11.01.2014 Views

Equivariant Embeddings of Algebraic Groups

Equivariant Embeddings of Algebraic Groups

Equivariant Embeddings of Algebraic Groups

SHOW MORE
SHOW LESS

You also want an ePaper? Increase the reach of your titles

YUMPU automatically turns print PDFs into web optimized ePapers that Google loves.

c○ Copyright by David Charles Murphy, 2004


EQUIVARIANT EMBEDDINGS OF ALGEBRAIC GROUPS<br />

BY<br />

DAVID CHARLES MURPHY<br />

B.A., Western Michigan University, 1996<br />

M.S., University <strong>of</strong> Illinois at Urbana-Champaign, 1998<br />

DISSERTATION<br />

Submitted in partial fulfillment <strong>of</strong> the requirements<br />

for the degree <strong>of</strong> Doctor <strong>of</strong> Philosophy in Mathematics<br />

in the Graduate College <strong>of</strong> the<br />

University <strong>of</strong> Illinois at Urbana-Champaign, 2004<br />

Urbana, Illinois


Abstract<br />

We classify embeddings <strong>of</strong> algebraic groups as open orbits in affine varieties, generalizing results<br />

from toric geometry to connected reductive groups. In particular, we show that an embedding is<br />

determined by the set <strong>of</strong> one-parameter subgroups that have a limit in the embedding. We then<br />

investigate the structure <strong>of</strong> such sets <strong>of</strong> one-parameter subgroups and how their properties reflect<br />

properties <strong>of</strong> the associated embedding.<br />

iii


To my parents in gratitude for all <strong>of</strong> their support and encouragement<br />

iv


Acknowledgements<br />

This thesis would not have been possible without the help, patience, and wisdom <strong>of</strong> my advisers,<br />

Robert M. Fossum and William J. Haboush. Specifically, I would like to thank Pr<strong>of</strong>essor Haboush<br />

for his great help in developing my understanding <strong>of</strong> the algebraic geometry and representation<br />

theory used in this dissertation and Pr<strong>of</strong>essor Fossum for his algebraic insights.<br />

I also wish to<br />

thank the other members <strong>of</strong> my committee, Pr<strong>of</strong>essors Maarten Bergvelt, Phillip Griffith and Rinat<br />

Kedem.<br />

I would like to thank Pr<strong>of</strong>essor Griffith, in his capacity as the Director <strong>of</strong> Graduate Studies. He<br />

and his staff, Lori Dick and Marci Blocher, have constantly helped and encouraged me during my<br />

years as a graduate student.<br />

Finally, I am very grateful to my fellow graduate student and friend, Joshua Mullet, for the<br />

many conversations we have had about our work and the new ideas these discussions generated.<br />

v


Table <strong>of</strong> Contents<br />

Chapter 1 Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1<br />

1.1 Toric varieties . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 2<br />

1.2 Luna–Vust theory and spherical varieties . . . . . . . . . . . . . . . . . . . . . . . . . 3<br />

1.3 The “wonderful compactification” . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 6<br />

1.4 Affine group embeddings . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 7<br />

Chapter 2 Toric varieties and group embeddings . . . . . . . . . . . . . . . . . . . . . . . . 11<br />

2.1 Toric varieties as group embeddings . . . . . . . . . . . . . . . . . . . . . . . . . . . 12<br />

2.1.1 Affine toric varieties and one-parameter subgroups . . . . . . . . . . . . . . . 12<br />

2.1.2 Toric varieties from fans . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 14<br />

2.1.3 The torus action . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 15<br />

2.1.4 <strong>Equivariant</strong> divisors in toric varieties . . . . . . . . . . . . . . . . . . . . . . . 16<br />

2.2 Toric varieties in group embeddings . . . . . . . . . . . . . . . . . . . . . . . . . . . . 17<br />

2.2.1 Regular compactifications . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 17<br />

2.2.2 A family <strong>of</strong> left-equivariant SL n -embeddings . . . . . . . . . . . . . . . . . . 19<br />

2.3 Group embeddings from toric varieties . . . . . . . . . . . . . . . . . . . . . . . . . . 22<br />

2.3.1 Mumford’s group embeddings . . . . . . . . . . . . . . . . . . . . . . . . . . . 22<br />

2.3.2 Group embeddings constructed using flag varieties . . . . . . . . . . . . . . . 23<br />

Chapter 3 Affine group embeddings . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 34<br />

3.1 Group actions on affine varieties . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 34<br />

3.2 Classification <strong>of</strong> affine group embeddings . . . . . . . . . . . . . . . . . . . . . . . . . 39<br />

3.2.1 Limits <strong>of</strong> one-parameter subgroups . . . . . . . . . . . . . . . . . . . . . . . . 39<br />

3.2.2 States and strongly convex lattice cones . . . . . . . . . . . . . . . . . . . . . 48<br />

3.2.3 Classification theorem . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 54<br />

3.3 Functoriality <strong>of</strong> our classification . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 58<br />

3.3.1 <strong>Equivariant</strong> morphisms between affine G-embeddings . . . . . . . . . . . . . . 58<br />

3.3.2 Biequivariant resolutions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 60<br />

References . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 67<br />

Vita . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 71<br />

vi


Chapter 1<br />

Introduction<br />

A basic problem in algebraic geometry is the study <strong>of</strong> algebraic actions. Even actions with a dense<br />

orbit are not well understood. A quasihomogeneous variety is a normal G-variety X possessing an<br />

open orbit isomorphic to G/H, for some closed subgroup H <strong>of</strong> G. Toric varieties are an important<br />

family <strong>of</strong> quasihomogeneous varieties, where G is an algebraic torus and H is its trivial subgroup.<br />

A partial classification <strong>of</strong> quasihomogeneous varieties was obtained in the important paper <strong>of</strong> Luna<br />

and Vust [30], written twenty years ago.<br />

Their classification seeks to generalize that <strong>of</strong> toric<br />

varieties, but it is only feasible in special cases.<br />

The goal <strong>of</strong> this dissertation is to solve the<br />

equivariant classification problem for one case not covered in [30], namely the classification <strong>of</strong><br />

quasihomogeneous varieties in which H is trivial.<br />

In this introduction, we review the history <strong>of</strong> and some <strong>of</strong> the results that are essential to our<br />

problem. As mentioned above, toric varieties are the natural place to begin, so we discuss them<br />

in the first section. After that, we present a short overview <strong>of</strong> Luna and Vust’s paper [30] and<br />

some <strong>of</strong> the work that has been done in an attempt to exploit the Luna–Vust classification in cases<br />

that initially appear to be outside the scope <strong>of</strong> their original work. This chapter concludes with an<br />

overview <strong>of</strong> our results.<br />

First, we establish a few conventions that will be used throughout this thesis and notation<br />

that we use freely hereafter. We will always work over a ground field k, which we assume to be<br />

algebraically closed and <strong>of</strong> characteristic zero. The term variety will refer to a separated, integral<br />

scheme <strong>of</strong> finite type over k. All algebraic groups are assumed to be linear and defined over k, and<br />

will be denoted by letters such as G and H. In particular, G will refer to a connected reductive<br />

linear algebraic group defined over k, unless otherwise stipulated.<br />

1


1.1 Toric varieties<br />

The theory <strong>of</strong> toric varieties is a well-established and valuable class <strong>of</strong> examples for the study<br />

<strong>of</strong> equivariant algebraic geometry. Toric varieties were introduced by Demazure [13] in the early<br />

1970s, in which he classified all smooth toric varieties. The classification <strong>of</strong> all toric varieties was<br />

accomplished three years later by Kempf–Knudsen–Mumford–Saint-Donat [28] using fans in the<br />

vector space associated to the set <strong>of</strong> one-parameter subgroups <strong>of</strong> a torus.<br />

We begin by recalling the definition <strong>of</strong> toric varieties. Suppose T is an algebraic torus. Since<br />

we are assuming that our ground field k is algebraically closed, T ∼ = G r m for some r, where G m =<br />

Spec k[x, x −1 ] is the multiplicative group scheme <strong>of</strong> k.<br />

Definition 1. A toric variety is a normal variety X containing an algebraic torus T as an open<br />

subset such that the translations <strong>of</strong> T on itself extend to give an action <strong>of</strong> T on X.<br />

Example 1 (A toric variety and its fan). Consider T = G 2 m. Then X = P 2 is a toric variety for<br />

T as it is normal, contains T as the open subvariety {[t 0 : t 1 : t 2 ] : t 0 , t 1 , t 2 ≠ 0}, and has T -action<br />

given by [t 0 : t 1 : t 2 ] · [x 0 : x 1 : x 2 ] = [t 0 x 0 : t 1 x 1 : t 2 x 2 ]. There are seven open affine T -subvarieties<br />

<strong>of</strong> X, which are glued together along common open affine subvarieties within X. This is recorded<br />

by the “fan” below, consisting <strong>of</strong> seven cones: 3 are two dimensional (cones 1 , 2 , 3 ), 3 are one<br />

dimensional (cones 4 , 5 , 6 ) and one is zero dimensional (cone 7 ).<br />

• • 5 • 1<br />

2 • • • •<br />

• • 7 • 4<br />

• • • • •<br />

<br />

6 • • 3 •<br />

The two dimensional cones each correspond to T -stable affine spaces in X, U i = {[x 0 : x 1 : x 2 ] :<br />

2


x i ≠ 0} for i = 0, 1, 2, the one dimensional cones correspond to the T -stable affine subvarieties<br />

U ij = {[x 0 : x 1 : x 2 ] : x i , x j ≠ 0}, and the zero dimensional cone corresponds to T itself, viewed as<br />

the subvariety {[t 0 : t 1 : t 2 ] : t 0 , t 1 , t 2 ≠ 0}.<br />

Toric varieties play an important role both in algebraic and combinatorial geometry due to the<br />

nature <strong>of</strong> their classification. In particular, toric varieties are useful for the calculation <strong>of</strong> the number<br />

<strong>of</strong> lattice points in polytopes in combinatorial geometry. Their applications in algebraic geometry<br />

are numerous as well, especially as a “source <strong>of</strong> interesting, yet manageable examples” [17]. Toric<br />

varieties also play a critical role in mirror symmetry [11].<br />

Due to the relationship between algebraic geometry and combinatorial geometry afforded by<br />

the classification <strong>of</strong> toric varieties, many have recently preferred to introduce and even define toric<br />

varieties not in terms <strong>of</strong> the torus within, but instead in terms <strong>of</strong> their combinatorial classification.<br />

However, this separation from the roots <strong>of</strong> the subject has had negative consequences. By studying<br />

toric varieties as varieties whose transition functions are monomials or via Cox’s homogeneous coordinate<br />

ring [10], as many now do, the role toric varieties play as part <strong>of</strong> the equivariant classification<br />

program has been diminished.<br />

In the introduction to [12], after motivating the importance <strong>of</strong> toric varieties and their relationship<br />

with combinatorial geometry, Danilov returns to their original definition and proposes “An<br />

extension <strong>of</strong> this theory would seem possible, in which the torus T is replaced by an arbitrary<br />

reductive group G” ([12], page 101).<br />

1.2 Luna–Vust theory and spherical varieties<br />

In [30], Luna and Vust began a program to describe up to G-isomorphism varieties admitting an<br />

action <strong>of</strong> an algebraic group G. A birational classification <strong>of</strong> G-varieties (i.e., classifying how G acts<br />

on fields) may be obtained using Galois cohomology [38]. The second part <strong>of</strong> the problem requires<br />

one to describe all G-models <strong>of</strong> a given function field K with G acting birationally on K.<br />

Let K be a field <strong>of</strong> finite type over k. A model <strong>of</strong> K is a variety X with k(X) = K. We remark<br />

that all models <strong>of</strong> K may be glued together to form a large (non-Noetherian, non-separated) scheme<br />

X = X(K/k). The group G acts on X in a natural way, but this is not an action in the category <strong>of</strong><br />

3


schemes. However, there is a largest open subset X G ⊆ X such that G acts on X G as an algebraic<br />

group.<br />

In the paper [30], G is assumed to be reductive and <strong>of</strong> simply connected type (that is, G is the<br />

direct product <strong>of</strong> a torus and a simply connected semisimple group). Furthermore, all G-models<br />

are assumed to be normal.<br />

The following theorem <strong>of</strong> Sumihiro is the basic tool for the local study <strong>of</strong> G-varieties.<br />

Theorem 1 ([42], Lemma 8). Let G be a connected linear algebraic group and let X be a normal<br />

G-variety. Then any point x ∈ X has a G-stable neighborhood X 0 ⊆ X admitting an equivariant<br />

locally closed embedding in a projective space.<br />

Therefore, every regular action <strong>of</strong> a connected linear algebraic group on a normal variety is<br />

obtained by patching finitely many linear actions on normal quasi-projective varieties. Thus, if we<br />

are interested in local properties <strong>of</strong> a G-model X in a neighborhood <strong>of</strong> a point x or a subvariety Y ,<br />

we may assume that X is quasi-projective.<br />

Let X be a quasihomogeneous G-variety with open orbit G/H and k(X) = k(G/H) = K. Let<br />

B be a Borel subgroup <strong>of</strong> G. We introduce the following terminology.<br />

The complexity <strong>of</strong> the variety X or <strong>of</strong> the field K, denoted c(X) or c(K) respectively, is the<br />

transcendence degree <strong>of</strong> K B .<br />

Hence c(X) is equal to the codimension <strong>of</strong> a B-orbit in general<br />

position on X.<br />

The set <strong>of</strong> G-stable discrete Q-valued valuations <strong>of</strong> K/k that are geometric (that is, correspond<br />

to prime divisors on a suitable model <strong>of</strong> K) is denoted V = V(K). Elements <strong>of</strong> V are called G-<br />

valuations. Let D = D(G/H) be the set <strong>of</strong> irreducible subvarieties <strong>of</strong> G/H <strong>of</strong> codimension one that<br />

are not G-stable. The valuation corresponding to a divisor D ∈ D is denoted by v D . Prime divisors<br />

that are B-stable but not G-stable are called B-divisors. The set <strong>of</strong> B-divisors is denoted D B .<br />

