to the case Y = X1 → X, i.e. B = A[f ] for some
integral f = f1 ∈ B. Write B = A[t]/(F (t)) where F ∈ A[t] is the minimal polynomial
over A of f . Since ϕ is separable, k(Y )/k(X) is a separable extension of fields and thus
(F 0 , F ) = 1 in k(X)[t]. Thus there exist G1 , G2 ∈ k(X)[t] such that G1 F 0 + G2 F = 1.
Cancelling denominators, there exists some g ∈ A such that H1 = G1 g and H2 = G2 g in
A[t], and so we have
H1 F 0 + H2 F = g in A[t].

623
29.2. Finite Branched Covers Chapter 29. Fundamental Groups of Algebraic Curves

Let U = D(g) = {P ∈ X | g(P ) 6= 0} which is a nonempty open subset of X. We claim ϕ is


étale over U . For any P ∈ U , take the reduction of the above equation in κ(P )[t] to see that
0
H 1 F + H 2 F = ḡ 6= 0 in κ(P )[t].
0
Thus for all α ∈ A, F (α) = 0 implies F (α) 6= 0, or in other words, F has no multiple roots.
Therefore we can write
s
= κ(P )[t]/(F ) ∼
B/P B ∼
Y
= κ(P )[t]/(F i )
i=1

where F 1 , . . . , F s are the irreducible factors of F in κ(P )[t]. The work above shows that each
κ(P )[t]/(F i ) is separable, so B/P B is a finite étale κ(P )-algebra. The result follows.
In light of Proposition 29.2.5, the following definitions make sense:

Definition. We call a morphism ϕ : Y → X of integral affine curves a finite branched


cover if it is finite and separable. Further, we call ϕ a (finite) Galois branched cover if
the field extension k(X) ,→ k(Y ) is Galois.

If, in addition to ϕ being a finite Galois cover, X and Y are normal curves, then
Gal(k(Y )/k(X)) acts transitively on the fibres ϕ−1 (P ) of ϕ, just as in the case of a Ga-
lois topological cover.

Proposition 29.2.6. If ϕ : Y → X is a Galois branched cover over a perfect field k and X


and Y are normal affine curves, then ϕ is étale at P ∈ X if and only if IQi = {1} for all
Qi ∈ ϕ−1 (P ), where IQi is the inertia subgroup of Gal(k(Y )/k(X)).

Proof. Recall the definitions of the decomposition and inertia groups for a point Qi ∈ ϕ−1 (P ):

DQi = {σ ∈ Gal(k(Y )/k(X)) | σ(Qi ) = Qi }


IQi = {σ ∈ Gal(k(Y )/k(X)) | σ acts trivially on κ(Qi )}.

Since k is perfect, we know from algebraic number theory that ei = [κ(Qi ) : κ(Qi )IQi ], where
κ(Qi )IQi is the subfield of κ(Qi ) fixed by the inertia group, but this index is precisely the order
of IQi . Hence ei = 1 if and only IQi = {1}, so the result follows by Proposition 29.2.1.
For a normal affine curve X and a tower of fields (finite over k(X))

k(X) ⊂ K1 ⊂ K2 ⊂ · · ·

set L = ∞
S
j=1 Kj and let Aj be the integral closure of O(X) in Kj for each j ≥ 1. From this
we get a sequence of morphisms of curves

· · · → X2 → X1 → X.

Fix P ∈ X. Choose maximal ideals {Pj }j≥1 with Pj ∈ Aj such that Pj ∩ O(X) = P and
Pj+1 ∩ Aj = Pj for all j ≥ 1. If Ij is the inertia subgroup of Pj in Gj := Gal(Ki /k(X)), then

624
29.2. Finite Branched Covers Chapter 29. Fundamental Groups of Algebraic Curves

the Ij form an inverse system with the natural maps coming from each quotient of Galois
groups. Define the inertia subgroup of P to be the inverse limit

IP := lim Ij .
←−

Since each Ij is a normal subgroup of Gj , it follows that IP is a closed subgroup of the


profinite group G := Gal(L/k(X)). Note that IP depends on the choices of P1 , P2 , . . . but a
different choice of points lying over P yields a conjugate subgroup.

Corollary 29.2.7. Let ϕ : Y → X be a Galois branched cover of normal affine curves over
a perfect field and assume k(Y ) ⊆ L where L is as above. Then ϕ is étale at P ∈ X if and
only if the image of IP in Gal(k(Y )/k(X)) = G/ Gal(L/k(Y )) is trivial.

Proof. Pick ` ≥ 1 so that k(Y ) ⊆ K` . Let Q = P` ∩ O(Y ). Then the image of IP in


Gal(k(Y )/k(X)) is precisely the inertia subgroup IQ ≤ Gal(k(Y )/k(X)). Now apply Propo-
sition 29.2.6.
We now extend the notion of finite branched covers to proper normal curves. Suppose ϕ :
Y → X is a finite separable morphism of proper normal curves (as defined in Section 29.1).

Definition. We say ϕ : Y → X is étale at P ∈ X if there exists an affine open neighborhood


U ⊆ X of P such that ϕ|ϕ−1 (U ) : ϕ−1 (U ) → U is étale at P as a morphism of affine curves.
As in the affine case, call ϕ étale over a subset S ⊆ X if it is étale at every point P ∈ S,
and simply étale if it is étale at every P ∈ X.

Note that by Lemma 29.2.1, this definition does not depend on the neighborhood U of
P chosen. The last few results for affine curves generalize to proper normal curves as an
immediate consequence of the definition. Again, these also follow from results in Section 15.6.

Theorem 29.2.8. Let ϕ : Y → X be a morphism of proper normal curves. Then

(1) There exists a nonempty open set U ⊆ X such that ϕ is étale over U .

(2) If ϕ is a Galois branched cover and k is perfect, then ϕ is étale at P ∈ X if and only
if the inertia group is trivial for each point Qi over P .

(3) If ϕ is a Galois branched cover, k is perfect L = ∞


S
j=1 Kj and k(Y ) ⊆ L, then ϕ is
étale at P ∈ X if and only if IP maps trivially into Gal(k(Y )/k(X)).

Example 29.2.9. For a proper normal curve X over C, X(C) has the structure of a compact
Riemann surface such that the cover X(C) → P1 (C) corresponding to X → P1 is proper and
holomorphic.

Proposition 29.2.10. For a proper normal curve X over C, the following categories are
anti-equivalent:
 
     compact connected Riemann 
proper normal curves Y finite extensions
←→ ←→ surfaces Y with proper .
with finite morphism Y → X k(X) ,→ L
holomorphic maps Y → X(C)
 

625
29.2. Finite Branched Covers Chapter 29. Fundamental Groups of Algebraic Curves

Example 29.2.11. Consider the squaring map

ρ2 : P1R −→ P1R , t 7→ t2

corresponding to the ring extension R[t2 ] ,→ R[t]. The maximal ideals of R[t] are of the form

(1) (t − a) for a ∈ R

(2) (t2 + bt + c) for b2 − 4c < 0.

For (1), we have the following ramification behavior:



R × R,
 a>0
2 ∼ 2
R[t]/(t − a) = R[t]/(t ), a = 0

a<0

C,

while for (2), we get


R[t]/(t4 + bt2 + c) ∼
= C × C.
Note that κ(t − a) = R, while κ(t2 + bt + c) = C so we see that ρ2 is étale off of {0, ∞}. (To
see that ρ2 is not étale at ∞, repeat the above argument replacing t with 1/t.)

Example 29.2.12. Let E be an elliptic curve over an algebraically closed field k of char-
acteristic char k = p. Take m ∈ Z such that p - m and consider the multiplication-by-m
isogeny:
[m] : E −→ E, P 7→ mP.
We know that [m] is separable when p - m, and moreover that E[m] = (Z/mZ)2 and thus
#E[m] = m2 . For each P ∈ E, we have
m 2 m 2

O(E)/[m]P ∼ κ(P ) ∼
Y Y
= = k
i=1 i=1

and hence [m] is étale. In general, any degree-m isogeny ϕ : E1 → E2 of elliptic curves over
an algebraically closed field of characteristic not dividing m is separable and therefore étale
by the same argument.

Example 29.2.13. Let k be any field of characteristic p > 0 and consider the field

L = k(x)[y]/(y p − (f (x)g(x))p−1 y − f (x)).

Note that y p − (f (x)g(x))p−1 y − f (x) is irreducible, e.g. by Eisenstein’s criterion, so L is


indeed a field. Let B be the integral closure of k[x] in L and let X be the proper normal curve
corresponding (via Proposition 29.1.4) to L. Then there is a morphism of curves ϕ : X → P1k
coming from k(x) ,→ L. For a ∈ k, consider the maximal ideal (x − a) ∈ B. We consider
three cases:

626
29.2. Finite Branched Covers Chapter 29. Fundamental Groups of Algebraic Curves

(1) If f (a) = 0, then (x − a) = (y p ) in B, so we have

B/(x − a)B ∼
= B/(y p )B.

Thus ϕ is not étale at (x − a).

(2) If f (a) 6= 0 but g(a) = 0, (x − a) = (y p − f (a)) so


(
B/(y − f (a)1/p )p , f (a)1/p ∈ k
B/(x − a)B ∼ =
B/(y p − f (a)), otherwise.

In the first case, B/(y − f (a)1/p )p is not a field extension of κ(x − a) = k, while in
the second, B/(y p − f (a)) is an inseparable field extension of k. Thus ϕ is not étale
at any of these (x − a).

(3) Finally, if f (a) 6= 0 and g(a) 6= 0, then (x − a) = (y p − (f (a)g(a))p−1 y − f (a)). If there


is a c ∈ k such that f (a) = cp − c, then
p−1 p−1

Y Y
(x − a) = (x − (c + if (a)g(a))) =⇒ B/(x − a)B = k.
i=0 i=0

Thus ϕ is étale at (x − a) in this case. On the other hand, if there does not exist such
a c, then (x − a) remains a prime ideal in B and thus B/(x − a) is a finite separable
extension of k.

Hence ϕ is étale at the point (x − a) if and only if f (a) 6= 0 and g(a) 6= 0.

627
29.3. Fundamental Group of CurvesChapter 29. Fundamental Groups of Algebraic Curves

29.3 Fundamental Group of Curves


Let k be a perfect field, X an integral proper normal curve over k, K = k(X) its function
field and fix a separable closure Ks of K. For a nonempty open subset U ⊆ X, define
KU to be the compositum of all finite subextensions Ks ⊇ L ⊇ K corresponding to finite
morphisms of proper normal curves Y → X which are étale over U .

Proposition 29.3.1. For any nonempty open subset U ⊆ X, KU /K is a Galois extension


and each finite subextension corresponds to a cover of X which is étale over U .

Proof. To show KU /K is a Galois extension, it’s enough to see that KU is stable under the
Gal(Ks /K)-action, but this follows from the definition of the field KU . To prove the second
statement, it’s enough to show for a finite subextension L/K of KU /K that:

(i) If L comes from a cover of X which is étale over U , then so does any subfield L ⊇ L0 ⊇
K.

(ii) If M/K is any other finite subextension of KU /K coming from a cover which is étale
over U , then LM/K also comes from a cover étale over U .

Then the result will follow since KU is the compositum of a tower of finite extensions L1 ⊆
L1 L2 ⊆ L1 L2 L3 ⊆ · · ·
ϕ
For (i), let L/K correspond to a cover Y − → X which is étale over U . Then for any
ϕ0
L ⊇ L0 ⊇ K, we get a composition of covers Y → Y 0 − → X. Applying Lemma 29.2.4,
0
we get that ϕ is étale over U . For (ii), fix P ∈ U and consider the inertia group IP ≤
Gal(Ks /K) as defined in Section 29.2. Then by (3) of Theorem 29.2.8, IP is trivial in
both Gal(L/K) and Gal(M/K). Since Gal(L/K) = Gal(Ks /K)/ Gal(Ks /L), this means
IP ⊆ Gal(Ks /L); likewise IP ⊆ Gal(Ks /M ). Hence IP ⊆ Gal(Ks /LM ), which means IP is
trivial in Gal(LM/K), so by Theorem 29.2.8 again, the curve corresponding to LM/K is
étale at P .

Definition. For a proper normal curve X over a perfect field k and a nonempty open subset
U ⊆ X, we define the algebraic fundamental group over U to be

π1 (U ) := Gal(KU /K).

Note that π1 (U ) is a profinite group which depends on the choice of separable closure
Ks . Here, this choice of Ks acts as the “basepoint” in analogy with the topological case (see
Section 28.1).

Definition. Extending the notion of curves from the last section, we define a proper nor-
mal curve to be a finite disjoint union of integral proper normal curves.
`
The ring of rational functions on a proper normal curve X L = Xi is defined to be the
direct sum of the function fields of each component, O(X) = k(Xi ); this is naturally a
finite dimensional algebra over any of the components k(Xi ). In general, we say a morphism
Y → X, with X integral, is separable if k(Y ) is étale over k(X).

628
29.3. Fundamental Group of CurvesChapter 29. Fundamental Groups of Algebraic Curves

Theorem 29.3.2. Suppose X is an integral proper normal curve over a perfect field k and
U ⊆ X is a nonempty open subset. Then there is an equivalence of categories
 ϕ 
covers of proper normal curves Y − →X ∼

→ {finite continuous left π1 (U )-sets}.
where ϕ is finite, separable, étale over U

L sums of fields, so for a cover Y → X


Proof. Propositions 29.1.4 and 29.3.1`extend to direct
which is a proper normal curve Y = Yi , let A = Li be the finite étale K-algebra corre-
sponding to these Yi → X. Finally, by Corollary 28.2.4, Hom(A, Ks ) is a finite continuous
Gal(Ks /K)-set.
Now, for an integral normal affine curve U over k, we know U is an affine open subset of
an integral proper normal curve X = U ∪ U − . Thus π1 (U ) is defined.
Corollary 29.3.3. For any integral normal affine curve U over a perfect field k, there is an
equivalence of categories
 
normal affine curves with ∼

→ {finite continuous left π1 (U )-sets}.
finite étale covers V → U
Proof. Any finite étale cover of normal affine curves V → U extends by Proposition 29.1.6
to a cover of proper normal curves Y L → X K which is in turn finite by Proposition 29.1.8.
Finally, Theorem 29.3.2 gives the desired correspondence.
We now discuss how to capture the notion of a “universal cover” as we have in the
topological case (Section 28.1). Let U ⊆ X be a nonempty open subset of an integral proper
normal curve and let A = O(U ) be its ring of rational functions. Take A e be the integral
closure of A in KU . Then for a finite étale cover V → U with fraction field K(V ), we have
O(V ) = Ae ∩ K(V ).

Ks

KU A
e

K(V ) O(V )

K A

Note that for any maximal ideal m ⊆ A, e m ∩ O(V ) is a closed point in V . Although Ae is
not itself a finitely generated ring over A, it is a compositum of such rings, i.e. the O(V ) are
finite over A. Define U e = MaxSpec A e and equip this with the inverse limit topology with
respect to the inverse system V → V 0 → U . We define a sheaf of rings OUe on U e stalkwise
by setting OUe ,Qe = A e∈U
e e for any Q
Q
e . Then for any open set Ve ⊆ Ue , we put
\
OUe (Ve ) = OUe ,Qe .
Q∈
e Ue

629
29.3. Fundamental Group of CurvesChapter 29. Fundamental Groups of Algebraic Curves

Definition. U
e is called the pro-étale cover of U .
e , O e ) is a locally ringed space.
Lemma 29.3.4. (U U

Remark. Note that U e is not an object in the category of proper normal curves (or even
in the category of schemes, as we shall see!) However, constructing a “universal cover” is
useful to illustrate the analogy with the topological case. For example, we have the following
theorem which will be proven in Section 30.3.
Theorem 29.3.5. Let X be an integral proper normal curve over C and U ⊆ X a nonempty
open subset. Then π1 (U ) is isomorphic to π1top\
(U (C)), the profinite completion of the topo-
logical fundamental group of the Riemann surface U (C). In particular, π1 (U ) is the profinite
completion of
ha1 , b1 , . . . , ag , bg , γ1 , . . . , γn | [a1 , b1 ] · · · [ag , bg ]γ1 · · · γn = 1i,
where n is the number of points in X r U and g is the genus of the Riemann surface U (C).
Let X/C and U be as above. If G is any finite quotient of π1 (U ), then G corresponds
to a finite Galois branched cover Y → X. Here, the image of each generator γi ∈ π1 (U ) in
G is a cyclic generator of the stabilizer of a point Qi → Pi , where Pi ∈ X r U is the point
corresponding to γi .
Let X be an integral proper normal curve over an arbitrary (perfect) field k. If L/k is
a finite extension, then k(X) ⊗k L is a finite dimensional algebra over L(t) which is not, in
general, a field. Assume k(X) ⊗k L is a finite direct product of fields L1 , . . . , Ln , each of
which is a finitely generated extension of L(t) of transcendence degree 1 over L. Then each
Li corresponds to an ` integral proper normal curve Xi /L by Proposition 29.1.4. Define the
base change XL = Xi . This comes equipped with a natural morphism of proper normal
curves XL → X. We will prove:
Theorem 29.3.6. If k is an algebraically closed field of characteristic 0, L/k is a finite
extension, X is an integral proper normal curve over k and U ⊆ X is open, then the base
change functor X 7→ XL induces an equivalence of categories
   
finite branched Galois covers Y → X ∼ finite branched Galois covers YL → XL

→ .
étale over U étale over UL
Corollary 29.3.7. For U ⊆ X as above, there is an isomorphism π1 (UL ) ∼
= π1 (U ) for all
finite extensions L/k.
Theorem 29.3.5 generalizes to curves over an algebraically closed field k of characteristic
0 in the following sense.
Corollary 29.3.8. If k is algebraically closed of characteristic 0 and U ⊆ X is a nonempty
open subset of an integral proper normal curve X over k, then π1 (U ) is isomorphic to the
profinite completion of a group with presentation
ha1 , b1 , . . . , ag , bg , γ1 , . . . , γn | [a1 , b1 ] · · · [ag , bg ]γ1 · · · γn = 1i,
where g = g(X) is the genus of X and n is the number of points in X r U .

630
29.3. Fundamental Group of CurvesChapter 29. Fundamental Groups of Algebraic Curves

Example 29.3.9. If k is an algebraically closed field of characteristic 0 and X → P1k is a


finite étale cover, the Riemann-Hurwitz formula (Theorem 17.7.2) says that

2g(X) − 2 = n(0 − 2) + 0

but this is only possible if g(X) = 0 and n = 1. Thus X → P1k is a birational isomorphism,
but since X is complete, it must be a regular isomorphism. Hence there are no nontrivial
extensions L of k(P1k ) = k which proves π1 (P1k ) = 1.

Example 29.3.10. A similar argument shows π1 (A1k ) = 1 when k is algebraically closed and
char k = 0. Both this and the previous result also follow easily from Theorem 29.3.5 and
Corollary 29.3.8.

Example 29.3.11. Theorem 29.3.5 and Corollary 29.3.8 also imply that π1 (P1k r {0, ∞}) ∼ =
Z, the profinite completion of the integers. Using Theorem 29.3.2, we get a finite étale cover
b
Xn → P1k r {0, ∞} for each n ≥ 1 having Galois group Z/nZ.

Example 29.3.12. By Theorem 29.3.5 and Corollary 29.3.8, π1 (P1k r {0, 1, ∞}) is the free
profinite group on two generators.

631
29.4. The Outer Galois Action Chapter 29. Fundamental Groups of Algebraic Curves

29.4 The Outer Galois Action


Let X be an integral proper normal curve over a perfect field k which is not necessarily alge-
braically closed. Let K = k(X). Fixing an algebraic closure k̄, we assume X is geometrically
integral, that is, K ⊗k k̄ is a field. Then Ks , the separable closure of K containing k̄, is defined.
For any nonempty open U ⊆ X, the base change Uk̄ is integral and we have K k̄ ⊆ KU , where
the compositum is taken in KU . By construction, for any finite extension L/k, the cover
XL → X is finite étale with Gal(k(XL )/K) ∼ = Gal(L/k). Thus Gal(K k̄/K) ∼ = Gal(k̄/k) so
the latter may be thought of as a quotient of π1 (U ) for any open U ⊆ X.

KU

π1 (U ) K k̄

Gal(k̄/k)

Proposition 29.4.1. For X a geometrically integral, proper normal curve over a perfect
field k and for any nonempty open subset U ⊆ X, there is a short exact sequence of profinite
groups
1 → π1 (Uk̄ ) → π1 (U ) → Gal(k̄/k) → 1.

Proof. The surjection π1 (U ) → Gal(k̄/k) was described above, so it remains to identify


π1 (Uk̄ ) with the kernel of this map. Suppose G is a finite quotient of π1 (Uk̄ ). Then there is
a corresponding finite extension K0 /K k̄:
KUk̄

π1 (Uk̄ ) K0

K k̄

This also determies a finite branched cover Y0 → Xk̄ which is étale over Uk̄ ⊆ Xk̄ . Let
f ∈ K k̄[t] be the minimal polynomial of K0 /K k̄. Then we can find a finite extension L/k
such that KL/k contains all the coefficients of f . Let L0 /KL be the finite extension with
Gal(L0 /KL) = G.

632
29.4. The Outer Galois Action Chapter 29. Fundamental Groups of Algebraic Curves

K0
G
π1 (U ) L0 K k̄
G
KL k̄

K L

By construction, L0 k̄ = K0 and the tower L0 ⊇ KL ⊇ L determines a composition of covers


Y → XL → X which is étale over U , and such that Yk̄ = Y0 . This implies L0 ⊆ KU and then
K0 = L0 k̄ ⊆ KU . This shows that we can identify any such G = Gal(L0 /KL) with a finite
quotient of ker(π1 (U ) → Gal(K k̄/K)). Varying the finite quotients G of π1 (Uk̄ ), we conclude
that KUk̄ ⊆ KU and hence there is a surjection Gal(KU /K k̄) → Gal(KUk̄ /K k̄) = π1 (Uk̄ )
which is bijective on finite quotients. Taking inverse limits – which is a left exact functor –
gives π1 (Uk̄ ) = ker(π1 (U ) → Gal(K k̄/K)) as required.

Corollary 29.4.2. There is a continuous homomorphism pU : Gal(k̄/k) → Out(π1 (Uk̄ ) for


any nonempty open set U ⊆ X.

Proof. Any short exact sequence of groups 1 → N → G → P → 1 defines a homomorphism


G → Aut(N ) whose restriction to the subgroup N ≤ G has image lying in Inn(N ). By
definition, Out(N ) = Aut(N )/ Inn(N ) so we get a map P → Out(N ). Applying this to the
short exact sequence from Proposition 29.4.1 gives the result.

Definition. For an open subset U of a geometrically integral proper normal curve X over a
perfect field k, the subgroup π1 (Uk̄ ) ≤ π1 (U ) is called the geometric fundamental group
of U . The map pU : Gal(k̄/k) → Out(π1 (Uk̄ ) is called the outer Galois action on the
geometric fundamental group.

We next study the action of π1 (U ) on X,


e the pro-étale cover of X defined in Section 29.3.
Let Qe be a pro-point in X e lying over a point P ∈ X. We define D e to be the stabilizer
Q
of Q under the π1 (U )-action. Then κ(Q) = OX,
e e eQe /QOX,
e eQ e = κ(P ), the algebraic closure of

the residue field κ(P ) = OX,P /P OX,P . We have a natural surjection DQe → Gal(κ(Q)/κ(P
e ))
with kernel denoted IQe , called the inertia group of Q.e We may alternatively view X e as
a profinite Galois cover of Xk̄ with Galois group π1 (Uk̄ ), corresponding to the short exact
sequence from Proposition 29.4.1:

633
29.4. The Outer Galois Action Chapter 29. Fundamental Groups of Algebraic Curves

X
e

π1 (Uk̄ )

π1 (U ) Xk̄

Gal(k̄/k)

Definition. Let P ∈ X be a closed point such that κ(P ) ∼


= k. Then we say P is k-rational.
Lemma 29.4.3. Let P ∈ X be a k-rational (closed) point. Then the stabilizer of any
e∈X
pro-point Q e lying over P in π1 (Uk̄ ) is equal to the inertia group I e .
Q

Proof. We have a surjection DQe → Gal(κ(Q)/κ(P


e )) = Gal(k̄/k) by assumption, but this
is just the restriction of the map π1 (U ) → Gal(k̄/k) to the action on Q.
e By definition the
kernel of the latter map is IQe .
Corollary 29.4.4. If U contains a k-rational point of X then the short exact sequence
1 → π1 (Uk̄ ) → π1 (U ) → Gal(k̄/k) → 1
is split.
Proof. If Q e is a pro-point over some k-rational point P ∈ U , then the stabilizer of Q
e in π1 (Uk̄ )
is trivial. Therefore Lemma 29.4.3 says that IQe = 1, but by definition of the inertia group
this implies DQe ∼ = Gal(k̄/k). Thus there is a map Gal(k̄/k) → π1 (U ) mapping isomorphically
onto DQe ⊆ π1 (U ) which defines the required splitting.
Example 29.4.5. Let X = P1k and consider the open set U = P1k r {0, ∞}. By Exam-
ple 29.3.11, π1 (U ) ∼ b a free abelian profinite group. Hence Gal(k̄/k) acts on π1 (Uk̄ ) directly,
= Z,
not just by outer automorphisms. Let n ≥ 1. Then the quotient π1 (Uk̄ )/nπ1 (Uk̄ ) ∼ = Z/nZ
n
corresponds to Gal(Ln /k̄(t)) where Ln /k̄(t) is the field extension defined by x − t. Consider
the short exact sequence of profinite groups
1 → Gal(Ln /k̄(t)) → Gal(Ln /k(t)) → Gal(k̄/k) → 1.
Identifying π1 (Uk̄ )/nπ1 (Uk̄ ) = Z/nZ, we see that this sequence √ defines√the Gal(k̄/k)-action
on π1 (Uk̄ )/nπ1 (Uk̄ ). Explicitly, Gal(Ln /k̄(t)) is generated by n t 7→ ζn n t, where ζn ∈ k̄ is a
primitive nth root of unity. Then for σ ∈ Gal(k̄/k), the Galois action is given by

n
√n
√n

n
σ · ( t 7→ ζn t) = t 7→ σ(ζn ) t.
Fixing a compatible system of primitive roots of unity (ζn )n∈N , we get isomorphisms
 
Z = Gal lim Ln /k̄(t) = Gal(k̄(t)cyc /k̄(t)) ∼
b ∼ = π1 (P1k r {0, ∞}).
←−

This defines a representation Gal(k̄/k) → Aut(Z)b =Z b × called the cyclotomic character of


Gal(k̄/k). Notice that this descends to a character Gal(k̄/k) → (Z/nZ)× for each n ≥ 1.

634
29.4. The Outer Galois Action Chapter 29. Fundamental Groups of Algebraic Curves

One of the most important objects in arithmetic geometry is π1 (P1Q r {0, 1, ∞}). If
U = P1Q r {0, 1, ∞}, there is an outer Galois action ρ : Gal(Q/Q) → Out(π1 (UQ )). We will
prove that this action is faithful, but first we need:
Theorem 29.4.6 (Belyi). Suppose X is an integral proper normal curve defined over an
algebraic closed field k of characteristic 0. Then there exists a morphism X → P1k which is
étale over P1k r {0, 1, ∞} if and only if X is defined over Q.
Proof. ( =⇒ ) follows from Theorem 29.3.6.
( ⇒= ) If X is defined over Q, there exists a map π : X → P1Q which is étale over P1Q r S
1
for some finite set S ⊆ PQ . We first show S consists of Q-rational points. Since S is finite,
we can find a point P ∈ S with [κ(P ) : Q] = n maximal among the points in S. Let
f ∈ Q[t] be the minimal polynomial of κ(P )/Q. Define ϕf : P1Q → P1Q by x 7→ F (x), where
F is the homogenization of f . Then ϕf is defined over Q and is étale over P1Q r Sf , where
Sf = {R ∈ AQ 1
| f 0 (R) = 0}. Notice that deg f 0 = n − 1 so for all R ∈ Sf , [κ(R) : Q] < n.
The composition ϕf ◦ π : X → P1Q is now étale away from S 0 := ϕf (S) ∪ {∞} ∪ ϕf (Sf ). But
now ∞ has degree 1, points in ϕf (S) have degree at most n and points in ϕf (Sf ) have degree
at most n − 1, so because ϕf (P ) = 0, there are strictly less points of degree n in S 0 than in
S. Repeat this procedure until n = 1, at which time all points in S 0 will be Q-rational.
Next, suppose S has more than three points. We may assume 0, 1, ∞ ∈ S and take
α ∈ S r {0, 1, ∞}. Define the Belyi function
1
ϕ : PQ −→ P1Q
x 7−→ xA (x − 1)B

where A, B ∈ Z r {0} are such that α = A+B A


. As above, ϕ is étale away from S 0 =
ϕ(S) ∪ {∞} ∪ ϕ(Sϕ ) where Sϕ = {R ∈ A1Q | ϕ0 (R) = 0}. Note that ϕ0 (x) = AxA−1 (x − 1)B +
BxA (x − 1)B−1 so ϕ0 (R) = 0 precisely when R ∈ {0, 1, ∞, α}. This shows that ϕ(S) = S 0 .
Composing with π, we get a map ϕ ◦ π : X → P1Q which is étale outside ϕ(S) and maps
{0, 1, ∞} to {0, ∞}. Hence |ϕ(S)| < |S|, so repeating this process reduces to the case when
S 0 has (at most) three elements.
Theorem 29.4.7. The outer Galois representation ρ : Gal(Q/Q) → Out(π1 (P1Q r{0, 1, ∞}))
is faithful.
ker ρ
Proof. Let L = Q be the subfield fixed by the kernel of the outer Galois representation.
Then for U = P1Q r {0, 1, ∞}, Proposition 29.4.1 gives us a short exact sequence of profinite
groups
1 → π1 (UQ ) → π1 (UL ) → Gal(Q/L) → 1.
By construction, ρUL : Gal(Q/L) → Out(π1 (UQ )) has trivial kernel, which means any auto-
morphism of π1 (UQ ) in π1 (UL ) is given by conjugation by some y ∈ π1 (UQ ). That is, for any
x ∈ π1 (UL ), there exists a y ∈ π1 (UQ ) such that for all a ∈ π1 (UQ ), we have xax−1 = yay −1 .
This is equivalent to y −1 xax−1 y = a, i.e. y −1 x belongs to C, the centralizer of π1 (UQ ) in
π1 (UL ). Hence π1 (UL ) is generated by π1 (UQ ) and C. However, Corollary 29.3.8 says that

635
29.4. The Outer Galois Action Chapter 29. Fundamental Groups of Algebraic Curves

π1 (UQ ) is a free profinite group on 2 generators, and such a group has trivial center, so we
must have π1 (UQ ) ∩ C = {1}. Thus π1 (UL ) ∼ = π1 (UQ ) × C as a direct product, or in other
words, π1 (UL ) → π1 (UQ ) is a continuous retraction of profinite groups. This says that any
continuous left π1 (UQ )-set comes from a continuous left π1 (UL )-set, but by Theorem 29.3.6,
any finite cover of UQ is obtained by base change from some finite cover of UL . Finally,
Belyi’s theorem says that any such cover defined over Q can be defined over L, but we will
demonstrate that this is impossible unless L = Q, in which case ker ρ is trivial.
If L 6= Q, take x ∈ Q r L. Let E 0 be an elliptic curve over Q with j-invariant j(E 0 ) = x.
If E 0 can be obtained by base change from some genus 1 curve X defined over L, as Belyi’s
theorem indicates, then X has a Jacobian E/L which is an elliptic curve. There exists an
embedding φ : E → P2L sending the distinguished L-point of E to [0, 1, 0] ∈ P2L . Moreover,
it is known that j(E) ∈ L and this j-invariant is preserved by any Q-isomorphism. Since E 0
is the base change of X, we have E 0 ∼ = XQ ,→ E, but this implies j(E) = j(E 0 ) = x 6∈ L, a
contradiction. Hence L = Q, so the outer Galois representation of U is faithful.

Example 29.4.8. One can similarly prove that for any elliptic curve E over Q (or any
number field K), the outer Galois representation ρ : Gal(Q/Q) → π1 (E r {0}) is faithful.

636
29.5. The Inverse Galois Problem Chapter 29. Fundamental Groups of Algebraic Curves

29.5 The Inverse Galois Problem


For a field k, let Gk = Gal(k̄/k) be the absolute Galois group. We showed in the previous sec-
tion (Theorem 29.4.7) that there is a faithful representation GQ ,→ Out(π1 (P1Q r {0, 1, ∞})),
but this group of automorphisms is inordinately complicated so we don’t have much under-
stand of GQ . A related question we might answer is: what are the finite quotients of GQ .
That is,

Question (Inverse Galois Problem). Which groups arise as Galois groups Gal(L/Q) for
some finite Galois extension L/Q?

Definition. If L/Q has Galois group G, we say L is a G-extension of Q.

Example 29.5.1. Let Cn = Z/nZ be a cyclic group of order n. We know that the cyclotomic
extension Q(ζp ) has Galois group Gal(Q(ζp )/Q) ∼
= Z/(p−1)Z, so by group theory, if n | p−1,
this Galois group has a subgroup H of order (p−1)/n. Hence the subextension Q(ζp )H /Q has
group Gal(Q(ζp )H /Z) = Z/nZ. Thus the inverse Galois problem for cyclic groups reduces to
finding a prime p such that p ≡ 1 (mod n), but by Dirichlet’s theorem, there are infinitely
many such primes p. Thus all cyclic groups arise as Galois groups over Q. Using the theory
of cyclotomic extensions, this can be generalized to give a positive solution to the inverse
Galois problem for all finite abelian groups.

Combined with the Kronecker-Weber theorem (every finite abelian extension of Q lies
in some cyclotomic extension Q(ζm )/Q), the conclusion of Example 29.5.1 gives a complete
description of abelian extensions of Q, namely that S the maximal abelian extension of the
rationals Qab is equal to the compositum Qcyc := m≥1 Q(ζm ). The study of Gal(Qab /Q)
and its subgroups is the content of global class field theory.
If a group G arises as a Galois group over Q, we may then ask for all G-extensions of Q.
This may be a much harder question to answer.

