Вы находитесь на странице: 1из 863

Toric Varieties

David Cox
John Little
Hal Schenck
DEPARTMENT OF MATHEMATICS, AMHERST COLLEGE, AMHERST, MA
01002
E-mail address: dac@math.amherst.edu
DEPARTMENT OF MATHEMATICS AND COMPUTER SCIENCE, COLLEGE OF
THE HOLY CROSS, WORCESTER, MA 01610
E-mail address: little@mathcs.holycross.edu
DEPARTMENT OF MATHEMATICS, UNIVERSITY OF ILLINOIS AT URBANA-
CHAMPAIGN, URBANA, IL 61801
E-mail address: schenck@math.uiuc.edu
c (2010, David Cox, John Little and Hal Schenck
Contents
Preface vii
Notation xiii
Part I. Basic Theory of Toric Varieties 1
Chapter 1. Afne Toric Varieties 3
1.0. Background: Afne Varieties 3
1.1. Introduction to Afne Toric Varieties 10
1.2. Cones and Afne Toric Varieties 23
1.3. Properties of Afne Toric Varieties 35
Appendix: Tensor Products of Coordinate Rings 48
Chapter 2. Projective Toric Varieties 49
2.0. Background: Projective Varieties 49
2.1. Lattice Points and Projective Toric Varieties 54
2.2. Lattice Points and Polytopes 62
2.3. Polytopes and Projective Toric Varieties 74
2.4. Properties of Projective Toric Varieties 86
Chapter 3. Normal Toric Varieties 93
3.0. Background: Abstract Varieties 93
3.1. Fans and Normal Toric Varieties 105
3.2. The Orbit-Cone Correspondence 114
3.3. Toric Morphisms 125
3.4. Complete and Proper 139
iii
iv Contents
Appendix: Nonnormal Toric Varieties 150
Chapter 4. Divisors on Toric Varieties 155
4.0. Background: Valuations, Divisors and Sheaves 155
4.1. Weil Divisors on Toric Varieties 170
4.2. Cartier Divisors on Toric Varieties 176
4.3. The Sheaf of a Torus-Invariant Divisor 189
Chapter 5. Homogeneous Coordinates on Toric Varieties 195
5.0. Background: Quotients in Algebraic Geometry 195
5.1. Quotient Constructions of Toric Varieties 205
5.2. The Total Coordinate Ring 218
5.3. Sheaves on Toric Varieties 226
5.4. Homogenization and Polytopes 231
Chapter 6. Line Bundles on Toric Varieties 245
6.0. Background: Sheaves and Line Bundles 245
6.1. Ample and Basepoint Free Divisors on Complete Toric Varieties 262
6.2. Polytopes and Projective Toric Varieties 277
6.3. The Nef and Mori Cones 286
6.4. The Simplicial Case 298
Appendix: Quasicoherent Sheaves on Toric Varieties 309
Chapter 7. Projective Toric Morphisms 313
7.0. Background: Quasiprojective Varieties and Projective Morphisms 313
7.1. Polyhedra and Toric Varieties 318
7.2. Projective Morphisms and Toric Varieties 328
7.3. Projective Bundles and Toric Varieties 335
Appendix: More on Projective Morphisms 345
Chapter 8. The Canonical Divisor of a Toric Variety 347
8.0. Background: Reexive Sheaves and Differential Forms 347
8.1. One-Forms on Toric Varieties 358
8.2. Differential Forms on Toric Varieties 365
8.3. Fano Toric Varieties 379
Chapter 9. Sheaf Cohomology of Toric Varieties 387
9.0. Background: Sheaf Cohomology 387
9.1. Cohomology of Toric Divisors 398
Contents v
9.2. Vanishing Theorems I 409
9.3. Vanishing Theorems II 420
9.4. Applications to Lattice Polytopes 430
9.5. Local Cohomology and the Total Coordinate Ring 443
Part II. Topics in Toric Geometry 457
Chapter 10. Toric Surfaces 459
10.1. Singularities of Toric Surfaces and Their Resolutions 459
10.2. Continued Fractions and Toric Surfaces 467
10.3. Gr obner Fans and M
c
Kay Correspondences 485
10.4. Smooth Toric Surfaces 495
10.5. Riemann-Roch and Lattice Polygons 502
Chapter 11. Toric Resolutions and Toric Singularities 513
11.1. Resolution of Singularities 513
11.2. Other Types of Resolutions 525
11.3. Rees Algebras and Multiplier Ideals 534
11.4. Toric Singularities 546
Chapter 12. The Topology of Toric Varieties 561
12.1. The Fundamental Group 561
12.2. The Moment Map 568
12.3. Singular Cohomology of Toric Varieties 577
12.4. The Cohomology Ring 592
12.5. The Chow Ring and Intersection Cohomology 612
Chapter 13. Toric Hirzebruch-Riemann-Roch 623
13.1. Chern Characters, Todd Classes, and HRR 624
13.2. Brions Equalities 632
13.3. Toric Equivariant Riemann-Roch 641
13.4. The Volume Polynomial 654
13.5. The Khovanskii-Pukhlikov Theorem 663
Appendix: Generalized Gysin Maps 672
Chapter 14. Toric GIT and the Secondary Fan 677
14.1. Introduction to Toric GIT 677
14.2. Toric GIT and Polyhedra 685
14.3. Toric GIT and Gale Duality 699
vi Contents
14.4. The Secondary Fan 712
Chapter 15. Geometry of the Secondary Fan 725
15.1. The Nef and Moving Cones 725
15.2. Gale Duality and Triangulations 734
15.3. Crossing a Wall 747
15.4. Extremal Contractions and Flips 762
15.5. The Toric Minimal Model Program 772
Appendix A. The History of Toric Varieties 787
A.1. The First Ten Years 787
A.2. The Story Since 1980 794
Appendix B. Computational Methods 797
B.1. The Rational Quartic 798
B.2. Polyhedral Computations 799
B.3. Normalization and Normaliz 803
B.4. Sheaf Cohomology and Resolutions 805
B.5. Sheaf Cohomology on the Hirzebruch Surface H
2
806
B.6. Resolving Singularities 808
B.7. Intersection Theory and Hirzebruch-Riemann-Roch 809
B.8. Anticanonical Embedding of a Fano Toric Variety 810
Appendix C. Spectral Sequences 811
C.1. Denitions and Basic Properties 811
C.2. Spectral Sequences Appearing in the Text 814
Bibliography 817
Index 831
Preface
The study of toric varieties is a wonderful part of algebraic geometry. There are
elegant theorems and deep connections with polytopes, polyhedra, combinatorics,
commutative algebra, symplectic geometry, and topology. Toric varieties also have
unexpected applications in areas as diverse as physics, coding theory, algebraic
statistics, and geometric modeling. Moreover, as noted by Fulton [105], toric
varieties have provided a remarkably fertile testing ground for general theories.
At the same time, the concreteness of toric varieties provides an excellent context
for someone encountering the powerful techniques of modern algebraic geometry
for the rst time. Our book is an introduction to this rich subject that assumes only
a modest background yet leads to the frontier of this active area of research.
Brief Summary. The text covers standard material on toric varieties, including:
(a) Convex polyhedral cones, polytopes, and fans.
(b) Afne, projective, and abstract toric varieties.
(c) Complete toric varieties and proper toric morphisms.
(d) Weil and Cartier divisors on toric varieties.
(e) Cohomology of sheaves on toric varieties.
(f) The classical theory of toric surfaces.
(g) The topology of toric varieties.
(h) Intersection theory on toric varieties.
These topics are discussed in earlier texts on the subject, such as [93], [105] and
[218]. One difference is that we provide more details, with numerous examples,
gures, and exercises to illustrate the concepts being discussed. We also provide
background material when needed. In addition, we cover a large number of topics
previously available only in the research literature.
vii
viii Preface
The Fifteen Chapters. To give you a better idea of what is in the book, we now
highlight a few topics from each chapter.
Chapters 1, 2 and 3 cover the basic material mentioned in items (a)(c) above.
The toric varieties encountered include:
The afne toric variety Y
A
of a nite set A M Z
n
(Chapter 1).
The afne toric variety U

of a polyhedral cone N
R
R
n
(Chapter 1).
The projective toric variety X
A
of a nite set A M Z
n
(Chapter 2).
The projective toric variety X
P
of a lattice polytope P M
R
R
n
(Chapter 2).
The abstract toric variety X

of a fan in N
R
R
n
(Chapter 3).
Chapter 4 introduces Weil and Cartier divisors on toric varieties. We compute
the class group and Picard group of a toric variety and dene the sheaf O
X

(D)
associated to a Weil divisor D on a toric variety X

.
Chapter 5 shows that the classical construction P
n
=

C
n+1
`0

/C

can be
generalized to any toric variety X

. The homogeneous coordinate ring C[x


0
, . . . , x
n
]
of P
n
also has a toric generalization, called the total coordinate ring of X

.
Chapter 6 relates Cartier divisors to invertible sheaves on X

. We introduce
ample, basepoint free, and nef divisors and discuss their relation to convexity. The
stucture of the nef cone and its dual, the Mori cone, are described in detail, as is
the intersection pairing between divisors and curves.
Chapter 7 extends the relation between polytopes and projective toric varieties
to a relation between polyhedra and projective toric morphisms : X

. We
also discuss projective bundles over a toric variety and use these to classify smooth
projective toric varieties of Picard number 2.
Chapter 8 relates Weil divisors to reexive sheaves of rank one and denes
Zariski p-forms. For p = dim X, this gives the canonical sheaf
X
and canonical
divisor K
X
. In the toric case we describe these explicitly and study the relation be-
tween reexive polytopes and Gorenstein Fano toric varieties, meaning that K
X

is ample. We nd the 16 reexive polygons in R


2
(up to equivalence) and note the
relation [PM[ +[P

N[ = 12 for a reexive polygon P and its dual P

.
Chapter 9 is about sheaf cohomology. We give two methods for computing
sheaf cohomology on a toric variety and prove a dizzying array of cohomology
vanishing theorems. Applications range from showing that normal toric varieties
are Cohen-Macaulay to the Dehn-Sommerville equations for a simple polytope and
counting lattice points in multiples of a polytope via the Ehrhart polynomial.
Chapter 10 studies toric surfaces, where we add a few twists to this classical
subject. After using Hirzebruch-Jung continued fractions to compute the minimal
resolution of a toric surface singularity, we discuss the toric meaning of ordinary
continued fractions. We then describe unexpected connections with Gr obner fans
and the M
c
Kay correspondence. Finally, we use the Riemann-Roch theorem on a
Preface ix
smooth complete toric surface to explain the mysterious appearance of the number
12 in Chapter 8 when counting lattice points in reexive polygons.
Chapter 11 proves the existence of toric resolutions of singularities for toric
varieties of all dimensions. This is more complicated than for surfaces because of
the existence of toric ips and ops. We consider simple normal crossing, crepant,
log, and embedded resolutions and study how Rees algebras and multiplier ideals
can be applied in the resolution problem. We also discuss toric singularities and
show that a fan is simplicial if and only if X

has at worst nite quotient singular-


ities and hence is rationally smooth. We also explain what canonical and terminal
singularities mean in the toric context.
Chapters 12 and 13 describe the singular and equivariant cohomology of a
complete simplicial toric variety X

and prove the Hirzebruch-Riemann-Roch and


equivariant Riemann-Roch theorems when X

is smooth. We compute the funda-


mental group of X

and study the moment map, with a brief mention of topological


models of toric varieties and connections with symplectic geometry. We describe
the Chow ring and intersection cohomology of a complete simplicial toric variety.
After proving Riemann-Roch, we give applications to the volume polynomial and
lattice point enumeration in polytopes.
Chapters 14 and 15 explore the rich connections that link geometric invariant
theory, the secondary fan, the nef and moving cones, Gale duality, triangulations,
wall crossings, ips, extremal contractions, and the toric minimal model program.
Appendices. The book ends with three appendices:
Appendix A: The History of Toric Varieties.
Appendix B: Computational Methods.
Appendix C: Spectral Sequences.
Appendix A surveys the history of toric geometry since its origins in the early
1970s. It is fun to see how the concepts and terminology evolved. Appendix B
discusses some of the software packages for toric geometry and gives examples
to illustrate what they can do. Appendix C gives a brief introduction to spectral
sequences and describes the spectral sequences used in Chapters 9 and 12.
Prerequisites. We assume that the reader is familiar with the material covered in
basic graduate courses in algebra and topology, and to a somewhat lesser degree,
complex analysis. In addition, we assume that the reader has had some previous
experience with algebraic geometry, at the level of any of the following texts:
Ideals, Varieties and Algorithms by Cox, Little and OShea [69].
Introduction to Algebraic Geometry by Hassett [133].
Elementary Algebraic Geometry by Hulek [151].
Undergraduate Algebraic Geometry by Reid [238].
x Preface
Computational Algebraic Geometry by Schenck [246].
An Invitation to Algebraic Geometry by Smith, Kahanp a a, Kek al ainen and
Traves [253].
Chapters 9 and 12 assume knowledge of some basic algebraic topology. The books
by Hatcher [135] and Munkres [210] are useful references here.
Readers who have studied more sophisticated algebraic geometry texts such as
Harris [130], Hartshorne [131], or Shafarevich [245] certainly have the background
needed to read our book. For readers with a more modest background, an important
prerequisite is a willingness to absorb a lot of algebraic geometry.
Background Sections. Since we do not assume a complete knowledge of algebraic
geometry, Chapters 19 each begin with a background section that introduces the
denitions and theorems from algebraic geometry that are needed to understand
the chapter. References where proofs can be found are provided. The remaining
chapters do not have background sections. For some of those chapters, no further
background is necessary, while for others, the material is more sophisticated and
the requisite background is given by careful references to the literature.
What Is Omitted. We work exclusively with varieties dened over the complex
numbers C. This means that we do not consider toric varieties over arbitrary elds
(see [92] for a treatment of this topic), nor do we consider toric stacks (see [39] for
an introduction). Moreover, our viewpoint is primarily algebro-geometric. Thus,
while we hint at some of the connections with symplectic geometry and topology in
Chapter 12, we do not do justice to this side of the story. Even within the algebraic
geometry of toric varieties, there are many topics we have had to omit, though we
provide some references that should help readers who want to explore these areas.
We have also omitted any discussion of how toric varieties are used in physics and
applied mathematics. Some pointers to the literature are given in our discussion of
the recent history of toric varieties in A.2 of Appendix A.
The Structure of the Text. We number theorems, propositions, and equations
based on the chapter and the section. Thus 3.2 refers to Section 2 of Chapter 3,
and Theorem 3.2.6, equation (3.2.6) and Exercise 3.2.6 all appear in this section.
Denitions, theorems, propositions, lemmas, remarks, and examples are numbered
together in one sequence within each section.
Some individual chapters have appendices. Within a chapter appendix the
same numbering system is used, except that the section number is a capital A.
This means that Theorem 3.A.3 is in the appendix to Chapter 3. On the other hand,
the three appendices at the end of the book are treated in the numbering system as
chapters A, B, and C. Thus Denition C.1.1 is in the rst section of Appendix C.
The end (or absence) of a proof is indicated by , and the end of an example
is indicated by .
Preface xi
For the Instructor. There is much more material here than you can cover in any
one-semester graduate course, probably more than you can cover in a full year
in most cases. So choices will be necessary depending on the background and
the interests of the student audience. We think it is reasonable to expect to cover
most of Chapters 16, 8 and 9 in a one-semester course where the students have
a minimal background in algebraic geometry. More material can be covered, of
course, if the students know more algebraic geometry. If time permits, you can
use toric surfaces (Chapter 10) to illustrate the power of the basic material and
introduce more advanced topics such as the resolution of singularities (Chapter 11)
and the Riemann-Roch theorem (Chapter 13).
Finally, we emphasize that the exercises are extremely important. We have
found that when the students work in groups and present their solutions, their en-
gagement with the material increases. We encourage instructors to consider using
this strategy.
For the Student. The book assumes that you will be an active reader. This means
in particular that you should do tons of exercisesthis is the best way to learn
about toric varieties. If you have a modest background in algebraic geometry, then
reading the book requires a commitment to learn both toric varieties and algebraic
geometry. It will be a lot of work but is worth the effort. This is a great subject.
Send Us Feedback. We greatly appreciate hearing from instructors, students, or
general readers about what worked and what didnt. Please notify one or all of us
about any typographical or mathematical errors you might nd.
Acknowledgements. We would like to thank to Dan Abramovich, Matt Baker,
Matthias Beck, TomBraden, Volker Braun, Sandra Di Rocco, Dan Edidin, Matthias
Franz, Dan Grayson, Paul Hacking, J urgen Hausen, Al Kasprzyk, Diane Maclagan,
Anvar Mavlyutov, Uwe Nagel, Andrey Novoseltsev, Sam Payne, Matthieu Ram-
baud, Raman Sanyal, Thorsten Rahn, Greg Smith, Bernd Sturmfels, Zach Teitler,
Michele Vergne, Mark Walker, G unter Ziegler, and Dylan Zwick for many helpful
conversations and emails as we worked on the book.
We would also like to thank all the participants of the 2009 MSRI Summer
Graduate Student Workshop on Toric Varieties and especially Dustin Cartwright
and Daniel Erman, our able assistants.
Finally we thank Ina Mette for her support and advice and Barbara Beeton for
her help with L
A
T
E
X.
September 2010 David Cox
John Little
Hal Schenck
Notation
The notation used in the book is organized by topic. The number in parentheses at
the end of an entry indicates the chapter in which the notation rst appears.
Basic Sets
Z, Q, R, C integers, rational numbers, real numbers, complex numbers
N semigroup of nonnegative integers 0, 1, 2, . . .
The Torus
C

multiplicative group of nonzero complex numbers C` 0 (1)


(C

)
n
standard n-dimensional torus (1)
M,
m
character lattice of a torus and character of m M (1)
N,
u
lattice of one-parameter subgroups of a torus and
one-parameter subgroup of u N (1)
T
N
torus N
Z
C

= Hom
Z
(M, C

) associated to N and M (1)


M
R
, M
Q
vector spaces M
Z
R, M
Z
Q built from M (1)
N
R
, N
Q
vector spaces N
Z
R, N
Z
Q built from N (1)
'm, u` pairing of m M or M
R
with u N or N
R
(1)
Hyperplanes and Half-Spaces
H
m
hyperplane in N
R
dened by 'm, ` = 0, m M
R
` 0 (1)
H
+
m
half-space in N
R
dened by 'm, ` 0, m M
R
` 0 (1)
H
u,b
hyperplane in M
R
dened by ', u` = b, u N
R
` 0 (2)
H
+
u,b
half-space in M
R
dened by ', u` b, u N
R
` 0 (2)
xiii
xiv Notation
Cones
Cone(S) convex cone generated by S (1)
rational convex polyhedral cone in N
R
(1)
Span() subspace spanned by (1)
dim dimension of (1)

dual cone of (1)


Relint() relative interior of (1)
Int() interior of when Span() = N
R
(1)

set of m M
R
with 'm, ` = 0 (1)
_, is a face or proper face of (1)

face of

dual to , equals

(1)
Rays
1-dimensional strongly convex cone (a ray) in N
R
(1)
u

minimal generator of N, a rational ray in N


R
(1)
(1) rays of a strongly convex cone in N
R
(1)
Lattices
ZA lattice generated by A (1)
Z

A elements

s
i=1
a
i
m
i
ZA with

s
i=1
a
i
= 0 (2)
N

sublattice Z( N) = Span() N (3)


N() quotient lattice N/N

(3)
M() dual lattice of N(), equals

M (3)
Fans
fan in N
R
(2,3)
(r) r-dimensional cones of (3)

max
maximal cones of (3)
Star() star of , a fan in N() (3)

() star subdivision of for (3)

(v) star subdivision of for v [[ N primitive (11)


Polytopes and Polyhedra

n
standard n-simplex in R
n
(2)
Conv(S) convex hull of S (1)
dim P dimension of a polyhedron P (2)
Q _P, Q P Q is a face or proper face of P (2)
P

dual or polar of a polytope (2)


A+B Minkowski sum (2)
k P multiple of a polytope or polyhedron (2)
Notation xv
Cones Built From Polyhedra
C
v
Cone(Pv) for a vertex v of a polytope or polyhedron (2)

Q
cone of a face Q _P in the normal fan
P
(2)

P
normal fan of a polytope or polyhedron P (2)
C(P) cone over a polytope or polyhedron (1)
S
P
semigroup algebra of C(P) (MZ) (7)
Combinatorics and Lattice Points of Polytopes
f
i
number of i-dimesional faces of P (9)
h
p

n
i=p
(1)
ip

i
p

f
i
, equals Betti number b
2p
(X
P
) when P simple (9)
L(P) number of lattice points of a lattice polytope (9)
L

(P) number of interior lattice points of a lattice polytope (9)


Ehr
P
() Ehrhart polynomial of a lattice polytope (9)
Ehr
p
P
() p-Ehrhart polynomial of a lattice polytope (9)
Semigroups
S, C[S] afne semigroup and its semigroup algebra (1)
NA afne semigroup generated by A (1)
S

= S
,N
afne semigroup

M (1)
H Hilbert basis of S

(1)
Rings
R
f
, R
S
, R
p
localization of R at f, a multiplicative set S, a prime ideal p (1)
R

integral closure of the integral domain R (1)

R completion of local ring R (1)


R
C
S tensor product of rings over C (1)
R
G
ring of invariants of G acting on R (1,5)
R[a] Rees algebra of an ideal a R (11)
R
[]
Veronese subring of a graded ring R (14)
Specic Rings
C[x
1
, . . . , x
n
] polynomial ring in n variables (1)
C[[x
1
, . . . , x
n
]] formal power series ring in n variables (1)
C[x
1
1
, . . . , x
1
n
] ring of Laurent polynomials (1)
I(V) ideal of an afne or projective variety (1,2)
C[V] coordinate ring of an afne or projective variety (1,2)
C[V]
d
graded piece in degree d when V is projective (2)
C(V) eld of rational functions when V is irreducible (1)
O
V,p
, m
V,p
local ring of a variety at a point and its maximal ideal (1)
xvi Notation
Varieties
V(I) afne or projective variety of an ideal (1,2)
V
f
subset of an afne variety V where f = 0 (1)
S Zariski closure of S in a variety (1,3)
T
p
(X) Zariski tangent space of a variety at a point (1,3)
dim X, dim
p
X dimension of a variety and dimension at a point (1,3)
Spec(R) afne variety of coordinate ring R (1)
Proj(S) projective variety of graded ring S (7)
X Y product of varieties (1,3)
X
S
Y ber product of varieties (3)

X afne cone of a projective variety X (2)


Toric Varieties
Y
A
, X
A
afne and projective toric variety of A M (1,2)
U

=U
,N
afne toric variety of a cone N
R
(1)
X

= X
,N
toric variety of a fan in N
R
(3)
X
P
projective toric variety of a lattice polytope or polyhedron (2.7)
lattice homomorphismof a toric morphism : X
1
X
2
(1,3)

R
real extension of (1)

distinguished point of U

(3)
O() torus orbit corresponding to (3)
V() = O() closure of orbit of , toric variety of Star() (3)
U
P
afne toric variety of recession cone of a polyhedron (7)
U

afne toric variety of a fan with convex support (7)


Specic Varieties
C
n
, P
n
afne and projective n-dimensional space (1,2)
P(q
0
, . . . , q
n
) weighted projective space (2)

C
d
, C
d
rational normal cone and curve (1,2)
Bl
0
(C
n
) blowup of C
n
at the origin (3)
Bl
V()
(X

) blowup of X

along V(), toric variety of

() (3)
H
r
Hirzebruch surface (3)
S
a,b
rational normal scroll (3)
Total Coordinate Ring
S total coordinate ring of X

(5)
x

variable in S corresponding to (1) (5)


S

graded piece of S in degree Cl(X

) (5)
deg(x

) degree in Cl(X

) of a monomial in S (5)
Notation xvii
x

monomial

/ (1)
x

for (5)
B() irrelevant ideal of S, generated by the x

(5)
x
m
Laurent monomial

x
m,u

, m M (5)
x
m,D
homogenization of
m
, m P
D
M (5)
x
F
facet variable of a facet F P (5)
x
m,P
P-monomial associated to m PM (5)
x
v,P
vertex monomial associated to vertex v PM (5)
M graded S-module (5)
M() shift of M by Cl(X

) (5)
Quotient Construction
X/G good geometric quotient (5)
X//G good categorical quotient (5)
Z() exceptional set in quotient construction, equals V(B()) (5)
G group in quotient construction, equals Hom
Z
(Cl(X

), C

) (5)
Divisors
O
X,D
local ring of a variety at a prime divisor (4)

D
discrete valuation of a prime divisor D (4)
div( f ) principal divisor of a rational function (4)
D E linear equivalence of divisors (4)
D 0 effective divisor (4)
Div
0
(X) group of principal divisors on X (4)
Div(X) group of Weil divisors on X (4)
CDiv(X) group of Cartier divisors on X (4)
Cl(X) divisor class group of a normal variety X (4)
Pic(X) Picard group of a normal variety X (4)
Pic(X)
R
Pic(X)
Z
R (6)
Supp(D) support of a divisor (4)
D[
U
restriction of a divisor to an open set (4)
(U
i
, f
i
) local data of a Cartier divisor on X (4)
[D[ complete linear system of D (6)
D, D round down and round up of a Q-divisor (9)
Torus-Invariant Divisors
D

= O() torus-invariant prime divisor on X

of ray (1) (4)


D
F
torus-invariant prime divisor on X
P
of facet F P (4)
Div
TN
(X

) group of torus-invariant Weil divisors on X

(4)
CDiv
TN
(X

) group of torus-invariant Cartier divisors on X

(4)
m

Cartier data of a torus-invariant Cartier divisor on X

(4)
xviii Notation
D
P
Cartier divisor of a polytope or polyhedron (4,7)
P
D
polyhedron of a torus-invariant divisor (4)
X
D
toric variety of a basepoint free divisor (6)

D
fan of X
D
(6)

D pullback of a Cartier divisor (6)


Support Functions

D
support function of a Cartier divisor (4)

P
support function of a polytope or polyhedron (4)
SF() support functions for (4)
SF(, N) support functions for integral with respect to N (4)
Sheaves
(U, F) sections of a sheaf over an open set (4)
F[
U
restriction of a sheaf to an open set (4)
F
p
stalk of a sheaf at a point (6)
F
OX
G tensor product of sheaves of O
X
-modules (6)
Hom
OX
(F, G ) sheaf of homomorphisms (6)
F

dual sheaf of F, equals Hom


OX
(F, O
X
) (6)
f

F direct image sheaf


Specic Sheaves
O
X
structure sheaf of a variety X (3)
O

X
sheaf of invertible elements of O
X
(4)
K
X
constant sheaf of rational functions for X irreducible (6)
O
X
(D) sheaf of a Weil divisor D on X (4)
I
Y
ideal sheaf of a subvariety Y X (3)

M sheaf on Spec(R) of an R-module M (4)

M sheaf on X

of the graded S-modules M (5)


O
X
() sheaf of the S-module S() (5)
Vector Bundles and Locally Free Sheaves
L, E invertible sheaf (line bundle) and locally free sheaf (6)
: V X vector bundle (6)
: V
L
X rank 1 vector bundle of an invertible sheaf L (6)
f

L pullback of an invertible sheaf (6)

L,W
map to projective space determined by W (X, L) (6)
P(V), P(E ) projective bundle of vector bundle or locally free sheaf (7)
D fan for rank 1 vector bundle V
L
for L =O
X
(D) (7)
Notation xix
Intersection Theory
deg(D) degree of a divisor on a smooth complete curve (6)
D C intersection product of Cartier divisor and complete curve (6)
D D

, C C

numerically equivalent Cartier divisors and complete curves (6)


N
1
(X), N
1
(X) (CDiv(X)/)
Z
R and (proper 1-cycles on X/)
Z
R (6)
Nef(X) cone in N
1
(X) generated by nef divisors (6)
Mov(X) moving cone of a variety X in N
1
(X) (15)
Eff(X) pseudoeffective cone of a variety X in N
1
(X) (15)
NE(X) cone in N
1
(X) generated by complete curves (6)
NE(X) Mori cone, equals the closure of NE(X) (6)
Differential Forms and Sheaves

R/C
module of K ahler differentials of a C-algebra R (8)

1
X
, T
X
cotangent and tangent sheaves of a variety X (8)
I
Y
/I
2
Y
, N
Y/X
conormal and normal sheaves of Y X (8)

p
X
,

p
X
sheaves of p-forms and Zariski p-forms on X (8)
K
X
,
X
canonical divisor and canonical sheaf

n
X
, n = dim X (8)

1
X
(logD) sheaf of 1-forms with logarithmic poles on D (8)
Sheaf Cohomology
H
0
(X, F) global sections (X, F) of a sheaf F on X (9)
H
p
(X, F) p-th sheaf cohomology group of a sheaf F on X (9)
R
p
f

F higher direct image sheaf (9)


Ext
p
OX
(G, F) Ext groups of sheaves of O
X
-modules G, F (9)

(U , F)

Cech complex for sheaf cohomology (9)
(F) Euler characteristic of F, equals

p
(1)
p
dim H
p
(X, F) (9)
Sheaf Cohomology of a Toric Variety
H
p
(X

, L)
m
graded piece of sheaf cohomology of L =O
X
(D) for m M (9)
V
D,m
, V
supp
D,m
subsets of [[ used to compute H
p
(X

, L)
m
(9)
Local Cohomology
H
p
I
(M) p-th local cohomology of an R-module M for the ideal I R (9)

(f, M)

Cech complex for local cohomology when I ='f` (9)
Ext
p
R
(N, M) Ext groups of R-modules N, M (9)
Resolution of Singularities
X
sing
singular locus of a variety (11)
Exc() exceptional locus of a resolution of singularities (11)
xx Notation
.(c I) multiplier ideal sheaf (11)
(X, D) log pair, D =

i
a
i
D
i
, a
i
[0, 1] Q (11)
Singularities of Toric Varieties
mult() multiplicity of a simplicial cone, equals [N

: Zu
1
+ +Zu
d
] (6,11)
P

parallelotope of a simplicial cone, equals

i
u
i
[ 0
i
< 1 (11)

polytope related to canonical and terminal singularities of U

(11)

the polyhedron Conv( N ` 0) (10, 11)

can
fan over bounded faces of

, reduces to canonical singularities (11)


Topology of a Toric Variety
N

sublattice of N generated by [[ N (12)

1
(X

) fundamental group of X

, isomorphic to N/N

(12)
S
N
real torus N
Z
S
1
= Hom
Z
(M, S
1
) (S
1
)
n
(12)
(X

)
0
nonnegative real points of a toric variety (12)
f , algebraic and symplectic moment maps X
P
M
R
(12)

symplectic moment map C


(1)
Cl(X

)
R
(12)
Singular Homology and Cohomology
H
i
(X, R) ith singular cohomology of X with coefcients in a ring R (9)

H
i
(X, R) i-th reduced cohomology of X (9)
H
i
c
(X, R) ith cohomology of X with compact supports (12)
H
i
(X, R) ith singular homology of X (12)
H
BM
i
(X, R) ith Borel-Moore homology of X (13)
b
i
(X) ith Betti number of X, equals dim H
i
(X, Q) (12)
e(X) Euler characteristic of X, equals

i
(1)
i
b
i
(X) (9,10,12)
, cap and cup products (12)
H

(X, R) cohomology ring

p
H
p
(X, R) under cup product (12)
[W] cohomology class of a subvariety W in H
2n2k
(X, Q) (12,13)
[W]
r
rened cohomology class of W in H
2n2k
(X, X `W, Q) (12,13)
f
!
generalized Gysin map (13)

X
integral

X
: H

(X, Q) Q, equals Gysin map of X pt (12,13)


Equivariant Cohomology for a Group Action
EG a contractible space on which G acts freely (12)
BG the quotient EG/G (12)
EG
G
X quotient of EGX modulo relation (e g, x) (e, g x) (12)
H

G
(X, R) equivariant cohomology ring, equals H

(EG
G
X, R) (12)

G
, (
G
)
Q
integral and rational equivariant cohomology ring of a point (12)
X
G
xed point set for action of G on X (12)
Notation xxi
Equivariant Cohomology for a Torus Action
Sym
Z
(M) symmetric algebra of M over Z (12)
Sym
Q
(M) rational symmetric algebra on M, equals Sym
Z
(M)
Z
Q (12)
s isomorphisms : Sym
Q
(M) (
T
)
Q
(12)
[D]
T
equivariant cohomology class of a T-invariant divisor D (12)

X
eq equivariant integral

X
eq : H

T
(X, Q) (
T
)
Q
(13)

T
(X, Q) completion

k=0
H
k
T
(X, Q) of equivariant cohomology of X (13)

completion of the equivariant cohomology of a point (13)


Chow Groups and the Chow Ring
A
k
(X) Chow group of k-cycles modulo rational equivalence (12)
A
k
(X) codimension k cycles modulo rational equivalence (12)
A

(X) integral Chow ring of X smooth and complete (12)


A

(X)
Q
rational Chow ring of X quasismooth and complete (12)
Intersection Cohomology
IH
p
i
(X) ith intersection homology of X for perversity p (12)
IH
i
(X) ith intersection cohomology of X for middle perversity (12)
IH
i
(X)
Q
ith rational intersection cohomology of X (12)
Cohomology Ring of a Complete Simplicial Toric Variety
I Stanley-Reisner ideal of the fan , ideal in Q[x
1
, . . . , x
r
] (12)
. ideal '

r
i=1
'm, u
i
`x
i
[ m M` Q[x
1
, . . . , x
r
] (12)
R
Q
() Jurkiewicz-Danilov ring Q[x
1
, . . . , x
r
]/(I +.) H

(X

, Q) (12)
SR
Q
() Stanley-Reisner ring Q[x
1
, . . . , x
r
]/I H

T
(X

, Q) (12)
Hirzebruch-Riemann-Roch
c
i
(E ) ith Chern class of a locally free sheaf E (13)
ch(L) Chern character of a line bundle L (13)
Td(X) Todd class of the variety X (13)
B
k
kth Bernoulli number (13)
c
i
= c
i
(T
X
) ith Chern class of the tangent bundle (13)
T
i
ith Todd polynomial in the c
i
(13)
K(X) Grothendieck group of classes of coherent sheaves on X (13)

T
(L) equivariant Euler characteristic (13)

(L) local contribution of (n) to


T
(L) (13)
ch
T
(L) equivariant Chern character of L (13)
Td
T
(X) equivariant Todd class of X (13)
Todd(x) formal Todd differential operator for the variable x (13)
xxii Notation
Brions Equalities
Z[M] integral semigroup algebra of M (13)
Z[[M]] formal semigroup module of M, formal sums

mM
a
m

m
(13)
Z[[M]]
Sum
summable elements in Z[[M]] (13)
o( f ) sum of an element f Z[[M]]
Sum
(13)
(L)

mM

n
i=0
dimH
i
(X, L)
m

m
Z[[M]] (13)
Geometric Invariant Theory

G character group of algebraic subgroup G (C

)
r
(14)
L

sheaf of sections of rank 1 line bundle on C


r
for character

G (14)
(C
r
)
ss

, (C
r
)
s

semistable and stable points for (14)


R

graded ring

d=0
(C
r
, L

d )
G
(14)
C
r
//

G GIT quotient of C
r
by G for , equals Proj(R

) = (C
r
)
ss

//G (14)
B() irrelevant ideal of (14)
Z() exceptional set of , equals V(B()) (14)
P

, P
a
polyhedra in R
r
and M
R
for =
a
(14)
F
i,
, F
i,a
ith virtual facet of P

, P
a
(14)
The Secondary Fan
, lists of r vectors in

G
R
and N
R
(14)
C

, C

cones generated by and (14)

,I

,
,I

GKZ cones determined by , I

(14)
B(, I

) irrelevant ideal determined by , I

(14)

GKZ
secondary fan of (14)
Mov
GKZ
moving cone of the secondary fan (15)
P
GKZ
secondary polytope, normal fan is
GKZ
(15)
Toric Minimal Model Program
1 extremal ray of the Mori cone (15)
D 1 intersection product D C for [C] 1` 0 (15)
f

D birational transform of a divisor by a birational map (15)


J

, J
+
index sets determined by a wall relation (15)

,
+
fans determined by a wall relation (15)

,
+
toric morphisms determined by a wall relation (15)
Miscellaneous

d
multiplicative group of dth roots of unity in C (1)
[a
1
, . . . , a
s
] ordinary continued fraction of a rational number (10)
[[b
1
, . . . , b
r
]] Hirzebruch-Jung continued fraction of a rational number (10)
Part I. Basic Theory of
Toric Varieties
Chapters 1 to 9 introduce the theory of toric varieties. This part of the
book assumes only a minimal amount of algebraic geometry, at the level
of Ideals, Varieties and Algorithms [69]. Each chapter begins with a back-
ground section that develops the necessary algebraic geometry.
1
Chapter 1
Afne Toric Varieties
1.0. Background: Afne Varieties
We begin with the algebraic geometry needed for our study of afne toric varieties.
Our discussion assumes Chapters 15 and 9 of [69].
Coordinate Rings. An ideal I S =C[x
1
, . . . , x
n
] gives an afne variety
V(I) =p C
n
[ f (p) = 0 for all f I
and an afne variety V C
n
gives the ideal
I(V) =f S [ f (p) = 0 for all p V.
By the Hilbert basis theorem, an afne variety V is dened by the vanishing of
nitely many polynomials in S, and for any ideal I, the Nullstellensatz tells us that
I(V(I)) =

I = f S [ f

I for some 1 since C is algebraically closed.


The most important algebraic object associated to V is its coordinate ring
C[V] = S/I(V).
Elements of C[V] can be interpreted as the C-valued polynomial functions on V.
Note that C[V] is a C-algebra, meaning that its vector space structure is compatible
with its ring structure. Here are some basic facts about coordinate rings:
C[V] is an integral domain I(V) is a prime ideal V is irreducible.
Polynomial maps (also called morphisms) : V
1
V
2
between afne varieties
correspond to C-algebra homomorphisms

: C[V
2
] C[V
1
], where

(g) =
g for g C[V
2
].
Two afne varieties are isomorphic if and only if their coordinate rings are
isomorphic C-algebras.
3
4 Chapter 1. Afne Toric Varieties
A point p of an afne variety V gives the maximal ideal
f C[V] [ f (p) = 0 C[V],
and all maximal ideals of C[V] arise this way.
Coordinate rings of afne varieties can be characterized as follows (Exercise 1.0.1).
Lemma 1.0.1. A C-algebra R is isomorphic to the coordinate ring of an afne
variety if and only if R is a nitely generated C-algebra with no nonzero nilpotents,
i.e., if f R satises f

= 0 for some 1, then f = 0.


To emphasize the close relation between V and C[V], we sometimes write
(1.0.1) V = Spec(C[V]).
This can be made canonical by identifying V with the set of maximal ideals of
C[V] via the fourth bullet above. More generally, one can take any commutative
ring R and dene the afne scheme Spec(R). The general denition of Spec uses
all prime ideals of R, not just the maximal ideals as we have done. Thus some
authors would write (1.0.1) as V = Specm(C[V]), the maximal spectrum of C[V].
Readers wishing to learn about afne schemes should consult [90] and [131].
The Zariski Topology. An afne variety V C
n
has two topologies we will use.
The rst is the classical topology, induced from the usual topology on C
n
. The
second is the Zariski topology, where the Zariski closed sets are subvarieties of V
(meaning afne varieties of C
n
contained in V) and the Zariski open sets are their
complements. Since subvarieties are closed in the classical topology (polynomials
are continuous), Zariski open subsets are open in the classical topology.
Given a subset S V, its closure S in the Zariski topology is the smallest
subvariety of V containing S. We call S the Zariski closure of S. It is easy to give
examples where this differs from the closure in the classical topology.
Afne Open Subsets and Localization. Some Zariski open subsets of an afne
variety V are themselves afne varieties. Given f C[V] `0, let
V
f
=p V [ f (p) = 0 V.
Then V
f
is Zariski open in V and is also an afne variety, as we now explain.
Let V C
n
have I(V) =' f
1
, . . . , f
s
` and pick g C[x
1
, . . . , x
n
] representing f .
Then V
f
=V `V(g) is Zariski open in V. Now consider a new variable y and let
W = V( f
1
, . . . , f
s
, 1gy) C
n
C. Since the projection map C
n
C C
n
maps
W bijectively onto V
f
, we can identify V
f
with the afne variety W C
n
C.
When V is irreducible, the coordinate ring of V
f
is easy to describe. Let C(V)
be the eld of fractions of the integral domain C[V]. Recall that elements of C(V)
give rational functions on V. Then let
(1.0.2) C[V]
f
=g/f

C(V) [ g C[V], 0.
1.0. Background: Afne Varieties 5
In Exercise 1.0.3 you will prove that Spec(C[V]
f
) is the afne variety V
f
.
Example 1.0.2. The n-dimensional torus is the afne open subset
(C

)
n
=C
n
`V(x
1
x
n
) C
n
,
with coordinate ring
C[x
1
, . . . , x
n
]
x
1
xn
=C[x
1
1
, . . . , x
1
n
].
Elements of this ring are called Laurent polynomials.
The ring C[V]
f
from (1.0.2) is an example of localization. In Exercises 1.0.2
and 1.0.3 you will show how to construct this ring for all afne varieties, not just
irreducible ones. The general concept of localization is discussed in standard texts
in commutative algebra such as [10, Ch. 3] and [89, Ch. 2].
Normal Afne Varieties. Let R be an integral domain with eld of fractions K.
Then R is normal, or integrally closed, if every element of K which is integral over
R (meaning that it is a root of a monic polynomial in R[x]) actually lies in R. For
example, any UFD is normal (Exercise 1.0.5).
Denition 1.0.3. An irreducible afne variety V is normal if its coordinate ring
C[V] is normal.
For example, C
n
is normal since its coordinate ring C[x
1
, . . . , x
n
] is a UFD and
hence normal. Here is an example of a nonnormal afne variety.
Example 1.0.4. Let C =V(x
3
y
2
) C
2
. This is an irreducible plane curve with a
cusp at the origin. It is easy to see that C[C] =C[x, y]/'x
3
y
2
`. Now let x and y be
the cosets of x and y in C[C] respectively. This gives y/ x C(C). A computation
shows that y/ x / C[C] and that ( y/ x)
2
= x. Consequently C[C] and hence C are not
normal. We will see below that C is an afne toric variety.
An irreducible afne variety V has a normalization dened as follows. Let
C[V]

= C(V) : is integral over C[V].


We call C[V]

the integral closure of C[V]. One can show that C[V]

is normal and
(with more work) nitely generated as a C-algebra (see [89, Cor. 13.13]). This
gives the normal afne variety
V

= Spec(C[V]

)
We call V

the normalization of V. The natural inclusion C[V] C[V]

= C[V

]
corresponds to a map V

V. This is the normalization map.


Example 1.0.5. We saw in Example 1.0.4 that the curve C C
2
dened by x
3
=y
2
has elements x, y C[C] such that y/ x / C[C] is integral over C[C]. In Exer-
cise 1.0.6 you will show that C[ y/ x] C(C) is the integral closure of C[C] and
that the normalization map is the map C C dened by t (t
2
, t
3
).
6 Chapter 1. Afne Toric Varieties
At rst glance, the denition of normal does not seem very intuitive. Once we
enter the world of toric varieties, however, we will see that normality has a very
nice combinatorial interpretation and that the nicest toric varieties are the normal
ones. We will also see that normality leads to a nice theory of divisors.
In Exercise 1.0.7 you will prove some properties of normal domains that will
be used in 1.3 when we study normal afne toric varieties.
Smooth Points of Afne Varieties. In order to dene a smooth point of an afne
variety V, we rst need to dene local rings and Zariski tangent spaces. When V
is irreducible, the local ring of V at p is
O
V,p
=f /g C(V) [ f , g C[V] and g(p) = 0.
Thus O
V,p
consists of all rational functions on V that are dened at p. Inside of
O
V,p
we have the maximal ideal
m
V,p
= O
V,p
[ (p) = 0.
In fact, m
V,p
is the unique maximal ideal of O
V,p
, so that O
V,p
is a local ring.
Exercises 1.0.2 and 1.0.4 explain how to dene O
V,p
when V is not irreducible.
The Zariski tangent space of V at p is dened to be
T
p
(V) = Hom
C
(m
V,p
/m
2
V,p
, C).
In Exercise 1.0.8 you will verify that dim T
p
(C
n
) = n for every p C
n
. According
to [131, p. 32], we can compute the Zariski tangent space of a point in an afne
variety as follows.
Lemma 1.0.6. Let V C
n
be an afne variety and let p V. Also assume that
I(V) =' f
1
, . . . , f
s
` C[x
1
, . . . , x
n
]. For each i, let
d
p
( f
i
) =
f
i
x
1
(p)x
1
+ +
f
i
x
n
(p)x
n
.
Then the Zariski tangent space T
p
(V) is isomorphic to the subspace of C
n
dened
by the equations d
p
( f
1
) = = d
p
( f
s
) = 0. In particular, dim T
p
(V) n.
Denition 1.0.7. A point p of an afne variety V is smooth or nonsingular if
dim T
p
(V) = dim
p
V, where dim
p
V is the maximum of the dimensions of the irre-
ducible components of V containing p. The point p is singular if it is not smooth.
Finally, V is smooth if every point of V is smooth.
Points lying in the intersection of two or more irreducible components of V are
always singular (see [69, Thm. 8 of Ch. 9, 6]).
Since dim T
p
(C
n
) = n for every p C
n
, we see that C
n
is smooth. For an
irreducible afne variety V C
n
of dimension d, x p V and write I(V) =
1.0. Background: Afne Varieties 7
' f
1
, . . . , f
s
`. Using Lemma 1.0.6, it is straightforward to show that V is smooth
at p if and only if the Jacobian matrix
(1.0.3) J
p
( f
1
, . . . , f
s
) =

f
i
x
j
(p)

1is,1jn
has rank nd (Exercise 1.0.9). Here is a simple example.
Example 1.0.8. As noted in Example 1.0.4, the plane curve C dened by x
3
= y
2
has I(C) ='x
3
y
2
` C[x, y]. A point p = (a, b) C has Jacobian
J
p
= (3a
2
, 2b),
so the origin is the only singular point of C.
Since T
p
(V) =Hom
C
(m
V,p
/m
2
V,p
, C), we see that V is smooth at p when dimV
equals the dimension of m
V,p
/m
2
V,p
as a vector space over O
V,p
/m
V,p
. In terms of
commutative algebra, this means that p V is smooth if and only if O
V,p
is a
regular local ring. See [10, p. 123] or [89, 10.3].
We can relate smoothness and normality as follows.
Proposition 1.0.9. A smooth irreducible afne variety V is normal.
Proof. In 3.0 we will see that C[V] =

pV
O
V,p
. By Exercise 1.0.7, C[V] is
normal once we prove that O
V,p
is normal for all p V. Hence it sufces to show
that O
V,p
is normal whenever p is smooth.
This follows from some powerful results in commutative algebra: O
V,p
is a
regular local ring when p is a smooth point of V (see above), and every regular
local ring is a UFD (see [89, Thm. 19.19]). Then we are done since every UFD is
normal. A direct proof that O
V,p
is normal at a smooth point p V is sketched in
Exercise 1.0.10.
The converse of Propostion 1.0.9 can fail. We will see in 1.3 that the afne
variety V(xy zw) C
4
is normal, yet V(xy zw) is singular at the origin.
Products of Afne Varieties. Given afne varieties V
1
and V
2
, there are several
ways to show that the cartesian product V
1
V
2
is an afne variety. The most direct
way is to proceed as follows. Let V
1
C
m
= Spec(C[x
1
, . . . , x
m
]) and V
2
C
n
=
Spec(C[y
1
, . . . , y
n
]). Take I(V
1
) =' f
1
, . . . , f
s
` and I(V
2
) ='g
1
, . . . , g
t
`. Since the f
i
and g
j
depend on separate sets of variables, it follows that
V
1
V
2
= V( f
1
, . . . , f
s
, g
1
, . . . , g
t
) C
m+n
is an afne variety.
A fancier method is to use the mapping properties of the product. This will
also give an intrinsic description of its coordinate ring. Given V
1
and V
2
as above,
8 Chapter 1. Afne Toric Varieties
V
1
V
2
should be an afne variety with projections
i
: V
1
V
2
V
i
such that
whenever we have a diagram
W
1

V
1
V
2

V
1
V
2
where
i
: W V
i
are morphisms from an afne variety W, there should be a unique
morphism (the dotted arrow) that makes the diagram commute, i.e.,
i
=
i
.
For the coordinate rings, this means that whenever we have a diagram
C[V
2
]

C[V
1
]

1

C[V
1
V
2
]

C[W]
with C-algebra homomorphisms

i
: C[V
i
] C[W], there should be a unique C-
algebra homomorphism

(the dotted arrow) that makes the diagram commute. By


the universal mapping property of the tensor product of C-algebras, C[V
1
]
C
C[V
2
]
has the mapping properties we want. Since C[V
1
]
C
C[V
2
] is a nitely generated
C-algebra with no nilpotents (see the appendix to this chapter), it is the coordinate
ring C[V
1
V
2
]. For more on tensor products, see [10, pp. 2427] or [89, A2.2].
Example 1.0.10. Let V be an afne variety. Since C
n
= Spec(C[y
1
, . . . , y
n
]), the
product V C
n
has coordinate ring
C[V]
C
C[y
1
, . . . , y
n
] =C[V][y
1
, . . . , y
n
].
If V is contained in C
m
with I(V) =' f
1
, . . . , f
s
` C[x
1
, . . . , x
m
], it follows that
I(V C
n
) =' f
1
, . . . , f
s
` C[x
1
, . . . , x
m
, y
1
, . . . , y
n
].
For later purposes, we also note that the coordinate ring of V (C

)
n
is
C[V]
C
C[y
1
1
, . . . , y
1
n
] =C[V][y
1
1
, . . . , y
1
n
].
Given afne varieties V
1
and V
2
, we note that the Zariski topology on V
1
V
2
is usually not the product of the Zariski topologies on V
1
and V
2
.
Example 1.0.11. Consider C
2
= CC. By denition, a basis for the product of
the Zariski topologies consists of sets U
1
U
2
where U
i
are Zariski open in C. Such
a set is the complement of a union of collections of horizontal and vertical lines
1.0. Background: Afne Varieties 9
in C
2
. This makes it easy to see that Zariski closed sets in C
2
such as V(y x
2
)
cannot be closed in the product topology.
Exercises for 1.0.
1.0.1. Prove Lemma 1.0.1. Hint: You will need the Nullstellensatz.
1.0.2. Let R be a commutative C-algebra. A subset S R is a multipliciative subset pro-
vided 1 S, 0 / S, and S is closed under multiplication. The localization R
S
consists of all
formal expressions g/s, g R, s S, modulo the equivalence relation
g/s h/t u(tg sh) = 0 for some u S.
(a) Show that the usual formulas for adding and multiplying fractions induce well-dened
binary operations that make R
S
into C-algebra.
(b) If R has no nonzero nilpotents, then prove that the same is true for R
S
.
For more on localization, see [10, Ch. 3] or [89, Ch. 2].
1.0.3. Let R be a nitely generated C-algebra without nilpotents as in Lemma 1.0.1 and
let f R be nonzero. Then S =1, f , f
2
, . . . is a multiplicative set. The localization R
S
is
denoted R
f
and is called the localization of R at f .
(a) Show that R
f
is a nitely generated C-algebra without nilpotents.
(b) Show that R
f
satises Spec(R
f
) = Spec(R)
f
.
(c) Show that R
f
is given by (1.0.2) when R is an integral domain.
1.0.4. Let V be an afne variety with coordinate ring C[V]. Given a point p V, let
S =g C[V] [ g(p) = 0.
(a) Show that S is a multiplicative set. The localization C[V]
S
is denoted O
V,p
and is
called the local ring of V at p.
(b) Show that every O
V,p
has a well-dened value (p) and that
m
V,p
= O
V,p
[ (p) = 0
is the unique maximal ideal of O
V,p
.
(c) When V is irreducible, show that O
V,p
agrees with the denition given in the text.
1.0.5. Prove that a UFD is normal.
1.0.6. In the setting of Example 1.0.5, show that C[ y/ x] C(C) is the integral closure of
C[C] and that the normalization C C is dened by t (t
2
, t
3
).
1.0.7. In this exercise, you will prove some properties of normal domains needed for 1.3.
(a) Let R be a normal domain with eld of fractions K and let S R be a multiplicative
subset. Prove that the localization R
S
is normal.
(b) Let R

, A, be normal domains with the same eld of fractions K. Prove that the
intersection

A
R

is normal.
1.0.8. Prove that dim T
p
(C
n
) = n for all p C
n
.
1.0.9. Use Lemma 1.0.6 to prove the claim made in the text that smoothness is determined
by the rank of the Jacobian matrix (1.0.3).
10 Chapter 1. Afne Toric Varieties
1.0.10. Let V be irreducible and suppose that p V is smooth. The goal of this exercise
is to prove that O
V,p
is normal using standard results from commutative algebra. Set n =
dimV and consider the ring of formal power series C[[x
1
, . . . , x
n
]]. This is a local ring with
maximal ideal m ='x
1
, . . . , x
n
`. We will use three facts:
C[[x
1
, . . . , x
n
]] is a UFD by [280, p. 148] and hence normal by Exercise 1.0.5.
Since p V is smooth, [207, 1C] proves the existence of a C-algebra homomorphism
O
V,p
C[[x
1
, . . . , x
n
]] that induces isomorphisms
O
V,p
/m

V,p
C[[x
1
, . . . , x
n
]]/m

for all 0. This implies that the completion (see [10, Ch. 10])

O
V,p
= lim

O
V,p
/m

V,p
is isomorphic to a formal power series ring, i.e.,

O
V,p
C[[x
1
, . . . , x
n
]]. Such an iso-
morphism captures the intuitive idea that at a smooth point, functions should have
power series expansions in local coordinates x
1
, . . . , x
n
.
If I O
V,p
is an ideal, then
I =

=1
(I +m

V,p
).
This theorem of Krull holds for any ideal I in a Noetherian local ring A and follows
from [10, Cor. 10.19] with M = A/I.
Now assume that p V is smooth.
(a) Use the third bullet to show that O
V,p
C[[x
1
, . . . , x
n
]] is injective.
(b) Suppose that a, b O
V,p
satisfy b[a in C[[x
1
, . . . , x
n
]]. Prove that b[a in O
V,p
. Hint:
Use the second bullet to show a bO
V,p
+m

V,p
and then use the third bullet.
(c) Prove that O
V,p
is normal. Hint: Use part (b) and the rst bullet.
This argument can be continued to show that O
V,p
is a UFD. See [207, (1.28)]
1.0.11. Let V andW be afne varieties and let S V be a subset. Prove that SW =S W.
1.0.12. Let V and W be irreducible afne varieties. Prove that V W is irreducible. Hint:
Suppose V W =Z
1
Z
2
, where Z
1
, Z
2
are closed. Let V
i
=v V [ vW Z
i
. Prove
that V =V
1
V
2
and that V
i
is closed in V. Exercise 1.0.11 will be useful.
1.1. Introduction to Afne Toric Varieties
We rst discuss what we mean by torus and then explore various constructions
of afne toric varieties.
The Torus. The afne variety (C

)
n
is a group under component-wise multiplica-
tion. A torus T is an afne variety isomorphic to (C

)
n
, where T inherits a group
structure from the isomorphism.
The term torus is taken from the language of linear algebraic groups. We
will use (without proof) basic results about tori that can be found in standard texts
on algebraic groups such as [37], [152], and [256]. See also [36, Ch. 3] for a
self-contained treatment of tori.
1.1. Introduction to Afne Toric Varieties 11
We begin with characters and one-parameter subgroups.
A character of a torus T is a morphism : T C

that is a group homo-


morphism. For example, m = (a
1
, . . . , a
n
) Z
n
gives a character
m
: (C

)
n
C

dened by
(1.1.1)
m
(t
1
, . . . , t
n
) = t
a
1
1
t
an
n
.
One can show that all characters of (C

)
n
arise this way (see [152, 16]). Thus the
characters of (C

)
n
form a group isomorphic to Z
n
.
For an arbitrary torus T, its characters form a free abelian group M of rank
equal to the dimension of T. It is customary to say that m M gives the character

m
: T C

.
We will need the following result concerning tori (see [152, 16] for a proof).
Proposition 1.1.1.
(a) Let T
1
and T
2
be tori and let : T
1
T
2
be a morphism that is a group homo-
morphism. Then the image of is a torus and is closed in T
2
.
(b) Let T be a torus and let H T be an irreducible subvariety of T that is a
subgroup. Then H is a torus.
Assume that a torus T acts linearly on a nite dimensional vector space W over
C, where the action of t T on w W is denoted t w. A basic result is that the
linear maps w t w are diagonalizable and can be simultaneously diagonalized.
We describe this as follows. Given m M, dene the eigenspace
W
m
=w W [ t w =
m
(t)w for all t T.
If W
m
= 0, then every w W
m
`0 is a simultaneous eigenvector for all t T,
with eigenvalue given by
m
(t). See [256, Thm. 3.2.3] for a proof of the following.
Proposition 1.1.2. In the above situation, we have W =

mM
W
m
.
A one-parameter subgroup of a torus T is a morphism : C

T that is a
group homomorphism. For example, u = (b
1
, . . . , b
n
) Z
n
gives a one-parameter
subgroup
u
: C

(C

)
n
dened by
(1.1.2)
u
(t) = (t
b
1
, . . . , t
bn
).
All one-parameter subgroups of (C

)
n
arise this way (see [152, 16]). It follows
that the group of one-parameter subgroups of (C

)
n
is naturally isomorphic to Z
n
.
For an arbitrary torus T, the one-parameter subgroups form a free abelian group N
of rank equal to the dimension of T. As with the character group, an element u N
gives the one-parameter subgroup
u
: C

T.
There is a natural bilinear pairing ' , ` : MN Z dened as follows.
(Intrinsic) Given a character
m
and a one-parameter subgroup
u
, the com-
position
m

u
: C

is a character of C

, which is given by t t

for
some Z. Then 'm, u` = .
12 Chapter 1. Afne Toric Varieties
(Concrete) If T =(C

)
n
with m=(a
1
, . . . , a
n
) Z
n
, u =(b
1
, . . . , b
n
) Z
n
, then
one computes that
(1.1.3) 'm, u` =
n

i=1
a
i
b
i
,
i.e., the pairing is the usual dot product.
It follows that the characters and one-parameter subgroups of a torus T form
free abelian groups M and N of nite rank with a pairing ' , ` : MN Z that
identies N with Hom
Z
(M, Z) and M with Hom
Z
(N, Z). In terms of tensor prod-
ucts, one obtains a canonical isomorphism N
Z
C

T via ut
u
(t). Hence
it is customary to write a torus as T
N
.
From this point of view, picking an isomorphism T
N
(C

)
n
induces dual
bases of M and N, i.e., isomorphisms M Z
n
and N Z
n
that turn characters
into Laurent monomials (1.1.1), one-parameter subgroups into monomial curves
(1.1.2), and the pairing into dot product (1.1.3).
The Denition of Afne Toric Variety. We now dene the main object of study of
this chapter.
Denition 1.1.3. An afne toric variety is an irreducible afne variety V contain-
ing a torus T
N
(C

)
n
as a Zariski open subset such that the action of T
N
on itself
extends to an algebraic action of T
N
on V. (By algebraic action, we mean an action
T
N
V V given by a morphism.)
Obvious examples of afne toric varieties are (C

)
n
and C
n
. Here are some
less trivial examples.
Example 1.1.4. The plane curve C = V(x
3
y
2
) C
2
has a cusp at the origin.
This is an afne toric variety with torus
C`0 =C(C

)
2
=(t
2
, t
3
) [ t C

,
where the isomorphism is t (t
2
, t
3
). Example 1.0.4 shows that C is a nonnormal
toric variety.
Example 1.1.5. The variety V = V(xy zw) C
4
is a toric variety with torus
V (C

)
4
=(t
1
, t
2
, t
3
, t
1
t
2
t
1
3
) [ t
i
C

(C

)
3
,
where the isomorphism is (t
1
, t
2
, t
3
) (t
1
, t
2
, t
3
, t
1
t
2
t
1
3
). We will see later that V is
normal.
Example 1.1.6. Consider the surface in C
d+1
parametrized by the map
: C
2
C
d+1
dened by (s, t) (s
d
, s
d1
t, . . . , st
d1
, t
d
). Thus is dened using all degree d
monomials in s, t.
1.1. Introduction to Afne Toric Varieties 13
Let the coordinates of C
d+1
be x
0
, . . . , x
d
and let I C[x
0
, . . . , x
d
] be the ideal
generated by the 22 minors of the matrix

x
0
x
1
x
d2
x
d1
x
1
x
2
x
d1
x
d

,
so I = 'x
i
x
j+1
x
i+1
x
j
[ 0 i < j d 1`. In Exercise 1.1.1 you will verify that
(C
2
) = V(I), so that

C
d
= (C
2
) is an afne variety. You will also prove that
I(

C
d
) = I, so that I is the ideal of all polynomials vanishing on

C
d
. It follows that I
is prime since V(I) is irreducible by Proposition 1.1.8 below. The afne surface

C
d
is called the rational normal cone of degree d and is an example of a determinantal
variety. We will see below that I is a toric ideal.
It is straightforward to show that

C
d
is a toric variety with torus
((C

)
2
) =

C
d
(C

)
d+1
(C

)
2
.
We will study this example from the projective point of view in Chapter 2.
We next explore three equivalent ways of constructing afne toric varieties.
Lattice Points. In this book, a lattice is a free abelian group of nite rank. Thus
a lattice of rank n is isomorphic to Z
n
. For example, a torus T
N
has lattices M (of
characters) and N (of one-parameter subgroups).
Given a torus T
N
with character lattice M, a set A = m
1
, . . . , m
s
M gives
characters
m
i
: T
N
C

. Then consider the map


(1.1.4)
A
: T
N
C
s
dened by

A
(t) =

m
1
(t), . . . ,
ms
(t)

C
s
.
Denition 1.1.7. Given a nite set A M, the afne toric variety Y
A
is dened
to be the Zariski closure of the image of the map
A
from (1.1.4).
This denition is justied by the following proposition.
Proposition 1.1.8. Given A M as above, let ZA M be the sublattice gener-
ated by A. Then Y
A
is an afne toric variety whose torus has character lattice
ZA. In particular, the dimension of Y
A
is the rank of ZA.
Proof. The map (1.1.4) can be regarded as a map

A
: T
N
(C

)
s
of tori. By Proposition 1.1.1, the image T =
A
(T
N
) is a torus that is closed in
(C

)
s
. The latter implies that Y
A
(C

)
s
= T since Y
A
is the Zariski closure of
the image. It follows that the image is Zariski open in Y
A
. Furthermore, T is
irreducible (it is a torus), so the same is true for its Zariski closure Y
A
.
14 Chapter 1. Afne Toric Varieties
We next consider the action of T. Since T (C

)
s
, an element t T acts on
C
s
and takes varieties to varieties. Then
T = t T t Y
A
shows that t Y
A
is a variety containing T. Hence Y
A
t Y
A
by the denition of
Zariski closure. Replacing t with t
1
leads to Y
A
= t Y
A
, so that the action of T
induces an action on Y
A
. We conclude that Y
A
is an afne toric variety.
It remains to compute the character lattice of T, which we will temporarily
denote by M

. Since T =
A
(T
N
), the map
A
gives the commutative diagram
T
N

(C

)
s
T
?

where denotes a surjective map and an injective map. This diagram of tori
induces a commutative diagram of character lattices
M Z
s
b

.
0P

Since

A
: Z
s
M takes the standard basis e
1
, . . . , e
s
to m
1
, . . . , m
s
, the image of

A
is ZA. By the diagram, we obtain M

ZA. Then we are done since the


dimension of a torus equals the rank of its character lattice.
In concrete terms, x a basis of M, so that we may assume M =Z
n
. Then the s
vectors in A Z
n
can be regarded as the columns of an ns matrix A with integer
entries. In this case, the dimension of Y
A
is simply the rank of the matrix A.
We will see below that every afne toric variety is isomorphic to Y
A
for some
nite subset A of a lattice.
Toric Ideals. Let Y
A
C
s
= Spec(C[x
1
, . . . , x
s
]) be the afne toric variety com-
ing from a nite set A = m
1
, . . . , m
s
M. We can describe the ideal I(Y
A
)
C[x
1
, . . . , x
s
] as follows. As in the proof of Proposition 1.1.8,
A
induces a map of
character lattices

A
: Z
s
M
that sends the standard basis e
1
, . . . , e
s
to m
1
, . . . , m
s
. Let L be the kernel of this
map, so that we have an exact sequence
0 L Z
s
M.
In down to earth terms, elements = (
1
, . . . ,
s
) of L satisfy

s
i=1

i
m
i
= 0 and
hence record the linear relations among the m
i
.
1.1. Introduction to Afne Toric Varieties 15
Given = (
1
, . . . ,
s
) L, set

+
=

i
>0

i
e
i
and

i
<0

i
e
i
.
Note that =
+

and that
+
,

N
s
. It follows easily that the binomial
x

+
x

i
>0
x

i
i

i
<0
x

i
i
vanishes on the image of
A
and hence on Y
A
since Y
A
is the Zariski closure of
the image.
Proposition 1.1.9. The ideal of the afne toric variety Y
A
C
s
is
I(Y
A
) =

+
x

[ L

[ , N
s
and L

.
Proof. We leave it to the reader to prove equality of the two ideals on the right
(Exercise 1.1.2). Let I
L
denote this ideal and note that I
L
I(Y
A
). We prove
the opposite inclusion following [264, Lem. 4.1]. Pick a monomial order > on
C[x
1
, . . . , x
s
] and an isomorphism T
N
(C

)
n
. Thus we may assume M = Z
n
and
the map : (C

)
n
C
s
is given by Laurent monomials t
m
i
in variables t
1
, . . . , t
n
.
If I
L
= I(Y
A
), then we can pick f I(Y
A
) ` I
L
with minimal leading monomial
x

s
i=1
x
a
i
i
. Rescaling if necessary, x

becomes the leading term of f .


Since f (t
m
1
, . . . , t
ms
) is identically zero as a polynomial in t
1
, . . . , t
n
, there must
be cancellation involving the term coming from x

. In other words, f must contain


a monomial x

s
i=1
x
b
i
i
< x

such that
s

i=1
(t
m
i
)
a
i
=
s

i=1
(t
m
i
)
b
i
.
This implies that
s

i=1
a
i
m
i
=
s

i=1
b
i
m
i
,
so that =

s
i=1
(a
i
b
i
)e
i
L. Then x

I
L
by the second description
of I
L
. It follows that f x

+x

also lies in I(Y


A
) ` I
L
and has strictly smaller
leading term. This contradiction completes the proof.
Given A M, there are several ways to compute the ideal I(Y
A
) = I
L
of
Proposition 1.1.9. In simple cases, the rational implicitization algorithm of [69,
Ch. 3, 3] can be used. One can also compute I
L
using a basis of L and ideal
quotients (Exercise 1.1.3). For more on computing I
L
, see [264, Ch. 12].
Inspired by Proposition 1.1.9, we make the following denition.
Denition 1.1.10. Let L Z
s
be a sublattice.
(a) The ideal I
L
=

[ , N
s
and L

is called a lattice ideal.


(b) A prime lattice ideal is called a toric ideal.
16 Chapter 1. Afne Toric Varieties
Since toric varieties are irreducible, the ideals appearing in Proposition 1.1.9
are toric ideals. Examples of toric ideals include:
Example 1.1.4 : 'x
3
y
2
` C[x, y]
Example 1.1.5 : 'xz yw` C[x, y, z, w]
Example 1.1.6 : 'x
i
x
j+1
x
i+1
x
j
[ 0 i < j d 1` C[x
0
, . . . , x
d
].
(The latter is the ideal of the rational normal cone

C
d
C
d+1
.) In each example,
we have a prime ideal generated by binomials. As we now show, such ideals are
automatically toric.
Proposition 1.1.11. An ideal I C[x
1
, . . . , x
s
] is toric if and only if it is prime and
generated by binomials.
Proof. One direction is obvious. So suppose that I is prime and generated by bino-
mials x

i
x

i
. Then observe that V(I) (C

)
s
is nonempty (it contains (1, . . . , 1))
and is a subgroup of (C

)
s
(easy to check). Since V(I) C
s
is irreducible, it fol-
lows that V(I) (C

)
s
is an irreducible subvariety of (C

)
s
that is also a subgroup.
By Proposition 1.1.1, we see that T = V(I) (C

)
s
is a torus.
Projecting on the ith coordinate of (C

)
s
gives a character T (C

)
s
C

,
which by our usual convention we write as
m
i
: T C

for m
i
M. It follows
easily that V(I) = Y
A
for A = m
1
, . . . , m
s
, and since I is prime, we have I =
I(Y
A
) by the Nullstellensatz. Then I is toric by Proposition 1.1.9.
We will later see that all afne toric varieties arise from toric ideals. For more
on toric ideals and lattice ideals, the reader should consult [204, Ch. 7].
Afne Semigroups. A semigroup is a set S with an associative binary operation
and an identity element. To be an afne semigroup, we further require that:
The binary operation on S is commutative. We will write the operation as +
and the identity element as 0. Thus a nite set A S gives
NA =

mA
a
m
m [ a
m
N

S.
The semigroup is nitely generated, meaning that there is a nite set A S
such that NA = S.
The semigroup can be embedded in a lattice M.
The simplest example of an afne semigroup is N
n
Z
n
. More generally, given
a lattice M and a nite set A M, we get the afne semigroup NA M. Up to
isomorphism, all afne semigroups are of this form.
Given an afne semigroup S M, the semigroup algebra C[S] is the vector
space over C with S as basis and multiplication induced by the semigroup structure
1.1. Introduction to Afne Toric Varieties 17
of S. To make this precise, we think of M as the character lattice of a torus T
N
, so
that m M gives the character
m
. Then
C[S] =

mS
c
m

m
[ c
m
C and c
m
= 0 for all but nitely many m

,
with multiplication induced by

m

m

=
m+m

.
If S =NA for A =m
1
, . . . , m
s
, then C[S] =C[
m
1
, . . . ,
ms
].
Here are two basic examples.
Example 1.1.12. The afne semigroup N
n
Z
n
gives the polynomial ring
C[N
n
] =C[x
1
, . . . , x
n
],
where x
i
=
e
i
and e
1
, . . . , e
n
is the standard basis of Z
n
.
Example 1.1.13. If e
1
, . . . , e
n
is a basis of a lattice M, then M is generated by
A = e
1
, . . . , e
n
as an afne semigroup. Setting t
i
=
e
i
gives the Laurent
polynomial ring
C[M] =C[t
1
1
, . . . , t
1
n
].
Using Example 1.0.2, one sees that C[M] is the coordinate ring of the torus T
N
.
Afne semigroup rings give rise to afne toric varieties as follows.
Proposition 1.1.14. Let S M be an afne semigroup. Then:
(a) C[S] is an integral domain and nitely generated as a C-algebra.
(b) Spec(C[S]) is an afne toric variety whose torus has character lattice ZS, and
if S =NA for a nite set A M, then Spec(C[S]) =Y
A
.
Proof. As noted above, A =m
1
, . . . , m
s
implies C[S] =C[
m
1
, . . . ,
ms
], so C[S]
is nitely generated. Since C[S] C[M] follows from S M, we see that C[S] is
an integral domain by Example 1.1.13.
Using A =m
1
, . . . , m
s
, we get the C-algebra homomorphism
: C[x
1
, . . . , x
s
] C[M]
where x
i

m
i
C[M]. This corresponds to the morphism

A
: T
N
C
s
from (1.1.4), i.e., = (
A
)

in the notation of 1.0. One checks that the kernel


of is the toric ideal I(Y
A
) (Exercise 1.1.4). The image of is C[
m
1
, . . . ,
ms
] =
C[S], and then the coordinate ring of Y
A
is
(1.1.5)
C[Y
A
] =C[x
1
, . . . , x
s
]/I(Y
A
)
=C[x
1
, . . . , x
s
]/Ker() Im() =C[S].
18 Chapter 1. Afne Toric Varieties
This proves that Spec(C[S]) =Y
A
. Since S =NA implies ZS =ZA, the torus of
Y
A
= Spec(C[S]) has the desired character lattice by Proposition 1.1.8.
Here is an example of this proposition.
Example 1.1.15. Consider the afne semigroup S Z generated by 2 and 3, so
that S = 0, 2, 3, . . . . To study the semigroup algebra C[S], we use (1.1.5). If we
set A = 2, 3, then
A
(t) = (t
2
, t
3
) and the toric ideal is I(Y
A
) = 'x
3
y
2
` by
Example 1.1.4. Hence
C[S] =C[t
2
, t
3
] C[x, y]/'x
3
y
2
`
and the afne toric variety Y
A
is the curve x
3
= y
2
from Example 1.1.4.
Equivalence of Constructions. Before stating our main result, we need to study
the action of the torus T
N
on the semigroup algebra C[M]. The action of T
N
on
itself given by multiplication induces an action on C[M] as follows: if t T
N
and
f C[M], then t f C[M] is dened by p f (t
1
p) for p T
N
. The minus
sign will be explained in 5.0.
The following lemma will be used several times in the text.
Lemma 1.1.16. Let A C[M] be a subspace stable under the action of T
N
. Then
A =

m
A
C
m
.
Proof. Let A

m
A
C
m
and note that A

A. For the opposite inclusion,


pick f = 0 in A. Since A C[M], we can write
f =

mB
c
m

m
,
where BM is nite and c
m
= 0 for all m B. Then f BA, where
B = Span(
m
[ m B) C[M].
An easy computation shows that t
m
=
m
(t
1
)
m
. It follows that B and
hence BA are stable under the action of T
N
. Since BA is nite-dimensional,
Proposition 1.1.2 implies that BA is spanned by simultaneous eigenvectors of T
N
.
This is taking place in C[M], where simultaneous eigenvectors are characters. It
follows that BAis spanned by characters. Then the above expression for f BA
implies that
m
A for m B. Hence f A

, as desired.
We can now state the main result of this section, which asserts that our various
approaches to afne toric varieties all give the same class of objects.
Theorem 1.1.17. Let V be an afne variety. The following are equivalent:
(a) V is an afne toric variety according to Denition 1.1.3.
(b) V =Y
A
for a nite set A in a lattice.
1.1. Introduction to Afne Toric Varieties 19
(c) V is an afne variety dened by a toric ideal.
(d) V = Spec(C[S]) for an afne semigroup S.
Proof. The implications (b) (c) (d) (a) follow from Propositions 1.1.8,
1.1.9 and 1.1.14. For (a) (d), let V be an afne toric variety containing the
torus T
N
with character lattice M. Since the coordinate ring of T
N
is the semigroup
algebra C[M], the inclusion T
N
V induces a map of coordinate rings
C[V] C[M].
This map is injective since T
N
is Zariski dense in V, so that we can regard C[V] as
a subalgebra of C[M].
Since the action of T
N
on V is given by a morphism T
N
V V, we see that
if t T
N
and f C[V], then p f (t
1
p) is a morphism on V. It follows that
C[V] C[M] is stable under the action of T
N
. By Lemma 1.1.16, we obtain
C[V] =

m
C[V]
C
m
.
Hence C[V] =C[S] for the semigroup S =m M [
m
C[V].
Finally, since C[V] is nitely generated, we can nd f
1
, . . . , f
s
C[V] with
C[V] = C[ f
1
, . . . , f
s
]. Expressing each f
i
in terms of characters as above gives a
nite generating set of S. It follows that S is an afne semigroup.
Here is one way to think about the above proof. When an irreducible afne
variety V contains a torus T
N
as a Zariski open subset, we have the inclusion
C[V] C[M].
Thus C[V] consists of those functions on the torus T
N
that extend to polynomial
functions on V. Then the key insight is that V is a toric variety precisely when the
functions that extend are determined by the characters that extend.
Example 1.1.18. We have seen that V = V(xy zw) C
4
is a toric variety with
toric ideal 'xyzw` C[x, y, z, w]. Also, the torus is (C

)
3
via the map (t
1
, t
2
, t
3
)
(t
1
, t
2
, t
3
, t
1
t
2
t
1
3
). The lattice points used in this map can be represented as the
columns of the matrix
(1.1.6)

1 0 0 1
0 1 0 1
0 0 1 1

.
The corresponding semigroup S Z
3
consists of the N-linear combinations of the
column vectors. Hence the elements of S are lattice points lying in the polyhe-
dral region in R
3
pictured in Figure 1 on the next page. In this gure, the four
vectors generating S are shown in bold, and the boundary of the polyhedral region
is partially shaded. In the terminology of 1.2, this polyhedral region is a rational
20 Chapter 1. Afne Toric Varieties
(0,0,1)
(0,1,0)
(1,0,0)
(1,1,1)
Figure 1. Cone containing the lattice points corresponding to V =V(xy zw)
polyhedral cone. In Exercise 1.1.5 you will show that S consists of all lattice points
lying in the cone in Figure 1. We will use this in 1.3 to prove that V is normal.
Exercises for 1.1.
1.1.1. As in Example 1.1.6, let
I ='x
i
x
j+1
x
i+1
x
j
[ 0 i < j d 1` C[x
0
, . . . , x
d
]
and let

C
d
be the surface parametrized by
(s, t) = (s
d
, s
d1
t, . . . , st
d1
, t
d
) C
d+1
.
(a) Prove that V(I) = (C
2
) C
d+1
. Thus

C
d
= V(I).
(b) Prove that I(

C
d
) is homogeneous.
(c) Consider lex monomial order with x
0
> x
1
> > x
d
. Let f I(

C
d
) be homogeneous
of degree and let r be the remainder of f on division by the generators of I. Prove
that r can be written
r = h
0
(x
0
, x
1
) +h
1
(x
1
, x
2
) + +h
d1
(x
d1
, x
d
)
where h
i
is homogeneous of degree . Also explain why we may assume that the
coefcient of x

i
in h
i
is zero for 1 i d 1.
(d) Use part (c) and r(s
d
, s
d1
t, . . . , st
d1
, t
d
) = 0 to show that r = 0.
(e) Use parts (b), (c) and (d) to prove that I = I(

C
d
). Also explain why the generators of
I are a Gr obner basis for the above lex order.
1.1.2. Let L Z
s
be a sublattice. Prove that
'x
+
x

[ L` ='x

[ , N
s
, L`.
Note that when L, the vectors
+
,

N
s
have disjoint support (i.e., no coordinate is
positive in both), while this may fail for arbitrary , N
s
with L.
1.1. Introduction to Afne Toric Varieties 21
1.1.3. Let I
L
be a toric ideal and let
1
, . . . ,
r
be a basis of the sublattice L Z
s
. Dene

I
L
='x

i
+
x

[ i = 1, . . . , r`.
Prove that I
L
=

I
L
: 'x
1
x
s
`

. Hint: Given , N
s
with L, write =

r
i=1
a
i

i
, a
i
Z. This implies
x

1 =

ai >0

i
+
x

ai

ai <0

i
+

ai
1.
Show that multiplying this by (x
1
x
s
)
k
gives an element of

I
L
for k 0. (By being more
careful, one can show that this result holds for lattice ideals. See [204, Lem. 7.6].)
1.1.4. Fix an afne variety V. Then f
1
, . . . , f
s
C[V] give a polynomial map : V C
s
,
which on coordinate rings is given by

: C[x
1
, . . . , x
s
] C[V], x
i
f
i
.
Let Y C
s
be the Zariski closure of the image of .
(a) Prove that I(Y) = Ker(

).
(b) Explain how this applies to the proof of Proposition 1.1.14.
1.1.5. Let m
1
= (1, 0, 0), m
2
= (0, 1, 0), m
3
= (0, 0, 1), m
4
= (1, 1, 1) be the columns of
the matrix in Example 1.1.18 and let
C =

i=1

i
m
i
[
i
R
0

R
3
be the cone in Figure 1. Prove that CZ
3
is a semigroup generated by m
1
, m
2
, m
3
, m
4
.
1.1.6. An interesting observation is that different sets of lattice points can parametrize the
same afne toric variety, even though these parametrizations behave slightly differently. In
this exercise you will consider the parametrizations

1
(s, t) = (s
2
, st, st
3
) and
2
(s, t) = (s
3
, st, t
3
).
(a) Prove that
1
and
2
both give the afne toric variety Y = V(xz y
3
) C
3
.
(b) We can regard
1
and
2
as maps

1
: C
2
Y and
2
: C
2
Y.
Prove that
2
is surjective and that
1
is not.
In general, a nite subset A Z
n
gives a rational map
A
: C
n
Y
A
. The image of

A
in C
s
is called a toric set in the literature. Thus
1
(C
2
) and
2
(C
2
) are toric sets. The
papers [169] and [239] study when a toric set equals the corresponding afne toric variety.
1.1.7. In Example 1.1.6 and Exercise 1.1.1 we constructed the rational normal cone

C
d
using all monomials of degree d in s, t. If we drop some of the monomials, things become
more complicated. For example, consider the surface parametrized by
(s, t) = (s
4
, s
3
t, st
3
, t
4
) C
4
.
This gives a toric variety Y C
4
. Show that the toric ideal of Y is given by
I(Y) ='xwyz, yw
2
z
3
, xz
2
y
2
w, x
2
z y
3
` C[x, y, z, w].
22 Chapter 1. Afne Toric Varieties
The ideal

C
4
has quadratic generators; by removing s
2
t
2
, we now get cubic generators. See
Example B.1.1 for a computational approach to this exercise. See also Example 2.1.10,
where we will study the paramaterization from the projective point of view.
1.1.8. Instead of working over C, we will work over an algebraically closed eld k of
characteristic 2. Consider the afne toric variety V k
5
parametrized by
(s, t, u) = (s
4
, t
4
, u
4
, s
8
u, t
12
u
3
) k
5
.
(a) Find generators for the toric ideal I = I(V) k[x
1
, x
2
, x
3
, x
4
, x
5
].
(b) Show that dimV = 3. You may assume that Proposition 1.1.8 holds over k.
(c) Show that I =

x
4
4
+x
8
1
x
3
, x
4
5
+x
12
2
x
3
3

.
It follows that V k
5
has codimension two and can be dened by two equations, i.e., V
is a set-theoretic complete intersection. The paper [12] shows that if we replace k with an
algebraically closed eld of characteristic p > 2, then the above parametrization is never a
set-theoretic complete intersection.
1.1.9. Prove that a lattice ideal I
L
for L Z
s
is a toric ideal if and only if Z
s
/L is torsion-
free. Hint: When Z
s
/L is torsion-free, it can be regarded as the character lattice of a torus.
The other direction of the proof is more challenging. If you get stuck, see [204, Thm. 7.4].
1.1.10. Prove that I ='x
2
1, xy 1, yz 1` is the lattice ideal for the lattice
L =(a, b, c) Z
3
[ a +b +c 0 mod 2 Z
3
.
Also compute the primary decomposition of I to show that I is not prime.
1.1.11. Let T
N
be a torus with character lattice M. Then every point t T
N
gives an evalua-
tion map
t
: M C

dened by
t
(m) =
m
(t). Prove that
t
is a group homomorphism
and that the map t
t
induces a group isomorphism
T
N
Hom
Z
(M, C

).
1.1.12. Consider tori T
1
and T
2
with character lattices M
1
and M
2
. By Example 1.1.13, the
coordinate rings of T
1
and T
2
are C[M
1
] and C[M
2
]. Let : T
1
T
2
be a morphism that is
a group homomorphism. Then induces maps

: M
2
M
1
and

: C[M
2
] C[M
1
]
by composition. Prove that

is the map of semigroup algebras induced by the map



of
afne semigroups.
1.1.13. A commutative semigroup S is cancellative if u +v = u +w implies v = w for all
u, v, w S and torsion-free if nu = nv implies u = v for all n N` 0 and u, v S. Prove
that S is afne if and only if it is nitely generated, cancellative, and torsion-free.
1.1.14. The requirement that an afne semigroup be nitely generated is important since
lattices contain semigroups that are not nitely generated. For example, let > 0 be irra-
tional and consider the semigroup
S =(a, b) N
2
[ b a Z
2
.
Prove that S is not nitely generated. (The generators of S are related to continued frac-
tions. For example, when = (1 +

5)/2 is the golden ratio, the minimal generators of


S are (0, 1) and (F
2n
, F
2n+1
) for n = 1, 2, . . . , where F
n
is the nth Fibonacci number. See
[231] and [259]. Continued fractions will play an important role in Chapter 10.
1.2. Cones and Afne Toric Varieties 23
1.1.15. Suppose that : M M is a group isomorphism. Fix a nite set A M and
let B = (A ). Prove that the toric varieties Y
A
and Y
B
are equivariantly isomorphic
(meaning that the isomorphism respects the torus action).
1.2. Cones and Afne Toric Varieties
We begin with a brief discussion of rational polyhedral cones and then explain how
they relate to afne toric varieties.
Convex Polyhedral Cones. Fix a pair of dual vector spaces M
R
and N
R
. Our dis-
cussion of cones will omit most proofswe refer the reader to [105] for more
details and [218, App. A.1] for careful statements. See also [51, 128, 241].
Denition 1.2.1. A convex polyhedral cone in N
R
is a set of the form
= Cone(S) =

uS

u
u [
u
0

N
R
,
where S N
R
is nite. We say that is generated by S. Also set Cone() =0.
A convex polyhedral cone is in fact convex, meaning that x, y implies
x +(1 )y for all 0 1, and is a cone, meaning that x implies
x for all 0. Since we will only consider convex cones, the cones satisfying
Denition 1.2.1 will be called simply polyhedral cones.
Examples of polyhedral cones include the rst quadrant in R
2
or rst octant in
R
3
. For another example, the cone Cone(e
1
, e
2
, e
1
+e
3
, e
2
+e
3
) R
3
is pictured
in Figure 2. It is also possible to have cones that contain entire lines. For example,
z
y
x
Figure 2. Cone in R
3
generated by e1, e2, e1 +e3, e2 +e3
24 Chapter 1. Afne Toric Varieties
Cone(e
1
, e
1
) R
2
is the x-axis, while Cone(e
1
, e
1
, e
2
) is the closed upper half-
plane (x, y) R
2
[ y 0. As we will see below, these last two examples are not
strongly convex.
We can also create cones using polytopes, which are dened as follows.
Denition 1.2.2. A polytope in N
R
is a set of the form
P = Conv(S) =

uS

u
u [
u
0,

uS

u
= 1

N
R
,
where S N
R
is nite. We say that P is the convex hull of S.
Polytopes include all polygons in R
2
and bounded polyhedra in R
3
. As we will
see in later chapters, polytopes play a prominent role in the theory of toric varieties.
Here, however, we simply observe that a polytope P N
R
gives a polyhedral cone
C(P) N
R
R, called the cone of P and dened by
C(P) = (u, 1) N
R
R[ u P, 0.
If P = Conv(S), then we can also describe this as C(P) = Cone(S1). Figure 3
shows what this looks when P is a pentagon in the plane.
P
Figure 3. The cone C(P) of a pentagon P R
2
The dimension dim of a polyhedral cone is the dimension of the smallest
subspace W = Span() of N
R
containing . We call Span() the span of .
Dual Cones and Faces. As usual, the pairing between M
R
and N
R
is denoted ' , `.
Denition 1.2.3. Given a polyhedral cone N
R
, its dual cone is dened by

=m M
R
[ 'm, u` 0 for all u .
Duality has the following important properties.
Proposition 1.2.4. Let N
R
be a polyhedral cone. Then

is a polyhedral cone
in M
R
and (

= .
1.2. Cones and Afne Toric Varieties 25
Given m = 0 in M
R
, we get the hyperplane
H
m
=u N
R
[ 'm, u` = 0 N
R
and the closed half-space
H
+
m
=u N
R
[ 'm, u` 0 N
R
.
Then H
m
is a supporting hyperplane of a polyhedral cone N
R
if H
+
m
,
and H
+
m
is a supporting half-space. Note that H
m
is a supporting hyperplane of
if and only if m

` 0. Furthermore, if m
1
, . . . , m
s
generate

, then it is
straightforward to check that
(1.2.1) = H
+
m
1
H
+
ms
.
Thus every polyhedral cone is an intersection of nitely many closed half-spaces.
We can use supporting hyperplanes and half-spaces to dene faces of a cone.
Denition 1.2.5. A face of a cone of the polyhedral cone is =H
m
for some
m

, written _ . Using m = 0 shows that is a face of itself, i.e., _ .


Faces = are called proper faces, written .
The faces of a polyhedral cone have the following obvious properties.
Lemma 1.2.6. Let = Cone(S) be a polyhedral cone. Then:
(a) Every face of is a polyhedral cone.
(b) An intersection of two faces of is again a face of .
(c) A face of a face of is again a face of .
You will prove the following useful property of faces in Exercise 1.2.1.
Lemma 1.2.7. Let be a face of a polyhedral cone . If v, w and v +w ,
then v, w .
A facet of is a face of codimension 1, i.e., dim =dim 1. An edge of
is a face of dimension 1. In Figure 4 on the next page we illustrate a 3-dimensional
cone with shaded facets and a supporting hyperplane (a plane in this case) that cuts
out the vertical edge of the cone.
Here are some properties of facets.
Proposition 1.2.8. Let N
R
R
n
be a polyhedral cone. Then:
(a) If = H
+
m
1
H
+
ms
for m
i

, 1 i s, then

= Cone(m
1
, . . . , m
s
).
(b) If dim = n, then in (a) we can assume that the facets of are
i
= H
m
i
.
(c) Every proper face is the intersection of the facets of containing .
Note how part (b) of the proposition renes (1.2.1) when dim = dim N
R
.
26 Chapter 1. Afne Toric Varieties

supporting
hyperplane
H
H

Figure 4. A cone R
3
and a hyperplane H supporting an edge H
When working in R
n
, dot product allows us to identify the dual with R
n
. From
this point of view, the vectors m
1
, . . . , m
s
in part (a) of the proposition are facet
normals, i.e., perpendicular to the facets. This makes it easy to compute examples.
Example 1.2.9. It is easy to see that the facet normals to the cone R
3
in
Figure 2 are m
1
= e
1
, m
2
= e
2
, m
3
= e
3
, m
4
= e
1
+e
2
e
3
. Hence

= Cone(e
1
, e
2
, e
3
, e
1
+e
2
e
3
) R
3
.
This is the cone of Figure 1 at the end of 1.1 whose lattice points describe the
semigroup of the afne toric variety V(xy zw) (see Example 1.1.18). As we will
see, this is part of how cones describe normal afne toric varieties.
Now consider

, which is the cone in Figure 1. The reader can check that the
facet normals of this cone are e
1
, e
2
, e
1
+e
3
, e
2
+e
3
. Using duality and part (a) of
Proposition 1.2.8, we obtain
= (

= Cone(e
1
, e
2
, e
1
+e
3
, e
2
+e
3
).
Hence we recover our original description of . See also Example B.2.1.
In this example, facets of the cone correspond to edges of its dual. More gen-
erally, given a face _ N
R
, we dene

=m M
R
[ 'm, u` = 0 for all u

=m

[ 'm, u` = 0 for all u


=

.
We call

the dual face of because of the following proposition.


Proposition 1.2.10. If is a face of a polyhedral cone and

, then:
(a)

is a face of

.
1.2. Cones and Afne Toric Varieties 27
(b) The map

is a bijective inclusion-reversing correspondence between the


faces of and the faces of

.
(c) dim +dim

= n.
Here is an example of Proposition 1.2.10 when dim < dim N
R
.
Example 1.2.11. Let = Cone(e
1
, e
2
) R
3
. Figure 5 shows and

. You

x
y
z

x
y
z
Figure 5. A 2-dimensional cone R
3
and its dual

R
3
should check that the maximal face of , namely itself, gives the minimal face

of

, namely the z-axis. Note also that


dim +dim

= 3
even though has dimension 2.
Relative Interiors. As already noted, the span of a cone N
R
is the smallest
subspace of N
R
containing . Then the relative interior of , denoted Relint(), is
the interior of in its span. Exercise 1.2.2 will characterize Relint() as follows:
u Relint() 'm, u` > 0 for all m

.
When the span equals N
R
, the relative interior is just the interior, denoted Int().
For an example of how relative interiors arise naturally, suppose that _ .
This gives the dual face

of

. Furthermore, if m

, then one
easily sees that
m

H
m
.
In Exercise 1.2.2, you will show that if m

, then
m Relint(

) = H
m
.
Thus the relative interior Relint(

) tells us exactly which supporting hyperplanes


of cut out the face .
28 Chapter 1. Afne Toric Varieties
Strong Convexity. Of the cones shown in Figures 15, all but

in Figure 5 have
the nice property that the origin is a face. Such cones are called strongly convex.
This condition can be stated several ways.
Proposition 1.2.12. Let N
R
R
n
be a polyhedral cone. Then:
is strongly convex 0 is a face of
contains no positive-dimensional subspace of N
R
() =0
dim

= n.
You will prove Proposition 1.2.12 in Exercise 1.2.3. One corollary is that if a
polyhedral cone is strongly convex of maximal dimension, then so is

. The
cones pictured in Figures 14 satisfy this condition.
In general, a polyhedral cone always has a minimal face that is the largest
subspace W contained in . Furthermore:
W = ().
W = H
m
whenever m Relint(

).
= /W N
R
/W is a strongly convex polyhedral cone.
See Exercise 1.2.4.
Separation. When two cones intersect in a face of each, we can separate the cones
with the following result, often called the separation lemma.
Lemma 1.2.13 (Separation Lemma). Let
1
,
2
be polyhedral cones in N
R
that
meet along a common face =
1

2
. Then
= H
m

1
= H
m

2
for any m Relint(

1
(
2
)

).
Proof. Given A, B N
R
, we set AB =ab [ a A, b B. A standard result
from cone theory tells us that

1
(
2
)

= (
1

2
)

.
Now x m Relint(

1
(
2
)

). The above equation and Exercise 1.2.4 imply


that H
m
cuts out the minimal face of
1

2
, i.e.,
H
m
(
1

2
) = (
1

2
) (
2

1
).
However, we also have
(
1

2
) (
2

1
) = .
One inclusion is obvious since =
1

2
. For the other inclusion, write u
(
1

2
) (
2

1
) as
u = a
1
a
2
= b
2
b
1
, a
1
, b
1

1
, a
2
, b
2

2
.
1.2. Cones and Afne Toric Varieties 29
Then a
1
+b
1
= a
2
+b
2
implies that this element lies in =
1

2
. Since a
1
, b
1

1
, we have a
1
, b
1
by Lemma 1.2.7, and a
2
, b
2
follows similarly. Hence
u = a
1
a
2
, as desired.
We conclude that H
m
(
1

2
) = . Intersecting with
1
, we obtain
H
m

1
= ( )
1
= ,
where the last equality again uses Lemma 1.2.7 (Exercise 1.2.5). If instead we
intersect with
2
, we obtain
H
m
(
2
) = ( ) (
2
) =,
and H
m

2
= follows.
In the situation of Lemma 1.2.13 we call H
m
a separating hyperplane.
Rational Polyhedral Cones. Let N and M be dual lattices with associated vector
spaces N
R
= N
Z
R and M
R
= M
Z
R. For R
n
we usually use the lattice Z
n
.
Denition 1.2.14. A polyhedral cone N
R
is rational if = Cone(S) for some
nite set S N.
The cones appearing in Figures 1, 2 and 5 are rational. We note without proof
that faces and duals of rational polyhedral cones are rational. Furthermore, if =
Cone(S) for S N nite and N
Q
= N
Z
Q, then
(1.2.2) N
Q
=

uS

u
u [
u
0 in Q

.
One new feature is that a strongly convex rational polyedral cone has a
canonical generating set, constructed as follows. Let be an edge of . Since is
strongly convex, is a ray, i.e., a half-line, and since is rational, the semigroup
N is generated by a unique element u

N. We call u

the ray generator of


. Figure 6 shows the ray generator of a rational ray in the plane. The dots are
the lattice N =Z
2
and the white ones are N.

ray generator u

Figure 6. A rational ray R


2
and its unique ray generator u
Lemma 1.2.15. A strongly convex rational polyhedral cone is generated by the ray
generators of its edges.
30 Chapter 1. Afne Toric Varieties
It is customary to call the ray generators of the edges the minimal generators
of a strongly convex rational polyhedral cone. Figures 1 and 2 show 3-dimensional
strongly convex rational polyhedral cones and their ray generators.
In a similar way, a rational polyhedral cone of maximal dimension has unique
facet normals, which are the ray generators of the dual

, which is strongly con-


vex by Proposition 1.2.12.
Here are some especially important strongly convex cones.
Denition 1.2.16. Let N
R
be a strongly convex rational polyhedral cone.
(a) is smooth or regular if its minimal generators form part of a Z-basis of N,
(b) is simplicial if its minimal generators are linearly independent over R.
The cone pictured in Figure 5 is smooth, while those in Figures 1 and 2 are
not even simplicial. Note also that the dual of a smooth (resp. simplicial) cone of
maximal dimension is again smooth (resp. simplicial). Later in the section we will
give examples of simplicial cones that are not smooth.
Semigroup Algebras and Afne Toric Varieties. Given a rational polyhedral cone
N
R
, the lattice points
S

M M
form a semigroup. A key fact is that this semigroup is nitely generated.
Proposition 1.2.17 (Gordans Lemma). S

M is nitely generated and


hence is an afne semigroup.
Proof. Since

is rational polyhedral,

=Cone(T) for a nite set T M. Then


K =

mT

m
m [ 0
m
< 1 is a bounded region of M
R
R
n
, so that K M is
nite since M Z
n
. Note that T (KM) S

.
We claim T (KM) generates S

as a semigroup. To prove this, take w S

and write w =

mT

m
m where
m
0. Then
m
=
m
+
m
with
m
N
and 0
m
< 1, so that
w =

mT

m
m+

mT

m
m.
The second sum is in K M (remember w M). It follows that w is a nonnegative
integer combination of elements of T (KM).
Since afne semigroups give afne toric varieties, we get the following.
Theorem 1.2.18. Let N
R
R
n
be a rational polyhedral cone with semigroup
S

M. Then
U

= Spec(C[S

]) = Spec(C[

M])
is an afne toric variety. Furthermore,
dimU

= n the torus of U

is T
N
= N
Z
C

is strongly convex.
1.2. Cones and Afne Toric Varieties 31
Proof. By Gordans Lemma and Proposition 1.1.14, U

is an afne toric variety


whose torus has character lattice ZS

M. To study ZS

, note that
ZS

= S

=m
1
m
2
[ m
1
, m
2
S

.
Now suppose that km ZS

for some k > 1 and m M. Then km = m


1
m
2
for
m
1
, m
2
S

M. Since m
1
and m
2
lie in the convex set

, we have
m+m
2
=
1
k
m
1
+
k1
k
m
2

.
It follows that m = (m+m
2
) m
2
ZS

, so that M/ZS

is torsion-free. Hence
(1.2.3) the torus of U

is T
N
ZS

= M rankZS

= n.
Since is strongly convex if and only if dim

=n (Proposition 1.2.12), it remains


to show that
dimU

= n rankZS

= n dim

= n.
The rst equivalence follows since the dimension of an afne toric variety is the
dimension of its torus, which is the rank of its character lattice. We leave the proof
of the second equivalence to the reader (Exercise 1.2.6).
Remark 1.2.19.
(a) For the rest of the book, we will always assume that N
R
is strongly convex
since we want T
N
to be the torus of the afne toric variety U

.
(b) The reader may ask why we focus on N
R
since U

= Spec(C[

M])
makes

M
R
seem more important. The answer will become clear once we
understand how normal toric varieties are constructed from afne pieces. The
discussion following Proposition 1.3.16 gives a rst hint of how this works.
Example 1.2.20. Let = Cone(e
1
, e
2
, e
1
+e
3
, e
2
+e
3
) N
R
= R
3
with N = Z
3
.
This is the cone pictured in Figure 2. By Example 1.2.9,

is the cone pictured


in Figure 1, and by Example 1.1.18, the lattice points in this cone are generated
by columns of matrix (1.1.6). It follows from Example 1.1.18 that U

is the afne
toric variety V(xy zw).
Example 1.2.21. Fix 0 r n and set = Cone(e
1
, . . . , e
r
) R
n
. Then

= Cone(e
1
, . . . , e
r
, e
r+1
, . . . , e
n
)
and the corresponding afne toric variety is
U

= Spec(C[x
1
, . . . , x
r
, x
1
r+1
, . . . , x
1
n
]) =C
r
(C

)
nr
(Exercise 1.2.7). This implies that if N
R
R
n
is a smooth cone of dimension r,
then U

C
r
(C

)
nr
. Figure 5 from Example 1.2.11 shows r =2 and n =3.
Example 1.2.22. Fix a positive integer d and let = Cone(de
1
e
2
, e
2
) R
2
.
This has dual cone

= Cone(e
1
, e
1
+de
2
). Figure 7 on the next page shows

when d = 4. The afne semigroup S

Z
2
is generated by the lattice points
32 Chapter 1. Afne Toric Varieties
(1, i) for 0 i d. When d = 4, these are the white dots in Figure 7. (You will
prove these assertions in Exercise 1.2.8.)
Figure 7. The cone

when d =4
By 1.1, the afne toric variety U

is the Zariski closure of the image of the


map : (C

)
2
C
d+1
dened by
(s, t) = (s, st, st
2
, . . . , st
d
).
This map has the same image as the map (s, t) (s
d
, s
d1
t, . . . , st
d1
, t
d
) used in
Example 1.1.6. Thus U

is isomorphic to the rational normal cone



C
d
C
d+1
whose ideal is generated by the 22 minors of the matrix

x
0
x
1
x
d2
x
d1
x
1
x
2
x
d1
x
d

.
Note that the cones and

are simplicial but not smooth.


We will return to this example often. One thing evident in Example 1.1.6 is the
difference between cone generators and semigroup generators: the cone

has
two generators but the semigroup S

Z
2
has d +1.
When N
R
has maximal dimension, the semigroup S

M has a
unique minimal generating set constructed as follows. Dene an element m= 0 of
S

to be irreducible if m = m

+m

for m

, m

implies m

= 0 or m

= 0.
Proposition 1.2.23. Let N
R
be a strongly convex rational polyhedral cone of
maximal dimension and let S

M. Then
H =m S

[ m is irreducible
has the following properties:
(a) H is nite and generates S

.
(b) H contains the ray generators of the edges of

.
(c) H is the minimal generating set of S

with respect to inclusion.


1.2. Cones and Afne Toric Varieties 33
Proof. Proposition 1.2.12 implies that

is strongly convex, so we can nd an


element u N `0 such that 'm, u` N for all m S

and 'm, u` = 0 if and


only if m = 0.
Now suppose that m S

is not irreducible. Then m = m

+m

where m

and
m

are nonzero elements of S

. It follows that
'm, u` ='m

, u` +'m

, u`
with 'm

, u`, 'm

, u` N`0, so that
'm

, u` <'m, u` and 'm

, u` <'m, u`.
Using induction on 'm, u`, we conclude that every element of S

is a sum of irre-
ducible elements, so that H generates S

. Furthermore, using a nite generating


set of S

, one easily sees that H is nite. This proves part (a). The remaining
parts of the proof are covered in Exercise 1.2.9.
The set H S

is called the Hilbert basis of S

and its elements are the


minimal generators of S

. More generally, Proposition 1.2.23 holds for any afne


semigroup S satisfying S(S) = 0. Algorithms for computing Hilbert bases
are discussed in [204, 7.3] and [264, Ch. 13], and Hilbert bases can be computed
using the computer program Normaliz [57]. See Examples B.3.1 and B.3.2.
Exercises for 1.2.
1.2.1. Prove Lemma 1.2.7. Hint: Write = H
m
for m

.
1.2.2. Here are some properties of relative interiors. Let N
R
be a cone.
(a) Show that if u , then u Relint() if and only if 'm, u` > 0 for all m

.
(b) Let _ and x m

. Prove that
m

H
m

m Relint(

) = H
m
.
1.2.3. Prove Proposition 1.2.12.
1.2.4. Let N
R
be a polyhedral cone.
(a) Use Proposition 1.2.10 to show that has a unique minimal face with respect to _.
Let W denote this minimal face.
(b) Prove that W = (

.
(c) Prove that W is the largest subspace contained in .
(d) Prove that W = ().
(e) Fix m

. Prove that m Relint(

) if and only if W = H
m
.
(f) Prove that = /W N
R
/W is a strongly convex polyhedral cone.
1.2.5. Let _ N
R
and let be dened as in the proof of Lemma 1.2.13. Prove
that = ( ) . Also show that = Span(), i.e., is the smallest subspace
of N
R
containing .
34 Chapter 1. Afne Toric Varieties
1.2.6. Fix a lattice M and let Span(S) denote the span over R of a subset S M
R
.
(a) Let S M be nite. Prove that rankZS = dimSpan(S).
(b) Let S M
R
be nite. Prove that dimCone(S) = dim Span(S).
(c) Use parts (a) and (b) to complete the proof of Theorem 1.2.18.
1.2.7. Prove the assertions made in Example 1.2.21.
1.2.8. Prove the assertions made in Example 1.2.22. Hint: First show that when a cone is
smooth, the ray generators of the cone also generate the corresponding semigroup. Then
write the cone

of Example 1.2.22 as a union of such cones.


1.2.9. Complete the proof of Proposition 1.2.23. Hint for part (b): Show that the ray
generators of the edges of

are irreducible in S

. Given an edge of

, it will help to
pick u N ` 0 such that = H
u

.
1.2.10. Let N
R
be a cone generated by a set of linearly independent vectors in N
R
.
Show that is strongly convex and simplicial.
1.2.11. Explain the picture illustrated in Figure 8 in terms of Proposition 1.2.8.

Figure 8. A cone in the plane and its dual


1.2.12. Let PN
R
be a polytope lying in an afne hyperplane (= translate of a hyperplane)
not containing the origin. Generalize Figure 3 by showing that P gives a strongly convex
polyhedral cone in N
R
. Draw a picture.
1.2.13. Consider the cone = Cone(3e
1
2e
2
, e
2
) R
2
.
(a) Describe

and nd generators of

Z
2
. Draw a picture similar to Figure 7.
(b) Compute the toric ideal of the afne toric variety U

and explain how this exercise


relates to Exercise 1.1.6.
1.2.14. Consider the simplicial cone = Cone(e
1
, e
2
, e
1
+e
2
+2e
3
) R
3
.
(a) Describe

and nd generators of

Z
3
.
(b) Compute the toric ideal of the afne toric variety U

.
1.2.15. Let be a strongly convex polyhedral cone of maximal dimension. Here is an
example taken from [105, p. 132] to show that and

need not have the same number


of edges. Let R
4
be the cone generated by 2e
i
+e
j
for all 1 i, j 4, i = j.
(a) Show that has 12 edges.
(b) Show that

is generated by e
i
and e
i
+2

j=i
e
j
, 1 i 4 and has 8 edges.
1.3. Properties of Afne Toric Varieties 35
1.3. Properties of Afne Toric Varieties
The nal task of this chapter is to explore the properties of afne toric varieties.
We will also study maps between afne toric varieties.
Points of Afne Toric Varieties. We rst consider various ways to describe the
points of an afne toric variety.
Proposition 1.3.1. Let V = Spec(C[S]) be the afne toric variety of the afne
semigroup S. Then there are bijective correspondences between the following:
(a) Points p V.
(b) Maximal ideals mC[S].
(c) Semigroup homomorphisms S C, where C is considered as a semigroup
under multiplication.
Proof. The correspondence between (a) and (b) is standard (see [69, Thm. 5 of Ch.
5, 4]). The correspondence between (a) and (c) is special to the toric case.
Given a point p V, dene S C by sending m S to
m
(p) C. This
makes sense since
m
C[S] =C[V]. One easily checks that S C is a semigroup
homomorphism.
Going the other way, let : S C be a semigroup homomorphism. Since

mS
is a basis of C[S], induces a surjective linear map C[S] C which is
a C-algebra homomorphism. The kernel of the map C[S] C is a maximal ideal
and thus gives a point p V by the correspondence between (a) and (b).
We construct p concretely as follows. Let A = m
1
, . . . , m
s
generate S, so
that V =Y
A
C
s
. Let p = ((m
1
), . . . , (m
s
)) C
s
. Let us prove that p V. By
Proposition 1.1.9, it sufces to show that x

vanishes at p for all exponent


vectors = (a
1
, . . . , a
s
) and = (b
1
, . . . , b
s
) satisfying
s

i=1
a
i
m
i
=
s

i=1
b
i
m
i
.
This is easy, since being a semigroup homomorphism implies that
s

i=1
(m
i
)
a
i
=

i=1
a
i
m
i

i=1
b
i
m
i

=
s

i=1
(m
i
)
b
i
.
It is straightforward to show that this point of V agrees with the one constructed in
the previous paragraph (Exercise 1.3.1).
As an application of this result, we describe the torus action on V. In terms of
the embedding V =Y
A
C
s
, the proof of Proposition 1.1.8 shows that the action
of T
N
on Y
A
is induced by the usual action of (C

)
s
on C
s
. But how do we see the
action intrinsically, without embedding into afne space? This is where semigroup
36 Chapter 1. Afne Toric Varieties
homomorphisms prove their value. Fix t T
N
and p V, and let p correspond to
the semigroup homorphism m (m). In Exercise 1.3.1 you will show that t p
is given by the semigroup homomorphism m
m
(t)(m). This description will
prove useful in Chapter 3 when we study torus orbits.
From the point of view of group actions, the action of T
N
on V is given by a
map T
N
V V. Since both sides are afne varieties, this should be a morphism,
meaning that it should come from a C-algebra homomorphism
C[S] =C[V] C[T
N
V] =C[T
N
]
C
C[V] =C[M]
C
C[S].
This homomorphism is given by
m

m
for m S (Exercise 1.3.2).
We next characterize when the torus action on an afne toric variety has a
xed point. An afne semigroup S is pointed if S (S) = 0, i.e., if 0 is the
only invertible element of S. This is the semigroup analog of strongly convex.
Proposition 1.3.2. Let V be an afne toric variety. Then:
(a) If we write V = Spec(C[S]), then the torus action has a xed point if and only
if S is pointed, in which case the unique xed point is given by the semigroup
homomorphism S C dened by
(1.3.1) m

1 m = 0
0 m = 0.
(b) If we write V = Y
A
C
s
for A S ` 0, then the torus action has a xed
point if and only if 0 Y
A
, in which case the unique xed point is 0.
Proof. For part (a), let p V be represented by the semigroup homomorphism
: S C. Then p is xed by the torus action if and only if
m
(t)(m) = (m)
for all m S and t T
N
. This equation is satised for m = 0 since (0) = 1, and if
m = 0, then picking t with
m
(t) = 1 shows that (m) = 0. Thus, if a xed point
exists, then it is unique and is given by (1.3.1). Then we are done since (1.3.1) is a
semigroup homomorphism if and only if S is pointed.
For part (b), rst assume that V = Y
A
C
s
has a xed point, in which case
S = NA is pointed and the unique point p is given by (1.3.1). Then A S`0
and the proof of Proposition 1.3.1 imply that p is the origin in C
s
, so that 0 Y
A
.
The converse follows since 0 C
s
is xed by (C

)
s
, hence by Y
A
(C

)
s
.
Here is a useful corollary of Proposition 1.3.2 (Exercise 1.3.3).
Corollary 1.3.3. Let U

be the afne toric variety of a strongly convex rational


polyhedral cone N
R
. Then the torus action on U

has a xed point if and


only if dim = dim N
R
, in which case the xed point is unique and is given by the
maximal ideal
'
m
[ m S

`0` C[S

],
where as usual S

M.
1.3. Properties of Afne Toric Varieties 37
We will see in Chapter 3 that this corollary is part of the correspondence be-
tween torus orbits of U

and faces of .
Normality and Saturation. We next study the question of when an afne toric
variety V is normal. We need one denition before stating our normality criterion.
Denition 1.3.4. An afne semigroup S M is saturated if for all k N`0 and
m M, km S implies m S.
For example, if N
R
is a strongly convex rational polyhedral cone, then
S

M is easily seen to be saturated (Exercise 1.3.4).


Theorem 1.3.5. Let V be an afne toric variety with torus T
N
. Then the following
are equivalent:
(a) V is normal.
(b) V = Spec(C[S]), where S M is a saturated afne semigroup.
(c) V =Spec(C[S

]) (=U

), where S

M and N
R
is a strongly convex
rational polyhedral cone.
Proof. By Theorem 1.1.17, V =Spec(C[S]) for an afne semigroup S contained in
a lattice, and by Proposition 1.1.14, the torus of V has the character lattice M =ZS.
Also let n = dimV, so that M Z
n
.
(a) (b): If V is normal, then C[S] = C[V] is integrally closed in its eld of
fractions C(V). Suppose that km S for some k N`0 and m M. Then
m
is a polynomial function on T
N
and hence a rational function on V since T
N
V is
Zariski open. We also have
km
C[S] since km S. It follows that
m
is a root
of the monic polynomial X
k

km
with coefcients in C[S]. By the denition of
normal, we obtain
m
C[S], i.e., m S. Thus S is saturated.
(b) (c): Let A S be a nite generating set of S. Then S lies in the rational
polyhedral cone Cone(A) M
R
, and rankZA = n implies dim Cone(A) = n by
Exercise 1.2.6. It follows that = Cone(A)

N
R
is a strongly convex ratio-
nal polyhedral cone such that S

M. In Exercise 1.3.4 you will prove that


equality holds when S is saturated. Hence S = S

.
(c) (a): We need to show that C[S

] =C[

M] is normal when N
R
is
a strongly convex rational polyhedral cone. Let
1
, . . . ,
r
be the rays of . Since
is generated by its rays (Lemma 1.2.15), we have

=
r

i=1

i
.
Intersecting with M gives S

r
i=1
S

i
, which easily implies
C[S

] =
r

i=1
C[S

i
].
38 Chapter 1. Afne Toric Varieties
By Exercise 1.0.7, C[S

] is normal if each C[S

i
] is normal, so it sufces to prove
that C[S

] is normal when is a rational ray in N


R
. Let u

N be the ray
generator of . Since u is primitive, i.e,
1
k
u

/ N for all k > 1, we can nd a


basis e
1
, . . . , e
n
of N with u

= e
1
(Exercise 1.3.5). This allows us to assume that
= Cone(e
1
), so that
C[S

] =C[x
1
, x
1
2
, . . . , x
1
n
]
by Example 1.2.21. But C[x
1
, . . . , x
n
] is normal (it is a UFD), so its localization
C[x
1
, . . . , x
n
]
x
2
xn
=C[x
1
, x
1
2
, . . . , x
1
n
]
is also normal by Exercise 1.0.7. This completes the proof.
Example 1.3.6. We saw in Example 1.2.20 that V = V(xy zw) is the afne toric
variety U

of the cone = Cone(e


1
, e
2
, e
1
+e
3
, e
2
+e
3
) pictured in Figure 2. Then
Theorem 1.3.5 implies that V is normal, as claimed in Example 1.1.5.
Example 1.3.7. By Example 1.2.22, the rational normal cone

C
d
C
d+1
is the
afne toric variety of a strongly convex rational polyhedral cone and hence is nor-
mal by Theorem 1.3.5.
It is instructive to view this example using the parametrization

A
(s, t) = (s
d
, s
d1
t, . . . , st
d1
, t
d
)
from Example 1.1.6. Plotting the lattice points in A for d = 2 gives the white
squares in Figure 9 (a) below. These generate the semigroup S = NA, and the
proof of Theorem 1.3.5 gives the cone

= Cone(e
1
, e
2
), which is the rst quad-
rant in the gure. At rst glance, something seems wrong. The afne variety

C
2
is normal, yet in Figure 9 (a) the semigroup generated by the white squares misses
some lattice points in

. This semigroup does not look saturated. How can the


afne toric variety be normal?
(a)

(b)

Figure 9. Lattice points for the rational normal cone


b
C2
The problem is that we are using the wrong lattice! Proposition 1.1.8 tells us
to use the lattice ZA, which gives the white dots and squares in Figure 9 (b). This
1.3. Properties of Afne Toric Varieties 39
gure shows that the white squares generate the semigroup of lattice points in

.
Hence S is saturated and everything is ne.
This example points out the importance of working with the correct lattice.
The Normalization of an Afne Toric Variety. The normalization of an afne toric
variety is easy to describe. Let V = Spec(C[S]) for an afne semigroup S, so that
the torus of V has character lattice M = ZS. Let Cone(S) denote the cone of any
nite generating set of S and set = Cone(S)

N
R
. In Exercise 1.3.6 you will
prove the following.
Proposition 1.3.8. The above cone is a strongly convex rational polyhedral cone
in N
R
and the inclusion C[S] C[

M] induces a morphism U

V that is the
normalization map of V.
The normalization of an afne toric variety of the form Y
A
is constructed by
applying Proposition 1.3.8 to the afne semigroup NA and the lattice ZA.
Example 1.3.9. Let A =(4, 0), (3, 1), (1, 3), (0, 4) Z
2
. Then

A
(s, t) = (s
4
, s
3
t, st
3
, t
4
)
parametrizes the surface Y
A
C
4
considered in Exercise 1.1.7. This is almost
the rational normal cone

C
4
, except that we have omitted s
2
t
2
. Using (2, 2) =
1
2

(4, 0) +(0, 4)

, we see that NA is not saturated, so that Y


A
is not normal.
Applying Proposition 1.3.8, one sees that the normalization of Y
A
is

C
4
. You
can check this using Normaliz [57] as explained in Example B.3.2. Note also that

C
4
is an afne variety in C
5
, and the normalization map is induced by the obvious
projection C
5
C
4
.
Proposition 1.3.8 tells us that

M is the saturation of the semigroup S. In


the appendix to Chapter 3 we will see that the normalization map U

V con-
structed in Proposition 1.3.8 is onto but not necessarily one-to-one.
Smooth Afne Toric Varieties. Our next goal is to characterize when an afne toric
variety is smooth. Since smooth afne varieties are normal (Proposition 1.0.9),
we need only consider toric varieties U

coming from strongly convex rational


polyhedral cones N
R
.
We rst study U

when has maximal dimension. Then

is strongly convex,
so that S

M has a Hilbert basis H . Furthermore, Corollary 1.3.3 tells us


that the torus action on U

has a unique xed point, denoted here by p

. The
point p

and the Hilbert basis H are related as follows.


Lemma 1.3.10. Let N
R
be a strongly convex rational polyhedral cone of max-
imal dimension and let T
p
(U

) be the Zariski tangent space to the afne toric


variety U

at the above point p

. Then dim T
p
(U

) =[H [.
40 Chapter 1. Afne Toric Varieties
Proof. By Corollary 1.3.3, the maximal ideal of C[S

] corresponding to p

is m =
'
m
[ m S

`0`. Since
m

mS
is a basis of C[S

], we obtain
m =

m,=0
C
m
=

m irreducible
C
m

m reducible
C
m
=


mH
C
m

m
2
.
It follows that dim m/m
2
=[H [. To relate this to the maximal ideal m
U,p
in the
local ring O
U,p
, we use the natural map
m/m
2
m
U,p
/m
2
U,p
which is always an isomorphism (Exercise 1.3.7). Since T
p
(U

) is the dual space


of m
U,p
/m
2
U,p
, we see that dim T
p
(U

) =[H [.
The Hilbert basis H of S

gives U

=Y
H
C
s
, where s =[H [. This afne
embedding is especially nice. Given any afne embedding U

, we have
dim T
p
(U

) by Lemma 1.0.6. In other words, dim T


p
(U

) is a lower bound
on the dimension of an afne embedding. Then Lemma 1.3.10 shows that when
has maximal dimension, the Hilbert basis of S

gives the most efcient afne


embedding of U

.
Example 1.3.11. In Example 1.2.22, we saw that the rational normal cone

C
d

C
d+1
is the toric variety coming from = Cone(de
1
e
2
, e
2
) R
2
and that S

Z
2
is generated by (1, i) for 0 i d. These generators form the Hilbert basis
of S

, so that the Zariski tangent space of 0



C
d
has dimension d +1. Hence C
d+1
in the smallest afne space in which we can embed

C
d
.
We now come to our main result about smoothness. Recall from 1.2 that a
rational polyhedral cone is smooth if it can be generated by a subset of a basis of
the lattice.
Theorem 1.3.12. Let N
R
be a strongly convex rational polyhedral cone. Then
U

is smooth if and only if is smooth. Furthermore, all smooth afne toric


varieties are of this form.
Proof. If an afne toric variety is smooth, then it is normal by Proposition 1.0.9
and hence of the form U

. Also, Example 1.2.21 implies that if is smooth as


a cone, then U

is smooth as a variety. It remains to prove the converse. So x


N
R
such that U

is smooth. Let n = dimU

= dim N
R
.
First suppose that has dimension n and let p

be the point studied in


Lemma 1.3.10. Since p

is smooth in U

, the Zariski tangent space T


p
(U

) has
dimension n by Denition 1.0.7. On the other hand, Lemma 1.3.10 implies that
dim T
p
(U

) is the cardinality of the Hilbert basis H of S

M. Thus
n =[H [ [edges

[ n,
where the rst inequality holds by Proposition 1.2.23 (each edge

con-
tributes an element of H ) and the second holds since dim

= n. It follows
1.3. Properties of Afne Toric Varieties 41
that has n edges and H consists of the ray generators of these edges. Since
M = ZS

by (1.2.3), the n edge generators of

generate the lattice M Z


n
and
hence form a basis of M. Thus

is smooth, and then = (

is smooth since
duality preserves smoothness.
Next suppose dim = r < n. We reduce to the previous case as follows. Let
N
1
N be the smallest saturated sublattice containing the generators of . Then
N/N
1
is torsion-free, which by Exercise 1.3.5 implies the existence of a sublattice
N
2
N with N = N
1
N
2
. Note rankN
1
= r and rankN
2
= nr.
The cone lies in both (N
1
)
R
and N
R
. This gives afne toric varieties U
,N
1
and U
,N
of dimensions r and n respectively. Furthermore, N = N
1
N
2
induces
M =M
1
M
2
, so that (N
1
)
R
and N
R
give the afne semigroups S
,N
1
M
1
and S
,N
M respectively. It is straighforward to show that
S
,N
= S
,N
1
M
2
,
which in terms of semigroup algebras can be written
C[S
,N
] C[S
,N
1
]
C
C[M
2
].
The right-hand side is the coordinate ring of U
,N
1
T
N
2
. Thus
(1.3.2) U
,N
U
,N
1
T
N
2
,
which in turn implies that
U
,N
U
,N
1
(C

)
nr
U
,N
1
C
nr
.
Since we are assuming that U
,N
is smooth, it follows that U
,N
1
C
nr
is smooth
at any point (p, q) in U
,N
1
(C

)
nr
. In Exercise 1.3.8 you will show that
(1.3.3) U
,N
1
C
nr
is smooth at (p, q) = U
,N
1
is smooth at p.
Letting p = p

U
,N
1
, the previous case implies that is smooth in N
1
since
dim = dim (N
1
)
R
. Hence is clearly smooth in N = N
1
N
2
.
Equivariant Maps between Afne Toric Varieties. We next study maps V
1
V
2
between afne toric varieties that respect the torus actions on V
1
and V
2
.
Denition 1.3.13. Let V
i
= Spec(C[S
i
]) be the afne toric varieties coming from
the afne semigroups S
i
, i = 1, 2. Then a morphism : V
1
V
2
is toric if the cor-
responding map of coordinate rings

: C[S
2
] C[S
1
] is induced by a semigroup
homomorphism

: S
2
S
1
.
Here is our rst result concerning toric morphisms.
Proposition 1.3.14. Let T
N
i
be the torus of the afne toric variety V
i
, i =1, 2. Then:
(a) A morphism : V
1
V
2
is toric if and only if
(T
N
1
) T
N
2
and [
T
N
1
: T
N
1
T
N
2
is a group homomorphism.
42 Chapter 1. Afne Toric Varieties
(b) A toric morphism : V
1
V
2
is equivariant, meaning that
(t p) = (t) (p)
for all t T
N
1
and p V
1
.
Proof. Let V
i
= Spec(C[S
i
]), so that the character lattice of T
N
i
is M
i
= ZS
i
. If
comes from a semigroup homomorphism

: S
2
S
1
, then

extends to a group
homomorphism

: M
2
M
1
and hence gives a commutative diagram
C[S
2
]

C[S
1
]

C[M
2
] C[M
1
].
Applying Spec, we see that (T
N
1
) T
N
2
, and [
T
N
1
: T
N
1
T
N
2
is a group homo-
morphism since T
N
i
= Hom
Z
(M
i
, C

) by Exercise 1.1.11. Conversely, if sat-


ises these conditions, then [
T
N
1
: T
N
1
T
N
2
induces a diagram as above where
the bottom map comes from a group homomorphism

: M
2
M
1
. This, com-
bined with

(C[S
2
]) C[S
1
], implies that

induces a semigroup homomorphism

: S
2
S
1
. This proves part (a) of the proposition.
For part (b), suppose that we have a toric map : V
1
V
2
. The action of T
N
i
on V
i
is given by a morphism
i
: T
N
i
V
i
V
i
, and equivariance means that we
have a commutative diagram
T
N
1
V
1

[
T
N
1

V
1

T
N
2
V
2

V
2
.
If we replace V
i
with T
N
i
in the diagram, then it certainly commutes since [
T
N
1
is a group homomorphism. Then the whole diagram commutes since T
N
1
T
N
1
is
Zariski dense in T
N
1
V
1
.
We can also characterize toric morphisms between afne toric varieties coming
from strongly convex rational polyhedral cones. First note that a homomorphism
: N
1
N
2
of lattices gives a group homomorphism : T
N
1
T
N
2
of tori. This
follows from T
N
i
= N
i

Z
C

, and one sees that is a morphism. Also, tensoring


with R gives
R
: (N
1
)
R
(N
2
)
R
.
Here is the result, whose proof we leave to the reader (Exercise 1.3.9).
Proposition 1.3.15. Suppose we have strongly convex rational polyhedral cones

i
(N
i
)
R
and a homomorphism : N
1
N
2
. Then : T
N
1
T
N
2
extends to a
map of afne toric varieties : U

1
U

2
if and only if
R
(
1
)
2
.
1.3. Properties of Afne Toric Varieties 43
Faces and Afne Open Subsets. Let N
R
be a strongly convex rational poly-
hedral cone and let _ be a face. Then we can nd m

M such that
= H
m
. This allows us to relate the semigroup algebras of and as follows.
Proposition 1.3.16. Let be a face of and as above write = H
m
, where
m

M. Then the semigroup algebra C[S

] =C[

M] is the localization of
C[S

] =C[

M] at
m
C[S

]. In other words,
C[S

] =C[S

m.
Proof. The inclusion implies S

, and since 'm, u` = 0 for all u , we


have m

. It follows that
S

+Z(m) S

.
This inclusion is actually an equality, as we now prove. Fix a nite set S N with
= Cone(S) and pick m

. Set
C = max
uS
['m

, u`[ N.
It is straightforward to show that m

+Cm S

. This proves that


S

+Z(m) = S

,
from which C[S

] =C[S

m follows immediately.
This interprets nicely in terms of toric morphisms. By Proposition 1.3.15, the
identity map N N and the inclusion give the toric morphismU

that
corresponds to the inclusion C[S

] C[S

]. By Proposition 1.3.16,
(1.3.4) U

= Spec(C[S

]) = Spec(C[S

m) = Spec(C[S

])

m = (U

m U

.
Thus U

becomes an afne open subset of U

when _ . In particular, if two


cones and

intersect in a common face =

, then we have inclusions


U

.
We will use this in Chapters 2 and 3 when we glue together afne toric varieties to
create more general toric varieties.
The role of faces is the key reason why we describe afne toric varieties using
N
R
rather than

M
R
. This answers the question raised in Remark 1.2.19.
Sublattices of Finite Index and Rings of Invariants. Another interesting class of
toric morphisms arises when we keep the same cone but change the lattice. Here is
an example we have already seen.
Example 1.3.17. In Example 1.3.7 the dual of = Cone(e
1
, e
2
) R
2
interacts
with the lattices shown in Figure 10 on the next page. To make this precise, let us
name the lattices involved: the lattices
N

=Z
2
N =(a/2, b/2) [ a, b Z, a+b 0 mod 2
44 Chapter 1. Afne Toric Varieties
(a)

(b)

Figure 10. Lattice points of

relative to two lattices


have N

R
N
R
, and the dual lattices
M

=Z
2
M =(a, b) [ a, b Z, a+b 0 mod 2
have

R
M
R
. Note that duality reverses inclusions and that M and N are
indeed dual under dot product. In Figure 10 (a), the black dots in the rst quadrant
form the semigroup S
,N

, and in Figure 10 (b), the white dots in the


rst quadrant form S
,N
=

M.
This gives the afne toric varieties U
,N
and U
,N
. Clearly U
,N
=C
2
since
is smooth for N

, while Example 1.3.7 shows that U


,N
is the rational normal cone

C
2
. The inclusion N

N gives a toric morphism


C
2
=U
,N
U
,N
=

C
2
.
Our next task is to nd a nice description of this map.
In general, suppose we have lattices N

N, where N

has nite index in N,


and let N

R
= N
R
be a strongly convex rational polyhedral cone. Then the
inclusion N

N gives the toric morphism


: U
,N
U
,N
.
The dual lattices satisfy M

M, so that corresponds to the inclusion


C[

] C[

M]
of semigroup algebras. The idea is to realize C[

M] as a ring of invariants of a
group action on C[

].
Proposition 1.3.18. Let N

have nite index in N with quotient G = N/N

and let
N

R
= N
R
be a strongly convex rational polyhedral cone. Then:
(a) There are natural isomorphisms
GHom
Z
(M

/M, C

) = ker(T
N
T
N
).
1.3. Properties of Afne Toric Varieties 45
(b) G acts on C[

] with ring of invariants


C[

]
G
=C[

M].
(c) G acts on U
,N
, and the morphism : U
,N
U
,N
is constant on G-orbits
and induces a bijection
U
,N
/G U
,N
.
Proof. Since T
N
= Hom
Z
(M, C

) by Exercise 1.1.11, applying Hom


Z
(, C

) to
0 M M

/M 0
gives the sequence
1 Hom
Z
(M

/M, C

) T
N
T
N
1.
This is exact since Hom
Z
(, C

) is left exact and C

is divisible. Note also that


since N

has nite index in N, we have inclusions


N

N N
Q
and M M

M
Q
.
The pairing between M and N induces a pairing M
Q
N
Q
Q. Hence the map
M

/MN/N

([m

], [u]) e
2i/m

,u)
is well-dened and induces G = N/N

Hom
Z
(M

/M, C

) (Exercise 1.3.10).
The action of T
N
on U
,N
induces an action of G on U
,N
since G T
N
.
Using Exercise 1.3.1, one sees that if g G and U
,N
, then g is dened by
the semigroup homomorphism m

g([m

])(m

) for m

. It follows that
the corresponding action on the coordinate ring is given by
g
m

= g([m

])
1

, m

.
(Exercise 5.0.1 explains why we need the inverse.) Since m

lies in M if and
only if g([m

]) = 1 for all g G, the ring of invariants


C[

]
G
=f C[

] [ g f = f for all g G,
is precisely C[

M], i.e.,
C[

]
G
=C[

M].
This proves part (b).
When a nite group G acts algebraically on C
n
, [69, Thm. 10 of Ch. 7, 4]
shows that the ring of invariants C[x
1
, . . . , x
n
]
G
C[x
1
, . . . , x
n
] gives a morphism of
afne varieties
C
n
= Spec(C[x
1
, . . . , x
n
]) Spec(C[x
1
, . . . , x
n
]
G
)
that is constant on G-orbits and induces a bijection
C
n
/G Spec(C[x
1
, . . . , x
n
]
G
).
46 Chapter 1. Afne Toric Varieties
The proof extends without difculty to the case when G acts algebraically on V =
Spec(R). Here, R
G
R gives a morphism of afne varieties
V = Spec(R) Spec(R
G
)
that is constant on G-orbits and induces a bijection
V/G Spec(R
G
).
From here, part (c) follows immediately from part (b).
We will give a careful treatment of these ideas in 5.0, where we will show
that the map Spec(R) Spec(R
G
) is a geometric quotient.
Here are some examples of Proposition 1.3.18.
Example 1.3.19. In the situation of Example 1.3.17, one computes that G is the
group
2
= 1 acting on U
,N
= Spec(C[s, t]) C
2
by 1 (s, t) = (s, t).
Thus the rational normal cone

C
2
is the quotient
C
2
/
2
=U
,N
/
2
U
,N
=

C
2
.
We can see this explicitly as follows. The invariant ring is easily seen to be
C[s, t]

2
=C[s
2
, st, t
2
] =C[

C
2
] C[x
0
, x
1
, x
2
]/'x
0
x
2
x
2
1
`,
where the last isomorphism follows from Example 1.1.6. From the point of view
of invariant theory, the generators s
2
, st, t
2
of the ring of invariants give a morphism
: C
2
C
3
, (s, t) (s
2
, st, t
2
)
that is constant on
2
-orbits. This map also separates orbits, so it induces
C
2
/
2
(C
2
) =

C
2
,
where the last equality is by Example 1.1.6. But we can also think about this in
terms of semigroups, where the exponent vectors of s
2
, st, t
2
give the Hilbert basis
of the semigroup S
,N
pictured in Figure 10 (b). Everything ts together very
nicely.
In Exercise 1.3.11 you will generalize Example 1.3.19 to the case of the ratio-
nal normal cone

C
d
for arbitrary d.
Example 1.3.20. Let N
R
R
n
be a simplicial cone of dimension n with ray
generators u
1
, . . . , u
n
. Then N

n
i=1
Zu
i
is a sublattice of nite index in N.
Furthermore, is smooth relative to N

, so that U
,N
= C
n
. It follows that G =
N/N

acts on C
n
with quotient
C
n
/G =U
,N
/G U
,N
.
Hence the afne toric variety of a simplicial cone is the quotient of afne space by
a nite abelian group. In the literature, varieties like U
,N
are called orbifolds and
are said to be Q-factorial.
1.3. Properties of Afne Toric Varieties 47
Exercises for 1.3.
1.3.1. Consider the afne toric variety Y
A
= Spec(C[S]), where A = m
1
, . . . , m
s
and
S =NA. Let : S C be a semigroup homomorphism. In the proof of Proposition 1.3.1
we showed that p = ((m
1
), . . . , (m
s
)) lies in Y
A
.
(a) Prove that the maximal ideal f C[S] [ f (p) = 0 is the kernel of the C-algebra
homomorphismC[S] C induced by .
(b) The torus T
N
of Y
A
has character lattice M =ZA and x t T
N
. As in the discussion
following Proposition 1.3.1, t p comes from the semigroup homomorphism m

m
(t)(m). Prove that this corresponds to the point
(
m1
(t), . . . ,
ms
(t)) ((m
1
), . . . , (m
s
)) = (
m1
(t)(m
1
), . . . ,
ms
(t)(m
s
)) C
s
coming from the action of t T
N
(C

)
s
on p Y
A
C
s
.
1.3.2. Let V = Spec(C[S]) with T
N
= Spec(C[M]), M = ZS. The action T
N
V V
comes from a C-algebra homomorphismC[S] C[M]
C
C[S]. Prove that this homomor-
phism is given by
m

m
. Hint: Show that this formula determines the C-algebra
homomorphismC[M] C[M]
C
C[M] that gives the group operation T
N
T
N
T
N
.
1.3.3. Prove Corollary 1.3.3.
1.3.4. Let A M be a nite set.
(a) Prove that the semigroup NA is saturated in M if and only if NA = Cone(A) M.
Hint: Apply (1.2.2) to Cone(A) M
R
.
(b) Complete the proof of (b) (c) from Theorem 1.3.5.
1.3.5. Let N be a lattice.
(a) Let N
1
N be a sublattice such that N/N
1
is torsion-free. Prove that there is a sublat-
tice N
2
N such that N = N
1
N
2
.
(b) Let u N be primitive as dened in the proof of Theorem 1.3.5. Prove that N has a
basis e
1
, . . . , e
n
such that e
1
= u.
1.3.6. Prove Proposition 1.3.8.
1.3.7. Let p be a point of an irreducible afne variety V. Then p gives the maximal ideal
m = f C[V] [ f (p) = 0 as well as the maximal ideal m
V,p
O
V,p
dened in 1.0.
Prove that the natural map m/m
2
m
V,p
/m
2
V,p
is an isomorphism of C-vector spaces.
1.3.8. Prove (1.3.3). Hint: Use Lemma 1.0.6 and Example 1.0.10.
1.3.9. Prove Proposition 1.3.15.
1.3.10. Prove the assertions made in the proof of Proposition 1.3.18 concerning the pairing
M

/MN/N

dened by ([m

], [u]) e
2im

,u
.
1.3.11. Let
d
= C

[
d
= 1 be the group of dth roots of unity. Then
d
acts on C
2
by (x, y) = (x, y). Adapt Example 1.3.19 to show that C
2
/
d


C
d
. Hint: Use lattices
N

=Z
2
N =(a/d, b/d) [ a, b Z, a +b 0 mod d.
1.3.12. Prove that the normalization map in Proposition 1.3.8 is a toric morphism.
1.3.13. Let
1
(N
1
)
R
and
2
(N
2
)
R
be strongly convex rational polyhedral cones. This
gives the cone
1

2
(N
1
N
2
)
R
. Prove that U
12
U
1
U
2
. Also explain how
this result applies to (1.3.2).
48 Chapter 1. Afne Toric Varieties
1.3.14. By Proposition 1.3.1, a point p of an afne toric variety V = Spec(C[S]) is repre-
sented by a semigroup homomorphism : S C. Prove that p lies in the torus of V if and
only if never vanishes, i.e., (m) = 0 for all m S.
Appendix: Tensor Products of Coordinate Rings
In this appendix, we will prove the following result used in 1.0 in our discussion of prod-
ucts of afne varieties.
Proposition 1.A.1. If R and S are nitely generated C-algebras without nilpotents, then
the same is true for R
C
S.
Proof. The tensor product is clearly a nitely generated C-algebra. Hence we need only
prove that R
C
S has no nilpotents. If we write R C[x
1
, . . . , x
n
]/I, then I is radical and
thus has a primary decomposition I =

s
i=1
P
i
, where each P
i
is prime (see [69, Ch. 4, 7]).
This gives
R C[x
1
, . . . , x
n
]/I
s

i=1
C[x
1
, . . . , x
n
]/P
i
where the map to the direct sum is injective. Each quotient C[x
1
, . . . , x
n
]/P
i
is an integral
domain and hence injects into its eld of fractions K
i
. This yields an injection
R
s

i=1
K
i
,
and since tensoring over a eld preserves exactness, we get an injection
R
C
S
s

i=1
K
i

C
S.
Hence it sufces to prove that K
C
S has no nilpotents when K is a nitely generated eld
extension of C. A similar argument using S then reduces us to showing that K
C
L has no
nilpotents when K and L are nitely generated eld extensions of C.
Since C has characteristic 0, the extension C L has a separating transcendence basis
(see [159, p. 519]). This means that we can nd y
1
, . . . , y
t
L such that y
1
, . . . , y
t
are
algebraically independent over C and F =C(y
1
, . . . , y
t
) L is a nite separable extension.
Then
K
C
L K
C
(F
F
L) (K
C
F)
F
L.
But C = K
C
F = K
C
C(y
1
, . . . , y
t
) = K(y
1
, . . . , y
t
) is a eld, so that we are reduced to
considering
C
F
L
where C and L are extensions of F and F L is nite and separable. The latter and the
theorem of the primitive element imply that L F[X]/' f (X)`, where f (X) has distinct
roots in some extension of F. Then
C
F
L C
F
F[X]/' f (X)` C[X]/' f (X)`.
Since f (X) has distinct roots, this quotient ring has no nilpotents. Our result follows.
A nal remark is that we can replace C with any perfect eld since nitely generated
extensions of perfect elds have separating transcendence bases (see [159, p. 519]).
Chapter 2
Projective Toric Varieties
2.0. Background: Projective Varieties
Our discussion assumes that the reader is familiar with the elementary theory of
projective varieties, at the level of [69, Ch. 8].
In Chapter 1, we introduced afne toric varieties. In general, a toric variety is
an irreducible variety X over C containing a torus T
N
(C

)
n
as a Zariski open set
such that the action of (C

)
n
on itself extends to an action on X. We will learn in
Chapter 3 that the concept of variety is somewhat subtle. Hence we will defer
the formal denition of toric variety until then and instead concentrate on toric
varieties that live in projective space P
n
, dened by
(2.0.1) P
n
= (C
n+1
`0)/C

,
where C

acts via homotheties, i.e., (a


0
, . . . , a
n
) =(a
0
, . . . , a
n
) for C

and
(a
0
, . . . , a
n
) C
n+1
. Thus (a
0
, . . . , a
n
) are homogeneous coordinates of a point in
P
n
and are well-dened up to homothety.
The goal of this chapter is to use lattice points and polytopes to create toric
varieties that lie in P
n
. We will use the afne semigroups and polyhedral cones
introduced in Chapter 1 to describe the local structure of these varieties.
Homogeneous Coordinate Rings. A projective variety V P
n
is dened by the
vanishing of nitely many homogeneous polynomials in the polynomial ring S =
C[x
0
, . . . , x
n
]. The homogeneous coordinate ring of V is the quotient ring
C[V] = S/I(V),
where I(V) is generated by all homogeneous polynomials that vanish on V.
49
50 Chapter 2. Projective Toric Varieties
The polynomial ring S is graded by setting deg(x
i
) = 1. This gives the decom-
position S =

d=0
S
d
, where S
d
is the vector space of homogeneous polynomials
of degree d. Homogeneous ideals decompose similarly, and the above coordinate
ring C[V] inherits a grading where
C[V]
d
= S
d
/I(V)
d
.
The ideal I(V) S = C[x
0
, . . . , x
n
] also denes an afne variety

V C
n+1
, called
the afne cone of V. The variety

V satises
(2.0.2) V = (

V `0)/C

,
and its coordinate ring is the homogeneous coordinate ring of V, i.e.,
C[

V] =C[V].
Example 2.0.1. In Example 1.1.6 we encountered the ideal
I ='x
i
x
j+1
x
i+1
x
j
[ 0 i < j d 1` C[x
0
, . . . , x
d
]
generated by the 22 minors of the matrix

x
0
x
1
x
d2
x
d1
x
1
x
2
x
d1
x
d

.
Since I is homogeneous, it denes a projective variety C
d
P
d
that is the image of
the map
: P
1
P
d
dened in homogeneous coordinates by (s, t) (s
d
, s
d1
t, . . . , st
d1
, t
d
) (see Ex-
ercise 1.1.1). This shows that C
d
is a curve, called the rational normal curve of
degree d. Furthermore, the afne cone of C
d
is the rational normal cone

C
d
C
d+1
discussed in Example 1.1.6.
We know from Chapter 1 that

C
d
is an afne toric surface; we will soon see
that C
d
is a projective toric curve.
Example 2.0.2. The afne toric variety V(xy zw) C
4
studied in Chapter 1 is
the afne cone of the projective surface V = V(xy zw) P
3
. Recall that this
surface is isomorphic to P
1
P
1
via the Segre embedding
P
1
P
1
P
3
given by (s, t; u, v) (su, tv, sv, tu). We will see below that V P
1
P
1
is the
projective toric variety coming from the unit square in the plane.
As in the afne case, a projective variety V P
n
has the classical topology,
induced from the usual topology on P
n
, and the Zariski topology, where the Zariski
closed sets are subvarieties of V (meaning projective varieties of P
n
contained in V)
and the Zariski open sets are their complements.
2.0. Background: Projective Varieties 51
Rational Functions on Irreducible Projective Varieties. A homogeneous polyno-
mial f S of degree d > 0 does not give a function on P
n
since
f (x
0
, . . . , x
n
) =
d
f (x
0
, . . . , x
n
).
However, the quotient of two such polynomials f , g S
d
gives the well-dened
function
f
g
: P
n
`V(g) C.
provided g = 0. We write this as f /g : P
n
C and say that f /g is a rational
function on P
n
.
More generally, suppose that V P
n
is irreducible, and let f , g C[V] =C[

V]
be homogeneous of the same degree with g = 0. Then f and g give functions on
the afne cone

V and hence an element f /g C(

V). By (2.0.2), this induces a


rational function f /g : V C. Thus
C(V) =f /g C(

V) [ f , g C[V] homogeneous of the same degree, g = 0


is the eld of rational functions on V. It is customary to write the set on the left as
C(

V)
0
since it consists of the degree 0 elements of C(

V).
Afne Pieces of Projective Varieties. A projective variety V P
n
is a union of
Zariski open sets that are afne. To see why, let U
i
=P
n
`V(x
i
). Then U
i
C
n
via
the map
(2.0.3) (a
0
, . . . , a
n
)

a
0
a
i
, . . . ,
a
i1
a
i
,
a
i+1
a
i
, . . . ,
an
a
i

,
so that in the notation of Chapter 1, we have
U
i
= Spec

x
0
x
i
, . . . ,
x
i1
x
i
,
x
i+1
x
i
, . . . ,
xn
x
i

.
Then V U
i
is a Zariski open subset of V that maps via (2.0.3) to the afne variety
in C
n
dened by the equations
(2.0.4) f

x
0
x
i
, . . . ,
x
i1
x
i
, 1,
x
i+1
x
i
, . . . ,
xn
x
i

= 0
as f varies over all homogeneous polynomials in I(V).
We call V U
i
an afne piece of V. These afne pieces cover V since the U
i
cover P
n
. Using localization, we can describe the coordinate rings of the afne
pieces as follows. The variable x
i
induces an element x
i
C[V], so that we get the
localization
(2.0.5) C[V]
x
i
=f / x
k
i
[ f C[V], k 0
as in Exercises 1.0.2 and 1.0.3. Note that C[V]
x
i
has a well-dened Z-grading given
by deg( f / x
k
i
) = deg( f ) k when f is homogeneous. Then
(2.0.6) (C[V]
x
i
)
0
=f / x
k
i
C[V]
x
i
[ f is homogeneous of degree k
is the subring of C[V]
x
i
consisting of all elements of degree 0. This gives an afne
piece of V as follows.
52 Chapter 2. Projective Toric Varieties
Lemma 2.0.3. The afne piece V U
i
of V has coordinate ring
C[V U
i
] (C[V]
x
i
)
0
.
Proof. We have an exact sequence
0 I(V) C[x
0
, . . . , x
n
] C[V] 0.
If we localize at x
i
, we get the exact sequence
(2.0.7) 0 I(V)
x
i
C[x
0
, . . . , x
n
]
x
i
C[V]
x
i
0
since localization preserves exactness (Exercises 2.0.1 and 2.0.2). These sequences
preserve degrees, so that taking elements of degree 0 gives the exact sequence
0 (I(V)
x
i
)
0
(C[x
0
, . . . , x
n
]
x
i
)
0
(C[V]
x
i
)
0
0.
Note that (C[x
0
, . . . , x
n
]
x
i
)
0
= C

x
0
x
i
, . . . ,
x
i1
x
i
,
x
i+1
x
i
, . . . ,
xn
x
i

. If f I(V) is homoge-
neous of degree k, then
f /x
k
i
= f

x
0
x
i
, . . . ,
x
i1
x
i
, 1,
x
i+1
x
i
, . . . ,
xn
x
i

(I(V)
x
i
)
0
.
By (2.0.4), we conclude that (I(V)
x
i
)
0
maps to I(V U
i
). To show that this map
is onto, let g

x
0
x
i
, . . . ,
x
i1
x
i
,
x
i+1
x
i
, . . . ,
xn
x
i

I(V U
i
). For k 0, x
k
i
g = f (x
0
, . . . , x
n
)
is homogeneous of degree k. It then follows easily that x
i
f vanishes on V since
g = 0 on V U
i
and x
i
= 0 on the complement of U
i
. Thus x
i
f I(V), and then
(x
i
f )/(x
k+1
i
) (I(V)
x
i
)
0
maps to g. The lemma follows immediately.
One can also explore what happens when we intersect afne pieces V U
i
and
V U
j
for i = j. By Exercise 2.0.3, V U
i
U
j
is afne with coordinate ring
(2.0.8) C[V U
i
U
j
] (C[V]
x
i
x
j
)
0
.
We will apply this to projective toric varieties in 2.2. We will also see later in
the book that Lemma 2.0.3 is related to the Proj construction, where Proj of a
graded ring gives a projective variety, just as Spec of an ordinary ring gives an
afne variety.
Products of Projective Spaces. One can study the product P
n
P
m
of projective
spaces using the bigraded ring C[x
0
, . . . , x
n
, y
0
, . . . , y
m
], where x
i
has bidegree (1, 0)
and y
i
has bidegree (0, 1). Then a bihomogeneous polynomial f of bidegree (a, b)
gives a well-dened equation f = 0 in P
n
P
m
. This allows us to dene varieties
in P
n
P
m
using bihomogeneous ideals. In particular, the ideal I(V) of a variety
V P
n
P
m
is a bihomogeneous ideal.
Another way to study P
n
P
m
is via the Segre embedding
P
n
P
m
P
nm+n+m
dened by mapping (a
0
, . . . , a
n
, b
0
, . . . , b
m
) to the point
(a
0
b
0
, a
0
b
1
, . . . , a
0
b
m
, a
1
b
0
, . . . , a
1
b
m
, . . . , a
n
b
0
, . . . , a
n
b
m
).
2.0. Background: Projective Varieties 53
This map is studied in [69, Ex. 14 of Ch. 8, 4]. If P
nm+n+m
has homogeneous
coordinates x
i j
for 0 i n, 0 j m, then P
n
P
m
P
nm+n+m
is dened by the
vanishing of the 22 minors of the (n+1) (m+1) matrix

x
00
x
0m
.
.
.
.
.
.
x
n0
x
nm

.
This follows since an (n +1) (m+1) matrix has rank 1 if and only if it is a
product A
t
B, where A and B are nonzero row matrices of lengths n+1 and m+1.
These approaches give the same notion of what it means to be a subvariety of
P
n
P
m
. A homogeneous polynomial F(x
i j
) of degree d gives the bihomogeneous
polynomial F(x
i
y
j
) of bidegree (d, d). Hence any subvariety of P
nm+n+m
lying
in P
n
P
m
can be dened by a bihomogeneous ideal in C[x
0
, . . . , x
n
, y
0
, . . . , y
m
].
Going the other way takes more thought and is discussed in Exercise 2.0.5.
We also have the following useful result proved in Exercise 2.0.6.
Proposition 2.0.4. Given subvarieties V P
n
and W P
m
, the product V W is
a subvariety of P
n
P
m
.
Weighted Projective Space. The graded ring associated to projective space P
n
is
C[x
0
, . . . , x
n
], where each variable x
i
has degree 1. More generally, let q
0
, . . . , q
n
be
positive integers with gcd(q
0
, . . . , q
n
) = 1 and dene
P(q
0
, . . . , q
n
) = (C
n+1
`0)/ ,
where is the equivalence relation
(a
0
, . . . , a
n
) (b
0
, . . . , b
n
) a
i
=
q
i
b
i
, i = 0, . . . , n for some C

.
We call P(q
0
, . . . , q
n
) a weighted projective space. Note that P
n
=P(1, . . . , 1).
The ring corresponding to P(q
0
, . . . , q
n
) is the graded ring C[x
0
, . . . , x
n
], where
x
i
now has degree q
i
. A polynomial f is weighted homogeneous of degree d if every
monomial x

appearing in f satises (q
0
, . . . , q
n
) = d. The equation f = 0 is
well-dened on P(q
0
, . . . , q
n
) when f is weighted homogeneous, and one can dene
varieties in P(q
0
, . . . , q
n
) using weighted homogeneous ideals of C[x
0
, . . . , x
n
].
Example 2.0.5. We can embed the weighted projective plane P(1, 1, 2) in P
3
using
the monomials x
2
0
, x
0
x
1
, x
2
1
, x
2
of weighted degree 2. In other words, the map
P(1, 1, 2) P
3
given by
(a
0
, a
1
, a
2
) (a
2
0
, a
0
a
1
, a
2
1
, a
2
)
is well-dened and injective. One can check that this map induces
P(1, 1, 2) V(y
0
y
2
y
2
1
) P
3
,
where y
0
, y
1
, y
2
, y
3
are homogeneous coordinates on P
3
.
54 Chapter 2. Projective Toric Varieties
Later in the book we will use toric methods to construct projective embeddings
of arbitrary weighted projective spaces.
Exercises for 2.0.
2.0.1. Let R be a commutative C-algebra. Given f R` 0 and an exact sequence of
R-modules 0 M
1
M
2
M
3
0, prove that
0 M
1

R
R
f
M
2

R
R
f
M
3

R
R
f
0
is also exact, where R
f
is the localization of R at f dened in Exercises 1.0.2 and 1.0.3.
2.0.2. Let V P
n
be a projective variety. If we set S =C[x
0
, . . . , x
n
], then V has coordinate
ring C[V] = S/I(V). Let x
i
be the image of x
i
in C[V].
(a) Note that C[V] is an S-module. Prove that C[V]
xi
C[V]
S
S
xi
.
(b) Use part (a) and the previous exercise to prove that (2.0.7) is exact.
2.0.3. Prove the claim made in (2.0.8).
2.0.4. Let V P
n
be a projective variety. Take f
0
, . . . , f
m
S
d
such that the intersection
V V( f
0
, . . . , f
m
) is empty. Prove that the map
(a
0
, . . . , a
n
) ( f
0
(a
0
, . . . , a
n
), . . . , f
m
(a
0
, . . . , a
n
))
induces a well-dened function : V P
m
.
2.0.5. Let V P
n
P
m
be dened by f

(x
i
, y
j
) = 0, where f

(x
i
, y
j
) is bihomogenous of
bidegree (a

, b

), = 1, . . . , s. The goal of this exercise is to show that when we embed


P
n
P
m
in P
nm+n+m
via the Segre embedding described in the text, V becomes a subvariety
of P
nm+n+m
.
(a) For each , pick an integer d

maxa

, b

and consider the polynomials f


,,
=
x

(x
i
, y
j
) where = 1, . . . , s and [[ = d

, [[ = d

. Note that f
,,
is
bihomogenous of bidegree (d

, d

). Prove that V P
n
P
m
is dened by the vanishing
of the f
,,
.
(b) Use part (a) to show that V is a subvariety of P
nm+n+m
under the Segre embedding.
2.0.6. Prove Proposition 2.0.4
2.0.7. Consider the Segre embedding P
1
P
1
P
3
. Show that after relabeling coordi-
nates, the afne cone of P
1
P
1
in P
3
is the variety V(xy zw) C
4
featured in many
examples in Chapter 1.
2.1. Lattice Points and Projective Toric Varieties
We rst observe that P
n
is a toric variety with torus
T
P
n =P
n
`V(x
0
x
n
) =(a
0
, . . . , a
n
) P
n
[ a
0
a
n
= 0
=(1, t
1
, . . . , t
n
) P
n
[ t
1
, . . . , t
n
C

(C

)
n
.
The action of T
P
n on itself clearly extends to an action on P
n
, making P
n
a toric
variety. To describe the lattices associated to T
P
n , we use the exact sequence of tori
1 C

(C

)
n+1

T
P
n 1
2.1. Lattice Points and Projective Toric Varieties 55
coming from the denition (2.0.1) of P
n
. Hence the character lattice of T
P
n is
(2.1.1) M
n
=(a
0
, . . . , a
n
) Z
n+1
[

n
i=0
a
i
= 0,
and the lattice of one-parameter subgroups N
n
is the quotient
N
n
=Z
n+1
/Z(1, . . . , 1).
Lattice Points and Projective Toric Varieties. Let T
N
be a torus with lattices M
and N as usual. In Chapter 1, we used a nite set of lattice points of A =
m
1
, . . . , m
s
M to create the afne toric variety Y
A
as the Zariski closure of
the image of the map

A
: T
N
C
s
, t (
m
1
(t), . . . ,
ms
(t)).
To get a projective toric variety, we regard
A
as a map to (C

)
s
and compose
with the homomorphism : (C

)
s
T
P
s1 to obtain
(2.1.2) T
N

A
C
s

T
P
s1 P
s1
.
Denition 2.1.1. Given a nite set A M, the projective toric variety X
A
is the
Zariski closure in P
s1
of the image of the map
A
from (2.1.2).
Proposition 2.1.2. X
A
is a toric variety of dimension equal to the dimension of the
smallest afne subspace of M
R
containing A.
Proof. The proof that X
A
P
s1
is a toric variety is similar to the proof given
in Proposition 1.1.8 of Chapter 1 that Y
A
C
s
is a toric variety. The assertion
concerning the dimension of X
A
will follow from Proposition 2.1.6 below.
More concretely, X
A
is the Zariski closure of the image of the map
T
N
P
s1
, t (
m
1
(t), . . . ,
ms
(t))
given by the characters coming from A = m
1
, . . . , m
s
M. In particular, if
M =Z
n
, then
m
i
is the Laurent monomial t
m
i
and X
A
is the Zariski closure of the
image of
T
N
P
s1
, t (t
m
1
, . . . , t
ms
).
In the literature, A Z
n
is often given as an ns matrix A with integer entries, so
that the elements of A are the columns of A.
Here is an example where the lattice points themselves are matrices.
Example 2.1.3. Let M =Z
33
be the lattice of 33 integer matrices and let
P
3
=33 permutation matrices Z
33
.
Write C[M] =C[t
1
1
, . . . , t
1
9
], where the variables give the generic 33 matrix

t
1
t
2
t
3
t
4
t
5
t
6
t
7
t
8
t
9

56 Chapter 2. Projective Toric Varieties


with nonzero entries. Also let P
5
have homogeneous coordinates x
i jk
indexed by
triples such that

1 2 3
i j k

is a permutation in S
3
. Then X
P
3
P
5
is the Zariski
closure of the image of the map T
N
P
5
given by the Laurent monomials t
i
t
j
t
k
for

1 2 3
i j k

S
3
. The ideal of X
P
3
is
I(X
P
3
) ='x
123
x
231
x
312
x
132
x
321
x
213
` C[x
i jk
],
where the relation comes from the fact that the sum of the permutation matrices
corresponding to x
123
, x
231
, x
312
equals the sum of the other three (Exercise 2.1.1).
Ideals of the toric varieties arising from permutation matrices have applications to
sampling problems in statistics (see [264, p. 148]).
The Afne Cone of a Projective Toric Variety. The projective variety X
A
P
s1
has an afne cone

X
A
C
s
. How does

X
A
relate to the afne toric variety Y
A
C
s
constructed in Chapter 1?
Recall from Chapter 1 that when A = m
1
, . . . , m
s
M, the map e
i
m
i
induces an exact sequence
(2.1.3) 0 L Z
s
M
and that the ideal of Y
A
is the toric ideal
I
L
=

[ , N
s
and L

(Proposition 1.1.9). Then we have the following result.


Proposition 2.1.4. Given Y
A
, X
A
and I
L
as above, the following are equivalent:
(a) Y
A
C
s
is the afne cone

X
A
of X
A
P
s1
.
(b) I
L
= I(X
A
).
(c) I
L
is homogeneous.
(d) There is u N and k > 0 in N such that 'm
i
, u` = k for i = 1, . . . , s.
Proof. The equivalence (a) (b) follows from the equalities I(X
A
) = I(

X
A
) and
I
L
= I(Y
A
), and the implication (b) (c) is obvious.
For (c) (d), assume that I
L
is a homogeneous ideal and take x

I
L
for L. If x

and x

had different degrees, then x

, x

I
L
= I(Y
A
) would
vanish on Y
A
. This is impossible since (1, . . . , 1) Y
A
by (2.1.2). Hence x

and
x

have the same degree, which implies that (1, . . . , 1) = 0 for all L. Now
tensor (2.1.3) with Q and take duals to obtain an exact sequence
N
Q
Q
s
Hom
Q
(L
Q
, Q) 0.
The above argument shows that (1, . . . , 1) Q
s
maps to zero in Hom
Q
(L
Q
, Q) and
hence comes from an element u N
Q
. In other words, 'm
i
, u` =1 for all i. Clearing
denominators gives the desired u N and k > 0 in N.
2.1. Lattice Points and Projective Toric Varieties 57
Finally, we prove (d) (a). Since Y
A


X
A
and

X
A
is irreducible, it sufces
to show that

X
A
(C

)
s
Y
A
.
Let p

X
A
(C

)
s
. Since X
A
T
P
s1 is the torus of X
A
, it follows that
p = (
m
1
(t), . . . ,
ms
(t))
for some C

and t T
N
. The element u N from part (d) gives a one-parameter
subgroup of T
N
, which we write as
u
() for C

. Then
u
()t T
N
maps
to the point q Y
A
given by
q =

m
1
(
u
()t), . . . ,
ms
(
u
()t)

/m
1
,u)

m
1
(t), . . . ,
/ms,u)

ms
(t)

,
since
m
(
u
()) =
/m,u)
by the description of ' , ` given in 1.1. The hypothesis
of part (d) allows us to rewrite q as
q =
k
(
m
1
(t), . . . ,
ms
(t)).
Using k > 0, we can choose so that p = q Y
A
. This completes the proof.
The condition 'm
i
, u` = k, i = 1, . . . , s, for some u N and k > 0 in N means
that A lies in an afne hyperplane of M
Q
not containing the origin. When M =Z
n
and A consists of the columns of an n s integer matrix A, this is equivalent to
(1, . . . , 1) lying in the row space of A (Exercise 2.1.2).
Example 2.1.5. We will examine the rational normal curve C
d
P
d
using two
different sets of lattice points.
First let A consist of the columns of the 2(d +1) matrix
A =

d d 1 1 0
0 1 d 1 d

.
The columns give the Laurent monomials dening the rational normal curve C
d
in Example 2.0.1. It follows that C
d
is a projective toric variety. The ideal of C
d
is the homogeneous ideal given in Example 2.0.1, and the corresponding afne
hyperplane of Z
2
containing A (= the columns of A) consists of all points (a, b)
satisfying a+b = d. It is equally easy to see that (1, . . . , 1) is in the row space of
A. In particular, we have
X
A
=C
d
and Y
A
=

C
d
.
Now let B=0, 1, . . . , d 1, d Z. This gives the map

B
: C

P
d
, t (1, t, . . . , t
d1
, t
d
).
The resulting projective variety is the rational normal curve, i.e., X
B
= C
d
, but
the afne variety of B is not the rational normal cone, i.e., Y
B
=

C
d
. This is
because I(Y
B
) C[x
0
, . . . , x
d
] is not homogeneous. For example, x
2
1
x
2
vanishes
at (1, t, . . . , t
d1
, t
d
) C
d+1
for all t C

. Thus x
2
1
x
2
I(Y
B
).
58 Chapter 2. Projective Toric Varieties
Given any A M, there is a standard way to modify A so that the conditions
of Proposition 2.1.4 are met: use A 1 MZ. This lattice corresponds to the
torus T
n
C

, and since
(2.1.4)
A|1
(t, ) = (
m
1
(t), . . . ,
ms
(t)) = (
m
1
(t), . . . ,
ms
(t)),
it follows immediately that X
A|1
= X
A
P
s1
. Since A 1 lies in an afne
hyperplane missing the origin, Proposition 2.1.4 implies that X
A
has afne cone
Y
A|1
=

X
A
. When M = Z
n
and A is represented by the columns of an n s
integer matrix A, we obtain A 1 by adding the row (1, . . . , 1) to A.
The Torus of a Projective Toric Variety. Our next task is to determine the torus
of X
A
. We will do so by identifying its character lattice. This will also tell us the
dimension of X
A
. Given A =m
1
, . . . , m
s
M, we set
Z

A =

s
i=1
a
i
m
i
[ a
i
Z,

s
i=1
a
i
= 0

.
The rank of Z

A is the dimension of the smallest afne subspace of M


R
containing
the set A (Exercise 2.1.3).
Proposition 2.1.6. Let X
A
be the projective toric variety of A M. Then:
(a) The lattice Z

A is the character lattice of the torus of X


A
.
(b) The dimension of X
A
is the dimension of the smallest afne subspace of M
R
containing A. In particular,
dim X
A
=

rankZA 1 if A satises the conditions of Proposition 2.1.4


rankZA otherwise.
Proof. To prove part (a), let M

be the character lattice of the torus T


X
A
of X
A
. By
(2.1.2), we have the commutative diagram
T
N

T
P
s1



P
s1
T
X
A
?

which induces the commutative diagram of character lattices


M M
s1

1Q

since M
s1
= (a
1
, . . . , a
s
) Z
s
[

s
i=0
a
i
= 0 is the character lattice of T
P
s1 by
(2.1.1). The map M
s1
M is induced by the map Z
s
M that sends e
i
to m
i
.
Thus Z

A is the image of M
s1
M and hence equals M

by the above diagram.


The rst assertion of part (b) follows from part (a) and the observation that
rankZ

A is the dimension of the smallest afne subspace of M


R
containing A.
2.1. Lattice Points and Projective Toric Varieties 59
Furthermore, if Y
A
equals the afne cone of X
A
, then there is u N with 'm
i
, u` =
k > 0 for all i by Proposition 2.1.4. This implies that '

s
i=1
a
i
m
i
, u` = k(

s
i=1
a
i
),
which gives the exact sequence
0 Z

A ZA
/ ,u)
kZ 0.
Then k > 0 implies rankZA 1 = rankZ

A = dim X
A
. However, if Y
A
=

X
A
,
then the ideal I
L
is not homogeneous. Thus some generator x

is not homo-
geneous, so that () (1, . . . , 1) = 0. But L, where L is dened by
0 L Z
s
ZA 0.
This implies that in the exact sequence
0 M
s1
Z
s
Z 0
(see (2.1.1)), the image of L Z
s
is Z Z for some > 0. This gives a diagram
0

0

LM
s1

0
0

M
s1

Z
s

0
0

Z

ZA

Z/Z

0
0 0 0
with exact rows and columns. Hence rankZA = rankZ

A = dim X
A
.
Example 2.1.7. Let A = e
1
, e
2
, e
1
+2e
2
, 2e
1
+e
2
Z
2
. One computes that
ZA = Z
2
but Z

A = (a, b) Z
2
[ a+b 0 mod 2. Thus Z

A has index 2 in
ZA. This means that Y
A
=

X
A
and the map of tori
T
Y
A
T
X
A
is two-to-one, i.e., its kernel has order 2 (Exercise 2.1.4).
Afne Pieces of a Projective Toric Variety. So far, our treatment of projective toric
varieties has used lattice points and toric ideals. Where are the semigroups? There
are actually lots of semigroups, one for each afne piece of X
A
P
s1
.
The afne open set U
i
=P
s1
`V(x
i
) contains the torus T
P
s1 . Thus
T
X
A
= X
A
T
P
s1 X
A
U
i
.
Since X
A
is the Zariski closure of T
X
A
in P
s1
, it follows that X
A
U
i
is the Zariski
closure of T
X
A
in U
i
C
s1
. Thus X
A
U
i
is an afne toric variety.
Given A =m
1
, . . . , m
s
M
R
, the afne semigroup associated to X
A
U
i
is
easy to determine. Recall that U
i
C
s1
is given by
(a
1
, . . . , a
s
) (a
1
/a
i
, . . . , a
i1
/a
i
, a
i+1
/a
i
, . . . , a
s
/a
i
).
60 Chapter 2. Projective Toric Varieties
Combining this and
m
j
/
m
i
=
m
j
m
i
with the map (2.1.2), we see that X
A
U
i
is the Zariski closure of the image of the map
T
N
C
s1
given by
(2.1.5) t

m
1
m
i
(t), . . . ,
m
i1
m
i
(t),
m
i+1
m
i
(t), . . . ,
msm
i
(t)

.
If we set A
i
=A m
i
=m
j
m
i
[ j = i, it follows that
X
A
U
i
=Y
A
i
= Spec(C[S
i
]),
where S
i
=NA
i
is the afne semigroup generated by A
i
. We have thus proved the
following result.
Proposition 2.1.8. Let X
A
P
s1
for A = m
1
, . . . , m
s
M
R
. Then the afne
piece X
A
U
i
is the afne toric variety
X
A
U
i
=Y
A
i
= Spec(C[S
i
])
where A
i
=A m
i
and S
i
=NA
i
.
We also note that the results of Chapter 1 imply that the torus of Y
A
i
has char-
acter lattice ZA
i
. Yet the torus is T
X
A
, which has character lattice Z

A by Propo-
sition 2.1.6. These are consistent since ZA
i
=Z

A for all i.
Besides describing the afne pieces X
A
U
i
of X
A
P
s1
, we can also de-
scribe how they patch together. In other words, we can give a completely toric
description of the inclusions
X
A
U
i
X
A
U
i
U
j
X
A
U
j
when i = j. For instance, U
i
U
j
consists of all points of X
A
U
i
where x
j
/x
i
= 0.
In terms of X
A
U
i
= Spec(C[S
i
]), this means those points where
m
j
m
i
= 0.
Thus
(2.1.6) X
A
U
i
U
j
= Spec

C[S
i
]

m
j
m
i
= Spec

C[S
i
]

m
j
m
i

,
so that if we set m = m
j
m
i
, then the inclusion X
A
U
i
U
j
X
A
U
i
can be
written
(2.1.7) Spec(C[S
i
])

m Spec(C[S
i
]).
This looks very similar to the inclusion constructed in (1.3.4) using faces of cones.
We will see in 2.3 that this is no accident.
We next say a few words about how the polytope P = Conv(A) M
R
relates
to X
A
. As we will learn in 2.2, the dimension of P is the dimension of the small-
est afne subspace of M
R
containing P, which is the same as the smallest afne
subspace of M
R
containing A. It follows from Proposition 2.1.6 that
dim X
A
= dim P.
Furthermore, the vertices of P give an especially efcient afne covering of X
A
.
2.1. Lattice Points and Projective Toric Varieties 61
Proposition 2.1.9. Given A =m
1
, . . . , m
s
M, let P = Conv(A) M
R
and set
J =

j 1, . . . , s [ m
j
is a vertex of P

. Then
X
A
=

jJ
X
A
U
j
.
Proof. We will prove that if i 1, . . . , s, then X
A
U
i
X
A
U
j
for some j J.
The discussion of polytopes from 2.2 below implies that
PM
Q
=

jJ
r
j
m
j
[ r
j
Q
0
,

jJ
r
j
= 1

.
Given i 1, . . . , s, we have m
i
PM, so that m
i
is a convex Q-linear combi-
nation of the vertices m
j
. Clearing denominators, we get integers k > 0 and k
j
0
such that
km
i
=

jJ
k
j
m
j
,

jJ
k
j
= k.
Thus

jJ
k
j
(m
j
m
i
) = 0, which implies that m
i
m
j
S
i
when k
j
> 0. Fix
such a j. Then
m
j
m
i
C[S
i
] is invertible, so C[S
i
]

m
j
m
i
= C[S
i
]. By (2.1.6),
X
A
U
i
U
j
= Spec(C[S
i
]) = X
A
U
i
, giving X
A
U
i
X
A
U
j
.
Projective Normality. An irreducible variety V P
n
is called projectively normal
if its afne cone

V C
n+1
is normal. A projectively normal variety is always
normal (Exercise 2.1.5). Here is an example to show that the converse can fail.
Example 2.1.10. Let A Z
2
consist of the columns of the matrix

4 3 1 0
0 1 3 4

,
giving the Laurent monomials s
4
, s
3
t, st
3
, t
4
. The polytope P=Conv(A) is the line
segment connecting (4, 0) and (0, 4), with vertices corresponding to s
4
and t
4
. The
afne piece of X
A
corresponding to s
4
has coordinate ring
C[s
3
t/s
4
, st
3
/s
4
, t
4
/s
4
] =C[t/s, (t/s)
3
, (t/s)
4
] =C[t/s],
which is normal since it is a polynomial ring. Similarly, one sees that the coordinate
ring corresponding to t
4
is C[s/t], also normal. These afne pieces cover X
A
by
Proposition 2.1.9, so that X
A
is normal.
Since (1, 1, 1, 1) is in the row space of the matrix, Y
A
is the afne cone of X
A
by Proposition 2.1.4. The afne variety Y
A
is not normal by Example 1.3.9, so that
X
A
is normal but not projectively normal. See Example B.1.2 for a sophisticated
proof of this fact that uses sheaf cohomology from Chapter 9.
The notion of normality used in this example is a bit ad-hoc since we have not
formally dened normality for projective varieties. Once we dene normality for
abstract varieties in Chapter 3, we will see that Example 2.1.10 is fully rigorous.
We will say more about projective normality when we explore the connection
with polytopes suggested by the above results.
62 Chapter 2. Projective Toric Varieties
Exercises for 2.1.
2.1.1. Consider the set P
3
Z
33
of 33 permutation matrices dened in Example 2.1.3.
(a) Prove the claim made in Example 2.1.3 that three of the permutation matrices sum to
the other three and use this to explain why x
123
x
231
x
312
x
132
x
321
x
213
I(X
P3
).
(b) Show that dim X
P3
= 4 by computing Z

P
3
.
(c) Conclude that I(X
P3
) ='x
123
x
231
x
312
x
132
x
321
x
213
`.
2.1.2. Let A Z
n
consist of the columns of an n s matrix A with integer entries. Prove
that the conditions of Proposition 2.1.4 are equivalent to the assertion that (1, . . . , 1) Z
s
lies in the row space of A over R or Q.
2.1.3. Given a nite set A M, prove that the rank of Z

A equals the dimension of the


smallest afne subspace (over Q or R) containing A.
2.1.4. Verify the claims made in Example 2.1.7. Also compute I(Y
A
) and check that it is
not homogeneous.
2.1.5. Let V P
n
be projectively normal. Use (2.0.6) to prove that the afne pieces V U
i
of V are normal.
2.1.6. Fix a nite subset A M. Given m M, let A +m =m

+m [ m

A. This is
the translate of A by m.
(a) Prove that A and its translate A +m give the same projective toric variety, i.e., X
A
=
X
A+m
.
(b) Give an example to show that the afne toric varieties Y
A
and Y
A+m
can differ. Hint:
Pick A so that it lies in an afne hyperplane not containing the origin. Then translate
A to the origin.
2.1.7. In Proposition 2.1.4, give a direct proof that (d) (c).
2.1.8. In Example 2.1.5, the rational normal curve C
d
P
d
was parametrized using the
homogeneous monomials s
i
t
j
, i + j = d. Here we will consider the curve parametrized by
a subset of these monomials corresponding to the exponent vectors
A =(a
0
, b
0
), . . . , (a
n
, b
n
)
where a
0
> a
1
> > a
n
and a
i
+b
i
= d for every i. This gives the projective curve
X
A
P
n
. We assume n 2.
(a) If a
0
< d or a
n
> 0, explain why we can obtain the same projective curve using mono-
mials of strictly smaller degree.
(b) Assume a
0
=d and a
n
=0. Use Proposition 2.1.8 to show that C is smooth if and only
if a
1
= d 1 and a
n1
= 1. Hint: For one direction, it helps to remember that smooth
varieties are normal.
2.2. Lattice Points and Polytopes
Before we can begin our exploration of the rich connections between toric varieties
and polytopes, we rst need to study polytopes and their lattice points.
2.2. Lattice Points and Polytopes 63
Polytopes. Recall from Chapter 1 that a polytope P M
R
is the convex hull of
a nite set S M
R
, i.e., P = Conv(S). Similar to what we did for cones, our
discussion of polytopes will omit proofs. Detailed treatments of polytopes can be
found in [51], [128] and [281].
The dimension of a polytope P M
R
is the dimension of the smallest afne
subspace of M
R
containing P. Given a nonzero vector u in the dual space N
R
and
b R, we get the afne hyperplane H
u,b
and closed half-space H
+
u,b
dened by
H
u,b
=m M
R
[ 'm, u` = b and H
+
u,b
=m M
R
[ 'm, u` b.
A subset QP is a face of P, written Q _P, if there are u N
R
`0, b R with
Q = H
u,b
P and P H
+
u,b
.
We say that H
u,b
is a supporting afne hyperplane in this situation. Figure 1 shows
a polygon with the supporting lines of its 1-dimensional faces. The arrows in the
gure indicate the vectors u.

Figure 1. A polygon P and four of its supporting lines


We also regard P as a face of itself. Every face of P is again a polytope, and
if P = Conv(S) and Q = H
u,b
P as above, then Q = Conv(S H
u,b
). Faces of
P of special interest are facets, edges and vertices, which are faces of dimension
dim P1, 1 and 0 respectively. Facets will usually be denoted by the letter F.
Here are some properties of faces.
Proposition 2.2.1. Let P M
R
be a polytope. Then:
(a) P is the convex hull of its vertices.
(b) If P = Conv(S), then every vertex of P lies in S.
(c) If Q is a face of P, then the faces of Q are precisely the faces of P lying in Q.
(d) Every proper face Q P is the intersection of the facets F containing Q.
A polytope P M
R
can also be written as a nite intersection of closed half-
spaces. The converse is true provided the intersection is bounded. In other words,
if an intersection
P =
s

i=1
H
+
u
i
,b
i
64 Chapter 2. Projective Toric Varieties
is bounded, then P is a polytope. Here is a famous example.
Example 2.2.2. Add matrix MR
dd
is doubly-stochastic if it has nonnegative
entries and its row and column sums are all 1. If we regard R
dd
as the afne
space R
d
2
with coordinates a
i j
, then the set M
d
of all doubly-stochastic matrices
is dened by the inequalites
a
i j
0 (all i, j)

d
i=1
a
i j
1,

d
i=1
a
i j
1 (all j)

d
j=1
a
i j
1,

d
j=1
a
i j
1 (all i).
(We use two inequalities to get one equality.) These inequalities easily imply that
0 a
i j
1 for all i, j, so that M
d
is bounded and hence is a polytope.
Birkhoff and Von Neumann proved independently that the vertices of M
d
are
the d! permutation matrices and that dimM
d
= (d 1)
2
. In the literature, M
d
has
various names, including the Birkhoff polytope and the transportation polytope.
See [281, p. 20] for more on this interesting polytope.
When P is full dimensional, i.e., dim P = dim M
R
, its presentation as an inter-
section of closed half-spaces has an especially nice form because each facet F has
a unique supporting afne hyperplane. We write the supporting afne hyperplane
and corresponding closed half-space as
H
F
=m M
R
[ 'm, u
F
` =a
F
and H
+
F
=m M
R
[ 'm, u
F
` a
F
,
where (u
F
, a
F
) N
R
R is unique up to multiplication by a positive real number.
We call u
F
an inward-pointing facet normal of the facet F. It follows that
(2.2.1) P =

F facet
H
+
F
=m M
R
[ 'm, u
F
` a
F
for all facets F P.
In Figure 1, the supporting lines plus arrows determine the supporting half-planes
whose intersection is the polygon P. We write (2.2.1) with minus signs in order to
simplify formulas in later chapters.
Here are some important classes of polytopes.
Denition 2.2.3. Let P M
R
be a polytope of dimension d.
(a) P is a simplex or d-simplex if it has d +1 vertices.
(b) P is simplicial if every facet of P is a simplex.
(c) P is simple if every vertex is the intersection of precisely d facets.
Examples include the Platonic solids in R
3
:
A tetrahedron is a 3-simplex.
The octahedron and icosahedron are simplicial since their facets are triangles.
The cube and dodecahedron are simple since three facets meet at every vertex.
2.2. Lattice Points and Polytopes 65
Polytopes P
1
and P
2
are combinatorially equivalent if there is a bijection
faces of P
1
faces of P
2

that preserves dimensions, intersections, and the face relation _. For example,
simplices of the same dimension are combinatorially equivalent, and in the plane,
the same holds for polygons with the same number of vertices.
Sums, Multiples, and Duals. Given a polytope P = Conv(S), its multiple rP =
Conv(rS) is again a polytope for any r 0. If P is dened by the inequalities
'm, u
i
` a
i
, 1 i s,
then rP is given by
'm, u
i
` ra
i
, 1 i s.
In particular, when P is full dimensional, then P and rP have the same inward-
pointing facet normals.
The Minkowski sum of subsets A
1
, A
2
M
R
is
A
1
+A
2
=m
1
+m
2
[ m
1
A
1
, m
2
A
2
.
Given polytopes P
1
=Conv(S
1
) and P
2
=Conv(S
2
), their Minkowski sumP
1
+P
2
=
Conv(S
1
+S
2
) is again a polytope. We also have the distributive law
rP+sP = (r +s)P.
When P M
R
is full dimensional and 0 is an interior point of P, we dene the
dual or polar polytope
P

=u N
R
[ 'm, u` 1 for all m P N
R
.
Figure 2 shows an example of this in the plane.
P P
Figure 2. A polygon P and its dual P

in the plane
When we write P =m M
R
[ 'm, u
F
` a
F
, F facet as in (2.2.1), we have
a
F
> 0 for all F since 0 is in the interior. Then P

is the convex hull of the vectors


(1/a
F
)u
F
N
R
(Exercise 2.2.1). We also have (P

= P in this situation.
66 Chapter 2. Projective Toric Varieties
Lattice Polytopes. Now let M and N be dual lattices with associated vector spaces
M
R
and N
R
. A lattice polytope P M
R
is the convex hull of a nite set S M. It
follows easily that a polytope in M
R
is a lattice polytope if and only if its vertices
lie in M (Exercise 2.2.2).
Example 2.2.4. The standard n-simplex in R
n
is

n
= Conv(0, e
1
, . . . , e
n
).
Another simplex in R
3
is P = Conv(0, e
1
, e
2
, e
1
+e
2
+3e
3
), shown in Figure 3.
e
1
+e
2
+3e
3
e
2
e
1
0
Figure 3. The simplex P =Conv(0, e1, e2, e1 +e2 +3e3) R
3
The lattice polytopes
3
and P are combinatorially equivalent but will give very
different projective toric varieties.
Example 2.2.5. The Birkhoff polytope dened in Example 2.2.2 is a lattice poly-
tope relative to the lattice of integer matrices Z
dd
since its vertices are the permu-
tation matrices, whose entries are all 0 or 1.
One can show that faces of lattice polytopes are again lattice polytopes and that
Minkowski sums and integer multiples of lattice polytopes are lattice polytopes
(Exercise 2.2.2). Furthermore, every lattice polytope is an intersection of closed
half-spaces dened over M, i.e., P =

s
i=1
H
+
u
i
,b
i
where u
i
N and b
i
Z.
When a lattice polytope P is full dimensional, the facet presentation given in
(2.2.1) has an especially nice form. If F is a facet of P, then the inward-pointing
facet normals of F lie on a rational ray in N
R
. Let u
F
denote the unique ray gener-
ator. The corresponding a
F
is integral since 'm, u
F
` = a
F
when m is a vertex of
F. It follows that
(2.2.2) P =m M
R
[ 'm, u
F
` a
F
for all facets F P
is the unique facet presentation of the lattice polytope P.
2.2. Lattice Points and Polytopes 67
Example 2.2.6. Consider the square P=Conv(e
1
e
2
) R
2
. The facet normals
of P are e
1
and e
2
and the facet presentation of P is given by
'm, e
1
` 1
'm, e
2
` 1.
Since the a
F
are all equal to 1, it follows that P

= Conv(e
1
, e
2
) is also a lattice
polytope. The polytopes P and P

are pictured in Figure 2.


It is rare that the dual of a lattice polytope is a lattice polytopethis is related
to the reexive polytopes that will be studied later in the book.
Example 2.2.7. The 3-simplex P = Conv(0, e
1
, e
2
, e
1
+e
2
+3e
3
) R
3
pictured in
Example 2.2.4 has facet presentation
'm, e
3
` 0
'm, 3e
1
e
3
` 0
'm, 3e
2
e
3
` 0
'm, 3e
1
3e
2
+e
3
` 3
(Exercise 2.2.3). However, if we replace 3 with 1 in the last inequality, we get
integer inequalities that dene (1/3)P, which is not a lattice polytope.
The combinatorial type of a polytope is an interesting object of study. This
leads to the question Is every polytope combinatorially equivalent to a lattice
polytope? If the given polytope is simplicial, the answer is yesjust wiggle
the vertices to make them rational and clear denominators to get a lattice polytope.
The same argument works for simple polytopes by wiggling the facet normals.
This will enable us to prove results about arbitrary simplicial or simple polytopes
using toric varieties. But in general, the answer is nothere exist polytopes in
every dimension 8 not combinatorially equivalent to any lattice polytope. An
example is described in [281, Ex. 6.21].
A First Guess for the Toric Variety of Polytope. A lattice polytope P gives a nite
set of lattice points PM, which in turn gives a projective toric variety X
PM
. This
is a natural candidate for the toric variety of P. However, X
PM
may fail to reect
the properties of P when there are too few lattice points.
Example 2.2.8. In Example 2.2.4, we dened the standard 3-simplex
3
and the
3-simplex P = Conv(0, e
1
, e
2
, e
1
+e
2
+3e
3
). Both have only four lattice points
(their vertices). Thus the toric varieties X

3
Z
3 and X
PZ
3 live in P
3
, and in fact
we have X

3
Z
3 = X
PZ
3 = P
3
(Exercise 2.2.3). Yet
3
and P are very different
lattice polytopes. For example, the nonzero vertices of
3
form a basis of Z
3
, but
this is not true for P.
68 Chapter 2. Projective Toric Varieties
We will see below that the construction P PM X
PM
works ne when
the lattice polytope P has enough lattice points. Hence our rst task is to explore
what this means.
Normal Polytopes. Here is one way to say that P has enough lattice points.
Denition 2.2.9. A lattice polytope P M
R
is normal if
(kP) M+(P) M = ((k +)P) M
for all k, N.
The inclusion (kP) M+(P) M ((k +)P) M is automatic. Thus nor-
mality means that all lattice points of (k +)P come from lattice points of kP and
P. In particular, a lattice polytope is normal if and only if
PM+ +PM
. .. .
k times
= (kP) M.
for all integers k 1. So normality means that P has enough lattice points to
generate the lattice points in all integer multiples of P.
Lattice polytopes of dimension 1 are normal (Exercise 2.2.4). Here is another
class of normal polytopes.
Denition 2.2.10. A simplex P M
R
is basic if P has a vertex m
0
such that the
differences mm
0
, for vertices m = m
0
of P, form a subset of a Z-basis of M.
This denition is independent of which vertex m
0
P is chosen. The standard
simplex
n
R
n
is basic, and any basic simplex is normal (Exercise 2.2.5). More
general simplices, however, need not be normal.
Example 2.2.11. We noted in Example 2.2.8 that the only lattice points of P =
Conv(0, e
1
, e
2
, e
1
+e
2
+3e
3
) R
3
are its vertices. It follows easily that
e
1
+e
2
+e
3
=
1
6
(0) +
1
3
(2e
1
) +
1
3
(2e
2
) +
1
6
(2e
1
+2e
2
+6e
3
) 2P
is not the sum of lattice points of P. This shows that P is not normal. In particular,
P is not basic.
Here is an important result on normality.
Theorem 2.2.12. Let P M
R
be a full dimensional lattice polytope of dimension
n 2. Then kP is normal for all k n1.
Proof. This result was rst explicitly stated in [55], though as noted in [55], its
essential content follows from [94] and [187]. We will use ideas from [187] and
[223] to show that
(2.2.3) (kP) M+PM = ((k +1)P) M for k Z, k n1.
2.2. Lattice Points and Polytopes 69
In Exercise 2.2.6 you will show that (2.2.3) implies that kP is normal for all integers
k n1. Note also that for (2.2.3), it sufces to prove that
((k +1)P) M (kP) M+PM
since the other inclusion is obvious.
First consider the case where P is a simplex with no interior lattice points.
Let the vertices of P be m
0
, . . . , m
n
and take k n1. Then (k +1)P has vertices
(k+1)m
0
, . . . , (k+1)m
n
, so that a point m((k+1)P)M is a convex combination
m =

n
i=0

i
(k +1)m
i
, where
i
0,

n
i=0

i
= 1.
If we set
i
= (k +1)
i
, then
m =

n
i=0

i
m
i
, where
i
0,

n
i=0

i
= k +1.
If
i
1 for some i, then one easily sees that mm
i
(kP) M. Hence m =
(mm
i
) +m
i
is the desired decomposition. On the other hand, if
i
< 1 for all i,
then
n = (n1) +1 k +1 =

n
i=0

i
< n+1,
so that k = n1 and

n
i=0

i
= n. Now consider the lattice point
m =

m
0
+ +m
n
) m =

n
i=0
(1
i
)m
i
.
The coefcients are positive since
i
< 1 for all i, and their sum is

n
i=0
(1
i
) =
n+1n = 1. Hence m is a lattice point in the interior of P since 1
i
> 0 for all
i. This contradicts our assumption on P and completes the proof when P is a lattice
simplex containing no interior lattice points.
To prove (2.2.3) for the general case, it sufces to prove that P is a nite union
of n-dimensional lattice simplices with no interior lattice points (Exercise 2.2.7).
For this, we use Carath eodorys theorem (see [281, Prop. 1.15]), which asserts that
for a nite set A M
R
, we have
Conv(A) =

Conv(B),
where the union is over all subsets BA consisting of dim Conv(A)+1 afnely
independent elements. Thus each Conv(B) is a simplex. This enables us to write
our lattice polytope P as a nite union of n-dimensional lattice simplices.
If an n-dimensional lattice simplex Q=Conv(w
0
, . . . , w
n
) has an interior lattice
point v, then
Q =
n

i=0
Q
i
, Q
i
= Conv(w
0
, . . . , w
i
, . . . , w
n
, v)
is a nite union of n-dimensional lattice simplices, each of which has fewer interior
lattice points than Q since v becomes a vertex of each Q
i
. By repeating this process
on those Q
i
that still have interior lattice points, we can eventually write Q and
hence our original polytope P as a nite union of n-dimensional lattice simplices
with no interior lattice points. You will verify the details in Exercise 2.2.7.
70 Chapter 2. Projective Toric Varieties
This theorem shows that for the nonnormal 3-simplex P of Example 2.2.11, its
multiple 2P is normal. Here is another consequence of Theorem 2.2.12.
Corollary 2.2.13. Every lattice polygon P R
2
is normal.
We can also interpret normality in terms of the cone of P, dened by
C(P) = Cone(P1) M
R
R.
This was introduced in Figure 3 of Chapter 1. The key feature of this cone is that
kP is the slice of C(P) at height k, as illustrated in Figure 4. It follows that lattice
points m kP correspond to points (m, k) C(P) (MZ).
P
2P
height = 1
height = 2
C(P)
Figure 4. The cone C(P) sliced at heights 1 and 2
In Exercise 2.2.8 you will show that the semigroup C(P) (MZ) of lattice
points in C(P) relates to normality as follows.
Lemma 2.2.14. Let P M
R
be a lattice polytope. Then P is normal if and only if
(PM) 1 generates the semigroup C(P) (MZ).
This lemma tells us that P M
R
is normal if and only if (PM) 1 is the
Hilbert basis of C(P) (MZ).
Example 2.2.15. In Example 2.2.11, the simplex P = Conv(0, e
1
, e
2
, e
1
+e
2
+3e
3
)
gives the cone C(P) R
4
. The Hilbert basis of C(P) (MZ) is
(0, 1), (e
1
, 1), (e
2
, 1), (e
1
+e
2
+3e
3
, 1), (e
1
+e
2
+e
3
, 2), (e
1
+e
2
+2e
3
, 2)
(Exercise 2.2.3). Since the Hilbert basis has generators of height greater than 1,
Lemma 2.2.14 gives another proof that P is not normal.
2.2. Lattice Points and Polytopes 71
In Exercise 2.2.9, you will generalize Lemma 2.2.14 as follows.
Lemma 2.2.16. Let P M
R
R
n
be a lattice polytope of dimension n 2 and let
k
0
be the maximum height of an element of the Hilbert basis of C(P). Then:
(a) k
0
n1.
(b) kP is normal for any k k
0
.
The Hilbert basis of the simplex P of Example 2.2.15 has maximum height 2.
Then Lemma 2.2.16 gives another proof that 2P is normal. The paper [187] gives
a version of Lemma 2.2.16 that applies to Hilbert bases of more general cones.
Very Ample Polytopes. Here is a slightly different notion of what it means for a
polytope to have enough lattice points.
Denition 2.2.17. A lattice polytope P M
R
is very ample if for every vertex
m P, the semigroup S
P,m
= N(P Mm) generated by the set PMm =
m

m[ m

PM is saturated in M.
This denition relates to normal polytopes as follows.
Proposition 2.2.18. A normal lattice polytope P is very ample.
Proof. Fix a vertex m
0
P and take m M such that km S
P,m
0
for some integer
k 1. To prove that m S
P,m
0
, write km S
P,m
0
as
km =

PM
a
m
(m

m
0
), a
m
N.
Pick d N so that kd

PM
a
m

. Then
km+kdm
0
=

PM
a
m
m

kd

PM
a
m

m
0
kdP.
Dividing by k gives m+dm
0
dP, which by normality implies that
m+dm
0
=
d

i=1
m
i
, m
i
PM for all i.
We conclude that m =

d
i=1
(m
i
m
0
) S
P,m
0
, as desired.
Combining this with Theorem 2.2.12 and Corollary 2.2.13 gives the following.
Corollary 2.2.19. Let P M
R
R
n
be a full dimensional lattice polytope. Then:
(a) If dim P 2, then kP is very ample for all k n1.
(b) If dim P = 2, then P is very ample.
Part (a) was rst proved in [94]. We will soon see that very ampleness is
precisely the property needed to dene the toric variety of a lattice polytope.
The following example taken from [52, Ex. 5.1] shows that very ample poly-
topes need not be normal, i.e., the converse of Proposition 2.2.18 is false.
72 Chapter 2. Projective Toric Varieties
Example 2.2.20. Given 1 i < j < k 6, let [i jk] denote the vector in Z
6
with 1
in positions i, j, k and 0 elsewhere. Thus [123] = (1, 1, 1, 0, 0, 0). Then let
A =

[123], [124], [135], [146], [156], [236], [245], [256], [345], [346]

Z
6
.
The lattice polytope P = Conv(A) lies in the afne hyperplane of R
6
where the
coordinates sum to 3. As explained in [52], this conguration can be interpreted in
terms of a triangulation of the real projective plane.
The points of A are the only lattice points of P (Exercise 2.2.10), so that A is
the set of vertices of P. Number the points of A as m
1
, . . . , m
10
. Then
(1, 1, 1, 1, 1, 1) =
1
5
10

i=1
m
i
=
10

i=1
1
10
(2m
i
)
shows that v = (1, 1, 1, 1, 1, 1) 2P. Since v is not a sum of lattice points of P
(when [i jk] A, the vector v [i jk] is not in A), we conclude that P is not a
normal polytope.
To show that P is very ample, we rst prove that A 1(v, 2) R
6
Ris
a Hilbert basis of the semigroup C(P) Z
7
, where C(P) R
6
R is the cone over
P1. We show how do this using Normaliz [57] in Example B.3.1. Another
approach would be to use 4ti2 [140].
Now x i and let S
P,m
i
be the semigroup generated by the m
j
m
i
. Take m
Z
6
such that km S
P,m
i
. As in the proof of Proposition 2.2.18, this implies that
m+dm
i
dP for some d N. Thus (m+dm
i
, d) C(P) Z
7
. Expressing this in
terms of the above Hilbert basis easily implies that
m = a(v 2m
i
) +
10

j=1
a
j
(m
j
m
i
), a, a
j
N.
If we can show that v2m
i
S
P,m
i
, then mS
P,m
i
follows immediately and proves
that S
P,m
i
is saturated. When i = 1, one can check that
v +[123] = [124] +[135] +[236],
which implies that
v 2m
1
= (m
2
m
1
) +(m
3
m
1
) +(m
6
m
1
) S
P,m
1
.
One obtains similar formulas for i = 2, . . . , 10 (Exercise 2.2.10), which completes
the proof that P is very ample.
The polytope P has further interesting properties. For example, up to afne
equivalence, P can be described as the convex hull of the 10 points in Z
5
given by
(0, 0, 0, 0, 0), (0, 0, 0, 0, 1), (0, 0, 1, 1, 0), (0, 1, 0, 1, 1), (0, 1, 1, 1, 0)
(1, 0, 1, 0, 1), (1, 0, 1, 1, 1), (1, 1, 0, 0, 0), (1, 1, 0, 1, 1), (1, 1, 1, 0, 0).
2.2. Lattice Points and Polytopes 73
Of all 5-dimensional polytopes whose vertices lie in 0, 1
5
, this polytope has the
most edges, namely 45 (see [4]). Since it has 10 vertices and 45 =

10
2

, every pair
of distinct vertices is joined by an edge. Such polytopes are 2-neighborly.
Exercises for 2.2.
2.2.1. Let P M
R
be a full dimensional polytope with the origin as an interior point.
(a) Write P =m M
R
[ 'm, u
F
` a
F
for all facets F. Prove that a
F
> 0 for all F and
that P

= Conv((1/a
F
)u
F
[ F a facet).
(b) Prove that the dual of a simplicial polytope is simple and vice versa.
(c) Prove that (rP)

= (1/r)P

for all r > 0.


(d) Use part (c) to construct an example of a lattice polytope whose dual is not a lattice
polytope.
2.2.2. Let P M
R
be a polytope.
(a) Prove that P is a lattice polytope if and only if the vertices of P lie in M.
(b) Prove that P is a lattice polytope if and only if P is the convex hull of its lattice points,
i.e., P = Conv(PM).
(c) Prove that every face of a lattice polytope is a lattice polytope.
(d) Prove that Minkowski sums and integer multiples of lattice polytopes are again lattice
polytopes.
2.2.3. Let P =Conv(0, e
1
, e
2
, e
1
+e
2
+3e
3
) R
3
be the simplex studied in Examples 2.2.4,
2.2.7, 2.2.8, 2.2.11 and 2.2.15.
(a) Verify the facet presentation of P given in Example 2.2.7.
(b) Show that the only lattice points of P are its vertices.
(c) Show that the toric variety X
PZ
3 is P
3
, as claimed in Example 2.2.8.
(d) Show that the vectors given in Example 2.2.15 form the Hilbert basis of the semigroup
C(P) (MZ).
2.2.4. Prove that every 1-dimensional lattice polytope is normal.
2.2.5. Recall the denition of basic simplex given in Denition 2.2.10.
(a) Show that if a simplex satises Denition 2.2.10 for one vertex, then it satises the
denition for all vertices.
(b) Show that the standard simplex
n
is basic.
(c) Prove that a basic simplex is normal.
2.2.6. Let P M
R
R
n
be an n-dimensional lattice polytope.
(a) Prove that (2.2.3) implies that
(kP) M+(P) M = ((k +)P) M
for all integers k n 1 and 0. Hint: When = 2, we have
(kP) M+PM+PM (kP) M+(2P) M((k +2)P) M.
Apply (2.2.3) twice on the right.
(b) Use part (a) to prove that kP is normal when k n 1 and P satisifes (2.2.3).
74 Chapter 2. Projective Toric Varieties
2.2.7. Let P M
R
R
n
be an n-dimensional lattice polytope.
(a) Follow the hints given in the text to give a careful proof that P is a nite union of
n-dimensional lattice simplices with no interior lattice points.
(b) In the text, we showed that (2.2.3) holds for an n-dimensional lattice simplex with no
interior lattice points. Use this and part (a) to show that (2.2.3) holds for P.
2.2.8. Prove Lemma 2.2.14.
2.2.9. In this exercise you will prove Lemma 2.2.16. As in the lemma, let k
0
be the maxi-
mum height of a generator of the Hilbert basis of C(P) (MZ).
(a) Adapt the proof of Gordans Lemma (Proposition 1.2.17) to show that if H is the
Hilbert basis of the semigroup of lattice points in a strongly convex cone Cone(A),
then the lattice points in the cone can be written as the union
NA

mH

m+NA

.
(b) If the Hilbert basis of C(P) (MZ) is (m
1
, h
1
), . . . , (m
s
, h
s
), then conclude that
C(P) (MZ) = S

s
i=1

(m
i
, h
i
) +S

,
where S =N((PM) 1).
(c) Use part (b) to show that (2.2.3) holds for k k
0
.
2.2.10. Consider the polytope P = Conv(A) from Example 2.2.20.
(a) Prove that A is the set of lattice points of P.
(b) Complete the proof begun in the text that P is very ample.
2.2.11. Prove that every proper face of a simplicial polytope is a simplex.
2.2.12. In Corollary 2.2.19 we proved that kP is very ample for k n 1 using Theo-
rem 2.2.12 and Proposition 2.2.18. Give a direct proof of the weaker result that kP is very
ample for k sufciently large. Hint: A vertex m P gives the cone C
P,m
generated by the
semigroup S
P,m
dened in Denition 2.2.17. The cone C
P,m
is strongly convex since m is
a vertex and hence C
P,m
M has a Hilbert basis. Furthermore, C
P,m
=C
kP,km
for all k N.
Now argue that when k is large, (kP) Mkm contains the Hilbert basis of C
P,m
M. A
picture will help.
2.2.13. Fix an integer a 1 and consider the 3-simplex P = Conv(0, ae
1
, ae
2
, e
3
) R
3
.
(a) Work out the facet presentation of P and verify that the facet normals can be labeled
so that u
0
+u
1
+u
2
+au
3
= 0.
(b) Show that P is normal. Hint: Show that PZ
3
+(kP) Z
3
= ((k +1)P) Z
3
.
We will see later that the toric variety of P is the weighted projective space P(1, 1, 1, a).
2.3. Polytopes and Projective Toric Varieties
Our next task is to dene the toric variety of a lattice polytope. As noted in 2.2,
we need to make sure that the polytope has enough lattice points. Hence we begin
with very ample polytopes. Strongly convex rational polyhedral cones will play an
important role in our development.
2.3. Polytopes and Projective Toric Varieties 75
The Very Ample Case. Let P M
R
be a full dimensional very ample polytope
relative to the lattice M, and let dim P = n. If PM = m
1
, . . . , m
s
, then X
PM
is
the Zariski closure of the image of the map T
N
P
s1
given by
t

m
1
(t), . . . ,
ms
(t)

P
s1
.
Fix homogeneous coordinates x
1
, . . . , x
s
for P
s1
.
We examine the structure of X
PM
P
s1
using Propositions 2.1.8 and 2.1.9.
For each m
i
PM consider the semigroup
S
i
=N(PMm
i
)
generated by m
j
m
i
for m
j
P M. In P
s1
we have the afne open subset
U
i
C
s1
consisting of those points where x
i
= 0. Proposition 2.1.8 showed that
the afne open piece X
PM
U
i
of X
PM
is the afne toric variety
X
PM
U
i
Spec(C[S
i
]),
and Proposition 2.1.9 showed that
X
PM
=

m
i
vertex of P
X
PM
U
i
.
Here is our rst major result about X
PM
.
Theorem 2.3.1. Let X
PM
be the projective toric variety of the very ample polytope
P M
R
, and assume that P is full dimensional with dim P = n. Then:
(a) For each vertex m
i
PM, the afne piece X
PM
U
i
is the afne toric variety
X
PM
U
i
=U

i
= Spec(C[

i
M])
where
i
N
R
is the strongly convex rational polyhedral cone dual to the cone
Cone(PMm
i
) M
R
. Furthermore, dim
i
= n.
(b) The torus of X
PM
has character lattice M and hence is the torus T
N
.
Proof. Let C
i
= Cone(PMm
i
). Since m
i
is a vertex, it has a supporting hy-
perplane H
u,a
such that P H
+
u,a
and PH
u,a
= m
i
. It follows that H
u,0
is a
supporting hyperplane of 0 C
i
(Exercise 2.3.1), so that C
i
is strongly convex. It is
also easy to see that dimC
i
=dim P (Exercise 2.3.1). It follows that C
i
and
i
=C

i
are strongly convex rational polyhedral cones of dimension n.
We have S
i
C
i
M =

i
M. By hypothesis, P is very ample, which means
that S
i
M is saturated. Since S
i
and C
i
=

i
are both generated by PMm
i
,
part (a) of Exercise 1.3.4 implies S
i
=

i
M. (This exercise was part of the proof
of the characterization of normal afne toric varieties given in Theorem 1.3.5.)
Part (a) of the theorem follows immediately.
For part (b), Theorem 1.2.18 implies that T
N
is the torus of U

i
since
i
is
strongly convex. Then T
N
U

i
= X
PM
U
i
X
PM
shows that T
N
is also the
torus of X
PM
.
76 Chapter 2. Projective Toric Varieties
The afne pieces X
PM
U
i
and X
PM
U
j
intersect in X
PM
U
i
U
j
. In order
to describe this intersection carefully, we need to study how the cones
i
and
j
t
together in N
R
. This leads to our next topic.
The Normal Fan. The cones
i
N
R
appearing in Theorem 2.3.1 t together in a
remarkably nice way, giving a structure called the normal fan of P.
Let P M
R
be a full dimensional lattice polytope, not necessarily very ample.
Faces, facets and vertices of P will be denoted by Q, F and v respectively. Hence
we write the facet presentation of P as
(2.3.1) P =m M
R
[ 'm, u
F
` a
F
for all F.
A vertex v P gives cones
C
v
= Cone(PMv) M
R
and
v
=C

v
N
R
.
(When v =m
i
, these are the cones C
i
and
i
studied above.) Faces QP containing
the vertex v correspond bijectively to faces Q
v
C
v
via the maps
(2.3.2)
Q Q
v
= Cone(QMv)
Q
v
Q = (Q
v
+v) P
which are inverses of each other. These maps preserve dimensions, inclusions, and
intersections (Exercise 2.3.2), as illustrated in Figure 5.
0
Q
v
C
v
v
P
Q
Figure 5. The cone Cv of a vertex v P
In particular, all facets of C
v
come from facets of P containing v, so that
C
v
=m M
R
[ 'm, u
F
` 0 for all F containing v.
2.3. Polytopes and Projective Toric Varieties 77
By the duality results of Chapter 1, it follows that the dual cone
v
is given by

v
= Cone(u
F
[ F contains v).
This construction generalizes to arbitrary faces Q _P by setting

Q
= Cone(u
F
[ F contains Q).
Thus the cone
F
is the ray generated by u
F
, and
P
= 0 since 0 is the cone
generated by the empty set. Here is our main result about these cones.
Theorem 2.3.2. Let P M
R
be a full dimensional lattice polytope and set
P
=

Q
[ Q _P. Then:
(a) For all
Q

P
, each face of
Q
is also in
P
.
(b) The intersection
Q

Q
of any two cones in
P
is a face of each.
Remark 2.3.3. A nite collection of strongly convex rational polyhedral cones
that satises (a) and (b) of Theorem 2.3.2 is called a fan. General fans will be
introduced in Chapter 3. Since the cones in the fan
P
are built from the inward-
pointing normal vectors u
F
, we call
P
the normal fan or inner normal fan of P.
Theorem 2.3.2 will follow from the results proved below. But rst, we give a
simple example of a polytope and its normal fan.
Example 2.3.4. The 2-simplex
2
R
2
has vertices 0, e
1
, e
2
. Let P = k
2
for an
integer k > 0. Figure 6 shows P and its normal fan
P
. At each vertex v
i
of P, we
P
v
0
v
1
v
2

P
Figure 6. The triangle P =k2 R
2
and its normal fan P
show the cone
i
=C

v
i
generated by the normal vectors of the facets containing v
i
.
The reassembled cones appear on the right as
P
.
Notice that the cones C
v
i
M
R
do not t together nicely; rather, it is their duals

i
N
R
that give the fan
P
. This explains why cones in N
R
are the key players in
toric geometry.
Here is the rst of the results we need to prove Theorem 2.3.2.
78 Chapter 2. Projective Toric Varieties
Lemma 2.3.5. Let Q be a face of P and let H
u,b
be a supporting afne hyperplane
of P. Then u
Q
if and only if Q H
u,b
P.
Proof. First suppose that u
Q
and write u =

QF

F
u
F
,
F
0. Then setting
b
0
=

QF

F
a
F
easily implies that P H
+
u,b
0
and Q H
u,b
0
P. Recall that
the integers a
F
come from the facet presentation (2.3.1). It follows that H
u,b
0
is a
supporting hyperplane of P. Since H
u,b
is a supporting hyperplane by hypothesis,
we must have b
0
= b, hence QH
u,b
P.
Going the other way, suppose that Q H
u,b
P. Take a vertex v Q. Then
P H
+
u,b
and v H
u,b
imply that C
v
H
+
u,0
. Hence u C

v
=
v
, so that
u =

vF

F
u
F
,
F
0.
Let F
0
be a facet of P containing v but not Q, and pick p Q with p / F
0
. Then
p, v Q H
u,b
imply that
b ='p, u` =

vF

F
'p, u
F
`
b ='v, u` =

vF

F
'v, u
F
` =

vF

F
a
F
,
where the last equality uses 'v, u
F
` =a
F
for v F. These equations imply

vF

F
'p, u
F
` =

vF

F
a
F
.
However, p / F
0
gives 'p, u
F
0
` >a
F
0
, and since 'p, u
F
` a
F
for all F, it follows
that
F
0
= 0 whenever Q F
0
. This gives u
Q
and completes the proof of the
lemma.
Corollary 2.3.6. If Q_P and F P is a facet, then u
F

Q
if and only if QF.
Proof. One direction is obvious by the denition of
Q
, and the other direction
follows from Lemma 2.3.5 since H
u
F
,a
F
is a supporting afne hyperplane of P
with H
u
F
,a
F
P = F.
Theorem 2.3.2 is an immediate consequence of the following proposition.
Proposition 2.3.7. Let Q and Q

be faces of a full dimensional lattice polytope


P M
R
. Then:
(a) Q Q

if and only if
Q

Q
.
(b) If Q Q

, then
Q
is a face of
Q
, and all faces of
Q
are of this form.
(c)
Q

=
Q

, where Q

is the smallest face of P containing Q and Q

.
Proof. To prove part (a), note that if Q Q

, then any facet containing Q

also
contains Q, which implies
Q

Q
. The other direction follows easily from
Corollary 2.3.6 since every face is the intersection of the facets containing it by
Proposition 2.2.1.
2.3. Polytopes and Projective Toric Varieties 79
For part (b), x a vertex v Q and note that by (2.3.2), Q determines a face Q
v
of C
v
. Using the duality of Proposition 1.2.10, Q
v
gives the dual face
Q

v
=C

v
Q

v
=
v
Q

v
of the cone
v
. Then using
v
= Cone(u
F
[ v F) and Q
v
C
v
=

v
, one obtains
Q

v
= Cone(u
F
[ v F, Q
v
H
u
F
,0
).
Since v Q, the inclusion Q
v
H
u
F
,0
is equivalent to Q H
u
F
,a
F
, which in turn
is equivalent to Q F. It follows that
(2.3.3) Q

v
= Cone(u
F
[ QF) =
Q
,
so that
Q
is a face of
v
, and all faces of
v
arise in this way.
In particular, Q Q

means that
Q

is also a face of
v
, and since
Q


Q
by part (a), we see that
Q
is a face of
Q
. Furthermore, every face of
Q
is a face
of
v
by Proposition 1.2.6 and hence is of the form
Q
for some face Q

. Using
part (a) again, we see that Q Q

, and part (b) follows.


For part (c), let Q

be the smallest face of P containing Q and Q

. This exists
because a face is the intersection of the facets containing it, so that Q

is the inter-
section of all facets containing Q and Q

(if there are no such facets, then Q

= P).
By part (b)
Q
is a face of both
Q
and
Q
. Thus
Q

Q

Q
.
It remains to prove the opposite inclusion. If
Q

Q
=0 =
P
, then Q

=P
and we are done. If
Q

Q
= 0, any nonzero u in the intersection lies in both

Q
and
Q
. The proof of Proposition 2.3.8 given below will show that H
u,b
is a
supporting afne hyperplane of P for some b R. By Lemma 2.3.5, u
Q
and
u
Q
imply that Q and Q

lie in H
u,b
P. The latter is a face of P containing
Q and Q

, so that Q

H
u,b
P since Q

is the smallest such face. Applying


Lemma 2.3.5 again, we see that u
Q

.
Proposition 2.3.7 shows that there is a bijective correspondence between faces
of P and cones of the normal fan
P
. Here are some further properties of this
correspondence.
Proposition 2.3.8. Let P M
R
be a full dimensional lattice polytope of dimension
n and consider the cones
Q
in the normal fan
P
of P. Then:
(a) dim Q+dim
Q
= n for all faces Q _P.
(b) N
R
=

v vertex of P

v
=

Q
.
Proof. Suppose Q _ P and take a vertex v of Q. By (2.3.2) this gives a face Q
v
of
the cone C
v
, which has a dual face Q

v
of the dual cone C

v
=
v
. Since Q

v
=
Q
by (2.3.3), we have
dim Q+dim
Q
= dim Q
v
+dim Q

v
= n,
80 Chapter 2. Projective Toric Varieties
where the rst equality uses Exercise 2.3.2 and the second follows from Propo-
sition 1.2.10. This proves part (a). For part (b), let u N
R
be nonzero and set
b = min'v, u` [ v vertex of P. Then P H
+
u,b
and v H
u,b
for at least one ver-
tex of P, so that u
v
by Lemma 2.3.5. The nal equality of part (b) follows
immediately.
A fan satisfying the condition of part (b) of Proposition 2.3.8 is called com-
plete. Thus the normal fan of a lattice polytope is always complete. We will learn
more about complete fans in Chapter 3.
In general, multiplying a polytope by a positive integer has no effect on its
normal fan, and the same is true for translations by lattice points. We record these
properties in the following proposition (Exercise 2.3.3).
Proposition 2.3.9. Let P M
R
be a full dimensional lattice polytope. Then for
any lattice point m M and any integer k 1, the polytopes m+P and kP have
the same normal fan as P.
Examples of Normal Fans. Here are some more examples of normal fans.
Example 2.3.10. Figure 7 shows a lattice hexagon P in the plane together with
its normal fan. The vertices of P are labeled v
1
, . . . , v
6
, with corresponding cone

1
, . . . ,
6
in the normal fan. In the gure, P is shown on the left, and at each vertex
v
i
, we have drawn the normal vectors of the facets containing v
i
and shaded the
cone
i
they generate. On the right, these cones are assembled at the origin to give
the normal fan.
P
v
1
v
2
v
3
v
4
v
5
v
6

P
Figure 7. A lattice hexagon P and its normal fan P
Notice how one can read off the structure of P from the normal fan. For exam-
ple, two cones
i
and
j
share a ray in
P
if and only if the vertices v
i
and v
j
lie
on an edge of P.
The hexagon P is an example of a zonotope since it is a Minkowski sum of
line segments (three in this case). Notice also that
P
is determined by three lines
2.3. Polytopes and Projective Toric Varieties 81
through the origin. In 6.2 we will prove that the normal fan of any zonotope is
determined by an arrangement of hyperplanes containing the origin.
Example 2.3.11. Consider the cube P R
3
with vertices (1, 1, 1). The facet
normals are e
1
, e
2
, e
3
, and the facet presentation of P is
'm, e
i
` 1.
The origin is an interior point of P. By Exercise 2.2.1, the facet normals are the
vertices of the dual polytope P

, the octahedron in Figure 8.


y
x
z
P
y
x
z
P
Figure 8. A cube P R
3
and its dual octahedron P

However, the facet normals also give the normal fan of P, and one can check
that in the above gure, the maximal cones of the normal fan are the octants of R
3
,
which are just the cones over the facets of the dual polytope P

.
As noted earlier, it is rare that both P and P

are lattice polytopes. However,


whenever P M
R
is a lattice polytope containing 0 as an interior point, it is still
true that maximal cones of the normal fan
P
are the cones over the facets of
P

N
R
(Exercise 2.3.4).
The special behavior of the polytopes P and P

discussed in Examples 2.2.6


and 2.3.11 leads to the following denition.
Denition 2.3.12. A full dimensional lattice polytope P M
R
is reexive if its
facet presentation is
P =m M
R
[ 'm, u
F
` 1 for all facets F.
If P is reexive, then 0 is a lattice point of P and is the only interior lattice
point of P (Exercise 2.3.5). Since a
F
= 1 for all F, Exercise 2.2.1 implies that
P

= Conv(u
F
[ F facet of P).
82 Chapter 2. Projective Toric Varieties
Thus P

is a lattice polytope and is in fact reexive (Exercise 2.3.5).


We will see later that reexive polytopes lead to some very interesting toric
varieties that are important for mirror symmetry.
Intersection of Afne Pieces. Let P M
R
be a full dimensional very ample poly-
tope and set s =[PM[. This gives
X
PM
P
s1
.
If X
PM
U
v
is the afne piece corresponding to a vertex v P, then
X
PM
U
v
=U
v
= Spec(C[

v
M])
by Theorem 2.3.1. Thus the afne piece X
PM
U
v
is the toric variety of the cone

v
in the normal fan
P
of P.
Our next task is to describe the intersection of two of these afne pieces.
Proposition 2.3.13. Let P M
R
be full dimensional and very ample. If v = w are
vertices of P and Q is the smallest face of P containing v and w, then
X
PM
U
v
U
w
=U

Q
= Spec(C[

Q
M])
and the inclusions
X
PM
U
v
X
PM
U
v
U
w
X
PM
U
w
can be written
(2.3.4) U
v
(U
v
)

wv =U

Q
= (U
w
)

vw U
w
.
Proof. We analyzed the intersection of afne pieces of X
PM
in 2.1. Translated
to the notation being used here, (2.1.6) and (2.1.7) imply that
X
PM
U
v
U
w
= (U
v
)

wv = (U
w
)

vw.
Thus all we need to show is that
(U
v
)

wv =U

Q
.
However, we have wv C
v
=

v
, so that = H
wv

v
is a face of
v
. In this
situation, Proposition 1.3.16 and equation (1.3.4) imply that
(U
v
)

wv =U

.
Thus the proposition will follow once we prove =
Q
, i.e., H
wv

v
=
Q
.
Since
Q
=
v

w
by Proposition 2.3.7, it sufces to prove that
H
wv

v
=
v

w
.
Let u H
wv

v
. If u = 0, there is b R such H
u,b
is a supporting afne hyper-
plane of P. Then u
v
implies v H
u,b
by Lemma 2.3.5, so that w H
u,b
since
u H
wv
. Applying Lemma 2.3.5 again, we get u
w
. Going the other way, let
u
v

w
. If u = 0, pick b R as above. Then u
v

w
and Lemma 2.3.5
2.3. Polytopes and Projective Toric Varieties 83
imply that v, w H
u,b
, from which u H
wv
follows easily. This completes the
proof.
This proposition and Theorem 2.3.1 have the remarkable result that the normal
fan
P
completely determines the internal structure of X
PM
: we build X
PM
from
local pieces given by the afne toric varieties U
v
, glued together via (2.3.4). We
do not need the ambient projective space P
s1
for any of thiseverything we need
to know is contained in the normal fan.
The Toric Variety of a Polytope. We can now give the general denition of the
toric variety of a polytope.
Denition 2.3.14. Let P M
R
be a full dimensional lattice polytope. Then we
dene the toric variety of P to be
X
P
= X
(kP)M
where k is any positive integer such that kP is very ample.
Such integers k exist by Corollary 2.2.19, and if k and are two such integers,
then kP and P have the same normal fan by Proposition 2.3.9, namely
kP
=

P
=
P
. It follows that while X
(kP)M
and X
(P)M
lie in different projective
spaces, they are built from the afne toric varieties U
v
glued together via (2.3.4).
Once we develop the language of abstract varieties in Chapter 3, we will see that
X
P
is well-dened as an abstract variety.
We will often speak of X
P
without regard to the projective embedding. When
we want to use a specic embedding, we will say X
P
is embedded using kP,
where we assume that kP is very ample. In Chapter 6 we will use the language of
divisors and line bundles to restate this in terms of a divisor D
P
on X
P
such that
kD
P
is very ample precisely when kP is.
Here is a simple example to illustrate the difference between X
P
as an abstract
variety and X
P
as sitting in a specic projective space.
Example 2.3.15. Consider the n-simplex
n
R
n
. We can dene X
n
using k
n
for any integer k 1 since
n
is normal and hence very ample. The lattice points
in k
n
correspond to the s
k
=

n+k
k

monomials of C[t
1
, . . . , t
n
] of total degree k.
This gives an embedding X
n
P
s
k
1
. When k = 1,
n
Z
n
= 0, e
1
, . . . , e
n

implies that
X
n
=P
n
.
The normal fan of
n
is described in Exercise 2.3.6. For an arbitrary k 1, we can
regard X
n
P
s
k
1
as the image of the map

k
: P
n
P
s
k
1
84 Chapter 2. Projective Toric Varieties
dened using all monomials of total degree k in C[x
0
, . . . , x
n
] (Exercise 2.3.6). It
follows that this map is an embedding, usually called the Veronese embedding. But
when we forget the embedding, the underlying toric variety is just P
n
.
The Veronese embedding allows us to construct some interesting afne open
subsets of P
n
. Let f C[x
0
, . . . , x
n
] be nonzero and homogeneous of degree k and
write f =

[[=k
c

. We write the homogeneous coordinates of P


s1
as y

for
[[ = k. Then L =

[[=k
c

is a nonzero linear form in the variables y

, so that
P
s
k
1
`V(L) is a copy of C
s
k
1
(Exercise 2.3.6). If follows that
P
n
`V( f )
k
(P
n
)

P
s
k
1
`V(L)

is an afne variety (usually not toric). This shows that P


n
has a richer supply of
afne open subsets than just the open sets U
i
=P
n
`V(x
i
) considered earlier in the
chapter.
When we explain the Proj construction of P
n
later in the book, we will see the
intrinsic reason why P
n
`V( f ) is an afne open subset of P
n
.
Example 2.3.16. The 2-dimensional analog of the rational normal curve C
d
is the
rational normal scroll S
a,b
, which is the toric variety of the polygon
P
a,b
= Conv(0, ae
1
, e
2
, be
1
+e
2
) R
2
,
where a, b N satisfy 1 a b. The polygon P = P
2,4
and its normal fan are
pictured in Figure 9.
P = P
2,4
v
1
= 0 v
4
= 2e
1
v
2
= e
2
v
3
= 4e
1
+ e
2

P
Figure 9. The polygon of a rational normal scroll and its normal fan
In general, the polygon P
a,b
has a+b+2 lattice points and gives the map
(C

)
2
P
a+b+1
, (s, t) (1, s, s
2
, . . . , s
a
, t, st, s
2
t, . . . , s
b
t)
such that S
a,b
= X
P
a,b
is the Zariski closure of the image. To describe the image, we
rewrite the map as
CP
1
P
a+b+1
, (s, , ) (, s, s
2
, . . . , s
a
, , s, s
2
, . . . , s
b
).
2.3. Polytopes and Projective Toric Varieties 85
When (, ) = (1, 0), the map is s (1, s, s
2
, . . . , s
a
, 0, . . . , 0), which is the rational
normal curve C
a
mapped to the rst a +1 coordinates of P
a+b+1
. In the same
way, (, ) = (0, 1) gives the rational normal curve C
b
mapped to the last b +1
coordinates of P
a+b+1
. If we think of these two curves as the edges of a scroll,
then xing s gives a point on each edge, and letting (, ) P
1
vary gives the line
of the scroll connecting the two points. So it really is a scroll!
An important observation is that the normal fan depends only on the difference
ba, since this determines the slope of the slanted edge of P
a,b
. If we denote the
difference by r N, it follows that as abstract toric varieties, we have
X
P
1,r+1
= X
P
2,r+2
= X
P
3,r+3
=
since they are all constructed from the same normal fan. In Chapter 3, we will see
that this is the Hirzebruch surface H
r
.
But if we think of the projective surface S
a,b
P
a+b+1
, then a and b have a
unique meaning. For example, they have a strong inuence on the dening equa-
tions of S
a,b
. Let the homogeneous coordinates of P
a+b+1
be x
0
, . . . , x
a
, y
0
, . . . , y
b
and consider the 2(a+b) matrix

x
0
x
1
x
a1
x
1
x
2
x
a

y
0
y
1
y
b1
y
1
y
2
y
b

.
One can show that I(S
a,b
) C[x
0
, . . . , x
a
, y
0
, . . . , y
b
] is generated by the 22 minors
of this matrix (see [130, Ex. 9.11], for example).
Example 2.3.16 is another example of a determinantal variety, as is the rational
normal curve from Example 2.0.1. Note that the rational normal curve C
d
comes
from the polytope [0, d] = d
1
, where the underlying toric variety is just P
1
.
Exercises for 2.3.
2.3.1. This exercise will use the same notation as the proof of Theorem 2.3.1.
(a) Let H
u,a
be a supporting hyperplane of a vertex m
i
P. Prove that H
u,0
is a supporting
hyperplane of 0 C
i
(b) Prove that dimC
i
= dim P.
2.3.2. Consider the maps dened in (2.3.2).
(a) Show that these maps are inverses of each other and dene a bijection between the
faces of the cone C
v
and the faces of P containing v.
(b) Prove that these maps preserve dimensions, inclusions, and intersections.
(c) Explain how this exercise relates to Exercise 2.3.1.
2.3.3. Prove Proposition 2.3.9.
2.3.4. Let P M
R
be a full dimensional lattice polytope containing 0 as an interior point,
and let P

N
R
be its dual polytope. Prove that the normal fan
P
consists of the cones
over the faces of P

. Hint: Exercise 2.2.1 will be useful.


86 Chapter 2. Projective Toric Varieties
2.3.5. Let P M
R
be a reexive polytope.
(a) Prove that 0 is the only interior lattice point of P.
(b) Prove that P

N
R
is reexive.
2.3.6. This exercise is concerned with Example 2.3.15.
(a) Let e
1
, . . . , e
n
be the standard basis of R
n
. Prove that the normal fan of the standard
n-simplex consists of the cones Cone(S) for all proper subsets S e
0
, e
1
, . . . , e
n
,
where e
0
=

n
i=1
e
i
. Draw pictures of the normal fan for n = 1, 2, 3.
(b) For an integer k 1, show that the toric variety X
kn
P
s
k
1
is given by the map

k
: P
n
P
s
k
1
dened using all monomials of total degree k in C[x
0
, . . . , x
n
].
2.3.7. Let P M
R
R
n
be an n-dimensional lattice polytope and let Q P be a face.
Prove the following intrinsic description of the cone
Q

P
:

Q
=u N
R
[ 'm, u` 'm

, u` for all m Q, m

P.
2.3.8. Prove that all lattice rectangles in the plane with edges parallel to the coordinate
axes have the same normal fan.
2.4. Properties of Projective Toric Varieties
We conclude this chapter by studying when the projective toric variety X
P
of a
polytope P is smooth or normal.
Normality. Recall from 2.1 that a projective variety is projectively normal if its
afne cone is normal.
Theorem 2.4.1. Let P M
R
be a full dimensional lattice polytope. Then:
(a) X
P
is normal.
(b) X
P
is projectively normal under the embedding given by kP if and only if kP is
normal.
Proof. Part (a) is immediate since X
P
is the union of afne pieces U
v
, v a vertex
of P, and U
v
is normal by Theorem 1.3.5. In Chapter 3 we will give an intrinsic
denition of normality that will make this argument completely rigorous.
It remains to prove part (b). The discussion following (2.1.4) shows that the
projective embedding of X
P
given by X
(kP)M
has Y
((kP)M)|1
as its afne cone.
By Theorem 1.3.5, this is normal if and only if the semigroup N

((kP) M) 1

is saturated in MZ. Since ((kP) M) 1 generates the cone C(P), this is


equivalent to the assertion that the semigroup C(P) (MZ) is generated by
((kP) M) 1. Then we are done by Lemma 2.2.14.
Smoothness. Given the results of Chapter 1, the smoothness of X
P
is equally easy
to determine. We need one denition.
2.4. Properties of Projective Toric Varieties 87
Denition 2.4.2. Let P M
R
be a lattice polytope.
(a) Given a vertex v of P and an edge E containing v, let w
E
be the rst lattice
point of E different from v encountered as one tranverses E starting at v. In
other words, w
E
v is the ray generator of the ray Cone(E v).
(b) P is smooth if for every vertex v, the vectors w
E
v, where E is an edge of P
containing v, form a subset of a basis of M. In particular, if dim P = dim M
R
,
then the vectors w
E
v form a basis of M.
We can now characterize when X
P
is smooth.
Theorem 2.4.3. Let P M
R
be a full dimensional lattice polytope. Then the fol-
lowing are equivalent:
(a) X
P
is a smooth projective variety.
(b)
P
is a smooth fan, meaning that every cone in
P
is smooth in the sense of
Denition 1.2.16.
(c) P is a smooth polytope.
Proof. Smoothness is a local condition, so that a variety is smooth if and only if its
local pieces are smooth. Thus X
P
is smooth if and only if U
v
is smooth for every
vertex v of P, and U
v
is smooth if and only if
v
is smooth by Theorem 1.3.12.
Since faces of smooth cones are smooth and
P
consists of the
v
and their faces,
the equivalence (a) (b) follows immediately.
For (b) (c), rst observe that
v
is smooth if and only if its dual C
v
=

v
is
smooth. The discussion following (2.3.2) makes it easy to see that the ray genera-
tors of C
v
are the vectors w
E
v from Denition 2.4.2. It follows immediately that
P is smooth if and only if C
v
is smooth for every vertex v, and we are done.
The theorem makes it easy to check the smoothness of simple examples such
as the toric variety of the hexagon in Example 2.3.10 or the rational normal scroll
S
a,b
of Example 2.3.16 (Exercise 2.4.1).
We also note the following useful fact, which you will prove in Exercise 2.4.2.
Proposition 2.4.4. Every smooth full dimensional lattice polytope P M
R
is very
ample.
One can also ask whether every smooth lattice polytope is normal. This is an
important open problem in the study of lattice polytopes.
Here is an example of a smooth reexive polytope whose dual is not smooth.
Example 2.4.5. Let P = (n +1)
n
(1, . . . , 1) R
n
, where
n
is the standard
n-simplex. Thus P is the translate of (n+1)
n
by (1, . . . , 1). Proposition 2.3.9
implies that P and
n
have the same normal fan, so that P and X
P
are smooth. Note
also that X
P
= X
n
=P
n
.
88 Chapter 2. Projective Toric Varieties
The polytope P has the following interesting properties (Exercise 2.4.3). First,
P has the facet presentation
x
i
1, i = 1, . . . , n,
x
1
x
n
1,
so that P is reexive with dual
P

= Conv(e
0
, e
1
, . . . , e
n
), e
0
=e
1
e
n
.
The normal fan of P

consists of cones over the faces of P. In particular, the cone


of
P
corresponding to the vertex e
0
P

is the cone

e
0
= Conv(v
1
, . . . , v
n
), v
i
= e
0
+(n+1)e
i
.
Figure 10 shows P and the cone
e
0
when n = 2.
P
v
1
v
2

e
0
e
0
e
1
e
2
P
Figure 10. The cone e
0
of the normal fan of P

For general n, observe that v


i
v
j
= (n +1)(e
i
e
j
). This makes it easy to
see that Zv
1
+ +Zv
n
has index (n+1)
n1
in Z
n
. Thus
e
0
is not smooth when
n 2. It follows that the dual toric variety X
P
is singular for n 2. Later we
will construct X
P
as the quotient of P
n
under the action of a nite group G
(Z/(n+1)Z)
n1
.
Example 2.4.6. Consider P = Conv(0, 2e
1
, e
2
) R
2
. Since P is very ample, the
lattice points PZ
2
=0, e
1
, 2e
1
, e
2
give the map (C

)
2
P
3
dened by
(s, t) (1, s, s
2
, t)
such that X
P
is the Zariski closure of the image. If P
3
has homogeneous coordinates
y
0
, y
1
, y
2
, y
3
, then we have
X
P
= V(y
0
y
2
y
2
1
) P
3
.
Comparing this to Example 2.0.5, we see that X
P
is the weighted projective space
P(1, 1, 2). Later we will learn the systematic reason why this is true.
2.4. Properties of Projective Toric Varieties 89
The variety X
P
is not smooth. By working on the afne piece X
P
U
3
, one can
check directly that (0, 0, 0, 1) is a singular point of X
P
.
We can also use Theorem 2.4.3 and the normal fan of P, shown in Figure 11.
One can check that the cones
0
and
1
are smooth, but
2
is not, so that
P
P
v
0
v
1
v
2

P
Figure 11. The polygon giving P(1, 1, 2) and its normal fan
is not a smooth fan. In terms of P, note that the vectors from v
2
to the rst lattice
points along the edges containing v
2
do not generate Z
2
. Either way, Theorem 2.4.3
implies that X
P
is not smooth.
If you look carefully, you will see that
2
is the only nonsmooth cone of the
normal fan
P
. Once we study the correspondence between cones and orbits in
Chapter 3, we will see that the nonsmooth cone
2
corresponds to the singular
point (0, 0, 0, 1) of X
P
.
Products of Projective Toric Varieties. Our nal task is to understand the toric
variety of a product of polytopes. Let P
i
(M
i
)
R
R
n
i
be lattice polytopes with
dim P
i
= n
i
for i = 1, 2. This gives a lattice polytope P
1
P
2
(M
1
M
2
)
R
of
dimension n
1
+n
2
.
Replacing P
1
and P
2
with suitable multiples, we can assume that P
1
and P
2
are
very ample. This gives projective embeddings
X
P
i
P
s
i
1
, s
i
=[P
i
M
i
[,
so that by Proposition 2.0.4, X
P
1
X
P
2
is a subvariety of P
s
1
1
P
s
2
1
. Using the
Segre embedding
P
s
1
1
P
s
2
1
P
s1
, s = s
1
s
2
,
we get an embedding
(2.4.1) X
P
1
X
P
2
P
s1
.
We can understand this projective variety as follows.
90 Chapter 2. Projective Toric Varieties
Theorem 2.4.7. If P
1
and P
2
are very ample, then:
(a) P
1
P
2
(M
1
M
2
)
R
is a very ample polytope with lattice points
(P
1
P
2
) (M
1
M
2
) = (P
1
M
1
) (P
2
M
2
).
Thus the integer s dened above is s =[(P
1
P
2
) (M
1
M
2
)[.
(b) The image of the embedding X
P
1
P
2
P
s1
determined by the very ample
polytope P
1
P
2
equals the image of (2.4.1).
(c) X
P
1
P
2
X
P
1
X
P
2
.
Proof. For part (a), the assertions about lattice points are clear. The vertices of
P
1
P
2
consist of ordered pairs (v
1
, v
2
) where v
i
is a vertex of P
i
(Exercise 2.4.4).
Given such a vertex, we have
(P
1
P
2
) (M
1
M
2
) (v
1
, v
2
) = (P
1
M
1
v
1
) (P
2
M
2
v
2
).
Since P
i
is very ample, we know that N(P
i
M
i
v
i
) is saturated in M
i
. From here,
it follows easily that P
1
P
2
is very ample.
For part (b), let T
N
i
be the torus of X
P
i
. Since T
N
i
is Zariski dense in X
P
i
, it fol-
lows that T
N
1
T
N
2
is Zariski dense in X
P
1
X
P
2
(Exercise 2.4.4). When combined
with the Segre embedding, it follows that X
P
1
X
P
2
is the Zariski closure of the
image of the map
T
N
1
T
N
2
P
s
1
s
2
1
given by the characters
m

, where m ranges over the s


1
elements of P
1
M
1
and m

ranges over the s


2
elements of P
2
M
2
. When we identify T
N
1
T
N
2
with
T
N
1
N
2
, the product
m

becomes the character


(m,m

)
, so that the above map
coincides with the map
T
N
1
N
2
P
s1
coming from the product polytope P
1
P
2
(M
1
M
2
)
R
. Part (b) follows, and
part (c) is an immediate consequence.
Here is an obvious example.
Example 2.4.8. Since P
n
is the toric variety of the standard n-simplex
n
, it fol-
lows that P
n
P
m
is the toric variety of
n

m
.
This also works for more than two factors. Thus P
1
P
1
P
1
is the toric
variety of the cube pictured in Figure 8.
To have a complete theory of products, we need to know what happens to the
normal fan. Here is the result, whose proof is left to the reader (Exercise 2.4.5).
Proposition 2.4.9. Let P
i
(M
i
)
R
be full dimensional lattice polytopes for i =1, 2.
Then

P
1
P
2
=
P
1

P
2
.
2.4. Properties of Projective Toric Varieties 91
Here is an easy example.
Example 2.4.10. The normal fan of an interval [a, b] R, where a < b in Z, is
given by
s
0
1

0
The corresponding toric variety is P
1
. The cartesian product of two such intervals
is a lattice rectangle whose toric variety is P
1
P
1
by Theorem 2.4.7. If we set

i j
=
i

j
, then Proposition 2.4.9 gives the normal fan given in Figure 12.

00

10

11

01
Figure 12. The normal fan of a lattice rectangle giving P
1
P
1
We will revisit this example in Chapter 3 when we construct toric varieties
directly from fans.
Proposition 2.4.9 suggests a different way to think about the product. Let v
i
range over the vertices of P
i
for i = 1, 2. Then the
v
i
are the maximal cones in the
normal fan
P
i
, which implies that
(2.4.2) X
P
i
=

v
i
U
v
i
, i = 1, 2.
Thus
X
P
1
X
P
2
=

v
1
U
v
1

v
2
U
v
2

(v
1
,v
2
)
U
v
1
U
v
2
=

(v
1
,v
2
)
U
v
1
v
2
=

(v
1
,v
2
)
U

(v
1
,v
2
)
= X
P
1
P
2
.
In this sequence of equalities, the rst follows from (2.4.2), the second is obvious,
the third uses Exercise 1.3.13, the fourth uses Proposition 2.4.9, and the last follows
since (v
1
, v
2
) ranges over all vertices of P
1
P
2
.
This argument shows that we can construct cartesian products of varieties using
afne open covers, which reduces to the cartesian product of afne varieties dened
in Chapter 1. We will use this idea in Chapter 3 to dene the cartesian product of
abstract varieties.
92 Chapter 2. Projective Toric Varieties
Exercises for 2.4.
2.4.1. Show that the hexagon P = Conv(0, e
1
, e
2
, 2e
1
+e
2
, e
1
+2e
2
, 2e
1
+2e
2
) pictured in
Figure 6 and the trapezoid P
a,b
pictured in Figure 9 are smooth polygons. Also, of the
polytopes shown in Figure 8, determine which ones are smooth.
2.4.2. Prove Proposition 2.4.4.
2.4.3. Consider the polytope P = (n +1)
n
(1, . . . , 1) from Example 2.4.5.
(a) Verify the facet presentation of P given in the example.
(b) What is the facet presentation of P

? Hint: You know the vertices of P.


(c) Let v
i
= e
0
+(n +1)e
i
, where i = 1, . . . , n and e
0
= e
1
e
n
, and then set L =
Zv
1
+ +Zv
n
. Use the hint given in the text to prove Z
n
/L (Z/(n+1)Z)
n1
. This
shows that the index of L in Z
n
is (n +1)
n1
, as claimed in the text.
2.4.4. Let P
i
(M
i
)
R
R
ni
be lattice polytopes with dim P
i
= n
i
for i = 1, 2. Also let S
i
be the set of vertices of P
i
.
(a) Use supporting hyperplanes to prove that every element of S
1
S
2
is a vertex of P
1
P
2
.
(b) Prove that P
1
P
2
= Conv(S
1
S
2
) and conclude that S
1
S
2
is the set of vertices of
P
1
P
2
.
2.4.5. The goal of this exercise is to prove Proposition 2.4.9. We knowfromExercise 2.4.4
that the vertices of P
1
P
2
are the ordered pairs (v
1
, v
2
) where v
i
is a vertex of P
i
.
(a) Adapt the argument of part (a) of Theorem 2.4.7 to show that C
(v1,v2)
= C
v1
C
v2
.
Taking duals, we see that the maximal cones of
P1P2
are
(v1,v2)
=
v1

v2
.
(b) Given rational polyhedral cones
i
(N
i
)
R
and faces
i

i
, prove that
1

2
is a
face of
1

2
and that all faces of
1

2
arise this way.
(c) Prove that
P1P2
=
P1

P2
.
2.4.6. Consider positive integers 1 = q
0
q
1
q
n
with the property that q
i
[

n
j=0
q
j
for i = 0, . . . , n. Set k
i
=

n
j=0
q
j

/q
i
for i = 1, . . . , n and let
P
q0,...,qn
= Conv(0, k
1
e
1
, k
2
e
2
, . . . , k
n
e
n
) (1, . . . , 1).
Prove that P
q0,...,qn
is reexive and explain how it relates to Example 2.4.5. We will prove
later that the toric variety of this polytope is the weighted projective space P(q
0
, . . . , q
n
).
2.4.7. The Sylvester sequence is dened by a
0
= 2 and a
k+1
= 1 +a
0
a
1
a
k
. It begins
2, 3, 7, 43, 1807, . . . and is described in [251, A000058]. Now x a positive integer n 3
and dene q
0
, . . . , q
n
by q
0
= q
1
= 1 and q
i
= 2(a
n1
1)/a
ni
for i = 2, . . . , n. For n = 3
and 4 this gives 1, 1, 4, 6 and 1, 1, 12, 28, 42. Prove that q
0
, . . . , q
n
satises the conditions
of Exercise 2.4.6 and hence gives a reexive simplex, denoted S
Q

n
in [215]. This paper
proves that when n 4, S
Q

n
has the largest volume of all n-dimensional reexive simplices
and conjectures that it also has the largest number of lattice points.
Chapter 3
Normal Toric Varieties
3.0. Background: Abstract Varieties
The projective toric varieties studied in Chapter 2 are unions of Zariski open sets,
each of which is an afne variety. We begin with a general construction of abstract
varieties obtained by gluing together afne varieties in an analogous way. The
resulting varieties will be abstract in the sense that they do not come with any
given ambient afne or projective space. We will see that this is exactly the idea
needed to construct a toric variety using the combinatorial data contained in a fan.
Sheaf theory, while important for later chapters, will make only a modest ap-
pearance here. For a more general approach to the concept of abstract variety, we
recommend standard books such as [90], [131] or [245].
Regular Functions. Let V = Spec(R) be an afne variety. In 1.0, we dened the
Zariski open subset V
f
=V ` V( f ) V for f R and showed that V
f
=Spec(R
f
),
where R
f
is the localization of R at f . The open sets V
f
form a basis for the Zariski
topology on V in the sense that every open set U is a (nite) union U =

f S
V
f
for some S R (Exercise 3.0.1).
For an afne variety, a morphism V C is called a regular map, so that the
coordinate ring of V consists of all regular maps fromV to C. We now dene what
it means to be regular on an open subset of V.
Denition 3.0.1. Given an afne variety V = Spec(R) and a Zariski open U V,
we say a function : U C is regular if for all p U, there exists f
p
R such
that p V
fp
U and [
V
f p
R
fp
. Then dene
O
V
(U) = : U C [ is regular.
93
94 Chapter 3. Normal Toric Varieties
The condition p V
fp
means that f
p
(p) = 0, and saying [
V
f p
R
fp
means that
= a
p
/f
np
p
for some a
p
R and n
p
0.
Here are some cases where O
V
(U) is easy to compute.
Proposition 3.0.2. Let V = Spec(R) be an afne variety. Then:
(a) O
V
(V) = R.
(b) If f R, then O
V
(V
f
) = R
f
.
Proof. It is clear from Denition 3.0.1 that elements of R dene regular functions
on V, hence elements of O
V
(V). Conversely, if O
V
(V), then for all p V there
is f
p
R such that p V
fp
and = a
p
/f
np
p
R
fp
. The ideal I =' f
np
p
[ p V` R
satises V(I) = since f
p
(p) = 0 for all p V. Hence the Nullstellensatz implies
that

I = I(V(I)) = R, so there exists a nite set S V and polynomials g
p
for
p S such that
1 =

pS
g
p
f
np
p
.
Hence =

pS
g
p
f
np
p
=

pS
g
p
a
p
R, as desired.
For part (b), let U V
f
be Zariski open. Then U is Zariski open in V, and
whenever g R satises V
g
U, we have V
g
=V
f g
with coordinate ring
R
f g
= (R
f
)
g/f

for all 0. These observations easily imply that


(3.0.1) O
V
(U) =O
V
f
(U).
Then setting U =V
f
gives
O
V
(V
f
) =O
V
f
(V
f
) = R
f
,
where the last equality follows by applying part (a) to V
f
= Spec(R
f
).
Local Rings. When V = Spec(R) is an irreducible afne variety, we can describe
regular functions using the local rings O
V,p
introduced in 1.0. A rational func-
tion in C(V) is contained in the local ring O
V,p
precisely when it is regular in a
neighborhood of p. It follows that whenever U V is open, we have

pU
O
V,p
=O
V
(U).
Thus regular functions onU are rational functions onV that are dened everywhere
on U. In particular, when U =V, Proposition 3.0.2 implies that
(3.0.2)

pV
O
V,p
=O
V
(V) = R =C[V].
3.0. Background: Abstract Varieties 95
The Structure Sheaf of an Afne Variety. Given an afne variety V, the mapping
U O
V
(U), U V open,
has the following useful properties:
When U

U, Denition 3.0.1 shows that there is an obvious restriction map

U,U
: O
V
(U) O
V
(U

)
dened by
U,U
() = [
U
. It follows that
U,U
is the identity map and that

,U

U,U
=
U,U
whenever U

U.
If U

is an open cover of U V, then the sequence


0 O
V
(U)

O
V
(U

,
O
V
(U

)
is exact. Here, the second arrow is dened by the restrictions
U,U
and the
double arrow is dened by
U,UU

and
U

,UU

. Exactness at O
V
(U)
means that regular functions are determined locally, i.e., two regular functions
on U are equal if their restrictions to all U

are equal. For the middle term,


exactness means that we have an equalizer: an element ( f

O
V
(U

)
comes from f O
V
(U) if and only if the restrictions f
[
UU

= f
[
UU

are
equal for all , . This is true because regular functions on the U

agreeing on
the overlaps U

patch together to give a regular function on U.