Suppose X is a G-model <strong>of</strong> K, so k(X) = K. Then the complexity measures the codimension<br />

<strong>of</strong> a B-orbit in general position on X. When the complexity is ≤ 1, D B is particularly simple. If<br />

c(X) = 0, G has an open orbit in X, say G/H. In this case, D B is the finite set <strong>of</strong> the irreducible<br />

components <strong>of</strong> the complement <strong>of</strong> the open orbit <strong>of</strong> B in G/H. Suppose c(X) = 1. Let G/H be a<br />

sufficiently generic orbit and let U be the open subset <strong>of</strong> G/H formed <strong>of</strong> B-orbits <strong>of</strong> codimension<br />

1. Then D ∈ D B implies either D is the closure <strong>of</strong> a B-orbit in U or D is a component <strong>of</strong> the<br />

4


complement <strong>of</strong> U in G/H.<br />

Remark 1 ([30], page 220). The Luna–Vust classification strategy is only feasible when c(K) ≤ 1.<br />

Let P be the set <strong>of</strong> f ∈ k(G) which are simultaneously eigenvectors <strong>of</strong> B (acting by left<br />

translations) and <strong>of</strong> H (acting by right translations).<br />

f = g ∏ D∈D B f v D(f)<br />

D<br />

, where g is some element <strong>of</strong> k[G] × .<br />

Any f ∈ P may be written in the form<br />

In order to state the classification result <strong>of</strong> [30] when c(G/H) = 0, we need two more ideas.<br />

Let X ∗ (H) be the group <strong>of</strong> characters <strong>of</strong> H.<br />

If f ∈ k(G) is an eigenfunction <strong>of</strong> H, let χ f<br />

be its character so χ f ∈ X ∗ (H). For E ⊂ k(G), let X ∗ E (H) be the set <strong>of</strong> χ ∈ X∗ (H) such that<br />

E χ ≠ 0, where E χ is the set <strong>of</strong> H-eigenvectors in E <strong>of</strong> character χ. In order to describe the final<br />

requirement <strong>of</strong> Luna and Vust’s result, we focus on two such E’s. The first is E = k[G] and the<br />

second is E = P(E) = {f ∈ P : v D (f) = 0, for all D ∈ E} for a subset E <strong>of</strong> D B .<br />

Denote by V the finite dimensional Q-vector space obtained by tensoring with Q the free Z-<br />

module (P ∩k(G/H))/G m . The elements <strong>of</strong> V and the v D , D ∈ D B , may be thought <strong>of</strong> as elements<br />

in the dual vector space V ∗ . If E ⊂ D B and W ⊂ V, let C(W, E) be the convex cone in V ∗ generated<br />

by W and {v D : D ∈ E}.<br />

Proposition 1 ([30], Proposition 8.10). Let H be a closed subgroup <strong>of</strong> G such that c(G/H) = 0.<br />

Let E ⊂ D B and let W be a finite subset <strong>of</strong> V. Then there exists a G/H-embedding X and a closed G-<br />

subvariety Y <strong>of</strong> X such that W is the set <strong>of</strong> valuations corresponding to the irreducible components<br />

<strong>of</strong> X − G/H which are <strong>of</strong> codimension 1 in X containing Y and E is the set <strong>of</strong> v D ∈ D B such that<br />

the closure <strong>of</strong> D in X contains Y if and only if the following conditions are satisfied:<br />

1. The cone C(W, E) is strongly convex.<br />

2. The lines Qw (w ∈ W) are the extremal lines <strong>of</strong> C(W, E) and do not coincide with any <strong>of</strong> the<br />

Qv D (D ∈ E).<br />

3. C(W, E) 0 ∩ V ≠ ∅, where C 0 denotes the relative interior <strong>of</strong> a cone C.<br />

4. The group <strong>of</strong> characters <strong>of</strong> H is generated, as a monoid, by X ∗ k[G] (H) and X∗ P(E) (H).<br />

Moreover, in terms <strong>of</strong> cones and sets as above, a pair (X ′ , Y ′ ) is an open G-subvariety <strong>of</strong> (X, Y )<br />

5


(i.e., O X ′ ,Y ′ is a localization <strong>of</strong> O X,Y ) if and only if C X ′ ,Y ′ is a face <strong>of</strong> C X,Y and E X ′ ,Y ′ = {D ∈<br />

E X,Y : v D ∈ C X ′ ,Y ′}.<br />

Therefore, to determine a G-model X is the same thing as to determine the set <strong>of</strong> its pairs<br />

(V Y , DY B ) as Y varies over the closed G-subvarieties <strong>of</strong> X, as then we may glue cones together to<br />

form colored fans (with “colors” being the sets D B Y<br />

⊂ DB ) and in this way characterize G/Hembeddings<br />

in general.<br />

Varieties <strong>of</strong> complexity zero are called spherical. These include flag varieties (and horospherical<br />

varieties in general), toric varieties, and symmetric varieties. The theory <strong>of</strong> these spaces can be<br />

unified with the concept <strong>of</strong> spherical varieties.<br />

For example, all three types <strong>of</strong> spaces have a<br />

nice compactification theory, respectively due to Pauer (horospherical varieties); Demazure, Kempf<br />

et al., Miyake–Oda, etc. (toric varieties); and De Concini–Procesi, etc. (symmetric varieties).<br />

Spherical varieties have been extensively studied over the past twenty years, so Knop reworked<br />

and improved the results <strong>of</strong> [30] in the special case <strong>of</strong> spherical varieties, extending the Luna–Vust<br />

classification to positive characteristic [29]. There are a number <strong>of</strong> results concerning the spherical<br />

subgroups <strong>of</strong> G, cohomology computations, local structure theorems, the B-orbit structure, and<br />

relations to symplectic geometry. Here is just one example:<br />

Proposition 2 ([30], Proposition 7.5). Suppose c(K) = 0. Then the number <strong>of</strong> G-orbits in any<br />

G-model <strong>of</strong> K is finite. In fact, any such G-model can only have finitely many B-orbits.<br />

1.3 The “wonderful compactification”<br />

A partial extension <strong>of</strong> toric geometry to embeddings <strong>of</strong> reductive groups may be obtained by viewing<br />

the group G as a homogeneous variety for G×G under left and right translations [6], [7], [8], [9], [46].<br />

In this way, all G × G-equivariant embeddings <strong>of</strong> G are spherical, so may be classified by means <strong>of</strong><br />

the Luna–Vust theory described above. For instance, the wonderful compactification <strong>of</strong> an adjoint<br />

group, introduced and studied by De Concini and Procesi [7], [8], [9], is a special case <strong>of</strong> a G × G-<br />

spherical compactification <strong>of</strong> G.<br />

The wonderful compactification <strong>of</strong> an adjoint group G is constructed as follows. For the adjoint<br />

action <strong>of</strong> G × G in its Lie algebra g ⊕ g, the isotropy group <strong>of</strong> the diagonal diag(g ⊕ g) is equal to<br />

6


G. The wonderful compactification <strong>of</strong> G is the closure <strong>of</strong> the (G × G)-orbit <strong>of</strong> diag(g ⊕ g) in the<br />

Grassmannian Grass dim G (g ⊕ g).<br />

Example 2 (The wonderful compactication <strong>of</strong> P GL n ). For G = P GL n , P(M n ) is its wonderful<br />

compactification.<br />

Example 3 (A spherical compactification <strong>of</strong> SL 2 ). The group SL 2 has a nice biequivariant<br />

compactification, namely the quadric in P 4 , Q = {[M : t] ∈ P(M 2 (k) × k) : det M = t 2 } with<br />

SL 2 → Q given by M ↦→ [M : 1] and the action is by left and right multiplications on the first<br />

coordinate, (g 1 , g 2 ) · [M : t] = [g 1 Mg −1<br />

2 : t]. This is a spherical compactification <strong>of</strong> SL 2 .<br />

In [5], Brion gave a combinatorial classification <strong>of</strong> regular compactifications [3] <strong>of</strong> semi-simple<br />

groups in terms subdivisions <strong>of</strong> a Weyl chamber for the group, which is simpler than the Luna–Vust<br />

classification for regular compactifications as complexity zero (G×G)-varieties. In general, however,<br />

the complexity <strong>of</strong> G as a G-variety is too large (i.e., c(G) > 1) for the Luna–Vust approach to be<br />

effective and, without a right G-action, Brion’s classification doesn’t apply. Thus the systematic<br />

study <strong>of</strong> all equivariant group embeddings proposed by Danilov has been left virtually undone for<br />

the past twenty five years.<br />

1.4 Affine group embeddings<br />

We study normal varieties X with a left G-action which possess an open subset isomorphic to an<br />

algebraic group G so that the action <strong>of</strong> G on itself by (left) translations extends to the action <strong>of</strong><br />

G on all <strong>of</strong> X. As noted in [30], the combinatorial machinery <strong>of</strong> Luna and Vust cannot be easily<br />

applied in this situation (as the complexity will be larger than 1 for most groups G). In its place,<br />

we use the structure theory <strong>of</strong> the group G, and in particular its set <strong>of</strong> one-parameter subgroups,<br />

X ∗ (G), to reduce the problem to classifying combinatorial data that we naturally attach to an<br />

embedding <strong>of</strong> the group.<br />

Let G be a connected reductive algebraic group defined over an algebraically closed field k. By a<br />

G-variety, we mean a k-variety X together with a morphism ρ : G × X → X, written (g, x) ↦→ g · x,<br />

satisfying g 1 · (g 2 · x) = (g 1 g 2 ) · x and e · x = x for all g 1 , g 2 ∈ G and all x ∈ X, where e ∈ G denotes<br />

the identity element. We are interested in the following class <strong>of</strong> G-varieties.<br />

7


Definition 2. A G-embedding is a normal G-variety X that contains an open orbit Ω isomorphic<br />

to G. The closed subvariety ∂X = X − Ω is called the boundary <strong>of</strong> X. Since Ω is an affine G-orbit,<br />

∂X is a G-stable divisor <strong>of</strong> X unless Ω = X.<br />

Any toric variety is a T -embedding, in our terminology. Similarly, the wonderful compactification<br />

<strong>of</strong> an adjoint group G defined by De Concini and Procesi [7] is a G-embedding. All <strong>of</strong> these<br />

examples have both a left and a right G-action. Yet our definition <strong>of</strong> a G-embedding allows the<br />

consideration <strong>of</strong> G-varieties which only possess a left action <strong>of</strong> the group. For instance, if Z is a<br />

toric variety for a maximal torus T <strong>of</strong> a reductive group G, then the variety X = G × T Z is a<br />

normal G-variety for the action g · [h, z] = [gh, z], and the orbit G · [e G , e T ] ⊂ X is open and isomorphic<br />

to G. However, we can’t define a right action <strong>of</strong> G on X since we can’t do so in a way that<br />

preserves the equivalence relation (g, z) ∼ (gt −1 , tz) defined by the diagonal action <strong>of</strong> T on G × Z.<br />

Thus, our definition <strong>of</strong> G-embeddings includes all <strong>of</strong> the biequivariant compactifications already<br />

in the literature [3], [6], [7], [28], [46] and many more. In Chapter 2, we discuss these biequivariant<br />

G-embeddings as well as other constructions <strong>of</strong> G-embeddings derived from toric varieties.<br />

An important consequence <strong>of</strong> this discussion is the pro<strong>of</strong> that non-isomorphic G-embeddings can<br />

produce isomorphic toric varieties T , which is impossible in the (G × G)-equivariant case. This<br />

is accomplished by our construction <strong>of</strong> a family <strong>of</strong> pairwise non-isomorphic SL n -embeddings in<br />

Section 2.2.2 in which two or more <strong>of</strong> the induced toric varieties T are isomorphic whenever n ≥ 4.<br />

Suppose X is a G-embedding. Identify G with the open orbit Ω in X by selecting a base point<br />

x 0 ∈ Ω and considering the isomorphism G → Ω given by g ↦→ g · x 0 . The choice <strong>of</strong> base point<br />

x 0 ∈ Ω, and hence the identification <strong>of</strong> G with Ω, is not canonical. If x ′ is any other point <strong>of</strong> Ω,<br />

then it could serve just as well to identify G with Ω. For such x ′ ∈ Ω, there is a unique element<br />

h ∈ G such that x ′ = h · x 0 . Thus any base point differs from a fixed one by a unique element <strong>of</strong><br />

G. The boundary is unchanged, but the manner in which G is identified with Ω changes the way<br />

in which one-parameter subgroups <strong>of</strong> G can approach ∂X, as we will see below. We will want our<br />

classification to be independent <strong>of</strong> this choice <strong>of</strong> base point x 0 , but for now our constructions will<br />

rely upon it.<br />

Suppose that X is an affine G-embedding with base point x 0 , so X is a normal affine G-variety<br />

containing the open orbit Gx 0 isomorphic to G. Our primary method for classifying embeddings<br />

8


X is to use one-parameter subgroups <strong>of</strong> G. Assign to the pair (X, x 0 ) the set<br />

Γ(X, x 0 ) = {γ ∈ X ∗ (G) : lim<br />

t→0<br />

γ(t)x 0 exists in X}.<br />

Recall, lim t→0 γ(t)x 0 exists in X if and only if the composition <strong>of</strong> γ : G m → G with ψ x0 : G → X<br />

extends to a morphism ˜γ : A 1 → X, where ψ x0 (g) = g · x 0 . In this case, lim t→0 γ(t)x 0 is defined to<br />

be ˜γ(0).<br />

The classification <strong>of</strong> affine quasihomogeneous G-varieties is equivalent to the classification, up<br />

to right translations, <strong>of</strong> left-invariant finitely generated subalgebras <strong>of</strong> k[G]. We prove that affine<br />

