geir ellingsrud and john christian ottem october 15, 2018€¦ · between algebra and geometry: a...

308
Introduction to Schemes Geir Ellingsrud and John Christian Ottem October 15, 2018

Upload: others

Post on 22-Sep-2020

1 views

Category:

Documents


0 download

TRANSCRIPT

Page 1: Geir Ellingsrud and John Christian Ottem October 15, 2018€¦ · between algebra and geometry: a ne varieties correspond to nitely generated k-algebras without zero-divisors, and

Introduction to Schemes

Geir Ellingsrud and John Christian Ottem

October 15, 2018

Page 2: Geir Ellingsrud and John Christian Ottem October 15, 2018€¦ · between algebra and geometry: a ne varieties correspond to nitely generated k-algebras without zero-divisors, and
Page 3: Geir Ellingsrud and John Christian Ottem October 15, 2018€¦ · between algebra and geometry: a ne varieties correspond to nitely generated k-algebras without zero-divisors, and

Contents

0.1 Introduction to the schemes formalism . . . . . . . . . . . . . . 11

0.1.1 Nilpotents and zero-divisors . . . . . . . . . . . . . . . . 11

0.1.2 Gluing . . . . . . . . . . . . . . . . . . . . . . . . . . . . 12

0.1.3 General base rings . . . . . . . . . . . . . . . . . . . . . 12

0.1.4 Prime ideals, rather than maximal ideals. . . . . . . . . . 13

1 Sheaves 15

1.1 Sheaves and presheaves . . . . . . . . . . . . . . . . . . . . . . . 15

1.1.1 Presheaves . . . . . . . . . . . . . . . . . . . . . . . . . . 16

1.1.2 Sheaves . . . . . . . . . . . . . . . . . . . . . . . . . . . 17

1.1.3 Subsheaves and saturation of subpresheaves . . . . . . . 18

1.2 A bunch of examples . . . . . . . . . . . . . . . . . . . . . . . . 19

1.3 Stalks . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 21

1.3.1 The germ of a section . . . . . . . . . . . . . . . . . . . . 22

1.3.2 A primer on limits . . . . . . . . . . . . . . . . . . . . . 22

1.4 Morphisms between (pre)sheaves . . . . . . . . . . . . . . . . . 23

1.5 Kernels and images . . . . . . . . . . . . . . . . . . . . . . . . . 25

1.6 Exact sequences of sheaves. . . . . . . . . . . . . . . . . . . . . 27

1.7 B-sheaves . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 30

1.8 A family of examples – Godement sheaves . . . . . . . . . . . . 31

1.8.1 Skyscraper sheaves. . . . . . . . . . . . . . . . . . . . . . 32

1.8.2 The Godement sheaf associated with a presheaf . . . . . 33

1.9 Sheafification . . . . . . . . . . . . . . . . . . . . . . . . . . . . 34

1.10 Quotient sheaves and cokernels . . . . . . . . . . . . . . . . . . 36

1.11 The pushforward of a sheaf . . . . . . . . . . . . . . . . . . . . . 36

1.12 Inverse image sheaf . . . . . . . . . . . . . . . . . . . . . . . . . 38

1.12.1 Adjoint property . . . . . . . . . . . . . . . . . . . . . . 38

2 Schemes 41

2.1 The spectrum of a ring . . . . . . . . . . . . . . . . . . . . . . . 41

2.1.1 Generic points . . . . . . . . . . . . . . . . . . . . . . . . 44

2.1.2 Irreducible closed subsets . . . . . . . . . . . . . . . . . . 44

2.2 Functoriality . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 45

3

Page 4: Geir Ellingsrud and John Christian Ottem October 15, 2018€¦ · between algebra and geometry: a ne varieties correspond to nitely generated k-algebras without zero-divisors, and

Contents

2.3 First examples . . . . . . . . . . . . . . . . . . . . . . . . . . . . 46

2.3.1 The world of schemes vs the world of varieties . . . . . . 48

2.4 Distinguished open sets . . . . . . . . . . . . . . . . . . . . . . . 49

2.4.1 Basic properties of distinguished sets . . . . . . . . . . . 50

2.5 Examples . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 50

2.6 The structure sheaf on SpecA . . . . . . . . . . . . . . . . . . . 53

2.6.1 Motivation . . . . . . . . . . . . . . . . . . . . . . . . . . 54

2.6.2 Definition of the structure sheaf OSpecA . . . . . . . . . . 54

2.6.3 Maps between the structure sheaves of two spectra . . . 57

2.7 Schemes . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 58

2.8 Open immersions and open subschemes . . . . . . . . . . . . . . 60

2.9 Residue fields . . . . . . . . . . . . . . . . . . . . . . . . . . . . 60

2.10 Relative schemes . . . . . . . . . . . . . . . . . . . . . . . . . . 62

3 Gluing 65

3.1 Gluing sheaves . . . . . . . . . . . . . . . . . . . . . . . . . . . 65

3.2 Gluing maps of sheaves . . . . . . . . . . . . . . . . . . . . . . . 67

3.3 The gluing of schemes . . . . . . . . . . . . . . . . . . . . . . . 67

3.4 Global sections of glued schemes . . . . . . . . . . . . . . . . . . 69

3.5 The gluing of morphisms . . . . . . . . . . . . . . . . . . . . . . 70

3.6 The category of affine schemes . . . . . . . . . . . . . . . . . . . 70

3.7 Universal properties of maps into affine schemes . . . . . . . . . 71

3.8 The functor from varieties to schemes . . . . . . . . . . . . . . . 73

4 More examples 75

4.1 A scheme that is not affine . . . . . . . . . . . . . . . . . . . . . 75

4.2 The projective line . . . . . . . . . . . . . . . . . . . . . . . . . 76

4.2.1 Affine line a doubled origin . . . . . . . . . . . . . . . . . 77

4.3 The blow-up of the affine plane . . . . . . . . . . . . . . . . . . 77

4.3.1 Blow-up as a variety . . . . . . . . . . . . . . . . . . . . 77

4.3.2 The blow-up as a scheme . . . . . . . . . . . . . . . . . . 78

4.4 Hyperelliptic curves . . . . . . . . . . . . . . . . . . . . . . . . . 80

5 Geometric properties of schemes 81

5.1 Some topological properties . . . . . . . . . . . . . . . . . . . . 81

5.2 Properties of the scheme structure . . . . . . . . . . . . . . . . . 82

5.2.1 Integral schemes . . . . . . . . . . . . . . . . . . . . . . 83

5.2.2 Noetherian spaces and quasi-compactness . . . . . . . . . 84

5.3 The dimension of a scheme . . . . . . . . . . . . . . . . . . . . . 86

5.4 More examples . . . . . . . . . . . . . . . . . . . . . . . . . . . 87

5.4.1 A more fancy example . . . . . . . . . . . . . . . . . . . 88

4

Page 5: Geir Ellingsrud and John Christian Ottem October 15, 2018€¦ · between algebra and geometry: a ne varieties correspond to nitely generated k-algebras without zero-divisors, and

Contents

6 Projective schemes 91

6.1 Motivation . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 91

6.2 Basic remarks on graded rings . . . . . . . . . . . . . . . . . . . 92

6.2.1 Localization . . . . . . . . . . . . . . . . . . . . . . . . . 93

6.3 The Proj construction . . . . . . . . . . . . . . . . . . . . . . . 93

6.3.1 ProjR as a scheme . . . . . . . . . . . . . . . . . . . . . 96

6.4 Functoriality . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 98

6.4.1 Closed immersions . . . . . . . . . . . . . . . . . . . . . 99

6.5 The Veronese embedding . . . . . . . . . . . . . . . . . . . . . . 99

6.5.1 Remark on rings generated in degree 1 . . . . . . . . . . 100

6.5.2 The Blow-up as a Proj . . . . . . . . . . . . . . . . . . . 100

7 Fiber products 103

7.1 Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 103

7.1.1 Fiber products of sets. . . . . . . . . . . . . . . . . . . . 103

7.1.2 The fiber product in general categories . . . . . . . . . . 104

7.2 Products of affine schemes . . . . . . . . . . . . . . . . . . . . . 105

7.3 A useful lemma . . . . . . . . . . . . . . . . . . . . . . . . . . . 107

7.4 The gluing process . . . . . . . . . . . . . . . . . . . . . . . . . 108

7.5 The final reduction . . . . . . . . . . . . . . . . . . . . . . . . . 109

7.6 Notation. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 110

7.7 Examples . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 110

7.7.1 Varieties versus Schemes . . . . . . . . . . . . . . . . . . 110

7.7.2 Non-algebraically closed fields . . . . . . . . . . . . . . . 111

7.8 Base change . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 113

7.9 Scheme theoretic fibres . . . . . . . . . . . . . . . . . . . . . . . 114

7.10 Separated schemes . . . . . . . . . . . . . . . . . . . . . . . . . 117

7.10.1 The diagonal . . . . . . . . . . . . . . . . . . . . . . . . 117

7.10.2 Separated schemes . . . . . . . . . . . . . . . . . . . . . 118

7.11 The valuative criterion . . . . . . . . . . . . . . . . . . . . . . . 119

7.11.1 Varieties vs schemes again . . . . . . . . . . . . . . . . . 121

8 Quasi-coherent sheaves 123

8.1 Sheaves of modules . . . . . . . . . . . . . . . . . . . . . . . . . 124

8.1.1 The support of a sheaf . . . . . . . . . . . . . . . . . . . 127

8.2 Pushforward and Pullback of OX-modules . . . . . . . . . . . . 127

8.2.1 Pushforward . . . . . . . . . . . . . . . . . . . . . . . . . 127

8.2.2 Pullback . . . . . . . . . . . . . . . . . . . . . . . . . . . 127

8.2.3 Adjoint properties of f∗, f∗ . . . . . . . . . . . . . . . . . 128

8.2.4 Pullback of sections . . . . . . . . . . . . . . . . . . . . . 128

8.2.5 Tensor products . . . . . . . . . . . . . . . . . . . . . . . 129

8.3 Quasi-coherent sheaves . . . . . . . . . . . . . . . . . . . . . . . 129

5

Page 6: Geir Ellingsrud and John Christian Ottem October 15, 2018€¦ · between algebra and geometry: a ne varieties correspond to nitely generated k-algebras without zero-divisors, and

Contents

8.4 Quasi-coherent sheaves on affine schemes . . . . . . . . . . . . . 130

8.4.1 Functoriality . . . . . . . . . . . . . . . . . . . . . . . . 133

8.5 Quasi-coherent sheaves on general schemes . . . . . . . . . . . . 134

8.6 Coherent sheaves . . . . . . . . . . . . . . . . . . . . . . . . . . 137

8.7 Functoriality . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 138

8.7.1 Quasi-coherence of pullbacks . . . . . . . . . . . . . . . . 139

8.7.2 Quasi-coherence of the direct image . . . . . . . . . . . . 140

8.7.3 Coherence of the direct image . . . . . . . . . . . . . . . 142

8.8 The categories of coherent and quasi-coherent sheaves . . . . . . 142

8.9 Closed immersions and subschemes . . . . . . . . . . . . . . . . 143

8.9.1 Morphisms to a closed subscheme . . . . . . . . . . . . . 145

8.9.2 Induced reduced scheme structure . . . . . . . . . . . . . 145

9 Vector bundles and locally free sheaves 147

9.1 Locally free sheaves and projective modules (work in progress) . 149

9.2 Invertible sheaves and the Picard group . . . . . . . . . . . . . . 149

9.2.1 Operations on locally free sheaves . . . . . . . . . . . . . 151

9.3 Vector bundles . . . . . . . . . . . . . . . . . . . . . . . . . . . 152

9.3.1 The sheaf of sections of a vector bundle . . . . . . . . . . 154

9.3.2 Equivalence between vector bundles and locally free sheaves154

9.4 Extended example: The tautological bundle on Pnk . . . . . . . . 155

10 Sheaves on projective schemes 159

10.1 The graded ∼-functor . . . . . . . . . . . . . . . . . . . . . . . . 160

10.2 Serre’s twisting sheaf O(1) . . . . . . . . . . . . . . . . . . . . . 161

10.3 Extending sections of invertible sheaves . . . . . . . . . . . . . . 163

10.4 The associated graded module . . . . . . . . . . . . . . . . . . . 164

10.4.1 Sections of the structure sheaf . . . . . . . . . . . . . . . 165

10.4.2 The homomorphism α. . . . . . . . . . . . . . . . . . . . 166

10.4.3 The map β . . . . . . . . . . . . . . . . . . . . . . . . . 167

10.5 Closed subschemes of ProjR . . . . . . . . . . . . . . . . . . . . 169

10.6 The Segre embedding . . . . . . . . . . . . . . . . . . . . . . . . 170

10.7 Two important exact sequences . . . . . . . . . . . . . . . . . . 172

10.7.1 Hypersurfaces . . . . . . . . . . . . . . . . . . . . . . . . 172

10.7.2 Complete intersections . . . . . . . . . . . . . . . . . . . 172

11 Differentials 175

11.1 Tangent vectors and derivations . . . . . . . . . . . . . . . . . . 175

11.2 Regular schemes and the Zariski tangent space . . . . . . . . . . 177

11.2.1 Zariski tangent space and the ring of dual numbers . . . 178

11.3 Derivations and Kahler differentials . . . . . . . . . . . . . . . . 179

11.3.1 The module of Kahler differentials . . . . . . . . . . . . . 180

6

Page 7: Geir Ellingsrud and John Christian Ottem October 15, 2018€¦ · between algebra and geometry: a ne varieties correspond to nitely generated k-algebras without zero-divisors, and

Contents

11.4 Properties of Kahler differentials . . . . . . . . . . . . . . . . . . 181

11.4.1 Localization . . . . . . . . . . . . . . . . . . . . . . . . . 182

11.4.2 Base change . . . . . . . . . . . . . . . . . . . . . . . . . 183

11.4.3 Exact sequences . . . . . . . . . . . . . . . . . . . . . . . 183

11.4.4 The diagonal and ΩB/A . . . . . . . . . . . . . . . . . . . 184

11.5 Some explicit computations . . . . . . . . . . . . . . . . . . . . 185

11.6 The sheaf of differentials . . . . . . . . . . . . . . . . . . . . . . 187

11.7 The sheaf of differentials of PnA . . . . . . . . . . . . . . . . . . . 189

11.8 Relation with the Zariski tangent space . . . . . . . . . . . . . . 190

11.8.1 Smooth morphisms . . . . . . . . . . . . . . . . . . . . . 193

11.9 The conormal sheaf . . . . . . . . . . . . . . . . . . . . . . . . . 194

12 First steps in sheaf cohomology 195

12.1 Some homological algebra . . . . . . . . . . . . . . . . . . . . . 196

12.2 Cech cohomology . . . . . . . . . . . . . . . . . . . . . . . . . . 198

12.3 Cohomology of sheaves on affine schemes . . . . . . . . . . . . . 202

12.3.1 Cech cohomology and affine coverings . . . . . . . . . . . 205

12.4 Chomology and dimension . . . . . . . . . . . . . . . . . . . . . 205

12.5 Cohomology of sheaves on projective space . . . . . . . . . . . . 206

12.6 Extended example: Plane curves . . . . . . . . . . . . . . . . . . 209

12.7 Extended example: The twisted cubic in P3 . . . . . . . . . . . 210

13 More on sheaf cohomology 213

13.1 Flasque sheaves . . . . . . . . . . . . . . . . . . . . . . . . . . . 213

13.2 The Godement resolution . . . . . . . . . . . . . . . . . . . . . . 215

13.2.1 Functoriality . . . . . . . . . . . . . . . . . . . . . . . . 216

13.2.2 Sheaf cohomology . . . . . . . . . . . . . . . . . . . . . . 217

13.2.3 Acyclic sheaves . . . . . . . . . . . . . . . . . . . . . . . 220

13.2.4 The Cech resolution . . . . . . . . . . . . . . . . . . . . 222

13.3 Godement vs. Cech . . . . . . . . . . . . . . . . . . . . . . . . 223

14 Divisors and linear systems 225

14.1 Weil divisors . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 227

14.1.1 The divisor of a rational function . . . . . . . . . . . . . 228

14.1.2 The divisor of a section of an invertible sheaf . . . . . . . 229

14.1.3 The class group . . . . . . . . . . . . . . . . . . . . . . . 230

14.1.4 Projective space . . . . . . . . . . . . . . . . . . . . . . . 230

14.1.5 Weil divisors and invertible sheaves . . . . . . . . . . . . 231

14.1.6 The class group and unique factorization domains . . . . 232

14.1.7 A useful exact sequence . . . . . . . . . . . . . . . . . . 233

14.2 Cartier divisors . . . . . . . . . . . . . . . . . . . . . . . . . . . 234

14.2.1 Cartier divisors and invertible sheaves . . . . . . . . . . 235

7

Page 8: Geir Ellingsrud and John Christian Ottem October 15, 2018€¦ · between algebra and geometry: a ne varieties correspond to nitely generated k-algebras without zero-divisors, and

Contents

14.2.2 CaCl(X) and Pic(X) . . . . . . . . . . . . . . . . . . . . 236

14.2.3 From Cartier to Weil . . . . . . . . . . . . . . . . . . . . 237

14.2.4 Projective space . . . . . . . . . . . . . . . . . . . . . . . 238

14.3 Effective divisors . . . . . . . . . . . . . . . . . . . . . . . . . . 239

14.3.1 Linear systems . . . . . . . . . . . . . . . . . . . . . . . 240

14.4 Examples . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 241

14.4.1 The quadric surface . . . . . . . . . . . . . . . . . . . . . 241

14.4.2 The quadric cone . . . . . . . . . . . . . . . . . . . . . . 243

14.4.3 Quadric hypersurfaces in higher dimension . . . . . . . . 245

14.4.4 Hirzebruch surfaces . . . . . . . . . . . . . . . . . . . . . 246

15 Two applications of cohomology 249

15.1 Vector bundles on the projective line . . . . . . . . . . . . . . . 249

15.2 Non-split vector bundles on Pn . . . . . . . . . . . . . . . . . . . 251

16 Maps to projective space 253

16.1 Globally generated sheaves . . . . . . . . . . . . . . . . . . . . . 254

16.1.1 Pullbacks and pushforwards . . . . . . . . . . . . . . . . 254

16.2 Morphisms to projective space . . . . . . . . . . . . . . . . . . . 255

16.2.1 Example: The twisted cubic . . . . . . . . . . . . . . . . 255

16.2.2 Morphisms and globally generated invertible sheaves . . 256

16.3 Application: Automorphisms of Pn . . . . . . . . . . . . . . . . 258

16.4 Projective embeddings . . . . . . . . . . . . . . . . . . . . . . . 259

16.5 Projective space as a functor . . . . . . . . . . . . . . . . . . . . 261

16.5.1 Generalizations . . . . . . . . . . . . . . . . . . . . . . . 262

17 More on projective schemes 265

17.1 Ample invertible sheaves and Serre’s theorems . . . . . . . . . . 265

17.1.1 Euler characteristic . . . . . . . . . . . . . . . . . . . . . 267

17.2 Proper and projective schemes . . . . . . . . . . . . . . . . . . . 267

18 The Riemann–Roch theorem and Serre duality 269

18.1 Curves . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 269

18.2 Divisors on Curves . . . . . . . . . . . . . . . . . . . . . . . . . 269

18.2.1 The genus of a curve . . . . . . . . . . . . . . . . . . . . 270

18.3 The canonical divisor . . . . . . . . . . . . . . . . . . . . . . . . 271

18.4 Hyperelliptic curves . . . . . . . . . . . . . . . . . . . . . . . . . 271

18.5 The Riemann Roch theorem for curves . . . . . . . . . . . . . . 274

18.6 Proof of Serre duality for curves . . . . . . . . . . . . . . . . . . 277

18.6.1 Laurent tails . . . . . . . . . . . . . . . . . . . . . . . . . 280

18.6.2 Residues on A . . . . . . . . . . . . . . . . . . . . . . . . 280

18.6.3 Residues on X . . . . . . . . . . . . . . . . . . . . . . . 281

8

Page 9: Geir Ellingsrud and John Christian Ottem October 15, 2018€¦ · between algebra and geometry: a ne varieties correspond to nitely generated k-algebras without zero-divisors, and

18.6.4 The final part of the proof . . . . . . . . . . . . . . . . . 283

19 Applications of the Riemann–Roch theorem 287

19.1 Morphisms of curves . . . . . . . . . . . . . . . . . . . . . . . . 287

19.1.1 Morphisms of curves are finite . . . . . . . . . . . . . . . 288

19.1.2 Pullbacks of divisors . . . . . . . . . . . . . . . . . . . . 289

19.1.3 Morphisms to P1 . . . . . . . . . . . . . . . . . . . . . . 290

19.2 Very ampleness criteria . . . . . . . . . . . . . . . . . . . . . . . 291

19.3 Curves on P1 × P1 . . . . . . . . . . . . . . . . . . . . . . . . . 292

19.4 Curves of genus 0 . . . . . . . . . . . . . . . . . . . . . . . . . . 293

19.5 Curves of genus 1 . . . . . . . . . . . . . . . . . . . . . . . . . . 294

19.5.1 Divisors on X . . . . . . . . . . . . . . . . . . . . . . . . 295

19.6 Curves of genus 2 . . . . . . . . . . . . . . . . . . . . . . . . . . 297

19.7 Curves of genus 3 . . . . . . . . . . . . . . . . . . . . . . . . . . 298

19.8 Curves of Genus 4 . . . . . . . . . . . . . . . . . . . . . . . . . . 301

19.8.1 Classifying curves of genus 4 . . . . . . . . . . . . . . . . 301

A To be deleted 303

A.1 L.E.S. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 303

A.2 Pic(X) and H1(X,O×X) . . . . . . . . . . . . . . . . . . . . . . . 307

Page 10: Geir Ellingsrud and John Christian Ottem October 15, 2018€¦ · between algebra and geometry: a ne varieties correspond to nitely generated k-algebras without zero-divisors, and
Page 11: Geir Ellingsrud and John Christian Ottem October 15, 2018€¦ · between algebra and geometry: a ne varieties correspond to nitely generated k-algebras without zero-divisors, and

Introduction

0.1 Introduction to the schemes formalism

If X is an affine variety over an algebraically closed field k, then it has an

affine coordinate ring A(X) of regular functions X → k, and A(X) completely

characterises X up to isomorphism. This sets up a bijective correspondence

between algebra and geometry: affine varieties correspond to finitely generated

k-algebras without zero-divisors, and morphism between varieties corresponds to

k-algebra homomorphisms. The modern formulation of algebraic geometry, due

to Grothendieck, vastly generalizes this picture, replacing the k-algebras without

zero-divisors with the more general category of commutative rings.

This has the tremendous advantage that many of the arguments and intuition

carry over from the classical setting, and can be used to study more general

problems in other branches of mathematics. For instance, studying diophantine

equations over non-closed fields (such as Q) is clearly of interest in number

theory.

We should emphasise that the main goal of algebraic geometry is still to

understand algebraic varieties over a field. However, enlarging our category to

schemes offers a tremendous flexibility. Let us give a few examples of this.

0.1.1 Nilpotents and zero-divisors

Nilpotents and zero-divisors appear already in Bezout’s theorem: The conic

C = Z(y − x2) intersects a general line in two points, but the line y = 0 only at

the origin (with ‘multiplicity’ 2). Looking at the ideals, it is intuitive that the

intersection should correspond to the ideal (y − x2, y) = (x2, y) which gives us

the intersection number (’multiplicity 2’), but that ideal does not correspond to

an algebraic variety, as it is not radical.

Similar phenomena appear naturally when you take ‘limits’ (or ‘deformations’)

11

Page 12: Geir Ellingsrud and John Christian Ottem October 15, 2018€¦ · between algebra and geometry: a ne varieties correspond to nitely generated k-algebras without zero-divisors, and

0.1. Introduction to the schemes formalism

of algebraic varieties. As a basic example, consider the family of curves in A2

given by Vt = Z(y2− tx). Then for t 6= 0, Vt is a smooth conic, whereas for t = 0

it is a ‘double line’. The limit as t→ 0 is problematic; if you insist that it should

correspond to the affine variety Z(y), which has degree 1, then the degree of Vtis not a continuous function in t.

These problems vanish when passing to schemes: it is completely unproblem-

atic to work with geometric objects (affine schemes) associated to the coordinate

rings k[x, y]/(x2, y) or k[x, y]/(y2) respectively. Note however that we have to

adjust our viewpoint about regular functions: For instance, in the second example

the element ‘y’ is non-zero in the ring k[x, y]/(y2) - but it takes the value 0 at

each of the closed points of the underlying topological space (the x-axis).

0.1.2 Gluing

Just like in differential geometry, we do not like to consider manifolds as embedded

in some Rn (although Whitney’s theorem says that such an embedding exists). It

is much more fruitful to study manifolds M locally, using how it is glued together

by a bunch of open subsets in Rn. For instance, the concept of a tangent vector of

M at a point p can be defined using equivalence classes of locally parameterized

curves passing through p; glueing these sets together for each p, we obtain the

tangent bundle of M .

In a similar fashion, we will see that we get a lot of flexibility by talking

about varieties or schemes as glued objects rather than embedded objects in

some An or Pn. We can define the normalization of a variety by normalizing

each chart; we can talk about tangent and cotangent bundles by gluing together

local trivializations.

Moreover, by gluing together affine schemes we obtain much more interesting

spaces which may indeed fail to be embeddable in any Pn. And a posterori

non-quasi-projective varieties are not so obscure as one would think - they really

do occur in many contexts in algebraic geometry.

0.1.3 General base rings

Working with a non-algebraically closed base field is obviously important in

number theory: If we want to talk about rational points on the curve given

by the equation x3 + y3 = z3 we need to be able to think of elliptic curves as

schemes over the field Q. There is a naive fix, to just go to the algebraic closure,

but this misses the main points: The curves x3 + y3 = z3 and x3 + y3 = 9z3 are

isomorphic over Q, but the first has only a few trivial points over Q and the

second has infinitely many.

Going from fields to more general rings offers even greater flexibility. For

instance, the variety Z(y2 − x(x − 1)(x − t)) can be viewed simultaneously a

12

Page 13: Geir Ellingsrud and John Christian Ottem October 15, 2018€¦ · between algebra and geometry: a ne varieties correspond to nitely generated k-algebras without zero-divisors, and

Contents

surface S in A3k and a curve C of genus 1 curve over the ring k[t]. The k[t]-points

here correspond to rational curves on S. This sort of interplay is very useful in

many situations.

0.1.4 Prime ideals, rather than maximal ideals.

By the Nullstellensatz, the points of an affine variety are in bijection with the

maximal ideals of its coordinate ring. We will see that it is in fact much more

natural to include all the prime ideals of this ring. This is something that is

almost forced on us by functoriality: Think about the inclusion k[x] → k(x).

The maximal ideal (0) of k(x) pulls back to a prime ideal of k[x] which is not

maximal. Including the set of all prime ideals, we obtain a very nice category

which behaves essentially just like the category of commutative rings.

Beside of these categorical advantages, there are also a number of places in

more classical settings where thinking of points defined by prime ideals provides

an extra conceptual clarity. We will try to highlight a few examples where this

happens in Chapter 2.

13

Page 14: Geir Ellingsrud and John Christian Ottem October 15, 2018€¦ · between algebra and geometry: a ne varieties correspond to nitely generated k-algebras without zero-divisors, and
Page 15: Geir Ellingsrud and John Christian Ottem October 15, 2018€¦ · between algebra and geometry: a ne varieties correspond to nitely generated k-algebras without zero-divisors, and

Chapter 1

Sheaves

The concept of a sheaf was conceived in the German camp for prisoners of war

called Oflag XVII where French officers taken captive during the fighting in

France in the spring 1940 were imprisoned. Among them was the mathematician

and lieutenant Jean Leray. In the camp he gave a course in algebraic topology(!!)

during which he introduced some version of the theory of sheaves. He was aimed

at calculating the cohomology of a total space of a fibration in terms of invariants

of the base and the fibres (and naturally the fibration). To achieve this, in

addition to the concept of sheaves, he also invented spectral sequences.

After the war, Henri Cartan and Jean Pierre Serre developed the theory

further, and finally the theory was brought to the state as we know it today by

Alexandre Grothendieck.

1.1 Sheaves and presheaves

A common theme in mathematics is to study spaces by describing them through

their local properties. A manifold is a space which looks locally like Euclidean

space; a complex manifold is a space which is locally biholomorphic to Cn;

an algebraic variety is a space that looks locally like the zero set of a set

of polynomials. Here it is clear that point set topology alone is not enough

to fully capture these three notions: In each case, the space comes equipped

with a natural set of functions: C∞-functions, holomorphic functions, and the

polynomials respectively, and these are preserved under the local identifications.

Sheaves are introduced to make these ideas precise. To explain this better,

let us consider the primary example of a sheaf: The sheaf of continuous maps on

15

Page 16: Geir Ellingsrud and John Christian Ottem October 15, 2018€¦ · between algebra and geometry: a ne varieties correspond to nitely generated k-algebras without zero-divisors, and

1.1. Sheaves and presheaves

a topological space X. By definition, X comes with a collection of ‘open sets’,

and these encode what it means for a map f : X → Y to another topological

space Y to be continuous: For every open U ⊆ Y , the set f−1(U) should be

open in X. For two topological spaces X and Y , we can define for each open

U ⊆ X, a set of continuous maps

C(U, Y ) = f : U → Y | f is continuous.

Note that if V ⊆ U is open, then the restriction f |V of a continuous function f

to V is again continuous, so we obtain a map

ρUV : C(U, Y )→ C(V, Y )

f 7→ f |V .

Moreover, note that if W ⊆ V ⊆ U , we can restrict to W by first restricting to

V , and so ρUW = ρVW ρUV . The collection of the sets C(U, Y ) together with

their restriction maps ρUV constitutes the sheaf of continuous maps from X to

Y .

An essential feature of continuity is that it is a local property; f is continuous

if and only it is continuous in a neighbourhood of every point, and of course

two continuous maps that are equal in a neighbourhood of every point, have to

be equal everywhere. Moreover, continuous functions can be glued: Given an

open covering Uii∈I of an open set U , and continuous functions fi ∈ C(Ui, Y ),

so that for each x ∈ X the value fi(x) does not depend on i, we can patch the

maps fi together to form a continuous map f : U → Y , so that f |Ui = fi for

each i: Just define f(x) = fi(x) for any i such that x ∈ Ui.Essentially, a sheaf on a topological space X is a structure that encodes these

properties. On a topological space, a sheaf can be thought of as a distinguished

set of functions, but they can also be more general objects that behave as sets

of functions. The main aspect is that we want the distinguished properties to

be preserved under restrictions to open sets, and that they are determined from

their local properties.

1.1.1 Presheaves

We begin with the definition of a presheaf.

Definition 1.1. Let X be a topological space. A presheaf of abelian groups Fon X consists of the following two sets of data:

(i) for each open U ⊆ X, an abelian group F(U);

16

Page 17: Geir Ellingsrud and John Christian Ottem October 15, 2018€¦ · between algebra and geometry: a ne varieties correspond to nitely generated k-algebras without zero-divisors, and

Chapter 1. Sheaves

(ii) for each pair of nested opens V ⊆ U a restriction map

ρUV : F(U)→ F(V ).

The restriction maps must furthermore satisfy the following two conditions

(a) for any open U ⊆ X, one has ρUU = idF(U);

(b) for any nested opens W ⊆ V ⊆ U , one has ρUW = ρVW ρUV .

We will usually write s|V for ρUV (s) when s ∈ F(U). The elements of F(U) are

usually called ‘sections’ (or ‘sections over U ’). You may perhaps think of them

as ‘abelian group-valued functions on U ’.

We will often write Γ(U,F) for the group F(U); here Γ is the ‘global sections’-

functor (it is functorial in both U and F).

Of course the notion of a presheaf is not confined to presheaves of abelian

groups. One may speak about presheaves of sets, rings, vector spaces or whatever

you want: Indeed, for any category C one may define presheaves with values

in C . The definition goes just like for abelian groups, the only difference being

that one requires the ‘spaces’ of sections F(U) over open sets U to be objects

in the category C and no longer abelian groups, and of course the restriction

maps are all required to be morphisms in C . One may phrase this definition

purely in categorical terms by introducing the small category openX of open sets

in X whose objects are the open sets, and the morphisms are the inclusion maps

between open sets. With that definition up our sleeve, a presheaf with values in

the category C is just a contravariant functor

F : openX → C .

We are certainly going to meet sheaves with a lot more structure than the mere

structure of abelian groups e.g., like sheaves of rings, but they will all have an

underlying structure of abelian group, so we start with those. That being said,

sheaves of sets play a great role in mathematics, and in algebraic geometry, so

we should not completely wipe them under the rug. Most results we establish

for sheaves of abelian groups can be proved mutatis mutandis for sheaves of sets

as well, as long as they can be formulated in terms of sets.

1.1.2 Sheaves

We are now ready to give the main definition of this chapter:

Definition 1.2. A presheaf F is a sheaf if it satisfies the two conditions:

17

Page 18: Geir Ellingsrud and John Christian Ottem October 15, 2018€¦ · between algebra and geometry: a ne varieties correspond to nitely generated k-algebras without zero-divisors, and

1.1. Sheaves and presheaves

(i) (Locality axiom) Given an open subset U ⊆ X with an open covering

U = Uii∈I , and a section s ∈ F(U), then if s|Ui = 0 for all i, then

s = 0 ∈ F(U).

(ii) (Gluing axiom) If U and U are as in (i), and if si ∈ F(Ui) is a collection of

matching on the overlaps, i.e.,

si|Ui∩Uj = sj|Ui∩Uj ∀i, j ∈ I,

then there exists a section s ∈ F(U) so that s|Ui = si for all i.

Note that the condition (i) says that sections are uniquely determined from

their restrictions to smaller open sets. The condition (ii) says that you can patch

together local sections to a global one, provided they agree on overlaps, just like

in the example of continuous functions.

Remark 1.3. There is a nice concise way of formulating the two sheaf axioms

at once. For each open over U = Ui of an open set U ⊆ X there is a sequence

0→ F(U)α−→∏i

F(Ui)β−→∏i,j

F(Ui ∩ Uj)

where the maps α, β are defined by α(s) = (s|Ui)i, and β(si) = (si|Ui∩Uj −sj|Ui∩Uj)i,j, Then F is a sheaf if and only if these sequences are exact. Indeed,

you should check that the Locality axiom is equivalent to α being injective, and

the Gluing axiom to ker β = imα. This reformulation is sometimes handy when

proving that a given presheaf is a sheaf.

1.1.3 Subsheaves and saturation of subpresheaves

If F is a presheaf on X, a subpresheaf G is a presheaf such that G(U) ⊆ F(U)

for every open U , and such that the restriction maps of G are induced by those

of F . If F and G are sheaves, of course G is called a subsheaf.

Let F be a sheaf on X and G ⊆ F a subpresheaf. We say that a section

s ∈ F(U) locally lies in G if for some open covering Ui i∈I one has s|Ui ∈ G(Ui)

for each i.

Definition 1.4. We define the sheaf saturation G of G in F by letting the

sections of G over U be the sections of F over U that locally lie in G.

The sheaf saturation G is again a subpresheaf of F (with restriction maps

being the ones induced from F). In fact, G is, almost by definition, a sheaf. The

locality axiom is satisfied for G, since it holds for F . Furthermore, given a set

of patching data for G; that is, an open covering Uii∈I of an open set U and

18

Page 19: Geir Ellingsrud and John Christian Ottem October 15, 2018€¦ · between algebra and geometry: a ne varieties correspond to nitely generated k-algebras without zero-divisors, and

Chapter 1. Sheaves

sections si over Ui matching on intersections, the si’s can be glued together in

F since F is a sheaf, and since they are born to locally lie in G, the patch lies

locally in G as well and is a section of G over U . So also the Gluing axiom holds

for G.

So G is basically the smallest subsheaf of F which contains G. If G already is

a sheaf, we don’t get anything new, so that G = G.

1.2 A bunch of examples

Example 1.5. Take X = Rn and let C(X,R) be the sheaf whose sections over

an open set U is the ring of continuous real valued functions on U , and the

restriction maps ρUV are just the good old restriction of functions. Then C(X,R)

is a sheaf of rings (functions can be added and multiplied), and both the sheaf

axioms are satisfied. You should convince yourself that this is true.

Example 1.6. For a second familiar example let X ⊆ C be any open set. On

X one has the sheaf OX of holomorphic functions. That is, for any open U ⊆ X

the sections OX(U) is the ring of holomorphic (i.e., complex analytic) functions

on U . One can relax the condition of holomorphy to get the larger sheaf KX

of meromorphic functions on X. This sheaf contains OX , and the sections over

an open U are the meromorphic functions on U . In a similar way, one can

get smaller sheaves contained in OX by imposing vanishing conditions on the

functions. For example if p ∈ X is any point, one has the sheaf denoted mp of

holomorphic functions vanishing at p. As the name indicates the sections of mp

over U are holomorphic functions in U , and if p ∈ U , one requires additionally

that they should vanish at p. Convince yourself that this indeed is a subsheaf of

OX .

Exercise 1. Let X ⊆ C be an open set, and assume a1, . . . , an are distinct

points in X and n1, . . . , nr be natural numbers. Define F(U) to be the set of

those meromorphic functions f ∈ Γ(U,KX) holomorphic away from the ai’s and

having a pole order bounded by ni at ai. Show that F is a sheaf of abelian

groups. Is it a sheaf of rings?

Example 1.7 (A presheaf that is not a sheaf). The sheaf C(X,R) from Example

1.5 has a subpresheaf F which is not a sheaf: For each open U let F(U) =

Cb(U,R), the group of bounded continuous functions. F is not a sheaf, because

you can’t glue; e.g., the function f(x) = x is bounded on any open ball Br(0) =

x ∈ Rn|‖x‖ < r, but the restrictions f |Bn do not give you a bounded continuous

function on all of⋃n≥0Bn = Rn.

Example 1.8. (Constant presheaf) For any space X and any abelian group A

one has the constant presheaf whose group of sections over any nonempty open

19

Page 20: Geir Ellingsrud and John Christian Ottem October 15, 2018€¦ · between algebra and geometry: a ne varieties correspond to nitely generated k-algebras without zero-divisors, and

1.2. A bunch of examples

set U equals A and equals 0 if U = ∅. This is not a sheaf, since if U ∪ U ′ is a

disjoint union, any choice of elements a, a′ ∈ A will give sections over U and

U ′ respectively, and they match on the empty intersection! But if a 6= a′, they

cannot be glued. In fact, this is a sheaf if and only if any two non-empty open

subsets of X have non-empty intersection. Varieties with Zariski topology are

examples of such spaces.

Example 1.9 (Skyscraper sheaf). Let A be a group. For x ∈ X define a presheaf

Ax by Ax(U) = A if x ∈ U and Ax(U) = 0 otherwise. This is a sheaf, usually

called a ‘skyscraper’ sheaf.

The two next examples suggest another way of thinking about sheaves: They

are ‘sections’ of maps Y → X into X:

Example 1.10 (Tautological sheaf on projective space). Let P (V ) ' Pn be the

projective space of lines in an (n+ 1)-dimensional vector space V . Let

L ⊆ P (V )× V

be the set of pairs ([l], v) where [l] denotes the class of a line l ⊆ V and v ∈ l.Let π be the projection to the first factor. Note that the fiber over any point

p ∈ P (V ) is a 1-dimensional vector space (π−1(p) is the line in V corresponding

to p). This is a basic example of a non-trivial vector bundle. We will discuss

these later in the course.

For now, we can use it to define the so-called tautological sheaf on P (V ): For

any U ⊆ P (V ), let

F(U) = σ : U → L|π σ = idU

This is indeed a sheaf, usually denoted by O(−1).

Example 1.11. Here is a related example from topology. Consider the Mobius

strip M with its projection π : M → S1 to S1. This is an example of a fiber

bundle; any fiber of π is homeomorphic to the unit interval [0, 1]. The sheaf of

sections of π is given over an open set U ⊆ S1 by

F(U) = s : U →M |π s = idU

Again, this is a sheaf.

There is also the trivial fiber bundle π : S1×[0, 1]→ S1 given by the projection

to the first factor. This is not homeomorphic to the bundle above. [Brain teaser:

Prove this.]

Our main interest in this course will still be the following:

20

Page 21: Geir Ellingsrud and John Christian Ottem October 15, 2018€¦ · between algebra and geometry: a ne varieties correspond to nitely generated k-algebras without zero-divisors, and

Chapter 1. Sheaves

Example 1.12. Let V be an algebraic variety (quasi-projective, as in Hartshorne

Ch. I) with the Zariski topology. For each open U ⊆ X, define OV (U) to be

the ring of regular functions U → k. This is certainly a presheaf, and in fact, a

sheaf.

1.3 Stalks

Given a presheaf F of abelian groups on X. With every point x ∈ X there is an

associated abelian group Fx called the stalk 1 of F at x. The elements of Fx are

called germs of sections2 near x.

The definition goes as follows: We begin with the disjoint union∐

x∈U F(U)

whose elements we think of as pairs (s, U) where U is any open neighbourhood of

x and s is a section of F over U . We want to identify sections that coincide near x;

that is, we declare (s, U) and (s′, U ′) to be equivalent, and write (s, U) ∼ (s′, U ′),

if there is an open V ⊆ U ∩ U ′ with x ∈ V such that s and s′ coincide on V ;

that is one has

s|V = s′|V .

This is clearly a reflexive and symmetric relation. And it is transitive as well.

Indeed, if (s, U) ∼ (s′, U ′) and (s′, U ′) ∼ (s′′, U ′′), one may find open neighbour-

hoods V ⊆ U ∩ U ′ and V ′ ⊆ U ′ ∩ U ′′ of x over which s and s′, respectively s′

and s′′, coincide. Clearly s and s′′ then coincide over the intersection V ′ ∩ V ,

and the relation ∼ is an equivalence relation.

Definition 1.13. The stalk Fx at x ∈ X is by definition the set of equivalence

classes

Fx =∐x∈U

F(U)/ ∼ .

In case F is a sheaf of abelian groups, the stalks Fx are all abelian groups.

This is not a priori obvious, since sections over different open sets can not be

added. However if (s, U) and (s′, U ′) are given, the restrictions s|V and s′|V to

1Norsk: ‘stilk’2Norsk: ‘seksjonskimer’ eller bare ‘kimer’

21

Page 22: Geir Ellingsrud and John Christian Ottem October 15, 2018€¦ · between algebra and geometry: a ne varieties correspond to nitely generated k-algebras without zero-divisors, and

1.3. Stalks

any open V ⊆ U ∩ U ′ can be added, and this suffices to define an abelian group

structure on the stalks.

1.3.1 The germ of a section

For any neighbourhood U of x ∈ X, there is a map F(U) → Fx sending a

section s to the equivalence class where the pair (s, U) belongs. This class is

called the germ of s at x, and a common notation for it is sx. The map is

a homomorphism of abelian groups (rings, modules,..) as one easily verifies.

Clearly one has sx = (s|V )x for any other open neighbourhood V of x contained

in U—or expressed in the lingo of diagrams—the following diagram commutes:

F(U) //

ρUV

Fx

F(V ).

<<

When working with sheaves and stalks, it is important to remember the three

following properties; the third one is easily deduced from the two first.

• The germ sx of s vanishes if and only s vanishes on a neighbourhood of x,

i.e., there is an open neighbourhood U of x with s|U = 0.

• All elements of the stalk Fx are germs, i.e., of the shape sx for some section

s over an open neighbourhood of x.

• The abelian sheaf F is the zero sheaf if and only if all stalks are zero, i.e.,

Fx = 0 for all x ∈ X.

Example 1.14. Let X = C, and let OX(U) be the ring of holomorphic functions

on U . If f and g are two sections of OX over a neigbourhood of p having the

same germ at p, i.e., f = g ∈ OX,p, then f = g where they are both defined.

If you replace ‘holomorphic’ by C∞, then f and g having the same germ at p

only implies that the derivatives of f and g at p of all orders of coincide, i.e.,

f (n)(p) = g(n)(p) for all n ≥ 0, but not much else.

1.3.2 A primer on limits

Another notation for the stalk of F at x is

Fx = lim−→U3xF(U).

22

Page 23: Geir Ellingsrud and John Christian Ottem October 15, 2018€¦ · between algebra and geometry: a ne varieties correspond to nitely generated k-algebras without zero-divisors, and

Chapter 1. Sheaves

This is the direct limit (or ‘colimit’) of all F(U) when U runs over the partially

ordered set of open sets containing x. Taking the direct limit is a general

construction in algebra, so let us give a few more details here for future reference.

A directed set I is a partially ordered set such that for each pair of elements

i, j ∈ I there is a third element k such that i < k and j < k. If I is a directed

set and C is a category, a directed system of objects in C is a collection Gii∈Iof objects in C, such that for all i < j there is a morphism fij : Gi → Gj, with

fii = id and fjk fij = fik.

Direct limits

The direct limit of Gi, denoted by G = lim−→i∈I Gi is an object in C, equipped

with morphisms gi : Gi → G which satisfy the following universal property: For

any object H ∈ C with maps hi : Gi → H such that hi = hj fij for each i ≤ j,

there is a unique map h : G→ H making the following diagram commute:

Gihi //

fi

H

Gh

>>

Heuristically, two elements in the direct limit represent the same element in the

direct limit if they are ‘eventually equal.’

If the Gi are sets (or groups, rings, ..), an explicit construction for this is

the quotient∐

i∈I Gi/ ∼, where and g ∼ h, with g ∈ Gi and h ∈ Gj, if there

exists a k ∈ I such that fik(g) = fjk(h). In the case I is the set of opens, and

GU = F(U), we recover the previous definition of the stalk Fx.

Inverse limits

By reversing all the arrows, we can define the ‘inverse limit’ or ‘projective limit’

of a directed system Gi is defined as above. That is, the maps Gi → Gj are

defined for j < i, and lim←−i∈I Gi is an element of C equipped with universal maps

to each of the Gi.

1.4 Morphisms between (pre)sheaves

A morphism φ : F → G of (pre)sheaves on a space X is collection of maps

φU : F(U)→ G(U) so that the following diagram commutes for each inclusion

23

Page 24: Geir Ellingsrud and John Christian Ottem October 15, 2018€¦ · between algebra and geometry: a ne varieties correspond to nitely generated k-algebras without zero-divisors, and

1.4. Morphisms between (pre)sheaves

V ⊆ U :

F(U)φU //

ρUV

G(U)

ρUV

F(V )φV // G(V )

In this way the abelian sheaves on X form a category AbShX whose objects are

the abelian sheaves on X and the morphisms being the maps between them.

The composition of two maps of sheaves is defined in the obvious way as the

composition of the maps on sections.

Remark 1.15. If one considers the sheaves F and G as contravariant func-

tors on the category openX , a map between them is what is called a natural

transformation (naturlig transformasjon) between the two functors.

As usual, a map φ between two (pre)sheaves F and G is an isomorphism

if it has a two-sided inverse, i.e., a map ψ : G → F such that φ ψ = idG and

ψ φ = idF .

A map φ : F → G between two presheaves F and G induces for every point

x ∈ X a map φx : Fx → Gx between the stalks. Indeed, one may send a pair

(s, U) to the pair (φU (s), U), and since φ behaves well with respect to restrictions,

this assignment is compatible with the equivalence relations; that is, if (s, U)

and (s′, U ′) are equivalent and s and s′ coincide on an open set V ⊆ U ∩U ′, one

has

φU(s)|V = φV (s|V ) = φV (s′|V ) = φU ′(s′)|V .

Obviously (φ ψ)x = φx ψx and (idF)x = idFx , so the assignment φ→ φx is a

functor from the category of abelian sheaves to the category of abelian groups.

Example 1.16. In the case X = R let Cr(X) be the subsheaf of C(X,R)

consisting of functions being r times continuously differentiable (check that

this is indeed a subsheaf !). The differential operator D = d/dx defines a map

D : Cr → Cr−1.

Exercise 2. In the same vein, the differential operator gives a mapD : OX → OX ,

where as previously X ⊆ C is an open set. Show that the assignment

A (U) = f ∈ OX(U) | Df = 0

defines a subsheaf A of OX . Show that if U is a connected open subset of X,

ones has A (U) = C. In general for a not necessarily connected set U , show

that A (U) =∏

π0UC where the product is taken over the set π0U of connected

components of U .

24

Page 25: Geir Ellingsrud and John Christian Ottem October 15, 2018€¦ · between algebra and geometry: a ne varieties correspond to nitely generated k-algebras without zero-divisors, and

Chapter 1. Sheaves

1.5 Kernels and images

Let φ : F → G be a map between two abelian sheaves on X.

Definition 1.17. The kernel kerφ of φ is a subsheaf of F whose sections over

U are just kerφU , i.e., the sections in F(U) mapping to zero under φU : F(U)→G(U).

The requirement in the definition is compatible with the restriction maps

since φV (s|V ) = φU(s)|V , for any section s over the open set U and any open

V ⊆ U . Thus we have defined a subpresheaf of F which, indeed, is a subsheaf:

If si i are gluing data for the kernel, one may glue the si’s to a section s of

F over U . One has φ(s)|Ui = φ(s|Ui) = φ(si) = 0, and from the locality axiom

for G it follows that φ(s) = 0. We leave it to the reader to verify that the stalk

(kerφ)x of kerφ at x equals kerφx. We have proven

Lemma 1.18. Let φ : F → G be a map of abelian sheaves. The kernel kerφ is

a subsheaf of F having the two properties

• Taking the kernel commutes with taking sections: Γ(U, kerφ) = kerφU ,

• Forming the kernel commutes with forming stalks: (kerφ)x = kerφx.

One says that the map φ is injective if kerφ = 0. This is, in view of the

previous lemma, equivalent to the condition kerφx = 0 for all x, i.e., that all φxare injective. One often expresses this in a slightly imprecise manner by saying

that φ is injective on all stalks.

When it comes to images the situation is not as nice as for kernels. One

defines the image presheaf contained in G by letting the sections over U be equal

to imφU . However this is not necessarily a sheaf. If si = φUi(ti) are gluing data

for the image presheaf there is no reason for the ti’s to match on the intersections

Uij = Uj ∩ Uj even if the si’s do; the differences ti|Uij − tj|Uij may very well

be non-zero sections of the kernel of φUij ! In fact, we will see several explicit

examples of this later.

To remedy this situation, we simply make the following definition:

Definition 1.19. For a morphism φ : F → G we define the sheaf imφ to be

the saturation of the image presheaf U 7→ imφU , i.e., the smallest subsheaf

containing the images.

Forming the image of a map of sheaves does not always commute with taking

sections, but as we shall verify in the upcoming lemma, forming images commutes

with forming stalks. One has

25

Page 26: Geir Ellingsrud and John Christian Ottem October 15, 2018€¦ · between algebra and geometry: a ne varieties correspond to nitely generated k-algebras without zero-divisors, and

1.5. Kernels and images

Lemma 1.20. Let φ : F → G be a map of abelian sheaves. The image imφ is a

subsheaf of G.

• For all open subsets U of X one has imφU ⊆ Γ(U, imφ).

• For all x ∈ X one has (imφ)x = imφx.

Proof. We only have to verify the last statement, so let tx ∈ imφx and pick

an sx ∈ Fx with φx(sx) = tx. This means that over some open neighbourhood

V , one has φV (s) = t, and t is a section of imφ over V . This shows that

imφx ⊆ (imφ)x. The other way around, if t is a section of G over an open U

containing x locally lying in image presheaf, the restriction t ⊆ V lies in imφVfor some smaller neighbourhood V of x, hence the germ tx lies in imφx.

The map φ : F → G is said to be surjective if the image sheaf imφ = G. This

is equivalent to all the stalk-maps φx being surjective (one says φ is surjective on

stalks), but it does not imply that all maps φU on sections are surjective. Here

is a counterexample:

Example 1.21. Let OX be the sheaf of holomorphic functions on X = CP1

and let p, q ∈ X be two distinct points. Let OX(−p) and OX(−q) be the two

subsheaves of OX of holomorphic functions vanishing at p and q respectively and

consider the map

φ : OX(−p)⊕ OX(−q)→ OX

defined by (f, g) 7→ f + g. Then φx is surjective for every x ∈ X [Check this!],

but φX cannot be surjective, since by Liouville’s theorem, the only globally

homolophic functions are the constants, and so OX(−p)(X) = OX(−q)(X) = 0

and OX(X) = C.

In particular, since imφ(X) = 0, this shows that the naive image presheaf

F(U) = imφU is not a sheaf.

However, for the map φ to be an isomorphism, one has the following

Proposition 1.22. Let φ : F → G be a map of abelian sheaves. Then the

following four conditions are equivalent

• The map φ is an isomorphism.

• For every x ∈ X the map on stalks φx : Fx → Gx is an isomorphism,

• One has kerφ = 0 and imφ = G,

• For all open subsets U ⊆ X the map on sections φU : F(U)→ G(U) is an

isomorphism.

26

Page 27: Geir Ellingsrud and John Christian Ottem October 15, 2018€¦ · between algebra and geometry: a ne varieties correspond to nitely generated k-algebras without zero-divisors, and

Chapter 1. Sheaves

Proof. Most of the implications are straightforward from what we have done so

far and are left to student. We comment just on the two most the salient points.

Firstly, assume that all the stalk maps φx : Fx → Gx are isomorphism, and

let us deduce from this that all the maps φU : F(U) → G(U) on sections are

isomorphisms. It is clear that φU is injective since forming kernels commute with

taking sections as in lemma 1.18 on page 25. So assume that s ∈ G(U). For

each x ∈ U there is a germ t(x) induced by a section tx of F over some open

neighbourhood Ux of x satisfying φx(t(x)) = s(x). Let sx = s|Ux .After possibly having shrunken the neighbourhood Ux, one has φUx(tx) = sx.

The sx’s match on the intersections Ux ∩ Ux′—they are all restrictions of the

section s—and therefore the tx’s match as well because φUx is injective as we

just observed. Hence, the tx’s patch together to a section t of F that must map

to s since it does so locally.

Secondly, if all the φU ’s are isomorphisms, we have all the inverse maps φ−1U

at our disposal. They commute with restrictions since the maps φU do. Indeed,

from φV ρUV = ρUV φU one obtains ρUV φ−1U = φ−1

V ρUV and the φ−1U ’s

thus define a map φ−1 : G → F of sheaves, which of course, is inverse to φ. We

conclude that φ is an isomorphism.

1.6 Exact sequences of sheaves.

Given a sequence

. . .φi−2 // Fi−1

φi−1 // Fiφi // Fi+1

φi+1 // . . .

of maps of abelian sheaves, we say that the sequence is exact at stage i if

kerφi = imφi−1. The short exact sequences are the ones one most frequently

encounters. They are sequences of the form

0 // F1φ1 // F2

φ2 // F3// 0 (1.6.1)

that are exact at each stage. This is just another and very convenient way

of simultaneously saying that φ1 is injective, that φ2 is surjective and that

imφ1 = kerφ2. Forming the sections of the short exact sequence (1.6.1) over

some open set U one finds the sequence

0 // Γ(U,F1) // Γ(U,F2) // Γ(U,F3) (1.6.2)

which is exact at the three leftmost stages, but the map Γ(U,F2) // Γ(U,F3)

to the right is not necessarily surjective (although it of course can be). One

way of phrasing this is to say that taking sections over an open set U—that is

27

Page 28: Geir Ellingsrud and John Christian Ottem October 15, 2018€¦ · between algebra and geometry: a ne varieties correspond to nitely generated k-algebras without zero-divisors, and

1.6. Exact sequences of sheaves.

Γ(U, ∗)—is a left exact functor (venstreeksakt funktor). However this is not right

exact in general. The defect of this lacking surjectivity is a fundamental problem

in every part of mathematics where sheaf theory is used, and to cope with it one

has cohomology!

Proposition 1.23. For a short exact sequence 0 → F ′ φ−→ F ψ−→ F ′′ → 0, we

have the following induced exact sequence

0 −→ Γ(U,F ′)φU−→ Γ(U,F )

ψU−→ Γ(U,F ′′),

Proof. The map φ is injective as a map of sheaves, hence injective on all open

sets U , so the sequence above is exact at Γ(U,F1). To see that it is also exact in

the middle, we show thatker(ψU) = im(φU). (The image presheaf is then given

by the kernel of a morphism of sheaves, which is indeed a sheaf.)

Consider for each x ∈ U the the diagram

0 F ′(U) F(U) F ′′(U) 0

0 F ′x Fx F ′′x 0

φU ψU

φx ψx

the bottom row is exact, since the above sequence is exact.

ker(ψU) ⊇ im(φU): Let s ∈ Γ(U,F1) and consider for each x ∈ U the germ

of ψU(φU(s)) in the stalk F ′′x :

(ψU(φU(s)))x = ψx(φx(sx)).

But by exactness, ψx(φx(sx)) = 0 for all x ∈ U . Hence ψU(φU(s)) = 0, so

im(φU) ⊆ ker(ψU).

ker(ψU) ⊆ im(φU) : Let t ∈ ker(ψU), so ψU(t) = 0. Then for all x ∈ U

we have that ψx(tx) = (ψU(t))x = 0, so the germ of t at x is an element in

ker(ψx) = im(φx) by exactness again. Hence for every x ∈ U there is a s′x ∈ F ′x,

say of the form [(s′(x), V(x))] for some open neighborhood V(x) ⊆ U of x and

s′(x) ∈ Γ(V(x),F′), such that φx(s

′x) = tx. Then we have that for x, y ∈ U

φV(x)∩V(y)(s′(x)|V(x)∩V(y)) = t|V(x)∩V(y) = φV(x)∩V(y)(s

′(y)|V(x)∩V(y)),

so that by the injectivity of φV(x)∩V(y) (already proved), we get the required

condition

s′(x)|V(x)∩V(y) = s′(y)|V(x)∩V(y)

for the gluing of the s′(x) for x ∈ U . Therefore we have a section s ∈ Γ(U,F )

28

Page 29: Geir Ellingsrud and John Christian Ottem October 15, 2018€¦ · between algebra and geometry: a ne varieties correspond to nitely generated k-algebras without zero-divisors, and

Chapter 1. Sheaves

with the property that for all x ∈ U

s|V(x) = s′(x).

Now we can conclude that for every x ∈ U

(φU(s))x = φx(sx) = φx(s′x) = tx,

since sx = s′x, which gives φU(s) = t as desired.

Let us give a few examples where the surjectivity on the right fails:

Example 1.24 (The exponential sequence). Let X = C − 0. The non-

vanishing holomorphic functions in an open set U ⊆ X form a multiplicative

group, and there is a sheaf O∗X with these groups as sections. For any f

holomorphic in U the exponential exp f(z) is a section of O∗X . Hence there is an

exact sequence

0 // Z // OXexp // O∗X // 0,

since non-vanishing functions locally have logarithms. However unless U is

simply connected, exp(U) is not surjective. In that case there will always be

some non-vanishing function in U without a logarithm defined throughout U .

Example 1.25 (Differential operators). Let X = C and recall the sheaf OX of

holomorphic functions and the map D : OX → OX sending f(z) to the derivative

f ′(z). There is an exact sequence

0 // C // OXD // OX

// 0.

This hinges on the two following facts. Firstly, a function whose derivative

vanishes identically is locally constant; hence the kernel kerD equals the constant

sheaf CX . Secondly, in small open disks any holomorphic function has a primitive

function, i.e., is a derivative, e.g., if f(z) =∑

n≥0 an(z − a)n in a small disk

around a, the function g(z) =∑

n≥0 an(n+ 1)−1(z− a)n+1 has f(z) as derivative.

However, taking sections over open sets U we merely obtain the sequence

0 // Γ(U,C) // Γ(U,OX)DU // Γ(U,OX).

Whether DU is surjective or not, depends on the topology of U . If U is simply

connected, one deduces from Cauchy’s integral theorem that every holomorphic

function in U is a derivative, so in that case DU is surjective. On the other

hand, if U not simply connected, DU is not surjective; e.g., if U = C− 0 , the

function z−1 is not a derivative in U .

29

Page 30: Geir Ellingsrud and John Christian Ottem October 15, 2018€¦ · between algebra and geometry: a ne varieties correspond to nitely generated k-algebras without zero-divisors, and

1.7. B-sheaves

Example 1.26. Consider an algebraic variety X with the sheaf OX of regular

functions on X. For any point x ∈ X let k(x) denote the skyscraper sheaf (see

Example 1.9) whose only non-zero stalk is the field k located at x. There is a map

of sheaves evx : OX → k(x) sending a function that is regular in a neighbourhood

of x to the value it takes at x. This maps sits in the exact sequence of sheaves

0 // mx// OX

evx // k(x) // 0,

where mx by definition is the kernel of evx (the sections of mx are the functions

vanishing at x). Taking two distinct points x and y in X we find a similar exact

sequence

0 // Ix,y// OX

evx,y // k(x)⊕ k(y) // 0,

where Ix,y is the sheaf of functions vanishing on x and y, and

evx,y(f) =

(0, 0) if x, y /∈ U,(f(x), 0) if x ∈ U but y /∈ U,(0, f(y)) if y ∈ U but x /∈ U,(f(x), f(y)) if x, y ∈ U.

If for example X = P1 (or any other complete variety), there are no global

regular functions on X other than the constants, and hence Γ(X,OX) = k. But

of course, Γ(P1, k(x)⊕ k(y)) = k ⊕ k, so the map evx,y can not be surjective on

global sections.

1.7 B-sheaves

Recall that a basis for a topology on X is a collection of open subsets B such

that any open set of X can be written as a union of elements of B. In many

situations it turns out to be convenient to define a sheaf by saying what it should

be on a specific basis for the topology on X. In this section we will see that

there exists a unique way to produce a sheaf, given sections that glue over open

subsets in B.

Let us first make the following definition:

Definition 1.27. A B-presheaf F consists of the following data: (i) For each

U ∈ B, an abelian group F(U); and (ii) for all U ⊆ V , with U, V ∈ B, a

restriction map ρUV : F(U) → F(V ). As usual, these are required to satisfy

the relations ρUU = idF(U) and ρWU = ρV U ρWV . A B-sheaf is a B-presheaf

satisfying the Locality and Gluing axioms for open sets in B.

30

Page 31: Geir Ellingsrud and John Christian Ottem October 15, 2018€¦ · between algebra and geometry: a ne varieties correspond to nitely generated k-algebras without zero-divisors, and

Chapter 1. Sheaves

Proposition 1.28. Let X be a topological space and let B be a basis for the

topology on X. Then

(i) Every B-sheaf F extends uniquely to a sheaf on X.

(ii) If φ : F → G is a morphism of B-sheaves, then φ extends uniquely to a

morphism between the corresponding sheaves.

Indeed, for any open set U ⊆ X, we can write U as a union of open sets

Ui ∈ B, and then we can define F(U) to be the set of elements si ∈∏

iF(Ui)

such that si|Ui∩Uj = sj|Ui∩Uj . This is indeed gives a well-defined sheaf, and it

must be unique, since we can see stalks using open sets in B.

1.8 A family of examples – Godement sheaves

We will consider a class of rather peculiar sheaves, called Godement sheaves, to

demonstrate the versatility and the generality of the notion of sheaves.

Let X any topological space. Assume that we for each point x ∈ X are given

an abelian group Ax. The groups Ax can be chosen in a completely arbitrary

way, at random if you will. The choice of these groups gives rise to a sheaf Aon X whose sections over an open set U ⊆ X are given as

Γ(U,A ) =∏x∈U

Ax,

and whose restriction maps are defined as the natural projections

ρUV :∏x∈U

Ax →∏x∈V

Ax,

where V ⊆ U is any pair of open subsets of X. The restriction map just “throws

away” the components at points in U not lying in V .

Proposition 1.29. A is a sheaf.

Proof. The Locality condition holds since if the family Ui i of open sets covers

U , any point x0 ∈ U lies in some Ui0 , so if s = (ax)x∈U ∈ Γ(U,A ) is a section,

the component ax0 survives in the projection onto Γ(Ui,A ) =∏

x∈Ui Ax. Hence

if s|Ui = 0 for all i, it follows that s = 0.

The Gluing condition holds: Assume given an open cover Ui i of U and

sections si ∈ Ui matching on the intersections Ui ∩ Uj. The matching conditions

mean that the component of si at a point x is the same whatever i is as long as

x ∈ Ui. Hence we get a section s of A over U by using this common component

as the component of s at x. It is clear that s|Ui = si.

31

Page 32: Geir Ellingsrud and John Christian Ottem October 15, 2018€¦ · between algebra and geometry: a ne varieties correspond to nitely generated k-algebras without zero-divisors, and

1.8. A family of examples – Godement sheaves

Definition 1.30. The sheaf A is sometimes called the Godement sheaf of the

collection Ax .

The construction is not confined to abelian groups, but works for any category

where general products exist; like sets, rings, etc.

So what is the stalk of A at a point x? Don’t ask that question! The group

can be fairly complicated. Of course there is a map Ax → Ax, but that is in

general the best you can say. For example, suppose Ay = Z/2 for all y ∈ X, and

that x = limxn with xn 6= xm for all n 6= m. Then Ax maps onto the (infinite)

set of tails of sequences of 0s and 1s. Namely, every open neighbourhood of x

contains almost all of the xn. On the other hand, if every neighbourhood of x

contains a point y such that Ay = 0, then Ax = 0.

1.8.1 Skyscraper sheaves.

This is a very special instance of a Godement sheaf where all the abelian groups

Ax are zero except for one. So we fix just one abelian group A, and we choose a

point x ∈ X. The corresponding Godement sheaf is denoted A(x) and is called

a skyscraper sheaf (skyskraperknippe). The sections are described by

Γ(U,A(x)) =

A if x ∈ U,0 otherwise .

Contrary to the general case, in this case the stalks are easy to describe. One

verifies easily that they are zero everywhere except at x, where the stalk equals

A. Indeed, if y 6= x, then y lies in the open set X − x over which all sections

of x∗A vanish.

In most spaces concerning us there are plenty of non-closed points. For such

points one still has the Godement sheaf A(x), but the argument above does not

work, and the description of the stalks are somehow more complicated. One

would hardly call A(x) a skyscraper sheaf.

Exercise 3. Assume that x is not closed and let Z = x be the closure of

the singleton x . Show that the stalks of A(x) are A(x)y = 0 if y 6∈ Z and

A(x)y = A for points y belonging to Z

Slightly generalising the construction of a skyscraper sheaf, one may form

the Godement sheaf A defined by a finite set of distinct closed points x1, . . . , xrand corresponding abelian groups A1, . . . , Ar. Then one sees, having in mind

that an empty direct sum is zero, that the sections of A over an open set U is

32

Page 33: Geir Ellingsrud and John Christian Ottem October 15, 2018€¦ · between algebra and geometry: a ne varieties correspond to nitely generated k-algebras without zero-divisors, and

Chapter 1. Sheaves

given as Γ(U,A ) =⊕

xi∈U Ai. The stalks of A are

Ax =

0 when x 6= xi for all i

Ai when x = xi.

One is tempted to call such a sheaf a barcode sheaf (barcodeknippe)!

Example 1.31 (A set of peculiar examples). Of course if Z is a discrete subset

of X, not necessarily finite, one may associate a Godement sheaf A with any

family Az z∈Z of abelian groups indexed by Z (letting Ax = 0 if x /∈ Z) and

the sections over an open set U is by definition of a Godement sheaf given by

Γ(U,A ) =∏

z∈Z∩U

Az

Just as in the finite case the stalks will be

Ax = Ax

However if Z has accumulation points this is no longer true. Assume for simplicity

that zi is a sequence converging to z. Two sequences of elements ai and a′i are

said to define the same tail if there is some N such that ai = a′i for i ≥ N . Then

the stalk at z will be the space of tails from the Ai, that is, Az =∏

iAi/ ∼.

1.8.2 The Godement sheaf associated with a presheaf

Assume F is a given abelian presheaf on X. The stalks Fx of F of course give a

collection of abelian groups indexed by points in X, good as any other, and we

may form the corresponding Godement sheaf which we denote by Π(F). The

sections of Π(F) are given by

Γ(U,Π(F)) =∏x∈U

Fx, (1.8.1)

and the restriction maps are the projections like for any Godement sheaf. Note

that this sheaf can be very big.

There is an obvious and canonical map

κF : F → Π(F)

sending a section s ∈ F(U) to the element (sx)x∈U of the product in (1.8.1). This

map is functorial in F , for if φ : F → G is a map, one has the stalkwise maps

φx : Fx → Gx, and by taking appropriate products of these, we obtain a map

33

Page 34: Geir Ellingsrud and John Christian Ottem October 15, 2018€¦ · between algebra and geometry: a ne varieties correspond to nitely generated k-algebras without zero-divisors, and

1.9. Sheafification

Π(φ) : Π(F)→ Π(G). This sheaf is sometimes called the sheaf of discontinuous

sections of F . On sections over U it has the effect

Π(φ)((sx)x∈U

)=(φx(sx)

)x∈U

and there is a commutative diagram

F κF //

φ

Π(F)

Π(φ)

G κG// Π(G).

(1.8.2)

It is not hard to check that Π(φ ψ) = Π(φ) Πψ and that Π(idF) = idΠ(F).

1.9 Sheafification

Given any presheaf F on X, there is a canonical way of defining a sheaf F+ that

in some sense is the sheaf that best approximates it. What prevents the presheaf

F from being a sheaf is of course the failure of one or both of the sheaves axioms.

To remedy this one must factor out all sections of F whose germs are everywhere

zero, and one has to enrich F by adding enough new sections to be able to glue

from local sections.

A nice and canonical way of doing this is by using the Godement sheaf Π(F)

associated with F . Recall the canonical map κ : F → Π(F) that sends a section s

of F over an open U to the sequence of germs (sx)x∈U ∈∏

x∈U Fx = Γ(U,Π(F)).

This map clearly kills the doomed sections, i.e., those whose germs all vanish.

Now we can get an actual sheaf, by taking the image of κ in Π(F):

Definition 1.32. For a presheaf F on X, we define its sheafification F+ as the

image sheaf imκ in Π(F) (in other words, F+ is the saturation of the subpresheaf

U 7→ imκU in Π(F)).

Abusing notation slightly, we also write κ : F → F+ for the canonical map.

So what are the sections of F+? Over an open set U ⊆ X they are charac-

terized by those locally coming from F . To be completely explicit about it, F+

is the subsheaf of Π(F) given by

F+(U) = (sx) ∈∏x∈U

Fx|(sx) locally lies in F

where, as before, the sentence in the bracket means the following: for each p ∈ U ,

there exists an open neighbourhood V ⊆ U containing p, and a section t ∈ F(V )

such that for all v ∈ V we have sv = tv in Fv.

34

Page 35: Geir Ellingsrud and John Christian Ottem October 15, 2018€¦ · between algebra and geometry: a ne varieties correspond to nitely generated k-algebras without zero-divisors, and

Chapter 1. Sheaves

Lemma 1.33. The sheafification F+ depends functorially on F . Moreover, if

F is a sheaf, κ : F → F+ is an isomorphism, so that F and F+ are canonically

isomorphic.

Proof. Assume that φ : F → G is a map between two presheaves. Let s be

section of Π(F) over some open set U , so that s locally lies in F . In other words

there is a covering Ui of U and sections s′i of F over Ui with s|Ui = κF(s′i).

Hence by (1.8.2) one has

Π(φ)(s|Ui) = Π(φ)(κF(s′i)) = κG(φ(s′i)).

We see that Π(φ)(s) lies locally in G, and Π(φ) takes F+ into G+. There is a

commutative diagram

F //

φ

F+ //

φ+

Π(F)

Π(φ)

G // G+ // Π(G)

In case F is a sheaf, the map κF maps F injectively into Π(F) and F = imκFis its own saturation, hence κ is an isomorphism.

Proposition 1.34. Given a presheaf F on X. Then the sheaf F+ and the

natural map κ : F → F+ enjoys the universal property that any map F → Gwhere G is sheaf, factors through F+ in a unique way. This characterises F+ up

to unique isomorphisms.

Proof. If G in the diagram above is a sheaf, the map G → G+ is an isomorphism

and φ+ takes values in G and provides the wanted factorization. The uniqueness

statement is a general feature of solutions to universal problems.

From what we have proved above, we have that for a presheaf F and a sheaf

GHomPrSh(F , i(G)) = HomSh(F+,G)

where on the right hand side, i(G) denotes G but considered as a presheaf. A

fancy way of restating this is that the sheafification functor F 7→ F+ is an

adjoint to the forgetful functor i : Sh→ PrSh from sheaves to presheaves.

Lemma 1.35. Sheafification preserves stalks: Fx = (F+)x via κx.

Proof. The map κx : Fx → (F+)x is injective, since Fx → (Π(F))x is injective.

To show that it is surjective, suppose that s ∈ (F+)x. We can find an open

neighbourhood U of x such that s is the equivalence class of (s, U) with s ∈F+(U). By definition, this means there exists an open neighbourhood V ⊆ U of

x and a section t ∈ F(V ) such that s|V is the image of t in Π(F)(V ). Clearly

the class of (t, V ) defines an element of Fx mapping to s.

35

Page 36: Geir Ellingsrud and John Christian Ottem October 15, 2018€¦ · between algebra and geometry: a ne varieties correspond to nitely generated k-algebras without zero-divisors, and

1.10. Quotient sheaves and cokernels

Example 1.36 (Constant sheaves). Recall Example 1.8 which showed that the

constant presheaf U 7→ A is usually not a sheaf (A is an abelian group). However,

the sheafification of this constant presheaf is a sheaf, and that is what we call

the constant sheaf with value A and denote by AX .

That being said, the sheaf AX is not quite worthy of the name as is is not

quite constant. One has

Γ(U,AX) =∏π0(U)

A, (1.9.1)

where π0(U) denotes the set of connected components of the open set U .

1.10 Quotient sheaves and cokernels

The main application of the sheafification process we just described is to show

that one may form quotient sheaves. So assume that G ⊆ F are sheaves,

and define a presheaf whose sections over U is the quotient F(U)/G(U). The

restriction maps of F and G respect the inclusions G(U) ⊆ F(U) and hence pass

to the quotients F(U)/G(U). We use these maps as the restriction maps for the

quotient presheaf. The quotient sheaf F/G is by definition the sheafification of

this quotient presheaf. The quotient sits in the exact sequence

0 // G // F // F/G // 0 .

The cokernel cokerφ of the map φ : F → G of abelian sheaves is then just the

quotient sheaf G/ imφ; it fits in the exact sequence

0 // kerφ // F φ // G // cokerφ // 0

Example 1.37. To illustrate why we have to sheafify in these constructions,

consider again the exponential map in Example 1.24. The naive presheaf U 7→coker exp(U) is not a sheaf: the class of the function f(z) = z restricts to 0 in

coker exp on sufficiently small open sets, but it is itself not zero (since otherwise

we would be able to define a global logarithm on C− 0)

1.11 The pushforward of a sheaf

So far we have been interested in various constructions of sheaves on a fixed space

X. These constructions become more interesting when one involve continuous

maps between spaces, and try to transfer sheaves between them.

Let X and Y be two topological spaces with a continuous map f : X → Y

between them. Assume that F is an abelian sheaf on X. This allows us to define

36

Page 37: Geir Ellingsrud and John Christian Ottem October 15, 2018€¦ · between algebra and geometry: a ne varieties correspond to nitely generated k-algebras without zero-divisors, and

Chapter 1. Sheaves

an abelian sheaf f∗F on Y in by specifying the sections of f∗F over the open set

U ⊆ Y to be:

(f∗F)(U) = F(f−1U),

and the restriction maps are those of F .

The sheaf f∗F is called the pushforward sheaf or the direct image of F . It

is straightforward to see that f∗F is a sheaf and not just a presheaf. Indeed,

if Ui i is an open covering of U a set of patching data consists of section

si ∈ Γ(Ui, f∗F) = Γ(f−1U,F). That they match on the intersection means that

they coincide in Γ(Ui ∩ Uj, f∗F) = Γ(f−1Ui ∩ f−1Uj, F ), and therefore can be

glued to a section in Γ(f−1U,F) = Γ(U, f∗F) since F is a sheaf. The Locality

axiom also follows for f∗F because it holds on F .

If you want, the pushforward sheaf f∗F is just the restriction of F to the

subcategory of the open sets in X which are inverse images of open sets in Y .

Example 1.38. Let i : x → X be the inclusion of a point in X. If G is the

constant sheaf of a group G on x, then i∗G is the skyscraper sheaf G(x) from

Example 1.9.

Example 1.39. Consider an affine variety X ⊆ An and let i : X → An be the

inclusion. For each open U ⊆ An define IX(U) = f ∈ OAN (U)|f(x) = 0∀x ∈X. Then IX is a sheaf (of ideals) and we have an exact sequence

0→ IX → OAn → i∗OX → 0

This is a functorial construction:

Lemma 1.40. If g : X → Y and f : Y → Z are continuous maps between

topological spaces, and F is a sheaf on X, one has

(f g)∗F = f∗(g∗F).

Remark, this is indeed an equality, not merely an isomorphism.

Proof. A good exercise in manipulating sections.

Lemma 1.41. The functor f∗ is left exact. That is: Given an exact sequence of

sheaves on X

0 // F ′ // // F ′′ // 0

then the following sequence is exact

0 // f∗F ′ // f∗F // f∗F ′′.

Proof. As f∗F are sections of F over inverse images of opens in Y , the lemma

follows readily from the exactness of the sequence (1.6.2) on page 27 above.

37

Page 38: Geir Ellingsrud and John Christian Ottem October 15, 2018€¦ · between algebra and geometry: a ne varieties correspond to nitely generated k-algebras without zero-divisors, and

1.12. Inverse image sheaf

Exercise 4. Denote by ∗ a one point set. Let f : X → ∗ be the one and

only map. Show that f∗F = Γ(X,F) (where strictly speaking Γ(X,F) stands

for the constant sheaf on ∗ ).

1.12 Inverse image sheaf

If F is a sheaf on X and U ⊆ X is an open subset, then F defines a sheaf on U

by restriction of sections. For an arbitrary subset Z ⊆ X, this naive restriction

does not give a sheaf on Z so easily, because an open subset V ⊆ Z, is usually

not open in X. However, as in the definition of stalks, we can take limits of

F(U) as U runs over smaller and smaller open subsets containing V . This gives

the following definition:

Definition 1.42. If i : Z → X is the inclusion of a closed subset Z ⊆ X, we

define the restriction F|Z as the sheafification of the following presheaf:

V 7→ lim−→U⊇VF(U)

We can extend this idea to any continuous map f : X → Y and a sheaf G on

Y . This gives the inverse image of G, denoted f−1G, which we define as follows.