In the language of sheaf theory, these properties imply that O
V
is a sheaf of C-
algebras, called the structure sheaf of V. We call (V, O
V
) a ringed space over C.
Also, since (3.0.1) holds for all open sets U V
f
, we write
O
V [
V
f
=O
V
f
.
In terms of ringed spaces, this means (V
f
, O
V [
V
f
) = (V
f
, O
V
f
).
Morphisms. By 1.0, a polynomial mapping : V
1
V
2
between afne vari-
eties corresponds to the C-algebra homomorphism

: C[V
2
] C[V
1
] dened by

() = for C[V
2
]. We now extend this to open sets of afne varieties.
Denition 3.0.3. Let U
i
V
i
be Zariski open subsets of afne varieties for i = 1, 2.
A function : U
1
U
2
is a morphism if denes a map

: O
V
2
(U
2
) O
V
1
(U
1
).
Thus : U
1
U
2
is a morphism if composing with regular functions on U
2
gives regular functions on U
1
. Note also that

is a C-algebra homomorphism
since it comes from composition of functions.
Example 3.0.4. Suppose that : V
1
V
2
is a morphism according to Deni-
tion 3.0.3. If V
i
= Spec(R
i
), then the above map

gives the C-algebra homo-


morphism
R
2
=O
V
2
(V
2
) O
V
1
(V
1
) = R
1
.
96 Chapter 3. Normal Toric Varieties
By Chapter 1, the C-algebra homomorphism R
2
R
1
gives a map of afne va-
rieties V
1
V
2
. In Exercise 3.0.3 you will show that this is the original map
: V
1
V
2
we started with.
Example 3.0.4 shows that when we apply Denition 3.0.3 to maps between
afne varieties, we get the same morphisms as in Chapter 1. In Exercise 3.0.3 you
will verify the following properties of morphisms:
If U is open in an afne variety V, then
O
V
(U) = : U C [ is a morphism.
Hence regular functions on U are just morphisms from U to C.
A composition of morphisms is a morphism.
An inclusion of open sets W U of an afne variety V is a morphism.
Morphisms are continuous in the Zariski topology.
We say that a morphism : U
1
U
2
is an isomorphism if is bijective and its
inverse function
1
: U
2
U
1
is also a morphism.
Gluing Together Afne Varieties. We now are ready to dene abstract varieties by
gluing together open subsets of afne varieties. The model is what happens for P
n
.
Recall from 2.0 of that P
n
is covered by open sets
U
i
=P
n
`V(x
i
) = Spec

x
0
x
i
, . . . ,
x
i1
x
i
,
x
i+1
x
i
, . . . ,
xn
x
i

for i = 0, . . . , n. Each U
i
is a copy of C
n
that uses a different set of variables. For
i = j, we glue together these copies as follows. We have open subsets
(3.0.3) (U
i
)
x
j
x
i
U
i
and (U
j
)
x
i
x
j
U
j
,
and we also have the isomorphism
(3.0.4) g
ji
: (U
i
)
x
j
x
i

(U
j
)
x
i
x
j
since both give the same open set U
i
U
j
in P
n
. The notation g
ji
was chosen so
that g
ji
(x) means x U
i
since the index i is closest to x, hence g
ji
(x) U
j
. At the
level of coordinate rings, g
ji
comes from the isomorphism
g

ji
: C

x
0
x
j
, . . . ,
x
j1
x
j
,
x
j+1
x
j
, . . . ,
xn
x
j

x
i
x
j
C

x
0
x
i
, . . . ,
x
i1
x
i
,
x
i+1
x
i
, . . . ,
xn
x
i

x
j
x
i
dened by
x
k
x
j

x
k
x
i
/
x
j
x
i
(k = j) and

x
i
x
j

x
j
x
i
.
We can turn this around and start from the afne varieties U
i
C
n
given above
and glue together the open sets in (3.0.3) using the isomorphisms g
ji
from (3.0.4).
This gluing is consistent since g
i j
= g
1
ji
and g
ki
= g
k j
g
ji
wherever all three maps
are dened. The result of this gluing is the projective space P
n
.
3.0. Background: Abstract Varieties 97
To generalize this, suppose we have a nite collection V

of afne varieties
and for all pairs , we have Zariski open sets V

and isomorphisms g

:
V

satisfying the following compatibility conditions:


g

= g
1

for all pairs , .


g

(V

) =V

and g

= g

on V

for all , , .
The notation g

means that in the expression g

(x), the point x lies in V

since
is the index closest to x, and the result g

(x) lies in V

.
We are now ready to glue. Let Y be the disjoint union of the V

and dene
a relation on Y by a b if and only if a V

, b V

for some , with b =


g

(a). The rst compatibility condition shows that is reexive and symmetric;
the second shows that it is transitive. Hence is an equivalence relation and we
can form the quotient space X =Y/ with the quotient topology. For each , let
U

=[a] X [ a V

.
Then U

X is an open set and the map h

(a) = [a] denes a homeomorphism


h

: V

X. Thus X locally looks like an afne variety.


Denition 3.0.5. We call X the abstract variety determined by the above data.
An abstract variety X comes equipped with the Zariski topology whose open
sets are those sets that restrict to open sets in each U

. The Zariski closed subsets


Y X are called subvarieties of X. We say that X is irreducible if it is not the union
of two proper subvarieties. One can show that X is a nite union of irreducible
subvarieties X =Y
1
Y
s
such that Y
i
Y
j
for i = j. We call the Y
i
the irreducible
components of X.
Here are some examples of Denition 3.0.5.
Example 3.0.6. We saw above that P
n
can be obtained by gluing together the
open sets (3.0.3) using the isomorphisms g
i j
from (3.0.4). This shows that P
n
is an abstract variety with afne open subsets U
i
P
n
. More generally, given a
projective variety V P
n
, we can cover V with afne open subsets V U
i
, and the
gluing implicit in equation (2.0.8). We conclude that projective varieties are also
abstract varieties.
Example 3.0.7. In a similar way, P
n
C
m
can be viewed as gluing afne spaces
U
i
C
m
C
n+m
along suitable open subsets. Thus P
n
C
m
is an abstract variety,
and the same is true for subvarieties V P
n
C
m
.
Example 3.0.8. Let V
0
=C
2
= Spec(C[u, v]) and V
1
=C
2
= Spec(C[w, z]), with
V
10
=V
0
`V(v) = Spec(C[u, v]
v
)
V
01
=V
1
`V(z) = Spec(C[w, z]
z
)
98 Chapter 3. Normal Toric Varieties
and gluing data
g
10
: V
10
V
01
coming from the C-algebra homomorphism
g

10
: C[w, z]
z
C[u, v]
v
dened by w uv and z 1/v
and
g
01
: V
01
V
10
coming from the C-algebra homomorphism
g

01
: C[u, v]
v
C[w, z]
z
dened by u wz and v 1/z.
One checks that g
01
= g
1
10
, and the other compatibility condition is satised since
there are only two V
i
. It follows that we get an abstract variety X.
The variety X has another description. Consider the product P
1
C
2
with
homogeneous coordinates (x
0
, x
1
) on P
1
and coordinates (x, y) on C
2
. We will
identify X with the subvariety W = V(x
0
y x
1
x) P
1
C
2
, called the blowup of
C
2
at the origin, and denoted Bl
0
(C
2
). First note that P
1
C
2
is covered by
U
0
C
2
= Spec(C[x
1
/x
0
, x, y]) and U
1
C
2
= Spec(C[x
0
/x
1
, x, y]).
Then W is covered by W
0
=W (U
0
C
2
) and W
1
=W (U
1
C
2
). Also,
W
0
= V(y (x
1
/x
0
)x) U
0
C
2
,
which gives the coordinate ring
C[x
1
/x
0
, x, y]/'y (x
1
/x
0
)x` C[x, x
1
/x
0
] via y (x
1
/x
0
)x.
Similarly, W
1
= V(x (x
0
/x
1
)y) U
1
C
2
has coordinate ring
C[x
0
/x
1
, x, y]/'x (x
0
/x
1
)y` C[y, x
0
/x
1
] via x (x
0
/x
1
)y.
You can check that these are glued together in W in exactly the same way V
0
and
V
1
are glued together in X. We will generalize this example in Exercise 3.0.8.
Morphisms Between Abstract Varieties. Let X and Y be abstract varieties with
afne open covers X =

and Y =

. A morphism : X Y is a Zariski
continuous mapping such that the restrictions
[
U
1
(U

)
: U

1
(U

) U

are morphisms in the sense of Denition 3.0.3.


The Structure Sheaf of an Abstract Variety. Let U be an an open subset of an
abstract variety X and set W

= h
1

(U U

) V

. Then a function : U C is
regular if
h
[
W
: W

C
is regular for all . The compatibility conditions ensure that this is well-dened,
so that one can dene
O
X
(U) = : U C [ is regular.
3.0. Background: Abstract Varieties 99
This gives the structure sheaf O
X
of X. Thus an abstract variety is really a ringed
space (X, O
X
) with a nite open covering U

such that (U

, O
X[
U
) is isomor-
phic to the ringed space (V

, O
V
) of the afne variety V

. (We leave the denition


of isomorphism of ringed spaces to the reader.)
Open and Closed Subvarieties. Given an abstract variety X and an open subset
U, we note that U has a natural structure of an abstract variety. For an afne
open subset U

X, U U

is open in U

and hence can be written as a union


U U

f S
(U

)
f
for a nite subset S C[U

]. It follows that U is covered


by nitely many afne open subsets and thus is an abstract variety. The structure
sheaf O
U
is simply the restriction of O
X
to U, i.e., O
U
= O
X[
U
. Note also that a
function : U C is regular if and only if is a morphism as dened above.
In a similar way, a closed subset Y X also gives an abstract variety. For an
afne open set U X, Y U is closed in U and hence is an afne variety. Thus
Y is covered by nitely many afne open subsets and thus is an abstract variety.
This justies the term subvariety for closed subsets of an abstract variety. The
structure sheaf O
Y
is related to O
X
as follows. The inclusion i : Y X is a mor-
phism. Let i

O
Y
be the sheaf on X dened by i

O
Y
(U) = O
Y
(U Y). Restricting
functions on X to functions on Y gives a map of sheaves O
X
i

O
Y
whose kernel
is the subsheaf I
Y
O
X
of functions vanishing on Y, meaning
I
Y
(U) =f O
X
(U) [ f (p) = 0 for all p Y U.
In the language of Chapter 6, we have an exact sequence of sheaves
0 I
Y
O
X
i

O
Y
0.
All of the types of variety introduced so far can be subsumed under the con-
cept of abstract variety. From now on, we will usually be thinking of abstract
varieties. Hence we will usually say variety rather than abstract variety.
Local Rings and Rational Functions. Let p be a point of an afne variety V.
Elements of the local ring O
V,p
are quotients f /g in a suitable localization with
f , g C[V] and g(p) = 0. It follows that V
g
is a neighborhood of p in V and f /g is
a regular function on V
g
. In this way, we can think of elements of O
V,p
as regular
functions dened in a neighborhood of p.
This idea extends to the abstract case. Given a point p of a variety X and
neighborhoods U
1
,U
2
of p, regular functions f
i
: U
i
Care equivalent at p, written
f
1
f
2
, if there is a neighbhorhood p U U
1
U
2
such that f
1[
U
= f
2[
U
.
Denition 3.0.9. Let p be a point of a variety X. Then
O
X,p
=f : U C [ U is a neighborhood of p in X/
is the local ring of X at p.
100 Chapter 3. Normal Toric Varieties
Every O
X,p
has a well-dened value (p). It is not difcult to see that
O
X,p
is a local ring with unique maximal ideal
m
X,p
= O
X,p
[ (p) = 0.
The local ring O
X,p
can also be dened as the direct limit
O
X,p
= lim

pU
O
X
(U)
over all neighborhoods of p in X (see Denition 6.0.1).
When X is irreducible, we can also dene the eld of rational functions C(X).
A rational function on X is a regular function f : U C dened on a nonempty
Zariski open set U X, and two rational functions on X are equivalent if they
agree on a nonempty Zariski open subset. In Exercise 3.0.4 you will show that this
relation is an equivalence relation and that the set of equivalence classes is a eld,
called the function eld of X, denoted C(X).
Normal Varieties. We return to the notion of normality introduced in Chapter 1.
Denition 3.0.10. A variety X is called normal if it is irreducible and the local
rings O
X,p
are normal for all p X.
At rst glance, this looks different from the denition given for afne varieties
in Denition 1.0.3. In fact, the two notions are equivalent in the afne case.
Proposition 3.0.11. Let V be an irreducible afne variety. Then C[V] is normal if
and only if the local rings O
V,p
are normal for all p V.
Proof. If O
V,p
is normal for all p, then (3.0.2) shows that C[V] is an intersection
of normal domains, all of which have the same eld of fractions. Since such an
intersection is normal by Exercise 1.0.7, it follows that C[V] is normal.
For the converse, suppose that C[V] is normal and let C(V) satisfy

k
+a
1

k1
+ +a
k
= 0, a
i
O
V,p
.
Write a
i
= g
i
/f
i
with g
i
, f
i
C[V] and f
i
(p) = 0. The product f = f
1
f
k
has
the properties that a
i
C[V]
f
and f (p) = 0. The localization C[V]
f
is normal by
Exercise 1.0.7 and is contained in O
V,p
since f (p) = 0. Hence C[V]
f
O
V,p
.
This completes the proof.
Here is a consequence of Proposition 3.0.11 and Denition 3.0.10.
Proposition 3.0.12. Let X be an irreducible variety with a cover consisting of
afne open sets V

. Then X is normal if and only if each V

is normal.
3.0. Background: Abstract Varieties 101
Smooth Varieties. For an afne variety V, the denition of a smooth point p V
(Denition 1.0.7) used T
p
(V), the Zariski tangent space of V at p, and dim
p
V,
the maximum dimension of an irreducible component of V containing p. You will
show in Exercise 3.0.2 that T
p
(X) and dim
p
X are well-dened for a point p X of
a general variety.
Denition 3.0.13. Let X be a variety. A point p X is smooth if dim T
p
(X) =
dim
p
X, and X is smooth if every point of X is smooth.
Products of Varieties. As another example of abstract varieties and gluing, we
indicate why the product X
1
X
2
of varieties X
1
and X
2
also has the structure of
a variety. In 1.0 we constructed the product of afne varieties. From here, it is
relatively routine to see that if X
1
is obtained by gluing together afne varieties U

and X
2
is obtained by gluing together afnes U

, then X
1
X
2
is obtained by gluing
together the U

in the corresponding fashion. Furthermore, X


1
X
2
has the
correct universal mapping property. Namely, given a diagram
W
1

X
1
X
2

X
1
X
2
where
i
: W X
i
are morphisms, there is a unique morphism : W X
1
X
2
(the dotted arrow) that makes the diagram commute.
Example 3.0.14. Let us construct the product P
1
C
2
. Write P
1
=V
0
V
1
where
V
0
= Spec(C[u]) and V
1
= Spec(C[v]), with the gluing given by
C[v]
v
C[u]
u
, v 1/u.
Then P
1
C
2
is constructed from
U
0
C
2
= Spec(C[u]
C
C[x, y]) C
3
U
1
C
2
= Spec(C[v]
C
C[x, y]) C
3
,
with gluing given by
(U
0
C
2
)
u
(U
1
C
2
)
v
corresponding to the obvious isomorphism of coordinate rings.
Separated Varieties. From the point of view of the classical topology, arbitrary
gluings can lead to varieties with some strange properties.
Example 3.0.15. In Example 3.0.14 we saw how to construct P
1
from afne va-
rieties V
0
= Spec(C[u]) C and V
1
= Spec(C[v]) C with the gluing given by
v 1/u on open sets C

(V
0
)
u
V
0
and C

(V
1
)
v
V
1
. This expresses P
1
as
102 Chapter 3. Normal Toric Varieties
consisting of C

plus two additional points. But now consider the abstract variety
arising from the gluing map
(V
0
)
u
(V
1
)
v
that corresponds to the map of C-algebras dened by v u. As before, the glued
variety X consists of C

together with two additional points. However here we


have a morphism : X C whose ber
1
(a) over a C

contains one point,


but whose ber over 0 consists of two points, p
1
corresponding to 0 V
0
and p
2
corresponding to 0 V
1
. If U
1
,U
2
are classical open sets in X with p
1
U
1
and
p
2
U
2
, then U
1
U
2
=. So the classical topology on X is not Hausdorff.
Since varieties are rarely Hausdorff in the Zariski topology (Exercise 3.0.5), we
need a different way to think about Example 3.0.15. Consider the product X X
and the diagonal mapping : X X X dened by (p) = (p, p) for p X. For
X from Example 3.0.15, there is a morphism X X C whose ber over over 0
consists of the four points (p
i
, p
j
). Any Zariski closed subset of X X containing
one of these four points must contain all of them. The image of the diagonal
mapping contains (p
1
, p
1
) and (p
2
, p
2
), but not the other two, so the diagonal is
not Zariski closed. This example motivates the following denition.
Denition 3.0.16. We say a variety X is separated if the image of the diagonal
map : X X X is Zariski closed in X X.
For instance, C
n
is separated because the image of the diagonal in C
n
C
n
=
Spec(C[x
1
, . . . , x
n
, y
1
, . . . , y
n
]) is the afne variety V(x
1
y
1
, . . . , x
n
y
n
). Similarly,
any afne variety is separated.
The connection between failure of separatedness and failure of the Hausdorff
property in the classical topology seen in Example 3.0.15 is a general phenomenon.
Theorem 3.0.17. A variety is separated if and only if it is Hausdorff in the classical
topology.
Here are some additional properties of separated varieties (Exercise 3.0.6).
Proposition 3.0.18. Let X be a separated variety. Then:
(a) If f , g : Y X are morphisms, then y Y [ f (y) =g(y) is Zariski closed in Y.
(b) If U,V are afne open subsets of X, then U V is also afne.
The requirement that X be separated is often included in the denition of an
abstract variety. When this is done, what we have called a variety is sometimes
called a pre-variety.
Fiber Products. Finally in this section, we will dene ber products of varieties,
a construction required for the discussion of proper morphisms in 3.4. First, if we
3.0. Background: Abstract Varieties 103
have mappings of sets f : X S and g : Y S, then the ber product X
S
Y is
dened to be
(3.0.5) X
S
Y =(x, y) X Y [ f (x) = g(y).
The ber product construction gives a very exible language for describing or-
dinary products, intersections of subsets, bers of mappings, the set where two
mappings agree, and so forth:
If S is a point, then X
S
Y is the ordinary product X Y.
If X,Y are subsets of S and f , g are the inclusions, then X
S
Y X Y.
If Y =s S, then X
S
Y f
1
(s).
The third property is the reason for the name. All are easy exercises that we leave
to the reader.
In analogy with the universal mapping property of the product discussed above,
the ber product has the following universal property. Whenever we have map-
pings
1
: W X and
2
: W Y such that f
1
= g
2
, there is a unique
: W X
S
Y that makes the following diagram commute:
W
1

X
S
Y

X
f

Y
g

S.
Equation (3.0.5) denes X
S
Y as a set. To prove that X
S
Y is a variety, we
assume for simplicity that S is separated. Then f : X S and g : Y S give a
morphism ( f , g) : X Y SS, and one easily checks that
X
S
Y = ( f , g)
1
((S)),
where (S) S S is the diagonal. This is closed in S S since S is separated,
and it follows that X
S
Y is closed in X Y and hence has a natural structure as
a variety. From here, it is straightforward to show that X
S
Y has the desired uni-
versal mapping property. Proving that X
S
Y is a variety when S is not separated
takes more work and will not be discussed here.
In the afne case, we can also describe the coordinate ring of X
S
Y. Let
X = Spec(R
1
), Y = Spec(R
2
), and S = Spec(R). The morphisms f , g correspond
to ring homomorphisms f

: R R
1
, g

: R R
2
. Hence both R
1
, R
2
have the
structure of R-modules, and we have the tensor product R
1

R
R
2
. This is also a
nitely generated C-algebra, though it may have nilpotents (Exercise 3.0.9). To
get a coordinate ring, we need to take the quotient by the ideal N of all nilpotents.
104 Chapter 3. Normal Toric Varieties
Then one can prove that
X
S
Y = Spec(R
1

R
R
2
/N).
We can avoid worrying about nilpotents by constructing X
S
Y as the afne scheme
Spec(R
1

R
R
2
). Interested readers can learn about the construction of ber prod-
ucts as schemes in [90, I.3.1] and [131, pp. 8789].
Exercises for 3.0.
3.0.1. Let V = Spec(R) be an afne variety.
(a) Show that every ideal I R can be written in the form I = ' f
1
, . . . , f
s
`, where f
i
R.
(This is the Hilbert basis theorem in R.)
(b) Let W V be a subvariety. Show that the complement of W in V can be written as a
union of a nite collection of open afne sets of the formV
f
.
(c) Deduce that every open cover of V (in the Zariski topology) has a nite subcover.
(This says that afne varieties are quasicompact in the Zariski topology.)
3.0.2. As in the afne case, we want to say a variety X is smooth at p if dim T
p
(X) =
dim
p
X. In this exercise, you will show that this is a well-dened notion.
(a) Show that if p X is in the intersection of two afne open sets V

, then the
Zariski tangent spaces T
V,p
and T
V

,p
are isomorphic as vector spaces over C.
(b) Show that dim
p
X is a well-dened integer.
(c) Deduce that the proposed notion of smoothness at p is well-dened.
3.0.3. This exercise explores some properties of the morphisms dened in Denition 3.0.3.
(a) Prove the claim made in Example 3.0.4. Hint: Take a point p V
1
and dene m
p
=
f R
1
[ f (p) = 0. Then describe (

)
1
(m
p
) in terms of (p).
(b) Prove the properties of morphisms listed on page 96.
3.0.4. Let X be an irreducible abstract variety.
(a) Let f , g be rational functions on X. Show that f g if f [
U
= g[
U
for some nonempty
open set U X is an equivalence relation.
(b) Show that the set of equivalence classes of the relation in part (a) is a eld.
(c) Show that if U X is a nonempty open subset of X, then C(U) C(X).
3.0.5. Show that a variety is Hausdorff in the Zariski topology if and only if it consists of
nitely many points.
3.0.6. Consider Proposition 3.0.18.
(a) Prove part (a) of the proposition. Hint: Show rst that if F : Y X X is dened by
F(y) = ( f (y), g(y)), then Z = F
1
((X)).
(b) Prove part (b) of the proposition. Hint: Show rst that U V can be identied with
(X) (U V) X X.
3.0.7. Let V = Spec(R) be an afne variety. The diagonal mapping : V V V cor-
responds to a C-algebra homomorphism R
C
R R. Which one? Hint: Consider the
universal mapping property of V V.
3.1. Fans and Normal Toric Varieties 105
3.0.8. In this exercise, we will study an important variety in P
n1
C
n
, the blowup of
C
n
at the origin, denoted Bl
0
(C
n
). This generalizes Bl
0
(C
2
) from Example 3.0.8. Write
the homogeneous coordinates on P
n1
as x
0
, . . . , x
n1
, and the afne coordinates on C
n
as
y
1
, . . . , y
n
. Let
(3.0.6) W = Bl
0
(C
n
) = V(x
i1
y
j
x
j1
y
i
[ 1 i < j n) P
n1
C
n
.
Let U
i1
, i = 1, . . . , n, be the standard afne opens in P
n1
:
U
i1
=P
n1
` V(x
i1
),
i = 1, . . . , n (note the slightly non-standard indexing). So the U
i1
C
n
form a cover of
P
n1
C
n
.
(a) Show that for each i = 1, . . . , n, W
i1
=W (U
i1
C
n
)
Spec

x
0
x
i1
, . . . ,
x
i2
x
i1
,
x
i
x
i1
, . . . ,
x
n1
x
i1
, y
i

using the equations (3.0.6) dening W.


(b) Give the gluing data for identifying the subsets W
i1
` V(x
j1
) and W
j1
` V(x
i1
).
3.0.9. Let V = V(y
2
x) C
2
and consider the morphism : V C given by projection
onto the x-axis. We will study the bers of .
(a) As noted in the text, the ber
1
(0) = (0, 0) can be represented as the ber
product 0
C
V. In terms of coordinate rings, we have 0 = Spec(C[x]/'x`),
C = Spec(C[x]) and V = Spec(C[x, y]/'y
2
x`. Prove that
C[x]/'x`
C[x]
C[x, y]/'y
2
x` C[y]/'y
2
`.
Thus, the coordinate rings C[x]/'x`, C[x] and C[x, y]/'y
2
x` lead to a tensor product
that has nilpotents and hence cannot be a coordinate ring.
(b) If a = 0 in C, then
1
(a) =(a,

a). Show that the analogous tensor product is


C[x]/'x a`
C[x]
C[x, y]/'y
2
x` C[y]/'y
2
a`
C[y]/'y

a` C[y]/'y +

a`.
This has no nilpotents and hence is the coordinate ring of
1
(a).
What happens in part (a) is that the two square roots coincide, so that we get only one
point with multiplicity 2. The multiplicity information is recorded in the afne scheme
Spec(C[y]/'y
2
`). This is an example of the power of schemes.
3.1. Fans and Normal Toric Varieties
In this section we construct the toric variety X

corresponding to a fan . We will


also relate the varieties X

to many of the examples encountered previously, and


we will see how properties of the fan correspond to properties such as smoothness
and compactness of X

.
106 Chapter 3. Normal Toric Varieties
The Toric Variety of a Fan. A toric variety continues to mean the same thing as in
Chapters 1 and 2, although we now allow abstract varieties as in 3.0.
Denition 3.1.1. A toric variety is an irreducible variety X containing a torus
T
N
(C

)
n
as a Zariski open subset such that the action of T
N
on itself extends to an
algebraic action of T
N
on X. (By algebraic action, we mean an action T
N
X X
given by a morphism.)
The other ingredient in this section is a fan in the vector space N
R
.
Denition 3.1.2. A fan in N
R
is a nite collection of cones N
R
such that:
(a) Every is a strongly convex rational polyhedral cone.
(b) For all , each face of is also in .
(c) For all
1
,
2
, the intersection
1

2
is a face of each (hence also in ).
Furthermore, if is a fan, then:
The support of is [[ =

N
R
.
(r) is the set of r-dimensional cones of .
We have already seen some examples of fans. Theorem 2.3.2 shows that the
normal fan
P
of a full dimensional lattice polytope P M
R
is a fan in the sense
of Denition 3.1.2. However, there exist fans that are not equal to the normal fan
of any lattice polytope. An example of such a fan will be given in Example 4.2.13.
We now show how the cones in any fan give the combinatorial data necessary
to glue a collection of afne toric varieties together to yield an abstract toric variety.
By Theorem 1.2.18, each cone in gives the afne toric variety
U

= Spec(C[S

]) = Spec(C[

M]).
Recall from Denition 1.2.5 that a face _ is given by =H
m
, where m

and H
m
= u N
R
[ 'm, u` = 0 is the hyperplane dened by m. In Chapter 1, we
proved two useful facts:
First, Proposition 1.3.16 used the equality
(3.1.1) S

= S

+Z(m)
to show that C[S

] is the localization C[S

m. Thus U

= (U

m when _.
Second, if =
1

2
, then Lemma 1.2.13 implies that
(3.1.2)
1
H
m
= =
2
H
m
,
for some m

1
(
2
)

M. This shows that


(3.1.3) U

1
(U

1
)

m =U

= (U

2
)

m U

2
.
The following proposition gives an additional property of the S

and their
semigroup rings that we will need.
3.1. Fans and Normal Toric Varieties 107
Proposition 3.1.3. If
1
,
2
and =
1

2
, then
S

= S

1
+S

2
.
Proof. The inclusion S

1
+S

2
S

follows directly from the general fact that

1
+

2
= (
1

2
)

. For the reverse inclusion, take p S

and assume
that m

1
(
2
)

M satises (3.1.2). Then (3.1.1) applied to


1
gives p =
q+(m) for some q S

1
and N. But m

2
implies m S

2
, so that
p S

1
+S

2
.
This result is sometimes called the separation lemma and is a key ingredient in
showing that the toric varieties X

are separated in the sense of Denition 3.0.16.


Example 3.1.4. Let
1
= Cone(e
1
+e
2
, e
2
) (as in Exercise 1.2.11), and let
2
=
Cone(e
1
, e
1
+e
2
) in N
R
= R
2
. Then =
1

2
= Cone(e
1
+e
2
). We show the
dual cones

1
= Cone(e
1
, e
1
+e
2
),

2
= Cone(e
1
e
2
, e
2
), and

1
+

2
in Figure 1.

Figure 1. The cones 1, 2, and their duals


The dark shaded region on the right is

2
. Note =
1
H
m
=
2
H
m
,
where m = e
1
+e
2

1
and m = e
1
e
2

2
. Since S

is the set of all sums


m+m

with m

1
M and m

2
M, we see that S

= S

1
+S

2
.
Now consider the collection of afne toric varieties U

= Spec(C[S

]), where
runs over all cones in a fan . Let
1
and
2
be any two of these cones and let
=
1

2
. By (3.1.3), we have an isomorphism
g

2
,
1
: (U

1
)

m (U

2
)

m
which is the identity on U

. By Exercise 3.1.1, the compatibility conditions as in


3.0 for gluing the afne varieties U

along the subvarieties (U

m are satised.
Hence we obtain an abstract variety X

associated to the fan .


Theorem 3.1.5. Let be a fan in N
R
. The variety X

is a normal separated toric


variety.
108 Chapter 3. Normal Toric Varieties
Proof. Since each cone in is strongly convex, 0 N is a face of all
. Hence we have T
N
= Spec(C[M]) (C

)
n
U

for all . These tori are all


identied by the gluing, so we have T
N
X

. We know from Chapter 1 that each


U

has an action of T
N
. The gluing isomorphism g

2
,
1
reduces to the identity
mapping on C[S

2
]. Hence the actions are compatible on the intersections of
every pair of sets in the open afne cover, and patch together to give an algebraic
action of T
N
on X

.
The variety X

is irreducible because all of the U

are irreducible afne toric


varieties containing the torus T
N
. Furthermore, U

is a normal afne variety by


Theorem 1.3.5. Hence the variety X

is normal by Proposition 3.0.12.


To see that X

is separated it sufces to show that for each pair of cones


1
,
2
in , the image of the diagonal map
: U

1
U

2
, =
1

2
is Zariski closed (Exercise 3.1.2). But comes from the C-algebra homomor-
phism

: C[S

1
]
C
C[S

2
] C[S

]
dened by
m

m+n
. By Proposition 3.1.3,

is surjective, so that
C[S

] (C[S

1
]
C
C[S

2
])/ker(

).
Hence the image of is a Zariski closed subset of U

1
U

2
.
Toric varieties were originally known as torus embeddings, and the variety X

would be written T
N
emb() in older references such as [218]. Other commonly
used notations are X(), or X(), if the fan is denoted by . When we want to
emphasize the dependence on the lattice N, we will write X

as X
,N
.
Many of the toric varieties encountered in Chapters 1 and 2 come from fans.
For example, Theorem 1.3.5 implies that a normal afne toric variety comes from
a fan consisting of a single cone together with all of its faces. Furthermore, the
projective toric variety associated to a lattice polytope in Chapter 2 comes from a
fan. Here is the precise result.
Proposition 3.1.6. Let P M
R
be a full dimensional lattice polytope. Then the
projective toric variety X
P
X

P
, where
P
is the normal fan of P.
Proof. When P is very ample, this follows immediately from the description of the
intersections of the afne open pieces of X
P
in Proposition 2.3.13 and the denition
of the normal fan
P
. The general case follows since the normal fans of P and kP
are the same for all positive integers k.
In general, every separated normal toric variety comes from a fan. This is a
consequence of a theorem of Sumihiro from [265].
3.1. Fans and Normal Toric Varieties 109
Theorem 3.1.7 (Sumihiro). Let the torus T
N
act on a normal separated variety X.
Then every point p X has a T
N
-invariant afne open neighborhood.
Corollary 3.1.8. Let X be a normal separated toric variety with torus T
N
. Then
there exists a fan in N
R
such that X X

.
Proof. The proof will be sketched in Exercise 3.2.11 after we have developed the
properties of T
N
-orbits on toric varieties.
Examples. We now turn to some concrete examples. Many of these are toric vari-
eties already encountered in previous chapters.
Example 3.1.9. Consider the fan in N
R
= R
2
in Figure 2, where N = Z
2
has
standard basis e
1
, e
2
. This is the normal fan of the simplex
2
as in Example 2.3.4.
Here we show all points in the cones inside a rectangular viewing box (all gures
of fans in the plane in this chapter will be drawn using the same convention.)

1
Figure 2. The fan for P
2
From the discussion in Chapter 2, we expect X

P
2
, and we will show
this in detail. The fan has three 2-dimensional cones
0
= Cone(e
1
, e
2
),
1
=
Cone(e
1
e
2
, e
2
), and
2
= Cone(e
1
, e
1
e
2
), together with the three rays

i j
=
i

j
for i = j, and the origin. The toric variety X

is covered by the
afne opens
U

0
= Spec(C[S

0
]) Spec(C[x, y])
U

1
= Spec(C[S

1
]) Spec(C[x
1
, x
1
y])
U

2
= Spec(C[S

2
]) Spec(C[xy
1
, y
1
]).
Moreover, by Proposition 3.1.3, the gluing data on the coordinate rings is given by
g

10
: C[x, y]
x
C[x
1
, x
1
y]
x
1
g

20
: C[x, y]
y
C[xy
1
, y
1
]
y
1
g

21
: C[x
1
, x
1
y]
x
1
y
C[xy
1
, y
1
]
xy
1 .
110 Chapter 3. Normal Toric Varieties
It is easy to see that if we use the usual homogeneous coordinates (x
0
, x
1
, x
2
) on P
2
,
then x
x
1
x
0
and y
x
2
x
0
identies the standard afne open U
i
P
2
with U

i
X

.
Hence we have recovered P
2
as the toric variety X

.
Example 3.1.10. Generalizing Example 3.1.9, let N
R
= R
n
, where N = Z
n
has
standard basis e
1
, . . . , e
n
. Set
e
0
=e
1
e
2
e
n
and let be the fan in N
R
consisting of the cones generated by all proper subsets
of e
0
, . . . , e
n
. This is the normal fan of the n-simplex
n
, and X

P
n
by Ex-
ample 2.3.15 and Exercise 2.3.6. You will check the details to verify that this gives
the usual afne open cover of P
n
in Exercise 3.1.3.
Example 3.1.11. We classify all 1-dimensional normal toric varieties as follows.
We may assume N =Z and N
R
=R. The only cones are the intervals
0
= [0, )
and
1
= (, 0] and the trivial cone = 0. It follows that there are only four
possible fans, which gives the following list of toric varieties:
, which gives C

0
, and
1
, , both of which give C

0
,
1
, , which gives P
1
.
Here is a picture of the fan for P
1
:
s
0
1

0
This is the fan of Example 3.1.10 when n = 1.
Example 3.1.12. By Example 2.4.8, P
n
P
m
is the toric variety of the polytope

m
. The normal fan of
n

m
is the product of the normal fans of each
factor (Proposition 2.4.9). These normal fans are described in Example 3.1.10. It
follows that the product fan gives X

P
n
P
m
.
When n = m = 1, we obtain the fan R
2
N
R
pictured in Figure 3 on the
next page. Here, we can use an elementary gluing argument to show that this fan
gives P
1
P
1
. Label the 2-dimensional cones
i j
=
i

j
as above. Then
Spec(C[S

00
]) C[x, y]
Spec(C[S

10
]) C[x
1
, y]
Spec(C[S

11
]) C[x
1
, y
1
]
Spec(C[S

01
]) C[x, y
1
].
We see that if U
0
and U
1
are the standard afne open sets in P
1
, then U

i j
U
i
U
j
and it is easy to check that the gluing makes X

P
1
P
1
.
3.1. Fans and Normal Toric Varieties 111

00

10

11

01
Figure 3. A fan with X P
1
P
1
Example 3.1.13. Let N = N
1
N
2
, with N
1
= Z
n
and N
2
= Z
m
. Let
1
in (N
1
)
R
be the fan giving P
n
, but let
2
be the fan consisting of the cone Cone(e
1
, . . . , e
m
)
together with all its faces. Then =
1

2
is a fan in N
R
and the the corre-
sponding toric variety is X

P
n
C
m
. The case P
1
C
2
was studied in Exam-
ple 3.0.14.
Examples 3.1.12 and 3.1.13 are special cases of the following general con-
struction, whose proof will be left to the reader (Exercise 3.1.4).
Proposition 3.1.14. Suppose we have fans
1
in (N
1
)
R
and
2
in (N
2
)
R
. Then

2
=
1

2
[
i

i

is a fan in (N
1
)
R
(N
2
)
R
= (N
1
N
2
)
R
and
X

2
X

1
X

2
.
Example 3.1.15. The two cones
1
and
2
in N
R
= R
2
from Example 3.1.4 (see
Figure 1), together with their faces, form a fan . By comparing the descriptions
of the coordinate rings of V

i
given there with what we did in Example 3.0.8, it is
easy to check that X

W, where W P
1
C
2
is the blowup of C
2
at the origin,
dened as W = V(x
0
y x
1
x) (Exercise 3.1.5).
Generalizing this, let N = Z
n
with standard basis e
1
, . . . , e
n
and then set e
0
=
e
1
+ +e
n
. Let be the fan in N
R
consisting of the cones generated by all subsets
of e
0
, . . . , e
n
not containing e
1
, . . . , e
n
. Then the toric variety X

is isomorphic
to the blowup of C
n
at the origin (Exercise 3.0.8).
Example 3.1.16. Let r N and consider the fan
r
in N
R
= R
2
consisting of the
four cones
i
shown in Figure 4 on the next page, together with all of their faces.
112 Chapter 3. Normal Toric Varieties

2
(1, r)
Figure 4. A fan r with Xr
Hr
The corresponding toric variety X
r
is covered by open afne subsets,
U

1
= Spec(C[x, y]) C
2
U

2
= Spec(C[x, y
1
]) C
2
U

3
= Spec(C[x
1
, x
r
y
1
]) C
2
U

4
= Spec(C[x
1
, x
r
y]) C
2
,
and glued according to (3.1.3). We call X
r
the Hirzebruch surface H
r
.
Example 2.3.16 constructed the rational normal scroll S
a,b
using the polygon
P
a,b
with b a 1. The normal fan of P
a,b
is the fan
ba
dened above, so that
as an abstract variety, S
a,b
H
ba
. Note also that H
0
P
1
P
1
.
The Hirzebruch surfaces H
r
will play an important role in the classication of
smooth projective toric surfaces given in Chapter 10.
Example 3.1.17. Let q
0
, . . . , q
n
Z
>0
satisfy gcd(q
0
, . . . , q
n
) = 1. Consider the
weighted projective space P(q
0
, . . . , q
n
) introduced in Chapter 2. Dene the lattice
N =Z
n+1
/Z (q
0
, . . . , q
n
) and let u
i
, i = 0, . . . , n, be the images in N of the standard
basis vectors in Z
n+1
, so the relation
q
0
u
0
+ +q
n
u
n
= 0
holds in N. Let be the fan made up of the cones generated by all the proper sub-
sets of u
0
, . . . , u
n
. When q
i
= 1 for all i, we obtain X

P
n
by Example 3.1.10.
And indeed, X

P(q
0
, . . . , q
n
) in general. This will be proved in Chapter 5 using
the toric generalization of homogeneous coordinates in P
n
.
Here, we will consider the special case P(1, 1, 2), where u
0
= u
1
2u
2
. The
fan in N
R
is pictured in Figure 5 on the next page, using the plane spanned by
u
1
, u
2
. This example is different from the ones we have seen so far. Consider
2
=
Cone(u
0
, u
1
) = Cone(u
1
2u
2
, u
1
). Then

2
=Cone(u
2
, 2u
1
u
2
) M, so the
situation is similar to the case studied in Example 1.2.22. Indeed, there is a change
3.1. Fans and Normal Toric Varieties 113

2
Figure 5. A fan with X P(1, 1, 2)
of coordinates dened by a matrix in GL(2, Z) that takes to the cone with d = 2
from that example. It follows that there is an isomorphism U

2
V(xz y
2
) C
3
(Exercise 3.1.6). This is the rational normal cone

C
2
, hence has a singular point
at the origin. The toric variety X

is singular because of the singular point in this


afne open subset.
In Example 2.4.6, we saw that the polytope P = Conv(0, 2e
1
, e
2
) R
2
gives
X
P
P(1, 1, 2) and that the normal fan
P
coincides with the fan shown above.
There is a dictionary between properties of X

and properties of that gener-


alizes Theorem 1.3.12 and Example 1.3.20. We begin with some terminology. The
rst two items parallel Denition 1.2.16.
Denition 3.1.18. Let N
R
be a fan.
(a) We say is smooth (or regular) if every cone in is smooth (or regular).
(b) We say is simplicial if every cone in is simplicial.
(c) We say is complete if its support [[ =

is all of N
R
.
Theorem 3.1.19. Let X

be the toric variety dened by a fan N


R
. Then:
(a) X

is a smooth variety if and only if the fan is smooth.


(b) X

is an orbifold (that is, X

has only nite quotient singularities) if and only


if the fan is simplicial.
(c) X

is compact in the classical topology if and only if is complete.


Proof. Part (a) follows from the corresponding statement for afne toric varieties,
Theorem 1.3.12, because smoothness is a local property (Denition 3.0.13). In
part (b), Example 1.3.20 gives one implication. The other implication will be
proved in Chapter 11. A proof of part (c) will be given in 3.4.
The blowup of C
2
at the origin (Example 3.1.15) is not compact, since the
support of the cones in the corresponding fan is not all of R
2
. The Hirzebruch
114 Chapter 3. Normal Toric Varieties
surfaces H
r
from Example 3.1.16 are smooth and compact because every cone in
the corresponding fan is smooth, and the union of the cones is R
2
. The variety
P(1, 1, 2) from Example 3.1.17 is compact but not smooth. It is an orbifold (it has
only nite quotient singularities) since the corresponding fan is simplicial.
Exercises for 3.1.
3.1.1. Let be a fan in N
R
. Show that the isomorphisms g
1,2
satisfy the compatibility
conditions from 0 for gluing the U

together to create X

.
3.1.2. Let X be a variety obtained by gluing afne open subsets V

along open subsets


V

by isomorphisms g

: V

. Show that X is separated when the image of


: V

dened by (p) = (p, g

(p)) is Zariski closed for all , .


3.1.3. Verify that if is the fan given in Example 3.1.10, then X

P
n
.
3.1.4. Prove Proposition 3.1.14.
3.1.5. Let N Z
n
, let e
1
, . . . , e
n
N be the standard basis and let e
0
= e
1
+ +e
n
. Let
be the set of cones generated by all subsets of e
0
, . . . , e
n
not containing e
1
, . . . , e
n
.
(a) Show that is a fan in N
R
.
(b) Construct the afne open subsets covering the corresponding toric variety X

, and
give the gluing isomorphisms.
(c) Show that X

is isomorphic to the blowup of C


n
at the origin, described earlier in
Exercise 3.0.8. Hint: The blowup is the subvariety of P
n1
C
n
given by W =
V(x
i
y
j
x
j
y
i
[ 1 i < j n). Cover W by afne open subsets W
i
= W
xi
and com-
pare those afnes with your answer to part (b).
3.1.6. In this exercise, you will verify the claims made in Example 3.1.17.
(a) Show that there is a matrix A GL(2, Z) dening a change of coordinates that takes
the cone in this example to the cone from Example 1.2.22, and nd the mapping that
takes

2
to the dual cone.
(b) Show that Spec(C[S
2
]) V(xz y
2
) C
3
.
3.1.7. In N
R
= R
2
, consider the fan with cones 0, Cone(e
1
), and Cone(e
1
). Show
that X

P
1
C

.
3.2. The Orbit-Cone Correspondence
In this section, we will study the orbits for the action of T
N
on the toric variety X

.
Our main result will show that there is a bijective correspondence between cones
in and T
N
-orbits in X

. The connection comes ultimately from looking at limit


points of the one-parameter subgroups of T
N
dened in 1.1.
A First Example. We introduce the key features of the correspondence between
orbits and cones by looking at a concrete example.
3.2. The Orbit-Cone Correspondence 115
Example 3.2.1. Consider P
2
X

for the fan from Figure 2 of 3.1. The torus


T
N
=(C

)
2
P
2
consists of points with homogeneous coordinates (1, s, t), s, t =0.
For each u = (a, b) N =Z
2
, we have the corresponding curve in P
2
:

u
(t) = (1, t
a
, t
b
).
We are abusing notation slightly; strictly speaking, the one-parameter subgroup
u
is a curve in (C

)
2
, but we view it as a curve in P
2
via the inclusion (C

)
2
P
2
.
We start by analyzing the limit of
u
(t) as t 0. The limit point in P
2
depends
on u = (a, b). It is easy to check that the pattern is as follows:
limit is (1,0,0)
limit is (0,0,1)
limit is (0,1,0)

limit is (0,1,1)

limit is (1,1,0)
limit is (1,0,1)

limit is (1,1,1)
a
b
Figure 6. limt0
u
(t) for u =(a, b) Z
2
For instance, suppose a, b >0 in Z. These points lie in the rst quadrant. Here,
it is obvious that lim
t0
(1, t
a
, t
b
) = (1, 0, 0). Next suppose that a = b < 0 in Z,
corresponding to points on the diagonal in the third quadrant. Note that
(1, t
a
, t
b
) = (1, t
a
, t
a
) (t
a
, 1, 1)
since we are using homogeneous coordinates in P
2
. Then a > 0 implies that
lim
t0
(t
a
, 1, 1) = (0, 1, 1). You will check the remaining cases in Exercise 3.2.1.
The regions of N described in Figure 6 correspond to cones of the fan . In
each case, the set of u giving one of the limit points equals N Relint(), where
Relint() is the relative interior of a cone . In other words, we have recovered
the structure of the fan by considering these limits!
Now we relate this to the T
N
-orbits in P
2
. By considering the description P
2

(C
3
`0)/C

, you will see in Exercise 3.2.1 that there are exactly seven T
N
-orbits
116 Chapter 3. Normal Toric Varieties
in P
2
:
O
1
=(x
0
, x
1
, x
2
) [ x
i
= 0 for all i (1, 1, 1)
O
2
=(x
0
, x
1
, x
2
) [ x
2
= 0, and x
0
, x
1
= 0 (1, 1, 0)
O
3
=(x
0
, x
1
, x
2
) [ x
1
= 0, and x
0
, x
2
= 0 (1, 0, 1)
O
4
=(x
0
, x
1
, x
2
) [ x
0
= 0, and x
1
, x
2
= 0 (0, 1, 1)
O
5
=(x
0
, x
1
, x
2
) [ x
1
= x
2
= 0, and x
0
= 0 =(1, 0, 0)
O
6
=(x
0
, x
1
, x
2
) [ x
0
= x
2
= 0, and x
1
= 0 =(0, 1, 0)
O
7
=(x
0
, x
1
, x
2
) [ x
0
= x
1
= 0, and x
2
= 0 =(0, 0, 1).
This list shows that each orbit contains a unique limit point. Hence we obtain a
correspondence between cones and orbits O by
corresponds to O lim
t0

u
(t) O for all u Relint().
We will soon see that these observations generalize to all toric varieties X

.
Points and Semigroup Homomorphisms. It will be convenient to use the intrinsic
description of the points of an afne toric variety U

given in Proposition 1.3.1.


We recall how this works and make some additional observations:
Points of U

are in bijective correspondence with semigroup homomorphisms


: S

C. Recall that S

M and U

= Spec(C[S

]).
For each cone we have a point of U

dened by
m S

1 m S

M
0 otherwise.
This is a semigroup homomorphism since

is a face of

. Thus, if
m, m

and m+m

, then m, m

. We denote this point


by

and call it the distinguished point corresponding to .


The point

is xed under the T


N
-action if and only if dim =dim N
R
(Corol-
lary 1.3.3).
If _ is a face, then

. This follows since

.
Limits of One-Parameter Subgroups. In Example 3.2.1, the limit points of one-
parameter subgroups are exactly the distinguished points for the cones in the fan
of P
2
(Exercise 3.2.1). We now show that this is true for all afne toric varieties.
Proposition 3.2.2. Let N
R
be a strongly convex rational polyhedral cone and
let u N. Then
u lim
t0

u
(t) exists in U

.
Moreover, if u Relint(), then lim
t0

u
(t) =

.
3.2. The Orbit-Cone Correspondence 117
Proof. Given u N, we have
lim
t0

u
(t) exists in U

lim
t0

m
(
u
(t)) exists in C for all m S

lim
t0
t
/m,u)
exists in C for all m S

'm, u` 0 for all m

M
u (

= ,
where the rst equivalence is proved in Exercise 3.2.2 and the other equivalences
are clear. This proves the rst assertion of the proposition.
In Exercise 3.2.2 you will also show that when u N, lim
t0

u
(t) is the
point corresponding to the semigroup homomorphism S

C dened by
m

M lim
t0
t
/m,u)
.
If u Relint(), then 'm, u` >0 for all mS

(Exercise 1.2.2), and 'm, u` =0


if m S

. Hence the limit point is precisely the distinguished point

.
Using this proposition, we can recover the fan from X

cone by cone as in
Example 3.2.1. This is also the key observation needed for the proof of Corol-
lary 3.1.8 from the previous section.
Let us apply Proposition 3.2.2 to a familiar example.
Example 3.2.3. Consider the afne toric variety V = V(xyzw) studied in a num-
ber of examples from Chapter 1. For instance, in Example 1.1.18, we showed that
V is the normal toric variety corresponding to a cone whose dual cone is
(3.2.1)

= Cone(e
1
, e
2
, e
3
, e
1
+e
2
e
3
),
and V = Spec(C[

M]).
In Example 1.1.18, we introduced the torus T = (C

)
3
of V as the image of
(3.2.2) (t
1
, t
2
, t
3
) (t
1
, t
2
, t
3
, t
1
t
2
t
1
3
).
Given u = (a, b, c) N =Z
3
, we have the one-parameter subgroup
(3.2.3)
u
(t) = (t
a
, t
b
, t
c
, t
a+bc
)
contained in V, and we proceed to examine limit points using Proposition 3.2.2.
Clearly, lim
t0

u
(t) exists in V if and only if a, b, c 0 and a +b c. These
conditions determine the cone N
R
given by
(3.2.4) = Cone(e
1
, e
2
, e
1
+e
3
, e
2
+e
3
).
One easily checks that (3.2.1) is the dual of this cone (Exercise 3.2.3). Note
also that u Relint() means a, b, c > 0 and a +b > c, in which case the limit
lim
t0

u
(t) = (0, 0, 0, 0), which is the distinguished point

.
118 Chapter 3. Normal Toric Varieties
The Torus Orbits. Now we turn to the T
N
-orbits in X

. We saw above that each


cone has a distinguished point

. This gives the torus orbit


O() = T
N

.
In order to determine the structure of O(), we need the following lemma, which
you will prove in Exercise 3.2.4.
Lemma 3.2.4. Let be a strongly convex rational polyhedral cone in N
R
. Let N

be the sublattice of N spanned by the points in N, and let N() = N/N

.
(a) There is a perfect pairing
' , ` :

MN() Z,
induced by the dual pairing ' , ` : MN Z.
(b) The pairing of part (a) induces a natural isomorphism
Hom
Z
(

M, C

) T
N()
,
where T
N()
= N()
Z
C

is the torus associated to N().


To study O() U

, we recall howt T
N
acts on semigroup homomorphisms.
If p U

is represented by : S

C, then by Exercise 1.3.1, the point t p is


represented by the semigroup homomorphism
(3.2.5) t : m
m
(t)(m).
Lemma 3.2.5. Let be a strongly convex rational polyhedral cone in N
R
. Then
O() = : S

C [ (m) = 0 m

M
Hom
Z
(

M, C

) T
N()
,
where N() is the lattice dened in Lemma 3.2.4.
Proof. The set O

= : S

C [ (m) = 0 m

M contains

and is
invariant under the action of T
N
described in (3.2.5).
Next observe that

is the largest vector subspace of M


R
contained in

.
Hence

M is a subgroup of S

M. If O

, then restricting to
m S

M yields a group homomorphism :

M C

(Exer-
cise 3.2.5). Conversely, if :

M C

is a group homomorphism, we obtain


a semigroup homomorphism O

by dening
(m) =

(m) if m

M
0 otherwise.
It follows that O

Hom
Z
(

M, C

).
Now consider the exact sequence
(3.2.6) 0 N

N N() 0.
3.2. The Orbit-Cone Correspondence 119
Tensoring with C

and using Lemma 3.2.4, we obtain a surjection


T
N
= N
Z
C

T
N()
= N()
Z
C

Hom
Z
(

M, C

).
The bijections
T
N()
Hom
Z
(

M, C

) O

are compatible with the T


N
-action, so that T
N
acts transtively on O

. Then

implies that O

= T
N

= O(), as desired.
The Orbit-Cone Correspondence. Our next theorem is the major result of this
section. Recall that the face relation _ holds when is a face of .
Theorem 3.2.6 (Orbit-Cone Correspondence). Let X

be the toric variety of the


fan in N
R
. Then:
(a) There is a bijective correspondence
cones in T
N
-orbits in X

O() Hom
Z
(

M, C

).
(b) Let n = dim N
R
. For each cone , dim O() = ndim.
(c) The afne open subset U

is the union of orbits


U

O().
(d) _ if and only if O() O(), and
O() =

O(),
where O() denotes the closure in both the classical and Zariski topologies.
For instance, Example 3.2.1 tells us that for P
2
, there are three types of cones
and torus orbits:
The trivial cone =(0, 0) corresponds to the orbit O() = T
N
P
2
, which
satises dim O() = 2 = 2 dim. This is a face of all the other cones in
, and hence all the other orbits are contained in the closure of this one by
part (d). Note also that U

= O() (C

)
2
by part (c), since there are no
cones properly contained in .
The three 1-dimensional cones give the torus orbits of dimension 1. Each is
isomorphic to C

. The closures of these orbits are the coordinate axes V(x


i
) in
P
2
, each a copy of P
1
. Note that each is contained in two maximal cones.
The three maximal cones
i
in the fan correspond to the three xed points
(1, 0, 0), (0, 1, 0), (0, 0, 1) of the torus action on P
2
. There are two of these in
the closure of each of the 1-dimensional torus orbits.
120 Chapter 3. Normal Toric Varieties
Proof of Theorem 3.2.6. Let O be a T
N
-orbit in X

. Since X

is covered by the
T
N
-invariant afne open subsets U

and U

1
U

2
=U

2
, there is a unique
minimal cone with OU

. We claim that O = O(). Note that part (a) will


follow immediately once we prove this claim.
To prove the claim, let O and consider those m S

satisfying (m) = 0.
In Exercise 3.2.6, you will show that these ms lie on a face of

. But faces of

are all of the form

for some face _ by Proposition 1.2.10. In other


words, there is a face _ such that
m S

[ (m) = 0 =

M.
This easily implies U

(Exercise 3.2.6), and then = by the minimality of .


Hence m S

[ (m) = 0 =

M, and then O() by Lemma 3.2.5. This


implies O = O() since two orbits are either equal or disjoint.
Part (b) follows from Lemma 3.2.5 and (3.2.6).
Next consider part (c). We know that U

is a union of orbits. If is a face of ,


then O() U

implies that O() is an orbit contained in U

. Furthermore,
the analysis of part (a) easily implies that any orbit contained in U

must equal
O() for some face _.
We now turn to part (d). We begin with the closure of O() in the classical
topology, which we denote O(). This is invariant under T
N
(Exercise 3.2.6) and
hence is a union of orbits. Suppose that O() O(). Then O() U

, since
otherwise O() U

= , which would imply O() U

= since U

is open in
the classical topology. Once we have O() U

, it follows that _ by part (c).


Conversely, assume _ . To prove that O() O(), it sufces to show that
O() O() = . We will do this by using limits of one-parameter subgroups as
in Proposition 3.2.2.
Let

be the semigroup homomorphism corresponding to the distinguished


point of U

, so

(m) = 1 if m

M, and 0 otherwise. Let u Relint(), and


for t C

dene (t) =
u
(t)

. As a semigroup homomorphism, (t) is


m
m
(
u
(t))

(m) = t
/m,u)

(m).
Note that (t) O() for all t C

since the orbit of

is O(). Now let t 0.


Since u Relint(), 'm, u` > 0 if m

, and = 0 if m

. It follows
that (0) = lim
t0
(t) exists as a point in U

by Proposition 3.2.2, and represents


a point in O(). But it is also in the closure of O() by construction, so that
O() O() =. This establishes the rst assertion of (d), and
O() =

O()
follows immediately for the classical topology.
3.2. The Orbit-Cone Correspondence 121
It remains to show that this set is also the Zariski closure. If we intersect O()
with an afne open subset U

, parts (c) and (d) imply that


O() U

O().
In Exercise 3.2.6, you will show that this is the subvariety V(I) U

for the ideal


(3.2.7) I ='
m
[ m

M` C[(

M] = S

.
This easily implies that the classical closure O() is a subvariety of X

and hence
is the Zariski closure of O().
Orbit Closures as Toric Varieties. In the example of P
2
, the orbit closures O()
also have the structure of toric varieties. This holds in general. We use the notation
V() = O().
By part (d) of Theorem 3.2.6, we have _ if and only if O() V(), and
V() =

O().
The torus O() = T
N()
is an open subset of V(), where N() is dened in
Lemma 3.2.4. We will show that V() is a normal toric variety by constructing
its fan. For each cone containing , let be the image cone in N()
R
under
the quotient map
N
R
N()
R
in (3.2.6). Then
(3.2.8) Star() = N()
R
[ _
is a fan in N()
R
(Exercise 3.2.7).
Proposition 3.2.7. For any , the orbit closure V() = O() is isomorphic to
the toric variety X
Star()
.
Proof. This follows from parts (a) and (d) of Theorem 3.2.6 (Exercise 3.2.7).
Example 3.2.8. Consider the fan in N
R
=R
3
shown in Figure 7 on the next page.
The support of is the cone in Figure 2 of Chapter 1, and is obtained from this
cone by adding a new 1-dimensional cone in the center and subdividing. The
orbit O() has dimension 2 by Theorem 3.2.6. By Proposition 3.2.7, the orbit
closure V() is constructed from the cones of containing and then collapsing
to a point in N()
R
= (N/N

)
R
R
2
. This clearly gives the fan for P
1
P
1
, so
that V() P
1
P
1
.
122 Chapter 3. Normal Toric Varieties
z
y
x

Figure 7. The fan and its 1-dimensional cone in Example 3.2.8


A nice example of orbit closures comes from the toric variety X
P
of a full
dimensional lattice polytope P M
R
. Here, we use the normal fan
P
of P, which
by Theorem 2.3.2 consists of cones
(3.2.9)
Q
= Cone(u
F
[ F is a facet of P containing Q)
for each face Q _P. Recall that u
F
is the facet normal of F.
The basic idea is that the orbit closure of V(
Q
) is the toric variety of the lattice
polytope Q. Since Q need not be full dimensional in M
R
, we need to be careful.
The idea is to translate P by a vertex of Q so that the origin is a vertex of Q. This
affects neither
P
nor X
P
, but Q is now full dimensional in Span(Q) and is a lattice
polytope relative to Span(Q) M. This gives the toric variety X
Q
, which is easily
seen to be independent of how we translate to the origin. Here is our result.
Proposition 3.2.9. For each face Q _P, we have V(
Q
) X
Q
.
Proof. We sketch the proof and leave the details to reader (Exercise 3.2.8). Let
P =m M
R
[ 'm, u
F
` a
F
for all facets F P
be the facet presentation of P. The facets of P containing Q also contain the origin,
so that a
F
= 0 for these facets. This implies that

Q
= Span(Q),
and then N(
Q
) is dual to Span(Q) M. Note also that N(
Q
)
R
= N
R
/Span(
Q
).
To keep track of which polytope we are using, we will write the cone (3.2.9)
associated to a face Q _P as
Q,P
. Then X
P
and X
Q
are given by the normal fans

P
=
Q

,P
N
R
[ Q

Q
=
Q

,Q
N(
Q,P
)
R
[ Q

Q.
3.2. The Orbit-Cone Correspondence 123
By Proposition 3.2.7, the toric variety V(
Q
) =V(
Q,P
) is determined by the fan
Star(
Q,P
) = [
Q,P

P

=
Q

,P
[
Q,P

,P

P
=
Q

,P
[ Q

_Q.
Then the proposition follows once one proves that
Q

,P
=
Q

,Q
.
Final Comments. The technique of using limit points of one-parameter subgroups
to study a group action is also a major tool in Geometric Invariant Theory as in
[209], where the main problem is to construct varieties (or possibly more general
objects) representing orbit spaces for the actions of algebraic groups on varieties.
We will apply ideas from group actions and orbit spaces to the study of toric vari-
eties in Chapters 5 and 14.
We also note the observation made in part (d) of Theorem 3.2.6 that torus orbits
have the same closure in the classical and Zariski topologies. For arbitrary subsets
of a variety, these closures may differ. A torus orbit is an example of a constructible
subset, and we will see in 3.4 that constructible subsets have the same classical
and Zariski closures since we are working over C.
Exercises for 3.2.
3.2.1. In this exercise, you will verify the claims made in Example 3.2.1 and the following
discussion.
(a) Show that the remaining limits of one-parameter subgroups P
2
are as claimed in the
example.
(b) Show that the (C

)
2
-orbits in P
2
are as claimed in the example.
(c) Show that the limit point equals the distinguished point

of the corresponding cone


in each case.
3.2.2. Let N
R
be a strongly convex rational polyhedral cone. This exercise will con-
sider lim
t0
f (t), where f : C

T
N
is an arbitrary function.
(a) Prove that lim
t0
f (t) exists in U

if and only if lim


t0

m
( f (t)) exists in C for all
m S

. Hint: Consider a nite set of characters A such that S

=NA.
(b) When lim
t0
f (t) exists in U

, prove that the limit is given by the semigroup homo-


morphism that maps m S

to lim
t0

m
( f (t)).
3.2.3. Consider the situation of Example 3.2.3.
(a) Show that the cones in (3.2.1) and (3.2.4) are dual.
(b) Identify the limits of all one-parameter subgroups in this example, and describe the
Orbit-Cone Correspondence in this case.
(c) Show that the matrix
A =

1 1 1
1 0 0
1 0 1

denes an automorphismof N Z
3
and the corresponding linear map on N
R
maps the
cone

to .
124 Chapter 3. Normal Toric Varieties
(d) Deduce that the afne toric varieties U

and U

are isomorphic. Hint: Use Proposi-


tion 1.3.15.
3.2.4. Prove Lemma 3.2.4.
3.2.5. Let O

be as dened in the proof of Lemma 3.2.5. In this exercise, you will complete
the proof that O

is a T
N
-orbit in U

.
(a) Show that if O

, then :

M C

is a group homomorphism.
(b) Deduce that O

has the structure of a group.


(c) Verify carefully that we have an isomorphism of groups O

Hom
Z
(

M, C

).
3.2.6. This exercise is concerned with the proof of Theorem 3.2.6.
(a) Let : S

C be a semigroup homomorphism giving a point of U

. Prove that
m S

[ (m) = 0 = M for some face _

.
(b) Show O() is invariant under the action of T
N
.
(c) Prove that O() U

is the variety of the ideal I dened in (3.2.7).


3.2.7. Let be a cone in a fan , and let Star() be as dened in (3.2.8).
(a) Show that Star() is a fan in N()
R
.
(b) Prove Proposition 3.2.7.
3.2.8. Supply the details omitted in the proof of Proposition 3.2.9.
3.2.9. Consider the action of T
N
on the afne toric variety U

. Use parts (c) and (d) of


Theorem 3.2.6 to show that O() is the unique closed orbit of T
N
acting on U

.
3.2.10. In Proposition 1.3.16, we saw that if is a face of the strongly convex rational poly-
hedral cone in N
R
then U

= Spec(C[S

]) is an afne open subset of U

= Spec(C[S

]).
In this exercise, you will prove the converse, i.e., that if and the induced map of
afne toric varieties : U

is an open immersion, then _, i.e., is a face of .


(a) Let u, u

N , and assume u +u

. Show that
lim
t0

u
(t) lim
t0

(t) U

.
(b) Show that lim
t0

u
(t) and lim
t0

(t) are each in U

. Hint: Use the description of


points as semigroup homomorphisms.
(c) Deduce that u, u

, so is a face of .
3.2.11. In this exercise, you will use Proposition 3.2.2 and Theorem3.2.6 to deduce Corol-
lary 3.1.8 from Theorem 3.1.7.
(a) By Theorem 3.1.7, and the results of Chapter 1, a separated toric variety has an open
cover consisting of afne toric varieties U
i
=U
i
for some collection of cones
i
. Show
that for all i, j, U
i
U
j
is also afne. Hint: Use the fact that X is separated.
(b) Show that U
i
U
j
is the afne toric variety corresponding to the cone =
i

j
.
Hint: Exercise 3.2.2 will be useful.
(c) If =
i

j
, then show that is a face of both
i
and
j
. Hint: Use Exercise 3.2.10.
(d) Deduce that X X

for the fan consisting of the


i
and all their faces.
3.3. Toric Morphisms 125
3.3. Toric Morphisms
Recall from 3.0 that if X and Y are varieties with afne open covers X =

and Y =

, then a morphism : X Y is a Zariski-continuous mapping such


that the restrictions
[
U
1
(U

)
: U

1
(U

) U

are morphisms in the sense of Denition 3.0.3 for all , .


In 1.3 we dened toric morphisms between afne toric varieties and studied
their properties. When applied to arbitrary normal toric varieties, these results yield
a class of morphisms whose construction comes directly from the combinatorics of
the associated fans. The goal of this section is to study these special morphisms.
Denition 3.3.1. Let N
1
, N
2
be two lattices with
1
a fan in (N
1
)
R
and
2
a fan in
(N
2
)
R
. A Z-linear mapping : N
1
N
2
is compatible with the fans
1
and
2
if
for every cone
1

1
, there exists a cone
2

2
such that
R
(
1
)
2
.
Example 3.3.2. Let N
1
= Z
2
with basis e
1
, e
2
and let
r
be the fan from Figure 4
in 3.1. By Example 3.1.16, X
r
is the Hirzebruch surface H
r
. Also let N
2
= Z
and consider the fan giving P
1
:
s
0
1

0
as in Example 3.1.11. The mapping
: N
1
N
2
, ae
1
+be
2
a
is compatible with the fans
r
and since each cone of
r
maps onto a cone of .
If r = 0, on the other hand, the mapping
: N
1
N
2
, ae
1
+be
2
b
is not compatible with these fans since
3

r
does not map into a cone of .
The Denition of Toric Morphism. In 1.3, we dened a toric morphism in the
afne case and gave an equivalent condition in Proposition 1.3.14. For general
toric varieties, it more convenient to take the result of Proposition 1.3.14 as the
denition of toric morphism.
Denition 3.3.3. Let X

1
, X

2
be normal toric varieties, with
1
a fan in (N
1
)
R
and
2
a fan in (N
2
)
R
. A morphism : X

1
X

2
is toric if maps the torus
T
N
1
X

1
into T
N
2
X

2
and [
T
N
1
is a group homomorphism.
126 Chapter 3. Normal Toric Varieties
The proof of part (b) of Proposition 1.3.14 generalizes easily to show that any
toric morphism : X

1
X

2
is an equivariant mapping for the T
N
1
- and T
N
2
-
actions. That is, we have a commutative diagram
(3.3.1)
T
N
1
X

1

[
T
N
1

T
N
2
X

2

X

2
where the horizontal maps give the torus actions.
Our rst result shows that toric morphisms : X

1
X

2
correspond to Z-
linear mappings : N
1
N
2
that are compatible with the fans
1
and
2
.
Theorem 3.3.4. Let N
1
, N
2
be lattices and let
i
be a fan in (N
i
)
R
, i = 1, 2.
(a) If : N
1
N
2
is a Z-linear map that is compatible with
1
and
2
, then there
is a toric morphism : X

1
X

2
such that [
T
N
1
is the map
1 : N
1

Z
C

N
2

Z
C

.
(b) Conversely, if : X

1
X

2
is a toric morphism, then induces a Z-linear
map : N
1
N
2
that is compatible with the fans
1
and
2
.
Proof. To prove part (a), let
1
be a cone in
1
. Since is compatible with
1
and
2
, there is a cone
2

2
with
R
(
1
)
2
. Then Proposition 1.3.15 shows
that induces a toric morphism

1
: U

1
U

2
. Using the general criterion for
gluing morphisms from Exercise 3.3.1, you will show in Exercise 3.3.2 that the

1
glue together to give a morphism : X

1
X

2
. Moreover, is toric be-
cause taking
1
=0 gives
|0
: T
N
1
T
N
2
, which is easily seen to be the group
homomorphism induced by the Z-linear map : N
1
N
2
.
For part (b), we show rst that the toric morphism induces a Z-linear map
: N
1
N
2
. This follows since [
T
N
1
is a group homomorphism. Hence, given
u N
1
, the one-parameter subgroup
u
: C

T
N
1
can be composed with [
T
N
1
to
give the one-parameter subgroup [
T
N
1

u
: C

T
N
2
. This denes an element
(u) N
2
. It is straightforward to show that : N
1
N
2
is Z-linear.
It remains to show that is compatible with
1
and
2
. Take
1

1
. By
the Orbit-Cone Correspondence (Theorem 3.2.6), this gives the T
N
1
-orbit O
1
=
O(
1
) X

1
. Because of the equivariance (3.3.1), there is a T
N
2
-orbit O
2
X

2
with (O
1
) O
2
. Using Theorem 3.2.6 again, we have O
2
= O(
2
) for some cone

2

2
. Thus (O(
1
)) O(
2
). Furthermore, if
1
_
1
is a face, then by the
same reasoning, there is some cone
2

2
such that (O(
1
)) O(
2
).
We claim that in this situation
2
must be a face of
2
. This follows since
O(
1
) O(
1
) by part (d) of Theorem 3.2.6. Since is continuous in the Zariski
topology,

O(
1
)

O(
2
). But the only orbits contained in the closure of O(
2
)
3.3. Toric Morphisms 127
are the orbits corresponding to cones which have
2
as a face. So
2
is a face of
2
.
It follows from part (c) of Theorem 3.2.6 that also maps the afne open subset
U

1
X

1
into U

2
X

2
, i.e.,
(3.3.2) (U

1
) U

2
.
Hence induces a toric morphismU

1
U

2
, which by Proposition 1.3.15 implies
that
R
(
1
)
2
. Hence is compatible with the fans
1
and
2
.
First Examples. Here are some examples of toric morphisms dened by mappings
compatible with the corresponding fans.
Example 3.3.5. Let N
1
=Z
2
and N
2
=Z, and let
: N
1
N
2
, ae
1
+be
2
a,
be the rst mapping in Example 3.3.2. We saw that is compatible with the fans

r
of the Hirzebruch surface H
r
and of P
1
. Theorem 3.3.4 implies that there is
a corresponding toric morphism : H
r
P
1
. We will see later in the section that
this mapping gives H
r
the structure of a P
1
-bundle over P
1
.
Example 3.3.6. Let N =Z
n
and be a fan in N
R
. For Z
>0
, the multiplication
map

: N N, a a
is compatible with . By Theorem 3.3.4, there is a corresponding toric morphism

: X

whose restriction to T
N
X

is the group endomorphism

[
T
N
(t
1
, . . . , t
n
) = (t

1
, . . . , t

n
).
For a concrete example, let be the fan in N
R
=R
2
from Figure 2 and take = 2.
Then we obtain the morphism
2
: P
2
P
2
dened in homogeneous coordinates
by
2
(x
0
, x
1
, x
2
) = (x
2
0
, x
2
1
, x
2
2
). We will use

in Chapter 9.
Sublattices of Finite Index. We get an interesting toric morphism when a lattice
N

has nite index in a larger lattice N. If is a fan in N


R
, then we can view as
a fan either in N

R
or in N
R
, and the inclusion N

N is compatible with the fan


in N

R
and N
R
. As in Chapter 1, we obtain toric varieties X
,N
and X
,N
depending
on which lattice we consider, and the inclusion N

N induces a toric morphism


: X
,N
X
,N
.
Proposition 3.3.7. Let N

be a sublattice of nite index in N and let be a fan in


N
R
= N

R
. Let G = N/N

. Then
: X
,N
X
,N
induced by the inclusion N

N presents X
,N
as the quotient X
,N

/G.
128 Chapter 3. Normal Toric Varieties
Proof. Since N

has nite index in N, Proposition 1.3.18 shows that the nite group
G = N/N

is the kernel of T
N
T
N
. It follows that G acts on X
,N
. This ac-
tion is compatible with the inclusion U
,N
X
,N
for each cone . Using
Proposition 1.3.18 again, we see that U
,N
/G U
,N
, which easily implies that
X
,N
/G X
,N
.
We will revisit Proposition 3.3.7 in Chapter 5, where we will show that the
map : X
,N
X
,N
is a geometric quotient.
Example 3.3.8. Let N = Z
2
, and be the fan shown in Figure 5, so X
,N
gives
the weighted projective space P(1, 1, 2). Let N

be the sublattice of N given by


N

=(a, b) N [ b 0 mod 2, so N

has index 2 in N. Note that N

is generated
by u
1
= e
1
, u
2
= 2e
2
and that
u
0
=e
1
2e
2
=u
1
u
2
N

.
Let : N

N be the inclusion map. It is not difcult to see that with respect to


the lattice N

, X
,N
P
2
(Exercise 3.3.3). By Theorem 3.3.4, the Z-linear map
induces a toric morphism : P
2
P(1, 1, 2), and by Proposition 3.3.7, it follows
that P(1, 1, 2) P
2
/G for G = N/N

Z/2Z.

2
Figure 8. The semigroups

2
M and

2
M

The cone
2
from Figure 5 has the dual cone

2
shown in Figure 8. It is
instructive to consider how

2
interacts with the lattice M

dual to N

. One checks
that M

(a, b/2) : a, b Z and

2
= Cone(2e
1
e
2
, e
2
). In Figure 8, the
points in

2
M are shown in white, and the points in

2
M

not in

2
M
are shown in black. Note that the picture in

2
M is the same (up to a change
of coordinates in GL(2, Z)) as Figure 10 from Chapter 1. This shows again that
P(1, 1, 2) contains the afne open subset U

2
,N
isomorphic to the rational normal
cone

C
2
. On the other hand U

2
,N
C
2
is smooth. The other afne open subsets
corresponding to
1
and
0
are isomorphic to C
2
in both P
2
and in P(1, 1, 2).
3.3. Toric Morphisms 129
Torus Factors. A toric variety X

has a torus factor if it is equivariantly isomor-


phic to the product of a nontrivial torus and a toric variety of smaller dimension.
Proposition 3.3.9. Let X

be the toric variety of the fan . Then the following are


equivalent:
(a) X

has a torus factor.


(b) There is a nonconstant morphism X

.
(c) The u

, (1), do not span N


R
.
Recall that (1) consists of the 1-dimensional cones of , i.e., its rays, and
that u

is the minimal generator of a ray (1).


Proof. If X

(C

)
r
for r > 0 and some toric variety X

, then a nontrivial
character of (C

)
r
gives a nonconstant morphism X

(C

)
r
C

.
If : X

is a nonconstant morphism, then Exercise 3.3.4 implies that


[
T
N
= c
m
where c C

and m M`0. Multiplying by c


1
, we may assume
that [
T
N
=
m
. Then is a toric morphism coming from a nonzero homomorphism
: N Z. Since C

comes from the trivial fan,


R
maps all cones of to the
origin. Hence u

ker(
R
) for all (1), so the u

do not span N
R
.
Finally, suppose that the u

, (1) span a proper subspace of N


R
. Then N

=
Span(u

[ (1)) N is proper sublattice of N such that N/N

is torsion-free, so
N

has a complement N

with N = N

. Furthermore, can be regarded as a


fan

in N

R
, and then is the product fan =

, where

is the trivial
fan in N

R
. Then Proposition 3.1.14 gives an isomorphism
X

,N

T
N

,N

(C

)
nk
,
where dim N
R
= n and dim N

R
= k.
In later chapters, toric varieties without torus factors will play an important
role. Hence we state the following corollary of Proposition 3.3.9.
Corollary 3.3.10. Let X

be the toric variety of the fan . Then the following are


equivalent:
(a) X

has no torus factors.


(b) Every morphism X

is constant, i.e., (X

, O
X

=C

.
(c) The u

, (1), span N
R
.
We can also think about torus factors from the point of view of sublattices.
Proposition 3.3.11. Let N

N be a sublattice with dim N


R
= n, dim N

R
= k. Let
be a fan in N

R
, which we can regard as a fan in N
R
. Then:
(a) If N

is spanned by a subset of a basis of N, then we have an isomorphism


: X
,N
X
,N
T
N/N
X
,N
(C

)
nk
.
130 Chapter 3. Normal Toric Varieties
(b) In general, a basis for N

can be extended to a basis of a sublattice N

N of
nite index. Then X
,N
is isomorphic to the quotient of
X
,N

X
,N

T
N

/N
X
,N

(C

)
nk
by the nite abelian group N/N

.
Proof. Part (a) follows from the proof of Proposition 3.3.9, and part (b) follows
from part (a) and Proposition 3.3.7.
Renements of Fans and Blowups. Given a fan in N
R
, a fan

renes if
every cone of

is contained in a cone of and [

[ =[[. Hence every cone of


is a union of cones of

. When

renes , the identity mapping on N is clearly


compatible with

and . This yields a toric morphism : X

.
Example 3.3.12. Consider the fan

in N Z
2
pictured in Figure 1 from 3.1.
This is a renement of the fan consisting of Cone(e
1
, e
2
) and its faces. The
corresponding toric varieties are X

C
2
and X

W = V(x
0
yx
1
x) P
1
C
2
,
the blowup of C
2
at the origin (see Example 3.1.15). The identity map on N induces
a toric morphism : W C
2
. This blowdown morphism maps P
1
0 W to
0 C
2
and is injective outside of P
1
0 in W.
We can generalize this example and Example 3.1.5 as follows.
Denition 3.3.13. Let be a fan in N
R
R
n
. Let = Cone(u
1
, . . . , u
n
) be a
smooth cone in , so that u
1
, . . . , u
n
is a basis for N. Let u
0
= u
1
+ +u
n
and let

() be the set of all cones generated by subsets of u


0
, . . . , u
n
not containing
u
1
, . . . , u
n
. Then

() = (`)

()
is a fan in N
R
called the star subdivision of along .
Example 3.3.14. Let = Cone(u
1
, u
2
, u
3
) N
R
R
3
be a smooth cone. Figure 9
on the next page shows the star subdivision of into three cones
Cone(u
0
, u
1
, u
2
), Cone(u
0
, u
1
, u
3
), Cone(u
0
, u
2
, u
3
).
The fan

() consists of these cones, together with their faces.


Proposition 3.3.15.

() is a renement of , and the induced toric morphism


: X

()
X

makes X

()
the blowup of X

at the distinguished point

corresponding to the
cone .
Proof. Since and

() are the same outside the cone , without loss of gener-


ality, we may reduce to the case that is the fan consisting of and all of its faces,
and X

is the afne toric variety U

C
n
.
3.3. Toric Morphisms 131
u
1
u
3
u
2
u
0
Figure 9. The star subdivision

() in Example 3.3.14
Under the Orbit-Cone Correspondence (Theorem 3.2.6), corresponds to the
distinguished point

, the origin (the unique xed point of the torus action). By


Theorem 3.3.4, the identity map on N induces a toric morphism
: X

()
U

C
n
.
It is easy to check that the afne open sets covering X

()
are the same as for the
blowup of C
n
at the origin from Exercise 3.0.8, and they are glued together in the
same way by Exercise 3.1.5.
The blowup X

at

is sometimes denoted Bl

(X

). In this notation, the


blowup of C
n
at the origin is written Bl
0
(C
n
).
The point blown up in Proposition 3.3.15 is a xed point of the torus action.
In some cases, torus-invariant subvarieties of larger dimension have equally nice
blowups. We begin with the afne case. The standard basis e
1
, . . . , e
n
of Z
n
gives
= Cone(e
1
, . . . , e
n
) with U

= C
n
, and the face = Cone(e
1
, . . . , e
r
), 2 r n,
gives the orbit closure
V() = O() =0C
nr
.
To construct the blowup of V(), let u
0
= u
1
+ +u
r
and consider the fan
(3.3.3)

() =Cone(A) [ A u
0
, . . . , u
n
, u
1
, . . . , u
r
A.
Example 3.3.16. Let = Cone(e
1
, e
2
, e
3
) N
R
R
3
and = Cone(e
1
, e
2
). The
star subdivision relative to subdivides into the cones
Cone(e
0
, e
1
, e
3
), Cone(e
0
, e
2
, e
3
),
as shown in Figure 10 on the next page. The fan

() consists of these two cones,


together with their faces.
132 Chapter 3. Normal Toric Varieties
e
1
e
3
e
2
e
0

Figure 10. The star subdivision

() in Example 3.3.16
For the fan (3.3.3), the toric variety X

()
is the blowup of 0C
nr
C
n
.
To see why, observe that

() is a product fan. Namely, Z


n
=Z
r
Z
nr
, and

() =
1

2
,
where
1
is the fan for Bl
0
(C
r
) (coming from a renement of Cone(u
1
, . . . , u
r
))
and
2
is the fan for C
nr
(coming from Cone(u
r+1
, . . . , u
n
)). It follows that
X

()
= Bl
0
(C
r
) C
nr
.
Since Bl
0
(C
r
) is built by replacing 0 C
r
with P
r1
, it follows that X

()
=
Bl
0
(C
r
) C
nr
is built by replacing 0 C
nr
C
n
with P
r1
C
nr
. The
intuitive idea is that Bl
0
(C
r
) separates directions through the origin in C
r
, while
the blowup Bl
|0C
nr (C
n
) = X

()
separates normal directions to 0C
nr
in
C
n
. One can also study Bl
|0C
nr (C
n
) by working on afne pieces given by the
maximal cones of

()see [218, Prop. 1.26].


We generalize (3.3.3) as follows.
Denition 3.3.17. Let be a fan in N
R
R
n
and assume has the property
that all cones of containing are smooth. Let u

(1)
u

and for each cone


containing , set

() =Cone(A) [ A u

(1), (1) A.
Then the star subdivision of relative to is the fan

() = [

().
The fan

() is a renement of and hence induces a toric morphism


: X

()
X

.
3.3. Toric Morphisms 133
Under the map , X

()
becomes the blowup Bl
V()
(X

) of X

along the orbit


closure V().
In Chapters 10 and 11 we will use toric morphisms coming from a generalized
version of star subdivision to resolve the singularities of toric varieties.
Exact Sequences and Fibrations. Next, we consider a class of toric morphisms
that have a nice local structure. To begin, consider a surjective Z-linear mapping
: N N

.
If in N
R
and

in N

R
are compatible with , then we have a corresponding toric
morphism
: X

.
Now let N
0
= ker(), so that we have an exact sequence
(3.3.4) 0 N
0
N

N

0.
It is easy to check that

0
= [ (N
0
)
R

is a subfan of whose cones lie in (N


0
)
R
N
R
. By Proposition 3.3.11,
(3.3.5) X

0
,N
X

0
,N
0
T
N

since N/N
0
N

. Furthermore, is compatible with


0
in N
R
and the trivial fan
0 in N

R
. This gives the toric morphism
[
X

0
,N
: X

0
,N
T
N
.
In fact, by the reasoning to prove Proposition 3.3.4,
(3.3.6)
1
(T
N

) = X

0
,N
X

0
,N
0
T
N

.
In other words, the part of X

lying over T
N

is identied with the product


of T
N
and the toric variety X

0
,N
0
. We say this subset of X

is a ber bundle over


T
N
with ber X

0
,N
0
.
When the fan has a suitable structure relative to , we can make a similar
statement for every torus-invariant afne open subset of X

.
Denition 3.3.18. In the situation of (3.3.4), we say is split by

and
0
if there
exists a subfan

such that:
(a)
R
maps each cone

bijectively to a cone

such that ( N) =

. Furthermore, the map

denes a bijection

.
(b) Given cones

and
0

0
, the sum +
0
lies in , and every cone of
arises this way.
134 Chapter 3. Normal Toric Varieties
Theorem 3.3.19. If is split by

and
0
as in Denition 3.3.18, then X

is a
locally trival ber bundle over X

with ber X

0
,N
0
, i.e., X

has a cover by afne


open subsets U satisfying

1
(U) X

0
,N
0
U.
In particular, all bers of X

are isomorphic to X

0
,N
0
.
Proof. Fix

in

and let (

) = [ ()

. Then

1
(U

) = X
(

)
.
It remains to show that X
(

)
X

0
,N
0
U

. Since (

) is split by
0
(

)
and

(

), we may assume X

=U

. In other words, we are reduced to the


case when

consists of

and its proper faces.


A Z-linear map : N

N splits the exact sequence (3.3.4) provided is


the identity on N

. A splitting induces an isomorphism


N
0
N

N.
By Denition 3.3.18, there is a cone

such that ( N) =

and
R
maps bijectively to

. Using , one can nd a splitting with the property that

R
maps

to for all

(Exercise 3.3.5). Using Denition 3.3.18 again, we
see that
(N
0
)
R
N

R
N
R
carries the product fan (
0
, (N
0
)
R
) (

, N

R
) to the fan (, N
R
). By Proposi-
tion 3.1.14, we conclude that
X

0
,N
0
X

0
,N
0
U

,
and the theorem is proved.
Example 3.3.20. To complete the discussion begun in Examples 3.3.2 and 3.3.5,
consider the toric morphism : H
r
P
1
induced by the mapping
: Z
2
Z, ae
1
+be
2
a.
The fan
r
of H
r
is split by the fan of P
1
and
0
=
r
[
R
() = 0
because of the subfan

of
r
consisting of the cones
Cone(e
1
+re
2
), 0, Cone(e
1
).
These cones are mapped bijectively to the cones in

under
R
. Note also that
0
consists of the cones
Cone(e
2
), 0, Cone(e
2
).
The fans

and
0
are shown in Figure 11 on the next page.
As we vary over all

and
0

0
, the sums +
0
give all cones of H
r
.
Hence Theorem 3.3.19 shows that H
r
is a locally trivial bration over P
1
, with
bers isomorphic to
X

0
,N
0
P
1
,
3.3. Toric Morphisms 135

2
(1, r)

Figure 11. The Splitting of the Fan r


where N
0
=ker() gives the vertical axis in Figure 11. This bration is not globally
trivial when r > 0, i.e., it is not true that H
r
P
1
P
1
. There is some twisting
on the bers involved when we try to glue together the
1
(U

) U

P
1
to
obtain H
r
.
We will give another, more precise, description of these ber bundles and the
twisting mentioned above using the language of sheaves in Chapter 7.
The denition of splitting fan in Denition 3.3.18 requires that ( N) =

when

maps to

. Exercise 3.3.6 will give an example of how


Theorem 3.3.19 can fail if this condition is not met, and Exercise 3.3.7 will explore
how to modify the theorem when this happens.
Images of Distinguished Points. Each orbit O() in a toric variety X

contains
a distinguished point

, and each orbit closure V() is a toric variety in its own


right. These structures are compatible with toric morphisms as follows.
Lemma 3.3.21. Let : X

be the toric morphism coming from a map :


N N

that is compatible with and

. Given , let

be the minimal
cone of

containing
R
(). Then:
(a) (

) =

, where

O() and

O(

) are the distinguished points.


(b) (O()) O(

) and

V()

V(

).
(c) The induced map [
V()
: V() V(

) is a toric morphism.
136 Chapter 3. Normal Toric Varieties
Proof. First observe that if

1
,

contain
R
(), then so does their intersec-
tion. Hence

has a minimal cone containing


R
().
To prove part (a), pick u Relint() and observe that (u) Relint(

) by the
minimality of

. Then
(

) =

lim
t0

u
(t)

= lim
t0
(
u
(t)) = lim
t0

(u)
(t) =

,
where the rst and last equalities use Proposition 3.2.2.
The rst assertion of part (b) follows immediately from part (a) by the equiv-
ariance, and the second assertion follows by continuity (as usual, we get the same
closure in the classical and Zariski topologies).
For (c), observe that [
O()
: O() O(

) is a morphism that is also a group


homomorphismthis follows easily from equivariance. Since the orbit closures
are toric varieties by Proposition 3.2.7, the map [
V()
: V() V(

) is a toric
morphism according to Denition 3.3.3.
Exercises for 3.3.
3.3.1. Let X be a variety with an afne open cover U
i
, and let Y be a second variety. Let

i
: U
i
Y be a collection of morphisms. We say that a morphism : X Y is obtained
by gluing the
i
if [
Ui
=
i
for all i. Show that there exists such a : X Y if and only if
for every pair i, j,

i[
UiUj
=
j [
UiUj
.
3.3.2. Let N
1
, N
2
be lattices, and let
1
in (N
1
)
R
,
2
in (N
2
)
R
be fans. Let : N
1
N
2
be
a Z-linear mapping that is compatible with the corresponding fans. Using Exercise 3.3.1
above, show that the toric morphisms
1
: U
1
U
2
constructed in the proof of Theo-
rem 3.3.4 glue together to form a morphism : X
1
X
2
.
3.3.3. This exercise asks you to verify some of the claims made in Example 3.3.8.
(a) Verify that X
,N
P
2
with respect to the lattice N

.
(b) Verify carefully that the afne open subset U
2,N


C
2
, where

C
2
is the rational normal
cone

C
d
with d = 2.
3.3.4. A character
m
, m M, gives a morphism T
N
C

. Here you will determine all


morphisms T
N
C

.
(a) Explain why morphisms T
N
C

correspond to invertible elements in the coordinate


ring of T
N
.
(b) Let c C

and Z
n
. Prove that ct

is invertible in C[t
1
1
, . . . , t
1
n
] and that all
invertible elements of C[t
1
1
, . . . , t
1
n
] are of this form.
(c) Use part (a) to show that all morphisms T
N
C

on T
N
are of the form c
m
for c C

and m M.
3.3.5. Let : N N

be a surjective Z-linear mapping and let and

be cones in N
R
and N

R
respectively with the property that
R
maps bijectively onto

. Prove that has


a splitting : N

N such that maps

to .
3.3. Toric Morphisms 137
3.3.6. Let be the fan in R
2
with minimal generators u
1
= (0, 1), u
2
= (2, 1) and u
3
=
(0, 1) and maximal cones Cone(u
1
, u
2
) and Cone(u
2
, u
3
). Let : Z
2
Z be projection onto
the rst factor and let
0
be the subfan of dened in the discussion following (3.3.4).
Also let

be the fan in R with maximal cone

=R
0
.
(a) Prove that is not split by

and
0
. Hint: Show that X
0,N0
and X

are smooth and


then show that Theorem 3.3.19 must fail.
(b) Let

be the subfan of with maximal cone =Cone(u


2
). Show that this satises all
parts of Denition 3.3.18 except for the requirement that ( Z
2
) =

Z. Draw a
picture similar to Figure 11 in Example 3.3.20.
3.3.7. In the situation of Denition 3.3.18, we say that is weakly split by

and
0
if

satises Denition 3.3.18 except that we no longer require ( N) =

.
(a) Explain why Exercise 3.3.6 is an example of a weak splitting that is not a splitting.
(b) In the situation of a weak splitting, prove that all bers of : X

are isomorphic
to X
0,N0
. Hint: First assume X

= U

. Prove that there is a sublattice of N

of nite index such that splits when we use the lattices N

and N

. Then show that


there is a commutative diagram
X
0,N0
U

,N

,N

U

.
such that X
0,N0
U

,N
is the ber product X

,N
as dened in (3.0.5).
Thus, while Theorem 3.3.19 may fail for a weak splitting, at least part remains true.
3.3.8. Let

be the fan obtained from the fan for P


2
in Example 3.1.9 by the following
process. Subdivide the cone
2
into two new cones
21
and
22
by inserting an edge
Cone(e
2
).
(a) Show that the resulting toric variety X

is smooth.
(b) Show that X

is the blowup of P
2
at the point V(
2
).
(c) Show that X

is isomorphic to the Hirzebruch surface H


1
.
3.3.9. Let X

be the toric variety obtained from P


2
by blowing up the points V(
1
) and
V(
2
) (see Figure 2 in Example 3.1.9). Show that X

is isomorphic to the blowup of


P
1
P
1
at the point V(
11
) (see Figure 3 in Example 3.1.12).
3.3.10. Let

be the fan obtained from the fan for P(1, 1, 2) in Example 3.1.17 by the
following process. Subdivide the cone
2
into two new cones
21
and
22
by inserting an
edge Cone(u
2
).
(a) Show that the resulting toric variety X

is smooth.
(b) Construct a morphism : X

and determine the ber over the unique singular


point of X

.
(c) One of our smooth examples is isomorphic to X

. Which one is it?


3.3.11. Consider the action of the group G = (,
3
) [
5
= 1 (C

)
2
on C
2
. We will
study the quotient C
2
/G and its resolution of singularities using toric morphisms.
138 Chapter 3. Normal Toric Varieties
(a) Let N

=Z
2
and N =(a/5, b/5) [ a, b Z, b 3a mod 5. Also let
5
=e
2i/5
. Prove
that the map N (C

)
2
dened by (a/5, b/5) (
a
5
,
b
5
) induces an exact sequence
0 N

N G 0.
(b) Let = Cone(e
1
, e
2
) N

R
= N
R
= R
2
. The inclusion N

N induces a toric mor-


phism U
,N
U
,N
. Prove that this is the quotient map C
2
C
2
/G for the above
action of G on C
2
.
(c) Find the Hilbert basis (i.e., the set of irreducible elements) of the semigroup N.
Hint: The Hilbert basis has four elements.
(d) Use the Hilbert basis from part (c) to subdivide . This gives a fan with [[ = .
Prove that is smooth relative to N and that the resulting toric morphism
X
,N
U
,N
=C
2
/G
is a resolution of singularities. See Chapter 10 for more details.
(e) The group G gives the nite set G (C

)
2
C
2
with ideal I(G) = 'x
5
1, y x
3
`.
Read about the Gr obner fan in [70, Ch. 8, 4] and compute the Gr obner fan of I(G).
The answer will be identical to the fan described in part (d). This is no accident, as
shown in the paper [156] (see also 10.3). There is a lot of interesting mathematics
going on here, including the M
c
Kay correspondence and the G-Hilbert scheme. See
also [206] for the higher dimensional case.
3.3.12. Consider the fan in R
3
shown in Figure 12. This fan has ve 1-dimensional
z
y
x
Figure 12. The fan for Exercise 3.3.12
cones with four upward ray generators (1, 0, 1), (0, 1, 1) and one downward gen-
erator (0, 0, 1). There are also nine 2-dimensional cones. Figure 12 shows ve of the
3.4. Complete and Proper 139
2-dimensional cones; the remaining four are generated by the combining the downward
generator with the four upward generators.
(a) Show that projection onto the y-axis induces a toric morphism X

P
1
.
(b) Show that X

P
1
is a locally trivial ber bundle over P
1
with ber P(1, 1, 2). Hint:
Theorem 3.3.19 and (1, 0, 1) +(1, 0, 1) +2(0, 0, 1) = 0. See Example 3.1.17.
(c) Explain how you can see the splitting (in the sense of Denition 3.3.18) in Figure 12.
Also explain why the gure makes it clear that the ber is P(1, 1, 2).
3.3.13. Consider the fan in R
2
with ray generators
u
0
= e
1
+e
2
, u
1
= e
1
, u
2
= e
2
, u
3
=e
1
and 1-dimensional cones Cone(u
0
, u
1
), Cone(u
0
, u
2
), Cone(u
2
, u
3
).
(a) Draw a picture of and prove that X

is the blowup of P
1
C at one point.
(b) Show that the map ae
1
+be
2
b induces a toric morphism : X

C such that

1
() P
1
for C

and
1
(0) is a union of two copies of P
1
meeting at a point.
Hint: Once you understand
1
(0), show that the fan for X

`
1
(0) gives P
1
C

.
(c) To get a better picture of X

, consider the map : (C

)
2
P
3
C dened by
(s, t) = ((s
3
, s
2
, st, t
2
), t).
Let X = ((C

)
2
) P
3
C be the closure of the image. Prove that X X

and
that the restriction of the projection P
3
C C to X gives the toric morphism of
part (b).
(d) Let x, y, z, w be coordinates on P
3
. Prove that X P
3
C is dened by the equations
ywz
2
= 0, xz ty
2
= 0, xwtyz = 0.
Also use these equations to describe the bers
1
() for C, and explain how this
relates to part (b). Hint: The twisted cubic is relevant.
This is a semi-stable degeneration of toric varieties. See [148] for more details.
3.4. Complete and Proper
The Compactness Criterion. We begin by proving part (c) of Theorem 3.1.19.
Theorem 3.4.1. Let X

be a toric variety. Then the following are equivalent:


(a) X

is compact in the classical topology.


(b) The limit lim
t0

u
(t) exists in X

for all u N.
(c) is complete, i.e., [[ =

= N
R
.
Proof. First observe that since X

is separated (Theorem 3.1.5), it is Hausdorff in


the classical topology (Theorem 3.0.17). In fact, since the classical topology on
each afne open set U

is a metric topology, X

is compact if and only if every


sequence of points in X

has a convergent subsequence.


For (a) (b), assume that X

is compact and x u N. Given a sequence


t
k
C

converging to 0, we get the sequence


u
(t
k
) X

. By compactness, this se-


quence has a convergent subsequence. Passing to this subsequence, we can assume
140 Chapter 3. Normal Toric Varieties
that lim
k

u
(t
k
) = X

. Because X

is the union of the afne open subsets


U

for , we may assume U

. Now take m

M. The character
m
is
a regular function on U

and hence is continuous in the classical topology. Thus

m
() = lim
k

m
(
u
(t
k
)) = lim
k
t
/m,u)
k
.
Since t
k
0, the exponent must be nonnegative, i.e., 'm, u` 0 for all m

M.
This implies 'm, u` 0 for all m

, so that u (

= . Then Proposi-
tion 3.2.2 implies that lim
t0

u
(t) exists in U

and hence in X

.
To prove (b) (c), take u N and consider the limit lim
t0

u
(t). This lies
in some afne open U

, which implies u N by Proposition 3.2.2. Thus every


lattice point of N
R
is contained in a cone of . It follows that is complete.
We will prove (c) (a) by induction on n = dim N
R
. In the case n = 1, the
only complete fan is the fan in Rpictured in Example 3.1.11. The corresponding
toric variety is X

= P
1
. This is homeomorphic to S
2
, the 2-dimensional sphere,
and hence is compact.
Now assume the statement is true for all complete fans of dimension strictly
less than n, and consider a complete fan in N
R
R
n
. Let
k
X

be a sequence.
We will show that
k
has a convergent subsequence.
Since X

is the union of nitely many orbits O(), we may assume the se-
quence
k
lies entirely within an orbit O(). If =0, then the closure of O() in
X

is the toric variety V() = X


Star()
of dimension n1 by Proposition 3.2.7.
Since is complete, it is easy to check that the fan Star() is complete in N()
R
(Exercise 3.4.1). Then the induction hypothesis implies that there is a convergent
subsequence in V(). Hence, without loss of generality again, we may assume that
our sequence lies entirely in the torus T
N
X

.
Recall from the discussion following Lemma 3.2.5 that
T
N
Hom
Z
(M, C

).
Moreover, when we regard T
N
as a group homomorphism : MC

, then for
any , restriction yields a semigroup homomorphism

M C and hence
a point in U

.
A key ingredient of the proof will be the logarithm map L : T
N
N
R
dened
as follows. Given a point : M C

of T
N
, consider the map M R dened by
the formula
m log[(m)[.
This is a homomorphism and hence gives an element L() Hom
Z
(M, R) N
R
.
For more properties of this mapping, see Exercise 3.4.2 below.
For us, the most important property of L is the following. Suppose that a point
T
N
satises L() for some . If m

M, then the denition of


3.4. Complete and Proper 141
L implies that
(3.4.1) log[(m)[ ='m, L()`,
which is 0 since m

and L() . Hence [(m)[ 1. Thus we have


proved that
(3.4.2) L() = [(m)[ 1 for all m

M.
Now apply L to our sequence, which gives a sequence L(
k
) N
R
. Since is
complete, the same is true for the fan consisting of the cones for . Hence,
by passing to a subsequence, we may assume that there is such that
L(
k
)
for all k. By (3.4.2), we conclude that [
k
(m)[ 1 for all m

M. It follows
that the
k
are a sequence of mappings to the closed unit disk in C. Since the closed
unit disk is compact, there is a subsequence
k

which converges to a point U

.
You will check the details of this nal assertion in Exercise 3.4.3.
Complete Varieties. The compactness criterion proved in Theorem 3.4.1 uses the
classical topology. It is natural to ask for an algebraic version of this theorem that
uses only the Zariski topology. The crucial idea is the notion of completeness.
To motivate the denition of completeness, we rst reformate the topological
notion of compactness. You will prove the following in Exercise 3.4.4.
Proposition 3.4.2. Let X be a locally compact Hausdorff topological space. Then
the following are equivalent:
(a) X is compact.
(b) For every topological space Z, the projection map
Z
: X Z Z is closed,
i.e.,
Z
(W) Z is closed for all closed subsets W X Z.
We now dene the algebraic analog of compactness.
Denition 3.4.3. A variety X is complete if for every variety Z, the projection map

Z
: X Z Z is a closed mapping in the Zariski topology.
Here are two examples to illustrate this denition.
Example 3.4.4. Consider the afne variety C. We claim that C is not complete.
To see this, consider the projection map
2
: CC = C
2
C. The closed subset
V(xy 1) C
2
does not map to a Zariski-closed subset of C under
2
. Hence
2
is not a closed mapping, so that C is not complete.
Example 3.4.5. The Projective extension theorem [69, Thm. 6 of Ch. 8, 5] shows
that for X =P
n
, the mapping

C
m
: P
n
C
m
C
m
142 Chapter 3. Normal Toric Varieties
is closed in the Zariski topology for all m. It follows that if V C
m
is any afne
variety, the projection

V
: P
n
V V
is a closed mapping in the Zariski topology. Then the gluing construction shows
that
Z
: P
n
Z Z is closed for any variety Z, so P
n
is a complete variety. In fact,
one can think of P
n
as the prototypical complete variety. Moreover, any projective
variety is complete (Exercise 3.4.5). However, there are complete varieties that are
not projectivewe will give a toric example in Chapter 6.
Completeness is the algebraic version of compactness, and it can be shown
that a variety is complete if and only if it is compact in the classical topology. This
is proved in Serres famous paper G eom etrie alg ebrique et g eom etrie analytique,
called GAGA for short. See [248, Prop. 6, p. 12]. As a consequence, we get the
following improved version of Theorem 3.4.1.
Theorem 3.4.6. Let X

be a toric variety. Then the following are equivalent:


(a) X

is compact in the classical topology.


(b) X

is complete.
(c) The limit lim
t0

u
(t) exists in X

for all u N.
(d) is complete, i.e., [[ =

= N
R
.
Proper Mappings. In algebraic geometry, many concepts that apply to varieties
have relative versions that apply to morphisms. To see how this works for complete
varieties, we will begin in the topological category with the relative version of
compactness.
Denition 3.4.7. A continuous mapping f : X Y is proper if the inverse image
f
1
(T) is compact in X for every compact subset T Y.
It is immediate that X is compact if and only if the constant mapping from X
to the space Y = pt consisting of a single point is proper. This relative version
of compactness may also be reformulated, for reasonably nice topological spaces,
in the following way.
Proposition 3.4.8. Let f : X Y be a continuous mapping of locally compact rst
countable Hausdorff spaces. Then the following are equivalent:
(a) f is proper.
(b) f is a closed mapping, i.e., f (W) Y is closed for all closed subsets W X,
and all bers f
1
(y), y Y, are compact.
(c) Exery sequence x
k
X such that f (x
k
) Y converges in Y has a subsequence
x
k

that converges in X.
3.4. Complete and Proper 143
Proof. A proof of (a) (b) can be found in [122, Ch. 9, 4]. See Exercise 3.4.6
for (a) (c).
Before we can give a denition of properness that works for morphisms, we
rst need to reformulate the topological notion of properness. Recall from 3.0
that morphisms f : X S and g : Y S give the ber product X
S
Y. Fiber
products can also be dened for continuous maps between topological spaces. You
will prove in Exercise 3.4.6 that properness can be described using ber products
as follows.
Proposition 3.4.9. Let f : X Y be a continuous map between locally compact
Hausdorff spaces. Then f is proper if and only if f is universally closed, meaning
that for all spaces Z and all continuous mappings g : Z Y, the projection
Z
dened by the commutative diagram
X
Y
Z

X

X
f

Z
g

Y
is a closed mapping.
In algebraic geometry, we will use the following denition of properness for
morphisms between varieties.
Denition 3.4.10. A morphism of varieties : X Y is proper if it is universally
closed, in the sense that for all varieties Z and morphisms : Z Y, the projection

Z
dened by the commutative diagram
X
Y
Z

X

Y
is a closed mapping in the Zariski topology.
It is easy to see that a variety X is complete if and only if the constant morphism
: X pt is proper. Furthermore, if X is complete, then the projection map

Z
: X Z Z
is proper for any variety Z. You will prove these assertions in Exercise 3.4.7.
The Properness Criterion. Theorem 3.4.6 can be understood as a special case of
the following statement for toric morphisms.
144 Chapter 3. Normal Toric Varieties
Theorem 3.4.11. Let : X

be the toric morphism corresponding to a


homomorphism : N N

that is compatible with fans in N


R
and

in N

R
.
Then the following are equivalent:
(a) : X

is proper in the classical topology (Denition 3.4.7).


(b) : X

is a proper morphism (Denition 3.4.10).


(c) If u N and lim
t0

(u)
(t) exists in X

, then lim
t0

u
(t) exists in X

.
(d)
1
R
([

[) =[[.
Proof. The proof of (a) (b) uses two fundamental results in algebraic geometry.
First, given any morphism of varieties f : X Y and a Zariski closed subset
W X, a theorem of Chevalley tells us that the image f (W) Y is constructible,
meaning that it can be written as a nite union f (W) =

i
(V
i
`W
i
), where V
i
and
W
i
are Zariski closed in Y. A proof appears in [131, Ex. II.3.19].
Second, given any constructible subset C of a variety Y, its closure in Y in the
classical topology equals its closure in the Zariski topology. When C is open in the
Zariski topology, a proof is given in [207, Thm. (2.33)], and when C is the image
of a morphism, a proof can be found in GAGA [248, Prop. 7, p. 12].
Now suppose that : X

is proper in the classical topology and let


: Z X

be a morphism. This gives the commutative diagram


X

.
Let Y X

Z be Zariski closed. We need to prove that


Z
(Y) is Zariski closed
in Z. First observe that Y is also closed in the classical topology, so that
Z
(Y)
is closed in Z in the classical topology by Proposition 3.4.9. However,
Z
(Y) is
constructible by Chevalleys theorem, and then, being classically closed, it is also
Zariski closed by GAGA. Hence
Z
is a closed map in the Zariski topology for any
morphism : Z X

. It follows that is a proper morphism.


To prove (b) (c), let u N and assume that

=lim
t0

(u)
(t) exists in X

.
We rst prove lim
t0

u
(t) exists in X

under the extra assumption that (u) = 0.


This means that
(u)
is a nontrivial one-parameter subgroup in X

.
Let
u
(C

) X

be the closure of
u
(C

) X

in the classical topology.


Our earlier remarks imply that this equals the Zariski closure. Since is proper,
it is closed in the Zariski topology, so that

u
(C

is closed in X

in both
topologies. It follows that

(u)
(C

u
(C

.
3.4. Complete and Proper 145
Hence there is
u
(C

) mapping to

. Thus there is a sequence of points t


k
C

such that
u
(t
k
) . Then

= () = lim
k
(
u
(t
k
)) = lim
k

(u)
(t
k
).
This, together with

= lim
t0

(u)
(t) and (u) = 0, imply that t
k
0. From
here, the arguments used to prove (a) (b) (c) of Theorem 3.4.1 easily imply
that lim
t0

u
(t) exists in X

.
For the general case when we no longer assume (u) = 0, consider the map
(, 1
C
) : X

C X

C. This is proper since is proper (Exercise 3.4.8).


Furthermore, X

C and X

C are toric varieties by Proposition 3.1.14, and


the corresponding map on lattices is (, 1
Z
) : N Z N

Z. Then applying the


above argument to (u, 1) N Z shows that lim
t0

u
(t) exists in X

. We leave
the details to the reader (Exercise 3.4.8).
For (c) (d), rst observe that the inclusion
[[
1
R
([

[)
is automatic since is compatible with and

. For the opposite inclusion,


take u
1
R
([

[) N. Then (u) [

[, which by Proposition 3.2.2 implies that


lim
t0

(u)
(t) exists in X

. By assumption, lim
t0

u
(t) exists in X

. Using
Proposition 3.2.2, we conclude that u N for some . Because all the
cones are rational, this immediately implies
1
R
([

[) [[.
Finally, we prove (d) (a). We begin with two special cases.
Special Case 1. Suppose that a toric morphism : X

T
N
satises (d) and
has the additional property that : N N

is onto. The fan of T


N
consists of the
trivial cone 0, so that (d) implies
(3.4.3) [[ =
1
R
(0) = ker(
R
).
When we think of as a fan

in ker(
R
) N
R
, (3.3.5) implies that
X

T
N
.
Then corresponds to the projection X

T
N

T
N

. The fan

is complete in
ker(
R
) by (3.4.3), so that X

is compact by Theorem 3.4.1. Thus X

pt is
proper, which easily implies that X

T
N
T
N
is proper. We conclude that is
proper in the classical topology.
Special Case 2. Suppose that a homomorphism of tori : T
N
T
N
has the
additional property that : N N

is injective. Then (d) is obviously satised. An


elementary proof that is proper is given in Exercise 3.4.9.
Now consider a general toric morphism : X

satisfying (d). We will


prove that is proper in the classical topology using part (c) of Proposition 3.4.8.
146 Chapter 3. Normal Toric Varieties
Thus assume that
k
X

is a sequence such that (


k
) converges in X

. We need
to prove that a subsequence of
k
converges in X

.
Since X

has only nitely many T


N
-orbits, we may assume that the sequence
lies in an orbit O(). As in Lemma 3.3.21, let

be the minimal cone of

containing
R
(). The restriction
[
V()
: V() V(

)
is a toric morphism by Lemma 3.3.21, and the fans of V() and V(

) are given by
Star() in N()
R
and Star(

) in N

)
R
respectively. Furthermore, one can check
that since and

satisfy (d), the same is true for the fans Star() and Star(

)
(Exercise 3.4.10). Hence we may assume that
k
T
N
and (
k
) T
N
for all k.
The limit

= lim
k
(
k
) lies in an orbit O(

) for some

. Thus the
sequence (
k
) and its limit

all lie in U

. Note that [ ()

is the
fan giving
1
(U

). Since (d) implies that

1
R
(

) =

R
()

,
we can assume that X

=U

, i.e., : X

and
1
(

) =[[.
If

=0, then O(

) =U

= T

N
. If we write as the composition
N (N) N

,
then : X

T
N
factors as X

T
(N)
T
N
. Special Cases 1 and 2 imply that
these maps are proper, and since the composition of proper maps is proper, we
conclude that is proper.
It remains to consider the case when

=0. When we think of

as a
semigroup homomorphism

: (

MC, Lemma 3.2.5 tells us that

(m

) = 0 for all m

`(

.
Since the (
k
) : M C

converge to

in U

, we see that
lim
k
(
k
)(m

) = 0 for all m

`(

.
Since (

is nitely generated, it follows that we may pass to a subsequence


and assume that
(3.4.4) [(
k
)(m

)[ 1 for all k and all m

`(

.
3.4. Complete and Proper 147
The logarithm map from the proof of Theorem 3.4.1 gives maps L
N
: T
N
N
R
and L
N
: T
N
N

R
linked by a commutative diagram:
T
N
L
N

[
T
N

N
R

T
N

L
N

R
.
Let

: M

M be dual to : N N

. Then m

`(

implies
that for all k, we have
(3.4.5)
'

(m

), L
N
(
k
)` ='m

,
R
(L
N
(
k
))`
='m

, L
N
((
k
))` = log[(
k
)(m

)[ 0,
where the rst equality is standard, the second follows from the above commutative
diagram, the third follows from (3.4.1), and the nal inequality uses (3.4.4).
Now consider the following equivalences:
u
1
R
(

)
R
(u)

'm

,
R
(u)` 0 for all m

'

(m

), u` 0 for all m

,
where the rst and third equivalences are obvious and the second uses

= (

and the rationality of

. But we also know that

= 0, which means that (

is a cone whose maximal subspace (

is a proper subset. This implies that


u
1
R
(

) '

(m

), u` 0 for all m

`(

(Exercise 3.4.11). Using (3.4.5), we conclude that L


N
(
k
)
1
R
(

) for all k.
But, as noted above, (d) means
1
(

) =[[. It follows that


L
N
(
k
) [[
for all k. Passing to a subsequence, we may assume that there is such that
L
N
(
k
)
for all k. From here, the proof of (c) (a) in Theorem 3.4.1 implies that there is
a subsequence
k

which converges to a point U

. This proves that is


proper in the classical topology. The proof of the theorem is now complete.
We noted earlier that a variety is complete if and only if it is compact. In a
similar way, a morphism f : X Y of varieties is a proper morphism if and only if
it is proper in the classical topology. This is proved in [126, Prop. 3.2 of Exp. XII].
Thus the equivalence (a) (b) of Theorem 3.4.11 is a special case of this result.
148 Chapter 3. Normal Toric Varieties
Theorems 3.4.6 and 3.4.11 show that properness and completeness can be
tested using one-parameter subgroups. In the case of completeness, we can for-
mulate this as follows. Given u N, the one-parameter subgroup gives a map

u
: C`0 T
N
X

, and saying that lim


t0

u
(t) exists in X

means that
u
extends to a morphism
u
0
: C X

. In other words, whenever we have a commu-


tative diagram with solid arrows
C`0

u

u
0

(u)
0

pt,
the dashed arrow
u
0
exists so that the diagram remains commutative. The existence
of
u
0
tells us that the variety X

is not missing any points, which is where the


term complete comes from. In a similar way, the properness criterion given in
part (c) of Theorem 3.4.11 can be formulated as saying that whenever u N gives
a commutative diagram,
C`0

u

u
0

(u)
0

,
the dashed arrow
u
0
exists so that the diagram remains commutative.
For general varieties, there are similar criteria for completeness and properness
that replace
u
: C`0 X

and
u
0
: C X

with maps coming from discrete


valuation rings, to be discussed in Chapter 4. An example of a discrete valuation
ring is the ring of formal power series R = C[[t]], whose eld of fractions is the
eld of formal Laurent series K =C((t)). By replacing Cwith Spec(R) and C`0
with Spec(K) in the above diagrams, where R is now an arbitrary discrete valuation
ring, one gets the valuative criterion for properness (see [131, Ex. II.4.11 and Thm.
II.4.7]). This requires the full power of scheme theory since Spec(R) and Spec(K)
are not varieties as dened in this book. Using the valuative criterion of properness,
one can give a direct, purely algebraic proof of (d) (b) in Theorem 3.4.11 and
Corollary 3.4.6 (see [105, Sec. 2.4] or [218, Sec. 1.5]).
Example 3.4.12. An important class of proper morphisms are the toric morphisms
: X

induced by a renement

of . Condition (d) of Theorem 3.4.11 is


obviously fullled since : N N is the identity and every cone of is a union
of cones of

. In particular, the blowups


: X

()
X

studied in Proposition 3.3.15 are always proper.


3.4. Complete and Proper 149
Exercises for 3.4.
3.4.1. Let be a complete fan in N
R
and let be a cone in . Show that the fan Star()
dened in (3.2.8) is a complete fan in N()
R
.
3.4.2. In this exercise, you will develop some additional properties of the logarithm map-
ping L : T
N
N
R
dened in the proof of Theorem 3.4.1.
(a) Let S
1
be the unit circle in the complex plane, a subgroup of the multiplicative group
C

. Show that there is an isomorphism of groups


: C

S
1
R
z ([z[, log[z[),
where the operation in the second factor on the right is addition.
(b) Show that the compact real n-dimensional torus (S
1
)
n
can be viewed as a subgroup of
T
N
and that L : T
N
N
R
induces an isomorphism T
N
/(S
1
)
n
N
R
. Hint: Use from
part (a).
(c) Let be a fan in N. Show that the action of the compact real torus (S
1
)
n
T
N
on T
N
extends to an action on the toric variety X

and that the quotient space


(X

)/(S
1
)
n

N()
R
,
where

= denotes homeomorphism of topological spaces, and the union is over all
cones in the fan. Hint: Use the Orbit-Cone Correspondence (Theorem 3.2.6).
(d) Let in R
2
be the fan from Example 3.1.9, so that X

P
2
. Show that under the
action of (S
1
)
2
(C

)
2
as in part (c), P
2
/(S
1
)
2
=
2
, the 2-dimensional simplex.
We will say more about the topology of toric varieties in Chapter 12.
3.4.3. This exercise will complete the proof of Theorem 3.4.1. Let Hom(

M, C) be
the set of semigroup homomorphisms

M C. Assume that
k
Hom(

M, C) is
a sequence such that [
k
(m)[ 1 for all m

M and all k. We want to show that there


is a subsequence
k

that converges to a point Hom(

M, C).
(a) The semigroup S

M is generated by a nite set m


1
, . . . , m
s
. Use this fact
and the compactness of z C [ [z[ 1 to show that there exists a subsequence
k

such that the sequences


k

(m
j
) converge in C for all j.
(b) Deduce that the subsequence
k

converges to a Hom(

M, C).
3.4.4. Prove Proposition 3.4.2. Hint: For (b) (a), let Z be the one-point compactication
of X and consider the projection of =(x, x) [ x X X Z.
3.4.5. Show that any projective variety is complete according to Denition 3.4.10.
3.4.6. Here you will prove some characterizations of properness stated in the text.
(a) Prove (a) (c) from Proposition 3.4.8.
(b) Prove Proposition 3.4.9. Hint: First show that if f is proper, then so is
Z
: X
Y
Z Z
for any morphism Z Y. Then use (a) (b) of Proposition 3.4.8, which does not
require rst countable. If f : X Y is universally closed, then prove that f
1
(y) y
is universally closed for any y Y. Then use Proposition 3.4.2 and (a) (b) of
Proposition 3.4.8.
150 Chapter 3. Normal Toric Varieties
3.4.7. Prove that X is complete if and only X pt is proper, and that if X is complete,
then
Z
: X Z Z is proper for any variety Z.
3.4.8. Complete the proof of (b) (c) of Theorem 3.4.11 begun in the text.
3.4.9. Let : T
N
T
N
be a map of tori corresponding to an injective homomorphism
: N N

. Also let

: M

M be the dual map. Finally, let


k
T
N
be a sequence such
that (
k
) converges to a point of T
N
.
(a) Prove that im(

) M has nite index. Hence we can pick an integer d > 0 such that
dM im(

).
(b) Show that
m
(
k
) converges for all m im(

). Conclude that
m
(
d
k
) converges for
all m M, where d is as in part (a).
(c) Pick a basis of M so that T
N
(C

)
n
and write
k
= (
1,k
, . . . ,
n,k
) (C

)
n
. Show
that (
d
1,k
, . . . ,
d
n,k
) converges to a point (
1
, . . . ,
n
) (C

)
n
.
(d) Show that the dth roots
1/d
i
can be chosen so that a subsequence of the sequence

k
= (
1,k
, . . . ,
n,k
) converges to a point = (
1/d
1
, . . . ,
1/d
n
) T
N
.
(e) Explain why this implies that T
N
T
N
is proper in the classical topology.
3.4.10. To nish the proof of (d) (a) of Theorem 3.4.11, suppose we have a toric mor-
phism : X

and a cone . Let

be the smallest cone containing


R
().
(a) Prove that induces a homomorphism

: N() N(

).
(b) Assume further that
1
R
([

[) =[[. Prove that (

)
1
R
([Star(

)[) =[Star()[.
3.4.11. Let

=0 be a strongly convex polyhedral cone in N

R
. Prove that
u

'm

, u

` 0 for all m

M` (

M
and then apply this to u

=
R
(u) to complete the argument in the text. Hint: To prove
, rst show that the right hand side of the equivalence implies that 'm

, u

` 0 for all
m

M
Q
` (

M
Q
. Then show that

= 0 implies that any element of


(

M is a limit of elements in (

M
Q
` (

M
Q
.
3.4.12. Give a second argument for the implication
X

compact complete
from part (c) of Theorem 3.1.19 using induction on the dimension n of N. Hint: If is not
complete and n > 1, then there is a 1-dimensional cone in the boundary of the support of
. Consider the fan Star() and the corresponding toric subvariety of X

.
3.4.13. Let

, be fans in N
R
compatible with the identity map N N. Prove that the
toric morphism : X

is proper if and only if

is a renement of .
Appendix: Nonnormal Toric Varieties
In this appendix, we discuss toric varieties that are not necessarily normal. We begin with
an example to show that Sumihiros theorem (Theorem 3.1.7) on the existence of a torus-
invariant afne open cover can fail in the nonnormal case.
Appendix: Nonnormal Toric Varieties 151
Example 3.A.1. Consider the nodal cubic C P
2
dened by y
2
z = x
2
(x +z). The only
singularity of C is p = (0, 0, 1). We claim that C is a toric variety with C` p C

as
torus. Assuming this for the moment, consider a torus-invariant neighborhood of p. It
contains p and the torus and hence is the whole curve! We conclude that p has no torus-
invariant afne open neighborhood. Thus Sumihiros theorem fails for C.
To see that C is a toric variety, we begin with the standard parametrization obtained
by intersecting lines y = t x with the afne curve y
2
= x
2
(x +1). This easily leads to the
parametrization
x =t
2
1, y =t(t
2
1).
The values t = 1 map to the singular point p. To get a parametrization that looks more
like a torus, we replace t with
t+1
t1
to obtain
x =
4t
(t 1)
2
, y =
4t(t +1)
(t 1)
3
.
Then t = 0, map to p and t C

maps bijectively to C` p.
Using this parametrization, we get C

C, and the action of C

on itself given by
multiplication extends to an action on C by making p a xed point of the action. With
some work, one can show that this action is algebraic and hence gives a toric variety. (For
readers familiar with elliptic curves, the basic idea is that the description of the group law
in terms of lines connecting points on the curve reduces to multiplication in C

C for
our curve C.)
In contrast, the projective toric varieties constructed in Chapter 2 satisfy Sumihiros
theorem by Proposition 2.1.8. Since these nonnormal toric varieties have a good local
structure, it is reasonable to expect that they share some of the nice properties of nor-
mal toric varieties. In particular, they satisfy a version of the Orbit-Cone Correspondence
(Theorem 3.2.6).
We begin with the afne case. Given M and a nite subset A = m
1
, . . . , m
s
M,
we get the afne toric variety Y
A
C
s
whose torus has character group ZA (Proposi-
tion 1.1.8). Assume M = ZA and let N
R
be dual to Cone(A) M
R
. By Proposi-
tion 1.3.8, the normalization of Y
A
is the map
U

Y
A
induced by the inclusion of semigroup algebras
C[NA] C[

M].
Recall that C[

M] is the integral closure of C[NA] in its eld of fractions. We now


apply standard results in commutative algebra and algebraic geometry:
Since the integral closure C[

M] is a nitely generated C-algebra, it is a nitely


generated module over C[NA] (see [10, Cor. 5.8]).
Thus the corresponding morphismU

Y
A
is nite as dened in [131, p. 84].
A nite morphism is proper with nite bers (see [131, Ex. II.3.5 and II.4.1]).
Since U

Y
A
is the identity on the torus, the image of the normalization is Zariski dense
in Y
A
. But the image is also closed since the normalization map is proper. This proves that
the normalization map is onto.
Here is an example of how the normalization map can fail to be one-to-one.
152 Chapter 3. Normal Toric Varieties
Example 3.A.2. The set A =e
1
, e
1
+e
2
, 2e
2
Z
2
gives the parametrization
A
(s, t) =
(s, st, t
2
), and one can check that
Y
A
= V(y
2
x
2
z) C
3
.
Furthermore, ZA = Z
2
and = Cone(A)

= Cone(e
1
, e
2
). It follows easily that the
normalization is given by
C
2
Y
A
(s, t) (s, st, t
2
).
This map is one-to-one on the torus (the torus of Y
A
is normal and hence is unchanged
under normalization) but not on the t-axis, since here the map is (0, t) (0, 0, t
2
). We will
soon see the intrinsic reason why this happens.
We now determine the orbit structure of Y
A
.
Theorem 3.A.3. Let Y
A
be an afne toric variety with M = ZA and let N
R
be as
above. Then:
(a) There is a bijective correspondence
faces of T
N
-orbits in Y
A

such that a face of of dimension k corresponds to an orbit of dimension dimY


A
k.
(b) If O

Y
A
is the orbit corresponding to a face of , then O

is the torus with


character group Z(

A).
(c) The normalization U

Y
A
induces a bijection
T
N
-orbits in U

T
N
-orbits in Y
A

such that if OU

and O

Y
A
are the orbits corresponding to a face of , then the
induced map O O

is the map of tori corresponding to the inclusion Z(

A)

M of character groups.
Proof. We will sketch the main ideas and leave the details for the reader. The proof uses
the Orbit-Cone Correspondence (Theorem3.2.6). We regard points of U

and Y
A
as semi-
group homomorphisms, so that :

M C in U

maps to [
NA
: NA C in Y
A
.
Note also that U

Y
A
is equivariant with respect to the action of T
N
.
By Lemma 3.2.5, the orbit O() U

corresponding to a face of is the torus


consisting of homomorphisms :

M C

. Thus

M is the character group of


O(). The normalization maps this orbit onto an orbit O

() Y
A
, where a point of
O() maps to its restriction to NA. Since
(

M) ZA =

ZA =Z(

A),
it follows that Z(

A) is the character group of O

(). This proves part (b), and the


nal assertion of part (c) follows easily.
Since

M is the saturation of NA, it follows that there is an integer d > 0 such


that d

M NA. It follows easily that Z(

A) has nite index in

M, so that
dim O

() = dim O() = dimU

dim = dimY
A
dim,
proving the nal assertion of part (a).
Appendix: Nonnormal Toric Varieties 153
Finally, every orbit in Y
A
comes from an orbit in U

since U

Y
A
is onto. If orbits
O(
1
), O(
2
) map to the same orbit of Y
A
, then
Z(

1
A) =Z(

2
A).
This implies

1
=

2
, so that
1
=
2
. The bijections in parts (a) and (c) now follow.
We leave it to the reader to work out other aspects of the Orbit-Cone Correspondence
(specically, the analogs of parts (c) and (d) of Theorem 3.2.6) for Y
A
.
Let us apply Theorem 3.A.3 to our previous example.
Example 3.A.4. Let A = e
1
, e
1
+e
2
, 2e
2
Z
2
as in Example 3.A.2. The cone =
Cone(A)

= Cone(e
1
, e
2
) has a face such that

= Span(e
2
). Thus
Z(

A) =Z(2e
2
)

M =Ze
2
.
It follows that Z(

A) has index 2 in

M, which explains why the normalization


map is two-to-one on the orbit corresponding to .
We now turn to the projective case. Here, A =m
1
, . . . , m
s
M gives the projective
toric variety X
A
P
s1
whose torus has character group Z

A (Proposition 2.1.6). Recall


that Z

A =

s
i=1
a
i
m
i
[ a
i
Z,

s
i=1
a
i
= 0

.
One observation is that translating A by m M leaves the corresponding projective
variety unchanged. In other words, X
m+A
= X
A
(see part (a) of Exercise 2.1.6). Thus, by
translating an element of A to the origin, we may assume 0 A. Note that the torus of
X
A
has character lattice Z

A =ZA when 0 A.
We dened the normalization of an afne variety in 1.0. Using a gluing construction,
one can dene the normalization of any variety (see [131, Ex. II.3.8]). We can describe the
normalization of a projective toric variety X
A
as follows.
Theorem 3.A.5. Let X
A
be a projective toric variety where 0 A and M = ZA. If
P = Conv(A) M
R
, then the normalization of X
A
is the toric variety X
P
of the normal
fan of P with respect to the lattice N = Hom
Z
(M, Z).
Proof. Again, we sketch the proof and leave the details to the reader. We use the local
description of X
A
given in Propositions 2.1.8 and 2.1.9. There, we saw that X
A
has an
afne open covering given by the afne toric varieties Y
Av
= Spec(NA
v
), where v A is
a vertex of P = Conv(A) and A
v
=A v =mv [ m A.
For the moment, assume that P is very ample. Then Theorem2.3.1 implies that X
P
has
an afne open cover given by the afne toric varieties U
v
= Spec(

v
M), where v A
is a vertex of P and

v
= Cone(PMv). One can check that

v
M is the saturation of
NA
v
, so that U
v
is the normalization of Y
Av
. The gluings are also compatible by equations
(2.1.6), (2.1.7) and Proposition 2.3.13. It follows that we get a natural map X
P
X
A
that
is the normalization of X
A
.
In the general case, we note that k
0
P is very ample for some integer k
0
1 and that
P and k
0
P have the same normal fan. Since
v
is a maximal cone of the normal fan, the
above argument now applies in general, and the theorem is proved.
154 Chapter 3. Normal Toric Varieties
Combining this result with the Orbit-Cone Correspondence and Theorem 3.A.3 gives
the following immediate corollary.
Corollary 3.A.6. With the same hypotheses as Theorem 3.A.5, we have:
(a) There is a bijective correspondence
cones of
P
T
N
-orbits in X
A

such that a cone of dimension k corresponds to an orbit of dimension dim X


A
k.
(b) If O

X
A
is the orbit corresponding to a cone of
P
, then O

is the torus with


character group Z(

A).
(c) The normalization X
P
X
A
induces a bijection
T
N
-orbits in X
P
T
N
-orbits in X
A

such that if O X
P
and O

X
A
are the orbits corresponding to
P
, then the
induced map O O

is the map of tori corresponding to the inclusion Z(

A)

M of character groups.
We leave it to the reader to work out other aspects of the Orbit-Cone Correspondence
for X
A
. A different approach to the study of X
A
appears in [113, Ch. 5].
Chapter 4
Divisors on Toric Varieties
4.0. Background: Valuations, Divisors and Sheaves
Divisors are dened in terms of irreducible codimension one subvarieties. In this
chapter, we will consider Weil divisors and Cartier divisors. These classes coincide
on a smooth variety, but for a normal variety, the situation is more complicated. We
will also study divisor classes, which are dened using the order of vanishing of
a rational function on an irreducible divisor. We will see that normal varieties are
the natural setting to develop a theory of divisors and divisor classes.
First, we give a simple motivational example.
Example 4.0.1. If f (x) C(x) is nonzero, then there is a unique n Z such that
f (x) = x
n
g(x)
h(x)
, where g(x), h(x) C[x] are not divisible by x. This works because
C[x] is a UFD. The integer n describes the behavior of f (x) at 0: if n > 0, f (x)
vanishes to order n at 0, and if n <0, f (x) has a pole of order [n[ at 0. Furthermore,
the map from the multiplicative group C(x)

to the additive group Z dened by


f (x) n is easily seen to be a group homomorphism. This works in the same way
if we replace 0 with any point of C.
Discrete Valuation Rings. The simple construction given in Example 4.0.1 applies
in far greater generality. We begin by reviewing the algebraic machinery we will
need.
Denition 4.0.2. A discrete valuation on a eld K is a group homomorphism
: K

Z
that is onto and satises (x+y) min((x), (y)) when x, y, x+y K

=K`0.
Note also that (xy) = (x) +(y). The corresponding discrete valuation ring is
R =x K

[ (x) 00.
155
156 Chapter 4. Divisors on Toric Varieties
One can check that a DVR is indeed a ring. Here are some properties of DVRs.
Proposition 4.0.3. Let R be a DVR with valuation : K

Z. Then:
(a) x R is invertible in R if and only if (x) = 0.
(b) R is a local ring with maximal ideal m =x R [ (x) > 00.
(c) R is normal.
(d) R is a principal ideal domain (PID).
(e) R is Noetherian.
(f) The only proper prime ideals of R are 0 and m.
Proof. First observe that since is a homomorphism, we have
(4.0.1) (x
1
) =(x)
for all x K

. If x R is a unit, then (x), (x


1
) 0 since x, x
1
R. Thus
(x) = 0 by (4.0.1). Conversely, if (x) = 0, then (x
1
) = 0 by (4.0.1), so that
x
1
R. This proves part (a).
For part (b), note that m = x R [ (x) > 0 0 is an ideal of R (this
follows directly from Denition 4.0.2). Then part (a) easily implies that R is local
with maximal ideal m (Exercise 4.0.1).
To prove part (c), suppose x K

= K `0 satises
x
n
+r
n1
x
n1
+ +r
0
= 0,
with r
i
R. If x R, we are done, so suppose x / R. Then n > 1 and (x) < 0.
Using (4.0.1) again, we see that x
1
R. So x
1n
= (x
1
)
n1
R and hence
x
1n
(x
n
+r
n1
x
n1
+ +r
0
) = 0,
showing that x =(r
n1
+r
n2
x
1
+ +r
0
x
1n
) R.
Let R satisfy () = 1 and let I = 0 be an ideal of R. Pick x I `0
with k = (x) minimal. Then y = x
k
K satises (y) = (x) k() = 0, so
that y is invertible in R. From here, one proves without difculty that I = '
k
`.
This proves part (d), and part (e) follows immediately.
For part (f), it is obvious that 0 and the maximal ideal m are prime. Note
also that m = '`. Now let P = 0 be a proper prime ideal. By the previous
paragraph, P = '
k
` for some k > 0. If k > 1, then
k1
P and ,
k1
/ P
give a contradiction.
This shows that every DVR is a Noetherian local domain of dimension one.
In general, the dimension dim R of a Noetherian ring R is one less than the length
of the longest chain P
0
P
d
of proper prime ideals contained in R. Among
Noetherian local domains of dimension one, DVRs are characterized as follows.
4.0. Background: Valuations, Divisors and Sheaves 157
Theorem 4.0.4. If (R, m) is a Noetherian local domain of dimension one, then the
following are equivalent:
(a) R is a DVR.
(b) R is normal.
(c) m is principal.
(d) (R, m) is a regular local ring.
Proof. The implications (a) (b) and (a) (c) follow from Proposition 4.0.3, and
the equivalence (c) (d) is covered in Exercise 4.0.2. The remaining implications
can be found in [10, Prop. 9.2].
DVRs and Prime Divisors. DVRs have a natural geometric interpretation. Let X
be an irreducible variety. A prime divisor D X is an irreducible subvariety of
codimension one, meaning that dim D = dim X 1. Recall from 3.0 that X has
a eld of rational functions C(X). Our goal is to dene a ring O
X,D
with eld
of fractions C(X) such that O
X,D
is a DVR when X is normal. This will give
a valuation
D
: C(X)

Z such that for f C(X)

,
D
( f ) gives the order of
vanishing of f along D.
Denition 4.0.5. For a variety X and prime divisor D X, O
X,D
is the subring of
C(X) dened by
O
X,D
= C(X) [ is dened on U X open with U D=.
We will see below that O
X,D
is a ring. Intuitively, this ring is built from rational
functions on X that are dened somewhere on D (and hence dened on most of D
since D is irreducible).
Since X is irreducible, Exercise 3.0.4 implies that C(X) = C(U) whenever
U X is open and nonempty. If we further assume that U D is nonempty, then
(4.0.2) O
X,D
=O
U,UD
follows easily (Exercise 4.0.3).
Hence we can reduce to the afne case X = Spec(R) for an integral domain
R. The codimension of a prime ideal p, also called its height, is dened to be
codimp = dim Rdim V(p). It follows easily that p V(p) induces a bijection
codimension one prime ideals of R prime divisors of X.
Given a prime divisor D = V(p), we can interpret O
X,D
in terms of R as follows.
The eld of rational functions C(X) is the eld of fractions K of R, and a rational
function = f /g K, f , g R, is dened somewhere on D=V(p) precisely when
g / I(D) = p. It follows that
O
X,D
=f /g K [ f , g R, g / p,
158 Chapter 4. Divisors on Toric Varieties
which is the localization R
p
of R at the multiplicative subset R`p (note that R`p
is closed under multiplication because p is prime). This localization is a local ring
with maximal ideal pR
p
(Exercise 4.0.3). It follows that
(4.0.3) O
X,D
= R
p
when X = Spec(R) and p is a codimension one prime ideal of R.
Example 4.0.6. In Example 4.0.1, we constructed a discrete valuation on C(x) by
sending f (x) C(x)

to n Z, provided
f (x) = x
n
g(x)
h(x)
, g(x), h(x) C[x], g(0) = 0, h(0) = 0.
The corresponding DVR is the localization C[x]
/x)
. It follows that the prime divisor
0 = V(x) C = Spec(C[x]) has the local ring
O
C,|0
=C[x]
/x)
which is a DVR.
More generally, a normal ring or variety gives a DVR as follows.
Proposition 4.0.7.
(a) Let R be a normal domain and p R be a codimension one prime ideal. Then
the localization R
p
is a DVR.
(b) Let X be a normal variety and DX a prime divisor. Then the local ring O
X,D
is a DVR.
Proof. By Proposition 3.0.12, part (b) follows immediately from part (a) together
with (4.0.2) and (4.0.3).
It remains to prove part (a). The maximal ideal of R
p
is the ideal m
p
= pR
p
generated by p in R
p
. The localization of a Noetherian ring is Noetherian (Exer-
cise 4.0.4), and the same is true for normality by Exercise 1.0.7. It follows that the
local domain (R
p
, m
p
) is Noetherian and normal.
We compute the dimension of R
p
as follows. Since dim X = dim R (see [69,
Ex. 17 and 18 of Ch. 9, 4]), our hypothesis on D = V(p) implies that there are no
prime ideals strictly between 0 and p in R. By [10, Prop. 3.11], the same is true
for 0 and m
p
in R
p
. It follows that R
p
has dimension one. Then R
p
is a DVR by
Theorem 4.0.4.
When D is a prime divisor on a normal variety X, the DVR O
X,D
means that
we have a discrete valuation

D
: C(X)

Z,
where O
X,D
consists of 0 and those nonzero rational functions satisfying
D
( f ) 0.
Given f O
X,D
`0, we call
D
( f ) the order of vanishing of f along the divisor
D. Thus the maximal ideal m
X,D
O
X,D
consists of 0 and those rational functions
4.0. Background: Valuations, Divisors and Sheaves 159
that vanish on D. When f C(X)

satises
D
( f ) = < 0, we say that f has a
pole of order [[ along D.
Weil Divisors. Recall that a prime divisor on an irreducible variety X is an irre-
ducible subvariety of codimension one.
Denition 4.0.8. Div(X) is the free abelian group generated by the prime divisors
on X. A Weil divisor is an element of Div(X).
Thus a Weil divisor DDiv(X) is a nite sum D=

i
a
i
D
i
Div(X) of prime
divisors D
i
with a
i
Z for all i. The divisor D is effective, written D 0, if the a
i
are all nonnegative. The support of D is the union of the prime divisors appearing
in D:
Supp(D) =

a
i
,=0
D
i
.
The Divisor of a Rational Function. An important class of Weil divisors comes
from rational functions. If X is normal, any prime divisor D on X corresponds to a
DVR O
X,D
with valuation
D
: C(X)

Z. Given f C(X)

, the integers
D
( f )
tell us how f behaves on the prime divisors of X. Here is an important property of
these integers.
Lemma 4.0.9. If X is normal and f C(X)

, then
D
( f ) is zero for all but a nite
number of prime divisors D X.
Proof. If f is constant, then it is a nonzero constant since f C(X)

. It follows
that
D
( f ) = 0 for all D. On the other hand, if f is nonconstant, then we can nd
a nonempty open subset U X such that f : U C is a nonconstant morphism.
Then V = f
1
(C

) is a nonempty open subset of X such that f [


V
: V C

. The
complement X `V is Zariski closed and hence is a union of irreducible compo-
nents of dimension < n. Denote the irreducible components of codimension one
by D
1
, . . . , D
s
.
Now let D be prime divisor in X. If V D = , then D X `V, so that D is
contained in an irreducible component of X `V since D is irreducible. Dimension
considerations imply that D = D
i
for some i. On the other hand, if V D=, then
f is an invertible element of O
X,D
=O
V,VD
, which implies that
D
( f ) = 0.
Denition 4.0.10. Let X be a normal variety.
(a) The divisor of f C(X)

is
div( f ) =

D
( f )D,
where the sum is over all prime divisors DX.
(b) div( f ) is called a principal divisor, and the set of all principal divisors is de-
noted Div
0
(X).
160 Chapter 4. Divisors on Toric Varieties
(c) Divisors D and E are linearly equivalent, written D E, if their difference is
a principal divisor, i.e., DE = div( f ) Div
0
(X) for some f C(X)

.
Lemma 4.0.9 implies that div( f ) Div(X). If f , g C(X)

, then div( f g) =
div( f ) + div(g) and div( f
1
) = div( f ) since valuations are group homomor-
phisms on C(X)

. It follows that Div


0
(X) is a subgroup of Div(X).
Example 4.0.11. Let f = c(x a
1
)
m
1
(x a
r
)
mr
C[x] be a polynomial of de-
gree m > 0, where c C

and a
1
, . . . , a
r
C are distinct. Then:
When X =C, div( f ) =

r
i=1
m
i
a
i
.
When X =P
1
=C, div( f ) =

r
i=1
m
i
a
i
m.
The divisor of f C(X)

can be written div( f ) = div


0
( f ) div

( f ), where
div
0
( f ) =

D
( f )>0

D
( f )D
div

( f ) =

D
( f )<0

D
( f )D.
We call div
0
( f ) the divisor of zeros of f and div

( f ) the divisor of poles of f .


Note that these are effective divisors.
Cartier Divisors. If D =

i
a
i
D
i
is a Weil divisor on X and U X is a nonempty
open subset, then
D[
U
=

UD
i
,=
a
i
U D
i
is a Weil divisor on U called the restriction of D to U.
We now dene a special class of Weil divisors.
Denition 4.0.12. A Weil divisor D on a normal variety X is Cartier if it is locally
principal, meaning that X has an open cover U
i

iI
such that D[
U
i
is principal in
U
i
for every i I. If D[
U
i
= div( f
i
)[
U
i
for i I, then we call (U
i
, f
i
)
iI
the local
data for D.
A principal divisor is obviously locally principal. Thus div( f ) is Cartier for all
f C(X)

. One can also show that if D and E are Cartier divisors, then D+E and
D are Cartier (Exercise 4.0.5). It follows that the Cartier divisors on X form a
group CDiv(X) satisfying
Div
0
(X) CDiv(X) Div(X).
Divisor Classes. For Weil and Cartier divisors, linear equivalence classes form the
following important groups.
4.0. Background: Valuations, Divisors and Sheaves 161
Denition 4.0.13. Let X be a normal variety. Its class group is
Cl(X) = Div(X)/Div
0
(X),
and its Picard group is
Pic(X) = CDiv(X)/Div
0
(X).
We will give a more sophisticated denition of Pic(X) in Chapter 6. Note that
since CDiv(X) is a subgroup of Div(X), we get a canonical injection
Pic(X) Cl(X).
In [131, II.6], Hartshorne writes The divisor class group of a scheme is a very
interesting invariant. In general it is not easy to calculate. Fortunately, divisor
class groups of normal toric varieties are easy to describe, as we will see in 4.1.
More Algebra. Before we can derive further properties of divisors, we need to
learn more about normal domains. Equation (3.0.2) shows that if X = Spec(R) is
irreducible, then
R =

pX
O
X,p
.
If a point p X corresponds to a maximal ideal m R, then the local ring O
X,p
is
the localization R
m
. Hence the above equality can be written
R =

m maximal
R
m
.
When R is normal, we get a similar result using codimension one prime ideals.
Theorem 4.0.14. If R is a Noetherian normal domain, then
R =

codimp=1
R
p
.
Proof. Let K be the eld of fractions of R and assume that a/b K, a, b R, lies
in R
p
for all codimension one prime ideals p. It sufces to prove that a 'b`. This
is obviously true when b is invertible in R, so we may assume that 'b` is a proper
ideal of R. Then we have a primary decomposition (see [69, Ch. 4, 7])
(4.0.4) 'b` = q
1
q
s
,
and each prime ideal p
i
=

q
i
is of the form p
i
= 'b` : c
i
for some c
i
R. In the
terminology of [195, p. 38], the p
i
are the prime divisors of 'b`.
Since R is Noetherian and normal, the Krull principal ideal theorem states that
every prime divisor of 'b` has codimension one (see [195, Thm. 11.5] for a proof).
This implies that in the primary decomposition (4.0.4), the prime divisors p
i
have
codimension one and hence are distinct.
162 Chapter 4. Divisors on Toric Varieties
Note that a/b R
p
i
for all i by our assumption on a/b. This implies a bR
p
i
.
Since (q
j
)
p
i
= R
p
i
for j = i (Exercise 4.0.6), localizing (4.0.4) at p
i
shows that for
all i, we have
a bR
p
i
= q
i
R
p
i
.
Since q
i
R
p
i
R = q
i
(Exercise 4.0.6), we obtain a

s
i=1
q
i
='b`.
This result has the following useful corollary.
Corollary 4.0.15. Let X be a normal variety and let f : U C be a morphism
dened on an open set U X. If X `U has codimension 2 in X, then f extends
to a morphism dened on all of X.
Proof. Since X has an afne open cover, we can assume that X = Spec(R), where
R is a Noetherian normal domain. If D X is a prime divisor, then U D = for
dimension reasons. It follows that f O
U,UD
=O
X,D
, so that
(4.0.5) f

D
O
U,UD
=

D
O
X,D
=

codimp=1
R
p
= R,
where the nal equality is Theorem 4.0.14.
These results enable us to determine when the divisor of a rational function is
effective.
Proposition 4.0.16. Let X be a normal variety. If f C(X)

, then:
(a) div( f ) 0 if and only if f : X C is a morphism, i.e., f O
X
(X).
(b) div( f ) = 0 if and only if f : X C

is a morphism, i.e., f O

X
(X).
In general, O

X
is the sheaf on X dened by
O

X
(U) =invertible elements of O
X
(U).
This is a sheaf of abelian groups under multiplication.
Proof. If f : X Cis a morphism, then f O
X,D
for every prime divisor D, which
in turn implies
D
( f ) 0. Hence div( f ) 0. Going the other way, suppose that
div( f ) 0. This remains true when we restrict to an afne open subset, so we may
assume that X is afne. Then div( f ) 0 implies
f

D
O
X,D
,
where the intersection is over all prime divisors. By (4.0.5), we conclude that f is
dened everywhere. This proves part (a), and part (b) follows immediately since
div( f ) = 0 if and only if div( f ) 0 and div( f
1
) 0.
4.0. Background: Valuations, Divisors and Sheaves 163
Singularities and Normality. The set of singular points of a variety X is denoted
Sing(X) X.
We call Sing(X) the singular locus of X. One can show that Sing(X) is a proper
closed subvariety of X (see [131, Thm. I.5.3]). When X is normal, things are even
nicer.
Proposition 4.0.17. Let X be a normal variety. Then:
(a) Sing(X) has codimension 2 in X.
(b) If X is a curve, then X is smooth.
Proof. You will prove part (b) in Exercise 4.0.7. A proof of part (a) can be found
in [245, Vol. 2, Thm. 3 of II.5].
Computing Divisor Classes. There are two results, one algebraic and one geomet-
ric, that enable us to compute class groups in some cases.
We begin with the algebraic result.
Theorem 4.0.18. Let R be a UFD and set X = Spec(R). Then:
(a) R is normal and every codimension one prime ideal is principal.
(b) Cl(X) = 0.
Proof. For part (a), we know that a UFD is normal by Exercise 1.0.5. Let p be a
codimension one prime ideal of R and pick a p`0. Since R is a UFD,
a = c
s

i=1
p
a
i
i
,
with the p
i
prime and c is invertible in R. Because p is prime, this means some
p
i
p, and since codimp = 1, this forces p ='p
i
`.
Turning to part (b), let D X be a prime divisor. Then p = I(D) is a codi-
mension one prime ideal and hence is principal, say p =' f `. Then f generates the
maximal ideal of the DVR R
p
, which implies
D
( f ) = 1 (see the proof of Propo-
sition 4.0.3). It follows easily that div( f ) = D. Then Cl(X) = 0 since all prime
divisors are linearly equivalent to 0.
In fact, more is true: a normal Noetherian domain is a UFD if and only if every
codimension one prime ideal is principal (Exercise 4.0.8).
Example 4.0.19. C[x
1
, . . . , x
n
] is a UFD, so Cl(C
n
) = 0 by Theorem 4.0.18.
Before stating the geometric result, note that if U X is open and nonempty,
then restriction of divisors D D[
U
induces a well-dened map Cl(X) Cl(U)
(Exercise 4.0.9).
164 Chapter 4. Divisors on Toric Varieties
Theorem 4.0.20. Let U be a nonempty open subset of a normal variety X and let
D
1
, . . . , D
s
be the irreducible components of X `U that are prime divisors. Then
the sequence
s

j=1
ZD
j
Cl(X) Cl(U) 0
is exact, where the rst map sends

s
j=1
a
j
D
j
to its divisor class in Cl(X) and the
second is induced by restriction to U.
Proof. Let D

i
a
i
D

i
Div(U) with D

i
a prime divisor in U. Then the Zariski
closure D

i
of D

i
in X is a prime divisor in X, and D =

i
a
i
D

i
satises D[
U
= D

.
Hence Cl(X) Cl(U) is surjective.
Since each D
j
restricts to 0 in Div(U), the composition of the two maps is
trivial. To nish the proof of exactness, suppose that [D] Cl(X) restricts to 0 in
Cl(U). This means that D[
U
is the divisor of some f C(U)

. Since C(U) =C(X)


and the divisor of f in Div(X) restricts to the divisor of f in Div(U), it follows that
we have f C(X)

such that
D[
U
= div( f )[
U
.
This implies that the difference Ddiv( f ) is supported on X `U, which means
that Ddiv( f )

s
j=1
ZD
j
by the denition of the D
j
.
Example 4.0.21. Write P
1
=C and note that is a prime divisor on P
1
.
Then Theorem 4.0.20 and Example 4.0.19 give the exact sequence
Z Cl(P
1
) Cl(C) = 0.
Hence the map Z Cl(P
1
) dened by a [a] is surjective. This map is
injective since a = div( f ) implies div( f )[
C
= 0, so that f (C, O
C
)

=C

by Proposition 4.0.16. Hence f is constant, which forces a = 0. If follows that


Cl(P
1
) Z.
Later in the chapter we will use similar methods to compute the class group of
an arbitrary normal toric variety.
Comparing Weil and Cartier Divisors. Once we understand Cartier divisors on
normal toric varieties, it will be easy to give examples of Weil divisors that are not
Cartier. On the other hand, there are varieties where every Weil divisor is Cartier.
Theorem 4.0.22. Let X be a normal variety. Then:
(a) If the local ring O
X,p
is a UFD for every p X, then every Weil divisor on X
is Cartier.
(b) If X is smooth, then every Weil divisor on X is Cartier.
4.0. Background: Valuations, Divisors and Sheaves 165
Proof. If X is smooth, then O
X,p
is a regular local ring for all p X. Since every
regular local ring is a UFD (see 1.0), part (b) follows from part (a).
For part (a), it sufces to show that prime divisors are locally principal. This
condition is obviously local on X, so we may assume that X = Spec(R) is afne.
Let D = V(p) be a prime divisor on X, where p R is a codimension one prime
ideal. Note that D is obviously principal on U = X `D since D[
U
= 0. It remains
to show that D is locally principal in a neighborhood of a point p D.
The point p corresponds to a maximal ideal mR. Thus p D implies p m.
Since p R has codimension one, it follows that the prime ideal pR
m
R
m
also has
codimension one (this follows from [10, Prop. 3.11]). Then Theorem 4.0.18 im-
plies that pR
m
is principal since R
m
is a UFD by hypothesis. Thus pR
m
= (a/b)R
m
where a, b R and b / m. Since b is invertible in R
m
, we in fact have pR
m
= aR
m
.
Now suppose p ='a
1
, . . . , a
s
` R. Then a
i
pR
m
=aR
m
, so that a
i
=(g
i
/h
i
)a,
where g
i
, h
i
R and h
i
/ m, i.e., h
i
(p) = 0. If we set h = h
1
h
s
, then pR
h
= aR
h
follows easily. Then U = Spec(R
h
) is a neighborhood of p, and from here, it is
straightforward to see that D = div(a) on U.
Example 4.0.23. Since P
1
is smooth, Theorem 4.0.22 and Example 4.0.21 imply
that Pic(P
1
) = Cl(P
1
) Z.
Sheaves of O
X
-modules. Weil and Cartier divisors on X lead to some important
sheaves on X. Hence we need a brief excursion into sheaf theory (we will go deeper
into the subject in Chapter 6). The sheaf O
X
was dened in 3.0. The denition
of a sheaf F of O
X
-modules is similar: for each open subset U X, there is an
O
X
(U)-module F(U) with the following properties:
When U

U, there is a restriction map

U,U

: F(U) F(U

)
such that
U,U
is the identity and
U

,U

U,U
=
U,U
when U

U.
Furthermore,
U,U
is compatible with the restriction map O
X
(U) O
X
(U

).
If U

is an open cover of U X, then the sequence


0 F(U)

F(U

,
F(U

)
is exact, where the second arrow is dened by the restrictions
U,U
and the
double arrow is dened by
U,UU

and
U

,UU

. Exactness means the


same as in 3.0.
When U F(U) satises just the rst bullet, we say that F is a presheaf.
Given a sheaf of O
X
-modules F, elements of F(U) are called sections of F
over U. The module of sections of F over U X is expressed in several ways:
F(U) = (U, F) = H
0
(U, F).
166 Chapter 4. Divisors on Toric Varieties
We will use in this chapter and switch to H
0
in later chapters. Traditionally,
(X, F) is called the module of global sections of F.
Example 4.0.24. Let f : X Y be a morphism of varieties and let F be a sheaf
of O
X
-modules on X. The direct image sheaf f

F on Y is dened by
U F( f
1
(U))
for U Y open. Then f

F is a sheaf of O
Y
-modules. For i : Y X, the direct
image i

O
Y
was mentioned in 3.0.
If F and G are sheaves of O
X
-modules, then a homomorphism of sheaves
: F G consists of O
X
(U)-module homomorphisms

U
: F(U) G(U),
such that the diagram
F(U)

U,V

G(U)

U,V

F(V)

V

G(V)
commutes whenever V U. It should be clear what it means for sheaves F, G of
O
X
-modules to be isomorphic, written F G.
Example 4.0.25. Let f : X Y be a morphism of varieties. If U Y is open, then
composition with f induces a natural map
O
Y
(U) O
X
( f
1
(U)) = f

O
X
(U).
This denes a sheaf homomorphism O
Y
f

O
X
.
Over an afne variety X = Spec(R), there is a standard way to get sheaves of
O
X
-modules. Recall that a nonzero element f R gives the localization R
f
such
that X
f
= Spec(R
f
) is the open subset X `V( f ). Given an R-module M, we get the
R
f
-module M
f
= M
R
R
f
. Then there is a unique sheaf

M of O
X
-modules such
that

M(X
f
) = M
f
for every nonzero f R (see [131, Prop. II.5.1]). This globalizes as follows.
Denition 4.0.26. Let F be a sheaf of O
X
-modules on a variety X.
(a) Let U X be open. Then the restriction F[
U
is the sheaf of O
U
-modules
dened by F[
U
(V) =F(V) for V U open.
(b) F is quasicoherent if X has an afne open cover U

, U

= Spec(R

), such
that for each , there is an R

-module M

satisfying F[
U

.
(c) If in addition each M

is a nitely generated R

-module, then we say that F


is coherent.
4.0. Background: Valuations, Divisors and Sheaves 167
The Sheaf of a Weil Divisor. Let D be a Weil divisor on a normal variety X. We
will show that D determines a sheaf O
X
(D) of O
X
-modules on X. Recall that if
U X is open, then O
X
(U) consists of all morphisms U C. Proposition 4.0.16
tells us that an arbitrary element f C(X)

is a morphism on U if and only if


div( f )[
U
0. It follows that the sheaf O
X
is dened by
U O
X
(U) =f C(X)

[ div( f )[
U
00.
In a similar way, we dene the sheaf O
X
(D) by
U O
X
(D)(U) =f C(X)

[ (div( f ) +D)[
U
00. (4.0.6)
Proposition 4.0.27. Let D be a Weil divisor on a normal variety X. Then the sheaf
O
X
(D) dened in (4.0.6) is a coherent sheaf of O
X
-modules on X.
Proof. In Exercise 4.0.10 you will show that O
X
(D) is a sheaf of O
X
-modules.
The proof is a nice application of the properties of valuations.
To show that O
X
(D) is coherent, we may assume that X = Spec(R). Let K be
the eld of fractions of R. It sufces to prove the following two assertions:
M = (X, O
X
(D)) = f K [ div( f ) +D 0 0 is a nitely generated
R-module.
(X
f
, O
X
(D)) = M
f
for all nonzero f R.
For the rst bullet, we will prove the existence of an element h R`0 such
that h(X, O
X
(D)) R. This will imply that h(X, O
X
(D)) is an ideal of R and
hence has a nite basis since R is Noetherian. It will follow immediately that
(X, O
X
(D)) is a nitely generated R-module.
Write D =

s
i=1
a
i
D
i
. Since supp(D) is a proper subvariety of X, we can nd
g R`0 that vanishes on each D
i
. Then
D
i
(g) >0 for every i, so there is m N
with m
D
i
(g) >a
i
for all i. Since div(g) 0, it follows that mdiv(g)D0. Now
let f (X, O
X
(D)). Then div( f ) +D 0, so that
div(g
m
f ) = mdiv(g) +div( f ) = mdiv(g) D+div( f ) +D0
since a sum of effective divisors is effective. By Proposition 4.0.16, we conclude
that g
m
f O
X
(X) = R. Hence h = g
m
R has the desired property.
To prove the second bullet, observe that M K and f R`0 imply that
M
f
=

g
f
m
[ g (X, O
X
(D)), m0

.
It is also easy to see that M
f
(X
f
, O
X
(D)). For the opposite inclusion, let D =

s
i=1
a
i
D
i
and write 1, . . . , s = I J where D
i
X
f
= for i I and D
j
V( f )
for j J. Given h (X
f
, O
X
(D)), (div(h) +D)[
X
f
0 implies that
D
i
(h) a
i
for i I. There is no constraint on
D
j
(h) for j J, but f vanishes on D
j
for j J,
so that
D
j
( f ) > 0. Hence we can pick m N sufciently large such that
m
D
j
( f ) +
D
j
(h) > 0 for j J.
168 Chapter 4. Divisors on Toric Varieties
Since div( f ) 0, it follows easily that div( f
m
h) +D 0 on X. Thus g = f
m
h
(X, O
X
(D)), and then h = g/f
m
has the desired form.
The sheaves O
X
(D) are more than just coherent; they have the additional prop-
erty of being reexive. Furthermore, when D is Cartier, O
X
(D) is invertible. The
denitions of invertible and reexive will be given in Chapters 6 and 8 respectively.
For now, we give two results about the sheaves O
X
(D). Here is the rst.
Proposition 4.0.28. Distinct prime divisors D
1
, . . . , D
s
on a normal variety X give
the divisor D = D
1
+ +D
s
and the subvariety Y = Supp(D) = D
1
D
s
.
Then O
X
(D) is the ideal sheaf I
Y
of Y, i.e.,
(U, O
X
(D)) =f O
X
(U) [ f vanishes on Y
for all open subsets U X.
Proof. Since sheaves are local, we may assume that X = Spec(R). Then note that
f (X, O
X
(D)) implies div( f )D0, so div( f ) D0 since D is effective.
Thus f R by Proposition 4.0.16 and hence (X, O
X
(D)) is an ideal of R.
Let p
i
= I(D
i
) R be the prime ideal of D
i
. Then, for f R, we have

D
i
( f ) > 0 f p
i
R
p
i
f p
i
,
where the last equivalence uses the easy equality p
i
R
p
i
R = p
i
. Hence div( f ) D
if and only if f vanishes on D
1
, . . . , D
s
, and the proposition follows.
Linear equivalence of divisors tells us the following interesting fact about the
associated sheaves.
Proposition 4.0.29. If D E are linearly equivalent Weil divisors, then O
X
(D)
and O
X
(E) are isomorphic as sheaves of O
X
-modules.
Proof. By assumption, we have D = E +div(g) for some g C(X)

. Then
f (X, O
X
(D)) div( f ) +D 0
div( f ) +E +div(g) 0
div( f g) +E 0
f g (X, O
X
(E)).
Thus multiplication by g induces an isomorphism (X, O
X
(D)) (X, O
X
(E))
which is clearly an isomorphism of (X, O
X
)-modules.
The same argument works over any Zariski open set U, and the isomorphisms
are easily seen to be compatible with the restriction maps.
The converse of Proposition 4.0.29 is also true, i.e., an O
X
-module isomor-
phism O
X
(D) O
X
(E) implies that D E. The proof requires knowing more
about the sheaves O
X
(D) and hence will be postponed until Chapter 8.
4.0. Background: Valuations, Divisors and Sheaves 169
Exercises for 4.0.
4.0.1. Complete the proof of part (b) of Proposition 4.0.3.
4.0.2. Prove (c) (d) in Theorem 4.0.4. Hint: Let m be the maximal ideal of R. Since R
has dimension one, it is regular if and only if m/m
2
has dimension one as a vector space
over R/m. For (d) (c), use Nakayamas Lemma (see [10, Props. 2.6 and 2.8]).
4.0.3. This exercise will study the rings O
X,D
and R
p
.
(a) Prove (4.0.2).
(b) Let p be a prime ideal of a ring R and let R
p
denote the localization of R with respect
to the multiplicative subset R` p. Prove that R
p
is a local ring and that its maximal
ideal is the ideal pR
p
R
p
generated by p.
4.0.4. Let S be a multiplicative subset of a Noetherian ring R. Prove that the localization
R
S
is Noetherian.
4.0.5. Let D and E be Weil divisors on a normal variety.
(a) If D and E are Cartier, show that D+E and D are also Cartier.
(b) If D E, show that D is Cartier if and only if E is Cartier.
4.0.6. Complete the proof of Theorem 4.0.14.
4.0.7. Prove that a normal curve is smooth.
4.0.8. Let R be a Noetherian normal domain. Prove that the following are equivalent:
(a) R is a UFD.
(b) Cl(Spec(R)) = 0.
(c) Every codimension one prime ideal of R is principal.
Hint: For (b) (c), assume that D = div( f ) corresponds to p. Use Theorem 4.0.14 to
show f R and use the Krull principal ideal theorem to show ' f ` is primary in R. Then
pR
p
= f R
p
and [10, Prop. 4.8] imply p = ' f `. For (c) (a), let a R be noninvertible
and let D
1
, . . . , D
s
be the codimension one irreducible components of V(a). If I(D
i
) ='a
i
`,
compare the divisors of a and

s
i=1
a
D
i
(a)
i
using Proposition 4.0.16.
4.0.9. Prove that the restriction map D D[
U
induces a well-dened homomorphism
Cl(X) Cl(U).
4.0.10. Let D be a Weil divisor on a normal variety X. Prove that (4.0.6) denes a sheaf
O
X
(D) of O
X
-modules.
4.0.11. For each of the following rings R, give a careful description of the eld of fractions
K and show that the ring is a DVR by constructing an appropriate discrete valuation on K.
(a) R =a/b Q[ a, b Z, b = 0, gcd(b, p) = 1, where p is a xed prime number.
(b) R =Cz, the ring consisting of all power series in z with coefcients in C that have
a positive radius of convergence.
4.0.12. The plane curve V(x
3
y
2
) C
2
has coordinate ring R = C[x, y]/'x
3
y
2
`. As
noted in Example 1.1.15, this is the coordinate ring of the afne toric variety given by
the afne semigroup S = 0, 2, 3, . . . . This semigroup is not saturated, which means that
R C[S] = C[t
2
, t
3
] is not normal by Theorem 1.3.5. It follows that R is not a DVR by
170 Chapter 4. Divisors on Toric Varieties
Theorem 4.0.4. Give a direct proof of this fact using only the denition of DVR. Hint: The
eld of fractions of C[t
2
, t
3
] is C(t). If C[t
2
, t
3
] comes from the discrete valuation , what
is (t)?
4.0.13. Let X be a normal variety. Use Proposition 4.0.16 to prove that there is an exact
sequence
1 O
X
(X)

C(X)

Div(X) Cl(X) 0,
where the map C(X)

Div(X) is f div( f ) and Div(X) Cl(X) is D[D]. Similarly,


prove that there is an exact sequence
1 O
X
(X)

C(X)

CDiv(X) Pic(X) 0.
4.0.14. Let D=

codimp=1
a
p
D
p
be a Weil divisor on a normal afne variety X =Spec(R).
As usual, let K be the eld of fractions of R. Here you give an algebraic description of
(X, O
X
(D)) in terms of the prime ideals p.
(a) Let p be a codimension one prime of R, so that R
p
is a DVR. Hence the maximal ideal
pR
p
is principal. Use this to dene p
a
R
p
K for all a Z.
(b) Prove that
(X, O
X
(D)) =

codimp=1
p
ap
R
p
.
(c) Now assume that D is effective, i.e., a
p
0 for all p. Prove that (X, O
X
(D)) is the
ideal of R given by
(X, O
X
(D)) =

codimp=1
p
ap
R
p
.
4.0.15. Let R be an integral domain with eld of fractions K. A nitely generated R-
submodule of K is called a fractional ideal. If R is normal and D is a Weil divisor on
X = Spec(R), explain why (X, O
X
(D)) K is a fractional ideal.
4.1. Weil Divisors on Toric Varieties
Let X

be the toric variety of a fan in N


R
with dim N
R
= n. Then X

is normal
of dimension n. We will use torus-invariant prime divisors and characters to give a
lovely description of the class group of X

.
The Divisor of a Character. The order of vanishing of a character along a torus-
invariant prime divisor is determined by the polyhedral geometry of the fan.
By the Orbit-Cone Correspondence (Theorem 3.2.6), k-dimensional cones
of correspond to (n k)-dimensional T
N
-orbits in X

. As in Chapter 3, (1)
is the set of 1-dimensional cones (i.e., the rays) of . Thus (1) gives the
codimension 1 orbit O() whose closure O() is a T
N
-invariant prime divisor on
X

. To emphasize that O() is a divisor we will denote it by D

rather than V().


Then D

= O() gives the DVR O


X

,D
with valuation

=
D
: C(X

Z.
4.1. Weil Divisors on Toric Varieties 171
Recall that the ray (1) has a minimal generator u

N. Also note that


when m M, the character
m
: T
N
C

is a rational function in C(X

since T
N
is Zariski open in X

.
Proposition 4.1.1. Let X

be the toric variety of a fan . If the ray (1) has


minimal generator u

and
m
is character corresponding to m M, then

(
m
) ='m, u

`.
Proof. Since u

N is primitive, we can extend u

to a basis e
1
= u

, e
2
, . . . , e
n
of
N, then we can assume N = Z
n
and = Cone(e
1
) R
n
. By Example 1.2.21, the
corresponding afne toric variety is
U

= Spec(C[x
1
, x
1
2
, . . . , x
1
n
]) =C(C

)
n1
and D

is dened by x
1
= 0. It follows easily that the DVR is
O
X

,D
=O
U,UD
=C[x
1
, . . . , x
n
]
/x
1
)
.
Similar to Example 4.0.6, f C(x
1
, . . . , x
n
)

has valuation

( f ) = Z when
f = x

1
g
h
, g, h C[x
1
, . . . , x
n
] `'x
1
`.
To relate this to

(
m
), note that x
1
, . . . , x
n
are the characters of the dual basis
of e
1
= u

, e
2
, . . . , e
n
N. It follows that given any m M, we have

m
= x
/m,e
1
)
1
x
/m,e
2
)
2
x
/m,en)
n
= x
/m,u)
1
x
/m,e
2
)
2
x
/m,en)
n
.
Comparing this to the previous equation implies that

(
m
) ='m, u

`.
We next compute the divisor of a character. As above, a ray (1) gives:
A minimal generator u

N.
A prime T
N
-invariant divisor D

= O() on X

.
We will use this notation for the remainder of the chapter.
Proposition 4.1.2. For m M, the character
m
is a rational function on X

, and
its divisor is given by
div(
m
) =

(1)
'm, u

`D

.
Proof. The Orbit-Cone Correspondence (Theorem 3.2.6) implies that the D

are
the irreducible components of X ` T
N
. Since
m
is dened and nonzero on T
N
, it
follows that div(
m
) is supported on

(1)
D

. Hence
div(
m
) =

(1)

D
(
m
)D

.
Then we are done since
D
(
m
) ='m, u

` by Proposition 4.1.1.
172 Chapter 4. Divisors on Toric Varieties
Computing the Class Group. Divisors of the form

(1)
a

are precisely the


divisors invariant under the torus action on X

(Exercise 4.1.1). Thus


Div
T
N
(X

) =

(1)
ZD

Div(X

)
is the group of T
N
-invariant Weil divisors on X

. Here is the main result of this


section.
Theorem 4.1.3. We have the exact sequence
M Div
T
N
(X

) Cl(X

) 0,
where the rst map is mdiv(
m
) and the second sends a T
N
-invariant divisor to
its divisor class in Cl(X

). Furthermore, we have a short exact sequence


0 M Div
T
N
(X

) Cl(X

) 0
if and only if u

[ (1) spans N
R
, i.e., X

has no torus factors.


Proof. Since the D

are the irreducible components of X

` T
N
, Theorem 4.0.20
implies that we have an exact sequence
Div
T
N
(X

) Cl(X

) Cl(T
N
) 0.
Since C[x
1
, . . . , x
n
] is a UFD, the same is true for C[x
1
1
, . . . , x
1
n
]. This is the co-
ordinate ring of the torus (C

)
n
, which is isomorphic to the coordinate ring C[M]
of the torus T
N
. Hence C[M] is also a UFD, which implies Cl(T
N
) = 0 by Theo-
rem 4.0.18. We conclude that Div
T
N
(X

) Cl(X

) is surjective.
The composition M Div
T
N
(X

) Cl(X

) is obviously zero since the rst


map is mdiv(
m
). Now suppose that DDiv
T
N
(X

) maps to 0 in Cl(X

). Then
D = div( f ) for some f C(X

. Since the support of D misses T


N
, this implies
that div( f ) restricts to 0 on T
N
. When regarded as an element of C(T
N
)

, f has
zero divisor on T
N
, so that f C[M]

by Proposition 4.0.16. Thus f = c


m
for
some c C

and m M (Exercise 3.3.4). It follows that on X

,
D = div( f ) = div(c
m
) = div(
m
),
which proves exactness at Div
T
N
(X

).
Finally, suppose that m M with div(
m
) =

(1)
'm, u

`D

is the zero
divisor. Then 'm, u

` = 0 for all (1), which forces m = 0 when the u

span
N
R
. This gives the desired exact sequence. Conversely, if the sequence is exact,
then one easily sees that the u

span N
R
, which by Corollary 3.3.10 is equivalent
to X

having no torus factors.


In particular, we see that Cl(X

) is a nitely generated abelian group.


4.1. Weil Divisors on Toric Varieties 173
Examples. It is easy to compute examples of class groups of toric varieties. In
practice, one usually picks a basis e
1
, . . . , e
n
of M, so that MZ
n
and (via the dual
basis) N Z
n
. Then the pairing 'm, u` becomes dot product. We list the rays of
as
1
, . . . ,
r
with corresponding ray generators u
1
, . . . , u
r
Z
n
. We will think
of u
i
as the column vector ('e
1
, u
i
`, . . . , 'e
n
, u
i
`)
T
, where the superscript denotes
transpose.
With this setup, the map M Div
T
N
(X

) in Theorem 4.1.3 is the map


A : Z
n
Z
r
represented by the matrix whose rows are the ray generators u
1
, . . . , u
r
. In other
words, A = (u
1
, . . . , u
r
)
T
. By Theorem 4.1.3, the class group of X

is the cokernel
of this map, which is easily computed from the Smith normal form of A.
When we want to think in terms of divisors, we let D
i
be the T
N
-invariant prime
divisor corresponding to
i
(1).
Example 4.1.4. The afne toric surface described in Example 1.2.22 comes from
the cone = Cone(de
1
e
2
, e
2
). For d = 3, is shown in Figure 1. The resulting

2
u
2

1
u
1
Figure 1. The cone when d =3
toric variety U

is the rational normal cone



C
d
. Using the ray generators u
1
=
de
1
e
2
= (d, 1) and u
2
= e
2
= (0, 1), we get the map Z
2
Z
2
given by the
matrix
A =

d 1
0 1

.
This makes it easy to compute that
Cl(

C
d
) Z/dZ.
We can also see this in terms of divisors as follows. The class group Cl(

C
d
) is
generated by the classes of the divisors D
1
, D
2
corresponding to
1
,
2
, subject to
174 Chapter 4. Divisors on Toric Varieties
the relations coming from the exact sequence of Theorem 4.1.3:
0 div(
e
1
) ='e
1
, u
1
`D
1
+'e
1
, u
2
`D
2
= d D
1
0 div(
e
2
) ='e
2
, u
1
`D
1
+'e
2
, u
2
`D
2
=D
1
+D
2
.
Thus Cl(

C
d
) is generated by [D
1
] with d[D
1
] = 0, giving Cl(

C
d
) Z/dZ.
Example 4.1.5. In Example 3.1.4, we saw that the blowup of C
2
at the origin is
the toric variety Bl
0
(C
2
) given by the fan shown in Figure 2.

0
u
1
u
0
u
2
Figure 2. The fan for the blowup of C
2
at the origin
The ray generators are u
1
= e
1
, u
2
= e
2
, u
0
= e
1
+e
2
corresponding to divisors
D
1
, D
2
, D
0
. By Theorem 4.1.3, the class group is generated by the classes of the D
i
subject to the relations
0 div(
e
1
) = D
1
+D
0
0 div(
e
2
) = D
2
+D
0
.
Thus Cl(Bl
0
(C
2
)) Z with generator [D
1
] = [D
2
] = [D
0
]. This calculation can
also be done using matrices as in the previous example.
Example 4.1.6. The fan of P
n
has ray generators given by u
0
=e
1
e
n
and
u
1
= e
1
, . . . , u
n
= e
n
. Thus the map M Div
T
N
(P
n
) can be written as
Z
n
Z
n+1
(a
1
, . . . , a
n
) (a
1
a
n
, a
1
, . . . , a
n
).
Using the map
Z
n+1
Z
(b
0
, . . . , b
n
) b
0
+ +b
n
,
one gets the exact sequence
0 Z
n
Z
n+1
Z 0,
4.1. Weil Divisors on Toric Varieties 175
which proves that Cl(P
n
) Z, generalizing Example 4.0.21. It is easy to redo this
calculation using divisors as in the previous example.
Example 4.1.7. The class group Cl(P
n
P
m
) is isomorphic to Z
2
. More generally,
Cl(X

1
X

2
) Cl(X

1
) Cl(X

2
).
You will prove this in Exercise 4.1.2.
Example 4.1.8. The Hirzebruch surfaces H
r
are described in Example 3.1.16.
The fan for H
r
appears in Figure 3, along with the ray generators u
1
= e
1
+re
2
,
u
2
= e
2
, u
3
= e
1
, u
4
=e
2
.
u
2
u
4
u
3
u
1
= (1, r)
Figure 3. A fan r with Xr
Hr
The class group is generated by the classes of D
1
, D
2
, D
3
, D
4
, with relations
0 div(
e
1
) =D
1
+D
3
0 div(
e
2
) = r D
1
+D
2
D
4
.
It follows that Cl(H
r
) is the free abelian group generated by [D
1
] and [D
2
]. Thus
Cl(H
r
) Z
2
.
In particular, r = 0 gives Cl(H
0
) = Cl(P
1
P
1
) Z
2
, which is a special case of
Example 4.1.7.
Exercises for 4.1.
4.1.1. This exercise will determine which divisors are invariant under the T
N
-action on
X

. Given t T
N
and p X

, the T
N
-action gives t p X

. If D is a prime divisor,
the T
N
-action gives the prime divisor t D. For an arbitrary Weil divisor D =

i
a
i
D
i
,
t D =

i
a
i
(t D
i
). Then D is T
N
-invariant if t D = D for all t T
N
.
(a) Show that

(1)
a

is T
N
-invariant.
(b) Conversely, show that any T
N
-invariant Weil divisor can be written as in part (a). Hint:
Consider Supp(D) and use the Orbit-Cone Correspondence.
176 Chapter 4. Divisors on Toric Varieties
4.1.2. Given fans
1
in (N
1
)
R
and
2
in (N
2
)
R
, we get the product fan

2
=
1

2
[
i

i
,
which by Proposition 3.1.14 is the fan of the toric variety X
1
X
2
. Prove that
Cl(X
1
X
2
) Cl(X
1
) Cl(X
2
).
Hint: The product fan has rays
1
0 and 0
2
for
1

1
(1) and
2

2
(1).
4.1.3. Redo the divisor class group calculation given in Example 4.1.5 using matrices, and
redo the calculation given in Example 4.1.6 using divisors.
4.1.4. The blowup of C
n
at the origin is the toric variety Bl
0
(C
n
) of the fan described in
Example 3.1.15. Prove that Cl(Bl
0
(C
n
)) Z.
4.1.5. The weighted projective space P(q
0
, . . . , q
n
), gcd(q
0
, . . . , q
n
) = 1, is built from a fan
in N =Z
n+1
/Z(q
0
, . . . , q
n
). The dual lattice is
M =(a
0
, . . . , a
n
) Z
n+1
[ a
0
q
0
+ +a
n
q
n
= 0.
Let u
0
, . . . , u
n
N denote the images of the standard basis e
0
, . . . , e
n
Z
n+1
. The u
i
are the
ray generators of the fan giving P(q
0
, . . . , q
n
). Dene maps
M Z
n+1
: m ('m, u
0
`, . . . , 'm, u
n
`)
Z
n+1
Z : (a
0
, . . . , a
n
) a
0
q
0
+ +a
n
q
n
.
Show that these maps give an exact sequence
0 M Z
n+1
Z 0
and conclude that Cl(P(q
0
, . . . , q
n
)) Z.
4.2. Cartier Divisors on Toric Varieties
Let X

be the toric variety of a fan . We will use the same notation as in 4.1,
where each (1) gives a minimal ray generator u

and a T
N
-invariant prime
divisor D

. In what follows, we write


for a summation over the rays


(1) when there is no danger of confusion.
Computing the Picard Group. A Cartier divisor D on X

is also a Weil divisor


and hence
D

, a

Z,
by Theorem 4.1.3. Then

is Cartier since D is (Exercise 4.0.5). Let


CDiv
T
N
(X

) Div
T
N
(X

)
denote the subgroup of Div
T
N
(X

) consisting of T
N
-invariant Cartier divisors. Since
div(
m
) CDiv
T
N
(X

) for all m M, we get the following immediate corollary of


Theorem 4.1.3.
4.2. Cartier Divisors on Toric Varieties 177
Theorem 4.2.1. We have an exact sequence
M CDiv
T
N
(X

) Pic(X

) 0,
where the rst map is dened above and the second sends a T
N
-invariant divisor to
its divisor class in Pic(X

). Furthermore, we have a short exact sequence


0 M CDiv
T
N
(X

) Pic(X

) 0
if and only if u

[ (1) spans N
R
.
Our next task is to determine the structure of CDiv
T
N
(X

). In other words,
which T
N
-invariant divisors are Cartier? We begin with the afne case.
Proposition 4.2.2. Let N
R
be a strongly convex polyhedral cone. Then:
(a) Every T
N
-invariant Cartier divisor on U

is the divisor of a character.


(b) Pic(U

) = 0.
Proof. Let R = C[

M]. First suppose that D =

is an effective T
N
-
invariant Cartier divisor. Using Proposition 4.0.16 as in the proof of Proposi-
tion 4.0.28, we see that
(U

, O
U
(D)) =f K [ f = 0, or f = 0 and div( f ) D
is an ideal I R. Furthermore, I is T
N
-invariant since D is. Hence
(4.2.1) I =

m
I
C
m
=

div(
m
)D
C
m
by Lemma 1.1.16.
Under the Orbit-Cone Correspondence (Theorem 3.2.6), a ray (1) gives
an inclusion O() O() = D

. Thus
O()

.
Nowx a point p O(). Since D is Cartier, it is locally principal, and in particular
is principal in a neighborhood U of p. Shrinking U if necessary, we may assume
that U = (U

)
h
= Spec(R
h
), where h R satises h(p) = 0.
Thus D[
U
= div( f )[
U
for some f C(U

. Since D is effective, f R
h
by
Proposition 4.0.16, and since h is invertible on U, we may assume f R. Then
(4.2.2) div( f ) =

D
( f )D

E,=D

E
( f )E

D
( f )D

= D.
Here,

E,=D
denotes the sum over all prime divisors different from the D

. The
rst equality is the denition of div( f ), the second inequality follows since f R,
and the nal equality follows from D[
U
= div( f )[
U
since p U D

for all
(1). Then f I since div( f ) D by (4.2.2).
178 Chapter 4. Divisors on Toric Varieties
Using (4.2.1), we can write f =

i
a
i

m
i
with a
i
C

and div(
m
i
) D.
Restricting to U, this becomes div(
m
i
)[
U
div( f )[
U
, which implies that
m
i
/f is
a morphism on U by Proposition 4.0.16. Then
1 =

i
a
i

m
i
f
=

i
a
i

m
i
f
and p U imply that (
m
i
/f )(p) = 0 for some i. Hence
m
i
/f is nonvanishing in
some open set V with p V U. It follows that
div(
m
i
)[
V
= div( f )[
V
= D[
V
.
Since div(
m
i
) and D have support contained in

and every D

meets V (this
follows from p V D

), we have div(
m
i
) = D.
To nish the proof of (a), let D be an arbitrary T
N
-invariant Cartier divisor on
U

. Since dim

= dim M
R
( is strongly convex), we can nd m

M such
that 'm, u

` > 0 for all (1). Thus div(


m
) is a positive linear combination of
the D

, which implies that D

= D+div(
km
) 0 for k N sufciently large. The
above argument implies that D

is the divisor of a character, so that the same is true


for D. This completes the proof of part (a), and part (b) follows immediately using
Theorem 4.2.1.
Example 4.2.3. The rational normal cone

C
d
is the afne toric variety of the cone
= Cone(de
1
e
2
, e
2
) R
2
. We saw in Example 4.1.4 that Cl(U

) Z/dZ.
The edges
1
,
2
of give prime divisors D
1
, D
2
on

C
d
, and the computations of
Example 4.1.4 show that [D
1
] = [D
2
] generates Cl(U

). Since Pic(U

) = 0 by
Proposition 4.2.2, it follows that the Weil divisors D
1
, D
2
are not Cartier if d > 1.
Next consider the fan
0
consisting of the cones
1
,
2
, 0. This is a subfan of
the fan giving

C
d
, and the corresponding toric variety is X


C
d
`

, where

is the distinguished point that is the unique xed point of the T


N
-action on

C
d
. The variety X

0
is smooth since every cone in
0
is smooth (Theorem 3.1.19).
Since
0
and have the same 1-dimensional cones, they have the same class group
by Theorem 4.1.3. Thus
Pic(X

0
) = Cl(X

0
) = Cl(X

) = Cl(

C
d
) Z/dZ.
It follows that X

0
is a smooth toric surface whose Picard group has torsion.
Example 4.2.4. One of our favorite examples is X =V(xyzw) C
4
, which is the
toric variety of the cone =Cone(e
1
, e
2
, e
1
+e
3
, e
2
+e
3
) R
3
. The ray generators
are
u
1
= e
1
, u
2
= e
2
, u
3
= e
1
+e
3
, u
4
= e
2
+e
3
.
Note that u
1
+u
4
= u
2
+u
3
. Let D
i
X be the divisor corresponding to u
i
. In
Exercise 4.2.1 you will verify that
a
1
D
1
+a
2
D
2
+a
3
D
3
+a
4
D
4
is Cartier a
1
+a
4
= a
2
+a
3
4.2. Cartier Divisors on Toric Varieties 179
and that Cl(X) Z. Since Pic(X) = 0, we see that the D
i
are not Cartier, and in
fact no positive multiple of D
i
is Cartier.
Example 4.2.3 shows that the Picard group of a normal toric variety can have
torsion. However, if we assume that has a cone of maximal dimension, then the
torsion goes away. Here is the precise result.
Proposition 4.2.5. Let X

be the toric variety of a fan in N


R
R
n
. If contains
a cone of dimension n, then Pic(X

) is a free abelian group.


Proof. By the exact sequence in Theorem 4.2.1, it sufces to show that if D is a
T
N
-invariant Cartier divisor and kD is the divisor of a character for some k > 0,
then the same is true for D. To prove this, write D =

and assume that


kD = div(
m
), m M.
Let have dimension n. Since D is Cartier, its restriction to U

is also Cartier.
Using the Orbit-Cone Correspondence, we have
D[
U
=

(1)
a

.
This is principal on U

by Proposition 4.2.2, so that there is m

M such that
D[
U
= div(
m

)[
U
. This implies that
a

='m

, u

` for all (1).


On the other hand, kD = div(
m
) implies that
ka

='m, u

` for all (1).


Together, these equations imply
'km

, u

` = ka

='m, u

` for all (1).


The u

span N
R
since dim = n. Then the above equation forces km

= m, and
D = div(
m

) follows easily.
This proposition does not contradict the torsion Picard group in Example 4.2.3
since the fan
0
in that example has no maximal cone.
Comparing Weil and Cartier Divisors. Here is an application of Proposition 4.2.2.
Proposition 4.2.6. Let X

be the toric variety of the fan . Then the following are


equivalent:
(a) Every Weil divisor on X

is Cartier.
(b) Pic(X

) = Cl(X

).
(c) X

is smooth.
180 Chapter 4. Divisors on Toric Varieties
Proof. (a) (b) is obvious, and (c) (a) follows from Theorem 4.0.22. For
the converse, suppose that every Weil divisor on X

is Cartier and let U

be
the afne open subset corresponding to . Since Cl(X

) Cl(U

) is onto
by Theorem 4.0.20, it follows that every Weil divisor on U

is Cartier. Using
Pic(U

) = 0 from Proposition 4.2.2 and the exact sequence from Theorem 4.1.3,
we conclude that m div(
m
) induces a surjective map
M Div
T
N
(U

) =

(1)
ZD

.
Writing (1) =
1
, . . . ,
s
, this map becomes
(4.2.3)
M Z
s
m('m, u

1
`, . . . , 'm, u
s
`).
Now dene : Z
s
N by (a
1
, . . . , a
s
) =

s
i=1
a
i
u

i
. The dual map

: M = Hom
Z
(N, Z) Hom
Z
(Z
s
, Z) =Z
s
is easily seen to be (4.2.3). In Exercise 4.2.2 you will show that
(4.2.4)

is surjective is injective and N/(Z


s
) is torsion-free.
u

1
, . . . , u
s
can be extended to a basis of N.
The rst part of the proof shows that

is surjective. Then (4.2.4) implies that the


u

for (1) can be extended to a basis of N, which implies that is smooth.


Then X

is smooth by Theorem 3.1.19.


Proposition 4.2.6 has a simplicial analog. Recall that X

is simplicial when
every is simplicial, meaning that the minimal generators of are linearly
independent over R. You will prove the following result in Exercise 4.2.2.
Proposition 4.2.7. Let X

be the toric variety of the fan . Then the following are


equivalent:
(a) Every Weil divisor on X

has a positive integer multiple that is Cartier.


(b) Pic(X

) has nite index in Cl(X

).
(c) X

is simplicial.
In the literature, a Weil divisor is called Q-Cartier if some positive integer mul-
tiple is Cartier. Thus Proposition 4.2.7 characterizes those normal toric varieties for
which all Weil divisors are Q-Cartier.
Describing Cartier Divisors. We can use Proposition 4.2.2 to characterize T
N
-
invariant Cartier divisors as follows. Let
max
be the set of maximal cones
of , meaning cones in that are not proper subsets of another cone in .
4.2. Cartier Divisors on Toric Varieties 181
Theorem 4.2.8. Let X

be the toric variety of the fan and let D =

.
Then the following are equivalent:
(a) D is Cartier.
(b) D is principal on the afne open subset U

for all .
(c) For each , there is m

M with 'm

, u

` =a

for all (1).


(d) For each
max
, there is m

M with 'm

, u

` =a

for all (1).


Furthermore, if D is Cartier and m

is as in part (c), then:


(1) m

is unique modulo M() =

M.
(2) If is a face of , then m

mod M().
Proof. Since D[
U
=

(1)
a

, the equivalences (a) (b) (c) follow im-


mediately from Proposition 4.2.2. The implication (c) (d) is clear, and (d) (c)
follows because every cone in is a face of some
max
and if m


max
works for , it also works for all faces of .
For (1), suppose that m

M satises 'm, u

` = a

for all (1). Then,


given m

M, we have
'm

, u

` =a

for all (1) 'm

, u

` = 0 for all (1)


'm

, u` = 0 for all u
m

M = M().
It follows that m

is unique modulo M(). Since m

works for any face of ,


uniqueness implies that m

mod M(), and (2) follows.


The m

of part (c) of the theorem satisfy D[


U
= div(
m
)[
U
for all .
Thus (U

,
m
)

is local data for D in the sense of Denition 4.0.12. We


call m

the Cartier data of D.


The minus signs in parts (c) and (d) of the theorem are related to the minus
signs in the facet presentation of a lattice polytope given in (2.2.2), namely
P =m M
R
[ 'm, u
F
` a
F
for all facets F of P.
We will say more about this below. The minus signs are also related to support
functions, to be discussed later in the section.
When is a complete fan in N
R
R
n
, part (d) of Theorem 4.2.8 can be recast
as follows. Let (n) = [ dim = n. In Exercise 4.2.3 you will show that
a Weil divisor D =

is Cartier if and only if:


(d)

For each (n), there is m

M with 'm, u

` =a

for all (1).


Part (1) of Theorem 4.2.8 shows that these m

s are uniquely determined.


In general, each m

in Theorem 4.2.8 is only unique modulo M(). Hence we


can regard m

as a uniquely determined element of M/M(). Furthermore, if is


a face of , then the canonical map M/M() M/M() sends m

to m

.
182 Chapter 4. Divisors on Toric Varieties
There are two ways to turn these observations into a complete description of
CDiv
T
N
(X

). For the rst, write

max
=
1
, . . . ,
r

and consider the map

i
M/M(
i
)

i<j
M/M(
i

j
)
(m
i
)
i
(m
i
m
j
)
i<j
.
In Exercise 4.2.4 you will prove the following.
Proposition 4.2.9. There is a natural isomorphism
CDiv
T
N
(X

) ker

i
M/M(
i
)

i<j
M/M(
i

j
)

.
For readers who know inverse limits (see [10, p. 103]), a more sophisticated
description of CDiv
T
N
(X

) comes from the directed set (, _), where _is the face
relation. We get an inverse system where _ gives M/M() M/M(), and
the inverse limit gives an isomorphism
(4.2.5) CDiv
T
N
(X

) lim

M/M().
The Toric Variety of a Polytope. In Chapter 2, we constructed the toric variety
X
P
of a full dimensional lattice polytope P M
R
. If M
R
R
n
, this means that
dim P = n. As noted above, P has a canonical presentation
(4.2.6) P =m M
R
[ 'm, u
F
` a
F
for all facets F of P,
where a
F
Z and u
F
N is the inward-pointing facet normal that is the minimal
generator of the ray
F
= Cone(u
F
). The normal fan
P
consists of cones
Q
indexed by faces Q_P, where

Q
= Cone(u
F
[ F contains Q).
Proposition 2.3.8 implies that the fan
P
is complete. Furthermore, the vertices of
P correspond to the maximal cones in
P
(n), and the facets of P correspond to the
rays in
P
(1).
The ray generators of the normal fan
P
are the facet normals u
F
. The corre-
sponding prime divisors in X
P
will be denoted D
F
. Everything is now indexed by
the facets F of P. The normal fan tells us the facet normals u
F
in (4.2.6), but
P
cannot give us the integers a
F
in (4.2.6). For these, we need the divisor
(4.2.7) D
P
=

F
a
F
D
F
.
As we will see in later chapters, this divisor plays a central role in the study of
projective toric varieties. For now, we give the following useful result.
Proposition 4.2.10. D
P
is a Cartier divisor on X
P
and D
P
0.
4.2. Cartier Divisors on Toric Varieties 183
Proof. A vertex v P corresponds to a maximal cone
v
, and a ray
F
lies in
v
(1)
if and only if v F. But v F implies that 'v, u
F
` = a
F
. Note also that v M
since P is a lattice polytope. Thus we have v M such that 'v, u
F
` = a
F
for all

F

v
(1), so that D
P
is Cartier by Theorem 4.2.8. You will prove that D
P
0 in
Exercise 4.2.5.
In the notation of Theorem 4.2.8, m
v
is the vertex v. Thus the Cartier data of
the Cartier divisor D
P
is the set
(4.2.8) m
v

v
P
(n)
=v [ v is a vertex of P.
This is very satisfying and explains why the minus signs in (4.2.6) correspond to
the minus signs in Theorem 4.2.8.
The divisor class [D
P
] Pic(X
P
) also has a nice interpretation. If DD
P
, then
D = D
P
+div(
m
) for some m M. In Proposition 2.3.9 we saw that P and its
translate Pm have the same normal fan and hence give the same toric variety,
i.e., X
P
= X
m+P
. We also have
D = D
P
+div(
m
) = D
Pm
(Exercise 4.2.5), so that the divisor class of D
P
gives all translates of P.
The divisor D
P
has many more wonderful properties. We will get a glimpse
of this in 4.3 and learn the full power of D
P
in Chapter 6 when we study ample
divisors on toric varieties.
Support Functions. The Cartier data m

that describes a torus-invariant


Cartier divisor can be cumbersome to work with. Here we introduce a more ef-
cient computational tool. Recall that has support [[ =

N
R
.
Denition 4.2.11. Let be a fan in N
R
.
(a) A support function is a function : [[ R that is linear on each cone of .
The set of all support functions is denoted SF().
(b) A support function is integral with respect to the lattice N if
([[ N) Z.
The set of all such support functions is denoted SF(, N).
Let D =

be Cartier and let m

be the Cartier data of D as in


Theorem 4.2.8. Thus
(4.2.9) 'm

, u

` =a

for all (1).


We now describe Cartier divisors in terms of support functions.
184 Chapter 4. Divisors on Toric Varieties
Theorem 4.2.12. Let be a fan in N
R
. Then:
(a) Given D =

with Cartier data m

, the function

D
: [[ R
u
D
(u) ='m

, u` when u
is a well-dened support function that is integral with respect to N.
(b)
D
(u

) =a

for all (1), so that


D =

D
(u

)D

.
(c) The map D
D
induces an isomorphism
CDiv
T
N
(X

) SF(, N).
Proof. Theorem 4.2.8 tells us that each m

is unique modulo

M and that
m

mod (

M. It follows easily that


D
is well-dened. Also,
D
is linear on each since
D[

(u) ='m

, u` for u , and it is integral with respect


to N since m

M. This proves part (a), and part (b) follows from the denition of

D
and (4.2.9).
It remains to prove part (c). First note that
D
SF(, N) by part (a). Since
D, E CDiv
T
N
(X

) and k Z imply that

D+E
=
D
+
E

kD
= k
D
,
the map CDiv
T
N
(X

) SF(, N) is a homomorphism, and injectivity follows from


part (b). To prove surjectivity, take SF(, N). Fix . Since is integral
with respect to N, it denes a N-linear map [
N
: N Z, which extends to
N-linear map

: N

Z, where N

= Span() N. Since
Hom
Z
(N

, Z) M/M(),
it follows that there is m

M such that [

(u) = 'm

, u` for u . Then D =

D
(u

)D

is a Cartier divisor that maps to .


In terms of support functions, the exact sequence of Theorem 4.2.1 becomes
(4.2.10) M SF(, N) Pic(X

) 0,
where m M maps to the linear support function dened by u 'm, u` and
SF(, N) maps to the divisor class [

(u

)D

] Pic(X

). Be sure you
understand the minus signs.
Here is an example of how to compute with support functions.
Example 4.2.13. The eight points e
1
e
2
e
3
are the vertices of a cube in R
3
.
Taking the cones over the six faces gives a complete fan in R
3
. Modify this fan by
replacing e
1
+e
2
+e
3
with e
1
+2e
2
+3e
3
. The resulting fan has the surprising
4.2. Cartier Divisors on Toric Varieties 185
property that Pic(X

) = 0. In other words, X

is a complete toric variety whose


Cartier divisors are all principal.
We will prove Pic(X

) = 0 by showing that all support functions for are


linear. Label the ray generators as follows, using coordinates for compactness:
u
1
= (1, 2, 3), u
2
= (1, 1, 1), u
3
= (1, 1, 1), u
4
= (1, 1, 1)
u
5
= (1, 1, 1), u
6
= (1, 1, 1), u
7
= (1, 1, 1), u
8
= (1, 1, 1).
The ray generators are shown in Figure 4. The gure also includes three maximal
cones of :

1
= Cone(u
1
, u
2
, u
3
, u
5
)

2
= Cone(u
1
, u
3
, u
4
, u
7
)

3
= Cone(u
1
, u
2
, u
4
, u
6
).
The shading in Figure 4 indicates
1

2
,
1

3
,
2

3
. Besides
1
,
2
,
3
, the
fan has three other maximal cones, which we call left, down, and back. Thus
the cone left has ray generators u
2
, u
5
, u
6
, u
8
, and similarly for the other two.

3
u
1
u
2
u
3
u
4
u
5
u
6
u
7
u
8
Figure 4. A fan with Pic(X) =0
Take SF(, Z
3
). We show that is linear as follows. Since [

1
is linear,
there is m
1
Z
3
such that (u) ='m
1
, u` for u
1
. Hence the support function
u (u) 'm
1
, u`
vanishes identically on
1
. Replacing with this support function, we may assume
that [

1
= 0. Once we prove = 0 everywhere, it will follow that all support
functions are linear, and then Pic(X

) = 0 by (4.2.10).
Since u
1
, u
2
, u
3
, u
5

1
and vanishes on
1
, we have (u
1
) = (u
2
) =
(u
3
) = (u
5
) = 0. It sufces to prove (u
4
) = (u
6
) = (u
7
) = (u
8
) = 0.
186 Chapter 4. Divisors on Toric Varieties
To do this, we use the fact that each maximal cone has four generators, which
must satisfy a linear relation. Here are the cones and the corresponding relations:
cone relation

1
2u
1
+5u
5
= 4u
2
+3u
3

2
2u
1
+4u
7
= 3u
3
+5u
4

3
2u
1
+3u
6
= 4u
2
+5u
4
left u
2
+u
8
= u
5
+u
6
down u
3
+u
8
= u
5
+u
7
back u
4
+u
8
= u
6
+u
7
Since is linear on each cone and (u
1
) =(u
2
) =(u
3
) =(u
5
) =0, the second,
third, fourth and fth relations imply
4(u
7
) = 5(u
4
)
3(u
6
) = 5(u
4
)
(u
8
) = (u
6
)
(u
8
) = (u
7
).
The last two equations give (u
6
) = (u
7
), and substituting these into the rst two
shows that (u
4
) = (u
6
) = (u
7
) = (u
8
) = 0.
Since the toric variety of a polytope P has the non-principal Cartier divisor
D
P
, its follows that the fan of Example 4.2.13 is not the normal fan of any 3-
dimensional lattice polytope. As we will see later, this implies that X

is complete
but not projective.
A full dimensional lattice polytope P M
R
leads to an interesting support
function on the normal fan
P
.
Proposition 4.2.14. Assume P M
R
is a full dimensional lattice polytope with
normal fan
P
. Then the function
P
: N
R
R dened by

P
(u) = min('m, u` [ m P)
has the following properties:
(a)
P
is a support function for
P
and is integral with respect to N.
(b) The divisor corresponding to
P
is the divisor D
P
dened in (4.2.7).
Proof. First note that minimum used in the denition of
P
exists because P is
compact. Now write
P =m M
R
[ 'm, u
F
` a
F
for all facets F of P.
Then D
P
=

F
a
F
D
F
is Cartier by Proposition 4.2.10, and Theorem 4.2.12 shows
that the corresponding support function maps u
F
to a
F
.
4.2. Cartier Divisors on Toric Varieties 187
It remains to show that
P
(u) SF(
P
) and
P
(u
F
) =a
F
. Recall that maxi-
mal cones of
P
correspond to vertices of P, where the vertex v gives the maximal
cone
v
= Cone(u
F
[ v F). Take u =

vF

F
u
F

v
, where
F
0. Then
m P implies
(4.2.11) 'm, u` =

vF

F
'm, u
F
`

vF

F
a
F
.
Thus
P
(u)

vF

F
a
F
. Since equality occurs in (4.2.11) when m = v, we
obtain

P
(u) =

vF

F
a
F
='v, u`.
This shows that
P
SF(
P
, N). Furthermore, when v F, we have
P
(u
F
) =
'v, u
F
` =a
F
, as desired.
We will return to support functions in Chapter 6, where we will use them to
give elegant criteria for a divisor to be ample or generated by its global sections.
Exercises for 4.2.
4.2.1. Prove the assertions made in Example 4.2.4.
4.2.2. Prove (4.2.4) and Proposition 4.2.7.
4.2.3. When is complete, prove that D =

is Cartier if and only if it satises


condition (d)

stated in the discussion following Theorem 4.2.8.


4.2.4. Prove Proposition 4.2.9.
4.2.5. A lattice polytope P gives the toric variety X
P
and the divisor D
P
from (4.2.7).
(a) Prove that D
P
+div(
m
) = D
Pm
for any m M.
(b) Prove that D
P
0. Hint: The normal fan of P is complete.
4.2.6. Let D be a T
N
-invariant Cartier divisor on X

. By Theorem 4.2.8, D is determined


by its Cartier data m

. Given any m M, show that D+div(


m
) has Cartier data
m

. Be sure to explain where the minus sign comes from.


4.2.7. Let X

be the toric variety of the fan . Prove the following consequences of the
Orbit-Cone Correspondence (Theorem 3.2.6).
(a) O() =

(1)
D

.
(b) Rays u
1
, . . . , u
r
(1) lie in a cone of if and only if D
1
D
r
=.
4.2.8. Let be a fan in N
R
R
n
and assume that has a cone of dimension n.
(a) Fix a cone of dimension n. Prove that
Pic(X

) SF(, N) [ [

= 0.
(b) Explain how part (a) relates to Example 4.2.13.
(c) Use part (a) to give a different proof of Proposition 4.2.5.
188 Chapter 4. Divisors on Toric Varieties
4.2.9. Let be as in Example 4.2.4, but instead of using the lattice generated by e
1
, e
2
, e
3
,
instead use N = Z
1
2b
e
1
+Z
1
b
e
2
+Z
1
a
e
3
+Z
1
2b
(e
1
+e
2
+e
3
), where a, b are relatively
prime positive integers with a > 1. Prove that no multiple of D
1
+D
2
+D
3
+D
4
is Cartier.
Hint: The rst step will be to nd the minimal generators (relative to N) of the edges of .
4.2.10. Let X
P
be the toric variety of the octahedron P = Conv(e
1
, e
2
, e
3
) R
3
.
(a) Show that Cl(X
P
) Z
5
(Z/2Z)
2
.
(b) Use support functions and the strategy of Example 4.2.13 to show that Pic(X
P
) Z.
4.2.11. In Exercise 4.1.5, you showed that the weighted projective space P(q
0
, . . . , q
n
) has
class group Cl(P(q
0
, . . . , q
n
)) Z. Prove that Pic(P(q
0
, . . . , q
n
)) Cl(P(q
0
, . . . , q
n
)) maps
to the subgroup mZ Z, where m = lcm(q
0
, . . . , q
n
). Hint: Show that

n
i=0
b
i
D
i
generates
the class group, where

n
i=0
b
i
q
i
= 1. Also note that m M
Q
lies in M if and only if
'm, u
i
` Z for all i, where the u
i
are from Exercise 4.1.5.
4.2.12. Let X

be a smooth toric variety and let have dimension 2. This gives the
orbit closure V() = O() X

. In 3.3 we dened the blowup Bl


V()
(X

). Prove that
Pic(Bl
V()
(X

)) Pic(X

) Z.
4.2.13. A nonzero polynomial f =

mZ
n c
m
x
m
C[x
1
, . . . , x
n
] has Newton polytope
P( f ) = Conv(m [ c
m
= 0) R
n
.
When P( f ) has dimension n, Proposition 4.2.14 tells us that the function
P( f )
(u) =
min('m, u` [ m P( f )) is the support function of a divisor on X
P( f )
. Here we interpret

P( f )
as the tropicalization of f .
The tropical semiring (R, , ) has operations
a b = min(a, b) (tropical addition)
a b = a +b (tropical multiplication).
A tropical polynomial in real variables x
1
, . . . , x
n
is a nite tropical sum
F = c
1
x
a1,1
1
x
a1,n
n
c
r
x
ar,1
1
x
ar,n
n
where c
i
R and x
a
i
= x
i
x
i
(a times). For a more compact representation, dene
a tropical monomial to be x
m
= x
a1
1
x
an
n
for m = (a
1
, . . . , a
n
) N
n
. Then, using the
tropical analog of summation notation, the tropical polynomial F is
F =

r
i=1
c
i
x
mi
, m
i
= (a
i,1
, . . . , a
i,n
).
(a) Show that F = min
1ir
(c
i
+a
i,1
x
1
+ +a
i,n
x
n
).
(b) The tropicalization of our original polynomial f is the tropical polynomial
F
f
=

cm=0
0 x
m
.
Prove that F
f
=
P( f )
. (The 0 is explained as follows. In tropical geometry, one
often works in a larger ring where the coefcients of f are Puiseux series, and the
tropicalization uses the order of vanishing of the coefcients. Here, we use a smaller
ring where the coefcients of f are nonzero constants, with order of vanishing 0.)
(c) The tropical variety of a tropical polynomial F is the set of points in R
n
where F is
not linear. For f = x +2y +3x
2
xy
2
+4x
2
y, compute the tropical variety of F
f
and
show that it consists of the rays in the normal fan of P( f ).
A nice introduction to tropical algebraic geometry can be found in [240].
4.3. The Sheaf of a Torus-Invariant Divisor 189
4.3. The Sheaf of a Torus-Invariant Divisor
If D=

is a T
N
-invariant divisor on the normal toric variety X

, we get the
sheaf O
X

(D) dened in 4.0. We will study these sheaves in detail in Chapters 6


and 8. In this section we will focus primarily on global sections.
We begin with a classic example of the sheaf O
X

(D).
Example 4.3.1. For P
n
, the divisors D
0
, . . . , D
n
correspond to the ray generators of
the usual fan for P
n
. The computation Cl(P
n
) Z from Example 4.1.6 shows that
D
0
D
1
D
n
. These linear equivalences give isomorphisms
O
P
n (D
0
) O
P
n (D
1
) O
P
n (D
n
)
by Proposition 4.0.29. These sheaves are denoted O
P
n (1), and simiarly the sheaves
O
P
n(kD
i
), k Z, are denoted O
P
n (k). We will see the intrinsic reason for this
notation in Example 5.3.8.
Global Sections. Let D be a T
N
-invariant divisor on a toric variety X

. We will
give two descriptions of the global sections (X

, O
X

(D)). Here is the rst.


Proposition 4.3.2. If D is a T
N
-invariant Weil divisor on X

, then
(X

, O
X

(D)) =

div(
m
)+D0
C
m
.
Proof. If f (X

, O
X

(D)), then div( f ) +D 0 implies div( f )[


T
N
0 since
D[
T
N
= 0. Since C[M] is the coordinate ring of T
N
, Proposition 4.0.16 implies
f C[M]. Thus
(X

, O
X

(D)) C[M].
Furthermore, (X

, O
X

(D)) is invariant under the T


N
-action on C[M] since D is
T
N
-invariant. By Lemma 1.1.16, we obtain
(X

, O
X

(D)) =

m
(X

,O
X

(D))
C
m
.
Since
m
(X

, O
X

(D)) if and only if div(


m
) +D0, we are done.
The Polyhedron of a Divisor. For D =

and m M, div(
m
) +D 0 is
equivalent to
'm, u

` +a

0 for all (1),


which can be rewritten as
'm, u

` a

for all (1). (4.3.1)


This explains the minus signs! To emphasize the underlying geometry, we dene
(4.3.2) P
D
=m M
R
[ 'm, u

` a

for all (1).


190 Chapter 4. Divisors on Toric Varieties
We say that P
D
is a polyhedron since it is an intersection of nitely many closed
half spaces. This looks very similar to the canonical presentation of a polytope
(see (4.2.6), for example). However, the reader should be aware that P
D
need not
be a polytope, and even when it is a polytope, it need not be a lattice polytope. All
of this will be explained in the examples given below.
For now, we simply note that (4.3.1) is equivalent to m P
D
M. This gives
our second description of the global sections.
Proposition 4.3.3. If D is a T
N
-invariant Weil divisor on X

, then
(X

, O
X

(D)) =

mP
D
M
C
m
,
where P
D
M
R
is the polyhedron dened in (4.3.2).
As noted above, a polyhedron is an intersection of nitely many closed half
spaces. A polytope is a bounded polyhedron.
Examples. Here are some examples to illustrate the kinds of polyhedra that can
occur in Proposition 4.3.3.
Example 4.3.4. The fan for the blowup Bl
0
(C
2
) of C
2
at the origin has ray
generators u
0
= e
1
+e
2
, u
1
= e
1
, u
2
= e
2
and corresponding divisors D
0
, D
1
, D
2
.
For the divisor D = D
0
+D
1
+D
2
, a point m = (x, y) lies in P
D
if and only if
'm, u
0
` 1 x +y 1
'm, u
1
` 1 x 1
'm, u
2
` 1 y 1.
u
0
u
1
u
2

e
1
e
2
P
D
Figure 5. The fan and the polyhedron PD
The fan and the polyhedron P
D
are shown in Figure 5. Note that P
D
is not
bounded. By Proposition 4.3.3, the lattice points of P
D
(the dots in Figure 5) give
characters that form a basis of (Bl
0
(C
2
), O
Bl
0
(C
2
)
(D)).
4.3. The Sheaf of a Torus-Invariant Divisor 191
Example 4.3.5. The fan
2
for the Hirzebruch surface H
2
has ray generators
u
1
= e
1
+ 2e
2
, u
2
= e
2
, u
3
= e
1
, u
4
= e
2
. The corresponding divisors are
D
1
, D
2
, D
3
, D
4
, and Example 4.1.8 implies that the classes of D
1
and D
2
are a
basis of Cl(H
2
) Z
2
.
Consider the divisor aD
1
+D
2
, a Z, and let P
a
R
2
be the corresponding
polyhedron, which is a polytope in this case. A point m = (x, y) lies in P
a
if and
only if
'm, u
1
` a y
1
2
x
a
2
'm, u
2
` 1 y 1
'm, u
3
` 0 x 0.
'm, u
4
` 0 y 0.
Figure 6 shows
2
, together with shaded areas marked A, B, C. These are related
u
2
u
4
u
3
u
1

2
e
2
e
1
e
2
0
A
B
C
a = 1 a = 2 a = 3
Figure 6. The fan 2 and the polyhedra Pa
to the polygons P
a
for a = 1, 2, 3 by the equations
P
1
= A
P
2
= AB
P
3
= ABC.
Notice that as we increase a, the line y =
1
2
x
a
2
corresponding to u
1
moves to the
right and makes the polytope bigger. In fact, you can see that
2
is the normal fan
of the lattice polytope P
a
for any a 3. For a = 2, we get a lattice polytope P
2
, but
its normal fan is not
2
you can see how the facet with inward normal vector
u
2
collapses to a point of P
2
. For a = 1, P
1
is not a lattice polytope since
1
2
e
2
is a
vertex.
Chapters 6 and 7 will explain how the geometry of the polyhedron P
D
relates to
the properties of the divisor D. In particular, we will see that the divisor aD
1
+D
2
from Example 4.3.5 is ample if and only if a 3 since these are the only as for
which
2
is the normal fan P
a
.
192 Chapter 4. Divisors on Toric Varieties
Example 4.3.6. By Example 4.3.1, the sheaf O
P
n (k) can be written O
P
n (kD
0
),
where the divisor D
0
corresponds to the ray generator u
0
from Example 4.1.6. It is
straightforward to show that the polyhedron of D = kD
0
is
P
D
=

k < 0
k
n
k 0,
where
n
R
n
is the standard n-simplex. We can think of characters as Laurent
monomials t
m
= t
a
1
1
t
an
n
, where m = (a
1
, . . . , a
n
). It follows that
(P
n
, O
P
n (k)) f C[t
1
, . . . , t
n
] [ deg( f ) k.
The homogenization of such a polynomial is
F = x
k
0
f (x
1
/x
0
, . . . , x
n
/x
0
) C[x
0
, . . . , x
n
].
In this way, we get an isomorphism
(P
n
, O
P
n (k)) f C[x
0
, . . . , x
n
] [ f is homogeneous with deg( f ) = k.
The toric interpretation of homogenization will be discussed in Chapter 5.
Example 4.3.7. Let X
P
be the toric variety of a full dimensional lattice polytope
P M
R
. The facet presentation of P gives the Cartier divisor D
P
dened in (4.2.7),
and one checks easily that the polyhedron P
D
P
is the polytope P that we began with
(Exercise 4.3.1). It follows from Proposition 4.3.3 that
(X
P
, O
X
P
(D
P
)) =

mPM
C
m
.
Recall from Chapter 2 that the characters
m
for m PM give the projective
toric variety X
PM
. The divisor kD
P
gives the polytope kP (Exercise 4.3.2), so that
(X
P
, O
X
P
(kD
P
)) =

m(kP)M
C
m
.
In Chapter 2 we proved that kP is very ample for k sufciently large, in which case
X
(kP)M
is the toric variety X
P
. So the characters
m
that realize X
P
as a projective
variety come from global sections of O
X
P
(kD
P
). In Chapter 6, we will pursue these
ideas when we study ample and very ample Cartier divisors.
Note also that dim (X
P
, O
X
P
(kD
P
)) gives number of lattice points in multiples
of P. This will have important consequences in later chapters.
The operation sending a T
N
-invariant Weil divisor D X

to the polyhedron
P
D
M
R
dened in (4.3.2) has the following properties:
P
kD
= kP
D
for k 0.
P
D+div(
m
)
= P
D
m.
P
D
+P
E
P
D+E
.
You will prove these in Exercise 4.3.2. The multiple kP
D
and Minkowski sum
P
D
+P
E
are dened in 2.2, and Pm is translation.
4.3. The Sheaf of a Torus-Invariant Divisor 193
Complete Fans. When the fan is complete, we have the following niteness
result that you will prove in Exercise 4.3.3.
Proposition 4.3.8. Let X

be the toric variety of a complete fan in N


R
. Then:
(a) (X

, O
X

) =C, so the only morphisms X

C are the constant ones.


(b) P
D
is a polytope for any T
N
-invariant Weil divisor D on X

.
(c) (X

, O
X

(D)) has nite dimension as a vector space over C for any Weil
divisor on X

.
The assertions of parts (a) and (c) are true in greater generality: if X is any
complete variety, then (X, O
X
) = C, and if F is any coherent sheaf on X, then
dim (X, F) <(see [245, Vol. 2, VI.1.1 and VI.3.4]).
Exercises for 4.3.
4.3.1. Prove the assertion P
DP
= P made in Example 4.3.7.
4.3.2. Prove the properties of D P
D
listed above.
4.3.3. Prove Proposition 4.3.8. Hint: For part (a), use completeness to show that m = 0
when 'm, u

` 0 for all . For part (b), assume M


R
= R
n
and suppose m
i
P
D
satisfy
[[m
i
[[ . Then consider the points
mi
||mi||
on the sphere S
n1
R
n
.
4.3.4. Let be a fan in N
R
with convex support. Then [[ N
R
is a convex polyhedral
cone with dual [[

M
R
.
(a) Prove that [[

is the polyhedron associated to the divisor D = 0 on X

.
(b) Conclude that (X

, O
X
) =

m||

M
C
m
.
(c) Use part (b) to prove part (a) of Proposition 4.3.8.
4.3.5. Example 4.3.5 studied divisors on the Hirzebruch surface H
2
. This exercise will
consider the divisors D = D
4
and D

= D+D
2
= D
2
+D
4
.
(a) Show that D gives the same polygon as D

, i.e., P
D
= P
D
.
(b) Since H
2
is smooth, D and D

are Cartier. Compute their respective Cartier data


m

2(2)
and m

2(2)
.
(c) Show that P = Conv(m

[
2
(2)) and that P = Conv(m

[
2
(2)).
Thus D and D

give the same polygon but differ in how their Cartier data relates to the
polygon. In Chapter 6 we will use this to prove that O
H2
(D) is generated by global sections
while O
H2
(D

) has base points.


Chapter 5
Homogeneous Coordinates
on Toric Varieties
5.0. Background: Quotients in Algebraic Geometry
Projective space P
n
is usually dened as the quotient
P
n
= (C
n+1
`0)/C

,
where C

acts on C
n+1
by scalar multiplication, i.e.,
(a
0
, . . . , a
n
) = (a
0
, . . . , a
n
).
The above representation denes P
n
as a set; making P
n
into a variety requires
the notion of abstract variety introduced in Chapter 3. The main goal of this chapter
is to prove that every toric variety has a similar quotient construction as a variety.
Group Actions. Let G be a group acting on a variety X. We always assume that
for every g G, the map
g
(x) = g x denes a morphism
g
: X X. When
X = Spec(R) is afne,
g
: X X comes from a homomorphism

g
: R R. We
dene the induced action of G on R by
g f =

g
1
( f )
for f R. In other words, (g f )(x) = f (g
1
x) for all x X. You will check in
Exercise 5.0.1 this gives an action of G on R. Thus we have two objects:
The set G-orbits X/G =G x [ x X.
The ring of invariants R
G
=f R [ g f = f for all g G.
To make X/G into an afne variety, we need to dene its coordinate ring, i.e., we
need to determine the polynomial functions on X/G. A key observation is that if
195
196 Chapter 5. Homogeneous Coordinates on Toric Varieties
f R
G
, then

f (G x) = f (x)
gives a well-dened function

f : X/G C. Hence elements of G give obvious
polynomial functions on X/G, which suggests that
as an afne variety, X/G = Spec(R
G
).
As shown by the following examples, this works in some cases but fails in others.
Example 5.0.1. Let
2
= 1 act on C
2
= Spec(C[s, t]), where 1
2
acts
by multiplication by 1. Note that every orbit consists of two elements, with the
exception of the orbit of the origin, which is the unique xed point of the action.
The ring of invariants C[s, t]

2
=C[s
2
, st, t
2
] is the coordinate ring of the afne
toric variety V(xz y
2
). Hence we get a map
: C
2
/
2
Spec(C[s, t]

2
) = V(xz y
2
) C
3
where the orbit
2
(a, b) maps to (a
2
, ab, b
2
). This is easily seen to be a bijection,
so that Spec(C[s, t]

2
) is the perfect way to make C
2
/
2
into an afne variety.
This is actually an example of the toric morphism induced by changing the
latticesee Examples 1.3.17 and 1.3.19.
Example 5.0.2. Let C

act on C
4
= Spec(C[x
1
, x
2
, x
3
, x
4
]), where C

acts via
(a
1
, a
2
, a
3
, a
4
) = (a
1
, a
2
,
1
a
3
,
1
a
4
).
In this case, the ring of invariants is
C[x
1
, x
2
, x
3
, x
4
]
C

=C[x
1
x
3
, x
2
x
4
, x
1
x
4
, x
2
x
3
],
which gives the map
: C
4
/C

Spec(C[x
1
, x
2
, x
3
, x
4
]
C

) = V(xy zw) C
4
where the orbit C

(a
1
, a
2
, a
3
, a
4
) maps to (a
1
a
3
, a
2
a
4
, a
1
a
4
, a
2
a
3
). Then we have
(Exercise 5.0.2):
is surjective.
If p V(xy zw) ` 0, then
1
(p) consists of a single C

-orbit which is
closed in C
4
.

1
(0) consists of all C

-orbits contained in C
2
(0, 0) (0, 0) C
2
.
Thus
1
(0) consists of innitely many C

-orbits.
This looks bad until we notice one further fact (Exercise 5.0.2):
The xed point 0 C
4
gives the unique closed orbit mapping to 0 under .
If (a, b) = (0, 0), then an example of a non-closed orbit is given by
C

(a, b, 0, 0) =(a, b, 0, 0) [ C

since lim
0
(a, b, 0, 0) = 0. However, restricting to closed orbits gives
closed C

-orbits V(xy zw).


5.0. Background: Quotients in Algebraic Geometry 197
We will see that this is the best we can do for this group action.
Example 5.0.3. Let C

act on C
n+1
=Spec(C[x
0
, . . . , x
n
]) by scalar multiplication.
Then the ring of invariants consists of polynomials satsifying
f (x
0
, . . . , x
n
) = f (x
0
, . . . , x
n
)
for all C

. Such polynomials must be constant, so that


C[x
0
, . . . , x
n
]
C

=C.
It follows that the quotient is Spec(C), which is just a point. The reason for this
is that the only closed orbit is the orbit of the xed point 0 C
n+1
.
Examples 5.0.2 and 5.0.3 show what happens when there are not enough in-
variant functions to separate G-orbits.
The Ring of Invariants. When G acts on an afne variety X = Spec(R), a natural
question concerns the structure of the ring of invariants. The coordinate ring R is a
nitely generated C-algebra without nilpotents. Is the same true for R
G
? It clearly
has no nilpotents since R
G
R. But is R
G
nitely generated as a C-algebra? This
is related to Hilberts Fourteenth Problem, which was settled by a famous example
of Nagata that R
G
need not be a nitely generated C-algebra! An exposition of
Hilberts problem and Nagatas example can be found in [83, Ch. 4].
If we assume that R
G
is nitely generated, then Spec(R
G
) is an afne variety
that is the best candidate for a quotient in the following sense.
Lemma 5.0.4. Let G act on X = Spec(R) such that R
G
is a nitely generated C-
algebra, and let : X Y = Spec(R
G
) be the morphism of afne varieties induced
by the inclusion R
G
R. Then:
(a) Given any diagram
X

Z
Y

where is a morphism of afne varieties such that (g x) = (x) for g G


and x X, there is a unique morphism making the diagram commute, i.e.,
= .
(b) If X is irreducible, then Y is irreducible.
(c) If X is normal, then Y is normal.
Proof. Suppose that Z = Spec(S) and that is induced by

: S R. Then

(S) R
G
follows easily from (g x) = (x) for g G, x X. Thus

fac-
tors uniquely as
S

R
G

R.
198 Chapter 5. Homogeneous Coordinates on Toric Varieties
The induced map : Y Z clearly has the desired properties.
Part (b) is immediate since R
G
is a subring of R. For part (c), let K be the eld
of fractions of R
G
. If a K is integral over R
G
, then it is also integral over R and
hence lies in R since R is normal. It follows that a RK, which obviously equals
R
G
since G acts trivially on K. Thus R
G
is normal.
This shows that Y = Spec(R
G
) has some good properties when R
G
is nitely
generated, but there are still some unanswered questions, such as:
Is : X Y surjective?
Does Y have the right topology? Ideally, we would like U Y to be open if
and only if
1
(U) X is open. (Exercise 5.0.3 explores how this works for
group actions on topological spaces.)
While Y is the best afne approximation of the quotient X/G, could there be a
non-afne variety that is a better approximation?
We will see that the answers to these questions are all yes once we work with the
correct type of group action.
Good Categorical Quotients. In order to get the best properties of a quotient map,
we consider the general situation where G is a group acting on a variety X and
: X Y is a morphism that is constant on G-orbits. Then we have the following
denition.
Denition 5.0.5. Let G act on X and let : X Y be a morphism that is constant
on G-orbits. Then is a good categorical quotient if:
(a) If U Y is open, then the natural map O
Y
(U) O
X
(
1
(U)) induces an
isomorphism
O
Y
(U) O
X
(
1
(U))
G
.
(b) If W X is closed and G-invariant, then (W) Y is closed.
(c) If W
1
,W
2
are closed, disjoint, and G-invariant in X, then (W
1
) and (W
2
) are
disjoint in Y.
We often write a good categorical quotient as : X X//G. Here are some
properties of good categorical quotients.
Theorem 5.0.6. Let : X X//G be a good categorical quotient. Then:
(a) Given any diagram
X

Z
X//G

where is a morphism such that (g x) = (x) for g G and x X, there is


a unique morphism making the diagram commute, i.e., = .
5.0. Background: Quotients in Algebraic Geometry 199
(b) is surjective.
(c) A subset U X//G is open if and only if
1
(U) X is open.
(d) If U X//G is open and nonempty, then [

1
(U)
:
1
(U) U is a good
categorical quotient.
(e) Given points x, y X, we have
(x) = (y) G xG y =.
Proof. The proof of part (a) can be found in [83, Prop. 6.2]. The proofs of the
remaining parts are left to the reader (Exercise 5.0.4).
Algebraic Actions. So far, we have allowed G to be an arbitrary group acting on
X, assuming only that for every g G, the map x g x is a morphism
g
: X X.
We now make the further assumption that G is an afne variety. To dene this
carefully, we rst note that the group GL
n
(C) of n n invertible matrices with
entries in C is the afne variety
GL
n
(C) =A C
nn
=C
n
2
[ det(A) = 0.
A subgroup G GL
n
(C) is an afne algebraic group if it is also a subvariety of
GL
n
(C). Examples include GL
n
(C), SL
n
(C), (C

)
n
, and nite groups.
If G is an afne algebraic group, then the connected component of the identity,
denoted G

, has the following properties (see [152, 7.3]):


G

is a normal subgroup of nite index in G.


G

is an irreducible afne algebraic group.


An afne algebraic group G acts algebraically on a variety X if the G-action
(g, x) g x denes a morphism
GX X.
Examples of algebraic actions include toric varieties since the torus T
N
X acts
algebraically on X. Examples 5.0.1, 5.0.2 and 5.0.3 are also algebraic actions.
Algebraic actions have the property that G-orbits are constructible sets in X.
This has the following nice consequence for good categorical quotients.
Proposition 5.0.7. Let an afne algebraic group G act algebraically on a variety
X, and assume that a good categorical quotient : X X//G exists. Then:
(a) If p X//G, then
1
(p) contains a unique closed G-orbit.
(b) induces a bijection
closed G-orbits in X X//G.
Proof. For part (a), rst note that uniqueness follows immediately from part (e) of
Theorem 5.0.6. To prove the existence of a closed orbit in
1
(p), let G

G be
200 Chapter 5. Homogeneous Coordinates on Toric Varieties
the connected component of the identity. Then
1
(p) is stable under G

, so we
can pick an orbit G

x
1
(p) such that G

x has minimal dimension.


Note that G

x is irreducible since G

is irreducible, and since G

x is con-
structible, there is a nonempty Zariski open subset U of G

x such that U G

x.
If G

x is not closed, then G

x contains an orbit G

y = G

x. Thus
G

y G

x `G

x G

x `U.
However, G

x is irreducible, so that
dim (G

x `U) < dim G

x.
Hence G

y has strictly smaller dimension, a contradiction. Thus G

x is closed.
If g
1
, . . . , g
t
are coset representatives of G/G

, then
G x =
t

i=1
g
i
G

x
shows that G x is also closed. This proves part (a) of the proposition, and part (b)
follows immediately from part (a) and the surjectivity of .
For the rest of the section, we will always assume that G is an afne algebraic
group acting algebraically on a variety X.
Geometric Quotients. The best quotients are those where points are orbits. For
good categorical quotients, this condition is captured by requiring that orbits be
closed. Here is the precise result.
Proposition 5.0.8. Let : X X//G be a good categorical quotient. Then the
following are equivalent:
(a) All G-orbits are closed in X.
(b) Given points x, y X, we have
(x) = (y) x and y lie in the same G-orbit.
(c) induces a bijection
G-orbits in X X//G.
(d) The image of the morphism GX X X dened by (g, x) (g x, x) is the
ber product X
X//G
X.
Proof. This follows easily from Theorem 5.0.6 and Proposition 5.0.7. We leave
the details to the reader (Exercise 5.0.5).
In general, a good categorical quotient is called a geometric quotient if it
satises the condtions of Proposition 5.0.8. We write a geometric quotient as
: X X/G since points in X/G correspond bijectively to G-orbits in X.
5.0. Background: Quotients in Algebraic Geometry 201
We have yet to give an example of a good categorical or geometric quotient.
For instance, it is not clear that Examples 5.0.1, 5.0.2 and 5.0.3 satisfy Deni-
tion 5.0.5. Fortunately, once we restrict to the right kind of algebraic group, exam-
ples become abundant.
Reductive Groups. An afne algebraic group G is called reductive if its maximal
connected solvable subgroup is a torus. Examples of reductive groups include nite
groups, tori, and semisimple groups such as SL
n
(C).
For us, actions by reductive groups have the following key properties.
Proposition 5.0.9. Let G be a reductive group acting algebraically on an afne
variety X = Spec(R). Then:
(a) R
G
is a nitely generated C-algebra.
(b) The morphism : X Spec(R
G
) induced by R
G
R is a good categorical
quotient.
Proof. See [83, Prop. 3.1] for part (a) and [83, Thm. 6.1] for part (b).
In the situation of Proposition 5.0.9, we can write Spec(R)//G = Spec(R
G
).
Examples 5.0.1, 5.0.2 and 5.0.3 involve reductive groups acting on afne varieties.
Hence these are good categorical quotients that have all of the properties listed in
Theorem 5.0.6 and Proposition 5.0.7. Furthermore, Example 5.0.1 (the action of

2
on C
2
) is a geometric quotient. This last example generalizes as follows.
Example 5.0.10. Given a strongly convex rational polyhedral cone N
R
and
a sublattice N

N of nite index, part (b) of Proposition 1.3.18 implies that the


nite group G = N/N

acts on U
,N
such that the induced map on coordinate rings
is given by
C[

M]

C[

]
G
C[

].
It follows that : U
,N
U
,N
is a good categorical quotient. In fact, is a
geometric quotient since the G-orbits are nite and hence closed. This completes
the proof of part (c) of Proposition 1.3.18.
Almost Geometric Quotients. Let us examine Examples 5.0.2 and 5.0.3 more
closely. As noted above, both give good categorical quotients. However:
(Example 5.0.3) Here we have the quotient
C
n+1
//C

= Spec(C[x
0
, . . . , x
n
]
C

) = Spec(C) =pt.
So the good categorical quotient C
n+1
C
n+1
//C

=pt is really bad.


(Example 5.0.2) In this case, the quotient is
: C
4
C
4
//C

= V(xy zw).
Let U =V(xyzw)`0 and U
0
=
1
(U). Then [
U
0
: U
0
U is a good cat-
egorical quotient by Theorem 5.0.6, and by Example 5.0.2, orbits of elements
202 Chapter 5. Homogeneous Coordinates on Toric Varieties
in U
0
are closed in C
4
. Then [
U
0
is a geometric quotient by Proposition 5.0.8,
so that C
4
//C

= V(xy zw) is a geometric quotient outside of the origin.


The difference between these two examples is that the second is very close to being
a geometric quotient. Here is a result that describes this phenomenon in general.
Proposition 5.0.11. Let : X X//G be a good categorical quotient. Then the
following are equivalent:
(a) X has a G-invariant Zariski dense open subset U
0
such that G x is closed in X
for all x U
0
.
(b) X//G has a Zariski dense open subset U such that [

1
(U)
:
1
(U) U is
a geometric quotient.
Proof. Given U
0
satisfying (a), then W = X `U
0
is closed and G-invariant. For
x U
0
, the orbit G x U
0
is also closed and G-invariant. These are disjoint,
which implies (x) / (W) since is a good categorical quotient. Since is onto,
we see that X//G = (U
0
) (W) is a disjoint union. If we set U = (U
0
), then
U
0
=
1
(U). Note also that U is open since (W) is closed and Zariski dense
since U
0
is Zariski dense in X. Then [
U
0
: U
0
U is a good categorical quotient
by Theorem 5.0.6, and by assumption, orbits of elements in U
0
are closed in C
4
and hence in U
0
. It follows that [
U
0
is a geometric quotient by Proposition 5.0.8.
The proof going the other way is similar and is omitted (Exercise 5.0.6).
In general, a good categorical quotient is called an almost geometric quotient
if it satises the conditions of Proposition 5.0.11. Example 5.0.2 is an almost
geometric quotient while Example 5.0.3 is not.
Constructing Quotients. Now that we can handle afne quotients in the reductive
case, the next step is to handle more general quotients. Here is a useful result.
Proposition 5.0.12. Let G act on X and let : X Y be a morphism of varieties
that is constant on G-orbits. If Y has an open cover Y =

such that
[

1
(V)
:
1
(V

) V

is a good categorical quotient for every , then : X Y is a good categorical


quotient.
Proof. The key point is that the properties listed in Denition 5.0.5 can be checked
locally. We leave the details to the reader (Exercise 5.0.7).
Example 5.0.13. Consider a lattice N and a sublattice N

N of nite index, and


let be a fan in N

R
= N
R
. This gives a toric morphism
: X
,N
X
,N
.
5.0. Background: Quotients in Algebraic Geometry 203
By Proposition 1.3.18, the nite group G = N/N

is the kernel of T
N
T
N
, so that
G acts on X
,N
. Since

1
(U
,N
) =U
,N

for , Example 5.0.10 and Proposition 5.0.12 imply that is a geometric


quotient. This strengthens the result proved in Proposition 3.3.7.
It is sometimes possible to construct the quotient of X by G by taking rings
of invariants for a suitable afne open cover. If the local quotients patch together
to form a separated variety Y, then the resulting morphism : X Y is a good
categorical quotient by Proposition 5.0.12. Here are two examples that illustrate
this strategy.
Example 5.0.14. Let C

act on C
2
` 0 by scalar multiplication, where C
2
=
Spec(C[x
0
, x
1
]). Then C
2
`0 =U
0
U
1
, where
U
0
=C
2
`V(x
0
) = Spec(C[x
1
0
, x
1
])
U
1
=C
2
`V(x
1
) = Spec(C[x
0
, x
1
1
])
U
0
U
1
=C
2
`V(x
0
x
1
) = Spec(C[x
1
0
, x
1
1
]).
The rings of invariants are
C[x
1
0
, x
1
]
C

=C[x
1
/x
0
]
C[x
0
, x
1
1
]
C

=C[x
0
/x
1
]
C[x
1
0
, x
1
1
]
C

=C[(x
1
/x
0
)
1
].
It follows that the V
i
=U
i
//C

glue together in the usual way to create P


1
. Since
C

-orbits are closed in C


2
`0, it follows that
P
1
= (C
2
`0)/C

is a geometric quotient.
This example generalizes to show that
P
n
= (C
n+1
`0)/C

is a good geometric quotient when C

acts on C
n+1
by scalar multiplication. At
the beginning of the section, we wrote this quotient as a set-theoretic construction.
It is now an algebro-geometric construction.
Our second example shows the importance of being separated.
Example 5.0.15. Let C

act on C
2
`0 by (a, b) =(a,
1
b). Then C
2
`0 =
U
0
U
1
, where U
0
, U
1
and U
0
U
1
are as in Example 5.0.14. Here, the rings of
204 Chapter 5. Homogeneous Coordinates on Toric Varieties
invariants are
C[x
1
0
, x
1
]
C

=C[x
0
x
1
]
C[x
0
, x
1
1
]
C

=C[x
0
x
1
]
C[x
1
0
, x
1
1
]
C

=C[(x
0
x
1
)
1
].
Gluing together V
i
= U
i
//C

along U
0
U
1
//C

gives the variety obtained from


two copies of C by identifying all nonzero points. This is the non-separated variety
constructed in Example 3.0.15.
In Exercise 5.0.8 you will draw a picture of the C

-orbits that explains why the


quotient cannot be separated in this example.
In this book, we usually use the word variety to mean separated variety.
For example, when we say that : X Y is a good categorical or geometric quo-
tient, we always assume that X and Y are separated. So Example 5.0.15 is not a
good categorical quotient. In algebraic geometry, most operations on varieties pre-
serve separatedness. Quotient constructions are one of the few exceptions where
care has to be taken to check that the resulting variety is separated.
Exercises for 5.0.
5.0.1. Let G act on an afne variety X = Spec(R) such that
g
(x) =g x is a morphism for
all g G.
(a) Show that g f =

g
1
( f ) denes an action of G on R. Be sure you understand why
the inverse is necessary.
(b) The evaluation map RX C is dened by ( f , x) f (x). Show that this map is
invariant under the action of G on RX given by g ( f , x) = (g f , g x).
5.0.2. Prove the claims made in Example 5.0.2.
5.0.3. Let G be a group acting on a Hausdorff topological space, and let X/G be the set of
G-orbits. Dene : X X/G by (x) = G x. The quotient topology on X/G is dened
by saying that U X/G is open if and only if
1
(U) X is open.
(a) Prove that if X/G is Hausdorff, then the G-orbits are closed subsets of X.
(b) Prove that if W X is closed and G-invariant, then (W) X/G is closed.
(c) Prove that if W
1
,W
2
are closed, disjoint, and G-invariant in X, then (W
1
) and (W
2
)
are disjoint in X/G.
5.0.4. Prove parts (b), (c), (d) and (e) of Theorem 5.0.6. Hint for part (b): Part (a) of Def-
inition 5.0.5 implies that O
X//G
(U) injects into O
X
(
1
(U)) for all open sets U X//G.
Use this to prove that (X) is Zariski dense in X//G. Then use part (b) of Denition 5.0.5.
5.0.5. Prove Proposition 5.0.8.
5.0.6. Complete the proof of Proposition 5.0.11.
5.0.7. Prove Proposition 5.0.12.
5.0.8. Consider the C

action on C
2
` 0 described in Example 5.0.15.
5.1. Quotient Constructions of Toric Varieties 205
(a) Show that with two exceptions, the C

-orbits are the hyperbolas x


1
x
2
= a, a = 0. Also
describe the two remaining C

-orbits.
(b) Give an intuitive explanation, with picture, to showthat the limit of the orbits x
1
x
2
=
a as a 0 consists of two distinct orbits.
(c) Explain how part (b) relates to the non-separated quotient constructed in the example.
5.0.9. Give an example of a reductive G-action on an afne variety X such that X has
a nonempty G-invariant afne open set U X with the property that the induced map
U//GX//G is not an inclusion.
5.0.10. Let a nite group G act on X. Then a good categorical quotient : X X//G
exists since nite groups are reductive. Explain why is a geometric quotient.
5.1. Quotient Constructions of Toric Varieties
Let X

be the toric variety of a fan in N


R
. The goal of this section is to construct
X

as an almost geometric quotient


X

(C
r
`Z)//G
for an appropriate choice of afne space C
r
, exceptional set Z C
r
, and reductive
group G. We will use our standard notation, where each (1) gives a minimal
generator u

N and a T
N
-invariant prime divisor D

.
No Torus Factors. Toric varieties with no torus factors have the nicest quotient
constructions. Recall from Proposition 3.3.9 that X

has no torus factors when N


R
is spanned by u

, (1), and when this happens, Theorem 4.1.3 gives the short
exact sequence
0 M

ZD

Cl(X

) 0,
where m M maps to div(
m
) =

'm, u

`D

and Cl(X

) is the class group


dened in 4.0. We use the convention that in expressions such as

and

, the index ranges over all (1).


We write the above sequence more compactly as
(5.1.1) 0 M Z
(1)
Cl(X

) 0.
Applying Hom
Z
(, C

) gives
1 Hom
Z
(Cl(X

), C

) Hom
Z
(Z
(1)
, C

) Hom
Z
(M, C

) 1,
which remains a short exact sequence since Hom
Z
(, C

) is left exact and C

is
divisible. We have natural isomorphisms
Hom
Z
(Z
(1)
, C

) (C

)
(1)
Hom
Z
(M, C

) T
N
,
and we dene the group G by
G = Hom
Z
(Cl(X

), C

).
206 Chapter 5. Homogeneous Coordinates on Toric Varieties
This gives the short exact sequence of afne algebraic groups
(5.1.2) 1 G (C

)
(1)
T
N
1.
The Group G. The group G dened above will appear in the quotient construction
of the toric variety X

. For the time being, we assume that X

has no torus factors.


The following result describes the structure of G and gives explicit equations
for G as a subgroup of the torus (C

)
(1)
.
Lemma 5.1.1. Let G (C

)
(1)
be as in (5.1.2). Then:
(a) Cl(X

) is the character group of G.