Example 29.5.2. One can show that the splitting field of f (x) = x3 + x2 − 2x − 1 defines
a C3 -extension of Q. How do we determine all C3 -extensions? Well notice that C3 embeds
as a subgroup of P GL2 (Q) = Aut(P1Q ) via

C3 = hσi −→ Aut(P1 )
 
1
σ 7−→ x 7→ .
1−x

Then P1 /C3 is birationally equivalent to P1 and the rational expression T = x + σx + σ 2 x =


x3 − 3x − 1
is an invariant under the C3 -action. Then the fixed subfield Q(x)C3 = Q(T )
x2 − x
corresponds to a cover

P1 −→ P1 /C3 ∼
= P1
x3 − 3x − 1
x 7−→ T = .
x2 − x

637
29.5. The Inverse Galois Problem Chapter 29. Fundamental Groups of Algebraic Curves

By construction, Gal(Q(x)/Q(T )) ∼
= C3 and this extension has minimal polynomial

f (x, T ) = x3 − T x2 + (T − 3)x + 1.

Notice that the discriminant of f is (T 2 − 3T − 9)2 , so for any t ∈ Q such that t2 − 3t − 9 6= 0,


f (x, t) becomes an irreducible polynomial over Q with Galois group Gal(f (x, t)) ∼ = C3 . In
fact, one can show that f (x, T ) parametrizes all C3 -extensions of Q.

Definition. We say a polynomial f (x, T ), with T = (T1 , . . . , Tn ) an n-tuple of indetermi-


nates, is a generic polynomial for a group G over a field K if for all extensions F/K,

(1) The splitting field Ff /F (T ) has group Gal(Ff /F (T )) ∼


= G.
(2) Every G-extension L/F is obtained as the splitting field of f (x, t) for some t =
(t1 , . . . , tn ) ∈ F n .

A polynomial f (x, T ) that satisfies these conditions for some fixed F is called a versal family
for G-extensions of F . Thus a generic polynomial for K is one that is a versal family for
all extensions of K.

Example 29.5.3. We saw in Example 29.5.2 that f (x, T ) = x3 − T x2 + (T − 3)x + 1 is a


versal family for C3 -extensions of Q. One can show (see Theorem 2.2.1 of Jensen, Ledet and
Yui’s Generic Polynomials: Constructive Aspects of the Inverse Galois Problem) that in fact
this f (x, T ) is a generic polynomial for C3 -extensions of Q.

Example 29.5.4. For C8 -extensions of Q, there is a versal family for Q, but there is no
generic polynomial for all extensions of Q.

Remark. Why could we find a one-parameter generic polynomial for C3 -extensions? It is


a fact that if there exists a one-parameter generic polynomial, i.e. f (x, T1 ) = f (x, T ), for
G-extensions of K, then G embeds as a subgroup of P GL2 (K).

Example 29.5.5. By the Remark, the fact that C4 does not embed as a subgroup of
P GL2 (Q) implies that there does not exist a 1-parameter generic polynomial for C4 -extensions
of Q. However, there does exist an embedding C4 ,→ P GL2 (Q(i)), so there may be a 1-
parameter generic polynomial over Q(i).

Let G be a finite group. By Cayley’s theorem, G embeds as a subgroup of some Sn ,


inducing a G-action on the set {x1 , . . . , xn }. Define E = Q(x1 , . . . , xn )G and let π : AnQ →
An /G be the corresponding covering map.

Question (Noether’s Problem). For which groups G is the field E = Q(x1 , . . . , xn )G tran-
scendental over Q?

The inverse Galois problem is known to be a consequence of the so-called regular inverse
Galois problem:

Question (Regular Inverse Galois Problem). Which groups G arise as Galois groups of
finite Galois extensions K/Q(t) which are regular over Q?

638
29.5. The Inverse Galois Problem Chapter 29. Fundamental Groups of Algebraic Curves

(A regular extension K/Q(t) is one which does not contain a subextension of the form
L(t) for L/Q a nontrivial extension.) To see this implication, we need Hilbert’s irreducibility
theorem:
Theorem 29.5.6 (Hilbert Irreducibility). Let G be a finite group acting on {x1 , . . . , xn }. If
AnQ /G is rational over Q, then there exist infinitely many points P ∈ AnQ /G with Gal(Q(Q)/Q) ∼ =
−1
G for each Q ∈ π (P ).
Example 29.5.7. Let G = Sn be the symmetric group on n symbols. Let F = Q(x1 , . . . , xn ).
Then E = F Sn = Q(σ1 , . . . , σn ) where σi is the ith elementary symmetric polynomial and
the extension F/E has minimal polynomial

f (x, σ) = xn − σ1 xn−1 + . . . + (−1)n σn

for σ = (σ1 , . . . , σn ). This defines an Sn -cover AnQ → AnQ /Sn . By Hilbert’s irreducibility
theorem, f (x, s) = xn − s1 xn−1 + . . . + (−1)n sn is irreducible for infinitely many tuples
s = (s1 , . . . , sn ) ∈ Qn . It is a fact that

#{s = (s1 , . . . , sn ) ∈ Qn | 1 ≤ si ≤ N, f (x, s) is irreducible} = O(N n−1/2 log N ).

Using this, one can prove that this polynomial f (x, σ) is a generic polynomial for Sn -
extensions of Q.
Similar proofs using Hilbert’s irreducibility theorem can be used to verify generic poly-
nomials for other classes of G-extensions. A question related to Noether’s Problem and the
regular IGP is:
Question. For which groups G and fields K is AnK /G rational?
The answer to this was demonstrated to be no the general case: Lenstra provided coun-
terexamples for K = Q and G = C8 , and Saltman even showed a counterexample for K = C.
The question is even unknown for G = An , n > 5.
Here is a strategy to solving the regular Inverse Galois Problem for a given group G.
(1) If G has a generating set consisting of n − 1 elements, then G arises as a quotient of
the topological fundamental group π1 (P1C r {P1 , . . . , Pn }), since this is the free group on
n − 1 elements. Set π top (n) = π1 (P1C r {P1 , . . . , Pn }). In particular, suppose we can find
g1 , . . . , gn ∈ G such that G = hg1 , . . . , gn i, g1 · · · gn = 1 and create a map ϕ : π top (n) → G
with ϕ(γi ) = gi for 1 ≤ i ≤ n. We call (g1 , . . . , gn ) ∈ G a generating n-tuple for G.
1
(2) Set π(n) = π1 (PQ top (n). We may
r {P1 , . . . , Pn }) so that by Corollary 29.3.8, π(n) = π\
even pick P1 , . . . , Pn ∈ P1Q (Q).

(3) Also set Π(n) = π1 (P1Q r {P1 , . . . , Pn }) so that Proposition 29.4.1 gives us a short exact
sequence of profinite groups

1 → π(n) → Π(n) → GQ → 1.

Given ϕ : π(n) → G as in Step 1, we want to find a lift of the following diagram:

639
29.5. The Inverse Galois Problem Chapter 29. Fundamental Groups of Algebraic Curves

γi π(n) Π(n)

ϕ
ϕ
b

gi G

In particular, can we extend a continuous, surjective homomorphism ϕ : π(n) → G to a


b : Π(n) → G?
continuous, surjective ϕ

In general, consider a lifting problem of the form

N Γ

where N is a normal subgroup of a profinite group Γ and G is any finite group. The set
Homcts (N, G) comes equipped with two actions:

G × Homcts (N, G) −→ Homcts (N, G)


(g, ϕ) 7−→ gϕ(−)g −1
and Homcts (N, G) × Γ −→ Homcts (N, G)
(ϕ, σ) 7−→ (ϕσ : n 7→ ϕ(σnσ −1 ))

which are compatible in the sense that gϕσ = (gϕ)σ .

Lemma 29.5.8. Let S ⊆ Homcts (N, G) be a set of maps which are stable under the G- and
Γ-actions and such that G acts freely and transitively on S. Then any ϕ ∈ S extends to a
continuous homomorphism ϕ b : Γ → G.

Proof. For ϕ ∈ S and σ ∈ Γ, ϕσ ∈ S by stability, but ϕσ (n) = ϕ(σnσ −1 ) = gσ ϕ(n)gσ−1 for


some gσ ∈ G, since G acts transitively. Moreover, freeness of the G-action implies this gσ is
unique. Define ϕ(σ)
b = gσ . We claim that this is the desired extension. First, for σ, τ ∈ Γ,

ϕστ (n) = ϕ(στ nτ −1 σ −1 ) = ϕσ (τ nτ −1 ) = ϕσ ϕτ (n)

b is a group homomorphism. Next, for σ ∈ N ⊆ Γ,


so ϕ

ϕσ (n) = ϕ(σnσ −1 ) = ϕ(σ)ϕ(n)ϕ(σ)−1

since ϕ is a homomorphism on N , so by uniqueness we have gσ = ϕ(σ). Thus ϕ


b extends ϕ
as claimed.

640
29.5. The Inverse Galois Problem Chapter 29. Fundamental Groups of Algebraic Curves

Now to access Lemma 29.5.8, we want to construct such a set S ⊆ Homcts (π(n), G).
First, we need S to be stable under the action of π(n). Note that for ϕ ∈ S, having ϕσ ∈ S
for all σ ∈ π(n) is equivalent to having ϕσ (n) = ϕ(σ)ϕ(n)ϕ−1 = (ϕ(σ) · ϕ)(n), under the
left action of G on ϕ. Let (g1 , . . . , gn ) be a generating n-tuple for G. Then we must have
gi = ϕ(γi ) stable under the conjugacy action of G. In particular, we will specify conjugacy
classes C1 , . . . , Cn in G and consider sets

S = {ϕ ∈ Homcts (π(n), G) | ϕ(γi ) ∈ Ci , (ϕ(γ1 ), . . . , ϕ(γn )) is a gen. n-tuple}.

By construction, this S is stable under the π(n)- and G-actions. Now we impose further
conditions on the Ci to force
(i) transitivity;

(ii) free action;

(iii) Π(n)-stability.
Definition. Let G be a finite group. We say a set of conjugacy classes C1 , . . . , Cn in G is a
rigid system if there exists a generating n-tuple (g1 , . . . , gn ) ∈ Gn for G with gi ∈ Ci and
G acts transitively on the set of all such tuples.
To ensure (i), we pick a rigid system of conjugacy classes in G represented by the ϕ(γi ).
To guarantee (ii), it is enough to assume that Z(G) = 1 (G has trivial center). Next, set

Σ = {(g1 , . . . , gn ) ∈ Gn | g1 · · · gn = 1 and gi ∈ Ci }
Σ = {(g1 , . . . , gn ) ∈ Σ | hg1 , . . . , gn i = G}.

Definition. If Σ = Σ, we say the Ci are strictly rigid.


G acts on Σ and Σ, so supposing Z(G) = 1, it is clear that rigidity is equivalent to
|G| = |Σ|. One can even write down formulas relating |G| and |Σ| (see Serre).
Example 29.5.9. Consider G = Sn for n ≥ 3. For 1 ≤ k ≤ n, let kA be the conjugacy
class of k-cycles in Sn . Also, let C (k) be the conjugacy class of (1 2 · · · k)(k + 1 · · · n). In
particular, C (1) = (n − 1)A. We claim that the system of conjugacy classes {nA, 2A, C (1) }
is strictly rigid. By definition,

Σ = {(x, y, z) ∈ Sn3 | xyz = 1, x ∈ nA, y ∈ 2A, z ∈ C (1) }.

Thus determining the size of Σ comes down to determining when xy is an (n − 1)-cycle.


In general, one can see through elementary calculations that xy ∈ C (n−k) if y = (y1 y2 )
with y1 and y2 exactly k numbers apart (e.g. if y = (2 4) then k = 2). One then shows
that the conjugacy action of Sn is transitive on these 3-tuples, so we get |Σ| = |G|. But
now it is obvious that this system of conjugacy classes is strictly rigid, since by group
theory hnA, 2Ai = G. Hence {nA, 2A, C (1) } is a good choice of conjugacy classes. Even
{nA, 2A, C (k) } will work in most situations.
To ensure that the chosen S is stable under the action of Π(n), we make a final definition.

641
29.5. The Inverse Galois Problem Chapter 29. Fundamental Groups of Algebraic Curves

Definition. A conjugacy class C in G is said to be rational if for every g ∈ C, g m ∈ C for


every m ∈ Z coprime to |G|.

Lemma 29.5.10. If C1 , . . . , Cn are rational conjugacy classes of G and ϕ : π(n) → G is a


continuous homomorphism with ϕ(γi ) ∈ Ci for each 1 ≤ i ≤ n, then for all σ ∈ Π(n), we
have ϕσ (γi ) ∈ Ci .

Proof. Topologically, each γi generates an inertia group IQei in π(n) for some pro-point Qei →
ei ) is another pro-point over Pi and σγi σ −1 generates I e .
Pi . Also, for each σ ∈ Π(n), σ(Q σ(Qi )
1 1
But Pi ∈ P (Q) so Qi and σ(Qi ) are both pro-points above the same point P i ∈ P which
Q
e e
Q
1
lies over Pi . Thus IQei and Iσ(Qei ) are both stabilizers in π1 (PQ ) of points above some P i .
This implies these groups are conjugate in π(n). Write Iσ(Qei ) = αIQei α−1 for α ∈ π(n). Then
ϕ(Iσ(Qei ) ) = ϕ(α)ϕ(IQei )ϕ(α)−1 so ϕ(IQei ) and ϕ(Iσ(Qei ) ) are conjugate in G and cyclically
generated by ϕ(γi ) and ϕ(σγi σ −1 ), respectively. So for some g ∈ G and m prime to |G|, we
have ϕ(σγi σ −1 ) = gϕ(γi )m g −1 . Finally, since Ci is rational, this means ϕ(γi )m ∈ Ci and so
ϕ(σγi σ −1 ) ∈ Ci . Hence ϕ is Π(n)-stable.

Theorem 29.5.11. For a finite group G with Z(G) = 1, suppose there exists a rigid system
of rational conjugacy classes C1 , . . . , Cn in G. Then for any P1 , . . . , Pn ∈ P1Q (Q), there is a
continuous, surjective homomorphism

ϕ : π1 (P1Q r {P1 , . . . , Pn }) −→ G

with ϕ(γi ) ∈ Ci for 1 ≤ i ≤ n. In particular, G = Gal(F/Q(T )) for some regular Galois


extension F/Q(T ).

Example 29.5.12. For Sn , we saw that {nA, 2A, C (1) } is a strictly rigid system. Also, any
conjugacy class in Sn is rational since g 7→ g m preserves cycle type. In general, one can
prove this holds for {nA, 2A, C (k) }, 1 ≤ k ≤ n. This choice of rigid system corresponds to
the cover

P1 −→ P1
x 7−→ xk (x − 1)n−k

with t = 0 corresponding to C (k) , t = ∞ corresponding to nA and t = k n (k − n)n−k n−n


corresponding to 2A. Taking a splitting field of xk (x − 1)n−k − T determines an Sn -extension
of Q(T ), so the regular Inverse Galois Problem is solvable for G = Sn .

642
Chapter 30

Riemann’s Existence Theorem

643
30.1. Riemann Surfaces Chapter 30. Riemann’s Existence Theorem

30.1 Riemann Surfaces


Riemann surfaces are a mix of the topology of covering spaces and the complex analysis
of analytic continuation. The main problem one encounters in the latter setting is that a
holomorphic function does not always admit a uniquely defined analytic continatuion. The
normal strategy then is to employ ‘branch cuts’, but this tactic seems ad hoc and not suited
to generalization. Riemann’s idea was to replace the branches of a function with a covering
space on which the analytic continuation is an actual function.
Definition. Let X be a surface, i.e. a two-dimensional manifold. A complex atlas on X
is a choice of open covering {Ui } of X together with homeomorphisms ϕi : Ui → ϕi (Ui ) ⊆ C
such that for each pair of overlapping charts Ui , Uj , the transition map

ϕij := ϕj ◦ ϕ−1
i : ϕi (Ui ∩ Uj ) −→ ϕj (Ui ∩ Uj )

and its inverse are holomorphic. A complex structure on X is the choice of a complex
atlas, up to holomorphic equivalence of charts, defined by a similar condition to the above.
A connected surface which admits a complex structure is called a Riemann surface.
Example 30.1.1. The complex plane C is a trivial Riemann surface. Any connected open
subset U in C is also a Riemann surface via the given embedding U ,→ C.
Example 30.1.2. The complex projective line P1 = P1C = C ∪ {∞} admits a complex
structure defined by the open sets U0 = P1 r {∞} = C and U1 = P1 r {0} = C× ∪ {∞},
together with charts
1
ϕ0 : U0 → C, z 7→ z and ϕ1 : U1 → C, z 7→ ,
z
1
where ∞ = 0 by convention. Note that ϕ1 ◦ ϕ−1
0 is the function z 7→ 1
z
on C× which is
holomorphic.
Example 30.1.3. Let Λ ⊆ C be a lattice with basis [ω1 , ω2 ].

ω1

ω2

644
30.1. Riemann Surfaces Chapter 30. Riemann’s Existence Theorem

Then the quotient C/Λ admits a complex structure as follows. Let π : C → C/Λ be the
quotient map and suppose Π ⊆ C is a fundamental domain for Λ, meaning no two points in
Π are equivalent mod Λ. Set U = π(Π) ⊆ C/Λ. Then π|Π : Π → U is a homeomorphism,
so let ϕ : U → Π be its inverse. Letting {Ui } be the collection of all images under π of
fundamental domains for Λ, we get a complex atlas on C/Λ (one can easily check that the
transition functions between the Ui are locally constant, hence holomorphic). Topologically,
C/Λ is homeomorphic to a torus.

Definition. A function f : U → C on an open subset U of a Riemann surface X is holo-


morphic if for every complex chart ϕ : V → ϕ(V ) ⊆ C, the function f ◦ ϕ−1 : ϕ(U ∩ V ) →
U ∩ V → C is holomorphic.

Let O(U ) denote the set of all holomorphic functions U → C.

Lemma 30.1.4. For any open set U of a Riemann surface X, O(U ) is a commutative
C-algebra.

Proposition 30.1.5 (Holomorphic Continuation). For any open set U ⊆ X of a Riemann


surface and any x ∈ U , if f ∈ O(U r {x}) is bounded in a neighborhood of x, then f extends
uniquely to some f˜ ∈ O(U ).

More generally, we can define holomorphic maps between two Riemann surfaces.

Definition. A continuous map f : X → Y between Riemann surfaces is called holomorphic


if for every pair of charts ϕ : U → ϕ(U ) ⊆ C on X and ψ : V → ψ(V ) ⊆ C on Y such that
f (U ) ⊆ V , the map
ψ ◦ f ◦ ϕ−1 : ϕ(U ) → U → V → ψ(V )
is holomorphic. We say f is biholomorphic if it is a bijection and its inverse f −1 is also
holomorphic. In this case X and Y are said to be isomorphic as Riemann surfaces.
f g
Lemma 30.1.6. If X → − Y → − Z are holomorphic maps between Riemann surfaces, then
g ◦ f : X → Z is also holomorphic.

Proposition 30.1.7. Let f : X → Y be a holomorphic map. Then for all open U ⊆ X,


there is an induced C-algebra homomorphism

f ∗ : O(U ) −→ O(f −1 (U ))
ψ 7−→ f ∗ ψ := ψ ◦ f.

Proof. The fact that f ∗ ψ is an element of O(f −1 (U )) follows from the above definitions of
O and a holomorphic map between Riemann surfaces. The ring axioms are also easy to
verify.

Theorem 30.1.8. Suppose f, g : X → Y are holomorphic maps between Riemann surfaces


such that there exist a set A ⊆ X containing a limit point a ∈ A and f |A = g|A . Then f = g.

645
30.1. Riemann Surfaces Chapter 30. Riemann’s Existence Theorem

Proof. Let U ⊆ X be the set of all x ∈ X with an open neighborhood W on which f |W = g|W .
Then U is open and a ∈ U ; we will show it is also closed. If x ∈ ∂U , we have f (x) = g(x) since
f and g are continuous. Choose a neighborhood W ⊆ X of x and charts ϕ : W → ϕ(W ) ⊆ C
and ψ : W 0 → ψ(W 0 ) ⊆ C in Y with f (W ) ⊆ W 0 and g(W ) ⊆ W 0 . Consider

F = ψ ◦ f ◦ ϕ−1 : ϕ(W ) → ψ(W 0 ) and G = ψ ◦ g ◦ ϕ−1 : ϕ(W ) → ψ(W 0 ).

Then F and G are holomorphic and W ∩ U = 6 ∅, so we must have F = G. Therefore


f |W = g|W , so x ∈ U after all. This implies U = X.
Definition. A meromorphic function on an open set U ⊆ X consists of an open subset
V ⊆ U and a holomorphic function f : V → C such that U r V contains only isolated points,
called the poles of f , and limx→p |f (x)| = ∞ for every pole p ∈ U r V .
Denote the set of meromorphic functions on U by M(U ). Then M(U ) is a C-algebra,
where f + g and f g are defined by meromorphic continuation.
Example 30.1.9. Any polynomial f (z) = c0 + c1 z + . . . + cn z n is a holomorphic function
C → C. Viewing C ⊆ P1 , f is a meromorphic function on P1 with only a pole at ∞ of order
n (assuming cn 6= 0).
Example 30.1.10. Any meromorphic function f ∈ M(X) may be represented by a Laurent
series expansion about any of its poles p by choosing a complex chart U → C containing p,
lifting z to a parameter t on U and writing

X
f (t) = cn tn for some cn ∈ C.
n=−N

Theorem 30.1.11. Suppose X is a Riemann surface. Then the set of meromorphic func-
tions M(X) is in bijection with the set of holomorphic maps X → P1 .
Proof. If f ∈ M(X) is a meromorphic function, then setting f (p) = ∞ for every pole p of
f defines a holomorphic map f : X → P1 . Indeed, it is clear that f is continuous. Let P be
the set of its poles. If ϕ : U → ϕ(U ) ⊆ C is a chart on X and ψ : V → ψ(V ) ⊆ C is a chart
on P1 with f (U ) ⊆ V , then since f is holomorphic on X r P , ψ ◦ f ◦ ϕ−1 : ϕ(U ) → ψ(V ) is
holomorphic on ϕ(U ) r ϕ(P ). By Proposition 30.1.5, ψ ◦ f ◦ ϕ−1 is actually holomorphic on
ϕ(U ), so f is a holomorphic map of Riemann surfaces.
Conversely, if g : X → P1 is holomorphic, then by Theorem 30.1.8, either g(X) = {∞}
or g −1 (∞) is a set of isolated points in X. It is then easy to see that g : X r g −1 (∞) → C
is meromorphic.
Corollary 30.1.12 (Meromorphic Continuation). For any open set U ⊆ X and any x ∈ U ,
if f ∈ M(U r {x}) is bounded in a neighborhood of x, then f extends uniquely to some
f˜ ∈ M(U ).
Proof. Apply Proposition 30.1.5 and Theorem 30.1.11.
Corollary 30.1.13. Any nonzero function in M(X) has only isolated zeroes. In particular,
M(X) is a field.

646
30.1. Riemann Surfaces Chapter 30. Riemann’s Existence Theorem

Theorem 30.1.14. Let f : X → Y be a nonconstant holomorphic map between Riemann


surfaces. Then for every x ∈ X with y = f (x) ∈ Y , there exists k ∈ N and complex charts
ϕ : U → ϕ(U ) ⊆ C of X and ψ : V → ψ(V ) ⊆ C of Y with f (U ) ⊆ V such that
(1) x ∈ U with ϕ(x) = 0 and y ∈ V with ψ(y) = 0.
(2) F = ψ ◦ f ◦ ϕ−1 : ϕ(U ) → ψ(V ) is given by F (z) = z k for all z ∈ ϕ(U ).
Proof. It is easy to arrange (1) by replacing (U, ϕ) with another chart obtained by composing
ϕ with an automorphism of C taking ϕ(x) 7→ 0. So without loss of generality assume (1) is
satisfied. By Theorem 30.1.8, F = ψ ◦ f ◦ ϕ−1 is nonconstant. Thus since f (0) = 0, we may
write F (z) = z k g(z) for some k ≥ 1 and some g ∈ O(ϕ(U )) with g(0) 6= 0. Then g(z) = h(z)k
for some holomorphic function h on ϕ(U ), and H(z) = zh(z) defines a biholomorphic map
α of some open neighborhood W ⊆ ϕ(U ) of 0 onto another open neighborhood of 0. Finally,
replace (U, ϕ) by (ϕ−1 (W ), α ◦ ϕ). By construction, F = ψ ◦ f ϕ−1 is now of the form
F (z) = z k .
Definition. The integer k for which F can be written F (z) = z k about x ∈ X is called the
multiplicity of f at x.
Corollary 30.1.15. If f : X → Y is a nonconstant holomorphic map between Riemann
surfaces, then f takes open sets to open sets.
Corollary 30.1.16. If f : X → Y is an injective holomorphic map, then f is biholomorphic
X → f (X).
Proof. If f is injective, then locally F (z) = z k with k = 1. Hence f −1 is holomorphic.
Theorem 30.1.17. If f : X → Y is a nonconstant holomorphic map and X is compact,
then Y is also compact and f is surjective.
Proof. By Corollary 30.1.15, f (X) is open but since X is compact, f (X) is also compact
and in particular closed. Therefore f (X) = Y .
Corollary 30.1.18. Every holomorphic function on a compact Riemann surface is constant.
Proof. C is not compact, so Theorem 30.1.17 implies that every holomorphic function from
a compact space into C must be constant.
Corollary 30.1.19. Every meromorphic function on P1 is rational.
Proof. First, note that the only way for such an f ∈ M(P1 ) to have infinitely many poles
is if it had a limit point, but then Theorem 30.1.8 would imply f ≡ ∞. Thus f has finitely
many poles, say a1 , . . . , an ∈ P1 ; we may assume ∞ is not one of the poles, or else consider
the function f1 instead. For 1 ≤ i ≤ n, expand f as a Laurent series about ai :
mi
X
fi (z) = cij (z − ai )−j for cij ∈ C.
j=1

Then g = f − (f1 + . . . + fn ) is holomorphic on P1 and thus constant by Corollary 30.1.18


since P1 is compact. This shows f is rational.

647
30.1. Riemann Surfaces Chapter 30. Riemann’s Existence Theorem

Corollary 30.1.20 (Liouville’s Theorem). Every bounded holomorphic function on C is


constant.
Proof. By Proposition 30.1.5, f has a holomorphic continuation to f˜ : P1 → C, but by
Corollary 30.1.18, f˜ must be constant.
The idea in the rest of the section is to relate holomorphic maps between Riemann
surfaces to covering space theory.
Theorem 30.1.21. If p : Y → X is a nonconstant holomorphic map between Riemann
surfaces then p is open and has discrete fibres.
Proof. By Corollary 30.1.15, p is open and Theorem 30.1.8 implies each fibre is discrete.
Let p : Y → X be a cover of Riemann surfaces. Traditionally, holomorphic functions
f : Y → C are treated as multi-valued functions on X by setting f (x) = {f (y1 ), . . . , f (yn )}
where p−1 (x) = {y1 , . . . , yn }.
Example 30.1.22. Let exp : C → C× be the exponential map z 7→ ez and f = id : C →
C the identity map. Then the resulting multi-valued function C× → C is the complex
logarithm, which is only defined as a function after making a particular choice of branch of
the function. We can describe this idea more cleanly with Riemann surfaces and branched
covers.
Definition. Suppose p : Y → X is a nonconstant holomorphic map. A ramification point
of p is a point y ∈ Y such that for every neighborhood V ⊆ Y of y, p|V : V → p(V ) is not
injective. The image x = p(y) is called a branch point of p. If p has no ramification points
(and hence no branch points), then we call p an unramified map.
Theorem 30.1.23. A nonconstant holomorphic map p : Y → X is unramified if and only
if it is a local homeomorphism.
Proof. Suppose p is unramified. Then for any y ∈ Y , there exists a neighborhood V ⊆ Y of
y such that p|V : V → p(V ) is injective and open. Therefore p|V is a homeomorphism onto
p(V ). The converse follows from basically the same argument.
Example 30.1.24. For each n ≥ 2, the map pn : C → C defined by pn (z) = z n is ramified
at 0 ∈ C and unramified everywhere else. Therefore pn : C× → C is an unramified cover.
Moreover, Theorem 30.1.14 says that every ramified cover of Riemann surfaces Y → X is
locally of the form C → C, z 7→ z n .
Example 30.1.25. The exponential map exp : C → C× , z 7→ ez is an unramified cover. In
fact, as in the topological case, exp gives a universal cover of C via the inverse system of the
covers pn .
Example 30.1.26. The quotient map π : C → C/Λ from Example 30.1.3 is an unramified
cover of Riemann surfaces.
Theorem 30.1.27. Suppose p : Y → X is a local homeomorphism of Hausdorff topological
spaces and X is a Riemann surface. Then Y admits a unique complex structure making p a
holomorphic map.

648
30.1. Riemann Surfaces Chapter 30. Riemann’s Existence Theorem

Proof. Let ϕ : V → C be a chart of X. Then there exists an open subset U ⊆ V over which
p|U : p−1 (U ) → U is a homeomorphism. Set U e = p−1 (U ) and ϕ
e = ϕ ◦ p|U : Ue → C. Then
e is a complex chart on Y and the collection {U
ϕ e , ϕ}
e obtained in this way forms a complex
atlas on Y . Since p : Y → X is locally biholomorphic by construction, it is a holomorphic
map between Riemann surfaces. Uniqueness is easy to check.
Example 30.1.28. Now that we can view nonconstant holomorphic maps as local homeo-
morphisms, and in most cases covering spaces, we can rephrase the language of branch cuts
as a lifting problem. For example, let exp : C → C× be the exponential map and suppose
f : X → C× is a holomorphic map of Riemann surfaces, with X simply connected. Then by
covering space theory, for each fixed x0 ∈ X and z0 ∈ C such that f (x0 ) = ez0 , there exists
a unique lift F : X → C making the diagram
C
F exp

X C×
f

commute. Theorem 30.1.27 can be used to show that any such F is holomorphic. Moreover,
any other lift G of f differs from F by 2πin for some n ∈ Z. For the special case of a simply
connected open set X ⊆ C× , any lift F is a branch of the complex logarithm on X.
Example 30.1.29. Similarly, one can construct the complex root functions z 7→ z 1/n , n ≥ 2,
as lifts along the cover pn : C× → C.
Let f : Y → X be a nonconstant holomorphic map that is proper, i.e. the preimage of
any compact set in X is compact in Y . For each x ∈ X, define the multiplicity of f at x to
be X
ordx (f ) = vy (f )
y∈f −1 (x)

where vy (f ) is the multiplicity of f at y.


Example 30.1.30. If f : Y → X is unbranched at x ∈ X, then p−1 (x) = {y1 , . . . , yn } for
some n and vyi (f ) = 1 for each 1 ≤ i ≤ n. Thus ordx (f ) = n.
Theorem 30.1.31. If f : Y → X is a proper, nonconstant holomorphic map between
Riemann surfaces, then there exists a number n ∈ N such that for every x ∈ X, ordx (f ) = n.
Proof. By Theorem 30.1.21, the set B of ramification points of f is a closed, discrete subset
of Y . Let A = f (B) ⊆ X. Then since f is proper, A is also closed and discrete. The
restriction f |Y rB : Y r B → X r A is unramified, so it is a finite-sheeted covering space; say
n is the number of sheets of f |Y rB , i.e. the size of any fibre f −1 (x) for an unbranched point
x ∈ X. By the above example, f has multiplicity n at every y ∈ Y r B. Suppose a ∈ A
and write f −1 (a) = {b1 , . . . , bk } ⊆ B and mi = vbi (f ). For each 1 ≤ i ≤ k, we may choose
neighborhoods Vi ⊂ Y of bi and Ui ⊂ X of a such that for all x ∈ Ui r {a}, f −1 (x) ∩ Vi
consists of exactly mi points. Then there is a neighborhood U ⊆ U1 ∩ · · · ∩ Uk of a such that
f −1 (U ) ⊆ V1 ∪· · ·∪Uk and for every x ∈ U ∩(X rA), f −1 (x) consists of exactly m1 +. . .+mk
points. However we showed that |f −1 (x)| = n, so n = m1 + . . . + mk as required.

649
30.1. Riemann Surfaces Chapter 30. Riemann’s Existence Theorem

Corollary 30.1.32. Let X be a compact Riemann surface and f : X → C a nonconstant


meromorphic function. Then the number of zeroes of f equals the number of poles of f ,
counted with multiplicity.

Proof. View f as a holomorphic function X → P1 . Since X and P1 are compact, f is a


proper map so ord0 (f ) = ord∞ (f ). But ord0 (f ) is precisely the number of zeroes of f , while
ord∞ (f ) is the number of poles.

Corollary 30.1.33. Any complex polynomial f (z) ∈ C[z] of degree n has exactly n zeroes,
counted with multiplicity.

Proof. We may view f as a holomorphic map P1 → P1 . Then it is easy to see ord∞ (f ) = n,


so once again ord0 (f ) = n.

650
30.2. The Existence Theorem over P1C Chapter 30. Riemann’s Existence Theorem

30.2 The Existence Theorem over P1C


Consider the complex projective line P1C = C ∪ {∞}, also known as the Riemann sphere,
and take a finite set of points P ⊂ P1C . Then P1C r P is both a topological space and an
algebraic curve. In the topological case, the first invariant that describes the projective
line is Γ := π1top (P1 r P ), while in the algebraic case we now have an algebraic invariant
π1 (P1 r P ) = π1ét (P1 r P ). We will prove the following case of Theorem 29.3.5:
Theorem 30.2.1 (Riemann Existence). For any finite set P ⊂ P1 , there is an isomorphism
π1ét (P1 r P ) ∼
= Γ,
b where Γ
b is the profinite completion of the topological fundamental group
top
Γ = π1 (P1 r P ).
Set K = C(P1 ) = C(t) be the function field of the projective line and let L(P )/K be the
maximal Galois extension of K that is unramified outside P . Then by the definition of the
algebraic fundamental group in Section 29.3, an equivalent statement to Theorem 30.2.1 is
that Gal(L(P )/K) ∼ = Γ. b This statement is quite useful for applications to the Inverse Galois
Problem over K = C(t), as we saw in Section 29.5.
Recall that each point Pi ∈ P corresponds to a generator in the fundamental group
Γ = hγ1 , . . . , γr | γ1 · · · γr = 1i, for a fixed basepoint x ∈ P1 :

γ1 γ2 γr

P1 P2 ··· Pr

Specify a generating r-tuple g1 , . . . , gr ∈ G – recall from Section 29.5 that this means
hg1 , . . . , gr i = G and g1 · · · gr = 1. Then G is obtained as a quotient of Γ by mapping
γi 7→ gi for each 1 ≤ i ≤ r. We have seen that the inertia groups at the points Pi ∈ P are
cyclicly generated, IPi = hgi i. The following lemma allows us to dispense with worries about
“canonical” generators γ1 , . . . , γr for Γ.
Lemma 30.2.2. If Γ1 and Γ2 are two finitely generated groups with the same homomorphic
images, then their profinite completions Γ
b1 and Γ
b2 are isomorphic.