As above, we know the values of G on open subsets V on Y , but we want to

define a sheaf on X. Here the image of an open set U , f(U) need not be open in

Y , but we can look at equivalence classes of G(V ) as V gets closer and closer to

f(U). This gives the following definition:

Definition 1.43. Let f : X → Y be a continuous map; G a presheaf on Y . We

define the inverse image f−1G as the sheafification of the following presheaf:

f−1p G : U 7→ lim−→

V⊇f(U)

G(V ). (1.12.1)

Note in particular that the stalk of f−1G at a point x ∈ X is isomorphic to

Gf(x). Indeed, it suffices to verify this on the level of presheaves:

(f−1p G)x = lim−→

U3xf−1p G(U) = lim−→

U3xlim−→

V⊇f(U)

G(V )

= lim−→V 3f(x)

G(V ) = Gf(x)

1.12.1 Adjoint property

The definition of f−1G is natural, but a little bit hard to work with for actual

computations, since it involves both taking a direct limit over open sets and a

38

Page 39: Geir Ellingsrud and John Christian Ottem October 15, 2018€¦ · between algebra and geometry: a ne varieties correspond to nitely generated k-algebras without zero-divisors, and

Chapter 1. Sheaves

sheafification. What’s more important is what this sheaf does: It is the adjoint

of the functor f∗ as a functor from ShX to ShY . The precise statement is the

following:

Theorem 1.44. Let f : X → Y be a morphism, and let F be a sheaf on X and

let G be a presheaf on Y . Then we have a natural bijection

HomY (G, f∗F) ' HomX(f−1G,F)

which is functorial in F and G.

So sheaf morphisms φ : G → f∗F correspond bijectively to maps f−1G → F .

In particular, applying this to F = f−1G, and G = f∗F with the identity maps,

we see that there are canonical maps

η : G → f∗f−1G,

which is functorial in G, and

ν : f−1f∗F → F

which is functorial in F .

To prove this theorem, it will be convenient to introduce some notation. Let

us define an f -map Λ : G → F to be a collection of maps ΛV : G(V )→ F(f−1(V ))

indexed by open subsets V ⊆ Y such that

G(V )ΛV//

ρG

F(f−1V )

ρF

G(V ′)ΛV ′ // F(f−1V ′)

commutes for all V ′ ⊆ V ⊆ Y open. With this definition in mind, we can now

prove the following lemma, which implies the above theorem:

Lemma 1.45. Let f : X → Y be a continuous map. Let F be a sheaf on X and

let G be a sheaf on Y . There are canonical bijections between the following four

sets:

(1) The set of f -maps Λ : G → F .

(2) The set of sheaf maps G → f∗F .

(3) The set of sheaf maps f−1G → F .

(4) The set of presheaf maps f−1p G → F .

39

Page 40: Geir Ellingsrud and John Christian Ottem October 15, 2018€¦ · between algebra and geometry: a ne varieties correspond to nitely generated k-algebras without zero-divisors, and

1.12. Inverse image sheaf

Proof. The bijection between (3) and (4) follows by the adjoint property of

sheafification (becuase F is a sheaf), so it suffices to consider the sets in (1), (2)

and (4).

Let us first consider (1) and (2). If we are given a map of sheaves φ : G → f∗F ,

we have a map φV : G(V ) → f∗F(V ) = F(f−1V ) for each open set V ⊆ Y .

By the definition of a sheaf map, this commutes with the various restriction

mappings to smaller opens V ′ ⊆ V , so we get a well-defined f -map Λ : G → F .

Conversely, it is clear that any f -map Λ appears from a map of sheaves in this

manner, so we have established the desired bijection.

For the bijection between (1) and (4), suppose we are given a map of

presheaves f−1p G → F , so that we have a map

lim−→W⊇f(U)

G(W )→ F(U).

Applying this to U = f−1(V ), and compositing with the map G(V ) into the

direct limit, we get a map G(V )→ F(f−1V ). Again, this is compatible with the

restriction maps to smaller open sets V ′ ⊆ V , so we get a well-defined f -map

Λ : G → F . Conversely, it is clear that any f -map Λ arises in this manner, i.e.,

each ΛV factors as

G(V ) F(f−1V )

lim−→W⊇V G(W )

ΛV

φf−1V

for some map of presheaves φ : f−1p G → F : Just define φU for U ⊆ X by

composing ΛW with the restriction map to get a map G(W )→ F(f−1W )→ F(U)

for W ⊇ f(U) – the fact that Λ is an f -map means that we get an induced map

in the direct limit. Over U = f−1V , this φ makes the above diagram commute.

This completes the proof of the lemma.

40

Page 41: Geir Ellingsrud and John Christian Ottem October 15, 2018€¦ · between algebra and geometry: a ne varieties correspond to nitely generated k-algebras without zero-divisors, and

Chapter 2

Schemes

Affine schemes are the building blocks in the theory of schemes. Every scheme is

locally an affine scheme, i.e., every point has a neighbourhood which is isomorphic

to an affine scheme. This is similar to the notion of a manifold, where topological

space is locally diffeomorphic to an open set in Rn. However, affine schemes are

infinitely more complicated than open sets in Rn and can have extremely bad

singularities.

The affine schemes are basically just a geometric embodiment of rings (com-

mutative with unit). There is a one-to-one correspondence between rings and

affine scheme, and even more, given two rings the ring homomorphisms between

them correspond in a one-to-one with the scheme-maps between the two corre-

sponding scheme, but the maps they change direction. The situation is in perfect

analogy with what happens for affine varieties. The same two correspondences

holds true for those as well. However, in that case there are heavy constraints on

the rings involved, they are confined to be integral domains of finite type over

an algebraically closed field.

2.1 The spectrum of a ring

Following the Grothendieck maxim of always working in the maximal generality

the theory allows, we start out working with a commutative1 ring A with a

unit element. We are going to define a scheme written SpecA, called the prime

spectrum (primspekteret) of A or just the spectrum (spekteret) of A for short.

1Many attempts have been made on defining non-commutative geometry, and there aresome versions around. None however are so far completely satisfying.

41

Page 42: Geir Ellingsrud and John Christian Ottem October 15, 2018€¦ · between algebra and geometry: a ne varieties correspond to nitely generated k-algebras without zero-divisors, and

2.1. The spectrum of a ring

As for any scheme the structure of SpecA has two layers. There is an underlying

topological space on which lives a sheaf of rings.

Recall the situation for affine varieties and assume for a moment that A is

an algebra of finite type over the algebraically closed field k that is an integral

domain. In other words, A is the coordinate ring of a variety, say X. So, the

elements of A are the regular functions on X.

We know by Hilbert’s Nullstellensatz that points of X correspond to the

maximal ideals in A, a point x being associated with the ideal mx of regular

function vanishing at x, and this relation is biunivoque—as the french say—every

maximal ideal is the vanishing ideal of a point.

Algebraically closed sets contained in X are loci where the element of ideals a

in A vanish; i.e., they are of the shape V (a) = x ∈ X | f(x) = 0 all f ∈ a , or

with the concordance of points and maximal ideal in mind this may be phrased as

V (a) = m | m ⊆ A maximal and a ⊆ m . The variety X is supplied with the

Zariski topology whose open sets are the just the complements of the algebraically

closed sets V (a).

The guide line when generalising this is simple: Replace maximal ideals with

prime ideals! This motivates the following definition:

Definition 2.1. For a ring 2 A we define its spectrum as

SpecA = p | p ⊆ A is a prime ideal .

The set SpecA has a topology, which generalizes the Zariski topology on a

variety, and the definitions are verbatim the same. The closed sets in SpecA are

to be those of the shape

V (a) = p ∈ SpecA | p ⊇ a

where a is any ideal in A. Of course one has verify that the axioms for a topology

are satisfied. For closed sets the wording of the axioms is that the union of two

closed and the intersection of any number (finite or infinite) must be closed. And

of course, both the whole space and the empty set must be closed. The family of

subsets of the form V (a) satisfies these axioms, as follows from the next lemma:

Lemma 2.2. Let A be a commutative ring with unit and assume that ai i∈i is

a family of ideals in A. Let a and b be two ideals in A. Then the following three

statements hold true:

• V (a ∩ b) = V (a) ∪ V (b) = V (ab),

• V (∑

i ai) =⋂i V (ai),

2From now on the term ‘ring’ stands for a commutative ring having a unit element.

42

Page 43: Geir Ellingsrud and John Christian Ottem October 15, 2018€¦ · between algebra and geometry: a ne varieties correspond to nitely generated k-algebras without zero-divisors, and

Chapter 2. Schemes

• V (A) = ∅ and V (0) = SpecA.

Proof. Prime ideals are by definition proper ideals, so V (A) = ∅, and of course

the zero-ideal is contained in any ideal so V (0) = SpecA. This proves the last

statement. The second follows as easily, the sum of a family of ideals being

contained in an ideal if and only each of the ideals is.

The first of the three statements in the lemma needs an argument. The

inclusion V (a)∪ V (b) ⊆ V (a∩ b) is clear, so our task is to show that V (a∩ b) ⊆V (a) ∪ V (b). Let p be a prime ideal such that a ∩ b ⊆ p. If b * p, there is

an element b ∈ b with b /∈ p, but since ab ∈ a ∩ b for all a ∈ a, and p contains

a ∩ b, one has ab ∈ p. The ideal p being prime, one deduces that a ∈ p, and

consequently one has the inclusion a ⊆ p.

Corollary 2.3. The set V (a) where a runs through all the ideals in A is the

family of closed sets for a topology on SpecA.

The next lemma is about what inclusions between the closed sets of SpecA

hold, and we recognise them all from the theory of varieties.

Lemma 2.4. One has

• V (a) ⊆ V (b) if and only if√a ⊇√b. In particular one has V (a) = V (

√a).

• V (a) = ∅ if and only if a = A.

• V (a) = SpecA if and only if a ⊆√A.

Proof. The main point of the proof is that radical of an ideal equals the intersec-

tion all the prime ideals containing it, or expressed with a formula:

√a =

⋂p⊇a

p. (2.1.1)

In particular one deduces that a and√a are contained in the same prime ideals

so that V (a) = V (√a). To show the first claim in the lemma we begin with

assuming that V (a) ⊆ V (b). From (2.1.1) we then obtain

√a =

⋂p∈V (a)

p ⊇⋂

p∈V (b)

p =√b.

If now p lies in V (a) and√a ⊇√b, the chain of inclusions p ⊇

√a ⊇√b ⊇ b

imply the inclusion p ∈ V (b) as well. And by that, the first claim is proved.

The second claim follows from the fact that√a = A if and only if a = A

(one has 1n = 1!), and the last holds as the radical of A by definition satisfies√A =

√(0).

43

Page 44: Geir Ellingsrud and John Christian Ottem October 15, 2018€¦ · between algebra and geometry: a ne varieties correspond to nitely generated k-algebras without zero-divisors, and

2.1. The spectrum of a ring

Remark 2.5 (A few words about notation). Many authors, unlike Hartshorne,

like to denote points in SpecA in a way that distinguish points in SpecA from

the corresponding ideals in A, although they are identical. For example by letting

geometric points be indicated by latin letters like x and y, and the corresponding

prime ideals by px or py. This might seem superfluous, but there is a reason.

When we start working with general schemes, points do no more correspond to

prime ideals in a canonical way, so in that case its not only natural but required.

2.1.1 Generic points

The Zariski topology on SpecA differs greatly from the standard topology in

the sense that points (eg., prime ideals) can fail to be closed. In fact, the

next proposition implies that a point x ∈ SpecA is closed if and only if the

corresponding ideal is maximal.

Proposition 2.6. If p is a prime ideal of A, the closure p of the one-point

set p in SpecA equals the closed set V (p).

Proof. If the point p is contained in a smaller closed set than V (p), there is

an ideal a with p ∈ V (a) ⊆ V (p). By lemma 2.4 above this implies that√p ⊆√a ⊆ p, from which it follows that p =

√a, and hence we conclude that

V (a) = V (p).

In general there are of course typically a lot of prime ideals which are not

maximal. In fact, the rings having the property that every prime ideal is maximal

are quite special: in the noetherian case they corresponds exactly to the artinian

rings (in which case the spectrum consists of a finite set of points).

Definition 2.7. We say that a point x is a generic point of the closure V (x) =

x of the singleton x .

2.1.2 Irreducible closed subsets

Recall that a topological space X is said to be irreducible if it can not be written

as the union two proper closed subsets; that is, if X = Z∪Z ′ with Z and Z ′ both

being closed subsets, then either Z = X or Z ′ = X holds true. This concept is

well known from the theory of varieties where it is shown that a closed algebraic

subset V (a) of the affine space An(k) is irreducible if and only if the radical√a

is a prime. The corresponding statement is true in SpecA:

Lemma 2.8. A closed subset Z ⊆ SpecA is irreducible if and only if Z is of

the form Z = V (p) for a prime ideal p.

44

Page 45: Geir Ellingsrud and John Christian Ottem October 15, 2018€¦ · between algebra and geometry: a ne varieties correspond to nitely generated k-algebras without zero-divisors, and

Chapter 2. Schemes

Proof. We have seen that the closure p is irreducible. For the reverse direction,

we let V (a) ⊆ SpecA be an irreducible closed subset. Recall that one has√a =

⋂a⊆p p, and if

√a is not prime, there are more than one prime involved in

the intersection. We may divide them into two different groups thus representing√a as the intersection

√a = b ∩ b′ where b and b′ are ideals both different from

a. One concludes that V (a) = V (b) ∪ V (b′).

2.2 Functoriality

Let A and B be two rings and assume that φ : A→ B is a ring homomorphism.

Let us check that the inverse image of a prime ideal p ⊆ B really is a prime

ideal: That ab ∈ φ−1(p) means that φ(ab) = φ(a)φ(b) ∈ p, so at least one of

φ(a) or φ(b) lies in p. Hence sending p to φ−1(p) gives us a well defined map

SpecB → SpecA,and we shall denote this map by φ∗.

Lemma 2.9. Assume that φ : A→ B is a map of rings. Then the induced map

between the ring spectra φ∗ : SpecB → SpecA is continuous.

Proof. We are to verify that inverse images of closed sets are closed, so let a ⊆ A

be an ideal. The the following equalities hold

(φ∗)−1(V (a)) = p ⊆ B | φ−1(p) ⊇ a = p ⊆ B | p ⊇ φ(a) = V (φ(a)B),

since p ⊇ φ(a) if and only if φ−1(p) ⊇ a as φ−1(φ(a)) ⊇ a.

This is also a functorial construction since clearly φ−1(ψ−1(p)) = (ψφ)−1(p)

where φ and ψ are two composable ring homomorphism, and of course one has

id∗A = idSpecA.

This sets up a contravariant functor from the category rings of rings to the

category top of topological spaces. It sends a ring A to the prime spectrum SpecA

and a map φ : A→ B of rings to the continuous map φ∗ : SpecB → SpecA.

We include two prototype examples:

Example 2.10 (Spec(A/a)). If a ⊆ A is an ideal, the ring homomorphism

A→ A/a induces a map Spec(A/a)→ SpecA. Note that prime ideals in A/a

correspond exactly to prime ideals p in A containing a: This is exactly our closed

set V (a). Thus the induced map on spectra corresponds to the inclusion of V (a)

in SpecA. This is the prototype of a closed immersion.

Example 2.11 (SpecAf ). For an element f ∈ A we can consider the localization

Af of A in which f is inverted, and the corresponding ring homomorphism

A → Af . The prime ideals in the localized ring Af stand in a natural one-

to-one correspondence with the prime ideals p of A not containing f , that is,

45

Page 46: Geir Ellingsrud and John Christian Ottem October 15, 2018€¦ · between algebra and geometry: a ne varieties correspond to nitely generated k-algebras without zero-divisors, and

2.3. First examples

the complement D(f) = SpecA − V (f). The induced map SpecAf → SpecA

corresponds to the inclusion of the open set D(f) of SpecA. This is an example

of an open immersion.

2.3 First examples

Example 2.12 (Fields). If K is a field, the prime spectrum SpecK has only

one element the zero-ideal being the only ideal i K. This also holds true for

local rings with the property that all elements in the maximal ideal are nilpotent,

i.e., the radical√

(o) of the ring is a maximal ideal. For noetherian rings this is

equivalent to the ring being an artinian local rings.

Example 2.13. Show that the converse of the last statement in the example

above is true. That is if SpecA has just one element, A is a local ring all whose

non-units are nilpotent.

Example 2.14. The ring R = C[x]/(x2) is not a field, but it has only one prime

ideal (namely the ideal (x)). Note that (0) is not prime, since x2 = 0, but

x 6∈ (0).

Example 2.15 (Discrete valuation rings). A discrete valuation ring A has only

two prime ideals, the maximal ideal m and the zero ideal (0). So its prime

spectrum SpecA has just two points, and SpecA = η, x with x corresponding

to the maximal ideal m and η corresponding to (0). The point x is closed in

SpecA, and therefore the set η being its complement is open. So η is an open

point! The point η is the generic point of SpecA; its closure is the whole SpecA.

Certainly the prime spectrum SpecA is not Hausdorff, as η is contained in the

only open set containing x, the whole space.

The spectrum of a DVR

Example 2.16 (SpecZ). There are two types of prime ideals in Z . There

is the zero-ideal and there are the maximal ideals (p)Z, one for each prime p.

These are closed points in SpecZ, however one has V (0) = SpecZ, so the point

corresponding to the zero-ideal is a generic point, the closure is the whole of

SpecZ.

46

Page 47: Geir Ellingsrud and John Christian Ottem October 15, 2018€¦ · between algebra and geometry: a ne varieties correspond to nitely generated k-algebras without zero-divisors, and

Chapter 2. Schemes

The spectrum SpecZ— a la Mumford

Example 2.17. If R is a domain, if an ideal of the form (r) r ∈ R is prime,

then r is irreducible (meaning, if r = ab, then either a or b is a unit). Indeed, if

r = ab, then since (r) is prime, then without loss of generality a ∈ (r), so a = xr

for some x ∈ R, and hence r = ab = xrb, so bx = 1 (since R is a domain).

The converse however is not true. Let R = Z[√−5]. Then 2 ∈ R is irreducible,

and 2 · 3 = 6 = (1 +√−5)(1−

√−5), so the ideal (2) is not prime (the factors

involving√−5 are not multiples of 2).

Example 2.18 (PIDs). If R is a principal ideal domain (PID), any ideal is of

the form (r) for some r ∈ R. Also, every PID is a unique factorization domain

(same argument as over Z). So the points of SpecR corresponds to (0) and the

set of irreducible elements (modulo units).

Example 2.19. For R = C[[x]], the non-zero ideals are (xn). R is a PID, so the

prime ideals are (0) and (x).

Example 2.20. The ring R = C[[x2, x3]] is not a PID. The non-zero ideals are

(xn + cxn+1) and (xn, xn+1). Of these, only (x2, x3) is prime (Prove this!).

Exercise 5. Let A′ → A be a surjection of commutative rings whose kernel is

nilpotent. Show that the map SpecA→ SpecA′ is a homeomorphism.

Exercise 6 (Direct products of rings). Assume that e1, . . . , er is a complete set of

orthogonal idempotents in the ring A, meaning that one has 1 = e1 +e2 + · · ·+er,

that eiej = 0 when i 6= j, and that e2i = ei. Such a set of idempotents

corresponds to a decomposition of A as the direct product A = A1 ×A2 × . . . Arwhere Ai = eiA.

(i) Let a be an ideal in A and let ai = aei. Show that ai is an ideal in Ai and

that a = a1 + a2 + · · ·+ ar.

(ii) Show that a is a prime ideal if and only ai = Ai for all but one index i0and ai0 is a prime ideal in Ai0 .

(iii) Show that SpecA = SpecA1 ∪ SpecA2 ∪ · · · ∪ SpecAr, the union being

disjoint and each SpecAi being open in SpecA.

Exercise 7. Assume that A1 and A2 are rings and let A = A1 ×A2. Show that

SpecA is not connected; i.e., SpecA = SpecA1 ∪ SpecA2 and the union being

disjoint with both SpecA1 and SpecA2 being open in SpecA.

47

Page 48: Geir Ellingsrud and John Christian Ottem October 15, 2018€¦ · between algebra and geometry: a ne varieties correspond to nitely generated k-algebras without zero-divisors, and

2.3. First examples

Show that every ideal a in A can be written as a = a1+a2 where a1 ⊆ A1× 0 and a2 ⊆ 0 × A2. Hint: 1A = (1A1 , 0) + (0, 1A2).

Show that a is prime if and only if ai’s is prime in A1× 0 (or in 0 ×A2)

and the other one equals Aj.

Assume that A is a ring. Then SpecA is disconnected, i.e., has more than

two connected components, if and only if A has two idempotents e1 and e2 with

e1 + e2 = 1.

Exercise 8. Show that SpecA is connected if and only if the only two idempo-

tents in A are 1 and 0.

Exercise 9. Show that the closure p of a point p ∈ SpecA is given by p = V (p).

Show that a point p in SpecA is closed if and only if p is a maximal ideal.

Exercise 10. Let a ⊆ A be an ideal. Show that√a =

⋂p⊇a p. Hint: If f /∈

√a

the ideal aAf is an ideal in the localization Af , hence contained in a maximal

ideal.

Exercise 11. Let k be an algebraically closed field. Let X ⊆ Ank be the set of

closed points with the induced topology. Show that X is homeomorphic to the

variety kn, that is kn with the Zariski topology.

2.3.1 The world of schemes vs the world of varieties

Comparing the world of varieties with the world of schemes, one notices many

similarities, but there are also quite a few dramatic differences. And in In

some sense the category of varieties over an algebraically closed field k is a full

subcategory of the category of schemes, as we are going to elucidate in full later

on. Now we just give a glimpse into the situation taking a look at the simplest

case, namely the affine varieties An(k).

If one lets A = k[x1, . . . , xn] denote the ring of polynomials over k in the

variables x1, . . . , xn, one knows thanks to Hilbert’s Nullstellensatz that then

maximal ideals in A stand in a one-to-one correspondence with the points of the

affine space An(k); they are all of the form (x1 − a1, . . . , xn − an) with the ai’s

being elements in k.

The affine variety An(k) is the subset of the scheme Ank = SpecA consisting

of the closed points, or the points in SpecA corresponding to maximal ideals.

The good old Zariski topology on the variety An(k) is the induced topology.

Indeed, the closed sets of the induced topology are by definition all of the form

V (a) ∩ An(k), which clearly equals the “old” V (a) from the world of varieties.

The scheme SpecA has however many more points that are not closed. There

are a great many prime ideals in A not being maximal; at least if n ≥ 2.

48

Page 49: Geir Ellingsrud and John Christian Ottem October 15, 2018€¦ · between algebra and geometry: a ne varieties correspond to nitely generated k-algebras without zero-divisors, and

Chapter 2. Schemes

Example 2.21 (The affine line A1k = Spec k[x]). In the polynomial ring k[x]

all ideals are principal, and all non-zero prime ideals are maximal. They are

of the form (f(x)) where f(x) is an irreducible polynomial, hence of the form

(x− a) since we have assumed that k is algebraically closed. There is only one

non-closed point in Spec k[x], the point η corresponding to the zero-ideal. The

closure η is the whole line A1k. We called η the generic point.

Example 2.22 (The affine plane A2k = Spec k[x1, x2]). In this case the maximal

ideals are of the shape (x1−a1, x2−a2) and constitute all the closed points of A2k.

Every irreducible polynomial f(x1, x2) generates a prime ideal ηf = (f) which

together with the zero ideal are all non-maximal prime ideals. In addition to the

point ηf , The members of the closed set V (f(x1, x2)) are all the closed points

corresponding to ideals (x1 − a1, x2 − a2) containing f(x1, x2). This condition

being equivalent to f(a1, a2) = 0, the closed points of V (f(x1, x2)) correspond

to what one in the world of varieties would call the curve given by the equation

f(x1, x2) = 0.

2.4 Distinguished open sets

There is no way to describe the open sets in SpecA as simply and elegantly

as the closed sets can be described. However there is a natural basis for the

topology on SpecA whose sets are easily defined, and it turns out to be very

useful. For every element f ∈ A we let D(f) be the complement of the closed

set V (f), that is

D(f) = p | f /∈ p .

These are clearly open sets and are called distinguished open sets.

Lemma 2.23. The open sets D(f) form a basis for the topology of SpecA when

f runs trough A.

Proof. We are to show that any open subset U of SpecA can be written as

the union of distinguished open sets, that is, of sets of the form D(f). Let the

complement U c of U be given as U c = V (a), and choose a set fi i of generators

for the ideal a (not necessarily a finite set in general). Then we have

U = V (a)c = V (∑i

(fi))c = (

⋂i

V (fi))c =

⋃i

D(fi).

Exercise 12. Show that D(f) = ∅ if and only if f is nilpotent. Hint: Use that√(0) =

⋂p∈SpecA p.

49

Page 50: Geir Ellingsrud and John Christian Ottem October 15, 2018€¦ · between algebra and geometry: a ne varieties correspond to nitely generated k-algebras without zero-divisors, and

2.5. Examples

2.4.1 Basic properties of distinguished sets

We proceed by giving some of the basic properties of the distinguished open sets

needed in the theory.

Lemma 2.24. A family D(fi) i forms an open covering of SpecA if and only

if one may write 1 =∑

i aifi with the ai’s being elements from A only a finite

number of which are non-zero. The topology of SpecA is quasi-compact.

Proof. One has V (∑

i(fi))c = (

⋂i V (fi))

c =⋃iD(fi), so the open sets D(fi)

constitute a covering if and only if the ideal generated by the fi’s is the whole

ring A, that is 1 belongs there.

This shows in particular that any covering by distinguished open sets may be

reduced to a finite one. As the distinguished sets form a basis for the topology,

any open cover may be refined to one whose sets all are distinguished, and hence

it can be reduced to a finite covering. Thus SpecA is quasi-compact.3

Lemma 2.25. One has D(f) ∩D(g) = D(fg).

Proof. If p is a prime ideal, f /∈ p and g /∈ p hold trues simultaneously if and

only if fg /∈ p.

Lemma 2.26. One has D(g) ⊆ D(f) if and only if gn ∈ (f) for a suitable

natural number n. In particular one has D(f) = D(fn) for all natural numbers

n.

Proof. The inclusion D(g) ⊆ D(f) holds if and only if V (f) ⊆ V (g), and by

lemma 2.4 om page 43 this is true if and only if (g) ⊆√

(f), i.e., if and only if

gn ∈ (f) for a suitable n.

In fact, the inclusion D(g) ⊆ D(f) is equivalent to the localization map

τ : A→ Ag extending to a map Af → Ag. Indeed, τ extends if and only if τ(f),

i.e., f regarded as en element in Ag is invertible, that is to say there is an c ∈ Aand an m ∈ N with gm(fc− 1), in turn, this is equivalent to gm = cf for some c

and some m.

2.5 Examples

One may think about the elements f in A as some sort of functions on SpecA.

They are not functions in the usual sense so we speak about an analogy. If x is

a point in SpecA corresponding to the prime ideal p, one has the fraction field

3In many contexts a compact spaces is by definition also Hausdorff. In our context whereHausdorff topologies are rare, one uses the term quasi-compact for spaces such that every opencovering can be reduced to a finite one.

50

Page 51: Geir Ellingsrud and John Christian Ottem October 15, 2018€¦ · between algebra and geometry: a ne varieties correspond to nitely generated k-algebras without zero-divisors, and

Chapter 2. Schemes

k(x) = K(A/p) of the quotient A/p, and the element f reduced modulo p gives

an element f(x) ∈ k(p), which may considered as the “value” of a at x; clearly

f(x) = 0 if and only if f ∈ p. Be aware that “values” the element f take at

points, are in lying in different fields that might vary with the point. There is

also then pathology of nilpotents. A nilpotent element f in A is contained in all

prime ideals and vanishes everywhere on SpecA, but is still a non-zero function!

Example 2.27 (SpecZ again). Assume that f ∈ SpecZ is a number. In the

analogy with functions, the “value” f takes at a prime p is the residue class of f

modulo p; that is, the of f in the finite field Fp = Z/pZ.

The reduction mod p-map Z → Fp induces a map SpecFp → SpecZ. The

one and only point in SpecFp is sent to the point in SpecZ corresponding to the

maximal ideal (p) generated by p. On the level of structure sheaves, the map is

just the restriction map.

The inclusion Z ⊆ Q of the integers in the field of rational numbers induces

likewise a map SpecQ→ SpecZ, that sends the unique point in SpecQ to the

generic point η of SpecZ.

Every ring A has (since by our convention it contains a unit element) a prime

ring, i.e., the subring generated by 1. Hence there is a canonical (and in fact,

unique) map SpecA→ SpecZ.

This canonical map factors through the map SpecFp → SpecZ described

above if and only if A is of characteristic p. This is clear if one considers the

diagram on the ring level:

Z //

Fp

A

the canonical map Z→ A goes via Fp if and only if A is of characteristic p.

Exercise 13. In the same vain, show that a ring A is Q-algebra (that is, contains

a copy of Q) is and only if the canonical map SpecA→ SpecZ factors through

the generic point SpecQ→ SpecZ.

Example 2.28 (SpecZ[i]). The inclusion Z ⊆ Z[i] induces a continuous map

φ : SpecZ[i]→ SpecZ.

We will study SpecZ[i] by studying the fibers of this map. If p ∈ Z is a prime,

the fiber over (p)Z consists of those primes that contain (p)Z[i]. These come in

three flavours:

• p stays prime in Z[i] and the fiber over (p)Z has one element, namely

(p)Z[i]. This happens if and only if p ≡ 3 mod 4.

51

Page 52: Geir Ellingsrud and John Christian Ottem October 15, 2018€¦ · between algebra and geometry: a ne varieties correspond to nitely generated k-algebras without zero-divisors, and

2.5. Examples

• p splits into a product of two different prime ideals, and the fiber consists

of these two points. This happens if and only if p ≡ 1 mod 4.

• p factors into a product of repeated primes (such a prime is said to ‘ramify’

and only (2) does this here).

Let us explain the last point. The ideal (2)Z[i] is not radical: Note that

(2)Z[i] = (2i)Z[i] = (1 + i)2Z[i].

So the fiber consists of the single prime (1 + i)Z[i].

The following picture shows Spec(Z[i]):

The spectrum Spec(Z[i])

Now here’s where the Galois group comes into play. The group G =

Gal(Q[i]/Q) permutes the primes in Spec(Z[i]) sitting over any p in Spec(Z). In

this case

G = Z/2Z = 〈g〉

where g is the order two element given by complex conjugation. So for instance

if you look at the primes sitting over say (5), namely (2 + i) and (2− i), you see

that complex conjugation maps one into the other.

So we think of Spec(Z[i]) as sitting over Spec(Z) and G permuting the fibers

over in the base space Spec(Z).

Exercise 14. (i) Show that the fiber of φ over a prime p is homeomorphic to

SpecFp[x]/(x2 + 1)

and that dimFp Fp[x]/(x2 + 1) =2. Hint: Use that Z[i] = Z[x]/(x2 + 1).

(ii) Show that Fp[x]/(x2 + 1) is a field if and only if x2 + 1 does not have a

root in Fp.(iii) Show that Fp[x]/(x2 + 1) is a field if and only if (p)Z[i] is a prime ideal.

Example 2.29 (SpecR[t]). It is instructive to take a closer look at the prime

spectrum SpecR[t]. The ring R[t] being a principal ideal domain, any prime

52

Page 53: Geir Ellingsrud and John Christian Ottem October 15, 2018€¦ · between algebra and geometry: a ne varieties correspond to nitely generated k-algebras without zero-divisors, and

Chapter 2. Schemes

ideal has the shape (f(t)) where f(t) is an irreducible polynomial, and we may

well assume f(t) to be monic. From the the fundamental theorem of algebra it

follows that either f is linear, that is, f(t) = t− a for α ∈ R, or f is quadratic

with to conjugate complex and non-real roots; That is, f(t) = (t − a)(t − a)

with a ∈ C but a /∈ R. This shows that the closed points in SpecR[t] may be

identified the set of pairs a, a for a ∈ C. And of course, there is the generic

point η corresponding to the zero ideal.

The Galois group G = Gal(C/R) = Z/2Z acts on C[t] via the conjugation

map σ; that is, the map that sends a polynomial f(t) =∑ait

i into∑

i aiti. The

corresponding scheme map from SpecC[t] to SpecC[t] defines an action of Z/2Zon SpecC[t], and the points of SpecR[t] are the quotient of points of SpecC[t]

by this action. We just saw this holds for closed points, and clearly the generic

point of SpecC[t] is invariant and corresponds to the generic point of SpecR[t].

The quotient map is the map SpecC[t] → SpecR[t] induced by the inclusion

R[t] ⊆ C[t].

Example 2.30. Generalizing the previous example, let K ⊆ L be a Galois

extension of fields with Galois group G; i.e., G acts on L and K = LG, the

subfield of invariant elements. This action induces and action on L[t] by letting the

action of a group element g on the polynomial f(t) =∑ait

i be g(f) =∑g(ai)t

i.

The map K[t] ⊆ L[t] induces a map π : SpecL[t] → SpecK[t]. Show that

π has the following universal property: For any affine scheme SpecA and any

invariant morphism ψ : SpecL[t]→ SpecA, i.e., a morphism such that ψ g = ψ

for all g, the map ψ factors through π; that is, there is a ψ′ : SpecK[t]→ SpecA

with ψ = ψ′ π. The diagram looks like

SpecL[t]g //

&&

SpecL[t]

xx

ψ // SpecA

SpecK[t]

44

2.6 The structure sheaf on SpecA

We have now come to point where we define a structure sheaf on the topological

space SpecA. This is a sheaf of rings OSpecA whose stalks all are local rings, so

that the pair (SpecA,OSpecA) is a ringed space.

The two paramount properties of the structure sheaf OSpecA are the following:

• Sections over distinguished opens: Γ(D(f),OSpecA) = Af .

• Stalks: OSpecA,x = Apx .

53

Page 54: Geir Ellingsrud and John Christian Ottem October 15, 2018€¦ · between algebra and geometry: a ne varieties correspond to nitely generated k-algebras without zero-divisors, and

2.6. The structure sheaf on SpecA

These are two properties that one uses all the time when working with

the structure sheaf on an affine scheme. Moreover, as we will see, they even

characterize the structure sheaf uniquely.

2.6.1 Motivation

The model for the structure sheaf OSpecA on the prime spectrum SpecA of a

ring A is the sheaf of regular functions on an affine variety X, and when we

subsequently define it we mimic what happens on varieties, so it is worthwhile

recalling that situation.

Let X be a variety with coordinate A; that is, A is the ring of globally defined

regular functions on X. The fraction field K of A is the field of rational functions

on X, and the regular functions on any open set are elements in K.

For different open sets U the set of functions regular in U form different

subrings of K, and if V ⊆ U is an open contained in U , the ring of regular

functions Γ(V,OX) on V of course contains the ring Γ(U,OX) of those regular on

the bigger set U . The restriction map is nothing but the inclusion Γ(U,OX) ⊆Γ(V,OX); it does nothing to the functions in K; just considers the functions in

Γ(U,OX) to lie in Γ(V,OX).

Those function regular on the distinguished open set D(f) are allowed to

have powers of f in the denominator, and they constitute the subring Af ⊆ K

of elements of the form af−n. If D(g) is another distinguished open set that

is contained in D(f), i.e., D(g) ⊆ D(f), one has f = cgn, and hence Ag ⊆ Af(since f−1 = cg−n).

The general situation differs only significantly from the situation of varieties

in that the localization maps Af → Ag are no more necessarily injective, but the

formalism is the same.

The standard example of restriction maps not being injective is the union of

two varieties; to make it simple say the union of two line in the plane given by

the equation xy = 0; i.e., X = Spec k[x, y]/(xy). The regular function x vanishes

identically on the open D(y) where y 6= 0, and the regular function y vanishes

on D(x). So this is by no means a big mystery, it naturally appears once we

allow reducible spaces into the mix.

2.6.2 Definition of the structure sheaf OSpecA

One may straight away write down a definition of the structure sheaf, but many

students experience this as coming out of the blue. So we find it instructive to

make a detour and start with the definition of a B-presheaf that has a virtue of

being intuitive.

54

Page 55: Geir Ellingsrud and John Christian Ottem October 15, 2018€¦ · between algebra and geometry: a ne varieties correspond to nitely generated k-algebras without zero-divisors, and

Chapter 2. Schemes

Definition 2.31. Let B be the collection of distinguished open subsets D(f).

We define the B-presheaf O by

O(D(f)) = Af

and for D(g) ⊆ D(f) we let the restriction map be localization map Af → Ag.

Let SD(f) be the multiplicative system s ∈ A | s 6∈ p ∀p ∈ D(f). There is

a localization map τ : Af → S−1D(f)A since f ∈ SD(f). The following lemma says

that the ring O(D(f)) is independent of the representative f for D(f):

Lemma 2.32. The map τ is an isomorphism, permitting us to identify Af =

S−1D(f)A

Proof. The point is that any element s ∈ SD(f) does not lie in p for any p ∈ D(f),

in other words, one has D(s) ⊆ D(f). This is equivalent to√

(f) ⊆√

(s), and

one may write fn = cs for an appropriate c ∈ A and n ∈ N . Assume that

af−m ∈ Af maps to zero in S−1D(f)A. This means that sa = 0 for some s ∈ SD(f).

But then fna = csa = 0, and therefore a = 0 in Af . This shows that the map τ

is injective. To see that is surjective, take any as−1 in S−1D(f)A and write is as

as−1 = ca(cfn)−1 = caf−n.

Proposition 2.33. O is a B-sheaf of rings.

The situation is as follows. We are given a distinguished set D(f) and an

open covering D(f) =⋃i∈I D(fi). Of course then D(fi) ⊆ D(f), and we have

localization maps τi : Af → Afi and τij : Afi → Afifj . The statement in the

Proposition is equivalent to the exactness of the following sequence

0 // Afα //∏

iAfiρ //∏

i,j Afifj (2.6.1)

where α(a)i = τi(a) and ρ((ai))i,j = (τij(ai)− τji(aj)). It is clear that α ρ = 0

since τij τi = τji τj.

Lemma 2.34. The sequence (2.6.1) is exact.

Proof. We start by observing that we assume A = Af—that is f = 1— indeed,

one has (Af)fi = Afi and (Af)fifj = Afifj since fnii = hif for suitable natural

numbers ni.

Then to the proof: That α(a) = 0 means that a is mapped to zero in each of

the localizations Afi , just means that a power of fi kills a. So for each i one has

fnii a = 0 for natural numbers ni. The open sets D(fi) covers D(f) which then

is covered by the D(fnii )’s as well. Thus one we may write 1 =∑

i bifnii , which

55

Page 56: Geir Ellingsrud and John Christian Ottem October 15, 2018€¦ · between algebra and geometry: a ne varieties correspond to nitely generated k-algebras without zero-divisors, and

2.6. The structure sheaf on SpecA

upon multiplication by a gives

a =∑i

bifnii a = 0.

In down-to-earth terms, the equality ker ρ = imα means the following:

Assume given a sequence of elements ai ∈ Afi such that ai and aj are mapped

to the same element in Afifj for every pair i, j of indices. Then there ought to

be an a ∈ A, such that every ai is the image of a in Afi , i.e., τi(a) = ai.

Each ai can be written as ai = bi/fnii where bi ∈ A, and since the indices are

finite in number one may replace ni with n = maxi n. That ai and aj induce the

same element in the localization Afifj means that we have the equations

fNi fNj (bif

nj − bjfni ) = 0, (2.6.2)

where N a priori depends on i and j, but again due to there being only finitely

many indices, it can be chosen independent of the indices. Equation (2.6.2) gives

bifNi f

mj − bjfNj fmi = 0 (2.6.3)

where m = N + n. Putting b′i = bifNi we see that ai equals b′i/f

mi in Afi , and

equation (2.6.3) takes the form

b′ifmj − b′jfmi = 0. (2.6.4)

Now D(fmi ) = D(fi), and the distinguished open sets D(fmi ) form an open

covering of SpecA. Therefore we may write 1 =∑

i cifmi . Letting a =

∑i cib

′i,

we find

afmj =∑i

cib′ifmj =

∑i

cib′jf

mi = b′j

∑i

cifmi = b′j

and hence a = b′j/fmi .

Using Proposition 1.28 of Chapter 1, we may now make the following defini-

tion:

Definition 2.35. We let OSpecA be the unique sheaf extending the B-sheaf O.

To end this section, we prove that our sheaf indeed satisfies the two properties

we want:

Proposition 2.36. The sheaf OSpecA on SpecA as defined above is a sheaf of

rings satisfying the two paramount properties above, namely

• Sections over distinguished opens: Γ(D(f),OSpecA) = Af .

56

Page 57: Geir Ellingsrud and John Christian Ottem October 15, 2018€¦ · between algebra and geometry: a ne varieties correspond to nitely generated k-algebras without zero-divisors, and

Chapter 2. Schemes

• Stalks: OSpecA,x = Apx.

In particular, Γ(X,OX) = A.

Proof. We defined O so that the first property would hold, so only the second

point needs proof. Using the first part, and the fact that you can compute stalks

by running through open sets of the form D(f), it suffices to show that the

canonical homomorphism

φ : lim−→f 6∈p

Af → Ap

isomorphism. φ is surjective: Any element in Ap is of the form a/f with f 6∈ p.

This element lies in Af and so it lies in the image of φ. φ is injective: if af−n ∈ Afis mapped to 0 in Ap, then ∃g 6∈ p such that ga = 0, hence af−n = 0 ∈ Agf and

so φ is injective.

The last statement follows by taking f = 1.

2.6.3 Maps between the structure sheaves of two spectra

Previously, we assigned to any map φ : A → B of rings a continuous map

φ∗ : SpecB → SpecA.

We climb one step in the hierarchy of structures and shall associate to φ a

map of sheaves of rings

φ# : OSpecA → φ∗OSpecB.

By a simple patching argument, it suffices to tell what φ# should do to the

sections over open sets belonging to a basis for the topology as long as this

definition is compatible with the restriction maps; and of course, the basis we

shall use is the basis of the distinguished open sets. Everything follows from the

following lemma:

Lemma 2.37. Let φ : A → B be a map of rings and let f ∈ A be an element.

Then (φ∗)−1(D(f)) = D(φ(f))

Proof. We have

(φ∗)−1(D(f)) = p ⊆ B | f /∈ φ−1(p) = p ⊆ B | φ(f) /∈ p = D(φ(f)).

This means that Γ(D(f), φ∗OSpecB) = Bφ(f), and we know that Γ(D(f),OSpecA) =

Af . The original map of rings φ : A → B now localizes to a map Af → Bφ(f),

sending af−n to φ(a)φ(f)−n, and that shall be the map φ# on sections over the

distinguished open set D(f).

57

Page 58: Geir Ellingsrud and John Christian Ottem October 15, 2018€¦ · between algebra and geometry: a ne varieties correspond to nitely generated k-algebras without zero-divisors, and

2.7. Schemes

It is obvious, but anyhow the matter of some easy work, to check that this

definition is compatible with the restriction maps among distinguished open sets:

Indeed, if D(g) ⊆ D(f), we can as usual write gn = cf , and the localization map

Af → Ag sends af−m to acmg−nm. One has φ(g)n = φ(c)φ(f), and the diagram

below is commutative:

Af //

Ag

Bφ(f)

// Bφ(g),

which is the required compatibility.

2.7 Schemes

So far, we have defined a topological space SpecA together with a sheaf of rings

on it. This is the prototype of a ringed space:

Definition 2.38. A ringed space is a pair (X,OX) where X is a topological

space and OX is a sheaf of rings on X.

A morphism of ringed spaces is a pair (f, f#) : (X,OX) → (Y,OY ) where

f : X → Y is continuous, and

f# : OY → f∗OX

is a map of sheaves of rings on Y (so that f#(U) is a ring homomorphism for

each open U ⊆ X).

Example 2.39. Consider a manifold X with the sheaf C∞(X) of smooth func-

tions. The pair (X,C∞(X)) forms a ringed space. Indeed, any morphism

f : X → Y of manifolds is smooth if and only if for every local section h of

C∞(Y ) the composition f#(h) = h f is a local section of C∞(X). This gives a

map

f# : C∞(Y )→ f∗C∞(X)

Thus a map f gives rise in a natural way to a morphism of ringed spaces

f : (X,C∞(X)) −→ (Y,C∞(Y ))

Let us consider what happens to the germs of functions near a point: For a point

x ∈ X, the ring of germs of functions C∞(X)x is a local ring with maximal ideal

mx the functions which vanish at x (similarly for C∞(Y )y where y = f(x) ∈ Y ).

Note that the map on stalks

f# : C∞(Y )y → C∞(X)x

58

Page 59: Geir Ellingsrud and John Christian Ottem October 15, 2018€¦ · between algebra and geometry: a ne varieties correspond to nitely generated k-algebras without zero-divisors, and

Chapter 2. Schemes

maps the maximal ideal into the maximal ideal: this is simply because f(x) = y.

This example motivates the following definition:

Definition 2.40. A locally ringed space is a pair (X,OX) as above, but with

the additional requirement that for every x ∈ X, the stalk OX,x is a local ring.

A morphism of locally ringed spaces is a pair (f, f#) : (X,OX)→ (Y,OY ) as

above, with the additional requirement that for every x ∈ X, f#x : OY,f(x) → OX,x

is a local homomorphism; that is the map

f#Y,f(x) : OY,f(x) → OX,x

satisfies

(f#Y,f(x))

−1(mx) = mf(x)

where mx ⊆ OX,x and mf(x) ⊆ OY,f(x) are the maximal ideals.

The intuition behind the local homomorphism condition comes from regarding

the sections of the structure sheaf as regular functions: we want g ∈ OY to

vanish at a point y = f(x) if and only if f#(g) vanishes at x.

Remark 2.41. Explicitly, the map f#Y,f(x) : OY,f(x) → OX,x in the definition, is

the map defined by the composition

OY,f(x) = lim−→f(x)∈V

OY (V )→ lim−→f(x)∈V

OX(f−1V )→ OX,x

Finally, we can give the formal definition of a scheme.

Definition 2.42. An affine scheme is a locally ringed space (X,OX) which is

isomorphic to (SpecA,OSpecA) for some ring A. A scheme is a locally ringed

spaced (X,OX) that is locally isomorphic to an affine scheme.

So a scheme is a topological space X and OX is a sheaf of rings such that all

stalks are local rings. A locally ringed space being locally affine means that it

possesses an open cover Ui of X with each open (Ui,OX |Ui) being isomorphic

(as locally ringed spaces) to an affine scheme.

A morphism or map for short between two schemes X and Y is simply a map

φ between X and Y regarded as locally ringed spaces. It has two components, a

continuous map, that we denote φ as well slightly abusing the language, and a

map of sheaves of rings

φ# : OY → φ∗OX

with the additional requirement that φ# induce local homomorphisms on the

stalks. This is to say, for all x ∈ Y the map φ#x : OY, φ(x) → φ∗OX,x maps the

maximal ideal in OY, x into the one in φ#x OX,x.

In this way the schemes form a category, which we denote by Sch. We denote

by AffSch the subcategory of affine schemes.

59

Page 60: Geir Ellingsrud and John Christian Ottem October 15, 2018€¦ · between algebra and geometry: a ne varieties correspond to nitely generated k-algebras without zero-divisors, and

2.8. Open immersions and open subschemes

2.8 Open immersions and open subschemes

If X is a scheme and U ⊆ X is an open subset, there restriction OX |U is a sheaf

on U , making (U,OX |U ) into a locally ringed space. This is even a scheme, since

if X is covered by affines Vi = SpecAi, then each U ∩ Vi is open in Vi, hence

can be covered by affine schemes. It follows that there is a canonical scheme

structure on U , and we call (U,OX |U) an open subscheme of X and say that U

has the induced scheme structure.

As a special case, consider U = SpecAf ⊆ SpecA = X. It follows from the

definition of the sheaf OX that the restriction OX |U coincides with the structure

sheaf on SpecAf .

Remark 2.43. A word of warning: An open subscheme of an affine scheme

might not not itself be affine (we will see an example of this in Chapter 3).

2.9 Residue fields

Built into the definition of a scheme is the condition that the stalks OX,x are

local rings. In particular, for x ∈ X, we have a maximal ideal mx ⊆ OX,x, and a

corresponding field OX,x/mx.

Definition 2.44. We define the residue field of x to be k(x) = OX,x/mx.

If U ⊆ X is an open set containing x, and s ∈ OX(U) (or if s is an element

of OX,x), we let s(x) denote the class of s modulo mx in k(x).

The point of the above definition is this: For varieties, we construct the sheaf

OX from the regular functions on which we think of as continuous maps X → k.

However, in the world of schemes, we do not have the luxury of having a field k

to map into – all we know is that locally OX comes from a ring A. This means

that we have to tweak our notion of a ‘regular function’ on X; they are not maps

into some fixed field, but rather maps into the disjoint union∐

x∈X k(x). Thus

on X = SpecZ, the element s = 17 ∈ Z defines a function on X. Some of its

”values” are given by

s((2)) = 1, s((3)) = 1, s((7)) = 3, s((11)) = 6, s((17)) = 0, s(((19)) = 17

Of course the values have to be interpreted in the appropriate residue field Z/p.It is instructive to consider the simplest examples of affine schemes, namely

spectra of fields. If K is a field, then of course the underlying topological

SpecK is very simple – it is a singleton set corresponding to the zero ideal.

However, the structure sheaf O on SpecK carries more information: Morphisms

SpecL→ SpecK correspond exactly to field extensions L ⊇ K. In particular,

SpecK and SpecL are isomorphic if and only if K ' L.

60

Page 61: Geir Ellingsrud and John Christian Ottem October 15, 2018€¦ · between algebra and geometry: a ne varieties correspond to nitely generated k-algebras without zero-divisors, and

Chapter 2. Schemes

More generally, for a scheme X it is interesting to study morphisms SpecK →X for fields K. Such morphisms will be referred to as K-valued points, and will

be denoted by X(K) (this jargon will be explained below). The residue fields

satisfy the following ‘universal’ property with respect to such morphisms:

Proposition 2.45. Let K be a field. Then to give a morphism of schemes

SpecK → X is equivalent to giving a point x ∈ X plus an injection k(x)→ K

Proof. Let (f, f#) : SpecK → X be the morphism and let x be the image of

f (SpecK consists of only one point, so this is a well-defined closed point of

X.) The map on stalks is just f#x : OX,x → OSpecK,0 = K, which gives a map on

residue fields k(x) = OX,x/mx → OSpecK,0 = K. As always with non-zero maps

of fields, this has to be an injection.

Conversely, let x ∈ X and k(x) → K. We construct a map of topological

spaces f : SpecK → X, which takes SpecK to x ∈ X. We also construct a map

of structure sheaves f# : OX → f∗OSpecK : for opens U ⊆ X not containing x,

f#(U) is the zero map as f−1(U) = ∅, while for opens U ⊆ X containing x,

we need maps OX(U)→ K. These maps are constructed using the given map

k(x) → K via the compositions

OX(U) OX,x k(x) K.

This yields a morphism of schemes (f, f#) : SpecK → X.

In particular, we have for X a variety over an algebraically closed field k

X(k) = closed points of X

More generally still, we make the following definition:

Definition 2.46. Let X, S be schemes. An S-valued point is a morphism

f : S → X. If S = SpecR, we call this an R-valued point. We denote the set of

such points by X(R).

The name R-valued point can be understood from the following example.

Example 2.47. Let An = SpecZ[x1, . . . , xn] and let R be a ring. What are

the R-valued points of An? Well, a morphism g : SpecR → SpecZ[x1, . . . , xn]

is determined by the n-tuple ai = g∗(x1) for i = 1, . . . , n. Hence, An(R) = Rn.

More generally, we have for any scheme S

An(S) = HomSch(S,An) = Γ(S,OS)n.

In fancy language, this says that An represents the functor taking a scheme to

n-tuples of elements of Γ(S,OS).

61

Page 62: Geir Ellingsrud and John Christian Ottem October 15, 2018€¦ · between algebra and geometry: a ne varieties correspond to nitely generated k-algebras without zero-divisors, and

2.10. Relative schemes

Now, let

X = SpecZ[x1, . . . , xn]/I

where I = (f1, . . . , fr). The set of R-points can be found similarly: Indeed, any

morphism

g : SpecR→ SpecZ[x1, . . . , xn]/I

is determined by the n-tuple ai = g∗(x1) for i = 1, . . . , n, and those n-tuples that

occur are exactly those such that f 7→ f(a1, . . . , an) defines a homomorphism

Z[x1, . . . , xn]/I → R.

In other words, the ai are elements in R which are solutions of f1 = · · · = fr = 0.

In other words, this justifies the notation X(R) – X(R) and X(S) naturally

generalize the solution set of the polynomials f1 = · · · = fr = 0.

2.10 Relative schemes

There is also the notion of relative schemes (relative skjemaer) where a bases

scheme S is chosen. A scheme over S (et skjema over S) is scheme X with a

morphism π : X → S called the structure map. If two such schemes over S are

given, say X → S and Y → S , then a map between them is a map X → Y

compatible with the two structure maps; that is, rendering the diagram below

commutative

X //

Y

S

The schemes over S form a category Sch/S, and the set morphisms as defined

above is denoted by HomS(X, Y ). Composition of maps makes it is a functor

in both X and Y . It is a functor from Sch/S to Sets, contravariant in X and

covariant in Y . If the base scheme S is affine, say S = SpecA, a short hand

notation for Sch/Spec A is Sch/A.

As there is a canonical map from any scheme X to SpecZ, every scheme is a

Z-scheme. On the level of categories one may express this a Sch = Sch/Z.

In case of affine schemes say S = SpecA and X = SpecB, giving a map

X → S is giving a ring homomorphism A→ B, or said differently giving B the

structure of an A-algebra. So if S = SpecA the category Aff/S of affine schemes

over S is equivalent to the category Alg/A of A-algebras.

Definition 2.48. 1. A morphism f : X → Y is of locally finite type if Y has

a cover consisting of affines Vi = SpecBi such that f−1Vi can be covered

62

Page 63: Geir Ellingsrud and John Christian Ottem October 15, 2018€¦ · between algebra and geometry: a ne varieties correspond to nitely generated k-algebras without zero-divisors, and

Chapter 2. Schemes

by affine subsets of the form SpecAij, where each Aij is finitely generated

as a Bi-module.

2. f is of finite type if, in (i), one can do with a finite number of SpecAij.

3. f is finite if f−1Vi = SpecAi, where Ai is a finite Bi-module.

We say that a scheme over k is of locally finite type (resp. of finite type, resp.

finite) over k, if the morphism X → Spec k has this property.

For an affine scheme these conditions translate into the following:

• SpecA is a scheme over k if and only if there is a homomorphism k → A,

or equivalently, that A is a k-algebra.

• SpecB → SpecA is a morphism of schemes over k if and only if the induced

map A→ B is a homorphism of k-algebras.

• SpecA is of finite type over k if and only if A is a finitely generated

k-algebra.

63

Page 64: Geir Ellingsrud and John Christian Ottem October 15, 2018€¦ · between algebra and geometry: a ne varieties correspond to nitely generated k-algebras without zero-divisors, and
Page 65: Geir Ellingsrud and John Christian Ottem October 15, 2018€¦ · between algebra and geometry: a ne varieties correspond to nitely generated k-algebras without zero-divisors, and

Chapter 3

Gluing

3.1 Gluing sheaves

The setting is a scheme X with an open covering Ui i∈I with a sheaf Fi on each

open subset Ui. As usual the sheaves can take values in any category, but the

principal situation we have in mind is when the sheaves are sheaves of abelian

groups. The intersections Ui ∩ Uj are denoted by Uij, and triple intersections

Ui ∩ Uj ∩ Uk are written as Uijk.

The gluing data in this case consist of isomorphisms τji : Fi|Uij → Fj|Uij . The

idea is to identify sections of Fi|Uij with Fj|Uij using the isomorphisms τij. For

the gluing process to be feasible, the τij’s must satisfy the three conditions

• τii = idFi ,

• τji = τ−1ij ,

• τki = τkj τji,

where the last identity takes place where it can; that is, on the triple intersection

Uijk. Observe that the three conditions parallel the three requirements for a

relation being an equivalence relation; the first reflects reflectivity, the second

symmetry and the third transitivity.

The third requirement is obviously necessary for gluing to be possible. A

section si of Fi|Uijk will be identified with its image sj = τij(si) in Fj|Uijk , and in

its turn, sj is going to be equal to sk = τkj(sj). Then, of course, si and sk are

enforced to be equal, which means that τki = τkj τji.

65

Page 66: Geir Ellingsrud and John Christian Ottem October 15, 2018€¦ · between algebra and geometry: a ne varieties correspond to nitely generated k-algebras without zero-divisors, and

3.1. Gluing sheaves

Proposition 3.1. In the setting as above, there exists a unique sheaf F on X

such that there are isomorphisms υi : F |Ui → Fi satisfying υj = τji υi over the

intersections Uij.

Proof. Let V ⊆ X be an open set and let Vi = Ui ∩ V and Vij = Uij ∩ V . We

are going to define the sections of F over V , and they are of course obtained by

gluing sections of the Fi’s along Vi using the isomorphisms τij. We define

Γ(V, F ) = (si)i∈I | τji(si|Vij) = sj|Vij ⊆∏i∈I

Γ(Vi, Fi). (3.1.1)

The τij’s are maps of sheaves and therefore are compatible with all restriction

maps, so if W ⊆ V is another open set, one has τij(si|Wij) = sj|Wij

if τji(si|Vij ) =

sj|Vij . By this, the defining condition (3.1.1) is compatible with componentwise

restrictions, and they can therefore be used as the restriction maps in F . We

have thus defined a presheaf.

The first step in what remains of the proof, is to establish the isomorphisms

υi : F |Ui → Fi. To avoid getting confused by the names of the indices, we work

with a fixed index α ∈ I. Suppose V ⊆ Uα is an open set. Then naturally one

has V = Vα, and projecting from the product∏

i Γ(Vi, Fi) onto the component

Γ(Vα, Fα) = Γ(V, Fα) gives us a map of presheaves1 υα : F |Vα → Fα. This map

sends (si)i∈I to sα. We treat lovers of diagrams with a display:

Γ(V, F ) //

υα ''

∏i∈I Γ(Vi, Fi)

Γ(V, Fα),

and proceed to verify that the map just defined gives us the searched for

isomorphism:

To begin with, on the intersections Vαβ the requirement in the proposition

that υβ = τβα υα is fulfilled. This follows immediately since by the second

property of the glue, one has sβ = τβα(sα)

It is surjective: Take a section σ of Fα over some V ⊆ Uα and define

s = (τiα(σ|Viα))i∈I . Then s satisfies the condition in (3.1.1), and is a legitimate

element of Γ(V, F ); indeed, by the third gluing condition we obtain

τji(τiα(σ|Vjiα)) = τjα(σ|Vjiα)

for every i, j ∈ I, and that is just the condition in (3.1.1). As ταα(σ|Vαα) = σ by

the first property of the glue, the element s projects to the section σ of Fα.

1Restrictions operating componentwise, it is straight forward to verify this map beingcompatible with restrictions.

66

Page 67: Geir Ellingsrud and John Christian Ottem October 15, 2018€¦ · between algebra and geometry: a ne varieties correspond to nitely generated k-algebras without zero-divisors, and

Chapter 3. Gluing

It is injective: This is clear, since if sα = 0 if follows that si|Viα = τiα(sα) = 0

for all i ∈ I. Now Fα is a sheaf and the Viα constitute an open covering of Vα.

We conclude that s = 0 by the locality axiom for sheaves.

The next step is to show that F is a sheaf, and we start with the patching

axiom: So assume that Vα is an open covering of V ⊆ X and that sα ∈ Γ(Vα, F )

is a bunch of sections matching on the intersections Vαβ. Since F |Ui∩V is a sheaf—

we just checked that F |Ui is isomorphic to Fi—the sections sα|Vα∩Ui patch together

to give sections si in Γ(Ui∩V, F ) matching on Uij ∩V . This last condition means

that τij(si) = sj . By definition (si)i∈I then this is a section in Γ(V, F ) restricting

to si, and we are done. The locality axiom is easy to verify and is left to reader

to verify (do it!).

Exercise 15. Show the uniqueness statement in the proposition.

3.2 Gluing maps of sheaves

This is the easiest gluing situation we encounter in this course. The setting is

as follows. We are given two sheaves F and G on the scheme X and an open

covering Ui i of X. On each open set Ui we are given a map φi : F |Ui → G|Uiof sheaves, and the gluing conditions

• φi|Uij = φj|Uij

are assumed to be satisfied for all i, j ∈ I. In this context we have

Proposition 3.2. There exists a unique map of sheaves φ : F → G such that

φ|Ui = φi.

Proof. The salient point is this: Take any s ∈ Γ(V, F ) where V ⊆ X is open,

and let Vi = Ui ∩ V . Then φi(s|Vi) is a well defined element in Γ(Vi, G), and it

holds true that φi(s|Vij) = φj(s|Vij) by the gluing condition. Hence the sections

φi(s|Vi)’s of G|Vij patch together to a section of G over V which we define to be

φ(s). The checking of all remaining details is hassle-free and left to the zealous

student.

3.3 The gluing of schemes

The possibility of gluing different schemes together along open subschemes, gives

rise to many new schemes. The most prominent one being the projective spaces.

The gluing process is also an important part in many general existence proofs,

like in the construction of the fiber-product, which as we are going to show,

exists without restrictions in the category of schemes.

67

Page 68: Geir Ellingsrud and John Christian Ottem October 15, 2018€¦ · between algebra and geometry: a ne varieties correspond to nitely generated k-algebras without zero-divisors, and

3.3. The gluing of schemes

In the present context of scheme gluing we are given a family Xi i∈I of

schemes indexed by the set I. In each of the schemes Xi we are given a collection

of open subschemes Xij, where the second index j runs through I. They form

the glue lines in the process; i.e., the contacting surfaces that are to be glued

together. In the glued scheme they will be identified and will be equal to the

intersections of Xi and Xj. The identifications of the different pairs of the Xij’s

are encoded by a family of scheme-isomorphisms τji : Xij → Xji. Furthermore,

we let Xijk = Xik ∩Xij— this are the different parts of the triple intersections

before the gluing has been done—and we have to assume that τij(Xijk) = Xjik.

Notice that Xijk is an open subscheme of Xi.

The three following gluing condition, very much alike the ones we saw for

sheaves, must be satisfied for the gluing to be doable:

• τii = idXi .

• τij = τ−1ji .

• The isomorphism τij takes Xijk into Xjik and one has τki = τkj τji over

Xijk.

Proposition 3.3. Given gluing data as above. Then there exists a scheme X

with open immersions ψi : Xi → X such that ψi|Xij = ψj|Xji τji, and such

that the images ψi(Xi) form an open covering of X. Furthermore one has

ψi(Xij) = ψi(Xi) ∩ ψj(Xj). The scheme X is uniquely characterized by these

properties

Proof. To build the scheme X we first build the underlying topological space

X and subsequently equip it with a sheaf of rings. For the latter, we rely on

the patching technic for sheaves presented in proposition 3.1. And finally, we

need to show that X is locally affine, this follows however immediately once the

immersions ψi are in place—the Xi’s are schemes and therefore locally affine.

On the level of topological spaces, we start out with the disjoint union∐

iXi

and proceed by introducing an equivalence relation on it. We require that two

points x ∈ Xij and x′ ∈ Xji be equivalent if x′ = τji(x)—observe that if the

point x does not lie in any Xij with i 6= j, we leave it alone, and it will not be

equivalent to any other point.

The three gluing conditions imply readily that we obtain an equivalence

relation. The first requirement entails that the relation is reflexive, the second

that it is symmetric, and the third ensures it is transitive. The topological space

X is then defined to be the quotient of∐

iXi by this relation, and we declare

the topology on X to be the quotient topology. If π denotes the quotient map, a

subset U of X is open if and only if π−1(U) is open.

Topologically the maps ψi : Xi → X are just the maps induced by the open

inclusions of the Xi’s in the disjoint union∐

iXi. They are clearly injective

68

Page 69: Geir Ellingsrud and John Christian Ottem October 15, 2018€¦ · between algebra and geometry: a ne varieties correspond to nitely generated k-algebras without zero-divisors, and

Chapter 3. Gluing

since a point x ∈ Xi never is equivalent to another point in Xi. Now, X has the

quotient topology so a subset U of X is open if and only if π−1(U) is open, and

this holds if and only if ψ−1i (U) = Xi ∩ π−1(U) is open for all i. In view of the

formula

π−1(ψi(U)) =⋃j

τji(U ∩Xij)

we conclude that each ψi is and open immersion.

To simplify notation we now identify Xi and ψi(Xi), which is in concordance

with our intuitive picture of X as being the union of the Xi’s with points in

the Xij’s identified according to the τij’s. Then Xij becomes Xi ∩Xj ant Xijk

becomes the triple intersection Xi ∩Xj ∩Xk.

On Xij we have the isomorphisms τ#ji : OXj |Xij → OXi |Xij ; the sheaf-parts of

the scheme-isomorphisms τji : Xij → Xji. In view of the third gluing condition

τki = τkj τji, valid on Xijk, we obviously have τ#ki = τ#

ji τ#kj. The two first

gluing conditions translate into τ#ii = id and τ#

ji = (τ#ji )−1. The end of the story

is that the gluing properties needed to apply proposition 3.1 are satisfied, and

we are enabled to glue the different OXi ’s together and thus to equip X with a

sheaf of rings. This sheaf of rings restricts to OXi on each of the open subsets

Xi, and therefore it is a sheaf local rings. So (X,OX) is a locally ringed space

that is locally affine, hence a scheme.

The uniqueness property is, as usual, left to the industrious student.

Before and after gluing

3.4 Global sections of glued schemes

The standard exact sequence for computing global sections from an open covering

is a valuable tool in the setting of glued schemes. If X is obtained by gluing the

open subschemes Xi along τji : Xij → Xji it reads:

0 // Γ(X,OX) α //⊕

i Γ(Xi,OXi)ρ //⊕

i,j Γ(Xij,OXij)

where ρ(si)i∈I = (si|Xij − τ#ij (sj|Xji)) and α(s) = (ψ#

i (s))i∈I .

69

Page 70: Geir Ellingsrud and John Christian Ottem October 15, 2018€¦ · between algebra and geometry: a ne varieties correspond to nitely generated k-algebras without zero-divisors, and

3.5. The gluing of morphisms

3.5 The gluing of morphisms

Assume given schemes X and Y and an open covering Ui i∈I of X. Assume

further that there is given a family of morphisms φ?i : Ui → Y which match on

the intersections Uij = Ui ∩Uj . The aim of this paragraph is show that they can

be glued together to give a morphism X → Y :

Proposition 3.4. In the situation just described, there exists a unique map of

schemes φ : X → Y such that φ|Ui = φi.

Proof. Clearly the underlying topological map is well defined, so if U ⊆ Y is

an open set, we have to define φ# : Γ(U,OY )→ Γ(U, φ∗OX) = Γ(φ−1U,OX). So

take any section s of OY over U . This gives sections ti = φ#i (s) of OX |Ui , but

since φ#i and φ#

j restrict to the same map on Uij, one has ti|Uij = tj|Uij . The titherefore patch together to a section of OX over U , which is the section we are

aiming at: We define φ#(s) to be t. The checking of the remaining details is left

to student (as usual).

3.6 The category of affine schemes

The assignment A → SpecA and φ → Spec(φ) gives a contravariant functor

from the category Rings of rings to the category AffSch of affine schemes. There

is also a contravariant functor the other way around. Taking the global sections

of the structure sheaf OSpecA gives us the ring A back. Furthermore, a map of

affine schemes f : SpecB → SpecA, comes with a map of sheaves f# : OSpecA →f∗OSpecB. Taking global sections gives a ring homomorphism

A = Γ(SpecA,OSpecA)→ Γ(SpecA, φ∗OSpecB) = B.

which coincides with φ in the case f = Spec(φ).

Let X,T be two schemes and let HomAffSch(X, Y ) denote the set of morphisms

f : X → Y between them. We can define a canonical map

ρ : HomAffSch(X, Y )→ HomRings(OY (Y ),OX(X)) (3.6.1)

which sends (f, f#) to the map f#(Y ) : OY (Y )→ OX(X).

Lemma 3.5. If X and Y are affine, the map ρ is bijective.

Proof. Write X = SpecB and Y = SpecA. By definition, we have A = OY (Y )

and B = OX(X). Given a morphism f : X → Y , let φ = ρ(f) = f#(Y ) :

B → A be the corresponding ring homomorphism. As we noted above, any ring

homomorphism induces a morphism of the corresponding ring spectra, so we

70

Page 71: Geir Ellingsrud and John Christian Ottem October 15, 2018€¦ · between algebra and geometry: a ne varieties correspond to nitely generated k-algebras without zero-divisors, and

Chapter 3. Gluing

obtain a morphism of affine scheme fφ : X → Y . To prove the lemma, we need

only prove that fφ = f .

Let x ∈ X be a point, corresponding to the prime ideal q ⊆ B, and let p ⊆ A

be the prime ideal corresponding to f(x) ∈ Y . We have a diagram

Aφ //

B

Ap

// Bq

where the vertical maps are the localization maps. We have φ(A− p) ⊆ B − q.

Hence φ−1(q) ⊆ p. Now f#x is a local homomorphism, and so in fact φ−1(q) = p.

Hence f and fφ induce the same map on the underlying topological spaces.

Moreover, we have f#x = f#

φ,x for every x ∈ X, and so also f# = f#φ .

We have established the all important theorem:

Theorem 3.6. The two functors Spec and Γ are mutually inverse and define

an equivalence between the categories Rings and AffSch.

Thus affine schemes X are characterized by their rings of global sections

Γ(X,OX), and morphisms between affine schemes X → Y are in bijective

correspondence with ring homomorphisms Γ(Y,OY )→ Γ(X,OX).

3.7 Universal properties of maps into affine schemes

For a general schemeX one can consider the associated affine scheme Spec Γ(X,OX),

but this is in general very different from X (for the projective line introduced in

the next section, Spec Γ(X,OX) is just a point). There is however still a canonical

morphism X → Spec Γ(X,OX) enjoying the following universal property:

Proposition 3.7. Let X be any scheme. Then there is a canonical (unique)

map of schemes X → Spec Γ(X,OX) inducing the identity on global sections of

the structure sheaves.

This will follow by applying the following theorem to A = Γ(X,OX):

Theorem 3.8. For any scheme X, the canonical map

ΦX : HomSch(X, SpecA)→ HomRings(A,Γ(X,OX))

given by f → f#, is bijective.

71

Page 72: Geir Ellingsrud and John Christian Ottem October 15, 2018€¦ · between algebra and geometry: a ne varieties correspond to nitely generated k-algebras without zero-divisors, and

3.7. Universal properties of maps into affine schemes

Proof. Let Ui be an affine covering of X, and consider the commutative

diagram

Hom(X, SpecA)ΦX //

_

Hom(A,Γ(X,OX)) _

∏i Hom(Ui, SpecA)

ΠΦUi//∏

i Hom(A,Γ(Ui,OX))

where the vertical maps are induced by the inclusion maps Ui → X. By the

affine schemes case (Theorem 3.6), we know that ΠΦUi is bijective. In particular,

this shows that ΦX is injective.

To show that ΦX is surjective, let β : A→ Γ(X,OX) be a ring homomorphism.

This induces maps βi : A → Γ(X,OX) → Γ(Ui,OX), and hence, morphisms

fi : Ui → SpecA. We check that the fi glue to a map f : X → SpecA. This is a

consequence of the fact that the following diagram commutes:

Γ(Ui,OUi)

**A

βi::

βj $$

β // Γ(X,OX) //

OO

Γ(Ui ∩ Uj,OX)→ Γ(V,OX)

Γ(Uj,OUj)

44

Indeed, note that for V ⊆ Ui∩Uj affine, the diagram implies that the restrictions

fi|V and fj|V induce the same element in Hom(A,Γ(V,OX)), and so they are

equal on V (using Theorem 3.6). Since this is true for any V , they are equal on

all of Ui ∩ Uj. So by gluing the fi, we obtain a morphism f : X → SpecA. We

must have ΦX(f) = β, by injectivity of ΦX and since f maps to∏

i βi via ΠΦUi .

This completes the proof.

Corollary 3.9. The canonical map ψ : X → Spec Γ(X,OX) is universal among

the maps from X to affine schemes; i.e., any map α : X → SpecA factors as

α = α′ ψ for a unique map α′ : Spec Γ(X,OX)→ SpecA.

Proof. In the theorem above, ψ corresponds to the identity map idΓ(X,OX) on

the right hand side. The morphism α′ is the map of Spec’s induced by the

ring map α# : A → Γ(X,OX). We check that it factors α: The morphism

(α′ ψ) : X → SpecA satisfies (α′ ψ)# = ψ# α# = α# and hence it coincides

with α by the above theorem.

Remark 3.10. Let X be any scheme. By the above we have a bijection

HomSch(X, SpecZ) = HomRings(Z,Γ(X,OX)).

72

Page 73: Geir Ellingsrud and John Christian Ottem October 15, 2018€¦ · between algebra and geometry: a ne varieties correspond to nitely generated k-algebras without zero-divisors, and

Chapter 3. Gluing

The set on the right is clearly a one point set and thus we find that SpecZ is

final object in the category Sch.

Sch also has a initial object: The spectrum of the zero ring, Spec 0, which

has the empty set as underlying topological space. Given any scheme X there

is clearly a unique morphism Spec 0→ X, which on the level of sheaves sends

every section to zero.

Example 3.11 (The morphism Spec OX,x → X). Let X be a scheme and let

x ∈ X be a point. There is a canonical morphism

f : Spec(OX,x)→ X

which is induced from all the morphisms SpecAp → SpecA. The image I =

im f ⊆ X of f is exactly the intersection of all neighbourhoods of x in X: it is a

kind of microgerm of X at x.

More geometrically consider the irreducible closed subset V = x ⊆ X

whose generic point is x. Let Y ⊆ X be a closed irreducible subset on which

V lies: V ⊆ Y and let ηY be the generic point of Y . Then our subset im f is

exactly the set of all those generic points ηY . We say that im f is the set of

specializations of x.

For example, if X = A2k = Spec(k[x, y]) and x is a closed point corresponding

to the maximal ideal m = (x − a, x − b), then I is the set consisting of x, the

generic point of A2k and the generic points of all irreducible curves passing through

x, like for example the curve (y − b)2 − (x− a)3 = 0.

3.8 The functor from varieties to schemes

We have already stated that schemes form a generalization of algebraic varieties.

On the other hand, we have also seen that even the simplest schemes (e.g.,

A2k = Spec k[x, y]) behave differently than varieties in the sense that they usually

have many non-closed points. Thus for this generalization to make sense, we

should expect that there to be a canonical way to ‘add non closed points’ to an

algebraic variety V , so that the resulting topological space has the structure of a

scheme. Let us explain what this means more precisely.

Let k be an algebraically closed field, and let V be an algebraic variety over

k. Let us first consider the case where V is an affine variety, i.e., V ⊆ Ank is

the zero-set of some ideal I ⊆ k[x1, . . . , xn]. The coordinate ring A = A(V ) =

k[x1, . . . , xn]/I is the ring of regular functions f : V → k. The scheme V sh =

SpecA is an affine scheme whose closed points is in bijection with the points of

V . So any affine variety determines a canonical affine scheme.

In the general case, a variety has an open cover by affine varieties, and gluing

can be performed in both the category of varieties, as well as the categories of

73

Page 74: Geir Ellingsrud and John Christian Ottem October 15, 2018€¦ · between algebra and geometry: a ne varieties correspond to nitely generated k-algebras without zero-divisors, and

3.8. The functor from varieties to schemes

schemes, and it is a matter of checking that this gives a well-defined scheme (see

Chapter 3).

This assignment works also for morphisms; a morphism of varieties φ : V → W

is determined by a ring homomorphism A(W ) → A(V ), and so taking Spec

we obtain a morphism φsh : V sh → W sh, which extends φ. It follows that the

assignment V 7→ V sh in fact determines a functor t : Var/k → Sch/k. It is a

theorem (Hartshorne Chapter II), that this functor is fully faithful, in the sense

that

HomVar/k(V,W ) = HomSch/k(Vsh,W sh)

is bijective. So two varieties give rise to the same scheme if and only if they

are isomorphic by a unique isomorphism. In particular, this tells us that the

category of varieties Var/k is equivalent to a full subcategory of Sch/k.

74

Page 75: Geir Ellingsrud and John Christian Ottem October 15, 2018€¦ · between algebra and geometry: a ne varieties correspond to nitely generated k-algebras without zero-divisors, and

Chapter 4

More examples

4.1 A scheme that is not affine

Consider A = k[u, v] and let X = SpecA. Let U = X − V (u, v); this is the

affine plane A2k minus the origin. We claim that U is not an affine scheme. The

point of this example is that the restriction map Γ(X,OX) → Γ(U,OU) is an

isomorphism. This shows that U can not be an affine scheme, since in that

case, by Theorem 3.8 above, the inclusion map would be an isomorphism which

obviously is not true, since it is not surjective.

Let us check that that restriction map really is an isomorphism. The two

distinguished open sets D(u) and D(v) form an open covering of U . The proof

is based on the sequence from before associated to that covering:

0 // Γ(U,OU) // Av × Auρ // Auv

A

i#

OO

where ρ is the difference between the two localization maps; that is; it maps

a pair (av−m, bu−n) to av−m − bu−n. We have included the restriction map i#

in the diagram, which is the just the action of the inclusion map i : U → X on

global sections of the structure sheaves. It sends an element a ∈ A to the pair

(a, a).

If the pair (av−m, bu−n) lies in kernel of ρ, we have the relation

aun = bvm,

75

Page 76: Geir Ellingsrud and John Christian Ottem October 15, 2018€¦ · between algebra and geometry: a ne varieties correspond to nitely generated k-algebras without zero-divisors, and

4.2. The projective line

which (since A is a UFD) implies that there is an element c ∈ A with a = cvm

and b = cun; that is, av−m = bu−n. This shows that i# is surjective, and hence

it is an isomorphism since it obviously is injective.

Note that the same example works for any is an integral domain A with an

ideal a generated by two elements u and v having the following property: If

there is a relation aun = bvm with a, b ∈ A and n,m ∈ N , one may find a c ∈ Asuch that a = cvm and b = cun.

4.2 The projective line

In elementary courses on complex function theory one learns about the Riemann

sphere. That is the complex plane with one point added, the point at infinity.

If z is the complex coordinate centered at the origin, the inverse 1/z is the

coordinate centered at infinity. Another name for the Riemann sphere is the

complex projective line, denoted P1C.

The construction of P1C can be vastly generalized, and works in fact over any

ring A. Let u be a variable (“the coordinate function at the origin”). and let

U1 = SpecA[u]. The inverse u−1 is a variable as good as u (“the coordinate at

infinity”), and we let U2 = SpecA[u−1]. Both are copies of the affine line A1A

over A.