(b) Gis isomorphic to a product of a torus and a nite abelian group. In particular,
G is reductive.
(c) Given a basis e
1
, . . . , e
n
of M, we have
G =

(t

) (C

)
(1)
[

t
/m,u)

= 1 for all m M

(t

) (C

)
(1)
[

t
/e
i
,u)

= 1 for 1 i n

.
Proof. Since Cl(X

) is a nitely generated abelian group, Cl(X

) Z

H, where
H is a nite abelian group. Then
G = Hom
Z
(Cl(X

), C

) Hom
Z
(Z

H, C

) (C

Hom
Z
(H, C

).
This proves part (b) since Hom
Z
(H, C

) is a nite abelian group. For part (a), note


that Cl(X

) gives the map that sends g G=Hom


Z
(Cl(X

), C

) to g() C

.
Thus elements of Cl(X

) give characters on G, and the above isomorphisms make


it easy to see that all characters of G arise this way.
For part (c), the rst description of G follows from (5.1.2) since M Z
(1)
is dened by m M ('m, u

`) Z
(1)
, and the second description is an easy
consequence of the rst.
Example 5.1.2. The ray generators of the fan for P
n
are u
0
=

n
i=1
e
i
, u
1
=
e
1
, . . . , u
n
= e
n
. By Lemma 5.1.1, (t
0
, . . . , t
n
) (C

)
n+1
lies in G if and only if
t
/m,e
1
en)
0
t
/m,e
1
)
1
t
/m,en)
n
= 1
for all m M =Z
n
. Taking m equal to e
1
, . . . , e
n
, we see that G is dened by
t
1
0
t
1
= = t
1
0
t
n
= 1.
Thus
G =(, . . . , ) [ C

,
which is the action of C

on C
n+1
given by scalar multiplication.
5.1. Quotient Constructions of Toric Varieties 207
Example 5.1.3. The fan for P
1
P
1
has ray generators u
1
= e
1
, u
2
= e
1
, u
3
=
e
2
, u
4
= e
2
in N = Z
2
. By Lemma 5.1.1, (t
1
, t
2
, t
3
, t
4
) (C

)
4
lies in G if and
only if
t
/m,e
1
)
1
t
/m,e
1
)
2
t
/m,e
2
)
3
t
/m,e
2
)
4
= 1
for all m M =Z
2
. Taking m equal to e
1
, e
2
, we obtain
t
1
t
1
2
= t
3
t
1
4
= 1.
Thus
G =(, , , ) [ , C

(C

)
2
.
Example 5.1.4. Let = Cone(de
1
e
2
, e
2
) R
2
, which gives the rational normal
cone

C
d
. Example 4.1.4 shows that Cl(

C
d
) Z/dZ, so that
G = Hom
Z
(Z/dZ, C

)
d
,
where
d
C

is the group of dth roots of unity. To see how G acts on C


2
, one
uses the ray generators u
1
= de
1
e
2
and u
2
= e
2
to compute that
G =(, ) [
d
= 1
d
(Exercise 5.1.1). This shows that G can have torsion.
The Exceptional Set. For the quotient representation of X

, we have the group G


and the afne space C
(1)
. All that is missing is the exceptional set Z C
(1)
that
we remove from C
(1)
before taking the quotient by G.
One useful observation is that G and C
(1)
depend only on (1). In order to
get X

, we need something that encodes the rest of the fan . We will do this using
a monomial ideal in the coordinate ring of C
(1)
. Introduce a variable x

for each
(1) and let
S =C[x

[ (1)].
Then Spec(S) =C
(1)
. We call S the total coordinate ring of X

.
For each cone , dene the monomial
x

=

/ (1)
x

.
Thus x

is the product of the variables corresponding to rays not in . Then dene
the irrelevant ideal
B() ='x

[ ` S.
A useful observation is that x

is a multiple of x

whenever _. Hence, if
max
is the set of maximal cones of , then
B() ='x

[
max
`.
Furthermore, one sees easily that the minimal generators of B() are precisely the
x

for
max
. Hence, once we have (1), B() determines uniquely.
208 Chapter 5. Homogeneous Coordinates on Toric Varieties
Now dene
Z() = V(B()) C
(1)
.
The variety of a monomial ideal is a union of coordinate subspaces. For B(), the
coordinate subspaces can be described in terms of primitive collections, which are
dened as follows.
Denition 5.1.5. A subset C (1) is a primitive collection if:
(a) C (1) for all .
(b) For every proper subset C

C, there is with C

(1).
Proposition 5.1.6. The Z() as a union of irreducible components is given by
Z() =

C
V(x

[ C),
where the union is over all primitive collections C (1).
Proof. It sufces to determine the maximal coordinate subspaces contained in
Z(). Suppose that V(x

1
, . . . , x
s
) Z() is such a subspace and take .
Since x

vanishes on Z() and 'x

1
, . . . , x
s
` is prime, the Nullstellensatz implies
x

is divisible by some x

i
, i.e.,
i
/ (1). It follows that C =
1
, . . . ,
s
sat-
ises condition (a) of Denition 5.1.5, and condition (b) follows easily from the
maximality of V(x

1
, . . . , x
s
). Hence C is a primitive collection.
Conversely, every primitive collection C gives a maximal coordinate subspace
V(x

[ C) contained in Z(), and the proposition follows.


In Exercise 5.1.2 you will show that the algebraic analog of Proposition 5.1.6
is the primary decomposition
B() =

C
'x

[ C`.
Here are some easy examples.
Example 5.1.7. The fan for P
n
consists of cones generated by proper subsets of
u
0
, . . . , u
n
, where u
0
, . . . , u
n
are as in Example 5.1.2. Let u
i
generate
i
, 0 i n,
and let x
i
be the corresponding variable in the total coordinate ring. We compute
Z() in two ways:
The maximal cones of the fan are given by
i
= Cone(u
0
, . . . , u
i
, . . . , u
n
). Then
x

i
= x
i
, so that B() ='x
0
, . . . , x
n
`. Hence Z() =0.
The only primitive collection is
0
, . . . ,
n
, so Z() = V(x
0
, . . . , x
n
) = 0
by Proposition 5.1.6.
Example 5.1.8. The fan for P
1
P
1
has ray generators u
1
= e
1
, u
2
= e
1
, u
3
=
e
2
, u
4
= e
2
. See Example 3.1.12 for a picture of this fan. Each u
i
gives a ray
i
and a variable x
i
. We compute Z() in two ways:
5.1. Quotient Constructions of Toric Varieties 209
The maximal cone Cone(u
1
, u
3
) gives the monomial x
2
x
4
, and similarly the
other maximal cones give the monomials x
1
x
4
, x
1
x
3
, x
2
x
3
. Thus
B() ='x
2
x
4
, x
1
x
4
, x
1
x
3
, x
2
x
3
`,
and one checks that Z() =0C
2
C
2
0.
The only primitive collections are
1
,
2
and
3
,
4
, so that
Z() = V(x
1
, x
2
) V(x
3
, x
4
) =0C
2
C
2
0
by Proposition 5.1.6. Note also that B() ='x
1
, x
2
` 'x
3
, x
4
`.
A nal observation is that (C

)
(1)
acts on C
(1)
by diagonal matrices and
hence induces an action on C
(1)
`Z(). It follows that G(C

)
(1)
also acts on
C
(1)
`Z(). We are now ready to take the quotient.
The Quotient Construction. To represent X

as a quotient, we rst construct a


toric morphism C
(1)
`Z() X

. Let e

[ (1) be the standard basis of


the lattice Z
(1)
. For each , dene the cone
= Cone(e

[ (1)) R
(1)
.
It is easy to see that these cones and their faces form a fan

= [ _ for some
in (Z
(1)
)
R
=R
(1)
. This fan has the following nice properties.
Proposition 5.1.9. Let

be the fan dened above. Then:
(a) C
(1)
`Z() is the toric variety of the fan

.
(b) The map e

denes a map of lattices Z


(1)
N that is compatible with
the fans

in R
(1)
and in N
R
.
(c) The resulting toric morphism
: C
(1)
`Z() X

is constant on G-orbits.
Proof. For part (a), let

0
be the fan consisting of Cone(e

[ (1)) and its


faces. Note that

is a subfan of

0
. Since

0
is the fan of C
(1)
, we get the
toric variety of

by taking C
(1)
and then removing the orbits corresponding to
all cones in

0
`

. By the Orbit-Cone Correspondence (Theorem 3.2.6), this is


equivalent to removing the orbit closures of the minimal elements of

0
`

. But
these minimal elements are precisely the primitive collections C (1). Since the
corresponding orbit closure is V(x

[ C), removing these orbit closures means


removing
Z() =

C
V(x

[ C).
210 Chapter 5. Homogeneous Coordinates on Toric Varieties
For part (b), dene : Z
(1)
N by (e

) =u

. Since the minimal generators


of are given by u

, (1), we have
R
( ) = by the denition of .
Hence is compatible with respect to the fans

and .
The map of tori induced by is the map (C

)
(1)
T
N
from the exact se-
quence (5.1.2) (you will check this in Exercise 5.1.3). Hence, if g G (C

)
(1)
and x C
(1)
`Z(), then
(g x) = (g) (x) = (x),
where the rst equality holds by equivariance and the second holds since G is the
kernel of (C

)
(1)
T
N
. This proves part (c) of the proposition.
In Exercise 5.1.4 you will prove the following lemma, which will be used in
the proof of the quotient construction.
Lemma 5.1.10. Assume that V is an afne toric variety, not necessarily normal,
with torus T. Given a point p V, there is a point q T and a one-parameter
subgroup : C

T such that p = lim


t0
(t)q.
We can now give the quotient construction of X

.
Theorem 5.1.11. Let X

be a toric variety without torus factors and consider the


toric morphism : C
(1)
`Z() X

from Proposition 5.1.9. Then:


(a) is an almost geometric quotient for the action of G on C
(1)
`Z(). Thus
X

C
(1)
`Z()

//G.
(b) is a geometric quotient if and only if is simplicial.
Proof. We begin by studying the map
(5.1.3) [

1
(U)
:
1
(U

) U

for . First observe that if , , then


R
( ) is equivalent to _ .
It follows that
1
(U

) is the toric variety U


e
of = Cone(e

[ (1)). This
shows that (5.1.3) is the toric morphism

: U
e
U

,
where for simplicity we write

instead of [

1
(U)
.
Our rst task is to show that

is a good categorical quotient. Since G is


reductive, Proposition 5.0.9 reduces this to showing that the map

on coordinate
rings induces an isomorphism
(5.1.4) C[U

] C[U
e
]
G
.
The map

can be described as follows:


5.1. Quotient Constructions of Toric Varieties 211
For U
e
, the cone gives the semigroup

Z
(1)
=(a

) Z
(1)
[ a

0 for all (1).


Hence the coordinate ring of U
e
is the semigroup algebra
C[U
e
] =C

x
a

[ a

0 for all (1)

= S
x
,
where S
x
is the localization S =C[x

[ (1)] at x

=

/ (1)
x

.
For U

, the coordinate ring is the usual semigroup algebra C[

M].
The map : Z
(1)
N dualizes to the map M Z
(1)
sending m M to
('m, u

`) Z
(1)
. It follows that

: C[

M] S
x
is given by

(
m
) =

x
/m,u)

.
Note that 'm, u

` 0 for all (1), so that the expression on the right really


lies in S
x
.
Thus

can be written

: C[

M] S
x
, and since

is constant on G-orbits,

factors as
C[

M]

S
x

G
S
x
.
The map

has Zariski dense image in U

since

((C

)
(1)
) = T
N
by the
exact sequence (5.1.2). It follows that

is injective. To show that its image is


(S
x
)
G
, take f S
x
and write it as
f =

a
c
a
x
a
where each x
a
=

x
a

satises a

0 for all (1). Then f is G-invariant if


and only if for all t = (t

) G, we have

a
c
a
x
a
=

a
c
a
t
a
x
a
.
Thus f is G-invariant if and only if t
a
= 1 for all t G whenever c
a
= 0. The
map t t
a
is a character on G and hence is an element of its character group
Cl(X

) (Lemma 5.1.1). This character is trivial when c


a
= 0, so that by (5.1.1), the
exponent vector a = (a

) must come from an element m M, i.e., a

='m, u

` for
all (1). But x
a
S
x
, which implies that
'm, u

` = a

0 for all (1).


Hence m

M, which implies that f is in the image of

. This proves (5.1.4).


We conclude that

is a good categorical quotient.


We next follow ideas from [83, Prop. 12.1] to prove that
(5.1.5)

: U

U

is a geometric quotient is simplicial.


212 Chapter 5. Homogeneous Coordinates on Toric Varieties
First suppose that is simplicial. By Proposition 5.0.8, it sufces to show that
G-orbits are closed in U

. Let G

G be the connected component of the identity.


Since G

has nite index in G, it sufces to show that G

-orbits are closed in U



.
Take p U

and p G

p, where the closure is taken in U



. Note that
G

p is an afne toric variety, possibly nonnormal, with torus T G

/G

p
. By
Lemma 5.1.10, there are

: C

T and q

T such that p = lim


t0

(t)q

p.
Lifting these to G

gives a one-parameter subgroup : C

and a point q G

such that
(5.1.6) p = lim
t0
(t)q p.
Using G

(C

)
(1)
, we can write (t) = (t
a
) for exponents a

Z. Since is
a one-parameter subgroup of G, we have
(5.1.7)

= 0.
This follows easily from Lemma 5.1.1 (Exercise 5.1.5). Write p = (p

), p = (p

),
and q = (q

). Then (5.1.6) implies


p

= lim
t0
t
a
q

.
Since p, p U
e
and q G

, their th coordinates are nonzero for / (1). Then


the above limit implies a

= 0 for these s, so that (5.1.7) becomes

(1)
a

= 0.
But is simplicial, which means that the u

, (1), are linearly independent.


Hence a

=0 for all , so that is the trivial one-parameter subgroup. Then (5.1.6)


implies p = q p G

p. We conclude that G

p is closed in U

.
For the other implication of (5.1.5), suppose that is non-simplicial. Then
there is a relation

(1)
a

= 0 where a

Z and a

> 0 for at least one . If


we set a

= 0 for / (1), then the one-parameter subgroup


(t) = (t
a
) (C

)
(1)
is a one-parameter subgroup of G by Exercise 5.1.5. Dene p = (p

) U
e
by
p

1 a

0
0 a

< 0
and consider lim
t0
(t) p. The limit exists in C
(1)
since p

= 0 for a

< 0.
Furthermore, if / (1), the th coordinate of (t) p is 1 for all t, so that the limit
p = lim
t0
(t) p lies in U
e
. Since there is
0
(1) with a

0
> 0, we have:
Since the
0
th coordinate of p is nonzero, the same is true for all of G p.
Since a

0
> 0, the
0
th coordinate of p = lim
t0
(t) p is zero.
Then G p is not closed in U
e
since its Zariski closure contains p U
e
`G p. This
shows that

is not a geometric quotient and completes the proof of (5.1.5).


5.1. Quotient Constructions of Toric Varieties 213
We can now prove the theorem. Since the maps (5.1.3) are good categorical
quotients, the same is true for : C
(1)
` Z() X

by Proposition 5.0.12. To
prove part (a), let

be the subfan of simplicial cones of . Then X

is open
in X

, and since

(1) = (1), X

and X

have the same total coordinate ring S


and same group G. In Exercise 5.1.5, you will show that
(5.1.8)
1
(X

) =C
(1)
`Z(

) =

U
e
.
As above, [

1
(X

)
:
1
(X

) X

is a good categorical quotient, and

is
a geometric quotient for each

by (5.1.5). It follows easily that [

1
(X

)
is a geometric quotient, and then Proposition 5.0.11 implies that is an almost
geometric quotient. This argument also implies that is a geometric quotient
when is simplicial, which proves half of part (b). The other half of part (b)
follows from (5.1.5). The proof of the theorem is now complete.
One nice feature of the quotient X

=(C
(1)
`Z())//G is that it is compatible
with the tori, meaning that we have a commutative diagram
X

(C
(1)
`Z())//G

T
N
(C

)
(1)
/G
where the isomorphism on the bottom comes from (5.1.2) and the vertical arrows
are inclusions.
Examples. Here are some examples of the quotient construction.
Example 5.1.12. By Examples 5.1.2 and 5.1.7, P
n
has quotient representation
P
n
= (C
n+1
`0)/C

,
where C

acts by scalar multiplication. This is a geometric quotient since is


smooth and hence simplicial.
Example 5.1.13. By Examples 5.1.3 and 5.1.8, P
1
P
1
has quotient representation
P
1
P
1
=

C
4
`(0C
2
C
2
0)

/(C

)
2
,
where (C

)
2
acts via (, ) (a, b, c, d) =(a, b, c, d). This is again a geometric
quotient.
Example 5.1.14. Fix positive integers q
0
, . . . , q
n
with gcd(q
0
, . . . , q
n
) = 1 and let
N be the lattice Z
n+1
/Z(q
0
, . . . , q
n
). The images of the standard basis in Z
n+1
give
primitive elements u
0
, . . . , u
n
N satisfying q
0
u
0
+ +q
n
u
n
= 0. Let be the fan
consisting of all cones generated by proper subsets of u
0
, . . . , u
n
.
As in Example 3.1.17, the corresponding toric variety is denoted P(q
0
, . . . , q
n
).
Using the quotient construction, we can now explain why this is called a weighted
projective space.
214 Chapter 5. Homogeneous Coordinates on Toric Varieties
We have C
(1)
=C
n+1
since has n+1 rays, and Z() =0 by the argument
used in Example 5.1.7. It remains to compute G(C

)
n+1
. In Exercise 4.1.5, you
showed that the maps m M ('m, u
0
`, . . . , 'm, u
n
`) Z
n+1
and (a
0
, . . . , a
n
)
Z
n+1
a
0
q
0
+ +a
n
q
n
Z give the short exact sequence
(5.1.9) 0 M Z
n+1
Z 0.
This shows that the class group is Z. Note also that e
i
Z
n+1
maps to q
i
Z. In
Exercise 5.1.6 you will compute that
G =(t
q
0
, . . . , t
qn
) [ t C

.
This is the action of C

on C
n+1
given by
t (u
0
, . . . , u
n
) = (t
q
0
u
0
, . . . , t
qn
u
n
).
Since is simplicial (every proper subset of u
0
, . . . , u
n
is linearly independent in
N
R
), we get the geometric quotient
P(q
0
, . . . , q
n
) = (C
n+1
`0)/C

.
This gives the set-theoretic denition of P(q
0
, . . . , q
n
) from 2.0 and also gives its
structure as a variety since we have a geometric quotient.
Example 5.1.15. Consider the cone = Cone(e
1
, e
2
, e
1
+e
3
, e
2
+e
3
) R
3
. To
nd the quotient representation of U

, we label the ray generators as


u
1
= e
1
, u
2
= e
2
+e
3
, u
3
= e
2
, u
4
= e
1
+e
3
.
Then C
(1)
= C
4
and Z() = since x

= 1. To determine the group G (C

)
4
,
note that the exact sequence (5.1.1) becomes
0 Z
3
Z
4
Z 0,
where (a
1
, a
2
, a
3
, a
4
) Z
4
a
1
+a
2
a
3
a
4
Z. This makes it straightforward
to show that
G =(, ,
1
,
1
) [ C

.
Hence we get the quotient presentation
U

=C
4
//C

.
In Example 5.0.2, we gave a naive argument that the quotient was V(xy zw). We
now see that the intrinsic meaning of Example 5.0.2 is the quotient construction of
U

given by Theorem 5.1.11. This example is not a geometric quotient since is


not simplicial.
Example 5.1.16. Let Bl
0
(C
2
) be the blowup of C
2
at the origin, whose fan is
shown in Figure 1 on the next page. By Example 4.1.5, Cl(Bl
0
(C
2
)) Z with
generator [D
1
] = [D
2
] = [D
0
]. Hence G = C

and the irrelevant ideal is B() =


'x, y`. This gives the geometric quotient
Bl
0
(C
2
)

C
3
`(C0, 0)

/C

,
5.1. Quotient Constructions of Toric Varieties 215

0
u
1
u
0
u
2
Figure 1. The fan for the blowup of C
2
at the origin
where the C

-action is given by (t, x, y) = (


1
t, x, y).
We also have C[t, x, y]
C

=C[tx, ty]. Then the inclusion


C
3
`(C0, 0) C
3
induces the map on quotients
: Bl
0
(C
2
)

C
3
`(C0, 0)

/C

C
3
//C

C
2
,
where the nal isomorphism uses
C
3
//C

= Spec(C[t, x, y]
C

) = Spec(C[tx, ty]).
In terms of homogeneous coordinates, (t, x, y) = (tx, ty). This map is the toric
morphism Bl
0
(C
2
) C
2
induced by the renement of Cone(u
1
, u
2
) given by .
The quotient representation makes it easy to see why Bl
0
(C
2
) is the blowup of
C
2
at the origin. Given a point of Bl
0
(C
2
) with homogeneous coordinates (t, x, y),
there are two possibilities:
t = 0, in which case t (t, x, y) = (1, tx, ty). This maps to (tx, ty) C
2
and
is nonzero since x, y cannot both be zero. It follows that the part of Bl
0
(C
2
)
where t = 0 looks like C
2
`0, 0.
t = 0, in which case (0, x, y) maps to the origin in C
2
. Since (0, x, y) =
(x, y) and x, y cannot both be zero, it follows that the part of Bl
0
(C
2
) where
t = 0 looks like P
1
.
This shows that Bl
0
(C
2
) is a built from C
2
by replacing the origin with a copy
of P
1
, which is called the exceptional locus E. Since E =
1
(0, 0), we see that
: X

C
2
induces an isomorphism
Bl
0
(C
2
) `E C
2
`(0, 0).
Note also that E is the divisor D
0
corresponding to the ray
0
. You should be able
to look at Figure 1 and see instantly that D
0
P
1
.
We can also check that lines through the origin behave properly. Consider the
line L dened by ax +by = 0, where (a, b) = (0, 0). When we pull this back to
216 Chapter 5. Homogeneous Coordinates on Toric Varieties
Bl
0
(C
2
), we get the subvariety dened by
a(tx) +b(ty) = 0.
This is the total transform of L. It factors as t(ax+by) = 0. Note that t = 0 denes
the exceptional locus, so that once we remove this, we get the curve in Bl
0
(C
2
)
dened by ax +by = 0. This is the proper transform of L, which meets the excep-
tional locus E at the point with homogeneous coordinates (0, b, a), corresponding
to (b, a) P
1
. In this way, we see how blowing up separates tangent directions
through the origin.
The General Case. So far, we have assumed that X

has no torus factors. When


torus factors are present, X

still has a quotient construction, though it is no longer


canonical.
Let X

be a toric variety with a torus factor. By Proposition 3.3.9, the ray


generators u

, (1), span a proper subspace of N


R
. Let N

be the intersection
of this subspace with N, and pick a complement N

so that N =N

. The cones
of all lie in N

R
and hence give a fan

in N

R
. As in the proof of Proposition 3.3.9.
we obtain
X

,N
(C

)
r
where N

Z
r
. Theorem 5.1.11 applies to X

,N

since u

(1) =(1), span


N

R
by construction. Note also that B(

) = B() and Z(

) = Z(). Hence
X

,N
(C
(1)
`Z())//G.
It follows that
(5.1.10)
X

,N
(C

)
r

(C
(1)
`Z())//G

(C

)
r

(C
(1)
(C

)
r
) `(Z() (C

)
r
)

//G,
In the last line, we use the trivial action of G on (C

)
r
. You will verify the last
isomorphism in Exercise 5.1.7.
We can rewrite (5.1.10) as follows. Using (C

)
r
=C
r
`V(x
1
x
r
), we obtain
(C
(1)
(C

)
r
) `(Z() (C

)
r
) =C
(1)+r
`Z

(),
where C
(1)+r
= C
(1)
C
r
and Z

() = (Z() C
r
) (C
(1)
V(x
1
x
r
)).
Hence the quotient presentation of X

is the almost geometric quotient


(5.1.11) X

C
(1)+r
`Z

()

//G.
This differs from Theorem 5.1.11 in two ways:
The representation (5.1.11) is non-canonical since it depends on the choice of
the complement N

.
5.1. Quotient Constructions of Toric Varieties 217
Z

() contains V(x
1
x
r
) C
(1)
and hence has codimension 1 in C
(1)+r
.
In constrast, Z() always has codimension 2 in C
(1)
(this follows from
Proposition 5.1.6 since every primitive collection has at least two elements).
In practice, (5.1.11) is rarely used, while Theorem 5.1.11 is a common tool in toric
geometry.
Exercises for 5.1.
5.1.1. In Example 5.1.4, verify carefully that G =(, ) [
d
.
5.1.2. Prove that B() =

C
'x

[ C`, where the intersection ranges over all primitive


collections C (1).
5.1.3. In Proposition 5.1.9, we dened : Z
(1)
N, and in the proof we use the map of
tori (C

)
(1)
T
N
induced by . Show that this is the map appearing in (5.1.2).
5.1.4. This exercise will prove Lemma 5.1.10. In parts (a) and (b), we consider a normal
afne toric variety U

and a point p U

. By Theorem 3.2.6, there is a face _ such


that p O() U

. Also take u Relint() N.


(a) Prove lim
t0

u
(t) =

, where

O() is the distinguished point dened in 3.2.


Hint: Proposition 3.2.2.
(b) Find q T
N
such that lim
t0

u
(t)q = p. Hint: T
N
acts transitively on O().
(c) Prove Lemma 5.1.10. Hint: Let U

V be the normalization map. Then apply


Theorem 3.A.3.
5.1.5. This exercise is concerned with the proof of Theorem 5.1.11.
(a) Prove that a one-parameter subgroup (t) = (t
a
) (C

)
(1)
takes values in G if and
only if

= 0. Hint: Use Lemma 5.1.1. You can give a more conceptual proof
by taking the dual of (5.1.1).
(b) Prove (5.1.8) and conclude that the quotient construction of X

is the map [

1
(X

)
:

1
(X

) X

used in the proof of Theorem 5.1.11.


5.1.6. Show that the group G in Example 5.1.14 is given by G = (t
q0
, . . . , t
qn
) [ t C

.
Hint: Pick integers b
i
such that

n
i=0
b
i
q
i
= 1. Given (t
0
, . . . , t
n
) G, set t =

n
i=0
t
ai
i
. Also
note that if e
0
, . . . , e
n
is the standard basis of Z
n+1
, then q
i
e
j
q
j
e
i
Z
n+1
maps to 0 Z
in (5.1.9).
5.1.7. Let X be a variety with trivial G action. Prove that (X U
e
)//G X U

and use
this to verify the nal line of (5.1.10).
5.1.8. Consider the usual fan for P
2
with the lattice N =(a, b) Z
2
[ a+b 0 mod d,
where d is a positive integer.
(a) Prove that the ray generators are u
1
= (d, 0), u
2
= (0, d) and
u
0
=

(d, d) d odd
(d/2, d/2) d even.
(b) Prove that the dual lattice is M =(a/d, b/d) [ a, b Z, a b 0 mod d.
(c) Prove that Cl(X

) =ZZ/dZ (d odd) or ZZ/


d
2
Z (d even).
218 Chapter 5. Homogeneous Coordinates on Toric Varieties
(d) Compute the quotient representation of X

.
5.1.9. Find the quotient representation of the Hirzeburch surface H
r
in Example 3.1.16.
5.1.10. Prove that G acts freely on C
(1)
`Z() when the fan is smooth. Hint: Let
and suppose that g = (t

) G xes u = (u

) U
e
. Show that t

= 1 for / and then


use Lemma 5.1.1 to show that t

= 1 for all .
5.1.11. Prove that G acts with nite isotropy subgroups on C
(1)
`Z() when the fan is
simplicial. Hint: Adapt the argument used in Exercise 5.1.10.
5.1.12. Prove that 2 codim(Z()) [(1)[. When is a complete simplicial fan, a
stronger result states that either
(a) 2 codim(Z())
1
2
dim X

+1, or
(b) [(1)[ = dimX

+1 and Z() =0.


This is proved in [19, Prop. 2.8]. See the next exercise for more on part (b).
5.1.13. Let be a complete fan such that [(1)[ = n +1, where n = dim X

. Prove that
there is a weighted projective space P(q
0
, . . . , q
n
) and a nite abelian group H acting on
P(q
0
, . . . , q
n
) such that
X

P(q
0
, . . . , q
n
)/H.
These are called fake weighted projective spaces in [60] and [167]. Also prove that the
following are equivalent:
(a) X

is a weighted projective space.


(b) Cl(X

) Z.
(c) N is generated by u

, (1).
Hint: Label the ray generators u
0
, . . . , u
n
. First show that is simplicial and that there
are positive integers q
0
, . . . , q
n
satisfying

n
i=0
q
i
u
i
= 0 and gcd(q
0
, . . . , q
n
) = 1. Then
consider the sublattice of N generated by the u
i
and use Example 5.1.14. You will also
need Proposition 3.3.7. If you get stuck, see [19, Lem. 2.11].
5.1.14. In the proof of Theorem 5.1.11, we showed that a non-simplicial cone leads to a
non-closed G-orbit. Show that the non-closed G-orbit exhibited in Example 5.0.2 is an
example of this construction. See also Example 5.1.15.
5.1.15. Example 5.1.16 gave the quotient construction of the blowup of 0 C
2
and used
the quotient construction to describe the properties of the blowup. Give a similar treatment
for the blowup of C
r
C
n
using the star subdivision described in 3.3.
5.2. The Total Coordinate Ring
In this section we assume that X

is a toric variety without torus factors. Its total


coordinate ring
S =C[x

[ (1)]
was dened in 5.1. This ring gives C
(1)
= Spec(S) and contains the irrelevant
ideal
B() ='x

[ `
used in the quotient construction of X

. In this section we will explore how this


ring relates to the algebra and geometry of X

.
5.2. The Total Coordinate Ring 219
The Grading. An important feature of the total coordinate ring S is its grading by
the class group Cl(X

). We have the exact sequence (5.1.1)


0 M Z
(1)
Cl(X

) 0,
where a = (a

) Z
(1)
maps to the divisor class

Cl(X

). Given a
monomial x
a
=

x
a

S, dene its degree to be


deg(x
a
) =

Cl(X

).
For Cl(X

), we let S

denote the corresponding graded piece of S.


The grading on S is closely related to the group G = Hom
Z
(Cl(X

), C

). Re-
call that Cl(X

) is the character group of G, where as usual Cl(X

) gives the
character

: G C

. The action of G on C
(1)
induces an action on S with the
property that given f S, we have
(5.2.1)
f S

g f =

(g
1
) f for all g G
f (g x) =

(g) f (x) for all g G, x C


(1)
(Exercise 5.2.1). Thus the graded pieces of S are the eigenspaces of the action of
G on S. We say that f S

is homogeneous of degree .
Example 5.2.1. The total coordinate ring of P
n
is C[x
0
, . . . , x
n
]. By Example 4.1.6,
the map Z
n+1
Z = Cl(P
n
) is (a
0
, . . . , a
n
) a
0
+ +a
n
. This gives the grad-
ing on C[x
0
, . . . , x
n
] where each variable x
i
has degree 1, so that homogeneous
polynomial has the usual meaning.
In Exercise 5.2.2 you will generalize this by showing that the total coordi-
nate ring of the weighted projective space P(q
0
, . . . , q
n
) is C[x
0
, . . . , x
n
], where the
variable x
i
now has degree q
i
. Here, homogeneous polynomial means weighted
homogeneous polynomial.
Example 5.2.2. The fan for P
n
P
m
is the product of the fans of P
n
and P
m
, and
by Example 4.1.7, the class group is
Cl(P
n
P
m
) Cl(P
n
) Cl(P
m
) Z
2
.
The total coordinate ring is C[x
0
, . . . , x
n
, y
0
, . . . , y
m
], where
deg(x
i
) = (1, 0) deg(y
i
) = (0, 1)
(Exercise 5.2.3). For this ring, homogeneous polynomial means bihomogeneous
polynomial.
Example 5.2.3. Example 5.1.16 gave the quotient representation of the blowup
Bl
0
(C
2
) of C
2
at the origin. The fan of Bl
0
(C
2
) is shown in Example 5.1.16 and
has ray generators u
0
, u
1
, u
2
, corresponding to variables t, x, y in the total coordinate
ring S = C[t, x, y]. Since Cl(Bl
0
(C
2
)) Z, one can check that the grading on S is
given by
deg(t) =1 and deg(x) = deg(y) = 1
220 Chapter 5. Homogeneous Coordinates on Toric Varieties
(Exercise 5.2.4). Thus total coordinate rings can have some elements of positive
degree and other elements of negative degree.
The Toric Ideal-Variety Correspondence. For n-dimensional projective space P
n
,
a homogeneous ideal I C[x
0
, . . . , x
n
] denes a projective variety V(I) P
n
. This
generalizes to more general toric varieties X

as follows.
We rst assume that is simplicial, so that we have a geometric quotient
: C
(1)
`Z() X

by Theorem 5.1.11. Given p X

, we say a point x
1
(p) gives homogeneous
coordinates for p. Since is a geometric quotient, we have
1
(p) = G x. Thus
all homogeneous coordinates for p are of the form g x for some g G.
Now let S be the total coordinate ring of X

and let f S be homogeneous for


the Cl(X

)-grading on S, say f S

. Then
f (g x) =

(g) f (x)
by (5.2.1), so that f (x) = 0 for one choice of homogeneous coordinates of p X

if and only if f (x) = 0 for all homogeneous coordinates of p. It follows that the
equation f = 0 is well-dened in X

. We can use this to dene subvarieties of X

as follows.
Proposition 5.2.4. Let S be the total coordinate ring of the simplicial toric variety
X

. Then:
(a) If I S is a homogeneous ideal, then
V(I) =(x) X

[ f (x) = 0 for all f I


is a closed subvariety of X

.
(b) All closed subvarieties of X

arise this way.


Proof. Given I S as in part (a), notice that
W =x C
(1)
`Z() [ f (x) = 0 for all f I
is a closed G-invariant subset of C
(1)
`Z(). By part (b) of the denition of good
categorical quotient (Denition 5.0.5), V(I) = (W) is closed in X

.
Conversely, given a closed subset Y X

, its inverse image

1
(Y) C
(1)
`Z()
is closed and G-invariant. Then the same is true for the Zariski closure

1
(Y) C
(1)
.
It follows without difculty that I = I(
1
(Y)) S is a homogeneous ideal satis-
fying V(I) =Y.
5.2. The Total Coordinate Ring 221
Example 5.2.5. The equation x

= 0 denes the T
N
-invariant closed subvariety
V(x

) X

which is easily seen to be the prime divisor D

. This shows that D

always has a global equation, though it fails to have local equations when D

is not
Cartier (see Example 4.2.3).
Classically, the Weak Nullstellensatz gives a necessary and sufcient condition
for the variety of an ideal to be empty. This applies to C
n
and P
n
as follows:
For C
n
: Given an ideal I C[x
1
, . . . , x
n
],
V(I) = in C
n
1 I.
For P
n
: Given a homogeneous ideal I C[x
0
, . . . , x
n
],
V(I) = in P
n
'x
0
, . . . , x
n
`

I for some 0.
For a toric version of the weak Nullstellensatz, we use the irrelevant ideal B() =
'x

[ ` S.
Proposition 5.2.6 (The Toric Weak Nullstellensatz). Let X

be a simplicial toric
variety with total coordinate ring S and irrelevant ideal B() S. If I S is a
homogeneous ideal, then
V(I) = in X

B()

I for some 0.
Proof. Let V
a
(I) C
(1)
denote the afne variety dened by I S. Then:
V(I) = in X

V
a
(I)

C
(1)
`Z()

=
V
a
(I) Z() = V
a
(B())
B()

I for some 0,
where the last equivalence uses the Nullstellensatz in C
(1)
.
For C
n
and P
n
, the irrelevant ideal is '1` C[x
1
, . . . , x
n
] and 'x
0
, . . . , x
n
`
C[x
0
, . . . , x
n
] respectively. For C
n
, the grading on C[x
1
, . . . , x
n
] by Cl(C
n
) = 0
is trivial, so that every ideal is homogeneous. Thus the toric weak Nullstellensatz
implies the classical version of the weak Nullstellensatz for both C
n
and P
n
.
The relation between ideals and varieties is not perfect because different ideals
can dene the same subvariety. In C
n
and P
n
, we avoid this by using radical ideals:
For C
n
: There is a bijective correspondence
closed subvarieties of C
n
radical ideals I C[x
1
, . . . , x
n
].
For P
n
: There is a bijective correspondence
closed subvarieties of P
n

radical homogeneous ideals


I 'x
0
, . . . , x
n
` C[x
0
, . . . , x
n
]

.
Here is the toric version of this correspondence.
222 Chapter 5. Homogeneous Coordinates on Toric Varieties
Proposition 5.2.7 (The Toric Ideal-Variety Correspondence). Let X

be a simpli-
cial toric variety. Then there is a bijective correspondence
closed subvarieties of X

radical homogeneous
ideals I B() S

.
Proof. Given a closed subvariety Y X

, we can nd a homogeneous ideal I S


with V(I) = Y by Proposition 5.2.4. Then

I is also homogeneous and satises
V(

I) = V(I) =Y, so we may assume that I is radical. Since


V
a
(I B()) = V
a
(I) V
a
(B()) = V
a
(I) Z()
in C
(1)
, we see that I B() B() is a radical homogeneous ideal satisfying
V(I B()) =Y. This proves surjectivity.
To prove injectivity, suppose that I, J B() are radical homogeneous ideals
with V(I) = V(J) in X

. Then
V
a
(I)

C
(1)
`Z()

= V
a
(J)

C
(1)
`Z()

.
However, I, J B() implies that Z() is contained in V
a
(I) and V
a
(J). Hence
the above equality implies
V
a
(I) = V
a
(J),
so that I = J by the Nullstellensatz since I and J are radical.
For general ideals, another way to recover injectivity is to work with closed
subschemes rather than closed subvarieties. We will say more about this in the
appendix to Chapter 6.
When X

is not simplicial, there is still a relation between ideals in the total


coordinate ring and closed subvarieties of X

.
Proposition 5.2.8. Let S be the total coordinate ring of the toric variety X

. Then:
(a) If I S is a homogeneous ideal, then
V(I) =p X

[ there is x
1
(p) with f (x) = 0 for all f I
is a closed subvariety of X

.
(b) All closed subvarieties of X

arise this way.


Proof. The proof is identical to the proof of Proposition 5.2.4.
The main difference between Propositions 5.2.4 and 5.2.8 is the phrase there
is x
1
(p). In the simplicial case, all such x are related by the group G, while
this may fail in the non-simplicial case. One consequence is that the ideal-variety
correspondence of Proposition 5.2.7 breaks down in the nonsimplicial case. Here
is a simple example.
5.2. The Total Coordinate Ring 223
Example 5.2.9. In Example 5.1.15 we described the quotient representation of
U

= C
4
//C

for the cone = Cone(e


1
, e
2
, e
1
+e
3
, e
2
+e
3
) R
3
, and in Exam-
ple 5.0.2 we saw that the quotient map
: C
4
U

= V(xy zw) C
4
is given by (a
1
, a
2
, a
3
, a
4
) =(a
1
a
3
, a
2
a
4
, a
1
a
4
, a
2
a
3
). Note that the irrelevant ideal
is B() =C[x
1
, x
2
, x
3
, x
4
].
The ideals I
1
='x
1
, x
2
` and I
2
='x
3
, x
4
` are distinct radical homogeneous ideals
contained in B() that give the same subvariety in U

:
V(I
1
) = (V
a
(I
1
)) = (C
2
0) =0 U

V(I
2
) = (V
a
(I
2
)) = (0C
2
) =0 U

.
Thus Proposition 5.2.7 fails to hold for this toric variety.
Local Coordinates. Let X

be an n-dimensional toric variety. When contains a


smooth cone of dimension n, we get an afne open set
U

with U

C
n
The usual coordinates for C
n
are compatible with the homogeneous coordinates
for X

in the following sense. The cone gives the map

: C
(1)
C
(1)
that
sends (a

)
(1)
to the point (b

)
(1)
dened by
b

(1)
1 otherwise.
Proposition 5.2.10. Let be a smooth cone of dimension n = dim X

and let

: C
(1)
C
(1)
be dened as above. Then we have a commutative diagram
C
(1)

C
(1)
`Z()

,
where the vertical maps are the quotient maps from Theorem 5.1.11. Furthermore,
the vertical map on the left is an isomorphism.
Proof. We rst show commutativity. In the proof of Theorem 5.1.11 we saw that

1
(U

) =U
e
. Since the image of

lies in U
e
, we are reduced to the diagram
C
(1)

U
e

.
224 Chapter 5. Homogeneous Coordinates on Toric Varieties
Since everything is afne, we can consider the corresponding diagram of coordi-
nate rings
C[x

[ (1)] C[x

[ (1)]
x

C[

M],

where

(
m
) =

(1)
x
/m,u)

and

(
m
) =

(1)
x
/m,u)

for m

M. It
is clear that

, and commutativity follows.


For the nal assertion, note that

is an isomorphism since the u

, (1),
form a basis of N by our assumption on . This completes the proof.
It follows that if a closed subvariety Y X

is dened by an ideal I S,
then the afne piece Y U

C
(1)
is dened by the dehomogenized ideal

I C[x

[ (1)] obtained by setting x

= 1, / (1), in all polynomials of I.


We will give examples of this below, and in 5.4, we will explore the corresponding
notion of homogenization.
Proposition 5.2.10 can be generalized to any cone satisfying dim =
dim X

(Exercise 5.2.5).
Example 5.2.11. In Example 5.1.16 we described the quotient construction of the
blowup of C
2
at the origin. This variety can be expressed as the union Bl
0
(C
2
) =
U

1
U

2
, where
1
,
2
are as in Example 5.1.16.
The map Bl
0
(C
2
) C
2
is given by (t, x, y) (tx, ty) in homogeneous coordi-
nates. Combining this with the local coordinate maps from Proposition 5.2.10, we
obtain
U

1
X

C
2
: (t, x) (t, x, 1) (tx, t)
U

2
X

C
2
: (t, y) (t, 1, y) (t, ty).
Consider the curve f (x, y) = 0 in the plane C
2
, where f (x, y) = x
3
y
2
. We study
this on the blowup Bl
0
(C
2
) using local coordinates as follows:
On U

1
, we get f (tx, t) = 0, i.e., (tx)
3
t
2
= t
2
(tx
3
1) = 0. Since t = 0
denes the exceptional locus, we get the proper transform tx
3
1 = 0.
On U

2
, we get f (t, ty) = 0, i.e., t
3
(ty)
2
= t
2
(t y
2
) = 0, with proper trans-
form t y
2
= 0.
Hence the proper transform is a smooth curve in Bl
0
(C
2
). This method of studying
the blowup of a curve is explained in many elementary texts on algebraic geometry,
such as [236, p. 100].
We relate this to the homogeneous coordinates of Bl
0
(C
2
) as follows. Using
the above map X

C
2
, we get the curve in X

dened by f (tx, ty) = 0, i.e.,


(tx)
3
(ty)
2
=t
2
(tx
3
y
2
) = 0. Hence the proper transform is tx
3
y
2
= 0. Then:
5.2. The Total Coordinate Ring 225
Setting y = 1 gives the proper transform tx
3
1 = 0 on U

1
.
Setting x = 1 gives the proper transform t y
2
= 0 on U

2
.
Hence the local proper transforms computed above are obtained from the homo-
geneous proper transform by setting appropriate coordinates equal to 1.
Exercises for 5.2.
5.2.1. Prove (5.2.1).
5.2.2. Show that the total coordinate ring of the weighted projective space P(q
0
, . . . , q
n
) is
C[x
0
, . . . , x
n
] where deg(x
i
) = q
i
. Hint: See Example 5.1.14.
5.2.3. Prove the claims made about the total coordinate ring of the product P
n
P
m
made
in Example 5.2.2.
5.2.4. Prove the claims made about the class group and the total coordinate ring of the
blowup of P
2
at the origin made in Example 5.2.3.
5.2.5. Let X

be the toric variety of the fan and assume as usual that X

has no torus
factors. A subfan

is full if

= [ (1)

(1). Consider a full subfan

with the property that X

has no torus factors.


(a) Dene the map : C

(1)
C
(1)
by sending (a

(1)
to the point (b

)
(1)
given by
b

(1)
1 otherwise.
Prove that there is a commutative diagram
C

(1)
` Z(

C
(1)
` Z()

,
where the vertical maps are the quotient maps from Theorem 5.1.11.
(b) Explain how part (a) generalizes Proposition 5.2.10.
(c) Use part (a) to give a version of Proposition 5.2.10 that applies to any cone
satisfying dim = dim X

.
5.2.6. The quintic y
2
= x
5
in C
2
has a unique singular point at the origin. We will resolve
the singularity using successive blowups.
(a) Show that the proper transform of this curve in Bl
0
(C
2
) is dened by y
2
t
3
x
5
= 0.
This uses the homogeneous coordinates t, x, y from Example 5.2.3.
(b) Show that the proper transform is smooth on U
1
but singular on U
2
.
(c) Subdivide
2
to obtain a smooth fan

. The toric variety X

has variables u, t, x, y,
where u corresponds to the ray that subdivides
2
. Show that Cl(X

) Z
2
with
deg(u) = (0, 1), deg(t) = (1, 0), deg(x) = (1, 1), deg(y) = (1, 2).
(d) Show that (u, t, x, y) (utx, u
2
ty) denes a toric morphism X

C
2
and use this to
show that the proper transform of the quintic in C
2
is dened by y
2
ut
3
x
5
= 0.
(e) Show that the proper transform is smooth by inspecting it in local coordinates.
226 Chapter 5. Homogeneous Coordinates on Toric Varieties
5.2.7. Adapt the method Exercise 5.2.6 to desingularize y
2
= x
2n+1
, n 1 an integer.
5.2.8. Given an ideal I in a commutative ring R, its Rees algebra is the graded ring
R[I] =

i=0
I
i
t
i
R[t],
where t is a new variable and I
0
= R. There is also the extended Rees algebra
R[I, t
1
] =

iZ
I
i
t
i
R[t, t
1
],
where I
i
= R for i 0. These rings are graded by letting deg(t) = 1, so that elements of R
have degree 0. See [56, 4.4] and 11.3 for more about Rees algebras.
(a) When I = 'x, y` R = C[x, y], prove that the extended Rees algebra R[I, t
1
] is the
polynomial ring C[xt, yt, t
1
]
(b) Prove that the ring of part (a) is isomorphic to the total coordinate ring of the blowup
of C
2
at the origin.
(c) Generalize parts (a) and (b) to the case of I ='x
1
, . . . , x
n
` R =C[x
1
, . . . , x
n
].
5.3. Sheaves on Toric Varieties
Given a toric variety X

, we show that graded modules over the total coordinate


ring S =C[x

[ (1)] give quasicoherent sheaves on X

. We continue to assume
that X

has no torus factors.


Graded Modules. The grading on S gives a direct sum decomposition
S =

Cl(X

)
S

such that S

S
+
for all , Cl(X

).
Denition 5.3.1. An S-module M is graded if it has a decomposition
M =

Cl(X

)
M

such that S

M
+
for all , Cl(X

). Given Cl(X

), the shift M()


is the graded S-module satisfying
M()

= M
+
for all Cl(X

).
The passage from a graded S-module to a quasicoherent sheaf on X

requires
some tools from the proof of Theorem 5.1.11. A cone gives the monomial
x

=

,(1)
x

S, and by (5.1.4), the map


m
x
/m)
=

x
/m,u)

induces an
isomorphism

: C[

M]

(S
x
)
G
S
x
,
5.3. Sheaves on Toric Varieties 227
where S
x
is the localization of S at x

. Since monomials are homogeneous, S
x
is
also graded by Cl(X

), and its elements of degree 0 are precisely its G-invariants


(Exercise 5.3.1), i.e., (S
x
)
0
= (S
x
)
G
. Hence the above isomorphism becomes
(5.3.1)

: C[

M]

(S
x
)
0
.
These isomorphisms glue together just as we would hope.
Lemma 5.3.2. Let = m

be a face of . Then (S
x
)
0
= ((S
x
)
0
)

(
m
)
, and
there is a commutative diagram of isomorphisms
(S
x
)
0
//

((S
x
)
0
)

(
m
)

C[

M]
//
C[

M]

m.
Proof. Since =m

, we have 'm, u

` =0 when (1) and 'm, u

` >0 when
(1)`(1). This means that S
x
=(S
x
)

(
m
)
. Taking elements of degree zero
commutes with localization, hence (S
x
)
0
= ((S
x
)
0
)

(
m
)
. The vertical maps
in the diagram come from (5.3.1), and the horizontal maps are localization. In
Exercise 5.3.2 you will chase the diagram to show that it commutes.
From Modules to Sheaves. We now construct the sheaf of a graded module.
Proposition 5.3.3. Let M be a graded S-module. Then there is a quasicoherent
sheaf

M on X

such that for every , the sections of



M over U

are
(U

M) = (M
x
)
0
.
Proof. Since M is a graded S-module, it is immediate that M
x
is a graded S
x
-
module. Hence (M
x
)
0
is an (S
x
)
0
-module, which induces a sheaf (M
x
)
0
on
U

= Spec(C[

M]) = Spec((S
x
)
0
). The argument of Lemma 5.3.2 applies
verbatim to show that
(M
x
)
0
= ((M
x
)
0
)

(
m
)
.
Thus the sheaves (M
x
)
0
patch to give a sheaf

M on X

which is quasicoherent by
construction.
Example 5.3.4. The total coordinate ring of P
n
is S = C[x
0
, . . . , x
n
] with the stan-
dard grading where every variable has degree 1. The quasicoherent sheaf on P
n
associated to a graded S-module was rst described by Serre in his foundational
paper Faisceaux alg ebriques coh erents [247], called FAC for short.
An important special case is when M is a nitely generated graded S-module.
We will need the following niteness result to understand the sheaf

M.
Lemma 5.3.5. (S
x
)

is nitely generated as a (S
x
)
0
-module for all and
Cl(X

).
228 Chapter 5. Homogeneous Coordinates on Toric Varieties
Proof. Write = [

] and consider the rational polyhedral cone


=(m, ) M
R
R[ 0, 'm, u

` a

for all M
R
R.
By Gordans Lemma, (MZ) is a nitely generated semigroup. Let the gen-
erators with last coordinate equal to 1 be (m
1
, 1), . . . , (m
r
, 1). Then you will prove
in Exercise 5.3.3 that the monomials

x
/m
i
,u)+a

, i = 1, . . . , r, generate (S
x
)

as a (S
x
)
0
-module.
Here are some coherent sheaves on X

.
Proposition 5.3.6. The sheaf

M on X

is coherent when M is a nitely generated


graded S-module.
Proof. Because M is graded, we may assume its generators are homogeneous of
degrees
1
, . . . ,
r
. Given , it follows immediately that M
x
is nitely gen-
erated over S
x
with generators in the same degrees. However, we need to be
careful when taking elements of degree 0. Multiply a generator of degree
i
by
the (S
x
)
0
-module generators of (S
x
)

i
(nitely many by the previous lemma).
Doing this for all i gives nitely many elements in (M
x
)
0
that generate (M
x
)
0
as
an (S
x
)
0
-module (Exercise 5.3.3). It follows that

M is coherent.
Given Cl(X

), the shifted S-module S() gives a coherent sheaf on X

denoted O
X

(). This is a sheaf we already know.


Proposition 5.3.7. Fix Cl(X

). Then:
(a) There is a natural isomorphism S

(X

, O
X

()).
(b) If D =

is a Weil divisor satisfying = [D], then


O
X

(D) O
X

().
Proof. By denition, the sections of O
X

() over U

are
(U

, O
X

()) = (S()
x
)
0
= (S
x
)

for . Since the open cover U

of X

satises U

= U

, the
sheaf axiom gives the exact sequence
0 (X

, O
X

())

(S
x
)

,
(S
x

.
The localization (S
x
)

has a basis consisting of all Laurent monomials

x
b

of
degree such that b

0 for all (1). Then the exact sequence implies that


(X

, O
X

()) has a basis consisting of all Laurent monomials

x
b

of degree
such that b

0 for all (1). These are precisely the monomials in S of degree


, which gives the desired isomorphism S

(X

, O
X

()).
We turn to part (b). Given a Weil divisor D =

with = [D], we need


to construct a sheaf isomorphism O
X

(D) O
X

(). By the above description of


5.3. Sheaves on Toric Varieties 229
the sections over U

, it sufces to prove that for every , we have isomor-


phisms
(5.3.2) (U

, O
X

(D)) (S
x
)

.
compatible with inclusions U

induced by _ in .
To construct this isomorphism, we apply Proposition 4.3.3 to U

to obtain
(U

, O
X

(D)) =

mM
/m,u)a, (1)
C
m
.
A lattice point m M gives the Laurent monomial
(5.3.3) x
/m,D)
=

x
/m,u)+a

.
When 'm, u

` a

for (1), this lies in S


x
, and in fact x
/m,D)
(S
x
)

since
deg(x
/m,D)
) =

('m, u

` +a

)D

div(
m
) +D

= [D] = .
We claim that map
m
x
/m,D)
induces the desired isomorphism (5.3.2).
Suppose that
m
,
m

map to the same monomial. Then 'm, u

` = 'm

, u

`
for all . This implies m = m

since X

has no torus factors. Furthermore, if


x
b
=

x
b

(S
x
)

, then [

] = = [

], so that there is m M
such that b

= 'm, u

` +a

for all . Since x


b
is a monomial in S
x
, b

0 for
(1), hence 'm, u

` a

for (1). Then


m
(U

, O
X

(D)) maps to
x
b
. This denes an isomorphism (5.3.2) which is easily seen to be compatible with
the inclusion of faces.
Example 5.3.8. For P
n
we have S = C[x
0
, . . . , x
n
] with the standard grading by
Z = Cl(P
n
). Then O
P
n (k) is the sheaf associated to S(k) for k Z. The classes of
the toric divisors D
0
D
n
correspond to 1 Z, so that
O
P
n (k) O
P
n (kD
0
) O
P
n (kD
n
).
Thus O
P
n (k) is a canonical model for the sheaf O
P
n (kD
i
). This justies what we
did in Example 4.3.1.
Also note that when k 0, we have
(P
n
, O
P
n (k)) = S
k
.
Hence global sections of O
P
n (k) are homogeneous polynomials in x
0
, . . . , x
n
of
degree k, which agrees with what we computed in Example 4.3.6.
230 Chapter 5. Homogeneous Coordinates on Toric Varieties
Sheaves versus Modules. An important result is that all quasicoherent sheaves on
X

come from graded modules.


Proposition 5.3.9. Let F be a quasicoherent sheaf on X

. Then:
(a) There is a graded S-module M such that

MF.
(b) If F is coherent, then M can be chosen to be nitely generated over S.
The proof will be given in the appendix to Chapter 6 since it involves tensor
products of sheaves from 6.0.
Although the map M

M is surjective (up to isomorphism), it is far from
injective. In particular, there are nontrivial graded modules that give the trivial
sheaf. This phenomenon is well-known for P
n
, where a nitely generated graded
module M over S =C[x
0
, . . . , x
n
] gives the trivial sheaf on P
n
if and only if M

= 0
for 0 (see [131, Ex. II.5.9]). This is equivalent to
'x
0
, . . . , x
n
`

M = 0
for 0 (Exercise 5.3.4). Since 'x
0
, . . . , x
n
` is the irrelevant ideal for P
n
, this
suggests a toric generalization. In the smooth case, we have the following result.
Proposition 5.3.10. Let B() S be the irrelevant ideal of S for a smooth toric
variety X

, and let M be a nitely generated graded S-module. Then



M = 0 if and
only if B()

M = 0 for 0.
Proof. First observe that

M=0 if and only if it vanishes on each afne open subset
U

. But on an afne variety, the correspondence between quasicoherent


sheaves and modules is bijective (see [131, Cor. II.5.5]). Hence

M = 0 if and only
if (M
x
)
0
= 0 for all .
Next suppose that B()

M=0 for some 0. Then (x



)

M=0, which easily


implies that M
x
= 0. Then

M = 0 follows from the previous paragraph. This part
of the argument works for any toric variety.
For the converse, we have (M
x
)
0
= 0 for all . Given h M

, we will
show that (x

)

h =0 for some 0, which will imply B()

M=0 for 0 since


M is nitely generated. Let = [D], where D =

. Since is smooth, D is
Cartier, so there is m

M such that 'm

, u

` = a

for all (1) (this is part


of the Cartier data for D). Replacing D with D+div(
m
), we may assume that
D =

/ (1)
a

. Now set k = max(0, a

[ / (1)) and observe that


x
b
= (x

)
k

/ (1)
x
a

/ (1)
x
ka

S.
Furthermore, x
b
h/(x

)
k
M
x
has degree 0. Hence x
b
h/(x

)
k
= 0 in M
x
, which
by the denition of localization implies that there is s 0 with
(x

)
s
x
b
h = 0 in M.
5.4. Homogenization and Polytopes 231
Since x
b
involves only x

for / (1), we can nd x


a
S such that x
a
x
b
is a
power of x

. Hence multiplying the above equation by x
a
implies (x

)

h = 0 for
some 0, as desired.
Unfortunately, the situation is more complicated when X

is not smooth. Here


is an example to show what can go wrong when X

is simplicial.
Example 5.3.11. The weighted projective space P(1, 1, 2) has total coordinate
ring S = C[x, y, z], where x, y have degree 1 and z has degree 2, and the irrele-
vant ideal is B() = 'x, y, z`. The graded S-module M = S(1)/(xS(1) +yS(1))
has only elements of odd degree. Then (M
z
)
0
= 0 since z has degree 2, and it is
clear that (M
x
)
0
= (M
y
)
0
= 0. It follows that

M = 0, yet one easily checks that
B()

M = z

M= 0 for all 0. Thus Proposition 5.3.10 fails for P(1, 1, 2).


Exercise 5.3.5 explores a version of Proposition 5.3.10 that applies to simpli-
cial toric varieties. The condition that B()

M = 0 is replaced with the weaker


condition that B()

= 0 for all Pic(X

).
We will say more about the relation between quasicoherent sheaves and graded
S-modules in the appendix to Chapter 6.
Exercises for 5.3.
5.3.1. As described in 5.0, the action of G on C
(1)
induces an action of G on the total
coordinate ring S. Also recall that g G is a homomorphism g : Cl(X

) C

.
(a) Given x
a
S and g G, show that g x
a
= g
1
()x
a
, where = deg(x
a
).
(b) Show that S
G
= S
0
and that a similar result holds for the localization S
x
b .
5.3.2. Complete the proof of Lemma 5.3.2.
5.3.3. Complete the proofs of Lemma 5.3.5 and Proposition 5.3.6.
5.3.4. Let S = C[x
0
, . . . , x
n
] where deg(x
i
) = 1 for all i, and let M be a nitely generated
graded S-module. Show that M

= 0 for 0 if and only if 'x


0
, . . . , x
n
`

M= 0 for 0.
5.3.5. Let X

be a simplicial toric variety and let M be a nitely generated graded S-


module. Prove that

M= 0 if and only if B()

= 0 for all 0 and Pic(X

).
5.3.6. Let X

be a smooth toric variety. State and prove a version of Proposition 5.3.10


that applies to arbitrary graded S-modules M. Also explain what happens when X

is
simplicial, as in Exercise 5.3.5.
5.4. Homogenization and Polytopes
The nal section of the chapter will explore the relation between torus-invariant
divisors on a toric variety X

and its total coordinate ring. We will also see that


when X

comes from a polytope P, the quotient construction of X

relates nicely
to the denition of projective toric variety given in Chapter 2.
232 Chapter 5. Homogeneous Coordinates on Toric Varieties
Homogenization. When working with afne and projective space, one often needs
to homogenize polynomials. This process generalizes nicely to the toric context.
The full story involves characters, polyhedra, divisors, sheaves, and graded pieces
of the total coordinate ring.
A Weil divisor D =

on X

gives the polyhedron


P
D
=m M
R
[ 'm, u

` a

for all (1).


Proposition 4.3.3 tells us that the global sections of the sheaf O
X

(D) are spanned


by characters coming from lattice points of P
D
, i.e.,
(X

, O
X

(D)) =

mP
D
M
C
m
.
This relates to the total coordinate ring S =C[x

[ (1)] as follows. Given


m P
D
M, the D-homogenization of
m
is the monomial
x
/m,D)
=

x
/m,u)+a

dened in (5.3.3). The inequalities dening P


D
guarantee that x
/m,D)
lies in S. Here
are the basic properties of these monomials.
Proposition 5.4.1. Assume that X

has no torus factors. If D and P


D
are as above
and = [D] Cl(X

) is the divisor class of D, then:


(a) For each m P
D
M, the monomial x
/m,D)
lies in S

.
(b) The map sending the character
m
of m P
D
M to the monomial x
/m,D)
induces an isomorphism
(X

, O
X

(D)) S

.
Proof. Part (a) follows from the proof of Proposition 5.3.7. As for part (b), we use
the same proposition to conclude that
(X

, O
X

(D)) (X

, O
X

()) S

.
One easily sees that this isomorphism is given by
m
x
/m,D)
.
Here are some examples of homogenization.
Example 5.4.2. The fan for P
n
has ray generators u
0
=

n
i=1
e
i
and u
i
= e
i
for
i = 1, . . . , n. This gives variables x
i
and divisors D
i
for i = 0, . . . , n. Since M = Z
n
,
the character of m = (b
1
, . . . , b
n
) Z
n
is the Laurent monomial t
m
=

n
i=1
t
b
i
i
.
5.4. Homogenization and Polytopes 233
For a positive integer d, the divisor D= d D
0
has polyhedron P
D
= d
n
, where

n
is the standard n-simplex. Given m= (b
1
, . . . , b
n
) d
n
, its homogenization is
x
/m,D)
= x
/m,u
0
)+d
0
x
/m,u
1
)+0
1
x
/m,un)+0
n
= x
b
1
bn+d
0
x
b
1
1
x
bn
n
= x
d
0

x
1
x
0

b
1

x
n
x
0

bn
,
which is the usual way to homogenize t
m
=

n
i=1
t
b
i
i
with respect to x
0
.
This monomial has degree d =[dD
0
] Cl(P
n
) =Z, in agreement with Proposi-
tion 5.4.1. The proposition also implies the standard fact that monomials of degree
d in x
0
, . . . , x
n
correspond to lattice points in d
n
.
Example 5.4.3. For P
1
P
1
, we have ray generators u
1
= e
1
, u
2
= e
1
, u
3
=
e
2
, u
4
= e
2
with corresponding variables x
i
and divisors D
i
. Given nonnegative
integers k, , we get the divisor D = kD
2
+D
4
. The polyhedron P
D
is the rectan-
gle with vertices (0, 0), (k, 0), (0, ), (k, ), and given (a, b) P
D
Z
2
, the Laurent
monomial t
a
1
t
b
2
homogenizes to
x
a
1
x
ka
2
x
b
3
x
b
4
= x
k
2
x

x
1
x
2

x
3
x
4

b
,
which is the usual way of turning a two-variable monomial into a bihomoge-
neous monomial of degree (k, ) (remember that deg(x
1
) = deg(x
2
) = (1, 0) and
deg(x
3
) = deg(x
4
) = (0, 1)). Thus monomials of degree (k, ) correspond to lattice
points in the rectangle P
D
.
Example 5.4.4. The fan for Bl
0
(C
2
) is shown in Example 5.1.16, and its total
coordinate ring S =C[t, x, y] is described in Example 5.2.3. If we pick D = 0, then
the polyhedron P
D
R
2
is dened by the inequalities
'm, u
i
` 0, i = 0, 1, 2.
Since u
1
, u
2
form a basis of N =Z
2
and u
0
= u
1
+u
2
, P
D
is the rst quadrant in R
2
.
Given m = (a, b) P
D
Z
2
, the monomial t
a
1
t
b
2
homogenizes to
t
/m,u
0
)
x
/m,u
1
)
y
/m,u
2
)
= t
a+b
x
a
y
b
= (tx)
a
(ty)
b
.
where the ray generators u
0
, u
1
, u
2
correspond to the variables t, x, y.
For example, the singular cubic t
3
1
t
2
2
= 0 homogenizes to (tx)
3
(ty)
2
= 0,
which is the equation enountered in Example 5.2.11 when resolving the singularity
of this curve.
One thing to keep in mind when doing toric homogenization is that characters

m
(in general) or Laurent monomials t
m
(in specic examples) are intrinsically
dened on the torus T
N
or (C

)
n
. The homogenization process produces a global
object x
/m,D)
relative to a divisor D that lives in the total coordinate ring or, via
Proposition 5.4.1, in the global sections of O
X

(D).
234 Chapter 5. Homogeneous Coordinates on Toric Varieties
We next study the isomorphisms S

(X

, O
X

(D)) from Proposition 5.4.1.


We will see that they are compatible with linear equivalence and multiplication.
First suppose that D and E are linearly equivalent torus-invariant divisors. This
means that D = E +div(
m
) for some m M. Proposition 4.0.29 implies that
f f
m
induces an isomorphism
(5.4.1) (X

, O
X

(D)) (X

, O
X

(E)).
Turning to the associated polyhedra, we proved P
E
= P
D
+m in Exercise 4.3.2. An
easy calculation shows that if m

P
D
, then
x
/m

,D)
= x
/m

+m,E)
(Exercise 5.4.1). Hence (5.4.1) ts into a commutative diagram of isomorphisms
(5.4.2)
(X

, O
X

(D))

(X

, O
X

(E))

.
Here, = [D] = [E] Cl(X

) and the diagonal maps are the isomorphisms from


Proposition 5.4.1. You will verify these claims in Exercise 5.4.1.
It follows that S

gives a canonical model for (X

, O
X

(D)), since the latter


depends on the particular choice of divisor D in the class . It is also possible to
give a canonical model for the polyhedron P
D
(Exercise 5.4.2).
Next consider multiplication. Let D and E be torus-invariant divisors on X

and set = [D], = [E] in Cl(X

). Then f g f g induces a C-linear map


(X

, O
X

(D))
C
(X

, O
X

(E)) (X

, O
X

(D+E))
such that the isomorphisms of Proposition 5.4.1 give a commutative diagram
(5.4.3)
(X

, O
X

(D))
C
(X

, O
X

(E))

(X

, O
X

(D+E))

C
S


S
+
where the bottom map is multiplication in the total coordinate ring (Exercise 5.4.3).
Thus homogenization turns multiplication of sections into ordinary multiplication.
Polytopes. A full dimensional lattice polytope P M
R
gives a toric variety X
P
.
Recall that X
P
can be constructed in two ways:
As the toric variety X

P
of the normal fan
P
of P (Chapter 3).
As the projective toric variety X
(kP)M
of the set of characters (kP) M for
k 0 (Chapter 2).
We will see that both descriptions relate nicely to homogeneous coordinates and
the total coordinate ring.
5.4. Homogenization and Polytopes 235
Given P as above, set n = dim P and let P(i) denote the set of i-dimensional
faces of P. Thus P(0) consists of vertices and P(n1) consists of facets. The facet
presentation of P given in equation (2.2.2) can be written as
(5.4.4) P =m M
R
[ 'm, u
F
` a
F
for all F P(n1).
In terms of the normal fan
P
, we have bijections
P(0)
P
(n) (vertices maximal cones)
P(n1)
P
(1) (facets rays).
When dealing with polytopes we index everything by facets rather than rays. Thus
each facet F P(n1) gives:
The facet normal u
F
, which is the ray generator of the corresponding cone.
The torus-invariant prime divisor D
F
X
P
.
The variable x
F
in the total coordinate ring S. We call x
F
a facet variable.
We also have the divisor
D
P
=

F
a
F
D
F
from (4.2.7). The polytope P
D
P
of this divisor is the polytope P we began with
(Exercise 4.3.1). Hence, if we set = [D
P
] Cl(X
P
), then we get isomorphisms
S

(X
P
, O
X
P
(D
P
))

mPM
C
m
.
In this situation, we write the D
P
-homogenization of
m
as
x
/m,P)
=

F
x
/m,u
F
)+a
F
F
.
We call x
/m,P)
a P-monomial.
The exponent of the variable x
F
in x
/m,P)
gives the lattice distance from m
to the facet F. To see this, note that F lies in the supporting hyperplane dened
by 'm, u
F
` +a
F
= 0. If the exponent of x
F
is a 0, then to get from the sup-
porting hyperplane to m, we must pass through the a parallel hyperplanes, namely
'm, u
F
` +a
F
= j for j = 1, . . . , a. Here is an example.
Example 5.4.5. Consider the toric variety X
P
of the polygon P R
2
with vertices
(1, 1), (1, 1), (1, 0), (0, 1), (1, 1), shown in Figure 2 on the next page. In
terms of (5.4.4), we have a
1
= = a
5
= 1, where the indices correspond to the
facet variables x
1
, . . . , x
5
indicated in Figure 2. The 8 points of PZ
2
give the
P-monomials
x
2
x
2
3
x
2
4
x
1
x
2
2
x
2
3
x
4
x
2
1
x
3
2
x
2
3
x
3
x
2
4
x
5
x
1
x
2
x
3
x
4
x
5
x
2
1
x
2
2
x
3
x
5
x
1
x
4
x
2
5
x
2
1
x
2
x
2
5
,
236 Chapter 5. Homogeneous Coordinates on Toric Varieties
d
d
d
x
1
x
5
x
2
x
4
x
3
Figure 2. A polygon with facets labeled by variables
where the position of each P-monomial x
/m,P)
corresponds to the position of the
lattice point m PZ
2
. The exponents are easy to understand if you think in
terms of lattice distances to facets.
The lattice-distance interpretation of the exponents in x
/m,P)
shows that lattice
points in the interior int(P) of P correspond to those P-monomials divisible by

F
x
F
. For example, the only P-monomial in Example 5.4.5 divisible by x
1
x
5
corresponds to the unique interior lattice point.
We next relate the constructions of toric varieties given in Chapter 2 and in
5.1. In Chapter 2, we wrote the lattice points of P as PM = m
1
, . . . , m
s
and
considered the map
(5.4.5) : T
N
P
s1
, t (
m
1
(t), . . . ,
ms
(t)).
The projective (possibly non-normal) toric variety X
PM
is the Zariski closure of
the image of .
On the other hand, we have the quotient construction of X
P
X
P

C
r
`Z(
P
)

//G,
where we write C
r
= C

P
(1)
. Also, the exceptional set Z(
P
) can be described in
terms of the P-monomials coming from the vertices of the polytope.
Lemma 5.4.6. The vertex monomials x
/v,P)
, v a vertex of P, have the following
properties:
(a)

'x
/v,P)
[ v P(0)` =B(
P
), where B(
P
) ='x

[ (n)` is the irrelevant
ideal of S.
(b) Z(
P
) = V(x
/v,P)
[ v P(0)).
Proof. We saw above that vertices v P(0) correspond bijectively to cones
v
=
Cone(u
F
[ v F)
P
(n). Then the lattice-distance interpretation of x
/v,P)
shows
the facet variables x
F
appearing in x
/v,P)
are precisely the variables appearing in
x
v
. This implies part (a), and part (b) follows immediately.
5.4. Homogenization and Polytopes 237
If we set = [D
P
] as above, then the P-monomials x
/m
i
,P)
, i = 1, . . . , s, form a
basis of S

and give a map


(5.4.6) : C
r
`Z(
P
) P
s1
p (p
/m
1
,P)
, . . . , p
/ms,P)
),
where p
/m
i
,P)
is the evaluation of the monomial x
/m
i
,P)
at the point p C
r
`Z(
P
).
This map is well-dened since for each p C
r
`Z(
P
), Lemma 5.4.6 implies that
at least one P-monomial (in fact, at least one vertex monomial) must be nonzero.
The maps (5.4.5) and (5.4.6) t into a diagram
(C

)
r

C
r
`Z(
P
)

T
N



X
P

P
s1
.
Here, the map (C

)
r
T
N
is from (5.1.2) and : C
r
`Z(
P
) X
P
is the quotient
map. This diagram has the following properties.
Proposition 5.4.7. There is a morphism : X
P
P
s1
represented by the dotted
arrow in the above diagram that makes the entire diagram commute. Furthermore,
the image of is precisely the projective toric variety X
PM
.
Proof. When we regard the x
F
as characters on (C

)
r
= (C

P
(1)
, the exact se-
quence (5.1.1) tells us that
(5.4.7)
m
=

F
x
/m,u
F
)
F
for m M. Multiplying each side by

F
x
a
F
F
, we obtain

F
x
a
F
F

m
= x
/m,D)
.
If we let m=m
i
, i =1, . . . , s and apply this to a point in p (C

)
r
, we see that (p)
and (p) give the same point in projective space since the vector for (p) equals

F
p
a
F
F
times the vector for (p). It follows that, ignoring for the moment, the
rest of the above diagram commutes.
We next show that is constant on G-orbits. This holds since P-monomials
are homogeneous of the same degree. In more detail, x points t = (t
F
) G,
p =(p
F
) C
r
`Z(
P
) and a P-monomial x
/m,D)
=

F
x
/m,u
F
)+a
F
F
. Then evaluating
238 Chapter 5. Homogeneous Coordinates on Toric Varieties
x
/m,D)
at t p gives
(t p)
/m,D)
=

F
(t
F
p
F
)
/m,u
F
)+a
F
=

F
t
/m,u
F
)
F

F
t
a
F
F

p
/m,D)
=

F
t
a
F
F

p
/m,D)
,
where the last equality follows from the description of G given in Lemma 5.1.1.
Arguing as in the previous paragraph, it follows that (t p) and (p) give the
same point in P
s1
. This proves the existence of since is a good categorical
quotient, and this choice of makes the entire diagram commute.
The nal step is to show that the image of : X
P
P
s1
is the Zariski closure
X
PM
of the image of : T
N
P
s1
. First observe that
(X
P
) = (T
N
) (T
N
) = (T
N
) = X
PM
since is continuous in the Zariski topology and [
T
N
= by commutativity of the
diagram. However, (X
P
) is Zariski closed in P
s1
since X
P
is projective. You will
give two proofs of this in Exercise 5.4.4, one topological (using constructible sets
and compactness) and one algebraic (using completeness and properness). Once
we know that (X
P
) is Zariski closed, (T
N
) (X
P
) implies
X
PM
= (T
N
) (X
P
),
and (X
P
) = X
PM
follows.
In Chapter 2, we used the map , constructed from characters, to parametrize
a big chunk of the projective toric variety X
PM
. In contrast, Proposition 5.4.7 uses
the map , constructed from P-monomials, to parametrize all of X
PM
.
If the lattice polytope P is very ample, then the results of Chapter 2 imply that
X
PM
is the toric variety X
P
. So in the very ample case, the P-monomials give an
explicit construction of the quotient

C
r
` Z(
P
)

//G by mapping C
r
` Z(
P
) to
projective space via the P-monomials. It follows that we have two ways to take the
quotient of C
r
by G:
At the beginning of the chapter, we took G-invariant polynomialselements
of S
0
to construct an afne quotient.
Here, we use P-monomialselements of S

to construct a projective quo-


tient, after removing a set Z(
P
) of bad points.
The P-monomials are not G-invariant but instead transform the same way under G.
This is why we map to projective space rather than afne space. We will explore
these ideas further in Chapter 14 when we discuss geometric invariant theory.
5.4. Homogenization and Polytopes 239
When P is very ample, we have a projective embedding X
P
P
s1
given by
the P-monomials in S

. If y
1
, . . . , y
s
are homogeneous coordinates of P
s1
, then
the homogeneous coordinate ring of X
P
P
s1
is
C[X
P
] =C[y
1
, . . . , y
s
]/I(X
P
)
as in 2.0. We also have the afne cone

X
P
C
s
of X
P
, and C[X
P
] is the ordinary
coordinate ring of

X
P
, i.e.,
C[X
P
] =C[

X
P
].
Recall that C[X
P
] is an N-graded ring since I(X
P
) is a homogeneous ideal.
Another N-graded ring is

k=0
S
k
. This relates to C[X
P
] as follows.
Theorem 5.4.8. Let P be a very ample lattice polytope with = [D
P
] Cl(X
P
).
Then:
(a)

k=0
S
k
is normal.
(b) There is a natural inclusion C[X
P
]

k=0
S
k
such that

k=0
S
k
is the nor-
malization of C[X
P
].
(c) The following are equivalent:
(1) X
P
P
s1
is projectively normal.
(2) P is normal.
(3)

k=0
S
k
=C[X
P
].
(4)

k=0
S
k
is generated as a C-algebra by its elements of degree 1.
Proof. Consider the cone
C(P) = Cone(P1) M
R
R.
This cone is pictured in Figure 4 of 2.2. Recall that kP is the slice of C(P)
at height k. Since the divisor D
kP
associated to kP is kD
P
, homogenization with
respect to kP induces an isomorphism
S
k
(X
P
, O
X
P
(kD
P
))

m(kP)M
C
m
.
Now consider the dual cone
P
= C(P)

N
R
R. The semigroup algebra
C[C(P) (MZ)] is the coordinate ring of the afne toric variety U

P
. Given
(m, k) C(P) (MZ), we write the corresponding character as
m
t
k
.
The algebra C[C(P)(MZ)] is graded using the last coordinate, the height.
Since (m, k) C(P) (MZ) if and only if m kP (this is the slice observation
made above), we have
C[C(P) (MZ)]
k
=

m(kP)M
C
m
t
k
.
240 Chapter 5. Homogeneous Coordinates on Toric Varieties
Using (5.4.3), we obtain a graded C-algebra isomorphism

k=0
S
k
C[C(P) (MZ)].
This proves that

k=0
S
k
is normal.
We next claim that U

P
is the normalization of the afne cone

X
P
. For this,
we let A = (PM) 1 MZ. As noted in the proof of Theorem 2.4.1, the
afne cone of X
P
= X
PM
is

X
P
= Y
A
. Since P is very ample, one easily checks
that A generates MZ, i.e., ZA = MZ (Exercise 5.4.5). It is also clear that A
generates the cone C(P) =

P
. Hence U

P
is the normalization of

X
P
by Proposi-
tion 1.3.8. This immediately implies part (b).
For part (c), we observe that (1) (2) follows from Theorem 2.4.1, and (1)
(3) follows from parts (a) and (b) since the projective normality of X
P
P
s1
is equivalent to the normality of C[X
P
]. Also (3) (4) is obvious since C[X
P
] is
generated by the images of y
1
, . . . , y
s
, which have degree 1. Finally, you will show
in Exercise 5.4.6 that (4) (2), completing the proof.
Further Examples. We begin with an example of that illustrates how there can be
many different polytopes that give the same toric variety.
Example 5.4.9. The toric surface in Example 5.4.5 was dened using the polygon
shown in Figure 2. In Figure 3 we see four polygons A, AB, AC, AD, all
d
d
d
d
d
d
d
d
d
A
B
C
D
Figure 3. Four polygons A, AB, AC, ADwith the same normal fan
of which have the same normal fan and hence give the same toric variety. Since
we are in dimension 2, these polygons are very ample (in fact, normal), so that
Theorem 5.4.8 applies.
These four polygons give four different projective embeddings, each of which
has its own homogeneous coordinate ring as a projective variety. By Theorem 5.4.8,
these homogeneous coordinate rings all live in the total coordinate ring S. This ex-
plains the total in total coordinate ring.
5.4. Homogenization and Polytopes 241
Our next example involves torsion in the grading of the total coordinate ring.
Example 5.4.10. The fan for P
4
has ray generators u
0
=

4
i=1
e
i
and u
i
= e
i
for i = 1, . . . , 4 in N =Z
4
and is the normal fan of the standard simplex
4
R
4
.
Another polytope with the same normal fan is
P = 5
4
(1, 1, 1, 1) M
R
=R
4
,
so that X
P
= P
4
. We saw that P is reexive in Example 2.4.5. One checks that
D
P
= D
0
+ +D
4
has degree 5 Z Cl(P
4
). Since P is a translate of 5
4
,
(5.4.2) implies that the P-monomials for m PZ
4
coincide with the homoge-
nizations coming from 5
4
, which are homogeneous polynomials of degree 5 in
S =C[x
0
, . . . , x
4
].
Since P is reexive, its dual P

is also a lattice polytope. Furthermore,


P

= Conv(u
0
, . . . , u
4
) N
R
=R
4
since the ray generators of the normal fan of P

are the vertices of P by duality for


reexive polytopes (be sure you understand thisExercise 5.4.7). The vertices of
P are
(5.4.8)
v
0
= (1, 1, 1, 1), v
1
= (4, 1, 1, 1), v
2
= (1, 4, 1, 1)
v
3
= (1, 1, 4, 1), v
4
= (1, 1, 1, 4).
The v
i
generate a sublattice M
1
M =Z
4
. In Exercise 5.4.7 you will show that the
map M Z
5
dened by
m M ('m, u
0
`, . . . , 'm, u
4
`) Z
5
induces an isomorphism
(5.4.9) M/M
1

(a
0
, a
1
, a
2
, a
3
, a
4
) (Z/5Z)
5
:

4
i=0
a
i
= 0

/(Z/5Z)
where Z/5Z (Z/5Z)
5
is the diagonal subgroup. Then M/M
1
(Z/5Z)
3
, so that
M
1
is a lattice of index 125 in M.
The dual toric variety X
P
is determined by the normal fan

of P

. The
ray generators of

are the vectors v


0
, . . . , v
4
from (5.4.8). The only possible
complete fan in R
4
with these ray generators is the fan whose cones are generated
by all proper subsets of v
0
, . . . , v
4
. Since v
0
+ +v
4
= 0 and the v
i
generate M
1
,
the toric variety of

relative to M
1
is P
4
, i.e., X

, M
1
= P
4
. (Remember that

is a fan in (M
1
)
R
= M
R
.) Since M
1
M has index 125, Proposition 3.3.7 implies
X
P
= X
P

, M
X
P

, M
1
/(M/M
1
) =P
4
/(M/M
1
).
Hence the dual toric variety X
P
is the quotient of P
4
by a group of order 125.
The total coordinate ring S

is the polynomial ring C[y


0
, . . . , y
4
], graded by
Cl(X
P
). The notation is challenging, since by duality N is the character lattice of
the torus of X
P
. Thus (5.1.1) becomes the short exact sequence
0 N Z
5
Cl(X
P
) 0,
242 Chapter 5. Homogeneous Coordinates on Toric Varieties
where N Z
5
is u ('v
0
, u`, . . . , 'v
4
, u`). If we let N
1
= Hom
Z
(M
1
, Z), then
M
1
M dualizes to N N
1
of index 125. Now consider the diagram
0

0

N

Z
5

Cl(X
P
)

0
0

N
1

Z
5
Z

0
N
1
/N

0
with exact rows and columns. In the middle row, we use Cl(X

,M
1
) =Cl(P
4
) =Z.
By the snake lemma, we obtain the exact sequence
0 N
1
/N Cl(X
P
) Z 0,
so Cl(X
P
) ZN/N
1
. Thus the class group has torsion.
The polytope P

has only six lattice points in N: the vertices u


0
, . . . , u
4
and the
origin (Exercise 5.4.7). When we homogenize these, we get six P

-monomials
y
/0,D)
=
4

j=0
y
/v
j
,0)+1
j
= y
0
y
4
y
/u
i
,D)
=
4

j=0
y
/v
j
,u
i
)+1
j
= y
5
i
, i = 0, . . . , 4
since 'v
j
, u
i
` = 5
i j
1 (Exercise 5.4.7).
The equation
c
0
y
5
0
+ +c
4
y
5
4
+c
5
y
0
y
4
= 0
denes a hypersurface Y X
P
since it is built from P

-monomials. If we want an
irreducible hypersurface, we must have c
0
, . . . , c
4
= 0, in which case Y is isomor-
phic (via the torus action) to a hypersurface of the form
y
5
0
+ +y
5
4
+y
0
y
4
= 0.
This is the quintic mirror family, which played a crucial role in the development of
mirror symmetry. See [68] for an introduction to this astonishing subject.
Exercises for 5.4.
5.4.1. Let D, E be linearly equivalent torus-invariant divisors with D = div(
m
) +E.
(a) If m

P
D
M, then prove that x
m

,D
= x
m

+m,E
.
(b) Prove (5.4.2).
5.4. Homogenization and Polytopes 243
5.4.2. Fix a torus-invariant divisor D =

and consider its associated polyhedron


P
D
=m M
R
[ 'm, u

` a

for all . Dene

D
: M
R
R
(1)
by
D
(m) = ('m, u

` +a

) R
(1)
.
(a) Prove that
D
embeds M
R
as an afne subspace of R
(1)
. Hint: Remember that X

has no torus factors.


(b) Prove that
D
induces a bijection

D[
PD
: P
D

D
(M
R
) R
(1)
0
.
This realizes P
D
as the polyhedron obtained by intersecting the positive orthant R
(1)
0
of R
(1)
with an afne subspace.
(c) Let D = div(
m
) +E. Prove that
D
(P
D
) =
E
(P
E
). Thus the polyhedron in R
(1)
constructed in part (b) depends only on the divisor class of D. This is the canonical
model of P
D
.
5.4.3. Prove that the diagram (5.4.3) is commutative.
5.4.4. The proof of Proposition 5.4.7 claimed that the image of : X
P
P
s1
was Zariski
closed. This follows from the general fact that if : X Y is a morphism of varieties and
X is complete, then (X) is Zariski closed in Y. You will prove this two ways.
(a) Give a topological proof that uses constructible sets and compactness. Hint: Remem-
ber that projective space is compact.
(b) Give an algebraic proof that uses completeness and properness from 3.4. Hint: Show
that X Y Y is proper and use the graph of .
5.4.5. Let P M
R
be a very ample lattice polytope and let A = (PM) 1 MZ.
Prove that ZA = MZ. Hint: First show that Z

A = M0, where Z

A is dened in
the discussion preceding Proposition 2.1.6.
5.4.6. Prove of (4) (2) in part (c) of Theorem 5.4.8. Hint: (4) implies that the map
S

C
S
k
S
(k+1)
is onto for all k 0.
5.4.7. This exercise is concerned with Example 5.4.10.
(a) Prove that if P R
n
is reexive, then the vertices of P are the ray generators of the
normal fan of P

.
(b) Prove (5.4.9).
(c) Prove 'v
j
, u
i
` = 5
i j
1, where v
j
, u
i
are dened in Example 5.4.10.
(d) Let G = Hom
Z
(Cl(X
P
), C

) (C

)
5
. Use Proposition 1.3.18 to prove
G =(
0
, . . . ,
4
) [ C

,
i

5
,
0

4
= 1 C

M/M
1
.
(e) Use part (e) and the quotient construction of X
P
to give another proof that X
P
=
P
4
/(M/M
1
). Also give an explicit description of the action of M/M
1
on P
4
.
5.4.8. This exercise will give another way to think about homogenization. Let e
1
, . . . , e
n
be a basis of M, so that t
i
=
ei
, i = 1, . . . , n, are coordinates for the torus T
N
.
(a) Adapt the proof of (5.4.7) to show that t
i
=

x
ei ,u

when we think of the x

as
characters on (C

)
(1)
.
244 Chapter 5. Homogeneous Coordinates on Toric Varieties
(b) Given m P
D
M, part (a) tells us that the Laurent monomial t
m
can be regarded as a
Laurent monomial in the x

. Show that we can clear denominators by multiplying


by

x
a

to obtain a monomial in the polynomial ring S =C[x

[ (1)].
(c) Show that this monomial obtained in part (b) is the homogenization x
m,D
.
5.4.9. Consider the toric variety X
P
of Example 5.4.5.
(a) Compute Cl(X
P
) and nd the classes of the four polygons appearing in Figure 3.
(b) Show that X
P
is the blowup of P
1
P
1
at one point.
5.4.10. Consider the reexive polytope P = 4
3
(1, 1, 1) R
3
. Work out the analog of
Example 5.4.10 for P.
5.4.11. Fix an integer a 1 and consider the 3-simplex P = Conv(0, ae
1
, ae
2
, e
3
) R
3
.
In Exercise 2.2.13, we claimed that the toric variety of P is the weighted projective space
P(1, 1, 1, a). Prove this.
5.4.12. Consider positive integers 1 =q
0
q
1
q
n
with the property that q
i
[

n
j=0
q
j
for i = 0, . . . , n. Set k
i
=

n
j=0
q
j

/q
i
for i = 1, . . . , n and let
P
q0,...,qn
= Conv(0, k
1
e
1
, k
2
e
2
, . . . , k
n
e
n
) (1, . . . , 1) R
n
.
This lattice polytope is reexive by Exercise 2.4.6. Prove that the associated toric variety
is the weighted projective space P(q
0
, q
2
, . . . , q
n
).
Chapter 6
Line Bundles on
Toric Varieties
6.0. Background: Sheaves and Line Bundles
Sheaves of O
X
-modules on a variety X were introduced in 4.0. Recall that for
an afne variety V = Spec(R), an R-module M gives a sheaf

M on V such that

M(V
f
) = M
f
for all f = 0 in R. Globalizing this leads to quasicoherent sheaves
on X. These include coherent sheaves, which locally come from nitely generated
modules. In this section we develop the language of sheaf theory and discuss vector
bundles and line bundles.
The Stalk of a Sheaf at a Point. Since sheaves are local in nature, we need a
method for inspecting a sheaf at a point p X. This is provided by the notion of
direct limit over a directed set.
Denition 6.0.1. A partially ordered set (I, _) is a directed set if
for all i, j I, there exists k I such that i _k and j _k.
If R
i
is a family of rings indexed by a directed set (I, _) such that whenever i _ j
there is a homomorphism

ji
: R
i
R
j
satisfying
ii
= 1
R
i
and
k j

ji
=
ki
, then the R
i
form a directed system. Let S
be the submodule of

iI
R
i
generated by the relations r
j

ji
(r
i
), for r
i
R
i
and
i _ j. Then the direct limit is dened as
lim

iI
R
i
=

iI
R
i

/S.
245
246 Chapter 6. Line Bundles on Toric Varieties
For simplicity, we often write the direct limit as lim

R
i
. Note also that references
such as [10] write
i j
instead of
ji
.
For every i I, there is a natural map R
i
lim

R
i
such that whenever i _ j,
the elements r R
i
and
ji
(r) R
j
have the same image in lim

R
i
. More generally,
two elements r
i
R
i
and r
j
R
j
are identied in lim

R
i
if there is a diagram
R
i
ki

R
k
R
j

k j

such that
ki
(r
i
) =
k j
(r
j
).
Example 6.0.2. Given p X, the denition of sheaf shows that the rings O
X
(U),
indexed by neighborhoods U of p, form a directed system under inclusion, so that
the
ji
are the restriction maps
U,U

for p U

U. The direct limit is the local


ring O
X,p
. For a quasicoherent sheaf F, take an afne open subset V = Spec(R)
containing p so that F(V) = M, where M is an R-module. If m
p
= I(p) R is the
corresponding maximal ideal, then O
X,p
is the localization R
mp
, and
lim

pU
F(U) = M
mp
,
where M
mp
is the localization of M at the maximal ideal m
p
.
The term sheaf has agrarian origins: farmers harvesting their wheat tied a rope
around a big bundle, and left it standing to dry. Think of the footprint of the bundle
as an open set, so that increasingly smaller neighborhoods around a point on the
ground pick out smaller and smaller bits of the bundle, narrowing to a single stalk.
Denition 6.0.3. The stalk of a sheaf F at a point p X is F
p
= lim

pU
F(U).
Injective and Surjective. A homomorphism : F G of O
X
-modules was de-
ned in 4.0. We can also dene what it means for to be injective or surjective.
The denition is a bit unexpected, since we need to take into account the fact that
sheaves are built to convey local data.
Denition 6.0.4. A sheaf homomorphism
: F G
is injective if for any point p X and open subset U X containing p, there exists
an open subset V U containing p, with
V
injective. Also, is surjective if for
any point p and open subset U containing p and any g G(U), there is an open
subset V U containing p and f F(V) such that
V
( f ) =
U,V
(g).
6.0. Background: Sheaves and Line Bundles 247
In Exercise 6.0.1 you will prove that for a sheaf homomorphism : F G,
U ker(
U
: F(U) G(U))
denes a sheaf denoted ker(). You will also show that is injective exactly when
the naive idea works, i.e., ker() = 0. On the other hand, surjectivity of a sheaf
homomorphism need not mean that the maps
U
are surjective for all U. Here is
an example.
Example 6.0.5. On P
1
=C, consider the Weil divisor D = 0 C P
1
.
If we write of P
1
= U
0
U
1
with U
0
= Spec(C[t]) and U
1
= Spec(C[t
1
]), then
C(P
1
) =C(t). Since
(P
1
, O
P
1 (D)) =f C(t)

[ div( f ) +D 00,
it follows easily that we have global sections
1, t
1
(P
1
, O
P
1 (D)).
For any f (P
1
, O
P
1 (D)), multiplication by f gives a sheaf homomorphism
O
P
1(D) O
P
1 . Doing this for 1, t
1
(P
1
, O
P
1 (D)) gives
O
P
1(D) O
P
1(D) O
P
1 .
(Direct sums of sheaves will be dened below.) In Exercise 6.0.2 you will check
that this sheaf homomorphism is surjective. However, taking global sections gives
00 = (P
1
, O
P
1 (D)) (P
1
, O
P
1 (D)) (P
1
, O
P
1 ) =C,
which is clearly not surjective.
There is an additional point to make here. Given : F G, the presheaf
U im(
U
: F(U) G(U))
need not be a sheaf. Fortunately, this can be rectied. Given a presheaf F, there is
an associated sheaf F
+
, the sheacation of F, which is dened by
F
+
(U) = f : U

pU
F
p
[ for all p U, f (p) F
p
and there is
p V
p
U and t F(V
p
) with f (x) = t
p
for all x V
p
.
See [131, II.1] for a proof that F
+
is a sheaf with the same stalks as F
p
. Hence
U im(
U
)
has a natural sheaf associated to it, denoted im().
248 Chapter 6. Line Bundles on Toric Varieties
Exactness. We dene exact sequences of sheaves as follows.
Denition 6.0.6. A sequence of sheaves
F
i1
d
i1
F
i
d
i
F
i+1
is exact at F
i
if there is an equality of sheaves
ker(d
i
) = im(d
i1
).
The local nature of sheaves is again highlighted by the following result, whose
proof may be found in [131, II.1].
Proposition 6.0.7. The sequence in Denition 6.0.6 is exact if and only if
F
i1
p
d
i1
p
F
i
p
d
i
p
F
i+1
p
is exact for all p X.
It follows from Example 6.0.5 that if
(6.0.1) 0 F
1
d
1
F
2
d
2
F
3
0
is a short exact sequence of sheaves, the corresponding sequence of global sections
may fail to be exact. However, we always have the following partial exactness,
which you will prove in Exercise 6.0.3.
Proposition 6.0.8. Given a short exact sequence of sheaves (6.0.1), taking global
sections gives the exact sequence
0 (X, F
1
)
d
1
(X, F
2
)
d
2
(X, F
3
).
In Chapter 9 we will use sheaf cohomology to extend this exact sequence.
Example 6.0.9. For an afne variety V = Spec(R), an R-module M gives a quasi-
coherent sheaf

M on V. This operation preserves exactness, i.e., an exact sequence
of R-modules
0 M
1
M
2
M
3
0
gives an exact sequence of sheaves
0

M
1


M
2


M
3
0
(see [131, Prop. II.5.2]).
Here is a toric generalization of this example.
Example 6.0.10. Let S = C[x

[ (1)] be the total coordinate ring of a toric


variety X

without torus factors. We saw in 5.3 that a graded S-module M gives


the quasicoherent sheaf

M on X.
6.0. Background: Sheaves and Line Bundles 249
Then an exact sequence 0 M
1
M
2
M
3
0 of graded S-modules gives
an exact sequence
0

M
1


M
2


M
3
0
on X

. To see why, note that for , the restriction of



M
i
to U

is the
sheaf associated to ((M
i
)
x
)
0
, the elements of degree 0 in the localization of M
i
at
x

S. Localization preserves exactness, as does taking elements of degree 0. The
desired exactness then follows from Example 6.0.9.
Another example is the following exact sequence of sheaves from 3.0.
Example 6.0.11. A closed subvariety i : Y X gives two sheaves:
The sheaf I
Y
, dened by I
Y
(U) =f O
X
(U) [ f (p) = 0 for p Y U.
The direct image sheaf i

O
Y
, dened by i

O
Y
(U) =O
Y
(Y U).
These are coherent sheaves on X and are related by the exact sequence
0 I
Y
O
X
i

O
Y
0.
Operations on Quasicoherent Sheaves of O
X
. Operations on modules over a ring
have natural analogs for quasicoherent sheaves. In particular, given quasicoherent
sheaves F, G , it is easy to show that U F(U)G (U) denes the quasicoherent
sheaf F G. We can also dene Hom
O
X
(F, G) via
U Hom
O
X
(U)
(F(U), G (U)).
In Exercise 6.0.4 you will show that Hom
O
X
(F, G) is a quasicoherent sheaf.
On the other hand, U F(U)
O
X
(U)
G(U) is only a presheaf, so the tensor
product F
O
X
G is dened to be the sheaf associated to this presheaf. This sheaf
is again quasicoherent and satises
(U, F
O
X
G) =F(U)
O
X
(U)
G(U)
whenever U X is an afne open set (see [131, Prop. II.5.2]).
Global Generation. For a module M over a ring, there is always a surjection from
a free module onto M. This is true for a sheaf F of O
X
-modules when (X, F) is,
in a certain sense, large enough.
Denition 6.0.12. A sheaf F of O
X
-modules is generated by global sections if
there exists a set s
i
(X, F) such that at any point p X, the images of the s
i
generate the stalk F
p
.
Any global section s (X, F) gives a sheaf homomorphism O
X
F. It
follows that if F is generated by s
i

iI
, there is a surjection of sheaves

iI
O
X
F.
In the next section we will see that when X is toric, there is a particularly nice way
of determining when the sheaves O
X
(D) are generated by global sections.
250 Chapter 6. Line Bundles on Toric Varieties
Locally Free Sheaves and Vector Bundles. We begin with locally free sheaves.
Denition 6.0.13. A sheaf F of O
X
-modules is locally free of rank r if there
exists an open cover U

of X such that for all , F[


U
O
r
U
.
Locally free sheaves are closely related to vector bundles.
Denition 6.0.14. A variety V is a vector bundle of rank r over a variety X if there
is a morphism
: V X
and an open cover U
i
of X such that:
(a) For every i, there is an isomorphism

i
:
1
(U
i
)

U
i
C
r
such that
i
followed by projection onto U
i
is [

1
(U
i
)
.
(b) For every pair i, j, there is g
i j
GL
r
((U
i
U
j
, O
X
)) such that the diagram
U
i
U
j
C
r

1
(U
i
U
j
)

i[

1
(U
i
U
j
)

.
.
.
.
.
.
.
.
.

j [

1
(U
i
U
j
)

U
i
U
j
C
r
1g
i j

commutes.
Data (U
i
,
i
) satisfying properties (a) and (b) is called a trivialization. The
map
i
:
1
(U
i
) U
i
C
r
gives a chart, where
1
(p) C
r
for p U
i
. We call

1
(p) the ber over p. See Figure 1 on the next page.
For p U
i
U
j
, the isomorphisms
C
r
pC
r

1
(p)

pC
r
C
r
given by
i
and
j
are related by the linear map g
i j
(p). Hence the ber
1
(p)
has a well-dened vector space structure. This shows that a vector bundle really is
a bundle of vector spaces.
On a vector bundle, the g
i j
are called transition functions and can be regarded
as a family of transition matrices that vary as p U
i
U
j
varies. Just as there
is no preferred basis for a vector space, there is no canonical choice of basis for
a particular ber. Note also that the transition functions satisfy the compatibility
conditions
(6.0.2)
g
ik
= g
i j
g
jk
on U
i
U
j
U
k
g
i j
= g
1
ji
on U
i
U
j
.
6.0. Background: Sheaves and Line Bundles 251
g
ij
(p)
p

U
i
U
j
X
U
i
C
r
U
j
C
r

i
:
1
(p)

{p}C
r

j
:
1
(p)

{p}C
r
Figure 1. Visualizing a vector bundle
Denition 6.0.15. A section of a vector bundle V over U X open is a morphism
s : U V
such that ( s)(p) = p for all p U. A section s : X V is a global section.
A section s picks out a point s(p) in each ber
1
(p), as shown in Figure 2.
p
( ) s x
X
( )
( )
p
s p
1
Figure 2. For a section s, s(p)
1
(p)
252 Chapter 6. Line Bundles on Toric Varieties
We can describe a vector bundle and its global sections purely in terms of the
transition functions g
i j
as follows.
Proposition 6.0.16. Let X be a variety with an afne open cover U
i
, and assume
that for every i, j, we have g
i j
GL
r
((U
i
U
j
, O
X
)) satisfying the compatibility
conditions (6.0.2). Then:
(a) There is a vector bundle : V X of rank r, unique up to isomorphism, whose
transition functions are the g
i j
.
(b) A global section s : X V is uniquely determined by a collection of r-tuples
s
i
O
r
X
such that for all i, j,
s
i[
U
i
U
j
= g
i j
s
j [
U
i
U
j
.
Proof. One easily checks that the g
1
i j
satisfy the gluing conditions from 3.0. It
follows that the afne varieties U
i
C
r
glue together to give a variety V. Fur-
thermore, the projection maps U
i
C
r
U
i
glue together to give a morphism
: V X. It follows easily that the open set of V corresponding to U
i
C
r
is

1
(U
i
), which gives an isomorphism
i
:
1
(U
i
) U
i
C
r
. Hence V is a vector
bundle with transition functions g
i j
.
Given a section s : X V,
i
s[
U
i
is a section of U
i
C
r
U
i
. Thus

i
s[
U
i
(p) = (p, s
i
(p)) U
i
C
r
,
where s
i
O
X
(U
i
)
r
. By Denition 6.0.14, the s
i
satisfy the desired compatibility
condition, and since every global section arises this way, we are done.
Let F(U) denote the set of all sections of V over U. One easily sees that F is
a sheaf on X and in fact is a sheaf of O
X
-modules since the bers are vector spaces.
In fact, F is an especially nice sheaf.
Proposition 6.0.17. The sheaf of sections of a vector bundle is locally free.
Proof. For a trivial vector bundle U C
r
U, the proof of Proposition 6.0.16
shows that a section is determined by a morphism U C
r
, i.e., an element of
O
U
(U)
r
. Thus the sheaf associated to a trivial vector bundle over U is O
r
U
.
For a general vector bundle : V X with trivialization (U
i
,
i
), each U
i
gives an isomorphism of vector bundles

1
(U
i
)

1
(U
i
)

U
i
C
r
.

U
i
.
Since isomorphic vector bundles have isomorphic sheaves of sections, it follows
that if F is the sheaf of sections of : V X, then F[
U
i
O
r
U
i
.
6.0. Background: Sheaves and Line Bundles 253
Line Bundles and Cartier Divisors. Since a vector space of dimension one is a
line, a vector bundle of rank 1 is called a line bundle. Despite the new terminology,
line bundles are actually familiar objects when X is normal.
Theorem 6.0.18. The sheaf L =O
X
(D) of a Cartier divisor D on a normal variety
X is the sheaf of sections of a line bundle V
L
X.
Proof. Recall from Chapter 4 that a Cartier divisor is locally principal, so that
X has an afne open cover U
i

iI
with D[
U
i
= div( f
i
)[
U
i
, f
i
C(X)

. Thus
(U
i
, f
i
)
iI
is local data for D. Note also that
div( f
i
)[
U
i
U
j
= div( f
j
)[
U
i
U
j
,
which implies f
i
/f
j
O
X
(U
i
U
j
)

by Proposition 4.0.16.
We use this data to construct a line bundle as follows. Since
GL
1
(O
X
(U
i
U
j
)) = O
X
(U
i
U
j
)

,
the quotients g
i j
= f
i
/f
j
may be regarded as transition functions. These satisfy the
hypotheses of Proposition 6.0.16 and hence give a line bundle : V
L
X.
A global section f (X, O
X
(D)) satises div( f ) +D 0, so that on U
i
,
div( f f
i
)[
U
i
= div( f )[
U
i
+div( f
i
)[
U
i
= (div( f ) +D)[
U
i
0.
This shows that s
i
= f
i
f O
X
(D)(U
i
). Then
g
i j
s
j
= f
i
/f
j
f
j
f = f
i
f = s
i
,
which by part (b) of Proposition 6.0.16 gives a global section of : V
L
X.
Conversely, the proposition shows that a global section of V
L
X gives functions
s
i
O
X
(D)(U
i
) such that g
i j
s
j
=s
i
. It follows that f =s
i
/f
i
C(X) is independent
of i. One easily checks that f (X, O
X
(D)). The same argument works when we
restrict to any open subset of X. It follows that L =O
X
(D) is the sheaf of sections
of : V
L
X.
We will see shortly that this process is reversible, i.e., there is a one-to-one
correspondence between line bundles and sheaves coming from Cartier divisors.
First, we give an important example.
Example 6.0.19. When we regard P
n
as the set of lines through the origin in C
n+1
,
each point p P
n
corresponds to a line
p
C
n+1
. We assemble these lines into
a line bundle as follows. Let x
0
, . . . , x
n
be homogeneous coordinates on P
n
and
y
0
, . . . , y
n
be coordinates on C
n+1
. Dene
V P
n
C
n+1
as the locus where the matrix

x
0
x
n
y
0
y
n

254 Chapter 6. Line Bundles on Toric Varieties


has rank one. Thus V is dened by the vanishing of x
i
y
j
x
j
y
i
. Then dene the
map : V P
n
to be projection on the rst factor of P
n
C
n+1
. To see that V
is a line bundle, consider the open subset C
n
U
i
P
n
where x
i
is invertible. On

1
(U
i
) the equations dening V become
x
j
x
i
y
i
= y
j
, for all j = i.
Thus (x
0
, . . . , x
n
, y
0
, . . . , y
n
) (x
0
, . . . , , x
n
, y
i
) denes an isomorphism

i
:
1
(U
i
)

U
i
C.
In other words, y
i
is a local coordinate for the line C over U
i
. Switching to the
coordinate system over U
j
, we have the local coordinate y
j
, which over U
i
U
j
is
related to y
i
via
x
i
x
j
y
j
= y
i
.
Hence the the transition function from U
i
U
j
C to U
i
U
j
C is given by
g
i j
=
x
i
x
j
O
P
n (U
i
U
j
)

.
This bundle is called the tautological bundle on P
n
. In Example 6.0.21 below, we
will describe the sheaf of sections of this bundle.
Projective spaces are the simplest type of Grassmannian, and just as in this
example, the construction of the Grassmannian shows that it comes equipped with
a tautological vector bundle. In Exercise 6.0.5 you will determine the transition
functions for the Grassmannian G(1, 3).
Invertible Sheaves and the Picard Group. Propositions 6.0.17 and 6.0.18 imply
that the sheaf O
X
(D) of a Cartier divisor is locally free of rank 1. In general, a
locally free sheaf of rank 1 is called an invertible sheaf.
The relation between Cartier divisors, line bundles and invertible sheaves is
described in the following theorem.
Theorem 6.0.20. Let L be an invertible sheaf on a normal variety X. Then:
(a) There is a Cartier divisor D on X such that L O
X
(D).
(b) There is a line bundle V
L
X whose sheaf of sections is isomorphic to L.
Proof. The part (b) of the theorem follows from part (a) and Proposition 6.0.18. It
remains to prove part (a).
Since X is irreducible, any nonempty open U X gives a domain O
X
(U) with
eld of fractions C(U). By Exercise 3.0.4, C(U) = C(X), so that U C(U)
denes a constant sheaf on X, denoted K
X
. This sheaf is relevant since O
X
(D) is
dened as a subsheaf of K
X
.
6.0. Background: Sheaves and Line Bundles 255
First assume that L is a subsheaf of K
X
. Pick an open cover U
i
of X such
that L[
U
i
O
X[
U
i
for every i. Over U
i
, this gives homomorphisms
O
X
(U
i
) L(U
i
) C(X).
Let f
1
i
C(X) be the image of 1 O
X
(U
i
). One can show without difculty that
f
i
/f
j
O
X
(U
i
U
j
)

. Then (U
i
, f
i
) is local data for a Cartier divisor D on X
satisfying L =O
X
(D).
For the general case, observe that on an irreducible variety, every locally con-
stant sheaf is globally constant (Exercise 6.0.6). Now let L be any invertible sheaf
on X. On a small enough open set U, L(U) O
X
(U), so that
L(U)
O
X
(U)
K
X
(U) O
X
(U)
O
X
(U)
K
X
(U) K
X
(U) =C(X).
Thus L
O
X
K
X
is locally constant and hence constant. This easily implies that
L
O
X
K
X
K
X
, and composing this with the inclusion
L L
O
X
K
X
expresses L as a subsheaf of K
X
.
We note without proof that the line bundle corresponding to an invertible sheaf
is unique up to isomorphism. Because of this result, algebraic geometers tend to
use the terms line bundle and invertible sheaf interchangeably, even though strictly
speaking the latter is the sheaf of sections of the former.
We next discuss some properties of invertible sheaves coming from Cartier
divisors. A rst result is that if D and E are Cartier divisors on X, then
(6.0.3) O
X
(D)
O
X
O
X
(E) O
X
(D+E).
This follows because f g f g induces a sheaf homomorphism
O
X
(D)
O
X
O
X
(E) O
X
(D+E)
which is clearly an isomorphism on any open set where O
X
(D) is trivial.
By standard properties of tensor product, the isomorphism (6.0.3) induces an
isomorphism
O
X
(E) Hom
O
X
(O
X
(D), O
X
(D+E)).
In particular, when E =D, we obtain
O
X
(D)
O
X
O
X
(D) O
X
and O
X
(D) O
X
(D)

,
where O
X
(D)

=Hom
O
X
(O
X
(D), O
X
) is the dual of O
X
(D).
More generally, the tensor product of invertible sheaves is again invertible, and
if L is invertible, then L

=Hom
O
X
(L, O
X
) is invertible and
L
O
X
L

O
X
.
This explains why locally free sheaves of rank 1 are called invertible.
256 Chapter 6. Line Bundles on Toric Varieties
Example 6.0.21. There is a nice relation between the tautological bundle on P
n
and the invertible sheaf O
P
n (1) introduced in Example 4.3.1. Recall that the T
N
-
invariant divisors D
0
, . . . , D
n
on P
n
are all linearly equivalent, and so dene iso-
morphic sheaves, usually denoted O
P
n (1). The local data for the Cartier divisor D
0
is easily seen to be (U
i
,
x
0
x
i
), where U
i
P
n
is the open set where x
i
= 0. Thus
the transition functions for O
X
(D
0
) are given by
g
i j
=
x
0
x
i
x
0
x
j
=
x
j
x
i
.
These are the inverses of the transition functions for the tautological bundle from
Example 6.0.19. It follows that the sheaf of sections of the tautological bundle is
O
P
n(1)

=O
P
n (1).
We can also explain when Cartier divisors give isomorphic invertible sheaves.
Proposition 6.0.22. Two Cartier divisors D, E give isomorphic invertible sheaves
O
X
(D) O
X
(E) if and only if D E.
Proof. By Proposition 4.0.29, linearly equivalent Cartier divisors give isomorphic
sheaves. For the converse, we rst prove that O
X
(D) =O
X
implies D = 0.
Assume O
X
(D) =O
X
. Then 1 (X, O
X
) =(X, O
X
(D)), so D0. If D=0,
then we can pick an irreducible divisor D
0
that appears in D with positive coef-
cient. The local ring O
X,D
0
is a DVR, so we can nd h O
X,D
0
with
D
0
(h) = 1.
Set U = X `W, where W is the union of all irreducible divisors D

= Z
0
with

(h) = 0. There are only nitely many such divisors, so that U is a nonempty
open subset of X with U D
0
=. Then h (U, O
X
), and h
1
/ (U, O
X
) since
h vanishes on U D
0
. However,
(D+div(h
1
))[
U
= (Ddiv(h))[
U
= (DD
0
)[
U
0,
so that h
1
(U, O
X
(D)) = (U, O
X
). This contradiction proves D = 0.
Now suppose that Cartier divisors D, E satisfy O
X
(D) O
X
(E). Tensoring
each side with O
X
(E) and applying (6.0.3), we see that O
X
(DE) O
X
. If
1 (X, O
X
) maps to g (X, O
X
(DE)) via this isomorphism, then
gO
X
=O
X
(DE)
as subsheaves of K
X
. Thus
O
X
= g
1
O
X
(DE) =O(DE +div(g)),
where the last equality follows from the proof of Proposition 4.0.29. By the previ-
ous paragraph, we have DE +div(g) = 0, which implies that D E.
In Chapter 4, the Picard group was dened as the quotient
Pic(X) = CDiv(X)/Div
0
(X).
6.0. Background: Sheaves and Line Bundles 257
We can interpret this in terms of invertible sheaves as follows. Given L invertible,
Theorem 6.0.20 tells us that L O
X
(D) for some Cartier divisor D, which is
unique up to linear equivalence by Proposition 6.0.22. Hence we have a bijection
Pic(X) isomorphism classes of invertible sheaves on X.
The right-hand side has a group structure coming from tensor product of invertible
sheaves. By (6.0.3), the above bijection is a group isomorphism.
In more sophisticated treatments of algebraic geometry, the Picard group of an
arbitrary variety is dened using invertible sheaves. Also, Cartier divisors can be
dened on an irreducible variety in terms of local data, without assuming normality
(see [131, II.6]), though one loses the connection with Weil divisors. Since most
of our applications involve toric varieties coming from fans, we will continue to
assume normality when discussing Cartier divisors.
Stalks, Fibers, and Sections. From here on, we will think of a line bundle L on
X as the sheaf of sections of a rank 1 vector bundle : V
L
X. Given a section
s L(U) and p U, we get the following:
Since V
L
is a vector bundle of rank 1, we have the ber
1
(p) C. Then
s : U V
L
gives s(p)
1
(p).
Since L is a locally free sheaf of rank 1, we have the stalk L
p
O
X,p
. Then
s L(U) gives s
p
L
p
.
In Exercise 6.0.7 you will show that these are related via the equivalences
(6.0.4)
s(p) = 0 in
1
(p) s
p
/ m
p
L
p
s
p
generates L
p
as an O
X,p
-module
A section s vanishes at p X if s(p) = 0 in
1
(p), i.e., if s
p
m
p
L
p
.
Basepoints. It can happen that many sections of a line bundle vanish at a point p.
This leads to the following denition.
Denition 6.0.23. A subspace W (X, L) has no basepoints or is basepoint
free if for every p X, there is s W with s(p) = 0.
As noted earlier, a global section s (X, L) gives a sheaf homomorphism
O
X
L. Thus a subspace W (X, L) gives
W
C
O
X
L
dened by s h hs. Then (6.0.4) and Proposition 6.0.7 imply the following.
Proposition 6.0.24. A subspace W (X, L) has no basepoints if and only if
W
C
O
X
L is surjective.
258 Chapter 6. Line Bundles on Toric Varieties
For a line bundle L = O
X
(D) of a Cartier divisor D on a normal variety, the
vanishing locus of a global section has an especially nice interpretation. The local
data (U
i
, f
i
) of D gives the rank 1 vector bundle : V
L
X with transition
functions g
i j
= f
i
/f
j
. Hence we can think of a nonzero global section of O
X
(D) in
two ways:
A rational function f C(X)

satisfying D+div( f ) 0.
A morphism s : X V
L
whose composition with is the identity on X.
The relation between s and f is given in the proof of Theorem 6.0.18: over U
i
, the
section s looks like (p, s
i
(p)) for s
i
= f
i
f O
X
(U
i
). It follows that s = 0 exactly
when s
i
= 0. Since D[
U
i
= div( f
i
)[
U
i
, the divisor of s
i
on U
i
is given by
div( f
i
f )[
U
i
= (D+div( f ))[
U
i
.
These patch together in the obvious way, so that the divisor of zeros of s is
div
0
(s) = D+div( f ).
Thus the divisor of zeros of a global section is an effective divisor that is linearly
equivalent to D. It is also easy to see that any effective divisor linearly equivalent
to D is the divisor of zeros of a global section of O
X
(D) (Exercise 6.0.8).
In terms of Cartier divisors, Proposition 6.0.24 has the following corollary.
Corollary 6.0.25. The following are equivalent for a Cartier divisor D:
(a) O
X
(D) is generated by global sections in the sense of Denition 6.0.12.
(b) D is basepoint free, meaning that (X, O
X
(D)) is basepoint free.
(c) For every p X there is s (X, O
X
(D)) with p / Supp(div
0
(s)).
The Pullback of a Line Bundle. Let L be a line bundle on X and V
L
X the
associated rank 1 vector bundle. A morphism f : Z X gives the bered product
f

V
L
=V
L

X
Z from 3.0 that ts into the commutative diagram
f

V
L

V
L

Z
f

X.
It is easy to see that f

V
L
is a rank 1 vector bundle over Z.
Denition 6.0.26. The pullback f

L of the sheaf L is the sheaf of sections of


the rank 1 vector bundle f

V
L
dened above.
Thus the pullback of a line bundle is again a line bundle. Furthermore, there is
a natural map on global sections
f

: (X, L) (Z, f

L)
6.0. Background: Sheaves and Line Bundles 259
dened as follows. A global section s : X V
L
gives the commutative diagram:
Z
f

1
Z

(s)

X
s

V
L

V
L

Z
f

X.
The universal property of bered products guarantees the existence and uniqueness
of the dotted arrow f

(s) : Z f

V
L
that makes the diagram commute. It follows
that f

(s) (Z, f

L).
Example 6.0.27. Let X P
n
be a projective variety. If we write the inclusion as
i : X P
n
, then the line bundle O
P
n (1) gives the line bundle i

O
P
n (1) on X. When
the projective embedding of X is xed, this line bundle is often denoted O
X
(1).
Thus a projective variety always comes equipped with a line bundle. However,
it is not unique, since the same variety may have many projective embeddings. You
will work out an example of this in Exercise 6.0.9.
In general, given a sheaf F of O
X
-modules on X and a morphism f : Z X,
one gets a sheaf f

F of O
Z
-modules on Z. The denition is more complicated, so
we refer the reader to [131, II.5] for the details.
Line Bundles and Maps to Projective Space. We now reverse Example 6.0.27 by
using a line bundle L on X to create a map to projective space.
Fix a nite-dimensional subspace W (X, L) with no basepoints and let
W

= Hom
C
(W, C) be its dual. The projective space of W

is
P(W

) = (W

`0)/C

.
We dene a map
L,W
: X P(W

) as follows. Fix p X and pick a nonzero


element v
p

1
(p) C, where : V
L
X is the rank 1 vector bundle associated
to L. For each s W, there is
s
C such that s(p) =
s
v
p
. Then the map dened
by
p
(s) =
s
is linear and nonzero since W has no basepoints. Thus
p
W

, and
since v
p
is unique up to an element of C

, the same is true for


p
. Then

L,W
(p) =
p
denes the desired map
L,W
: X P(W

).
Lemma 6.0.28. The map
L,W
: X P(W

) is a morphism.
Proof. Let s
0
, . . . , s
m
be a basis of W and let U
i
=p X [ s
i
(p) = 0. These open
sets cover X since W has no basepoints. Furthermore, the natural map
U
i
C
1
(U
i
), (p, ) s
i
(p)
260 Chapter 6. Line Bundles on Toric Varieties
is easily seen to be an isomorphism. Since all sections of U
i
C C are of the
form p (p, h(p)) for h O
X
(U
i
), it follows that for all 0 j m, we can write
s
j[
U
i
= h
i j
s
i[
U
i
, h
i j
O
X
(U
i
).
The denition of
L,W
uses a nonzero vector v
p

1
(p). Over U
i
, we can
use s
i
(p)
1
(p). Then s
j
(p) = h
i j
(p)s
i
(p) implies
p
(s
j
(p)) = h
i j
(p). Since
((s
0
), . . . , (s
m
)) gives an isomorphism P(W

) P
m
,
L,W[
U
i
can be written
(6.0.5) U
i
P
m
, p (h
i0
(p), . . . , h
im
(p)),
which is a morphism since h
ii
= 1.
When W has no basepoints and s
0
, . . . , s
m
span W,
L,W
is often written
(6.0.6) X P
m
, p (s
0
(p), . . . , s
m
(p)) P
m
with the understanding that this means (6.0.5) on U
i
=p X [ s
i
(p) = 0.
Furthermore, when L = O
X
(D), we can think of the global sections s
i
as
rational functions g
i
such that D+div(g
i
) 0. Then
L,W
can be written
(6.0.7) X P
m
, p (g
0
(p), . . . , g
m
(p)) P
m
.
Since g
i
(p) may be undened, this needs explanation. The local data (U
j
, f
j
) of
D implies that f
j
g
0
, . . . , f
j
g
m
O
X
(U
j
). Then (6.0.7) means that
L,W[
U
j
is
U
j
P
m
, p ( f
j
g
0
(p), . . . , f
j
g
m
(p)) P
m
.
This is a morphism on U
j
since the global sections corresponding to g
0
, . . . , g
m
have
no base points.
Exercises for 6.0.
6.0.1. For a sheaf homomorphism : F G, show that
U ker(
U
)
denes a sheaf. Also prove that the following are equivalent:
(a) The kernel sheaf is identically zero.
(b)
U
is injective for every open subset U.
(c) is injective as dened in Denition 6.0.4.
6.0.2. In Example 6.0.5, prove that O
P
1 (D) O
P
1 (D) O
P
1 is surjective.
6.0.3. Prove Proposition 6.0.8.
6.0.4. Let F, G be quasicoherent sheaves on X. Prove that U Hom
OX (U)
(F(U), G (U))
denes a quasicoherent sheaf Hom
OX
(F, G ).
6.0.5. The Grassmannian G(1, 3) is dened as the space of lines in P
3
, or equivalently, of
2-dimensional subspaces of V = C
4
. This exercise will construct the tautological bundle
on G(1, 3), which assembles these 2-dimensional subspaces into a rank 2 vector bundle
over G(1, 3). A point of G(1, 3) corresponds to a full rank matrix
p =

0

1

2

3

0

1

2

3

6.0. Background: Sheaves and Line Bundles 261


up to left multiplication by elements of GL
2
(C). Then dene
V G(1, 3) C
4
to consist of all pairs ((

), v) such that v Span(, ).


(a) A pair ((

), v) gives the 3 4 matrix


A =

v
0
v
1
v
2
v
3

0

1

2

3

0

1

2

3

.
Prove that ((

), v) is a point of V if and only if the maximal minors of A vanish. This


shows that V G(1, 3) C
4
is a closed subvariety.
(b) Projection onto the rst factor gives a morphism : V G(1, 3). Explain why the
ber over p G(1, 3) is the 2-dimensional subspace of C
4
corresponding to p.
(c) Given 0 i < j 3, dene
U
i j
=(

) G(1, 3) [
i

i
= 0.
Prove that U
i j
C
4
and that the U
i j
give an afne open cover of G(1, 3).
(d) Given 0 i < j 3, pick k < l such that i, j, k, l = 0, 1, 2, 3. Prove that the map
(p, v) (p, v
k
, v
l
) gives an isomorphism

1
(U
i j
)

U
i j
C
2
.
(e) By part (d), V is a vector bundle over G(1, 3). Determine its transition functions.
6.0.6. Prove that a locally constant sheaf on an irreducible variety is constant.
6.0.7. Prove (6.0.4).
6.0.8. Prove that an effective divisor linearly equivalent to a Cartier divisor D is the divisor
of zeros of a global section of O
X
(D).
6.0.9. Let
d
: P
1
P
d
be the Veronese mapping dened in Example 2.3.15. Prove that

d
O
P
d (1) =O
P
1 (d).
6.0.10. Let f : Z X be a morphism and let L be a line bundle on X that is generated by
global sections. Prove that the pullback line bundle f

L is generated by global sections.


6.0.11. Let D be a Cartier divisor on a complete normal variety X.
(a) f , g (X, O
X
(D)) ` 0 give effective divisors D+div( f ), D+div(g) on X. Prove
that these divisors are equal if and only if f = g, C

.
(b) The complete linear system of D is dened to be
[D[ =E CDiv(X) [ E D, E 0.
Thus the complete linear system of D consists of all effective Cartier divisors on X
linearly equivalent to D. Use part (a) to show that [D[ can be identied with the
projective space of (X, O
X
(D)), i.e., there is a natural bijection
[D[ =P((X, O
X
(D))) = ((X, O
X
(D)) ` 0)/C

.
(c) Assume that D has no basepoints and set W = (X, O
X
(D)). Then we can identify
P(W

) with the set of hyperplanes in P(W) =[D[. Prove that the morphism
OX (D),W
:
X P(W

) is given by

OX (D),W
=E [D[ [ p Supp(E) [D[.
262 Chapter 6. Line Bundles on Toric Varieties
6.1. Ample and Basepoint Free Divisors on Complete Toric Varieties
In this section we will study two special classes of Cartier divisors on complete
toric varieties. We begin with the basepoint free case.
Basepoint Free Divisors. Consider the toric variety X

of a fan in N
R
R
n
and
let D =

be a torus-invariant Cartier divisor on X

. By Propositions 4.3.3
and 4.3.8, we have the global sections
(X

, O
X

(D)) =

mP
D
M
C
m
,
where P
D
M
R
is the polyhedron dened by
(6.1.1) P
D
=m M
R
[ 'm, u

` a

for all (1).


Since D =

is Cartier, there are m

M for such that


(6.1.2) 'm

, u

` =a

, (1).
Furthermore, when
max
= (n), D is uniquely determined by the Cartier data
m

(n)
. Then we have the following preliminary result.
Proposition 6.1.1. If
max
= (n), then the following are equivalent:
(a) D has no basepoints, i.e., O
X

(D) is generated by global sections.


(b) m

P
D
for all (n).
Proof. First suppose that D is generated by global sections and take (n).
The T
N
-orbit corresponding to is a xed point p of the T
N
-action, and by the
Orbit-Cone Correspondence,
p =

(1)
D

.
By Corollary 6.0.25, there is a global section s such that p is not in the support
of the divisor of zeros div
0
(s) of s. Since (X

, O
X

(D)) is spanned by
m
for
m P
D
M, we can assume that s is given by
m
for some m P
D
M. The
discussion preceding Corollary 6.0.25 shows that the divisor of zeros of s is
div
0
(s) = D+div(
m
) =

(a

+'m, u

`)D

.
The point p is not in the support of div
0
(s) yet lies in D

for every (1). This


forces a

+'m, u

` = 0 for (1). Since is n-dimensional, we conclude that


m

= m P
D
.
For the converse, take (n). Since m

P
D
, the character
m
gives a
global section s whose divisor of zeros is div
0
(s) = D+div(
m
). Using (6.1.2),
one sees that the support of div
0
(s) misses U

, so that s is nonvanishing on U

.
Then we are done since the U

cover X

.
6.1. Ample and Basepoint Free Divisors on Complete Toric Varieties 263
Here is an example to illustrate Proposition 6.1.1.
Example 6.1.2. The fan for the Hirzebruch surface H
2
is shown in Figure 3. Let
u
2
u
4
u
3
u
1
= (1,2)

4
Figure 3. A fan 2 with X
2
=H2
D
i
be the divisor corresponding to u
i
. We will study the divisors
D = D
4
and D

= D
2
+D
4
.
Write the Cartier data for D and D

with respect to
1
, . . . ,
4
as m
i
and m

respectively. Figure 4 shows P


D
and m
i
(left) and P
D
and m

i
(right) (see also
P
D
m
2
m
1
= m
4
m
3
2
1
P
D
m
2
m
3
m
1
m
4
2
1
Figure 4. PD and mi (left) and P
D
and m

i
(right)
Exercise 4.3.5). This gure and Proposition 6.1.1 make it clear that D is basepoint
free while D

is not.
Support Functions and Their Graphs. Let D=

be a Cartier divisor on a
toric variety X

. As in Chapter 4, its support function


D
: [[ R is determined
by the following properties:

D
is linear on each cone .

D
(u

) =a

for all (1).


264 Chapter 6. Line Bundles on Toric Varieties
This is where the m

from (6.1.2) appear naturally, since the explicit formula


for
D[

is given by
D
(u) ='m

, u` for all u .
When M = Z
2
and is complete, it is easy to visualize the graph of
D
in
M
R
R = R
3
: imagine a tent, with centerpole extending from (0, 0, 0) down the
z-axis, and tent stakes placed at positions (u

, a

). Here is an example.
Example 6.1.3. Take P
1
P
1
and consider the divisor D = D
1
+D
2
+D
3
+D
4
.
This gives the support function where
D
(u
i
) = 1 for the four ray generators
u
1
, u
2
, u
3
, u
4
of the fan of P
1
P
1
. The graph of
D
is shown in Figure 5. This
u
1
u
3
u
4
u
2
Figure 5. The graph of D
should be visualized as an innite Egyptian pyramid, with apex at the origin and
edges going through (u
i
, 1) for 1 i 4.
Convex Functions. We now introduce the key concept of convexity.
Denition 6.1.4. Let S N
R
be convex. A function : S R is convex if
(tu+(1t)v) t(u) +(1t)(v),
for all u, v S and t [0, 1].
We caution the reader that some books dene convexity with the inequality
going the other way.
Continuing with the tent analogy, a support function
D
is convex exactly if
there are unimpeded lines of sight inside the tent. It is clear that for Example 6.1.3,
the support function is convex.
6.1. Ample and Basepoint Free Divisors on Complete Toric Varieties 265
Full Dimensional Convex Support. In this chapter, our main focus is on complete
fans. However, the natural setting for convexity is the class of fans in N
R
which
satisfy the following two conditions:
[[ N
R
is convex.
dim[[ = n = dim N
R
.
We say that has convex support of full dimension. Such fans satisfy
(6.1.3) [[ = Cone(u

[ (1)) =

(n)
.
In particular, the maximal cones of have dimension n, so that we can focus on
the cones (n), just as in the complete case.
Support Functions and Convexity. The following lemma describes when a sup-
port function is convex. Given a fan in N
R
R
n
, a cone (n1) is called a
wall when it is the intersection of two n-dimensional cones ,

(n), i.e, when


=

forms the wall separating and

. If is complete, every (n1)


is a wall.
Lemma 6.1.5. Let D be a Cartier divisor on a toric variety whose fan has convex
support of full dimension. Then the following are equivalent:
(a) The support function
D
: [[ R is convex.
(b)
D
(u) 'm

, u` for all u [[ and (n).


(c)
D
(u) = min
(n)
'm

, u` for all u [[.


(d) For every wall =

, there is u
0

` with
D
(u
0
) 'm

, u
0
`.
Proof. First assume (a) and x v in the interior of (n). Given u [[, we can
nd t (0, 1) with tu+(1t)v . By convexity, we have
'm

, tu+(1t)v` =
D
(tu+(1t)v)
t
D
(u) +(1t)
D
(v) = t
D
(u) +(1t)'m

, v`.
This easily implies 'm

, u`
D
(u), proving (b). The implication (b) (c) is
immediate since
D
(u) = 'm

, u` for u , and (c) (a) follows because the


minimum of a nite set of linear functions is always convex (Exercise 6.1.1).
Since (b) (d) is obvious, it remains to prove the converse. Assume (d) and
x a wall =

. Then

lies on one side of the wall. We claim that


(6.1.4) 'm

, u` 'm

, u`, when u,

are on the same side of .


This is easy. The wall is dened by 'm

, u` = 0. Then (d) implies that the


halfspace containing

is dened by 'm

, u` 0, and (6.1.4) follows.


Now take u [[ and (n). Since [[ is convex, we can pick v in the
interior of so that the line segment uv intersects every wall of in a single point,
as shown in Figure 6 on the next page. Using (6.1.4) repeatedly, we obtain
266 Chapter 6. Line Bundles on Toric Varieties
s s s s
u
. .. .

. .. .

wall

wall
. .. .

wall
Figure 6. Crossing walls from u to v along uv
'm

, u` 'm

, u` 'm

, u` .
When we arrive at the cone containing u, the pairing becomes
D
(u), so that
'm

, u`
D
(u). This proves (b).
In terms of the tent analogy, part (b) of the lemma means that if we have a
convex support function and extend one side of the tent in all directions, the rest of
the tent lies below the resulting hyperplane. Then part (d) means that it sufces to
check this locally where two sides of the tent meet.
The proof of our main result about convexity will use the following lemma that
describes the polyhedron of a Cartier divisor in terms of its support function.
Lemma 6.1.6. Let be a fan and D =

be a Cartier divisor on X

. Then
P
D
=m M
R
[
D
(u) 'm, u` for all u [[.
Proof. Assume
D
(u) 'm, u` for all u [[. Applying this with u = u

gives
a

=
D
(u

) 'm, u

`,
so that m P
D
by the denition of P
D
. For the opposite inclusion, take m P
D
and
u [[. Thus u , so that u =

(1)

0. Then
'm, u` =

(1)

'm, u

(1)

(a

)
=

(1)

D
(u

) =
D
(u),
where the inequality follows from m P
D
, and the last two equalities follow from
the dening properties of
D
.
We now expand Proposition 6.1.1 to give a more complete characterization of
when a divisor is basepoint free. Recall that P
D
is a polytope when is complete.
Theorem 6.1.7. Assume [[ is convex of full dimension n and let
D
be the support
function of a Cartier divisor D on X

. Then the following are equivalent:


(a) D is basepoint free.
(b) m

P
D
for all (n).
(c)
D
(u) = min
(n)
'm

, u` for all u [[.


(d)
D
: [[ R is convex.
If addition is complete, then (a)(d) are equivalent to the following:
6.1. Ample and Basepoint Free Divisors on Complete Toric Varieties 267
(e) P
D
= Conv(m

[ (n)).
(f) m

[ (n) is the set of vertices of P


D
.
(g)
D
(u) = min
mP
D
'm, u` for all u N
R
.
Proof. The equivalences (a) (b) and (c) (d) were proved in Proposition 6.1.1
and Lemma 6.1.5. Furthermore, Lemmas 6.1.5 and 6.1.6 imply that

D
is convex
D
(u) 'm

, u` for all (n), u [[


m

P
D
for all (n).
This proves (d) (b), so that (a), (b), (c) and (d) are equivalent.
Assume (b) and note that P
D
is a polytope since is complete. Then m

P
D
and
D
(u) = min
(n)
'm

, u`. Combining these with Lemma 6.1.6, we obtain

D
(u) min
mP
D
'm, u` min
(n)
'm

, u` =
D
(u),
proving (g). The implication (g) (d) follows since the minimum of a compact
set of linear functions is convex (Exercise 6.1.1). So (a) (b) (c) (d) (g).
The implications (f) (e) (b) are clear. It remains to prove (b) (f). Take
(n). Let u be in the interior of and set a =
D
(u). By Exercise 6.1.2,
H
u,a
=m M
R
[ 'm, u` = a is a supporting hyperplane of P
D
and
(6.1.5) H
u,a
P
D
=m

.
This implies that m

is a vertex of P
D
. Conversely, let H
u,a
be a supporting hyper-
plane of a vertex v P
D
. Thus 'm, u` a for all m P
D
, with equality if and only if
m=v. Since (b) holds, we also have (c) and (g). By (g),
D
(u) =min
mP
D
'm, u` =
'm, v` = a. Combining this with (c), we obtain

D
(u) = min
(n)
'm

, u` = a.
Hence 'm

, u` = a must occur for some (n), which forces v = m

.
Example 6.1.8. In Example 6.1.2 we showed that on the Hirzebruch surface H
2
,
D=D
4
is basepoint free while D

=D
2
+D
4
is not. Theorem 6.1.7 gives a different
proof using support functions. Figure 7 on the next page shows the graph of the
support function
D
. Notice that the portion of the roof containing the points
u
1
, u
2
, u
3
and the origin lies in the plane z = 0, and it is clear that for
D
, there are
unimpeded lines of sight within the tent. In other words,
D
is convex.
The support function
D
is shown in Figure 8 on the next page. Here, the line
of sight from u
1
to u
3
lies in the plane z = 0, yet the ridgeline going from the origin
to the point (u
2
, 1) on the tent lies below the plane z = 0. Hence this line of sight
does not lie inside the tent, so that
D
is not convex.
268 Chapter 6. Line Bundles on Toric Varieties
u
2
u
4
u
1
u
3
Figure 7. The graph of D =D
4
in Example 6.1.8
u
2
u
4
u
1
u
3
not inside tent

Figure 8. The graph of


D
=D
2
+D
4
in Example 6.1.8
When D is basepoint free, Theorem 6.1.7 implies that the vertices of P
D
are
the lattice points m

, (n). One caution is that in general, the m

need not
be distinct, i.e., =

can have m

= m

. An example is given by the divisor


D = D
4
considered in Example 6.1.2see Figure 4. As we will see later, this
behavior illustrates the difference between basepoint free and ample.
It can also happen that P
D
has strictly smaller dimension than the dimension of
X

. You will work out a simple example of this in Exercise 6.1.3.


Ample Divisors. We now introduce the second key concept of this section.
Denition 6.1.9. Let D be a Cartier divisor on a complete normal variety X. As
we noted in 4.3, W = (X, O
X
(D)) is nite-dimensional.
(a) The divisor D and the line bundle O
X
(D) are very ample when D has no base-
points and
D
=
O
X
(D),W
: X P(W

) is a closed embedding.
(b) D and O
X
(D) are ample when kD is very ample for some integer k > 0.
6.1. Ample and Basepoint Free Divisors on Complete Toric Varieties 269
We will see that support functions give a simple, elegant characterization of
when a torus-invariant Cartier divisor is ample. But rst, we explore how the very
ample polytopes from Denition 2.2.17 relate to Denition 6.1.9.
Very Ample Polytopes. Let P M
R
R
n
a full dimensional lattice polytope with
facet presentation
P =m M
R
[ 'm, u
F
` a
F
for all facets F.
This gives the complete normal fan
P
and the toric variety X
P
. Write
PM =m
1
, . . . , m
s
.
A vertex m
i
P corresponds to a maximal cone
(6.1.6)
i
= Cone(PMm
i
)


P
(n).
Proposition 4.2.10 implies that D
P
=

F
a
F
D
F
is Cartier since 'm
i
, u
F
` = a
F
when m
i
F.
Recall from Denition 2.2.17 that P is very ample if for every vertex m
i
P,
the semigroup N(PMm
i
) is saturated in M. The denition of X
P
given in
Chapter 2 used very ample polytopes. This is no accident.
Proposition 6.1.10. Let X
P
and D
P
be as above. Then:
(a) D
P
is ample and basepoint free.
(b) If n 2, then kD
P
is very ample for every k n1.
(c) D
P
is very ample if and only if P is a very ample polytope.
Proof. First observe that the polytope of the divisor D
P
is the polytope P we began
with, i.e., P
D
P
= P. This has two consequences:
D
P
is basepoint free by Proposition 6.1.1, which proves the nal assertion of
part (a).
If PM =m
1
, . . . , m
s
, then the characters
m
i
span W = (X
P
, O
X
P
(D
P
)).
Since D
P
is basepoint free, these global sections give the morphism

D
P
=
O
X
P
(D
P
),W
P
s1
by Lemma 6.0.28. As explained in (6.0.7),
D
can be written
(6.1.7)
D
P
(p) = (
m
1
(p), . . . ,
ms
(p)).
It follows that
D
P
factors as
X
P
X
PM
P
s1
,
where X
PM
is the projective toric variety of PM M from 2.1. We need to
understand when X
P
X
PM
is an isomorphism.
270 Chapter 6. Line Bundles on Toric Varieties
Fix coordinates x
1
, . . . , x
s
of P
s1
and let I 1, . . . , s be the set of indices
such that m
i
is a vertex of P. Hence each i I gives a vertex m
i
and a corresponding
maximal cone
i
in the normal fan of P.
If i I, then 'm
i
, u
F
` = a
F
for every facet F containing m
i
. For all other
facets F, 'm
i
, u
F
` > a
F
. Hence, if s
i
is the global section corresponding to
m
i
,
then the support of div(s
i
)
0
= D
P
+div(
m
i
) consists of those divisors missing the
afne open toric variety U

i
X
P
of
i
. It follows that U

i
is the nonvanishing
locus of s
i
.
Under
D
P
, this nonvanishing locus maps to the afne open subset U
i
P
s1
where x
i
= 0. Since X
P
=

iI
U

i
, and X
PM

iI
U
i
by Proposition 2.1.9, it
sufces to study the maps
U

i
X
PM
U
i
of afne toric varieties. By Proposition 2.1.8,
X
PM
U
i
= Spec(C[N(PMm
i
)]).
Since

i
= Cone(PMm
i
) by (6.1.6), we have an inclusion of semigroups
N(PMm
i
)

i
M.
This is an equality precisely when N(PMm
i
) is saturated in M. Since U

i
=
Spec(C[

i
M]), we obtain the equivalences:
D
P
is very ample X
P
X
PM
is an isomorphism
U

i
X
PM
U
i
is an isomorphism for all i I
C[N(PMm
i
)] C[

M] is an
isomorphism for all i I
N(PMm
i
) is saturated for all i I
P is very ample.
This proves part (c) of the proposition. For part (b), recall that if n 2 and P
is arbitrary, then kP is very ample when k n 1 by Corollary 2.2.19. Hence
kD
P
= D
kP
is very ample. This implies that D
P
is ample (the case n = 1 is trivial),
which completes the proof of part (a).
Example 6.1.11. In Example 2.2.11, we showed that the simplex
P = Conv(0, e
1
, e
2
, e
1
+e
2
+3e
3
) R
3
is not normal. We show that P is not very ample as follows. From Chapter 2
we know that the only lattice points of P are its vertices, so that
D
P
: X
P
P
3
.
Since X
P
is singular (Exercise 6.1.4) of dimension 3, it follows that
D
P
cannot be
a closed embedding. Hence P and D
P
are not very ample. However, 2P and 2D
P
are very ample by Proposition 6.1.10.
6.1. Ample and Basepoint Free Divisors on Complete Toric Varieties 271
Ampleness and Strict Convexity. We next determine when a Cartier divisor D =

on X

is ample. Our criterion will involve the support function


D
of D.
Recall that the Cartier data m

(n)
of D satises
'm

, u` =
D
(u), for all u .
Denition 6.1.12. Assume that has full dimensional convex support. Then the
support function
D
of a Cartier divisor D on X

is strictly convex if it is convex


and for every (n) satises
'm

, u` =
D
(u) u .
The following lemma, which you will prove in Exercise 6.1.5, shows that there
are many ways to think about strict convexity.
Lemma 6.1.13. Let D Cartier divisor on a toric variety whose fan has convex
support of full dimension. Then the following are equivalent:
(a) The support function
D
: [[ R is strictly convex.
(b)
D
(u) <'m

, u` for all u [[ ` and (n).


(c) For every wall =

, there is u
0

` with
D
(u
0
) <'m

, u
0
`.
(d)
D
is convex and m

= m

when =

in (n) and

is a wall.
(e)
D
is convex and m

= m

when =

in (n).
(f) 'm

, u

` >a

for all (1) `(1) and (n).


(g)
D
(u+v) >
D
(u) +
D
(v) for all u, v [[ not in the same cone of .
We now relate strict convexity to ampleness.
Theorem 6.1.14. Assume that
D
is the support function of a Cartier divisor D =

on a complete toric variety X

. Then
D is ample
D
is strictly convex.
Furthermore, if n 2 and D is ample, then kD is very ample for all k n1.
Proof. First suppose that D is very ample. Very ample divisors have no basepoints,
so
D
is convex by Theorem 6.1.7. If strict convexity fails, then Lemma 6.1.13
implies that has a wall =

with m

= m

. Let V() = O() X

.
Let P
D
be the polyhedron of D from (6.1.1), which is a polytope since is
complete. Let P
D
M =m
1
, . . . , m
s
, so that
D
: X

P
s1
can be written

D
(p) = (
m
1
(p), . . . ,
ms
(p))
as in (6.1.7). In this enumeration, m

= m

= m
i
0
for some i
0
. We will study
D
on the open subset U

.
272 Chapter 6. Line Bundles on Toric Varieties
First consider U

. Theorem 6.1.7 implies that m

P
D
, so that the section
corresponding to
m
is nonvanishing on U

by the proof of Proposition 6.1.1. It


follows that on U

,
D
is given by

D
(p) = (
m
1
m
(p), . . . ,
msm
(p)) U
i
0
C
s1
,
where U
i
0
P
s1
is the open subset where x
i
0
= 0.
Since m

= m

, the same argument works on U

. This gives a morphism

D[
UU

: U

U
i
0
C
s1
.
The only n-dimensional cones of containing are ,

since is a wall. Hence


V() U

by the Orbit-Cone Correspondence. Note also V() P


1
since is a wall. Since
P
1
is complete, Proposition 4.3.8 implies that all morphisms fromP
1
to afne space
are constant. Thus
D
maps V() to a point, which is impossible since D is very
ample. Hence
D
is strictly convex when D is very ample.
If D is ample, then kD is very ample for k 0. Thus
kD
= k
D
must be
strictly convex, which implies that
D
is strictly convex.
For the converse, assume
D
is strictly convex. Let m

(n)
be the Cartier
data of D. Since is convex, Theorem 6.1.7 shows that the m

are the vertices of


P
D
. Hence P
D
is a lattice polytope.
If P
D
is not full dimensional, then there are u = 0 in N
R
and k R such that
'm

, u` = k for all (n). Then Theorem 6.1.7 implies

D
(u) ='m

, u` = k
for all (n). Using strict convexity and Denition 6.1.12, we conclude that
u for all (n). Hence u = 0 since is complete. This contradicts u = 0
and proves that P
D
is full dimensional.
Hence P
D
gives the toric variety X
P
D
with normal fan
P
D
. Furthermore, X
P
D
has the ample divisor D
P
D
from Proposition 6.1.10. We studied the support function
of this divisor in Proposition 4.2.14, where we showed that it is the function

P
D
(u) = min
mP
D
'm, u`.
However, this is precisely
D
by Theorem 6.1.7. Hence
P
D
=
D
is strictly convex
with respect to (by hypothesis) and
P
D
(by the rst part of the proof).
Denition 6.1.13 implies that the maximal cones of the fan are the maximal
subsets of N
R
on which a strictly convex support function is linear. This, combined
with the previous paragraph, implies that =
P
D
. Thus
(6.1.8) X

= X
P
D
.
Furthermore, we also have
(6.1.9) D = D
P
D
6.1. Ample and Basepoint Free Divisors on Complete Toric Varieties 273
since the divisors have the same support function. Since D
P
D
is an ample divisor
by Proposition 6.1.10, it follows that D is also ample.
The nal assertion of the theorem follows from Proposition 6.1.10.
The relation between polytopes and ample divisors given by (6.1.8) and (6.1.9)
will be explored in 6.2. These facts also give the following nice result.
Theorem 6.1.15. On a smooth complete toric variety X

, a divisor D is ample if
and only if it is very ample.
Proof. If D is ample, then X

is the toric variety of P


D
by (6.1.8). Since X

is
smooth, P
D
is very ample by Theorem 2.4.3 and Proposition 2.4.4. Since D is the
divisor of P
D
by (6.1.9), D is very ample by Proposition 6.1.10.
Computing Ample Divisors. Given a wall (n1), write =

and pick

(1) `(1). Then a Cartier divisor D =

gives the wall inequality


(6.1.10) 'm

, u

` >a

.
Lemma 6.1.13 and Theorem 6.1.14 imply that D is ample if and only if it satises
the wall inequality (6.1.10) for every wall of .
In terms of divisor classes, recall the map CDiv
T
(X

) Pic(X

) whose kernel
consists of divisors of characters. If we x
0
(n), then we have an isomor-
phism
(6.1.11)

D =

CDiv
T
(X

) [ a

= 0 for all
0
(1)

Pic(X

)
(Exercise 6.1.6). Then (6.1.10) gives inequalities for determining when a divisor
class is ample. Here is a classic example.
Example 6.1.16. Let us determine the ample divisors on the Hirzebruch surface
H
r
. The fan for H
2
is shown in Figure 3 of Example 6.1.2, and this becomes
the fan for H
r
by redening u
1
to be u
1
= (1, r). Hence we have ray generators
u
1
, u
2
, u
3
, u
4
and maximal cones
1
,
2
,
3
,
4
.
In Examples 4.3.5 and 4.1.8, we used D
1
and D
2
to give a basis of Pic(H
r
) =
Cl(H
r
). Here, it is more convenient to use D
3
and D
4
. More precisely, applying
(6.1.11) for the cone
4
, we obtain
Pic(H
r
) aD
3
+bD
4
[ a, b Z.
To determine when aD
3
+bD
4
is ample, we compute m
i
= m

i
to be
m
1
= (a, 0), m
2
= (a, b), m
3
= (rb, b), m
4
= (0, 0).
Then (6.1.10) gives four wall inequalities which reduce to a, b > 0. Thus
(6.1.12) aD
3
+bD
4
is ample a, b > 0.
274 Chapter 6. Line Bundles on Toric Varieties
For an arbitrary divisor D =

4
i=1
a
i
D
i
, the relations
0 div(
e
1
) =D
1
+D
3
0 div(
e
2
) = r D
1
+D
2
D
4
show that D (a
1
ra
2
+a
3
)D
3
+(a
2
+a
4
)D
4
. Hence

4
i=1
a
i
D
i
is ample a
1
+a
3
> ra
2
, a
2
+a
4
> 0.
Sometimes ampleness is easier to check if we think geometrically in terms of
support functions. For D = aD
3
+bD
4
, look back at Figure 7 and imagine moving
the vertex at u
3
downwards. This gives the graph of
D
, which is strictly convex
when a, b > 0.
Here is an example of how to determine ampleness using support functions.
Example 6.1.17. The fan for P
1
P
1
P
1
has the eight orthants of R
3
as its maxi-
mal cones, and the ray generators are e
1
, e
2
, e
3
. Take the positive orthant R
3
0
and subdivide further by adding the new ray generators
a = (2, 1, 1), b = (1, 2, 1), c = (1, 1, 2), d = (1, 1, 1).
We obtain a complete fan by lling the rst orthant with the cones in Figure 9,
which shows the intersection of R
3
0
with the plane x +y +z = 1. You will check
that is smooth in Exercise 6.1.7.
e
1
e
2
e
3
a b
c
d
Figure 9. Cones of lying in R
3
0
Let D =

be a Cartier divisor on X

. Replacing D with D+div(


m
)
for m = (a
e
1
, a
e
2
, a
e
3
), we can assume that
D
satises

D
(e
1
) =
D
(e
2
) =
D
(e
3
) = 0.
Now observe that e
1
+b = (2, 2, 1) = e
2
+a. Since e
1
and b do not lie in a cone of
, part (g) of Lemma 6.1.13 implies that

D
(e
1
+b) >
D
(e
1
) +
D
(b) =
D
(b).
6.1. Ample and Basepoint Free Divisors on Complete Toric Varieties 275
However, e
2
and a generate a cone of , so that

D
(a) =
D
(e
2
) +
D
(a) =
D
(e
2
+a) =
D
(e
1
+b).
Together, these imply
D
(a) >
D
(b). By symmetry, we obtain

D
(a) >
D
(b) >
D
(c) >
D
(a),
an impossibility. Hence has no strictly convex support functions, which shows
that X

is a smooth complete nonprojective variety. See also Example B.2.2 for a


computational approach using the Polyhedra package of Macaulay2 [123].
We will say more computing ample divisors later in the chapter.
The Toric Chow Lemma. Recall from Chapter 3 that a renement

of gives a
proper birational toric morphism X

. We will now use the methods of this


section to prove the toric Chow lemma, which asserts that any complete fan has a
renement that gives a projective toric variety. Here is the precise result.
Theorem 6.1.18. A complete fan has a renement

such that X

is projective.
Proof. Suppose is a fan in N
R
R
n
. Let

be obtained from by considering


the complete fan obtained from

(n1)
Span().
So for each wall , we take the entire hyperplane spanned by the wall. This yields
a subdivison

with the property that

(n1)

(n1)
Span(),
i.e., each hyperplane Span() is a union of walls of

, and all walls of

arise
this way.
Choosing m

M so that
u N
R
[ 'm

, u` = 0 = Span(),
dene the map : N
R
R by
(u) =

(n1)
['m

, u`[.
Note that takes integer values on N and is convex by the triangle inequality (this
explains the minus sign).
Let us show that is piecewise linear with respect to

. Fix (n1) and


note that each cone of

is contained in one of the closed half-spaces bounded by


Span(). This implies that u ['m

, u`[ is linear on each cone of

. Hence the
same is true for .
276 Chapter 6. Line Bundles on Toric Varieties
Finally, we prove that is strictly convex. Suppose that

2
is a wall
of

. Then

Span(
0
),
0
(n1). We label

1
and

2
so that
[

1
(u) ='m

0
, u`

,=
0
in (n1)
['m

, u`[, u

1
[

2
(u) = 'm

0
, u`

,=
0
in (n1)
['m

, u`[, u

2
.
The sum

,=
0
in (n1)
['m

, u`[ is linear on

2
, so is represented by dif-
ferent linear functions on each side of the wall

. Since is convex, it is strictly


convex by Lemma 6.1.13. Then X

is projective since D

(u

)D

is
ample by Theorem 6.1.14.
Using the results of Chapter 11, one can improve this result by showing that
X

can be chosen to be smooth and projective.


Exercises for 6.1.
6.1.1. Let S M
R
be a compact set and dene : N
R
R by (u) = min
mS
'm, u`.
Explain carefully why the minimum exists and prove that is convex.
6.1.2. Let H
u,a
be as in the proof of (b) (d) of Theorem 6.1.7. Prove that H
u,a
is a
supporting hyperplane of P
D
that satises (6.1.5). Hint: Write u =

(1)

> 0.
Then show m P
D
implies 'm, u` =

(1)

'm, u

`
D
(u).
6.1.3. As noted in the text, the polytope P
D
of a basepoint free Cartier divisor on a complete
toric variety X

can have dimension strictly less than dim X

. Here are some examples.


(a) Let D be one of the four torus-invariant prime divisors on P
1
P
1
. Show that P
D
is a
line segment.
(b) Consider (P
1
)
n
and x an integer d with 0 <d < n. Find a basepoint free divisor D on
(P
1
)
n
such that dimP
D
= d. Hint: See Exercise 6.1.9 below.
6.1.4. Show that the toric variety X
P
of the polytope P in Example 6.1.11 is singular.
6.1.5. This exercise is devoted to proving that the statements (a)(g) of Lemma 6.1.13 are
equivalent. Many of the implications use Lemma 6.1.5.
(a) Prove (a) (b) and (c) (d).
(b) Prove (b) (e) and (b) (f) (c).
(c) Prove (c) (b) by adapting the proof of (d) (b) from Lemma 6.1.5.
(d) Prove (b) (g) and use the obvious implication (e) (d) to complete the proof of
the lemma.
6.1.6. Let X

be the toric variety of a fan in N


R
R
n
and x
0
(n). Prove that the
natural map CDiv
T
(X

) Pic(X

) induces an isomorphism

D =

CDiv
T
(X

) [ a

= 0 for all
0
(1)

Pic(X

).
6.1.7. Prove that the toric variety X

of Example 6.1.17 is smooth.


6.1.8. For the following toric varieties X

, compute Pic(X

) and describe which torus-


invariant divisors are ample and which are basepoint free.
6.2. Polytopes and Projective Toric Varieties 277
(a) X

is the toric variety of the smooth complete fan in R


2
with
(1) =e
1
, e
2
, e
1
+e
2
.
(b) X

is the blowup Bl
p
(P
n
) of P
n
at a xed point p of the torus action.
(c) X

is the toric variety of the fan fromExercise 3.3.12. See Figure 12 fromChapter 3.
(d) X

is the toric variety of the fan obtained from the fan of Figure 12 from Chapter 3 by
combining the two upward pointing cones.
6.1.9. The toric variety (P
1
)
n
has ray generators e
1
, . . . , e
n
. Let D

1
, . . . , D

n
denote the
corresponding torus-invariant divisors. Consider D =

n
i=1
(a
+
i
D
+
i
+a

i
D

i
).
(a) Show that D is basepoint free if and only if a
+
i
+a

i
0 for all i.
(b) Show that D is ample if and only if a
+
i
+a

i
> 0 for all i.
6.1.10. Let D =

be an ample divisor on a complete toric variety X

. Dene
= Cone((u

, a

) [ (1)) N
R
R.
(a) Prove that is strongly convex.
(b) Prove that the boundary of is the graph of the support function
D
.
(c) Prove that is the set of cones obtained by projecting proper faces of onto M
R
.
6.1.11. Let be the fan from Example 4.2.13. Prove the X

is not projective.
6.2. Polytopes and Projective Toric Varieties
We begin with the set of polytopes

P M
R
[ P is a full dimensional lattice polytope

and the set of pairs

(X

, D) [ a complete fan in N
R
, D a torus-invariant ample divisor on X

.
These sets are related as follows.
Theorem 6.2.1. The maps P (X
P
, D
P
) and (X

, D) P
D
dene bijections be-
tween the above sets that are inverses of each other.
Proof. The map P (X
P
, D
P
) comes from Proposition 6.1.10, where we showed
that D
P
is an ample divisor on X
P
. Also recall from Proposition 3.1.6 that X
P
is
the toric variety of the normal fan
P
, which is a fan in N
R
. For (X

, D) P
D
,
we showed that P
D
M
R
is a full dimensional lattice polytope in the proof of
Theorem 6.1.14.
It remains to prove that these maps are inverses of each other. One direction is
easy, since P (X
P
, D
P
) P
D
P
= P, where the equality is Exercise 4.3.1. Going
the other way, we have (X

, D) P
D
(X
P
D
, D
P
D
) = (X

, D), where the equality


follows from (6.1.8) and (6.1.9) in the proof of Theorem 6.1.14.
The goal of this section is to look more deeply into the above relationship. In
particular, we are interested in the following questions:
278 Chapter 6. Line Bundles on Toric Varieties
Suppose P and Q are full dimensional lattice polytopes with X
P
= X
Q
. How
are P and Q related?
Suppose D is a torus-invariant Cartier divisor on X
P
that is basepoint free. How
are P and P
D
related?
The answers to these questions will involve generalized fans, pullbacks of divisors,
and Minkowski sums of polytopes.
Generalized Fans. The polytope P
D
of a basepoint free Cartier divisor D is a lattice
polytope by Theorem 6.1.7, but need not be full dimensional (see Exercise 6.1.3).
If we want P
D
to have a normal fan, we need to allow for more general fans. Here
is the denition we will use.
Denition 6.2.2. A generalized fan in N
R
is a nite collection of cones N
R
such that:
(a) Every is a rational polyhedral cone.
(b) For all , each face of is also in .
(c) For all
1
,
2
, the intersection
1

2
is a face of each (hence also in ).
This agrees with the denition of fan given in Denition 3.1.2, except that the
cones are no longer required be strongly convex. The denitions of support and
complete extend to generalized fans in the obvious way. A generalized fan that
is a ordinary fan is called nondegenerate; otherwise is degenerate. Generalized
fans will play an important role in Chapters 14 and 15.
Let be a generalized fan. Then
0
=

is the minimal cone in . It


has no proper faces and hence must be a subspace of N
R
. Let N = N/(
0
N) with
quotient map : N N. You will prove the following in Exercise 6.2.1:
is a fan if and only if
0
=0.
For , = /
0
N
R
/
0
= N
R
is a strongly convex rational polyhedral
cone such that =
1
R
().
= [ is a fan in N
R
.
The toric variety X

of the generalized fan is dened to be the toric variety of


the usual fan , i.e., X

= X

.
The Normal Fan of a Lattice Polytope. Some of most interesting generalized fans
come from polyhedra. Let P M
R
be a lattice polytope. We do not assume that P
is full dimensional. A vertex v P gives the cone
C
v
= Cone(PMv) M
R
.
Similar to 2.3, the dual cone
v
= C

v
N
R
is a rational polyhedral cone, and
these cones give a generalized fan as follows (Exercise 6.2.2).
6.2. Polytopes and Projective Toric Varieties 279
Proposition 6.2.3. Given a lattice polytope P M
R
, the set

P
= [ _
v
, v is a vertex of P
is a complete generalized fan in N
R
. Furthermore:
(a) The minimal cone of
P
is the dual of Span(mm

[ m, m

PM) M
R
.
(b)
P
is a fan if and only if P M
R
is full dimensional.
We call
P
the normal fan of P. The toric variety X
P
is then dened to be the
toric variety of the generalized fan X

P
, i.e., X
P
= X

P
.
Example 6.2.4. Let P M
R
be a line segment whose vertices are lattice points.
The cone C
v
at each vertex is a ray, so that the normal fan
P
consists of two
closed half-spaces and the hyperplane where they intersect. Taking the quotient by
this hyperplane gives the usual fan for P
1
, so that X
P
=P
1
.
The Normal Fan of a Basepoint Free Divisor. If is a complete fan in N
R
R
n
and D =

has no basepoints and Cartier data m

(n)
, then P
D
is a
lattice polytope with the m

as vertices. We can describe the normal fan


P
D
of P
D
as follows.
Proposition 6.2.5. Let D =

be a basepoint free Cartier divisor on X

with polytope P
D
. Then:
(a) If v P
D
is a vertex, then the corresponding cone
v
=C

v
in the normal fan

P
D
is the union

v
=

(n)
m=v
.
(b) is a renement
P
D
.
Proof. Part (b) follows immediately from part (a). Let v P
D
be a vertex. Since
C
v
= Cone(P
D
Mv) is strongly convex, its dual
v
= C

v
has dimension n in
N
R
. It follows that part (a) is equivalent to the assertion
(6.2.1) for all (n), Int() Int(
v
) = implies m

=v,
where Int denotes the interior (Exercise 6.2.3). Also note that any u
v
satises
'mv, u` 0, for all m P
D
M.
In particular, m

P
D
for (n) since D is basepoint free, so that
(6.2.2) 'm

, u` 'v, u`, for all (n).


We now prove (6.2.1). Assume Int() Int(
v
) = and let u be an element of
the intersection. Since v = m

for some

(n), we have
'v, u`
D
(u) ='m

, u`
280 Chapter 6. Line Bundles on Toric Varieties
by convexity and part (b) of Lemma 6.1.5. Combining this with (6.2.2), we see that
'm

, u` ='v, u`, for all u Int() Int(


v
).
Since Int() Int(
v
) is open, this forces m = m

, proving (6.2.1).
This proposition gives a nice way to think about the normal fan
P
D
. One
begins with the Cartier data m

(n)
of Dand then combines all cones (n)
whose m

s give the same vertex of P


D
. These combined cones and their faces
satisfy the conditions for being a fan, except that strong convexity fails when P
D
is
not full dimensional. Here is an example of how this works.
Example 6.2.6. For the Hirzebruch surface H
2
, consider the divisors D = D
4
and
D

= D
1
. The polytope P
D
from Figure 4 of Example 6.1.2 is shown on the left in
Figure 10 on the next page. By Proposition 6.2.5, m
1
= m
4
tells us to combine
1
and
4
, as shown on the right in Figure 10. Thus the normal fan of P
D
is a fan with
three maximal cones.
P
D
m
2
m
1
= m
4
m
3
2
1
u
2
u
4
u
3
u
1
= (1,2)

4
Figure 10. PD (left) and its normal fan (right)
The polytope P
D
is the line segment shown on the left in Figure 11. Here, we
combine
1
and
2
(since m
1
= m
2
) and also combine
3
and
4
(since m
3
= m
4
).
This gives the degenerate normal fan shown on the right in Figure 11. Thus the
toric variety of P
D
is P
1
.
P
D
m
3
= m
4
m
1
= m
2
1
u
2
u
4
u
3
u
1
= (1,2)

4
Figure 11. P
D
(left) and its degenerate normal fan (right)
6.2. Polytopes and Projective Toric Varieties 281
Pulling Back via Toric Morphisms. In order to understand the full implications of
Proposition 6.2.5, we need the following description of pullbacks of torus-invariant
Cartier divisors by toric morphisms.
Proposition 6.2.7. Assume that : X

1
X

2
is the toric morphism induced by
: N
1
N
2
, and let D
2
be a torus-invariant Cartier divisor with support function

D
2
: [
2
[ R. Then there is a unique torus-invariant Cartier divisor D
1
on X

1
with the following properties:
(a) O
X

1
(D
1
)

O
X

2
(D
2
).
(b) The support function
D
1
is the composition
[
1
[

[
2
[

D
2
R.
Proof. Let the local data of D
2
be (U

,
m
)

2
, where now refers to an
arbitrary cone of
2
. Recall that the minus sign comes from 'm

, u

` =a

when
(1). Then the proof of Theorem 6.0.18 shows that O
X

2
(D
2
) is the sheaf of
sections of a rank 1 vector bundle V X

2
with transition functions
g

=
m m
.
Now take


1
and let
2
be the smallest cone satisfying
R
(

) .
Using the dual map

: M
2
M
1
, we set
m

(m

).
Since (U

) U

, one can show without difculty that


g

=
m

O
X

1
(U

.
Then (U

,
m

1
is the local data for a Cartier divisor D
1
on X

1
. It is
straightforward to verify that D
1
has the required properties (Exercise 6.2.4).
In the situation of Proposition 6.2.7, we call D
1
is the pullback of D
2
via
since O
X

1
(D
1
) is the pullback of O
X

2
(D
2
) via . We denote this by D
1
=

D
2
.
The Structure of Basepoint Free Divisors. Proposition 6.2.5 shows that renes
the normal fan
P
D
. Hence we should have a toric morphism X

X
P
D
. This is
certainly true when
P
D
is nondegenerate, and as we will see below, it remains
true when
P
D
is degenerate. More importantly, D is (up to linear equivalence) the
pullback of an ample divisor on X
P
D
via this morphism.
Theorem 6.2.8. Let D be a basepoint free Cartier divisor on a complete toric va-
riety, and let X
D
be the toric variety of the polytope P
D
M
R
. Then the renement
of
P
D
induces a proper toric morphism
: X

X
P
D
.
Furthermore, D is linearly equivalent to the pullback via of the ample divisor on
X
P
D
coming from P
D
.
282 Chapter 6. Line Bundles on Toric Varieties
Proof. The minimal cone
0
of
P
D
is a subspace of N
R
. Let N = N/(
0
N),
with quotient map : N N. Since renes
P
D
and
P
D
projects to a genuine
fan in N
R
, it follows that induces a toric morphism as claimed. Note also that
is proper since X

and X
P
D
are complete.
Let M M be dual to : N N. Part (a) of Proposition 6.2.3 implies that
M
R
= Span(mm

[ m, m

P
D
M. Translating P
D
by a lattice point, we may
assume that P
D
M
R
. This changes our original divisor D by a linear equivalence.
The polytope P
D
gives the ample divisor D = D
P
D
on X
P
D
. Since D is basepoint
free, Theorem 6.1.7 implies that

D
(u) = min
mP
D
'm, u`.
Using P
D
M
R
, one sees that
D
factors through : N N, and in fact,

D
=
D

R
(Exercise 6.2.5). By Proposition 6.2.7, D is the pullback of D = D
P
D
via .
Theorem 6.2.8 implies that a Cartier divisor without basepoints on a complete
toric variety has a very nice structure: it is linearly equivalent to the pullback (via
a toric morphism) of an ample divisor on a projective toric variety of possibly
smaller dimension. This will be useful when we study the geometric invariant
theory of toric varieties in Chapters 14 and 15.
Here are two examples to illustrate what can happen in Theorem 6.2.8.
Example 6.2.9. The toric variety X

of Example 6.1.17 has no ample divisors, but


it does have nontrivial basepoint free divisors. The ray generators of are
e
1
, e
2
, e
3
, a, b, c, d,
with corresponding toric divisors
D

1
, D

2
, D

3
, D
a
, D
b
, D
c
, D
d
.
Then one can show that
D = 2D

1
+2D

2
+2D

3
D
a
D
b
D
c
D
d
is basepoint free (Exercise 6.2.6). Thus the support function
D
is convex.
Figure 9 in Example 6.1.17 shows that Cone(e
1
, e
2
, d) is a union of three cones
of . Using
D
(e
1
) =
D
(e
2
) = 0 and
D
(a) =
D
(b) =
D
(d) = 1, one sees that
these three cones all have m

= e
3
(Exercise 6.2.6). Hence we should combine
these three cones. The same thing happens in Cone(e
1
, e
3
, d) and Cone(e
2
, e
3
, d).
In the rst orthant, the fan of X
P
D
looks like Figure 12 on the next page when
intersected with x +y +z = 1. Hence X
P
D
is the blowup of (P
1
)
3
at the point
corresponding to the rst orthant (Exercise 6.2.6). Also, : X

X
P
D
is a proper
birational toric morphism since renes the (nondegenerate) normal fan
P
D
.
6.2. Polytopes and Projective Toric Varieties 283
e
1
e
2
e
3
d
Figure 12. Combined cones of lying in R
3
0
in Example 6.2.9
Example 6.2.10. Consider the divisor D = 3D
1
+D
2
D
4
= D
1
+div(
e
2
) on
the Hirzebruch surface H
2
. Then P
D
= Conv(e
2
, e
1
e
2
) = Conv(0, e
1
) e
2
.
This gives the degenerate normal fan shown in Figure 11 of Example 6.2.6, and
: X

X
P
D
= P
1
is the toric morphism from Example 3.3.5. Then D D
1
,
which is the pullback of an ample divisor on P
1
.
N-Minkowski Summands. We now return to the questions asked at the beginning
of the section. In terms of normal fans, the answers are easy to give:
Full dimensional lattice polytopes P and Q in M
R
give the same toric variety if
and only if they have the same normal fan.
If D is a torus-invariant basepoint free Cartier divisor on X
P
, then the normal
fan of P renes the normal fan of P
D
by Proposition 6.2.5.
By rephrasing this in terms of Minkowski sums, we can state both of these purely
in the language of polytopes. Here is the denition.
Denition 6.2.11. Given lattice polytopes P and Q in M
R
, Q is an N-Minkowski
summand of P if
Q+Q

= kP,
where k N is positive and Q

M
R
is a lattice polytope.
Example 6.2.12. The rectangle Q = Conv(0, 2e
1
, e
2
, 2e
1
+e
2
) is an N-Minkowski
summand of the hexagon P=Conv(0, e
1
, e
2
, 2e
1
+e
2
, e
1
+2e
2
, 2e
1
+2e
2
), as shown
by Figure 13.
0
+
0
=
0
Q Q 2P
Q
Q+e
2
Figure 13. Q is an N-Minkowski summand of P since Q+Q

=2P
284 Chapter 6. Line Bundles on Toric Varieties
Minkowski sums are related to normal fans as follows [28, Prop. 1.2].
Proposition 6.2.13. Let P and Q be lattice polytopes in M
R
. Then:
(a) Q is an N-Minkowski summand of P if and only if
P
renes
Q
.
(b)
P+Q
is the coarsest common renement of
P
and
Q
, i.e., any fan that
renes
P
and
Q
also renes
P+Q
.
Proposition 6.2.13 does not assume that the lattice polytopes P and Q have full
dimension, so the normal fans
P
and
Q
in the proposition may be degenerate.
Also note that
P+Q
is common renement of
P
and
Q
by part (a). So the point
of part (b) is that
P+Q
is the most efcient common renement.
We can now describe when two polytopes give the same toric variety.
Corollary 6.2.14. Full dimensional lattice polytopes in M
R
give the same toric
variety if and only if each is an N-Minkowski summand of the other.
Proof. This follows immediately from Proposition 6.2.13 since two fans are equal
if and only if each renes the other.
We also have the following lovely result about basepoint free divisors.
Corollary 6.2.15. Let P be a full dimensional lattice polytope in M
R
. Then a
polytope Q M
R
is an N-Minkowski summand of P if and only if there is a torus-
invariant basepoint free Cartier divisor D on X
P
such that Q = P
D
.
Proof. If D is basepoint free on X
P
, then Propositions 6.2.5 and 6.2.13 imply that
P
D
is an N-Minkowski summand of P. For the converse, suppose that Q is an N-
Minkowski summand of P. Then
P
renes
Q
. We will write the maximal cones
of
Q
as
v
for v Q a vertex. Also let n = dim X
P
. We dene D as follows. Each

P
(n) is contained in
v
for some vertex v Q. Then D is the Cartier divisor
on X
P
whose Cartier data m

P
(n)
is dened by m

= v when
v
.
Thus D =

P
(1)
a

, where a

= 'v, u

` when u


v
. To prove that
P
D
= Q, take m P
D
, so that 'm, u

` a

for all
P
(1). This implies
'mv, u

` ='m, u

` +a

0 for all u


v
.
These u

s generate
v
since
P
renes
Q
, so that mv

v
=C
v
. Hence
m

v is a vertex of Q
(C
v
+v) = Q,
where the equality follows from Exercise 6.2.7. The opposite inclusion Q P
D
is
straightforward and hence is left to the reader. This proves P
D
= Q, and then D is
basepoint free by Proposition 6.1.1.
Example 6.2.16. Consider the rectangle Q and the hexagon P dened in Exam-
ple 6.2.12. Since Q is an N-Minkowski summand of P, it gives a basepoint free
divisor D on X
P
. Let the ray generators u
1
, . . . , u
6
of
P
be arranged clockwise
6.2. Polytopes and Projective Toric Varieties 285
around the origin, starting with u
1
= e
2
. Then the recipe for D given in the proof
of Corollary 6.2.15 makes it easy to show that
D = D
3
+D
4
+2D
5
+D
6
,
where D
i
is the toric divisor corresponding to u
i
(Exercise 6.2.8).
Zonotopes. Recall from Example 2.3.10 that a zonotope is a Minkowski sum of
line segments. Here we show that zonotopes have especially nice normal fans. A
central hyperplane arrangement in N
R
consists of nitely many rational hyper-
planes H N
R
whose intersection is the origin. This determines a fan in N
R
whose
maximal cones are the closures of the connected components of the complement
of the arrangement.
Example 6.2.17. The hexagon P from Example 6.2.12 is a zonotope since P =
Conv(0, e
1
) +Conv(0, e
2
) +Conv(0, e
1
+e
2
). Figure 14 reproduces Figure 7 from
P
v
1
v
2
v
3
v
4
v
5
v
6

P
Figure 14. A zonotope P and its normal fan P
Example 2.3.10. As you can see, the normal fan of P comes from an arrangement
of three lines through the origin in R
2
.
Proposition 6.2.18. The normal fan of a full dimensional lattice zonotope P comes
from a central hyperplane arrangement.
Proof. First note that a Minkowski sum of parallel line segements is again a line
segment. Thus we can write P=L
1
+ +L
s
as a Minkowski sum of line segments
where no two segments are parallel. Each normal fan
L
i
is determined by the
hyperplane normal H
i
to L
i
, as explained in Example 6.2.4. By Proposition 6.2.13,

P
=
L
1
++Ls
is the coarsest common renement of
L
1
,. . . ,
Ls
. This is clearly
the fan determined by the central hyperplane arrangement H
1
, . . . , H
s
. Note that
that H
1
H
s
=0 since
P
is a nondegenerate fan.
See [281, Thm. 7.16] for a different proof of Proposition 6.2.18 that uses linear
programming.
286 Chapter 6. Line Bundles on Toric Varieties
Exercises for 6.2.
6.2.1. Prove the properties of generalized fans stated in three bullets in the discussion
following Denition 6.2.2.
6.2.2. Prove Proposition 6.2.3.
6.2.3. Prove that part (a) of Propostion 6.2.5 follows from (6.2.1).
6.2.4. Complete the proof of Proposition 6.2.7.
6.2.5. Complete the proof of Theorem 6.2.8.
6.2.6. This exercise deals with Example 6.2.9.
(a) Let D = 2D

1
+2D

2
+2D

3
D
a
D
b
D
c
D
d
be the divisor from Example 6.2.9.
Prove that P
D
is the polytope with 10 vertices
e
1
, e
2
, e
3
, 2e
1
, 2e
2
, 2e
3
, 2e
1
+2e
2
, 2e
1
+2e
3
, 2e
2
+2e
3
, 2e
1
+2e
2
+2e
3
and conclude that D is basepoint free.
(b) In Example 6.2.9, we asserted that certain maximal cones of must be combined to
get the maximal cones of
PD
. Prove that this is correct.
(c) Show that X
PD
is the blowup of (P
1
)
3
at the point corresponding to the rst orthant.
6.2.7. This exercise is concerned with the proof of Corollary 6.2.15.
(a) Given a lattice polytope Q M
R
, let C
v
= Cone(QMv) for v Q a vertex. Prove
that

v is a vertex of Q
(C
v
+v) = Q.
(b) Complete the proof of the corollary by showing Q P
D
.
6.2.8. In Example 6.2.16, the rectangle Q is an N-Minkowski summand of the hexagon P.
(a) In the example, we claimed that D = D
3
+D
4
+2D
5
+D
6
. Prove this.
(b) Let Q

= Conv(0, e
2
, 2e
1
+2e
2
, 2e
1
+3e
2
). Prove carefully that Q+Q

=2P and com-


pute the basepoint free divisor D

determined by Q

.
6.2.9. Suppose that full dimensional lattice polytopes P, QM
R
give the same toric variety
X

. Prove that P+Q also gives X

.
6.2.10. Let P M
R
be a full dimensional lattice polytope. A face Q_P determines a cone

Q
in the normal fan of P. This gives the orbit closure V(
Q
) X
P
, and V(
Q
) X
Q
by
Proposition 3.2.9. This gives an inclusion i : X
Q
X
P
which is not a toric morphism when
Q P. Prove that i

O
XP
(D
P
) O
XQ
(D
Q
).
6.3. The Nef and Mori Cones
In 6.1, we gave some nice criteria for when a Cartier divisor D is basepoint free
or ample. We now study the structure of the set of basepoint free divisors and the
set of ample divisors inside Pic(X

)
R
= Pic(X

)
Z
R.
The main concept of this section is that of numerical effectivity. Roughly
speaking, the goal is to dene a pairing between divisors and curves, such that
for a Cartier divisor D and complete curve C on a variety X, the number D C
counts the number of points of DC, with appropriate multiplicity.
6.3. The Nef and Mori Cones 287
Example 6.3.1. Suppose X = P
2
with homogeneous coordinates x, y, z, and let
D = V(y) and C = V(zyx
2
). Then D and C meet at the single point p = (0, 0, 1),
where they share a common tangent. If we replace D with the linearly equivalent
divisor E = V(y z), then clearly E and C meet in two points. This suggests that
the point p = DC should be counted twice, since it is a tangent point. Hence
we should have D C = 2.
Despite this encouraging example, there are several technical hurdles to over-
come in order to make this precise in a general setting. Note that in C
2
, two lines
may or may not meet, so to get a reasonable theory, we will work with complete
curves C on a normal variety X. We also need to restrict to Cartier divisors D on X.
With these assumptions, the intersection product DC should possess the following
properties:
(D+E) C = D C+E C.
D C = E C when DE.
Let D be a prime divisor on X such that DC is nite. Assume each p DC
is smooth in C, D, X and that the tangent spaces T
p
(C) T
p
(X) and T
p
(D)
T
p
(X) meet transversely. Then D C =[DC[.
Using these properties, one can give a rigorous proof of the computation D C = 2
from Example 6.3.1.
The Degree of a Line Bundle. The key tool we will use is the notion of the degree
of a divisor on an irreducible smooth complete curve C. Such a divisor can be
written as a nite sum D =

i
a
i
p
i
where a
i
Z and p
i
C.
Denition 6.3.2. Let D =

i
a
i
p
i
be a divisor on an irreducible smooth complete
curve C. Then the degree of D is the integer
deg(D) =

i
a
i
Z.
Note the obvious property deg(D+E) = deg(D) +deg(E). The following key
result is proved in [131, Cor. II.6.10].
Theorem 6.3.3. Every principal divisor on an irreducible smooth complete curve
has degree zero.
In other words, deg(div( f )) = 0 for all nonzero rational functions f on an
irreducible smooth complete curve C. Thus
deg(D) = deg(E) when D E on C,
and the degree map induces a surjective homomorphism
deg : Pic(C) Z.
Note that all Weil divisors are Cartier since C is smooth.
288 Chapter 6. Line Bundles on Toric Varieties
In 6.0 we showed that Pic(C) is the set of isomorphism classes of line bundles
on C. Hence we can dene the degree deg(L) of a line bundle L on C. This leads
immediately to the following result.
Proposition 6.3.4. Let C be an irreducible smooth complete curve. Then a line
bundle L has a degree deg(L) such that L deg(L) has the following prop-
erties:
(a) deg(L L

) = deg(L) +deg(L

).
(b) deg(L) = deg(L

) when L L

.
(c) deg(L) = deg(D) when L O
C
(D).
The Normalization of a Curve. We dened the normalization of an afne variety
in 1.0, and by gluing together the normalizations of afne pieces, one can dene
the normalization of any variety (see [131, Ex. II.3.8]). In particular, an irreducible
curve C has a normalization map
: C C,
where C is an normal variety. Here are the key propeties of C.
Proposition 6.3.5. Let C be the normalization of an irreducible curve C. Then:
(a) C is smooth.
(b) C is complete whenever C is complete.
Proof. Since C is a curve, Proposition 4.0.17 implies that C is smooth. Part (b) is
covered by [131, Ex. II.5.8].
One can prove that every irreducible smooth complete curve is projective. See
[131, Ex. II.5.8].
The Intersection Product. We now have the tools needed to dene the intersec-
tion product. Let X be a normal variety. Given a Cartier divisor D on X and an
irreducible complete curve C X, we have
The line bundle O
X
(D) on X.
The normalization : C C.
Then

O
X
(D) is a line bundle on the irreducible smooth complete curve C.
Denition 6.3.6. The intersection product of D and C is D C = deg(

O
X
(D)).
Here are some properties of the intersection product.
Proposition 6.3.7. Let C be an irreducible complete curve and D, E Cartier divi-
sors on a normal variety X. Then:
(a) (D+E) C = D C+E C.
(b) D C = E C when D E.
6.3. The Nef and Mori Cones 289
Proof. The pullback of line bundles is compatible with tensor product, so that
part (a) follows from (6.0.3) and Proposition 6.3.4. Part (b) is an easy consequence
of Propositions 6.0.22 and 6.3.4.
The intersection product extends to Q-Cartier divisors as follows. Recall from
Chapter 4 that a Weil divisor D is Q-Cartier if D is Cartier for some integer > 0.
Given an irreducible complete curve C X, let
(6.3.1) D C =
1

(D) C Q.
In Exercise 6.3.1 you will show that this intersection product is well-dened and
satises Propostion 6.3.7.
Intersection Products on Toric Varieties. In the toric case, DC is easy to compute
when D and C are torus-invariant in X

. In order for C to be torus-invariant and


complete, we must have C =V() =O(), where =

(n1) is the wall


separating cones ,

(n), n = dim X

. We do not assume is complete.


In this situation, we have the sublattice N

=Span()N N and the quotient


N() = N/N

. Let and

be the images of and

in N()
R
. Since is a wall,
N() Z and ,

are rays that correspond to the rays in the usual fan for P
1
. In
particular, V() P
1
is smooth, so no normalization is needed when computing
the intersection product.
Proposition 6.3.8. Let C = V() be the complete torus-invariant curve in X

coming from the wall =

. Let D be a Cartier divisor with Cartier data


m

, m

M corresponding to ,

(n). Also pick u

N that maps to the


minimal generator of

N()
R
. Then
D C ='m

, u` Z.
Proof. Since V() U

, we can assume X

= U

and is the fan


consisting of ,

and their faces. We also have


D[
U
= div(
m
)[
U
, D[
U

= div(
m

)[
U

.
The proof of Proposition 6.2.7 shows that the line bundle O
X

(D) is determined by
the transition function g

=
mm

. Thus
D C = deg(i

O
X

(D)),
where i : V() X

is the inclusion map. The pullback bundle is determined by


the restriction of g

to
V() U

=V() U

= O(),
where O() is the T
N
-orbit corresponding to . This is also the torus of the toric
variety V() = O(). In Lemma 3.2.5, we showed that

M is the dual of N()


and that
O() Hom
Z
(M

, C

) T
N()
.
290 Chapter 6. Line Bundles on Toric Varieties
Now comes the key observation: since the linear functions given by m

, m

agree
on , we have m

M. Thus i

O
X

(D) is the line bundle on V()


whose transition function is g

=
mm

for m

M.
It follows that i

O
X

(D) O
V()
(D), where D is the divisor on V() given by
the Cartier data
m

= 0, m

= m

.
Let p

, p

be the torus xed points corresponding to ,

. Since u

N maps
to the minimal generator u

N(), we have
D ='m

, u` p

+'m

, u` p

='m

, u` p

,
where the second equality follows from m

= m

M. Hence
D C = deg(i

O
X

(D)) = deg(D) ='m

, u`.
Example 6.3.9. Consider the toric surface whose fan in R
2
has ray generators
u
1
= e
1
, u
2
= e
2
, u
0
= 2e
1
+3e
2
and maximal cones
= Cone(u
1
, u
0
),

= Cone(u
2
, u
0
).
The support of is the rst quadrant and =

=Cone(u
0
) gives the complete
torus-invariant curve C =V() X

.
If D
1
, D
2
, D
0
are the divisors corresponding to u
1
, u
2
, u
0
, then
D = aD
1
+bD
2
+cD
0
is Cartier 2a+3b c mod 6.
When this condition is satised, we have
m

=ae
1
+
2ac
3
e
2
, m

=
3bc
2
e
1
be
2
.
Also, u =e
1
+2e
2

maps to the minimal generator of

since u, u
0
form a basis
of Z
2
. (You will check these assertions in Exercise 6.3.2.) Thus
D C ='m

, u` =
2a+3bc
6
by Proposition 6.3.8. Note that D is Q-Cartier since is simplicial. Then (6.3.1)
shows that the formula for D C holds for arbitrary integers a, b, c. In particular,
D
1
C =
1
3
, D
2
C =
1
2
, D
0
C =
1
6
.
In the next section we will see that these intersection products follow directly from
the relation u
0
+2u
1
+3u
2
= 0 and the fact that Zu
0
= Span() Z
2
.
6.3. The Nef and Mori Cones 291
Nef Divisors. We now dene an important class of Cartier divisors.
Denition 6.3.10. Let X be a normal variety. Then a Cartier divisor D on X is nef
(short for numerically effective) if
D C 0
for every irreducible complete curve C X.
A divisor linearly equivalent to a nef divisor is nef. Here is another result.
Proposition 6.3.11. Every basepoint free divisor is nef.
Proof. The pullback of a line bundle generated by global sections is generated by
global sections (Exercise 6.0.10). Thus, given : C C and D basepoint free,
the line bundle L =

(O
X
(D)) is generated by global sections. This allows us to
write L = O
C
(D

) for a basepoint free divisor D

on C. A nonzero global section


of O
C
(D

) gives an effective divisor E

linearly equivalent to D

. Then
D C = deg(

(O
X
(D))) = deg(O
C
(D

)) = deg(D

) = deg(E

) 0,
where the last inequality follows since E

is effective.
In the toric case, nef divisors are especially easy to understand.
Theorem 6.3.12. Let D be a Cartier divisor on a toric variety X

whose fan has


convex support of full dimension. The following are equivalent:
(a) D is basepoint free, i.e., O
X
(D) is generated by global sections.
(b) D is nef.
(c) D C 0 for all torus-invariant irreducible complete curves C X.
Proof. The rst item implies the second by Proposition 6.3.11, and the second
item implies the third by the denition of nef. So suppose that D C 0 for all
torus-invariant irreducible curves C. We can replace D with a linearly equivalent
torus-invariant divisor. Then, by Theorem 6.1.7, it sufces to show that
D
is
convex.
Take a wall =

of and set C = V(). If we pick u

N as in
Proposition 6.3.8, then
'm

, u` = D C 0,
so that
'm

, u` 'm

, u` =
D
(u).
Note that u / since the image of u is nonzero in N() = N/(Span() N). Then
Lemma 6.1.5 implies that
D
is convex.
A variant of the above proof leads to the following ampleness criterion, which
you will prove in Exercise 6.3.3.
292 Chapter 6. Line Bundles on Toric Varieties
Theorem 6.3.13 (Toric Kleiman Criterion). Let D be a Cartier divisor on a com-
plete toric variety X

. Then Dis ample if and only if DC>0 for all torus-invariant


irreducible curves C X

.
Note that one direction of the proof follows from the general fact that on any
complete normal variety, an ample divisor D satises D C > 0 for all irreducible
curves C X (Exercise 6.3.4).
Theorems 6.3.12 and 6.3.13 were well-known in the smooth case and proved
more recently (and independently) in [185], [197] and [212] in the complete case.
Numerical Equivalence of Divisors. The intersection product leads to an impor-
tant equivalence relation on Cartier divisors.
Denition 6.3.14. Let X be a normal variety.
(a) A Cartier divisor D on X is numerically equivalent to zero if D C = 0 for all
irreducible complete curves C X.
(b) Cartier divisors D and E are numerically equivalent, written D E, if DE
is numerically equivalent to zero.
What does this say in the toric case?
Proposition 6.3.15. Let D be a Cartier divisor on a toric variety X

whose fan
has convex support of full dimension. Then D 0 if and only if D 0.
Proof. Clearly if D is principal then D is numerically equivalent to zero. For the
converse, assume D 0 and let =

be a wall of . If we pick u

as in
Proposition 6.3.8, then
0 = D C ='m

, u`
for C =V(). This forces m

= m

since m

and u / . From here,


one sees that m

= m

for all ,

(n), and it follows that D is principal.


Numerical Equivalence of Curves. We also get an interesting equivalence relation
on curves. Let Z
1
(X) be the free abelian group generated by irreducible complete
curves C X. An element of Z
1
(X) is called a proper 1-cycle.
Denition 6.3.16. Let X be a normal variety.
(a) A proper 1-cycle C on X is numerically equivalent to zero if D C = 0 for all
Cartier divisors D on X.
(b) Proper 1-cycles C and C

are numerically equivalent, written C C

, if CC

is numerically equivalent to zero.


The intersection product (D,C) D C extends naturally to a pairing
CDiv(X) Z
1
(X) Z.
6.3. The Nef and Mori Cones 293
between Cartier divisors and 1-cycles. In order to get a nondegenerate pairing, we
work over R and mod out by numerical equivalence.
Denition 6.3.17. For a normal variety X, dene
N
1
(X) = (CDiv(X)/)
Z
R and N
1
(X) = (Z
1
(X)/)
Z
R.
It follows easily that we get a well-dened nondegenerate bilinear pairing
N
1
(X) N
1
(X) R.
A deeper fact is that N
1
(X) and N
1
(X) have nite dimension over R. Thus N
1
(X)
and N
1
(X) are dual vector spaces via intersection product.
The Nef and Mori Cones. The vector spaces N
1
(X) and N
1
(X) contain some in-
teresting cones.
Denition 6.3.18. Let X be a normal variety.
(a) Nef(X) is the cone in N
1
(X) generated by classes of nef Cartier divisors. We
call Nef(X) the nef cone.
(b) NE(X) is the cone in N
1
(X) generated by classes of irreducible complete
curves.
(c) NE(X) is the closure of NE(X) in N
1
(X). We call NE(X) the Mori cone.
Here are some easy observations about the nef and Mori cones.
Lemma 6.3.19.
(a) Nef(X) and NE(X) are closed convex cones and are dual to each other, i.e.,
Nef(X) = NE(X)

and NE(X) = Nef(X)

.
(b) NE(X) has full dimension in N
1
(X).
(c) Nef(X) is strongly convex in N
1
(X).
Proof. It is obvious that Nef(X), NE(X) and NE(X) are convex cones, and Nef(X)
is closed since it is dened by inequalities of the form D C 0. In fact,
Nef(X) = NE(X)

by the denition of nef. Then Nef(X) =NE(X)

follows easily. In general, NE(X)


need not be closed. However, since the closure of a convex cone is its double dual,
we have
NE(X) = NE(X)

= Nef(X)

.
Note that the cone NE(X) has full dimension since N
1
(X) is spanned by the classes
of irreducible complete curves. Hence the same is true for its closure NE(X). Then
Nef(X) is strongly convex since its the dual has full dimension.
294 Chapter 6. Line Bundles on Toric Varieties
Let X

be a toric variety whose fan has convex support of full dimension


and set Pic(X

)
R
= Pic(X

)
Z
R. We have an inclusion
Pic(X

) Pic(X

)
R
since Pic(X

) is torsion-free by Proposition 4.2.5. Thus Pic(X

) is a lattice in the
vector space Pic(X

)
R
. Note also that
(6.3.2) Pic(X

)
R
= N
1
(X

)
since numerical and linear equivalence coincide by Proposition 6.3.15. In this case,
we will write Pic(X

)
R
instead of N
1
(X

).
Theorem 6.3.20. Let X

be a toric variety whose fan has convex support of full


dimension. Then:
(a) Nef(X

) is a rational polyhedral cone in Pic(X

)
R
.
(b) NE(X

) = NE(X

) is a rational polyhedral cone in N


1
(X

). Furthermore,
NE(X

) =

a wall of
R
0
[V()],
where [V()] N
1
(X

) is the class of V().


Proof. Part (a) is an immediate consequence of part (b). For part (b), let =

a wall of
R
0
[V()] and note that is a rational polyhedral cone contained in
NE(X

). Furthermore, Theorem 6.3.12 easily implies Nef(X

) =

. Then
NE(X

) = Nef(X

= NE(X

) NE(X

),
where the third equality follows since is polyhedral.
The formula from part (b) of Theorem 6.3.20, namely
NE(X

) =

a wall of
R
0
[V()],
is called the Toric cone theorem. Although the Mori cone equals NE(X

) in this
case, we will continue to write NE(X

) for consistency with the literature. For the


same reason we write R
0
[V()] instead of Cone([V()]).
The Mori cone of an arbitrary normal variety can have a complicated structure.
The cone theorem shows that some parts of the Mori cone are locally polyhedral.
See [179, Ch. 3] and [194, Ch. 7] for a discussion of this important result.
Since every irreducible complete curve C X

gives a class in NE(X

), we
get the following corollary of the toric cone theorem.
Corollary 6.3.21. Assume the fan has convex support of full dimension. Then
any irreducible complete curve on X

is numerically equivalent to a non-negative


linear combination of torus-invariant complete curves.
When X

is projective we can say more about the nef and Mori cones.
6.3. The Nef and Mori Cones 295
Theorem 6.3.22. Let X

be a projective toric variety. Then:


(a) Nef(X

) and NE(X

) are dual strongly convex rational polyhedral cones of


full dimension.
(b) A Cartier divisor D is ample if and only if its class in Pic(X

)
R
lies in the
interior of Nef(X

).
Proof. By hypothesis, X

has an ample divisor D. Then D C > 0 for every irre-


ducible curve in X

. This easily implies that the class of D lies in the interior of


Nef(X

). Thus Nef(X

) has full dimension and hence its dual NE(X

) is strongly
convex. When combined with Lemma 6.3.19, part (a) follows easily.
The strict inequality D C > 0 also shows that every irreducible curve gives a
nonzero class in N
1
(X

). Now suppose that the class of D is in the interior of the


nef cone. Then [D] denes a supporting hyperplane of the origin of the dual cone
NE(X

). Since 0 = [C] NE(X

) for every irreducible curve C X

, we have
D C > 0 for all such C. Hence D is ample by Theorem 6.3.13.
It follows that NE(X

) is strongly convex in the projective case. The rays of


NE(X

) are called extremal rays, which by the toric cone theorem are of the form
R
0
[V()]. The corresponding walls are called extremal walls.
Here is an example of the cones Nef(X

) and NE(X

).
Example 6.3.23. For the Hirzebruch surface H
r
, we showed in Example 6.1.16
that Pic(H
r
) =a[D
3
] +b[D
4
] [ a, b Z. Figure 15 shows Nef(H
r
) and NE(H
r
).
[D
4
]
[D
3
]
(1,0)
(0,1)
nef cone
[V(
4
)]
[V(
3
)] = [V(
1
)]
[V(
2
)]
Mori cone
(r,1)
(1,0)
Figure 15. The nef and Mori cones of Hr
Here,
i
= Cone(u
i
), so that D
i
= V(
i
). Using both notations helps distinguish
between Nef(H
r
) (built from divisors) and NE(H
r
) (built from curves).
The description of the nef cone follows from the characterization of ample
divisors on H
r
given in Example 6.1.16. The Mori cone is generated by the classes
of the V(
i
) by the toric cone theorem. Using the the basis given by D
3
=V(
3
),
296 Chapter 6. Line Bundles on Toric Varieties
D
4
=V(
4
) and the linear equivalences
D
1
D
3
, D
2
rD
3
+D
4
from Example 6.1.16, we get the picture of NE(H
r
) shown in Figure 15. It follows
that [V(
2
)] and [V(
3
)] = [V(
1
)] generate extremal rays, while [V(
4
)] does not.
Thus
1
,
2
,
3
are extremal walls.
The explicit duality between the cones Nef(X

) and NE(X

) in Figure 15 will
be described in the next section.
Theorem 6.3.22 tells us that ample divisors correspond to lattice points in the
interior of Nef(H
r
). Thus lattice points on the boundary correspond to divisors
that are basepoint free but not ample. We can see this vividly by looking at the
polytopes P
D
associated to divisors D whose classes lie in Nef(H
r
).
P
D
=
P
D
=
nef cone
P
D
=
Figure 16. Polytopes PD associated to divisors D in nef cone of Hr
Figure 16 shows that when D is in the interior of the nef cone, P
D
is a polygon
whose normal fan is the fan of H
r
. On the boundary of the nef cone, however,
things are different: P
D
is a triangle on the vertical ray and and a line segment on
the horizontal ray. This follows from Figures 10 and 11 in Example 6.2.6.
The Simplicial Case. When X

is complete and simplicial, a result to be proved


in 6.4 gives the following criterion for X

to be projective.
Proposition 6.3.24. A complete simplicial toric variety X

is projective if and only


if its nef cone Nef(X

) Pic(X

)
R
has full dimension in Pic(X

)
R
.
Proof. One direction is an immediate consequence of Theorem 6.3.22. For the
converse, suppose that Nef(X

) has full dimension. Then we can nd a Cartier


divisor D whose class lies in the interior of Nef(X

). Since Nef(X

) =NE(X

, it
follows that DC > 0 for every curve C whose class in NE(X

) is nonzero. Hence,
if we can show that the torus-invariant curves V(), a wall, give nonzero
classes in NE(X

), then Theorem 6.3.13 will imply that D is ample, proving that


X

is projective.
6.3. The Nef and Mori Cones 297
A wall is a facet of some maximal cone , and since is simplicial,
there is (1) such that (1) = (1) . Then Lemma 6.4.2 implies that
D

V() > 0. Hence the class of V() in NE(X

) is nonzero.
Here is a nice application of this result.
Proposition 6.3.25. A complete toric surface X

is projective.
Proof. Picking a basis of N, we may assume N = Z
2
and N
R
= R
2
. Let (1) =

1
, . . . ,
r
and pick
i

i
with [[
i
[[ = 1. Then dene : R
2
R such that
is linear on the cones of and satises (
i
) =1 for all i. The tent analogy (see
Figure 5 in Example 6.1.3) shows that is strictly convex with respect .
Let D =

r
i=1
(u
i
)D
i
=

r
i=1
[[u
i
[[D
i
, where u
i
Z
2
is the minimal gen-
erator of
i
. Note that [D] Pic(X

)
R
since is simplicial. Strict convexity and
the proof of Theorem 6.3.12 imply that D C > 0 for every torus-invariant curve
C X

, so that [D] Nef(X

). The strict inequalities show that [D] is an interior


point, so that X

is projective by Proposition 6.3.24.


When X

is not simplicial, the criterion given in Proposition 6.3.24 can fail.


Here is an example due to Fujino [100].
Example 6.3.26. Consider the complete fan in R
3
with six minimal generators
u
1
= (1, 0, 1), u
2
= (0, 1, 1), u
3
= (1, 1, 1)
u
4
= (1, 0, 1), u
5
= (0, 1, 1), u
6
= (1, 1, 1)
and six maximal cones
Cone(u
1
, u
2
, u
3
), Cone(u
1
, u
2
, u
4
), Cone(u
2
, u
4
, u
5
)
Cone(u
1
, u
3
, u
4
, u
6
), Cone(u
2
, u
3
, u
5
, u
6
), Cone(u
4
, u
5
, u
6
).
You will draw a picture of this fan in Exercise 6.3.5 and show that the resulting
complete toric variety satises
Pic(X

) a(D
1
+D
4
) [ a 3Z Z.
The cones = Cone(u
1
, u
2
, u
4
) and

= Cone(u
2
, u
4
, u
5
) meet along the wall
=

= Cone(u
2
, u
4
).
However, any Cartier divisor D =

6
i=1
a
i
D
i
satises m

= m

(Exercise 6.3.5),
so that the irreducible complete curve C =V() satises
D C = 0
by Proposition 6.3.8. This holds for all Cartier divisors on X

, so C 0. Then X

has no ample divisors by the toric Kleiman criterion, so that X

is nonprojective.
By Exercise 6.3.5, the nef cone of X

is the half-line
Nef(X

) =a[D
1
+D
4
] [ a 0.
298 Chapter 6. Line Bundles on Toric Varieties
This has full dimension in Pic(X

)
R
, yet X

is not projective. Note also that the


Cartier divisor D = 3(D
1
+D
4
) gives a class in the interior of the nef cone, yet D
is not ample. Hence part (b) of Theorem 6.3.22 also fails for X

. The failure is due


to the existence of irreducible curves in X

that are numerically equivalent to zero.


This shows that numerical equivalence of curves can be badly behaved in complete
toric varieties that are neither projective nor simplicial.
Exercises for 6.3.
6.3.1. Let X be a normal variety. Prove that (6.3.1) gives a well-dened pairing between
Q-Cartier divisors and irreducible complete curves. Also show that this pairing satises
Propostion 6.3.7.
6.3.2. Derive the formulas for m

and m

given in Example 6.3.9.


6.3.3. Prove Theorem 6.3.13.
6.3.4. Prove that on a complete normal variety, an ample divisor D satises D C > 0 for
all irreducible curves C X.
6.3.5. Consider the fan from Example 6.3.26.
(a) Draw a picture of this fan in R
3
.
(b) Prove that Pic(X

) a(D
1
+D
4
) [ a 3Z.
(c) Prove that 3(D
1
+D
4
) is nef.
6.4. The Simplicial Case
Here we assume that is a simplicial fan in N
R
R
n
. Then Proposition 4.2.7
implies that every Weil divisor is Q-Cartier. Since we will be working in Pic(X

)
R
,
it follows that we can drop the adjective Cartier when discussing divisors.
Relations Among Minimal Generators. We begin our discussion of the simplicial
case with another way to think of elements of N
1
(X

). Recall from Theorem 4.1.3


that we have an exact sequence
(6.4.1) M

Z
(1)

Cl(X

) 0
where (m) = ('m, u

`)
(1)
and sends the standard basis element e

Z
(1)
to [D

] Cl(X

).
Proposition 6.4.1. Let be a simplicial fan in N
R
with convex support of full
dimension. Then there are dual exact sequences
0 M
R

R
(1)

Pic(X

)
R
0
and
0 N
1
(X

R
(1)

N
R
0,
6.4. The Simplicial Case 299
where

(e

) = u

, e

a standard basis vector of R


(1)

([C]) = (D

C)
(1)
, C X

an irreducible complete curve.


Thus N
1
(X

) can be interpreted as the space of linear relations among the minimal


generators of . Furthermore, given D =

and a relation

= 0,
the intersection pairing of [D] Pic(X

)
R
and R = (b

)
(1)
N
1
(X

) is
[D] R =

.
Proof. Since is simplicial, all Weil divisors are Q-Cartier. Hence
Pic(X

)
R
= Pic(X

)
Z
R= Cl(X

)
Z
R.
Tensoring with R preserves exactness, so exactness of the rst sequence follows
from (6.4.1). Note also that Pic(X

)
R
= N
1
(X

) by (6.3.2).
The dual of an exact sequence of nite-dimensional vector spaces is still exact.
Then the perfect pairings
M
R
N
R
R : (m, u) 'm, u`
Pic(X

)
R
N
1
(X

) R : ([D], [C]) D C
easily imply that for m M
R
and [C] N
1
(X

), we have
(m) = ('m, u

`)
(1)
=

(e

) = u

and
(e

) = [D

] =

([C]) = (D

C)
(1)
.
Finally, the duality between the two exact sequences reduces to dot product on the
middle terms R
(1)
. This proves the nal assertion of the proposition.
The map

: N
1
(X

) R
(1)
in Proposition 6.4.1 implies that an irreducible
complete curve C X

gives the relation

(D

C)u

= 0 in N
R
.
This can be proved directly as follows. First observe that m M gives

'm, u

`D

= div(
m
) 0.
Taking the intersection product with C, we see that

'm, u

`(D

C) = 0
holds for all m M
R
. Writing this as

m,

(D

C)u

= 0, we obtain
(6.4.2)

(D

C)u

= 0 in N
R
.
This argument shows that (6.4.2) holds for any simplicial toric variety.
300 Chapter 6. Line Bundles on Toric Varieties
Intersection Products. Our next task is to compute D

C when C is a torus-
invariant complete curve in X

. This means C = V(), where (n 1) is a


wall, i.e., the intersection of two cones in (n). Here, we only assume that is a
simplicial fan in N
R
R
n
, with no hypotheses on its support.
We begin with an easy case. Fix a wall
=

.
Since is simplicial, we can label the minimal generators of so that
= Cone(u

1
, u

2
, . . . , u
n
)
= Cone(u

2
, . . . , u
n
).
Thus is the facet of opposite to
1
. We will compute the intersection prod-
uct D

1
V() in terms of the multiplicity (also called the index) of a simplicial
cone. This is dened as follows. If is a simplicial cone with minimal generators
u
1
, . . . , u
k
, then mult() is the index of the sublattice
Zu
1
+ +Zu
k
N

= Span() N = (Ru
1
+ +Ru
k
) N.
Lemma 6.4.2. If , and
1
are as above, then
D

1
V() =
mult()
mult()
.
Proof. Since u

1
, . . . , u
n
is a basis of N
Q
, we can nd m M
Q
such that
'm, u

i
` =

1 i = 1
0 i = 2, . . . , n.
Pick a positive integer such that m M. On U

, D

1
is the Cartier divisor
determined by m

= m and m

= 0. By (6.3.1) and Proposition 6.3.8,


D

1
V() =
1

(D

1
) V() =
1

'm, u` ='m, u`,


where u

maps to a generator of

N(). Recall that N() = N/N

.
When we combine u with a basis of N

, we get a basis of N. Thus there is a


positive integer such that u

1
=u+v, v N

. The minus sign appears because


u and u

1
lie on opposite sides of . By considering the sublattices
Zu

1
+Zu

2
+ +Zu
n
Zu

1
+N

Zu+N

= N,
one sees that = mult()/mult(). Thus
u =
1

1
v

=
mult()
mult()

1
v

.
Since m

, it follows that
D

1
V() ='m, u` =
mult()
mult()
'm, u

1
` =
mult()
mult()
.
6.4. The Simplicial Case 301
Corollary 6.4.3. Let be a simplicial fan in N
R
R
n
and (n1) be a wall.
If (1) and generate a smooth cone of (n), then
D

V() = 1
Given a wall (n1), our next task is to compute D

V() for the other


rays (1). Let =

and write
(6.4.3)
= Cone(u

1
, . . . , u
n
)

= Cone(u

2
, . . . , u

n+1
)
= Cone(u

2
, . . . , u
n
).
This situation is pictured in Figure 17.
u

n
u

n+1
u

1
u

Figure 17. =

The n+1 minimal generators u

1
, . . . , u

n+1
are linearly dependent. Hence they
satisfy a linear relation, which we write as
(6.4.4) u

1
+
n

i=2
b
i
u

i
+u

n+1
= 0.
We may assume , > 0 since u

1
and u

n+1
lie on opposite sides of the wall .
Then (6.4.4) is unique up to multiplication by a positive constant since u

1
, . . . , u
n
are linearly independent. We call (6.4.4) a wall relation.
On the other hand, setting C =V() in (6.4.2) gives the linear relation
(6.4.5)

(D

V())u

= 0
302 Chapter 6. Line Bundles on Toric Varieties
As we now prove, the two relations are the same up to a positive constant.
Proposition 6.4.4. The relations (6.4.4) and (6.4.5) are equal after multiplication
by a positive constant. Furthermore:
(a) D

V() = 0 for all /


1
, . . . ,
n+1
.
(b) D

1
V() =
mult()
mult()
and D

n+1
V() =
mult()
mult(

)
.
(c) D

i
V() =
b
i
mult()
mult()
=
b
i
mult()
mult(

)
for i = 2, . . . , n.
Proof. Part (b) follows immediately from Lemma 6.4.2. Also observe that when
/
1
, . . . ,
n+1
, and never lie in the same cone of , so D

V() = by
the Orbit-Cone Correspondence. This easily implies D

V() =0 (Exercise 6.4.1).


This proves part (a) and implies that (6.4.5) reduces to the equation
(D

1
V())u

1
+
n

i=2
(D

i
V())u

i
+ (D

n+1
V())u

n+1
= 0.
The coefcients of u

1
and u

n+1
are positive by part (b), so up to a positive con-
stant, this must be the wall relation (6.4.4). The rst assertion of the lemma follows.
Since the above relation equals (6.4.4) up to a nonzero constant, we obtain
b
i
(D

1
V()) = (D

i
V()), b
i
(D

n+1
V()) = (D

i
V()),
for i =2, . . . , n. Then the formulas for D

i
V() in part (c) follow from part (b).
For a simplicial toric variety, Proposition 6.4.4 provides everything we need to
compute D V() when is a wall of .
Example 6.4.5. Consider the fan in R
2
from Example 6.3.9. We have the wall
= Cone(u
0
) =

= Cone(u
1
, u
0
) Cone(u
2
, u
0
),
where u
1
= e
1
, u
2
= e
2
and u
0
= 2e
1
+3e
2
. Computing multiplicities gives
mult() = 1, mult() = 3, mult(

) = 2.
Then part (b) of Proposition 6.4.4 implies
D
1
V() =
1
3
, D
2
V() =
1
2
,
and the relation
2 u
1
+(1) u
0
+3 u
2
= 0
implies
D
0
V() =
1 1
2 3
=
1 1
3 2
=
1
6
by part (c) of the proposition. Hence we recover the intersection products computed
in Example 6.3.9.
6.4. The Simplicial Case 303
When X

is smooth, all multiplicities are 1. Hence the wall relation (6.4.4) can
be written uniquely as
(6.4.6) u

1
+
n

i=2
b
i
u

i
+u

n+1
= 0, b
i
Z,
and then the intersection formulas of Proposition 6.4.4 reduce to
(6.4.7) D

1
V() = D

n+1
V() = 1, D

i
V() = b
i
, i = 2, . . . , n.
Example 6.4.6. For the Hirzebruch surface H
r
, the four curves coming from walls
are also divisors. Recall that the minimal generators are
u
1
=e
1
+re
2
, u
2
= e
2
, u
3
= e
1
, u
4
=e
2
,
arranged clockwise around the origin (see Figure 3 from Example 6.1.2 for the case
r = 2). Hence the wall generated by u
1
gives the relation
u
2
0 u
3
+u
4
= 0,
which implies
D
1
D
1
= 0
by (6.4.7). On the other hand, the wall generated by u
2
gives the relation
u
1
r u
2
+u
3
= 0.
Then (6.4.7) implies
D
2
D
2
=r.
Similarly, one can check that
D
3
D
3
= 0, D
4
D
4
= r,
and by Corollary 6.4.3 or (6.4.7) we also have
D
1
D
2
= D
2
D
3
= D
3
D
4
= D
4
D
1
= 1.
These computations give an explicit description of the duality between the nef and
Mori cones shown in Figure 15 of Example 6.3.23 (Exercise 6.4.2).
In general, a Q-Cartier divisor D on a complete surface has self-intersection
D D = D
2
. Self-intersections will play a prominent role in Chapter 10 when we
study toric surfaces.
Example 6.4.7. Let X

be a complete toric surface. Write (1) =


1
, . . . ,
r
,
where the
i
are arranged clockwise around the origin in N
R
R
2
. Each
i
gives a
minimal generator u
i
and a toric divisor D
i
. Note also that Pic(X

)
R
R
r2
.
Proposition 6.3.25 tells us that X

is projective, so that its Mori cone NE(X

)
is strongly convex of dimension r 2. Hence a minimal generating set has at
least r 2 elements. Since the r classes [D
i
] = [V(
i
)] generate by the toric cone
theorem, we almost know the minimal generators.
304 Chapter 6. Line Bundles on Toric Varieties
Now suppose that X

is smooth. Then the wall relation for D


i
= V(
i
) is
u
i1
+b
i
u
i
+u
i+1
= 0 by (6.4.6), where b
i
= D
2
i
by (6.4.7). Given a divisor D =

r
i=1
a
i
D
i
, Proposition 6.4.1 implies that
D D
i
= a
i1
+b
i
a
i
+a
i+1
,
so that D is nef (resp. ample) if and only if
a
i1
+b
i
a
i
+a
i+1
0 (resp. > 0)
for i = 1, . . . , r. This makes it easy to study nef and ample divisors on X

.
Primitive Collections. In the projective case, there is a beautiful criterion for a
Cartier divisor to be nef or ample in terms of the primitive collections introduced
in Denition 5.1.5. Recall that
P =
1
, . . . ,
k
(1)
is a primitive collection if P is not contained in (1) for all but any proper
subset is. Since is simplicial, primitive means that P does not generate a cone of
but every proper subset does. This is the denition given by Batyrev in [14].
Example 6.4.8. Consider the complete fan in R
3
shown in Figure 18. One can
check that

1
,
3
,
0
,
2
,
4

are the only primitive collections of .


z
y
x

2
Figure 18. A complete fan in R
3
with two primitive collections
Here is the promised characterization, due to Batyrev [14] in the smooth case.
6.4. The Simplicial Case 305
Theorem 6.4.9. Let X

be a projective simplicial toric variety. Then:


(a) A Cartier divisor D is nef if and only if its support function
D
satises

D
(u

1
+ +u

k
)
D
(u

1
) + +
D
(u

k
)
for all primitive collections P =
1
, . . . ,
k
of .
(b) A Cartier divisor D is ample if and only if its support function
D
satises

D
(u

1
+ +u

k
) >
D
(u

1
) + +
D
(u

k
)
for all primitive collections P =
1
, . . . ,
k
of .
Before we discuss the proof of Theorem 6.4.9, we need to study the relations
that come from primitive collections.
Denition 6.4.10. Let P =
1
, . . . ,
k
(1) be a primitive collection for the
complete simplicial fan . Hence

k
i=1
u

i
lies in the relative interior of a cone
. Thus there is a unique expression
u

1
+ +u

k
=

(1)
c

, c

Q
>0
.
Then u

1
+ +u

(1)
c

= 0 is the primitive relation of P.


The coefcient vector of this relation is r(P) = (b

)
(1)
R
(1)
, where
(6.4.8) b

1 P, / (1)
1c

P(1)
c

(1), / P
0 otherwise.
Then

= 0, so that r(P) gives an element of N


1
(X

) by Proposition 6.4.1.
In Exercise 6.4.3, you will prove that c

< 1 when P(1). This means that P


is determined by the positive entries in the coefcient vector r(P).
The Mori cone for X

has a nice description in terms of primitive relations.


Theorem 6.4.11. If X

is a projective simplicial toric variety, then


NE(X

) =

P
R
0
r(P),
where the sum is over all primitive collections P of .
Proof. Given a Cartier divisor D =

, Proposition 6.4.1 shows that


[D] r(P) =

=
k

i=1
a

(1)
a

.
306 Chapter 6. Line Bundles on Toric Varieties
Since the support function of D satises
D
(u

) = a

and is linear on , we can


rewrite this as
(6.4.9) [D] r(P) =
D
(u

1
)
D
(u

k
) +
D
(u

1
+ +u

k
).
If D is nef, then it is basepoint free (Theorem 6.3.12), so
D
is convex. It follows
that [D] r(P) 0, which proves r(P) NE(X

). Note also that r(P) is nonzero.


To nish the proof, we need to show that NE(X

) is generated by the r(P).


In the discussion following the proof of Theorem 6.3.22, we noted that NE(X

)
is generated by its extremal rays, each of which is of the form R
0
[V()] for an
extremal wall . It sufces to show that [V()] is a positive multiple of r(P) for
some primitive collection P.
We rst make a useful observation about nef divisors. Given a cone , we
claim that any nef divisor is linearly equivalent to a divisor of the form
(6.4.10) D =

, a

= 0, (1) and a

0, / (1).
To prove this, rst recall that any nef divisor is linearly equivalent to a torus-
invariant nef divisor D =

. Then we have m

M with 'm

, u

` = a

for (1). Since D is nef, it is also basepoint free, so that


'm

, u

`
D
(u

) =a

, (1),
by Theorem 6.1.7. Replacing D with D+div(
m
), we obtain (6.4.10).
Now assume we have an extremal wall and let C =V(). Consider the set
P = [ D

C > 0.
We will prove that P is a primitive collection whose primitive relation is the class
of C, up to a positive constant. Our argument is taken from [71], which is based on
ideas of Kresch [181].
We rst prove by contradiction that P (1) for all . Suppose P (1)
and take an ample divisor D (remember that X

is projective). Then in particular D


is nef, so we may assume that D is of the form (6.4.10). Since a

= 0 for (1),
we have
D C =

/ (1)
a

C.
However, a

0 by (6.4.10), and P (1) implies D

C 0 for / (1). It
follows that D C 0, which is impossible since D is ample. Thus no cone of
contains all rays in P.
It follows that some subset Q P is a primitive collection. This gives the
primitive relation with coefcient vector r(Q) N
1
(X

), and we also have the


class [C] N
1
(X

). Let
= [C] r(Q) N
1
(X

),
where > 0. We claim that if is sufciently small, then
(6.4.11) [ [D

] < 0 [ D

C < 0.
6.4. The Simplicial Case 307
To prove this, rst observe that the denition of implies
D

C = [D

] r(Q) +[D

] .
Suppose [D

] < 0 and D

C 0. This forces [D

] r(Q) > 0. Proposition 6.4.1


implies that [D

] r(Q) is the coefcent of u

in the primitive relation of Q, which


by (6.4.8) is positive only when Q. Then Q P implies D

C > 0 by the
denition of P. But we can clearly choose sufciently small so that
D

C > [D

] r(Q) whenever D

C > 0.
This inequality and the above equation imply [D

] >0, which is a contradiction.


We next claim that NE(X

). By (6.4.11), we have
[ [D

] < 0 [ D

C < 0 (1),
where the second inclusion follows from C =V() and Proposition 6.4.4. Now let
D be nef, and by (6.4.10) with = , we may assume that
D =

, a

= 0, (1) and a

0, / (1).
Then
[D] =

/ (1)
a

[D

] 0,
where the nal inequality follows since a

0 and [D

] < 0 can happen only


when (1). This proves that NE(X

).
We showed earlier that r(Q) NE(X

). Thus the equation


[C] = r(Q) +
expresses [C] as a sum of elements of NE(X

). But [C] is extremal, i.e., it lies in


a 1-dimensional face of NE(X

). By Lemma 1.2.7, this forces r(Q) and to lie


in the face. Since r(Q) is nonzero, it generates the face, so that [C] is a positive
multiple of r(Q).
The relation corresponding to C has coefcients (D

C)
(1)
, and P is the
set of s where D

C > 0. But this relation is a positive multiple of r(Q), whose


positive entries correspond to Q. Thus P = Q and the proof is complete.
It is now straightforward to prove Theorem 6.4.9 using Theorem 6.4.11 and
(6.4.9) (Exercise 6.4.4). We should also mention that these results hold more gen-
erally for any projective toric variety (see [71]).
Example 6.4.12. Let be the fan shown in Figure 18 of Example 6.4.8. The
minimal generators of
0
, . . . ,
4
are
u
0
= (0, 0, 1), u
1
= (0, 1, 1), u
2
= (1, 0, 1), u
3
= (0, 1, 1), u
4
= (1, 0, 1).
308 Chapter 6. Line Bundles on Toric Varieties
The computations you did for part (c) of Exercise 6.1.8 imply that X

is pro-
jective. Since the only primitive collections are
1
,
3
and
0
,
2
,
4
, Theo-
rem 6.4.9 implies that a Cartier divisor D is nef if and only if

D
(u
1
+u
3
)
D
(u
1
) +
D
(u
3
)

D
(u
0
+u
2
+u
4
)
D
(u
0
) +
D
(u
2
) +
D
(u
4
)
and ample if and only if these inequalities are strict. One can also check that
Pic(X

) a[D
1
] +b[D
2
] [ a, b 2Z
and aD
1
+bD
2
is nef (resp. ample) if and only if a b 0 (resp. a > b > 0).
Exercise 6.4.5 will relate this example to the proof of Theorem 6.4.11.
Besides , the minimal generators u
0
, . . . , u
4
support two other complete fans
in R
3
: rst, the fan

obtained by replacing Cone(u


2
, u
3
) with Cone(u
1
, u
3
) in
Figure 18, and second, the fan
0
obtained by removing this wall to create the cone
Cone(u
1
, u
2
, u
3
, u
4
). Since (1) =

(1) =
0
(1), the toric varieties X

, X

, X

0
have the same class group, though X

0
has strictly smaller Picard group since it is
not simplicial. Thus
Pic(X

0
)
R
Pic(X

)
R
= Pic(X

)
R
R
2
.
This allows us to draw all three nef cones in the same copy of R
2
. In Exercise 6.4.5
you show that we get the cones shown in Figure 19. The ideas behind this gure
Nef(X

)
Nef(X
0
)
Nef(X

)
Figure 19. The nef cones of X, X

, X
0
will be developed in Chapters 14 and 15 when we study geometric invariant theory
and the minimal model program for toric varieties.
Exercises for 6.4.
6.4.1. This exercise will describe a situation where D C is guaranteed to be zero.
(a) Let X be normal and assume that C is a complete irreducible curve disjoint from the
support of a Cartier divisor D. Prove that D C = 0. Hint: Use U = X ` Supp(D).
Appendix: Quasicoherent Sheaves on Toric Varieties 309
(b) Let be a wall of a fan and pick (1) such that and do not lie in the same
cone of . Use the Cone-Orbit Correspondence to prove that D

V() = , and
conclude that D

V() = 0.
6.4.2. As in Example 6.4.2, the classes [D
3
], [D
4
] give a basis of Pic(H
r
)
R
. Since H
r
is a
smooth complete surface, the intersection product D
i
V(
j
) is written D
i
D
j
.
(a) Give an explicit formula for (a[D
3
] +b[D
4
]) (a[D
3
] +b[D
4
]) using the computations
of Example 6.4.2.
(b) Use part (a) to show that the cones shown in Figure 15 in Example 6.3.23 are dual to
each other.
6.4.3. In the primitive relation dened in Denition 6.4.10, prove c

< 1 when P
(1). Hint: If
1
(1) and c
1
1, then cancel u
1
and show that u
2
, . . . , u

k
.
6.4.4. Prove Theorem 6.4.9 using Theorem 6.4.11 and (6.4.9).
6.4.5. Consider the fan from Examples 6.4.8 and 6.4.12. Every wall of is of the
form
i j
= Cone(u
i
, u
j
) for suitable i < j. Let r(
i j
) R
5
denote the wall relation of
i j
.
Normalize by a positive constant so that the entries of r(
i j
) are integers with gcd = 1.
(a) Show the nine walls give the three distinct wall relations r(
02
), r(
03
), r(
24
).
(b) Show that r(
03
) +r(
24
) = r(
02
) and conclude that
03
and
24
are extremal walls
whose classes generate the Mori cone of X

.
(c) For each extremal wall of part (b), determine the corresponding primitive collection.
You should be able to read the primitive collection directly from the wall relation.
(d) Show that the nef cones of X

, X

, X
0
give the cones shown in Figure 19.
6.4.6. Let X

be the blowup of P
n
at a xed point of the torus action. Thus Pic(X

) Z
2
.
(a) Compute the nef and Mori cones of X

and draw pictures similar to Figure 15 in


Example 6.3.23.
(b) Determine the primitive relations and extremal walls of X

.
6.4.7. Let {
r
be the toric surface obtained by changing the ray u
1
in the fan of the Hirze-
bruch surface H
r
from (1, r) to (r, 1). Assume r > 1.
(a) Prove that {
r
is singular. How many singular points are there?
(b) Determine which divisors a
1
D
1
+a
2
D
2
+a
3
D
3
+a
4
D
4
are Cartier and compute D
i
D
j
for all i, j.
(c) Determine the primitive relations and extremal walls of {
r
.
Appendix: Quasicoherent Sheaves on Toric Varieties
Now that we know more about sheaves (specically, tensor products and exactness), we
can complete the discussion of quasicoherent sheaves on toric varieties begun in 5.3. In
this appendix, X

will denote a toric variety with no torus factors. The total coordinate
ring of X

is S =C[x

[ (1)], which is graded by Cl(X

).
Recall from 5.3 that for Cl(X

), the shifted S-module S() gives the sheaf


O
X
() satisfying O
X
() O
X
(D) for any Weil divisor with = [D]. In 6.0 we con-
structed a sheaf homomorphism O
X
(D)
OX
O
X
(E) O
X
(D+E). In a similar way, one
310 Chapter 6. Line Bundles on Toric Varieties
can dene
(6.A.1) O
X
()
OX

O
X
() O
X
(+).
for , Cl(X

) such that if = [D] and = [E], then the diagram


O
X
(D))
OX

O
X
(E))

O
X
(D+E))

O
X
()
OX

O
X
()

O
X
(+)
commutes, where the vertical maps are isomorphisms.
From Sheaves to Modules. The main construction of 5.3 takes a graded S-module Mand
produces a quasicoherent sheaf

Mon X

. We now go in the reverse direction and show that


every quasicoherent sheaf on X

arises in this way. We will use the following construction.


Denition 6.A.1. For a sheaf F of O
X
-modules on X

and Cl(X

), dene
F() =F
OX

O
X
()
and then set

(F) =

Cl(X)
(X

, F()).
For example,

(O
X
) = S since (X

, O
X
()) S

by Proposition 5.3.7. Using


this and (6.A.1), we see that

(F) is a graded S-module.


We want to show that F is isomorphic to the sheaf associated to

(F) when F is
quasicoherent. We will need the following lemma due to Mustat a [212]. Recall that for
, we have the monomial x

=

/ (1)
x

S. Let

= deg(x

) Cl(X

).
Lemma 6.A.2. Let F be a quasicoherent sheaf on X

. Then:
(a) If v (U

, F), then there are 0 and u (X

, F(

)) such that u restricts to


(x

)

v (U

, F(

)).
(b) If u (X

, F) restricts to 0 in (U

, F), then there is 0 such that (x



)

u = 0 in
(X

, F(

)).
Proof. For part (a), x and take v (U

, F). Given , let v

be the restriction
of v to U

. By (3.1.3), we can nd m(

M such that U

=(U

m =
Spec(C[

M]

m ). In terms of the total coordinate ring S, we have C[

M] (S
x
)
0
by (5.3.1). Hence the coordinate ring of U

is the localization

(S
x
)
0

x
m
,
where x
m
=

x
m,u

(S
x
)
0
since m

M. This enables us to write


U

= (U

)
x
m .
Since F is quasicoherent, F[
U
is determined by its sections G = (U

, F), and then


(U

, F) is the localization G
x
m .
It follows that v

(U

, F) equals u

/(x
m
)
k
, where k 0 and u

(U

, F).
Hence u

restricts to (x
m
)
k
v (U

, F). Since m (

), we see that
(6.A.2) x
a
= (x

)

(x
m
)
k
S
Appendix: Quasicoherent Sheaves on Toric Varieties 311
for 0. This monomial has degree

. Then u

= x
a
u

(U

, F(

)) restricts
to (x

)

(U

, F(

)). By making sufciently large, we can nd one that


works for all .
To study whether the u

patch to give a global section of F(

), take
1
,
2
and
set =
1

2
. Thus U

=U
1
U
2
, and
(6.A.3) w = u
1
[
U
u
2
[
U
(U

, F(

))
restricts to 0 (U

, F(

)). Arguing as above, this group of sections is the lo-


calization (U

, F(

))
x
m , where m

) M such that U

= (U

m .
Since w gives the zero element in this localization, there is k 0 with (x
m
)
k
w = 0 in
(U

, F(

)). If we multiply by x
b
= (x

)

(x
m
)
k
for

0, we obtain (x

)

w = 0
in (U

, F((

+)

)). Another way to think of this is that if we made in (6.A.2) big


enough to begin with, then in fact w = 0 in (U

, F(

)) for all ,

. Given the deni-


tion (6.A.3) of w, it follows that the u

patch to give a global section u (X

, F(

))
with the desired properties.
The proof of part (b) is similar and is left to the reader.
Proposition 6.A.3. Let F be a quasicoherent sheaf on X

. Then F is isomorphic to the


sheaf associated to the graded S-module

(F).
Proof. Let M =

(F) and recall from 5.3 that for every , the restriction of

M to
U

is the sheaf associated to the (S


x
)
0
-module (M
x
)
0
.
We rst construct a sheaf homomorphism

MF. Elements of (M
x
)
0
are u/(x

)

for u (X

, F(

)). Since (x

)

is a section of O
X
(

) over U

, the map
(U

, O
X
(

))
C
(U

, F(

)) (U

, F)
induces a homomorphismof (S
x
)
0
-modules
(6.A.4) (M
x
)
0
(U

, F).
This gives compatible sheaf homomorphisms

M[
U
F[
U
that patch to give

MF.
Since F is quasicoherent, it sufces to show that (6.A.4) is an isomorphism for every
. First suppose that u/(x

)
k
(M
x
)
0
maps to 0 (U

, F). It follows easily that


u restricts to zero in (U

, F(k

)). By Lemma 6.A.2 applied to F(k

), there is 0
such that (x

)

u = 0 in (X

, F(( +k)

)). Then
u
(x

)
k
=
(x

)

u
(x

)
+k
= 0 in (M
x
)
0
,
which shows that (6.A.4) is injective. To prove surjectivity, take v (U

, F) and apply
Lemma 6.A.2 to nd 0 and u (X

, F(

)) such that u restricts to (x



)

v. It
follows immediately that u/(x

)

(M
x
)
0
maps to v.
This result proves part (a) of Proposition 5.3.9. We now turn our attention to part (b)
of the proposition, which applies to coherent sheaves.
Proposition 6.A.4. Every coherent sheaf F on X

is isomorphic to the sheaf associated


to a nitely generated graded S-module.
312 Chapter 6. Line Bundles on Toric Varieties
Proof. On the afne open subset U

, we can nd nitely many sections f


i,
(U

, F)
which generate F over U

. By Lemma 6.A.2, we can nd 0 such that (x



)

f
i,
comes
from a global section g
i,
of F(

). Now consider the graded S-module M

(F)
generated by the g
i,
. Proposition 6.A.3 gives an isomorphism

(F) F.
Hence M

(F) gives a sheaf homomorphism



MF which is injective by the exact-
ness proved in Example 6.0.10. Over U

, we have f
i,
= g
i,
/(x

)

(M
x
)
0
, and since
these sections generate F over U

, it follows that

M F. Then we are done since M is
clearly nitely generated.
The proof of Proposition 6.A.4 used a submodule of

(F) because the full module


need not be nitely generated when F is coherent. Here is an easy example.
Example 6.A.5. A point p P
n
gives a subvariety i : p P
n
. The sheaf F = i

O
{p}
can be thought of as a copy of C sitting over the point p. The line bundle O
P
n (a) is free in
a neighborhood of p, so that F(a) F for all a Z. Thus

(F) =

aZ
(P
n
, F(a)) =

aZ
C.
This module is not nitely generated over S since it is nonzero in all negative degrees.
Subschemes and Homogeneous Ideals. For readers who know about schemes, we can
apply the above results to describe subschemes of a toric variety X

with no torus factors.


Let I S be a homogeneous ideal. By Proposition 6.0.10, this gives a sheaf of ideals
I O
X
whose quotient is the structure sheaf O
Y
of a closed subscheme Y X

. This
differs from the subvarieties considered in the rest of the book since the structure sheaf O
Y
may have nilpotents.
Proposition 6.A.6. Every subscheme Y X

is dened by a homogeneous ideal I S.


Proof. Given an ideal sheaf I O
X
, we get a homomorphismof S-modules

(I)

(O
X
) = S.
If I S is the image of this map, then the map factors

(I) I S, where the rst


arrow is surjective and the second injective. By Example 6.0.10 and Proposition 6.A.3, the
inclusion I O
X
factors as I

I O
X
. It follows immediately that I =

I.
In the case of P
n
, it is well-known that different graded ideals can give the same ideal
sheaf. The same happens in the toric situation, and as in 5.3, we get the best answer in
the smooth case. Not surprisingly, the irrelevant ideal B() S plays a key role in the
following result from [65, Cor. 3.8].
Proposition 6.A.7. Homogeneous ideals I, J S in the total coordinate of a smooth toric
variety X

give the same ideal sheaf of O


X
if and only if I : B()

= J : B()

.
There is a less elegant version of this result that applies to simplicial toric varieties. See
[65] for a proof and more details about the relation between graded modules and sheaves.
See also [204] for more on multigraded commutative algebra.
Chapter 7
Projective Toric
Morphisms
7.0. Background: Quasiprojective Varieties and Projective
Morphisms
Many results of Chapter 6 can be generalized, but in order to do so, we need to
learn about quasiprojective varieties and projective morphisms.
Quasiprojective Varieties. Besides afne and projective varieties, we also have the
following important class of varieties.
Denition 7.0.1. A variety is quasiprojective if it is isomorphic to an open subset
of a projective variety.
Here are some easy properties of quasiprojective varieties.
Proposition 7.0.2.
(a) Afne varieties and projective varieties are quasiprojective.
(b) Every closed subvariety of a quasiprojective variety is quasiprojective.
(c) A product of quasiprojective varieties is quasiprojective.
Proof. You will prove this in Exercise 7.0.1.
Projective Morphisms. In algebraic geometry, concepts that apply to varieties
sometimes have relative versions that apply to morphisms between varieties. For
example, in 3.4, we dened completeness and properness, where the former ap-
plies to varieties and the latter applies to morphisms. Sometimes we say that the
313
314 Chapter 7. Projective Toric Morphisms
relative version of a complete variety is a proper morphism. In the same way, the
relative version of a projective variety is a projective morphism.
We begin with a special case. Let f : X Y be a morphism and L a line
bundle on X with a basepoint free nite-dimensional subspace W (X, L). Then
combining f : X Y with the morphism
L,W
: X P(W

) from 6.0 gives a


morphism X Y P(W

) that ts into a commutative diagram


(7.0.1)
X
f
L,W

Y P(W

)
p
1

Y.
If f
L,W
is a closed embedding (meaning that its image Z Y P(W

) is
closed and the induced map X Z is an isomorphism), then you will show in
Exercise 7.0.2 that f has the following nice properties:
f is proper.
For every p Y, the ber f
1
(p) is isomorphic to a closed subvariety of
P(W

) and hence is projective.


The general denition of projective morphism must include this special case.
In fact, going from the special case to the general case is not that hard.
Denition 7.0.3. A morphism f : X Y is projective if there is a line bundle L
on X and an afne open cover U
i
of Y with the property that for each i, there is
a basepoint free nite-dimensional subspace W
i
( f
1
(U
i
), L) such that
f
1
(U
i
)
f
i

L
i
,W
i
U
i
P(W

i
)
is a closed embedding, where f
i
= f [
f
1
(U
i
)
and L
i
= L[
f
1
(U
i
)
. We say that
f : X Y is projective with respect to L.
The general case has the properties noted above in the special case.
Proposition 7.0.4. Let f : X Y be projective. Then:
(a) f is proper.
(b) For every p Y, the ber f
1
(p) is a projective variety.
Here are some further properties.
Proposition 7.0.5.
(a) The composition of projective morphisms is projective.
(b) A closed embedding is a projective morphism.
(c) A variety X is projective if and only if X pt is a projective morphism.
7.0. Background: Quasiprojective Varieties and Projective Morphisms 315
Proof. Parts (a) and (b) are proved in [127, (5.5.5)]. For part (c), one direction
follows immediately from the previous proposition. Conversely, let i : X P
n
be
projective, and assume that X is nondegenerate, meaning that X is not contained in
any hyperplane of P
n
. Now let L =O
X
(1) = i

O
P
n (1). Then
i

: (P
n
, O
P
n (1)) (X, L)
is injective since X is nondegenerate. In Exercise 7.0.3 you will show that
(P
n
, O
P
n (1)) = Span(x
0
, . . . , x
n
)
and that if W (X, L) is the image of i

, then
L,W
is the embedding we began
with. Hence Denition 7.0.3 is satised for X pt and L.
When the domain is quasiprojective, the relation between proper and projective
is especially easy to understand.
Proposition 7.0.6. Let f : X Y be a morphism where X is quasiprojective. Then:
f is proper f is projective.
Proof. One direction is obvious since projective implies proper. For the opposite
direction, X is quasiprojective, which implies that there is a morphism
g : X Z
such that Z is projective, g(X) Z is open, and X g(X) via g. Then one can
prove without difculty that the product map
(7.0.2) f g : X Y Z
induces an isomorphism X ( f g)(X).
Since f : X Y is proper, f g : X Y Z is also proper (Exercise 7.0.4).
Hence the image of f g is closed in X Z since proper morphisms are universally
closed. Thus X ( f g)(X) and ( f g)(X) is closed in Y Z. This proves that
(7.0.2) is a closed embedding.
Now take a closed embedding Z P
s
. Arguing as above, we get a closed
embedding of X into Y P
s
. From here, it is straightforward to show that f is
projective (Exercise 7.0.4).
To complicate matters, there are two denitions of projective morphism used
in the literature. In [131, II.4], a projective morphism is dened as the special
case considered in (7.0.1), while [127, (5.5.2)] and [273, 5.3] give a much more
general denition. Theorem 7.A.4 of the appendix to this chapter shows that the
more general denition is equivalent to Denition 7.0.3.
Projective Bundles. Vector bundles give rise to an interesting class of projective
morphisms.
316 Chapter 7. Projective Toric Morphisms
Let : V X be a vector bundle of rank n 1. Recall from 6.0 that V has
a trivialization (U
i
,
i
) with
i
:
1
(U
i
) U
i
C
n
. Furthermore, the transition
functions g
i j
GL
n
((U
i
U
j
, O
X
)) make the diagram
U
i
U
j
C
n

1
(U
i
U
j
)

i[

1
(U
i
U
j
)

.
.
.
.
.
.
.
.
.

j [

1
(U
i
U
j
)

U
i
U
j
C
n
1g
i j

commute. Note that 1g


i j
induces an isomorphism
1g
i j
: U
i
U
j
P
n1
U
i
U
j
P
n1
.
This gives gluing data for a variety P(V). It is clear that induces a morphism
: P(V) X and that
i
induces the trivialization

i
:
1
(U
i
) U
i
P
n1
.
The discussion following Theorem 7.A.4 in the appendix to this chapter shows that
: P(V) X is a projective morphism. We call P(V) the projective bundle of V.
Example 7.0.7. Let W be a nite-dimensional vector space over C of positive
dimension. Then, for any variety X, the trivial bundle X W X gives the trivial
projective bundle X P(W) X.
There is also a version of this construction for locally free sheaves. If E is
locally free of rank n, then E is the sheaf of sections of a vector bundle V
E
X of
rank n. When n = 1, we proved this in Theorem 6.0.20. Then dene
(7.0.3) P(E ) =P(V

E
),
where V

E
is the dual vector bundle of V
E
. Here are some properties of P(E ).
Lemma 7.0.8.
(a) P(L) = X when L is locally free of rank 1.
(b) P(E
O
X
L) =P(E ) when E is locally free and L is a line bundle.
(c) If a homomorphism E F of locally free sheaves is surjective, then the in-
duced map P(F) P(E ) of projective bundles is injective.
Proof. You will prove this in Exercise 7.0.5. The dual in (7.0.3) is the reason why
E F gives P(F) P(E ).
The appearance of the dual in (7.0.3) can be explained as follows. Let L be a
line bundle with W (X, L) basepoint free of nite dimension. As in 6.0, this
gives a morphism

L,W
: X P(W

).
7.0. Background: Quasiprojective Varieties and Projective Morphisms 317
Let E =W
C
O
X
. The corresponding vector bundle is V
E
= X W, so
(7.0.4) P(E ) =P(V

E
) = X P(W

).
By Proposition 6.0.24, the natural map E L is surjective since W has no base-
points. By Lemma 7.0.8, we get an injection of projective bundles
P(L) P(E ).
The lemma also implies P(L) = X. Using this and (7.0.4), we get an injection
X X P(W

).
Projection onto the second factor gives a morphism X P(W

), which is the
morphism
L,W
from 6.0 (Exercise 7.0.6).
Proj of a Graded Ring. As described in [90, III.2] and [131, II.2], a graded ring
S =

d=0
S
d
gives the scheme Proj(S) such that for every non-nilpotent f S
d
, d > 0, we have
the afne open subset D
+
( f ) Proj(S) with
D
+
( f ) Spec(S
( f )
),
where S
( f )
is the homogeneous localization of S at f , i.e.,
S
( f )
=

g
f

[ g S
d
, N

.
Furthermore, if homogeneous elements f
1
, . . . , f
s
S satisfy

' f
1
, . . . , f
s
` = S
+
=

d>0
S
d
,
then the afne open subsets D
+
( f
1
), . . . , D
+
( f
s
) cover Proj(S). Thus we construct
Proj(S) by gluing together the afne varieties D
+
( f
i
), just as we construct P
n
by
gluing together copies of C
n
.
The inclusions S
0
S
( f )
for all f give a natural morphism Proj(S) Spec(S
0
).
We have the following important result from [131, Prop. II.7.10].
Theorem 7.0.9. The morphism Proj(S) Spec(S
0
) is projective.
Example 7.0.10. Let U = Spec(R) and consider the graded ring
S = R[x
0
, . . . , x
n
]
such that each x
i
has degree 1. Then
Proj(S) =U P
n
,
where Proj(S) Spec(S
0
) = Spec(R) =U is projection onto the rst factor.
318 Chapter 7. Projective Toric Morphisms
In [131, II.7], the projective bundle P(E ) of a locally free sheaf E on X is
constructed via a relative version of the Proj construction. More generally, one can
dene the projective bundle P(E ) for any coherent sheaf E on X.
Exercises for 7.0.
7.0.1. Prove Proposition 7.0.2.
7.0.2. Prove Proposition 7.0.4. Hint: First prove the special case given by (7.0.1). Recall
from 3.4 that P
n
is complete, so that P
n
pt is proper.
7.0.3. Complete the proof of Proposition 7.0.5.
7.0.4. Let : X Y and : Y Z be morphisms such that : X Z is proper. Prove
that : X Y is also proper. Hint: As noted in the comments following Corollary 3.4.6,
being proper is equivalent to being topologically proper (Denition 3.4.7). Also, T Y
implies
1
(T) ( )
1
((T)).
7.0.5. Prove Lemma 7.0.8. Hint: Work on an open cover where the bundles are trivial.
7.0.6. In the discussion following (7.0.4), we constructed a morphism X P(W

) using
the surjection E =W
C
O
X
L. Prove that this coincides with the morphism
L,W
.
7.0.7. Show that C
2
` (0, 0) is quasiprojective but neither afne nor projective.
7.1. Polyhedra and Toric Varieties
This section and the next will study quasiprojective toric varieties and projective
toric morphisms. Our starting point is the observation that just as polytopes give
projective toric varieties, polyhedra give projective toric morphisms.
Polyhedra. Recall that a polyhedron P M
R
is the intersection of nitely many
closed half-spaces
P =m M
R
[ 'm, u
i
` a
i
, i = 1, . . . , s.
A basic structure theorem tells us that P is a Minkowski sum
P = Q+C,
where Q is a polytope and C is a polyhedral cone (see [281, Thm. 1.2]). If P is
presented as above, then the cone part of P is
(7.1.1) C =m M
R
[ 'm, u
i
` 0, i = 1, . . . , s.
(Exercise 7.1.1). Following [281], we call C the recession cone of P.
Similar to polytopes, polyhedra have supporting hyperplanes, faces, facets,
vertices, edges, etc. One difference is that some polyhedra have no vertices.
Lemma 7.1.1. Let P M
R
be a polyhedron with recession cone C.
(a) The set V = v P [ v is a vertex is nite and is nonempty if and only if C is
strongly convex.
(b) If C is strongly convex, then P = Conv(V) +C.
7.1. Polyhedra and Toric Varieties 319
Proof. You will prove this in Exercises 7.1.27.1.5.
Example 7.1.2. The polyhedron P =(a
1
, . . . , a
n
) R
n
[ a
i
0,

n
i=1
a
i
1 has
vertices e
1
, . . . , e
n
and recession cone C = Cone(e
1
, . . . , e
n
).
Lattice Polyhedra. We now generalize the notion of lattice polytope.
Denition 7.1.3. A polyhedron P M
R
is a lattice polyhedron with respect to the
lattice M M
R
if
(a) The recession cone of P is a strongly convex rational polyhedral cone.
(b) The vertices of P lie in the lattice M.
A full dimensional lattice polyhedron has a unique facet presentation
(7.1.2) P =m M
R
[ 'm, u
F
` a
F
for all facets F,
where u
F
N is a primitive inward pointing facet normal. This was dened in
Chapter 2 for full dimensional lattice polytopes but applies equally well to full
dimensional lattice polyhedra. Then the cone of P is the cone C(P) M
R
R by
C(P) =(m, ) M
R
R[ 'm, u
F
` a
F
for all F, 0.
When P is a polytope, this reduces to the cone C(P) = Cone(P1) considered
in 2.2.
Example 7.1.4. The blowup of C
2
at the origin is given by the fan in R
2
with
minimal generators u
0
= e
1
+e
2
, u
1
= e
1
, u
2
=e
2
and maximal cones Cone(u
0
, u
1
),
Cone(u
0
, u
2
). For the divisor D = D
0
+D
1
+D
2
, we computed in Figure 5 from
Example 4.3.4 that the polyhedron P
D
is a 2-dimensional lattice polyhedron whose
recession cone C is the rst quadrant.
Figure 1 on the next page shows the 3-dimensional cone C(P
D
) with P
D
at
height 1. Notice how the cone C of P
D
appears naturally at height 0 in Figure 1.
Some of the properties suggested by Figure 1 hold in general.
Lemma 7.1.5. Let P be a full dimensional lattice polyhedron in M
R
with recession
cone C. Then C(P) is a strongly convex cone in M
R
R and
C(P) (M
R
0) =C.
Proof. The nal assertion of the lemma follows from (7.1.1) and the denition of
C(P). For strong convexity, note that C(P) M
R
R
0
implies
C(P) (C(P)) M
R
0.
Then we are done since C is strongly convex.
320 Chapter 7. Projective Toric Morphisms
z
x
y
1
C
P
D
Figure 1. The cone C(PD) in Example 7.1.4
We say that a point (m, ) C(P) has height . Furthermore, when > 0, the
slice of C(P) at height is P. If we write P = Q+C, where Q is a polytope,
then for > 0,
P = Q+C
since C is a cone. It follows that as 0, the polytope shrinks to a point so that at
height 0, only the cone C remains, as in Lemma 7.1.5. You can see how this works
in Figure 1.
The Toric Variety of a Polyhedron. In Chapter 2, we constructed the normal fan
of a full dimensional lattice polytope. We now do the same for a full dimensional
lattice polyhedron P. Given a vertex v P, we get the cone
C
v
= Cone(PMv) M
R
.
Note that v M since P is a lattice polyhedron. It follows easily that C
v
is a strongly
convex rational polyhedral cone of maximal dimension, so that the same is true for
its dual

v
=C

v
= Cone(PMv)

N
R
.
These cones t together nicely.
Theorem 7.1.6. Given a full dimensional lattice polyhedron P M
R
with reces-
sion cone C, the set

P
= [ _
v
, v is a vertex of P
is a fan in N
R
. Furthermore:
7.1. Polyhedra and Toric Varieties 321
(a) The support of
P
is [
P
[ =C

.
(b)
P
has full dimensional convex support in N
R
.
Proof. The proof that we get a fan is similar to the proof for the polytope case (see
2.3) and hence is omitted. To prove part (a), we need to show

vV

v
=C

,
where V is the set of vertices of P. Now take v V and m CM. Then m =
(v+m) v PMv, which easily implies C Cone(PMv). Taking duals,
we obtain
v
C

. For the opposite inclusion, take u C

and pick v V such


that 'v, u` 'w, u` for all w V. We show u
v
as follows. Any m PM can
be written m =

wV

w
w+m

where
w
0,

wV

w
= 1 and m

C. Then
'm, u` =

wV

w
'w, u` +'m

, u`

wV

w
'v, u` ='v, u`.
Thus 'mv, u` 0 for all mv PMv, which proves u
v
.
Part (b) now follows since C

is clearly convex, and has full dimension since


C is strongly convex.
The fan
P
of Theorem 7.1.6 is the normal fan of P. We dene X
P
to be the
toric variety X

P
of
P
. Here is an example.
Example 7.1.7. The polyhedron P = (a
1
, . . . , a
n
) R
n
[ a
i
0,

n
i=1
a
i
1 of
Example 7.1.2 has vertices e
1
, . . . , e
n
. The facet of P dened by

n
i=1
a
i
= 1 has
e
1
+ +e
n
as inward normal. Then the vertex e
i
gives the cone

e
i
= Cone(e
1
+ +e
n
, e
1
, . . . , e
i
, . . . , e
n
).
These cones form the fan of the blowup of C
n
at the origin, so X
P
= Bl
0
(C
n
).
Note that X
P
is not complete in this example. In general, the normal fan has
support [
P
[ =C

. We measure the deviation from completeness as follows.