On the other hand, with K = C(t), we have that Gal(K/K) = Fbω is free profinite of
(uncountably) infinite rank. By definition Gal(K/K) = lim Gal(L(P )/K) where the inverse
←−
limit is taken over all finite sets P ⊂ P1 , ordered by P ⊆ P 0 which induces L(P ) ⊆ L(P 0 )
and thus Gal(L(P 0 )/K) → Gal(L(P )/K). In particular, for each containment of finite sets
P ⊆ P 0 we have a diagram

651
30.2. The Existence Theorem over P1C Chapter 30. Riemann’s Existence Theorem

Gal(L(P 0 )/K) Gal(L(P )/K)


= ∼
=

Fb|P 0 |−1 Fb|P |−1

The topological side of the proof of Theorem 30.2.1 is essentially contained in the following
theorem.
Theorem 30.2.3. Let P ⊂ P1C be a finite set and π : R → P1C r P a finite-sheeted Galois
covering of topological spaces. Then there exists a smooth projective curve W with a regular
map f : W → P1C which is unramified outside P and such that the covering

W r f −1 (P ) → P1C r P

is equivalent to π. Furthermore, under this identification, the deck transformations of π


become regular automorphisms of W → P1C .
The proof requires a sequence of reduction steps, given by the following lemmas.
π
Lemma 30.2.4. For a topological cover R →− P1 r P , R may be given the structure of a
compact Riemann surface such that π becomes analytic.
π
− P1 r P , we first give R the structure of a Riemann surface. This can
Proof. Given R →
be done using Theorem 30.1.27, but the following explicit description will be of use in later
arguments.

Ui
q R

P2 B(z, ε) P3
P1 z P1

Take q ∈ R r π −1 (P ) and set z = π(q). Since P is finite, we can choose a ball B(z, ε) which
does not contain any point of P . Then π −1 (B(z, ε)) = U1 q · · · q Un for disjoint open sets Ui
for which π|Ui : Ui → B(z, ε) is a homeomorphism. One of these Ui contains q; let this Ui be
the coordinate chart defining the surface structure on R. For two of these charts, U = Ui and
V = Vj , the transition maps are just given by lifting the transition maps B(z, ε) → B(z 0 , ε0 )

652
30.2. The Existence Theorem over P1C Chapter 30. Riemann’s Existence Theorem

along the local homeomorphism π, and these are analytic on the disks B(z, ε), B(z 0 , ε0 ) so
they are analytic on the U, V . This gives R the structure of a Riemann surface, but it may
not be compact.
In the next step, we show that there exists a compact Riemann surface R with an analytic
map π̄ : R → P1 and an open embedding R ,→ R such that the diagram
R R

π π̄

P1 r P P1

commutes. For any ball B = B(z, ε) ⊆ P1 , let B 0 = B 0 (z, ε) = B r {z} be the disk
punctured at its center. Then π1top (B 0 ) ∼
= Z and in particular B 0 has (up to isomorphism)
only one connected, degree d cover for each d ≥ 1. Call this cover µd . When B = B(0, 1) is
the unit ball, this degree d cover is explicitly given by
µd : B 0 −→ B 0
z 7−→ z d .
Moreover, µd extends to a branched cover of the entire disk B:
µd
B0 B0

µd
B B

The deck transformations of this cover are also easy to write down: they are generated by
τ : z 7→ ζz where ζ = ζd is a primitive dth root of unity.
Let z0 ∈ P and take a disk D = B(z0 , ε) around z0 which does not contain any other
points in P . Let D0 = D r {z0 }. Then as above, the preimage of D0 along π can be written
as a disjoint union of connected components,
π −1 (D0 ) = D10 q · · · q Dm
0

where m ≤ n, such that for each 1 ≤ i ≤ m, π|Di0 : Di0 → D0 is a covering map. We now
construct a Riemann surface R1 together with an analytic map π1 : R1 → P1 r (P r {z0 })
and an embedding R ,→ R1 such that |R1 r R| = m and the diagram
R R1

π π1

P1 r P P1 r (P r {z0 })

653
30.2. The Existence Theorem over P1C Chapter 30. Riemann’s Existence Theorem

commutes. After scaling by some λ ∈ C, we may replace D0 with B 0 = B(0, 1) r {0}. Then
there exists an integer di and a homeomorphism τi : Di → B 0 making the following diagram
commute:
τi
Di0 B0

π µdi

λ
D0 B0

It is an easy exercise to show that τi is in fact biholomorphic. Notice that while τi is not
unique, it is defined up to a deck transformation of µdi . We now construct R1 by taking m
0
copies of B, say B1 , . . . , Bm , and the corresponding punctured disks B10 , . . . , Bm and setting
T = B1 q · · · q Bm . Obtain R1 by gluing R to T along the homeomorphisms

D10 q · · · q Dm
0
−→ B10 q · · · q Bm
0
.

Repeating this process, we obtain R after a finite number of steps, i.e. whenever all points
of π −1 (P ) have been “filled in”.

Lemma 30.2.5. If π : R → P1 r P is a Galois covering, then π̄ : R → P1 is Galois as well.

Proof. Suppose σ is a deck transformation of π : R → P1 r P . We claim σ extends uniquely


0
to a biholomorphic automorphism of π̄ : R → P1 . Let D0 , D10 , . . . , Dm be as above. On the
disjoint union π (D ) = D1 q · · · q Dm , σ just acts by permuting the Di0 . For one of the
−1 0 0 0

Di0 , set σ(Di0 ) = Dj0 . Then the diagram


σ
Di0 Dj0

σ
Bi0 Bj0

commutes but by Proposition 30.1.5, since σ has bounded image it extends to Bi → Bj and
hence to σ̄ : Di → Dj . Now for any z0 ∈ P , let D0 = D r {z0 }. Then each Di0 in the disjoint
union π −1 (D0 ) = D10 q · · · q Dm
0
corresponds to a point qi ∈ R such that π(qi ) = z0 , and
which we glue to R to obtain R. Moreover, for each qi , qj , we may choose points xi ∈ Di0 and
xj ∈ Dj0 such that xi 6= qi and xj 6= qj . Now π is Galois, so there exists a deck transformation
σ ∈ ∆(π) mapping σ(xi ) = xj . Then we must have σ(Di0 ) = Dj0 and hence by the above, the
extension σ̄ maps σ̄(qi ) = qj . This proves π̄ is a Galois covering.
Suppose π : R → P1 r P is a Galois cover and set G = ∆(π) = Aut(π), the group of deck
transformations. For any analytic curve Y , let M(Y ) be the field of meromorphic functions
on Y . For Y = R as above, we have an action of G on M(R) by (σf )(x) = f (σ −1 (x)).

654
30.2. The Existence Theorem over P1C Chapter 30. Riemann’s Existence Theorem

Lemma 30.2.6. For any f ∈ M(R) which is invariant under the G-action described above,
then f ∈ M(P1 ) ⊆ M(R). In other words, f = g ◦ π for some g ∈ M(P1 ).

Proof. By the construction in Lemma 30.2.4, there exists a function g : P1 → P1 such that
f = g ◦ π. Thus we need only show that g is meromorphic. Take p ∈ R and consider
q = π(p) ∈ P1 . If q 6∈ P , the set of branch points of π, then locally f is given by f = g ◦ π :
D(p) → D(q) on local trivializations D(p) ⊆ R and D(q) ⊆ P1 of p, q, respectively. But
locally, π is biholomorphic on these neighborhoods so we can invert π and write g = f ◦ π −1
and it follows that g is meromorphic.
If q ∈ P , look at the punctured neighborhoods D(p) = D(p) r {0}, D(q) = D(q) r {0}
and the cover π : D(p) → D(q). This is not holomorphic, but we know that it’s a degree
m map that in fact just corresponds to the multiplication-by-m map on the unit punctured
disk D(0, 1):
π
D(p) D(q)


= ∼
=
[m]
D D
P∞
Write f as a Laurent series f (z) = j=−N aj z j for aj ∈ C and N ≥ 0. Then we must have
f (ζm z) = f (z) for any root of unity ζm . Thus aj is nonzero only when m | j and we can
rewrite ∞
X
f (z) = akm z km .
k=−K

Let w = z m . Then w is a local parameter on D and we have



X
f (z) = akm wk = g(w).
k=−K

Thus g(w) is meromorphic as required.


In the next step, we need the so-called analytic version of the existence theorem. A proof
will be provided in Section 30.4.

Theorem 30.2.7 (Riemann Existence – Analytic Version). Let X be a compact Riemann


surface. Then for any distinct points p1 , . . . , pn ∈ X and any numbers c1 , . . . , cn ∈ C, possibly
not distinct, there exists a meromorphic function g ∈ M(X) such that g(pi ) = ci for each
1 ≤ i ≤ n.

Lemma 30.2.6 shows that there exists a finite group G acting on M(R) such that
M(R)G = M(P1 ). By Serre’s GAGA principle, M(P1 ) = C(t), so therefore we have proven
that [M(R) : C(t)] < ∞.

Theorem 30.2.8. M(R)/C(t) is a Galois extension with Galois group G = ∆(π).

655
30.2. The Existence Theorem over P1C Chapter 30. Riemann’s Existence Theorem

Proof. The paragraph above implies, by Artin’s Lemma, that M(R)/C(t) is Galois. Further,
Gal(M(R)/C(t)) = G/N where N is the kernel of the action of G on M(R). To prove this,
pick a point p ∈ P1 and write π −1 (p) = {p1 , . . . , pn }. By the analytic version of Riemann’s
existence theorem (30.2.7), there exists a meromorphic g ∈ M(R) such that the values
g(p1 ), . . . , g(pn ) are all distinct. Suppose σ ∈ N is nontrivial. Then σ(pi ) = pj 6= pi for some
i 6= j, but σ ∈ N implies σ(g) = g, or g ◦ σ = g. Thus g(pi ) = g(σ(pi )) = g(pj ), which is
impossible. Therefore N = {1} so Gal(M(R)/C(t)) = G as claimed.
The final step in the proof of Theorem 30.2.3 is contained the following lemma.

Lemma 30.2.9. There exists a smooth projective curve W over C and a biholomorphic
isomorphism R ∼
= W (C).
Proof. We know M(R) is a transcendence degree 1 extension of C, so by Corollary 29.1.9
there exists a proper normal curve W with C(W ) = M(R). Abstractly, there is a map
ϕ : R → W which sends a point r ∈ R to the associated valuation vr on the field M(R),
which is now identified with a valuation v on C(W ) that is trivial on C, and hence determines
a point of W .
We may assume W is smooth and projective over C. Let W ,→ PN for large enough N .
Write PN = {[x0 , x1 , . . . , xN ] : xi ∈ C, some xi 6= 0}. Then for any x ∈ R, there is an affine
patch on which the map ϕ : R → W (C) is defined, so on this patch, ϕ = (f1 , . . . , fN ) where
fi = xx0i . But by properness, this must in fact be holomorphic. Moreover, Theorem 30.2.7
implies ϕ is also injective. Further, properness also implies ϕ(R) ⊆ W is closed. Therefore
to prove ϕ is a biholomorphic isomorphism, it suffices to show W (C) is connected.
This is a rather standard property, but we repeat the proof here. Given that W is
irreducible, smooth and projective over C, let us assume W (C) = M1 ∪ M2 for some open
sets M1 , M2 ⊂ W . Pick x0 ∈ M1 . Then by the Riemann-Roch theorem, there exists a
rational function f ∈ C(W ) that has a pole at x0 and is holomorphic everywhere else. View
f : M2 → P1 ; this cannot be constant since M2 is infinite and in particular dense in W (C).
On the other hand, f (M2 ) ⊆ P1 is open and W (C) is compact, so M2 is compact and
therefore f (M2 ) is also closed. Thus f (M2 ) = P1 so some point m ∈ M2 has f (m) = ∞,
contradicting the fact that f was holomorphic on M2 . Hence W (C) is connected, so we are
finished.
We derive the following important consequences.

Corollary 30.2.10. Let K = C(t) and E = C(W ). Then the embedding f ∗ : K ,→ E


satisfies
[E : K] = deg π = |∆(π)|
where ∆(π) is the group of deck transformations of π.

Corollary 30.2.11. E/K is Galois and | Gal(E/K)| = |∆(π)|.

Proof. By Theorem 30.2.3, ∆(π) is a subgroup of Aut(E/K) but in general we have | Aut(E/K)| ≤
[E : K]. Therefore Corollary 30.2.10 implies Aut(E/K) = ∆(π) and therefore the extension
is Galois.

656
30.2. The Existence Theorem over P1C Chapter 30. Riemann’s Existence Theorem

Corollary 30.2.12. There is an anti-equivalence of categories

{compact Riemann surfaces} −→ {finite étale C(t)-algebras}


Y 7−→ M(Y ).

Corollary 30.2.13. Let K = C(t). Then there is an equivalence of categories



{compact Riemann surfaces} −
→ {finite, continuous left Gal(K/K)-sets}.

Proof. Combine Corollaries 30.2.12 and 28.2.4.

657
30.3. The General Case Chapter 30. Riemann’s Existence Theorem

30.3 The General Case


Let X and Y be Riemann surfaces and let M(X) and M(Y ) be their corresponding fields
of meromorphic functions, as in the previous section. Suppose ϕ : Y → X is a holomorphic
map which is not constant on any connected component. This induces a field homomorphism
ϕ∗ : M(X) → M(Y ) defined by f 7→ ϕ∗ (f ) = f ◦ ϕ.
Proposition 30.3.1. If ϕ : Y → X is a nonconstant holomorphic map of degree d between
compact, connected Riemann surfaces, then M(Y )/ϕ∗ M(X) is a field extension of degree d.
Proof. Let f ∈ M(Y ) be a meromorphic function. We first prove f satisfies a polynomial
equation of degree d over M(X). Let S be the set of branch points of the cover. Then
for each x 6∈ ϕ(S), there is a neighborhood U ⊆ X of x such that ϕ−1 (U ) = di=1 Vi for
`
V1 , . . . , Vd ⊆ Y homeomorphic to U via ψi : U → Vi . Set fi = f ◦ ψi , which is meromorphic
on U . Define
d
Y
F (t) = (t − fi (t)) = td + an−1 td−1 + . . . + a0
i=1

for ai meromorphic on U . One shows further that ai extend meromorphically to all of X, so


F (t) is defined over M(X). Lastly, (ϕ∗ F )(f ) = 0 since it restricts to F (fi ) = 0 on each Vi .
This proves the claim.
Now we show that some f ∈ M(Y ) satisfies an irreducible polynomial of degree d over
M(X). If x 6∈ ϕ(S), then ϕ−1 (x) = {y1 , . . . , yd }. By the analytic existence theorem (30.2.7),
choose f ∈ M(Y ) which is holomorphic at y1 , . . . , yd and has f (yi ) all distinct. Then by the
first paragraph, there exist ai ∈ M(X) such that

(ϕ∗ an )f n + . . . + ϕ∗ a0 = 0.

(Here, n ≤ d.) Suppose one of the ai has a pole at x. Then we may choose a small enough
neighborhood about x not containing any branch points such that f is holomorphic and
assigns distinct values to all preimages of each point. Replacing x with any of the points of
this neighborhood, we may assume each ai is holomorphic at x. Finally, take g ∈ M(Y ).
Then by the primitive element theorem, M(X)(f, g) = M(X)(h) for some h ∈ M(Y ). But
then M(X)(f ) ⊆ M(X)(h) and by the first paragraph, h has degree ≤ d over M(X), so
M(X)(f ) = M(X)(h). In particular, M(Y ) = M(X)(f ) so the proof is complete.
We now generalize the results from Section 30.2 to arbitrary covers of Riemann surfaces.
Lemma 30.3.2. Suppose X is a Riemann surface, S ⊂ X a discrete set of points and
ϕ : Y → X r S a proper holomorphic cover of Riemann surfaces. Then there is a Riemann
surface Y and proper holomorphic map ϕ̄ : Y → X such that Y r ϕ−1 (S) = Y .
Proof. This is only a slight generalization of the proof of Lemma 30.2.4. For details, see
Forster.
Lemma 30.3.3. If ϕ : Y → X r S is a Galois covering, then ϕ̄ : Y → X is Galois as well.
Proof. A generalization of Lemma 30.2.5.

658
30.3. The General Case Chapter 30. Riemann’s Existence Theorem

Theorem 30.3.4. Let X be a connected, compact Riemann surface with field of meromorphic
functions M(X). Then for every finite étale M(X)-algebra A, there is a compact Riemann
surface Y with M(Y ) ∼
= A and a holomorphic map Y → X.
Corollary 30.3.5. For any connected, compact Riemann surface X, there is an anti-equivalence
of categories
 
proper holomorphic maps of ∼

→ {finite étale M(X)-algebras}.
Riemann surfaces Y → X

Corollary 30.3.6. Let X be a connected, compact Riemann surface. Then there is an


equivalence of categories
 
compact Riemann surfaces ∼

→ {finite, continuous left Gal(M(X)/M(X)-sets}.
Y →X

We can now give a proof of Theorem 29.3.5.


Proof. Let K = k(X) be the function field of X and KU the compositum in Ks of all
finite extensions L/K coming from proper normal curves Y → X étale over U . Then
by Theorem 30.3.4, for each of these L/K there is a compact connected Riemann surface
YL → X étale over U such that M(Y ) ∼ = L. Let Mf be the compositum of these M(Y ).
Then π1 (U ) = Gal(KU /K) ∼ = Gal(M f/M(X)) so it suffices to show Gal(M f/M(X)) is the
top
profinite completion of G := π1 (U (C)). But by Corollary 28.1.5, every finite quotient of G b
0
uniquely determines a finite Galois cover Y → U ; thus a covering Y → X by Lemma 30.3.2
which is Galois by Lemma 30.3.3; thus a finite Galois field extension M(Y )/M(X) by
Proposition 30.3.1. It is clear that this correspondence is functorial on both sides, so taking
inverse limits, we get the desired identification G
b = Gal(Mf/M(X)).

659
30.4. The Analytic Existence Theorem Chapter 30. Riemann’s Existence Theorem

30.4 The Analytic Existence Theorem


In this section we give a proof of Theorem 30.2.7. First, recall the following results for sheaf
cohomology.
Theorem 30.4.1 (Leray). Let X be a space with a sheaf F and an open cover U. If the
cover is acyclic, i.e. Ȟ 1 (U, F) = 0, then H 1 (X, F) = Ȟ 1 (U, F).
For a cover U = {Ui }i∈I , we can view Ui as a ball B(ai , εi ) – using the biholomorphic
map z. Writing
B(ai , εi ) = ∪∞ 1

n=k B ai , εi − n ,

compactness of X implies there is a finite subcover of each B(ai , εi ) and therefore each Ui .
Taking a finite subcover consisting of such sets, we obtain a cover W = {Wj }N j=1 of X such
that each W j is compactly contained in some Ui . We say the cover W is compactly contained
in U.
Proposition 30.4.2. If W is compactly contained in U then the restriction map Ȟ 1 (U, O) →
Ȟ 1 (W, O) has finite-dimensional image.
Proof. Consider the sheaf E of smooth functions on X. It is a fact that H 1 (B, E) = 0 for
any disk B ⊆ C – in technical terms, E is a flabby sheaf. Given a cocycle (fij ) ∈ Z 1 (B, O),
we can resolve it using smooth functions: fij = gj − gi for gi , gj ∈ C 0 (B, E). Now these may
not be analytic, but this won’t be a problem. Consider the Cauchy-Riemann operator
 
∂ 1 ∂f ∂f
: f 7→ +i .
∂ z̄ 2 ∂x ∂y

By the Cauchy-Riemann equations, ∂ z̄
is 0 on analytic functions, so in particular
∂fij ∂gj ∂gi
0= = − .
∂ z̄ ∂ z̄ ∂ z̄
Now the functions ∂g

i
∂ z̄
agree on any Ui ∩ Uj , so we get a lift h ∈ Ž 1 (B, E). By Dolbeault’s
Lemma from complex analysis, there is some g ∈ E(B) such that ∂g ∂ z̄
= h. By the above,
∂(gj −g
∂ z̄
= 0 for all j. Writing fij = (gj − g) − (gi − g), we now see that (fij ) is a coboundary.
Hence Ȟ 1 (B, O) = 0.
To extend this to covers, suppose W is compactly contained in U. Take U ∈ U and
W ∈ W with W ⊆ U . Then for any f ∈ O(U ), f may be unbounded but by compactness,
f |W is bounded. Let || · || be the L2 -norm, so that
ZZ
||f || = |f (x + iy)|2 dx dy.
U

This may be infinite, but define


L2 (U, O) = {f ∈ O(U ) : ||f || < ∞}.
As spaces, L2 (U, O) ⊆ L2 (U ), the ordinary Hilbert space of square-integrable complex func-
tions on U . Thus the above says that O(U ) → O(W ) has image lying in L2 (W, O). We
need:

660
30.4. The Analytic Existence Theorem Chapter 30. Riemann’s Existence Theorem

Lemma 30.4.3. If W 0 is compactly contained in U 0 ⊆ C, then for every ε > 0 there exists
a closed subspace A ⊆ L2 (U, OC ) of finite codimension such that for all f ∈ A,

||f ||L2 (W 0 ) ≤ ε||f ||L2 (U 0 ) .

We next construct a norm on the Čech cochain groups of X. First, each Ui ∈ U is biholo-
morphic to some disk Di ⊆ C via a map zi . For g ∈ O(Ui ), set ||g||L2 (Ui ) := ||zi (g)||L2 (Di ) .
For a 0-cocycle η = (fi ) ∈ Č 0 (U, O), where fi ∈ O(Ui ), define

n
!1/2
X
||η|| := ||fi ||2L2 (Ui ) .
i=1

Likewise, for ξ = (fij ) ∈ Č 1 (U, O), define


!1/2
X
||ξ|| := ||fij ||2L2 (Ui ∩Uj ) .
i,j

Continue in this fashion, we get a norm for every element of Č q (U, O), q ≥ 0. Write
ČLq 2 (U, O) for the subspace of Č q (U, O) where this norm is finite. Now for the compactly con-
tained covers W ⊆ U and for any ε > 0, Lemma 30.4.3 provides a closed, finite-codimension
subspace A ⊆ ŽL1 2 (U, O) such that for all ξ ∈ A, ||ξ||L2 (W) ≤ ε||ξ||L2 (U ) .
On the other hand, suppose we have a containment of covers W ⊆ V ⊆ U, with compact
containments Wi ⊆ Vi ⊆ Ui for each i.
Lemma 30.4.4. Given compactly contained covers W ⊆ V ⊆ U, there exists c > 0 such that
for every cocycle ξ ∈ ŽL1 2 (V, O), there is some ζ ∈ ŽL1 2 (U, O) such that

ζ = ξ + d(0) η

for some η ∈ ČL0 2 (W, O) with ||ζ||, ||η|| ≤ c||ξ||.


Now we play Lemmas 30.4.3 and 30.4.4 against each other. Use Lemma 30.4.4 to “extend”
1
the norm to U subject to the bound c > 0. Let ε = 2c and apply Lemma 30.4.3 to produce
a closed subspace A ⊆ ŽL1 2 (U, O) of finite codimension for which ||ξ||L2 (V) ≤ ε||ξ||L2 (U ) for
all ξ ∈ A. Then S = A⊥ is finite dimensional, where ⊥ is taken with respect to the L2 -norm
on Ž 1 (U, O).
Consider the restriction map Ȟ 1 (U, O) → Ȟ 1 (W, O). We claim the image is equal to
the image of S. Take ξ ∈ Ž 1 (U, O) and restrict to a bounded cocycle ξ|V ∈ ŽL1 2 (V, O). Then
extend to U with Lemma 30.4.4 to get ζ0 ∈ ŽL1 2 (U, O) such that ζ0 = ξ + d(0) η on W. Write
ζ = ξ0 + σ0 for ξ0 ∈ A and σ0 ∈ S. By orthogonal decomposition, ||ξ0 ||, ||σ0 || ≤ ||ζ0 ||. Now
ξ0 + σ0 = ζ0 = ξ + d(0) η, so we can restrict ξ0 to V and extend to
1
ζ1 = ξ0 + d(0) η1 , with ||ζ1 || ≤ cε||ξ0 || = ||ξ0 ||.
2

661
30.4. The Analytic Existence Theorem Chapter 30. Riemann’s Existence Theorem

Further decompose ζ1 = ξ1 + σ1 for ξ1 ∈ A, σ1 ∈ S and repeat. After k steps, we have

ξk + (σ0 + . . . + σk ) = ξ0 + d(0) (η0 + . . . + ηk )


1
with ||ξk || ≤ k ||ξ0 ||
2
1
and ||σi ||, ||ηi || ≤ i ||ζ0 || for all i.
2
As k → ∞, σ0 + . . . + σk converges in L2 -norm to some σ ∈ S, while η0 + . . . + ηk converges
to some η ∈ ČL0 2 (W, O). Thus we have σ = ξ + d(0) η on W and hence the cohomology class
of ξ is in the image of S.
Theorem 30.4.5 (Finiteness). If X is a compact Riemann surface with sheaf of analytic
functions O = OX , then dimC H 1 (X, O) < ∞.
Proof. For any open cover U of X, we have a commutative diagram
id
Ȟ 1 (X, O) Ȟ 1 (X, O)

Ȟ 1 (U, O) Ȟ 1 (U, O)

By Leray’s theorem (30.4.1), the vertical arrows are isomorphisms but by Proposition 30.4.2,
the restriction map along the bottom has finite dimensional image. This implies the identity
has finite dimensional image, i.e. dim H 1 (X, O) < ∞.
As a preliminary step to proving the analytic existence theorem, we have:
Theorem 30.4.6. Let X be a compact Riemann surface. Then for any a0 ∈ X, there exists
a meromorphic function f ∈ M(X) that has a pole at a0 and is holomorphic elsewhere.
Proof. Fix a0 ∈ X and choose an open cover {Up }p∈X such that p ∈ Up and a0 6∈ Up for
any p 6= a0 . Since X is compact, there is a finite subcover X = U0 ∪ U1 ∪ · · · ∪ Ur with
z
a0 ∈ U0 = Ua0 . Set U = {Ui }ri=0 . We know the map U0 →
− B(0, 1) is biholomorphic.
(m)
Now for each m ∈ N, we define a cocycle ξ (m) = (fij ) ∈ Ž 1 (U, O) by
(m)
fij = 0 when i, j 6= 0
(m)
f00 = 0
(m) (m)
f0j = −fj0 = z −m .

Note that since a0 ∈ U0 , z −m is holomorphic. Moreover, we have the restriction map


Ȟ 1 (U, O) → Ȟ 1 (X, O) and H 1 (X, O) is finite dimensional by Theorem 30.4.5, so there
is a relation
n
X
cm ξ (m) = d(0) η (*)
m=1

662
30.4. The Analytic Existence Theorem Chapter 30. Riemann’s Existence Theorem

where η = (gi ) ∈ Č 0 (U, O) with gi ∈ O(Ui ) and d(0) is the 0th boundary operator on the
Čech complex. Define f : X r {a0 } → C by
(
gi (z), z ∈ Ui and i ≥ 1
f (z) = Pn −m
g0 (z) + m=1 cm z , z ∈ U0 .

By construction, f (z) will be meromorphic so we need only check that the definition is
consistent on overlapping elements of the cover U. Suppose i, j 6= 0 and take z ∈ Ui ∩ Uj .
(m)
Then by definition the ijth part of the left side of (∗) is 0 since (ξ (m) )ij = fij = 0, whereas
on
Pnthe right it’s gi − gj (or the reverse). Hence gi = gj on Ui ∩ Uj . When j = 0, we have
−m
c
m=1 m z = gi − g0 on Ui ∩ U0 so again the definition of f (z) is consistent.
We now prove Theorem 30.2.7.
Proof. Suppose we can find meromorphic functions h1 , . . . , hn ∈ M(X) such that hi (pj ) = δij
for each 1 ≤ i, j ≤ n. Then the function g(z) = c1 h1 (z)+. . .+cn hn (z) will satisfy the desired
property. To produce these hi , first use Theorem 30.4.6 to find meromorphic functions
f1 , . . . , fn ∈ M(X) such that for each i, fi has a pole at pi and is holomorphic everywhere
else. Choose dij ∈ C such that fi (z) − fi (pi ) + dij is nonzero whenever we evaluate at z = pj .
Then define
Y fi (z) − fi (pi )
hi (z) = .
j6=i
fi (z) − fi (pi ) + dij

By the choice of dij , these functions are well-defined, meromorphic and clearly hi (pi ) = 1
and hi (pj ) = 0 for all j 6= i. Therefore the theorem is proved.

663
Chapter 31

Fundamental Groups of Schemes

In this chapter we define the étale fundamental group of a scheme. This will generalize the
construction for curves in Section 29.3 and allow us to fully translate the topological theory
of covering spaces to an algebraic setting. We first recall the so-called Galois theory for
covering spaces.
Let p : Y → X be a topological cover and let Aut(Y /X) be the group of automorphisms
of the cover, i.e. the homeomorphisms Y → Y making the diagram

Y Y

commute. If p is a finite, degree n cover, then # Aut(Y /X) ≤ n. When # Aut(Y /X) = n,
we call p a regular, or Galois cover. In this case, we write Gal(Y /X) = Aut(Y /X).

Proposition 31.0.1. For a covering space p : Y → X of degree n, there is a bijective


correspondence

{subgroups of Gal(Y /X)} ←→ {intermediate covers Y → Y 0 → X}



H 7−→ (Y /H →
− X)
0 0
Aut(Y /Y ) →−7 (Y → X).

Example 31.0.2. Let X = Y = C r {0}, the punctured complex plane, with complex
parameters x and y, respectively. Then an example of a Galois cover of degree n is the nth
power map f : Y → X, y 7→ y n . For example, when n = 2 the real coordinates of ths cover
take the form of a parabola covering the line in all but one point:

664
Chapter 31. Fundamental Groups of Schemes

In general, f : y 7→ y n is a Galois cover with Galois group Gal(Y /X) ∼


= Z/nZ generated by
the automorphism ϕ : Y → Y, y 7→ ζn y, where ζn is a primitive nth root of unity. Viewing
X and Y as complex varieties, f and ϕ are in fact morphisms of varieties which determine
corresponding field embeddings:

C(x) ,−→ C(y) and C(y) ,−→ C(y)


x 7−→ y n y 7−→ ζn y.

We want to generalize this to any variety or scheme.

665
31.1. Galois Theory for Schemes Chapter 31. Fundamental Groups of Schemes

31.1 Galois Theory for Schemes


Let X and Y be schemes and suppose a morphism ϕ : Y → X is finite and locally free. If
the fibre YP over a point P ∈ X is equal to the ring spectrum Spec A of some finite étale
κ(P )-algebra A, then ϕ is étale at P . If this holds for each point P ∈ X, then ϕ is a finite
étale morphism of schemes. Finally, if ϕ is also surjective, we will call it a finite étale cover
of schemes.
Remark. Note that if A is a local ring, a finitely generated A-module M is free if and only
if it is flat. It follows that ϕ : Y → X is locally free if and only if ϕ∗ OY is a sheaf of flat
OX -modules.
Lemma 31.1.1. Let ϕ : Y → X be a finite étale morphism. Then
(a) If ψ : Z → Y is finite étale, then so is ϕ ◦ ψ : Z → X.
(b) If ψ : Z → Y is any morphism, then the base change Y ×X Z → Z is finite étale.
The category FétX is defined to be the full subcategory of SchX consisting of finite étale
ϕ
covers Y −→ X. To compare to the topological case of a covering space (see Section 28.1),
we define:
Definition. A geometric point of X is a morphism x̄ : Spec Ω → X for some algebraically
closed field Ω.
Concretely, the image of x̄ is some point x ∈ X for which κ(x) ⊆ Ω.
Definition. Let ϕ : Y → X be a morphism and x̄ : Spec Ω → X a geometric point. Then
the geometric fibre of x̄ is the fibre product Yx̄ := Y ×X Spec Ω.
Proposition 31.1.2. For a morphism ϕ : Y → X, ϕ is étale at each P ∈ X if and only
if every geometric fibre Yx̄ of ϕ is of the form Spec(Ω × · · · × Ω), where Ω ⊇ κ(P ) is an
algebraically closed field.
Proof. This comes from the algebraic fact (Lemma 15.5.1) that for any field k, a finitely
generated k-algebra A is finite étale if and only if A ⊗k k̄ is isomorphic to a finite direct sum
of copies of k̄, where k̄ is an algebraic closure of k.
Example 31.1.3. The cover C r {0} → C r {0}, y 7→ y n in Example 31.0.2 is a finite étale
cover.
Example 31.1.4. Let X be a normal scheme of dimension 1 and let ϕ : Y → X be a finite
morphism. In the affine case, with X = Spec A and Y = Spec B, this corresponds to an
extension of Dedekind rings ϕ∗ : A → B. Then points in the fibre YP of a point P ∈ X
correspond to a prime factorization of ideals:
r
Y
PB = Qei i
i=1

where Q1 , . . . , Qr are prime ideals of B. Then YP is a finite étale algebra if and only if ei = 1,
i.e. the prime P is unramified in the language of algebraic number theory. In the general
case, a finite morphism of normal schemes of dimension 1 is étale if and only if each affine
piece corresponds to a finite, unramified extension of Dedekind rings.