Inside U1 we have the open set U12 = D(u) which is canonically equal to

the prime spectrum SpecA[u, u−1], the isomorphism coming from the inclusion

A[u] ⊆ A[u, u−1]. In the same way, inside U2 there is the open set U21 = D(u−1).

This is also canonical isomorphic to the spectrum SpecA[u−1, u], the isomorphism

being induced by the inclusion A[u−1] ⊆ A[u−1, u]. Hence U12 and U21 are

isomorphic schemes (even canonically), and we may glue U1 to U2 along U12.

The result is called the projective line over A and is denoted by P1A.

Gluing two affine lines to get P1

In this example, we will carry out our first computations with sheaf cohomology

using the Cech complex:

Proposition 4.1. We have Γ(P1,OP1A

) = A.

76

Page 77: Geir Ellingsrud and John Christian Ottem October 15, 2018€¦ · between algebra and geometry: a ne varieties correspond to nitely generated k-algebras without zero-divisors, and

Chapter 4. More examples

Proof. The sheaf axiom exact sequence gives us

Γ(P1,OP1A

) // Γ(U1,OP1A

)⊕ Γ(U2,OP1A

) //

'

Γ(U12,OP1A

)

'

A[u]⊕ A[u−1] ρ// A[u, u−1]

,

where the map ρ sends a pair (f(u), g(u−1)) of polynomials with coefficients in

A, one in the variable u and one in u−1, to their difference. We claim that the

kernel of ρ equals A; i.e., the polynomials f and g must both be constants.

So assume that f(u) = g(u−1), and let f(u) = aun + lower terms in u , and

in a similar way, let g(u−1) = bu−m+ lower terms in u−1 , where both a 6= 0 and

b 6= 0, and without loss of generality we may assume that m ≥ n. Now, assume

that m ≥ 1. Upon multiplication by um we obtain b + uh(u) = umf(u) for

some polynomial h(u), and putting u = 0 we get b = 0, which is a contradiction.

Hence m = n = 0 and we are done.

In particular, the global sections of OX of X = P1C is just C. In particular,

knowing Γ(X,OX) does not recover X, in constrast to the case when X = SpecA

is affine.

4.2.1 Affine line a doubled origin

Keeping the notation from the previous example, if we now glue U1 = U2 =

SpecA[v] along U12 = SpecA[u, u−1] using the identity map on the overlap, we

obtain the ‘affine line with two origins’.

This scheme is not affine: Using the sheaf axiom sequence, we find that

Γ(X,OX) = A[u]. Note that both of the points above the origin would correspond

to the same ideal, which is not possible on an affine scheme.

More strikingly, the cokernel of the map A[u] ⊗ A[u−1] → A[u, u−1] is

A[u, u−1]/A[u], and so H1(X,OX) is not finitely generated as an A-module.

We will see that this cannot happen on affine schemes. In fact, for X affine, all

higher cohomology groups vanish: H i(X,OX) = 0 i > 0.

4.3 The blow-up of the affine plane

4.3.1 Blow-up as a variety

Recall the classical construction of a blow-up. There is a rational map f :

A2k 99K P1

k sending a point (x, y) to (x : y) (in homogeneous coordinates on P1k).

77

Page 78: Geir Ellingsrud and John Christian Ottem October 15, 2018€¦ · between algebra and geometry: a ne varieties correspond to nitely generated k-algebras without zero-divisors, and

4.3. The blow-up of the affine plane

This map is not defined at the origin, but we can still associate to it its graph

X ⊆ A2k×P1

k, consisting of all pairs (x, y)× (s : t) so that f(x, y) = (s : t). From

the definition, we see that X is defined by a single equation in A2k × P1

k, namely

X = Z(xt− ys) ⊆ A2k × P1

k

We also have two projection maps p : X → A2 and q : X → P1. The situation is

depicted in Figure 4.1

Figure 4.1: The blow-up of the plane at a point

Let us analyze the fibers of these maps.

The fibers of p are easy to describe. If (x, y) ⊆ A2k is not the origin, then

p−1(x, y) consists of a single point; the equation xt = ys allows us to determine

the point (s : t) uniquely. However, when (x, y) = (0, 0), any s and t satisfy

the equation, so p−1(0, 0) = (0, 0)× P1. In particular, that this inverse image is

1-dimensional - it is called the exceptional divisor of X, and is usually denoted

by E.

Similarly, if (s : t) ∈ P1k is a point, the the fiber q−1(s : t) = (x, y) × (s :

t)|xt = ys ⊆ A2k × (s : t) is a line in A2. The map q is an example of a line

bundle; all of its fibers are A1k’s. We will see this later on in the book.

The standard covering of P1k as a union of two A1

k’s gives an affine cover of X:

If U ⊆ P1k is the open set where s 6= 0, we can normalize so that s = 1, and solve

the equation xt = sy for y. Hence x and t give coordinates on q−1(U) ' A2x,t.

In these coordinates, the morphism p : X → A2k, restricts to a map A2

x,t → A2x,y

given by (x, t) 7→ (x, xt). Similarly, q−1(V ) = A2y,s and the map p is given here

as (y, s) 7→ (sy, y).

4.3.2 The blow-up as a scheme

From the above discussion, we can define the scheme-analogue of the blow-up

of A2k at a point. We will define this as a scheme over Z, rather than over a

78

Page 79: Geir Ellingsrud and John Christian Ottem October 15, 2018€¦ · between algebra and geometry: a ne varieties correspond to nitely generated k-algebras without zero-divisors, and

Chapter 4. More examples

field k. Also, in addition to the scheme X, we also want a morphism of schemes

p : X → A2Z having similar properties to that above.

Consider the affine plane A2 = SpecZ[x, y]. The prime ideal p = (x, y) ⊆Z[x, y] corresponds to the origin in A2. Consider the diagram

Z[x, y]

Z[x, t] Z[y, s]

R = Z[x, y, s, t]/(xt− y, st− 1)

Here the diagonal maps on the top are given by (x, y) 7→ (x, xt) and (x, y) 7→(ys, y) respectively.

Note that the ring R is isomorphic to Z[x, s, t]/(st− 1) = Z[x, s, s−1], as well

as Z[y, s, t]/(st− 1) = Z[y, s, s−1]. Since this is a localization of both Z[x, t] and

Z[y, s] we can identity its spectrum as an open set of SpecZ[x, t] and SpecZ[y, s]

respectively. This gives a digram

SpecZ[x, y]

U = SpecZ[x, t] V = SpecZ[y, s]

SpecR

Hence we can glue these two affine spaces along SpecR to a new scheme

X. By construction, the maps SpecZ[x, t] → SpecZ[x, y] and SpecZ[y, s] →SpecZ[x, y] coincide with the map SpecR → SpecZ[x, y] which is induced by

Z[x, y]→ R. Hence they glue to a morphism

p : X → A2Z.

To complete the discussion, we should define the corresponding morphism

q : X → P1. Again we work locally. On the affine open U = SpecZ[x, t], we have

a map U 7→ A1 = SpecZ[t] induced by the inclusion Z[t] ⊆ Z[x, t]. Similarly, on

V = SpecZ[y, s], we have a map V 7→ A1 = SpecZ[s]. To see if we can glue, we

have to see what happens on the overlap U ∩ V = SpecR. However, on SpecR,

t = s−1, so using the standard gluing on P1, we see that the maps Z[t]→ R and

Z[s]→ R induce the desired morphism q : X → P1.

79

Page 80: Geir Ellingsrud and John Christian Ottem October 15, 2018€¦ · between algebra and geometry: a ne varieties correspond to nitely generated k-algebras without zero-divisors, and

4.4. Hyperelliptic curves

4.4 Hyperelliptic curves

Let k be a field and consider the scheme X glued together by the affine schemes

U = SpecA and V = SpecB, where

A =k[x, y]

(−y2 + a2g+1x2g+1 + · · ·+ a1x)and B =

k[u, v]

(−v2 + a2g+1u+ · · ·+ a1u2g+1)

Here we glue D(x) = SpecA(x) to D(u) = SpecB(u) using the identifications

u = x−1 and v = x−g−1y. These curves are so called hyperelliptic curves or

double covers’ of P1. In the case g = 1 they are called elliptic curves. Here is an

illustration of one of the affine charts:

The term ‘double cover’ comes from the fact that it admits a map X → P1

coming from the inclusions k[x] ⊆ A and k[u] ⊆ B; note that u = x−1 gives the

standard gluing of P1 by the two affine schemes Spec k[x] and Spec k[u]. The

corresponding morphism X → P1 is finite of degree 2, as we are adjoining a

square root of an element of k[x].

80

Page 81: Geir Ellingsrud and John Christian Ottem October 15, 2018€¦ · between algebra and geometry: a ne varieties correspond to nitely generated k-algebras without zero-divisors, and

Chapter 5

Geometric properties of schemes

5.1 Some topological properties

Many basic geometric concepts of schemes are formulated just in terms of their

underlying topological space. For instance, a scheme (X,OX) is said to be

connected if the topological space of X connected (ie it cannot be written as a

union of two proper open sets.)

Example 5.1. If R = Z×Z. Then SpecR is not connected. Indeed, the proper

prime ideals of R are of the form p× Z or Z× p, for p ⊆ Z a prime ideal. And

so SpecR = SpecZ ∪ Spec(Z).

More generally, SpecA is disconnected if and only if A ' B × C for some

non-zero rings B,C.

Likewise, we say that X is irreducible (irredusibelt) if its space is not the

union of two proper, closed subsets. That is, if X = Z ∪ Z ′ with Z and Z ′ both

closed one has either Z = X or Z ′ = X. An equivalent formulation is that any

two nonempty open subsets of X has a non empty intersection; indeed if U and

U ′ are open, it holds that U ∩ U ′ = ∅ is equivalent to U c ∪ (U ′)c = X.

From the the theory of varieties we know that the coordinate ring of a variety

(which is assumed to be irreducible), is an integral domain, and very simple

examples illustrate that reducibility is strongly connected to zero divisors in ring

of functions:

Example 5.2. The scheme X = Spec k[x, y]/(xy) is the prime example of

something that is not irreducible. The coordinate functions x and y are zero-

divisors in the ring k[x, y]/(xy), and their zero-sets V (x) and V (y) show that X

has two components.

81

Page 82: Geir Ellingsrud and John Christian Ottem October 15, 2018€¦ · between algebra and geometry: a ne varieties correspond to nitely generated k-algebras without zero-divisors, and

5.2. Properties of the scheme structure

We have seen that the irreducible subsets of an affine scheme SpecA cor-

respond exactly to the prime ideals p in A. Specializing to the irreducible

components of X, we find:

Proposition 5.3. Let X = SpecA be an affine scheme.

1. The irreducible components of X are precisely the closed subsets of the

form V (p) where p ⊆ A is a minimal prime.

2. X is irreducible if and only if A has a unique minimal prime, if and only

if√

0 is prime.

Note that the ring R = k[t]/(t2) has only one prime ideal (t), so X = SpecR

is just point, hence irreducible. However k[t]/(t2) is not an integral domain.

5.2 Properties of the scheme structure

A ring A is said to be reduced if it has no nilpotent elements. A scheme (X,OX)

is said to be reduced if for every x ∈ X, the local ring OX,x is reduced. This

condition holds if and only if for every open U ⊆ X, the ring OX(U) has no

nilpotents. Indeed, suppose (X,OX) is reduced. Clearly, then, for every x ∈ X,

the local ring OX,x has no nilpotents via the description of its elements as germs

of OX around x. The converse follows similarly: a nilpotent section of OX(U)

would pass to a nilpotent element in the local ring at any point x ∈ U .

Given (X,OX), we can define a new ringed space (X, (OX)red) whose under-

lying topological space is the same as that of X, but the structure sheaf of OXred

is defined as the sheafification of the presheaf

F(U) = OX(U)red

where OX(U)red = OX(U)/N (U) is the quotient of OX(U) by its nilradical N (U)

of nilpotent elements. The collection N (U) forms the so-called nilradical sheaf

N . You should convince yourself that N is indeed a sheaf.

Lemma 5.4. (X,OXred) is a scheme.

Proof. We find that OXred,x = (OX,x)red = OX,x/√

0 is the quotient of a local ring

and hence local. It follows that (X, (OX)red) is a locally ringed space. Further-

more, if (X,OX) is locally isomorphic to affine schemes (SpecAi,OSpecAi), and

then (X, (OX)red) is locally isomorphic to affine schemes (SpecAi/Ni,OSpecAi/Ni),

where Ni is the nilradical of Ai. Hence (X, (OX)red) is a scheme.

We refer to (X,OXred) as simply Xred, the reduced scheme associated to X.

Xred comes with a canonical morphism of schemes r : Xred → X defined as

82

Page 83: Geir Ellingsrud and John Christian Ottem October 15, 2018€¦ · between algebra and geometry: a ne varieties correspond to nitely generated k-algebras without zero-divisors, and

Chapter 5. Geometric properties of schemes

follows. The underlying topological space of Xred is that of X, and hence there is a

homeomorphism r : Xred → X. There is also a map of sheaves r# : OX → r∗OXred

that on each open U ⊆ X is the quotient map OX(U) → (OX)red(U). As the

induced map on stalks is a quotient map as well, it is a local homomorphism,

and we obtain a morphism (r, r#) : Xred → X of schemes.

The scheme Xred and the morphism r satisfy the following universal property:

Proposition 5.5. Let f : Y → X be a morphism of schemes, with Y reduced.

Then f factors uniquely through the natural map r : Xred → X, i.e. there exists

a unique g : Y → Xred such that f = r g.

Proof. The only difference between X and Xred is the structure sheaf, so define

g on the level of topological spaces by f . On the level of sheaves we find that

(over any open U ⊆ X) the map f# : OX(U)→ f∗OY (U) takes all nilpotents to

zero as Y is reduced. By the universal property of quotients there must exist a

unique morphism of rings g#(U) : OXred(U)→ f∗OX(U) such that f# = g# r#.

This gives the required morphism (g, g#) of schemes.

5.2.1 Integral schemes

Definition 5.6. A scheme is integral if it is both irreducible and reduced.

In particular, an affine scheme X = SpecA is integral if and only if A is an

integral domain. Indeed, X is reduced if and only if A has no nilpotents. X

is irreducible if and only if the nilradical of A N = ∩pp is prime. These two

statements imply that the zero-ideal (=∩p) is prime, and so A is an integral

domain.

Moreover, it is not so hard to prove the following:

Proposition 5.7. A scheme X is integral if and only if OX(U) is an integral

domain for each open U ⊆ X.

If X is an integral scheme, there is a unique generic point η of X which

is dense in X. Indeed, take any U = SpecA ⊆ X and let η be the point

corresponding to the zero ideal (which is prime). The local ring OX,η is equal to

the field of fractions of A. We define the function field K(X) of X as the field of

fractions of OX,η at the generic point.

Example 5.8. The function field of Ank = Spec k[x1, . . . , xn] is k(x1, . . . , xn).

Example 5.9. The function field of SpecZ is Q.

83

Page 84: Geir Ellingsrud and John Christian Ottem October 15, 2018€¦ · between algebra and geometry: a ne varieties correspond to nitely generated k-algebras without zero-divisors, and

5.2. Properties of the scheme structure

5.2.2 Noetherian spaces and quasi-compactness

Definition 5.10. A topological space is called noetherian (noetersk) if every

descending chain of closed subsets is eventually constant; that is, if Zi i∈N is a

family of closed subsets Zi with Zi+1 ⊆ Zi, there is an i0 such that Zi = Zi+1 for

i ≥ i0.

An equivalent formulation is any any non-empty family of closed subsets has

a minimal element. This happens if and only if every non-empty family of open

subsets has a maximal element.

Clearly if A is a noetherian ring, the prime spectrum SpecA is a noetherian

topological space. The converse however, is not true. A simple example is the

following: Take the polynomial ring k[t1, t2, . . . ] in countably many variables tiand mod out by the square m2 of the maximal ideal generated by the variables,

i.e., with m = (t1, t2, . . . ). The resulting ring A has just one prime ideal, the

one generated by the ti’s, so SpecA has just one point. The ring A, however, is

clearly not noetherian; the sole prime ideal requiring infinitely many generators,

namely all the ti’s.

Given this example, we take a different route to define noetherianness for

schemes:

Definition 5.11. (i) A scheme is quasi-compact if every open cover of X has

a finite subcover.

(ii) A scheme is locally noetherian if it can be covered by open affine subsets

SpecAi where each Ai is a noetherian ring

(iii) A scheme is noetherian if it is both locally noetherian and quasi-compact.

An affine scheme X = SpecA is always quasi-compact: If X =⋃

SpecAi,

each SpecAi can be covered by the distinguished opens D(fij), and X is covered

by such subsets if and only if 1 belongs to the ideal generated by the fij; and in

this case finitely many of them will do. Hence a general scheme is noetherian if

and only if it can be covered by finitely many SpecAi where each Ai is noetherian.

In fact, now we have

Proposition 5.12. SpecA is noetherian (as a scheme) if and only if A is

noetherian.

This is really a purely algebraic fact: Refining the cover, we may assume that

each Ai = Afi . By a theorem in commutative algebra, A is noetherian provided

that each Afi is noetherian and 1 ∈ (f1, . . . , fr).

Proposition 5.13. If X is a noetherian scheme, then its underlying topological

space is noetherian.

84

Page 85: Geir Ellingsrud and John Christian Ottem October 15, 2018€¦ · between algebra and geometry: a ne varieties correspond to nitely generated k-algebras without zero-divisors, and

Chapter 5. Geometric properties of schemes

By covering X with open affine subsets, it suffices to show the proposition for

X = SpecA, for A a noetherian ring. In ths case a descending chain of closed

subsets is of the form V (I1) ⊇ V (I2) ⊇ · · · , where we may assume that the

ideals In are radical. Then the condition that V (In) is decreasing, corresponds

the sequence (In) being increasing, and so it has to be stationary, because A is

noetherian.

Proposition 5.14. Let X be a (locally) noetherian scheme. Then any open

subscheme of X is also (locally) noetherian.

This is clear for the locally noetherian case. For the noetherianness, we need

to show that U is also quasi-compact. Since X is noetherian, U is noetherian as

a topological space. Now if (Vi)i∈I is a cover of U , then the collection of unions

of the form ∪i∈FVi for F finite, has a maximal element, which gives a finite

subcover of U .

We also see that if X is locally noetherian, then all the local rings OX,x are

noetherian. The converse however, does not hold:

Example 5.15. Let A = C[y, x1, x2, . . . ]/I where I is the ideal generated by

xi−(y−i)xi+1, x2i for i = 1, 2, . . . . Then A is not noetherian, but Ap is noetherian

for every prime ideal p. The nilradical N is (x1, x2, . . . ) – the xi are nilpotent,

and A/(x1, . . . ) is a domain. In particular, A is not noetherian. We have

A/N ' C[y], so the maximal ideals of A are of the form m = (y − a, x1, x2, . . . )

for some a ∈ C. If a is not a positive integer, then Am is the localization of a ring

generated by y and x1, and hence is noetherian. If a = n is a positive integer,

we have xi+1 = xi/(y − i) for i 6= n in m. Hence Am is the localization of a ring

generated by y, xn, and hence is Noetherian.

Proposition 5.16. Let X be a noetherian scheme. Then any closed subset

Y ⊆ X can be written as a finite union Y =⋃Yi where each Yi is irreducible.

Proof. If W is the set of closed subsets that cannot be written as a finite union

of closed irreducible subsets, and W is non-empty, it has a minimal element

Z ∈ W . Z is not irreducible, so it can be written as a union of two proper closed

subsets which do not lie in W , a contradiction.

The Yi are called the irreducible components of Y . They are a priori defined

only at the level of topological spaces, but we will see in Chapter 7 (Lemma

8.37) that one can define a scheme structure on them in a natural way, so that

they become ‘closed subschemes’ of X.

85

Page 86: Geir Ellingsrud and John Christian Ottem October 15, 2018€¦ · between algebra and geometry: a ne varieties correspond to nitely generated k-algebras without zero-divisors, and

5.3. The dimension of a scheme

5.3 The dimension of a scheme

Recall that the Krull dimension of a ring A is defined as the suprenum of the

length of all chains of prime ideals in A. For a scheme, we make the following

similar definition:

Definition 5.17. Let X be a scheme. The dimension of X is the suprenum of

all integers n such that there exists a chain

Y0 ⊆ Y1 ⊆ · · · ⊆ Yn

of distinct closed irreducible closed subsets of X.

Note that this suprenum might not be a finite number, in which case we

say that dimX =∞. Note also that the dimension of X only depends on the

underlying topological space.

In the case where X = SpecA is affine, we know that the closed irreducible

subsets are of the form V (p) where p is a prime ideal. Using this observation we

find

Proposition 5.18. The dimension of X = SpecA equals the Krull dimension

of A.

Example 5.19. 1. The dimension of AnA = SpecA[x1, . . . , xn] is n + dimA.

In particular, when A = k is a field, Ank has dimension n.

2. dim SpecZ is 1.

3. dim Spec k[ε]/ε2 = 0.

Remark 5.20. Having finite dimension does not guarantee that the scheme is

noetherian. For an explicit counterexample, take SpecR where R is the ring of

integers in Qp.

More surprisingly, there might may be noetherian rings having infinite Krull

dimension! First counterexamples were constructed by Nagata. In short the

counterexample involves taking the polynomial ring in infinitely many variables

k[x1, x2, . . . ] and dividing out by a certain ideal, ensuring that all prime chains

are finite, but there is a sequence of prime chains whose lengths tend to infinity!

Definition 5.21. Let Y ⊆ X be a closed subset of X. We define the codimension

of Y as the suprenum of all integers n such that there exists a chain

Y = Y0 ⊆ Y1 ⊆ · · ·Yn

of distinct irreducible closed subsets of X.

86

Page 87: Geir Ellingsrud and John Christian Ottem October 15, 2018€¦ · between algebra and geometry: a ne varieties correspond to nitely generated k-algebras without zero-divisors, and

Chapter 5. Geometric properties of schemes

The codimension of V (p) in SpecA is the height of the prime p in A.

Note that the formula dim Y + codimY = dimX does not always hold, even

if X is the spectrum of a ring. Indeed, let X = SpecR[t] where R = k[[t]]. The

prime p = (tu− 1) has height 1, but A/p ' R[1/t] is a field, hence of dimension

0. However, dimA = dimR + 1 = 2.

On the other hand, for integral schemes, we can say something:

Theorem 5.22. Let X be an integral scheme of finite type over a field k, with

function field K. Then

(i) dimX equals the trancendence degree of K over k (in particular, dimX <

∞).

(ii) For each U ⊆ X open, dim U = dimX.

(iii) For p ∈ X closed, dimX = dim OX,p.

Proof. Note that (ii) follows from (i) since X and an open subset U have the

same function field. To prove (i) we may assume that X = SpecA is affine. The

hypothesis on X gives that A is a finitely generated k-algebra with quotient field

K. In this case, (i) is a consequence of Noether normalization (see Chapter 12

of Atiyah–MacDonald). To prove (iii), we may again assume that X = SpecA,

and use the formula

dimA/p + ht p = dimA

which folds for prime ideals in finitely generated k-algebras.

For schemes which aren’t integral but still of finite type, we still have a good

control over its dimension. First of all, the dimension of X is the same as Xred,

so we may assume that X is reduced. Then if X =⋃Xi is the decomposition

into irreducible components, we have Xi integral, and dimX is the suprenum of

all dimXi.

5.4 More examples

The rings Z(2) and Z(3) are both discrete valuation rings with maximal ideal

being (2) and (3) respectively. Their fraction field are both equal to Q. Let

X1 = SpecZ(2) and A2 = SpecZ(3). Both have a generic point that is open, so

there is a canonical open immersion SpecQ → Xi for i = 1, 2. Hence we can

glue. We obtain a scheme X with one open point η and two closed points. Let

us compute the global sections of OX using the now classical restriction sequence

87

Page 88: Geir Ellingsrud and John Christian Ottem October 15, 2018€¦ · between algebra and geometry: a ne varieties correspond to nitely generated k-algebras without zero-divisors, and

5.4. More examples

for the open covering X1, X2 :

Γ(X,OX) // Γ(X1,OX)⊕ Γ(X2,OX) //

=

Γ(X1 ∩X2,OX)

=

Z(2) ⊕ Z(3) ρ

// Q

.

The map ρ sends a pair (an−1, bm−1) to the difference an−1 − bm−1, hence the

kernel consists of the diagonal, so to speak, in Z(2) ⊕ Z(3), which is isomorphic

to the intersection Z(2) ∩ Z(3). This is a semi local ring with the two maximal

ideals (2) and (3). Hence there is a map X → SpecZ(2) ∩ Z(3) and it is left as

an exercise to show this is an isomorphism.

More generally, if P = p1, . . . , pr is a finite set different prime numbers,

one may let Xp = SpecZ(p) for p ∈ P . There is, as in the previous case, canonical

open embedding SpecQ → Xp. Let the image be ηp . Obviously the gluing

conditions are all satisfied (the transition maps are all equal to idSpecQ and

Xpq = ηp for all p). We do the gluing and obtain a scheme X. Again, to

compute the global sections, we use the sequence

Γ(X,OX) //⊕

p∈P Γ(Xp,OX) //

=

⊕p,q∈P Γ(X1 ∩X2,OX)

=

⊕p∈P Z(p) ρ

//⊕

p,q∈P Q.

The map ρ sends the sequence (ap)p∈P to the sequence (ap − aq)p,q∈P and the

kernel of ρ is the intersection AP =⋂p∈P Z(p), which is a semilocal ring whose

maximal ideals are the (p)Ap’s for p ∈ P . There is a canonical morphism

X → SpecAp, and again we leave it to the industrious student to verify that

this is an isomorphism.

5.4.1 A more fancy example

In the previous example we worked with a finite set of primes, but the hypothesises

of the gluing theorem impose no restrictions on the number of schemes to be

glued together, and we are free to take P infinite, for example we can use the set

P of all primes! The glued scheme XP is a peculiar animal: It is neither affine

nor noetherian, but it is locally noetherian. There is a map φ : XP → SpecZwhich is bijective and continuous, but not a homeomorphism, and it has the

property that for all open subsets U ⊆ SpecZ the map induced on sections

φ# : Γ(U,OSpecZ)→ Γ(φ−1U,OXP ) is an isomorphism!

As before we construct the scheme XP by gluing the different SpecZ(p)’s

together along the generic points. However, when computing the global sections,

88

Page 89: Geir Ellingsrud and John Christian Ottem October 15, 2018€¦ · between algebra and geometry: a ne varieties correspond to nitely generated k-algebras without zero-divisors, and

Chapter 5. Geometric properties of schemes

we see things change. The kernel of ρ is still⋂p∈P Z(p), but now this intersection

equals Z! Indeed, a rational number α = a/b lies in Z(p) precisely when the

denominator b does not have p as factor, so lying in all Z(p), means that b has

no non-trivial prime-factor. That is, b = ±1, and α ∈ Z.

There is a morphism XP → SpecZ which one may think about as follows.

Each of the schemes SpecZ(p) maps in a natural way into SpecZ, the mapping

being induced by the inclusions Z ⊆ Z(p). The generic points of the SpecZp’sare all being mapped to the generic point of SpecZ. Hence they patch together

to give a map XP → SpecZ. This is a continuous bijection by construction, but

it is not a homeomorphism! Indeed, the subsets SpecZ(p) are open in XP by the

gluing construction, but they are not open in SpecZ, since their complements

are infinite, and the closed sets in SpecZ are just the finite sets of maximal

ideals.

The topology of the scheme XP is not noetherian since the subschemes

SpecZ(p) form an open cover that obviously can not be reduced to a finite

cover. However, it is locally noetherian the open subschemes SpecZ(p) being

noetherian. The sets Up = XP − (p) map bijectively to D(p) ⊆ SpecZ and

Γ(Up,OXP ) = Zp, but Up and D(p) are not isomorphic.

Exercise 16. Glue X2 to itself along the generic point to obtain a scheme X.

Show that X is not affine. Hint: Show that Γ(X,OX) = Z(2).

89

Page 90: Geir Ellingsrud and John Christian Ottem October 15, 2018€¦ · between algebra and geometry: a ne varieties correspond to nitely generated k-algebras without zero-divisors, and
Page 91: Geir Ellingsrud and John Christian Ottem October 15, 2018€¦ · between algebra and geometry: a ne varieties correspond to nitely generated k-algebras without zero-divisors, and

Chapter 6

Projective schemes

6.1 Motivation

Let us consider the usual construction of complex projective space: As a topo-

logical space, CPn is the quotient space

CPn =(Cn+1 − 0

)/C∗

where C∗ acts on Cn+1 by scaling the coordinates. We can translate this into

algebra as follows: If f is a function on Cn+1 and λ ∈ C∗ then we get a (right)

action of C∗ on f by defining f.λ(x) = f(λx). In other words, C∗ acts on the

ring C[x0, . . . , xn]. Now, the actions C∗ are well-understood: Any time C∗ acts

on a complex vector space V , we can decompose that vector space as a direct

sum

V =∑d∈Z

Vd

where λ ∈ C∗ acts on v ∈ Vd by λ.v = λdv. Thus we obtain a grading on

C[x0, . . . , xn]. More generally, affine schemes over C with an action of C∗correspond to graded C-algebras.

We want to take the quotient of An+1C −0 = SpecC[x0, . . . , xn]−V (x0, . . . , xn)

by this action. We write PnC for the corresponding quotient space. The notation

PnC, rather than CPn, is used to emphasise that the quotient is taken with the

Zariski topology, and not the usual topology.

We can try to put a scheme structure on PnC by looking for reasonable open

covers. Note that the open subsets of PnC correspond to C∗-invariant open subsets

of An+1C − 0. If is not too hard to see that D(f) ⊆ An+1

C is C∗-invariant if and

91

Page 92: Geir Ellingsrud and John Christian Ottem October 15, 2018€¦ · between algebra and geometry: a ne varieties correspond to nitely generated k-algebras without zero-divisors, and

6.2. Basic remarks on graded rings

only if f is a homogeneous polynomial. We write D+(f) ⊆ PnC for the open

subset corresponding to D(f) ⊆ An+1C − 0.

To define a structure sheaf we should figure out what OPnC(D+(f)) should

be. While it is true that D(f), being an affine scheme, has a structure sheaf,

we have to take more care in deciding which sections to take, to make things

compatible with the C∗-action: A function on D+(f) should be a function on

D(f) that is invariant under the action of C∗. That is, we should have g.λ = g,

which means precisely that g has graded degree zero. The functions of with this

property are exactly the ones of graded degree zero, thus we define

OP (D+(f)) = C[x0, . . . , xn, f−1]0.

where the degree means that we take the degree 0 part.

We can generalize the above for any C-scheme with an action of C∗. Such a

scheme corresponds to a graded C-algebra R. To make a reasonable quotient,

we must remove the locus in SpecR that is fixed by C∗. It is not too hard to

prove the following:

Lemma 6.1. The fixed locus of C∗ acting on SpecR is V (R+), where R+ denotes

the ideal generated by element of positive degree.

So we then attempt to take a quotient of SpecR− V (R+) by C∗. Again, the

C∗-invariant open subsets are of the form D(f) where f is homogeneous, and

these correspond to open subsets D+(f) ⊆ P = (SpecR− V (R+)) /C∗. We can

then define a B-sheaf on P by setting OP (D+(f)) = OSpecS(D(f))0, and check

that we get a scheme P .

Beside of inducing a grading on S, the action of C∗ plays very little role here.

Realizing this, we can in fact build a scheme P from any graded ring R: We

construct the topological space of P from the set of homogeneous prime ideals of

R (with the induced Zariski topology), and define a structure sheaf on it by the

formula like to the one above. This is essentially the ‘Proj’-construction.

6.2 Basic remarks on graded rings

A graded ring R is a ring with a decomposition

R = R0 ⊕R1 ⊕ · · ·

such that Rm · Rn ⊆ Rm+n for each m,n ≥ 0, and such that R0 is a subring

of R. We say that an R-module M is graded if it has a similar decomposition

M =⊕

n≥0Mn such that Rm ·Mn ⊆Mm+n for each m,n ≥ 0. As usual we say

that an element f ∈ R (resp. m ∈M) is homogeneous of degree d if it lies in Rd

92

Page 93: Geir Ellingsrud and John Christian Ottem October 15, 2018€¦ · between algebra and geometry: a ne varieties correspond to nitely generated k-algebras without zero-divisors, and

Chapter 6. Projective schemes

(resp. m ∈Md). An ideal I ⊆ R is homogeneous if the homogeneous components

of any element in I belongs to I.

We will write R+ for the sum⊕

n>0Rn; this is naturally a homogeneous ideal

of R, which we call the irrelevant ideal.

We let R(d) denote the subring of R given by⊕

n≥0Rnd

6.2.1 Localization

Given a homogeneous prime ideal p let T (p) be the set of homogeneous elements

f ∈ R−p. We obtain a grading on the localization T (p)−1R by setting deg(f/g) =

deg f − deg g.

Definition 6.2. For a homogeneous prime ideal p ⊆ R and f ∈ R homogeneous,

define for an R-module M

• M(p) = (T (p)−1M)0

• M(f) = (1, f, f 2, . . . , −1M)0

where the subscript means the degree 0 part.

Example 6.3. For R = A[x0, . . . , xn], we have R(xi) = A[x−1i xj]0≤j≤n.

6.3 The Proj construction

Definition 6.4. Let ProjR denote the set of homogeneous prime ideals of R

that do not contain the irrelevant ideal R+.

We can put a topology on ProjR by setting, for a homogeneous ideal b,

V (b) = p ∈ ProjR : p ⊇ b.

These sets satisfy the following identities

1. V (∑

bi) =⋂V (bi).

2. V (ab) = V (a) ∪ V (b).

3. V (√a) = V (a).

In particular, the V ’s do in fact yield a topology on ProjR (setting the open

sets to be complements of the V ’s). Notice that this is nothing by the induced

topology from the inclusion ProjR ⊆ SpecR.

As with the affine case, we can define the distinguished open sets. For f

homogeneous of positive degree, define D+(f) to be the collection of homogeneous

93

Page 94: Geir Ellingsrud and John Christian Ottem October 15, 2018€¦ · between algebra and geometry: a ne varieties correspond to nitely generated k-algebras without zero-divisors, and

6.3. The Proj construction

ideals (not containing R+) that do not contain f . These are open sets with

respect to the Zariski topology.

Let a be a homogeneous ideal. Then we claim that:

Lemma 6.5. V (a) = V (a ∩R+).

Proof. Indeed, suppose p is a homogeneous prime not containing R+ such that

all homogeneous elements of positive degree in a (i.e., anything in a∩R+) belongs

to p. We will show that a ⊆ p.

Choose a ∈ a ∩R0. It is sufficient to show that any such a belongs to p since

we are working with homogeneous ideals. Let f be a homogeneous element of

positive degree that is not in p. Then af ∈ a ∩ R+, so af ∈ p. But f /∈ p, so

a ∈ p.

Thus, when constructing these closed sets V (a), it suffices to work with ideals

contained in the irrelevant ideal. In fact, we could take a in any prescribed power

of the irrelevant ideal, since taking radicals does not affect V . Note incidentally

that we would not get any more closed sets if we allowed all ideals b, since to

any b we can consider its “homogenization.”

Proposition 6.6. We have D+(f)∩D+(g) = D+(fg). Also, the D+(f) form a

basis for the topology on ProjR.

Proof. The first part is evident, by the definition of a prime ideal. We prove the

second. Note that V (a) is the intersection of the V ((f)) for the homogeneous

f ∈ a ∩R+. Thus ProjR− V (a) is the union of these D+(f). So every open set

is a union of sets of the form D+(f).

Proposition 6.7. Let f ∈ R be homogeneous of degree d. There is a canonical

homeomorphism φf : D+(f)→ SpecR(f) given by

φf : p→ pRf ∩R(f)

sending homogeneous prime ideals of R not containing f into primes of R(f).

Moreover,

(i) For g ∈ R such that D+(g) ⊆ D+(f). If u = gdf− deg g ∈ R(f), then

φ(D+(g)) = D(u).

(ii) For any graded R-module M , there is a canonical homomorphism M(f) →M(g), which induces an isomorphism (M(f))u 'M(g).

(iii) If I ⊆ R is a homogeneous ideal, then φ(V (I)∩D+(f)) = V (IRf ∩R(f)).

94

Page 95: Geir Ellingsrud and John Christian Ottem October 15, 2018€¦ · between algebra and geometry: a ne varieties correspond to nitely generated k-algebras without zero-divisors, and

Chapter 6. Projective schemes

Proof. Note that φ is just the restriction of the canonical map SpecRf →SpecR(f) induced by the inclusion map R(f) ⊆ Rf . Therefore it is continuous.

We now check that it is a bijection.

We construct the inverse map ψ : Spec(R(f))→ D+(f) as follows. Given a

prime ideal p ∈ Spec(R(f)), define ψ(p) =⊕

n≥0 ψ(p)n where

ψ(p)n = x ∈ Rn|xd/fn ∈ p

(Recall that d = deg f). We need to check that this is a homogeneous prime ideal.

Given x, y ∈ ψ(p)n, we have, by definition, that xd/fn ∈ p and yd/fn ∈ p, and

so by the Binomial theorem, (x+ y)2d/f 2n ∈ p. Therefore, (x+ y)d/fn ∈ p, and

so ψ(p) is closed under addition. The other conditions for being a homogeneous

ideal are proved similarly. To show that ψ(p) is prime, let x ∈ Rm and y ∈ Rn

so that (xy)d/fm+n ∈ p. Then (xy)d ∈ p, and so either x ∈ p or y ∈ p since p is

prime. Hence either xd/fm ∈ ψ(p)m or yd/fn ∈ ψ(p)n, and so ψ(p) is prime.

The fact that φ and ψ are mutually inverse follows almost by construction.

Indeed, to verify ψ(pRf ∩ R(f)) = p, note that x ∈ pn ⊆ Rn satisfies xd/fn ∈pRf ∩R(f). Conversely, if x ∈ ψ(pRf ∩R(f))n, then xd/fn ∈ pRf ∩R(f), and so

there is an N > 0 so that fNxd ∈ p. As p is prime, and f 6∈ p, we have xd ∈ p,

and hence x ∈ p. Proving φ ψ = id is similar. This shows that φ is bijective.

The fact that φ is a homeomorphism follows from the fact that it is continuous,

and the point (i) showing that it is open.

(i) Let g ∈ R be an element with D+(g) ⊆ D(f). Then for p ∈ D+(f), we

have p ∈ D+(g) ⇔ gr/f s 6∈ pRf for some r, s > 0 ⇔ gr/f s 6∈ pRf for all r, s > 0

⇔ gdeg f/fdeg g 6∈ p ∩R(f) = φ(p). Hence φ(D+(g)) = D(u).

(ii) Write gk = af . The localization map Mf → Mg is given by xfm7→ amx

gmk,

where x ∈ M . This induces a map M(f) → M(g) (since deg x + m(deg a −k deg g) = deg x−m deg f). The element u acts as an invertible element on R(g),

so the map M(f) →M(g) factors via a map

ρ : (M(f))u →M(g)

We claim that this is an isomorphism.

ρ surjective: Explicitly, we have

ρ

(xf−n

um

)=f tm−nx

gdm

where t = deg g. Take an element y/gl ∈M(g) where deg y = tl. Choose m large

so that dm ≥ l. Define x = gdm−lyf tm−n

∈ M(f). We have deg x = nd, and hencexf−n

um∈ (M(f))u is an element that maps to the given y

gl.

ρ is injective: If x/fn ∈ M(f) maps to 0 in M(g), then there is an l > 0

95

Page 96: Geir Ellingsrud and John Christian Ottem October 15, 2018€¦ · between algebra and geometry: a ne varieties correspond to nitely generated k-algebras without zero-divisors, and

6.3. The Proj construction

so that gldanx = 0 ∈ M . Multiplying up by powers of a and f , we get a

relation of the form g(l+n)dx = 0 ∈M , and hence u(l+n)dx = 0 ∈M(f). But then

x/fn = 0 ∈ (M(f))u.

(iii) Follows by the definition of φ. Let p ∈ V (I)∩D+(f), so p ⊇ I and p 63 f .

Hence pRf∩R(f) ⊇ IRf∩R(f) which gives the ‘⊆′ containment. Conversely, given

a prime ideal p ⊆ R(f) such that p ⊇ IRf ∩R(f), then its preimage p′ in R is a

homogeneous prime ideal not containing f , and φ(p′) = p′Rf ∩R(f) ⊇ IRf ∩R(f).

6.3.1 ProjR as a scheme

We shall now make X = ProjR into a locally ringed space. Let B be the base

of ProjR made up by the distinguished open subsets. For each D+(f) we let

OX(D+(f)) = R(f)

The previous proposition shows that this gives a well-defined B-presheaf of rings,

and using the homeomorphism φ to SpecR(r), we see that it is a actually a

B-sheaf. (Alternatively, we could modify the proof for the case of Spec to see

this directly). We will denote the unique sheaf extension by OX .

It follows that X is has the structure of a ringed space. This is in fact a locally

ringed space, because the stalk Op is just R(p), which is a local ring. Indeed,

the unique maximal ideal is generated by p. Moreover, the previous discussion

has shown that the basic open sets D+(f) are each isomorphic as locally ringed

spaces to SpecR(f), which are affine schemes, and so ProjR is a scheme.

Definition 6.8. For a graded ring R, we call the scheme (ProjR,O) the projec-

tive spectrum of R.

In fact, ProjR is naturally a scheme over R0: The homomorphisms R0 → R(f)

induce a morphism

ProjR→ Spec(R0)

Moreover, if R is a finitely generated over R0, this is of finite type over SpecR0.

This follows by looking at the distinguished opens, D+(f) - each R(f) is finitely

generated as an R0-algebra if R is.

Example 6.9. Let A be a ring and let R = A[t] with the grading given by

deg t = 1 and deg a = 0 for all a ∈ A. Then ProjR ' SpecA.

Example 6.10. Let us study the case of a polynomial ring in two variables.

R = Z[s, t] where s and t have degree 1, and k is any ring. To see that

X = ProjR coincides with P1 as defined in Chapter 3, we see how it is glued

from two affine schemes. Note that X is covered by D+(s) and D+(t) (since

96

Page 97: Geir Ellingsrud and John Christian Ottem October 15, 2018€¦ · between algebra and geometry: a ne varieties correspond to nitely generated k-algebras without zero-divisors, and

Chapter 6. Projective schemes

these generate the irrelevant ideal). Write for simplicitly U = D+(s) ' R(s)

and V = D+(t) ' R(t). Let us check how X is glued together from U, V along

U ∩ V = D+(st) ' SpecR(st).

Note first that Rs ' Z[s, s−1, t] has degree 0 part equal to Z[s−1t] ⊆ Rs.

Similarly, R(t) = Z[st−1]. The intersection D+(st) is the degree 0 part of Rst, i.e.,

R(st) = Z[s−1t, st−1]. In other words, if we write u = s−1t, we have R(s) = Z[u],

R(t) = Z[u−1] and R(st) = Z[u, u−1] = Z[u]u. It follows that U ' SpecZ[u] = A1Z

and V ' SpecZ[u−1] ' A1Z are glued along SpecR(st) which is an open set in

both. From this we see that ProjR coincides with out previous defintion of P1.

Example 6.11. Let R = k[x, y]/(xy). SpecR is the union of the x- and y-axes.

So SpecR− V (x, y) is the axes excluding the origin. ProjR consists of just two

points: We obtain ProjR by gluing Spec(R(x)) and Spec(R(y)). Now,

R(x) = k[x, y](x)/xy = k[x, y](x)/y = k[x](x) = k

and the corresponding chart of ProjR is Spec k. Similarly, the other chart is

Spec k. We have R(xy) = 0, so the overlap is empty and ProjR is two points.

Definition 6.12. We define the projective n-space to be the scheme

Pn = ProjZ[x0, . . . , xn].

More generally, for a ring A, the projective space over A

PnA = ProjA[x0, . . . , xn]

In this notation, Pnk is a scheme whose closed k-points Pn(k) coincides with

the variety of projective n-space.

Proposition 6.13. Let R be a graded ring.

(i) If R is noetherian, then ProjR is noetherian.

(ii) If R is finitely generated over R0, then ProjR is of finite type over SpecR0.

(iii) If R is an integral domain, then ProjR is integral.

Proof. The properties listed are properties one can verify on an affine covering.

In out case ProjR is covered by the affines SpecR(f) which are noetherian (resp.

of finite type, integral) provided R is noetherian (resp. finitely generated, an

integral domain).

97

Page 98: Geir Ellingsrud and John Christian Ottem October 15, 2018€¦ · between algebra and geometry: a ne varieties correspond to nitely generated k-algebras without zero-divisors, and

6.4. Functoriality

6.4 Functoriality

Unlike the case of affine schemes, a graded ring homomorphism φ : R→ S does

not induce a morphism between the projective spectra ProjS and ProjR. The

reason is that some primes in S may pullback to R to contain the irrelevant ideal

R+. However, as we will see shortly, this is the only obstruction to defining a

morphism.

Given a homomorphism φ : R→ S, we define the set G(φ) ⊆ ProjS to be the

set of prime ideals p such that p ⊆ S that do not contain φ(R+), in particular,

such that φ−1(p) is in ProjR. This sets up a map

ψ : G(φ) → ProjR

p 7→ φ−1(p)

Note that G(φ) is an open set of ProjS, so in particular it has an induced scheme

structure. We claim that the map G(φ)→ ProjR is a morphism of schemes.

First of all, the map ψ is continuous, as the Zariski topologies are induced

from those of SpecS and SpecR. More explicitly, the inverse image φ−1a (D+(f))

is equal to G(φ) ∩D+(φ(f)), which is open.

Write X = G(φ) and Y = ProjR. We now define the map ψ on the level of

sheaves, i.e., a map

ψ# : OY → ψ∗OX

On the distinguished open set D+(f) ⊆ ProjR, we can define this map using the

isomorphism to SpecR(f). We have a map R(f) → Sφ(f) induced by φ. Moreover,

since ψ−1(D+(f)) is open in SpecS(φf), we get the map

OY (D+(f)) = R(f) → OX(ψ−1(D+(f))

by restriction. We have shown the following

Proposition 6.14. Let φ : R → S be a morphism of graded rings. The map

ψ : G(φ)→ ProjR is a morphism of schemes.

Example 6.15. To see why restriction to the open set G(φ) is necessary, we

consider the case where R = k[x0, x1], S = k[x0, x1, x2] and φ is the inclusion

map. Note that the prime ideal a = (x0, x1) defines an element in ProjS, but its

restriction to R is the whole irrelevant ideal of R. In fact, G(φ) = ProjS− V (a),

and the map

ψ : P2k − V (a)→ P1

k

is nothing but the projection from the point [0, 0, 1]. It is a good exercise to

prove that there can be no morphisms Pmk → Pnk for m > n in general.

98

Page 99: Geir Ellingsrud and John Christian Ottem October 15, 2018€¦ · between algebra and geometry: a ne varieties correspond to nitely generated k-algebras without zero-divisors, and

Chapter 6. Projective schemes

6.4.1 Closed immersions

The primary example of the above construction is considering the graded quotient

homomorphism φ : R → R/I, where I ⊆ R is a homogeneous ideal. Here we

have G(φ) = Proj(R/I), so we get a map

ψ : Proj(R/I)→ ProjR.

We claim that this is a closed immersion. As usual, homogeneous primes

in R/I not containing R+ correspond to homogeneous primes in R containing

I but not R+. It follows that ψ is injective, mapping onto the closed subset

V (I) in ProjR. Finally, ψ# is surjective on stalks (where it is just the map

R(p) → (R/I)(p)), and so ψ is a closed immersion. We will see later that there is

a converse to this statement, under some mild assumptions on R.

6.5 The Veronese embedding

Let R be a graded ring and let d > be an integer. The inclusion φ : R(d) → R

induces a morphism

νd : ProjR→ ProjR(d)

Indeed, in this case G(φ) = ProjR, since any prime p such that p ⊇ R+ ∩R(d)

must also contain all of R+ – if r ∈ R+, note that rd ∈ R+ ∩R(d) and so r ∈ p

as well! This map is called the Veronese embedding, or d-uple embedding of X.

Proposition 6.16. The Veronese embedding νd is an isomorphism.

Proof. νd is injective: If p, q ∈ ProjR such that p ∩R(d) = q ∩R(d). Then for a

homogeneous element x ∈ R we have

x ∈ p⇔ xd ∈ p⇔ xd ∈ q⇔ x ∈ q

And so p = q. To show that νd is surjective, let q ∈ ProjR(d), and define the

homogeneous ideal in R by

p =∞⊕n=0

x ∈ Rn|xd ∈ q

.

It is not too hard to check that p is prime, and that p ∩ R(d) = q. Hence

νd is bijective, and even a homeomorphism. νd induces an isomorphism when

restricted to the open affines D+(f) as well, and so we are done.

99

Page 100: Geir Ellingsrud and John Christian Ottem October 15, 2018€¦ · between algebra and geometry: a ne varieties correspond to nitely generated k-algebras without zero-divisors, and

6.5. The Veronese embedding

6.5.1 Remark on rings generated in degree 1

We will frequently assume that the ring R is generated in degree 1, that is, R

is generated as an R0-algebra by R1. The reason for this will become clear in

the next section. Intuitively, it is because we want ProjR to be covered by the

‘affine coordinate charts’ D+(x) where x should have degree 1.

We remark that this assumption is in fact not not too restrictive: Any

projective spectrum of a finitely generated ring is isomorphic to the Proj of a

ring generated in degree 1. This is because of the basic algebraic fact that if R

is finitely generated, then some subring R(d) will have all of its generators in one

degree, and since ProjR(d) ' ProjR, we don’t change the Proj by replacing R

with R(d).

6.5.2 The Blow-up as a Proj

Consider the ring A = k[x, y] and the ideal I = (x, y). We can form a new

graded ring by introducing a formal variable t and setting

R =⊕m≥0

Iktk

where I0 = A. In R, t has degree 1, and all other variables have degree 0. This

gives a morphism

π : ProjR→ SpecA = A2k

This ProjR is glued together by two affine schemes SpecR(xt) and SpecR(yt)

which are both isomorphic to A2k. It is a nice exercise to check that this coincides

with the gluing of the blow-up of A2k, and that π is the blow-up morphism.

Claim: We have R ' A[u, v]/(xv − yu).

Consider the map φ : A[u, v] → R given by φ(u) = xt, φ(v) = yt. This is

clearly surjective. It suffices to show that kerφ = I where I = (xv − yu). The

‘ ⊇′-direction is clear. Conversely, we can write, modulo I, any element p of

k[x, u, y, v] as ∑ai,j,kx

iujvk +∑

bi′,j′,k′xi′yj

′vk′.

If p ∈ kerφ, we have

0 = φ(p) =∑

ai,j,kxi+jyktj+k +

∑bi′,j′,k′x

j′yi′+j′tk

′.

Hence the coefficients ai,j,k, bi′,j′,k′ vanish except possibly when we have the same

monomials appearing in each sum i.e. when i+ j = i′, k = j′ + k′, i = i′ + j′ and

j + k = k′, in which case we have ai,j,k = −bi′,j′,k. These conditions imply that

j = −j′ which must then both be 0, and so i = i′, k = k′. Hence p = 0 mod I

and so kerφ ⊆ I.

100

Page 101: Geir Ellingsrud and John Christian Ottem October 15, 2018€¦ · between algebra and geometry: a ne varieties correspond to nitely generated k-algebras without zero-divisors, and

Chapter 6. Projective schemes

Now, writing R = A[u, v]/(xv − yu), we see that ProjR is covered by the

two open sets D+(u) = SpecR(v) and D+(v) = SpecR(v). Here

R(u) ' (A[u, v]u/(xv − yu))0 = k[x, v/u]

and

R(v) ' (A[u, v]v/(xv − yu))0 = k[y, u/v].

These are glued along

R(uv) ' (A[u, v]uv/(xv − yu))0 = k[x, y, u/v, v/u]/(x · v/u− y) ' k[x, u/v, v/u]

so we see that ProjR coincides with the previous blow-up description.

101

Page 102: Geir Ellingsrud and John Christian Ottem October 15, 2018€¦ · between algebra and geometry: a ne varieties correspond to nitely generated k-algebras without zero-divisors, and
Page 103: Geir Ellingsrud and John Christian Ottem October 15, 2018€¦ · between algebra and geometry: a ne varieties correspond to nitely generated k-algebras without zero-divisors, and

Chapter 7

Fiber products

7.1 Introduction

One of the most fundamental properties of the category of schemes is that all

fiber products exist. The fiber product is extremely useful in many situations

and takes on astonishingly versatile roles.

The aim of this chapter is to prove the existence theorem for fiber products

and explain the various contexts where fiber products appear. We begin with

recalling the definition of the fibre product of sets, then transition into a very

general situation to discuss fibre product in general categories, for then to return

to the present context of schemes. At the end of the chapter we will go through

a series of examples.

7.1.1 Fiber products of sets.

As a warming up we use some lines on recalling the fiber product in the category

Sets of sets. The points of departure is two sets X1 and X2 both equipped with

a map to a third set S; i.e., we are given a diagram

X1

ψ1

X2

ψ2~~S.

103

Page 104: Geir Ellingsrud and John Christian Ottem October 15, 2018€¦ · between algebra and geometry: a ne varieties correspond to nitely generated k-algebras without zero-divisors, and

7.1. Introduction

The fibre product X1×SX2 is the subset of the cartesian product X×Y consisting

of the pairs whose two components have the same image in S; that is, we have

X1 ×S X2 = (x1, x2) | ψ1(x1) = ψ2(x2) .

Clearly the diagram below where π1 and π2 denote the restrictions of the two

projections to the fiber product—that is, πi(x1, x2) = xi—is commutative,

X1 ×S X2

π2

%%

π1

yyX1

ψ1 %%

X2

ψ2yyS.

(7.1.1)

And more is true; the fibre product enjoys a universal property: Given any two

maps φ1 : Z → X1 and φ2 : Z → X2 such that ψ1 φ1 = ψ2 φ2 there is a unique

map φ : Z → X1 ×S X2 satisfying π1 φ = φ1 and π2 φ = φ2. To lay your

hands on such a φ, just use the map whose two components are φ1 and φ2 and

observe that it takes values in X1×SX2 since the relation ψ1 φ1 = ψ2 φ2 holds.

Giving the two φi’s is to give a commutative diagram like 7.1.1 above with Z

replacing the product X ×S Y , and the universal property is to say that 7.1.1 is

universal—a more precise usage would be to say it is final among such diagrams.

7.1.2 The fiber product in general categories

The notion of a fiber product—formulated as the solution to a universal problem

as above—is mutatis mutandis meaningful in any category C. Given any two

arrows ψi : Xi → S in the category C, an object—that we shall denote by

X1 ×S X2—is said to be the fiber product (fiberproduktet) of the objects Xi, or

more precisely of the two arrows ψi : Xi → S, if the following two conditions are

fulfilled:

• There are two arrows πi : X1 ×S X2 → Xi in C such that ψ1 π1 = ψ2 π2

(called the projections).

• For any two arrows φi : Z → Xi in C such that φ1 ψ1 = φ2 ψ2, there is

a unique arrow φ : Z → X1 ×S X2 satisfying πi φ = φi for i = 1, 2.

The two arrows π1 φ and π2 φ that determine the arrow φ : Z → X1×SX2,

are called the components (komponentene) of φ, and the notation φ = (φ1, φ2) is

sometimes used. If φ1 : Y1 → X1 and φ2 : Y2 → X2 are two arrows over S, there

is a unique arrow denoted φ1× φ2 from Y1×S Y2 to X1×S X2 whose components

are φ1 πY1 and φ2 πY2 .

104

Page 105: Geir Ellingsrud and John Christian Ottem October 15, 2018€¦ · between algebra and geometry: a ne varieties correspond to nitely generated k-algebras without zero-divisors, and

Chapter 7. Fiber products

If the fiber product exists, it is unique up to a unique isomorphism, as is true

for solutions to any universal problem. However, it is a good exercise to check

this in detail in this specific situation.

Exercise 17. Show that if the fiber product exists in the category C, it is unique

up to a unique isomorphism.

It is not so hard to come up with examples of categories where fiber products

do not exist. For instance, consider the category C of subsets X of the integers

with an even number of elements, and morphisms given by inclusions Y ⊆ X.

The fiber product of Y ⊆ X and Z ⊆ X over X would be Y ∩Z ⊆ X – however

Y ∩ Z does not necessarily have an even number of elements!

What is of course much more disappointing is that fiber products fail to

exist in our good old categories like manifolds or affine varieties. This shows yet

another reason why we need to make the transition from varieties to schemes.

Exercise 18. (i) Give an example showing that the fiber product does not

always exist in the category of manifolds.

(ii) Give an example showing that the fiber product does not always exist in

the category of affine varieties.

7.2 Products of affine schemes

The category Aff of affine schemes is, more or less by definition, equivalent to the

category of rings, and in the category of rings we have the tensor product. The

tensor product enjoys a universal property dual to the one of the fiber product.

To be precise, assume A1 and A2 are B-algebras, i.e., we have two maps of rings

αiA1 A2

B.

α1

==

α2

aa

There are two maps βi : Ai → A1⊗B A2 sending a1 ∈ A1 to a1⊗ 1 and a2

to 1⊗ a2, respectively. These are both ring homomorphisms since aa′⊗ 1 =

(a⊗ 1)(a′⊗ 1) respectively 1⊗ aa′ = (1⊗ a)(1⊗ a′), and they fit into the follow-

ing commutative diagram as α1(b)⊗ 1 = 1⊗ α2(b) by the definition of the tensor

product A1⊗B A2 (this is the significance of the tensor product being taken over

105

Page 106: Geir Ellingsrud and John Christian Ottem October 15, 2018€¦ · between algebra and geometry: a ne varieties correspond to nitely generated k-algebras without zero-divisors, and

7.2. Products of affine schemes

B; one can move elements in B from one side of the ⊗-glyph to the other):

A1⊗B A2

A1

β1::

A2

β2dd

B.

α1

99

α2

ee

(7.2.1)

Moreover, the tensor product is universal in this respect. Indeed, assume that

γi : Ai → C are B-algebra homomorphisms, i.e., γ1 α1 = γ2 α2; or said

differently, they fit into the commutative diagram analogous to (7.2.1) with the

βi’s replaced by the γi’s. The association a1⊗ a2 → γ1(a1)γ(a2) is B bilinear,

and hence it extends to a B-algebra homomorphism γ : A1⊗B A2 → C, that

obviously have the property that γ βi = γi.

Applying the Spec-functor to all this, we get the diagram

Spec(A1⊗B A2)

π1vv

π2

((SpecA1

((

SpecA2

vvSpecB

(7.2.2)

and the affine scheme Spec(A1⊗B A2) enjoys the property of being universal

among affine schemes sitting in a diagram like 7.2.2. Hence Spec(A1⊗B A2)

equipped with the two projections π1 and π2 is the fiber product in the category

Aff of affine schemes. One even has the stronger statement; it is the fiber product

in the bigger category Sch of schemes.

Proposition 7.1. Given φi : SpecAi → SpecB. Then Spec(A1⊗B A2) with the

two projection π1 and π2 defined as above is the fiber product of the SpecAi’s

in the category of schemes. That is, if Z is a scheme and ψi : Z → SpecAiare morphisms with ψ1 π1 = ψ2 φ2, there exists a unique morphism ψ : Z →Spec(A1⊗B A2) such that πi ψ = ψi for i = 1, 2.

Proof. We know that the proposition is true whenever Z is an affine scheme;

so the salient point is that Z is not necessarily affine. For short, we let X =

Spec(A1⊗B A2). The proof is just an application of the gluing lemma for

morphisms. One covers Z by open affines Uα and covers the intersections

Uαβ = Uα ∩ Uβ by open affine subsets Uαβγ as well. By the affine case of the

proposition, for each Uα we get a map ψα : Uα → X, such that ψα πi = ψi|Uα .

106

Page 107: Geir Ellingsrud and John Christian Ottem October 15, 2018€¦ · between algebra and geometry: a ne varieties correspond to nitely generated k-algebras without zero-divisors, and

Chapter 7. Fiber products

By the uniqueness part of the affine case, these maps coincide on the open affines

Uαβγ , and therefore on the intersections Uαβ. They can thus be patched together

to a map ψ : Z → X, which is is unique since the ψα’s are unique.

Example 7.2. Let k be a field and x and y two variables. Consider the tensor

product A = k(x)⊗k k(y). We can regard this as a localization of k[x, y] where

we invert everything in the multiplicative set S = p(x)q(x)|p(x), q(y) 6= 0.Suppose that m ⊆ A is a maximal ideal; it has the form S−1p for some prime

ideal p ⊆ k[x, y] which is maximal among the primes that do not intersect S.

In this case we must have pk[x] = 0, since otherwise there will be a non-zero

p(x) ⊆ p ∩ S. Similarly p ∩ k[y] = 0, which implies that p has height at most 1.

Hence either p = (0), of p = (f) for some irreducible polynomial f ∈ k[x, y] not

contained in k[x] ∪ k[y]. If follows that A has dimension 1, and A has infinitely

many maximal ideals.

This example shows how strange the fiber product really is – SpecA is an

infinite set, even though it is the fiber product of two schemes with one-point

sets. We will see more examples like this in the end of this chapter.

7.3 A useful lemma

Lemma 7.3. If X ×S Y exists and U ⊆ X is an open subscheme, then U ×S Yexists and is (canonically isomorphic to) an open subset of X×SY and projections

restrict to projections. Indeed π−1X (U) with the two restrictions πY |π−1

X (U) and

πX |π−1X (U) as projections is a fiber product.

Proof. Displayed the situation appears like

Y

π−1X (U)

99

// X ×s YπX

πY

OO

U // X,

and we are to verify that π−1X (U) together with the restriction of the two projec-

tions to π−1X (U) satisfy the universal property. If Z is a scheme and φX : Z → U

and φY : Z → Y are two morphisms over S we may consider φU as a map into X,

and therefore they induce a map of schemes φ : Z → X ×S Y whist φX = πX φand φY = πY φ. Clearly πX φ = φU takes values in U and therefore φ takes

values in π−1X (U). It follows immediately that φ is unique (see the exercise below),

and we are through.

107

Page 108: Geir Ellingsrud and John Christian Ottem October 15, 2018€¦ · between algebra and geometry: a ne varieties correspond to nitely generated k-algebras without zero-divisors, and

7.4. The gluing process

Exercise 19. Assume that U ⊆ X is an open subscheme and let ι : U → X be

the inclusion map. Let φ1 and φ2 be two maps of a schemes from a scheme Z to

U and assume that ι φ1 = ι φ2. Then φ1 = φ2.

When identifying π−1X (U) with U ×S Y , the inclusion map π−1

X (U) ⊆ X ×S Ywill correspond to the map ι × idY where ι : U → X is the inclusion, so a

reformulation of the lemma is that open immersions stay open immersion under

change of basis.

7.4 The gluing process

The following proposition will be basis for all gluing necessary for the construction:

Proposition 7.4. Let ψX : X → S and ψY : Y → S be two maps of schemes,

and assume that there is an open covering Ui i∈I of X such that Ui ×S Y exist

for all i ∈ I. Then X ×S Y exists. The products Ui ×S Y form an open covering

of X ×S Y and projections restrict to projections.

Proof. We need some notation. Let Uij = Ui ∩ Uj be the intersections of the

Ui’s, and let πi : Ui ×S Y → Ui denote the projections. By lemma 7.3 there

are isomorphisms θji : π−1i (Uij) → Uij ×S Y , and gluing functions we shall use

τji = θ−1ij θji that identifies π−1

i (Uij) with π−1j (Uij). The picture is like this

Ui ×S Y ⊇ π−1i (Uij)

θji

'//Uij ×S Y

θ−1ij

'//π−1j (Uji) ⊆ Uj ×S Y.

The gluing maps τij clearly satisfy the gluing conditions being compositions

of that particular form, and the scheme emerging from gluing process is X ×S Y .

The two projections are essential parts of the product and mut not be

forgotten: The projection onto Y is there all the time since Y is nevener touched

during the construction. The projection onto X is obtained by gluing the

projections πi along the π−1i (Uij). By lemma 7.3 we know that when identifying

π−1i (Uij) with the product Uij×SY , the projection πij onto Uij corresponds to the

restriction πi|π−1i (Uij)

. This means that πi|π−1i (Uij)

= πijθji. Saying that πi|π−1i (Uij)

and πj|π−1j (Uij)

become equal after gluing, is to say that πi|π−1i (Uij)

= πj|π−1j (Uij)

τji(remember that in the gluing process points x and τji(x) are identified), but this

holds true since

πj|π−1j (Uij)

τji = πij θij τji = πij θij θ−1ij θji = πij θji = πi|π−1

i (Uij).

Hence we can glue the πi’s together to obtain πX .

It is a matter of easy verification that the the glued scheme with the two

projections has the universal property.

108

Page 109: Geir Ellingsrud and John Christian Ottem October 15, 2018€¦ · between algebra and geometry: a ne varieties correspond to nitely generated k-algebras without zero-divisors, and

Chapter 7. Fiber products

It is worth while commenting that the product X ×S Y is not defined

as a particular scheme, it is just an isomorphism class of schemes (having

the fundamental property that there is a unique isomorphism respecting the

projections between any two). In the proof above both π−1i (Uij) and π−1

j (Uij) are

products Uij ×S Y , they are however not equal, merely canonically isomorphic.

In the construction we could have use any of them, or as we in fact did, any

non-specified representative of the isomorphism class, as this makes the situation

more symmetric in i and j.

An immediate consequence of the gluing proposition 7.4 is that fibre products

exist over an affine base S.

Lemma 7.5. Assume that S is affine, then X ×S Y exists.

Proof. First if Y as well is affine, we are done. Indeed, cover X by open affine

sets Ui. Then Ui ×S Y exists by the affine case, and we are in the position to

apply proposition 7.4 above. We then cover Y by affine open sets Vi. As we just

verified, the products X ×S Vi all exist, and applying proposition 7.4 once more,

we can conclude that X ×S Y exists.

7.5 The final reduction

Let Si be an open affine covering of S and let Ui = ψ−1(Si) and Vi = ψ−1Y (Si).

By lemma 7.5 the products Ui ×Si Vi all exist. Using the following lemma and,

for the third time, the gluing proposition 7.4 we are trough:

Lemma 7.6. With current notation, we have the equality Ui ×Si Vi = Ui ×S Y .

Proof. The key diagram is

Zf

g

Ui

ψX |Ui

Y

ψY

⊇ Vi

S

where f and g are given maps. If one follows the left path in the diagram, one

ends up in Si, and hence the same must hold following the right path. But then,

Vi being equal the inverse image ψ−1Y (Si), it follows that g necessarily factors

through Vi, and we are done.

109

Page 110: Geir Ellingsrud and John Christian Ottem October 15, 2018€¦ · between algebra and geometry: a ne varieties correspond to nitely generated k-algebras without zero-divisors, and

7.6. Notation.

7.6 Notation.

If S = SpecA on often writes X×AY in short for X×SpecAY . If S = SpecZ, one

writes X × Y . In case Y = SpecB the shorthand notation X ⊗AB is frequently

seen as well—it avoids writing Spec twice.

Diagrams arising from fiber products are frequently called cartesian diagram

(kartesiske diagrammer); that is, the diagram

Zπ1 //

π2

X

ψX

YψY// S

is said to be a cartesian diagram if there is an isomorphism Z ' X ×S Y with

π1 and π2 corresponding to the two projections.

Exercise 20. Let X, Y and Z be three schemes over S. Show that X×S S = X,

that X ×S Y ' Y ×S X and that (X ×S Y )×S Z ' X ×S (Y ×S Z). If T is a

scheme over S, show that X ×S T ×T Y ' X ×S Y .

Exercise 21. Show by using the universal property that if φ : X ′ → X and

g : Y ′ → Y are morphisms over S, then there is morphism f × g : X ′ ×S Y ′ →X ×S Y such that

X ′ ×S Y ′f×g //

πX′

X ×S YπX

X ′ // X

and a corresponding diagram involving Y and Y ′ commute

7.7 Examples

7.7.1 Varieties versus Schemes

In the important case that X and Y are integral schemes of finite type over the

algebraically closed field k the product of the two as varieties coincides with

their product as schemes over k, with the usual interpretation that the variety

associated to the scheme X is the set closed points X(k) with induced topology.

The product X ×k Y will be a variety (i.e., an integral scheme of finite type

over k) and the closed points of the product X ×k Y will be the direct product

of the closed points in X and Y .

It is not obvious that A⊗k B is an integral domain when A and B are, and

in fact, in general, even if k is a field, it is by no means true. But it holds true

110

Page 111: Geir Ellingsrud and John Christian Ottem October 15, 2018€¦ · between algebra and geometry: a ne varieties correspond to nitely generated k-algebras without zero-divisors, and

Chapter 7. Fiber products

whenever A and B are of finite type over k and k is an algebraically closed field.

The standard reference for this is Zariski and Samuel’s book Commutative algebra

I which is the Old Covenant for algebraists. It is also implicit in Hartshorne’s

book, exercise 3.15 b) on page 22.

However that the tensor product A⊗k B is of finite type over when A and B

are, is straightforward. If u1, . . . , um generate A over k and v1, . . . , vm generate

B over k the products ui⊗ 1 and 1⊗ vj generate A⊗k B.

7.7.2 Non-algebraically closed fields

This case the situation is more subtle when one works over fields that are not

algebraically closed. To illustrate some of the phenomena that can occur, we

study a few basic examples.

Example 7.7. A simple but illustrative example is the product SpecC×SpecRSpecC. This scheme has two distinct closed points, and it is not integral—it is

not even connected!

The example also shows that the underlying set of the fiber product is not

necessarily equal to the fiber product of the underlying sets, although this was

true for varieties over an algebraically closed field. In the present case the three

schemes involved all have just one element and the their fibre product has just

one point. So we issue warnings: The product of integral schemes is in general

not necessarily integral! The underlying set of the fiber product is not always

the fiber product of the underlying sets.

The tensor product C⊗R C is in fact isomorphic to the direct product C×C of

two copies of the complex field C; indeed, we compute using that C = R[t]/(t2+1)

and find

C⊗R C = R[t]/(t2 + 1)⊗RC = C[t]/(t2 + 1) = C[t]/(t− i)(t+ i) = C× C

where for the last equation we use the Chinese remainder theorem and that the

rings C[t]/(t± i) both are isomorphic to C.

Example 7.8. This little example can easily be generalized: Assume that L is

a simple, separable field extension of K of degree d; that is L = K(α) where

the minimal polynomial f(t) of α over K is separable and of degree d. Let

Ω be a field extension of K in which the polynomial f(t) splits completely—

e.g., a normal extension of L or any algebraically closed field containing K—

then by an argument completely analogous to the one above one finds that

L⊗K Ω=Ω× . . .×Ω where the product has d factors. Consequently the product

scheme SpecL ×Spec k Spec Ω has an underlying set with d points, even if the

three sets of departure all are prime spectra of fields and thus singletons.

111

Page 112: Geir Ellingsrud and John Christian Ottem October 15, 2018€¦ · between algebra and geometry: a ne varieties correspond to nitely generated k-algebras without zero-divisors, and

7.7. Examples

One may push this further and construct examples where SpecK ⊗Spec k Spec Ω

is not even noetherian and has infinity many points!

Exercise 22. With the assumptions of the example above, check the statement

that L⊗K Ω ' Ω× . . .× Ω, the product having d factors.

Exercise 23. Assume that A is an algebra over the field k having a countable

set e1, e2, . . . , ei, . . . of mutually orthogonal idempotents, i.e., eiej = 0 if i 6= i

and eiei = 1, and assume that eiA ' k. Assume also that every element is a

finite linear combination of the ei’s.

• Show that the ideal Ij generated by the ei’s with i 6= j is a maximal ideal.

Example 7.9. In this example we let L ∈ C[x, y] be a linear form that is not real,

for example L = x+ iy + 1, and we introduce the real algebra A = R[x, y]/(LL).

The product LL of L and its complex conjugate is a real irreducible quadric;

which in the concrete example is (x+ 1)2 + y2. The prime spectrum SpecA is

therefore an integral scheme. However, the fiber product SpecA ×R SpecC is

not irreducible being the union of the two conjugate lines L = 0 and L = 0 in

SpecC[x, y].

The scheme SpecA has just one real point, namely the point (−1, 0) (i.e.,

corresponding to the maximal ideal (x + 1, y)). The C-points however, are

plentiful.

They are contained in the C-points A2R(C), which are of orbits (a, b), (a, b)

of the complex conjugation with (a, b) non-real, and form the subset of those

(a, b) such that L(a, b) = 0.

Example 7.10. Another example along same lines as example 7.8 shows that

the fiber product X ×S Y is not necessarily reduced even if both X and Y are;

the point being to use an inseparable polynomial f(t) in stead of a the separable

one in 7.8. Let k be a non-perfect field in characteristic p which means that

there is an element a ∈ k not being a p-th power of any element in k. Let L be

the field extension L = k(b) where bp = a. That is, L = k[t]/(tp − a). which is

a field since tp − a is an irreducible polynomial over k. However, upon being

tensored by itself over k, it takes the shape

L⊗k L = L[t]/(tp − a) = L[t]/(tp − bp) = L[t]/((t− b)p)

which is not reduced, the non-zero element t− b being nilpotent. So we issue a

third warning: the fiber product of integral schemes is not in general reduced!

One can elaborate these example and construct an example of two noetherian

schemes X and Y such that X ×S Y is not noetherian, even if S is the spectrum

of a field. The next examples shows that if X, Y and S are fields, the product

X ×S Y can even have an uncountable number of elements!

112

Page 113: Geir Ellingsrud and John Christian Ottem October 15, 2018€¦ · between algebra and geometry: a ne varieties correspond to nitely generated k-algebras without zero-divisors, and

Chapter 7. Fiber products

Example 7.11. In this example we take L to the subfield of the field Q of

algebraic numbers that is generated by all the elements a such that a2n = 2 for

some n. The field L is the union of the ascending chain of fields

Q ⊆ Q(√

2) ⊆ Q(4√

2) ⊆ Q(8√

2) ⊆ . . . ⊆ Q(2r√

2) ⊆ . . . ⊆ L ⊆ Q

We denote r-th field in the chain Q( 2r√

2) by Lr. The next field Lr+1 is the

quadratic extension of Lr obtained simply by adjoining 2r+1√2, or, in other words,

the square root of 2r√

2.

We let Ar = Lr⊗QQ. This is a finite algebra of rank 2r over Q. It has a

refined algebraic structure given by the smaller algebras As for s ≤ r which are

all subalgebras of Ar. the algebra Ar splits as the direct product of two copies of

Ar−1; indeed, one has Lr⊗QQ = Lr⊗Lr−1 Q⊗Q Lr−1⊗Q Q = Ar−1 ⊕ Ar−1 since

Lr⊗Lr−1 Q ' Q⊕Q.

The two idempotents in Ar that induce this splitting are denoted by er,0and er,1. They are orthogonal and their sum equals one. Each of the two

subalgebras er,εAr are isomorphic to Ar−1 with 1 ∈ Ar−1 corresponding to er,ε,

they contain the idempotens er−1,ε which will correspond to the product er,εer−1,ε′

in Ar. Working our way down in Ar, this yields a sequence of idempotents

eI = er,εrer−1,εr−1 . . . e1,ε1 where I = (ε1, . . . , εr) is a sequence of 1’s and 0’s.

Take any sequence σ = (σi)i ∈ N . Let I ⊆ A be generated by the ei,εi with

εi /∈ σ. Then I is maximal. It contains all product except∏

i ei,σi . And these all

generate the same copy of Q ⊆ A!

Then of course L being a field is noetherian as is Q, but L⊗QQ is not!

Indeed, Q( 2r√

2)⊗Q is isomorphic to the direct product of 2r-copies of Q so

L⊗QQ is the union of a sequence of subrings each being a direct product of a

steadily increasing number of copies of Q.

7.8 Base change

The fiber product is in constant use in algebraic geometry, and it is an aston-

ishingly versatile and flexible instrument. In different situations it serves quite

different purposes and appears under different names. We shall comment on

some of the most frequently encountered applications, and we begin with notion

of base change.

In its simples and earliest appearances base change is just extending the

field over which one works; e.g., in Galois theory, or even in the theory of real

polynomials, when studying an equation with coefficients in a field k one often

finds it fruitful to study the equation over a bigger field K. Generalizing this to

extensions of algebras over which one works, and then to schemes, one arrives

naturally at the fiber product.

113

Page 114: Geir Ellingsrud and John Christian Ottem October 15, 2018€¦ · between algebra and geometry: a ne varieties correspond to nitely generated k-algebras without zero-divisors, and

7.9. Scheme theoretic fibres

If X is a scheme over S and T → S is map. Considering T → S as change of

base schemes one frequently writes XT = X ×S T and says that XT is obtained

from X by base change (basisforandring) or frequently that XT is the pull back

(tilbaketrekningen) of X along T → S. This is a functorial construction, since

if φ : X → Y is a morphism over S, there is induced a morphism φT = φ× idTfrom XT to YT over T , and one easily checks that (φ φ′)T = φT φ′T . The

defining properties of φT are πY φT = φ πX and πT φ = πT , as depicted in

the diagram:

T

XT = X ×S TφT //

πT55

φπX ))

Y ×S TπY

πT

OO

Y.

Exercise 24. Verify in detail that (φ φ′)T = φT φ′T .

If P is a property of morphisms, one says that P is stable under base change

if for any T over S, the map fT has the property P whenever f has it. For

example, another way of phrasing lemma 7.3 on page 107 is to say that being an

open immersion is stable under base change.

7.9 Scheme theoretic fibres

In most parts of mathematics, when one studies a map of some sort, a knowledge

of what the fibres of the map are, is of great help. This is also true in the theory

of schemes.

Suppose that φ : X → Y is a map of schemes and that y ∈ Y is a point. On

the level of topological spaces, we are interested in the preimage φ−1(y), and

we aim at giving a scheme theoretic definition of the fiber φ−1(y). Having the

fiber product at our disposal, nothing is more natural than defining the fiber

to be the fiber product φ−1(Y ) = Spec k(y)×Y X, where Spec k(y)→ Y is the

map corresponding to the point y. Recall that k(y) = OY,y/my and the map is

the composition Spec k(y)→ Spec OY,y → Y . For diagrammoholics, the scheme

theoretic fiber of φ over y fits into the cartesian diagram

φ−1(y) = X ×Y Spec k(y) //

X

φ

Spec k(y) // Y

.