The support [
P
[ is a rational polyhedral cone but need not be strongly convex.
Recall that W = [
P
[ ([
P
[) is the largest subspace contained in [
P
[. Hence
[
P
[ gives the following:
The sublattice W N N and the quotient lattice N
P
= N/(W N).
The strongly convex cone
P
=[
P
[/W N
R
/W = (N
P
)
R
.
The afne toric variety U
P
of
P
.
The projection map : N N
P
is compatible with the fans of X
P
and U
P
since

R
([
P
[) =
P
. Hence we get a toric morphism
: X
P
U
P
.
Since [
P
[ =
1
R
(
P
) (Exercise 7.1.6), Theorem 3.4.11 implies that is proper.
322 Chapter 7. Projective Toric Morphisms
The key result of this section is that : X
P
U
P
is a projective morphism. We
rst give an elementary proof in Theorem 7.1.10. We will also give a more sophis-
ticated proof that applies Proj construction from 7.0 to the semigroup algebra
(7.1.3) S
P
=C[C(P) (MZ)].
The character associated to (m, k) C(P) (MZ) is written
m
t
k
, and S
P
is
graded by height, i.e, deg(
m
t
k
) = k. In Proposition 7.1.13 we will prove that
X
P
Proj(S
P
).
Then standard properties of Proj will imply that : X
P
U
P
is projective.
The Divisor of a Polyhedron. Let P be a full dimensional lattice polyhedron in
M
R
. As in the polytope case, facets of P correspond to rays in the normal fan
P
,
so that each facet F gives a prime torus-invariant divisor D
F
X
P
. Thus the facet
presentation (7.1.2) of P gives the divisor
(7.1.4) D
P
=

F
a
F
D
F
,
where the sum is over all facets of P. Results from Chapter 4 (Proposition 4.2.10
and Example 4.3.7) easily adapt to the polyhedral case to show that D
P
is Cartier
(with m
v
= v for every vertex v) and the polyhedron of D
P
is P, i.e., P = P
D
P
.
Then Proposition 4.3.3 implies that
(7.1.5) (X
P
, O
X
P
(D
P
)) =

mPM
C
m
.
The denition of projective morphism given in 7.0 involves a line bundle L
and a nite-dimensional subspace W of global sections. The line bundle will be
O
X
P
(kD
P
) for a suitably chosen integer k 1 and W will be determined by certain
carefully chosen lattice points of kP. The reason we need a multiple is that P might
not have enough lattice points.
Normal and Very Ample Polyhedra. In Chapter 2, we dened normal and very
ample polytopes, which are different ways of saying that there are enough lattice
points. For a lattice polyhedron P, the denitions are the same.
Denition 7.1.8. Let P M
R
be a lattice polyhedron. Then:
(a) P is normal if for all integers k 1, every lattice point of kP is a sum of k
lattice points of P.
(b) P is very ample if for every vertex v P, the semigroup N(PMv) generated
by PMv is saturated in M.
We have the following result about normal and very ample polyhedra.
Proposition 7.1.9. Let P M
R
be a lattice polyhedron. Then:
(a) If P is normal, then P is very ample.
(b) If dim P = n 2, then kP is normal and hence very ample for all k n1.
7.1. Polyhedra and Toric Varieties 323
Proof. Part (a) follows from the proof of Proposition 2.2.18. For part (b), write
P = Q+C, where Q is a lattice polytope and C is the recession cone of P. Let
C = Cone(m
1
, . . . , m
s
), m
i
M. In Exercise 7.1.7 you will show that
(7.1.6) C =

mCM
Conv(0, sm
1
, . . . , sm
s
) +m.
Since P = Q+C is a full dimensional polyhedron, Q+Conv(0, sm
1
, . . . , sm
s
) is
a full dimensional lattice polytope. It follows that P = Q+C is a union of full
dimensional lattice polytopes. Then part (b) follows by applying Theorem 2.2.12
to each of these polytopes.
The Projective Morphism. Let P be a full dimensional lattice polyhedron in M
R
.
Assume P is very ample and pick a nite set A PM such that:
A contains the vertices of P.
For every vertex v P, A v generates Cone(PMv) M =

v
M.
We can always satsify the rst condition, and the second is possible since P is very
ample. Using (7.1.5), we get the subspace
W = Span(
m
[ m A) (X
P
, O
X
P
(D
P
)).
We claim that W has no basepoints since A contains the vertices of P. To prove
this, let v be a vertex. Recall that D
P
+div(
v
) is the divisor of zeros of the global
section given by
v
. One computes that
D
P
+div(
v
) =

F
(a
F
+'v, u
F
`)D
F
.
Since 'v, u
F
` = a
F
for all facets containing v and 'v, u
F
` > a
F
for all other
facets, the support of D
P
+div(
v
) is the complement of the afne open subset
U
v
X
P
, i.e., the nonvanishing set of the section is precisely U
v
. Then we are
done since the U
v
cover X
P
.
It follows that we get a morphism

L,W
: X
P
P(W

)
for L =O
X
P
(D
P
). Here is our result.
Theorem 7.1.10. Let P be a full dimensional lattice polyhedron. Then:
(a) The toric variety X
P
is quasiprojective.
(b) : X
P
U
P
is a projective morphism.
Proof. First suppose that P is very ample. The proof of part (a) is similar to
the proof of Proposition 6.1.10. Let L, W and A be as above and write A =
m
1
, . . . , m
s
. Consider the projective toric variety
X
A
P
s1
=P(W

).
324 Chapter 7. Projective Toric Morphisms
Let I 1, . . . , s be the set of indices corresponding to vertices of P. So i I gives
a vertex m
i
and a corresponding cone
i
=
m
i
in
P
. Also let U
i
P
s1
be the
afne open subset where the ith coordinate is nonzero. By our choice of A, the
proof of Proposition 6.1.10 shows that
L,W
induces an isomorphism
U

i
X
A
U
i
.
Since X
P
is the union of the U

i
for i I, it follows that
(7.1.7)
L,W
: X
P

X
A

iI
U
i
.
Since X
A
is projective, this shows that X
P
is quasiprojective. Part (b) now follows
immediately from Proposition 7.0.6 since : X
P
U
P
is proper.
When P is not very ample, we know that a positive multiple kP is. Since P
and kP have the same normal fan and same recession cone, the maps X
P
U
P
and X
kP
U
kP
are identical. Hence the general case follows immediately from the
very ample case.
Here are two examples to illustrate Theorem 7.1.10.
Example 7.1.11. The polyhedron P from Example 7.1.7 is very ample (in fact, it
is normal), and the set A used in the proof of Theorem 7.1.10 can be chosen to
be A = e
1
, . . . , e
n
, 2e
1
, . . . , 2e
n
(Exercise 7.1.8). This gives X
A
P
2n1
, where
P
2n1
has variables x
1
, . . . , x
n
, w
1
, . . . , w
n
corresponding to the elements e
1
, . . . , e
n
,
2e
1
, . . . , 2e
n
of A. Then X
A
P
2n1
is dened by the equations x
2
i
w
j
= x
2
j
w
i
for 1 i < j n (Exercise 7.1.8). Since X
P
= Bl
0
(C
n
) by Example 7.1.7, the
isomorphism (7.1.7) implies
Bl
0
(C
n
) (x
1
, . . . , x
n
, w
1
, . . . , w
n
) P
2n1
[ (x
1
, . . . , x
n
) = (0, . . . , 0)
and x
2
i
w
j
= x
2
j
w
i
for 1 i < j n.
We get a better description of Bl
0
(C
n
) using the vertices B = e
1
, . . . , e
n
of P.
This gives a map X
P
P
n1
which, when combined with X
P
U
P
= C
n
, gives
a morphism
: X
P
P
n1
C
n
.
Let P
n1
and C
n
have variables x
1
, . . . , x
n
and y
1
, . . . , y
n
respectively. Then is an
embedding onto the subvariety of P
n1
C
n
dened by x
i
y
j
=x
j
y
i
for 1 i < j n
(Exercise 7.1.8). Hence
Bl
0
(C
n
) (x
1
, . . . , x
n
, y
1
, . . . , y
n
) P
n1
C
n
[ x
i
y
j
= x
j
y
i
, 1 i < j n.
This description of the blowup Bl
0
(C
n
) can be found in many books on algebraic
geometry and appeared earlier in this book as Exercise 3.0.8. Note also that the
projective morphism of Theorem 7.1.10 is the blowdown map Bl
0
(C
n
) C
n
.
Example 7.1.12. Consider the full dimensional lattice polyhedron P R
2
dened
by the inequalities
x 2, 0 y 2, y x +1.
7.1. Polyhedra and Toric Varieties 325
This polyhedron has vertices v
1
= (1, 0), v
2
= (2, 1), v
3
= (2, 2) shown in Figure 2.
The left side of the gure also shows the recession cone C and the decompostion
P = Q+C, where Q is the convex hull of the vertices.
P
v
1
v
2
v
3
(0,0)

P
Q C
P = Q + C
v
1
v
2
v
3
(0,0)

P
Figure 2. The polyhedron P =Q+C, the normal fan P, and the cone P
The normal vectors at each vertex v
i
are reassembled on the right to give the
maximal cones
i
of the normal fan
P
. Note also that [
P
[ is not strictly convex,
so we mod out by its maximal subspace to get the strictly convex cone
P
. The
projection map on the right of Figure 2 gives the projective morphism X
P
U
P
,
where U
P
C is the toric variety of
P
.
The Proj Construction. We conclude this section by explaining how construct X
P
using Proj. Let P M
R
be a full dimensional lattice polyhedron. By (7.1.3), the
cone C(P) M
R
R gives the semigroup algebra
S
P
=C[C(P) (MZ)].
We use the height grading dened by deg(
m
t
k
) = k. Then we can relate Proj(S
P
)
to X
P
as follows.
Theorem 7.1.13. There is a natural isomorphism X
P
Proj(S
P
).
Proof. Let V be the set of vertices of P. In Exercise 7.1.9 you will prove:
326 Chapter 7. Projective Toric Morphisms

'
v
t [ v V` = (S
P
)
+
=

d>0
(S
P
)
d
.
If v V, then (S
P
)
(
v
t)
=C[

v
M], where
v
= Cone(PMv)

.
By the rst bullet, Proj(S
P
) is covered by the afne open subsets Spec((S
P
)
(
v
t)
),
and the second shows that Spec((S
P
)
(
v
t)
) is the afne toric variety of the cone
v
.
These patch together in the correct way to give X
P
Proj(S
P
).
We can also interpret the morphism : X
P
U
P
in terms of Proj. The idea is to
compute (S
P
)
0
, the degree 0 part of the graded ring S
P
=C[C(P) (MZ)]. The
slice of C(P) at height 0 is the recession cone C of P. Recall that N
P
= N/(WN),
where W C

is the largest subspace contained in C

and that U
P
is the afne
toric variety of
P
, which is the image of C

in (N
P
)
R
. Then the inclusion M
P
M
dual to N N
P
gives

P
=C (M
P
)
R
M
R
.
It follows that (S
P
)
0
=C[CM] =C[

P
M]. This implies Spec((S
P
)
0
) =U
P
, so
that Proj(S
P
) Spec((S
P
)
0
) becomes : X
P
U
P
. It follows that is projective
by Proposition 7.0.9. This gives a second proof of Theorem 7.1.10.
It is also possible to prove directly that Proj(S
P
) Spec((S
P
)
0
) is projective
without using Proposition 7.0.9. See Exercise 7.1.10.
Exercises for 7.1.
7.1.1. Prove (7.1.1). Hint: Fix m
0
P and take any m C. Show that m
0
+m P for
> 0, so 'm
0
+m, u
i
` a
i
. Then divide by and let .
7.1.2. Let P = Q+C be a polyhedron in M
R
where Q is a polytope and C is a polyhedral
cone. Dene
P
(u) = min
mP
'm, u` for u C

.
(a) Show that
P
(u) = min
mQ
'm, u` for u C

and conclude that


P
: C

R is well-
dened.
(b) Show that
P
(u) = min
vVQ
'v, u` for u C

, where V
Q
is the set of vertices of Q.
(c) Show that P = m M
R
[
P
(u) 'm, u` for all u C

. Hint: For the non-obvious


direction, represent P as the intersection of closed half-spaces coming fromsupporting
hyperplanes.
7.1.3. Let P be a polyhedron in M
R
with recession cone C and let W =C(C) be the
largest subspace contained in C. Prove that every face of P contains a translate of W and
conclude that P has no vertices when C is not strongly convex.
7.1.4. Let P = Q+C be a polyhedron in M
R
where Q is a polytope and C is a strongly
convex polyhedral cone. Let V
Q
be the set of vertices of Q. Assume that there is v V
Q
and u in the interior of C

such that 'v, u` < 'w, u` for all w = v in V


Q
. Prove that v is a
vertex of P. Hint: Show that H
u,a
, a = 'v, u`, is a supporting hyperplane of P such that
H
u,a
P = v. Also show if v and u satisfy the hypothesis of the problem, then so do v and
u

for any u

sufciently close to u. Finally, Exercise 7.1.2 will be useful.


7.1.5. Let P = Q+C be a polyhedron in M
R
where Q is a polytope and C be a strongly
convex polyhedral cone. Let V
Q
be the set of vertices of Q and let
U
0
=u Int(C

) [ 'v, u` ='w, u` whenever v = w in V


Q
.
7.1. Polyhedra and Toric Varieties 327
(a) Show that U
0
is open and dense in C

. Hint: dimC

= dim N
R
.
(b) Use Exercise 7.1.4 to show that for every u U
0
, there is a vertex v of P such that

P
(u) ='v, u`. Conclude that the set V
P
of vertices of P is nonempty and nite.
(c) Show that
P
(u) = min
vVP
'v, u` for u C

.
(d) Conclude that P = Conv(V
P
) +C. Hint: The rst step is to show that
P
=
P
, where
P

= Conv(V
P
) +C. Then use part (c) of Exercise 7.1.2.
7.1.6. Let C N
R
be a polyhedral cone. Set W =C(C) and let = (C) N
R
/W for
the quotient map : N
R
N
R
/W. Show that is strongly convex and C =
1
().
7.1.7. Prove (7.1.6). Hint: Given

s
i=1

i
m
i
C, let m =

s
i=1

i
m
i
CM.
7.1.8. Prove the claims made in Example 7.1.11.
7.1.9. Supply proofs of the two bullets from the proof of Theorem 7.1.13.
7.1.10. Here you will give an elementary proof that Proj(S
P
) Spec((S
P
)
0
) is projective,
where S
P
is the graded semigroup algebra from (7.1.3).
(a) Explain why we can assume that P is normal.
(b) Showthat C(P)(MZ) is generated by its elements of height 1 when P is normal.
(c) Assume P is normal and let H be a Hilbert basis of C(P) (MZ). Then H =
H
0
H
1
, where elements of H
i
have height i and write H
1
= (m
1
, 1), . . . , (m
s
, 1).
Prove that S
P
is generated as an (S
P
)
0
-algebra by
m1
t, . . . ,
ms
t and conclude that
there is a surjective homomorphismof graded rings
(S
P
)
0
[x
1
, . . . , x
s
] S
P
, x
i

mi
t.
(d) Prove that when P is normal, there is a commutative diagram
X
P

U
P
P
s1
p1

U
P
such that is a closed embedding and is a projective morphism.
7.1.11. In this exercise, you will prove a stronger version of part (b) of Theorem 7.1.10.
Let X
A
and W be as in the proof of the theorem. Prove that there is a commutative diagram
(7.1.8)
X
P

L,W

U
P
P(W

)
p1

U
P
such that
L,W
: X
P
U
P
P(W

) is a closed embedding. Hint: Proposition 7.0.6.


7.1.12. Let N
R
be a strongly convex rational polyhedral cone. This gives the semi-
group algebra C[S

] = C[

M]. Given a monomial ideal a = '


m1
, . . . ,
ms
` C[S

],
we get the polyhedron
P = Conv(m M [
m
a),
Prove that P = Conv(m
1
, . . . , m
s
) +

.
328 Chapter 7. Projective Toric Morphisms
7.2. Projective Morphisms and Toric Varieties
We begin our study of projective toric morphisms with a toric variety X

whose fan
has full dimensional convex support. We construct an afne toric variety from
[[ as follows. The largest subspace contained in [[ is W =[[ ([[). Similar
to 7.1, we have:
The sublattice W N N and the quotient lattice N

= N/(W N).
The strongly convex cone

=[[/W N
R
/W = (N

)
R
.
The afne toric variety U

=U

.
The projection map : N N

is compatible with the fans of X

and U

since

R
([[) =

. This gives a toric morphism


(7.2.1) : X

.
which as in 7.1 is easily seen to be proper. The difference between here and
7.1 is that : X

may fail to be projective. Our rst goal is to understand


when this happens. As you might suspect, the answer involves polyhedra, support
functions, and convexity.
The Polyhedron of a Divisor. A Weil divisor D =

on X

gives the poly-


hedron
P
D
=m M
R
[ 'm, u

` a

for all .
When is complete, this is a polytope, but in general we have
P
D
= Q+C,
where Q is a polytope and C is the recession cone of P
D
.
Lemma 7.2.1. Assume [[ is convex of full dimension and let D =

be a
Weil divisor on X

. If P
D
=, then:
(a) The recession cone of P
D
is [[

.
(b) The set V =v P
D
[ v is a vertex is nonempty and nite.
(c) P
D
= Conv(V) +[[

.
Proof. Combining (7.1.1) with the denition of P
D
, we see that the recession cone
of P
D
is
m M
R
[ 'm, u

` 0 for all =[[

since [[ = Cone(u

[ (1)) by (6.1.3). This proves part (a). The recession


cone is strongly convex since [[ has full dimension, so that parts (b) and (c) follow
from Lemma 7.1.1.
7.2. Projective Morphisms and Toric Varieties 329
Divisors and Convexity. Now that we know about recession cones, the convexity
result proved in Theorem 6.1.7 can be improved as follows.
Theorem 7.2.2. Assume [[ is convex of full dimension n and let
D
be the support
function of a Cartier divisor D on X

. Then the following are equivalent:


(a) D is basepoint free.
(b) m

P
D
for all (n).
(c)
D
(u) = min
(n)
'm

, u` for all u [[.


(d)
D
: [[ R is convex.
(e) P
D
= Conv(m

[ (n)) +[[

.
(f) m

[ (n) is the set of vertices of P


D
.
(g)
D
(u) = min
mP
D
'm, u` for all u [[.
In particular, P
D
is a lattice polyhedron when D is basepoint free.
Proof. Parts (a)(d) are equivalent by Theorem 6.1.7. Furthermore, (b) (f) and
(b) (g) follow as in the proof of Theorem 6.1.7, and (f) (e) follows from
Lemma 7.2.1. Also, (e) (b) is obvious. Finally, (g) (d) follows from part (b)
of Exercise 7.1.2.
Strict Convexity. Our next task is to show that : X

is projective if and only


if X

has a Cartier divisor D with strictly convex support function. We continue


to assume that has full dimensional convex support. As in 6.1,
D
is strictly
convex if it is convex and for each (n), the equation
D
(u) = 'm

, u` holds
only on . The strict convexity criteria from Lemma 6.1.13 apply to this situation.
When
D
is strictly convex, the polyhedron P
D
has an especially nice relation
to the fan .
Proposition 7.2.3. Assume that [[ is convex of full dimension and D =

has a strictly convex support function. Then:


(a) P
D
is a full dimensional lattice polyhedron.
(b) is the normal fan of P
D
.
Proof. Theorem 7.2.2 and Lemma 6.1.13 imply that the m

, (n), are distinct


and give the vertices of the polyhedron. As in 7.1, a vertex m

P
D
gives the
cone C
m
= Cone(P
D
Mm

). We claim that
=C

m
.
This easily implies that P
D
has full dimension and that is the normal fan of P
D
.
Fix (n) and let (1). Then m P
D
M implies
(7.2.2) 'm, u

`
D
(u

) ='m

, u

`,
330 Chapter 7. Projective Toric Morphisms
where the inequality holds by Lemma 6.1.6 and the equality holds since u

.
Thus 'mm

, u

` 0 for all m P
D
M, so that u

m
for all (1). Hence
C

m
.
Since [[

is the recession cone of P


D
, the proof of Theorem 7.1.6 implies
C

m
[[ =

(n)
.
Now take u Int(C

m
). Hence u

for some

(n). Then u C

m
and
m

C
m
imply
'm

, u` 0, so 'm

, u` 'm

, u`.
On the other hand, if we apply (7.2.2) to the cone

and m = m

, we obtain
'm

, u

` 'm

, u

`. We conclude that
'm

, u` ='m

, u`,
and the same equality holds for all elements of Int(C

m
)

. This easily implies


that m

= m

. Then =

by strict convexity, so that u .


Here is the rst major result of this section.
Theorem 7.2.4. Let : X

be the proper toric morphism where U

is afne.
Then [[ is convex. Furthermore, the following are equivalent:
(a) X

is quasiprojective.
(b) is a projective morphism.
(c) X

has a torus-invariant Cartier divisor with strictly convex support function.


Proof. Since is proper, Theorem 3.4.11 implies that [[ =
1
R
(). Thus [[ is
convex. To prove (a) (b) (c), rst assume that [[ has full dimension.
If (c) is true, then is the normal fan of the full dimensional lattice polyhedron
P
D
by Proposition 7.2.3. It follows that X

= X
P
D
, which is quasiprojective by
Theorem 7.1.10, proving (a). Furthermore, (a) (b) by Proposition 7.0.6.
If (b) is true, we will use the theory of ampleness developed in [127]. The
essential facts we need are summarized in the appendix to this chapter. Since
is projective, there is a line bundle L on X

that satises Denition 7.0.3. Then,


since U

is afne, Theorem 7.A.4 and Proposition 7.A.6 imply that


L
k
=L
O
X


O
X

L (k times) is generated by global sections for some


integer k > 0.
The nonvanishing set of a global section of L is an afne open subset of X

.
By 7.0, L O
X

(D) for some Cartier divisor on X, and since linearly equiv-


alent Cartier divisors give isomorphic line bundles, we may assume that D is torus-
invariant (this follows from Theorem 4.2.1). Then O
X

(kD) is generated by global


7.2. Projective Morphisms and Toric Varieties 331
sections for some k > 0. This implies that
kD
= k
D
is convex by Theorem 7.2.2,
so that
D
is convex as well. We show that
D
is strictly convex by contradiction.
If strict convexity fails, then Lemma 6.1.13 implies that there is a wall =

in with m

= m

. Then m = m

= m

corresponds to a global section


s, which by the proof of Proposition 6.1.1 is nonvanishing on U

(since m = m

)
and on U

(since m = m

). Thus the nonvanishing set contains U

, which
contains the complete curve V() U

. But being afne, the nonvanishing


set cannot contain a complete curve (Exercise 7.2.1). This completes the proof of
the theorem when [[ has full dimension.
Finally, suppose that [[ fails to have full dimension. Let N
1
= Span([[) N
and pick N
0
N such that N = N
0
N
1
. The cones of lie in (N
1
)
R
and hence
give a fan
1
in (N
1
)
R
. If N
0
has rank r, then Proposition 3.3.11 implies that
(7.2.3) X

(C

)
r
X

1
.
It follows that
D
: [[ = [
1
[ R is the support function of a Cartier divisor D
1
on X

1
. Note also that [
1
[ is convex of full dimension in (N
1
)
R
. Since (C

)
r
is
quasiprojective, this allows us to reduce to the case of full dimensional support.
You will supply the omitted details in Exercise 7.2.2.
f -Ample and f -Very Ample Divisors. The denitions of ample and very ample
from 6.1 generalize to the relative setting as follows. Recall from Denition 7.0.3
that a morphism f : X Y is projective with respect to the line bundle L when for
a suitable open cover U
i
of Y, we can nd global sections s
0
, . . . , s
k
i
of L over
f
1
(U
i
) that give a closed embedding
f
1
(U
i
) U
i
P
k
i
.
Then we have the following denition.
Denition 7.2.5. Let D be a Cartier divisor on X and f : X Y be proper.
(a) The divisor D and the line bundle O
X
(D) are f -very ample if f is projective
with respect to the line bundle L =O
X
(D).
(b) D and O
X
(D) are f -ample when kD is f -very ample for some integer k > 0.
Hence f : X Y is projective if and only if X has an f -ample line bundle. In
the toric case, Proposition 7.1.9 and Theorem 7.2.4 give the following result.
Theorem 7.2.6. Let : X

be a proper toric morphism where U

is afne,
and let D =

be a Cartier divisor on X

. Also let n = dim X

. Then:
(a) D is -ample if and only if
D
is strictly convex.
(b) If n 2 and D is -ample, then kD is -very ample for all k n.
332 Chapter 7. Projective Toric Morphisms
Here is are two examples of Theorem 7.2.6.
Example 7.2.7. Consider the blowdown morphism : Bl
0
(C
n
) C
n
. The fan for
Bl
0
(C
n
) has minimal generators u
0
= e
1
+ +e
n
and u
i
= e
i
for 1 i n. Let
D
0
be the divisor corresponding to u
0
. The support function
D
0
of D
0
is easily
seen to be strictly convex (Exercise 7.2.4). Thus D
0
is -ample by Theorem 7.2.6.
Note also that the polyhedron P
D
0
is the polyhedron P from Example 7.1.7.
Example 7.2.8. The P be a full dimensional lattice polyhedron in M
R
. The map
: X
P
U
P
is projective by Theorem 7.1.10. We also have the Cartier divisor D
P
on X
P
dened in (7.1.4). As noted in the discussion following (7.1.4), P = P
D
P
and the vertices of P give the Cartier data of D
P
, so that
D
P
is strictly convex by
Theorem 7.2.2 and Lemma 6.1.13. Hence D
P
is -ample by Theorem 7.2.6.
Semiprojective Toric Varieties. Following [137], we say that X

is semiprojective
if the natural map : X

Spec((X

, O
X

)) is projective and X

has a torus
xed point. We can characterize semiprojective toric varieties as follows.
Proposition 7.2.9. Given a toric variety X

, the following are equivalent:


(a) X

is semiprojective.
(b) X

is quasiprojective and has full dimensional convex support in N


R
.
(c) X

= X
P
is the toric variety of a full dimensional lattice polyhedron P M
R
.
Proof. By the Orbit-Cone Correspondence (Theorem 3.2.6), X

has a torus xed


point if and only if has a full dimensional cone, which is equivalent to having
full dimensional support. In Exercise 7.2.3 you will show that Spec((X

, O
X

))
is a normal afne toric variety U

. Then (a) (b) follows from Theorem 7.2.4.


The equivalence (b) (c) follows from Proposition 7.2.3 and Theorem 7.2.4.
This completes the proof.
A semiprojective toric X

variety comes equipped with a projective morphism


: X

Spec((X

, O
X

)), and a full dimensional lattice polyhedron P comes


with a projective morphism : X
P
U
P
by Theorem 7.1.10. These maps are the
same by Exercise 7.2.3.
We can also extend the relation between polytopes and ample divisors on com-
plete toric varieties described in 6.2. Consider the set of polyhedra

P M
R
[ P is a full dimensional lattice polytope

and the set of pairs

(X

, D) [ is a fan in N
R
, X

is semiprojective, and
D is a torus-invariant -ample divisor on X

.
These sets are related as follows.
7.2. Projective Morphisms and Toric Varieties 333
Theorem 7.2.10. The maps P (X
P
, D
P
) and (X

, D) P
D
dene bijections be-
tween the above sets that are inverses of each other.
Proof. First note that X
P
is semiprojective by Proposition 7.2.9, and D
P
is -ample
by Example 7.2.8. Going the other way, suppose that X

is semiprojective and D
is -ample. Then Theorem 7.2.6 and Proposition 7.2.9 imply that D
P
has a strictly
convex support function, so that P
D
is a full dimensional lattice polyhedron by
Proposition 7.2.3.
Using Proposition 7.2.3 and P = P
D
P
, it is easy to see that the two maps are
inverses of each other.
Projective Toric Morphisms. Suppose we have fans in N
R
and

in N

R
. Recall
from 3.3 that a toric morphism
: X

is induced from a map of lattices


: N N

compatible with and

, i.e., for each there is

with
R
()

.
We rst determine when a torus-invariant Cartier divisor on X

is -ample.
Since projective morphisms are proper, we can assume that is proper, which by
Theorem 3.4.11 is equivalent to
(7.2.4) [[ =
1
R
([

[).
Here is our result.
Theorem 7.2.11. Let : X

be a proper toric morphism and let D =

be a Cartier divisor on X

. Also let n = dim X

. Then:
(a) D is -ample if and only if for every

,
D
is strictly convex on
1
R
(

).
(b) If n 2 and D is -ample, then kD is -very ample for all k n1.
Proof. The idea is to study what happens over the afne open subsets U

for

. Observe that
1
(U

) is the toric variety corresponding to the fan

= [
R
()

.
Thus
1
(U

) = X

. Let

= [

1
(U

)
and consider the diagram
X

1
(U

)
?

.
334 Chapter 7. Projective Toric Morphisms
Also let D

be the restriction of D to
1
(U

) = X

.
By Proposition 7.A.5, D is -ample if and only if the restriction D[

1
(U

)
is
[

1
(U

)
-ample for all

. Using the above notation, this becomes


D is -ample D

is

-ample for all

.
However, Theorem 7.2.6 implies that
D

is

-ample
D

is strictly convex.
This completes the proof of the theorem.
It is now easy to characterize when a toric morphism is projective.
Theorem 7.2.12. Let : X

be a toric morphism. Then the following are


equivalent:
(a) is projective.
(b) is proper and X

has a torus-invariant Cartier divisor D whose support


function
D
is strictly convex on
1
R
(

) for all

.
You will prove Theorem 7.2.12 in Exercise 7.2.5. The rst proof of this theo-
rem was given in [172, Thm. 13 of Ch. I]. In Chapter 11 we will use this result to
construct interesting examples of projective toric morphisms.
Exercises for 7.2.
7.2.1. Prove that an afne variety cannot contain a complete variety of positive dimension.
Hint: If X is complete and irreducible, then (X, O
X
) =C.
7.2.2. This exercise will complete the proof of Theorem 7.2.4. Let : X

satisfy
the hypothesis of the theorem and write X

as in (7.2.3). We also have the Cartier divisors


D on X

and D
1
on X
1
as in the proof of the theorem.
(a) Assume that is projective. Prove that X

is quasiprojective and conclude that X


1
is
quasiprojective. Now use the rst part of the proof to show that
D
is strictly convex.
Hint: See Exercise 7.0.1.
(b) Assume that
D
is strictly convex. Prove that X
1
is quasiprojective and conclude that
X

is quasiprojective. Then use Proposition 7.0.6.


7.2.3. Given a toric variety X

, let C = m M
R
[ 'm, u

` 0 for all (1), and let


be the strongly convex cone obtained by taking the quotient of C

by its minimal face.


Prove that U

Spec((X

, O
X
)).
7.2.4. Prove that the support function
D0
in Example 7.2.7 is strictly convex. We will
generalize this result considerably in Chapter 11.
7.2.5. Prove Theorem 7.2.12.
7.3. Projective Bundles and Toric Varieties 335
7.3. Projective Bundles and Toric Varieties
Given a vector bundle or projective bundle over a toric variety, the nicest case is
when the bundle is also a toric variety. This will lead to some lovely examples of
toric varieties.
Toric Vector Bundles and Cartier Divisors. A Cartier divisor D =

on a
toric variety X

gives the line bundle L = O


X

(D), which is the sheaf of sections


of the rank 1 vector bundle : V
L
X

.
We will show that V
L
is a toric variety and is a toric morphism by construct-
ing the fan of V
L
in terms of and D. To motivate our construction, recall that for
m M, we have

m
(X

, O
X

(D)) m P
D
'm, u`
D
(u) for all u [[
the graph of u 'm, u` lies
above the graph of
D
.
The rst equivalence follows from Proposition 4.3.3 and the second from Propos-
tion 6.1.6. The key word is above: it tells us to focus on the part of N
R
R that
lies above the graph of
D
.
We dene the fan D in N
R
R as follows. Given , set
=(u, ) [ u ,
D
(u)
= Cone((0, 1), (u

, a

) [ (1)),
where the second equality follows since
D
(u

) =a

and
D
is linear on . Note
that is a strongly convex rational polyhedral cone in N
R
R. Then let D be
the set consisting of the cones for and their faces. This is a fan in N
R
R,
and the projection : N Z N is clearly compatible with D and . Hence
we get a toric morphism
: X
D
X

.
Proposition 7.3.1. : X
D
X

is a rank 1 vector bundle whose sheaf of sec-


tions is O
X

(D).
Proof. We rst show that is a toric bration as in Theorem 3.3.19. The kernel
of : NZ N is N
0
=0Z, and the fan
0
= D[ (N
0
)
R
has

0
= Cone((0, 1)) as its unique maximal cone. Also, for let
= Cone((u

, a

) [ (1)).
This is the face of consisting of points (u, ) where
D
(u) = . Thus D
and in fact

= [ is a subfan of D. Since = +
0
and
R
336 Chapter 7. Projective Toric Morphisms
maps bijectively to , we see that D is split by and
0
in the sense of
Denition 3.3.18. Since X

0
,N
0
=C, Theorem 3.3.19 implies that

1
(U

) U

C.
To see that this gives the desired vector bundle, we study the transition func-
tions. First note that
1
(U

) =U
e
, so that the above isomorphism is
U
e
U

C,
which by projection induces a map U
e
C. It is easy to check that this map is

(m,1)
, where
D
(u) ='m

, u` for u (Exercise 7.3.1). Note that


(m

, 1)

(MZ),
follows directly from the denition of . Then, given another cone , the
transition map fromU

CU

Cto U

CU

Cis given by (u, t)


(u, g

(u)t), where g

=
m m
(Exercise 7.3.1).
We are now done, since the proof of Proposition 6.2.7 shows that O
X

(D) is
the sheaf of sections of a rank 1 vector bundle over X

whose transition functions


are g

=
m m
.
This construction is easy but leads to some surprisingly rich examples.
Example 7.3.2. Consider P
n
with its usual fan and let D
0
correspond to the min-
imal generator u
0
= e
1
e
n
. Recall that O
P
n (D
0
) is denoted O
P
n (1).
This gives the rank 1 vector bundle V P
n
described in Proposition 7.3.1 whose
fan in R
n
R=R
n+1
has minimal generators
e
1
, . . . , e
n+1
, e
1
e
n
+e
n+1
.
You will check this in Exercise 7.3.2.
We can also describe this vector bundle geometrically as follows. Consider the
lattice polyhedron in R
n+1
given by
P = Conv(0, e
1
, . . . , e
n
) +Cone(e
n+1
, e
1
+e
n+1
, . . . , e
n
+e
n+1
).
The normal fan of P is the fan (Exercise 7.3.2), so that X
P
is the above vector
bundle V. Note also that [[ is dual to the recession cone of P.
It is easy to see that [[ is a smooth cone of dimension n+1, so that the pro-
jective toric morphism X
P
U
P
constructed in 7.1 becomes X
P
C
n+1
. When
combined with the vector bundle map X
P
=V P
n
, we get a morphism
X
P
P
n
C
n+1
.
When the coordinates of P
n
and C
n+1
are ordered correctly, the image is precisely
the variety dened by x
i
y
j
= x
j
y
i
(Exercise 7.3.2). In this way, we recover the
description of V P
n
given in Example 6.0.19.
7.3. Projective Bundles and Toric Varieties 337
Proposition 7.3.1 extends easily to decomposable toric vector bundles. Sup-
pose we have r Cartier divisors D
i
=

a
i
D

, i = 1, . . . , r. This gives the locally


free sheaf
(7.3.1) O
X

(D
1
) O
X

(D
r
)
of rank r. To construct the fan of the corresponding vector bundle, we work in
N
R
R
r
. Let e
1
, . . . , e
r
be the standard basis of R
r
and write elements of N
R
R
r
as u+
1
e
1
+ +
r
e
r
. Then, given , we get the cone
=u+
1
e
1
+ +
r
e
r
[ u ,
i

D
i
(u) for i = 1, . . . , r
= Cone(u

a
1
e
1
a
r
e
r
[ (1)) +Cone(e
1
, . . . , e
r
).
One can show without difculty that the set consisting of the cones for
and their faces is a fan in N
R
R
r
such that the toric variety of this fan is the vector
bundle over X

whose sheaf of sections is (7.3.1) (Exercise 7.3.3).


Besides decomposable vector bundles, one can also dene a toric vector bundle
: V X

. Here, rather than assume that V is a toric variety, one makes the weaker
assumption the torus of X

acts on V such that the action is linear on the bers and


is equivariant. Toric vector bundles have been classied by Klyachko [178] and
otherssee [225] for the historical background. Oda noted in [217, p. 41] that if a
toric vector bundle is a toric variety in its own right, then the bundle is a direct sum
of line bundles, as above. This can be proved using Klyachkos results.
Toric Projective Bundles. The decomposable toric vector bundles have associated
toric projective bundles. Cartier divisors D
0
, . . . , D
r
give the locally free sheaf
E =O
X

(D
0
) O
X

(D
r
),
of rank r +1. Then P(E ) X

is a projective bundle whose bers look like P


r
.
To describe the fan of P(E ), we rst give a new description of the fan of P
r
. In
R
r+1
, we use the standard basis e
0
, . . . , e
r
. The rst orthant Cone(e
0
, . . . , e
r
) has
r +1 facets
F
i
= Cone(e
0
, . . . , e
i
, . . . , e
r
), i = 0, . . . , r.
Now set N =Z
r+1
/Z(e
0
+ +e
r
). Then the images e
i
of e
i
sum to zero in N and
the images F
i
of F
i
give the fan for P
r
in N
R
.
The construction of P(E ) given in 7.0 involves taking the dual vector bundle.
Thus P(E ) =P(V
E
), where V
E
is the vector bundle whose sheaf of sections is
O
X

(D
0
) O
X

(D
r
).
The fan of V
E
is built from cones
Cone(u

+a
0
e
0
+ +a
r
e
r
[ (1)) +Cone(e
0
, . . . , e
r
)
338 Chapter 7. Projective Toric Morphisms
and their faces, as ranges over the cones . To get the fan for P(E ) =P(V
E
),
take and let F
i
be a facet of Cone(e
0
, . . . , e
r
). This gives the cone
Cone(u

+a
0
e
0
+ +a
r
e
r
[ (1)) +F
i
N
R
R
r+1
,
and one sees that
i
N
R
N
R
is the image of this cone under the projection map
N
R
R
r+1
N
R
N
R
.
Proposition 7.3.3. The cones
i
[ , i = 0, . . . , r and their faces form a fan

E
in N
R
N
R
whose toric variety X
E
is the projective bundle P(E ).
Proof. Consider the fan
0
in N
R
given by the F
i
and their faces. Also, for ,
let be the image of Cone(u

+a
0
e
0
+ +a
r
e
r
[ (1)) in N
R
N
R
. Then
one easily adapts the proof of Proposition 7.3.1 to show that the toric variety X
E
of
E
is a bration over X

with ber P
r
. Furthermore, working over an afne
open subset of X

, one sees that X


E
is obtained from V
E
by the process described
in 7.0. We leave the details as Exercise 7.3.4.
In practice, one usually replaces N = Z
r+1
/Z(e
0
+ +e
r
) with Z
r
and the
basis e
1
, . . . , e
r
. Then set e
0
=e
1
e
r
and we redene F
i
as
(7.3.2) F
i
= Cone(e
0
, . . . , e
i
, . . . , e
r
) R
r
and for a cone , redene
i
as
(7.3.3)
i
= Cone(u

+(a
1
a
0
)e
1
+ +(a
r
a
0
)e
r
[ (1)) +F
i
in N
R
R
r
. This way,
E
is a fan in N
R
R
r
. Here is a classic example.
Example 7.3.4. The fan for P
1
has minimal generators u
1
and u
0
= u
1
. Also
let O
P
1 (1) =O
P
1 (D
0
), where D
0
is the divisor corresponding to u
0
. Fix an integer
a 0 and consider
E =O
P
1 O
P
1 (a).
As above, we get a fan
E
in RR=R
2
. The minimal generators u
0
, u
1
live in the
rst factor. In the second factor, the vectors e
0
=

r
i=1
e
i
, e
1
, . . . , e
r
in the above
construction reduce to e
0
= e
1
, e
1
. Thus F
0
= Cone(e
1
) and F
1
= Cone(e
0
). We
will use u
1
, e
1
as the basis of R
2
.
The maximal cones for the fan of P
1
are = Cone(u
1
) and

= Cone(u
0
).
Then
E
has four cones:

0
= Cone(u
1
+(00)e
1
) +F
0
= Cone(u
1
, e
1
)

1
= Cone(u
1
+(00)e
1
) +F
1
= Cone(u
1
, e
1
)

0
= Cone(u
0
+(a0)e
1
) +F
0
= Cone(u
1
+ae
1
, e
1
)

1
= Cone(u
0
+(a0)e
1
) +F
1
= Cone(u
1
+ae
1
, e
1
).
This is the fan for the Hirzebruch surface H
a
. Thus
H
a
=P(O
P
1 O
P
1(a)).
7.3. Projective Bundles and Toric Varieties 339
Note also that the toric morphism H
a
P
1
constructed earlier is the projection
map for the projective bundle.
This example generalizes as follows.
Example 7.3.5. Given integers s, r 1 and 0 a
1
a
r
, consider the projec-
tive bundle
P(E ) =P(O
P
s O
P
s (a
1
) O
P
s (a
r
)).
The fan
E
of P(E ) has a nice description. We will work in R
s
R
r
, where R
s
has basis u
1
, . . . , u
s
and R
r
has basis e
1
, . . . , e
r
. Also set u
0
=

s
j=1
u
j
and e
0
=

s
i=1
e
i
. As usual, u
0
corresponds to the divisor D
0
of P
s
such that O
P
s (a
i
) =
O
P
s (a
i
D
0
).
The description (7.3.3) of the cones in uses generators of the form
(7.3.4) u

+(a
1
a
0
)e
1
+ +(a
r
a
0
)e
r
,
where the u

are minimal generators of the fan of the base of the projective bundle.
Here, the u

s are u
0
, . . . , u
s
. Since we are using the divisors 0, a
1
D
0
, . . . , a
r
D
0
, the
formula (7.3.4) simplies dramatically, giving minimal generators
u

= u
0
: v
0
= u
0
+a
1
e
1
+ +a
r
e
r
u

= u
j
: v
j
= u
j
, j = 1, . . . , s.
Since the maximal cones of P
s
are Cone(u
0
, . . . , u
j
, . . . , u
s
), (7.3.2) and (7.3.3) im-
ply that the maximal cones of are
Cone(v
0
, . . . , v
j
, . . . , v
s
) +Cone(e
0
, . . . , e
i
, . . . , e
r
)
for all j = 0, . . . , s and i = 0, . . . , r. It is also easy to see that the minimal generators
v
0
, . . . , v
s
, e
0
, . . . , e
r
have the following properties:
v
1
, . . . , v
s
, e
1
, . . . , e
r
form a basis of Z
s
Z
r
.
e
0
+ +e
r
= 0.
v
0
+ +v
s
= a
1
e
1
+ +a
r
e
r
.
The rst two bullets are clear, and the third follows from

s
j=0
u
j
= 0 and the
denition of the v
j
.
One also sees that X
E
= P(E ) is smooth of dimension s +r. Since
E
has
(s+1)+(r +1) =s+r +2 minimal generators, the description of the Picard group
given in 4.2 implies that
Pic(P(E )) Z
2
.
(Exercise 7.3.5). Also observe that v
0
, . . . , v
s
and e
0
, . . . , e
r
give primitive col-
lections of
E
. We will see below that these are the only primitive collections
of
E
. Furthermore, they are extremal in the sense of 6.4 and their primitive
relations generate the Mori cone of P(E ).
This is a very rich example!
340 Chapter 7. Projective Toric Morphisms
A Classication Theorem. Kleinschmidt [177] classied all smooth projective
toric varieties with Picard number 2, i.e., with Pic(X

) Z
2
. The rough idea is
that they are the toric projective bundles described in Example 7.3.5. Following
ideas of Batyrev [14], we will use primitive collections to obtain the classication.
We begin with some results from [14] about primitive collections. Recall from
6.4 that a primitive collection P=
1
, . . . ,
k
(1) gives the primitive relation
(7.3.5) u

1
+ +u

(1)
c

= 0, c

Q
>0
,
where is the minimal cone containing u

1
+ +u

k
. When X

is smooth
and projective, these primitive relations have some nice properties.
Proposition 7.3.6. Let X

be a smooth projective toric variety. Then:


(a) In the primitive relation (7.3.5), P(1) = and c

Z
>0
for all (1).
(b) There is a primitive collection P with primitive relation u

1
+ +u

k
= 0.
Proof. The c

are integral since is smooth. Let the minimal generators of be


u
1
, . . . , u

, so the primitive relation becomes


u

1
+ +u

k
= c
1
u
1
+ +c

.
To prove part (a), suppose for example that u

1
= u
1
. Then
u

2
+ +u

k
= (c
1
1)u
1
+c
2
u
2
+c

.
Note that u

2
, . . . , u

k
generate a cone of since P is a primitive collection. So the
above equation expresses an element of a cone of in terms of minimal generators
in two different ways. Since is smooth, these must coincide. To see what this
means, we consider two cases:
c
1
> 1. Then u

2
, . . . , u

k
=u
1
, u
2
, . . . , u

, so that u

i
= u
1
for some i > 1.
This is impossible since u

1
= u
1
.
c
1
=1. Then u

2
, . . . , u

k
=u
2
, . . . , u

. Since u

1
=u
1
, we obtain P(1),
which is impossible since P is a primitive collection.
Since c
1
must be positive, we conclude that u

1
=u
1
leads to a contradiction. From
here, it is easy to see that P(1) =.
Turning to part (b), let be the support function of an ample divisor on X

.
Thus is strictly convex. Since is complete, we can nd an expression
(7.3.6) b
1
u

1
+ +b
s
u
s
= 0
such that b
1
, . . . , b
s
are positive integers. Note that u

1
, . . . , u
s
cannot lie in a cone
of . By strict convexity and Lemma 6.1.13, it follows that
(7.3.7) 0 = (0) = (b
1
u

1
+ +b
s
u
s
) > b
1
(u

1
) + +b
s
(u
s
).
Pick a relation (7.3.6) so that the right-hand side is as big as possible.
7.3. Projective Bundles and Toric Varieties 341
The set u

1
, . . . , u
s
is not contained in a cone of and hence has a subset
that is a primitive collection. By relabeling, we may assume that u

1
, . . . , u

k
,
k s, is a primitive collection. Using (7.3.6) and the primitive relation (7.3.5), we
obtain the nonnegative relation

(1)
c

+
k

i=1
(b
i
1)u

i
+
s

i=k+1
b
i
u

i
= 0.
Since is linear on and strictly convex,

(1)
c

(u

) =


(1)
c

i=1
u

>
k

i=1
(u

i
),
which implies that

(1)
c

(u

) +
k

i=1
(b
i
1)(u

i
) +
s

i=k+1
b
i
(u

i
)
>
k

i=1
(u

i
) +
k

i=1
(b
i
1)(u

i
) +
s

i=k+1
b
i
(u

i
) =
s

i=1
b
i
(u

i
).
This contradicts the maximality of the right-hand side of (7.3.7), unless k = s and
b
1
= = b
k
= 1, in which case we get the desired primitive collection.
We now prove Kleinschmidts classication theorem.
Theorem 7.3.7. Let X

be a smooth projective toric variety with Pic(X

) Z
2
.
Then there are integers s, r 1, s +r = dim X

and 0 a
1
a
r
with
X

O
P
s O
P
s (a
1
) O
P
s (a
r
)

.
Proof. Let n = dim X

. Then Pic(X

) Z
2
and Theorem 4.2.1 imply that (1)
has n+2 elements. We recall two facts about divisors D on X

:
If D is nef and (n), then D

where a

= 0 for (1) and


a

0 for / (1).
If D 0 and D 0, then D = 0 since X

is complete.
The rst bullet was proved in (6.4.10), and the second is an easy consequence of
Propositions 4.0.16 and 4.3.8.
By assumption, X

has an ample divisor D which lies in the interior of the


nef cone Nef(X

). Changing D if necessary, we can assume that D is effective


and [D] Pic(X

)
R
is not a scalar multiple of any [D

] for (1). The line


determined by [D] divides Pic(X

)
R
R
2
into closed half-planes H
+
and H

.
Then dene
P = (1) [ [D

] H
+

Q = (1) [ [D

] H

.
342 Chapter 7. Projective Toric Morphisms
Note that PQ = (1), and PQ = by our choice of D. We claim that
(7.3.8)
(n) =
,
[ P,

Q, where

,
= Cone(u

[ (1) `,

).
To prove this, rst take (n). Since [(1)[ = n and [(1)[ = n+2, we have
(7.3.9) (1) = (1) ,

.
Applying the rst bullet above to D and , we get [D] = a[D

] +b[D

] where
a, b > 0 since [D] is a multiple of neither [D

] nor [D

]. It follows that [D

] and
[D

] lie on opposite sides of the line determined by [D]. We can relabel so that
P and

Q, and then has the desired form by (7.3.9).


For the converse, take P and

Q. Since Pic(X

)
R
R
2
, we can nd a
linear dependence
a
0
[D

] +b
0
[D

] +c
0
[D] = 0, a
0
, b
0
, c
0
Z not all 0.
We can assume that a
0
, b
0
0 since [D

] and [D

] lie on opposite sides of the


line determined by [D]. Note also that c
0
< 0 by the second bullet above, and then
a
0
, b
0
> 0 by our choice of D. It follows that D

= a
0
D

+b
0
D

is ample. In
Exercise 7.3.6 you will show that
X

`Supp(D

) = X

`(D

)
is the nonvanishing set of a global section of O
X

(D

) and hence is afne. This set


is also torus-invariant and hence is an afne toric variety. Thus it must be U

for
some . In other words,
X

=U

.
Since U

(D

) = , the Orbit-Cone correspondence (Theorem 3.2.6) im-


plies that satsises (7.3.9) and hence gives an element of (n). This completes
the proof of (7.3.8).
An immediate consequence of this description of (n) is that P and Q are
primitive collections. Be sure you understand why. It is also true that P and Q
are the only primitive collections of . To prove this, suppose that we had a third
primitive collection R. Then P R, so there is P` R, and similarly there is

Q` R since Q P. By (7.3.8), the rays of R all lie in


,
(n), which
contradicts the denition of primitive collection.
Since X

is projective and smooth, Proposition 7.3.6 guarantees that has a


primitive collection whose elements sum to zero. We may assume that P is this
primitive collection. Let [P[ = r +1 and [Q[ = s +1, so r, s 1 since primitive
collections have at least two elements, and r +s = n since [P[ +[Q[ = n+2.
Now rename the minimal generators of the rays in P as e
0
, . . . , e
r
. Thus
e
0
+ +e
r
= 0.
7.3. Projective Bundles and Toric Varieties 343
The next step is to rename the minimal generators of the rays in P as v
0
, . . . , v
s
.
Proposition 7.3.6 implies that

s
j=0
v
j
lies in a cone whose rays lie in the
complement of Q, which is P. Since P is a primitive collection, must omit at
least one element of P, which we may assume to be the ray generated by e
0
. Then
the primitive relation of Q can be written
v
0
+ +v
s
= a
1
e
1
+ +a
r
e
r
,
and by further relabeling, we may assume 0 a
1
a
r
. Finally, observe that
v
1
, . . . , v
s
, e
1
, . . . , e
r
generate a maximal cone of by (7.3.8). Since is smooth,
it follows that these r +s vectors form a basis of N. Comparing all of this to
Example 7.3.5, we conclude that the toric variety of is the projective bundle
P

O
P
s O
P
s (a
1
) O
P
s (a
r
)

.
The classication result proved in [177] is more general than the one given in
Theorem 7.3.7. By using a result from [191] on sphere triangulations with few
vertices, Kleinschmidt does not need to assume that X

is projective. Another
proof of Theorem 7.3.7 that does not assume projective can be found in [14, Thm.
4.3]. We should also mention that (7.3.8) can be proved using the Gale transforms
discussed in [93, II.46] and [281, Ch. 6]. We will explain this is 15.2.
Exercises for 7.3.
7.3.1. Here you will supply some details needed to prove Theorem 7.3.1.
(a) In the proof we constructed a map U
e
C. Show that this map is
(m,1)
, where

D
(u) ='m

, u` for u .
(b) Given cones , , the transition map from U

C U

C to U

C
U

C is given by (u, t) (u, g

(u)t). Prove that g

=
m m
.
7.3.2. In Example 7.3.2, we study the rank 1 vector bundle V P
n
whose sheaf of sections
is O
P
n (1). Let be the fan of V in R
n+1
.
(a) Prove that e
1
, . . . , e
n+1
, e
1
e
n
+e
n+1
are the minimal generators of .
(b) Prove that is the normal fan of
P = Conv(0, e
1
, . . . , e
n
) +Cone(e
n+1
, e
1
+e
n+1
, . . . , e
n
+e
n+1
).
(c) The example constructs a morphismV P
n
C
n+1
. Prove that the image of this map
is dened by x
i
y
j
= x
j
y
i
and explain how this relates to Example 6.0.19.
7.3.3. Consider the locally free sheaf (7.3.1) and the cones N
R
R
r
dened in the
discussion following (7.3.1). Prove that these cones and their faces give a fan in N
R
R
r
whose toric variety is the vector bundle with (7.3.1) as sheaf of sections.
7.3.4. Complete the proof of Proposition 7.3.3.
7.3.5. Let P(E ) P
s
be the toric projective bundle constructed in Example 7.3.5. Prove
that Pic(P(E )) Z
2
.
7.3.6. Let D be an ample effective divisor on a complete normal variety X. The goal of
this exercise is to prove that X ` Supp(D) is afne.
344 Chapter 7. Projective Toric Morphisms
(a) Assume that D is very ample. Let s (X, O
X
(D)) be nonzero and consider the
nonvanishing set of s dened by U = s X [ s(x) = 0. Prove that U is afne.
Hint: Show that a basis s = s
0
, s
1
, . . . , s
m
of (X, O
X
(D)) gives a closed embedding
X P
m
. Let P
m
have homogeneous coordinates x
0
, . . . , x
m
and regard X as a subset
of P
m
. Prove that U = X U
0
, where U
0
P
m
is where x
0
= 0.
(b) Explain why part (a) remains true when D is ample but not necessarily very ample.
Hint: s
k
(X, O
X
(kD)).
(c) Since D is effective, 1 (X, O
X
(D)) is a global section. Prove that the nonvanish-
ing set of this global section is X ` Supp(D). Hint: For s (X, O
X
(D)), recall the
denition of div
0
(s) given in 4.0.
Parts (b) and (c) imply that X ` Supp(D) is afne when D is ample, as desired. Note also
that part (b) is a special case of Proposition 7.A.6.
7.3.7. By Example 2.3.16, the rational normal scroll S
a,b
is the toric variety of
P
a,b
= Conv(0, ae
1
, e
2
, be
1
+e
2
) R
2
,
where a, b N satisfy 1 a b, and S
a,b
H
ba
by Example 3.1.16. Thus rational
normals scroll are Hirzebruch surfaces. Here you will explore an n-dimensional analog.
Take integers 1 d
0
d
1
d
n1
. Then P
d0,...,dn1
is the lattice polytope in R
n
having the 2n lattice points
0, d
0
e
1
, e
2
, e
2
+d
1
e
1
, e
3
, e
3
+d
2
e
1
, . . . , e
n
, e
n
+d
n1
e
1
as vertices. The toric variety of P
d0,...,dn1
is denoted S
d0,...,dn1
.
(a) Explain why P
d0,...,dn1
is a truncated prism whose base in 0R
n1
is the standard
simplex
n1
, and above the vertices of
n1
there are edges of lengths d
0
, . . . , d
n1
.
Here, above means the e
1
direction. Draw a picture when n = 3.
(b) Prove that S
d0,...,dn1
P(O
P
1 (d
0
) O
P
1 (d
n1
)).
(c) S
d0,...,dn1
is smooth by part (b), so that P
d0,...,dn1
is very ample and hence gives a
projective embedding of S
d0,...,dn1
. Explain how this embedding consists of n embed-
dings of P
1
such that for each point p P
1
, the resulting n points in projective space
are connected by an (n 1)-dimensional plane that lies in S
d0,...,dn1
.
(d) Explain how part (c) relates to the scroll discussion in Example 2.3.16.
(e) Show that the (n1)-dimensional plane associated to p P
1
in part (c) is the ber of
the projective bundle P(O
P
1 (d
0
) O
P
1 (d
n1
)) P
1
.
7.3.8. Consider the toric variety P(E ) constructed in Example 7.3.5.
(a) Prove that P(E ) is projective. Hint: Proposition 7.0.5.
(b) Show that P(E ) P(O
P
s (1) O
P
s (a
1
+1) O
P
s (a
r
+1)). Hint: Part (b) of
Lemma 7.0.8.
(c) Find a lattice polytope in R
s
R
r
whose toric variety is P(E ). Hint: In the polytope
of Exercise 7.3.7, each vertex of 0
n1
RR
n1
is attached to a line segment
in the normal direction. Also observe that a line segment is a multiple of
1
. Adapt
this by using 0
r
R
s
R
r
as base and then, at each vertex of
r
, attach a
positive mutliple of
s
in the normal direction.
Appendix: More on Projective Morphisms 345
7.3.9. Let X

be a projective toric variety and let D


0
, . . . , D
r
be torus-invariant ample divi-
sors on X

. Each D
i
gives a lattice polytope P
i
= D
Pi
whose normal fan is . Prove that
the projective bundle P(O
X
(D
0
) O
X
(D
r
)) is the toric variety of the polytope in
N
R
R
Conv(P
0
0 P
1
e
1
P
r
e
r
).
Hint: If you get stuck, see [61, Sec. 3]. Do you see how this relates to Exercise 7.3.8?
7.3.10. Use primitive collections to showthat P
n
is the only smooth projective toric variety
with Picard number 1.
Appendix: More on Projective Morphisms
In this appendix, we discuss some technical details related to projective morphisms.
Ampleness. A comprehensive treatment of ampleness appears in Volume II of

El ements
de g eom etrie alg ebrique (EGA) by Grothendieck and Dieudonn e [127]. The results we
need from EGA are spread out over several sections. Here we collect the denitions and
theorems we will use in our discusion of ampleness.
1
Denition 7.A.1. A line bundle L on a variety X is absolutely ample if for every coherent
sheaf F on X, there is an integer k
0
such that F
OX
L
k
is generated by global sections
for all k k
0
.
By [127, (4.5.5)], this is equivalent to what EGA calls ample in [127, (4.5.3)]. We
use the name absolutely ample to prevent confusion with Denition 6.1.9, where ample
is dened for line bundles on complete varieties. Here is another denition from EGA.
Denition 7.A.2. Let f : X Y be a morphism. A line bundle L on X is relatively ample
with respect to f if Y has an afne open cover U
i
such that for every i, L[
f
1 (U
i
) is
absolutely ample on f
1
(U
i
).
This is [127, (4.6.1)]. When mapping to an afne variety, relatively ample and abso-
lutely ample coincide. More precisely, we have the following result from [127, (4.6.6)].
Proposition 7.A.3. Let f : X Y be a morphism, where Y is afne, and let L be a line
bundle on X. Then:
L is relatively ample with respect to f L is absolutely ample.
The reader should be warned that in EGA, relatively ample with respect to f and
f -ample are synonyms. In this text, they are slightly different, since relatively ample
with respect to f refers to Denition 7.A.2 while f -ample refers to Denition 7.2.5.
Fortunately, they coincide when the map f is proper.
Theorem 7.A.4. Let f : X Y be a proper morphism and L a line bundle on X. Then
the following are equivalent:
(a) L is relatively ample with respect to f in the sense of Denition 7.A.2.
1
The theory developed in EGA applies to very general schemes. The varieties and morphisms appearing in
this book are nicely behavedthe varieties are quasi-compact and noetherian, the morphisms are of nite type,
and coherent is equivalent to quasicoherent of nite type. Hence most of the special hypotheses needed for some
of the results in [127] are automatically true in our situation.
346 Chapter 7. Projective Toric Morphisms
(b) L is f -ample in the sense of Denition 7.2.5.
(c) There is an integer k > 0 such that f is projective with respect to L
k
in the sense of
Denition 7.0.3.
Proof. First observe that (b) and (c) are equivalent by Denition 7.2.5. Now suppose that
f is projective with respect to L
k
. Then there is an afne open covering U
i
of Y such
that for each i, there is a nite-dimensional subspace W (U
i
, L
k
) that gives a closed
embedding of f
1
(U
i
) into U
i
P(W

) for each i.
The locally free sheaf E = W

C
O
Ui
is the sheaf of sections of the trivial vector
bundle U
i
W

U
i
. This gives the projective bundle P(E ) = U
i
P(W

), so that we
have a closed embedding
f
1
(U
i
) P(E ).
By denition [127, (4.4.2)], L
k
[
f
1
(Ui)
is very ample for f [
f
1
(Ui)
. Then [127, (4.6.9)]
implies that L[
f
1
(Ui)
is relatively ample with respect to f [
f
1
(Ui)
, and hence absolutely
ample by Proposition 7.A.3. Then L is relatively ample with respect to f by Deni-
tion 7.A.2.
Finally, suppose that L is relatively ample with respect to f and let U
i
be an afne
open covering of Y. Then [127, (4.6.4)] implies that L[
f
1
(Ui )
is relatively ample with
respect to f [
f
1
(Ui)
. By [127, (4.6.9)], L
k
[
f
1
(Ui)
is very ample for f [
f
1
(Ui)
, which by
denition [127, (4.4.2)] means that f
1
(U
i
) can be embedded in P(E ) for a coherent sheaf
E on U
i
. Then the proof of [273, Thm. 5.44] shows how to nd nitely many sections of
L
k
over f
1
(U
i
) that give a suitable embedding of f
1
(U
i
) into U
i
P(W

).
In EGA [127, (5.5.2)], the denition of when a morphism f : X Y is projective
involves two equivalent conditions stated in [127, (5.5.1)]. The rst condition uses the
projective bundle P(E ) of a coherent sheaf E on Y, and the second uses Proj(S), where
S is a quasicoherent graded O
Y
-algebra such that S
1
is coherent and generates S. By
[127, (5.5.3)], projective is equivalent to proper and quasiprojective, and by the dention
of quasiprojective [127, (5.5.1)], this means that X has a line bundle relatively ample with
respect to f . Hence Theorem 7.A.4 shows that the denition of projective morphism given
in EGA is equivalent to Denition 7.0.3.
We close with two further results about projective morphisms. Proofs can be found in
[127, (4.6.4)] and [127, (5.5.7)] respectively.
Proposition 7.A.5. Let f : X Y be a proper morphism and L a line bundle on X. Given
an afne open cover U
i
of Y, the following are equivalent:
(a) L is f -ample.
(b) For every i, L[
f
1
(Ui)
is f [
f
1
(Ui )
-ample.
Proposition 7.A.6. Let f : X Y be a projective morphism with Y afne and let L be an
f -ample line bundle on X. Then:
(a) Given a global section s (X, L), let X
s
X be the open subset where s is nonvan-
ishing. Then X
s
is an afne open subset of X.
(b) There is an integer k
0
such that L
k
is generated by global sections for all k k
0
.
Chapter 8
The Canonical Divisor
of a Toric Variety
8.0. Background: Reexive Sheaves and Differential Forms
This chapter will study the canonical divisor of a toric variety. The theory devel-
oped in Chapters 6 and 7 dealt with Cartier divisors and line bundles. As we will
see, the canonical divisor of a normal toric variety is a Weil divisor that is not
necessarily Cartier. We will also study the sheaves associated to Weil divisors.
Reexive Sheaves. A Weil divisor D on a normal variety X gives the sheaf O
X
(D)
dened by
(U, O
X
(D)) =f C(X)

[ (div( f ) +D)[
U
00.
Our rst task is to characterize these sheaves.
Recall that the dual of a sheaf of O
X
-modules F is F

= Hom
O
X
(F, O
X
).
We say that F is reexive if the natural map
F F

is an isomorphism. It is easy to see that locally free sheaves are reexive. Here are
some properties of reexive sheaves.
Proposition 8.0.1. Let F be a coherent sheaf on a normal variety X and consider
the inclusion j : U
0
X where U
0
is open with codim(X `U
0
) 2. Then:
(a) F

and hence F

are reexive.
(b) If F is reexive, then F j

(F[
U
0
).
(c) If F[
U
0
is locally free, then F

(F[
U
0
).
347
348 Chapter 8. The Canonical Divisor of a Toric Variety
Proof. Recall from 4.0 that the direct image j

G of a sheaf G on U
0
is dened by
(U, j

G) = (U U
0
, G) for U X open.
Parts (a) and (b) of the proposition are proved in [132, Cor. 1.2 and Prop. 1.6].
For part (c), we rst observe that restriction is compatible with taking the dual, i.e.,
(G

)[
U
0
= (G[
U
0
)

for any coherent sheaf G on X. Then


F

((F

)[
U
0
) = j

((F[
U
0
)

) j

(F[
U
0
),
where the rst isomorphism follows from parts (a) and (b), and the last follows
since F[
U
0
is locally free and hence reexive.
Later in the section we will study the sheaf
p
X
of p-forms on X. This sheaf is
locally free when X is smooth. For X normal, however,
p
X
may be badly behaved,
though it is locally free on the smooth locus of X. Hence we can use part (c) of
Proposition 8.0.1 to create a reexive version of
p
X
.
For more on reexive sheaves, the reader should consult [132] and [235].
Reexive Sheaves of Rank One. We rst dene the rank of a coherent sheaf on an
irreducible variety X. Recall that K
X
is the constant sheaf on X given by C(X).
Denition 8.0.2. Given a F coherent sheaf on irreducible variety X, the global
sections of F
O
X
K
X
form a nite-dimensional vector space over C(X) whose
dimension is the rank of F.
For a locally free sheaf, the rank is just the rank of the associated vector bundle.
Other properties of the rank will be studied in Exercise 8.0.1.
In the smooth case, reexive sheaves of rank 1 are easy to understand.
Proposition 8.0.3. On a smooth variety, a coherent sheaf of rank 1 is reexive if
and only if it is a line bundle.
This is proved in [132, Prop. 1.9]. We now have all of the tools needed to
characterize which coherent sheaves on a normal variety come from Weil divisors.
Theorem 8.0.4. Let L be a coherent sheaf on a normal variety X. Then the
following are equivalent:
(a) L is reexive of rank 1.
(b) There is an open subset j : U
0
X such that codim(X `U
0
) 2, L[
U
0
is a
line bundle on U
0
, and L j

(L[
U
0
).
(c) L O
X
(D) for some Weil divisor D on X.
Proof. (a) (b) Since X is normal, its singular locus Y = Sing(X) has codimen-
sion at least two in X by Proposition 4.0.17. Then U
0
= X `Y is smooth, which
implies that L[
U
0
is a line bundle by Proposition 8.0.3. Hence L j

(L[
U
0
) by
Proposition 8.0.1.
8.0. Background: Reexive Sheaves and Differential Forms 349
(b) (c) The line bundle L[
U
0
can be written as O
U
0
(E) for some Cartier
divisor E =

i
a
i
E
i
on U
0
. Consider the Weil divisor D =

i
a
i
D
i
, where D
i
is the
Zariski closure of E
i
in X. Given f C(X)

, note that
div( f ) +D 0 (div( f ) +D)[
U
0
0
since codim(X `U
0
) 2, and the same holds over any open set of X. Combining
this with E = D[
U
0
, we obtain
O
X
(D) j

O
U
0
(E) = j

(L[
U
0
) L.
(c) (a) The proof of (b) (c) shows that O
X
(D) j

(O
X
(D)[
U
0
). But
codim(X `U
0
) 2, and O
X
(D)[
U
0
=O
U
0
(D[
U
0
) is locally free since U
0
is smooth.
Thus j

(O
X
(D)[
U
0
) O
X
(D)

by Proposition 8.0.1, so O
X
(D) O
X
(D)

is
reexive and has rank 1 since it is a line bundle on U
0
.
Tensor Products and Duals. Given Weil divisors D, E on a normal variety X, the
map f g f g denes a sheaf homomorphism
(8.0.1) O
X
(D)
O
X
O
X
(E) O
X
(D+E).
This is an isomorphism when D or E is Cartier but may fail to be an isomorphism
in general.
Example 8.0.5. Consider the afne quadric cone X = V(y
2
xz) C
3
. From
examples in previous chapters, we know that this is a normal toric surface. The
line L = V(y, z) gives a Weil divisor that is not Cartier, though 2L is Cartier (this
follows from Example 4.2.3). The coordinate ring of X is R = C[x, y, z]/'y
2
xz`.
Let x, y, z denote the images of the variables in R. In Exercise 8.0.2 you will show
the following:
(X, O
X
(L)) is the ideal 'y, z` R.
(X, O
X
(2L)) is the ideal 'z` R (principal since 2L is Cartier).
On global sections, the image of the map
O
X
(L)
O
X
O
X
(L) O
X
(2L)
is 'y, z`
2
, which is a proper subset of (X, O
X
(2L)) ='z`.
It follows that O
X
(L)
O
X
O
X
(L) O
X
(2L).
If we apply (8.0.1) when E =D, we get a map
O
X
(D)
O
X
O
X
(D) O
X
,
which in turn induces a map
(8.0.2) O
X
(D) O
X
(D)

.
As noted in 6.0, this is an isomorphism when D is Cartier. In general, we have the
following result about the maps (8.0.1) and (8.0.2).
350 Chapter 8. The Canonical Divisor of a Toric Variety
Proposition 8.0.6. Let D, E be Weil divisors on a normal variety X. Then (8.0.1)
induces an isomorphism
(O
X
(D)
O
X
O
X
(E))

O
X
(D+E).
Furthermore, (8.0.2) is an isomorphism, i.e.,
O
X
(D) O
X
(D)

.
Proof. The rst isomorphism follows from Proposition 8.0.1 since O
X
(D+E) is
reexive and (8.0.1) is an isomorphism on the smooth locus of X. The second
isomorphism follows similarly since both sheaves are reexive and (8.0.2) is an
isomorphism on the smooth locus.
Divisor Classes. Recall that Weil divisors D and E on X are linearly equivalent,
written D E, if D = E +div( f ) for some f C(X)

.
Proposition 8.0.7. Let X be a normal variety.
(a) If D and E are Weil divisors on X, then
O
X
(D) O
X
(E) D E.
(b) If D is a Weil divisor on X, then
D is Cartier O
X
(D) is a line bundle.
Proof. Linearly equivalent divisors give isomorphic sheaves by Proposition 4.0.29.
Conversely, O
X
(D) O
X
(E) implies
O
X
(D)
O
X
O
X
(E) O
X
(E)
O
X
O
X
(E).
Taking the double dual and using Proposition 8.0.6, we get O
X
(DE) O
X
. From
here, showing that D E follows exactly as in the proof of Proposition 6.0.22.
One direction of part (b) was proved in Chapter 6 (see Proposition 6.0.17 and
Theorem 6.0.18). Conversely, if O
X
(D) is a line bundle on X, then Theorem 6.0.20
shows that O
X
(D) O
X
(E) for some Cartier divisor E. Thus D E by part (a).
Then we are done since any divisor linearly equivalent to a Cartier divisor is Cartier
by Exercise 4.0.5.
In Chapter 4 we dened the class group Cl(X) and Picard group Pic(X) in
terms of Weil and Cartier divisors. Then, in Chapter 6, we reinterpreted Pic(X)
as the group of isomorphism classes of line bundles, where the group operation
was tensor product and the inverse was the dual. We can now reinterpret Cl(X) as
the group of isomorphism classes of reexive sheaves of rank 1, where the group
operation is the double dual of the tensor product and the inverse is the dual. This
follows immediately from Propositions 8.0.6 and 8.0.7.
8.0. Background: Reexive Sheaves and Differential Forms 351
K ahler Differentials. In order to give an algebraic denition of differential forms
on a variety, we begin with the case of a C-algebra.
Denition 8.0.8. Let R be a C-algebra. The module of K ahler differentials of R
over C, denoted
R/C
, is the R-module generated by the formal symbols df for
f R, modulo the relations
(a) d(c f +g) = cdf +dg for all c C, f , g R.
(b) d( f g) = f dg+gdf for all f , g R.
Example 8.0.9. If R =C[x
1
, . . . , x
n
], then

R/C

i=1
Rdx
i
.
This follows because the relations dening
R/C
imply df =

n
i=1
f
x
i
dx
i
for all
f R (Exercise 8.0.3).
A C-algebra homomorphism R S induces a natural homomorphism

R/C

S/C
.
When we regard
S/C
as an R-module, we obtain a homomorphism of S-modules
S
R

R/C

S/C
.
Here is a case when this map is easy to understand. See [195, Thm. 25.2] for a
proof.
Proposition 8.0.10. Let R S be a surjection of C-algebras with kernel I. Then
there is an exact sequence of S-modules
I/I
2
S
R

R/C

S/C
0,
where [ f ] I/I
2
maps to 1df S
R

R/C
.
Example 8.0.11. Let R = C[x
1
, . . . , x
n
] and S = R/I, where I = ' f
1
, . . . , f
s
`. The
generators of I give a surjection R
s
I and hence a surjection S
s
I/I
2
. Combin-
ing this with Proposition 8.0.10 and Example 8.0.9, we obtain an exact sequence
S
s

S
n

S/C
0,
where is given by the reduction of the ns Jacobian matrix
(8.0.3)

f
1
x
1

fs
x
1
.
.
.
.
.
.
f
1
xn

fs
xn

modulo the ideal I (Exercise 8.0.4). This presentation of


S/C
is very useful for
computing examples.
352 Chapter 8. The Canonical Divisor of a Toric Variety
K ahler differentials also behave well under localization, as you will prove in
Exercise 8.0.5.
Proposition 8.0.12. Let R
f
be the localization of R at a non-nilpotent element
f R. Then
R
f
/C

R/C
R
f
.
Cotangent and Tangent Sheaves. Now we globalize Denition 8.0.8.
Denition 8.0.13. Let X be a variety. The cotangent sheaf
1
X
is the sheaf of
O
X
-modules dened via

1
X
(U) =
O
X
(U)/C
on afne open sets U. The tangent sheaf T
X
is the dual sheaf
T
X
= (
1
X
)

=Hom
O
X
(
1
X
, O
X
).
The reason for the superscript in the notation for the cotangent sheaf will be-
come clear later in this chapter. In Exercise 8.0.6 you will use Example 8.0.11 and
Proposition 8.0.12 to show that
1
X
is a coherent sheaf. See [131, II.8] for a slightly
different approach to dening the sheaf
1
X
, and [131, II.8, Comment 8.9.2] for the
connection between these methods.
When U = Spec(R) is an afne open of X, the denition of the tangent sheaf
implies that
T
X
(U) = Hom
R
(
R/C
, R).
This can also be described in terms of derivationssee Exercises 8.0.7 and 8.0.8.
When X is smooth, these sheaves are nicely behaved, as shown by the follow-
ing result from [131, Thm. II.8.15].
Theorem 8.0.14. A variety X is smooth if and only if
1
X
is locally free. When this
happens,
1
X
and T
X
are locally free sheaves of rank n, n = dim X.
In the smooth toric case, it is easy to see that the cotangent sheaf is locally free.
Example 8.0.15. A smooth cone N
R
R
n
of dimension r gives the afne toric
variety
U

C
r
(C

)
nr
C
n
.
Then Example 8.0.9 and Proposition 8.0.12 imply that
1
U
is locally free of rank n.
It follows immediately that
1
X

is locally free for any smooth toric variety X

.
We know from Chapter 6 that a locally free sheaf is the sheaf of sections of a
vector bundle. When X is smooth, the vector bundles corresponding to
1
X
and T
X
are called the cotangent bundle and tangent bundle respectively.
8.0. Background: Reexive Sheaves and Differential Forms 353
Example 8.0.16. We construct the cotangent bundle for P
2
. Recall that P
2
has a
covering by the afne open sets
U

0
Spec(C[x, y])
U

1
Spec(C[yx
1
, x
1
])
U

2
Spec(C[xy
1
, y
1
]).
where
0
,
1
,
2
are the maximal cones in the usual fan for P
2
.
Let C
2
=Spec(R) for R =C[x, y]. The module
R/C
is a free R-module of rank
2 with generators dx, dy by Example 8.0.9. Thus a 1-form on C
2
may be written
uniquely as f
1
dx + f
2
dy, where f
i
R. To generalize this to P
2
, we require that
after changing coordinates, dx and dy transform via the Jacobian matrix described
in Example 8.0.11. More precisely, the matrix for the transition function
i j
will
be the Jacobian of the map U

j
U

i
.
On U

2
, the coordinates (a
1
, a
2
) are represented in terms of the (x, y) coordi-
nates on U

0
as (xy
1
, y
1
), yielding

20
=

1/y x/y
2
0 1/y
2

.
Next, we compute
12
. Things get messy if we keep everything in (x, y) coordi-
nates, so we rst translate to coordinates (a
1
, a
2
) on U

2
, and then translate back.
On U

2
we identify (a
1
, a
2
) with (xy
1
, y
1
). Then U

1
has coordinates
(yx
1
, x
1
) =

1
a
1
,
a
2
a
1

.
So in terms of (a
1
, a
2
), we have

12
=

1/a
2
1
0
a
2
/a
2
1
1/a
1

.
Rewriting this in terms of (x, y) yields

12
=

y
2
/x
2
0
y/x
2
y/x

.
Finally, computing
10
directly, we obtain

10
=

y/x
2
1/x
1/x
2
0

.
A check shows that
10
=
12

20
. Similar computations show that the compati-
bility criteria are satised for all i, j, k, i.e.,

ik
=
i j

jk
.
Since det(
i j
) is invertible on U

i
U

j
, the same is true for
i j
. Hence we obtain
a rank 2 vector bundle on P
2
whose sheaf of sections is
1
P
2
.
354 Chapter 8. The Canonical Divisor of a Toric Variety
Relation with the Zariski Tangent Space. The denition of the tangent sheaf
T
X
seems far removed from the denition of the Zariski tangent space T
p
(X) =
Hom
C
(m
X,p
/m
2
X,p
, C) given in Chapter 1. Here we explain (without proof) the
connection.
The stalk (T
X
)
p
of the tangent sheaf at p X can be described as follows. The
stalk of
1
X
at p is the module of K ahler differentials
(
1
X
)
p
=
O
X,p
/C
,
where O
X,p
is the local ring of X at p. Since O
X,p
/m
X,p
C and
C/C
= 0 (easy
to check), the exact sequence of Proposition 8.0.10 gives a surjection
m
X,p
/m
2
X,p

O
X,p
/C

O
X,p
C
which is an isomorphism of vector spaces over C by [131, Prop. II.8.7]. Since T
X
is dual to
1
X
, taking the dual of the above isomorphism gives
(8.0.4) (T
X
)
p

O
X,p
C Hom
C
(m
X,p
/m
2
X,p
, C) = T
p
(X).
This omits many details but should help you understand why T
X
is the correct
denition of tangent sheaf.
Example 8.0.17. Let V C
n
be dened by I = I(V) =' f
1
, . . . , f
s
` C[x
1
, . . . , x
n
].
The coordinate ring of V is S = C[x
1
, . . . , x
n
]/I, so that Example 8.0.11 gives the
exact sequence
S
s
S
n

S/C
0.
Now take p V and tensor with O
V,p
to obtain the exact sequence
O
s
V,p
O
n
V,p

O
V,p
/C
0
(Exercise 8.0.9). If we tensor this with C and dualize, (8.0.4) and the isomorphism

O
X,p
/C

O
X,p
C m
X,p
/m
2
X,p
give the exact sequence
0 T
p
(V) C
n

C
s
,
where comes from the s n Jacobian matrix

f
i
x
j
(p)

(Exercise 8.0.9). This


explains the description of T
p
(V) given in Lemma 1.0.6.
Conormal and Normal Sheaves. Given a closed subvariety i : Y X, it is natural
to ask how their cotangent sheaves relate. We begin with the exact sequence
0 I
Y
O
X
i

O
Y
0,
which we write more informally as
0 I
Y
O
X
O
Y
0.
The quotient sheaf I
Y
/I
2
Y
has a natural structure as a sheaf of O
Y
-modules, as
does F
O
X
O
Y
for any sheaf F of O
X
-modules. The following basic result is
proved in [131, Prop. II.8.12 and Thm. II.8.17].
8.0. Background: Reexive Sheaves and Differential Forms 355
Theorem 8.0.18. Let Y be a closed subvariety of a variety X. Then:
(a) There is an exact sequence of O
Y
-modules
I
Y
/I
2
Y

1
X

O
X
O
Y

1
Y
0.
(b) If X and Y are smooth, then this sequence is also exact on the left and I
Y
/I
2
Y
is locally free of rank equal to the codimension of Y.
Note that part (a) of this theorem is a global version of Proposition 8.0.10. We
call I
Y
/I
2
Y
the conormal sheaf of Y in X and call its dual
N
Y/X
= (I
Y
/I
2
Y
)

=Hom
O
Y
(I
Y
/I
2
Y
, O
Y
)
the normal sheaf of Y in X. When X andY are smooth, we can dualize the sequence
appearing in Theorem 8.0.18 to obtain the exact sequence
0 T
Y
T
X

O
X
O
Y
N
Y/X
0.
The vector bundle associated to N
Y/X
is the normal bundle of Y in X. Then the
above sequence says that when the tangent bundle of X is restricted to the subvari-
ety Y, it contains the tangent bundle of Y with quotient given by the normal bundle.
This is the algebraic analog of what happens in differential geometry, where the
normal bundle is the orthogonal complement of the tangent bundle of Y.
Differential Forms. We call
1
X
the sheaf of 1-forms, and we dene the sheaf of
p-forms to be the wedge product

p
X
=

1
X
.
For any sheaf F of O
X
-modules, the exterior power

p
F is the sheaf associated
to the presheaf which to each open set U assigns the O
X
(U)-module

p
F(U).
Example 8.0.19. For C
n
,
1
C
n is the sheaf associated to the free R-module
R/C
=

n
i=1
Rdx
i
, R =C[x
1
, . . . , x
n
]. Then
p
C
n is the sheaf associated to

R/C
=

1i
1
<<ipn
Rdx
i
1
dx
ip
.
It follows that
p
C
n is free of rank

n
p

.
More generally, Theorem 8.0.14 implies that
p
X
is locally free of rank

n
p

when X is smooth of dimension n. In particular,


n
X
is a line bundle in this case.
Zariski p-Forms and the Canonical Sheaf . For a normal variety X, the sheaf of
p-forms
p
X
may fail to be locally free. However, this sheaf is locally free on the
smooth locus of X, and the complement of the smooth locus has codimension 2
since X is normal. Hence we can use Proposition 8.0.1 to dene the sheaf of Zariski
p-forms
(8.0.5)

p
X
= (
p
X
)

= j

p
U
0
,
356 Chapter 8. The Canonical Divisor of a Toric Variety
where j : U
0
X is the inclusion of the smooth locus of X. It follows that

p
X
is a
reexive sheaf of rank

n
p

, where n = dim X.
For later purposes, we note that by Proposition 8.0.1, (8.0.5) is valid for any
smooth open subset U
0
X whose complement has codimension 2.
The case p = n is especially important.
Denition 8.0.20. The canonical sheaf of a normal variety X is

X
=

n
X
,
where n is the dimension of X. This is a reexive sheaf of rank 1, so that

X
O
X
(D)
for some Weil divisor D on X. We call this divisor a canonical divisor of X, often
denoted K
X
.
Proposition 8.0.7 shows that the canonical divisor K
X
is well-dened up to
linear equivalence and hence gives a unique divisor class in Cl(X), known as the
canonical class of X. In the toric case, we will see later in the chapter that there is
a natural choice for the canonical divisor.
When X is smooth, we call
X
the canonical bundle since it is a line bundle. In
this case, the canonical divisor is Cartier. There are also singular varieties whose
canonical divisors are Cartierthese are the Gorenstein varieties to be studied later
in the chapter.
While it often sufces to know
X
up to isomorphism, there are situations
where a unique model of
X
is required. One such construction uses

C(X)/C
,
where C(X) is the eld of rational functions on X. We can regard

n

C(X)/C
as
the constant sheaf of rational n-forms on X, similar to the way that C(X) gives the
constant sheaf K
X
of rational functions on X. There is an obvious sheaf map

n
X

C(X)/C
.
You will prove the following result in Exercises 8.0.10 and 8.0.11.
Proposition 8.0.21. When X is normal, image of the map
n
X

C(X)/C
is

C(X)/C
.
The canonical sheaf can be dened for any irreducible variety X as a subsheaf
of

C(X)/C
, though the denition is more sophisticated (see [184, 9]). When
X is projective, another approach is given in [131, III.7], where
X
is called the
dualizing sheaf. We will see in Chapter 9 that
X
plays a key role in Serre duality.
Exercises for 8.0.
8.0.1. The rank of a coherent sheaf on an irreducible variety was dened in Denition 8.0.2.
Here are some properties of the rank.
8.0. Background: Reexive Sheaves and Differential Forms 357
(a) Let an irreducible afne variety have coordinate ring R with eld of fractions K. Let
M be a nitely generated R-module. Show that M
R
K is a nite-dimensional vector
space over K whose dimension equals the rank of the coherent sheaf

M on Spec(R).
(b) Let F be a coherent sheaf on X let U X be an nonempty open subset. Prove that F
and F[
U
have the same rank.
(c) Let 0 F G H be an exact sequence of sheaves on X. Prove that rank(G) =
rank(F) +rank(H ).
8.0.2. Prove the claims made in Example 8.0.5.
8.0.3. Let R =C[x
1
, . . . , x
n
]. In Example 8.0.9 we claimed that df =

n
i=1
f
xi
dx
i
in
R/C
for all f R. Prove this.
8.0.4. Prove that the map in the exact sequence from Example 8.0.11 comes from the
Jacobian matrix (8.0.3).
8.0.5. Prove Proposition 8.0.12.
8.0.6. Prove that the cotangent sheaf
1
X
dened in Denition 8.0.13 is a coherent sheaf.
8.0.7. Given a C-algebra R and an R-module M, a C-derivation : R M is a C-linear
map that satises the Leibniz rule, i.e., ( f g) = f (g) +g( f ) for all f , g R.
(a) Show that f df denes a C-derivation d : R
R/C
.
(b) More generally, show that if :
R/C
M is an R-module homomorphism, then
d : R M is a C-derivation.
8.0.8. Continuing Exercise 8.0.7, we let Der
C
(R, M) denote the set of all C-derivations
: R M. This is an R-module where (r)( f ) = r( f ).
(a) Use part (b) of Exercise 8.0.7 to construct an R-module isomorphism Der
C
(R, M)
Hom
R
(
R/C
, M). Explain why d : R
R/C
is called the universal derivation.
(b) Let T
X
be the tangent sheaf of a variety X and let U = Spec(R) be an afne open
subset of X. Prove that T
X
(U) = Der
C
(R, R).
8.0.9. Fill in the details omitted in Example 8.0.17.
8.0.10. Let j : U X be the inclusion of a nonempty open subset of a variety X.
(a) Show that there is a sheaf map F j

(F[
U
) for any sheaf F on X.
(b) Show that the map of part (a) is an isomorphism when X is irreducible and F is a
constant sheaf.
8.0.11. Prove Proposition 8.0.21. Hint:
1
X
is locally free when restricted to the smooth
locus of X. Exercise 8.0.10 will be useful.
8.0.12. Let 1 T
N
be the identity element of the torus T
N
and let m C[M] be the corre-
sponding maximal ideal. Set N
C
= N
Z
C and M
C
= M
Z
C.
(a) Let f =

m
c
m

m
C[M]. Show that f m
2
if and only if

m
c
m
= 0 in C and

m
c
m
m = 0 in M
C
. Hint: Pick a basis e
1
, . . . , e
n
of M and set t
i
=
ei
, so that f is
a Laurent monomial in t
1
, . . . , t
n
. Then show that f m
2
if and only if f m and
f
ti
(1) = 0 for all i.
(b) Use part (a) to construct an isomorphism m/m
2
M
C
, and conclude that the Zariski
tangent space of T
N
at the identity is naturally isomorphic to N
C
.
358 Chapter 8. The Canonical Divisor of a Toric Variety
8.0.13. Let F be a free module of rank n over a ring R and x 0 p n. Recall that wedge
product induces an isomorphism

np
F Hom
R
(

p
F,

n
F).
(a) If X is smooth of dimension n, then show that
np
X
Hom
OX
(
p
X
,
n
X
).
(b) Show that if X is a normal variety of dimension n, then there is an isomorphism

np
X
Hom
OX
(

p
X
,
X
). Hint: If you get stuck, see [76, Prop. 4.7].
(c) In a similar vein, show that the tangent sheaf T
X
of a normal variety X satises T
X

Hom
OX
(

1
X
, O
X
).
8.1. One-Forms on Toric Varieties
In this section we will describe two interesting exact sequences that involve the
sheaves
1
X

and

1
X

on a normal toric variety X

.
The Torus. The coordinate ring of the torus T
N
is the semigroup algebra C[M].
Then the map

C[M]/C
M
Z
C[M]
dened by d
m
m
m
is easily seen to be an isomorphism. It follows that
(8.1.1)
1
T
N
M
Z
O
T
N
,
and dualizing, we obtain
T
T
N
N
Z
O
T
N
= N
C

C
O
T
N
.
This makes intuitive sense since T
N
= N
Z
C

as a complex Lie group. Thus its


tangent space at the identity is N
Z
C = N
C
via the exponential map. This is also
true algebraically, as shown in Exercise 8.0.12. The group action transports the
tangent space N
C
over the whole torus, which explains the above trivialization of
the tangent bundle T
T
N
.
As a consequence, the 1-form
d
m

m
is a global section of
1
T
N
that maps to m1
in (8.1.1) and hence is invariant under the action of T
N
. See Exercise 8.1.1 for more
on invariant 1-forms on the torus.
The First Exact Sequence. Now consider the toric variety X

of the fan . For


(1), the inclusion i : D

gives the sheaf i

O
D
on X

, which following
8.0 we write as O
D
. Using the map M Z given by m'm, u

`, we obtain the
composition
M
Z
O
X

Z
Z
O
X

=O
X

O
D
.
This gives a natural map
(8.1.2) : M
Z
O
X

O
D
,
where the direct sum is over all (1). We also have a canonical map
(8.1.3) :
1
X

M
Z
O
X

8.1. One-Forms on Toric Varieties 359


constructed as follows. On the afne piece U

= Spec(C[

M]), is dened by
d
m

C[

M]/C
m
m
M
Z
C[

M].
These C[

M]-module homomorphisms
C[

M]/C
M
Z
C[

M] patch
to give the desired map (8.1.3) (Exercise 8.1.2). Note that over the torus T
N
, the
map of (8.1.3) reduces to the isomorphism (8.1.1).
Theorem 8.1.1. For a smooth toric variety X

, the sequence
0
1
X

M
Z
O
X

O
D
0
formed using (8.1.2) and (8.1.3) is exact.
Proof. We rst verify that is the zero map. On the afne piece U

, the
subvariety D

is dened by the ideal I

= I(D

) C[

M]. By
Propositions 4.0.28 and 4.3.2, we have
(8.1.4) I

= (U

, O
X

(D

)) =

div(
m
)[
U
D[
U
C
m
=

M,/m,u)>0
C
m
.
Over U

, the composition
1
X

M
Z
O
X

O
D
takes a 1-form d
m
, m

M, to 'm, u

`
m
C[

M]/I

. This is obviously zero if 'm, u

` = 0, and
if 'm, u

` = 0, it vanishes since
m
I

in this case.
We now verify that the sequence is exact over U

. Since is smooth, we
may assume = Cone(e
1
, . . . , e
r
), where r n and e
1
, . . . , e
n
is a basis of N.
Then U

= C
r
(C

)
nr
. Let x
1
, . . . , x
n
denote the characters of the correspond-
ing dual basis of M, also denoted e
1
, . . . , e
n
. The coordinate ring of U

is R =
C[x
1
, . . . , x
r
, x
1
r+1
, . . . , x
1
n
], and the 1-forms on U

form the free R-module


R/C
=

n
i=1
Rdx
i
by Example 8.0.9 and Proposition 8.0.12. Since takes dx
i
to e
i
x
i
,
we see that can be regarded as the map

R/C
=
n

i=1
Rdx
i
M
R
R =
n

i=1
R
that sends

n
i=1
f
i
dx
i
to ( f
1
x
1
, . . . , f
n
x
n
). This gives the exact sequence
0
R/C

i=1
R
r

i=1
R/'x
i
` 0
since x
r+1
, . . . , x
n
are units in R, and the theorem follows.
Logarithmic Forms. The exact sequence of Theorem 8.1.1 has a lovely interpreta-
tion in terms of residues of logarithmic 1-forms. The idea is that M
Z
O
X

can be
thought of as the sheaf
1
X

(logD) of 1-forms on X

with logarithmic poles along


D =

. We begin with an example.


360 Chapter 8. The Canonical Divisor of a Toric Variety
Example 8.1.2. The coordinate ring of C
n
is R =C[x
1
, . . . , x
n
], and the divisor D is
the sum of the coordinate hyperplanes D
i
= V(x
i
). As above,
R/C
=

n
i=1
Rdx
i
.
Now introduce some denominators: a rational 1-form has logarithmic poles
along D if
=
n

i=1
f
i
dx
i
x
i
, f
i
R.
These form the free R-module

n
i=1
R
dx
i
x
i
, and the corresponding sheaf is dened
to be
1
C
n (logD). The formal calculation
d
m

m
=
n

i=1
'm, e
i
`
dx
i
x
i
shows that the map
d
m

m
m1 induces an isomorphism of sheaves

1
C
n (logD) M
C
n
O
C
n
such that the map :
1
C
n M
C
n
O
C
n
of (8.1.3) is induced by the inclusion of
1-forms
1
C
n
1
C
n (logD).
This construction works for any smooth afne toric variety U

, and the sheaves


of logarithmic 1-forms on U

patch to give the sheaf


1
X

(logD) for any smooth


toric variety X

. Furthermore, we have a canonical isomorphism


(8.1.5)
1
X

(logD) M
Z
O
X

such that the map of (8.1.3) comes from the inclusion of 1-forms.
The construction of
1
X

(logD) can be done more generally. Let X be a smooth


variety. A divisor D =

i
D
i
on X has simple normal crossings if every D
i
is
smooth and irreducible, and for every p X, the divisors containing p meet nicely.
More precisely, if I
p
= i [ p D
i
, we require that the tangent spaces T
p
(D
i
)
T
p
(X) meet transversely, i.e.,
codim

iIp
T
p
(D
i
)

=[I
p
[.
For example, the divisor D =

is a simple normal crossing divisor on any


smooth toric variety X

. A nice discussion of
1
X
(logD) for complex manifolds
can be found in [125, p. 449].
The Poincar e Residue Map. Let f (z) be an analytic (also called holomorphic)
function dened in a punctured neighborhood of a point p C. Take a counter-
clockwise loop C around p on X such that f (z) has no other poles inside C. Then
the residue of the 1-form = f (z)dz at p is the contour integral
res
p
() =
1
2i

C
.
8.1. One-Forms on Toric Varieties 361
In particular, if p = 0 and f (z) =
g(z)
z
, with g(z) analytic at zero, then the residue
theorem tells us that res
p
() is the coefcient of
1
z
in the Laurent series for f (z) at
0, and is equal to g(0). Note that the 1-form = f (z)dz =g(z)
dz
z
has a logarithmic
pole at p = 0.
When there are several variables, we can do the same construction by working
one variable at a time. Here is an example.
Example 8.1.3. Given f = f (x
1
, . . . , x
n
) R = C[x
1
, . . . , x
n
], we get the logarith-
mic 1-form = f
dx
1
x
1
. In terms of the above discussion of residues, we can regard
f (0, x
2
, . . . , x
n
) as the residue of at V(x
1
). Note also that f (0, x
2
, . . . , x
n
) repre-
sents the class of f in R/'x
1
`. Doing this for every variable shows that the map

1
C
n (logD) M
Z
O
C
n

i=1
O
D
i
can be interpreted as a sum of residue maps.
More generally, if X is a smooth variety and D =

i
D
i
a simple normal cross-
ing divisor, one can dene the Poincar e residue map
P
r
:
1
X
(logD)

i
O
D
i
(see [227, p. 254]) such that we have an exact sequence
(8.1.6) 0
1
X

1
X
(logD)
Pr

i
O
D
i
0.
When applied to a smooth toric variety X

and the divisor D =

, this gives
the exact sequence of Theorem 8.1.1 via the isomorphism (8.1.5).
The Normal Case. When X

is normal, we get an analog of Theorem 8.1.1 that


uses the sheaf

1
X

of Zariski 1-forms in place of


1
X

. Since M
Z
O
X

is reexive,
taking the double dual of (8.1.3) gives a map

1
X

M
Z
O
X

.
Theorem 8.1.4. Let X

be a normal toric variety. Then:


(a) The sequence
0

1
X

M
Z
O
X

O
D
is exact.
(b) If X

is simplicial, then the map on the right is surjective, so that


0

1
X

M
Z
O
X

O
D
0
is exact.
362 Chapter 8. The Canonical Divisor of a Toric Variety
Proof. Let j : U
0
X

be the inclusion map for U


0
=

. Note that U
0
is
a smooth toric variety whose fan has the same 1-dimensional cones as , and
codim(X `U
0
) 2 by the Orbit-Cone Correspondence. By Theorem 8.1.1, we
have an exact sequence
0
1
U
0
M
Z
O
U
0

O
DU
0
0,
so that applying j

gives the exact sequence


0 j

1
U
0
j

(M
Z
O
U
0
)

O
DU
0
since j

is left exact (Exercise 8.1.3). However, j

1
U
0
=

1
X

by the remarks
following (8.0.5), and j

(M
Z
O
U
0
) = M
Z
O
X

by Proposition 8.0.1. Hence we


get an exact sequence
0

1
X

M
Z
O
X

O
DU
0
.
In Exercise 8.1.4 you will show that the maps M
Z
O
X

O
DU
0
factor as
M
Z
O
X

O
D
j

O
DU
0
,
where O
D
j

O
DU
0
is injective. The exact sequence of part (a) then follows
immediately.
It remains to show that M
Z
O
X

O
D
is surjective when X

is simpli-
cial. Given , we need to show that
M
Z
O
U

(1)
O
DU
is surjective. Fix (1) and pick m M such that 'm, u

` = 0 and 'm, u

` = 0
for all

= in (1). Such an m exists since is simplicial. Then m1 maps


to a nonzero constant function on O
DU
and to the zero function on O
D

U
for

= . The desired surjectivity now follows easily.


When X

has no torus factors, we learned in 5.3 that graded modules over the
total coordinate ring S = C[x

[ (1)] give quasicoherent sheaves on X

. It is
easy to describe a graded S-module that gives

1
X

. For each , there are two maps


M
Z
S Z
Z
S = S S/'x

`,
where the rst map comes from m 'm, u

` and the second map is obvious. This


gives a homomorphism M
Z
S

S/'x

`, and we dene

1
S
to be the kernel
of this map. Hence we have an exact sequence of graded S-modules
(8.1.7) 0

1
S
MS

S/'x

`.
Using Example 6.0.10 and Theorem 8.1.4, we obtain the following result.
Corollary 8.1.5. When X

has no torus factors,


1
X

is the sheaf associated to the


graded S-module

1
S
.
8.1. One-Forms on Toric Varieties 363
The Euler Sequence. In [131, Thm. II.8.13], Hartshorne constructs an exact se-
quence
(8.1.8) 0
1
P
n O
P
n (1)
n+1
O
P
n 0,
called the Euler sequence of P
n
. He goes on to say This is a fundamental result,
upon which we will base all future calculations involving differentials on projective
varieties. Of course, P
n
is toric, and there is a toric generalization of this result,
due to Batyrev and Mel

nikov [20] and Jaczewski [160] in the smooth case and


Batyrev and Cox [19, Thm. 12.1] in the simplicial case.
Theorem 8.1.6. Let X

be a simplicial toric variety with no torus factors, i.e.,


u

[ (1) spans N
R
. Then there is an exact sequence
0

1
X

O
X

(D

) Cl(X

)
Z
O
X

0.
Furthermore, if X

is smooth, then the sequence can be written


0
1
X

O
X

(D

) Pic(X

)
Z
O
X

0.
Proof. Consider the following diagram:
0

1
X

M
Z
O
X

O
D

0
0

O
X

(D

O
X

O
D

0
0

Cl(X

)
Z
O
X

Cl(X

)
Z
O
X

0
0 0 0
The top row is from Theorem 8.1.4 and is exact since X

is simplicial. Also, by
Proposition 4.0.28, each (1) gives an exact sequence
0 O
X

(D

) O
X

O
D
0,
and the middle row is the direct sum of these exact sequences. The third row is the
obvious exact sequence that uses the identity map on Cl(X

)
Z
O
X

.
Since X

has no torus factors, we have the exact sequence


0 M

Z Cl(X

) 0
from Theorem 4.1.3, and tensoring this with O
X

gives the middle column. The


column on the right is the another obvious exact sequence, and one can check
without difculty that the solid arrows in the diagram commute (Exercise 8.1.5).
364 Chapter 8. The Canonical Divisor of a Toric Variety
Then commutativity and exactness imply the existence of the dotted arrows in the
diagram, which give the desired exact sequence by a standard diagram chase.
The exact sequence of sheaves in Theorem 8.1.6 is the (generalized) Euler
sequence of the toric variety X

. We will use it in the next section to determine the


canonical sheaf of X

. The Euler sequence also encodes relations generalizing the


classical Euler relation for homogeneous polynomials (see Exercise 8.1.8). Note
also that in [160], Jaczewski shows that smooth toric varieties can be characterized
as smooth varieties which admit a generalized Euler sequence.
Exercises for 8.1.
8.1.1. We will study invariant 1-forms and derivations on the torus. Since the torus T
N
=
Spec(C[M]) is afne, we know from Exercise 8.0.8 that the derivations Der
C
(C[M], C[M])
give the global sections of the tangent sheaf T
TN
.
(a) For u N, dene
u
: C[M] C[M] by

u
(
m
) ='m, u`
m
.
Prove that
u
Der
C
(C[M], C[M])
(b) Let x
1
, . . . , x
n
be the characters corresponding to the elements of M dual to some par-
ticular basis e
1
, . . . , e
n
of N. Thus C[M] =C[x
1
1
, . . . , x
1
n
]. Prove that
ei
= x
i

xi
and
that (T
N
, T
TN
) is the free C[M]-module generated by x
1

x1
, . . . , x
n

xn
.
(c) Dualizing, conclude that (T
N
,
1
TN
) is the free C[M]-module generated by the T
N
-
invariant differentials
dx1
x1
, . . . ,
dxn
xn
.
8.1.2. Consider the afne toric variety U

= Spec(C[

M]).
(a) Prove that the map
d
m

C[

M]/C
m
m
M
Z
C[

M]
denes a C[

M]-module homomorphism.
(b) For a toric variety X

, prove that these homomorphisms patch together to give the map


:
1
X
M
Z
O
X
in (8.1.3).
8.1.3. Let 0 F G H 0 be an exact sequence of sheaves on X and let f : X Y
be a morphism.
(a) Prove that 0 f

F f

G f

H is exact on Y.
(b) Suppose that Y = pt and f : X Y is the obvious map. Use part (a) to give a new
proof of Proposition 6.0.8.
8.1.4. In the proof of Theorem 8.1.4, show that the map M
Z
O
X
j

O
DU0
factors as
M
Z
O
X
O
D
j

O
DU0
,
where O
D
j

O
DU0
is injective.
8.2. Differential Forms on Toric Varieties 365
8.1.5. The proof of Theorem 8.1.6 contains the square
M
Z
O
X

O
D

O
X

O
D
.
Describe the maps in this square carefully and prove that it commutes.
8.1.6. Show that the Euler sequence from Theorem 8.1.6 reduces to (8.1.8) when X =P
n
.
8.1.7. Sometimes the name Euler sequence is used to refer to an exact sequence for the
tangent sheaf T
X
of a smooth toric variety.
(a) Show that for P
n
, we have an exact sequence
0 O
P
n O
P
n (1)
n+1
T
P
n 0.
Hint: Use (8.1.8).
(b) What is the corresponding sequence for a general smooth toric variety X

for as in
Theorem 8.1.6?
8.1.8. Let f be a homogeneous polynomial of degree d in C[x
1
, . . . , x
n
]. The classical Euler
relation is the equation
(8.1.9)
n

i=1
x
i
f
x
i
= d f .
In this exercise, you will prove this relation and consider generalizations encoded by the
generalized Euler sequence from a toric variety.
(a) Prove (8.1.9). Hint: Differentiate the equation
f (tx
1
, . . . , tx
d
) =t
d
f (x
1
, . . . , x
n
)
with respect to t.
(b) To see how the classical Euler relation generalizes, recall from Chapter 5 that given
a toric variety X

with no torus factors (i.e., u

[ (1) spans N
R
), we have the
total coordinate ring
S = C[x

[ (1)],
graded by Cl(X

). The graded pieces S

for Cl(X

) consist of homogeneous
polynomials as described by (5.2.1) from Chapter 5. If Hom
Z
(Cl(X

), Z) and
f S

show that we have a generalized Euler relation

(1)
([D

])x

f
x

= () f .
Hint: Follow what you did for part a, which is the case X =P
n1
.
(c) When Cl(X

) has rank greater than 1, there will be several distinct generalized Euler
relations on homogeneous elements of S. Compute the Euler relations on X =P
1
P
1
.
8.2. Differential Forms on Toric Varieties
For a toric variety X

, we have the sheaf of p-forms


p
X

and the sheaf of Zariski


p-forms

p
X

. By 8.0, the canonical sheaf of X

is
X

n
X

, n = dim X

.
366 Chapter 8. The Canonical Divisor of a Toric Variety
Properties of Wedge Products. We will need the following properties of wedge
products of free R-modules.
Proposition 8.2.1. Let F, G, H be free R-modules of nite rank.
(a) An R-module homomorphism : F G induces a homomorphism

p
:

p
F

p
G.
(b) Let 0 F GH 0 be an exact sequence with rankF = m and rankH =
n. Then rankG = m+n and there is a natural isomorphism

m+n
G

m
F
R

n
H.
Proof. Part (a) is straightforward (see Exercise 8.2.1 for an explicit description
of

p
), as is the rank assertion in part (b). For the isomorphism of part (b),
we assume n > 0 and dene a map

m
F
R

n
H

m+n
G as follows. If the
maps in the exact sequence are : F G and : G H, then one checks that

m
:

m
F

m
G is injective and

n
:

n
G

n
H is surjective. Then

m
F
R

n
H maps to

m
()

m+n
G, where

= . This map is well-


dened and gives the desired isomorphism (Exercise 8.2.1).
A corollary of this proposition is that if 0 F G H 0 is an exact
sequence of locally free sheaves on a variety X with rankF = m and rankH = n,
then rankG = m+n and there is a natural isomorphism
(8.2.1)

m+n
G

m
F
O
X

n
H .
Example 8.2.2. Suppose that Y X is a smooth subvariety of a smooth variety,
and let n = dim X, m = dimY. Then we have the exact sequence
0 I
Y
/I
2
Y

1
X

O
X
O
Y

1
Y
0
from Theorem 8.0.18, where I
Y
O
X
is the ideal sheaf of Y. By (8.2.1), we obtain

n
(
1
X

O
X
O
Y
)

nm
(I
Y
/I
2
Y
)
O
Y

1
Y
.
One can check that

n
(
1
X

O
X
O
Y
) (

1
X
)
O
X
O
Y
. Now recall that
X
=

1
X
and
Y
=

1
Y
and that the normal sheaf of Y X is N
Y/X
= (I
Y
/I
2
Y
)

.
Hence the above isomorphism implies

Y

X

O
X

nm
N
Y/X
.
This isomorphism is called the adjunction formula.
The Canonical Sheaf of a Toric Variety. Our rst major result gives a formula for
the canonical sheaf of a toric variety.
Theorem 8.2.3. For a toric variety X

, the canonical sheaf


X

is given by

O
X

.
Thus K
X

is a torus-invariant canonical divisor on X

.
8.2. Differential Forms on Toric Varieties 367
Proof. We rst assume that X

is smooth with no torus factors. Then we have the


Euler sequence
0
1
X

O
X

(D

) Pic(X

)
Z
O
X

0
from Theorem 8.1.6. Each O
X

(D

) is a line bundle since X

is smooth, and if
we set r = [(1)[, then one sees easily that Pic(X

)
Z
O
X

O
rn
X

. Hence we
can apply part (b) of Proposition 8.2.1 to obtain
(8.2.2)

n

1
X

O
X

rn
O
rn
X

O
X

(D

.
It follows by induction from Proposition 8.2.1 that the right-hand side of (8.2.2) is
isomorphic to

O
X

(D

) O
X

.
Turning to the left-hand side of (8.2.2), note that

rn
O
rn
X

O
X

, so that the
left-hand side is isomorphic to

1
X

=
n
X

=
X

since X

is smooth. This proves the result when X

is smooth without torus factors.


In Exercise 8.2.2 you will deduce the result for an arbitrary smooth toric variety.
Now suppose that X

is normal but not necessarily smooth. Let j : U


0
X

be the inclusion map for U


0
=

. We saw in the proof of Theorem 8.1.4 that


U
0
is a smooth toric variety satsifying codim(X `U
0
) 2. Now consider
X

and
O
X

). Since the fans for U


0
and X

have the same 1-dimensional cones,


these sheaves become isomorphic over U
0
by the smooth case. Since these sheaves
are reexive and codim(X `U
0
) 2, we conclude that
X

O
X

) by
Proposition 8.0.1.
Here are some examples.
Example 8.2.4. Theorem 8.2.3 implies that the canonical bundle of P
n
is

P
n O
P
n (n1)
for all n 1 since Cl(P
n
) Z and D
0
D
1
D
n
. In Exercise 8.2.3, you will
see another way to understand and derive this isomorphism.
Example 8.2.5. The previous example shows that
P
2 O
P
2 (3). We will com-
pute this directly using

P
2 =
2
P
2
=

1
P
2
and the description of
1
P
2
as a rank 2 vector bundle given in Example 8.0.16.
Recall that the transition functions for this bundle are given by:

20
=

1/y x/y
2
0 1/y
2

,
12
=

y
2
/x
2
0
y/x
2
y/x

,
10
=

y/x
2
1/x
1/x
2
0

.
368 Chapter 8. The Canonical Divisor of a Toric Variety
By Exercise 8.2.1, the corresponding maps on

2
are given by the determinants of
these 22 matrices:

20
=
1
y
3
,

12
=
y
3
x
3
,

10
=
1
x
3
.
Note that each is a cube. It is also evident that

10
=

12

20
,
so that these give the transition functions for a line bundle on P
2
.
On the other hand,

2

1
P
2
O
P
2 (3) says the canonical bundle of P
2
is
the third tensor power of the tautological bundle described in Examples 6.0.19
and 6.0.21. To see this directly, we rst need to calibrate the coordinate systems.
Example 8.0.16 used coordinates x, y fromU

0
, and Example 6.0.19 used homoge-
neous coordinates x
0
, x
1
, x
2
for P
2
, with the standard open cover U
i
=P
2
`V(x
i
).
Letting x = x
0
/x
2
and y = x
1
/x
2
gives an isomorphism U

0
=U
2
. Translating
coordinates for U

1
, we have

1/x, y/x

1/(x
0
/x
2
), (x
1
/x
2
)/(x
0
/x
2
)

x
2
/x
0
, x
1
/x
0

,
hence U

1
=U
0
. A similar computation shows U

2
=U
1
. We are now set for the
nal calculation. Keep in mind that the
i j
are in the coordinate system with charts
U

i
. We will use
i j
to denote the same transition function, but using the U
i
charts.
Thus we have

20
= 1/y
3
= (x
2
/x
1
)
3
=
12

12
= y
3
/x
3
= (x
1
/x
0
)
3
=
01

10
= 1/x
3
= (x
2
/x
0
)
3
=
02
.
Up to a sign, these are indeed the cubes of the transition functions that we computed
for the tautological bundle O
P
n (1) in Example 6.0.19. In Exercise 8.2.4, you will
work through the denition of the canonical bundle directly to nd the transition
functions given in Example 8.0.16.
Example 8.2.6. When we computed the class group of the Hirzebruch surface H
r
in Example 4.1.8, we wrote the divisors D

as D
1
, D
2
, D
3
, D
4
and showed that
D
3
D
1
D
4
rD
1
+D
2
.
Thus Cl(H
r
) = Pic(H
r
) Z
2
is freely generated by the classes of D
1
and D
2
. It
follows that the canonical bundle can be written

Hr
O
Hr
(D
1
D
2
D
3
D
4
) O
Hr
((r +2)D
1
2D
2
)
by Theorem 8.2.3.
8.2. Differential Forms on Toric Varieties 369
The Canonical Module. For a toric variety X

without torus factors, the canonical


sheaf
X

comes from a graded S-module, where S = C[x

[ (1)] is the total


coordinate ring of X

. This module is easy to describe explicitly.


Each variable x

S has degree deg(x

) = [D

] Cl(X

). Dene

0
= deg

Cl(X

).
Then S(
0
) is the graded S-module where S(
0
)

= S

0
for Cl(X

). As
in 5.3, the coherent sheaf associated to S(
0
) is denoted O
X

(
0
). We have
the following result.
Proposition 8.2.7. O
X

(
0
)
X

.
Proof. According to Proposition 5.3.7, O
X

(
0
) O
X

(D) for any Weil divisor


with
0
= [D] Cl(X

). The denition of
0
allows us to pick D =

.
Then Theorem 8.2.3 implies
O
X

(
0
) O
X

.
We call S(
0
) the canonical module of S.
Corollary 8.2.8. For any normal toric variety X

we have an exact sequence


0
X

O
X

O
D
.
Proof. First suppose that X

has no torus factors. Multiplication by

induces
an exact sequence of graded S-modules
0 S(
0
) S

S/'x

`
since
0
=deg(

). The sheaf associated to S/'x

` is O
D
(Exercise 8.2.5), and
then we are done by Example 6.0.10 and Proposition 8.2.7. When X

has a torus
factor, the result follows using the strategy described in Exercise 8.2.2.
We can also describe
X

in terms of n-forms in the x

. Fix a basis e
1
, . . . , e
n
of
M. For each n-element subset I =
1
, . . . ,
n
(1), we get the nn determinant
det(u
I
) = det('e
i
, u

j
`).
This depends on the ordering of the
i
, as does the n-form dx

1
dx
n
, though
the product
det(u
I
)dx

1
dx
n
depends only on e
1
, . . . , e
n
. It follows that the n-form
(8.2.3)
0
=

[I[=n
det(u
I
)

/ I
x

dx

1
dx
n
is well-dened up to 1. If we set deg(dx

) = deg(x

), then
0

S/C
is
homogeneous of degree
0
. This gives the submodule
S
0

S/C
,
370 Chapter 8. The Canonical Divisor of a Toric Variety
which is isomorphic to S(
0
) since deg(
0
) =
0
.
To see where
0
comes from, let L =C(x

[ (1)) be the eld of fractions


of S. Then L is the function eld of C
(1)
, and the surjective toric morphism
: C
(1)
`Z() X

from Proposition 5.1.9 induces an injection on function elds

: C(X

) L.
Furthermore, the proof of Theorem 5.1.11 implies that for any mM, the character

m
C(X

) maps to

(
m
) =

x
/m,u)

.
We regard C(X

) as a subeld of L via

, so that C(X

) L and
m
=

x
/m,u)

.
This induces an inclusion of K ahler n-forms
(8.2.4)

n

C(X

)/C

L/C
.
A basis e
1
, . . . , e
n
of M gives coordinates t
i
=
e
i
for the torus T
N
, and we know
from 8.1 that
dt
i
t
i
is a T
N
-invariant section of
1
T
N
. Then
dt
1
t
1

dt
n
t
n
is a T
N
-invariant section of
n
T
N
which pulls back to a rational n-form in

n

L/C
via (8.2.4). Note that
dt
i
t
i
=

'e
i
, u

`
dx

since t
i
=

x
/e
i
,u)

. When we mutiply by

to clear denominators, we obtain

dt
1
t
1

dt
n
t
n
=

'e
1
, u

`
dx

'e
n
, u

`
dx

[I[=n
det(u
I
)

/ I
x

dx

1
dx
n
=
0
.
Hence
0
arises in a completely natural way.
This can be interpreted in terms of sheaves as follows. We regard (8.2.4) as an
inclusion of constant sheaves on X

. Then the inclusion

C(X

)/C
from Proposition 8.0.21 induces an inclusion

L/C
.
On the other hand, it is easy to see that S
0

S/C
induces an inclusion

S
0

L/C
.
8.2. Differential Forms on Toric Varieties 371
Using the above derivation of
0
, one can prove that

S
0
=
X

as subsheaves of the constant sheaf

L/C
see [184, Prop. 14.14]. Note that this
is an equality of sheaves, not just an isomorphism. This shows that from the point
of view of differential forms, S
0
deserves to be called the canonical module. It is
isomorphic to the earlier version S(
0
) of the canonical module via the map that
takes
0
S
0
to 1 S(
0
).
The Afne Case. The canonical sheaf
U
of a normal afne toric variety U

is
determined by its module of global sections, the canonical module. When we think
of
U
as the ideal sheaf O
U
(

), we get the ideal


(U

,
U
) (U

, O
U
) =C[

M].
Proposition 8.2.9. (U

,
U
)C[

M] is the ideal generated by the characters

m
for all m M in the interior of

.
Proof. Corollary 8.2.8 gives the exact sequence
0 (U

,
U
) C[

M]

C[

M]/I

,
where I

= I(D

) is the ideal of D

. Since
I

M,/m,u)>0
C
m

M
C
m
=C[

M]
by (8.1.4), it follows that (U

,
X

) is the direct sum of C


m
for all m M such
that 'm, u

` > 0 for all . Since the u

generate , this is equivalent to saying that


m is in the interior of

.
The Projective Case. A full dimensional lattice polytope P M
R
gives two inter-
esting graded rings, each with its own canonical module. For the rst, we use an
auxiliary variable t so that for every k N, a lattice point m (kP) M gives a
character
m
t
k
on T
N
C

. These characters span the semigroup algebra


(8.2.5) S
P
=C[C(P) (MZ)] =

m(kP)M
C
m
t
k
,
where C(P) = Cone(P1) M
R
R is the cone of P. Recall that the slice
of C(P) at height k is kP (see Figure 4 is 2.2). The ring S
P
is graded by setting
deg(
m
t
k
) = k. Then Proposition 8.2.9 tells us that the canonical sheaf of the
associated afne toric variety comes from the ideal
I
P
=

mInt(kP)M
C
m
t
k
S
P
.
This is the canonical module of S
P
.
372 Chapter 8. The Canonical Divisor of a Toric Variety
The second ring associated to P uses the total coordinate ring S of the projective
toric variety X
P
and the corresponding ample divisor D
P
. Let = [D
P
] Cl(X
P
)
be the divisor class of D
P
. The graded pieces S
k
S, k N, form the graded ring
S

k=0
S
k
.
The canonical module of S is S(
0
), which is isomorphic to the ideal

S
via multiplication by

since
0
= deg

). When restricted to S

, this
gives the ideal
I

k=0
I
k
S

,
where I
k
S
k
is generated by all monomials of degree k in which every variable
appears to a positive power.
It follows that the polytope P gives graded rings S
P
and S

, each of which has


an ideal representing the canonical module. These are related as follows.
Theorem 8.2.10. The graded rings S
P
and S

are naturally isomorphic via an


isomorphism that takes I
P
to I

.
Proof. The proof of Theorem 5.4.8 used homogenization to construct an isomor-
phism S
P
S

. It is straightforward to see that this isomorphism carries the ideal


I
P
S
P
to the ideal I

(Exercise 8.2.6).
Recall from Theorem 7.1.13 that X
P
Proj(S
P
) via the Proj construction from
7.0. Furthemore, graded S
P
-modules give quasicoherent sheaves on Proj(S
P
), as
described in [131, II.5] (this is similar to the construction given in 5.3). Then the
following is true (Exercise 8.2.7).
Proposition 8.2.11. The sheaf on X
P
associated to the ideal I
P
S
P
by the Proj
construction is the canonical sheaf of X
P
.
The situation becomes even nicer when P is a normal polytope. The ample
divisor D
P
is very ample and hence embeds X
P
into a projective space, which gives
the homogeneous coordinate ring C[X
P
]. As we learned in 2.0, C[X
P
] is also the
ordinary coordinate ring of its afne cone

X
P
, i.e.,
C[X
P
] =C[

X
P
].
Furthermore, since P is normal, Theorem 5.4.8 gives isomorphisms
S
P
S

C[X
P
]
and implies that

X
P
is the normal afne toric variety given by

X
P
= Spec(S
P
) = Spec(S

).
8.2. Differential Forms on Toric Varieties 373
It follows that the canonical sheaf of the afne cone of X
P
comes from the ideal
I
P
S
P
, and when we use the grading on S
P
and I
P
, they give X
P
and its canonical
sheaf via X
P
Proj(S
P
). Everything ts together very nicely.
There is a more general notion of canonical module that applies to any graded
Cohen-Macaulay ring S

k=0
S
k
where S
0
is a eldsee [56, Sec. 3.6]. The
canonical modules of S
P
S

constructed above are canonical in this sense.


When the Canonical Divisor is Cartier. For a toric variety X

, the Weil divisor


K
X

is called the canonical divisor. Note that K


X

need not be a Cartier


divisor if X

is not smooth. In fact, from Theorem 4.2.8, we have the following


characterization of the cases when the canonical divisor is Cartier.
Proposition 8.2.12. Let X

be a normal toric variety. Then K


X

is Cartier if and
only if for each maximal cone , there exists m

M such that
'm

, u

` = 1 for all (1).


Similarly, K is Q-Cartier if and only if for each maximal , there exists
m

M
Q
such that 'm

, u

` = 1 for all (1).


Example 8.2.13. Let = Cone(de
1
e
2
, e
2
) R
2
. The afne toric variety U

is the rational normal cone



C
d
. We computed Cl(

C
d
) Z/dZ in Example 4.1.4,
where the Weil divisors D
1
, D
2
coming from the rays satisfy D
2
D
1
, dD
1
0.
The canonical divisor K
b
C
d
= D
1
D
2
has divisor class corresponding to
[2] Z/dZ. Since the Picard group of a normal afne toric variety is trivial
(Proposition 4.2.2), it follows that K
b
C
d
is Cartier if and only if d 2.
Another way to see this is via Proposition 8.2.12, where one easily computes
that m

= (2/d)e
1
+e
2
satises 'm

, de
1
e
2
` ='m

, e
2
` = 1. This lies in M =Z
2
if and only if d 2.
We will use the following terminology.
Denition 8.2.14. Let K
X
be a canonical divisor on a normal variety X. We say
that X is Gorenstein if K
X
is a Cartier divisor.
Since the canonical sheaf
X
=O
X
(K
X
) is reexive, we have
X is Gorenstein K
X
is Cartier
X
is a line bundle
by Proposition 8.0.7. All smooth varieties are Gorenstein, of course. We will study
further examples of singular Gorenstein varieties in the next section of the chapter.
Renements. A renement

of a fan induces a proper birational toric mor-


phism : X

. The canonical sheaves of X

and X

are related as follows.


374 Chapter 8. The Canonical Divisor of a Toric Variety
Theorem 8.2.15. Let : X

be the toric morphism induced by a renement

of a fan in N
R
R
n
. Then

.
In particular, if

is a smooth renement of , then

n
X

.
Proof. First assume X

=U

, so that

renes . Since K

(1)
D

,
the description of global sections given in Proposition 4.3.3 implies that
(X

,
X

) =

mP

M
C
m
,
where
P

=m M
R
[ 'm, u

` 1 for all

(1).
We clearly have P

MInt(

) M since (1)

(1). The opposite inclusion


also holds, as we now prove. Given m Int(

) M, we have 'm, u` > 0 for all


u = 0 in . In particular, 'm, u

` > 0 for all

(1). This is an integer, so that


'm, u

` 1, which implies m P

M, as desired. Since
(U

,
U
) =

mInt(

)M
C
m
by Proposition 8.2.9, we conclude that (X

,
X

) = (U

,
U
). Then we have


U
since U

is afne.
In the general case, X

is covered by afne open subsets U

for , and

1
(U

) is the toric variety of the renement of induced by

. The above
paragraph gives isomorphisms

[
U

[
U
which are compatible with the inclusion U

when is a face of . Hence


these isomorphisms patch to give the desired isomorphism

.
In Chapter 11, we will prove the existence of smooth renements. It follows
that for any n-dimensional toric variety X

, there are two ways of constructing the


canonical sheaf
X

from the sheaf of n-forms of a smooth toric variety:


(Internal)
X

= j

n
U
0
, where j : U
0
X

is the inclusion of the smooth locus


of X

.
(External)
X

n
X

, where : X

is the toric morphism coming


from a smooth renement

of .
8.2. Differential Forms on Toric Varieties 375
Sheaves of p-forms. The sheaves

p
X

for 1 < p <n =dim (X

) are also important


for the geometry of X

. We construct a sequence
(8.2.6) 0

p
X

p
M
Z
O
X

p1
(

M)
Z
O
D
as follows. The map :
1
X

M
Z
O
X

from (8.1.3) induces

p
:
p
X

1
X

p
M
Z
O
X

,
and then
p
=

. To dene
p
, recall that for u N, the contraction map
i
u
:

p
M

p1
M has the property that
(8.2.7) i
u
(m
1
m
p
) =
p

i=1
(1)
i1
'm
i
, u`m
1
m
i
m
p
when m
1
, . . . , m
p
M. Note also that im(i
u
)

p1
(u

M) (Exercise 8.2.8).
An element (1) gives u

as usual. Using i
u
:

p
M

p1
(

M) and
O
X

O
D
, we obtain the composition

p
M
Z
O
X

p1
(

M)
Z
O
X

p1
(

M)
Z
O
D
.
This gives a natural map

p
:

p
M
Z
O
X

p1
(

M)
Z
O
D
.
Theorem 8.2.16. The sequence (8.2.6) is exact for any normal toric variety X

.
Proof. We begin with the smooth case. Since exactness of a sheaf sequence is
local, we can work over an afne toric variety U

= C
r
(C

)
nr
, where is
generated by the rst r elements of a basis of e
1
, . . . , e
n
of N. By abuse of notation,
e
1
, . . . , e
n
will denote the corresponding dual basis of M. Setting x
i
=
e
i
, we get
U

= Spec(R), R =C[x
1
, . . . , x
r
, x
1
r+1
, . . . , x
1
n
].
Then (8.2.6) comes from the exact sequence of R-modules
0

R/C
p

p
M
Z
R
p

i=1

p1
(e

i
M)
Z
R/'x
i
`,
where
p
maps dx
i
1
dx
ip
to e
i
1
e
ip
x
i
1
x
ip
by the description of
given in (8.1.3).
It is thus obvious that
p

p
= 0. To prove exactness, we regard

p
M
Z
R
as

p
F, where F is the free R module with basis e
1
, . . . , e
n
. Now suppose that

p
() = 0 for some

p
F. We can write uniquely as
=

i
1
<<ip
f
i
1
...ip
e
i
1
e
ip
.
Exactness will follow once we prove x
i
1
x
ip
divides f
i
1
...ip
for all i
1
< < i
p
.
376 Chapter 8. The Canonical Divisor of a Toric Variety
If an index i > r appears in f
i
1
...ip
, then x
i
automatically divides f
i
1
...ip
since x
i
is invertible in R. Now suppose that an index i r appears in f
i
1
...ip
. We can write
= e
i

1
+
2
, where e
i
does not appear in
2
and all f
i
1
...ip
involving the index
i appear in
1
. Since i
e
1
() =
1
, the only way for
p
() to vanish is for f
i
1
...ip
to
be zero in R/'x
i
`, i.e., for x
i
to divide f
i
1
...ip
. This completes the proof of exactness
in the smooth case.
The proof for the general case follows from the smooth case by an argument
similar to what we did in the proof of Theorem 8.1.4.
Note that when p = 1, the exact sequence (8.2.6) reduces to the sequence ap-
pearing in Theorem 8.1.4, and when p = n, we get the sequence in Corollary 8.2.8
since
X

n
X

.
When X

has no torus factors, it is easy to nd a graded S-module whose


associated sheaf is

p
X

. Adapting the denition of


p
in (8.2.6) to the module
case, we get a homomorphism

p
M
Z
S

p1
(

M)
Z
S/'x

`
of graded S-modules whose kernel we denote

p
S
. This gives an exact sequence of
graded S-modules
0

p
S

p
M
Z
S

p1
(

M)
Z
S/'x

`.
Using Example 6.0.10, we obtain the following corollary of Theorem 8.2.16.
Corollary 8.2.17. If X

has no torus factors, then


p
X

is the sheaf associated to


the graded S-module

p
S
.
The Afne Case. For an afne toric variety U

, the sheaf

p
U
is determined by its
global sections, which can be described using Theorem 8.2.16. We will need the
following notation. If m M and , let
V

(m) = Span
C
(m
0
M [ m
0
the minimal face of

containing m) M
C
,
where M
C
= M
Z
C. Here is a result due to Danilov [76, Prop. 4.3].
Proposition 8.2.18. Let U

be the toric variety of the cone . Then we have an


isomorphism
(U

p
U
)

p
V

(m)
m
.
Proof. Using the inclusion

MM and the exact sequence of Theorem 8.2.16


over U

, we obtain the exact sequence


0 (U

p
U
)

p
M
Z
(U

, O
U
)
p

p1
M
Z
(U

, O
D
)
8.2. Differential Forms on Toric Varieties 377
of modules over (U

, O
X

) =C[

M] =

M
C
m
. Also note that
(8.2.8) (U

, O
D
) =

M,/m,u)=0
C
m
by the analysis given in the proof of Theorem 8.1.1. It follows that we have a direct
sum decomposition (U

p
U
) =

M
(U

p
U
)
m
, where
(8.2.9) 0 (U

p
U
)
m

p
M
C
p

/m,u)=0

p1
M
C
is exact for m

M and
p
is the sum of the contraction maps i
u
for all
satisfying 'm, u

` = 0.
Thus

p
M
C
is in the kernel of
p
if and only if i
u
() = 0 for all with
'm, u

` = 0. You will show in Exercise 8.2.8 that i


u
() = 0 if and only if

p
(

)
C
. It follows that the kernel of
p
in (8.2.9) is the intersection
(8.2.10)

/m,u)=0

p
(

)
C
=


/m,u)=0
(

)
C

.
However, the intersection
F =

/m,u)=0

is the minimal face of

containing m, and one sees easily that


Span
R
(m
0
M [ m
0
F) =

/m,u)=0

.
It follows that V

(m) =

/m,u)=0
(

)
C
. This plus (8.2.10) imply that the kernel
of
p
in (8.2.9) is

p
V

(m), as claimed.
The Simplicial Case. When X

is simplicial, the sequence (8.2.6) is exact on the


right when p = 1 by part (b) of Theorem 8.1.4. For p > 1, similar though longer
exact sequences exist in the simplicial case, as we now describe without proof.
Given a fan , consider the sheaf on X

dened by
K
j
(, p) =

( j)

pj
(

M)
Z
O
V()
,
where V() = O() X

is the orbit closure corresponding to . Thus


K
0
(, p) =

p
M
Z
O
X

K
1
(, p) =

(1)

p1
(

M)
Z
O
V()
=

(1)

p1
(

M)
Z
O
D
,
and the exact sequence (8.2.6) can be written
0

p
X

K
0
(, p) K
1
(, p).
When X

is simplicial, one can extend this exact sequence as follows.


378 Chapter 8. The Canonical Divisor of a Toric Variety
Theorem 8.2.19. When X

is simplicial, there is an exact sequence


0

p
X

K
0
(, p) K
1
(, p) K
p
(, p) 0.
A proof of this result can be found in [218, Sec. 3.2], where K

(, p) is called
the pth Ishida complex. We will use this exact sequence in Chapter 9 to prove a
vanishing theorem for sheaf cohomology on simplicial toric varieties.
Exercises for 8.2.
8.2.1. Given a map of free R-modules : F G, we can pick bases of F, G and represent
by a n m matrix with entries in R, where m = rank(F), n = rank(G). These bases
induce bases of the wedge products

p
F and

p
G. Then prove that the induced map

p
:

p
F

p
G is given by the p p minors of (with appropriate signs).
8.2.2. Complete the proof of Theorem 8.2.3 in the smooth case by showing how to reduce
to the case when X

is smooth with no torus factors. Hint: First prove that any smooth
toric variety is equivariantly isomorphic to a product X

(C

)
r
, where X

is a smooth
toric variety with no torus factors. Then consider X

C
r
.
8.2.3. Let T
N
be the torus of P
n
. If x
0
, . . . , x
n
are the usual homogeneous coordinates, then
y
i
= x
i
/x
0
for i = 1, . . . , n are afne coordinates for the open subset where x
0
= 0. The
differential n-form =
dy1
y1

dyn
yn
spans (T
N
,
n
TN
) = (T
N
,

1
TN
) and has poles of
order 1 along each D
i
= V(x
i
) for i = 1, . . . , n. Show that if we write z
j
= x
j
/x
i
for the
afne coordinates on the complement of V(x
i
), and change coordinates to the z
j
, j = i,
then we can see also has a pole of order 1 along V(x
0
). Hence on P
n
, denes a section
of O
P
n (

n
i=0
D
i
).
8.2.4. Using the open subsets U
i
P
2
(see Figure 2 in Example 3.1.9), compute
1
P
2
(U
i
),
and write down the transition functions on U
i
U
j
. Compare to Example 8.0.16. If the
result of your computation differs, describe how you can explain this via a change of basis.
8.2.5. Let S be the total coordinate ring of a toric variety X

without torus factors. Prove


that O
D
is the sheaf associated to the graded S-module S/'x

`. Hint: Consider the exact


sequence 0 'x

` S S/'x

` 0 and apply Proposition 5.3.7 and Example 6.0.10.


8.2.6. Prove Theorem 8.2.10.
8.2.7. Prove that the sheaf associated to the ideal I
P
S
P
from Theorem 8.2.10 is the
canonical sheaf of X
P
= Proj(S
P
). Hint: Let S be the total coordinate ring of X
P
. We saw
in 5.3 that a graded S-module such as S(
0
) gives a sheaf on X
P
. Compare this to how
I

gives a sheaf on Proj(S

). See the proof of Theorem 7.1.13.


8.2.8. Let F be a free module of nite rank over a domain R. Given u F

, we get the
kernel u

F and the contraction map i


u
:

p
F

p1
F. Assume u = re
1
where r R
and e
1
, . . . , e
n
is a basis of F

.
(a) Prove that the image of i
u
is contained in

p1
u

.
(b) Given

p
F, prove that i
u
() = 0 if and only if

p
u

.
8.2.9. Assume that X

has no torus factors. For Cl(X

), dene

p
X
() to be the sheaf
associated to the graded S-module

p
S
(). Prove that (X

p
X
()) (

p
S
)

. Hint: If
you get stuck, see [19, Prop. 8.5].
8.3. Fano Toric Varieties 379
8.2.10. Compute the n-form
0
dened in (8.2.3) for P
n
, P(q
0
, . . . , q
n
), and P
1
P
1
.
8.2.11. Prove that our favorite afne toric variety V = V(xy zw) C
4
is Gorenstein.
Hint: Example 1.2.20.
8.2.12. Prove that a product of two Gorenstein toric varieties is again Gorenstein. Hint:
Use Propositions 3.1.14 and 8.2.12.
8.2.13. We will see in Chapter 10 that every 2-dimensional rational cone in R
2
is lattice
equivalent to a cone of the form = Cone(e
2
, de
1
ke
2
) for integers d > 0 and 0 k < d
with gcd(d, k) = 1. Prove that U

is Gorenstein if and only if d = k +1.


8.2.14. A strongly convex rational polyhedral cone is Gorenstein if its associated afne
toric variety is Gorenstein. In the discussion of the algebra (8.2.5), we saw that a full
dimensional lattice polytope P M
R
gives the cone
C(P) = Cone(P1) M
R
R.
(a) Prove that C(P) is Gorenstein.
(b) Prove that the dual cone C(P)

N
R
R is Gorenstein if and only if P is reexive.
In general, a cone is reexive Gorenstein if both the cone and its dual are Gorenstein.
Reexive Gorenstein cones play an important role in mirror symmetrysee [18, 204].
8.2.15. Explain why S
0
=

k=0
S
k0
gives another model for the canonical module
of the graded ring S

discussed in the text.


8.3. Fano Toric Varieties
We nish this chapter with a discussion of an interesting class of projective toric
varieties and their corresponding polytopes.
Denition 8.3.1. A complete normal variety X is said to be a Gorenstein Fano
variety if the anticanonical divisor K
X
is Cartier and ample.
Thus Gorenstein Fano varieties are projective. When X is smooth, we will
simply say that X is Fano.
Example 8.3.2. Example 8.2.5 shows that O
P
2 (K
P
2) O
P
2 (3) is ample, so P
2
is
a Fano variety. We continue this example to introduce the key ideas that will lead
to a classication of 2-dimensional Gorenstein Fano toric varieties.
The standard fan for P
2
= X

has minimal generators u


0
= e
1
e
2
, u
1
= e
1
and u
2
= e
2
. The polytope corresponding to the anticanonical divisor of P
2
is
P =m R
2
[ 'm, u
i
` 1, i = 0, 1, 2.
In Exercise 8.3.1 you will check that
P = Conv(e
1
e
2
, 2e
1
e
2
, e
1
+2e
2
).
This lattice polygon is shown in Figure 1 on the next page. The open circle in the
gure represents the origin, which is the unique interior lattice point of P. Also,
380 Chapter 8. The Canonical Divisor of a Toric Variety
t
t
t
t
t
d
t
t
t
t
d
d
d
d
d
d
P
t
d
t
t

d
d
P

Figure 1. The anticanonical polytope P and its dual P

for P
2
by Exercise 8.3.1, the dual polytope P

= u N
R
[ 'm, u` 1 for all m P is
given by
P

= Conv(e
1
, e
2
, e
1
e
2
),
which is the polytope generated by the ray generators of the fan of P
2
.
Example 8.3.3. In Exercise 8.3.2, you will generalize Example 8.3.2 by showing
that the weighted projective space P(q
0
, . . . , q
n
) is Gorenstein Fano if and only if
q
i
[ q
0
+ +q
n
for all 0 i n.
The special features of Example 8.3.2 involve an unexpected relation between
Fano toric varieties and the reexive polytopes introduced in Chapter 2.
Fano Toric Varieties and Reexive Polytopes. Recall from Denition 2.3.12 that
a lattice polytope in M
R
is reexive if its facet presentation is
(8.3.1) P =m M
R
[ 'm, u
F
` 1 for all facets F.
It follows that if P is reexive, the origin is the unique interior lattice point of P
(Exercise 2.3.5). Since a
F
= 1 for all facets F, the dual polytope is
(8.3.2) P

= Conv(u
F
[ F is a facet of P)
(Exercise 2.2.1). Finally, P

is a lattice polytope and is reexive (Exercise 2.3.5).


The polytopes pictured in Figure 1 are reexive.
The following result gives the connection between projective Gorenstein Fano
toric varieties and reexive polytopes generalizing what we saw above for P
2
.
Theorem 8.3.4. Let X

be a normal toric variety. If X

is a projective Gorenstein
Fano variety, then the polytope associated to the anticanonical divisor K
X

is reexive. Conversely, if X
P
is the projective toric variety associated to a
reexive polytope P, then X
P
is a Gorenstein Fano variety.
Proof. If X

is Gorenstein Fano, then K


X

is Cartier and ample. This implies


that polytope associated to K
X

has facet presentation


P =m M
R
[ 'm, u

` 1 for all (1).


8.3. Fano Toric Varieties 381
and hence is reexive by (8.3.1). Conversely, let P be a reexive polytope in M
R
.
The normal fan
P
of P denes the variety X
P
=X

P
. The facet presentation (8.3.1)
of P has a
F
= 1 for every facet F of P. Hence the Cartier divisor corresponding to
P is D
P
=

F
D
F
= K
X
P
. We proved that D
P
is ample in Proposition 6.1.10, so
that K
X
P
is ample. Hence X
P
is Gorenstein Fano.
Classication. By Theorem 8.3.4, classifying toric Gorenstein Fano varieties is
equivalent to classifying reexive polytopes P in M
R
. Since reexive polytopes
contain the origin as an interior point, classify means up to invertible linear maps
of M
R
induced by isomorphisms of M. This is called lattice equivalence. The
rst interesting case is in dimension two, where toric Gorenstein Fano surfaces
correspond to 16 equivalence classes of reexive polygons.
We begin with some general facts about a reexive polytope P. First note that
the lattice points of P are the origin and the lattice points lying on the boundary.
Furthermore, any boundary lattice point is primitive. We will also need two less
obvious results from [214].
The rst result concerns projections of reexive polytopes. Given a reexive
polygon P and a primitive element m M, there is a projection map

m
: M
R
M
R
/Rm
whose image is a polytope whose vertices lie in M/Zm.
Lemma 8.3.5. Let P be a reexive polytope in M
R
R
n
and let m be a lattice point
m in the boundary of P. Then
m
(P) is a lattice polytope in M
R
/Rm containing the
origin as an interior lattice point, and

m
(P) =
m

mF facet of P
F

.
Proof. For each p
m
(P),
1
m
(p) is a line parallel to the line Rm. By taking the
point in P that maps to p and is farthest along this line in the direction of m, we see
that the points in
m
(P) are in bijective correspondence with the points in
U =x P [ x +m / P for all > 0.
If F is a facet of P containing m, one easily sees that F U. Conversely, let x U.
One can show without difculty there exists some facet F of P, not parallel to m,
containing x and such that 'u
F
, m` < 0. Since P is reexive and m M, 'u
F
, m` is
an integer 1. Hence 'u
F
, m` = 1, so that x, m F. Thus U

mF
F, and
from here the lemma follows easily.
The second result is the following fact about pairs of lattice points on the
boundary of a reexive polytope. The vertices are primitive vectors in M, but
we can have other lattice points on the boundary of P as well.
382 Chapter 8. The Canonical Divisor of a Toric Variety
Lemma 8.3.6. Let m, m

be distinct lattice points on the boundary of a reexive


polytope P. Then exactly one of the following holds:
(a) m and m

lie in a common edge of P,


(b) m+m

= 0, or
(c) m+m

is also on the boundary of P.


The proof is left to the reader as Exercise 8.3.3.
The 2-Dimensional Case. The following theorem classies reexive polygons in
the plane M
R
R
2
, up to lattice equivalence. Figure 2 shows 16 lattice polygons,
where the open circle in the center of each polygon is the origin and the solid circles
are the lattice points on the boundary. The numbers in the labels give the number
of boundary lattice points. Polygons 3 and 9 are the dual pair from Example 8.3.2.
t
d
t
t

d
d
3
t
t
d
t
t

d
d
d
d

4a
t t
d
t
t
e
e
e
e
d
d

4b
t t t
d
t
e
e
e
e

4c
t t
t d t
t
d
d
d
d

5a
t t t
t d
t
d
d

5b
t t
t d t
t t
d
d
d
d
6a
t t t
t d t
t
d
d

6b
t t t
t d t
t

6c
t t t
t d
t
t

6d
t t t
t d t
t t

7a
t t t
t d
t t
t

7b
t
t
t
t
d
t
t
t
t
8a
t
t
t
t
t
d
t
t
t

8b
t
t
t
t
t
t
d
t
t

8c
t
t
t
t
t
d
t
t
t
t
d
d
d
d
d
d
9
Figure 2. The 16 equivalence classes of reexive lattice polygons in R
2
8.3. Fano Toric Varieties 383
Theorem 8.3.7. There are exactly 16 equivalence classes of reexive polygons in
the plane, shown in Figure 2.
Proof. We will sketch the proof following [215, Proposition 4.1], and leave the
details for the reader to verify (Exercise 8.3.4). We consider several cases.
Case A. First assume that P is a reexive polygon such that each edge contains
exactly two lattice points, and x one such pair. These lattice points must form a
basis for M since the triangle formed by these two vertices and the origin has lattice
points only at the vertices (part (a) of Exercise 8.3.4). Hence we can use a lattice
equivalence to place the two of them at e
1
and e
2
.
Subcase A.1. If P has exactly three vertices, the third is located at ae
1
+be
2
for some a, b Z. Since 0 is the only lattice point in the interior, we must have
a = b =1 (part (b) of Exercise 8.3.4). This gives the polygon of type 3.
Subcase A.2. Next, still in Case A, assume that there are three distinct vertices
m, m

, m

such that m+m

= m

. There must be edges of P containing the pairs


m

, m

and m

, m

(part (c) of Exercise 8.3.4). Hence each pair must yield a basis
for M and we can place m = e
1
, m
2
=e
1
+e
2
, and then m

= e
2
. Project P from
m

= e
2
using Lemma 8.3.5. It follows that P can only contain points e
1
, 0, e
1
on
the line y = 0 and e
2
, e
1
e
2
on the line y =1. Moreover, P cannot contain any
points with y 2. It follows that the only other possible polygons in this case are
4b, 5a, 6a (part (d) of Exercise 8.3.4) up to lattice equivalence.
Subcase A.3. Finally, in Case A, if we are not in subcases A.1 or A.2, then by
Lemma 8.3.6, if m and m

are not in the same edge, we must have m+m

= 0. The
only possibility here is clearly polygon 4a.
Case B. Assume there are exactly three lattice points on some edge of P. By an
isomorphism of M, we can place these at e
1
+e
2
, e
2
, e
1
+e
2
. Projecting from m=
e
2
and using Lemma 8.3.5, we see that P must be contained in the strip 1 x 1.
Moreover, P cannot contain any points with y < 3, or else e
2
= 0 would be an
interior lattice point. Up to lattice equivalence, the possibilities are 4c, 5b, 6b, 6c,
6d, 7a, 7b, 8a, 8b, and 8c (part (e) of Exercise 8.3.4).
Case C. Assume no edge of P has exactly three lattice points and some edge has
four or more. Place the vertex of this edge at e
1
e
2
so that e
1
, e
1
+e
2
and
e
1
+2e
2
also lie on this edge. Projecting from e
1
via Lemma 8.3.5 shows that
P lies above the line y =1, and since the origin is an interior point, there must be
a lattice point m = ae
1
+be
2
with a > 0 and b 1.
Subcase C.1 If b = 1, then 2e
1
e
2
must lie in P. If e
1
+2e
2
and 2e
1
e
2
do not lie on an edge of P, their sum e
1
+e
2
would lie in P by Lemma 8.3.6 and
force e
1
, e
2
to be interior points. This is impossible, so e
1
+2e
2
and 2e
1
e
2
lie
on an edge of P. Hence P has type 9.
Subcase C.2 If b = 0, then m = e
1
, for otherwise e
1
would be an interior point.
Applying Lemma 8.3.6 to e
1
and e
1
+e
2
shows that e
2
P. If no edge connects
384 Chapter 8. The Canonical Divisor of a Toric Variety
e
1
, e
2
, then Lemma 8.3.6 would imply e
1
+e
2
P. This point and e
1
+2e
2
would
force e
2
to be interior, again impossible. Hence e
1
, e
2
lie on an edge, which forces P
to be contained in the polygon of type 9. Then our hypotheses on P force equality.
Subcase C.3 If b 1, then e
2
becomes an interior point of P (part (f) of Exer-
cise 8.3.4). Hence this subcase cannot occur.
There are many other proofs of the classication given in Theorem 8.3.7, in-
cluding [74, 214, 230]. References to further proofs are given in [74, Thm. 6.10].
The collection of 2-dimensional reexive polygons also exhibits some very
interesting symmetries and regularities. For instance, we know that if P is reexive,
then its dual is also reexive. In Exercise 8.3.5 you will determine which are the
dual pairs in the list above. There is also a very interesting relation between the
numbers of boundary lattice points in P and P

: in all cases, we have


(8.3.3) [PM[ +[P

N[ = 12.
We will see in Chapter 10 that there is an explanation for this coincidence coming
from the cohomology theory of sheaves on surfaces.
Higher Dimensions. Reexive polytopes of dimension greater than two and their
associated toric varieties are being actively studied. See for instance [215] and the
references therein. One reason for the interest in these varieties is the relation with
mirror symmetry. It is known that there are a nite number of equivalence classes
of reexive polytopes in all dimensions. Using their computer program PALP,
Kreuzer and Skarke [183] determined that there are 4319 classes of 3-dimensional
reexive polytopes and 473800776 classes of 4-dimensional reexive polytopes.
Since these numbers grow so quickly, most more recent work has focused on sub-
classes, for instance the polytopes giving smooth Fano toric varieties. These poly-
topes have a deceptively simple description (Exercise 8.3.6). There are 5 types of
such polytopes in dimension 2 (Exercise 8.3.7), and Batyrev [17] and Sato [244]
have shown that there are 18 types in dimension 3 and 124 types in dimension 4.
See [216] for a classication algorithm and further references.
Exercises for 8.3.
8.3.1. Verify the claims about the polygons P and P

dened in Example 8.3.2.


8.3.2. Prove the claim about P(q
0
, . . . , q
n
) made in Example 8.3.3. Hint: Exercise 4.2.11.
8.3.3. Prove Lemma 8.3.6. Hint: Assume that (a) and (b) do not hold. Show that for any
facet F of P, 'u
F
, m` +'u
F
, m

` >2, and use this to conclude that (c) must hold.


8.3.4. In this exercise, you will supply the details for the proof of Theorem 8.3.7.
(a) Let m, m

M Z
2
and assume that Conv(0, m, m

) has no lattice points other than


the vertices. Prove that m, m

form a basis of M.
8.3. Fano Toric Varieties 385
(b) Show that if P is a reexive triangle with vertices at e
1
, e
2
, then the third vertex must
be e
1
e
2
. Hint: One way to show this succinctly is to use the projections from the
vertices e
1
and e
2
as in Lemma 8.3.5.
(c) In the case that each facet contains exactly two vertices, show that if there are vertices
m, m

, m

with m+m

= m

, then the pairs m, m

and m

, m

must lie in edges.


(d) Complete the proof that the polygons of types 4b, 5a, 6a are the only possibilities in
Subcase A.2. Hint: Show that Conv(e
2
, (e
2
e
1
), e
1
) is lattice equivalent to 5a.
(e) Show that every reexive polygon in Case B is equivalent to one of type 4c, 5b, 6b,
6c, 6d, 7a, 7b, 8a, 8b, or 8c.
(f) Prove the claim made in Subcase C.3.
8.3.5. Consider the 16 reexive polygons in Figure 2. Since each polygon in the gure is
reexive, its dual must also appear in the gure, up to lattice equivalence.
(a) For each polygon in the Figure 2, determine its dual polygon and where the dual ts
in the classication. In some cases, you will need to nd an isomorphism of M that
takes the dual to a polygon in the gure. Also, some of the polygon are self-dual, up
to lattice equivalence.
(b) Show that the relation (8.3.3) holds for each polar pair.
8.3.6. A lattice polytope is called Fano if the origin is an interior point and the vertices of
every facet form a basis of the lattice. Let Q N
R
be a Fano polytope.
(a) Prove that Q is reexive.
(b) Part (a) implies that Q

M
R
is a lattice polytope. Prove that X
Q
is a smooth Fano
toric variety whose normal fan is formed by taking cones over proper faces of Q.
(c) Prove that every smooth Fano toric variety arises this way.
8.3.7. Of the 16 Gorenstein Fano varieties classied in Theorem 8.3.7, exactly ve are
smooth. Find them.
8.3.8. Some of the polygons in Figure 2 give well-known toric varieties. For example,
type 9 gives P
2
and type 8a gives P
1
P
1
. This follows easily by computing the normal
fan. Here you will describe the toric surfaces coming from some of the other polygons in
Figure 2. This exercise is based on [74, Rem. 6.11].
(a) Show that type 8b corresponds to the blow-up of P
2
at one of the torus xed points.
Also show that this is the Hirzebruch surface H
1
. Hint: Remember the description of
blowing up given in 2.3.
(b) Similarly show that type 7a (resp. 6a) corresponds to the blow-up of P
2
at two (resp.
three) torus xed points.
(c) Showthat type 3 corresponds to a quotient P
2
/(Z/3Z) and is isomorphic to the surface
in P
3
dened by w
3
= xyz. Hint: Compute the normal fan and show that its minimal
generators span a sublattice of index 3 in Z
2
. Proposition 3.3.7 will be helpful. For
the nal assertion, use the characters coming from the lattice points of the polygon.
(d) Show that type 8c gives the weighted projective space P(1, 1, 2) and type 4c gives the
quotient P(1, 1, 2)/(Z/2Z). Hint: See Example 3.1.17.
(e) Show that type 4a gives the quotient P
1
P
1
/(Z/2Z).
(f) Show that type 6d gives the weighted projective space P(1, 2, 3).
386 Chapter 8. The Canonical Divisor of a Toric Variety
(g) Show that type 7b (resp. 6b) corresponds to the blow-up of P(1, 1, 2) at one (resp. two)
smooth torus xed points.
(h) Showthat type 5b gives the blow-up of P(1, 2, 3) at its unique smooth torus xed point.
8.3.9. In this exercise you will consider a toric surface that is not quite Fano. The lattice
polygon P = Conv(3e
1
, 3e
2
, 2e
1
2e
2
) M
R
= R
2
gives a toric surface X
P
. Let
D =

be the anticanonical divisor of X


P
.
(a) Prove that the ample Cartier divisor D
P
associated to P is given by D
P
=6D. Conclude
that D is not Cartier and that 6D is the smallest integer multiple of D that is Cartier.
(b) Show that the normal fan of P has minimal generators e
1
2e
2
, 2e
1
e
2
and that
the minimal generators are the vertices of Q = Conv(e
1
2e
2
, 2e
1
e
2
) N
R
.
(c) Show that the dual of Q is Q

=
1
6
P and conclude that 6 is the smallest integer multiple
of Q

that is a lattice polytope.


8.3.10. A complete toric surface X

is log del Pezzo if some integer multiple of its anti-


canonical divisor K
X
is an ample Cartier divisor. The index of X

is the smallest positive


integer such that such that K
X
is Cartier. This exercise and the next will consider this
interesting class of toric surfaces.
(a) Prove that a complete toric surface is log del Pezzo of index 1 if and only if it is
Gorenstein Fano.
(b) Prove that the toric surface of Exercise 8.3.9 is log del Pezzo of index 6.
8.3.11. A lattice polygon Q N
R
is called LDP if the origin is an interior point of Q and
the vertices of Q are primitive vectors in N. The dual Q

M
R
of a LDP polygon contains
the origin as an interior point but may fail to be a lattice polygon. The index of Q is the
smallest positive integer such that such that Q

is a lattice polygon. This exercise will


explore the relation between LDP polygons and toric log del Pezzo surfaces.
(a) Show that the polygon Q of Exercise 8.3.9 is LDP of index 6.
(b) An LDP polygon Q N
R
gives a fan in N
R
by taking the cones over the faces of Q.
Show that the minimal generators of are the vertices of Q and that the toric surface
X

is log del Pezzo of index equal to the index of Q.


(c) Conversely, let X

be a toric log del Pezzo surface and let Q be the convex hull of
the minimal generators of . Note that Q is LDP and that is the fan obtained by
taking the cones over the faces of Q. Hint: The key point is to show that every minimal
generator of is a vertex of Q. This can be proved using the strict convexity of the
support function of K
X
.
This exercise shows that classifying toric log del Pezzo surfaces up to isomorphism is
equivalent to classifying LDP polygons up to lattice equivalence. This is an active area of
researchsee [168].
8.3.12. Let X

be a smooth projective toric variety. As in Denition 6.4.10, a primitive


colection P =
1
, . . . ,
k
(1) gives a primitive relation
u
1
+ +u

k
=

(1)
c

, c

Z
>0
,
where is the minimal cone containing

k
i=1
u
i
. Following [17], the degree of P is
(P) = k

(1)
c

. Prove that X

is Fano if and only if (P) > 0 for all primitive


collections P. This is [17, Prop. 2.3.6]. Hint: Use (6.4.9) and Theorem 6.4.9.
Chapter 9
Sheaf Cohomology of
Toric Varieties
9.0. Background: Sheaf Cohomology
By Proposition 6.0.8, a short exact sequence of sheaves on X
(9.0.1) 0 F G H 0
gives rise to an exact sequence of global sections
(9.0.2) 0 (X, F) (X, G) (X, H ).
The failure of (X, G) (X, H ) to be surjective is measured by a sheaf coho-
mology group. The main goal of this chapter is to understand sheaf cohomology
on a toric variety.
Sheaves and Cohomology. Asheaf F on a variety X has sheaf cohomology groups
H
p
(X, F). The abstract denition uses an exact sequence of sheaves
0 F I
0
d
0
I
1
d
1
I
2
d
2
,
where I
0
, I
1
, . . . are injective. This term is from homological algebra: a sheaf I
is injective if given a sheaf homomorphismH

I and an injection : H G,
there exists a sheaf homomorphism making the diagram below commute:
I
0

H

G.

387
388 Chapter 9. Sheaf Cohomology of Toric Varieties
We say that I

is an injective resolution of F. From I

we get the complex of


global sections
(X, I

) : (X, I
0
)
d
0
(X, I
1
)
d
1
(X, I
2
)
d
2
.
The term complex refers to the fact that d
p+1
d
p
= 0 for all p 0. Then the pth
sheaf cohomology group of F is dened to be
(9.0.3) H
p
(X, F) = H
p
((X, I

)) = ker(d
p
)/im(d
p1
),
where for p = 0, we dene d
1
to be the zero map 0 (X, I
0
). One can prove
that injective resolutions always exist and that two different injective resolutions of
F give the same sheaf cohomology groups.
This denition of H
p
(X, F) has some very nice properties, including:
H
0
(X, F) = (X, F).
A sheaf homomorphism F G induces a homomorphism of cohomology
groups H
i
(X, F) H
i
(X, G) that is compatible with composition and takes
the identity to the identity, i.e., F H
i
(X, F) is a functor.
A short exact sequence of sheaves (9.0.1) gives a long exact sequence
0 H
0
(X, F) H
0
(X, G) H
0
(X, H )

0

H
1
(X, F) H
1
(X, G) H
1
(X, H)

1

p1

H
p
(X, F) H
p
(X, G) H
p
(X, H)
p
.
We call
p
: H
p
(X, H ) H
p+1
(X, F) a connecting homomorphism.
You will prove the rst bullet in Exercise 9.0.1. For the second bullet, the key step
is to show that given a sheaf homomorphism : F G and injective resolutions
F A

, G B

, there are sheaf homomorphisms


p
: A
p
B
p
such that the
diagram
F

A
0
d
0

A
1
d
1


G

A
0
d
0

A
1
d
1


commutes. We say that

: A

is a map of complexes. Then


p
induces the
desired map H
p
(X, F) H
p
(X, G). Finally, for the last bullet, an exact sequence
(9.0.1) lifts to an exact sequence of injective resolutions
0

A

B

C

0
0

F

0,
and one can show that taking global sections gives an exact sequence of complexes
(9.0.4) 0 (X, A

) (X, B

) (X, C

) 0.
9.0. Background: Sheaf Cohomology 389
It is a general fact in homological algebra that any exact sequence of complexes
0 A

0
gives a long exact sequence
(9.0.5)
0 H
0
(A

) H
0
(B

) H
0
(C

)

0

H
1
(A

) H
1
(B

) H
1
(C

)

1

p1

H
p
(A

) H
p
(B

) H
p
(C

)
p
.
Applied to (9.0.4), we get the desired long exact sequence in sheaf cohomology.
In the language of homological algebra, is left exact since (9.0.2) is exact.
Then the sheaf cohomology groups are the derived functors of . The texts [125,
131, 273] discuss sheaf cohomology and homological algebra in more detail. We
especially recommend Appendix 3 of [89] and Chapters 2 and 3 of [158].

Cech Cohomology. While the abstract denition of sheaf cohomology has nice
properties, it is not useful for explicit computations. Fortunately, there is a down-
to-earth way of viewing sheaf cohomology, in terms of the

Cech complex, which
we now describe.
Let U = U
i

i=1
be an open cover of X. The denition of a sheaf F on X
shows that H
0
(X, F) = (X, F) is the kernel of the map
(9.0.6)

1i
F(U
i
)
d
0

1i<j
F(U
i
U
j
),
where d
0
is dened as follows: if =(
i
)

1i
F(U
i
), then the component of
d
0
() in F(U
i
U
j
) for i < j is given by
j [
U
i
U
j

i[
U
i
U
j
. Here,
j

j[
U
i
U
j
is the restriction map F(U
j
) F(U
i
U
j
), and similarly for
i

i[
U
i
U
j
.
To get higher information about how sections of F t together, we extend
(9.0.6) to the

Cech complex. We will use the following notation. Let [] =1, . . . ,
be the index set for the open cover and let []
p
denote the set of all (p+1)-tuples
(i
0
, . . . , i
p
) of elements of I satisfying i
0
< < i
p
.
Denition 9.0.1. The group of pth

Cech cochains is

C
p
(U , F) =

(i
0
,...,ip)[]p
F(U
i
0
U
ip
).
One can think of an element of

C
p
(U , F) as a function that assigns an ele-
ment of F(U
i
0
U
ip
) to each (i
0
, . . . , i
p
) []
p
. Then we dene a differential

C
p
(U , F)
d
p


C
p+1
(U , F)
390 Chapter 9. Sheaf Cohomology of Toric Varieties
by describing how d
p
() operates on elements of []
p+1
:
d
p
()(i
0
, . . . , i
p+1
) =
p+1

k=0
(1)
k
(i
0
, . . . ,

i
k
, . . . , i
p+1
)[
U
i
0
U
i
p+1
.
As above, (i
0
, . . . ,

i
k
, . . . , i
p+1
)[
U
i
0
U
i
p+1
is obtained by applying the restriction
map
F(U
i
0

U
i
k
U
i
p+1
) F(U
i
0
U
i
p+1
)
to (i
0
, . . . ,

i
k
, . . . , i
p+1
). In Exercise 9.0.2 you will verify that d
p
d
p1
= 0.
Denition 9.0.2. Given a sheaf F on X and an open cover U =U
i

iI
of X, the

Cech complex is

(U , F) : 0

C
0
(U , F)
d
0


C
1
(U , F)
d
1


C
2
(U , F)
d
2
,
and the pth

Cech cohomology group is

H
p
(U , F) = H
p
(

(U , F)) = ker(d
p
)/im(d
p1
).
Notice that

H
0
(U , F) = H
0
(X, F) = (X, F) since F is a sheaf. However,

H
p
(U , F) need not equal H
p
(X, F) for p > 0. We will soon see that there is a
nice case where equality occurs for all p.
Cohomology of a Quasicoherent Sheaf . To compute the sheaf cohomology of a
quasicoherent sheaf F on a variety X using

Cech cohomology, the idea is to use an
open cover U of X such that

H
p
(U , F) equals H
p
(X, F) for all p. The following
vanishing theorem of Serre is very useful in this regard. Proofs can be found in
[125], [131] and [273].
Theorem 9.0.3 (Serre Vanishing for Afne Varieties). Let F be a quasicoherent
sheaf on an afne variety U. Then H
p
(U, F) = 0 for all p > 0.
By Theorem 9.0.3, we can compute the cohomology of a quasicoherent sheaf
F using any afne open cover. Here is the rough intuition:
Consider an arbitrary open cover U =U
i
of X. Since we can construct X by
gluing together the U
i
, the cohomology of X should be obtained from the co-
homologies of the U
i
and their various intersections. In other words, H

(X, F)
should be determined by the cohomology groups H

(U
i
0
U
i
p+1
, F) as we
vary over all p.
Now suppose that U = U
i
is an afne open cover. Then U
i
0
U
i
p+1
is
afne and hence has vanishing higher cohomology by Serres theorem. So all
that is left is H
0
(U
i
0
U
i
p+1
, F). This gives the

Cech complex, which
thus computes the sheaf cohomology of F.
This suggests the following result.
9.0. Background: Sheaf Cohomology 391
Theorem 9.0.4. Let U = U
i
be an afne open cover of a variety X and let F
be a quasicoherent sheaf on X. Then there are natural isomorphisms

H
p
(U , F) H
p
(X, F)
for all p 0.
The proof will be given later in the section after we introduce a spectral se-
quence that makes the above intuition rigorous. A more elementary proof that does
not use spectral sequences can be found in [131, Thm. III.4.5].
Here is an application of Theorem 9.0.4.
Example 9.0.5. We compute the cohomology of O
P
1 , O
P
1 (1) and
1
P
1
on P
1
.
Consider the afne open cover U =U
0
,U
1
of P
1
, where
U
0
= Spec(C[x]) and U
1
= Spec(C[x
1
]).
Note also that
U
0
U
1
= Spec(C[x, x
1
]).
For any quasicoherent sheaf F on P
1
, the

Cech complex is
F(U
0
) F(U
1
)
d
0
F(U
0
U
1
).
It follows that H
p
(P
1
, F) = 0 for p 2. Hence we need only consider H
0
(P
1
, F)
and H
1
(P
1
, F). For simplicity, we write these sheaf cohomology groups as H
0
(F)
and H
1
(F) respectively.
For F =O
P
1, the

Cech complex becomes
(9.0.7) C[x] C[x
1
]
d
0
C[x, x
1
]
where d
0
( f (x), g(x
1
)) = f (x) g(x
1
). Then
H
0
(O
P
1 ) = ker(d
0
) =C
H
1
(O
P
1 ) = coker(d
0
) = 0.
The assertion for H
0
is clear since f (x) g(x
1
) = 0 implies that f and g are the
same constant, and the assertion for H
1
follows since an element of C[x, x
1
] can
be written
a
m
x
m
+ +a
1
x
1
. .. .
g(x
1
)
+a
0
+a
1
x + +a
n
x
n
. .. .
f (x)
.
We can represent O
P
1 (1) as the line bundle O
P
1 (D), where the divisor D is
one of the xed points of the torus action on P
1
. It follows that we have an exact
sequence of sheaves
0 O
P
1(1) O
P
1 O
D
0.
The long exact sequence in sheaf cohomology gives
0 H
0
(O
P
1 (1)) H
0
(O
P
1 ) H
0
(O
D
) H
1
(O
P
1 (1)) H
1
(O
P
1 ) .
392 Chapter 9. Sheaf Cohomology of Toric Varieties
The map H
0
(O
P
1 ) H
0
(O
D
) is the isomorphism that sends a constant function on
P
1
to the same constant function on D. Since H
1
(O
P
1 ) =0, the long exact sequence
implies that
(9.0.8) H
0
(O
P
1 (1)) = H
1
(O
P
1 (1)) = 0.
Finally, for
1
P
1
, we use the Euler sequence
0
1
P
1
O
P
1(1) O
P
1 (1) O
P
1 0
from (8.1.8). Since H
p
(X, F G) H
p
(X, F) H
p
(X, G) (Exercise 9.0.3), the
vanishing (9.0.8) and the long exact sequence for cohomology imply that
H
0
(
1
P
1
) = 0
H
1
(
1
P
1
) H
0
(O
P
1 ) =C.
Earlier, in Example 6.0.5, we showed that the surjective sheaf homomorphism
O
P
1 (1) O
P
1 (1) O
P
1
is not surjective on global sections. This is what forces H
1
(
1
P
1
) to be nonzero.
Akey part of Example 9.0.5 was the surjectivity of (9.0.7). This is the algebraic
analog of a Cousin problem in complex analysis. A discussion of Cousin problems
and their relation to sheaves and cohomology can be found in [180, Ch. 13].
Serre Vanishing for Projective Varieties. The Serre vanishing theorem for afne
varieties (Theorem 9.0.3) has a projective version.
Theorem 9.0.6 (Serre Vanishing for Projective Varieties). Let L be an ample line
bundle on a projective variety X. Then for any coherent sheaf F on X, we have
H
p
(X, F
X
L

) = 0
for all p > 0 and 0.
A proof can be found in [131, Prop. II.5.3]. We will use this result later in the
chapter to prove some interesting vanishing theorems for toric varieties.
Higher Direct Images. Given a morphism f : X Y of varieties and a sheaf F
of O
X
-modules on X, the direct image is the sheaf f

F on Y dened by
U F( f
1
(U))
for U Y open. We noted in Example 4.0.24 that f

F is a sheaf of O
Y
-modules.
The denition of f

F implies in particular that


H
0
(Y, f

F) = H
0
(X, F)
since f
1
(Y) = X. More generally, there are homomorphisms
(9.0.9) H
p
(Y, f

F) H
p
(X, F),
9.0. Background: Sheaf Cohomology 393
which need not be isomorphisms for p > 0.
In this situation, we also get the higher direct image R
p
f

F, which is the sheaf


on Y associated to the presheaf dened by
U H
p
( f
1
(U), F).
Proposition 9.0.7. Let f : X Y be a morphism and F be a quasicoherent sheaf
on X. Then:
(a) The higher direct images R
p
f

F are quasicoherent sheaves on Y.


(b) If U Y is afne, then R
p
f

F[
U
is the sheaf associated to the O
Y
(U)-module
H
p
( f
1
(U), F).
A proof can be found in [131, Propositions II.5.8 and III.8.5]. One especially
nice case is when the higher direct images R
p
f

F vanish for p > 0. We will


see below that the maps (9.0.9) are isomorphisms when this happens. The proof
involves our next topic, spectral sequences.
Spectral Sequences. Readers not familar with spectral sequences should glance at
Appendix C before proceeding farther. Here we discuss two spectral sequences
relevant to this section.
First suppose that f : X Y is a morphism and F is a quasicoherent sheaf
on X. As above, we get the higher direct images R
p
f

F, which are sheaves on


Y. Proposition 9.0.7 shows that these sheaves compute the cohomology of F
over certain open subsets of X. So when we put these together, i.e., compute
H
p
(Y, R
q
f

F), we should get the cohomology of F on all of X. The precise form


of this intuition is the Leray spectral sequence:
E
p,q
2
= H
p
(Y, R
q
f

F) H
p+q
(X, F).
Furthermore, the map H
p
(Y, f

F) H
p
(X, F) from (9.0.9) is the edge homo-
morphism E
p,0
2
H
p
(X, F) (see Denition C.1.6). This spectral sequence is dis-
cussed in Theorem C.2.1 and in more detail in [115, II.4.17] and [125, p. 463].
Now assume R
q
f

F = 0 for q > 0. Then


E
p,q
2
=

H
p
(Y, f

F) q = 0
0 q > 0.
Then Proposition C.1.7 implies that (9.0.9) is an isomorphism. Thus we have
proved the following.
Proposition 9.0.8. Suppose f : X Y is a morphism and F is a quasicoherent
sheaf on X such that R
q
f

F =0 for q >0. Then the map (9.0.9) is an isomorphism


H
p
(Y, f

F) H
p
(X, F).
For our second spectral sequence, let U =U
i
be an open cover of X and F
be a sheaf on X. In the discussion leading up to Theorem 9.0.4, we asserted that
394 Chapter 9. Sheaf Cohomology of Toric Varieties
H

(X, F) is determined by H

(U
i
0
U
i
p+1
, F) as we vary over all p. The
precise meaning is given by the E
1
spectral sequence
(9.0.10) E
p,q
1
=

(i
0
,...,ip)[]p
H
q
(U
i
0
U
ip
, F) H
p+q
(X, F),
where the differential d
p,q
1
: E
p,q
1
E
p+1,q
1
is induced by inclusion, with signs sim-
ilar to the differential in the

Cech complex. This spectral sequence is constructed
in [115, II.5.4]. See also Theorem C.2.2.
We now have the tools needed to prove Theorem 9.0.4.
Proof of Theorem 9.0.4. We are assuming that U =U
i
is an afne open cover
of X. First observe that the q = 0 terms of (9.0.10) are given by
E
p,0
1
=

(i
0
,...,ip)[]p
H
0
(U
i
0
U
ip
, F) =

C
p
(U , F),
and the differentials d
p,0
1
are the differentials in the

Cech complex. Hence
E
p,0
2
= ker(d
1
: E
p,0
1
E
p+1,0
1
)

im(d
1
: E
p1,0
1
E
p,0
1
)
= H
p
(

(U , F)) =

H
p
(U , F).
Since F is quasicoherent and the intersections U
i
0
U
ip
are afne for all p 0,
it follows from Theorem 9.0.3 that E
p,q
1
=0 for q >0. This implies that E
p,q
2
=0 for
q > 0. Using Proposition C.1.7 again, we conclude that the edge homomorphism
E
p,0
2
=

H
p
(U , F) H
p
(X, F)
is an isomorphism for all p 0.
Cohen-Macaulay Varieties. We next discuss a class of varieties that play a crucial
role in duality theory and are interesting in their own right.
We rst dene what it means for a local ring (R, m) to be Cohen-Macaulay.
Elements f
1
, . . . , f
s
m form a regular sequence if f
i
is not a zero divisor in
R/' f
1
, . . . , f
i1
` for all i, and the depth of R is the maximal length of a regular
sequence. Then R is Cohen-Macaulay if its depth equals its dimension. Examples
include regular local rings. A nice discussion of what Cohen-Macaulay means can
be found in [246, 10.2]. See also [179, pp. 153155].
A variety X is Cohen-Macaulay if its local rings (O
X,p
, m
X,p
) are Cohen-
Macaulay for all p X. Thus smooth varieties are Cohen-Macaulay. Later in
the chapter we will prove that normal toric varieties are Cohen-Macaulay.
Serre Duality. Cohen-Macaulay varieties provide the natural setting for a basic
duality theorem of Serre. Here is the simplest version.
9.0. Background: Sheaf Cohomology 395
Theorem 9.0.9 (Serre Duality I). Let
X
be the canonical sheaf of a complete
normal Cohen-Macaulay variety X of dimension n. Then for every locally free
sheaf F of nite rank on X, there are natural isomorphisms
H
p
(X, F)

H
np
(X,
X

O
X
F

).
In particular, when D is a Cartier divisor on X and K
X
is a canonical divisor, we
have isomorphisms
H
p
(X, O
X
(D))

H
np
(X, O
X
(K
X
D)).
A proof for the projective case can be found in [131, Thm. III.7.6]. The asser-
tion for divisors follows from

O
X
O
X
(D)

O
X
(K
X
)
O
X
O
X
(D) O
X
(K
X
D),
where the last isomorphism holds since D is Cartier.
There is also a more general version of Serre duality that applies when F
is coherent but not necessarily locally free. The cohomology groups H
p
(X, F)
are the derived functors of the global section functor F (X, F), and in the
same way, the Ext groups Ext
p
O
X
(G, F) are the derived functors of the Hom functor
F Hom
O
X
(G, F) for xed G. Then we have the following result.
Theorem 9.0.10 (Serre Duality II). Let
X
be the canonical sheaf of a complete
normal Cohen-Macaulay variety X of dimension n. Then for every coherent sheaf
F on X, there are natural isomorphisms
H
p
(X, F)

Ext
np
O
X
(F,
X
).
When X is projective, this is proved in [131, Thm. III.7.6]. We will give a
version of Serre duality especially adapted to the toric case in 9.2.
When X fails to be Cohen-Macaulay, there is a more general duality theorem
where canonical sheaf
X
is replaced with the dualizing complex

X
. A discussion
of this version of duality can be found in [218, 3.2].
Singular Cohomology. Our discussion of the sheaf cohomology of a toric variety
in 9.1 will use some algebraic topology. Here we review the topological invariants
we will need, beginning with the singular cohomology groups
H
p
(Z, R),
where Z is a topological space and R is a commutative ring, usually Z, Q, R or C.
These are dened using continuous maps
:
n
Z,
where
n
is the standard n-simplex. There are several good introductions to sin-
gular cohomology, including [124], [135] and [210].
Here are some important properties of singular cohomology:
396 Chapter 9. Sheaf Cohomology of Toric Varieties
A continuous map f : Z W induces f

: H
p
(W, R) H
p
(Z, R) such that
homotopic maps induce the same map on cohomology and the identity map
induces the identity on cohomology.
If i : A Z is a deformation retract (i.e., there is a continuous map r : Z A
such that r i = 1
A
and i r is homotopic to 1
Z
), then i

: H
p
(Z, R) H
p
(A, R)
is an isomorphism.
If Z is contractible (i.e, there is z Z such that z Z is a deformation
retract), then
H
p
(Z, R) =

R p = 0
0 otherwise.
For n > 1, the singular cohomology of the (n1)-sphere S
n1
R
n
is
H
p
(S
n1
, R) =

R p = 0, n1
0 otherwise.
We will always assume that Z is a locally contractible metric space. This allows
us to interpret the singular cohomology of Z in terms of sheaf cohomology as
follows. The ring R denes a presheaf on Z where R is the group of sections over
every nonempty open U Z. The corresponding sheaf is the constant sheaf of R.
By [45, III.1], the sheaf cohomology of the constant sheaf of R on Z is the singular
cohomology H
p
(Z, R).
The Cohomology Ring. For a commutative ring R, H

(Z, R) is a graded R-algebra


with multiplication given by cup product. A continuous map induces a ring homo-
morphism on cohomology, so in particular, a deformation retract i : A Z induces
a ring isomorphism i

: H

(Z, R) H

(A, R).
Here are some examples.
Example 9.0.11. The real torus (S
1
)
n
has cohomology ring
H

((S
1
)
n
, R) = R[
1
, . . . ,
n
]
where the
i
lie in H
1
((S
1
)
n
, R) and satisfy the relations
i

j
=
j

i
and
2
i
= 0
(see Examples 3.11 and 3.15 of [135]). Thus
H

((S
1
)
n
, R)

R
n
,
i.e., the cohomology ring is the exterior algebra of a free R-module.
Example 9.0.12. The torus (C

)
n
contains the real torus (S
1
)
n
as a deformation
retract via (t
1
, . . . , t
n
) (t
1
/[t
1
[, . . . , t
n
/[t
n
[). Hence
H

((C

)
n
, R) H

((S
1
)
n
, R)

R
n
.
More canonically, the torus T
N
= NC

has cohomology ring


H

(T
N
, R)

M
Z
R,
9.0. Background: Sheaf Cohomology 397
where M is the dual of N. Thus a lattice homomorphism N N

gives a map
T
N
T
N
of tori, and the induced map H

(T
N
, R) H

(T
N
, R) is the map

Z
R

M
Z
R
determined by the dual homomorphism M

M.
Example 9.0.13. The cohomology ring of P
n
over Z is
H

(P
n
, Z) Z[]/'
n+1
`,
where H
2
(P
n
, Z) (see [135, Thm. 3.12]). In Theorem 12.4.4, we will give a
similar quotient description of the cohomology ring of any smooth complete toric
variety.
Reduced Cohomology. Given a topological space Z, the canonical map Z pt
sends all elements of Z to the single point denoted pt. This induces the map
R = H
0
(pt, R) H
0
(Z, R)
whose cokernel is denoted

H
0
(Z, R). Thus, assuming R = 0,

H
0
(Z, R) = 0 Z is path connected.
The reduced cohomology of Z with coefcients in R is dened for p 0 by

H
p
(Z, R) =

H
0
(Z, R) p = 0
H
p
(Z, R) p > 0,
and for p =1 by

H
1
(Z, R) =

0 Z =
R Z =.
The denition of

H
1
may look strange but means that the sequence
(9.0.11) 0

H
1
(Z, R) R H
0
(Z, R)

H
0
(Z, R) 0
is exact for all Z, including Z = (Exercise 9.0.4). We will use

H
1
and (9.0.11)
in 9.1 when we compute sheaf cohomology on a toric variety.
Exercises for 9.0.
9.0.1. Use (9.0.3) to show that H
0
(X, F) = (X, F).
9.0.2. Check that the map d
p
dened on the

Cech cochains satises d
p
d
p1
= 0.
9.0.3. Let F, G be sheaves on X. Prove H
p
(X, F G) H
p
(X, F) H
p
(X, G).
9.0.4. Prove the exactness of (9.0.11).
9.0.5. A morphism f : X Y is afne if Y has an afne open cover U
i
such that f
1
(U
i
)
is afne for all i. Nowassume that f : X Y is afne and let F be a quasicoherent sheaf on
X. Use Theorem 9.0.3, Proposition 9.0.7 and Theorem 9.0.8 to prove that H
p
(Y, f

F)
H
p
(X, F) for all p 0.
398 Chapter 9. Sheaf Cohomology of Toric Varieties
9.0.6. Use Exercise 9.0.5 to prove the following isomorphisms in cohomology:
(a) H
p
(X, i

F) H
p
(Y, F) when i : Y X is closed in X and F is quasicoherent.
(b) H
p
(X,

F) H
p
(V, F) when : V X is a vector bundle and F is quasicoherent.
9.0.7. Let X be a variety that is a union of afne open subsets X =U
1
U
s
.
(a) Let F be quasicoherent on X. Prove that H
p
(X, F) = 0 for p > s 1.
(b) Let F be quasicoherent on P
n
. Prove that H
p
(P
n
, F) = 0 for p > n.
9.0.8. Consider the (n 1)-sphere S
n1
R
n
. Show that

H
p
(S
n1
, R)

R p = n 1
0 otherwise.
Hint: Remember that S
0
consists of two points.
9.1. Cohomology of Toric Divisors
Demazure [82] gave a concrete description of the sheaf cohomology of the sheaf
O
X

(D) of a torus-invariant Cartier divisor D on a toric variety X

. We generalize
this to torus-invariant Q-Cartier divisors, inspired by papers of Eisenbud, Mustat a
and Stillman [91], Hering, K uronya and Payne [141], and Perling [226].
The Toric

Cech Complex. When we compute sheaf cohomology using

Cech co-
homology, the obvious choice of open cover is
U =U

max
,
where
max
is the set of maximal cones in . We write these as
i
and order them
according to their indices.
Given a torus-invariant Cartier divisor D =

on X

, the

Cech complex
is given by

C
p
(U , O
X

(D)) =

=(i
0
,...,ip)[]p
H
0
(U

i
0
U

ip
, O
X

(D)).
If we set

=
i
0

ip
for = (i
0
, . . . , i
p
) []
p
, then we rewrite this as
(9.1.1)

C
p
(U , O
X

(D)) =

[]p
H
0
(U

, O
X

(D)).
The Grading on Cohomology. For an afne open subset U

, Proposition 4.3.3
implies that the sections of O
X

(D) over U

can be written
H
0
(U

, O
X

(D)) =

m
C
m
,
where the direct sum is over all m M such that 'm, u

` a

for all (1).


We can write this as
H
0
(U

, O
X

(D)) =

mM
H
0
(U

, O
X

(D))
m
,
9.1. Cohomology of Toric Divisors 399
where for m M,
(9.1.2) H
0
(U

, O
X

(D))
m
=

C
m
'm, u

` a

for all (1)


0 otherwise.
This induces a grading of the

Cech complex (9.1.1). The

Cech differential
is built from the restriction maps and hence respects the grading of the complex.
Since H
p
(X

, O
X
(D)) =

H
p
(U , O
X
(D)), we obtain a natural decomposition of
sheaf cohomology
H
p
(X

, O
X
(D)) =

mM
H
p
(X

, O
X
(D))
m
.
When decomposing cohomology this way, we often refer to the m M as weights.
Here is an example of how weights can be used to compute sheaf cohomology.
Example 9.1.1. On P
2
, label the rays as usual: u
0
= e
1
e
2
, u
1
= e
1
, u
2
= e
2
,
and maximal cones
i
, starting with
0
in the rst quadrant and going counter-
clockwise. We will compute H
p
(P
2
, O
P
2 (a)) for a Z, where O
P
2(a) =O
P
2 (aD
0
)
for the divisor D
0
corresponding to u
0
.
Let U
i
be the afne open corresponding to
i
. Then U
i j
is U
i
U
j
and U
012
is
the triple intersection. This allows us to write the

Cech complex as
(9.1.3) 0

C
0
(a)
0
B
B
@
1 1 0
1 0 1
0 1 1
1
C
C
A


C
1
(a)

1 1 1


C
2
(a) 0,
where

C
0
(a) =
2

i=0
H
0
(U
i
, O
P
2 (aD
0
))

C
1
(a) =

i<j
H
0
(U
i j
, O
P
2 (aD
0
))

C
2
(a) = H
0
(U
012
, O
P
2 (aD
0
)).
We will compute the cohomology of this complex using the graded pieces for
m M = Z
2
. For this purpose, let l
0
, l
1
, l
2
be the lines in M
R
= R
2
dened by
'm, u
0
` = a, 'm, u
1
` = 0, 'm, u
2
` = 0. These lines divide the plane into various
regions called chambers. To get a disjoint decomposition, we dene
C
++
=m M
R
[ 'm, u
0
` <a, 'm, u
1
` 0, 'm, u
2
` 0.
Note that a minus sign corresponds to strict inequality (< 0) while a plus sign
corresponds to a weak inequality (0). The other chambers C
++
, C
+
, etc. are
dened similarly.
Suppose a < 0. The corresponding chamber decomposition of M
R
is shown in
Figure 1 on the next page. The labels l
i
are placed on the plus side of the lines, i.e.,
400 Chapter 9. Sheaf Cohomology of Toric Varieties
where 'm, u
0
` a, 'm, u
1
` 0, 'm, u
2
` 0. Each chamber is labeled with its sign
pattern, and the shading indicates which chamber the points on the lines belong to.
l
1
l
2
l
0
+ +

+ +
+ +
+
+ +
0
a
a
Figure 1. The chamber decomposition for a <0
The cone
i
is generated by u
j
, u
k
, where i, j, k =0, 1, 2. Thus
(9.1.4)
H
0
(U
0
, O
P
2 (a))
m
= 0 mC
++
M
H
0
(U
1
, O
P
2 (a))
m
= 0 mC
++
M
H
0
(U
2
, O
P
2 (a))
m
= 0 mC
++
M.
This follows from (9.1.2) (Exercise 9.1.1). Hence we know when

C
0
(a)
m
= 0.
The cone
1

2
is generated by u
0
, so its dual is the half-plane of M
R
where
'm, u
0
` a, i.e., on the plus side of l
0
. Using Figure 1 and (9.1.2) again, we get
H
0
(U
12
, O
P
2 (a))
m
= 0 m (C
++
C
+
C
++
) M.
Similar results hold for U
01
and U
02
, so we know when

C
1
(a)
m
= 0. Finally, since
U
012
is the torus of P
2
, we have

C
2
(a)
m
= H
0
(U
012
, O
P
2 (a))
m
= 0 for all m M.
Putting everything together, we get Table 1 on the next page, which shows the
dimension of

C
p
(a) for m M and a < 0. For example, dim

C
1
(a)
m
= 2 when
mC
++
since both U
01
and U
02
contribute 1-dimensional subspaces.
When a < 0, it is now easy to understand the

Cech complex
(9.1.5) 0

C
0
(a)
m


C
1
(a)
m


C
2
(a)
m
0.
The rst line of Table 1 corresponds to mC
++
C
++
C
++
. One can check
without difculty that for these ms, the complex (9.1.5) is exact, and the same is
true for the second row as well.
9.1. Cohomology of Toric Divisors 401
m M is in dim

C
0
(a)
m
dim

C
1
(a)
m
dim

C
2
(a)
m
C
++
C
++
C
++
1 2 1
C
+
C
+
C
+
0 1 1
C

0 0 1
Table 1. Dimension of

C
p
(a)m for a <0
These remarks imply that if m M is not in the interior of the triangle
a
2
= Conv(0, 0), (a, 0), (0, a),
then
H
p
(P
2
, O
P
2 (a))
m
= 0 for all p.
However, if m is a lattice point in the interior of a
2
, then H
2
(P
2
, O
P
2 (a))
m
is
nonzero. Summing up, we have:
(9.1.6) dim H
p
(P
2
, O
P
2 (a)) =

0 p = 2

Int(a
2
) M

a1
2

p = 2
when a < 0. In what follows, we write h
p
(a) = dim H
p
(P
2
, O
P
2 (a)) for a Z.
In Exercise 9.1.2 you will adapt the above methods to show that
(9.1.7) h
p
(a) =

a
2
M

a+2
2

p = 0
0 p > 0
when a 0. Thus, for any line bundle on P
2
, we have completely determined the
dimensions of the cohomology groups H
p
(P
2
, O
P
2 (a)). The values for 7 a 4
are depicted in Table 2. Note the symmetry in the table. In Exercise 9.1.3 you will
explore how this symmetry relates to Serre duality.
a h
0
(a) h
1
(a) h
2
(a)
7 0 0 15
6 0 0 10
5 0 0 6
4 0 0 3
3 0 0 1
2 0 0 0
1 0 0 0
0 1 0 0
1 3 0 0
2 6 0 0
3 10 0 0
4 15 0 0
Table 2. Sheaf cohomology of O
P
2 (a) on P
2
402 Chapter 9. Sheaf Cohomology of Toric Varieties
Computing Cohomology. Given a divisor D =

on X

and m M, dene
two subsets of [[:
(When is general) V
D,m
=

Conv(u

[ (1), 'm, u

` <a

).
(When D is Q-Cartier) V
supp
D,m
= u [[ [ 'm, u` <
D
(u), where
D
is the
support function of D (see Exercise 9.1.4).
Here is an example of these sets.
Example 9.1.2. Let X

be the toric surface obtained by blowing up P


2
at the three
xed points of the torus action. The minimal generators u
1
, . . . , u
6
of the fan give
divisors D
1
, . . . , D
6
on X

. Let D =D
3
+2D
5
D
6
and m = e
1
. Since
'm, u
1
` 0, 'm, u
2
` 0, 'm, u
3
` < 1, 'm, u
4
` < 0, 'm, u
5
` 2, 'm, u
6
` < 1,
we get the sets V
D,m
and V
supp
D,m
shown in Figure 2 (Exercise 9.1.5). The open circle
u
1
u
4
u
2
u
5
u
3
u
6
V
D,m

V
D,m
u
1
u
4
u
2
u
5
u
3
u
6
V
D,m
supp
V
D,m
supp
Figure 2. VD,m and V
supp
D,m
for D =D3 +2D5 D6 and m =e1
at the origin on the right side of the gure is a reminder that V
supp
D,m
does not contain
the origin.
The sets V
D,m
and V
supp
D,m
enable us to compute sheaf cohomology as follows.
Theorem 9.1.3. Let D =

be a Weil divisor on X

. Fix m M and p 0.
(a) H
p
(X

, O
X

(D))
m


H
p1
(V
D,m
, C).
(b) If D is Q-Cartier, then H
p
(X

, O
X

(D))
m


H
p1
(V
supp
D,m
, C).
Proof. For part (a), rst note that if , then
(9.1.8) V
D,m
= Conv(u

[ (1), 'm, u

` <a

).
9.1. Cohomology of Toric Divisors 403
One inclusion is trivial; the other follows easily using Lemma 1.2.7. Hence, for all
, we have
(9.1.9)
H
0
(U

, O
X

(D))
m
=0 'm, u

` a

for all (1)


V
D,m
=,
where the rst equivalence uses (9.1.2) and the second follows from (9.1.8).
The equation (9.1.8) shows that V
D,m
is convex and hence connected when
it is nonempty. It follows that H
0
(V
D,m
, C) = C when V
D,m
= . Since
H
0
(U

, O
X

(D))
m
= C
m
when it is nonzero, the equivalences (9.1.9) give a
canonical exact sequence
(9.1.10) 0 H
0
(U

, O
X

(D))
m
C H
0
(V
D,m
, C) 0
for all . It follows that for every p 0, we get an exact sequence
0

[]p
H
0
(U

, O
X

(D))
m

[]p
C

[]p
H
0
(V
D,m

, C) 0,
which we write as
0

C
p
(U , O
X

(D))
m
B
p
C
p
0.
The formula for the differential in the

Cech complex can be used to dene differ-
entials B
p
B
p+1
and C
p
C
p+1
. Then we get an exact sequence of complexes
0

C

(U , O
X

(D))
m
B

0.
Since

H
p
(U , O
X

(D)) H
p
(X

, O
X

(D)), the long exact sequence from (9.0.5)


becomes
0 H
0
(X

, O
X

(D))
m
H
0
(B

) H
0
(C

) H
1
(X

, O
X

(D))
m
.
In Exercises 9.1.6 and 9.1.7 you will use the theory of Koszul complexes to show
that the complex B

has very simple cohomology:


(9.1.11) H
p
(B

) =

C p = 0
0 p > 0.
Thus our long exact sequence breaks up into an exact sequence
0 H
0
(X

, O
X

(D))
m
C H
0
(C

) H
1
(X

, O
X

(D))
m
0
and isomorphisms
H
p1
(C

) H
p
(X

, O
X

(D))
m
, p 2.
We will show below that
(9.1.12) H
p
(C

) H
p
(V
D,m
, C), p 0.
For p 2, this gives the desired isomorphism

H
p1
(V
D,m
, C) = H
p1
(V
D,m
, C) H
p
(X

, O
X

(D))
m
.
404 Chapter 9. Sheaf Cohomology of Toric Varieties
When p = 1, we obtain the exact sequence
0 H
0
(X

, O
X

(D))
m
C H
0
(V
D,m
, C) H
1
(X

, O
X

(D))
m
0.
Since H
0
(X

, O
X

(D))
m
= 0 implies V
D,m
= (Exercise 9.1.8), we obtain
(9.1.13)

H
0
(V
D,m
, C) H
1
(X

, O
X

(D))
m

H
1
(V
D,m
, C) H
0
(X

, O
X

(D))
m
,
where the last line uses the exact sequence (9.0.11).
It remains to prove (9.1.12). Since [[ is the union of the maximal cones and
V
D,m
[[, we get the closed cover C =V
D,m

i
of V
D,m
, where the
i
are the
maximal cones of . Furthermore, C

is the

Cech complex

C

(C, C) for the


constant sheaf C on V
D,m
, where we use quotation marks since C is a closed cover
rather than an open cover.
Similar to the spectral sequence of an open covering discussed in 9.0, there
is a spectral sequence for the closed covering C = V
D,m

i
of V
D,m
. This E
1
spectral sequence is
E
p,q
1
=

[]p
H
q
(V
D,m

, C) H
p+q
(V
D,m
, C).
As explained in Theorem C.2.3 in Appendix C or [115, II.5.2], the differential d
p,q
1
is determined by the

Cech differential. In particular,
E
p,0
1
=

C
p
(C, C) =C
p
.
Also, we noted earlier in the proof that each V
D,m

is convex. It follows that


V
D,m

is contractible, so that
E
p,q
1
=

[]p
H
q
(V
D,m

, C) = 0, q > 0.
As in the proof of Theorem 9.0.4, this implies that
E
p,q
2
=

H
p
(

(C, C)) = H
p
(C

) q = 0
0 q > 0.
Similar to proof of Theorem 9.0.4, Proposition C.1.3 of Appendix C implies that
the edge homomorphisms are isomorphisms, i.e.,
H
p
(C

) H
p
(V
D,m
, C).
The proves (9.1.12) and completes the proof of part (a).
For part (b), we assume that D is Q-Cartier with support function
D
. Then
one can modify the proof of (9.1.9) to obtain the equivalence
H
0
(U

, O
X

(D))
m
=0 V
supp
D,m
=
9.1. Cohomology of Toric Divisors 405
(Exercise 9.1.9). This allows us to replace (9.1.10) with the exact sequence
0 H
0
(U

, O
X

(D))
m
C H
0
(V
supp
D,m
, C) 0.
From here, the proof is identical to what we did in part (a) (Exercise 9.1.9).
Example 9.1.4. Consider the divisor D =D
3
+2D
5
D
6
on the surface X

from
Example 9.1.2. Let us compute H
1
(X

, O
X

(D))
m
for m = e
1
. The sets V
D,m
and
V
supp
D,m
are shown in Figure 2 of Example 9.1.2. Since both sets have two connected
components, Theorem 9.1.3 implies that
H
1
(X

, O
X

(D))
m


H
0
(V
D,m
, C)

H
0
(V
supp
D,m
, C) C.
Each representation H
p
(X

, O
X

(D))
m
in Theorem 9.1.3 has its advantages.
For example, the vanishing theorems in 9.2 use V
supp
D,m
because of its relation to
convexity, while for surfaces, we will see below that V
D,m
is easier to work with.
Other treatments of H
p
(X

, O
X

(D))
m
can be found in [91], [141] and [226].
Also note that Theorem 9.1.3 can be stated using relative cohomology instead
of reduced cohomology. Consider, for example, the case when D is Q-Cartier.
Then the inclusion V
supp
D,m
[[ gives the long exact sequence
H
p
([[,V
supp
D,m
, C) H
p
([[, C) H
p
(V
supp
D,m
, C) .
Since [[ is contractible (it is a cone), this sequence and Theorem 9.1.3 imply
(9.1.14) H
p
(X

, O
X

(D))
m
H
p
([[,V
supp
D,m
, C), p 0
(Exercise 9.1.11). Since V
supp
D,m
is open in [[, algebraic geometers often write this
relative cohomology group as H
p
Z
D,m
([[, C), where Z
D,m
is the closed set
Z
D,m
=[[ `V
supp
D,m
=u [[ [ 'm, u`
D
(u).
Then part (b) of Theorem 9.1.3 can be restated as
H
p
(X

, O
X

(D))
m
H
p
Z
D,m
([[, C).
This is the version that appears in [105, p. 74].
The Surface Case. When X

is a complete toric surface, the set V


D,m
appearing
in Theorem 9.1.3 has an especially simple topology that can be encoded by a sign
sequence. We illustrate this with an example.
Example 9.1.5. We return to the situation of Example 9.1.4. Write the divisor
D = D
3
+2D
5
D
6
as D =

i
a
i
D
i
, and label the ray generated by u
i
with + if
'm, u
i
` a
i
and with if 'm, u
i
` <a
i
. Then the picture of V
D,m
from Figure 2
of Example 9.1.2 gives the sign pattern shown in Figure 3 on the next page.
If we start at u
1
and go counterclockwise around the origin, we get the sign
pattern +++, which we regard as cyclic. The positions of the s determine
V
D,m
, so that connected components of V
D,m
correspond to strings of consecutive
s in the sign pattern.
406 Chapter 9. Sheaf Cohomology of Toric Varieties
u
1
u
4
u
2
u
5
u
3
u
6
V
D,m

V
D,m
+

+
+
Figure 3. The sign pattern for D =D3 +2D5 D6 and m =e1 in Example 9.1.5
These observations hold in general. Let X

be a complete toric surface with


minimal generators u
1
, . . . , u
r
arranged counterclockwise around the origin. Given
a torus-invariant Weil divisor D=

i
a
i
D
i
on X

and mMZ
2
, the sign pattern
sign
D
(m) is the string of length r whose ith entry is + if 'm, u
i
` a
i
and if
'm, u
i
` < a
i
. We regard sign
D
(m) as cyclic. Thus, for example, ++++
has only one string of consecutive s.
We have the following result from [142] and [154] (Exercise 9.1.12).
Proposition 9.1.6. Given a Weil divisor D =

i
a
i
D
i
on a complete toric surface
X

, the dimension of H
p
(X

, O
X

(D))
m
is given by
dim H
0
(X

, O
X

(D))
m
=

1 if sign
D
(m) = + +(V
D,m
=)
0 otherwise
dim H
1
(X

, O
X

(D))
m
= max(0, #connected components of V
D,m
1)
= max(0, #strings of consecutive s in sign
D
(m) 1)
dim H
2
(X

, O
X

(D))
m
=

1 if sign
D
(m) = (V
D,m
is a cycle)
0 otherwise.
Example 9.1.7. Let us revisit Example 9.1.1. For each m M = Z
2
, the minimal
generators u
0
, u
1
, u
2
give a sign pattern for the line bundle O
P
2 (a) = O
P
2 (aD
0
).
When a < 0, all possible sign patterns are recorded in Figure 1 of Example 9.1.1.
Using Proposition 9.1.6, we obtain:
H
0
(P
2
, O
P
2 (a)) = 0 since +++ does not appear.
H
1
(P
2
, O
P
2 (a)) = 0 since all patterns have one string of consecutive s.
H
2
(P
2
, O
P
2 (a)) =

mInt(a
2
)
C
m
since labels the interior of a
2
.
Thus the computation of Example 9.1.1 follows immediately from Figure 1.
9.1. Cohomology of Toric Divisors 407
Example 9.1.8. Consider the surface X

and divisor D = D
3
+2D
5
D
6
from
Example 9.1.4. The sign patterns sign
D
(m) for all m M Z
2
give the chamber
decomposition shown in Figure 4 (Exercise 9.1.10).
+ +
+ + + + + + + + + +
+ + + +
+ + +
+ + +
+ +
+ + + +
+ + +
+ + +
+ + +
+ + +
++
l
6
l
2
l
5
l
3
l
4
= l
1
e
1 0
2e
1
e
1
+ e
2

+ + + +

+ + + +

+ + + +
+ + +

+ + + + +

Figure 4. The chamber decompostion for D =D3 +2D5 D6


As in Figure 1, we have lines l
i
dened by 'm, u
i
` = a
i
, where D =

i
a
i
D
i
.
We put the label l
i
on the side where 'm, u
i
` a
i
, and the shading indicates where
the boundary points lie. Each chamber is labeled with its sign pattern (some labels
are rather small), and we get ve 1-dimensional chambers since l
1
= l
4
.
We now compute cohomology using Proposition 9.1.6. First,
H
0
(X

, O
X

(D)) = H
2
(X

, O
X

(D)) = 0
since neither ++++++ nor appear anywhere. Furthermore, of the
six chambers with sign patterns having more than one string of consecutive s,
three of them (+++, ++, plus the 1-dimensional +++)
have no lattice points. Of the remaining three, the lattice points are:
m = 0 from ++++(1-dimensional).
m = e
1
, 2e
1
from +++.
m = e
1
+e
2
from ++++.
These lattice points are indicated with white dots in Figure 4. Hence
H
1
(X

, O
X

(D)) C
0
C
e
1
C
2e
1
C
e
1
+e
2
.
In particular, dim H
1
(X

, O
X

(D)) = 4. Exercise B.8.2 explains how to check this


computation using Macaulay2 [123].
408 Chapter 9. Sheaf Cohomology of Toric Varieties
The paper [142] applies these methods to classify line bundles with vanishing
higher cohomology on the toric surface coming from three consecutive blowups of
the Hirzebruch surface H
2
.
Exercises for 9.1.
9.1.1. In Example 9.1.1, verify the sign patterns in Figure 1.
9.1.2. Use the methods of Example 9.1.1 to prove (9.1.7).
9.1.3. The canonical bundle of P
2
is
P
2 =O
P
2 (3). Use this and Serre duality to explain
the symmetry in Table 2 in Example 9.1.1.
9.1.4. Recall that a Weil divisor D =

on X

is Q-Cartier if some positive integer


multiple of D is Cartier. Let
D
be the support function of D as in Theorem 4.2.12.
(a) Show that (1/)
D
: [[ R is a support function (Denition 4.2.11) and depends
only on D. We dene the support function of D to be
D
= (1/)
D
.
(b) Show that D is Q-Cartier if and only if for every , there is m

M
Q
such that
'm

, u

` =a

for all (1). When this is satised, show that


D
(u) ='m

, u` for
all u .
9.1.5. Verify Figure 2 in Example 9.1.2
9.1.6. Given a ring R and elements f
1
, . . . , f

R, dene
d
p
:

p
R

p+1
R

by d
p
() = (

i=1
f
i
e
i
) , where e
1
, . . . , e

are the standard basis of R

. Setting K
p
=

p
R

, we get the complex


K

: K
0
d
0
K
1
d
1

d
2
K
1
d
1
K

,
and for an R-module M, we get the Koszul complex
K

( f
1
, . . . , f

; M) = K

R
M.
Thus K

= K

( f
1
, . . . , f

; R).
(a) Given : R

R, show that there are maps s


p
: K
p
K
p1
such that s
1
= and
s
p+q
() = s
p
() +(1)
p
s
q
() for all K
p
and K
q
.
(b) Show that for all K
p
, the maps s
p
satisfy
d
p1
(s
p
()) +s
p+1
(d
p
()) = (

i=1
f
i
e
i
).
(c) Now assume that ' f
1
, . . . , f

` = R. Prove that K

( f
1
, . . . , f

; M) is exact for every R-


module M. Hint: Dene (e
i
) = g
i
where

i=1
g
i
f
i
= 1.
9.1.7. When f
1
= = f

= 1 and M is an R-module, the Koszul complex K

(1, . . . , 1; M)
of Exercise 9.1.6 becomes
(9.1.15) M

1i
M

1i<j
M

1i<j<k
M ,
where the entries of the matrices representing the differentials are 0 or 1.
(a) Use the previous exercise to prove that (9.1.15) is exact.
9.2. Vanishing Theorems I 409
(b) Prove that the complex B

dened in the proof of Theorem 9.1.3 satises (9.1.11).


Hint: Show B

is the Koszul complex (9.1.15) with M=C, minus its rst term.
9.1.8. Prove that H
0
(X

, O
X
(D))
m
= 0 if and only if V
m,D
=. Then prove (9.1.13).
9.1.9. Complete the proof of part (b) of Theorem 9.1.3.
9.1.10. Verify Figure 4 in Example 9.1.8.
9.1.11. Prove (9.1.14).
9.1.12. Prove Proposition 9.1.6.
9.1.13. Prove that H
0
(X

, O
X
(D))
m
=0 implies that H
p
(X

, O
X
(D))
m
=0 for all p >0.
Hint: Exercise 9.1.8.
9.2. Vanishing Theorems I
In this section we prove two basic vanishing theorems for Q-Cartier divisors on a
toric variety X

. We then apply these results to show that normal toric varieties are
Cohen-Macaulay and hence satisfy Serre duality. We also give a version of Serre
duality that is special to the toric case.
Nef Q-Cartier Divisors. A Weil divisor D on a normal variety X is Q-Cartier if
some positive integer multiple is Cartier. Also recall that the intersection product
D C is dened whenever D is Q-Cartier and C is a complete irreducible curve in
X. Then D is nef if D C 0 for all complete irreducible curves C X. This
generalizes the denition of nef Cartier divisor given in 6.3.
In the toric case, we characterize Q-Cartier nef divisors as follows.
Lemma 9.2.1. Let D =

be a Q-Cartier divisor on X

. If has convex
support, then the following are equivalent:
(a) D is nef.
(b) For some integer > 0, D is Cartier and basepoint free, i.e., O
X

(D) is a
line bundle generated by its global sections.
(c) The support function
D
: [[ R is convex.
Proof. The proof combines ideas from Theorems 6.1.7, 6.3.12 and 7.2.2. We leave
the details to the reader (Exercise 9.2.1).
In the Cartier case, we know that D is nef if and only if O
X

(D) is generated
by global sections (Theorems 6.1.7 and 6.3.12). This fails in the Q-Cartier case, as
shown by the following example.
Example 9.2.2. Let X

=P(2, 3, 5) and let D


0
, D
1
, D
2
be the divisors given by the
minimal generators u
0
, u
1
, u
2
. Note that 2u
0
+3u
1
+5u
2
= 0. Then Cl(X

) Z,
where the classes of D
0
, D
1
, D
2
map to 2, 3, 5 respectively. Let D = D
1
D
0
and
note that D
0
2D, D
1
3D, D
2
5D. It also easy to see that D is Q-Cartier and
410 Chapter 9. Sheaf Cohomology of Toric Varieties
nef. However, you will check in Exercise 9.2.2 that H
0
(X

, O
X

(D)) = 0. Thus
O
X

(D) is not generated by its global sections.


Demazure Vanishing. The most basic vanishing theorem for toric varieties is the
following result, rst proved by Demazure [82] for Cartier divisors.
Theorem 9.2.3 (Demazure Vanishing). Let D be a Q-Cartier divisor on X

. If [[
is convex and D is nef, then
H
p
(X

, O
X

(D)) = 0 for all p > 0.


Proof. The support function
D
: [[ R is convex by Lemma 9.2.1. Fix m M
and let V
supp
D,m
= u [[ [ 'm, u` <
D
(u) as in Theorem 9.1.3. Let u, v V
supp
D,m
,
so that 'm, u` <
D
(u) and 'm, v` <
D
(u). Then, for 0 t 1, we have
'm, (1t)u+tv` = (1t)'m, u` +t'm, v`
< (1t)
D
(u) +t
D
(v)

D
((1t)u+tv).
The last line follows since
D
is convex. This implies that V
supp
D,m
is convex and
hence contractible. Combining this with Theorem 9.1.3, we obtain
H
p
(X

, O
X

(D))
m


H
p1
(V
supp
D,m
, C) = 0 for all p > 0.
In particular, if D is a basepoint free Cartier divisor on X

, then D is nef, so
that Theorem 9.2.3 implies
H
p
(X

, O
X

(D)) = 0 for all p > 0.


Example 9.2.4. For P
n
, aD
0
is basepoint free for a 0. Hence the higher co-
homology of O
P
n (a) = O
P
n (aD
0
) vanishes by Theorem 9.2.3, which agrees with
what we found in Example 9.1.1 when n = 2.
Vanishing of Higher Direct Images. Here is an easy application of Demazure
vanishing.
Theorem 9.2.5. Let : X

1
X

2
be a proper toric morphism and let L be a line
bundle on X

2
Then:
(a) R
p

L = 0 for all p > 0.


(b) If the map : N
1
N
2
on the lattices of one-parameter subgroups is surjective,
then

L L.
Proof. First assume L = O
X

2
, so that

L = O
X

1
. Let U

2
be the afne
open corresponding to
2
. By Proposition 9.0.7, R
p

O
X

1
[
U
is the sheaf
associated to the module H
p
(
1
(U

), O
X

1
). Since is proper, Theorem 3.4.11
tells us that
1
(U

) is a toric variety whose fan is supported on the convex set


9.2. Vanishing Theorems I 411

1
R
(). Thus H
p
(
1
(U

), O
X

1
) = 0 for p > 0 by Theorem 9.2.3. It follows that
R
p

O
X

1
= 0 for p > 0 by Proposition 9.0.7.
The morphism induces a homomorphism of sheaves O
X

O
X

1
on X

2
(Example 4.0.25). It sufces to show that this map is an isomorphism on each afne
open U

2
,
2
. Note also that the restriction
1
(U

) U

is again a
proper toric morphism. Hence we may assume that X

2
=U

. Then [
1
[ =
1
R
()
is a convex cone, so that
H
0
(X

1
, O
X

1
) =

m[
1
[

M
1
C
m
by Exercise 4.3.4. We analyze [
1
[

M
1
as follows.
The surjection N
1
N
2
gives an injection M
2
M
1
on character lattices. Then
(N
2
)
R
has dual

(M
2
)
R
(M
1
)
R
and [
1
[ (N
1
)
R
has dual [
1
[

(M
1
)
R
. Also, [
1
[ =
1
R
() implies

=[
1
[

. It follows that
[
1
[

M
1
=

M
1
=

(M
2
)
R
M
1
=

M
2
,
where the last equality follows since M
1
is saturated in M
2
by the surjectivity of
N
1
N
2
. Hence
H
0
(X

1
, O
X

1
) =

M
2
C
m
= H
0
(U

, O
U
).
This shows that O
U

O
X

1
induces an isomorphism on global sections, which
gives the desired sheaf isomorphism since we are working with quasicoherent
sheaves over an afne variety.
For a line bundle L on X

2
, take U

2
. Then L[
U
O
X

2
[
U
, so that
R
p

L[
U
R
p

O
X

2
[
U
R
p

O
X

1
[
U
= 0
for p > 0. Furthermore, when : N
1
N
2
is surjective, we have

O
X

O
X

2
L L,
where the rst isomorphism is part (a) of Exercise 9.2.3 and the second follows
from

O
X

1
O
X

2
.
The Injectivity Lemma. We introduce a method that will be used several times in
this section and the next. The general framework uses a morphism f : Y X and
coherent sheaves G on Y and F on X such that:
f is an afne morphism, meaning f
1
(U) X is afne when U Y is afne.
There is a split injection i : F f

G of O
X
-modules. This means that there is
a homomorphism r : f

G F such that r i = 1
F
.
By functoriality, a split injection i : F f

G induces a split injection


H
p
(X, F) H
p
(X, f

G).
412 Chapter 9. Sheaf Cohomology of Toric Varieties
However, since f is afne, we also have
H
p
(X, f

G) H
p
(Y, G)
by Exercise 9.0.5. Hence we get an injection
H
p
(X, F) H
p
(Y, G).
In particular, H
p
(Y, G) = 0 implies H
p
(X, F) = 0. This method is used in the
proofs of many vanishing theoremssee [186, Ch. 4].
In the toric context, Fujino [101] identied the best choice for the morphism
f . Given a fan in N
R
and a positive integer , let

: N N be multiplication
by . This maps to itself and hence induces a toric morphism

: X

. The
dual of

is multiplication by on M, and the restriction of

to T
N
= M
Z
C

is the group homomorphism given by raising to the th power. Furthermore, given


any , we have
1

(U

) =U

since (

)
1
R
() = . This shows that

is an
afne morphism.
Here is our rst injectivity lemma.
Lemma 9.2.6. Let D be a Weil divisor on X

and let be a positive integer. Then


there is an injection
H
p
(X

, O
X

(D)) H
p
(X

, O
X

(D)) for all p 0.


Proof. We will construct a split injection O
X

(D)

O
X

(D). The lemma


then follows immediately from the above discussion.
For D=

and , let P=mM


R
[ 'm, u` a

for all (1).


Over U

, the sheaves O
X

(D) and O
X

(D) come from the C[

M]-modules

mPM
C
m
and

m(P)M
C
m
.
Since
1

(U

) = U

O
X

(D) is determined by

m(P)M
C
m
with the
module structure given by
m
a =
m
a. This implies that the map
(9.2.1)

mPM
C
m

m(P)M
C
m
that sends
m
to
m
is a homomorphism of C[

M]-modules (Exercise 9.2.4).


Furthermore, one can show that the map given by
(9.2.2) r(
m
) =

0 m / M

m = m

, m

M
denes a C[

M]-module splitting of (9.2.1) (Exercise 9.2.4).


The map (9.2.1) and the splitting r are easily seen to be compatible with the
inclusion U

when is a face of . It follows that the maps (9.2.1) patch to


give the split injection O
X

(D)

O
X

(D).
9.2. Vanishing Theorems I 413
Batyrev-Borisov Vanishing. In Example 9.1.1, we saw that O
P
2 (a) has nontrivial
H
2
when a 3. If we write a =b for b > 0, we can rewrite (9.1.6) as
dim H
p
(P
2
, O
P
2 (b)) =

0 p = 2

Int(b
2
) M

p = 2.
This has been generalized by Batyrev and Borisov [18]. Recall that a Weil divisor
D =

gives the polyhedron


P
D
=m M
R
[ 'm, u

` a

for all (1),


which is a polytope when is complete (Proposition 4.3.8).
Theorem 9.2.7 (Batyrev-Borisov Vanishing). Let D =

be a Q-Cartier
divisor on a complete toric variety X

. If D is nef, then
(a) H
p
(X

, O
X

(D)) = 0 for all p = dim P


D
.
(b) When p = dim P
D
, H
p
(X

, O
X

(D))

mRelint(P
D
)M
C
m
.
Proof. First assume that D is Cartier and dim P
D
= dim N
R
= dim X

. Then D is
basepoint free since it is nef (Theorem 6.3.12). By Proposition 6.2.5 and Theo-
rem 6.2.8, combining cones of that share the same linear functional relative to

D
gives a fan
D
such that renes
D
and
D
is strictly convex relative to
D
.
For m M, H
p
(X

, O
X

(D))
m


H
p1
(V
supp
D,m
, C) by Theorem 9.1.3. Set
Z
D,m
= N
R
`V
supp
D,m
=u N
R
[ 'm, u`
D
(u).
Since
D
is strictly convex relative to
D
, the proof of Theorem 9.2.3 implies that
Z
D,m
is convex. In fact, it is strongly convex. To see why, note that for any nonzero
u N
R
, the strict convexity of
D
implies 0 =
D
(u+(u)) >
D
(u) +
D
(u)
since u and u do not lie in the same cone of
D
. Thus

D
(u) =
D
(u) >
D
(u) for all u = 0 in N
R
.
Now assume u Z
D,m
`0. Then 'm, u`
D
(u), so that

D
(u) 'm, u`.
Combining the displayed inequalities gives
D
(u) > 'm, u`, which implies
that u V
supp
D,m
= N
R
`Z
D,m
. Hence Z
D,m
is strongly convex.
We next prove that
(9.2.3) Z
D,m
=0 m Int(P
D
) M.
If m Int(P
D
) M, then m Int(P
D
), so that
'm, u

` >a

=
D
(u

) for all (1).


Thus
(9.2.4) 'm, u

` <
D
(u

) for all (1),


414 Chapter 9. Sheaf Cohomology of Toric Varieties
which easily implies that Z
D,m
= 0 (Exercise 9.2.5). On the other hand, if we
have m / Int(P
D
) M, then one of the inequalities of (9.2.4) must fail, i.e.,
'm, u

0
`
D
(u

0
) for some
0
(1).
Then u

0
Z
D,m
, so that Z
D,m
=0. This proves (9.2.3).
We are now ready to compute H
p
(X

, O
X

(D))
m
. If m Int(P
D
) M, then
V
supp
D,m
= N
R
`0 is homotopic to S
n1
. Hence
H
p
(X

, O
X

(D))
m


H
p1
(V
supp
D,m
, C)

H
p1
(S
n1
, C) =

0 p = n
C p = n.
If m / Int(P
D
) M, then V
supp
D,m
= R
n
`Z
D,m
, where Z
D,m
is a closed strongly
convex cone of positive dimension. If u = 0 in Z
D,m
, we showed above that
u V
supp
D,m
. It is easy to see that V
supp
D,m
contracts to u (Exercise 9.2.6). Hence
H
p
(X

, O
X

(D))
m


H
p1
(V
supp
D,m
, C) = 0 for all p 0.
This completes the proof when D is Cartier and P
D
has dimension n.
Now suppose D is Cartier but dim P
D
< n. We will use Proposition 6.2.5 and
Theorem 6.2.8. After translation, P
D
spans (M
D
)
R
M
R
, where M
D
M is dual to
a surjection : N N
D
, and the normal fan of P
D
lies in (N
D
)
R
. If X
D
is the toric
variety of P
D
(M
D
)
R
, then Theorem 6.2.8 shows that induces a toric morphism
: X

X
D
such that D is the pullback of the ample divisor D

on X
D
determined
by P
D
. In particular, O
X

(D)

O
X
D
(D

).
Since is proper and : N N
D
is surjective, we have
R
p

O
X

(D) = 0, p > 0

O
X

(D) =O
X
D
(D

)
by Theorem 9.2.5. Then Proposition 9.0.8 implies that
(9.2.5) H
p
(X
D
, O
X
D
(D

)) H
p
(X

, O
X

(D)).
This proves the theorem, as we now explain. Since M
D
is saturated in M by the
surjectivity of N N
D
, we have
P
D
M = P
D
(M
D
)
R
M = P
D
M
D
,
Since P
D
= P
D
is full dimensional in (M
D
)
R
, we are done by (9.2.5) and the full
dimensional case considered above.
Finally, we need to consider what happens when Dis Q-Cartier. Pick an integer
> 0 such that D is Cartier. By Lemma 9.2.6, we have an injection
H
p
(X

, O
X

(D)) H
p
(X

, O
X

(D)).
9.2. Vanishing Theorems I 415
Since D is Cartier and nef with P
D
= P
D
, the Cartier case proved above implies
that H
p
(X

, O
X

(D)) = 0 for p = dim P


D
. Furthermore, when p = dim P
D
,
H
p
(X

, O
X

(D)) =

mM

H
p1
(V
supp
D,m
, C)
m
=

mRelint(P
D
)M
C
m
,
from which we conclude

H
p1
(V
supp
D,m
, C) =

C m Relint(P
D
) M
0 otherwise.
Since V
supp
D,m
=V
supp
D,m
and m Relint(P
D
) M m Relint(P
D
) M,
H
p
(X

, O
X

(D)) =

mM

H
p1
(V
supp
D,m
, C)
m
=

mRelint(P
D
)M
C
m
.
Example 9.2.8. We saw in Example 9.2.4 that the higher cohomology of O
P
n (a)
vanishes when a 0. When a < 0, O
P
n (a) = O
P
n ([a[D
0
). Since [a[D
0
is ample
with polytope [a[
n
, Theorem 9.2.7 implies
dim H
p
(P
n
, O
P
n (a)) =

0 p = n

Int(a
n
) M

p = n
when a < 0. This generalizes (9.1.6) from Example 9.1.1.
The Cohen-Macaulay Property. Here is a surprising consequence of Batyrev-
Borisov vanishing.
Theorem 9.2.9. A normal toric variety is Cohen-Macaulay.
Proof. First suppose that X

is projective and let D be a torus-invariant ample


divisor on X

. Corollary 5.72 of [179, p. 182] implies that X

is Cohen-Macaulay
if and only if
(9.2.6) H
p
(X

, O
X

(D)) = 0 for all 0, p < n.


Now take any integer > 0. Since D is ample, the divisor D is nef and its
polytope P
D
has dimension n =dim X

. Then (9.2.6) follows from Theorem 9.2.7.


We conclude that X

is Cohen-Macaulay.
Next consider an afne toric variety U

. We can nd a projective toric variety


that contains U

as an afne open subset (Exercise 9.2.7). Since being Cohen-


Macaulay is a local property, it follows that U

is Cohen-Macaulay, and then any


normal toric variety X

is Cohen-Macaulay.
This result was originally proved by Hochster [146]. Another proof can be
found in [76, Thm. 3.4]. A projective variety is called arithmetically Cohen-
Macaulay (aCM for short) if its afne cone is Cohen-Macaulay. In Exercise 9.2.8
you will show that normal lattice polytopes give aCM toric varieties.
416 Chapter 9. Sheaf Cohomology of Toric Varieties
Serre Duality. Theorem 9.2.9 shows that Serre duality holds for any normal toric
variety. However, the version stated in Theorem 9.0.9 only applies to locally free
sheaves, while the more general version Theorem 9.0.10 uses Ext groups. Many
of the sheaves we deal with, such as O
X

(D) and

p
X

, fail to be locally free when


X

is not smooth. Fortunately, there is an Ext-free version of Serre duality that


holds for these sheaves.
Theorem 9.2.10 (Toric Serre Duality). Assume that X

is a complete toric variety


of dimension n.
(a) If D is a Q-Cartier divisor and K
X

is the canonical divisor, then we have


natural isomorphisms
H
p
(X

, O
X

(D))

H
np
(X

, O
X

(K
X

D)).
(b) If X

is simplicial and F is locally free, then there are natural isomorphisms


H
p
(X

q
X

O
X

F)

H
np
(X

nq
X

O
X

).
Remark 9.2.11. Here are two interesting aspects of Theorem 9.2.10:
(a) O
X

(K
X

D) need not be isomorphic to

O
X

O
X

(D)

=O
X

(K
X

)
O
X

O
X

(D)
when D is Q-Cartier. So Theorem 9.2.10 does not follow from Theorem 9.0.9.
(b) Every Weil divisor is Q-Cartier on a simplicial toric variety. Theorem 9.2.10
holds for all Weil divisors in this case.
Before beginning the proof of Theorem 9.2.10, we introduce some tools from
commutative algebra. This is typical of algebraic geometryonce a variety ceases
to be smooth, the theory becomes more technically demanding.
In 9.0 we dened the depth of a local ring (R, m). More generally, the depth of
a nitely generated R-module F is the maximal length of a sequence f
1
, . . . , f
s
m
such that f
i
is not a zero divisor in F/' f
1
, . . . , f
i1
`F for all i, and the dimension of
F is dim R/Ann(F), where Ann(F) =f R [ f F = 0.
Then we dene a coherent sheaf F on a variety X to be maximal Cohen-
Macaulay (MCM for short) if for every point x X, the stalk F
x
is an O
X,x
-
module whose depth and dimension both equal dim X. For example, if X is Cohen-
Macaulay, then every locally free sheaf on X is MCM.
The following result from [226, Prop. 4.24] shows that MCM sheaves satisfy a
nice version of Serre duality. See also [179, Thm. 5.71].
Theorem 9.2.12 (Serre Duality III). Let F be a MCM sheaf on a complete normal
Cohen-Macaulay variety X of dimension n. There there are natural isomorphisms
H
p
(X, F)

H
np
(X, Hom
O
X
(F,
X
)).
9.2. Vanishing Theorems I 417
Proof. We will use the sheaf version of Ext, where the sheaves Ext
q
O
X
(F,
X
) are
the derived functors of F Hom
O
X
(F,
X
). The stalk at x X is
(9.2.7) Ext
q
O
X
(F,
X
)
x
= Ext
q
O
X,x
(F
x
, (
X
)
x
).
The relation between the Ext groups Ext
q
O
X
(F,
X
) and Ext sheaves Ext
q
O
X
(F,
X
)
is described by the spectral sequence
E
p,q
2
= H
p
(X, Ext
q
O
X
(F,
X
)) Ext
p+q
O
X
(F,
X
)
from Theorem C.2.4 of Appendix C.
Since F is MCM, [56, Cor. 3.5.11] implies that Ext
q
O
X,x
(F
x
, (
X
)
x
) = 0 for
q > 0 and x X. By (9.2.7), we obtain Ext
q
O
X
(F,
X
) = 0 for q > 0. Hence
H
p
(X, Hom
O
X
(F,
X
)) Ext
p
O
X
(F,
X
)
by Proposition C.1.7 of Appendix C. Since H
p
(X, F)

Ext
np
O
X
(F,
X
) by the
version of Serre duality given in Theorem 9.0.10, we are done.
We can now prove the toric version of Serre duality.
Proof of Theorem 9.2.10. We rst show that O
X

(D) is MCM when D is Q-


Cartier. Pick > 0 such that D is Cartier. We will use splitting methods, though
for clarity we replace

: N N with the sublattice N N. The dual lattice

1
M M gives semigroup algebras S
,N
=C[

M] S
,N
=C[

1
M].
Over an afne open subset U

, we can pick m


1
M such that
D
(u) =
'm

, u` for u . Now consider the S


,N
-module A = (U

, O
X

(D)). If we set
P = m

, then one easily sees that


(9.2.8) A =

mPM
C
m
B =

mP
1
M
C
m
=
m
S
,N
,
where the last equality uses m


1
M. Note that B =
m
S
,N
is free over S
,N
and hence is MCM over S
,N
. However, S
,N
is nitely generated as a S
,N
-
module (Exercise 9.2.9), so that B is MCM over S
,N
by [56, Ex. 1.2.26].
Another property of MCM modules is that any nontrivial direct summand of an
MCM module is again MCM. This follows from [56, Thm. 3.5.7]. Hence it sufces
to split (9.2.8). In this situation, we use the map r : B A that sends
m
B to
r(
m
) =

m
if
m
A
0 otherwise.
Similar to what we did in Lemma 9.2.6, r is a homomorphism of S
,N
-modules.
From here, it follows easily that O
X

(D) is MCM when D is Q-Cartier.


418 Chapter 9. Sheaf Cohomology of Toric Varieties
Now we prove duality. Since O
X

(D) is MCM, Theorem 9.2.12 implies that


H
p
(X

, O
X

(D))

H
np
(X

, Hom
O
X

(O
X

(D),
X

))
= H
np
(X

, Hom
O
X

(O
X

(D), O
X

(K
X

))).
However, Hom
O
X
(O
X
(D), O
X
(E)) O
X
(E D) for any Weil divisors D, E on a
normal variety X (Exercise 9.2.10). This gives the desired duality for O
X

(D).
For part(b), we rst show that

q
X

is MCM when X

is simplicial. Since is
complete, we can work locally over U

, dim = n. The minimal generators of


form a basis of a sublattice N

N of nite index such that is smooth relative to


N

. Let M

be the dual of N

. As above, we get semigroup algebras S


,N
S
,N
.
By Proposition 8.2.18, the restriction of

q
X

to U

= U
,N
is determined by
the S
,N
-module
A =

q
V

(m)
m
,
where V

(m) = Span
C
(m
0
M[ m
0
the minimal face of

containing m). Now


consider the larger S
,N
-module dened by
B =

q
V

(m)
m
.
Since V

(m) is unaffected when M is replaced by M

, we see that as a S
,N
-module,
B = (U
,N
,

q
U
,N

).
However,

q
U
,N

is locally free since U


,N
is smooth by our choice of N

. Hence
B is MCM over S
,N
. Arguing as in part (a), B is MCM over S
,N
and we have a
splitting map r dened by
r(
m
) =

m
if m

M
0 otherwise.
Thus

q
X

is MCM, and then


q
X

O
X

F is MCM since F is locally free.


The proof is now easy to nish, since Theorem 9.2.12 implies
H
p
(X

q
X

O
X

F)

H
np
(X

, Hom
O
X

q
X

O
X

F,
X

))
However, using the local freeness of F and Exercise 8.0.13, we have
Hom
O
X

q
X

O
X

F,
X

) Hom
O
X

q
X

,
X

)
O
X

nq
X

O
X

.
From here, the theorem follows easily.
The proof of part (a) of the theorem was inspired by [76, Lem. 3.4.2]. Other
proofs that Q-Cartier divisors give MCM sheaves can be found in [54, Cor. 4.2.2]
and [226, Prop. 4.22]. In part (b) we followed [76, Prop. 4.8].
9.2. Vanishing Theorems I 419
We should also mention that besides Q-Cartier divisors, there can be other
Weil divisors that give MCM sheaves. For example,
X
is MCM on any normal
Cohen-Macaulay variety (see [56, Def. 3.3.1]), so that
X

is MCM on any toric


variety X

, even when the canonical class K


X

fails to be Q-Cartier. You will give


a toric proof of this in Exercise 9.2.11.
In Exercise 9.2.12 you will use Alexander duality to give a purely toric proof
of Serre duality for simplicial toric varieties.
Exercises for 9.2.
9.2.1. Prove Proposition 9.2.1.
9.2.2. Verify the claims made in Example 9.2.2.
9.2.3. We saw in Example 4.0.25 a morphism of varieties f : X Y induces a homomor-
phism O
Y
f

O
X
of O
Y
-modules.
(a) Let L be a line bundle on Y. Construct an isomorphism f

L ( f

O
X
)
OY
L
of O
Y
-modules. Hint: Construct a homomorphism and then study the homomorphism
over open subsets of Y where L is trivial.
(b) Generalize part (b) by showing that for any sheaf G of O
X
-modules on X and any line
bundle L on Y, there is an isomorphismf

(G
OX
f

L) f

G
OY
L. This is the
projection formula.
9.2.4. Consider the map

mPM
C
m

m(P)M
C
m
from (9.2.1), where the
C[

M]-module structure on

m(P)M
C
m
is given by
m
a =
m
a. Prove that
(9.2.1) and (9.2.2) are C[

M]-module homomorphisms.
9.2.5. Prove that (9.2.4) implies Z
D,m
=0, as claimed in the proof of Theorem 9.2.7.
9.2.6. As in the proof of Theorem 9.2.7, assume that Z
D,m
is strongly convex and that u
Z
D,m
is nonzero. Thus u V
supp
D,m
. Prove that for every v V
supp
D,m
, the line segment uv
is contained in V
supp
D,m
(this means that V
supp
D,m
is star shaped with respect to u). Conclude
that the constant map : V
supp
D,m
u is a contraction.
9.2.7. Prove that every afne toric variety U

is contained in a projective toric variety X

.
9.2.8. The lattice points of a normal lattice polytope P give a projective embedding of the
toric variety X
P
. Prove that X
P
is aCM in this embedding.
9.2.9. Assume that N
1
N
2
has nite index, with dual M
2
M
1
, and let be a cone in
(N
1
)
R
= (N
2
)
R
. This gives semigroup algebras S
,N2
=C[

M
2
] S
,N1
=C[

M
1
].
Prove that S
,N1
is nitely generated as a module over S
,N2
. Hint: Let m
1
, . . . , m
r
M
2
generate

and consider

r
i=1

i
m
i
[ 0
i
< 1 M
1
.
9.2.10. Let D, E be Weil divisors on a normal variety X and let U X be the smooth locus.
Then (X, O
X
(D)) (U, O
X
(D)[
U
), and the same holds for E.
(a) Construct a natural map
X
: (X, O
X
(E D)) Hom
OX
(O
X
(D), O
X
(E)) and prove
that
X
is an isomorphism when X is smooth.
(b) Prove that Hom
OX
(O
X
(D), O
X
(E)) is reexive. Hint: Let U is the smooth locus and
study the restriction map Hom
OX
(O
X
(D), O
X
(E)) Hom
OU
(O
X
(D)[
U
, O
X
(D)[
U
).
420 Chapter 9. Sheaf Cohomology of Toric Varieties
(c) Show that
X
is an isomorphism and that O
X
(E D) Hom
OX
(O
X
(D), O
X
(E)).
9.2.11. Following [76, Cor. 3.5], you will show that
X
is MCM on a normal toric variety.
Fix a cone N
R
and pick a basis e
1
, . . . , e
n
of M such that e
n
is in the interior of

. Let
=lcm('e
n
, u

` [ (1)) and let M be the lattice with basis e


1
, . . . , e
n1
,
1
e
n
, with dual
N N. By Proposition 8.2.9,
U
comes fromthe ideal

mInt(

)M
C
m
C[

M].
(a) Prove that u

= (/'e
n
, u

`)u

lies in N. Then use '


1
e
n
, u

` = 1 to show that u

is
the minimal generator of with respect to N.
(b) Conclude that the canonical divisor of U
,N
is Cartier.
(c) Construct a splitting of

mInt(

)M
C
m

mInt(

)M
C
m
and conclude that
the canonical sheaf of U
,N
is MCM.
9.2.12. Alexander duality (see [210, 71]) states that if A S
n1
is a closed subset such
that the pair (S
n1
, A) is triangulable (see [210, p. 150]), then

H
p1
(A, C)


H
np1
(S
n1
` A, C).
You will prove Serre duality when D =

on a complete simplicial toric variety X

of dimension n. Let K = K
X
be the canonical divisor of X

. Theorem 9.1.3 implies


H
p
(X

, O
X
(D))
m


H
p1
(V
D,m
, C), m M.
Set A
D,m
=V
D,m
and

= Conv(u

[ (1)). Your goal is to prove that


(9.2.9) H
p
(X

, O
X
(D))

m
H
np
(X

, O
X
(K D))
m
.
(a) Explain why S =

is homeomorphic to S
n1
. Note that A
D,m
, A
KD,m
S.
(b) Prove that A
D,m
A
KD,m
= and that for , is contained in neither A
D,m
nor A
KD,m
if and only if there are
1
,
2
(1) such that 'm, u
1
` < a
1
and
'm, u
2
` a
2
. We say that is intermediate when this happens.
(c) Fix an intermediate cone . Let

be the face of

generated by u

s with
'm, u

` < a

and
+

be the face generated by u

s with 'm, u

` a

. Show
that every u

can be written uniquely as u = (1 t)u


+
+tu

where u
+

+

,
u

and 0 t 1. Then show that


+

is a deformation retract.
(d) Prove that A
KD,m
S ` A
D,m
is a deformation retract.
(e) Finally, prove (9.2.9) by applying Alexander duality to A
D,m
S.
This exercise was inspired by [76, Prop. 7.7.1] and [105, Sec. 4.4].
9.3. Vanishing Theorems II
Vanishing theorems play an important role in algebraic geometry. A glance at the
index of Lazarsfelds two-volume treatise [186] lists vanishing theorems due to
Bogomolov, Demailly, Fujita, Grauert-Riemenschneider,
Grifths, Kawamata-Viehweg, Kodaira, Koll ar, Le Potier,
Manivel, Miyoka, Nadel, Nakano, and Serre.
We will explore toric versions of several of these results. In some cases, the toric
version is stronger, which is to be expected since toric varieties are so special.
9.3. Vanishing Theorems II 421
Twisting. If D is a Cartier divisor and F a coherent sheaf on a normal variety X,
then the twist of F by D is the sheaf
F(D) =F
O
X
O
X
(D).
For example, if K
X
is a canonical divisor on X, then

X
(D) =
D

O
X
O
X
(D) O
X
(K
X
)
O
X
O
X
(D) O
X
(K
X
+D).
This notation will be used some of the vanishing theorems stated below. However,
when we start working with non-Cartier divisors, we will drop the twist notation.
Kodaira and Nakano. Two of the earliest vanishing theorems are due to Kodaira
and Nakano. For an ample divisor D on a smooth projective variety X of dimension
n, Kodaira vanishing asserts that
H
p
(X,
X
(D)) = 0, p > 0,
and Nakano vanishing states that
H
p
(X,
q
X
(D)) = 0, p+q > n.
Nakanos theorem generalizes Kodairas since
X
=
n
X
in the smooth case.
In the toric case, we get more vanishing, in what has become known as the
Bott-Steenbrink-Danilov vanishing theorem.
Theorem 9.3.1 (Bott-Steenbrink-Danilov Vanishing). Let D be an ample divisor
on a projective toric variety X

. Then for all p > 0 and q 0, we have


H
p
(X

q
X

(D)) = 0.
Proof. We give a proof due to Fujino [101] that uses the morphism

: X

from Lemma 9.2.6. We can assume that D is torus-invariant with support function

D
: N
R
R. Let L =O
X

(D) be the associated line bundle.