666
31.1. Galois Theory for Schemes Chapter 31. Fundamental Groups of Schemes

We next translate the ‘local triviality’ condition from covering space theory to schemes.

Proposition 31.1.5. Let X be a connected scheme and ϕ : Y → X an affine surjective


morphism. Then ϕ is a finite étale cover if and only if there exist a locally free, surjective
morphism ψ : Z → X such that Y ×X Z is a trivial cover of Z, i.e. a disjoint union of
copies of Z with the morphism Y ×X Z → Z equal to the identity on each piece.

Proof. ( =⇒ ) The assumption that X is connected means every fibre of ϕ has the same car-
dinality, say n. If n = 1, the property is trivial, so we may induct on n.`By Theorem 15.5.2,
the diagonal Y → Y ×X Y induces a decomposition Y ×X Y = ∆(Y ) Y 0 for Y 0 an open
and closed subscheme of Y ×X Y . The maps Y 0 ,→ Y ×X Y (obvious) and Y ×X Y → Y
(Proposition 15.6.4(a)) are finite étale. Thus Proposition 15.6.4(b) implies their composition
Y 0 → Y is finite étale as well. But by construction the fibres of Y 0 → Y have cardinality
n − 1 so by induction, there exists a finite, locally free, surjective morphism ψ 0 : Z → Y
such that Y 0 ×Y Z consists of n − 1 disjoint copies of Z. Then finally the composition
ψ = ϕ ◦ ψ 0 : Z → X is finite, locally free and surjective (it is the composition of two such
maps) and (Y ×X Y ) ×Y Z ∼ = Y ×X Z is the disjoint union of n copies of Z.
( ⇒= ) Since ψ is locally free, each P ∈ X has an open neighborhood U = Spec A such
that ψ −1 (U ) = Spec C for some finitely generated free A-algebra C. Let ϕ−1 (U ) ⊆ Y be
ϕ−1 (U ) = Spec B for B an A-module. Then the base change of ϕ to ψ over ϕ−1 (U ) is given
by Spec(B ⊗A C) → Spec B. By linear algebra, B ⊗A C is a finitely generated, free C-module
so it is also finitely generated and free as an A-module. Hence B must be finitely generated
and free as an A-module, too.
We next show ϕ is étale using Proposition 31.1.2. Let z̄ : Spec Ω → Z be a geometric
point of Z; then x̄ = ψ ◦ z̄ : Spec Ω → X is a geometric point of X. There is a natural
isomorphism of fibres

Yx̄ = Spec Ω ×X Y ∼
= Spec Ω ×Z (Y ×X Z) = (Y ×X Z)z̄

but the hypothesis implies (Y ×X Z)z̄ is a disjoint union of copies of Spec Ω. Thus Yx̄ is a
disjoint union of points, and since ψ is surjective, every geometric fibre of ϕ is as well. Hence
Proposition 31.1.2 implies ϕ is étale.
Recall from Section 14.3 that the codimension of a point P ∈ X is defined as the di-
mension of the local ring OX,P (or the height of the prime associated to P in any affine
neighborhood). The powerful Zariski-Nagata purity theorem says that to show a map is
étale, it suffices to check the étale property at all codimension 1 points of the base scheme.

Theorem 31.1.6 (Zariski-Nagata Purity Theorem). If ϕ : Y → X is a finite surjective


morphism of integral schemes, with Y normal and X regular, such that the fibre YP of ϕ
above each codimension 1 point P ∈ X is étale over κ(P ), then ϕ itself is étale.

Corollary 31.1.7. Let X be a regular integral scheme and U ⊆ X an open subscheme


consisting of points of codimension at least 2. Then there is an equivalence of categories

FétX −→ FétU
Y 7−→ Y ×X U.

667
31.1. Galois Theory for Schemes Chapter 31. Fundamental Groups of Schemes

Proposition 31.1.8. Let ϕ : Y → X be a finite étale cover, z̄ : Spec Ω → z a geometric


point and Z a connected scheme over X. If f, g : Z → Y are two X-morphisms such that
f ◦ z̄ = g ◦ z̄, then f = g.
Proof. By Proposition 15.6.4(a), we may assume X = Z, so that f and g are two sections
of ϕ : Y → X. It follows that f and g are finite étale morphisms and each induces an
isomorphism of Z = X with an open and closed subscheme of Y . Since Z is connected, the
images of f and g are determined by the images of any geometric point, hence f ◦ z̄ = g ◦ z̄
implies f = g.
Corollary 31.1.9. Let ϕ : Y → X be a finite étale cover. Then Aut(Y /X) is finite.
Proof. Take σ ∈ Aut(Y /X), σ 6= 1, and set f = ϕ and g = ϕ◦σ. Then by Proposition 31.1.8,
f and g send some geometric point of Y to different points. In other words, the action of
Aut(Y /X) on any geometric fibre is free, or the permutation representation of Aut(Y /X)
on any of the geometric fibres is faithful. Now since the map is finite étale, each geometric
fibre is finite as a set, so this implies Aut(Y /X) is itself finite.
Now let ϕ : Y → X be a finite étale cover and let G be a group scheme over X such that
Y is a left G-torsor (see Section 14.5). Let Y /G be the quotient space with projection map
π : Y → Y /G. We define a sheaf on Y /G by OY /G := (π∗ OY )G , the subsheaf of G-invariants
of the pushforward of OY to Y /G along π. This makes Y /G into a ringed space.
Proposition 31.1.10. The ringed space (Y /G, OY /G ) is a scheme over X. Moreover, ϕ :
Y → X factors through a finite morphism ψ : Y /G → X.
The following are analogues of the basic Galois theory of covering spaces (see Section 28.1,
e.g.).
Proposition 31.1.11. If ϕ : Y → X is a connected, finite étale cover and G ≤ Aut(Y /X)
is any finite subgroup of automorphisms, then π : Y → Y /G is a finite étale cover.
Definition. A connected, finite étale cover ϕ : Y → X is a Galois cover if Aut(Y /X) acts
transitively on every geometric fibre of ϕ.
Theorem 31.1.12. Let ϕ : Y → X be a Galois cover and suppose ψ : Z → X is a connected,
finite étale cover such that Z is a scheme over Y and the diagram
Y Z

ϕ ψ

X
commutes. Then
(1) Y → Z is a Galois cover and Z ∼
= Y /G for some subgroup G ≤ Aut(Y /X).
(2) There is a bijection
{subgroups G ≤ Aut(Y /X)} ←→ {intermediate covers Y → Z → X}.

(3) The correspondence is bijective on normal subgroups of Aut(Y /X) and Galois covers
Z → X, and in this case Aut(Z/X) ∼ = Aut(Y /X)/G as groups.

668
31.2. The Étale Fundamental Group Chapter 31. Fundamental Groups of Schemes

31.2 The Étale Fundamental Group


Let X be a scheme and FétX the category of finite étale covers of X. Fix a geometric point
x̄ : Spec Ω → X and let

Fibx̄ : FétX −→ Sets


Y 7−→ Yx̄ = Y ×X Spec Ω

be the fibre functor over x̄. By Lemma 14.3.6(c), Fibx̄ is indeed a functor. Using the
definition of the automorphism group of a functor from Section 28.1, we can now define the
algebraic, or étale, fundamental group of a scheme.

Definition. The algebraic, or étale fundamental group of a scheme of X at a geometric


point x̄ : Spec Ω → X is the automorphism group of the fibre functor over x̄:

π1 (X, x̄) := Aut(Fibx̄ ).

By Theorem 28.1.3, this coincides with the topological fibre functor and fundamental
group. We will also see below that this definition for schemes coincides with the definition
given for curves in Section 29.3.

Theorem 31.2.1 (Grothendieck). Let X be a connected scheme and x̄ : Spec Ω → X a


geometric point. Then

(1) π1 (X, x̄) is a profinite group and its action on Fibx̄ (Y ) is continuous for all Y ∈ FétX .

(2) Fibx̄ induces an equivalence of categories



FétX −→ {finite, continuous, left π1 (X, x̄)-sets}
Y 7−→ Fibx̄ (Y ).

Example 31.2.2. Let k be a field and consider X = Spec k. Then finite étale covers
Y → Spec k are precisely Y = Spec A for A = L1 × · · · × Lr a finite étale k-algebra. Here,
the fibre functor over any x̄ : Spec k̄ → Spec k (equivalent to a choice of algebraic closure k̄
of k) is exhibited by

Fibx̄ (Y ) = Spec(A ⊗k k̄) = Spec(Ω × · · · × Ω),

and Spec(Ω × · · · × Ω) is a finite set of r closed points indexed by the homomorphisms


A → k̄. Indeed, as we saw in Section 28.2, Fibx̄ (Y ) ∼
= Homk (A, ks ), where ks is the separable
closure of k in k̄ via the embedding k ,→ k̄. The action of Aut(Homk (−, ks )) on any
given Homk (A, ks ) is given by T · σ = σ ◦ T , or precisely the action of Gk = Gal(ks /k)
on Homk (A, ks ) from Example 28.3.2. Therefore

π1 (X, x̄) = Aut(Fibx̄ ) ∼


= Aut(Homk (−, ks )) ∼
= Gal(ks /k).

Moreover, there is an identification (via Prop. 14.2.2) Homk (A, ks ) = HomSpec k (Spec ks , Spec k).

669
31.2. The Étale Fundamental Group Chapter 31. Fundamental Groups of Schemes

The above shows that although Fibx̄ is not a representable functor – Spec ks is not a
finite étale k-scheme – it is pro-representable in FétSpec k . Explicitly,

Fibx̄ (Spec A) = lim HomSpec k (Y 0 , Spec A)


−→

where the direct limit is over all finite étale Galois covers Y 0 → Spec k ordered by the
existence of a Spec k-morphism Y 0 → Y 00 .

In fact, there was nothing special about X being Spec k or even affine in the last para-
graph.

Proposition 31.2.3. For a connected scheme X and any geometric point x̄ : Spec Ω → X,
the fibre functor Fibx̄ is pro-representable. Explicitly, for any finite étale cover Y → X,

Fibx̄ (Y ) = lim HomX (Y 0 , Y )


−→

where the direct limit is over all finite étale Galois covers Y 0 → Y .

Lemma 31.2.4. Every automorphism of Fibx̄ is determined by a unique automorphism of


the direct system (Y 0 → X) of finite Galois covers of X.

We can now give the proof of Theorem 31.2.1.


Proof. (1) By Proposition 31.2.3, Fibx̄ is pro-representable by the direct system of HomX (Y 0 , Y )
where Y 0 ranges over all finite Galois covers of X. Now Lemma 31.2.4 implies

Aut(Fibx̄ ) = lim Aut(Y 0 /X)


←−

where again Y 0 ranges over all finite Galois covers. Since Corollary 31.1.9 says that each
Aut(Y 0 /X) is a finite group, we see that π1 (X, x̄) = Aut(Fibx̄ ) is a profinite group. Note
that π1 (X, x̄) has a natural action on each Fibx̄ (Y ) for any finite étale cover Y → X.
It remains to show this action is continuous. Take a geometric point ȳ ∈ Fibx̄ (Y ) lying
over x̄ : Spec Ω → X. If ȳ comes from an element of HomX (Y 0 , Y ) by the direct limit in
Proposition 31.2.3, then the action of π1 (X, x̄) factors through Aut(Y 0 /X). This implies
continuity.
(2) Take a finite, continuous, left π1 (X, x̄)-set S. The action of π1 (X, x̄) is transitive on
each orbit of S, so we may assume it is transitive on S to begin with. Let H be the stabilizer
of any point in S. Then H is an open subgroup of π1 (X, x̄), so it contains the open normal
subgroup corresponding to the kernel of π1 (X, x̄) → Aut(Y 0 /X) for some finite Galois cover
Y 0 → X. Let H be the image of H in Aut(Y 0 /X). Then one proves S ∼ = Y 0 /H. This proves
essential surjectivity; fully faithfulness is routine.
The following shows that our definition of the fundamental group for schemes properly
captures the notion we described for curves in Section 29.3.

Theorem 31.2.5. Let X be an integral normal scheme with function field K and fix a
separable closure Ks /K. Let KX be the compositum of all finite subextensions L/K in Ks
such that the normalization XL of X in L is étale over X. Then

670
31.2. The Étale Fundamental Group Chapter 31. Fundamental Groups of Schemes

(1) KX /K is Galois.

(2) For any geometric point x̄ : Spec K → X, we have π1 (X, x̄) ∼


= Gal(KX /K).
Proof. The proof of (1) is basically the same as the proof of Proposition 29.3.1. Then (2)
follows from (2) of Theorem 31.2.1.
Example 31.2.6. Let X = Y = C r {0} and define the finite étale cover

f : Y −→ X
y 7−→ πy n .

(Here, we use π to ensure there are no algebraic relations on this map.) As in Example 31.1.3,
f is a finite étale Galois cover of degree n, with Galois group G = hσi ∼= Z/nZ, where σ is
the automorphism σ : y 7→ ζn y for ζn a primitive nth root of unity. Notice that by setting
z = π 1/n y, we may define f over Q. We make use of the following important result.
Theorem 31.2.7. If a scheme X/C is defined over Q, then any finite étale cover of X is
also defined over Q.
In particular, the finite cover f : y 7→ z n is defined over some number field K/Q. A
classic question asks for a description of K in terms of the topology of this cover.
Set GQ = Gal(Q/Q) and fix an automorphism ω ∈ GQ . Let G be any group and suppose
Y → X is a G-Galois cover which is defined over a number field K. Then Gal(Q/K) ≤ GQ
and if ω lies in Gal(Q/K), then ω carries this cover Y → X to itself (the action on Y being
the action of ω on the coefficients of the map). The field of moduli for the cover Y → X is
defined to be the subfield M of Q/Q fixed by all ω taking Y → X to itself.
Proposition 31.2.8. If G is an abelian group, then the unique maximal field of definition
for Y → X is precisely M.
Note that Proposition 31.2.8 is completely open when G is nonabelian. In any case, the
consequences for abelian covers is that we can see the Galois theory of finite étale covers in
two parallel ways:
ˆ (Geometric) Every Galois cover of schemes has a group of automorphisms and a cor-
responding field extension for which this is the Galois group.

ˆ (Arithmetic) Every cover of schemes has a field of definition over Q, and in the abelian
case there is even a unique minimal such field of definition.
As in Section 29.4, these related pieces are governed by a short exact sequence

1 → π1 (X) → π1 (X) → Gal(k̄/k) → 1

(here, the base point can be omitted and k can be any field). Accordingly, π1 (X) is called
the geometric fundamental group of X, π1 (X) the algebraic fundamental group and Gal(k̄/k)
the arithmetic fundamental group. As in Corollary 29.4.4, the sequence splits if and only if
X contains a k-rational point.

671
31.3. Properties of the Fundamental Group Chapter 31. Fundamental Groups of Schemes

31.3 Properties of the Fundamental Group


Let X be a connected scheme and take two geometric points x̄ : Spec Ω → X and x̄0 :
Spec Ω0 → X. In the topological case (Section 28.1), there was a nice notion of a path
between two points. We now describe an analogue for schemes.

Proposition 31.3.1. For any two geometric points of X, x̄ : Spec Ω → X and x̄0 : Spec Ω0 →

X, there is a natural isomorphism Fibx̄ −
→ Fibx̄0 .

Proof. We saw in Proposition 31.2.3 that each fibre functor is pro-representable by an inverse
system (Pα ) of finite étale Galois covers of X, but the choice of X-morphisms in each case
depends on the choices of basepoints in each Pα lying over x̄ or x̄0 . Let the covering morphisms
be denoted ϕαβ , ψαβ : Pβ → Pα for x̄, x̄0 , respectively. Then by Proposition 31.2.3, for any
Y ∈ FétX , Fibx̄ (Y ) ∼
= lim Hom(Pα , Y ) where the connecting homomorphisms in the direct
−→
limit are those induced by the ϕαβ ; and similarly Fibx̄0 (Y ) ∼
= lim Hom(Pα , Y ) with the maps
−→
induced by the ψαβ . Therefore is suffices to exhibit an isomorphism of inverse systems
(Pα , ϕαβ ) ∼
= (Pα , ψαβ ).
Let α ≤ β and take λβ ∈ AutX (Pβ ). Let xα ∈ Pα and xβ ∈ Pβ be the specified points
over x̄ and set x0α = ψαβ ◦ λβ (xβ ):
xβ λβ λβ (xβ )
Pβ Pβ

ϕαβ ψαβ

Pα Pα
xα λα x0α

Since Pα is a Galois cover of X, there exist a (unique) automorphism λα ∈ AutX (Pα )


such that λ(xα ) = x0α . It follows from Proposition 31.1.8 that this λα makes the diagram
above commute. Define ραβ : AutX (Pβ ) → AutX (Pα ) by sending λβ ∈ AutX (Pβ ) to the λα
described above. This gives an inverse system (AutX (Pα , ραβ ) in which each AutX (Pα ) is
finite and nonempty, so the corresponding inverse limit lim AutX (Pα ) is nonempty. That is,
←−
there is at least one system of isomorphisms of (Pα ) compatible with the ϕαβ and ψαβ .

Corollary 31.3.2. For any pair of geometric points x̄ : Spec Ω → X and x̄0 : Spec Ω0 → X,
there is a continuous isomorphism of profinite groups π1 (X, x̄0 ) → π1 (X, x̄).

Proof. Any natural isomorphism λ : Fibx̄ − → Fibx̄0 from Proposition 31.3.1 determines an
isomorphism λ : π1 (X, x̄ ) → π1 (X, x̄) by λ∗ (ϕ) = λ−1 ◦ ϕ ◦ λ.
∗ 0

Definition. A natural isomorphism λ : Fibx̄ → Fibx̄0 is called a path from x̄ to x̄0 .

Note that the isomorphism λ∗ : π1 (X, x̄0 ) → π1 (X, x̄) depends on the choice of path λ,
but only up to inner automorphism.

672
31.3. Properties of the Fundamental Group Chapter 31. Fundamental Groups of Schemes

Now let X and X 0 be two connected schemes, take a geometric point in each, x̄ : Spec Ω →
X and x̄0 : Spec Ω → X 0 , and suppose there is a covering morphism ϕ : X 0 → X such that
ϕ ◦ x̄0 = x̄. Then ϕ induces a base change functor

Bx̄,x̄0 : FétX −→ FétX 0


Y 7−→ Y ×X X 0 .

Recall from the proof of Proposition 31.1.5 that there is a natural isomorphism Fibx̄ ∼
=
Fibx̄0 ◦Bx̄,x̄0 . This induces a map

ϕ∗ : π1 (X 0 , x̄0 ) −→ π1 (X, x̄).

Lemma 31.3.3. ϕ∗ : π1 (X 0 , x̄0 ) → π1 (X, x̄) is a continuous homomorphism of profinite


groups.

Proposition 31.3.4. Let ϕ : X 0 → X be a cover of connected schemes. Then

(1) ϕ∗ is trivial if and only if for all connected Y ∈ FétX , the base change Y ×X X 0 → X 0
is a trivial cover.

(2) ϕ∗ is surjective if and only if for all connected Y ∈ FétX , the base change Y ×X X 0 →
X 0 is a connected cover.

Proof. (1) First note that an X 0 -cover is trivial if and only if π1 (X 0 , x̄0 ) acts trivially on its
fibre over x̄0 . Thus if, for any λ ∈ π1 (X 0 , x̄0 ) and Y ∈ FétX , λ acts trivially on Fibx̄0 (Y ×X
X 0) ∼
= Fibx̄ (Y ), then ϕ∗ (λ) must be trivial. Conversely, if some Y ×X X 0 is a nontrivial
cover, there exists λ ∈ π1 (X 0 , x̄0 ) acting nontrivially on Fibx̄ (Y ) ∼
= Fibx̄0 (Y ×X X 0 ). Hence
ϕ∗ (λ) is nontrivial.
(2) Suppose Y ∈ FétX (connected) is such that Y ×X X 0 is not connected. Then there
exist x1 , x2 ∈ Fibx̄ (Y ) such that ϕ∗ (λ)(x1 ) 6= x2 for any λ ∈ π1 (X 0 , x̄0 ). However, the fact
that Y is connected implies there exists µ ∈ π1 (X, x̄) such that µ(x1 ) = x2 . Thus µ is not
in the image of ϕ∗ , so ϕ∗ is not surjective. Going the other direction, if ϕ∗ is not surjective
then im ϕ∗ is a proper, closed subgroup of the profinite group π1 (X, x̄), so there exists a
proper, open subgroup U ⊆ π1 (X, x̄) such that U ⊇ im ϕ∗ . Thus π1 (X 0 , x̄0 ) acts trivially on
the coset space π1 (X, x̄)/U , or in other words, the connected cover Y → X corresponding
to π1 (X, x̄)/U from Theorem 31.1.12 base changes to a trivial cover Y ×X X 0 → X 0 . But if
U 6= π1 (X, x̄), then Y ×X X 0 6= X 0 and is therefore a disconnected cover.

Remark. Suppose Y → X is connected, ȳ is a geometric point over x̄ and U = Stabπ1 (X,x̄) (ȳ).
Then U is an open subgroup of π1 (X, x̄) and Fibx̄ (Y ) ∼
= π1 (X, x̄)/U as left π1 (X, x̄)-sets.
As above, let ϕ : X 0 → X be a cover of connected schemes such that ϕ ◦ x̄0 = x̄.

Proposition 31.3.5. Let U ⊆ π1 (X, x̄) be any open subgroup and let Y → X be the cor-
responding connected cover; let ȳ ∈ Fibx̄ (Y ) be the geometric point such that U = Stab(ȳ).
Then U ⊇ im ϕ∗ if and only if the base change Y ×X X 0 → X 0 has a section such that
y 0 := x̄0 (Spec Ω) ⊆ X 0 maps to y := ȳ(Y ×X Spec Ω) ⊂ Y ×X X 0 .

673
31.3. Properties of the Fundamental Group Chapter 31. Fundamental Groups of Schemes

Proof. Note that

U ⊇ im ϕ∗ ⇐⇒ every element of π1 (X, x̄) fixes ȳ


⇐⇒ the connected component of y is fixed under the π1 (X, x̄)-action
⇐⇒ the connected component of y is isomorphic to X 0 .

Therefore under Y ×X X 0 → X 0 , the component of y maps isomorphically to X 0 , so we get


a section X 0 ,→ Y ×X X 0 . Conversely, any such section determines a π1 (X 0 , x̄0 )-equivariant
map Spec Ω → Fibx̄0 (Y ×X X 0 ) whose image must be fixed. Hence U ⊇ im ϕ∗ .
Proposition 31.3.6. Let U 0 ⊇ π1 (X 0 , x̄0 ) be an open subgroup, Y 0 → X 0 the corresponding
connected cover and ȳ 0 the geometric point such that Stabπ1 (X 0 ,x̄0 ) (ȳ 0 ) = U 0 . Then U 0 ⊇ ker ϕ∗
if and only if there exists a connected cover Y → X and an X 0 -morphism Y0 → Y 0 such that
Y0 is a connected component of Y ×X X 0 → X 0 .
Proof. ( =⇒ ) Suppose U 0 ⊇ ker ϕ∗ . We know im ϕ∗ is a subgroup in π1 (X, x̄) and V 0 :=
ϕ∗ (U 0 ) is open in im ϕ∗ , so by profinite group theory, there exists an open subgroup V ⊆
π1 (X, x̄) such that V ∩ im ϕ∗ = V 0 . This V corresponds to a cover Y → X. Let Y0 be the
connected component of Y ×X X 0 → X 0 containing y = ȳ(Y ×X Spec Ω) ⊆ Y ×X X 0 . Then
ȳ ∈ Fibx̄ (Y0 ) and there is an isomorphism of π1 (X 0 , x̄0 )-sets

Fibx̄ (Y0 ) ∼
= π1 (X 0 , x̄0 )/U 00 where U 00 = Stab(ȳ).

Now an X 0 -morphism Y0 → Y 0 exists if and only if there is an equivariant map π1 (X 0 , x̄0 )/U 00 →
π1 (X 0 , x̄0 )/U 0 . This in turn occurs if and only if U 00 ⊆ U 0 . If λ ∈ ker ϕ∗ , then λ fixes ȳ in
Fibx̄ (Y ) = Fibx̄0 (Y ×X X 0 ), hence λ fixes ȳ in Fibx̄0 (Y0 ). This means λ ∈ U 00 and so we see
that U 00 ⊇ ker ϕ∗ . By hypothesis, U 0 ⊇ ker ϕ∗ as well, so by the correspondence theorem,
U 00 ⊆ U 0 if and only if ϕ∗ (U 00 ) ⊆ ϕ∗ (U 0 ), but this holds by construction. Therefore an
X 0 -morphism Y0 → Y 0 exists.
( ⇒= ) We must show U 0 ⊇ ker ϕ∗ . Applying the fibre functor Fibx̄0 to the morphism
Y0 → Y 0 gives a map Fibx̄0 (Y0 ) → Fibx̄0 (Y 0 ). Note that ȳ 0 ∈ Fibx̄0 (Y 0 ); choose any lift
ȳ ∈ Fibx̄0 (Y0 ) and set U = Stab(ȳ). Identifying Y0 → Y 0 with a subgroup U 00 ⊆ π1 (X 0 , x̄0 ),
we get U 00 ⊆ U 0 by the above. Then once again, ker ϕ∗ has to fix ȳ, so ker ϕ∗ ⊆ U 00 ⊆ U 0 .
Corollary 31.3.7. ϕ∗ is injective if and only if for all connected covers Y 0 → X 0 , there
exists a cover Y → X and a morphism Y0 → Y 0 over X 0 such that Y0 is isomorphic to a
connected component of Y ×X X 0 .
Corollary 31.3.8. If every connected cover Y 0 → X 0 arises as a base change Y 0 = Y ×X X 0
of some Y → X, then ϕ∗ is injective.
ψ ϕ
Corollary 31.3.9. Let X 00 − → X0 −→ X be a pair of morphisms of connected schemes and
let x̄ : Spec Ω → X, x̄ : Spec Ω → X 0 and x̄00 : Spec Ω → X 00 be geometric points such that
0

ϕ ◦ x̄0 = x̄ and ψ ◦ x̄00 = x̄0 . Then the sequence of profinite groups


ψ∗ ϕ∗
π1 (X 00 , x̄00 ) −→ π1 (X 0 , x̄0 ) −→ π1 (X, x̄)

is exact if and only if both of the following conditions hold:

674
31.3. Properties of the Fundamental Group Chapter 31. Fundamental Groups of Schemes

(1) For all covers Y → X, the base change Y ×X X 00 → X 00 is a trivial cover.

(2) For all connected covers Y 0 → X 0 such that Y 0 ×X 0 X 00 → X 00 admits a section


X 00 → Y 0 ×X 0 X 00 , there exists a connected cover Y → X and an X 0 -morphism Y0 → Y 0
for Y0 a connected component of Y ×X X 0 .

Remark. It suffices to check condition (2) on Galois covers, since


\
ker ϕ∗ = U.
open normal
U ⊇ker ϕ∗

675
31.4. Structure Theorems Chapter 31. Fundamental Groups of Schemes

31.4 Structure Theorems


In this section we prove several fundamental results about π1 (X) for schemes over C, and
then generalize. In particular, we will state an analogue of Theorem 29.3.5. First, let us
show that for any smooth, projective, integral scheme X over C, π1 (X) is finitely generated.
To do this, we need:

Theorem 31.4.1 (Bertini). Let k be an algebraically closed field and let X ⊆ Pnk be a smooth,
closed subscheme. Then there exists a hyperplane H ⊂ Pnk not containing X such that X ∩ H
is regular.

Proof. Hartshorne, II.8.1.8.

Lemma 31.4.2. Let k be algebraically closed and let X ⊆ Pnk be a smooth, connected, closed
subscheme of dimension dim X ≥ 2 and let Y → X be a connected, finite étale cover. Then
there exists a hyperplane H ⊆ Pnk not containing X so that X ∩ H is smooth and connected
and Y ×X (X ∩ H) is connected.

Proof. Let H be as in Bertini’s theorem. When dim X ≥ 2, one can use sheaf cohomology
(cf. Hartshorne, III.7.9.1) to show that X ∩ H is not just regular but also smooth and
connected. Then the same proof applies to the closed immersion Y ×X (X ∩ H) ,→ Y to
show that Y ×X (X ∩ H) is connected.

Theorem 31.4.3. For any smooth, projective, integral scheme X over C and every geometric
point x̄ : Spec C → X, π1 (X, x̄) is finitely generated.

Proof. Suppose dim X ≥ 2. Then by Lemma 31.4.2, for any connected Y → X there
exists a hyperplane H ⊆ PnC such that X ∩ H is smooth, connected and Y ×X (X ∩ H) is
connected. By Proposition 15.6.4(a), Y× (X ∩ H) → X ∩ H is a (connected) finite étale cover,
so Proposition 31.3.4 implies π1 (X ∩ H, x̄) → π1 (X, x̄) is surjective. Now note that X ∩ H
is smooth, projective and dim X ∩ H < dim X, so repeating this process, we eventually
obtain a surjection π1 (C, x̄) → π1 (X, x̄) for some smooth projective curve C over C. By
Theorem 29.3.5, we know π1 (C, x̄) is finitely generated so this implies π1 (X, x̄) is as well.
Even stronger than the surjections π1 (X ∩ H) → π1 (X) in the above proof, we have the
following result of Lefschetz.

Theorem 31.4.4 (Lefschetz’s Hyperplane Theorem). Let X be a smooth, closed projective


scheme over C and H ⊆ PnC a hyperplane such that X ∩ H is smooth. Then

Hk (X ∩ H) −→ Hk (X)
H k (X ∩ H) −→ H k (X)
πk (X ∩ H) −→ πk (X)

are all isomorphisms for k < dim X − 1 and are surjective for k = dim X − 1.

676
31.4. Structure Theorems Chapter 31. Fundamental Groups of Schemes

To compare π1 and π1top as we did in Chapter 30, we associate to any C-scheme X of finite
an
S
type a complex analytic space X as follows. Write X = Spec Ai for C-algebras Ai of finite
type. Then Ai ∼ = C[x1 , . . . , xn ]/(f1 , . . . , fr ) for polynomials f1 , . . . , fr ∈ C[x1 , . . . , xn ]. Then
each fj may be regarded as a holomorphic function on Cn , so the zero set of {f1 , . . . , fr }
in Cn is a ringed space (Yi , OYi ), where OYi is the ring of holomorphic functions on Yi (in
the sense of complex geometry). We regard this (Yi , OYi ) as the basic model S of a complex
analytic space. Now glue the Yi together using the same gluing as for Spec Ai to get a
ringed space (X an , OX
an
). Note that for any morphism of schemes Y → X, there is a natural
induced morphism of complex analytic spaces Y an → X an . The complex analytic space X an
has the following properties relative to X.
Proposition 31.4.5. Let X/C be a scheme of finite type. Then
(1) X is separated if and only if X an is Hausdorff.

(2) X is connected if and only if X an is connected.

(3) X is reduced if and only if X an is reduced.

(4) X is smooth if and only if X an is a complex manifold.

(5) A morphism Y → X is proper (i.e. of finite type and the base change is closed, as
in Section 14.3) if and only if Y an → X an is proper (i.e. compact sets pull back to
compact sets).

(6) In particular, X is proper as a C-scheme if and only if X an is compact.

(7) There is a morphism of locally ringed spaces (ϕ, ϕ# ) : (X an , OX


an
) → (X, OX ) sending
an
X bijectively to the closed points of X.
As in Chapter 30 for curves, Grothendieck proved the following equivalence of categories.
Theorem 31.4.6 (Grothendieck). Let X be a connected scheme over C of finite type. Then
there is an equivalence of categories

FétX −→ FCovX an (finite-sheeted topological covers)


(Y → X) 7−→ (Y an → X an ).

Therefore, for any geometric point x̄ : Spec C → X with image x = x̄(Spec C), the induced
map
π1top\
(X an , x) −→ π1 (X, x̄)
is an isomorphism.
Just like in Chapter 30, the difficult part of this proof is establishing the essential sur-
jectivity of the functor Y 7→ Y an . This is essentially the main idea behind Serre’s GAGA
principle.
Question (Serre). Does there exist a scheme X over C of finite type such that π1 (X) = 1
but π1top (X an ) 6= 1?

677
31.4. Structure Theorems Chapter 31. Fundamental Groups of Schemes

To extend Theorem 31.4.6 to any algebraically closed field k of characteristic 0, we need


the following generalization of Proposition 29.4.1.
Proposition 31.4.7. Let X be a noetherian integral scheme, k an algebraically closed field,
ϕ : Y → X a proper flat morphism with geometrically integral fibres and suppose x̄ : Spec k →
X is a geometric point such that k is the algebraic closure of the residue field κ(x), where
x = im x̄. Fix ȳ ∈ Yx̄ . Then there is an exact sequence
π1 (Yx̄ , ȳ) → π1 (Y, ȳ) → π1 (X, x̄) → 1.
Corollary 31.4.8. Let k be algebraically closed and suppose X, Y are noetherian, connected
schemes over k, with X proper and geometrically integral. Then for any geometric points
x̄ : Spec k → X, ȳ : Spec k → Y , the natural map π1 (X ×k Y, (x̄, ȳ)) → π1 (X, x̄) × π1 (Y, ȳ) is
an isomorphism.
Proof. From Proposition 31.4.7, we get a diagram (with basepoints suppressed):
1 π1 (X) π1 (X ×k Y ) π1 (Y ) 1

id id

1 π1 (X) π1 (X) × π1 (Y ) π1 (Y ) 1

with the 1 on the left in the top row coming from the natural section X ,→ X ×k Y of
the projection morphism X ×k Y → X, and the entire bottom row representing the direct
product of profinite groups. Since the left and right vertical arrows are identities, the Five
Lemma implies the middle vertical arrow is an isomorphism.
Corollary 31.4.9. Let K ⊇ k be an extension of algebraically closed fields, X a proper
integral scheme over k and XK its base change. Then for any geometric points x̄ : Spec k →
X, x̄0 : Spec K → XK , the induced map π1 (XK , x̄0 ) → π1 (X, x̄) is an isomorphism.
Proof. We first show that π1 (XK , x̄0 ) → π1 (X, x̄) is surjective. Suppose Y → X is connected.
Then since k is algebraically closed, the K-algebra k(Y )⊗k K is a field, but since this is equal
to the function field of YK , we see that YK is connected. Therefore by Proposition 31.3.4,
the induced map on fundamental groups is surjective.
For injectivity, suppose Y → XK is a connected cover. It is possible (see Szamuely) to find
a subfield k 0 ⊆ K such that k 0 /k is finitely generated, as well as an integral affine k-scheme T
such that k(T ) = k 0 and there is a connected finite étale cover Y 0 → X ×k T . Fix a geometric
point t̄ : Spec k → T . By Corollary 31.4.8, there is an isomorphism π1 (X ×k T, (x̄, t̄)) ∼ =
π1 (X, x̄) × π1 (T, t̄), so Corollary 31.3.7 implies there exist connected covers Z → X and
T 0 → T and Z ×T T 0 → Y 0 a morphism of T -schemes. Take a k-point t ∈ T ; then the fibre
(Z ×T T 0 )t of the morphism Z ×T T 0 → T is a finite étale cover of X and there is a morphism
Z ×T T 0 → Y of schemes over XK . Thus by Corollary 31.3.7, π1 (XK , x̄0 ) → π1 (X, x̄) is
injective.
Corollary 31.4.10. Let k be an algebraically closed field of characteristic 0, X a smooth,
connected, projective scheme over k and x̄ : Spec k → X a geometric point. Then π1 (X, x̄)
is finitely generated.