As the next lemma will show, the underlying topological space of φ−1(y) is

the topological fiber, but in addition there is a scheme structure on it. In in

114

Page 115: Geir Ellingsrud and John Christian Ottem October 15, 2018€¦ · between algebra and geometry: a ne varieties correspond to nitely generated k-algebras without zero-divisors, and

Chapter 7. Fiber products

many cases it is not reduced, and this a mostly a good thing since it makes

certain continuity results true.

Proposition 7.12. The inclusion Xy → X of the scheme theoretic fiber is a

homeomorphism onto the topological fiber φ−1(y).

Proof. We start with the affine case. Obviously Y can always without loss of

generality be assumed to be affine, say Y = SpecB, but to begin with we adopt

the additional assumption that X is affine as well, let’s say X = SpecA.

The map φ of affine schemes is induced by a map of rings α : B → A. Let

p ⊆ B be a prime ideal. We have the following equality between sets

q ⊆ A | q prime ideal , α−1(q) ⊇ p = q ⊆ A | q prime ideal , q ⊇ pA .

In the particular case that p is a maximal ideal, the inclusion α−1(q) ⊇ q is

necessarily an equality, and the sets above describe the fiber set-theoretically:

φ−1(p) = q ⊆ A | q ⊇ pA ' SpecA/pA.

But this also describes the good old embedding of SpecA/pA into SpecA identi-

fying it with the closed subscheme V (pA), and therefore this yields a homeomor-

phism between SpecA/pA and the topological fiber φ−1(p). On the other hand

by standard equalities between tensor products one has

A/pA = A⊗B B/pB = A⊗B k(y),

and so the scheme theoretical fiber φ−1(y) = Xy = X×Y Spec k(y) = SpecA⊗B k(y)

is in a canonical way homeomorphic to the topological fiber.

If p is not a maximal ideal, the set SpecA/pA is strictly bigger than the fiber,

the superfluous prime ideals being those for which α−1q strictly bigger than p.

When localising in the multiplicative system S = B − p ⊆ B, these superfluous

prime ideals go non-proper, since they all contain elements of the form α(s) with

s ∈ S. Hence the points in the fiber correspond to the primes in the localized

ring (A/pA)p. Standard formulas for the tensor product gives on the other hand

the equality

(A/pA)p = A⊗B Bp/pBp = A⊗B k(y).

The topologies coincides as well, since Spec(A/pA)p naturally is a subscheme

of SpecA/pA; induced topology being the one a prime spectrum.

In the general case, i.e., when X is no longer affine, we cover X by open, affine

Ui’s. By the universal property of fiber products, we know that U ∩Xs = Us (see

the diagram below). This shows that the scheme theoretical and the topological

115

Page 116: Geir Ellingsrud and John Christian Ottem October 15, 2018€¦ · between algebra and geometry: a ne varieties correspond to nitely generated k-algebras without zero-divisors, and

7.9. Scheme theoretic fibres

fiber coincide as topological spaces.

Us //

U

Xs

//

X

Spec k(y) // Y

Example 7.13. We take a look at a simple but classic example: The map

Spec k[x, y]/(x− y2)→ Spec k[x]

induced by the injection B = k[x] → k[x, y]/(x − y2) = A. Geometrically one

would say it is just the projection of the parabola onto the x-axis.

If a ∈ k computing the fiber yields, where k(a) denotes the field k(a) =

k[t]/(t− a) (which of course is just a copy of k).

k[x, y]/(x− y2)⊗k[x] k[x]/(x− a) ' k[y]/(y2 − a).

(Here we are using the isomorphism R/I ⊗RM ' M/IM for an ideal I in

an R-module M .)

Several cases can occur, apart from the characteristic two case being special.

• If a does not have a square root in k, the fiber is Spec k(√a) where k(

√a)

is a quadratic field extension of k.

• In case a has a square root in K, way b2 = a, the polynomial y2−a factors as

(y − b)(y + b), and the fiber becomes Spec k[y]/(y − b)× Spec k[y]/(y + b),

the disjoint union of two copies of Spec k

• Finally, the case appears when a = 0. The the fiber is not reduced, but

equals Spec k[y]/y2.

We also notice that the generic fiber is the quadratic extension k(x)(√x) of

the function field k(x).

Over perfect fields k of characteristic two, the picture is completely different.

Then a is a square, say a = b2 and as (y2 − b2) = (y − b)2 non of the fibers are

reduced, they equal Spec k[y]/(y − b)2, except the generic one which is k(x)(√x).

One observes interestingly enough, that all the non-reduced fibers deform into a

field!

116

Page 117: Geir Ellingsrud and John Christian Ottem October 15, 2018€¦ · between algebra and geometry: a ne varieties correspond to nitely generated k-algebras without zero-divisors, and

Chapter 7. Fiber products

Example 7.14. A similar example can be obtained from the map

f : SpecA→ SpecB

where A = Spec k[x, y, z]/(xy−z) and B = k[z] (The map k[z]→ k[x, y, z]/(xy−z) is the obvious one). As before, we pick an element a ∈ SpecB, and consider

the fiber

Xa = Spec (A⊗B k(a)) = Spec k[x, y]/(xy − a)

Again, two cases occur. If a 6= 0, in which case xy−a is an irreducible polynomial,

and so Xa is an integral scheme. This is intuitive, since it corresponds to the

hyperbola xy = a. If a = 0, we are left with X0 = Spec k[x, y]/(xy), which is

not irreducible; it has two components corresponding to V (x) and V (y). (X0 is

reduced however).

For good measure, we also consider the fiber over the generic point η of

SpecB. This corresponds to

k[x, y, z]/(xy − z)⊗k[z] k(z) = k(z)[x, y]/(xy − z)

which is an integral domain. Hence Xη is integral.

7.10 Separated schemes

We have previously seen that the topology on schemes behaves very differently

from the usual euclidean topoology. In particular, schemes are not Hausdorff,

except in trivial cases – the open sets in the Zariski topology are just too large.

Still we would like to find an analogous property that can serve as a substitute

for this property. The route we take is to impose that the diagonal should be

closed.

7.10.1 The diagonal

Let X/S be a scheme over S. There is a canonical map ∆X/S : X → X ×S X of

schemes over S called the diagonal map or the diagonal morphism (diagonalav-

bildningen). The two components maps of ∆ are both equal to the identity idX ;

that is, the defining properties of ∆X/S are πi ∆X/S = idX for i = 1, 2 where

the πi’s denote the two projections.

In the case X and S are affine schemes, the diagonal has a simple and natural

interpretation in terms of algebras; it corresponds to most natural map, the

multiplication map:

µ : A⊗B A→ A.

It sends a⊗ a′ to the product aa′ and then extends to A⊗B A by linearity. The

117

Page 118: Geir Ellingsrud and John Christian Ottem October 15, 2018€¦ · between algebra and geometry: a ne varieties correspond to nitely generated k-algebras without zero-divisors, and

7.10. Separated schemes

projections correspond to the two maps ιi : A → A⊗B A sending a to a⊗ 1

respectively to 1⊗ a. Clearly it holds that µ ιi = idA, and on the level of

schemes this translates into the defining relations for diagonal map. Moreover µ

is clearly surjective, so we have established the following:

Proposition 7.15. If X an affine scheme over the affine scheme S, then the

diagonal ∆X/S : X → X ×S X is a closed imbedding.

This is not generally true for schemes and shortly we shall give examples,

however from the proposition we just proved, it follows readily that the image

∆X/S(X) is locally closed—i.e., the diagonal is locally a closed embedding:

Proposition 7.16. The diagonal ∆X/S is locally a closed embedding.

Proof. Begin with covering S by open affine subset and subsequently cover each

of their inverse images in X by open affines as well. In this way one obtains a

covering of X by affine open subsets Ui whose images in S are contained in affine

open subsets Si. The products Ui×Si Ui = Ui×SUi are open and affine, and their

union is an open subset containing the image of the diagonal. By proposition

7.15 above the diagonal restricts to a closed embedding of Ui in Ui ×Si Ui.

Exercise 25. In the setting of the previous proof, show that ∆X/S|Ui = ∆Ui/S.

7.10.2 Separated schemes

Definition 7.17. The scheme X/S is separated (separert) over S, or that the

structure map X → S is separated if the diagonal map is a closed imbedding. If

X is separated over SpecZ one says for short that it is separated.

Example 7.18. The simplest example of a scheme that is not separated is

obtained by glueing the prime spectrum of a discrete valuation ring to itself

along the generic point.

To give more details let R be the DVR with fraction field K. Then SpecR =

x, η where x is the closed point corresponding to the maximal ideal, and η

is the generic point corresponding to the zero ideal. The generic point η is an

open point (the complement of η is the closed point x) and the support of

the open subscheme η = SpecK. By by the glueing lemma, we may glue one

copy of SpecR to another copy of SpecR by identifying the generic points—that

is, the open subschemes SpecK—in the two copies.

In this manner we construct a scheme ZR together with two open embeddings

ιi : SpecR→ ZR. They send the generic point η to the same point, which is an

open point in ZR, but they differ on the closed point x.

118

Page 119: Geir Ellingsrud and John Christian Ottem October 15, 2018€¦ · between algebra and geometry: a ne varieties correspond to nitely generated k-algebras without zero-divisors, and

Chapter 7. Fiber products

It follows that the diagonal is not closed. Indeed, the subscheme of SpecR×SpecR where the two maps ιi agree is the preimage of the diagonal. But this

subscheme has exactly one point which is open.

Exercise 26. Show that ZR × ZR is obtained by glueing four copies of SpecR

together along their generic points. Show that the diagonal is open and not

closed.

Example 7.19. The affine line with two origins from Chapter 3 is not separated.

One of the nice properties affine schemes enjoy, is the following:

Proposition 7.20. Assume that X is separated and that U and V are to

open affine subscheme. Then the intersection U ∩ V is affine and the map

Γ(U,OU)⊗ Γ(V,OV )→ Γ(U ∩ V,OU×V ) is surjective

Proof. The product U × V is an open and affine subset of X × V , and U ∩ V =

∆X(X) ∩ (U × V ). So if the diagonal is closed, U ∩ V is a closed subset of the

affine set U ×V hence affine. It is a general fact about products of affine schemes

that one has

Γ(U × V,OU×V ) = Γ(U,OU)⊗ Γ(V,OV ),

and as U ∩ V is a closed subscheme of U × V , the restriction map

Γ(U × V,OU×V )→ Γ(U ∩ V,OU∩V )

is surjective.

7.11 The valuative criterion

In some sense, these tiny schemes ZR together with some of their bigger cousins

are always at the root of a non separated scheme. Using the properties of these

subschemes, one can prove the following theorem, which gives a convenient

description of separated schemes.

Proposition 7.21. Assume that X is a quasi-compact scheme. Then X/S is

non-separated if and only if it contains a subscheme ZR/S for some valuation

ring R.

119

Page 120: Geir Ellingsrud and John Christian Ottem October 15, 2018€¦ · between algebra and geometry: a ne varieties correspond to nitely generated k-algebras without zero-divisors, and

7.11. The valuative criterion

This is usually formulated as follows:

Theorem 7.22 (Valuative criterion for separateness). A quasi-compact scheme

X is separated if and only if the following condition is satisfied: For any DVR

R with fraction field K, a map SpecK → X over S has at most one extension

to SpecR→ X.

In other words, a morphism X → S is separated if for every DVR R with

fraction field K and diagram

Spec(K) //

X

Spec(R) //

;;

S

there can be at most one extension SpecR → X (the dotted arrow in the

diagram) such that everything commutes.

This looks like a technical statement, but it is tremendously powerful for

proving theorems about separateness. Here is a sample:

Corollary 7.23. (i) Open and closed immersions are separated

(ii) A composition of two separated morphisms is again separated

(iii) Separatedness is stable under base change: If f : X → S is separated, and

S ′ → S is any morphism, then f ′ : X ×S S ′ → S ′ is separated.

The proof is straightforward using the Valuative Criterion.

For instance, consider item (iii). Taking a base change S ′ → S gives us a

diagram

Spec(K) //

X ×S S ′

g

// X

f

Spec(R) //

8888

S ′ // S

giving two arrows SpecR ⇒ X (by composition), which have to be equal, by

separatedness of f . Now, the two morphisms SpecR ⇒ X ×S S ′ have to be

equal by the universal property of the fiber product.

It is not true that maps SpecK → X like in the criterion always can be

extended—schemes with that property are called proper schemes—so the criterion

says there can never be two different extensions in case X is separated.

One doesn’t frequently encounter non-separated scheme in practice, but some

very nice properties are only true for separated schemes, and this legitimates the

notion. Of course, one needs good criteria to be sure we have a large class of

120

Page 121: Geir Ellingsrud and John Christian Ottem October 15, 2018€¦ · between algebra and geometry: a ne varieties correspond to nitely generated k-algebras without zero-divisors, and

Chapter 7. Fiber products

separated schemes. We have already seen that all affine schemes are separated,

and when we come to that point projective schemes will turn out to be separated

as well.

7.11.1 Varieties vs schemes again

Finally we can state the definition of a variety:

Definition 7.24. A variety X is an integral, separated scheme of finite type

over an algebraically closed field.

This definition requires some explanation. Recall that we defined for a field

k a functor

t : Var/k → Sch/k

which associates a variety V to a scheme t(V ) over k, such that V is homeomorphic

to the subset of k-points of t(V ). We stated that this functor is fully faithful,

making the category of varieties Var/k equivalent to a full subcategory of Sch/k.

It is a theorem (Hartshorne II.4.10) that this subcategory is exactly the category

of schemes satisfying Definition 7.24. For the proof of this statement, see See

Hartshorne II.4.10.

Theorem 7.25. The functor t is fully faithful, and gives and equivalence between

the category of varieties Var/k and the subcategory of Sch/k of schemes satisfying

Definition 7.24.

121

Page 122: Geir Ellingsrud and John Christian Ottem October 15, 2018€¦ · between algebra and geometry: a ne varieties correspond to nitely generated k-algebras without zero-divisors, and
Page 123: Geir Ellingsrud and John Christian Ottem October 15, 2018€¦ · between algebra and geometry: a ne varieties correspond to nitely generated k-algebras without zero-divisors, and

Chapter 8

Quasi-coherent sheaves

When you study commutative algebra may be you are primarily interested in

the rings and ideals, but probably you start turning your interest towards the

modules pretty quickly; they are an important part of the world of rings, and to

get the results one wants, one can hardly do without them. The category ModAof A-modules is a fundamental invariant of a ring A; and in fact, is the principal

object of study in commutative algebra. This is true also with schemes; for

which the so called quasi-coherent Ox-modules form an important attribute of

the scheme, if not a decisive part of the structure. They form a category QCohXwith many properties parallelling those of the category ModA. In fact, in case

the schemeX is affine, i.e., X = SpecA, the two categories ModA and ModXare equivalent, as we will prove shortly. Imposing finiteness conditions on the

OX-modules one arrives at the category CohX of so called coherent OX-modules

that in the noetherian case parallel the finitely generated A-modules.

We start out by describing the much broader concept of an OX-module. The

theory here is presented for schemes, but the concept is meaningful for any ringed

space.

In the literature one finds different approaches to the quasi-coherent sheaves.

We follow EGA and Hartshorne and introduce the quasi-coherent module first for

affine schemes. If X = SpecA one defines an OX-module M associated with an

A-module M , and in the general case these modules serve as the local models for

the quasi-coherent ones. There is a notion of quasi-coherence for OX-modules on

a general locally ringed space. In some other branches of mathematics they are

important, e.g., analytic geometry, but we concentrate our efforts on schemes.

At the end of the chapter we will discuss locally free sheaves and vector

bundles, which are in many respects the most interesting coherent sheaves on a

123

Page 124: Geir Ellingsrud and John Christian Ottem October 15, 2018€¦ · between algebra and geometry: a ne varieties correspond to nitely generated k-algebras without zero-divisors, and

8.1. Sheaves of modules

scheme.

8.1 Sheaves of modules

A module over a ring is just an additive abelian group equipped with a multi-

plicative action of A. Loosely speaking we can multiply members of the module

by elements from the ring, and of course, the usual set of axioms must be

satisfied. In a similar way, if X is a scheme, an OX-module is an abelian sheaf

F whose sections can be multiplied by sections of OX ; the multiplicator and the

multiplicand of course being sections over the same open subset.

Formally, an OX-module structure on the abelian sheaf F is defined as a family

of multiplication maps Γ(U,F)×Γ(U,OX)→ Γ(U,F)—one for each open subset

U of X—making the space of sections Γ(U,F) into a Γ(U,OX)-module in a way

compatible with all restrictions. That is, for every pair of open subsets V ⊆ U ,

the natural diagram below—where vertical arrows are made up of appropriate

restrictions, and the horizontal arrows are multiplications—commutes

Γ(U,F)× Γ(U,OX)

// Γ(U,F)

Γ(V,F)× Γ(V,OX) // Γ(V,F).

Maps, or homomorphisms, of OX-modules are just maps α : F → G between

OX-modules considered as abelian sheaves, respecting the multiplication by

sections of OX . That is, for any open U the map αU : Γ(U,F) → Γ(U,G) is a

Γ(U,OX)-homomorphism.

We have now explained what OX-modules are and told what their homomor-

phisms should be, and this organizes the OX-modules into a category that we

denote by ModX .

The category ModX is an additive category : The sum of two OX-homomorphisms

as maps of abelian sheaves is again an OX-homomorphism. So for all F and

G the set HomOX (F ,G) is an abelian group, and the compositions maps are

bilinear. Moreover, the direct sum of two OX-modules as abelian sheaves has

an obvious OX-structure, with multiplication being defined componentwise. In

fact, this argument works for arbitrary direct sums (or coproducts as they also

are called). For any family Fi of OX-modules,⊕

i∈I Fi is an OX-module (see

Exercise 31 below).

The notions of kernels, cokernels and images of OX-module homomorphisms

now appear naturally. All the three corresponding abelian constructions are

invariant under multiplication by sections of OX , and therefore they have OX-

module structures. The respective defining universal properties (in the category

124

Page 125: Geir Ellingsrud and John Christian Ottem October 15, 2018€¦ · between algebra and geometry: a ne varieties correspond to nitely generated k-algebras without zero-divisors, and

Chapter 8. Quasi-coherent sheaves

of OX-modules) come for free, and one easily checks that this makes ModX an

abelian category; i.e., the kernels of the cokernels equal the cokernels of the

kernels.

There is tensor product of OX-modules denoted by F ⊗OX G. As in many

other cases, the tensor product F ⊗OX G is defined by first describing a presheaf

that subsequently is sheafified. The sections of the presheaf, temporarily denoted

by F ⊗′OX G, are defined in the natural way by

Γ(U,F ⊗′OX G) = Γ(U,F)⊗Γ(U,OX) Γ(U,G).

There is also a sheaf of OX-homomorphisms between F and G. The definition

goes along lines slightly different from the definition of the tensor product. Recall

the sheaf Hom (F ,G) of homomorphisms between the abelian sheaves F and Gwhose sections over an open set U is the group Hom(F|U ,G|U ) of homomorphisms

between the restrictions F|U and G|U . Inside this group one has the subgroup

of the maps being OX-homomorphisms, and these subgroups, for different open

sets U , are respected by the restriction map. So they form the sections of a

presheaf, that turns out to be a sheaf, and that is the sheaf HomOX (F ,G) of

OX-homomorphisms from F to G.

Example 8.1 (Modules on spectra of DVR’s). Modules on the prime spectrum

of a discrete valuation ring R are particularly easy to describe.

Recall that the scheme X = SpecR has only two non-empty open sets, the

whole space X it self and the singleton η where η denotes the generic point.

The singleton η is underlying set of the open subscheme SpecK, where K

denotes the fraction field of R.

We claim that to give an OX-module is equivalent to giving an R-module M ,

a K-vector space N and a homomorphism ρ : M ⊗R K → N .

Indeed, given an OX module F , we get an R module M = Γ(X,F), and

a vector space N = Γ( η ,F) over K. The homomorphism ρ is just the

restriction map F(X)→ F( η ). Conversely, given the data M,N, ρ, we can

define F(X) = M and F(η) = N . The map ρ : F(X)→ F( η ) makes F into

an OX-module.

Note that the restriction map can be just any R-module homomorphism

M → N . In particular, it can be the zero homomorphism, and in that case M

and N can be completely arbitrary modules. Again this illustrates the versatility

of general OX-modules.

Example 8.2 (A bunch of wild examples). The OX-modules (or at least some

of them) play a leading leading role in the theory of schemes, and shortly we

shall see a long series of examples. These will all be so called quasi-coherent

sheaves. The examples we now describe are a bunch of wild examples, intended

125

Page 126: Geir Ellingsrud and John Christian Ottem October 15, 2018€¦ · between algebra and geometry: a ne varieties correspond to nitely generated k-algebras without zero-divisors, and

8.1. Sheaves of modules

to show that OX-modules without any restrictive hypothesis are very general

and often unmanageable objects.

Recall the Godement construction from Chapter 1. Given any collection of

abelian groups Ax x∈X indexed by the points x of X. We defined a sheaf Awhose sections over an open subset U was

∏x∈U Ax, and whose restriction maps

to smaller open subsets were just the projections onto the corresponding smaller

products. Requiring that each Ax be a module over the stalk OX,x makes A into

an OX-module; indeed, the space of sections Γ(U,A) =∏

x∈U Ax is automatically

an Γ(U,OX)-module, the multiplication being defined componentwise with the

help of the stalk maps Γ(U,OX) → OX,x. Clearly this module structures is

compatible with the projections, and thus makes A into an OX-module.

Exercise 27. Assume that F and G are OX-modules and that α : F → G is a

map between them. Show that the kernel, cokernel and image of α as a map

of abelian sheaves indeed are OX-modules, and that they respectively are the

kernel, cokernel and image of α in the category of OX-modules as well. Show

that a complex of OX-modules is exact if and only it is exact as a complex of

abelian sheaves.

Exercise 28. Let F and G be two OX-modules on the scheme X. Show that

the stalk (F ⊗OX G)x at the point x ∈ X is naturally isomorphic to the tensor

product Fx⊗OX,x Gx of the stalks Fx and Gx. Show that the tensor product is

right exact in the category of OX-modules.

Exercise 29. Show that the sheaf-hom HomOX (F ,G) of two OX-modules as

defined above is a sheaf. Show that HomOX (F ,G) is right exact in the second

variable and left exact in the first.

Exercise 30 (Adjunction between Hom and ⊗). Show that for OX-modules

F ,G,HHomOX (F ,HomOX (G,H) ' HomOX (F ⊗OX G,H)

(this is easier than it looks – reduce to the usual tensor product for modules over

rings).

Exercise 31. Show that the category ModX has arbitrary products and coprod-

ucts, by showing that the products and coproducts in the category of abelian

sheaves ShX are OX-modules and are the products, respectively the coproducts,

in the category ModX .

Exercise 32. Assume that p1, . . . , pr is a set of primes, and let Z(pi) as usual

denote the localization at prime ideal (pi). Let X be the scheme obtained by

gluing the schemes Xi = SpecZ(pi) together along their common open subschemes

SpecQ. Describe the OX-modules on X.

Exercise 33. Let A =∏

1≤i≤nKi be the product of finitely many fields. Describe

the category ModX .

126

Page 127: Geir Ellingsrud and John Christian Ottem October 15, 2018€¦ · between algebra and geometry: a ne varieties correspond to nitely generated k-algebras without zero-divisors, and

Chapter 8. Quasi-coherent sheaves

8.1.1 The support of a sheaf

Definition 8.3. Let (X,OX) be a ringed space. Let F be a sheaf of OX-modules.

The support of F , Supp(F), is the set of points x ∈ X such that Fx 6= 0. For

s ∈ Γ(X,F) we define the support of s as the set of points x ∈ X such that the

image sx ∈ Fx of s is not zero. We denote this by Supp(s).

In general the support of a sheaf of modules is not closed. Indeed, as before,

we can get strange sheaves by taking any non-closed Z set of your favorite ringed

space, and define a Godement sheaf A by setting Ax 6= 0 if and only if x ∈ Z.

However, for sheaves of rings, the support always closed: this is because a

ring is 0 if and only if 1 = 0, and so the support of a sheaf of rings is the support

of the section ‘1’ - which is closed.

8.2 Pushforward and Pullback of OX-modules

8.2.1 Pushforward

Let f : (X,OX)→ (Y,OY ) be a morphism of schemes. If F is a sheaf on X the

pushforward f∗F is naturally a f∗OX-module via the addition and multiplication

maps

f∗F × f∗F → f∗F , f∗OX × f∗F → f∗F .

Via f# : OY → f∗OX we obtain the structure of a OY -module on f∗F .

Definition 8.4. The above OY -module is called the direct image of F under f .

This condition is clearly functorial in F and so we obtain a functor f∗ :

ModX → ModY . By the properties of f∗ in Chapter 1, we see that this is left

exact.

8.2.2 Pullback

The pullback of a sheaf of OY -modules is a little bit more subtle to define. Recall

that we in Chapter 1 defined the inverse image f−1G of a sheaf G by sheafifying

the presheaf given by assigning an open U to the direct limit of all G(V ) where V

contains f(U). This sheaf is naturally a f−1OY -module. We can make it into an

OX-module using the map f−1OY → OX (which makes OX an f−1OY -algebra),

and taking the tensor product:

f ∗G = OX ⊗f−1OY f−1G

The association G 7→ f ∗G is functorial, so we get a functor f ∗ : ModOY → ModOX .

The above OX-module is called the pullback of G under f . Note in particular

that f ∗OY = OX .

127

Page 128: Geir Ellingsrud and John Christian Ottem October 15, 2018€¦ · between algebra and geometry: a ne varieties correspond to nitely generated k-algebras without zero-divisors, and

8.2. Pushforward and Pullback of OX-modules

Proposition 8.5. For a point x ∈ X we have the following expression for the

stalk

(f ∗G)x = Gf(x) ⊗OY,f(x) OX,x.

Proof. This follows from the facts that taking stalks commutes with tensor

products, and (f−1G)x = Gf(x).

8.2.3 Adjoint properties of f∗, f∗

At first sight, the definition of the pullback might seem a bit out of the blue.

It is defined from f−1G, tensoring with OX over f−1OY to rig it into being a

OX-module. However, as like in the case of the inverse image functor f−1, the

important point is what the sheaf does, rather than how it is explicitly defined.

Proposition 8.6. The functors f∗, f∗ between ModOX ,ModOY are adjoint. In

particular, if F ∈ ModOX ,G ∈ ModOY , there is a functorial isomorphism

HomOX (f ∗G,F) ' HomOY (G, f∗F).

Proof. Suppose φ : G → f∗F is an OY -module homomorphism. Then by the

adjoint property of f∗ and f−1 (in the categories ShX and ShY ), we get a map

f−1G → F , which is f−1OY -linear. Now F is an OX-module, so we get a

OX-linear map f−1G ⊗f−1OY OX → G by the universal property of the tensor

product.

Conversely, let φ : f ∗G → F be OX-linear. There is a map f−1G → f ∗Gwhich is f−1OY -linear. Consequently there is a f−1OY -linear map f−1G → F .

This induces a OY -linear map G → f∗F by the earlier adjointness property of f∗and f−1.

In particular, we the maps idf∗G and idf∗F provide us with the canonical

maps

η : G → f∗f∗G, ν : f ∗f∗F → F

We already saw previously that f∗ is a left-exact functor. Now f ∗ is right-

exact, by general properties of adjoint functors.

Corollary 8.7. f ∗ is right exact and f∗ is left-exact.

8.2.4 Pullback of sections

We can also pull back sections of G. If G is an OY -module, and s ∈ Γ(V,G),

then we get a section f ∗(s) = η(s) ∈ Γ(f−1(V ), f ∗G) by the map η : G → f∗f∗G.

If we think of sections of G as local maps from X into the various stalks, the

section f ∗s simply associates to each x ∈ X the germ sy ⊗ 1 ∈ Gy ⊗Oy OX,x for

y = f(x).

128

Page 129: Geir Ellingsrud and John Christian Ottem October 15, 2018€¦ · between algebra and geometry: a ne varieties correspond to nitely generated k-algebras without zero-divisors, and

Chapter 8. Quasi-coherent sheaves

8.2.5 Tensor products

From its definition, it is straightforward to check that applying f ∗ commutes

with taking tensor products of sheaves. On the level of presheaves, we have

f ∗(G ⊗H) = f−1(G ⊗OY H)⊗f−1OY OX

= (f−1G ⊗f−1OY f−1H)⊗f−1OY OX

= (f−1G ⊗f−1OY OX)⊗OX ⊗(f−1H)⊗f−1OY OX)

= f ∗G ⊗OX f∗H.

and sheafifying, we get an isomorphism of the corresponding sheaves.

However, the pushforward f∗ rarely commutes with taking tensor products of

sheaves (we will see several examples of this later). There is however, at least, a

map f∗(F)⊗ f∗(G)→ f∗(F ⊗ G) for for OX-modules F ,G: If U ⊆ Y is an open,

and s ∈ f∗(F), t ∈ f∗(G), then s ⊗ t is an element of F ⊗ G over f−1(U), and

hence s⊗ t defines a section of f∗(F ⊗ G) over U .

8.3 Quasi-coherent sheaves

In The Oxford English Dictionary there several nuances of the word coherent

are given, but the one at top is:

That sticks or clings firmly together; esp. united by the force of cohesion. Said

of a substance, material, or mass, as well as of separate parts, atoms, etc.

So the coherence of a sheaf should mean that there are some strong relations

between the sections over different open sets, at least over sets from some

sufficiently large collection of open sets. In our context the open affine subsets

stand out as obvious candidates to form such a collection, and indeed, a quasi-

coherent sheaf on the scheme X will turn out1 to have the following coherence

property: If V ⊆ U are two affine open sets in X there is a canonical map

Γ(U,F)⊗Γ(U,OX) Γ(V,OV )→ Γ(V,F) (8.3.1)

sending s⊗ f to ρUV (s) · f , where ρUV as usual indicates the restriction maps

in OX . The salient point is that for a quasi-coherent sheaf F this map is an

isomorphism. In fact the converse holds true as well, so the map in (8.3.1) being

an isomorphism for all pairs V ⊆ U of open affine sets is equivalent to F being

quasi-coherent.

To pin down the quasi-coherent sheaves, one first establishes a collection

of model-sheaves on affine schemes. To each A-module M one associates an

1One might have used this coherence property as the definition, but because of obscurereasons we choose another definition.

129

Page 130: Geir Ellingsrud and John Christian Ottem October 15, 2018€¦ · between algebra and geometry: a ne varieties correspond to nitely generated k-algebras without zero-divisors, and

8.4. Quasi-coherent sheaves on affine schemes

OX-module M on X = SpecA, and on a general scheme X, a quasi-coherent

module F is going to be one that locally is of the form M .

The coherent modules are the quasi-coherent modules that satisfy a certain

finiteness condition that in the noetherian case boils down to the OX-module

being finitely generated. These coherent OX-modules are the OX-modules most

frequently encountered in algebraic geometry and hence they have gotten the

shortest name, even though it is being quasi-coherent that means satisfying a

coherence property.

8.4 Quasi-coherent sheaves on affine schemes

For each A-module M we shall exhibit an OX-module M ; the construction of

which completely parallels what we did when constructing the structure sheaf

OX on X = SpecA. Letting B again be the base of the topology of distinguished

open subsets on X, we can define a B-presheaf M as follows:

(M)(D(f)) = Mf

The restriction maps are defined as before: D(g) ⊆ D(f) we have a canonical

localization map Mf →Mg. The same proof as for OX shows that this is actually

a B-sheaf, and hence gives rise to a unique sheaf M on X. By the properties of

sheafification, this is functorial in M : A homomorphism φ : M → N gives rise

to a canonical morphism of OX-modules M → N .

For any A-module homomorphism α : M → N there is an obvious way

of obtaining an OX-module homomorphism α : M → N ; indeed, the maps

α⊗ idΓ(U ;OX) are Γ(U,OX)-homomorphisms from Γ(U, M) to Γ(U, N) compatible

with the restrictions, and thus induce a map between M and N . The map α is

the associated map between the sheafifications. Clearly one has φ ψ = φ ψ,

and the “tilde-operation” is therefore a functor ModA → ModOX .

The three main properties of M are listed below. They are completely

analogous to the statements in proposition 2.36 on page 56 in Chapter 2 about

the structure sheaf OX ; and as well, the proofs are mutatis mutandis the same.

• Stalks: Let x ∈ SpecA be a point whose corresponding prime ideal is p.

The stalk Mx of M at x ∈ X is Mx = Mp = M ⊗AAp.

• Sections over distinguished open sets: If f ∈ A one has Γ(D(f), M) =

Mf = M ⊗AAf . In particular it holds true that Γ(X, M) = M .

• Sections over arbitrary open sets. For any open subset U of SpecA covered

130

Page 131: Geir Ellingsrud and John Christian Ottem October 15, 2018€¦ · between algebra and geometry: a ne varieties correspond to nitely generated k-algebras without zero-divisors, and

Chapter 8. Quasi-coherent sheaves

by the distinguished sets D(fi) i∈I , there is an exact sequence

0 // Γ(U, M) α //∏

iMfi

ρ //∏

i,jMfifj . (8.4.1)

These statements follow formally from the construction. The first follows

since both the stalks Mx and the localizations Mp are direct limits of the same

modules over the same inductive system (indexed by the distinguished open

subsets D(f) containing x); the second follows from the way we defined M ; and

the third is just the general exact sequence expressing the space of sections of a

sheaf over an open set in terms of the space of sections over members of an open

covering.

Lemma 8.8. For any two A-modules M and N , the association φ → φ is an

isomorphism HomA(M,N) ' HomOX (M, N).

In the lingo of category theory the lemma is expressed by saying that the

tilde-operation is a fully faithful functor (trofast, full funktor), fully meaning

the map in the lemma is surjective and faithful that it is injective. Loosely

speaking, the “tilde-operation” gives an isomorphism of ModA with a subcategory

of ModOX . It is a strict subcategory; most of the OX-modules are not of the

form M , but those that are, form the category of quasi-coherent OX-modules,

that we shortly return to.

Assume that an OX-module F is given on X = SpecA, and let M denote

the global sections of F ; that is, M is the A-module M = Γ(X,F). There is

a natural map M → F of OX-modules. As usual, it suffices to tell what the

map does to sections over members of a basis for the topology, e.g., over open

distinguished sets. The sheaf F being an OX-module, multiplication by f−1

in the space of sections Γ(D(f),F) has a meaning since Γ(D(f),OX) = Af .

Hence we may send the section mf−n ∈Mf of M to the section of F over D(f)

obtained by multiplying the restriction of m to D(f) by f−n; i.e., we send m to

f−n ·m|D(f) . For later reference we state this observation as a lemma, leaving

the task of checking the details to the zealous student:

Lemma 8.9. Given a quasi-coherent sheaf F on the affine scheme X = SpecA.

Then there is a unique OX-module homomorphism

θF : Γ(X,F ) → F

inducing the identity on the spaces of global sections. Moreover, it is natural in the

sense that if α : F → G is a map of OX-module inducing the map a : Γ(X,F)→Γ(X,G) on global sections, one has θG a = α θF .

Exercise 34. Check that this is a well defined map (there is a choice involved

in the definition).

131

Page 132: Geir Ellingsrud and John Christian Ottem October 15, 2018€¦ · between algebra and geometry: a ne varieties correspond to nitely generated k-algebras without zero-divisors, and

8.4. Quasi-coherent sheaves on affine schemes

Lemma 8.10. In the canonical identification of the distinguished open subset

D(f) with SpecAf , the OX-module M restricts to Mf .

Proof. As Γ(D(f), M) = Mf , there is a map Mf → M |D(f) that on distinguished

open subsets D(g) ⊆ D(f) induces an isomorphism between the two spaces of

sections, both being equal to the localization Mg.

The “tilde-functor” is really a no-nonsense functor having almost all properties

one can desire. It is fully faithful, as we have seen, but it also is an additive as

well as an exact functor. Moreover, it takes the tensor product M ⊗AN of two

A-modules to the tensor product M ⊗OX N of the corresponding OX-modules,

and if M is of finite presentation, the A-module of homomorphisms HomA(M,N)

to the sheaf of homomorphisms HomOX (M, N). However, this is not true if M

is not of finite presentation, the only lacking desirable property of the functor

(−).

To verify the first of these claims, assume given an exact sequence of A-

modules:

0 //M ′ //M //M ′′ // 0.

That the induced sequence of OX-modules

0 // M ′ // M // M ′′ // 0

is exact is a direct consequence of the three following facts. The stalk of a tilde-

module M at the point x with corresponding prime ideal p is Mp, localization is

an exact functor, and finally, a sequence of abelian sheaves is exact if and only if

the sequence of stalks at every point is exact.

For the tensor product, let T denote the presheaf U 7→ M(U)⊗OX(U) N(U).

We have the map on presheaves T → M ⊗A N (on U = D(f) it is the isomor-

phism Mf ⊗Af Nf ' (M ⊗AN)f ; it sends m/fa⊗ n/f b to (m⊗ n)/fa+b). After

sheafifying, we get a the desired map of sheaves

M ⊗OX N → M ⊗A N.

This is an isomorphism, since it is an isomorphism over every stalk.

When it comes to hom’s however, the situation is somehow more subtle. If

M is not of finite presentation, it is not true that HomA(M,N) localizes. There

is always a canonical map

HomA(M,N)⊗AAf → HomAf (Mf , Nf ), (8.4.2)

but without some finiteness condition (like being of finite presentation) on M ,

it is not necessarily an isomorphism. Even in the simplest case of a infinitely

132

Page 133: Geir Ellingsrud and John Christian Ottem October 15, 2018€¦ · between algebra and geometry: a ne varieties correspond to nitely generated k-algebras without zero-divisors, and

Chapter 8. Quasi-coherent sheaves

generated free module M =⊕

i∈I Aei that map is not surjective. An element

in HomA(M,N)⊗AAf is of the form α⊗ f−n. An element in HomAf (Mf , Nf),

however, is given by its values mi⊗ f−ni on the free generators ei, and the salient

point is that the ni’s may tend to infinity, and no n working for all i’s can be

found.

Exercise 35. Show that the map (8.4.2) above is an isomorphism when M is of

finite presentation. Hint: First observe that this holds when M = A. Then use

a presentation of M to reduce to that case.

IfM is of finite presentation, the global sections of the sheaf-hom HomOX (M, N)

equals HomA(M,N), and there is a map

HomA(M,N)˜→HomOX (M, N).

In case M is of finite presentation, the maps in (8.4.2) are isomorphisms, and

the map above induces isomorphisms between the spaces of sections of the two

sides over any distinguished open subset. One concludes that the map is an

isomorphism, and one has HomA(M,N)˜'HomOX (M, N).

In short, we have established the important proposition:

Proposition 8.11. Assume that A is a ring and let X = SpecA. The functor

from the category ModA of A-modules to the category ModOX of OX-modules

given by M → M enjoys the following three properties

• It is a fully faithful additive and exact functor.

• One has a canonical isomorphism (M ⊗AN ) ' M ⊗OX N .

• If M is of finite presentation, one has a canonical isomorphism HomA(M,N)˜'HomOX (M, N).

Exercise 36. Show that the tilde-functor is additive; i.e., takes direct sums of

modules to the direct sum and sums of maps to the corresponding sums.

Example 8.12. Assume that A is an integral domain and that K is the field of

fractions of A. Show that the OX-module K is a constant sheaf in the strong

sense, that is Γ(U, K) = K for any non-empty open U ⊆ X and the restriction

maps all equal the identity idK .

8.4.1 Functoriality

Suppose we are given a morphism between the two affine schemes X and Y ; say

φ : X → Y . We let X = SpecA and Y = SpecB and we let φ# : B → A be the

map corresponding to φ.

133

Page 134: Geir Ellingsrud and John Christian Ottem October 15, 2018€¦ · between algebra and geometry: a ne varieties correspond to nitely generated k-algebras without zero-divisors, and

8.5. Quasi-coherent sheaves on general schemes

If M is an A-module, can one describe the sheaf φ∗M on Y ? The answer

is not only yes but the description is very simple. The A-module M can be

considered a B-module via the map B → A and as such is denoted by MB.

Clearly this is a functorial construction in M . In this setting one has

Proposition 8.13. One has φ∗M = MB.

Proof. There is an obvious map MB → φ∗(M) as Γ(Y, φ∗(M)) = Γ(X, M) = M .

The argument that follows shows it is an isomorphism, and it is as usual sufficient

to verify that sections over distinguished open sets of the two sides coincide. The

crucial observation is that φ−1D(f) = D(φ#f)—indeed, a prime ideal p ⊆ A

satisfies f ∈ φ#−1(p) if and only if φ#(f) ∈ p. As f acts on MB as multiplication

by φ#(f), one clearly has (MB)f = Mφ#(f) = Γ(D(φ#(f)), M).

8.5 Quasi-coherent sheaves on general schemes

Having established the sheaves on affine space that serve as local models for the

quasi-coherent sheaves, we are now ready for the general definition.

Definition 8.14. If X is a scheme and F an OX-module, one says that F is

quasi-coherent (kvasikoherent) OX-module, or quasi-coherent sheaf for short, if

there is an open affine covering Ui i∈I of X, say Ui = SpecAi, and modules

Mi over Ai such that F|Ui ' Mi.

Phrased in slightly different manner, OX-module F is quasi-coherent if the

restriction F|Ui of F to each Ui is of type tilde of an Ai-module. In particular,

the modules M on affine schemes SpecA are all quasi-coherent.

The restriction of a quasi-coherent sheaf F to any open set U ⊆ X is quasi-

coherent. Indeed, it will suffices to verify this for X an affine scheme, and by

lemma 8.10 the restriction of a sheaf of tilde-type to a distinguished open set is

of tilde-type. As any open U in an affine scheme is the union of distinguished

open subsets, it follows that F|U is quasi-coherent.

For F to be quasi-coherent, we require that F be locally of tilde-type for just

one open affine cover. However, it turns out that this will hold for any open

affine cover, or equivalently, that F|U is of tilde-type for any open affine subset

U ⊆ X. This is a much stronger than the requirement in the definition, and it

is somehow subtle to prove. As a first corollary we arrive at the a priori not

obvious conclusion that the modules of the form M are the only quasi-coherent

OX-modules on an affine scheme. We shall also see that quasi-coherent modules

enjoy the coherence property (8.3.1) on page 129 that was the point of departure

for our discussion.

134

Page 135: Geir Ellingsrud and John Christian Ottem October 15, 2018€¦ · between algebra and geometry: a ne varieties correspond to nitely generated k-algebras without zero-divisors, and

Chapter 8. Quasi-coherent sheaves

The story starts with a lemma that establishes the coherence property (8.3.1)

in a very particular case; i.e., for sections over distinguished open sets of a quasi-

coherent OX-module on an affine scheme X = SpecA. For any distinguished

open set D(f) ⊆ X it holds that Γ(D(f),OX) = Af , and consequently there

is for any OX-module a canonical map Γ(X,F)⊗AAf → Γ(D(f),F) sending

s⊗ af−n to af−n · s|D(f). It turns out to be an isomorphism whenever F is

quasi-coherent:

Lemma 8.15. Suppose that X = SpecA is an affine scheme and that F is a

quasi-coherent OX-module. Let D(f) ⊆ X be a distinguished open set. Then one

has

• Γ(D(f),F) ' Γ(X,F)⊗AAf .

• Let s ∈ Γ(X,F) be a global section of F and assume that s|D(f) = 0, then

sufficiently big powers of f kill s, that is, for sufficiently big integers n one

has fns = 0.

• Let s ∈ Γ(D(f),F) be a section. Then for a sufficiently large n, the section

fns extends to a global section of F . That is, there exists an n and a global

section t ∈ Γ(X,F) such that t|D(f) = fns

Proof. The first statement is by the definition of localization equivalent to the

two others.

The sheaf F being quasi-coherent by hypothesis, and the affine scheme

X = SpecA being quasi-compact, there is a finite open affine covering of X by

distinguished sets D(gi) such that F|D(gi) ' Mi for some Agi-modules Mi. The

section s of F restricts to sections si of F|D(gi) over D(gi), that is, to elements

si of Mi.

Further restricting F to the intersections D(f) ∩D(gi) = D(fgi) yields the

equality F|D(fgi) = (Mi)f , and by hypothesis, the section s restricts to zero in

Γ(D(fgi),F) = (Mi)f . This means that the localization map sends si to zero

in (Mi)f . Hence si is killed by some power of f , and since there is only finitely

many gi’s, there is an n with fnsi = 0 for all i; that is, (fns)|D(gi) = 0 for all i.

By the locality axiom for sheaves, it follows that fns = 0.

Assume now a section s ∈ Γ(D(f),F) is given. We are to see that fns extends

to a global section of F for large n. Each restriction s|D(fgi) ∈ Γ(D(fgi),F) =

(Mi)f is of the form f−nsi with si ∈Mi = Γ(D(gi),F), and by the usual finiteness

argument, n can be chosen uniformly for all i. This means that si = fns and

sj = fns mach on the intersection D(f)∩D(gi)∩D(gj), and by the first part of

the lemma applied to SpecAgigj , one has fN (si − sj) = 0 on D(gi) ∩D(gj) for a

sufficiently large integers N . Hence the different fNsi’s patch together to give

the desired global section t of F .

135

Page 136: Geir Ellingsrud and John Christian Ottem October 15, 2018€¦ · between algebra and geometry: a ne varieties correspond to nitely generated k-algebras without zero-divisors, and

8.5. Quasi-coherent sheaves on general schemes

Theorem 8.16. Let X be a scheme and F an OX-module. Then F is quasi-

coherent if and only if for all open affine subsets U ⊆ X = SpecA, the restriction

F|U is isomorphic to a an OX-module of the form M for an A-module M .

Proof. As quasi-coherence is conserved when restricting OX-modules to open sets,

we may surely assume that X itself is affine; say X = SpecA. Let M = Γ(X,F).

We saw in lemma 8.9 on page 131 that there is a natural map M → F that

on distinguished open sets sends mf−n to f−nm|U , but by the fundamental

lemma 8.15 above, this is an isomorphism between the spaces of sections over

the distinguished open sets. Hence the two sheaves are isomorphic.

Applying this to an affine scheme, yields the important fact that any quasi-

coherent sheaf F on an affine scheme X = SpecA is of the form M for an

A-module M .

Proposition 8.17. Assume that X = SpecA. The tilde-functor M 7→ M is an

equivalence of categories ModA and QCohX with the global section functor as an

inverse.

When speaking about mutually inverse functors one should be very careful;

in most cases such a statement is an abuse of language. Two functors F and

G are mutually inverses when there are natural transformations, both being an

isomorphisms, between the compositions F G and G F and the appropriate

identity functors. In the present case one really has an equality Γ(X, M) = M ,

so that Γ (−) = idModA . On the other hand, the natural transformation

Γ(X,F ) → F from lemma 8.9 on page 131 furnishes the required isomorphism

of functors.

Theorem 8.18. Let X be a scheme and let F be an OX-module on X. Then

F is quasi-coherent of and only if for any pair V ⊆ U open affine subsets, the

natural map

Γ(U,F)⊗Γ(U,OX) Γ(V,OX)→ Γ(V,F) (8.5.1)

is an isomorphism.

Proof. We may clearly assume that X is affine, say X = SpecA. Assume that

the maps (8.5.1) are isomorphisms. We may take V = D(f) and U = X and

M = Γ(X,F). Then from (8.5.1) it follows that Γ(D(f),F) = Mf which shows

that the canonical map M → F is an isomorphism over all distinguished open

subsets, and therefore an isomorphism. Hence F is quasi-coherent.

To argue for the implication the other way, assume that U = SpecB and that

F is quasi-coherent; that is, F = M for some A-module M after theorem 8.18.

The restriction of a quasi-coherent sheaf is quasi-coherent, so M |U = N for some

B-module N , and the map in (8.5.1) is just a map have map M ⊗AB → N . It

136

Page 137: Geir Ellingsrud and John Christian Ottem October 15, 2018€¦ · between algebra and geometry: a ne varieties correspond to nitely generated k-algebras without zero-divisors, and

Chapter 8. Quasi-coherent sheaves

induces an isomorphism over all local rings Ap = Bp (where p ∈ U = SpecB)

since M |U ' N and therefore is an isomorphism of B-modules.

Example 8.19. The example of an discrete valuation ring is always useful to

consider, and we continue exploring example 8.1 above. The OX module given

by the data M,N, ρ. We claim that F is quasi-coherent if and only if ρ is an

isomorphism.

If F is quasi-coherent, then every point has a neighbourhood on which Fis the˜of some module. The only neighbourhood of the unique closed point is

X itself, and so F = M . Therefore, N = F(U) = M(0) = M ⊗R K and ρ is an

isomorphism. Conversely, of ρ : M ⊗R K → K is an isomorphism, then F is

given by F(X) = M and F( η ) = M ⊗R N , and so F ' M is quasi-coherent.

8.6 Coherent sheaves

Let A be a ring and let M be an A-module. The module M is of finite presentation

if for some integers n and m there is an exact sequence

An // Am //M // 0 .

One says that M is coherent if the following two requirements are fulfilled

• M is finitely generated.

• The kernel of every surjection An →M is finitely presented.

The second statement is equivalent to every finitely generated submodule N ⊆M

being of finite presentation. In the case A is a noetherian ring—which frequently

is the case in algebraic geometry—a module M being coherent is equivalent

to M being finitely generated. The condition comes from the theory analytic

functions where coherent non-noetherian rings are frequent.

When A is noetherian, then the three conditions of coherence, finitely gen-

eration and being of finite presentation on an A-module M coincide. The key

point is that for a finitely generated module M over a noetherian ring A, every

submodule N ⊆M is also finitely generated. Indeed, the set V of finitely gener-

ated submodules N ′ ⊆ N has a maximal element N, which has to equal N : If

not, there is an n ∈ N −N, and a finitely generated submodule N ′′ = N ′ +An

which is strictly bigger than N.

So to show that M is finitely presented, we can take a presentation Am →M → 0 and let L be the kernel, regarded as a submodule of An. Then L is

finitely generated, so there is a surjection An → L, and we get a presentation

An → Am → M → 0. Applying this argument to any finitely generated

submodule N ⊆M shows that M is coherent.

137

Page 138: Geir Ellingsrud and John Christian Ottem October 15, 2018€¦ · between algebra and geometry: a ne varieties correspond to nitely generated k-algebras without zero-divisors, and

8.7. Functoriality

Exercise 37. Fill in the details of the proof that if A is a noetherian ring,

then the three conditions of coherence, finitely generation and being of finite

presentation on an A-module M coincide.

Definition 8.20. On a scheme X a quasi-coherent OX-module is coherent if

there is a covering of X by open affine sets Ui = SpecAi such that F|Ui = Mi

with the Mi’s being coherent Ai-modules. F is finitely presented if each Mi are

finitely presented as Ai-modules.

So if X is noetherian (or locally noetherian), the condition that M is coherent

is equivalent to the weaker condition that M is finitely generated.

Remark 8.21. The definition above is the one that appears in EGA and the

Stacks project. However, it differs slightly from the notation used in Hartshorne’s

book. In that book, one says that F is coherent if there is a covering by open

affines Ui = SpecAi such that F|Ui = Mi with the Mi’s being finitely generated

Ai-modules (that is, the condition about kernels is left out). In the case X is

(locally) noetherian, these two notions coincide, but they are not equivalent in

general.

The notion of coherent sheaf was actually introduced by Henri Cartan in

the theory of holomorphic functions of several varables around 1944. In 1946

Oka proved that OCn is coherent and this is a very difficult theorem (although it

seems like a trivial statement if the other definition of ‘coherence’ is used).

One benefit of using coherent modules rather than finitely generated ones

is that the category of coherent modules is an abelian category, even in the

non-noetherian setting. However, a problem is that coherence is very difficult to

check in general and actually for some schemes, even affine ones, the structure

sheaf OX is not coherent!

8.7 Functoriality

Proposition 8.22. Suppose that α : F → G is a map of quasi-coherent sheaves

on the scheme X. The kernel, cokernel and the image of α are all quasi-coherent.

The category QCohX is closed under extensions; that is, if

0 //M ′ //M //M ′′ // 0 (8.7.1)

is a short exact sequence of OX-modules with the two extreme sheaves M ′ and

M ′′ being quasi-coherent, the middle sheaf M is quasi-coherent as well.

Proof. If α : F → G is a map of quasi-coherent OX-modules, on any open affine

subsets U = SpecA of X it may be described as α|U = a where a : M → N

is a A-module homomorphism and M and N are A-modules with F|U = M

138

Page 139: Geir Ellingsrud and John Christian Ottem October 15, 2018€¦ · between algebra and geometry: a ne varieties correspond to nitely generated k-algebras without zero-divisors, and

Chapter 8. Quasi-coherent sheaves

and G|U = N . Since the tilde-functor is exact, one has kerα|U = (ker a) .

Moreover, by the same reasoning, it holds true that cokerα|U = (coker a) and

imα|U = (im a) .

Suppose now that an extension like (8.7.1) is given. The leftmost sheaf M ′

being quasi-coherent lemma 13.4 entails that the induced sequence of global

sections is exact; that is, the upper horizontal sequence in the diagram below.

The three vertical maps in the diagram are the natural maps from lemma 8.9 on

page 131. Since M ′ and M ′′ both are quasi-coherent sheaves, the two flanking

vertical maps are isomorphisms, and the snake lemma implies that the middle

vertical map is an isomorphism as well. Hence M is quasi-coherent.

0 // Γ(X,M ′) //

Γ(X,M ) //

Γ(X,M ′′) //

0

0 //M ′ //M //M ′′ // 0

The category QCohX is an abelian category with tensor products and internal

hom’s.

8.7.1 Quasi-coherence of pullbacks

Recall the notion of pullback of a sheaf via a morphism f : X → Y . This

is a relatively complicated operation, since it involves taking a direct limit, a

tensor product, and finally a sheafification. The next result tells us that for G a

quasi-coherent sheaf on Y , we have a much simpler description of the pullback

f ∗G, which will allow us to do local computations more easily.

Theorem 8.23. Let f : X → Y be a morphism of schemes.

(i) If X = SpecB, Y = SpecA, and M is an A-module, then

f ∗(M) = (M ⊗A B)

(ii) If G is a quasi-coherent sheaf on Y , then f ∗G is quasi-coherent on X.

(iii) If X and Y are noetherian, then f ∗G is coherent if G is.

Proof. (ii) and (iii) follows by (i), since quasi-coherence is a local property, and

since M ⊗A B is a finitely generated B if M is a finitely generated A-module.

So let us proceed to prove the first point.

First, note that the theorem holds in the special case when M = AI is a free

module (here the index set I is allowed to be infinite) – this is simply because

139

Page 140: Geir Ellingsrud and John Christian Ottem October 15, 2018€¦ · between algebra and geometry: a ne varieties correspond to nitely generated k-algebras without zero-divisors, and

8.7. Functoriality

f ∗OY = OX and f ∗ commutes with taking direct sums. To prove it in general,

we pick a presentation of M of the form

AJγ−→ AI →M → 0

Applying ∼ and then f ∗ we get a sequence

f ∗AJν−→ f ∗AI → f ∗M → 0

which is exact on the right, since f ∗ is right-exact. From this we get that

f ∗M = coker ν = coker ˜(γ ⊗A B)

= ((coker γ)⊗A B))∼ = (M ⊗A B)∼.

It is instructive to see the first point via the adjoint property of f∗ and f ∗. If

F = M for a B-module M and G = N for an A-module N , we have the following

isomorphisms induced by the adjoint property of Hom and ⊗:

HomX(f ∗G,F) = HomY (f ∗N , M)

= Hom((M ⊗A B)∼, N)

= HomA(M ⊗A B,N)

= HomB(M,HomA(B,N))

= HomB(M,NB)

= HomY (M, NB)

= HomY (G, f∗F)

8.7.2 Quasi-coherence of the direct image

Recall that we showed that for a map φ : SpecA → SpecN , the pushforward

φ∗F is quasi coherent, of F is quasi-coherent (since φ∗M = MB). The following

result applies to more general morphisms:

Theorem 8.24. Let φ : X → Y be a morphism of schemes and that F is a

quasi-coherent sheaf on X. If X is noetherian, then the direct image φ∗F is

quasi-coherent on Y .

Proof. We may assume that Y = SpecA. Then since X is quasi-compact, we

may cover it by open affines Ui. The intersection Ui ∩Uj is again quasi-compact,

so we can cover it with open affines Uijk.

140

Page 141: Geir Ellingsrud and John Christian Ottem October 15, 2018€¦ · between algebra and geometry: a ne varieties correspond to nitely generated k-algebras without zero-divisors, and

Chapter 8. Quasi-coherent sheaves

For any open V ⊆ Y , one has the exact sequence

0 // Γ(φ−1V,F) //∏

i Γ(Ui ∩ φ−1V,F) //∏

i,j,k Γ(Uijk ∩ φ−1V,F).

(8.7.2)

The sequence is compatible with the restriction maps induced from an inclusion

V ′ ⊆ V , hence gives rise to the following exact sequence of sheaves on X:

0 // φ∗F //∏

i φi∗F|Ui //∏

i,j,k φijk∗F|Uijk (8.7.3)

where φi = φ|Ui and φijk = φ|Uijk . Now, each of the sheaves φi∗F|Ui and φij∗F|Uijare quasi-coherent by the affine case of the theorem (proposition 8.13 on page

134). They are finite in number as the covering Ui is finite. Hence∏

i φi∗F|Ui and∏i,j φij∗F|Uij are finite products of quasi-coherent OX-modules and therefore

they are quasi-coherent. Now the φ∗F is the kernel of a homomorphism between

two quasi-coherent sheaves, and so the theorem the follows from proposition 8.22

on page 138.

The following example shows that some of the hypotheses are neccessary for

this statement to hold:

Example 8.25. Let X =∐

i∈I SpecZ be the disjoint union of countably in-

finitely many copies of SpecZ and let φ : X → SpecZ be the morphism that

equals the identity on each of the copies of SpecZ that constitute X. Then

φ∗OX is not quasi-coherent. Indeed, the global sections of φ∗OX satisfy

Γ(SpecZ, φ∗OX) = Γ(X,OX) =∏i∈I

Z.

On the other hand if p is any prime, one has

Γ(D(p), φ∗OX) = Γ(φ−1D(p),OX) =∏i∈I

Z[p−1].

It is not true that Γ(D(p), φ∗OX) = Γ(SpecZ, φ∗OX)⊗Z Z[p−1] hence φ∗OX is

not quasi-coherent. Indeed, elements in∏

i∈I Z[p−1] are sequences of the form

(zip−ni)i∈I where zi ∈ Z and ni ∈ N. Such an element lies in (

∏i∈I)⊗Z Z[p−1])

only if the ni’s form a bounded sequence, which is not the case for general

elements of shape (zip−ni)i∈I when I is infinite.

Exercise 38. Show that X =⋃i∈I SpecZ is not affine. Show that in the

category Sch of schemes X is the coproduct of the infinitely many copies of

SpecZ involved . Show that Spec∏

i Z is the coproduct of the copies of SpecZin the category Aff of affine schemes. Show that these gadgets are substantially

different.

141

Page 142: Geir Ellingsrud and John Christian Ottem October 15, 2018€¦ · between algebra and geometry: a ne varieties correspond to nitely generated k-algebras without zero-divisors, and

8.8. The categories of coherent and quasi-coherent sheaves

8.7.3 Coherence of the direct image

For morphisms of schemes f : X → Y , it is not expected that the pushforward

of a coherent sheaf is again coherent, even for ‘nice’ morphisms f . A simple

example is the following:

Example 8.26. Let X = Spec k[t] and consider the structure morphism f :

X → Spec k (induced by k ⊆ k[t]). The sheaf OX is of course coherent, but

f∗OX is not. Indeed, this is k[t], and k[t] is clearly not finitely generated as a

k-module.

However, for finite morphisms, we have a positive result:

Lemma 8.27. Let f : X → Y be a finite morphism of schemes. If F be a

quasi-coherent sheaf on X, then f∗F is quasi-coherent on Y . If X and Y are

noetherian, f∗F is even coherent if F is.

Proof. Since f is finite, we can cover Y by open affines SpecA such that each

f−1 SpecA = SpecB is also affine, where B is a finite A-module. We then have

f∗F(SpecA) = F(SpecB). Now, since F is quasicoherent, we have F|SpecB = M

for some B-module, which we can view as an A-module via f . Hence f∗F is

quasi-coherent. If X and Y are noetherian, and F is coherent, the module M is

finitely generated as an B-module, and hence as an A-module, since B is a finite

A-module.

8.8 The categories of coherent and quasi-coherent

sheaves

Our work in the previous sections imply the following statement

Theorem 8.28. The category QCoh(X) is an abelian category and is closed

under direct limits.

By definition, being an abelian category entails that the hom-sets Hom(F ,G)

are abelian groups; finite direct sums exist; kernels and cokernels of morphisms

exist; every monomorphism F → G is the kernel of its cokernel; every epimor-

phism is the cokernel of its kernel; and every morphism can be factored into an

epimorphism followed by a monomorphism. The hard part is thus in the last

part of the statement, that any direct limit of quasi-coherent sheaves is again

quasi-coherent.

One reason why we prefer the notion of ‘coherence’ used here (rather than

the one in Harthorne) is that the category of coherent sheaves Coh(X) is also an

abelian category, even in the non-noetherian case. Note that it does not contain

142

Page 143: Geir Ellingsrud and John Christian Ottem October 15, 2018€¦ · between algebra and geometry: a ne varieties correspond to nitely generated k-algebras without zero-divisors, and

Chapter 8. Quasi-coherent sheaves

all its direct limits, simply because an arbitrary product of coherent A-modules

is typically not coherent (not even finitely generated!)

Coherent sheaves can still be regarded as the building blocks of the category

QCoh(X). In fact, a common technique is to prove statements about quasi-

coherent sheaves by approximating them with coherent sheaves. This is justified

by the following

Theorem 8.29. Any quasi-coherent sheaf on a noetherian scheme is the direct

limit of its coherent subsheaves.

We will not go into details about this statement here, but remark that the

proof is not too difficult (see EGA I, Section 6.9).

8.9 Closed immersions and subschemes

If A is a ring and I is an ideal of A, we have seen that there is a morphism of

schemes f : Spec(A/I) → SpecA, inducing an homeomorphism of Spec(A/I)

onto the closed subset V (I). In other words, f is what’s known as a closed

immersion of topological spaces. Conversely, such a morphism f : SpecB →SpecA which is an homeomorphism onto a closed subset, the induced map

on rings φ : A → B is surjective and SpecB is isomorphic to the scheme

Spec(A/ kerφ).

The aim of this section is to globalize these observations for general schemes.

This will involve replacing ideals with (quasi-coherent) ideal sheaves.

Lemma 8.30. Let X be a scheme and let I ⊆ OX be a quasi-coherent sheaf

of ideals. Then the ringed space Z = (Supp(OX/I),OX/I) is a scheme with a

canonical morphism i : Z → X.

Indeed, to prove this, we may assume that X = SpecA is affine. In this case

I is the ∼ of some ideal I ⊆ A, and the support of this is exactly the primes p

such that (A/I)p 6= 0, or equivalently p ∈ V (I). Hence Z is the closed subset

V (I), which is homeomorphic to Spec(A/I). The sheaf of rings on Spec(A/I) is

the same as OX/I on Z and hence Z is the scheme SpecA/I. The map i is just

induced by the inclusion Z ⊆ X and the natural map OX → i∗(OX/I) is just ∼of the quotient map A→ A/I.

Definition 8.31. A closed subscheme of a schemeX is a scheme (Supp(OX/I),OX/I)

for some quasi-coherent sheaf of ideals I on X

Definition 8.32. A morphism of schemes f : Y → X is called an closed

immersion if there is an isomorphism Y ' Z onto a closed subscheme Z =

143

Page 144: Geir Ellingsrud and John Christian Ottem October 15, 2018€¦ · between algebra and geometry: a ne varieties correspond to nitely generated k-algebras without zero-divisors, and

8.9. Closed immersions and subschemes

(Supp(OX/I),OX/I) for some quasi-coherent sheaf of ideals I on X, with a

commutative diagram

Yf //

X

Zi

>>

If φ : A→ B is a surjective homomorphism of rings, then SpecB → SpecA

is a closed immersion. Conversely, we have the following:

Lemma 8.33. Let f : Z → X be a closed immersion. If X is affine, then so is

Z and in fact Z ' SpecA/I for some ideal I of A.

Indeed, we can view Z as a closed subscheme of X = SpecA, with the locally

ringed structure being given by OX/I for some quasi-coherent sheaf of ideals

I on X. We have seen that this is the sheaf associated to an ideal I of A.

Since Z = Supp(OX/I), we have that Z = V (I) and in fact Z is isomorphic to

Spec(A/I), because the sheaves of rings are isomorphic.

Proposition 8.34. Let f : Y → X be a morphism of schemes. Then f is a

closed immersion if and only if f is a homeomorphism of the topological space of

Y onto a closed subset of X and f# : OX → f∗OY is surjective.

Proof. A closed immersion by a quasi-coherent sheaf of ideals clearly satisfies

the two conditions, so we need only prove the ‘if’ direction.

We want to associate to f a quasi-coherent ideal sheaf I on X. Since

OX → f∗OY is surjective, it is natural to take the kernel of this map. However,

we have to work a bit to prove that this is actually quasi-coherent.

Let us first show that f is affine, i.e., that for any x ∈ f(Y ), there is an

open affine neighborhood U of y with f−1(U) affine. Let y ∈ Y be the (uniquely

defined) preimage of x and choose affine open subsets V ⊆ Y,W ⊆ X with

f(V ) ⊆ W containing y and x respectively. f is a homeomorphism, so f(V )

can be written as W ′ ∩ f(X) for some W ′ open set W ′ ⊆ W containing y. Now

choose a section s of OX(W ) such that D(s) ⊆ W ′ and contains x. Then the

open set f−1(D(s)) is contained in the affine set V (it is the intersection of V

with D(f ∗s)). So f−1(D(s)) is also affine, and hence we see that f is affine.

Now we claim that f∗OY is quasi-coherent on X. By the above, X is covered

by open affines U = SpecA such that f−1(U) = SpecB is also affine. In this case,

f∗OY |U is the ∼ of B considered as an A-module, and hence is quasi-coherent.

Now, letting I be the kernel of f# (which is quasi-coherent!), we have an

exact sequence

0→ I → OX → f∗(OY )→ 0.

In particular, we see that f∗(OY ) is isomorphic to the sheaf OX/I. Moreover,

the OX/I has support on f(Y ). So the map f : Y → X is isomorphic to

144

Page 145: Geir Ellingsrud and John Christian Ottem October 15, 2018€¦ · between algebra and geometry: a ne varieties correspond to nitely generated k-algebras without zero-divisors, and

Chapter 8. Quasi-coherent sheaves

the inclusion (f(Y ),OX/I)→ (X,OX), which proves that it is indeed a closed

immersion.

8.9.1 Morphisms to a closed subscheme

If f : Y → X is a map of schemes, it is natural to ask when it factors through a

closed immersion of X. Here we need only work up to isomorphism, so we can

assume that the closed immersion is given by (Supp OX/I,OX/I)→ (X,OX).

Proposition 8.35. Let Z be a closed subscheme of X given by sheaf of ideals I.

Suppose f : Y → X is a morphism of schemes. Then f factors through a map

g : Y → Z if and only if

(i) f(Y ) ⊆ Z.

(ii) I ⊆ ker(OX → f∗(OY )).

Proof. The condition (i) is clearly necessary. If there is a sequence Y → Z → X,

then there is a sequence of sheaves OX → OX/I → f∗(OY ), which means that

the map OX → f∗(OY ) factors through OX/I, and so also (ii) holds.

Conversely, we define the map g on topological spaces by the inclusion (i).

To define it on sheaves, we use the map OX → f∗(OY ). This annihilates I, so

we thus get a map OX/I → f∗(OY ) = g∗(OY ). This gives us the map g : Y → Z

factoring f .

Remark 8.36 (Scheme-theoretic image). Let f : Y → X be a morphism of

schemes. Then we can define the scheme-theoretic image of f as a subscheme

Z ⊆ X satisfying the universal property that if f factors through a subscheme

Z ′ ⊆ Z, then Z ⊆ Z ′. To define Z it is is tempting to use the ideal sheaf

I = ker(OX → f∗(OY )) – but this may fail to be quasi-coherent for a general

morphism f . However, one can show that there is a largest quasi-coherent sheaf

of ideals J contained in I, and we then define Z to be associated to J .

8.9.2 Induced reduced scheme structure

For an open subscheme U ⊆ X, we saw that there was a natural scheme structure

on U induced from that of X. For W ⊆ X a closed subset, there can be several

quasi-coherent ideal sheaves I corresponding to W . For instance, Spec k[x]/x

is naturally a subscheme of Spec k[x]/x2, but of course they have the same

underlying topological space. So, in contrast with the ‘open subschemes’ the

underlying topological space does not determine the scheme structure. However

there is one, which is in some sense the ‘smallest’ one:

145

Page 146: Geir Ellingsrud and John Christian Ottem October 15, 2018€¦ · between algebra and geometry: a ne varieties correspond to nitely generated k-algebras without zero-divisors, and

8.9. Closed immersions and subschemes

Proposition 8.37 (Induced reduced scheme structure). Let X be a scheme. Let

W ⊆ X be a closed subset. There exists a unique closed subscheme Z ⊆ X such

that

• Z is reduced;

• The underlying topological space of Z is W .

Proof. Let I ⊆ OX be the sheaf of ideals defined by

I(U) = f ∈ OX(U) | fx ∈ mx ⊆ OX,x for all x ∈ W ∩ U

We claim that this is quasi-coherent. On U = SpecA, we have W ∩U = V (I) for

a unique radical ideal I ⊆ R. Here we have I(U) = I: If f ∈ A is an element such

that f/1 ∈ mp ⊆ Ap then f ∈ p, and so if f ∈ I(U) then f ∈ ∩p∈V (I)p =√I = I.

Moreover, for D(g) ⊆ U , we have I(D(g)) = Ig by the same argument, and so I

and I are equal as sheaves on U , and hence I is a quasi-coherent sheaf of ideals.

Now define Z to be the closed subscheme associated with the ideal sheaf

I. Z is reduced, and has same underlying topological space as W ; this holds

because it is true on the open affines U = SpecA as above. Finally, if Z,Z ′ are

two subschemes satisfying the two points, then their ideal sheaves I, I ′ define the

same radical ideal I(U) = I ′(U) = I on U = SpecA, and so they are equal.

146

Page 147: Geir Ellingsrud and John Christian Ottem October 15, 2018€¦ · between algebra and geometry: a ne varieties correspond to nitely generated k-algebras without zero-divisors, and

Chapter 9

Vector bundles and locally free

sheaves

The most important examples of quasi-coherent sheaves are the locally free

sheaves.

Definition 9.1. A locally free sheaf F is an OX-module such that there exists

an open cover Ui and an index set I such that F|Ui '⊕

I OUi for each i. We

call Ui a trivialization of F . If I is finite, we say that F has finite rank, and

that r = |I| is called the rank of F . A locally free sheaf of rank 1 is called an

invertible sheaf.

If F is a locally free sheaf, the stalk Fx at a point x ∈ X is a free OX,x-module,

i.e., Fx ' OrX,x where r is the rank of F . In fact the converse is also true:

Lemma 9.2. Suppose that X is locally noetherian. A coherent sheaf F having

the property that Fx ' OrX,x for some fixed r is locally free.

Proof. We can assume that X = Spec(A), where A is noetherian, and F = M ,

where M is a finitely generated A-module. Let x1, . . . , xn be generators for M

as an A module. We have Fx = Mp, for the prime ideal p ⊆ A corresponding to

x ∈ X. By assumption, Mp ' Arp is free, so let m1, . . . ,mr be a basis of Mp as

an Ap-module. We can write in Mp

xi =∑

cijmj

Clearing denominators, we see that some multiple of dixi (with di ∈ A − p)

is generated by the elements mi with coefficients in A. Let s = d1 · · · dr, and

147

Page 148: Geir Ellingsrud and John Christian Ottem October 15, 2018€¦ · between algebra and geometry: a ne varieties correspond to nitely generated k-algebras without zero-divisors, and

consider the open set D(s) ⊆ X. s is invertible in Ms, so there is a surjective

map Ars → Ms. This is also injective, since any relation between the mi in

Ms must survive in Mp (since s 6∈ p). Hence Ms ' Ars. It then follows that

F |D(s) ' Ms ' OrX |D(s), is free on an open neighbourhood of x.

The coherence condition in this lemma is essential. Indeed, if X is a scheme,

and ix : x→ X denotes the inclusion of a point, the sheaf

F =⊕x∈X

ix∗Ox

is naturally an OX-module with Fx = Ox, but yet it is certainly not locally free.

Example 9.3. Here is an example of a locally free sheaf of rank 2. Consider

the morphism f : P1k → P1

k given by k[u] 7→ k[t] u 7→ t2 on U = Spec k[t], and

k[u−1]→ k[t−1] u−1 7→ t−2 on V = Spec k[t−1]. Then f∗OP1 is a locally free sheaf.

Indeed, on U , f∗OP1|U is the ∼ of k[t] as a k[u]-module, which equals k[u]⊕k[u]t.

We get a similar expression on V = Spec k[t−1]. It follows that f∗OP1 is locally

free of rank 2.

However, the pushforward of a locally free sheaf is not locally free in general.

For instance, if i : Y → is a closed immersion, then F = i∗OY has Fy = OY,y for

y ∈ Y , but zero stalks outside of Y .

For pullbacks however, we have the following:

Lemma 9.4. Let f : X → Y be a morphism of schemes. If G is a locally free

OY -module, then f ∗G is a locally free OX-module.

Proof. Let Ui be trivialization of F on Y , such that F|Ui '⊕

I OUi . Then, since

f ∗OY = OX , we see that f−1(Ui) is a trivialization of f ∗F .

Example 9.5. (The tangent bundle of the n-sphere) Let X = SpecA where

A = R[x0, . . . , xn]/(x20 + · · ·+x2

n−1), and consider the A-module homomorphism

f : An+1 → A given by f(ei) = xi. Then M = ker f gives rise to a quasi-coherent

sheaf F = M . Any element in the kernel corresponds to a vector of elements

a = (a0, . . . , an) ∈ An+1 so that

a0x0 + · · ·+ anxn = 0

On U = D(x0) we may divide by x0, and solve for a0, so a is uniquely determined

by the elements (a1, . . . , an). Conversely, given any such an n-tuple of elements

in A, we may define an element a ∈Mx0 using the above relation. In particular,

Mx0 ' An. A similar argument works for the other xi, showing that F is locally

free.

It is a non-trivial theorem that F 6= OnX if n 6∈ 0, 1, 3, 7.

148

Page 149: Geir Ellingsrud and John Christian Ottem October 15, 2018€¦ · between algebra and geometry: a ne varieties correspond to nitely generated k-algebras without zero-divisors, and

Chapter 9. Vector bundles and locally free sheaves

9.1 Locally free sheaves and projective mod-

ules (work in progress)

On an affine scheme X = SpecA, any quasi-coherent OX-module F is isomorphic

to M for an A-module M . In this section we investigate which A-modules give

rise to locally free sheaves. The main result is that F is locally free if and only

if M is finitely generated and projective.

Defn. projective module

Defn. Locally free module

Lemma 9.6. Let (A,m) be a local ring. If M is a finitely generated projective

A-module, then M is free.

Let M be a projective A-module, so there is a Q so that M⊕Q = F is finitely

generated, and free. Taking localizations, we then see that Mp is a projective

Ap-module for every prime p. Hence Mp is free for every p.

Lemma 9.7. Let M be a finitely generated A-module. then M is projective if

and only if M is finitely presented and locally free.

Theorem 9.8. Let M be a finitely generated A-module, and assume A is a

Noetherian ring. Then F = M is locally if and only if M is projective.

Corollary 9.9. Let X be a noetherian scheme and let F be a coherent sheaf.

Then the following are equivalent:

(i) F is locally free

(ii) Fx is a free OX,x-module for every x ∈ X.

(iii) For every open affine U = SpecA ⊆ X, we have F|U ' M , for some

finitely generated projective A-module M .

Proof. Suppose Fx is free for every x ∈ X. Fix x ∈ X and let U = SpecA be

an open subset containing x. We have F|U ' M for some finitely generated

A-module M . Then Mq is free for every q ∈ SpecA, and hence M is projective.

Hence M is locally free, and hence so is F , since x was arbitrary.

9.2 Invertible sheaves and the Picard group

An invertible sheaf on a scheme X is a locally free sheaf of rank 1. We usually

write L for such sheaves (for reasons that will become clear later). This means

that there exists a covering U = Ui and isomorphisms φi : OUi → L|Ui . We say

149

Page 150: Geir Ellingsrud and John Christian Ottem October 15, 2018€¦ · between algebra and geometry: a ne varieties correspond to nitely generated k-algebras without zero-divisors, and

9.2. Invertible sheaves and the Picard group

that gi = (φi)Ui(1) ∈ L(Ui) is a local generator for L and that Ui is a trivialization

of L. By Lemma 9.2, a coherent OX-module L is invertible if and only if Lx is

isomorphic to OX,x for every x ∈ X.

Proposition 9.10. For L,M invertible sheaves, we have

(i) L⊗M is also an invertible sheaf. If g, h are local generators for L and M

respectively, then g ⊗ h a local generator for L⊗OX M .

(ii) Hom(L,OX) is invertible and Hom(L,OX) ⊗ L ' L. If g is a local

generator for L, then ψg defined by ψg(ag) = a is a local generator for

Hom(L,OX).

(iii) Hom(L,M) 'Hom(L,OX)⊗M

Proof. (i) We can find a common trivialization of L and M , such that X is covered

by open sets U where we have isomorphisms φ : OU → L|U and ψ : OU →M |U .

Over such a U , we have an isomorphism OU ' OU ⊗ OY ' L|U ⊗M |U given by

1 7→ 1⊗ 1 7→ φ(1)⊗ ψ(1) (all tensor products in this section are over OX . This

shows (i).

For (ii), as above the fact that Hom(L,OX) is invertible can be seen by re-

stricting to an open where L|U ' OU . The identity for the tensor product follows

from (iii) below. Finally, the fact that ψg is a local generator for Hom(L,OX)

follows from the isomorphisms

OX |U 'Hom(OX ,OX)|U 'Hom(L|U ,OX)

given by 1 7→ idOX and α 7→ α ψ−1. This means that 1 maps to ψ−1 = ψg.

(iii) By the universal property of the tensor product, we have a canonical

isomorphism

HomOX(U)(L(U),M(U))) ' HomOX(U)(HomOX(U)(L(U),OX(U))⊗OX(U) M(U))

This shows (iii) on the level of presheaves. Sheafify to get the result.