The morphism

has two key properties. The rst concerns the pullback

O
X

(D). Proposition 6.2.7 implies that

O
X

(D) comes from a divisor whose


support function is
D

. Since

is multiplication by ,
D

is the support
function of D. Hence

L =

O
X

(D) O
X

(D) O
X

(D)

=L

.
The second key property of

is that there is a split injection


(9.3.1)

q
X

q
X

.
We will assume this for now.
Given these properties of

, the theorem follows easily. Tensoring (9.3.1) with


L gives a split injection

q
X

q
X

L.
422 Chapter 9. Sheaf Cohomology of Toric Varieties
Since

q
X

q
X

L) (this is the projection formula from


Exercise 9.2.3) and

L L

, we obtain a split injection

q
X

q
X

).
As in the discussion leading up to Lemma 9.2.6, this gives the injection
H
p
(X

q
X

L) H
p
(X

q
X

).
When p >0, the right-hand side vanishes for sufciently large by Serre vanishing
(Theorem 9.0.6). Hence the left-hand side also vanishes for p > 0, which is what
we want.
It remains to prove (9.3.1). If we set D = 0 in the proof of Lemma 9.2.6, we
get a split injection i : O
X

O
X

. Recall that locally,


O
X

looks like O
X

, with module structure is given by


m
a =
m
a.
i sends
m
to
m
.
The splitting r :

O
X

O
X

is dened by
r(
m
) =

0 m / M

m = m

, m

M.
If we tensor i : O
X

O
X

with

q
M, we get a sheaf homomorphism
(9.3.2)

q
M
Z
O
X

q
M
Z
O
X

together with a splitting map


(9.3.3)

q
M
Z
O
X

q
M
Z
O
X

.
Thus (9.3.2) is a split injection.
By Theorem 8.2.16,

q
X

sits inside

q
M
Z
O
X

, so that

q
X

is a subsheaf
of

q
M
Z
O
X

. These subsheaves relate to (9.3.2) and (9.3.3) as follows.


Over U

, Theorem 8.2.18 implies


(U

q
X

q
V

(m)
m

q
M
C

m
= (U

q
M
Z
O
X

),
where V

(m) M
C
is spanned by the lattice points lying in the same minimal face
of

as m. Then over U

, we have:

q
X

looks like

q
X

with module structure


m
a =
m
a.
The map (9.3.2) takes

q
V

(m)
m
to

q
V

(m)
m
=

q
V

(m)
m
(U

q
X

)
since V

(m) =V

(m). Thus (9.3.2) induces a map (9.3.1).


9.3. Vanishing Theorems II 423
The map (9.3.2) takes

q
V

(m)
m
to 0 or, when m = m

, to

q
V

(m)
m

q
V

(m

)
m

(U

q
X

)
since V

(m) =V

(m

). Thus (9.3.3) induces a map

q
X

q
X

.
This gives the desired split injection (9.3.1).
Although Theorem 9.3.1 generalizes the vanishing theorems of Kodaira and
Nakano, its name Bott-Steenbrink-Danilov reects the more special vanishing
that happens for projective spaces (Bott [40]), weighted projective spaces (Steen-
brink [260]), and projective toric varieties (Danilov [76], stated without proof).
Theorem 9.3.1 was rst proved by Batyrev and Cox [19] in the simplicial case and
by Buch, Thomsen, Lauritzen and Mehta [58] in general. Further proofs have been
given by Fujino [101] (noted above) and Mustat a [212].
The Simplicial Case. When X

is simplicial, there is another vanishing theorem


involving the sheaves

q
X

.
Theorem 9.3.2. If X

is a complete simplicial toric variety, then


H
p
(X

q
X

) = 0 for all p = q.
Proof. Since X

is simplicial, we have the Ishida complex


(9.3.4) 0

q
X

K
0
(, q) K
1
(, q) K
q
(, q) 0
which is exact by Theorem 8.2.19. Recall that
K
j
(, q) =

( j)

qj
(

M)
Z
O
V()
and V() = O() X

is the orbit closure corresponding to . Since V() is


a toric variety, its structure sheaf has vanishing higher cohomology by Demazure
vanishing, so that H
p
(X

, K
j
(, q)) = 0 for p > 0. Since the above sequence has
q +1 terms with vanishing higher cohomology, an easy argument using the long
exact sequence in cohomology implies H
p
(X

q
X

) =0 for p >q (Exercise 9.3.1).


Now suppose p < q. Since X

is simplicial, the version of Serre duality given


in Theorem 9.2.10 implies
H
p
(X

q
X

H
np
(X

nq
X

)
The right-hand side vanishes since np > nq, and the result follows.
When X

complete but not necessarily simplicial, Danilov [76, Cor. 12.7]


proves that H
p
(X

q
X

) = 0 when q > p.
In [198], Mavlyutov discusses a vanishing theorem for

q
X

(D), where D is a
nef Cartier divisor. His result goes as follows.
424 Chapter 9. Sheaf Cohomology of Toric Varieties
Theorem 9.3.3. Let X

be a complete simplicial toric variety. If D is a nef Cartier


divisor on X

, then
H
p
(X

q
X

(D)) = 0
whenever p > q or q > p+dim P
D
.
The proof for p >q is relatively easy (Exercise 9.3.2). Note that Theorem 9.3.2
is the case D = 0 of Theorem 9.3.3. The paper [198] computes H
p
(X

q
X

(D))
explicitly for all p, q.
Q-Weil Divisors. AQ-Weil divisor or D on a normal variety X is a formal Q-linear
combination of prime divisors. Thus a positive integer multiple of D is an ordinary
Weil divisor, often called integral in this context. A Q-Weil divisor is Q-Cartier if
some positive multiple is integral and Cartier. In the literature (see [186, 1.1.4]),
Q-Cartier Q-Weil divisors are often called Q-divisors. These divisors and their
close cousins, R-divisors, are essential tools in modern algebraic geometry.
For x R, x is the greatest integer x and x is the least integer x. Then,
given a Q-Weil divisor D =

i
a
i
D
i
, we get integral Weil divisors
D =

i
a
i
D
i
(the round down of D)
D =

i
a
i
D
i
(the round up of D).
We now prove an injectivity lemma for Q-Weil divisors due to Fujino [101].
Lemma 9.3.4. Let D be a Q-Weil divisor on a toric variety X

and let > 0 be an


integer such that D is integral. Then for all p 0 there is an injection
H
p
(X

, O
X

(D)) H
p
(X

, O
X

(D)).
Proof. We adapt the proof of Lemma 9.2.6 to our situation. Write D =

,
where a

= b

for b

Z and 0

< 1. Thus D =

. With as
above, we claim that

: X

from Lemma 9.2.6 gives a split injection


(9.3.5) O
X

(D)

O
X

(D).
Assuming (9.3.5), the lemma follows from the remarks leading up to Lemma 9.2.6.
It remains to prove the existence of a split injection (9.3.5). Take an afne open
subset U

and consider
P

=m M
R
[ 'm, u

` a

for all (1).


Then
(U

, O
X

(D)) =

mP
1
M
C
m
, (U

, O
X

(D)) =

mP

M
C
m
,
9.3. Vanishing Theorems II 425
where the rst equality uses a

= b

, 0

< 1. Since P

M = (P
1
M),
the map m m induces an inclusion

mP
1
M
C
m

mP

M
C
m
.
This is a C[

M]-module homomorphism, provided that the right-hand side has


module structure
m
a =
m
a, and the usual formula for the splitting map r is
also a C[

M]-module homomorphism. From here, we get the required split


injection (9.3.5) without difculty.
Here are Q-divisor versions of Demazure and Batyrev-Borisov vanishing.
Theorem 9.3.5. Let D be a Q-Cartier Q-Weil divisor on a toric variety X

.
(a) If [[ is convex and D is nef, then
H
p
(X

, O
X

(D)) = 0 for all p > 0.


(b) If is complete and D is nef, then
H
p
(X

, O
X

(D)) = 0 for all p = dim P


D
.
Proof. Pick > 0 with D Cartier. For part (a), H
p
(X

, O
X

(D)) = 0 for p > 0


by Theorem 9.2.3. The desired vanishing follows immediately from Lemma 9.3.4.
For part (b), replacing D with D in Lemma 9.3.4 gives
H
p
(X

, O
X

(D)) = H
p
(X

, O
X

(D)) H
p
(X

, O
X

(D)),
and then we are done by Theorem 9.2.7.
Here is an example that uses part (a) of Theorem 9.3.5 to extend Demazure
vanishing beyond nef Q-Cartier divisors.
Example 9.3.6. The fan for the Hirzebruch surface H
r
has minimal generators
u
1
= (1, r), u
2
= (0, 1), u
3
= (1, 0), u
4
= (0, 1), giving divisors D
1
, . . . , D
4
. By
Example 6.1.16,
Pic(H
r
)
R
aD
3
+bD
4
[ a, b R
and the nef cone is generated by D
3
and D
4
. Since D
1
D
3
and D
2
D
4
rD
3
, it
follows that a Q-Weil divisor D = a
1
D
1
+ +a
4
D
4
is nef if and only
(9.3.6) a
1
+a
3
ra
2
and a
2
+a
4
0.
We will show that the divisors aD
3
+bD
4
, a, b 1, have vanishing higher
cohomology when r > 0. Given such a divisor, pick a rational number 0 < <
1
2
and consider the Q-Weil divisor
D = 2D
1
+

r
D
2
+(a+1)D
3
+(b+1

r
)D
4
.
This satises (9.3.6) and hence is nef. Then D =aD
3
+bD
4
has vanishing higher
cohomology by Theorem 9.3.5. Taking a =1 or b =1, we get non-nef divisors
whose higher cohomology vanishes.
426 Chapter 9. Sheaf Cohomology of Toric Varieties
Lemma 9.3.4 also leads to the following result due to Mustat a [212].
Theorem 9.3.7. Let X

be a projective toric variety and let


1
, . . . ,
r
(1) be
distinct. Then for p > 0 and any ample Cartier divisor D on X

, we have
H
p
(X

, O
X

(DD

1
D
r
)) = 0.
Proof. Let B =D

1
+ +D
r
. Since D is ample, Serre vanishing (Theorem 9.0.6)
implies that we can nd > 0 such that H
p
(X

, X

, O
X

(DB)) = 0 for p > 0.


Now let E = D
1
B. Since E = DB, the inclusion
H
p
(X

, O
X

(E)) H
p
(X

, O
X

(E))
from Lemma 9.3.4 implies
H
p
(X

, O
X

(DB)) H
p
(X

, O
X

(DB)) = 0, p > 0.
Here is a non-Q-Cartier divisor with vanishing higher cohomology.
Example 9.3.8. Consider the complete fan in R
3
shown in Figure 5. It has
z
y
x

2
Figure 5. A complete fan in R
3
divisors D
0
, . . . , D
4
corresponding to
0
, . . . ,
4
. In Exercise 9.3.3 you will show
that for a Q-Weil divisor D = a
0
D
0
+ +a
4
D
4
, we have:
D is Q-Cartier if and only a
1
+a
3
= a
2
+a
4
.
D is Q-Cartier and nef if and only a
1
+a
3
= a
2
+a
4
0.
In particular, D = D
3
+D
4
is Q-Cartier and nef, so that D
4
= DD
3
has
vanishing higher cohomology by Theorem 9.3.7. Yet D
4
is not Q-Cartier.
9.3. Vanishing Theorems II 427
Iitaka Dimension and Big Divisors. Given a nef Cartier divisor D on a complete
toric variety X

and an integer > 0, the global sections W

= H
0
(X

, O
X

(D))
give a morphism
W

: X

P(W

) as in Lemma 6.0.28. Also recall that since D


is nef, its polytope P
D
is a lattice polytope. The map
W

and the polytope P


D
are
related as follows.
Lemma 9.3.9. For 0, the image of
W

is isomorphic to the toric variety of


P
D
. In particular, dim
W

(X

) = dim P
D
for 0.
Proof. This follows from Proposition 6.2.8 (Exercise 9.3.4).
This situation is a special case of the denition of the Iitaka dimension (X, D)
(see [186, Def. 2.1.3]) of a Cartier divisor D on a complete irreducible variety X. In
this terminology, Lemma 9.3.9 implies that a nef Cartier divisor D on a complete
toric variety X

has Iitaka dimension


(X

, D) = dim P
D
.
In general, a Cartier divisor is big if it has maximal Iitaka dimension, which for a
nef Cartier divisor D on a complete toric variety X

means
D is big dim P
D
= dim X

.
It should be clear what it means for a Q-Cartier Q-Weil divisor to be big and nef.
Kawamata-Viehweg. The classic version of Kodaira vanishing can be stated as
H
p
(X, O
X
(K
X
+D)) = 0 for all p > 0
when D is an ample line bundle on a smooth projective variety X. In the 1982,
Kawamata and Viehweg independently weakened the hypotheses on D. Here is the
toric version of their result.
Theorem 9.3.10 (Toric Kawamata-Viehweg). Let X

be a complete toric variety


and let D be a Q-Cartier Q-Weil divisor on X

that is big and nef. Then


H
p
(X

, O
X

(K
X

+D)) = 0 for all p > 0.


Proof. Let n = dim X

. When D is Cartier, Serre duality implies


H
p
(X

, O
X

(K
X

+D))

H
np
(X

, O
X

(D)).
Since D is nef, the latter vanishes for np = dim P
D
by Theorem 9.2.7, and since
D is big, the condition on p becomes np = n, i.e, p = 0.
In general, write D =

where a

= b

for b

Z and 0

< 1.
Pick such that D is Cartier and 0 <

+
1
< 1 for all . Let E = D+
1
K
X

.
Since b

1
= b

1, we have E =D +K
X

. Thus the inclusion


H
p
(X

, O
X

(E)) H
p
(X

, O
X

(E))
428 Chapter 9. Sheaf Cohomology of Toric Varieties
from Lemma 9.3.4 implies
H
p
(X

, O
X

(K
X

+D)) H
p
(X

, O
X

(K
X

+D)).
Then we are done by the Cartier case already proved.
Another approach to Theorem 9.3.10 is to prove a version of Serre duality that
implies H
p
(X

, O
X

(K
X

+D))

H
np
(X

, O
X

(D)) (Exercise 9.3.5).


Grauert-Riemenschneider. Our next vanishing result features the higher direct
images of the canonical sheaf. We will need the following preliminary result.
Proposition 9.3.11. Let be a simplicial fan whose support [[ is strongly convex.
Then H
p
(X

,
X

) = 0 for all p > 0.


Proof. Fix p 1 and m M. Theorem 9.1.3 with D = K
X

implies
H
p
(X

,
X

)
m
= H
p
(X

, O
X

(D))
m


H
p1
(V
D,m
, C),
where
V
D,m
=

Conv(u

[ (1), 'm, u

` < 1)
=

Conv(u

[ (1), 'm, u

` 0).
The last equality follows since 'm, u

` Z. Now consider the set


W =u [[ `0 [ 'm, u` 0 = ([[ `0) u N
R
[ 'm, u` 0,
and observe that W is convex since [[ is strongly convex. Note also that
(9.3.7) V
D,m
W.
The proposition will follow once we prove that (9.3.7) is a deformation retract.
Given with W = , we construct r

: W V
D,m
as follows.
The assumption W = implies that
A = (1) [ 'm, u

` 0 =.
Since is simplicial, u can be uniquely written u =

(1)

0, and
when u W , dene r

(u) = (

)
1

. The sum

A

is
nonzero since u W , so that r

(u) Conv(u

[ A) V
D,m
. It is also easy
to see that r

is compatible with r

whenever is a face of . Thus we get a map


r : W V
D,m
, which is the desired retraction by Exercise 9.3.6.
Here is a toric version of Grauert-Riemenschneider vanishing.
Theorem 9.3.12 (Toric Grauert-Riemenschneider Vanishing). Let : X

be
a surjective proper toric morphism between toric varieties of the same dimension.
If X

is simplicial, then
R
p

= 0 for all p > 0.


If in addition is birational, then

.
9.3. Vanishing Theorems II 429
Proof. Our hypothesis on implies that the associated lattice map : N N

induces an isomorphism
R
: N
R
N

R
such that is the inverse image of

.
Thus, given

, we see that
1
R
() is strongly convex. This is the support of
the fan of
1
(U

), so that
H
p
(
1
(U

),
X

) = 0
for p >0 by Proposition 9.3.11. Then R
p

=0 for p >0 by Proposition 9.0.7.


The nal assertion of the theorem is an easy consequence of Theorem 8.2.15
(Exercise 9.3.7).
Most versions of Theorem 9.3.12 in the literature assume that X

is smooth.
Other Vanishing Theorems. There are many more toric vanishing theorems. Both
Fujino [101] and Mustat a [212] state general vanishing theorems that imply ver-
sions of Theorems 9.3.1, 9.3.5, 9.3.7 and 9.3.10. Further vanishing results can be
found in [102] and [226].
Exercises for 9.3.
9.3.1. A sheaf G on a variety X is acyclic if H
p
(X, G) = 0 for all p >0. Now suppose that
we have an exact sequence of sheaves
0 F G
0
G
1
G
r
0,
where G
0
, . . . , G
r
are acyclic. Prove H
p
(X, F) = 0 for all p > r by induction on r.
9.3.2. Use (9.3.4) and the previous exercise to prove Theorem 9.3.3 when p > q.
9.3.3. Let D
0
, . . . , D
4
be the divisors from Example 9.3.8 and consider a Q-Weil divisor
D = a
0
D
0
+ +a
4
D
4
.
(a) Show that D
0
2D
3
+2D
4
, D
1
D
3
, D
2
D
4
.
(b) Show that D is Q-Cartier if and only a
1
+a
3
= a
2
+a
4
.
(c) Show that D is Q-Cartier and nef if and only a
1
+a
3
= a
2
+a
4
0.
9.3.4. Complete the proof of Lemma 9.3.9.
9.3.5. Let D be a Q-Cartier Q-Weil divisor on a complete toric variety X

.
(a) Pick >0 such that Dis Cartier and consider O
X
(D)

O
X
(D) from(9.3.5).
Use this to prove that O
X
(D) is MCM. Hint: Replace

: N N with N N as
in the proof of Theorem 9.2.10.
(b) Adapt the proof of Theorem 9.2.10 to show that
H
p
(X

, O
X
(D))

H
np
(X

, O
X
(K
X
D)).
(c) Prove H
p
(X

, O
X
(D))

H
np
(X

, O
X
(K
X
+D)). Hint: D =D.
(d) Prove Theorem 9.3.10 using part (c) and Theorem 9.3.5.
9.3.6. Consider the map r

: W V
D,m
dened in the proof of Proposition 9.3.11.
(a) Prove that these maps patch to give a retraction r : W V
D,m
.
430 Chapter 9. Sheaf Cohomology of Toric Varieties
(b) Prove that when regarded as a map fromW to itself, r

is homotopic to the identity.


Then formulate and prove a similar result for r.
9.3.7. Let : X

be a proper birational toric morphism.


(a) Prove that : N N

is an isomorphism.
(b) Let
0
be the fan in N
R
consisting of the cones
1
R
() for

. Prove that is a
renement of
0
.
(c) Complete the proof of Theorem 9.3.12.
9.3.8. Given a toric variety X

, let D be a Weil divisor and E =

a Q-Weil divisor
such that 0 a

1 for all and E is integral for some integer > 0. Prove that there is
an injection
H
p
(X

, O
X
(D)) H
p
(X

, O
X
(D+(D+E))) for all p 0.
This is a strengthened version of Theorem 0.1 from [212]. Hint: As suggested by [101],
apply Lemma 9.3.4 to the Q-Weil divisor D+

+1
E.
9.4. Applications to Lattice Polytopes
In this section we use vanishing theorems to study lattice polytopes.
The Euler Characteristic of a Sheaf . Let F be a coherent sheaf on a complete
variety X. Its Euler characteristic (F) is dened to be the alternating sum
(F) =

p0
(1)
p
dim H
p
(X, F).
Our hypotheses on X and F guarantee that dim H
p
(X, F) < for all p, and
H
p
(X, F) =0 for p >dim X by [131, Thm. III.2.7]. Hence (F) is a well-dened
integer. The Euler characteristic satises
(9.4.1) (G) = (F) +(H )
whenever we have an exact sequence 0 F G H 0 of coherent sheaves.
This follows from the long exact sequence in cohomology (Exercise 9.4.1).
Given a line bundle L on X, dene
L

L
O
X

O
X
L
. .. .
times
> 0
(L

)
()
< 0
O
X
= 0,
where as usual L

=Hom
O
X
(L, O
X
). A basic result [176] states that
(F
O
X
L

), Z,
9.4. Applications to Lattice Polytopes 431
is a polynomial in , called the Hilbert polynomial. Exercises 9.4.2 and 9.4.3 sketch
a proof in the ample case. When L = O(D) for a Cartier divisor D, the Hilbert
polynomial is written
(F(D)), Z,
via the twisting convention introduced in 9.2.
Example 9.4.1. We will compute (O
P
n ()). When 0, Example 9.2.4 implies
that H
p
(P
n
, O
P
n ()) = 0 for p > 0, so that
(9.4.2) (O
P
n ()) = dim H
0
(P
n
, O
P
n ()) =[
n
Z
n
[ =

+n
n

.
The second equality follows from Proposition 4.3.3 since the polytope associated
to D
0
is
n
, where
n
is the standard n-simplex from Example 4.3.6. You will
prove the last equality in Exercise 9.4.4. When < 0, Example 9.2.8 implies
(9.4.3)
(O
P
n ()) = (1)
n
dim H
n
(P
n
, O
P
n ()) = (1)
n
[Int(
n
) Z
n
[
= (1)
n

1
n

,
where the last equality uses Exercise 9.4.4.
Now consider the polynomial
p(x) =
(x +n)(x +n1) (x +1)
n!
Q[x]
and observe that p() =

+n
n

when N.
We claim that p() = (O
P
n ()) for all Z. For 0 this follows easily
from (9.4.2). When < 0, note that
p() = (1)
n

1
n

(Exercise 9.4.4). Then p() = (O


P
n ()) for < 0 by (9.4.3).
Note also that when > 0, (9.4.2) and (9.4.3) imply that
p() =[
n
Z
n
[
p() = (1)
n
[Int(
n
) Z
n
[.
We will see below that this is a special case of Ehrhart reciprocity.
The Ehrhart Polynomial. Given a full dimensional lattice polytope P M
R
, the
functions
L(P) =[PM[
L

(P) =[Int(P) M[
count lattice points in P or in its interior. For example, if > 0 is an integer, then
(9.4.2) and (9.4.3) imply that L(
n
) =

+n
n

and L

(
n
) =

1
n

.
432 Chapter 9. Sheaf Cohomology of Toric Varieties
Theorem 9.4.2 (Ehrhart Reciprocity). Let P M
R
R
n
be a full dimensional
lattice polytope. There is a polynomial Ehr
P
(x) Q[x] such that if N, then
Ehr
P
() = L(P).
Furthermore, if N is positive, then
Ehr
P
() = (1)
n
L

(P).
Proof. Let X
P
be the toric variety of P and D
P
the associated ample divisor. We
will show that the Hilbert polynomial
Ehr
P
() = (O
X
P
(D
P
))
has the desired properties. The case when 0 is easy since D
P
is basepoint free
by Proposition 6.1.10. Thus
(O
X
P
(D
P
)) = dim H
0
(X
P
, O
X
P
(D
P
)) =[PM[ = L(P)
by Demazure vanishing (Theorem 9.2.3) and Example 4.3.7.
Now assume > 0. Then P
D
P
= P is full dimensional, hence
(O
X
P
(D
P
)) = (1)
n
dim H
n
(X
P
, O
X
P
(D
P
)) = (1)
n
[Int(P) M[
= (1)
n
L

(P)
by Batyrev-Borisov vanishing (Theorem 9.2.7).
The polynomial Ehr
P
in Theorem 9.4.2 is called the Ehrhart polynomial of P.
An elementary approach to Ehrhart polynomials and Ehrhart Reciprocity can be
found in [22].
When P is very ample, the Ehrhart polynomial has a nice interpretation. The
very ample divisor D
P
on X
P
gives a projective embedding i : X
P
P
s1
, where
s = [PM[. Its homogeneous ideal I(X
P
) C[x
1
, . . . , x
s
] gives the homogeneous
coordinate ring
C[X
P
] =C[x
1
, . . . , x
s
]/I(X
P
).
This graded ring has a Hilbert function
dimC[X
P
]

, 0,
which is a polynomial for 0 (see [70, Ch. 6, 4]). This is the Hilbert polynomial
of X
P
P
s1
,
Proposition 9.4.3. If P is very ample, then the Ehrhart polynomial Ehr
P
equals the
Hilbert polynomial of the toric variety X
P
under the projective embedding given by
the very ample divisor D
P
.
9.4. Applications to Lattice Polytopes 433
Proof. Consider the exact sequence
0 I
X
P
O
P
s1 O
X
P
0.
Since O
X
P
(D
P
) is the restriction of O
P
s1 (1) to X
P
, tensoring with O
P
s1 () gives
an exact sequence
0 I
X
P
() O
P
s1 () O
X
P
(D
P
) 0.
In Exercise 9.4.5 you will show that the resulting long exact sequence is
0 I(X
P
)

C[x
1
, . . . , x
s
]

H
0
(X
P
, O
X
P
(D
P
)) H
1
(P
s1
, I
X
P
()) .
The H
1
term vanishes for 0 by Serre vanishing (Theorem 9.0.6). Hence for
large, we get an isomorphism
C[X
P
]

H
0
(X
P
, O
X
P
(D
P
)).
This implies that the Hilbert polynomial of X
P
is the Ehrhart polynomial of P.
We can describe the degree and leading coefcient of the Ehrhart polynomial.
For the leading coefcient, we use the normalized volume in M
R
. Let e
1
, . . . , e
n
be a basis of M and consider the simplex
n
= Conv(0, e
1
, . . . , e
n
) M
R
. Then
the normalized volume is the usual n-dimensional Lebesgue measure, scaled so
that
n
has volume equal to 1. Thus the unit cube

n
i=1

i
e
i
[ 0
i
1 has
normalized volume n!.
Given a full dimensional lattice polytope P M
R
, its normalized volume is
denoted Vol(P) and is computed by the limit
(9.4.4)
Vol(P)
n!
= lim

L(P)

n
.
This is proved in many places, including [22, Lem. 3.19]. Since L(P) is given
by the polynomial Ehr
P
(), it follows easily that Ehr
P
has degree n and its leading
coefcient Vol(P)/n! (Exercise 9.4.7).
Here is a classic application of these ideas.
Example 9.4.4. Let P R
2
be a lattice polygon with Ehrhart polynomial Ehr
P
.
The leading coefcient of Ehr
P
is
1
2
Vol(P), which is the usual Euclidean area
Area(P) since the normalized volume of the unit square is 2. The constant term is
also easy to compute, since Ehr
P
(0) = L(0 P) = 1 by Theorem 9.4.2. Thus
Ehr
P
(x) = Area(P)x
2
+
1
2
Bx +1,
where B is yet to be determined. The reason for the
1
2
will soon become clear.
By Ehrhart reciprocity, we have
(9.4.5)
Area(P) +
1
2
B+1 = Ehr
P
(1) = L(P)
Area(P)
1
2
B+1 = Ehr
P
(1) = (1)
2
L

(P) = L

(P).
434 Chapter 9. Sheaf Cohomology of Toric Varieties
Solving for B gives B = L(P) L

(P) =[PM[, where P is the boundary of P.


Thus the Ehrhart polynomial of P is
Ehr
P
(x) = Area(P)x
2
+
1
2
[PM[ x +1.
Furthermore, solving the bottom equation of (9.4.5) for the area gives
Area(P) = L

(P) +
1
2
[PM[ 1.
This equation, called Picks formula, shows how to compute the area of a lattice
polygon in terms of its lattice points.
Example 9.4.5. Let P = Conv(0, 0, 0), (1, 0, 0), (0, 1, 0), (1, 1, 3) be the lattice
simplex shown in Figure 3 of Example 2.2.4. Since the normalized volume is 3 and
Ehr
P
(0) = L(0 P) = 1, the Ehrhart polynomial is Ehr
P
(x) =
1
2
x
3
+ax
2
+bx +1.
We noted in Example 2.2.8 that the only lattice points of P are its vertices. By
Theorem 9.4.2, we obtain
Ehr
P
(1) =
1
2
+a+b+1 = L(P) = 4 (4 lattice points)
Ehr
P
(1) =
1
2
+ab+1 = (1)
3
L

(P) = 0 (no interior lattice points)


Hence a = 1, b =
3
2
, so that Ehr
P
(x) =
1
2
x
3
+x
2
+
3
2
x +1. In Example B.3.3 we
will explain how to compute Ehr
P
(x) using Normaliz [57].
Examples 9.4.4 and 9.4.5 used Ehr
P
(0) = L(0 P) = 1. The latter equality is
obvious since 0 P = 0. However, the equality Ehr
P
(0) = L(0 P) (proved in
Theorem 9.4.2) is more subtle since it can fail when we replace the polytope P
with a polytopal complex (a collection of polytopes where the intersection of any
two is a face of each). Here is a simple example.
Example 9.4.6. Consider the boundary
2
of the 2-simplex
2
R
2
. Thus
2
consists of three line segments. One easily computes that
L(
2
) =[(
2
) Z
2
[ = 3
when >0 is an integer. This gives the Ehrhart polynomial Ehr

2
(x) = 3x with
the property that Ehr

2
() = L(
2
) for integers > 0. However,
Ehr

2
(0) = 3 0 = 0 = L(0
2
) = L(0) = 1.
Further examples (and an explanation) will be given in Execise 9.4.8.
The p-Ehrhart Polynomials. Following Materov [193], we use the sheaves

p
X
P
to generalize the Ehrhart polynomial when the lattice polytope P is simple. Recall
that a full dimensional lattice polytope P M
R
R
n
is simple if every vertex is the
intersection of exactly n facets. Hence each vertex has n facet normals that gener-
ate the corresponding cone in the normal fan
P
. It follows that
P
is simplicial
whenever P is simple. Hence the toric variety X
P
is simplicial.
9.4. Applications to Lattice Polytopes 435
We proved earlier that the Ehrhart polynomial Ehr
P
(x) of P satises
Ehr
P
() = (O
X
P
(D
P
)), Z,
where D
P
is the ample divisor coming from P. More generally, given an integer
0 p n, the Euler characteristic (

p
X
P
(D
P
)) is a polynomial function of .
Then the p-Ehrhart polynomial of P, denoted Ehr
p
P
(x), is the unique polynomial in
Q[x] that satises
Ehr
p
P
() = (

p
X
P
(D
P
)), Z.
Note that Ehr
0
P
(x) is the ordinary Ehrhart polynomial Ehr
P
(x). To state the
properties of Ehr
p
P
(x), we will use the following notation:
P(i) =Q [ Q is an i-dimensional face of P
f
i
=[P(i)[ = #i-dimensional faces of P
h
p
=
n

i=p
(1)
ip

i
p

f
i
.
The f
i
are the face numbers of P. Furthermore, if Q is a face of P, we dene
L(Q) =[QM[
L

(Q) =[Relint(Q) M[.


These invariants are related to the p-Ehrhart polynomials by the following result of
Materov [193].
Theorem 9.4.7. Let X
P
be the toric variety of a full dimensional simple lattice
polytope P M
R
R
n
.
(a) If > 0 is an integer, then
Ehr
p
P
() =
n

i=p

i
p


QP(i)
L

(Q).
(b) If 0 is an integer, then
Ehr
p
P
() =
p

i=0
(1)
i

ni
np


QP(ni)
L(Q).
(c) If 0 p n, then
Ehr
p
P
(x) = (1)
n
Ehr
np
P
(x).
(d) If 0 p n, then
Ehr
p
P
(0) = (1)
p
h
p
.
436 Chapter 9. Sheaf Cohomology of Toric Varieties
We will defer the proof for now. Theorem 9.4.7 has some nice consequences.
First, setting x = 0 in part (c) and using part (d), we see that
(1)
p
h
p
= (1)
n
(1)
np
h
np
.
This proves that
(9.4.6) h
p
= h
np
for 0 p n. These are called the Dehn-Sommerville equations.
Also, if we write Ehr
p
P
() using part (b) of the theorem and Ehr
np
P
() using
part (a), then part (c) gives the following formulas for > 0:
Ehr
p
P
() =
p

i=0
(1)
i

ni
np


QP(ni)
L(Q),
Ehr
p
P
() = (1)
n
Ehr
np
P
() = (1)
n
n

i=np

i
np


QP(i)
L

(Q).
For p = 0 and > 0, these equations reduce to Ehrhart reciprocity:
Ehr
0
P
() = L(P)
Ehr
0
P
() = (1)
n
L

(P).
Thus Theorem 9.4.7 simultaneously generalizes the Dehn-Sommerville equations
and Ehrhart reciprocity. The proof of the theorem will use the following lemma.
Lemma 9.4.8. Let P M
R
R
n
be a full dimensional lattice polytope with toric
variety X
P
and ample divisor D
P
. Given an integer > 0 and m (P) M, let
V
P
(m) be the subspace of M
C
such that V
P
(m) +m is the smallest afne subspace
of M
C
containing the minimal face of P containing m. Then
H
0
(X
P
,

p
X
P
(D
P
)) =

m(P)M

p
V
P
(m)
m
.
Proof. We adapt the strategy used in the proof of Proposition 8.2.18. To simplify
notation, set D = D
P
. Using

M M and tensoring (8.2.6) with O


X
P
(D), we
obtain the exact sequence
0

p
X
P
(D)

p
M
Z
O
X
P
(D)

p1
M
Z
O
D
(D).
Take global sections over X
P
and consider the graded piece for m M. Since
H
0
(X
P
, O
X
P
(D)) =

m(P)M
C
m
, we get the exact sequence
0 H
0
(X
P
,

p
X
P
(D))
m

p
M
C

p1
M
Z
H
0
(X
P
, O
D
(D))
m
when m(P)M. To compute H
0
(X
P
, O
D
(D))
m
, recall from Proposition 4.0.28
that we have an exact sequence
0 O
X
P
(D

) O
X
P
O
D
0.
9.4. Applications to Lattice Polytopes 437
Tensor this with O
X
P
(D) and take global sections to obtain
0 H
0
(X
P
, O
X
P
(DD

)) H
0
(X
P
, O
X
P
(D)) H
0
(X
P
, O
D
(D)) 0,
where the exactness on right follows from H
1
(X
P
, O
X
P
(DD

)) = 0 courtesy of
Theorem 9.3.7. Comparing the polytopes of DD

and D =

shows that
H
0
(X
P
, O
D
(D))
m
=

C 'm, u

` =a

0 otherwise.
Let F

be the facet of P corresponding to . For m P, 'm, u

` = a

if and
only if m F

. Hence we get an exact sequence


0 H
0
(X
P
,

p
X
P
(D))
m

p
M
C
p

mF

p1
M
C
,
where
p
is a sum of contraction maps i
u
dened in (8.2.7).
Thus

p
M
C
is in the kernel of
p
if and only if i
u
() = 0 for all with
m F

. We know from Exercise 8.2.8 that i


u
() = 0 if and only if

p
(

)
C
.
It follows that the kernel of
p
in (8.2.9) is the intersection
(9.4.7)

mF

p
(

)
C
=

mF
(

)
C

.
Since F =

mF
F

is the minimal face of P containing m, one sees easily that


Span
R
(m
0
m M [ m
0
F) =

mF

.
Thus V
P
(m) =

mF
(

)
C
. This and (9.4.7) imply ker(
p
) =

p
V
P
(m).
We are now ready to prove our main result.
Proof of Theorem 9.4.7. We begin with part (a). Lemma 9.4.8 implies that
dim H
0
(X
P
,

p
X
P
(D
P
)) =

m(P)M

dim V
P
(m)
p

when > 0. Given a face Q of P and m (P) M, note that


m Relint(Q) M Q is the minimal face of P containing m.
When this happens, we have dim Q = dim V
P
(m). Since the i-dimensional faces
of P are Q for Q P(i), we obtain
dim H
0
(X
P
,

p
X
P
(D
P
)) =

i0

i
p


QP(i)
L

(Q) =
n

i=p

i
p


QP(i)
L

(Q).
We also have H
q
(X
P
,

p
X
P
(D
P
)) = 0 for q > 0 by Theorem 9.3.1. Hence
Ehr
p
P
() = (

p
X
P
(D
P
)) = dim H
0
(X
P
,

p
X
P
(D
P
)) =
n

i=p

i
p


QP(i)
L

(Q),
438 Chapter 9. Sheaf Cohomology of Toric Varieties
which proves part (a).
Now consider Ehr
p
P
(0). If > 0 and Q P(i), then L

(Q) = (1)
i
Ehr
Q
()
by Ehrhart reciprocity. Since the previous display holds for all >0, we obtain the
polynomial identity
(9.4.8) Ehr
p
P
(x) =
n

i=p

i
p


QP(i)
(1)
i
Ehr
Q
(x).
Setting x = 0 gives
Ehr
p
P
(0) =
n

i=p
(1)
i

i
p


QP(i)
Ehr
Q
(0) = (1)
p
h
p
by the denition of h
p
, and part (d) follows.
For part (c), we use Serre duality as given in Theorem 9.2.10, which implies
(9.4.9) H
q
(X

p
X
P
(D
P
))

H
nq
(X

np
X
P
(D
P
))
since D
P
is Cartier for any Z. This easily implies
(

p
X
P
(D
P
)) = (1)
n
(

np
X
P
(D
P
))
(Exercise 9.4.9). Thus Ehr
p
P
() = (1)
n
Ehr
np
P
() for Z, proving part (d).
Finally, for part (b), take 0 and consider
Ehr
p
P
() = (1)
n
Ehr
np
P
() = (1)
n
n

i=np

i
np


QP(i)
(1)
i
Ehr
Q
(),
where the rst equality uses part (d) and the second uses (9.4.8) with p replaced by
np. Since 0 implies Ehr
Q
(Q) = L(Q), we obtain
Ehr
p
P
() =
n

i=np
(1)
ni

i
np


QP(i)
L(Q) =
p

j=0
(1)
j

n j
np


QP(i)
L(Q),
where the last equality follows by setting j = ni. This completes the proof.
Identities equivalent to Theorem 9.4.7 were discovered by McMullen in 1977
in a very different context. See [22, Ch. 5] for details and references, including a
non-toric version of Theorem 9.4.7 in [22, Ex. 5.85.10].
Examples. We rst give a classic example of the Dehn-Sommerville equations.
Example 9.4.9. Let P be a simple 3-dimensional lattice polytope in R
3
. The Dehn-
Sommerville equations can be written
h
3
= h
0
, so f
3
= f
0
f
1
+ f
2
f
3
h
2
= h
1
, so f
2
3f
3
= f
1
2f
2
+3f
3
.
9.4. Applications to Lattice Polytopes 439
Since f
3
= 1 (P is the only 3-dimensional face of itself), we obtain
f
0
f
1
+ f
2
= 2
f
1
= 3f
2
6.
The rst equation Eulers celebrated formula V E +F = 2, which holds for any
3-dimensional polytope. The second seems more mysterious, but when combined
with the rst reduces to 2f
1
= 3f
0
. This holds because is P is simpleevery vertex
meets three edges and every edge meets two vertices.
There is a version of the Dehn-Sommerville equations for the dual polytope
P

N
R
, which is simplicial since P is simple (Exercise 9.4.10). More on the
Dehn-Sommerville equations and toric varieties can be found in [105, Sec. 5.6].
We also recommend [22, Ch. 5] and [281, Sec. 8.3].
We next give an application of Theorem 9.4.7.
Example 9.4.10. The standard n-simplex
n
has P
n
as its associated toric variety.
Given > 0, note that

Qn(i)
L

(
n
) =

n+1
ni

1
i

because
An n-simplex has

n+1
i+1

n+1
ni

faces of dimension i.
Each Q
n
(i) is lattice isomorphic to the standard i-simplex and hence has

1
i

interior lattice points by (9.4.3).


Then part (a) of Theorem 9.4.7 gives the formula
Ehr
p
n
() =
n

i=p

i
p

n+1
ni

1
i

.
Since

i
p

1
i

1
p

p1
ip

, this becomes
Ehr
p
n
() =

1
p

i0

n+1
ni

p1
i p

1
p

n+ p

,
where the last equality uses the Vandermonde identity discussed in Exercise 9.4.11.
Hence
dim H
0
(P
n
,
p
P
n
()) =

1
p

n+ p

.
This formula was rst proved by Bott [40] in 1957.
440 Chapter 9. Sheaf Cohomology of Toric Varieties
Cohomology of p-Forms. Given a simple polytope P as above, the sheaf

p
X
P
has
Euler characteristic
(9.4.10) (

p
X
P
) = Ehr
p
P
(0) = (1)
p
h
p
by Theorem 9.4.7. The factor of (1)
p
is explained as follows.
Theorem 9.4.11. Let X
P
be the toric variety of a full dimensional simple lattice
polytope in P M
R
R
n
. Then
dim H
q
(X
P
,

p
X
P
) =

h
p
q = p
0 q = p.
Proof. The toric variety X
P
is simplicial, so that Theorem 9.3.2 applies to X
P
. This
shows that H
q
(X
P
,

p
X
P
) = 0 for q = p. It follows that
(

p
X
P
) = (1)
p
dim H
p
(X
P
,

p
X
P
).
The theorem follows by comparing this with (9.4.10).
A different proof of dim H
p
(X
P
,

p
X
P
) =h
p
will be sketched in Exercise 9.4.12.
Theorem 9.4.11 has a nice application to the singular cohomology of X
P
because
of the the Hodge decomposition
(9.4.11) H
k
(X
P
, C)

p+q=k
H
q
(X
P
,

p
X
P
).
When X
P
is smooth, this is a classical factsee, for example, [125, p. 116]. In the
simplicial case, see [261] for a proof. Combined with Theorem 9.4.11, we obtain
(9.4.12) dim H
k
(X
P
, C) =

h
p
k = 2p
0 k odd.
We will give a topological proof of this formula in 12.4.
Exercises for 9.4.
9.4.1. Prove (9.4.1).
9.4.2. Here are some cases where it is easy to show that Euler characteristics give Hilbert
polynomials.
(a) Let F be a nitely generated graded module over the polynomial ring S =C[x
0
, . . . , x
n
].
By [70, Ch. 6, Thm. (3.8)], there is an exact sequence
0 F
r
F
1
F
0
F 0,
where each F
i
is a nite direct sum of modules for the form S(a). Now let F be a
coherent sheaf on P
n
. Prove that there is an exact sequence of sheaves
0 F
r
F
1
F
0
F 0,
where each F
i
is a nite direct sum of sheaves of the form O
P
n (a). Hint: Apply
Example 6.0.10 with X

=P
n
.
9.4. Applications to Lattice Polytopes 441
(b) Use part (a) together with (9.4.1) and Example 9.4.1 to show that (F()) is a poly-
nomial in Z, where F() =F
O
P
n
O
P
n ().
(c) Let F be a coherent sheaf on a complete variety X and let D be a very ample divisor
on X. Use part (b) and Exercise 9.0.6 to show that (F(D)) is a polynomial in Z.
9.4.3. Let F be a coherent sheaf on a projective variety X, and let D be an ample divisor
on X. Thus there is k
0
> 0 such that kD is very ample for all k k
0
.
(a) Let a, k be integers with k k
0
. Use Exercise 9.4.2 to show that there is a polynomial
p
a,k
(x) Q[x] such that p
a,k
() = (F((a +k)D)) for all Z.
(b) Show that the polynomials p
a,k
(x/k) and p
a,m
(x/m) are equal when k, m k
0
. Con-
clude that there is a polynomial p
a
(x) such that p
a
() =(F((a+)D)) for all k
0
.
(c) Show that the polynomials p
a
(x a) and p
b
(x b) are equal for any a, b Z and
conclude that there is a polynomial p(x) Q[x] such that p() = (F(D)) for all
Z.
9.4.4. In this exercise will always denote an integer.
(a) Prove that [(
n
) Z
n
[ =

+n
n

for nonnegative. Hint: Lattice points in


n
give
monomials of degree in x
0
, . . . , x
n
by Exercise 4.3.6. Here is a combinatorial proof.
Begin with a list of 1s of length +n. In n of the positions, convert the 1 to a 0. This
divides the remaining 1s into n +1 groups. The number of elements in each group
gives the exponents of x
0
, . . . , x
n
. See also [69, Ex. 13 of Ch. 9, 3].
(b) Prove that [Int(
n
) Z
n
[ =

1
n

for positive. Hint: Show that shifting the interior


lattice points by (1, . . . , 1) gives the lattice points in ( n 1)
n
.
(c) Prove that p(x) = (x +n)(x +n 1) (x +1)/n! satises p() = (1)
n

1
n

for
negative.
9.4.5. A projective variety X P
n
gives 0 I
X
O
P
n O
X
0. After tensoring with
O
P
n (), we have an exact sequence 0 I
X
() O
P
n () O
X
() 0 whose long exact
sequence in cohomology begins
0 H
0
(P
n
, I
X
()) H
0
(P
n
, O
P
n ()) H
0
(P
n
, O
X
()) H
1
(P
n
, I
X
()) .
We know from Example 4.3.6 that H
0
(P
n
, O
P
n ()) = S

, where S = C[x
0
, . . . , x
n
] with the
standard grading. Also recall that as a sheaf on P
n
, O
X
() stands for i

O
X
(), where
i : X P
n
is the inclusion map. It follows that H
0
(P
n
, O
X
()) = H
0
(X, O
X
()).
(a) Show that H
0
(P
n
, I
X
()) = I(X)

, where I(X) S is the ideal of X.


(b) Show that the Hilbert polynomial of the coordinate ring C[X] = S/I(X) is the Euler
characteristic (O
X
()).
9.4.6. Recall that X P
n
is projectively normal if and only if its afne cone is normal.
By [131, Ex. II.5.14], this is equivalent to saying that X is normal and H
0
(P
n
, O
P
n ())
H
0
(X, O
X
()) is surjective for all 0. Combine this with the previous exercise to show
that X is projectively normal if and only if X is normal and H
1
(P
n
, I
X
()) =0 for all 0.
See Example B.1.2 for a computational example.
9.4.7. Let p(x) be a polynomial such that lim

p()/
n
exists and is nonzero.
(a) Prove that n = deg(p(x)) and that the above limit is the leading coefcient of p(x).
(b) Use part (a) to prove the assertion made following (9.4.4) concerning the degree and
leading coefcient of the Ehrhart polynomial.
442 Chapter 9. Sheaf Cohomology of Toric Varieties
9.4.8. Here we compute the Ehrhart polynomials of some simple polytopal complexes.
(a) As in Example 9.4.6, let
2
be the boundary of the standard simplex
2
R
2
. Prove
that [(
2
) Z
2
[ = 3, so that Ehr
2
(x) = 3x, with constant term 0.
(b) Consider the buttery B = Conv(0, e
1
e
2
) Conv(0, e
1
e
2
) with boundary B.
Prove that [( B) Z
2
[ = 8 1, so that Ehr
B
(x) = 8x 1, with constant term 1.
(c) The Euler characteristic of a well-behaved topological space Z is dened to be e(Z) =

p
(1)
p
rankH
p
(Z, Z). Show that e(
2
) = 0 and e(B) =1.
The answer to part (c) is not a coincidencegiven any polytopal complex P whose ver-
tices are lattice points, the constant term of its Ehrhart polynomial Ehr
P
is e(P) (see
[189]). Since a polytope P is contractible, this gives a different way of seeing that the
constant term of Ehr
P
is 1.
9.4.9. Use (9.4.9) to prove that (

p
XP
(D
P
)) = (1)
n
(

np
XP
(D
P
)).
9.4.10. Let P M
R
R
n
be a full dimensional simplicial lattice polytope and dene
h
simp
p
=
n

i=p
(1)
ip

i
p

f
ni1
.
Rescaling and translating P does not affect the face numbers f
i
. So we may that assume the
origin is an interior point of P. Let P

N
R
be the dual polytope of P. Use Exercise 2.3.4
and Proposition 2.3.8 to prove the following:
(a) P

is simple.
(b) The face numbers f
P
i
of P and f
P

i
of P

are related by f
P
ni1
= f
P

i
.
(c) h
simp
p
= h
simp
np
.
9.4.11. The goal of this exercise is to prove the identity used in Example 9.4.10.
(a) Prove the Vandermonde identity, which states that

i+j=k

a
i

b
j

a+b
k

for a, b N.
Hint: (1 +x)
a
(1 +x)
b
= (1 +x)
a+b
.
(b) Use part (a) to showthat

i0

n+1
ni

p1
ip

n+p

, as claimed in Example 9.4.10.


9.4.12. For a full dimensional simple polytope P M
R
R
n
, Theorem 8.2.19 gives
0

p
XP
K
0
(
P
, p) K
1
(
P
, p) K
p
(
P
, p) 0,
where K
i
(
P
, p) =

P(i)

pi
(

M)
Z
O
V()
and V() X
P
is the closure of the
orbit corresponding to
P
. You will use this to compute dim H
p
(X
P
,

p
XP
).
(a) Show that (

p
XP
) =

p
j=0
(1)
j
(K
j
(
P
, p)).
(b) Use Theorem 9.2.3 to show that (K
j
(
P
, p)) = dim H
0
(X
P
, K
j
(
P
, p)).
(c) Use Proposition 2.3.8 to showthat dim H
0
(X
P
, K
j
(
P
, p)) =

nj
np

f
nj
. Hint:

M
has rank n j for
P
( j).
(d) Use Theorem 9.3.2 to conclude that dim H
p
(X
P
,

p
XP
) = h
np
. Hint: Set j = n i.
9.4.13. When a polytope P has rational but not integral coordinates, the counting function
L(P) is almost a polynomial. Here you will study P = Conv(0, e
1
,
1
2
e
2
) R
2
.
9.5. Local Cohomology and the Total Coordinate Ring 443
(a) Given N, prove that
L(P) =

1
4

2
+ +1 even
1
4

2
+ +
3
4
odd.
This is an example of a quasipolynomial. See [22, Sec. 3.7] for a discussion of the
Ehrhart quasipolynomial of a rational polytope.
(b) The weighted projective plane P(1, 1, 2) is given by the fan in R
2
with minimal gen-
erators u
0
= e
1
2e
2
, u
1
= e
1
, u
2
= e
2
. Show that X
2P
= P(1, 1, 2) with D
2P
= 2D
0
,
where D
0
is the divisor corresponding to u
0
. Also show that
(O
P(1,1,2)
(D
0
)) = L(P).
The Euler characteristic is not a polynomial in . The reason is that D
0
is not Cartier.
9.4.14. Let P M
R
R
n
be a full dimensional lattice poltyope, not necessarily simple.
Let Ehr
p
P
(x) be the unique polynomial satisfying Ehr
p
P
() = (

p
XP
(D
P
)) for all Z.
(a) Prove that if > 0 in an integer, then
Ehr
p
P
() =
n

i=p

i
p


QP(i)
L

(Q).
Hint: Look at the hypotheses of Theorem 9.3.1 and Lemma 9.4.8.
(b) Prove that Ehr
0
P
() = (1)
n
Ehr
n
P
(). Hint: Serre duality.
(c) Prove that Ehr
p
P
(0) = (1)
p
h
p
. Hint: Follow the proof of Theorem 9.4.7.
This shows that some parts of Theorem 9.4.7 hold for arbitrary lattice polytopes. In the
next exercise you will see that other parts can fail.
9.4.15. This exercise will show how things can go wrong for a non-simple polytope. Let
P = Conv(e
1
, e
2
, e
3
) R
3
. Note that P is not simple, so that X
P
is not simplicial.
(a) Show that h
1
= h
2
.
(b) Conclude that for this polytope, (9.4.9) cannot hold for all p, q, . So the version of
Serre duality stated in part (b) of Theorem 9.2.10 can fail for non-simplicial toric
varieties. Hint: The Dehn-Sommerville equations follow from Theorem 9.4.7.
9.4.16. Suppose that P R
3
is a lattice simplex whose only lattice points are its vertices,
and let k be the normalized volume of P. Prove that 2P has k 1 interior lattice points.
Hint: Adapt Exercise 9.4.5.
9.5. Local Cohomology and the Total Coordinate Ring
In this section, we use local cohomology and Ext to compute the cohomology of a
coherent sheaf on a toric variety X

. Our treatment is based on [91].


We rst review the basics of local cohomology.
444 Chapter 9. Sheaf Cohomology of Toric Varieties
Local Cohomology. Let R be a nitely generated C-algebra, I R an ideal, and
M an R-module. First dene

I
(M) =a M[ I
k
a = 0 for some k N.
This is the I-torsion submodule of M. One checks easily that M
I
(M) is left
exact. Hence, just as we did for global sections of sheaves in 9.0, we get the
derived functors MH
p
I
(M) such that
H
0
I
(M) =
I
(M).
A short exact sequence 0 MN P 0 gives a long exact sequence
0 H
0
I
(M) H
0
I
(N) H
0
I
(P) H
1
I
(M) H
1
I
(N) H
1
I
(P) .
The Local

Cech Complex. As with sheaf cohomology, there is a

Cech complex
for local cohomology. The sheaf case used an afne open cover; here we use
generators of the ideal. More precisely, if I = ' f
1
, . . . , f

`, let [] = 1, . . . , be
the index set and as in 9.0, let []
p
denote the set of all (p+1)-tuples (i
0
, . . . , i
p
)
of elements of I satisfying i
0
< < i
p
. Also set f = ( f
1
, . . . , f

)
Given an R-module M, dene

C
p
(f, M) =

(i
0
,...,i
p1
)[]
p1
M
f
i
0
f
i
p1
,
where M
f
i
0
f
i
p1
is the localization of M at f
i
0
f
i
p1
. An element of

C
p
(f, M) is
a function that assigns an element of M
f
i
0
f
i
p1
to each (i
0
, . . . , i
p1
) []
p1
.
Then dene a differential
d
p
:

C
p
(f, M)

C
p+1
(f, M)
by
d
p
()(i
0
, . . . , i
p
) =
p

k=0
(1)
k
(i
0
, . . . ,

i
k
, . . . , i
p
),
where we regard (i
0
, . . . ,

i
k
, . . . , i
p
) as an element of M
f
i
0
f
ip
via the map
M
f
i
0

c
f
i
k
f
ip
M
f
i
0
f
ip
given by localization at f
i
k
. Similar to Exercise 9.0.2, we have d
p
d
p1
= 0.
Denition 9.5.1. Given an R-module M and generators f = ( f
1
, . . . , f

) of I, the
local

Cech complex is

(f, M) : 0

C
0
(f, M)
d
0


C
1
(f, M)
d
1


C
2
(f, M)
d
2
.
Just as the

Cech complex computes sheaf cohomology, the local

Cech complex
computes local cohomology. See [158, Thm. 7.13] for a proof of the following.
9.5. Local Cohomology and the Total Coordinate Ring 445
Theorem 9.5.2. Given an R-module M and generators f = ( f
1
, . . . , f

) of I, there
are natural isomorphisms
H
p
I
(M) H
p
(

(f, M))
for all p 0.
Example 9.5.3. For I ='x, y` S =C[x, y], the local

Cech complex is
0 S S
x
S
y
S
xy
0.
Theorem 9.5.2 implies that H
2
I
(S) is the cokernel of S
x
S
y
S
xy
. Consider
x
1
y
1
C[x
1
, y
1
] = Span
C
(x
a
y
b
[ a, b > 0) with S-module structure given by
x
i
y
j
x
a
y
b
=

x
(ai)
y
(bj)
i < a, j < b
0 otherwise.
Then the map S
xy
x
1
y
1
C[x
1
, y
1
] dened by f (x, y)/(xy)
k
f (x, y) x
k
y
k
gives an exact sequence
(9.5.1) S
x
S
y
S
xy
x
1
y
1
C[x
1
, y
1
] 0
(Exercise 9.5.1). Thus H
2
I
(S) x
1
y
1
C[x
1
, y
1
].
It follows that if x and y have degree 1, then H
2
I
(S) is graded with
H
2
I
(S) = H
2
I
(S)
4
H
2
I
(S)
3
H
2
I
(S)
2
0
and dim H
2
I
(S)

= 1 for > 0 (Exercise 9.5.1).


The number 1 just computed is also the dimension of H
1
(P
1
, O
P
1 ()) (see
Example 9.4.1). This is no accident, since we will prove below that
H
2
I
(S)
a
H
1
(P
1
, O
P
1 (a))
for all a Z.
Relation with Ext. For us, an especially useful aspect of local cohomology is its
relation to Ext. In 9.0 we introduced Ext in the context of sheaves. There is
also a module version, where for a xed R-module N, the derived functors of
M Hom
R
(N, M) are denoted Ext
p
R
(N, M). They have the expected properties,
and more importantly, are easy to compute by computer algebra systems such as
Macaulay2.
To see the relation between Ext and local cohomology, we begin with the ob-
servation that an R-module homomorphism : R/I
k
M is determined by (1),
and choosing any a Mgives a homomorphism, provided that I
k
a = 0. It follows
that Hom
R
(R/I
k
, M) consists of those elements of M annihilated by I
k
. Comparing
this to the denition of
I
(M), we obtain

I
(M) = lim

k
Hom
R
(R/I
k
, M).
446 Chapter 9. Sheaf Cohomology of Toric Varieties
Since direct limit is an exact functor (Exercise 9.5.2), it follows without difculty
that the derived functors are related the same way (see [158, Thm. 7.8] for a proof).
Theorem 9.5.4. H
p
I
(M) = lim

k
Ext
p
R
(R/I
k
, M).
We will need a variant of this result. For an ideal I =' f
1
, . . . , f

`, let
I
[k]
=' f
k
1
, . . . , f
k

`.
Theorem 9.5.5. H
p
I
(M) = lim

k
Ext
p
R
(R/I
[k]
, M).
Proof. This follows from [158, Rem. 7.9] since I
[k]
I
k
and I
k
I
[k]
.
Here is a simple example.
Example 9.5.6. Let I = 'x, y` S = C[x, y]. We know by Example 9.5.3 that
H
2
I
(S)
5
has dimension 4. To compute this using Ext, let A = S/I
[k]
= S/'x
k
, y
k
`.
Since A and S are graded S-modules, the Ext group Ext
2
S
(A, S) is also graded. As
described in Example B.4.1, we can compute the graded piece Ext
2
S
(A, S)
5
for
various values of k using Macaulay2. When k = 3, we nd that
dim Ext
2
S
(S/I
[3]
, S)
5
= 2.
Since dim H
2
I
(S)
5
= 4, we see that k = 3 is not big enough. If we switch to k 4,
then the answer stabilizes at the number 4.
In this example, Ext
2
S
(S/I
[k]
, S)
5
= H
2
I
(S)
5
for k sufciently large. We will
prove below that for any degree a Z, we have
Ext
2
S
(S/I
[k]
, S)
a
= H
2
I
(S)
a
provided that k is sufciently large. The problem is that Ext
2
S
(S/I
[k]
, S) is nitely
generated over S but H
2
I
(S) is not (Exercise 9.5.4). So we cannot use the same k
for all degrees a. Explicit bounds on k in terms of a are needed in order to turn the
method of Example 9.5.6 into an algorithm.
This concludes our overview of local cohomology; for more, see [89, App. 4]
or [158]. Our next task to is apply local cohomology to toric varieties.
The Toric Case. The total coordinate ring S = C[x

[ (1)] of a toric variety


X

is graded by the class group Cl(X

), where a monomial

x
a

has degree
[

] Cl(X

). We also have the irrelevant ideal


B() ='x

[
max
`, x

=

,(1)
x

,
introduced in Chapter 5. We assume that X

has no torus factors.


We proved in Chapter 5 that a nitely generated graded S-module M and a
divisor class Cl(X

) give the following:


9.5. Local Cohomology and the Total Coordinate Ring 447
The coherent sheaf F =

M on X

. Every coherent sheaf on X

arises in this
way (Proposition 5.3.9).
The shifted module M(), where M()

= M
+
for Cl(X

).
The sheaf associated to S() is denoted O
X

(), and Proposition 5.3.7 tells us that


O
X

() O
X

(D) when D is a Weil divisor with divisor class . More generally,


if F =

M, then F() will denote the sheaf associated to M(). When is the
class of a Cartier divisor, one can prove that
(9.5.2) F() F
X

O
X

()
(Exercise 9.5.3). Our rst main result is that the local cohomology for the irrelevant
ideal B() computes the sheaf cohomology of all twists of a coherent sheaf on X

.
Theorem 9.5.7. Let M be a nitely generated graded S-module with associated
coherent sheaf F =

M on X

. If p 2, then
H
p
B()
(M)

Cl(X

)
H
p1
(X

, F()).
Furthermore, we have an exact sequence
0 H
0
B()
(M) M

Cl(X

)
H
0
(X

, F()) H
1
B()
(M) 0.
Proof. Let
max
=
1
, . . . ,

, so that B() is generated by the monomials f


i
=
x

i
, i [] = 1, . . . , . Then the terms of the local

Cech complex

C

(f, M) are
direct sums of localizations M at products of various f
i
s. These localizations are
Cl(X

)-graded since M is graded and the f


i
are monomials. The differentials also
preserve the grading, which by Theorem 9.5.2 implies that H
p
B()
(M) has a natural
Cl(X

)-grading such that for all Cl(X

), we have
H
p
B()
(M)

= H
p
(

(f, M)

).
We will relate

C
p
(f, M)

to the

Cech complex for F() given in (9.1.1).
To compute

C
p
(f, M)

, rst observe that

C
p
(f, M)

(i
0
,...,i
p1
)[]
p1

M
f
i
0
f
i
p1

(i
0
,...,i
p1
)[]
p1

M()
f
i
0
f
i
p1

0
since shifting commutes with localization. For =(i
0
, . . . , i
p1
) []
p1
, note that
(9.5.3) f
i
0
f
i
p1
= x

i
0
x

i
p1
=

/ (1)
x
a

,
where a

= [i
j
[ /
i
j
[ > 0. If we set

=
i
0

i
p1
, then x

gives
the same localization as (9.5.3). Since F() is the sheaf associated to M(),
Proposition 5.3.3 implies that

M()
f
i
0
f
i
p1

0
=

M()
x

0
(U

, F()).
448 Chapter 9. Sheaf Cohomology of Toric Varieties
Comparing this with the

Cech complex (9.1.1) for the open cover U = U

i[]
,
we obtain

C
p
(f, M)

[]
p1
(U

, F())

C
p1
(U , F()).
This isomorphism is compatible with the differentials. It follows that the local

Cech
complex

C

(f, M)

is obtained from the



Cech complex

C

(U , F()) by deleting
the rst term and shifting the remaining terms. When p 2, this implies
H
p
B()
(M)

H
p1
(X

, F()),
and with a little work (Exercise 9.5.5) we also get an exact sequence
0 H
0
B()
(M)

H
0
(X

, F()) H
1
B()
(M)

0.
Theorems 9.5.5 and 9.5.7 imply that when p 1,
H
p
(X

, F()) lim

k
Ext
p+1
S
(S/B()
[k]
, M)

when F =

M and Cl(X

). We can compute Ext


p+1
S
(S/B()
[k]
, M)

by the
methods of Example 9.5.6. The problem is the direct limit. We tackle this next.
Stabilization of Ext. In the toric case, Ext has a nice relation to local cohomology.
Lemma 9.5.8. Let M be a nitely generated Cl(X

)-graded S-module and x


Cl(X

). If is a complete fan, then there exists k


0
N such that for all k k
0
, the
natural map S/B()
[k+1]
S/B()
[k]
induces an isomorphism
Ext
p
S
(S/B()
[k]
, M)

Ext
p
S
(S/B()
[k+1]
, M)

.
In particular, k k
0
implies
Ext
p
S
(S/B()
[k]
, M)

H
p
B()
(M)

.
Proof. We give a proof only for M=S following Mustat a [211, Thm. 1.1]. For the
general case, one replaces M with a free resolution and uses a spectral sequence
argument. See [91, Prop. 4.1] for the details.
Earlier we described Ext using the derived functors of MHom
R
(N, M) for
xed N. Ext also comes from the derived functors of NHom
R
(N, M) for xed M,
where one uses a projective resolution of N instead of an injective resolution of M
(see, for example, [89, A3.11]). In particular, we can compute Ext
p
S
(S/B()
[k]
, S)
using any free resolution of S/B()
[k]
. An especially nice resolution is given by
the Taylor resolution, which is described as follows.
We begin with S/B(). The minimal generators of B() are f
i
=x

i
for i [].
Let F
s
be a free S module with basis e
I
for all I [] with [I[ = s. To dene the
9.5. Local Cohomology and the Total Coordinate Ring 449
differential d
s
: F
s
F
s1
, take I, J [] with [I[ =[J[ +1 = s and list the elements
of I as i
1
< < i
s
. Then dene
c
IJ
=

0 J I
(1)
r
f
I
/f
J
I = J i
r
,
where f
I
= lcm( f
i
[ i I) and similarly for f
J
. Then
d
s
(e
I
) =

J
c
IJ
e
J
.
Since F
0
= S, we have an obvious map F
0
S/B(). One can prove that
0 F


d
2
F
1
d
1
F
0
S/B() 0
is exact (see [89, Ex. 17.11]). This is the Taylor resolution (F

, d

) of S/B().
This construction applies to any monomial ideal. In particular, it works for
B()
[k]
= ' f
k
1
, . . . , f
k

`. Let (F
[k]

, d
[k]

) denote the Taylor resolution of S/B()


[k]
.
It has the same modules F
[k]
s
= F
s
, and since the f
i
are square-free, we have f
k
I
=
lcm( f
k
i
[ i I). Thus the differentials d
[k]
s
in the Taylor resolution of S/B()
[k]
are
given by
d
[k]
s
(e
I
) =

J
c
k
IJ
e
J
, c
IJ
as above.
We now compare the Taylor resolutions of S/B()
[k]
and S/B()
[k+1]
. The
maps
s
: F
[k+1]
s
F
[k]
s
dened by
s
(e
I
) = f
I
e
I
induce a commutative diagram
F
[k+1]
s
s

d
[k+1]
s

F
[k]
s
d
[k]
s

F
[k+1]
s1

s1

F
[k]
s1
.
These maps are compatible with the surjection S/B()
[k+1]
S/B()
[k]
. Hence
(9.5.4) H
p
(Hom
S
(F
[k+1]

, S)) H
p
(Hom
S
(F
[k]

, S))
induced by the
s
can be identied with the canonical map
Ext
p
S
(S/B()
[k]
, S) Ext
p
S
(S/B()
[k+1]
, S).
The next key idea involves the use of a ner grading than the Cl(X

)-grading
used so far. Recall that this grading is induced by the map
(9.5.5) Z
(1)
Cl(X

)
where x
a
=

x
a

has degree [

]. The ring S also has a Z


(1)
-grading
where deg(x
a
) = a. Since B()
[k]
is a monomial ideal, the quotient ring S/B()
[k]
is Z
(1)
-graded. Then Ext
i
S
(S/B()
[k]
, S) inherits a natural Z
(1)
-grading.
Nowgrade the Taylor resolution (F
[k]

, d
[k]

) by setting deg(e
I
) =kdeg( f
I
). This
guarantees two things:
450 Chapter 9. Sheaf Cohomology of Toric Varieties
The differential d
[k]
s
has degree 0, so the isomorphism Ext
p
S
(S/B()
[k]
, S)
H
p
(Hom
S
(F
[k]

, S) is Z
(1)
-graded.
The map
s
has degree 0, so (9.5.4) is Z
(1)
-graded.
It follows that (F
[k]
s
)

=Hom
S
(F
[k]
s
, S) has dual basis e

I
with deg(e

I
) =kdeg( f
I
).
Given a, b Z
(1)
, dene
a b if and only if a

for all (1).


We claim that if a (k, . . . , k), then
(9.5.6) (

s
)
a
: (F
[k]
s
)

a
(F
[k+1]
s
)

a
is an isomorphism for all s.
To prove this, rst observe that

s
(e

I
) = f
I
e

I
. It follows that (

s
)
a
is injective.
Now take x
b
e

I
(F
[k+1]
s
)

a
. Then b(k +1)deg( f
I
) = a, so
b = (k +1)deg( f
I
) +a (k +1)deg( f
I
) +(k, . . . , k).
Since f
I
is square-free, (k +1)deg( f
I
) is a vector with whose th entry is (k +1) if
x

divides f
I
and 0 otherwise. Then the above inequality implies that f
I
divides x
b
,
so that x
b
e

I
is in the image of (

s
)
a
. It follows that (9.5.6) is an isomorphism.
The local cohomology H
p
B()
(S) has a Z
(1)
-grading which is compatible with
its Cl(X

)-grading via (9.5.5). This follows easily from Theorem 9.5.5 and the
Taylor resolution. If Cl(X

) and p 2, then
H
p
B()
(S)

H
p1
(X

, O
X

()).
by Theorem 9.5.7. The right-hand side is nite-dimensional since is complete.
This implies that when we decompose H
p
B()
(S) into its nonzero Z
(1)
-graded
pieces H
p
B()
(S)
a
, only nitely many can appear in H
p
B()
(S)

. For these nitely


many as, pick k
0
such that they all satisfy a (k
0
, . . . , k
0
). This k
0
has the
required properties. The argument for p = 0, 1 is covered in Exercise 9.5.6.
Theorem 9.5.7 and Lemma 9.5.8 imply that for p 1 and k 0,
H
p
(X

, O
X

()) H
p+1
B()
(S)

Ext
p+1
S
(S/B()
[k]
, S).
Notice how these isomorphisms generalize Examples 9.5.3 and 9.5.6. Our nal
task is to give an explicit method for deciding when k is big enough.
Bounds. For a complete fan , graded S-module M, and divisor class Cl(X

),
the paper [91] gives an explicit value for the number k
0
appearing in Lemma 9.5.8.
We will state the results of [91] without proof, though the following example sug-
gests some of the ideas involved.
9.5. Local Cohomology and the Total Coordinate Ring 451
Example 9.5.9. Let be a complete fan in M
R
R
2
with minimal generators
u
1
, . . . , u
r
arranged counterclockwise around the origin. Suppose we have a divisor
D =

i
a
i
D
i
on X

and m M such that H


1
(X

, O
X

(D))
m
= 0. It follows from
Proposition 9.1.6 that the sign pattern of m (determined by 'm, u
i
` +a
i
0 or < 0)
has at least two strings of consecutive s. The set I = i [r] [ 'm, u
i
` +a
i
< 0
records the locations of the s in the sign pattern of m.
Since H
1
(X

, O
X

(D))
m
=0 is nite-dimensional, there are only nitely many
ms with the same sign pattern. In other words, the inequalities
'm, u
i
` +a
i
< 0, i I
'm, u
i
` +a
i
0, i [r] `I
have only nitely many integer solutions, which means that the region of the plane
dened by these inequalities is bounded. You can see and example of this in Fig-
ure 4 from Example 9.1.8. Since this region determined by the u
i
and a
i
, it should
be possible to bound the size of the ms that appear in terms of the u
i
and a
i
.
To relate this to Lemma 9.5.8, let b = ('m, u
1
` +a
1
, . . . , 'm, u
r
` +a
r
). Then it
is easy to see that
H
1
(X

, O
X

(D))
m
H
2
B()
(S)
b
.
Thus bounding the ms with H
1
(X

, O
X

(D))
m
= 0 is equivalent to bounding the
bs with H
2
B()
(S)
b
= 0. And once we nd this bound, we know how to pick k
0
so
that b (k
0
, . . . , k
0
) for all such bs. The proof of Lemma 9.5.8 shows that this
k
0
works.
In the general case, we proceed as follows. Pick a basis e
1
, . . . , e
n
of M and
let A be the matrix whose rows give the coefcients of the u

s with respect to the


chosen basis. Then A is an r n matrix, where r =[(1)[. For this matrix, dene
(9.5.7)
q
n
= min([nonzero nn minors of A[)
Q
1
= max([entries of A[)
Q
n1
= max([(n1) (n1) minors of A[).
The following bound is proved in [91, Cor. 3.3].
Theorem 9.5.10. Given a complete fan in N
R
R
n
and a divisor class =
[

] Cl(X

), we have
Ext
p+1
S
(S/B()
[k]
, S)

H
p
B()
(S)

for all k k
0
, where
k
0
= n
2
max

([a

[)
Q
1
Q
n1
q
n
.
452 Chapter 9. Sheaf Cohomology of Toric Varieties
For a nitely generated graded S-module M, the formula for k
0
involves the
minimal free resolution of M. Write the resolution as
F
2
F
1
F
0
M0
and write
F
j
=

Cl(X

)
S()
r
j,
.
Finally, for each with r
j,
=0, write = [

a
,
D

] Cl(X

). Then [91, Prop.


4.1] gives the following bound.
Theorem 9.5.11. Let and be as in Theorem 9.5.10, and let M be a nitely
generated graded S-module. Then
Ext
p+1
S
(S/B()
[k]
, M)

H
p
B()
(M)

for all k k
0
, where
k
0
= n
2
max
, j,r
j,
,=0
([a

a
,
[)
Q
1
Q
n1
q
n
.
The Cotangent Bundle. We end this section by using our methods to calculate the
cohomology of the cotangent bundle of a smooth complete toric variety X

. The
rst step is to nd a graded S-module M with
1
X


M.
In (8.1.7) we constructed an exact sequence
0
1
S
M
Z
S

S/'x

`
where M
Z
S

S/'x

` is dened by m f

'm, u

`[ f ], and in Corol-
lary 8.1.5 we showed that
1
X

is the sheaf of
1
S
. However, we need a description
of
1
S
that is easier to implement on a computer. We do this as follows.
Pick bases of M and Pic(X

) and order the elements of (1) as


1
, . . . ,
r
.
Then the basic exact sequence
0 M Z
(1)
Pic(X

) 0
can be written
(9.5.8) 0 Z
n
A
Z
r
B
Z
rn
0,
where A is an r n matrix whose ith row consists of the coefcients of u

i
in
the basis of M. This is the same matrix A that appears in Theorem 9.5.10. The
(r n) r matrix B is called the Gale dual of A.
Lemma 9.5.12. The (r n) r matrix
= B

1
0 0
0 x

2
0
.
.
.
.
.
.
.
.
.
.
.
.
0 0 x
r

9.5. Local Cohomology and the Total Coordinate Ring 453


induces a graded homomorphism :

r
i=1
S(deg(x

i
)) S
rn
of degree 0 such
that
1
X

is the sheaf associated to ker().


Proof. Let x be the r r diagonal matrix of variables appearing in the statement of
the lemma. Similar to the proof of Theorem 8.1.6, we have a commutative diagram
0

0


1
S

Z
n

Z
S

A

r
i=1
S/'x

i
`

r
i=1
S(deg(x

i
))
x

Z
r

Z
S

B

r
i=1
S/'x

i
`

0
0

Z
rn

Z
S

Z
rn

Z
S

0
0 0
The rows are exact, as are the center and right columns. By the diagram chase from
the proof of Theorem 8.1.6, the dotted arrows are exact, and by commutivity, the
dotted arrow from

r
i=1
S(deg(x

i
)) to Z
rn

Z
S is given by = B x.
Example 9.5.13. For P
2
, the matrices A and B of (9.5.8) are given by
A =

1 0
0 1
1 1

, B =

1 1 1

.
Thus = (x y z), which gives the exact sequence
0
1
S
S(1)
3
(x y z)
S.
Using Macaulay2 as in Example B.4.2, one computes the free resolution
(9.5.9) 0 S(3)
0
@
z
x
y
1
A
S(2)
3

1
S
0.
The numbers from (9.5.7) are q
2
= Q
1
= 1, so that for a Z, the formula for k
0
from Theorem 9.5.11 is
k
0
= 4max([a2[, [a3[).
For a in the range 4 a 4, we can use k
0
= 28. This implies that for p 1,
H
p
(P
2
,
1
P
2
(a)) Ext
p+1
S
(S/'x
28
, y
28
, z
28
`,
1
S
)
a
when 4 a 4. This can be computed by the methods of Example 9.5.6. We
can also compute H
0
(P
2
,
1
P
2
(a)) directly from
1
S
(Exercise 9.5.7). The results
are shown in Table 3 on the next page.
454 Chapter 9. Sheaf Cohomology of Toric Varieties
a`p 0 1 2
4 0 0 15
3 0 0 8
2 0 0 3
1 0 0 0
0 0 1 0
1 0 0 0
2 3 0 0
3 8 0 0
4 15 0 0
Table 3. dim H
p
(P
2
,
1
P
2
(a)) for 4 a 4
In Exercise 9.5.8 you will calculate this table by hand. Notice also that the symme-
try of the table comes from Serre duality. See Example B.5.2 and Exercise B.5.1
for different way to compute Table 3.
Here is a slightly more complicated example.
Example 9.5.14. Let us compute the rst cohomology group of various twists of
the cotangent sheaf of the Hirzebruch surface H
2
. We will use the notation of
Example 9.3.6. The goal is to compute the cohomology groups of
1
H
2
(a, b) =

1
H
2
(aD
3
+bD
4
) for all a, b.
One way to proceed would be to follow the approach of the previous example
using Macaulay2. This would involve the following steps:
Construct the total coordinate ring S = Q[x
1
, x
2
, x
3
, x
4
] of H
2
is with its Z
2
-
grading and irrelevant ideal B ='x
1
x
3
, x
2
x
4
`.
Use the 24 matrix to dene the S-module
1
S
as a kernel as in Lemma 9.5.8.
Use a free resolution of
1
S
to compute the bound k
0
from Theorem 9.5.11.
Compute various graded pieces of Ext
p+1
S
(S/B
[k
0
]
,
1
S
) as explained in B.4 of
Appendix B.
Alternatively, we could use the Macaulay2 package NormalToricVarieties
described in Examples B.5.1 and B.5.2. This package automates most of the above
steps and makes it easy to calculate H
p
(H
2
,
1
H
2
(a, b)) directly and efciently.
Table 4 records some computations of H
1
done this way (see Example B.5.2 for
the details). There are many things we can see from this table, including:
Theorem 9.4.11 implies that dim H
1
(H
2
,
1
H
2
) = h
1
= f
1
2f
2
= 42 1 = 2
since H
2
comes from a quadrilateral (see Examples 2.3.16 and 3.1.16). This
explains the 2 when (a, b) = (0, 0).
Serre duality implies that H
1
(H
2
,
1
H
2
(a, b))

H
1
(H
2
,
1
H
2
(a, b)). This
explains the symmetry in the table.
9.5. Local Cohomology and the Total Coordinate Ring 455
a`b 2 1 0 1 2
2 0 0 3 4 3
1 0 0 2 2 2
0 1 1 2 1 1
1 2 2 2 0 0
2 3 4 3 0 0
Table 4. dim H
1
(H2,
1
H
2
(a, b)) for 2 a, b 2
Bott-Danilov-Steenbrink vanishing implies that H
1
(H
2
,
1
H
2
(a, b)) = 0 for
a, b > 0. This explains the 0s in the lower right corner of the table.
See [91, Sec. 5] for an algorithmic approach to these calculations. Using
[190, Cor. 3.4], one can describe an especially efcient method for computing the
sheaf cohomology of a toric variety. This method is implemented in Greg Smiths
Macaulay2 package NormalToricVarieites [252]. A equivalent method was
conjectured independently [35].
Exercises for 9.5.
9.5.1. Prove the exactness of (9.5.1).
9.5.2. Suppose that (A
i
,
i
), (A

i
,

i
), (A

i
,

i
) are directed systems, and that for each i,
there exist d
i
and
i
commuting with the s, such that we have an exact sequence
0 A
i
di
A

i
i
A

i
0.
Prove the exactness of the sequence
0 lim

i
A
i
lim

i
A

i
lim

i
A

i
0.
9.5.3. Prove (9.5.2) when Cl(X

) is the class of a Cartier divisor. Hint: Look carefully


at Proposition 5.3.3 and remember that the restriction of O
X
() to U

is trivial.
9.5.4. For I ='x, y` S =C[x, y], prove that H
2
I
(S) is not nitely generated as an S-module.
Hint: Use Example 9.5.3 to show that H
2
I
(S)
a
= 0 for all a 2.
9.5.5. Complete the proof of Theorem 9.5.7 by proving that there is an exact sequence
0 H
0
B()
(M)

H
0
(X

, F()) H
1
B()
(M)

0
for all Cl(X

).
9.5.6. Prove that H
0
B()
(S) = H
1
B()
(S) = 0. Hint: Proposition 5.3.7.
9.5.7. Here are some details to check from Example 9.5.13.
(a) Use the exact sequence from Lemma 9.5.12 to show that (
1
S
)

H
0
(X

,
1
X
())
for all Pic(X

) when X

is smooth and complete.


(b) Verify the rst column of Table 3.
(c) Show that the rst column agrees with the Bott formula from Example 9.4.10.
456 Chapter 9. Sheaf Cohomology of Toric Varieties
9.5.8. We can check Table 3 of Example 9.5.13 with a barehanded approach. Observe
that we have exact sequences involving
1
P
2
:
0
1
P
2 O
3
P
2 (1) O
P
2 0
0 O
P
2 (3) O
3
P
2 (2)
1
P
2 0.
The rst comes from Lemma 9.5.12, and the second comes by sheafying the free resolu-
tion of
1
S
computed in Example 9.5.13. Twist these sequences by a Z and consider the
resulting long exact sequences in cohomology. Using the vanishing theorems we know,
conclude that the nonzero values for H
p
(P
2
,
1
P
2
(a)) are:
dim H
0
(P
2
,
1
P
2 (a)) = a
2
1 if a 2
dimH
1
(P
2
,
1
P
2 (a)) = 1 if a = 1
dimH
2
(P
2
,
1
P
2 (a)) = a
2
1 if a 2.
These formulas give the numbers in Table 3.
Part II. Topics in
Toric Geometry
Chapters 10 to 15 explore further topics in theory of toric varieties. This
part of the book assumes a wider knowledge of algebraic geometry, though
we give careful references for all of the tools used in our study of toric
geometry. The topics presented here are just of few of many rich areas of
inquiry encountered in the study of toric varieties.
457
Chapter 10
Toric Surfaces
In this chapter, we will apply the theory developed so far to study the structure
of 2-dimensional normal toric varieties (toric surfaces). We will describe their
singularities, introduce the idea of a resolution of singularities, and also classify
smooth complete toric surfaces. Along the way, we will encounter two types of
continued fractions, Hilbert bases, the Gr obner fan, the M
c
Kay correspondence,
the Riemann-Roch theorem, the sectional genus, and the number 12.
10.1. Singularities of Toric Surfaces and Their Resolutions
Singular Points of Toric Surfaces. If X

is the toric surface of a fan in N


R
R
2
,
then minimal generators of the rays (1) are primitive and hence extend to a
basis of N. Then Theorem 3.1.19 implies that the toric surface obtained by re-
moving the xed points of the torus action (i.e., the points corresponding to the
2-dimensional cones under the Orbit-Cone Correspondence) is smooth. There are
only nitely many such points, so X

has at most nitely many singular points.


Moreover, 2-dimensional cones are always simplicial, so from Example 1.3.20,
each of these singular points is a nite abelian quotient singularity (isomorphic to
the image of the origin in the quotient C
2
/G where G is a nite abelian group).
All cones are assumed to be rational and polyhedral. A 2-dimensional strongly
convex cone in N
R
R
2
has the following normal form that will facilitate our study
of the singularities of toric surfaces.
Proposition 10.1.1. Let N
R
R
2
be a 2-dimensional strongly convex cone.
Then there exists a basis e
1
, e
2
for N such that
= Cone(e
2
, de
1
ke
2
),
where d > 0, 0 k < d, and gcd(d, k) = 1.
459
460 Chapter 10. Toric Surfaces
Proof. We will need the following modied division algorithm here and at several
other points in this chapter (Exercise 10.1.1):
(10.1.1)
Given integers l and d > 0, there are unique integers
s and k such that l = sd k and 0 k < d.
Say = Cone(u
1
, u
2
), where u
i
are primitive vectors. Since u
1
is primitive, we
can take it as part of a basis of N, and we let e
2
= u
1
. Since is strongly convex,
for any basis e

1
, e
2
for N, it will be true that
u
2
= de

1
+le
2
for some d = 0. By replacing e

1
by e

1
if necessary, we can assume d > 0. By
(10.1.1), there are integers s, k such that l = sd k, where 0 k < d. Using this
integer s, let e
1
= e

1
+se
2
. Then e
1
, e
2
is also a basis for N and
u
2
= de
1
+(l sd)e
2
= de
1
ke
2
.
Hence = Cone(e
2
, de
1
ke
2
) as claimed, and gcd(d, k) = 1 follows since u
2
is
primitive.
We will call the integers d, k in this statement the parameters of the cone ,
and e
1
, e
2
is called a normalized basis for N relative to . The uniqueness of d, k
will be studied in Proposition 10.1.3 below.
Using the normal form, we next describe the local structure of the point p

in the afne toric variety U

. Recall from Example 1.3.20 that if N

N is the
sublattice generated by the ray generators of , then U

C
2
/G, where G=N/N

.
In our situation, N =Ze
1
Ze
2
, and
N

=Ze
2
Z(de
1
ke
2
) = dZe
1
Ze
2
,
so it follows easily that
(10.1.2) G = N/N

Z/dZ.
In particular, for singularities of toric surfaces, the nite group G is always cyclic.
The action of G on C
2
is determined by the integers d, k as follows. We write

d
= C [
d
= 1
for the group of dth roots of unity in C. Then a choice of a primitive dth root of
unity denes an isomorphism of groups
d
Z/dZ.
Proposition 10.1.2. Let M

be the dual lattice of N

and let m
1
, m
2
M

be dual
to u
1
, u
2
in N

. Using the coordinates x =


m
1
and y =
m
2
of C
2
, the action of

d
N/N

on C
2
is given by
(x, y) = (x,
k
y).
Furthermore, U

C
2
/
d
with respect to this action.
10.1. Singularities of Toric Surfaces and Their Resolutions 461
Proof. The general discussion in 1.3 shows that the quotient N/N

Z/dZ acts
on the coordinate ring of C
2
via
(10.1.3) (u+N

)
m

= e
2i/m

,u)

,
where m

and u = je
1
for 0 j d 1.
An easy calculation shows that 'm
1
, e
1
` = 1/d and 'm
2
, e
1
` = k/d. Hence if
we set up the isomorphism
d
N/N

by mapping e
2i j/d
je
1
+N

, then for all


= e
2i j/d

d
, we have
(x, y) = (e
2i j/d
x, e
2i jk/d
y) = (x,
k
y)
by (10.1.3). This is what we wanted to show.
We next describe the slight but manageable ambiguity in the normal form for
2-dimensional cones. Two cones are lattice equivalent if there is a bijective Z-
linear mapping : N N taking one cone to the other. After choice of basis for
N, such mappings are dened by matrices in GL(2, Z).
Proposition 10.1.3. Let = Cone(e
2
, de
1
ke
2
) and = Cone(e

2
,

de

ke

2
) be
cones in normal form that are lattice equivalent. Then

d = d and either

k = k or
k

k 1 mod d.
Proof. Since the cones are lattice equivalent, writing N

and

N

for the sublattices


as in (10.1.2), there is a bijective Z-linear mapping : N N such that (N

) =

.
Hence N/

N/N

, so

d = d. The statement about k and

k is left to the reader in


Exercise 10.1.2.
Here are two examples to illustrate Proposition 10.1.2.
Example 10.1.4. First consider a cone
= Cone(e
2
, de
1
e
2
)
with parameters d > 1 (so the cone is not smooth) and k = 1. This is precisely
the cone considered in Example 1.2.22. The corresponding toric surface U

is
the rational normal cone

C
d
C
d+1
. The quotient

C
d
C
2
/
d
was studied in
the special case d = 2 in Example 1.3.19, and the general case was described in
Exercise 1.3.11. With the notation of Proposition 10.1.2,
d
acts on (x, y) C
2
via (x, y) = (x, y) and the ring of invariants is
C[x, y]

d
=C[x
d
, x
d1
y, . . . , xy
d1
, y
d
],
so
U

C
2
/
d
Spec(C[x
d
, x
d1
y, . . . , xy
d1
, y
d
]).
On the other hand, from Example 1.2.22, we also have the description
U

Spec(C[s, st, st
2
, . . . , st
d
]).
462 Chapter 10. Toric Surfaces
Exercise 10.1.3 studies the relation between these representations of the coordinate
ring of U

.
Example 10.1.5. Next consider a cone with parameters d and k = d 1, so
d = k +1. We will express everything in terms of the parameter k in the following.
Unlike the previous example, this is a case we have not encountered previously.
Note that k 1 mod d. Hence by Proposition 10.1.2, the action of G = N/N

on
C
2
is given by
(x, y) = (x,
1
y).
It is easy to check that the ring of invariants here is
C[x, y]

k+1
=C[x
k+1
, y
k+1
, xy].
Moreover we have an isomorphism of rings
: C[X,Y, Z]/'Z
k+1
XY` C[x
k+1
, y
k+1
, xy]
X x
k+1
Y y
k+1
Z xy,
so we may identify the toric surface U

with the variety V(Z


k+1
XY) C
3
.
The origin is the unique singular point of the afne variety of Example 10.1.5
and is called a rational double point (or Du Val singularity) of type A
k
. Another
standard form of these singularities is given in Exercise 10.1.4. They are called
double points because the lowest degree nonzero term in the dening equation has
degree two (i.e., the multiplicity of the singularity is two). The rational double
points are the simplest singularities from a certain point of view. The exact deni-
tion, which we will give in 10.4, depends on the notion of a resolution of singular-
ities, which will be introduced shortly. All rational double points appear as singu-
larities of quotient surfaces C
2
/G where G is a nite subgroup of SU(2, C). There
is a complete classication of such points in terms of the Dynkin diagrams of types
A
k
, D
k
, E
6
, E
7
, and E
8
. The groups corresponding to the diagrams D
k
, E
6
, E
7
, E
8
are
not abelian, so by the comment after (10.1.2), such points do not appear on toric
surfaces. We will see one way that the Dynkin diagram A
k
appears from the geom-
etry of the toric surface U

in Exercise 10.1.5, and we will return to this example


in 10.4. More details on these singularities can be found in [245, Ch. VI] and in
the article [85].
Here is another interesting aspect of Example 10.1.5. Recall that a normal
variety X is Gorenstein if its canonical divisor is Cartier (Denition 8.2.14). The
following result was proved in Exercise 8.2.13 of Chapter 8.
Proposition 10.1.6. For a cone = Cone(e
2
, de
1
ke
2
) in normal form, the afne
toric surface U

is Gorenstein if and only if k = d 1.


10.1. Singularities of Toric Surfaces and Their Resolutions 463
Toric Resolution of Singularities. Let X be a normal toric surface, and denote by
X
sing
the nite set of singular points of X (possibly empty).
Denition 10.1.7. A proper morphism : Y X is a resolution of singularities
of X if Y is a smooth surface and induces an isomorphism of varieties
(10.1.4) Y `
1
(X
sing
) X ` X
sing
.
Such a mapping modies X to produce a smooth variety without changing the
smooth locus X ` X
sing
. One of the most appealing aspects of toric varieties is the
way that many questions that are difcult for general varieties admit simple and
concrete solutions in the toric case. The problem of nding resolutions of singu-
larities is a perfect example. We illustrate this by constructing explicit resolutions
of singularities of the toric surfaces from Examples 10.1.4 and 10.1.5.
Example 10.1.8. Consider the rational normal cone of degree d, the afne toric
surface U

for = Cone(e
2
, de
1
e
2
) studied in Example 10.1.4. Let be the fan
in Figure 1 obtained by inserting a new ray = Cone(e
1
) subdividing into two
2-dimensional cones:

1
= Cone(e
2
, e
1
)

2
= Cone(e
1
, de
1
e
2
).

1

2

Figure 1. The cone and the renement given by 1, 2,
We now use some results from Chapter 3. The identity mapping on the lattice
N is compatible with the fans and as in Denition 3.3.1. By Theorem 3.3.4,
we have a corresponding toric blowup morphism
(10.1.5) : X

.
Note that both
1
and
2
(as well as all of their faces) are smooth cones. Hence The-
orem 3.1.19 implies that X

is a smooth surface. In addition, the toric morphism


is proper by Theorem 3.4.11 since is a renement of . Finally, we claim that
satises (10.1.4). This follows from the Orbit-Cone Correspondence on the two
464 Chapter 10. Toric Surfaces
surfaces: if p

is the distinguished point corresponding to the 2-dimensional cone


(the singular point of U

at the origin), then restricts to an isomorphism


X

`
1
(p

) U

`p

= (U

)
smooth
.
The inverse image E =
1
(p

) is the curve on X

given by the closure of the


T
N
-orbit O() corresponding to the ray . That is, the singular point blows up
to E P
1
on the smooth surface. It follows that X

and the morphism (10.1.5)


give a toric resolution of singularities of the rational normal cone. We call E the
exceptional divisor on the smooth surface. We will say more about how E sits
inside the surface X

in 10.4.
Example 10.1.9. We consider the case d = 4 of Example 10.1.5, for which the
surface U

has a rational double point of type A


3
. We will leave the details, as well
as the generalization to all d 2, to the reader (Exercise 10.1.5). It is easy to nd
subdivisions of
= Cone(e
2
, 4e
1
3e
2
)
yielding collections of smooth cones. The most economical way to do this is to
insert three new rays
1
= Cone(e
1
),
2
= Cone(2e
1
e
2
),
3
= Cone(3e
1
2e
2
)
to obtain a fan consisting of four 2-dimensional cones and their faces.
The fan produced by this subdivision is somewhat easier to visualize if we draw
the cones relative to a different basis u
1
, u
2
for N. For u
1
= e
2
and u
2
= e
1
e
2
, the
cone = Cone(u
1
, u
1
+4u
2
) and the fan with maximal cones
(10.1.6)

1
= Cone(u
1
, u
1
+u
2
)

2
= Cone(u
1
+u
2
, u
1
+2u
2
)

3
= Cone(u
1
+2u
2
, u
1
+3u
2
)

4
= Cone(u
1
+3u
2
, u
1
+4u
2
)
appear in Figure 2.

1

4
Figure 2. The cone and the renement
10.1. Singularities of Toric Surfaces and Their Resolutions 465
You will check that each of these cones is smooth. Hence X

is a smooth
surface. Since is a renement of , we have a proper toric morphism
: X

.
As in the previous example, restricts to an isomorphism from X

`
1
(p

) to
X

`p

. In this case, the exceptional divisor E =


1
(p

) is the union
E =V(
1
) V(
2
) V(
3
)
on X

. The curves V(
i
) are isomorphic to P
1
. The rst two intersect transversely
at the xed point of the T
N
-action on X

corresponding to the cone


2
, while the
second two intersect transversely at the xed point corresponding to
3
.
In these examples, we constructed toric resolutions of afne toric surfaces with
just one singular point. The same techniques can be applied to any normal toric
surface X

.
Theorem 10.1.10. Let X

be a normal toric surface. There exists a smooth fan

rening such that the associated toric morphism : X

is a toric
resolution of singularities.
Proof. It sufces to show the existence of the smooth fan

rening . The
reasoning given in Example 10.1.8 applies to show that the corresponding toric
morphism is proper and birational, hence a resolution of singularities of X

.
We will prove this by induction on an integer invariant of fans that measures
the complexity of the singularities on the corresponding surfaces. Let
1
, . . . ,

denote the 2-dimensional cones in a fan . For each i, we will write N


i
for the
sublattice of N generated by the ray generators of
i
. Then we dene
s() =

i=1
(mult(
i
) 1) ,
where mult(
i
) = [N : N
i
] as in 6.3. If s() = 0, then = 0 or mult(
i
) = 1 for
all i. It is easy to see that this implies that is a smooth fan. Hence X

is already
smooth and we take this as the base case for our induction.
For the induction step, we assume that the existence of smooth renements has
been established for all fans with s() < s, and consider a fan with s() = s.
If s 1, then there exists some nonsmooth cone
i
in . By Proposition 10.1.1,
there is a basis e
1
, e
2
for N such that
i
= Cone(e
2
, de
1
ke
2
) with parameters
d > 0, 0 k < d, and gcd(d, k) = 1. Consider the renement

of obtained by
subdividing the cone
i
into two new cones

i
= Cone(e
2
, e
1
)

i
= Cone(e
1
, de
1
ke
2
)
466 Chapter 10. Toric Surfaces
with a new 1-dimensional cone = Cone(e
1
). We must show that s(

) < s() to
invoke the induction hypothesis and conclude the proof.
In s(), the terms corresponding to the other cones
j
for j = i are unchanged.
The cone

i
is smooth since e
1
, e
2
is the normalized basis of N relative to
i
. So it
contributes a zero term in s(

). Now consider the cone

i
. In order to compute
its contribution to s(

), we must determine the parameters of

i
.
In terms of the basis e
1
, e
2
for N, the Z-linear mapping dened by the matrix
A =

0 1
1 0

(a 90-degree rotation) takes

i
to Cone(e
2
, ke
1
+de
2
). Since A GL(2, Z), it
denes an automorphism of N, and hence

i
will have the same parameters as
Cone(e
2
, ke
1
+de
2
). But now we apply (10.1.1) to write
(10.1.7) d = sk l
where 0 l < k. Since gcd(d, k) = 1, we have gcd(k, l) = 1 as well. Hence the
cone

i
has parameters k and l obtained from (10.1.7). Since k < d, if N

i
is the
sublattice generated by the ray generators of

i
, then by (10.1.2),
[N : N

i
] = k < [N : N
i
] = d.
It follows that s(

) < s(), and the proof is complete by induction.


We will see in the next section that in the afne case, the renement that gives
the resolution of singularities of U

has a very nice description. As a preview,


notice that in Examples 10.1.8 and 10.1.9, the renement of the given cone was
produced by subdividing along the rays through the Hilbert basis (the irreducible
elements) of the semigroup N.
A resolution of a non-normal toric surface singularity can be constructed by
rst saturating the associated semigroup as in Theorem 1.3.5, then applying the
results of this section. Toric resolutions of singularities for toric varieties of di-
mension three and larger also exist. However, we postpone the higher-dimensional
case until Chapter 11.
Exercises for 10.1.
10.1.1. Adapt the usual proof of the integer division algorithm to prove (10.1.1).
10.1.2. In this exercise, you will develop further properties of the parameters d, k in the
normal form for cones from Proposition 10.1.1 and prove part of Proposition 10.1.3.
(a) Show that if is obtained from a cone by parameters d, k by a Z-linear mapping of
N dened by a matrix in GL(2, Z), then the parameter

k of satises either

k = k, or
k

k 1 mod d. Hint: There is a choice of orientation to be made in the normalization


process. Recall that gcd(d, k) = 1, so there are integers

d,

k such that d

d +k

k = 1.
10.2. Continued Fractions and Toric Surfaces 467
(b) Show that if is a cone with parameters d, k, then the dual cone

M
R
has param-
eters d, d k. Hint: Use the normal form for , write down

in the corresponding
dual basis in M, then change bases in M to normalize

.
10.1.3. With the notation in Example 10.1.4, show that
C[s, st, st
2
, . . . , st
d
] C[x
d
, x
d1
y, . . . , xy
d1
, y
d
]
under s x
d
and t y/x, and use Proposition 10.1.2 to explain where these identications
come from in terms of the semigroup S

. Hint: We have s =
m1
and t =
m2
where e
1
, e
2
is the normalized basis for N and m
1
, m
2
is the dual basis for M.
10.1.4. In Example 10.1.5, we gave one form of the rational double point of type A
k
,
namely the singular point at (0, 0, 0) on the surface V = V(Z
k+1
XY) C
3
. Another
commonly used normal form for this type of singularity is the singular point at (0, 0, 0) on
the surface W =V(X
k+1
+Y
2
+Z
2
). Show that V and W are isomorphic as afne varieties,
hence the singularities at the origin are analytically equivalent. Hint: There is a linear
change of coordinates in C
3
that does this.
10.1.5. In this exercise, you will check the claims made in Example 10.1.9 and show how
to extend the results there to the case = Cone(e
2
, de
1
(d 1)e
2
) for general d.
(a) Check that each of the four cones in (10.1.6) is smooth, so that the toric surface X

is
smooth by Theorem 3.1.19.
(b) For general d, show how to insert new rays
i
to subdivide and obtain a fan whose
associated toric surface is smooth. Try to do this with as few new rays as possible.
Hence we obtain toric resolutions of singularities : X

for all d.
(c) Identify the inverse image C =
1
(p

) in general. For instance, how many irre-


ducible components does C have? How are they connected? Hint: One way to repre-
sent the structure is to drawa graph with vertices corresponding to the components and
connect two vertices by an edge if and only if the components intersect on X

. Do you
notice a relation between this graph and the Dynkin diagram A
k
= A
d1
mentioned
before? We will discuss the relation in detail in 10.4.
10.2. Continued Fractions and Toric Surfaces
To relate continued fractions to toric surfaces, we begin with the afne toric surface
U

of a cone N
R
R
2
in normal form with parameters d, k. We will always
assume d > k > 0, so that U

has a unique singular point.


Hirzebruch-Jung Continued Fractions. When we construct a resolution of singu-
larities of U

by following the proof of Theorem 10.1.10, the rst step is to rene


the cone = Cone(e
2
, de
1
ke
2
) to a fan containing the 2-dimensional cones

= Cone(e
2
, e
1
) and

= Cone(e
1
, de
1
ke
2
).
The rst is smooth, but the second may not be. However, we saw in the proof of
Theorem 10.1.10 that the cone

has parameters k, k
1
satsifying
d = b
1
k k
1
,
468 Chapter 10. Toric Surfaces
where b
1
2, 0 k
1
< k as in (10.1.1). We used slightly different notation before,
writing s rather than b
1
and l rather than k
1
; the new notation will help us keep
track of what happens as we continue the process and rene the cone

.
Using the normalized basis for N relative to

, we insert a new ray and obtain


a new smooth cone and a second, possibly nonsmooth cone with parameters k
1
, k
2
,
where
k = b
2
k
1
k
2
using (10.1.1). Doing this repeatedly yields a modied Euclidean algorithm
(10.2.1)
d = b
1
k k
1
k = b
2
k
1
k
2
.
.
.
k
r3
= b
r1
k
r2
k
r1
k
r2
= b
r
k
r1
that computes the parameters of the new cones produced as we successively sub-
divide to produce the fan giving the resolution of singularities. The process termi-
nates with k
r
= 0 for some r as shown, since as in the usual Euclidean algorithm,
the k
i
are a strictly decreasing sequence of nonnegative numbers. Also, by (10.1.1),
we have b
i
2 for all i.
The equations (10.2.1) can be rearranged:
(10.2.2)
d/k = b
1
k
1
/k
k/k
1
= b
2
k
2
/k
1
.
.
.
k
r3
/k
r2
= b
r1
k
r1
/k
r2
k
r2
/k
r1
= b
r
and spliced together to give a type of continued fraction expansion for the rational
number d/k, with minus signs:
(10.2.3) d/k = b
1

1
b
2

1

1
b
r
.
This is the Hirzebruch-Jung continued fraction expansion of d/k. For obvious
typographical reasons, it is desirable to have a more compact way to represent
these expressions. We will use the notation
d/k = [[b
1
, b
2
, . . . , b
r
]].
10.2. Continued Fractions and Toric Surfaces 469
The integers b
i
are the partial quotients of the Hirzebruch-Jung continued fraction,
and the truncated Hirzebruch-Jung continued fractions
[[b
1
, b
2
, . . . , b
i
]], 1 i r,
are the convergents.
Example 10.2.1. Consider the rational number 17/11. The Hirzebruch-Jung con-
tinued fraction expansion is
17/11 = [[2, 3, 2, 2, 2, 2]],
as may be veried directly using the modied Euclidean algorithm (10.2.1).
Proposition 10.2.2. Let d > k > 0 be integers with gcd(d, k) = 1 and let d/k =
[[b
1
, . . . , b
r
]]. Dene sequences P
i
and Q
i
recursively as follows. Set
(10.2.4)
P
0
= 1, Q
0
= 0
P
1
= b
1
, Q
1
= 1,
and for all 2 i r, let
(10.2.5)
P
i
= b
i
P
i1
P
i2
Q
i
= b
i
Q
i1
Q
i2
.
Then the P
i
, Q
i
satisfy:
(a) The P
i
and Q
i
are increasing sequences of integers.
(b) [[b
1
, . . . , b
i
]] = P
i
/Q
i
for all 1 i r.
(c) P
i1
Q
i
P
i
Q
i1
= 1 for all 1 i r.
(d) The convergents form a strictly decreasing sequence:
d
k
=
P
r
Q
r
<
P
r1
Q
r1
< <
P
1
Q
1
.
Proof. The proof of part (a) is left to the reader (Exercise 10.2.1).
To prove part (b), rst observe that the expression on the right side of (10.2.3)
makes sense when the b
j
are any rational numbers (not just integers) such that all
denominators in (10.2.3) are nonzero. We will show that the sequences dened by
(10.2.5) satisfy
[[b
1
, . . . , b
s
]] =
P
s
Q
s
for all such lists b
1
, . . . , b
s
. The proof is by induction on the length s of the list.
When s = 1, we have [[b
1
]] = b
1
=
P
1
Q
1
by (10.2.4). Now assume that the result has
been proved for all lists of of length t and consider the expression
[[b
1
, . . . , b
t+1
]] = [[b
1
, . . . , b
t

1
b
t+1
]],
470 Chapter 10. Toric Surfaces
where the right side comes from a list of length t. By the induction hypothesis, this
equals

b
t

1
b
t+1

P
t1
P
t2

b
t

1
b
t+1

Q
t1
Q
t2
.
By the recurrences (10.2.5), this equals
P
t

1
b
t+1
P
t1
Q
t

1
b
t+1
Q
t1
=
b
t+1
P
t
P
t1
b
t+1
Q
t
Q
t1
=
P
t+1
Q
t+1
,
which is what we wanted to show.
Part (c) will be proved by induction on i. The base case i = 1 follows directly
from (10.2.4). Now assume that the result has been proved for i s, and consider
i = s +1. Using the recurrences (10.2.5), we have
P
s
Q
s+1
P
s+1
Q
s
= P
s
(b
s
Q
s
Q
s1
) (b
s
P
s
P
s1
)Q
s
= P
s1
Q
s
P
s
Q
s1
= 1
by the induction hypothesis.
Finally, from part (b), for each 1 i r 1, we have
P
i1
Q
i1
=
P
i
Q
i
+
1
Q
i1
Q
i
.
Hence
P
i
Q
i
<
P
i1
Q
i1
since Q
i1
Q
i
> 0 by part (a). Hence part (d) follows.
Hirzebruch-Jung Continued Fractions and Resolutions. When is a cone with
parameters d > k > 0, the process of computing the Hirzebruch-Jung continued
fraction of d/k yields a convenient method for nding a renement of such
that : X

is a toric resolution of singularities.


Theorem 10.2.3. Let = Cone(e
2
, de
1
ke
2
) be in normal form. Let u
0
= e
2
and
use the integers P
i
and Q
i
from Proposition 10.2.2 to construct vectors
u
i
= P
i1
e
1
Q
i1
e
2
, 1 i r +1.
Then the cones

i
= Cone(u
i1
, u
i
), 1 i r +1,
have the following properties:
(a) Each
i
is a smooth cone and u
i1
, u
i
are its ray generators.
(b) For each i,
i+1

i
= Cone(u
i
).
10.2. Continued Fractions and Toric Surfaces 471
(c)
1

r+1
= , so the fan consisting of the
i
and their faces gives a
smooth renement of .
(d) The toric morphism : X

is a resolution of singularities.
Proof. Both statements in part (a) follow easily from part (c) of Proposition 10.2.2.
For part (b), we note that the ratio Q
i1
/P
i1
represents the slope of the line
through u
i
in the coordinate system relative to the normalized basis e
1
, e
2
for . By
part (d) of Proposition 10.2.2, these slopes form a strictly decreasing sequence for
i 0, which implies the statement in part (b).
Part (c) follows from part (b) by noting that u
0
= e
2
and P
r
/Q
r
= d/k, so
u
r+1
= de
1
ke
2
. Hence the cones
i
ll out .
Part (d) now follows by the reasoning used in Examples 10.1.8 and 10.1.5.
Example 10.2.4. Consider the cone = Cone(e
2
, 7e
1
5e
2
) in normal form. To
construct the resolution of singularities of the afne toric surface U

, we simply
compute the Hirzebruch-Jung continued fraction expansion of the rational number
d/k = 7/5 using the modied Euclidean algorithm:
7 = 2 53
5 = 2 31
3 = 3 1.
Hence b
0
= b
1
= 2, b
2
= 3, and
(10.2.6) 7/5 = [[2, 2, 3]].
Then from Proposition 10.2.2 we have
P
0
= 1, Q
0
= 0
P
1
= 2, Q
1
= 1
P
2
= b
2
P
1
P
0
= 3, Q
2
= b
2
Q
1
Q
0
= 2
P
3
= b
3
P
2
P
1
= 7, Q
3
= b
3
Q
2
Q
1
= 5.
Theorem 10.2.3 gives the vectors
u
0
= e
2
, u
1
= e
1
, u
2
= 2e
1
e
2
, u
3
= 3e
1
2e
2
, u
4
= 7e
1
5e
2
and the cones
(10.2.7)

1
= Cone(e
2
, e
1
)

2
= Cone(e
1
, 2e
1
e
2
)

3
= Cone(2e
1
e
2
, 3e
1
2e
2
)

4
= Cone(3e
1
2e
2
, 7e
1
5e
2
)
shown in Figure 3 on the next page. The cones
i
give a smooth renement of .
You will see another example of this process in Exercise 10.2.2.
472 Chapter 10. Toric Surfaces

4
Figure 3. The renement with open circles at ui =Pi1e1 Qi1e2 in Example 10.2.4
Next we show that the vectors u
i
from Theorem 10.2.3 determine the partial
quotients in the Hirzebruch-Jung continued fraction expansion of d/k.
Theorem 10.2.5. Let = Cone(e
2
, de
1
ke
2
) be in normal form, and let
d/k = [[b
1
, b
2
, . . . , b
r
]].
Then the vectors u
0
, u
1
, . . . , u
r+1
constructed in Theorem 10.2.3 satisfy
(10.2.8) u
i1
+u
i+1
= b
i
u
i
, b
i
2,
for 1 i r.
Proof. By the recurrences (10.2.5),
u
i1
+u
i+1
= (P
i2
e
1
Q
i2
e
2
) +(P
i
e
1
Q
i
e
2
)
= (P
i2
+P
i
)e
1
(Q
i2
+Q
i
)e
2
= b
i
(P
i1
e
1
Q
i1
e
2
) = b
i
u
i
.
Later in this chapter we will see several important consequences of (10.2.8)
connected with the geometry of smooth toric surfaces.
The nonuniqueness of Proposition 10.1.3 has a nice relation to Theorem 10.2.3.
For instance, Example 10.2.4 used the Hirzebruch-Jung expansion 7/5 = [[2, 2, 3]].
Since 5 3 1 mod 7, the cone of Example 10.2.4 also has parameters d =7, k =3.
We leave it to the reader to check that
7/3 = [[3, 2, 2]],
with the partial quotients the same as those in (10.2.6), but listed in reverse order.
This pattern holds for all Hirzebruch-Jung continued fractions. We give a proof
that uses the properties of the associated toric surfaces.
10.2. Continued Fractions and Toric Surfaces 473
Proposition 10.2.6. Let 0 <k,

k <d and assume k

k 1 mod d. If the Hirzebruch-


Jung continued fraction expansion of d/k is
d/k = [[b
1
, b
2
, . . . , b
r
]],
then the Hirzebruch-Jung continued fraction expansion of d/

k is
d/

k = [[b
r
, b
r1
, . . . , b
1
]].
Proof. Let = Cone(e
2
, de
1
ke
2
) and = Cone(e
2
, de
1

ke
2
) be the corre-
sponding cones in normal form. Since k

k 1 mod m, there is an integer



d such
that d

d +k

k =1. The Z-linear mapping : N N dened with respect to the basis


e
1
, e
2
by the matrix
A =

k d

d k

is bijective, maps to , and is orientation-reversing. Thus (de


1

ke
2
) = e
2
and
(e
2
) = de
1
ke
2
. If we apply Theorem 10.2.3 to , then we obtain vectors u
i
satisfying the equations
u
i1
+u
i+1
= b
i
u
i
for all 1 i r. We claim that when we apply the mapping
1
dened by the
inverse of the matrix A above, then the vectors u
i
are taken to corresponding vec-
tors u
i
for the cone . But since and
1
are orientation-reversing, the partial
quotients in the Hirzebruch-Jung continued fraction will be listed in the opposite
order. You will complete the proof of this assertion in Exercise 10.2.3.
Hilbert Bases and Convex Hulls. Our next result gives two alternative ways to
understand the vectors u
i
in Theorem 10.2.3. The idea is that gives two objects:
The semigroup N. Since is strongly convex, its irreducible elements
form the unique minimal generating set called the Hilbert basis of N. (See
Proposition 1.2.23.)
The convex hull

= Conv( (N `0)). This is an unbounded polygon in


the plane whose bounded edges contain nitely many lattice points.
Example 10.2.7. Consider the cone =Cone(e
2
, 7e
1
5e
2
) from Example 10.2.4.
Figure 4 on the next page shows the convex hull

, where the white circles rep-


resent the lattice points on the bounded edges. We will see below that these lattice
points give the Hilbert basis of N.
Here is the general result suggested by Example 10.2.7.
Theorem 10.2.8. Let = Cone(e
2
, de
1
ke
2
) be in normal form and let
S =u
0
, u
1
, . . . , u
r+1

be the set of vectors constructed in Theorem 10.2.3. Then:


474 Chapter 10. Toric Surfaces

Figure 4. Convex hull and lattice points on bounded edges in Example 10.2.7
(a) S is the Hilbert basis of the semigroup N.
(b) S is the set of lattice points on the bounded edges of

.
Proof. For part (a), we use the notation of Theorem 10.2.3, where the cone
i
is
generated by u
i1
, u
i
. Then
i
N is generated as a semigroup by u
i1
, u
i
since
i
is smooth. Using =
1

r+1
, one sees easily that S generates N.
We claim next that all the u
i
are irreducible elements of N. This is clear for
u
0
=e
2
and u
r+1
=de
1
ke
2
since they are the ray generators for . If 1 i r and
u
i
is not irreducible, then u
i
would have to be a linear combination of the vectors
in S`u
i
with nonnegative integer coefcients, i.e.,
u
i
= P
i1
e
1
Q
i1
e
2
=

j,=i
c
j
u
j
=

j,=i
c
j
P
j1

e
1

j,=i
c
j
Q
j1

e
2
with c
j
0 in Z. Hence
P
i1
=

j,=i
c
j
P
j1
, Q
i1
=

j,=i
c
j
Q
j1
.
Since the P
i
and Q
i
are strictly increasing by part (a) of Proposition 10.2.2, we must
have c
j
= 0 for all j > i. But this would imply that u
i
is a linear combination with
nonnegative integer coefcients of the vectors in u
0
, . . . , u
i1
. This contradicts
the observation made in the proof of Theorem 10.2.3 that the slopes of the u
i
are
strictly decreasing. It follows that the u
i
are irreducible elements of N.
Finally, we must show that there are no other irreducible elements in N.
But this follows from what we have already said. Since =
1

r+1
, if u
is irreducible, then u
i
N for some i. But then u = c
i1
u
i1
+c
i
u
i
for some
c
i1
, c
i
0 in Z. Thus u is irreducible only if u = u
i1
or u
i
.
10.2. Continued Fractions and Toric Surfaces 475
For part (b), rst observe that by Proposition 10.2.2, we have
P
i2
Q
i
P
i
Q
i2
= P
i2
(b
i
Q
i1
Q
i2
) (b
i
P
i1
P
i2
)Q
i2
= b
i
(P
i2
Q
i1
P
i1
Q
i2
) = b
i
2.
Combining this with part (c) of Proposition 10.2.2, one obtains the inequality
(Q
i1
Q
i2
)
P
i1
P
i2

(Q
i
Q
i1
)
P
i
P
i1
.
Since u
i
= P
i1
e
1
Q
i1
e
2
, this inequality tells us that the slopes of the line seg-
ments u
i1
u
i
and u
i
u
i+1
are related by
slope of u
i1
u
i
slope of u
i
u
i+1
.
This implies that these line segments lie on boundary of

. From here, it is easy


to see that the u
i
are the lattice points of the bounded edges of

.
Ordinary Continued Fractions. The Hirzebruch-Jung continued fractions studied
above are less familiar than ordinary continued fraction expansions in which the
minus signs are replaced by plus signs. If d > k > 0 are integers, then the or-
dinary continued fraction expansion of d/k may be obtained by performing the
same sequence of integer divisions used in the usual Euclidean algorithm for the
gcd. Starting with k
1
= d and k
0
= k, we write a
i
for the quotient and k
i
for the
remainder at each step, so that the ith division is given by
(10.2.9) k
i2
= a
i
k
i1
+k
i
,
where 0 k
i
< k
i1
.
Let k
s1
be the nal nonzero remainder (which equals gcd(d, k)). The resulting
equations splice together to form the ordinary continued fraction
d/k = a
1
+
1
a
2
+
1
+
1
a
s
.
To distinguish these from Hirzebruch-Jung continued fractions, we will use the
notation
(10.2.10) d/k = [a
1
, a
2
, . . . , a
s
]
for the ordinary continued fraction. The a
i
are the (ordinary) partial quotients of
d/k, and the truncated continued fractions
[a
1
, a
2
, . . . , a
i
], 1 i s,
are the (ordinary) convergents of d/k.
476 Chapter 10. Toric Surfaces
Example 10.2.9. By Example 10.2.1, the Hirzebruch-Jung continued fraction of
17/11 is
17/11 = [[2, 3, 2, 2, 2, 2]],
and the ordinary continued fraction is
17/11 = [1, 1, 1, 5].
The partial quotients and the lengths are different. However, each expansion deter-
mines the other, and there are methods for computing the Hirzebruch-Jung partial
quotients b
j
in terms of the ordinary partial quotients a
i
and vice versa. See [75,
Prop. 3.6], [145, p. 257], [231, Prop. 2.3], and Exercise 10.2.4.
The following result is mostly parallel to Proposition 10.2.2, but shows that
ordinary continued fractions are slightly more complicated than Hirzebruch-Jung
continued fractions. The proof is left to the reader (Exercise 10.2.5).
Proposition 10.2.10. Let d > k > 0 be integers with gcd(d, k) = 1, and let d/k =
[a
1
, . . . , a
s
]. Dene sequences p
i
and q
i
recursively as follows. First set
(10.2.11)
p
0
= 1, q
0
= 0
p
1
= a
1
, q
1
= 1,
and for all 2 i s, let
(10.2.12)
p
i
= a
i
p
i1
+ p
i2
q
i
= a
i
q
i1
+q
i2
.
Then the p
i
, q
i
satisfy:
(a) [a
1
, . . . , a
i
] = p
i
/q
i
for all 1 i s.
(b) p
i
q
i1
p
i1
q
i
= (1)
i
for all 1 i s.
(c) The convergents converge to d/k, but in an oscillating fashion:
p
1
q
1
<
p
3
q
3
<
d
k
<
p
4
q
4
<
p
2
q
2
.
Ordinary Continued Fractions and Convex Hulls. Felix Klein discovered a lovely
geometric interpretation of ordinary continued fractions. Given a basis u
o
1
, u
o
0
of
N Z
2
and relatively prime integers d > k > 0, compute the continued fraction
d/k = [a
1
, . . . , a
s
] and set
(10.2.13) u
o
i
= q
i
u
o
1
+ p
i
u
o
0
, 1 i s.
In this notation, the superscript o stands for ordinary. Then u
o
s
= ku
o
1
+du
o
0
,
and part (c) of Proposition 10.2.10 implies that the u
o
i
lie on one side of the ray
Cone(u
o
s
) for even indices and on the other side for odd indices. To give a careful
description of what is happening, we introduce the cones

1
= Cone(u
o
1
, u
o
s
),
0
= Cone(u
o
0
, u
o
s
)
and associated convex hulls
i
=

i
= Conv(
i
(N`0)), i =1, 0.
10.2. Continued Fractions and Toric Surfaces 477

1
u
0
o
u
1
o
u
1
o
u
3
o
u
2
o
Figure 5. The cones 1, 0, the convex hulls 1, 0, and their vertices
Example 10.2.11. For d = 7, k = 5, the expansion 7/5 = [1, 2, 2] gives the vectors
u
o
1
, . . . , u
o
3
shown in Figure 5. In this gure, it is clear that the u
o
i
are the vertices
of the convex hulls
1
and
0
.
This example is a special case of the following general result.
Theorem 10.2.12. For u
o
1
, . . . , u
o
s
and
1
,
0
as above, we have:
(a)
1
has vertex set u
o
2j1
[ 1 j s/2u
o
s
.
(b)
0
has vertex set u
o
2j
[ 0 j s/2u
o
s
.
(c) For 1 i s, u
o
i2
u
o
i
is an edge of
1
(resp.
0
) for i odd (resp. even) with
a
i
+1 lattice points.
Proof. First note that by Proposition 10.2.10, the vectors u
o
i
satisfy the recursion
(10.2.14) u
o
i
= a
i
u
o
i1
+u
o
i2
, 1 i s.
Since u
o
i1
is primitive (Proposition 10.2.10), part (c) follows from parts (a) and (b).
We now prove the theorem using induction on the length s of the continued
fraction expansion. Consider Figure 6 on the next page, which shows the rst quad-
rant determined by u
o
1
, u
o
0
, together with the vector u
o
s
. In the picture,
0
(darker)
lies above the ray determined by u
o
s
and
1
(lighter) lies below. The gure also
shows u
o
1
and the smaller cone
1
= Cone(u
o
1
, u
o
s
)
1
= Cone(u
o
1
, u
o
s
).
Since u
o
s
= ku
o
1
+du
o
0
, the ray starting from u
o
1
through u
o
1
passes through
the upper edge of
1
at a point between u
o
1
+d/ku
o
1
and u
o
1
+(d/k+1)u
o
1
.
But by the computation of the ordinary continued fraction, d/k = a
1
. Therefore
the segment from u
o
1
to u
o
1
is the rst bounded edge of
1
. It follows that

1
Cone(u
o
1
, u
o
1
)
478 Chapter 10. Toric Surfaces

1
u
0
o
u
1
o
u
1
o
= a
1
u
0
o
+ u
1
o
u
s
o
Figure 6. The cones 0, 1 1 and vectors u
o
1
, u
o
0
, u
o
1
, u
o
s
has vertices u
o
1
, u
o
1
. It remains to understand
0
and
1

1
. This is where
induction comes into play.
Apply the above construction to the new basis u
o
0
, u
o
1
and the continued fraction
k
d a
1
k
=
1
d
k
a
1
= [a
2
, . . . , a
s
]
of length s 1. If we start numbering at 0 rather than at 1, then the vectors
u
o
0
, u
o
1
, . . . , u
o
s
are the same as before by the recursion (10.2.14). This gives the
cones
0
,
1
shown in Figure 6. By induction, the vectors u
o
0
, u
o
1
, . . . , u
o
s
give
the vertices of the corresponding convex hulls
0
= Conv(
0
(N ` 0)) and

1
= Conv(
1
(N `0)). It follows easily that the theorem holds for continued
fraction expansions of length s.
A discussion of Kleins formulation of Theorem 10.2.12 can be found in [231].
In Exercise 10.2.6 you will apply the theorem to the resolution of pairs of singular
points on certain toric surfaces. Geometric pictures similar to Figure 5 have also
appeared in recent work of McDuff on symplectic embeddings of 4-dimensional
ellipsoids (see [202]).
Ordinary Continued Fractions and the Supplementary Cone. To relate the above
theorem to toric geometry, we follow the approach of [231]. Given a cone =
Cone(e
2
, de
1
ke
2
) N
R
R
2
in normal form, its supplement is the cone

=
Cone(e
2
, de
1
ke
2
). Thus

is the right half-plane, and the cones ,

give
the convex hulls

= Conv( (N `0))

= Conv(

(N`0)).
Example 10.2.13. When d = 7, k = 5, Figure 7 on the next page shows the cones
,

and the convex hulls

. The open circles in the gure are the vertices


10.2. Continued Fractions and Toric Surfaces 479

7e
1
5e
2
e
2
e
1
Figure 7. The cones ,

, the convex hulls ,

, and their vertices


of

and

. Also observe that the fourth quadrant portion of Figure 7 becomes


Figure 5 after a 90

counterclockwise rotation.
For = Cone(e
2
, de
1
ke
2
), the ordinary continued fraction expansion d/k =
[a
1
, a
2
, . . . , a
s
] gives the sequences p
i
, q
i
, 0 i s, dened in Proposition 10.2.10.
Then dene the vectors
(10.2.15) u
o
1
=e
2
, u
o
i
= p
i
e
1
q
i
e
2
, 0 i s.
These vectors enable us to describe the vertices of

as follows.
Theorem 10.2.14. Let = Cone(e
2
, de
1
ke
2
) be a cone in normal form with
supplement

. Also let u
o
i
, 1 i s, be as dened in (10.2.15). Then:
(a) The set of vertices of

is
u
o
2j1
[ 1 j s/2u
o
s
.
(b) If a
1
= 1, then the set of vertices of

is
e
2
u
o
2j
[ 1 j s/2u
o
s
.
(c) If a
1
> 1, then the set of vertices of

is
e
2
u
o
2j
[ 0 j s/2u
o
s
.
Proof. First note that since u
o
1
=e
2
and u
o
0
= e
1
, we can rewrite (10.2.15) as
u
o
i
= p
i
e
1
q
i
e
2
= q
i
u
1
+ p
i
u
0
, 0 i s.
Thus we are in the situation of Theorem 10.2.12, where we have the cones
1
=
Cone(u
o
1
, u
o
s
) and
0
= Cone(u
o
0
, u
o
s
) and associated convex hulls
1
=

1
and
0
=

0
. Then Theorem 10.2.12 implies that the vectors u
o
1
, u
o
0
, . . . , u
o
s
give
the vertices of
1
and
0
.
480 Chapter 10. Toric Surfaces
However, = Cone(e
1
, e
2
)
0
and

=
1
. In particular,

=
1
, so
that part (a) of the theorem follows immediately. For parts (b) and (c), note that

= (

Cone(e
1
, e
2
))
0
.
The intersection

Cone(e
1
, e
2
) has vertices e
1
= u
o
0
, e
2
, while
0
has vertices
u
o
0
, u
o
2
, . . . , u
o
s
. If a
1
= 1, then e
2
, u
o
0
, u
o
2
are collinear, so that u
o
0
is not a vertex.
This proves part (b) of the theorem. Finally, if a
1
> 1, then one can prove without
difculty that u
o
0
is a vertex (Exercise 10.2.7), and part (c) follows.
Example 10.2.15. Figure 7 above illustrates Theorem 10.2.12 for the cone =
Cone(e
2
, 7e
1
5e
2
). Since 7/5 = [1, 2, 2], we use part (b) of the theorem.
Ordinary Continued Fractions and the Dual Cone. For a cone in normal form,
one surprise is that its supplement

is essentially the dual of .


Lemma 10.2.16. Given a cone = Cone(e
2
, de
1
ke
2
) N
R
in normal form, its
supplementary cone

N
R
is isomorphic to

M
R
.
Proof. Let e

1
, e

2
be the basis of M dual to the basis e
1
, e
2
of N. Then the isomor-
phism dened by e
1
e

2
and e
2
e

1
takes

= Cone(e
2
, de
1
ke
2
) to
Cone((e

1
), d(e

2
) k(e

1
)) = Cone(e

1
, ke

1
+de

2
) =

.
The isomorphism

from this lemma leads to some nice results about

and the associated convex hull

= Conv(

(M`0)). Specically, this


isomorphism takes the vectors
u
o
1
=e
2
, u
o
i
= p
i
e
1
q
i
e
2
, 0 i s
from (10.2.15) to the dual vectors
m
1
= e

1
, m
i
= q
i
e

1
+ p
i
e

2
, 0 i s.
In particular, m
s
= ke

1
+de

2
, so that

= Cone(m
1
, m
s
) in this notation. Then
Theorem 10.2.14 implies that the vertex set of the convex hull

is
(10.2.16) m
2j1
[ 1 j s/2m
s
.
Hence, in the language of 7.1,

M
R
is a full dimensional lattice polyhedron.
We can describe its recession cone and normal fan as follows.
Proposition 10.2.17. Let = Cone(e
2
, de
1
ke
2
) N
R
R
2
be in normal form
and set P =

= Conv(

(M`0)). Then:
(a) P M
R
is a full dimensional lattice polyhedron with recession cone

.
(b) The normal fan
P
of P is the renement of obtained by adding the minimal
generators
u
o
0
, u
o
2
, u
o
4
, . . . , u
o
s1
(s odd)
u
o
0
, u
o
2
, u
o
4
, . . . , u
o
s2
, u
o
s
u
o
s1
(s even).
10.2. Continued Fractions and Toric Surfaces 481
(c) The toric morphism X
P
U

is projective and X
P
is Gorenstein with at worst
rational double points.
Remark 10.2.18. When s is even, we can explain u
o
s
u
o
s1
as follows. By Theo-
rem 10.2.14, u
o
s2
and u
o
s
give an edge

. The vector determined by this edge is


u
o
s
u
o
s2
=a
s
u
o
s1
. Hence the edge has a
s
+1 lattice points since u
o
s1
is primitive.
Thus, if we start from u
o
s
, the next lattice point along the edge is
u
o
s
u
o
s1
.
Since d > k > 0 and d/k = [a
1
, . . . , a
s
] is computed using the Euclidean algorithm,
we have a
s
2 (Exercise 10.2.8). It follows that u
o
s
u
o
s1
is not a vertex of

.
Proof. Part (a) is straightforward (Exercise 10.2.9). For part (b), we will assume
that s is even and leave the case when s is odd to the reader (Exercise 10.2.9).
If we write s = 2, then the vertices (10.2.16) of P are m
1
, m
1
, . . . , m
21
, m
2
.
Thus the bounded edges of P =

are
m
1
m
1
, m
1
m
3
, . . . , m
23
m
21
, m
21
m
2
.
The inward-pointing normals of these edges give the rays that rene in the normal
fan of P.
The m
i
s satisfy the same recursion m
i
= a
i
m
i1
+m
i2
, 1 i s, as the u
o
i
s
in (10.2.14). Hence, for the edges m
2j1
m
2j+1
, 1 j s/2, we have
m
2j+1
m
2j1
= a
2j+1
m
2j
= a
2j+1
(q
2j
e

1
+ p
2j
e

2
).
Since u
o
2j
= p
2j
e
1
q
2j
e
2
, one easily computes that 'm
2j+1
m
2j1
, u
o
2j
` = 0. It
follows that u
o
2j
is the inward-pointing normal of this edge since it lies in and is
primitive. This takes care of all of the bounded edges except for m
21
m
2
. Here,
we compute
m
2
m
21
= (q
2
q
21
)e

1
+(p
2
p
21
)e

2
.
This is clearly normal to u
o
2
u
o
21
=(p
2
p
21
)e
1
(q
2
q
21
)e
2
. The latter
vector is easily seen to be primitive by part (b) of Proposition 10.2.10. Furthermore,
u
o
2
u
o
21
by Remark 10.2.18. Hence this is the inward-pointing normal of
the nal bounded edge m
21
m
2
.
For part (c), note that X
P
U

is projective by Theorem 7.1.10. To complete


the proof, we need to show that each maximal cone of
P
gives a Gorenstein afne
toric variety. For simplicity, we assume that s is odd (see Exercise 10.2.9 for the
even case). Since = Cone(e
2
, de
1
ke
2
) = Cone(e
2
, u
o
s
), the maximal cones of

P
consist of two boundary cones Cone(e
1
, u
o
0
) and Cone(u
o
s1
, u
o
s
), plus the
interior cones Cone(u
2j2
, u
2j
)
P
, 1 j s/2. The boundary cones are
easily seen to be smooth and hence Gorenstein (Exercise 10.2.9). For an interior
482 Chapter 10. Toric Surfaces
cone Cone(u
2j2
, u
2j
), we use part (b) of Proposition 10.2.10 to compute
'm
2j1
, u
o
2j2
` ='q
2j1
e

1
+ p
2j1
e

2
, p
2j2
e
1
q
2j2
e
2
`
=(p
2j1
q
2j2
p
2j2
q
2j1
) =(1)
2j1
= 1,
and a similar computation gives 'm
2j1
, u
o
2j
` = p
2j
q
2j1
p
2j1
q
2j
= (1)
2j
= 1.
By Proposition 8.2.12, we conclude that the corresponding afne toric variety is
Gorenstein, as desired. Then the singular points of X
P
are rational double points
by Example 10.1.5 and Proposition 10.1.6.
In 11.3 we will revisit this result, where we will learn that the morphism
X
P
U

from Proposition 10.2.17 is the blowup of the singular point of U

.
Just as the morphism X
P
U

is projective, one can show more generally


that the resolution of singularities X

from Theorem 10.2.3 is a projective


morphism. This requires nding a lattice polyhedron with the correct normal fan.
You will explore one way of doing this in Exercise 10.2.10.
We next consider the Hilbert basis H of

M. Recall from Lemma 1.3.10


that [H [ is the dimension of the Zariski tangent space at the singular point of U

and is the dimension of the most efcient embedding of U

into afne space.


Theorem 10.2.8 tells us that the Hilbert basis of N is computed using the
Hirzebruch-Jung continued fraction expansion of d/k. Since

has parameters
d, d k (part (b) of Exercise 10.1.2), it follows that we need the Hirzebruch-Jung
continued fraction expansion of d/(d k) to get the Hilbert basis of

M. By
Exercise 10.2.4, the ordinary continued fraction
d/k = [a
1
, . . . , a
s
]
gives the Hirzebruch-Jung continued fraction
d/(d k) =

[[(2)
a
1
1
, a
2
+2, (2)
a
3
1
, a
4
+2, . . . , (2)
a
s1
1
, a
s
+1]] s even
[[(2)
a
1
1
, a
2
+2, (2)
a
3
1
, a
4
+2, . . . , a
s1
+2, (2)
as1
]] s odd.
Theorem 10.2.8, applied to this expansion, gives the Hilbert basis of

M. To see
the underlying geometry, we need the following observation (Exercise 10.2.11):
(10.2.17)
In Theorem 10.2.5, three consecutive lattice points u
i1
, u
i
, u
i+1
are collinear if and only if u
i1
+u
i+1
= 2u
i
, i.e., b
i
= 2.
This means that a string of consecutive 2s in a Hirzebruch-Jung continued fraction
of d/(d k) gives a string of lattice points in the relative interior of a bounded
edge of the convex hull

. For the vertices m


1
, m
1
, . . . of

, this gives two


ways to think about lattice points on an edge connecting two adjacent vertices. For
example, the edge m
1
m
1
has a
1
1 lattice points in its relative interior because:
The Hirzebruch-Jung expansion of d/(dk) starts with a
1
1 consecutive 2s.
m
1
m
1
= a
1
u
0
, u
0
primitive, gives a
1
+1 lattice points on the edge.
This pattern continues for the other bounded edges of

.
10.2. Continued Fractions and Toric Surfaces 483
Oda gives a different argument for this pattern in [218, Sec. 1.6] and uses it to
relate the Hirzebruch-Jung continued fractions for d/k and d/(d k). His relation
also follows from our approach (see part (d) of Exercise 10.2.4). The connection
between continued fractions and toric surfaces is surprisingly rich and varied and
is one of the reasons why toric varieties are so much fun to study. See [75] and
[231] for a further discussion of this wonderful topic.
Exercises for 10.2.
10.2.1. Prove part (a) of Proposition 10.2.2: Show that the sequences P
i
and Q
i
from
(10.2.5) are increasing sequences of nonnegative numbers. Hint: Use b
i
2 for all i.
10.2.2. In 10.1, we constructed several resolutions of singularities in a rather ad hoc way.
In this exercise, we will see that the resolutions given by Theorem 10.2.3 are the same as
what we saw before.
(a) When has parameters d, 1, showthat the Hirzebruch-Jung continued fraction method
gives the same resolution of U

as the one given in Example 10.1.8.


(b) Do the same for Example 10.1.9. Hint: First show that
d
d 1
= [[2, 2, . . . , 2]],
where there are d 1 2s.
10.2.3. Verify the last claim in the proof of Proposition 10.2.6.
10.2.4. This exercise will consider some relations between ordinary continued fraction
expansions and Hirzebruch-Jung continued fraction expansions.
(a) Given integers a
1
, a
2
> 0 and a variable x, prove that
[a
1
, a
2
, x] = [[a
1
+1, (2)
a21
, x +1]],
where for any l 0, (2)
l
denotes a string of l 2s. Hint: Argue by induction on a
2
.
(b) Use part (a) to prove the equality [1, 1, 1, 5] = [[2, 3, 2, 2, 2, 2]] from Example 10.2.9.
(c) Given d/k = [a
1
, . . . , a
s
], prove that d/(d k) has the Hirzebruch-Jung expansion
given in the discussion leading up to (10.2.17). Hint: You will want to consider the
cases a
1
= 1 and a
1
> 1 separately, but the formula can be written as in (10.2.17) in
either case. If you get stuck, see [231].
(d) Starting fromthe ordinary continued fraction for d/k, use parts (a) and (c) to show that
if d/k =[[b
1
, . . . , b
r
]] and d/(d k) =[[c
1
, . . . , c
s
]], then (

r
i=1
b
i
)r =(

s
j=1
c
j
)s.
10.2.5. In this exercise we will consider Proposition 10.2.10.
(a) Prove the proposition. Hint: For part (a), argue by induction on the length s 1 of
the expansion. The expression [a
1
, a
2
, . . . , a
i
] is well-dened when the a
i
are positive
rational numbers, so we can write
[a
1
, a
2
, . . . , a
i1
, a
i
] = [a
1
, a
2
, . . . , a
i1
+1/a
i
].
Then use (10.2.12) and follow the reasoning from the proof of Proposition 10.2.2.
(b) Suppose we modify the initialization and the recurrences (10.2.12) as follows. Let
(r
1
, s
1
) = (1, 0) and (r
0
, s
0
) = (0, 1).
484 Chapter 10. Toric Surfaces
Then, for all 1 i s, compute
r
i
= r
i2
a
i
r
i1
s
i
= s
i2
a
i
s
i1
(note the change in sign!). What is true about r
i
d +s
i
k for all i? What do we get with
i = s? Hint: This fact is the basis for the extended Euclidean algorithm.
10.2.6. In this exercise, you will show that the ordinary continued fraction expansion of a
rational number d/k can be used to construct simultaneous resolutions of pairs of singular-
ities of certain toric surfaces. Let 1 < k < d be relatively prime integers, and let P be the
triangle Conv(0, de
1
, ke
2
) in R
2
.
(a) Draw the normal fan
P
and show that X
P
has exactly two singular points. Note that
X
P
is isomorphic to the weighted projective plane P(1, k, d).
(b) Adapt Theorem 10.2.12 to produce a resolution of singularities of X
P
from the or-
dinary continued fraction expansion of d/k. Hint: First rene
P
by introducing 1-
dimensional cones Cone(e
1
) and Cone(e
2
). Then apply Theorem 10.2.12 to the
third quadrant of your drawing.
10.2.7. Complete the proof of Theorem 10.2.14 by showing that u
o
0
is a vertex of

if
and only if a
1
> 1.
10.2.8. For relative prime integers d >k >0, we used the Euclidean algorithmto construct
d/k = [a
1
, . . . , a
s
]. Prove that a
s
2.
10.2.9. Prove part (a) of Proposition 10.2.17. Also prove part (b) for s odd and part (c) for
s even.
10.2.10. In this exercise, given a cone in normal form with parameters d, k, you will
see how to construct an unbounded polyhedron P in M
R
R
2
whose recession cone is

= Cone(e
1
, ke
1
+de
2
), and whose normal fan denes the resolution of U

from The-
orem 10.2.3. Let P
i
, Q
i
be the sequences constructed in Proposition 10.2.2. Let m be the
smallest positive integer such that md >P
1
+ +P
r1
and mk >Q
1
+ +Q
r1
. Starting
from the point A
r
= (mk, md) on the upper boundary ray of

, construct the points


A
r1
= (mk Q
r1
, md P
r1
)
A
r2
= (mk Q
r1
Q
r2
, md P
r1
P
r2
)
.
.
.
A
1
= (mk

r1
i=1
Q
i
, md

r1
i=1
P
i
)
A
0
= (mk

r1
i=1
Q
i
, 0).
Then let P be the polyhedron with one edge along the positive x-axis starting at A
0
, edges
A
i
A
i+1
for i = 0, . . . , r 1, and one edge along the upper edge of

starting from A
r
.
(a) Drawthe polyhedron P for the cone with parameters (d, k) =(7, 5). What is the integer
m in this case?
(b) Show that

is the recession cone of P.


(c) Show that the normal fan of P is the fan giving the resolution of U

constructed in
Theorem 10.2.3.
10.2.11. Prove (10.2.11).
10.3. Gr obner Fans and M
c
Kay Correspondences 485
10.2.12. Let d/k be a rational number in lowest terms with 0 < k < d, and let
d/k = [[b
1
, . . . , b
r
]] = [a
1
, . . . , a
s
]
be its continued fraction expansions. You will show that the sequences P
i
, Q
i
and p
i
, q
i
considered in this section can be expressed in matrix form.
(a) Let M

(b) =

b 1
1 0

. Show that for all 1 i r,

P
i
P
i1
Q
i
Q
i1

= M

(b
1
)M

(b
2
) M

(b
i
).
(b) Let M
+
(a) =

a 1
1 0

. Show that for all 1 i s,

p
i
p
i1
q
i
q
i1

= M
+
(a
1
)M
+
(a
2
) M
+
(a
i
).
10.2.13. In this exercise, you will apply the results of this section to the weighted projective
plane P(q
0
, q
1
, q
2
) from 2.0 and Example 3.1.17.
(a) Construct a resolution of singularities for any P(1, 1, q
2
), where q
2
2. What smooth
toric surface is obtained in this way? A complete classication of the smooth complete
toric surfaces will be developed in 10.4.
(b) Do the same for P(1, q
1
, q
2
) in general. Hint: Exercise 10.2.6.
10.3. Gr obner Fans and M
c
Kay Correspondences
The fans obtained by resolving the singularities of the afne toric toric surfaces U

have unexpected descriptions that involve Gr obner bases and representation theory.
In this section we will present these ideas, following [156] and [157].
Gr obner Bases and Gr obner Fans. We assume the reader knows about Gr obner
bases (see [69]) and Gr obner fans (see [70, Ch. 8, 4] or [264]). A nonzero ideal
I C[x
1
, . . . , x
n
] has a unique reduced Gr obner basis with respect to each mono-
mial order > on the polynomial ring. However, the set of distinct reduced marked
Gr obner bases for I (i.e., reduced Gr obner bases with marked leading terms in each
polynomial) is nite. Hence the ideal I has a nite universal Gr obner basis, i.e., a
nite subset U I that is a Gr obner basis for all monomial orders simultaneously.
Let w R
n
0
be a weight vector in the positive orthant (so w could be taken as
the rst row of a weight matrix dening a monomial order). Let
G =g
1
, . . . , g
t

be one of the reduced marked Gr obner bases for I, where


g
i
= x
(i)
+

c
i
x

,
and x
(i)
is marked as the leading term of g
i
. If w (i) > w whenever c
i
= 0,
then I will have Gr obner basis G with respect to any monomial order dened by a
486 Chapter 10. Toric Surfaces
weight matrix with rst row w. The set
(10.3.1) C
G
=w R
n
0
[ w (i) w whenever c
i
= 0
is the intersection of a nite collection of half-spaces, hence has the structure of a
closed convex polyhedral cone in R
n
0
. The cones C
G
as G runs over all distinct
marked Gr obner bases of I, together with all of their faces, have the structure of a
fan in R
n
0
called the Gr obner fan of I. In particular, for each pair G, G

of marked
Gr obner bases, the cones C
G
and C
G
intersect along a common face where the
w-weights of terms in some polynomials in G (and in G

) coincide.
A First Example. Let be a cone in normal form with parameters d, k, and recall
gcd(d, k) = 1 by hypothesis. By Proposition 10.1.2, the group
G
d,k
=(,
k
) (C

)
2
[
d
= 1, 0 k d 1
d
acts on C
2
by componentwise multiplication
(10.3.2) (,
k
) (x, y) = (x,
k
y),
with quotient C
2
/G
d,k
U

.
Let I(G
d,k
) be the ideal dening G
d,k
as a variety in C
2
. In the next extended
example, we will introduce the rst main result of this section.
Example 10.3.1. Let d = 7, k = 5 and I = I(G
7,5
). It is easy to check that
I ='x
7
1, y x
5
`.
Moreover, for lexicographic order with y > x, the set
G
(1)
=x
7
1, y x
5

is the reduced marked Gr obner basis for I, where the underlines indicate the leading
terms. The corresponding cone in the Gr obner fan of I is
C
G
(1) =w = (a, b) R
2
0
[ b 5a = Cone(e
2
, e
1
+5e
2
).
There are three other marked reduced Gr obner bases of I:
G
(2)
=x
5
y, x
2
y 1, y
2
x
3
,
G
(3)
=x
3
y
2
, x
2
y 1, y
3
x,
G
(4)
=y
7
1, x y
3
.
It is easy to check that each of these sets is a Gr obner basis for I using Buchbergers
criterion. The corresponding cones are
C
G
(2) =(a, b) R
2
0
[ b 5a, 2b 3a = Cone(e
1
+5e
2
, 2e
1
+3e
2
),
C
G
(3) =(a, b) R
2
0
[ 2b 3a, 3b a = Cone(2e
1
+3e
2
, 3e
1
+e
2
),
C
G
(4) =(a, b) R
2
0
[ 3b a = Cone(3e
1
+e
2
, e
1
).
10.3. Gr obner Fans and M
c
Kay Correspondences 487
Since these three cones ll out rest of the rst quadrant in R
2
, the Gr obner fan of
I consists of the four cones C
G
(i) and their faces, as shown in Figure 8. We will
denote this fan by in the following.
(3,1)
(2,3)
(1,5)
Figure 8. The Gr obner fan
Next, let us consider the resolution of singularities
X

for the cone with parameters d = 7, k = 5 computed in Example 10.2.4 in the


last section. The reader can check that the linear transformation T : N
R
N
R
with
matrix relative to the basis e
1
, e
2
given by
(10.3.3) A =

7 0
5 1

maps the cones C


G
(i) in the Gr obner fan to the corresponding cones
i
in the fan
. The matrix in (10.3.3) is invertible, but its inverse is not an integer matrix. The
image of the lattice N = Z
2
under T is the proper sublattice 7Ze
1
Ze
2
, and T
1
maps N to the lattice
N

=(a/7, b/7) [ a, b Z, b 5a mod 7 = N +Z

1
7
e
1
+
5
7
e
2

.
There is an exact sequence
0 N N

G 0
induced by the map N

(C

)
2
dened by (a/7, b/7) (e
2ia/7
, e
2ib/7
). Letting
= Cone(e
1
, e
2
), the corresponding toric morphism U
,N
U
,N

is the quotient
mapping C
2
C
2
/G.
It is easy to check that w
0
= (0, 1) and w
1
= (1/7, 5/7) form a basis of the
lattice N

. In Figure 9 on the next page, the fan dened by the cones with ray
488 Chapter 10. Toric Surfaces
(1,0) = w
4
= 7w
1
- 5w
0
(3/7,1/7) = w
3
= 3w
1
- 2w
0
(2/7,3/7) = w
2
= 2w
1
- w
0
(1/7,5/7) = w
1
(0,1) = w
0
u
0
u
1
u
2
u
3
u
4

Figure 9. The toric variety X


,N
is the resolution of U in Example 10.3.1
generators on the left is the same as in Figure 8 above, and the fan on the right is
the same as in Figure 3 above. Note that T(w
i
) = u
i
for 0 i 4 in Figure 9.
With respect to N

, the cones in the Gr obner fan are smooth cones, and it


follows from the discussion of toric morphisms in 3.3 that the toric surfaces X
,N

and X

are isomorphic. In other words, the Gr obner fan of the ideal I encodes
the structure of the resolution of singularities of U

.
Example B.2.4 shows how to compute this example using GFan [161].
A Tale of Two Fans. We next show that the last observation in Example 10.3.1
holds in general. As in the example, consider the ideal I = I(G
d,k
) and the action
of G
d,k
given in (10.3.2). Each monomial in C[x, y] is equivalent modulo I to one
of the monomials x
j
, j = 0, . . . , d 1. This may be seen, for instance, from the
remainders on division by the lexicographic Gr obner basis x
d
1, y x
k
. As
a result, we have a direct sum decomposition of the coordinate ring of the variety
G
d,k
as a C-vector space:
(10.3.4) C[x, y]/I
d1

j=0
V
j
,
where V
j
is the 1-dimensional subspace spanned by x
j
mod I.
The following result establishes a rst connection between the ideal I and the
resolution of singularities of U

described in Theorem 10.2.3.


Proposition 10.3.2. Let I = I(G
d,k
) where 0 < k < d and gcd(d, k) = 1. Consider
u
0
= e
2
u
i
= P
i1
e
1
Q
i1
e
2
, i = 1, . . . , r +1,
from Theorem 10.2.3. Let T : N
R
N
R
be the linear transformation with matrix
A =

d 0
k 1

10.3. Gr obner Fans and M


c
Kay Correspondences 489
and let w
i
= T
1
(u
i
) for i = 0, . . . , r +1. Write w
i
=
1
d
(a
i
e
1
+b
i
e
2
) and dene
g
i
= x
b
i
y
a
i
, 0 i r +1.
(a) The polynomials g
0
, . . . , g
r+1
are contained in the ideal I.
(b) S =ae
1
+be
2
N [ a, b 0, x
b
y
a
I N is an additive semigroup.
(c) dw
i
[ i = 0, . . . , r +1 is the Hilbert basis of the semigroup S of part (b).
Proof. It is an easy calculation to show T
1
maps = Cone(e
2
, de
1
ke
2
) to the
rst quadrant R
2
0
. Moreover, w
0
= e
2
, and for i = 1, . . . , r +1,
w
i
=
1
d

P
i1
e
1
+(kP
i1
dQ
i1
)e
2

.
Therefore, g
0
= x
d
1 and
g
i
= x
kP
i1
dQ
i1
y
P
i1
for i =1, . . . , r +1. Since
d
=1, these polynomials clearly vanish at (,
k
) G
d,k
.
Therefore g
i
I(G
d,k
) = I for all i.
The proof of part (b) is left to the reader as Exercise 10.3.3. For part (c), it
follows from parts (a) and (b) that dw
i
is contained in S. On the other hand, let
ae
1
+be
2
S. Then x
b
y
a
I, which implies that b ak mod d. Hence
T

1
d
(ae
1
+be
2
)

= ae
1
+
bak
d
e
2
must be an element of N. Since the u
i
are the Hilbert basis for the semigroup
N by Theorem 10.2.8, this vector is a nonnegative integer combination of the
u
i
. Hence ae
1
+be
2
is a nonnegative integer combination of the dw
i
. It follows that
S is generated by the dw
i
. The dw
i
are irreducible in S because the corresponding
u
i
= T(w
i
) are irreducible in the semigroup N.
We also have a rst result about reduced Gr obner bases of the ideal I.
Lemma 10.3.3. Every element of a reduced Gr obner basis of I = I(G
d,k
) is either
of the form x
b
y
a
or of the form x
s
y
t
1 for s, t > 0.
Proof. Since I is generated by x
d
1, y x
k
, the Buchberger algorithm implies
that a reduced Gr obner basis G of I consists of binomials. By taking out common
factors, every g G can be written g = x
i
y
j
h, where h = x
b
y
a
or x
s
y
t
1. Then
h I since it vanishes on G
d,k
. Its leading term divisible by the leading term of an
element of G, which is impossible in a reduced Gr obner basis unless g = h.
The Gr obner bases in Example 10.3.1 give a nice illustration of Lemma 10.3.3.
Our next lemma relates the polynomials g
i
= x
b
i
y
a
i
from Proposition 10.3.2 to
the reduced Gr obner bases of I.
Lemma 10.3.4. Let g
i
= x
b
i
y
a
i
be as in Proposition 10.3.2 and x a monomial
order > on C[x, y].
490 Chapter 10. Toric Surfaces
(a) The a
i
are increasing and the b
i
are decreasing with i.
(b) There is some index i = i
0
(depending on >) such that
LT
>
(g
i
) = x
b
i
for all i i
0
, and LT
>
(g
i
) = y
a
i
for all i > i
0
.
(c) If i =i
0
is the index from part (b), then g
i
0
and g
i
0
+1
are elements of the reduced
Gr obner basis of I = I(G
d,k
) with respect to >.
Proof. You will prove parts (a) and (b) in Exercise 10.3.4. For part (c), let G
be the reduced Gr obner basis of I with respect to >. Since g
i
0
I by part (a) of
Proposition 10.3.2, there is g G whose leading term divides LT
>
(g
i
0
) = x
b
i
0
. By
Lemma 10.3.3, it follows that g = x
b
y
a
with LT
>
(g) = x
b
. In particular, b b
i
0
and ae
1
+be
2
S, where S is the semigroup from part (b) of Proposition 10.3.2.
Then part (c) of the same proposition implies that ae
1
+be
2
must be a nonnegative
integer combination
(10.3.5) ae
1
+be
2
=

r+1
i=0

i
dw
i
=

r+1
i=0

i
(a
i
e
1
+b
i
e
2
),
i
N.
Since b b
i
0
and the b
i
decrease with i, (10.3.5) can include only the dw
i
with
i i
0
. Suppose that dw
i
appears in (10.3.5) with i > i
0
. Then a a
i
and y
a
i
> x
b
i
,
so that LT
>
(g
i
) = y
a
i
divides y
a
. Since g
i
I, LT
>
(g
i
) is divisible by the leading
term of some h G. Hence LT
>
(h) divides y
a
, which is a term of g = x
b
y
a
G.
This is impossible in a reduced Gr obner basis, hence i >i
0
cannot occur in (10.3.5).
From here, it follows easily that g = g
i
0
, giving g
i
0
G as desired.
The statement for g
i
0
+1
follows by an argument parallel to the one above. The
details are left to the reader (Exercise 10.3.4).
We are now ready for the rst major result of this section.
Theorem 10.3.5. Let be the fan in R
2
with maximal cones

i
= Cone(a
i1
e
1
+b
i1
e
2
, a
i
e
1
+b
i
e
2
), 1 i r +1,
for a
i
, b
i
as in Proposition 10.3.2. Then is the Gr obner fan of the ideal I(G
d,k
).
Proof. The cones
i
ll out the rst quadrant R
2
0
. Take any w = ae
1
+be
2
lying
in the interior of some
i
and let > be a monomial order dened by a weight matrix
with w as rst row. This gives a reduced Gr obner basis G. The theorem will follow
once we prove that
i
is the Gr obner cone C
G
of G.
First observe that for >, we must have must have i
0
= i 1 in part (b) of
Lemma 10.3.4. It follows from part (c) of the lemma that g
i1
and g
i
are elements
of G. Since a reduced Gr obner basis has only one element with leading term a
power of x and only one with leading term a power of y, all other elements of G
will have the form x
s
y
t
1, s, t > 0, by Lemma 10.3.3. Therefore, the Gr obner
cone C
G
is exactly
i
and we are done.
As in Example 10.3.1, the following statement is an immediate consequence.
10.3. Gr obner Fans and M
c
Kay Correspondences 491
Corollary 10.3.6. Let N

be the lattice
N

=(a/d, b/d) [ a, b Z, b k mod d =Z

1
d
e
1
+
k
d
e
2

Ze
2
.
The toric surfaces X

and X
,N

are isomorphic.
In other words, the Gr obner fan of I(G
d,k
) can be used to construct a resolution
of singularities of the afne toric surface U

when has parameters d, k.


Connections with Representation Theory. We now consider the above results
from a different point of view. We assume the reader is familiar with the beginnings
of representation theory for nite abelian groups.
The group G = G
d,k
from (10.3.2) acts on V =C
2
by the 2-dimensional linear
representation of the group
d
of dth roots of unity dened by
(10.3.6)
:
d
GL(V) = GL(2, C)

0
0
k

.
Since
d
is abelian, its irreducible representations are 1-dimensional over C, and
hence each is dened by a character

j
:
d
C


j
for j = 0, . . . , d 1. The reason for the minus sign will soon become clear.
Via (10.3.2), we get the induced action of
d
on the polynomial ring C[x, y] by
x =
1
x, y =
j
y,
as explained in 5.0. Each monomial x
a
y
b
spans an invariant subspace where
the action of
d
is given by the irreducible representation with character
j
for
j a +kb mod d. We call a +kb mod d the weight of the monomial x
a
y
b
with
respect to this action of
d
. Since the ideal of the group G(C

)
2
is invariant, the
action descends to the quotient C[x, y]/I(G), and we have a representation of
d
on C[x, y]/I(G). The direct sum decomposition (10.3.4) shows that the irreducible
representation with character
j
appears exactly once in this representation, as
the subspace V
j
in (10.3.4). This means that the representation on C[x, y]/I(G) is
isomorphic to the regular representation of
d
(Exercise 10.3.5).
A 2-dimensional M
c
Kay Correspondence. In 1979, M
c
Kay pointed out that there
is a one-to-one correspondence between the irreducible representations of
d
and
the components of the exceptional divisor in the resolution : X

when
was a cone in normal form with parameters k =d 1, as in Example 10.1.5. In this
case, the singular point of U

is a rational double point, and the image of the rep-


resentation from (10.3.6) lies in SL(2, C). A great deal of research was devoted
to explaining the original M
c
Kay correspondence in representation-theoretic and
492 Chapter 10. Toric Surfaces
geometric terms (work of Gonzalez-Sprinberg, Artin, and Verdier). However, for
1 < k < d 1, (
d
) is a subgroup of GL(2, C), not SL(2, C), and there are more
irreducible representations of
d
than components of the exceptional divisor. The
M
c
Kay correspondence can be extended to these cases by identifying certain spe-
cial representations that correspond to the components of the exceptional divisor
(work of Wunram, Esnault, Ito and Nakamura, Kidoh, and others).
We will describe a generalized M
c
Kay correspondence that applies for all d, k.
Writing G= G
d,k
as before, consider the ring of invariants C[x, y]
G
. You will prove
the following in Exercise 10.3.6.
Lemma 10.3.7. Let V
j
be an irreducible representation of
d
with character
j
,
and consider the action of G = G
d,k

d
on C[x, y]
C
V
j
. Then the subspace of
invariants (C[x, y]
C
V
j
)
G
has the structure of a module over the ring C[x, y]
G
.
Example 10.3.8. Let G = G
7,5
as in Example 10.3.1. The ring of invariants is
C[x, y]
G
=C[x
7
, x
2
y, xy
4
, y
7
] in this case. If v
j
is the basis of the representation V
j
,
then it is easy to check that x
a
y
b
v
j
is invariant under G if and only if a+kb j
0 mod 7, or in other words if and only if x
a
y
b
has weight j under this action of
7
.
First consider the case j = 1 in Lemma 10.3.7. The monomials in the comple-
ment of the monomial ideal
M ='x
7
, x
2
y, xy
4
, y
7
`
that have weight 1 are x and y
3
. Then x v
1
and y
3
v
1
generate the module
(C[x, y] V
1
)
G
. Since x and y
3
have the same weight with respect to this action
of
7
, the difference x y
3
is an element of the ideal I(G), and this is one of the
polynomials g
i
as in the proof of Proposition 10.3.2.
On the other hand, if j = 2, then there are three monomials with weight 2 in
the complement of M, and these give three generators of (C[x, y] V
2
)
G
, namely
x
2
v
2
, xy
3
v
2
, and y
6
v
2
. It is still true that x
2
xy
3
, x
2
y
6
, and xy
3
y
6
are
elements of I(G), but these polynomials cannot appear in a reduced Gr obner basis
for I(G). Moreover, no proper subset of the three generators generates the whole
module (C[x, y] V
2
)
G
.
Denition 10.3.9. Let G = G
d,k

d
as above. We say that the representation V
j
is special with respect to k if (C[x, y] V
j
)
G
is minimally generated as a module
over the invariant ring C[x, y]
G
by two elements.
Hence, in Example 10.3.8, V
1
is special while V
2
is not. According to our
denition, the trivial representation V
0
is never special, since (C[x, y] V
0
)
G
is
generated by the single monomial 1 over the invariant ring. Our next theorem
gives a rudimentary form of a M
c
Kay correspondence for the group G
d,k
.
10.3. Gr obner Fans and M
c
Kay Correspondences 493
Theorem 10.3.10 (M
c
Kay Correspondence). Let be a cone with parameters
d, k, where 0 < k < d and gcd(d, k) = 1. Then there is a one-to-one correspon-
dence between the representations of
d
that are special with respect to k and the
components of the exceptional divisor for the minimal resolution : X

.
Proof. Write G = G
d,k
as above and consider the set B of monomials in the com-
plement of the ideal M generated by the G-invariant monomials. This set contains
L =1, x, x
2
, . . . , x
d1
, y, y
2
, . . . , y
d1
.
Since gcd(d, k) = 1, for each 1 j d 1, there is an integer 1 a
j
d 1 such
that x
j
and y
a
j
have equal weight (equal to j) for the action of G. The representa-
tion V
j
is special with respect to k if and only if these are the only two monomials
of weight j in the set B, and nonspecial if and only if there is some monomial
x
a
y
b
with a, b > 0 in B which also has weight j. Since x
j
y
a
j
I(G), saying
V
j
is special with respect to k is in turn equivalent to saying that the corresponding
vector a
j
e
1
+ je
2
is an irreducible element in the semigroup from part (b) of Propo-
sition 10.3.2 (Exercise 10.3.8). By Theorem 10.3.5, Cone(a
j
e
1
+ je
2
) is one of the
1-dimensional cones of the Gr obner fan of I(G). Then V
j
corresponds to one of the
1-dimensional cones in the fan and hence to one of the irreducible components
of the exceptional divisor.
The original M
c
Kay correspondence is the following special case.
Corollary 10.3.11. When k = d 1, there is a one-to-one correspondence be-
tween the set of all irreducible representations of
d
and the components of the
exceptional divisor of the minimal resolution : X

.
Proof. In this case, the invariant ring is C[x
d
, xy, y
d
], so the sets L and B in the
proof of the theorem coincide.
There has also been much work devoted to extend the M
c
Kay correspondence
to nite abelian subgroups G GL(n, C) for n 3, and several other ways to
understand these constructions have also been developed, including the theory of
G-Hilbert schemes. See Exercise 10.3.10 for the beginnings of this.
Exercises for 10.3.
10.3.1. In this exercise, you will verify the claims made in Example 10.3.1, and extend
some of the observations there.
(a) Show that each of the G
(i)
is a Gr obner basis of I(G
7,5
).
(b) Show that G =x
5
1, y x
3
, x
2
y 1, x y
2
, y
5
1 is a universal Gr obner basis for
I(G
7,5
).
(c) Determine the cones C
G
(i) using (10.3.1).
(d) Verify the nal claim that linear transformation dened by the matrix A from (10.3.3)
maps the Gr obner cones C
G
(i) to the
i
for i = 1, 2, 3.
494 Chapter 10. Toric Surfaces
10.3.2. Verify the conclusions of Proposition 10.3.2 and Theorem 10.3.5 for the case d =
17, k = 11.
10.3.3. Show that the set S dened in part (b) of Proposition 10.3.2 is an additive semi-
group. Hint: A direct proof starts from two general elements ae
1
+be
2
and a

e
1
+b

e
2
in
S. Consider (x
b
y
a
)(x
b

+y
a

) and (x
b
+y
a
)(x
b

y
a

).
10.3.4. In this exercise you will complete the proof of Lemma 10.3.4.
(a) Show that for each i, kP
i
dQ
i
= k
i
, where the k
i
are produced by the modied Eu-
clidean algorithm from (10.2.1).
(b) Prove part (a) of Lemma 10.3.4.
(c) Verify that there is an index i
0
as in part (b) of Lemma 10.3.4.
(d) Verify that if i
0
is as in part (c), then g
i0+1
is contained in the reduced Gr obner basis.
10.3.5. If G is any nite group, the (left) regular representation of G is dened as follows.
Let W be a vector space over C of dimension [G[ with a basis e
h
[ h G indexed by the
elements of G. For each g G let (g) : W W be dened by (g)(e
h
) = e
gh
.
(a) Show that g (g) is a group homomorphismfrom G to GL(W).
(b) Now let G be the cyclic group
d
of order d. Show that W is the direct sum of 1-
dimensional invariant subspaces W
j
, j = 0, . . . , d 1 on which G acts by the character

j
dened in the text, so that W decomposes as W
d1
j=0
V
j
.
10.3.6. In this exercise you will consider the module structures from Lemma 10.3.7.
(a) Prove Lemma 10.3.7.
(b) Verify the claims made in Example 10.3.8.
10.3.7. In this exercise you will prove an alternate characterization of the special represen-
tations with respect to k from Denition 10.3.9. We write G = G
d,k

d
as usual.
(a) Let
2
C
2
= f dx dy [ f C[x, y]. Show that x
a
y
b
dx dy =
a+b+1+k
x
a
y
b
dx dy
denes an action of G on
2
C
2
.
(b) Show that the spaces of G-invariants (
2
C
2
)
G
and (
2
C
2
V
j
)
G
have the structure of
modules over the invariant ring C[x, y]
G
.
(c) Show that V
j
is special with respect to k if and only if the multiplication map
(
2
C
2 )
G
(C[x, y] V
j
)
G
(
2
C
2 V
j
)
G
is surjective.
10.3.8. In the proof of the M
c
Kay correspondence, show that a
j
e
1
+ je
2
is irreducible in
the semigroup S from Proposition 10.3.2 if and only if the representation V
j
is special with
respect to k.
10.3.9. Let g
i
, 0 i r +1, be the binomials constructed in Proposition 10.3.2. Show that
U =g
1
, . . . , g
r
x
a
y
b
1 [ x
a
y
b
is G
d,k
-invariant
is a universal Gr obner basis for I(G
d,k
) (not always minimal, however).
10.3.10. Let G =G
d,k
act on C
2
as in (10.3.2). As a point set, G can be viewed as the orbit
of the point (1, 1) under this action. The ideal I = I(G) is invariant under the action of G
on C[x, y] and as we have seen, the corresponding representation on C[x, y]/I is isomorphic
to the regular representation of G.
10.4. Smooth Toric Surfaces 495
(a) Showthat if p =(, ) is any point in C
2
other than the origin, the ideal I of the orbit of
p is another G-invariant ideal and the corresponding representation of G on C[x, y]/I
is also isomorphic to the regular representation of G.
(b) The G-Hilbert scheme can be dened as the set of all G-invariant ideals in C[x, y] such
that the representation of G on C[x, y]/I is isomorphic to the regular representation of
G. Show that every such ideal has a set of generators of the form
x
a
y
c
, y
b
x
d
, x
ad
y
bc

for some , C and where x


a
and y
c
(resp. y
b
and x
d
) have equal weights for the
action of G. It can be seen fromthis result that the G-Hilbert scheme is also isomorphic
to the minimal resolution of singularities of U

.
10.4. Smooth Toric Surfaces
This section will use 10.1 and 10.2 to classify smooth complete toric surfaces
and study the relation between continued fractions and intersection products of
divisors on the resulting resolutions of singularities.
Classication of Smooth Toric Surfaces. We will show that smooth complete
toric surfaces are all obtained by toric blowups from either P
2
, P
1
P
1
, or one
of the Hirzebruch surfaces H
r
with r 2 from Example 3.1.16. The proof will be
based on the following facts.
First, Proposition 3.3.15 implies that if = Cone(u
1
, u
2
) is a smooth cone and
we rene by inserting the new 1-dimensional cone = Cone(u
1
+u
2
), then on
the resulting toric surface, the smooth point p

is blown up to a copy of P
1
.
For the second ingredient of the proof, we introduce the following notation. If
is a smooth complete fan, then list the ray generators of the 2-dimensional cones
in as u
0
, u
1
, . . . , u
r1
in clockwise order around the origin in N
R
, and we will
consider the indices as integers modulo r, so u
r
= u
0
. Then we have the following
statement parallel to (10.2.8).
Lemma 10.4.1. Let u
0
, . . . , u
r
be the ray generators for a smooth complete fan
in N
R
R
2
. There exist integers b
i
, i = 0, . . . , r 1, such that
(10.4.1) u
i1
+u
i+1
= b
i
u
i
.
Proof. This is a special case of the wall relation (6.4.4).
We also have the following result.
Lemma 10.4.2. Let be a smooth fan that renes a smooth cone . Then is
obtained from by a sequence of star subdivisions as in Denition 3.3.13.
Proof. Suppose has r cones of dimension 2, with ray generators u
0
, . . . , u
r
, listed
clockwise starting from u
0
. We argue by induction on r = [(2)[. If r = 1, there
is only one cone in and there is nothing to prove. Assume the result has been
496 Chapter 10. Toric Surfaces
proved for all with [(2)[ = r, and consider a fan with [(2)[ = r +1. There
are r interior rays that by (6.4.4) give wall relations u
i1
+u
i+1
= b
i
u
i
, b
i
Z,
for 1 i r. Note that is strongly convex so b
i
> 0 for all i. We claim that there
exists some i such that b
i
= 1. If not, i.e., if b
i
2 for all i, then as in 10.2, the
Hirzebruch-Jung continued fraction
[[b
1
, b
2
, . . . , b
r
]]
represents a rational number d/k and the cone has parameters d, k with d >k >0.
But then d 2, which would contradict the assumption that is a smooth cone.
Hence there exists an i, 1 i r, such that
u
i1
+u
i+1
= u
i
.
In this situation, Cone(u
i1
, u
i+1
) is also smooth (Exercise 10.4.1). Moreover,
Cone(u
i1
, u
i
) and Cone(u
i
, u
i+1
) are precisely the cones in the star subdivision
of Cone(u
i1
, u
i+1
). Then we are done by induction.
We are now ready to state our classication theorem.
Theorem 10.4.3. Every smooth complete toric surface X

is obtained from either


P
2
, P
1
P
1
, or H
r
, r 2
by a nite sequence of blowups at xed points of the torus action.
Proof. We follow the notation of Lemma 10.4.1. As in the proof of Lemma 10.4.2,
if b
i
= 1 in (10.4.1) for some i, then our surface is a blowup of the smooth surface
corresponding to the fan where u
i
is removed. Hence we only need to consider the
case in (10.4.1) where b
i
= 1 for all i.
Suppose rst that u
j
= u
i
for some i < j. We relabel the vertices to make
u
j
= u
1
for some j. Note that j > 2 since the cones must be strongly convex.
Then from (10.4.1),
u
0
=u
2
+b
1
u
1
.
Using the basis u
1
, u
2
of N, we get the picture shown in Figure 10 on the next page.
Comparing this with the fans for P
2
, P
1
P
1
and H
a
(Figures 2, 3 and 4 from
3.1), we see that is a renement of the fan of H
r
if r = b
1
> 2. The same
follows if b
1
<2 and r =[b
1
[ (see Exercise 10.4.2). Since b
1
= 1, the remaining
possibilities are b
1
= 0 or 1, where we get a renement of the fan of P
1
P
1
or
P
2
respectively. Then the theorem follows from Lemma 10.4.2 in this case.
We will complete the proof using primitive collections (Denition 5.1.5). The
rst step is to note that X

is projective by Proposition 6.3.25. This will allow


us to use Proposition 7.3.6, which asserts that has a primitive collection whose
minimal generators sum to 0.
If (1) has only three elements, it is easy to see that X

= P
2
since is
smooth and complete (Exercise 10.4.3). If [(1)[ > 3, every primitive collection
10.4. Smooth Toric Surfaces 497
u
0
u
1
u
2
u
j
Figure 10. The ray generators u0, u1, u2, uj when uj =u1
of has exactly two elements (Exercise 10.4.3). By Proposition 7.3.6, one of these
primitive collections must have minimal generators u
i
, u
j
that satisfy u
i
+u
j
= 0.
Hence u
j
=u
i
, and we are done by the earlier part of the proof.
Since there is also a Hirzebruch surface H
1
, the statement of this theorem
might seem puzzling. The reason that H
1
is not included is that this surface is
actually a blowup of P
2
(Exercise 10.4.4).
The problem of classifying smooth compete toric varieties of higher dimension
is much more difcult. We did this when rankPic(X

) = 2 in Theorem 7.3.7. See


[14] for the case when rankPic(X

) = 3.
Intersection Products on Smooth Surfaces. A fundamental feature of the theory
of smooth surfaces is the intersection product on divisors. In 6.3, we dened DC
when D is a Cartier divisor and C is a complete irreducible curve. On a smooth
complete surface, this means that the intersection product D C is dened for all
divisors D and C. In particular, taking D =C gives the self-intersection D D = D
2
.
Here is a useful result about intersection numbers on a smooth toric surface.
Theorem 10.4.4. Let D

be the divisor on a smooth toric surface X

corresponding
to = Cone(u) which is the intersection of 2-dimensional cones Cone(u, u
1
) and
Cone(u, u
2
) in . Then:
(a) D

=b, where u
1
+u
2
= bu as in (10.4.1).
(b) For a divisor D

= D

, we have
D

= Cone(u
i
), i = 1, 2
0 otherwise.
Proof. Since X

is smooth, part (a) follows from Lemma 6.4.4 once you compare
(10.4.1) to (6.4.4). Part (b) follows from Corollary 6.4.3 and Lemma 6.4.4.
498 Chapter 10. Toric Surfaces
Example 10.4.5. Let = Cone(u
1
, u
2
) be a cone in a smooth fan , and consider
the star subdivision, in which = Cone(u
1
+u
2
) is inserted to subdivide into two
cones. Call the rened fan

. Then the exceptional divisor E = D

of the blowup
: X

satises E E =1 on X

.
Complete curves with self-intersection number 1 on a smooth surface are
called exceptional curves of the rst kind. They can always be contracted to a
smooth point on a birationally equivalent surface, as in the above example.
One of the foundational results in the theory of general algebraic surfaces is
that every smooth complete surface S has at least one relatively minimal model.
This means that there is a birational morphism S S, where S is a smooth surface
with the property that if : S S

is a birational morphism to another smooth


surface S

, then is necessarily an isomorphism. This is proved in [131, V.5.8].


Interestingly, the possible relatively minimal models for rational surfaces are pre-
cisely the surfaces P
2
, P
1
P
1
, and H
r
, r 2, from Theorem 10.4.3.
On a smooth complete toric surface X

, the intersection product can be re-


garded as a Z-valued symmetric bilinear form on Pic(X

). Here is an example.
Example 10.4.6. Consider the Hirzebruch surface H
r
. Using the fan shown in
Figure 3 of Example 4.1.8, we get divisors D
1
, . . . , D
4
corresponding to minimal
generators u
1
= e
1
+re
2
, u
2
= e
2
, u
3
= e
1
, u
4
= e
2
. By Theorem 10.4.4, we
have the self-intersections
D
1
D
1
= D
3
D
3
= 0, D
2
D
2
=r, D
4
D
4
= r.
The Picard group Pic(H
r
) is generated by the classes of D
3
and D
4
. Note also that
D
3
D
4
= D
4
D
3
= 1
by Theorem 10.4.4. The intersection product is described by the matrix

D
3
D
3
D
3
D
4
D
4
D
3
D
4
D
4

0 1
1 r

.
If D aD
3
+bD
4
and E cD
3
+dD
4
are any two divisors on the surface, then
(10.4.2) D E =

a b

0 1
1 r

c
d

= bc +ad +rbd.
For instance, with D = E = D
2
rD
3
+D
4
, we obtain
D
2
D
2
= 1 (r) +(r) 1+r 1 1 =r.
The self-intersection numbers D
1
D
1
= D
3
D
3
= 0 reect the bration structure
on H
r
studied in Example 3.3.20. The divisors D
1
and D
3
are bers of the map-
ping H
r
P
1
. Such curves always have self-intersection equal to zero. You will
compute several other intersection products on H
r
in Exercise 10.4.5.
10.4. Smooth Toric Surfaces 499
Resolution of Singularities Reconsidered. Another interesting class of smooth
toric surfaces consists of those that arise from a resolution of singularities of the
afne toric surface U

of a 2-dimensional cone . Here is a simple example.


Example 10.4.7. Let = Cone(e
2
, de
1
e
2
) have parameters d, 1 where d > 1.
The resolution of singularities X

constructed in Example 10.1.8 uses the


smooth renement of obtained by adding Cone(e
1
). This gives the exceptional
divisor E on X

. Since
e
2
+(de
1
e
2
) = de
1
,
we see that
E E =d
is the self-intersection number of E.
More generally, suppose that the smooth toric surface X

is obtained via a
resolution of singularities of U

, where the 2-dimensional cone has parameters


d, k with d > 1. Let the Hirzebruch-Jung continued fraction expansion of d/k be
d/k = [[b
1
, b
2
, . . . , b
r
]].
Recall from Theorem 10.2.3 that is obtained from = Cone(u
0
, u
r+1
) by adding
rays generated by u
1
, . . . , u
r
, and by Theorem 10.2.5, we have
u
i1
+u
i+1
= b
i
u
i
, 1 i r.
It follows that D
1
, . . . , D
r
are complete curves in X

. They are the irreducible


components of the exceptional ber, with self-intersections
D
i
D
i
=b
i
, 1 i r,
by Theorem 10.4.4. Then the intersection matrix (D
i
D
j
)
1i, jr
is given by
(10.4.3) D
i
D
j
=

b
i
if j = i
1 if [i j[ = 1
0 otherwise.
In Exercise 10.4.6 you will show that the associated quadratic form is negative
denite. This condition is necessary for the contractibility of a complete curve C
on a smooth surface S, i.e., the existence of a proper birational morphism : S S,
where (C) is a (possibly singular) point on S.
The resolutions described here have another important property.
Denition 10.4.8. A resolution of singularities : Y X is minimal if for every
resolution of singularities : Z X, there is a morphism : Z Y such that
Y

X
500 Chapter 10. Toric Surfaces
is a commutative diagram, i.e., = .
It is easy to see that if a minimal resolution of X exists, then it is unique up to
isomorphism. If X has a unique singular point p, then using the theory of birational
morphisms of surfaces it is not difcult to show that a resolution of singularities
: Y X is minimal if the exceptional ber contains no irreducible components E
with E E =1 (see Exercise 10.4.7). By Theorem 10.4.4 and the fact that b
i
2
in Hirzebruch-Jung continued fractions, this holds for the resolutions constructed
in Theorem 10.2.3. Hence we have the following.
Corollary 10.4.9. The resolution of singularities of the afne toric surface U

constructed in Theorem 10.2.3 is minimal.


Rational Double Points Reconsidered. If has parameters d, d 1, then from
Exercise 10.2.2, the Hirzebruch-Jung continued fraction expansion of d/(d 1) is
given by
d/(d 1) = [[2, 2, . . . , 2]],
with d 1 terms. Hence b
i
= 2 for all i, and (10.4.3) gives the (d 1) (d 1)
matrix

2 1 0 0
1 2 1 0
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
0 1 2 1
0 0 1 2

representing the intersection product on the subgroup of Pic(X

) generated by the
components of the exceptional divisor for the resolution of a rational double point
of type A
d1
. We can now fully explain the terminology for these singularities.
The problem of classifying lattices
Ze
1
Ze
s
with negative denite bilinear forms B satisfying B(e
i
, e
i
) = 2 for all i arises
in many areas within mathematics, most notably in the classication of complex
simple Lie algebras via root systems. The matrix above is the (negative of) the
Cartan matrix for the root system of type A
d1
, which is often represented by the
Dynkin diagram:
t t t t t q q q
with d 1 vertices. The vertices represent the lattice basis vectors. The edges
connect the pairs with B(e
i
, e
j
) = 0 and B(e
i
, e
i
) = 2 for all i as above. In our
case, the vertices represent the components D
i
of the exceptional divisor, and the
bilinear form is the intersection product.
A precise denition of a surface rational double point follows.
10.4. Smooth Toric Surfaces 501
Denition 10.4.10. A singular point p of a normal surface X is a rational double
point or Du Val singularity if X has a minimal resolution of singularities : Y X
such that if K
Y
is a canonical divisor on Y, then every irreducible component E
i
of
the exceptional divisor E over p satises
K
Y
E
i
= 0.
We can relate these concepts to the toric case as follows.
Proposition 10.4.11. Assume has parameters d > k > 0 and let : X

be the resolution of singularities constructed in Theorem 10.2.3. Then the singular


point of U

is a rational double point if and only if k = d 1.


Proof. The canonical divisor of X

is K
X

r+1
i=0
D
i
, and one computes that
K
X

D
i
= b
i
2, 1 i r.
Thus the singular point is a rational double point if and only if b
i
= 2 for all i. This
easily implies k = d 1. You will verify these claims in Exercise 10.4.8.
There is much more to say about rational double points. For example, one can
show that E =E
1
+ +E
r
satises E E =2 (you will prove this in the toric case
in Exercise 10.4.8). From a more sophisticated point of view, E E = 2 implies
that the canonical sheaf on Y is the pullback of the canonical sheaf on X under .
We will explore this in Proposition 11.2.8. See [85] for more on rational double
points.
Exercises for 10.4.
10.4.1. Here you will verify several statements made in the proof of Lemma 10.4.2.
(a) Show that if the cone is strictly convex, then the integers b
i
in (10.4.1) must be
strictly positive.
(b) Show that if u
i1
+u
i+1
= u
i
, then Cone(u
i1
, u
i+1
) must also be smooth.
10.4.2. In the proof of Theorem 10.4.3, verify that if u
j
= u
1
and u
0
= u
2
+b
1
u
1
with
b
1
<2, then is a renement of a fan

with X

H
r
, where r =[b
1
[.
10.4.3. In this exercise, you will prove some facts used in the proof of Theorem 10.4.3.
Let be a smooth complete fan in N
R
R
2
.
(a) If [(1)[ = 3, prove that X

P
2
.
(b) If [(1)[ > 3, prove that every primitive collection of (1) has two elements.
10.4.4. In the statement of Theorem 10.4.3, you might have noticed the absence of the
Hirzebruch surface H
1
. Show that this surface is isomorphic to the blowup of P
2
at one of
its torus-xed points. See Exercise 3.3.8 for more details.
10.4.5. This exercise studies several further examples of the intersection product on H
r
.
(a) Compute D
1
D
1
using (10.4.2) and also directly from Theorem 10.4.4.
(b) Compute K
2
= K K on H
r
, where K = K
Hr
is the canonical divisor.
502 Chapter 10. Toric Surfaces
10.4.6. Show that the matrix dened by (10.4.3) has a negative-denite associated qua-
dratic form. Hint: Recall that if B(x, y) is a bilinear form, the associated quadratic form is
Q(x) = B(x, x).
10.4.7. Let X have a unique singular point p and let : Y X be a resolution of singu-
larities such that no component E of the exceptional ber
1
(p) has E E = 1. In this
exercise, you will show that Y is a minimal resolution of X according to Denition 10.4.8.
Let : Z X be another resolution of singularities and consider the possibly singular
surface S = Z
X
Y. Let R be a resolution of S. Then we have a commutative diagram of
morphisms
R

X.
(a) Explain why it sufces to show that must be an isomorphism.
(b) If not, apply [131, V.5.3] to show that factors as a sequence of blowups of points.
Hence R must contain curves L with L L =1 in the exceptional ber over p.
(c) Let L be an irreducible curve on R with L L = 1 and show that E = (L) satises
E E =1.
(d) Deduce that is an isomorphism, hence : Y X is a minimal resolution.
10.4.8. This exercise deals with the proof of Proposition 10.4.11.
(a) Show that K
X
D
i
= b
i
2 for 1 i r.
(b) Show that d/k = [[2, . . . , 2]] if and only if k = d 1. Hint: Exercise 10.2.2.
(c) Show that E = D
1
+ +D
r
satises E E =2.
10.4.9. Let have parmeters d, d 1, so that the singular point of U

is a rational double
point. By Proposition 10.1.6, U

is Gorenstein, so that its canonical sheaf


U
is a line
bundle. Let : X

be the resolution constructed in Theorem 10.2.3. Prove that

U
is the canonical sheaf of X

.
10.4.10. Another interesting numerical fact about the integers b
i
from (10.4.1) is the fol-
lowing. Suppose a smooth fan has 1-dimensional cones labeled as in Lemma 10.4.1.
Then
(10.4.4) b
0
+b
1
+ +b
r1
= 3r 12.
This exercise will sketch a proof of (10.4.4).
(a) Show that (10.4.4) holds for the standard fans of P
2
, P
1
P
1
and H
r
, r 2.
(b) Show that if (10.4.4) holds for a smooth fan , then it holds for the fan obtained by
performing a star subdivision on one of the 2-dimensional cones of .
(c) Deduce that (10.4.4) holds for all smooth fans using Theorem 10.4.3.
10.5. Riemann-Roch and Lattice Polygons
Riemann-Roch theorems are a class of results about the dimensions of sheaf co-
homology groups. The original statement along these lines was the theorem of
10.5. Riemann-Roch and Lattice Polygons 503
Riemann and Roch concerning sections of line bundles on algebraic curves. This
result and its generalizations to higher-dimensional varieties can be formulated
most conveniently in terms of the Euler characteristic of a sheaf, dened in 9.4.
Riemann-Roch for Curves. A modern form of the Riemann-Roch theorem for
curves states that if D is a divisor on a smooth projective curve C, then
(10.5.1) (O
C
(D)) = deg(D) +(O
C
),
where the degree deg(D) is dened in Denition 6.3.2. This equality can be rewrit-
ten using Serre duality as follows. Namely, if K
C
is a canonical divisor on C, then
we have
H
1
(C, O
C
(D)) H
0
(C, O
C
(K
C
D))

H
1
(C, O
C
) H
0
(C, O
C
(K
C
))

.
The integer g = dim H
0
(C, O
C
(K
C
)) is the genus of the curve C. Then (10.5.1) can
be rewritten in the form commonly used in the theory of curves:
(10.5.2) dim H
0
(C, O
C
(D)) dim H
0
(C, O
C
(K
C
D)) = deg(D) +1g.
A proof of this theorem and a number of its applications are given in [131, Ch. IV].
Also see Exercise 10.5.1 below. As a rst consequence, note that if D = K
C
is a
canonical divisor, then
(10.5.3) deg(K
C
) = 2g2.
We will need to use (10.5.2) most often in the simple case X P
1
. Then g = 0
and the Riemann-Roch theorem for P
1
is the statement for all divisors D on P
1
,
(10.5.4) (O
P
1 (D)) = deg(D) +1.
The Adjunction Formula. For a smooth curve C contained in a smooth surface X,
the canonical sheaves
C
of the curve and
X
of the surface are related by
(10.5.5)
C

X
(C)
O
X
O
C
.
This follows without difculty from Example 8.2.2 (Exercise 10.5.2) and has the
following consequence for the intersection product on X.
Theorem 10.5.1 (Adjunction Formula). Let C be a smooth curve contained in a
smooth complete surface X. Then
K
X
C+C C = 2g2,
where g is the genus of the curve C.
Proof. Let i : C X be the inclusion map. Then

X
(C)
O
X
O
C
= i

X
(C) = i

O
X
(K
X
+C),
so that
2g2 = deg(
C
) = deg(i

O
X
(K
X
+C)) = (K
X
+C) C,
504 Chapter 10. Toric Surfaces
where the rst equality is (10.5.3) and the last is the denition of (K
X
+C)C given
in 6.3.
Riemann-Roch for Surfaces. The statement for surfaces corresponding to (10.5.1)
is given next.
Theorem 10.5.2 (Riemann-Roch for Surfaces). Let D be a divisor on a smooth
projective surface X with canonical divisor K
X
. Then
(O
X
(D)) =
D DD K
X
2
+(O
X
).
We will only prove this for X a smooth complete toric surface; there is a simple
and concrete proof in this case.
Proof. The theorem certainly holds for D = 0 since O
X
(D) =O
X
in this case. Our
proof will use the special properties of smooth complete toric surfaces. Recall that
if X = X

, then Pic(X) is generated by the classes of the divisors D


i
, i = 1, . . . , r,
corresponding to the 1-dimensional cones in . Hence, to prove the theorem, it
sufces to show that if the theorem holds for a divisor D, then it also holds for
D+D
i
and DD
i
for all i.
Assume the theorem holds for D. By Proposition 4.0.28, the sequence
0 O
X
(D
i
) O
X
O
D
i
0
is exact. Tensoring this with O
X
(D+D
i
) gives the exact sequence
0 O
X
(D) O
X
(D+D
i
) O
D
i
(D+D
i
) 0.
By (9.4.1), it follows that
(O
X
(D+D
i
)) = (O
X
(D)) +(O
D
i
(D+D
i
)).
By the induction hypothesis,
(10.5.6) (O
X
(D)) =
D DD K
X
2
+(O
X
).
For (O
D
i
(D+D
i
)), recall that D
i
P
1
. Hence, by the Riemann-Roch theorem
for P
1
given in (10.5.4), we have
(10.5.7) (O
D
i
(D+D
i
)) = D D
i
+D
i
D
i
+1.
Combining (10.5.6) and (10.5.7), we obtain
(O
X
(D+D
i
)) =
D DD K
X
2
+D D
i
+D
i
D
i
+1+(O
X
)
=
(D+D
i
) (D+D
i
) +D
i
D
i
D K
X
+2
2
+(O
X
)
However, using K
X
=(D
1
+ +D
r
) and Theorem 10.4.4, one computes that
(10.5.8) D
i
K
X
=D
i
D
i
2.
10.5. Riemann-Roch and Lattice Polygons 505
Substituting this into the above expression for (O
X
(D+D
i
)) and simplifying, we
obtain
(O
X
(D+D
i
)) =
(D+D
i
) (D+D
i
) (D+D
i
) K
X
2
+(O
X
),
which shows that the theorem holds for D+D
i
The proof for DD
i
is similar and is left to the reader (Exercise 10.5.3).
The following statement is sometimes considered as the topological part of the
Riemann-Roch theorem for surfaces.
Theorem 10.5.3 (Noethers Theorem). Let X be a smooth projective surface with
canonical divisor K
X
. Then
(O
X
) =
K
X
K
X
+e(X)
12
,
where e(X) is the topological Euler characteristic of X dened by
e(X) =
4

k=0
(1)
k
dim H
k
(X, C).
As before, we will give a proof only for a smooth complete toric surface. We
will also use the Hodge decomposition
H
k
(X, C)

p+q=k
H
q
(X,
p
X
)
from (9.4.11).
Proof. Demazure vanishing (Theorem 9.2.3) implies that for a smooth complete
toric surface X = X

,
(10.5.9)
(O
X
) = dim H
0
(X, O
X
) dim H
1
(X, O
X
) +dim H
2
(X, O
X
)
= 10+0 = 1.
Thus Noethers theorem for a smooth complete toric surface is equivalent to
(10.5.10) K
X
K
X
+e(X) = 12.
We prove this as follows. Set r =[(1)[ and let the minimal generators of the rays
be u
0
, . . . , u
r1
as in Lemma 10.4.1. Since K
X
=

r1
i=0
D
i
, (10.5.8) implies
K
X
K
X
=
r1

i=0
D
i
K
X
=
r1

i=0
(D
i
D
i
2) = 2r +
r1

i=0
D
i
D
i
.
If u
i1
+u
u+i
= b
i
u
i
as in (10.4.1), then D
i
D
i
=b
i
by Theorem 10.4.4. Hence
K
X
K
X
= 2r
r1

i=0
b
i
= 2r (3r 12) = 12r,
506 Chapter 10. Toric Surfaces
where the equality

r1
i=0
b
i
= 3r 12 is from Exercise 10.4.10.
We next compute e(X). Proposition 11.2.8 shows that is the normal fan of
a polygon with r = [(1)[ sides. Then the formula for dim H
q
(X,
p
X
) given in
Theorem 9.4.11 implies
dim H
0
(X, C) = dim H
0
(X, O
X
) = 1
dim H
1
(X, C) = dim H
0
(X,
1
X
) +dim H
1
(X, O
X
) = 0+0 = 0
dim H
2
(X, C) = dim H
0
(X,
2
X
) +dim H
1
(X,
1
X
) +dim H
2
(X, O
X
)
= 0+ f
1
2f
2
= r 2
dim H
3
(X, C) = dim H
1
(X,
2
X
) +dim H
2
(X,
1
X
) = 0+0 = 0
dim H
4
(X, C) = dim H
2
(X,
2
X
) = 1.
where f
1
= r and f
2
= 1 are the face numbers of a polygon with r sides. It follows
that e(X) = 1 +(r 2) +1 = r. Then (10.5.10) follows easily from the above
computation of K
X
K
X
.
We will give a topological proof of e(X) = r in Chapter 12, and in Chapter 13,
we will interpret e(X) in terms of the Chern classes of the tangent bundle.
The Riemann-Roch theorems for curves and surfaces have been vastly gen-
eralized by results of Hirzebruch and Grothendieck, and the precise relation of
Noethers theorem to the Riemann-Roch theorem for surfaces is a special case of
their approach. We will discuss Riemann-Roch theorems for higher-dimensional
toric varieties in Chapter 13.
Lattice Polygons. For the remainder of this section, we will explore the relation
between toric surfaces and the geometry and combinatorics of lattice polygons. We
will see that the results from 9.4 for lattice polytopes have an especially nice form
for lattice polygons.
Let X = X

be a smooth complete toric surface. Since (O


X
) = 1 by (10.5.9),
Riemann-Roch for a divisor D on X becomes
(10.5.11) (O
X
(D)) =
D DD K
X
2
+1,
Thus, for any Z,
(10.5.12) (O
X
(D)) =
D DD K
X
2
+1 =
1
2
(D D)
2

1
2
(D K
X
) +1.
The theory developed in 9.4 guarantees that (O
X
(D)) is a polynomial in ; the
above computation gives explicit formulas for the coefcients in terms of intersec-
tion products.
Here is an example of how this formula works.
10.5. Riemann-Roch and Lattice Polygons 507
Example 10.5.4. Let X be the Hirzebruch surface H
2
. We will use the notation of
Example 10.4.6. The divisor D = D
1
+D
2
is clearly effective, but the inequalities
(9.3.6) dening the nef cone show that D is not nef. Using K
X
= D
1
D
4
and Example 10.4.6, one computes that
D D = 0, D K
X
=2.
Then (10.5.12) implies that
(O
X
(D)) = +1.
An easy application of Serre duality gives H
2
(X, O
X
(D)) = 0 (Exercise 10.5.4).
Thus
dim H
0
(X, O
X
(D)) dim H
1
(X, O
X
(D)) = +1.
Things get more surprising when we compute dim H
0
(X, O
X
(D)). Using the ray
generators u
1
, . . . , u
4
from Example 10.4.6, the polygon P
D
corresponding to D =
D
1
+D
2
is dened by the inequalities
'm, u
1
` 1, 'm, u
2
` 1, 'm, u
3
` 0, 'm, u
4
` 0.
The polygon P
D
is shown in Figure 11.
P
D
u
2
u
4
u
3
u
1
Figure 11. The polygon of the divisor D and the fan of H2
Even though P
D
is not a lattice polytope, Proposition 4.3.3 still applies. Thus
dim H
0
(X, O
X
(D)) =[P
D
M[ =[P
D
M[ =

1
4

2
+ +1 even
1
4

2
+ +
3
4
odd,
where the nal equality follows from Exercise 9.4.13. Combining this with the
above computation of (O
X
(D)), we obtain
dim H
1
(X, O
X
(D)) =

1
4

2
even
1
4

1
4
odd.
This is a vivid example how the Euler characteristic smooths out the complicated
behavior of the individual cohomology groups.
508 Chapter 10. Toric Surfaces
On the other hand, if D is nef, the higher cohomology is trivial by Demazure
vanishing, so that the Euler characteristic reduces to dim H
0
. We exploit this as
follows. Suppose that P M
R
R
2
is a lattice polygon. By Theorem 9.4.2 and
Example 9.4.4, the Ehrhart polynomial Ehr
P
(x) Q[x] of P satises
(10.5.13) Ehr
P
() =[(P) M[ = Area(P)
2
+
1
2
[PM[ +1
for N. We next describe this polynomial in terms of intersection products.
By the results of 2.3, we get the projective toric surface X
P
coming from the
normal fan
P
of P. In general X
P
will not be smooth, so we compute a minimal
resolution of singularities
: X

X
P
using the methods of this chapter. Recall that X
P
has the ample divisor D
P
whose
associated polygon is P.
Proposition 10.5.5. There is unique torus-invariant nef divisor D on X

such that
(a) The support function of D equals the support function of D
P
.
(b) (O
X

(D)) is the Ehrhart polynomial of P.


Proof. Proposition 6.2.7 implies that X

has a divisor D that satises part (a). As


in 6.1, we call D the pullback of D
P
. Since D
P
has a convex support function, the
same is true for D, so that D is nef. Furthermore, P is the polytope associated to
D
P
and hence is the polytope associated to D since the polytope of a nef divisor is
determined by its support function (Theorem 6.1.7).
It follows that dim H
0
(X

, O
X

(D)) = [P
D
M[ = [(P) M[ when 0,
so that H
0
(X

, O
X

(D)) equals the Ehrhart polynomial of P when 0. How-


ever, D is nef when 0 and hence has trivial higher cohomology by Demazure
vanishing (Theorem 9.2.3). Thus 0 implies
(O
X

(D)) = dim H
0
(X

, O
X

(D)) =[(P) M[.


Since (O
X

(D)) is a polynomial in , it must be the Ehrhart polynomial of P.


Theorem 10.5.5 and (10.5.12) imply that the Ehrhart polynomial of P is
Ehr
P
() =
1
2
(D D)
2

1
2
(D K
X

) +1.
Comparing this to the formula (10.5.13) for Ehr
P
(), we get the following result.
Proposition 10.5.6. Let P be a lattice polygon and let D be the pullback of D
P
constructed in Proposition 10.5.5. Then
D D = 2Area(P)
D K
X

=[PM[.
10.5. Riemann-Roch and Lattice Polygons 509
Example 10.5.7. Take the fan of H
2
shown in Figure 11 from Example 10.5.4
and combine the two 2-dimensional cones containing u
2
into a single cone. The
resulting fan has minimal generators u
1
, u
3
, u
4
that satisfy u
1
+u
3
+2u
4
= 0, so the
resulting toric variety is P(1, 1, 2).
Let P = Conv(0, 2e
1
, e
2
) M
R
, which is the double of the polytope shown
in Figure 11. The normal fan of P is the fan of P(1, 1, 2). The minimal generators
u
1
, u
3
, u
4
of this fan give divisors D

1
, D

3
, D

4
on P(1, 1, 2), and the divisor D
P
is
easily seen to be the ample divisor 2D

1
.
Since the fan of H
2
renes the fan of P(1, 1, 2), the resulting toric morphism
H
2
P(1, 1, 2) is a resolution of singularities. By considering the support function
of D
P
, we nd that the pullback of D
P
= 2D

1
is D = 2D
1
+2D
2
. We leave it as
Exercise 10.5.5 to compute D D and D K
H
2
and verify that they give the numbers
predicted by Proposition 10.5.6.
Sectional Genus. The divisor D
P
on X
P
is very ample since dim P = 2. Hence it
gives a projective embedding X
P
P
s
such that O
P
s (1) restricts to O
X

(D
P
). In
geometric terms, this means that hyperplanes H P
s
give curves X
P
H X
P
that
are linearly equivalent to D
P
. For some hyperplanes, the intersection X
P
H can be
complicated. Since X
P
has only nitely many singular points, the Bertini theorem
(see [131, II 8.18 and III 7.9.1]) guarantees that when H is generic, C = X
P
H is
a smooth connected curve contained in the smooth locus of X
P
. The genus g of C
is called the sectional genus of the surface X
P
.
We will compute g in terms of the geometry of P using the adjunction formula.
Since we need a smooth surface for this, we use a resolution : X

X
P
and note
that C can be regarded as a curve in X

since is an isomorphism away from the


singular points of X
P
. Since C D
P
on X
P
, we have C D on X

, where D is the
pullback of D
P
. Then the adjunction formula (Theorem 10.5.1) implies
2g2 = K
X

C+C C = K
X

D+D D,
so that
(10.5.14) g =
1
2
D (K
X

+D) +1.
Then we have the following result.
Proposition 10.5.8. The sectional genus of X
P
is g =[Int(P) M[.
Proof. Picks formula from Example 9.4.4 can be written as
[Int(P) M[ = Area(P)
1
2
[PM[ +1,
which by Proposition 10.5.6 becomes
[Int(P) M[ =
1
2
D D+
1
2
D K
X

+1.
The right-hand side is g by (10.5.14), completing the proof.
510 Chapter 10. Toric Surfaces
Example 10.5.9. Let P = d
2
= Conv(0, de
1
, de
2
). Then X
P
is the projective
plane P
2
in its dth Veronese embedding, and D
P
dL, where L P
2
is a line. The
hyperplane sections are the curves of degree d in P
2
, and the smooth ones have
genus
g =[Int(d
2
) M[ =
(d 1)(d 2)
2
.
You will check this assertion and another example in Exercise 10.5.6.
The curves C X

studied here can be generalized to the study of hypersur-


faces in projective toric varieties coming from sections of a nef line bundle. The
geometry and topology of these hypersurfaces have been studied in many papers,
including [15], [19], [77] and [197].
Reexive Polygons and The Number 12. Our nal topic gives a way to understand
a somewhat mysterious formula we noted in the last section of Chapter 8. Recall
from Theorem 8.3.7 that there are exactly 16 equivalence classes of reexive lattice
polytopes in R
2
, shown in Figure 3 of 8.3. The article [230] gives four different
proofs of the following result.
Theorem 10.5.10. Let P be a reexive lattice polygon in M
R
R
2
. Then
[PM[ +[P

N[ = 12.
One proof consists of a case-by-case verication of the statement for each of
16 equivalence classes. You proved the theorem this way in Exercise 8.3.5. The
argument was straightforward but not very enlightening! Here we will give another
proof using Noethers theorem.
Proof. Since Noethers theorem requires a smooth surface, we need to rene the
normal fan
P
of P M
R
. Since P is reexive, we can do this using the dual
polygon P

N
R
. We know from (8.3.2) that the vertices of P

are the minimal


generators of
P
. Let be the renement of
P
whose 1-dimensional cones are
generated by the rays through the lattice points on the boundary of P

. This is
illustrated in Figure 12.
s
s
s
s
s
s
s
s
s
s
d
d
d
d
d
d
P

s
s
s
s

d
d
P

e
e
e
e
r
r
r
r
Figure 12. A reexive polygon P and its dual P

10.5. Riemann-Roch and Lattice Polygons 511


The fan has the following properties:
For each cone of , its minimal generators and the origin form a triangle whose
only lattice points are the vertices. Thus is smooth by Exercise 8.3.4.
The minimal generators of are the lattice points of P

lying on the boundary.


Thus [(1)[ =[P

N[.
From the rst bullet, we get a resolution : X

X
P
. Recall that D
P
=K
X
P
since
P is reexive. The wonderful fact is that its pullback via is again anticanonical,
i.e., D =K
X

. To prove this, recall that D and D


P
have the same support function
, which takes the value 1 at the vertices of P

since D
P
= K
X
P
. It follows that
= 1 on the boundary of P

. Then D =K
X

because the minimal generators of


all lie on the boundary.
Now apply Noethers theorem to the toric surface X

. By (10.5.10), we have
K
X

K
X

+e(X

) = 12.
We analyze each term on the left as follows. First, D =K
X

implies
K
X

K
X

=D K
X

=[PM[,
where the last equality follows from Proposition 10.5.6. Second, e(X

) is the num-
ber of minimal generators of by the proof of Theorem 10.5.3. In other words,
e(X

) =[(1)[ =[P

N[,
where the second equality follows from the above analysis of . Hence the theorem
is an immediate consequence of Noethers theorem.
A key step in the above proof was showing that the pullback of the canonical
divisor on X
P
was the canonical divisor on X
P
. This may fail for a general resolution
of singularities. We will say more about this when we study crepant resolutions in
Chapter 11.
Exercises for 10.5.
10.5.1. The Riemann-Roch theorem for curves, in the form (10.5.1), can be proved by
much the same method as used in the proof of Theorem 10.5.2. Namely, show that if
(10.5.1) holds for a divisor D then it also holds for the divisors D+P and DP, where P
is an arbitrary point on the curve.
10.5.2. Prove the adjunction formula (Theorem 10.5.1) using (10.5.3) and (10.5.5).
10.5.3. Complete the proof of Theorem 10.5.2 by showing that if the theorem holds for
D, then it also holds for DD
i
where D
i
is any one of the divisors corresponding to the
1-dimensional cones in .
10.5.4. Let D=

be an effective Q-Cartier Weil divisor on a complete toric variety


X

of dimension n. Use Serre duality (Theorem9.2.10) to prove that H


n
(X

, O
X
(D)) =0.
10.5.5. In Example 10.5.7, compute D D and D K
H2
and check that they agree with the
numbers given by Proposition 10.5.6.
512 Chapter 10. Toric Surfaces
10.5.6. This exercise studies the sectional genus of toric surfaces.
(a) Verify the formula given in Example 10.5.9 for the sectional genus of P
2
in its dth
Veronese embedding.
(b) Let P = Conv(0, ae
1
, be
2
, ae
1
+be
2
). What is the smooth toric surface X
P
in this case?
Show that its sectional genus is (a 1)(b 1).
10.5.7. Let P be a reexive polygon.
(a) Prove that the singularities (if any) of the toric surface X
P
are rational double points.
Hint: Proposition 10.1.6.
(b) Prove that X
P
has sectional genus g = 1. This means that smooth anticanonical curves
in X
P
are all elliptic curves.
(c) Explain how part (b) relates to Exercise 10.5.6.
10.5.8. According to Theorem 10.4.3, every smooth toric surface is a blowup of either P
2
,
P
1
P
1
, or H
r
for r 2. For each of the 16 reexive polygons in Figure 3 of 8.3, the
process described in the proof of Theorem 10.5.10 produces a smooth toric surface X

.
Where does X

t in this classication in each case? (This gives a classication of smooth


toric Del Pezzo surfaces.)
10.5.9. As in 9.3, the p-Ehrhart polynomials of a lattice polygon P M
R
are dened by
Ehr
p
P
() = (

p
XP
(D
P
)), p = 0, 1, 2.
We know that Ehr
0
P
is the usual Ehrhart polynomial Ehr
P
, and then Ehr
2
P
(x) = Ehr
P
(x)
by Theorem 9.4.7. The remaining case is Ehr
1
P
. Prove that
Ehr
1
P
(x) = 2Area(P)x
2
+ f
1
2,
where f
1
is the number of edges of P. Hint: Use Theorem 9.4.11 for the constant term and
part (c) of Theorem 9.4.7 for the coefcient of x. For the leading coefcient, tensor the
exact sequence of Theorem 8.1.6 with O
XP
(D
P
), take the Euler characteristic, and then let
.
Chapter 11
Toric Resolutions and
Toric Singularities
This chapter will study the singularities of a normal toric variety. We begin in
11.1 with the existence of toric resolutions of singularities. In 11.2 we consider
more special resolutions, including simple normal crossing resolutions, crepant
resolutions, log resolutions, and embedded resolutions. There are also relations
with ideal sheaves, Rees algebras, and multiplier ideals, to be studied in 11.3, and
nally, in 11.4 we consider some important classes of toric singularities.
11.1. Resolution of Singularities
The singular locus of an irreducible variety X, denoted X
sing
, is the set of all sin-
gular points of X. One can prove that X
sing
is a proper closed subvariety of X (see
[131, Thm. I.5.3]). We call the complement X `X
sing
the smooth locus of X.
Denition 11.1.1. Given an irreducible variety X, a resolution of singularities of
X is a morphism f : X

X such that:
(a) X

is smooth and irreducible.


(b) f is proper.
(c) f induces an isomorphism of varieties f
1
(X `X
sing
) X `X
sing
.
Furthermore, f : X

X is a projective resolution if f is a projective morphism.


A resolution of singularities modies X to make it smooth without changing
the part that is already smooth. In particular, a resolution of singularities is a bi-
rational morphism. Hironaka [143] proved the existence of a resolution of singu-
larities over an algebraically closed eld of characteristic 0. An introduction to
Hironakas proof and more recent developments can be found in [138].
513
514 Chapter 11. Toric Resolutions and Toric Singularities
Fortunately, the toric case is much simpler, partly because toric varieties cannot
have arbitrarily bad singularities. Hence we will be able to give a complete proof
of resolution of singularities for toric varieties. The key idea is that given a fan
in N
R
, the resolution will be the toric morphism
: X

coming from a suitable renement

of .
The Singular Locus. The rst step is to determine the singular locus of X

. This
has the following purely toric description.
Proposition 11.1.2. Let X

be the toric variety of the fan . Then:


(X

)
sing
=

not smooth
V()
X

`(X

)
sing
=

smooth
U

,
where V() = O() is the closure of the T
N
-orbit corresponding to .
Proof. Recall that is smooth if its minimal generators can be extended to a basis
of N and that is smooth if and only if U

is smooth. Also observe that


if is smooth, then so is every face of , hence
if is not smooth, then so is every cone of containing .
The rst bullet tells us that the smooth cones of form a fan whose toric variety is

smooth
U

. This open set of X

is clearly smooth, and its complement in X

is

not smooth
O() by the Orbit-Cone correspondence.
We also have

not smooth
V() =

not smooth
O() by the second bullet and
the Orbit-Cone correspondence. Hence the proof will be complete once we show
that every point of O() is singular in X

when is not smooth. It sufces to work


in the afne open set U

. Let N

= Span() N and pick a complement N


2
N
such that N = N

N
2
. By (1.3.2), we have
U
,N
U
,N
T
N
2
.
Since dim =rankN

, the orbit O
N
() U
,N
consists of the unique xed point
of the action of T
N
. Since is not smooth with respect to N

, the proof of Theo-


rem 1.3.12 shows that this xed point is singular in U
,N
. Then (1.3.3) shows that
every point of O
N
() T
N
2
is singular in U
,N
T
N
2
. Since the above product
decomposition induces
O
N
() O
N
() T
N
2
(Exercise 11.1.1), we see that all points of O() = O
N
() are singular in X

.
It follows that when constructing renements of a fan , we need to be sure to
leave the smooth cones of unchanged.
11.1. Resolution of Singularities 515
Star Subdivisions. We will construct renements using a generalization of the star
subdivision

() dened in 3.3. Given a fan in N


R
and a primitive element
v [[ N, let

(v) be the set of the following cones:


(a) , where v / .
(b) Cone(, v), where v / and v .
We call

(v) the star subdivision of at v. This terminology is justied by the


following result.
Lemma 11.1.3.

(v) is a renement of .
Proof. Let Cone(, v) be a cone as in (b) above. Then observe that v / Span()
since v / and = Span(). It follows that dim Cone(, v) = dim +1.
To show that

(v) is a fan, rst consider a face of a cone in

(v). If the
cone arises from (a) above, then the face clearly lies in

(v). If the cone arises


from (b), then we have a face of Cone(, v). Write = H
m
Cone(, v) for
m Cone(, v)

and observe that m

and 'm, v` 0. There are two cases:


'm, v` = 0, in which case = Cone(H
m
, v)

(v).
'm, v` > 0, in which case = H
m

(v).
You will provide the details of these cases in Exercise 11.1.2.
Next consider two cones in

(v). If neither contains v, then the intersection


is clearly in

(v). Now suppose that one does not contain v and the other does.
This gives cones v / and Cone(, v) as in (a) and (b) above. We claim that
(11.1.1) Cone(, v) = ,
which will imply that Cone(, v)

(v). One inclusion of (11.1.1) is obvious;


for the other inclusion, let u Cone(, v) and write u = u
0
+v for u
0
and
0. Also let

be a cone containing v. Then


u = u
0
+v Cone(, v)

.
Since u
0
, v

and

is a face of

, it follows that u
0
, v

. But v / ,
so that = 0 and hence u = u
0
. Thus u , as desired.
The remaining case to consider is when both cones contain v. If the cones are
Cone(
1
, v) and Cone(
2
, v), then we claim that
(11.1.2) Cone(
1
, v) Cone(
2
, v) = Cone(
1

2
, v),
which will imply that Cone(
1
, v) Cone(
2
, v)

(v). As above, one inclusion


is obvious, and for the other, take u Cone(
1
, v) Cone(
2
, v). Then
u = u
1
+
1
v = u
2
+
2
v,
where u
i

i
and
i
0. We may assume that
1

2
, in which case u
2
= u
1
+
(
2

1
) Cone(
1
, v). Thus u
2

2
Cone(
1
, v), which by (11.1.1) implies
u
2

2
Cone(
1
, v) =
1

2
.
516 Chapter 11. Toric Resolutions and Toric Singularities
Thus u = u
2
+
2
v Cone(
1

2
, v), as desired.
Finally, we need to show that

(v) is a renement of . Since every cone of

(v) is clearly contained in a cone of , we need only show that each is


a union of cones in

(v). When v / , this is obvious since

(v). On the
other hand, when v , pick cone generators m
i

, i = 1, . . . , r, and note that


v implies 'm
i
, v` 0 for all i. Then set
(11.1.3) = min
/m
i
,v)>0
'm
i
, u`
'm
i
, v`
.
We claim that u
0
= u v lies in . To prove this, take m =

r
i=1

i
m
i

i
0, and compute
'm, u
0
` =

r
i=1

i
m
i
, uv

/m
i
,v)=0

i
'm
i
, u` +

/m
i
,v)>0

i
'm
i
, v`

'm
i
, u`
'm
i
, v`

.
The rst sum is 0 since u , and the second is 0 by the denition of . Hence
u
0
(

= , as claimed.
The denition (11.1.3) implies that we can pick i with 'm
i
, v` > 0 and =
/m
i
,u)
/m
i
,v)
. Then 'm
i
, u
0
` = 0, so that u
0
lies in the face =H
m
i
of . This face does
not contain v since 'm
i
, v` > 0, hence u Cone(, v)

(v).
Example 11.1.4. Let = Cone(u
1
, . . . , u
r
) be a smooth cone in N
R
R
n
.
In 3.3, we dened the star subdivision

() using u
0
= u
1
+ +u
r
. It is easy
to check that this is the star subdivision

(u
0
) since u
0
N is primitive. See
Figure 9 in 3.3 for a picture when n = 3.
Example 11.1.5. Figure 1 shows a fan in R
3
consisting of two 3-dimensional
cones (one simplicial, the other not) that meet along a 2-dimensional face . The
v

Figure 1. A fan and its star subdivision

(v)
11.1. Resolution of Singularities 517
gure also shows the star subdivision induced by a primitive vector v Relint().
The fan

(v) has ve 3-dimensional cones.


The denition of

(v) implies that v generates a 1-dimensional cone of

(v)
and hence gives a divisor on X

(v)
. Recall that a Weil divisor is Q-Cartier if some
positive integer multiple is Cartier.
Proposition 11.1.6. The star subdivision

(v) has the following properties:


(a) The 1-dimensional cones of

(v) consist of the 1-dimesional cones of plus


the ray
v
= Cone(v) generated by v, i.e.,

(v)(1) = (1)
v
.
(b) The torus-invariant divisor prime D
v
is Q-Cartier on X

(v)
.
(c) The induced toric morphism : X

(v)
X

is projective.
Proof. The rst assertion is Exercise 11.1.3. To show that D
v
is Q-Cartier, we
construct a support function : [

(v)[ R as follows. For Cone(, v)

(v),
consider the linear function dened on Span(Cone(, v)) =Span()+Rv that maps
Span() to 0 and v to 1. This is well-dened since v / Span(). Then dene
[
Cone(,v)
to be the restriction of this linear map to Cone(, v). The remaining
cones in

(v) are the cones not containing v. We set [

= 0 for each of
these. By (11.1.1) and (11.1.2), we get a well dened map : [

(v)[ R that is
linear on each cone of

(v). Furthermore, is clearly dened over N


Q
since
is rational and v N. It follows that some positive integer multiple k is integral
with respect to N, i.e., k SF(

(v), N). Then by Theorem 4.2.12,


D
k
=

(v)(1)
k(u

)D

is a Cartier divisor on X

(v)
. However, v is primitive and hence is the ray generator
of
v
. Since vanishes on all other ray generators, we have D
k
=k(v)D
v
=
kD
v
. This proves that D
v
and hence D
v
are Q-Cartier.
For the nal assertion, we use the projectivity criterion from Theorem 7.2.12.
Since

(v) renes , it sufces to nd a torus-invariant Cartier divisor on X

(v)
whose support function is strictly convex with respect to the fan

(v)

(v) [

for every . The required Cartier divisor is easy to nd: it is the divisor kD
v
,
where k > 0 is from the previous paragraph, with support function k. We can
ignore the factor of k when thinking about strict convexity. Hence the proof will be
complete once we prove that has the desired strict convexity.
When v / , this is obvious since

(v) in this case. So it remains to con-


sider what happens when v . By restricting to Span(), we can assume that is
518 Chapter 11. Toric Resolutions and Toric Singularities
full dimensional. Then, as noted in the discussion leading up to Proposition 7.2.3,
the strict convexity criteria from Lemma 6.1.13 apply to this situation.
The analysis of

(v) given in the proof of Lemma 11.1.3 shows that =

Cone(, v), where the union is over all facets of not containing v. Now x
such a facet and consider the cone

= Cone(, v)

(v). Pick m

M
such that H
m
= . Since v `, we have
(11.1.4) 'm, u

` = 0, (1) and 'm, v` > 0.


We claim that after rescaling m by a positive constant, we have (u) = 'm, u` for
all u

. To see why, note that is linear on

, vanishes on , and takes the value


1 on v. Comparing this to (11.1.4), our claim follows immediately.
Given any (1) `(1), the way we picked m implies that
'm, u

` > 0.
However, m represents on

, and
(1) `(1) =

(v)

(1) `

(1)
by part (a) and the denitions of

(v) and

(v)

. By (f) (a) of Lemma 6.1.13,


it follows that [

is strictly convex with respect to

(v)

.
Simplicialization. As an application of star subdivisions, we prove that every fan
has an efcient simplicial renement. Here is the precise result.
Proposition 11.1.7. Every fan has a renement

with the following properties:


(a)

is simplicial.
(b)

(1) = (1).
(c)

contains every simplicial cone of .


(d)

is obtained from by a sequence of star subdivisions.


(e) The induced toric morphism X

is projective.
Proof. First observe that part (c) follows from part (b). To see why, suppose that

is a renement of satisfying

(1) = (1) and let be simplicial. If

lies in , then

(1) = (1) implies

(1) (1), and hence

is a face
of because the latter is simplicial. Since =

, it follows that some

in this union must equal , i.e.,

.
Note also that part (e) is a consequence of part (d). This follows because, rst,
star subdivisions give projective morphisms by Proposition 11.1.6, and second,
compositions of projective morphisms are projective by Proposition 7.0.5.
Hence it sufces to nd a sequence of star subdivisions that lead to a simpli-
cial fan with the same 1-dimensional cones as . Our proof, based on an idea of
Thompson [270], uses complete induction on
r =[ (1) [ D

is not Q-Cartier on X

[.
11.1. Resolution of Singularities 519
If r = 0, then every D

is Q-Cartier. Since the D

generate Cl(X

), every Weil
divisor on X

is Q-Cartier. By Proposition 4.2.7, we conclude that is simplicial.


If r > 0, then we can pick (1) such that D

is not Q-Cartier. Consider


the star subdivision

(u

), where u

is the ray generator of . Proposition 11.1.6


implies that

(u

)(1) = (1) = (1)


and that D

is Q-Cartier on X

(u)
. Note also that if D

is Q-Cartier on X

, then
it remains Q-Cartier on X

(u)
(think support functions). It follows easily that

(u

)(1) [ D

is not Q-Cartier on X

(u)

has strictly fewer than r elements. Our inductive hypothesis gives a renement

of

(u

) which is easily seen to be the desired renement of .


In Exercise 11.1.4 you will show that the number of star subdivisions needed
to create the simplicial renement

of described in Proposition 11.1.7 is at


most the rank of Cl(X

)/Pic(X

). A different proof of Proposition 11.1.7 can be


found in [99]. See also [93, Thm. V.4.2].
The Multiplicity of a Simplicial Cone. When we have a fan that is simplicial but
not smooth, we need to subdivide the non-smooth cones to make them smooth.
The elegant approach used in Chapter 10 does not generalize to higher dimensions.
Instead, we will use the multiplicity of a simplicial cone from 6.4 and show how a
carefully chosen star subdivision of a non-smooth cone can lower its multiplicity.
Let us recall the denition of multiplicity. Given a simplicial cone N
R
with
generators u
1
, . . . , u
d
, let N

=Span()N and note that Zu


1
+ +Zu
d
N

has
nite index. Then the multiplicity or index of is the index
(11.1.5) mult() = [N

: Zu
1
+ +Zu
d
].
The multiplicity of a simplicial cone has the following properties.
Proposition 11.1.8. Let N
R
be a simplicial cone and let u
1
, . . . , u
d
and N

be
as in (11.1.5). Then:
(a) is smooth if and only if mult() = 1.
(b) mult() is the number of points in P

N, where
P

d
i=1

i
u
i
: 0
i
< 1.
(c) Let e
1
, . . . , e
d
be a basis of N

and write u
i
=

d
j=1
a
i j
e
j
. Then
mult() =[det(a
i j
)[.
(d) mult() mult() whenever _.
520 Chapter 11. Toric Resolutions and Toric Singularities
Proof. The proof of part (a) is straightforward, and part (b) is equally easy since
the composition
P

N N

/(Zu
1
+ +Zu
d
)
is a bijection (Exercise 11.1.5). Part (c) is the standard fact that [det(a
i j
)[ is the
index of Zu
1
+ +Zu
d
in N

(see [242, Corollary 9.63]). Finally, if _ , then


(1) (1). This implies P

, and then part (d) follows from part (b).


Some further properties of mult() are discussed in Exercise 11.1.5.
The Resolution. We can now construct a fan that resolves the singularities of a
normal toric variety X

.
Theorem 11.1.9. Every fan has a renement

with the following properties:


(a)

is smooth
(b)

contains every smooth cone of .


(c)

is obtained from by a sequence of star subdivisions.


(d) The toric morphism : X

is a projective resolution of singularities.


Proof. We rst observe that part (d) follows from parts (a), (b) and (c). Since
the renement

comes from a sequence of star subdivisions, : X

is
projective by the proof of Proposition 11.1.7. Furthermore, is the identity on
the smooth locus of X

by Proposition 11.1.2 and X

is smooth Theorem 3.1.19.


Hence is a resolution of singularities.
To produce a renement that satises (a), (b) and (c), we begin with Proposi-
tion 11.1.7, which uses star subdivisions to construct a simplicial renement with
the same simplicial cones as . This renement does not change the smooth cones
since smooth cones are simplicial.
Hence we may assume that is simplicial. To measure how close is to being
smooth, we dene
mult() = max

mult().
Note that mult() =1 if and only if is a smooth fan. We will show that whenever
mult() > 1, we can nd a star subdivision

(v) that does not change the smooth


cones of and satises either
(11.1.6)
mult(

(v)) <mult(), or
mult(

(v)) =mult() and

(v) has fewer cones of this multiplicity.


Once this is done, the theorem will follow immediately.
Suppose mult() >1 and let
0
have maximal multiplicity. By part (b) of
Proposition 11.1.8, there is v P

0
N `0. The star subdivision

(v) consists
11.1. Resolution of Singularities 521
of the cones of not containing v together with the cones Cone(, v) where v /
and is a face of a cone containing v. We claim that
(11.1.7) mult(Cone(, v)) < mult().
Assume (11.1.7) for the moment. Part (b) of Proposition 11.1.8 shows that v
lies in no smooth cone of , so that

(v) contains all smooth cones of . Since all


cones of

(v) either lie in or satisfy the inequality (11.1.7), it follows that when
we replace to

(v), we lose at least one cone of maximum multiplicity (namely,

0
) and create no new cones of this multiplicity. Thus

(v) satises (11.1.6).


It remains to prove (11.1.7). Given v P

0
N 0 as above, we can write
v =
1
u
1
+ +
d
u
d
, 0 <
i
< 1,
where u
1
, . . . , u
d
are the ray generators of the minimal face
0
of
0
containing v.
We also have Cone(, v) . Since v implies that
0
is a face of , we can
write the ray generators of as
u
1
, . . . , u
d
, u
d+1
, . . . , u
s
, s = dim.
Also, v / implies that there is i 1, . . . , d such that u
i
/ . Then
mult(Cone(, v)) mult(Cone(u
1
, . . . , u
i
, . . . , u
d
, u
d+1
, . . . , u
s
, v)
=[det(u
1
, . . . , u
i
, . . . , u
d
, u
d+1
, . . . , u
s
, v)[
=[det(u
1
, . . . , u
i
, . . . , u
d
, u
d+1
, . . . , u
s
,
i
u
i
)[
=
i
[det(u
1
, . . . , u
s
)[ =
i
mult() < mult().
The rst line follows from part (d) of Proposition 11.1.8 and the second follows
from part (c). Here, det(w
1
, . . . , w
d
) = det(b
i j
) where w
i
=

d
j=1
b
i j
e
j
. The last
two lines follow from 0 <
i
< 1 and standard properties of determinants. This
completes the proof of (11.1.7), and the theorem follows.
The Exceptional Locus. The resolution : X

from Theorem 11.1.9 is an


isomorphism above the smooth locus of X

. It remains to consider what happens


above the singular locus. The exceptional locus of is Exc() =
1
((X

)
sing
).
Proposition 11.1.2 implies that the irreducible components of (X

)
sing
are given
by the orbit closures
(X

)
sing
=V(
1
) V(
s
)
where
1
, . . . ,
s
are the minimal non-smooth cones of . Hence, to understand the
exceptional locus Exc(), it sufces to describe
1
(V(
i
)) for 1 i s. We will
use the following more general result.
Proposition 11.1.10. Let

be a renement of with induced toric morphism


: X

, and x . Then the irreducible components of


1
(V()) are

1
(V()) =V(

1
) V(

r
),
where

1
, . . . ,

r
are the minimal cones of

that meet the relative interior of .


522 Chapter 11. Toric Resolutions and Toric Singularities
Proof. Since is equivariant, it maps T
N
-orbits into T
N
-orbits. More precisely, we
know from Lemma 3.3.21 that for , a cone

satises (O(

)) O()
if and only if is the minimal cone of containing

. Furthermore, the latter


happens if and only if

intersects the relative interior of . From here, the proof


is an easy application of the Orbit-Cone correspondence (Exercise 11.1.6).
Example 11.1.11. Our rst example of a toric resolution was in Example 10.1.9,
where we considered the renement of the cone shown in Figure 2. Here, U


Figure 2. A cone with smooth renement
has a unique singular point, and the minimal cones of that meet the interior of
are the interior rays of . Hence the exceptional locus of X

is a divisor.
Example 11.1.12. Consider = Cone(e
1
, e
2
, e
1
+e
3
, e
2
+e
3
) R
3
. Figure 3 on
the next page shows and three smooth renements
1
,
2
,
3
that give
X

For
1
,
2
, the smallest cone meeting the interior of is 2-dimensional, so the
exceptional locus is P
1
in each case. On the other hand, the ray in Figure 3 is the
smallest cone of
3
meeting the interior of . Hence the exceptional locus is V(),
which by Example 3.2.8 is P
1
P
1
.
The exceptional loci in the above resolutions are related by the diagram
(11.1.8)
P
1
P
1

P
1


P
1

pt
where the nontrivial maps in (11.1.8) are the projections onto the factors of P
1
P
1
(Exercise 11.1.7).
Example B.6.1 shows how to nd the renement
1
of using Sage [262].
11.1. Resolution of Singularities 523
z
y
x

z
y
x

3
z
y
x

1
z
y
x

2
Figure 3. The cone with smooth renements 1, 2, 3
524 Chapter 11. Toric Resolutions and Toric Singularities
Example 11.1.12 was rst described by Atiyah [8] in 1958. The rational map
X
1
X
2
in (11.1.8) is an isomorphism outside of a set of codimension 2. In the
minimal model program, such rational maps are called ips or ops, depending on
the context. We will say more about this in Chapter 15.
In 10.4, we noted that when X

is a toric surface, there is a unique minimal


toric resolution of singularities X

. Example 11.1.12 shows that unique


minimal resolutions do not exist in dimensions 3.
Exercises for 11.1.
11.1.1. The proof of Proposition 11.1.2 constructed an isomorphism U
,N
U
,N
T
N2
and claimed that it induced an isomorphism of orbits
O
N
() O
N
() T
N2
.
Prove this.
11.1.2. Let Cone(, v) be cone of the star subdivision

(v) containing v, and let be the


face of Cone(, v) determined by m Cone(, v)

. Thus = H
m
Cone(, v). As noted in
the text, we have m

and 'm, v` 0.
(a) If 'm, v` = 0, prove that = Cone(H
m
, v)

(v).
(b) If 'm, v` > 0, in which case = H
m

(v).
11.1.3. Let

(v) be the star subdivision determined by v and let


v
=Cone(v). Prove that

(v)(1) = (1)
v
.
11.1.4. Recall that if G is a nitely generated abelian group and G
tor
is the subgroup of
elements of nite order, then G/G
tor
Z
r
, where r is called the rank of G. Prove that the
number of star subdivisions need to create the simplicial renement

of described in
Proposition 11.1.7 is at most the rank of Cl(X

)/Pic(X

).
11.1.5. This exercise will study the multiplicity mult() of a simplicial cone .
(a) Prove part (a) of Proposition 11.1.8.
(b) Prove part (b) of Proposition 11.1.8.
(c) Prove that mult() is the normalized volume of P

(N

)
R
= Span(), where nor-
malized means that the parallelotope determined by a basis of N

has volume 1.
Hint: Use part (c) of Proposition 11.1.8 and remember how to compute the volume of
a parallelotope (see [263, Theorem 15.2.1]).
(d) Let be a face of and let u
1
, . . . , u
d
be the minimal generators of . Prove that
mult() = mult()[N

: N

+Zu
1
+ +Zu
d
]
and use this to give another proof of part (d) of Proposition 11.1.8.
11.1.6. Complete the proof of Proposition 11.1.10.
11.1.7. Verify the claims made in Example 11.1.12.
11.1.8. Let be the fan in R
n+1
consisting of Cone(e
0
, . . . , e
n
) R
n+1
and its faces. Let
v =

n
i=0
q
i
e
i
, where the q
i
are positive integers satisfying gcd(q
0
, . . . , q
n
) = 1. The star
subdivision

(v) gives a toric morphism : X

(v)
X

= C
n+1
. Prove that
1
(0)
P(q
0
, . . . , q
n
).
11.2. Other Types of Resolutions 525
11.1.9. Consider the complete fan in R
3
with six 3-dimensional cones and minimal
generators (1, 1, 1). Thus consists of the cones over the faces of a cube centered at
the origin in R
3
.
(a) Show that the simplicialization process described in Theorem 11.1.7 requires exactly
two star subdivisions to produce a simplicial renement

of .
(b) Show that

is smooth, so that X

is a projective resolution of singularities.


(c) Find a smooth renement

of such that

(1) = (1) but X

is not projective.
Conclude that X

is not a projective resolution of singularities.


11.1.10. The simplicial fan constructed in Proposition 11.1.7 is efcient (no new rays) but
noncanonical (no canonical choice for the star subdivisions used in construction). Here
you will explore a different method for simplicialization that is canonical but introduces
many new rays. The barycenter v

of a nonzero cone N
R
is the minimal generator
of Cone(

(1)
u

) N. Now list the nonzero cones of the fan as


1
, . . . ,
r
, where
dim
1
dim
r
. Then the barycentric subdivision of is the fan
() =

(v
r
)

(v
r1
)

(v
1
).
Thus () is obtained fromby a sequence of star subdivisions where we use the barycen-
ters of the cones, starting with the biggest cones and working down.
(a) Consider the fan in R
3
from Exercise 11.1.9. Draw two pictures to illustrate the
construction of (). The rst picture, drawn on the surface of the cube, should
illustrate the intermediate fan obtained by taking star subdivisions at the barycenters
of the 3-dimensional cones. The second picture should show how this gets further
subdivided to obtain ().
(b) When the cones are listed by increasing dimension, cones of the same dimension can
appear in any order. Hence the list is not unique. However, they all give the same
barycentric subdivision (). Prove this.
(c) Prove that () is a simplicial fan.
Part (a) of this exercise was inspired by [93, p. 74].
11.2. Other Types of Resolutions
We next consider simple normal crossing, crepant, log, and embedded resolutions,
which are important renements of what it means to be a resolution of singularities.
SNC Resolutions. Example 11.1.12 illustrates that the exceptional locus of a res-
olution of singularities need not always be a divisor. Yet in many situations in
algebraic geometry, one wants a divisor, and in fact, one often wants a simple
normal crossing divisor, as mentioned in 8.1. Recall that a divisor D =

i
D
i
on a smooth variety X has simple normal crossings (SNC for short) if every D
i
is
smooth and irreducible, and for all p X, the divisors containing p meet as follows:
if I
p
=i [ p D
i
, then the tangent spaces T
p
(D
i
) T
p
(X) meet transversely, i.e.,
codim

iIp
T
p
(D
i
)

=[I
p
[.
526 Chapter 11. Toric Resolutions and Toric Singularities
For example, given a smooth toric variety and a nonempty subset A (1), the
divisor D =

A
D

is SNC.
Denition 11.2.1. A resolution of singularities f : X

X is called SNC if the


exceptional locus Exc( f ) is a divisor with simple normal crossings.
Theorem 11.2.2. Every fan has a renement

as in Theorem 11.1.9 such that


the toric morphism : X

is a projective SNC resolution of singularities.


Proof. By Theorem 11.1.9, has a smooth renement
0
that gives a resolution
of singularities. Let V(
i
) be an irreducible component of (X

)
sing
and let
0
be a minimal cone that meets the relative interior of
i
. If is not a ray, then
consider the star subdivision

0
(v

), where v

(1)
u

. Since
0
is smooth,
this is the fan denoted

0
() in Denition 3.3.17. In the discussion that followed
the denition in 3.3, we saw that corresponding toric morphism is the blowup
Bl
V()
(X

0
) X

0
. Since v

lies in the relative interior of , the inverse image of


V() is the divisor corresponding to the ray of

0
(v

) generated by v

.
The composed map Bl
V()
(X

0
) X

is a resolution of singularities whose


exceptional locus has one less non-divisorial component. Repeating this nitely
many times gives a projective resolution (star subdivisions give projective mor-
phisms) whose exceptional locus is a divisor, automatically SNC by the remark
following Denition 11.2.1.
Example 11.2.3. The resolution X

1
U

constructed in Example 11.1.12 is not


SNC. Applying the process described in the proof of Theorem 11.2.2 gives the
resolution X

3
U

, which is SNC.
Log Resolutions. There are several types of log resolutions. Here we discuss the
one for Q-divisors.
Denition 11.2.4. Let D =

i
d
i
D
i
be a Q-divisor on a normal variety X. Then
a projective resolution of singularities f : X

X is a log resolution of D if

i
f
1
(D
i
) Exc( f ) is a SNC divisor, where as usual Exc( f ) is the exceptional
locus of f .
For a toric variety X

, the natural choice for D is a torus-invariant Q-divisor


D =

, a

Q. In this case, a projective SNC resolution X

is au-
tomatically a log resolution of (X, D) since any collection of torus-invariant prime
divisors on X

is SNC.
In 11.3 we will discuss a different type of log resolution, where the given data
consists of a normal variety X and a sheaf of ideals a O
X
. See [186, 9.1.B] for
more on log resolutions.
Crepant Resolutions. Given the importance of the canonical sheaf
X
, it is natural
to ask how
X
changes in a resolution of singularities f : X

X. The nicest case is


11.2. Other Types of Resolutions 527
when X is normal and Q-Gorenstein, i.e.,
X
=O
X
(K
X
) for a Q-Cartier canonical
divisor K
X
. (Note that if one canonical divisor is Q-Cartier, then they all are.)
Suppose that K
X
is Cartier, where > 0 is an integer. It follows that f

(K
X
)
is a Cartier divisior, and then f

K
X
is dened to be the Q-divisor
f

K
X
=
1

(K
X
).
Example 11.2.7 below will show that f

K
X
need not be integral.
Denition 11.2.5. Let X be a normal Q-Gorenstein variety. Then a resolution of
singularities f : X

X is crepant if f

(K
X
) is integral and K
X

(K
X
).
Since K
X
is only unique up to linear equivalence, the denition of crepant is
often stated as K
X

= f

K
X
. The term crepant is due to Reid [235] and signals
that there is no discrepancy between K
X

and f

K
X
.
Example 11.2.6. The toric variety U

of = Cone(e
1
, e
2
, e
1
+e
3
, e
2
+e
3
) R
3
is Gorenstein since the minimal generators of lie in the plane 'm, u` = 1, for
m =e
1
e
2
. Thus the canonical divisor is K
U
= div(
m
).
For the fan
1
from Figure 3 of Example 11.1.12, the resolution X

1
U

is
crepant since

(K
U
) =

(div
U
(
m
)) = div
X

(
m
) = K
X

.
The subscripts indicate where the divisor is being computed. You will verify this
computation in Exercise 11.2.1, where you will also show that the SNC resolution
X

3
U

from Example 11.1.12 is not crepant.


Example 11.2.7. Consider = Cone(e
1
, e
2
, e
3
) N
R
=R
3
relative to the lattice
N =(a, b, c) Z
3
[ a+b+c 0 mod 2.
The minimal generators are u
i
= 2e
i
, 1 i 3, and the class group of U

is Z/2Z,
where the torus-invariant divisors D
i
all map to the generator. It follows that the
canonical divisor K
U
= D
1
D
2
D
3
is not Cartier, but 2K
U
= div(
m
) for
m =e
1
e
2
e
3
M
R
(Exercise 11.2.2).
The star subdivision with respect to u
0
=e
1
+e
2
+e
3
N
R
gives a resolution
of singularities : X

. The four minimal generators of give divisors


denoted D
0
, D
1
, D
2
, D
3
by abuse of notation. Then

(2K
U
) =

(div
U
,N
(
m
)) = div
X

(
m
)
=3D
0
2D
1
2D
2
2D
3
= 2K
X

D
0
,
so that K
X

(K
U
) +
1
2
D
0
(Exercise 11.2.2). Thus is not crepant. Note also
that

(K
U
) is not an integral divisor even though K
U
is.
For a toric surface, we have the minimal resolution X

constructed in
10.2. It is easy to characterize when this resolution is crepant.
528 Chapter 11. Toric Resolutions and Toric Singularities
Proposition 11.2.8. Let X

be a toric surface and consider the minimal resolution


of singularities : X

from 10.2. Then the following are equivalent:


(a) is crepant.
(b) X

is Gorenstein.
(c) The singularties of X

are at worst rational double points.


Proof. The equivalence (b) (c) follows from Proposition 10.1.6. For (a)
(c), one can reduce to the case of an afne surface U

with ray generators u


1
, u
2
.
Here, the key observation is that (a) is equivalent to saying that the minimal smooth
renement of has ray generators lying on the line segment connecting u
1
and u
2
,
u
2
u
1

Figure 4. Cone with renement ray generators on a line segement


as illustrated in Figure 4. Using the methods of 10.2, one sees that the rational
double points are the only toric surface singularities where is happens. We leave
the details as Exercise 11.2.3.
Here is an example that we have already seen.
Example 11.2.9. Let P be a reexive polygon in R
2
. In Theorem 10.5.10, we
proved that the minimal resolution X

X
P
comes from the fan given by the lattice
points on the boundary of P

. Figure 5 makes it clear that X

X
P
is crepant, a
fact we used in the proof of Theorem 10.5.10.
s
s
s
s
s
s
s
s
s
s
d
d
d
d
d
d
P

s
s
s
s

d
d
P

e
e
e
e
r
r
r
r
Figure 5. A reexive polygon P and its dual P

11.2. Other Types of Resolutions 529


More generally, crepant resolutions of toric varieties coming from reexive
polytopes are important in mirror symmetry because they give rise to Calabi-Yau
varieties. A smooth projective variety Y is Calabi-Yau if
Y
is trivial and
(11.2.1) H
1
(Y, O
Y
) = = H
d1
(Y, O
Y
) = 0, d = dimY.
Nowx a reexive polyope PM
R
of dimension 2. We know from 8.3 that
X
P
is a Gorenstein Fano toric variety, so that K
X
P
is Cartier and ample. Assume
also that : X

X
P
is a crepant resolution. Then K
X

(K
X
P
) is basepoint free.
By the Bertini theorem [131, Cor. III.10.9 and Ex. III.11.3], a generic section of
O
X

(K
X

) gives a smooth connected hypersurface Y X

such that Y K
X

.
We call Y an anticanonical hypersurface.
Proposition 11.2.10. With the above hypotheses, Y is Calabi-Yau.
Proof. Since I
Y
= O
X

(Y) by Proposition 4.0.28, we have I


Y
O
X

(K
X

n
X

, n = dim X

. Then the desired vanishing (11.2.1) follows from the long exact
cohomology sequence associated to
0 I
Y
O
X

O
Y
0
plus the vanishing of H
p
(X

,
n
X

) for p < n from Theorem 9.3.2.