G-embeddings X may be recovered from the set Γ(X, x 0 ) by reconstructing k[X] from Γ(X, x 0 ),<br />

up to right translation. To each one-parameter subgroup γ <strong>of</strong> G, we associate a G-stable discrete<br />

valuation v γ <strong>of</strong> k(G) and prove<br />

Theorem. X ∼ = Spec A Γ(X,x0 ), where A Γ(X,x0 ) = {f ∈ k[G] : v γ (f) ≥ 0 for all γ ∈ Γ(X, x 0 )} is a<br />

finitely generated subalgebra <strong>of</strong> k[G] that is invariant under left translations by G.<br />

Therefore, the classification <strong>of</strong> affine G-embeddings is equivalent to characterizing sets <strong>of</strong> oneparameter<br />

subgroups <strong>of</strong> G that can be obtained from affine G-embeddings. The next observation<br />

is that the set Γ(X, x 0 ) for an affine G-embedding, or for any affine G-variety, is determined by a<br />

state [27]. A state Ξ is an assignment <strong>of</strong> a nonempty subset Ξ(R) ⊆ X ∗ (R), for each torus R <strong>of</strong> G,<br />

satisfying certain compatibility conditions. In particular, the set Γ(X, x 0 ) defines a state Γ(X, x 0 ) ∨<br />

by<br />

Γ(X, x 0 ) ∨ (R) = {χ ∈ X ∗ (R) : 〈χ, γ〉 ≥ 0 for all γ ∈ Γ(X, x 0 ) ∩ X ∗ (R)}.<br />

Similarly, a state Ψ determines a set <strong>of</strong> one-parameter subgroups by Ψ ∨ = {γ ∈ X ∗ (G) : 〈χ, γ〉 ≥ 0<br />

for all χ ∈ Ψ(γ(G m ))}. So, if a set Γ <strong>of</strong> one-parameter subgroups is obtained from an affine G-<br />

embedding, it must be strongly convex: Γ = (Γ ∨ ) ∨ and γ, γ −1 ∈ Γ if and only if γ is the trivial<br />

one-parameter subgroup, ε. Moreover, the associated state Γ ∨ must be a finitely generated monoid<br />

state which will be defined in Chapter 3. Thus,<br />

Theorem. Affine G-embeddings are classified, up to conjugation, by strongly convex lattice cones<br />

<strong>of</strong> one-parameter subgroups. Conjugation corresponds to the change <strong>of</strong> base point.<br />

9


We conclude by describing G-equivariant morphisms between affine G-embeddings in terms<br />

<strong>of</strong> this classification.<br />

In particular, if X 1 → X 2 is a morphism <strong>of</strong> G-embeddings and x i ∈ X i<br />

are base points such that x 1 ↦→ x 2 , then Γ(X 1 , x 1 ) ⊆ Γ(X 2 , x 2 ) and conversely. We characterize<br />

biequivariant affine G-embeddings in terms <strong>of</strong> their cones and construct a canonical “biequivariant<br />

resolution” X G → X <strong>of</strong> any given affine G-embedding X, where X G is a (G × G)-equivariant affine<br />

G-embedding satisfying a universal property.<br />

10


Chapter 2<br />

Toric varieties and group embeddings<br />

Throughout this chapter, k is an algebraically closed field <strong>of</strong> characteristic zero, G will denote a<br />

connected reductive algebraic group defined over k, and T will denote a maximal torus <strong>of</strong> G.<br />

Suppose X is a G-embedding. That is, assume X is a normal G-variety which contains G as<br />

an open subvariety in such a way that the action <strong>of</strong> G on X extends the left translation action<br />

on G. Note that this is a generalization <strong>of</strong> the notion <strong>of</strong> toric varieties to more general algebraic<br />

groups. The hope for such a theory was mentioned by Danilov in [12]. However, little work in this<br />

generality appears in the literature, where preference is given to the class <strong>of</strong> embeddings that are<br />

also spherical varieties for G × G.<br />

Among the biequivariant embeddings <strong>of</strong> G, i.e., G-embeddings which have both a left and a right<br />

action by G, the regular compactifications <strong>of</strong> G have been studied extensively [3], [2], [5], [46], [44]<br />

and, more specifically, the wonderful compactification <strong>of</strong> the group [7], [8], [9], [41], [6]. In such<br />

embeddings, the closure <strong>of</strong> a maximal torus is always a toric variety [46] and this toric variety<br />

determines the compactification [5].<br />

In this chapter, we investigate the relationship between toric varieties and embeddings <strong>of</strong> reductive<br />

groups in a number <strong>of</strong> contexts. First, we will briefly review the theory <strong>of</strong> toric varieties,<br />

borrowing heavily from [28], [12], [33], [17], and [14]. With this background in hand, we next consider<br />

toric varieties in group embeddings. Finally, we conclude the chapter with several constructions <strong>of</strong><br />

group embeddings using toric varieties. Using one <strong>of</strong> these constructions, in Section 2.2.2 we prove<br />

that the closure <strong>of</strong> a single maximal torus <strong>of</strong> G in a G-embedding is insufficient for determining the<br />

embedding, in contrast to the result mentioned above for regular compactifications.<br />

11


2.1 Toric varieties as group embeddings<br />

Let T = G r m be a split algebraic torus defined over k. A toric variety X is a normal T -variety which<br />

contains T as a dense open subvariety so that the action <strong>of</strong> T on X extends the regular action <strong>of</strong><br />

T on itself by multiplication. As the rest <strong>of</strong> this paper hinges upon the ideas originally used to<br />

classify toric varieties in [28], we present a brief overview <strong>of</strong> those ideas here. Other references for<br />

this subject are [12], [33], [17]. We borrow much from their treatment <strong>of</strong> the topic and refer to<br />

them for pro<strong>of</strong>s.<br />

Let T be an algebraic torus and X a toric variety for T , both defined over an algebraically<br />

closed field k. By a theorem <strong>of</strong> Sumihiro [42], there is a covering <strong>of</strong> X by T -stable, affine, open<br />

subvarieties which are therefore also toric varieties.<br />

Hence every toric variety may be obtained<br />

by gluing together affine toric varieties, so the classification <strong>of</strong> toric varieties reduces to the two<br />

problems <strong>of</strong> classifying affine toric varieties and describing how affine toric varieties may be glued<br />

together. We review the solutions to these problems in the following subsections.<br />

2.1.1 Affine toric varieties and one-parameter subgroups<br />

Suppose X is an affine toric variety, say X = Spec A. The open embedding T → X corresponds to<br />

an injective homomorphism A → k[T ] ∼ = k[x 1 , x −1<br />

1 , . . . , x r, x −1<br />

r ]. The action <strong>of</strong> T on X induces an<br />

action <strong>of</strong> T on A as well via t · f : x ↦→ f(t −1 x) for all f ∈ A.<br />

Let X ∗ (T ) denote the set <strong>of</strong> one-parameter subgroups <strong>of</strong> T , X ∗ (T ) = Hom(G m , T ). This is a<br />

free abelian group, which is dual to the group <strong>of</strong> characters <strong>of</strong> T , X ∗ (T ), with respect to the perfect<br />

pairing 〈·, ·〉 : X ∗ (T ) × X ∗ (T ) → Z determined by<br />

χ(γ(a)) = a 〈χ,γ〉 (2.1)<br />

for all χ ∈ X ∗ (T ), γ ∈ X ∗ (T ), a ∈ k × .<br />

Let S be a semi-group in X ∗ (T ), which is finitely generated. Set k[S] equal to the k-subalgebra<br />

<strong>of</strong> k[T ] = k[X ∗ (T )] generated by characters χ ∈ S. Since S is finitely generated, k[S] is a k-algebra<br />

<strong>of</strong> finite type. If S generates X ∗ (T ) as a group, then the inclusion k[S] ⊂ k[T ] induces an equivariant<br />

embedding T ⊂ Spec k[S], and every equivariant T -embedding arises in this way. Therefore,<br />

12


Proposition 3 ([28], Proposition I.1). The correspondence S ↦→ Spec k[S] defines a bijection<br />

between the set <strong>of</strong> finitely generated semi-groups S ⊂ X ∗ (T ) which generate X ∗ (T ) as a group and<br />

the set <strong>of</strong> isomorphism classes <strong>of</strong> equivariant affine embeddings <strong>of</strong> T . Moreover, the morphisms<br />

<strong>of</strong> equivariant affine embeddings correspond (in a contravariant way) to the inclusions between<br />

semi-groups contained in X ∗ (T ).<br />

We say a semi-group S ⊂ X ∗ (T ) is saturated if χ n ∈ S for some positive n ≥ 1 and χ ∈ X ∗ (T )<br />

implies that χ ∈ S. This saturated condition corresponds to the normality <strong>of</strong> the associated torus<br />

embedding, so that<br />

Theorem 2 ([28], Theorem I.1). The correspondence S ↦→ Spec k[S] defines a bijection between<br />

the set <strong>of</strong> finitely generated semi-groups S ⊂ X ∗ (T ) which generate X ∗ (T ) as a group and are<br />

saturated and the set <strong>of</strong> equivariant normal affine embeddings <strong>of</strong> T .<br />

Hence, affine toric varieties correspond to saturated semi-groups S <strong>of</strong> X ∗ (T ) which generate<br />

X ∗ (T ) as a group. Now we classify all such saturated semi-groups <strong>of</strong> X ∗ (T ).<br />

Regard X ∗ (T ) as a Z-lattice in the real vector space X ∗ (T ) R = X ∗ (T ) ⊗ Z R.<br />

Definition 3. Let σ ⊂ X ∗ (T ) R . We call σ a convex rational polyhedral cone if σ = { ∑ N<br />

i=1 λ ix i :<br />

λ i ∈ R ≥0 , for all i} for some finite collection <strong>of</strong> elements x i ∈ X ∗ (T ).<br />

We say σ is a strongly<br />

convex rational polyhedral cone if it is a convex rational polyhedral cone and it does not contain<br />

any non-zero linear subspace <strong>of</strong> X ∗ (T ) R .<br />

We associate to a strongly convex rational polyhedral cone σ its dual cone, σ ∨ ⊂ X ∗ (T ) R , which<br />

is defined as σ ∨ = {u ∈ X ∗ (T ) R : 〈u, v〉 ≥ 0 for all v ∈ σ}. Then σ ∨ is a convex, rational, polyhedral<br />

cone in X ∗ (T ) R and (σ ∨ ) ∨ = σ. The correspondence σ ↦→ σ ∨ ∩ X ∗ (T ) defines a bijection between<br />

the set <strong>of</strong> strongly convex rational polyhedral cones and the set <strong>of</strong> finitely generated semi-groups<br />

in X ∗ (T ) which are saturated and generate X ∗ (T ) as a group by<br />

Lemma 1 ([28], Gordan’s Lemma). Given a finite set <strong>of</strong> homogeneous linear integral inequalities,<br />

the semi-group <strong>of</strong> integral solutions is finitely generated.<br />

Therefore, we obtain the classification <strong>of</strong> affine toric varieties below.<br />

13


Theorem 3 ([28], Theorem I.1 ′ ). The correspondence σ ↦→ Spec k[σ ∨ ∩ X ∗ (T )] = X σ defines<br />

a bijection between the set <strong>of</strong> strongly convex, rational, polyhedral cones in X ∗ (T ) R and the set <strong>of</strong><br />

affine toric varieties <strong>of</strong> T . Moreover, for γ ∈ X ∗ (T ), γ ∈ σ if and only if lim t→0 γ(t) exists in X σ .<br />

Therefore, affine toric varieties X are all <strong>of</strong> the form X σ , where<br />

σ = {γ ∈ X ∗ (T ) : lim<br />

t→0<br />

γ(t) exists in X}<br />

is the intersection <strong>of</strong> a strongly convex, rational, polyhedral cone with X ∗ (T ).<br />

2.1.2 Toric varieties from fans<br />

In this subsection, we explain how affine toric varieties are glued together to form new toric varieties<br />

using the classification <strong>of</strong> affine toric varieties given above. As we will only be considering strongly<br />

convex, rational, polyhedral cones in the remainder <strong>of</strong> this chapter, we shall simply use the term<br />

“cone” and suppress the adjectives.<br />

Recall σ ∨ ⊂ X ∗ (R) R is the set <strong>of</strong> u ∈ X ∗ (T ) R such that 〈u, v〉 ≥ 0 for all v ∈ σ. For each<br />

χ ∈ σ ∨ ∩ X ∗ (T ), define χ ⊥ = {x ∈ X ∗ (T ) R : 〈χ, x〉 = 0}. We call τ = σ ∩ χ ⊥ a face <strong>of</strong> σ and write<br />

τ < σ.<br />

Definition 4. A fan Σ in X ∗ (T ) R is a finite collection <strong>of</strong> (strongly convex, rational, polyhedral)<br />

cones σ such that<br />

1. every face <strong>of</strong> a cone <strong>of</strong> Σ is itself a cone <strong>of</strong> Σ;<br />

2. if σ and σ ′ are cones in Σ, then σ ∩ σ ′ is a common face <strong>of</strong> σ and σ ′ .<br />

Toric varieties for T are classified by fans as follows. Let Σ be a fan in X ∗ (T ) R . For each<br />

σ ∈ Σ, we have an affine toric variety X σ as constructed above. We may naturally glue X σ and<br />