This proposition explains the term ‘invertible’. Indeed, the tensor product

acts as a sort of binary operation on the set of invertible sheaves; L ⊗M is

invertible if L and M are. Tensoring an invertible sheaf by OX does nothing,

so OX serves as the identity. Moreover, for an invertible sheaf L we will define

L−1 = Hom(L,OX); by the proposition, L−1 is again invertible, and serves as a

multiplicative inverse of L under ⊗. We can make the following definition:

Definition 9.11. Let X be a scheme. The Picard group Pic(X) is the group of

invertible sheaves L modulo isomorphism.

150

Page 151: Geir Ellingsrud and John Christian Ottem October 15, 2018€¦ · between algebra and geometry: a ne varieties correspond to nitely generated k-algebras without zero-divisors, and

Chapter 9. Vector bundles and locally free sheaves

Notice it is the set of isomorphism classes of line bundles that form a group –

not the line bundles themselves (L⊗ L−1 is isomorphic but strictly speaking not

equal to OX).

Note also that Pic(X) is also an abelian group, because L⊗M is canonically

isomorphic to M ⊗ L.

Example 9.12 (Invertible sheaves on the affine line). If F is a coherent sheaf on

A1k, then F = M for some finitely generated k[t]-module. The structure theorem

of finitely generated modules over a PID, tells us that M ' k[t]r ⊕ T where

r ≥ 0 and T is a torsion module (which is in turn a direct sum of modules of the

form k[t]/(t− a)n). From this we find

Proposition 9.13. Any invertible sheaf over A1k is trivial; Pic(A1

k) = 0.

More generally, in the correspondence between quasi-coherent sheaves on

X = SpecA and A-modules M , one can show that an OX-module is locally free

if and only if it is associated to a projective module M , that is, there exists N

such that M ⊕N is free.

9.2.1 Operations on locally free sheaves

Most constructions for vector spaces and modules globalize and carry over to

locally free sheaves. For instance, for a locally free sheaf F we can define

the tensor algebra T (F) as the graded algebra⊕

n≥0 Tn(F) where T n(F) =

F⊗n. We let Sym(F) = T (F)/I where I is the ideal generated by expressions

m⊗m′ −m′ ⊗m where m,m′ ∈ F . This inherits a Z-grading from T (F), and

we let SnF denote the n-th symmetric power of F .

Similarly, we define∧F as the OX-module T (F)/J where J is the ideal

generated by all products m ⊗m where m ∈ F . This is sometimes called the

exterior algebra. This is again graded, and a graded piece∧nF is called the

n-th wedge product. When F has rank r,∧r+1F = 0 and

∧r F is called the

determinant of F - it is a locally free sheaf of rank 1, i.e. an invertible sheaf.

Proposition 9.14 (Facts about locally free sheaves). (i) Locally free sheaves are

closed under direct sums, tensor products, symmetric products, exterior products,

duals, and pullbacks.

If F is locally free of rank r then

• T n(F) is locally free or rank rn (spanned by all elements m1 ⊗ · · · ⊗mn

where mi ∈ F)

• Sn(F) is locally free or rank(n+r−1r−1

)(spanned by all elements mn1

1 · · · ⊗mr

where mi ∈ F and ∼ ni = n)

151

Page 152: Geir Ellingsrud and John Christian Ottem October 15, 2018€¦ · between algebra and geometry: a ne varieties correspond to nitely generated k-algebras without zero-divisors, and

9.3. Vector bundles

•∧nF is locally free of rank

(rn

)(spanned by elements mi1 ∧ · · ·min where

i1 < i2 < · · · < in.

(ii) Let 0→ F ′ → F → F ′′ → 0 be an exact sequence of locally free sheaves

of ranks n′, n, n′′ respectively. Then

∧nF ' ∧n′F ′ ⊗OX ∧n′′F ′′

9.3 Vector bundles

Locally free sheaves are in many respects the most basic sheaves on a scheme X;

they are obtained by locally gluing together copies of the structure sheaf OX in

various ways. What makes these sheaves particularly interesting is the link to

the theory of vector bundles.

A vector bundle is essentially a family of vector spaces parameterized over

a base scheme X. This means that we have a morphism π : E → X, such

that the scheme theoretic fibers π−1(x) are isomorphic to an affine space Ar.

The prototype example of a vector bundle is obtained by simply taking the

product E = X × Ar and letting π be the first projection. This is the so called

trivial bundle on X. A vector bundle is more generally a scheme together with a

morphism π : E → X which is locally isomorphic to the trivial bundle:

Definition 9.15. Let X be a scheme. A vector bundle E of rank r on X is a

scheme with a morphism π : E → X and an open cover U = Ui of X such that

for each i, there is an isomorphism φi : π−1(Ui)→ Ui × Ar making the following

diagram commutative:

π−1(Ui)φi //

π|π−1(Ui) ((

Ui × Ar

pr1

Ui

such that for each affine V = SpecA ⊆ Uij the automorphism φj φ−1i of V ×Ar

is linear; i.e., induced by an automorphism θ : A[x1, . . . , xn] → A[x1, . . . , xn]

such that θ(a) = a,∀a ∈ A, and θ(xi) =∑aijxj for aij ∈ A.

152

Page 153: Geir Ellingsrud and John Christian Ottem October 15, 2018€¦ · between algebra and geometry: a ne varieties correspond to nitely generated k-algebras without zero-divisors, and

Chapter 9. Vector bundles and locally free sheaves

So a vector bundle is obtained by gluing together trivial bundles Ui × Ar via

linear gluing maps. A convenient way of rephrasing this is in terms of transition

functions: Given a vector bundle π : E → X defined by the data (Ui, φi) we have

for each i, j an element gij ∈ GLr(Γ(Uij,OX)) such that the diagram

Uij × Ar

π−1(Uij)φj |Uij

&&

φi|Uij88

Uij × Ar

id×gij

OO

commutes.

Definition 9.16. The elements gij are the transition functions of E .

The gluing axioms for E shows that the transition functions satisfy the

following compatibility conditions (or ‘cocycle conditions’)

gik = gij gjk on Uijk (9.3.1)

gij = g−1ji on Uij

Proposition 9.17. Let X be a scheme with an open cover Ui and assume that

gij are a collection of elements of GLr(Γ(Uij,OX)) satisfying the compatibility

conditions (9.3.1). Then there is a vector bundle π : E → X, unique up to

isomorphism, whose transition functions are the gij.

Proof. The compatibility conditions ensure that the maps (id×gij) glue the affine

schemes Ui × Ar along (Ui ∩ Uj)× Ar to a scheme E . Moreover, the projection

maps Ui × Ar → Ui glue to give a morphism π : E → X. By construction, the

open set of E corresponding to Ui×Ar is identified with π−1(Ui), which gives an

isomorphism φi : π−1(Ui)→ Ui×Ar. Hence E is a vector bundle with transition

functions gij.

153

Page 154: Geir Ellingsrud and John Christian Ottem October 15, 2018€¦ · between algebra and geometry: a ne varieties correspond to nitely generated k-algebras without zero-divisors, and

9.3. Vector bundles

Example 9.18 (Facts about vector bundles). Let π : E → X be a vector bundle

or rank r associated with the transition functions gij ∈ Γ(Uij,O∗X) and let Odenote the trivial rank 1 bundle.

(i) There is a vector bundle E ∗, the dual bundle of E , with corresponding

transition functions g−1ij = gji. E ∗ = Homvb(E ,O) is also a vector bundle

(ii) If E ,E ′ are vector bundles with transition functions gij hij respectively,

one can define the tensor product bundle E⊗E ′ as the vector bundle corresponding

to the transition functions gij⊗hij . If E and E ′ have ranks r, r′, then E ⊗E ′ has

rank rr′. As a special case of this, we have the tensor power E ⊗n which is a vector

bundle of rank rn, with corresponding transition functions gij ⊗ gij ⊗ · · · ⊗ gij.

9.3.1 The sheaf of sections of a vector bundle

A section of a vector bundle π : E → X over U ⊆ X is a morphism s : U → Esuch that π s = idU . So s picks out a single vector in each fiber π−1(x) for

x ∈ X a closed point.

Proposition 9.19. Let π : E → X be a vector bundle given by the data

(Ui, φi, gij). A section s : X → E is determined uniquely by a collection of

r-tuples si ∈ OrX such that for each i, j

si|Uij = gijsj|Uij .

Proof. Given s : X → E , φi s|Ui is a section of Ui×Ar → Ui. Hence φi s|Ui =

(idU ×si) for si ∈ O(Ui)r. By construction, these satisfy the given compatibility

conditions si|Uij = gijsj|Uij . Conversely, any such section s : X → E defines a

set of such sections si = s|Ui satisfying this condition.

Given E , we can define a sheaf OX(E ) by defining

OX(E )(U) = sections s : U → E

The above proposition shows that this set is naturally a group: If s : U → E ,

t : U → E are two sections given by the data si and ti respectively, we can define

s+ t, by si + ti ∈ OrX . In fact, by multiplying si with elements of OX(U), we see

that OX(E )(U) has the structure of an OX(U)-module. This shows that OX(E )

is a sheaf of OX-modules, locally free of rank r.

9.3.2 Equivalence between vector bundles and locally free

sheaves

A vector bundle π : E → X gives rise to a locally free sheaf F of rank r. In fact,

there is a way to reverse this process; that is, to a locally free sheaf F one can

154

Page 155: Geir Ellingsrud and John Christian Ottem October 15, 2018€¦ · between algebra and geometry: a ne varieties correspond to nitely generated k-algebras without zero-divisors, and

Chapter 9. Vector bundles and locally free sheaves

associate a scheme, denoted V(F) with a morphism π : V(F)→ X making V(F)

into a vector bundle under which F is the sheaf of sections. In particular, there

is an equivalence between vector bundles and locally free sheaves of finite rank.

The construction goes as follows. If F is locally free of rank r. Choose an

open cover Ui of X such that there is an isomorphism fi : F|Ui → OUi for

each i. If we take two fi : F|Ui → OUi and fj : F|Uj → OUi and restrict to

Uij = Ui ∩ Uj, we get two different isomorphisms gi, gj : F|Uij → OrUij

. Let

gij = gj g−1i , which gives an automorphism of Or

Uij. We can identity this with

an r × r matrix of regular functions on Uij.

Now we glue Ui × Ar and Uj × Ar along Uij × Ar by the map sending

(x, v) ∈ Uij × Ar to (x, gij(v)). The resulting scheme which we denote by

V(F) comes with a morphism π : V(F) → X (obtained by gluing all the first

projections Ui×Ar → Ui), whose fibers are linear spaces. Over each Ui ⊆ X the

maps gi also give isomorphisms from π−1(Ui) to Ui × Ar. The maps gj g−1i are

all linear - this follows from the fact that gj g−1i is a morphism of OX-modules.

Finally, it is a matter of checking that F coincides with the sheaf of sections of

π.

In particular, there is a correspondence between vector bundles and locally

free sheaves of finite rank. Under this correspondence the trivial bundle X × A1

corresponds to the structure sheaf OX .

Remark 9.20. There is also an alternate construction of V(F) using the relative

spectrum of the symmetric algebra of a sheaf. We direct the interested reader to

Exc. II.5.16 and II.5.18 in Hartshorne’s book for more details.

9.4 Extended example: The tautological bun-

dle on PnkLet k be a field, and let Pnk denote the projective n-space over k. The closed

points of Pnk parameterizes lines through the origin in An+1k . We define the

tautological sheaf, denoted by O(−1) as follows. Let L ⊆ Pn ×k An+1 denote the

variety of pairs ([l], v), where l ⊆ An+1 is a line, and v ∈ l. In terms of equations

L is defined by the 2× 2 minors of the matrix(x0 x1 . . . xny0 y1 . . . yn

)where x0, . . . , xn are homogeneous coordinates on P1 and y0, . . . , yn are affine

coordinates on An+1.

Let π : L → Pnk be the projection to the first factor. Then for a closed

point [a] = (a0 : . . . , an) ∈ Pnk , the preimage π−1([l]) is exactly the line l ⊆ An+1

155

Page 156: Geir Ellingsrud and John Christian Ottem October 15, 2018€¦ · between algebra and geometry: a ne varieties correspond to nitely generated k-algebras without zero-divisors, and

9.4. Extended example: The tautological bundle on Pnk

corresponding to [a]. We claim that L is a rank 1 vector bundle on Pn.

On Ui = D(xi) the equations for L become

xjxiyi = yj

for each j 6= i and hence (x0 : . . . , xn, y0, . . . , yn) 7→ (x0 : . . . , xn, yi) defines an

isomorphism

φi : π−1(Ui)→ Ui × A1

In other words, yi is a local coordinate on the A1 on the right hand side. On the

intersection Ui ∩ Uj, yi is related to yj via

xixjyj = yi

This means that the transition function

Uij × A1 → Uij × A1

is given by idUij × gij where gij = xixj

. Hence π is a vector bundle. We denote

the corresponding sheaf of sections by O(−1).

What are the sections of O(−1) ? Note that a section s : Pn → L is given by

a set of n+ 1 sections si, such that for each i, j we have

si|Uij = gijsj|Uij =xixjsj|Uij (9.4.1)

The functions si : D(xi)→ L can be represented by polynomials in x0xi, . . . , xn

xi.

Note that on the right hand side of (9.4.2), the Laurent polynomial xixjsj has only

xj in the denominator, wheras si, on the left, has only xi’s. It follows that the

only way this equation in polynomials can hold is that both sides are identically

0. In particular,

Proposition 9.21. Γ(Pnk ,O(−1)) = 0.

In particular, the sheaf O(−1) is invertible, but not isomorphic to OPnk .

Consider now the vector bundle σ : L∗ → X with transition functions

gij =xjxi

Note that these transition functions are almost as before, only that we have

inverted the gij . The corresponding vector bundle is the dual bundle of π : L→ X;

the fiber σ−1(x) is identified with the dual vector space (π−1(x))∗.

In this case we find that the bundle does in fact have global sections: A

156

Page 157: Geir Ellingsrud and John Christian Ottem October 15, 2018€¦ · between algebra and geometry: a ne varieties correspond to nitely generated k-algebras without zero-divisors, and

Chapter 9. Vector bundles and locally free sheaves

section s : Pn → L∗ is given by a set of sections si : Ui → W such that

si|Uij = gijsj|Uij =xjxisj|Uij (9.4.2)

As before, the left-hand side is a a polynomial in x0xi, . . . , xn

xi- in particular it is

a Laurent polynomial with xi in the denominators. This is ok with respect to

the right-hand side, as long as si has degree 1, that is it has a pole of order at

most one at xi = 0. Conversely, you can start with any section s0 which is linear

in x1x0, . . . , xn

x0, and use (9.4.2) to define s1, . . . , sn as Laurent polynomials. These

glue together to a global section s : Pn → L∗.

Proposition 9.22. The space of global sections of L∗ can be identified with the

vector space of linear forms in n+ 1 variables.

157

Page 158: Geir Ellingsrud and John Christian Ottem October 15, 2018€¦ · between algebra and geometry: a ne varieties correspond to nitely generated k-algebras without zero-divisors, and
Page 159: Geir Ellingsrud and John Christian Ottem October 15, 2018€¦ · between algebra and geometry: a ne varieties correspond to nitely generated k-algebras without zero-divisors, and

Chapter 10

Sheaves on projective schemes

Projective schemes are to affine schemes what projective varieties are to affine

varieties. The construction of the projective spectrum ProjR was similar to that

of Spec: the underlying topological space is defined via prime ideals and the

structure sheaf from localizations of R. However, there were some fundamental

differences between the two: We only consider graded rings R, and we only

consider homogeneous prime ideals that do not contain the irrelevant ideal R+.

As we saw, this reflects the construction of ProjR as a quotient space

π : SpecR− V (R+)→ ProjR

Given this, we can pull back a quasi-coherent sheaf to SpecR − V (R+) and

extend it to a sheaf on SpecR via the inclusion map. Thus, it is natural to

expect that quasi-coherent sheaves on ProjR should be in correspondence with

equivariant modules on SpecR, i.e., with graded R-modules. The irrelevant

subscheme V (R+) complicates the picture a little bit, so the classification is a

little bit more involved than the one for affines schemes. In particular, we will

see that different graded R-modules can correspond to the same quasi-coherent

sheaf on ProjR.

Another important feature of ProjR is that it comes equipped with a canonical

invertible sheaf that we will denote by O(1). This is the geometric manifestation

of the fact that R is graded. Unlike the case of affine schemes, ProjS can

typically not be recovered from the global sections of the structure sheaf. It is

the sheaf O(1), or rather, the various tensor powers O(d) = O(1)⊗d, that will

play the role of the affine coordinate ring in the affine case.

159

Page 160: Geir Ellingsrud and John Christian Ottem October 15, 2018€¦ · between algebra and geometry: a ne varieties correspond to nitely generated k-algebras without zero-divisors, and

10.1. The graded ∼-functor

10.1 The graded ∼-functor

Let R be a graded ring and let GrModR denote the category of graded R-modules.

Just like in the case of SpecA we will define a tilde-construction to construct

sheaves on ProjR from graded S-modules, giving a functor GrModR to ModOX .

However, as we will see, this is not an equivalence of categories.

Recall that for f, g ∈ R+, such that D+(g) ⊆ D+(f), there is a canonical

localization homomorphism ρf,g : M(f) →M(g) where as before M(f) denotes the

degree 0 part of the localization 1, f, f 2, . . . −1M . It follows that we can define

a B-presheaf M by defining for each D+(f),

M(D+(f)) = M(f).

Note that M |D(f) ' (M(f)) via the isomorphism of D+(f) with SpecR(f). It

follows that this in fact gives a B-sheaf, and hence a sheaf on ProjR. Moreover,

as in the Spec case, the assignment M 7→ M is functorial. The following

proposition summarizes the properties of M :

Proposition 10.1. The contravariant functor M 7→ M has the following prop-

erties:

• ∼ is exact, commutes with direct sums and limits.

• The stalks satisfy Mp = M(p) for each p ∈ ProjR.

• If R is noetherian, and M is finitely generated, then M is coherent.

Proving these properties is straightforward, since most of them can be checked

locally on stalks. Using the isomorphisms between D+(f) and SpecR(f) we reduce

immediately to the affine case.

However, it is important to note that, unlike the affine case, the functor is not

faithful, as several modules can correspond to the same module. For instance,

take any graded R-module M such that Md = 0 for all large d. Then M(f) = 0

for all f ∈ R+, and so M = 0, even though M is not the 0-module. We will

however see shortly that it is only the modules of this form that cause the lack

of faithfulness.

The following is useful for working with the localization of M . It says

essentially that we are allowed to ‘substitute in 1’ when restricting a module to

an affine chart D+(f) ⊆ ProjR,

Lemma 10.2. Suppose that M is a graded R-module and f ∈ R homogeneous

of degree 1. Then

M(f) 'M/(f − 1)M 'M ⊗R R/(f − 1)

160

Page 161: Geir Ellingsrud and John Christian Ottem October 15, 2018€¦ · between algebra and geometry: a ne varieties correspond to nitely generated k-algebras without zero-divisors, and

Chapter 10. Sheaves on projective schemes

Proof. Define a map M → M(f) by m 7→ m/fdegm. The submodule (f − 1)M

is annihilated by this map, so we get an induced map φ : M/(f − 1)M →M(f).

We can construct an inverse to this map as follows. Let m/fn ∈ M(f) for

m ∈ Md. Map this to the class of m in M/(f − 1)M . This is well-defined: If

m/fn = m′/fn′, then there is an N > 0 with

fN(mfn′ −m′fn) = 0.

so since f gets reduced to 1 modulo (f − 1), we find that they have the same

images in M/(f − 1)M .

Let us use this to compare M ⊗R N with M ⊗OX N . Let f ∈ R be a

homogeneous element. We have a map M(f) × N(f) → (M ⊗R N)(f) sending

m/fa × n/f b to (m⊗ n)/fa+b. As this is R(f)-bilinear, we get an induced map

M(f) ⊗R(f)N(f) → (M ⊗R N)(f). Since a map of B-sheaves induces a map of

sheaves, we get a natural map

M ⊗OX N → M ⊗R N. (10.1.1)

Proposition 10.3. Suppose R is generated in degree 1. Then the natural map

(10.1.1) is an isomorphism.

Proof. By assumption, X = ProjR is covered by open affines of the form D+(f)

where f has degree 1. For such an f , the functor M → M(f) is the same as

tensoring with R/(f − 1) ' R(f) by the previous lemma. Furthermore,

(M ⊗R R(f))⊗R(f)(N ⊗R R(f)) ' (M ⊗R N)⊗R R(f)

This isomorphism provides the inverse to the natural map M(f) ⊗R(f)N(f) →

(M ⊗R N)(f) defined above. Then, since the map (10.1.1) restricts to an isomor-

phism on all D+(f) for f ∈ R1, it is an isomorphism.

10.2 Serre’s twisting sheaf O(1)

Arguably the most interesting sheaf on X = ProjR is the so-called Serre’s

twisting sheaf, denoted by OX(1). This is a generalization of the tautological

sheaf on Pnk , and constitutes a geometric manifestation of the fact that R is a

graded ring.

For an integer n, we will define an R-module R(n) as follows: As an underlying

module R(n) is just R, but with the grading shifted by n:

R(n)d = Rd+n

161

Page 162: Geir Ellingsrud and John Christian Ottem October 15, 2018€¦ · between algebra and geometry: a ne varieties correspond to nitely generated k-algebras without zero-divisors, and

10.2. Serre’s twisting sheaf O(1)

This is naturally a graded R-module and hence gives rise to a quasi-coherent

OX-module on X.

Definition 10.4. For an integer n, we define

OX(n) = R(n).

For a sheaf of OX modules F on X, we define the twist by n by F ⊗OX OX(n).

For an element r ∈ Rd, there is a corresponding section in Γ(X,OX(d)). This

is because we can think of an element of Γ(X,OX(d)) as a collection of elements

(rf , D+(f)) with rf ∈ (Rf )d matching on the overlaps D+(f)∩D+(g). Hence we

can define an R0-module homomorphism

Rd → Γ(X,OX(d))

by r 7→ (r/1, D+(f)). (On the overlaps D+(fg) it is clear that the two elements

(r/1, D+(f)) and (r/1, D+(g)) become equal, so this defines an actual global

section of O(d)). Abusing notation, we will also denote the section by r.

Note that if f ∈ R1, then R(n)(f) = fnR(f). Thus, on the affine D+(f), we

have OX(n)|D+(f) = fnOX |D+(f). In particular, OX(n)|D+(f) ' OD+(f). In other,

words OX(n) is a locally free sheaf of rank 1, that is, an invertible sheaf. By the

previous chapter, we know that this gives rise to a line bundle L→ ProjR.

Proposition 10.5. If R is generated in degree 1, then OX(n) is an invertible

sheaf for every n. Moreover, there is a canonical isomorphism

OX(m+ n) ' OX(m)⊗OX OX(n)

Indeed, if R is generated in degree 1, Proposition 10.3 shows that OX(m)⊗OX(n) is the sheaf associated to R(m)⊗R R(n) ' R(n+m), i.e., OX(n+m).

So this is a big difference between affine schemes and projective schemes:

ProjR comes equipped with lots of invertible sheaves.

Example 10.6. Consider again the example of projective space Pnk = ProjR

whereR = k[x0, . . . , xn]. We use the covering Ui = D(xi) ' Spec(k[x0, . . . , xn](xi)

)=

Spec(k[x0

xi, . . . , xn

xi])

.Then R (l)xi = (k[x0, . . . , xn]xi)l = xlik[x0xi, . . . , xn

xi], and so

Γ (Ui,O (l)) = xlik[x0

xi, . . . ,

xnxi

]

On the overlaps,(R (l)xi

)xjxi

= R (l)xixj =(R (l)xi

)xixj

we find that two regular

sections

xlisi

(x0

xi, . . . ,

xnxi

)162

Page 163: Geir Ellingsrud and John Christian Ottem October 15, 2018€¦ · between algebra and geometry: a ne varieties correspond to nitely generated k-algebras without zero-divisors, and

Chapter 10. Sheaves on projective schemes

and

xljsj

(x0

xj, . . . ,

xnxj

)restrict to the same section of O (l) |Ui∩Uj if and only if

sj

(x0

xj, . . . ,

xnxj

)=

(xixj

)lsi

(x0

xi, . . . ,

xnxi

)You should compare this to the descriptions of the transition functions of the

tautological bundle in Chapter 7. In particular, we see that the invertible sheaf

O(1) coincides with the tautological sheaf defined there.

10.3 Extending sections of invertible sheaves

Let F be an OX-module and let x ∈ X be a point. We define the fiber of F at

x to be the k(x)-vector space

F(x) : Fx/mxFx ' Fx ⊗OX,x k(x)

If U is an open set containing x and s ∈ Γ(U,F), we denote by s(x) the image

of the germ sx ∈ Fx in F(x).

Definition 10.7. Let L be an invertible sheaf and f ∈ Γ(X,L). We define the

open set Xs as

Xs = x ∈ X|s(x) 6= 0.

Equivalently, Xf is the set of points where f 6∈ mxLx.

Xs is indeed an open set of X: L is locally free, so through every point there

is a neighbourhood U such that L|U ' OX |U . To show that it is open, we can

therefore assume that L = OX . If x ∈ Xs there exists a tx ∈ OX,x such that

sxtx = 1. Choose an open neighbourhood V of x such tx is represented by a

section t ∈ Γ(V,OX). By shrinking V we can assume that (s|V )t = 1 on V , and

so V ⊆ Xs is open.

Lemma 10.8. Suppose X is a noetherian scheme. Let F be a quasi-coherent

sheaf on X, and L an invertible sheaf. Suppose f ∈ Γ(X,L). Then:

1. If a section t ∈ Γ(X,F) restricts to zero on Xf , then there is an integer

N such that t⊗ fN ∈ Γ(X,F ⊗ LN) is zero (on all of X).

2. Suppose t ∈ Γ(Xf ,F). Then there is an integer N such that t⊗fN extends

to a global section of F ⊗ LN .

163

Page 164: Geir Ellingsrud and John Christian Ottem October 15, 2018€¦ · between algebra and geometry: a ne varieties correspond to nitely generated k-algebras without zero-divisors, and

10.4. The associated graded module

Proof of the lemma. (i) Suppose t restricts to zero on Xf . We may cover X

by finitely many open affines Ui such that there is a collection of trivialization

isomorphisms φi : L|Ui → OX |Ui . Then t maps to zero in Γ(Ui,F)f because Ui is

affine (regarding f as an element of OX(Ui) via the isomorphism φi). So there is

a power of f that annihilates t on Ui. In other words, this says that t⊗ fNi = 0

on Ui. Taking N = maxNi, we find that t⊗ fN is zero on all of X, as desired.

(ii) For each i, we know that some t⊗ fNi extends to a section over all of Ui(because Ui is affine and taking sections over basic open subsets corresponds to

localization). Take M = maxNi. Then t⊗fM extends to sections ti ∈ Γ(Ui,F ⊗LM). A potential problem is that the ti might not necessarily glue on Ui ∩ Uj.However, ti = tj on Ui ∩ Uj ∩Xf since they are extending something defined on

Xf (namely t). Since Ui∩Uj∩Xf is also noetherian, the first part (i) now implies

that there are powers Mij such that (ti− tj)⊗fMij = 0 ∈ Γ(Ui∩Uj,F⊗LM+Mij ).

Taking N = M + maxMij now does the trick: the ti ⊗ fmaxMij will glue and

extend t⊗ fN .

Remark 10.9. The same proof works also for an invertible sheaf L on a quasi-

compact and separated scheme X.

10.4 The associated graded module

We have associated to a graded R-module M a sheaf M on X = ProjR. To

classify quasi-coherent sheaves on X we would, like in the case of affine schemes,

give some sort of inverse to this assignment. However, unlike the case for

X = SpecA, we can not simply take the global sections functor. Indeed, even for

F = OX on X = P1k, Γ(X,OX) = k, form which we cannot recover F . However,

the remedy is to look at the various Serre twists F(m) – in fact all of them at

once:

Definition 10.10. Let R be a graded ring and let F be an OX-module on

X = ProjR. We define the graded R-module associated to F , Γ∗(F) as

Γ∗(F) =⊕n∈Z

Γ(X,F(n)).

This has the structure of an R-module: If r ∈ Rd, we have a corresponding

section r ∈ Γ(X,OX(d)) (abusing notation, as before). So if σ ∈ Γ(X,F(n))

then we define r · σ ∈ Γ(F(n+ d)) by the tensor product r ⊗ σ, and using the

isomorphism F(n)⊗ O(d) ' F(n+ d). In particular, Γ∗(OX) is a graded ring.

164

Page 165: Geir Ellingsrud and John Christian Ottem October 15, 2018€¦ · between algebra and geometry: a ne varieties correspond to nitely generated k-algebras without zero-divisors, and

Chapter 10. Sheaves on projective schemes

10.4.1 Sections of the structure sheaf

Proposition 10.11. Let R be a graded integral domain, finitely generated in

degree 1 by elements x1, . . . , xn, and let X = ProjR. Then

1. Γ∗(OX) =⋂ni=1R(xi) ⊆ K(R)

2. If each xi is a prime element, then R = Γ∗(OX).

Proof. Cover X by opens Ui = D+(xi). We have, since Γ(D+(xi),O(m)) '(Rxi)m, that the sheaf axiom sequence takes the following form

0→ O(m)→n⊕i=0

(Rxi)m →⊕i,j

(Rxixj)m

Taking directs sums over all m, we get

0→ Γ∗(OX)→n⊕i=0

(Rxi)→⊕i,j

(Rxixj)

So a section of Γ∗(OX) corresponds to an (n+ 1)-tuple (t0, . . . , tn) ∈⊕n

i=0(Rxi)

such that ti and tj coincide in Rxixj for each i 6= j. Now, the xi are not zero-

divisors in R, so the localization maps R → Rxi are injective. It follows that

we can view all the localizations Rxi as subrings of Rx0...xn , and then Γ∗(OX)

coincides with the intersection

n⋂i=0

Rxi ⊆ A[x±10 , . . . , x±1

n ].

In the case each xi is prime, this intersection is just R.

Corollary 10.12. Let X = PnA = ProjA[x0, . . . , xn] for a ring A. Then

Γ∗(OX) ' A[x0, . . . , xn]

In particular we can identify Γ(PnA,O(d)) with the A-module generated by homo-

geneous degree d polynomials.

When R is not a polynomial ring, it can easily happen that Γ∗(OX) is different

than R. Here is an explicit example:

Example 10.13 (A quartic rational curve). Let k be a field and let R =

k[s4, s3t, st3, t4] ⊆ k[s, t]. Note that the monomial s2t2 is missing from the

generators of R. Define the grading such that R1 = k · s4, s3t, st3, t4.

165

Page 166: Geir Ellingsrud and John Christian Ottem October 15, 2018€¦ · between algebra and geometry: a ne varieties correspond to nitely generated k-algebras without zero-divisors, and

10.4. The associated graded module

We can also think of R as the graded ring

R = k[x0, x1, x2, x3]/〈x20x2 − x3

1, x1x23 − x3

2, x0x3 − x1x2〉.

We have a covering ProjR = U0 ∪ U1, where

U0 = Spec(R(x0)) and U1 = Spec(R(x3)).

Here R(x0) = k[ ts, t

3

s3, t

4

s4] = k[ t

s] and R(x3) = k[ s

t]. So ProjR is simply P1. We

have shown that X embeds as a smooth, rational (degree 4) curve in P3.

What is Γ(X,OX(1)? On the opens we find OX(1)(U0) = k[ts

]· s4 and

OX(1)(U1) = k[st

]· t4. So using the sheaf sequence, we get

0→ Γ(X,OX(1))→ k[st

]s4 ⊕ k

[t

s

]t4 → k

[s

t,t

s

]u4

Note that the monomial s2t2 belongs to both k[st

]s4 and k

[ts

]t4, and so defines

an element in Γ(X,OX(1)). In fact,

Γ(X,OX(1)) = ks4, s3t, s2t2, st3, t4

even though R1 = ks4, s3t, st3, t4.In this example, the graded ring Γ∗(OX) is the integral closure of R. We will

see later that this is not a coincidence.

10.4.2 The homomorphism α.

Let X = ProjR, where R be a graded ring and let M be a graded R-module.

We will define a homomorphism of graded R-modules called the saturation map

α : M → Γ∗(M)

As before, it is useful to think of elements in Γ(X, M(n)) as a collection

of elements (mf , D+(f)) for m ∈ (M(f))n and f ∈ R matching on the various

overlaps.

Proposition 10.14. When R is generated in degree 1, there is a graded R-module

homomorphism

α : M → Γ∗(M)

Indeed, we can define α by sending an element m ∈Md to the collection given

by (m/1, D+(f)), where f ranges over R1. On the overlaps D+(f) ∩D+(g) =

D+(fg) it is clear that the two elements (m/1, D+(f)) and (m/1, D+(g)) become

166

Page 167: Geir Ellingsrud and John Christian Ottem October 15, 2018€¦ · between algebra and geometry: a ne varieties correspond to nitely generated k-algebras without zero-divisors, and

Chapter 10. Sheaves on projective schemes

equal so this defines an actual global section of M(n). We see that this is a

graded homomorphism. Moreover, it is functorial in M .

Lemma 10.15. If R is a noetherian integral domain generated in degree 1. Then

R′ = Γ∗(OX) is integral over R.

Proof. Let x1, . . . , xr be degree 1 generators of R. Let α : R→ Γ∗(OX), be the

map above, and let s ∈ R′ be a homogeneous element of non-negative degree.

We can find an n > 0, so that α(xni )s ∈ α(R) for every i. Rm is generated by

monomials in xi of degree m, so α(Rm)s ⊆ α(R) for m large (e.g., m ≥ rn. Let

R≥rn be the ideal of R generated by elements of degree ≥ rn. We have that

α(R≥rn)s ⊆ α(R≥rn). Moreover, since R is noetherian, R≥rn is finitely generated,

so applying Cayley–Hamilton, we get that s satisfies an integral equation over R.

Hence R′ is integral over R.

10.4.3 The map β

Let F be an OX-module. We will define a map of OX-modules

β : Γ∗(F)→ F

as follows. Let f ∈ R1. We will define β over D+(f). A section of Γ∗(F) is

represented on D+(f) by a fraction m/fd where m ∈ Γ(X,F(d)). If we think

of f−d as a section in O(−d)(D+(f)), then we can consider the tensor product

m⊗ f−d which is a section of F via the isomorphism F(d)⊗ O(−d) ' F . This

is compatible with the module structures, so we obtain a homomorphism of

OX-modules

β : Γ∗(F)→ F

by associating m/fd to m⊗ f−d.

Proposition 10.16. Suppose R is a graded ring, finitely generated in degree 1

over R0. Suppose F is a quasi-coherent sheaf on ProjR. Then the map

β : Γ∗(F)→ F (10.4.1)

is an isomorphism.

Proof. Since R is generated by R1 over R0, the open sets D+(f) with f ∈ R1

cover X. To show that (10.4.1) is an isomorphism, it is sufficient to prove it on

such an open.

Let f ∈ R1, and consider it as a section of Γ(X,O(1)). Then taking L = O(1)

in Lemma 10.8, (i) there says that if an element s of Γ(D+(f),F) is given, we

167

Page 168: Geir Ellingsrud and John Christian Ottem October 15, 2018€¦ · between algebra and geometry: a ne varieties correspond to nitely generated k-algebras without zero-divisors, and

10.4. The associated graded module

can find some element t of Γ∗(F)N (for N sufficiently large) that t ⊗ f−N ∈Γ(D+(f),F) equals s. This implies that the map (10.4.1) is surjective.

For injectivity, suppose s ∈ Γ(X,F(n)) is such that s⊗ f−n = 0 on D+(f),

i.e. s/fn ∈ Γ∗(F)(f) is in the kernel of (10.4.1) on the D+(f)-sections. Then

the lemma implies that there is a power fN with s⊗ fN ∈ Γ(X,F(n+N)) = 0.

This states that s/fn = 0 in Γ∗(F)(f) by the definition of localization and so the

map is injective.

We have defined two functors

∼: GrModR → QCohX

and

Γ∗ : QCohX → GrModR

Since β : Γ∗(F)→ F is an isomorphism, it follows that ∼ is essentially surjective.

However, unlike the affine case, the functors do not give mutual inverses. This is

because, as we have seen, that ∼ is not faithful; the ∼ of any module M which

is finite over R0 is the zero sheaf.

It turns out that if we restrict to a subcategory of GrModR, the category of

saturated R-modules, we do get an equivalence. Here we say that a graded R

module is saturated if α : M → Γ∗(M) is an isomorphism. This makes sense, as

the module M = Γ∗(F) is a saturated R-module. We denote this category by

GrModsatR . We then have a diagram

GrModRM 7→M

&&α

QCohX

F7→Γ∗(F)xxGrModsatR

KK

We have seen that the map β is an isomorphism, so any quasi-coherent sheaf

F is isomorphic to the ∼ of some module (Γ∗(F)).

Putting everything together, we find

Theorem 10.17. Let R be a graded ring, finitely generated in degree 1 over R0

and let X = ProjR. Then the functors

∼: GrModsatR → QCohX

and

Γ∗ : QCohX → GrModsatR

168

Page 169: Geir Ellingsrud and John Christian Ottem October 15, 2018€¦ · between algebra and geometry: a ne varieties correspond to nitely generated k-algebras without zero-divisors, and

Chapter 10. Sheaves on projective schemes

are mutually quasi-inverse, and establish an equivalence of categories.

So we have a full classification of quasi-coherent sheaves on X in terms of

modules on R.

10.5 Closed subschemes of ProjR

The notion of a saturated module comes from commutative algebra and has a

more conceptual definition there. To give some flavour of what it entails, we will

explain what this means for a graded ideal I ⊆ R. Here the saturation of I with

respect to an ideal B is defined as the ideal

I : B∞ :=⋃i≥0

I : Bi = r ∈ R|Bnr ∈ I for some n > 0.

We say that I is B-saturated if I = I : B∞ and in our case, saturated if it is

R+-saturated. We will here denote I : (R+)∞ by I. Note that the ideal I is

homogeneous if I is.

Example 10.18. In R = k[x0, x1], the (x0, x1)-saturation of (x20, x0x1) is the

ideal (x0). Note that both (x0) and (x20, x0x1) define the same subscheme of

P1k, but in some sense the latter ideal is inferior, since it has a component in

the irrelevant ideal (x0, x1). This example is typical; the saturation is a process

which throws away components of I supported in the irrelevant ideal.

Proposition 10.19. Let A be a ring and let R = A[x0, . . . , xn].

(i) To each subscheme Y of PnA, there is a corresponding homogeneous saturated

ideal I ⊆ R. Such that Y corresponds to the subscheme Proj(R/I) →ProjR.

(ii) Two ideals I, J defined the same subscheme if and only if they have the

same saturation.

(iii) If Y ⊆ PnA is a closed subscheme with ideal sheaf I, then Γ∗(I) is a

saturated ideal of R. In fact, the ideal Γ∗(I) is the largest ideal that defines

the subscheme Y .

In particular, there is a 1-1 correspondence between closed subschemes i : Y → PnAand saturated homogeneous ideals I ⊆ R.

Proof. (i) Let i : Y → PnA be a subscheme of PnA = ProjR and let I⊆OPnAdenote the ideal sheaf of Y . Using the fact that global sections is left-exact,

we have Γ∗(I) ⊆ Γ∗(OPnA). This is naturally an R-module, so I = Γ∗(I) is a

(homogeneous) ideal of R.

169

Page 170: Geir Ellingsrud and John Christian Ottem October 15, 2018€¦ · between algebra and geometry: a ne varieties correspond to nitely generated k-algebras without zero-divisors, and

10.6. The Segre embedding

Any such ideal I gives rise to a closed subscheme i′ : Proj(R/I)→ PnA and

hence an ideal sheaf J satisfying I = J . By Proposition 10.16, we also have

I = IY , so the two quasi-coherent sheaves coincide and i is indeed the same as

i′. By construction I = Γ∗(I), so I is saturated.

(ii) If I, J define the same subscheme, they have the same ideal sheaf I = Jon PnA. Let s ∈ Id, then on Ui = D+(xi), the fraction rx−di defines an element

of Γ(Ui, I) = Γ(Ui, J). Since also I corresponds to J , we have rx−di = tix−di for

some ti ∈ Jd of degree d. Hence there is a power ni such that xnii (r − ti) = 0 in

R. This shows that r is in the saturation of J . By symmetry, we have I = J .

(iii) Let r ∈ R be such that xnii r ∈ Γ∗(I) (that is r ∈ Γ∗(I)). Let m = maxni.

We want to show that r ∈ Γ∗(I)d = Γ(X, I(d)). On Ui = D+(xi), we see that

x−mi ⊗ xmi r defines a section of I(d + m) ⊗ O(−n). The latter is isomorphic

to I(d) and x−mi ⊗ xmi r = r via this isomorphism. So r ∈ Γ(Ui, I(d)). Hence

r ∈ Γ(X, I(d)) ⊆ Γ∗(I), and Γ∗(I) is saturated.

10.6 The Segre embedding

Theorem 10.20. Let R,R′ be graded rings with R0 = R′0 = A. Let S =⊕n≥0(Rn ⊗R′n). Then

ProjS ' ProjR×A ProjR′

Proof. Let X = ProjR, Y = ProjR′ and Z = ProjS. Let f ∈ R be a

homogeneous element. Define

Zf =⋃g∈R′

SpecSfdeg g⊗gdeg f

We claim that there is a natural isomorphism

Sf ′⊗g′ → R(f) ⊗A R′(g)r ⊗ r′

(f ′ ⊗ g′)s7→ r

f ′s⊗ r′

g′s

where f ′ = fdeg g and g′ = gdeg f . Indeed, the inverse is given by the map

R(f) ⊗A R′(g) → Sf ′⊗g′ defined by

r

f r⊗ r′

gt7→ rtdeg g ⊗ r′r deg f

(f ′ ⊗ g′)rt

170

Page 171: Geir Ellingsrud and John Christian Ottem October 15, 2018€¦ · between algebra and geometry: a ne varieties correspond to nitely generated k-algebras without zero-divisors, and

Chapter 10. Sheaves on projective schemes

Hence we see that

Zf =⋃g∈R′

Spec(R(f) ⊗A R′(g))

On the overlaps, we have

Spec(R(f) ⊗A R′(g)) ∩ Spec(R(f) ⊗A R′(h)) = Spec(Sfdeg g+deg h⊗(gh)deg f

)= Spec

(R(f) ⊗A R′gh

)From this is is clear that

Zf = D+(f)×R Y

Moreover, for any other f ′ ∈ R we have Zf ′ = Zff ′ = X(ff ′) ×R Y . Hence

Z =⋃f∈R

Zf = X ×R Y.

Corollary 10.21. Let A be a ring and let m,n ≥ 1 be integers. Then there is a

closed immersion

σ : PmA ×A PnA → Pmn+m+n

Proof. Consider theA-algebra S =⊕

n≥0(Rn⊗R′n) above, whereR = A[x0, . . . , xm]

and R′ = A[y0, . . . , yn] are the polynomial rings. Consider the following morphism

of graded A-algebras.

A[zij]0≤i≤m,0≤j≤n → A[x0, . . . , xm]⊗ A[y0, . . . , yn]

zij 7→ xi ⊗ yj.

It is clear that S is generated as an R0 ⊗R′0-algebra by the products xi ⊗ yj , so

the map is surjective and thus we get the desired closed immersion.

Example 10.22. Let R = k[x0, x1], R′ = k[y0, y1]. Then uij = xiyi defines an

isomorphism

S =⊕n≥0

(Rn ⊗R′n)→ k[u00, u01, u10, u11]/(u00u11 − u01u10)

This recovers the usual embedding of P1k ×k P1

k as a quadric surface in P3k.

171

Page 172: Geir Ellingsrud and John Christian Ottem October 15, 2018€¦ · between algebra and geometry: a ne varieties correspond to nitely generated k-algebras without zero-divisors, and

10.7. Two important exact sequences

A smooth quadric surface

10.7 Two important exact sequences

10.7.1 Hypersurfaces

Let R = k[x0, . . . , xn] and Pnk = ProjR. Let F ∈ R denote a homogeneous

polynomial of degree d > 0. F determines a projective hypersurface X = V (F ),

which has dimension n− 1. Note that I(X) = (F ) by the Nullstellensatz. We

then have an isomorphism

R(−d)→ I(X)

given by multiplication with F . Note the shift here: The constant ‘1’ gets sent

to F should have degree d on both sides! This gives the sequence of R-modules

0→ R(−d)→ R→ R/(F )→ 0

We have R(−d) = OPnk (−d) and (R/F ) = i∗OX , where i : X → Pnk is the

inclusion, so we get the exact sequence of sheaves

0→ OPnk (−d)→ OPnk → i∗OX → 0

10.7.2 Complete intersections

Let F,G be two homogeneous polynomials without common factors of degrees

d, e respectively. Let I = (F,G) and X = V (I) ⊆ Pnk . X is called a ‘complete

intersection’ – it is the intersection of the two hypersurfaces V (F ) and V (G). To

study X we have exact sequences

0→ R(−d− e) α−→ R(−d)⊕R(−e) β−→ I → 0

and applying ∼:

0→ OPnk (−d− e)→ OPnk (−d)⊕ OPnk (−e)→ IX → 0

172

Page 173: Geir Ellingsrud and John Christian Ottem October 15, 2018€¦ · between algebra and geometry: a ne varieties correspond to nitely generated k-algebras without zero-divisors, and

Chapter 10. Sheaves on projective schemes

The maps here are defined by α(h) = (−hG, hF ) and β(h1, h2) = h1F + h2G.

These maps preserve the grading. To prove exactness, we start by noting that α

is injective (since R is an integral domain) and β is surjective (by the defintion

of I). Then if (h1, h2) ∈ ker β, we have h1F = −h2G, which by the coprimality

of F,G means that there is an element h so that h1 = −hG, h2 = hF .

These sequences are fundamental in computing the geometric invariants from

X. We will see several examples of this later.

173

Page 174: Geir Ellingsrud and John Christian Ottem October 15, 2018€¦ · between algebra and geometry: a ne varieties correspond to nitely generated k-algebras without zero-divisors, and
Page 175: Geir Ellingsrud and John Christian Ottem October 15, 2018€¦ · between algebra and geometry: a ne varieties correspond to nitely generated k-algebras without zero-divisors, and

Chapter 11

Differentials

So far we have defined schemes and surveyed a few of their basic properties

(e.g., how to study sheaves on them). In this chapter, we introduce Kahler

differentials, which allow us in some sense to do calculus on schemes. This in

turn will allow us to define the most important sheaves in algebraic geometry,

namely, the cotangent sheaf, the tangent sheaf, and the sheaves of n-forms.

Differentials appear prominently throughout many areas of mathematics,

e.g., multivariable analysis, manifolds and differential geometry. In algebraic

geometry they are introduced algebraically using their formal properties and are

usually refered to as Kahler differentials after the German mathematician Erich

Kahler (1906–2000).

11.1 Tangent vectors and derivations

Consider a real manifold M of dimension n. To each point p ∈M , we can attach

its tangent space of TpM , which is a real vector space of dimension n. When M

is embedded as a submanifold in some RN , we can think of tangent vectors v

as vectors in the ambient space RN that ‘stick out’ from p, and naively define

TpM as the vector space of vectors that tangentially pass through p. However,

we do not like to think of manifolds as embedded in some RN , and then giving a

precise definition of TpM becomes a little bit subtle.

When defining tangent spaces in the abstract setting, there are a few basic

properties we want. First of all, each R-vector space TpM should have dimension

n, and the elements should have some sort of geometric interpretation. Moreover,

the assignment p→ TpM should vary continuously in p, so that we can be able

175

Page 176: Geir Ellingsrud and John Christian Ottem October 15, 2018€¦ · between algebra and geometry: a ne varieties correspond to nitely generated k-algebras without zero-divisors, and

11.1. Tangent vectors and derivations

to talk about continuous families of tangent vectors (or vector fields). The most

precise way to phrase this is to say that the collection TM = TpMp∈M should

be a vector bundle, i.e., TM has the structure of a 2n-dimensional manifold,

and there is a projection map π : TM →M sending (p, v) tp p, which is locally

diffeomorphic to U ×Rn in a neighbourhood of p. This is what is usually known

as the tangent bundle. In this setting a vector field is simply a map ν : M → TM

of π so that π ν = idM .

There are a number of ways to rigorously define the tangent bundle TM .

The simplest, and perhaps most flexible method is using derivations. Here we

think about tangent vectors as ways to differentiate functions defined at p. So we

consider the local ring Op of C∞-functions defined at p and think about elements

of Tp as linear functionals d : Op → R, satisfying the Leibniz rule

d(fg) = fd(g) + gd(f).

Note that the set of such derivatives naturally form a real vector space; we can

add them and multiply them with real numbers and the result will be a new

derivation.

In the prototype example, when M ⊆ RN is an embedded submanifold, we

can consider the directional derivative

∇vf = limh→0

f(p+ hv)− f(p)

h.

which we may think of as a derivation C∞(M,R) → R. It is not hard to see

that every derivation arises this way, and that conversely a tangent vector v is

determined by the functional ∇v.

An alternative viewpoint to tangent spaces uses the maximal ideal m ⊆ Op

of functions vanishing at the point p. We would like to consider functions f

vanishing at p, but identify two functions f1, f2 if they have the same linear

Taylor expansion around p, i.e., if f1 − f2 ∈ m2. In this way we can define the

cotangent space m/m2, and then set Tp = (m/m2)∨ (vector space dual). That is,

176

Page 177: Geir Ellingsrud and John Christian Ottem October 15, 2018€¦ · between algebra and geometry: a ne varieties correspond to nitely generated k-algebras without zero-divisors, and

Chapter 11. Differentials

we can think about elements of Tp as linear functionals

v : m/m2 → R.

It is not hard to see that these two perspectives give the same vector space.

Given a derivation d : Op → R, we must have d(m2) = 0, and so d induces via

restriction a map v : m/m2 → R. Conversely, any element of (m/m2)∨ arises this

way.

So for instance, ifM has local coordinates x1, . . . , xn near p, we let dx1, . . . , dxndenote their images in m/m2. Each xi determines dually a derivation ∂

∂xi: Op →

R, so that ∂∂x1, . . . , ∂

∂xnforms a basis Tp. Note that by definition, ∂

∂xi(dxi) = 1

and ∂∂xi

(dxj) = 0 for i 6= j.

11.2 Regular schemes and the Zariski tangent

space

Let X be a scheme and let x ∈ X be a point. We have the local ring R = OX,x

with maximal ideal m. Motivated by the case of manifolds, we will then consider

mx/m2x, which is in a natural way a vector space over the residue field k(x) =

R/mx. This is what we would like to define as the cotangent space of X at x.

Intuitively, the cotangent space is what you get by differentiating functions which

vanish at that point, but differentiating functions that vanish at x at higher

order should give zero.

Definition 11.1. The cotangent space of X at x is the quotient mx/m2x.

The Zariski tangent space TX,x to X at x is the dual vector space of mx/m2x.

That is,

TX,x = Homk(x)(mx/m2x, k(x))

and an element of TX,x is called a tangent vector; it is a linear functional

mx/m2x → k(x).

X is called regular at x if the local ring OX,x is a regular local ring, i.e.,

dim OX,x = dimk(x) m/m2.

X is regular if this condition holds for every x ∈ X.

Example 11.2. A1k and P1

k are both regular schemes. Indeed, in both cases, the

local ring at a point is isomorphic to a localization of a polynomial ring, which

is regular.

A non-normal example is X = SpecA where A = k[x, y]/(x2 − y3). In

the point m = (x, y) corresponding to the origin, the local ring is not normal:

177

Page 178: Geir Ellingsrud and John Christian Ottem October 15, 2018€¦ · between algebra and geometry: a ne varieties correspond to nitely generated k-algebras without zero-divisors, and

11.2. Regular schemes and the Zariski tangent space

x/y satisfies a monic equation, (x/y)3 − x = 0, but it is not contained in the

localization Ap. In this example, the vector space m/m2 is 2-dimensional at

the origin – it is spanned by the residue classes of x and y (which are linearly

independent since all relations in A happen in degree at least two).

11.2.1 Zariski tangent space and the ring of dual num-

bers

Let k be a field. The ring R = k[ε]/(ε2) is called the ring of dual numbers over k.

The spectrum of R is a very simple scheme: Its underlying topological space is

a single point. However, the non-reduced structure on SpecR shows that it is

more interesting than Spec k. We picture it as follows:

Proposition 11.3. Let X be a scheme over k. To give a k-morphism Spec(k[ε]/ε2)→X is equivalent to giving a point x ∈ X which is rational over k (meaning that

k(x) = k), and an element of TX,x.

Proof. Let Y = Spec(k[ε]/(ε2). The map Y → X is a map of schemes over k,

and so f# induces a map of k-algebras:

OX,x// k[ε]/ε2//

k

OO ::

This is a local homomorphism, so we can mod out by the ideals mx and (ε) to

get

k(x) // k

k

OO ==

where the diagonal arrow is an isomorphism; hence k(x) = k and x is a k-

rational point. Moreover, restricting f#x to mx ⊆ OX,x we get a k-algebra map

178

Page 179: Geir Ellingsrud and John Christian Ottem October 15, 2018€¦ · between algebra and geometry: a ne varieties correspond to nitely generated k-algebras without zero-divisors, and

Chapter 11. Differentials

α : mx → (ε) ' k. This has to vanish on m2x, and hence induces an element in

TX,x.

Conversely, suppose that x ∈ X is a k-rational point and let α : m/m2 → k

be a k-linear map (m is the maximal ideal in OX,x). We want to construct

(f, f#) : (Y,OY ) → (X,OX). Define f((ε)) = x; this takes care of the map on

topological spaces. Now we define f#. Set for U ⊆ X not containing x, we

define f# : OX(U) → OY (∅) = 0 to be the zero map. For U containing x, we

argue as follows. The point is that each OX(U) has a structure of a k-algebra

(ie there is a map k → OX(U)), and so we can write any element a ∈ OX,x as

a = a(x) + (a− a(x)) where a(x) is the image of a via

OX,x → OX,x/mx → k → OX,x.

Note that the germ of a − a(x) in OX,x lies in mx. Define f# : OX(U) →OY (f−1(U)) = k[ε]/ε2 by sending a to the class of

a(x) + α((a− a(x))|x + m2)ε

One can check that this is a ring homomorphism, using k-linearity and the fact

that the square of (a− a(x))|x lies in m2. Everything here is compatible with

the inclusions V ⊆ U , and so sets up a map of sheaves f#.

11.3 Derivations and Kahler differentials

So far we have considered a scheme X and a point x ∈ X, and the cotangent

space m/m2 at x. We would like to generalize this construction, and instead of

fixing x, consider all points at once. That is, we would like to form a sheaf on X

with stalks m/m2 at each point. This is what will be the cotangent sheaf on X.

We will work over a base ring A. B will be an A-algebra and M an B-

module. The picture to have in mind is that A = k, where k is a field, and

X = SpecB → Spec k, where B is a k-algebra.

Definition 11.4. An A-derivation (from B with values in M) is an A-linear

map d : B →M satisfying the product rule:

d(b1b2) = b1 d(b2) + b2 d(b1).

Given that the product rule holds, it is easy to see that d is A-linear if and

only if d vanishes on all elements of the form a · 1 where a ∈ A. We can therefore

think of the elements in B of the form a · 1 as the ‘constants’ – however a

derivation can also be zero on other elements on B.

Example 11.5. A simple example is the classical derivation of polynomial ring

179

Page 180: Geir Ellingsrud and John Christian Ottem October 15, 2018€¦ · between algebra and geometry: a ne varieties correspond to nitely generated k-algebras without zero-divisors, and

11.3. Derivations and Kahler differentials

B = k[x] over k to itself given by P (x) 7→ P ′(x). As usual, P (x) must be defined

formally by P (x) =∑

i iaixi−1 if P (x) =

∑i aix

i.

More generally, the differential operators ∂∂x1. . . . , ∂

∂xn, as well as their linear

combinations, define derivations on the polynomial ring k[x1, . . . , xn].

The set of derivations d : B →M is usually denoted with DerA(B,M). This

inherits a B-module structure from M and it is as such naturally a submodule of

HomA(B,M). This gives rise to a covariant functor DerA(B,−) from mod B to

itself. More precisely, if φ : M →M ′ is a B-module homomorphism, we can map

a derivation d ∈ DerA(B,M) to φ d : B →M ′, which is in turn an A-derivation

of B with values in M ′.

The derivation DerA(B,M) is also functorial in the base ringA and the algebra

B; in both of these cases, it is covariant. If A → A′ is a ring homomorphism,

any A′-derivation B → M is in turn an A-derivation. We therefore obtain an

inclusion DerA′(B,M) ⊆ DerA(B,M).

11.3.1 The module of Kahler differentials

The covariant functor Derk(A,−) on the category of A-modules is representable.

In down-to-earth terms, this just means that there exist a B-module ΩB/A and

an isomorphism of functors

DerA(B,−) ' HomB(ΩB/A,−).

By formal properties of functors and representability, this condition is equivalent

by having a universal derivation dB : B → ΩB/A, having the following property:

For any A-derivation d′ : B →M there exists a unique B-module homomorphism

α : ΩB/A →M such that d′ = α dB. In terms of diagrams, we have

BdB //

d′

ΩB/A

α

M

To see directly why such a module exists, we as construct it explicitly as follows.

ΩB/A is the A-module defined as the quotient of the free module⊕

b∈B Bdb by

the submodule generated by the expressions d(b+b′)−db−db′, d(bb′)−bdb′−b′db,da, for b, b′ ∈ B, a ∈ A. Then the universal derivation dB : B → ΩB/A is given

by b 7→ db.

It is not hard to see that this module indeed satisfies the property above:

Given an A-derivation D : B →M , we define the homomorphism α : ΩB/A →M

180

Page 181: Geir Ellingsrud and John Christian Ottem October 15, 2018€¦ · between algebra and geometry: a ne varieties correspond to nitely generated k-algebras without zero-divisors, and

Chapter 11. Differentials

by α(db) = D(b) (which is well-defined precisely because D is a derivation!).

Conversely, given α : ΩB/A → M , we get a derivation D : B → M defined by

D(b) = φ(db).

Definition 11.6. The elements of the module ΩB/A are called the Kahler

differentials, or simply differentials of B over A.

Example 11.7. Let B = k[t] be the polynomial ring over k in the variable t.

Then ΩB/k is a free module over k generated by dBt, i.e., ΩB/k = B · dBt. This

follows since an element m in an arbitrary A-module M uniquely determines

a derivation A→ M with dt = m, and so there is an isomorphism of functors

Derk(B,−) = Hom(B,−). By the Yoneda lemma, it follows that ΩB/k = BdBt.

Alternatively, we can see this from the explicit construction of Ω1A/k, because

for any f ∈ k[t], we have df = f ′(t)dt.

More generally we have:

Proposition 11.8. Let A be any ring and let B = A[x1, . . . , xn]. Then ΩB/A is

the free B-module generated by dx1, . . . , dxn.

The universal derivation d : B → ΩB/A is the map f 7→∑ ∂f

∂xidxi. Since B is

generated as an A-algebra by the xi, ΩB/A is generated as a B-module by the

dxi and there is a surjection Br → ΩB/A taking the ith basis vector to dxi.

On the other hand, the partial derivative ∂/∂xi is an A-linear derivation from

B to B, and thus induces a B-module map ∂i : ΩB/A → B carrying dxi to 1 and

all the other xj to 0. Putting these maps together we get the inverse map.

Example 11.9. If B = S−1A is a localization of A, then ΩB|A = 0. Indeed, take

b ∈ B, and choose s ∈ S so that sb ∈ A, and hence sdb = d(sb) = 0 This implies

that db = 0, since s is invertible in B.

Example 11.10. If B = A/I, then ΩB|A = 0. More generally, if φ : A→ B is

surjective, then ΩB|A. This follows because if b = φ(a), then db = a · d(1) = 0 in

ΩB|A.

Example 11.11. Suppose k is a field of characteristic p > 0. Let B = k[x] and

let A = k[xp]. Then ΩB/A is the free B-module of rank 1 generated by dx.

Example 11.12. Let B = Q[√

2] and A = Q. Then 0 = d(2) = d(√

2√

2) =

2√

2d(√

2) so d(√

2) = 0. Thus ΩQ[√

2]/Q = 0.

11.4 Properties of Kahler differentials

There are a few useful ways for computing modules of differentials when changing

rings. The proofs of the following propositions are not difficult, and involve only

basic commutative algebra.

181

Page 182: Geir Ellingsrud and John Christian Ottem October 15, 2018€¦ · between algebra and geometry: a ne varieties correspond to nitely generated k-algebras without zero-divisors, and

11.4. Properties of Kahler differentials

11.4.1 Localization

The aim of this section is to show that Kahler differentials behave very well

under localization.

Proposition 11.13. Let S ⊆ A be a multiplicative subset mapping into the

group of units in B. Then

ΩB|S−1A = ΩB|A

Proof. We have a natural mapA→ S−1A, so we get DerS−1(B,M) ⊆ DerA(B,M).

Conversely, if d : B → B is an A-derivation, then for each s ∈ S, we have

0 = d(1) = sd(1/s) + 1/sd(s) = sd(1/s)

As s is invertible in B, it follows that d(1/s) = 0, and hence by the product rule,

d is also an S−1A-derivation of B.

Proposition 11.14. Suppose S is a multiplicative system in B and let i : B →S−1B be the localization map. Then

di : ΩB/A⊗A S−1B → ΩS−1B/A

is an isomorphism

Proof. Using the usual formula of differentiating a fraction, we can define an

A-derivation d : S−1B → ΩB/A⊗A S−1A. This will give rise to an S−1B-module

homomorphism ΩS−1B/A → ΩB/A⊗A S−1B which will turn out to give the inverse

to di.

Let us first define the derivation d. We define

d(b/s) = (sdB(b)− adB(s))s−2

We need to check that this does not depend on the representative a/s. Using

the product rule for dB, we have

d(b1b2/st) = b1s−1d(b2/t) + b2t

−1d(b1/s), (11.4.1)

Suppose now that tb1 = sb2. Using this, the product rule, the identity tdBb1 +

b1dBt = sdBb2 + b2dBs, and solving for tdBb1 and substituting into (11.4.1), we

find

t2(sdBb1 − b1dBs) = ts(sdBb2 + b2dBs− b1dBt)− t2b1dBs = s2(tdBb2 − b2dBt)

where the last equality follows by tb1 = sb2. This shows that d(a/s) = d(b/t).

Finally, if b1s−1 = b2t

−1, there is an element u ∈ S such that utb1 = usb2. We

182

Page 183: Geir Ellingsrud and John Christian Ottem October 15, 2018€¦ · between algebra and geometry: a ne varieties correspond to nitely generated k-algebras without zero-divisors, and

Chapter 11. Differentials

then find

d(b2/t)u−1 + b2t

−1d(1/u) = d(b2/tu) = d(b1/su) = d(b1/s)u−1 + b1s

−1d(1/u)

and since b1s−1 = b2t

−1, it follows that d(b1/s) = d(b2/t), and d is well defined.

Moreover, it is an A-derivation by (11.4.1).

11.4.2 Base change

Proposition 11.15. Let A,B be as above and let A′ be an A-algebra. Define

B′ = B ⊗A A′. Then there is a canonical isomorphism

ΩB′|A′ ' ΩB|A ⊗B B′

Proof. The universal derivation d : B → ΩB|A induces an A′-derivation

d′ = d⊗ idA′ : B′ → ΩB|A ⊗A A′ = ΩB|A ⊗B B′

One can check that this is the required universal derivation of ΩB′|A′ which gives

the claim.

11.4.3 Exact sequences

Let A be a ring and let ρ : B → C be a homomorphism of A-algebras. Then

there are universal homomorphisms of (C-modules)

α : ΩB|A ⊗B C → ΩC|A

defined by α(db⊗ c) = cdρ(b). And

β : ΩC|A → ΩC|B

Proposition 11.16. Then there is an exact sequence of C-modules

ΩB/A ⊗B Cα−→ ΩC/A

β−→ ΩC/B → 0.

Proof. It is sufficient to show that for any C-module N , the dual sequence

0→ HomC(ΩC|B, N)→ HomC(ΩC|A, N)→ HomC(ΩB/A ⊗B C,N)

is exact. Note that HomC(ΩB/A ⊗B C,N) = HomB(ΩB/A, N), so the sequence

can be written as

0→ DerB(C,N)→ DerA(C,N)→ DerA(B,N)

183

Page 184: Geir Ellingsrud and John Christian Ottem October 15, 2018€¦ · between algebra and geometry: a ne varieties correspond to nitely generated k-algebras without zero-divisors, and

11.4. Properties of Kahler differentials

The map on the right hand side is the one induced by composition A→ B → C.

The fact that it is exact on the left is clear: If d : C → N maps to the zero

derivation, it is the zero homomorphism and hence 0 in DerB(C,N). To check

exactness in the middle, note that if d : C → N is mapped to the zero derivation

in DerA(B,N), then d : C → N is in fact B-linear: This follows by the Leibniz

rule d(bc) = b ·dc+db · c = b ·dc+ 0 = bdc. Hence d is in fact a B-derivation.

Applying this to C = S−1B, we obtain a new proof of Proposition 11.14 that

S−1ΩB/A ' ΩS−1B|A

for the localization in a multiplicative subset S ⊆ B.

Proposition 11.17 (Conormal sequence). Suppose that B is an A-algebra, and

C = B/I for some ideal I ⊆ B. Then there is an exact sequence of C-modules

I/I2 d−→ ΩB/A ⊗B Cα−→ ΩC/A → 0

Here the first map sends [f ] 7→ df ⊗ 1 and the second sends c⊗ db 7→ cdb.

Proof. Note that I/I2 = I ⊗B C. As in the previous proposition it suffices to

check that for each C-module N , the dual sequence

0→ DerA(C,N)→ DerA(B,N)→ HomC(I/I2, N) = Hom(I,N)

is exact. The rightmost map associates to a derivation d′ : B → N its restriction

to I. Note that this is indeed a homorphism of C-modules since IN = 0.

That this sequence is exact is a matter of checking, the case of exactness at

the left being easy. To check exactness in the middle, we suppose that d : B → N

restricts to 0 on I. Then d factors via a homomorphism d′ : B/I → N , which is

C-linear (since d is B-linear). This is the required element in DerA(C,N).

Corollary 11.18. Let A be a ring and let B be a finitely generated A-algebra

(or a localization of such). Then ΩB|A is finitely generated over B.

Proof. Write B = A[x1, . . . , xn]/I for some variables x1, . . . , xn and apply the

above propositions.

11.4.4 The diagonal and ΩB/A

Suppose now that B is an A-algebra. There is an exact sequence of A-modules

0→ I → B ⊗A B∆−→ B → 0

184

Page 185: Geir Ellingsrud and John Christian Ottem October 15, 2018€¦ · between algebra and geometry: a ne varieties correspond to nitely generated k-algebras without zero-divisors, and

Chapter 11. Differentials

where ∆ is the diagonal map b1⊗b2 7→ b1b2 and I is the kernel of ∆. We can make

I/I2 into a B-module by letting B act on the first factor by b(x⊗ y) = (bx)⊗ y)

and then define a map d : B → I/I2 by db = 1⊗ b− b⊗ 1.

Proposition 11.19. The module I/I2 along with the map d is the module of

differentials for B over A. E

Proof. Let us check that the map d is indeed a derivation: Suppose b1, b2 ∈ B,

then

d(b1b2) = 1⊗ b1b2 − b1b2 ⊗ 1.

One the other hand,

b1db2 + b2db1 = b1(1⊗ b2 − b2 ⊗ 1) + b2(1⊗ b1 − b1 ⊗ 1)

= b1 ⊗ b2 − b1b2 ⊗ 1 + b2 ⊗ b1 − b1b2 ⊗ 1

The difference between these two elements is

1⊗ b1b2 − b1b2 ⊗ 1− b1 ⊗ b2 + b1b2 ⊗ 1− b2 ⊗ b1 + b1b2 ⊗ 1

= (1⊗ b1 − b1 ⊗ 1)(1⊗ b2 − b2 ⊗ 1) ∈ I2

Now one has to check that (I/I2, d) satisfies the required universal property. We

leave this as a task for the reader (See [Matsumura p.192] for details).

11.5 Some explicit computations

We can use the previous exact sequences to do explicit computations with

ΩB/A. If B is a finitely generated A-algebra, say B = A[x1, . . . , xr]/I where

I = (f1, . . . , fr). Then we have

B⊗A ΩA[x1,...,xr]/A 'n⊕i=1

Bdxi

Note that I/I2 is generated by the classes of the f1, . . . , fr modulo I2. This gives

a surjection Br → I/I2. Now, by the conormal sequence (Proposition 11.17), we

have

ΩB/A = coker(d : I/I2 →n⊕i=1

Bdxi)

= coker(d : Bn →n⊕i=1

Bdxi)

185

Page 186: Geir Ellingsrud and John Christian Ottem October 15, 2018€¦ · between algebra and geometry: a ne varieties correspond to nitely generated k-algebras without zero-divisors, and

11.5. Some explicit computations

Explicitly, the map Bn →⊕n

i=1Bdxi is given by the n × r Jacobian matrix

J =(∂fj∂xi

).

Theorem 11.20. Let A be a ring, and let B = A[x1, . . . , xn]/(f1, . . . , fr). Then

ΩB|A =

⊕iBdxi∑

j B(∑

i∂fj∂xidxi

)together with the A-derivation dB|Af =

∑i∂f∂xidxi.

Example 11.21. Let k be a field and let X = SpecR where R = k[x, y]/(f).

Let us compute the module of differentials. Let fx, fy denote the (images of the)

partial derivatives of f in R. Then

ΩR|A =Rdx⊕Rdy

(fxdx+ fydy)

If X is non-singular, e.g., V (f, fx, fy) = ∅, then ΩR|A is locally free of rank 1.

The bases are given by

dy

fxon D(fx); and − dx

fyon D(fy).

(Note that on the overlap D(fx) ∩D(fy) we have fxdx+ fydy = 0 in ΩR|A).

Example 11.22. The curve X given by y2 = x2(x+1) in A2k is the so-called nodal

cubic. It has a singular point at the origin (0, 0). Let B = k[x, y]/(y2− x2(x+ 1).

Then

ΩB/k =Bdx⊕Bdy

(2ydy − (3x2 + 2x)dx)

In this case ΩX|k has rank 1 for every point (x, y) 6= (0, 0). At the origin, the

relation 2ydy − (3x2 + 2x)dx is identically zero, so ΩX|k has rank two there.

We can also view B as an algebra over A = k[x]. In that case, we get

ΩB/A =Bdy

(2ydy)

Example 11.23. Let B = k[x, y]/(x2 + y2). If k has characteristic 6= 2, then

ΩB|k = (Bdx+Bdy) /(xdx+ ydy)

If k has characteristic 2, then ΩB|k is the free B module Bdx+Bdy.

186

Page 187: Geir Ellingsrud and John Christian Ottem October 15, 2018€¦ · between algebra and geometry: a ne varieties correspond to nitely generated k-algebras without zero-divisors, and

Chapter 11. Differentials

11.6 The sheaf of differentials

For us, the primary motivation for studying ΩB/A is that it gives us an intrisic

module ΩB/A associated to a ring homomorphism A→ B. By taking ∼, we thus

get an intrinsic sheaf on X = SpecB associated to the map of affine schemes

SpecB → SpecA. We would like to globalize this construction to an arbitrary

morphism of schemes f : X → S. This will lead us to form the sheaf of relative

differentials ΩX|S which will be a quasicoherent OX-module.

This sheaf is locally built out of the various ΩB|A on local affine charts. These

not arbitrary modules that just happen to glue together to a sheaf; each of them

come with a universal property of classifying derivations d : B →M . For this

reason, we would like to say that the ΩX|Y should satisfy a similar universal

property. We make the following definition:

Definition 11.24. Let F be an OX module. A morphism d : OX → F of

sheaves is an S-derivation if for all open affine subsets V ⊆ S and U ⊆ X

with f(U) ⊆ V , the map dU |V is a OS(V )-derivation of OX(U). The set of all

S-derivations is denoted by DerS(OX , F ).

Definition 11.25. The sheaf of relative differentials is a pair (ΩX|S, dX|S) of

an OX-module ΩX|S and a S-derivation dX|S : OX → ΩX|S that satisfies the

following universal property: For any OX-module F , and each S-derivation

d : OX → F there exists a unique OX-linear map φ : ΩX|S → F such that

d = φ dX|S.

When S = SpecA, we sometimes write ΩX|A for ΩX|S.

In other words, ΩX|S is a sheaf that represents the sets of S-derivations, in

the sense that

HomOX (ΩX|S,F) = DerS(OX , F ).

As usual, the universal property gives that this sheaf is unique up isomorphism,

if it exists.

Theorem 11.26. Let f : X → S be a morphism of schemes. Then there is a

sheaf of relative differentials ΩX|S, which is a quasicoherent sheaf on X.

Moreover, ΩX|S has the property that for each open affine open V = SpecA,

and open affine U = SpecB ⊆ f−1(V ),

ΩX|S|U ' ΩB|A

Also for each x ∈ X,

(ΩX|S)x ' ΩOX,x|OS,f(x)

187

Page 188: Geir Ellingsrud and John Christian Ottem October 15, 2018€¦ · between algebra and geometry: a ne varieties correspond to nitely generated k-algebras without zero-divisors, and

11.6. The sheaf of differentials

Proof. Fix an open subset V = SpecA of S, and let U = SpecB be an affine

open subset in X so that f(U) ⊆ V . For these two, we define

ΩU |V = ΩB|A

on U . We first show that the ΩU |V glue to a Of−1V -module Ωf−1(V )|V . This

comes down to showing that if U ′ = SpecB′ is a distinguished open affine subset

of U , then

ΩU |V |U ′ ' ΩU ′|V .

But this follows from Proposition 11.14, as B′ is a localization of B.

Then we show that the sheaves Ωf−1V |V for all affine opens V ⊆ S glue to

a OX-module ΩX|S. This amounts to showing that for each distinguished open

V ′ = SpecA′ ⊆ V , and all open U = SpecB of f−1(V ′), we have

ΩU |V = ΩU |V ′

But this follows from Proposition 11.13, as A′ is a localization of A in a single

element (which maps to an invertible element in B).

This means that we get an OX-module ΩX|S. Let us check that it satisfies

the above universal property. So we need to define the universal derivation

dX|S : OX → ΩX|S.

Let V = SpecA ⊆ S and U = SpecB ⊆ X be an affine open subset such

that f(U) ⊆ V . Define dX|S(U) = dB|A. By the gluing construction above, this

map does not depend on the chosen affine open V , and it can be checked that

the rule is compatible with restriction maps. Hence this gives a morphism of

sheaves dX|S : OX → ΩX|S, which by construction is a S-derivation.

To check that this is universal, we again work locally. Let d : OX → Fbe a S-derivation, where F is an OX-module. Let U = SpecA ⊆ S and

V = SpecB ⊆ X so that f(U) ⊆ V . By the universal property of ΩB|A,

we get an A-derivation d(V ) : B → F(V ), and hence a unique B-linear map

φ(V ) : ΩX|S(V ) = ΩB|A → F such that d(V ) = φ(V )dX|V (V ). One has to check

that these maps are compatible with restriction maps (use the universal property

of ΩB/A), but after that, we obtain a unique OX-linear map φ : ΩX|S → F so

that d = φ dX|S.

Note that the sheaf ΩX|S is always quasi-coherent (it is by definition locally

of the form M for some module). Moreover, when X is of finite type over a field,

ΩB/A is finitely generated, and so ΩX|k is even coherent.

Example 11.27. Let A be a ring and let X = AnS = SpecS[x1, . . . , xn] be affine

n-space over S = SpecA. Then ΩX/S ' OnX is the free OX-module generated by

dx1, . . . , dxn.

188

Page 189: Geir Ellingsrud and John Christian Ottem October 15, 2018€¦ · between algebra and geometry: a ne varieties correspond to nitely generated k-algebras without zero-divisors, and

Chapter 11. Differentials

Remark 11.28. If X is a separated scheme over S then one could also define

ΩX/S as follows. Let ∆ : X → X ×S X be the diagonal morphism and let I∆

be the ideal sheaf of the image of ∆. Then ΩX/S = ∆∗(I∆/I2∆). This does in

fact give the sheaf above, since these two definitions coincide when X and S are

both affine (Section 11.4.4). This definition gives a quick way of obtaining the

sheaf ΩX/S, but it is not very well suited for computations.

The properties of ΩB/A translate into the following results for ΩX|Y :

Proposition 11.29 (Base change). Let f : X → S be a morphism of schemes

and let S ′ be a S-scheme. Let X ′ = X×S S ′ and let p : X ′ → X be the projection.

Then

ΩX′|S′ ' p∗ΩX|S

Proposition 11.30. Let X, Y , and Z be schemes along with maps Xf−→ Y

g−→ Z.

Then there is an exact sequence of OX-modules

f ∗(ΩY/Z)→ ΩX/Z → ΩX/Y → 0.

Proposition 11.31 (Conormal bundle sequence). Let Y be a closed subscheme

of a scheme X over S. Let IY be the ideal sheaf of Y on X. Then there is an

exact sequence of sheaves of OX-modules

IY /I2Y → ΩX/S ⊗OY → ΩY/S → 0.

11.7 The sheaf of differentials of PnAWe have seen that the cotangent sheaf of An is trivial, i.e., isomorphic to On

An . In

this sections we give a concrete description of the cotangent bundle of projective

space, suitable for explicit computations.

Theorem 11.32. Let X = PnA, then there is an exact sequence

0→ ΩX/A → OX(−1)n+1 → OX → 0.

Proof. Let R = A[x0, . . . , xn] so that X = ProjR. Let T be the R-module

R(−1)n+1 with basis elements e0, . . . , en. Define the map α : T → R by sending

ei 7→ xi and let M = kerα. We have an exact sequence

0→M → Tα−→ R

Taking ∼ this gives

0→ M → OX(−1)n+1 → OX → 0

189

Page 190: Geir Ellingsrud and John Christian Ottem October 15, 2018€¦ · between algebra and geometry: a ne varieties correspond to nitely generated k-algebras without zero-divisors, and

11.8. Relation with the Zariski tangent space

which is surjective on the right, because the map α is surjective in degrees 0

(the cokernel C of α is a sheaf with Cd = 0 for large d, and so it has C = 0). We

claim that M = ΩX/A.