To compute the canonical bundle of Y, we use the adjunction formula from
Example 8.2.2:

Y

X

O
X

(I
Y
/I
2
Y
)

(Y)
O
X

O
Y
,
where we use I
Y
/I
2
Y
I
Y

O
X

(O
X

/I
Y
) =O
X

(Y)
O
X

O
Y
. Since

=O
X

(K
X

) O
X

(Y),
one sees immediately that
Y
=
n1
Y
is trivial.
Example 11.2.11. Continuing with Example 11.2.9, let P be a reexive polygon
with minimal resolution X

X
P
. In this case, a generic section of O
X

(K
X

) is
a smooth connected curve C X

. We have two ways to compute its genus g:


(a) Recall from 10.5 that g is the sectional genus of X
P
, which equals [Int(P)M[
by Proposition 10.5.8. Since P is reexive, the origin is its unique interior
point. Hence g = 1.
(b) The canonical bundle of C is trivial by Proposition 11.2.10. The canonical
divisor has degree 0, so 0 = 2g2 by (10.5.3), hence g = 1.
Thus C is an elliptic curve, which is a Calabi-Yau variety of dimension 1.
Unfortunately, for a reexive polytope of dimension 3, X
P
need not have a
crepant resolution. Here, the best one can do is the maximal projective crepant
partial desingularization described by Batyrev in [16]. The resulting anticanonical
hypersurfaces are singular Calabi-Yau varieties (see [68, Def. 1.4.1]).
We will study toric singularities in more detail in 11.4.
530 Chapter 11. Toric Resolutions and Toric Singularities
Embedded Resolutions. Many treatments of the resolution of singularities use em-
bedded resolutions. For an embedded resolution, we have an irreducible variety X
contained in a smooth variety W, which we think of as the ambient space. The
idea is to repeatedly blow up the ambient space along smooth subvarieties (thereby
changing the ambient space) until the proper transform of X becomes smooth.
Here is a more careful description. An irreducible smooth subvariety Z W
gives the blowup f : Bl
Z
(W) W, where Bl
Z
(W) is smooth and the exceptional
locus Exc( f ) (the subset of Bl
Z
(W) where f is not an isomorphism) is the divisor
given by the projective bundle of the normal bundle of Z W. We call Z the center
of the blowup. Then X W has a proper transform in Bl
Z
(W) dened as the
Zariski closure
f
1
(X `Z) Bl
Z
(W).
Then one seeks a series of smooth centers Z
i
W
i
, 0 i , such that W
0
=W,
W
i+1
= Bl
Z
i
(W
i
), and Z
i+1
is transverse to the exceptional locus of W
i+1
W.
Furthermore, the nal map f

: W

W needs to have the following properties:


The induced map f [
X

: X

X is a resolution of singularities, where X

is the
proper transform of X in W

.
X

is transverse to the exceptional locus of f

.
The last bullet implies that f [
X

: X

X is a SNC resolution of singularities, and


one can show that f [
X

is a projective morphism. A careful explanation of what


this means can be found in [138], along with a discussion of further properties one
can impose on an embedded resolution.
In the toric version of this, we have an equivariant embedding X W of toric
varieties, with W smooth, and at each stage of the above process, W
i
is toric, and
the center Z
i
W
i
is an orbit closure. Then Z
i
is smooth, W
i+1
= Bl
Z
i
(W
i
) is toric,
and W
i+1
W
i
is a toric morphism. When such an embedded resolution exists, the
result is a toric SNC resolution of singularities X

X achieved entirely through


blowups of torus-invariant centers in smooth toric ambient spaces. The papers [25],
[32] and [117] prove the existence of embedded toric resolution, and more impor-
tantly, they give algorithms for nding the centers Z
0
, Z
1
, . . . needed at each stage
of the resolution process. See also [268]. In contrast, the resolution constructed in
Theorem 11.1.9 depended on many choices and is far from unique.
The proof of toric embedded resolution, though much simpler than the general
case, is still not easy. Hence we will conne ourselves to some examples, together
with some remarks about nonnormal toric varieties.
Example 11.2.12. The toric variety V(xy zw) C
4
is the afne toric variety U

of the cone from Example 11.1.12. We can resolve the singularity of U

by
blowing up the origin in C
4
. The proper transform of V(xyzw) in Bl
0
(C
4
) is the
toric variety X

3
from Example 11.1.12. There are two ways to see this.
11.2. Other Types of Resolutions 531
First, the isomorphism U

V(xy zw) induces the isomorphism (C

)
3

V(xyzw) (C

)
4
given by (t
1
, t
2
, t
3
) (t
1
, t
2
, t
3
, t
1
t
2
t
1
3
) (see Example 1.1.5). At
the level of one-parameter subgroups, this gives the embedding
: Z
3
Z
4
, (a, b, c) (a, b, c, a+bc).
Since =
1
R
(R
4
0
) and R
4
0
is the cone for C
4
, the corresponding toric morphism
is the embedding : U

C
4
with image V(xy zw).
Let be the star subdivision of R
4
0
at (1, 1, 1, 1). This gives the toric variety
X

= Bl
0
(C
4
). Via
R
, induces the renement
3
of described in Exam-
ple 11.1.12 (Exercise 11.2.4). This gives a toric morphism : X

3
Bl
0
(C
4
)
whose image is the proper transform (we will say more about this below).
Second, the proper transform can studied by the method of Example 5.2.11.
The blowup has quotient representation
Bl
0
(C
4
) = (C
5
`C(0, 0, 0, 0))/C

,
where (t, x, y, z, w) = (
1
t, x, y, z, w), and the blowup map Bl
0
(C
4
) C
4
is given by
(t, x, y, z, w) (tx, ty, tz, tw).
Via this map, 0 = xy zw becomes 0 = (tx)(ty) (tz)(tw) = t
2
(xy zw), so that
the proper transform is dened by xy zw = 0 in Bl
0
(C
4
). There are two cases:
If t = 0, then t (t, x, y, z, w) = (1, tx, ty, tz, tw), so that this part of the proper
transform is isomorphic to V(xy zw) `0 C
4
.
If t = 0, then this part of the proper transform is the ber over 0 V(xy zw)
and is the subvariety of P
3
dened by xy = zw. This quadric surface is the
Segre embedding of P
1
P
1
.
One can also show that the proper transform is smooth by checking locally on the
afne pieces of Bl
0
(C
4
), along the lines of Example 5.2.11 (Exercise 11.2.4).
The general situation has an interesting wrinkle. Given a smooth toric variety
X

0
with torus T
N
0
, any subtorus T T
N
0
gives the toric variety
(11.2.2) X = T X

0
with torus T. However, X need not be normal, as shown by the simple example
where T =(t
2
, t
3
) [ t C

(C

)
2
and X = T = V(x
3
y
2
) C
2
.
For this reason, the embedded resolution of toric varieties described in [25],
[32] and [117] works with an equivariant embedding X W of toric varieties where
W is smooth but X is not assumed to be normal. These papers use cones and fans
for the ambient space W and its blowups, but use binomial equations for X and its
proper transforms, just as we did in the second half of Example 11.2.12.
532 Chapter 11. Toric Resolutions and Toric Singularities
Example 11.2.13. In Example 5.2.11 we showed that the proper transform of the
nonnormal toric variety V(x
3
y
2
) C
2
is smooth in Bl
0
(C
2
). This not an embed-
ded resolution since the proper transform is not transverse to the exceptional locus.
Two more blowups are needed to get an embedded resolution (Exercise 11.2.5).
While the toric variety X from (11.2.2) need not be normal, its normalization is
easy to construct. Suppose in general that X

0
is a normal toric variety, where
0
is
a fan in (N
0
)
R
. Then subtori of T
N
0
correspond bijectively to sublattices of N
0
with
torsion-free quotient (Exercise 11.2.6). Let T
N
be the subtorus coming from N N
0
and let X = T
N
X

0
be its Zariski closure. Using the subspace N
R
(N
0
)
R
, the
fan of the normalization of X is described as follows.
Proposition 11.2.14. Let X

0
be the toric variety of a fan
0
in (N
0
)
R
. Fix a
subtorus T
N
T
N
0
and let X = T
N
X

0
. Then X is a (possibly nonnormal) toric
variety with torus T
N
whose normalization is the toric variety X

of the fan
= N
R
[
0

in N
R
. In particular, if X is normal, then X X

.
Proof. The proof is similar to the proof of Theorem 3.A.5.
Example 11.2.15. As in Example 11.2.12, consider U

V(xyzw) C
4
, and let
X be the proper transform of V(xyzw) in Bl
0
(C
4
). Then X is the Zariski closure
of its torus in Bl
0
(C
4
), and we also know that X is smooth by the second half of
Example 11.2.12. Using Proposition 11.2.14, it is now easy to show that X is the
toric variety X

3
, as claimed in Example 11.2.12 (Exercise 11.2.4).
Exercises for 11.2.
11.2.1. In Exercise 11.2.6,
(a) Prove that

(div
U
(
m
)) = div
X
(
m
) for any m M. Hint: Use Proposition 6.2.7.
(b) Show that the resolution X
3
U

from Example 11.1.12 is not crepant.


11.2.2. Fill in the details omitted in Example 11.2.7.
11.2.3. Complete the proof of Proposition 11.2.8. Hint: If has parameters d, k, then the
results of 10.2 give a smooth renement of with minimal generators u
0
, . . . , u
r+1
such
that u
0
= e
2
, u
1
= e
1
, u
r+1
= de
1
ke
2
. When are u
0
, u
1
, u
r+1
collinear?
11.2.4. This exercise concerns Example 11.2.12.
(a) In the rst half of the example, prove the claim that
3
is the renement of induced
by the fan for Bl
0
(C
4
) via the map
R
.
(b) Fill in the details omitted in the second half of the example.
(c) Use parts (a) and (b) together Proposition 11.2.14 to show that X
3
is isomorphic to
the proper transform of V(xy zw) in Bl
0
(C
4
).
11.2.5. Explain why the blowup Bl
0
(C
2
) C
2
does not give an embedded resolution of
singularities of V(x
3
y
2
) C
2
. Then construct a blowup of Bl
0
(C
2
) that does.
11.2. Other Types of Resolutions 533
11.2.6. Fix a lattice N
0
. Prove that there is a bijection between sublattices of N
0
with
torision-free quotient and subtori of T
N0
.
11.2.7. Consider the complete fan in R
3
consisting of the cones over the faces of the
cube with vertices (1, 1, 1). In Exercise 11.1.9 you constructed a renement

giving
a projective resolution of singularities.
(a) Explain why X

is not SNC.
(b) Find a nice renement of

that gives a projective SNC resolution of singularities.


11.2.8. The goal of this exercise and the next is to construct an embedded resolution of the
rational double point singularity A
k
fromExample 10.1.5. We will change notation slightly
and consider V(xzy
k+1
) C
3
, k 1. The blowup Bl
0
(C
3
) has homogeneous coordinates
t, x, y, z in the sense of Chapter 5 such that Bl
0
(C
3
) C
3
is given by (t, x, y, z) (tx, ty, tz).
(a) Show that the proper transform of V(xz y
k+1
) is dened by xz =t
k1
z
k+1
.
(b) Bl
0
(C
3
) is covered by three afne open subsets isomorphic to C
3
. By the discussion
of local coordinates in 5.2, these afne charts are given by setting one of x, y, z equal
to 1. Show that the proper transform is smooth in the x = 1 and z = 1 charts and is
dened by xz =t
k1
in the y = 1 chart. Conclude that the proper transform is normal.
(c) Show that the embedded resolution of V(xz y
k+1
) C
3
can be done using k/2
blowups at smooth points of the ambient space. Hint: Use part (b) and recursion.
11.2.9. We continue the study of V(xz y
k+1
) C
3
begun in Exercise 11.2.8, this time
focusing on the fan. First observe that V(xz y
k+1
) is the Zariski closure of the subtorus
(t
1
, t
2
, t
1
1
t
k+1
2
) [ t
1
, t
2
C

(C

)
3
corresponding to the sublattice
N =Z(1, 0, 1) +Z(0, 1, k +1) Z
3
.
Since V(xz y
k+1
) is normal (Example 10.1.5), Proposition 11.2.14 implies that it is the
afne toric variety of the cone = N
R
R
3
0
since R
3
0
is the cone for C
3
.
(a) Show that = Cone((k +1, 1, 0), (0, 1, k +1)).
(b) Bl
0
(C
3
) is the toric variety of the star subdivision of R
3
0
at =(1, 1, 1). This induces
a renement of , shown for k = 2 in Figure 6 on the next page. Prove that for
k 1, is the fan of the proper transform. Hint: Part (b) of the previous exercise.
(c) When k = 1, show that is smooth with exactly two 2-dimensional cones. Draw a
picture similar to Figure 6 and explain why it is simpler.
(d) When k 2, show that the fan has three 2-dimensional cones, where the outer two
cones are smooth and the middle cone gives A
k1
.
The methods of 10.2 resolve U

by subdividing into k +1 smooth cones (see Exer-


cise 10.2.2). When k 2, the fan constructed in part (b) has the outermost cones of the
smooth renement plus a middle cone that combines the remaining cones of the smooth
renement. When k 3, blowing up Bl
0
(C
3
) renes by adding two more cones fromthe
smooth renement, and each successive blowup adds in two more cones from the smooth
renement until the process is complete.
534 Chapter 11. Toric Resolutions and Toric Singularities
induced
refinement of
Figure 6. for k =2 and renement induced by blowing up 0 C
3
in Exercise 11.2.9
11.3. Rees Algebras and Multiplier Ideals
So far, we have constructed resolution of singularities using repeated blowups. By
using the blowup of a sheaf of ideals, it is often possible to construct the resolu-
tion in a single blowup. From an abstract point of view, this is a standard fact.
Given any birational projective morphism f : X Y of irreducible varieties with Y
quasiprojective, there is an ideal sheaf I O
Y
such that f : X Y is the blowup
map : Bl
I
(Y) Y. This is proved in [131, Thm. II.7.17]. In the toric case, we
will see that the blowup has an especially nice interpretation.
Blowups and Polyhedra. We begin with a full dimensional lattice polyhedron P
in M
R
. Thus P = Q+C, where Q is a lattice polytope and C is strongly convex.
In order to get a blowup, we assume in addition that C is full dimensional, so that
=C

N
R
is also full dimensional and strongly convex. Then P = Q+

and
the normal fan
P
is a renement of . Hence we have a birational toric morphism
: X
P
U

,
which is projective by Theorem 7.1.10. Our goal is to represent this morphism as
the blowup of an ideal a in the coordinate ring R =C[

M] of U

.
Translating P by an element of M has no effect on X
P
. For m Int(

) M,
one sees easily that P+m

for integers 0. Hence we may assume that


P

. This gives the ideal


(11.3.1) a =

mPM
C
m
R =C[

M] =

M
C
m
,
and from a R, we get the Rees algebra
R[a] =

k=0
a
k
t
k
,
11.3. Rees Algebras and Multiplier Ideals 535
where t is a dummy variable that keeps track of the grading. This graded R-algebra
is generated by its elements of degree 1 as an R-algebra and satises R[a]
0
= R.
Using the Proj construction described in the appendix to Chapter 7, we get a
projective morphism
(11.3.2) X = Proj(R[a]) Spec(R) =U

.
This is the blowup of Spec(R) with respect to a, denoted Bl
a
(Spec(R)). For us, a
key property of this blowup involves the ideal sheaf aO
X
, dened as follows. On
an afne open subset U = Spec(A) X, the map U Spec(R) comes from a ring
homomorphism R A, and aO
U
is the sheaf associated to the ideal aA A. For
the blowup morphism (11.3.2) the sheaf aO
X
is a line bundlethis is [131, Prop.
II.7.13]. See [131, II.7] for a discussion of blowing up ideals and ideal sheaves.
Besides the Rees algebra, the polyhedron P gives a second graded algebra.
Recall from 7.1 that from P M
R
we get the cone C(P) M
R
R, where the
slice at height k > 0 is kP and the slice at height 0 is

(see Figure 1 from 7.1).


The associated semigroup algebra S
P
= C[C(P) (MZ)] is graded by height,
and (S
P
)
0
= R by the discussion following the proof of Theorem 7.1.13.
Proposition 11.3.1. With the above setup, there is an inclusion R[a] S
P
that
makes S
P
the integral closure of R[a]. Furthermore, this inclusion is an equality if
and only P is normal.
Proof. First observe that R[a] is a semigroup algebra. Let
A =

M) 0

(PM) 1

MZ.
We leave it to reader to show that R[a] =C[NA] (Exercise 11.3.1). Also note:
We have ZA = MZ since

is full dimensional and P

.
Cone(A) =C(P).
Then Proposition 1.3.8 implies that S
P
=C[C(P) (MZ)] is the integral closure
of R[a] =C[NA].
Since R[a] and S
P
are equal in degrees 1, the nal assertion of the proposition
is an easy consequence of the denition of normal polyhedron.
To translate Proposition 11.3.1 into geometric terms, we will use the Cartier
divisor D
P
of the polyhedron P. Recall that the Cartier data of D
P
is given by the
vertices v P. This means that the support function of D
P
satises (u) = 'v, u`
when u lies in the cone
v

P
corresponding to v. Thus (u) 0 since v P

and u
v
. It follows that D =

(u

)D

is an effective divisor. Since


D
P
=

(u

)D

, we conclude that D
P
=D, where D 0.
In particular, O
X
P
(D
P
) = O
X
P
(D) is an ideal sheaf of O
X
P
. This compares to
aO
X
P
as follows.
536 Chapter 11. Toric Resolutions and Toric Singularities
Corollary 11.3.2. With the same hypotheses as Proposition 11.3.1, we have:
(a) The map X
P
U

factors
X
P
Bl
a
(U

) U

,
where the rst morphism is normalization and the second is the blowup of a.
(b) aO
X
P
=O
X
P
(D
P
), where D
P
is the Cartier divisor associated to P.
Proof. For part (a), recall from Theorem 7.1.13 that X
P
=Proj(S
P
), where S
P
is the
integral closure of R[a] by Proposition 11.3.1. You will prove in Exercise 11.3.2
that the inclusion R[a] S
P
induces the normalization map
X
P
= Proj(S
P
) Proj(R[a]).
For part (b), take a vertex v P. This gives the cone
v

P
, and since
P = Conv(m [
m
a), one sees easily that

v
= Cone(mv [
m
a). Then any

m
a can be written

m
=
v

mv

v
C[

v
M],
from which we conclude that aC[

v
M] =
v
C[

v
M] since
v
a. Since X
P
is covered by the afne open subsets U
v
, it follows that aC[

v
M] is the line
bundle of the Cartier divisor whose Cartier data is given by the vertices of P. As
noted above, this divisor is D
P
, which completes the proof.
This corollary shows that the map X
P
U

is a single blowup followed by a


normalization. Moreover, replacing P with a sufciently high multiple, we may
assume that P is normal (Theorem 7.1.9). This does not change the normal fan, so
that in this case, the map X
P
U

is the blowup of the ideal a.


By Theorem 7.2.4, it follows that any projective toric resolution X

is
the blowup of a suitable ideal a C[

M]. For the special case of blowing up


the origin 0 C
n
or more generally a coordinate subspace 0C
nr
C
n
, the
ideal a is easy to describe (Exercise 11.3.3).
More on the Ideal. If we x the cone N
R
, then we can give a purely algebraic
description of the ideals that arise from (11.3.1). We rst need a denition.
Denition 11.3.3. An ideal a in a ring R is integrally closed if whenever an element
r R satises an equation
r
k
+a
1
r
k1
+ +a
k1
r +a
k
= 0
with a
i
a
i
for 1 i k, we have r a.
Then we have the following result.
11.3. Rees Algebras and Multiplier Ideals 537
Proposition 11.3.4. Fix a full dimensional strongly convex cone N
R
. Then the
maps
P a =

mPM
C
m
a P = Conv(m [
m
a)
induce a bijection

lattice polyhedra P

with recession cone

nonzero integrally closed


ideals a C[

M]
generated by characters

.
Proof. Given a lattice polyhedron P

with recession cone

, we need to
show that a is integrally closed. Standard arguments (see [89, p. 137] or [266, Cor.
1.3.1]) imply that
a =r C[

M] [ r
k
+a
1
r
k1
+ +a
k
= 0 for some a
i
a
i

is an integrally closed ideal containing a. Hence it sufces to prove a = a.


Since a is generated by characters, it is invariant under the T
N
-action on R,
which easily implies that a is T
N
-invariant. Then Lemma 1.1.16 implies that a is
also generated by characters (see also [266, Prop. 1.4.2]). Thus, to prove a = a, we
need to show that if a character
m
a satises

km
+a
1

(k1)m
+ +a
k1

m
+a
k
= 0
with a
i
a
i
, then
m
a. Solving for the rst term of the above equation gives

km
=a
1

(k1)m
a
k1

m
a
k
.
Writing a
i
as a linear combination of i-fold products of characters in a, it follows
that for some 1 i k, we must have an equality

km
=
m
1

m
i

(ki)m
with
m
1
, . . . ,
m
i
a, i.e., m
1
, . . . , m
i
P. This implies
m =
1
i
(m
1
+ +m
i
),
which lies in P since P is convex. Hence
m
a, proving that a is integrally closed.
In the other direction, a nonzero ideal a = '
m
1
, . . . ,
ms
` C[

M] gives
the convex set P = Conv(m [
m
I)

. In Exercise 7.1.12 you showed that


P = Conv(m
1
, . . . , m
s
) +

. Hence P is a lattice polyhedron contained in I with


recession cone

.
It remains to show that these maps are inverses of each other. One direction is
easy, for if Pa, then a Conv(m[ mPM), which equals P since Pis a lattice
polyhedron. The other direction takes more thought. A nonzero integrally closed
ideal a gives P = Conv(m [
m
a). We need to show that m PM implies

m
a. To prove this, rst note that m PM implies m Conv(m
1
, . . . , m
s
)
where
m
1
, . . . ,
ms
a. By Carath eodorys theorem (see [281, Prop. 1.15]), we
538 Chapter 11. Toric Resolutions and Toric Singularities
can assume that m
1
, . . . , m
s
are afnely independent. Since m, m
1
, . . . , m
s
M, it
follows that
m =

s
i=1

i
m
i
, where
i
0,

s
i=1

i
= 1,
i
Q.
Alfter clearing denominators, we get km =

s
i=1

i
m
i
where k,
i
, . . . ,
s
N and

s
i=1

i
= k > 0. Then r =
m
satises the equation
r
k

s
i=1
(
m
i
)

i
= 0.
Since

s
i=1
(
m
i
)

i
a
k
and a is integrally closed, we must have r =
m
a.
Example 11.3.5. For the cone R
3
of Example 11.1.12, Figure 7 shows the
polyhedron P

obtained by pushing one of the facets of

one unit to the right.


translate
this facet
one unit
to the
right
P

Figure 7. The polyhedron P and the dual cone

The corresponding ideal a is generated by the vertices of P (a rare occurrence, but


true here), so that
a ='
e
2
,
e
1
+e
2
e
3
` C[

M].
Looking at P, one sees that its normal fan is the fan
2
from Figure 3 in Exam-
ple 11.1.12. Since P is normal (Exercise 11.3.4), it follows that X

2
= Bl
a
(U

).
We can also think about this from the point of view of the simplicialization
process described in the proof of Proposition 11.1.7. Since is not simplicial,
we can take (1) such that D

is not Cartier, say = Cone(e


2
). The star
subdivision at e
2
gives the fan
2
such that D

is becomes Cartier on X

2
. Then
one can check in that P is the polyhedron of D

, so that
aO
X

2
=O
X

2
(D

)
(Exercise 11.3.4).
11.3. Rees Algebras and Multiplier Ideals 539
Example 11.3.6. The maximal ideal m =

(M\|0)
C
m
C[

M] is
clearly integrally closed. The polyhedron associated to m is the convex hull P
m
=
Conv(

(M`0)).
When has dimension 2, this convex hull was denoted

in 10.2, i.e.,
P
m
=

. In Proposition 10.2.17, we described the properties of the associated


projective toric morphism X
Pm
U

. Since lattice polyhedra are normal in dimen-


sion 2, Proposition 11.3.1 and Corollary 11.3.2 imply that this morphism is the
blowup of the ideal m, i.e., X
Pm
= Bl
m
(U

).
Example 11.3.7. Consider the graph G in Figure 8, where the vertices are labeled
with variables x, y, z. This is gives the edge ideal a ='xy, yz, zx` C[x, y, z] whose
generators correspond to the edges of the graph. Note that a is radical since it is
Graph G
x
z
y

Ideal

xy, yz, zx

Fan
e
3
e
1
e
2
x
z
y
Polyhedron P

a
Figure 8. Graph, ideal, polyhedron, and fan
generated by square-free monomials and hence is integrally closed since radical
ideals are automatically integrally closed (Exercise 11.3.5).
The ideal a gives the polyhedron P shown in Figure 8. Since P sits inside the
rst octant in R
3
, the shaded facets form the back of P from our point of view.
The vertices of P correspond to the generators of a, though in general, the ideal of
a polyhedron usually has more generators than just the characters of the vertices.
540 Chapter 11. Toric Resolutions and Toric Singularities
One can show that P is normal (Exercise 11.3.5), so that
X

= X
P
= Bl
a
(C
3
) = Proj(R[a]), R =C[x, y, z],
by Proposition 11.3.1 and Corollary 11.3.2. Another proof of normality comes
from [250, Thm. 6.3], which implies that R[a] is normal since G is an odd cycle.
Hence P is normal Proposition 11.3.1.
When a C[

M] is generated by characters but not integrally closed, we


have the following description of its integral closure.
Proposition 11.3.8. Let a C[

M] be an ideal generated by characters and


set P = Conv(m [
m
a). Then:
(a) a =

mPM
C
m
, so P a in the correspondence of Proposition 11.3.4.
(b) aO
X
P
= aO
X
P
=O
X
P
(D
P
).
Proof. We leave the proof of part (a) as Exercise 11.3.6. For part (b), applying
Corollary 11.3.2 to P and a gives aO
X
P
= O
X
P
(D
P
). Furthermore, the proof of the
corollary shows that
aC[

v
M] =
v
C[

v
M]
when v is a vertex of P. However, P = Conv(m [
m
a) implies that the vertices
come from a, i.e.,
v
a. Hence aC[

v
M] = aC[

v
M], and part (b) follows
immediately.
Part (a) of the proposition is well-known when a C[x
1
, . . . , x
n
] is a monomial
ideal. See, for example, [266, Prop. 1.4.6]. Another nice result of monomial
commutative algebra is the monomial Briancon-Skoda theorem, which asserts that
if a C[

M] is an ideal generated by characters, then


a
n+1
a

where n = rankM and 1 is arbitrary. A proof can be found in [268, 4]. Mono-
mial ideals are an active area of research in commutative algebra (see [274] and the
references therein).
Global Aspect. Given a projective toric resolution of singularities X

, it is
natural to ask if X

can be represented as the blowup Bl


I
(X

) for some torus-


invariant sheaf of ideals I O
X

. When X

is quasiprojective, this can be done


by adapting the technique used in the proof of [131, Thm. II.7.17]. In general, it
might not be possible to nd a sheaf of ideals, but one can always nd a sheaf of
fractional ideals that works. This is explained in [172, Ch. I, Thms. 10 and 11].
11.3. Rees Algebras and Multiplier Ideals 541
Log Resolutions. Suppose that we have a sheaf of ideals I O
X
, where X is
smooth. Then IO
Y
is a line bundle on Y = Bl
I
(X), though Y may be rather
singular. A log resolution for I asks for a smooth version of this. Here is the
precise denition.
Denition 11.3.9. Let I O
X
be an ideal sheaf on a smooth variety X. Then a
projective birational morphism f : X

X is a log resolution of I if X

is smooth,
IO
X
=O
X
(D) for a divisor D on X

, and Supp(D) Exc( f ) is a SNC divisor.


This type of log resolution is often called a principalization. In the toric case,
we have the following principalization result of Howald [147] (see also [268]).
Theorem 11.3.10. Let a C[x
1
, . . . , x
n
] be a monomial ideal. Then there is a toric
morphism : X

C
n
that is a log resolution of a.
Proof. Let P = Conv(m [
m
a) as in Proposition 11.3.8. This gives X
P
C
n
with aO
X
P
= O
X
P
(D
P
). Since X
P
may be singular, we take a projective resolution
of singularities X

X
P
. The composed map : X
P
C
n
is projective, toric, and
birational, though its exceptional locus Exc() need not be a divisor. This is easily
xed by adapting the strategy used in the proof of Theorem 11.2.2. Hence we may
assume that Exc() is a divisor.
Then let D be the pullback of D
P
via X

X
P
. One easily checks that
aO
X

= O
X

(D), and Supp(D) Exc() is SNC since it is a union of torus-


invariant prime divisors on a smooth toric variety.
A stronger principalization result, proved by Goward in [121], asserts that the
log resolution : X

C
n
can be chosen to be a composition of blowups with
smooth torus-invariant centers. This is analogous to embedded resolution of a
toric variety, which also uses blowups with smooth torus-invariant centers. See
Exercise 11.3.7 for an example.
Multiplier Ideals. Given a proper birational morphism f : X

X where X

and
X are smooth, we have the relative canonical divisor
K
X

/X
= K
X
f

K
X
,
As noted in [186, 9.1.B], this divisor is supported on the exceptional locus Exc( f )
and hence satises
(11.3.3) f

O
X
(K
X

/X
) O
X
.
In the toric context, these assertions are easy to understand and are covered in
Exercise 11.3.8.
One consequence of (11.3.3) is that if E is any effective divisor on X

, then
O
X
(K
X

/X
E) O
X
(K
X

/X
),
542 Chapter 11. Toric Resolutions and Toric Singularities
so that f

O
X
(K
X

/X
E) f

O
X
(K
X

/X
) O
X
is an ideal sheaf. When applied
to the log resolution of an ideal sheaf, this gives the following denition.
Denition 11.3.11. Let I O
X
be a nonzero ideal sheaf on a smooth variety X
and let f : X

X be a log resolution of I such that IO


X

= O
X

(D) for an
effective divisor D on X

. Given any rational number c > 0, the multiplier ideal


.(c I) is the ideal sheaf dened by
.(c I) = f

O
X

(K
X

/X
cD).
One can show that this denition is independent of the log resolution [186,
Thm. 9.2.18]. See [34] for an introduction to multiplier ideals and [186, Ch. 910]
for a thorough discussion and many applications.
In the toric context, Howald [147] showed how to compute the multiplier ideal
.(c a) of a monomial ideal a C[x
1
, . . . , x
n
].
Theorem 11.3.12. Let P = Conv(m [
m
a) be the polyhedron associated to a
monomial ideal a C[x
1
, . . . , x
n
]. Then, for any c > 0 in Q, we have
.(c a) =

m+e
0
Int(cP)
C
m
.
were e
0
= (1, . . . , 1) Z
n
.
Proof. We use the log resolution : X

C
n
constructed in Theorem 11.3.10.
Thus aO
X

= O
X

(D), where D is the pullback of D


P
on X
P
, and it follows
that the polyhedron of D is P
D
= P. If we write D =

, a

0, then
P =m M
R
[ 'm, u

` a

for all (1),


hence
cP =m M
R
[ 'm, u

` ca

for all (1). (11.3.4)


Also note that div
C
n
(
e
0
) = K
C
n
, so that

K
C
n
= div
X

(
e
0
). Hence
K
X

/C
n cD = K
X

K
C
n
cD =

(1+'e
0
, u

` ca

)D

.
The multiplier ideal .(c a) is given by the global sections of the sheaf of this
divisor. By Proposition 4.3.3, we obtain

m
.(c a) 'm, u

` 1'e
0
, u

` +ca

for all (1)


'm+e
0
, u

` 1+ca

for all (1)


'm+e
0
, u

` > ca

for all (1).


By (11.3.4), we conclude that
m
.(c a) if and only if m+e
0
lies in the interior
of cP. This completes the proof.
11.3. Rees Algebras and Multiplier Ideals 543
Theorem 11.3.12 shows that the characters in the multiplier ideal .(c a) come
from translates by e
0
=(1, . . . , 1) of interior points of c times the polyhedron.
This result is also discussed in [186, Thm. 9.3.27].
Example 11.3.13. Consider the monomial ideal a = 'x
4
, x
2
y, y
3
` C[x, y]. The
multiplier ideal .(1 a) =.(a) is given by
.(a) ='x
2
, xy, y
2
`.
This follows from Figure 9, which shows the polyhedron P (light shading), its
P
interior
points of P
translated interior
points of P
Figure 9. The polyhedron P, its interior lattice points, and their translates
interior lattice points (in the dark shading), and the result of translating these lattice
points by (1, 1) (bounded by the dashed line).
Example 11.3.14. For the ideal a ='xy, yz, zx` C[x, y, z] of Example 11.3.7, one
sees easily that (1, 1, 1) lies in the interior of P. It follows immediately that the
multiplier ideal .(a) is trivial.
Exercises for 11.3.
11.3.1. Prove the claim made in the proof of Proposition 11.3.1 that R[a] =C[NA].
11.3.2. In Proposition 11.3.1, we showed that S
P
= C[C(P) (MZ)] is the integral
closure of R[a] = C[NA]. Prove that Proj(S
P
) Proj(R[a]) is the normalization map.
Hint: Look at the proof of Theorem 7.1.13. A character
m
a gives an element
m
t of
degree 1 in R[a] and S
P
, which in turn gives open sets
U = Spec(R[a]
(
m
t)
) and U

= Spec((S
P
)
(
m
t)
)
of Proj(R[a]) and Proj(S
P
) respectively. Show that these open sets cover Proj(R[a]) and
Proj(S
P
) and that U

is the normalization of U.
11.3.3. For the blowup of the origin 0 C
n
, nd an explicit ideal a R=C[x
1
, . . . , x
n
] such
that Bl
0
(C
n
) = Proj(R[a]). Then do the same for the blowup of the coordinate subspace
0 C
nr
C
n
.
11.3.4. Fill in the details omitted in the discussion of Example 11.3.5.
11.3.5. This exercise is concerned with Example 11.3.7.
544 Chapter 11. Toric Resolutions and Toric Singularities
(a) Prove that an ideal in C[x
1
, . . . , x
n
] generated by square-free monomials is radical.
(b) Prove that every radical ideal is integrally closed.
(c) Prove that the polytope P pictured in Figure 8 of Example 11.3.7 is normal.
11.3.6. Prove Proposition 11.3.8.
11.3.7. Let a ='x
3
, y
2
` C[x, y].
(a) Construct a log resolution : X

C
2
of a using the method described in the proof
of Theorem 11.3.10.
(b) Interpret the fan X

constructed in part (a) as a sequence of three star subdivisions,


each of which blows up a xed point of the torus action.
(c) Let E
i
be the proper transform in X

exceptional locus of the ith blowup. Show that


aO
X
=O
X
(2E
1
3E
2
6E
3
).
This exercise is taken from [34].
11.3.8. Let : X

be the toric morphism coming from a smooth renement

of a
smooth fan .
(a) Prove that K
X

K
X
is supported on the exceptional locus Exc() and is 0.
(b) Let D be an effective divisor supported on Exc(). Prove that

O
X

(D) O
X
. Hint:
Reduce to the afne case.
11.3.9. Compute the multiplier ideal .(a) of the following monomial ideals:
(a) a ='x
8
, y
6
` C[x, y] (see [147, Ex. 2]).
(b) a ='x
7
, x
3
y, xy
2
, y
6
` C[x, y] (see [186, Ex. 9.3.28]).
11.3.10. Let a ='x
a1
1
, . . . , x
an
n
` C[x
1
, . . . , x
n
], where a
1
, . . . , a
n
are positive integers. Prove
that .(c a) is trivial if and only if c <

n
i=1
1
ai
(see [147, Ex. 3]).
11.3.11. Consider the monomial ideal a = 'xy, yz, zx` C[x, y, z] and polyhedron P from
Figure 8 in Example 11.3.7. The toric morphism X
P
C
3
satises aO
XP
= O
XP
(D
P
). To
get a log resolution of a as in Theorem11.3.10, we need a smooth renement of the normal
fan of P. Figure 10 shows and two smooth renements
1
and
2
.

e
3
e
1
e
2

1
e
3
e
1
e
2

2
e
3
e
1
e
2
Figure 10. The fan and smooth renements 1 and 2
(a) Explain how
1
and
2
can be obtained from by a series of star subdivisions. Hence
X
1
C
3
and X
2
C
3
are birational projective toric morphisms.
11.3. Rees Algebras and Multiplier Ideals 545
(b) Show that X
1
C
3
is a composition of blowups at smooth torus-invariant centers.
(c) Showthat X
2
C
3
is not a composition of blowups at smooth torus-invariant centers.
11.3.12. In this exercise and the next we follow [33] and explore the toric case of two
variants of multiplier ideals. Let U

= Spec(C[

M]) be an afne toric variety. Fix an


effective torus-invariant Q-divisor D and an ideal a C[

M] generated by characters.
Also assume that K
U
+D is Q-Cartier. Thus there are m
0
M and r > 0 in Z with
div(
m0
) = r(K
U
+D). Also let P

be the polyhedron of a.
(a) Prove that there is a toric morphism : X

which is a log resolution for both D


and a. Write aO
X
=O
X
(A), where A is an effective divisor.
(b) Given c > 0 in Q, prove that we have a natural inclusion

O
X
(K
X

(K
U
+D) +cA) O
U
.
This gives a multiplier ideal .(D, c a) C[

M].
(c) Adapt the proof of Theorem 11.3.12 to show that
.(D, c a) =

m+
1
r
m0Int(cP)
C
m
.
Part (c) is [33, Thm. 1]. Note that Theorem 11.3.12 is a special case of this result. Earlier,
we dened the multiplier ideal .(c a) for C
n
. This denition extends to Q-Gorenstein
afne toric varieties. However, when U

fails to be Q-Gorenstein, then we use the log


version, since K
U
+D could be Q-Cartier for some Q-divisor D. See [186, 9.3.G].
11.3.13. To dene a multiplier ideal that works for any afne toric variety U

, we proceed
as follows. Let a C[

M] be generated by characters and let : X

be a log
resolution such that aO
X
=O
X
(A), A 0. Following [33], the multiplier module is

O
X
(K
X
cA)
U
, c > 0 in Q.
By Proposition 8.2.9, we can regard
U
as an ideal of C[

M] generated by characters
coming from lattice points in the interior of

. Then the multiplier module gives an ideal


of C[

M] denoted .

(c a). Prove that this ideal is given by


.

(c a) =

mInt(cP)
C
m
,
where as usual P is the polyhedron of a. This is [33, Thm. 2].
11.3.14. Let a R=C[x
1
, . . . , x
n
] be a monomial ideal. In the text, we used a log resolution
to dene the multiplier ideal .(c a). This exercise will present a different construction of
.(c a). Let S be the integral closure of the Rees algebra R[a]. This gives the projective
morphism : X = Proj(S) C
n
. Prove that
.(c a) =

O
X
(K
X

K
C
n cD),
where aO
X
= O
X
(D). Hint: Use Proposition 11.3.1. (For experts, this works because X
has rational singularities.)
11.3.15. This exercise is for readers who know about symbolic powers of ideal sheaves
(see, for example, [186, Def. 9.3.4]). Let D =

i
D
i
be a sum of distinct prime divisors on
a normal variety X. Then O
X
(D) O
X
is a sheaf of radical ideals. For an integer n 1,
show that O
X
(nD) equals the nth symbolic power O
X
(D)
n
. Hint: Exercise 4.0.14.
546 Chapter 11. Toric Resolutions and Toric Singularities
11.4. Toric Singularities
We next discuss the singularities of normal toric varieties. This is an active area of
research, and our treatment will omit many important topics.
General Properties of Toric Singularities. Although a normal toric variety need
not be smooth, it is at least Cohen-Macaulay. Furthermore, its singularities are of
the following special type.
Denition 11.4.1. A normal variety X has rational singularities if for every reso-
lution of singularities f : X

X, we have
R
p
f

O
X
= 0 for all p > 0.
One can show that if the above vanishing holds for one resolution of singular-
ities, then it holds for all [179, Thm. 5.10]. Hence, in the toric case, we can use a
toric resolution of singularities.
Theorem 11.4.2. A normal toric variety X

has rational singularities.


Proof. Let : X

be a toric resolution of singularities. In particular, is


proper, so that R
p

O
X

= 0 for p > 0 by Theorem 9.2.5. Then we are done by


the above remark.
The reader should consult [179, Sec. 5.1] for a careful discussion of rational
singularities.
The Simplicial Case. We next give several characterizations of the singularities of
a simplicial toric variety. The rst is topological, based on the following idea. If X
is a smooth irreducible variety of dimension n, then any x X has a neighborhood
x U in the classical topology that is homeomorphic to an open ball in C
n
R
2n
.
By excision and the long exact sequence for relative cohomology, one obtains
H
p
(X, X `x, Z) H
p
(U,U `x, Z)

0 p = 2n
Z p = 2n.
When we weaken the coefcients from Z to Q, we get the following denition.
Denition 11.4.3. An irreducible variety X of dimension n is rationally smooth if
for every x X, we have
H
p
(X, X `x, Q)

0 p = 2n
Q p = 2n.
Smooth varieties are obviously rationally smooth. Here is a example that will
be useful below.
Example 11.4.4. If GGL(n, C) is a nite group, then C
n
/G is rationally smooth
by [49, Prop. A.1].
11.4. Toric Singularities 547
Before giving our second characterization, we need some terminology. The
structure sheaf O
X
of a variety X is a sheaf in the Zariski topology. When we
switch to the classical topology, the corresponding sheaf is the sheaf of analytic
functions on X, denoted O
an
X
(see [131, App. B] for a brief description). Then
any classical open subset U X gives the analytic variety (U, O
an
U
) = (U, O
an
X
[
U
),
and varieties X
1
and X
2
are locally analytically equivalent at p
1
X
1
and p
2
X
2
if
there are classical neighborhoods p
1
U
1
X
1
and p
2
U
2
X
2
such that U
1
U
2
as analytic varieties, where the isomorphism take p
1
to p
2
.
Denition 11.4.5. Let X be an irreducible variety of dimension n.
(a) A point p X is a nite quotient singularity if there is a nite subgroup G
GL(n, C) such that p X is locally analytically equivalent to 0 C
n
/G.
(b) X is an orbifold or is quasismooth or has nite quotient singularities if every
point of X is a nite quotient singularity.
(c) X has abelian nite quotient singularities if every point of X is a nite quotient
singularity such that the nite group G in part (a) is abelian.
Here is a lemma whose proof is almost obvious.
Lemma 11.4.6. Let GGL(n, C) be a nite subgroup.
(a) C
n
/G is an orbifold.
(b) If G is abelian, then C
n
/G has abelian nite quotient singularities.
Proof. Denition 11.4.5 guarantees that 0 C
n
/G is a nite quotient singularity.
But what about the other points of C
n
/G? For v C
n
, let G
v
=g G [ g v = v
be its isotropy subgroup. We will show that 0 C
n
/G
v
is locally analytically
equivalent to v C
n
/G.
Note that w w+v is equivariant with respect to G
v
and hence induces a
local analytic equivalence from 0 C
n
/G
v
to v C
n
/G
v
. To complete the proof,
we need to nd a local analytic equivalence from v C
n
/G
v
to v C
n
/G.
Pick coset representatives so that G =

i
g
i
G
v
. The points g
i
v are distinct in
C
n
/G
v
, so there is a classical neighborhood v U such that the g
i
U are disjoint,
and replacing U with

gG
gU, we can assume that U is G
v
-invariant. Then V =

i
g
i
U is a G-invariant neighborhood of v, and one sees easily that U/G
v
V/G.
This gives the desired local analytic equivalence.
On C
n
/G, the coordinates x
1
, . . . , x
n
on C
n
become local coordinates on the
quotient. The intuition is that anything G-invariant in the x
i
descends to C
n
/G. For
example, one can show that (C
n
/G,

p
C
n
/G
) (C
n
,
p
C
n )
G
, so that

p
C
n
/G
comes
from G-invariant p-forms in the local coordinates.
548 Chapter 11. Toric Resolutions and Toric Singularities
Example 11.4.7. If G SL(n, C), then dx
1
dx
n
is G-invariant. It follows
that
C
n
/G

n
C
n
/G
is trivial, i.e., C
n
/G has trivial canonical divisor. This is why
nite subgroups of SL(n, C) are so important.
In the analytic category, the existence of local coordinates enables one to do
analysis on orbifolds. See [68, App. A.3] for further discussion and references.
We now characterize simplicial toric varieties in terms of their singularities.
Theorem 11.4.8. Given a normal toric variety X

, the following are equivalent:


(a) is simplicial.
(b) X

has abelian nite quotient singularities.


(c) X

is an orbifold.
(d) X

is rationally smooth.
Proof. First assume that is simplicial and take . If N

is the sublattice of
N Z
n
generated by the minimal generators of , then U
,N
C
k
, where k =
dim. Then Proposition 3.3.11 implies that there is a nite abelian group G with
U

=U
,N
(U
,N

(C

)
nk
)/G (C
k
(C

)
nk
)/G C
n
/G.
Since G is a subgroup of (C

)
n
GL(n, C), Lemma 11.4.6 implies that U

and
hence X

have abelian nite quotient singularities.


This proves (a) (b). Then (b) (c) is obvious, and (c) (d) follows
from Example 11.4.4. It remains to prove (d) (a). For this, we can assume that
N
R
R
n
is a cone such that U

is rationally smooth. A face _ gives the


following objects:
The sublattice N

= Span() N and quotient lattice N() = N/N

.
The exact sequence 1 T
N
T
N
T
N()
1 of tori.
The orbit closure V() = O() U

of dimension ndim.
Since O() T
N()
by Theorem 3.2.6, it follows easily that V() is the xed point
set of the action of T
N
on U

(Exercise 11.4.1). In the standard notation for xed-


point sets, this means U
T
N

=V(). Hence
(11.4.1) dimU
T
N

= ndim.
Now let
1
, . . . ,
r
be the facets of . Then each T
i
= T
N
i
is a subtorus of
T = T
N
of codimension 1 (Exercise 11.4.1). Furthermore, for any p V(),
(11.4.2)
dim
p
U

dim
p
U
T

= n(ndim) = dim
dim
p
U
T
i

dim
p
U
T

= (ndim
i
) (ndim) = 1
11.4. Toric Singularities 549
by (11.4.1). Since U

is rationally smooth at p, a result of Brion [49, Thm., p. 130]


implies that
dim
p
U

dim
p
U
T

dim
p
U
T

dim
p
U
T

,
where the sum is over all subtori T

T of codimension 1. Using (11.4.2), we


obtain
dim
r

i=1

dim
p
U
T
i

dim
p
U
T

= r.
For any cone, the number of facets is at least its dimension, with equality if and
only if the cone is simplicial (Exercise 11.4.1). Hence the above inequality implies
that is simplicial, and the proof is complete.
We note one other important characterization of simplicial toric varieties: a
normal toric variety is simplicial if and only if it is Q-factorial, meaning that every
Weil divisor is Q-Cartier. We proved this in Proposition 4.2.7.
Terminal and Canonical Singularities. Given a normal variety X, recall from
11.2 that a resolution of singularities f : X

X is crepant if X is Q-Gorenstein
and K
X
= f

K
X
. Having a crepant resolution is rare, so for most varieties, we
have K
X
= f

K
X
no matter which resolution we use. Measuring the difference
between these divisorstheir discrepancyleads to some interesting classes of
singularities.
In general, let K
X
be a canonical divisor on a normal Q-Gorenstein variety X.
Given a proper birational morphism f : X

X with X

smooth, one can nd a


canonical divisor K
X
on X

such that
(11.4.3) K
X
= f

K
X
+

i
a
i
E
i
,
where a
i
Q and the E
i
are the irreducible divisors lying in the exceptional locus
Exc( f ). This is proved in [194, Rem. 4-1-2].
Denition 11.4.9. Let X be a normal Q-Gorenstein variety.
(a) X has terminal singularities if there is a proper birational f : X

X with X

smooth, such that the coefcients a


i
in (11.4.3) satisfy a
i
> 0 for all i.
(b) X has canonical singularities if there is a proper birational f : X

X with X

smooth, such that the coefcients a


i
in (11.4.3) satisfy a
i
0 for all i.
If X has terminal (resp. canonical) singularities, then the inequalities a
i
> 0
(resp. a
i
0) hold for all proper birational morphisms f : X

X with X

smooth
[194, Rem. 4-1-2 and 4-2-2]. Note that f : X

X may fail to be a resolution of


singularities since (for instance) some smooth points of X may get blown up by f .
However, it is often useful to be able to work with these more general morphisms.
550 Chapter 11. Toric Resolutions and Toric Singularities
If X is not smooth and has terminal singularities, then X cannot have a SNC
crepant resolution. Terminal singularities are very important in the minimal model
program, since a minimal model is a Q-factorial projective variety that has only
terminal singularities. Canonical singularities are also relevant, since (roughly
speaking) canonical singularities are the those that appear on canonical models
of varieties of general type. See [194, Ch. 4] for a general discussion of terminal
and canonical singularities. We will use terminal singularities in Chapter 15 when
we discuss the toric minimal model program.
Let us apply these ideas in the toric context. Given a normal Q-Gorenstein
toric variety X

, we can nd a toric resolution of singularities : X

. Here,
we have the toric canonical divisors K
X

and K
X

, and the relation (11.4.3) can be


described explicitly as follows.
Lemma 11.4.10. Let be the support function of the Q-Cartier divisor K
X

and
let : X

be the toric morphism coming from a renement

of . Then:
K
X

K
X

(1)\(1)
((u

) 1)D

.
Proof. The key observation is that K
X

and its pullback

K
X

have the same


support function. Thus

K
X

(u

)D

. Then the desired equation


follows immediately since (u

) = 1 for all (1).


Here is an easy consequence of this lemma.
Proposition 11.4.11. Let X

be a normal Gorenstein toric variety. Then X

has
canonical singularities.
Proof. The support function of K
X

is integral with respect to N since X

is
Gorenstein. Now let

be a smooth renement of . Given u

(1), we have:
(u

) Z since u

[[ N.
(u

) >0 since u

lies in a cone and >0 on since takes the value


1 on every minimal generator of .
It follows that (u

) 1, and we are done by Lemma 11.4.10.


Our next result shows that we can determine when an afne toric variety U

has terminal or canonical singularities by studying the lattice points of the polytope

= Conv(0, u

[ (1)).
Proposition 11.4.12. Let U

and

be as above.
(a) The following are equivalent:
(i) U

is Q-Gorenstein
(ii) There is m M
Q
such that 'm, u

` = 1 for all (1).


(iii)

has a unique facet not containing the origin.


11.4. Toric Singularities 551
(b) If U

is Q-Gorenstein, then:
(i) U

has terminal singularities the only lattice points of

are given
by its vertices.
(ii) U

has canonical singularities the only nonzero lattice points of

lie in the facet not containing the origin.


Proof. We begin with part (a). On an afne toric variety, the canonical divisor is
Q-Cartier if and only some multiple is the divisor of a character. This easily gives
the equivalence of (i) and (ii). Furthermore, given m M
Q
as in (ii),

lies on one
side of the hyperplane dened by 'm, u` = 1. The face cut out by this hyperplane
contains all of the u

and hence is the unique facet of

not containing the origin,


proving (iii). We leave (iii) (ii) as Exercise 11.4.2.
For part (b), rst note that the support function of K
U
is (u) = 'm, u` for
m M
Q
is as in the previous paragraph. Then

=u [ 'm, u` 1.
Now take any primitive vector v N different from the u

. By taking a smooth
renement of the star subdivision given by v, we get a proper birational mor-
phism : X

such that v is the minimal generator of = Cone(v) (1).


By Lemma 11.4.10, the coefcient of D

in K
X

K
U
is
(11.4.4) (v) 1 ='m, v` 1.
If U

has terminal singularities, then (11.4.4) is positive. Hence 'm, v` > 1, so


that v /

. It follows that the only lattice points of

are its vertices. Conversely,


suppose that

satises this condition and let be a smooth renement of .


Given any u

(1) `(1), we have u

, so that (11.4.4) is positive. Hence


U

has terminal singularities.


The proof for canonical singularities is similar. The only difference is that the
coefcient (11.4.4) is allowed to be zero, which happens only for lattice points
lying on the facet of

not containing the origin. You will supply the details in


Exercise 11.4.2.
Example 11.4.13. Consider the lattice N = (a, b, c) Z
3
[ a+b+c 0 mod 2
and cone =Cone(e
1
, e
2
, e
3
) from Example 11.2.7. There are two ways to see that
U

has terminal singularities:


The resolution : X

constructed in Example 11.2.7 satises K


X

K
U
+
1
2
D
0
. Since
1
2
> 0 and D
0
is the exceptional locus, U

has terminal
singularities by Denition 11.4.9.
The minimal generators of with respect to N are 2e
1
, 2e
2
, 2e
3
, so that

=
Conv(0, 2e
1
, 2e
2
, 2e
3
). One easily checks that

N consists only of vertices,


so that U

has terminal singularities by Proposition 11.4.12.


In the surface case, terminal and canonical singularities are easy to understand.
552 Chapter 11. Toric Resolutions and Toric Singularities
Theorem 11.4.14. Let X

be a toric surface. Then:


(a) X

has terminal singularities X

is smooth.
(b) X

has canonical singularities X

is Gorenstein X

has at worst rational


double points.
Proof. Part (a) follows because a 2-dimensional cone = Cone(u
1
, u
2
) such that

N = Conv(0, u
1
, u
2
) N =0, u
1
, u
2
is smooth (Exercise 8.3.4).
For part (b), Proposition 11.2.8 implies that being Gorenstein is equivalent
to having rational double points, and if X

is Gorenstein, then it has canonical


singularities by Proposition 11.4.11. To complete the proof, we will show that an
afne toric surface U

with canonical singularities has a rational double point. Let


d, k be the parameters of , so that has minimal generators e
2
, de
1
ke
2
. Then

=u [ 'm, u` 1, m =
k+1
d
e
1
+e
2
.
If k < d 1, then 'm, e
1
` =
k+1
d
< 1, so that e
1
is an interior point of

. This
is impossible since U

has canonical singularities. The only remaining case is


k = d 1, which gives a rational double point by Example 10.1.5.
Reduction to Canonical and Terminal Singularities. We next discuss how any
toric singularity can be made rst canonical and then terminal. We begin with the
canonical case.
Proposition 11.4.15. A strongly convex cone N
R
has a canonically determined
renement
can
such that X
can
has canonical singularities and the induced toric
morphism : X
can
U

is projective.
Proof. Let

= Conv( (N `0)). In Exercise 11.4.3 you will show that

is a lattice polyhedron with as recession cone. You will also show that taking
cones over bounded faces of

gives a fan
can
in N
R
such that the morphism
: X
can
U

is projective.
Take


can
and suppose

is the cone over the bounded face F of

.
Then F is the unique facet of

not containing the origin, so U

is Q-Gorenstein
by Proposition 11.4.12. Since F is a facet of

= Conv( (N `0)), the only


nonzero lattice points of

lie in F, so that U

has canonical singularities by


Proposition 11.4.12.
Example 11.4.16. When is 2-dimensional, the renement
can
has an elegant
description as follows. Recall from Proposition 8.2.9 that the canonical sheaf
U
comes from the ideal
(U

,
U
) =

mInt(

)M
C
m
.
This ideal is integrally closed with associated polyhedron
P

= Conv(Int() N)
11.4. Toric Singularities 553
as in Proposition 11.3.4. The reasoning in Example 11.3.6 shows that X
P
U

is
the blowup of this ideal. We claim that

can
is the normal fan of P

,
i.e., X
can
= X
P
. This will imply that X
P
has at worst rational double points by
Proposition 11.4.14 and Theorem 11.4.15.
In dimension 2, a renement of is determined uniquely by its ray generators.
Thus it sufces to prove that
(11.4.5) the vertices of

are the ray generators of the normal fan of P

.
We show this as follows. Recall that by Proposition 10.2.17, the ray generators of
the normal fan of P
m
=

= Conv(

(M` 0)) are almost the vertices of

. The difference arises from slight complications that can occur at the edges of
, as can be seen by looking at Theorem 10.2.12 and Proposition 10.2.17.
Let us compare P
m
and P

. Since the latter uses only interior lattice points,


these polyhedra differ along the edges of

. To see how this affects their normal


fans, consider Figure 11, which shows what can happen at an edge _

. Let a
1
be the lattice length (number of lattice points1) of the bounded edge of P
m
that
a
1
> 1

a
1
= 1

Figure 11. Two examples of Pm (outlined in bold) and P (shaded)


touches . As you can see, P
m
and P

have the same inner normal vectors near


when a
1
> 1, but P

has one less inner normal when a


1
= 1. The same thing hap-
pens at the other edge of

. Then (11.4.5) follows by comparing Theorem 10.2.12


and Proposition 10.2.17 to Figure 11 (Exercise 11.4.4).
We should also mention that P

is the polyhedron denoted

in [218, p. 28].
Before we can take a canonical singularity and make it terminal, we need to
extend the denition of crepant, which was originally dened only for resolutions
of singularities. Given normal varieties X,Y where Y is Q-Gorenstein, a morphism
f : X Y is crepant if K
X
= f

K
Y
. Then we have the following result.
554 Chapter 11. Toric Resolutions and Toric Singularities
Proposition 11.4.17. If X

has canonical singularities, then we can nd a simpli-


cial renement

of such that X

has terminal singularities and the induced


toric morphism : X

is projective and crepant. Furthermore, X

is Goren-
stein if and only if X

is Gorenstein.
Proof. Suppose that X

has canonical singularities. To measure how far X

is
from being terminal, set
t() =

max
[(

N) `vertices[.
By Proposition 11.4.12, X

has terminal singularities if and only if t() = 0.


If t() > 0, pick a non-vertex

N for some
max
. Then the star
subdivision

() satises the following:


X

has canonical singularities with t(

) = t() 1.
X

is Gorenstein if and only if X

is.
X

is crepant.
You will verify these properties in Exercise 11.4.5. Figure 12 shows what happens
for the afne toric variety of a cone R
3
with t() = 2. Here, is one of the

Figure 12. The polyedron and the star subdivision of coming from
two non-vertex lattice points of

and the star subdivision

has t(

) = 1.
By induction on t(), a sequence of these star subdivisions gives a renement

of such that X

has terminal singularities, and X

is Gorenstein if and only


if X

is. If

is simplicial, then we are done. If not, then the simplicial renement


constructed in Proposition 11.1.7 is easily seen to have the desired properties.
Terminal Singularities. Propositions 11.4.15 and 11.4.17 allow us to focus on
toric varieties with terminal singularities. We begin with the afne case.
11.4. Toric Singularities 555
Gorenstein afne toric varieties with terminal singularities can be classied in
terms of empty lattice polytopes, which are lattice polytopes whose only lattice
points are their vertices. Here is the precise result.
Proposition 11.4.18. Classifying Gorenstein afne toric varieties with terminal
singularities that come from n-dimensional cones N
R
R
n
is equivalent to
classifying (n1)-dimensional empty lattice polytopes.
Proof. First assume that U

is Gorenstein with terminal singularities. Then there


is m M such that 'm, u

` = 1 for all (1). Extend m to a basis of M and


consider the dual basis of N
R
R
n
. Since dim = n, it follows that = Cone(F),
where F R
n1
1 is the facet of

not containing the origin. Then F is a


lattice polytope of dimension n1 and is empty since U

has terminal singularities


(Proposition 11.4.12).
Conversely, given an empty lattice polytope P R
n1
of dimension n1, we
get the n-dimensional cone = Cone(P1) R
n
. In Exercise 11.4.6 you will
show that U

is Gorenstein and has terminal singularities since P is empty.


Proposition 11.4.18 has some nice consequences in the 3-dimensional case.
Proposition 11.4.19. Let X

be a 3-dimensional Gorenstein toric variety. Then:


(a) If X

is simplicial, then X

has terminal singularities X

is smooth.
(b) X

has a resolution of singularities : X

such that is projective and


crepant.
Proof. For part (a), assume that X

is simplicial with terminal singularities and


take (3). Since 2-dimensional empty lattice simplices are lattice equivalent
to the standard 2-simplex (Exercise 8.3.4), Proposition 11.4.18 implies that is
smooth. When (2), Proposition 11.4.12 implies that

is a 2-dimensional
empty lattice simplex, which as already noted is smooth. Hence X

is smooth.
For part (b), recall that X

has canonical singularities by Proposition 11.4.11.


Then Proposition 11.4.17 gives a simplicial renement

such that : X

is
projective and crepant, X

has terminal singularities, and X

is Gorenstein since
X

is. Then X

is smooth by part (a). One can also check that the renement
constructed in Proposition 11.4.17 does not affect the smooth cones of (Exer-
cise 11.4.7). Hence is the desired resolution of singularities.
Corollary 11.4.20. The toric variety of a 3-dimensional reexive polytope has a
projective crepant resolution.
Proof. This follows immediately from Proposition 11.4.19 since the toric variety
of a reexive polytope is a Gorenstein Fano variety by Theorem 8.3.4.
In the 3-dimensional case, there are also results about terminal singularities
that do not assume that the variety is Gorenstein. To state our result, we dene the
556 Chapter 11. Toric Resolutions and Toric Singularities
index of a Q-Gorenstein variety X to be the smallest positive integer r such that
rK
X
is Cartier. The X is Gorenstein if and only if it is Q-Gorenstein of index 1.
Then we can classify all 3-dimensional cones that give terminal singularities
as follows.
Theorem 11.4.21. Let N
R
R
3
be a 3-dimensional cone whose afne toric
variety U

has terminal singularities.


(a) If is simplicial, then N has a basis u
1
, u
2
, u
3
such that
= Cone(u
1
, u
2
, u
1
+ pu
2
+qu
3
),
where 0 p < q are relatively prime. Furthermore, U

is Q-Gorenstein of
index q, and
U

C
3
/
q
,
where
q
= C

[
q
= 1 acts on C
3
via (x, y, z) = (
1
x,
p
y, z).
(b) If is not simplicial, then N has a basis u
1
, u
2
, u
3
such that
= Cone(u
1
, u
2
, u
1
+u
3
, u
2
+u
3
).
Furthermore, U

is Gorenstein and U

V(xy zw) C
4
.
Proof. For part (a), rst note that

is a 3-dimensional empty lattice simplex.


By the terminal lemma [218, 1.6], N has a basis such that the vertices of

are 0, u
1
, u
2
, u
1
+ pu
2
+qu
3
, where 0 p < q are relatively prime. This gives the
desired formula for , and the quotient construction from Chapter 5 easily gives
the representation U

C
3
/
q
in the statement of the theorem (Exercise 11.4.8).
To compute the index, let m = (1, 1, p/q) relative to the dual basis of M. Then
'm, u
1
` ='m, u
2
` ='m, u
3
` = 1.
Since gcd(p, q) = 1, qm is the smallest positive multiple of m that is a lattice point,
hence the index is q. See [103, Thm. 2.2] for part (b).
In [218, 1.6], Oda discusses the terminal lemma, with references. In the
literature, the action of
q
on C
3
is often written differently. Given the action
(x, y, z) = (
1
x,
p
y, z) as in the theorem, let 0 a < q be the multiplicative
inverse of p modulo q. Using the automorphism of
q
given by
a
and
changing coordinates in C
3
, the action becomes (x, y, z) = (
a
x,
a
y, z). See
[218, 1.6] for references to the classication of 3-dimensional cones whose toric
varieties have canonical singularities. Also, [13] and the references therein discuss
the classication of empty lattice simplices in dimensions 3 and 4.
Part (b) of Theorem 11.4.21 is especially nice, since it shows that one of our
favorite examples, V(xy zw) C
4
, is even more interesting than we realized.
The singularities of a normal variety have codimension 2. In Exercise 11.4.9
you will prove that the following result, which tells us that the codimension is
higher when the singularities are terminal.
11.4. Toric Singularities 557
Proposition 11.4.22. Assume that X

has only terminal singularities.


(a) The singular locus (X

)
sing
of X

has codimension codim(X

)
sing
3.
(b) If in addition X

is Gorenstein and simplicial, then codim(X

)
sing
4.
At the end of 8.3, we noted that in dimension 3, there are 4319 isomorphism
classes of Fano toric varieties, corresponding to the 4319 classes of 3-dimensional
reexive polytopes. They all have canonical singularities by Proposition 11.4.11
since Fano varieties are Gorenstein. But only 100 of these have terminal singular-
ities by [165], and as noted in 8.3, only 18 are smooth. This indicates the special
nature of terminal singularities. See also [166].
Log Singularities. Besides terminal and canonical singularities, there are other
classes of singularities that are important for the minimal model program. We
will consider two, log canonical and Kawamata log terminal. The latter is usually
abbreviated klt. These singularities are dened for a pair (X, D), where X is a
normal variety and D is a Q-divisor with coefcients in the interval [0, 1]. See
[179, p. 98] or [194, Ch. 11] for a nice discussion of why pairs are useful. A more
recent reference is [30]. Pairs are sometimes called log pairs.
Let (X, D), D =

i
d
i
D
i
, be a pair. Also x a proper birational morphism
f : X

X such that X

is smooth and Exc( f )

i
f
1
(D
i
) is a SNC divisor. Note
that f may fail be a log resolution of D. We dene the birational transform D

of
D to be the divisor on X

dened by D

i
d
i
D

i
, where D

i
= f
1
(D
i
U) and
U X is the largest open subset where f
1
is a morphism (see also 15.4).
Let (X, D) and f : X

X be as above, and assume that K


X
+D is Q-Cartier.
By [179, Sec. 2.3], we can pick a canonical divisor K
X

such that
(11.4.6) K
X
+D

= f

(K
X
+D) +

i
a
i
E
i
,
where a
i
Q and the E
i
are the irreducible divisors lying in the exceptional locus.
This is the log version of (11.4.3).
Denition 11.4.23. Let (X, D) be a pair such that K
X
+D is Q-Cartier, and write
D =

i
d
i
D
i
with d
i
[0, 1] Q.
(a) (X, D) has log canonical singularities if there is f : X

X as above such that


the coefcients a
i
in (11.4.6) satisfy a
i
1 for all i.
(b) (X, D) has klt singularities if d
i
[0, 1) for all i and there is f : X

X as
above such that the coefcients a
i
in (11.4.6) satisfy a
i
>1 for all i.
One can show that this denition is independent of the morphism f : X

X.
In practice, one often says (X, D) is log canonical instead of (X, D) has log
canonical singularities, and (X, D) is klt has a similar meaning.
In the toric case, we consider pairs (X

, D) where D is torus-invariant.
558 Chapter 11. Toric Resolutions and Toric Singularities
Proposition 11.4.24. Let X

be a normal toric variety, and let D =

,
where d

[0, 1] Q. If K
X

+D is Q-Cartier, then:
(a) (X

, D) is log canonical.
(b) If in addition d

[0, 1) for all (1), then (X

, D) is klt.
Proof. Let be the support function of K
X

+D, so that (u

) = 1 d

for all
(1). Let : X

be a toric log resolution. For the rest of the proof,


D

will denote the divisor on X

corresponding to

(1). Thus the birational


transform of D is D

(1)
d

.
To simplify notation, let

(1) = AB, where A = (1) and B =

(1) ` A.
Since is the support function of

(K
X

+D), we obtain

(K
X

+D) =

AB
(u

)D

A
(1+d

)D

B
(u

)D

AB
D

A
d

B
(1(u

))D

= K
X

+ D

B
(1(u

))D

.
Hence
(11.4.7) K
X

+D

(K
X

+D) +

B
((u

) 1)D

.
This is the log version of Lemma 11.4.10.
To analyze the coefcients, take B =

(1) `(1). Since

renes , we
have u

for some . Thus u

(1)

, where

0. Hence
(11.4.8) (u

) =

(1)

(u

) =

(1)

(1d

).
We have d

1 by assumption, so that 1d

0. It follows that (u

) 0, and
thus the coefcients (u

) 1 in (11.4.7) are all 1. This proves that (X

, D) is
log canonical.
Now assume in addition that d

< 1 for all (1). Then 1 d

> 0 in
(11.4.8), and since the

are not all 0 (u

= 0), we conclude that (u

) > 0 for all


B. This shows that the coefcients (u

) 1 in (11.4.7) are all > 1. Hence


(X

, D) is klt, as claimed.
Here is an easy corollary that you will prove in Exercise 11.4.10.
Corollary 11.4.25. Let X

be a normal toric variety. Then:


(a) (X

) is log canonical.
(b) If X

is Q-Gorenstein, then (X

, 0) is klt.
11.4. Toric Singularities 559
Here is an example based on an observation of Chen and Shokurov [62].
Example 11.4.26. Recall from 8.3 that a projective normal variety X is Goren-
stein Fano if K
X
is ample. The log version says that a normal projective variety X
is of Fano type if there is Q-divisor D such that (K
X
+D) is Q-ample (meaning
that some positive multiple is Cartier and ample) and (X, D) is klt. Let us show
that every projective toric variety X

is of Fano type.
Since X

is projective, there is a lattice polytope P that gives an ample divisor


D
P
on X

. Multiplying P by a suitably large integer, we can assume that P has


an interior lattice point m, and translating by m (which gives a linearly equivalent
divisor), we can assume that 0 is an interior point. Since D
P
=

and
P =m M
R
[ 'm, u

` a

,
it follows that a

> 0 for all . Then set


D =

(1a

)D

,
where Q is positive and satises a

< 1 for all . Then the coefcients of D


lie in [0, 1) for all .
By construction, D satises K
X

+D = D
P
, so that K
X

+D is Q-Cartier.
Then (X

, D) is klt by Proposition 11.4.24. It follows that X

is of Fano type since


(K
X

+D) = D
P
is Q-ample.
In the minimal model program, many results assume that (X, D) is a klt pair.
See [30, Sec. 4.3] for some examples. In Chapter 15 we will see that klt pairs arise
naturally in the toric minimal model program.
Exercises for 11.4.
11.4.1. In this exercise you will supply some details omitted from the proof of (d) (a)
in Theorem 11.4.8.
(a) Prove that V() is the unique xed point of the action of T
N
on U

.
(b) Explain why T
N
is a subtorus of T
N
of codimension 1 when is a facet of .
(c) Prove that is simplicial if and only if its dimension equals the number of its facets.
11.4.2. This exercise is concerned with the proof of Proposition 11.4.12.
(a) Prove (iii) (ii) in part (a).
(b) Prove the characterization of canonical singularities given in part (b).
11.4.3. As in Proposition 11.4.15, let

= Conv( N ` 0).
(a) Prove that

is a lattice polyhedron with as recession cone.


(b) Prove that taking cones over bounded faces of

gives a fan
can
rening .
(c) Prove that
can
U

is projective. Hint: Suppose


can
comes from a facet of

. If the facet is dened by 'm, ` = a, then consider the support function on

=[
can
[ whose restriction to is given by
1
a
m.
560 Chapter 11. Toric Resolutions and Toric Singularities
11.4.4. Complete the proof of (11.4.5) sketched in Example 11.4.16.
11.4.5. Prove the properties of

() stated in the three bullets in the proof of Propo-


sition 11.4.17.
11.4.6. In the proof of Proposition 11.4.18, we saw that an elementary lattice polytope
P gives a cone of one dimension higher. Prove that U

is Gorenstein with terminal


singularities. Hint: P is the facet of

lying at height 1.
11.4.7. Let : X

be the terminalization constructed in Proposition 11.4.17.


Prove that all terminal simplices of lie in

and explain why this implies that is an


isomorphism above the Q-factorial terminal locus of X

.
11.4.8. Complete the proof of Theorem 11.4.21 by using the methods of Chapter 5 to
describe the quotient representation of U

. Hint: Lemma 5.1.1 will be useful.


11.4.9. Prove Proposition 11.4.22. Hint: Study the cones of dimension 3 using
the methods of the proof of Proposition 11.4.19.
11.4.10. Prove Corollary 11.4.25.
11.4.11. Here is a partial converse to Theorem 11.4.17. Assume that X

is Q-Gorenstein.
Prove that if has a renement

such that X

has terminal singularities and X

is crepant, then X

has canonical singularities.


11.4.12. Explain how Propositions 11.4.15 and 11.4.17 give the minimal resolution of
singularities of a toric surface X

. Draw some pictures to illustrate what is happening.


11.4.13. Let X

be a Q-Gorenstein toric variety with terminal singularities.


(a) Give an example where X

has a crepant resolution of singularities. Hint: Try one of


our favorite examples.
(b) If X

is simplicial but not smooth, then prove that all resolutions of singularities are
not crepant.
(c) If X

is not smooth, then prove that all SNC resolutions of singularities are not crepant.
Chapter 12
The Topology of
Toric Varieties
In this chapter, we will study some topological invariants of a toric variety X,
always for the classical topology. The fundamental group
1
(X) is studied in
12.1, and 12.2 addresses the moment map and alternative topological mod-
els of X. Methods for computing the singular cohomology groups H
k
(X, Z) are
discussed in 12.3. A complete description of the cup product ring structure on
H

(X, Q) =

k
H
k
(X, Q) for simplicial complete X is developed in 12.4 using
the equivariant cohomology of X for the action of T
N
. The chapter concludes with
12.5, where we consider the Chow ring and intersection cohomology.
Our goal is to understand the information these invariants provide about the
topology and geometry of toric varieties and their applications to polytopes. We
will freely use the denitions and various properties of homotopy, homology, and
cohomology groups, referring to [135], [210], and [255] as our primary references.
12.1. The Fundamental Group
Topology of Tori. The most basic toric varieties are the tori T
N
, and their topol-
ogy is correspondingly simple to understand. A choice of basis in N determines a
homeomorphism and isomorphism of groups T
N
(C

)
n
. Let S
1
denote the unit
circle in C

. The usual polar coordinate system in each factor gives a homeomor-


phism and isomorphism of multiplicative groups
(C

)
n
(R
>0
)
n
(S
1
)
n
.
In the following, we will use the notation S
N
for the compact real torus in T
N
.
561
562 Chapter 12. The Topology of Toric Varieties
Proposition 12.1.1. Let N be a lattice of rank n. Then:
(a) S
N
is a deformation retract of T
N
(C

)
n
.
(b) There is an isomorphism

1
(T
N
) Z
n
.
Proof. Since R
>0
is contractible to the point 1 R
>0
, we easily construct the
deformation retraction. Part (b) follows from the standard facts
1
(S
1
) Z and

1
(X Y)
1
(X)
1
(Y).
There is also a more intrinsic way to understand this isomorphism. Recall that
each u N denes a one-parameter subgroup
u
: C

T
N
. Restricting
u
to
S
1
C

gives a closed path in T


N
, whose homotopy class represents an element of

1
(T
N
). In Exercise 12.1.1, you will show that the resulting map N
1
(T
N
) is an
isomorphism. Hence we will usually use this intrinsic form of Proposition 12.1.1:
(12.1.1)
1
(T
N
) N.
To determine the fundamental group of other toric varieties, we will use the
fact that the algebraic condition of normality has a very important, though quite
nontrivial, topological consequence.
Topological Implications of Normality. To see the pattern, we consider two toric
examples we have encountered before.
Example 12.1.2. In Example 3.A.2, we studied the nonnormal afne toric surface
Y
A
= V(y
2
x
2
z) C
3
with A = e
1
, e
1
+e
2
, 2e
2
Z
2
. Note that the z-axis is clearly contained in Y
A
.
In Exercise 12.1.2 you will show that the z-axis is precisely the singular locus of
Y
A
. Letting p = (0, 0, z) with z = 0, you will also show that if U is any sufciently
Figure 1. The surface V(y
2
x
2
z)
small open neighborhood of p in Y
A
, then U `(U Sing(Y
A
)) has two connected
components, as suggested by the picture of the real points of the surface in Figure 1.
These are called the branches of Y
A
at p.
12.1. The Fundamental Group 563
Example 12.1.3. Consider the quadric

C
2
= V(xz y
2
) C
3
, the normal afne
toric surface from the cone = Cone(e
2
, 2e
1
e
2
). The only singularity of this
surface is the point p = (0, 0, 0). From our standard parametrization,

C
2
is the
image of the map
: C
2
C
3
(t
1
, t
2
) (t
1
, t
1
t
2
, t
1
t
2
2
).
The singular point p is the image of the line V(t
1
) in C
2
. If we remove that line,
then C
2
`V(t
1
) is still connected, so

C
2
`p = (C
2
`V(t
1
))
is the continuous image of a connected set, hence connected. The same will be true
if we intersect

C
2
`p with any connected open neighborhood U of p in C
3
.
These examples illustrate a general phenomenon. While a nonnormal variety
can have more than one branch at a singular point, at each point x of a normal
variety X, singular or nonsingular, X is locally irreducible or unibranch in the
following sense.
Proposition 12.1.4. Let X be a normal variety and take x X. Then there is a
basis V

[ A of open neighborhoods of x in X such that V

`(V

Sing(X))
is connected for all .
This is a topological version of the fundamental result known as Zariskis main
theorem. We refer to [208, III.9] for algebraic versions of the result and their
relation to Proposition 12.1.4.
Comparing
1
(X

) and
1
(T
N
). Let be a fan in N
R
and consider the normal
toric variety X

and the inclusion i : T


N
X

. There is a general fact that applies


to the induced map i

:
1
(T
N
)
1
(X

) in this situation (see [109]).


Theorem 12.1.5. Let X be a normal variety and let i : U X be the inclusion of
an open subvariety. Then the induced map i

:
1
(U)
1
(X) is surjective.
The proof of the theorem will consist of the following two lemmas. We recall
that any suitably nice topological space X has a universal covering space, a simply-
connected space

X with an (unramied) covering map p :

X X. The universal
cover of a variety over C is not always a variety, but it inherits the structure of a
complex analytic space from X, which will be sufcient for our purposes.
Lemma 12.1.6. If X is a normal variety and i : U X is the inclusion of an open
subvariety U, then U
X

X is path-connected.
564 Chapter 12. The Topology of Toric Varieties
Proof. Let U = X `Y, where Y is Zariski closed in X. Then U is connected and
path-connected in the classical topology. By Proposition 12.1.4 and the construc-
tion of the universal covering space,

X is also locally irreducible. Since it is also
connected, it is irreducible as an analytic space. Hence

X ` p
1
(Y) = p
1
(X `Y) = p
1
(U)
is also connected and path-connected in the classical topology. The analytic space
U
X

X =(u, x) [ p( x) = u
is essentially the graph of p restricted to p
1
(U). Since p is continuous, this is also
connected and path-connected.
Lemma 12.1.7. Let X, Z be topological spaces and assume X has a universal cov-
ering space p :

X X. Let f : Z X be continuous. If Z
X

X is path-connected,
then f

:
1
(Z)
1
(X) is a surjection.
Proof. We must show that for each homotopy class []
1
(X), there exists a
closed curve in Z such that f

([]) = [], or equivalently that f and are


homotopic in X. Pick a base point x
0
on X with x
0
= f (z
0
). Let x
0
p
1
(x
0
) in

X, and lift to in

X starting from the point x
0
. The nal point of will be some
x
1
p
1
(x
0
). Hence both (z
0
, x
0
) and (z
0
, x
1
) are points in
Z
X

X =(z, x) Z

X [ f (z) = p( x).
By hypothesis, we can nd a path : [0, 1] Z
X

X with (0) = (z
0
, x
0
) and
(1) = (z
0
, x
1
). Let p
1
, p
2
be the projections from Z

X to the two factors. The


projection p
2
is homotopic to since

X is simply-connected. Hence p p
2

is homotopic to . Moreover p
1
= is a loop in Z such that f = p p
2
.
This shows that f

is surjective.
Because of Theorem 12.1.5, to determine
1
(X

) we must determine the kernel


of the homomorphism i

:
1
(T
N
)
1
(X

), and then

1
(X

)
1
(T
N
)/ker(i

) N/ker(i

).
In particular, the fundamental group of a normal toric variety is always a nitely-
generated abelian group. Here is one case where the kernel is easy to see.
Proposition 12.1.8. Let = Cone(u

) be a 1-dimensional cone with u

the primi-
tive ray generator in N. Then ker(i

:
1
(T
N
)
1
(U

)) = N

=Zu

.
Proof. We use (12.1.1) and view i

as a map from N to
1
(U

). Applying Propo-
sition 3.2.2 to the cone , we see that for all u N,
(12.1.2) u lim
z0

u
(z) exists in U

.
12.1. The Fundamental Group 565
Since u

, the limit point lim


z0

u
(z) =

is in U

. Restrict z to S
1
C

and
take t [0, 1]. Then the map
(t, z) : [0, 1] S
1
X

(t, z)

u
((1t)z) 0 t < 1

t = 1
is continuous. Moreover gives a homotopy from a loop representing the homo-
topy class on T
N
corresponding to u

N to a constant path at the distinguished


point

, where

is dened on page 116. This shows N

ker(i

). The
opposite inclusion follows since if ker(i

) contained any u N not in N

, then
rank
1
(U

) = rank(N/ker(i

))
would be n2 or smaller. But U

= (C

)
n1
C, so
1
(U

) Z
n1
.
The Afne Case. The idea behind the homotopy constructed in the proof of the
last proposition can be generalized to prove the following fact.
Proposition 12.1.9. Let be a strongly convex rational polyhedral cone in N
R
.
(a) The torus orbit O() is a T
N
-equivariant deformation retract of U

.
(b) There are isomorphisms

1
(U

)
1
(T
N()
) N(),
where N() = N/N

as usual.
Proof. Part (b) follows immediately from part (a) because O() is equal to the
torus T
N()
by Lemma 3.2.5. To prove part (a), view the points x U

as semigroup
homomorphisms x : S

C, where S

M. Choose any lattice point u in


Relint() N and dene the map
: U

[0, 1] U

by
(x, t)(m) =

m
(
u
(1t))x(m) if 0 t < 1

(m)x(m) if t = 1,
where

is the distinguished point corresponding to .


For all (x, t), it is not hard to see that (x, t) is a semigroup homomorphism
fromS

to C, hence a point of U

(Exercise 12.1.3). Moreover, (x, t) is continuous


in x and t.
We have
m
(
u
(1 t)) = (1 t)
/m,u)
, so (x, 0) is the identity map on U

.
Moreover if m

M,
m
(
u
(1t)) = 1 for all 0 t < 1. From Lemma 3.2.5,
O() = : S

C [ (m) = 0 m

M.
Chasing the denitions, one sees that (x, t) = x for all t when x O().
566 Chapter 12. The Topology of Toric Varieties
Finally,
lim
t1
(1t)
/m,u)
=

1 if m S

M
0 otherwise,
which implies that (x, 1) O() for all x. Combining all of this, we see that is
a deformation retraction. You will show T
N
-equivariance in Exercise 12.1.3.
The General Case. For a general fan we will use the last proposition and the
Van Kampen theorem to complete the computation of
1
(X

).
Theorem 12.1.10. Let be a fan in N
R
and let N

be the sublattice of N generated


by [[ N. Then
1
(X

) N/N

.
Proof. We will prove the theorem by induction on the number of cones in . If
consists of a single cone 0, then the claim follows from (12.1.1).
Now assume the result has been proved for all fans with k 1 cones or fewer
and let contain k cones. Pick any cone of maximal dimension in , and let

= `. Then
X

= X

,
with X

= X

for the fan

consisting of the proper faces of . Since X

,
U

, and X

are all path connected, we can apply the Van Kampen theorem [135,
Thm. 1.20] to determine
1
(X

). From the diagram of inclusions


X

j
1

i
1

i
2

j
2

we obtain the corresponding diagram of fundamental groups

1
(X

)
j
1

1
(X

)
i
1

i
2


1
(X

).

1
(U

)
j
2

The Van Kampen theorem identies


1
(X

) as the free product of


1
(X

) and

1
(U

), modulo the normal subgroup generated by all elements of the form


( j
1
i
1
)

([])(( j
2
i
2
)

([]))
1
,
12.1. The Fundamental Group 567
where []
1
(X

). Our induction hypothesis gives isomorphisms

1
(X

) N/N

1
(X

) N/N

1
(U

) N/N

.
By using presentations of these groups in terms of generators and relations, it is
easy to see that

1
(X

) N/(N

+N

) = N/N

.
You will complete the details in Exercise 12.1.4.
Theorem 12.1.10 implies that X

is simply connected if and only if the support


of contains a basis for N. This is the case, for instance, if contains an n-
dimensional cone. It follows that a simply connected toric variety has no torus
factors. The converse of this assertion is not true, though, as the following example
shows.
Example 12.1.11. Let N have rank 2, and let be the fan consisting of the cones
0, Cone(e
1
), Cone(e
1
+de
2
).
The corresponding toric variety X

is the complement of the origin in the rational


normal cone

C
d
. Clearly, N

=Ze
1
+dZe
2
and
1
(X

) N/N

Z/dZ.
Exercise 12.1.6 shows that for any nitely generated abelian group G, there
exists a normal toric variety X

with
1
(X

) G. Some information on the higher


homotopy group
2
(X

) will be obtained later in Exercise 12.3.10.


Exercises for 12.1.
12.1.1. Let [] be the homotopy class of a closed path . Show that the map
N
1
(T
N
)
u [
u
[
S
1 ]
is an isomorphism of groups.
12.1.2. In this exercise you will verify the claims made in Example 12.1.2 about the surface
Y
A
= V(y
2
x
2
z) C
3
, where A =e
1
, e
1
+e
2
, 2e
2
.
(a) Show that the singular locus of Y
A
is the z-axis.
(b) Show that if p is any point of Y
A
with z = 0 and V is a sufciently small open neigh-
borhood of p in Y
A
, then V ` (V Sing(Y
A
)) has two connected components.
12.1.3. In this exercise, you will complete the details of the proof of Proposition 12.1.9.
(a) Show that for all x U

and t [0, 1], (x, t) : S

C is a semigroup homomorphism,
hence a point of U

.
(b) Verify the claim in the proof of Proposition 12.1.9 that (x, t) is continuous in x and t.
(c) Show that the deformation retraction is equivariant for the T
N
-action.
568 Chapter 12. The Topology of Toric Varieties
12.1.4. Complete the details of the argument using the Van Kampen theorem in the proof
of Theorem 12.1.10.
12.1.5. Let be the fan in N
R
R
2
with cones
0, Cone(e
1
+e
2
), Cone(e
1
e
2
), Cone(e
1
+e
2
), Cone(e
1
e
2
).
Show that
1
(X

) Z/2Z.
12.1.6. Show that given any nitely generated abelian group G, there exists a normal toric
variety whose fundamental group is isomorphic to G.
12.1.7. Prove that
1
(X

) is nite if and only if X

has no torus factors.


12.2. The Moment Map
The compact real torus S
N
(S
1
)
n
is a subgroup of T
N
and hence also acts on the
toric variety X

. We will show rst that the quotient space X

/S
N
can be identied
with a certain subset (X

)
0
of X

.
The Nonnegative Part of a Toric Variety. Recall the interpretation of points of
an afne toric variety U

as semigroup homomorphisms : S

C. We can
dene subsets of U

by placing restrictions on the image of . For instance, the


nonnegative part of U

is dened formally as
(U

)
0
= Hom
Z
(S

, R
0
).
The absolute value map z [z[ on C gives a retraction of C onto R
0
. So there
is a corresponding retraction U

(U

)
0
dened by mapping [[. If X

is the toric variety of a fan , it is easy to check that the (U

)
0
for glue
together to form a closed subset (X

)
0
of X

in the classical topology. Moreover


the retraction maps glue properly to give a retraction
X

(X

)
0
.
By considering semigroup homomorphisms with images in R or R
>0
, we can
dene real or positive real points of toric varieties as well. The real points of toric
varieties are discussed for instance in [254].
If : X Y is a toric morphism, then the above discussion implies that
restricts to a map

0
: X
0
Y
0
.
Moreover,
0
ts in a commutative diagram
(12.2.1)
X

X
0

Y
0
12.2. The Moment Map 569
where the vertical arrows are the above retractions (Exercise 12.2.1). This implies
that a character
m
of T
N
(C

)
n
maps (T
N
)
0
(R
>0
)
n
to (C

)
0
=R
>0
.
We can also apply (12.2.1) to the quotient construction of toric varieties from
Chapter 5. By Proposition 5.1.9, the quotient map
: C
(1)
`Z() X

.
is a toric morphism, so that
0
is dened.
Proposition 12.2.1. Let X

be a normal toric variety without torus factors. Then

0
: (C
(1)
`Z())
0
(X

)
0
is surjective.
Proof. Theorem 5.1.11 implies that is a good categorical quotient. Hence is
surjective by Theorem 5.0.6. Surjectivity of
0
follows from the commutativity
of the diagram (12.2.1).
Here is one way to use Proposition 12.2.1.
Example 12.2.2. Consider X

P
1
P
1
for the fan in N
R
R
2
given in Exam-
ple 3.1.12. By Example 5.1.8, we have
P
1
P
1
=U/(C

)
2
,
where U =C
4
`(((0, 0)C
2
) (C
2
(0, 0))) and (C

)
2
acts via
(, ) (a, b, c, d) = (a, b, c, d).
By Proposition 12.2.1, a point in (P
1
P
1
)
0
comes from (a, b, c, d) U R
4
0
.
Since a, b 0 cannot both vanish, we can rescale by a positive number so that
a+b = 1. Doing the same for c, d, we see that (P
1
P
1
)
0
can be written as
(P
1
P
1
)
0
=(a, b, c, d) R
4
: a, b, c, d 0 and a+b = c +d = 1,
which is a square in a two-dimensional afne subspace in R
4
. Note that this is
the same (combinatorially, at least) as a plane polygon P whose normal fan
P
coincides with the original fan . We will see shortly that this is no coincidence.
In homogeneous coordinates, the retraction P
1
P
1
(P
1
P
1
)
0
is given by
(a, b, c, d)

[a[
[a[ +[b[
,
[b[
[a[ +[b[
,
[c[
[c[ +[d[
,
[d[
[c[ +[d[

.
Note the structure of the bers of this map. Over the four corners of the square
(that is (a, b) = (1, 0) or (0, 1) and similarly for (c, d)), the ber consists of a single
point. On an edge but not at a vertex, the ber consists of a copy of S
1
. Finally at
an interior point of the square, the ber consists of a copy of S
1
S
1
= T
2
. Thus
P
1
P
1
has a stratication by ber bundles with compact real torus bers.
570 Chapter 12. The Topology of Toric Varieties
The Quotient of X

by S
N
. Now we consider the quotient of X

by the compact
real torus S
N
. To prepare for our next result, note that the isomorphism C

R
>0
S
1
used in 12.1, composed with the real logarithm map R
>0
R on the
rst factor, yields an isomorphism C

RS
1
. In Exercise 12.2.2, you will show
that S
N
can be identied with Hom
Z
(M, S
1
) Hom
Z
(M, C

), where S
1
is the unit
circle in C

. Putting all of this together, we obtain isomorphisms


(12.2.2) T
N
= Hom
Z
(M, C

) Hom
Z
(M, R) Hom
Z
(M, S
1
) N
R
S
N
.
The structure of the quotient X

/S
N
is as follows.
Proposition 12.2.3. Let X

be a normal toric variety.


(a) The retraction X

(X

)
0
induces a homeomorphism X

/S
N

= (X

)
0
.
(b) For each cone in , the ber of X

(X

)
0
over a point in O()
0
can
be identied with S
N()
, a compact real torus of dimension ndim .
Proof. For part (a), consider how S
N
acts on each torus orbit O(). Recall from
Lemma 3.2.5 that O() Hom
Z
(

M, C

) T
N()
, where N() = N/N

is the
dual lattice of

M. Then (12.2.2) implies that


O() N()
R
S
N()
.
Using the real logarithm again, O() retracts to
O()
0
= Hom
Z
(

M, R
>0
) Hom
Z
(

M, R) = N()
R
.
However, S
N
acts on O() via the compact real torus S
N()
in T
N()
. As a result,
O()/S
N

= N()
R

= O()
0
. The assertion for X

follows. Part (b) follows by


similar reasoning.
Example 12.2.4. The toric variety P
1
is homeomorphic to the real 2-sphere S
2
.
The torus S
N
in this case is S
1
, a circle acting on P
1
xing two points shown as
the north and south poles in Figure 2. Every other point has a circle as orbit. The
quotient space is homeomorphic to a closed interval.

Figure 2. The quotient P


1
/S
1

=(P
1
)
0
12.2. The Moment Map 571
Note also that the map P
1
P
1
(P
1
P
1
)
0
from Example 12.2.2 can be
visualized as the Cartesian product of two copies of this picture.
The Moment Map. Recall that in Example 12.2.2 above we saw that (P
1
P
1
)
0
was combinatorially the same as a polygon with normal fan equal to the standard
fan for P
1
P
1
. We now turn to a general connection between (X
P
)
0
and the
polytope P.
Let P be a full dimensional lattice polytope in M
R
. Corresponding to P, we
have the normal fan
P
and the toric variety X
P
= X

P
. Replacing P by a multiple
P if necessary, we will assume that PM =m
0
, . . . , m
s
denes an embedding
: X
P
P
s
x (
m
0
(x), . . . ,
ms
(x)).
Following the online version of [254], we dene the algebraic moment map by
(12.2.3)
f : X
P
M
R
x
1

mPM
[
m
(x)[

mPM
[
m
(x)[m.
The symplectic moment map corresponding to the action of the compact real torus
S
N
(as dened in symplectic geometry) is the closely related map
(12.2.4)
: X
P
M
R
x
1

mPM
[
m
(x)[
2

mPM
[
m
(x)[
2
m.
Note that f and are invariant under the action of S
N
T
N
. The behavior we saw
in Example 12.2.2 is a special case of the next result.
Theorem 12.2.5. Let P be a full dimensional lattice polytope in M
R
. The restricted
algebraic moment map
f = f [
(X
P
)
0
: (X
P
)
0
M
R
,
is a homeomorphism from (X
P
)
0
to P.
Proof. The characters
m
for m P M map x (X
P
)
0
to nonnegative real
numbers, so that [
m
(x)[ =
m
(x). Rescaling as in Example 12.2.2 gives (x) =
(a
0
, . . . , a
s
) P
s
, where a
0
+ +a
s
= 1 and a
i
0 for all i. This implies that
f (x) = a
0
m
0
+ +a
s
m
s
.
It is clear from (12.2.3) that f is continuous. We will show that f is bijec-
tive and leave the verication that f
1
is continuous as an exercise. The proof
of bijectivity will be accomplished by showing that f maps each torus orbit in X
P
572 Chapter 12. The Topology of Toric Varieties
bijectively to the relative interior of the corresponding face of P. We will give the
details for T
N
(C

)
n
in X
P
and show that (T
N
)
0
maps bijectively to the interior
of P. The result for the other orbits will then follow by similar reasoning.
First we establish surjectivity. Fixing an isomorphism M Z
n
, write m
i
=
(m
i1
, . . . , m
in
) where m
i j
Z for all i, j. Given v in the interior of P, there are
in general many different ways to write v as a convex linear combination of the
m
i
. If a = (a
0
, . . . , a
s
) is one vector of coefcients in such a combination and
a

= (a

0
, . . . , a

s
) is a second, then
v = a
0
m
0
+ +a
s
m
s
= a

0
m
0
+ +a

s
m
s
with a
i
, a

i
0 and a
0
+ +a
s
= a

0
+ +a

s
= 1. It follows that the b
i
= a

i
a
i
satisfy the linear equations b
0
m
0
+ +b
s
m
s
= 0 and b
0
+ +b
s
= 0. With the
notation
W
P
=(b
0
, . . . , b
s
) R
s+1
[ b
0
m
0
+ +b
s
m
s
= 0 and b
0
+ +b
s
= 0,
the vectors a

= (a

0
, . . . , a

s
) satisfying v =

s
i=0
a

i
m
i
are precisely the elements of
(a+W
P
) R
s+1
0
.
Our proof will show, in fact, that v can be written as v = a
0
m
0
+ + a
s
m
s
,
where a
i
=
m
i
(x) where > 0 and x = (x
1
, . . . , x
n
) T
N
with x
i
> 0 for all i.
This will show the surjectivity of f as a map from the positive real points in T
N
in
X
P
to the interior of P. The particular representation of v we want will come from
minimizing a certain function on (a+W
P
) R
s+1
0
.
First note that g(x) = xlog(x) x can be dened for all x 0 in R since
LH opitals rule implies that lim
x0
+ g(x) = 0. Using g, we dene
G : (R
0
)
s+1
R
(x
0
, . . . , x
s
) g(x
0
) + +g(x
s
).
Since g

(x) =
1
x
, the Hessian (second derivative) of G is positive denite at any
point (x
0
, . . . , x
s
) with x
i
> 0 for all i. In other words, G is concave up at all such
points. You will show in Exercise 12.2.7 that G, restricted to (a+W
P
) R
s+1
0
, has
a unique critical point a = ( a
0
, . . . , a
s
), necessarily a minimum.
We claim that a
i
> 0 for all i. If not, then I = i [ a
i
= 0 is nonempty. In
Exercise 12.2.7, you will show that there exists b W
P
satisfying b
i
> 0 for all
i I. Let (t) = a+tb be the line through a with direction vector b, and let
i
(t)
be the components. Then 0 <
i
(t) < 1 for all i and all t in some sufciently small
open interval (0, ). It follows that
lim
t0
+
d
dt
(G)(t) = lim
t0
+
s

i=0
b
i
log( a
i
+tb
i
) =.
But this contradicts the concavity properties of G. Hence a must have all nonzero
components.
12.2. The Moment Map 573
For an arbitrary b in W
P
, let (t) = a +bt and consider the restriction of G to
this line. Since a
i
>0 for all i, we have 0 <
i
(t) <1 when t lies in some symmetric
interval about 0. Then (G)(t) has a local minimum at t =0, so by the chain rule,
0 =
d
dt
(G)[
t=0
=
s

i=0
b
i
log( a
i
).
Since this is true for all b W
P
, linear algebra implies that the overdetermined
system of linear equations
(12.2.5)
log( a
0
) = m
01
y
1
+ +m
0n
y
n
+c
.
.
.
log( a
s
) = m
s1
y
1
+ +m
sn
y
n
+c
in n+1 variables (y
1
, . . . , y
n
, c) has a real solution. Let y
i
= log(x
i
) and c = log()
where x
i
, are real and positive. Exponentiating, a
i
= x
m
i1
1
x
m
in
n
=
m
i
(x) for
all i = 0, . . . , s. Since

s
i=0
a
i
= 1, the constant must be
=
1

s
i=0

m
i
(x)
.
This completes the proof of surjectivity of the restricted algebraic moment map.
The injectivity of
f : (X
P
)
0
M
R
is now a consequence of the constructions already made. Suppose v is in the in-
terior of P and v = f (x) for x a positive real point in the T
N
in X
P
. Then the
denition of f gives v = f (x) = a
0
m
0
+ + a
s
m
s
with a
i
= x
m
i1
1
x
m
in
n
and
= (

s
i=0

m
i
(x))
1
. Writing y
i
= log(x
i
) and c = log(), we have a system
log(a
0
) = m
01
y
1
+ +m
0n
y
n
+c
.
.
.
log(a
s
) = m
s1
y
1
+ +m
sn
y
n
+c
of the same form as (12.2.5). Since the system is consistent, we must have
(12.2.6) 0 =
s

i=0
b
i
log(a
i
)
for all b W
P
. However, by the same computations done before, the sum on the
right of (12.2.6) is the derivative at t = 0 of (G)(t), where (t) = a+bt. By the
concavity of G, this implies that a = a, the unique critical point of G on a+W
P
. But
then, since there are n linearly independent vectors among the m
i
, it follows that the
a
i
=x
m
i1
1
x
m
in
n
are uniquely determined. Since we assume that
m
for mPM
dene an embedding of X
P
, this shows that x is also uniquely determined.
574 Chapter 12. The Topology of Toric Varieties
This proof basically follows the presentation in [93, VII.7]; [105] gives a differ-
ent argument. The result of Theorem 12.2.5 remains true if we replace the algebraic
moment map by the symplectic moment map from (12.2.4) (Exercise 12.2.8).
Topological Models of Toric Varieties. Combining Theorems 12.2.3 and 12.2.5
gives a way to understand the underlying topological space of the projective toric
variety X
P
of a polytope P, similar to what we observed in Example 12.2.2. Indeed,
it is not difcult to see that there is an S
N
-equivariant homeomorphism
(12.2.7) X
P

= (S
N
P)/ ,
where two points (s
1
, x
1
) and (s
2
, x
2
) are identied if x
1
= x
2
, and s
1
and s
2
are
congruent modulo the subtorus S
N
of S
N
for the cone
P
corresponding to the
minimal face of P containing x
1
(Exercise 12.2.9).
An analogous topological model can be given for any complete normal toric
variety X

using the unit ball B


n
. Namely, a complete fan in N
R
determines
a spherical complex C

on the unit sphere S


n
in N
R
by intersecting each cone
with the sphere. Then each determines a spherical dual in B
n
as follows.
By a process analogous to that described in Exercise 11.1.10, one constructs a
barycentric subdivision of C

. If =0 then =B
n
. Otherwise, is the union of
all spherical simplices in the barycentric subdivision whose vertices are barycenters
of S
n
with . Then
(12.2.8) X

= (S
N
B
n
)/ ,
where two points (s
1
, x
1
) and (s
2
, x
2
) are identied if and only if x
1
= x
2
and s
1
is
congruent to s
2
modulo S
N
for the unique such that x
1
is in the relative interior
of . See [162], where this construction of a space homeomorphic to X

is attrib-
uted to MacPherson, and [97]. These references also discuss generalizations for
toric varieties associated to noncomplete fans. Some recent work in toric topology
essentially takes this topological construction as the denition of a toric variety.
The article [59] gives a nice overview of toric topology.
Symplectic Geometry and Toric Varieties. We conclude with a brief discussion
of the relations between moment maps, symplectic geometry, and toric varieties,
without proofs. In symplectic geometry, one studies Hamiltonian actions of a com-
pact connected Lie group G
R
on a symplectic manifold X. Such an action has a
symplectic moment map
: X g

R
,
where g
R
is the Lie algebra of G
R
. For example, the action of S
N
on the toric
variety X
P
gives the symplectic moment map (12.2.4). This follows because the
the dual of the Lie algebra of S
N
is N

R
= M
R
.
Another example comes from quotient construction of a projective simplicial
toric variety
X

(C
(1)
`Z())/G
12.2. The Moment Map 575
from Chapter 5. The maximal compact subgroup of G = Hom
Z
(Cl(X

), C

) is
G
R
=Hom
Z
(Cl(X

), S
1
) and the Lie algebra of G
R
can be identied with Cl(X

)
R
.
Here, the symplectic moment map

: C
(1)
Cl(X

)
R
factors as
C
(1)

R
(1)

Cl(X

)
R
.
The map is given by
(z
1
, . . . , z
r
) =
1
2
([z
1
[
2
, . . . , [z
r
[
2
),
and the map comes from the exact sequence
0 M
R
R
(1)

Cl(X

)
R
0
obtained by tensoring (5.1.1) by R.
Let [D] be the class of an ample divisor on a complete simplicial toric variety
X

. Then
1

([D]) C
(1)
`Z() and results of Guillemin imply that the natural
map
(12.2.9)
1

([D])/G
R
X

is a diffeomorphism. The divisor D determines a symplectic structure on X

. The
natural symplectic structure fromC
(1)
descends to the quotient
1

([D])/G
R
and
the diffeomorphism in (12.2.9) preserves the cohomology classes of the symplectic
forms. Discussions of these results and references to detailed proofs can be found
in [66, 4].
Example 12.2.6. Let X

= P
n
, and recall that Cl(P
n
)
R
R. The symplectic mo-
ment map

above is given by

(z
1
, . . . , z
n+1
) =
1
2
n+1

i=1
[z
i
[
2
.
The class of an ample [D] corresponds to a positive real value, so that
1

([D]) is
diffeomorphic to S
2n+1
. The group G
R
S
1
in this case, and the diffeomorphism
(12.2.9) is
S
2n+1
/S
1

P
n
.
We obtain the identication of P
n
as the base space of the Hopf bration with total
space S
2n+1
and ber S
1
[135, p. 337].
For symplectic geometers, (12.2.9) shows that toric varieties can be dened by
a construction known as symplectic reduction.
Results along the lines of Theorem 12.2.5 were originally developed in a more
general setting. A compact connected real 2n-dimensional symplectic manifold
(M, ) is said to be toric if it has an effective Hamiltonian (S
1
)
n
action. Results
576 Chapter 12. The Topology of Toric Varieties
of Atiyah, Guillemin and Sternberg show that the image of the symplectic moment
map of such a manifold is a polytope P in R
n
. Moreover, the polytopes that appear
have been characterized by Delzant in [81]. They are polytopes having n edges
incident at each vertex p, of the form p+tu
i
for some u
i
Z
n
forming a Z-basis
of Z
n
. The vertices need not be lattice points, but if they are then P is a smooth
polytope. In any case, M is diffeomorphic to a smooth projective toric variety.
Exercises for 12.2.
12.2.1. This exercise supplies some details for the construction of the nonnegative part of
a toric variety and the behavior of a toric morphism on the nonnegative part.
(a) Let X

be a normal toric variety. Show that the (U

)
0
for all in glue together to
form a closed subset (X

)
0
of X

(in the classical topology).


(b) Show that the retraction maps U

(U

)
0
for glue together properly to give
a retraction X

(X

)
0
.
(c) Let : X Y be a toric morphism. Show that restricts to a map
0
: X
0
Y
0
commuting with the retractions on X, Y as in (12.2.1).
12.2.2. Let M, N be dual lattices Show that the compact torus S
N
T
N
is identied with
Hom
Z
(M, S
1
) Hom
Z
(M, C

).
12.2.3. A generating set m
1
, . . . , m
s
of S

gives an embedding U

C
s
. Prove that
(U

)
0
=U

R
s
0
. Hint: See the proof of Proposition 1.3.1.
12.2.4. Check the claims about the bers of the retraction P
1
P
1
(P
1
P
1
)
0
made in
Example 12.2.2.
12.2.5. Determine the image of the algebraic moment maps for each of the following toric
varieties directly from the associated polytope P without using Theorem 12.2.5.
(a) The projective plane P
2
, with P = Conv(0, e
1
, e
2
). Generalize to P
n
.
(b) The rational normal scroll X
P
for P = Conv(0, 3e
1
, e
2
, e
1
e
2
) (isomorphic to the
Hirzebruch surface H
2
).
12.2.6. Complete the proof that f in Theorem12.2.5 is a homeomorphismby showing that
f
1
is continuous.
12.2.7. This exercise concerns some details in the proof of Theorem 12.2.5. Let G =

s
i=0
x
i
log(x
i
)x
i
be the function on R
s+1
dened in that proof. Recall that G has positive
denite Hessian (second derivative) at all points with all positive coordinates.
(a) Show that if W is a translate of a linear subspace of R
s+1
, and W
0
is the subset
of W consisting of points with all coordinates nonnegative, then G[
W
0
has a unique
minimum when W
0
= . Hint: Argue by contradiction. Let a and b be two minima
and consider G restricted to the line containing a and b.
(b) Show that if G attains its minimum on (a +W
P
) R
s+1
0
at a such that I =i [ a
i
= 0
is nonempty, then there exists b W
P
with b
i
> 0 for all i I. Hint: Recall that the
point v is assumed to lie in the interior of P.
12.2.8. Show that the result of Theorem 12.2.5 is still true if we replace the algebraic
moment map by the symplectic moment map from (12.2.4).
12.2.9. Use Theorems 12.2.5 and 12.2.3 to construct a homeomorphism (12.2.7).
12.3. Singular Cohomology of Toric Varieties 577
12.3. Singular Cohomology of Toric Varieties
In this section, we will study the singular cohomology groups of a toric variety
X

. We rst describe these groups using the singular cohomology of the afne
toric varieties U

for
max
. We then give a different description that uses
the singular cohomology of the torus orbits O() for . In both cases, the
machinery of spectral sequences (see Appendix C) will establish the connection.
The Picard Group of a Toric Variety and H
2
. We begin with a lovely application
of spectral sequences that connects the Picard group of a toric variety to H
2
(X

, Z).
Let be a fan in N
R
R
n
. In Chapter 4, we constructed the map

i
max
M/M(
i
)

i<j
M/M(
i

j
)
(m
i
)
i
(m
i
m
j
)
i<j
,
where M() =

M. Proposition 4.2.9 provides a natural isomorphism


CDiv
T
N
(X

) ker

i
M/M(
i
)

i<j
M/M(
i

j
)

,
where CDiv
T
N
(X

) is the group of torus-invariant Cartier divisors on X

. Then
Theorem 4.2.1 relates this to Pic(X

) via the exact sequence


M CDiv
T
N
(X

) Pic(X

) 0.
The map MCDiv
T
N
(X

) is given by mdiv(
m
). Our goal is to relate Pic(X

)
to the topological object H
2
(X

, Z).
We begin by computing the singular cohomology of an afne toric variety.
Proposition 12.3.1. Let be a cone in N
R
. Then
H

(U

, Z) H

(T
N()
, Z)

M().
Proof. By Proposition 12.1.9, T
N()
is a deformation retract of U

, and this implies


the rst isomorphism. The cohomology of a torus was given in Example 9.0.12 and
the second isomorphism follows from the duality of N N() and M() M.
The Picard group relates to the singular cohomology of X

as follows.
Theorem 12.3.2. Let be a fan in N
R
R
n
with all maximal cones n-dimensional.
Then
Pic(X

) H
2
(X

, Z).
Proof. Consider the open cover of X

given by
U =U

max
=U

(n)
.
As noted in 9.0, for the spaces we are considering, singular cohomology with
coefcients in Z is the sheaf cohomology of the constant sheaf (in the classical
578 Chapter 12. The Topology of Toric Varieties
topology) given by Z. Hence the spectral sequence of the covering U (see (9.0.10)
and Theorem C.2.2) becomes
E
p,q
1
=

=(i
0
,...,ip)Ip
H
q
(U

i
0
U

ip
, Z) H
p+q
(X

, Z).
Our strategy will be to compute E
p,q
2
for small values of p, q.
The rst observation is that
E
p,0
1
=

(i
0
,...,ip)Ip
Z
since U

i
0
U

ip
=U

=
i
0

ip
, is connected for all . Hence we
get the Koszul complex (9.1.15) with M =Z, minus its rst term. Thus
E
p,0
2
=

0 p > 0
Z p = 0.
Since the maximal cones in are n-dimensional, Proposition 12.3.1 implies
E
0,q
1
=

i
(n)
H
q
(U

i
, Z) = 0, for all q > 0.
It follows that
E
0,q
2
= 0 for all q > 0.
Thus the E
2
sheet of the spectral sequence (with differentials shown only in the
p = 1 column) is:
0 E
1,2
2
d
1,2
2

E
2,2
2
E
3,2
2
0 E
1,1
2
d
1,1
2

E
2,1
2
E
3,1
2
Z 0 0 0.
Then E
2,0
r
and E
0,2
r
must be zero for all r 2. Moreover, the differentials into and
out of E
1,1
r
for all r 2 must be zero and as a result,
E
1,1
2
= E
1,1

H
2
(X

, Z).
However, we also know that E
1,1
2
is the kernel of the map
E
1,1
1
=

i<j
H
1
(U

i
U

j
, Z) E
2,1
1
=

i<j<k
H
1
(U

i
U

j
U

k
, Z)
since E
0,1
1
= 0. By Proposition 12.3.1,
H
1
(U

i
U

j
, Z) M(
i

j
), H
1
(U

i
U

j
U

k
, Z) M(
i

k
).
12.3. Singular Cohomology of Toric Varieties 579
Hence we get the commutative diagram:
0

0

H
2
(X

, Z)

i<j
M(
i

j
)

i<j<k
M(
i

k
)

i
M

i<j
M

i<j<k
M

i<j
M/M(
i

j
)

i<j<k
M/M(
i

k
)

0 0
The last two columns are obviously exact, and the rst row is exact by the analysis
of H
2
(X

, Z) given above. The second row is also exact since it is the Koszul
complex from (9.1.15). Then an easy diagram chase gives the exact sequence
M ker() H
2
(X

, Z) 0.
However,
i
has dimension n for all i, so that M(
i
) = 0 for all i. Thus
ker() = ker

i
M/M(
i
)

i<j
M/M(
i

j
)

CDiv
T
N
(X

),
and it follows immediately that H
2
(X

, Z) Pic(X

).
This is a wonderful illustration of how to use spectral sequences.
Computing the Other Cohomology Groups. The spectral sequence of the cover
U = U

max
can be used to study the H
k
(X

, Z) for all k. However, if there


are many cones in it becomes somewhat unwieldy to derive detailed information
about H
k
for k > 2 this way. In remainder of this section, we will see how to
use the decomposition of X

into torus orbits to compute the cohomology groups


H
k
(X

, Z), and hence the additive structure of H

(X

, Z), more efciently. The


ring structure of H

(X

, Z) will be discussed in 12.4.


A Family of Complexes. The method for computing the cohomology groups of
X

that we present comes from [162]. To begin, we discuss a notion of orientation


for a pair of cones with dim = dim +1. First, for each cone , we may
arbitrarily pick an orientation of the linear subspace (N

)
R
spanned by , deter-
mined by a choice of basis. For instance, in the special case that is a simplicial
fan, we can number the one-dimensional cones
i
= Cone(u
i
) in a xed way and
if =
i
1
+ +
ip
with i
1
< < i
p
, then we can specify the orientation via the
element
u
i
1
u
ip

p
N

.
580 Chapter 12. The Topology of Toric Varieties
Now if dim = dim +1, let v be any vector in not contained in . Then v
together with a basis for (N

)
R
forms a basis for (N

)
R
, and denes an orientation.
We dene an orientation coefcient
c
,
=

+1 if the orientation of determined by agrees with the chosen one


1 if not
0 if is not a face of .
Given any fan , x an integer q, 0 q n, and consider the abelian groups
and maps
C

(,

q
) =(C
p
(,

q
),
p
) [ p Z
dened as follows. First, we take
C
p
(,

q
) =

(np)

q
M(),
where M() =

M as usual. This is free abelian with


rankC
p
(,

q
) =

p
q

[(np)[.
Then

p
: C
p
(,

q
) C
p+1
(,

q
)
is the map dened on the components corresponding to cones (, ) in the two
direct sums by the following rule. If is not a face of , that component of
p
is
dened to be zero. On the other hand, if , then
p
is dened by
c
,
i
q
,
,
where c
,
are the orientation coefcients, and
i
q
,
:

q
M()

q
M()
is induced by the inclusion

. In other words, the component of


p
in the
summand for the cone in C
p+1
(,

q
) is given by

c
,
i
q
,
.
Note that C
p
(,

q
) is nonzero only for 0 q p n.
Lemma 12.3.3. C

(,

q
) is a complex, i.e.,
p+1

p
= 0 for all p.
Proof. If is a codimension 2 face of a cone , then there are exactly two facets
of containing , and you will show in Exercise 12.3.1 that
(12.3.1)

c
,
c
,
= 0.
12.3. Singular Cohomology of Toric Varieties 581
Now, for any cone (np),

p+1

p
[
M()
=
p+1

(np1)

c
,
i
q
,
=

(np2)


(np1)

c
,
c
,
i
q
,
i
q
,

(np2)


(np1)

c
,
c
,

i
q
,
= 0,
using (12.3.1).
Example 12.3.4. Consider the fan dening P
2
, shown for instance in Figure 2
from Example 3.1.9. Denote
i
= Cone(e
i
) for i = 0, 1, 2. We show the complexes
(C

(,

q
),

) for q = 0, 1, 2 in the following diagram:


(12.3.2)
q = 2 :
0

0

0

Z

0


q = 1 :
0

0

Z
3
C

Z
2
0


q = 0 :
0

Z
3
A

Z
3
B

Z

0

.
For instance, on the row for q = 1, the group C
1
(,

1
) is
C
1
(,

1
) =

(1)

1
M() Z
3
,
since

1
M(
0
) =Z(e
1
e
2
)

1
M(
1
) =Ze
2

1
M(
2
) =Ze
1
.
Similarly,
C
2
(,

1
) =

1
M(0) = M Z
2
.
Use orientation coefcients determined by the numbering of the one-dimensional
cones in the order
0
,
1
,
2
and the two-dimensional cones listed in the order

1
,
2
,
0
. The map denoted by C in (12.3.2) is dened by the matrix
C =

1 0 1
0 1 1

.
The maps denoted by A and B in (12.3.2) are
A =

1 1 0
0 1 1
1 0 1

, B =

1 1 1

.
582 Chapter 12. The Topology of Toric Varieties
Note that BA = 0 in agreement with Lemma 12.3.3.
The cohomology groups
H
p
(,

q
) = ker(
p
)/im(
p1
)
of these complexes will appear shortly.
Spectral Sequence of a Filtered Topological Space. For each integer p, the toric
variety X

contains the union of the closures of torus orbits of dimension p,


X
p
=

(np)
V() =

(), np
O().
By denition, these subsets give an increasing ltration of X

,
(12.3.3) = X
1
X
0
X
1
X
n
= X

.
For technical reasons, when working with general , where X

may not be
compact, we will consider cohomology with compact supports, denoted H
k
c
(X

, Z)
(see [135, p. 242]). When is complete the groups H
k
c
(X

, Z) and the ordinary


singular cohomology groups H
k
(X

, Z) coincide. For orientable noncompact man-


ifolds of real dimension d, it is the cohomology groups with compact supports that
appear in the statement of Poincar e duality (see [135, Thm. 3.35]). Thus for in-
stance, if X is homeomorphic to an open ball in R
d
,
(12.3.4) H
k
c
(X, Z) =

Z if k = d
0 otherwise.
Corresponding to the ltration (12.3.3), we have a rst quadrant cohomology
spectral sequence E
p,q
r
with
(12.3.5) E
p,q
1
= H
p+q
c
(X
p
, X
p1
, Z) H
p+q
c
(X

, Z).
See Theorem C.2.5 in Appendix C, and [136, Ch. 1] or [255, 9.4] for more details
on the construction.
Proposition 12.3.5. For p, q 0, the spectral sequence (12.3.5) has
E
p,q
1

(np)

q
M() =C
p
(,

q
).
Moreover, the differentials d
p,q
1
: E
p,q
1
E
p+1,q
1
agree with the coboundary maps
in the complex C

(,

q
), so that
E
p,q
2
= H
p
(,

q
).
Proof. By the excision property of cohomology with compact supports, we have
E
p,q
1

(np)
H
p+q
c
(O(), Z).
12.3. Singular Cohomology of Toric Varieties 583
Furthermore, the homeomorphism O()

= R
p
>0
S
N()
and the K unneth formula
imply that
H
p+q
c
(O(), Z)

k+=p+q
H
k
c
(R
p
>0
, Z)
Z
H

c
(S
N()
, Z).
To analyze the terms in this direct sum, note that by (12.3.4)
H
k
c
(R
p
>0
, Z) =

Z if k = p
0 otherwise.
Using Proposition 12.3.1, this shows that for each cone of dimension np,
H
p+q
c
(O(), Z) H
q
(S
N()
, Z)

q
M().
Hence E
p,q
1
C
p
(,

q
) as claimed.
To complete the proof, we need to show that the coboundary maps in the com-
plex C

(,

q
) agree with the differentials in the spectral sequence. By the con-
struction of the spectral sequence as in [255], the differential
d
p,q
1
: E
p,q
1
E
p+1,q
1
comes from the connecting homomorphism in the long exact cohomology sequence
of the triple (X
p+1
, X
p
, X
p1
):
H
p+q
c
(X
p
, X
p1
, Z) H
p+q+1
c
(X
p+1
, X
p
, Z).
By considering how this connecting homomorphism arises, it can be seen that d
p,q
1
coincides with
p
from the complex C

(,

q
) (Exercise 12.3.2).
Since E
p,q
1
is zero for all p, q outside the triangular region 0 q p n, it
follows easily that for r sufciently large, all of the higher differentials
d
p,q
r
: E
p,q
r
E
p+r,qr+1
r
will be zero and the spectral sequence will degenerate with
E
p,q
r
= E
p,q
r+1
= = E
p,q

for all p, q. When this happens, the E


p,q

with p+q = k are successive quotients in


a ltration of H
k
c
(X

, Z). In relatively small dimensions, and in many other good


cases as well, this information can be used to determine these cohomology groups
completely. We will illustrate this with the following examples.
Example 12.3.6. We continue from (12.3.2) to compute the E
2
sheet of the spectral
sequence (12.3.5) arising from the fan for P
2
. By a direct computation, the q = 0
row is E
0,0
2
=Z and E
1,0
2
= E
2,0
2
= 0 (see also Exercise 12.3.3).
On the second row, the kernel of C is 1-dimensional and the image of C is Z
2
.
Hence E
1,1
2
Z, and E
2,1
2
= 0. Finally E
2,2
2
=Z. Hence the E
2
sheet of the spectral
584 Chapter 12. The Topology of Toric Varieties
sequence is reduced to:
(12.3.6)
0 0 E
2,2
2
=Z 0

0 E
1,1
2
=Z 0 0

E
0,0
2
=Z 0 0 0
.
The E
2
differentials are d
p,q
2
: E
p,q
2
E
p+2,q1
2
. It can be seen directly in this case
that the spectral sequence degenerates at E
2
. Since there is at most one nonzero
group for each possible value of p+q, it follows that
H
0
(P
2
, Z) Z, H
2
(P
2
, Z) Z, H
4
(P
2
, Z) Z,
and all of the other H
k
(P
2
, Z) (including all the odd-numbered ones) are zero (see
Proposition C.1.5). The computations here can be generalized without difculty to
the case of X

=P
n
dened by the fan with one-dimensional cones
i
= Cone(e
i
),
i = 0, . . . , n and e
0
=e
1
e
n
. The result is that
H
2k
(P
n
, Z) Z, k = 0, . . . , n,
and the odd-numbered cohomology groups are all zero.
Our next examples show that the integral cohomology groups of a toric variety
can have torsion.
Example 12.3.7. Let in N
R
R
2
be the complete fan with 1-dimensional cones

i
and ray generators u
i
for i = 1, . . . , r, listed counterclockwise around the origin.
The 2-dimensional cones are
i
=
i
+
i+1
, i = 1, . . . , r if we take indices modulo
r. The rows of the E
1
sheet of the spectral sequence (12.3.5) for q = 0, 1, 2 are
(12.3.7)
0

0

0

Z

0


0

0

Z
r
C

Z
2
0


0

Z
r
A

Z
r
B

Z

0

.
By Exercise 12.3.3, E
p,0
2
= 0 for p = 1, 2 and E
0,0
2
Z.
The map C on the q = 1 row is the dened by the 2r matrix with columns
given by the vectors u
i
= (a
i
, b
i
). Since each pair u
i
, u
i+1
spans a two-dimensional
cone
i
, the kernel of C has rank r 2 and the image of C is a rank 2 sublattice of
M. The E
2
sheet of the spectral sequence has the form
(12.3.8)
0 0 E
2,2
2
=Z
0 E
1,1
2
=Z
r2
E
2,1
2
=Z/mZ
E
0,0
2
=Z 0 0.
12.3. Singular Cohomology of Toric Varieties 585
The integer m is given by
m = gcd

det

a
i
a
j
b
i
b
j

1 i < j r

.
Once again, because of the placement of the nonzero terms in E
2
, all of the higher
differentials must be zero, so E
2
= E

and there is at most one nonzero E


p,q

for
each value of p+q. Hence
H
0
(X

, Z) Z, H
1
(X

, Z) = 0, H
2
(X

, Z) Z
r2
,
H
3
(X

, Z) Z/mZ, H
4
(X

, Z) Z.
By the universal coefcient theorem for cohomology [135, Thm. 3.2], the torsion
also appears in H
2
(X

, Z). Also, note that H


1
(X

, Z) = 0 as we expect from 12.1


since X

is simply connected and H


1
(X

, Z) is the abelianization of
1
(X

). You
will study some special cases in Exercise 12.3.6.
We conclude with a nonsimplicial example that illustrates most of the general
behavior of these cohomology groups.
Example 12.3.8. Consider the fan in R
3
whose maximal cones are the cones
over the faces of the cube with vertices (1, 1, 1). This cube is shown in Figure
8 of Example 2.3.11 and is the normal fan of the octahedron shown in the same
gure. The E
1
sheet in the spectral sequence (12.3.5) is:
0

0

0

0

Z

0


0

0

0

Z
8
F

Z
3
0


0

0

Z
12
D

Z
16
E

Z
3
0


0

Z
6
A

Z
12
B

Z
8
C

Z

0

.
Suitable orientation coefcients can be obtained by placing any orientations on the
edges of the cube in R
3
. We leave it to the reader to prove the following claims.
First, the matrices A, B,C on the q = 0 row of the E
1
sheet make that row exact
except at the left and A has a rank 1 kernel, so E
0,0
2
Z, while
E
1,0
2
= E
2,0
2
= E
3,0
2
= 0.
For the q = 1 and q = 2 rows, we must determine the M() for the cones in .
When this is done, it is routine to show that D has rank 11 and E has rank 3. The
image of E is a sublattice of index 2 in

1
M Z
3
. Hence, E
1,1
2
Z, E
2,1
2
Z
2
,
and E
3,1
2
Z/2Z. With respect to obvious choices of bases, the matrix F is
F =

1 1 1 1 1 1 1 1
1 1 1 1 1 1 1 1
1 1 1 1 1 1 1 1

.
586 Chapter 12. The Topology of Toric Varieties
It is easy to check that F has rank 3, but that the image is a sublattice of index 4 in

2
M Z
3
. The quotient is isomorphic to Z/2ZZ/2Z. As a result the E
2
sheet
of the spectral sequence is:
(12.3.9)
0 0 0 E
3,3
2
=Z
0 0 E
2,2
2
=Z
5
E
3,2
2
=Z/2ZZ/2Z
0 E
1,1
2
=Z E
2,1
2
=Z
2
E
3,1
2
=Z/2Z
E
0,0
2
=Z 0 0 0.
In this case too, all of the higher differentials are zero, so the spectral sequence
degenerates at this point, and the E

sheet coincides with (12.3.9). You will verify


the details in Exercise 12.3.7 and show that
H
0
(X

, Z) Z, H
1
(X

, Z) = 0
H
2
(X

, Z) Z, H
3
(X

, Z) Z
2
H
4
(X

, Z) Z
5
Z/2Z, H
5
(X

, Z) Z/2ZZ/2Z
H
6
(X

, Z) Z.
Note that unlike the previous examples, some of the odd cohomology groups
would be nonzero in this example even if coefcients in Q were used. One can do
these computations using the Maple package torhom developed by Franz [98].
The Topological Euler Characteristic. The spectral sequence (12.3.5) can be used
to deduce several connections between the topology of the toric variety X

and the
combinatorics of the fan . First, we consider the topological Euler characteristic
of X

, which by [105, p. 95] and [162, Prop. 3.1.2] equals


(12.3.10) e(X

) =
2n

k=0
(1)
k
rank H
k
c
(X

, Z) =
2n

k=0
(1)
k
dim H
k
c
(X

, Q).
Theorem 12.3.9. For an n-dimensional toric variety X

, the Euler characteristic


is the number of n-dimensional cones in , i.e.,
e(X

) =[(n)[.
Proof. For each sheet of the spectral sequence (12.3.5), dene
e(E
r
) =

p,qZ
(1)
p+q
rank E
p,q
r
.
Since E
r+1
is the cohomology of E
r
with respect to d
p,q
r
, it follows that
e(E
r+1
) = e(E
r
)
12.3. Singular Cohomology of Toric Varieties 587
for all r. The E
p,q

with p+q = k are the quotients of a ltration of H


k
(X

, Z), and
by Proposition 12.3.5,
e(E
1
) =

p,qZ
(1)
p+q
rank


(np)

q
M()

p,qZ
(1)
p+q

p
q

[(np)[
since M() Z
p
when (np). Hence
e(X

) = e(E

) = e(E
1
) =
n

p=0

q=0
(1)
p+q

p
q

[(np)[ =[(n)[,
where the last equality follows since the inner sums vanish for p > 0 by properties
of binomial coefcients.
When X

is complete and simplicial, the individual Betti numbers (the ranks


of the H
k
(X

, Z)) can also be determined from the structure of .


Rational Coefcients. To suppress torsion in cohomology, we switch coefcients
from Z to Q. In this case, there is also a signicant simplication in the behavior
of the cohomology spectral sequence
(12.3.11) E
p,q
1
= H
p+q
c
(X
p
, X
p1
, Q) H
p+q
c
(X

, Q).
The same argument given in Proposition 12.3.5 shows that
(12.3.12) E
p,q
1

(np)

q
M()
Q
.
Proposition 12.3.10. The spectral sequence (12.3.11) degenerates at E
2
.
Proof. We show that d
p,q
r
=0 for all r 2 and all (p, q), so that E
p,q
2
=E
p,q

. Recall
from Example 3.3.6 that for any positive integer , the multiplication map

: N N, a a
is compatible with so there is a corresponding toric morphism

: X

whose restriction to T
N
X

is the group homomorphism

[
T
N
(t
1
, . . . , t
n
) = (t

1
, . . . , t

n
),
and similarly on each torus orbit. Because

respects the orbit decomposition of


X

, it respects the ltration from (12.3.3) and induces homomorphisms

: E
p,q
r
E
p,q
r
for each r. These commute with the differentials since (12.3.11) is functorial with
respect to maps that preserve the ltration.
In Exercise 12.3.8 you will use E
p,q
1

(np)
H
q
c
(O(), Q) to show that

acts on E
p,q
1
by multiplication by
q
. Then the same holds for all r since E
p,q
r+1
is
a quotient of a subspace of E
p,q
r
.
588 Chapter 12. The Topology of Toric Varieties
Let E
p,q
r
for r 2. Since d
p,q
r
() E
p+r,qr+1
r
, we have

qr+1
d
p,q
r
() =

(d
p,q
r
())
= d
p,q
r
(

())
= d
p,q
r
(
q
)
=
q
d
p,q
r
().
Since we use coefcients in Q, this implies d
p,q
r
() = 0 for all .
Proposition 12.3.10 shows that computing H
k
c
(X

, Q) using the cohomology


spectral sequence is especially simple. Franz has shown in [96] that if X

is
smooth, the spectral sequence with integer coefcients also degenerates at E
2
, and
no additional extension data is needed to determine H

c
(X

, Z).
A Vanishing Theorem for Singular Cohomology. We will now focus on complete
simplicial toric varieties. In this case, the following result of Oda [220, Thm. 4.1]
implies that the spectral sequence (12.3.11) simplies even further.
Theorem 12.3.11. If X

is complete and simplicial, then E


p,q
2
= 0 when p = q in
the spectral sequence (12.3.11). Thus:
(a) H
2k+1
(X

, Q) = 0 for all k.
(b) H
2k
(X

, Q) E
k,k
2
for all k.
Proof. Our proof will use some results from Chapter 9. Theorem 9.3.2 tells us that
(12.3.13) H
p
(X

q
X

) = 0
when p = q, and the Hodge decomposition (9.4.11) states that
(12.3.14) H
k
(X

, C)

p+q=k
H
p
(X

q
X

).
Here,

q
X

is the sheaf of Zariski q-forms on X

dened in 8.0.
It is now easy to show that E
p,q
2
= 0 when p +q is odd, since (12.3.13) and
(12.3.14) imply that H
k
(X

, C) = 0 for odd k. When we combine this with the


degeneration proved in Proposition 12.3.10, we obtain
0 = dim H
k
(X

, Q) =

p+q=k
dim E
p,q
2
.
When p+q is even, we change notation slightly and consider p+q = 2k. The
idea is to study how the maps

: X

from the proof of Proposition 12.3.10


act on each side of (12.3.14). On the left-hand side, the degeneration implies that
H
2k
(X

, C) has a ltration 0 = F
2k
F
0
= H
2k
(X

, C) such that
F
p
/F
p1
(E
p,q
2
)
C
= E
p,q
2

Q
C, p+q = 2k.
By the proof of Proposition 12.3.10,

: (E
p,q
2
)
C
(E
p,q
2
)
C
is multiplication by

q
. Hence

: H
2k
(X

, C) H
2k
(X

, C) is multiplication by
q
on F
p
/F
p1
.
12.3. Singular Cohomology of Toric Varieties 589
Now consider the right-hand side of (12.3.14) and assume for the moment that
X

is smooth. Recall from 8.1 that

1
X

M
Z
O
X


1
X

(logD).
Since
m

=
m
, we have

(d
m
/
m
) = d
m
/
m
= d
m
/
m
. It follows
that over the afne open subset U

, ,

acts on sections by multiplication


by . Hence, for
q
X

1
X

acts on sections over U

by multiplication by

q
. Computing cohomology via the

Cech complex, we conclude that

acts on
H
p
(X

,
q
X

) by multiplication by
q
. In Exercise 12.3.9, you will show that in the
general case,

continues to be multiplication by
q
on H
p
(X

q
X

).
Since H
2k
(X

, C) H
k
(X

k
X

) by (12.3.13) and (12.3.14), it follows that

is multiplication by
k
on H
2k
(X

, C). Comparing this with our earlier analysis,


we see that the ltration F
p
must collapse, giving (E
p,q
2
)
C
= 0 when p+q = 2k
and p = q. This proves the theorem.
You should check that Examples 12.3.6 and 12.3.7 illustrate the vanishing of
odd cohomology asserted by part (a) of Theorem 12.3.11. On the other hand, we
see from Example 12.3.8 that this can fail for nonsimplicial X

. There is also a
more rened vanishing statement due to Brion [48, p. 5], which states that if X

is complete of dimension n and all cones of dimension m are simplicial,


then E
p,q
2
= 0 unless 0 pq nm.
Some Combinatorial Consequences. There are interesting relations between the
numbers of cones of various dimensions in simplicial fans and the Betti numbers
of the corresponding toric varieties, which are the numbers
b
2k
(X

) = dim H
2k
(X

, Q).
Theorem 12.3.12. Let be a complete simplicial fan in N
R
R
n
. Then the Betti
numbers of X

are given by
b
2k
(X

) =
n

i=k
(1)
ik

i
k

[(ni)[
and satisfy
b
2k
(X

) = b
2n2k
(X

).
Proof. The E
p,q
2
terms of the spectral sequence (12.3.11) are the cohomology of
the E
p,q
1
terms, and we also have E
p,q
1
= 0 for p < q by (12.3.12). Since E
p,q
2
= 0
unless p = q by Theorem 12.3.11, it follows that
0 E
k,k
2
E
k,k
1
E
k+1,k
1

590 Chapter 12. The Topology of Toric Varieties
is exact. Hence
b
2k
(X

) = dim E
k,k
2
=
n

i=k
(1)
ik
dim E
i,k
1
=
n

i=k
(1)
ik

i
k

[(ni)[,
where the last equality uses (12.3.12). The second assertion follows from Poincar e
duality, which is discussed in more detail in 12.4.
If P is a full dimensional simple lattice polytope and X
P
is the corresponding
simplicial toric variety, then [(ni)[ = f
i
, the number of i-dimensional faces of
P. An immediate corollary of the theorem is the formula
(12.3.15) b
2k
(X
P
) =
n

i=k
(1)
ik

i
k

f
i
= h
k
,
where the h
k
are the combinations of the face numbers introduced in 9.4. The
Dehn-Sommerville equations h
k
= h
nk
are a consequence of the symmetry of the
Betti numbers in Theorem 12.3.12.
A nal comment is that as in the proof of Theorem 12.3.11, (12.3.13) and
(12.3.14) imply that
H
2k
(X
P
, C) H
k
(X
P
,

k
X
P
)
for P as above. Since we also have dim H
k
(X
P
,

k
X
P
) = h
k
by Theorem 9.4.11, we
get another proof of (12.3.15). This was noted earlier in (9.4.12).
Exercises for 12.3.
12.3.1. Show that if is a face of of codimension 2, then

c
,
c
,
= 0,
where the sum has two terms corresponding to the two facets of containing .
12.3.2. Complete the proof of Proposition 12.3.5 by showing that
d
p,q
1
: H
p+q
(X
p
, X
p1
, Z) H
p+q+1
(X
p+1
, X
p
, Z)
coincides with
p
: C
p
(,

q
) C
p+1
(,

q
) under the isomorphisms given in the rst
part of the proof. Hint: If you get stuck, see [162].
12.3.3. This exercise will show that when X

is complete, the q = 0 row of the E


2
sheet of
the spectral sequence (12.3.5) is given by
E
p,0
2
=

Z if p = 0
0 otherwise.
Let S
n1
be the unit sphere in N
R
and let
( = S
n1
[ , =0
be the spherical cell complex determined by .
(a) Show that the complex (C

(,

0
),

) is isomorphic to the augmented cellular chain


complex of the spherical complex ( (see [135, p. 139]).
12.3. Singular Cohomology of Toric Varieties 591
(b) Show that E
p,0
2
is isomorphic to the reduced homology group

H
np1
((, Z).
(c) If is complete, then ( is a subdivision of S
n1
. Deduce the above formula for E
p,0
2
.
12.3.4. Let Bl
0
(C
2
) denote the blowup of C
2
at the origin. Compute the cohomology
groups H
k
(Bl
0
(C
2
), Z) directly from the spectral sequence (12.3.5). Generalize this by
computing H
k
(Bl
0
(C
n
), Z) for all n 2.
12.3.5. Let be the complete fan in N
R
R
2
with four maximal cones Cone(e
1
, e
2
),
so X

= P
1
P
1
. Compute H
k
(X

, Z) directly via the spectral sequence (12.3.5), and


compare with the result in Example 12.3.7. Generalize your computations to the complete
fans in N
R
R
n
with maximal cones Cone(e
1
, . . . , e
n
).
12.3.6. This exercise deals with Example 12.3.7.
(a) Show that if X

is a smooth toric surface, then the cohomology groups H


k
(X

, Z) are
torsion free.
(b) Let X

be the toric variety of the complete fan in R


2
whose minimal ray generators
are u = (1, 1). Show that H
3
(X

, Z) Z/2Z.
12.3.7. In this exercise, you will verify details of the computation in Example 12.3.8.
(a) Show that the E

sheet of the cohomology spectral sequence is given by (12.3.9).


(b) Show that the cohomology groups are as given in the example. Note that Proposi-
tion C.1.5 applies for all of the groups except H
4
(X

, Z). Show that H


4
(X

, Z)
Z
5
Z/2Z directly by considering the ltration with quotients given by E
p,q

with
p +q = 4.
12.3.8. Show that

is the th power map on O() and conclude that

is multiplication
by
q
on H
q
c
(O(), Q) H
q
(S
N()
, Q).
12.3.9. Here are some details from the proof of Theorem 12.3.11
(a) Use the properties of reexive sheaves from 8.0 to prove the assertion in the text that
over U

acts on sections

q
X
by multiplication by
q
.
(b) Explain carefully how computing cohomology via the

Cech complex implies that

acts on H
p
(X

q
X
) by multiplication by
q
.
12.3.10. In this exercise, you will show that if in N
R
R
n
is smooth and complete, then
the higher homotopy group
2
(X

) can be identied with Pic(X

.
(a) Show that
2
(X

) H
2
(X

, Z). Hint: Apply the Hurewicz theorem as stated in [135,


Thm. 4.32].
(b) Prove that H
2
(X

, Z) (H
2
(X

, Z))

under these hypotheses. Hint: You will need to


show that H
2
(X

, Z) has no torsion.
(c) Conclude that
2
(X

) is isomorphic to Pic(X

.
12.3.11. To what extent is the cohomology of a toric variety X

determined by the com-


binatorics of the fan (the numbers of cones of each dimension, and their intersection
relations)? In this exercise, drawing on [200], you will see that for n 3, the Betti num-
bers of a toric variety are not necessarily determined by the combinatorial type of .
(a) To begin, deduce from Example 12.3.7 that the Betti numbers of a complete toric
surface are determined by the number of 2-dimensional cones in , and that this is the
only combinatorial invariant of .
592 Chapter 12. The Topology of Toric Varieties
(b) Now as in Example 12.3.8, consider the fan over the faces of the polytope P in R
3
given by
P = Conv

e
1
, e
2
, e
3
,
1
2
e
1

1
2
e
2

1
2
e
3

,
where we take all possible choices of signs, so there are 14 points in all in the set.
Show that the Betti numbers of X

are
1, 0, 2, 3, 11, 0, 1.
Note that X

is not simplicial and Poincar e duality does not hold.


(c) Now let P

be the convex hull of the 14 points


e
1
, e
2
,
1
2
e
1
+
1
2
e
3
, e
1
, e
2
, e
3
,
2
5
e
1
+
3
5
e
2
+
1
5
e
3
,
1
2
e
1
+
1
2
e
2

1
2
e
3
,
2
5
e
1

3
5
e
2
+
1
5
e
3
,
1
2
e
1

1
2
e
2

1
2
e
3
,
2
3
e
1
+
1
3
e
2
+
1
3
e
3
,
1
2
e
1
+
1
2
e
2

1
2
e
3
,

2
3
e
1

1
3
e
2
+
1
3
e
3
,
1
2
e
1

1
2
e
2

1
2
e
3
.
If

is the fan over the faces of P

, show that and

are combinatorially equivalent,


but X

has Betti numbers


1, 0, 1, 2, 11, 0, 1.
(d) Modify this example to produce similar examples in all dimensions n 3.
12.3.12. The Poincar e polynomial P
X
(t) =

2n
k=0
b
k
(X)t
k
of a variety X of dimension n is
the generating function for its Betti numbers. Show that Theorem 12.3.12 is equivalent to
the assertion that P
X
(t) =

(t
2
1)
dim
.
12.4. The Cohomology Ring
In this section, we will make the standing assumption that X

is complete and
simplicial. Our goal is to prove a general theorem describing the ring structure
given by cup product on
H

(X

, Q) =
2n

k=0
H
k
(X

, Q), n = dim X

.
Along the way, we will also describe the equivariant cohomology ring H

T
N
(X

, Q).
Rationally Smooth Varieties. By Theorem 11.4.8, simplicial toric varieties are
rationally smooth. The basic intuition is that rationally smooth varieties behave
like smooth varieties, provided that one works over Q.
We begin with some properties of the cohomology ring of an n-dimensional
complete rationally smooth variety X:
Poincar e duality holds between cohomology and homology, so that
H
k
(X, Q) H
2nk
(X, Q).
The isomorphism H
2n
(X, Q) H
0
(X, Q) =Q induces a map

X
: H

(X, Q) Q,
where elements of H
k
(X, Q) map to zero when k < 2n.
12.4. The Cohomology Ring 593
Poincar e duality implies that the cup product pairing
H
k
(X, Q) H
2nk
(X, Q) Q
dened by (, )

X
is nondegenerate. Thus H
2nk
(X, Q) is isomor-
phic to the dual vector space of H
k
(X, Q) and their dimensions are equal.
An irreducible subvariety W X of codimension k has a rened cohomology
class [W]
r
H
2n2k
(X, X `W, Q). By mapping this to H
2n2k
(X, Q), W has a
cohomology class [W] H
2n2k
(X, Q).
The cohomology class of a divisor D =

i
a
i
D
i
is [D] =

i
a
i
[D
i
] H
2
(X, Q),
and linearly equivalent divisors give the same cohomology class.
See [135, Sec. 3.3] for the manifold case. For rationally smooth varieties these
statements follow, for instance, from the properties of intersection cohomology
developed by Goresky and MacPherson in [119] since over Q, intersection coho-
mology coincides with ordinary cohomology for rationally smooth varieties. The
assertions about rened cohomology classes and linearly equivalent Cartier divi-
sors can be found in [107, Ch. 19]. We also note that if X is smooth, then the above
properties hold over Z. See also the appendix to Chapter 13.
These properties enable us to generalize the intersection products dened in
6.3. Let V,W X be irreducible subvarieties with dimV +dimW = n. The cup
product [V] [W] lies in H
2n
(X, Q) and hence gives the intersection product
(12.4.1) V W =

X
[V] [W] Q.
When D is an irreducible Cartier divisor and C is an irreducible curve, this agrees
with the intersection product D C from Denition 6.3.6. See Exercise 12.5.9 and
[107, 2.3].
Statement of the Main Theorem. Let X

be a complete simplicial toric variety


and x a numbering
1
, . . . ,
r
for the rays in (1). Also let u
i
be the minimal
generator of
i
and introduce a variable x
i
for each
i
. In the ring Q[x
1
, . . . , x
r
], let
I be the monomial ideal with square-free generators as follows:
(12.4.2) I ='x
i
1
x
is
[ i
j
are distinct and
i
1
+ +
is
is not a cone of `.
We call I the Stanley-Reisner ideal. Also let . be the ideal generated by the
linear forms
(12.4.3)
r

i=1
'm, u
i
`x
i
,
where m ranges over M (or equivalently, over some basis for M). We dene
R
Q
() =Q[x
1
, . . . , x
r
]/(I +.).
Since I +. is homogeneous for the standard grading on Q[x
1
, . . . , x
r
], R
Q
() is
a graded ring. Let R
Q
()
d
denote the graded piece in degree d.
594 Chapter 12. The Topology of Toric Varieties
Our next task is to show that x
i
[D
i
] H
2
(X

, Q) induces a well-dened
ring homomorphism
(12.4.4) R
Q
() H

(X

, Q).
First note that

r
i=1
'm, u
i
`D
i
= div(
m
) 0, which by the above properties of
cohomology classes implies that

r
i=1
'm, u
i
`[D
i
] = 0 H
2
(X

, Q). This shows


that the ideal . maps to zero in H

(X

, Q).
Moreover, if
i
1
+ +
is
is not a cone of , then D
i
1
D
is
= in X

.
Since cup product in relative cohomology satises
H
k
(X, A, Q) and H

(X, B, Q) = H
k+
(X, AB, Q)
(see [210, Ch. 5]), the cup product of the rened classes [D
i
j
]
r
is
[D
i
1
]
r
[D
is
]
r
H
2s
(X

s
j=1
(X

`D
i
j
), Q).
This relative cohomology group vanishes since D
i
1
D
is
=. Then
[D
i
1
] [D
is
] = 0 H
2s
(X

, Q)
since [D
i
j
]
r
maps to [D
i
j
]. Hence the Stanley-Reisner ideal I also maps to zero in
H

(X

, Q). This gives the ring homomorphism (12.4.4).


Our main theorem will show that R
Q
() and H

(X

, Q) are isomorphic.
Theorem 12.4.1. Let be complete and simplicial. Then the map (12.4.4) is an
isomorphism:
R
Q
() H

(X

, Q).
Thus, in even degrees, H
2k
(X

, Q) is isomorphic to R
Q
()
k
, and in odd degrees,
H
2k+1
(X

, Q) is zero.
The proof of this theorem will be given later in the section.
Example 12.4.2. As a toric variety, P
n
comes from the fan with ray generators
u
i
= e
i
for i = 1, . . . , n and u
0
=e
1
e
n
. Then it is easy to check that
I ='x
0
x
n
`,
and using the basis e
1
, . . . , e
n
for M to obtain generators for ., we have
. ='x
1
x
0
, . . . , x
n
x
0
`.
Then Theorem 12.4.1 gives an isomorphism
H

(P
n
, Q) Q[x
0
, x
1
, . . . , x
n
]/'x
0
x
n
, x
1
x
0
, . . . , x
n
x
0
`
Q[x
0
]/'x
n+1
0
`.
This agrees with Example 9.0.13. See also Example 12.3.6.
12.4. The Cohomology Ring 595
Example 12.4.3. Consider the Hirzebruch surface H
r
and label the cones in the
fan as in Example 10.4.6, so
1
=Cone(e
1
+re
2
),
2
=Cone(e
2
),
3
=Cone(e
1
),
and
4
= Cone(e
2
). Theorem 12.4.1 gives
H

(H
r
, Q) Q[x
1
, x
2
, x
3
, x
4
]/'x
1
x
3
, x
2
x
4
, x
1
+x
3
, r x
1
+x
2
x
4
`.
The two linear relations from . can be used to eliminate x
1
, x
2
, giving
H

(H
r
, Q) Q[x
3
, x
4
]/'x
2
3
, x
2
4
r x
3
x
4
`.
Note that 1, x
3
, x
4
, x
3
x
4
form a Q-basis of H

(H
r
, Q). Hence
H
0
(H
r
, Q) Q, H
2
(H
r
, Q) Q
2
, and H
4
(H
r
, Q) Q,
as we expect from 12.3. Also note that cup product in H

(H
r
, Q) is dened by
x
2
3
= 0, x
2
4
= r x
3
x
4
.
By (12.4.1), we recover the intersection form on the divisor classes, described by
the matrix
(D
i
D
j
) =

0 1
1 r

in Example 10.4.6.
Theorem 12.4.1 also holds over Z when X

is smooth. The examples just given


show this for P
n
and the Hirzebruch surface H
r
. Here is the precise result.
Theorem 12.4.4 (Jurkiewicz-Danilov). Let X

be a smooth complete toric variety.


For the polynomial ring Z[x
1
, . . . , x
r
] with variables indexed by
1
, . . . ,
r
(1),
let I and . be the ideals in Z[x
1
, . . . , x
r
] generated by the polynomials in (12.4.2)
and (12.4.3), and dene
R() =Z[x
1
, . . . , x
r
]/(I +.).
Then x
i
[D

i
] induces a ring isomorphism R() H

(X

, Z).
See [76, Thm. 10.8], [105, Sec. 5.2] (which gives a proof under an additional
hypothesis that is satised, for instance, if X

is projective), and [218, Sec. 3.3].


Note that Theorem 12.4.1 is simultaneously less general (coefcients in Q instead
of Z) and more general (X

is simplicial instead of smooth) than Theorem 12.4.4.


Let us make one nal comment about the cohomology ring H

(X

, Q). Recall
from Theorem 12.3.12 that the Betti numbers b
k
(X

) = dim H
k
(X

, Q) depend
only on the combinatorial structure of the fan in the simplicial case. Does this
extend to the ring structure on H

(X

, Q)? Sometimes the answer is yes.


Example 12.4.5. Given nonnegative integers r = s, the Hirzebruch surfaces H
r
and H
s
have cohomology rings
H

(H
r
, Q) Q[x, y]/'x
2
, y
2
r xy`, H

(H
s
, Q) Q[x, y]/'x
2
, y
2
sxy`
by Example 12.4.3. One easily checks that (x, y) (x, y +
1
2
(s r)x) induces a
ring isomorphism H

(H
r
, Q) H

(H
s
, Q).
596 Chapter 12. The Topology of Toric Varieties
In general, however, the ring structure on H

(X

, Q) is not a combinatorial
invariant, even when is simplicial (Exercise 12.4.1).
Equivariant Cohomology. We will prove Theorem 12.4.1 by rst computing the
equivariant cohomology of X

for the action of the torus T


N
and then passing from
equivariant cohomology to ordinary singular cohomology. This method of proof
comes from an extensive study of equivariant cohomology by many authors over
the past 30 years. Our presentation draws mostly on [47] and [108].
Let G be a Lie group acting on a topological space X on the left. When the
action of G on X is not free, the quotient X/G can be badly behaved. As a replace-
ment for X/G when the action is not free, in [38], Borel introduced the following
idea. Let EG be a contractible space on which G acts freely on the right, and let
BG = EG/G. Then EGX is homotopy equivalent to X and G acts freely on this
space. Hence we can form the quotient
EG
G
X = EGX/ , where (e g, x) (e, g x) for g G,
as replacement for X/G. Moreover, EG
G
X has the structure of a ber bundle
over BG with ber equal to X. It is a theorem that suitable, possibly innite-
dimensional, EG always exist. The homotopy type of EG
G
X is also independent
of the EG used. Hence we have a well-dened new cohomology theory.
Denition 12.4.6. Let X be a topological space with a left action of the Lie group
G. The equivariant cohomology of X with respect to G with coefcients in a ring
R, denoted H

G
(X, R), is dened as
H

G
(X, R) = H

(EG
G
X, R),
where the right-hand side is ordinary singular cohomology.
We note an important special case of this denition that has some useful con-
sequences. If X =pt is a single point with the trivial action of G, then
EG
G
pt

= EG/G = BG.
Hence H

G
(pt, Z) = H

(BG, Z) is an algebraic invariant depending only on the


group G. We will call this ring
G
. If X is any other space on which G acts, the
constant map X pt induces a map
(12.4.5)
G
= H

G
(pt, Z) H

G
(X, Z)
that makes H

G
(X, Z) into a module over the ring
G
. Since the product in
G
is
the cup product,
G
may fail to be commutative if H
k
(BG, Z) = 0 for some odd k.
The Torus. For us, the group G will always be a torus T (C

)
n
. Here, the spaces
EG and BG and the ring
G
can be described very concretely.
First consider the 1-dimensional torus C

. Let C

be the space of all vectors


(a
0
, a
1
, . . .) with a
i
C for i N, and a
i
= 0 for i 0. Note that C

is the union
12.4. The Cohomology Ring 597
of C
+1
for 1 if we embed C
+1
as the set of vectors with zeros after the rst
+1 entries. In Exercise 12.4.2 you will show the unexpected fact that
(12.4.6) EC

=C

`0
is a contractible space. The torus C

acts freely on EC

in the obvious way:


t (a
0
, a
1
, . . .) = (ta
0
, ta
1
, . . .).
By analogy with the nite-dimensional construction, BC

is denoted by P

:
BC

= EC

/C

=P

.
Note that P

=0
P

. Thus, when computing cohomology in a xed degree,


we can replace BC

with the nite-dimensional approximation P

for 0. To
see how this works, note that
H

(P

, Z) Z[t]/'t
+1
`
since the computations in Example 12.4.2 work over Z. Letting , we obtain
(12.4.7)
C
= H

(BC

, Z) = H

(P

, Z) Z[t],
where the generator t is in degree 2. Note that in this case
C
is a commutative
ring since H
k
(P

, Z) = 0 when k is odd.
For the n-dimensional torus T = (C

)
n
, we can take
ET = EC

EC

,
which implies
BT = BC

BC

.
Therefore by the K unneth formula,
(12.4.8) H

T
(pt, Z) = H

(BT, Z) =
T
N
Z[t
1
, . . . , t
n
].
Note that
T
N
is commutative and the t
i
all have degree 2.
For a torus T
N
= Hom
Z
(M, C

), we can describe (12.4.8) more intrinsically as


follows. Recall that M has the symmetric algebra
Sym
Z
(M) =ZMSym
2
(M) .
To map this to H

(BT
N
, Z), take m M. By the functorial properties of G BG
(see [271]),
m
: T
N
C

induces B
m
: BT
N
BC

. Taking H
2
, we obtain
Z t H
2
(BC

, Z)
(B
m
)

H
2
(BT
N
, Z).
Then dene s : M H
2
(BT
N
, Z) by m (B
m
)

(t). This gives a canonical iso-


morphism
s : Sym
Z
(M)

H

(BT
N
, Z).
To simplify notation, we will often write T instead of T
N
in what follows. Then,
when we switch from Z to Q, these isomorphisms become
(12.4.9) s : Sym
Q
(M)

H

T
(pt, Q) = H

(BT, Q) = (
T
)
Q
,
598 Chapter 12. The Topology of Toric Varieties
where Sym
Q
(M) is our simplied notation for Sym
Z
(M)
Z
Q= Sym
Q
(M
Q
).
The Fixed Points. Now let T = T
N
be the torus of X

and let X
T

denote the set


of xed points for the torus action. The Orbit-Cone Correspondence implies that
there are nitely many xed points x

, one for each cone (n).


The inclusion X
T

induces a ring homomorphism


(12.4.10) H

T
(X

, Z) H

T
(X
T

, Z).
Since X
T

is nite, we have
H

T
(X
T

, Q) =

(n)
H

T
(x

, Q) =

(n)
(
T
)
Q
,
so that (12.4.10) can be written
(12.4.11) : H

T
(X

, Q)

(n)
(
T
)
Q
.
The map is used in the localization theorem, which is a key tool in equivariant
cohomology. We will say more about this in Corollary 12.4.9.
Comparing Equivariant and Ordinary Cohomology. We next study the relation
between equivariant and singular cohomology for a toric variety. Let T = T
N
be
the torus of X

and note that H

T
(X

, Q) is a (
T
)
Q
-module by (12.4.5). This is
compatible with the ring structure, so that H

T
(X

, Q) is a (
T
)
Q
-algebra.
The next proposition shows that the equivariant cohomology of a toric variety
can be obtained from its singular cohomology by a change of scalars.
Proposition 12.4.7. Let X

be a complete simplicial toric variety and let T = T


N
.
Then there is a isomorphism of (
T
)
Q
-modules
H

T
(X

, Q) (
T
)
Q

Q
H

(X

, Q).
In particular, H

T
(X

, Q) is a free (
T
)
Q
-module of nite rank.
Proof. By the Borel construction, ET
T
X

is a ber bundle over BT = (P

)
n
with ber equal to X

. The Serre spectral sequence for singular cohomology (see


[136, Ch. 1], [199, Thm. 5.2], [255, Sec. 9.4], and Theorem C.2.6) computes the
cohomology of the total space ET
T
X

in terms of the cohomologies of the base


and the ber. Since we use rational coefcients, the E
2
sheet has the form
E
p,q
2
= H
p
(BT, H
q
(X

, Q)) H
p
(BT, Q)
Q
H
q
(X

, Q)
by the universal coefcient theorem, and the spectral sequence converges to
H
p+q
(ET
T
X

, Q) = H
p+q
T
(X

, Q).
However, H
q
(X

, Q) = 0 for odd q by Theorem 12.3.11, and similarly H


p
(BT, Q)
vanishes for odd p by the comments following (12.4.8). Since the rth differential
is d
r
: E
p,q
r
E
p+r,qr+1
r
, it follows that all of the d
r
are zero for r 2. Hence
12.4. The Cohomology Ring 599
the Serre spectral sequence degenerates at E
2
. Then the Leray-Hirsch theorem (see
[135, Thm. 4D.1]) implies that
H

(ET
T
X

, Q) = H

T
(X

, Q) H

(BT, Q)
Q
H

(X

, Q).
This is a (
T
)
Q
-module isomorphism since ET
T
X

BT induces the edge


homomorphism E
p,0
= H
p
(BT, Q) H
p
T
(X

, Q) by Theorem C.2.6.
Example 12.4.15 below will show that the isomorphism in Proposition 12.4.7
may fail to preserve the ring structure.
In general, the action of a torus T on a space X is called equivariantly formal
if the spectral sequence in Proposition 12.4.7 degenerates at E
2
. Starting in [38],
equivariantly formal spaces have been studied extensively in topology and sym-
plectic geometry. A good reference is the paper [118] by Goresky, Kottwitz and
MacPherson. For toric varieties, the paper [97] by Franz gives conditions under
which Proposition 12.4.7 remains true over Z.
We next describe two useful consequences of Proposition 12.4.7. The rst
uses the ideal I
T
(
T
)
Q
generated by elements of positive degree. The quotient
(
T
)
Q
/I
T
Q gives Q the structure of a (
T
)
Q
-module. Here is the result.
Corollary 12.4.8. The ordinary cohomology ring H

(X

, Q) is determined by the
(
T
)
Q
-algebra structure of H

T
(X

, Q) via the isomorphism


H

(X

, Q) H

T
(X

, Q)/I
T
H

T
(X

, Q) H

T
(X

, Q)
(
T
)
Q
Q.
Proof. We have to be careful since H

T
(X

, Q) (
T
)
Q

Q
H

(X

, Q) need not
be a ring isomorphism. When we combine this isomorphism with (
T
)
Q
Q, we
get a surjective homomorphism of graded Q-vector spaces
(12.4.12) H

T
(X

, Q) H

(X

, Q)
with kernel I
T
H

T
(X

, Q). If we can show that (12.4.12) is a ring homomorphism,


then the corollary will follow.
The isomorphism of Proposition 12.4.7 comes from the Serre spectral sequence
used in the proof. Degeneration at E
2
implies that
H
q
T
(X

, Q) =

a+b=q
E
a,b
2
=

a+b=q
H
a
(BT, Q)
Q
H
b
(X

, Q).
Since I
T
(
T
)
Q
=H

(BT, Q) is generated by the elements of positive degree, the


map (12.4.12) in degree q is the edge homomorphism H
q
T
(X

, Q) E
0,q
2
, which
by Theorem C.2.6 is the map induced by the inclusion of the ber
i
X

: X

ET
T
X

in the ber bundle given by the Borel construction. Hence (12.4.12) is the ring
homomorphism
(12.4.13) i

: H

T
(X

, Q) H

(X

, Q).
The map i

is often called taking the nonequivariant limit.


600 Chapter 12. The Topology of Toric Varieties
The second corollary concerns the map : H

T
(X

, Q)

(n)
(
T
)
Q
that
we constructed in (12.4.10) using the xed points of the torus action.
Corollary 12.4.9. : H

T
(X

, Q)

(n)
(
T
)
Q
is injective.
Proof. By (12.4.10) and (12.4.11), is the map H

T
(X

, Q) H

T
(X
T

, Q) induced
by the inclusion of the xed point set X
T

. Let K be the eld of fractions of


(
T
)
Q
Q[t
1
, . . . , t
n
]. The localization theorem in equivariant cohomology states
that becomes an isomorphism after tensoring with K (see [9]). This implies that
the kernel of is a (
T
)
Q
-torsion module. However, H

T
(X

, Q) is free over (
T
)
Q
by Proposition 12.4.7, so ker() is torsion-free. Hence the kernel vanishes.
Other proofs of this corollary can be found in [50, 3.4] and [118, Thm. 7.2].
Both papers explain how one can choose a single element f (
T
)
Q
such that
H

T
(X, Q)
f
H

T
(X
T
, Q)
f
is an isomorphism. See also Exercise 12.4.3.
The Stanley-Reisner Ring. Our strategy for proving Theorem 12.4.1 will be to
compute H

T
(X

, Q) and then take its nonequivariant limit via Corollary 12.4.8.


For this purpose, we introduce a modication of the ring R
Q
() appearing in the
statement of Theorem 12.4.1.
Denition 12.4.10. The Stanley-Reisner ring of a fan in N
R
is
SR
Q
() =Q[x
1
, . . . , x
r
]/I,
where (1) =
1
, . . . ,
r
and as in (12.4.2), I is the Stanley-Reisner ideal
I ='x
i
1
x
is
[ i
j
distinct and
i
1
+ +
is
is not a cone of `.
Example 12.4.11. Let
r
be the usual fan dening the Hirzebruch surface H
r
as
in Example 12.4.3. Then SR
Q
(
r
) =Q[x
1
, x
2
, x
3
, x
4
]/'x
1
x
3
, x
2
x
4
`.
The Stanley-Reisner ring SR
Q
() also has a natural structure as an algebra
over Sym
Q
(M). This comes from the ring homomorphism
Sym
Q
(M) SR
Q
()
induced by m [

r
i=1
'm, u
i
`x
i
] for m M, where [ f ] Sym
Q
(M) is the image
of f Q[x
1
, . . . , x
r
]. As a Sym
Q
(M)-module, an element m M acts on SR
Q
()
by multiplication by the class of the linear form

r
i=1
'm, u
i
`x
i
. The unexpected
minus sign will become clear once we discuss equivariant cohomology classes at
the end of the section.
Here is a simple but illuminating example.
Example 12.4.12. Let us compute the Sym
Q
(M)-module structure on SR
Q
(
r
) =
Q[x
1
, x
2
, x
3
, x
4
]/'x
1
x
3
, x
2
x
4
` from Example 12.4.11. Let e
1
, e
2
be the standard basis
12.4. The Cohomology Ring 601
of M =Z
2
. Then
e
1
acts by multiplication by

4
i=1
'e
1
, u
i
`x
i
= x
1
x
3
e
2
acts by multiplication by

4
i=1
'e
2
, u
i
`x
i
=r x
1
x
2
+x
4
.
For r 0, the surfaces H
r
all have isomorphic Stanley-Reisner rings. But once
we introduce the Sym
Q
(M)-module structure, then the action of e
2
tells us which
Hirzebruch surface we have.
The Equivariant Cohomology Class of a Divisor. A divisor D on a toric variety
X

has a cohomology class [D] H


2
(X

, Q). If D is torus-invariant, then it also


has an equivariant cohomology class [D]
T
H
2
T
(X

, Q) as follows.
Proposition 12.4.13. Let X

be a simplicial toric variety. Then a torus-invariant


divisor has a class [D]
T
H
2
T
(X

, Q) with the following properties:


(a) [D
1
+D
2
]
T
= [D
1
]
T
+[D
2
]
T
.
(b) [div(
m
)]
T
=s(m) 1, where 1 H
0
T
(X

, Q) and s(m) (
T
)
Q
by (12.4.9).
(c) i

[D]
T
= [D[
U
]
T
, where i : U X

is the inclusion of a torus-invariant open


subset of X

.
(d) i

[D]
T
=[D], where i

: H

T
(X

, Q) H

(X

, Q) is the nonequivariant limit


map from (12.4.13).
We defer the proof until the the end of the section.
The Equivariant Cohomology Ring of a Toric Variety. Recall that H

T
(X

, Q) is a
(
T
)
Q
-algebra. Then the isomorphism s : Sym
Q
(M) (
T
)
Q
from (12.4.9) makes
H
2
T
(X

, Q) into a Sym
Q
(M)-algebra. Then we have the following theorem.
Theorem 12.4.14. The map x
i
[D
i
]
T
H
2
T
(X

, Q) induces an isomorphism of
Sym
Q
(M)-algebras SR
Q
() H

T
(X

, Q).
Before starting the proof, we give two applications. The rst is a proof of the
main result of this section.
Proof of Theorem 12.4.1. The ideal 'M` Sym
Q
(M) corresponds to I
T
(
T
)
Q
.
Thus Theorem 12.4.14 gives an isomorphism
SR
Q
()/'M` H

T
(X

, Q)/I
T
H

T
(X

, Q).
The denition of Sym
Q
(M)-module structure on SR
Q
() easily implies that
SR
Q
()/'M` Q[x
1
, . . . , x
r
]/(I +.) = R
Q
(),
and H

T
(X

, Q)/I
T
H

T
(X

, Q) is isomorphic to H

(X

, Q) by Corollary 12.4.8.
This gives a ring isomorphism R
Q
() H

(X

, Q) that takes the class of x


i
to
the nonequivariant limit of [D
i
]
T
, which is [D
i
] by Proposition 12.4.12.
We next give a concrete example of Theorem 12.4.14.
602 Chapter 12. The Topology of Toric Varieties
Example 12.4.15. Combining Example 12.4.12 and Theorem 12.4.14, we see that
the Hirzebruch surface H
r
has equivariant cohomology ring
H

T
(H
r
, Q) SR
Q
(
r
) =Q[x
1
, x
2
, x
3
, x
4
]/'x
1
x
3
, x
2
x
4
`,
where
r
is the usual fan for H
r
. This has two interesting consequences:
H

T
(H
r
, Q) (
T
)
Q

Q
H

(H
r
, Q) by Proposition 12.4.7. This is not a ring
isomorphism, for it were, we would have an injective ring homomorphism
H

(H
r
, Q) H

T
(H
r
, Q).
This is impossible since elements of H
2
(H
r
, Q) are nilpotent in H

(H
r
, Q),
yet nonzero elements of H
2
T
(H
r
, Q) are never nilpotent in H

T
(H
r
, Q).
As a ring, H

T
(H
r
, Q) Q[x
1
, x
2
, x
3
, x
4
]/'x
1
x
3
, x
2
x
4
` does not depend on r.
However, r appears in Sym
Q
(M)-module structure from Example 12.4.12.
In Exercise 12.4.4 you will show that r is an isomorphism invariant of the
Sym
Q
(M)-algebra structure of H

T
(H
r
, Q).
In contrast, we saw in Example 12.4.5 that r is not an isomorphism invariant of the
ring structure of H

(H
r
, Q).
Preparation for the Proof . We need two lemmas and a diagram before we can
prove Theorem 12.4.14. The rst lemma embeds SR
Q
() into an exact sequence
of SR
Q
()-modules. For a cone =
i
1
+ +
it
in , let
Q[] =Q[x
i
1
, . . . , x
it
] =Q[x
i
[
i
(1)] Q[x
1
, . . . , x
r
]/'x
j
[
j
/ (1)`.
Thus we can view Q[] as a module over Q[x
1
, . . . , x
r
] where x
j
Q[] = 0 whenever

j
/ (1). Hence Q[] may also be viewed as a module over SR
Q
(), since each
generator of the ideal I acts by zero on Q[]. Then consider the sequence
(12.4.14) 0 SR
Q
()

(n)
Q[]

(n1)
Q[],
where and are dened as follows:
Since f

f [
x
i
=0 for
i
/ (1)

(n)
sends the Stanley-Reisner ideal to zero, we
get a well-dened map
([ f ]) =

f [
x
i
=0 for
i
/ (1)

(n)
for [ f ] SR
Q
().
Given f = ( f

)
(n)
and (n1), we dene the -component (f)

as
follows. Let the two cones of (n) containing be = +
i
and

= +
j
where i < j. Then
(f)

= ( f

)[
x
i
=0
( f

)[
x
j
=0
.
It is easy to see that and are SR
Q
()-module homomorphisms.
Lemma 12.4.16. The sequence (12.4.14) is exact.
12.4. The Cohomology Ring 603
Proof. We leave it to the reader to show that is injective and that = 0
(Exercise 12.4.5). If we write the sequence as 0 SR
Q
() C
n
C
n1
, then it
remains to prove exactness at C
n
.
We will use the N
r
-grading on Q[x
1
, . . . , x
r
] where the graded piece in degree
a N
r
is Q x
a
. When we regard (12.4.14) as a sequence of Q[x
1
, . . . , x
r
]-modules,
the N
r
-grading is preserved since the Stanley-Reisner ideal is generated by mono-
mials. Hence it sufces to check exactness in degree a N
r
. The case a = 0 is
easy, since an element of (C
n
)
0
is f = (

)
(n)
where

Q. Then (f) = 0
forces the

to be all equal since is a complete fan.


If a = (a
1
, . . . , a
r
) = 0, set =

a
i
>0

i
. If / , exactness holds trivially
since all graded pieces in degree a vanish. So we may assume . As in (3.2.8),
let = Star() be the fan in N()
R
= (N/N

)
R
consisting of the projections of
cones of containing . Note that is complete and simplicial since is.
By Exercise 12.4.5, SR
Q
()
a
(C
n
)
a
(C
n1
)
a
is isomorphic to
SR
Q
()
0

(nk)
Q[]
0

(nk1)
Q[]
0
,
where k = dim . This is exact by our previous observation about degree 0.
We next consider the following large diagram that will be used in proof of
Theorem 12.4.14:
(12.4.15)
0

SR
Q
()

(n)
Q[]

(n1)
Q[]
B

0

H

T
(X

(n)
H

T
(U

(n1)
H

T
(U

).
All cohomology is computed over Q in this diagram. The dotted arrow is the
isomorphism we are trying to construct. The top row is exact by Lemma 12.4.14,
and we will prove below that the square on the right commutes and that A and B
are isomorphisms. As for the bottom row, exactness can be proved using [118,
(6.3)]. Our approach is different and will in particular give an elementary proof of
the exactness of the bottom row of (12.4.15).
Let us describe the maps in the diagram. We know and , and

is induced
by the inclusions U

for (n). Similarly, if (n 1) is a facet of


(n), then on the and components in the bottom row of (12.4.15),

is
induced by the inclusion U

, with the same sign as in the denition of .


For the vertical map A, let (n). For each
i
(1), the torus-invariant
divisor D
i
U

has a class [D
i
U

]
T
H
2
T
(U

, Q). Then x
i
[D
i
U

]
T
gives a map Q[] H

T
(U

). This denes A, and B is dened similarly.


604 Chapter 12. The Topology of Toric Varieties
Lemma 12.4.17. In the diagram (12.4.15), we have:
(a) The square on the right commutes.
(b)

= 0.
(c)

is injective.
(d) The vertical maps A and B are isomorphisms.
Proof. Proposition 12.4.13 implies that [D
i
U

]
T
maps to [D
i
U

]
T
, and part (a)
follows immediately. For part (b), note that (n1) is a face of exactly two
different maximal cones and

since is complete. The commutative diagram


of inclusions
X

induces the commutative diagram of maps on equivariant cohomology


H

T
(X

, Q)

T
(U

, Q)

T
(U

, Q)

T
(U

, Q).
In Exercise 12.4.6, you will check that the component of

mapping to Q[] is the


difference of these two restrictions. It follows that

vanishes on the image of

,
which is what we wanted to show.
For part (c), the inclusions x

of the xed points for (n) give a


commutative diagram
H

T
(X

, Q)

(n)
H

T
(U

, Q)

(n)
H

T
(x

, Q)
where is the injective map from Lemma 12.4.9. Hence

is injective.
For part (d), take (n) and for simplicity assume (1) =
1
, . . . ,
n
.
Since is simplicial, D
i
U

is Q-Cartier, so that
i
D
i
U

is Cartier for some


integer
i
> 0. The Picard group of an afne toric variety is trivial, so
i
D
i
U

=
div(
m
i
) for some m
i
M. By Proposition 12.4.13, we obtain
(12.4.16)
i
[D
i
U

]
T
= [
i
D
i
U

]
T
= [div(
m
i
)]
T
=s(m
i
) 1 H
2
T
(U

, Q).
Note also that D
i
U

= div(
m
i
) implies that 'm
i
, u
j
` =
i

i j
. The u
i
are a basis
of N
Q
since (n), and hence the
1
i
m
i
are a basis of M
Q
. Thus x
i

1
i
m
i
denes an isomorphism

A

: Q[] =Q[x
1
, . . . , x
n
] Sym
Q
(M).
12.4. The Cohomology Ring 605
Let A

be the -component of A and let i

: H
2
T
(U

, Q) H
2
T
(x

, Q) =
(
T
)
Q
be induced by the inclusion of the xed point i : x

. Using the
isomorphism s : Sym
Q
(M) (
T
)
Q
from (12.4.9), we get a diagram
Q[]

T
(U

, Q)
i

Sym
Q
(M)
s

(
T
)
Q
which is commutative by (12.4.16). By Proposition 12.1.9, there is an equivariant
deformation retraction from U

to x

. Thus i

is an isomorphism. Since

A

and s
are also isomorphisms, the same is true for A

.
The proof that B is an isomorphism is similar (with some small differences)
and will be covered in Exercise 12.4.7.
We now have everything needed to compute the equivariant cohomology of X

.
Proof of Theorem 12.4.14. Consider the map Q[x
1
, . . . , x
r
] H

T
(X

, Q) dened
by x
i
[D
i
]
T
. To show that this map kills the Stanley-Reisner ideal, take distinct
indices i
1
, . . . , i
s
such that
i
1
+ +
is
/ . We need to prove that
[D
i
1
]
T
[D
is
]
T
= 0 H

T
(X

, Q).
For each (n), this cup product maps to [D
i
1
U

]
T
[D
is
U

]
T
in
H

T
(U

, Q) by Proposition 12.4.13. This vanishes in H

T
(U

, Q) since
i
j
/ (1) for
some j 1, . . . , s. Thus the map

from (12.4.15) maps [D


i
1
]
T
[D
is
]
T
to
zero. Since

is injective by Lemma 12.4.17, the cup product is zero in H

T
(X

, Q).
Thus we have a well-dened ring homomorphism : SR
Q
() H

T
(X

, Q).
Also, m M acts on SR
Q
() by multiplication by the image of

r
i=1
'm, u
i
`x
i
in
SR
Q
(). This maps to

r
i=1
'm, u
i
`D
i

T
=[div(
m
)]
T
= s(m) 1,
where the last equality follows from Proposition 12.4.13. It follows that is a
Sym
Q
(M)-algebra homomorphism.
Now add to the diagram (12.4.15):
0

SR
Q
()

(n)
Q[]

(n1)
Q[]
B

0

H

T
(X

(n)
H

T
(U

(n1)
H

T
(U

).
The argument used in in Proposition 12.4.13 shows that the square on the left
commutes. Hence we have a commutative diagram where the top row is exact
606 Chapter 12. The Topology of Toric Varieties
(Lemma 12.4.16), the maps A and B are isomorphisms (Lemma 12.4.17), and

is injective (Lemma 12.4.17). From here, an easy diagram chase shows that is
bijective, and the theorem is proved.
The above proof implies that the bottom row of our big diagram
0 H

T
(X

, Q)

(n)
H

T
(U

, Q)

(n1)
H

T
(U

, Q)
is exact. There are equivariant deformation retractions U

O() = x

and
U

O() by Proposition 12.1.9. Thus we can rewrite the exact sequence as


(12.4.17) 0 H

T
(X

, Q)

(n)
H

T
(O(), Q)

(n1)
H

T
(O(), Q).
This sequence, called a Chang-Skjelbred sequence, tells us that the structure of
H

T
(X

, Q) (and hence that of H

(X

, Q)) is completely determined by three things:


The equivariant cohomology of the xed point set
X
0
= X
T

(n)
O() =

(n)
x

.
The equivariant cohomology of the 1-dimensional orbits
X
1
=

(n1)
O().
The inclusions X
0
X
1
X

.
In [118, 7.2], Goresky, Kottwitz and MacPherson show that the same is true when
a variety X is equivariantly formal for the action of a torus and there are nitely
many xed points and 1-dimensional orbits. If X

is smooth, then the same argu-


ments can be carried out with coefcients in Z rather than in Q, and this gives a
proof of the Jurkiewicz-Danilov theorem (Theorem 12.4.4).
Piecewise-Polynomial Functions. The exact sequence (12.4.17) has an interesting
interpretation. For (n), H

T
(O(), Q) H

T
(U

, Q) Sym
Q
(M), and for
(n1), you will construct an isomorphism
H

T
(O(), Q) Sym
Q
(M

)
in Exercise 12.4.7, where M

= M/(

M). It follows that the exact sequence


for equivariant cohomology can be written
(12.4.18) 0 H

T
(X

, Q)

(n)
Sym
Q
(M)

(n1)
Sym
Q
(M

).
Now observe the following:
For each (n), Sym
Q
(M) is naturally isomorphic to the ring of polynomial
functions on with coefcients in Q.
Similarly, for each (n1), Sym
Q
(M

) is naturally isomorphic to the ring


of polynomial functions on with coefcients in Q.
12.4. The Cohomology Ring 607
The exactness of (12.4.18) means that an element of H

T
(X

, Q) can be thought
of as a collection f

(n)
, where f

is a polynomial function on , such that


f
[

= f

whenever

= (n1).
Thus we can identify the equivariant cohomology of X

with the ring of piecewise


polynomial functions with rational coefcients on the simplicial decomposition of
N
R
dened by with the usual pointwise sum and product. Such functions are also
known as polynomial splines and are an important tool in numerical analysis and
applications. This gives perhaps the most concrete way to compute and understand
the equivariant cohomology of X

.
Example 12.4.18. The torus T (C

)
2
acts as usual on P
2
. In Exercise 12.4.8,
you will verify the following claims. We write Sym
Q
(Z
2
) = Q[x, y]. First, by
(12.4.18), H

T
(P
2
, Q) is equal to the set of triples ( f
0
, f
1
, f
2
) Q[x, y]
3
such that
f
1
f
0
'x`
f
2
f
0
'y`
f
2
f
1
'y x`.
Therefore,
H

T
(P
2
, Q) =( f
0
, f
0
+xg, f
0
+yg+y(y x)h) [ f
0
, g, h Q[x, y].
This shows that as a module over Q[x, y], H

T
(P
2
, Q) is generated by (1, 1, 1),
(0, x, y), and (0, 0, y(y x)). Since we have one generator in each degree 0, 1, 2,
remembering the doubling of degrees in Theorem 12.4.1, we obtain an explicit
description of the (
T
)
Q
-module isomorphism
H

T
(P
2
, Q) (
T
)
Q

Q
H

(P
2
, Q) Sym
Q
(Z
2
)
Q
H

(P
2
, Q)
from Proposition 12.4.7 in this case.
Properties of Equivariant Cohomology Classes. The nal task of this section is to
dene the equivariant cohomology class of a torus-invariant divisor on a simplicial
toric variety (not necessarily complete) and to prove that these classes have the
properties listed in Proposition 12.4.13. Our construction will use Chern classes.
In 13.1 we give an overview of the Chern classes of locally free sheaves on
varieties. Here, we combine the algebraic approach with a more topological version
of Chern classes that applies to complex vector bundles over topological spaces.
Let D be a torus-invariant Cartier divisor on a simplicial toric variety X

. This
gives the invertible sheaf O
X

(D), which by Theorem 6.0.18 is the sheaf of sections


of a rank 1 vector bundle : V
L
X

for L =O
X

(D). Here, we write V


D
instead
of V
L
. In Proposition 7.3.1 we constructed a fan D in N
R
R such that V
D
is
the toric variety of D. Hence the torus of V
D
is T
NZ
= T
N
C

. In particular,
the torus T = T
N
of X

acts on V
D
such that the projection map is equivariant.
This makes V
D
into an equivariant vector bundle over X

.
608 Chapter 12. The Topology of Toric Varieties
In the classical topology, : V
L
X

has a Chern class c


1
(V
D
) H
2
(X

, Z),
which equals the algebraic Chern class c
1
(O
X

(D)) discussed in 13.1. Thus


(12.4.19) c
1
(V
D
) = c
1
(O
X

(D)) = [D] H
2
(X

, Z),
where the last equality uses the properties of Chern classes stated in 13.1.
To construct the cohomology class [D]
T
H
2
T
(X

, Z), recall that H

T
(X

, Z) =
H

(ET
T
X

, Z). Since : V
D
X

is T-equivariant, the rank 1 vector bundle


(V
D
)
T
= ET
T
V
D
ET
T
X

has Chern class [D]


T
= c
1

(V
D
)
T

H
2
(ET
T
X

, Z) = H

T
(X

, Z).
Proof of Proposition 12.4.13. We need to show that the classes [D]
T
satisfy:
(a) [D
1
+D
2
]
T
= [D
1
]
T
+[D
2
]
T
.
(b) [div(
m
)]
T
=s(m) 1, where 1 H
0
T
(X

, Q) and s(m) (
T
)
Q
by (12.4.9).
(c) i

[D]
T
= [D[
U
]
T
, where i : U X

is the inclusion of a torus-invariant open


subset of X

.
(d) i

[D]
T
=[D], where i

: H

T
(X

, Q) H

(X

, Q) is the nonequivariant limit


map from (12.4.13).
For part (a), note that V
D
1
+D
2
= V
D
1
V
D
2
, so (V
D
1
+D
2
)
T
= (V
D
1
)
T
(V
D
2
)
T
as vector bundles over ET
T
X

. Then we are done by the properties of Chern


classes of rank 1 vector bundles.
Part (c) is also easy. Given i : U X

, one checks that V


D[
U
is the pullback
of V
D
via i. Thus (V
D[
U
)
T
is the pullback of (V
D
)
T
via i
T
: ET
T
U ET
T
X

.
Then we are done since c
1
is compatible with pullbacks.
For part (d), recall that the map ET
T
X

BT is a bration with ber X

. If
i
X

: X

ET
T
X

is the inclusion of the ber, then V


D
is the pullback of (V
D
)
T
via i
X

. Since c
1
is compatible with pullbacks, we obtain
i

[D]
T
= i

c
1

(V
D
)
T

= c
1
(i

(V
D
)
T
) = c
1
(V
D
) = [D],
where the last equality uses (12.4.19).
The above proofs work over Z since we are assuming that D is Cartier. If D
is an arbitrary divisor on X

, then D is Cartier for some integer > 0 since X

is
simplicial. Then
[D]
T
=
1
[D]
T
H

(X

, Q)
is well-dened. It follows that properties (a), (c) and (d) hold over Q.
The proof of part (b) takes a bit more work. Take m M and set D = div(
m
).
Also let p : X

pt be the map that takes every element of X

to the same point,


and let V

m =C pt be the vector bundle over pt where T acts on C via


m
.
12.4. The Cohomology Ring 609
In Exercise 12.4.9 you will show that there is a pullback diagram
V
D

pt
of T-equivariant vector bundles. In concrete terms, this says that V
D
is the trivial
bundle X

C X

where T acts on C via


m
.
It follows that (V
D
)
T
is the pullback of (V

m)
T
via p
T
: ET
T
X

BT. For
the moment, set = c
1

(V

m)
T

H
2
T
(pt, Z) =
T
. Then
[div(
m
)]
T
= c
1

(V
D
)
T

= p

T
c
1

(V

m)
T

= p

T
= 1
The last equality follows from the denition of H

T
(X

, Z) as a
T
-module.
Hence it sufces to prove that =s(m), where s is dened in the discussion
leading up to (12.4.9). First observe that
m
: T C

induces B
m
: BT BC

.
By Exercise 12.4.9, there is a pullback diagram
(V

m)
T

(C)
C

BT
B
m

BC

,
where C pt where C

acts on C by multiplication. This implies that


(12.4.20) = c
1

(V

m)
T

= (B
m
)

c
1

(C)
C

.
So we need to compute c
1
((C)
C
) H
2
(BC

, Z). Recall that (C)


C
=EC

C
C,
where C

acts on C by multiplication. Also, EC

0
(C
+1
`0) and BC

0
P

=P

. When computing equivariant cohomology in a specic degree, one


can always restrict to a nite-dimensional approximation. Hence, for sufciently
large, c
1
((C)
C
) is the Chern class of the rank 1 vector bundle
V

= (C
+1
`0)
C
C P

.
We have seen this bundle before. Since C

acts on (C
+1
`0) C via t (p, ) =
(t
1
p, t ), the map (p, ) ([p], p) P

C
+1
is constant on C

orbits. Here
[p] denotes the point in P

whose homogeneous coordinates are p C


+1
` 0.
Hence we get a map
V

C
+1
.
In Exercise 12.4.9, you will prove that the image of this map is the vector bundle
V P

from Example 6.0.19. Thus V

V as vector bundles over P

. Since the
sheaf of sections of V is O
P
(1) by Example 6.0.21, we have
(12.4.21) c
1

(C)
C

= c
1
(V

) = c
1
(V) = c
1
(O
P
(1))
in the cohomology group
Z t = H
2
(P

, Z) = H
2
(P

, Z) = H
2
(BC

, Z).
610 Chapter 12. The Topology of Toric Varieties
Here we are using H

(P

, Z) = Z[t]/'t
+1
` and H

(P

, Z) = Z[t] from (12.4.7).


A standard fact about Chern classes is that
c
1
(O
P
(1)) = t Z t.
This is the normalization axiom for Chern classessee [144, I.4]. It follows that
c
1
(O
P
(1)) =t. Combining this with (12.4.20) and (12.4.21), we have
= c
1

(V

m)
T

= (B
m
)

c
1

(C)
C

= (B
m
)

c
1
(O
P
(1))

= (B
m
)

(t) =(B
m
)

(t)
=s(m),
where the nal equality is the denition of s(m) in (12.4.9). This completes the
proof of Proposition 12.4.13.
The proof of the proposition explains the minus sign in Proposition 12.4.13it
comes from the Chern class of the vector bundle over BC

= P

determined by
the equvariant bundle C pt where C

acts on C by multiplication.
Exercises for 12.4.
12.4.1. The following example is taken from [95].
(a) Let denote the complete fan in N
R
R
2
with minimal generators
u
1
= e
1
, u
2
= e
2
, u
3
=e
1
, u
4
=e
2
and

the complete fan with minimal generators


u

1
= e
1
, u

2
= e
2
, u

3
=e
1
+e
2
, u

4
=e
1
e
2
.
Use Theorem 12.4.1 to show that
H

(X

, Q) Q[x, y]/'xy, x
2
+y
2
`
H

(X

, Q) Q[x, y]/'xy, x
2
+2y
2
`.
(b) Deduce that the cohomology rings H

(X

, Q) and H

(X

, Q) are not isomorphic.


Hint: Show that H

(X

, Q) has no elements , of degree 2 satisng


2
=
2
= 0
and = 0.
(c) Show that if we extend scalars to R, however, then the cohomology rings become
isomorphic:
H

(X

, R) = H

(X

, Q)
Z
R H

(X

, Q)
Z
R = H

(X

, R).
The paper [95] shows that the same is true for all complete toric surfacesthe real
cohomology algebra depends only on the number of rays in the fan. This result does
not extend to higher-dimensional toric varieties.
12.4.2. In this exercise, you will show that C

` 0 is contractible.
(a) Show that C

` 0 is homotopy equivalent to the unit sphere S

, where S

is the set of all (a


0
, a
1
, . . . ) C

such that

i=0
[a
i
[
2
= 1.
(b) Show that the shift map T : S

dened by T(a
0
, a
1
, . . . ) = (0, a
0
, a
1
, . . . ) is
homotopic to the identity map id : S

. Hint: Consider the 1-parameter family


of maps (1 t)id+t T, t [0, 1], and note that T has no eigenvalues or eigenvectors.
12.4. The Cohomology Ring 611
(c) Finally, show that T is homotopic to the constant map U(a
0
, a
1
, . . . ) =(1, 0, . . .). Hint:
Consider the 1-parameter family
cos(t/2)T(a
0
, a
1
, . . .) +sin(t/2)U(a
0
, a
1
, . . .).
12.4.3. For each (n 1), let m

be a nonzero element of

M and let f (
T
)
Q
correspond to

(n1)
m

Sym
Q
(M). Prove that the map from (12.4.11) becomes
an isomorphism after localizing at f . Hint: Use (12.4.18).
12.4.4. Consider SR
Q
(
r
) = Q[x
1
, x
2
, x
3
, x
4
]/'x
1
x
3
, x
2
x
4
`, which for M = Z
2
becomes a
Sym
Q
(M)-algebra where e
1
acts as x
1
x
3
and e
1
acts as r x
1
x
2
+x
4
.
(a) SR
Q
(
r
) has graded ring automorphisms given by rescaling the variables by elements
of Q

, plus the maps induced by x


1
x
3
, by x
2
x
4
, and by (x
1
, x
3
) (x
2
, x
4
). Show
that these generate the group of all graded ring automorphisms of SR
Q
(
r
)
(b) Prove that if SR
Q
(
r
) SR
Q
(
s
) as graded Sym
Q
(M)-algebras, then r = s.
12.4.5. Complete the proof of Lemma 12.4.16.
12.4.6. In Lemma 12.4.17, show that if (n 1) is a face of ,

(n), then the


component of the map

corresponding to can be identied with the difference of the


restriction maps H

T
(U

, Q) H

T
(U

, Q) and H

T
(U

, Q) H

T
(U

, Q).
12.4.7. We will use the notation of Lemma 12.4.17. Let (n 1). Dene the map
B

: Q[] H

T
(U

, Q) by x
i
[D
i
U

]
T
for
i
(1). Also let N

= Span() N N,
with dual M M

= M/(

N).
(a) As in the proof of Lemma 12.4.17, there are
i
> 0 and m
i
M such that
i
D
i
U

=
div(
mi
). Explain why m
i
is not unique and show that x
i

1
i
[m
i
] (M

)
Q
denes
an isomorphism

B

: Q[] Sym
Q
(M

).
(b) Use Proposition 12.1.9 to show that H

T
(U

, Q) H

T
(O(), Q).
(c) Let x

be the identity element of O(). Show that T


Nt
is the isotropy subgroup of
x

O() under the action of T = T


N
.
(d) Show that N = N

, where N

Z. Conclude that T = T
N
C

, where T
Nt
acts
trivially on O() and C

acts freely and transitively on O().


(e) Suppose a product group GH acts on a space X such that G acts trivially and H acts
freely and transitively. Prove that H

GH
(X, Q) H

G
(pt, Q). Hint: Use E(GH) =
EGEH and regard GH as acting on pt X.
(f) Explain why H

T
(O(), Q) H

TN
(pt, Q) and conclude that B

is an isomorphism.
Hint: (12.4.16) holds for .
12.4.8. Verify the claims in Example 12.4.18.
12.4.9. In proof of part (b) of Proposition 12.4.13
(a) Prove that V
D
is the pullback of V

m via p : X pt.
(b) Prove that (V

m )
T
is the pullback of (C)
C
via B
m
: BT BC

.
(c) Let C

act on C by multiplication. Prove that (C


+1
` 0)
C
C P

is the vector
bundle V P

described in Example 6.0.19.


12.4.10. In the situation of Theorem 12.4.1, take and write (1) = i
1
, . . . , i
d
,
where d = dim . Then let
x

= [x
i1
x
i
d
] R
Q
()
d
.
612 Chapter 12. The Topology of Toric Varieties
In the text, was often an element of (n); here, it represents an arbitrary element of .
In this exercise you will prove that R
Q
() is spanned over Q by the x

for . (The x

may be linearly dependent, however.)


(a) Show that
x

x
+

if

=0 and +


0 if +

/ .
(b) Now consider a monomial x
a
= x
a1
i1
x
as
is
with i
1
< i
2
< < i
s
and a
j
1. We may
assume
i1
+ +
is
since otherwise [x
a
] = 0 in R
Q
(). If a
1
> 1, then use the
relations from. to show that [x
i1
] is a linear combination with coefcients in Q of the
[x

] for / i
1
, . . . , i
s
Hint: u
i1
, . . . , u
is
are part of a basis of N
Q
.
(c) Write x
a
=x
i1
x
a11
i1
x
as
is
and replace the rst x
i1
with the linear combination found in
part (b). Then repeat part (b) for each term the resulting linear combination. Show that
this process eventually expresses [x
a
] as a linear combination of cosets of square-free
monomials. This will prove our claim. Hint: Induction on

s
i=1
(a
i
1).
12.5. The Chow Ring and Intersection Cohomology
Here we explore two alternative descriptions of the cohomology ring of a complete
simplicial toric variety. The rst involves the Chow ring, while the second uses
intersection cohomology.
The Chow Ring. We will show that the description of the cohomology ring given
in Theorem 12.4.1 applies equally well to the rational Chow ring A

(X

)
Q
of a
complete simplicial toric variety X

. We begin by sketching the construction of


the Chow ring. We refer the reader to [107] for the details.
Let Z
k
(X) denote the group of k-dimensional algebraic cycles on a variety X.
In other words, Z
k
(X) is the additive group of nite formal linear combinations

i
n
i
[V
i
]
of irreducible subvarieties V
i
of dimension k with coefcients n
i
Z. A k-cycle
on X is dened to be rationally equivalent to zero if there are nitely many
(k+1)-dimensional irreducible subvarieties W
i
X and nonzero rational functions
f
i
C(W
i
) such that
=

i
[div
W
i
( f
i
)],
where div
W
i
( f
i
) is the divisor of the rational function f
i
on W
i
. Note that W
i
may
fail to be normal, so this requires a more general denition of div( f
i
) than the one
given in 4.0. The cycles rationally equivalent to zero form a subgroup Rat
k
(X) of
Z
k
(X) (Exercise 12.5.1). Then the group of k-cycles modulo rational equivalence
is dened to be
A
k
(X) = Z
k
(X)/Rat
k
(X).
12.5. The Chow Ring and Intersection Cohomology 613
When X is normal of dimension n, A
n1
(X) is the group Cl(X) of Weil divisors
modulo linear equivalence dened in 4.0 (Exercise 12.5.1).
For a general n-dimensional variety X, the group of cycles of codimension k
modulo rational equivalence is dened to be A
k
(X) = A
nk
(X). If in addition X is
smooth and projective, then with a substantial amount of work (see [107, Ch. 8]),
it can be shown that there is a product
A
k
(X) A

(X) A
k+
(X),
which coincides with the geometric intersection of cycles in the case of transverse
intersections. We will write the product as for A
k
(X) and A

(X).
It can be shown that this intersection product is commutative and associative, and
this makes A

(X) =

n
k=0
A
k
(X) into a graded ring, called the Chow ring of X.
A

(X) can be viewed as a sort of algebraic version of the integral cohomology ring
H

(X, Z). The intersection of cycles corresponds to cup product in cohomology,


so there is a ring homomorphism
(12.5.1) A

(X) H

(X, Z)
dened as follows. Let n = dim X. Then a cycle A
k
(X) = A
nk
(X) gives a
homology class in H
2n2k
(X, Z) whose Poincar e dual is in H
2k
(X, Z). In particular,
the map (12.5.1) doubles degrees.
The Toric Case. If X

is a complete simplicial toric variety of dimension n, then


the intersection product can be dened on rational cycles, making
A

(X

)
Q
= A

(X

)
Z
Q=
n

k=0
A
k
(X

)
Z
Q
into a graded ring. The key point here is that X

has a covering by open afne


varieties of the form C
n
/G for some nite group G. In this case,
A

(C
n
/G)
Q
A

(C
n
)
G
Q
,
the G-invariant subring. As in the smooth case, each subvariety of codimension
k in X

denes a class in H
2n2k
(X

, Q), hence by Poincar e duality, a class in


H
2k
(X

, Q). Hence we also have a ring homomorphism


(12.5.2) A

(X

)
Q
H

(X

, Q).
For each , the orbit closure V() is a subvariety of codimension dim.
Write [V()] for its rational equivalence class in A
dim
(X

).
Lemma 12.5.1. The [V()] for generate A

(X

) as an abelian group.
Proof. We will show that the [V()] for of dimension nk generate A
k
(X

). In
general, when Y X is a closed subvariety of a variety X, there is a short exact
sequence of Chow groups
(12.5.3) A
k
(Y) A
k
(X) A
k
(X `Y) 0,
614 Chapter 12. The Topology of Toric Varieties
where the rst map comes from the inclusion Y X, and the second comes by
restriction. See [107, Prop. 1.8]) for a proof. If X is normal of dimension n, then
(12.5.3) for k = n1 is precisely Theorem 4.0.20 (Exercise 12.5.2).
Consider the ltration of X

as in (12.3.3),
= X
1
X
0
X
1
X
n
= X

,
where X
p
is the union of the closures V() of the torus orbits of dimension p. By
the Orbit-Cone Correspondence, we have
X
p
`X
p1
=

(np)
O().
Then (12.5.3) gives an exact sequence
A
k
(X
p1
) A
k
(X
p
)

(np)
A
k
(O()) 0.
But O() is a p-dimensional torus when (np) and hence can be viewed as
an afne open subset of C
p
. In Exercise 12.5.3 you will show that
A
k
(C
p
) =

Z[C
p
] k = p
0 otherwise.
Using (12.5.3) for the inclusion O() C
p
, we see that
A
k
(O()) =

Z[O()] k = p
0 otherwise.
When p > k, it follows that the above exact sequence simplies to
A
k
(X
p1
) A
k
(X
p
) 0.
Now x k and let p increase from k to n. It follows that A
k
(X
k
) surjects onto
A
k
(X

). By Exercise 12.5.2, we also have


A
k
(X
k
) =

(nk)
Z[V()]
since X
k
=

(nk)
V() and each V() is irreducible of dimension k. The
lemma now follows easily.
When (1), the orbit closure V() is the divisor denoted D

. For m M,
recall from Proposition 4.1.2 that
div(
m
) =

'm, u

`D

.
Since A
1
(X

) = Cl(X

) by Exercise 12.5.2, we obtain


(12.5.4)

'm, u

`[D

] = 0 in A
1
(X

).
Here is an example of how to compute intersection products in the Chow ring.
12.5. The Chow Ring and Intersection Cohomology 615
Lemma 12.5.2. Assume that X

is complete and simplicial. If


1
, . . . ,
d
(1)
are distinct, then in A

(X

)
Q
, we have
[D

1
] [D

2
] [D

d
] =

1
mult()
[V()] if =
1
+ +
d

0 otherwise.
Proof. Let (1) and be a cone of not containing . We will show that
(12.5.5) [D

] [V()] =

mult()
mult()
[V()] if = +
0 otherwise.
In Exercise 12.5.4 you will prove that the lemma follows from (12.5.5). You should
also check that (12.5.5) generalizes the intersection formulas from parts (a) and (b)
of Proposition 6.4.4.
We noted earlier that intersection products can be subtle. Life is easier when
intersecting with a divisor, but even here, some care is needed. The relevant theory
is developed in [107, 2.3]. We will use the following special case. Let D be a
Cartier divisor on X

and let i : V X

be a normal subvariety of dimension k.


The line bundle O
X

(D) pulls back to a line bundle i

O
X

(D) on V, so that by
Theorem 6.0.20, i

O
X

(D) O
V
(D

) for some Cartier divisor D

on V. Then D

is
a (k 1)-cycle on X

and by [107, 2.3] gives the intersection product


(12.5.6) [D] [V] = [D

].
In Exercise 12.5.5 you will use this denition to show that [D] [V] = 0 when
Supp(D) V =.
Since X

is simplicial, D

is Q-Cartier, so there is some positive multiple D =


D

that is Cartier on X

. If + / , then Supp(D) V() = D

V() = by
the Orbit-Cone Correspondence. Then [D] [V()] = 0 by the previous paragraph,
which implies [D

] [V()] = 0. Hence we may assume that = + . We


need to prove that
(12.5.7) [D] [V()] =
mult()
mult()
[V()].
The rst step is to nd a Cartier divisor D

on V() such that i

O
X

(D)
O
V()
(D

), where i : V() X

is the inclusion. We will use the support function

D
: N
R
R of D, which satises

D
(u

) =

if

=
0 otherwise.
Note in particular that
D
vanishes on Span() since / (1).
Recall from Proposition 3.2.7 that V() is the toric variety of the fan Star()
in N()
R
, where N() =N/(Span()N). The previous paragraph shows that
D
616 Chapter 12. The Topology of Toric Varieties
factors through the quotient map N
R
N()
R
. By Exercise 12.5.6, the induced
map N()
R
R is the support function
D
of the divisor D

we want.
The rays of Star() are images of rays of , so that
D
vanishes on all rays of
Star() except for = + = . Thus D

is the divisor on V() given by


D

=
D

(u

)D

,
where u

is the minimal generator of N(). In particular, the image of u

in
N() is u

= u

for some integer . Thus

D
(u

) =
1


D
(u

) =
1


D
(u

) =
1


D
(u

) =

.
Since = mult()/mult() by the proof of Lemma 6.4.2, we obtain
D

=
mult()
mult()
D

.
However, since = +, the divisor D

in V() is the subvariety V() in V().


From here, (12.5.7) follows easily, and the proof is complete.
Similar to what happened in 6.4, Lemma 12.5.2 shows that rational numbers
appear naturally when doing intersection theory on simplicial toric varieties.
The Chow Ring of a Toric Variety. As in 12.4, write (1) =
1
, . . . ,
r
. This
gives the ring
R
Q
() =Q[x
1
, . . . , x
r
]/(I +.)
for I and . as in (12.4.2) and (12.4.3). Then Lemma 12.5.2 and (12.5.4) imply
that [x
i
] [D

i
] A
1
(X

)
Q
denes a ring homomorphism
(12.5.8) R
Q
()
Q
A

(X

)
Q
.
We also have the ring homomorphism A

(X

)
Q
H

(X

, Q) from (12.5.2).
Theorem 12.5.3. If X

is complete and simplicial, then


R
Q
()
Q
A

(X

)
Q
H

(X

, Q),
where the maps are given by (12.5.8) and (12.5.2).
Proof. The composition of the maps R
Q
()
Q
A

(X

)
Q
H

(X

, Q) from
(12.5.8) and (12.5.2) is given by [x
i
] [D

i
] H
2
(X

, Q). Since this denes the


isomorphism R
Q
()
Q
H

(X

, Q) from Theorem 12.4.1, it follows that the map


R
Q
()
Q
A

(X

)
Q
is injective.
For surjectivity, take =
i
1
+ +
i
d
. By Lemma 12.5.2, [x
i
1
x
i
d
] =
[x
i
1
] [x
i
d
] maps to the product
[D

i
1
] [D

i
d
] =
1
mult()
[V()].
12.5. The Chow Ring and Intersection Cohomology 617
Then surjectivity follows since the [V()] span A

(X

)
Q
by Lemma 12.5.1. Hence
(12.5.8) is an isomorphism, and then the same must be true for (12.5.2).
The isomorphism here explains the connection between the cup product in
cohomology and the intersection formula from (12.4.1). We should also note that
there are isomorphisms
(12.5.9) R() A

(X

) H

(X

, Z)
when X

is smooth, where R() is the ring from Theorem 12.4.4.


There is also an equivariant Chow ring A

T
(X

) associated to the action of


T = T
N
on X

. If X

is not simplicial, then Payne has shown in [224] that A

T
(X

)
still coincides with the ring of continuous piecewise polynomial functions on the
polyhedral decomposition of N
R
dened by , even over Z. However, this ring
need not coincide with equivariant cohomology ring in this case.
An Exact Sequence for the Chow Group. The proof of Lemma 12.5.1 gives a
surjective map :

(nk)
Q[V()] A
k
(X

)
Q
. In [111], Fulton and Sturmfels
describe generators of the kernel using the sequence
(12.5.10)

(nk1)
M()
Q

(nk)
Q[V()]

A
k
(X

)
Q
0.
To dene , take (nk 1), so dimV() = k +1, and note that M() is the
character group of the torus of the toric variety V(). If m M(), then
m
is a
rational function on V() and hence div(
m
) is a divisor on V(). This gives the
k-cycle (m) = [div(
m
)] on X

. In Exercise 12.5.7 you will show that


(12.5.11) (m) =

(nk),
'm, u
,
`[V()],
where u
,
generates N

/N

Z. Then the denition of rational equivalence


guarantees that (12.5.10) is a complex, i.e., = 0.
Theorem 12.5.4. The complex (12.5.10) is exact if X

is complete and simplicial.


Proof. Since is surjective, (12.5.10) gives a surjective map coker() A
k
(X

)
Q
.
Now consider the dual of , which can be written

(nk)

0
M()
Q

(nk1)

1
M()
Q

Since dim M()


Q
= k and dim M()
Q
= k +1, we have

0
M()
Q

k
M()
Q

k
N()
Q

1
M()
Q
)

k
M()
Q

k+1
N()
Q
.
618 Chapter 12. The Topology of Toric Varieties
Picking bases of N() and N() and thinking in terms of orientation coefcients,
one sees that

can be identied with the map


C
k
(,

k
)
Q
=

(nk)

k
M()
Q

k
C
k+1
(,

k
)
Q
=

(nk1)

k
M()
Q
,
where C

(,

k
) is the complex dened in the discussion preceding Lemma 12.3.3.
The spectral sequence E
p,q
1
H
p+q
(X

, Q) from (12.3.11) has E


p,q
1
=C
p
(,

q
)
Q
by (12.3.12). Since E
k1,k
1
=C
k1
(,

k
) = 0 for dimension reasons, we obtain
E
k,k
2
= ker(E
k,k
1
E
k+1,k
1
) = ker(
k
) = ker(

).
Theorem 12.3.11 implies that E
k,k
2
H
2k
(X

, Q), which in turn is isomorphic to


H
2n2k
(X

, Q) A
nk
(X

)
Q
by Poincar e duality and Theorem 12.5.3. Thus
dim coker() = dim ker(

) = dim A
nk
(X

)
Q
= dim A
k
(X

)
Q
.
It follows that the surjective map coker() A
k
(X

)
Q
is an isomorphism. This
proves that (12.5.10) is exact.
The result proved in [111, Prop. 2.1] is stronger than Theorem 12.5.4 since it
holds over Z and applies to arbitrary toric varieties. Note also that when k = n1,
the exact sequence (12.5.10) becomes the familiar exact sequence
M
Q

Q[D

]

A
n1
(X

)
Q
= Cl(X

)
Q
0
from Chapter 4. For general k, you should check that (12.5.10) can be regarded as
the purely toric version of rational equivalence.
Intersection Cohomology. As we saw in 12.3, most of the nice properties of
the cohomology ring H

(X

, Q) for simplicial X

(the vanishing of the odd-degree


groups, Poincar e duality, and so forth) can fail for nonsimplicial toric varieties. The
theory of intersection homology and cohomology was rst developed by Goresky
and MacPherson in the 1970s to study spaces with singularities more complicated
than nite quotient singularities, and many of the nice properties are recovered in
this cohomology theory. See [175] for a general introduction.
Intersection homology was originally dened for spaces with a special sort of
stratication (see [119] and [120]). In the following, cone(S) is the open topolog-
ical cone over S, namely the quotient space of S [0, 1) modulo the relation that
identies all (s, 0) for s S.
Denition 12.5.5. A topological pseudomanifold X is a paracompact Hausdorff
space with a ltration
X
0
X
1
X
n
= X
such that:
12.5. The Chow Ring and Intersection Cohomology 619
(a) For each point p in X
i
`X
i1
, there is a neighborhood N of p in X, a compact
Hausdorff space L with an ni 1 dimensional topological stratication
= L
1
L
0
L
ni1
= L,
and a homeomorphism : R
i
cone(L) N that maps each R
i
cone(L
j
)
homeomorphically to N X
i+j+1
.
(b) X
n1
= X
n2
.
When the rst condition holds, the second condition implies that X ` X
n2
is
an n-dimensional real manifold. It is known that all projective or quasiprojective
varieties over C are topological pseudomanifolds.
A perversity is a function p : 2, 3, . . . N such that p(2) = 0 and both p(k)
and k 2 p(k) are nonnegative and nondecreasing functions of k. The comple-
mentary perversity of p is q(k) =k2p(k). The constant function min(k) =0 is
the minimal perversity; its complement is the maximal perversity max(k) = k 2.
The middle perversity is p(k) =
k2
2
. Perversities are used to specify allowable
intersections of cycles with the strata X
i
in the denition of intersection homology.
Denition 12.5.6. The intersection homology groups IH
p
i
(X) of a topological
pseudomanifold X with respect to a given perversity p are the homology groups
of a subcomplex IC
p

(X) of the singular chains on X as follows. A singular i-


simplex :
i
X is said to be p-allowable if for every k,
1
(X
nk
`X
nk1
) is
contained in the union of the faces of
i
of dimension i k+p(k). Then a singular
chain is said to be p-allowable if it is a linear combination of p-allowable singular
simplices. Finally, IC
p
i
(X) consists of all p-allowable i-chains c such that c is
also p-allowable.
Following an idea of Deligne, this denition was later reformulated using the
derived category of the category of sheaves of Z-modules on X (see [120]). This
way of framing the theory has numerous technical advantages, but we will not
pursue this approach here. The denition of the IH
p
i
(X) appears to depend on the
stratication, but Goresky and MacPherson proved that the IH
p
i
(X) are actually
topological invariants of X.
Example 12.5.7. Let X be an irreducible variety of dimension n with a single
isolated singular point x. For the middle perversity, it is not difcult to see that
IH
i
(X) =

H
i
(X) i > n
im(H
i
(X `x) H
i
(X)) i = n
H
i
(X `x) i < n.
You will prove this in Exercise 12.5.8.
The following theorem summarizes the properties of intersection homology
that we will discuss.
620 Chapter 12. The Topology of Toric Varieties
Theorem 12.5.8. Let X be a normal projective variety of dimension n.
(a) IH
p
i
(X) is nitely generated for any i and any perversity p.
(b) For the minimal and maximal perversities, we have
IH
min
i
(X) H
2ni
(X, Z), IH
max
i
(X) H
i
(X, Z).
Now let p be the middle perversity and dene the intersection cohomology groups
as IH
i
(X)
Q
= IH
p
2ni
(X)
Z
Q. Then:
(c) Poincar e duality holds for the intersection cohomology. In particular,
dim IH
i
(X)
Q
= dim IH
2ni
(X)
Q
.
(d) IH

(X)
Q
is a module over the cohomology ring H

(X, Q).
(e) Let H
2
(X, Q) be the class of an ample divisor on X. Then for all 0 i n,
multiplication by
ni
denes an isomorphism

ni
: IH
i
(X)
Q
IH
2ni
(X)
Q
.
This is the hard Lefschetz theorem.
(f) Let be as in part (e). Then for 0 i < n, multiplication by
: IH
i
(X)
Q
IH
i+2
(X)
Q
is injective.
See [120] for parts (a)(d) of this theorem and [23] for the hard Lefschetz
theorem stated in part (e). Note that part (f) follows from part (e).
The intersection cohomology IH

(X)
Q
does not have a natural ring structure.
However, when X

is a complete simplicial toric variety, Comment 6.4 in [119]


implies that the odd-degree intersection cohomology groups are zero and that
(12.5.12) IH
2i
(X

)
Q
H
2i
(X

, Q) A
i
(X

)
Q
, i = 0, . . . , n.
Applications to Polytopes. In 1980, Stanley [257] used toric varieties and their
cohomology to prove the McMullen conjecture for a simplicial polytope P. His
proof applied hard Lefschetz to the toric variety X

, where is the fan consisting


of the cones over the proper faces of P. This is the normal fan of the dual polytope
P

, which is simple. Hence X

comes from a simplicial fan and we know hard


Lefschetz is true in this case by (12.5.12) and part (e) of Theorem 12.5.8. As
explained in Appendix A, Stanleys paper was an important event in the history of
toric geometry.
Let f
P
0
, . . . , f
P
n
denote the face numbers of an n-dimensional polytope P. If
P is simplicial, then (9.4.6), Exercise 9.4.10 and Theorem 12.3.12 imply that the
numbers
(12.5.13) h
k
(P) =
n

i=k
(1)
ik

i
k

f
P
ni1
=
n

i=k
(1)
ik

i
k

f
P

i
12.5. The Chow Ring and Intersection Cohomology 621
are the even Betti numbers b
2k
= dim H
2k
(X

, Q) of X

. Hence they satisfy the


Dehn-Sommerville equations h
k
(P) = h
nk
(P) by Poincar e duality for the simpli-
cial toric variety X

. Also observe that


1 = h
0
(P) h
1
(P) h
m
(P), m =
n
2

by the injectivity of part (f) of Theorem 12.5.8. This monotone property of the
h
k
(P) is used by Stanley in [257].
For a more general (i.e., nonsimplicial) polytope P, some of the h
k
(P) in
(12.5.13) can be negative, so a different approach is necessary if we want to relate
these numbers to cohomology groups. Following [258], we dene polynomials
h(P, t) and g(P, t) recursively as follows. Let g(, t) = 1 and dim = 1. Now
assume that g(Q, t) has been dened for all polytopes Q of dimension < n. Then,
for a polytope P of dimension n, dene
h(P, t) =

QP
g(Q, t)(t 1)
n1dim Q
,
where is sum is over all proper faces Q of P, including Q =. Then write h(P, t) =

n
k=0
h
k
(P)t
k
and dene
g(P, t) =
n/2

k=0
(h
k
(P) h
k1
(P))t
k
.
In particular, the numbers h
k
(P) are dened for any polytope P, and one can show
that they agree with (12.5.13) when P is simplicial. We call
h(P, t) =
n

k=0
h
k
(P)t
k
the h-polynomial of P. Here is a result from [258].
Theorem 12.5.9. Let P be a full dimensional lattice polytope in R
n
containing the
origin as an interior point and let be the fan whose cones are the cones over the
proper faces of P. Then:
(a) The h
k
(P) are the intersection cohomology Betti numbers of X

, i.e.,
h
k
(P) = dim IH
2k
(X

)
Q
.
(b) h
k
(P) = h
nk
(P) for k = 0, . . . , n.
(c) Let m =
n
2
. Then
1 = h
0
(P) h
1
(P) h
m
(P).
The equalities in part (b) of the theorem generalize the Dehn-Sommerville
equations. They follow from part (a) and Poincar e duality for intersection coho-
mology. Similarly, part (c) follows from part (f) of Theorem 12.5.8, which in turn
is a consequence of hard Lefschetz.
622 Chapter 12. The Topology of Toric Varieties
Motivated by this application to polytopes, a purely combinatorial version of
intersection cohomology for toric varieties has been studied by many researchers,
including Barthel, Brasselet, Bressler, Fieseler, Karu, Kaup and Lunts. See [42]
for a discussion of this topic and references to the original papers.
Exercises for 12.5.
12.5.1. This exercise will study rational equivalence on a variety X of dimension n.
(a) Show that Rat
k
(X) is a subgroup of Z
k
(X) for all k.
(b) When X is normal, prove that two (n 1) cycles on X are rationally equivalent if and
only if they are linearly equivalent as dened in 4.0. Conclude that A
n1
(X) =Cl(X).
(c) Explain why A
n1
(X) = Pic(X) when X is smooth.
12.5.2. Let X be a variety of dimension n.
(a) In general, X may be reducible and may have irreducible components of dimension
strictly smaller than n. Prove that A
n
(X) = Z
n
(X) =

Y
Z[Y], where the sum is over
all irreducible components Y X of dimension n.
(b) Assume in addition that X is normal. Then show that the exact sequence (12.5.3) is
Theorem 4.0.20 when k = n 1.
12.5.3. Show directly from the denitions that the Chow groups of afne space are
A
k
(C
p
) =

Z[C
p
] if k = p
0 otherwise.
12.5.4. Prove that Lemma 12.5.2 follows from (12.5.5).
12.5.5. Use (12.5.6) to prove that [D] [V] = 0 when Supp(D) V =. Hint: Explain why
Supp(D) V = implies that i

O
X
(D) is trivial on V.
12.5.6. In the proof of Lemma 12.5.2, we claimed that
D
: N
R
R is the composition of
N
R
N()
R
and
D
: N()
R
R. Prove this. Hint: Adapt the proof of Proposition 6.2.7.
12.5.7. Prove (12.5.11) when (n k 1) and m M().
12.5.8. Use the denition of intersection homology to prove the formula for IH
i
(X) given
in Example 12.5.7.
12.5.9. Let X be a complete smooth variety of dimension n. Given a Cartier divisor D and
a smooth curve C on X, we dened the intersection number D C in 6.3. In this exercise
you will use (12.5.6) to prove that
D C =

X
[D] [C],
where [D] and [C] are the corresponding cohomology classes in H

(X, Z).
(a) Explain why a point p X gives a class [p] H
2n
(X, Z) such that

X
[x] = 1.
(b) Let i : C X be the inclusion map and suppose that i

O
X
(D) = O
C

s
i=1

i
p
i

for

i
Z and p
i
C. Explain why D C =

s
i=1

i
and [D] [C] =

s
i=1

i
[p
i
] in A

(X).
(c) Use the ring homomorhpismA

(X) H

(X, Z) to prove D C =

X
[D] [C].
12.5.10. Explain how Exercise 12.4.10 relates to the results of this section.
Chapter 13
Toric Hirzebruch-
Riemann-Roch
The Hirzebruch-Riemann-Roch theorem (called HRR) shows how to compute the
Euler characteristic of a coherent sheaf F on a smooth complete n-dimensional
variety X using intersection theory. The formula is
(13.0.1) (F) =

X
ch(F)Td(X).
The term on the left is the Euler characteristic of F from Chapter 9, dened by
(F) = (X, F) =
n

i=0
(1)
i
dim H
i
(X, F).
The right-hand side of (13.0.1) introduces two objects we have not seen before, the
Chern character ch(F) H

(X, Q) and the Todd class Td(X) H

(X, Q). The


integrand on the right-hand side of (13.0.1) is the cup product ch(F)Td(X),
which is written ch(F)Td(X) or ch(F) Td(X) in practice.
This chapter will prove HRR for a line bundle L = O
X
(D) on a smooth
complete toric variety X = X

. Our strategy will be to rst prove an equivariant


Riemann-Roch theorem for X and then derive HRR via the nonequivariant limit.
In 13.1, we begin by dening the terms involved in (13.0.1) and working
out some specic examples of HRR. To prepare for the equivariant version of
Riemann-Roch, we discuss some remarkable equalities due to Brion in 13.2. Then
13.3 proves equivariant Riemann-Roch in the smooth case, following the paper
[50] of Brion and Vergne. The nal two sections 13.4 and 13.5 apply these ideas
to lattice points in polytopes, including the volume polynomial and the Khovanskii-
Pukhlikov theorem. We will also sketch a second proof of HRR.
623
624 Chapter 13. Toric Hirzebruch-Riemann-Roch
13.1. Chern Characters, Todd Classes, and HRR
In order to state our version of the toric HRR, we need to dene Chern characters
and Todd classes. Both of these require knowledge of Chern classes, which we
now discuss.
Chern Classes. A locally free sheaf E of rank r on a variety X has Chern classes
c
i
(E ) H
2i
(X, Z) for 0 i r. Its total Chern class is c(E ) =c
0
(E )+ +c
r
(E ).
Here are the key properties we will need:
c
0
(E ) = 1 for any E and c(O
X
) = 1 for any X.
c
i
( f

E ) = f

c
i
(E ) when f : X Y is a morphism and E is locally free on Y.
c(F) = c(E ) c(G) when 0 E F G 0 is an exact sequence.
c
i
(E

) = (1)
i
c
i
(E ), where E

is the dual of E .
c
1
(

r
E ) = c
1
(E ) when E has rank r.
If D is a Cartier divisor on a smooth complete variety X, then
c
1
(O
X
(D)) = [D] H
2
(X, Z).
We refer the reader to [107] or [144] for a further discussion of Chern classes. Later
in the section we will say a few words about Chern classes of coherent sheaves.
Here is a nice example that shows how to use the properties of Chern classes.
Example 13.1.1. Let X = X

be a smooth complete toric variety. Its cotangent


bundle
1
X
ts into the generalized Euler sequence
(13.1.1) 0
1
X

O
X
(D

) Pic(X)
Z
O
X
0
from Theorem 8.1.6. Since Pic(X)
Z
O
X
O
rn
X
for r = [(1)[, the total Chern
class of Pic(X)
Z
O
X
is 1. Hence
c(
1
X
) = c

O
X
(D

c(O
X
(D

)) =

(1[D

]),
where

means cup product in H

(X, Z). Thus c


1
(
1
X
) = [

].
Let dim X = n. Then the canonical bundle
X
=
n
X
has rst Chern class
c
1
(
X
) = c
1
(

1
X
) = c
1
(
1
X
) = [

].
This gives a Chern class proof that X has canonical divisor K
X
=

.
The Chern classes of the tangent bundle T
X
of a smooth complete variety X
are called the Chern classes of X. We often write c
i
= c
i
(T
X
). As we will soon
see, the Todd class of X is a special polynomial in the c
i
. The top Chern class is
c
n
= c
n
(T
X
), n = dim X, which is usually called the Euler class of X. The name
comes from the well-known formula
(13.1.2)

X
c
n
(T
X
) = e(X),
13.1. Chern Characters, Todd Classes, and HRR 625
where e(X) =

2n
i=0
(1)
i
rank H
i
(X, Z) is the topological Euler characteristic
of X. See, for example, [144, Thm. 4.10.1].
In the toric case, the Euler class is easy to compute. Recall from (12.5.9) that
H

(X, Z) is isomorphic to the Chow ring A

(X) when X is a smooth complete toric


variety. In particular, a point of X gives a class [pt] A
n
(X) = H
2n
(X, Z), where
n = dim X. Then we have the following result.
Proposition 13.1.2. Let X = X

be a smooth complete n-dimensional toric variety


with tangent bundle T
X
. Then:
(a) c(T
X
) =

(1+[D

]) =

[V()].
(b) c
1
= c
1
(T
X
) = [

] = [K
X
].
(c) c
n
= c
n
(T
X
) = [pt], where =[(n)[.
Proof. The rst equality of part (a) follows easily by taking the dual of the exact
sequence (13.1.1) and applying the method of Example 13.1.1. Next observe that
Lemma 12.5.2 implies that if
1
, . . . ,
n
(1) are distinct, then
[D

1
] [D
n
] =

[V()] =
1
+ +
n
(n)
0 otherwise,
since X is smooth. Then multiplying out

(1+[D

]) gives the second equality.


Part (b) follows, and for part (c), note that if (n), then V() is a point, so
that [V()] = [pt]. From here, the formula for c
n
follows immediately.
Since

X
[pt] = 1, (13.1.2) and Proposition 13.1.2 imply that the topological
Euler characteristic of X is
e(X) =

X
c
n
(T
X
) =

X
[pt] = =[(n)[.
This agrees with the result proved in Theorem 12.3.9.
Using Chern classes, it is easy to dene the Chern character of a line bundle.
Denition 13.1.3. Let X be a variety of dimension n. The Chern character of a
line bundle L is
(13.1.3) ch(L) = 1+c
1
(L) +
c
1
(L)
2
2!
+
c
1
(L)
3
3!
+ H

(X, Q).
Truncating the Taylor series for e
x
in degrees greater than n and then substituting
c
1
(L) H
2
(X, Q) for x shows that we may write this more compactly as e
c
1
(L)
.
We will see later that Chern characters can also be dened for coherent sheaves.
626 Chapter 13. Toric Hirzebruch-Riemann-Roch
The Todd Class. To dene the Todd class Td(X) of a smooth complete variety X
of dimension n, we begin by introducing symbolic objects
1
, . . . ,
n
such that
c(T
X
) =
n

i=1
(1+
i
).
We regard the
i
as living in H
2
(X, Q). The
i
are the Chern roots of the tangent
bundle T
X
. In practical terms, this means that c
i
= c
i
(T
X
) =
i
(
1
, . . . ,
n
), where

i
is the ith elementary symmetric polynomial.
Denition 13.1.4. Let X be a smooth complete n-dimensional variety and let

1
, . . . ,
n
be the Chern roots of its tangent bundle. Then the Todd class of X is
Td(X) =
n

i=1

i
1e

i
.
When we expand the right-hand side of this formula in H

(X, Q), we get a


symmetric polynomial in the
i
with rational coefcients. The theory of symmetric
polynomials implies that this is a polynomial expression in c
i
=
i
(
1
, . . . ,
n
) for
1 i n. Hence the Todd class of X is a polynomial in its Chern classes c
1
, . . . , c
n
.
The key fact is that the same polynomial is used for all smooth complete varieties
of dimension n.
To compute these polynomials, we use the well-known power series expansion
(13.1.4)
x
1e
x
=1+
1
2
x+

k=1
(1)
k1
B
k
(2k)!
x
2k
=1+
1
2
x+
1
12
x
2

1
720
x
4
+ ,
where B
1
=
1
6
, B
2
=
1
30
, B
3
=
1
42
, . . . are the Bernoulli numbers (see [144, p. 13]).
Here is a classic example.
Example 13.1.5. A smooth complete surface X has Chern classes c
1
=
1
+
2
and
c
2
=
1

2
, where
1
,
2
are the Chern roots. Since the Chern roots have degree 2
and all monomials in
1
,
2
of total degree 3 vanish (X is a surface), we obtain
Td(X) = (1+
1
2

1
+
1
12

2
1
)(1+
1
2

2
+
1
12

2
2
)
= 1+
1
2
(
1
+
2
) +
1
12
(
2
1
+3
1

2
+
2
2
)
= 1+
1
2
c
1
+
1
12
(c
2
1
+c
2
).
When X has dimension n, one can show that Td(X) = T
0
+ +T
n
such that T

is a weighted homogeneous polynomial of weighted degree in c


1
, . . . , c

, provided
13.1. Chern Characters, Todd Classes, and HRR 627
c
i
has weight i. We call T
i
the ith Todd polynomial. The rst few are
T
0
= 1, T
1
=
1
2
c
1
, T
2
=
1
12
(c
2
1
+c
2
), T
3
=
1
24
c
1
c
2
T
4
=
1
720
(c
4
1
+4c
2
1
c
2
+3c
2
2
+c
1
c
3
c
4
).
You will compute T
3
in Exercise 13.1.1. See [144, pp. 1314] for more on Todd
polynomials. At the end of the section we will say more about the intuition behind
Denition 13.1.4.
The Todd Class of a Smooth Toric Variety. In the toric setting, the cohomology
ring is generated by the classes [D

] for (1) by Theorems 12.4.1 and 12.4.4.


The Todd class can be expressed in terms of these classes as follows.
Theorem 13.1.6. The Todd class of a smooth complete toric variety X = X

is
Td(X) =

(1)
[D

]
1e
[D]
H

(X, Q).
Proof. Recall that c(T
X
) =

(1+[D

]) by Proposition 13.1.2. If we set (1) =

1
, . . . ,
r
and D
i
= D

i
, then
c
i
=
i
([D
1
], . . . , [D
r
]).
On the other hand, the Chern roots
i
satisfy c
i
=
i
(
1
, . . . ,
n
), where n = dim X.
So [D
1
], . . . , [D
r
] behave like the Chern roots, except that r > n since r = [(1)[
and is a complete fan in N
R
R
n
.
We can avoid this problem by computing

n
i=1

i
1e

i
and

r
i=1
[D
i
]
1e
[D
i
]
purely
symbolically. We rst consider the formal power series ring Q[[x
1
, . . . , x
r
]] with
maximal ideal m = 'x
1
, . . . , x
r
`. Then there is a polynomial P of n variables such
that
(13.1.5)
r

i=1
x
i
1e
x
i
P

1
(x
1
, . . . , x
r
), . . . ,
n
(x
1
, . . . , x
r
)

mod m
n+1
.
We can stop at n since
i
(x
1
, . . . , x
r
) m
n+1
for i > n. Setting x
i
= [D
i
], we obtain
r

i=1
[D
i
]
1e
[D
i
]
= P(c
1
, . . . , c
n
).
Now let n = 'x
1
, . . . , x
n
` Q[[x
1
, . . . , x
n
]]. Since
x
i
1e
x
i
= 1 +
1
2
x
i
+ equals 1
when x
i
= 0, we see that setting x
i
= 0 for n +1 i r in (13.1.5) gives the
identity
n

i=1
x
i
1e
x
i
P

1
(x
1
, . . . , x
n
, 0, . . . , 0), . . . ,
n
(x
1
, . . . , x
n
, 0, . . . , 0)

mod n
n+1
P

1
(x
1
, . . . , x
n
), . . . ,
n
(x
1
, . . . , x
n
)

mod n
n+1
628 Chapter 13. Toric Hirzebruch-Riemann-Roch
in Q[x
1
, . . . , x
n
]. The substitution x
i
=
i
gives
n

i=1

i
1e

i
= P(c
1
, . . . , c
n
),
and the theorem is proved.
The HRR Theorem. Now that we have dened Chern characters and the Todd
class, we can begin to play with HRR. If D is a divisor on a smooth complete
variety X, then HRR gives the equality
(O
X
(D)) =

X
ch(O
X
(D))Td(X)
Here is what this looks like for a surface.
Example 13.1.7. Let X be a smooth complete surface, so that
Td(X) = 1+
1
2
c
1
+
1
12
(c
2
1
+c
2
).
Also recall that c
1
= [K
X
], where K
X
is the canonical divisor.
We rst apply HRR to D = 0. Since ch(O
X
) = 1, we obtain
(O
X
) =

X
Td(X) =

1+
1
2
c
1
+
1
12
(c
2
1
+c
2
)

=
1
12

X
c
2
1
+c
2
since

X
is trivial away from the top degree. It follows that the formula for Td(X)
can be written
(13.1.6) Td(X) = 1
1
2
[K
X
] +(O
X
)[pt].
Furthermore, since

X
c
2
1
=

X
[K
X
]
2
= (K
X
) (K
X
) = K
X
K
X
by (12.4.1)
and

X
c
2
= e(X) by (13.1.2), the above formula for (O
X
) simplies to
(O
X
) =
1
12

K
X
K
X
+e(X)

,
which is Noethers theorem (Theorem 10.5.3). This result expresses the topological
invariant e(X) in terms of the algebro-geometric invariants (O
X
) and K
X
K
X
.
In Exercise 13.1.2 you will show more generally that HRR applied to O
X
(D)
gives the formula
(13.1.7) (O
X
(D)) =
1
2
(D DD K
X
) +(O
X
).
This is the version of Riemann-Roch stated in Theorem 10.5.2.
It should be no surprise that the Riemann-Roch theorem for smooth complete
curves is also a special case of HRR. See Exercise 13.1.3.
13.1. Chern Characters, Todd Classes, and HRR 629
Example 13.1.8. The Hirzebruch surface X = H
2
was studied in Example 4.1.8.
In particular, the Picard group Pic(X) is generated by the classes of D
3
and D
4
.
We will use the Riemann-Roch theorem to compute (O
X
(D)) for the divisor D =
3D
3
5D
4
on X. Example 4.1.8 shows that D
1
D
3
and D
2
2D
3
+D
4
. Thus
K
X
=(D
1
+D
2
+D
3
+D
4
) =2D
4
.
The intersection pairing on X = H
2
was determined in Example 10.4.6, which
showed that
D
2
3
= 0, D
3
D
4
= 1, D
2
4
= 2.
Also note that (O
X
) = 1 by Demazure vanishing. Then applying Riemann-Roch
from (13.1.7) to the divisor D = 3D
3
5D
4
, we obtain
(O
X
(D)) =
1
2
D (DK
X
) +1
=
1
2
(3D
3
5D
4
) (3D
3
3D
4
) +1
=
1
2
(9D
2
3
24D
3
D
4
+15D
2
4
) +1 = 4.
In Example B.7.1 we show how to do this computation using Sage [262].
Grothendieck-Riemann-Roch. Hirzebruch proved his version of Riemann-Roch
for vector bundles on smooth projective varieties in 1954. This was generalized by
Grothendieck in 1957 to proper morphisms f : X Y between smooth quasipro-
jective varieties and coherent sheaves on X. His proof was published by Borel
and Serre in 1958. Botts wonderful review [41] of their paper points out that
Grothendiecks theorem explains why the Todd class from Denition 13.1.4 makes
sense. For this reason, we will briey discuss Grothendieck-Riemann-Roch (called
GRR) in the special case that uses smooth complete varieties and replaces the Chow
ring used by Grothendieck with the cohomology ring.
For a smooth complete variety X, we rst dene the group K(X) generated by
coherent sheaves on X, modulo the relations
[G] = [F] +[H ]
for all short exact sequences 0 F G H 0. One can also prove that a
coherent sheaf F on X has a Chern character ch(F) H

(X, Q) such that


ch(L) = e
c
1
(L)
when L is a line bundle.
ch(G) = ch(F) +ch(H ) when 0 F G H 0 is exact.
It follows that the Chern character induces a group homomorphism
ch : K(X) H

(X, Q).
Now suppose that f : X Y is a proper morphism between smooth complete
varieties. If F is coherent on X, then one can prove that the higher direct images
630 Chapter 13. Toric Hirzebruch-Riemann-Roch
R
i
f

F are coherent on Y. By the long exact sequence in cohomology, the map


f
!
[F] =

i0
(1)
i
[R
i
f

F] K(Y)
denes a homomorphism f
!
: K(X) K(Y). On the cohomology side, we also have
a map f
!
: H

(X, Q) H

(Y, Q), which for f : X Y as above is the Poincar e


dual of f

: H

(X, Q) H

(Y, Q). We call f


!
a generalized Gysin map. Here are
two important examples of these maps:
When f : X pt, f
!
is the map

X
: H

(X, Q) Q from 12.4.


When i : X Y is an inclusion map, then i
!
: H
k
(X, Q) H
k+2codimX
(Y, Q)
is the Gysin map, which satises i
!
i

= [X] for H
k
(Y, Q).
We will say more about the maps f
!
in the appendix at the end of the chapter.
These denitions show that a proper morphism f : X Y gives a diagram
K(X)
ch

f
!

(X, Q)
f
!

K(Y)
ch

(Y, Q)
which unfortunately does not commute. This is why Todd classes are needed: GRR
states that the modied diagram
(13.1.8)
K(X)
chTd(X)

f
!

(X, Q)
f
!

K(Y)
chTd(Y)

(Y, Q)
does commute. In Exercise 13.1.4 you will prove that (13.1.8) reduces to HRR
when Y =pt.
Following [41], we now specialize to the case where i : X Y is the inclusion
of a smooth hypersurface in a smooth complete variety, and we apply (13.1.8) to
[O
X
] K(X). First observe that i
!
[O
X
] = [i

O
X
] since i is an afne morphism.
Then the exact sequence 0 O
Y
(X) O
Y
i

O
X
0 and the properties of
the Chern character imply that
ch(i
!
[O
X
]) Td(Y) = ch([i

O
X
]) Td(Y) = ch

[O
Y
] [O
Y
(X)]

Td(Y)
= (1e
[X]
) Td(Y).
This is encouraging since expressions like 1e
[X]
appear in the denominator in
the formula for the Todd class given in Denition 13.1.4. We can do more since
S = [X]/(1e
[X]
) is invertible in H

(Y, Q), so that


(13.1.9) (1e
[X]
) Td(Y) = (S
1
Td(Y)) [X] = i
!
i

(S
1
Td(Y)),
13.1. Chern Characters, Todd Classes, and HRR 631
where the last equality follows since i
!
is a Gysin map. In order for (13.1.8) to
commute, (13.1.9) must equal i
!

ch([O
X
]) Td(X)

= i
!
(Td(X)). In other words,
we want Td(X) = i

(S
1
Td(Y)), which is equivalent to the equation
(13.1.10) i

Td(Y) = Td(X) i

[X]
1e
[X]

.
In Exercise 13.1.5, you will show that the formula of Denition 13.1.4 guarantees
that Todd classes satisfy this identity.
Exercises for 13.1.
13.1.1. Use the method of Example 13.1.5 to compute the Todd polynomial T
3
.
13.1.2. Derive the Riemann-Roch theorem for surfaces (13.1.7) from (13.0.1).
13.1.3. Derive the Riemann-Roch theorem for curves (10.5.1) from (13.0.1).
13.1.4. Prove that HRR is the special case of the commutative diagram (13.1.8) when
Y =pt. Thus GRR implies HRR.
13.1.5. Let Y be a smooth complete variety and let i : X Y be a smooth hypersurface.
Then we have the exact sequence of locally free sheaves
0 T
X
i

T
Y
N
X/Y
0
from Theorem 8.0.18, where N
X/Y
= (I
X
/I
2
X
)

.
(a) Prove that N
X/Y
i

O
Y
(X). Hint: The ideal sheaf of i : X Y is I
X
= O
Y
(X)
since X has codimension 1.
(b) Prove the identity c(i

T
Y
) = c(T
X
) c(i

O
Y
(X)) of total Chern classes.
(c) Showthat c(i

O
Y
(X)) =1+i

[X] and use part (b) to explain why i

[X] can be regarded


as a Chern root of i

T
Y
.
(d) Finally, show that the Todd classes Td(X) and Td(Y) satisfy (13.1.10).
13.1.6. Let E be a locally free sheaf of rank r on a smooth variety X. Similar to what
we did with the tangent bundle, the Chern roots of E are symbolic objects
i
such that
c(E ) =

r
i=1
(1 +
i
). Then the Chern character ch(E ) is dened as
ch(E ) =

k=0
1
k!
(
k
1
+ +
k
r
).
(a) Explain why ch(E ) is a polynomial in the Chern class of E .
(b) Let E =L
1
L
r
be a sum of line bundles. Use the denition of Chern character
to show that ch(E ) = ch(L
1
) + +ch(L
r
).
(c) If E has rank 2, show that
ch(E ) = 2 +c
1
(E ) +
1
2

c
1
(E )
2
2c
2
(E )

+
1
6

c
1
(E )
3
3c
1
(E )c
2
(E )

+ .
632 Chapter 13. Toric Hirzebruch-Riemann-Roch
13.2. Brions Equalities
When we prove the toric version of equivariant Riemann-Roch in 13.3, we will
create equivariant versions of Euler characteristics, Chern characters and Todd
classes, and then decompose them into sums of local terms. In this section, we
will see how this works for the Euler characteristic of a line bundle.
When D is a torus-invariant divisor on a complete toric variety X = X

of
dimension n, the sheaf L =O
X
(D) has Euler characteristic
(L) = (X, L) =
n

i=0
(1)
i
dim H
i
(X

, L).
Also recall from 9.1 that we used the decomposition
H
i
(X, L) =

mM
H
i
(X, L)
m
to compute sheaf cohomology of L. Combining these two formulas, we dene
(L) = (X, L) =
n

i=0
(1)
i
dim H
i
(X

, L)
m

m
.
This lives in the semigroup algebra Z[M]. The goal of this section is to write
(X, L) as a sum of local terms that live in a certain localization of Z[M].
The Formal Semigroup Module. The local terms will involve cohomology over
afne open subsets, which can lead to innite sums. Following [46], we introduce
the abelian group Z[[M]] consisting of all formal sums

mM
a
m

m
for a
m
Z.
This is not a ring, but there is a multiplication map
Z[M] Z[[M]] Z[[M]]
that makes Z[[M]] into a Z[M]-module. We say that Z[[M]] is the formal semigroup
module of M.
Now let X = X

be an arbitrary n-dimensional toric variety, which need not be


complete. Given a torus-invariant Weil divisor D on X, let L =O
X
(D) and dene
(13.2.1) (X, L) =

mM
n

i=0
(1)
i
dim H
i
(X, L)
m

m
Z[[M]].
In Chapter 9, we computed H

(X, L) using the



Cech complex

C

(U , L).
For the open cover U =U

max
, let
max
=
1
, . . . ,

and dene

C
p
(U , L) =

[]p
H
0
(U

, L),
where we refer to (9.1.1) for the precise meaning of []
p
and

. This gives a

Cech formula for (X, L) as follows.


13.2. Brions Equalities 633
Lemma 13.2.1. (X, L) =

p0
(1)
p

[]p
(U

, L)

in Z[[M]].
Proof. Just as in 9.1, the M-grading on H

(X

, L) and

C

(U , L) implies that
H

(X, L)
m
is the cohomology of the complex

C

(U , L)
m
. Since a complex and
its cohomology have the same Euler characteristic, it follows that
n

i=0
(1)
i
dim H
i
(X, L)
m
=
n

p=0
(1)
p

[]p
dim H
0
(U

, L)
m
.
However dim H
0
(U

, L)
m
=

n
i=0
(1)
i
dim H
i
(U

, L)
m
since U

is afne.
From here, the lemma follows easily.
Summable Series. For any cone , the vanishing of H
i
(U

, L) for i > 0
implies that
(13.2.2) (U

, L) =

mM
dim H
0
(U

, L)
m

m
Z[[M]].
We will soon see that these sums are reasonably behaved in the following sense.
Denition 13.2.2. An element f Z[[M]] is summable if there exists g Z[M]
and a nite set I M`0 such that in Z[[M]], we have
f

mI
(1
m
) = g.
Also let Z[[M]]
Sum
Z[[M]] be the subset of summable elements of Z[[M]].
By Exercise 13.2.1, Z[[M]]
Sum
is a Z[M]-submodule of Z[[M]]. Here are some
examples of summable elements.
Example 13.2.3. Let M =Z and let =
1
. Then

=0

is summable since
(1)(1++
2
+ ) = 1.
A more surprising example is

, which is summable since


(1)( +
2
+
1
+1++
2
+ ) = 0.
In Example 13.2.7 we will use this to show that

sums to zero.
Here is a basic summability result.
Lemma 13.2.4. Let D be a torus-invariant Cartier divisor on X = X

and set
L =O
X
(D). Then (U

, L) is summable for every . Furthermore:


(a) If dim < n, then
(U

, L) (1
m
0
) = 0
for some m
0
M`0.
634 Chapter 13. Toric Hirzebruch-Riemann-Roch
(b) If dim = n and is smooth, then
(U

, L)
n

i=1
(1
m
,i
) =
m
,
where m

M satises D[
U
= div(
m
)[
U
and m
,1
, . . . , m
,n
M are the
minimal generators of

.
Proof. If dim < n, then

is not strongly convex, so that we can nd a nonzero


m
0

) M. Then
m
0
C[

M] is invertible, which means that


multiplication by
m
0
gives an isomorphism H
0
(U

, L) H
0
(U

, L). Hence
H
0
(U

, L)
m
H
0
(U

, L)
m+m
0
for all m M. This implies
m
0
(U

, L) = (U

, L), and part (a) follows.