X σ ′ together along their common open toric subvariety X σ∩σ ′, since if τ ≤ σ, then k[τ ∨ ∩ X ∗ (T )]<br />

is the localization <strong>of</strong> k[σ ∨ ∩ X ∗ (T )] at χ, where τ = σ ∩ χ ⊥ , so X τ is an open subvariety <strong>of</strong> X σ . In<br />

this way, we may form the toric variety associated to the fan Σ,<br />

X Σ = ⋃ σ∈Σ<br />

X σ ,<br />

14


and every toric variety for T arises in this way for some fan Σ.<br />

Moreover, we note that since T ∼ = X {0}<br />

⊂ X σ for all σ ∈ Σ (since {0} is a face <strong>of</strong> every<br />

cone σ), the torus T is naturally an open dense subset <strong>of</strong> X Σ . Thus we have an open embedding<br />

i Σ : T → X Σ . Therefore,<br />

Theorem 4 ([28], Theorem I.6). The correspondence Σ ↦→ X Σ defines a bijection between fans<br />

in X ∗ (T ) R and isomorphism classes <strong>of</strong> equivariant normal embeddings <strong>of</strong> T .<br />

Many <strong>of</strong> the algebro-geometric features <strong>of</strong> a toric variety may be described by corresponding<br />

features <strong>of</strong> its fan. For instance, the affine toric variety X σ is smooth if and only if the cone σ is<br />

generated by part <strong>of</strong> a basis for the vector space X ∗ (T ) R . In this case, we say that the cone σ is<br />

smooth. Thus any toric variety X Σ is smooth if and only if each cone σ ∈ Σ is smooth in this sense.<br />

Additionally, we may detect when X Σ is a complete variety using the support <strong>of</strong> the fan Σ. The<br />

support <strong>of</strong> Σ, denoted |Σ|, is defined to be<br />

|Σ| = ⋃ σ∈Σ<br />

σ.<br />

We call a fan Σ complete if its support is the entire vector space X ∗ (T ) R . Complete fans are so<br />

named because they are the ones that correspond to complete toric varieties.<br />

2.1.3 The torus action<br />

The definition <strong>of</strong> toric varieties we gave at the beginning <strong>of</strong> this section required them to be T -<br />

varieties, so we now show how the torus acts on its toric varieties in terms <strong>of</strong> the cones and fan<br />

which classify the embedding.<br />

First, we give a coordinate-free description <strong>of</strong> the action, making use <strong>of</strong> the isomorphism <strong>of</strong><br />

functors Spec ∼ = Hom k ( , k). Then T = Spec k[X ∗ (T )] = Hom k (k[X ∗ (T )], k) can be thought <strong>of</strong> as<br />

the collection <strong>of</strong> homomorphisms Hom(X ∗ (T ), G m ), while X σ = Spec k[σ ∨ ∩X ∗ (T )] = Hom k (k[σ ∨ ∩<br />

X ∗ (T )], k) consists <strong>of</strong> k-algebra homomorphisms from k[σ ∨ ∩ X ∗ (T )] to k. The multiplication in k<br />

gives us the ability to multiply such homomorphisms, and this determines the action <strong>of</strong> T on X σ ,<br />

for<br />

(t · x)(m) = t(m)x(m)<br />

15


for all t ∈ T and x ∈ X σ (where t(m) = m(t) and x(m) = m(x)).<br />

We may also describe this action in terms <strong>of</strong> coordinates for the affine toric variety X σ . Let<br />

(a 1 , . . . , a s ) be a system <strong>of</strong> generators for the monoid σ ∨ ∩ X ∗ (T ). Given a basis {x i = n i ⊗ 1 : i =<br />

1, . . . , r} for X ∗ (T ) R , each a i may be written in the form a i = (αi 1, . . . , αr i ) with αj i ∈ Z and t ∈ T<br />

is written as t = (t 1 , . . . , t r ) with t j ∈ G m (k). A point x ∈ X σ is written x = (x 1 , . . . , x s ). Then<br />

the action T × X σ → X σ <strong>of</strong> T on the affine subvariety X σ is<br />

(t, x) ↦→ (t a 1<br />

x 1 , . . . , t as x s )<br />

where t a i<br />

= t α1 i<br />

1 · · · tαr i<br />

r<br />

to give an action on the entire variety X Σ .<br />

∈ G m (k). The actions <strong>of</strong> T on each <strong>of</strong> the X σ , for σ ∈ Σ, clearly glue together<br />

As the action <strong>of</strong> T on a toric variety X Σ can be described in terms <strong>of</strong> the fan, the orbit structure<br />

<strong>of</strong> X Σ may also be characterized using Σ. In particular, for each cone σ ∈ Σ, there is a base point<br />

z σ ∈ X Σ which is the limit point <strong>of</strong> any <strong>of</strong> the one-parameter subgroups in the relative interior <strong>of</strong><br />

the cone σ. That is, z σ = lim t→0 γ(t) for any γ ∈ σ − ⋃ τ


To see this, suppose D is a T -stable divisor. The T -action on D corresponds to a grading <strong>of</strong> I by<br />

X ∗ (T ), so I = ⊕ kχ where the sum is taken over some subset <strong>of</strong> X ∗ (T ). Since I is principal at x σ ,<br />

the unique fixed point <strong>of</strong> T on X σ , which is the limit <strong>of</strong> any γ in the relative interior <strong>of</strong> σ, I/MI is<br />

one-dimensional, where M = ⊕ η≠0 kη. Hence, there is a unique χ with I = k[σ∨ ∩ X ∗ (T )] · χ, so<br />

D = div(χ) for some unique χ ∈ X ∗ (T ) as claimed.<br />

Lemma 2 ([17]). Let χ ∈ X ∗ (T ) and let v be the first lattice point along an edge ρ <strong>of</strong> Σ. Then<br />

ord Dρ (div(χ)) = 〈χ, v〉, so<br />

where D i = D ρi = O ρi .<br />

[div(χ)] =<br />

d∑<br />

〈χ, v i 〉D i<br />

i=1<br />

2.2 Toric varieties in group embeddings<br />

In the biequivariant theory [5], the G-compactification X is determined by the closure <strong>of</strong> a maximal<br />

torus T in X, T . Such a T is normal [46], and thus is a toric variety. In contrast, a single toric<br />

variety within an equivariant group embedding need not be sufficient to recover the compactification<br />

<strong>of</strong> the group, as we will demonstrate in Section 2.2.2.<br />

2.2.1 Regular compactifications<br />

Definition 5. A biequivariant compactification <strong>of</strong> G is any complete, normal variety which contains<br />

G as an open subvariety so that the action by left and right multiplication <strong>of</strong> G × G on G extends<br />

to the whole variety.<br />

Biequivariant compactifications are a special class <strong>of</strong> the equivariant compactifications that we<br />

study, as we do not require the right action <strong>of</strong> G on itself to extend. For instance, Mumford’s group<br />

embedding (discussed in Section 2.3.1) need not have a global right action [28].<br />

The notion <strong>of</strong> regular variety was introduced by Bifet, De Concini and Procesi [3].<br />

Definition 6. One says that a variety X, with an action <strong>of</strong> G, is regular if it satisfies:<br />

1. X is smooth and has an open, dense G-orbit XG 0 whose complement is a divisor with normal<br />

crossings. One calls the irreducible components <strong>of</strong> X − X 0 G<br />

the boundary divisors.<br />

17


2. Each G-orbit closure is the transverse intersection <strong>of</strong> the boundary divisors containing it.<br />

3. If x ∈ X, then in the normal space to G · x in X the isotropy group G x <strong>of</strong> x acts with an<br />

open orbit.<br />

Definition 7. The regular compactifications <strong>of</strong> G are the biequivariant compactifications X <strong>of</strong> G<br />

that are regular as (G × G)-varieties.<br />

Example 4 (Some regular compactifications). Smooth complete toric varieties are regular<br />

compactifications <strong>of</strong> tori. For an adjoint group G, De Concini and Procesi [7] have constructed the<br />

wonderful compactification <strong>of</strong> G, as in Section 1.3. One could also define it as the unique regular<br />

compactification <strong>of</strong> G which has a unique closed (G × G)-orbit, although the notion <strong>of</strong> regular<br />

compactifications arose later. More recently, Brion [6] used the Hilbert scheme to recreate this<br />

compactification. For example, the projective space P(M 2 (C)) is the wonderful compactification <strong>of</strong><br />

P GL 2 (C).<br />

From now on, X will be a regular compactification <strong>of</strong> G. Fix a maximal torus T <strong>of</strong> G.<br />

Let Φ(T, G) be the set <strong>of</strong> roots <strong>of</strong> G relative to T . Suppose Φ + is a set <strong>of</strong> positive roots <strong>of</strong><br />

Φ(T, G) and ∆ is its base. Let C + = {ν ∈ X ∗ (T ) R : ∀α ∈ ∆, 〈α, ν〉 ≥ 0}, the closure <strong>of</strong> the positive<br />

Weyl chamber.<br />

One now defines a subdivision <strong>of</strong> C + associated to X. For this, let T denote the closure <strong>of</strong> T in<br />

X. On G, the restriction <strong>of</strong> the action <strong>of</strong> G×G to the torus diag(T ×T ) is given by (t, t)·g = tgt −1 .<br />

This extends to all <strong>of</strong> X. The variety <strong>of</strong> fixed points for diag(T × T ) is smooth [23] and T is an<br />

irreducible component. For the left action <strong>of</strong> T , T is thus a smooth complete toric variety, so one<br />

has the associated fan Σ for T . As T is invariant by the diagonal action <strong>of</strong> the Weyl group, Σ is<br />

also W -invariant. Hence Σ = W Σ + , where Σ + is the subdivision <strong>of</strong> C + formed <strong>of</strong> the cones <strong>of</strong> Σ<br />

contained in C + .<br />

In a manner analogous to the toric case, the cones <strong>of</strong> the fan Σ + parametrize the (G × G)-orbits<br />

<strong>of</strong> X. For all σ ∈ Σ + , denote by z σ the corresponding base point in T and by O σ its (G × G)-orbit.<br />

One says that z σ is the base point <strong>of</strong> the orbit O σ . The closed (G × G)-orbits <strong>of</strong> X correspond<br />

to maximal dimensional cones <strong>of</strong> Σ + . When σ is such a cone, z σ has isotropy group B − × B, so<br />

(G × G) · z σ<br />

∼ = G/B − × G/B.<br />

18


Conversely, every subdivision <strong>of</strong> C + all <strong>of</strong> whose cones are smooth (that is to say, generated<br />

by a partial basis <strong>of</strong> X ∗ (T )) corresponds to a regular compactification <strong>of</strong> G.<br />

In particular, the<br />

trivial subdivision formed <strong>of</strong> C + and its faces defines the wonderful compactification <strong>of</strong> G when G<br />

is adjoint.<br />

Therefore, biequivariant embeddings <strong>of</strong> groups are classified by the toric variety they define,<br />

and so if G is a semisimple algebraic group with maximal torus T , the category <strong>of</strong> biequivariant<br />

G-embeddings may be realized as a full subcategory <strong>of</strong> the category <strong>of</strong> toric varieties for T . In<br />

general, however, T fails to determine X for embeddings that are not biequivariant.<br />

2.2.2 A family <strong>of</strong> left-equivariant SL n -embeddings<br />

Let n be an integer ≥ 2 and let π be a partition <strong>of</strong> the set {2, 3, . . . , n} into disjoint subsets<br />

whose union is all <strong>of</strong> {2, 3, . . . , n}. Thus π = {P 1 , . . . , P m } where the P i = {p i1 , . . . , p idi } ⊂<br />

{2, 3, . . . , n} satisfy ⋃ m<br />

i=1 P i = {2, 3, . . . , n} and P i ∩ P j = ∅ if i ≠ j. Set Y π = ∏ P ∈π P|P | , where<br />

P = {p 1 , p 2 , . . . , p d } with p 1 < p 2 < · · · < p d , and |P | = p 1 + p 2 + · · · + p d . Clearly Y π is an<br />

equivariant B-embedding, where B is the Borel subgroup <strong>of</strong> upper triangular matrices in SL n ,<br />

with embedding given by<br />

⎛<br />

⎞<br />

a 11 a 12 · · · a 1n<br />

0 a 22 · · · a 2n<br />

. ↦→ ([a 1p1 : a 2p1 : · · · : a p1 p<br />

⎜ . ..<br />

1<br />

: a 1p2 : a 2p2 : · · · : a p2 p 2<br />

: · · · : a pd p d<br />

: 1]) P ∈π ,<br />

⎟<br />

⎝<br />

⎠<br />

0 0 · · · a nn<br />

and the B action on Y π is the unique diagonal action <strong>of</strong> B in this product <strong>of</strong> projective spaces<br />

extending the action <strong>of</strong> B on itself.<br />

Then SL n × B Y π is an SL n -embedding for each partition<br />

π <strong>of</strong> {2, 3, . . . , n}, whose SL n -orbits are in one-to-one correspondence with the B-orbits in Y π .<br />

The number <strong>of</strong> such partitions, and thus the corresponding SL n -embeddings, may be described<br />

recursively as follows. For simplicity, reindex the columns <strong>of</strong> B ⊂ SL n starting with 0, so we may<br />

consider each π as a partition <strong>of</strong> the set {1, 2, . . . , n − 1}. Let f(n) be the number <strong>of</strong> partitions<br />

<strong>of</strong> this set. Then f(n) may be computed using the formulas f(n) = ∑ k≥1<br />

f(n, k) and f(n, k) =<br />

f(n−1, k−1)+k·f(n−1, k), where f(n, k) is the number <strong>of</strong> partitions <strong>of</strong> {1, 2, . . . , n−1} containing<br />