Let Ui = D(xi) = SpecR(xi) denote the standard open affines. We have

ΩX/A|Ui = ΩUi/A = ΩR(xi)/A

and ΩR(xi)/A is generated as an R(xi)-module by the expressions d

(xjxi

)for i 6= j.

Now define the maps

φi : ΩX/Y |Ui → M |Uid(xjxi

)7→ 1

x2i(xiej − xjei)

Note that xiej − xjei ∈ kerα. The factor 1x2i

is included to obtain an element of

degree zero. This is an isomorphism on Ui, so we are done if we can make sure

that the φi glue. Now, on the overlaps Ui ∩ Uj, we have xkxi

= xkxj

xjxi

, which gives

d

(xkxi

)=xkxjd

(xjxi

)+xjxi

(xkxj

)or, in other words,

xjxid

(xkxj

)= d

(xkxi

)− xkxjd

(xjxi

)Applying φi to both sides, gives the same answer ( 1

xixj(xjek−xkej)), so the maps

φi are compatible.

Since ΩPnA injects into OPnA(−1)n+1 (which has no global sections), we get:

Corollary 11.33. Γ(PnA,ΩPnA) = 0

11.8 Relation with the Zariski tangent space

Let X be a scheme over an algebraically closed field k. Recall that for a point

x ∈ X, the Zariski tangent space is defined as m/m2 where m is the maximal

ideal in the local ring B = OX,x. In this section we relate this to the module of

differentials ΩR|k. Note that since k is algebraically closed, the Nullstellensatz

gives that B/m = k.

Proposition 11.34. Suppose (B,m) is a local ring with residue field k = B/m

190

Page 191: Geir Ellingsrud and John Christian Ottem October 15, 2018€¦ · between algebra and geometry: a ne varieties correspond to nitely generated k-algebras without zero-divisors, and

Chapter 11. Differentials

and there is an injection k → B. Then there is an exact sequence

m/m2 d−→ ΩB/k ⊗B k → 0

and d is actually an isomorphism.

Proof. The exact sequence is obtained from the conormal sequence by letting

A = C = k.

The above map sends x ∈ m to dx. We construct an inverse ψ : ΩB|k ⊗B k →m/m2 as follows. Constructing such a map is equivalent to constructing a map

of B-modules ΩB|k → m/m2, or equivalently, a derivation ∂ : B → m/m2. We

define ∂ by

∂(a+ x) = x

for a ∈ k and x ∈ m. (This is well-defined since k + m = B). To complete the

proof we need only prove that ∂ is a derivation:

∂((a+ x)(a′ + x′)) = ∂(aa+ (ax′ + a′x) + x′x)

= ∂(aa) + ∂(ax′ + a′x) + ∂(x′x)

= ax′ + a′x

We get the same answer when we expand

(a′ + x′)∂(a+ x) + (a+ x)∂(a′ + x′)

(since xx′ ∈ m2). Hence the map is a derivation, and we get a map ψ. This

is indeed an inverse to the map d, since via the identification DerA(B,M) =

HomB(ΩB|A, A), ψ sends dx to x

Corollary 11.35. Let (B,m) be as above. Then B is a regular local ring if and

only if

dimBm = dimk ΩB|k ⊗B k

Definition 11.36. Let X be a scheme over a field k and let x ∈ X be a point.

We say that X is smooth at x if ΩX|k is locally free at x. X is called smooth if it

is smooth at every point.

Theorem 11.37. Let X be a variety over an (algebraically closed) field k of

characteristic 0 and let x ∈ X. Then the following are equivalent:

(i) X is smooth at x

(ii) (ΩX|k)x is free of rank dimxX

(iii) X is non-singular at x.

191

Page 192: Geir Ellingsrud and John Christian Ottem October 15, 2018€¦ · between algebra and geometry: a ne varieties correspond to nitely generated k-algebras without zero-divisors, and

11.8. Relation with the Zariski tangent space

Proof. (i)⇐⇒ (ii) is clear.

(ii) =⇒ (iii). If ΩB|k is free of rank n = dimR, then by the above proposition,

we have dimk(x) m/m2 = n and so Bm is a regular local ring.

(iii) =⇒ (i). This direction requires some commutative algebra of ΩB|A we

haven’t covered. See [Matsumura p. 191] or [Hartshorne p. 174] for details.

Definition 11.38. When X is smooth and ΩX|k is a locally free sheaf on X,

and we refer to it as the cotangent bundle or cotangent sheaf of X.

The sheaf of p-forms is defined as

ΩpX =

p∧ΩX|k

In particular, if X has dimension n, ωX = ΩnX is called the canonical bundle of

X. As ΩX|k has rank n, ωX is locally free of rank 1, i.e., an invertible sheaf.

The tangent sheaf, or tangent bundle is the sheaf

TX = HomOX (ΩX|k,OX)

Thus for X = P1, we get ΩP1 = O(−2) and TP1 = OP1(2).

Proposition 11.39. Let X = Pn. Then ΩnX = O(−n− 1).

Proof. Consider the Euler sequence for the cotangent bundle of Pn

0→ ΩPn → O(−1)n+1 → O → 0

In general, if 0→ E ′ → E → E ′′ → 0, we have∧e E =

∧e′ E ′ ⊗∧e′′ E ′′. Hence

we getn∧

ΩPn =n+1∧

O(−1) = O(−n− 1).

Note by the way that the tangent bundle TPn fits into the following sequence:

0→ OPn → OPn(1)n+1 → TX → 0

where the left-most map sends 1 to the vector (x0, . . . , xn).

Example 11.40. Let A be a ring and let X = P1A. Then we have ΩX|A '

OP1(−2). For this, we can use the standard covering of P1 = ProjA[x0, x1], given

by Ui = D+(xi). On U0, we have

ΩU0|A ' A[x1

x0

]d

(x1

x0

)192

Page 193: Geir Ellingsrud and John Christian Ottem October 15, 2018€¦ · between algebra and geometry: a ne varieties correspond to nitely generated k-algebras without zero-divisors, and

Chapter 11. Differentials

, and similarly on U1. On the intersection D+(x0) ∩D+(x1), we have

x20d

(x1

x0

)= −x2

1d

(x0

x1

)This gives a non-vanishing section of ΩX(2)⊗ ΩX|A and furthermore an isomor-

phism ΩX|A ' OX(−2).

11.8.1 Smooth morphisms

We can also use the sheaves of differentials to define a notion of smoothness for

morphisms. In short, we say that a morphism f : X → S at a point x ∈ X if

smooth if ΩX|S is locally free of rank dimxX − dimf(x) S there. If this is not

the case, we say that f is ramified at x, and that x is a ramification point. f is

smooth if it is smooth at every point.

Example 11.41. Let A = k[x] and let B = k[x, y]/(x− y2) ' k[y] where k is a

field of characteristic not equal to 2. Let X = SpecB and let Y = SpecA. Let

f : X → Y be the morphism induced by the inclusion A → B (thus x 7→ y2).

Since B ' k[y] it follows that ΩX = Bdy, the free B-module generated by

dy. Similarly ΩY = Adx. The conormal sequence gives us

→ ΩY ⊗A B → ΩX → ΩX/Y → 0

|| || o|Bdx → Bdy → Bdy/B(2ydy)

The point is that ΩX/Y = (k[y]/(2y))dy is a torsion sheaf supported on the

ramification locus of the map f : X → Y . (The only ramification point is above

0.) Note that ΩX/Y is the quotient of ΩX by the submodule generated by the

image of dx in ΩX = Bdy. The image of dx is 2ydy.

193

Page 194: Geir Ellingsrud and John Christian Ottem October 15, 2018€¦ · between algebra and geometry: a ne varieties correspond to nitely generated k-algebras without zero-divisors, and

11.9. The conormal sheaf

11.9 The conormal sheaf

Let Y ⊆ X be a closed subscheme defined by an ideal sheaf I. Then I/I2 is

naturally a OY -module via I/I2 = I ⊗ OX/I = I ⊗ OY . We call I/I2 the

conormal sheaf of Y . Its dual, NY = HomOY (I/I2,OY ) is the normal sheaf of

Y in X.

Proposition 11.42. When X and Y are non-singular schemes over a field k,

the sheaves I/I2 and NY are locally free of rank r = codim(Y,X).

In this case, they fit into the exact sequences

0→ I/I2 → ΩX|k|Y → ΩY |k → 0

0→ TY → TX |Y → NY → 0

(This the conormal bundle sequence and its dual respectively. The first sequence

is exact on the left, because I/I2 is locally free).

Taking∧

of these sequences we get the following result, which is very useful

for computing canonical bundles of subvarieties:

Proposition 11.43 (Adjunction formula). If X and Y are non-singular, we

have

ωY = ωX ⊗r∧NY = ωX ⊗ detNY

In particular, if Y ⊆ X is a smooth divisor, we get

ωY = ωX ⊗ OX(Y )|Y

Example 11.44. Let X ⊆ P2 be a non-singular plane curve of degree d. Then

IX = OP2(−d), and so I/I2 = I ⊗ OP2/I = OY (−d). Here∧1 I/I2 = OY (−d)

so the previous proposition shows that

ωY ' OY (d− 3)

For d = 1, this gives our previous computation that ΩP1|k = OP1(−2).

Also for d = 2, Y ' P1, and ΩY = OP2(−1)|Y . This is consistent with the

previous example, because OP2(1)|Y ' OP1(2).

For d ≥ 3, the invertible sheaf ωY has many global sections. In particular, we

recover the fact that a non-singular plane curve of degree d ≥ 3 is not rational

(i.e., not isomorphic to P1). In fact, it will follow from the results of the next

chapter that Γ(Y, ωY ) has dimension exactly 12(d− 1)(d− 2). So for d ≥ 3, no

two smooth curves of different degrees can be isomorphic.

194

Page 195: Geir Ellingsrud and John Christian Ottem October 15, 2018€¦ · between algebra and geometry: a ne varieties correspond to nitely generated k-algebras without zero-divisors, and

Chapter 12

First steps in sheaf cohomology

We have seen several examples of the global sections functor Γ failing to be exact

when applied to a short exact sequence. On the other hand, we have a sequence

0 Γ(X,F ′) Γ(X,F) Γ(X,F ′′)

which is exact at each stage except on the right. In many situations in algebraic

geometry, knowing that Γ(X,F) → Γ(X,F ′′) is surjective of fundamental im-

portance. We will see several examples of this phenomenon later in this chapter.

Essentially, cohomology is a tool that allows us to continue this sequence, so that

the sequence can be continued as follows:

0 Γ(X,F ′) Γ(X,F) Γ(X,F ′′)

H1(X,F ′) H1(X,F) H1(X,F ′′)

H2(X,F ′) H2(X,F) H2(X,F ′′) −→ · · ·

In other words, the failure of surjectivity of the above is controlled by the group

H1(X,F ′), and the other groups in the sequence.

Cohomology groups can be defined in a completely general setting, for any

topological space X and a sheaf F on it. With that as input, we define the

cohomology groups Hk(X,F), which will capture the main geometric invariants of

F . These should also be functorial in F , which means that we want to construct

195

Page 196: Geir Ellingsrud and John Christian Ottem October 15, 2018€¦ · between algebra and geometry: a ne varieties correspond to nitely generated k-algebras without zero-divisors, and

12.1. Some homological algebra

functors

Hq(X,−) : AbShX → Ab

F 7→ Hq(X,F)

satisfying the following properties:

(I) H0(X,F) = Γ(X,F) = F(X),

(II) A morphism of sheaves φ : F → G induces maps H i(X,F) → H i(X,G)

which are functorial and take the identity to the identity (e.g., H i(X,−−)

is a functor).

(III) Every short exact sequence 0 → F ′ → F → F ′′ → 0 gives a long exact

sequence as above.

There are several ways to define these groups. The modern approach, and

the one summarized in Hartshorne Chapter III, takes the approach of using

derived functors. This is in most respects this is the ‘right way’ define the groups

in general, but going through the whole machinery of derived functors and

homological algebra would take us too far astray. We take a more down-to-earth

approach using Cech cohomology and the Godement resolution. The Godement

resolution has the advantage that it is completely canonical, and we can prove the

main theorems we need by hand. On the other hand, the Cech resolution, which

depends on the choice of a covering of X, but is better suited for computations.

Of course, the two notions of cohomology of course turn out to be the same, see

section XX.

12.1 Some homological algebra

A complex of abelian groups A• is a sequence of groups Ai together with maps

between them

· · · di−2

−−→ Ai−1 di−1

−−→ Aidi−→ Ai+1 di+1

−−→ · · ·

such that di+1 di = 0 for each i. A morphism of two complexes A•fn−→ B• is a

collection of maps fn : An → Bn so that the following diagram commutes:

· · · Ai−1 Ai Ai+1 · · ·

· · · Bi−1 Bi Bi+1 · · ·

di−1

fi−1

di

fi

di+1

fi+1

di−1 di di+1

In this way, we can talk about kernels, images, cokernels, exact sequences of

complexes, etc.

196

Page 197: Geir Ellingsrud and John Christian Ottem October 15, 2018€¦ · between algebra and geometry: a ne varieties correspond to nitely generated k-algebras without zero-divisors, and

Chapter 12. First steps in sheaf cohomology

Let ZnA• = ker di and BiA• = im di−1. From this condition on the di, we

have ZnA• ⊇ BnA•. We define the cohomology groups of A•, denoted HpA• as

HpA• = Zn(A•)/Bn(A•) = ker di/ im di−1

A resolution of a group G is a complex of the form

0→ G→ A0 → A1 → A2 → · · ·

where the maps are exact at each Ai, i.e., HpA• = 0 for each p ≥ 0.

Proposition 12.1. Suppose that 0→ F •f−→ G•

g−→ H• → 0 is an exact sequence

of complexes. Then there is a long exact sequence

· · · HnF • HnG• HnH•

Hn+1F • Hn+1G• Hn+1H• −→ · · ·Proof. Consider the commutative diagram

0 F n Gn Hn 0

0 F n+1 Gn+1 Hn+1 0

fn

dn

gn

dn dn

fn+1 gn+1

where the rows are exact. By the snake lemma, we obtain a sequence

0→ Zn(F •)fn−→ Zn(G•)

gn−→ Zn(H•)δ−→ F n+1/Bn(F •)

fn+1−−→ Gn+1/Bn(G•)gn+1−−→ Hn+1/Bn(H•)→ 0

Consider now the diagram

0 F n/Bn(F •) Gn/Bn(G•) Hn/Bn(H•) 0

0 Zn+1(F •) Zn+1(G•) Zn+1(H•)

fn

dn

gn

dn dn

fn+1 gn+1

where the rows are exact. Applying the snake lemma one more time, we get the

desired exact sequence.

A complex of sheaves F• is a sequence of sheaves with maps between them

· · · di−2

−−→ Fi−1di−1

−−→ Fidi−→ Fi+1

di+1

−−→ · · ·

such that di+1di = 0 for each i. Given such a complex, we define the cohomology

sheaves HpF• as ker di/ im di−1. A resolution of a sheaf F is a complex of the

197

Page 198: Geir Ellingsrud and John Christian Ottem October 15, 2018€¦ · between algebra and geometry: a ne varieties correspond to nitely generated k-algebras without zero-divisors, and

12.2. Cech cohomology

form

0→ F → C 0 → C 1 → C 2 → · · ·

where the maps are exact at each C i, i.e., HpC • = 0 for each p ≥ 0.

12.2 Cech cohomology

Let X be a topological space, and let F be a sheaf on it. Let U = Ui be an

open cover of X, indexed by an ordered set I. As we saw previously, the sheaf

axioms gives that the following sequence is exact:

0→ F(X)→∏i

F(Ui)→∏i,j

F(Ui ∩ Uj).

The Cech complex is essentially the continuation of this sequence; it is a complex

obtained by adjoining all the groups F(Ui1 ∩ · · · ∩ Uir) over all intersections

Ui1 ∩ · · · ∩ Uir .

Definition 12.2. For a sheaf F on X, define the Cech complex of F (with

respect to U) as

C•(U ,F) : C0(U ,F)d0−→ C1(U ,F)

d1−→ C2(U ,F)d2−→ · · ·

where

C0(U ,F) =∏i

F(Ui)

C1(U ,F) =∏i0<i1

F(Ui0 ∩ Ui1)

...

Cp(U ,F) =∏|I|=p+1

F(Ui0 ∩ · · · ∩ Uip).

and the coboundary map d : Cp(U ,F)→ Cp+1(U ,F) by

(dpσ)i0,...,ip+1 =

p+1∑j=0

(−1)jσi0,...ij ,...,ip+1|Ui0∩···∩Uip

where i0, . . . ij, . . . , ip+1 means i0, . . . , ip+1 with the index ij omitted.

Example 12.3. Let us look at the first few maps in the complex:

198

Page 199: Geir Ellingsrud and John Christian Ottem October 15, 2018€¦ · between algebra and geometry: a ne varieties correspond to nitely generated k-algebras without zero-divisors, and

Chapter 12. First steps in sheaf cohomology

In degree 0: d0 : C0(U ,F)→ C1(U ,F). If σ = (σi)i, then

(d0σ)ij = σj − σi

In degree 1: d1 : C1(U ,F)→ C2(U ,F). If σ = (σij)i, then

(d1σ)ijk = σjk − σik + σij

Notice that in the example that d1 d0 = 0 (all the σij cancel). This happens

also in higher degrees by a basic computation using the definition of di.

Lemma 12.4. dp+1 dp = 0.

In particular, the C•(U ,F) forms a complex of abelian groups. We say that

an element σ ∈ Cp is a cocycle if dpσ = 0, and a coboundary if σ = dp−1τ . By

the lemma, all coboundaries are cocycles. The cohomology groups of F are set

up to measure the difference between these two notions:

Definition 12.5. The p-th Cech cohomology of F with respect to U is defined

as

Hp(U ,F) = (ker dp)/(im dp−1)

One can show that a sheaf homomorphism F → G induces a mapping of

Cech cohomology groups, so we obtain functors F → Hp(U ,F).

Example 12.6. Again it is instructive to consider the group H0(U ,F). Here

the map d0 : C0(U ,F)→ C1(U ,F), which is simply the usual map∏F (Ui)→

∏F(Ui ∩ Uj)

which has kernel F (X) by the sheaf axioms. It follows that H0(U ,F) = F(X).

Example 12.7. X = S1, U = U, V the standard covering of S1 (intersecting

in two intervals) and let F = ZX (the constant sheaf).

Here we have

C0(U ,F) = ZU × ZV C1(U ,Z) = ZU∩V = Z× Z

199

Page 200: Geir Ellingsrud and John Christian Ottem October 15, 2018€¦ · between algebra and geometry: a ne varieties correspond to nitely generated k-algebras without zero-divisors, and

12.2. Cech cohomology

The map C0 → C1 is the map d0 : Z2 → Z2 given by d0(a, b) = (b − a, b − a).

Hence

H0(U ,ZX) = ker d = Z(1, 1) ' Z

and

H1(U ,ZX) = coker d = Z× Z/Z(1, 1) ' Z

Students familiar with algebraic topology, might recognize that this is gives the

same answer as singular cohomology.

Exercise 39. Let X = S1 and let U be the covering of X with three pairwise

intersecting open sets with empty triple intersection. Show that the Cech

-complex is of the form

Z3 d0−→ Z3 → 0

Compute the map d0 and use it to verify again that H i(U ,ZX) = Z for i = 0, 1

as above.

Exercise 40. Let X = S2 and let U be the covering of X with four pairwise

intersecting open sets with empty triple intersection. Show that the Cech

-complex takes the form

Z4 d0−→ Z6 d1−→ Z4 → 0

Compute the matrices d0, d1 and show that H i(U ,ZX) = Z for i = 0, 2 and

H i(U ,ZX) = 0 for i 6= 0, 2.

The following proposition shows that constant sheaves are not so interesting

for our purposes.

Proposition 12.8. Let X be an irreducible topological space (so that any open

set U ⊆ X is connected). Then for any covering U of X we have for a constant

sheaf AXH i(U , AX) = 0

for i > 0.

Proof.

At this point, there are good news and bad news. As seen in the examples

above, the groups Hp(U ,F) are easily computable, if one is given a nice cover of

X. Indeed, the maps in the Cech complex are completely explicit, and computing

their kernels and cokernels are usually just a computation in linear algebra.

On the other hand, the definition is a little bit unsatisfactory for various

reasons. First of all, the groups Hp(U ,F) depend on the cover U , whereas we

want something canonical that only depends on F . More importantly, it is not

clear that the definition above should capture enough of the desired information

200

Page 201: Geir Ellingsrud and John Christian Ottem October 15, 2018€¦ · between algebra and geometry: a ne varieties correspond to nitely generated k-algebras without zero-divisors, and

Chapter 12. First steps in sheaf cohomology

of F . For instance, U could consist of the single open set X, and so H i(U ,F) = 0

for i ≥ 1! Finally, it is not at all clear if these groups satisfy the requirements

mentioned in the introduction.

There is a fix for these problems, which involves passing to finer and finer

‘refinements’ of the covering. We say that a covering V is a refinement of U if for

every V ∈ V there is a U ∈ U so that V ⊆ U . This defines an ordering on the

coverings which we denote by V ≤ U . One can then define a group Hp(X,F) to

be the direct limit of all Hp(U ,F) as U runs through all possible open covers Uordered by ≤. The resulting groups are indeed canonical, and turn out to give

the right answer:

Definition 12.9. The groups Hp(X,F) are called the cohomology groups of F .

In symbols,

Hp(X,F) = lim−→UHp(U ,F)

The main preoperties of Cech cohomology are summarized in the following

theorem:

Theorem 12.10. Let X be a topological space and let F be a sheaf on X.

• The Cech cohomology groups are functors H i(X,−) : ShX → Groups.

• H0(X,F) = Γ(X,F) = F(X),

• Short exact sequences of sheaves induce long exact sequences of cohomology.

• (Leray’s theorem) If F is a sheaf and U is a covering such that H i(Ui1 ∩· · ·Uip .F) = 0 for all i > 0 and multiindexes i1 < · · · < ip, then

H i(X,F) = H i(U ,F).

We have proved the first two of these properties. The next two require more

work, but the arguments are mostly formal. We will nevertheless postpone the

proof until Appendix 13.

Note that the last statement shows that it suffices to compute the Cech

cohomology groups on a covering which is ‘sufficiently fine’ in the sense that the

higher groups H i(Ui1 ∩ · · ·Uip .F) = 0 vanish for i > 0.

Remark 12.11. Let us at least say a few words about the first boundary map δ

in the long-exact sequence above. Let 0→ F ′ → F → F ′′ → 0 be a short exact

sequence of sheaves. We would like to construct the boundary map δ which is a

map

δ : H0(X,F ′′)→ H1(X,F ′)

201

Page 202: Geir Ellingsrud and John Christian Ottem October 15, 2018€¦ · between algebra and geometry: a ne varieties correspond to nitely generated k-algebras without zero-divisors, and

12.3. Cohomology of sheaves on affine schemes

If each restriction map

Γ(Ui,F)→ Γ(Ui,F ′) (12.2.1)

is surjective, then we can explicitly describe δ as follows: Represent a class

σ ∈ H0(X,F ′′) by a collection of elements fi ∈ Γ(Ui,F ′′) so that fi = fj on

Ui∩Uj . If the above restriction map is surjective, we can choose lifts gi ∈ Γ(Ui,F)

and define f ′ij = gi − gj. The f ′ij then form a 2-cocycle, since f ′ij − f ′ik + f ′jk = 0.

So we get a well defined class δ(σ) ∈ H1(X,F ′) (of course one needs to check

that this is independent of the choice of lift, but this is straightforward). Hence

we get a sequence

0→ H0(F ′)→ H0(F)δ−→ H0(F)→ H1(F)→ H1(F ′′)

which, by another diagram chase is exact.

In general, it can certainly happen that the restriction map (12.2.1) is not

surjective – one can for instance take the open covering of X with just one open

set X. This explains why the Cech cohomology groups H i(U ,F) do not give long

exact sequences in general. However, by passing to smaller refinements V ≤ U ,

we can arrange that any section lifts and we can use the above to construct δ.

12.3 Cohomology of sheaves on affine schemes

Theorem 12.12. Let X = SpecA and let F be a quasi-coherent sheaf on X.

Then

Hp(X,F) = 0 for all p > 0.

Proof. Recall that we defined the groups H i(X,F) by taking the direct limit of

H i(U ,F) finer and finer coverings U of X. Since the distinguished opens subsets

form a basis for the topology on X. It suffices to prove that

Hp(U ,F) = 0 for all p > 0.

for a covering U = D(gi) where the gi are finitely many elements of A generating

the ring. (We can choose finitely many gi, by the quasi-compactness of X).

As F is quasi-coherent, F = M for some A-module M , and M = Γ(X,F).

With this setup, the fact that Cech cohomology vanishes of a quasi-coherent

module on an affine scheme corresponds to the observation that for some com-

mutative ring A, some finite sequence of elements (gi)i∈I of A generating A as

an ideal, and some A-module M , the following sequence is exact:

0→M →∏i∈I

Mgi →∏i,j∈I

Mgigj →∏i,j,k∈I

Mgigjgk → . . .

202

Page 203: Geir Ellingsrud and John Christian Ottem October 15, 2018€¦ · between algebra and geometry: a ne varieties correspond to nitely generated k-algebras without zero-divisors, and

Chapter 12. First steps in sheaf cohomology

Here, the boundary maps are given as alternating sums of localization maps. For

example,

d :∏i,j∈I

Mgigj →∏i,j,k∈I

Mgigjgk

maps (σij)i,j ∈Mgigj to (σjk − σik + σij)i,j,k.

Notice that the beginning of the exact sequence

0→M →∏i

Mgi →∏i,j

Mgigj

appeared in the construction of the quasi-coherent module M . The proof for the

exactness of this sequence is similar to the general case.

To prove that the cohomology groups vanish, we must to each cocycle σ (such

that dσ = 0) find an element τ such making σ = dτ a boundary. The proof is a

direct calculation; one constructs an element τ by hand.

To see how this can be done, let us for simplicity consider the case p = 1 first.

Let σ ∈∏

i,jMgigj be in the kernel of d. We may write

σij =mij

(gigj)rwhere mij ∈M

for some r (since the index set I is finite, we may choose this independent of

i, j). The relation dσ = 0 gives the relation

mjk

(gjgk)r− mik

(gigk)r+

mij

(gigj)r= 0

or in other words, we have, in Mgjgk ,

gr+li mjk

(gjgk)r=glig

rjmik

(gjgk)r− glig

rkmij

(gjgk)r(12.3.1)

for some l ≥ 0. As the sets D(gi) = D(gr+li ) cover X, we have a relation

1 =∑i∈I

higr+li

where hi ∈ A. Now, define τ ∈∏Mgj by

τj =∑i∈I

higlimij

grj

203

Page 204: Geir Ellingsrud and John Christian Ottem October 15, 2018€¦ · between algebra and geometry: a ne varieties correspond to nitely generated k-algebras without zero-divisors, and

12.3. Cohomology of sheaves on affine schemes

In∏Mgjgk , we may write it as

τj =∑i∈I

higlig

rkmij

(gjgk)r

We want to show that dτ = σ. This is a basic computation, using the relation

(12.3.1) above:

(dτ)jk = τk − τj

=∑i∈I

higlig

rjmik

(gjgk)r−∑i∈I

higlig

rkmij

(gjgk)r

=∑i∈I

hi

(glig

rjmik

(gjgk)r− glig

rkmij

(gjgk)r

)

=∑i∈I

higr+li mjk

(gjgk)r=

mjk

(gjgk)r

∑i∈I

higr+li

=mjk

(gjgk)r= σjk

As desired. Hence H1(U ,F) = 0.

In the general case, let σ ∈∏

i0,...,ipMgi0 ···gip be in the kernel of d. We may

write

σi0,...,ip =mi0,...,ip

(gi0 · · · gip)rwhere mi0,...,ip ∈M

for some r (since the index set I is finite, we may choose this independent of

i, j). The relation dσ = 0 gives the relation in Mgi0 ···gip :

gr+li mi0,...,ip

(gi0 · · · gip)r=

p∑k=0

(−1)kglig

rikmii0,...,ik...ip

(gi0 · · · gip)r(12.3.2)

As the sets D(gi) = D(gr+li ) cover X, we have a relation

1 =∑i∈I

higr+li

where hi ∈ A. Now, define τ ∈∏Mgi0 ···gip−1

by

τi0,...,ip−1 =∑i∈I

higli

mii0,...ip

(gi0 · · · gip−1)r

204

Page 205: Geir Ellingsrud and John Christian Ottem October 15, 2018€¦ · between algebra and geometry: a ne varieties correspond to nitely generated k-algebras without zero-divisors, and

Chapter 12. First steps in sheaf cohomology

Localizing to∏Mgi0 ,...,gip

, we may write this as

τi0,...,ip =∑i∈I

higligrik

mii0,...ip

(gi0 · · · gip)r

We want to show that dτ = σ. As before, we can check this using the relation

(12.3.2) above:

(dτ)i0···ip =

p∑k=0

(−1)kτi0...ik...ip

=∑i∈I

higr+li σi0...ip = σi0...ip

This completes the proof.

12.3.1 Cech cohomology and affine coverings

As a Corollary of the previous theorem, we see that affine coverings of schemes

satisfy the conditions of Leray’s theorem (see Theorem 12.10). This implies

Corollary 12.13. Let X be a noetherian scheme and let U = Ui be an affine

covering such that all intersections Ui0 ∩ · · · ∩ Uis are affine. Then

H i(X,F) = H i(U ,F)

In particular, the theorem applies to any affine covering on a noetherian separated

scheme.

12.4 Chomology and dimension

Lemma 12.14. Let X be a topological space and let Z ⊆ X be a closed subset.

Then for any abelian sheaf F on Z we have Hp(Z,F) = Hp(X, i∗F).

Theorem 12.15. Let X be a quasi-projective scheme of dimension n. Then X

admits an open cover U consisting of at most n+ 1 affine open subsets.

This implies that

H i(X,F) = 0 for i > n

for any quasi-coherent sheaf F on X.

Proof. Let X be a quasi-projective scheme, i.e., X = Y −W , where Y,W ⊆ PnAare (closed) projective schemes. Consider the irreducible decomposition Y = ∪iYiand observe that IW 6⊆ ∪IYi where IT ⊆ A[xo, . . . , xN ] denotes the ideal of the set

205

Page 206: Geir Ellingsrud and John Christian Ottem October 15, 2018€¦ · between algebra and geometry: a ne varieties correspond to nitely generated k-algebras without zero-divisors, and

12.5. Cohomology of sheaves on projective space

T ⊆ PN . Pick a homogenous polynomial f of degree d such that f ∈ IW −(∪iIYi).Let H = Z(f). Then Pn −H is affine and hence so is Y −H.

By construction Y −H ⊆ Y −W = X and H 6⊇ Yi for any i by the choice of

f . Therefore dim(Yi ∩H) < dim Yi so we may use induction on dimX. Notice

that the affine subset is obtained as an affine subset of the ambient projective

space intersected with our scheme, so the affine schemes obtained subsequently

are restrictions of affine subschemes of the original X. This shows the first claim.

For the second, note that for ≤ n + 1 affines, there are at most n terms

Ci(X,F) in the Cech complex. From this it follows that H i(U ,F) = 0 for any

F and i > n.

In fact, this theorem holds in much greater generality:

Theorem 12.16. Let X be a topological space of dimension n, and let F be a

sheaf on X. Then

Hp(X,F) = 0

for all p > n.

12.5 Cohomology of sheaves on projective space

On projective space PnA, we can use the Cech complex associated with the

standard covering to compute the cohomology of any invertible sheaf on PnA.

Theorem 12.17. Let X = PnA = ProjR where R = A[x0, . . . , xn] where A is a

noetherian ring.

(i)

H0(X,OX(m)) =

Rm for m ≥ 0

0 otherwise

(ii)

Hn(X,OX(m)) =

A for m = −n− 1

0 for m > −n− 1

(iii) For m ≥ 0, there is a perfect pairing1 of A-modules

H0(X,O(m))×Hn(X,OX(−m− n− 1))→ Hn(X,OX(−n− 1)) ' A

1Recall that a bilinear map M × N → A is a perfect pairing if the induced map M 7→HomA(N,A) is an isomorphism

206

Page 207: Geir Ellingsrud and John Christian Ottem October 15, 2018€¦ · between algebra and geometry: a ne varieties correspond to nitely generated k-algebras without zero-divisors, and

Chapter 12. First steps in sheaf cohomology

(iv) For 0 < i < n and all n ∈ Z, we have

H i(X,OX(m)) = 0

Proof. Consider the OX-module F =⊕

m∈Z O(m). We would like to show for

instance that H1(X,F) = 0 for i 6= 0, n; this is equivalent to H i(X,OX(m)) = 0

for all m, but F has the advantage that it is a graded OX-algebra. Consider

the Cech complex associated with the standard covering U = Ui where Ui =

D+(xi) = SpecR(xi). This is simply∏i

Rxid0−→∏i,j

Rxixjd1−→ . . .

dn−1

−−−→ Rx0···xn

We have a graded isomorphism of R-modules:

H0(X,F) = ker d0

= (ri)i∈I |ri ∈ Rxi , ri = rj ∈ Rxixj' R.

This isomorphism preserves the grading, so we get (i).

For (ii): Note that Rx0···xn is a free graded A-module spanned by monomials

of the form

xa00 · · ·xannfor multidegrees (a0, . . . , an) ∈ Zn+1. The image of dn−1 is spanned by such

monomials where at least one ai is non-negative. Hence

Hn(X,F) = coker dn−1

= A xa00 · · ·xann |ai < 0∀i ⊆ Rx0...xn

Hence

Hn(X,OX(m)) = Hn(X,F)m

= Axa00 · · ·xann |ai < 0∀i,

∑ai = m

⊆ Rx0...xn

In degree −n− 1 there is only one such monomial, namely x−10 · · ·x−1

n .

(iii): If we identify Hn(X,OX(−m− n− 1)) with

Axa00 · · ·xann : ai < 0∀i,

∑ai = m

and H0(X,O(m)) = Rm, we can define the pairing via multiplication of Laurent

207

Page 208: Geir Ellingsrud and John Christian Ottem October 15, 2018€¦ · between algebra and geometry: a ne varieties correspond to nitely generated k-algebras without zero-divisors, and

12.5. Cohomology of sheaves on projective space

polynomials:

H0(X,O(m))×Hn(X,OX(−m− n− 1))→ Rx0···xn

(xm00 · · ·xmnn )× (xa00 · · ·xann ) 7→ xa0+m0

0 · · ·xan+mnn

Here the exponents satisfy mi ≥ 0, ai < 0∑ai = −m− n− 1,

∑mi = m. This

gives a map

H0(X,O(m))×Hn(X,OX(−m−n−1))→ Hn(X,OX(−n−1)) = Ax−10 · · ·x−1

n

sending (xm00 · · ·xmnn ) × (xa00 · · ·xann ) to zero if mi + ai ≥ 0 for some i. This

pairing is perfect: The dual of a monomial (xm00 · · ·xmnn ) is represented by

(x−m0−10 · · · x−mn−1

n ).

(iv) This point is more involved, and we proceed by induction on n. For

n = 1, there is nothing to prove. For n > 1, let H = V (xn) ' Pn−1 be the

hyperplane determined by xn. We have an exact sequence

0→ R(−1)·xn−−→ R→ R/(xn)→ 0 (12.5.1)

Applying ∼, we find

0→ OX(−1)→ OX → i∗OH → 0

where i : H → X is the inclusion. If we take the direct sum of all the twists of

this sequence, we get

0→ F(−1)→ F → i∗FH → 0

where FH =⊕

m∈Z OH(m). By induction on n, we have for 0 < i < n− 1 and

all m ∈ Z: H i(X, i∗OH(m)) = H i(H,OH(m)) = 0. So taking the long exact

seqence of cohomology, we get isomorphisms

H i(X,F(−1))·xn−−→ H i(X,F)

for 1 < i < n − 1. We claim that we have isomorphisms also for i = 1 and

i = n− 1. For i = 1, this follows because the sequence

0→ H0(X,F(−1))→ H0(X,F)→ H0(X, i∗FH)→ 0

is exact (this is the same sequence as (12.5.1)).

For i = n− 1 we need to show that

0→ Hn−1(X, i∗FH)δ−→ Hn(X,F(−1))

·xn−−→ Hn(X,F)

208

Page 209: Geir Ellingsrud and John Christian Ottem October 15, 2018€¦ · between algebra and geometry: a ne varieties correspond to nitely generated k-algebras without zero-divisors, and

Chapter 12. First steps in sheaf cohomology

is exact. The kernel of ·xn, is generated by monomials xa00 · · ·xann with ai < 0 for

all i. So it suffices to show that the connecting map δ is just multiplication by

x−1n . Define R′ = R/xn. Writing the arrows in the Cech complex, vertically we

get the diagram

0 //

∏iR(−1)x0···xi···xn

·xn //

∏iRx0···xi···xn

//

R′x0···xn−1//

0

0 // Rx0···xn(−1)·xn // Rx0···xn

// 0

If xa00 · · ·xan−1

n−1 is a monomial in Hn−1(H,FH) with ai < 0 for all 1 ≤ i ≤ n− 1,

then it comes from an (n+1)-tuple in∏

iRx0···xi···xn which maps to ±xa00 · · ·xan−1

n−1

in Rx0···xn , which is in turn mapped onto by the monomial xa00 · · ·xan−1

n−1 x−1n in

Rx0···xn(−1). So δ(xa00 · · ·xan−1

n−1 ) is represented by the monomial xa00 · · ·xan−1

n−1 x−1n

in Hn(X,F(−1)).

Now we claim that we have an isomorphism H∗(X,F)xn = H∗(Un,F|Un).

Indeed, the Cech complex of F|Un with respect to the covering Ui∩Un is just the

localization of C•(X,F) at xn. Localization is exact, so it preserves cohomology,

which gives the claim.

We know that H i(X,F)xn = H i(Un,F|Ur) = 0 for all i > 0, since Un is affine.

Hence for l 0, xlnHi(X,F) = 0 as an A-module. However, we have shown

that ·xn gives an isomorphism of H i(X,F) for 0 < i < n. This implies that

H i(X,F) = 0.

In the above proof, we used the following lemma:

Lemma 12.18. Let X = PnA over a noetherian ring A and let F be a quasi-

coherent sheaf on X. Then H i(X,F) = 0 for all i > n.

Proof. PnA can be covered with n+1 open affine schemes, all of whose intersections

are affine. Hence the Cech complex associated with F has length n+ 1, and so

there is no cohomology in degrees higher than n.

12.6 Extended example: Plane curves

Let X = V (f) denote an integral subvariety of P2, defined by an irreducible

homogeneous polynomial f of degree d.

LetX ⊆ P2 be a plane curve, defined by a homogeneous polynomial f(x0,1 , x2)

of degree d. Let us compute H i(X,OX).. We have the ideal sheaf sequence

0→ IX → OP2 → i∗OX → 0

209

Page 210: Geir Ellingsrud and John Christian Ottem October 15, 2018€¦ · between algebra and geometry: a ne varieties correspond to nitely generated k-algebras without zero-divisors, and

12.7. Extended example: The twisted cubic in P3

where the ideal sheaf IX is the kernel of the restriction OP2 → i∗OX . By Section

10.7.1, OP2(−X) ' OP2(−d). More explicitly, on Spec k[x1x0, x2x0

], we see IX is

a k[x1x0, x2x0

]-module generated by f(x1x0, x2x0, 1), and OP2(−d) is a k[x1

x0, x2x0

]-module

generated by x−d0 , and this isomorphism sends generator to generator. The above

exact sequence can thus be rewritten as

0 OP2(−d) OP2 OX 0.·f

From the short exact sequence, we get the long exact sequence as follows:

0 H0(P2,O(−d)) H0(P2,OP2) H0(X,OX)

H1(P2,O(−d)) H1(P2,OP2) H1(X,OX)

H2(P2,O(−d)) H2(P2,OP2) 0.

Using the results on cohomology of line bundles on P2, we deduce thatH0(X,OX) 'k and

H1(X,OX) ' k(d−12 ).

12.7 Extended example: The twisted cubic in

P3

Let k be a field and consider P3 = ProjR where R = k[x0, x1, x2, x3]. We will

consider the twisted cubic curve C = V (I) where I ⊆ R is the ideal generated

by the 2× 2-minors of the matrix

M =

(x0 x1 x2

x1 x2 x3

)i.e., I = (q0, q1, q2) = (x2

1 − x0x2, x0x3 − x1x2,−x22 + x1x3). Let us by hand

compute the group H1(X,OX). Of course we know what the answer should be,

since X ' P1, and H1(P1,OP1) = 0. Indeed, P1 = Proj k[s, t] is isomorphic to

the Veronese embedding Proj k[s, t](3), where k[s, t](3) = k[s3, s2t, st2, t3], and the

latter ring is isomorphic to S = R/I.

Now, to computeH1(X,OX) onX, it is convenient to relate it to a cohomology

group on P3. We have H1(X,OX) = H1(P3, i∗OX) where i : X → P3 is the

inclusion. This fits into the ideal sheaf sequence

0→ I → OP3 → i∗OX → 0.

210

Page 211: Geir Ellingsrud and John Christian Ottem October 15, 2018€¦ · between algebra and geometry: a ne varieties correspond to nitely generated k-algebras without zero-divisors, and

Chapter 12. First steps in sheaf cohomology

where I is the ideal sheaf of X in P3. Applying the long exact sequence in

cohomology we get

0 I OP3 i∗OX 0.

From the short exact sequence, we get the long exact sequence as follows:

· · · H1(P3, I) H1(P3,OP3) H1(P3, i∗OX)

H2(P3, I) H2(P3,OP3) · · ·

By our description of sheaf cohomology on P3, H1(P3,OP3) = H2(P3,OP3) = 0,

we find that H1(X,OX) = H2(P3, I). We can compute the latter cohomology

group as follows. Note that I = I. Consider the map R3 → I → 0 sending

ei 7→ qi. Let us consider the kernel of this map, that is, relations of the form

a0q0 + a1q1 + a2q2 = 0 for ai ∈ R. There are two obvious relations of this form,

i.e., the ones we get from expanding the determinants of the two matricesx0 x1 x2

x0 x1 x2

x1 x2 x3

x0 x1 x2

x1 x2 x3

x1 x2 x3

(So first matrix gives x0q2 − x1q1 + x2q2 = 0 for instance). These give a map

R2 ·M−→ R3, where M is the matrix above. This map is injective, and there is an

exact sequence

0→ R2 ·M−→ R3 → I → 0

If we want to be completely precise, and consider these as graded modules, we

must shift the degrees according to the degrees of the maps above

0→ R(−3)2 ·M−→ R(−2)3 → I → 0

Applying ∼, we get an exact sequence of sheaves

0→ OP3(−3)2 → OP3(−2)3 → I → 0

Now, this we can use to compute H2(X, I); taking the long exact sequence we

get

· · · H2(P3,O(−3)2) H2(P3,OP3(−2)3) H2(X, I)

H3(P3,O(−3)2) H3(P3,OP3(−2)3) H3(X, I)

211

Page 212: Geir Ellingsrud and John Christian Ottem October 15, 2018€¦ · between algebra and geometry: a ne varieties correspond to nitely generated k-algebras without zero-divisors, and

12.7. Extended example: The twisted cubic in P3

Here H2(P3,O(−2)) = 0 and H3(P3,O(−3)) = 0 (!) by our previsous computa-

tions. Hence by exactness, we find H2(X, I) = 0, as expected.

Exercise 41. Using the sequences above, show that

• H0(P3, I(2)) = k3 (find a basis!)

• H1(P3, I(m)) = 0 for all m ∈ Z.

• H2(P3, I(−1)) = k.

212

Page 213: Geir Ellingsrud and John Christian Ottem October 15, 2018€¦ · between algebra and geometry: a ne varieties correspond to nitely generated k-algebras without zero-divisors, and

Chapter 13

More on sheaf cohomology

13.1 Flasque sheaves

A sheaf F on a topological space X is said to be flasque if the restriction maps

F(X)→ F(U)

for every open subset U ⊆ X.

Example 13.1. If X is irreducible, then any constant sheaf is flasque.

Example 13.2. The skyskraper sheaves A(x) defined in Chapter 1 are flasque.

Example 13.3. OP1k

is not flasque, since Γ(P1,OP1k) = k, whereas Γ(U,OP1

k) is

typically much larger (e.g., for U = A1k ⊆ P1

k it is an infinite dimensional k-vector

space).

Recall that the global section functor Γ is left exact. The following lemma

says that it is also exact on the right, given that the left-most term in the

sequence is flasque:

Lemma 13.4. Given an exact sequence

0 // F // G //H // 0

of sheaves F , G where F is flasque. Then the corresponding sequence of global

sections

0 // F(U) // G(U) //H(U) // 0

is exact for every open set U ⊆ X.

213

Page 214: Geir Ellingsrud and John Christian Ottem October 15, 2018€¦ · between algebra and geometry: a ne varieties correspond to nitely generated k-algebras without zero-divisors, and

13.1. Flasque sheaves

Proof. By restricting F ,G,H to U , it suffices to prove the statement for U = X.

We only need to check that the sequence is exact on the right. Let s ∈ H(X)

and let V be the family of open subsets U ⊆ X such that s|U is represented by

an element of G(X). We need to show that X ∈ V .

Since V contains a covering of X and X is quasi-compact, it suffices to show

that V is closed under finite unions. So let U1, U2 ∈ V and let ti ∈ Γ(Ui,F)

represent s on Ui, i = 1, 2. Note that by flasque property of F , t1 − t2 on

U1 ∩ U2 lifts to a global section u of F . Moreover, by adding u|U2 to t2, we may

assume that t1, t2 agree on U1∩U2, so by gluing we get a section t ∈ G(X) which

represents s, and hence U1 ∪ U2 ∈ V .

Lemma 13.5. Suppose we are given an exact sequence of sheaves

0 // F // G //H // 0.

If F and G are flasque, then so is H.

Proof. Let U ⊆ X be a subset of X. Then any section h ∈ H(U) is represented

by a section g ∈ G(U) by the previous lemma. Since G is flasque, this can be

extended to an element g of G(X). Then g maps to an element h ∈ H(X) which

has the property that h|U = h.

Lemma 13.6. Suppose we are given an exact sequence of flasque sheaves

0 // E → F0 → F1 → F2 → · · · , (13.1.1)

then for each open set U ⊆ X, the sequence

0 // E (U)→ F0(U)→ F1(U)→ F2(U)→ · · · , (13.1.2)

is exact.

Proof. We chop (13.1.2) into short exact sequences as follows:

0→ E → F0 → Q0 → 0

0→ Q0 → F1 → Q1 → 0

· · ·

By induction, it follows that the sheaves Qi are flasque, being quotients of flasque

sheaves. Applying Γ we then get exact sequences

0→ E (U)→ F0(U)→ Q0(U)→ 0

0→ Q0(U)→ F1(U)→ Q1(U)→ 0

· · ·

214

Page 215: Geir Ellingsrud and John Christian Ottem October 15, 2018€¦ · between algebra and geometry: a ne varieties correspond to nitely generated k-algebras without zero-divisors, and

Chapter 13. More on sheaf cohomology

If we connect these together, we obtain the exact sequence (13.1.2).

13.2 The Godement resolution

We will define the cohomology groups Hp(X,F) using a special resolution of Fknown as the Godement resolution. To explain how this works, we again recall

the canonical map κ : F → Π(F), which over an open set U ⊆ X sends a section

s ∈ F(U) to the element (sx)x∈U ∈ Π(F)(U) =∏

x∈U Fx. Defining C 0F = Π(F)

and Z1F as the cokernel of κ, we get a canonical exact sequence

0→ F → C 0F → Z1F → 0

Taking global sections, we get

0→ F(X)→ C 0F(X)→ Z1F(X).

As we have seen, we cannot expect to be exact on the right. In fact, this failure

of exactness is how we will define cohomology groups: We define H1(X,F) as

the cokernel of the right most map C 0F(X)→ Z1F(X), i.e.,

H1(X,F) = Z1(X)/B0(X).

where B = im(C 0F(X)→ Z1(X)).

We note that the assignments F 7→ C 0F , F 7→ Z1F and F 7→ H1(X,F) are

indeed functors. Indeed, given a map F → G of sheaves, we get canonical maps

between the two sequences

0 F(X) C 0F(X) Z1F(X) H1(X,F)→ 0

0 G(X) C 0G(X) Z1G(X) H1(X,G)→ 0

and from this we see that we have a canonical map H1(X,F)→ H1(X,G).

What about the higher cohomology groups? Well, we can continue, replacing

F with the new sheaf Z1. We again have a canonical map κ : Z1 → Π(Z1), and

we can define C 1F = C 0Z1 and Z2F = coker(Z1F → C 1F). We then have an

exact sequence

0→ Z1F → C 1F → Z2F → 0

and we can define H2(X,F) to be the cokernel of the map on global sections

C 1F(X)→ Z2F(X).

215

Page 216: Geir Ellingsrud and John Christian Ottem October 15, 2018€¦ · between algebra and geometry: a ne varieties correspond to nitely generated k-algebras without zero-divisors, and

13.2. The Godement resolution

So to define cohomology groups in general, we define inductively the sheaves C i

C 0F = Π(F)

Z1F = coker(κ : F → C 0F)

C nF = C 0Zn for n ≥ 1

Zn+1F = Z1Zn = coker(κ : Zn → C n) for n ≥ 1

The way to visualise these is as follows:

0→ F C 0F C 1F C 2F · · ·

Z1F Z2F

κ d0 d1 d2

κ κ

(13.2.1)

Here the map dq : C q → C q+1 is defined by the composition

C qF → Zq+1F → C 0Zq+1 = C q+1F

The second map is injective, which means that the complex in (13.2.1) is exact.

It follows that the complex 0 → F → C 0 → C 1 → C 2 → · · · is actually a

resolution, and we call it the Godement resolution of F .

13.2.1 Functoriality

Given a morphism of sheaves φ : F → G, we have a canonical morphism

Π(F) → Π(G) which is compatible with the two maps κF : F → Π(F) and

κG : G → Π(G). Therefore, we get a morphism of quotient sheaves

C 0F/F → C 0G/G

or in other words, a morphism Z1F → Z1G. Taking again Godement sheaves,

we get

C 0Z1F → C 0Z1G

or, a morphism C 1F → C 1G. Thus, inductively, we construct canonical sheaf

maps

C kφ : C kF → C kG

These C k() define the Godement functors; they are functors from ShX to ShX .

The Godement resolution is also functorial, in the sense that we have a canonical

morphism of complexes C •F → C •G:

216

Page 217: Geir Ellingsrud and John Christian Ottem October 15, 2018€¦ · between algebra and geometry: a ne varieties correspond to nitely generated k-algebras without zero-divisors, and

Chapter 13. More on sheaf cohomology

0 F C 0F C 1F C 2F · · ·

0 G C 0G C 1G C 2G · · ·

13.2.2 Sheaf cohomology

If we take global sections in the Godement resolution C •F , we get a sequence

0→ F(X) C 0F(X) C 1F(X) C 2(X) · · ·d0(X) d1(X) d2(X)

(13.2.2)

This may or may not be exact, but since di+1(X) di(X) = 0 we still have a

complex of abelian groups, and we can talk about its cohomology.

Definition 13.7. The p-th cohomology group Hp(X,F) is defined to be the

p-th cohomology group of the complex of abelian groups (13.2.2), i.e.,

Hp(X,F) = (ker dp(X))/(im dp−1(X))

Equivalently,

Hp(X,F) = Zp(X)/Bp(X)

where Bp(X) = (im(C p−1(X)→ Zp(X))).

So what is H0(X,F)? By definition this group is the kernel of the map d0,

which is defined as the composition

C 0F(X)→ Z1F(X) → C 1F(X)

Hence H0(X,F) = ker(C 0F(X)→ Z1F(X)). On the other hand, we have the

sequence 0→ F → C 0 → C 1; taking global sections, we get the exact sequence

0→ F(X)→ C 0(X)d0(X)−−−→ Z1F(X)

from which we see that H0(X,F) = ker d0(X) = Γ(X,F).

By the recursive definition of the sheaves C pF and Zp, we only need to know

how to define H1 of a sheaf in order to see the higher cohomology groups. Indeed,

from the complex (13.2.2), it is clear that the C q also give a resolution of Z1, so

that

H i(X,F) =

Γ(X,F ) if i = 0

coker(C 0(X)→ Z1(X)) if i = 1

H1(X,Z i−1) if i ≥ 2

The geometric meaning of the elements of these higher cohomology groups

217

Page 218: Geir Ellingsrud and John Christian Ottem October 15, 2018€¦ · between algebra and geometry: a ne varieties correspond to nitely generated k-algebras without zero-divisors, and

13.2. The Godement resolution

Hp(X,F) is on the other hand far from obvious, and unfortunately this is a

general tendency when defining cohomology groups. What is important is what

sort of properties they have: at the moment these groups are completely canonical

in their definition, and again since κ and Π(F) are functorial, we find that the

assignments F 7→ H i(X,F) define functors from sheaves to groups. This takes

care of the two first properties of sheaf cohomology.

We now turn to prove the third property, that is, that we have a long exact

sequence. First we prove that the assignments F 7→ C iF and F 7→ Z iF are in

fact exact functors.

Lemma 13.8. Let 0→ F1 → F2 → F3 → 0 be a short exact sequence of sheaves.

Then we have exact sequences

0→ C iF1 → C iF2 → C iF3 → 0

and

0→ Z iF1 → Z iF2 → Z iF3 → 0

Proof. By definition we have C 0i (U) =

∏x∈U (Fi)x for an open set U ⊆ X. Since

the above sequence is exact on stalks, we get also that the sequence

0→ C 0F1 → C 0F2 → C 0F3 → 0

is also exact. (In fact this observation is one of main reasons why we use the

Godement sheaves in the definition of cohomology in the first place.) This row

fits into the following diagram:

0 0 0

0 F1 F2 F3 0

0 C 0F1 C 0F2 C 0F3 0

0 Z1F1 Z1F2 Z1F3 0

0 0 0

(13.2.3)

Since the two top rows are exact, the snake lemma implies that also the lower

row is exact. Now using the fact that the composition of exact functors is still

exact, and the fact that C iF = C 0Z i and Z i+1F = Z1Z iF , we conclude by

induction on i.

218

Page 219: Geir Ellingsrud and John Christian Ottem October 15, 2018€¦ · between algebra and geometry: a ne varieties correspond to nitely generated k-algebras without zero-divisors, and

Chapter 13. More on sheaf cohomology

Lemma 13.9. The sheaves C iF and Z iF are flasque for every i ≥ 0.

Proof. Note first that Π(F) is flasque: For any open set U ⊆ X it is clear that∏x∈X

Fx →∏x∈U

Fx

is surjective. Hence C 0 is flasque. This implies that Z1 is flasque as well, being

a quotient of C 0 by a subsheaf. Now we can conclude by induction, using

C iF = C 0Z i and Z i+1F = Z1Z iF

Corollary 13.10. If F is flasque, then for every U ⊆ X we have

H i(U,F) = 0

In particular, F is acyclic.

Proof. We take the Godement resolution 0→ F → C 0F → C 1F →. Since Fand each C iF are flasque, we get

0→ F(U)→ C 0(U)→ C 1(U)→ C 2(U)→ · · ·

Since this complex is exact, we obtain that H i(X,F) = 0 by the definition of

cohomology.

We are now ready to prove property (iii), namely the long exact sequence of

cohomology.

Theorem 13.11. Let 0 → F1 → F2 → F3 → 0 be a short exact sequence of

sheaves. Then there is a long exact sequence

0→ H0(X,F1)→ H0(X,F2)→ H0(X,F3)δ−→ H1(X,F1)→ H1(X,F2)→ H1(X,F3)→ · · ·

Proof. First of all, by the previous lemma, there are exact sequences

C • : 0→ C iF1 → C iF2 → C iF3 → 0

Since each C iF is flasque, we get also that

0→ C iF1(X)→ C iF2(X)→ C iF3(X)→ 0

is exact. Hence we get an exact sequence of complexes of abelian groups

0→ C •F1(X)→ C •F2

(X)→ C •F3(X)→ 0

Then, since Hp(X,Fi) is defined as the cohomology of this complex, we conclude

by Proposition 12.1 that we have the above long exact sequence.

219

Page 220: Geir Ellingsrud and John Christian Ottem October 15, 2018€¦ · between algebra and geometry: a ne varieties correspond to nitely generated k-algebras without zero-divisors, and

13.2. The Godement resolution

Exercise 42. Let X be a topological space and let i : Z → X be the inclusion

of a subset. Show that for a sheaf F on Z,

H i(X, i∗F) = H i(Z,F) (13.2.4)

for all i.

13.2.3 Acyclic sheaves

Definition 13.12. A sheaf F is called acyclic if H i(X,F) = 0 for all i > 0.

Given a resolution

0→ F → C 0 → C 1 → C 2 → · · ·

of F (which by definition means that the sequence is exact), the resulting

sequence

C •(X) : C 0(X)→ C 1(X)→ C 2(X)→ · · · ,

may fail to be exact, but is still a complex of abelian groups, so it makes sense

to ask about its cohomology.

Lemma 13.13. If the sheaves C i are acyclic, then there is a natural isomorphism

H i(X,F) ' H i(C •(X))

Proof. Define K−1 = F , and Ki = ker(C i+1 → C i+2) for i ≥ 0. By exactness of

C •, we have for each i ≥ 0 and exact sequence

0→ Ki−1 → C i → Ki → 0

Taking the long exact sequence, gives

0→ H0(Ki−1)→ H0(C i)→ H0(Ki)→ H1(Ki−1)→ H1(C i) = 0 (13.2.5)

where the right-most group it zero because C i is acyclic. Also, the same sequence

shows that Hp(Ki) = Hp+1(Ki−1) for every p ≥ 1. The maps in these sequences

fit into the diagram

220

Page 221: Geir Ellingsrud and John Christian Ottem October 15, 2018€¦ · between algebra and geometry: a ne varieties correspond to nitely generated k-algebras without zero-divisors, and

Chapter 13. More on sheaf cohomology

From this, we see that

im(H0(C i)→ H0(K i)

)= im

(H0(C i)→ H0(C i+1)

)We furthermore have

H0(Ki−1) = ker(H0(C i)→ H0(C i+1)

).

Indeed, if σ ∈ H0(C i) maps to zero in H0(F i+1), then it maps to zero in H0(Ki),

and hence it lies in the image of H0(Ki−1). Hence

H0(F) = ker(H0(C i)→ H0(C i+1)

)In particular,

H0(F) = ker(H0(C 0)→ H0(C 1)

)= H0(C •(X))

and the theorem holds in degree p = 0. By the same token, we have

H0(Ki) = ker(H0(C i+1)→ H0(C i+2)

).

From (13.2.5), and the isomorphisms Hp(Ki) ' Hp+1(Ki−1) we therefore get

H i+1(C •(X)) = ker(H0(C i+1)→ H0(C i+2)

)/ im(H0(C i)→ H0(C i+1))

= H0(Ki)/ im(H0(C i)→ H9(Ki))

= H1(Ki−1)

= H2(Ki−2)

= · · ·= H i+1(K−1)

= H i+1(F)

221

Page 222: Geir Ellingsrud and John Christian Ottem October 15, 2018€¦ · between algebra and geometry: a ne varieties correspond to nitely generated k-algebras without zero-divisors, and

13.2. The Godement resolution

13.2.4 The Cech resolution

We also have the following sheafified version of the Cech complex. Given a sheaf

F on X and an open cover U , we set

Cp(U ,F) =∏

i0<···<ip

i∗F|Ui0∩···∩Uip

where i : Ui0 ∩ · · · ∩ Uip → X denotes the inclusion.

We will now show that the sheaves Cp(U ,F) give a resolution of F . So we

have an exact sequence

C•(U ,F) : 0→ F → C0(U ,F)d0−→ C1(U ,F)

d1−→ C2(U ,F)→ · · · (13.2.6)

In particular, if we apply Γ, we get back the Cech complex from earlier.

We define the maps in the complex as follows. The first map F → C0(U ,F)

is takes a section and restricts it to Ui. The differentials di are constructed in the

same way as in the Cech complex. Thus if we apply the global sections functor

Γ to Cp(U ,F) we get the earlier defined Cech complex.

The main difference here is that the new complex is exact. We can check

exactness on stalks and the nicest way to do that is with a homotopy operator

k : Cp(U , F )→ Cp−1(U ,F) so that (dk + kd)(α) = α: If we are given such a k,

every cocycle is a coboundary and so HpC•(U ,F) = 0.

For each x ∈ X, and open set V 3 x, refine V to V ∩Uj for some Uj 3 x. So

now, V ⊆ Uj. Define k by

(kα)i0,...,ip−1 = αj,i0,...,ip−1

Then it is easy to check that (dk + fd)(α) = α.

Proposition 13.14. If F is flasque, then so are the sheaves Cp(U ,F) for p > 0.

Hence (13.2.6) is an acyclic resolution for F and

Hp(X,F) = Hp(C•(U ,F))

Proof. If F is flasque, then so is each restriction to each Ui0∩···∩Up , and products

of flasque sheaves are flasque, so∏

i0<···<ip i∗F|Ui0∩···∩Up is flasque.

222

Page 223: Geir Ellingsrud and John Christian Ottem October 15, 2018€¦ · between algebra and geometry: a ne varieties correspond to nitely generated k-algebras without zero-divisors, and

Chapter 13. More on sheaf cohomology

13.3 Godement vs. Cech

It remains to see why these two definitions are equivalent. So let U = Ui be a

covering for F . We will assume that this is Leray in the sense that H i(UI ,F) = 0

for all multi-indices I and i > 0. We claim that there is a natural isomorphism

H i(X,F) ' H i(U ,F),

where we, in order to avoid confusion, let H i(U ,F) denote the Cech cohomology

group. The statement is clearly true for i = 0, since both coincide with Γ(X,F).

Lemma 13.15. Let 0→ F → G → H → 0 be an exact sequence. If U is Leray,

there is a long exact sequence

0 H0(U ,F) H0(U ,G) H0(U ,H)

H1(U ,F) H1(U ,G) H1(U ,H) −→ · · ·

Proof. Since U is Leray, we have H1(UI ,F) = 0 for all multi-indexes I (in fact,

this is the only property we need from the covering U). Hence the following

sequences are exact:

0→ F(UI)→ G(UI)→ H(UI)→ 0

Then applying the Cech complex, we get an exact sequence of complexes

0→ C•(U ,F)→ C•(U ,G)→ C•(U ,H)→ 0

Now the claim follows from Proposition 12.1.

Hence we get our desired theorem:

Theorem 13.16 (Leray). Suppose U is a cover of X and Hq(U1∩· · ·∩Up,F ) = 0

for all p, q > 0 and all U1, . . . , Up ∈ U . Then there is a natural isomorphism

between cohomology and Cech cohomology:

Hp(X,F) ' Hp(U ,F)

Proof. We use induction on p. For p = 0, the claim is clear. Note that we have

the exact sequence

0→ F → Π(F)→ Z1 → 0

and H1(X,F) = coker(Γ(X,Π(F))→ Γ(X,Z1F)), and

Hp(X,F) = Hp−1(X,Z1F)

223

Page 224: Geir Ellingsrud and John Christian Ottem October 15, 2018€¦ · between algebra and geometry: a ne varieties correspond to nitely generated k-algebras without zero-divisors, and

13.3. Godement vs. Cech

for p ≥ 2. On the other hand, we also have the corresponding result for Cech

cohomology:

0 H0(U ,F) H0(U ,Π(F)) H0(U ,Z1)

H1(U ,F) H1(U ,Π(F)) = 0

where H1(U ,F) = 0 by Lemma 13.13, since Π(F) is acyclic. Hence also

H1(X,F) = coker(Γ(X,Π(F))→ Γ(X,Z1F)) = H1(X,F)

Hence the theorem also holds for p = 1.

We continue by induction on p. Since also H i(U ,Π(F )) = 0 for all i > 0,

same long exact sequence of Cech cohomology also shows that Hp(U ,F) =

Hp−1(U ,Z1). Moreover, the cover U is also Leray with respect to Z1: H i(UI ,Z1) =

H i+1(UI ,F) = 0). Hence replacing F with Z1, we get the desired conclusion.

224

Page 225: Geir Ellingsrud and John Christian Ottem October 15, 2018€¦ · between algebra and geometry: a ne varieties correspond to nitely generated k-algebras without zero-divisors, and

Chapter 14

Divisors and linear systems

When studying a scheme it is natural to ask about its closed subschemes. We

have seen that such subschemes are in correspondence with quasi-coherent ideal

sheaves. For (integral) subschemes of codimension 1, these ideal sheaves tend

to be much simpler than for higher codimension. This is essentially because of

Krull’s Hauptidealsatz, which says roughly that (under certain hypotheses), this

is generated locally by one element. This is essentially the prototype of a Cartier

divisor; a subscheme which is locally cut out by one equation.

The prototype of a divisor is a hypersurface of projective space, that is, an

integral subscheme of Pn of codimension one. Each such subscheme is defined by

a homogeneous ideal I of k[x0, . . . , xn], and since the codimension is assumed to

be one, Krull’s theorem implies that I is principal, i.e., I = (f) for some degree

d polynomial f ∈ k[x0, . . . , xn]. To give a concrete example, consider the case of

P2, and the curve

D0 = V (x3 + y3 + z3) ⊆ P2k

This is clearly integral, since the defining equation is irreducible. Similarly, we

can consider the subscheme

D1 = V (xyz)

This subscheme is reduced, but not irreducible: D has three irreducible compo-

nents V (x), V (y), V (z).

The main feature of divisors to talk about sums and differences of such

subschemes, thereby turning them into a group. This is illustrated in the

example above, by writing D1 = V (x)+V (y)+V (z). The sum here is completely

formal – it is an element in the free group on integral subschemes of codimension

one. Such a sum is by definition, a Weil divisor.

225

Page 226: Geir Ellingsrud and John Christian Ottem October 15, 2018€¦ · between algebra and geometry: a ne varieties correspond to nitely generated k-algebras without zero-divisors, and

There is an equivalence relation defined on such objects, designed to capture

when two divisors belong to the same family. In the example D and D′ belong

to the same family, namely

D[s:t] = V (sxyz + t(x3 + y3 + z3))

More precisely, there is a closed subscheme D ⊆ P1 × P2 defined by the above

equation, so that the fiber over a closed point [s : t] ∈ P1 is exactly the curve

D[s:t] of P2k. Geometrically, we have a family of ‘moving divisors’, parameterized

over the projective line. We say the two divisors are linearly equivalent. The key

feature is that there is this morphism f : D → P1, and f−1([0 : 1]) = D0 and

f−1([1 : 0]) = D1. Moreover, the quotient

g =x3 + y3 + z3

xyz

defines a rational function on P2. This is the pullback of the rational function

s/t on P1: g = f ∗(s/t).

There is a second approach to divisors, which is perhaps more algebraic in

nature, namely Cartier divisors. The definition is motivated by the fact that

integral subschemes of codimension one are typically defined by a single equation

locally. Note that in the above example, the special properties of projective space

imply that D0 and D1 are defined globally by a single equation. It is not hard to

come up with examples of schemes with subschemes of codimension one that are

not. In fact, for most schemes, the concept of a ‘globally defined equation’ does

not make sense, since we do not have global coordinates to work with. However,

locally this concept makes sense: We can consider subschemes Y ⊆ X so that

the ideal sheaf IY is locally generated by a single element fi ∈ Ai, on some affine

covering X =⋃

SpecAi. In other words, the ideal sheaf IY is an invertible sheaf.

While Weil divisors are conceptual and more geometric, Cartier divisors do

have some advantages. For instance, they are very closely related to invertible

sheaves and line bundles, which in turn makes computations with them easier.

For instance, given a morphism f : X → Y , we would like to define a ‘pullback’

of a divisor on Y to a divisor on X – this turns out to only be possible for Cartier

divisors. There are also other settings where the special properties of Cartier

divisors are essential, for instance defining intersection products.

That being said, for a smooth scheme we will see that the various notions of

a divisor are equivalent, and there are natural ways to switch between the four

226

Page 227: Geir Ellingsrud and John Christian Ottem October 15, 2018€¦ · between algebra and geometry: a ne varieties correspond to nitely generated k-algebras without zero-divisors, and

Chapter 14. Divisors and linear systems

categories in the diagram

Weil divisors line bundles

Cartier divisors invertible sheaves

The main interest in these concepts is of course that they give us tools to classify

varieties and schemes. So given X, perhaps as an abstract scheme, we can first

look for a divisor D on it. If the corresponding invertible sheaf L = OX(D) is

globally generated, we have a morphism to projective space f : X → Pn and

we can use this to study the geometry of X: We can ask about the fibers of f ,

whether it is an closed immersion, and if so, what the equations of the image is.

We will in this chapter assume that

All schemes are integral and noetherian.

Noetherianness is somewhat important, since we want to talk about the decom-

position of a closed subschemes into its irreducible components. The assumption

of integrality, especially irreducibility, is not essential for most of the statements,

but it makes the definitions more transparent, and more importantly, the proofs

considerably simpler (e.g., not having to worry about nilpotent elements allows

us to work with meromorphic functions as elements in a fraction field, rather

some more obscure localization). Still, we remark that the theory of Cartier

divisors work in a completely general setting, although there are some subtle

points one has to take into account. If you are curious about general background

on meromorphic functions on arbitrary schemes, see EGA IV4, section 20.

14.1 Weil divisors

In this section X will denote an integral, normal noetherian scheme. Recall that

this means that each local ring OX,x is an integral domain, which is integrally

closed in its function field K = K(X).

Definition 14.1. A prime divisor on X is a closed integral subscheme Y of

codimension 1. A Weil divisor on X is a finite formal sum

D =∑

niYi (14.1.1)

where ni ∈ Z and Yi are prime divisors.

We say D is effective if all the ni are non-negative in (14.1.1).

We call⋃i Yi the support of D.

We denote by Div(X) the group of Weil divisors; this is the free abelian

group on prime divisors.

227

Page 228: Geir Ellingsrud and John Christian Ottem October 15, 2018€¦ · between algebra and geometry: a ne varieties correspond to nitely generated k-algebras without zero-divisors, and

14.1. Weil divisors

If Y is a prime divisor on X and U ⊆ X is an open set, then Y ∩ U is

naturally a prime divisor on U . It follows that we obtain a presheaf U 7→ Div(U).

This is in fact a sheaf; it coincides with⊕x∈X,codim x=1

jx∗(Zx)

where jx : x → X is the inclusion of a point x of codimension 1 in X. We will

denote this sheaf by Div.

14.1.1 The divisor of a rational function

The main reason for the normality assumption is that if X is normal, then it

is regular in codimension one. This implies that for each point η ∈ X with

codim η = 1, each local ring OX,η (which has Krull dimension 1) is a discrete

valuation ring, with a corresponding valuation vY : K×Z. The concept of a

valuation is a generalization of the ‘order’ of a zero or a pole of a meromorphic

function in complex analysis. Intuitively, an element f ∈ K× has positive

valuation N if it vanishes to order N along Y , and negative valuation −N if it

has a pole of order N there.

To say what vY is explicitly, we define for an element f ∈ OX,η,

vY (f) = n

where n is the integer so that f ∈ mn −mn−1. In the function field, an element

f is represented by a fraction g/h and we define vY (f) = vY (g)− vY (h). (Check

that this is independent of the chosen representative). This means that OX,η =

v−1Y (Z≥0), and the maximal ideal m = v−1

Y (Z≥1).

Definition 14.2. If f ∈ K×, we define its divisor as

div (f) =∑

vY (f)Y

Divisors of the form div (f) are called principal divisors, and they form a subgroup

Div0(X) ⊆ Div(X).

This is well defined by the following lemma:

Lemma 14.3. Suppose that X is an integral normal noetherian scheme with

fraction field K and let f ∈ K. Then vY (f) = 0 for all but finitely many prime

divisors Y .

Proof. We first reduce to the case when X is affine. Let U = SpecA be an open

affine subset such that f |U ∈ Γ(U,OX). Since X is noetherian, the complement

228

Page 229: Geir Ellingsrud and John Christian Ottem October 15, 2018€¦ · between algebra and geometry: a ne varieties correspond to nitely generated k-algebras without zero-divisors, and

Chapter 14. Divisors and linear systems

Z = X−U is a closed subset of X which has finitely many irreducible components;

in particular, only finitely many prime divisors Y are supported in Z. So we

reduce to the affine case X = SpecA and f ∈ Γ(X,OX), by ignoring these

finitely many components. Then vY (f) ≥ 0 and vY (f) > 0 if and only if Y is

contained in V (f); and since V (f) has only finitely many irreducible components

of codimension 1, we are done.

Proposition 14.4. Let f ∈ K×, and let Y ⊆ X be a prime divisor with generic

point η. Then

1. vY (f) ≥ 0 if and only if f ∈ OX,η

2. vY (f) = 0 if and only if f ∈ O×X,η.

If X is projective, then vY (f) ≥ 0 for all Y if and only if vY (f) = 0 for all Y if

and only if f ∈ k∗.

Example 14.5. Consider X = A1k = Spec k[t] and let K = k(t). Here prime

divisors in X correspond to closed points p ∈ X. Let f = x2(x−1)x+1

. Then vp(f) = 0

for all Y except when p = 0,±1, where we have

v0(f) = 2, v1(f) = 1, v−1(f) = −1

Hence the divisor of f is 2V (t) + V (t− 1)− V (t+ 1).

Example 14.6. Consider X = P1k = Proj k[u, v] and let K be the fraction field.

Let f = u2−v2(u−2v)2

. On D(u) we use the coordinate t = v/u so that K = k(t). Here

f = 1−t2(1−2t)2

, and the only non-zero valuations are

v1(f) = 1, v−1(f) = 1, v1/2(f) = −2

The rational function f does not have any poles or zeroes at the point [0 : 1], so

the divisor is [1 : 1] + [1 : −1]− 2[2 : 1].

14.1.2 The divisor of a section of an invertible sheaf

Let L be an invertible sheaf and let s be a rational section (a section defined over

an open set). We can define a divisor div s as follows. For Y a prime divisor, let

y be the generic point. Let U ⊆ X be a neighbourhood of y such that there is a

trivialization φ : L|U → OX |U . On U , φ(s) defines a rational section of OX(U),

and hence an element of K. We define vY (s) = vY (f). It is not hard to see that

this is independent of U .

229

Page 230: Geir Ellingsrud and John Christian Ottem October 15, 2018€¦ · between algebra and geometry: a ne varieties correspond to nitely generated k-algebras without zero-divisors, and

14.1. Weil divisors

Example 14.7. Let L = O(2) on X = P1k = Proj k[u, v] and let K be the

fraction field. Let f = v2

u+v. On D(u) we use the coordinate t = v/u so that

K = k(t). Here f = t2

1+t, and the only non-zero valuations are

v0(f) = 2, v−1(f) = −1.

So on D(u) we get the divisor 2[1 : 0] − [1 : −1]. On D(v), we can use the

coordinate t = u/v, where f becomes f = tt+1

. The non-zero valuations are the

following:

vt=0(f) = 1, vt=−1(f) = −1.

So we get [0 : 1]− [−1 : 1] = [0 : 1]− [1 : −1]. On X = D(u) ∪D(v), we then

find the divisor is

2[1 : 0]− [1 : −1] + [0 : 1].

14.1.3 The class group

Definition 14.8. We define the class group of X as

Cl(X) = Div(X)/Div0(X)

Two Weil divisors D,D′ are linearly equivalent (written D ∼ D′) if they have

the same image in Cl(X), or equivalently, that D −D′ is principal.

Example 14.9. In fact, any divisor on A1k for an algebraically closed field k is

principal. Indeed, if D =∑n

i=1 ni[pi] where ni ∈ Z and pi ∈ k, then

f =n∏i=1

(t− pi)ni

is an element of k(t)× with div (f) = D. It follows that Cl(A1k) = 0 in this case.

14.1.4 Projective space

Proposition 14.10. Cl(Pnk) = Z.

Write Pnk = ProjR, R = k[x0, . . . , xn]. Prime divisors on Pnk correspond to

height one prime ideals in R, i.e., p = (g) where g is a homogeneous irreducible

polynomial. We can use this to define the degree of a divisor, by taking the

degrees of the corresponding polynomials:

Div(Pn)→ Z∑niV (gi) 7→

∑ni deg gi

230

Page 231: Geir Ellingsrud and John Christian Ottem October 15, 2018€¦ · between algebra and geometry: a ne varieties correspond to nitely generated k-algebras without zero-divisors, and

Chapter 14. Divisors and linear systems

If f ∈ K(Pnk) is a rational function, then f can be written as a quotient of two

homogeneous polynomials of the same degree. (Indeed, K = K(Pn) = K(U) =

K(R(x0)) and, so every rational non-zero function on Pn gives rise to an element

of F of degree 0, via the inclusion K(R(x0))× → K. Conversely, any fraction

in K of degree 0 gives an element in R(x0).) Now, if f ∈ K(Pnk) is a rational

function, we can write it as f =∏

i fnii where the fi are irreducible coprime

polynomials in R and ni ∈ Z. Let us first show that

div f =∑

ni[V (fi)]

If Y ⊆ X is a prime divisor, let y ∈ X be the generic point. Y = V (g) for

some irreducible polynomial g of degree d. For any other polynomial h of degree

d, the fraction g/h is a generator of myOX,y. We can write f = (g/h)rf ′ with

r = ni if fi divides g (and 0 if no fi divides g) and f ′ a rational function not

containing g in the numerator or the denominator. Hence vY (f) = r, and hence

(f) =∑ni[V (fi)]. Note that deg div f =

∑ni deg fi = 0, because f is a rational

function. It follows that the degree map descends to a map

deg : Cl(Pn)→ Z∑niV (gi) 7→

∑ni deg gi

We claim that this is an isomorphism. Note that degH = 1, where H = V (x0)

is a hyperplane, so deg is surjective. Now, any Z =∑ni[V (fi)] in the kernel

of deg, must have∑ni deg fi = 0. Consider the element f =

∏i f

nii , which

now defines an element of K. Using the formula for div f above, we see that

Z = div f , and hence Z is a principal divisor, and so deg is injective.

Example 14.11. Consider the curve X as in Figure , given by X = V (y2x−x3 − xz2) ⊆ P2. By the above discussion any two lines L,L′ in P2 are linearly

equivalent. Since X is integral, then there are well-defined restrictions L|X and

L′|X , which are linearly equivalent divisors on X. The figure below shows one

example where L|X = P +Q+R and L′ = 2S + T .

14.1.5 Weil divisors and invertible sheaves

Let D be a Weil divisor on X. We define the sheaf OX(D) as follows:

OX(D)(U) = f ∈ K|(div f +D)|U ≥ 0 ∪ 0= f ∈ K| vZ(f) ≥ −vZ(D), ∀Z ∈ U a point of codim 1 ∪ 0

This is a quasi-coherent sheaf on X. (We will see soon that it is invertible if and

only if D is a Cartier divisor).