678
31.4. Structure Theorems Chapter 31. Fundamental Groups of Schemes

Proof. We know the result for k = C by Theorem 31.4.3, so applying Corollary 31.4.9
to the extension C/Q shows that any Q-scheme has finitely generated fundamental group.
Finally, any algebraically closed field of characteristic 0 contains Q, so one more application
of Corollary 31.4.9 finishes the proof.

679
31.5. Results in Characteristic p Chapter 31. Fundamental Groups of Schemes

31.5 Results in Characteristic p


To study fundamental groups of schemes in characteristic p > 0, we give a brief account of
Grothendieck’s theory of specialization of the fundamental group. Suppose A is a complete
DVR with fraction field K and residue field k of characteristic p > 0. Set X = Spec A and
let η : Spec K ,→ X and x : Spec k ,→ X be the generic and closed points of X, respectively.
Also let η̄ : Spec K ,→ Spec K ,→ X and x̄ : Spec k̄ ,→ Spec k ,→ X be the corresponding
geometric points. For any X-scheme Y → X, set Yη = Y ×X Spec K, Yx = Y ×X Spec k, Yη̄ =
Y ×X Spec K and Yx̄ = Y ×X Spec k̄. Grothendieck proves:

Theorem 31.5.1. If ϕ : Y → X is a proper map, then

(1) For any geometric point ȳ 0 ∈ Yx , the natural map

π1 (Yx , ȳ 0 ) −→ π1 (Y, ȳ 0 )

is an isomorphism.

(2) Further, if k is algebraically closed, ϕ is flat and Yη̄ and Yx̄ are geometrically reduced,
then for any geometric point ȳ ∈ Yη̄ , the natural map

π1 (Yη̄ , ȳ) −→ π1 (Y, ȳ)

is an isomorphism.

For the moment, assume k is algebraically closed, ϕ : Y → X is a morphism of schemes


and ȳ ∈ Yη̄ and ȳ 0 ∈ Yx are geometric points as above. By Corollary 31.3.2, there is a

→ π1 (Y, ȳ 0 ) determined by picking a path λ : Fibȳ →
noncanonical isomorphism π1 (Y, ȳ) −
Fibȳ0 . The isomorphisms in Theorem 31.5.1 allow us to make the following definition.

Definition. The specialization map from ȳ to ȳ 0 (associated to ϕ : Y → X) is the


composite
∼ ∼
→ π1 (Y, ȳ 0 ) −
sp : π1 (Yη̄ , ȳ) → π1 (Y, ȳ) − → π1 (Yx̄ , ȳ 0 ).
0
For any profinite group G and prime integer p, we denote by G(p ) the maximal quotient
of G whose order (as a profinite group) is relatively prime to p. This is typically called the
prime-to-p part of G. We will prove an important result of Grothendieck which says that
specialization induces an isomorphism on prime-to-p parts of the fundamental groups of Yη̄
and Yx̄ . First, we need:

Lemma 31.5.2 (Abhyankar). Let (A, m, k) be a DVR with fraction field K and L/K and
M/K finite Galois extensions in which A has integral closures AL and AM , respectively, and
integral closure B in the compositum LM . Suppose pL ⊂ AL and pM ⊂ AM are maximal
ideals lying over m such that the inertia degrees e(pL | m) and e(pM | m) are relatively prime
to char k and e(pL | m) | e(pM | m). Then Spec B → Spec AM is a finite étale morphism of
affine schemes.

680
31.5. Results in Characteristic p Chapter 31. Fundamental Groups of Schemes

Proof. Set p = char k. Fix a maximal ideal P ⊂ B lying over m. We may assume P∩AL = pL
and P ∩ AM = pM . Then the injection Gal(LM/K) ,→ Gal(L/K) × Gal(M/K) restricts to
an injection of inertia groups IP ,→ IpL × IpM . Moreover, the induced projections IP → IpL
and IP → IpM are surjective. Since |Ip | = e(p | m) for any prime p lying over m, then
by hypothesis the inertia groups IP , IpL and IpM all have prime-to-p order and hence are
cyclic. Further, since |IpL | divides |IpM |, every element of IP ⊆ IpL × IpM has order dividing
e(pM | m), but since there is a surjection IP → IpM and the target has order e(pM | m), this
implies the map is really an isomorphism IP ∼ = IpM . By transitivity of inertia degrees, that
is e(P | m) = e(P | pM )e(pM | m), we get e(P | pM ) = 1, so P is unramified over M . Since
P was arbitrary, it follows from Proposition 29.2.6 that Spec B → Spec AM is étale.
Theorem 31.5.3 (Grothendieck). If k is algebraically closed of characteristic p > 0, ϕ :
Y → X = Spec A is smooth, proper and has geometrically connected fibres, then for any
geometric points ȳ ∈ Yη̄ and ȳ 0 ∈ Yx , the specialization map descends to an isomorphism of
prime-to-p parts
0 ∼ 0
π1 (Yη̄ , ȳ)(p ) −→ π1 (Yx̄ , ȳ 0 )(p ) .
0 0
Proof. By definition of sp, it’s enough to show π1 (Yη̄ , ȳ)(p ) → π1 (Y, ȳ)(p ) is an isomorphism
of prime-to-p profinite groups. To begin, suppose Z → Y is a connected finite étale cover
with base change Zη̄ = Z ×X Spec K over the generic point. Then

Zη̄ = lim Z ×X Spec L


←−

where the inverse limit is over all finite extensions L/K contained in a fixed algebraic closure
K corresponding to η̄. Then each Z ×X Spec L is an open subscheme of the base change
ZL = Z ×X XL , where XL = Spec AL and AL is the integral closure of A in L. We claim
ZL is connected. If ZL = Z1 ∪ Z2 for two closed subsets Z1 , Z2 , then the special fibre
Zx = Z ×X Spec k lies in one of them, say Z1 . Under the map ZL → XL , Zx is taken
to Yx , so Z2 must be open, contradicting the assumption that Y → X, and therefore also
ZL → XL , is proper and in particular closed. Thus ZL is connected, so Z ×X Spec L is as
well. Since L was arbitrary and the inverse limit of connected spaces is connected, we have
shown that Zη̄ is connected. Now (2) of Proposition 31.3.4 implies that π1 (Yη̄ , ȳ) → π1 (Y, ȳ)
is surjective and in particular it is surjective on prime-to-p parts.
For injectivity (on prime-to-p parts), Corollary 31.3.7 says that it’s enough to show that
every finite étale cover Z 0 → Yη̄ of prime-to-p degree is the base change of some cover Z → Y
of prime-to-p degree. In fact, the remark at the end of Section 31.3 implies that we can check
the condition for Galois covers Z 0 → Yη̄ . Applying Theorem 31.5.1, we have an isomorphism
π1 (Y ×X XL , z̄) ∼= π1 (Y, z̄) so replacing X with XL , we may assume Z 0 is obtained by base
change from a finite Galois cover Zη → Yη . Let Y be the normalization of Y in the function
field K(Zη ) of Zη . Then K(Zη )/K(Y ) is a finite Galois extension, say of degree d, and by the
Zariski-Nagata purity theorem (31.1.6), it’s enough to find a finite Galois extension K 0 /K
such that Y ×X XK 0 → Y ×X XK 0 is étale over all codimension 1 points. Further, since all
codimension 1 points of Y lie in Yη except for the generic point ξ of Yx , we need only check
that Y ×X XK 0 is étale over every point of Y ×X XK 0 lying over ξ. If A has maximal ideal
m = (π), then π also generates the √ maximal ideal of the local ring OY,ξ . It follows from
Kummer theory that K 0 = K(Y )( d π) is a finite Galois extension of K(Y ) of degree d and

681
31.5. Results in Characteristic p Chapter 31. Fundamental Groups of Schemes

since A is a DVR, ramification theory implies that there is a unique point ξ 0 ∈ YK 0 lying
above ξ, with inertia group I(ξ 0 | ξ) ∼= Z/dZ. Now Abhyankar’s lemma shows that YK(Zη )K 0
0 ∼
is étale over ξ . Moreover, YK(Zη )K 0 = Y ×X XK 0 by construction. Finally, this shows that
the base change of Z := Y ×X XK 0 → Y to Yη̄ is precisely Z 0 , completing the proof.
Now let X be any locally noetherian scheme and Y → X a proper morphism with
connected geometric fibres. Suppose x0 , x1 ∈ X such that x0 is a specialization of x1 , i.e.
x0 ∈ {x1 }. Let x̄0 , x̄1 be geometric points of X with images x0 , x1 , respectively, and fix
geometric points ȳ0 , ȳ1 ∈ Y lying over x̄0 , x̄1 , respectively. Set Z = {x1 }, consider the local
ring OZ,x0 and note that the geometric fibres Yx̄0 and Yx̄1 over X remain the same over
Spec OZ,x0 . Denote by A the completion of the localization at a height 1 prime ideal of
OZ,x0 . Then A is a complete DVR and we have a specialization map

spy1 : π1 (Yη̄ , ȳ1 ) −→ π1 (Yx̄1 , ȳ1 )

for the generic point η̄ of Spec A. On the other hand, by Corollary 31.4.9, the morphism
Yη̄ → Yx̄0 induces an isomorphism

π1 (Yη̄ , ȳ0 ) −→ π1 (Yx̄0 , y¯0 ).

Definition. The specialization map from ȳ0 to ȳ1 is the composition

π1 (Yx̄1 , ȳ1 ) ∼
= π1 (Yη̄ , ȳ0 ) −→ π1 (Yx̄0 , ȳ0 ).

Theorem 31.5.4. Suppose ϕ : Y → X is a proper flat morphism over a locally noethe-


rian scheme X with connected geometric fibres. Then for any geometric points ȳ0 , ȳ1 of Y
lying over x0 , x1 ∈ X with x0 ∈ {x1 }, the specialization map from ȳ0 to ȳ1 descends to an
isomorphism of prime-to-p parts

π1 (Yx̄1 , ȳ1 ) −→ π1 (Yx̄0 , ȳ0 ).

Grothendieck used this specialization result to prove the following analogue of Theo-
rem 31.4.6.

Theorem 31.5.5. Let X be an integral, proper, normal curve of genus g over an algebraically
closed field k of characteristic p > 0. Then for any geometric point x̄ : Spec k → X,

(1) π1 (X, x̄) is finitely generated as a profinite group.


0
(2) π1 (X, x̄)(p ) is isomorphic to the profinite prime-to-p completion of the topological fun-
damental group

Πg = ha1 , b1 , . . . , ag , bg | [a1 , b1 ] · · · [ag , bg ] = 1i.

Corollary 31.5.6. Let k be an algebraically closed field of characteristic p > 0, X a smooth,


connected, projective scheme over k and x̄ : Spec k → X a geometric point. Then π1 (X, x̄)
is finitely generated.

682
31.5. Results in Characteristic p Chapter 31. Fundamental Groups of Schemes

Proof. Same as the proof in characteristic 0, using the same reduction to the case of curves
as in Theorem 31.4.3.
Example 31.5.7. This result is false in characteristic p when X is not assumed to proper.
For example, Raynaud proved that for X = A1k = P1k r {∞}, π1 (X, x̄) is infinitely generated.
(Compare this to Example 29.3.10.) Any Artin-Schreier equation y p − y = f (x), where
f (x) ∈ k[x] has prime-to-p degree, defines a nontrivial Galois cover Yf → P1k that is étale
over A1k and has Galois group Z/pZ. If k is algebraically closed, this gives an infinitely family
of finite Galois covers of A1k . More generally, π1 (X, x̄) is infinitely generated for any affine
curve X over k.
Definition. Let ϕ : Y → X be a finite separable morphism of normal integral schemes.
Then ϕ is tamely ramified at a codimension 1 point P ∈ X if ϕ is ramified at P
and for each Q ∈ ϕ−1 (P ), the ramification index e(Q | P ) of ϕ at Q is relatively prime to
the characteristic of κ(P ). We say ϕ is tamely ramified, or simply tame, if it is tamely
ramified at every codimension 1 point of X.
Let X be an integral normal scheme, let U ⊆ X be an open subscheme and denote by
FéttU
the full subcategory of FétU consisting of finite separable covers ϕ : Y → X which are
tamely ramified over X r U . Note that the fibre functor Fibx̄ restricts to a functor on FéttU .
Definition. For a geometric point x̄ : Spec Ω ,→ X with image x ∈ U , the tame funda-
mental group of U at x̄ is
π1t (U, x̄) := Aut(Fibx̄ |FéttU ).
Theorem 31.5.8. Let x̄ be a geometric point in X and U ⊆ X an open subscheme containing
x̄. Then
(1) π1t (U, x̄) is a quotient of π1 (U, x̄) and in particular a profinite group and its action on
Fibx̄ (Y ) is continuous for all Y ∈ FéttU .
(2) The restriction of Fibx̄ to FéttU induces an equivalence of categories

FéttU −→ {finite, continuous, left π1t (U, x̄)-sets}
Y 7−→ Fibx̄ (Y ).

(3) Let KUt be the compositum inside a fixed algebraic closure of K = k(U ) of all finite
extensions L/K such that the normalization XL of X in L is étale over U and tamely
ramified over X r U .
(4) The tame fundamental group π1t (U, x̄) is finitely generated as a profinite group.
Statements (1) – (3) can be derived from similar arguments as for the full étale funda-
mental group. The proof of (4) relies on the deep fact that tame covers in characteristic
p lift to p-tame covers in characteristic 0, where a p-tame cover is one defined over a field
of characteristic 0 in which p does not divide the order of any inertia groups of the cover.
However, this more or less allows one to compute the tame fundamental group by lifting to
characteristic 0.
In contrast, the pro-p-part of the fundamental group in characteristic p is not completely
understood and is the subject of much current research. But we will end the discussion here.

683
Part VII

Étale Cohomology

684
Chapter 32

Introduction

Part VII is an introduction to étale cohomology, which I wrote as part of my dissertation


research in 2017 – 2020 at the University of Virginia. The main texts I use for reference are
Milne’s Lectures on Etale Cohomology and Etale Cohomology, as well as Arapura’s Introduc-
tion to Etale Cohomology. Some of the topics covered are:

ˆ Grothendieck topologies

ˆ Sheaves and cohomology on the étale site

ˆ Galois cohomology

ˆ Cohomology of curves

ˆ Étale versions of standard theorems in algebraic topology, including:

– The Gysin sequence


– Finiteness theorems
– The comparison theorems
– Künneth formulas
– Poincaré duality
– The Lefschetz fixed-point theorem

ˆ The Weil conjectures.

685
Chapter 33

Sites

This chapter covers the basic definitions and results in Grothendieck’s theory of sites, a
useful generalization of a topological space. The main motivation is to develop a working
sheaf theory on schemes that can detect the features of étale morphisms and more general
properties.

686
33.1. Grothendieck Topologies and Sites Chapter 33. Sites

33.1 Grothendieck Topologies and Sites


To every topological space X, we can associate a category Top(X) consisting of the open
subsets U ⊆ X with morphisms given by inclusions of open sets U ,→ V . A presheaf on X
is a functor F : Top(X)op → Set, i.e. a contravariant functor on the category Top(X). The
conditions for F to be a sheaf on X can beS summarized by saying that for every open set
U ∈ Top(X) and every open covering U = Ui , the set F (U ) is an equalizer in the following
diagram:
Y Y
F (U ) F (Ui ) F (Ui ∩ Uj )
i i,j

This generalizes as follows.


Definition. A Grothendieck topology on a category C is a set of collections of morphisms
Cov(X) = {{Xi → X}i } for every objects X ∈ C, called coverings, satisfying:
(i) Every isomorphism X 0 → X defines a covering {X 0 → X} in Cov(X).
(ii) For any covering {Xi → X} of X and any morphism Y → X in C, the fibre products
Xi ×X Y exist and the induced maps {Xi ×X Y → Y } are a covering of Y .
(iii) If {Xi → X}i is a covering of X and {Yij → Xi }j is a covering of Xi for each i,
then the compositions {Yij → Xi → X}i,j are a covering of X.
A category equipped with a Grothendieck topology is called a site.
Example 33.1.1. For a topological space X, the category Top(X) is a site with coverings
( )
[
Cov(U ) = {Ui ,→ U } : Ui ⊆ U are open and U = Ui .
i

When X is a scheme with the Zariski topology, Top(X) is called the (small) Zariski site on
X.
Example 33.1.2. The category Top of all topological spaces with continuous maps between
them is a site, called the big topological site, whose coverings are defined by
( )
[
Cov(X) = {fi : Xi ,→ X} : fi is an open embedding and X = Xi .
i

Example 33.1.3. Similarly, for a scheme X, let SchX be the category of X-schemes (the
category Sch of all schemes can be viewed in this framework by setting X = Spec Z since
this is a terminal object in Sch). Then SchX is a site, called the big Zariski site on X, with
coverings
( )
[
Cov(Y ) = {ϕi : Yi → Y } : ϕi is an open embedding and Y = Yi .
i

687
33.1. Grothendieck Topologies and Sites Chapter 33. Sites

Example 33.1.4. Let C be a site and X ∈ C be an object. Define the localized site (or
the slice category) C/X to be the category with objects Y → X ∈ HomC (Y, X), morphisms
Y → Z in C such that
Y Z

commutes. Then C/X can be equipped with a Grothendieck topology by defining


Cov(Y → X) = {{Yi → Y } : Yi → Y ∈ HomX (Yi , Y ), {Yi → Y } ∈ CovC (Y )}.
Example 33.1.5. Let X be a scheme and define the (small) étale site on X to be the
category Ét(X)
` of X-schemes with étale morphisms Y → X and covers {Yi → Y } ∈ Cov(Y )
such that Yi → Y is surjective.
Example 33.1.6. In contrast, we can equip the slice category Sch/X with a Grothendieck
topology
` by declaring {Yi → Y } to be a covering of Y → X if each Yi → Y is étale and
Yi → Y is surjective. The resulting site is referred to as the big étale site on X.
Example 33.1.7. Similar constructions can be made by replacing “étale” with other prop-
erties, such as:
ˆ The fppf site is the category Sch/X with coverings {Y `i → Y } ∈ Cov(Y ) such that
Yi → Y are flat and locally of finite presentation and Yi → Y is surjective.
ˆ The lisse-étale site LisÉt(X) is the category of X-schemes with smooth morphisms
between them,
` whose coverings are {Yi → Y } ∈ Cov(Y ) such that the Yi → Y are
étale and Yi → Y is surjective.
ˆ The smooth site Sm(X) is the category of X-schemes with smooth morphisms between
them and surjective families of smooth coverings.
ˆ Most generally, the flat site is Sch/X with surjective families of flat morphisms of finite
type as coverings.
Definition. A continuous map between sites f : C1 → C2 is a functor F : C2 → C1
that preserves fibre products and takes coverings in C2 to coverings in C1 .
Remark. Notice that a continuous map between sites is a functor in the opposite direction.
This is in analogy with the topological notion: a continuous map f : X → Y between
topological spaces induces a functor F : Top(Y ) → Top(X) given by V 7→ f −1 (V ).
Example 33.1.8. When X is a scheme, there are continuous maps between the various sites
we have defined on Sch/X. We collect some of these sites in the following table, along with
their relevant features. (The arrows between sites represent continuous maps between sites,
so the functors on the underlying categories go in the opposite direction. Note that when
we define sheaves in the next section, sheaves will pull back in the same direction as these
arrows.)

688
33.1. Grothendieck Topologies and Sites Chapter 33. Sites

Xf lat → Xf ppf → Xsmooth → Xét → XN is → XZar

name flat fppf smooth étale Nisnevich Zariski


étale, with
flat,
maps all smooth étale residue field all
locally f.p.
isomorphisms
Example 33.1.9. Let G be a profinite group and let CG be the category of ` all finite,
discrete G-sets. Then the collections of G-homomorphisms {Xi → X} such that i Xi → X
is surjective form a Grothendieck topology on CG . When G = Gal(k̄/k) for some field k, the
category CG is equivalent to Xét for X = Spec k.

689
33.2. Sheaves on Sites Chapter 33. Sites

33.2 Sheaves on Sites


In this section we generalize the notions of presheaf and sheaf to an arbitrary category with
a Grothendieck topology.
Definition. A presheaf on a category C is a functor F : Cop → Set, that is, a contravariant
functor from C to the category of sets. The category of presheaves on C (with natural
transformations between them) will be denoted PreShC .

` We say F is separated if for every collection of maps {Xi → X}, the map
Definition.
F (X) → i F (Xi ) is injective.
Definition. Let C be a site. A sheaf on C is a presheaf F : Cop → Set such that for every
object X ∈ C and every covering {Xi → X} ∈ Cov(X), the sequence of based sets
Y Y
F (X) F (Xi ) F (Xi ×X Xj )
i i,j

is exact, or equivalently, F (X) is an equalizer in the diagram. The category of sheaves on C


will be denoted ShC .
As in topology, we can consider sheaves on C with values in set categories with further
structure, e.g. Group, Ring,R−Mod, Algk .
Theorem 33.2.1 (Sheafification). The forgetful functor ShC → PreShC has a left adjoint
F 7→ F a .
Proof. First consider the forgetful functor SepC → PreShC defined on the subcategory of
separated presheaves on C. For a presheaf F on C, let F sep be the presheaf

X 7−→ F sep (X) := F (X)/ ∼

where, for a, b ∈ F (X), a ∼ b if there is a covering {Xi → X} of X such that a and b have
the same image under the map a
F (X) → F (Xi ).
i

By construction, F sep
is a separated presheaf on C and for any other separated presheaf F 0 ,
any morphism of presheaves F → F 0 factors through F sep uniquely. Hence F 7→ F sep is left
adjoint to the forgetful functor SepC → PreShC so it remains to construct a sheafification of
every separated presheaf on C.
For a separated presheaf F , define F a to be the presheaf

X 7−→ F a (X) := ({Xi → X}, {αi })/ ∼

where {Xi → X} ∈ CovC (X), {αi } is a collection of elements in the equalizer


 
Y Y
Eq  F (Xi ) F (Xi ×X Xj )  ,
i i,j

690
33.2. Sheaves on Sites Chapter 33. Sites

and ({Xi → X}, {αi }) ∼ ({Yj → Y }, {βj }) if αi and βj have the same image in F (Xi ×X Yj )
for all i, j. Then as above, F a is a sheaf which is universal with respect to all morphisms of
sheaves F → F 0 . Thus F 7→ F a defines a left adjoint to the forgetful functor ShC → SepC
and composition with the first construction proves the theorem.

Definition. A topos is a category T which is equivalent to ShC for some site C. The plural
of ‘topos’ is ‘topoi’.

Remark. Grothendieck held the view that the topos of a site is more important than the site
itself. In practice, many naturally occurring sites induce the same topos (that is, their sheaf
theories are equivalent). For example, the Zariski sites on Sch and AffSch (the categories
of schemes and affine schemes, respectively) induce the same topos. In general, AffSch is
easier to study since it is equivalent to the opposite category of commutative rings.

Definition. A morphism of topoi f : T → T 0 is an adjoint pair f = (f ∗ , f∗ ) of functors


f∗ : T → T 0 and f ∗ : T 0 → T such that the functors HomT (f ∗ (−), −) and HomT 0 (−, f∗ (−))
are naturally isomorphic.

Definition. Let T be a topos and A ∈ T an object. Then A is called a group object (or
a group in T ) if it possesses distinguished morphisms

m : A × A → A, e : ∗ → A, i:A→A

satisfying the following axioms:

(i) (Associativity) The diagram

1A × m
A×A×A A×A

m × 1A m

A×A m A

commutes.

(ii) (Identity) The composition m ◦ (1A × e) : A → A × A → A is equal to 1A .

(iii) (Inverses) The diagram

(i, 1A )
A A×A

∗ A
e

691
33.2. Sheaves on Sites Chapter 33. Sites

commutes.

In addition, A is called abelian if the diagram

A×A
m
p A
m
A×A

commutes, where p flips the factors of A × A. In this case, we will often write the map m
as a to denote addition.

Definition. An abelian group A in a topos T is called a ring if, in addition to a : A × A →


A, e : ∗ → A and i : A → A, it possesses morphisms

m : A × A → A and 1:∗→A

satisfying the usual axioms of associativity, identity and left and right distributivity. Further,
if m satisfies the axiom of commutativity, A will be called a commutative ring in T .

Remark. To make this more compatible with the language of categorical homotopy theory,
one might instead call a ring object a monoid or monoid object in T , meaning with respect
to the symmetric monoidal structure on T induced by ×.

Definition. A ringed topos is a pair (T , A) where T is a topos and A is a ring in T . A


morphism of ringed topoi is a pair (f, f # ) : (T , A) → (T 0 , A0 ) consisting of a morphismi
of topoi f = (f ∗ , f∗ ) : T → T 0 and a morphism of ring objects f # : A0 → f∗ A, or equivalently
by adjointness, f ∗ A0 → A.

Proposition 33.2.2. For every continuous map of sites f : C0 → C, where C and C0


are small categories, there exists a morphism of topoi f∗ : ShC → ShC0 with left adjoint
f ∗ : SchC0 → SchC .

Proof. Let T = SchC , T 0 = SchC0 and define the morphism f∗ : T → T 0 for each object
X 0 ∈ C0 and sheaf F ∈ T by
(f∗ F )(X 0 ) = F (f (X 0 )).
If {Xi0 → X 0 } is a covering in C0 , we have a commutative diagram
Y Y
(f∗ F )(X )0 (f ∗ F )(X 0
i ) (f∗ F )(Xi0 ×X 0 Xj0 )
i i,j

= = ∼
=
Y Y
0
F (f (X )) F (f (Xi0 )) F (f (Xi0 ) ×f (X 0 ) f (Xj0 ))
i i,j

692
33.2. Sheaves on Sites Chapter 33. Sites

Here, the bottom row is exact since F is a sheaf and f is continuous; the right vertical
arrow is an isomorphism since f is continuous; and the other vertical arrows are equalities
by definition. Hence by the Five Lemma, the top row is exact, so f∗ F defines a sheaf on C0 .
To define the left adjoint f ∗ : ShC0 → ShC , take an object X ∈ C and define a category
IX with objects (X 0 , ρ) where X 0 ∈ C0 and ρ ∈ HomC (X, f (X 0 )), and with morphisms
(X 0 , ρ) → (Y 0 , σ) given by a morphism g : X 0 → Y 0 in C0 making the following diagram
commute:
f (X 0 )
ρ
X f (g)
σ
f (Y 0 )

Now define f ∗ F for a sheaf F ∈ T 0 on an object X ∈ C by

(f ∗ F )(X) = lim F (X 0 )
−→

op
where the limit is over all objects (X 0 , ρ) in the opposite category IX . If h : X → Y is a
∗ ∗
morphism in C, then there is an induced map (f F )(Y ) → (f F )(X) given by the functor

IY −→ IX
(Y 0 , ρ) 7−→ (Y 0 , ρ ◦ h).

This shows that f ∗ F is a presheaf on C. Moreover, the maps

(f ∗ f∗ F )(X) = lim F (f (X 0 )) → F (X)


−→

for each F ∈ T , X ∈ C give a natural transformation

HomT 0 (F, f∗ G) −→ HomT (f ∗ F, G).

One can show that it is an isomorphism, which establishes that (f ∗ , f∗ ) is an adjoint pair.
Hence we have a morphism of topoi T → T 0 .

Example 33.2.3. Let f : C0 → C be a continuous map of sites. For an object X 0 ∈ C0 ,


consider the presheaf “ represented by X 0 ”:

hX 0 : (C0 )op −→ Set


Y 0 7−→ HomC0 (Y 0 , X 0 ).

It’s easy to see that f ∗ hX 0 ∼


= hf (X 0 ) as functors, that is, the induced morphism of topoi

(f , f∗ ) commutes with representable functors.

693
33.3. The Étale Site Chapter 33. Sites

33.3 The Étale Site


Let Xét denote the étale site on a scheme X. Fix a faithfully flat morphism ϕ : Y → X and
a group G acting on the morphism on the right via α : G → AutX (Y ).
Definition. We say ϕ : Y → X is a Galois cover with Galois group G, or a G-cover for
short, if the morphism

Y × G −→ Y ×X Y
(y, g) 7−→ (y, yg)

is an isomorphism.
Lemma 33.3.1. ϕ : Y → X is a G-cover if and only if ϕ is surjective, finite, étale and
deg ϕ = |G|.
Definition. A Galois cover ϕ : Y → X is said to be generically Galois if k(Y )/k(X) is
a Galois extension of fields.
Example 33.3.2. Let A be a ring, B an A-algebra and consider the corresponding morphism
of affine schemes
ϕ : Spec B −→ Spec A.
Then ϕ is a Galois cover with Galois group G if and only if A → B is faithfully flat and G
acts on B such that
Y
B ⊗A A −→ B × G = B
g∈G
0 0
b ⊗ b 7−→ (bgb )g

is an isomorphism.
Example 33.3.3. Let k be a field and f ∈ k[t] a monic irreducible polynomial. Set K =
k[t]/(f ). Then f = f1e1 · · · frer in K[t] and by the Chinese remainder theorem,
r
K ⊗k K ∼
= K[t]/(f ) ∼
Y
= K[t]/(fiei ).
i=1

Thus Spec K → Spec k is a Galois cover if and only if each fi is linear, fi 6= fj for any i 6= j
and ei = 1 for all i. That is, Spec K → Spec k is Galois if and only if f is separable with
splitting field K, just as in classical Galois theory.
Proposition 33.3.4. Suppose ϕ : Y → X is a G-cover and F is a presheaf on the étale
site Xét taking disjoint unions to products. Then F is a sheaf on Xét if and only if F (ϕ) :
F (X) → F (Y ) is an isomorphism onto the fixed set F (Y )G ⊆ F (Y ).
Proof. Consider the two maps Y × G Y given by (y, g) 7→ y and (y, g) 7→ yg. These
fit into a commutative diagram with the two coordinate projections Y ×X Y Y :

694
33.3. The Étale Site Chapter 33. Sites

Y ×G Y X


= id id

Y ×X Y Y X

Applying F to the diagram, we obtain a commutative diagram of sets


F (X) F (Y ) F (Y ×X Y )

=
id id
Y
F (X) F (Y ) F (Y )
g∈G

where the maps F (Y ) are given by s 7→ (s)g and s 7→ (gs)g . Then these
Q
g∈G F (Y )
maps agree precisely when gs = s for all g ∈ G, i.e. F (X) is the equalizer in the top row if
and only if it identifies with F (Y )G in the bottom row.

Proposition 33.3.5. Suppose F is a presheaf on Xét which satisfies the condition that
Y Y
F (U ) F (Ui ) F (Ui ×U Uj )
i i,j

is an equalizer diagram for all covers {Ui → U } of U ∈ XZar in the Zariski site on X and
for all étale affine covers {V → U } of U ∈ Xét consisting of a single morphism in the étale
site. Then F is a sheaf on Xét .
`
Proof. If U = i Ui for schemes Ui ∈ Xét , then the first condition implies that F (U ) =
a covering {Ui → U 0 } in Xét , the sequence
Q
i F (Ui ). Thus for
Y Y
0
F (U ) F (U i ) F (Ui ×U 0 Uj )
i i,j

is isomorphic to the sequence


F (U 0 ) F (U ) F (U ×U 0 U )

for the covering {U → U 0 }, using the fact that ( i Ui ) ×U 0 ( i Ui ) = i,j (Ui ×U 0 Uj ). Since
` ` `
the equalizer condition is assumed to hold for all étale affine covers, this argument shows the
condition holds for all {Ui → U }i∈I with I finite `and each Ui affine.
Let {Uj → U }j be a covering and set U 0 = j Uj and f : U 0 → U . Write U = i Vi
S

for open affine subschemes Vi ⊆ U and for each i, write f −1 (Vi ) = k Wik for open affine
S
subschemes Wik ⊆ U 0 . Fix one of the Vi . Then each f (Wik ) is open in Vi , so by quasi-
compactness, we may reduce to a finite cover {Wik → Vi }K k=1 . Now consider the diagram

695
33.3. The Étale Site Chapter 33. Sites

F (U ) F (U 0 ) F (U 0 ×U U 0 )

Y YY YY
F (Vi ) F (Wik ) F (Wik ×U Wi` )
i i k i k,`

Y YY
F (Vi ×U Vj ) F (Wik ×Vi Wj` )
i,j i,j k,`

The two columns correspond to the coverings {Vi → U }i and {Wik → U 0 }i,k which are
all coverings in the Zariski site on X and hence these columns are exact by hypothesis.
Moreover, the middle row corresponds to the coverings {Wik → f (Wik ) ⊆ Vi }k which for
each i is finite and affine, so by the above paragraph this row is exact. An easy diagram
chase then implies the top row of the diagram is also exact, which is what we want.