19


k sets. (The first formula is obviously true and the second is valid because there are f(n − 1, k − 1)<br />

partitions with k members where π consists <strong>of</strong> {n − 1} and k − 1 other subsets while all others have<br />

the form <strong>of</strong> a partition <strong>of</strong> {1, 2, . . . , n−2} into k sets with n−1 then placed into any <strong>of</strong> these (which is<br />

why we multiply f(n−1, k) by k).) Clearly, f(n, 0) = 0, f(n, 1) = 1, f(n, n−1) = 1 and f(n, k) = 0<br />

whenever k ≥ n for all n. A little work shows f(2) = 1, f(3) = 2, f(4) = 5, f(5) = 15, f(6) = 52<br />

and f(7) = 203.<br />

We must observe that these Y π are not the only compactifications <strong>of</strong> B ⊂ SL n . For instance,<br />

in the SL 2 case, P 2 is the only Y π . However, the blow-up <strong>of</strong> P 2 at the origin is another B-<br />

compactification, which produces a distinct SL 2 -compactification, SL 2 × B Bl 0 P 2 .<br />

Lemma 3. Let B be the Borel subgroup <strong>of</strong> upper triangular matrices in SL n . Let U be the unipotent<br />

radical <strong>of</strong> B and let T be the maximal torus in B consisting <strong>of</strong> the diagonal matrices <strong>of</strong> SL n . Let<br />

π be a partition <strong>of</strong> the set {2, 3, . . . , n} and let Y π be the associated B-embedding. Then the closure<br />

<strong>of</strong> T in Y π , and thus in SL n × B Y π , is T ∼ = ∏ P ∈π P#P , where #P denotes the cardinality <strong>of</strong> the<br />

set P .<br />

Therefore, the closure <strong>of</strong> T in any SL n × B Y π only depends on the cardinalities <strong>of</strong> the elements<br />

<strong>of</strong> π, and not specifically on the partition. As a result, only when n = 2 or 3 will the closure T<br />

<strong>of</strong> T in SL n differentiate between non-isomorphic embeddings. When n = 4, all three embeddings<br />

SL 4 × B Y {{2,3},{4}} , SL 4 × B Y {{2,4},{3}} and SL 4 × B Y {{2},{3,4}} have T ∼ = P 1 × P 2 . {{2, 5, 6}, {3, 4}}<br />

have T = P 2 × P 3 and<br />

Proposition 4. For all n ≥ 4, there are distinct partitions π 1 , π 2 <strong>of</strong> {2, . . . , n} whose associated<br />

SL n -embeddings are non-isomorphic while their associated toric varieties T are isomorphic.<br />

Pro<strong>of</strong>. Let T be the maximal torus <strong>of</strong> SL n consisting <strong>of</strong> diagonal matrices in SL n .<br />

Consider<br />

the partitions π 1 = {{2, 3}, {4}, {5, . . . , n}} and π 2 = {{2}, {3, 4}, {5, . . . , n}} <strong>of</strong> {2, . . . , n}. The<br />

corresponding B-embeddings are Y π1 = P 5 × P 4 × P 1 2 n(n+1)−10 and Y π2 = P 2 × P 7 × P 1 2 n(n+1)−10 ,<br />

which are non-isomorphic. Observe that when n = 4, 1 n(n + 1) − 10 = 0, as there are only two<br />

2<br />

terms in Y π1 = P 5 × P 4 and in Y π2 = P 2 × P 7 . Hence the associated SL n -embeddings, SL n × B Y π1<br />

and SL n × B Y π2 , are non-isomorphic as varieties and so also as SL n -embeddings.<br />

In SL n × B Y π1 , if we identify SL n with its open orbit using the base point [I n , ([0 : 1 :<br />

20


0 : 0 : 1 : 1], [0 : 0 : 0 : 1 : 1], [. . . ])], then T · [I n , ([0 : 1 : 0 : 0 : 1 : 1], [0 : 0 : 0 : 1 : 1], [. . . ])] =<br />

[I n , ([0 : t 2 : 0<br />

(<br />

: 0 : t 3 : 1], [0 : 0 : 0 : t 4 : 1], [. . . ])] ∼ = P 2 × P 1 × P n−4 .<br />

1 0 0 0<br />

)<br />

0 0 0 1<br />

Let s 24 = 0 0 −1 0 0 be the permutation matrix in SL n corresponding to the transposition<br />

0 1 0 0<br />

0 I n−4 ⎛<br />

⎞ ⎛<br />

⎞<br />

t 1 t 1<br />

t2 t4<br />

(24) <strong>of</strong> the set {1, 2, 3, 4, . . . , n}. Hence s 2 24 = I ⎜<br />

n and s<br />

t3 ⎟ ⎜<br />

24 ⎝ t4 ⎠ s 24 =<br />

t3 ⎟<br />

⎝ t2 ⎠. In<br />

... ...<br />

SL n × B Y π2 , identify SL n with its open orbit using the base point [s 24 , ([0 : 1 : 1], [0 : 0 : 1 : 0 : 0 :<br />

0 : 1 : 1], [. . . ])]. We see that the closure <strong>of</strong> T in SL n × B Y π2 is<br />

T = T · [s 24 , ([0 : 1 : 1], [0 : 0 : 1 : 0 : 0 : 0 : 1 : 1], [. . . ])]<br />

= [s 2 24 T s 24, ([0 : 1 : 1], [0 : 0 : 1 : 0 : 0 : 0 : 1 : 1], [. . . ])]<br />

= [s 24 , ([0 : t 4 : 1], [0 : 0 : t 3 : 0 : 0 : 0 : t 2 : 1], [. . . ])]<br />

∼= P 1 × P 2 × P n−4 .<br />

More importantly, we see that this is isomorphic, as toric varieties, to the closure <strong>of</strong> T in SL n × B Y π1<br />

computed above.<br />

Therefore, in contrast to biequivariant compactifications, the closure <strong>of</strong> a single maximal torus<br />

in an equivariant embedding <strong>of</strong> a connected reductive group need not determine the embedding.<br />

Note that isomorphic varieties may be non-isomorphic as B-compactifications. The varieties<br />

P 3 × P 6 × P 11 , P 4 × P 5 × P 11 , P 5 × P 15 , P 6 × P 14 , P 7 × P 13 and P 8 × P 12 all appear twice while<br />

the varieties P 9 × P 11 and P 5 × P 6 × P 9 both occur three times in the family <strong>of</strong> non-isomorphic<br />

B-embeddings when B ⊂ SL 6 .<br />

For instance, P 3 × P 6 × P 11 is the underlying variety <strong>of</strong> both<br />

Y {{2,4},{3},{5,6}} and Y {{2,4,5},{3},{6}} .<br />

As another example, consider the GL 2 -embeddings GL 2 × D 2<br />

A 2 and M 2 , where D 2 denotes the<br />

maximal torus <strong>of</strong> diagonal matrices in GL 2 and M 2 denotes the space <strong>of</strong> all 2×2 matrices. Then, in<br />

M 2 , the closure <strong>of</strong> D 2 is isomorphic to A 2 , so there is an equivariant morphism GL 2 × D 2<br />

A 2 → M 2<br />

given by the action [g, x] ↦→ g · x, where [g, x] denotes the equivalence class <strong>of</strong> the pair (g, x) in<br />

GL 2 × D 2<br />

A 2 . The GL 2 -orbits in GL 2 × D 2<br />

A 2 are in one-to-one correspondence with the D 2 -orbits<br />

in A 2 , <strong>of</strong> which there are four.<br />

There is the open D 2 -orbit through the point (1, 1) ∈ A 2 , two<br />

21


one-dimensional orbits through (1, 0) and (0, 1) ∈ A 2 and the zero-dimensional orbit (0, 0). Under<br />

the morphism GL 2 × D 2<br />

A 2 → M 2 , these orbits are sent to GL 2 , ( a c 0<br />

0 ) , ( )<br />

0 b<br />

0 d , (<br />

0 0<br />

0 0 ) in M 2. Therefore,<br />

while it is clear that D 2<br />

∼ = A 2 is the same in both GL 2 × D 2<br />

A 2 and M 2 , these GL 2 -embeddings are<br />

not isomorphic. For example, M 2 contains at least the fifth orbit through ( 1 1 1<br />

1<br />

), in addition to the<br />

four in the image <strong>of</strong> GL 2 × D 2<br />

A 2 . Therefore, there is no isomorphism between GL 2 × D 2<br />

A 2 and M 2<br />

as GL 2 -embeddings. Another point <strong>of</strong> interest which distinguishes the one-sided and biequivariant<br />

cases is that with only a one-sided action, it is possible that there are maximal tori T 1 , T 2 such<br />

that T 1<br />

≁ = T 2 . This is obviously impossible in the presence <strong>of</strong> both a left and a right action, as all<br />

maximal tori are conjugate in G.<br />

While we cannot always recover our G-embedding X from T , we can associate to T other group<br />

embeddings related to the original X in a number <strong>of</strong> ways. This is the topic <strong>of</strong> the next section.<br />

2.3 Group embeddings from toric varieties<br />

2.3.1 Mumford’s group embeddings<br />

Not all group embeddings will respect both the left and right translations <strong>of</strong> G on itself, as is<br />

demonstrated by the following construction due to Mumford [28]. He constructs an embedding<br />

G ⊂ Ĝ <strong>of</strong> a semisimple group G into a reduced, irreducible and separated scheme Ĝ locally <strong>of</strong> finite<br />

type over k satisfying the following conditions:<br />

1. Ĝ is a toroidal embedding and, if G has no center, Ĝ is non-singular;<br />

2. The left action <strong>of</strong> G on itself extends to an action <strong>of</strong> G on<br />

Ĝ: i.e., there is a morphism<br />

α : G × Ĝ → Ĝ extending the left multiplication in G;<br />

3. The right action <strong>of</strong> G on itself extends pointwise but not continuously to an action on Ĝ: i.e.,<br />

for all g ∈ G k , if R g : G → G is right multiplication by g, then R g extends to a morphism<br />

̂R g : Ĝ → Ĝ;<br />

4. For each stratum Y <strong>of</strong> Ĝ − G, {g ∈ G k : ̂R g (Y ) = Y } is the set <strong>of</strong> closed points <strong>of</strong> a parabolic<br />

subgroup P Y <strong>of</strong> G and Y ↦→ P Y sets up a bijection between the strata <strong>of</strong> Ĝ − G and the<br />

parabolics P ⊂ G.<br />

22


Mumford’s embedding is constructed as follows. For every Borel subgroup B ⊂ G, form ̂B in the<br />

following way. First, let U be the unipotent radical <strong>of</strong> B. Second, let T = B/U and α i ∈ X ∗ (T ) the<br />

roots <strong>of</strong> T in U. Third, let σ ⊂ X ∗ (T ) R be {x : 〈α i , x〉 ≥ 0, ∀i}. Fourth, split B = U × T . Finally,<br />

form ̂B = B × T T σ . Only the fourth step is non-canonical and ̂B does not depend on the choice <strong>of</strong><br />

the splitting. With these ̂B’s constructed, define Ĝ = ⋃ B G ×B ̂B where the gluing is the unique<br />

one so that all G × B ̂B’s, for all Borel subgroups B <strong>of</strong> G, are identified at least on their common<br />

open set G ∼ = G × B B ⊂ G × B ̂B and Ĝ is separated and G × B ̂B ⊂ Ĝ is an open subset. Then<br />

this embedding G ⊂ Ĝ satisfies the conditions above. In particular, Ĝ has only a left G-action, so<br />

it does not fall within the scope <strong>of</strong> the spherical approach <strong>of</strong> Sections 1.2, 1.3, 2.2.1.<br />

2.3.2 Group embeddings constructed using flag varieties<br />

Let G be a semi-simple simply connected algebraic group with maximal torus T and Borel subgroup<br />

B ⊃ T .<br />

In Section 2.2, we saw how B-embeddings B ⊂ Y induced G-embeddings G ⊂ G × B Y via<br />

g ↦→ [g, y 0 ]. If Y 1 → Y 2 is a B-equivariant morphism <strong>of</strong> B-embeddings, then there is a corresponding<br />

G-morphism G× B Y 1 → G× B Y 2 . Moreover, if X is any G-embedding and B denotes the closure <strong>of</strong><br />

B in X, then there is an equivariant morphism G× B B → X <strong>of</strong> G-embeddings. If X is projective, so<br />

that B is too, then G× B B is projective and the morphism G× B B → X is surjective. Furthermore,<br />

if X 1 → X 2 is a morphism <strong>of</strong> G-embeddings, then, setting Y i = B ⊂ X i and constructing G × B Y i ,<br />

we obtain a G-morphism G × B Y 1 → G × B Y 2 over X 1 → X 2 :<br />

G × B Y 1<br />

X 1<br />

G × B Y 2<br />

X 2 .<br />

Therefore, the category <strong>of</strong> B-embeddings may be viewed as a full subcategory <strong>of</strong> the category <strong>of</strong><br />

G-embeddings by the functor Y ↦→ G × B Y , which is adjoint to the restriction functor X ↦→ B ⊂ X<br />

from the category <strong>of</strong> G-embeddings to the category <strong>of</strong> B-embeddings. These results are true for<br />

arbitrary closed subgroups H <strong>of</strong> G [36].<br />

In this section, we give another construction <strong>of</strong> G-embeddings obtained using flag varieties and<br />

23


toric varieties. As [B, B] is the unipotent radical U <strong>of</strong> B and B/U ∼ = T , there is a unique, welldefined<br />

homomorphism β : B → T .<br />

Through this homomorphism, every T -variety Z is also a<br />

B-variety, where b · z := β(b)z for all b ∈ B and all z ∈ Z.<br />

Now let B − denote the Borel subgroup <strong>of</strong> G which is opposite to B, so that B ∩ B − = T .<br />

Define the homomorphism β ′ : B × B − → T by (b, c) ↦→ β(b). Via β ′ , any T -variety becomes a<br />