231

Page 232: Geir Ellingsrud and John Christian Ottem October 15, 2018€¦ · between algebra and geometry: a ne varieties correspond to nitely generated k-algebras without zero-divisors, and

14.1. Weil divisors

Example 14.12. Let X be the projective line P1k = Proj k[x0, x1] over k and let

D = V (x0) = [1 : 0]. Let U0 = Spec k[x1/x0] = Spec k[t], U1 = Spec k[x0/x1] =

Spec k[s] be the standard covering of P1k (so s = t−1 on U0 ∩ U1). Note that the

point [1 : 0] does not lie in U0 = D+(x0). This means that a rational function

f ∈ K such that f +D is effective on U0 must be regular on U0, i.e.,

Γ(U0,OX(D)) = k[t]

On U1, we are looking at elements f ∈ k(s) having valuation ≥ −1 at s = 0.

This implies that

Γ(U1,OX(D)) = k[s]⊕ k[s]s−1

Now we think of elements in Γ(X,OX(D)) as pairs (f, g) with f, g ∈ K sections

of OX(D) over U0 and U1 respectively, so that f = g on U0 ∩ U1. Here f = f(t)

is a polynomial in t, and

g(s) = p(s) + q(s)s−1 = p(t−1) + q(t−1)t.

If f = g in k[t, t−1] we must deg p, q ≤ 1. This implies that

Γ(X,OX(D)) = k ⊕ kt.

14.1.6 The class group and unique factorization domains

The term ‘class group’ comes from algebraic number theory and its origins be

traced back Kummer’s work on Fermat’s last theorem. If A is a Dedekind domain

then Cl(SpecA) coincides with the class group of A, which measures how far A

is from being a unique factorization domain.

Proposition 14.13. Let A be a noetherian integral domain and let X = SpecA.

232

Page 233: Geir Ellingsrud and John Christian Ottem October 15, 2018€¦ · between algebra and geometry: a ne varieties correspond to nitely generated k-algebras without zero-divisors, and

Chapter 14. Divisors and linear systems

Then A is a UFD if and only if Cl(SpecA) = 0 and X is normal.

Proof. If A is a UFD, then it is integrally closed, and hence normal. Moreover,

A is a UFD if and only if every prime ideal p of height 1 is principal. We want

to show that this condition is equivalent to Cl(X) = 0.

⇒: If Y ⊆ X is a prime divisor, then Y = V (p) for p of height 1. Hence

Y = V (f) = div f for some f , and so Cl(X) = 0.

⇐: If Cl(X) = 0, let p be a prime of height 1, and let Y = V (p) ⊆ X. By

assumption, there is a f ∈ K× such that div (f) = Y . We want to show that

f ∈ A and that p = (f).

vY (f) = 1, so f ∈ Ap and f generates pAp. If q ⊆ A is any prime ideal not

equal to p, then Y ′ = V (q) is a prime divisor not equal to Y . vY ′(f) = 0, since

(f) = Y , and so f ∈ Aq. So in all, we have

f ∈⋂

q∈SpecA

Aq = A

where the equality comes from the fact that A is an integrally closed domain and

the ‘algebraic Hartog’s theorem’: A function which is regular in codimension 1

is regular (see Matsumura p. 124). Hence f ∈ A, and in fact f ∈ A ∩ pAp = p.

We show that f generates p: Take any f ∈ p. Then vY (g) ≥ 1 amd vY ′(g) ≥ 0

for all Y ′ 6= Y . It follows that vY ′(g/f) ≥ 0 for all prime divisors Y ′. Hence

g/f ∈ Aq for all q prime of height 1, and hence g/f ∈ A, by the above. It follows

that g ∈ fA and so p = fA is principal.

14.1.7 A useful exact sequence

Given a scheme X and an open subset U , the restriction of a prime divisor on

X is a prime divisor on U , so it is natural to ask how the two class groups are

related. The answer is given by the following theorem:

Theorem 14.14. Let X be a normal, integral scheme, let Z ⊆ X be a closed

subscheme and let U = X−Z. If Z1, . . . , Zr are the prime divisors corresponding

to the codimension 1 components of Z, then there is an exact sequence

r⊕i=1

ZZi → Cl(X)→ Cl(U)→ 0 (14.1.2)

where the map Cl(X)→ Cl(U) is defined by [Y ] 7→ [Y ∩ U ].

Here we regard each Zi as a subscheme of X with the reduced scheme structure

(cf. Proposition 8.37)

233

Page 234: Geir Ellingsrud and John Christian Ottem October 15, 2018€¦ · between algebra and geometry: a ne varieties correspond to nitely generated k-algebras without zero-divisors, and

14.2. Cartier divisors

Proof. If Y is a prime divisor on U , the closure in X is a prime divisor in X, so

the map is surjective.

We just need to check exactness in the middle. Suppose D is a prime divisor

which is principal on U . Then D|U = div f for some f ∈ K(U) = K = K(X).

Now div f is a divisor on X such that D|U = div (f)|U . Hence D − div (f) is

supported in X − U , and hence it is a linear combination of the Zi.

As a special case, we see that removing a codimension 2 subset does not

change the group of Weil divisors. So for instance Cl(A2 − 0) = Cl(A2).

Example 14.15. Consider the projective line P1k over a field k, and let p be a

point. We have the exact sequence

Z[p]→ Cl(P1)→ Cl(A1)→ 0

We saw that Cl(A1) = 0, so the map Z→ Cl(P1) is surjective. It is also injective:

If [nP ] = 0 in Cl(P1) for some n. Then nP = div f for some f ∈ k(P1), so if

div f |A1 = 0, we have f ∈ Γ(A1,O) = k∗. Hence f is constant, and so n = 0. It

follows that Cl(P1) = Z.

14.2 Cartier divisors

Let X be a noetherian integral scheme with function field K. Let KX denote

the constant sheaf with value K. The constant sheaf K ×X with value K× (the

group of non-zero elements of K) is a subsheaf, and contains O×X , the sheaf of

units in O×X .

Definition 14.16. A Cartier divisor D on X is an element of Γ(X,K ×X /O

×X).

This definition of a Cartier divisor is pretty obscure, and one rarely thinks of

a divisor in this way. We can shed some light on the definition by considering an

open cover Ui of X. The sheaf axiom sequence takes the following form

0→ Γ(X,K ×X /O

×X)→

∏i

Γ(Ui,K×X /O

×X)→

∏i,j

Γ(Uij,K×X /O

×X)

From this, we see that a Cartier divisor is therefore given as a set of sections of

K ×X /O

×X over the open sets Ui that agree on the overlaps Uij.

If we shrink the Ui we may assume that si are induced by a section fi of

K ×X over Ui, that is, an element of K×. To see what it means that two such

data Ui, fi agree on Uij we keep in mind that they are sections of K ×X /O

×X .

The condition si|Uij = sj|Uij translates into the statement that fi|Uijf−1j |Uij is a

section of O×X over Uij. In other words, there are units cij ∈ OX |Uij such that

234

Page 235: Geir Ellingsrud and John Christian Ottem October 15, 2018€¦ · between algebra and geometry: a ne varieties correspond to nitely generated k-algebras without zero-divisors, and

Chapter 14. Divisors and linear systems

fi = cijfj over Uij. In K× we have cij = fifj

, which implies that the cij satisfy

the ‘cocyle conditions’

cik = cijcjk; cji = c−1ij ; cii = 1

The idea here is that the fi give the local equations for a subscheme of X. On

an affine open set U where fi is regular, the zero-set V (fi) defines a codimension

1 closed subset of U . If both fj and fi are regular on U , their zero-sets must be

the same since they differ by a unit.

Definition 14.17. The pairs (Ui, fi) are called the local defining data or the

local equations for the divisor D (with respect to the covering Ui).

Such defining data are not unique: (Ui, fi)i∈I and (Vj, gj)i∈I give the

same Cartier divisor if fig−1j ∈ Γ(Ui ∩ Vj,O×X) for all i, j.

Now, the set of Cartier divisors naturally forms an abelian group: Given

D and D′ represented by the data (Ui, fi)i∈I and (Vi, gi)i∈I , we can define

D +D′ as the Cartier divisor associated to the data

(Ui ∩ Vj, figj)i,j

Moreover, the inverse −D will be defined as (Ui, f−1i )i∈I , and the identity is

defined by the data (Ui, fi)i∈I where fi ∈ Γ(Ui,O×X). We denote the group of

Cartier divisors by CaDiv(X) = Γ(X,KX/O×X).

Definition 14.18. We say that a Cartier divisor is principal if it is equal (as an

element of CaDiv(X)) to (X, f) where f ∈ K×.

The set of principal Cartier divisors form a subgroup of CaDiv(X), which

is typically much smaller than CaDiv(X). However, by definition, any Cartier

divisor is ‘locally principal’.

We define CaCl(X) to be the group of Cartier divisors modulo principal

divisors:

CaCl(X) = CaDiv(X)/(X, f)|f ∈ K×

We say that two Cartier divisors D,D′ are linearly equivalent, if D−D′ = (X, f)

for some principal divisor (X, f), or equivalently, [D] = [D′] in CaCl(X).

14.2.1 Cartier divisors and invertible sheaves

Let D be a Cartier divisor on a scheme X given by the data (Ui, fi). We can

associate to it an invertible sheaf, which we denote by OX(D). On Ui we take

the the subsheaf f−1i OUi ⊆ KUi of KX . This subsheaf is isomorphic to OUi and

has f−1i as a local generator. So over an affine subset U = SpecA ⊆ Ui, the

sheaf is the sheaf associated to f−1i A ⊆ K(A).

235

Page 236: Geir Ellingsrud and John Christian Ottem October 15, 2018€¦ · between algebra and geometry: a ne varieties correspond to nitely generated k-algebras without zero-divisors, and

14.2. Cartier divisors

On the intersection, Ui ∩ Uj, we have fi = cijfh where cij is an invertible

section of OUij . This means that f−1i OUij = f−1

j OUij as subsheaves of KUij . We

have therefore, constructed a sheaf on each Ui, and the elements coincide on

the intersections Uij. In order to be able to glue to a sheaf, there is a cocycle

condition that has to be satisfied. But since these sheaves are all subsheaves of

a fixed sheaf K , the gluing maps are actually identity maps, and the cocycle

condition is automatically satisfied. It follows that the sheaves f−1i OUi glue to

a sheaf OX(D) defined on all of X. It is by construction invertible, since it is

invertible on each Ui.

Explicitly, OX(D) is defined as the subsheaf of KX given by

Γ(V,OX(D)) = f ∈ K|fif ∈ Γ(Ui ∩ V,OX)∀i

Two different data (Ui, fi) and (Vj, gj) for the same divisor D give rise to the

same invertible sheaf. This is because over Ui ∩ Vj, we have fi = dijgj for some

sections dij ∈ OX(Ui ∩ Vj)×. This means that f−1i OUi∩Vj = g−1

i OUi∩Vj , and so

the sheaf is uniquely determined.

Moreover, by how we defined the sum D +D′ we have

Proposition 14.19. Let X be an integral noetherian scheme and let D and D′

be two Cartier divisors.

(i) OX(D +D′) ' OX(D)⊗ OX(D′)

(ii) OX(D) = OX(D′) if and only if D and D′ are linearly equivalent.

Proof. We can pick a covering Ui so that both D and D′ are both represented

by data (Ui, fi), (Ui, f′i). Then D +D′ is defined by (Ui, fif

′i). Locally, over Ui

the sheaf OX(D +D′) is defined as the subsheaf of K given by (fif′i)−1OUi =

f−1i f ′−1

i OUi . The tensor product is locally(!) given as f−1OUi ⊗ f−1OUi , which

is clearly isomorphic to f−1i f ′−1

i OUi via af−1i ⊗ bf ′i

−1 7→ abfi−1fj

−1.

For the second claim, it suffices (by point (i)) to show that OX(D) ' OX if

and only if D is a principal Cartier divisor. So suppose that OX(D) ⊆ KX is a

sub OX-module which is isomorphic to OX . Then the image of 1 in OX will be a

section of OX(D) which generates OX(D) everywhere. This means that (X, f)

is the local defining data, and D is a principal Cartier divisor.

Conversely, if D = (X, f), then OX(D) = f−1OX , and multiplication by f

gives an isomorphism OX(D) ' OX .

14.2.2 CaCl(X) and Pic(X)

By the item (ii) in Proposition 14.19. We see that the map CaDiv(X)→ Pic(X),

which sends D to the class of OX(D) in Pic(X) is additive and has the subgroup

236

Page 237: Geir Ellingsrud and John Christian Ottem October 15, 2018€¦ · between algebra and geometry: a ne varieties correspond to nitely generated k-algebras without zero-divisors, and

Chapter 14. Divisors and linear systems

of principal divisors CaDiv0(X) as its kernel. This means that the induced map

ρ : CaCl(X)→ Pic(X) is injective. In this section we will show that this map is

also surjective, so that CaCl(X) = Pic(X).

Proposition 14.20. When X is integral, the map ρ : CaCl(X)→ Pic(X) is an

isomorphism.

Proof. We need to show that ρ is surjective. It suffices to show that any invertible

OX-module L is a submodule of KX : If L ⊆ K , let Ui be a trivializing cover of

L and let gi be its local generators. Then we have L|Ui = giOUi ⊆ KUi and the giare rational functions on Ui. On Uij = Ui ∩ Uj, we have giOUij = L|Uij = gjOUij ,

and it follows that gi = cijgj for units cij ∈ O×Uij . Consequently, (Ui, g−1i ) forms

a set of local defining data for a Cartier divisor D, and of course we have

L = OX(D).

Let L be an invertible sheaf and consider the sheaf L⊗OX KX . Let Ui ⊆ X

be a set of open subsets such that L|Ui = OX |Ui . Note that the restriction of

L ⊗OX KX ' KX to each Ui is a constant sheaf. Since X is irreducible, any

sheaf whose restriction to opens in a covering is constant, is in fact a constant

sheaf, and therefore L⊗OX KX ' KX as sheaves on X. Now we can regard L

as a rank 1 subsheaf of KX using the composition L→ L⊗KX ' KX . Hence

we get a Cartier divisor D on X such that L = OX(D).

14.2.3 From Cartier to Weil

Let D be a Cartier divisor given by the data (Ui, fi). If Y is a prime divisor on

X, with generic point η, then since Ui is a cover, η lies in some Ui. We can then

define

vY (D) = vY (fi)

This is independent of the choice of Ui: If η ∈ Ui ∩ Uj, then fif−1j is an element

of O×X(Ui ∩ Uj), and so vY (fif−1j ) = 0, and hence vY (fi) = v(fj).

This gives a map φ : CaDiv(X)→ Div(X) defined by

φ(D) =∑Y

vY (D)Y

Proposition 14.21. φ is injective.

Indeed, if φ(D) = 0, then all vY (fi) = 0 for all i (so that ηY ∈ Ui). Then the

Hartogs’ lemma implies that fi ∈ O×(Ui), so that D = 0 (as a Cartier divisor).

Note that by the definition of (f), a principal Cartier divisor (X, f) gives rise

to a principal Weil divisor. It follows that we get an induced map CaCl(X)→Cl(X)

237

Page 238: Geir Ellingsrud and John Christian Ottem October 15, 2018€¦ · between algebra and geometry: a ne varieties correspond to nitely generated k-algebras without zero-divisors, and

14.2. Cartier divisors

Proposition 14.22. CaCl(X) = Div(X) if and only if X is locally factorial

(all the local rings OX,x are UFDs).

Proof. The inclusion CaDiv(X) ⊆ Div(X) is an isomorphism if and only if the

map of sheaves K ×X → Z1 is surjective, i.e., the map is surjective for every stalk.

However, at the stalk at x ∈ X, this map is simply

K× → Div(Spec OX,x)

which sends f to the sum∑vp(f)V (p) where p is a height 1 prime ideal of OX,x.

This is surjective if and only if every p ⊆ OX,x is a principal ideal. This happens

if and only if OX,x is a UFD.

Corollary 14.23. On a non-singular variety X, there is a bijective correspon-

dence between

(i) Line bundles

(ii) Invertible sheaves

(iii) Cartier divisors

(iv) Weil divisors

up to isomorphism.

From this, we get the following theorem:

Theorem 14.24. Let k be a field. Then Pic(An) = Cl(An) = CaCl(An) = 0.

Proof. The polynomial ring is a UFD.

14.2.4 Projective space

Let us take a closer look at the projective space Pnk over a field k. Write Pnk =

ProjR where R = k[x0, . . . , xn]. Consider the standard covering Ui = D+(xi) of

Pnk .

Let F (x0, . . . , xn) ∈ Rd denote a homogeneous polynomial of degree d. Then

the function F (x/xi) = F (x0xi, . . . , xi−1

xi, 1, xi+1

xi, . . . , xn

xi) defines a non-zero regular

function on Ui, and the collection

(Ui, F (xj/xi))

forms a Cartier divisor on X. Indeed, on the overlap Ui∩Uj we have the relation

F (x/xi) = (xj/xi)n F (x/xj)

238

Page 239: Geir Ellingsrud and John Christian Ottem October 15, 2018€¦ · between algebra and geometry: a ne varieties correspond to nitely generated k-algebras without zero-divisors, and

Chapter 14. Divisors and linear systems

and xj/xi is a regular, and invertible function. The corresponding invertible sheaf

is OPnk (d). Two homogeneous polynomials F,G of the same degree d give linearly

equivalent divisors, because the quotient F (x)/G(x) is a global rational function

on Pnk . In particular, this applies to F and xd0, and we get OPnk (n) = OPk(1)⊗n.

Corollary 14.25. On Pnk any invertible sheaf is isomorphic to some O(m).

In the case of P1 over a field k, the above result is not so difficult to prove

directly. Indeed, recall that P1 is obtained via gluing U0 = Spec k[s] and

U1 = Spec k[t] along U01 = Spec k[s, s−1] = Spec k[t, t−1] using the identification

t = s−1. So given an invertible sheaf L on P1, the restriction of it to each open

must be trivial (since Pic(A1k) = 0), we must have isomorphisms φi : L|Ui → OUi .

In particular, over U012 we obtain two isomorphisms φi|U01 : L|U01 → OU01 . In

particular, the map ψ : φ1φ−10 : OU01 → OU01 is an isomorphism. This is induced

by a map k[s, s−1]→ k[s, s−1] which is just multiplication by some polynomial

p(s, s−1). Note that p(s, s−1) cannot vanish at any point on U01 (otherwise ψ

would not be an isomorphism over the stalk at that point). Hence p(s, s−1) = sn

for some n ∈ Z. But that implies that L ' OP1(n).

14.3 Effective divisors

We say that a Cartier divisor D is effective (and write D ≥ 0) if D can be

represented by local data (Ui, fi) where the rational functions (which a priori

are just assumed to be sections of KX over Ui) actually lie in OX(Ui). It is

straightforward to verify that if this is true for one set of data, then it holds for

any other set as well.

We will write D ≥ D′ if D −D′ ≥ 0, i.e., the difference D −D′ is effective.

Note also that if D and D′ are both effective, then so is D + D′. This means

that Pic(X) is an ordered group.

Effective divisors are of paramount importance in algebraic geometry; they

carry lots of essential geometric information. With the set up above, D being

effective is equivalent to the statement that OX(−D) (regarded as a subsheaf of

KX is contained in OX . The OX-module OX(−D) is a coherent sheaf of ideals

which is locally generated by one element (in other words, it is locally principal).

We will usually denote the corresponding closed subscheme also by D (this is a

somewhat bad example of abusing the notation, but it has its advantages).

The inclusion OX(−D) ⊆ OX induces an exact sequence

0→ OX(−D)→ OX → OD → 0

where the right hand side can be interpreted both as the cokernel of the left-most

map and as the structure sheaf of the subscheme associated to D. Moreover, the

239

Page 240: Geir Ellingsrud and John Christian Ottem October 15, 2018€¦ · between algebra and geometry: a ne varieties correspond to nitely generated k-algebras without zero-divisors, and

14.3. Effective divisors

support of the OD (i.e., the set of points x ∈ X, such that OD,x 6= 0) coincides

with underlying topological space of D.

So effective divisors give rise to closed subschemes of codimension one. The

converse is also true: If D is a closed subscheme which is locally defined by

one equation (which is not a zero-divisor), then we can find, for any x ∈ X,

an open affine neighbourhood Ux and an element fx ∈ A = Γ(Ux,OX) so

that D = V (fx) = SpecA/(fxA). Two such elements fx, f′x generate the same

principal ideal, so there must be a relation fx = cxf′x for some cx ∈ A×. From this,

we see that (Ux, fx) form the defining data for a Cartier divisor, which we will

also denote by D. Indeed, on an overlap Uxy = Ux∩Uy we have fx|Uxy = cxyfy|Uxyfor a section cxy of O×Uxy . We have therefore proved the following theorem:

Theorem 14.26. Let X be an noetherian, integral scheme. Then there is a

one-to-one correspondence between closed subschemes D ⊆ X locally defined by

a principal ideal, and effective Cartier divisors on X.

The inclusion OX(−D) ⊆ OX can be dualized, i.e., we apply the functor

HomOX (−,OX), and obtain a map α : OX → OX(D). Such a map is uniquely

determined by its value on 1, i.e., the global section σ = α(1) ∈ Γ(X,OX(D)).

Conversely, given a global section Γ(X,OX(D)), we dually a map OX(−D)→OX . When X is integral, then this gives us a divisor, which is of course the

effective divisor D. We have therefore proved:

Theorem 14.27. Let X be an noetherian, integral scheme and let D be a Cartier

divisor on X. Then D is effective if and only if OX(D) has a non-zero global

section. For each section σ, we get a divisor denoted by div σ. Two such sections

σ, σ′ give rise to the same divisor if and only if σ = cσ′ where c ∈ Γ(X,OX)×.

14.3.1 Linear systems

Definition 14.28. The set of effective divisors D′ linearly equivalent to D is

denoted by |D|. This is called the complete linear system of D.

The name ‘linear system’ comes from the special case when X is a projective

variety X over a field k (thus X is integral, separated of finite type over k). In

this case, we have Γ(X,OX)× = k×, and the previous discussion shows that the

linear system |D| is given by

|D| = D′|D′ ≥ 0 and D′ ∼ D= (Γ(X,OX(D))− 0) /k×

= PΓ(X,OX(D))

So the linear system |D| is (as a set) a projective space. When X is projective,

the cohomology groups H0(X,OX(D)) is a finite dimensional (we will prove

240

Page 241: Geir Ellingsrud and John Christian Ottem October 15, 2018€¦ · between algebra and geometry: a ne varieties correspond to nitely generated k-algebras without zero-divisors, and

Chapter 14. Divisors and linear systems

this fact in Chapter 17), so the set of divisors D′ linearly equivalent to D is

parameterized by a projective space Pnk .

Definition 14.29. A linear system of divisors is a linear subspace of a complete

linear system |D|.

Example 14.30. Consider the case X = Pnk and D = dH, where H is the

hyperplane divisor (so H is a Cartier divisor with OX(H) ' OX(1). In this case

the linear system of D associated to OX(dH) is given by

|D| = PRd ' PN

where N =(n+dd

)− 1. The points of this projective space correspond to degree

d hypersurfaces.

14.4 Examples

14.4.1 The quadric surface

Let k be a field, and let Q = P1k×P1

k. Recall that Q embeds as a quadric surface

in P3k via the Segre embedding:

So we can view Q both as a fiber product P1×P1 and the quadric V (xy−zw) ⊆ P3.

Weil divisors on Q

Since Q is a product of two P1s there are natural ways of constructing Weil

divisors on Q from those on P1. For instance, we can let

L1 = [0 : 1]× P1 ⊆ Q,

which is a prime divisor on Q corresponding to the ‘vertical fiber’ of Q. Similarly,

L2 = P1 × [0 : 1] is a Weil divisor on Q.

ZL1 ⊕ ZL2 → Cl(Q)→ Cl(Q− L1 − L2)→ 0

241

Page 242: Geir Ellingsrud and John Christian Ottem October 15, 2018€¦ · between algebra and geometry: a ne varieties correspond to nitely generated k-algebras without zero-divisors, and

14.4. Examples

We claim that the first map is injective, and in fact that

Cl(Q) = ZL1 ⊕ ZL2.

For this, we need to analyse the two divisors L1, L2 a bit closer.

Cartier divisors on Q

To study Cartier divisors on Q, we first need a covering. Q is covered by four

affines

U00 = Spec k[x, y] U10 = Spec k[x−1, y]

U01 = Spec k[x, y−1] U11 = Spec k[x−1, y−1]

Consider P1k = W0 ∩ W1, where W0 = Spec k[t],W1 = Spec k[t−1]. The first

projection p1 : Q→ P1k is induced by the ring maps

k[t]→ k[x, y] k[t−1]→ k[x−1, y]

t 7→ x t−1 7→ x−1

k[t−1]→ k[x, y−1]; k[t−1]→ k[x−1, y−1];

t 7→ x t−1 7→ x−1

Let p = [0 : 1] be the Weil divisor on P1. The Cartier data is given by

(W0, t), (W1, 1). The pullback D = p∗1(p) is a Cartier divisor on Q, corresponding

to the Weil divisor [0 : 1]×P1. Writing K = k(x, y), the Cartier data is given by

(U00, x), (U10, 1)

(U01, x), (U11, 1)

Let L1 = [0 : 1] × P1k and L2 = P1

k × [0 : 1]. Consider the restriction of D to

L2. L2 is covered by the two open subsets V0 = U00 ∩ L2 = Spec k[x, y]/y, V1 =

U10 ∩ L2 = Spec k[x−1, y]/y. In terms of these opens, D|L2 has Cartier data

(V0, x), (V1, 1)

obtained by restricting the data above. In particular, identifying L2 ' P1, we

see that OQ(D)|L1 ' OP1(1). In particular, since Cl(P1) = Z, no multiple nD is

equivalent to 0 in Cl(Q) – if that were the case, we would have OQ(nD) ' OQ,

and hence OQ(nD)|L2 ' OQ|L2 ' OP1 , a contradiction.

In all, this shows that the first map in the exact sequence (14.4.1) is injec-

tive. We also have that Cl(Q − L1 − L2) = 0: Indeed Q − L1 − L2 = U11 =

242

Page 243: Geir Ellingsrud and John Christian Ottem October 15, 2018€¦ · between algebra and geometry: a ne varieties correspond to nitely generated k-algebras without zero-divisors, and

Chapter 14. Divisors and linear systems

Spec k[x−1, y−1] which is the Spec of a UFD. Hence

Cl(Q) ' ZL1 ⊕ ZL2

If D is a divisor on Q, D ∼ aL1 + bL2 and we call (a, b) the ‘type’ of D.

The canonical divisor of Q

We saw in the previous section that Cl(Q) ' Z2. Thus it makes sense to ask how

to represent the canonical divisor of Q in terms of L1, L2. Since Q = P1 × P1

and KP1 = −2p, it should perhaps not come as a big surprise that

KQ = −2L1 − 2L2

In fact, ΩQ = p∗1ΩP1 ⊕ p∗2ΩP1 . So taking∧2, we get

ωQ =2∧

(p∗1ΩP1 ⊕ p∗2ΩP1)

= p∗1ΩP1 ⊗ p∗2ΩP1

= OQ(−2, 0)⊗ OQ(0,−2) = OQ(−2,−2)

A second way of seeing this, is to use the fact that Q is embedded by the Segre

embedding Q ⊆ P3 as a smooth quadric surface. Then the Adjunction formula

of Proposition 11.43 on page 194 tells us that

ΩQ = ωP3 ⊗ OP3(Q)|Q = OP3(−4)⊗ OP3(2)|Q = OQ(−2)

This gives the same answer as before, since OP3(1)|Q = OQ(L1 + L2).

14.4.2 The quadric cone

Let X = SpecR where R = k[x, y, z]/(xy − z2), and k has characteristic 6= 2.

Let Z = V (x, z) be the closed subscheme corresponding to the line x = z = 0.Note that Z ' Spec k[x, y, z]/(xy−z2, x, z) = Spec k[y] is integral of codimension

1.

243

Page 244: Geir Ellingsrud and John Christian Ottem October 15, 2018€¦ · between algebra and geometry: a ne varieties correspond to nitely generated k-algebras without zero-divisors, and

14.4. Examples

A singular quadric surface

Note that X −Z = Spec k[x, y, y−1, z]/(xy− z2) = Spec k[y]y[t, u]/(t− u2) =

Spec k[y]y[u] which is a a UFD. It follows that Cl(X − Z) = 0. Recall now the

sequence

Z→ Cl(X)→ Cl(X − Z)→ 0

where the first map sends 1 to [Z]. Hence Cl(X) is generated by [Z].

We first show that 2Z = 0 in Cl(X). This is because we can consider the

divisor of x. This can only be non-zero along Z. The valuation here is 2: The

local ring is

(k[x, y, z]/(xy − z2))(x,z)

since y is invertible here, we see that x ∈ (z2) and that z is the uniformizer.

Now we show that Z is not a principal divisor. It suffices to prove that this

is not principal in Spec OX,x where x ∈ X is the singular point of X. The local

ring here is

(k[x, y, z]/(xy − z2))(x,y,z)

In this ring p = (x, z) is a height 1 prime ideal, but it is not principal: Let

m ⊆ OX,x be the maximal ideal. Note that x, y ∈ m, since x, y are not units.

However, we can compute that the vector space m/m2 (which is the Zariski

cotangent space at x) is 3-dimensional, spanned by x, y, z. Then I = (y, z) ⊆ m,

and x, y give a 2-dimensional subspace I/m2 of m/m2. Hence, since x and y

are linearly independent here, there couldn’t be an element f ∈ OX,x for which

x = af, y = bf . Hence

Cl(X) = Z/2.

Note that X − 0, 0, 0 is factorial. Hence removing a codimension 2 subset

has an effect on CaCl(X). Recall however, that the class group of Weil divisors

Cl(X) stays unchanged under removing a codimension 2 subset.

Projective quadric cone

Let X = ProjR where R = k[x, y, z, w]/(xy − z2). Let H = V (w) be the

hyperplane determined by w. We have

0→ ZH → Cl(X)→ Cl(X −H)→ 0

(Here H is an ample divisor, hence non-torsion in Cl(X), so the first map is

injective). X −H is isomorphic to the affine quadric cone from before, hence

Cl(X −H) = Z/2. Using this sequence, we see that Cl(X) = Z, generated by12H.

244

Page 245: Geir Ellingsrud and John Christian Ottem October 15, 2018€¦ · between algebra and geometry: a ne varieties correspond to nitely generated k-algebras without zero-divisors, and

Chapter 14. Divisors and linear systems

14.4.3 Quadric hypersurfaces in higher dimension

Theorem 14.31 (Nagata’s lemma). Let A be a noetherian integral domain, and

let x ∈ A− 0. Suppose that (x) is prime, and that Ax is a UFD. Then A is a

UFD.

Proof. We first show that A is normal. Of course Ax is normal, being a UFD.

So if t ∈ K(A) is integral in A, it lies in Ax. We need to check that if a/xn ∈ Axis integral over A and x 6 |a, then n = 0. If we have an integral relation

(a/xn)N + b1(a/xn)N−1 + · · ·+ bN = 0

Multiplying by xnN give get aN ∈ xA, so x|a, because A is an integral domain.

This shows that the divisor D = div x is an effective divisor and so there is an

exact sequence

ZD → Cl(SpecA)→ Cl(SpecAx) = 0→ 0

The image of the left-most map is 0, so Cl(A) = 0, and so A is a UFD.

Let A = k[x1, . . . , xn, y, z]/(x21 + · · · + x2

m − yz). We will prove that A is a

UFD for m ≥ 3. A is a domain, since the defining ideal is prime. Apply Nagata’s

lemma with the element y:

Ay = k[x1, . . . , xn, y, y−1, z]/(y−1(x2

1 + · · ·+ x2m)− z) ' k[x1, . . . , xn, y, y

−1, z]

which is a UFD. We show that y is prime: Taking the quotient we get

A/y = k[x1, . . . , xn, x]/(x21 + · · ·+ x2

m)

which is an integral domain, because x21 + . . . , x2

m is irreducible (for m ≥ 3).

Note that for m = 2, we get the quadric cone, which we have seen is not a

UFD.

Applying a change of variables, we find the following:

Proposition 14.32. Let k be a field containing√−1 and let X = V (x2

0 + · · ·+x2m) ⊆ An+1

k = Spec k[x0, . . . , xn].

1. m = 2, Cl(X) = Z/2

2. m = 3, Cl(X) = Z

3. m ≥ 4, Cl(X) = 0

Proposition 14.33. Let X = V (x20 + · · ·+ x2

m) ⊆ Pn = Proj k[x0, . . . , xn].

1. m = 2, Cl(X) = Z

245

Page 246: Geir Ellingsrud and John Christian Ottem October 15, 2018€¦ · between algebra and geometry: a ne varieties correspond to nitely generated k-algebras without zero-divisors, and

14.4. Examples

2. m = 3, Cl(X) = Z2

3. m ≥ 4, Cl(X) = Z

14.4.4 Hirzebruch surfaces

Let r ≥ 0 be an integer and consider the scheme X which is glued together by

the four affine scheme charts

U00 = Spec k[x, y]

U01 = Spec k[x, y−1]

U10 = Spec k[x−1, xry]

U11 = Spec k[x−1, x−ry−1]

This is a non-singular, integral 2-dimensional scheme over k. When k = C, this

are the so-called r-th Hirzebruch surface. In many ways, these behave as the

’mobius strips’ in algebraic geometry.

Note in particular, when r = 0, we get P1k × P1

k.

Divisors

Let us define two divisors D1, D2 by the following Cartier divisors:

Writing K = k(x, y), the Cartier data is given by

D1 =

[(U00, x), (U01, x)

(U10, 1), (U11, 1)

], D2 =

[(U00, y), (U01, 1)

(U10, y), (U11, 1)

]We will show that D1, D2 generate Pic(X). Note that both of these divisors are

effective, since they are defined by rational functions which are regular on the

Uij . In fact D1 ' P1, since it is glued together by V0 = U00∩x = 0 = Spec k[y]

and V1 = U01 ∩ x = 0 = Spec k[y−1]. Moreover, D2 restricted to D1 is given

by the Cartier divisor (V0, y), (V1, 1), which corresponds to a point on D1. Hence

OD1(D2) ' OP1(1). This shows that D2 is non-torsion in Cl(X). A similar

argument shows that D1 is non-torsion in Cl(X).

Let D′1 be the Cartier divisor

D′1 =

[(U00, 1), (U01, 1)

(U10, x−1), (U11, x

−1)

],

We can compute that D′1|D1 the divisor (V0, 1), (V1, 1) which is principal. In fact,

div x = D1 −D′1

246

Page 247: Geir Ellingsrud and John Christian Ottem October 15, 2018€¦ · between algebra and geometry: a ne varieties correspond to nitely generated k-algebras without zero-divisors, and

Chapter 14. Divisors and linear systems

So, D1 = D′1 in Cl(X). This shows that D1 and D2 are independent in Cl(X) –

D′1 restricts to 0 on D1 by O(1) on D2.

Now let U = X −D1 −D2. This is isomorphic to U11 which is the spectrum

of k[x−1, x−ry−1] which is a UFD. The exact sequence

ZD1 ⊕ ZD2 → Cl(X)→ 0

and the previous analysis shows that Cl(X) = ZD1 ⊕ ZD2.

Sheaf cohomology

We want to compute H i(X,OX(D)) for a divisor D = aD1 + bD2. As usual, we

utilize a Cech complex.

We want to mimick the computation for Pn. In that proof, the polynomial

ring k[x0, . . . , xn] played an important role – the groups in the Cech complex

corresponded to degree 0 localizations of it. For X there is no such Z-graded

ring lying around, but we can get by by introducing the bigraded ring

R = k[x0, x1, y0, y1]

where the degrees of the variables are defined by

deg x0 = (1, 0), deg x1 = (1, 0), deg x2 = (0, 1), deg x3 = (−r, 1).

By identifying x = x1x0

, y =xr0y1y0

, we find that the Cech complex of Uij can be

writtenR[x0x1] R[x0x1y0]

⊕ ⊕R[x0x1] R[x0x1y0]

⊕ → ⊕ → R[x0x1y0y1]

R[x0x1] R[x0x1y0]

⊕ ⊕R[x0x1] R[x0x1y0]

where the bracket means that we take the (0, 0)-part of the localization. So for

instance,

R[x0y0] = R

[x1

x0

,xr0y1

y0

]As in the Pn case, we now have a bigraded isomorphism⊕

a,b∈Z

H0(X,OX(aD1 + bD2)) '⋂i,j

R[xiyj ] = R

In particular, H0(X,OX(aD1 + bD2)) can be identified with polynomials of

247

Page 248: Geir Ellingsrud and John Christian Ottem October 15, 2018€¦ · between algebra and geometry: a ne varieties correspond to nitely generated k-algebras without zero-divisors, and

14.4. Examples

bidegree (a, b) in R. So for example H0(X,O(D1)) corresponds to degree (1, 0)-

polynomials, e.g., polynomials in x0, x1. These two sections corresponds to the

sections 1, x above. Similarly, H0(X,O(D2)) is 1-dimensional.

Perhaps the most interesting divisor is E = D2 − rD1,which is effective.

This has H0(X,OX(nE)) = k for every n ≥ 0, by E is not the trivial divisor.

Therefore E couldn’t possibly by globally generated. In fact, E ' P1 and

OX(E)|E ' OP1(−r)

Maps to projective space

Note that OX(D1) is globally generated by the sections x0, x1. These define a

map

f : X → P1

This is an example of a P1-fibration – the closed fibers here are all isomorphic to

P1. In fact, one can show that any such fibration over P1 is isomorphic to this

example for some choice of r.

248

Page 249: Geir Ellingsrud and John Christian Ottem October 15, 2018€¦ · between algebra and geometry: a ne varieties correspond to nitely generated k-algebras without zero-divisors, and

Chapter 15

Two applications of cohomology

15.1 Vector bundles on the projective line

As an instructive and not too complicated example of how cohomology groups

can give important results, we give a proof that all locally free sheaves on the

projective line P1k decompose as direct sums of invertible sheaves, and as all

invertible sheaves are of the form OP1k(n) for some integer n, we have the following

theorem:

Theorem 15.1. Assume that k is a field and that E is a locally free sheaf of

rank r on the projective line P1k. Then there is an isomorphism

E ' OP1k(α1)⊕ · · · ⊕ OP1

k(αr),

where the αi are integers, uniquely defined up to ordering.

We need some preperatory results about zeros of sections of locally free

sheaves om curves. Assume for a moment that C is a non-singular curve and

that E is locally free sheaf of finite rank on C. Let σ : OC → E be an injective

map of sheaves on C; such maps correspond bijectively to sections of E, that we

as well denote by σ (the correspondence is σ ↔ σ(1)). The dualized the map σ∨,

sits in an exact sequence

E∗ // OC// OZ(σ)

// 0 (15.1.1)

where Z(σ) is a closed subscheme of C. Generically σ is injective and therefore

σ∨ is generically surjective, and hence Z is not equal to the whole curve C.

249

Page 250: Geir Ellingsrud and John Christian Ottem October 15, 2018€¦ · between algebra and geometry: a ne varieties correspond to nitely generated k-algebras without zero-divisors, and

15.1. Vector bundles on the projective line

It follows that Z(σ) is finite C being a curve. The scheme Z(σ) is called the

zero-scheme of φ.

As C is non-singular, any non-zero coherent ideal I ⊆ OC is an invertible

sheaf. The ideal sheaf of Z(σ) – or the image of σ∨, if you want – is therefore an

invertible sheaf Lσ, and we may tensor the sequence (15.1.1) by Lσ−1 to obtain

the following factorization of σ∨

E∗⊗ Lσ−1 σZ // // OC // Lσ

−1.

For later references we formulate this as a proposition:

Proposition 15.2. Let C be a non-singular curve over a field k and E a locally

free sheaf of finite rank on C. Assume that σ is a non-zero section of E. Then

there is a finite subscheme Zσ in C and a surjective map E∗⊗ Lσ−1 σZ // // OC

where Ls denotes the invertible sheaf being the ideal of Z(σ), and a factorization

σ∨ = (ι⊗ Lσ−1) σZ where ι denotes the inclusion of the ideal Lσ in OC.

Proof of the theorem. The first step in the proof relies on two fundamental

facts about locally free sheaves of finite rank on projective spaces. Firstly,

sufficiently high positive twists have global sections; stated slightly differently

Γ(P1k, E(n)) 6= 0 if n 0 (this property locally free sheaves of finite rank share

with all coherent sheaves).

Secondly, sufficiently negative twists do not have global sections, i.e., it holds

true that Γ(P1k, E(−n)) = 0 for n >> 0. The third salient point is that any

invertible sheaf on the projective line is isomorphic to OP1k(α) for some integer α,

in other words, the theorem is true in rank one. The proof goes by induction on

the rank of E, and this will be the start of the induction.

We now let n0 be the greatest integer such that Γ(P1k, E(−n0)) 6= 0, and

after having replaced E by E(−n0), we may thus assume that Γ(P1k, E) 6= 0, but

Γ(P1k, E(−α)) = 0 for all α > 0. Let σ be a global section of E (it is called a

minimal section).

Lemma 15.3. The minimal section σ does not vanish anywhere.

Proof. Assume that Zσ 6= 0; then its sheaf of ideals equals OP1k(−α) for some

α > 0, and by (15.2) there is a surjective map E∗(α) → OP1k, whose dual is a

section of E(−α). This contradicts the fact that σ is a minimal section.

To prove the theorem we proceed as announced by induction on the rank r

of E, the r = 1 being taken care of by the fact that every invertible sheaf on

P1k is isomorphic to OP1

k(α) for some α ∈ Z. Choose a minimal section σ of E.

Then there is an exact sequence

0 // F // E∗ // OP1k

// 0 (15.1.2)

250

Page 251: Geir Ellingsrud and John Christian Ottem October 15, 2018€¦ · between algebra and geometry: a ne varieties correspond to nitely generated k-algebras without zero-divisors, and

Chapter 15. Two applications of cohomology

where the kernel F is locally free of rank r−1. By induction F '⊕

1≤i≤r−1 OP1k(αi).

Lemma 15.4. H1(P1k, F ) = 0

Proof. Taking the dual of (15.1.2) and twisting by O(−1) gives:

0 // OP1k(−1) // E(−1) // F ∗(−1) // 0

Then the long exact sequence of cohomology shows H0(F ∗(−1)) = 0. But then

also H0(F ∗(−2)) = 0 = H0(F ∗ΩP1), so H1(F ) = 0 by Serre duality. (Note that

we only need Serre duality for a direct sum of invertible sheaves of the form

OP1k(αi) in this proof).

Since H1(P1k, F ) = 0, the long exact sequence associated to 15.1.2 shows that

we have a surjection

Γ(P1k, E

∗) // Γ(P1k,OP1

k) // 0

and the section 1 can be lifted to a section τ of E∗. This is translates into the

diagram

0 // F // E∗ // OP1k

// 0

OP1k

τ

OO

=

==

It follows that the sequence (15.1.2) is split, and as a consequence we have

E∗ ' OP1k⊕ F,

that is E ' OP1k⊕⊕

1≤i≤r−1 OP1k(−αi) and the proof is complete.

15.2 Non-split vector bundles on Pn

A locally free sheaf is said to be split if it is isomorphic to a direct sum of

invertible sheaves. For instance, we showed before that any locally free sheaf

on on P1k is split. Let us give one example of a non-split locally free sheaf. We

consider again the sheaf E of rank n on Pnk sititing in the exact sequence

0→ E → On+1Pnk→ OPnk (1)→ 0.

Let us show that E is not split, i.e., E is not isomorphic to a direct sum of

invertible sheaves. Since Pic(Pnk) is generated by the class of O(1), this would

mean that E ' OPnk (a1)⊕ · · ·OPnk (an) for some integers a1, . . . an.

251

Page 252: Geir Ellingsrud and John Christian Ottem October 15, 2018€¦ · between algebra and geometry: a ne varieties correspond to nitely generated k-algebras without zero-divisors, and

15.2. Non-split vector bundles on Pn

Recall that for n ≥ 2, we have H1(Pnk ,O(m)) = 0 for any m ∈ Z. So if

we could show that H1(Pnk ,E ) 6= 0, we would have a contradiction. Actually,

H1(Pnk ,E ) = 0, but we can instead consider E (−1), which fits into

0→ E (−1)→ On+1Pnk

(−1)→ OPnk → 0.

Taking the long exact sequence in cohomology, we get

· · · → H0(O(−1)n+1)→ H0(OPnk )→ H1(E )→ H1(O(−1)n+1)→ · · ·

Here the two outer cohomology groups are zero, since O(−1) does not have any

global sections, and H1(Pnk ,OPnk (−1)) = 0. Hence, by exactness, we find that

H1(E (−1)) ' H0(OPnk ) = k. This implies that E (−1), and hence E cannot be a

sum of invertible sheaves, and we are done.

However, coming up with examples of non-split vector bundles of low rank

is a notoriously difficult problem, even for projective space. In fact, a famous

conjecture of Hartshorne says that any rank 2 vector bundle on Pn for n ≥ 5 is

split. (On P4 this statement does not hold, as shown by the so-called Horrocks–

Mumford bundle). This is related to the conjecture that any smooth codimension

2 subvariety of Pn is a complete intersection.

252

Page 253: Geir Ellingsrud and John Christian Ottem October 15, 2018€¦ · between algebra and geometry: a ne varieties correspond to nitely generated k-algebras without zero-divisors, and

Chapter 16

Maps to projective space

Given a scheme X it is natural to ask when there is a morphism to a projective

space

f : X → Pn,

or when there is a closed immersion X → Pn. Given such a morphism, we get

geometric information about X using this map, e.g., using the fibers f−1(y) or

pulling back sheaves from Pn.

The corresponding question for An has already been answered. Morphisms

X → An are in one-to-one correspondence with elements of Γ(X,OX)n. In fancy

terms, An represents the functor

AffSch → Sets

X 7→ Γ(X,OX)

So which functor does projective space represent? Intuitively we would like

to associate to each morphism f : X → Pn to a set of data on X. As we have

seen, there is not so much information Γ(X,OX) in the context of projective

schemes. However, we do have something canonical associated to Pn, namely

sections of OPn(1). Taking the usual global sections x0, . . . , xn on OPn(1), we

can pull back via the map f ∗ and get n+ 1 sections si = f ∗xi of the invertible

sheaf f ∗O(1). Note that there is no point of Pn where the xi simultaneously

vanish. More precisely, for every y ∈ Pn, the stalk OPn(1)y is generated by one

of the xi (as an OPn-module). So my the properties of the pullback, we see that

the same statement holds for the si and X.

In this chapter we will se that there is a way to reverse this process, i.e., that

from a given invertible sheaf L and n+ 1 global sections si ∈ Γ(X,L) with the

253

Page 254: Geir Ellingsrud and John Christian Ottem October 15, 2018€¦ · between algebra and geometry: a ne varieties correspond to nitely generated k-algebras without zero-divisors, and

16.1. Globally generated sheaves

above property, we can uniquely reconstruct a morphism f : X → Pn so that

f ∗OPn(1) = L and f ∗xi = si.

16.1 Globally generated sheaves

Definition 16.1. Let X be a scheme and let F be an OX-module. We say F is

globally generated (or generated by global sections) if there is a family of sections

si ∈ F(x), i ∈ I, such that the images of sx generate Fx as an OX,x-module.

Equivalently, F is globally generated if there is a surjection

OIX → F → 0

for some index set I.

Let us consider a few examples:

Example 16.2. On an affine scheme any quasi-coherent sheaf is globally gener-

ated. Indeed, if X = SpecA, F = M , for some A-module M , then picking any

presentation AI →M → 0 for M shows that F is globally generated.

Example 16.3. X = ProjR. Then F = O(1) is globally generated if R

is generated in degree 1. Indeed, if this is the case, we have a surjection

R ⊗ R1 → R(1). Taking ∼, we get OX ⊗ R1 → OX(1) → 0, which shows that

O(1) is globally generated.

Example 16.4. The tangent bundle of Pn is globally generated. To see why, we

recall the Euler sequence on Pn:

0→ O → O(1)n+1 → TPn → 0

Here the leftmost map is given by multiplication by the vector (x0, . . . , xn), and

the rightmost sends xi to ∂∂xi

. So in fact, even TPn(−1) is globally generated.

16.1.1 Pullbacks and pushforwards

Let us recall a few things about pulling back sheaves and sections. If f : X → Y

is a morphism of schemes, and G is an OY -module, we define f ∗G as the tensor

product f−1G ⊗f−1OY OX . The stalk of f ∗G at a point x ∈ X is isomorphic to

Gf(x)⊗OY,f(x)OX,x .

If G is an invertible sheaf, the pullback f ∗G is also invertible. Indeed, if

U ⊆ X and V =⊆ Y are open affine subsets such that f(U) ⊆ V , and L|U ' OU ,

then f ∗L|V ' OV (recall that a morphism pulls back the structure sheaf of the

target gives to the structure sheaf of the source).

254

Page 255: Geir Ellingsrud and John Christian Ottem October 15, 2018€¦ · between algebra and geometry: a ne varieties correspond to nitely generated k-algebras without zero-divisors, and

Chapter 16. Maps to projective space

We can pull back sections s ∈ Γ(Y,G) as well. Let us consider the affine

case first, where U = SpecB and V = SpecA. In this case, G = M for some

A-module, and f ∗G = M ⊗A B. So finally, if s ∈ Γ(U,G) = M , then the pullback

is given by f ∗(s) = s⊗ 1 ∈M ⊗A B.

Proposition 16.5. Let f : X → Y be a morphism of schemes. If F is an

OY -module which is globally generated by sections sii∈I , then the pullbacks

ti = f ∗(si) generate f ∗F .

Proof. The si determine a surjection OIY → F → 0, and the pullback OI

X →f ∗F → 0 (which is surjective since f ∗ is right-exact) is the map determined by

the ti. Hence f ∗F is globally generated by the ti.

Remark 16.6. For the pushforward, f∗F is typically not globally generated even

when F is the structure sheaf. For example, if f : P1 → P1 is the map induced

by k[u2, v2] ⊆ k[u, v], then f∗OP1 ' OP1 ⊕ OP1(−1). The latter is not globally

generated: If it were, we would in particular get a surjection OI → O(−1)→ 0,

however O(−1) is clearly not globally generated, since it has no global sections

at all.

16.2 Morphisms to projective space

16.2.1 Example: The twisted cubic

Consider the ring R = k[u3, u2v, uv2, v3] with the grading such that the mono-

mials have degree 1. We have seen that ProjR = P1k (since R is the Veronese

subring k[u, v](3) of k[u, v]). R is isomorphic to the ring

k[x0, x1, x2, x3]/(x21 − x0x2, x0x3 − x1x2, x

22 − x1x3).

This shows that ProjR embeds as a cubic curve in P3k.

Intuitively, we would like to say that this morphism f : X → P3 is given by

something like

[u, v] 7→ [u3, u2v, uv2, v3].

However, here u, v are not regular functions (they however define rational func-

tions), so to define an actual morphism of schemes, we need to be a little bit more

rigorous. Note, however, that on D+(u) = SpecR(u) = Spec k[x1x0, x2x0, x3x0

], the

ratio t = vu

is a regular section Γ(D+(u),O). Moreover, the ring homomorphism

φ : k[x1x0, x2x0, x3x0

] → k[t]x1x0

7→ tx2x0

7→ t2

x3x0

7→ t2

255

Page 256: Geir Ellingsrud and John Christian Ottem October 15, 2018€¦ · between algebra and geometry: a ne varieties correspond to nitely generated k-algebras without zero-divisors, and

16.2. Morphisms to projective space

gives a morphism of affine schemes

D+(u)→ D+(x0) ⊆ P3k

Note that this corresponds to the usual affine twisted cubic A1 → A3, t 7→(t, t2, t3). Similarly, on D+(v), the ratio s = u

vdefines a morphism D+(x0) →

D+(x3). On the overlaps, these maps are compatible with the standard gluing

construction of P1k, and so we get finally a morphism P1

k → P3k.

16.2.2 Morphisms and globally generated invertible sheaves

The morale of the previous example is the morphism is not specified using a set

of regular functions, but rather, sections of an invertible sheaf. Using the same

type of gluing argument, we will prove the following general result:

Theorem 16.7. Let X be a scheme over a ring A.

(i) If f : X → PnA is a morphism over A, then f ∗(O(1)) is an invertible sheaf

which is globally generated by the sections si = f ∗(xi) for i = 0, . . . , n.

(ii) If L is an invertible sheaf on X which is globally generated by n+ 1 sections

s0, . . . , sn, then there exist a unique morphism f : X → PnA such that

L ' f ∗(O(1)) and si = f ∗(xi) under this isomorphism.

Proof. (i) This is a consequence of Proposition 16.5.

(ii) Given n+ 1 sections s0, . . . , sn of an invertible sheaf L, we will construct

a morphism to PnA, such that si is the pullback of xi. Let

Xi = Xsi = x ∈ X|si(x) 6= 0.

(Recall that s(x) is the residue class of s in Lx/mxLx.) As we have seen, this is

the complement of the support of the section si, and is therefore an open set.

By the assumption that these sections generate L, we see that X is covered by

the sets Xi.

Now, we have given local isomorphisms φi : L|Ui → OX |Ui , which we can use

to view the fractionsjsi

can be regarded as a regular section on Xi, that is, an

element of Γ(Xi,OX).

Write R = A[x0, . . . , xn] and consider the ring homomorphism given by

R(xi) → Γ(Xi,OXi)xjxi7→ sj

si

By the correspondence between ring homomorphisms and maps into affine

schemes, we obtain a morphism of schemes fi : Xi → D+(xi). It is not so hard

256

Page 257: Geir Ellingsrud and John Christian Ottem October 15, 2018€¦ · between algebra and geometry: a ne varieties correspond to nitely generated k-algebras without zero-divisors, and

Chapter 16. Maps to projective space

to see that on the overlaps Xi ∩Xj, two maps are given by the same ratios of

sections, and so they glue to a morphism f : X → Pn. By construction of the

map f , we have f ∗xi = si.

Abusing notation, we will refer to a morphism φ : X → PnA as ‘given by the

data (L, s0, . . . , sn) and write

X → PnAx 7→ [s0(x) : · · · : sn(x)]

One should still keep in mind that the sections si are sections of L, rather than

regular functions.

Given a scheme X with s0, . . . , sn of a line bundle L, there is a maximal open

subset U such that the sections generate L for all points in U . We then get a

rational map φ : X 99K PnA, that is φ is a morphism on the smaller open set U .

Example 16.8. Let X = P1k = Proj k[s, t] and L = O(2). Then L is globally

generated by s2, st, t2 and the corresponding morphism

X → P2k

[s : t] 7→ [s2 : st : t2]

has image V (x0x2 − x21) which is a smooth conic.

Example 16.9 (Cuspidal cubic). Let X = A1k and L = OX . Then, Γ(X,L) =

k[t] is infinite dimensional over k. Choosing the three sections 1, t2, t3, we get a

map of schemes

X → P2k

t 7→ [1 : t2 : t3]

whose image in P2 is the cuspidal cubic minus the point at ∞.

Example 16.10 (P1 as a quotient space). Let X = A2, and L = OX . Then,

Γ(X,L) = k[x, y]. If we take the sections x, y, then they generate L outside

V (x, y). Hence we get a morphism of schemes

A2k − V (x, y)→ P1

k

(x, y) 7→ [x : y]

which we is the quotient map used to define P1.

Example 16.11 (Projection from a point). Consider the projective space X =

PnA and sections x1, . . . , xn of O(1), then these sections generate O(1) outside

257

Page 258: Geir Ellingsrud and John Christian Ottem October 15, 2018€¦ · between algebra and geometry: a ne varieties correspond to nitely generated k-algebras without zero-divisors, and

16.3. Application: Automorphisms of Pn

the point p corresponding to I = (x1, . . . , xn) (that is, what we would classically

write as the point p = (1 : 0 : · · · : 0)). The induced morphism PnA−V (I)→ Pn−1A

is the projection from p.

Example 16.12 (Cremona transformation). Consider the projective space

X = P2A and sections x0, x1, x2 of O(1), then the sections x0x1, x0x2, x1x2

generate O(2) outside V (x0x1, x0x2, x1x2) corresponding to the three points

(1, 0, 0), (0, 1, 0), (0, 0, 1). The induced rational map P2A 99K P2

A is the Cremona

transformation.

Example 16.13 (The Veronese surface). Consider X = P2, and L = OP2(2). If

x0, x1, x2 are projective coordinates on X, then the quadratic monomials

x20, x

21, x

22, x0x1, x0x2, x1x2

form a basis for H0(X,L), and generate L at every point. The corresponding

map X → P5 is an embedding; the image is the Veronese surface. It is a classical

fact that the image is defined by the 2× 2 minors of the matrixu0 u1 u2

u1 u3 u4

u2 u4 u5

Example 16.14. Let us consider again the case Q = P1 × P1. Keeping the

notation from Section 14.4.1, we have two divisors, L1 = [0 : 1] × P1, L2 =

P1 × [0 : 1]. Note that eacj Li is globally generated (being the pullback of a base

point free divisor on P1). The corresponding map is of course the i-th projection

map pi : Q→ P1.

If x0, x1 is a basis for Γ(X,L1), and y0, y1 is a basis for Γ(X,L2), we find that

Γ(X,L1 + L2) is spanned by the sections

s0 = x0y0, s1 = x0y1, s2 = x1y0, s3 = x1y1

Moreover, these sections generate OQ(D) everywhere, and so we get a map

Q→ P3

This is of course nothing but the Segre embedding; note the quadratic relation

between the four sections s0s3 − s1s2 = 0.

16.3 Application: Automorphisms of Pn

If k is a field, then any invertible (n+1)×(n+1) matrix m gives an automorphism

Pnk → Pnk (since it induces an automorphism of k[x0, . . . , xn]). Moreover, two

258

Page 259: Geir Ellingsrud and John Christian Ottem October 15, 2018€¦ · between algebra and geometry: a ne varieties correspond to nitely generated k-algebras without zero-divisors, and

Chapter 16. Maps to projective space

matrices m and m′ determine the same automorphism if and only if m = λm′

for some non-zero scalar λ ∈ k∗ (. So we are led to consider the projective linear

group

PGLn(k) = GLn(k)/k∗

Theorem 16.15. Autk(Pn) = PGLn(k).

Proof. The above shows that there is a containment ⊇. To show the reverse

inclusion, let φ : Pnk → Pnk be any automorphism. Then we get an induced map

φ∗ : Pic(Pn)→ Pic(Pn)

which is an isomorphism. Since Pic(Pn) = Z, we must have either φ∗OPn(1) =

OPn(1) or φ∗OPn(1) = O(−1). The latter case is impossible, since φ∗(OPn(1))

has a lot of global sections, whereas OPn(−1) has none. So φ∗(OPn(1)) = OPn(1).

In particular, φ∗ gives a map

Γ(Pn,OPn(1))→ Γ(Pn,OPn(1)),

which is a isomorphism of k-vector spaces. However, these vector spaces can be

identified with 〈x0, . . . , xn〉, and so φ∗ gives rise to an invertible (n+ 1)× (n+ 1)-

matrix m. By construction m induces the map φ, and so φ comes from an

element of PGLn(k).

16.4 Projective embeddings

We have seen how morphisms X → Pn corresponds to invertible sheaves L plus

n+ 1 global sections s0, . . . , sn that generate it. Given this it is natural to ask

which data (L, si) that corresponds to special types of morphisms, e.g., closed

immersions. The following criterion tells us how to check this locally:

Proposition 16.16. Let φ : X → PnA be a morphism corresponding to the data

(L, s0, . . . , sn). Then φ is a closed immersion if and only if for each i = 0, . . . , n,

Xi = Xsi is affine, and the homomorphism A[y0, . . . , yn]→ Γ(Xi,OX) yj 7→ sjsi

is surjective.

Proof. ⇒: If φ is a closed immersion, we may view X as a closed subscheme

of PnA (so it is given by some ideal in A[y0, . . . , yn]). Note that in this case

Xi = X ∩ D+(xi) is a closed subscheme of the affine scheme D+(xi). By our

results on closed subschemes of affine schemes, Xi corresponds to an ideal in

A[y0, . . . , yn](x0), and the map A[y0, . . . , yn]→ Γ(Xi,OX) yj 7→ sjsi

is surjective.

⇐: Since Xi is affine, and the map to the right is surjective, we see that Xi

corresponds to a closed subscheme of D+(xi). As X is glued together by the Xi

we see that X is a closed subscheme of PnA as well.

259

Page 260: Geir Ellingsrud and John Christian Ottem October 15, 2018€¦ · between algebra and geometry: a ne varieties correspond to nitely generated k-algebras without zero-divisors, and

16.4. Projective embeddings

In practice however, this condition is not so hard to verify (that is, not much

easier than checking directly that the induced map is an embedding). To get a

better criterion we first introduce some more notation.

Definition 16.17. Let X be a scheme over a field k.

1. A linear series V is a subspace V ⊆ Γ(X,L) where L is an invertible sheaf

on X.

2. A linear series V separates points if for any two points x, y ∈ X there is a

section s ∈ V such that s(x) = 0 but s(y) 6= 0

3. A linear series V separates tangent vectors if for any x ∈ X the set

s ∈ V |s(x) 6= 0 spans mxLx/m2x.

The definitions here come from the following theorem

Theorem 16.18. Let X be a projective scheme over an algebraically closed field

k and let φ : X → Pnk be a morphism over k corresponding to (L, s0, . . . , sn).

Then φ is a closed immersion if and only if V = spans0, . . . , sn separates points

and tangent vectors.

We will not prove this theorem here, but let us at least say a few words about

it.

If φ is a closed immersion, we can consider X ⊆ Pnk as closed subscheme, and

the si are simply restrictions of th xi to X. Now the theorem is essentially a

consequence of the easy geometric fact that the x0, . . . , xn separate points and

tangent vectors on Pn. Indeed, we can pick a linear form l = a0x0 + · · ·+ anxnsuch that the hyperplane H = V (l) meets x but not not y (here we are using

that k is algebraically closed!). l|X is a linear combination of the si, and so V

separates points. Similarly, linear forms on Pn separate tangent vectors, so they

also separate tangent vectors on the subscheme X.

The hard part is showing that the converse holds. If V is a linear series

separating points and tangent vectors, then the morphism φ is injective (since

you can use linear forms for distinguish the images of two points). The main

ingredient we need is that φ is proper, which implies that φ is closed, and

furthermore that φ gives a homeomorphism onto a closed set in Pn. Given this,

the rest of the proof is relatively straightforward: We need only check that

OPn → φ∗OX is surjective, which can be done on stalks.

What makes this theorem more powerful than the previous is that the

conditions on sections to separate points and tangent vectors can be turned in

to statements about restriction maps of certain sheaves being surjective. The

latter task is something we can approach using the machinery of cohomology.

And indeed, using the long exact sequence in cohomology we can turn this into

simple, computable numerical criteria that guarantee that the corresponding

map is an embedding.

260

Page 261: Geir Ellingsrud and John Christian Ottem October 15, 2018€¦ · between algebra and geometry: a ne varieties correspond to nitely generated k-algebras without zero-divisors, and

Chapter 16. Maps to projective space

16.5 Projective space as a functor

Recall that we defined a functor F : Schop → Sets to be representable if there is

a scheme X and an isomorphism of functors Φ : hX ' F ; i.e., for each S ∈ Schop

a bijection Φ(S) : Hom(S,X)→ F (S).

We saw in Chapter 6 that affine space An = SpecZ[x1, . . . , xn] represented

the functor F taking a scheme S to Γ(X,OX)n; this just amounts to saying

that a morphism X → An is determined by n regular functions. More generally

Corollary 3.7 says that an affine scheme SpecA is represented by the functor

F (S) = HomRings(A,Γ(X,OX)).

So which functor does projective space represent? Intuitively we would like

to associate to each morphism f : X → Pn to a set of data on X. As we have

seen, there is not so much information Γ(X,OX) in the context of projective

schemes. However, we do have something canonical associated to Pn, namely

sections of O(1).

As a first question, we should ask what F (Spec k) should correspond to.

Motivated by the very definition of projective space over a field, we would like

to say that the value of F on closed points should correspond to lines in a vector

space. More precisely, morphisms Spec k → Pn should be in correspondence with

lines l ⊆ kn+1 through the origin. This assignment should also be continuous,

e.g., for each open U ⊆ Pn, we should have a sub-line bundle LU of the trivial

bundle Pnk ×An+1k . By the correspondence between vector bundles and invertible

sheaves, this is equivalent to a subsheaf L ⊆ On+1U which is an invertible sheaf.

This motivates the following definition:

Definition 16.19. Define a functor F : Schop → Sets by defining for a scheme

X

F (X) =

invertible subsheaves L ⊆ On+1X

=

invertible sheaf quotients On+1X → L→ 0

/ ∼

where the ∼ says that quotients On+1X → L→ 0, On+1

X →M → 0 are equivalent

if have the same kernel.

Note that on projective space Pn = ProjZ[x0, . . . , xn], we do have such a

quotient

On+1Pn → OPn(1)→ 0.

Indeed, on a D+(f) we can define it by sending an element (t0, . . . , tm) ∈ R(f)

to x0t0 + · · · + xntn ∈ R(1)(f). Of course this quotient is not canonical, as it

depends on the choice x0, . . . , xn.

261

Page 262: Geir Ellingsrud and John Christian Ottem October 15, 2018€¦ · between algebra and geometry: a ne varieties correspond to nitely generated k-algebras without zero-divisors, and

16.5. Projective space as a functor

Theorem 16.20. The functor F is represented by the scheme Pn.

Proof. We need to define natural transformations of the two functors hPn to F .

For a scheme X, we define

Φ(X) : Hom(X,Pn)→ F (X)

by sending a morphism f : X → Pn to the (equivalence class of the) quotient

On+1X → f ∗O(1)→ 0

which is the pullback of On+1 → O(1)→ 0 on Pn. It is clear that this assignment

is functorial, i.e., it gives a natural transformation. Moreover, by definition Φ

sends the identity map idPn to the equivalence class of the sequence above.

To prove the theorem, we need to construct an inverse Ψ to Φ. In other

words, to each quotient On+1X → L → 0, we need to produce a morphism of

schemes f : X → Pn so that On+1Pn → OPn(1)→ 0 pulls back to On+1

X → L→ 0.

However, from the quotient On+1X → L → we obtain n + 1 sections s0, . . . , sn

by taking the n + 1 maps OX → On+1X → L (where the first map corresponds

to a ‘coordinate inclusion’). We then define Ψ by associating On+1X → L→ to

the morphism f : X → Pn ∈ Hom(S,Pn). One can check that two choices of

quotients induce the same map f : X → Pn, so this is well-defined.

Finally, in the construction of the morphisms in Theorem 16.7, we have

L = f ∗O(1) and si = f ∗xi, and so Ψ is the inverse to Φ.

Why should one care about such a statement? There are several good reasons,

but perhaps the most basic is that it tells us how to think about points of the

scheme projective space as corresponding to something geometric. Indeed, a

priori, the only thing we know about PnZ is that it is glued together by n + 1

coordinate affine n-spaces. The theorem above shows that in fact does what it is

supposed to do; the k-points Pn(k) correspond to lines in the vector space kn+1

for any field k.

The case of P1 is particularly vivid. The functor of points of A1 shows that a

morphism X → A1 is equivalent to giving an element of Γ(X,OX). In less precise

terms this is what we think of as a ‘regular function’ on X. The corresponding

statement for P1 is the following. A map X → P1 corresponds to a section s of

a line bundle L on X, or in other words a ‘meromorphic function’, which is only

required to be regular on Xs. .

16.5.1 Generalizations

If we modify of the functor of points of Pn, we obtain other interesting examples

of schemes parameterizing geometric objects. The following examples are only

262

Page 263: Geir Ellingsrud and John Christian Ottem October 15, 2018€¦ · between algebra and geometry: a ne varieties correspond to nitely generated k-algebras without zero-divisors, and

Chapter 16. Maps to projective space

meant as basic illustrations of this point - they will not appear later in the notes.

Example 16.21 (The Grassmannian). Consider the functor

F (S) = rank r locally free quotients On → Q→ 0 / ∼

where two quotients again are defined to be equivalent if they have the same

kernel. Then F is represented by a scheme, known as the Grassmannian Gr(r, n).

This is scheme can realized as the projective scheme in P(n+rr )−1

Z = ProjZ[pi1,...,ir ]

where 0 ≤ i1 < i2 < . . . ir ≤ n. Explicitly, Gr(r, n) is given by the Plucker

equationsr+1∑k=0

(−1)kpi1,i2,...,ir−1,jk · pi1,...,jk,...,jr+1= 0

(over all sequences of indices i1, . . . , ir and j1, . . . , jr+1). We can define the map

Gr(k, n)→ P(n+rr )−1

Z using the Yoneda lemma, by giving a natural transformation

between the functors. If S is a scheme, this transformation takes an S-valued

point of Gr(k, n), that is, a quotient On → Q→ 0, to ∧rOn → ∧rQ→ 0, which

since ∧rQ has rank 1, defines a point in P(n+rr )−1

Z (S).

Proving that this scheme actually represents Gr(k, n) is similar to what we

did for projective n-space, although it is slightly more involved, due to the

Plucker equations.

Example 16.22 (Projective bundles). Let E be a locally free sheaf on a scheme

X. Consider the functor on Sch/X given by

F (h : Y → X) = invertible sheaf quotients h∗E → L→ 0

Then F is represented by a scheme P(E )→ X over X called the projectivization

of E . One can think of closed points of this scheme as hyperplanes in the fibers

of E .

Example 16.23 (Proj of an OX-module). Let X be a scheme and let F be a

quasi-coherent sheaf of OX-modules. Consider the functor on Sch/X given by

F (h : Y → X) = invertible sheaf quotients h∗F → L→ 0

Then F is represented by a scheme Proj(F). Many geometric constructions can

be formulated as such schemes. For instance, if I is a sheaf of ideals on X, then

Proj(F) can be identified with the blow-up of X along I (see Hartshorne II.7).

263

Page 264: Geir Ellingsrud and John Christian Ottem October 15, 2018€¦ · between algebra and geometry: a ne varieties correspond to nitely generated k-algebras without zero-divisors, and
Page 265: Geir Ellingsrud and John Christian Ottem October 15, 2018€¦ · between algebra and geometry: a ne varieties correspond to nitely generated k-algebras without zero-divisors, and

Chapter 17

More on projective schemes

17.1 Ample invertible sheaves and Serre’s the-

orems

The prototype of an invertible sheaf is the sheaf O(1) on X = ProjR, which is

globally generated if R is generated in degree 1. As we have seen, this corresponds

to a morphism f : X → Pn, with the property that f ∗OPnZ (1) = OProjR(1). In

the case where f is a closed immersion, we can regard X as a subscheme of

Pn, and O(1) is simply the restriction OPnZ (1)|X . This restricted invertible sheaf

has a special property - R and hence X is completely recovered by the global

sections of it and the polynomial relations between them. This motivates the

following definition:

Definition 17.1. Let X be a scheme over S. A invertible sheaf L on X is said

to be very ample over S if there is a closed immersion i : X → PnS such that

L ' i∗OPnS(1). L is ample if L⊗N is very ample for some N > 0.

Theorem 17.2 (Serre). Let X ⊆ PnA be a projective scheme over a noetherian

ring A and let L be an ample invertible sheaf. Let F be a coherent sheaf on X.

Then there is an integer m0 such that F ⊗ Lm is globally generated (by a finite

set of global sections) for all m ≥ m0.

Proof. By replacing L with a multiple, we may assume that L is very ample

and even that L = i∗O(1) for a projective embedding i : X → Pn. Since X is

noetherian, i is finite and F is coherent, we have that i∗F is coherent on Pn.

Moreover, (i∗F)(m) = i∗(F ⊗ Lm), so F ⊗ Lm is globally generated if and only

if (i∗F)(m) is. Hence we may assume that X = PnA.

265

Page 266: Geir Ellingsrud and John Christian Ottem October 15, 2018€¦ · between algebra and geometry: a ne varieties correspond to nitely generated k-algebras without zero-divisors, and

17.1. Ample invertible sheaves and Serre’s theorems

Write PnA = ProjR where R = ProjA[x0, . . . , xn]. We can cover X by open

sets D+(xi) where i = 0, . . . , n, such that F|D(xi) = Mi some finitely generated

R(xi)-module. For each i we can choose finitely many elements sij generating Mi.

Regarding sij as sections of the sheaf Mi, we see that there are integers mij such

that xmiji sij extend to global sections of F(mij). Take m = maxijmij, then for

each i, j we have a section tij ∈ Γ(X,F(m)).

We claim that the global sections tij generate F(m) locally. It is sufficient to

prove this on the opens D+(xi) over which F(m) is isomorphic to the R(xi)-module

Mi(m) 'Mi ⊗R R(m), which is isomorphic to xmi Mi as a graded R(xi)-module.

Since sij generates Mi, we see that xmi sij generate F(m)|D+(xi).

This shows that F(m0) is globally generated for the chosen m0 = m above.

Then also F(m) is globally generated for m ≥ m0, by multiplying the generating

sections of F(m0) by the various monomials in the xi.

In particular, we see that any coherent sheaf can be written as a quotient of

a direct sum of invertible sheaves of the form O(−m): Take any F(m) which is

globally generated; then there is a surjection OI → F(m)→ 0 (with I finite!),

which becomes

O(−m)I → F → 0 (17.1.1)

after tensoring by O(−n).

Theorem 17.3 (Serre). Let X = ProjR be a noetherian projective scheme over

SpecA, where A is a noetherian ring, and let F be a coherence sheaf on X.

Then:

(i) The cohomology groups

H i(X,F)

are finitely generated A-modules for every i ≥ 0.

(ii) There exists an n0 > 0 such that

H i(X,F (n)) = 0.

for all n ≥ n0.

Proof. As in the above proof, we immediately reduce to the case where X = PnA.

Note that both of the conclusions hold for the twisting sheaves OX(n). To

prove it for any coherent F , take a quotient of the form (17.1.1) and let K be

the kernel, so that we have an exact sequence

0→ K → E → F → 0

where E = O(−n)I , with I finite. Note that K is again coherent.

266

Page 267: Geir Ellingsrud and John Christian Ottem October 15, 2018€¦ · between algebra and geometry: a ne varieties correspond to nitely generated k-algebras without zero-divisors, and

Chapter 17. More on projective schemes

(i): Take the l.e.s of cohomology to get

H i(X,E )→ H i(X,F)→ H i+1(X,K )

We can now prove the theorem by downwards induction on i: H i+1(X,K )

and H i(X,E ) are both finitely generated, and hence so is H i(X,F), since A is

noetherian.

(ii): Twist the above sequence by OX(m) and take the long exact sequence

in cohomology to get

H i(X,E (m))→ H i(X,F(m))→ H i+1(X,K (m))

By downward induction on i, and the fact that H i(X,E (m)) for any m, we find

that H i(X,F(m)) = 0.

17.1.1 Euler characteristic

Let X be a noetherian scheme over a field k and let F be a coherent sheaf on

X. By Serre’s theorem, the cohomology groups H i(X,F) are finite dimensional

k-vector spaces, so it makes sense to ask about their dimensions. It turns out

that the alternating sum of these dimensions is the right thing to look at:

Definition 17.4. Let X be a projective scheme of dimension n. We define the

Euler characteristic of F as

χ(F) =n∑k=0

(−1)k dimHk(X,F)

Proposition 17.5. Let X be a projective scheme and let O(1) be the very ample

invertible sheaf. Then the function

PF(m) = χ(F(m))

is a polynomial in m

This polynomial is called the Hilbert polynomial of F . When F = M , this

coincides with the Hilbert polynomial of M as defined in commutative algebra.

17.2 Proper and projective schemes

Let S be a scheme and let X be a scheme over S. We say call X is projective

over S (or that the structure morphism f : X → S is projective) if f : X → S

267

Page 268: Geir Ellingsrud and John Christian Ottem October 15, 2018€¦ · between algebra and geometry: a ne varieties correspond to nitely generated k-algebras without zero-divisors, and

17.2. Proper and projective schemes

factors as f = π i where i : X → PnS is a closed immersion and π : PnS → S is

the projection.

The primary examples are of course X = ProjR where R is a graded R0-

algebra generated in degree 1 and S = SpecR0. In this case, we can define the

projective immersion i by taking a surjection R0[x0, . . . , xn]→ R, which upon

taking Proj, a closed immersion X → PnR0.

Theorem 17.6. A projective morphism X → S is proper.

268

Page 269: Geir Ellingsrud and John Christian Ottem October 15, 2018€¦ · between algebra and geometry: a ne varieties correspond to nitely generated k-algebras without zero-divisors, and

Chapter 18

The Riemann–Roch theorem and

Serre duality

18.1 Curves

We will in this chapter assume that k is an algebraically closed field. Recall that

a variety over k is a separated, integral scheme of finite type over k. Throughout

this section, a curve will mean a 1-dimensional variety over k.

18.2 Divisors on Curves

Let X be a non-singular curve. Since X has dimension 1, a Weil divisor D on X

is a finite formal combination of points on X,

D =∑

nipi

where pi ∈ X. We say that D is effective if each ni ≥ 0.

The degree of D is defined as the sum degD =∑ni. Equivalently, if we view

D as a Cartier divisor given by the data (Ui, fi), then

degD =∑x∈X

vx(D)

where vx(D) = vx(fi) as before.

Lemma 18.1. Let D be a non-zero effective divisor on X. Then degD =

dimkH0(D,OD).

269

Page 270: Geir Ellingsrud and John Christian Ottem October 15, 2018€¦ · between algebra and geometry: a ne varieties correspond to nitely generated k-algebras without zero-divisors, and

18.2. Divisors on Curves

Proof. Note that D is a finite scheme, hence finite. Let A = H0(D,OD). Then

A =⊕x∈D

OD(x) =⊕x∈X

OD,x

Hence dimk A =∑

x∈D dim OD,x. Now, vx(D) equals the length of OD,x. Since

k is algebraically closed, this length coincides with dimk OD,x, which gives the

claim.

Recall that each Weil divisor determines an invertible sheaf OX(D), which

over an open set U takes the value

OX(D)(U) = f ∈ K|(div f +D)|U ≥ 0

Then D is effective if and only if Γ(OX(D)) 6= 0. In particular, if Γ(X,OX(D))

has dimension at least 2, there is a second effective divisor D′ =∑miqi such

that D and D′ are linearly equivalent.