Example 33.3.6. Let A → B be a ring homomorphism such that Spec B → Spec A is


surjective and étale. In particular, A → B is faithfully flat and unramified. We claim that
the sequence

0 → A → B → B ⊗A B (*)

is exact, where the second map is b 7→ 1 ⊗ b − b ⊗ 1. First note that the map g : B →
B ⊗A B, b 7→ b ⊗ 1 has a section s : b ⊗ b0 7→ bb0 . Then consider

h : (B ⊗A B) ⊗B (B ⊗A B) −→ B ⊗A B, x ⊗ y 7−→ xgs(y).

We have h(1 ⊗ x − x ⊗ 1) = gs(x) − x, so if 1 ⊗ x − x ⊗ 1 = 0, we get x = gs(x) ∈ im g and


the sequence
g
0→B→ − B ⊗A B → (B ⊗A B) ⊗B (B ⊗A B)
is exact. Tensoring (∗) with B induces the vertical arrows in the following diagram:

0 A B B ⊗A B

0 B B ⊗A B (B ⊗A B) ⊗B (B ⊗A B)

Therefore the top row is exact as claimed.

696
33.3. The Étale Site Chapter 33. Sites

Example 33.3.7. Let X be a scheme and Y → X an étale morphism. Set OXét (Y ) =


Γ(Y, OY ), which defines a sheaf OXét for the Zariski topology by standard calculations. To
check that OXét is a sheaf for the étale site, it suffices to check the conditions of Proposi-
tion 33.3.5, but the Zariski condition was just seen to hold. If {Y → Z} is an étale affine
covering in Xét , with Y = Spec B and Z = Spec A, then the corresponding ring map A → B
satisfies the condition of Example 33.3.6, meaning

0 → A → B → B ⊗A B

is exact and hence so is the sequence

OXét (Z) → OXét (Y ) → OXét (Y ×Z Y ).

(Note that this is precisely the same as the equalizer condition since the map B → B ⊗A B
is b 7→ 1 ⊗ b − b ⊗ 1 and Γ(Y, OY ) are abelian groups.)
Example 33.3.8. Any scheme Z → X defines a presheaf

FZ : Xét −→ Set
Y 7−→ HomX (Y, Z).

Then FZ is a sheaf for XZar so once again, to show it is a sheaf for the étale topology, by
Proposition 33.3.5 it will suffice to check the sheaf condition for single étale affine covers.
For such a cover {Spec B → Spec A}, the map A → B is faithfully flat with

0 → A → B → B ⊗A B

exact. Take an open affine subscheme U = Spec C ,→ Z. Applying HomA (C, −) to the
above sequence gives

HomA (C, A) → HomA (C, B) → HomA (C, B ⊗A B)

which is exact since HomA (C, −) is a left-exact functor. Generalizing to HomX (−, Z) is
straightforward using patching, so ultimately we conclude that FZ is a sheaf on Xét . As
usual, by Yoneda’s Lemma the assignment Z 7→ FZ is injective so we will often write FZ
simply by Z.
Example 33.3.9. Let n ≥ 1 be an integer and let µn be the group scheme defined (locally)
by tn − 1 = 0. Then for any Y ∈ Xét , µn (Y ) coincides with the set of nth roots of unity in
Γ(Y, OY ).
Example 33.3.10. Let X be a k-scheme and let Ga be the affine group scheme defined by
the additive group of k. Then for each Y ∈ Xét , Ga (Y ) = Γ(Y, OY ).
Example 33.3.11. Similarly, when Gm is the affine multiplicative group scheme defined by
k × , then for each Y ∈ Xét , Gm (Y ) = Γ(Y, OY )× .
Example 33.3.12. When char k = p > 0, let αp denote the group scheme defined by tp = 0.
Note that αp is not an étale group scheme (but it is flat). For Y ∈ Xét , αp (Y ) corresponds
to the set of nilpotent elements in Γ(Y, OY ).

697
33.3. The Étale Site Chapter 33. Sites

Example 33.3.13. Consider the ring k[ε] = k[t]/(t2 ). Write T = Spec k[ε]. The functor
T = Homk (−, T ) is called the (étale) tangent space functor since for any Y ∈ Xét , T (Y ) is
the tangent space to Y , which is locally given by T (Y )x = Tx Y := (mx /m2x ).

Example 33.3.14. Let R be a set and let FR be the sheaf on Xét defined by
Y
FR : Y 7−→ FR (Y ) := R
π0 (Y )

where π0 (Y ) denotes the set of connected components of Y . Then FR is called the constant
sheaf on Xét associated to R.

Example 33.3.15. Let M be a sheaf of coherent OX -modules on the Zariski site XZar .
This gives us an étale sheaf Mét as follows. If ϕ : Y → X is an étale morphism, then ϕ∗ M
is a coherent OY -module on YZar which on affine patches U = Spec A ⊆ X, V = Spec B ⊆ Y
takes the form
M ⊗A B M

B A

Let Mét be the presheaf Y 7→ Γ(Y, ϕ∗ M) on Xét . By a similar proof to the one in Exam-
ple 33.3.7 for OXét , one can show that Mét is then a sheaf on Xét . As a special case, note
that (OXZar )ét = OXét .

Example 33.3.16. Let X be a k-scheme, ϕ : Y → X a morphism and consider the exact


sequence of sheaves
ϕ∗ Ω1X/k → Ω1Y /k → Ω1Y /X → 0.
If ϕ is an étale morphism, then by Proposition 14.4.17, Ω1Y /X = 0 so ϕ∗ Ω1X/k → Ω1Y /k is
surjective. In fact, both are locally free sheaves of the same rank, so ϕ∗ Ω1X/k ∼
= Ω1Y /k . It
follows that (Ω1X/k )ét |YZar = Ω1Y /k .

Example 33.3.17. Let k be a field, X = Spec k and consider the étale site Xét . If G-Modd
denotes the category of discrete G-modules, then there is an equivalence of categories

Sh(Xét ) ←→ G-Modd
F 7−→ MF
FM →−7 M

where MF = lim F (L) is the direct limit over all finite, Galois extensions L/k and FM is the
−→
sheaf A 7→ HomG (Homk (A, k sep ), M ).

We next describe the category Sh(Xét ). To begin, first consider the category Presh(Xét )
of presheaves on the étale site.

698
33.3. The Étale Site Chapter 33. Sites

Lemma 33.3.18. Presh(Xét ) is an abelian category.


Let F 0 → F → F 00 be a sequence of presheaves on Xét . Then this sequence is exact if
and only if the sequence
F 0 (Y ) → F (Y ) → F 00 (Y )
is exact for all étale morphisms Y → X. Let Sh(Xét ) be the full subcategory of Presh(Xét )
of sheaves of abelian groups on the étale site Xét . Then Sh(Xét ) is an additive category; we
will prove that it is abelian.
Definition. A morphism of sheaves T : F → F 0 on Xét is locally surjective if for every
Y ∈ Xét and s ∈ F 0 (Y ), there exists a covering {Yi → Y } in the étale topology such that for
each i, s|Yi lies in the image of F (Yi ) → F 0 (Yi ).
Proposition 33.3.19. For a morphism of sheaves T : F → F 0 on the étale site, the following
are equivalent:
T
→ F 0 → 0 is exact.
(a) F −

(b) T is locally surjective.

(c) For every geometric point x̄ : Spec k̄ → X, the map on stalks Tx̄ : Fx̄ → Fx̄0 is
surjective.
Proposition 33.3.20. For a sequence of sheaves 0 → F 0 → F → F 00 → 0 in Sh(Xét ), the
following are equivalent:
(a) The sequence is exact.

(b) For all Y ∈ Xét , the sequence 0 → F 0 (Y ) → F (Y ) → F 00 (Y ) → 0 is exact.

(c) For all geometric points x̄ : Spec k̄ → X, the sequence of stalks 0 → Fx̄0 → Fx̄ → Fx̄00 →
0 is exact.
Corollary 33.3.21. For any scheme X, Sh(Xét ) is an abelian category.
Example 33.3.22. Let n ∈ N and assume n is invertible on X, i.e. char k - n for any residue
fields k of X. Consider the following sequence of sheaves, called the Kummer sequence for
X:
n
0 → µn →
− Gm → − Gm → 0
n
where Gm → − Gm is the morphism induced by t 7→ tn . We claim the Kummer sequence is
exact. By Proposition 33.3.20, it suffices to show exactness on stalks at geometric points.
For such a point x̄, set A = OX,x̄ . Then µn,x̄ = µn (A) and Gm,x̄ = A× and moreover,
0 → µn (A) → A× is clearly exact. For the right map, notice that tn − a splits in A[t] for
every a ∈ A× , since dtd (tn − a) = ntn−1 6= 0 when n is invertible on X. Thus every a is an
nth root in A× , and it follows that the sequence
n
− A× →
0 → µn (A) → − A× → 0

is exact as required.

699
33.3. The Étale Site Chapter 33. Sites

Example 33.3.23. When char k | n for some residue field k of X, the above example fails
since étale locally, we have
d p
(t − a) = ptp−1 = 0 in characteristic p > 0.
dt
Note however that the equation t 7→ tp does define a (locally) flat covering Gm → Gm , so
the sequence
p
0 → µp →− Gm →− Gm → 0
is exact in the flat topology on X. On the étale site, the appropriate characteristic p
replacement for the Kummer sequence is the Artin-Schreier sequence

0 → Z/pZ →
− Ga →
− Ga → 0

where Z/pZ is the constant group scheme defined by the same group and ℘ : Ga → Ga is
induced by the map t 7→ tp − t. To see that the AS sequence is exact, it suffices once again to
check exactness on stalks at geometric points. For a geometric point x̄, again let A = OX,x̄
so that Z/pZx̄ = Z/pZ(A) and Ga,x̄ = A. As before, 0 → Z/pZ(A) → A is exact and the
kernel of t 7→ tp − t : A → A is precisely Z/pZ(A). For surjectivity, note that for any a ∈ A,

d p
(t − t − a) = ptp−1 − 1 = −1 6= 0
dt
so tp − t − a splits in A[t] for all a ∈ A. Thus the end of the sequence

0 → Z/pZ(A) →
− A→
− A→0

is exact and we are done.

700
Chapter 34

Cohomology

The étale site on a scheme has a sheaf cohomology theory which mimics the ordinary (sin-
gular) cohomology for schemes X/C with their natural complex structure. To define coho-
mology, we must first show that Sh(Xét ) has enough injectives. Along the way, we develop
a few useful algebraic constructions on sheaf categories that will play a role in our study of
cohomology and beyond. We will assume all sheaves have values in abelian groups.

701
34.1. Direct and Inverse Image Functors Chapter 34. Cohomology

34.1 Direct and Inverse Image Functors


Given a morphism of schemes f : Y → X, we have several ways of mapping sheaves on Xét
to sheaves on Yét and vice versa. The simplest to define (although not always the simplest
to understand) is the pushfoward functor.
Definition. For a morphism f : Y → X and a presheaf F on Ét(Y ) (the category of étale
X-schemes), define the direct image presheaf (or pushforward presheaf) of F along
f to be the functor f∗ F : Ét(X)op → Ab sending U 7→ f∗ F (U ) := F (Y ×X U ) for all U → X
étale.
Lemma 34.1.1. Direct image restricts to a functor f∗ : Sh(Yét ) → Sh(Xét ).
Proof. This follows from the axiom (see Section 33.1) that if {Ui → U } is a covering in Xét
then {Y ×X Ui → Y ×X U } is a covering in Yét .
Example 34.1.2. If x̄ : ∗ → X is a geometric point of X, then any sheaf on ∗ is an abelian
group, say A, and x̄∗ A = Ax̄ is a skyscraper sheaf supported at the image of x̄ with stalk A.
This explains the usual notation x̄∗ A for such a skyscraper sheaf.
Proposition 34.1.3. Let f : Y → X be a morphism and x̄ : ∗ → X a geometric point of X
with image x ∈ X. Then for any sheaf F on Yét ,
(1) If f is an open immersion, then (f∗ F )x̄ = Fx̄ if x ∈ Y .
(2) If f is a closed immersion, then
(
Fx̄ , it x ∈ Y
(f∗ F )x̄ =
0, if x 6∈ Y.

(3) If f is finite, then


⊕[k(z):k(x)]s
M
(f∗ F )x̄ = Fz̄
z7→x

where the sum is over all points z ∈ Y mapping to x and [k(z) : k(x)]s denotes the
separable degree of the residue field extension at z. In particular, if f is finite étale of
degree n, then (f∗ F )x̄ = Fȳ⊕n .
Corollary 34.1.4. If f : Y → X is finite or a closed immersion, then f∗ : Sh(Yét ) → Sh(Xét )
is an exact functor.
g f
Proposition 34.1.5. Suppose Z →
− Y →
− X are morphisms of schemes. Then (f ◦ g)∗ =
f∗ ◦ g∗ .
The direct image functor on presheaves is exact, so f∗ : Sh(Yét ) → Sh(Xét ) is left exact –
in general, it is not right exact. Thus we can define:
Definition. Let f : Y → X be a morphism with pushfoward functor f∗ : Sh(Yét ) → Sh(Xét ).
The inverse image functor (or pullback) along f is the left adjoint f ∗ : Sh(Xét ) → Sh(Yét )
to f∗ .

702
34.1. Direct and Inverse Image Functors Chapter 34. Cohomology

Lemma 34.1.6. For a morphism f : Y → X and any sheaf F on Xét , f ∗ F is the sheafifi-
cation of the presheaf f −1 F on Yét defined by
f −1 F (V ) := lim F (U )
−→

where V → Y is étale and the direct limit is over all U → X étale such that
V Y

U X

commutes.
Proof. It is easy to show that f −1 is left adjoint to f∗ on presheaves. Explicitly, for any
presheaf Q on Yét , there are bijections
HomPresh(Yét ) (f −1 F, Q) ←→ lim HomPresh(Xét ) (F (U ), Q(U ×X Y )) ←→ HomPresh(Xét ) (F, f∗ Q)
−→

which are functorial in F and Q. Therefore, passing to the sheafification yields


HomSh(Yét ) ((f −1 F )sh , Q) ∼
= HomSh(Xét ) (F, f∗ Q)
for any sheaf Q on Yét . Since left adjoints are unique, this proves (f −1 F )sh = f ∗ F .
g f
Proposition 34.1.7. For any morphisms Z → − X, (f ◦ g)∗ = g ∗ ◦ f ∗ .
− Y →
Proof. Each of (f ◦ g)∗ and g ∗ ◦ f ∗ is a left adjoint of (f ◦ g)∗ = f∗ ◦ g∗ .
Example 34.1.8. When f : Y → X is an étale morphism, the inverse image f ∗ : Sh(Xét ) →
Sh(Yét ) is just the restriction functor F 7→ F |Yét .
Proposition 34.1.9. For all morphisms f : Y → X, f ∗ is exact and f∗ preserves injectives.
Proof. Let ȳ : Spec k → Y be a geometric point of Y , let x̄ = f ◦ ȳ and denote their images
respectively by y ∈ Y and x ∈ X. Then for any sheaf F ∈ Sh(Xét ), Example 34.1.2 and
Proposition 34.1.7 give us
(f ∗ F )y = ȳ ∗ (f ∗ F ) = (f ◦ ȳ)∗ = x̄∗ F = Fx̄ .
Therefore f ∗ is exact on stalks at geometric points, so it is exact on stalks and therefore
exact by Proposition 33.3.20. The statement for f∗ follows from the more general statement
that the right adjoint of an exact functor preserves injectives.
Theorem 34.1.10. For any scheme X, the category Sh(Xét ) has enough injectives.
Proof. Take a geometric point x̄ : Spec k → X with image x ∈ X and let F be a sheaf on
Xét . Since Fx is an abelian group, there is some injective Ex such that Fx ,→ Ex . Now Ex
defines a sheaf on Q k, necessarily injective, so that by Proposition 34.1.9, x̄∗ Ex is injective.
Therefore E = x∈X x̄∗ Ex is injective and by construction there is a monomorphism of
sheaves F → E.

703
34.1. Direct and Inverse Image Functors Chapter 34. Cohomology

A slight annoyance in Proposition 34.1.3 is that the direct image of a sheaf along an open
immersion is not necessarily supported on the image of the immersion. Indeed, consider the
following example.
Example 34.1.11. Let X = A1 be the affine line over an algebraically closed field k and
consider the open immersion j : A1 r {0} ,→ A1 . Recall (Example 29.3.11) that G :=
π1ét (A1 r {0}, ȳ) ∼
=Zb for any geometric point ȳ. Let S be a module over this fundamental
group, which determines a locally constant sheaf F = FS . Then (j∗ F )0 = S G , the submodule
of G-invariants of S. This shows that in (1) of Proposition 34.1.3, (f∗ F )x need not be 0.
However, we can define a version of the direct image which plays nicely with open im-
mersions.
Definition. Let X be a scheme and j : Y ,→ X an open immersion. For a sheaf F on Yét ,
define the extension along j of F to be the sheafification of the presheaf P on Xét which
sends an étale scheme U → X to
(
F (U ), if im(U ) ⊆ Y
P (U ) =
0, otherwise.
This defines a functor j! : Sh(Yét ) → Sh(Xét ).
Proposition 34.1.12. For any open immersion j : Y ,→ X, j! is exact and j ∗ preserves
injectives.
Proof. We first observe that (j! , j ∗ ) is an adjoint pair: for all F ∈ Sh(Yét ) and G ∈ Sh(Xét ),
HomSh(Xét ) (j! F, G) ∼
= HomPresh(Xét ) (P, G) where P is as above

= HomSh(Yét ) (F, G|Yét )
= HomSh(Yét ) (F, j ∗ G) by Example 34.1.8.
Now take a geometric point x̄ : Spec k → X with image x and note that
(
Fx , if x ∈ Y
(j! F )x =
0, if x 6∈ Y
by definition of j! F . Then the first statement follows from Proposition 33.3.20 and the second
follows from a similar categorical statement as in the proof of Proposition 34.1.9.
Proposition 34.1.13. Let j : Y ,→ X be an open immersion, Z = X r j(Y ) and i : Z ,→ X
the corresponding closed immersion. Then for any sheaf F on Xét , the sequence
0 → j! j ∗ F → F → i∗ i∗ F → 0
is exact.
Proof. By Proposition 33.3.20, it suffices to check this on stalks, where for x ∈ X we have
( id
0 → Fx − → Fx →− 0 → 0, x ∈ Y
id
0→0→
− Fx −
→ Fx → 0, x ∈ Z.

704
34.2. Étale Cohomology Chapter 34. Cohomology

34.2 Étale Cohomology


Let X be a scheme and Sh(Xét ) the abelian category of sheaves on the étale site, which has
enough injectives by Theorem 34.1.10. For a sheaf F on Xét , let Γ(Xét , F ) := F (X) be the
abelian group of global sections of F . Then Γ(Xét , −) : Sh(Xét ) → Ab is a left exact functor,
so we can take its right derived functors to define a cohomology theory on Xét .

Definition. For i ≥ 0, the ith étale cohomology of X with coefficients in a sheaf F is


i
Hét (X, F ) := Ri Γ(Xét , F ).

Explicitly, if F → E • is an injective resolution of F in Sh(Xét ), then


i
Hét (X, F ) = H i (Γ(Xét , E • )).

Note the following properties which are immediate from the definition/the theory of
derived functors:

ˆ For all sheaves F on Xét , Hét


0
(X, F ) = Γ(Xét , F ).

ˆ If E is an injective sheaf, then Hét


i
(X, E) = 0 for all i > 0.

ˆ Every short exact sequence of sheaves 0 → F 00 → F → F 0 → 0 determines a long exact


sequence in étale cohomology:
0
0 → Hét (X, F 00 ) → Hét
0 0
(X, F ) → Hét (X, F 0 ) → Hét
1
(X, F 00 ) → · · ·

ˆ For an étale morphism f : Y → X, the right derived functors of the composite functor
f∗ Γ(Yét ,−) i
Γ(Y, −) : Sh(Xét ) −→ Sh(Yét ) −−−−−→ Ab are Hét (Y, (−)|Y ), which we will simply write
i
Hét (Y, −) when the context is clear.

We now verify that the Eilenberg-Steenrod axioms hold for Hét (X, −), showing it is a

cohomology theory. Note that additivity is guaranteed since Hét is constructed as a right
derived functor.

Theorem 34.2.1 (Dimension). Let x = Spec k be a geometric point, i.e. k is separably


i
closed. Then for any sheaf F on x, Hét (x, F ) = 0 for all i > 0.

Proof. If k is an arbitrary field, then by Example 33.3.17 there is an equivalence of categories



Sh(Spec két ) −→ G-Modd
F 7−→ Fx

where G = Gal(k sep /k). Therefore Hét i


(x, F ) ∼
= H i (G, Fx ), the group cohomology of G with
coefficients in Fx , so when k is separably closed, these are all trivial for i > 0.
Fix a sheaf F on Xét and consider the right derived functors of the left exact functor
HomSh(Xét ) (F, −) : Sh(Xét ) → Ab, which are denoted Exti (F, −).

705
34.2. Étale Cohomology Chapter 34. Cohomology

Example 34.2.2. For the constant sheaf Z on X, there are isomorphisms



HomSh(Xét ) (Z, G) −→ G(X) = Γ(Xét , G)
α 7−→ α(1)

for all G ∈ Sh(Xét ). Therefore Exti (Z, −) ∼ i


= Hét (X, −).

Theorem 34.2.3 (Exactness). Let Z ⊆ X be a closed subscheme and F any sheaf on Xét .
Then there is a long exact sequence
i−1
· · · → Hét (X r Z, F ) → HZi (X, F ) → Hét
i i
(X, F ) → Hét (X r Z, F ) → · · ·

where HZi (X, −) denotes the ith right derived functor of F 7→ ΓZ (X, F ) := ker(Γ(Xét , F ) →
Γ((X r Z)ét , F )). Moreover, the sequence is functorial in the pair (X, X r Z) and in F .

Proof. Let i : Z ,→ X and j : X r Z ,→ X be the (closed and open, resp.) immersions. For
the constant sheaf Z on X, Proposition 34.1.13 gives us a short exact sequence

0 → j! j ∗ Z → Z → i∗ i∗ Z → 0.

For F ∈ Sh(Xét ), applying HomSh(Xét ) (−, F ), we get an exact sequence

0 → Hom(i∗ i∗ Z, F ) → Hom(Z, F ) → Hom(j! j ∗ Z, F ).

The adjointness of (j! , j ∗ ) from Proposition 34.1.12 implies that

HomSh(Xét ) (j! j ∗ Z, F ) = HomSh((XrZ)ét ) (j ∗ Z, j ∗ F ) = F (X r Z) = Γ((X r Z)ét , F ).

Taking derived functors gives Hét i


(X r Z, F ) = Exti (j! j ∗ Z, F ) for all i ≥ 0. Therefore
i
Hom(i∗ i∗ Z, F ) = ΓZ (X, F ), Ext (i∗ i∗ Z, F ) = HZi (X, F ) for all i ≥ 0, and the long exact
sequence in Ext for the above Hom sequence is precisely the long exact sequence for the pair
(X, X r Z) that we are looking for.

Theorem 34.2.4 (Excision). Let π : Y → X be an étale morphism and W ⊆ Y a closed


subscheme such that

(1) Z = π(W ) is closed in X and π|W : W → Z is an isomorphism.

(2) π(Y r W ) ⊆ X r Z.

Then for any sheaf F on Xét , the induced map


i
HZi (X, F ) −→ HW (Y, F |Y )

is an isomorphism for all i ≥ 0.

Proof. The maps in the commutative diagram

706
34.2. Étale Cohomology Chapter 34. Cohomology

Y rW Y W

π ∼
=

X rZ X Z

induce a commutative diagram


0 ΓW (Y, π ∗ F ) Γ(Y, π ∗ F ) Γ(Y r W, π ∗ F )

0 ΓZ (X, F ) Γ(X, F ) Γ(X r Z, F )

By Proposition 34.1.9 and Example 34.1.8, π ∗ is exact and preserves injectives, so it’s enough
to prove the isomorphism for i = 0, that is, to show α is an isomorphism.
Take s ∈ ΓZ (X, F ) and suppose α(s) = 0. Then s|XrZ = 0 by exactness of the bottom
row and α(s) = 0 in Γ(Y, π ∗ F ), but since {Y → X, X r Z → X} is an étale covering of X,
the sheaf axioms imply s = 0. Therefore α is injective.
On the other hand, if t ∈ ΓW (Y, π ∗ F ), we may view it as an element of Γ(Y, π ∗ F ). Let
t0 ∈ Γ(Y, π ∗ F ) be the zero section. Note that (X r Z) ×X Y = π −1 (X r Z) ⊆ Y r W and
we have t|Y rW = 0 by exactness of the top row and t0 |Y rW = 0 because it’s the zero section.
Thus the sheaf condition on π ∗ F says that both t and t0 come from a section s ∈ Γ(X, F )
which must satisfy s|XrZ = 0 by commutativity. Hence by exactness of the bottom row,
s ∈ ΓZ (X, F ) and α(s) = t, proving surjectivity.
Corollary 34.2.5. For any closed point x ∈ X and any sheaf F on Xét , there is an isomor-
phism
i ∼ i h
H{x} (X, F ) −→ H{x} (Spec OX,x ,F)
h
for all i ≥ 0, where OX,x is the henselization of OX,x .
Proof. Let π : U → X be an étale neighborhood of X with u = π −1 (x). Taking Z = {x}
and W = {u}, Theorem 34.2.4 gives us isomorphisms
i ∼ i
H{x} (X, F ) −→ H{u} (U, F |U )
h
for all i ≥ 0. Since OX,x = lim Γ(U, OU ) where the limit is over all étale neighborhoods of x,
−→
and since cohomology commutes with direct limits, we get
i
H{x} h
(Spec OX,x ,F) ∼ i
= lim H{u} (U, F |U ) ∼ i
= H{x} (X, F )
−→

as desired.
Finally, to prove the homotopy axiom, we need a suitable analogue of “homotopic maps”
in the category of schemes. There are various concepts of homotopy that can be introduced
in algebraic geometry, some giving quite different theories, but we will focus on one that
gives an analogue of the homotopy axiom for étale cohomology.

707
34.2. Étale Cohomology Chapter 34. Cohomology

Let k be a field (perhaps algebraically closed), X a scheme over k and suppose Z ⊆ X ×P1
is a closed subscheme whose image under the projection π : X × P1 → P1 is dense. For each
closed point t ∈ P1 , set Z(t) = Z ×X (X × {t}), so that each Z(t) may be viewed as an
algebraic cycle on X. If Z and Z 0 are algebraic cycles on X, we say that they are rationally
equivalent, denoted Z ∼ Z 0 , if there are finitely many such families of algebraic cycles
Z1 (t), Z2 (t), . . . , Zr (t) interpolating Z1 and Z2 in the following sense: for each 1 ≤ i ≤ r − 1,
there are closed points ti , t0i ∈ P1 such that Zi (ti ) = Zi+1 (t0i ), and in addition there are
t, t0 ∈ P1 such that Z = Z1 (t) and Z 0 = Zr (t0 ). We say two morphisms f, g : Y → X are
rationally equivalent if their graphs Γf ∼ Γg as algebraic cycles in X ×k Y .

Theorem 34.2.6 (Homotopy). If f, g : Y → X are rationally equivalent, then f ∗ = g ∗ :


• •
Hét (X, F ) → Hét (Y, F ) for any sheaf F ∈ Sh(Xét ).

Proof. See Milne’s LEC, Ch. 9.

708
Part VIII

Algebraic Spaces and Stacks

709
Chapter 35

Introduction

Part VIII introduces algebraic stacks (following Olsson’s Algebraic Spaces and Stacks) and
is taken from part of my dissertation research in 2017 – 2020 at the University of Virginia.
The following topics are considered:

ˆ Grothendieck topologies

ˆ Sheaves on a site and their cohomology

ˆ Fibred categories and categories fibred in groupoids

ˆ Descent theory and stacks

ˆ Algebraic spaces

ˆ Algebraic stacks

ˆ Deligne-Mumford stacks

ˆ Moduli spaces

ˆ Stacky curves.

For the necessary background in algebraic geometry, the reader can consult Hartshorne’s
Algebraic Geometry.

710
Chapter 36

Categories Fibred in Groupoids

One of the main motivations for Grothendieck’s use of fibred categories in the study of
algebraic spaces and stacks is to allow for the construction of universal objects. Here’s an
example to keep in mind. Suppose f : X → Y is a morphism in the category of topological
spaces (this problem will also arise in a category of schemes). Then for any sheaf F on Y , one
way to define a pullback sheaf f ∗ F on X is as a solution to the universal mapping problem
F → f∗ G of sheaves on Y , where G is a sheaf on X. This object f ∗ F is not unique, it is only
defined up to canonical isomorphism. Similar problems occur where universal objects are
present (any direct limit construction might pose a problem) so we must find a way around.
Here’s a brief discussion of how nontrivial automorphisms can get in the way of solving
a “moduli problem”, which is loosely defined as a classification problem with some natural
notion of geometry attached to the collection of all objects to be classified. A more rigorous
definition is that a moduli problem is a functor F : Schop → Set. If F is representable, say
F ∼= Hom(−, M ) for some scheme M , we say M is a (fine) moduli space for F . In general
no such space exists.
Example 36.0.1. There are many interesting moduli problems in classical geometry which
aren’t represented by schemes. For example, the problem of classifying all circles in R2 is
represented by R2 × R>0 , while circles up to isometry are parametrized by R>0 , neither of
which is a scheme.
Example 36.0.2. In differential geometry, all complex (or equivalently, hyperbolic) struc-
tures on a surface of genus g ≥ 1 are specified by Fenchel-Nielsen coordinates, and these
form a moduli space homeomorphic to R6g−6 .
Example 36.0.3. One can think of algebraic varieties themselves (over an algebraically
closed field) as moduli spaces whose points correspond to solutions to a given set of polyno-
mial equations.
Example 36.0.4. The classical Cayley-Salmon theorem says that the moduli problem of
finding lines on a smooth cubic surface is represented by a 0-dimensional scheme with 27
components.
Example 36.0.5. Let k be a field. The moduli problem of finding r-dimensional vector
subbundles of a rank n bundle on X ∈ Schk is represented by the Grassmannian variety

711
Chapter 36. Categories Fibred in Groupoids

Gr(r, n). In particular, Pnk = Gr(1, n) classifies line bundles inside rank n bundles. The
moduli problem of rank n bundles themselves has the infinite Grassmannian Gr(n, ∞) as a
moduli space.

Example 36.0.6. More generally, there is a scheme BG which is a moduli space for iso-
morphism classes of principal G-bundles.

Example 36.0.7. Let M1,1 : Schop C → Set be the moduli problem parametrizing complex
elliptic curves, i.e. M1,1 (S) is the set of isomorphism classes of smooth curves E → S whose
geometric fibres are all complex curves of genus 1 with a marked point. For a morphism
S → T , we get a functor M1,1 (T ) → M1,1 (S) defined by pullback: for a family of elliptic
curves E → T , f ∗ (E → T ) is the family E 0 → S which is the pullback in the diagram

E0 E

f
S T

Suppose M1,1 were representable, say M1,1 ∼= Hom(−, M ) for some scheme M . Let E0 ∈
M1,1 (M ) correspond to the identity idM ∈ Hom(M, M ). Then every elliptic curve E → S
would be the pullback

E E0

S M

by a unique morphism S → M – that is, E0 → M would be a “universal” elliptic curve.


However, this is impossible: every elliptic curve E → S has a nontrivial degree 2 automor-
phism (corresponding to the map (x, y) 7→ (x, −y) if E is locally given by y 2 = x3 +ax+b) so
the map E → E0 cannot be unique. (Even worse, in M1,1 (C) the isomorphism classes of el-
liptic curves with j-invariant 0 and 1728 have additional nontrivial automorphisms to worry
about.) So M1,1 is not representable, but it does in fact have a coarse moduli space M1,1 ,
that is, a scheme M1,1 and a functor M1,1 → Hom(−, M1,1 ) which is a bijection when evalu-
ated on algebraically closed fields and such that for any scheme S and natural transformation
M1,1 → Hom(−, S), there is a unique morphism M1,1 → S making the diagram

M1,1 Hom(−, M1,1 )

Hom(−, S)

commute. In fact, M1,1 is none other than the j-line, A1j = Spec C[j].

712
Chapter 36. Categories Fibred in Groupoids

Example 36.0.8. Consider the moduli problem parametrizing Riemann surfaces of genus
g ≥ 2, or smooth proper complex curves X over C with X(C) ∼ = Cg /Λ for a full lattice Λ.
Let Mg : SchC → Set be the functor sending S to the set of isomorphism classes of smooth
op

schemes X → S whose geometric fibres are all Riemann surfaces of genus g. Such an X
is often called a family of Riemann surfaces, parametrized by the base S. As above, for a
morphism f : S → T we get a pullback functor f ∗ : Mg (T ) → Mg (S). By a similar proof to
the elliptic curve case, Mg is not a representable functor – that is, there is no (fine) moduli
space of genus g Riemann surfaces. However, as above, there is a moduli space Mg for Mg
which is of considerable complexity and has inspired much research in algebraic geometry.

The standard way to approach the issue of automorphisms in a moduli problem F :


Schop → Set is to replace the category Set with the category Gpd of groupoids.

Definition. A category C is a groupoid if every morphism in C is an isomorphism.

Example 36.0.9. Let G be a group. Then G determines a category – usually also denoted
by G – in which there is only one object ∗ and for each g ∈ G, there is an automorphism
g
∗→− ∗. Thus G is a groupoid.

In Examples 36.0.7 and 36.0.8, the pullback functor f ∗ was crucial for studying the moduli
problems of curves of genus g. In order to allow for pullbacks to play a role in our study of
moduli spaces, we will replace a functor Schop → Set with a category fibred in groupoids. In
order to make this precise, we need to introduce fibred categories.