(B × B − )-variety, so in particular every toric variety for T is also a (B × B − )-variety.<br />

Suppose X = X Σ is a toric variety for T associated to the fan, Σ, in X ∗ (T ) R . Then X is a<br />

(B × B − )-variety as described above, so define Ind(X) := (G × G) × B×B− X. That is, Ind(X) =<br />

[(G×G)×X]/ ∼ where we define ((g 1 , g 2 ), x) ∼ ((g 1 b 1 , g 2 b 2 ), β ′ (b 1 , b 2 ) −1 x) for all (b 1 , b 2 ) ∈ B ×B − .<br />

We call the variety Ind(X) the induced toric variety associated to X. Let [(g 1 , g 2 ), x] denote the<br />

equivalence class <strong>of</strong> ((g 1 , g 2 ), x) in Ind(X). Then Ind(X) is a left (G × G)-variety via the left action<br />

<strong>of</strong> G × G on the first component: (h 1 , h 2 ) · [(g 1 , g 2 ), x] = [(h 1 g 1 , h 2 g 2 ), x].<br />

We wish to elaborate on this construction. Given a T -variety X, we define the induced variety<br />

Ind ∗ (X) to be the principal fiber space over the flag variety G/B × G/B − ∼ = G × G/B × B − with<br />

fiber X under the (B × B − )-action just described. That is, there is a surjection p : Ind ∗ (X) →<br />

G/B × G/B − all <strong>of</strong> whose fibers are isomorphic to X, and we may embed X in Ind ∗ (X) as the<br />

fiber over (eB, eB − ).<br />

Lemma 4. Suppose X is a toric variety for T . Then Ind(X) ∼ = Ind ∗ (X) as (G × G)-varieties.<br />

Pro<strong>of</strong>. Let X be a toric variety for T . Then consider the (G × G)-variety Ind(X) and the map<br />

π 1 : Ind(X) → G × G/B × B − given by [(g 1 , g 2 ), x] ↦→ [g 1 , g 2 ]. This is well-defined, for if<br />

(g 1 , g 2 , x) ∼ (g 1 b 1 , g 2 b 2 , β ′ (b 1 , b 2 ) −1 x), then π 1 ([(g 1 b 1 , g 2 b 2 ), β ′ (b 1 , b 2 ) −1 x]) = [g 1 b 1 , g 2 b 2 ] = [g 1 , g 2 ] =<br />

π 1 ([(g 1 , g 2 ), x]).<br />

Moreover, π −1<br />

1 ([g 1, g 2 ]) = {[(h 1 , h 2 ), y] ∈ Ind(X) : π([(h 1 , h 2 ), y]) = [g 1 , g 2 ]} =<br />

{[(g 1 b 1 , g 2 b 2 ), y] ∈ Ind(X)} = {[(g 1 , g 2 ), β ′ (b 1 , b 2 ) −1 y]} ∼ = X, since y is arbitrary. Therefore, Ind(X)<br />

is fiber space over G × G/B × B − whose fiber over each point is isomorphic to X. Thus using<br />

the isomorphism between G × G/B × B − and G/B × G/B − , we obtain the desired isomorphism<br />

Ind(X) ∼ = Ind ∗ (X), which is clearly (G × G)-equivariant.<br />

From now on we will identify Ind(X) with Ind ∗ (X) and refer to both as the induced toric<br />

variety associated to X. However, we will see that this second description <strong>of</strong> the induced variety is<br />

<strong>of</strong> greater value when determining the isomorphism classes <strong>of</strong> line bundles on Ind(X).<br />

24


Moreover, Ind(X) is a G-embedding, for the image <strong>of</strong> diag(G × G) ∼ = G in Ind(X) is<br />

G ∼ = G × T T<br />

⊆ G × T X ∼ = p −1 (diag(G/B × G/B − )) ⊂ Ind(X)<br />

open<br />

open<br />

since T is open in X and diag(G/B × G/B − ) ∼ = G/T is open in G/B × G/B − [21].<br />

Using this description <strong>of</strong> the induced toric variety Ind(X), we can readily verify that induction<br />

is a functor:<br />

Lemma 5. Induction as defined above is a covariant functor<br />

(( toric varieties/T )) → (( (G × G)-varieties )).<br />

Pro<strong>of</strong>. By the existence <strong>of</strong> fiber products, Ind(X) is a variety and we have already described its<br />

(G × G)-action, so Ind maps toric varieties to (G × G)-varieties. Furthermore, if f : X 1 → X 2 is a<br />

T -equivariant morphism <strong>of</strong> toric varieties, it determines a (G × G)-equivariant morphism Ind(f) :<br />

Ind(X 1 ) → Ind(X 2 ) as follows. Consider the map id G×G × f : G × G × X 1 → G × G × X 2 :<br />

(id G×G × f)(g 1 b 1 , g 2 b 2 , β ′ (b 1 , b 2 ) −1 x) = (g 1 b 1 , g 2 b 2 , f(β ′ (b 1 , b 2 ) −1 x))<br />

= (g 1 b 1 , g 2 b 2 , β ′ (b 1 , b 2 ) −1 f(x))<br />

∼ (g 1 , g 2 , f(x))<br />

= (id G×G × f)(g 1 , g 2 , x)<br />

for all (b 1 , b 2 ) ∈ B × B − and (g 1 , g 2 , x) ∈ G × G × X. Since Ind(X i ) = [(G × G) × X i ]/ ∼, it<br />

follows that Ind(f) : Ind(X 1 ) → Ind(X 2 ) is well-defined. Thus induction is the functor given by<br />

X ↦→ Ind(X) := (G×G)× B×B− X and sending the morphism f : X 1 → X 2 to Ind(f) : [(g 1 , g 2 ), x] ↦→<br />

[(g 1 , g 2 ), f(x)].<br />

Now we only have left to show that Ind(id X ) = id Ind(X) for all toric varieties X and that<br />

Ind(g) ◦ Ind(f) = Ind(g ◦ f) whenever f : X 1 → X 2 and g : X 2 → X 3 are T -equivariant morphisms.<br />

The first <strong>of</strong> these requirements is clear. The second follows since Ind(g ◦ f) is the map Ind(X 1 ) →<br />

Ind(X 3 ) associated to the map id G×G × (g ◦ f) = (id G×G × g) ◦ (id G×G × f), the latter map<br />

corresponding to the composition Ind(g) ◦ Ind(f). Therefore, induction is a covariant functor from<br />

25


the category <strong>of</strong> toric varieties to the category <strong>of</strong> (G × G)-varieties as claimed.<br />

Via the diagonal map ∆ : G → G×G, we have a functor (( (G×G)-varieties )) → ((G-varieties )),<br />

so it follows that<br />

Corollary 1. X ↦→ Ind(X) is a covariant functor (( T -varieties )) → (( G-varieties )).<br />

As an aside, we describe the cohomology groups <strong>of</strong> line bundles on Ind(X) when it is projective<br />

in terms <strong>of</strong> the cohomology groups <strong>of</strong> the toric variety X and <strong>of</strong> the flag variety G/B × G/B − .<br />

Therefore, we first recall how to compute cohomology for toric varieties and flag varieties.<br />

Cohomology <strong>of</strong> Line Bundles over Toric Varieties: Suppose X = X Σ is a toric variety.<br />

The equivariant Picard group <strong>of</strong> X Σ corresponds with the set <strong>of</strong> Σ-linear support functions on<br />

X ∗ (T ): a function h : |Σ| → R is called a Σ-linear support function if it is Z-valued on X ∗ (T ) ∩ |Σ|<br />

and is linear on each σ ∈ Σ, i.e., there are linear maps l σ ∈ X ∗ (T ) ∨ such that h(v) = 〈l σ , v〉 for all<br />

v ∈ σ and 〈l σ , v〉 = 〈l τ , v〉 whenever v ∈ σ ∩ τ. We denote the set <strong>of</strong> Σ-linear support functions<br />

SF (X ∗ (T ), Σ). Then SF (X ∗ (T ), Σ) ∼ = Z Σ(1) by h ↦→ (h(σ 1 )) σ∈Σ(1) , where Σ(1) = {σ ∈ Σ : dim σ =<br />

1} and σ 1 denotes the unique element <strong>of</strong> the ray σ ∈ Σ(1) such that σ = {nσ 1 : n ∈ N 0 }. The Σ-<br />

linear support functions h determine T -equivariant divisors D h = − ∑ σ∈Σ(1) h(σ 1)D σ on X Σ , where<br />

D σ is the closure <strong>of</strong> the orbit through z σ = lim t→0 σ 1 (t). Every T -equivariant line bundle L on X<br />

is thus <strong>of</strong> the form O X (D h ) for some Σ-linear support function h, so SF (X ∗ (T ), Σ) ∼ = Pic T (X Σ ).<br />

Suppose X = X Σ is a complete toric variety (i.e., Σ is a complete fan in X ∗ (T ) R ). Then the<br />

cohomology groups H p (X, O X (D h )) are T -representations via the action <strong>of</strong> T on X. As such, since<br />

T is linearly reductive, they split into a direct sum <strong>of</strong> T -eigenspaces:<br />

H p (X, O X (D h )) ∼ =<br />

⊕<br />

H p (X, O X (D h )) λ ,<br />

λ∈X ∗ (T )<br />

where λ runs over the set <strong>of</strong> characters <strong>of</strong> T . Now define Z(h, λ) to be the closed subset {n ∈<br />

X ∗ (T ) R : 〈 u λ , n 〉 ≥ h(n)} ⊂ X ∗ (T ) R , where u λ ∈ X ∗ (T ) ∨ corresponds to the character λ ∈ X ∗ (T ).<br />

Then we have the following result.<br />

Proposition 5 ([12], Theorem 7.2). For each λ ∈ X ∗ (T ), the eigenspace H q (X, O X (D h )) λ <strong>of</strong><br />

26


the T -action with respect to the character λ can be identified canonically as<br />

H q (X, O X (D h )) λ = H q Z(h,λ) (X ∗(T ) R , k)e λ<br />

so that we have a direct sum decomposition<br />

H q (X, O X (D h )) =<br />

⊕<br />

H q Z(h,λ) (X ∗(T ) R , k)e λ .<br />

λ∈X ∗ (T )<br />

Furthermore, if h is upper convex, i.e., h(v 1 ) + h(v 2 ) ≤ h(v 1 + v 2 ) for all v 1 , v 2 ∈ X ∗ (T ) R ,<br />

H i (X, O X (D h )) = 0 for all i > 0. In this case, we have<br />

⎧<br />

⎪⎨ #{χ ∈ X ∗ (T ) : 〈χ, v〉 ≥ h(v) for all v ∈ X ∗ (T ) R } i = 0,<br />

dim C H i (X, O X (D h )) =<br />

⎪⎩ 0 i > 0.<br />

Cohomology <strong>of</strong> Line Bundles over Flag Varieties: The flag variety G/B plays an important<br />

role in the representation theory <strong>of</strong> the algebraic group G. Representations <strong>of</strong> the Borel<br />

subgroup B may be lifted to representations <strong>of</strong> the group G, and all finite-dimensional representations<br />

are obtained in this way. This lifting is realized as a cohomology group <strong>of</strong> the flag variety<br />

G/B with coefficients in a line bundle L(λ) induced from a representation λ <strong>of</strong> B. The work <strong>of</strong><br />

Weyl, Borel–Weil, and later Bott completely described the cohomology groups H p (G/B, L(λ)) in<br />

characteristic zero. The results are stated in the following two theorems:<br />

Theorem 5 ([26], Bott–Borel–Weil Theorem). If λ is a fundamental dominant weight, then,<br />

for all w ∈ W (T, G) and i ∈ N 0 ,<br />

⎧<br />

⎪⎨ H 0 (G/B, L(λ)) i = l(w),<br />

H i (G/B, L(w · λ)) =<br />

⎪⎩ 0 otherwise.<br />

Define the character <strong>of</strong> a G-module M to be ch(M) = ∑ µ m µ(M)e µ ∈ Ze X∗ (T ) , where m µ (M)<br />

is the multiplicity <strong>of</strong> µ as a weight in M. The Euler characteristic <strong>of</strong> a character λ is χ(λ) =<br />

∑<br />

i≥0 (−1)i ch(H i (G/B, L(λ))).<br />

27


Theorem 6 ([26], Weyl’s Character Formula). For any character λ ∈ X ∗ (T ),<br />

χ(λ) = J(λ + ρ)/J(ρ),<br />

where J(µ) = ∑ w∈W (−1)l(w) e w·µ and ρ = 1 2<br />

∑<br />

α∈Φ + (T,G) α.<br />

Line Bundles on Induced Toric Varieties: Suppose L is a (G × G)-equivariant line bundle<br />

on Ind(X). Then consider its restriction i ∗ L to the toric variety X. Clearly this is a (G × G)-<br />

equivariant line bundle on X. Now consider the diagram:<br />

G × G × X<br />

µ<br />

<br />

π 3<br />

<br />

X<br />

i<br />

Ind(X)<br />

p<br />

<br />

G/B × G/B −<br />

Since i ∗ L is a (G × G)-equivariant line bundle on X and π 3 is a (G × G)-equivariant morphism,<br />

π3 ∗(i∗ L) is a (G × G)-equivariant line bundle on G × G × X. We wish to push this bundle forward<br />

via µ : (g 1 , g 2 , x) ↦→ [(g 1 , g 2 ), x] onto Ind(X). Yet, µ ∗ needn’t send line bundles to line bundles.<br />