18.2.1 The genus of a curve

Definition 18.2. The arithmetic genus of X is the defined as

pa(X) = dimkH1(X,OX)

The geometric genus of X is defined as

pg(X) = dimkH0(X,ΩX)

Note that χ(OX) = dimH0(X,OX) − dimH1(X,OX), so we get pa(X) =

1− χ(OX).

At first sight, these numbers are defined using different sheaves, and there

is no a priori reason to expect that they should have anything to do with each

other. However, we shall see later in the chapter that there is a strong relation

between them: pa = pg. For the time being we will still refer to pa as the genus

of X.

Example 18.3. When X = P1, we have H1(P1,O) = 0 and H0(P1,ΩP1) =

H0(P1,OP1(−2)) = 0, so both genera are zero.

Example 18.4. Let X ⊆ P2 be a plane curve, defined by a homogeneous poly-

nomial f(x0,1 , x2) of degree d. In Chapter 12, we computed that H1(X,OX) 'k(d−1

2 ). Hence the arithmetic genus of X is (d−1)(d−2)2

.

270

Page 271: Geir Ellingsrud and John Christian Ottem October 15, 2018€¦ · between algebra and geometry: a ne varieties correspond to nitely generated k-algebras without zero-divisors, and

Chapter 18. The Riemann–Roch theorem and Serre duality

18.3 The canonical divisor

Let X be a non-singular curve. Then the sheaf of differentials ΩX is a locally

free sheaf of rank 1, i.e., an invertible sheaf. Thus ΩX gives rise to a divisor,

which we denote by KX . We can describe KX as a Cartier divisor as follows.

Over an open subset U ⊆ X a local section of ΩX is given by fUdx. We can

then define a Cartier divisor with the data

(U, fU)

This is well-defined, because on the overlaps U ∩ V , two sections fUdx, fV dy are

related by fU = dydxfV , and dy

dxis a unit in OX(U ∩ V ).

Any other section ω′ of ΩX has the form fω for some f ∈ K×. Therefore,

div (ω′) = div (ω) + div (f)

so the divisor class [KX ] := [div (ω)] ∈ Cl(X) associated to ω is independent of

the chosen ω. We call this the canonical class of X. Also, any divisor D so that

[D] = [KX ] is called a canonical divisor and it is denoted by KX .

Definition 18.5. A canonical divisor is a divisor KX , so that OX(KX) ' ΩX .

Example 18.6. Let X = P1k, and consider the open set U0 = D+(x0). On U0 we

have a local coordinate t = x1x0

, and we consider the differential form dt = d(x1x0

).

On the overlap U0 ∩ U1, we have

d

(x0

x1

)= d(t−1) = −t−2dt = −

(x0

x1

)2

d

(x1

x0

)It follows that ΩX ' OP1(−2).

18.4 Hyperelliptic curves

Let us recall the hyperelliptic curves defined in Chapter 3. For an integer g ≥ 1

we consider the scheme X glued together by the affine schemes U = SpecA and

V = SpecB, where

A =k[x, y]

(−y2 + a2g+1x2g+1 + · · ·+ a1x)and B =

k[u, v]

(−v2 + a2g+1u+ · · ·+ a1u2g+1)

As before, we glue D(x) to D(u) using the identifications u = x−1 and v = x−g−1y.

271

Page 272: Geir Ellingsrud and John Christian Ottem October 15, 2018€¦ · between algebra and geometry: a ne varieties correspond to nitely generated k-algebras without zero-divisors, and

18.4. Hyperelliptic curves

There is a morphism f : X → P1 coming from the inclusions k[x] ⊆ A and

k[u] ⊆ B; note that u = x−1 gives the standard gluing of P1 by the two affine

schemes Spec k[x] and Spec k[u]. The corresponding morphism X → P1 is finite,

of degree 2, as we are adjoining a square root of an element of k[x].

Let us compute the genus of X using Cech cohomology. We will use the

covering U = U, V above. Viewing the first ring as a k[x]-module, we can write

k[x, y]

(−y2 + a2g+1x2g+1 + · · ·+ a1x)= k[x]⊕ k[x]y

and similarly k[u]⊕ k[u]v as a k[u]-module for the ring on the right.

The first map in the Cech complex of OX has the following form:

d0 : (k[x]⊕ k[x]y)⊕(k[x−1]⊕ k[x−1]x−g−1y

)→ k[x±1]⊕ k[x±1]y

(p(x) + q(x)y, r(x−1) + s(x−1)x−g−1y) 7→ p(x)− r(x−1) + (q(x)− s(x−1)x−g−1)y

From this we can read off H0(X,O) = ker d0 = k and H1(X,O) ' kg, with

basis yx−1, yx−2, . . . , yx−g. In particular, g is exactly the arithmetic genus of the

curve.

For g = 2, we get a particularly interesting curve – a smooth projective curve

which cannot be embedded in P2. Indeed, any smooth curve in P2 has a degree d

and corresponding arithmetic genus 12(d− 1)(d− 2). However, there is no integer

solution to 12(d− 1)(d− 2) = 2!

Proposition 18.7. There exist non-singular projective curves which cannot be

embedded in P2.

To prove this, we need to verify that the curve is in fact projective. That

is, we need to find some projective embedding of it. To do this, we will need to

work out the cohomology groups H0(X,OX(nP )) for a point p ∈ X.

Let us for simplicity assume that a2g+1 = 1. Let p be the unique point in

u = v = 0 in X. In the local ring at p, we have

u = v2(1 + a2gu+ · · ·+ a1u2g)−1 = v2(unit),

272

Page 273: Geir Ellingsrud and John Christian Ottem October 15, 2018€¦ · between algebra and geometry: a ne varieties correspond to nitely generated k-algebras without zero-divisors, and

Chapter 18. The Riemann–Roch theorem and Serre duality

and hence vOp generates mp. We compute the some valuations of elements in

Op:

νp(v) = 1, νp(u) = 2, νp(x) = −2, νp(y) = 1 + (g + 1)(−2) = −(2g + 1)

We’ve seen that Γ(X,OX) = k, which agrees with our expectation that there

are no non-constant regular function on a projective curve. Let us consider the

case where the functions are allowed to have poles at p (and only at p). In other

words, we are interested in elements s ∈ Γ(X,OX(p)). Note that the point p

does not lie in U ; this means that s is regular there, and hence is a polynomial

in x, y. Now,

k[x, y]

−y2 + a2g+1x2g+1 + · · ·+ a1x= k[x] + k[x]y (as a k[x]-module)

So, any element s can be expressed as f(x) + h(x)y. We then can calculate

νp(f(x) + h(x)y) = minνp(f(x)), νp(h(x))νp(y)= min−2 deg f,−(2 deg h+ 2g + 1)

Thus, even if we allow a pole or order 1 at p, we still have Γ(X,OX(p)) = k.

On the other hand for 2p we gain an extra section, corresponding to f(x) = x:

Γ(X,OX(2p)) = k1, x

Note that O(2p)p = Op · x. The section x ∈ Γ(X,OX(2p)) is nonvanishing at

p, while the section 1 ∈ Γ(X,OX(2p)) is vanishing at p, since 1 = u · x and

u ∈ m ⊆ Op. Note that the linear series generated by 1, x generates OX(2p)

everywhere, inducing the morphism

Xϕ→ P1

(x, y) 7→ [1 : x]

which is the same morphism as the double cover above.

We can allow even higher order poles at p. The computation above shows

that

Γ(X,O(3p)) =

k1, x, y if g = 1

k1, x if g > 1

273

Page 274: Geir Ellingsrud and John Christian Ottem October 15, 2018€¦ · between algebra and geometry: a ne varieties correspond to nitely generated k-algebras without zero-divisors, and

18.5. The Riemann Roch theorem for curves

Using the embedding criterion of Chapter 10, one can show that in the case

g = 1, the sections x0 = 1, x1 = x, x2 = y give an embedding

X → P2k

(x, y) 7→ [1 : x : y]

The image is even seen to be a non-singular cubic curve: One computes that

Γ(X,O(6p)) is 6-dimensional, but we have 7 global sections: 1, x, y, x2, xy, x3, y2.

That means that there must be some relation between them. But of course this

is the relation

y2 = a3x3 + a2x

2 + a1x

giving the following relation in P2:

x22x0 = a3x

31 + a2x0x

21 + a1x

20x1

For g = 2, this map is not an embedding, since the image is P1. However, the

map given by 5p is: We obtain

Γ(X,OX(5p)) = k1, x, x2, y

These sections generate OX(5p), so we obtain a morphism

φ : X → P3

given by u0 = 1, u1 = x, u2 = x2, u3 = y. Notice that u0u2− u21 = 0, so X lies on

a quadric surface. In fact, the image of φ is precisely the relations between the

sections:

C = V(u2

1 − u0u2, u20u2 + u3

2 − u1u23, u

20u1 + u1u

22 − u0u

23

)⊆ P3

k

This shows that X is projective.

18.5 The Riemann Roch theorem for curves

When X is a curve over a field k, the cohomology groups H i(X,F ) are k-vector

spaces and we define will write

hi(X,F) := dimkHi(X,F)

Note that in this case, hi(X,F) = 0 for all i ≥ 2, so we have two invariants

h0(X,F) and h1(X,F) to work with.

Recall, that we defined for a sheaf F , the Euler characteristic χ(F) as the

274

Page 275: Geir Ellingsrud and John Christian Ottem October 15, 2018€¦ · between algebra and geometry: a ne varieties correspond to nitely generated k-algebras without zero-divisors, and

Chapter 18. The Riemann–Roch theorem and Serre duality

alternating sum of the hi(X,F). One useful property of χ(X,−) is that it is

additive on exact sequences:

Lemma 18.8. Let 0 → F ′ → F → F ′′ → 0 be an exact sequence of sheaves.

Then

χ(F) = χ(F ′) + χ(F ′′)

This follows because if 0→ V0 → V1 → · · · → Vn → 0 is an exact sequence

of k-vector spaces, then∑

i(−1)i dimk V = 0. Applying this to the long exact

sequence in cohomology gives the claim.

The most important sequence we will encounter is the ideal sheaf sequence

of a point p ∈, which takes the form

0→ OX(−p)→ OX → k(p)→ 0 (18.5.1)

where the first map is the inclusion and the second is evaluation at p. Here we

have identified the ideal sheaf mp ⊆ OX by the invertible sheaf OX(−p), and the

sheaf i∗Op with the skyscraper sheaf with value k(p) at p. If L is an invertible

sheaf, we can tensor (18.5.1) by L and get

0→ L(−p)→ L→ k(p)→ 0 (18.5.2)

where L(−p) is the invertible sheaf of sections of L vanishing at p. Taking χ, we

get

χ(L(−p)) = χ(L)− χ(k(−p)) = χ(L)− 1.

Theorem 18.9 (Easy Riemann–Roch). Let X be a smooth projective curve of

genus g and let D be a Cartier divisor on X. Then

χ(OX(D)) = h0(OX(D))− h1(X,OX(D)) = degD + 1− g

Proof. Let p ∈ X be a point and consider the sequence (18.5.2) with L =

OX(D + p). Then, as we just saw, χ(OX(D + p)) = χ(OX(D)) + 1. Also the

right-hand side of the equation above increases by 1 by adding p to D (since

deg(D + p) = degD + 1). This means that the theorem holds for a divisor D if

and only if it holds for D+p for any closed point p. So by adding and subtracting

points, we can reduce to the case when D = 0. But in that case, the left hand

side of the formula is by definition dimkH0(X,OX)− dimkH

1(X,OX) = 1− g,

which equals the right hand side.

The formula above is extremely useful because the right hand side is so easy to

compute. The number we are really after is the number h0(X,OX(D)), since this

is the dimension of global sections of OX(D). This in turn would help us to study

X geometrically, since we could use sections of OX(D) to define rational maps

275

Page 276: Geir Ellingsrud and John Christian Ottem October 15, 2018€¦ · between algebra and geometry: a ne varieties correspond to nitely generated k-algebras without zero-divisors, and

18.5. The Riemann Roch theorem for curves

X 99K Pn. So if we, for some reason, could argue that say, H1(X,OX(D)) = 0,

we would have a formula for the dimension of the space of global sections of

OX(D).

In any case, we can certainly say that h1(X,OX(D)) ≥ 0, so we get the

following bound on h0(X,OX(D)):

Corollary 18.10. h0(X,OX(D)) ≥ degD + 1− g

Example 18.11. A typical feature is that H1(X,OX(D)) = 0 provided that

the degree degD is large enough. This is essentially a consequence of Serre’s

theorem. To give an example, consider again the case where X is a hyperelliptic

curve of genus 2, as in the introduction. We have the following table of the

various cohomology groups H i(X,OX(np)) for the point p = (u, v):

D 0 1p 2p 3p 4p 5p 6p 7p

χ(OX(D)) -1 0 1 2 3 4 5 6

H0(X,OX(D)) 1 1 2 2 3 4 5 6

H1(X,OX(D)) 2 1 1 0 0 0 0 0

and it is not so hard to prove directly using the Cech complex thatH1(X,OX(np)) =

0 for all n ≥ 4.

Fortunately, there are more general results which tell us whenH1(X,OX(D)) =

0. This is due to the following fundamental theorem:

Theorem 18.12 (Serre duality). Let X be a smooth projective variety of dimen-

sion n and let D be a Cartier divisor on X. Then for each 0 ≤ p ≤ n,

dimkHp(X,OX(D)) = dimkH

n−p(X,OX(KX −D))

So if X is a curve, we get that H1(X,OX(D)) = H0(X,OX(KX −D)) and

the Riemann–Roch theorem takes the following form:

Theorem 18.13 (Riemann–Roch). Let X be a projective curve of genus g and

let D ∈ Div(X) be a divisor. Then

h0(X,OX(D))− h0(X,OX(KX −D)) = degD + 1− g

This is a much stronger statement than the Riemann–Roch formula we had

before, since group H0(X,OX(KX −D)) is easier to interpret: it is the space of

global sections of the sheaf associated to the divisor KX −D, or equivalently

ΩX(−D). It is usually easier to argue that there can be no such global sections

of this divisor. For instance, in the case degD < dimKX then KX −D cannot

be effective: if s ∈ H0(X,OX(KX −D)), then the divisor of s defines an effective

276

Page 277: Geir Ellingsrud and John Christian Ottem October 15, 2018€¦ · between algebra and geometry: a ne varieties correspond to nitely generated k-algebras without zero-divisors, and

Chapter 18. The Riemann–Roch theorem and Serre duality

divisor on X, that is, a sum div (s) =∑nipi where ni ≥ 0, contradicting our

assumption degKX < dimD.

So what is this degree of the canonical divisor KX? From Serre duality,

we get that H0(X,OX(KX)) and H1(X,OX) have the same dimension, so the

geometric genus and arithmetic genus agree:

pg = pa = g.

Then applying the Riemann–Roch formula to D = KX , we get

g − 1 = dimkH0(X,OX(K))− dimkH

0(X,OX(KX −K)) = degK + 1− g

and so degKX = 2g − 2. This observation gives us

Corollary 18.14. Suppose that D is a Cartier divisor of degree > 2g− 2. Then

H1(X,OX(D)) = 0, and

dimkH0(X,OX(D)) = degD + 1− g

Moreover, if degD = 2g − 2, then H1(X,OX(D)) 6= 0 only if D ∼ KX .

18.6 Proof of Serre duality for curves

Riemann–Roch and Serre duality are two of the very deep theorems in math-

ematics. The statement holds true in any dimension and they have numerous

consequences.

In this section we will outline the proof of Theorem 18.12 for curves. We

follow the proof in of Serre’s Groupes algebriques et corps de classes, with some

changes to make the proof as self-contained as possible.

So let X be a smooth projective curve over an algebraically closed field k

and let K denote the function field of X.

The product∏

x∈X K will play an important role in the following proof. It has

a natural structure of a (big!) k-vector space. An element r = (rx) ∈∏

x∈X K is

a rational function rx ∈ K for each closed point x ∈ X. These rational functions

can a priori be chosen in a completely arbitrary way, but or course for special

elements r, the rx can be linked together more closely. For instance, the function

field K is ‘diagonally embedded’ into∏

x∈X K: For each f ∈ K we get the

element r(f) by rx = f for all x ∈ X.

We follow Serre and call an element r ∈∏

x∈X K a repartition if rx ∈ OX,x

for all but finitely many x. Thus r consists of infinitely many rational functions

rx, indexed by the closed points in X, but all but finitely many of them are

regular near x for each x. We denote the set of repartitions by R. This is clearly

277

Page 278: Geir Ellingsrud and John Christian Ottem October 15, 2018€¦ · between algebra and geometry: a ne varieties correspond to nitely generated k-algebras without zero-divisors, and

18.6. Proof of Serre duality for curves

a subring of∏

x∈X K, as rxsx is regular near x if both rx are sx are. Also R

contains the ‘diagonally embedded’ K ⊆∏

x∈X K as a subring: If f ∈ K then f

lies in OX,x for almost all x and hence r(f) is a repartition.

Consider now a divisor D =∑npp on X. This defines an OX-module

OX(D) ⊆ K (where K is the constant sheaf with value K on X). This is similar

to the definition of OX(D), but we treat each point and each rational function

rx separately. The stalk of OX(D) at a point x consists of all elements f ∈ Ksuch that vx(f) ≥ −vx(D).

Let R(D) ⊆ R denote the additive subgroup of repartitions r ∈ R so that

vx(rx) ≥ −vx(D). Equivalently, we pick the r = (rx) so that each rx lies in the

stalk OX(D)x ⊆ K. Then R(D) has the structure of a sub-k-vector space of R.

If D ≤ D′ (or in other words, D′−D is effective), then clearly R(D) ⊆ R(D′),

as −vx(D′) ≤ −vx(D). Note also that R(0) coincides with the repartitions

r = (rx) so that rx ∈ OX,x is regular for every x ∈ X.

The quotient R/R(D) is naturally contained in the product∏

x∈X R/OX(D)x,

and it can be identified with the direct sum⊕

x∈X R/OX(D)x since the elements

of R are contained in OX,x for almost all x.

Proposition 18.15. We have a canonical isomorphism

H1(X,OX(D)) = R/(R(D) +K)

We will from now on define I(D) := R/(R(D) +K).

Proof. We start out with the short exact sequence

0→ OX(D)→ K → S → 0

where S = K/OX(D) is the quotient sheaf. The sheaf K is a constant sheaf, and

hence all of its higher cohomology groups vanish; in partcular H1(X,K) = 0.

Taking the long exact sequence in cohomology, we get

K → H0(X,S)→ H1(X,OX(D))→ 0

So it remains to identify the global sections of S.

For this, we need a simpler description of S. At a point x ∈ X, we have

Sx = (K/OX(D))x = f ∈ K|vx(f) + vx(D) < 0

and hence R/R(D) =⊕

x∈X Sx.We would like to show that S equals the the direct sum its skyscraper sheaves,

that is, S =⊕Sx and so sections of S coincides with a selection of elements

of Sx for x ∈ X, almost all of which are zero. Given this, the space of global

sections of S is indeed⊕

x Sx = R/R(D), which gives the proposition.

278

Page 279: Geir Ellingsrud and John Christian Ottem October 15, 2018€¦ · between algebra and geometry: a ne varieties correspond to nitely generated k-algebras without zero-divisors, and

Chapter 18. The Riemann–Roch theorem and Serre duality

In any case, we recall that an element of Sx is represented by an element

(r, U), where r ∈ S(U) and U ⊆ X is an open set containing x. Since K → S is

surjective, we may assume that r is represented by an element of K.

The element r may or not satisfy vx(r) < −vx(D) (i.e., r 6∈ OX(D)x): This

happens if the class [r] ∈ Sx is non-zero. Note that vy(r) = 0 for y in a

neighbourhood V of U − x: This is because the rational function is regular in a

neighbourhood of x (here we use that X is a curve), and we can assume that

it is has no zero there. We also have that vx(D) = 0 in a neighbourhood W of

U − x. Hence [r] vanishes in the stalks Sy for y ∈ V ∩W .

Now we conclude the proof using the following lemma:

Lemma 18.16. Let F be a sheaf on a quasicompact topological space X so that

the closed points are dense in x. Suppose that for each point x ∈ X and each

element s ∈ Fx there exists an open set U 3 x and a section σ ∈ F(U) which

induces s in Fx and vanishes on U − x. Then

F(X) =⊕x∈X

Fx

where the sum is over the closed point of X.

Proof. We have an injective map

F(X)→⊕x∈X

Fx

which sends a section σ ∈ F (X) to σx in each stalk. We will show that the image

is the whole direct sum of the stalks Fx (i.e., that a global section of F is zero

for all but a finite number of points x, and there are no conditions on the points

x and the ‘values’ sx).

Let σ ∈ F(X), and choose an open set Ux of each point x ∈ X so that σ|Ux−xvanishes. Since X is quasicompact, X can be covered by finitely many of the

sets Ux, say Ux1 , . . . , UXr . So σ is zero outside of the points x1, . . . , xr.

Conversely, let x1, . . . , xr be points in X and suppose we are given an element

si ∈ Fxi for each i = 1, . . . , r. Each si can be extended to a section σi of F in

some neighbourhood Ui of xi which vanishes on Ui−xi. If we let Vi = Ui−xjj 6=i,then the Vi form an open cover of X, which is such that Vi ∩ Vj contains none

of the points x1, . . . , xr. Therefore, σ|Vi∩Vj = 0 = σj|Vi∩Vj and the σi glue to a

global section of F with the desired property.

The take-away from this is that we have an explicit description ofH1(X,OX(D))

in terms of rational functions. This is already interesting in the case D = 0. In

279

Page 280: Geir Ellingsrud and John Christian Ottem October 15, 2018€¦ · between algebra and geometry: a ne varieties correspond to nitely generated k-algebras without zero-divisors, and

18.6. Proof of Serre duality for curves

this case, H1(X,OX) = 0 means that for any collection of rational functions rxwhich are allowed to have poles at a finite number of points in X, there is a

rational function f ∈ K so that rx − f is regular for every x ∈ X.

18.6.1 Laurent tails

In this section A will be a discrete valuation ring, which is an algebra over its

residue field k = A/m, where of course m is the maximal ideal of A. Our main

example will be when A = OX,x where X is a curve over an algebraically closed

field and x is a closed point, or more generally A = ΩX,η where η is a point of

codimension 1.

Let K be the fraction field of A and let v : K → Z denote the normalized

valuation. The field K has a decreasing filtration (infinite in both directions) of

A-modules Ki given by

Ki = f ∈ K|v(f) ≥ i

Note that K0 = A, and Ki = mi for i ≥ 0. Moreover, the quotient Ki/Ki+1 has

the structure of a k-vector space of dimension 1: it is spanned by [ti] where t is

a uniformizing parameter of A.

Lemma 18.17. Any element f ∈ K can be written as

f =∑i>0

a−it−i + g

where g ∈ A.

Proof. This is obvious if f ∈ A. More generally, take f ∈ F and suppose

n = v(f) < 0. Then f ∈ Kn and the class [f ] in Kn/Kn+1 is a scalar multiple of

the generator [tn], say [f ] = α[tn] where α ∈ k. In particular, f − αrn maps to

zero in Kn/Kn+1, and hence v(f − αtn) ≥ n+ 1. So by induction on |v(f)| we

can write f =∑

i>0 α−it−i + g and we are done.

18.6.2 Residues on A

Let t be a uniformizing parameter in A, and consider a differential ω = fdt which

lies inM = ΩK|k. We define the residue of f (relative to t) as the coefficient a−1

in the above decomposition of f . This is denoted by Rest f , Rest fdt or Rest ω.

We have

fdt = a−nt−ndt+ · · ·+ a−2t

−2dt+ Rest f · t−1dt+ g dt

An important point, which we will address below, is that this definition does not

depend on the choice of uniformizing parameter t.

280

Page 281: Geir Ellingsrud and John Christian Ottem October 15, 2018€¦ · between algebra and geometry: a ne varieties correspond to nitely generated k-algebras without zero-divisors, and

Chapter 18. The Riemann–Roch theorem and Serre duality

Here are some standard properties of residues

Lemma 18.18. (i) Rest is a k-linear map Rest : ΩA|k → k.

(ii) Rest(fdt) = 0 if f ∈ A

(iii) Rest(df) = 0 for f ∈ K.

(iv) Rest(g−1dg) = v(g) for all g ∈ K − k.

Proof. These properties are easy to verify. (i) is clear. (ii) follows since all

elements of A have zero residues by definition. For (iii), we write

f =∑i>0

a−it−i + g

where g ∈ A. This gives that df =∑

i>0(−i)a−it−i−1dt + dg where dg ∈ ΩA|k.

The coefficient of t−1 vanishes.

For (iv), write g = tnα, where n = v(g) and α ∈ A×. Differentiation gives

that dg = ntn−aαdr + tndα, which gives that

g−1dg = nt−1dt+ α−1dα

As dα ∈ ΩA|k and α is invertible in A, so α−1dα ∈ ΩA|k which gives the claim.

18.6.3 Residues on X

Let now X be a smooth projective curve over k, and let ΩX = ΩX|k denote

the sheaf of differentials. For each divisor D we also have the invertible sheaf

ΩX(D) = ΩX ⊗ OX(D) of differentials with poles at worst order vx(D) on D for

each x ∈ X.

We let M = ΩK|k denote the space of meromorphic differentials on X. Note

that M ' Kdt is a 1-dimensional K-vector space. Given an element ω ∈ M,

and a point x ∈ X, K is the fraction field of OX,x so we can use the residue map

from the previous section. In particular, we get a map

Resx :M→ k.

Proposition 18.19. Let X be a smooth projective curve over k. Then

(i) (Independence of uniformizer) Resx(ω) does not depend on the choice of t.

(ii) (Residue theorem) For any ω ∈ ΩK|k, Resx(ω) = 0 for all but finitely many

x. Moreover,∑

x∈X Resx ω = 0.

281

Page 282: Geir Ellingsrud and John Christian Ottem October 15, 2018€¦ · between algebra and geometry: a ne varieties correspond to nitely generated k-algebras without zero-divisors, and

18.6. Proof of Serre duality for curves

Over the field k = C, these are indeed the same residues you know from

complex analysis. Here the theorem is easy to prove: (i) follows since the residue

is uniquely determined by

Resx(ω) =

∫σ

ω

for a small circle σ around x. The point (ii) follows just as easily, by the Residue

theorem. Indeed, if Di are small disks around each pole of ω, ω is holomorphic

in U = X −⋃Di, and so

∫∂Uω = 0, which means that the residues sum to zero.

One can prove the above properties (i) and (ii) over any field, completely

algebraically, but it is a bit more involved (especially in positive characteristic),

so we will not do so here.

We now return to the proof of Serre duality. We recall the set of repartitions

R, which form a vector space over K. (If f ∈ K and r is a repartition, the

product fr is given by (fr)x = frx: Again this makes sense since a rational

function is regular outside a finite number of points.)

Definition 18.20. Let I(D) = R/(R(D) +K) and let

J(D) = Homk(I(D), k)

denote the space of k-linear functionals on I(D).

We can (and will!) think of the elements of J(D) as the linear functionals on

the huge ring of repartitions R that vanish on the subspace R(D) +K. So we

have an inclusion

J(D) ⊆ Homk(R, k).

If D′ ≥ D, we saw before that R(D) ⊆ R(D′), which implies that J(D′) ⊆ J(D).

We define

J =⋃D

J(D)

where the sum is over all divisors on D. This is yet another horribly big vector

space over k.

Actually, J also has the structure of a vector space over K, via the product

fα(r) = α(fr). Note that α(fr) is a k-linear functional that vanishes on K.

If α vanishes on R(D), and div f = D′, then fα vanishes on R(D − D′): If

r ∈ R(D −D′), then fr ∈ R(D) because

vx(fr) = vx(f) + vx(r) ≥ vx(f)− (vx(D) + vx(f)) = −vx(D)

and hence α(fr) = 0. Hence fα belongs to J(D −D′) ⊆ J .

Proposition 18.21. We have dimK J ≤ 1

282

Page 283: Geir Ellingsrud and John Christian Ottem October 15, 2018€¦ · between algebra and geometry: a ne varieties correspond to nitely generated k-algebras without zero-divisors, and

Chapter 18. The Riemann–Roch theorem and Serre duality

Proof. Assume the contrary and let α, β be two linearly independent elements

of J . Let D be a divisor, such that α and β both lie in J(D). Let d = degD.

For each n ∈ Z we pick a divisor Dn so that degDn = n (for instance

Dn = nP for a point P will do). If f ∈ H0(X,OX(Dn)), then fα ∈ J(D −Dn),

by the argument above.

Define the map

H0(X,OX(Dn))⊗H0(X,OX(Dn)) → J(D −Dn)

(f, g) 7→ fα + gβ

This is injective since α, β are assumed to be linearly independent over K. In

particular,

dimk J(D −Dn) ≥ 2h0(X,OX(Dn))

We will show that this gives a contradiction by estimating these dimensions using

Easy Riemann–Roch.

On the left-hand side we get

dimk J(D −Dn) = dimk I(D −Dn)

= h1(X,OX(D −Dn))

= h0(X,OX(D −Dn)− (d− n) +O(1)

= n+O(1)

for n 0. Here O(1) means a constant independent of n and Dn.

On the right-hand side we use the inqueality h0(D) ≥ degD + 1− g:

2h0(X,OX(Dn)) ≥ 2 deg(Dn) +O(1) = 2n+ O(1)

So if n 0 we get a contradiction as the two sides cannot be equal. Hence α, β

cannot be linearly independent, which finishes the proof.

18.6.4 The final part of the proof

The statement of Serre duality is that two vectors are dual to each other. We

have to define the pairing explicitly and check that it is indeed perfect1.

Let ω ∈ M be a meromorphic differential on X. This defines a canonical

divisor

div (ω) :=∑x∈X

vx(ω)x

1Recall that a bilinear map M × N → A is a perfect pairing if the induced map M 7→HomA(N,A) is an isomorphism

283

Page 284: Geir Ellingsrud and John Christian Ottem October 15, 2018€¦ · between algebra and geometry: a ne varieties correspond to nitely generated k-algebras without zero-divisors, and

18.6. Proof of Serre duality for curves

Note that this means that the invertible sheaf ΩX(−D) is the sheaf of differentials

so that divω −D ≥ 0.

Definition 18.22. We define the Serre duality pairing as follows:

〈 , 〉 :M×R → k

(ω, r) 7→ 〈ω, r〉 =∑x∈X

Resx(rxω)

This has the following properties:

(1) 〈ω, r〉 = 0 if r ∈ K(by the Residue theorem)

(2) 〈ω, r〉 = 0 if r ∈ R(D) and ω ∈ ΩX(−D)(X).

(The product rxω cannot have a pole for any x ∈ X, because the zeroes must

at least cancel the poles by the assumptions on r and ω.

(3) 〈fω, r〉 = 〈ω, fr〉 if f ∈ K(Since both pairings evaluate to the same sum of the residues over fωr.)

Now, fixing a meromorphic differential ω ∈ ΩX(−D)(X), gives a linear

functional θ(ω) on R. By properties (1) and (2), this descends to a functional

on R/(R(D) +K). Hence we get a map

θ : ΩX(−D)→ J(D)

(Recall that J(D) is the dual of R/(R(D) +K)).

Lemma 18.23. Let ω ∈M such that θ(ω) ∈ J(D). Then ω ∈ ΩX(−D)(X).

Proof. Suppose this is not true, so that ω 6∈ ΩX(−D)(X). Hence there is a

point x ∈ X so that ω has a pole in x with a higher order than allowed by D:

vx(ω) < vx(−D).

Pick the repartition r ∈ R(D) by defining

rq =

0 for q 6= x

t−vx(ω)−1 for q = x

Then vx(rxω) = −1, and so

〈ω, r〉 =∑q∈X

Res(rqω) = Resx(rω) 6= 0

This means that θ(ω) 6= 0 on R(D). However, θ(ω) is supposed to vanish on all

of R(D) +K, so we have a contradiction.

284

Page 285: Geir Ellingsrud and John Christian Ottem October 15, 2018€¦ · between algebra and geometry: a ne varieties correspond to nitely generated k-algebras without zero-divisors, and

Chapter 18. The Riemann–Roch theorem and Serre duality

Theorem 18.24 (Serre duality). The map θ induces an isomorphism

θ : H0(X,ΩX(−D))→ H1(X,OX(D))∨

Proof. θ is injective. Let ω ∈ H0(X,ΩX(−D)) be so that θ(ω) = 0. Then by

the previous lemma, we have that ω ∈ ΩX(−D′)(X) for every divisor D′. This

implies that ω = 0, since it is impossible to have a non-zero ω ∈ M so that

divω −D′ ≥ 0 for every divisor D′ (just take D′ to have very high degree).

θ is surjective. θ is K-linear, and we can view it as a map from M to J .

By definition, dimKM = 1, and we also have dimK J ≤ 1. An injection of

finite dimensional vector spaces into a smaller dimensional vector space must be

surjective. Hence if α ∈ J(D), we get a meromorphic differential ω ∈M so that

θ(ω) = α, and so the previous lemma shows that ω ∈ Ω(−D).

285

Page 286: Geir Ellingsrud and John Christian Ottem October 15, 2018€¦ · between algebra and geometry: a ne varieties correspond to nitely generated k-algebras without zero-divisors, and
Page 287: Geir Ellingsrud and John Christian Ottem October 15, 2018€¦ · between algebra and geometry: a ne varieties correspond to nitely generated k-algebras without zero-divisors, and

Chapter 19

Applications of the

Riemann–Roch theorem

In this chapter we give a few of the (many) consequences of the Riemann–Roch

formula. We start by translating the results of Chapter 16 into concrete numerical

criteria for a divisor D to be base-point free or very ample. Then we use these

results to classify all curves of all genus ≤ 4.

19.1 Morphisms of curves

Let f : X → Y be a morphism of schemes. Recall that we say that f is said to

be finite if the following two conditions are satisfied:

(i) f is affine, i.e., if U ⊆ Y is open and affine, then f−1(U) ⊆ X is afiine.

(ii) If U = SpecA and V = f−1(U) = SpecB, then B is a finite A-algebra (via

f#).

For such morphisms, the pushforward f∗OX is a coherent OY -module.

If y ∈ Y is a closed point, corresponding to the sheaf of ideals my ⊆ OY , then

the scheme theoretic fibre Xy is defined in X by the ideal sheaf f ∗my ·OX ⊆ OX .

Note that in the case X = SpecB, and Y = SpecA, the scheme theoretic fibre

over y is naturally isomorphic to Spec(B/myB) where my is the maximal ideal

in the point y. If k(y) = OY,y/myOY,y = A/m denotes the residue field, then we

have f−1(x) ' Spec(B⊗A k(y)). Note that B⊗A k(y) is a finite algebra over the

field k(y). Such an algebra has only finitely many prime ideals (and all of these

287

Page 288: Geir Ellingsrud and John Christian Ottem October 15, 2018€¦ · between algebra and geometry: a ne varieties correspond to nitely generated k-algebras without zero-divisors, and

19.1. Morphisms of curves

are maximal). In particular, a finite morphism can only have a finite number of

preimages over a closed point y ∈ Y .

The converse of this statement is of course not true. If f : U = SpecAf →X = SpecA is an open immersion, e.g., f : Spec k[x, y]/(xy − 1) → Spec k[x],

the number of preimages are of course finite (0 or 1 in fact). However, Af is

usually not finite as an A-algebra.

However if f : X → Y is a projective morphisms (or more and its fibers are

finite, then f is finite. We will prove this in the next section for curves.

We can be a little bit more precise when talking about the ‘number of

preimages’ of f . Let Y = SpecA be an integral noetherian scheme; so that A is

a noetherian integral domain. Let f : X → Y be a finite morphism. Then X is

also affine, say X = SpecB for some finite A-algebra B. We finally suppose that

f is dominant. This means that the image of each component of X is dense in X.

As A is an integral domain, this is equivalent to A ⊆ B, and q ∩ A = 0 for each

associated primeideal q ⊆ B. It follows that we can extend the inclusion A ⊆ B

to an inclusion K(A) ⊆ K(B). Since B is a finite A-algebra, this extension is

finite. We define the degree of f as the degree of the field extension

[K(B) : K(A)] = dimK(A) K(B).

Example 19.1. Let Y = Spec k[x], X = Spec k[x, y]/(x−y2) and let f : X → Y

denote the projection map (induced by k[x] ⊆ k[x, y]/(x−y2). This is a morphism

of degree two, since K(X) = k(y) is a degree two extension of K(Y ) = k(x)

(since x = y2). Note that the fiber of f over the point 0 ∈ Y has set-theoretically

only one preimage, namely the point (0, 0) ∈ Y corresponding to (0, 0). However,

the scheme theoretic fiber here has a non-reduced structure that reflects the

number 2:

Xy = Spec k[x, y]/(x− y2)⊗k[x] k(x) ' Spec k[y]/y2

19.1.1 Morphisms of curves are finite

Let us now specialize to the case of curves. For a closed point x ∈ X, the local

ring OX,x is a discrete valuation ring and there is an element t ∈ OX,x generating

the maximal ideal m. We say that t is a local parameter or a uniformizing

parameter if vx(t) = 1. Note that we can always normalize the valuation so that

a generator of m has valuation 1.

Proposition 19.2. Let X be a projective non-singular curve over k and let Y

be any curve. Let f : X → Y be a morphism. Then either

(i) f(X) = a point in Y ; or

288

Page 289: Geir Ellingsrud and John Christian Ottem October 15, 2018€¦ · between algebra and geometry: a ne varieties correspond to nitely generated k-algebras without zero-divisors, and

Chapter 19. Applications of the Riemann–Roch theorem

(ii) f(X) = Y ;

In the case (ii) K(X) ⊇ K(Y ) is a finite extension, f is a finite morphism and

Y is also projective.

Proof. Since X is projective, the image f(X) must be closed in Y (Add this

theorem). On the other hand X is irreducible, and hence so is f(X). Hence

f(X) is either a point or all of Y .

If f(X) = Y , then f is dominant, so we get an inclusion of function fields

K(Y ) ⊆ K(X). Both of these fields are finitely generated over k of the same

transcendence degree (i.e., 1), so the extension is finite algebraic.

Let V = SpecB ⊆ Y be an open set of Y and let A denote the integral

closure of B in K(X). Then by [Hartshorne I.6.7] we have SpecA = f−1(V ). So

since A is a finite B-module, we have that f is finite.

We also have the following theorem of [Hartshorne Ch. I. 6.12]:

Theorem 19.3. There is an equivalence of categories

1. Non-singular projective curves, and dominant morphisms

2. Function fields K of dimension 1 over k and k-homomorphisms.

19.1.2 Pullbacks of divisors

Given f : X → Y (non-constant) we have a well-defined pullback map

f ∗ : Pic(Y )→ Pic(X)

sending the class of an invertible sheaf L on X to that of f ∗L. If L = OX(D)

for some divisor, we can make this pullback a bit more explicit as follows.

Let t ∈ OY,y denote the uniformizing parameter. We consider t as an element

of K(X) via the field extension K(X) ⊇ K(Y ). In particular, we can talk about

the valuations of t in each local ring OX,x and define the Weil divisor

f ∗(y) =∑

x∈f−1(y)

vx(t)x

(Since f is finite, there are only finitely many preimages of y, and so the sum is

finite). The expression is also independent of the choice of t: If t′ is a different

uniformizing parameter, we can in any case write t′ = ut where u is a unit, so

that vx(t) = vx(t′).

Thus we get a well defined map

f ∗ : Div(Y )→ Div(X)

289

Page 290: Geir Ellingsrud and John Christian Ottem October 15, 2018€¦ · between algebra and geometry: a ne varieties correspond to nitely generated k-algebras without zero-divisors, and

19.1. Morphisms of curves

which descends to the above map f ∗ : Pic(Y )→ Pic(X).

Lemma 19.4. We have deg f ∗D = deg f · degD.

Proof. TODO.

Definition 19.5. For a morphism f : X → Y , we call the number ex = vx(t)

the ramification index of f at x. If ex > 1, we say that f is ramified at x.

Example 19.6. Let A = k[x] and let B = k[x, y]/(x− y2) ' k[y] where k is a

field. Let X = SpecB and let Y = SpecA. Let f : X → Y be the morphism

induced by the inclusion A → B (thus x 7→ y2).

The morphism f is ramified only at the origin (0, 0), and here the ramification

index is two. Indeed, x is a uniformizing parameter of OY,y = k[x](x), while y is

the uniformizer of OX,x = B(x,y). Then we have v(0,0)(x) = v(0,0)(y2) = 2.

The reader might notice a resemblance between the previous example and

Example 11.41, where ramification was defined in terms of the relative sheaf of

differentials ΩX|Y . In that example, ΩX|Y was a torsion sheaf supported on the

single point (0, 0). This correspondence between the two notions of ramification

is a general fact (at least in characteristic 0), and we have the useful formula for

the ramification indexes of curves:

ep = length(ΩX|Y )p + 1

19.1.3 Morphisms to P1

Let X be a non-singular curve, and let f ∈ K. Then f induces a morphism

φ : X → P1 in the following way. Let U = X−Supp f−1(∞) and V = X−f−1(0),

so that f ∈ OX(U) and 1/f ∈ OX(V ).

Write P1 = Proj k[x0, x1]. On D+(x0), the map k[x1/x0]→ OX(U) sending

x1/x0 7→ f , induces a map φU : U → D+(x0) ⊆ P1. Similarly, on D+(x1), we

get a map φV : V → P1. These maps coincide on U ∩ V , and therefore we get a

morphism

φ : X → P1

This morphism is non-constant, hence finite.

290

Page 291: Geir Ellingsrud and John Christian Ottem October 15, 2018€¦ · between algebra and geometry: a ne varieties correspond to nitely generated k-algebras without zero-divisors, and

Chapter 19. Applications of the Riemann–Roch theorem

19.2 Very ampleness criteria

Recall the criterion of Theorem 16.18, that an invertible sheaf L is very ample if

and only if its linear system separates points and tangent vectors. Using Riemann–

Roch we can translate that result into a very simple, numerical criterion for very

ampleness on a curve:

Theorem 19.7. Let X be a non-singular projective curve and let D be a divisor

on X. Then

(i) |D| is base-point free if and only if

h0(D − P ) = h0(D)− 1 for every point P ∈ X.

(ii) |D| is very ample if and only if

h0(D − P −Q) = h0(D)− 2 for every two points P,Q ∈ X

(including the case P = Q)

(iii) A divisor D is ample iff degD > 0

Proof. (i) We take the cohomology of the following exact sequence

0→ OX(D − P )→ OX(D)→ k(P )→ 0

and get

0→ H0(X,OX(D − P ))→ H0(X,OX(D))→ k

From this sequence, we get h0(D)− 1 ≤ h0(D − P ) ≤ h0(D).

The right-most map takes a global section of OX(D) and evaluates it at P . To

prove that |D| is base point free, we must prove that there is a section s ∈ OX(D)

which does not vanish at P , or equivalently, that the map H0(X,OX(D))→ k

is surjective. But this happens if and only if h0(D − P ) = h0(D)− 1.

(ii) If the above inequality is satisfied, we see in particular that |D| is base-

point free. So it determines a morphism φ : X → Pn. We can use Theorem

16.18 that ensure that φ is an embedding. That is, φ separates (a) points and

(b) tangent vectors.

For (a), we are assuming that h0(D − P − Q) = h0(D) − 2, so the divisor

D − P is effective and does not have Q as a base point (by (i)). But this means

that there is a section of H0(X,OX(D − P )) which doesn’t vanish at Q. We

have H0(X,D − P ) ⊆ H0(X,D), so we get a section of OX(D) which vanishes

at P , but not at Q. Hence |D| separates points.

For (b), we need to show that |D| separates tangent vectors, i.e., the elements

of H0(X,OX(D)) should generate the k-vector space mPOX(D)/m2POX(D) at

291

Page 292: Geir Ellingsrud and John Christian Ottem October 15, 2018€¦ · between algebra and geometry: a ne varieties correspond to nitely generated k-algebras without zero-divisors, and

19.3. Curves on P1 × P1

every point P ∈ X. This condition is equivalent to saying that there is a divisor

D′ ∈ |D| with multiplicity 1 at P : Note that dim TP (X) = 1, dim TPD′ = 0 if

P has multiplicity 1 in D′ and dim Tp(D′) = 1 if P has higher multiplicity. But

this is equivalent to P not being a base-point of D − P . Applying (i) again, we

see that h0(D − 2P ) = h0(D)− 2, so we are done.

(iii) By definition, D is ample if mD is very ample for m 0. So the result

follows by the next result, since any divisor of degree ≥ 2g+ 1 is very ample.

Proposition 19.8. Let X be a non-singular projective curve and let D be a

divisor on X. Then

1. If degD ≥ 2g, then |D| is base-point free.

2. If degD ≥ 2g + 1, then |D| is very ample.

Proof. By Serre duality, h1(D) = h0(K−D) = 0 because degD > degK = 2g−2.

Similarly, h1(D − P ) = 0.

(i) Applying Riemann–Roch, we find that h0(D − p) = h0(D) − 1 for any

P ∈ X, so we are done by the above theorem.

(ii) In this case we also get h1(D − P − Q) = 0, so Riemann–Roch shows

that h0(D − P −Q) = h0(D)− 2, which is the conclusion we want.

19.3 Curves on P1 × P1

Let us consider one central example, namely curves on Q = P1×P1. Recall, that

Cl(Q) = ZL1 ⊕ ZL2 where L1 = [0 : 1]× P1 and L2 = P1 × [0 : 1].

We can use this to prove that Q contains non-singular curves of any genus

g ≥ 0. (This is in contrast with the case of P2, where only genera of the form(d−1

2

)were allowed).

To prove this, consider the divisor D = aL1 + bL2 where a, b ≥ 1. D is

effective, so let C ∈ |D| be a generic element.

Lemma 19.9. C is non-singular.

To compute the genus of C, we use the formula 2g− 2 = deg ΩC . So we need

to find ΩC and compute its degree. This is best computed using the Adjunction

formula of Proposition 11.43:

ΩC = ωQ ⊗ OQ(C)|X= OQ(−2L1 − 2L2)⊗ O(aL1 + bL2)|C= OC((a− 2)L1 + (b− 2)L2)

To compute the degree of this, we consider the degrees of L1|C and L2|C separately.

Note that the degree degL1|C is invariant under linear equivalence, so we can

292

Page 293: Geir Ellingsrud and John Christian Ottem October 15, 2018€¦ · between algebra and geometry: a ne varieties correspond to nitely generated k-algebras without zero-divisors, and

Chapter 19. Applications of the Riemann–Roch theorem

compute the degree of any [s : t]× P1 for a general point [s : t]× P1. The point

is that as a Weil divisor, L1|X is obtained by intersecting [s : t] × P1 with X.

When [s : t] ∈ P1 is a general point, the intersection X ∩ [s : t]× P1 is a reduced

subscheme of X, consisting of b points (as C ⊆ Q = P1 × P1 is a divisor of type

aL1 + bL2). Hence degL1|C = b and degL2|C = a. It follows that

2g − 2 = deg ΩC = (a− 2)b+ (b− 2)a = 2ab− 2a− 2b

Solving for g gives us the following theorem:

Theorem 19.10. Let Q = P1 × P1. Then a generic divisor C in |aL1 + bL2| is

a smooth projective curve of genus

g = (a− 1)(b− 1).

In particular, Q contains non-singular curves of any genus g ≥ 0.

19.4 Curves of genus 0

The results of the previous results are particularly strong when the genus is small.

For instance, when g = 0, any divisor of positive degree is very ample! We can

use this to classify all curves of genus 0.

Lemma 19.11. Let X be a non-singular curve. Then X ' P1 if and only if

there exists a Cartier divisor D such that degD = 1 and h0(X,OX(D)) ≥ 2. In

this case, the divisor D is even very ample.

Proof. This follows directly from Proposition 19.8. Still, let us give an alternative

proof, which is perhaps a little bit more enlightening.

Let g ∈ H0(X,OX(D)). Then D′ ∼ div g +D ≥ 0, so we may assume that

D is effective. Since degD = 1, we must have D = p for some point p ∈ X. Now

take f ∈ H0(X,OX(D))− k. As above, f induces a morphism φ : X → P1. This

morphism has degree equal to 1, so it is birational, and hence X is isomorphic

to P1.

Proposition 19.12. A non-singular curve X is isomorphic to P1 if and only if

Pic(X) ' Z.

Proof. We have seen that the Picard group of any Pnk is isomorphic to Z.

Conversely, suppose X is a curve with Pic(X) ' Z. Let p, q be two distinct

points on X. By assumption, OX(p) ' OX(q), so the linear system |p| has at

least two sections. Then X ' P1k by the previous lemma.

Theorem 19.13. Any curve of genus 0 over an algebraically closed field is

isomorphic to P1.

293

Page 294: Geir Ellingsrud and John Christian Ottem October 15, 2018€¦ · between algebra and geometry: a ne varieties correspond to nitely generated k-algebras without zero-divisors, and

19.5. Curves of genus 1

Proof. Let p ∈ X be a point and consider the divisor D = p. If X has genus 0,

then 1 = degD > 2g − 2 = −2, so H1(X,OX(D)) = 0. Then Riemann-Roch

tells us that

dimH0(X,OX(D)) = 1 + 1− 0 = 2

Hence X ' P1k by Lemma 19.11.

We conclude by yet another characterisation of P1:

Lemma 19.14. Let C be a non-singular projective curve and D any divisor of

degree d > 0. Then

dim |D| ≤ degD

with equality if and only if C ' P1.

Proof. This is (IV, Ex. 1.5) in Hartshorne. Although one might guess that this

lemma follows from Riemann-Roch this is not the case. Riemann-Roch gives a

different sort of relationship between the dimension and degree of a divisor. We

induct on d.

First suppose d = 1. There is an exact sequence

0→ OC → OC(P )→ k(P )→ 0.

Now h0(OC) = 1 and h0(k(P )) = 1 therefore h0(OC(P )) ≤ 2 so dim |P | ≤ 1. If

dim |P | = 1 then |P | has no base points so we obtain a morphism C → P1 of

degree degP = 1 which must be an isomorphism so C is rational.

Next suppose D = P1 + · · ·+Pd. Let D′ = P1 + · · ·+Pd−1. There is an exact

sequence

0→ OC(D′)→ OC(D)→ k(Pd)→ 0.

Now h0(OC(D′)) ≤ d by induction and h0(k(Pd)) = 1 so h0(OC(D)) ≤ d+1, there-

fore dim |D| ≤ d with equality iff h0(OC(D′)) = d. By induction h0(OC(D′)) = d

iff C is rational.

19.5 Curves of genus 1

A curve in P2k of degree three has genus 1. This follows from our earlier work on

the canonical bundle, which showed ωX ' OP2k(d− 3)|X ' OX . In this section,

we note that in fact every curve of genus 1 arises this way:

Theorem 19.15. Any projective curve of genus 1 can be embedded as a plane

cubic curve in P2k.

Proof. Pick a point P ∈ X and consider the divisor D = 3P . D has degree

3 ≥ 2g + 1, so it is very ample. Let φ : X → P2k denote the embedding. The

image φ(X) is a smooth curve of degree equal to deg φ∗OP2(1) = degD = 3.

294

Page 295: Geir Ellingsrud and John Christian Ottem October 15, 2018€¦ · between algebra and geometry: a ne varieties correspond to nitely generated k-algebras without zero-divisors, and

Chapter 19. Applications of the Riemann–Roch theorem

In contrast to the g = 0 case, there is however many non-isomorphic genus 1

curves. For instance, in the Ledengre family of curves in Xλ ⊆ P2 given by

y2z = x(x− z)(x− λz)

where λ ∈ k, each Xλ is isomorphic to at most a finite number of other Xλ′ ’s.

19.5.1 Divisors on X

Let us study the divisors on X. To make the discussion a bit more concrete, let

X ⊆ P2 be the curve given by y2z = x3 − xz2. We claim that there is an exact

sequence

0→ X(k)→ Cl(X)deg−−→ Z→ 0

This means that the Class group Cl(X) is very big – its elements are in bijection

with the k-points of X, of which there might be uncountably many. In particular

X cannot be isomorphic to P1.

If L ⊆ P2 is a line, we get a divisor L|X . That is, we take a section s ∈ OP2(1)

defining L and restrict it to X. The divisor of s ∈ OX(1) consists of three points

P,Q,R (counted with multiplicity). In particular, since any two lines are linearly

equivalent on P2, we get for every pair of lines L,L′ and corresponding triples

P,Q,R, a relation

P +Q+R ∼ P ′ +Q′ +R′

(where ∼ denotes linear equivalence).

Let us consider the point O = [0, 1, 0] on X. This is a special point on X: it

is an inflection point, in the sense that there is a line L = V (z) ⊆ P2 which has

multiplicity three at O, so that L restricts to 3O on X. This has the property

that any three collinear points P,Q,R in X satisfy

P +Q+R ∼ 3O

We will use these observations to define a group structure on the set of closed

points X(k), using the point O as the identity. The group structure will be

induced from that in Cl(X).

Consider the subgroup Cl0(X) ⊆ Cl(X) consisting of degree 0. This fits into

the exact sequence

0→ Cl0(X)→ Cl(X)deg−−→ Z→ 0

295

Page 296: Geir Ellingsrud and John Christian Ottem October 15, 2018€¦ · between algebra and geometry: a ne varieties correspond to nitely generated k-algebras without zero-divisors, and

19.5. Curves of genus 1

We now define a map

ξ : X(k) → Cl0(X)

P 7→ [P −O]

Lemma 19.16. ξ is a bijection.

Proof. ξ is injective: ξ(P ) = ξ(Q) implies that P ∼ Q. Then P = Q (otherwise

X would be rational, by Proposition 19.12). (Alternatively, it follows because

h0(X,OX(P )) = 1).

ξ is surjective: Take a divisor D =∑niPi ∈ Div(X) of degree 0. Then

D′ = D+O has degree 1, so by Riemann–Roch, H0(X,OC(D′)) is 1-dimensional.

Hence there exists an effective divisor of degree 1 in |D′|, which must then be

of the form D′ = Q. But that means that D + O ∼ Q, or, D ∼ Q − O, as

desired.

Using this bijection, we can put a group structure on the set X(k):

Theorem 19.17. The set of k-points X(k) on a genus 1 form a group.

The group law has the following famous geometric interpretation. Given two

points p1, p2 ∈ X, we draw the line L they span (see the figure below). This

intersects X in one more point, say p3. In the group Cl0(X) we have

p1 + p2 + p3 = 3O

To define the ‘sum’ p1 + p2 (which should be a new k-point of X), we then look

for a point p4 so that

p4 −O = (p1 −O) + (p2 −O)

or in other words, p4 +O = p1 +p2. By the above, this becomes p4 +O = 3O−p3

or, p3 + p4 + O = 3O. This tells us that we should define p4 as follows: We

draw the line L′ from O to p3 (shown as the dotted line in the figure), and define

p4 to be the third intersection point of L′ with X. By construction, we get

(p1 −O) + (p2 −O) = (p4 −O) in Cl0(X).

296

Page 297: Geir Ellingsrud and John Christian Ottem October 15, 2018€¦ · between algebra and geometry: a ne varieties correspond to nitely generated k-algebras without zero-divisors, and

Chapter 19. Applications of the Riemann–Roch theorem

Given the equation of X in P2, and coordinates for the points p1, p2, we can

of course write down explicit formulas for the coordinates of p4, and they are

rational functions in the coordinates of p1, p2. This is almost enough to justify

that X is a group variety, i.e., it is an algebraic variety equipped with morphisms

m : X ×X → X satisfying the usual group axioms.

19.6 Curves of genus 2

Let X be a non-singular projective curve of genus 2. We saw one example of

such a curve earlier in this chapter, namely the curve obtained by gluing two

copies of the affine curve y2 = p(x) where p(x) is a polynomial of degree five.

The condition that X is non-singular implies that p has distinct roots.

We already saw in Chapter XX that a genus 2 curve cannot be embedded in

the projective plane P2k (since 2 is not a triagonal number). However, we show

the following:

Theorem 19.18. Any curve of genus 2 is isomorphic to a hyperelliptic curve

Here, a curve C is said to be hyperelliptic if there is a base point free linear

system of degree 2 and dimension 1. Equivalently, there exists points P,Q ∈ Xso that the invertible sheaf L = OX(P + Q) is globally generated and by two

global sections.

It is classical notation that a base point free linear system of degree d and

dimension r is called a grd. So to say that a curve is hyperelliptic is to say that it

has a g12.

Example 19.19. If g = 0, then X ' P1. Let D = 2P , then H0(D) =

kx20 +kx0x1 +x2

1, so |D| ' P2 is identified with the space of quadratic polynomials

up to scaling. If we take two quadratic polynomials q0, q1 with no common zeroes,

we get a base-point free linear system g12 ⊆ |D|.

297

Page 298: Geir Ellingsrud and John Christian Ottem October 15, 2018€¦ · between algebra and geometry: a ne varieties correspond to nitely generated k-algebras without zero-divisors, and

19.7. Curves of genus 3

Example 19.20. If g = 1 any divisor of degree 2 gives a g12 by Riemann-Roch.

Indeed, if D has degree 2 then

h0(D)− h0(K −D) = 2 + 1− g = 2

and deg(K − D) = −2 so h0(D) = 2 and hence dim |D| = 1. This D is

base-point free, since D − p has degree 1, and hence since X is not rational,

h0(D − p) = 1 = h0(D)− 1.

Example 19.21. Let X ⊆ P1 × P1 be a smooth divisor of bidegree (2, g + 1).

Then KX ' OP1×P1(0, g − 1) and X has genus g. Moreover, the projection

p2 : X → P1 is finite of degree 2, which shows that X is hyperelliptic.

The projections p1, p2 : Q → P1 give rise to a degree 2 and a degree g + 1

morphism of X to P1. Thus there exists a 2:1 morphism f : X → P1. f

corresponds to a base point free linear system on X of degree 2 and dimension 1.

Thus X is hyperelliptic.

In this example, ΩX = OQ(X) ⊗ ωQ|X = OQ(2, g + 1) ⊗ OQ(−2,−2) =

OX(0, g − 1). The latter invertible sheaf has g independent global sections so

X has genus g. Moreover KX is base point free, but not very ample, since the

corresponding morphism X → P2 is not an embedding (it maps X onto a conic).

To prove the theorem, we must produce a degree two map φ : X → P1. We

have a natural candidate: the canonical divisor KX , which has degree 2g− 2 = 2.

We claim that KX is base point free.

Note that we cannot apply Proposition 19.8 directly to prove this, since the

degree is too small. However, we can use Riemman–Roch to check directly that

the conditions in Theorem 19.7 apply. That is, we need to show that for every

point P ∈ X, we have

h0(X,KX − P ) = h0(X,KX)− 1 = 2− 1 = 1

Applying Riemann–Roch to the divisor D = P , we also get h0(P )−h0(KX−P ) =

1 + 1− 2 = 0. As P is effective, and X is not rational, we have h0(P ) = 1, and

so also h0(X,KX − P ) = 1, as we want.

19.7 Curves of genus 3

The case of curves of genus 3 is especially interesting. We have seen two examples

of curves of genus 3 so far:

Example 19.22. A plane curve X ⊆ P2 of degree d = 4 has genus 12(d− 1)(d−

2) = 3.

298

Page 299: Geir Ellingsrud and John Christian Ottem October 15, 2018€¦ · between algebra and geometry: a ne varieties correspond to nitely generated k-algebras without zero-divisors, and

Chapter 19. Applications of the Riemann–Roch theorem

Notice that

ΩX = OP2(d− 3)|X = OX(1)

so ΩX is very ample, since it is the restriction of the very ample invertible sheaf

OP2(1) on P2. Hence KX is very ample, and the corresponding morphism is

exactly the given embedding X → P2.

Example 19.23. The curves in Section 18.4 on page 271 can be chosen to have

genus g = 3. In this case, X admits a 2:1 map to P1.

Example 19.24. A curve X on the quadric surface Q ' P1 × P1 in P3 of type

(2,4) is hyperelliptic. It is a curve of degree 6 and genus 3.

Thus these examples are a bit different. The curves in the first example are

‘canonical’ whereas the curves in the other two are ‘hyperelliptic’. We show that

this distinction is a general phenomenon for curves of genus three:

Proposition 19.25. Let X be a curve of genus 3. Then there are two possibili-

ties:

(i) KX is very ample. Then X embeds as a plane curve of degree 4.

(ii) KX is not very ample. Then X is a hyperelliptic curve, and it embeds as a

(2, 4) divisor in P1 × P1. Moreover, KX ∼ 2F , where F = L1|X .

We willl deduce this from a more general result:

Theorem 19.26. Let X be a curve of genus ≥ 2. Then K is very ample if and

only if X is not hyperelliptic.

Proof. K is very ample if and only if h0(K − P −Q) = h0(K)− 2 = g − 2 for

every P,Q ∈ X. By Riemann–Roch, we compute

h0(P +Q)− h0(K − P −Q) = 2 + 1− g = 3− g

Hence K is very ample if and only if h0(P +Q) = 1 for every P,Q.

If X is hyperelliptic, then there is a map φ : X → P1, so that φ∗([1 : 0]) =

P +Q for some points P,Q ∈ X (possibly equal). Here the linear system |P +Q|is 1-dimensional, so h0(X,P +Q) = 2, and hence KX is not very ample.

If X is not hyperelliptic, we have h0(X,P +Q) = 1 for any P,Q (otherwise

it is ≥ 2, and P + Q induces a map X → P1 of degree two), and hence KX is

very ample.

We still need to check the last part of the above theorem, namely that every

hyperelliptic curve arises as a curve of type (2,4) on Q ⊆ P3.

299

Page 300: Geir Ellingsrud and John Christian Ottem October 15, 2018€¦ · between algebra and geometry: a ne varieties correspond to nitely generated k-algebras without zero-divisors, and

19.7. Curves of genus 3

We proceed as follows. Let D = P1 + · · · + P4 denote a generic degree 4

divisor on X (so P1, . . . , P4 are general points of X). By Riemann–Roch, we get

h0(D)− h0(K −D) = 4 + 1− 3 = 2

We claim that h0(K −D) = 0, so that h0(D) = 2. Note that K −D has degree

2g − 2− 4 = 0, so K −D is a divisor of degree 0. This is effective if and only

if K ∼ D. However, there is a 4-dimensional family of divisors of the form

P1 + · · · + P4, wheras the space of effective canonical divisors has dimension

dim |K| = 2. Hence if the points Pi are chosen generically, K −D will not be

effective, and hence the claim holds.

From this, we obtain a linear system |D| of dimension 1. We claim that D is

base point free. We need to show that

h0(D − P ) = h0(D)− 1 = degD + 1− 3)− 1 = 1

for every point P . Suppose not, and let P be a base point of D. Since D =

P1 + P2 + P3 + P4 we may suppose that P = P4.

By Riemann–Roch, we are done if we can show h0(K−D+P ) = 0. However,

K − D + P = K − P1 − P2 − P3. There is a 3-dimensional space of effective

divisors of the form P1 + P2 + P3 for points Pi ∈ X, but only a 2-dimensional

linear system of effective canonical divisors |K|. Hence K−D+P is not effective.

We therefore have two morphisms from our hyperelliptic curve X; f : X → P1

(induced by the g12) and g : X → P1 (induced by D). By the universal property

of the fiber product, this gives a morphism

φ = (f × g) : X → P1 ×k P1

We claim that this is a closed immersion. Let F = P +Q ∈ |g12|. The map D+F

induces the map F : X → P3, which coincides with j φ where j : P1 × P1 is the

Segre embedding. To prove the claim, it suffices to show that F is an embedding,

or equivalently that D + F is very ample.

First claim that K ∼ 2F . Since both of these divisors have degree 4 it

suffices to show that K − 2F is effective. Note that in any case h0(X, 2F ) ≥ 3,

since if H0(X,F ) = 〈x, y〉, then x2, xy, y2 are linearly independent in H0(X, 2F )

(understand why!). Now applying Riemann–Roch to D = 2F , we get

h0(2F )− h0(K − 2F ) = 4 + 1− 3 = 2

so h0(K − 2F ) ≥ 1, and K ∼ 2F as we want.

Now, to show that D + F is very ample, we need to show that

h0(X,D + F − p− q) = h0(D + F )− 2

300

Page 301: Geir Ellingsrud and John Christian Ottem October 15, 2018€¦ · between algebra and geometry: a ne varieties correspond to nitely generated k-algebras without zero-divisors, and

Chapter 19. Applications of the Riemann–Roch theorem

for any pair of points p, q ∈ X. By Riemann–Roch again, we can conclude if we

know that h0(K −D − F + p+ q) = 0. But since K ∼ 2F , we have

K −D − F + p+ q ∼ F −D + p+ q

These are divisors of degree 0, so if this is effective, we must have D ∼ F + p+ q.

However, the space of effective divisors of the form D′ + p + q with D′ ∼ F

is 3-dimensional (since |F | has dimension 1, and p and q can be chosen freely

on X). On the other hand, as we have seen, the space of divisors of the form

D = P1 + · · · + P4 is of dimension 4, so choosing D generically means that

thisF −D+ p+ q is not effective. It follows that h1(D− p− q) = h0(D+F )− 2

and hence D is very ample.

19.8 Curves of Genus 4

Recall that curves of genus g ≥ 2 split up into two disjoint classes.

(I) Hyperelliptic curves: X admits a 2:1 to P1

(II) Canonical curves: KX is very ample

Here’s an example of a genus 4 curve in P1 × P1:

Example 19.27. Consider a type (2, 5) curve C on Q ⊆ P3. Then C has degree

7 = 2 + 5 and C is hyperelliptic (because of the degree 2 map coming from

projection onto the first fact p1 : Q→ P1). A type (3, 3) curve on Q is also of

genus 4. It is a degree 6 complete intersection of Q and a cubic surface. Curves

of type (3, 3) have at least two g13’s.

In fact, using the same strategy as for g = 3, one can show that any

hyperelliptic curve of genus 4 arises this way.

19.8.1 Classifying curves of genus 4

We start with an abstract curve X of genus 4. We may assume that X is not

hyperelliptic (since in that case it embeds as a (2, 5)-divisor on P1 × P1). So

we assume that KX is very ample. Therefore we have the canonical embedding

X → Pg−1 = P3. The degree of the embedded curve is degωX = 2g − 2 = 6.

Thus we can view X as a degree 6 genus 4 curve in P3.

What are the equations of X in P3? To answer this question we use a very

powerful technique in curve theory, namely we combine Riemann–Roch with the

301

Page 302: Geir Ellingsrud and John Christian Ottem October 15, 2018€¦ · between algebra and geometry: a ne varieties correspond to nitely generated k-algebras without zero-divisors, and

19.8. Curves of Genus 4

sheaf cohomology on Pn. Twisting the ideal sheaf sequence of X by OP3(2) and

taking cohomology gives the exact sequence

0→ H0(P3, IX(2))→ H0(P3,OP3(2))→ H0(X,OX(2))→ · · ·

Note that OP3(1)|X = KX . Applying Riemann-Roch states to the divisor

D = 2KX , we get

h0(OX(2)) = deg 2KX + 1− g + h1(OX(D)) = 12 + 1− 4 + 0 = 9.

(Note that h1(OX(2)) = h0(KX − 2KX) = h0(−KX) = 0 since KX is effective).

Since h0(P3,OP3(2)) = 10 it follows that the map H0(OP3(2)) → H0(OX(2))

must have a nontrivial kernel so h0(P3, IX(2)) > 0.

The upshot of this is that we now know that X lies in some surface of degree

2. Since X is integral, this surface cannot be a union of hyperplanes. So X lies

on either a singular quadric cone Q0 = V (xy − z2) or the nonsingular quadric

surface Q = V (xy − zw).

If C lies on Q then it must have a type (a, b) which must satisfy a+ b = 6 and

(a− 1)(b− 1) = 4. The only solution is a = b = 3. Since OQ(3, 3) ' OP3(3)|Q,

this implies that C is the restriction of a divisor on P3, that is, C = Q ∩ S for a

degree 3 surface S.

The other possibility is that C lies on Q0. Computing as before, we obtain

0→ H0(OX(3))→ H0(OP3(3))→ H0(OX(3))→ · · ·

As before one sees that h0(OX(3)) = 15 and h0(OP3(3)) = 20. Thus h0(OC(3)) ≥5. Let q ∈ H0(OC(2)) be the defining equation of Q0. Then xq, yq, zq, wq ∈H0(OC(3)). But h0(OC(3)) ≥ 5 so there exists an f ∈ H0(OC(3)) so that

the global sections xq, yq, zq, wq, f are independent. Thus there is an f not in

(q). Since f 6∈ (q) we see that S = Z(f) 6⊇ Q so C ′ = S ∩ Q is a degree 6

not necessarily nonsingular or irreducible curve. Since C ⊆ S and C ⊆ Q it

follows that C ⊆ C ′. Since these are both integral curves of the same degree

degC = 6 = degC ′, we must have C = C ′. Thus in the case that C lies on Q0

we see that C is also a complete intersection C = Q0 ∩ S for some cubic surface

S.

This proves the following theorem:

Theorem 19.28. Let X be a non-singular curve of genus 4. Then either

(i) X is hyperelliptic (in which case X embeds as a (2, 5)-divisor in P1 × P1);

or

(ii) X = Q ∩ S is the intersection of a quadric surface and a cubic surface in

P3.

302

Page 303: Geir Ellingsrud and John Christian Ottem October 15, 2018€¦ · between algebra and geometry: a ne varieties correspond to nitely generated k-algebras without zero-divisors, and

Appendix A

To be deleted

A.1 L.E.S.

Theorem A.1. Let 0 → F1 → F2 → F3 → 0 be a short exact sequence of

sheaves. Then there is a long exact sequence

0→ H0(X,F1)→ H0(X,F2)→ H0(X,F3)δ−→ H1(X,F1)→ H1(X,F2)→ H1(X,F3)→ · · ·

Proof. This follows essentially by repeated use of the snake lemma. To start, we

have a commutative diagram

0 0 0

0 F1(X) F2(X) F3(X)

0 C 01 (X) C 0

2 (X) C 03 (X) 0

0 Z11 (X) Z1

2 (X) Z13 (X)

A subtle but key point is that the middle row is exact on the right, since the

short exact sequence is exact on every stalk. (In fact this observation is one of

main reasons why we use the Godement sheaves in the definition of cohomology

in the first place.)

Note that H1(X,Fi) is by definition the cokernel of the vertical map in

303

Page 304: Geir Ellingsrud and John Christian Ottem October 15, 2018€¦ · between algebra and geometry: a ne varieties correspond to nitely generated k-algebras without zero-divisors, and

A.1. L.E.S.

column i. Applying the snake lemma, we obtain an exact sequence

0→ H0(X,F1)→ H0(X,F2)→ H0(X,F3)δ−→ H1(X,F1)→ H1(X,F2)→ H1(X,F3)

We will show how to continue this sequence inductively. Applying the above

reasoning to the exact sequence

0→ Z11 → Z1

2 → Z13 → 0

we get a diagram

0 0 0

0 Z11 (X) Z1

2 (X) Z13 (X)

0 C 11 (X) C 1

2 (X) C 13 (X) 0

0 Z21 (X) Z2

2 (X) Z23 (X) 0

(where the vertical maps Z1 → C1 are induced by κ’s, as before) and a new

6-term exact sequence

H0(X,Z1)→ H0(X,Z02 )→ H0(X,Z0

3 )δ−→ H1(X,Z0

1 )→ H1(X,Z02 )→ H1(X,Z0

3 )

Note that by the recursive definition of cohomology, we have

H1(X,Z0i ) = H2(X,Fi)

for i = 1, 2, 3. We claim that this δ factors through a map

H1(X,F3) = H0(X,Z13 )/(imH0(X,C 0

3 )→ H1(F3))→ H1(X,Z01 ).

Again by the snake lemma, we get a diagram

H0(C 01 ) //

H0(C 02 ) //

H0(C 03 ) //

0

H0(Z1

1 ) //

H0(Z12 ) //

H0(Z03 ) δ //

H1(Z01 )

H1(F1) // H1(F2) // H1(F3)

δ99

304

Page 305: Geir Ellingsrud and John Christian Ottem October 15, 2018€¦ · between algebra and geometry: a ne varieties correspond to nitely generated k-algebras without zero-divisors, and

Appendix A. To be deleted

We construct the dotted arrow δ in the diagram as follows: For t ∈ H1(F3) take

a lift s ∈ H0(Z03 ) and map it to δ(s). This is well-defined, since if s′ is another

lift, then s− s′ can be lifted to an element of H0(C 03 ), and the commutativity of

the upper right square shows that δs = δs′. Moreover, another diagram chase

shows that δ(s) = 0 if and only if s is in the image of H1(F2)→ H1(F3). Hence

by identifying H1(X,Z0i ) and H2(X,Fi), we get a sequence

· · · → H1(F2)→ H1(X,F3)δ−→ H2(X,F1)→ H2(X,F2)→ H2(X,F3)

which extends the previous sequence. If we now apply the same argument to

0→ Z i1 → Z i2 → Z i3 → 0

for each i ≥ 2, we see that we can define the remaining sequence inductively.

Write C•(U ,G) for the Cech complex of a sheaf G. From the Godement

resolution F → C 1 → C 2 → · · · , we can form a double complex of sheaves by

setting:

Kp,q = Cp(U ,C q)

We can picture this as a collection of sheaves on a 2D grid, with morphisms

going both in the upwards and the right direction:

......

......

0 C2(U ,F) C2(U ,C 0) C2(U ,C 1) C2(U ,C 2) · · ·

0 C1(U ,F) C1(U ,C 0) C1(U ,C 1) C1(U ,C 2) · · ·

0 C0(U ,F) C0(U ,C 0) C0(U ,C 1) C0(U ,C 2) · · ·

0 0 0 0

So if we set p = 0, the complex K0,• is simply Γ(X,C •), whose cohomology is,

by definition, Hq(X,F). If we set q = 0, we get a complex whose (p, 0)-entry is∏U1,...,Up∈U

Hq(U1 ∩ · · · ∩ Up,F |U1∩···∩Up).

In particular, we find the Cech cohomology as the cohomology of the q = 0

305

Page 306: Geir Ellingsrud and John Christian Ottem October 15, 2018€¦ · between algebra and geometry: a ne varieties correspond to nitely generated k-algebras without zero-divisors, and

A.1. L.E.S.

column. Suppose that Hq(U1 ∩ · · · ∩ Up,F ) = 0 for all p and all q > 0. Then

the morphisms above are exact at Cp(U ,C q) with p, q ≥ 0, and we can conclude

by the following:

Proposition A.2. Suppose that K is a double complex in the first quadrant.

Assume that for all p > 0 we have Hp,•K = 0 and that for all q > 0 we have

H•,qK = 0. (In other words, the columns and rows are all exact, except in

degree 0.) Then are natural isomorphisms HpH0,•K ' HpH•,0K for p ≥ 0.

Proof. This is again a rather straightforward diagram chase. To make the chase

a little bit simpler, we can carry out the following process. We fix p ≥ 0 and

sequentially modify K, and keep track of how the cohomology of K changes if

we replace an entry by 0 and replace all arrows out of that entry by cokernels

and all arrows into that entry by kernels. That is, we replace a sequence

A B Cφ ψ

by either

A kerψ 0φ ψ

or 0 kerψ Cφ ψ

Doing this, we can modify K to get the following diagram:

0 0

K0,p K1,p 0

0 K1,p−1 K2,p−1 0

0. . . . . . Kp−1,2 Kp−1,2 0

0 Kp−1,1 Kp,1 0

0 Kp,0 0

All the arrows are isomorphisms, except possibly the left-most vertical maps, or

the lower horizontal maps. From this diagram it is clear that the cohomology

along the y-axis HpK0,∗ = K0,p is isomorphic to the cohomology along the x-axis,

HpK0,∗ = Kp,0.

306

Page 307: Geir Ellingsrud and John Christian Ottem October 15, 2018€¦ · between algebra and geometry: a ne varieties correspond to nitely generated k-algebras without zero-divisors, and

Appendix A. To be deleted

A.2 Pic(X) and H1(X,O×X)

So far, the Picard group has been defined as an abstract group defined in terms

of the tensor product of sheaves. But how do we compute it for a given X?

We can make the definition a little bit more down to earth as follows. The

essential data of L thus lies in how the OX |Ui are patched together: For each Uithere should be isomorphisms φi : LUi → OUi such that φi = φj on the overlaps

Ui ∩ Uj. Note that φj φ−1i is an isomorphism OUi∩Uj → OUi∩Uj . Now, we have

an isomorphism

HomOX (OUi∩Uj ,OUi∩Uj) ' Γ(Ui ∩ Uj,OX) (A.2.1)

(given by h 7→ h(1) and s ∈ Γ(Ui ∩ Uj,OX) gives a homomorphism given by

multiplication by s).

Recall the sheaf of units in O×X ; over an open set U ⊆ X, these where the

elements s having a multiplicative inverse s−1 ∈ Γ(U,OX). Equivalently, for each

x ∈ U , the germ sx does not lie in the maximal ideal of OX,x.

Via the correspondence (A.2.1), we find that the group of isomorphisms

OUi∩Uj → OUi∩Uj correspond exactly to the units in Γ(Ui ∩ Uj → OX). In other

words,

IsomOX (OUi∩Uj ,OUi∩Uj) ' Γ(Ui ∩ Uj,O×X)

To specify L, we must say how we glue OUi and OUj along Ui ∩ Uj – in other

words, we specify a unit sij ∈ Γ(Ui ∩ Uj,O×X) for each i, j. These elements sijcannot be chosen completely randomly, as we need to make sure that things are

compatible on the triple overlaps Ui ∩ Uj ∩ Uk. This means that they satisfy the

one constraint that on Ui ∩ Uj ∩ Uk: sijsjkski = 1, or in other words

sjks−1ik sij = 1

in Γ(Ui ∩ Uj ∩ Uk,OX). A restatement of this is that the collection sij ∈∏i,j Γ(Uij,O

×X) is a 1-cocyle! This means that we get an element in H1(X,O×X).

We can check that this is independent of the choice of cover Ui and that the map

Pic(X)→ H1(X,O×X) is in fact a group homomorphism.

Conversely, any element s ∈ H1(X,O×X) is represented by a cover U = Uiof X and cocycles sij ∈ C1(U ,O×X). The cocyle condition implies that the sijdefine isomorphisms OUi |Uij → OUi|Uij which glue the OUi together to form an

invertible sheaf. This provides an inverse to the above map. We have therefore

shown:

307

Page 308: Geir Ellingsrud and John Christian Ottem October 15, 2018€¦ · between algebra and geometry: a ne varieties correspond to nitely generated k-algebras without zero-divisors, and

A.2. Pic(X) and H1(X,O×X)

Theorem A.3. Let X be a scheme. Then

Pic(X) ' H1(X,O×X).

308