713
36.1. Fibred Categories Chapter 36. Categories Fibred in Groupoids

36.1 Fibred Categories


Definition. A category over a category C is a category F and a functor π : F → C. The
fibre of an object X ∈ C is the subcategory F(X) of objects x ∈ F such that π(x) = X and
morphisms covering idX , i.e. morphisms ϕ : x → x0 such that π(ϕ) : X → X is the identity.
A morphism of categories over C is a functor T : F → G commuting with the functors
F → C and G → C.
Definition. Let π : F → C be a category over C. A morphism ϕ : x → x0 in F is cartesian if
for any other morphism ψ : y → x0 in F such that π(ψ) = π(ϕ) ◦ h for some h : π(y) → π(x),
there exists a unique morphism α : y → x covering h and making the diagram
ψ
α ϕ
y x x0

h π(ϕ)
π(y) π(x) π(x0 )

π(ψ)

commute. Then π : F → C is a fibred category if for every morphism f : X → X 0 in C


and object x0 ∈ F(X 0 ), there exists a cartesian morphism ϕ : x → x0 covering f .
In particular, note that if f : X → X 0 lifts to a cartesian morphism ϕ : x → x0 in F then
x ∈ F(X). One sometimes calls x the pullback of x0 along f , written x = f ∗ x0 .
Definition. A morphism of fibred categories over C is a morphism T : F → G of cate-
gories over C that takes cartesian morphisms to cartesian morphisms. A base-preserving
natural transformation between two morphisms S, T : F → G of fibred categories is a
natural transformation τ : S → T such that for all x ∈ F, τx : S(x) → T (x) covers the
identity morphism in C.
The collection of all morphisms of fibred categories F → G over C, together with the base-
preserving natural transformations between them, forms a category denoted HomC (F, G).
This shows that the category of all fibred categories over C in fact forms a 2-category.
Definition. A category fibred in groupoids is a fibred category F → C such that for
every object X ∈ C, the fibre category F(X) is a groupoid.
Let CFG(C) denote the full 2-subcategory of fibred categories over C which are fibred in
groupoids.
Example 36.1.1. Suppose C is a category and F : Cop → Set is a presheaf on C. Then
F determines a category π : CF → C with CF (X) = F (X) for each X ∈ C. There is a
morphism s → t in CF if π(s) = X, π(t) = X 0 and there is a morphism ϕ : X → X 0 in C
such that F (ϕ)(F (X)) = F (X 0 ). A set is naturally a groupoid with only identity morphisms
and it is easy to check the axioms of a fibred category hold for π : CF → C, so every presheaf
on C naturally determines a category fibred in groupoids over C.

714
36.1. Fibred Categories Chapter 36. Categories Fibred in Groupoids

Example 36.1.2. If X ∈ C is any object, then the slice category C/X is naturally a
category fibred in groupoids over C, where the projection C/X → C is just (Y → X) 7→ Y .
A category F fibred in groupoids over C is called representable if it is equivalent (as a category
fibred in groupoids) to a slice category C/X for some object X, in which case F is said to
be represented by X.
Example 36.1.3. Let SchX be the category of X-schemes and define a category π : F →
SchX whose objects are pullback squares
P Y

T X

in SchX and whose morphisms are pairs of morphisms (T 0 → T, P 0 → P ) making the


appropriate diagrams commute. The functor π sends the square above to T → X ∈ SchX ,
so that for any fixed X-scheme T , the fibre category F(T ) may be identified with the category
of pullbacks P = T ×X Y which exist (showing the fibred condition holds for F) and are
unique up to unique isomorphism. Therefore F(T ) is a groupoid so F is a category fibred in
groupoids over SchY .
Proposition 36.1.4. A fibred category F → C is fibred in groupoids if and only if every
morphism in F is cartesian.
We will use liberally the following 2-categorical version of the Yoneda lemma from cate-
gory theory.
Lemma 36.1.5 (2-Yoneda Lemma). For any object X ∈ C and fibred category π : F → C,
the functor η : HomC (C/X, F) → F(X) defined by sending S : C/X → F to S(idX ) ∈ F(X)
is an equivalence of 2-categories.
Proof. We define a functor ξ : F(X) → HomC (C/X, F) as follows. For each object x ∈ F(X)
and morphism ϕ : Y → X in C, choose a pullback ϕ∗ x ∈ F(Y ) and set
ξx : C/X −→ F
ϕ 7−→ ϕ∗ x.
On the level of morphisms, if ψ : Z → X is another morphism in C and α : Y → Z
is a morphism over X, then by the cartesian condition in F there is a unique morphism
ξx (α) : ϕ∗ x → ψ ∗ x which completes the diagram

ϕ∗ x ψ∗x x
ξx (α)

α ψ
Y Z X

715
36.1. Fibred Categories Chapter 36. Categories Fibred in Groupoids

This defines ξx ∈ HomC (C/X, F) for every object x ∈ F(X). If f : x → x0 is a morphism in


F(X), then for any ϕ : Y → X in C/X, choose pullbacks ϕ∗ x and ϕ∗ x0 ∈ F(Y ) of x and x0 ,
respectively. Then there is a unique morphism ξf (ϕ) completing the diagram
ξf (ϕ)
ϕ∗ x ϕ∗ x0

f
x x0

by the cartesian condition defining ϕ∗ x, ϕ∗ x0 . This defines ξ on morphisms, so ξ : F(X) →


HomC (C/X, F) is a functor.
It remains to check that η and ξ together give an equivalence of categories. On one hand,
ξ ◦ η : HomC (C/X, F) → HomC (C/X, F) sends S : C/X → F to ξS(idX ) : (ϕ : Y → X) 7→
ϕ∗ S(idX ). This is canonically isomorphic to S itself since idX is a final object in C/X and
thus there exists a unique cartesian morphism (ϕ : Y → X) → (idX : X → X), which makes
S(ϕ) → S(idX ) also cartesian and hence S(ϕ) is a pullback of S(idX ). This shows ξ ◦ η is
naturally isomorphic to the identity functor. On the other hand, η ◦ ξ : F(X) → F(X) takes
x ∈ F(X) to id∗X x which is canonically isomorphic to x itself, thus proving η ◦ ξ ' idF(X) .

Corollary 36.1.6. For any objects X, Y ∈ C, the functor

HomC (C/X, C/Y ) −→ HomC (X, Y )


S 7−→ S(idX )

is an equivalence of categories.

Corollary 36.1.7. For any category C, there is a fully faithful embedding of 2-categories

C ,−→ CFG(C), X 7−→ C/X,

where CFG(C) is the 2-category of categories fibred in groupoids over C. In a similar way,
there is a fully faithful embedding of 2-categories

PreshC ,−→ CFG(C), F 7−→ CF .

Example 36.1.8. Let Schk be the category of schemes over a field k. Then for each k, n ≥ 1,
we define a category fibred in groupoids Gr(k, n) → Schk called the (k, n)th Grassmannian
category, by putting Gr(k, n)(X) as the category of vector bundles E → X of rank k, together
with embeddings of bundles E ,→ AnX . A morphism in Gr(k, n) from E ∈ Gr(k, n)(X) to
E 0 ∈ Gr(k, n)(X 0 ) is a morphism of bundles E → E 0 covering a morphism of schemes X → X 0
and commuting with the natural map AnX → AnX 0 .

Example 36.1.9. For each X ∈ Schk , let M1,1 (X) be the category consisting of pairs (E, O)
where E → X is a smooth curve, O : X → E is a section and for every geometric point
x̄ : Spec k̄ ,→ X, the pullback (Ex̄ , Ox̄ ) is an elliptic curve over k̄. Then M1,1 → Schk is a
fibred category. Morphisms (E, O) → (E 0 , O0 ) in M1,1 are given by pullback diagrams

716
36.1. Fibred Categories Chapter 36. Categories Fibred in Groupoids

F
E E0

f
X X0

such that F ◦ O0 = O ◦ f . Since pullbacks are unique up to unique isomorphism, it follows


that M1,1 is a category fibred in groupoids over Schk .

Example 36.1.10. Likewise for each g = 0, g ≥ 2 and X ∈ Schk , define Mg (X) to be the
category consisting of smooth curves C → X whose geometric fibres Cx̄ are smooth curves
of genus g over k̄. Then Mg is a fibred category over Schk . As above, morphisms in Mg are
given by pullbacks, so Mg is a category fibred in groupoids.

Example 36.1.11. The following construction will be important in the definition of stacks.
Let π : F → C be a category fibred in groupoids. For X ∈ C and x, x0 ∈ F(X), define a
presheaf Isom(x, x0 ) : (C/X)op → Set by

Isom(x, x0 )(ϕ : Y → X) = HomF(Y ) (ϕ∗ x, ϕ∗ x0 )

where ϕ∗ x, ϕ∗ x0 ∈ F(Y ) are pullbacks of x and x0 , respectively, along ϕ. Moreover, any other
ψ : Z → Y induces a map

ψ ∗ : Isom(x, x0 )(ϕ : Y → X) −→ Isom(x, x0 )(ϕ ◦ ψ : Z → X)

by composition of pullbacks. In particular, Aut(x) := Isom(x, x) is a presheaf of groups on


C/X for any x ∈ F(X).

717
36.2. Fibre Products Chapter 36. Categories Fibred in Groupoids

36.2 Fibre Products


In this section we construct fibre products of categories fibred in groupoids. We first give
the construction for individual groupoids. Fix a groupoid G and suppose G1 and G2 are two
groupoids over G, meaning there are functors π1 : G1 → G and π2 : G2 → G. The fibre product
G1 ×G G2 is defined to be the category with objects (X, Y, ϕ) consisting of X ∈ G1 , Y ∈ G2 and
an isomorphism ϕ : π1 (X) → π2 (Y ) in G. We define a morphism (X, Y, ϕ) → (X 0 , Y 0 , ϕ0 ) in
G1 ×G G2 to be a pair of morphisms α : X → X 0 in G1 and β : Y → Y 0 in G2 making the
diagram
ϕ
π1 (X) π2 (Y )

ϕ(α) ϕ0 (β)
ϕ0
π1 (X 0 ) π2 (Y 0 )

commute. Then G1 ×G G2 is a groupoid by construction and there are functors p1 : G1 ×G G2 →


G1 and p2 : G1 ×G G2 → G2 given by p1 (X, Y, ϕ) = X, p2 (X, Y, ϕ) = Y and similar projections
on morphisms. Further, the diagram
p2
G1 ×G G2 G2

p1 π2
π1
G1 G

2-commutes, i.e. there is a natural isomorphism π1 ◦ p1 ' π2 ◦ p2 .


Proposition 36.2.1. For any groupoids π1 : G1 → G and π2 : G2 → G, the fibre product
G1 ×G G2 is universal with respect to groupoids P making the diagram
q2
P G2

q1 π2
π1
G1 G

2-commute. That is, for such a groupoid P there is a functor t : P → G1 ×G G2 , unique up to


unique natural isomorphism, and natural isomorphisms λ1 : q1 → p1 ◦ t and λ2 : q2 → p2 ◦ t
making the diagram of functors
π1 (λ1 )
π 1 ◦ q1 π 1 ◦ p1 ◦ t

' '
π2 (λ2 )
π 2 ◦ q2 π 2 ◦ p2 ◦ t

718
36.2. Fibre Products Chapter 36. Categories Fibred in Groupoids

commute.

This says that G1 ×G G2 is a 2-categorical fibre product of G1 and G2 . Now let C be any
category and π : F → C, π1 : F1 → F and π2 : F2 → F categories fibred in groupoids over
C. The construction of a 2-categorical fibre product F1 ×F F2 in the 2-category CFG(C) of
categories fibred in groupoids over C is similar to the construction above, but we give it here
for completeness.

Proposition 36.2.2. There exists a category F1 ×F F2 fibred in groupoids over C together


with morphisms of fibred categories p1 : F1 ×F F2 → F1 and p2 : F1 ×F F2 → F2 as well as
an isomorphism (a base-preserving natural isomorphism) σ : π1 ◦ p1 ' π2 ◦ p2 satisfying the
following universal property:

(i) For any category H → C fibred in groupoids, the morphism of groupoids

HomC (H, F1 ×F F2 ) −→ HomC (H, F1 ) ×HomC (H,F) HomC (H, F2 )


η 7−→ (p1 ◦ η, p2 ◦ η, σ ◦ η)

is an isomorphism.

(ii) For any morphisms q1 : G 0 → F1 , q2 : G 0 → F2 with an isomorphism τ : π1 ◦q1 ' π2 ◦q2


making
HomC (H, G 0 ) −→ HomC (H, F1 ) ×HomC (H,F) HomC (H, F2 )
an isomorphism for all H → C, there exists an equivalence of fibred categories t :
H → F1 ×F F2 , unique up to unique isomorphism, and isomorphisms λ1 : q1 → p1 ◦ t
and λ2 : q2 → p2 ◦ t making the diagram of morphisms

π1 ◦ λ1
π1 ◦ q 1 π1 ◦ p 1 ◦ t

τ σ◦t
π2 ◦ λ2
π2 ◦ q 2 π2 ◦ p 2 ◦ t

commute.

Proof. (Sketch; see Olsson 3.4.13) Define G = F1 ×F F2 to be the category with objects
(x, y, ϕ) consisting of x ∈ F1 , y ∈ F2 and an isomorphism ϕ : π1 (x) → π2 (y) in F. A
morphism (x, y, ϕ) → (x0 , y 0 , ϕ0 ) in G is a pair of morphisms α : x → x0 in F1 and β : y → y 0
in F2 making the diagram
ϕ
π1 (x) π2 (y)

ϕ(α) ϕ0 (β)

0
ϕ0
π1 (x ) π2 (y 0 )

719
36.2. Fibre Products Chapter 36. Categories Fibred in Groupoids

commute. Define p1 : G → F1 by (x, y, ϕ) 7→ x and the obvious projection on morphisms


and p2 : G → F2 similarly by (x, y, ϕ) 7→ y. A natural isomorphism σ : π1 ◦ p1 → π2 ◦ p2 is
induced by the isomorphisms ϕ. One can check that G is a category over C; it is clear that
G(X) is a groupoid for each X ∈ C.
(i) Suppose H → C is a category fibred in groupoids and (ξ1 , ξ2 , γ) is an object in the
groupoid HomC (H, F1 ) ×HomC (H,F) HomC (H, F2 ). This induces a morphism η ∈ HomC (H, G)
defined on objects by η(h) = (ξ1 (h), ξ2 (h), γh ) where γh is the isomorphism π1 ◦ ξ1 (h) →
π2 ◦ ξ2 (h) specified by γ. The definition of η on functors is similar and under the morphism

HomC (H, G) −→ HomC (H, F1 ) ×HomC (H,F) HomC (H, F2 ),

it is clear that η goes to (ξ1 , ξ2 , γ). On the other hand, if η ∈ HomC (H, G) then η is precisely
the image under the above morphism of the triple (p1 ◦ η, p2 ◦ η, σ ◦ η). Therefore

HomC (H, G) −→ HomC (H, F1 ) ×HomC (H,F) HomC (H, F2 )

is an isomorphism of groupoids.
(ii) Now suppose qi : G 0 → Fi , i = 1, 2, are morphisms and τ : π1 ◦ q1 ' π2 ◦ q2 a natural
isomorphism such that

HomC (H, G 0 ) −→ HomC (H, F1 ) ×HomC (H,F) HomC (H, F2 )


η 7−→ (q1 ◦ η, q2 ◦ η, τ ◦ η)

is an isomorphism of groupoids for any H. Then for H = G = F1 ×F F2 , we have an


isomorphism
HomC (G, G 0 ) −→ HomC (G, F1 ) ×HomC (G,F) HomC (G, F2 ).
Then the data (p1 , p2 , σ) on the right determines a morphism t−1 : G → G 0 on the left and
natural transformations λ−1 1 : p1 ◦ t
−1
→ q1 and λ−1
2 : p2 ◦ t
−1
→ q2 making the appropriate
diagram of functors commute. Applying this with G and G 0 reversed constructs t, λ1 and λ2
and shows that the first is an equivalence and the second and third are natural isomorphisms.
This completes the proof.

720
Chapter 37

Descent

The notion of descent can be phrased in a quite general context. Fix an object S in a
category C and recall that the localized or slice category at S is the category C/S whose
objects are arrows X → S in C and whose morphisms are commutative triangles

X Y

Given any other object S 0 ∈ C and a morphism S 0 → S, there is a base change functor
C/S → C/S 0 given by (X → S) 7→ (X 0 → S 0 ) where X 0 is the fibre product

X 0 = X ×S S 0 X

S0 S

This answers the question “when does an object over S determine an object over S 0 ?” The
opposite question, namely when an object over S 0 determines an object over S, is much
harder to answer in general. In fact there are two interesting questions one might ask: given
a morphism S 0 → S and an arrow (X 0 → S 0 ) ∈ C/S 0 , (1) is there an arrow (X → S) ∈ C/S
completing the diagram

X0 X

S0 S

and if so, (2) how many ways are there to complete the diagram? The process of answering
both of these questions is known as descent. In the first three sections of this chapter, we
outline some of the basics of descent theory in the algebraic geometry of schemes, following

721
Chapter 37. Descent

a series of lectures given by Dr. Andrei Rapinchuk in the Galois-Grothendieck Seminar at


University of Virginia in Fall 2018. Then we turn to descent conditions on different sites on
a category of schemes and ultimately give the definition of a stack.

722
37.1. Galois Descent Chapter 37. Descent

37.1 Galois Descent


Let K/k be a field extension and obj(k) some class of objects defined over k.

Example 37.1.1. obj(k) could be the class of k-vector spaces, or more specifically, the class
of Lie algebras over k, central simple algebras over k, etc. Notice that many examples like
this admit morphisms, and so form a category over k.

Example 37.1.2. A quadratic space over k is a k-vector space V together with a quadratic
form q, that is, a symmetric bilinear function q : V → k. The set of quadratic spaces
(V, q) may be an interesting class of objects over k. These too admit morphisms, where
(V, q) → (V 0 , q 0 ) consists of a k-linear isomorphism V → V 0 which commutes with q and q 0 .

Example 37.1.3. Important choices of obj(k) in algebraic geometry include the class of
algebraic varieties or algebraic groups over k, the class of schemes over Spec k, and more
generally things like algebraic spaces and algebraic stacks over k.

Many of these examples admit a base change functor as in the chapter introduction, i.e.
an assignment

obj(k) −→ obj(K)
X 7−→ XK .

Most of the time this functor can be built using the tensor product or fibre product in the
right category.

Example 37.1.4. For some of the examples above, we have the following base change
functors:
object over k base change to K
vector space V VK = V ⊗k K
central simple algebra A AK = A ⊗k K
quadratic space (V, q) (VK , qK ) where qK extends q
variety/scheme X XK = X ×k K (defined by ⊗ on affine patches)
algebraic group G GK = G ×k K
In the pattern of the introduction, there are two natural questions we would like to answer
about descending objects defined over K to objects over k:

(1) Given an object A ∈ obj(K), does there exist an object X ∈ obj(k) such that XK = A?

(2) What are all the possible objects X ∈ obj(k) with XK = A or XK ∼


= A in obj(K)?
When K/k is a Galois extension, this is known as Galois descent.

Definition. An object X ∈ obj(k) with XK ∼


= A in obj(K) is called a K/k-form of A.
Example 37.1.5. In some situations, descent is trivial. For example, if VK is a K-vector
space with basis {xi } then the same basis gives a k-vector space Vk for which Vk ⊗k K = VK .

723
37.1. Galois Descent Chapter 37. Descent


Example 37.1.6. In other situations, descent is √ impossible. Let k = Q, K = Q( 2) and
2 2 2
consider the quadratic form qK (x, y) = x − y 2 on V = K . Is there a quadratic form
qQ on VQ = Q2 which extends to qK on VQ ⊗Q K = V ? The answer in this case is no – we
will see soon that there are some elementary conditions that are necessary for descent to be
possible, and they are not satisfied in this situation.
Example 37.1.7. (Motivation) The Inverse Galois Problem asks: for which finite groups
G can one construct a field extension `/Q with Gal(`/Q) ∼ = G? By Hilbert’s Irreducibility
Theorem, it is enough to construct an extension L/Q(t) with Gal(L/Q(t)) ∼
= G. Over C, one
can construct Galois extensions L/C(t) by instead constructing branched covers of Riemann
surfaces X → P1C . It is a general fact that every such cover X is defined over Q, so the
question of constructing a G-extension of Q(t) comes down to being able to descend the
cover X → P1Q to a cover XQ → P1Q . A general solution to this would solve the Inverse
Galois Problem.
Let K/k be a Galois extension with Galois group G = Gal(K/k) and suppose Vk and Wk
are objects over k (e.g. k-algebras) that are isomorphic over K, that is, VK = Vk ×k K ∼ =
Wk ×k K = WK . Moreover, assume that under the natural Galois action on VK and WK ,
we have Vk = VKG and Wk = WKG (this is true e.g. for k-algebras). When there is a notion of
maps between objects over k (e.g. k-linear algebra homomorphisms), the G-action extends
to MapK (VK , WK ) by (σf )(v) = σ(f (σ −1 v)) for all v ∈ VK , σ ∈ G.
f
For example, let f be an equivalence VK →− WK , e.g. a K-algebra isomorphism. Set
ξ(σ) = f −1 ◦ σf ∈ GK := AutK (VK ). Notice that if ξ(σ) = 1 for all σ ∈ G, then σf = f σ

so f descends to an isomorphism fk : Vk = VKG −
→ WKG = Wk . That is, ξ = 1 is a necessary
condition for two (K/k)-forms of VK to be isomorphic over k. In general, the automorphism
ξ(σ) = f −1 · σf satisfies

ξ(στ ) = f −1 · (στ )f = (f −1 · σf )((σf )−1 στ f ) = ξ(σ)σ(ξ(τ )).

Such a map ξ : G → GK is called a 1-cocycle, and the set of all 1-cocycle is denoted Z 1 (G, GK ).
If K/k is an infinite extension, we are guaranteed to have f (Vk ) ⊆ Wk ⊗k ` ⊆ Wk ⊗k K = WK
for some finite extension `/k. If σ ∈ H := Aut(K/`), then for all v ∈ Vk ,

(σf )(v) = σ(f (σ −1 v)) = σσ −1 f (v) = f (v)

since f (v) ∈ Wk ⊗k `. Thus σf = f , so ξ(σ) = 1 for all σ ∈ H. Said another way, the
1-cocycle ξ is constant on cosets of H, which means ξ is in fact a continuous 1-cocycle on

G = lim Aut(`/k)
−→

over all finite extensions `/k, when this direct limit is equipped with the Krull topology
induced by discrete topologies on each Aut(`/k).

Turning back to our K-isomorphism f : VK − → WK , the definition of ξ may depend on

this f , but suppose f 0 were a different isomorphism VK −→ WK . Then f 0 = f ◦ g for some
g ∈ GK and the cocycle ξ 0 defined by f 0 has the form

ξ 0 (σ) = (f 0 )−1 · σf 0 = (f ◦ g)−1 · σ(f ◦ g) = g −1 f −1 · (σf )(σg) = g −1 ξ(σ)σg.

724
37.1. Galois Descent Chapter 37. Descent

Definition. Two 1-cocycles ξ, ξ 0 : G → GK are equivalent cocycles if there is some g ∈ GK


such that ξ 0 (σ) = g −1 ξ(σ)σg for all σ ∈ G. The set of equivalence classes of 1-cocycles is
denoted H 1 (G, GK ).

Fix an object VK over K, set GK = AutK (VK ) and let F (K/k, VK ) be the set of k-
isomorphism classes of (K/k)-forms of Vk . Then our work above shows that there is a map

Θ : F (K/k, VK ) −→ H 1 (G, GK ).

Lemma 37.1.8. Θ is injective.



Proof. Suppose Vk and Wk are (K/k)-forms of VK with isomorphisms f : VK − → Vk ⊗k K and

f 0 : VK −→ Wk ⊗k K. Let ξ, ξ 0 : G → GK be their corresponding 1-cocycles and suppose ξ and
ξ 0 are equivalent. Replacing f with f ◦ g, we may assume ξ = ξ 0 . Then (f 0 )−1 · σf 0 = f −1 · σf .
Rearranging this, we have

f ◦ (f 0 )−1 = σ(f ◦ (f 0 )−1 ) as a map Wk ⊗k K −→ Vk ⊗k K.



Therefore f ◦(f 0 )−1 descends to an isomorphism Wk −
→ Vk over k, so Vk and Wk are isomorphic
as (K/k)-forms of VK .

Theorem 37.1.9. When Vk is a k-algebra, Θ : F (K/k, VK ) → H 1 (G, GK ) is a bijection.

Proof. Let ξ ∈ Z 1 (G, GK ) be a 1-cocycle. We define a (K/k)-form Wk by defining a new


Galois action σ ∗ v on WK := Vk ⊗k K and taking Wk = (Vk ⊗k K)G∗ . We must do so in such
a way that the identity VK → Vk ⊗k K = WK determines the cocycle ξ. Note that for this
to happen, for any v ∈ VK we need to have

ξ(σ)(v) = f −1 (σf )(v) = f −1 (σf σ −1 (v))


=⇒ ξ(σ)σ(v) = f −1 (σf (v)) replacing v with σ(v)
=⇒ f (ξ(σ)σ(v)) = σf (v).

Then defining σ ∗v = ξ(σ)σ(v), i.e. twisting by the cocycle ξ, gives a new k-semilinear action
of G on Vk ⊗k K. Note that for any σ, τ ∈ G,

σ ∗ (τ ∗ v) = σ ∗ (ξ(τ )τ (v)) = ξ(σ)σ(ξ(τ )τ (v))


= ξ(σ)σ(ξ(τ ))σ(τ (v)) by the cocycle condition on ξ
= ξ(στ )στ (v) = (στ ) ∗ v,

where f is the identity VK = WK . Therefore ∗ is a G-action on WK . Set Wk = WKG∗ , the fixed


points of this Galois action. We will prove later that the identity f induces an isomorphism
Wk ⊗k K = WK ∼ = VK as G-sets, but for now, it is obvious that this isomorphism (really, the
identity) determines the cocycle ξ. Therefore Θ is surjective, so it is a bijection.
The proof of Theorem 37.1.9 suggests a general strategy for descent in other contexts:
define a new Galois action on the object VK over K and take Wk to be the fixed points of
VK under this action.

725
37.1. Galois Descent Chapter 37. Descent

It will be useful to define group cohomology and Galois cohomology in the nonabelian
case, so we recall the definitions here. Suppose G acts continuously on a group A (which
may be nonabelian) with the discrete topology – remember that this means the stabilizers
are open in G. Then a 1-cocycle of the action is a function ξ : G → A that is continuous and
satisfies
ξ(στ ) = ξ(σ)σξ(τ ) for all σ, τ ∈ G.
Let Z 1 (G, A) denote the set of all 1-cocycles ξ : G → A. We say two cocycles ξ, ξ 0 ∈ Z 1 (G, A)
are equivalent if there exists an a ∈ A such that

ξ 0 (σ) = a−1 ξ(σ)σ(a).

Let H 1 (G, A) be the set of equivalence classes of 1-cocycles. Then H 1 (G, A) is a pointed set
with distinguished element [1], where 1 is the trivial cocycle.
Suppose f : A → B is a morphism of G-groups, that is, a group homomorphism which
commutes with the G-action. Then there is a map of pointed sets

H 1 f : H 1 (G, A) −→ H 1 (G, B).

Set H 0 (G, A) = AG .

Proposition 37.1.10. For every short exact sequence of G-groups 1 → A → B → C → 1,


there is an exact sequence of pointed sets

1 → H 0 (G, A) → H 0 (G, B) → H 0 (G, C) → H 1 (G, A) → H 1 (G, B) → H 1 (G, C).

Furthermore, if A is central in B (and in particular is an abelian group), then the sequence


may be extended by
· · · → H 1 (G, B) → H 1 (G, C) → H 2 (G, A)
and this is exact.

Here, H 2 (G, A) is the ordinary 2nd group cohomology of G with coefficients in an abelian
G-group A. Suppose that α ∈ Z 1 (G, A). Define the twist of A by α to be the G-group Aα
whose underlying group is A but with G-action defined by

g ∗ a = α(g)g(a)α(g)−1 = α(g) · g(a)

for any g ∈ G and a ∈ A. (Here, x · y denotes the inner action of A on itself.) Let AG∗ denote
the fixed points of this action.

Lemma 37.1.11. The map ξ 0 7→ ξ := ξ 0 · α yields a bijection tα : Z 1 (G, Aα ) −
→ Z 1 (G, A)
which descends to a bijection of pointed sets

τα : H 1 (G, Aα ) −→ H 1 (G, A).
α β
Let 1 → A −
→B→
− C → 1 be a short exact sequence of G-groups and let
f
1 → H 0 (G, A) → H 0 (G, B) → H 0 (G, C) → H 1 (G, A) →
− H 1 (G, B)

726
37.1. Galois Descent Chapter 37. Descent

be the corresponding long exact sequence. For ξ ∈ H 1 (G, A), there is a twisted sequence

1 → Aξ → Bα(ξ) → Cξ → 1

which yields a long exact sequence



G
1 → AGξ → Bα(ξ) → CξG → H 1 (G, Aξ ) −
→ H 1 (G, Bα(ξ) ).

Further, this fits into a commutative diagram


[1] H 1 (G, Aξ ) H 1 (G, Bα(ξ) )

τξ τα(ξ)
f
ξ H 1 (G, A) H 1 (G, B)

whose columns are bijections by Lemma 37.1.11. This shows that the fibre of f : H 1 (G, A) →
H 1 (G, B) over f (ξ), namely f −1 (f (ξ)), may be identified with the fibre of fξ over [1], which
G
is ker fξ . Explicitly, ker fξ = Bα(ξ) /CξG which can be computed in many cases.
Example 37.1.12. Let K = k sep be the separable closure of k and consider the matrix
algebra Ak = M2 (k) ∈ Algk . Suppose Bk is a (K/k)-form of AK = M2 (K), with k-linear

isomorphism f : M2 (K) − → BK . Then Bk is a central simple k-algebra of dimension 4 over
k. Assuming char k 6= 2, it is well-known that any such Bk is a quaternion algebra given by
a basis {1, i, j, k} satisfying the relations i2 = a, j 2 = b, k = ij, ij = −ji
 for some a, b ∈ k.
Denote the quaternion algebra with this presentation by Bk = a,b k
. For Ak = M2 (k),
×
we have G  K = AutK (M2 (K)) = GL2 (K)/K = P GL2 (K). Given a quaternion algebra
Bk = a,bk
, set L = k(i) and define an embedding

Bk ,−→ M2 (L)
 
x0 + x1 i b(x2 + x3 i)
x0 + x1 i + x2 j + x3 k 7−→ .
x2 − x3 i x 0 − x1 i

This extends to an isomorphism h : BK = Bk ⊗k K − → M2 (K) = AK ; call its inverse f . For
x = x0 + x1 i + x2 j + x3 k ∈ Bk and σ ∈ Gal(L/k), we have

σ(h)(x) = σ(hσ −1 (x)) = σ(h(x)) since Bk is fixed by σ −1


 
x0 − x1 i b(x2 − x3 i)
=
x2 + x3 i x0 + x1 i
 
−1 0 b
= gh(x)g where g = .
1 0

This shows that


   
−1 0 1 0 b
ξ(σ) = g = = =g in P GL2 (K).
b−1 0 1 0

727
37.1. Galois Descent Chapter 37. Descent

In fact, (
1, σ|L = id
ξ(σ) =
g, σ|L = 6 id.

This gives us a 1-cocycle ξ corresponding to a,b



k
.
Alternatively, consider the exact sequence of G-groups

1 → K × → GL2 (K) → P GL2 (K) → 1.

Then the extended term in the long exact sequence from Proposition 37.1.10 is

H 1 (G, P GL2 (K)) −→ H 2 (G, K × )


ξ 7−→ [δ]

˜
where δ(σ, τ ) = ξ(σ)σ ˜ )ξ(στ
ξ(τ ˜ )−1 for σ, τ ∈ G and ξ˜ is any lift of ξ to GL2 (K). Notice
that H 2 (G, K × ) = Br(k) is the Brauer group of k, whose elements are equivalence classes
of central simple k-algebras. For the cocycle ξ coming from the quaternion algebra a,b

k
as
above, (
b, σ|L , τ |L 6= id
δ(σ, τ ) =
1, otherwise.
This identifies each element of F (K/k, M2 (k)) with a (Brauer class of a) central simple k-
algebra of degree 2. In general, the (K/k)-forms of Mn (k) may be identified with the Brauer
classes of central simple algebras of degree n.
Conversely, start with a quadratic extension L/k and a 1-cocycle ξ ∈ H 1 (G, P GL2 (K)).
One can also construct a central simple algebra corresponding to ξ by taking BL = M2 (L)
as an algebra and defining a new Galois action on BL according to σ ∗ a = ξ(σ)σ(a) for all
x y
σ ∈ G and a ∈ M2 (L). In particular, if M = then
z w
     
0 b σ(x) σ(y) 0 1 σ(w) bσ(z)
ξ(σ)σ(M ) = = −1
1 0 σ(z) σ(w) b−1 0 b σ(y) σ(x)

for some b ∈ k. Thus we can identify Bk = BLG∗ as the matrices M ∈ M2 (L) for which
σ ∗ M = M , i.e.    
σ(w) bσ(z) x y
= .
b−1 σ(y) σ(x) z w
That is,   
x y
Bk = x, y ∈ k .
b−1 σ(y) σ(x)
Example 37.1.13. Assume char k 6= 2 and consider the collection of quadratic spaces
(V, q) over k, with morphisms f : V1 ,→ V2 such that q2 |V1 = q1 . Fix a quadratic space
(V, q). Then GK = Aut(VK , qK ) = O(K, 2)q , the 2 × 2 orthogonal group of K preserving
the orientation induced by q. The set F (K/k, (VK , qK )) in this case classifies all quadratic
forms of a given rank (up to isomorphism). By Lemma 37.1.8, there is an injective map

728
37.1. Galois Descent Chapter 37. Descent

Θ : F (K/k, (VK , qK )) ,→ H 1 (G, O(K, 2)q ). Descent theory shows when this is a bijection.
Let q = x2 + y 2 be the quadratic form on V = and q 0 = ax2 + by 2 the quadratic form on V 0 .
Define

f : VK −→ VK0
 
x y
(x, y) 7−→ √ , √ .
a b
The corresponding 1-cocycle ξ ∈ H 1 (G, O(K, 2)q ) is
√ !
a

σ( a)
0
ξ(σ) = f −1 ◦ σ(f ) = √ .
0 √b
σ( b)

Conversely, for a cocycle ξ one can construct a quadratic space (V 0 , q 0 ) with V 0 = VKG∗ and
q 0 = qK |V 0 by defining the action σ ∗ v = ξ(σ)σ(v) for all σ ∈ G, v ∈ VK . Explicitly, for
v = (x, y), √ !
a
√ σ(x) 0
σ( a)
σ ∗ (x, y) = √ .
0 √b σ(y)
σ( b)
   
The fixed points under this action are (x, y) ∈ VK such that σ a = a and σ √yb = √yb ,
√x √x

√ √
so V 0 has k-basis {( a, 0), (0, b)}. The quadratic form in this case is precisely q 0 = ax2 +by 2 .
Thus all Galois twists of the quadratic form q = x2 + y 2 are of the form q 0 = ax2 + by 2 for
some a, b ∈ k.
Finally, consider the short exact sequence of G-groups
det
1 → SO(K, 2)q → O(K, 2)q −→ Z/2Z → 1.