Thus we push π3 ∗(i∗ L) forward onto Ind(X) via µ as<br />

S := [µ ∗ (π ∗ 3i ∗ L)] B×B− .<br />

We call S the line bundle induced from i ∗ L. (More generally, given any line bundle N on X, we<br />

can produce a line bundle [µ ∗ (π ∗ 3 N )]B×B−<br />

on Ind(X).)<br />

Now S is a (G × G)-equivariant line bundle on Ind(X) and agrees with L on the fibers <strong>of</strong><br />

p : Ind(X) → G/B×G/B − . Thus L⊗ OInd(X) S −1 is trivial on the fibers <strong>of</strong> p, so p ∗ (L⊗ OInd(X) S −1 ) is a<br />

line bundle M on G/B×G/B − . Since p ∗ is adjoint to p ∗ , however, this implies that L⊗ OInd(X) S −1 ∼ =<br />

p ∗ M, and hence that L ∼ = S ⊗ OInd(X) p ∗ M is the tensor product <strong>of</strong> a line bundle induced from X<br />

and one from the flag variety G/B × G/B − . We have just proven:<br />

Proposition 6. Let X be a toric variety for T and suppose L is a (G × G)-equivariant line bundle<br />

on the induced toric variety Ind(X). Then L decomposes as the tensor product <strong>of</strong> a line bundle<br />

28


from X with a line bundle from the flag variety G/B × G/B − .<br />

Corollary 2. There is an exact sequence <strong>of</strong> Picard groups associated to our fibration construction<br />

<strong>of</strong> Ind(X):<br />

0 Pic(G/B × G/B − )<br />

p ∗ Pic(Ind(X)) i∗ Pic(X) 0.<br />

Moreover, there is a section s : Pic(X) → Pic(Ind(X)) <strong>of</strong> i ∗ given by [L] ↦→ [(µ ∗ π ∗ 3 L)B×B− ].<br />

Pro<strong>of</strong>. We first show that p ∗ is injective.<br />

Suppose p ∗ [M] = p ∗ [N ] for line bundles M, N on<br />

G/B × G/B − . Then there is an isomorphism p ∗ M → p ∗ N . This pushes down to an isomorphism<br />

p ∗ p ∗ M → p ∗ p ∗ N . However, if L is a line bundle on G/B × G/B − , then p ∗ p ∗ L = L: by definition<br />

p ∗ p ∗ L is the sheaf associated to the presheaf U ↦→ p ∗ L(p −1 U) = lim −→V ⊃p(p<br />

L(V ) = L(U), since<br />

−1 U)<br />

p is surjective implies p(p −1 U) = U. Therefore, since L is already a sheaf, p ∗ p ∗ L = L. Hence p ∗ is<br />

an injection.<br />

We now show that the sequence is exact at Pic(Ind(X)). First suppose that M is a line bundle<br />

on G/B ×G/B − . Then i ∗ (p ∗ M) = (p◦i) ∗ M is (p◦i) −1 M⊗ (p◦i) −1 O G/B×G/B − O X, where (p◦i) −1 M<br />

is the sheafification <strong>of</strong> U ↦→ lim −→V ⊃(p◦i)(U)<br />

M(V ) = lim −→V ∋[e,e]<br />

M(V ) = lim −→V ⊂U[e,e]<br />

k = k. So i ∗ (p ∗ M)<br />

is the constant sheaf k tensored with the structure sheaf <strong>of</strong> X, which is isomorphic to the structure<br />

sheaf <strong>of</strong> X, O X . Hence the image <strong>of</strong> p ∗ is contained in the kernel <strong>of</strong> i ∗ . Moreover, if L is a line<br />

bundle on Ind(X) such that i ∗ L is trivial, then the restriction <strong>of</strong> L to the fibers <strong>of</strong> p is always<br />

trivial. Therefore p ∗ L is a line bundle on G/B × G/B − whose pull-back p ∗ p ∗ L is isomorphic to L<br />

by the adjointness <strong>of</strong> p ∗ and p ∗ . Hence the sequence is exact at Pic(Ind(X)).<br />

Finally, i ∗ is surjective, since we have already seen that s : [L] ↦→ [(µ ∗ π ∗ 3 i∗ L) B×B− ] is a section<br />

satisfying i ∗ ◦ s = 1.<br />

Since X = X Σ is a complete toric variety for T , its equivariant Picard group corresponds with<br />

the set <strong>of</strong> Σ-linear support functions on X ∗ (T ). Moreover, the Picard group <strong>of</strong> the flag variety<br />

G/B × G/B − is well-known to be isomorphic to the character group <strong>of</strong> the maximal torus T × T<br />

<strong>of</strong> G × G [21], namely X ∗ (T ) × X ∗ (T ). Therefore,<br />

Pic(Ind(X)) ∼ = SF (X ∗ (T ), Σ) ⋉ (X ∗ (T ) × X ∗ (T )), (2.2)<br />

29


where SF (X ∗ (T ), Σ) is the abelian group <strong>of</strong> Σ-linear support functions.<br />

Cohomology <strong>of</strong> Line Bundles over an Induced Toric Variety: From now on, we will<br />

assume that the variety X is a complete toric variety corresponding to the fan Σ in X ∗ (T ) ∼ = Z r .<br />

To compute the cohomology groups H • (Ind(X), L) <strong>of</strong> Ind(X) with coefficients in a (G × G)-<br />

equivariant line bundle L, we take advantage <strong>of</strong> the decomposition <strong>of</strong> line bundles as described<br />

in Proposition 6. Since Ind(X) may be viewed as a fibration <strong>of</strong> G/B × G/B − with fiber X, this<br />

cohomology may be computed using a Leray spectral sequence.<br />

Theorem 7 ([18], Leray Spectral Sequence). Given the fiber bundle p : Ind(X) → G/B ×<br />

G/B − , suppose L is a (G × G)-equivariant line bundle on Ind(X). Then, for each integer q ≥ 0,<br />

define the sheaf H q (X, L) to be the sheaf on G/B × G/B − generated by the presheaf<br />

U ↦→ H q (p −1 U, L).<br />

Then there exists a spectral sequence such that<br />

E pq<br />

2 = Hp (G/B × G/B − , H q (X, L)),<br />

and whose abutment E ∞ is the bigraded group associated to a suitable filtration <strong>of</strong> the graded group<br />

H • (Ind(X), L).<br />

Thus, to compute the cohomology groups H • (Ind(X), L) as we desire, we must first compute<br />

the sheaves H q (X, L) for each integer q ≥ 0.<br />

Consider the stalk <strong>of</strong> the sheaf H q (X, L) at the<br />

image <strong>of</strong> the identity element in the flag variety, [e, e]. Take an affine trivializing cover {U} <strong>of</strong><br />

G/B × G/B − . Then<br />

H q (X, L) [e,e] = inj lim<br />

[e,e]∈U<br />

= inj lim<br />

[e,e]∈U<br />

H q (p −1 U, L) = inj lim H q (X × U, S ⊗ OInd(X) p ∗ M)<br />

[e,e]∈U<br />

⊕<br />

r∈Z<br />

= H q (X, S| X ) ⊗ k M [e,e] ,<br />

H r (X, S| X ) ⊗ k H q−r (U, M) = inj lim H q (X, S| X ) ⊗ k H 0 (U, M)<br />

[e,e]∈U<br />

using the Künneth formula. Therefore H q (X, L) is the induced (G × G)-sheaf over the flag variety<br />

30


G/B × G/B − corresponding to the B × B − -representation on the module H q (X, S| X ) ⊗ k M [e,e] .<br />

Now S| X is a line bundle on the toric variety X, so S| X = O X (D h ) for some Σ-linear support<br />

function h = h S .<br />

First consider the B ×B − -representation on M [e,e] . Since M is a line bundle on the flag variety<br />

G/B × G/B − ∼ = G × G/B × B − , it corresponds to a pair <strong>of</strong> characters (ξ, η) ∈ X ∗ (T ) × X ∗ (T ) [21].<br />

Therefore B × B − ’s representation on M [e,e] corresponds to the one-dimensional representation <strong>of</strong><br />

B × B − on k associated to the character (ξ, η): (b 1 , b 2 ) · a = ξ(b 1 )aη(b −1<br />

2 ), where ξ, η ∈ X∗ (T ).<br />

Therefore the B ×B − -action on M [e,e] effectively twists its action on H q (X, S| X ), which we discuss<br />

next.<br />

Since B × B − acts on X via the map β : (b 1 , b 2 ) ↦→ β(b 1 ) and T ’s action on X, we obtain<br />

a B × B − -representation in the cohomology groups H q (X, S| X ) for all integers q ≥ 0.<br />

Let us<br />

first note that S| X = i ∗ S is a line bundle on the toric variety X, and so is <strong>of</strong> the form O X (D h )<br />

for some Σ-linear support function h. (Since X is complete, Pic(X) ∼ = SF (X ∗ (T ), Σ).) As a T -<br />

representation, H q (X, O X (D h )) is a direct sum <strong>of</strong> its T -eigenspaces H q (X, O X (D h )) λ , λ ∈ X ∗ (T ).<br />

Since the B × B − -action on X factors through that <strong>of</strong> T , this decomposition also holds as a<br />

B × B − -representation. Thus, we have<br />

⎡<br />

H q (X, S| X ) ⊗ k M [e,e] = ⎣<br />

⊕<br />

λ∈X ∗ (T )<br />

= ⊕<br />

λ∈X ∗ (T )<br />

= ⊕<br />

λ∈X ∗ (T )<br />

H q Z(h S ,λ) (X ∗(T ) R , k)e λ ⎤<br />

⎦ ⊗ k M [e,e]<br />

[H q Z(h S ,λ) (X ∗(T ) R , k)e λ ⊗ k M [e,e]<br />

]<br />

[<br />

H q Z(h S ,λ) (X ∗(T ) R , k)e (λ+ξ,η)] .<br />

To compute the sheaf H q (X, L), we use the following.<br />

Lemma 6 ([26], Propositions I.3.3, I.5.9). If V, W are two K-representations, where K is a<br />

closed subgroup <strong>of</strong> H, then I H/K (V ⊕ W ) = I H/K (V ) ⊕ I H/K (W ).<br />

Pro<strong>of</strong>. I H/K (V ⊕ W ) is the sheaf U ↦→ [O H (π −1 U) ⊗ (V ⊕ W )] K , where U is an open set in H/K<br />

and π : H → H/K is the natural projection. Now [O H (π −1 U) ⊗ (V ⊕ W )] K = [(O H (π −1 U) ⊗<br />

V ) ⊕ (O H (π −1 U) ⊗ W )] K = (O H (π −1 U) ⊗ V ) K ⊕ (O H (π −1 U) ⊗ W ) K , which is exactly (I H/K (V ) ⊕<br />

I H/K (W ))(U).<br />

31


Therefore, since H q (X, S| X ) = ⊕ λ∈X ∗ (T )H q (X, S| X ) λ , we have<br />

H q (X, L) = I G×G/B×B −(H q (X, S| X ) ⊗ k M [e,e] )<br />

⊕<br />

= I G×G/B×B −( [H q (X, S| X ) λ ⊗ M [e,e] ])<br />

= ⊕<br />

λ∈X ∗ (T )<br />

λ∈X ∗ (T )<br />

I G×G/B×B −(H q (X, S| X ) λ ⊗ k M [e,e] ),<br />

so E pq<br />

2 = Hp (G/B ×G/B − , H q (X, L)) = H p (G×G/B ×B − , ⊕ λ∈X ∗ (T )I G×G/B×B −(H q (X, S| X ) λ ⊗ k<br />

M [e,e] )) = ⊕ λ∈X ∗ (T )H p (G/B × G/B − , I G×G/B×B −(H q (X, S| X ) λ ⊗ k M [e,e] )). Thus it suffices to<br />

compute each<br />

H p (G/B × G/B − , I G×G/B×B −(H q (X, S| X ) λ ⊗ k M [e,e] ))<br />

separately and then to add them together. Now H q (X, S| X ) λ = H q Z(h,λ) (X ∗(T ) R , k)e λ , where<br />

S| X = O X (D h ) for some Σ-linear support function h, by Proposition 5. Since H q (X, S| X ) λ ⊗ k<br />

M [e,e]<br />

∼ = H<br />

q<br />

Z(h,λ) (X ∗(T ) R , k)e (λ+ξ,η) , where M [e,e]<br />

∼ = k · (ξ, η) as a B × B − -representation, we have<br />

a filtration<br />

H q Z(h,λ) (X ∗(T ) R , k)e (λ+ξ,η) = N r ⊃ N r−1 ⊃ · · · ⊃ N 2 ⊃ N 1 ⊃ N 0 = 0<br />

where each L i = N i /N i−1 is a one-dimensional representation <strong>of</strong> B × B − <strong>of</strong> character (λ + ξ, η) and<br />

r = dim H q Z(h,λ) (X∗ (T ) R , k). Hence we get short exact sequences<br />

0 → N i → N i+1 → L i+1 → 0,<br />

which yield long exact sequences <strong>of</strong> cohomology (where we suppress the G/B × G/B − in the<br />

notation)<br />

0 → H 0 (N i ) → H 0 (N i+1 ) → H 0 (L i+1 ) → H 1 (N i ) → H 1 (N i+1 ) → H 1 (L i+1 ) → · · ·<br />

for i = 1, 2, . . . , r − 1. Now N 1 = L 1 , L 2 , L 3 , . . . , L r are line bundles over G/B × G/B − all <strong>of</strong> weight<br />

(λ + ξ, η), so their cohomology groups are non-zero in at most one index, which we call p λ . Then,<br />