For part (b), assume is smooth and let u
1
, . . . , u
n
be the minimal generators
of
1
, . . . ,
n
(1). They form a basis of N since is smooth. Then the dual
basis m
1
, . . . , m
n
M consists of the minimal generators of

. If we write
D[
U
=

n
i=1
a
i
D

i
, then m

n
i=1
a
i
m
i
satises D[
U
= div(
m
)[
U
.
The polyhedron of L[
U
is P
D,
=m M [ 'm, u
i
` a
i
, 1 i n. Using
the above formula for m

, we obtain P
D,
M = m

M. We also have
H
0
(U

, L) =

mP
D,
M
C
m
by (4.3.3). Then (13.2.2) implies that
(13.2.3) (U

, L) =

mm+

m
=
m

m
.
Since

M = Nm
1
+ +Nm
n
, the sum on the right lives in the formal series
ring Q[[
m
1
, . . . ,
mn
]]. In this ring, one computes that

m
=
n

i=1

=0

m
i

=
n

i=1
(1
m
i
)
1
.
Part (b) of the proposition now follows easily.
Finally, suppose that (n) is not smooth. Since D is Cartier, we have
D[
U
= div(
m
)[
U
for some m

M. Then (13.2.3) still holds, so it sufces to


show that

m
is summable. The methods of 11.1 imply that

M
R
has a smooth renement. Let the maximal cones of this renement be C
1
, . . . ,C
s
.
By inclusion-exclusion, we obtain
(13.2.4)

m
=

mC
i
M

i<j

mC
i
C
j
M

m
+ .
Since an intersection C =C
i
1
C
i

is smooth, the minimal generators of CM


are a subset of a basis of M. Then the proof of part (b) implies that

mCM

m
is
summable. We conclude that (U

, L) is summable.
13.2. Brions Equalities 635
The Simplicial Case. When (n) is smooth, part (b) of Lemma 13.2.4 shows
that (U

, L) satises a nice identity. With a little work, this can be generalized


to the simplicial case.
Let (n) be simplicial and let m
,1
, . . . , m
,n
be the minimal generators of

M
R
. Recall from Proposition 11.1.8 that
(13.2.5) mult(

) =[P

[, P

n
i=1

i
m
,i
[ 0
i
< 1,
where mult(

) =[M: Zm
,1
+ +Zm
,n
]. You will explore the relation between
mult() and mult(

) in Exercise 13.2.3.
Lemma 13.2.5. If D is a torus-invariant Cartier divisor on X = X

and (n)
is simplicial, then for L =O
X
(D), we have
(U

, L)
n

i=1
(1
m
,i
) =
m

mP

m
,
where D[
U
= div(
m
)[
U
, and m
,1
, . . . , m
,n
and P

are as above.
Proof. For simplicity, write m
i
= m
,i
. Since (U

, L) =
m

m
by
(13.2.3), it sufces to prove that
(13.2.6)


m

i=1
(1
m
i
) =

mP

m
in Z[[M]]. Let A =Nm
1
+ +Nm
n
M. We claim that
(13.2.7) m

M m = m

+m

for m

A and m

M.
One direction is obvious. For the other direction, take m

M. Since the m
i
generate

, we can write m =

n
i=1
c
i
m
i
for c
i
0. Then writing c
i
= a
i
+b
i
for
a
i
N and 0 b
i
< 1 easily implies that m has the desired form, proving (13.2.7).
By Exercise 13.2.4, the decomposition m =m

+m

in (13.2.7) is unique when


it exists. It follows that
(13.2.8)

m
=

mA


mP

.
Arguing as in the proof of Theorem 13.2.8, we also have

mA

i=1
(1
m
i
) = 1.
Hence multiplying each side of (13.2.8) by

n
i=1
(1
m
i
) gives (13.2.6).
Example 13.2.6. Consider the cone = Cone(e
1
, e
1
2e
2
) N
R
, with dual

= Cone(e
2
, 2e
1
e
2
) M
R
. One computes easily that P

= 0, e
1
e
2
.
Setting t
i
=
e
i
, Lemma 13.2.5 implies that
(U

, O
U
) (1t
1
2
)(1t
2
1
t
1
2
) = 1+t
1
t
1
2
.
636 Chapter 13. Toric Hirzebruch-Riemann-Roch
In [50, 1.3], Brion and Vergne prove that (U

, L) is summable under the


weaker hypothesis that D is a Weil divisor, and in [50, 2.1], they give a version of
Lemma 13.2.5 that applies to Weil divisors and simplicial cones.
Brions Equalities. Consider the multiplicative set S Z[M] consisting of all nite
products of the form

s
i=1
(1
m
i
), where m
1
, . . . , m
s
M`0. Note that 1 S
because of the empty product. This gives the localization Z[M]
S
.
Given a summable element f Q[[M]]
Sum
, there are h S and g Z[M] such
that h f = g. Then set
o( f ) =
g
h
Z[M]
S
.
In Exercise 13.2.1 you will prove that the map
(13.2.9) o : Z[[M]]
Sum
Z[M]
S
is a well-dened Z[M]-module homomorphism. We call o the sum function.
Example 13.2.7. Let M =Z and be as in Example 13.2.3. Then
o

=0

=
1
1
since (1)(1++
2
+ ) = 1
o

= 0 since (1)( +
1
+1++ ) = 0.
This example shows that o is not injective. In fact, it has a rather large kernel,
which is key to proving the following remarkable equalities due to Brion.
Theorem 13.2.8. Let X = X

be a complete toric variety of dimension n. If L =


O
X
(D) for a torus-invariant Cartier divisor D with Cartier data m

(n)
, then:
(a) (X, L) =

(n)
o( (U

, L)) in Q[M]
S
.
(b) If (n) is simplicial and m
,1
, . . . , m
,n
M are the minimal generators of

M
R
, then
o( (U

, L)) =

mP

n
i=1
(1
m
,i
)
for P

as in (13.2.5). In particular, if is smooth, then


o( (U

, L)) =

m

n
i=1
(1
m
,i
)
.
Proof. First, o( (X, L)) = (X, L) since X is complete. Then Lemma 13.2.1
implies that
(X

, L) =

p0
(1)
p


[]p
o( (U

, L))

.
But we also know that o( (U

, L)) = 0 when dim

< n by Lemma 13.2.4.


Since X is complete, the open cover used in the

Cech complex is U =U

(n)
,
13.2. Brions Equalities 637
and the only place n-dimensional cones appear is in the rst term

C
0
(U , L) =

(n)
H
0
(U

, L). Hence all terms in the above sum drop out except for those
for p =0, which correspond to (n). This proves part (a). Part (b) now follows
immediately from Lemmas 13.2.4 and 13.2.5.
Example 13.2.9. For the cone = Cone(e
1
, e
1
2e
2
) N
R
of Example 13.2.6,
part (b) of Theorem 13.2.8 implies that
o( (U

, O
U
)) =
1+t
1
t
1
2
(1t
1
2
)(1t
2
1
t
1
2
)
.
This also follows from Example 13.2.6 by the denition of o.
Here is a wonderful application of Theorem 13.2.8.
Corollary 13.2.10 (Brions Equalities). Let P M
R
R
n
be a full dimensional
lattice polytope, and for each vertex v P, let C
v
= Cone(PMv) M
R
. Then:
(a)

mPM

m
=

v vertex

v
o


mCvM

.
(b) If P is simple and v P is a vertex, let m
v,1
, . . . , m
v,n
M be the minimal
generators of C
v
. Then

mPM

m
=

v vertex

mPvM

n
i=1
(1
m
v,i
)
for P
v
= P
Cv
as in (13.2.5). In particular, if P is smooth, then

mPM

m
=

v vertex

n
i=1
(1
m
v,i
)
.
Proof. We will apply Proposition 13.2.8 to the toric variety X
P
and the line bundle
L =O
X
P
(D
P
). By denition, the maximal cones of the normal fan
P
are
v
=C

v
for v P a vertex. Also recall from (4.2.8) that the vs are the Cartier data of D
P
.
The higher cohomology of L vanishes since D
P
is ample, so that
(X
P
, L) =

mPM

m
.
Furthermore,

v
=C
v
and (13.2.3) imply that
(U
v
, L) =
v

v
M

m
=
v

mCvM

m
.
Then part (a) follows from Theorem 13.2.8, and the same is true for part (b).
Brions original proof [46] of Corollary 13.2.10 used equivariant Riemann-
Roch. In [155], Ishida gave an elementary proof of Brions equalities and used it
prove a special case of HRR. Brion and Vergne prove Theorem 13.2.8 in [50, 1.3]
638 Chapter 13. Toric Hirzebruch-Riemann-Roch
as part of their toric proof of equivariant Riemann-Roch. A completely elementary
approach to Corollary 13.2.10 can be found in [22, Thm. 9.7].
Let us give a simple example of Brions equalities.
Example 13.2.11. Consider the interval [0, d] R with vertices v
1
=0 and v
2
=d.
Note that C
v
1
= [0, ) and C
v
2
= (, 0]. Since
o


[0,)Z

=o

=0

=
1
1
o


(,0]Z

=o

=0

=
1
1
1
,
the right-hand side of Brions equality is
1
1
+
d
1
1
1
=
1
1
+
d
(1)
(1)(1
1
)
=
(1
d+1
)(1
1
)
(1)(1
1
)
= 1++
2
+ +
d
.
The nal result is

[0,d]Z

, as predicted by Brions equality.


Here is a more substantial example that uses Theorem 13.2.8.
Example 13.2.12. For the Hirzebruch surface X =H
2
and divisor D=3D
3
5D
4
,
we found that (L) = 4 for L = O
X
(D) in Example 13.1.8. Here, M = Z
2
, and
u
2
u
4
u
3
u
1
= (1,2)

4
Figure 1. The fan for X =H2
we set t
1
=
e
1
and t
2
=
e
2
. In Exercise 13.2.2 you will check that with labellings
13.2. Brions Equalities 639
on (2) shown in Figure 1, we have
o( (U

1
, L)) =
t
3
1
(1t
1
)(1t
2
)
o( (U

2
, L)) =
t
3
1
t
5
2
(1t
1
)(1t
1
2
)
o( (U

3
, L)) =
t
10
1
t
5
2
(1t
1
1
)(1t
2
1
t
1
2
)
o( (U

4
, L)) =
1
(1t
1
1
)(1t
2
1
t
2
)
.
Adding these up and doing a lot of algebra yields
(13.2.10) t
4
1
t
3
2
+t
5
1
t
3
2
+t
4
1
t
4
2
+t
5
1
t
4
2
+t
6
1
t
4
2
+t
7
1
t
4
2
t
3
1
t
1
2
t
2
1
t
1
2
In Figure 2, the six exponent vectors m = (a, b) where t
a
1
t
b
2
appears with a positive
sign in (13.2.10) are plotted as solid dots, and the two exponent vectors where t
a
1
t
b
2
appears with a negative sign are plotted as hollow dots.
1
2
3
4
1 2 3 4 5 6 7
Figure 2. Exponent vectors of the monomials in (13.2.10)
In Exercise 13.2.2 you will show that P
D
=, so that H
0
(X, L) = 0. Hence
(X, L) =

mZ
2
dim H
1
(X, L)
m
t
m
+

mZ
2
dim H
2
(X, L)
m
t
m
.
where t
m
= t
a
1
t
b
2
for m = (a, b) Z
2
. This equals (13.2.10) by Theorem 13.2.8.
The sign patterns from Proposition 9.1.6 show that the same m cant appear in both
H
1
(X, L) and H
2
(X, L). Thus solid dots correspond to m with H
2
(X, L)
m
= 0
and hollow dots correspond to m with H
1
(X, L)
m
= 0. We will conrm these
cohomology computations using Macaulay2 [123] in Example B.5.1 and using
Sage [262] in Example B.7.1.

640 Chapter 13. Toric Hirzebruch-Riemann-Roch


When X is complete,
m
1 takes (X, L) to (X, L). We will call this
taking the nonequivariant limit in 13.3. For instance, Example 13.2.12 gives
(L) = t
4
1
t
3
2
+t
5
1
t
3
2
+t
4
1
t
4
2
+t
5
1
t
4
2
+t
6
1
t
4
2
+t
7
1
t
4
2
t
3
1
t
1
2
t
2
1
t
1
2
,
so that mapping t
i
=
e
i
to 1 gives
(L) = 1+1+1+1+1+111 = 4.
This agrees with the computation done in Example 13.1.8. The difference is that
previously, we used Riemann-Roch, while here, all we needed was the explicit
local decompostion of (L). In the next section, this local decompostion will be
an important part of our proof of equivariant Riemann-Roch for toric varieties.
Exercises for 13.2.
13.2.1. Consider Q[[M]]
Sum
Q[[M]] from Denition 13.2.2.
(a) Prove that Q[[M]]
Sum
is a Q[M]-submodule of Q[[M]].
(b) Prove that (13.2.9) is well-dened and is a homomorphism of Q[M]-modules.
13.2.2. Consider X =H
2
and L =O
X
(D), D = 3D
3
5D
4
, as in Example 13.2.12.
(a) Compute the polyhedron P
D
of D and conclude that H
0
(X, L) = 0.
(b) This example computed the ms that occur in H
1
(X, L) and H
2
(X, L). Compute the
chamber decomposition for D on X (see Example 9.1.8) and then use Proposition 9.1.6
to show that we get the same ms as in Figure 2.
13.2.3. Let N
R
R
n
be a simplicial cone of dimension n with minimal generators
u
1
, . . . , u
n
N, and let m

1
, . . . , m

n
M
R
be the dual basis in the sense of linear algebra.
Also let
i
be the smallest positive integer such that m
i
=
i
m

i
M.
(a) Prove that m
1
, . . . , m
n
are the minimal generators of

.
(b) Prove that mult()mult(

) =

n
i=1

i
. Hint: Show that MZm

1
+ +Zm

n
is dual
to Zu
1
+ +Zu
n
N.
13.2.4. Prove that the decomposition m = m

+m

in (13.2.7) is unique when it exists.


13.2.5. Let L =O
P
1 (d) for d > 0.
(a) Compute (P
1
, L) using Batyrev-Borisov vanishing (Theorem 9.2.7).
(b) Compute (P
1
, L) using Theorem 13.2.8.
(c) Compute (P
1
, L) using the Riemann-Roch theorem for curves (10.5.1) and explain
how your answer relates to (P
1
, L).
13.2.6. For X = X

and L as in Lemma 13.2.2, we have o( (U

, L)) = 0 when
satises dim < n = dim X. Here you will prove that o( (U

, L)) = 0 when dim = n.


To begin, assume that there are m
1
, . . . , m
s
M` 0 with
(13.2.11) (U

, O
U
)
s

i=1
(1
mi
) = 0.
(a) Show there is u Int( N) such that 'm
i
, u` = 0 for all i.
(b) Show that we can assume 'm
i
, u` > 0 for all i. Hint: If necessary, multiply (13.2.11)
by
mi
.
13.3. Toric Equivariant Riemann-Roch 641
(c) Show that the constant term 1 on the left-hand side of (13.2.11) cannot cancel. Then
conclude that o( (U

, L)) = 0.
13.3. Toric Equivariant Riemann-Roch
In 12.4 we described the cohomology ring of a complete simplicial toric variety
by rst computing its equivariant cohomology and then taking the nonequivariant
limit. We use the same strategy here: we will deduce HRR for a smooth complete
toric variety from the following equivariant Riemann-Roch theorem.
Theorem 13.3.1. For a smooth complete toric variety X = X

and a line bundle


L =O
X
(D) of a torus-invariant divisor D on X, we have

T
(L) =

X
eq
ch
T
(L)Td
T
(X).
In this theorem,
T
(L),

X
eq , ch
T
(L) and Td
T
(X) are equivariant versions of
the corresponding objects appearing in (13.0.1). The equality of Theorem 13.3.1
takes place in the completion of the equivariant cohomology ring of a point, which
we denote by

. All of this will be dened carefully in the course of the section.
We will follow [50], where Brion and Vergne prove equivariant Riemann-Roch
for complete simplicial varieties. We will discuss their simplicial version of the
theorem after completing the proof of the smooth case.
The Equivariant Euler Characteristic. As in 12.4, we set
(
T
)
Q
= H

T
(pt, Q),
where T = T
N
is the torus associated to M. Since exponentials are innite sums,
we will use the completion of H

T
(pt, Q) =

k=0
H
k
T
(pt, Q), written

=

H

T
(pt, Q) =

k=0
H
k
T
(pt, Q).
Recall from (12.4.9) that we have an isomorphism s : Sym
Q
(M) (
T
)
Q
,
where Sym
Q
(M) is our notation for

k=0
Sym
k
Q
(M
Q
). Hence, if m M, then
s(m) is a degree 2 element of

. It follows that m M gives the exponential
e
s(m)
= 1+s(m) +
1
2
s(m)
2
+
1
6
s(m)
3
+

.
We can now dene the equivariant Euler characteristic that appears on the left-
hand side of the equivariant Riemann-Roch theorem.
Denition 13.3.2. If X = X

is complete and L = O
X
(D) is the line bundle of a
torus-invariant divisor D on X, then the equivariant Euler characteristic of L is
(13.3.1)
T
(L) =

mM
n

i=0
(1)
i
dim H
i
(X, L)
m
e
s(m)

,
642 Chapter 13. Toric Hirzebruch-Riemann-Roch
where H
i
(X, L) =

mM
H
i
(X, L)
m
is the decomposition from 9.1.
We will see below that
T
(L) is closely related to (L) from 13.2.
The Nonequivariant Limit. Let i

pt
:

Q be the map that sends elements of
positive degree to zero. Then
i

pt
(
T
(F)) = (F) =
n

i=0
(1)
i
dim H
i
(X, F),
i.e., i

pt
takes the equivariant Euler characteristic to the ordinary Euler characteristic.
Later in the section we will show that applying i

pt
to the equivariant Riemann-
Roch theorem gives HRR from (13.0.1). As mentioned in 12.4, the maps i

pt
and
i

X
: H

T
(X, Q) H

(X, Q) are the nonequivariant limit that turns equivariant


cohomology into ordinary cohomology.
Equivariant Chern Characters and Todd Classes. To dene equivariant Chern
characters, we replace the equivariant cohomology H

T
(X, Q) with its completion

T
(X, Q) =

k=0
H
k
T
(X, Q).
The nonequivariant limit map i

X
: H

T
(X, Q) H

(X, Q) extends in the obvious


way to a ring homomorphism
i

X
:

H

T
(X, Q) H

(X, Q)
since elements of degree > 2dim X map to zero.
By Proposition (12.4.13), a torus-invariant divisor D =

gives the
equivariant cohomology class [D]
T
=

[D

]
T
H
2
T
(X, Q). Then we dene
the equivariant Chern character ch
T
(L) of L =O
X
(D) to be
ch
T
(L) = e
[D]
T
= 1+[D]
T
+
1
2!
[D]
2
T
+

H

T
(X, Q).
Furthermore, since X is smooth, Theorem 13.1.6 implies that the ordinary Todd
class of X is
Td(X) =
r

i=1
[D

]
1e
[D]
.
Hence it is natural to dene the equivariant Todd class of X to be
Td
T
(X) =

[D

]
T
1e
[D]
T
.
The power series (13.1.4) shows that Td
T
(X)

H

T
(X, Q). Since i

X
[D

]
T
= [D

]
by Proposition 12.4.13, we have
(13.3.2) i

X
ch
T
(L) = ch(L) and i

X
Td
T
(L) = Td(L).
13.3. Toric Equivariant Riemann-Roch 643
The Equivariant Integral. We saw in 13.1 that the map

X
: H

(X, Q) Q is
the generalized Gysin map of the constant function p : X pt. In the appendix
to this chapter, we will see that p also gives the equivariant Gysin map
p
!
: H

T
(X, Q) H

T
(pt, Q)
which maps H
k
T
(X, Q) to H
k2dim X
T
(pt, Q). We write

X
eq instead of p
!
, so that

X
eq
:

H

T
(X, Q)

H

T
(pt, Q) =

.
This is called the equivariant integral. We will prove later that

X
eq and

X
are
compatible with taking the nonequivariant limit. This will make it easy to derive
HRR from equivariant Riemann-Roch.
Strategy of the Proof . We have now dened everything needed to make sense of
the equivariant Riemann-Roch theorem for X = X

in Theorem 13.3.1:

T
(L) lives in

.
ch
T
(L) and Td
T
(X) live in the ring

H

T
(X, Q).

X
eq maps H

T
(X, Q) to

.
Thus equivariant RR is the equation
T
(L) =

X
eq ch
T
(L)Td
T
(X) in

.
The next step is to say a few words about how we will prove the theorem. The
main idea is to use the map : H

T
(X, Q) H

T
(X
T
, Q) from (12.4.11) induced by
the inclusion of the xed point set X
T
X. Recall that X
T
= x

[ (n),
where n = dim X. The completed version of is
:

H

T
(X, Q)

(n)

T
(x

, Q) =

(n)

.
This map is injective by Corollary 12.4.9, and the localization theorem (see [9])
implies that becomes an isomorphism after tensoring with the eld of fractions of
(
T
)
Q

. In our case, we can do better. Let S be the multiplicative set consisting


of all nite products of nonzero degree 2 elements of

. The empty product shows


that 1 S. Then Exercise 12.4.3 implies that the localized map
(13.3.3)
S
:

H

T
(X, Q)
S

(n)

S
.
is an isomorphism. Here,

H

T
(X, Q)
S
and

S
are the localizations in the sense of
commutative algebra. Thus we consider fractions whose denominators lie in S.
The strategy of the proof is to express each side of the desired equation as
a sum of local terms in

S
indexed by (n). The proof will then reduce to
checking that the local terms on each side of the equation are equal in

S
. Many
proofs in equivariant cohomology use this approach.
644 Chapter 13. Toric Hirzebruch-Riemann-Roch
Local Euler Characteristics. Following the strategy outlined above, we rst show
that the equivariant Euler characteristic
T
(L)

decomposes as a sum of local


terms

T
(L) =

(n)

(L),
T

(L)

S
.
When X = X

is complete, Brions equalities (Theorem 13.2.8) imply that


(13.3.4) (X, L) =

mM
n

i=0
(1)
i
dim H
i
(X, L)
m

m
Z[M]
from (13.2.1) as a sum of local terms
(13.3.5) (X, L) =

(n)
o( (U

, L)),
where each o( (U

, L)) lies in the localization Z[M]


S
and the sum function o is
dened in (13.2.9).
We translate (13.3.5) into a decomposition in

S
using the ring homomorphism
: Z[M]

,
m
e
s(m)
.
If m M is nonzero, then
1e
s(m)
s(m)
=
s(m)
1
2
s(m)
2

s(m)
=1
1
2
s(m)
is invertible in

. Thus
(1
m
) = 1e
s(m)
= s(m)
1e
s(m)
s(m)
= s(m) invertible element of

.
Since Q[M]
S
inverts all 1
m
and

S
inverts all s(m), it follows that extends to
a ring homomorphism
: Q[M]
S

S
.
The denitions of
T
(L) from (13.3.1) and (X

, L) from (13.3.4) make it


easy to see that
( (X, L)) =
T
(L)
when X = X

is complete. Using , we now dene the local version of


T
(L).
Denition 13.3.3. Let X = X

be complete of dimension n and set L = O


X
(D)
for a torus-invariant Cartier divisor D. Given (n), dene

(L) = (o( (U

, L))).
Here is the desired decomposition.
Theorem 13.3.4. Let X = X

be complete of dimension n and set L =O


X
(D) for
a torus-invariant Cartier divisor D with Cartier data m

(n)
. Then:
13.3. Toric Equivariant Riemann-Roch 645
(a)
T
(L) =

(n)

(L) in

S
.
(b) If (n) is simplicial and m
,1
, . . . , m
,n
M are the minimal generators of

M
R
, then

(L) =
e
s(m)

mP

M
e
s(m)

n
i=1
(1e
s(m
,i
)
)
for P

as in (13.2.5). In particular, if is smooth, then

(L) =
e
s(m)

n
i=1
(1e
s(m
,i
)
)
.
Proof. Apply to Theorem 13.2.8.
Fixed Points. After decomposing the equivariant Euler characteristic into local
factors, the next step is to decompose the equivariant integral. This requires that
we study the xed points of the torus action on X = X

. Here we will assume that


X is complete and simplicial.
Given (n), note that V() =x

is a xed point of the torus action. Let


i

: V() =x

X
be the inclusion map. We also have the constant map p : X pt. Since pi

is
the canonical map that takes x

to pt, we get a commutative diagram


(13.3.6)

T
(X, Q)
i

T
(pt, Q) =

(x

, Q),
where the isomorphism on the bottom is (pi

=i

. Using this isomorphism


to identify

H

(x

, Q) with

H

T
(pt, Q) =

, we can write
i

()

when

H

T
(X, Q).
Lemma 13.3.5. Let X = X

be complete and simplicial of dimension n. Also let


(n) and (1). Then:
(a) If / (1), then i

([D

]
T
) = 0.
(b) If (1), then i

([D

]
T
) =
1

s(m), where be the smallest positive integer


such that D

is Cartier and m M satises D

= div(
m
).
Proof. Factor i

: x

X as x

j
X. Proposition 12.4.13 implies that
for any (1),
j

([D

]
T
) = [D

]
T
.
646 Chapter 13. Toric Hirzebruch-Riemann-Roch
If / (1), this is zero since D

=. On the other hand, if (1), then on


U

, we have D

= div(
m
). Using Proposition 12.4.13 again, we obtain
j

([D

]
T
) = [D

]
T
=[div(
m
)]
T
= s(m) 1

H

T
(U

, Q).
Since s(m)

and everything is a module over

, mapping the above equation into

T
(x

, Q) implies i

([D

]
T
) =s(m) 1. The desired formula follows.
We also have the equivariant Gysin map i
!
:

H

T
(pt, Q)

H

T
(X, Q) from
the appendix to this chapter. Here is an important property of this map.
Proposition 13.3.6. Let X = X

be complete and simplicial. If (n), then


i
!
(1) = mult()

(1)
[D

]
T
.
Proof. Since we have not developed the full theory of equivariant cohomology
classes, our argument will be somewhat ad-hoc. Note that

(1)
[D

]
T
lies in
H
2n
T
(X, Q), and the same is true for i
!
(1) by Proposition 13.A.9.
The rst step is to show that
(13.3.7)

X
eq
i
!
(1) = 1 and

X
eq
mult()

(1)
[D

]
T
= 1.
Since

X
eq = p
!
, the integral on the left is p
!
(i
!
(1)), which by Proposition 13.A.9
is equal to (p i

)
!
(1) = 1 since p i

: x

pt. For the other integral in


(13.3.7), observe that

X
eq

(1)
[D

]
T


=

H

T
(pt, Q) has degree zero by
Proposition 13.A.9. Since

0
=Q, it sufces to consider
i

pt

X
eq

(1)
[D

]
T
=

X
i


(1)
[D

]
T

(1)
i

X
[D

]
T
=

(1)
[D

],
where the rst equality uses the commutative diagram from Proposition 13.A.11
and third uses Proposition 12.4.13. Then we are done by Lemma 12.5.2, which
implies that

(1)
[D

] = mult()
1
[V()] = mult()
1
[x

].
The second step is to show that
(13.3.8) i

(i
!
(1)) = 0 and i


(1)
[D

]
T

= 0 when

(n),

= .
In this case, the second equality is easy, since some (1) is not contained in

(1), so that i

[D

]
T
= 0 by Lemma 13.3.5. For the rst, x

= and
Proposition 13.A.10 give a commutative diagram
0


H

T
(x

, Q)

T
(x

, Q)

i!

(X, Q).
i

13.3. Toric Equivariant Riemann-Roch 647


It follows that i

i
!
= 0, and (13.3.8) is proved.
In Exercise 13.3.1 you will give the easy argument that (13.3.7) and (13.3.8)
imply that i
!
(1) and mult()

(1)
[D

]
T
are equal in H
2n
T
(X, Q).
Decomposing the Equivariant Integral. When X is smooth and complete, we get
the following important formula for the equivariant integral.
Theorem 13.3.7. If X =X

is an n-dimensional smooth complete toric variety and




H

T
(X, Q), then

X
eq
= (1)
n

(n)
i

()

n
i=1
s(m
,i
)
in

S
, where m
,1
, . . . , m
,n
are the minimal generators of

M
R
for (n).
Proof. We will work in the localization

H

T
(X, Q)
S
. For (n), let

i
!
(1)

H

T
(X, Q)
S
, where

=
(1)
n

n
i=1
s(m
,i
)

S
.
The two key properties of the

are

X
eq

() (13.3.9)

(n)

= 1

H

T
(X, Q)
S
. (13.3.10)
Note that the theorem follows immediately from these properties.
For (13.3.9), recall that

X
eq equals p
!
for the constant map p : X pt. Then

T
(X, Q) becomes a

-module via p

and p
!
is a

-module homomorphism since


p
!
(p

() ) = p
!
() by Proposition 13.A.9. It follows that

X
eq extends to a

S
-module homomorphism between the localizations at S.
By Proposition 13.A.9, the equivariant Gysin map i
!
satises i
!
(1) =
i
!
(i

()). Hence

i
!
(1) =

i
!
(i

()).
Then (13.3.9) follows from the equalities

X
eq

= p
!
(

i
!
(i

()) =

p
!
(i
!
(i

()) =

(),
where the last equality follows because p
!
i
!
= (pi

)
!
is the inverse of the map
(pi

used in (13.3.6) to identify



H

T
(x

, Q) with

H

T
(pt, Q) =

.
For (13.3.10), we will use the map :

H

T
(X, Q)
S

(n)

S
from (13.3.3).
The strategy is to show that for (n),
(13.3.11) (i
!
(1)) = (1)
n

n
i=1
s(m
,i
) e

.
648 Chapter 13. Toric Hirzebruch-Riemann-Roch
Once this is proved, then (

) = e

follows by the denition of

. Thus

(n)

(n)
e

= (1),
which implies (13.3.10) since is injective.
To prove (13.3.11), rst note that i

(i
!
(1)) = 0 for

= by (13.3.8). To
calculate i

(i
!
(1)), let u
i
be the minimal generator of
i
(1) for i i n.
Since is smooth, the dual basis m
i
= m
,i
gives the minimal generators of

.
Furthermore, D
i
= D

i
is Cartier since X is smooth. Then
i

(i
!
(1)) =
n

i=1
i

([D
i
]
T
) =
n

i=1
(s(m
i
)) = (1)
n
n

i=1
s(m
i
),
where the rst equality uses Proposition 13.3.6 and the second uses Lemma 13.3.5
and D
i
U

= div(
m
i
). This proves (13.3.11) and completes the proof.
Theorem 13.3.7 is a special case of a general formula described in [9]. From
this more sophisticated view, the denominator

n
i=1
s(m
,i
) is the equivariant Euler
class of the normal bundle of x

X. See Exercise 13.3.2 for further details.


We should also mention that there is a simplicial version of Theorem 13.3.7,
which states that
(13.3.12)

X
eq
= (1)
n

(n)
mult(

) i

()

n
i=1
s(m
,i
)
when X is complete and simplicial. You will prove this in Exercise 13.3.3.
Proof of Equivariant RR. We nally have the tools needed to prove our version of
the equivariant Riemann-Roch theorem for smooth complete toric varieties, stated
earlier as Theorem 13.3.1.
Proof of Theorem 13.3.1. We need to show that

T
(L) =

X
eq
, = ch
T
(L)Td
T
(X),
when L = O
X
(D) and D is a torus-invariant divisor on X = X

. Let the Cartier


data of D be m

(n)
. By Theorem 13.3.4, we have

T
(L) =

(n)
e
s(m)

n
i=1
(1e
s(m
,i
)
)
.
Comparing this to the decomposition of

X
eq given in Theorem 13.3.7, we see
that it sufces to prove that for (n), we have
(13.3.13)
e
s(m)

n
i=1
(1e
s(m
i
)
)
= (1)
n
i

()

n
i=1
s(m
i
)
, = ch
T
(L)Td
T
(X).
where for simplicity we write m
i
= m
,i
. The m
i
are dual to the minimal generators
n
i
of . Let D
i
= D

i
be the divisor corresponding to
i
= Cone(u
i
) (1).
13.3. Toric Equivariant Riemann-Roch 649
Since i

() = i

(ch
T
(L))i

(Td
T
(X)), we need to compute i

(ch
T
(L)) and
i

(Td
T
(X)). We begin with the former. The Chern character is ch
T
(L) = e
[D]
T
.
Adapting the proof of Lemma 13.3.5, one easily sees that
i

([D]
T
) = s(m

)
since the Cartier data of D satises D[
U
= div(
m
)[
U
. Hence
i

(ch
T
(L)) = i

(e
[D]
T
) = e
i

([D]
T
)
= e
s(m)
.
We next compute i

(Td
T
(X)). If / (1), then
i

[D

]
T
1e
[D]
T

= i

1+
1
2
[D

]
T
+

= 1
by Lemma 13.3.5. On the other hand, if =
i
for 1 i n, then D
i
U

=
div(
m
i
) on U

, so the same lemma implies that


i

[D

]
T
1e
[D]
T

=
i

[D

]
T
1e
i

[D]
T
=
s(m
i
)
1e
s(m
i
)
.
Since Td
T
(X) =

(1)
[D

]
T
/(1e
[D]
T
), it follows that
i

(Td
T
(X)) =
n

i=1
s(m
i
)
1e
s(m
i
)
= (1)
n
n

i=1
s(m
i
)
1e
s(m
i
)
.
It is now easy to prove (13.3.13), since
e
s(m)

n
i=1
(1e
s(m
i
)
)
=
(1)
n

n
i=1
s(m
i
)
e
s(m)
(1)
n
n

i=1
s(m
i
)
1e
s(m
i
)
=
(1)
n

n
i=1
s(m
i
)
i

(ch
T
(L)) i

(Td
T
(X)),
where we have used the above computations of i

(ch
T
(L)) and i

(Td
T
(X)). This
gives (13.3.13) since i

is a ring homomorphism.
Passing from Equivariant RR to HRR. After a lot of hard work, our reward is an
easy proof of the main result of this chapter.
Theorem 13.3.8. For an invertible sheaf L on a smooth complete toric variety
X = X

, we have
(L) =

X
ch(L)Td(X).
Proof. We may assume that L =O
X
(D) for a torus-invariant divisor D on X. This
ensures that
T
(L) and ch
T
(L) are dened. Then
(L) = i

pt
(
T
(L)) = i

pt

X
eq
ch
T
(L)Td
T
(X) =

X
i

ch
T
(L)Td
T
(X)

,
650 Chapter 13. Toric Hirzebruch-Riemann-Roch
where the rst equality is (13.3.1), the second is equivariant Riemann-Roch, and
the third is the commutative diagram from Proposition 13.A.11. Since i

X
is a ring
homomorphism, (13.3.2) implies that the integral on the right is

X
ch(L)Td(X).
We have proved the Hirzebruch-Riemann-Roch theorem!
To get a better sense of how HRR relates to equivariant RR, let us work out a
concrete example.
Example 13.3.9. Let L = O
P
1(dD
0
), where the minimal generators of the fan of
P
1
are u
0
= e
1
and u
1
= e
1
in N = Ze
1
. The completed equivariant cohomology
ring of X =P
1
is

T
(P
1
, Q) =Q[[x
0
, x
1
]]/'x
0
x
1
`
by Theorem 12.4.14. We write

= Q[[t]], where t = s(e
1
) for e
1
M = Ze
1
. As
explained in the discussion following Example 12.4.11, t acts on

H

T
(P
1
, Q) by
multiplication by ('e
1
, u
0
`x
0
+'e
1
, u
1
`x
1
) = x
0
x
1
.
Recall that (L) = 1 ++ +
d
Z[M] by Example 13.2.11. Since
maps to e
s(e
1
)
= e
t
, we obtain
(13.3.14)
T
(L) = 1+e
t
+ +e
dt
= d +1+

d+1
2

t + .
Using x
i
= [D
i
]
T
, we see that the equivariant Todd class of X =P
1
is
Td
T
(P
1
) =
x
0
1e
x
0
x
1
1e
x
1
= (1+
1
2
x
0
+
1
12
x
2
0
)(1+
1
2
x
1
+
1
12
x
2
1
)
= 1+
1
2
(x
0
+x
1
) +
1
12
(x
2
0
+x
2
1
) +
since x
0
x
1
= 0 in Q[[x
0
, x
1
]]/'x
0
x
1
`. The equivariant Chern character of L is
ch
T
(L) = e
[dD
0
]
T
= e
dx
0
= 1+dx
0
+
1
2
(dx
0
)
2
+ ,
and you will compute in Exercise 13.3.4 that
ch
T
(L)Td
T
(P
1
) = 1+(d +
1
2
)x
0
+
1
2
x
1
+(
1
2
d
2
+
1
2
d +
1
12
)x
2
0
+
1
12
x
2
1
+ .
The next step is to describe the equivariant integral

(P
1
)
eq :

H

T
(P
1
, Q) Q[[t]].
In Exercise 13.3.4 you will explain why

(P
1
)
eq x
0
=

(P
1
)
eq x
1
= 1. Since

(P
1
)
eq is a
Q[[t]]-module homomorphism, we have

(P
1
)
eq
x
2
0
=

(P
1
)
eq
(x
0
x
1
)x
0
=

(P
1
)
eq
t x
0
= t

(P
1
)
eq
x
0
= t.
A similar computation shows that

(P
1
)
eq x
2
1
=t (be sure you see where the minus
sign comes form). Applying

(P
1
)
eq to the above computation of ch
T
(L)Td
T
(P
1
),
we obtain

(P
1
)
eq
ch
T
(L)Td
T
(P
1
) = 0+d +
1
2
+
1
2
+(
1
2
d
2
+
1
2
d +
1
12
)t +
1
12
(t) +
= d +1+

d+1
2

t + .
13.3. Toric Equivariant Riemann-Roch 651
The rst two terms agree with what we computed in (13.3.14), and all terms agree
by the equivariant Riemann-Roch theorem. Note that HRR is the equality of the
constant terms.
The Simplicial Case. The equivariant Riemann-Roch theorem proved by Brion
and Vergne [50] applies when X = X

is complete and simplicial. The problem is


that the formula
Td
T
(X) =

[D

]
T
1e
[D]
T
for the equivariant Todd class no longer holds in the simplicial case. Similarly, for
HRR, the formula
(13.3.15) Td(X) =

[D

]
1e
[D]
needs to be modied when X is simplicial. Here is an example.
Example 13.3.10. Suppose that HRR holds for X =P(1, 1, 2). We will use the fan
shown in Figure 5 of Example 3.1.17, where the minimal generators are u
1
, u
2
and u
0
=u
1
2u
2
. The relations in the class group are
D
1
D
0
and D
2
2D
0
,
and the intersection pairing is determined by D
1
D
2
= 1 since u
1
, u
2
generate a
smooth cone. Note also that X = X
P
for P = Conv(0, 2e
1
, e
2
) and that D
P
= 2D
0
.
In Exercise 13.3.5 you will check the details of the following computations.
The cohomology ring of X is H

(X, Q) QQ[D
0
] Q[pt], where [D
0
]
2
=
1
2
[pt]. If we apply the formula (13.3.15) to X =P(1, 1, 2), then one computes that
(13.3.16) Td(X) = 1+2[D
0
] +
7
8
[pt].
The actual Todd class will have the form a +b[D
0
] +c[pt]. To determine the
constants a, b, c, we apply HRR to L =O
X
(D
P
), which gives
(O
X
(D
P
)) =

X
(1+[D
P
] +
1
2
[D
P
]
2
)(a+b[D
0
] +c[pt])
=

X
(1+[2D
0
] +
1
2
[2D
0
]
2
)(a+b[D
0
] +c[pt])
=

X
a+(b+2a)[D
0
] +(a
2
+b +c)[pt]
= a
2
+b +c.
On the other hand, (O
X
(kD
P
)) =[(P)M[ is the Ehrhart polynomial of P, which
by (10.5.13) is
Ehr
P
() = Area(P)
2
+
1
2
[PM[ +1 =
2
+2 +1.
652 Chapter 13. Toric Hirzebruch-Riemann-Roch
It follows that the actual Todd class is
Td(X) = 1+2[D
0
] +[pt].
This differs from Td(X) by
1
8
[pt]. We will soon see the theoretical reason for this
discrepancy.
Returning to the equivariant case, let X = X

be a complete simplicial toric


variety. To dene its equivariant Todd class Td
T
(X), consider the group G in the
quotient construction of X from 5.1, namely
G =

(t

) (C

)
(1)
[

t
/m,u)

= 1 for all m M

.
Each (1) gives the character

: (C

)
(1)
C

dened by projection on the th factor. Then for any cone , let


G

=g G [

(g) = 1 for all / (1)


=(t

) (C

)
(1)
[ t

= 1 for / (1),

(1)
t
/m,u)

= 1 for m M

.
One can show that
(13.3.17) G

/(

(1)
Zu

)
(Exercise 13.3.6), so that [G

[ = mult().
Then we set
G

G
and dene the equivariant Todd class of X to be
(13.3.18) Td
T
(X) =

gG

(1)
[D

]
T
1

(g)e
[D]
T
when X = X

is complete and simplicial. Here is the result proved in [50].


Theorem 13.3.11. Let X = X

be a complete simplicial toric variety. If L =


O
X
(D) is the line bundle of a torus-invariant Cartier divisor D on X, then

T
(L) =

X
eq
ch
T
(L)Td
T
(X),
where Td
T
(X) is dened in (13.3.18).
This result has also been proved by Edidin and Graham [87] using different
methods. Once we have a simplicial version of equivariant Riemann-Roch, the
proof of Theorem 13.3.8 gives HRR in the simplicial case as follows.
13.3. Toric Equivariant Riemann-Roch 653
Corollary 13.3.12. Let X = X

be a complete simplicial toric variety. If L is a


line bundle on X, then
(L) =

X
ch(L)Td(X),
where
Td(X) =

gG

(1)
[D

]
1

(g)e
[D]

Let us apply this to the previous example.
Example 13.3.13. Consider P(1, 1, 2) as in Example 13.3.10. In Exercise 13.3.5
you will show that G

= (1, 1, 1), (1, 1, 1) G = (t, t, t


2
) [ t C

. Thus
Td(X) is a sum of two terms. For g = (1, 1, 1), we gave

(g) = 1 for all , so that


this term is what we get when we use the formula for the smooth case. In (13.3.16),
we computed this to be
Td(X) = 1+2[D
0
] +
7
8
[pt].
For g = (1, 1, 1), we get the product
[D
0
]
1+e
[D
0
]

[D
1
]
1+e
[D
1
]

[D
2
]
1e
[D
2
]
,
which becomes
(
1
2
[D
0
] +
1
4
[D
0
]
2
) (
1
2
[D
1
] +
1
4
[D
1
]
2
) (1+
1
2
[D
2
] +
1
12
[D
1
]
2
) =
1
4
[D
0
][D
1
] =
1
8
[pt].
The terms for g = (1, 1, 1) and g = (1, 1, 1) sum to Td(X) = 1+2[D
0
] +[pt], in
agreement with Example 13.3.10.
In [229], Pommersheim denes the mock Todd class of a complete simplicial
toric variety to be the class computed using the formula for the smooth case, i.e.,
Td(X) =

[D

]
1e
[D]
.
He shows that the difference between the actual and mock Todd classes depends
on the codimension of the singular locus of X. He also expresses the difference in
codimension 2 in terms of Dedekind sums.
We should also mention that a nice discussion of Todd classes of general toric
varieties and their enumerative applications can be found in [105, Sec. 5.3].
Exercises for 13.3.
13.3.1. Let X = X

be complete and simplicial of dimension n, and for (n), let i

be
as in (13.3.6). Let , H
2n
T
(X, Q) satisfy

X
eq = 1 and

X
eq = 1 and assume there is
(n) such that i

() = 0 and i

() = 0 for all

= . Prove that = . Hint: Let


u = i

() and v = i

() in

. Then consider vu H
2n
T
(X, Q).
654 Chapter 13. Toric Hirzebruch-Riemann-Roch
13.3.2. The decomposition of the equivariant integral given in Theorem 13.3.7 involves
terms containing the product (1)
n

n
i=1
s(m
,i
) for (n). Here you will show that
this is the nth equivariant Chern class of the normal bundle of x

X.
(a) Use x

X to show that the normal bundle is C


n
where T acts on the ith
factor via
m,i
. Thus the normal bundle is the direct sum of the rank 1 equivariant
vector bundles given by the characters
m,1
, . . . ,
m,n
.
(b) If a rank n vector bundle is a direct sum of rank 1 bundles, show that its nth Chern
class is the product of the rst Chern classes of the factors.
(c) Explain why the proof of Proposition 12.4.13 implies that the rst equivariant Chern
of the ith factor of the normal bundle is s(m
,i
).
13.3.3. Prove (13.3.12). Hint: Adapt the proof of Theorem 13.3.7 using Lemma 13.3.5
and Exercise 13.2.3.
13.3.4. Supply the details omitted in Example 13.3.9.
13.3.5. Supply the details omitted in Examples 13.3.10 and 13.3.13.
13.3.6. Prove (13.3.17). Also show that G

acts on C
(1)
with quotient C
(1)
/G

.
13.3.7. Our discussion of the equivariant Euler characteristic used the map : Z[M]

dened by
m
e
s(m)
. This exercise will explore the canonical meaning of . We rst
replace Z[M] with Q[M] and note that Q[M] is the rational representation ring of T since
the
m
are the irreducible representations of T. The augmentation ideal of Q[M] is
I =

mM
a
m

m
Q[M] [

mM
a
m
= 0

.
We also replace

with

Sym =

k=0
Sym
k
Q
(M
Q
). Then we can think of as the map
: Q[M]

Sym,
m
e
m
.
Prove that induces an isomorphism between

Sym and the I-adic completion of Q[M] at
the augmentation ideal I. This is the canonical meaning of .
13.3.8. Let X =X
P
be a projective toric variety, where P M
R
is a full dimensional lattice
polytope, and let L =O
X
(D
P
). Then
T
(L) =

mPM
e
s(m)

2
.
(a) Show that the constant term of
T
(L) is [PM[.
(b) Show that the degree 2 term is

mPM
s(m) = s

mPM
m

. It is interesting to
note that
1
|PM|

mPM
m is the average lattice point of P.
In a similar way, the higher terms of
T
(L) can be interpreted as various moments of the
lattice points of P. The power of equivariant Riemann-Roch is that when P is smooth, all
of these terms can be computed using equivariant intersection theory on X = X
P
.
13.4. The Volume Polynomial
In this section, we discuss the relation between cup product and the volume of a
polytope, leading up to the volume polynomial. This will give a new description of
the cohomology ring of a complete simplicial toric variety.
13.4. The Volume Polynomial 655
Cup Product and Normalized Volume. By Proposition 10.5.6, the Euclidean area
of a lattice polygon P in the plane is given by the intersection formula
2Area(P) = D D,
where D is a torus-invariant nef divisor on a smooth complete toric surface X

such
that P = P
D
.
Recall from 9.5 that a polytope P M
R
R
n
has Euclidean volume vol(P),
where a fundamental parallelotope determined by a basis of M has volume 1. Then
we dene the normalized volume of P by
Vol(P) = n! vol(P).
Thus the standard n-simplex
n
R
n
has normalized volume Vol(
n
) = 1.
Theorem 13.4.1. Let D be a torus-invariant divisor on a smooth complete toric
variety X

of dimension n and set P = P


D
M
R
.
(a) If D is very ample and X

P
s
, s +1 = dim H
0
(X

, O
X

(D)) = [PM[, is
the projective embedding given by global sections of O
X

(D), then
deg(X

P
s
) =

[D]
n
= Vol(P).
(b) If D is nef, then

[D]
n
= Vol(P).
Proof. The degree of X

P
s
is the number of points in the intersection X

L,
where L P
s
is a generic linear subspace of codimension n. Bertinis theorem
(see [131, II.8.18]) and (12.5.6) imply that if V P
s
is smooth and H P
s
is a
sufciently generic hyperplane, then in the Chow ring A

(P
s
), we have [H] [V] =
[HV], where HV is smooth. Hence in H

(P
s
, Z) we have
deg(X

P
s
) =

P
s
[H
1
] [H
n
] [X

]
when H
1
, . . . , H
n
are sufciently generic. However, if i : X

P
s
is the inclusion,
the Gysin map i
!
: H
k
(X

, Z) H
k+2n
(P
s
, Z) satises i
!
i

= [X

] for all
H

(P
s
, Z). Setting = [H
1
] [H
n
], we get

P
s
[H
1
] [H
n
] [X

] =

P
s
[X

] =

P
s
i
!
i

[D]
n
by properties of Gysin maps discussed in Theorem 13.A.6. Thus the degree is the
intersection number

[D]
n
.
To bring volume into the picture, we note that the degree is also n! times the
leading coefcient of the Hilbert polynomial of the homogeneous coordinate ring
C[X

] = C[x
0
, . . . , x
s
]/I(X

) of X

P
s
. We proved in Proposition 9.4.3 that the
Hilbert polynomial is the Ehrhart polynomial of P, and by the discussion following
656 Chapter 13. Toric Hirzebruch-Riemann-Roch
Proposition 9.4.3, we see that the leading coefcient of the Ehrhart polynomial is
Vol(P)/n!. Hence the degree is Vol(P).
For part (b), the nef divisor D need not give a projective embedding. In fact,
X

need not be projective. So we instead use HRR. Let 1 be an integer. Then


[(P) M[ = dim H
0
(X

, O
X

(D)) = (O
X

(D)),
where the rst equality comes from Example 4.3.7, and the second follows from
Demazure vanishing. By HRR,
[(P) M[ = (O
X

(D)) =

ch(O
X

(D))Td(X

).
In ch(O
X

(D)), the termcontaining the highest power of is [D]


n
/n! =
n
[D]
n
/n!.
Since the degree zero term of Td(X

) is 1, we obtain

ch(O
X

(D))Td(X

) =

n
n!

[D]
n
+lower degree terms in .
Therefore, by (9.4.4), in the limit as , the normalized volume is given by
Vol(P) = n! lim

[(P) M[

n
= lim

n!

ch(O
X

(D))Td(X

) =

[D]
n
,
which is what we wanted to show.
If P is not full dimensional, then Vol(P) = 0. As in the discussion following
Lemma 9.3.9, D is a big divisor if Vol(P) > 0. A different proof that the degree
equals in the volume in the very ample case can be found in [113, Thm. 6.2.3].
We next extend part (b) of Theorem 13.4.1 to arbitrary complete toric varieties.
We will need the following fact.
Lemma 13.4.2. Assume that f : X Y is a birational morphism between complete
irreducible varieties of dimension n. If D is a Cartier divisor on Y and D

= f

D,
then

X
[D

]
n
=

Y
[D]
n
.
Proof. Set = [D]
n
and let p : Y pt be the constant map. Example 13.A.3
explains that

Y
= p

([Y]), where [Y] H


2n
(Y, Q) is the fundamental class
and p

: H

(Y, Q) H

(pt, Q) =Q. Then p

= p f is the constant map for X,


so that

X
[D

]
n
= p

( f

()[X]) = p

( f

()[X])
= p

( f

[X]) = p

([Y]) =

Y
[D]
n
.
Here, the second line uses Proposition 13.A.2, and in the third line, f

[X] = [Y]
follows from Proposition 13.A.5.
13.4. The Volume Polynomial 657
We can now prove a more general version of Theorem 13.4.1.
Theorem 13.4.3. Let D be a torus-invariant nef Cartier divisor on a complete toric
variety X

of dimension n. Then

[D]
n
= Vol(P
D
).
Proof. Let : X

be a toric resolution of singularities and set D

D.
The pullback D

D has the same support function as D by Proposition 6.2.7.


Then D

and D have the same polytope by Lemma 6.1.6, i.e., P


D
= P
D
. Thus
Vol(P
D
) = Vol(P
D
) =

[D

]
n
=

[D]
n
,
where the second equality uses Theorem 13.4.1 and the third equality follows from
Lemma 13.4.2.
The Cartier hypothesis can also be relaxed slightly; Theorem 13.4.3 is also true
for any nef Q-Cartier divisor D (Exercise 13.4.1). Here is an example illustrating
the theorem.
Example 13.4.4. See Figure 11 in Example 2.4.6 and consider the triangle P =
Conv(2e
1
+e
2
, 2e
1
e
2
, 2e
1
e
2
) in M
R
= R
2
. The toric variety X
P
is the
weighted projective plane P(1, 1, 2). Label the ray generators of the normal fan of
P as u
1
= e
1
, u
2
= e
2
, u
3
= e
1
2e
2
, and let
i
= Cone(u
i
), D
i
= D

i
. Then it is
easy to see that P = P
D
for the divisor D = 2D
1
+D
2
. On X
P
P(1, 1, 2), we have

X
P
[D]
2
= Vol(P) = 8
by Theorem 13.4.3. On the other hand,

X
P
[D]
2
= D
2
. Since D
2
1
=
1
2
(D
1
is not
Cartier), D
1
D
2
= 1, and D
2
2
= 2, one computes directly that D
2
= 8.
The proof of Theorem 13.4.3 shows that we can also do this computation using
a resolution of singularities. If we rene the normal fan of X
P
by introducing a
new ray
0
= Cone(e
2
), we get a smooth fan

. Indeed, the resulting surface


isomorphic to the Hirzebruch surface H
2
. Let : H
2
X
P
be the corresponding
morphism. If we let D

D = D
0
+2D
1
+D
2
, then one can verify directly that
P
D

= P, so that

H
2
[D

]
2
= Vol(P) = 8
by Theorem 13.4.1. On H
2
, D
1
D
3
, D
2
D
0
+2D
3
, so that D

2D
0
+4D
3
.
Since D
2
0
=2, D
2
3
= 0, and D
0
D
3
= 1, one computes that (

D)
2
= 8. You will
check these assertions in Exercise 13.4.2.
658 Chapter 13. Toric Hirzebruch-Riemann-Roch
The Volume Polynomial. Let X

be a complete simplicial toric variety of dimen-


sion n. The divisor

is Q-Cartier when the t

are all rational and hence has


a cohomology class in H
2
(X

, Q). The resulting integral


(13.4.1)

n
is a homogeneous polynomial of degree n in t

, (1). When D =

is
nef and Cartier, the integral equals Vol(P
D
) by Theorem 13.4.3. For this reason,
we call (13.4.1) the volume polynomial of X

.
Example 13.4.5. Let D
0
, . . . , D
n
be the torus-invariant prime divisors on P
n
. Then
the volume polynomial is

P
n
[t
0
D
0
+ +t
n
D
n
]
n
= (t
0
+ +t
n
)
n
since D
i
D
0
for 1 i n and

P
n
[D
0
]
n
= 1. Note that D = t
0
D
0
+ +t
n
D
n
is nef if and only if t
0
+ +t
n
0. In this case, the above computation and
Theorem 13.4.1 imply that P
D
has normalized volume (t
0
+ +t
n
)
n
.
Example 13.4.6. Let be the usual fan for (P
1
)
n
. Let the divisor D
2i1
correspond
to the ray Cone(e
i
) and D
2i
correspond to Cone(e
i
) for 1 i n. Then the
polytope of a general divisor D = t
1
D
1
+ +t
2n
D
2n
with t
1
, . . . , t
2n
0 is the
rectangular solid
P
D
= [t
1
, t
2
] [t
2n1
, t
2n
]
in M
R
= R
n
. The Euclidean volume is (t
1
+t
2
) (t
2n1
+t
2n
), which gives the
normalized volume
(13.4.2) Vol(P
D
) = n! (t
1
+t
2
) (t
2n1
+t
2n
).
The reason for the n! becomes clear when we compute the volume polynomial
in H

((P
1
)
n
, Q). By the calculations in 12.3 and 12.4,
H

((P
1
)
n
, Q) Q[x
1
, . . . , x
n
]/'x
2
1
, . . . , x
2
n
`,
with x
i
= [D
2i1
], 1 i n. The top-dimensional component of the cohomology
ring is H
2n
((P
1
)
n
, Q), generated by the product x
1
x
n
. The cohomology class of
D is [D] = (t
1
+t
2
)x
1
+ +(t
2n1
+t
2n
)x
n
H
2
((P
1
)
n
, Q). Then
((t
1
+t
2
)x
1
+ +(t
2n1
+t
2n
)x
n
)
n
= n! (t
1
+t
2
) (t
2n1
+t
2n
)x
1
x
n
,
since there are n! ways to order the factors in x
1
x
n
, and all other terms in the nth
power are zero. Since

(P
1
)
n
x
1
x
n
= 1, we get (13.4.2).
In general, as in these examples, if we start from a complete simplicial fan
with [(1)[ = r and the divisors D
1
, . . . , D
r
on X

corresponding to
1
, . . . ,
r
in
(1), then the volume polynomial (13.4.1) becomes
V(t
1
, . . . , t
r
) =

[t
1
D
1
+ +t
r
D
r
]
n
.
13.4. The Volume Polynomial 659
The classes [D
i
] are not linearly independent in H
2
(X

, Q). There is also a


more efcient reduced volume polynomial, which is constructed using divisors D
i
whose classes give a basis of H
2
(X

, Q) = Pic(X

)
Q
. In Example 13.4.5, for
instance, Pic(P
n
)
Q
has dimension 1, and the reduced volume polynomial is just
V(t) =t
n
. In Example 13.4.6, on the other hand, Pic((P
1
)
n
)
Q
has dimension n and
the reduced volume polynomial is V(t
1
, . . . , t
n
) = n! t
1
t
n
.
The Volume Polynomial and the Cohomology Ring. An alternative description of
the cohomology ring H

(X

, Q) for complete simplicial toric varieties was proved


by Khovanskii and Pukhlikov in [174]. We will sketch the ideas following the
presentation in [170] with some modications.
The relation between the volume polynomial and the cohomology ring will
come from the following algebraic fact.
Lemma 13.4.7. Let A =

m
i=0
A
i
be a nite-dimensional commutative graded
algebra over Q satisfying:
(a) A
0
A
m
Q.
(b) A is generated by A
1
as a Q-algebra.
(c) For all i = 0, . . . , m, the bilinear map
A
i
A
mi
A
m
Q
(u, v) uv
is nondegenerate.
Let s
1
, . . . , s
r
span A
1
and dene
P(t
1
, . . . , t
r
) = (t
1
s
1
+ +t
r
s
r
)
m
A
m
Q,
which we regard as a polynomial in Q[t
1
, . . . , t
r
]. Finally, dene
I =

f (x
1
, . . . , x
r
) Q[x
1
, . . . , x
r
]


t
1
, . . . ,

t
r

P = 0

.
Then the map x
i
s
i
induces an isomorphism of graded Q-algebras
Q[x
1
, . . . , x
r
]/I A.
Here f


t
1
, . . . ,

tr

is the differential operator obtained from f by replacing


each x
i
with

t
i
. We will postpone the proof of the lemma until we see how it
applies to H

(X

, Q).
The rings satisfying the hypotheses in Lemma 13.4.7 are a special class of
nite-dimensional graded Gorenstein rings. If X

is complete and simplicial, then


hypothesis (b) is satised by the cohomology ring H

(X

, Q) by Theorem 12.4.1.
It is easy to see that hypothesis (a) holds from the cohomology spectral sequence
in 12.3. Then hypothesis (c) follows from Poincar e duality on X

. As a result,
660 Chapter 13. Toric Hirzebruch-Riemann-Roch
the cohomology rings of complete simplicial toric varieties are Gorenstein rings of
this type. Here is the Khovanskii-Pukhlikov presentation of the cohomology ring.
Theorem 13.4.8. Let X

be a complete simplicial toric variety. Let r =[(1)[ and


s = dim H
2
(X

, Q) = dim Pic(X

)
Q
. Then:
(a) Let V(t
1
, . . . , t
r
) be the volume polynomial of X

. Then there is an isomorphism


of Q-algebras
H

(X

, Q) Q[x
1
, . . . , x
r
]/I,
where I =

f (x
1
, . . . , x
r
) Q[x
1
, . . . , x
r
]


t
1
, . . . ,

tr

V(t
1
, . . . , t
r
) = 0

. In
addition, I equals the ideal I +. from Theorem 12.4.1.
(b) Let V(t
1
, . . . , t
s
) be a reduced volume polynomial for X

. Then there is an
isomorphism of Q-algebras,
H

(X

, Q) Q[x
1
, . . . , x
s
]/J,
where J =

f (x
1
, . . . , x
s
) Q[x
1
, . . . , x
s
]


t
1
, . . . ,

ts

V(t
1
, . . . , t
s
) = 0

.
Proof. The isomorphisms in parts (a) and (b) follow from Lemma 13.4.7, and the
equality I =I +. then follows from Theorem 12.4.1 (Exercise 13.4.3).
Here are some examples.
Example 13.4.9. For instance, if X

=P
n
, then from Example 13.4.5
V(t
0
, . . . , t
n
) = (t
0
+ +t
n
)
n
.
It is easy to check directly that the ideal I from Theorem 13.4.8 is
I ='x
1
x
0
, . . . , x
n
x
0
, x
0
x
n
` =I +.
(Exercise 13.4.4). Using the reduced volume polynomial V(t) = t
n
, one computes
that J ='x
n+1
`. Hence we get isomorphisms
Q[x
0
, . . . , x
n
]/I Q[x]/'x
n+1
` H

(P
n
, Q).
Both isomorphisms were known previously; the surprise is seeing how they arise
from volume polynomials.
Example 13.4.10. The volume polynomial computed in Example 13.4.6 is
V(t
1
, . . . , t
2n
) = n! (t
1
+t
2
) (t
2n1
+t
2n
).
Here, one computes that the ideal I from Theorem 13.4.8 is
I ='x
1
x
2
, . . . , x
2n1
x
2n
, x
1
x
2
, . . . , x
2n1
x
2n
` =I +.
(Exercise 13.4.4). Furthermore, the reduced volume polynomial V(t
1
, . . . , t
n
) =
n! t
1
t
n
gives the ideal
J ='x
2
1
, . . . , x
2
n
`.
This gives the presentation of the cohomology ring used in Example 13.4.6.
We now turn to the proof of Lemma 13.4.7.
13.4. The Volume Polynomial 661
Proof of Lemma 13.4.7. It is easy to prove that
I =

f (x
1
, . . . , x
r
)


t
1
, . . . ,

t
r

P = 0

is an ideal in Q[x
1
, . . . , x
r
].
Using hypothesis (a) and spanning set s
1
, . . . , s
r
for A
1
, we have a surjection
: Q[x
1
, . . . , x
r
] A,
f f (s
1
, . . . , s
r
),
and we need to show ker() = I. Since P(t
1
, . . . , t
r
) is a homogeneous polynomial
of degree m, I will be generated by homogeneous polynomials, and I will contain
all homogeneous polynomials of degree strictly larger than m. Note also that we
have the expansion
P(t
1
, . . . , t
n
) = (t
1
s
1
+ +t
r
s
r
)
m
=

1
++r=m
m!

1
!
r
!
t

1
1
t
r
r
s

1
1
s
r
r
.
If f (x
1
, . . . , x
r
) is homogeneous of degree m, then a direct calculation shows that
f


t
1
, . . . ,

t
r

P(t
1
, . . . , t
r
) = m! f (s
1
, . . . , s
r
)
(Exercise 13.4.5). Hence
f I f (s
1
, . . . , s
r
) = 0 f ker()
in this case.
Now assume f is homogeneous of degree p < m. Suppose that f / ker(),
so f (s
1
, . . . , s
r
) = 0. By hypothesis (c) in the Proposition, there is a homogeneous
polynomial g of degree m p such that g(s
1
, . . . , s
r
) f (s
1
, . . . , s
r
) = 0 A
m
. Thus,
by the rst part of the proof, gf / I so f / I. Thus f I implies f ker(). Finally,
suppose f ker(), so f (s
1
, . . . , s
r
) = 0. You will show in Exercise 13.4.5 that
f


t
1
, . . . ,

t
r

P(t
1
, . . . , t
r
) = f (s
1
, . . . , s
r
)h(t
1
, . . . , t
r
, s
1
, . . . , s
r
),
where h(t
1
, . . . , t
r
, s
1
, . . . , s
r
) is given by

1
++r=mp
m!

1
!
r
!
(t
1
s
1
)

1
(t
r
s
r
)
r
.
Since we assume f (s
1
, . . . , s
r
) = 0, this shows that f I.
In [170], Kaveh shows that the results of this section can be generalized to
spherical varieties, which are varieties with an action of a reductive algebraic group
G such that some Borel subgroup of G has a dense orbit.
662 Chapter 13. Toric Hirzebruch-Riemann-Roch
Exercises for 13.4.
13.4.1. Show that Theorem 13.4.3 is also true if the divisor D is nef and Q-Cartier.
13.4.2. Check the assertions made in Example 13.4.4.
13.4.3. Complete the proof of part (a) of Theorem 13.4.8 by showing that if X

is a
complete simplicial toric variety, then the ideal I from the theorem is the ideal I +.
from Theorem 12.4.1. Hint: Think about the kernel of the homomorphismQ[x
1
, . . . , x
r
]
H

(X

, Q) that takes x
i
to [D
i
].
13.4.4. In this exercise, you will check some claims made in Examples 13.4.9 and 13.4.10.
(a) In Example 13.4.9, verify that
I ='x
1
x
0
, . . . , x
n
x
0
, x
0
x
n
`.
Hint: One inclusion is clear. For the other, show that the ideal on the right-hand
side is 'x
1
x
0
, . . . , x
n
x
0
, x
n+1
0
`. Also note that f Q[x
0
, . . . , x
n
] can be written
as f =

n
i=1
(x
i
x
0
)A
i
(x
0
, . . . , x
n
) +g(x
0
). Remember that I is homogeneous.
(b) In Example 13.4.10, verify that I is as claimed.
(c) In both examples, compute the ideal J coming from the reduced volume polynomial.
13.4.5. In this exercise, you will verify some details of the proof of Lemma 13.4.7.
(a) Show that if f (x
1
, . . . , x
r
) is homogeneous of degree m, then
f


t
1
, . . . ,

t
r

P(t
1
, . . . , t
r
) = m! f (s
1
, . . . , s
r
).
(b) Show that if f (x
1
, . . . , x
r
) is homogeneous of degree p < m, then
f


t
1
, . . . ,

t
r

P(t
1
, . . . , t
r
) = f (s
1
, . . . , s
r
)h(t
1
, . . . , t
r
, s
1
, . . . , s
r
),
where h(t
1
, . . . , t
r
, s
1
, . . . , s
r
) is given by

1++r =mp
m!

1
!
r
!
(t
1
s
1
)
1
(t
r
s
r
)
r
.
13.4.6. Consider the Hirzebruch surface H
r
.
(a) Determine the volume polynomial and verify the isomorphism from Theorem 13.4.8
in this case.
(b) In the notation of Example 12.4.3, compute the reduced volume polynomial V(x
3
, x
4
)
using the basis of H
2
(H
r
, Q) =Pic(H
r
)
Q
given by D
3
, D
4
. Then verify that this gives
the presentation of the cohomology ring constructed in Example 12.4.3.
13.4.7. Let be a complete simplicial fan in N
R
R
n
. Following [170], we present
another way to think about the reduced volume polynomial when X

is projective.
(a) Let S

be the set of all lattice polytopes with normal fan , modulo the equivalence
relation P P

if P

is a translate of P. Show that S

forms a semigroup under the


operation of Minkowski sum of polytopes. Hint: Use Proposition 6.2.13.
(b) Let V

be the vector space generated by S

over Q. Show that there is a natural


isomorphism
V

H
2
(X

, Q).
13.5. The Khovanskii-Pukhlikov Theorem 663
(c) The normalized volume function Vol : S

Q is homogeneous of degree n. Show


that there is V Sym
n
(V

) such that V : V

Q extends Vol, i.e., V[


S
= Vol. Hint:
You may nd it useful to rst do part (d) of the exercise.
(d) Let s = dimH
2
(X

, Q) and consider a reduced volume polynomial V(t


1
, . . . , t
s
) for
X

, which we construct using prime toric divisors D


i
that give a basis of H
2
(X

, Q).
Show that this basis gives an isomorphismV

Q
s
that takes V to V(t
1
, . . . , t
s
).
13.5. The Khovanskii-Pukhlikov Theorem
Given a positive integer n and a function f on the interval [0, n], the Euler-Maclaurin
summation formula relates the sum of the values of f at the integers 0, . . . , n to the
integral

n
0
f (x)dx. In [174], Khovanskii and Pukhlikov use Brions equalities and
Todd differential operators to give an analogous formula relating the sum of the
values of a suitable function f at the lattice points in a polytope to integrals over
polytopes. Specializing to the constant function f = 1 gives a formula for the
number of lattice points in a polytope. We will see that the Khovanskii-Pukhlikov
theorem gives another proof of Hirzebruch-Riemann-Roch for a smooth projective
toric variety.
The Euler-Maclaurin Formula. In 13.1, we saw that the Bernoulli numbers B
k
give the power series
(13.5.1)
x
1e
x
= 1+
1
2
x +

k=1
(1)
k1
B
k
(2k)!
x
2k
.
This series converges for all x C with [x[ < 2. Bernoulli numbers also play a
key role in the Euler-Maclaurin formula, which can be stated as follows.
Theorem 13.5.1 (Euler-Maclaurin Summation). If f is a C

function on [0, n],


then

n
0
f (x)dx =
1
2
f (0) + f (1) + f (2) + + f (n1) +
1
2
f (n)
+

k=1
(1)
k
B
k
(2k)!
( f
(2k1)
(n) f
(2k1)
(0)) +R
2
,
where R
2
is a remainder term. Furthermore, if there are positive constants C and
< 2 such that [ f
()
(x)[ C

for all x [0, n] and all 1, then

n
0
f (x)dx =
1
2
f (0) + f (1) + f (2) + + f (n1) +
1
2
f (n)
+

k=1
(1)
k
B
k
(2k)!
( f
(2k1)
(n) f
(2k1)
(0)).
664 Chapter 13. Toric Hirzebruch-Riemann-Roch
The explicit form of the remainder is not important for us here; it and a proof
of the summation formula are given in [5]. The Euler-Maclaurin formula is often
used as a generalization of the trapezoidal rule for approximating the integral of
f . We should mention that there are a number of different competing conventions
for writing Bernoulli numbers, so Euler-Maclaurin formulas in other sources may
look different from Theorem 13.5.1.
Todd Operators. We will also use formal Todd differential operators obtained from
the series expansion in (13.5.1). The denition is given by
(13.5.2) Todd(x) = 1+
1
2

x
+

k=1
(1)
k1
B
k
(2k)!

2k
x
2k
.
For notational convenience, we will write Todd operators as
Todd(x) =
/x
1e
/x
,
but an expression of this form should always be interpreted as the series (13.5.2).
A key property of the Todd operator for us will be the equation
(13.5.3) Todd(x)(e
xz
) =
ze
xz
1e
z
,
which follows easily by a direct calculation:
Todd(x)(e
xz
) = e
xz
+
1
2
ze
xz
+

k=1
(1)
k1
B
k
(2k)!
z
2k
e
xz
= e
xz

1+
1
2
z +

k=1
(1)
k1
B
k
(2k)!
z
2k

= e
xz

z
1e
z
.
This computation is valid for all z C satisfying [z[ < 2.
A Consequence of Brions Equalities. Let P be a simple lattice polytope in M
R
.
Recall from Brions equalities (Corollary 13.2.10) that
(13.5.4)

mPM

m
=

v vertex

mPvM

n
i=1
(1
m
v,i
)
.
Here, the m
v,i
are the minimal generators of the cone C
v
= Cone(PMv), and
P
v
=

n
i=1

i
m
v,i
[ 0
i
< 1
is the fundamental parallelotope of the simplicial cone C
v
. Equation (13.5.4) is an
equality of two elements of the localization Z[M]
S
. In 13.3, we mapped equations
of this form to the localized ring

S
that arises in equivariant cohomology via the
13.5. The Khovanskii-Pukhlikov Theorem 665
map
m
e
s(m)
, and we used the resulting equations in our proof of the equivariant
Riemann-Roch theorem.
In this section, on the other hand, we will consider a different consequence
of Brions equalities. Take a point z =

i
z
i
u
i
N
C
= N
Z
C, where z
i
C and
u
i
N. Each u
i
gives the one-parameter subgroup
u
i
: C

T
N
. The point p
z
=

u
i
(e
z
i
) T
N
depends only on z. Here, e is the base of the natural logarithm.
Then observe that for any m M, we have

m
(p
z
) = e
/m,z)
C

.
Choose z N
C
so that 'm
v,i
, z` / 2i Z for all m
v,i
in (13.5.4). Then evaluating
(13.5.4) at the point p
z
maps
m
to e
/m,z)
and hence gives the relation
(13.5.5)

mPM
e
/m,z)
=

v vertex
e
/v,z)

mPvM
e
/m,z)

n
i=1
(1e
/m
v,i
,z)
)
since none of the denominators vanish.
This shows the power of the what we did in 13.2. The version of Brions
equalities proved there yields different equalities depending on how we evaluate
the
m
, giving results that can be used in very different contexts.
From Discrete to Continuous. Brions equalities in the form (13.5.5) deal with
discrete sums over sets of lattice points in polytopes. Our next results deal with
continuous analogs of these sums, namely integrals over polytopes, and their rela-
tion with discrete sums. For the application to the Khovanskii-Pukhlikov theorem,
we need to consider a class of polytopes more general than lattice polytopes.
Theorem 13.5.2. Let P M
R
be a full dimensional simple polytope, and assume
the rays of the vertex cones C
v
are spanned by primitive vectors m
v,i
M for all
vertices v. If z N
C
satises 'm
v,i
, z` / 2i Z for all m
v,i
, then

P
e
/x,z)
dx = (1)
n

v vertex
e
/v,z)
mult(C
v
)

n
i=1
'm
v,i
, z`
.
Proof. The integral on the left is equal to the limit of a sum as follows. Scale the
lattice M by a factor of
1
k
, and consider the points m P
1
k
M. Approximate P
by a collection of small cubes of side
1
k
centered at all m P
1
k
M. Then evaluate
the exponential function e
/m,u)
at these points and multiply by the volumes of the
cubes to form a Riemann sum. In the limit as k we have
(13.5.6)

P
e
/x,z)
dx = lim
k
1
k
n

mP
1
k
M
e
/m,z)
.
The next step is to apply (13.5.5) to P and the scaled lattice
1
k
M. The original
fundamental parallelotope P
v
=P
v,M
depends on M. For
1
k
M, the primitive ray gen-
erators are
1
k
m
v,i
, and it follows that the
1
k
M-lattice points of the new fundamental
666 Chapter 13. Toric Hirzebruch-Riemann-Roch
parallelotope are given by
(13.5.7) P
v,
1
k
M

1
k
M =
1
k
(P
v
M)
(Exercise 13.5.1). When we combine this with (13.5.5), we see that the right-hand
side of (13.5.6) can be written as
lim
k
1
k
n

v vertex
e
/v,z)

mPvM
e
/m/k,z)

n
i=1
(1e
/m
v,i
/k,z)
)
= lim
k

v vertex
e
/v,z)

mPvM
e
/m/k,z)

n
i=1
k(1e
/m
v,i
/k,z)
)
.
For the term in the sum for the vertex v, the limit of the numerator is
lim
k
e
/v,z)

mPvM
e
/m/k,z)
= e
/v,z)
[P
v
M[ = e
/v,z)
mult(C
v
)
by Proposition 11.1.8. For the denominator

n
i=1
k(1e
/m
v,i
/k,z)
), note that
lim
k
k(1e
/m
v,i
/k,z)
) = lim
k
1e
/m
v,i
,z)/k
1/k
='m
v,i
, z`
by LH opitals Rule. Reassembling the different parts of the computation, we get
the desired expression for the integral.
The Khovanskii-Pukhlikov Theorem. Let P be a simple lattice polytope dened
by inequalities of the form
'm, u
F
` +a
F
0,
where u
F
are the facet normals. Let h = (h
F
)
F facet
be a vector with real entries
indexed by the facets F of P. Consider the polytope P(h) with shifted facets dened
by the inequalities
(13.5.8) 'm, u
F
` +a
F
+h
F
0.
Note that the shifting factors h
F
for the different facets are independent. It is not
difcult to see that if all entries of h are sufciently small in absolute value, then
P(h) is still simple (Exercise 13.5.2). If some of the h
F
are not rational, then the
vertices might not be rational points. We want to consider what happens when we
apply differential operators with respect to the h
F
to integrals over the correspond-
ing polytopes, and then set h = 0.
The differential operators alluded to above are the formal multivariate Todd
differential operators dened using (13.5.2):
Todd(h) =

F facet
/h
F
1e
/h
F
.
We are now ready to state the Khovanskii-Pukhlikov theorem for smooth lattice
polytopes, the main result of this section.
13.5. The Khovanskii-Pukhlikov Theorem 667
Theorem 13.5.3. Let P be a smooth lattice polytope. Then
Todd(h)

P(h)
e
/x,z)
dx

h=0
=

mPM
e
/m,z)
provided z N
C
satises 0 < ['m
v,i
, z`[ < 2 for all primitive ray generators m
v,i
of the vertex cones C
v
of P.
Before we give the proof, we want to indicate exactly how this result relates to
the Euler-Maclaurin formula from the start of the section.
Example 13.5.4. Let P = [0, n] R. The endpoints of P are its facets, so the facet
normals are 1, and the facet equations of P are
'm, 1` = m 0 and 'm, 1` =mn.
Hence, the shifted polytope P(h) is the interval P(h) = [h
0
, n+h
1
] where h
0
, h
1
are real-valued independent variables. Fix z C such that 0 < [z[ < 2 and let
f (x) = e
xz
. In Exercise 13.5.3, you will show that for h = (h
0
, h
1
),
(13.5.9)
Todd(h)

n+h
1
h
0
f (x)dx

h
0
=h
1
=0
=

n
0
f (x)dx +
1
2
( f (n) + f (0))
+

k=1
(1)
k1
B
k
(2k)!
( f
(2k1)
(n) f
(2k1)
(0)).
However, Theorem 13.5.3 implies that
Todd(h)

n+h
1
h
0
e
xz
dx

h
0
=h
1
=0
=

m[0,n]Z
e
mz
.
Since f (x) = e
xz
, we can write this as
Todd(h)

n+h
1
h
0
f (x)dx

h
0
=h
1
=0
=

m[0,n]Z
f (m) = f (0) + + f (n).
Combining this with (13.5.9) gives the Euler-Maclaurin formula

n
0
f (x)dx =
1
2
f (0) + f (1) + f (2) + + f (n1) +
1
2
f (n)
+

k=1
(1)
k
B
k
(2k)!
( f
(2k1)
(n) f
(2k1)
(0))
from Theorem 13.5.1 for f (x) = e
xz
when 0 <[z[ < 2.
From this point of view, we see that Theorem 13.5.3 is the polytope version
of the Euler-Maclaurin formula for exponential functions of the form f (x) = e
/x,z)
.
In [174], Khovanskii and Pukhlikov study more general functions (polynomials
times exponentials) and prove a related result in this case. We should also mention
668 Chapter 13. Toric Hirzebruch-Riemann-Roch
that there is a large and growing literature on various forms of generalized Euler-
Maclaurin formulas on polytopes. See the notes at the end of [22, Ch. 10] for
references.
Now we turn to the proof of Theorem 13.5.3.
Proof of Theorem 13.5.3. Since P is smooth, each vertex cone C
v
has multiplicity
mult(C
v
) = 1. Then Theorem 13.5.2 implies that

P
e
/x,z)
dx = (1)
n

v vertex
e
/v,z)

n
i=1
'm
v,i
, z`
.
Our rst goal is to determine the corresponding equation with P replaced by P(h).
For this, we need to determine how the vertices of P(h) are related to the vertices
of P. First we note that since each facet of P shifts to a parallel facet in P(h), the
m
v,i
do not change, and hence Theorem 13.5.2 will apply to all P(h) as well.
If v is the intersection of facets F
1
(v), . . . , F
n
(v), then the ray generators m
v,i
of
the vertex cone at v are the basis of M dual to the basis u
F
1
(v)
, . . . , u
Fn(v)
of N. It
follows that
'v

n
i=1
h
F
i
(v)
m
v,i
, u
F
j
(v)` =a
F
j
(v)
h
F
j
(v)
.
So the vertex v of P shifts to v

n
i=1
h
F
i
(v)
m
v,i
in P(h). For simplicity, we will
write h
F
i
(v)
as h
i
(v) in the following. Hence, applying Theorem 13.5.2 to P(h),
(13.5.10)

P(h)
e
/x,z)
dx = (1)
n

v vertex
e
/v
P
i
h
i
(v)m
v,i
,u)

n
i=1
'm
v,i
, z`
.
Now, we apply the Todd operator to both sides of (13.5.10). Since each sum-
mand of the right hand side factors as
e
/v,z)

n
i=1
'm
v,i
, z`
e
/
P
i
h
i
(v)m
v,i
,z)
= c
v
e
/
P
i
h
i
(v)m
v,i
,z)
,
where c
v
does not depend on h, we have Todd(h)

P(h)
e
/x,z)
dx

=
(13.5.11)
= (1)
n

v vertex
c
v
Todd(h)

e
/
P
i
h
i
(v)m
v,i
,z)

= (1)
n

v vertex
c
v

e
/
P
i
h
i
(v)m
v,i
,z)
n

i=1
'm
v,i
, z`
(1e
/m
v,i
,z)
)

,
where the second line follows by applying (13.5.3) for each h
F
(Exercise 13.5.4).
Setting h
F
= 0 for all F and some algebraic simplication between c
v
and the other
factor in each term yields

v vertex
e
/v,z)

n
i=1
(1e
/m
v,i
,z)
)
,
13.5. The Khovanskii-Pukhlikov Theorem 669
which is exactly the sum on the right in (13.5.5). (Recall that we are assuming
the vertex cones of P are smooth, so P
v
M = 0 for all v). This completes the
proof of the theorem because the other side of the equality in (13.5.5) is the sum

mPM
e
/m,z)
and this is what we wanted.
Theorem 13.5.3 has the following nice consequence.
Corollary 13.5.5. For a smooth lattice polytope P
Todd(h)

vol(P(h))

h=0
=[PM[,
where vol is the usual Euclidean volume.
Proof. Theorem 13.5.3 applies to the exponential function f (x) = e
/x,z)
for z N
C
satisfying 0 < ['m
v,i
, z`[ < 2 for all m
v,i
. Then, taking the limit as z 0 in the
theorem implies that
Todd(h)

P(h)
e
/x,0)
dx

h=0
=

mPM
e
/m,0)
,
which gives the desired result. Taking the limit inside the Todd operator and then
inside the integral takes some care. We omit the details.
Relation to HRR. It turns out that Corollary 13.5.5 can be used to prove HRR for
smooth projective toric varieties.
Theorem 13.5.6. Let L be a line bundle on a smooth projective toric variety
X = X

. Then
(L) =

X
ch(L)Td(X).
Proof. We will sketch the proof, though several points in the discussion require
justication that we will omit. We begin with a very ample divisor D =

and set P = P
D
. Then part (a) of Theorem 13.4.1 implies that
(13.5.12)

X
[D]
n
= Vol(P).
Unlike part (b) of Theorem 13.4.1, the proof of part (a) does not use HRR. Note
also that this equation holds when D =

is ample since replacing D with


D multiplies each side by
n
.
The next step is to let h = (h

)
(1)
, where each h

is real, and set D(h) =

(a

+h

)D

. The cohomology class [D(h)] lives in H

(X, R), and the polytope


associated to D(h) is dened by
'm, u

` +a

+h

0, (1),
670 Chapter 13. Toric Hirzebruch-Riemann-Roch
which as above is denoted by P(h). While P(h) is no longer a lattice polytope, its
normal fan is still when h is small. (Earlier, we indexed h by facets of P, which
correspond to the rays since =
P
.) For h sufciently small, (13.5.12) implies
(13.5.13)

X
[D(h)]
n
= Vol(P(h))
for h rational (by rescaling) and then for h arbitrary (by continuity).
Note that the Todd operators Todd(h

) act naturally on cohomology classes


in H

(X, R) that depend polynomially on h. An example is e


h[D]
, which is a
polynomial in h

since [D

] has degree 2. The basic relation (13.5.3) implies that


Todd(h

e
h[D]
) =
[D

] e
h[D]
1e
[D]
.
Then we compute
Todd(h)

e
[D(h)]

h=0
=

Todd(h

e
(a+h)[D]

h=0
=

[D

] e
a[D]
1e
[D]
= ch(L)Td(X)
since ch(L) = e
[
P

aD]
and Td(X) =

[D

]/(1e
[D]
) by Theorem 13.1.6.
The integral

X
: H

(X, R) R is a linear map, which implies that



X
and
Todd(h) commute when applied to e
[D(h)]
. Since

X
kills everything in degree
different from 2n, we obtain

X
ch(L)Td(X) =

X
Todd(h)

e
[D(h)]

h=0
= Todd(h)

X
e
[D(h)]

h=0
= Todd(h)

X
1
n!

D(h)

h=0
= Todd(h)

1
n!
Vol(P(h))

h=0
= Todd(h)

vol(P(h))

h=0
,
where the fourth equality follows from (13.5.13).
Note that P=P
D
is a smooth lattice polytope since D is ample and X is smooth.
Thus we can bring Corollary 13.5.5 into the picture, which implies
Todd(h)

vol(P(h))

h=0
=[PM[ = (O
X
(D)) = (L).
Combining this with the previous display gives the Hirzebruch-Riemann-Roch
equality
(13.5.14)

X
ch(L)Td(X) = (L)
13.5. The Khovanskii-Pukhlikov Theorem 671
in this special case when L =O
X
(D) and D is ample.
Now pick divisors D
1
, . . . , D
s
whose classes give a basis of Pic(X) and let
D(a) =

s
i=1
a
i
D
i
and L(a) =O
X
(D(a)). It sufces to prove that
(13.5.15)

X
ch(L(a))Td(X) = (L(a))
for all a =(a
1
, , a
s
) Z
s
. The left-hand side of (13.5.15) is clearly a polynomial
in a, and the same is true for the Euler characterstic (L(a)) on the right-hand
side. When D
1
is very ample, we proved in Exercise 9.4.2 that (O
X
(a
1
D
1
))
is a polynomial in a
1
, and a similar result for a
1
D
1
+ + a
s
D
s
(for arbitrary
D
1
, . . . , D
s
) is proved in [176].
If a Z
s
is chosen so that D(a) is ample, then equality holds for (13.5.15) by
(13.5.14). However, since X is projective, the ample divisor classes are the lattice
points in the interior of the nef cone in Pic(X)
R
. Since two polynomials on Pic(X)
R
that agree on the lattice points in an open cone must be equal (Exercise 13.5.5), we
conclude that (13.5.15) is true for all a Z
s
. This completes our second proof of
the Hirzebruch-Riemann-Roch theorem for toric varieties.
The version of HRR proved here is not as general as Theorem 13.3.8, which
only assumes that X is complete. Nevertheless, it is amazing how the relatively
elementary techniques leading to the Khovanskii-Pukhlikov theorem are so closely
related to a deep theorem about the geometry of toric varieties.
Exercises for 13.5.
13.5.1. Prove (13.5.7).
13.5.2. Show that if [h
F
[ is sufciently small for all F in (13.5.8), then P is still simple.
13.5.3. In this exercise, you will give two proofs of (13.5.9). Assume 0 <[z[ < 2.
(a) Prove (13.5.9) using the power series denition of Todd(h) = Todd(h
0
)Todd(h
1
).
(b) Compute

n+h1
h0
e
xz
dx explicitly and then use (13.5.3) to show that
Todd(h)

n+h1
h0
e
xz
dx

h0=h1=0
=
e
nz
1 e
z
+
1
1 e
z
.
Then apply (13.5.1) to the right-hand side to derive (13.5.9).
(c) Use Example 13.2.11 to show that right-hand side of the equation from part (b) is
equal to 1 +e
z
+ +e
nz
. Thus we get a direct proof of Theorem 13.5.3 in this case.
13.5.4. This exercise deals with some details in the proof of Theorem 13.5.3.
(a) Show that the second line of (13.5.11) equals the rst by using (13.5.3) for each h
F
.
(b) Showthat the second line in (13.5.11) simplies to yield the right-hand side of (13.5.5).
13.5.5. Assume that f R[x
1
, . . . , x
s
] vanishes on all lattice points in the interior of a cone
in R
s
of dimension s. Prove that f = 0.
672 Chapter 13. Toric Hirzebruch-Riemann-Roch
Appendix: Generalized Gysin Maps
Here we collect some facts we need about Borel-Moore homology and generalized Gysin
maps. This material is discussed briey in [105, Ch. 19] and in [110] in more detail. There
is also a nice treatment of Borel-Moore homology in [106, App. B].
Borel-Moore Homology. Besides ordinary homology, a suitably nice topological space X
has Borel-Moore homology groups dened by
H
BM
i
(X, Q) = H
ni
(R
n
, R
n
` X, Q)
for a closed embedding X R
n
. These groups are independent of the embedding and are
dened for any variety X in its classical topology.
Proposition 13.A.1 (Basic Properties).
(a) If X is compact, then H
BM
i
(X, Q) = H
i
(X, Q).
(b) There is a cap product operation
H
i
(X, Q)
Q
H
BM
j
(X, Q) H
BM
ji
(X, Q),
such that
() = ( )
for all , H

(X, Q) and H
BM

(X, Q).
The functorial properties of Borel-Moore homology are more complicated since a
continuous map f : X Y does not always induce a map f

: H
BM
i
(X, Q) H
BM
i
(Y, Q).
Proposition 13.A.2 (Functorial Properties).
(a) A proper map f : X Y induces f

: H
BM
i
(X, Q) H
BM
i
(Y, Q) such that
f

( f

() ) = f

()
for all H

(Y, Q) and H
BM

(X, Q).
(b) If f : X Y and g : Y Z are proper, then (g f )

= g

.
(c) An inclusion j : U Y of an open set induces j
!
: H
BM
i
(X, Q) H
BM
i
(U, Q) such that
j
!
() = j

() j
!
()
for all H

(X, Q) and H
BM

(X, Q).
(d) If j

: U

U and j : U X are open inclusions, then ( j j

)
!
= j
!
j
!
.
(e) If f : X Y is proper and j : U X is an open inclusion, then the diagram
H
BM

(X, Q)
j
!

H
BM

( f
1
(U), Q)
f

H
BM

(Y, Q)
j
!

H
BM

(U, Q)
commutes, where f

= f [
f
1
(U)
and j

: f
1
(U) X is inclusion.
Appendix: Generalized Gysin Maps 673
Fundamental Classes and Rened Cohomology Classes. A variety X of dimension n has
a canonically dened homology class [X] H
BM
2n
(X, Q) called the fundamental class of X.
Furthermore, if X is irreducible, then
H
BM
2n
(X, Q) =Q[X].
Example 13.A.3. Let X be a complete variety of dimension n. Then the constant map
p : X pt is proper. Hence we can dene

X
: H

(X, Q) Q such that the diagram


H

(X, Q)
[X]

R
X

H
BM

(X, Q)
p

H
BM

(pt, Q) =Q
commutes, i.e.,

X
= p

([X]). We use this map frequently in Chapters 12 and 13.


When Y X is a closed subset, the cap product in Proposition 13.A.1 generalizes to
H
i
(X, X `Y, Q)
Q
H
BM
j
(X, Q) H
BM
ji
(Y, Q), .
Applied to the fundamental class [X], this gives a map
H
i
(X, X `Y, Q) H
BM
ji
(Y, Q), [X].
When X is irreducible and rationally smooth, we get some classical duality theorems.
Proposition 13.A.4 (Duality). If X is irreducible and rationally smooth of dimension n
and Y X is closed, then we have isomorphisms
(a) Poincar e Duality: [X] : H
i
(X, Q) H
BM
2ni
(X, Q).
(b) Alexander Duality: [X] : H
i
(X, X `Y, Q) H
BM
2ni
(Y, Q).
This proposition explains why we are using coefcients in Q. If we want Poincar e and
Alexander duality to hold over Z, then we need to assume that X is smooth.
For X as in Proposition 13.A.4, let i : Y X be a d-dimensional irreducible subvariety.
Then the fundamental class [Y] H
BM
2d
(Y, Q) gives the following classes:
The rened cohomology class [Y]
r
H
2n2d
(X, X `Y, Q) maps to [Y] H
BM
2d
(Y, Q)
under Alexander duality.
The cohomology class [Y] H
2n2d
(X, Q) maps to i

[Y] H
BM
2d
(X, Q) under Poincar e
duality.
The natural map H
2n2d
(X, X `Y, Q) H
2n2d
(X, Q) takes [Y]
r
to [Y]. We use rened
cohomology classes in 12.4.
We will also need the following property of fundamental classes.
Proposition 13.A.5. If f : X Y is a proper birational morphism between irreducible
varieties, then f

[X] = [Y].
Generalized Gysin Maps. Let f : X Y be proper map such that Y is irreducible and
rationally smooth. Besides the usual contravariant map f

: H

(Y, Q) H

(X, Q), there


674 Chapter 13. Toric Hirzebruch-Riemann-Roch
is also a unique covariant map f
!
: H

(X, Q) H

(Y, Q) such that the diagram


H

(X, Q)
f
!

[X]

(Y, Q)
[Y]

H
BM

(X, Q)
f

H
BM

(Y, Q)
commutes. This follows since the vertical map on the right is an isomorphism. We call f
!
a generalized Gysin map. These maps behave nicely as follows.
Proposition 13.A.6. Assume that f : X Y is proper and Y is irreducible and rationally
smooth. Then:
(a) f
!
: H
k
(X, Q) H
k+2dim Y2dim X
(Y, Q).
(b) If g : Y Z is proper and Z is irreducible and rationally smooth, then (g f )
!
=g
!
f
!
.
(c) f
!
: H

(X, Q) H

(Y, Q) satises
f
!
( f

() ) = f
!
()
for all H

(Y, Q) and H

(X, Q).
In particular, taking = 1 in part (c) of the proposition gives f
!
( f

()) = f
!
(1).
Here are two cases where f
!
(1) is explicitly known for f as in Proposition 13.A.6:
If f : X Y is proper and birational, then f
!
(1) = 1. Thus f
!
( f

()) = for all


H

(Y, Q).
If i : Y X is the inclusion of an irreducible subvariety, then i
!
(1) = [Y]. Thus
i
!
(i

()) = [Y] for all H

(X, Q).
In the second bullet, the map i
!
: H

(Y, Q) H

(X, Q) is called the Gysin map.


Here is an example of a generalized Gysin map.
Example 13.A.7. Let X be complete, irreducible and rationally smooth of dimension n. If
p : X pt is the constant map, then p
!
: H

(X, Q) H

(pt, Q) Q is the map

X
from Example 13.A.3. This follows easily from the denitions of p
!
and

X
. Note also that
this agrees with the ad-hoc denition of

X
given in 12.4.
We will need the following compatibility result between pullbacks and generalized
Gysin maps.
Proposition 13.A.8. Suppose we have a commutative diagram of maps
(13.A.1)
X

X
f

Y
where f is proper, Y and Y

are irreducible and rationally smooth, and X

=X
Y
Y

. Then
we have a commutative diagram in cohomology
H

(X

, Q)
f

(Y, Q)
H

(X, Q)
f
!

(Y, Q).
g

Appendix: Generalized Gysin Maps 675


The diagram (13.A.1) is called Cartesian when X

= X
Y
Y

. It is a standard fact that


when (13.A.1) is Cartesian, f

is proper whenever f is.


The complicated behavior of Borel-Moore homology and Gysin maps were one of the
factors that inspired Fulton and MacPherson to develop the bivariant theories introduced
in [110]. This is related to the bivariant intersection theory discussed in [107, Ch. 17].
Equivariant Gysin Maps. Let T be a torus acting on varieties X and Y and let f : X Y
be T-equivariant and proper. Our goal is to dene an equivariant Gysin map
(13.A.2) f
!
: H

T
(X, Q) H

T
(Y, Q).
For simplicity, we will assume that Y is a simplicial toric variety where T acts on Y via a
homomorphismT T
N
= the torus of Y.
To dene f
!
, we follow [203, App.]. Recall that H

T
(Y, Q) = H

(ET
T
Y, Q), where
T (C

)
m
implies EG(C

`0)
m
. This follows from (12.4.6), where we showed that
EC

=C

` 0. Then dene the nite-dimensional approximation


ET

(C
+1
` 0)
m
of ET. Taking the quotient by T, we get the approximation BT

(P

)
m
of BT.
When Y is a simplicial toric variety, Y

=ET

T
Y is irreducible and rationally smooth
for all , so that the induced proper map f

: X

has a generalized Gysin map


f
!
: H

(X

, Q) H

(Y

, Q).
Furthermore, the inclusions X

X
+1
and Y

Y
+1
give a Cartesian diagram
X


_


_

X
+1
f
+1

Y
+1
It follows from Proposition 13.A.8 that the maps f
!
are compatible as . Hence we
get the desired map (13.A.2).
Here are the basic properties of equivariant Gysin maps.
Proposition 13.A.9. Assume that f : X Y is proper and T-equivariant, and assume that
Y is a simplicial toric variety. Then:
(a) f
!
: H
k
T
(X, Q) H
k+2dim Y2dim X
T
(Y, Q).
(b) If g : Y Z is proper and T-equivariant and Z is a simplicial toric variety, then
(g f )
!
= g
!
f
!
.
(c) f
!
: H

T
(X, Q) H

T
(Y, Q) satises
f
!
( f

() ) = f
!
()
for all H

T
(Y, Q) and H

T
(X, Q).
Equivariant Gysin maps also work nicely in Cartesian squares.
676 Chapter 13. Toric Hirzebruch-Riemann-Roch
Proposition 13.A.10. Suppose we have a commutative diagram of maps
X

X
f

Y
such that f is proper, f and g are equivariant, Y and Y

are simplicial toric varieties, and


X

= X
Y
Y

. Then we have a commutative diagram in equivariant cohomology


H

T
(X

, Q)
f

T
(Y

, Q)
H

T
(X, Q)
f
!

T
(Y, Q).
g

For the nal property we need, let X =X

be a complete simplicial toric variety. Then


the constant map p : X pt is proper and equivariant under the action of T =T
N
. Hence
we get

X
eq
= p
!
: H

T
(X, Q) H

T
(pt, Q).
This is the equivariant integral. We need to show that

X
eq is compatible with the ordinary
integral

X
= p
!
: H

(X, Q) H

(pt, Q) =Q.
Here is the precise result we will use in 13.4.
Proposition 13.A.11. In the above situation, we have a commutative diagram
H

(X, Q)
R
X

Q
H

T
(X, Q)
i

R
X
eq

T
(pt, Q).
i

pt

Proof. We will prove this using the commutative diagram


X
p


_

pt

_

BT

,
where X is a complete simplicial toric variety. Here, p

: X

= ET

T
X BT

is induced
by projection on the rst factor and pt is any point of BT

. This is a Cartesian diagram, so


that by Proposition 13.A.8, we get a commutative diagram
H

(X, Q)
p
!

Q
H

(X

, Q)
i

p
!

(BT

, Q).
i

pt

where i
X
: X X

and i
pt
: pt BT

are the inclusions. Letting , we get the


desired commutative diagram.
Chapter 14
Toric GIT and the
Secondary Fan
This chapter will explore a rich collection of ideas that give different ways to think
about toric varieties. We begin in 14.1 with the geometric invariant theory of a
closed subgroup G (C

)
r
acting on C
r
, which uses a character of G to lift the
G-action to a trivial line bundle over C
r
. In 14.2 we show that the GIT quotient
is a semiprojective toric variety in this situation, and in 14.3 we use Gale duality
to help us understand how the quotient depends on the character . The full story
of what happens as varies is controlled by the secondary fan, which is the main
topic of 14.4. The geometry of the secondary fan will be explored in Chapter 15.
14.1. Introduction to Toric GIT
Geometric invariant theory was invented by Mumford [209] in 1965 to prove the
existence of suitable projective compactications of the moduli spaces he was
studying. GIT, as it is now called, is a powerful tool in modern algebraic geometry.
See [83] for a nice introduction to the subject.
We will consider the special case of a closed subgroup G (C

)
r
acting on
C
r
= Spec(C[x
1
, . . . , x
r
]). Thus G (C

H, where H is nite. Since G is


reductive, Proposition 5.0.9 gives the good categorical quotient
C
r
//G = Spec(C[x
1
, . . . , x
r
]
G
).
However, such quotients can behave badly, as shown by the diagonal action of C

on C
r
. The only invariants are the constant polynomials, so that
C
r
//C

= Spec(C[x
1
, . . . , x
r
]
C

) = Spec(C) =pt.
677
678 Chapter 14. Toric GIT and the Secondary Fan
A better model for the kind of quotient we want comes from the quotient con-
struction of a toric variety X

without torus factors. Here, Theorem 5.1.11 gives


the almost geometric quotient
X

(C
(1)
`Z())//G.
Thus, if we take C
(1)
and throw away some points, we get an almost geometric
quotient, which by denition means that removing further points gives a geometric
quotient. As we will see, this is similar to what happens in GIT, where (roughly
speaking) we rst remove points that leave us with the set of semistable points, and
then we remove even more points to get the set of stable points.
Linearized Line Bundles. In GIT, deciding which points to remove is governed
by a lifting of the G-action on C
r
to the rank 1 trivial vector bundle C
r
C C
r
.
Liftings are described by characters of G. Let the character group of G be

G = : GC

[ is a homomorphism of algebraic groups.


Then a character

G gives the action of G on C
r
C dened by
g (p, t) = (g p, (g)t), g G, (p, t) C
r
C.
This lifts the G-action on C
r
, and all possible liftings arise this way.
Let L

denote the sheaf of sections of C


r
C with this G-action. We call L

the linearized line bundle with character . Note that for d Z, the tensor product
L
d

is the linearized line bundle with character


d
, i.e., L
d

=L

d .
If we forget the G-action, then L

O
C
r
as a line bundle on C
r
. It follows
that a global section s (C
r
, L

) can be written
s(p) = (p, F(p)) C
r
C, p C
r
,
for a unique F C[x
1
, . . . , x
r
]. The group G acts on global sections as follows.
Lemma 14.1.1. Let L

be the linearized line bundle on C


r
with character

G.
(a) If s is the global section of L

given by F, then for any g G, g s is the global


section dened by
(g s)(p) = (p, (g)F(g
1
p)), p C
r
.
(b) The G-invariant global sections are described by the isomorphism
(C
r
, L

)
G
F C[x
1
, . . . , x
r
] [ F(g p) = (g)F(p) for g G, p C
r
.
Proof. By Exercise 14.1.1, g G acts on a global section s by
(g s)(p) = g (s(g
1
p)), p C
r
.
Since s(p) = (p, F(p)), part (a) follows immediately, and then we are done since
part (b) is an easy consequence of part (a).
Denition 14.1.2. The polynomials in part (b) of the lemma are (G, )-invariant.
14.1. Introduction to Toric GIT 679
Example 14.1.3. Consider the usual action of C

on C
r
. Then

C

Z, where
d Z gives the character (t) = t
d
for t C

. One checks that (C


r
, L

)
G
is
isomorphic to the vector space of homogeneous polynomials of degree d.
Example 14.1.3 has a nice toric generalization.
Example 14.1.4. Let X

be a toric variety with no torus factors. The class group


Cl(X

) gives the algebraic group G = Hom


Z
(Cl(X

), C

), and using the map


Z
(1)
Cl(X

) from 4.1, we can regard G as a subgroup of C


(1)
.
The total coordinate ring of C
(1)
is S =C[x

[ (1)], which is graded by


Cl(X

) as in 5.2. Note also that



G Cl(X

), where a divisor class Cl(X

)
gives the character

G dened by evaluation at . In Exercise 14.1.2 you will
construct an isomorphism
(C
(1)
, L

)
G
S

=F S [ deg(F) = .
Thus S

is the set of all (G, )-invariant polynomials.


Here is an example that will appear several times in this section and the next.
Example 14.1.5. Let G = (t, t
1
, u) (C

)
3
[ t C

, u = 1 C

2
. One
easily sees that
C[x, y, z]
G
=C[xy, z
2
],
so that C
3
//G C
2
.
Now consider the character

G Z(Z/2Z) dened by (t, t
1
, u) = tu.
Then a monomial x
a
y
b
z
c
is (G, )-invariant if and only if
(tx)
a
(t
1
y)
b
(uz)
c
= tux
a
y
b
z
c
t
ab
u
c
= tu a = b+1, c 1 mod 2.
It follows easily that (C
3
, L

)
G
xz C[xy, z
2
].
Semistable and Stable Points. Given a global section s of L

, note that
(C
r
)
s
=p C
r
[ s(p) = 0
is an afne open subset of C
r
since s(p) =(p, F(p)) and s(p) =0 means F(p) =0.
Also observe that G acts on (C
r
)
s
when s is G-invariant.
Denition 14.1.6. Fix G (C

)
r
and

G, with linearized line bundle L

.
(a) p C
r
is semistable if there are d >0 and s (C
r
, L

d )
G
such that p (C
r
)
s
.
(b) p C
r
is stable if there are d > 0 and s (C
r
, L

d )
G
such that p (C
r
)
s
,
the isotropy subgroup G
p
is nite, and all G-orbits in (C
r
)
s
are closed in (C
r
)
s
.
(c) The set of all semistable (resp. stable) points is denoted (C
r
)
ss

(resp. (C
r
)
s

).
Since a global section s (C
r
, L

d )
G
corresponds to a (G,
d
)-invariant
polynomial F, one can determine semistability and stability using (G,
d
)-invariant
polynomials. We will do this frequently in the remainder of the chapter.
680 Chapter 14. Toric GIT and the Secondary Fan
In general treatments of GIT, where Gacts on a variety X, one must also require
that the nonvanishing subset of the section s be afne in the denition of semistable
and stable point. This is automatic in our case. See [83, Ch. 8] and [209].
Example 14.1.7. Consider the usual action of C

on C
r
and let
d
, d Z, be as in
Example 14.1.3. Then one can check without difculty that
d > 0 :(C
r
)
ss

d
= (C
r
)
s

d
=C
r
`0
d = 0 :(C
r
)
ss

d
=C
r
, (C
r
)
s

d
=
d < 0 :(C
r
)
ss

d
= (C
r
)
s

d
=.
This shows that the notions of semistable and stable depend strongly on which
character we use.
Example 14.1.8. For G = (t, t
1
, u) [ t C

, u = 1 and (t, t
1
, u) = tu from
Example 14.1.5, the (G,
d
)-invariant polynomials are
(14.1.1) (C
3
, L

d )
G

x
d
z C[xy, z
2
] d odd
x
d
C[xy, z
2
] d even.
In particular, x
2
is (G,
2
)-invariant, which implies that all points in C

C
2
are
semistable. With more work (Exercise 14.1.3), one can show that
(C
3
)
s

= (C
3
)
ss

=C

C
2
.
For a toric variety X

, the group G (C

)
(1)
acts on C
(1)
as described in
Example 14.1.4, and characters correspond to divisor classes since

G Cl(X

).
For the character of an ample class, semistable and stable have a nice meaning.
Proposition 14.1.9. Let X

be a projective toric variety, and let

be the
subfan consisting of all simplicial cones of . If

G comes from an ample
divisor class Cl(X

), then

C
(1)

ss

=C
(1)
`Z()

C
(1)

=C
(1)
`Z(

).
Proof. Let = [D], where D =

is ample. Then the polytope P


D
M
R
has facet presentation
P
D
=m M
R
[ 'm, u

` a

for all (1),


and the vertices of P
D
give the Cartier data m

(n)
of D. Also recall from 5.4
that if S =C[x

[ (1)], then the graded piece S

is spanned by the monomials


x
/m,D)
=

x
/m,u)+a

, m P
D
M.
Now take p C
(1)
` Z(). Since Z() is dened by the vanishing of x

=

/ (1)
x

, (n), there must be (n) such that x



does not vanish at p.
14.1. Introduction to Toric GIT 681
We claim that x
/m,D)
S

does not vanish at p. Since x


/m,D)
is (G, )-invariant
by Example 14.1.4, this will imply that p is semistable. To prove our claim, recall
from the discussion before Example 5.4.5 that the exponent of x

in x
/m,D)
is the
lattice distance from m to the facet of P
D
whose normal is u

. For m

, the lattice
distance is zero when (1) since these s give the facets containing the vertex
m

, and all other lattice distances are positive. Thus x


/m,D)
and x

involve the
same variables, so that x
/m,D)
does not vanish at p since the same is true for x

.
Next assume p C
(1)
is semistable. Then there is d > 0 such that some
element of S
d
does not vanish at p. It follows that x
/m,dD)
is nonzero at p for some
m (dP
D
) M. Let Q_dP
D
be the smallest face of dP
D
containing m and let dm

be a vertex of this face. We claim that x



divides x
/m,dD)
. Since x
/m,dD)
does not
vanish at p, this will imply that the same is true for x

, and p C
(1)
`Z() will
follow. To prove our claim, take / (1) and consider the facet F _dP
D
with facet
normal u

. If m F, then Q _ F, which would give dm

F. This contradicts
/ (1) by the denition of normal fan, so that m has positive lattice distance to
the facets corresponding to / (1). It follows that x

divides x
/m,dD)
.
It remains to consider stable points. Since

(1) = (1) (rays are simplicial),


X

and X

have the same class group and the same group G. The difference is that

implies an inclusion of irrelevant ideals B(

) B(), which in turn gives


the inclusion
C
(1)
`Z(

) C
(1)
`Z().
Since

is simplicial, Theorem 5.1.11 implies that X

(C
(1)
`Z(

))//G is a
geometric quotient. It follows that G p is closed for all p C
(1)
` Z(

). Fur-
thermore, Exercise 5.1.11 implies that the isotropy subgroup G
p
is nite for these
ps since

is simplicial. Thus p is stable, hence C


(1)
`Z(

C
(1)

.
For the opposite inclusion, let U =

C
(1)

be the set of stable points. Using


the trivial character, the action of (C

)
(1)
lifts to an action on L

that is easily
seen to commute with the action of G on L

. This induces an action of (C

)
(1)
on (C
(1)
, L

d )
G
for every d N. It follows that if p C
(1)
is stable, so is
its (C

)
(1)
-orbit. In other words, (C

)
(1)
acts on U. Note that we also have
U C
(1)
`Z() since stable points are semistable.
Now consider the quotient map
: C
(1)
`Z() X

= (C
(1)
`Z())//G.
Because G-orbits in U are closed, Theorem 5.0.6 and Proposition 5.0.7 imply that
(U) X

is open. By the previous paragraph, (U) is stable under the action of


T
N
= (C

)
(1)
/G and hence is a toric variety. This means it comes from a subfan
of . Note also that [
U
: U (U) is a geometric quotient since G-orbits are
closed in U. By Theorem 5.1.11, this subfan is simplicial and hence is contained
in

. It follows easily that U C


(1)
`Z(

), and we are done.


682 Chapter 14. Toric GIT and the Secondary Fan
GIT Quotients. Given G (C

)
r
and

G, our next task is to dene the GIT
quotient C
r
//

G. The basic idea is to take the quotient of



C
(1)

ss
under the
action of G. As shown by the non-separated quotient from Example 5.0.15, care
must be taken to ensure that the quotient is well-behaved.
The strategy will be to use the graded ring
(14.1.2) R

d=0
(C
r
, L

d )
G
.
We rst study the structure of this ring.
Lemma 14.1.10. The graded ring R

is a nitely generated C-algebra.


Proof. Consider the action of G on C
r
C given by g (p, t) = (g p,
1
(g)t).
Then G acts on f C[x
1
, . . . , x
r
, w] by
(g f )(p, t) = f (g
1
(p, t)) = f (g
1
p,
1
(g
1
)t) = f (g
1
p, (g)t).
In particular, if F C[x
1
, . . . , x
r
], then f = F w
d
is G-invariant if and only if for all
(p, t) C
r
C and all g G, we have
f (g
1
p, (g)t) = f (p, t) F(g
1
p)((g)t)
d
= F(p)t
d
F(g
1
p) = ((g))
d
F(p) =
d
(g
1
)F(p).
Replacing g with g
1
, we see that f = F w
d
is G-invariant if and only if F is
(G,
d
)-invariant. Using Lemma 14.1.1, we obtain an isomorphism
(14.1.3) R

C[x
1
, . . . , x
r
, w]
G
.
Since G is reductive, R

is a nitely generated C-algebra by Proposition 5.0.9


We now dene the GIT quotient using the Proj construction described in 7.0.
Denition 14.1.11. For G (C

)
r
and

G, the GIT quotient C
r
//

G is
C
r
//

G = Proj(R

).
Here are some easy properties of GIT quotients.
Proposition 14.1.12. For G (C

)
r
and

G, we have:
(a) There is a projective morphism C
r
//

G C
r
//G = Spec(C[x
1
, . . . , x
r
]
G
).
(b) C
r
//

G = if and only if (C
r
)
ss

=.
(c) The GIT quotient C
r
//

G is a good categorical quotient of (C


r
)
ss

under the
action of G, i.e., C
r
//

G (C
r
)
ss

//G.
(d) If (C
r
)
s

=, then the action of G on (C


r
)
s

has a geometric quotient (C


r
)
s

/G
isomorphic to a nonempty open subset of C
r
//

G. Thus C
r
//

G (C
r
)
ss

//G
is an almost geometric quotient and dimC
r
//

G = r dim G.
14.1. Introduction to Toric GIT 683
Proof. The map Proj(R

) Spec((R

)
0
) is projective by Proposition 7.0.9, and
then part (a) follows since (R

)
0
=C[x
1
, . . . , x
r
]
G
. For part (b), rst note that
(C
r
)
ss

= (R

)
d
= 0 for d > 0.
The proof of Lemma 14.1.10 shows that R

is an integral domain, and then one


sees easily that Proj(R

) = if and only if (R

)
d
= 0 for d > 0 (Exercise 14.1.5).
For part (c), let R = C[x
1
, . . . , x
r
, w]
G
be the ring of invariants introduced in
Lemma 14.1.10, which we grade by degree in w. The isomorphism of graded
rings R

R from (14.1.3) implies that C


r
//

G Proj(R). Recall from 7.0 that


Proj(R) is covered by open subsets D
+
( f ) = Spec(R
( f )
), where f R
d
is nonzero
with d > 0 and
(14.1.4) R
( f )
=

h
f

[ h S
d
, 0

.
Now suppose that f = F w
d
, where F C[x
1
, . . . , x
r
] is (G,
d
)-invariant. Then
F corresponds to global section s (C
r
, L

d )
G
by Lemma 14.1.1. Furthermore,
the nonvanishing set (C
r
)
s
C
r
of s is
(C
r
)
s
= (C
r
)
F
= Spec(C[x
1
, . . . , x
r
]
F
),
where C[x
1
, . . . , x
r
]
F
is localization at F. In Exercise 14.1.6 you will show that
(14.1.5)
H
F


Hw
d
(F w
d
)

=
Hw
d
f

induces an isomorphism (C[x


1
, . . . , x
r
]
F
)
G
R
( f )
. This implies that
D
+
( f ) = Spec(R
( f )
) Spec((C[x
1
, . . . , x
r
]
F
)
G
) = (C
r
)
s
//G.
Then part (c) follows since Proj(R) is covered by the open subsets D
+
( f ) and
(C
r
)
ss

is covered by the afne open subsets (C


r
)
s
. See [83, Sec. 8.2] for the details.
The rst assertion of part (d) is a standard result in GIT (see [83, Thm. 8.1]).
Thus the quotient map : (C
r
)
ss

C
r
//

G of part (c) is an almost geometric


quotient as dened in 5.0. Finally, if p is a stable point, then
1
((p)) = G p
G/G
p
, which has dimension dim G since p is stable. Then the dimension formula
dim (C
r
)
ss

= dimC
r
//

G + dim generic ber of


(see [245, Thm. I.6.7]) shows that dimC
r
//

G = r dim G.
Remark 14.1.13. Part (d) of Proposition 14.1.12 has two useful consequences:
(a) If stable points exist, then C
r
//

G has the expected dimension r dim G. In


general, we always have dimC
r
//

G r dim G (Exercise 14.1.4).


(b) G-orbits of stable points are closed in (C
r
)
ss

(Exercise 14.1.4).
The properties of C
r
//

G stated in Proposition 14.1.12 apply to more general


GIT quotients, as explained in [83] and [209].
Here are two illustrations of Proposition 14.1.12.
684 Chapter 14. Toric GIT and the Secondary Fan
Example 14.1.14. For G = (t, t
1
, u) [ t C

, u = 1 and (t, t
1
u) = tu, we
have (C
3
)
ss
=C

C
2
by Example 14.1.8. Since this is afne, Proposition 14.1.12
implies that
C
3
//

G (C

C
2
)//G = Spec(C[x
1
, y, z]
G
).
The ring of invariants is easily seen to be C[xy, z
2
], so that
C
3
//

G Spec(C[xy, z
2
]) C
2
.
Working directly from C
3
//

G = Proj(R

) is harder since
R

=C[xy, z
2
] xz C[xy, z
2
] x
2
C[xy, z
2
] x
3
z C[xy, z
2
] x
4
C[xy, z
2
]
by (14.1.1). It is not obvious that Proj of this ring gives C
2
. We will see in 14.2
that the complications of this example come from a polyhedron whose vertices are
not lattice points.
Example 14.1.15. Consider G = (t, t, u, u) [ t, u C

and (t, t, u, u) = t. Then


a polynomial F(x, y, z, w) is (G,
d
)-invariant if and only if
F(tx, ty, uz, uw) = t
d
F(x, y, z, w),
i.e., if and only if F has degree (d, 0) in the grading where x, y have degree (1, 0)
and z, w have degree (0, 1). It follows that as graded rings,
R

C[x, y].
Thus C
4
//

G=Proj(C[x, y]) =P
1
. This is disconcerting since C
4
has dimension 4
and G has dimension 2, yet the quotient only has dimension 1. The reason is that
there are no stable points. In fact,
(C
4
)
ss

=C
2
(C
2
`0), (C
4
)
s

=.
Note that (C
4
)
s

= follows from part (d) Proposition 14.1.12 since the quotient


does not have the expected dimension.
In contrast, if one uses the character dened by (t, t, u, u) = t
a
u
b
for a, b > 0,
then C
4
//

G = P
1
P
1
and (C
4
)
s
= . Once we introduce the secondary fan in
14.4, this example will be easy to understand.
The GIT quotients in Examples 14.1.14 and 14.1.15 are toric varieties. This is
no accident, as we will prove in the next section.
More General Toric GIT Quotients. So far, we have studied the quotient of C
r
by a subgroup G (C

)
r
. An obvious generalization would be to consider the
quotient of a toric variety X

by a subgroup GT
N
. This question has been studied
extensively in the literature. For more details about these quotients, interested
readers should consult the papers [3] by ACampo-Neuen and Hausen, [149] by
Hu, and [164] by Kapranov, Sturmfels and Zelevinsky.
14.2. Toric GIT and Polyhedra 685
Exercises for 14.1.
14.1.1. Let : V X a vector bundle over a variety X. Suppose that an afne algebraic
group G acts algebraically on X and V such that is equivariant.
(a) Given a global section s : X V of , show that
(g s)(p) = g (s(g
1
p))
denes a global section of .
(b) Show carefully that g (h s) = (gh) s for g, h G.
14.1.2. Prove the isomorphism (C
r
, L

)
G
S

from Example 14.1.4.


14.1.3. Prove that (C
3
)
s

= (C
3
)
ss

=C

C
2
in Example 14.1.8.
14.1.4. This exercise concerns Remark 14.1.13.
(a) Prove that dimC
r
//

G r dim G. Hint: Use [245, Thm. I.6.7].


(b) Prove that G-orbits of stable points are closed in (C
r
)
ss

. Hint: Use part (d) of Proposi-


tion 14.1.12 and remember that bers are closed.
14.1.5. Let R =

d=0
R
d
be a graded integral domain. Prove that Proj(R) = if and only
if R
d
= 0 for all d > 0.
14.1.6. Prove that the map (14.1.5) in the proof of Proposition 14.1.12 induces a ring
isomorphism (C[x
1
, . . . , x
r
]
F
)
G
R
( f )
.
14.1.7. Prove that if C[x
1
, . . . , x
r
]
G
= C, then C
r
//

G is projective. The converse is more


subtle, as you will learn in the next two exercises.
14.1.8. Let G =(t, t, 1) [ t C

(C

)
3
and (t, t, 1) =t
1
.
(a) Show that (R

)
0
=C[x, y, z]
G
=C[z] and that (R

)
d
= 0 for d > 0.
(b) Conclude that the converse of Exercise 14.1.7 is false. Hint: is a projective variety.
14.1.9. The converse of Exercise 14.1.7 is false by Exercise 14.1.8. Fortunately, once we
assume that C
r
//

G =, everything is ne.
(a) Let U V be a morphism of irreducible varieties such that the map C[V] C[U] is
injective. Prove that the image of U is Zariski dense in V.
(b) Suppose (R

)
d
=0 for some d >0. Use part (a) to prove that Proj(R

) Spec((R

)
0
)
has Zariski dense image. Hint: Construct a map (R

)
0
(R

)
( f )
for f = 0 in (R

)
d
.
(c) Prove that if C
r
//

G is projective and nonempty, then C[x


1
, . . . , x
r
]
G
=C.
14.1.10. Prove that C
r
//

G Spec(C[x
1
, . . . , x
r
]
G
) when is a torsion element of

G.
14.2. Toric GIT and Polyhedra
We continue to assume that G (C

)
r
is an algebraic subgroup and

G is a
character. A full understanding of the GIT quotients introduced in 14.1 involves
the interesting polyhedra associated to characters.
686 Chapter 14. Toric GIT and the Secondary Fan
The Group G. Before discussing polyhedra, we need to study G. Since Z
r
is
the character group of (C

)
r
, the inclusion G (C

)
r
induces a homomorphism
: Z
r


G whose kernel we denote by M. Hence we have an exact sequence
(14.2.1) 0 M

Z
r


G.
The image of a Z
r
in

G will be written (a) =
a
. The dual of is a map
Z
r
N, where N is the dual of M. The images of the standard basis e
1
, . . . , e
r
Z
r
give elements
1
, . . . ,
r
N, and the map in (14.2.1) can be written
(m) = ('m,
1
`, . . . , 'm,
r
`), m M.
Maps like this appear in the exact sequence (4.1.3) for the class group of a toric
variety, where the u

, (1), play the role of the


i
. However, we will see in
Example 14.2.4 below that the
i
are not always as nice as the u

.
Here is a basic result about G and its character group.
Lemma 14.2.1. Let T
N
= N
Z
C

= Hom
Z
(M, C

).
(a) The map in (14.2.1) is surjective, so that we have an exact sequence
(14.2.2) 0 M

Z
r


G 0.
(b) (C

)
r
/G T
N
, so that we have an exact sequence of algebraic groups
1 G(C

)
r
T
N
1.
(c) G =(t
1
, . . . , t
r
) (C

)
r
[

r
i=1
t
/m,
i
)
i
= 1 for all m M.
Proof. We begin with part (b). Since G is reductive and G-orbits are closed in
(C

)
r
, we have a geometric quotient (C

)
r
/G = Spec(C[x
1
1
, . . . , x
1
r
]
G
). One
checks that Laurent monomial x
b
is G-invariant if and only (b)

G is the trivial
character. By (14.2.1), it follows that C[x
1
1
, . . . , x
1
r
]
G
= C[M]. Part (b) follows
immediately since T
N
= Spec(C[M]).
For part (c), consider (C

)
r
T
N
= Hom
Z
(M, C

). In Exercise 14.2.1 you


will show that this map takes (t
1
, . . . , t
r
) (C

)
r
to Hom
Z
(M, C

) dened by
(m) =

r
i=1
t
/m,
i
)
i
. Then we are done since G is the kernel of this map by part (b).
Finally, for part (a), let n be the rank of M, which is also the rank of . Then
pick bases of M and Z
r
such that has Smith normal form
=

d
1
0
.
.
.
.
.
. d
n
.
.
.
.
.
.
0 0

,
where d
1
[d
2
[ [d
n
are positive and all other entries are zero. Using the dual basis
e
1
, . . . , e
n
of N, this means that is given by

i
= d
i
e
i
for 1 i n and

i
= 0 for
14.2. Toric GIT and Polyhedra 687
i > n. Then applying part (c) gives a description of G which makes it obvious that
characters of G extend to characters of (C

)
r
(Exercise 14.2.1).
The Polyhedron of a Character. We will give two models of the polyhedron of


G. The rst model is less intrinsic but has a nice relation to the treatment of
polytopes and polyhedra given in Chapters 2 and 7.
For a character =
a


G, where a =(a
1
, . . . , a
r
) Z
r
, dene the polyhedron
(14.2.3) P
a
=m M
R
[ 'm,
i
` a
i
, 1 i r M
R
.
For xed , the various choices of a differ by elements of (M), so the polyhedra
P
a
are translates of each other by elements of M.
Here is an important property of these polyhedra.
Lemma 14.2.2. If P
a
=, then the recession cone of P
a
is strongly convex.
Proof. Assume that P
a
from (14.2.3) is nonempty. Then its recession cone is C =
m M
R
[ 'm,
i
` 0, 1 i r, so that C(C) is dened by 'm,
i
` = 0 for
1 i r. Thus C(C) =0 since the map in (14.2.2) is injective.
Recall from 7.1 that P M
R
is a lattice polyhedron when its recession cone
is strongly convex and its vertices are lattice points. Furthermore, when dim P =
dim M
R
, we constructed the toric variety X
P
with torus T
N
. Lemma 14.2.2 shows
that P
a
has the right kind of recession cone. Unfortunately, it may have the wrong
dimension and its vertices may fail to be lattice points. Here are some examples.
Example 14.2.3. Consider G = (t, t, u, u) [ t, u C

and (t, t, u, u) = t. Then


=
a
for a = (1, 0, 0, 0), and the polyhedron P
a
is the line segment
P
a
= Conv(0, e
1
) M
R
=R
2
(Exercise 14.2.2). Hence P
a
is not full dimensional.
Example 14.2.4. For the group G = (t, t
1
, u) [ t C

, u = 1 (C

)
3
, the
exact sequence (14.2.2) is
0 Z
2

Z
3

Z(Z/2Z) 0,
where (m) = ('m, e
1
`, 'm, e
1
`, 'm, 2e
2
`) and (a, b, c) = (a b, c mod 2). Thus

1
=
2
= e
1
and
3
= 2e
2
, so these
i
are neither distinct nor primitive. This
contrasts with toric case, where the
i
are the ray generators of a fan and hence are
distinct and primitive.
The character (t, t
1
u) =tu from Example 14.1.14 is =
a
for a =(1, 0, 1).
Then P
a
M
R
=R
2
is dened by the inequalities
'm, e
1
` 1, 'm, e
1
` 0, 'm, 2e
2
` 1.
Figure 1 on the next page shows P
a
. Note that the vertex of P
a
is not a lattice point.
We also see that the rst inequality dening P
a
is redundant.
688 Chapter 14. Toric GIT and the Secondary Fan
(0,1/2)
P
a
Figure 1. The polyhedron Pa R
2
for a =(1, 0, 1) in Example 14.2.4
Example 14.2.5. For the usual action of C

on C
r
, the exact sequence (14.2.2) is
0 Z
r1

Z
r

Z 0,
where is dened using
i
= e
i
, 1 i r 1 and
r
=e
1
e
r1
.
Let =
a
for a = (0, . . . , 0, d) Z
r
. Then:
P
a
=

d
r1
d > 0
0 d = 0
d < 0.
Theorem 14.2.13 below will explain how this relates to Example 14.1.7.
The polyhedron of

G has a second, more intrinsic model P

described as
follows. Tensoring (14.2.2) with R gives an exact sequence of vector spaces
(14.2.4) 0 M
R

R
R
r

R


G
R
0,
where

G
R
=

G
Z
R. Also let 1

G
R
be the image of under the map

G

G
R
.
Note that this map is not injective when

G has torsion, which happens precisely
when G is not a torus.
Then we get the polyhedron P

R
r
dened by
(14.2.5) P

=b R
r
0
[
R
(b) = 1 =
1
R
(1) R
r
0
.
It is straightforward to show that if =
a
for a Z
r
, then
P

=
R
(P
a
) +a.
In other words, if a = (a
1
, . . . , a
r
), then
(14.2.6) P

=('m,
1
` +a
1
, . . . 'm,
r
` +a
r
) [ m P
a

(Exercise 14.2.3). The key point is that the inequalities (14.2.3) dening P
a
are
equivalent to saying that ('m,
1
` +a
1
, . . . 'm,
r
` +a
r
) lies in R
r
0
.
Via (14.2.6), lattice points of P
a
correspond to the points

R
(P
a
M) +a =
1
() N
r

1
R
(1) N
r
= P

Z
r
.
14.2. Toric GIT and Polyhedra 689
One subtle feature of P

is that the inclusion


R
(P
a
M) +a P

Z
r
is strict
when

G has torsion. Here is an example.
Example 14.2.6. For the polyhedron P
a
shown in Figure 1 of Example 14.2.4, we
have a = (1, 0, 1) and
1
=
2
= e
1
,
3
= 2e
2
. Thus
P

=('m, e
1
` +1, 'm, e
1
`, 'm, 2e
2
` +1) [ m P
a
.
Figure 2 shows P
a
and P

. The vertex (1, 0, 0) P

Z
3
does not come from a
lattice point in P
a
M.
(0,1/2)
P
a
(1,0,0)
P

Figure 2. The polyhedra Pa and P


In the literature, this difculty is often avoided by assuming that G is a torus,
so that

G is torsion-free (see, for example, [137]).
Our goal is to prove that C
r
//

G is a toric variety. The rough intuition is that


C
r
//

G is the toric variety of P


a
M
R
. However, we cant apply the theory of
Chapter 7 directly since P
a
may have the wrong kind of vertices and the wrong
dimension. Our next task is to develop some tools to address these problems.
Veronese Subrings. The polyhedron P
a
M
R
need not be a lattice polyhedron
since its vertices may not be lattice points. However, the vertices are rational, so
some integer multiple will be a lattice polyhedron. We will see that when translated
into algebra, this gives a Veronese subring of a graded semigroup algebra.
Before dening Veronese subrings in general, we recall a classic example that
explains where the name comes from.
Example 14.2.7. For a positive integer , the line bundle O
P
n () on P
n
gives the
th Veronese embedding P
n
X P
(
n+
n
)1
. If we let C[X] be the homogeneous
coordinate ring of X as a subvariety of P
(
n+
n
)1
, then
P
n
X Proj(C[X]).
However, the total coordinate ring of P
n
as a toric variety is S =C[x
0
, . . . , x
n
], and
the th Veronese embedding is determined by the polytope P =
n
, where
n
is
690 Chapter 14. Toric GIT and the Secondary Fan
the standard n-simplex. Since P is normal, Theorem 5.4.8 implies that
C[X] =C[X
P
]

d=0
S
d

d=0
S
d
= S.
This is the th Veronese subring of S. When we combine the above two displays,
we see that P
n
is the Proj of not only S but also of its Veronese subrings.
More generally, let R =

d=0
R
d
be a graded ring that is nitely generated as
a C-algebra. Given a positive integer , the th Veronese subring of R is
R
[]
=

d=0
R
d

d=0
R
d
= R,
so (R
[]
)
d
= R
d
. Then Example 14.2.7 generalizes as follows (Exercise 14.2.4).
Lemma 14.2.8. Given R and a Veronese subring R
[]
as above, there is a natural
isomorphism Proj(R
[]
) Proj(R).
Here is an example of how to use Lemma 14.2.8.
Example 14.2.9. The coordinate ring A of an afne variety gives the graded ring
A[x
0
, . . . , x
n
], where all variables have degree 1. In Example 7.0.10, we noted that
Proj(A[x
0
, . . . , x
n
]) = Spec(A) P
n
.
In particular, when n = 0, we have
Proj(A[x
0
]) = Spec(A) P
0
= Spec(A).
Now consider the complicated graded ring R

from Example 14.1.14. The


Veronese subring
R
[2]

=C[xy, z
2
] x
2
C[xy, z
2
] x
4
C[xy, z
2
] x
6
C[xy, z
2
]
is a polynomial ring in one variable over C[xy, z
2
], i.e., R
[2]

=C[xy, z
2
][x
2
]. Thus
C
3
//

G = Proj(R

) Proj(R
[2]

) = Spec(C[xy, z
2
]) C
2
.
We now understand this example from the Proj point of view.
The Normal Fan of a Polyhedron. Previously, we have given three constructions
of the normal fan:
From 2.3: The normal fan of a full dimensional lattice polytope P M
R
is a
complete fan in N
R
.
From 6.2: The normal fan of a lattice polytope P M
R
is a complete gener-
alized fan in N
R
. This fan is degenerate when P is not full dimensional.
From 7.1: The normal fan of a full dimensional lattice polyhedron P M
R
is
a fan in N
R
whose support is the dual of the recession cone of P.
14.2. Toric GIT and Polyhedra 691
It can happen that none of these constructions apply to the polyhedra P
a
M
R
considered here. Hence we need a fourth construction of a normal fan. You may
want to review the discussion of generalized fans given in 6.2.
Let P M
R
be a polyhedron with rational vertices such that the recession cone
of P is strongly convex and rational. A vertex v P gives the cone
C
v
= Cone(PM
Q
v) M
R
.
This is similar to our earlier constructions, except that we use PM
Q
instead of
PM since P may have few lattice points (in fact, it could have none). The dual
cone C

v
N
R
is a rational polyhedral cone, and these cones give a generalized fan
as follows (Exercise 14.2.5).
Proposition 14.2.10. Let C M
R
be the recession cone of P. Then

P
= [ _C

v
, v is a vertex of P
is a generalized fan in N
R
called the normal fan of P. Furthermore, [
P
[ =C

,
and
P
is a fan if and only if P M
R
is full dimensional.
The toric variety X
P
is then dened to be the toric variety of the generalized
fan X

P
, i.e., X
P
= X

P
. Here is an example we studied in 6.2.
Example 14.2.11. Let D be a basepoint free torus-invariant Cartier divisor on a
complete toric variety X

. From D we get P
D
M
R
, which is a lattice polytope by
Theorem 6.1.7, though it may fail to be full dimensional. The normal fan
P
D
is a
generalized fan that gives the toric variety X
P
D
featured in Theorem 6.2.8.
We note that generalized fans appear naturally in this situation. The vertices of
P
D
give the Cartier data m

(n)
of D, where n = dim X

. The m

are distinct
when D is ample, but some may coincide in the basepoint free case. Combining
those s for which m

has a common value, say v P


D
, gives the union

v
=

(n)
m=v

By Proposition 6.2.5, this union is a maximal cone in the normal fan of P


D
in N
R
.
You should reread the discussion following the proof of Proposition 6.2.5.
We will soon see that X
P
D
has a natural interpretation as a GIT quotient.
Similar to 7.1, a polyhedron PM
R
gives a cone C(P) M
R
Rwhose slice
at height > 0 (resp. = 0) is P (resp. the recession cone of P). The associated
semigroup algebra
S
P
=C[C(P) (MZ)]
is graded as usual using the last coordinate. We now generalize Theorem 7.1.13.
Proposition 14.2.12. For a polyhedron P M
R
as above, we have X
P
Proj(S
P
).
Furthermore, X
P
is semiprojective.
692 Chapter 14. Toric GIT and the Secondary Fan
Proof. The polyhedron P gives the (possibly degenerate) normal fan
P
. First note
that multiplying P by a positive integer has no effect on the normal fan
P
. Also,
Lemma 14.2.8 and the obvious isomorphism S
[]
P
S
P
imply that
Proj(S
P
) Proj(S
[]
P
) Proj(S
P
).
Hence we may assume that P is a lattice polyhedron. Note also that translating P
by m M has no effect on
P
. Since S
P+m
S
P
, we may assume that 0 P.
The minimal cone
0

P
is the largest subspace of N
R
contained in every
cone of the normal fan. As we saw in 6.2,
P
projects to a genuine fan
P
in
N
R
= (N/
0
)
R
, and X
P
is the toric variety of this fan. Since 0 P, it follows that:

0
is the smallest subspace of M
R
containing P.
M =

0
M M is dual to the map N N.
P is a full dimensional lattice polytope in M
R
whose normal fan in N
R
is given
by
P
.
The straightforward proofs are left to the reader as Exercise 14.2.6. The last bullet
and Theorem 7.1.13 imply that
X
P
Proj(C[C(P) (MZ)]).
Then the rst bullet implies
C(P) (MZ) =C(P) (MZ),
which gives X
P
Proj(S
P
).
Finally, we show that X
P
is semiprojective. As above, we may assume that
P M
R
is a lattice polyhedron containing the origin. Then P is a full dimensional
lattice polyhedron in M
R
, so X
P
is semiprojective by Proposition 7.2.9.
GIT Quotients are Toric. We can now prove that C
r
//

G is a toric variety.
Theorem 14.2.13. For G(C

)
r
and

G, pick a Z
r
such that =
a
. Then:
(a) The graded ring R

from (14.1.2) satises R

C[C(P
a
) (MZ)].
(b) The GIT quotient C
r
//

G = Proj(R

) is the toric variety of P


a
M
R
.
Proof. We rst note that R

is the semigroup algebra of the semigroup given by


the disjoint union
S

d=0

1
(
d
) N
r
,
where is from (14.2.2). To see why, note that x
b
C[x
1
, . . . , x
r
] is (G,
d
)-
invariant if and only if (b) =
d
. Then (14.1.3) gives an isomorphism
R

d=0
F C[x
1
, . . . , x
r
] [ Fis (G,
d
)-invariant =C[S

].
Lemma 14.2.2 implies that P
a
M
R
has a strongly convex rational recession
cone, and the description of P
a
given in (14.2.3) shows that the vertices are rational.
14.2. Toric GIT and Polyhedra 693
Hence Propositions 14.2.10 and 14.2.12 apply to P
a
. In particular, we get the cone
C(P
a
) M
R
R, and (m, k) (m) +ka induces a semigroup isomorphism
(14.2.7) C(P
a
) (MZ) S

(Exercise 14.2.7). It follows that R

C[C(P
a
) (MZ)]. Thus
C
r
//

G = Proj(R

) Proj(C[C(P
a
) (MZ)]) X
Pa
,
where the last isomorphism follows from Proposition 14.2.12.
In [204, Def. 10.8], Miller and Sturmfels dene a toric variety to be a GIT
quotient of C
r
by G. They use the notation C
r
//
a
G, which for us means C
r
//

a G.
Versions of Theorem 14.2.13 appear in [83, Ch. 12], [204, Ch. 10] and [232].
In Theorem 14.2.13, we used the polyhedron P
a
rather than the more intrinsic
model P

. The reason goes back to Example 14.2.4, where we learned that lattice
points in P

R
r
relative to Z
r
need not come from lattice points in P
a
. When

G
has no torsion (i.e., when G is a torus), this problem goes away and C
r
//

G is the
toric variety of the polyhedron P

R
r
.
Here are some examples that illustrate Theorem 14.2.13.
Example 14.2.14. Let X

be a projective toric variety with quotient construction


X

= (C

(1) `Z())/G. Then a divisor class [D] = [

] Cl(X

) gives a
character

G. Note also that =
a
for a = (a

) Z
(1)
. Furthermore,
P
a
= P
D
M
R
,
where P
D
is from (4.3.2). There are two cases where we know the GIT quotient:
When D is ample, is the normal fan of P
D
= P
a
. Hence
C
r
//

GX

by Theorem 14.2.13. This also follows from Propositions 14.1.9 and 14.1.12.
When D is nef, Theorem 14.2.13 implies that
C
r
//

G X
P
D
,
where X
P
D
is the toric variety from Example 14.2.11 and Theorem 6.2.8.
We will pursue this example in Chapter 15 when we study the secondary fan.
Example 14.2.15. The afne space Sym
3
(C) of 33 symmetric matrices contains
the torus G (C

)
3
consisting of the rank 1 matrices

t
1
t
2
t
3

t
1
t
2
t
3

t
2
1
t
1
t
2
t
1
t
3
t
1
t
2
t
2
2
t
2
t
3
t
1
t
3
t
2
t
3
t
2
3

, t
1
, t
2
, t
3
C

.
694 Chapter 14. Toric GIT and the Secondary Fan
We will compute Sym
3
(C)//

G for the character that sends the above matrix to


t
2
1
t
2
2
t
2
3
. Write Sym
3
(C) = Spec(C[x
11
, x
12
, x
13
, x
22
, x
23
, x
33
]) C
6
and note that
det

x
11
x
12
x
13
x
12
x
22
x
23
x
13
x
23
x
33

= x
11
x
22
x
33
+2x
12
x
23
x
13
x
11
x
2
23
x
22
x
2
13
x
33
x
2
12
.
The monomials appearing in the determinant form a basis of the (G, )-invariant
polynomials. This explains our choice . Label these monomials as
X
0
= x
12
x
23
x
13
, X
1
= x
11
x
2
23
, X
2
= x
22
x
2
13
, X
3
= x
33
x
2
12
, X
4
= x
11
x
22
x
33
and note that X
2
0
X
4
= X
1
X
2
X
3
. Using coordinates Y
0
, . . . ,Y
4
on P
4
, we claim that
(14.2.8) Sym
3
(C)//

G V(Y
2
0
Y
4
Y
1
Y
2
Y
3
) P
4
.
We will sketch the proof, leaving the details as Exercise 14.2.8. The map
Z
6


G Z
3
is given by e
i j
e
i
+e
j
, where e
11
, e
12
, e
13
, e
22
, e
23
, e
33
is the basis
of Z
6
. Also observe that corresponds to (2, 2, 2) Z
3
, so that =
a
for a =
(0, 1, 1, 0, 1, 0) Z
6
. In Example 14.2.18 below we will see that the
i
s are

1
= (1, 0, 0),
2
= (1, 1, 1),
3
= (1, 1, 1),

4
= (0, 1, 0),
5
= (1, 1, 1),
6
= (0, 0, 1).
This gives the polytope P
a
in Figure 3. The only lattice points of P
a
are its vertices
m
0
, . . . , m
4
, which map to (m
0
) +a, . . . , (m
4
) +a, the exponent vectors of the
(G, )-invariant monomials X
0
, . . . , X
4
. Furthermore, since P
a
is a normal polytope,
the monomials X
i
generate R

and give a projective embedding into P


4
. From here,
(14.2.8) follows easily. We thank Igor Dolgachev for showing us this example.
m
1
m
2
m
3
m
0
m
4
Figure 3. Pa and its lattice points m0, . . . , m4
Since P
a
P

, Proposition 14.2.12 and Theorem 14.2.13 have the following


immediate corollary.
14.2. Toric GIT and Polyhedra 695
Corollary 14.2.16. If G (C

)
r
and

G, then the GIT quotient C
r
//

G is a
semiprojective toric variety of dimension
dimC
r
//

G = dim P

.
Multigraded Polynomial Rings. This section began with an algebraic subgroup G
of (C

)
r
. Taking character groups, we obtained the exact sequence
0 M

Z
r


G 0
from (14.2.2). This gives a multigrading on the polynomial ring S = C[x
1
, . . . , x
r
]
by dening deg(x
a
) = (a)

G. An example is given by the grading on the total
coordinate ring of a toric variety without torus factors.
In particular, a character

G gives the graded piece S

of S, and one checks


that (C
r
, L

) S

. It follows that the graded ring R

used in the denition of


C
r
//

G is isomorphic to the ring


(14.2.9) R

d=0
S

d .
We will see below that determines an irrelevant ideal B() S and leads to a
quotient construction remarkably similar to what we did in Chapter 5.
Conversely, suppose that we start with a multigrading on C[x
1
, . . . , x
r
] given by
a surjective homomorphism : Z
r
A for a nitely generated abelian group A.
Applying Hom
Z
(, C

), we get an inclusion
G = Hom
Z
(A, C

) (C

)
r
= Hom
Z
(Z
r
, C

)
such that

G A. Hence the multigraded approach, featured in [190] and [204], is
fully equivalent to what we are doing here.
Matrices. When G is a torus, we have

GZ
s
. This implies M Z
rs
, so that the
exact sequence (14.2.2) can be written
(14.2.10) 0 Z
rs
B
Z
r
A
Z
s
0,
where B and A are integer matrices of respective sizes r (r s) and s r such
that AB = 0. Note that the r rows of B give
1
, . . . ,
r
N =Z
rs
.
Example 14.2.17. Let r N and consider the exact sequence
0 Z
2
B
Z
4
A
Z
2
0,
where
B =

1 r
0 1
1 0
0 1

, A =

1 r 1 0
0 1 0 1

.
696 Chapter 14. Toric GIT and the Secondary Fan
This is the sequence 0 M Z
4
Cl(H
r
) 0 that computes the class group
of the Hirzebruch surface H
r
. The rows
1
, . . . ,
4
of B are the minimal generators
u
i
of the fan of H
r
shown in Figure 3 of Example 4.1.8.
Here the group is G = (t, t
r
u, t, u) [ t, u C

. Note that the rst row of A


gives the t-exponent of elements of G and the second row gives the u-exponent.
Also, the last two columns of A show that the divisor classes [D
3
], [D
4
] correspond-
ing to u
3
, u
4
give a basis of the class group, and the rst two columns show how to
express [D
1
], [D
2
] in terms of this basis.
Example 14.2.18. In Example 14.2.15, the exact sequence (14.2.2) can be written
0 Z
3
B
Z
6
A
Z
3
0,
where
B =

1 0 0
1 1 1
1 1 1
0 1 0
1 1 1
0 0 1

, A =

2 1 1 0 0 0
0 1 0 2 1 0
0 0 1 0 1 2

.
The matrix A gives the map e
i j
e
i
+e
j
from Example 14.2.15, and given A, it
is easy to nd a matrix B that makes the sequence exact. We call B the Gale dual
of A. Note that the
i
s are the rows of B. See 14.2 for more on Gale duality.
As shown by Examples 14.2.17 and 14.2.18, the matrices B and A make it easy
to determine the vectors
i
and the group G.
Virtual Facets. The facets of a polyhedron contain a lot of information about its
geometry and combinatorics. For a GIT quotient C
r
//

G, we have the polyhedra


P

P
a
for =
a
. However, to capture the full story of what is going on, we
need not only their facets but also their virtual facets.
This is easiest to see in P
a
, which for a = (a
1
, . . . , a
r
) Z
r
is dened by
P
a
=m M
R
[ 'm,
i
` a
i
, 1 i r
as in (14.2.3). For 1 i r, we call the set
F
i,a
=m P
a
[ 'm,
i
` =a
i

a virtual facet of P
a
. The facets of P
a
all occur among the F
i,a
, while other virtual
facets may be quite differentsome may be faces of smaller dimension, some may
be empty, and some may be all of P
a
. We can also have F
i,a
= F
j,a
for i = j.
Example 14.2.19. In Example 14.2.4, we have M = Z
3
, a = (1, 0, 1), and
1
=

2
= e
1
,
3
= 2e
2
. Thus
P
a
=m M
R
[ 'm, e
1
` 1, 'm, e
1
` 0, 'm, 2e
2
` 1.
14.2. Toric GIT and Polyhedra 697
It follows that F
1,a
= while F
2,a
and F
3,a
are the facets of P
a
shown in Figure 1
from Example 14.2.4.
Example 14.2.20. Let G = (t, t
r
u, t, u) [ t, u C

as in Example 14.2.17. For


a =(0, 0, 0, 1), the polytope P
a
R
2
is shown in Figure 4. Here we have three gen-
F
4,a

F
2,a
F
1,a
F
3,a
P
a
Figure 4. The polytope Pa R
2
for a =(0, 0, 0, 1) and its virtual facets
uine facets F
1,a
, F
3,a
, F
4,a
and one virtual facet F
2,a
which is a vertex. We will see in
Example 14.3.7 below that P
a
is the polytope of the divisor D
4
on the Hirzebruch
surface H
r
.
Virtual facets can also be dened for P

. Using the description of P

given
in (14.2.6), one sees that F
i,a
P
a
corresponds to the subset of P

where the ith


coordinate vanishes. We will denote this set by F
i,
, so that
F
i,
= P

V(x
i
).
Semistable Points. Virtual facets determine the semistable points of C
r
as follows.
Proposition 14.2.21. Let p=(p
1
, . . . , p
r
) C
r
and set I(p) =i [ 1 i r, p
i
=0.
Then p (C
r
)
ss

if and only if

iI(p)
F
i,
=.
Proof. Let p be semistable, so that some (G,
d
)-invariant polynomial F does not
vanish at p. Writing F as a linear combination of (G,
d
)-invariant monomials
shows that some (G,
d
)-invariant monomial x
b
that does not vanish at p. Then
b = (b
1
, . . . , b
r
) lies in dP

. For I(p) as above, we have b


i
= 0 for all i I(p) since
x
b
does not vanish at p. This implies b

iI(p)
dF
i,
, hence

iI(p)
F
i,
=.
The converse is also easy, except that we have to be careful about torsion.
Assume that

iI(p)
F
i,
= . Then the intersection contains a rational point, so
that there is b (

iI(p)
dF
i,
) Z
r
for some integer d > 0. This implies two
things:
x
b
does not vanish at p.

R
(b) =
d
1, so that (b) and
d
differ by a torsion element of

G.
It follows that for some integer >0, x
b
is (G,
d
)-invariant. Then p is semistable
since x
b
does not vanish at p.
698 Chapter 14. Toric GIT and the Secondary Fan
When we think about semistable points algebraically, we get the ideal
(14.2.11) B() =

i / I
x
i
[ I 1, . . . , r,

iI
F
i,
=

S =C[x
1
, . . . , x
r
],
called the irrelevant ideal of . The vanishing locus of B() is the exceptional set
Z() =V(B()) C
r
. This gives a nice description of the set of semistable points,
which you will prove in Exercise 14.2.9.
Corollary 14.2.22. (C
r
)
ss

=C
r
`Z(), so that C
r
//

G (C
r
`Z())//G.
Exercises for 14.2.
14.2.1. Supply the details omitted in the proof of Lemma 14.2.1.
14.2.2. Supply the details omitted in Example 14.2.3.
14.2.3. Prove the description of P

given in (14.2.6).
14.2.4. The exercise will prove Lemma 14.2.8. As in the proof of Proposition 14.1.12,
Proj(R) is covered by the afne open subsets D
+
( f ) = Spec(R
( f )
), where f R
d
is non-
nilpotent and R
( f )
is the homogeneous localization (14.1.4).
(a) Note that f R
d
implies f

R
d
= (R
[]
)
d
. Prove that
(R
[]
)
( f

)
= R
( f )
.
(b) Pick f
1
, . . . , f
s
R homogeneous such that

' f
1
, . . . , f
s
` =R
+
=

d>0
R
d
. Prove that

1
, . . . , f

f
1
, . . . , f
s

= R
+
.
as ideals of R, and conclude that

1
, . . . , f

= (R
[]
)
+
as ideals of R
[]
.
(c) Prove that Proj(R
[]
) Proj(R). Hint: For f
1
, . . . , f
s
R as in part (b), D
+
( f
i
)
s
i=1
is
an open cover of Proj(R) by 7.0.
14.2.5. Prove Proposition 14.2.10.
14.2.6. Prove the assertions made in the three bullets in the proof of Proposition 14.2.12.
14.2.7. In (14.2.7) we claimed that the mapping (m, k) (m) +ka induces a semigroup
isomorphismC(P
a
) (MZ) S

. Prove this.
14.2.8. This exercise will ll in some of the details omitted in Example 14.2.15.
(a) Show that P
a
R
3
= M
R
is dened by x 0, y 0, z 0, x +y 1+z, x +z 1+y
and y +z 1 +x. Also show that the vertices are the only lattice points of P
a
.
(b) Prove that P
a
is normal. Hint: If (a, b, c) P
a
, study abc = 0 and abc > 0 separately.
(c) Prove (14.2.8)
14.2.9. Prove Corollary 14.2.22. Hint: Use Proposition 14.1.12.
14.2.10. Consider the group G =(t, t
1
, u) [ t C

, u =1 and character (t, t


1
, u) =
tu from Example 14.1.5. Note that =
a
for a =(1, 0, 1). By Example 14.1.8, the graded
pieces of R

have a complicated behavior, and by Example 14.2.4, multiples of P


a
have a
complicated behavior since P
a
is not a lattice polyhedron.
(a) Use the proof of Theorem 14.2.13 to explain how these complications are linked.
14.3. Toric GIT and Gale Duality 699
(b) In Example 14.2.9, we used the Veronese subring R
[2]

to show that Proj(R

) C
2
.
Explain how this relates to the lattice polyhedron 2P
a
.
14.2.11. Consider the Grassmannian G(1, 3) of lines in P
3
dened in Exercise 6.0.5. We
use the Pl ucker embedding G(1, 3) P
5
=P
(
4
2
)1
dened by
A =

0

1

2

3

0

1

2

3

(p
i j
) = (p
01
, p
02
, p
03
, p
12
, p
13
, p
23
),
where p
i j
is the 2 2 minor of A formed by columns i and j. An elementary treatment
of this Pl ucker embedding appears in [69, Ch. 8, 6]. The torus G = (C

)
4
acts on A
by right multiplication. This gives an action on G(1, 3) where g = (t
0
, t
1
, t
2
, t
3
) G acts
on Pl ucker coordinates via g (p
i j
) = (t
i
t
j
p
i j
). This lifts to an action on the afne cone

G(1, 3) C
6
= C
(
4
2
)
. The action of G extends to an action on C
6
. For the character
=
a
, a = (1, . . . , 1) Z
6
, prove that C
6
//

G P
2
. The more general GIT quotient

G(1, 3)//

G is described in [83, Ex. 12.4].


14.3. Toric GIT and Gale Duality
This section begins our study of what happens to the GIT quotient C
r
//

G as we
vary . A key player is Gale duality.
Gale Duality. A toric variety X

comes with two nite sets indexed by (1):


The minimal generators u

N.
The divisor classes [D

] Cl(X

).
When X

has no torus factors, these are related by the exact sequence


0 M Z
(1)
Cl(X

) 0,
where the rst map is dened by m ('m, u

`)
(1)
Z
(1)
and the second is
e

[D

] Cl(X

).
Gale duality generalizes this situation in the context of real vector spaces. Let
W be a vector space over R spanned by vectors
1
, . . . ,
r
. Some of the
i
may be
zero, and repetitions are allowedwe do not assume that the
i
are distinct. We
will use to denote this list of vectors. Then we have an exact sequence
(14.3.1) 0 V

R
r

W 0,
where (e
i
) =
i
and : V R
r
is the inclusion of the kernel of . Dualizing, we
obtain the exact sequence
(14.3.2) 0 W

R
r

0,
where we use dot product to identify R
r
with its dual. This gives vectors
i
=

(e
i
) V

. We will use to denote this list of vectors. As in 14.2, the map in


(14.3.1) is given by (v) = ('v,
1
`, . . . , 'v,
r
`) R
r
for v V.
Since (14.3.1) is the dual of (14.3.2), we could equally well start with the
vectors in V

and then recover the vectors by duality. This implies in particular


700 Chapter 14. Toric GIT and the Secondary Fan
that

in (14.3.2) is given by

(w) = ('
1
, w`, . . . , '
r
, w`) R
r
for w W

. The
symmetry between and will be used often in our discussion of Gale duality.
Our rst result describes what happens when a subset of or is a basis.
Lemma 14.3.1. Let I 1, . . . , r and set J =1, . . . , r`I. Then
i
, i I, form a
basis of W if and only if
j
, j J, form a basis of V

.
Proof. Note that [I[ = dimW implies [J[ = dimV = dimV

. Assume
j
, j J,
give a basis of V

and let v
j
, j J, be the dual basis of V. Then we have (v
j
) =
e
j
+

iI
'v
j
,
i
`e
i
for j J. Applying , we see that j J implies that

j
=

iI
'v
j
,
i
`
i
.
This proves that
i
, i I, span W, and since [I[ = dimW, they form a basis of W.
The other direction of the proof follows immediately by duality.
Using and we get cones C

= Cone() W and C

= Cone() V

.
Here is an example of how these cones are related.
Lemma 14.3.2. Consider the cones C

W and C

dened above.
(a) If C

= W, then C

is strongly convex in V

. Conversely, if consists of
nonzero vectors and C

is strongly convex in V

, then C

=W.
(b) If C

= V

, then C

is strongly convex in W. Conversely, if consists of


nonzero vectors and C

is strongly convex in W, then C

=V

.
Proof. Assume C

=W. Pick v C

(C

) and write
v =
r

i=1
a
i

i
=
r

i=1
b
i

i
with a
i
, b
i
0. Then

r
i=1
(a
i
+b
i
)
i
= 0. Since (14.3.2) is exact, there is w W

such that

(w) =

r
i=1
(a
i
+b
i
)e
i
, so that '
i
, w` =a
i
+b
i
by the description of

given in the discussion following (14.3.2). Since a


i
+b
i
0 for all i, we conclude
that w Cone()

= C

= 0, where the last equality follows from C

= W.
Then a
i
+b
i
= 0 for all i, so that a
i
= b
i
= 0 for all i, hence v = 0.
For the converse, take w C

. Then '
i
, w` 0 for all i. Since

(w) =

r
i=1
'
i
, w`e
i
, we have

r
i=1
'
i
, w`
i
=0. Strong convexity implies that 0 C

is a face, and then '


i
, w`
i
= 0 for all i by Lemma 1.2.7. Since the
i
are nonzero,
we must have '
i
, w` = 0 for all i, which implies w = 0 since the
i
span W. This
proves C

=0. Taking duals gives C

=W, completing the proof of part (a).


Part (b) of the lemma follow from part (a) by duality.
Gale duality also describes the faces of these cones. Here is the result for C

;
the corresponding statment for C

follows immediately by duality.


14.3. Toric GIT and Gale Duality 701
Lemma 14.3.3. Let I 1, . . . , r and set J = 1, . . . , r ` I. Then the following
are equivalent:
(a) There is a face F _C

such that
i
F if and only if i I.
(b) There are positive numbers b
i
, i J, such that

iJ
b
i

i
= 0.
Proof. Given (a), there is w C

such that '


i
, w` = 0 for i I and '
i
, w` > 0
for i J. Then (b) holds since 0 =

(w) =

iJ
'
i
, w`
i
. The proof of the
converse is equally easy (Exercise 14.3.1).
The introduction to Gale duality given here follows [222]. There is much more
to say about this subjectthe full story involves ideas such as circuits and oriented
matroids. See [281, Lec. 6] for a more complete treatment of Gale duality. Circuits
will play an important role in Chapter 15.
Characters as Vectors. We return to the GIT quotient C
r
//

G, where G (C

)
r
and

G. We will continue to regard

G as a multiplicative group, but since the
tensor product

G
R
=

G
Z
R is a vector space over R of dimension dim G, we will
switch to additive notation when working in

G
R
. Thus characters ,

G give
vectors 1, 1

G
R
such that () 1 = 1+1.
The inclusion G (C

)
r
gives characters
i


G dened by
i
(g) = t
i
for
g = (t
1
, . . . , t
r
) G. In terms of (14.2.2), this means
i
= (e
i
). These generate

G,
so that

G
R
is spanned by the vectors

i
=
i
1 =
R
(e
i
)

G
R
, 1 i r.
Note that a = (a
1
, . . . , a
r
) Z
r
gives the character

r
i=1

a
i
i
, which was written
(a) =
a
in the discussion following (14.2.2). It follows that

R
(a) =
a
1 =

r
i=1
a
i


G
R
.
The Two Cones. The vectors
i


G
R
give a list we denote by , and the vectors

i
=

(e
i
) N N
R
= M

R
give a list we denote by . Then the exact sequence
(14.3.3) 0 M
R

R
R
r

R


G
R
0
from (14.2.4) satises

R
(e
i
) =
i
for 1 i r.

R
(m) = ('m,
1
`, . . . , 'm,
r
`) for m M
R
.
Hence Gale duality applies to this situation. We have the additional structure given
by the lattices M M
R
and

G1

G
R
. Furthermore, the cones
C

= Cone() = Cone(
1
, . . . ,
r
)

G
R
C

= Cone() = Cone(
1
, . . . ,
r
) N
R
702 Chapter 14. Toric GIT and the Secondary Fan
are full dimensional in

G
R
and N
R
and are rational polyhedral with respect to the
lattices

G1 and N. These cones will play an important role in this section.
Example 14.3.4. Let X
P
be the toric variety of a full dimensional lattice polyhe-
dron P M
R
. Then (14.3.3) is obtained from the standard exact sequence
0 M Z

P
(1)
Cl(X
P
) 0
by tensoring with R. Thus:
The
i
s come from the divisor classes [D

], so that the cone C

is the cone of
torus-invariant effective R-divisor classes.
The
i
s are the minimal generators u

, so that the cone C

is the support of
the normal fan
P
.
Existence of Semistable and Stable Points. We rst show the cone C

controls
when the GIT quotient is nonempty, i.e., when semistable points exist.
Proposition 14.3.5. If

G is a character, then the following are equivalent:
(a) C
r
//

G=.
(b) (C
r
)
ss

=.
(c) 1 C

.
Proof. We have (a) (b) by Proposition 14.1.12. Using dimC
r
//

G = dim P

(Corollary 14.2.16) and P

=
1
R
(1) R
r
0
, (a) (c) follows from
C
r
//

G = P

= there is b = (b
1
, . . . , b
r
)
1
R
(1) R
r
0
1 =
r

i=1
b
i

i
, b
i
0 1 C

.
Our next task is to determine when stable points exist. Here, the interior of C

is the key player.


Proposition 14.3.6. If 1 C

, then the following are equivalent:


(a) (C

)
r
(C
r
)
s

.
(b) (C
r
)
s

=.
(c) 1 is contained in the interior of C

.
(d) F
i,
is a proper face of P

for all i.
Furthermore, the above conditions imply that
(e) dimC
r
//

G = r dim G,
and when consists of nonzero vectors, (e) is equivalent to (a)(d).
Proof. Throughout the proof we pick a = (a
1
, . . . , a
r
) Z
r
such that =
a
.
14.3. Toric GIT and Gale Duality 703
The implication (a) (b) is obvious. Assume (b) and suppose that 1 F,
where F is a proper face of C

. Let I = i [
i
F and J = 1, . . . , r`I. Note
that J is nonempty since F is a proper face. If b = (b
1
, . . . , b
r
) P

, then
1 =

r
i=1
b
i

i
.
Then Lemma 1.2.7 and 1 F imply that b
i

i
F for all i. Since
i
/ F for i J,
we must have b
i
= 0 for i J. Thus F
i,
= P

for i J. Relabeling if necessary, we


may assume 1 J. Then F
1,
=P

and (14.2.11) imply that the minimal generators


of the ideal B() do not involve x
1
. By Corollary 14.2.22, we have
(14.3.4) (C
r
)
ss

=CU, U C
r1
open.
Translated to P
a
, F
1,
= P

means that
P
a
m M
R
[ 'm,
1
` = a
1
.
Since P
a
has full dimension in M
R
by Proposition 14.1.12 and Theorem 14.2.13,
this inclusion forces
1
= 0. Using the description of G from Lemma 14.2.1, we
get a product decomposition
G =C

(C

)
r1
= (C

)
r
.
Thus (t) = (t, 1) C

= G is a one-parameter subgroup of G. Via (14.3.4),


pick p = (p
1
, p

) CU = (C
r
)
ss

. Then (t) p G p, and the limit


lim
t0
(t) p = lim
t0
(t p
1
, p

) = (0, p

) CU
exists as point of (C
r
)
ss

. But this limit is not in G p when p


1
= 0, so that G p
is not closed in (C
r
)
ss

when p
1
= 0. Thus p cannot be stable by Remark 14.1.13.
This is easily seen to contradict (C
r
)
s

=, and (b) (c) follows.


Assume (c) and let J =i [ F
i,
=P

. Suppose J = and set I =1, . . . , r`J.


Since F
i,
P

for each i I, we can pick c = (c


1
, . . . , c
r
) P

`(

iI
F
i,
). Then
the ith coordinate of c is positive for i I and is zero for i J. Hence
(14.3.5) 1 =

iI
c
i

i
Suppose that Cone(
i
[ i J) N
R
is strongly convex. Then 0 is a face,
which implies that there is m M
R
such that 'm,
i
` > 0 for all i J. This gives
c

= ('m,
1
`, . . . , 'm,
r
`)
1
R
(0) whose ith coordinate is positive for i J. Then
b = c +c


1
R
(0)
for all , and when > 0 is sufciently small, the ith coordinate of b is positive for
all i. This follows easily by treating the cases i I and i J separately. Thus b P

has positive coordinates, which contradicts J = i [ F


i,
= P

= . We conclude
that Cone(
i
[ i J) is not strongly convex. This easily gives a nonempty subset
J

J such that

iJ

i

i
= 0,
i
> 0
704 Chapter 14. Toric GIT and the Secondary Fan
(Exercise 14.3.2). By Gale duality (Lemma 14.3.3), it follows that there is a face
F _C

such that i [
i
F =1, . . . , r`J

. This is a proper face since J

=, and
it contains 1 by (14.3.5) and I =1, . . . , r`J. This contradicts 1 Int(C

)
and completes the proof of (c) (d).
Assume (d). Since F
i,
= P

for all i, P

has a rational point with positive


coordinates. Arguing as in the proof of Proposition 14.2.21, there is d > 0 and
b dP

Z
r
>0
such that x
b
is (G,
d
)-invariant. Since all exponents are positive,
the nonvanishing set of x
b
is (C

)
r
. It follows easily that points of (C

)
r
are stable.
This proves (d) (a).
The implication (b) (e) follows from Proposition 14.1.12. Now assume (e),
so that P
a
is full dimensional in M
R
. This implies that if
i
= 0, then m 'm,
i
`
cannot be constant on P
a
, and hence F
i,a
= P
a
. Thus (e) (d) when consists of
nonzero vectors, and the proof is complete.
Proposition 14.3.6 has many nice consequences concerning C
r
//

G when
consists of nonzero vectors, including:
Stable points exist if and only if the GIT quotient has the expected dimension.
The GIT quotient is nonempty of smaller than expected dimension if and only
if the character is on the boundary of C

.
Here is an example to illustrate the last bullet.
Example 14.3.7. Let G =(t, t
r
u, t, u) [ t, u C

(C

)
4
. This group appeared
in Example 14.2.17. For r = 1, the cone C


G
R
R
2
is the shaded region in
Figure 5. The gure also shows
1
,
2
,
3
,
4
. Three of the P

s in the gure include


dashed lines that represent virtual facets, two of which are empty and the other of
which is a vertex. Note that dim P

= dimC
4
//

G drops dimension precisely on


the boundary of C

, as predicted by Proposition 14.3.6.


P

=
P

=
P

=
P

=
P

1
=
3

2
Figure 5. C

and the polytopes P for all 1 C

14.3. Toric GIT and Gale Duality 705


Figure 5 is closely related to Figure 12 in Example 6.3.23. In that example, our
focus was on the Hirzebruch surface H
r
, so there we saw only the rst quadrant
portion of Figure 5. Looking at all of Figure 5, we see that away from the boundary,
C
r
//

G is either P
2
or H
1
, depending on which quadrant we are in.
In many of the most interesting cases, consists of nonzero vectors. We should
nevertheless give an example where one of the
i
s vanishes.
Example 14.3.8. Consider the group G = (t, u, u) [ t, u C

and the character


(t, u, u) = u. We leave it as Exercise 14.3.3 to show that:

1
= 0,
2
= e
1
,
3
=e
1
, where e
1
is a basis of N
R
Z.
C

=R
2
0
and 1 = (0, 1) C

.
P

= Conv((0, 1, 0), (0, 0, 1)) and F


1,
= P

.
(C
3
)
ss
=C(C
2
`0) and (C
3
)
s
=.
Note that parts (a), (b), (c), (d) of Proposition 14.3.6 are false while part (e) is true
since C
3
//

G P
1
has the expected dimension 32 = 1.
Projective and Birational GIT Quotients. By Proposition 14.1.12, the quotients
we are studying come with a projective morphism
(14.3.6) C
r
//

G C
r
//G = Spec(C[x
1
, . . . , x
r
]
G
).
Here we explore the extreme cases where C
r
//G is very small or very large. We
will need the following result about C

= Cone() N
R
.
Lemma 14.3.9.
(a) C[x
1
, . . . , x
r
]
G
=C[C

M], so that C
r
//G = Spec(C[C

M]).
(b) If 1 C

, then C

M
R
is the recession cone of P
a
when =
a
.
Proof. For part (a), note that x
b
is G-invariant if and only if (b) is trivial, which
is equivalent to b = (m) for some m M. Since b N
r
, the formula for (m)
shows that C[x
1
, . . . , x
r
]
G
is the semigroup algebra of lattice points in the cone
(14.3.7) m M
R
[ 'm,
i
` 0, 1 i r = Cone(
1
, . . . ,
r
)

=C

.
For part (b), let 1 C

. Then P
a
is nonempty by the proof of Proposition 14.3.5
and hence has recession cone (14.3.7) by the proof of Lemma 14.2.2.
Our rst extreme case is when C
r
//G = Spec(C[x
1
, . . . , x
r
]
G
) is very small.
Proposition 14.3.10. The following are equivalent:
(a) C
r
//

G is projective for all 1 C

.
(b) C
r
//

G is projective for some 1 C

.
(c) C[x
1
, . . . , x
r
]
G
=C.
(d) consists of nonzero vectors and C

is strongly convex.
706 Chapter 14. Toric GIT and the Secondary Fan
Proof. First, (a) (b) is clear, and (b) (c) follows from Exercise 14.1.9 since
1 C

implies C
r
//

G =. We also have (c) (a) by Exercise 14.1.7.


If (d) is true, then C

= N
R
by Gale duality (Lemma 14.3.2), and then we have
C

= 0. Assume 1 C

. Then C

= 0 is the recession cone of P


a
by
Lemma 14.3.9, so that P
a
is a polytope. It follows that C
r
//

GX
Pa
is projective.
Finally, suppose that (c) is true. If
i
= 0, then
i
=
i
1 implies that
i


G
has nite order, say . Since
i
: G C

is projection onto the ith coordinate,


it follows that C[x

i
] C[x
1
, . . . , x
r
]
G
, a contradiction. Hence the
i
are nonzero.
Also, (c) and Lemma 14.3.9 imply that C

= 0, so that C

= N
R
. Hence C

is
strongly convex by Gale duality (Lemma 14.3.2).
The other extreme is when C
r
//G is large. Since dimC
r
//G r dim G by
Remark 14.1.13, C
r
//G is as large as possible when dimC
r
//G = r dim G. Here
is our result.
Proposition 14.3.11. The following conditions are equivalent:
(a) dimC
r
//G = r dim G.
(b) C

N
R
is strongly convex.
(c) The map C
r
//

G C
r
//G from (14.3.6) is birational for all 1 C

.
Furthermore, when consists of nonzero vectors, we can replace (b) and (c) with:
(b

) C

=

G
R
.
(c

) C
r
//

G C
r
//G is birational for all .
Proof. Lemma 14.3.9 implies dimC
r
//G = dimC

. Since r dim G = dim M


R
,
we obtain
dimC
r
//G = r dim G dimC

= dim M
R
C

N
R
is strongly convex.
This proves (a) (b). Now assume (b). Then C

is strongly convex, so that


C
r
//G is the afne toric variety of C

by Lemma 14.3.9. Furthermore, P


a
is full
dimensional since its recession cone is C

by Lemma 14.3.9. Hence the normal


fan of P
a
renes C

by Theorem 7.1.6. Then (c) follows since C


r
//

GC
r
//G is
the birational map induced by this renement. On the other hand, if (c) holds, then
C
r
//

G C
r
//G is birational for 1 Int(C

), so that
dimC
r
//G = dimC
r
//

G = r dim G,
where the last equality uses Proposition 14.3.6. Hence (c) (a).
Assume that consists of nonzero vectors. By Gale duality (Lemma 14.3.2),
C

N
R
is strongly convex if and only if C

=

G
R
, which gives (b) (b

). Then
(c) (c

) follows immediately.
14.3. Toric GIT and Gale Duality 707
Example 14.3.12. For G =(t, t, t
1
) [ t C

, one gets
1
=
2
= e
1
,
3
=e
1
in

G
R
R and
1
= e
1
,
2
= e
2
,
3
= e
1
+e
2
in N
R
R
2
. Thus C

= Cone(e
1
, e
2
)
is strongly convex and
i
= 0 for all i. Hence all conditions of Proposition 14.3.11
apply, i.e., for all

G, we have a birational morphism
C
3
//

G C
3
//G = Spec(C[xz, yz]) =C
2
.
To understand C
r
//

G, consider Figure 6, which shows the polyhedron P

for all . For positive , C


r
//

G is the blowup of C
2
at the origin, while for
P

1
=
2
0
3
P

Figure 6. The polytopes P for all


negative , the GIT quotient is C
2
. The dashed lines in the gure indicate the
presence of a virtual facet, which is empty for negative and consists of the vertex
when is the origin. In Exercise 14.3.4 you will compare this to the quotient
construction of Bl
0
(C
2
).
Generic Characters. We next study what it means for a character to be generic.
Denition 14.3.13. A character

G is generic if 1 C

and for every subset

with dim Cone(

) < dimC

= dim G, we have 1 / Cone(

).
We can determine when a character is generic as follows.
Theorem 14.3.14. For 1 C

, the following are equivalent:


(a) is generic.
(b) Every vertex of P

has precisely dim G nonzero coordinates.


(c) P

is simple of dimension r dim G, every virtual facet F


i,
P

is either
empty or a genuine facet, and F
i,
= F
j,
if i = j and F
i,
, F
j,
are nonempty.
(d) (C
r
)
s

= (C
r
)
ss

.
Proof. We will prove that (a), (c) and (d) are each equivalent to (b).
For (a) (b), assume 1 C

is generic. Then it does not lie on the


boundary of C

, which implies dimC


r
//

G = r dim G by Proposition 14.3.6. In


other words, dim P

= r dim G.
Now take any vertex b P

. Then dim P

= r dim G implies that at least


this many facets meet at b. Since all facets of P

occur among the virtual facets


F
i,
= P

V(x
i
), it follows that at least r dim G coordinates of b vanish. Thus
b has at most dim G nonzero coordinates. Suppose that b has s < dim G nonzero
708 Chapter 14. Toric GIT and the Secondary Fan
coordinates. Writing b =

s
j=1

j
e
i
j
, u
j
>0, we obtain 1 =

s
j=1

i
j
, which
is impossible since is generic and s < dim G. This proves (a) (b).
To prove (b) (a), assume that is not generic. By Carath edorys theorem,
1 =

s
j=1

i
j
, where u
j
> 0 and s < dim G (Exercise 14.3.5). Then b =

s
j=1

j
e
i
j
P

has fewer than dim G nonzero coordinates. In Exercise 14.3.5 you


will prove that this gives a vertex of P

with the same property, contradicting (b).


To prove (b) (c), assume (b) and take a vertex b P

. Arguing as above, we
see that dim P

= r dim G and that the actual facets of P

make at least r dim G


coordinates of b vanish. But exactly r dim G vanish by (b), so that the only
virtual facets containing b correspond to actual facets. Then (c) follows easily.
The proof of (c) (b) is similar and is left to the reader (Exercise 14.3.5).
Before proving (b) (d), we rst describe the semistable points. Given a
subset I 1, . . . , r, note that

iI
F
i,
= if and only if there is a vertex b P

with b F
i,
for all i I. Hence the maximal such subsets are those given by
i [ b F
i,
, b P

a vertex. It follows that the ideal B() dened in (14.2.11) is


given by
(14.3.8) B() =

b/ F
i,
x
i
[ b is a vertex of P

.
By Corollary 14.2.22, it follows that
(14.3.9) (C
r
)
ss

b a vertex
C
r
b
.
where C
r
b
=(p
1
, . . . , p
r
) C
r
[ p
i
= 0 when b / F
i,
.
We now turn to (b) (d). The rst step is to show that G-orbits are closed in
(C
r
)
ss

. By (14.3.9), we need only show that G-orbits of points in C


r
b
are closed in
C
r
b
for every vertex of b. By hypothesis, b has precisely r dim G coordinates that
vanish, i.e., precisely this many virtual facets F
i,
contain b. Let I =i [ b F
i,
.
By the earlier part of the proof, we know that P

is simple of dimension r dim G.


Hence the F
i,
, i I, are the genuine facets containing b, so the facet normals
i
,
i I, form a basis of M
R
.
Since the connected component of the identity G

G has nite index, it suf-


ces to show that G

p is closed in C
r
b
for all p C
r
b
. Take p G

p C
r
b
.
Using Lemma 5.1.10 and arguing as in the proof of Theorem 5.1.11, we can nd a
one-parameter subgroup : C

and a point g G

such that
(14.3.10) p = lim
t0
(t)g p.
Since G

(C

)
r
, we can write (t) = (t
a
1
, . . . , t
ar
) for exponents a
i
Z, and
(14.3.11)

r
i=1
a
i

i
= 0,
follows from (14.2.2) since is a one-parameter subgroup of G (Exercise 14.3.5).
The coordinates of p, p, g are nonzero for i / I, hence the limit (14.3.10) implies
that a
i
= 0 for i / I. Thus (14.3.11) becomes

iI
a
i

i
= 0, so that a
i
= 0 for all
14.3. Toric GIT and Gale Duality 709
i since the
i
, i I, are linearly independent. Hence is the trivial one-parameter
subgroup. which implies p = g p G

p by (14.3.10). We conclude that G

p
is closed in C
r
b
.
In Exercise 14.3.5 you will show that the isotropy subgroup G
p
is nite for
p (C
r
)
ss

. This and the previous paragraphs imply that every semistable point is
stable, which completes the proof of (b) (d).
Finally, for (d) (b), rst note that (C
r
)
s

=, so that dim P

=dimC
r
//

G=
r dim G = dim M
R
by Proposition 14.1.12. As noted earlier, this implies that
every vertex of P

has at least r dim G coordinates that vanish. Now assume (b)


fails, so that some vertex b P

has I = i [ b F
i,
with [I[ > r dim G. Thus
the
i
M
R
, i I, are linearly dependent since dim M
R
= r dim G. Then there is
a relation

iI
a
i

i
= 0 where a
i
Z and a
i
> 0 for at least one i. If we set a
i
= 0
for i / I, then
(t) = (t
a
1
, . . . , t
ar
) (C

)
r
is a one-parameter subgroup of G by Exercise 14.3.5. Now dene the point p =
(p
1
, . . . , p
r
) C
r
b
by
p
i
=

1 a
i
0
0 a
i
< 0.
Consider lim
t0
(t) p. The limit exists in C
r
since p
i
= 0 for a
i
< 0, and if i / I,
then the ith coordinate of (t) p is 1 for all t, so that the limit p = lim
t0
(t) p
lies in C
r
b
. Since there is i
0
I with a
i
0
>0, the i
0
th coordinate of p =lim
t0
(t) p
is zero. However, the i
0
th coordinate of p is nonzero, so the same is true for G p.
Hence G p is not closed in C
r
b
, and it follows that p is a point that is semistable
but not stable. This contradicts (d) and completes the proof of the theorem.
The equivalence (a) (b) in Theorem 14.3.14 appears in [137], while (c) is
taken from [275]. Our treatment of (d) was inspired by [83, Prop. 12.1].
Example 14.3.15. In Example 14.3.12, every nontrivial character is generic. When
is trivial, Figure 6 shows that part (c) of Theorem 14.3.14 fails because there is a
nonempty virtual facet that is not a facet.
Example 14.3.16. Figure 5 of Example 14.3.7 makes it clear that the non-generic
elements of C

lie on the three rays generated by the


i
s. Hence the generic
elements form two chambers, a term that will be dened in 14.4. In the chamber
on the right in Figure 5, P

is a quadrilateral, while in the chamber on the left, P

is a triangle with an empty virtual facet indicated by the dashed line. When
lies on the vertical axis, P

is also a triangle but is non-generic because of the


nonempty virtual facet where the dashed line touches the triangle. This illustrates
the importance of the virtual facets, as described in part (c) of Theorem 14.3.14.
We will see in 14.4 that Figure 5 is an example of a secondary fan.
710 Chapter 14. Toric GIT and the Secondary Fan
Example 14.3.17. The polytope in Figure 3 of Example 14.2.15 is not simple, so
that the character in that example is not generic. Since was chosen because
it gave the monomials appearing in the determinant, this shows that geometrically
interesting characters may fail to be generic.
Let us give one more detailed example of the theorem.
Example 14.3.18. Consider G =(t, t, t
1
) [ t C

as in Example 14.3.12. If we
choose the trivial character = 1, then the GIT quotient is
C
3
//

G =C
3
//G = Spec(C[x, y, z]
G
) = Spec(C[xz, yz]) =C
2
.
Every part of Theorem 14.3.14 fails in this example. It is instructive to see how
this happens:
(a) The trivial character is never generic, as follows from the denition.
(b) P

is the middle polyhedron in Figure 6 of Example 14.3.12. All three virtual


facets meet at the vertex of P

, which means that too many coordinates vanish.


(c) As shown by Figure 6, P

has a nonempty virtual facet that is not a genuine


facet. Yet P

is simple and has the correct dimension.


(d) = 1 easily implies that (C
3
)
ss

=C
3
. Since G (1, 0, 0) =C

(0, 0) is not
closed in (C
3
)
ss

, it follows that (1, 0, 0) / (C


3
)
s

.
For the quotient X

(C
(1)
` Z())//G considered in Proposition 14.1.9, all
semistable points were stable when the toric variety is simplicial. For general GIT
quotients such as the one considered here, things can be more complicated.
See also Exercise 14.3.6 for an example where the same genuine facet occurs
for two distinct indices.
The following immediate corollary of Theorem 14.3.14 summarizes the nice
properties of GIT quotients when the character is generic.
Corollary 14.3.19. If

G is generic, then C
r
//

G is a simplicial semiprojective
toric variety of the expected dimension r dim G with (C
r
)
s

= (C
r
)
ss

.
Generic characters are clearly very nice. In the next section we will see that
generic characters determine a renement of C

, the so-called secondary fan, that


tells the full story of C
r
//

G as we vary .
Exercises for 14.3.
14.3.1. Complete the proof of Lemma 14.3.3.
14.3.2. Show that a cone Cone(A) is not strongly convex if and only there is a nonempty
subset A

A and
v
> 0 for v A

such that

vA

v
v = 0.
14.3.3. Supply the details omitted in Example 14.3.8.
14.3. Toric GIT and Gale Duality 711
14.3.4. Explain how Example 14.3.12 relates to the quotient construction of Bl
0
(C
2
) from
Example 5.1.16. See also Example 7.1.4.
14.3.5. This exercise is devoted to the proof of Theorem 14.3.14.
(a) In the proof of (b) (a), we claimed that Carath edorys theorem (see the proof of
Theorem 2.2.12) implies that if is not generic, then 1 =

s
j=1

i j
, where
u
j
> 0 and s < dim G. Prove this.
(b) Suppose that a point b P

has fewer than dim G nonzero coordinates. Prove that


there is a vertex of P

with the same property. Hint: Write b as a convex combination


of vertices plus an element of the recession cone.
(c) Given (a
1
, . . . , a
r
) Z
r
, prove that (t) = (t
a1
, . . . , t
ar
) denes a one-parameter sub-
group of G if and only if

r
i=1
a
i

i
= 0.
(d) Complete the proof of (b) (d) by showing that the isotropy subgroup G
p
is nite
for every p C
r
b
. Hint: Let I = i [ b F
i,
. Then show two things: rst, that
g = (t
1
, . . . , t
r
) G
p
has g
i
= 1 for i / I and second, that the
i
, i I, form a basis of
N
R
. Now use Exercise 14.2.1 to show that G has nite order. You may also want to
look at Exercise 5.1.11.
14.3.6. Consider G =(t, t
1
) [ t C

with the trivial character = 1.


(a) Show that
1
=
2
= e
1
and
1
= e
1
,
2
=e
1
.
(b) Show that P

= Cone(e
1
) M
R
R with F
1,
= F
2,
=0.
(c) Show that C
2
//

G =C.
(d) Show that (C
2
)
ss

=C
2
and (C
2
)
s

= (C

)
2
.
This is our rst example where a character fails to be generic because the same genuine
facet occurs for two separate indices.
14.3.7. In the situation of Gale duality, let I 1, . . . , r and J =1, . . . , r ` I. Prove that
the
i
, i I, are linearly independent in W if and only if the
j
, j J, span V

. Hint:
Enlarge I to get a basis of W.
14.3.8. Suppose that W R with basis e
1
. Then pick vectors
1
= e
1
and
2
= 0 and
dene V as in (14.3.1).
(a) Show that V has a basis e
1
such that
1
= 0 and
2
= e
1
.
(b) Use this example to explain why the nonzero vector hypothesis is needed in parts (b)
and (d) of Lemma 14.3.2.
(c) Adapt this situation to Z and give an example where part (e) of Proposition 14.3.6 is
true but the other parts of the proposition are false.
14.3.9. An augmented polyhedron consists of a polyhedron P and a list of faces F
i
of P for
1 i r. We allow F
i
= or F
i
= P, and we can also have F
i
= F
j
for i = j. However,
we require that all actual facets of P occur among the F
i
. An example of an augmented
polyhedron is given by P

with its virtual facets F


i,
, 1 i r.
(a) Dene what it means for an augmented polyhedron to be simple in terms of the number
of virtual facets containing a vertex.
(b) Restate part (c) of Theorem 14.3.14 using augmented polyhedra. It will be elegant.
(c) Use one of the examples given in the text to show that there is a non-simple augmented
polyhedron whose underlying polyhedron is simple.
712 Chapter 14. Toric GIT and the Secondary Fan
14.4. The Secondary Fan
For G (C

)
r
, we study the GIT quotient C
r
//

G as

G varies. The cones
C

= Cone() = Cone(
1
, . . . ,
r
)

G
R
C

= Cone() = Cone(
1
, . . . ,
r
) N
R
introduced in 14.3 will play a key role in our discussion. Recall that

G
R
and N
R
have the lattices

G1 and N.
A rst observation is that there are only nitely many distinct GIT quotients
up to isomorphism. To see why, recall from Corollary 14.2.22 that
C
r
//

C
r
`Z()

//G,
where Z() is the vanishing locus of the irrelevant ideal (14.2.11) dened by
(14.4.1) B() =

i / I
x
i
[ I 1, . . . , r,

iI
F
i,
=

C[x
1
, . . . , x
r
].
There are only nitely many such ideals, hence only nitely many GIT quotients.
This decomposes the lattice points of C

into nitely many disjoint subsets where


C
r
//

G is constant on each subset. The basic idea is that each subset lies in the
relative interior of a cone of the secondary fan, whose support is C

.
A Shift in Focus. Giving a rigorous construction of the secondary fan will require
some new ideas, and we will also need to work in R
r
rather than in

G
R
. Given
a Z
r
, the polyhedron P
a
M
R
gives two objects of interest:
The generalized fan , which is the normal fan of P
a
in N
R
.
The set I

=i [ F
i,a
=, which determines the empty virtual facets of P
a
.
We can characterize the pairs (, I

) that occur as follows.


Proposition 14.4.1. Given a generalized fan in N
R
and subset I

1, . . . , r,
then and I

come from some a Z


r
if and only if
(a) [[ =C

.
(b) X

is semiprojective.
(c) = Cone(
i
[
i
, i / I

) for every .
Proof. First suppose that and I

come from a Z
r
. Since is the normal fan
of P
a
, Lemma 14.2.2 and Proposition 14.2.10 imply [[ =C

. Furthermore, X

is
the toric variety of P
a
and hence is semiprojective by Proposition 14.2.12. Thus
satises conditions (a) and (b).
For (c), it sufces to consider maximal cones, which correspond to vertices
v P
a
. In Exercise 14.4.1 you will show that v gives the maximal cone
Cone(
i
[ 'v,
i
` =a
i
) = Cone(
i
[ v F
i,a
).
Then (c) follows since v F
i,a
implies i / I

,
14.4. The Secondary Fan 713
Conversely, suppose that and I

satisfy (a), (b) and (c). Let


0
be the minimal
cone of and set N = N/
0
. Then gives a genuine fan in N
R
. Since X

=
X

is semiprojective, Theorem 7.2.4 and Proposition 7.2.9 imply that X

has a
torus-invariant Cartier divisor whose support function is strictly convex on [[.
Composing this with the map N
R
N
R
gives a strictly convex support function
on [[ that takes integer values on N. Then dene a = (a
1
, . . . , a
r
) Z
r
by
a
i
=

(
i
) i / I

(
i
) 1 i I

.
In Exercise 14.4.1 you will show that and I

come from a.
GKZ Cones. Following Gel

fand, Kapranov and Zelevinsky [113, Ch. 7] (see also


[112]), we construct the secondary fan using cones dened in terms of support
functions. Our treatment is also inuenced by [141] and [222].
Support functions with respect to a fan were dened in Denition 4.2.11. This
denition extends to generalized fans without change. Let SF() denote the set of
all support functions with respect the generalized fan . Since has full dimen-
sional convex support by Proposition 14.4.1, we can dene
(14.4.2) CSF() = SF() [ is convex.
The convexity criteria from 6.1 and 7.2 apply to our situation.
Denition 14.4.2. Let and I

1, . . . , r be as in Proposition 14.4.1. Then the


GKZ cone of and I

is the set

,I

(a
1
, . . . , a
r
) R
r
[ there is CSF() such that
(
i
) =a
i
for i / I

and (
i
) a
i
for i I

.
We will see below that

,I

is a cone. Note that when a

,I

, the support
function is unique. This follows from Proposition 14.4.1 since is linear on the
cones of and satises (
i
) =a
i
for i / I

. Hence we write as
a
.
Here are some easy properties of GKZ cones.
Proposition 14.4.3. Let and I

be as above. Then:
(a)

,I

is a rational polyhedral cone in R


r
.
(b) The minimal face of

,I

is ker(
R
) = im(
R
) M
R
.
Proof. We begin with a different model of

,I

that will be useful later in the


section. Consider the direct sum W =R
r

max
M
R
. We write points of W as
(a, m

), where a = (a
1
, . . . , a
r
) and m

=m

max
. Then let (
,I

W be
the subset consisting of all points (a, m

) satisfying
(14.4.3)
'm

,
i
` =a
i
, for
i
and i / I

'm

,
i
` a
i
, for
i
/ or i I

.
714 Chapter 14. Toric GIT and the Secondary Fan
Let : W R
r
be the projection map. We claim that induces a bijection
(14.4.4) (
,I

,I

.
To prove this, take (a, m

) (
,I

. Then (u) =min


max
'm

, u` is convex
by Lemma 6.1.5, and by (14.4.3), we have (
i
) = a
i
, i / I

, and (
i
) a
i
,
i I

. If
max
, then (14.4.3) also implies that 'm

,
i
` =a
i
for
i
, i / I

.
Since these
i
s generate by Proposition 14.4.1, we have CSF() and thus
a

,I

. Conversely, let a

,I

. For each
max
there is m

M
R
such that

a
(u) ='m

, u` for u . By convexity and Lemma 6.1.5, 'm

, u`
a
(u) for all
u [[. From here, one sees easily that (a, m

) (
,I

.
For part (a), observe that (
,I

is a rational polyhedral cone in W relative to the


lattice Z
r

max
M since
i
N for all i. Then (14.4.4) implies that

,I

is a
rational polyhedral cone in R
r
.
For part (b), note that for m M
R
, the point
R
(m) = ('m,
1
`, . . . , 'm,
r
`) lies
in

,I

because of the convex support function dened by (u) = 'm, u`. Now
take a

,I

,I

). Then
a
and
a
=
a
are convex, so that
a
is linear,
say
a
(u) = 'm, u` for u C

. Since
a
(
i
) a
i
and
a
(
i
) a
i
for all i,
we obtain 'm,
i
` =a
i
for all i, i.e.,
R
(m) = a.
We will soon see that the GKZ cones form a generalized fan in R
r
. The rst
step is Proposition 14.4.3, which shows that GKZ cones are rational polyhedral
cones. The other properties we need will be covered in three lemmas that describe
the relative interiors, union, and faces of GKZ cones.
Three Lemmas. Our rst lemma describes the relative interior of a GKZ cone.
Lemma 14.4.4. If

,I

is a GKZ cone, then


Relint(

,I

) =a

,I

[
a
is strictly convex and
a
(
i
) >a
i
, i I

.
Furthermore, if a Relint(

,I

), then:
(a) is the normal fan of P
a
.
(b) I

=i [
a
(
i
) >a
i
=i [ F
i,a
=.
Proof. Let (
,I

W be the cone from the proof of Proposition 14.4.3. The rst


line of (14.4.3) denes a subspace W
0
W containing (
,I

, and then the second


line shows that (
,I

W
0
is dened by the inequalities
(14.4.5)
'm

,
i
` a
i
, for
i
/ and i / I

'm

,
i
` a
i
, for i I

.
By assumption, and I

come from a polytope P


a
, a Z
r
. Let the vertices of
P
a
be m

for
max
. This gives the point (a, m

) (
,I

that makes all of the


14.4. The Secondary Fan 715
inequalities in (14.4.5) strict. It follows easily that the relative interior of (
,I

is
the subset of (
,I

dened by
(14.4.6)
'm

,
i
` >a
i
, for
i
/ and i / I

'm

,
i
` >a
i
, for i I

.
We claim that these strict inequalities correspond via (14.4.4) to the subset
a

,I

[ is strictly convex and (


i
) >a
i
, i I

.
To prove this, note that an arbitrary point (a, m

) (
,I

gives a support function

a
such that
a
(u) ='m

, u` for u . Since
max
is built from the
i
for i / I

,
the rst line of (14.4.6) is equivalent to the strict convexity of
a
by (a) (f) of
Lemma 6.1.13, and the second line is equivalent to
a
(
i
) > a
i
for i I

. This
proves our claim, and our description of Relint(

,I

) follows.
Now take a Relint(

,I

) and suppose that (a, m

) (
,I

maps to a. Then
m

represents
a
on
max
, and the convexity of
a
implies that

a
(u) = min
max
'm

, u`.
by Lemma 6.1.5. Then Theorem 7.2.2 shows that the m

are the vertices of P


a
. The
cone of the normal fan corresponding to this vertex is
Cone(P
a
m

,
which is easily seen to be (Exercise 14.4.1). This proves part (a).
For part (b), take i I

. Then Proposition 14.4.3 implies that


a
(
i
) > a
i
since a is in the relative interior. On the other hand, if i / I

, then pick
max
with
i
. The rst line of (14.4.3) implies that
a
(
i
) = a
i
, and then part (b)
follows easily.
A nice consequence of Lemma 14.4.4 is that the relative interiors of two GKZ
cones are either equal or disjoint.
Our second lemma describes the union of the GKZ cones.
Lemma 14.4.5.

,I

,I

=
1
R
(C

), where the union is over all and I

sat-
isfying Proposition 14.4.1.
Proof. If a is in the union, then P
a
is nonempty, so that
R
(a) C

by the proof
of Proposition 14.3.5. Conversely, suppose that a
1
R
(C

) Q
r
. Then some
positive multiple a
1
R
(C

) Z
r
. By Proposition 14.4.1, a gives , I

with
a = (1/)(a)

,I

. Thus

1
R
(C

) Q
r

,I

,I

,
and then
1
R
(C

,I

,I

follows by taking the closure.


716 Chapter 14. Toric GIT and the Secondary Fan
The third lemma concerns the faces of a GKZ cone.
Lemma 14.4.6. Every face of a GKZ cone is again a GKZ cone, and

,I

is a
face of

,I

if and only if renes

and I

.
Proof. We will use the cone (
,I

W from the proof of Proposition 14.4.3. Recall


that (
,I

W
0
, where W
0
W = R
r

max
M
R
is dened by the rst line of
(14.4.3) and (
,I

W
0
is dened by the inequalities from the second line, namely
(14.4.7)
'm

,
i
` a
i
, for
i
/ and i / I

'm

,
i
` a
i
, for i I

.
The idea is to study the faces of (
,I

, which map to faces of

,I

by (14.4.4).
Recall that (a, m

) (
,I

gives a polyhedron P
a
with virtual facets F
i,a
,
normal fan
a
, and index set I
,a
=i [ F
i,a
=. Also note the following:
The m

are the vertices of P


a
by the proof of Lemma 14.4.4.
renes
a
since
a
SF().
I
,a
I

. To see why, take i / I

and pick
max
with
i
. This implies
'm

,
i
` =a
i
, so that m

F
i,a
. Hence i / I
,a
.
Nowsuppose that F _(
,I

is a proper face. Then F is dened by turning some


of the inequalities (14.4.7) into equalities. To keep track of which ones, dene
1
F
=(, i) [
i
/ or i I

, and 'm

,
i
` =a
i
for all (a, m

) F.
We claim that
a
and I
,a
are equal for all (a, m

) Relint(F). For I
,a
, note
that since the m

are the vertices of P


a
, we have
F
i,a
= 'm

,
i
` >a
i
for all
max
i I

and (, i) / 1
F
for all
max
.
You will prove the second equivalence in Exercise 14.4.2. This shows that I
,a
depends only on F when (a, m

) Relint(F).
The argument for
a
will take more work. Recall from 6.1 that convexity
can be detected by looking at walls. We say that 'm

,
i
` a
i
is a (,

,
i
)-wall
inequality if
,


max
,

is a wall,
i

`, i / I

.
Every wall inequality appears among the inequalities in the rst line of (14.4.7).
Furthermore, given (a, m

) ( and a (,

,
i
)-wall inequality, we have
(14.4.8) 'm

,
i
` =a
i
m

= m

.
This follows since

is a wall and 'm

, u` = 0 for all u

.
The face F gives an equivalence relation on
max
as follows. Given cones
,


max
, we say that
F

if either =

or there exists a chain of cones


=
0
,
1
, . . . ,
s1
,
s
=

in
max
and
i
1
, . . . ,
is
such that for 1 j s, we
14.4. The Secondary Fan 717
have a (
j1
,
j
,
i
j
)-wall inequality with (
j1
,
i
j
) 1
F
. Thus
F

means
that and

are connected by a chain of compatible wall inequalities that become


equalities in F. For (a, m

) Relint(F), we will show that


m

= m

.
One direction is easy: if
F

, then (14.4.8) and (


j1
,
i
j
) 1
F
imply that
m

j1
= m

j
for 1 j s, and m

= m

follows immediately. For the other


direction, suppose that m

= m

. Then m

and m

give the same vertex of P


a
,
which implies that and

are contained in the same maximal cone


a
. Then
a suitably chosen line segment connecting interior points of ,

lies in and
meets only walls of . This gives a compatible chain of wall inequalities that are
equalities. Since (a, m

) Relint(F), inequalities indexed by (,


i
) / 1
F
are
strict. Thus the wall inequalities in our chain come from 1
F
, hence
F

.
It follows that each maximal cone of
a
is the union of s contained in an
equivalence class of
F
. This proves that
a
depends only on the face F when
(a, m

) Relint(F). The same is true for I


,a
, so that via (14.4.4), we have
F

a,I
,a
when (a, m

) Relint(F). Thus every face of a GKZ cone is a GKZ cone. From


here, Lemma 14.4.4 makes it straightforward to prove the nal assertion of the
proposition. We leave the details to the reader (Exercise 14.4.2).
The Secondary Fan. The above results show that the set of GKZ cones

GKZ
=

,I

[ , I

as in Proposition 14.4.1
has many nice properties. To see that

GKZ
is a generalized fan, we need one further
observation: a set

of rational polyhedral cones is a generalized fan if and only if
Every face of an element of

lies in

.
The relative interiors of the elements of

are pairwise disjoint.
You will prove this Exercise 14.4.3.
It follows that

GKZ
is a generalized fan: GKZ cones are rational polyhedral by
Proposition 14.4.3, their faces are again GKZ cones by Lemma 14.4.6, and their
relative interiors are pairwise disjoint by Lemma 14.4.4. Note also that [

GKZ
[ =

1
R
(C

) by Lemma 14.4.5.
We know from 6.2 that we can turn a generalized fan into an actual fan by
taking the quotient by its minimal face, which is the minimal face of every cone in
the fan. For

GKZ
, this minimal face is ker(
R
) by Proposition 14.4.3. It follows
that the GKZ cones

,I

,I

/ker(
R
) R
r
/ker(
R
) =

G
R
718 Chapter 14. Toric GIT and the Secondary Fan
form the secondary fan

GKZ
=
,I

[ , I

as in Proposition 14.4.1.
Note that [
GKZ
[ =C

, so the maximal cones of


GKZ
have dimension dim G. The
maximal cones of
GKZ
are the chambers of the secondary fan. Another name for
the secondary fan is the GKZ decomposition, hence the notation
GKZ
.
Here are some properties of the secondary fan.
Theorem 14.4.7.
(a)
GKZ
is a fan in

G
R
with [
GKZ
[ =C

.
(b)

,I

_
,I

if and only if renes

and I

.
(c) C
r
//

GX

when 1 Relint(
,I

).
Remark 14.4.8. Part (c) of the theorem tells us that the GIT quotient is constant
on the relative interiors of the GKZ cones.
Proof. We proved part (a) above, and part (b) follows from Lemma 14.4.6. For
part (c), assume that 1 Relint(
,I

) and write =
a
for a Z
r
. The GIT
quotient C
r
//

G is the toric variety of P


a
by Theorem 14.2.13. However, is the
normal fan of P
a
by Lemma 14.4.4, and C
r
//

GX

follows.
Our approach to the secondary fan is more general than what one nds in the
literature. The papers [141] and [222] assume that the
i
are nonzero and generate
distinct rays in N
R
([141] also assumes that the
i
are primitive). We will study this
special case in 15.1, where we will see that the secondary fan has an especially
nice structure.
Generic Characters. Before giving examples of the secondary fan, we need to
relate GKZ cones to the generic characters introduced in 14.3.
Proposition 14.4.9. If 1 Relint(
,I

), then the following are equivalent:


(a) is generic.
(b) is simplicial and i Cone(
i
) induces a bijection 1, . . . , r`I

(1).
(c)
,I

is a chamber of the secondary fan.


Remark 14.4.10. Using this proposition, one can show that the generic characters
determine the chambers of the secondary fan uniquely (Exercise 14.4.4).
Proof. Write =
a
for a Z
r
. By Theorem 14.3.14, is generic if and only if
P
a
is simple of dimension r dim G and the nonempty virtual facets F
i,a
, i / I

, are
the actual facets of P
a
, with no duplications. Then (a) (b) follows easily since
is the normal fan of P
a
by Lemma 14.4.4.
14.4. The Secondary Fan 719
To prove (b) (c), note that
,I

is a chamber if and only if dim

,I

= r.
Furthermore, the proof of Lemma 14.4.4 shows that
dim

,I

= dim(
,I

= dimW
0
,
where W
0
R
r

max
M
R
is dened by
'm

,
i
` = a
i
, for
i
and i / I

.
Hence it sufces to prove that (b) implies that dimW
0
= r.
Assume (b) and take
max
. Then is simplicial of maximal dimension, so
that generators of its rays form a basis of N
R
. By assumption, the ray generators
can be chosen to be the
i
with i / I

. It follows that for any b R


r
and

max
, the equations
'm

,
i
` =b
i
, for
i
and i / I

have a unique solution m

M
R
. When we do this for all
max
, we see that for
any b R
r
, the equations dening W
0
have a unique solution. Thus dimW
0
= r.
Finally, assume (c). First observe that since (b) involves only and I

, the
equivalence (a) (b) means that if (a) holds for one

1 Relint(
,I

) =
Int(
,I

), then it holds for all such

1. Hence it sufces to nd one generic


character that gives an interior point of the chamber.
Let U = C

Cone(

, where the union is over all

such that
dim Cone(

) < dim G. Then U is open and dense in C

, and


G is generic if
and only if

1 U. Since
,I

is a chamber, its interior must have nonempty


intersection with U. This easily implies the existence of the required

, and the
proof is complete.
We can nally give an example of a secondary fan.
Example 14.4.11. Consider Figure 5 from Example 14.3.7, which we reproduce
as Figure 7 on the next page. As we noted in Example 14.3.16, the non-generic
characters of C

lie on the three rays generated by the


i
s. This shows that the
secondary fan has two chambers. You should compute and I

for each of the six


GKZ cones in the secondary fan, and you should check that part (b) of Proposi-
tion 14.4.9 holds only in the interior of the chambers.
Another secondary fan is Figure 6 in 14.3, and Exercises 14.4.514.4.8 give
more secondary fans. Further examples will be given in Chapter 15.
Degenerate Fans. Recall that a generalized fan in N
R
is called degenerate when
it is not an actual fan. Equivalently,
(14.4.9) is degenerate dim X

< dim N
R
.
It is easy to determine where degenerate fans occur in the GKZ decomposition.
720 Chapter 14. Toric GIT and the Secondary Fan
P

=
P

=
P

=
P

=
P

1
=
3

2
Figure 7. C

and the polytopes P for 1 C

in Example 14.4.11
Proposition 14.4.12. If 1 Relint(
,I

), then
(14.4.10) is degenerate dimC
r
//

G < r dim G.
Furthermore:
(a) If is degenerate, then
,I

is contained in the boundary of C

.
(b) If = (
1
, . . . ,
r
) consists of nonzero vectors, then
C

degenerate

,I

.
Proof. Take 1 Relint(
,I

). Since dim N
R
= r dim G and C
r
//

G X

(Theorem 14.4.7), the equivalence (14.4.10) follows from (14.4.9).


Recall from (c) (e) of Proposition 14.3.6 that dimC
r
//

G has the expected


dimension r dim G when 1 lies in the interior of the cone C

. Hence 1
must lie in the boundary when the dimension of dimC
r
//

Gdrops. Then when part


(a) follows from (14.4.10). Also recall that (c) (e) in Proposition 14.3.6 when
consists of nonzero vectors. In other words, when satises this condition, the
boundary of C

is precisely where the dimension of the GIT quotient drops. Then


part (b) follows from (14.4.10).
Irrelevant Ideals. At the beginning of the section, we motivated the existence of
the secondary fan by noting that there are only nitely many possible irrelevant
ideals (14.4.1). Here we will show that the secondary fan can be described com-
pletely in terms of irrelevant ideals.
Recall that 1 C

gives the polyhedron P

R
r
with virtual facets F
i,
=
P

V(x
i
), and by (14.4.1), we have the irrelevant ideal
B() =

i / I
x
i
[ I 1, . . . , r,

iI
F
i,
=

.
By (14.3.8), the vertices of P

determine the minimal generators of B(). Thus


(14.4.11) B() =

b/ F
i,
x
i
[ b is a vertex of P

.
14.4. The Secondary Fan 721
Example 14.4.13. Figure 8 shows an example of P

and its virtual facets F


i,
. This
picture and (14.4.11) make it easy to compute that B() = 'x
1
x
2
, x
2
x
3
, x
4
`. Note
that this is the polytope from Figure 7 when comes from the vertical axis.
F
4,

F
2,
F
1,
F
3,
P

Figure 8. A polytope P and its virtual facets


Given a GKZ cone
,I

, dene the ideal


(14.4.12) B(, I

) =

i
/ or iI

x
i
[
max

.
The ideals B() and B(, I

) relate to GKZ cones as follows.


Proposition 14.4.14. If 1 C

, then:
(a) 1
,I

if and only if B(, I

) B().
(b) 1 Relint(
,I

) if and only if B(, I

) = B().
(c) is generic if and only if every minimal generator of B() has degree dim G.
Proof. We begin with two easy observations about the ideals B(, I

):
and I

are easy to recover from B(, I

). To see why, observe rst that


I

= i [ B(, I

) 'x
i
` and second that the minimal generators of B(, I

)
determine the maximal cones of by (14.4.12) and Proposition 14.4.1.
B(, I

) B(

, I

) if and only if renes

and I

. We leave the easy


proof to the reader.
Now suppose that 1 Relint(
,I

) and write =
a
for a Z
r
. Then
is the normal fan of P
a
and I

= i [ F
i,a
= by Lemma 14.4.4, and it follows
easily from Proposition 14.4.1 that B(, I

) = B(). The converse follows easily


from the rst bullet above. This proves part (a) of the proposition, and part (b)
follows easily from the second bullet and the description of the face relation given
in Lemma 14.4.6.
For part (c), suppose that is generic. Then Proposition 14.3.14 implies that
every vertex of P

is contained in precisely r dim G virtual facets. By (14.4.11),


the corresponding minimal generator of B() has degree dim G. The converse is
equally straightforward.
722 Chapter 14. Toric GIT and the Secondary Fan
Propositions 14.3.14, 14.4.9 and 14.4.14 give a robust understanding of what
it means for a character to be generic. Proposition 14.4.14 also gives an alternate
description of the face relation in the GKZ fan.
Corollary 14.4.15. Given GKZ cones
,I

and

,I

, we have

,I

_
,I

B(, I

) B(

, I

).
It follows that the ideals B(, I

) completely determine the secondary fan.


Here is what this looks like for a familiar example.
Example 14.4.16. Figure 7 from Example 14.4.11 shows P

and its virtual facets


for ve different s. Computing the ideals B() (going from left to right) gives
'x
2
` 'x
1
x
2
, x
2
x
3
, x
2
x
4
` 'x
1
x
2
, x
2
x
3
, x
4
` 'x
1
x
2
, x
2
x
3
, x
3
x
4
, x
1
x
4
` 'x
1
, x
3
`.
These ve ideals are contained in the ideal of the trivial character, B(1) = '1`.
The inclusion relation of these ideals determines the face relation of the GKZ
fan, so that the chambers correspond to the minimal ideals 'x
1
x
2
, x
2
x
3
, x
2
x
4
` and
'x
1
x
2
, x
2
x
3
, x
3
x
4
, x
1
x
4
`. The minimal generators of these two ideals all have degree
equal to 2. It follows that generic characters give elements in the interiors of the
corresponding cones, exactly as predicted by Proposition 14.4.14.
Exercises for 14.4.
14.4.1. This exercise is concerned with the proof of Proposition 14.4.1.
(a) Let a R
r
and assume that P
a
=. For a (possibly non-rational) vertex v P
a
, dene
the cone C
v
= Cone(P
a
v). Prove that
C

v
= Cone(
i
[ 'v,
i
` =a
i
) = Cone(
i
[ v F
i,a
).
(b) Prove the nal assertion made in the proof of Proposition 14.4.1.
14.4.2. This exercise is concerned with the proof of Lemma 14.4.6.
(a) As in the proof of the lemma, let (a, m

) Relint(F). Prove that 'm

,
i
` >a
i
for
all
max
if and only if i I

and (, i) / 1
F
for all
max
.
(b) Complete the proof of the lemma.
14.4.3. Let

be a collection of polyhedral cones in a nite dimensional vector space W.
Assume that every face of a cone in

lies in

and that the relative interiors of the cones
in

are pairwise disjoint. The goal of this exercise is to prove that

is closed under
intersection. For this purpose, take ,

.
(a) Given u

, let (resp.

) be the minimal face of (resp.

) containing u.
Prove that =

.
(b) Conclude that

is a union
1

with
i

for 1 i .
(c) Show that there is some i such that

=
i
. Hint:

is closed under addition.


Take u
i
Relint(
i
) and apply Lemma 1.2.7 to u
1
+ +u

.
14.4. The Secondary Fan 723
14.4.4. Suppose that

is a fan in

G
R
consisting of strongly convex rational polyhedral
cones. Assume that [

[ =C

and that every generic 1 lies in the interior of a maximal


cone of

. Prove that

=
GKZ
. This justies the uniqueness claimed in Remark 14.4.8.
14.4.5. Use the secondary fan to explain what is happening in Example 14.1.15.
14.4.6. Assume that G (C

)
6
with

G Z
3
and
1
, . . . ,
6
correspond to (2, 1, 1),
(1, 0, 1) Z
3
. Determine the number of chambers in the secondary fan. Hint: Think
about generic characters and draw a picture.
14.4.7. Let G = (t, t
r
u, t, u) [ t, u C

(C

)
4
. In Examples 14.3.7 and 14.4.11, we
drewthe secondary fan when r =1 and showed that for generic characters, the GIT quotient
was either H
1
or P
2
. Draw the secondary fan for general r 2 and determine the generic
GIT quotients. (One will be H
r
, while the other will be a weighted projective plane.) You
should also draw the generalized fan for each cone in the secondary fan.
14.4.8. The projective bundle X

=P(O
P
2 O
P
2 (1)) is a 3-dimensional smooth projective
toric variety. The description given in Example 7.3.5 shows that has ve minimal gen-
erators. Since Pic(X

) Z
2
, the quotient construction of X

uses a subgroup G (C

)
5
with G (C

)
2
. In this situation, the
i
are the minimal generators u

.
(a) Determine the
i
. Hint: In terms of the matrices in the exact sequence (14.2.10), you
know the matrix B. Use this to nd A. Three of the
i
are equal.
(b) Use Lemma 14.2.1 to give an explicit description of the group G.
(c) Determine the secondary fan and compute all possible GIT quotients C
5
//

G as we
vary . Hint: There are ve distinct quotients.
(d) What happens when X

= P(O
P
2 O
P
2 (a)) for a 2? Do you see the similarity to
Exercise 14.4.7?
(e) Generalize part (d) by computing the secondary fan coming from the quotient con-
struction of X

=P(O
P
s O
P
s (a)). Hint: Treat the cases a = 0 and a 1 separately.
Chapter 15
Geometry of the
Secondary Fan
This chapter will continue our study of the GIT quotients C
r
//

G as

G varies.
Recall the key players introduced in Chapter 14:
G (C

)
r
gives = (
1
, . . . ,
r
),
i


G
R
and = (
1
, . . . ,
r
),
i
N
R
. We
also have the cones C

= Cone()

G
R
and C

= Cone() N
R
.
The secondary fan
GKZ
consists of GKZ cones
,I

and has support [


GKZ
[ =
C

. A chamber is a full dimensional GKZ cone.


If 1 Relint(
,I

), then C
r
//

GX

, where X

is semiprojective and
is a generalized fan with [[ =C

. Furthermore, the cones of are generated


by the
i
for i / I

, as described in Proposition 14.4.1.


The goal of this chapter is to understand the structure of the GKZ cones and what
happens to the associated toric varieties as we move around the secondary fan. We
will discuss a variety of topics, including nef cones, moving cones, Gale duality,
triangulations, wall crossings, and the toric minimal model program.
15.1. The Nef and Moving Cones
Before we can study GKZ cones, we rst need to generalize the results of 6.3 and
6.4 concerning the nef and Mori cones of toric varieties.
The Nef Cone of a Toric Variety. Given a generalized fan in N
R
, the vector
space of support functions SF() contains the lattice
SF(, N) = SF() [ ([[ N) Z.
When [[ is convex, we also have CSF() = SF() [ is convex.
725
726 Chapter 15. Geometry of the Secondary Fan
In 6.3 we studied the nef cone Nef(X

) for a fan with full dimensional


convex support. Our results apply to generalized fans as follows.
Theorem 15.1.1. Let be a generalized fan with full dimensional convex support
in N
R
. Then:
(a) There is an exact sequence
0 M SF(, N) Pic(X

) 0.
(b) A Cartier divisor on X

is basepoint free if and only if it is nef.


(c) The nef cone Nef(X

) is contained in Pic(X

)
R
, and its inverse image in SF()
is CSF().
(d) If X

is semiprojective, then Nef(X

) Pic(X

)
R
is full dimensional.
Proof. As explained in 6.2, the generalized fan in N
R
gives a genuine fan in
N
R
, where N = N/(
0
N) and
0
is the minimal cone of . Let M M be the
dual of N N. Since X

= X

, the correspondence between Cartier divisors and


support functions (Theorem 4.2.12) gives an exact sequence
0 M SF(, N) Pic(X

) 0.
To relate this to the generalized fan , note that composition with N
R
N
R
induces an injection SF(, N) SF(, N). This gives the commutative diagram
(ignore the dotted arrow for now):
(15.1.1)
0

M

SF(, N)

Pic(X

)

0
0

M

?

SF(, N)

?

Pic(X

)

0.
However, the vertical injections have isomorphic cokernels, because:
A element of SF(, N) comes from SF(, N) if and only if it vanishes on the
kernel of N
R
N
R
.
Every element of SF(, N) is linear on the kernel of N
R
N
R
since the kernel
is the minimal cone of .
It follows that the dotted arrow exists such that the above diagram commutes and
has exact rows. This proves part (a).
Part (b) follows from Theorem 6.3.12. For part (c), (6.3.2) implies
Nef(X

) N
1
(X

) = Pic(X

)
R
.
In Exercise 15.1.1 you will show that the convexity criteria of Theorem 7.2.2 apply
to generalized fans. This plus part (b) complete the proof of part (c).
If X

is semiprojective and is a fan, then Theorem 7.2.4 and Proposition 7.2.9


imply that CSF() contains a strictly convex support function. Strict convexity is
15.1. The Nef and Moving Cones 727
an open condition, so that CSF() is full dimensional in SF(). Then part (c)
implies that Nef(X

) is full dimensional Pic(X

)
R
, as claimed in part (d).
To complete the proof, we need to consider the case of a generalized fan. Here,
the proof follows easily from (15.1.1) and the previous paragraph.
The Toric Cone Theorem. We can also describe the dual of Nef(X

) in terms of
walls when is a generalized fan with full dimensional convex support. A wall
of such a fan is a facet of full dimensional cones =

. In Exercise 15.1.2
you will show that the orbit closure V() is a complete torus-invariant curve in X

and that all complete torus-invariant curves arise this way.


The toric cone theorem stated in Theorem 6.3.20 generalizes as follows.
Theorem 15.1.2. If is a generalized fan with full dimensional convex support,
then we have
NE(X

) = NE(X

) =

a wall
R
0
[V()].
Furthermore, NE(X

) is strongly convex when X

is semiprojective.
Proof. Let be the genuine fan determined by . Since walls of correspond
bijectively to walls of , Theorem 6.3.20 gives the desired formula for NE(X

).
If X

is semiprojective, then NE(X

) is strongly convex since its dual Nef(X

) is
full dimensional by Theorem 15.1.1.
When X

is semiprojective, NE(X

) is minimally generated by its edges, called


extremal rays. If 1is the extremal ray generated by the class of a curve C, then the
corresponding facet of Nef(X

) is dened by D C = 0 for [D] Nef(X

).
Also, in the situation of Theorem 15.1.2, we have a proper map : X

,
where U

is the afne toric variety associated to [[ as in (7.2.1). Then


N
1
(X

) = N
1
(X

/U

).
Here, N
1
(X

/U

) is generated by irreducible curves lying in bers of , modulo


numerical equivalence, which equals N
1
(X

) since is proper and U

is afne.
Hence we are in what is often called the relative case.
The Simplicial Case. When is simplicial with full dimensional convex support,
we have an exact sequence
(15.1.2) 0 M
R
R
(1)
Pic(X

)
R
0,
and since N
1
(X

) is dual to Pic(X

)
R
under intersection product, taking duals gives
the exact sequence
(15.1.3) 0 N
1
(X

) R
(1)
N
R
0.
728 Chapter 15. Geometry of the Secondary Fan
Thus we can interpret classes in N
1
(X

) as relations among the u

s. In particular,
the class of an irreducible complete curve C X

is represented by the relation


(15.1.4)

(D

C)u

= 0.
This follows from Proposition 6.4.1 and (6.4.2).
Nef Cones and the Secondary Fan. Now consider the secondary fan determined
by and as at the beginning of the chapter. A GKZ cone
,I

gives the toric


variety
C
r
//

GX

for all 1 Relint(


,I

). The nef cone Nef(X

) relates to
,I

as follows.
Proposition 15.1.3. For any GKZ cone
,I

, we have:
(a)
,I

Nef(X

) R
I

0
.
(b) dim
,I

= dim Pic(X

)
R
+[I

[.
Proof. Since GIT quotients are semiprojective, part (b) follows from part (a) by
Theorem 15.1.1. For part (a), recall from Denition 14.4.2 that a
,I

has con-
vex support function
a
CSF(). Then we have an isomorphism of cones

,I

CSF() R
I

0
,
where a
,I

maps to (
a
, (
a
(
i
) + a
i
)
iI

). Since

,I

/M
R

,I

and
CSF()/M
R
Nef(X

) by Theorem 15.1.1, part (a) follows immediately.


The Moving Cone of the Secondary Fan. The nicest case of Proposition 15.1.3 is
when I

=, for here the GKZ cone is


,
Nef(X

). There may be many GKZ


cones with I

=. Let
(15.1.5) Mov
GKZ
=

,
be the union of all GKZ cones with I

=. This union has a nice structure.


Proposition 15.1.4. Mov
GKZ
is a convex polyhedral cone in

G
R
.
Proof. We need only prove convexity. Take
1

1
,
and
2

2
,
and consider
=t
1
+(1t)
2
, t [0, 1].
Write
j
=
R
(a
j
) for a
j
R
r
and set a = ta
1
+(1 t)a
2
. By hypothesis, the
virtual facets F
i,a
1
P
a
1
and F
i,a
2
P
a
2
are nonempty. An easy calculation shows
that
tF
i,a
1
+(1t)F
i,a
2
F
i,a
,
which implies that the virtual facets of P
a
are also nonempty. If Relint(
,I

),
then I

is the set of indices of empty virtual facets of P


a
by Lemma 14.4.4. Hence
I

=, so that Mov
GKZ
.
15.1. The Nef and Moving Cones 729
We call Mov
GKZ
the moving cone of the secondary fan. The name moving
cone will be explained later in the section.
Here is an example where the moving cone is small.
Example 15.1.5. Let G=(t, u, tu) [ t, u C

. One easily computes that



G
R
R
2
with basis
1
,
2
, and
3
=
1
+
2
. Furthermore
1
=
2
= e
1
and
3
=e
1
in N.
The secondary fan is illustrated in Figure 1. Virtual facets are indicated by dotted
lines. Hence the moving cone is the diagonal ray generated by
3
.
P

=
F
1
F
2
F
3
P

=
F
2
F
1
F
3
P

=
F
1
=F
2
F
3
P

=
F
2
F
1
=F
3
P

=
F
1
F
2
=F
3

3
Figure 1. A secondary fan with small moving cone
The Geometric Case. Our next result characterizes when the secondary fan has a
chamber satisfying I

= . We say that = (
1
, . . . ,
r
) is geometric if every
i
is
nonzero and the
i
generate distinct rays in N
R
.
Proposition 15.1.6. The secondary fan has a chamber with I

= if and only if
is geometric.
Proof. Suppose we have a chamber of the form
,
and pick 1 Int(
,
).
Then is generic, so that by Proposition 14.4.9, i Cone(
i
) gives a bijection
1, . . . , r (1) since I

=. It follows immediately that is geometric.


Conversely, assume that is geometric and for each 1 i r, let u
i
be the
minimal generator of Cone(
i
) N. By hypothesis, this gives r primitive elements
of N. Now take any chamber

,I

. By Proposition 14.4.9, we have a bijection


1, . . . , r`I

(1) dened by i Cone(


i
). If I

=, then we are done. If not,


let be the renement of

obtained by successive star subdivisions (as dened


in 11.1) at u
i
for i I

. The fan will depend on the order in which we perform


the star subdivisions, but all such have the following properties:
is simplicial since

is simplicial and the star subdivision of a simplicial fan


is simplicial.
i Cone(
i
) is a bijection 1, . . . , r (1) since the star subdivisions add the
rays Cone(
i
), i I

, to the original fan

. This follows since is geometric.


730 Chapter 15. Geometry of the Secondary Fan
X

is semiprojective since X

is semiprojective and X

is projective,
being a composition of projective morphisms.
It follows that and I

= satisfy the conditions of Proposition 14.4.1 and hence


give the GKZ cone
,
. Furthermore, the rst two bullets and Proposition 14.4.9
imply that this cone is a chamber.
Besides being geometric, a stronger condition is when is primitive geometric,
which means that is geometric and each
i
N is primitive. Here is an example.
Example 15.1.7. Let X

be a projective toric variety. The quotient construction


X

(C
(1)
` Z())//G from 5.1 is a GIT quotient by Example 14.2.14. As
noted in Example 14.3.4,
i
s come from the divisor classes [D

] and the
i
s are
the minimal generators u

. Thus is primitive geometric.


The Pseudoeffective and Moving Cones of a Normal Variety. Let X be a normal
variety. In addition to the nef cone Nef(X), other interesting cones live in N
1
(X),
in particular, the pseudoeffective cone Eff(X) and the moving cone Mov(X). These
cones are related to the nef cone by the inclusions
Nef(X) Mov(X) Eff(X) N
1
(X).
The pseudoeffective cone Eff(X) is easy to dene: it is the closure of the cone
generated by effective R-Cartier divisor classes. Here is one case where the pseu-
doeffective cone is easy to describe.
Lemma 15.1.8. If X

is simplicial and semiprojective, then


Eff(X

) = Eff(X

) = Cone([D

] [ (1)).
Proof. We know that N
1
(X

) = Pic(X

)
R
since has full dimensional convex
support, and every D

is Q-Cartier since X

is simplicial. The classes [D

] are
clearly effective. Conversely, if D is effective, then H
0
(X

, O
X

(D)) = 0. We may
assume that D is torus-invariant, and then Proposition 4.3.2 implies that there is
m M with D+div(
m
) 0. Then [D] = [D+div(
m
)] Cone([D

]).
The moving cone Mov(X) was introduced in [171]. An effective Weil divisor
D 0 on X has a xed component if there is an effective divisor D
0
= 0 such that
every effective divisor D

D satises D

D
0
. This implies that D = E +D
0
,
E 0, where every effective divisor D

D of the form D

= E

+D
0
, E

0. In
other words, as we vary D = E +D
0
in its linear equivalence class, D
0
is xed.
We can formulate this in terms of sheaves as follows. If D D
0
0 and
D
0
= 0, then D
0
is a xed component of D if and only if the natural inclusion of
sheaves O
X
(DD
0
) O
X
(D) induces an isomorphism
(15.1.6) H
0
(X, O
X
(DD
0
)) H
0
(X, O
X
(D))
(Exercise 15.1.3). Here is an example of a divisor with a xed component.
15.1. The Nef and Moving Cones 731
Example 15.1.9. For the Hirzebruch surface H
2
, we have the usual divisors D
1
,
D
2
, D
3
, D
4
such that the classes of D
3
, D
4
generate the nef cone. Let D = D
4
and D

= D
2
+D
4
. Figure 4 from Example 6.1.2 shows that the divisors D and D

have the same polytopes and hence their sheaves have the same global sections by
Proposition 4.3.3. Since D

D
2
= D, the discussion leading up to (15.1.6) implies
that D
2
is a xed component of D

.
A Weil divisor D on X is movable if it doesnt have a xed component, and
then the moving cone of X is dened by
(15.1.7) Mov(X) = Cone([D] [ D is Cartier and movable) N
1
(X).
The Moving Cone in the Primitive Geometric Case. We now explain how the
moving cone Mov
GKZ
dened in (15.1.5) relates to the moving cone (15.1.7).
Theorem 15.1.10. If is primitive geometric, then the moving cone Mov
GKZ
is the
union of the GKZ chambers with I

=. Furthermore, if
,
is a chamber, then:
(a) X

is simplicial and semiprojective, with as minimal generators of (1).


(b) If 1 Relint(
,
), then the GIT quotient
C
r
//

G =

C
r
`Z()

//G X

is precisely the quotient construction described in Theorem 5.1.11.


(c) There is a natural isomorphism Pic(X

)
R


G
R
that takes the cones
Nef(X

) Mov(X

) Eff(X

)
to the cones

,
Mov
GKZ
C

.
Proof. Mov
GKZ
is the union of chambers with I

= by Propositions 15.1.4 and


15.1.6. Now let
,
be a chamber. By Proposition 14.4.9, is simplicial and
i
i
= Cone(
i
) gives a bijection 1, . . . , r (1). Thus (1) =
1
, . . . ,
r
,
and the corresponding divisors will be denoted D
1
, . . . , D
r
. This proves part (a).
Since is primitive, the map : M Z
r
from (14.2.2) is the same as the
corresponding map in the exact sequence
0 M Z
(1)
=Z
r
Cl(X

) 0
from (5.1.1) in Chapter 5. This gives a natural isomorphism Cl(X

)

G, which
shows that G is the group from the quotient construction in 5.1. Furthermore, if
B() is the irrelevant ideal from 5.1, then B() = B(, ) = B(), where the rst
equality follows from Proposition 14.4.14 and the second from (14.4.12). From
here, is it straightforward to complete the proof of part (b).
For part (c), the isomorphism Cl(X

)

G induces Pic(X

)
R


G
R
since X

is
simplicial. This takes Nef(X

) to
,
by Proposition 15.1.3. Since [D
i
]
i
, this
isomorphism takes Eff(X

) to C

by Lemma 15.1.8.
732 Chapter 15. Geometry of the Secondary Fan
Let Mov(X

) Pic(X

)
R
be the cone generated by classes of movable divisors
(remember that every Weil divisor is Q-Cartier since X

is simplicial). We prove
Nef(X

) Mov(X

) as follows. A nef Cartier divisor D=

r
i=1
a
i
D
i
has no base-
points by Theorem 15.1.1, so that its Cartier data m

max
lies in the polyhedron
P
D
by Theorem 7.2.2. This implies that the coefcient of D
i
in D+div(
m
) 0
vanishes when
i
(1), hence D is movable.
We also note that X

contains the toric variety X


(1)
whose fan consists of the
rays in (1), together with the trivial cone 0. Since X
(1)
X

is an open subset
whose complement has codimension 2, X

and X
(1)
have the same Weil divisors
via the restriction D DX
(1)
. They also have the same principal divisors and
same class group. It follows easily that
(15.1.8) X

and X
(1)
have the same movable divisors.
For the remainder of the proof, x a chamber
,
and identify Pic(X

)
R
with

G
R
. If

,
is another chamber, then

(1) =
1
, . . . ,
r
= (1), so that
(15.1.9) X

X
(1)
= X

(1)
X

.
When we identify Pic(X

)
R
and Pic(X

)
R
with

G
R
, these inclusions and (15.1.8)
imply that X

and X

have the same movable divisors, i.e., Mov(X

) =Mov(X

).
Since

,
=Nef(X

) Mov(X

), it follows

,
Mov(X

). This proves that


Mov
GKZ
Mov(X

) since Mov
GKZ
is the union of the chambers

,
.
For the opposite inclusion, take an effective Cartier divisor D =

r
i=1
a
i
D
i
and
assume that [D] / Mov
GKZ
. Then [D] Relint(

,I

) with I

= . We claim that
D
0
=

iI

D
i
is a xed component of D, which will imply [D] / Mov(X

).
By (15.1.6) and Proposition 4.3.3, the claim will follow once we prove that
P
DD
0
M = P
D
M . The inclusion P
DD
0
M P
D
M is trivial since D
0
0.
For the other direction, take any m P
D
M. In the notation of 14.2, P
D
= P
a
for
a = (a
1
, . . . , a
r
), so that the inequality 'm,
i
` a
i
must be strict for i I

since
I

consists of the indices of empty virtual facets. Since 'm,


i
` and a
i
are integers,
it follows that 'm,
i
` (a
i
1), hence m P
DD
0
M. This proves our claim.
We have now shown that Mov
GKZ
= Mov(X

). Since Mov
GKZ
is closed, the
same is true for Mov(X

). Thus Mov
GKZ
= Mov(X

) and part (c) is proved.


The proof of Theorem 15.1.10 shows that Mov(X

) = Mov(X

), and we saw
that Eff(X

) = Eff(X

) in Lemma 15.1.8. Furthemore, these cones are polyhedral,


as is Nef(X

). This is typical of how life is simpler in the toric case. In general,


these cones can be much more complicated.
For a simplicial semiprojective toric variety X

, Theorem 15.1.10 gives a vivid


description of the moving cone of X

. Namely,
Mov(X

) =

Nef(X

),
15.1. The Nef and Moving Cones 733
where the union is over all simplicial fans

in N
R
such that X

is semiprojec-
tive and

(1) = (1). Notice that X

is isomorphic to X

in codimension 1 by
(15.1.9). In 15.3 we will give a careful description of the birational isomorphism
X

.
Here is an easy example of the moving cone.
Example 15.1.11. For the Hirzebruch surface H
2
considered in Example 15.1.9,
we have the secondary fan pictured in Figure 2, where
i
= [D
i
]. We showed in
Example 15.1.9 that D
2
+D
4
has a xed component, so that the class [D
2
+D
4
] is
not in the moving cone. This is also clear since I

= in the left chamber. Thus


the moving cone is the right chamber, which is Nef(H
2
).
P

=
P

=
P

=
P

=
P

=
[D
1
] = [D
3
]
[D
4
] [D
2
]
Figure 2. The secondary fan of H2
In 15.2 we will give a more substantial example of a moving cone when we
study an alternate description of the moving cone and explore the relation between
the secondary fan and triangulations.
In the literature, the nef and moving cones are often dened in the relative
case of a projective morphism f : X S (see [171]). Here, numerical equivalence
is dened using complete curves in X that map to a point under f , and N
1
(X/S)
is dened using this notion of numerical equivalence. When the base S is afne,
every complete curve in X maps to a point, so that the denitions of numerical
equivalence and N
1
(X) given in 6.2 agree with the relative versions. This applies
in particular to a semiprojective toric variety, since such a variety is projective over
an afne toric variety. See also [71, Sec. 1.4].
Exercises for 15.1.
15.1.1. Prove that the convexity criteria of Theorem 7.2.2 apply to generalized fans with
full dimensional convex support.
15.1.2. Let X

be the toric variety of a generalized fan in N


R
.
(a) Prove that there is a bijection O() between cones of and T
N
-orbits in X

such
that dim O() = codim.
(b) Prove that if is a wall, then the orbit closure V() is isomorphic to P
1
.
734 Chapter 15. Geometry of the Secondary Fan
15.1.3. Let D and D
0
be effective Weil divisors on a normal variety X.
(a) Construct a natural inclusion of sheaves O
X
(DD
0
) O
X
(D).
(b) Prove that D
0
is a xed component of D if and only if the map of part (a) induces an
isomorphism of global sections as in (15.1.6).
15.1.4. Consider the exact sequence 0 Z
2
B
Z
4
A
Z
2
0, where
B =

0 1
1 1
1 0
1 1

, A =

1 0 1 1
0 1 2 1

.
As in 14.2, the
i
s are the rows of B and the
i
s are the columns of A.
(a) Determine the group G (C

)
4
.
(b) Compute the secondary fan and, for each chamber, draw the corresponding fan that
gives the GIT quotient in that chamber.
(c) What is the moving cone in this case?
15.1.5. Suppose that = (
1
, . . . ,
r
) consists of vectors in Z
2
. Also assume that is
geometric and C

R
2
is strongly convex with C

= Cone(
1
,
r
).
(a) Use Proposition 14.4.9 to prove that
GKZ
has 2
r2
chambers. Hint: First show that if

,I

is a chamber, then I

2, . . . , r 1 and that is uniquely determined by I

.
(b) Explain why Mov
GKZ
consists of a single chamber.
15.1.6. As in Example 15.1.7, a projective toric variety X

gives a GIT quotient where


consists of the u

, (1). The GKZ cone


,
equals Nef(X

) by Proposition 15.1.3.
Let Dbe a torus-invariant nef divisor on X

and set =[D] Pic(X

). By Theorem14.4.7,
1 Relint(

,
), where renes

. Prove that C
r
//

G X

is the toric variety


X
PD
in Example 14.2.11 and Theorem 6.2.8. See also Example 14.2.14.
15.1.7. Let X

be the toric variety of a generalized fan in N


R
. Parallel to the denition
of semiprojective, we say that X

is semicomplete if the map X

Spec(H
0
(X

, O
X
)) is
proper and X

has a torus xed point. Prove that X

is semicomplete if and only if has


full dimensional convex support. Hint: See the proof of Proposition 7.2.9.
15.1.8. Assume that is geometric. Prove that there is a bijective correspondence between
chambers of secondary fan and simplicial fans such that (1) Cone(
i
) [ 1 i r,
[[ =C

, and X

is semiprojective.
15.2. Gale Duality and Triangulations
In 15.1, we described various cones in C

in terms of divisors on toric varieties.


In this section we will instead focus on the combinatorial structure of these cones.
We begin with the chambers of the secondary fan. Our goal is to show that they can
be constructed in a purely combinational way that makes nice use of Gale duality.
Asubset J 1, . . . , r is a -basis if [J[ =dim Gand the vectors
i
, i J, form
a basis of

G
R
. Then Cone(
J
) = Cone(
i
[ i J) is a full dimensional simplicial
cone in

G
R
. These cones relate nicely to the chambers of
GKZ
.
15.2. Gale Duality and Triangulations 735
To see how this works, x a chamber
,I

. The lattice points in its interior


are all generic by Proposition 14.4.9. On the other hand, for a -basis J, the lattice
points in the boundary of Cone(
J
) are nongeneric by denition. It follows easily
that for any -basis J, we have
(15.2.1) either
,I

Cone(
J
) or Int(
,I

) Cone(
J
) =.
The surprise is that the -bases J that satisfy
,I

Cone(
J
) have a very nice
structure and that the chamber
,I

is uniquely determined by these bases. Here


is the precise statement.
Proposition 15.2.1. Let
,I

be a chamber of the secondary fan. Then:


(a) If
max
, then J

=i [
i
/ or i I

is a -basis.
(b) if J is a -basis, then
,I

Cone(
J
) if and only if J =J

for some
max
.
(c)
,I

max
Cone(
J
).
Proof. Our proof is based on [28, Lem. 4.2]. Since
,I

is a chamber, we know
that is simplicial and that i Cone(
i
) induces a bijection 1, . . . , r`I

(1).
Thus, given a maximal cone
max
, the
i
with i / I

form a basis of N
R
.
By Gale duality (Lemma 14.3.1), the
i
with
i
/ or i I

form a basis of

G
R
,
i.e., J

=i [
i
/ or i I

is a -basis. This proves part (a).


For part (b), we rst prove that if
max
, then

,I

Cone(
J
).
Take 1 Int(
,I

) and let b P

be the vertex corresponding to . This easily


implies that b =

iJ

i
e
i
with
i
0, so that 1 =
R
(b) Cone(
J
). Then
the desired inclusion follows from (15.2.1).
Next assume
,I

Cone(
J
) for a -basis J and take 1 Int(
,I

).
Write 1 =

iJ

i
e
i
with
i
> 0. Then b =

iJ

i
e
i
R
r
is a point of P

with precisely dim G nonzero coordinates. Since is generic, the vertices of P

have the same property by Proposition 14.3.14. It follows that b is a vertex and
hence corresponds to a maximal cone
max
. Since J is the set of indices of
nonzero coordinates of b, it follows easily that J = J

.
Part (b) implies
,I

max
Cone(
J
), and then part (c) will follow once
we prove the opposite inclusion. Take

max
Cone(
J
) and pick a R
r
with =
R
(a). It sufces to prove that a

,I

. This means nding a support


function CSF() such that (
i
) a
i
for all i and (
i
) =a
i
for i / I

.
For
max
, our hypothesis on implies that Cone(
J
), so that we can
write =

iJ

i
with
i
0. Hence there is m

M
R
such that a+
R
(m

) =

iJ

i
e
i
. Thus
(15.2.2)
'm

,
i
` a
i
for all i
'm

,
i
` =a
i
for all i / J

, i.e., for all


i
, i / I

.
736 Chapter 15. Geometry of the Secondary Fan
Now dene
(u) = min
max
'm

, u`, u C

.
This function is clearly convex, and by (15.2.2) it satises
(
i
) a
i
for all i
(
i
) ='m

,
i
` =a
i
for all
i
, i / I

.
The second line implies that SF(), and it follows that has the required
properties. This completes the proof.
Here is an example that will appear several times in this section.
Example 15.2.2. Consider the exact sequence 0 Z
3
B
Z
5
A
Z
2
0, where
B =

2 1 1
2 1 1
1 0 1
2 1 1
2 1 1

, A =

1 2 4 1 0
2 1 4 0 1

.
As in 14.2, the
i
s are the rows of B and the
i
s are the columns of A. The
secondary fan shown in Figure 3 has ve chambers. The shaded chamber
,
is

,
Figure 3. The secondary fan and a selected chamber
,
contained in the cones
Cone(
1
,
5
), Cone(
1
,
3
), Cone(
4
,
5
), Cone(
2
,
5
).
By Proposition 15.2.1, the maximal cones of are the complementary cones
Cone(
2
,
3
,
4
), Cone(
2
,
4
,
5
), Cone(
1
,
2
,
3
), Cone(
1
,
3
,
4
).
15.2. Gale Duality and Triangulations 737
These lie in R
3
, but since the
i
s lie on the plane z = 1, we can intersect with
this plane to get the triangulation of the rectangle Conv() shown in Figure 4.

1
Figure 4. The fan intersected with the plane z =1
Our second example is related to the classication theorem proved in 7.3.
Example 15.2.3. Let X

be a smooth projective toric variety of dimension n whose


fan in N
R
R
n
has n+2 rays, with ray generators u
1
, . . . , u
n+2
. Then we have a
geometric quotient X

(C
n+2
`Z())/G, where G(C

)
2
. Furthermore,

G
R
=
Pic(X

)
R
R
2
. The
i
, 1 i n+2, are the classes of the torus-invariant prime
divisors, and C

= Cone() is the effective cone of X

. The secondary fan renes


C

, and the GKZ cone


,
is the nef cone Nef(X

) (Proposition 15.1.3). Note


also that C

is strongly convex since is complete (Proposition 14.3.10).


Fix an ample class Nef(X

). Since Pic(X

)
R
R
2
, the line through
divides Pic(X

)
R
into two half-planes H
+
and H

. Let
P =i [
i
H
+
, [P[ = r +1
Q =j [
j
H

, [Q[ = s +1,
as illustrated by Figure 5. Take any i P and j Q. Then
i
,
j
form a basis of

B
g
g
g
g
g
g
g
gy
T



s +1 js with
j
H

r +1 is with
i
H
+
. .. .
H
+
H

Figure 5. The partition of induced by


Pic(X

)
R
, so that i, j is a -basis. Note also that Cone(
i
,
j
) since
i
and
738 Chapter 15. Geometry of the Secondary Fan

j
lie on opposite sides of the line determined by . By (15.2.1), it follows that

,
Cone(
i
,
j
). Using Proposition 15.2.1, we conclude that

i, j
= Cone(u
k
[ k = i, j)
is a maximal cone of and that all maximal cones of are of this form. This
proves (7.3.8), which is a key step in the proof of Kleinschmidts classication
theorem of smooth projective toric varieties of Picard number 2 (Theorem 7.3.7).
The theory developed here makes (7.3.8) easy to see.
The cones
J
,
max
, from Proposition 15.2.1 are an example of a bunch
of cones in the terminology of Berchtold and Hausen. Their paper [24] shows that
many aspects of toric geometry can be explained in the language of bunches. For
example, they give a bunch proof of Kleinschmidts classication theorem from
7.3 in [24, Prop. 10.1].
The Moving Cone. Proposition 15.2.1 shows how to represent the chambers of
the secondary fan as intersections of cones generated by certain subsets of . Our
next result, taken from [134], shows that the moving cone Mov
GKZ
has a similar
representation.
Proposition 15.2.4. The moving cone of the secondary fan is the intersection
Mov
GKZ
=
r

i=1
Cone(
1
, . . . ,

i
, . . . ,
r
).
Proof. Take 1 C

. The key observation is that


F
i,
= 1 Cone(
1
, . . . ,

i
, . . . ,
r
).
This follows from F
i,
= P

V(x
i
) and P

=
1
R
(1) R
r
0
. Now suppose
that 1 Relint(
,I

). Since I

=i [ F
i,
= by Lemma 14.4.4, we obtain
I

= 1

r
i=1
Cone(
1
, . . . ,

i
, . . . ,
r
).
Then we are done since Mov
GKZ
is the union of all GKZ cones with I

=.
In the primitive geometric case, Theorem 15.1.10 implies that the moving cone
is Mov(X

) for any chamber


,
. In this case,
i
= [D
i
], and a divisor D =

r
i=1
a
i
D
i
is movable if for each i, we can move it away from D
i
, i.e., we can nd
a linear equivalence D

j,=i
b
j
D
j
0. You should check that this translates
exactly into the condition of Proposition 15.2.4.
Here are two examples to illustrate this proposition.
Example 15.2.5. In Figure 2 from Example 15.1.11, we have
i
= [D
i
]. Looking
at the gure, we see that Cone(
1
,
3
,
4
) is the right chamber and
Cone(
1
,
2
,
3
) = Cone(
1
,
3
,
4
) = Cone(
2
,
3
,
4
) =C

.
15.2. Gale Duality and Triangulations 739
By Proposition 15.2.4, the moving cone is the intersection of these cones, which is
clearly the right chamber. This conrms the result of Example 15.1.11.
Example 15.2.6. In Figure 3 from Example 15.2.2, we see that Cone(
1
,
2
,
4
,
5
)
is the rst quadrant and that Cone(
1
, . . . ,

i
, . . . ,
5
) = R
2
for i = 1, 2, 4, 5. Then
Proposition 15.2.4 implies that the moving cone is the rst quadrant, which by
Figure 3 is the union of three GKZ chambers.
Triangulations. The triangulation in Example 15.2.2 generalizes nicely. Before
giving the general theory, let us explore this example in more detail.
Example 15.2.7. Let and be as in Example 15.2.2, and recall that
1
, . . . ,
5
lie
in the plane z = 1. The convex hull Q

= Conv() is a rectangle. Figures 3 and 4


illustrate the triangulation of Q

obtained from one of the GKZ chambers. Doing


this for every chamber gives the picture shown in Figure 6.