For the quadratic form ax2 +by 2 over k, let ξa,b be the corresponding cocycle in H 1 (G, O(K, 2)q ).
By Proposition 37.1.10, there is a map

H 1 (G, O(K, 2)q ) −→ H 1 (G, Z/2Z)


√ !
ab
ξa,b 7−→ det ξ : σ →
7 √ .
σ( ab)

Further, by Kummer theory (see below), H 1 (G, Z/2Z) ∼ = k × /(k × )2 , the set of nonzero ele-
ments of k modulo squares in k. Under this correspondence, ξa,b is identified with the coset √
ab(k × )2 . Similar analysis
√ justifies the assertion in Example 37.1.6 that q K = x 2
− y 2
2
defined over K = Q( 2) does not descend to a quadratic form on Q.
Let K/k be a Galois extension with Galois group G = Gal(K/k) and suppose G is a
linear algebraic group over k on which G acts continuously. Then

H 1 (G, G) = lim H 1 (G/U, G(K U ))


−→

where the limit is over all finite index open subgroups U ⊆ G and K U is the subfield of K
fixed by U.

729
37.1. Galois Descent Chapter 37. Descent

Theorem 37.1.14 (Hilbert’s Theorem 90). For a Galois extension K/k with group G,

(1) H i (G, K) = 0 for all i ≥ 1.

(2) H 1 (G, K × ) = 1.

By the direct limit interpretation of H 1 (G, G) above, it is enough to prove Hilbert’s


Theorem 90 for a finite Galois extension, where it is a classical consequence of Artin’s lemma
on linear independence of characters. (For the i > 2 case in (1), one uses Shapiro’s lemma
and the fact that K is induced from the representation k of the trivial subgroup {1} ⊆ G.)
Note that for i ≥ 2, H i (G, K × ) need not vanish – for example, H 2 (G, K × ) may be identified
with the Brauer group of K/k which is nontrivial in general.

Example 37.1.15. (Artin-Schreier Theory) Let k be a field of characteristic p > 0, K = k sep


and consider the Artin-Schreier sequence

0 → Fp →
− K→
− K→0

where ℘(x) = xp − x. Applying H • (G, −), we get an exact sequence



− k → H 1 (G, Fp ) → H 1 (G, K) = 0
0→k→

where the last term is 0 by Hilbert’s Theorem 90. Thus H 1 (G, Fp ) ∼ = k/℘(k). On the
1 ∼
other hand, H (G, Fp ) = Hom(G, Fp ) which classifies cyclic extensions of degree p over k
(up to k-isomorphism). This demonstrates the Artin-Schreier correspondence between cyclic
p-extensions of k and elements of k/℘(k); explicitly, every cyclic p-extension L/k is of the
form L = k[x]/(xp − x − a) for some a ∈ k/℘(k). The general case of cyclic extensions of a
characteristic p field of order divisible by p is given by Artin-Schreier-Witt theory.

Example 37.1.16. (Kummer Theory) There is an analogous description of cyclic extensions


of prime-to-p order: consider
n
1 → µn → K × →
− K× → 1
n
where K × →
− K × is the map x 7→ xn and µn is the group (scheme) of nth roots of unity.
The long exact sequence in Galois cohomology is
n
1 → k× →
− k × → H 1 (G, µn ) → H 1 (G, K × ) = 1

by Hilbert’s Theorem 90. Thus H 1 (G, µn ) ∼ = k × /(k × )n . At the same time, if k contains the
nth roots of unity µn (k), then H 1 (G, µn ) ∼
= Hom(G, µn ) classifies cyclic n-extensions, and
every such an extension L/k is of the form L = k[x]/(xn − b) for b ∈ k × /(k × )n .

Example 37.1.17. The additive and multiplicative groups of a field k are often written
k = Ga and k × = Gm . Consider the semidirect product
  
a b
G = Ga o Gm = ∈ GL2 (k) .
0 1

730
37.1. Galois Descent Chapter 37. Descent

The short exact sequence


0 → Ga → G → Gm → 1
induces
1 = H 1 (G, Ga ) → H 1 (G, G) → H 1 (G, Gm ) = 1
by Hilbert’s Theorem 90, so exactness implies H 1 (G, G) = 1 as well. Note that G = Aut(A1k )

→ H 1 (G, G) = 1 tells us that there is a unique (K/k)-form
so the injective Θ : F (K/k, A1k ) −
of A1K , namely A1k . In simpler terms, if X is a curve over k such that X ×k K ∼ = A1K , then
X∼ = A1k as curves. In particular, if B is a k-algebra such that B ⊗k K ∼ = K[x], then it must
∼ 2
be that B = k[x] as k-algebras. This result also holds for Ak , or equivalently the polynomial
ring k[x, y], by a result of Kambayashi.

We next observe that H 1 (G, GLn (K)) = 1. To do so, we need:

Lemma 37.1.18. Let W be a vector space over K and assume that G acts on W by semi-
linear transformations. Set W0 = W G . Then W0 ⊗k K ∼
= W.
Proof. By passing to the direct limit, we may assume K/k is finite. Enumerate G =
{σ1 , . . . , σn } and pick a basis {a1 , . . . , an } for K/k. The matrix A = (σi (aj ))ij ∈ Mn (K)
satisfies At A = (TrK/k (ai aj ))ij which is the Gram matrix of the trace form TrK/k and hence
invertible since the trace form is nondegenerate. This shows A ∈ GLn (k). Suppose w ∈ W
and set n n
X X
λj (w) = σi (aj w) = σi (aj )σi (w)
i=1 i=1

for each 1 ≤ j ≤ n. Then


       
λ1 (w) σ1 (w) σ1 (w) λ1 (w)
 ..  t .   .  t −1  . 
 .  = A  ..  =⇒  ..  = (A )  .. 
λn (w) σn (w) σn (w) λn (w)

so σi (w) is in the K-span of W0 . Hence W0 spans W . By dimension counting, W ∼


=
W0 ⊗k K.

Theorem 37.1.19. For any Galois extension K/k with group G, H 1 (G, GLn (K)) = 1.

Proof. In this case GLn (K) = Aut(VK ) where V = k n . Suppose ξ ∈ Z 1 (G, GLn (K)). This
defines a twisted action of G on WK := K n by σ ∗ w = ξ(σ)σ(w) for any σ ∈ G. Setting

Wk = WKG∗ , Lemma 37.1.18 shows that Wk ⊗k K = WK . Thus the isomorphism f : k n − → Wk

extends to an isomorphism fK : K n − → WK and additionally, f and fK each commutes with
the Galois action:
f (σ(v)) = σ ∗ f (v) = ξ(σ)σ(f (v))
for all v ∈ k n and σ ∈ G (likewise for fK and v ∈ K n ). Replacing v with σ −1 (v), we get
f (v) = ξ(σ)σf (σ −1 (v)) for all v, so f = ξ(σ)σ(f ) which may be written f −1 ξ(σ)σ(f ) = 1.
This shows that ξ is equivalent to 1 in H 1 (G, GLn (K)).

731
37.1. Galois Descent Chapter 37. Descent

Recall that by Theorem 37.1.9,



Θ : F (K/k, VK ) −→ H 1 (G, GLn (K)).

Here, the proof of Theorem 37.1.19 essentially demonstrates that F (K/k, VK ) = 1 and then
the bijection Θ is used to conclude that H 1 (G, GLn (K)).

Corollary 37.1.20. For any Galois extension K/k with group G, H 1 (G, SLn (K)) = 1.

Proof. Apply the long exact sequence in Galois cohomology to the sequence
det
1 → SLn (K) → GLn (K) −→ K × → 1

and use Hilbert’s Theorem 90 and Theorem 37.1.19.


Notice that SLn (K) = AutK (V, ω) where V = K n and ω is any nonzero element of n V ∗ ,
V
for V ∗ the vector space dual of V . Therefore Corollary 37.1.20 says that there is only the
trivial twist of (V, ω).

Example 37.1.21. Let K = k sep and fix a quadratic space (V, q) over K. We use de-
scent to show that Θ : F (K/k, (V, q)) → H 1 (G, Oq (K, n)) is always a bijection. Given
ξ ∈ Z 1 (G, Oq (K, n)), define a twisted action of G on WK := V ⊗k K by σ ∗ w = ξ(σ)σ(w).
Then Wk = WKG∗ is an n-dimensional vector space over k, so it remains to produce a quadratic
form on Wk to which q restricts. Such a form q 0 must satisfy q(w) = q 0 (w) ∈ k for all w ∈ Wk .
Indeed, if w ∈ Wk , σ ∗ w = w so

q 0 (w) = q 0 (σ ∗ w) = q 0 (ξ(σ)σ(w)) = q 0 (σ(w)) = σ(q 0 (w))

since ξ takes values in Oq0 (K, n) and q 0 is defined over k. This implies q restricts to q 0 on
Wk , so descent is always possible here.

Finally, we have a generalization of Hilbert’s Theorem 90 to finite dimensional k-algebras.

Theorem 37.1.22. Let K/k be a Galois extension with group G, let A be a finite dimensional
k-algebra and set AK = A ⊗k K. Then H 1 (G, A×
K ) = 1.

Proof. Since H 1 (G, A×K ) is a direct limit over finite Galois extensions, we may assume K/k
is finite. Set d = [K : k]. View A and AK as right A-modules. For a cocycle ξ ∈ Z 1 (G, A× K ),
define the twisted action of G on AK by σ ∗a = ξ(σ)σ(a) and put B = AG∗ K . As vector spaces,
AK ∼ = B ⊗k K, so A ∼ = B as vector spaces. Note that B is also a right A-module. We claim
A∼ = B as A-modules. By the Krull-Schmidt theorem, we may write

A = I1m1 ⊕ · · · ⊕ Irmr and B = I1n1 ⊕ · · · ⊕ Irnr

for indecomposable submodules I1 , . . . , Ir and integers mj , nj ≥ 1. But since A ⊗k K ∼


=
B ⊗k K as AK -modules and

A ⊗k K ∼
= I1m1 d ⊕ · · · ⊕ Irmr d and B ⊗k K ∼
= I1n1 d ⊕ · · · ⊕ Irnr d ,

732
37.1. Galois Descent Chapter 37. Descent

the uniqueness part of the Krull-Schmidt theorem implies mj d = nj d for all 1 ≤ j ≤ r.


Therefore mj = nj for each j which shows A ∼ = B as A-modules.

Fix an isomorphism of A-modules f : A − → B and let c = f (1) ∈ B. Then f (x) = cx for
all x ∈ A and the base change fK : AK → B ⊗k K is also an isomorphism, so c must be a
unit in AK . In addition, c ∈ B = AG∗
K means that c = σ ∗ c = ξ(σ)σ(c), which can be written
c−1 ξ(σ)σ(c) = 1. Therefore ξ is equivalent to 1 in H 1 (G, A×
K ).

Suppose A is a central simple k-algebra and K = k sep . Then AK = A ⊗k K ∼


= Mn (K).
The determinant map det : Mn (K) → K restricts to a map

Nrd = NrdK/k : A −→ k

called the reduced norm map. Further, there is a short exact sequence of algebraic groups
det
1 → SL1 (AK ) → GL1 (AK ) −→ K × → 1.

Taking Galois cohomology yields an exact sequence


Nrd
1 → A× −−→ k × → H 1 (G, SL1 (AK )) → H 1 (G, GL1 (AK )).

By Theorem 37.1.22, H 1 (G, GL1 (AK )) = 1 so k × / Nrd(A× ) ∼ = H 1 (G, SL1 (AK )).
To interpret this cohomology set, recall that for an algebraic group G, the set H 1 (G, G(K))
classifies torsors for G defined over k, up to isomorphism over K. That is, an equivalence
class in H 1 (G, G(K)) is represented by a G-space X for which the map

G ×k X −→ X ×k X
(g, x) 7−→ (x, gx)

is an isomorphism. In general, a torsor X is defined over k but need not have any rational
points over k.
Example 37.1.23. For G = SL1 (AK ), the G-torsors are given by solutions to the algebraic
equation Nrd(x) = a for some a ∈ k × . Further, such a torsor has a rational point over k
if and only if a ∈ Nrd(A× ). This exhibits the bijection k × / Nrd(A× ) ∼ = H 1 (G, SL1 (AK ))
explicitly. For an example of this correspondence, let H be the quaternion algebra over R.
Then Nrd(x0 + x1 i + x2 j + x3 k) = x20 + x21 + x22 + x23 and the equation

x20 + x21 + x22 + x23 = a

has solutions over R if and only if a > 0.


Example 37.1.24. Let q be a quadratic form over k which determines a nondegenerate
symmetric matrix Q ∈ Mn (k). Then the torsors for Oq (K, n), which are classified by
H 1 (G, Oq (K, n)), are given by equations

X t QX = B

for X = (xij ) and some B ∈ Mn (k). This equation has a k-rational point if and only if Q
and B are similar over k.

733
37.2. Fields of Definition Chapter 37. Descent

37.2 Fields of Definition


Let K be a field, I ⊂ K[x1 , . . . , xn ] an ideal and fix a subfield k ⊂ K.
Definition. We say I is defined over k, or k is a field of definition for I, if Ik :=
I ∩ k[x1 , . . . , xn ] generates I in K[x1 , . . . , xn ].
Theorem 37.2.1. For any ideal I ⊂ K[x1 , . . . , xn ], there exists a subfield k0 ⊂ K such that
I is defined over k0 and k0 ⊆ L for any other field of definition L of I. Further, for an
automorphism σ ∈ Aut(K), I σ = I if and only if σ|k0 = 1.
Proof. Let {Mα }α∈A be the set of all monomials in x1 , . . . , xn . Pick a subset B ⊆ A so that
{Mβ }β∈B is a maximal linearly independent subset mod I, i.e. a basis of K[x1 , . . . , xn ]/I.
Put C = A r B. Then for each γ ∈ C, we can write
X
Mγ ≡ gβγ Mβ mod I
β∈B

for some gβγ ∈ K. Let k0 be the subfield of K generated by {gβγ }β∈B,γ∈C . Take f ∈ I and
write it
X
f= aα Mα for aα ∈ K
α∈A
X X
= aβ Mβ + aγ Mγ .
β∈B γ∈C

Since f ∈ I, we have
!
X X X
f= a0β Mβ + aγ Mγ − gβγ Mβ
β∈B γ∈C β∈B

with a0β ∈ K. Each Mγ − β∈B gβγ Mβ lies in I ∩ k[x1 , . . . , xn ] = Ik0 , so β∈B a0β Mβ ≡ 0
P P
mod I. Hence by definition of B, we have a0β = 0 for each β ∈ B, so f is a linear combination
of elements in Ik0 . This shows k0 is a field of definition for I.
To show k0 is minimal, let L ⊆ K be any field of definition of I and write IL = (p1 , . . . , ps )
for polynomials pi ∈ L[x1 , . . . , xn ]. Given the same notation as above, we must show that
gβγ ∈ L for all β ∈ B, γ ∈ C. For a fixed γ0 ∈ C, we have
X
Mγ0 − gβγ0 Mβ = p1 q1 + . . . + ps qs
β∈B
P
for some qi ∈ K[x1 , . . . , xn ], say qi = α∈A yiα Mα for yiα ∈ K. For 1 ≤ i ≤ s, define
X
qi∗ = Yiα Mα ∈ K[Yiα | 1 ≤ i ≤ s, α ∈ A][x1 , . . . , xn ]
α∈A

(here, Yiα are indeterminates). Let Y = (Yiα )1≤i≤s,α∈A . Then


X
p1 q1∗ + . . . + ps qs∗ = ϕα (Y )Mα
α∈A

734
37.2. Fields of Definition Chapter 37. Descent

where ϕα (Y ) are linear functions of Y with coefficients in L itself. Consider the following
linear system over L:

ϕγ0 (Y ) = 1, ϕα (Y ) = 0 for α ∈ A r {γ0 }.

Then {yiα }1≤i≤s,α∈A gives a solution to the system over K, but by linear algebra, there also
exists a solution {ȳiα }1≤i≤s,α∈A over L. Evaluating the system at this solution yields
X
Mγ0 − ḡβγ0 Mβ = p1 q̄1 + . . . + ps q̄s ∈ I
β∈B

¯ 0 Mβ mod I so by linear independence,


where ḡβγ0 ∈ L and q̄i = qi∗ (Y ). Then Mγ0 ≡ β∈B βγ
P
ḡβγ0 = gβγ0 for all β ∈ B. Since γ0 was arbitrary, we conclude that gβγ = ḡβγ ∈ L for all
β ∈ B, γ ∈ C. Hence k0 ⊆ L.
For the final statement, note that σ ∈ Aut(K) acts on I via its action on the coefficients
gβγ , so the equations in the above paragraph show that I σ = I if and only if σ|k0 = 1.
In practice, this proof is useless for explicitly finding k0 . Instead, Galois descent can be
used to give a more meaningful description of k0 .
Definition. Such a field k0 as in Theorem 37.2.1 is called a minimal field of definition
for I.
Definition. For a K-vector space V and a subfield k ⊆ K, a k-structure on V is a k-
subspace Vk ⊆ V such that Vk ⊗k K ∼= V as K-vector spaces. Similarly, a k-structure on a
K-algebra A is a subalgebra Ak ⊆ A such that Ak ⊗k K ∼
= A as K-algebras.
Suppose K/k is a Galois extension with group G. From Theorem 37.1.9, we know that
giving a k-structure on a vector space V is the same as defining a semilinear G-action on
V . Similarly, giving a k-structure on an algebra A is equivalent to defining a multiplication-
preserving semilinear G-action on A. The following results are easily proven from the defi-
nitions of k-structure.
Lemma 37.2.2. Suppose I ⊂ K[x1 , . . . , xn ] is an ideal defined over a subfield k ⊆ K. Then
AK := k[x1 , . . . , xn ]/Ik is a k-structure on A := K[x1 , . . . , xn ]/I.
Definition. Suppose V is a K-vector space with k-structure Vk . Then a K-subspace W ⊆ V
is defined over k if W ∩ Vk spans W over K.
Lemma 37.2.3. Suppose K/k is a Galois extension with group G. Then for a vector space
V with k-structure Vk and a subspace W ⊆ V , W is defined over k if and only if W G = W .
Definition. Suppose V and W are K-vector spaces with k-structures Vk and Wk , respectively,
and f : V → W is a K-linear function. Then f is defined over k if f (Vk ) ⊆ Wk . The
same definition goes for homomorphisms of K-algebras.
Lemma 37.2.4. If K/k is Galois with group G, then a k-linear map f : V → W (or a
homomorphism f : A → B of K-algebras) is defined over k if and only if f σ = f for all
σ ∈ G.

735
37.3. Galois Descent for Varieties and Schemes Chapter 37. Descent

37.3 Galois Descent for Varieties and Schemes


Definition. Suppose X ⊆ AnK is an affine algebraic set with vanishing ideal I = I(X) ⊆
K[x1 , . . . , xn ]. Then X is defined over a subfield k ⊆ K, or k is a field of definition for
X, if I is defined over k.
Remark. Notice that the property of being defined over a subfield depends on the explicit
embedding X ,→ AnK . We will see later that this definition can be made intrinsic.
Lemma 37.3.1. Let k sep be the separable closure of a field k, set G = Gal(k sep /k), let k̄ be
the algebraic closure of k and suppose I ⊂ k̄[x1 , . . . , xn ] is an ideal. Then the following are
equivalent:
(a) I is defined over k.

(b) I is defined over k sep and is G-invariant.


Definition. Let k ⊆ K be a subfield and X ⊆ AnK an affine algebraic set. Then X is
k-closed if it can be defined by equations f1 = 0, . . . , fr = 0 for fi ∈ k[x1 , . . . , xn ].
Remark. The property of X being k-closed is in general not the same as X being defined over
k, which can be expressed by saying that the vanishing ideal I(X) = rad((f1 , . . . , fr )K[x1 , . . . , xn ])
is defined over k.
Example 37.3.2. When char k = 0 and K = k sep , the properties of being k-closed and
defined over k are equivalent.
Lemma 37.3.3. Suppose X ⊆ AnK is an affine algebraic set which is k-closed for some
k ⊆ K. Then X is defined over k if and only if X is reduced.
Proof. Saying X is defined over k is equivalent to saying the natural surjection
k[x1 , . . . , xn ] K[x1 , . . . , xn ] K[x1 , . . . , xn ]
= −→
(I(X)k K[x1 , . . . , xn ]) ⊗k K (I(X)k K[x1 , . . . , xn ]) I(X)
is an isomorphism. On the other hand, its kernel is precisely the nil radical of the quotient
ring K[x1 , . . . , xn ]/(I(X)k K[x1 , . . . , xn ]) which is 0 if and only if X is reduced.
Next, consider an affine scheme X = Spec A over Spec K and let k ⊂ K be a subfield.
Definition. If A has a k-structure Ak , we define a k-topology on X = Spec A by declaring
{V (a) | a ⊂ A is defined over k} to be the closed sets. Equivalently, this topology can be
specified by declaring {D(f ) | f ∈ Ak } to be a basis of open sets.
For each f ∈ Ak , set OX,k (D(f )) := (Ak )f . Since {D(f ) | f ∈ Ak } is a basis for the
k-topology on X, this defines a sheaf of rings OX,k on X with respect to the k-topology.
This proves:
Lemma 37.3.4. Suppose A is a K-algebra with k-structure Ak and X = Spec A. Then
(X, OX,k ) is a k-structure on the ringed space (X, OX ).

736
37.3. Galois Descent for Varieties and Schemes Chapter 37. Descent

In general, we define:
Definition. A k-structure on a scheme X over K is the data of a k-topology on X,
i.e. a collection of open sets Topk (X) ⊆ Top(X), and for each k-open set U ∈ Topk (X), a
k-structure on the K-algebra OX (U ), which are subject to the following conditions:
(1) For every inclusion of k-open sets V ,→ U , the morphism OX (U ) → OX (V ) is defined
over k.
S
(2) For any affine open cover X = Ui , the subspace k-topology on each Ui and the
specified k-structure on each OX (Ui ) agree with the k-structure on the affine scheme
(Ui , OUi ,k ) defined above.
Set G = Gal(K/k). For a K-scheme X, let F (K/k, X) denote the collection of k-
structures on X. Then there is a map
Θ : F (K/k, X) −→ H 1 (G, AutK (X)).
(To ensure the G-actions are continuous, we will always assume X is a scheme of finite type
over K.) By Lemma 37.1.8, Θ is always injective. To show Θ is surjective, one might try to
construct a Galois twist of X(K), but in general the fixed points of this space may be too
small to recover X under base change – in many cases X(K)G is even empty! The solution for
affine schemes at least is to define a new G-action on the algebra K[X] of regular functions
on X. Explicitly, for f ∈ K[X], σ ∈ G and ξ ∈ H 1 (G, AutK (K[X])), set
(σ ∗ f )(x) = σ(f (σ −1 ∗ x)) = σf (σ −1 (ξ(σ)−1 x)) = σ(f )ξ(σ)−1 x.

This can be written σ ∗ f = ξ(σ)σ(f


g ) where ∼ is the map AutK (X) → AutK (K[X]) sending
−1
g 7→ g̃ defined by (g̃ · f )(x) = f (g x). Now take Xk to be the affine K-scheme with algebra
of regular functions K[Xk ] = K[X]G∗ . Then Xk defines a k-structure on X.
Example 37.3.5. Consider the linear algebraic group
  
x 0
T = Gm = ∈ GL2 (K) ,
0 x−1
which as an affine scheme corresponds to the algebra K[T ] = K[x, x−1 ]. Then as an algebraic
group, T has automorphism group AutK (T ) = {1, γ} where γ(t) = t−1 . In this case, γ̃ : x 7→
x−1 . For a Galois extension K/k with G = Gal(K/k), a 1-cocycle ξ ∈ H 1 (G, AutK (T )) is
simply a group homomorphism ξ : G → Aut√K (T ). Set H = ker ξ ⊆ G, which is a subgroup
of index 2 as long as ξ 6= 1. Then K H = k( a) for some a ∈ k × and G acts on the algebra
K[T ] via σ ∗ a = aσ and σ ∗ f = ξ(σ)σ(f
g ), where
(
g = x 7→ x,
ξ(σ)
if ξ(σ) = 1
−1
x 7→ x , if ξ(σ) = γ.

So K[T ]G∗ = (K[T ]H∗ )G/H∗ = k( a)[x, x−1 ]G/H∗ . This quotient
√ group is a 2-group G/H =
−1
{1, τ } and the nontrivial element acts on the elements of k( a)[x, x ] by
d
X d
X
τ∗ n
an x = τ (an )x−n .
n=−d n=−d

737
37.3. Galois Descent for Varieties and Schemes Chapter 37. Descent

So dn=−d an xn is a fixed point of the twisted G/H-action if and only if τ (an ) =


P
n −n
√ a−n for all
n ∈ Z. The algebra of such Laurent series is generated by {bx +τ (b)x | b ∈ k( a), n ≥ 0}.
One can in fact show that this algebra is precisely k[s, t] where

x + x−1 x − x−1
s= and t= √ .
2 2 a

So K[T ]G∗ = k[s, t]. Noting that s2 − at2 = 1, we can rewrite this algebra as

k[s, t] = k[S, T ]/(S 2 − aT 2 − 1)

which is the ring of regular functions for the 1-dimensional k-scheme


  
u av 2 2
Ta := ∈ GL2 (k) : u − av = 1 .
v u
√ √ √
In fact, Ta (k( a)) ∼
= Gm (k( a)), but if a 6∈ k then Ta 6∼ = Gm (k). Thus each a ∈ k × r (k × )2
gives a distinct isomorphism class of twists of the torus Gm . One can even check that the
Hopf algebra structure on K[T ] descends to k[s, t], so Ta is even an algebraic group over k.

Example 37.3.6. Using the notation of Example 37.1.12, let A = a,b



k
be a quaternion
algebra over a field k of characteristic other√than 2. Recall that A is a K/k-form of M2 (L)
for the Galois extension L/k where L = k( a) and G = Gal(L/k) = {1, σ}. Explicitly, A
corresponds to an equivalence class of cocycles ξ ∈ H 1 (G, P GL2 (L)) = H 1 (G, AutL (M2 (L)))
where  
0 b
ξ(σ) = ∈ P GL2 (L).
1 0
At the same time, P GL2 = Aut(P1 ) so the cocycle ξ defines a twist of P1 , called a Severi-
Brauer curve. Here’s how to construct it. Take U = P1 r {0, ∞}, which is an affine curve
invariant under ξ(σ). Then as we saw in Example 37.3.5, L[U ] = L[x, x−1 ]; define a new
action of G on L[U ] by
√ √
σ ∗ a = σ( a) and σ ∗ x = bx−1 .

One can show that L[U ]G∗ = k[s, t] where

x + bx−1 x − bx−1
s= and t= √ .
2 2 a

Note that s2 − at2 = b, so we may instead write L[U ]G∗ = k[S, T ]/(S 2 − aT 2 − b). Further,
note that the affine equation S 2 − aT 2 = b has a solution over k if and only if A splits over
k. The homogeneous equation
X 2 − aY 2 − bZ 2 = 0
defines a smooth, projective curve SB(A) over k, which is the aforementioned Severi-Brauer
curve for A. Note that SB(A) is a k-structure for P1L and SB(A) has a k-rational point
exactly when A splits over k. More broadly:

738
37.3. Galois Descent for Varieties and Schemes Chapter 37. Descent

Theorem 37.3.7 (Amitsur). Suppose A and B are quaternion algebras over k with Severi-
Brauer curves SB(A) and SB(B), respectively. Then A ∼
= B over k if and only if SB(A)
and SB(B) are birationally equivalent over k.
Example 37.3.8. For a central simple algebra A of degree n > 2, one can similarly construct
a Severi-Brauer variety SB(A). The most general form of Amitsur’s Theorem says that if
SB(A) and SB(B) are birational as k-varieties, then [A] and [B] generate the same cyclic
subgroup in Br(k). Whether the converse to this statement holds is an open question.
Question. Suppose X is an algebraic variety over K and k ⊂ K is a subfield. Is there a
variety Y over K such that X ∼
= Y as K-varieties and Y is defined over K?
In the case of an affine algebraic variety X ⊆ AnK , if there exists such a variety Y ⊆ Am
K (a

priori m 6= n) that has a k-structure, then we can twist K[X] = K[Y ] to obtain a k-structure
on X. In general, this is not possible, but there are some conditions one can place on the
variety to allow for descent. The following examples illustrate these conditions for quadratic
forms.
Example 37.3.9. First consider the quadratic form x2 + 2iy 2 over Q(i). Notice that the
linear change of variables u = x, v = (1 + i)y
√ yields form uv + v 2 over Q. On
a quadratic √
2 2
the other hand, one can show that q = x − 5y defined over Q( 5) is not equivalent to a
quadratic form q0 over Q. Let

 
1 0

Q= ∈ GL2 (Q( 5))
0 − 5
be the symmetric matrix corresponding to q. If q were equivalent to a quadratic form
over Q, with √matrix Q0 ∈ GL2 (Q), then we would be able√ to write Q = B t Q0 B for some
B ∈ GL2 (Q( √ 5)). Taking determinants, we see that − 5 =√rs where r = det(B)2 is
a√square in Q(√ 5) and s = det(Q0 ) ∈ Q. Write r = (α + β 5)2√for α, β ∈ Q. Then
− 5 = (α + β 5)2 s and applying the nontrivial element τ ∈ Gal(Q( 5)/Q) yields
√ √
5 = (τ (α) − τ (β) 5)2 s.
√ 2 √ 2
Combining these
√ expressions, we get that (α + β 5) = −(τ (α) − τ (β) 5) , which is im-
possible in Q( 5). Therefore no such Q0 exists.
More generally, suppose K/k is a Galois extension with group G and Q is a symmetric
matrix in GLn (K) for which Q = B t Q0 B for some Q0 ∈ GLn (k) and B ∈ GLn (K). Then
for any σ ∈ G,
Qσ = (B σ )t Q0 B σ = A(σ)t QA(σ)
for A(σ) = B −1 B σ . Notice that the A(σ) satisfy A(στ ) = A(σ)σ(A(τ )) for all σ, τ ∈ G,
so they define a 1-cocycle in Z 1 (G, GLn (K)). By Theorem 37.1.19, any such cocycle is a
coboundary, which reflects the fact that A(σ) = B −1 B σ .
Example 37.3.10. For the quadratic form q = x2 + 2iy 2 over Q(i) above, its symmetric
matrix  
1 0
Q= ∈ GL2 (Q(i))
0 2i

739
37.3. Galois Descent for Varieties and Schemes Chapter 37. Descent

has Galois conjugate  


1 0
σ
Q =
0 −2i
where Gal(Q(i)/Q) = hσi ∼
= Z/2Z, and one can see that Q ∼ Qσ via the cocycle
 
1 0
A(σ) = .
0 i

Then A(1) = A(σ 2 ) = A(σ)σ(A(σ)) = I2 , the 2 × 2 identity matrix, so Q descends to a


matrix over Q and therefore√
descent is possible
√ for the quadratic form q.
2 2
Meanwhile, for q = x − 5y over Q( 5), the symmetric matrix
 
1 0

Q=
0 − 5

has Galois conjugate  


τ1 √0
Q =
0 5

and one can show that Q is not similar to Qτ over Q( 5). Thus descent is not possible in
this case.

Example 37.3.11. The previous example shows that having a similarity Q ∼ Qσ for some
Galois conjugate of the symmetric matrix is a necessary condition for descent of a quadratic
√ 2
2
√ sufficient. Consider the same quadratic form q = x − 5y
form, but it turns out that it is not
but now over the field K = Q( 5, i). Set k = Q(i) and G = Gal(K/k) = hτ i ∼ = Z/2Z. Then
Q and Qτ are similar via the cocycle
 
1 0
A(τ ) = ∈ GL2 (K).
0 i

Suppose q0 is a quadratic form which is equivalent to q over Q(i). Then taking determinants
of Q = B t Q0 B, we once again have equations
√ √
− 5 = (α + βi)2 s and 5 = (τ (α) + τ (β)i)2 s

for α, β ∈ Q( 5). These imply (α + βi)2 = −(τ (α) + τ (β)i)2 , or

α + βi = i(τ (α) + τ (β)i) or α + βi = −i(τ (α) + τ (β)i).

In the first case, we must have τ (α) = β and τ (β) = −α, which implies τ 2 = −1, a
contradiction. The second case produces the same contradiction. Therefore descent is not
possible in this case.

The correct characterization of descent for quadratic forms is contained in the following
proposition.

740
37.3. Galois Descent for Varieties and Schemes Chapter 37. Descent

Proposition 37.3.12. Let K/k be a Galois extension with Galois group G, let Q ∈ GLn (K)
be a symmetric matrix representing a quadratic form q over K and assume that for every
σ ∈ G, there exists a 1-cocycle A ∈ Z 1 (G, GLn (K)) such that Qσ = A(σ)t QA(σ). Then q
descends to a quadratic form on k.

Proof. Let V = K n . It will be enough to define a k-structure Vk on V such that Q = B t Q0 B


for some Q0 ∈ Aut(Vk ) and B ∈ GL2 (K), since then Q0 defines a quadratic form which is
equivalent to q|Vk . For σ ∈ G and v ∈ V , we have:

σ(q(v)) = σ(v t Qv) = σ(v)t Qσ σ(v)


= σ(v)t A(σ)t QA(σ)σ(v)
= q(A(σ)σ(v)).

This suggests defining a twisted G-action on V by σ ∗ v = A(σ)σ(v). Then as we have seen


before, Vk = V G∗ is a k-structure on V and the rest follows.

741

You might also like