1
Figure 6. The secondary fan and its associated triangulations in Example 15.2.7
Note that each GKZ chamber gives a triangulation of Q

such that the vertices


of the triangles lie in . Some elements of can be omitted provided that we
triangulate all of Q

. The triangulations that use every element of correspond to


the chambers that make up the moving cone.
To generalize this construction, take distinct elements
1
, . . . ,
r
N lying on
an integral afne hyperplane H N
R
with 0 / H. Thus there are m M and
a Z`0 such that 'm,
i
` = a for all i. This gives the lattice polytope
Q

= Conv() = Conv(
1
, . . . ,
r
)
lying in the hyperplane H. We assume that Q

has full dimension in H. Thus Q

has codimension 1 in N
R
and the cone C

= Cone() has full dimension in N


R
.
Note also that C

is strongly convex.
740 Chapter 15. Geometry of the Secondary Fan
A triangulation T of is a collection of simplices satisfying:
Each simplex in T has codimension 1 in N
R
with vertices in .
The intersection of any two simplices in T is a face of each.
The union of the simplices in T is Q

.
Figure 6 shows all triangulations of the vectors = (
1
, . . . ,
5
) given by the rows
of the matrix B in Example 15.2.2. See [80] for an introduction to triangulations.
There is bijective correspondence between triangulations T of and simplicial
fans such that [[ =C

and (1) Cone(


i
) [ 1 i r. The correspondence
is easy to describe:
Given T , the cones of are the cones over the simplices of T and their faces.
Given , the simplices of T are Q

for
max
.
This correspondence, written T = Q

, will be used frequently in the our


discussion of triangulations.
We next dene an especially nice class of triangulations, following [264].
Given nonnegative weights = (
1
, . . . ,
r
) R
r
0
, we get the cone
C
,
= Cone((
1
,
1
), . . . , (
r
,
r
)) N
R
R.
When consists of ve equally spaced points on a line in R
2
, Figure 7 shows C
,
for one choice of the weights . In the gure, the lengths of the dashed lines are

5
0
Figure 7. The cone C, determined by and weights
determined by , and (
4
,
4
) is in the interior of the cone. Projection onto N
R
maps C
,
onto C

.
The lower hull of C
,
consists of all facets of C
,
whose inner normal has a
positive last coordinate. Projecting the facets in the lower hull and their faces gives
a fan

in N
R
such that [

[ =C

and

(1) Cone(
i
) [ 1 i r.
Denition 15.2.8. A triangulation T of is regular if there are weights such
that

is simplicial and T =

.
15.2. Gale Duality and Triangulations 741
To see how this relates to the secondary fan, note that gives an injection
M Z
r
dened by m ('m,
1
`, . . . , 'm,
r
`) as in 14.2. The cokernel of this
map is the character group

G of a subgroup G(C

)
r
. Hence the theory developed
in Chapter 14 applies. In particular, Lemma 14.3.2 implies that C

=

G
R
, so that
the secondary fan is complete in this situation.
Proposition 15.2.9. For as above, the map
,I

T = Q

is a bijection
between chambers of the secondary fan and regular triangulations of .
Proof. If
,I

is a chamber, then is a simplicial fan such that [[ = C

and
(1) Cone(
i
) [ 1 i r. Thus T = Q

is a triangulation of . If another
chamber

,I

maps to the same triangulation T , then

= since the maximal


cones of the fan are the cones over the simplices of T . Then Proposition 14.4.9
implies that I

= I

. Hence the map from chambers to triangulations is injective.


We prove that T = Q

is regular as follows. Take Relint(


,I

) and
pick a =(a
1
, . . . , a
r
) R
r
0
that maps to . By Lemma 14.4.4, the support function

a
is strictly convex with respect to . Furthermore,

a
(
i
) =a
i
for i / I

,
a
(
i
) >a
i
for i I

.
For weights given by a, you will prove in Exercise 15.2.1 that

a
convex the lower hull of C
,a
is the graph of
a

a
strictly convex facets of the lower hull map to maximal cones of .
Hence =
a
, so that T = Q

is a regular triangulation of when


,I

is a
chamber.
It remains to prove that every regular triangulation of arises in this way.
Suppose that T =

. Then R
r
0
implies that Relint(

,I

) for
some GKZ cone

,I

. The associated support function

is strictly convex with


respect to and satises

(
i
) =
i
for i / I

(
i
) >
i
for i I

.
This implies =

. However,
,I

need not be a chamber, even though is


simplicial. Figure 8 on the next page gives an example where I

= 2 and has
maximal cones Cone(
1
,
3
) and Cone(
3
,
4
,
5
). This is easy to xfor any i / I

with Cone(
i
) / (1) (such as i =4 in Figure 8), increase
i
so that

(
i
) >
i
.
When we do this, stays the same but we now have I

= i [ Cone(
i
) / (1).
Since is also simplicial, Proposition 14.4.9 implies that
,I

is a chamber.
The Secondary Polytope. In the situation of Proposition 15.2.9, the secondary fan
is complete. It is natural to ask if the secondary fan is the normal fan a polytope.
Gel

fand, Kapranov and Zelevinsky [112] proved the existence of such a polytope,
called the secondary polytope. Other proofs have been given in [28] and [222]. We
will present a proof from [28] that uses Gale duality.
742 Chapter 15. Geometry of the Secondary Fan

5
0
Figure 8. Weights where is simplicial but ,I

is not a chamber in Proposition 15.2.9


A -basis J 1, . . . , r gives the cone Cone(
J
) = Cone(
i
[ i J) featured
in Proposition 15.2.1. Given a real number > 0, consider the polytope
P
J
= Conv(
i
,
j
[ i J, j / J)

G
R
.
We assume that is chosen sufciently small so that Conv(
J
) = Conv(
i
[ i J)
is a facet of P
J
. Since C

= Cone() =

G
R
, the origin is an interior point of P
J
.
Hence taking cones over proper faces of P
J
gives a complete fan
J
in

G
R
. Our
choice of guarantees that Cone(
J
) is a cone of
J
.
We will use the following notation. Given two complete fans
1
and
2
, the
set of all intersections
1

2
for
i

i
, i = 1, 2, is a fan denoted
1

2
. Thus

2
=
1

2
[
i

i
, i = 1, 2
Any common renement of
1
and
2
also renes
1

2
.
Lemma 15.2.10. As above, let
J
be the fan constructed from the -basis J. Then:
(a) The secondary fan
GKZ
renes
J
.
(b)
GKZ
=

J is a -basis

J
.
Proof. Extending Denition 14.3.13, we say that C

is generic if / Cone(

)
for all subsets

with dim Cone(

) < dim G.
Faces of P
J
are of the form Conv(
i
,
j
[ i A, j / B) for suitable sets A and
B. It follows that the cones of
J
are generated by subsets of . In particular, any
cone of
J
of dimension < dim G consists entirely of nongeneric elements.
Take any chamber
,I

. The interior points of


,I

are generic (this follows


easily from the proof of Proposition 14.4.9), so that Int(
,I

) is disjoint from
any cone of
J
of dimension < dim G. This implies that
,I

is contained in a
maximal cone of
J
, and part (a) follows without difculty.
15.2. Gale Duality and Triangulations 743
For part (b), rst note that by part (a),
GKZ
renes

J is a -basis

J
. Then
observe that every maximal cone of
GKZ
appears in

J is a -basis

J
by Proposi-
tion 15.2.1. It follows that the two fans must be equal, which completes the proof
of part (b).
We now construct the secondary polytope.
Proposition 15.2.11. There is a full dimensional lattice polytope P
GKZ

R
whose
normal fan in

G
R
is the secondary fan
GKZ
.
Proof. Let J be a -basis. Since P
J


G
R
contains the origin as an interior point,
we have the dual polytope P

J


G

R
dened in 2.2, and
J
is the normal fan of
P

J
by Exercise 2.3.4. Recall from Proposition 6.2.13 that if polytopes P
1
, P
2
have
normal fans
P
1
,
P
2
, then the Minkowski sum P
1
+P
2
has normal fan
P
1

P
2
.
Now consider the Minkowski sum
P
GKZ
=

J is a -basis
P

J


G

R
.
It follows that P
GKZ
has normal fan

J is a -basis

J
, which by Lemma 15.2.10 is
the secondary fan.
It is customary to call P
GKZ
the secondary polytope even though it is far
from unique. A specic choice for P
GKZ
, based on a different but quite elegant
construction, is given in [113] and [112] (see also [28]).
Since the vertices of a polytope correspond to maximal cones of the normal
fan, Propositions 15.2.9 and 15.2.11 have the following nice corollary.
Corollary 15.2.12. There is a bijective correspondence between vertices of the
secondary polytope P
GKZ

R
and regular triangulations of .
The secondary polytope P
GKZ
has dimension dim G. For = (
1
, . . . ,
r
), we
can compute this dimension as follows. First, the convex hull Q

= Conv() has
dimension dim Q

= dim N
R
1 by assumption. Since dim G = r dim N
R
, we
get the dimension formula
dim P
GKZ
=[[ dim Q

1.
Here is a famous example of a secondary polytope.
Example 15.2.13. The associahedron is the secondary polytope for a conguration
where N
R
= R
3
and consists of the vertices of a convex r-gon sitting in the
plane z = 1. This polytope has dimension is r 3 and is sometimes denoted /
r3
.
Another name for the associahedron is the Stasheff polytope because of its origins
in algebraic topology.
The number of vertices of /
r3
is the Catalan number C
r2
, where
C
n
=
1
n+1

2n
n

.
744 Chapter 15. Geometry of the Secondary Fan
Besides counting vertices of the associahedron, the Catalan number C
r2
counts
many things, including:
The number of triangulations of a convex r-gon that introduce no new vertices.
The number of rooted binary trees with r 1 leaves.
The number of ways pairs of parentheses can be introduced in a nonassociative
product x
1
x
r1
. For r 1 = 4, one gets
(((x
1
x
2
)x
3
)x
4
, (x
1
(x
2
x
3
))x
4
, (x
1
x
2
)(x
3
x
4
), x
1
((x
2
x
3
)x
4
), x
1
(x
2
(x
3
x
4
))).
See Exercise 15.2.2 and [80, Thms. 1.1.2 and 1.1.3].
The associahedron /
r3
is easy to work out when r =4 or 5 (Exercise 15.2.3).
When r = 6, /
3
has C
4
= 14 vertices. In Figure 9, we show /
3
together with
the triangulation at each vertex. The gure uses Lodays construction [188] of the
associahedron, and the labeling of the vertices in Figure 9 was inspired by [80,
Fig. 1.15]. See Example B.2.3 for more on how to construct this gure.
Figure 9. The associahedron K
3
with triangulations at the vertices
Final comments. There is a lot more to say about the secondary fan. Here are
some highlights:
By Proposition 15.2.11, the secondary fan is the normal fan of the secondary
polytope when lies on an afne hyperplane not passing through the origin in
N
R
. One can ask more generally if the secondary fan is the normal fan of some
polyhedron for arbitrary . By[29], the answer is yes when is geometric.
The secondary polytope is related to Chow polytopes and the Newton polytope
of the principal A-determinant. See [113] and [112].
15.2. Gale Duality and Triangulations 745
The secondary fan is also related to the Gr obner fan dened in 10.3. See [80,
Sec. 9.4] and [264].
The secondary fan is used in mirror symmetry. See [66] and [68].
Chapter 5 of the book [80] is devoted entirely to regular triangulations and the
secondary fan.
Another important topic is what happens to the toric variety X

of a chamber
,I

when we cross a wall in the secondary fan. This will be studied in the next section.
Exercises for 15.2.
15.2.1. Prove the claims about
a
made in the proof of Proposition 15.2.9.
15.2.2. The goal of this exercise is to relate triangulations, binary trees, and parentheses.
For r 3, x an r-gon with vertices labeled 1, . . . , r. Also consider symbols x
1
, . . . , x
r1
,
which we arrange in a diagramwith an empty box before and after each symbol. The boxes
are numbered 1, . . . , r corresponding to the vertices of the r-gon:
(15.2.3)
1 2 3 r 1 r
x
1
x
2
x
r1

A triangulation T of the r-gon is determined by adding certain diagonals. Each diagonal


is oriented so that we go from a smaller vertex to a bigger one. Also, at vertex i, we order
the diagonals containing i according to the label of the other vertex of the diagonal.
(a) Given T , ll in the ith box in the diagram (15.2.3) with open or close parentheses, one
for each diagonal containing i. The parentheses are placed in the same order as the
diagonals containing i, and we use an open parenthesis if the diagonal starts at i and a
close parenthesis if the diagonal ends at i. Prove that this denes a bijection between
triangulations and parenthesis placements in the symbols.
(b) Given a rooted binary tree with leaves x
1
, . . . , x
r1
, each internal node different from
the root is root of a subtree, and we put parentheses around the rightmost and left-
most leaves this subtree. Prove that this gives a bijection between binary trees and
parenthesis placements in the symbols.
Figure 10 shows an example of parts (a) and (b) of the exercise when r = 6. See also [80,
Thm. 1.1.3].
2
5
3
4
1
6
x
1
((x
2
(x
3
x
4
))x
5
)
x
1
x
2
x
3
x
4
x
5
triangulation parentheses binary tree
Figure 10. The correspondences described in Exercise 15.2.2
15.2.3. Consider the associahedron /
r3
from Example 15.2.13.
746 Chapter 15. Geometry of the Secondary Fan
(a) Draw /
1
and /
2
. At each vertex, give the corresponding triangulation of the square
and pentagon, similar to Figure 9.
(b) Consider the top pentagonal facet of /
3
in Figure 9. All triangulations lying in this
facet contain a common diagonal that forms a triangle with two edges of the hexagon.
Use this and part (a) to explain why this facet is a pentagon. Then explain why /
3
has
six such facets.
(c) Pick one of the four-sided facets of Figure 9. Show that all triangulations lying in this
facet contain a common diagonal that bisects the hexagon. Use this to explain why the
facet is a quadrilateral and why there are three such facets.
15.2.4. The secondary fan computed in Exercise 15.1.4 has four chambers. Apply the
method of Example 15.2.2 to each of these chambers and determine the maximal cones of
the corresponding fan.
15.2.5. Consider the cone =Cone(e
1
, e
1
+de
2
) R
2
and let be the smooth renement
whose ray generators are
i
=e
1
+ie
2
for 0 i d. We assume d 2. By Exercise 15.1.5,
the secondary fan has 2
d1
chambers corresponding to subsets I

1, . . . , d 1. This
exercise will study the chambers where I

= and I

=1, . . . , d 1.
(a) The
i
s are the rows of the (d +1) 2 matrix
B =

1 0
1 1
.
.
.
.
.
.
1 d

that ts into an exact sequence 0 Z


2
B
Z
d+1
A
Z
d1
0. Show that A can be
chosen to be the (d 1) (d +1) matrix
A =

1 2 1 0 0
0 1 2 1 0
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
0 0 1 2 1

.
Then the
i
s are the columns of A.
(b) Suppose that I

=1, . . . , d 1. Show that U

is the toric variety of this chamber and


use Proposition 15.2.1 to show that the chamber is the simplicial cone generated by
the columns 2, 3, . . . , d 1 of A.
(c) Suppose that I

= . Show that X

is the toric variety of this chamber. Proposi-


tion 15.2.1 represents this chamber as the intersection of d simplicial cones. The next
parts of the exercise will show that this chamber is R
d1
0
.
(d) Show that the chamber is Nef(X

) and show that D =

d
i=0
a
i
D
i
is nef if and only if
a
i1
2a
i
+a
i+1
0 for i = 1, . . . , d 1. Hint: Use Example 6.4.7.
(e) Show Nef(X

) =R
d1
0
. Hint: [D] =

d
i=0
a
i

i
, where
i
is the ith column of A.
This exercise was inspired by [64, App. A], where the toric variety of the secondary fan
has a nice moduli interpretation.
15.2.6. The denition of regular triangulation involved choosing a weight vector in R
r
0
.
15.3. Crossing a Wall 747
(a) Dene what it means for a weight vector to be generic. Hint: The weights used in
Figure 7 are generic while those in Figure 8 are not.
(b) Explain how the denition given in part (a) relates to denition of generic given in the
proof of Lemma 15.2.10.
15.2.7. Let = (
1
, . . . ,
6
) be the rows of the matrix
B =

0 0 0 1
1 0 0 1
0 1 0 1
0 0 1 1
1 0 1 1
0 1 1 1

.
(a) Show that Q

= Conv() is a 3-dimensional prism in the hyperplane H R


4
where
the last coordinate is 1. The secondary polytope has dimension r dim Q

1 = 2.
(b) Show that the secondary polytope is a hexagon and each triangulation has 3 simplices.
See [228, Fig. 11] for a splendid picture of the secondary polytope with the corresponding
triangulation of Q

at each vertex.
15.3. Crossing a Wall
The goal of this section is to understand what happens when we cross a wall in the
secondary fan. We begin with some examples.
Examples. In the general case when has elements that vanish or are repeated, it
is possible that very little happens when crossing a wall. Here is a simple example.
Example 15.3.1. In Example 15.1.5 we have
3
=
1
+
2
in

G
R
R
2
, and
1
=

2
= e
1
,
3
= e
1
in N
R
R. Figure 11 shows the secondary fan together with
P

=
F
1
F
2
F
3
P

=
F
2
F
1
F
3
P

=
F
1
=F
2
F
3

3
Figure 11. A boring wall crossing
P

and its virtual facets F


1
, F
2
, F
3
for three choices of . These s were chosen to
illustrate that very little happens when crossing the wall between the two chambers:
748 Chapter 15. Geometry of the Secondary Fan
the GIT quotient is P
1
on either side of the wall and is also P
1
on the wall itself.
The only thing that changes is which virtual facet is empty.
What allows Example 15.3.1 to be so boring is the equality
1
=
2
. For the
rest of this section, we will assume that is geometric, so that the
i
are nonzero
and generate distinct rays in N
R
. We will see that this assumption rules out the
behavior illustrated in Example 15.3.1.
Here are some more interesting wall crossings.
Example 15.3.2. Figure 6 from Example 15.2.7 shows a secondary fan with ve
chambers. Here, consists of ve vectors in R
3
lying in the plane z = 1. Each
triangulation in Figure 6 is the intersection of a fan in R
3
with z = 1.
We will study how these fans change as we go clockwise around the origin,
starting with the fan
0
in the large chamber to the left in in Figure 6. Figure 12

1
Figure 12. The fans to the left and right of the wall generated by 5
shows that as we cross the vertical wall generated by
5
,
0
changes to
1
, which
is the star division of
0
at the point
3
, as indicated by the shading.
Continuing our loop, we cross the wall generated by
1
. This takes us from
1
to the fan
2
shown in Figure 13. Here, the only change is that the shaded wall
in
1
ips to become the shaded wall in
2
in Figure 13. Weve seen this type
of ip beforethe large illustration given in Figure 3 of Example 11.1.12 has fans
(also called
1
and
2
) that are related by the same type of ip.

2
Figure 13. The fans to the left and right of the wall generated by 1
15.3. Crossing a Wall 749
To nish the clockwise loop, we have three more walls to cross in Figure 6:
Crossing the wall generated by
2
is another ip.
Crossing the wall generated by
4
undoes a star subdivision.
Crossing the wall generated by
3
is a ip.
Example 15.3.3. In the situation of Propositions 15.2.9 and 15.2.11, the secondary
fan is the normal fan of the secondary polytope. Thus a wall between two chambers
of
GKZ
corresponds to an edge of the secondary polytope P
GKZ
.
Now look at the associahedron /
3
shown in Figure 9 of Example 15.2.13.
Every edge comes from a ip of the triangulations at the vertices of the edge.
Example 15.3.4. Figure 2 of Example 15.1.11 shows a secondary fan with exactly
two chambers. You should check that the corresponding fans are related by a star
subdivision.
With the exception of Example 15.3.1, the wall crossings in these examples all
come from ips or star subdivisions. We will soon see that this is no accident.
For the rest of this section we assume that is geometric.
Facets of GKZ Cones. Before we can understand wall crossings, we need to learn
more about the facets of a GKZ cone. Recall from Proposition 15.1.3 that

,I

Nef(X

) R
I

0
.
Using this and the face relation described in Theorem 14.4.7, one easily obtains
the following result (Exercise 15.3.1).
Lemma 15.3.5. The facets of
,I

have one of two forms:


(a)

,I

(a facet of Nef(X

)) R
I

0
, where renes

, or
(b)
,I

\|i
Nef(X

) R
I

\|i
0
, where i I

.
Part (b) of the lemma describes exactly what the facets look like, while part (a)
does not. Later in the section we will use the toric cone theorem to say more the
facets from part (a).
Note also that if
,I

is a chamber of the secondary fan, then its facets are


either walls of the secondary fan or lie on the boundary of C

. Since is geometric,
Proposition 14.4.12 implies that the latter happens only for facets of the form

,I

where

is degenerate. This can only occur in part (a) of Lemma 15.3.5.


Star Subdivisions. There is one type of wall crossing in the secondary fan that is
very easy to describe.
Theorem 15.3.6. Assume that is geometric and let
,I

be a chamber of the
secondary fan with I

=. If i I

, then:
750 Chapter 15. Geometry of the Secondary Fan
(a)
,I

\|i
is a facet of
,I

and is a wall of the secondary fan.


(b) The chamber on the other side of the wall is

,I

\|i
, where

is the star
subdivision of at the minimal generator of Cone(
i
) N.
(c) The exceptional locus of the toric morphism X

has codimension 1.
Proof.
,I

\|i
is a facet of
,I

by Lemma 15.3.5 and is not contained in the


boundary of C

since is nondegenerate. Part (a) follows. Let

be the star
subdivision of at the generator of Cone(
i
) N, as dened in 11.1. The proof
of Proposition 15.1.6 shows that

is simplicial with X

semiprojective.
Since
,I

is a chamber, j Cone(
j
) is a bijection 1, . . . , r`I

(1) by
Proposition 14.4.9. This extends to a bijection 1, . . . , r`(I

`i)

(1) since
is geometric, and it follows that

,I

\|i
is a chamber by Proposition 14.4.9.
Furthermore,
,I

\|i
is a face of this chamber by Theorem 14.4.7. Thus

,I

_
,I

\|i
_

,I

\|i
.
The cone in the middle has codimension 1 (it is a facet of the chamber
,I

) and
hence must be the wall between the two chambers.
We get a birational toric morphism : X

since

is a renement of
. To describe the exceptional locus, let D
i
X

be the divisor corresponding to


Cone(
i
)

(1) and let be the smallest cone containing


i
. Then (D
i
) =
V() and induces an isomorphism X

`D
i
X

`V() by the denition of star


subdivision. Since codimV() 2, D
i
is the exceptional locus.
Types of Walls. Theorem 15.3.6 describes the walls of the secondary fan that arise
from star subdivisions, which we call divisorial walls because of part (c) of the
proposition. The remaining walls are the intersection of two GKZ chambers that
satisfy part (a) of Lemma 15.3.5. We will denote such a wall by

0
,I

, and the
chambers on either side will be written
(15.3.1)
,I

0
,I

,I

.
These are the only GKZ cones containing

0
,I

since it is a wall. We call

0
,I

a
ipping wall. This terminology will be justied below. Here is a preliminary result
about ipping walls.
Lemma 15.3.7. If

0
,I

is a ipping wall between

0
,I

and

0
,I

, then:
(a)
0
is not simplicial.
(b)
0
(1) = (1) =

(1).
(c)
0
is the coarsest common renement of and

.
(d) The exceptional loci of the toric morphisms X

0
and X

0
have
codimension 2.
15.3. Crossing a Wall 751
Proof. We rst prove part (b). Note
0
(1) (1) since renes
0
. Suppose
that Cone(
i
) (1) `
0
(1). Then

0
,I

0
,I

|i
by Theorem 14.4.7. This
is impossible by (15.3.1), and part (b) follows easily.
For part (a), we have 1, . . . , r`I

(1) by Proposition 14.4.9 since


,I

is a chamber. If
0
were simplicial, then
0
(1) = (1) and Proposition 14.4.9
would imply that

0
,I

is a chamber, which is clearly impossible. Hence


0
is not
simplicial.
Part (c) is straightforward and is left to the reader (Exercise 15.3.2).
Since is a renement of
0
satsifying
0
(1) = (1), the birational toric
morphism X

0
is an isomorphism in codimension 1 (see (15.1.9)). Hence
the exceptional locus must have codimension 2. The same holds for X

0
,
and part (d) follows.
Later we will give a careful description of the exceptional loci that occur in
part (d) of Lemma 15.3.7. The terms divisorial wall and ipping wall are
taken from the minimal model program. We will say more about this in 15.4.
The Geometry of Circuits. Our next task is to study a ipping wall given by
(15.3.1). The starting point is that

0
,I

_
,I

comes from a facet of the nef


cone Nef(X

) by Lemma 15.3.5. This corresponds to an edge of the Mori cone


NE(X

) N
1
(X

), which by the toric cone theorem is generated by the class of an


orbit closure V() of a wall =

of .
Since is simplicial, we can describe using the notation of Figure 17 from
6.4. Let n = dim N
R
. After renumbering the
i
s, we can assume that
(15.3.2)
= Cone(
1
, . . . ,
n
)

= Cone(
2
, . . . ,
n+1
)
= Cone(
2
, . . . ,
n
).
The vectors
1
, . . . ,
n+1
are linearly dependent with a nontrivial linear relation
(15.3.3)
n+1

i=1
b
i

i
= 0
that is unique up to multiplication by a nonzero constant. This relation played an
important role in 6.4, where it was called a wall relation (see (6.4.4)).
Here we will consider the more general situation where
1
, . . . ,
n+1
N Z
n
are nonzero and span N
R
. A result of Reid [236] describes some simplicial fans
associated to
1
, . . . ,
n+1
. Our treatment is based on [80, Lem. 2.4.2]. Using the
relation (15.3.3), we dene the sets
J

=i [ b
i
< 0, J
0
=i [ b
i
= 0, J
+
=i [ b
i
> 0.
The vectors
i
for i J

J
+
form a circuit since they are linearly dependent but
every proper subset is linearly independent. Its orientation is determined by J

752 Chapter 15. Geometry of the Secondary Fan


and J
+
. Multiplying (15.3.3) by a negative number switches J

and J
+
and hence
changes the orientation. Then dene the following two sets of simplicial cones
(15.3.4)

= [ _Cone(
1
, . . . ,
i
, . . . ,
n+1
), i J
+

+
= [ _Cone(
1
, . . . ,
i
, . . . ,
n+1
), i J

.
Let us give some examples of

and
+
.
Example 15.3.8. The rows of the matrix B from Example 15.2.2 are the points

1
= (2, 1, 1),
2
= (2, 1, 1),
3
= (1, 0, 1),
4
= (2, 1, 1),
5
= (2, 1, 1) in
R
3
. Then:

1
,
2
,
3
,
4
satisfy the relation
1
2
2
+4
3

4
= 0, so that
J

=1, 2, 4, J
0
=, J
+
=3.
Computing

and
+
, we obtain:

has maximal cone Cone(


1
,
2
,
4
)

+
has maximal cones

Cone(
2
,
3
,
4
), Cone(
1
,
3
,
4
),
Cone(
1
,
2
,
3
).
In Figure 12,

is a subfan of
0
while
+
is a subfan of
1
.

2
,
3
,
4
,
5
satisfy the relation 3
2
+4
3
2
4
+
5
= 0. Thus
J

=2, 4, J
0
=, J
+
=3, 5.
Computing
+
and

, we obtain:

has maximal cones Cone(


2
,
4
,
5
), Cone(
2
,
3
,
4
)

+
has maximal cones Cone(
3
,
4
,
5
), Cone(
2
,
3
,
5
).
In Figure 13,

is a subfan of
1
while
+
is a subfan of
2
.
Thus the fan changes observed in Figures 12 and 13 can be explained in terms of
replacing

with
+
.
Here is a 4-dimensional example where things are more complicated.
Example 15.3.9. Consider the ve points in R
4
given by
1
= (0, 1, 0, 1),
2
=
(0, 1, 0, 1),
3
= (1, 0, 1, 1),
4
= (1, 0, 1, 1),
5
= (0, 0, 2, 1). The relation is
3
1
3
2
+2
3
+2
4
+2
5
= 0, so
J

=1, 2, J
0
=, J
+
=3, 4, 5.
These points lie in the hyperplane H R
3
where the last coordinate is 1. If we
slice the fans

and
+
with H, we get Figure 14 on the next page. For the fan

, the three maximal cones spin about the common face corresponding to J

(the thick line segment on the left), and for


+
, the two maximal cones meet along
the common face corresponding to J
+
(the shaded triangle on the right).
15.3. Crossing a Wall 753

+
H

3
Figure 14. The intersections H and + H
In Figure 14, going from

to
+
is a ip where the three maximal cones
of

that meet along Cone(


1
,
2
) are replaced with the two maximal cones of

+
that meet along Cone(
3
,
4
,
5
).
The fans

and
+
dened in (15.3.4) have the following properties.
Lemma 15.3.10. Let
1
, . . . ,
n+1
N Z
n
be nonzero, and assume that
0
=
Cone(
1
, . . . ,
n+1
) is strongly convex of full dimension. Then:
(a) J

and J
+
are nonempty.
(b)

and
+
are simplicial fans with support
0
such that

(1) =
+
(1) =
Cone(
i
) [ 1 i n+1. Furthermore, X

and X

+
are semiprojective.
(c)

and
+
are the only fans with the properties listed in part (b).
Proof. Note that
1
, . . . ,
n+1
span N
R
since
0
has full dimension. We will rst
compute the Gale dual of
1
, . . . ,
n+1
. The relation

n+1
i=1
b
i

i
= 0 gives the exact
sequence
0 M
R

R
R
n+1

R
R 0,
where
R
(m) = ('m,
1
`, . . . , 'm,
n+1
`) and
R
(e
i
) = b
i
. Thus the Gale dual is the
vector of coefcients = (b
1
, . . . , b
n+1
).
Since Cone(
1
, . . . ,
n+1
) is strongly convex and the
i
are nonzero, C

= R
by Lemma 14.3.2. Part (a) follows immediately. Also, C

= R implies that the


secondary fan has two chambers, hence there are precisely two fans satisfying the
conditions of part (b). Thus part (c) will follow as soon as we prove part (b).
First take the GKZ cone R
0
R. We will use Proposition 15.2.1 to determine
the corresponding fan. The -bases are the nonzero b
i
s, and R
0
Cone(b
i
)
if and only if i J
+
. By Proposition 15.2.1, the maximal cones of the fan are
Cone(
1
, . . . ,
i
, . . . ,
n+1
) for i J
+
(you should check this carefully). This gives

, which proves that

is a fan with the required properties. Similarly, applying


Proposition 15.2.1 to R
0
shows that
+
also satises part (b).
The assumption in Lemma 15.3.10 that Cone(
1
, . . . ,
n+1
) is strongly convex
is needed to ensure that we have two fans. You will explore what happens when
strong convexity fails in Exercise 15.3.4.
754 Chapter 15. Geometry of the Secondary Fan
The proof of Lemma 15.3.10 showed that the secondary fan of
1
, . . . ,
n+1
is
the complete fan in R. Thus the origin is a wall between the two chambers. To
describe the resulting wall crossing, we will use the following notation:

J
= Cone(
i
[ i J) for J 1, . . . , n.
Then the fans

and
+
from (15.3.4) can be written as
(15.3.5)

=
J
[ J
+
J

+
=
J
[ J

J
(Exercise 15.3.5). Here is our result.
Lemma 15.3.11. Let
1
, . . . ,
n+1
N Z
n
be nonzero, and assume that
0
=
Cone(
1
, . . . ,
n+1
) is strongly convex of full dimension. Then:
(a) If J

= i, then the origin in the secondary fan of


1
, . . . ,
n+1
is a divisorial
wall. Furthermore,
0
is simplicial,
+
consists of
0
and its faces, and

is
the star subdivision of
0
at the minimal generator of Cone(
i
) N.
(b) If J
+
=i, then part (a) holds with the roles of + and reversed.
(c) If J

and J
+
have at least two elements, then the origin is a ipping wall.
Furthermore,
0
is nonsimplicial, and the renements

and
+
of
0
give a
commutative diagram of surjective toric morphisms
V(
J

)



V(
J
+
)
?
_

0
V(
J

J
+
)
?

such that

is birational with exceptional locus V(


J

). In addition,
dimV(
J

) = [J
0
[ +[J
+
[ 1 = n[J

[
dimV(
J
+
) = [J
0
[ +[J

[ 1 = n[J
+
[
dimV(
J

J
+
) = [J
0
[.
Proof. If J

= i, then
i
Cone(
1
, . . . ,
i
, . . . ,
n+1
), which implies that
0
=
Cone(
1
, . . . ,
i
, . . . ,
n+1
), so
0
is simplicial, and (15.3.4) shows that
0
is the
unique maximal cone of
+
. The relation (15.3.3) implies b
i

i
=

jJ
+
b
j

j
,
where b
i
<0 and b
j
>0 for j J
+
. Thus
J
+
is the minimal cone of
+
containing

i
. Using the denition of star subdivison from 11.1, one can check that

is the
star subdivison of
+
at the generator of Cone(
i
) N (Exercise 15.3.6). Parts (a)
and (b) follow easily.
For part (c), suppose the origin is a divisorial wall. Then some
i
is a nonnega-
tive combination of
j
for j =i. This implies J
+
=i or J

=i, a contradiction.
Hence the origin is a ipping wall and
0
is nonsimplicial by Lemma 15.3.7.
15.3. Crossing a Wall 755
To analyze

0
, rst note that
J

by (15.3.5). By the
Orbit-Cone Correspondence, X

`V(
J

) is the toric variety of the fan


(15.3.6)

[
J

_ =

`
J

[ J

J
=
J
[ J

J, J
+
J,
where the second equality uses (15.3.5). The next step is to describe the faces of

0
. If we set = (
1
, . . . ,
n+1
), then
0
=C

, so that we can use Lemma 14.3.3


to describe its faces. Given J 1, . . . , n +1, the lemma implies that there is a
face _
0
with J =i [
i
if and only if there are a
j
> 0 for j / J such that

j / J
a
j

j
= 0. Using the partition 1, . . . , n+1 = J

J
0
J
+
, you will prove in
Exercise 15.3.6 that
0
has two types of faces:
(15.3.7)

J
for J

J
+
J (all nonsimplicial)

J
for J

J, J
+
J (all simplicial).
The cones in the rst line of (15.3.7) are the ones we need to remove to get the fan
for U

0
`V(
J

J
+
), while the cones in the second line are the cones of (15.3.6).
This proves that

induces an isomorphism
(15.3.8) X

`V(
J

) U

0
`V(
J

J
+
).
Now we compute some dimensions. Since
J

is simplicial, we have
dimV(
J

) = ndim
J

= n[J

[ =[J
0
[ +[J
+
[ 1
since n+1 =[J
+
[ =[J
0
[ +[J

[. The
i
, i J

J
+
, form a circuit, so that
dimV(
J

J
+
) = ndim
J

J
+
= n([J

[ +[J
+
[ 1) =[J
0
[.
Then dimV(
J

) > dimV(
J
+
J

) since [J
+
[ 2. This and (15.3.8) show that
V(
J

) is the exceptional locus of

. Note also that the exceptional locus has


codimension [J

[ 2.
We get the same picture for
+
with the roles of and + reversed. This gives
the diagram in the statement of the lemma and completes the proof of part (c).
The Flipping Theorem. Before stating our main result, we need an example of a
more complicated wall crossing.
Example 15.3.12. Consider the six points in R
4
given by
1
= (0, 1, 0, 1),
2
=
(0, 1, 0, 1),
3
= (1, 0, 1, 1),
4
= (0, 0, 1, 1),
5
= (0, 0, 0, 1),
6
= (1, 0, 0, 1), which
lie in the hyperplane H R
3
where the last coordinate is 1. Figure 15 on the next
page shows the secondary fan in R
2
. There are four chambers, two divisorial walls,
and two ipping walls. In Exercise 15.3.7 you will compute this secondary fan and
the fan of each chamber.
We are most interested in the chambers

1
,
and

2
,
, which meet along the
ipping wall

0
,
in Figure 15. Similar to what we did in Example 15.3.9, we
visualize the fans
1
,
2
, and
0
in R
4
by slicing them with the hyperplane H to
get the picture shown in Figure 16 on the next page.
756 Chapter 15. Geometry of the Secondary Fan
E
T
c

1
,

2
,

0
,

4
,|5

3
,|5
s
s
s s

4
=
6

1
=
2
Figure 15. A secondary fan in R
2
with two ipping walls
To understand Figure 16, rst focus on the left-hand side of the intersections,
which involve
1
,
3
,
4
,
5
,
6
. Using the relation
3
+
4

5
+
6
= 0, we get
J

= 3, 5 and J
+
= 4, 6. Then Lemma 15.3.11 explains the left half of the
fans we see in Figure 16.

1
H

2
H

0
H

6
Figure 16. The intersections 1 H, 2 H, and 0 H
Now shift your focus to the right-hand side of the intersections in the gure,
which involve
2
,
3
,
4
,
5
,
6
. Lemma 15.3.11 still applies, which explains the
right half of each fan. The key observation is that both sides use the same relation

3
+
4

5
+
6
=0. In other words, the wall crossing is completely determined
by the single oriented circuit
3
+
4

5
+
6
= 0.
We now return to the general situation of two chambers in the secondary fan
separated by a ipping wall, which by (15.3.1) is written

,I

0
,I

,I

.
Recall that
0
is not simplicial by Lemma 15.3.7.
15.3. Crossing a Wall 757
Similar to what we did in Lemma 15.3.11, given = (
1
, . . . ,
r
), we set

J
= Cone(
i
[ i J) for J 1, . . . , r`I

.
Then we have the following result.
Theorem 15.3.13. Assume is geometric and let

0
,I

be a ipping wall between


chambers
,I

and

,I

of the secondary fan. Then there is an oriented circuit


(15.3.9)

iJ

b
i

i
+

iJ
+
b
i

i
= 0
with J

J
+
1, . . . , r`I

, J

J
+
=, and b
i
< 0 for i J

, b
i
> 0 for i J
+
.
Furthermore:
(a)
J

,
J
+

, and
J

J
+

0
.
(b) X

0
`V(
J

J
+
) is the simplicial locus of X

0
.
(c) The renements and

of
0
give a commutative diagram of surjective toric
morphisms
V(
J

)



V(
J
+
)
?
_

0
V(
J

J
+
)
?

such that and

are birational with exceptional loci V(


J

) and V(
J
+
).
Also, codimV(
J

) =[J

[ 2 and codimV(
J
+
) =[J
+
[ 2.
(d) If
0
(
0
)
max
is nonsimplicial, then there is J
0
1, . . . , r ` I

such that
the disjoint union J

J
0
J
+
has cardinality dim
0
+1 and satises
0
=

J
0
J
+
. Furthermore, [

0
=

and

0
=
+
, so that restricting the
diagram of part (c) to U

0
X

0
gives the diagram of Lemma 15.3.11
Proof. As noted in the discussion leading up to (15.3.3),

0
,I

_
,I

comes
from a facet Nef(X

0
) _ Nef(X

) by Lemma 15.3.5. This facet is dened by an


extremal ray 1of NE(X

) N
1
(X

), which by the toric cone theorem is generated


by the class of V() for a wall of .
Let n = dim N
R
and let
0

0
(n) be nonsimplicial.
Claim 1. If (n1) is any wall that meets the interior of
0
, then the class
[V()] lies in the ray 1.
To prove this, let D =

be a Cartier divisor on X

whose class lies in


the relative interior of Nef(X

0
) _ Nef(X

). Note that the support function of D


is linear on
0
. If =

is a wall of that meets Int(


0
), then the Cartier
data of D satises m

= m

since and

lie in
0
. Hence D V() = 0 by
Proposition 6.3.8. Our hypothesis on D then implies that [V()] 1.
758 Chapter 15. Geometry of the Secondary Fan
To construct the oriented circuit (15.3.9), let =

be a wall of that
meets the interior of
0
. We use the notation of (15.3.2), where the
i
are renum-
bered so that is generated by
2
, . . . ,
n
, and (resp.

) is obtained from by
adding
1
(resp.
n+1
). The linear relation satised by
1
, . . . ,
n+1
can be written
(15.3.10)
n+1

i=1
b
i

i
= 0, b
1
, b
n+1
> 0.
Then (15.3.9) is dened to be the relation

iJ

J
+
b
i

i
= 0, where
J

i 1, . . . , n+1 [ b
i
< 0

, J
+
=

i 1, . . . , n+1 [ b
i
> 0

.
Let D
i
be the divisor corresponding to Cone(
i
) (1) for i 1, . . . , r`I

.
Then we can describe J

and J
+
in terms of the intersections D
i
V() as follows:
(15.3.11)
J

=i [ D
i
V() < 0
J
+
=i [ D
i
V() > 0.
This follows from the intersection formulas of Proposition 6.4.4 since X

is sim-
plicial and b
1
, b
n+1
> 0 in (15.3.10). Here is an observation of Reid [236].
Claim 2. If i J
+
, then
i
= Cone(
1
, . . . ,
i
, . . . ,
n+1
) .
First note that 1, n +1 J
+
since b
1
, b
n+1
> 0 in (15.3.10). For i = 1, n +1,
the claim holds since
1
=

and
n+1
= , both of which lie in . Now take
i = 1, n +1 in J
+
. Then = Cone(
2
, . . . ,
i
, . . . ,
n
) is a facet of and is in the
subfan [

0
of consisting of cones of contained in
0
. Let N()
R
R
2
be the
quotient of N
R
by Span(). By Proposition 3.2.7, V() is the toric variety of
Star() =

N()
R
[ _

0
,
where u N
R
gives u N()
R
. Since [

0
has full dimensional convex support,
the same is true for Star(). Also, b
1

1
+b
i

i
+b
n+1

n+1
= 0 in N()
R
. Since
b
1
, b
i
, b
n+1
> 0, we conclude that Star() is complete.
Consider the ray

of generated by
1
. On one side of
1
we have =
Cone(
1
,
i
) Star() as shown in Figure 17 on the next page. By completeness,
there must be _

0
such that

is on the other side of the wall generated


by
1
. Hence there is = i such that 1 r and

= Cone(
1
,

). Thus

= +Cone(
1
) +Cone(

) = Cone(
1
,
2
, . . . ,
i
, . . . ,
n
,

) .
Note that

= Cone(
1
,
2
, . . . ,
i
, . . . ,
n
) is a wall of [

0
.
If = n+1, then

=
i
and the claim follows. So suppose > n+1. Then
(15.3.12)
D

V(

) > 0 by Proposition 6.4.4 since

+Cone(

)
D

V() = 0 by Proposition 6.4.4 since

/
1
, . . . ,
n+1
.
This is impossible since [V()], [V(

)] 1by Claim 1, and Claim 2 is proved.


Claim 3.
0
= Cone(
1
, . . . ,
n+1
) and [

0
=

.
15.3. Crossing a Wall 759
E e
e
e
e
e
eu
c

s
s
s s

n+1

i
Figure 17. The fan Star() in N()
R
R
2
By Lemma 15.3.10, Cone(
1
, . . . ,
n+1
) =

iJ
+

i

0
, hence
i
[

0
by
Claim 2. Suppose [

0
contains a maximal cone different from the
i
. Since
0
is convex, there must be a wall

=
i

where

0
(n) is different from
the
i
. Writing

+Cone(

), we have D

V(

) > 0 by the rst line of


(15.3.12). Then the second line of (15.3.12) and Claim 1 imply that

is one
of
1
, . . . ,
n+1
. It follows that

Cone(
1
, . . . ,
n+1
). This is impossible since
Cone(
1
, . . . ,
n+1
) =

iJ


i
and

is different from the


i
. Claim 3 follows.
We are now ready to prove the theorem. First note that
0
is nonsimplicial and
hence has a nonsimplicial maximal cone
0
. By Claim 3 and Lemma 15.3.11, we
have
J

= [

0
and
J

J
+
_
0

0
.
Also observe that 1is the extremal ray dening Nef(X

0
) _Nef(X

) since
the nef cones are on opposite sides of the wall dened by 1. Hence, if
0

0
(n) is
nonsimplicial, then walls

of

that meet the interior of


0
satisfy [V(

)] 1,
which is the

version of Claim 1. The other claims modify in similar ways; in


particular, the

version of Claim 3 states that

0
=
+
. Arguing as above, we
get
J
+

. This proves part (a).


For part (b), let

be any nonsimplicial cone of


0
. It is contained in a non-
simplicial maximal cone
0
. The above claims imply that Lemma 15.3.11 applies
to
0
. Then (15.3.7) from the proof of the lemma show that
J

J
+
_

. Thus

0
`


0
[
J

J
+
_

is the subfan of simplicial cones of


0
. It follows
from the Orbit-Cone Correspondence that X

0
`V(
J

J
+
) is the simplicial locus
of X

0
, proving part (b).
For part (d), let
0

0
(n) be nonsimplicial. By Claim 3, we have [

0
=

and

0
=
+
. It follows that over U

0
, the diagram of part (c) becomes the
diagram in Lemma 15.3.11. This proves part (d).
Finally, for (c), rst observe that and

induce isomorphisms
(15.3.13) X

`V(
J

) X

0
`V(
J

J
+
) X

`V(
J
+
).
760 Chapter 15. Geometry of the Secondary Fan
To see why, recall that (1) =
0
(1) =

(1) by Lemma 15.3.7. It follows that


any simplicial cone of
0
is a cone of since renes
0
(Exercise 15.3.8). The
same is true for

, and the isomorphisms (15.3.13) follow.


Hence, to study and

, it sufces to work over U

0
for
0

0
(n) nonsim-
plicial. The key point is that all such
0
use the same relation (15.3.9). Then we
are done by part (c) and Lemma 15.3.11.
The maps and

from Theorem 15.3.13 give a birational map


(15.3.14)
1
: X

called an elementary ip. This birational map is equivariant with respect to the
torus actions on X

and X

and is an isomorphism in codimension 1.


In 15.4 we will say more about the maps and

when we discuss the toric


minimal model program. We will show that they are the extremal contractions
determined by the extremal rays 1NE(X

) and 1NE(X

).
An Application. Any two toric varieties X

and X

of dimension n are birational


since they both contain a copy of (C

)
n
. However, if we require the birational map
to be equivariant and an isomorphism in codimension 1, then a much nicer picture
emerges.
Theorem 15.3.14. Let X

and X

be simplicial semiprojective toric varieties and


let : X

be an equivariant birational map that is an isomorphism in codi-


mension 1. Then is a composition of elementary ips and a toric isomorphism.
Proof. First note that induces an isomorphism : N N

since is equivariant
and birational, and ((1)) =

(1) since it is an isomorphism is codimension 1.


Changing X

by a toric isomorphism, we may assume that

is a fan in N
R
with

(1) = (1) and that is the identity on the torus T


N
.
Now consider the secondary fan where consists of the minimal generators
u

for (1) =

(1). The GKZ cones


,
and

,
are chambers of
GKZ
by Proposition 14.4.9 and lie in the moving cone Mov
GKZ
. This cone is convex by
Proposition 15.1.4, so that we can nd a line segment connecting interior points of

,
and

,
which meets all intermediate walls at points of their relative interior.
This gives chambers

,
,

1
,
, . . . ,

,
,

,
such that consecutive chambers share a common wall. These walls are ipping
walls since I

= on each side of the wall. It follows from Theorem 15.3.13 that


we get a composition of elementary ips

: X

1
X

1
X

.
Thus

is an equivariant birational map that is an isomorphism in codimension 1


and is the identity on T
N
.
15.3. Crossing a Wall 761
The composed birational map =
1
: X

is the identity on T
N
and hence is the identity as a birational map. It follows that the birational maps
and

are equal.
When is primitive and lies in an afne hyperplane not containing the origin,
Theorem 15.3.6, Lemmas 15.3.10 and 15.3.11 and Theorem 15.3.13 can all be
interpreted in terms of triangulations. This is explained nicely in the book [80],
though one caution is that in [80], the term ip applies to both divisorial walls
and ipping walls.
The material in this section is based on [194, Ch. 14], [222], [236] and [269].
Exercises for 15.3.
15.3.1. Prove Lemma 15.3.5. Hint: The facets of a product cone are easy to describe, as
are the facets of R
I

0
.
15.3.2. Prove part (c) of Lemma 15.3.7. Hint: Let

be a common renement of and

. Prove that
e
,I

is a face of
0,I

.
15.3.3. Prove part (d) of Proposition 15.3.10.
15.3.4. Let
1
, . . . ,
n+1
N span N
R
R
n
and assume that
0
=Cone(
1
, . . . ,
n+1
) is not
strongly convex.
(a) Prove that J
+
or J

is empty. For the rest of the exercise, assume that J

=.
(b) Prove that the secondary fan of
1
, . . . ,
n+1
consists has only one chamber.
(c) Conclude that
0
= Cone(
1
, . . . ,
n+1
) has a unique simplicial renement

with
[

[ =
0
and X

semiprojective. Also explain why the maximal cones of

are
given by Cone(
1
, . . . ,
i
, . . . ,
n+1
) for i J
+
.
(d) Prove that
J+
is the minimal face of
0
and that the quotient
0
/
J+
is simplicial.
(e) Draw a picture of what happens when n = 2, J
+
=1, 3, and J

=.
15.3.5. Prove that (15.3.5) follows from (15.3.4).
15.3.6. This exercise will cover some details needed for the proof of Lemma 15.3.11.
(a) In part (a) of the lemma, show carefully that

is the star division of


+
at the
generator of Cone(
i
) N when J

=i.
(b) Prove the description of the faces of
0
given in (15.3.7).
15.3.7. Compute the secondary fan shown in Figure 15.3.12. Also compute the fans

1
, . . . ,
4
mentioned in the gure. Hint: Use Proposition 15.2.1.
15.3.8. Suppose that we have fans
1
and
2
in N
R
such that
1
(1) =
2
(1) and
1
renes

2
. Prove that every simplicial cone of
2
is contained in
1
.
15.3.9. In the situation of Lemma 15.3.11, prove that the following are equivalent:
(a) Cone(
1
, . . . ,
n+1
) is simplicial.
(b)
i
Cone(
1
, . . . ,
i
, . . . ,
n+1
) for some i.
(c) J

or J
+
consists of one element.
762 Chapter 15. Geometry of the Secondary Fan
(d)

and
+
are related by a star subdivision.
(e)

(1) =
+
(1).
15.3.10. In the situation of Lemma 15.3.11, prove that the following are equivalent:
(a) Cone(
1
, . . . ,
n+1
) is not simplicial.
(b) J

and J
+
have at least two elements.
(c)

(1) =
+
(1).
15.3.11. In the situation of Lemma 15.3.11, assume that J

and J
+
have at least two
elements.
(a) Prove that J
+
is the unique primitive collection for

.
(b) Prove that J

is the unique primitive collection for


+
.
Hence the elementary ip X

X
+
interchanges the primitive collections.
15.4. Extremal Contractions and Flips
The minimal model progam, called the MMP, is a rst step toward the birational
classication of projective varieties. In dimension 1, birationally equivalent smooth
projective curves are isomorphic. This fails in dimension 2, where instead smooth
surfaces are successive blowups of smooth minimal surfaces. Minimal surfaces
play a key role in the Enriques-Kodaira classication of smooth projective surfaces.
In dimension 3, two complications arise. First, smooth minimal models need
not exist. Instead, a minimal model is a normal Q-Gorenstein projective variety
with only terminal singularities (Denition 11.4.9) and whose canonical divisor is
nef. The second complication is that constructing the minimal model of a smooth
projective variety may require successive blowdowns of divisors together with ips,
which are a type of birational map not present in dimension 2.
Brief Sketch of the MMP. The goal of the MMP is to show that if X is a Q-factorial
normal projective variety with only terminal singularities, then there is a sequence
of blowdowns of divisors and ips X X

Y such that either


Y is a minimal model, or
There is f : Y Z such that dim Z < dimY and all curves C Y mapping to
a point in Z satisfy K
Y
C < 0. We say that f is a Mori bration.
To apply the MMP to a variety X as above, we can assume that K
X
is not nef. Then
one can show the following:
There is a curve C X such that K
X
C < 0 and C generates an edge of the
Mori cone of X, a so-called extremal ray.
An extremal ray gives an extremal contraction f : X X

.
The contraction f is of one of three types: a Mori bration, a contraction of a
divisor, or a small contraction, where the exceptional locus of f has codimen-
sion 2.
15.4. Extremal Contractions and Flips 763
The type of f dictates the next step in the MMP. When f is a small contraction,
this leads to one of the ips mentioned above. Proving termination of this process
is a very difcult problem in general.
The MMP has been proved in dimension 3 by the work of Kawamata, Koll ar,
Mori, Reid, Shokurov, and others. Recently, minimal models have been proved
to exist for varieties of general type of arbitrary dimension [31]. The MMP is an
active area of research in algebraic geometry. See [179] or [194] for an introduction
to the minimal model program.
The Toric MMP. When we apply the MMP to toric varieties and toric morphisms,
we get the toric minimal model progam. In one sense, this is uninteresting, since
a projective toric variety X

is birationally equivalent to P
n
and its canonical class
K
X

can never be made nef, so the toric MMP will always end with a
Mori bration. Nevertheless, there are good reasons to study the toric MMP:
Many MMP constructions are easy to describe in the toric setting. For instance,
extremal contractions and ips are part of the rich geometry of wall crossings.
There are wonderful toric examples of the individual steps in the MMP.
When we switch from the MMP to the relative or log MMP in 15.5, we will
see that there are nontrivial toric examples of these versions of the MMP.
The goal of this section is to introduce the toric versions of extremal contrac-
tions and ips. Then, in 15.5, we will discuss the toric MMP.
Extremal Contractions. We begin with the following result that shows how an
extremal ray in the Mori cone gives an interesting toric morphism.
Proposition 15.4.1. Let be a simplicial fan in N
R
such that X

is semiprojective
and let 1NE(X

) be an extremal ray. Then there is a generalized fan


0
in N
R
with the following properties:
(a) renes
0
and X

0
is semiprojective.
(b) The toric morphism : X

0
has the property that for every wall of ,
(V()) =pt [V()] 1.
Proof. Consider the secondary fan when consists of the minimal generators u

for (1). Then


,
is chamber isomorphic to Nef(X

) by Proposition 15.1.3,
so that 1 denes a facet of
,
. By Proposition 14.4.6 this facet equals

0
,
where the generalized fan
0
satises part (a) of the proposition.
Let be the support function of a Cartier divisor D on X

whose class lies


in the relative interior of

0
,
. Then is strictly convex with respect to
0
. If
m

(n)
, n = dim N
R
, is the Cartier data of D, then m

= m

if and only if
,

lie in the same cone of


0
. This follows from the proof of Lemma 14.4.6.
764 Chapter 15. Geometry of the Secondary Fan
Now suppose that =

is a wall of . Then
(V()) =pt meets the interior of some
0

0
(n)
m

= m

D V() = 0
[V()] 1.
The rst equivalence is easy and the second follows from the previous paragraph.
The third equivalence follows from Proposition 6.3.8 plus standard arguments (see
the proof of Proposition 6.3.15), and the fourth follows since D lies in the relative
interior of the facet of Nef(X

) dened by 1. Thus
0
satises part (b).
When is complete, one can give an elementary proof of Proposition 15.4.1
using Proposition 6.2.5 and Theorem 6.2.8 (Exercise 15.4.1).
The morphism : X

0
constructed in Proposition 15.4.1 is called an
extremal contraction. It is of one of three types:
Fibering: dim X

0
< dim X

.
Divisorial: is birational and its exceptional locus is a divisor.
Flipping: is birational and its exceptional locus codimension 2.
This follows from Theorem 15.3.6 and Lemma 15.3.7.
Structure of Extremal Contractions. Our next task is to describe the contraction
: X

0
from Proposition 15.4.1. Since is simplicial, the discussion of
(15.1.4) shows that elements of N
1
(X

) come from relations among the u

.
Let 1NE(X

) be an extremal ray as in Proposition 15.4.1 and suppose that


1is generated by a relation

= 0.
We will construct a very simple toric variety X
1
that contains a lot of information
about the extremal contraction . The construction of X
1
will use only (1) and

= 0. The other cones of will play no role.


We begin by dening the sets
(15.4.1)
J

= J
1,
= (1) [ b

< 0
J
+
= J
1,+
= (1) [ b

> 0,
and for a subset J (1), let

J
= Cone(u

[ J).
Then we have the sublattices
N

= Span(
J

) N N

J
+
= Span(
J

J
+
) N N,
and we set N
1
= N

J
+
/N

. The map N

J
+
N
1
will be denoted u u.
15.4. Extremal Contractions and Flips 765
Lemma 15.4.2.
1
=
J
[ J J
+
is a complete simplicial fan in (N
1
)
R
whose
toric variety X
1
has the following properties:
(a) X
1
is Q-factorial and Q-Fano.
(b) X
1
has dimension [J
+
[ 1 and Picard number 1.
(c) X
1
is a nite abelian quotient of a weighted projective space.
Furthermore, we always have [J
+
[ 2.
Remark 15.4.3.
(a) A complete normal variety X is Q-Fano if K
X
is Q-Cartier and some posi-
tive integer multiple of K
X
is Cartier and ample. In the literature, Q-Fano
varieties with Picard number 1 are sometimes called unipolar [63].
(b) In the literature, the toric varieties of part (b) are often called fake weighted
projective spaces [60], [167].
Proof. The toric cone theorem tells us that has a wall =

such that [V()]


generates 1. This means that 1 is generated by the wall relation (15.1.4), which
we write as

= 0, b

= D

V().
Then [J
+
[ 2 since the rays (1) and

(1) opposite the common facet


satisfy D

V() >0 and D

V() >0 by Proposition 6.4.4. Also note that the u

for J
+
span N
1
and satisfy

J
+
b

= 0. Furthermore, dim N
1
=[J
+
[ 1
since the u

, J

J
+
, form a circuit.
As in the proof of Lemma 15.3.10, the Gale transform of u

, J
+
, consists of
the positive numbers b

> 0 for J
+
. It follows that R
0
is the unique chamber
of the secondary fan. Each b

is a -basis in the sense of Proposition 15.2.1, so that

J
+
\|
is a maximal cone of the fan corresponding to the chamber. Thus
1
is a
fan and X
1
is projective. Furthermore, since Pic(X
1
)
R
R, any effective divisor
(e.g., K
R
) has a positive integer multiple that is ample. Hence X
1
is Q-Fano.
This proves parts (a) and (b), and then (c) follows from Exercise 5.1.13.
Example 15.4.4. Suppose that is a fan in R
4
with minimal generators u
1
= e
1
,
u
2
= e
2
, u
3
= e
3
, u
4
= e
1
+e
2
+e
3
+2e
4
, u
5
= e
2
+e
3
2e
4
and maximal cones
Cone(u
2
, u
3
, u
4
, u
5
), Cone(u
1
, u
3
, u
4
, u
5
), and Cone(u
1
, u
2
, u
4
, u
5
). There are three
walls, all with the same relation
u
1
+2u
2
+2u
3
u
4
u
5
= 0.
This relation generates an extremal wall 1in NE(X

) where J
+
consists of the rays
generated by u
1
, u
2
, u
3
. Then N
1
Z
2
and u
1
, u
2
, u
3
N
1
generate a sublattice of
index 2 and satisfy u
1
+2u
2
+2u
3
= 0 (Exercise 15.4.2). It follows that
X
1
P(1, 2, 2)/(Z/2Z).
766 Chapter 15. Geometry of the Secondary Fan
Given an extremal ray 1, we have the extremal contraction
: X

0
from Proposition 15.4.1. The behavior of is determined by J

and J
+
from
(15.4.1) and X
1
from Proposition 15.4.2 as follows.
Proposition 15.4.5. Let : X

0
come from the extremal ray 1 with J

, J
+
and X
1
as above, and recall that [J
+
[ 2 by Lemma 15.4.2. Then:
(a) is bering
0
is a degenerate fan J

=. Here, X

0
is simplicial, and
the bers of are isomorphic to X
1
.
(b) is divisorial
0
is a simplicial fan [J

[ = 1. Here, the exceptional


locus V(
J

) is a divisor, and the bers of [


V(
J

)
are isomorphic to X
1
.
(c) is ipping
0
is a nonsimplicial fan [J

[ > 1. Here, the exceptional


locus V(
J

) has codimension [J

[, and the bers of [


V(
J

)
are isomorphic
to X
1
.
Proof. We begin with part (c). Pick a wall of so that [V()] generates 1, which
denes a ipping wall of Nef(X

). Then 1 comes from the relation (15.1.4) for


V(), and J

and J
+
are dened in (15.4.1). On the other hand, Theorem 15.3.13
also uses sets denoted J

and J
+
, which are the same as here by (15.3.11).
This gives the equivalences of part (c). Also, Theorem 15.3.13 implies that
induces an isomorphism
(15.4.2) X

`V(
J

) X

0
`V(
J

J
+
)
and that the exceptional locus V(
J

) has codimension [J
+
[. It follows that [
V(
J

)
is the map
(15.4.3) = [
V(
J

)
: V(
J

) V(
J

J
+
).
In the notation of Proposition 3.2.7, the fan Star(
J

) of V(
J

) lives in N(
J

)
R
=
(N/N

)
R
, and the fan Star(
J

J
+
) of V(
J

J
+
) lives in N(
J

J
+
)
R
, dened
similarly. Using the snake lemma, one easily obtains the exact sequence
0 N
1
N(
J

)

N(
J

J
+
) 0.
We will show that the Star(
J

) is weakly split by
1
and Star(
J

J
+
) in the
sense of Exercise 3.3.7. This will imply that the bers of are isomorphic to X
1
.
The proof of Theorem 15.3.13 shows that for J (1),

_
J
J
+
J and J

J
0
(1) for some
0

0
.
When we project these into N(
J

)
R
, the elements in J

map to zero. Thus


Star(
J

) =
K
[ J
+
K and K
0
(1) `J

for some
0

0
.
It follows that Star(
J

) has two subfans:


15.4. Extremal Contractions and Flips 767
The cones with K J
+
give
1
=
K
[ K J
+
.
The cones with KJ
+
= give

=
K
[ K
0
(1) `(J

J
+
),
0

0
.
It is easy to see that every cone of Star(
J

) can be written uniquely as a sum

K
1
+
K
2
where
K
1

1
and
K
2


. In Exercise 15.4.3 you will show that

R
maps
K


bijectively to
R
(
K
) Star(
J

J
+
) such that
K

R
(
K
)
denes a bijection


Star(
J

J
+
). This gives the desired weak splitting, and
part (c) follows.
For part (b), the equivalences follow easily. If J

= , then the results of


15.3 imply that /
0
(1) and that is the star subdivision of
0
at u

. A relation

= 0 generating 1can be chosen so that b

=1, hence
u

J
+
b

, b

> 0.
Thus
J
+
is the minimal cone of
0
containing u

, and
J

J
+
=
J
+
. It follows
that as in part (c), induces an isomorphism (15.4.2) and [

is (15.4.3). Then
the analysis of part (c) shows that the bers of [

are isomorphic to X
1
.
Finally, for part (a), suppose that J

=. Then 1is generated by a relation


(15.4.4)

J
+
b

= 0, b

> 0.
Let
0
be a maximal cone of the degenerate fan
0
. Then there is at least one wall
of whose interior meets
0
. The wall relation of must be (15.4.4). Then
Claims 1 and 2 from the proof of Theorem 15.3.13 remain true in this situation, as
does Claim 3, provided we interpret

as the fan constructed in Exercise 15.3.4.


By Exercise 15.3.4,
J
+
is the minimal face of
0
. Since
0
was an arbitrary
maximal cone of
0
, it follows that
J
+
is the minimal cone of the degenerate fan

0
. Exercise 15.3.4 also implies that
0
/
J
+
is simplicial, so that the genuine fan

0
constructed from
0
is simplicial. Hence X

0
= X

0
is simplicial.
Since J

= and
J
+
is the minimal cone of
0
, we have V(
J

J
+
) =
V(
J
+
) = X

0
, and V(
J

) = X

follows by a similar argument. Hence (15.4.3) is


the map : X

0
, and then the analysis of part (c) shows that the bers of
are isomorphic to X
1
.
Example 15.4.6. Consider the complete fan in R
3
pictured in Figure 18 on the
next page. The minimal generators of the rays
0
, . . . ,
4
are
u
0
= (0, 0, 1), u
1
= (0, 1, 1), u
2
= (1, 0, 1), u
3
= (0, 1, 1), u
4
= (1, 0, 1).
The wall = Cone(u
2
, u
4
) gives an extremal ray 1=R
0
[V()], whose extremal
contraction is the toric morphism X

P
1
induced by projection onto the y-axis
in Figure 18. The bers of are isomorphic to X
1
=P(1, 1, 2). In Exercise 15.4.4
768 Chapter 15. Geometry of the Secondary Fan
z
y
x

2
Figure 18. The fan from Example 15.4.6
you will prove these claims and explain why Figure 19 from Example 6.4.12 is the
secondary fan of this example. See also Example 15.4.12 below.
Birational Transforms. Before discussing ips, we need to study how divisors
move between birationally equivalent varieties. Suppose that f : X X

is a
birational map between normal varieties and let U X is the largest open subset
of X where f is dened. A prime divisor D X gives a divisor on X

dened by
f

D =

f (DU) if codim f (DU) = 1


0 otherwise.
Then the birational transform by f of a Weil divisor D =

i
a
i
D
i
on X is the Weil
divisor on X

dened by
f

D =

i
a
i
f

D
i
.
The birational transform of a Q-Weil divisor on X is dened similarly.
Example 15.4.7. Let D =

be an effective torus-invariant divisor on a


simplicial semiprojective toric variety X

. By Theorem 15.1.10, the u

for
(1) give a secondary fan where Nef(X

) is the chamber
,
and the class of D
lies in the support C

of the secondary fan. Hence there is a GKZ chamber with


[D]

,I

. Under the isomorphism

,I

Nef(X

) R
I

from Proposition 15.1.3, [D] projects to the class [D

] of a nef divisor D

on X

.
Since X

and X

have the same torus T


N
, we get a birational map : X

.
In Exercise 15.4.5 you will show that D

is the birational transform of D by .


15.4. Extremal Contractions and Flips 769
This example shows that an effective divisor D on X

becomes nef after a


suitable birational transform. The toric MMP discussed in 15.5 will explain how
to do this using only elementary ips and divisorial extremal contractions.
One has to be careful with birational transforms since they are sometimes not
well-behaved. For example, Cartier divisors need not be preserved.
Example 15.4.8. Consider the toric varieties X

and X

+
from Example 15.3.9.
The fans

and
+
are pictured in Figure 14, and Lemma 15.3.11 shows that we
have an elementary ip X

+
. In Exercise 15.4.6 you will show that
Cl(X

) = Cl(X

+
) Z(Z/2Z),
where
[

5
i=1
a
i
D
i
] = [

5
i=1
a
i
D

i
] (3a
1
+3a
2
2a
3
2a
4
2a
5
, a
3
+a
4
mod 2).
Furthermore, you will show that the Picard groups do not have the same image in
Z(Z/2Z):
Pic(X

) =Z[2D
5
], [2D
5
] (4, 0 mod 2)
Pic(X

+
) =Z[D
1
], [D
1
] (3, 0 mod 2).
So 2D
5
is Cartier on X

but its birational transform 2D

5
is not Cartier on X

+
.
Another problem with birational transforms is that they are not functorial.
Example 15.4.9. The blowup of a smooth point p of a variety Y of dimension n >1
gives a birational morphism f : X Y whose exceptional locus E is a divisor. Then
f
1
: Y X is also birational. However,
f
1

E = f
1

0 = 0, yet ( f
1
f )

E = E,
since the largest open set where f
1
f is dened is all of X.
Things are much nicer if we work with special types of birational maps.
Lemma 15.4.10. Let f : X X

be a birational map between normal varieties


that is either an isomorphism in codimension 1 or a proper morphism dened on
all of X. Then:
(a) If D
1
D
2
on X, then f

D
1
f

D
2
on X

.
(b) If g : X

is birational of one of the same two types, then g

= (g f )

.
Proof. Part (a) is easy if f is an isomorphism in codimension 1 (Exercise 15.4.7).
So assume that f : X X

is proper and birational. We claim that the birational


transform is the same as the push-forward dened in [107, Sec. 1.4]. Once we
prove this, then part (a) for f will follow from [107, Thm. 1.4].
770 Chapter 15. Geometry of the Secondary Fan
If DX is a prime divisor, then f (D) is closed in X

. The push-forward of D is
(15.4.5) f

D =

[C(D): C( f (D))] f (D) if codim f (D) = 1


0 otherwise,
where [C(D): C( f (D))] is the degree of the eld extension C( f (D)) C(D).
However, f
1
: X

X is dened outside of a set of codimension 2. This


is proved in [131, Lem. V.5.1] when X, X

are projective, but the same proof works


when f : X X

is proper and birational. In particular, f


1
is dened on an open
subset of f (D), so that f [
D
: D f (D) is birational. Hence the degree is 1, which
shows that (15.4.5) agrees with the birational transform dened above.
Part (b) is straightforward and is left to the reader (Exercise 15.4.7).
Note that Lemma 15.4.10 applies to elementary ips and divisorial extremal
contractions. This will be important in our discussion of the toric MMP.
If : X X

is an isomorphism in codimension 1, the birational transform


gives an isomorphism Div(X) Div(X

) that preserves linear equivalence and


hence induces an isomorphism Cl(X) Cl(X

). It follows that if X

and X

are
simplicial semiprojective toric varieties and : X

is equivariant and an
isomorphism in codimension 1, then induces isomorphisms
Pic(X

)
R
Pic(X

)
R
and N
1
(X

) N
1
(X

).
The left isomorphism was used implicitly in 15.1 and 15.3. Also note that a ray
1in N
1
(X

) gives a ray in N
1
(X

), which will be denoted by 1.


Existence of Toric Flips. A ipping contraction is a birational extremal contrac-
tions whose exceptional locus has codimension 2. An important part of the MMP
is proving the existence of ips for ipping contractions. Our previous results make
this easy to do in the toric setting.
Theorem 15.4.11. Let X

be a simplicial semiprojective toric variety and let 1


be an extremal ray that gives a ipping contraction : X

0
. Then there is a
commutative diagram of toric maps
X

0
with the following properties:
(a) The birational map is an elementary ip in the sense of (15.3.14) and in
particular is equivariant and an isomorphism in codimension 1.
(b) X

is simplicial and semiprojective.


(c) 1

=1is an extremal ray for X

.
(d)

is a ipping extremal contraction for 1

.
15.4. Extremal Contractions and Flips 771
Proof. As in the proof of Proposition 15.4.1, gives a chamber
,
Nef(X

)
in the secondary fan. The extremal ray 1 denes a ipping wall

0
,
, which has
a chamber

,
Nef(X

) on the other side. Then the diagram of the theorem


follows from Theorem 15.3.13, and the proof of the theorem shows that

is the
extremal contraction of X

for the extremal ray 1

=1.
In the situation of Theorem 15.4.11, we say that

: X

0
is the ip of
: X

0
. Here is an example.
Example 15.4.12. Consider the toric variety X

from Example 15.4.6. The left


side of Figure 19 shows the secondary fan, whose chambers are the nef cones of
the toric varieties X

and X

. Here,
i
= [D
i
], and

is obtained from by
replacing Cone(u
2
, u
4
) with Cone(u
1
, u
3
) in Figure 18 of Example 15.4.6. The
right side of Figure 19 shows the corresponding Mori cones.
E
T

Nef(X

)
Nef(X

)
s
s
s

2
=
4

1
=
3
E
T
d
d
d
d
d
ds
d
d
d
d
d
d
1
1

=1

NE(X

)
NE(X

Figure 19. Extremal rays Rfor X and R

=Rfor X

The ray generated by


0
on the left side of Figure 19 is a ipping wall between
X

and X

. This ipping wall is dened by the extremal rays 1 for X

and
1

=1for X

, as pictured on the right side of the gure.


Exercises for 15.4.
15.4.1. Here you will give an alternative proof of Proposition 15.4.1.
(a) Assume is complete and let 1 be an extremal ray. Pick a divisor class [D] in the
relative interior of the facet of Nef(X

) corresponding to 1. Use Proposition 6.2.5


and Theorem 6.2.8 to give an elementary proof of Proposition 15.4.1 in this case.
(b) Show that Proposition 6.2.5 and Theorem 6.2.8 apply when has full dimensional
convex support. This gives an elementary proof of Proposition 15.4.1 in general.
(c) Explain how Proposition 15.4.1 relates to to Examples 14.2.11 and 14.2.14.
Theorem 4.1 of [104] gives a yet different approach to Proposition 15.4.1 that avoids fans.
15.4.2. Verify the details of Example 15.4.4.
772 Chapter 15. Geometry of the Secondary Fan
15.4.3. Consider part (c) of Proposition 15.4.5. Prove that
R
maps
K

bijectively to

R
(
K
) Star(
JJ+
) such that
K

R
(
K
) gives a bijection


Star(
JJ+
).
15.4.4. In Example 15.4.6, show that X
R
P(1, 1, 2) and explain how the secondary fan
relates to Figure 19 of Example 6.4.12.
15.4.5. Prove the claim made in Example 15.4.7 that D

D.
15.4.6. Verify the details of Example 15.4.8.
15.4.7. In part (b) of Exercise 15.4.10, there are four cases to consider. When f and g are
both proper birational morphisms, g

= (g f )

follows from [107, Sec. 1.4]. Prove the


remaining three cases.
15.4.8. Let X

be a simplicial semiprojective toric variety. Recall that P (1) is a


primitive collection if P is not contained in (1) for all but any proper subset is.
(a) Show that the denition of primitive relation given in Denition 6.4.10 applies to X

.
This gives an element r(P) N
1
(X

) as explained in (6.4.8)
(b) Prove that r(P) NE(X

). Hint: Use (6.4.9).


(c) Let 1 be a ipping extremal ray. Use Lemma 15.3.11 and Theorem 15.3.13 to show
that the rays generated by the u
i
for i J
+
form a primitive collection whose primitive
relation generates 1.
(d) Prove a similar result for bering and divisorial extremal rays. Hint: See the proof of
Proposition 15.4.5.
(e) Conclude that NE(X

) =

P
R
0
r(P), where the sum is over all primitive collections
P of .
This shows that Theorem 6.4.11 holds for simplicial semiprojective toric varieties.
15.5. The Toric Minimal Model Program
Our discussion of the toric MMP is based on the papers [104] and [236] and the
books [179] and [194]. As mentioned at the beginning of 15.4, the MMP starts
with a Q-factorial variety X with only terminal singularities and tries to make K
X
nef using divisorial extremal contractions and ips. There are also relative, log,
and equivariant versions of the MMP described in [179, Sec. 2.2]. We will see that
all of these are relevant to the toric case.
Before giving the toric MMP, we rst consider the extremal contractions and
ips constructed in 15.4 from the point of view of the MMP.
Toric Flips via Proj. We will use the following terminology:
Given a ray 1 in N
1
(X

) and a Q-Cartier divisor D, we write D 1 to mean


D C for any generator [C] 1. This is well-dened up to a positive constant.
A ray 1is D-negative if D 1< 0
A -ample Cartier divisor was dened in Denition 7.2.5. Then a Q-Cartier
divisor D is -ample if a positive integer multiple is a -ample Cartier divisor.
Here is a different way to think about a toric ip.
15.5. The Toric Minimal Model Program 773
Proposition 15.5.1. Let

: X

0
be the ip of : X

0
for the extremal
ray 1of ipping type as in Theorem 15.4.11. If D is a divisor on X

with birational
transform D

on X

, then
D is -ample D 1> 0 D

is

-ample.
Furthermore, when D is -ample, we have isomorphisms
X

Proj

=0

O
X

(D)

Proj

=0

O
X

(D)

.
Proof. By Theorem 7.2.11, a Cartier divisor on X

is -ample if and only if for


every maximal cone
0

0
, its support function is strictly convex on the fan
[

0
= [
0
. The proof generalizes to Q-Cartier divisors and hence
applies to D since is simplicial.
Assume D 1> 0 and take
0
(
0
)
max
. If
0
, then the support function

D
is clearly strictly convex on [

0
. When
0
/ , let =

be a wall of
that meets the interior of
0
. Then maps V() to a point, so [V()] 1. Hence
D V() > 0.
However, picking u

N as in Proposition 6.3.8 implies


D V() ='m

, u`
where m

, m

are the Cartier data of D for ,

. Since u

, we obtain

D
(u) ='m

, u` <'m

, u`.
Hence
D
is strictly convex on [

0
by Lemma 6.1.13, proving that D is -ample.
The other direction is straightforward, so D is -ample if and only if D 1> 0.
Since

is the contraction associated to the extremal ray 1

, the equivalence
just proved implies that D

> 0 if and only if D

is

-ample. Since 1

=
1, the latter is equivalent to D 1> 0. This gives the desired equivalences.
Now assume that D is -ample on X

. In Exercise 15.5.1 you will give a toric


proof of X

Proj(

=0

O
X

(D)). Furthermore, D

is

-ample on X

, so
X

Proj

=0

O
X

(D

.
However, : X

0
and

: X

0
are isomorphisms in codimension 1.
This means that the divisors D on X

and D

on X

have the same birational


transform D
0
on X

0
. Then it is straightforward to show that
(15.5.1)

O
X

(D) O
X

0
(D
0
)

O
X

(D

)
(Exercise 15.5.1), and X

Proj(

=0

O
X

(D)) follows easily.


774 Chapter 15. Geometry of the Secondary Fan
The Canonical Divisor and Terminal Singularities. Suppose we apply the MMP
to a toric variety X

. Following strategy outlined at the beginning of 15.4, we


start with a K
X

-negative extremal ray 1 (i.e., K


X

1 < 0). If 1 is of ipping


type, then applying Proposition 15.5.1 with D = K
X

gives a nice description of


the resulting ip.
Proposition 15.5.2. Suppose that : X

0
is a ipping contraction for the
extremal ray 1satisfying K
X

1< 0. Then the ip of X

is given by
X

Proj

=0

O
X

(K
X

.
In the MMP, the ip of a ipping extremal contraction f : X Y satisfying
K
X
1 < 0 is constructed via Proj as in Proposition 15.5.2. However, in order to
apply Proj, one has to know that

=0

O
X
(K
X
) is nitely generated. This is
hard in general [31] but easy in the toric context (Exercise 15.5.1). Fujino and Sato
[104, Thm. 4.5] use the latter to give a quick proof of the existence of toric ips.
Another feature of the MMP is that it applies to Q-factorial varieties with only
terminal singularities. A key step is showing that divisorial extremal contractions
and ips preserves these singularities. In the toric case, we prove this as follows.
Proposition 15.5.3. Let : X

0
be the extremal contraction for the extremal
ray 1. If 1is K
X

-negative and X

has only terminal singularities, then:


(a) If is divisorial, then X

0
has only terminal singularities and

K
X

= K
X

0
.
(b) If is ipping with ip : X

, then X

has only terminal singularities


and

K
X

= K
X

.
Proof. We studied terminal singularities in 11.4. Given a simplicial cone N
R
,
recall from Proposition 11.4.12 that U

has only terminal singularities if and only


if the only is terminal, meaning that lattice points of

=Conv(0, u

[ (1))
are its vertices.
Now take a wall =

of such that [V()] generates 1 and set n =


dim N
R
. Similar to (15.3.2), we label the minimal generators of and

so that
= Cone(u
1
, . . . , u
n
)

= Cone(u
2
, . . . , u
n+1
)
= Cone(u
2
, . . . , u
n
).
Then [V()] is represented by the wall relation of the form

n+1
i=1
b
i
u
i
= 0 with
b
1
, b
n+1
> 0. Dene J

and J
+
as usual. Also let
0
= Cone(u
1
, . . . , u
n+1
) and

i
= Cone(u
1
, . . . , u
i
, . . . , u
n+1
) for 1 i n+1.
Since K
X

, it follows from (15.1.2) and (15.1.3) that


(15.5.2) K
X

V() =

n+1
i=1
b
i
(Exercise 15.5.2). Thus K
X

1< 0 is equivalent to

n+1
i=1
b
i
> 0.
15.5. The Toric Minimal Model Program 775
Now suppose that is a ipping contraction. By Theorem 15.3.13, [

0
has
maximal cones
i
for i J
+
, and

0
has maximal cones
i
for i J

. Since X

has only terminal singularities,


i
is terminal when i J
+
. We need to prove that
the same is true for
i
when i J

.
We begin with two consequences of

n+1
i=1
b
i
> 0. First,
(15.5.3) Conv(0, u
1
, . . . , u
n+1
) =

iJ
+
P

i
.
For the nontrivial inclusion, suppose u =

n+1
i=1

i
u
i
with
i
0 and

n+1
i=1

i
1.
Set = min
iJ
+

i
/b
i
. Since

n+1
i=1
b
i
> 0 and u =

n+1
i=1
(
i
b
i
)u
i
, one sees
easily that u P

i
for any i J
+
with =
i
/b
i
. This proves (15.5.3).
The second consequence of

n+1
i=1
b
i
> 0 is
(15.5.4) u
i
/ P

i
for i J

.
To prove (15.5.4), assume otherwise, so that u
i
=

j,=i

j
u
j
with
j
0 and

j,=i

j
1. This implies u
i
+

j,=i

j
u
j
= 0, where the coefcient of u
i
is
negative and the sum of the coefcients is 0. Since i J

, this relation is a
positive multiple of

n+1
i=1
b
i
u
i
= 0. Thus

n+1
i=1
b
i
0, a contradiction.
Using (15.5.3), we obtain
Conv(0, u
1
, . . . , u
n+1
) N =

iJ
+
P

i
N
=

iJ
+
0, u
1
, . . . , u
i
, . . . , u
n+1

=0, u
1
, . . . , u
n+1

since each
i
is terminal. This and (15.5.4) easily imply that
i
is terminal for
i J

. This shows that X

has only terminal singularities. We leave the proof of

K
X

= K
X

to the reader.
Finally, suppose that is a divisorial contraction, so that is a star subdivision
of
0
at a primitive element we will call u
i
. Suppose that u
i

0
(n) and let the
minimal generators of
0
be u
1
, . . . , u
i
, . . . , u
n+1
, where things are labeled so that
i = 1, n+1. Then u
i
=

j,=i

j
u
j
, b
j
0, so that we get a relation

n+1
j=1
b
j
u
j
= 0
with J

= i. Then one checks easily that (15.5.3) and (15.5.4) still hold and
that [

0
=

in the notation of Theorem 15.3.13. The argument of the previous


paragraph shows that
0
is terminal. It is also easy to see that

K
X

= K
X

0
.
Example 15.5.4. Consider the smooth complete fan in R
2
shown in Figure 20
on the next page. The walls
1
, . . . ,
5
give curves C
i
=V(
i
) with Mori cone
NE(X

) = Cone([C
2
], [C
3
], [C
4
]) N
1
(X

) R
3
(Exercise 15.5.3). The resulting extremal rays 1
2
, 1
3
, 1
4
give three divisorial
extremal contractions, as you can see by removing
2
or
3
or
4
. Also,
K
X

C
2
= 0, K
X

C
3
= K
X

C
4
=1
776 Chapter 15. Geometry of the Secondary Fan

E
T

!
4

2

5
Figure 20. A fan with three extremal contractions
(Exercise 15.5.3). By Proposition 15.5.3, the contractions for 1
3
, 1
4
map to
smooth varieties since terminal means smooth in dimension 2 (Theorem 11.4.14).
However, the extremal contraction for 1
2
maps to a singular variety. It still has
canonical singularities since K
X

C
2
=0, as you will prove in Exercise 15.5.4.
The Toric MMP. For a toric variety, it may be impossible to make the canonical
divisor nef. Hence we will modify the MMP for the toric case by replacing K
X

with an arbitrary divisor D. The goal is to make D nef by using elementary ips
and divisorial contractions. We do this as follows.
Procedure 15.5.5. Let X

be simplicial and semiprojective, and let D be a Weil


divisor on X

. Then do the following steps:


(a) If D is nef, then stop.
(b) If D is not nef, then by the toric cone theorem, there is an extremal ray 1with
D 1< 0. Let : X

0
be the corresponding extremal contraction.
(c) If is a bering contraction, then stop.
(d) If is a divisorial contraction, replace X

and D with X

0
and the birational
transform

D. Note that X

0
is simplicial and semiprojective with support
[
0
[ = [[, and as we saw in the proof of Lemma 15.4.10,

D in the push-
forward of D in the sense of [107, Sec. 1.4]. Return to step (a) and continue.
(e) If is a ipping contraction, then we have the ip
X

0
.
Note that X

is simplicial and semiprojective with [

[ =[[. Replace X

and
D with X

and the birational transform

D. Return to step (a) and continue.


15.5. The Toric Minimal Model Program 777
As we run this procedure on X

, the toric varieties that appear all have the


same convex support, namely [[. If U

is the afne toric variety of [[, then there


is a projective morphism X

. The other toric varieties involved also map


projectively to U

, and the elementary ips and extremal contractions that occur


commute with the projective morphisms to U

. Hence:
Procedure 15.5.5 is an example of the relative MMP, since everything is pro-
jective over the base U

.
Procedure 15.5.5 is also an example of the equivariant MMP, since the torus
T
N
acts on all toric varieties that occur and all maps are T
N
-equivariant.
The procedure terminates with a composition : X

of
D-negative divisorial extremal contractions and elementary ips such that either
There is a D-negative bering contraction from X

to a toric variety of smaller


dimension, or

D is nef on X

.
To explain what we mean by D-negative, let X

i
X

i+1
be an elementary ip
that arises from running Procedure 15.5.5 on X and D. Then the previous steps
of the MMP give a birational map
i
: X

i
, and the ip X

i
X

i+1
is
associated to a ipping extremal ray of X

i
that is negative with respect to the
birational transform
i
D. We say that X

i
X

i+1
is a D-negative elementary
ip in this situation, and the terms D-negative divisorial contraction and D-negative
bering contraction have similar meanings. Note that we are making frequent use
of the functoriality proved in Lemma 15.4.10.
Termination of Flips. The big question is whether Procedure 15.5.5 terminates.
The secondary fan makes it easy to choose the extremal rays so that this happens.
Proposition 15.5.6. Let D be a Weil divisor on a simplicial semiprojective toric
variety X

. Then the D-negative extremal rays in Procedure 15.5.5 for X

and D
can be chosen so that the procedure stops after nitely many iterations.
Proof. We use induction on the rank of Pic(X

). The base case is Exercise 15.5.5.


Now assume that Pic(X

) has rank > 1. Consider the secondary fan where


consists of the u

for (1). Then


,
=Nef(X

) is a chamber in the secondary


fan in

G
R
Pic(X

)
R
. Assume [D] / Nef(X

) and draw a line segment between


[D] and a generic point in the interior of Nef(X

) such that the segment always


crosses from one chamber to another at a relative interior point of a wall.
Consider the facet

0
,
where the line segment leaves
,
. This gives an
extremal ray 1 such that D 1< 0, and : X

0
is the extremal contraction
of 1. There are four possibilities to consider:
If

0
,
lies on the boundary the secondary fan, then
0
is degenerate by
Proposition 14.4.12, in which case we are done.
778 Chapter 15. Geometry of the Secondary Fan
If

0
,
is divisorial wall, then we replace X

and D with X

0
and

D. Since
the rank of the Picard drops by one, we are done by induction.
If

0
,
is ipping wall, then we have the ip : X

. If

D is nef on
X

, then we are done by Procedure 15.5.5.


The remaining case is when

0
,
is ipping wall and

D is not nef on X

.
Then our chosen line segment leaves

,
= Nef(X

) at a facet dened by an
extremal ray 1

of X

with (

D) 1

< 0. Then replace X

, D, 1 with X

D, 1

and continue, using the same line segment as before.


The fourth bullet can occur only nitely many times since the line segment meets
only nitely many chambers. Hence the process must terminate.
We will next improve Proposition 15.5.6 by showing that Procedure 15.5.5
terminates no matter which D-negative extremal rays are used in step (b) of the
procedure. Since a divisorial extremal contraction lowers the rank of the Picard
group by 1, only nitely many of these steps can occur. Hence the problem reduces
to the termination of ips. We will need the following lemma.
Lemma 15.5.7. Let : X

be an elementary ip for an extremal ray 1.


Pick a relation

= 0 generating 1and scale it so that


(15.5.5) u
0
=

J
+
b

is a primitive element of N. Then:


(a) We have an equality of star subdivisions

(u
0
) =

0
(u
0
) =

(u
0
).
(b) Let

be the fan of part (a). Then we have a commutative diagram


X

0
.
(c) Let D

D be the birational transform of a divisor D on X

. Then

D =

(D C)D

0
,
where D

0
is the toric divisor on X

corresponding to u
0
and C is a 1-cycle on
X

whose class is represented by the relation

= 0.
Proof. Part (a) is straightforward, and part (b) follows immediately from part (a)
(Exercise 15.5.6). For part (c), suppose that D =

. The fan

renes ,
which gives the proper birational map : X

. Since X

is simplicial, D is
Q-Cartier and hence has a support function
D
. By Proposition 6.2.7,

D and D
have the same support function
D
.
15.5. The Toric Minimal Model Program 779
This makes

D easy to compute. First note that

(1) = (1)
0
, where

0
= Cone(u
0
). Let D

(resp. D

0
) be the divisor on X

associated to (1)
(resp.
0
). Then

D =

(1)

D
(u

)D

D
(u
0
)D

0
=

(1)
a

D
(u
0
)D

0
,
where the second equality uses
D
(u

) = a

for (1). Theorem 15.3.13


implies that
J

= Cone(u

[ J

) . Since
D
is linear on
J

, (15.5.5)
yields
D
(u
0
) =

. Hence

D =

(1)
a

0
.
Furthermore,
J
+

by Theorem 15.3.13, so repeating the above computation


with D

on X

gives

(1)
a

J
+
a

0
.
We conclude that

D =

(1)
a

0
.
If C is a 1-cycle whose class is represented by the relation

= 0, then the
quantity in parentheses is D C by Exercise 15.5.2.
Example 15.5.8. Figure 3 from Example 11.1.12 is a classic example of part (b)
of Lemma 15.5.7. In 11.1 we mentioned that Figure 3 was an example of a ip.
But the ip only concerns the bottom half of the gurethe common subdivision

is what explains the top half.


Here is an immediate consequence of Lemma 15.5.7.
Corollary 15.5.9. Assume that D 1< 0 in the situation of Lemma 15.5.7. Then
there is a nonzero effective divisor D

on X

such that

D =

+D

.
Corollary 15.5.9 is a special case of a result that applies more generally. See,
for example, [104, Lem. 4.10], [179, Lem. 3.38], or [194, Lem. 9-1-3].
We can now prove termination.
Theorem 15.5.10 (Termination of Toric Flips). Given a divisor D
1
on a simplicial
semiprojective toric variety X

1
, there is no innite sequence of elementary ips
X

1
X

2
X

3

for ipping extremal rays 1
i
NE(X

i
) satisfying D
i
1
i
< 0, where for i 2,
D
i
= (
i1

1
)

D
1
.
780 Chapter 15. Geometry of the Secondary Fan
Proof. Assume that such an innite sequence exists. Since
1
(1) =
2
(1) =
and there are only nitely many fans with a given set of rays, the same fan must
occur twice in the sequence. If we start numbering at this fan, then we obtain
X

1
X

1
X

= X

1
.
Furthermore,
1

1
is the identity since it is the identity on T
N
X

1
.
We will show that this is impossible by proving the existence of a commutative
diagram
(15.5.6)
X
b

i
b

i1

1

X

i
such that
(15.5.7)

i
D
1
=

i
D
i
+

D
i
,

D
i
nonzero and effective.
Once this is proved, we get an easy contradiction. This is because when i = , the
diagram (15.5.6) implies that

since
1

1
is the identity. Then

D
1
=

D
1
+

since D

= D
1
. This forces

D

= 0, contradicting (15.5.7).
We prove (15.5.6) and (15.5.7) by induction on i. The base case i = 2 follows
from Corollary 15.5.9. For the inductive step, we will do i = 3 for simplicity.
Consider the commutative diagram
X
b

1

X

2

X

3
,
where the two lower triangles come from Lemma 15.5.7 and

2
is any common
renement of

1
and

2
. Then Corollary 15.5.7 implies that

1
D
1
=

1
D
2
+D

1
and

2
D
2
=

1
D
3
+D

2
,
where D

1
and D

2
are nonzero effective divisors. An easy diagram chase yields

2
D
1
=

2
D
3
+

1
+

2
.
Then we are done since the pullback by surjective map of a nonzero effective divi-
sor is nonzero and effective.
15.5. The Toric Minimal Model Program 781
In the general version of the MMP, termination of ips has been proved only
in dimension 3. However, a special case of termination is proved in [31], which
is sufcient to prove the existence of minimal models of n-dimensional projective
varieties of general type.
The Log Toric MMP. An important technical tool in the MMP is provided by pairs
(X, D), where X is normal and D =

i
a
i
D
i
is a Q-divisor such that a
i
[0, 1] Q.
These log varieties were introduced in 11.4, where we gave the references [179,
p. 98] and [194, Ch. 11] that explain their usefulness.
We also dened log canonical and klt (kawamata log terminal) singularities of
a pair (X, D) in Denition 11.4.23, which are important in the MMP. For a toric
variety X

, we use D=

. When K
X

+D is Q-Cartier, Proposition 11.4.24


tells us that:
If a

[0, 1] for all (1), then (X

, D) is log canonical.
If a

[0, 1) for all (1), then (X

, D) is klt.
The standard MMP for X uses the divisor K
X
and requires that X be Q-factorial
with only terminal singularities. Hence it should not be surprising that the log
MMP for the pair (X, D) uses the divisor K
X
+D and requires that X be Q-factorial
with only klt singularities. In order to run the log MMP, one needs to show that
these properties are preserved by the divisorial contractions and ips associated to
(K
X
+D)-negative extremal rays.
In the toric case, let X

be simplicial and semiprojective, and take a divisor


D=

with a

[0, 1)Q for all . Then K


X

+D is Q-Cartier and (X

, D)
is klt by Proposition 11.4.24. Then the log toric MMP for the pair (X

, D) is
described as follows.
Procedure 15.5.11. Let (X

, D) be as above. Then do the following steps:


(a) If K
X

+D is nef, then stop.


(b) If K
X

+D is not nef, then there is an extremal ray 1 with (K


X

+D) 1< 0.
Let : X

0
be the corresponding extremal contraction.
(c) If is bering, then stop.
(d) If is divisorial, then replace (X

, D) with (X

0
,

D). Return to step (a) and


continue.
(e) If is ipping with ip : X

, then replace (X

, D) with (X

D).
Return to step (a) and continue.
The log toric MMP for (X

, D) is the toric MMP for X and K


X

+D since

(K
X

+D) = K
X

0
+

D in step (d)

(K
X

+D) = K
X

D in step (e)
782 Chapter 15. Geometry of the Secondary Fan
by Proposition 15.5.3. Also note that klt singularities are preserved when we run
the log toric MMP since
(X

0
,

D) in step (d)
(X

D) in step (e)
are klt by Proposition 11.4.24. Finally, Proposition 15.5.6 (careful choice of 1 in
step (b)) and Theorem 15.5.10 (any (K
X

+D)-negative 1 in step (b)) guarantee


that the log toric MMP terminates.
Example 15.5.12. Suppose that 1 is any extremal ray for X

. Let us show that


the associated extremal contraction : X

0
occurs in Procedure 15.5.11 for
the pair (X

, D), provided that D is chosen correctly.


Pick a Cartier divisor D
0
whose class lies in the interior of Nef(X

). Replacing
D
0
with a positive integer multiple, we may assume that the polyhedron P
D
0
has
an interior point m, and then replacing D
0
with D
0
+div(
m
), we may assume that
D
0
=

, where a

> 0 for all . Now consider the Q-divisor


D =K
X

D
0
=

(1a

)D

,
where Q is positive and satises 0 < a

< 1 for all . Then K


X

+D =D
0
,
which makes it easy to see that every extremal ray 1is (K
X

+D)-negative.
Log Minimal Models and Flops. When we run the log toric MMP on (X

, D), the
result is a sequence : X

of (K
X

+D)-negative divisorial
extremal contractions and elementary ips such that either
There is a (K
X

+D)-negative bering extremal contraction from X

to a toric
variety of smaller dimension, or
K
X

+D

is nef on X

, where D

D.
In the latter case, observe that (X

, D

) is klt and K
X

+D

is nef. We say that


(X

, D

) is a log minimal model of (X

, D).
A basic result of the MMP states that log minimal models are isomorphic
in codimension 1 [179, Thm. 3.52]. In our situation, two log minimal models
(X

, D

) and (X

, D

) of (X

, D) satisfy
[

[ =[

[ =[[,

(1) (1),

(1) (1).
Since X

and X

are isomorphic in codimension 1, it follows that


(15.5.8)

(1) =

(1).
This will enable us to show that the map connecting X

and X

is built from a
special type of ip called a op. Here is the denition.
Denition 15.5.13. Let 1be a ipping extremal ray with ip : X

, and
let D be a Q-divisior on X

. Then is a D-op if D 1= 0.
15.5. The Toric Minimal Model Program 783
Here is a classic example.
Example 15.5.14. Figure 21 on the next page is Figure 3 of Example 11.1.12. For
now, focus on the bottom half of the gure, which involves the cone and two
smooth renements
1
,
2
that give toric morphisms
X

.
The birational map is a ip (see, for instance, Figure 13 from Example 15.3.2).
One easily computes that K
X

1
0. Hence is a K
X

1
-op.
We can now describe the log minimal models of a toric pair (X

, D).
Theorem 15.5.15. Assume that X

is simplicial and semiprojective, and let D =

, where a

[0, 1) Q for all . Then:


(a) Up to toric isomorphism, (X

, D) has only nitely many log minimal models.


(b) Any two log minimal models of (X

, D) are related by a nite sequence of


(K
X

+D)-ops.
Proof. Any two log minimal models share the same rays by (15.5.8), so niteness
is immediate.
To relate the various log minimal models of (X

, D), we start by xing one,


say (X

, D

). Consider the secondary fan where consists of u

for

(1).
Then the GKZ cone
,
= Nef(X

) is a chamber of the secondary fan. Now


let (X

, D

) be a second log minimal model. Then (15.5.8) implies that

(1) =

(1), so that

,
is also a chamber, which (via the birational transform) is
isomorphic to Nef(X

). Under this identication, we see that


,
and

,
both contain the class [K
X

+D

].
Hence the log minimal models of (X

, D) correspond to certain chambers of


the secondary fan that contain [K
X

+D

]. Suppose for the moment that


,
and

,
are chambers that contain [K
X

+D

] and share a common wall, regardless


of whether or not they are actual log minimal models. If 1 is an extremal ray
for X

that denes this wall, then (K


X

+D

) 1= 0 since K
X

+D

is in both
chambers and hence is contained in the wall where they intersect. The resulting
ip X

is a (K
X

+D

)-op. This is what we mean by (K


X

+D)-op
in part (b) of the theorem.
Since any two chambers containing [K
X

+D

] can be connected by a chain


of such wall crossings, we see that any two log minimal models can be connected
by a nite composition of (K
X

+D)-ops.
Here is an example of the log toric MMP.
784 Chapter 15. Geometry of the Secondary Fan
z
y
x

u
2
u
1
u
4
u
3
u
0

3
z
y
x

1
z
y
x

2
Figure 21. The cone with smooth renements 1, 2, 3 in Examples 15.5.14 and 15.5.16
15.5. The Toric Minimal Model Program 785
Example 15.5.16. In Figure 21 from Example 15.5.14, the toric variety X

3
on top
has ve minimal generators
u
0
= e
1
+e
2
+e
3
, u
1
= e
1
, u
2
= e
2
, u
3
= e
2
+e
3
, u
4
= e
1
+e
3
.
We also have the four walls
i
= Cone(u
0
, u
i
) for 1 i 4. The walls
1
and
3
share the wall relation u
2
u
0
+u
4
= 0, so that [V(
1
)] = [V(
3
)] in N
1
(X

3
). One
similarly sees that [V(
2
)] = [V(
4
)]. This gives extremal rays
1
1
=R
0
[V(
1
)] =R
0
[V(
3
)] and 1
2
=R
0
[V(
2
)] =R
0
[V(
4
)].
The intersection formulas from 6.4 make it easy to compute that
K
X

3
1
1
= K
X

3
1
2
=1.
Notice also that X

3
is smooth.
The log toric MMP for (X

3
, 0) is the same as the toric MMP for X

3
and K
X

3
.
We have two ways to begin, since both extremal rays are K
X

3
-negative. They are
also divisorial, and the resulting contractions give the commutative diagram
(15.5.9)
X

2
,
where the fans
1
,
2
are from Figure 21. You should check that X

3
X

1
is
contraction for 1
2
and X

3
X

2
is contraction for 1
1
.
We saw in Example 15.5.14 that K
X

1
0, and similarly K
X

2
0. It follows
that X

1
and X

2
are both minimal models of X

3
(we drop the log in this case
since we are doing the MMP for the canonical divisor). In particular, X

3
does not
have a unique minimal model. Also, Example 15.5.14 shows that the birational
map in (15.5.9) is the op connecting the two minimal models.
Figure 21 tells the full story of this example. You can also see the two minimal
models of X

3
by computing the secondary fan of X

1
.
Note that the cone at the bottom of Figure 21 gives the afne toric variety
V(xy zw) C
4
from Example 1.1.5. It is amazing that this simple toric variety
is still relevant 750 pages after we rst encountered it. This captures perfectly the
power of toric geometry to illustrate deep phenomena in algebraic geometry.
Exercises for 15.5.
15.5.1. This exercise will supply some details omitted in the proof of Proposition 15.5.1
(a) Prove X

Proj(

=0

O
X
(D)) when : X

X
0
is -ample. Hint: You can
assume that X
0
is afne. Use Proposition 7.2.3 and Theorems 7.2.4 and 7.1.13.
(b) Prove (15.5.1). Hint: and

are isomorphisms in codimension 1.


(c) Let D be any divisor on a toric variety X

. Show that

=0
H
0
(X

O
X
(D)) is
a nitely generated C-algebra. Hint: Interpret the ring in terms of the graded ring
C[C(P
D
) (MZ)].
786 Chapter 15. Geometry of the Secondary Fan
15.5.2. Use (15.1.2) and (15.1.3) to prove (15.5.2).
15.5.3. This exercise concerns the fan from Example 15.5.4.
(a) Show that the Mori cone is generated by the classes of the cones C
2
, C
3
and C
4
.
(b) Verify the intersection products computed in Example 15.5.4.
(c) Show that running the toric MMP with K
X
always leads to a bering contraction.
Most of the bering contractions map to P
1
, though there is one that maps to a point.
15.5.4. Here we explore the relation between divisorial and ipping extremal contractions
and the canonical class.
(a) Assume that X

has only terminal singularities and let : X

X
0
be a divisorial
extremal contraction for the extremal ray 1. Show that X
0
has only canonical singu-
larities if and only if K
X
10.
(b) Let : X

X
0
be a ipping extremal contraction for the extremal ray 1. Prove that
K
X
0
is Q-Cartier if and only if K
X
1= 0.
(c) As in part (b), let : X

X
0
be a ipping extremal contraction for the extremal
ray 1, but now assume that K
X
1 < 0. Prove that X
0
is neither Q-factorial nor
Q-Gorenstein, i.e.,
0
is not simplicial and K
X
0
is not Q-Cartier.
Part (c) shows that if 1 is a K
X
-negative extremal ray, then the associated contraction
maps to the badly behaved variety X
0
. This explains why ips are essential in the MMP.
15.5.5. Prove that the toric MMP always terminates when Pic(X

) has rank 1.
15.5.6. Prove parts (a) and (b) of Lemma 15.5.7.
15.5.7. Explain how Example 15.5.12 relates to Example 11.4.26.
15.5.8. Suppose that D is an effective divisor on a simplicial semiprojective toric variety
X

. Show that applying Procedure 15.5.5 to X

and D will never result in a D-negative


bering contraction.
Appendix A
The History of
Toric Varieties
This appendix discusses the origins of toric geometry, with brief remarks about
recent developments. We will describe how the subject evolved and what its early
successes were. We make no claim as to completeness and omit many topics.
A.1. The First Ten Years
Specic examples of toric varieties, such as C
n
, P
n
, P
n
P
m
, and the Hirzebruch
surfaces H
r
, have been known for a long time. The idea of a general toric variety
is more recent.
Demazure. The rst formal denition of toric variety came in 1970 in Demazures
paper Sous-groupes alg ebriques de rang maximum du groupe de Cremona [82].
The Cremona group is the (very large) group of birational automorphisms of P
n
.
Demazures main result is that automorphism groups of smooth toric varieties give
interesting algebraic subgroups of the Cremona group. For him, toric varieties are
certain Z-schemes with a cellular decomposition obtained by
adding certain points at innity to a split torus.
Demazure worked with schemes over Spec(Z), and if M is a lattice, then the split
torus Spec(Z[M]) is a group scheme over Spec(Z). This makes M the character
group of the torus. So the notation for M was there from the beginning.
Here is Demazures denition of fan [82, Def. 4.2.1].
Denition A.1.1. Let M

be a free abelian group of nite type. One calls a fan (in


French, eventail) a nite set of subsets of M

such that
787
788 Appendix A. The History of Toric Varieties
(a) Every element of is a subset of a basis of M

.
(b) Every subset of an element of belongs to .
(c) If K, L , one has NKNL =N(K L).
This looks odd until you realize that Demazure is only considering the smooth
case. His K is the set of minimal generators of a smooth cone . Also, M

is the
dual of M, which is N in our notation. Here are some further denitions from [82]:
The support of is [[ =

K
K.
=

K
NK.
is complete if = M

.
If [[ is the support of a fan in the modern sense, then [[ N is what Demazure
calls . Thus his notion of complete is equivalent to ours.
Given a fan as in Denition A.1.1, Demazure constructs its toric variety X
as a scheme over Spec(Z) by gluing together afne varieties, similar to what we
did in Chapter 3. Many basic results about smooth toric varieties appear in [82].
Here is a sample, with pointers to where the results appear in this book:
For a eld k, X
k
= X
Spec(Z)
Spec(k) is complete if and only if is complete
(Theorem 3.4.6).
The exact sequence M Z
(1)
Pic(X) 0 (Theorem 4.2.1).
If X is smooth and complete, then every ample line bundle on X is very ample
(Theorem 6.1.15).
H
i
(X, O
X
(D))
m


H
i1
(V
D,m
, C) when m M (Theorem 9.1.3).
When X is complete, Demazure applies the last bullet to prove that H
p
(X, O
X
) = 0
for p > 0, the rst toric vanishing theorem. He also gives criteria for a divisor
D=

with Cartier data m

max
to be basepoint free or ample. Support
functions are not mentioned, so that, for example, his criterion for ampleness is
D is ample 'm

, u

` >a

when (1), / (1).


For us, this follows from Lemma 6.1.13 and Theorem 6.1.14.
Given Demazures focus on the Cremona group, it is not surprising that one of
the major results of [82] is an explicit description of the automorphism group of a
smooth complete toric variety (generalized to the simplicial case in [65]).
After Demazure. Besides [82], other papers touched on some of the same ideas,
though often in very different contexts. An example is Satakes 1973 Bulletin
article On the arithmetic of tube domains [243]. It did not take long for people to
esh out the wonderful ideas implicit in these papers. Two particularly important
early works are the Springer lecture notes Toroidal Embeddings I by Kempf, Knud-
sen, Mumford and Saint-Donat, published in 1973 [172], and the paper Almost
homogeneous algebraic varieties under algebraic torus action by Miyake and Oda,
presented at a conference in 1973 and published in 1975 [205].
A.1. The First Ten Years 789
The spirit of the times is captured nicely in Mumfords introduction to [172],
which begins
The goal of these notes is to formalize and illustrate the power of
a technique which has cropped up independently in the work of
at least a dozen people, . . .
The aptness of the phrase cropped up independently is conrmed by a footnote
that Mumford added at the end of the introduction:
*) After this was written, I received a paper by K. Miyake and
T. Oda entitled Almost homogeneous algebraic varieties under
algebraic torus action also on this topic.
The introduction also shows that the elementary aspects of toric geometry were
already recognized:
When teaching algebraic geometry and illustrating simple singu-
larities, varieties, and morphisms, one almost invariably tends to
choose examples of a monomial type type: i.e., varieties de-
ned by equations
x
a
1
1
x
ar
r
= x
a
r+1
r+1
x
an
n
and morphisms f for which
f

(y
i
) = x
a
i1
1
x
a
in
n
.
In [172] we nd the rst appearance of M and N for the dual lattices used in
toric geometry. Here is their denition of fan.
Denition A.1.2. A nite rational partial polyhedral decomposition (we abbre-
viate this to f.r.p.p decomposition) of N
R
is a nite set

of convex rational
polyhedral cones in N
R
such that:
(a) if is a face of

, then =

for some
(b) , ,

is a face of

and

.
This accidentally omits the hypothesis of strong convexity, which is expressed
elsewhere in [172] by saying that does not contain any linear subspace. Given
the unwieldy phrase f.r.p.p decomposition, we should be grateful to Demazure
for the elegant word fan now in use.
A key result of [172] and [205] is that if a normal variety is toric in the sense
of Denition 3.1.1, then it is the toric variety of a fan. This is our Corollary 3.1.8,
which follows from Sumihiros theorem (Theorem 3.1.7), just as in [172] and
[205]. We also nd the Orbit-Cone Correspondence (Theorem 3.2.6) and the usual
criteria for completeness (Theorem 3.4.6) and smoothness (Theorem 3.1.19) in
[172] and [205].
790 Appendix A. The History of Toric Varieties
The properness criterion from Theorem 3.4.11 is proved in [172]. Kempf et
al. also treat convexity in the general setting of sheaves of torus-invariant complete
fractional ideals. The support functions dened in Chapters 4 and 6 appear in part
III.c of Theorem 9 on [172, pp. 2829]. A major result is Theorem 13 on [172,
pp. 48], which gives the criterion of Theorem 7.2.12 for a toric morphism to be
projective. Our proof of this theorem uses results from EGA [127] described in the
appendix to Chapter 7. These results are not mentioned in [172] since the authors,
like many algebraic geometers of the time, knew much of EGA by heart.
Another important result in [172] is that any toric variety has a resolution of
singularities given by a projective toric morphism, our Theorem 11.1.9.
The material on toric varieties appears in the rst chapter of [172]. The second
chapter introduces the more general toroidal varieties, which locally look like
toric varieties in a suitable sense. Toroidal varieties are used in the main result
of the book, the characteristic 0 semi-stable reduction theorem for surjective mor-
phisms f : X C, where C is a smooth curve. More recently, toroidal varieties
have been used to study the structure of birational morphismssee [1] and [279].
Turning to [205], we nd the rst statement of the classication of smooth
complete toric surfaces (Theorem 10.4.3), along with preliminary classication re-
sults for toric threefolds. Miyake and Oda also give an example of a 3-dimensional
smooth complete nonprojective toric variety X that is simpler than the complicated
one given by Demazure. The fan for their example is Figure 9 in Example 6.1.17.
The nonprojective toric threefold X reappears at the end of [205], where the
authors use a series of smooth blowups followed by a series of smooth blowdowns
to convert X into P
3
. They also conjecture that this can be done for any smooth
complete toric threefold. This conjecture is still open, as is its generalization called
the strong Oda conjecture, which applies to birational toric maps between smooth
complete toric varieties of arbitrary dimension. However, if we allow the blowups
and blowdowns to be intermixed, then one gets the weak Oda conjecture, which
has been proved in [2] and [278].
We should mention two other notable papers from this period:
Hochsters 1971 paper Rings of invariants of tori, Cohen-Macaulay rings gen-
erated by monomials, and polytopes [146] proved that the semigroup algebra
of a saturated afne semigroup is Cohen-Macaulay. This implies that normal
toric varieties are Cohen-Macaulay (Theorem 9.2.9). Hochsters paper initiated
the important interaction between toric geometry and commutative algebra.
Ehlers 1975 paper Eine Klasse komplexer Mannigfaltigkeiten und die Au os-
ung einiger isolierter Singularit aten [88] considers toric varieties as complex
manifolds. Our proof of the compactness criterion in Theorem 3.4.1 uses his
argument. Ehlers used innite fans to resolve some interesting singularities.
Similar innite fans were also used to construct smooth toroidal compactications
of bounded symmetric domains. See the books [6] from 1975 and [213] from 1980.
A.1. The First Ten Years 791
The Russian School. In the mid 1970s, Bernstein, Khovanskii and Kusnirenko
were studying subvarieties of (C

)
n
dened by the vanishing of Laurent polyno-
mials f
i
. The Newton polytopes of the f
i
play an important role their work. For
example, the number of solutions of a generic system f
1
= = f
n
= 0 is given
by the mixed volume of the Newton polytopes of the f
i
. This is described in [70,
7.5] and [105, Sec. 5.5].
Drawing on [88], [146] and especially [172], Khovanskii studies toric varieties
in his 1977 paper Newton polyhedra and toroidal varieties. This paper is notable
for several reasons:
It introduced the term support function in the toric context.
It proved the Demazure vanishing theorem H
p
(X, O
X
(D)) = 0 for p > 0 when
D is basepoint free (Theorem 9.2.3).
It gives the rst toric proof of the properties of the Ehrhart polynomial, similar
to what we did in Theorem 9.4.2.
Khovanskiis paper also makes it clear that there is a deep connection between toric
varieties and polytopes. We will soon see another important aspect of this paper.
Danilovs wonderful 1978 survey paper Geometry of toric varieties [76] is a
high point of the era. He drew on all of the references mentioned so far, with the
exception of [205], which was not known in Russia at the time. Danilov covers
an amazing amount of material in [76]. In addition to the basic facts about toric
varieties already mentioned, we also nd:
The inverse limit formula (4.2.5) for Cartier divisors.
The isomorphism
X

n
X
O
X

) from Theorem 8.2.3.


The formula for (U

p
U
) given in Proposition 8.2.18.
The Demazure vanishing theorem (Theorem 9.2.3).
The toric proof of Serre duality (Exercise 9.2.12).
The Bott-Steenbrink-Danilov vanishing theorem (Theorem 9.3.1).
The fundamental group of a toric variety (Theorem 12.1.10).
The Chow and cohomology rings of a smooth toric variety (Theorem 12.5.3).
Riemann-Roch and lattice points in polytopes (Chapter 13).
Danilovs paper remains to this day one of the best introductions to toric geometry.
Toric Varieties. The name toric variety was not used until 1977. In earlier works,
we nd a variety of names, such as:
In [82], Demazure says the scheme dened by the fan .
In [172], Kempf et al. say torus embedding.
In [205], Miyake and Oda say almost homogeneous algebraic variety under
torus action.
792 Appendix A. The History of Toric Varieties
The 1977 article of Khovanskii mentioned earlier originally appeared in Russian as
Mnogogranniki Ntona i toriqeskie mnogoobrazi. This was trans-
lated as Newton polyhedra and toroidal varieties, but toroidal is not the right
word for toriqeskie (toricheskie), because the toroidal varieties dened in [172]
are slightly different from toric varieties. Something different was needed.
The same problem occurred with Danilovs 1978 survey, whose Russian title
is Geometri toriqeskih mnogoobrazi i. One translation was Geometry
of toral varieties, but fortunately for us, toral did not stick. When Miles Reid
translated the paper into English for the Russian Math Surveys, he chose the title
Geometry of toric varieties. This is the origin of the name toric variety.
It took a while before toric variety became standard. For many years, the
term torus embedding introduced in [172] was more common, especially since
it was used in Odas wonderful books [217] from 1978 and [218] from 1988. The
switch to toric variety was conrmed by Odas use of the term in title of his
survey papers from 1991 [219] and 1994 [221].
Polytopes and Normal Fans. Polytopes have been objects of mathematical interest
for over 2000 years. The study of lattice points is more recent, dating from the 19th
century. For example, in 1844 Eisenstein gave a nice proof of quadratic reciprocity
that involved the interior lattice points in the triangle Conv(0,
p
2
e
1
,
p
2
e
1
+
q
2
e
2
) for
distinct odd primes p and q.
In toric geometry, the most interesting object associated to a lattice polytope
is its normal fan. For many years, normal fans were only implicit in the polytope
literature. Here are some examples:
Proposition 6.2.13 relates renements of normal fans and Minkowski sums.
This result is implicit in the 1963 paper Decomposable convex polyhedra by
Shephard [249].
Given a polytope P, the cones C
v
= Cone(Pv) for v a vertex of P are studied
in Exercise 3.4.9 of Gr unbaums classic 1967 book Convex Polytopes [128].
The duals of these cones are the maximal cones of the normal fan of P.
Proposition 6.2.18 relates normal fans of zonotopes to central hyperplane ar-
rangements. This is implicit in McMullens 1971 paper On zonotopes [201].
In the toric literature, normal fans are implicit in Khovanskiis 1977 paper [173]
and are described clearly for the rst time in Odas 1988 book [218], though the
term normal fan did not appear in print until the 1990s (see, for example, [28]).
Toric varieties came to the attention of the polytope community when people
started using them to prove theorems about polytopes, such as Khovanskiis 1977
toric proof of Ehrhart reciprocity [173]. Shortly thereafter, in 1979, Teissier [268]
and Khovanskii (unpublished) used toric varieties and the Hodge index theorem to
prove the Alexandrov-Fenchel inequality for mixed volumes, which states that
MV(P
1
, P
2
, P
3
, . . . , P
n
)
2
MV(P
1
, P
1
, P
3
. . . , P
n
)MV(P
2
, P
2
, P
3
, . . . , P
n
)
A.1. The First Ten Years 793
for rational polytopes P
1
, . . . , P
n
in R
n
. See [105, Sec. 5.4] for a toric proof.
The McMullen Conjecture. The deeper connection between toric varieties and
polytopes appears in Stanleys 1980 paper [257], which completed the proof of the
McMullen conjecture on the face numbers of an n-dimensional simplicial polytope
P. If f
i
is the number of i-dimensional faces of P, then Stanley dened
(A.1.1) h
i
=
i

j=0
(1)
ij

n j
ni

f
j1
=
n

=ni
(1)
(ni)


ni

f
n1
,
where f
1
= 0. Note that h
i
equals the number h
ni
(P) dened in (12.5.13). The
h
i
satisfy the Dehn-Sommerville equations
(A.1.2) h
i
= h
ni
,
and in 1971 McMullen conjectured that
(A.1.3) h
i
h
i1
0, 1 i
n
2
,
and that if
h
i
h
i1
=

n
i
i

n
i1
i 1

+ +

n
r
r

with 1 i
n
2
1 and n
i
> n
i1
> > n
r
r 1, then
(A.1.4) h
i+1
h
i

n
i
+1
i +1

n
i1
+1
i

+ +

n
r
+1
r +1

.
Furthermore, McMullen conjectured that if positive integers f
0
, . . . , f
n1
satisfy
(A.1.2)(A.1.4), then they are the face numbers of an n-dimensional simplicial
polytope. In other words, he conjectured that (A.1.2)(A.1.4) are necessary and
sufcient conditions for the f
i
to come from a simplicial polytope.
The sufciency of (A.1.2)(A.1.4) was proved by Billera and Lee [27]. It
thus remained to show necessity, specically that (A.1.3) and (A.1.4) hold for any
simplicial polytope. Stanleys paper [257] is three pages long, where the rst page
recalls the conjecture and the third is mostly references. It is a one-page proof!
Stanley rst explains why one can assume that P N
R
is rational with the
origin as an interior point. Then cones over the faces of P give a complete fan in
N
R
. The toric variety X

is projective, which for us is easy since is the normal


fan of the dual polytope P

M
R
. Stanley had to work a little harder to prove this
since the theory of normal fans was not fully developed at the time. Note also that
X

is an orbifold since P is simplicial. Hence:


(A.1.2) follows from Poincar e duality.
(A.1.3) follows from the hard Lefschetz theorem.
Stanley then proves (A.1.4) using hard Lefschetz and known results about graded
algebras. A complication is that his arguments require a version of hard Lefschetz
that applies to projective orbifolds, yet Steenbrinks proof in [261] has a gap. To
794 Appendix A. The History of Toric Varieties
get a complete proof of hard Lefschetz in this situation, one needs to use hard
Lefschetz for intersection cohomology, which was proved by Saito in 1990. We
discuss the intersection cohomology of toric varieties in 12.5 and give a careful
statement of hard Lefschetz in Theorem 12.5.8. See also [105, Sec. 5.2].
When we studied polytopes in 9.4, we focused on simple polytopes P M
R
and used their face numbers to dene the numbers h
i
in Theorem 9.4.7. These
differ from the h
i
dened in (A.1.1), but are closely related since the dual of a
simple polytope is simplicial. See Exercise 9.4.10.
A nice commentary on the rst decade of toric geometry was written by Reid
in 1983 [237], where he notes that the construction of the toric variety of a fan
has been of considerable use within algebraic geometry in the
last 10 years . . . and has also been amazingly successful as a
tool of algebro-geometric imperialism, inltrating areas of com-
binatorics. For example, the hard Lefschetz theorem on the co-
homology of projective varieties has been translated into combi-
natorics to complete the proof of a long-standing conjecture of
P. McMullen giving an if and only if condition for the existence
of a simplicial polytope with a given number f
i
of i-dimensional
faces.
Final Comment. We end with an observation from the social history of modern
mathematics. The foundational works on toric varieties were written in a variety of
languages, including English, French, German and Russian. This is quite different
from the current era, where most mathematics is published in English.
A.2. The Story Since 1980
After 1980, the study of toric varieties expanded rapidly, especially during the
1990s. We will say some brief words about some of these developments, though
we caution the reader that many important topics will be omitted.
The 1980s. There was a steady stream of papers about toric varieties in the 1980s,
some of which are featured in this book. Here are a few examples:
In 1983, Reid published Decomposition of toric morphisms [236], cited in
Chapter 15.
In 1986, Danilov and Khovanskii published Newton polyhedra and an algo-
rithm for calculating Hodge-Deligne numbers [77], whose results give a vast
generalization of the sectional genus formula proved in Proposition 10.5.8.
In 1988, Kleinschmidt published A classication of toric varieties with few
generators [177], discussed in 7.3.
A.2. The Story Since 1980 795
In 1989, Gel

fand, Kapranov and Zelevinsky published Newton polyhedra of


principal A-determinants [112], which introduced the GKZ decomposition that
we studied in Chapters 14 and 15.
The titles of these papers give a good sense of the wide range of topics relevant to
toric geometry. A nice overview of the evolution of toric varieties in the 1980s can
be found in Odas 1989 survey paper [219].
One notable event of the decade was the 1988 publication of Odas Convex
Bodies and Algebraic Geometry [218]. This mature and polished book has had a
major inuence on the eld.
The 1990s. In 1989, the rst author had the good fortune to attend a conference at
Washington University where Bill Fulton gave a series of lectures on toric varieties,
based on notes that eventually became his wonderful 1993 book Introduction to
Toric Varieties [105]. As a result, toric geometry began the 1990s with a rich body
of work and the superb expositions of Oda and Fulton. This, coupled with two
other events, led to an explosion of papers about toric varieties.
The rst event was the quotient representation of toric varieties given in 5.1,
which appeared in Audins book The Topology of Torus Actions on Symplectic
Manifolds [11] in 1991. As before, this construction cropped up independently in
several places in the 1990s. The 1995 paper [65] notes ve independent discoveries
of Theorem 5.1.11. The total coordinate ring from 5.2 was introduced in [65] and
has been used in non-toric situations, often under the name Cox ring (see [150]).
The second event was mirror symmetry in mathematical physics, which had a
huge impact on algebraic geometry and on the study of toric varieties in particular.
Here are three brief hints of the role played by toric geometry:
The 1993 paper Phases of N = 2 theories in two dimensions [277] by Witten
used the symplectic version of the quotient representation of a toric variety
(see the end of 12.2) to construct a quantum eld theory called a gauged
linear sigma model.
The 1994 paper Dual polyhedra and mirror symmetry for Calabi-Yau hyper-
surfaces in toric varieties by Batyrev [16] showed that the duality of reexive
polytopes (see 8.3 and Proposition 11.2.1) is related to mirror symmetry.
The 1994 paper Calabi-Yau moduli space, mirror manifolds and spacetime
topology change in string theory by Aspinwall, Greene and Morrison [7] used
ideas from the secondary fan (see Chapters 14 and 15) to study moduli spaces
associated to certain quantum eld theories.
The full story is summarized in the paper [66, Sec. 10] and developed in more
detail in the book [68].
One way to understand the growth of the eld during this decade is to consider
the four survey papers that have been written in recent years about toric varieties:
796 Appendix A. The History of Toric Varieties
Odas 1989 Geometry of toric varieties [219], with 114 references.
Odas 1994 Recent topics on toric varieties [221], with 64 references.
Coxs 1997 Recent developments in toric geometry [66], with 157 references.
Coxs 2001 Update on toric geometry [67], with 240 references.
We should also mention three notable books published in the 1990s:
The 1993 book Discriminants, Resultants and Multidimensional Determinants
[113] by Gel

fand, Kapranov and Zelevinsky included the rst careful treat-


ment of nonnormal toric varieties.
The 1996 book Combinatorial Convexity and Algebraic Geometry by Ewald
[93] was the rst attempt to explain the classic theory of normal toric varieties
to a wider audience than just algebraic geometers.
The 1996 book Gr obner Bases and Convex Polytopes by Sturmfels [264] cov-
ers the elementary theory of afne and projective toric varieties and explains
their relation to toric ideals, Gr obner Bases, and graded algebras.
Applications of Toric Varieties. Besides the applications to physics coming from
mirror symmetry, toric varieties began to be applied to other contexts starting in
the mid 1990s. Here are ve examples:
Relations between toric ideals and integer programming are discussed in the
1996 book Gr obner Bases and Convex Polytopes [264] mentioned above.
Applications to solving systems of polynomial equations are described in the
1998 book Using Algebraic Geometry [70].
Applications to geometric modeling are explored in the 2003 book Topics in
Algebraic Geometry and Geometric Modeling [116].
Applications to coding theory appear in the 2008 book Advances in Algebraic
Geometry Codes [192].
Applications to algebraic statistics are discussed in the 2009 book Lectures on
Algebraic Statistics [84].
Toric Varieties in the 21st Century. Since 2001, research on toric varieties has
continued at a rapid pace. One can measure the growth of the eld by looking
at the MathSciNet database. Starting in 1991, toric varieties have had their own
classication number, 14M25, and during the period between January 2001 and
August 2010, there were
229 papers that listed 14M25 as their primary classication, and
another 434 papers that listed 14M25 as their secondary classication.
It is our hope that this book will help you understand the reasons for this amaz-
ing activity and encourage some of you to make your own contributions to this
wonderful eld of mathematics.
Appendix B
Computational Methods
There are a wide range of packages available for computations in toric geometry.
We focus on packages supported by two open source systems:
(a) Macaulay2 [123], by Dan Grayson, Mike Stillman and collaborators.
(b) Sage [262], by William Stein and collaborators.
Both Magma and GAP also have toric packages, and the open source algebra sys-
tems Singular [78] and CoCoA [72] are functionally similar to Macaulay2. The
Macaulay2 toric package NormalToricVarieties [252] is by Greg Smith, and
the Sage toric package ToricVarieties [44] is by Volker Braun and Andrey
Novoseltsev. There are many other programs which are useful for computations in
toric and polyhedral geometry; an incomplete list might include:
(a) Normaliz by Bruns, Ichim and S oger [57].
(b) LattE by De Loera [79].
(c) Polymake by Gawrilow and Joswig [114].
(d) Polyhedra(Sage) by Braun, Hampton, Novoseltsev and collaborators [43].
(e) 4ti2 by Hemmecke, K oppe, Malkin and Walter [140].
(f) Gfan by Jensen [161].
(g) TOPCOM by Rambau [233, 234].
Many of the programs listed above have interfaces to Macaulay2 and Sage, as
well as interfaces to each other. There are also packages for applications, such as
coding theory (toriccodes by Ilten [153], or toric by Joyner [163]) and physics
(PALP, by Kreuzer and Skarke [182, 183]). In what follows, we will illustrate how
to compute with toric varieties by working through a series of examples. To save
space, we will often suppress superuous output.
797
798 Appendix B. Computational Methods
B.1. The Rational Quartic
Example B.1.1. We begin with a barehanded example. In Exercise 1.1.7, we
considered the map : C
2
C
4
dened by
(s, t) = (s
4
, s
3
t, st
3
, t
4
).
Since the monomials are homogeneous, also denes a map P
1
P
3
. Note that
these monomials correspond to the subset A Z
2
given by the columns of

4 3 1 0
0 1 3 4

.
To nd the image of via Macaulay2, we compute the kernel I of g =

: S R:
i1: R = QQ[s,t]
i2: S = QQ[x,y,z,w]
i3 : g = map(R,S,{s^4,s^3*t,s*t^3,t^4})
o3 : RingMap R <--- S
bi4 : I = kernel g 3 2 2 2 3 2
o4 = ideal (y*z - x*w, z - y*w , x*z - y w, y - x z)
These are the equations appearing in Exercise 1.1.7.
Example B.1.2. The map of the previous example gives the projective curve
X
A
P
3
. Example 2.1.10 showed that X
A
is normal but not projectively normal.
To see this computationally, we use Exercise 9.4.6, which implies that X
A
P
3
is
projectively normal if and only if X
A
is normal and H
1
(P
3
, I
X
A
()) = 0 for all
0. By local duality [89, Thm. A4.2] for the curve X
A
P
3
,
H
1
(P
3
, I
X
A
()) Ext
3
S
(S/I
X
A
, S)
4
.
Since X
A
is dened by the ideal I computed in Example B.1.1, we obtain:
i5 : E3 = Ext^3(coker gens I, S)
o5 = cokernel {-5} | w z y x |
i6 : hilbertFunction(-5,E3)
o6 = 1
Since dim H
1
(P
3
, I
X
A
(1)) =1, projective normality fails. This reects our failure
to use all lattice points of the polytope P = Conv(A) to map to projective space.
Our computational check works in general: a normal variety X P
n
is projectively
normal exactly when Ext
n
S
(S/I
X
, S) = 0, where S =C[x
0
, . . . , x
n
].
Exercises for B.1.
B.1.1. Use Chapter 9 to show that H
i
(P
n
, O
P
n ()) = 0 for all 1 i n 1 and all .
B.2. Polyhedral Computations 799
B.2. Polyhedral Computations
Example B.2.1. The cone = Cone(e
1
, e
2
, e
1
+e
3
, e
2
+e
3
) is depicted in Fig-
ure 2 of Chapter 1, and Example 1.2.9 showed that the dual cone is given by

= Cone(e
1
, e
2
, e
3
, e
1
+e
2
e
3
) R
3
. The Macaulay2 package Polyhedra
is one option for computing with cones: for the example above, we have
i1 : loadPackage "Polyhedra"
i2 : M= matrix{{1,0,0},{0,1,0},{1,0,1},{0,1,1}}
i3 : C=posHull transpose M
o3 = {ambient dimension => 3 }
dimension of lineality space => 0
dimension of the cone => 3
number of facets => 4
number of rays => 4
i4 : rays C
o4 = | 1 0 1 0 |
| 0 1 0 1 |
| 0 0 1 1 |
i5 : fVector C
o5 = {1, 4, 4, 1}
i6 : Cv = dualCone C
i7 : rays Cv
o7 = | 1 0 1 0 |
| 0 1 1 0 |
| 0 0 -1 1 |
i8 : hilbertBasis C
o8 = {| 1 |, | 0 |, | 1 |, | 0 |}
| 0 | | 1 | | 0 | | 1 |
| 0 | | 0 | | 1 | | 1 |
Example B.2.2. We now illustrate a fan computation. In Example 6.1.17 we stud-
ied the smooth fan F consisting of a subdivision of the positive orthant into cones
B1,..,B9, and the remaining 7 orthants C1,..,C7.
o9 : C1=posHull transpose matrix{{0,-1,0},{0,0,1},{-1,0,0}};
(input cones C2..B9, supressed)
i25 : F=fan{C1,C2,C3,C4,C5,C6,C7,B1,B2,B3,B4,B5,B6,B7,B8,B9}
o25 = {ambient dimension => 3 }
number of generating cones => 16
number of rays => 10
top dimension of the cones => 3
i26 : isPolytopal F
o26 = false

800 Appendix B. Computational Methods
Example B.2.3. In this example, we compute the secondary polytope for the con-
vex hull of six points in the plane. The secondary polytope is an associahedron, and
appears in Example 15.2.13. One package which computes the secondary polytope
is TOPCOM by Rambau. Braun has created a Sage version of TOPCOM, which we use
for the computation (you may need to install this package separately).
sage: pointmatrix = matrix([
[ 1, 1, 0, -1, -1, 0 ],
[ 0, 1, 1, 0, -1, -1 ]])
sage: pc = PointConfiguration(pointmatrix.columns())
sage: Tris = pc.triangulations_list()
sage: Tri = Tris[7]
sage: list(Tri)
[[0, 1, 4], [0, 4, 5], [1, 2, 3], [1, 3, 4]]
sage: show(Tris[7].plot(axes=False))
Figure 1. The seventh triangulation
Next, we generate the secondary polytope and face lattice
sage: K3 = pc.secondary_polytope(); K3
A lattice polytope: 3-dimensional, 14 vertices.
sage: K3.faces()
[[0],[1], (vertices: output suppressed) [13]],
[[1,6],[5,6],[0,1],[0,5],[6,10],[1,4],[4,13],[10,13],
[5,7],[7,9],[9,10],[3,4],[2,3],[0,2],[3,12],[12,13],
[2,8],[8,11],[11,12],[7,8],[9,11]],
[[0,1,5,6],[7,8,9,11],[0,2,5,7,8],[0,1,2,3,4],[3,4,12,13],
[2,3,8,11,12],[5,6,7,9,10],[1,4,6,10,13],[9,10,11,12,13]]
B.2. Polyhedral Computations 801
sage: plotK3 = K3.plot3d() #view the secondary polytope
sage: plotK3.rotateZ(60*pi/180).show(viewer=tachyon)
Figure 2. The secondary polytope K
3
Figure 2 is a simplied version of the Sage output optimized for viewing in black
and white. To see the 1-skeleton appearing in Figure 9 of Example 15.2.13, enter:
sage: Kp = Polyhedron(vertices=K3.vertices().columns())
sage: Egraph = Kp.graph()
sage: Egraph.show()
0
1
2
3
4
5
6
7
8
9
10
11
12
13
Figure 3. The edge graph of K
3

802 Appendix B. Computational Methods


Example B.2.4. In our nal polyhedral example, we compute the Gr obner fan of
the ideal 'x
7
1, yx
5
` which appears in Example 10.3.1, using the package Gfan,
which has both Macaulay2 and Sage interfaces. In Macaulay2:
i1 : R=QQ[x,y]; I=ideal(x^7-1, y-x^5); gfan(I)
+-------------+----------------------------+
| 7 | 7 3 |
o1 = |{y , x} |{y - 1, - y + x} |
+-------------+----------------------------+
| 3 2 3 | 3 2 3 2 |
|{y , x y, x }|{y - x, x y - 1, x - y } |
+-------------+----------------------------+
| 2 2 5 | 3 2 2 5 |
|{y , x y, x }|{- x + y , x y - 1, x - y}|
+-------------+----------------------------+
| 7 | 5 7 |
|{y, x } |{- x + y, x - 1} |
+-------------+----------------------------+
And in Sage:
sage: R.<x,y,z> = PolynomialRing(QQ)
sage: I=R.ideal([x^7-1,y-x^5]).groebner_fan()
sage: I.reduced_groebner_bases()
[[y^7 - 1, -y^3 + x], [y^3 - x, x^2*y - 1, x^3 - y^2],
[-x^3 + y^2, x^2*y - 1, x^5 - y], [-x^5 + y, x^7 - 1]]
sage: I.render()
2 4 6 8 10 12
2
4
6
8
10
Figure 4. Output of the render command
The render command expects input in three variables (hence the three variable
ring). The gure above is the intersection of the Gr obner fan with the standard two
simplex, which after an afne transformation gives the fan of Example 10.3.1.
B.3. Normalization and Normaliz 803
B.3. Normalization and Normaliz
The Normaliz package is designed for computations with afne monoids, vector
congurations, lattice polytopes, and rational cones.
As input, Normaliz expects a le containing matrices; each matrix is preceded
by two lines, the rst indicating the number of rows and the second the number of
columns.
Example B.3.1. We illustrate Normaliz for the very ample but not normal poly-
tope of Example 2.2.20.
10
6
1 1 1 0 0 0
(8 additional lattice points, suppressed)
0 0 1 1 0 1
1
Save the le above as VeryampleNonNormal.in. Normaliz supports ten types
of input, specied by the number on the line following the matrix. In the example
above the 1 on the last line indicates we want to normalize the monoid generated by
the rows of the matrix. To execute the code, use the graphical interface jNormaliz,
or type directly from the command line:
% norm64 VeryampleNonNormal.in
The output is saved in the le VeryampleNonNormal.out, which reads:
11 Hilbert basis elements
10 height 1 Hilbert basis elements
10 extreme rays
22 support hyperplanes
rank = 6 (maximal)
index = 1
original monoid is not integrally closed
extreme rays are homogeneous via the linear form:
1 1 1 1 1 1
Hilbert basis elements are not homogeneous
multiplicity = 21
h-vector:
1 4 11 4 1 0
Hilbert polynomial:
1/1 157/60 25/8 53/24 7/8 7/40
****************************************************
11 Hilbert basis elements:
(output suppressed: original rays and 1 1 1 1 1 1)
804 Appendix B. Computational Methods
10 extreme rays:
(output suppressed: original rays)
22 support hyperplanes:
-1 2 -1 2 2 -1
-1 -1 2 2 2 -1
(remaining 20 hyperplanes suppressed)
1 congruences:
1 1 1 1 1 1 3
10 height 1 Hilbert basis elements:
(output suppressed: original rays)
Normaliz has interfaces for both Macaulay2 and Singular. For example, start
Macaulay2 in the directory containing Normaliz, and enter the commands
i1 : installPackage "Normaliz"
i2 : A = matrix{{1,1,1,0,0,0},(suppressed),{0,0,1,1,0,1}}
i3 : normaliz(A,1)

Example B.3.2. Consider the binomial ideal I computed in Example B.1.1. To
nd the associated monoid and normalization, we use the option 10 in Normaliz.
We encode the rst generator yz xw of I as -1 1 1 -1, and similarly for the
other three generators. Hence our input le is
4
4
-1 1 1 -1
0 -1 3 -2
1 -2 2 -1
-2 3 -1 0
10
This yields output (with some items suppressed)
4 original generators:
0 4
3 1
1 3
4 0
5 Hilbert basis elements:
4 0
3 1
2 2
1 3
0 4
You should compare this to what we computed in Example 1.3.9.
B.4. Sheaf Cohomology and Resolutions 805
Example B.3.3. In Example 9.4.5 we found the Ehrhart polynomial for the lattice
simplex Conv(0, 0, 0), (1, 0, 0), (0, 1, 0), (1, 1, 3). Create an input le B3-3.in as
below:
4
3
0 0 0
1 0 0
0 1 0
1 1 3
2
The command norm64 -h B3-3.in produces an output le B3-3.out as below.
The -h option stipulates that the Ehrhart polynomial is to be included in the output
data.
4 lattice points in polytope
4 extreme points of polytope
4 support hyperplanes
polytope is not integrally closed
dimension of the polytope = 3
normalized volume = 3
h-vector:
1 0 2 0
Ehrhart polynomial:
1/1 3/2 1/1 1/2
The translates to the polynomial 1+
3
2
x +x
2
+
1
2
x
3
from Example 9.4.5.
B.4. Sheaf Cohomology and Resolutions
Example B.4.1. Example 9.5.6 computed H
1
(P
1
, O
P
1 (5)) via Ext. Let I =
'x, y` S =C[x, y]. Then
H
1
(P
1
, O
P
1 (5)) = H
2
I
(S)
5
= Ext
2
S
(A, S)
5
,
where A = S/I
[k]
= S/'x
k
, y
k
` for k sufciently large. Since dim H
2
I
(S)
2
= 4 by
Example 9.5.3, we can use Macaulay2 to nd a k that works. Suppressing some
outputs, one computes:
i1 : S = QQ[x,y]
i2 : A = coker matrix{{x^3,y^3}}
i3 : hilbertFunction(-5,Ext^2(A,S))
o3 = 2
So k = 3 is not big enough. Repeating this with k = 4 shows that k = 4 works.
806 Appendix B. Computational Methods
Example B.4.2. In Example 9.5.13, we needed a resolution of the cotangent bun-
dle of P
1
. This is done using the commands:
i1 : S = QQ[x,y,z]
i2 : OM = ker matrix{{x,y,z}}
i3 : F = res OM
i4 : F.dd
3 1
o4 = 0 : S <-------------- S : 1
{2} | z |
{2} | x |
{2} | -y |
The last two commands compute the resolution and display the differentials. This
gives the resolution shown in (9.5.9).
B.5. Sheaf Cohomology on the Hirzebruch Surface H
2
This example illustrates the NormalToricVarieties package for Macaulay2,
written by Greg Smith. Use viewHelp for documentation on the commands.
Example B.5.1. In Example 13.2.12 we studied the line bundle L = O
X
(D) for
D = 3D
3
5D
4
on X = H
2
, computing that H
0
(X, L) = 0, dim H
1
(X, L) = 2,
and dim H
2
(X, L) = 6. The command hirzebruchSurface(2) creates the toric
variety X =H
2
:
i1 : installPackage "NormalToricVarieties"
i2 : H2 = hirzebruchSurface(2)
i3 : rays H2
o3 = {{1, 0}, {0, 1}, {-1, 2}, {0, -1}}
i4 : max H2
o4 = {{0, 1}, {0, 3}, {1, 2}, {2, 3}}
i5 : isFano H2
o5 = false
Nowwe bring Dinto the picture. In NormalToricVarietiesa divisor is specied
by a tuple corresponding to the rays of the fan, so with the rays ordered as above,
D is 3D
0
5D
3
.
i6 : D = 3*H2_0 - 5*H2_3
o6 = 3*D - 5*D
0 3
i7 : isCartier D
o7 = true
i8 : isAmple D
o8 = false
B.5. Sheaf Cohomology on the Hirzebruch Surface H
2
807
To compute cohomology, we turn D into the sheaf L = O
X
(D), and use a loop to
display the ranks of H
i
(X, L):
i9 : L = OO D
1
o9 = OO (3,-5)
H2
i10 : for i to 2 list rank HH^i(H2, L)
o10 = {0, 2, 6}
Example B.5.2. We continue to study cohomology on X = H
2
, but now turn our
attention to
1
X
. In Table 4 of Example 9.5.14 we computed dim H
1
(X,
1
X
(a, b))
for (a, b) (2, 2), . . . , (2, 2). We will build a script called cohomologyTable
which takes as input an integer k reecting the H
k
to compute, a sheaf, and high and
low values for the range of degrees in the Picard group. Since cohomologyTable
builds on packages BoijSoederberg and BGG, we need to load these. Below, we
check the computations in Example 9.5.14
i11 : loadPackage "BoijSoederberg"
i12 : loadPackage "BGG"
i13 : OM = cotangentSheaf H2
o13 = cokernel {2, 0} | 2x_1x_3 |
{-1, 2} | x_0 |
{-1, 2} | -x_2 |
1 2
o13 : coherent sheaf on H2, quotient of OO (-2,0)++OO (1,-2)
H2 H2
i14 : cohomologyTable := (k,F,lo,hi) ->
(DegRange = toList(lo#0..hi#0);
new CohomologyTally from select(flatten apply(DegRange,
j -> apply(toList(lo#1..hi#1),
i -> {(j,i-j), rank HH^k(variety F, F(i,j))})),
p -> p#1 != 0));
i15 : cohomologyTable(1,OM,{-2,-2},{2,2})
-2 -1 0 1 2
o15 = 2: 3 2 1 . .
1: 4 2 1 . .
0: 3 2 2 2 3
-1: . . 1 2 4
-2: . . 1 2 3
Exercises for B.5.
B.5.1. Compute the cohomology table for H
i
(P
2
,
1
P
2
( j)) appearing in Example 9.5.13.
808 Appendix B. Computational Methods
B.6. Resolving Singularities
In this section we use the ToricVarieties package in Sage, written by Braun
and Novoseltsev, to resolve singularities. The package nds a simplicial resolution
automatically.
Example B.6.1. Let us revisit Example B.2.1. Consider the afne toric variety
dened by the cone = Cone(e
1
, e
2
, e
1
+e
3
, e
2
+e
3
), which is the nonsimplicial
afne threefold appearing in Example 11.1.12. We rst nd a simplicial resolution:
sage: C = Cone([(1,0,0), (0,1,0), (1,0,1), (0,1,1)]); C
3-d cone in 3-d lattice N
sage: QuadCone = AffineToricVariety(C)
sage: Orbif = QuadCone.resolve_to_orbifold()
sage: Orbif.fan().ngenerating_cones()
2
sage: [ cone.ambient_ray_indices() for cone in Orbif.fan() ]
[(2, 3, 0), (1, 3, 0)]
This is the renement
1
of appearing in Figure 3 of Example 11.1.12. Note that
the rays are indexed by 0..3 rather than 1..4.
Example B.6.2. In this example, we resolve weighted projective space P(1, 1, 2),
whose fan is pictured in Figure 11 of Example 2.4.6.
sage: s0 = Cone([(0,1), ( 1, 0)]);
sage: s1 = Cone([(0,1), (-1,-2)]);
sage: s2 = Cone([(1,0), (-1,-2)]);
sage: F = Fan([s0, s1, s2]);
sage: P112 = ToricVariety(F); P112
2-d toric variety covered by 3 affine patches
sage: P112.is_orbifold()
True
sage: P112.is_smooth()
False
sage: lp = LatticePolytope( matrix(F.rays()).transpose() )
sage: lp.points() # columns are all points in the convex hull
[ 0 1 -1 0 0]
[ 1 0 -2 -1 0]
sage: BlP112 = P112.resolve(new_rays=[(0,-1)]); BlP112
2-d toric variety covered by 4 affine patches
sage: BlP112.is_smooth()
True
We used this resolution of P(1, 1, 2) in Example 13.4.4.
B.7. Intersection Theory and Hirzebruch-Riemann-Roch 809
B.7. Intersection Theory and Hirzebruch-Riemann-Roch
Example B.7.1. In Example 13.2.12, we checked the Hirzebruch-Riemann-Roch
theorem for the divisor D=D
3
5D
4
on H
2
. Using the ToricVarietiespackage
in Sage and observing that the variety BlP112 in the previous example is X =H
2
,
we calculate

X
ch(L) Td(X), L =O
X
(D).
sage: BlP112.fan().ray_matrix()
[ 0 1 -1 0]
[ 1 0 -2 -1]
sage: BlP112.cohomology_ring().gens()
([2*z2 + z3], [z2], [z2], [z3])
sage: d4 = BlP112.fan().cone_containing(0);d4
sage: D4 = d4.cohomology_class()
[2*z2 + z3]
sage: d3 = BlP112.fan().cone_containing(1);d3
sage: D3 = d3.cohomology_class(); D3
[z2]
sage: chL = (3*D3-5*D4).exp(); chL
[-5*z3^2 - 7*z2 - 5*z3 + 1]
sage: tdX = BlP112.Todd_class(); tdX
[-1/2*z3^2 + 2*z2 + z3 + 1]
sage: BlP112.integrate(tdX*chL)
4
This agrees with our computation of (L) in B.5. We can also check by extract-
ing the degree two piece of ch(L) Td(X):
sage: chD*tdX
[-2*z3^2 - 5*z2 - 4*z3 + 1]
sage: Deg2part=(chD*tdX).part_of_degree(2); Deg2part
[-2*z3^2]
With labelling as in Example 4.1.8, z3 corresponds to the ray u
2
, and one checks
that the corresponding divisor D
2
has self-intersection 2. Thus 2D
2
2
= 4.
Example B.7.2. To compute sheaf cohomology using ToricVarieties, rst we
create a divisor from an list indexed by the rays, where the order corresponds to
the ray matrix of the fan:
sage: Ddiv = BlP112.divisor([-5,3,0,0]);Ddiv
sage: Ddiv.cohomology()
(0, 2, 6)
sage: M = BlP112.fan().dual_lattice()
810 Appendix B. Computational Methods
sage: Ddiv.cohomology(M(-2,1))
(0, 1, 0)
The last line gives dim H
i
(H
2
, O
H
2
(D))
(2,1)
; the weight (2, 1) corresponds to a
weight of (2, 1) in Figure 2 in Example 13.2.12 since the fans are ipped.
B.8. Anticanonical Embedding of a Fano Toric Variety
Example B.8.1. The toric variety X = X

in Examples 9.1.2 and 9.1.8 is the


blowup of P
2
at the three torus-xed points. We rst show how to create this toric
variety from scratch using the Macaulay2 package NormalToricVarieties.
The input is a list of rays and a list of maximal cones. The rays below are ordered
as in Figure 2 of Example 9.1.2.
i2 : rayList = {{1,0},{1,1},{0,1},{-1,0},{-1,-1},{0,-1}}
i3 : coneList = {{0,1},{1,2},{2,3},{3,4},{4,5},{5,0}}
i4 : BlP2 = normalToricVariety(rayList,coneList)
Now let D = K
X
be the anticanonical divisor, which is very ample since X
is a smooth Fano surface. We analyze the equations that arise when we embed
X into P(H
0
(X, O
X
(D))). Below, we create a script which takes as input a toric
divisor D, nds the sections, and returns the ideal dening X P(H
0
(X, L)) for
L =O
X
(D). Make a le called Embed as below:
projEmb = (D)->(X = variety D;
L = OO D;
m = rank HH^0(X, L);
S = ring X;
R = QQ[y_0..y_(m-1)];
phi = map(S, R, basis(- first degrees L, S));
kernel phi)
Now we reap the benets of our work:
i5 : D = BlP2_0+BlP2_1+BlP2_2+BlP2_3+BlP2_4+BlP2_5
i6 : load "Embed"
i7 : I = projEmb(D);
i8 : transpose mingens I
o8 = {-2} | u_4u_5-u_3u_6 | (more quadrics, output suppressed)
i9 : hilbertPolynomial(coker gens I, Projective=>false)
2
o9 = 3i + 3i + 1
Exercises for B.8.
B.8.1. Compute the Ehrhart polynomial of P
KX
using Normaliz as in Example B.3.3.
B.8.2. Compute H
i
(X, O
X
(D)) for the divisor D =D
3
+2D
5
D
6
from Example 9.1.8.
Appendix C
Spectral Sequences
Spectral sequences are important tools in modern algebraic geometry and algebraic
topology that are used several times in Chapters 9 and 12. Here we give some
background necessary for understanding these applications for readers who have
not encountered these objects before. Our discussion is very far from complete; we
refer the reader to [115], [136], [199], or [276] for more comprehensive treatments
of spectral sequences.
C.1. Denitions and Basic Properties
At rst glance, spectral sequences may seem rather complicated. Fortunately, they
are often straightforward to use in practice.
The Denition. The spectral sequences encountered in this book will all be of the
following type and will be relatively simple to understand.
Denition C.1.1. A (cohomology) spectral sequence is a collection of abelian
groups E
p,q
r
and homomorphisms d
p,q
r
with the following structure and properties:
(a) The groups are E
p,q
r
, indexed by integers p, q, r. Fixing r, we obtain one sheet
of the spectral sequence, which is visualized as a diagram of groups indexed
by integer lattice points in the plane.
(b) In the rth sheet, there are homomorphisms
d
p,q
r
: E
p,q
r
E
p+r,qr+1
r
such that d
p+r,qr+1
r
d
p,q
r
= 0 for all p, q, r. In other words, the rth sheet
splits up into a collection of cochain complexes in which the differentials are
all mappings of bidegree (r, 1r) for the indexing by p, q.
811
812 Appendix C. Spectral Sequences
(c) The (r +1)st sheet E
p,q
r+1
is the cohomology of (E
p,q
r
, d
p,q
r
), i.e.,
E
p,q
r+1
= ker(d
p,q
r
: E
p,q
r
E
p+r,qr+1
r
)

im(d
pr,q+r1
r
: E
pr,q+r1
r
E
p,q
r
).
In many of the examples in the text, the E
p,q
r
are vector spaces over Q or C
and the d
p,q
r
are linear mappings. There are also homology spectral sequences that
appear in many applications; the only differences are that the groups are usually
written E
r
p,q
and the differentials d
r
p,q
in a homology spectral sequence have bide-
gree (r, r 1).
We will always work with rst quadrant spectral sequences, for which E
p,q
r
=0
when p <0 or q <0. Thus in each sheet, the nonvanishing terms lie in the quadrant
where p, q 0. The minimum value of r is usually 1 or 2, in which case we say
that (E
p,q
r
, d
p,q
r
) is an E
1
or E
2
spectral sequence respectively.
Intuition and Meaning of Convergence. When working with an E
1
spectral se-
quence, for instance, we often know the differentials d
p,q
1
explicitly. In general,
however, the differentials d
p,q
r
for r 2 are much more difcult to describe. Fortu-
nately, there are many cases where the structure of the nonzero terms in a spectral
sequence forces many differentials to be zero. For instance, in the rst quadrant
spectral sequences we will encounter, the differentials mapping to and from E
p,q
r
for xed p, q vanish when r is sufciently large. It follows that for each p, q, there
exists some r (depending on p, q) such that
E
p,q
r
= E
p,q
r+1
= E
p,q
r+2
= .
This common value is dened to be E
p,q

.
Denition C.1.2. A rst quadrant spectral sequence (E
p,q
r
, d
p,q
r
) converges to a
sequence of abelian groups H
k
, k 0, if there is a ltration
0 = F
k+1
H
k
F
k
H
k
F
k1
H
k
F
1
H
k
F
0
H
k
= H
k
of H
k
by subgroups such that
E
p,q

F
p
H
p+q
/F
p+1
H
p+q
.
For an E
1
or E
2
spectral sequence, we write this as
E
p,q
1
H
p+q
or E
p,q
2
H
p+q
respectively.
According to [199, Ch. 1], the generic theorem about spectral sequences
says something like: there is a spectral sequence with E
p,q
1
or E
p,q
2
something
computable and converging to H
p+q
, something desirable. Several specic ex-
amples of this are discussed later. The intuition behind a convergent E
1
spectral se-
quence is that we can usually compute the E
2
terms from the d
1
differentials explic-
itly. But we are primarily interested in what the spectral sequence converges to. We
can think of E
p,q
2
as a rst approximation to the convergent, and then E
p,q
3
, E
p,q
4
, . . .
C.1. Denitions and Basic Properties 813
as better and better approximations. If the situation is favorable, enough of the
higher differentials d
p,q
r
for r 2 and enough of the E
p,q

may vanish so that the


H
p+q
can be determined.
Here is a simple rst example where we can carry out this sort of analysis.
Proposition C.1.3. Suppose E
p,q
2
H
p+q
is a rst quadrant E
2
spectral sequence
with the property that E
p,q
2
= 0 for all q > 0. Then
E
p,0
2
H
p
for all p 0.
Proof. First observe that all differentials d
p,q
r
vanish for r 2 since E
p,q
2
= 0 for
q > 0. It follows that E
p,q
2
= E
p,q

for all p, q. Then we have


E
p,0
2
F
p
H
p
/F
p+1
H
p
= F
p
H
p
and 0 = E
p1,1
2
= E
p2,2
2
= = E
0,p
2
implies
0 = F
p1
H
p
/F
p
H
p
= F
p2
H
p
/F
p1
H
p
= = F
0
H
p
/F
1
H
p
.
Hence F
p
H
p
=F
p1
H
p
= =F
1
H
p
=F
0
H
p
=H
p
. Thus E
p,0
2
H
p
, as claimed.

The hypothesis of Proposition C.1.3 implies that the differentials d


p,q
2
vanish
for all p, q. This is an instance of the following general idea.
Denition C.1.4. We say that a spectral sequence degenerates at the E
r
sheet if
the differentials d
p,q
s
= 0 for all p, q and all s r.
Note that degeneration at E
r
implies that E
p,q
r
E
p,q

for all p, q, so we have a


strong form of convergence in this case.
If we have a convergent spectral sequence, say E
p,q
2
H
p+q
, it is important
to note that there are situations where knowing E
p,q

for all p, q is not quite enough


to determine H
p+q
. This can happen for instance if there are several E
p,q

with
p+q = k for some k. Denition C.1.2 shows that we know the quotients
E
p,q

F
p
H
p+q
/F
p+1
H
p+q
in the ltration of H
k
= H
p+q
. However, additional extension data may be needed
to determine the isomorphism type of the abelian group H
k
in general. This situa-
tion does not arise when the E
p,q
r
are vector spaces over a eld. In all cases, though,
the proof of Proposition C.1.3 shows the following.
Proposition C.1.5. Let E
p,q
r
0
H
p+q
be a spectral sequence. If E
p,q

= 0 for all
p, q with p+q = k except possibly p, q = p
0
, q
0
, then H
k
E
p
0
,q
0

.
814 Appendix C. Spectral Sequences
Edge Homomorphisms. Another feature of a convergent rst quadrant E
2
spec-
tral sequence is that all differentials starting from E
p,0
r
vanish, so that the spectral
sequence gives maps
E
p,0
2
E
p,0
3
E
p,0

F
p
H
p
/F
p+1
H
p
= F
p
H
p
H
p
.
Denition C.1.6. The map E
p,0
2
H
p
is called an edge homomorphism.
This leads to a more precise version of Proposition C.1.3.
Proposition C.1.7. Suppose E
p,q
2
H
p+q
is a rst quadrant E
2
spectral sequence
with the property that E
p,q
2
= 0 for all q > 0. Then the edge homomorphism
E
p,0
2
H
p
is an isomorphism for all p 0.
A convergent rst quadrant E
2
spectral sequence has second edge homomor-
phism since all differentials ending at E
0,q
r
vanish, so that we have the map
H
q
H
q
/F
1
H
q
= F
0
H
q
/F
1
H
q
= E
0,q

E
0,q
3
E
0,q
2
.
Denition C.1.8. The map H
q
E
0,q
2
is called an edge homomorphism.
For many E
2
spectral sequences, we know some of the edge homomorphisms.
For example, we know E
p,0
2
H
p
in the two applications of Proposition C.1.7
given in 9.0, and H
q
E
0,q
2
is known when we apply the Serre spectral sequence
(Theorem C.2.6) in 12.4.
C.2. Spectral Sequences Appearing in the Text
In the remainder of this appendix we discuss the rst quadrant spectral sequences
used in Chapters 9 and 12 to derive facts about singular and sheaf cohomology of
toric varieties.
The Leray Spectral Sequence. The Leray spectral sequence of a continuous map
is used in 9.0.
Theorem C.2.1. Let f : X Y be a continuous map of topological spaces, and let
F be a sheaf of abelian groups on X. There is a spectral sequence
E
p,q
2
= H
p
(Y, R
q
f

F) H
p+q
(X, F).
Furthermore, when q = 0, the map H
p
(Y, f

F) H
p
(X, F) is the edge homo-
morphism E
p,0
2
H
p
(X, F) (see Denition C.1.6 above).
This spectral sequence is covered in [115, II.4.17] and [125, p. 463], and its
intuitive meaning is explained in the discussion leading up to Proposition 9.0.8.
C.2. Spectral Sequences Appearing in the Text 815
The Spectral Sequence of a Covering. Next, let U =U
i
be an open cover of X
and F be a sheaf on X. If U consists of two sets and F is a constant sheaf, then
the Mayer-Vietoris long exact sequence can be used to obtain information about
the cohomology of X from the cohomology of U
1
,U
2
and U
1
U
2
. More generally
H

(X, F) is determined by H

(U
i
0
U
i
p+1
, F) as we vary over all p by way
of the following spectral sequence of an open cover.
Theorem C.2.2. Let U = U
i
be an open cover of X and F be a sheaf on X.
Then there is a spectral sequence
E
p,q
1
=

(i
0
,...,ip)[]p
H
q
(U
i
0
U
ip
, F) H
p+q
(X, F),
where the differential d
p,q
1
: E
p,q
1
E
p+1,q
1
is induced by inclusion, with signs sim-
ilar to the differential in the

Cech complex.
This spectral sequence is discussed in [115, II.5.4]. We use it in 9.0 and 12.3.
A variant of Theorem C.2.2 applies to a locally nite closed cover C =C
i
. This
means X =

i
C
i
, each C
i
is closed in X, and each point x X has a neighborhood
meeting only nitely many C
i
. A nite cover is clearly locally nite.
Theorem C.2.3. Let C = C
i
be a locally nite closed cover of X and F be a
sheaf on X. Then there is a spectral sequence
E
p,q
1
=

(i
0
,...,ip)[]p
H
q
(C
i
0
C
ip
, F) H
p+q
(X, F),
where the differential d
p,q
1
: E
p,q
1
E
p+1,q
1
is induced by inclusion, with signs sim-
ilar to the differential in the

Cech complex.
This spectral sequence is constructed in [115, II.5.2]. We will use it in 9.1
when we compute the sheaf cohomology of a toric variety.
The Spectral Sequence for Ext. Let X be a variety. As in 9.0, the Ext groups
Ext
p
O
X
(F, G ) are the derived functors of the Hom functor G Hom
O
X
(F, G ) for
xed F. The Ext sheaves Ext
q
O
X
(F, G )) are dened similarly. These are related
by the following spectral sequence.
Theorem C.2.4. Let F and G be quasicoherent sheaves on a variety X. Then
there is a spectral sequence
E
p,q
2
= H
p
(X, Ext
q
O
X
(F, G )) Ext
p+q
O
X
(F, G).
We will use this in Theorem 9.2.12, which is part of our discussion of Serre
duality. The spectral sequences from Theorems C.2.1 and C.2.4 are special cases of
the Grothendieck spectral sequence of a composition of functors (see [276, 5.8]).
816 Appendix C. Spectral Sequences
The Spectral Sequences of a Filtered Toplogical Space. First, given an increasing
ltration of a topological space X,
(C.2.1) = X
1
X
0
X
1
X
n
= X,
suppose we have information about the singular cohomology of the pairs (X
p
, X
p1
)
for all p. This information can be stitched together to produce the E
1
sheet of the
spectral sequence of a ltration. See [136, Ch. 1] and [255, Sec. 9.4] for proofs.
Theorem C.2.5. Let X be a topological space with a ltration of the form (C.2.1)
and R be a ring. Then there is a spectral sequence
E
p,q
1
= H
p+q
(X
p
, X
p1
, R) H
p+q
(X, R),
where the ltration on H
p+q
(X, R) is given by
F
p
H
p+q
(X, R) = ker(H
p+q
(X, R) H
p+q
(X
p
, R))
for all p, q.
This is not always a rst quadrant spectral sequence, though it will be when
we apply Theorem C.2.5 in 12.3 to the ltration of a toric variety, where X
p
is the
union of the closures of the T
N
-orbits of dimension p.
The Serre Spectral Sequence. The spectral sequence for a ltration can be used to
derive the following Serre spectral sequence of a bration. We recall that a bra-
tion is a continuous mapping p : E B satisfying the homotopy lifting property
with respect to all spaces Y (see [199, 4.3]). It follows that if B is path-connected,
then all of the bers p
1
(b) for b B are homotopy equivalent. Hence we can
speak of the ber and denote it by F. Let i : F E be the inclusion of the ber.
Theorem C.2.6. Let p : E B be an orientable bration over a path-connected
and simply-connected base B with ber F, and let R be a ring. Then there is a
spectral sequence
E
p,q
2
= H
p
(B, H
q
(F, R)) H
p+q
(E, R)
for a suitable ltration on H
p+q
(E, R). Furthermore:
(a) The edge homomorphism E
p,0
2
H
p
(E, R) is p

: H
p
(B, R) H
p
(E, R).
(b) The edge homomorphism H
q
(E, R) E
0,q
2
is i

: H
q
(E, R) H
q
(F, R).
If B is not simply connected, then a similar result is true, but the fundamental
group
1
(B) acts on the the cohomology of the ber and the E
2
sheet involves
the more complicated cohomology with local coefcients. This is discussed in
[136, Ch. 1], [199, Thm. 5.2], and [255, Sec. 9.4]. We use the Serre spectral
sequence in 12.4 to study the equivariant cohomology of a toric variety. In our
case, B = (P

)
n
= (

=1
P

)
n
, so the hypotheses of Theorem C.2.6 are satised.
The Serre spectral sequence can be seen as a special case of the Leray spectral
sequence from Theorem C.2.6.
Bibliography
[1] D. Abramovich, K. Karu, K. Matsuki and J. Wodarczyk, Torication and factorization of
birational maps, J. Amer. Math. Soc. 15 (2002), 531572.
[2] D. Abramovich, K. Matsuk and S. Rashid, A note on the factorization theorem of toric bira-
tional maps after Morelli and its toroidal extension, Tohoku Math. J. 51 (1999), 489537.
[3] A. ACampo-Neuen and J. Hausen, Toric invariant theory, in Algebraic Group Actions and
Quotients, (J. Wi snewski, ed.), Hindawi Publ. Corp., Cairo, 2004 137.
[4] O. Aichholzer, Extremal properties of 0/1-polytopes of dimension 5, in Polytopes Combina-
torics and Computation, (G. Kalai and G. Ziegler, eds.), DMV Seminar 29, Birkh auser Verlag,
Basel Boston Berlin, 2000, 111130.
[5] T. Apostol, An Elementary View of Eulers Summation Formula, Am. Math. Monthly 106
(1999) 409-418.
[6] A. Ash, D. Mumford, M. Rapoport and Y. Tai, Smooth Compactication of Locally Symmetric
Varieties, Math. Sci. Press, Brookline, MA, 1975.
[7] P. Aspinwall, B. Greene and D. Morrison, Calabi-Yau moduli space, mirror manifolds and
spacetime topology change in string theory, Nucl. Phys. B416 (1994), 414480.
[8] M. F. Atiyah, On analytic surfaces with double points, Proc. Roy. Soc. London Ser. A 247
(1958) 237-244.
[9] M. F. Atiyah and R. Bott, The moment map and equivariant cohomology, Topology 23 (1984),
128.
[10] M. F. Atiyah and I. G. MacDonald, Introduction to Commutative Algebra, Addison-Wesley,
Reading, MA, 1969.
[11] M. Audin, The Topology of Torus Actions on Symplectic Manifolds, Progress in Math. 93,
Birkh auser Verlag, Basel Boston Berlin, 1991.
[12] M. Barile, On simplicial toric varieties of codimension 2, Rend. Istit. Mat. Univ. Trieste 39
(2007), 942
[13] M. Barile, D. Bernardi, A. Borisov and J.-M. Kantor, On empty lattice simplicies in dimension
4, preprint, 2009. Available electronically at arXiv:0912.5310 [math.AG].
[14] V. Batyrev, On the classication of smooth projective toric varieties, Tohoku Math. J. 43
(1991), 569585.
817
818 Bibliography
[15] V. Batyrev, Variations of the mixed Hodge structure of afne hypersurfaces in algebraic tori,
Duke Math. J. 69 (1993), 349409.
[16] V. Batyrev, Dual polyhedra and mirror symmetry for Calabi-Yau hypersurfaces in toric vari-
eties, J. Algebraic Geom. 3 (1994), 493535.
[17] V. Batyrev, On the classication of toric Fano 4-folds, J. Math. Sci. (New York) 94 (1999),
10211050.
[18] V. Batyrev and L. Borisov, On Calabi-Yau complete intersections in toric varieties, in Higher-
dimensional complex varieties (Trento, 1994), (M. Andreatta and T. Peternell, eds.), de
Gruyter, Berlin, 1996, 3965.
[19] V. Batyrev and D. Cox, On the Hodge structure of projective hypersurfaces in toric varieties,
Duke Math. J. 75 (1994), 293338.
[20] V. Batyrev and D. Mel

nikov, A theorem on nonextendability of toric varieties, Vestnik


Moskov. Univ. Ser. I Mat. Mekh. 1986, 2024, 118; English translation, Moscow Univ. Bull.
41 (1986), 2327.
[21] M. Beck, C. Haase, and F. Sottile, Formulas of Brion, Lawrence, and Varchenko on rational
generating functions for cones, Math. Intelligencer 31 (2009), no. 1, 917.
[22] M. Beck and S. Robins, Computing the Continuous Discretely, Springer, New York, 2007.
[23] A. A. Beilinson, J. Bernstein and P. Deligne, Faisceaux pervers, in Analysis and topology on
singular spaces, I (Luminy, 1981), Ast erisque 100 Soc. Math. France, Paris, 1982, 5171.
[24] F. Berchtold and J. Hausen, Bunches of cones in the divisor class groupa new combinatorial
language for toric varieties, Int. Math. Res. Not. (2004), 261302.
[25] E. Bierstone and P. Milman, Desingularization of toric and binomial varieties, J. Algebraic
Geom. 15 (2006), 443486.
[26] L. J. Billera, The algebra of continuous piecewise polynomials, Adv. Math. 76 (1989), 170
183.
[27] L. J. Billera and C. Lee, Sufciency of McMullens conditions for f -vectors of simplicial poly-
topes, Bull. Amer. Math. Soc. 2 (1980), 181185.
[28] L. J. Billera, P. Filliman and B. Sturmfels, Constructions and complexity of secondary poly-
topes, Adv. Math. 83 (1990), 155179.
[29] L. J. Billera, I. M. Gel

fand and B. Sturmfels, Duality and minors of secondary polyhedra, J.


Combin. Theory Ser. B 57 (1993), 258268.
[30] C. Birkar, Birational Geometry, lecture notes for a course taught at Cambridge University,
Winter, 2007. Available electronically at arXiv:0706.1794 [math.AG].
[31] C. Birkar, P. Cascini, C. D. Hacon and J. M
c
Kernan, Existence of minimal models for varieties
of log general type, J. Amer. Math. Soc. 23 (2010), 405468.
[32] R. Blanco and S. Encinas, Embedded desingularization of toric varieties, preprint, 2009.
Available electronically at arXiv:0901.2211 [math.AG].
[33] M. Blickle, Multiplier ideals and modules on toric varieties, Math. Z. 248 (2004), 113121.
[34] M. Blickle and R. Lazarsfeld, An informal introduction to multiplier ideals, in Trends in Com-
mutative Algebra, (L. Avramov, M. Green and C. Huneke, eds.), Math. Sci. Res. Inst. Publ.
51, Cambridge Univ. Press, Cambridge, 2004, 87114.
[35] R. Blumenhagen, B. Jurke, T. Rahn and H. Roschy, Cohomology of Line Bundles: A Compu-
tational Algorithm, preprint, 2010. Available electronically at arXiv:1003.5217 [math.AG].
[36] E. Bombieri and W. Gubler, Heights in Diophantine Geometry, Cambridge Univ. Press, Cam-
bridge, 2006.
[37] A. Borel, Linear Algebraic Groups, second edition, Graduate Texts in Math. 126, Springer,
New York, 1991.
Bibliography 819
[38] A. Borel, et al., Seminar on transformation groups, Annals of Math. Studies 46, Princeton
Univ. Press, Princeton, 1960.
[39] L. Borisov, L. Chen and G. Smith, The orbifold Chow ring of toric Deligne-Mumford stacks,
J. Amer. Math. Soc. 18 (2005), 193215.
[40] R. Bott, Homogeneous vectors bundles, Ann. Math. 66 (1957), 203248.
[41] R. Bott, Review of Le th eor` eme de Riemann-Roch, Bull. Soc. Math. France 86 (1958), 97136
by A. Borel and J.-P. Serre, in Mathematical Reviews, MR0116022 (22 #6817).
[42] T. Braden, Remarks on the combinatorial intersection cohomology of fans, Pure Appl. Math.
Q. 2 (2006), 11491186.
[43] V. Braun, M. Hampton, A. Novoseltsev, et.al., Sage polyhedral package, available electroni-
cally at www.sagemath.org, 2010.
[44] V. Braun, A. Novoseltsev, Sage toric varieties package, available electronically at www.
sagemath.org, 2010.
[45] G. Bredon, Sheaf Theory, second edition, Springer, New York, 1997.
[46] M. Brion, Points entiers dans les poly edres convexes, Ann. Sci.

Ecole Norm. Sup. 21 (1988),
653663.
[47] M. Brion, Piecewise polynomial functions, convex polytopes and enumerative geometry, in
Parameter spaces (Warsaw, 1994), Banach Center Publ. 36, Polish Acad. Sci., Warsaw, 1996,
2544.
[48] M. Brion, The structure of the polytope algebra, Tohoku Math. J. 49 (1997), 132.
[49] M. Brion, Rational smoothness and xed points of torus actions, Transform. Groups 4 (1999),
127156.
[50] M. Brion and M. Vergne, An equivariant Riemann-Roch theorem for complete, simplicial toric
varieties, J. Reine Angew. Math. 482 (1997), 6792.
[51] A. Brndsted, An Introduction to Convex Polytopes, Springer, New York, 1983.
[52] W. Bruns and J. Gubeladze, Polytopal linear groups, J. Algebra 218 (1999), 715737.
[53] W. Bruns and J. Gubeladze, Polytopes, Rings, and K-Theory, Monographs in Mathematics,
Springer, New York, 2009.
[54] W. Bruns and J. Gubeladze, Semigroup algebras and discrete geometry, in Geometry of toric
varieties, (L. Bonavero and M. Brion, eds.), S emin. Congr. 6, Soc. Math. France, Paris, 2002,
43127.
[55] W. Bruns, J. Gubeladze and N. V. Trung, Normal polytopes, triangulations, and Koszul alge-
bras, J. Reine Angew. Math. 485 (1997), 123160.
[56] W. Brun and J. Herzog, Cohen-Macaulay Rings, Cambridge Univ. Press, Cambridge, 1993.
[57] W. Bruns, B. Ichim and C. S oger, NORMALIZ: Algorithms for rational cones and afne
monoids, available electronically at www.math.uos.de/normaliz/, 2010.
[58] A. Buch, J. F. Thomsen, N. Lauritzen and V. Mehta, The Frobenius morphism on a toric
variety, Tohoku Math. J. 49 (1997), 355366.
[59] V. Buchstaber and N. Ray, An invitation to toric topology: vertex four of a remarkable tetra-
hedron, in Toric topology, (M. Harada, Y. Karshon, M. Masuda and T. Panov, eds.), Contemp.
Math. 460 (2008), Amer. Math. Soc., Providence, RI, 127.
[60] W. Buczy nska, Fake weighted projective spaces, English translation of Toryczne przestrzenie
rzutowe, Magister thesis, 2002. Available electronically at arXiv:0805.1211 [math.AG].
[61] E. Cattani, D. Cox and A. Dickenstein, Residues in toric varieties, Compositio Math. 108
(1997), 3576.
820 Bibliography
[62] Y. Chen and V. Shokurov, Strong rational connectedness of toric varieties, preprint, 2009.
Available electronically at arXiv:0905.1430 [math.AG].
[63] C. H. Clemens and Z. Ran, A new method in Fano geometry, Internat. Math. Res. Notices
(2000), 527549.
[64] T. Coates, A. Corti, H. Iritani and H.-H. Tseng, Computing genus-zero twisted Gromov-Witten
invariants, Duke Math. J. 147 (2009), 377438.
[65] D. Cox, The homogeneous coordinate ring of a toric variety, J. Algebraic Geom. 4 (1995),
1550.
[66] D. Cox, Recent developments in toric geometry, in Algebraic geometrySanta Cruz 1995, (J.
Koll ar, R. Lazarsfeld and D. Morrison, eds.), Proc. Sympos. Pure Math. 62.2, Amer. Math.
Soc., Providence, RI, 1997, 389436.
[67] D. Cox, Update on toric geometry, in Geometry of toric varieties, (L. Bonavero and M. Brion,
eds.), S emin. Congr. 6, Soc. Math. France, Paris, 2002, 141.
[68] D. Cox and S. Katz, Mirror Symmetry and Algebraic Geometry, Mathematical Surveys and
Monographs 68, Amer. Math. Soc., Providence, RI, 1999.
[69] D. Cox, J. Little and D. OShea, Ideals, Varieties and Algorithms, third edition, Undergraduate
Texts in Math., Springer, New York, 2007.
[70] D. Cox, J. Little and D. OShea, Using Algebraic Geometry, second edition, Graduate Texts
in Math. 185, Springer, New York, 2005.
[71] D. Cox and C. von Renesse, Primitive collections and toric varieties, Tohoku Math. J. 61
(2009), 309332.
[72] CoCoATeam, CoCoA Version 4.7.5Computations in commutative algebra, available elec-
tronically at cocoa.dima.unige.it, 1987present.
[73] D. Dais, Resolving 3-dimensional toric singularities, in Geometry of toric varieties, (L.
Bonavero and M. Brion, eds.), S emin. Congr. 6, Soc. Math. France, Paris, 2002, 213225
[74] D. Dais, Geometric combinatorics in the study of compact toric surfaces, in Algebraic and
Geometric Combinatorics, (C. Athanasiadis, V. Batyrev, D. Dais, M. Henk and F. Santos,
eds.), Contemp. Math. 423, Amer. Math. Soc., Providence, RI, 2006, 71123.
[75] D. Dais, U.-U. Haus and M. Henk, On crepant resolutions of 2-parameter series of Gorenstein
cyclic quotient singularities, Results Math. 33 (1998), 208265.
[76] V. Danilov, The geometry of toric varieties, Uspekhi Mat. Nauk. 33 (1978), 85134. English
translation, Russian Math. Surveys 33 (1978), 97154.
[77] V. Danilov and A. Khovanskii, Newton polyhedra and an algorithm for calculating Hodge-
Deligne numbers, Izv. Akad. Nauk SSSR Ser. Mat. 50 (1986), 925945. English translation
Math. USSR-Izv. 29 (1987), 279298.
[78] W. Decker, G.-M. Greuel, G. Pster, G. Sch onemann, Singular Version 3.1.1Commutative
algebra and algebraic geometry, available electronically at www.singular.uni-kl.de,
1989present.
[79] J. De Loera, D. Haws, R. Hemmecke, P. Huggins, J. Tauzer and R. Yoshida, LattE Version
1.2Lattice points in convex polytopes and integer programming, available electronically at
www.math.ucdavis.edu/
~
latte/software.htm, 2003present.
[80] J. De Loera, J. Rambau and F. Santos, Triangulations: Structures for Algorithms and Applica-
tions, Springer, Berlin, 2010.
[81] D. Delzant, Hamiltoniens p eriodiques et images convexes de lapplication moment, Bull. Soc.
Math. France 116 (1988), 315339.
[82] M. Demazure, Sous-groupes alg ebriques de rang maximum du groupe de Cremona, Ann. Sci.

Ecole Norm. Sup. 3 (1970), 507588.


Bibliography 821
[83] I. Dolgachev, Lectures on Invariant Theory, Cambridge Univ. Press, Cambridge, 2003.
[84] M. Drton, B. Sturmfels and S. Sullivant, Lectures on Algebraic Statistics, Oberwolfach Semi-
nars 39, Birkh auser Verlag, Basel Boston Berlin, 2009.
[85] A. Durfee, Fifteen characterizations of rational double points and simple critical points, En-
seign. Math. 25 (1979), 131163.
[86] D. Edidin and W. Graham, Equivariant intersection theory, Invent. Math. 131 (1998), 595
634.
[87] D. Edidin and W. Graham, Riemann-Roch for quotients and Todd classes of simplicial toric
varieties, Comm. Algebra 31 (2003), 37353752.
[88] F. Ehlers, Eine Klasse komplexer Mannigfaltigkeiten und die Au osung einiger isolierter Sin-
gularit aten, Math. Ann. 218 (1975), 127156.
[89] D. Eisenbud, Commutative Algebra with a View Toward Algebraic Geometry, Graduate Texts
in Math. 150, Springer, New York, 1995.
[90] D. Eisenbud and J. Harris, The Geometry of Schemes, Graduate Texts in Math. 197, Springer,
New York, 2000.
[91] D. Eisenbud, M. Mustat a and M. Stillman, Cohomology on toric varieties and local cohomol-
ogy with monomial supports, J. Symbolic Comput. 29 (2000), 583600.
[92] E. J. Elizondo, P. Lima-Filho, F. Sottile and Z. Teitler, Arithmetic toric varieties, preprint,
2010. Available electronically at arXiv:1003.5141 [math.AG].
[93] G. Ewald, Combinatorial Convexity and Algebraic Geometry, Graduate Texts in Math. 168,
Springer, New York, 1996.
[94] G. Ewald and U. Wessels, On the ampleness of invertible sheaves in complete projective toric
varieties, Results Math. 19 (1991), 275278.
[95] E. M. Feichtner, Rational versus real cohomology algebras of low-dimensional toric varieties,
Proc. Amer. Math. Soc. 131 (2003), 16951704.
[96] M. Franz, The integral cohomology of toric manifolds, Proc. Steklov Inst. Math. 252 (2006),
5362.
[97] M. Franz, Describing toric varieties and their equivariant cohomology, Colloq. Math. 121
(2010), 116.
[98] M. Franz, torhom, a Maple package for toric cohomology, available electronically at http:
//www.math.uwo.ca/
~
mfranz/torhom/, 2004.
[99] O. Fujino, Notes on toric varieties from Mori theoretic viewpoint, Tohoku Math. J. 55 (2003),
551564.
[100] O. Fujino, On the Kleiman-Mori cone, Proc. Japan Acad. Ser. A Math. Sci. 81 (2005), 8084.
[101] O. Fujino, Multiplication maps and vanishing theorems for toric varieties, Math. Z. 257
(2007), 631641.
[102] O. Fujino, Vanishing theorems for toric polyhedra, in Higher dimensional algebraic varieties
and vector bundles, (S. Mukai, ed.), RIMS Lecture Notes B9, Res. Inst. Math. Sci. (RIMS),
Kyoto, 2008, 8195.
[103] O. Fujino, Three-dimensional terminal toric ips, Cent. Eur. J. Math. 7 (2009), 4653.
[104] O. Fujino and H. Sato, Introduction to the toric Mori theory, Michigan Math. J. 52 (2004),
649665.
[105] W. Fulton, Introduction to Toric Varieties, Annals of Math. Studies 131, Princeton Univ. Press,
Princeton, NJ, 1993.
[106] W. Fulton, Young Tableaux, with Applications to Representation Theory and Geometry, Cam-
bridge Univ. Press, Cambridge, 1997.
822 Bibliography
[107] W. Fulton, Intersection theory, second edition, Ergebnisse der Mathematik und ihrer Grenzge-
biete, 3. Folge, Band 2, Springer, Berlin, 1998.
[108] W. Fulton, Equivariant Cohomology in Algebraic Geometry, Eilenberg lectures, Columbia
University, Spring 2007, notes by D. Anderson. Available electronically at www.math.lsa.
umich.edu/
~
dandersn/eilenberg.
[109] W. Fulton and R. Lazarsfeld, Connectivity and its applications in algebraic geometry, in Al-
gebraic Geometry: Proceedings University of Illinois at Chicago Circle, 1980, (A. Libgober
and P. Wagreich, eds), Lecture Notes in Math. 862, Springer, Berlin, 1981, 2692.
[110] W. Fulton and R. MacPherson, Categorical framework for the study of singular spaces, Mem.
Amer. Math. Soc. 31, (1981).
[111] W. Fulton and B. Sturmfels, Intersection theory on toric varieties, Topology 36 (1997), 335
353.
[112] I. M. Gel

fand, M. M. Kapranov and A. V. Zelevinsky, Newton polyhedra of principal A-


determinants, Dokl. Akad. Nauk SSSR 308 (1989), 2023; English translation, Soviet Math.
Dokl. 40 (1990), 278281.
[113] I. M. Gel

fand, M. M. Kapranov and A. V. Zelevinsky, Discriminants, Resultants and Multi-


dimensional Determinants, Birkh auser Boston, Boston Basel Berlin, 1994.
[114] E. Gawrilow, M. Joswig, Polymake Version 2.9.8Convex polytopes, polyhedra, simpli-
cial complexes, matroids, fans, and tropical objects, available electronically at wwwopt.
mathematik.tu-darmstadt.de/polymake/doku.php, 1997present.
[115] R. Godement, Th eorie ds Faisceaux, Hermann, Paris, 1964.
[116] R. Goldman and R. Krasauskas, eds., Topics in Algebraic Geometry and Geometric Modeling
(Proceedings of AGGM, Vilnius, Lithuania), Contemp. Math. 334, Amer. Math. Soc., Provi-
dence, RI, 2003.
[117] P. Gonz alez P erez and B. Teissier, Embedded resolutions of non necessarily normal afne
toric varieties, C. R. Math. Acad. Sci. Paris 334 (2002), 379382.
[118] M. Goresky, R. Kottwitz and R. MacPherson, Equivariant cohomology, Koszul duality, and
the localization theorem, Invent. Math. 131 (1998), 2583.
[119] M. Goresky and R. MacPherson, Intersection homology theory, Topology 19 (1980), 135162.
[120] M. Goresky and R. MacPherson, Intersection homology theory II, Invent. Math. 72 (1983),
77129.
[121] R. Goward, A simple algorithm for principalization of monomial ideals, Trans. Amer. Math.
Soc. 357 (2005), 48054812.
[122] H. Grauert and R. Remmert, Coherent Analytic Sheaves, Grundlehren der mathematischen
Wissenschaften 265, Springer, Berlin, 1984.
[123] D. Grayson and M. Stillman, Macaulay2 Version 1.4Algebraic geometry and commutative
algebra, available electronically at www.math.uiuc.edu/Macaulay2, 1992present.
[124] M. Greenberg and J. Harper, Algebraic Topology: A First Course, Benjamin/Cummings, Read-
ing, MA, 1981. Reprinted by Westview Press, Boulder, CO.
[125] P. Grifths and J. Harris, Principles of Algebraic Geometry, Wiley, New York, 1978
[126] A. Grothendieck, Rev etements etales et groupe fondamental (SGA 1), S eminaire de g eom etrie
alg ebrique du Bois Marie 196061 (with two papers by M. Raynaud), Lecture Notes in
Math. 224, Springer, Berlin, 1971. Updated and annotated reprint, Documents Math ematiques
(Paris) 3, Soci et e Math ematique de France, Paris, 2003.
[127] A. Grothendieck,

El ements de g eom etrie alg ebrique (r edig es avec la collaboration de Jean
Dieudonn e), II.

Etude globale el ementaire de quelques classes de morphismes, Inst. Hautes
tudes Sci. Publ. Math. 8 (1961), 5222.
Bibliography 823
[128] B. Gr unbaum, Convex Polytopes, Wiley Interscience, New York, 1967. Second edition pre-
pared by V. Kaibel, V. Klee and G. Ziegler, Springer, 2003.
[129] C. Haase and I. Melnikov, The reexive dimension of a lattice polytope, Ann. Comb. 10 (2006),
211217.
[130] J. Harris, Algebraic Geometry: A First Course, Graduate Texts in Math. 133, Springer, New
York, 1992.
[131] R. Hartshorne, Algebraic Geometry, Graduate Texts in Math. 52, Springer, New York, 1977.
[132] R. Hartshorne, Stable reexive sheaves, Math. Ann. 254 (1980), 121176.
[133] B. Hassett, Introduction to Algebraic Geometry, Cambridge Univ. Press, Cambridge, 2007.
[134] B. Hassett and Y. Tschinkel, Universal torsors and Cox rings, in Arithmetic of Higher-
Dimensional Algebraic Varieties, (B. Poonen and Y. Tschinkel, eds), Progress in Math. 226,
Birkh auser Verlag, Basel Boston Berlin, 2004, 149173.
[135] A. Hatcher, Algebraic Topology, Cambridge Univ. Press, Cambridge, 2002.
[136] A. Hatcher, Spectral Sequences in Algebraic Topology, draft available electronically at www.
math.cornell.edu/
~
hatcher/SSAT/SSATpage.html.
[137] T. Hausel and B. Sturmfels, Toric hyperK ahler varieties, Doc. Math. 7 (2002), 495534.
[138] H. Hauser, The Hironaka Theorem on resolution of singularities (Or: a proof that we always
wanted to understand), Bull. Amer. Math. Soc. 40 (2003), 323403.
[139] D. Hensley, Continued Fractions, World Scientic, Singapore, 2006.
[140] R. Hemmecke, R. Hemmecke and P. Malkin, 4ti2 Version 1.2Computation of Hilbert bases,
Graver bases, toric Gr obner bases, and more, available electronically at www.4ti2.de, 2005.
[141] M. Hering, A. K uronya and S. Payne, Asymptotic cohomological functions of toric divisors,
Adv. Math. 207 (2006), 634645.
[142] L. Hille and M. Perling, A counterexample to Kings conjecture, Compos. Math. 142 (2006),
15071521.
[143] H. Hironaka, Resolution of singularities of an algebraic variety over a eld of characteristic
zero I-II, Ann. Math. 79 (1964), 109326.
[144] F. Hirzebruch, New Methods in Algebraic Geometry, third enlarged edition, Die Grundlehren
der Mathematischen Wissenschaften, Band 131, Springer-Verlag, New York, 1966.
[145] F. Hirzebruch, Hilbert Modular Surfaces, Ens. Math. 19 (1973), 183281.
[146] M. Hochster, Rings of invariants of tori, Cohen-Macaulay rings generated by monomials, and
polytopes, Ann. Math. 96 (1972), 318337.
[147] J. Howald, Multiplier ideals of monomial ideals, Trans. Amer. Math. Soc. 353 (2001), 2665
2671.
[148] S. Hu, Semistable degeneration of toric varieties and their hypersurfaces, Comm. Anal.
Geom. 14 (2006), 5989.
[149] Y. Hu, Combinatorics and quotients of toric varieties, Discrete Comput. Geom. 28 (2002),
151174.
[150] Y. Hu and S. Keel, Mori dream spaces and GIT, Michigan Math. J. 48 (2000), 331348.
[151] K. Hulek, Elementary Algebraic Geometry, Amer. Math. Soc., Providence, RI, 2003.
[152] J. Humphreys, Linear Algebraic Groups, Graduate Texts in Math. 21, Springer, New York,
1975.
[153] N. Ilten, toriccodes.m2Macaulay2 package for divisorial polytopes, toric codes
and T-codes, available electronically at people.cs.uchicago.edu/
~
nilten/files/
toriccodes.m2, 2008present.
824 Bibliography
[154] N. Ilten, Deformations of smooth toric surfaces, Manuscripta Math. to appear. Available elec-
tronically at arXiv:0902.0529 [math.AG].
[155] M.-N. Ishida, Polyhedral Laurent series and Brions equalities, Internat. J. Math. 1 (1990),
251265.
[156] Y. Ito, Minimal resolution via Gr obner basis, in Algebraic Geometry in East Asia, (A.
Ohbuchi, K. Konno, S. Usui, A. Moriwaki and N. Nakayama, eds.), World Sci. Publ., River
Edge, NJ, 2002, 165174.
[157] Y. Ito, Special M
c
Kay Correspondence, in Geometry of toric varieties, (L. Bonavero and M.
Brion, eds.), S emin. Congr. 6, Soc. Math. France, Paris, 2002, 213225
[158] S. Iyengar, G. Leuschke, A. Leykin, C. Miller, E. Miller, A. Singh and U. Walther, Twenty-
four Hours of Local Cohomology, Grad. Stud. Math. 87, Amer. Math. Soc., Providence, RI,
2007.
[159] N. Jacobson, Basic Algebra II, Freeman, San Francisco, 1980.
[160] K. Jaczewski, Generalized Euler sequences and toric varieties, in Classication of Algebraic
Varieties (LAquila, 1992), (C. Ciliberto, E. L. Livorni and A. Sommese, eds.), Contemp.
Math. 162, Amer. Math. Soc., Providence, RI, 1994, 227247.
[161] A. Jensen, Gfan Version 0.4plusGr obner fans and tropical varieties, available electronically
at www.math.tu-berlin.de/
~
jensen/software/gfan/gfan.html, 2005present.
[162] A. Jordan, Homology and Cohomology of Toric Varieties, Ph.D. thesis, University of
Konstanz, 1998. Available electronically at citeseerx.ist.psu.edu/viewdoc/summary?
doi=10.1.1.45.7648.
[163] D. Joyner, GAP toric Version 1.6toric varieties and combinatorial geometry in GAP, avail-
able electronically at www.opensourcemath.org/toric, 2005present.
[164] M. Kapranov, B. Sturmfels and A. Zelevinsky, Quotients of toric varieties, Math. Ann. 290
(1991), 643655.
[165] A. Kasprzyk, Toric Fano 3-folds with terminal singularities, Tohoku Math. J. 58 (2006), 101
121.
[166] A. Kasprzyk, Canonical toric Fano threefolds, Canad. J. Math., to appear. Available electron-
ically at arXiv:0806.2604 [math.AG].
[167] A. Kasprzyk, Bounds on fake weighted projective space, Kodai Math. J. 32 (2009), 197208.
[168] A. Kasprzyk, M. Kreuzer and B. Nill, On the combinatorial classication of toric log del
Pezzo surfaces, LMS J. Comput. Math. 13 (2010), 3346.
[169] A. Katsabekis and A. Thoma, Toric sets and orbits on toric varieties, J. Pure Appl. Algebra
181 (2003), 7583.
[170] K. Kaveh, Note on cohomology rings of spherical varieties and volume polynomial, preprint,
2010. Available electronically at arXiv:math/0312503 [math.AG].
[171] Y. Kawamata, Crepant blowing-up of 3-dimensional canonical singularities and its applica-
tion to degeneration of surfaces, Ann. of Math. 127 (1988), 93163.
[172] G. Kempf, F. Knudsen, D. Mumford and B. Saint-Donat, Toroidal Embeddings I, Lecture
Notes in Math. 339, Springer, Berlin, 1973.
[173] A. G. Khovanskii, Newton polyhedra and toroidal varieties, Funktsional. Anal. i Prilozhen.
11 (1977), 5664; English translation, Functional Anal. Appl. 11 (1977), 289296.
[174] A. G. Khovanskii and A. V. Pukhlikov, The Riemann-Roch theorem for integrals and sums of
quasipolynomials on virtual polyhedra, Algebra i Analiz 4 (1992), 188216, English transla-
tion, St. Petersburg Math J. 4 (1993), 789812.
[175] F. Kirwan and J. Woolf, An Introduction to Intersection Homology Theory, second edition,
Chapman & Hall/CRC, Boca Raton, FL, 2006.
Bibliography 825
[176] S. Kleiman, Toward a numerical theory of ampleness, Ann. of Math. 84 (1966), 293344.
[177] P. Kleinschmidt, A classication of toric varieties with few generators, Aequationes Math. 35
(1988), 254266.
[178] A. Klyachko, Equivariant vector bundles over toric varieties, Izv. Akad. Nauk SSSRSer. Mat.
53 (1989),10011039, 1135; English translation, Math. USSR-Izv. 35 (1990), 337375.
[179] J. Koll ar and S. Mori, Birational Geometry of Algebraic Varieties, Cambridge Univ. Press,
Cambridge, 1998.
[180] S. Krantz, Geometric Function Theory: Explorations in Complex Analysis, Birkh auser
Boston, Boston Basel Berlin, 2005.
[181] A. Kresch, Gromov-Witten invariants of a class of toric varieties, Michigan Math. J. 48 (2000),
369391.
[182] M. Kreuzer and H. Skarke, PALP Version 1.1Lattice polytopes with a view towards physics,
available electronically at hep.itp.tuwien.ac.at/
~
kreuzer/CY/CYpalp.html, 2002
present.
[183] M. Kreuzer and H. Skarke, PALP: A package for analyzing lattice polytopes with applications
to toric geometry, Comput. Phys. Comm. 157 (2004), 87106.
[184] E. Kunz (with contributions by D. Cox and A. Dickenstein), Residues and Duality for Pro-
jective Algebraic Varieties, University Lecture Series 47, Amer. Math. Soc., Providence, RI,
2008.
[185] R. Laterveer, Linear systems on toric varieties, Tohoku Math. J. 48 (1996), 451458.
[186] R. Lazarsfeld, Positivity in Algebraic Geometry, Vols. I and II, Springer, New York, 2004.
[187] Y. Liu, L. E. Trotter and G. Ziegler, On the height of the minimal Hilbert basis, Results Math.
23 (1993), 374376.
[188] J.-L. Loday, Realization of the Stasheff polytope, Arch. Math. 83 (2004), 267278.
[189] I. G. Macdonald, The volume of a lattice polyhedron, Proc. Camb. Phil. Soc. 59 (1963), 719
726.
[190] D. Maclagan and G. Smith, Multigraded Castelnuovo-Mumford regularity, J. Reine Angew.
Math. 571 (2004), 179212.
[191] P. Mani, Spheres with few vertices, Journal of Comb. Theory Ser. A 13 (1992), 346352.
[192] E. Martnez-Moro, C. Munuera and D. Ruano, eds., Advances in Algebraic Geometry Codes,
Ser. Coding Theory Cryptol. 5, World Sci. Publ., Hackensack, NJ, 2008.
[193] E. Materov, The Bott formula for toric varieties, Mosc. Math. J. 2 (2002), 161182, 200.
[194] K. Matsuki, Introduction to the Mori Program, Universitext, Springer, New York, 2002.
[195] H. Matsumura, Commutative Ring Theory, Cambridge Univ. Press, Cambridge, 1986.
[196] A. Mavlyutov, Cohomology of complete intersections in toric varieties, Pacic J. Math. 191
(1999), 133144.
[197] A. Mavlyutov, Semiample hypersurfaces in toric varieties, Duke Math. J. 101 (2000), 85116.
[198] A. Mavlyutov, Cohomology of rational forms and a vanishing theorem on toric varieties, J.
Reine Angew. Math. 615 (2008), 4558.
[199] J. McCleary, A Users Guide to Spectral Sequences, Cambridge Univ. Press, Cambridge, 2001.
[200] M. McConnell, The rational homology of toric varieties is not an combinatorial invariant,
Proc. Amer. Math. Soc. 104 (1989), 986-991.
[201] P. McMullen, On zonotopes, Trans. Amer. Math. Soc. 159 (1971), 91109.
[202] D. McDuff, Symplectic embeddings of 4-dimensional ellipsoids, J. Topol. 2 (2009), 122.
826 Bibliography
[203] L. Mihalcea, Positivity in equivariant quantum Schubert calculus, Am. J. Math. 128 (2006),
787803.
[204] E. Miller and B. Sturmfels, Combinatorial Commutative Algebra, Graduate Texts in Math.
227, Springer, New York, 2005.
[205] K. Miyake and T. Oda, Almost homogeneous algebraic varieties under algebraic torus action,
in ManifoldsTokyo 1973 (Proc. Internat. Conf., Tokyo, 1973), (A. Hattori, ed.), Univ. Tokyo
Press, Tokyo, 1975, 373381.
[206] T. Morita, Hilbert schemes of nite abelian group orbits and Gr obner fans, Hokkaido Math.
J. 38 (2009), 249265.
[207] D. Mumford, Algebraic Geometry I: Complex Projective Varieties, Springer, Berlin, 1976.
[208] D. Mumford, The Red Book of Varieties and Schemes, second expanded edition, Lecture Notes
in Math. 1358, Springer, Berlin, 1999.
[209] D. Mumford, J. Fogarty and F. Kirwan, Geometric Invariant Theory, third edition, Ergebnisse
der Mathematik und ihrer Grenzgebiete 34, Springer, Berlin, 1994.
[210] J. Munkres, Elements of Algebraic Topology, Addison-Wesley, Reading, MA, 1984. Reprinted
by Westview Press, Boulder, CO.
[211] M. Mustat a, Local cohomology at monomial ideals, J. Symbolic Comput. 29 (2000), 709720.
[212] M. Mustat a, Vanishing theorems on toric varieties, Tohoku Math. J. 54 (2002), 451470.
[213] Y. Namikawa, Toroidal compactication of Siegel spaces, Lecture Notes in Math. 812,
Springer, Berlin, 1980.
[214] B. Nill, Gorenstein toric Fano varieties, Manuscripta Math. 116 (2005), 183210.
[215] B. Nill, Volume and lattice points of reexive simplices, Discr. Comp. Geom. 37 (2007), 301
320.
[216] M. bro, An algorithm for the classication of smooth Fano polytopes, preprint, 2007. Avail-
able electronically at arXiv:0704.0049 [math.CO].
[217] T. Oda, Torus Embeddings and Applications. Based on joint work with Katsuya Miyake, Tata
Institute of Fundamental Research, Bombay; Springer, Berlin, 1978.
[218] T. Oda, Convex Bodies and Algebraic Geometry, Ergebnisse der Mathematik und ihrer Gren-
zgebiete, 3. Folge, Band 15, Springer, New York, 1988.
[219] T. Oda, Geometry of toric varieties, in Proceedings of the Hyderabad Conference on Algebraic
Groups, (S. Ramanan, C. Musili and N. Mohan Kumar, eds.), Manoj Prakashan, Madras, 1991,
407440.
[220] T. Oda, Simple convex polytopes and the strong Lefschetz theorem, J. Pure Appl. Algebra 71
(1991), 265286.
[221] T. Oda, Recent developments in the theory of toric varieties, S ugaku 46 (1994), 323335;
English translation, Recent topics on toric varieties, in Selected papers on number theory and
algebraic geometry, Amer. Math. Soc. Transl. Ser. 2, 172, Amer. Math. Soc., Providence, RI,
1996, 7791.
[222] T. Oda and H. S. Park, Linear Gale transforms and Gelfand-Kapranov-Zelevinskij decompo-
sitions, Tohoku Math. J. 43 (1991), 375399.
[223] S. Ogata and K. Nakagawa, On generators of ideals dening projective toric varieties,
Manuscripta Math. 108 (2002), 33-42.
[224] S. Payne, Equivariant Chow cohomology of toric varieties, Math. Res. Lett. 13 (2006), 2941.
[225] S. Payne, Moduli of toric vector bundles, Compos. Math. 144 (2008), 11991213.
[226] M. Perling, Divisorial cohomology vanishing on toric varieties, preprint, 2009. Available elec-
tronically at arXiv:0711.4836v2 [math.AG].
Bibliography 827
[227] C. Peters and J. Steenbrink, Mixed Hodge Structures, Springer, New York, 2008.
[228] J. Pfeie and J. Rambau, Computing triangulations using oriented matroids, in Algebra, Ge-
ometry, and Sofware Systems, (M. Joswig and N. Takayma, eds.), Springer, Berlin, 2003,
4975.
[229] J. Pommersheim, Toric varieties, lattice points and Dedekind sums, Math. Ann. 295 (1993),
124.
[230] B. Poonen and F. Rodriguez-Villegas, Lattice polygons and the number 12, Am. Math.
Monthly 107 (2000), 238250.
[231] P. Popescu-Pampu, The geometry of continued fractions and the topology of surface singular-
ities, in Singularities in geometry and topology 2004, (J.-P. Brasselet and T. Suwa, eds.), Adv.
Stud. Pure Math., 46, Math. Soc. Japan, Tokyo, 2007, 119195.
[232] N. Proudfoot, Geometric invariant theory and projective toric varieties, in Snowbird Lectures
in Algebraic Geometry, (R. Vakil, ed.), Contemp. Math. 388, Amer. Math Soc., Providence,
RI, 2005, 161167.
[233] J. Rambau, TOPCOM Version 0.16.0Triangulations of point congurations, available elec-
tronically at www.rambau.wm.uni-bayreuth.de/TOPCOM/, 1987present.
[234] J. Rambau, TOPCOM: Triangulations of point congurations and oriented matroids, in Math-
ematical Software - ICMS 2002, (A. Cohen, X. Gao and N. Takayama, eds.), World Scientic
2002, 330340.
[235] M. Reid, Canonical 3-folds, in Journ ees de G eometrie Alg ebrique dAngers, Juillet 1979/Al-
gebraic Geometry, Angers, 1979, (A. Beauville, ed.), Sijthoff & Noordhoff, Alphen aan den
Rijn, 1980, 273310.
[236] M. Reid, Decomposition of Toric Morphisms, in Arithmetic and Geometry, (M. Artin and J.
Tate, eds.), Progress in Math. 36, Birkh auser Verlag, Basel Boston Berlin, 1983, 395418.
[237] M. Reid, Review of Geometry of convex bodies and algebraic geometry, S ugaku 33 (1981),
120133 by T. Oda, in Mathematical Reviews, MR0631487 (83f:14010).
[238] M. Reid, Undergraduate Algebraic Geometry, LMS Student Texts 12, Cambridge Univ. Press,
Cambridge, 1989.
[239] E. Reyes, R. Villarreal and L. Z arate, A note on afne toric varieties, Linear Algebra Appl.
318 (2000), 173179.
[240] J. Richter-Gebert, B. Sturmfels and T. Theobald, First steps in tropical geometry, in Idem-
potent Mathematics and Mathematical Physics, (G. Litvinov and V. Maslov, eds.), Contemp.
Math. 377, Amer. Math. Soc., Providence, RI, 2003.
[241] R. T. Rockafellar, Convex Analysis, Princeton Univ. Press, Princeton, NJ, 1970.
[242] J. J. Rotman, Advanced Modern Algebra, Prentice-Hall, Upper Saddle River, NJ, 2002.
[243] I. Satake, On the arithmetic of tube domains (blowing-up of the point at innity), Bull. Amer.
Math. Soc. 79 (1973), 10761094.
[244] H. Sato, Toward the classication of higher-dimensional toric Fano varieties, Tohoku Math.
J. 52 (2000), 383413.
[245] I. Shafarevich, Basic Algebraic Geometry, Volumes 1 and 2, Springer, New York, 1994 and
1996.
[246] H. Schenck, Computational Algebraic Geometry, Cambridge Univ. Press, Cambridge, 2003.
[247] J.-P. Serre, Faisceaux alg ebriques coh erents, Ann. Math. 61, (1955), 197278.
[248] J.-P. Serre, G eom etrie alg ebrique et g eom etrie analytique, Ann. Inst. Fourier (Grenoble) 6
(1956), 142.
[249] G. Shephard, Decomposable convex polyhedra, Mathematika 10 (1963), 8995.
828 Bibliography
[250] A. Simis, W. Vasconcelos and R. Villarreal, On the ideal theory of graphs, J. Algebra 167
(1994), 389416.
[251] N. J. A. Sloane, On-line Encyclopedia of Integers Sequences, 2010. Published electronically
at the web site oeis.org.
[252] G. G. Smith, Macaulay2 toric varieties package, available electronically at www.math.uiuc.
edu/Macaulay2, 2010.
[253] K. Smith, L. Kahanp a a, P. Kek al ainen and W. Traves, An Invitation to Algebraic Geometry,
Universitext, Springer, New York, 2000.
[254] F. Sottile, Toric ideals, real toric varieties, and the algebraic moment map, in Topics in Al-
gebraic Geometry and Geometric Modeling (Proceedings of AGGM, Vilnius, Lithuania), (R.
Goldman and R

Krasuaskas eds), Contemp. Math. 334, Amer. Math. Soc., Providence, RI,
2003, 225240. Corrected version available electronically at arXiv:math/0212044 [math.AG].
[255] E. Spanier, Algebraic Topology, McGraw-Hill, New York, 1966. Reprinted by Springer, New
York, 1994.
[256] T. A. Springer, Linear Algebraic Groups, second edition, Progress in Math. 9, Birkh auser
Verlag, Basel Boston Berlin, 1998.
[257] R. Stanley, The number of faces of a simplicial convex polytope, Adv. Math. 35 (1980), 236
238.
[258] R. Stanley, Generalized H-vectors, intersection cohomology of toric varieties, and related
results, in Commutative algebra and combinatorics (Kyoto 1985), (M. Nagata and H. Mat-
sumura, eds.), Adv. Stud. Pure Math. 11, North-Holland, Amsterdam, 1987, 187213.
[259] H. Stark, An Introduction to Number Theory, MIT Press, Cambridge, MA, 1978.
[260] J. Steenbrink, Intersection form for quasi-homogeneous singularities, Compositio Math. 34
(1977), 211223.
[261] J. Steenbrink, Mixed Hodge structure on the vanishing cohomology, in Real and complex
singularities (Proc. Ninth Nordic Summer School/NAVF Sympos. Math., Oslo, 1976), Sijthoff
and Noordhoff, Alphen aan den Rijn, 1977, 525563.
[262] W. Stein et.al., Sage Mathematics Software, Version 2.6Free open-source mathematics soft-
ware system, available electronically at www.sagemath.org, 2004present.
[263] R. S. Strichartz, The Way of Analysis, Jones and Bartlett, Boston, MA, 1995.
[264] B. Sturmfels, Gr obner Bases and Convex Polytopes, University Lecture Series 8, Amer. Math.
Soc., Providence, RI, 1996.
[265] H. Sumihiro, Equivariant completion, I, II, J. Math. Kyoto Univ. 14 (1974), 128; 15 (1975),
573605.
[266] I. Swanson and C. Huneke, Integral Closure of Ideals, Rings, and Modules, LMS Lecture
Note Series 336, Cambridge Univ. Press, Cambridge, 2006.
[267] B. Tessier, Du th eor` eme de lindex de Hodge aux in egalit es isop erim etriques, C. R. Acad. Sci.
Paris S er. A-B 288 (1979), A287A289.
[268] B. Tessier, Monomial ideals, binomial ideals, polynomial ideals, in Trends in Commutative
Algebra, (L. Avramov, M. Green and C. Huneke, eds.), Math. Sci. Res. Inst. Publ. 51, Cam-
bridge Univ. Press, Cambridge, 2004, 211246.
[269] M. Thaddeus, Toric quotients and ips, in Topology, Geometry, and Field Theory, (K. Fukaya,
M. Furuta, T. Kohno and D. Kotschick, eds.), World Scientic Publ., River Edge, NJ, 1994,
193213.
[270] H. Thompson, Comments on toric varieties, preprint, 2003. Available electronically at
arXiv:math/0310336 [math.AG].
Bibliography 829
[271] T. tomDieck, Transformation Groups, de Gruyter Studies in Mathematics 8, Walter de Gruyter
& Co., Berlin, 1987.
[272] B. Totaro, The Chow ring of a classifying space, in Algebraic K-theory (Seattle, WA, 1997),
(W. Raskind and C. Weibel, eds.), Proc. Sympos. Pure Math. 67, Amer. Math. Soc., Provi-
dence, RI, 1999, 249281.
[273] K. Ueno, Algebraic Geometry 2, Amer. Math. Soc., Providence, RI, 2001.
[274] R. Villarreal, Monomial Algebras, Marcel Dekker, Inc., New York, 2001.
[275] C. von Renesse, Combinatorial Aspects of Toric Varieties, Ph. D. Thesis, University of
Massachusetts, 2007. Available electronically at /www.wsc.ma.edu/renesse/research/
UMTHESIS.pdf.
[276] C. Weibel, An Introduction to Homological Algebra, Cambridge Univ. Press, Cambridge,
1994.
[277] E. Witten, Phases of N =2 theories in two dimensions, Nucl. Phys. B403 (1993), 159222.
[278] J. Wodarczyk, Decomposition of birational toric maps in blow-ups & blow-downs, Trans.
Amer. Math. Soc. 349 (1997), 373411.
[279] J. Wodarczyk, Toroidal varieties and the weak factorization theorem, Invent. Math. 154
(2003), 223331.
[280] O. Zariski and P. Samuel, Commutative Algebra, Volume II, Van Nostrand, Princeton, 1960.
Corrected reprint, Springer, New York, 1976.
[281] G. Ziegler, Lectures on Polytopes, Graduate Texts in Math. 152, Springer, 1995.
Index
absolutely ample, 345
abstract variety, 97, 99
adjunction formula, 366, 503
afne
piece of a variety, 51, 52
scheme, 4, 104
toric variety, 12, 18
of a cone, 30
variety, 3
afne cone
of a projective toric variety, 56
of a projective variety, 50
afne hyperplane, 63
afne morphism, 397
afne semigroup, 16
Alexander duality, 420, 673
Alexandrov-Fenchel inequality, 792
algebraic action, 12, 106, 199
algebraic cycle, 612
rationally equivalent to zero, 612
algebraic statistics, 796
algebro-geometric imperialism, 794
almost geometric quotient, 201, 202, 210, 682
almost homogeneous variety, 791
ample, see Cartier divisor, ample
absolutely, 345
relatively with respect to f , 345
analytic variety, 547
annihilator, 416
arithmetically Cohen-Macaulay (aCM), 415
associahedron, 743, 745, 749, 800
augmentation ideal, 654
-basis, 734
barycentric subdivision, 525
basepoint free, see also sheaf, generated by global
sections
Cartier divisor, see Cartier divisor, basepoint
free
subspace of global sections, 257
basic simplex, 68
Bernoulli numbers, 626, 663
Bertini theorem, 509
Betti numbers, 587, 589, 621
binomial, 16
birational transform of a divisor, 557, 768, 769
Birkhoff polytope, 64
bivariant intersection theory, 675
blowup, 98, 105, 111, 130, 131, 174, 176, 188,
190, 277, 286, 319, 591
of an ideal, 535
Borel construction, 596, 598, 599
for a torus, 596
Borel-Moore homology, 672
Bott formula, 439, 455
branches, 562
Brions equalities, 632, 636, 644, 664, 665
for a polytope, 637
bunch of cones, 738
Calabi-Yau variety, 529
cancellative semigroup, 22
canonical bundle, 356
canonical class, 356
canonical divisor, 356, 373
of a toric variety, 366, 791
relative, 541
canonical module
of a projective toric variety, 371
of an afne toric variety, 371
of the total coordinate ring, 369
831
832 Index
canonical sheaf, 356, 365
of a toric variety, 366
canonical singularities, 549
Carath eodorys theorem, 69, 537
Cartan matrix, 500
Cartesian diagram, 675
Cartier data, 181, 193, 791
Cartier divisor, 160
ample, 268, 269, 271, 305, 788
basepoint free, 258, 262, 266, 269, 291, 329,
726
big, 427, 656
f -ample, 331, 772
f -very ample, 331
local data of, 160
nef, 291, 305, 726
numerical equivalence of, 292
numerically equivalent to zero, 292
of a polyhedron, 322
of a polytope, 182
on an irreducible variety, 257
torus-invariant, see torus-invariant, Cartier
divisor
very ample, 268, 269

Cech cochains, 389

Cech cohomology, 389391

Cech Complex
toric, 398

Cech complex, 389, 390


ceiling , 424
central hyperplane arrangement, 285
chamber, 718, 725
character, 11
generic, 707, 718
Chern character, 623, 625, 629, 631, 809
equivariant, 642
Chern class, 607, 624
of a locally free sheaf, 624
of a smooth variety, 624
of a toric variety, 625
total, 624
Chern roots, 626, 631
of a locally free sheaf, 631
Chevalleys theorem, 144
Chow group, 612
Chow polytope, 744
Chow ring, 613
of a toric variety, 616, 791
rational, 613
circuits, 701, 751
oriented, 751
class group, 161, 350
classical topology, 4, 50, 102, 113, 139, 142
closed half-space, 25, 63
CoCoA, 797
codimension of a prime ideal, 157
coding theory, 796
Cohen-Macaulay variety, 394
toric, 415, 790
coherent sheaf, 166
cohomology class, 593, 673
equivariant, 601, 608
rened, 593, 673
combinatorially equivalent polytopes, 65
compact, 113, 139, 142
compatibility conditions, 97, 252
compatible map of lattices, 125
complete fan, 113, 139, 142, 193, 788
complete linear system, 261
complete variety, 141, 143
toric, 142
complex, 388
exact sequence of, 388, 389
Koszul, 403, 408
map of, 388
computer algebra
4ti2, see 4ti2
GAP, see GAP
Gfan, see Gfan
LattE, see LattE
Macaulay2, see Macaulay2
Magma, see Magma
Normaliz, see Normaliz
PALP, see PALP
Polymake, see Polymake
Sage, see Sage
Singular, see Singular
TOPCOM, see TOPCOM
cone
convex polyhedral, see polyhedral cone
Gorenstein, 379
of a polyhedron, 319, 325
of a polytope, 24, 70, 86
polyhedral, see polyhedral cone
rational polyhedral, see rational polyhedral cone
reexive Gorenstein, 379
simplicial, see simplicial cone
smooth, see smooth cone
strongly convex, see strongly convex cone
connecting homomorphism, 388
conormal sheaf, 355
constant sheaf, 396
constructible set, 123
continued fraction
Felix Klein, 476
Hirzebruch-Jung, 468, 469, 473
ordinary, 475, 476
contractible, 396
convergents, 469, 475
convex
function, 264
hull, 24
polyhedral cone, see polyhedral cone
convex support of full dimension, 265
Index 833
coordinate ring, 3, 4
homogeneous, 49, 372
total, see total coordinate ring
cotangent bundle, 352
of a toric variety, 452
cotangent sheaf, 352
of a toric variety, 359
Cousin problem, 392
Cox ring, 795
Cremona group, 787
crepant morphism, 553
crepant resolution, 527, 529
cup product, 593
D-op, 782, 783
Dedekind sums, 653
deformation retract, 396
degree
of a divisor on a curve, 287
of a line bundle on a curve, 288
of a projective variety, 655
Dehn-Sommerville equations, 436, 438, 439, 590,
621
depth, 394, 416
derivation, 357
torus-invariant, 364
universal, 357
derived functor, 389, 444, 445
determinantal variety, 13
diagonal map, 102
dimension at a point, 6, 101
dimension of a cone, 24
dimension of a module, 416
dimension of a polytope, 63
dimension of a ring, 156
direct image, 166, 392
higher, 393
left exact, 364
direct limit, 100, 245
directed set, 182, 245
directed system, 245
discrepancy, 549
discrete valuation, 155
discrete valuation ring (DVR), 155
distinguished point, 116, 118, 130, 135
division algorithm, 475
modied, 460, 467
divisor
anticanonical, 379
toric, 380
canonical, see canonical divisor
Cartier, see Cartier divisor
linear equivalence of, 160, 350
locally principal, see locally principal divisor
nef, see Cartier divisor, nef
numerically effective, see Cartier divisor, nef
of a character, 171
of a rational function, 159
of poles, 160
of zeros, 160
principal, 159
Weil, see Weil divisor
divisorial wall, 750
double dual, 347, 350
double point, 462
doubly-stochastic matrix, 64
Du Val singularity, 462, 501
dual cone, 24, 799
dual face, 26
dual polytope, 65
dual sheaf, 255, 347, 350
dualizing sheaf, 356
dualizing complex, 395
Dynkin diagram, 462, 467, 500
edge, 25, 63
edge ideal of a graph, 539
effective divisor, 159
Ehrhart polynomial, 432, 508, 655, 791, 805
of a polygon, 433, 434
of a polytopal complex, 434, 442
Ehrhart quasipolynomial, 443
Ehrhart reciprocity, 432, 436, 791
elementary ip, see ip, elementary
elementry symmetric polynomial, 626
Enriques-Kodaira classication, 762
equalizer, 95
equivariant cohomology, 596
of a toric variety, 601
equivariant Euler class, 648, 654
equivariant Gysin map, 646, 675, 676
equivariant integral, 643, 676
decomposition of, 647
equivariant map, 41, 126
equivariant Riemann-Roch, 623
for a toric variety, 641
proof of, 648
simplicial, 651, 652
equivariant Todd class, 642
simplicial, 652
equivariant vector bundle, 607
equivariantly formal, 599, 606
Euclidean algorithm, 468, 475
modifed, 468
Euler characteristic
decomposition in equivariant case, 643, 645
equivariant, 641
of a sheaf, 430, 440, 623
of a topological space, 442, 505, 586, 625
of a toric variety, 586
Euler class, 624
Euler relation, 365
generalized, 365
Euler sequence, 363, 365
834 Index
generalized, 364
toric, 363
Euler-Maclaurin formula, 663, 667
eventail, 787
exact sequence
long, 389
of complexes, 388
of sheaves, 248
exceptional curve of the rst kind, 498
exceptional divisor, 464
exceptional locus, 521, 762
exceptional set, 207
of a character, 698
expected dimension, 683, 704, 710, 720
Ext, 395, 445
exterior product, see wedge product
extremal wall, see wall, extremal
extremal contraction, 762, 763
divisorial, 764, 766, 774, 776
bering, 764, 766, 776
ipping, 764, 766, 770, 773, 774, 776
structure of, 764, 766
extremal ray, 295, 306, 727, 762, 770
D-negative, 772, 777
extremal walls, 295
f -ample, see Cartier divisor, f -ample
f -very ample, see Cartier divisor, f -very ample
4ti2, 72, 797
f.r.p.p decomposition, 789
face, 25, 63
face numbers, 435, 620, 793
facet, 25, 63
normal, 26, 64
presentation, 66
virtual, 696, 697
fake weighted projective spaces, 765
fan, 77, 106, 787, 789
complete, see complete fan
degenerate, 278, 690, 719, 720, 749, 766, 777
generalized, 278, 690, 691, 712, 719, 725, 726
innite, 790
nondegenerate, 278
normal, see normal fan
renement of, see renement of a fan
simplicial, see simplicial fan
smooth, see smooth fan
Fano type, 559
toric, 559
Fano variety, see variety, Fano
ber bundle, 133
ber product of varieties, 102
Fibonacci number, 22
nite quotient singularities, 113, 547
nitely generated semigroup, 16
xed component, 730
xed points, 598, 645
ip, 748, 761, 762
elementary, 760, 770
Proj construction, 772, 774
termination of, 777779, 781
toric, 770
ipping wall, 750, 757
oor , 424
op, 782, 783
formal power series, 10
formal semigroup module, 632
summable elements of, 633
fractional ideal, 170
sheaf of, 540, 790
full dimensional
convex support, 265
polyhedron, 319
polytope, 64
function eld, 100
functor, 388
derived, 389
fundamental class, 673
fundamental group
of a toric variety, 566, 791
of a torus, 562
of an afne toric variety, 565
G-Hilbert scheme, 493, 495
(G, )-invariant polynomial, 678, 679
GAGA, 142, 144
Gale duality, 452, 699701, 711, 734, 741
GAP, 797
toric package, 797
gauged linear sigma model, 795
generalized Euler sequence, 624
generalized Gysin map, 630, 672674, see also
equivariant Gysin map
generated by global sections, see sheaf, generated
by global sections
geometric case, 729, 744, 748, 749
primtive, 731
geometric invariant theory, see GIT
geometric modeling, 796
geometric quotient, 200, 202, 210, 682
almost, see almost geometric quotient
Gfan, 488, 797, 802
GIT, 677, 678, see also expected dimension
general quotient, 684, 699
quotient, 682, 692, 725
GKZ cone, 713717, 725, 728
facet of, 749
irrelevant ideal of, 720
GKZ decomposition, 718
global sections, 189
of a sheaf, 166, 248
of a toric sheaf, 190, 193
of a vector bundle, 251
gluing data, 97, 252
Index 835
good categorical quotient, 198202, 682
Gordans Lemma, 30
Gorenstein Fano variety, see variety, Gorenstein
Fano
Gorenstein ring, 660
Gr obner basis, 485, 796
marked, 485
reduced, 485
universal, 485
Gr obner fan, 486, 745, 802
resolution of singularities, 490
graded module, 226, 248, 310
shift of, 226
graded polynomial ring, 219, 695
Grassmannian, 254, 260, 699
Grothendieck-Riemann-Roch, 629, 630
Gysin map, 674
1om sheaf, 249, 255
h-polynomial, 621
hard Lefschetz theorem, 620, 621, 793
Hausdorff topological space, 102
height of a prime ideal, 157
higher direct image, 393
Hilbert basis, 32, 33, 473, 474, 799, 803
Hilbert basis theorem, 3
Hilbert polynomial, 655, 803
of a projective variety, 432
of a sheaf, 431, 440
Hirzebruch surface, 112, 175, 191, 263, 267, 273,
303, 338, 368, 408, 425, 454, 498, 509, 595,
600, 602, 611, 629, 638, 662, 696, 704, 731,
733, 787, 806, 807, 809
Hirzebruch-Riemann-Roch, 623, 628, 629, 656,
791, 809
proof of, 649
proof via Khovanskii-Pukhlikov, 669
simplicial, 653
Hodge decomposition, 440
Hodge index theorem, 792
homogeneous coordinate ring, see coordinate ring,
homogeneous
homogeneous coordinates, 49
homogenization, 192
homomorphism of sheaves, 166, 246
injective, 246
sujective, 246
Hopf bration, 575
Iitaka dimension, 427
image sheaf, 247
index of a Q-Gorenstein toric variety, 555
index of a simplicial cone, see multiplicity of a
simplicial cone
injective resolution, 387, 388
injective sheaf, 387
injectivity lemma, 411, 412, 424, 430
inner normal fan, see normal fan
integer programming, 796
integral closure, 5
of an ideal, 537, 540
integral of a cohomology class, 592, 623
equivariant, see equivariant integral
integrally closed ideal, 536
integrally closed ring, 5
intersection cohomology, 620, 621, 793
Betti numbers, 621
intersection homology, 618, 619
intersection product, 288, 593
on a toric surface, 497
on a toric variety, 289, 300
inverse limit, 182
inverse system, 182
invertible sheaf, 168, 254, 255
inward-pointing facet normal, see facet, normal
irreducible components, 97
irreducible variety, 97
irrelevant ideal, 207
of a character, 698
of a GKZ cone, 720
Ishida complex, 378, 423
isomorphism of varieties, 96
Jacobian matrix, 7, 351
Jurkiewicz-Danilov
ring, 593, 595
theorem, 594, 595, 606
k-cycle, see algebraic cycle
K ahler differentials, 351, 352, 354
K unneth formula, 597
Kawamata log terminal singularities, see klt
singularities
kernel sheaf, 247
Khovanskii-Pukhlikov theorem, 666, 667
proof of, 668
relation to HRR, 669
Kleinschmidts classication theorem, 341, 738,
794
klt singularities, 557, 781
Koszul complex, 403, 408
Krull principal ideal theorem, 161, 169
LattE, 797
lattice, 13
lattice polyhedron, 319
lattice equivalence, 381, 461
lattice ideal, 15
lattice polytope, 66
empty, 555
Laurent polynomial, 5
LDP polygon, 386
index of, 386
left exact, 389
836 Index
Leray-Hirsch theorem, 599
limit of one-parameter subgroup, 115, 116, 139,
142, 144
line bundle, 253, 255, 348, see also Cartier divisor
ample, 268
f -ample, 331
f -very ample, 331
linearized, 678
pullback of, 258
very ample, 268
linear equivalence, see divisor, linear equivalence
of
linearized line bundle, see line bundle, linearized
linearly equivalent divisors, 256
local

Cech complex, 444
local analytic equivalence, 547
local cohomology, 444447
local data, 160, 253, 257
toric, 181
local duality, 798
local ring, 6, 156
at a point, 6, 9, 94, 99
at a prime divisor, 157
localization, 5, 9
homogeneous, 317
localization theorem, 600, 643
locally principal divisor, 160
locally irreducible, 563
locally trival ber bundle, 134
log canonical singularities, 557, 781
log del Pezzo surface, 386
index of, 386
log minimal model, 782, 783
log pair, 557
log resolution, 526, 541
logarithmic 1-forms, 359, 360
logarithmic poles, 360
long exact sequence, 444
lower hull, 740
Macaulay2, 275, 407, 445, 639, 797, 798, 804,
805
BGG package, 807
BoijSoederberg package, 807
NormalToricVarieties package, 797, 806,
810
Polyhedra package, 275, 799
toriccodes package, 797
Magma, 797
maximal Cohen-Macaulay (MCM), 416
maximal projective crepant partial
desingularization, 529
maximal spectrum, 4
M
c
Kay correspondence, 491, 493
McMullen conjecture, 620, 793
minimal generator, 30
minimal model, 762
log, 782, 783
minimal model program, see MMP
minimal surface, 762
Minkowski sum, 65, 192, 318, 792
mirror symmetry, 384, 745, 795
mixed volume, 791, 792
MMP, 762, 781
equivariant, 777
log toric, 781, 783
relative, 777
toric, 772, 776
moment map
algebraic, 571
symplectic, 571, 574
monomial Briancon-Skoda theorem, 540
monomial ideal, 540
Mori cone, 293
of a toric variety, 294, 305, 727
Mori bration, 762
morphism of varieties, 3, 95
projective, see projective morphism
proper, see proper morphism
moving cone
of a toric variety, 731
of a variety, 730
of the secondary fan, 728, 729, 738
multipliciative subset, 9
multiplicity of a simplicial cone, 300, 465, 519,
520
properties of, 519
multiplier ideal, 542, 545
toric, 542, 545
multiplier module, 545
N-Minkowski summand, 283
Nakayamas Lemma, 169
nef cone, 293
of a toric variety, 294, 726, 728, 731
nef divisor, see Cartier divisor, nef and Q-Cartier
divisor, nef
Newton polytope, 188, 744, 791
nilpotents, 4, 8, 9, 48, 103, 105
Noethers theorem, 505, 511, 628
Noetherian, 156
nonassociative product, 744, 745
nonequivariant limit, 599, 601, 623, 642
nonnormal toric variety, 150
nonsingular point, 6
normal
afne toric variety, 37
polyhedron, 322
polytope, 68, 86, 372, 803
toric variety, 86, 107
variety, 5, 100
normal bundle, 355, 631, 648, 654
normal fan
of a polyhedron, 321, 690, 691
Index 837
of a polytope, 76, 77, 108, 279, 690, 792
normal ring, 5, 156, 157
normal sheaf, 355, 366
Normaliz, 33, 39, 72, 434, 797, 803805
normalization, 5, 151
of a projective toric variety, 153
of an afne toric variety, 39, 152
of an irreducible curve, 288
normalized basis, 460
normalized volume, 433, 655
Nullstellensatz, 3
numerically effective divisor, see Cartier divisor,
nef and Q-Cartier divisor, nef
numerically equivalent
divisors, see Cartier divisor, numerical
equivalence of
proper 1-cycles, see proper 1-cycle, numerical
equivalence of
to zero, see Cartier divisor, numerically
equivalent to zero and proper 1-cycle,
numerically equivalent to zero
one-parameter subgroup, 11
orbifold, 46, 113, 547
orbit closure, 121, 135
Orbit-Cone Correspondence, 119, 789
nonnormal case, 154
order of vanishing, 158
orientation coefcient, 580
oriented circuit, see circuits, oriented
oriented matroids, 701
p-allowable, 619
pair (X, D), 557, 781
PALP, 797
parameters of a cone, 460
partial quotients, 469, 475
p-Ehrhart polynomial, 435, 512
perfect eld, 48
permutation matrix, 55
perversity, 619
complementary, 619
maximal, 619
middle, 619
minimal, 619
Picard group, 161, 256, 350
of a toric variety, 176, 577
Picks formula, 434
piecewise polynomial functions, 606, 607
Poincar e duality, 593, 620, 673
Poincar e polynomial, 592
Poincar e residue, 360, 361
pointed afne semigroup, 36
polar polytope, 65
pole, 159
polyhedral cone, 23
polyhedron, 190, 318
augmented, 711
full dimensional, see full dimensional,
polyhedron
lattice, see lattice polyhedron
normal, see normal, polyhedron
of a character, 687, 688
of a torus-invariant divisor, 190, 266, 328
very ample, see very ample, polyhedron
Polymake, 797
polynomial splines, 607
polytopal complex, 434
polytope, 24, 63
combinatorially equivalent, see combinatorially
equivalent polytopes
full dimensional, see full dimensional, polytope
lattice, see lattice polytope
normal, see normal, polytope
simple, see simple polytope
simplicial, see simplicial polytope
smooth, see smooth polytope
very ample, see very ample, polytope
pre-variety, 102
presheaf, 165, 247
primary decomposition, 161
prime divisor, 157
primitive collection, 304, 340, 772
primitive relation, 305, 340, 772
principal A-determinant, 744
principal divisor, see divisor, principal
principal ideal domain (PID), 156, 157
principalization, 541
product variety, 7, 53, 101, 313
class group of, 175
toric, 47, 90, 111
Proj, 317, 322, 372, 682, 690, 691
projective bundle
of a coherent sheaf, 318
of a locally free sheaf, 316
of a vector bundle, 316
toric, 337, 338
projective morphism, 314, 315, 346
projective space, 49
projective toric variety, 55
projective variety, 49
projective with respect to a line bundle, 314, 345
projectively normal variety, 61, 86, 441, 798
proper 1-cycle, 292
numerical equivalence of, 292
numerically equivalent to zero, 292
proper continuous map, 142, 144
proper face, 25
proper morphism, 143, 144, 315
pseudoeffective cone, 730
of a toric variety, 731
Puiseux series, 188
pullback, see line bundle, pullback of
pullback of a torus-invariant Cartier divisor, 281
838 Index
Q-Cartier divisor, 180
nef, 409
Q-divisor, 424
Q-factorial, 46
Q-Fano variety, see variety, Q-Fano
quadratic reciprocity, 792
quasicoherent sheaf, 166
quasicompact, 104
quasiprojective toric variety, 330
quasiprojective variety, 313, 315
quotient
almost geometric, see almost geometric quotient
geometric, see geometric quotient
good categorical, see good categorical quotient
Q-Weil divisor, 424
round down of, 424
round up of, 424
rank
of a coherent sheaf, 348
of a nitely generated abelian group, 524
of a module, 366
rational double point, 462, 464, 467, 500
rational function, 51, 100
rational normal cone, 13, 32, 38, 40, 46, 50, 178,
461, 463
rational normal curve, 50, 57
rational normal scroll, 84, 112
rational polyhedral cone, 29
rational singularities, 546
toric, 546
rationally smooth, 546, 592
ray generator, 29
R-divisor, 424
real projective plane, 72
real torus, 561, 570
recession cone, 318
reduced cohomology, 397
reductive group, 201
Rees algebra, 534
rened cohomology class, see cohomology class,
rened
renement of a fan, 130, 373
reexive polygon, 382, 510
reexive polytope, 81, 87, 380, 381, 529, 795
classication, 384
reexive sheaf, 168, 347
of rank 1, 348
regular cone, see smooth cone
regular fan, 113
regular local ring, 7, 157
regular map, 93, 98
regular representation, 491
regular sequence, 394
relative case, 727, 733
relative cohomology, 405
relative interior, 27
relatively ample with respect to f , 345
representation ring, 654
representations of nite groups, 491
special, 492
resolution of singularities, 463, 513
crepant, 511, 527
embedded, 530
Gr obner fan, 490
Hironaka, 513
log, 526
minimal, 499
projective, 513
SNC, 526
toric, 520
toric surface, 465, 470
restriction of a divisor, 160
restriction of a sheaf, 166
Riemann-Roch, see also
Hirzebruch-Riemann-Roch
curves, 503, 628
surfaces, 504, 628
topological part, 505, 628
ring of invariants, 45
ringed space, 95, 99
rooted binary tree, 744, 745
Sage, 522, 629, 639, 797
PALP package, 797
Polyhedra package, 797
TOPCOM package, 800
ToricVarieties package, 797, 808, 809
saturated afne semigroup, 37
secondary fan, 712, 717, 725, 741, 744, 795
chamber of, see chamber
moving cone of, see moving cone, of the
secondary fan
properties of, 718
secondary polyhedron, 744
secondary polytope, 741, 743, 744, 749, 800
section
of a sheaf, 165
of a vector bundle, 251
sectional genus, 509, 794
of a toric surface, 509
Segre embedding, 52
self-intersection, 303
semi-stable reduction, 790
semicomplete toric variety, 734
semigroup, 16
semigroup algebra, 16
of a cone, 30
of a polyhedron, 322, 691
semigroup homomorphism, 35, 116
semiprojective toric variety, 332, 691, 695, 712
semistable points, 679, 697, 698, 702, 707
separated variety, 102, 204
separating transcendence basis, 48
Index 839
separation lemma, 28, 107
Serre duality, 395, 420
for a toric variety, 416, 791
MCM sheaves, 416
set-theoretic complete intersection, 22
sheaf, 95
constant, 255
generated by global sections, 249, 258, 346
locally constant, 255
locally free, 250
of 1-forms, 352
of O
X
-modules, 165, 245
of p-forms, 355
of a graded module, 227, 248, 310
of a torus-invariant divisor, 189
global sections of, 189, 190, 193
of sections of a vector bundle, 252
of Zariski p-forms, 355
sheaf cohomology, 387, 388, 391
long exact sequence, 389
toric, 402, 447, 809
toric surface, 406
sheacation, 247
sheet, see spectral sequence, sheet of
simple normal crossings, see SNC divisor
simple polytope, 64, 434
simplex, 64
simplicial cone, 30
multiplicity of, see multiplicity of a simplicial
cone
simplicial fan, 113
simplicial polytope, 64
simplicial toric variety, 180
simplicialization, 518, 520
Singular, 797, 804
singular cohomology, 395
of a simplicial toric variety, 588
of a toric variety, 582
of an afne toric variety, 577
reduced, 397
relative, 405
ring, see singular cohomology ring
with compact supports, 582
singular cohomology ring, 396
of a toric variety, 594, 595, 660, 791
of an afne toric variety, 577
singular locus, 163, 513
of a normal variety, 163
of a toric variety, 514
singular point, 6
of a toric surface, 459
singularities, see also toric singularities
canonical, see canonical singularities
nite quotient, see nite quotient singularities
klt, see klt singularities
log canonical, see log canonical singularities
terminal, see terminal singularities
small contraction, 762
Smith normal form, 173
smooth cone, 30, 40
smooth fan, 113
smooth Fano polytope, 385
smooth locus, 513
smooth point, 6, 101
smooth polytope, 87
smooth toric variety, 40, 87, 113, 179
smooth variety, 6, 101
SNC divisor, 360, 525
span of a cone, 24
Spec, 4, 317
spectral sequence, 393, 811
convergent, 812
degenerates, 813
differentials, 811
E
1
, 812
E
2
, 812
edge homomorphism, 814
Ext, 417, 815
rst quadrant, 812
Leray, 393, 814
of a closed cover, 404, 815
of a ltered space, 582, 816
of an open cover, 394, 578, 815
Serre, 598, 599, 816
sheet of, 811
singular cohomology of a toric variety, 582, 587
spherical dual, 574
spherical variety, 661
splitting fan, 133, see also weak splitting, 137
stable points, 679, 702, 707
stalk of a sheaf, 246, 257
standard n-simplex n, 66
Stanley-Reisner ideal, 593, 600
Stanley-Reisner ring, 600, 601
star subdivision, 130, 132, 515, 518, 520, 748, 749
properties of, 515, 517
Stasheff polytope, 743
strictly convex, see support function, of a Cartier
divisor, strictly convex
strong Oda conjecture, 790
strongly convex cone, 28
structure sheaf, 95, 99
analytic, 547
sublattice of nite index, 44
subvariety, 97
sum function, 636
Sumihiros theorem, 108, 151, 789
supplementary cone, 478, 479
isomorphic to dual cone, 480
support function, 183, 713, 725, 791
integral with respect to a lattice, 183
of a Q-Cartier divisor, 408
of a Cartier divisor, 184, 263
convex, 265, 266, 329, 725
840 Index
strictly convex, 271, 329331, 334, 518
of a polytope, 186
support of a divisor, 159
support of a fan, 106, 113, 788
supporting afne hyperplane, 63
supporting half-space, 25
supporting hyperplane, 25
Sylvester sequence, 92
symbolic powers, 545
symmetric algebra, 597
symplectic geometry, 574
symplectic reduction, 575
2-neighborly polytope, 73
tangent bundle, 352
tangent sheaf, 352
tautological bundle, 254, 256, 260
Taylor resolution, 448, 449
tensor product, 8, 48
of sheaves, 249, 255, 350
tent analogy, 264, 267
terminal lemma, 556
terminal singularities, 549, 762
Todd class, 623, 626, 630
equivariant, 642
mock, 653
of a simplicial toric variety, 653
of a smooth toric variety, 627, 809
Todd operators, 664
Todd polynomial, 627
TOPCOM, 797, 800
topological pseudomanifold, 618
torhom, 586
toric Chow lemma, 275
toric cone theorem, 294, 727
toric bration, 134
toric ideal, 15, 796
toric Kleiman criterion, 291
toric minimal model progam, see MMP, toric
toric morphism, 41, 42, 125, 135
projective, 330, 334, 790
proper, 144, 790
toric set, 21
toric singularities
canonical, 550552, 554
nite quotient, 548
klt, 558, 781
log canonical, 558, 781
simplicial, 548
terminal, 551, 554, 555, 774
toric surface
resolution of singularities, 465, 470
singular points, 459
smooth classication, 495, 496, 790
toric topology, 574
toric variety, 106
afne, see afne, toric variety
complete, see complete variety, toric
name, 791, 792
nonnegative points of, 568
normal, see normal, toric variety
of a fan, 107
of a polyhedron, 321, 692
of a polytope, 83, 108
positive points of, 568
projective, see projective toric variety
quasiprojective, see quasiprojective toric variety
semicomplete, 734
semiprojective, see semiprojective toric variety
topological model of, 574
toroidal variety, 790
torsion submodule, 444
torsion-free semigroup, 22
torus, 5, 10
embedding, 108, 791, 792
of a projective toric variety, 58
of an afne toric variety, 13
orbit, 118
torus-invariant
Cartier divisor, 176
prime divisor, 170
Weil divisor, 172, 175
total coordinate ring, 207, 218, 248, 309, 362, 446,
795
transportation polytope, 64
triangulation, 739, 744, 745, 761
regular, 740, 743, 745
tropical polynomial, 188
tropical variety, 188
tropicalization, 188
twist of a sheaf, 421
unibranch, 563
unipolar variety, 765
unique factorization domain (UFD), 7
universal coefcient theorem, 598
universal covering space, 563
universally closed, 143
valuative criterion for properness, 148
Van Kampen theorem, 566
Vandermonde identity, 439, 442
vanishing theorem
Batyrev-Borisov, 413, 415, 425
Bott-Steenbrink-Danilov, 421, 423, 791
Demazure, 410, 425, 788, 791
Grauert-Riemenschneider, 428
higher direct images, 410
Kawamata-Viehweg, 427
Kodaira, 421, 423
Mavlyutov, 424
Mustat a, 426
Nakano, 421, 423
Serre afne, 390
Index 841
Serre projective, 392
Zariski p-forms, 423, 440
variety
abstract, see abstract variety
afne, see afne, variety
complete, see complete variety, see complete
variety
Fano, 379
Gorenstein, 373
Gorenstein Fano, 379
toric, 380, 810
irreducible, see irreducible variety
normal, see normal, variety
of general type, 763
projective, see projective variety
projectively normal, see projectively normal
variety
Q-Fano, 765
quasiprojective, see quasiprojective variety
separated, see separated variety
toric, see toric variety
vector bundle, 250
chart of, 250
decomposable, 337
ber above a point, 250, 257
sheaf of sections of, 252
toric, 335, 337
transition functions of, 250, 252
trivialization of, 250
Veronese embedding, 84, 689
Veronese subring, 689, 690
vertex, 63
very ample
divisor, see Cartier divisor, very ample
polyhedron, 322
polytope, 71, 75, 87, 269, 803
volume polynomial, 654, 658660
reduced, 659, 660
wall, 265, 301
extremal, 306
wall crossing, 747, 749, 750, 757
wall relation, 301, 751
weak Oda conjecture, 790
weak splitting, 137, 766
wedge product, 355, 366
weighted homogeneous polynomial, 53
weighted projective space, 53, 112, 176, 188, 651,
653, 657, 808
Weil divisor, 159, 348
torus-invariant, see torus-invariant, Weil divisor
Zariski 1-forms, 361
on a simplicial toric variety, 363
on a toric variety, 361
Zariski p-forms, 355, 416, 421, 423, 424, 435, 440
on a simplicial toric variety, 378
on a toric variety, 375
on an afne toric variety, 376
Zariski closure, 4
Zariski tangent space, 6, 101, 354
Zariski topology, 4, 50, 97
Zariskis main theorem, 563
zonotope, 81, 285, 792

Вам также может понравиться