32


since G (and hence G × G) is semi-simple, the short exact sequences<br />

0 → H p λ<br />

(N i−1 ) → H p λ<br />

(N i ) → H p λ<br />

(L i ) → 0<br />

split, implying that H p λ(N i ) = H p λ(N i−1 ) ⊕ H p λ(L i ) = H p λ(L 1 ) ⊕ H p λ(L 2 ) ⊕ · · · ⊕ H p λ(L i ), by<br />

induction, for i = 1, 2, . . . , r. In particular, H p λ(N r ) = ⊕ r<br />

i=1 Hp λ(L i ), so that<br />

H p (G/B × G/B − ,I G×G/B×B −(H q (X, S| X ) λ ⊗ k M [e,e] )) = H p (N r )<br />

⎧<br />

⎪⎨ H p λ(G/B × G/B − , L(λ + ξ, η)) ⊕ dim Hq Z(h,λ) (X∗(T ) R,k)<br />

p = p λ<br />

=<br />

⎪⎩ 0 p ≠ p λ<br />

Therefore,<br />

E pq<br />

2 = Hp (G/B × G/B − , H q (X, L))<br />

= ⊕ [<br />

] H p λ<br />

(G/B × G/B − , L(λ + ξ, η)) ⊕ dim Hq Z(h,λ) (X∗(T ) R,k)<br />

.<br />

{λ:p λ =p}<br />

Since these cohomology groups will be (G × G)-representations, they are also T -representations via<br />

T → T × T → G × G. Furthermore, all the maps must respect this T -action, so a λ-eigenspace may<br />

only be sent to another λ-eigenspace. However, the λ-eigenspaces occur in at most one column <strong>of</strong><br />

the spectral sequence E pq<br />

2 , so all boundary maps must be zero. Therefore the sequence is degenerate,<br />

so we may compute the cohomology <strong>of</strong> Ind(X) as the abutment:<br />

H n (Ind(X), L) ∼ = ⊕<br />

p+q=n<br />

∼= ⊕<br />

p+q=n<br />

E pq<br />

2<br />

⎛<br />

⎝<br />

⊕<br />

{λ:p λ =p}<br />

[<br />

H p λ<br />

(G/B × G/B − , L(λ + ξ, η)) ⊕ dim Hq Z(h,λ) (X∗(T ) R,k) ]⎞ ⎠ .<br />

33


Chapter 3<br />

Affine group embeddings<br />

Let G be a connected reductive algebraic group defined over an algebraically closed field k <strong>of</strong><br />

characteristic 0. By a G-variety, we mean a k-variety X together with a morphism σ : G × X → X,<br />

written (g, x) ↦→ g · x, satisfying g 1 · (g 2 · x) = (g 1 g 2 ) · x and e · x = x for all g 1 , g 2 ∈ G and all x ∈ X,<br />

where e ∈ G denotes the identity element. We are interested in the following class <strong>of</strong> G-varieties.<br />

Definition 8. A G-embedding is a normal G-variety X that contains an open orbit Ω isomorphic<br />

to G. The closed subvariety ∂X = X − Ω is called the boundary <strong>of</strong> X. Since Ω is a G-orbit, ∂X<br />

is a G-stable divisor <strong>of</strong> X unless Ω = X. The irreducible components <strong>of</strong> ∂X are G-stable prime<br />

divisors <strong>of</strong> X.<br />

Before we begin our classification <strong>of</strong> affine G-embeddings in this chapter, we study the nature<br />

<strong>of</strong> an action <strong>of</strong> a reductive group G on an affine variety X, and, in particular, the induced actions<br />

<strong>of</strong> the one-parameter subgroups <strong>of</strong> G.<br />

3.1 Group actions on affine varieties<br />

In this section, we recall the results <strong>of</strong> [27] that, with their pro<strong>of</strong>s, will be exploited throughout the<br />

chapter. Some <strong>of</strong> this material also appears in [32].<br />

Let X be an affine variety with an action by a reductive group G, σ : G × X → X. This action<br />

corresponds to the coaction σ ◦ : k[X] → k[G] ⊗ k[X], which will be critical for the following.<br />

Lemma 7 ([27], Lemma 1.1). Let X be an affine G-variety and suppose that Y is a closed<br />

G-stable subvariety <strong>of</strong> X. Then:<br />

34


1. There is a finite-dimensional G-representation V and an equivariant closed embedding X ↩→<br />

V .<br />

2. There is a finite-dimensional G-representation W and an equivariant map f : X → W such<br />

that Y is the scheme-theoretic inverse image f −1 (0) <strong>of</strong> the reduced subscheme <strong>of</strong> W supported<br />

by 0.<br />

Pro<strong>of</strong>. As observed, the action <strong>of</strong> G on X corresponds to a coaction σ ◦ : k[X] → k[G] ⊗ k[X],<br />

giving k[X] the structure <strong>of</strong> a left k[G]-comodule. Recall that any finite subset <strong>of</strong> a k[G]-comodule<br />

is contained in a finite dimensional subcomodule. Therefore, if f 1 , . . . , f r generate the affine algebra<br />

k[X] over k, there is a finite dimensional subcomodule V ′ <strong>of</strong> k[X] that contains f 1 , . . . , f r and so<br />

generates k[X] as a k-algebra. Set V = Spec Sym k (V ′ ). This is a G-representation and the map<br />

X → V induced by the surjective map Sym k (V ′ ) → k[X] is an equivariant closed immersion.<br />

Now let I ⊂ k[X] denote the ideal defining the closed subvariety Y <strong>of</strong> X. Since Y is G-stable,<br />

σ ◦ : I → k[G] ⊗ I, so I is a subcomodule <strong>of</strong> k[X]. As k[X] is noetherian, I is finitely generated as<br />

a k[X]-module, so let f 1 ′, . . . , f s ′ be a set <strong>of</strong> generators for I. As before, there is a finite dimensional<br />

G-stable subspace W ′ ⊂ k[X] containing f 1 ′, . . . , f s. ′ Let W = Spec Sym k (W ′ ). This is a G-<br />

representation, and the map f : X → W corresponding to the homomorphism Sym k (W ′ ) → k[X]<br />

has the properties claimed in part 2 <strong>of</strong> the Lemma.<br />

An effective tool in the study <strong>of</strong> affine actions is the use <strong>of</strong> one-parameter subgroups.<br />

A<br />

one-parameter subgroup <strong>of</strong> G is a homomorphism <strong>of</strong> algebraic groups γ : G m → G, which thus<br />

corresponds to a map γ ◦ : k[G] → k[t, t −1 ]. Let X ∗ (G) denote the set <strong>of</strong> one-parameter subgroups<br />

<strong>of</strong> G. The inclusion <strong>of</strong> k[t, t −1 ] in k((t)) allows us to view X ∗ (G) as a subset <strong>of</strong> G k((t)) =<br />

Hom k (k[G], k((t))), the set <strong>of</strong> k((t))-points <strong>of</strong> G. Let 〈γ〉 ∈ G k((t)) denote the point corresponding<br />

to the one-parameter subgroup γ. The group G k((t)) contains the subgroup G k[[t]] , which consists<br />

<strong>of</strong> all k((t))-points <strong>of</strong> G that have a specialization in G as t → 0. The group G k((t)) is the disjoint<br />

union <strong>of</strong> the double cosets <strong>of</strong> G k[[t]] , as described by the Iwahori decomposition:<br />

Theorem 8 ([25], Cartan–Iwahori Decomposition). Let G be a reductive algebraic group over<br />

k. Every double coset <strong>of</strong> G k((t)) with respect to the subgroup G k[[t]] is represented by a point <strong>of</strong> the<br />

35


type 〈γ〉, for some one-parameter subgroup γ <strong>of</strong> G. That is,<br />

G k((t)) =<br />

⋃<br />

G k[[t]] 〈γ〉G k[[t]] (3.1)<br />

γ∈X ∗(G)<br />

Moreover, each double coset is represented by a unique dominant one-parameter subgroup.<br />

This decomposition will be essential for replacing k((t))-points <strong>of</strong> G with one-parameter subgroups<br />

in the pro<strong>of</strong> <strong>of</strong> Theorem 9 below.<br />

For a one-parameter subgroup γ <strong>of</strong> G, recall that, by definition, lim t→0 γ(t)x 0 exists in X if<br />

γ : G m → G, when composed with ψ x0<br />

: g ↦→ g · x 0 , extends to a morphism ˜γ : A 1 → X and<br />

lim t→0 γ(t)x 0 is defined to be ˜γ(0). That is, the composition <strong>of</strong> ψ ◦ x 0<br />

: k[X] → k[G] with γ ◦ : k[G] →<br />

k[t, t −1 ] factors through k[t], and the limit lim t→0 γ(t)x 0 is the k-point <strong>of</strong> X corresponding to the<br />

composite k[X] → k[t] → k sending t → 0. This is described by the diagrams:<br />

G m ⊂ <br />

γ<br />

A 1˜γ 0<br />

<br />

<br />

k[X]<br />

<br />

<br />

<br />

˜γ 0 ◦<br />

G<br />

ψx0<br />

X k[t]<br />

ψ ◦ x 0<br />

k[G]<br />

γ ◦<br />

⊂ k[t, t −1 ].<br />

Similarly, if λ is a k((t))-point <strong>of</strong> G, then lim t→0 λ(t)x 0 exists in X means λ ◦ | k[X] : k[X] → k[[t]].<br />

The following lemma is used frequently hereafter.<br />

Lemma 8. Suppose λ ∈ G k((t)) and α ∈ G k[[t]] , so that α has specialization α 0 ∈ G k . Let<br />

X be an affine G-embedding with base point x 0 . Then lim t→0 [λ(t)x 0 ] exists in X if and only if<br />

lim t→0 [α(t)λ(t)x 0 ] exists, in which case<br />

lim [α(t)λ(t)x 0] = α 0 · lim[λ(t)x 0 ]. (3.2)<br />

t→0 t→0<br />

Pro<strong>of</strong>. Since α has a specialization in G k , α corresponds to a ring homomorphism α ◦ : k[G] → k[[t]].<br />

Now consider λ ∈ G k((t)) . Suppose lim t→0 λ(t)x 0 ∈ X. Then λ ◦ ◦ψ ◦ x 0<br />

: k[X] → k[G] → k((t)) factors<br />

through k[[t]] ⊂ k((t)). Moreover, (αλ) ◦ ◦ ψ x0<br />

is given by the composition k[X] → k[G] ⊗ k k[X] →<br />

k[[t]] ⊗ k k[[t]] → k[[t]] in the diagram below, so αλ ∈ X k[[t]] . Hence, lim t→0 [α(t)λ(t)x 0 ] exists in X,<br />

and its value is given by composing (αλ) ◦ ◦ ψ ◦ x 0<br />

with the map k[[t]] → k sending t ↦→ 0. However,<br />

36


the following (obviously) commutative diagram<br />

k[X]<br />

ψ ◦ x 0<br />

k[G]<br />

(αλ) ◦ k((t))<br />

<br />

coaction<br />

(αλ) ◦ ◦ψ ◦ x 0<br />

⊂<br />

k[G] ⊗ k k[X] α◦ ⊗(λ ◦ ◦ψx ◦ 0<br />

) k[[t]] ⊗ k k[[t]]<br />

<br />

α ◦ 0 ⊗(λ◦ ◦ψx ◦ 0<br />

)<br />

k ⊗ k k[[t]]<br />

mult.<br />

(t↦→0)⊗id<br />

t↦→0<br />

k[[t]]<br />

k<br />

shows that we may evaluate this composition by first mapping α to α 0 and then taking the limit<br />

lim t→0 λ(t)x 0 . Therefore, lim t→0 [α(t)λ(t)x 0 ] = α 0 · lim t→0 λ(t)x 0 as claimed.<br />

Conversely, if lim t→0 [α(t)λ(t)x 0 ] ∈ X, apply the previous case <strong>of</strong> this lemma to α ′ = α −1 ∈ G k[[t]]<br />

and λ ′ = αλ ∈ G k((t)) , observing that λ = α ′ λ ′ :<br />

lim λ(t)x 0 = lim[α −1 (t)α(t)λ(t)x 0 ] = α −1<br />

t→0 t→0<br />

0 · lim[α(t)λ(t)x 0 ].<br />

t→0<br />

Therefore, lim t→0 λ(t)x 0 exists in X. Multiplying both sides by α 0 implies that lim t→0 [α(t)λ(t)x 0 ] =<br />

α 0 · [lim t→0 λ(t)x 0 ].<br />

Remark 2. Suppose X is a biequivariant G-embedding, so G has both a left and a right action<br />

on X and G may be identified with an open subvariety Ω which is stable for both actions. Then<br />

we could amplify Lemma 8 as follows: If α, β ∈ G k[[t]] and λ ∈ G k((t)) , then lim t→0 λ(t)x 0 ∈ X if<br />

and only if lim t→0 [α(t)λ(t)β(t)x 0 ] ∈ X, in which case<br />

lim [α(t)λ(t)β(t)x 0] = α 0 · [lim λ(t)x 0 ] · β ′<br />

t→0 t→0<br />

0,<br />

where β(t) · x 0 = x 0 · β ′ (t) for some β ′ ∈ G k[[t]] and α 0 , β ′ 0 ∈ G k denote the specializations <strong>of</strong><br />

α, β ′ , respectively. However, we must be careful, for lim t→0 [α(t)λ(t)β(t)x 0 ] does not have to equal<br />

lim t→0 [α 0 λ(t)β 0 x 0 ], where α 0 , β 0 are the specializations <strong>of</strong> α, β [27].<br />

The following theorem, which makes use <strong>of</strong> Lemmas 7 and 8 and Theorem 8, will be essential<br />

for the pro<strong>of</strong> <strong>of</strong> Theorem 10 in the next section.<br />

37


Theorem 9 ([27], Theorem 1.4). Let X be an affine G-variety. Suppose that Y is a closed<br />

G-stable subvariety <strong>of</strong> X and that x 0 ∈ X is a closed point such that the closure <strong>of</strong> the orbit Gx 0<br />

intersects Y . Then there is a one-parameter subgroup γ <strong>of</strong> G such that lim t→0 γ(t)x 0 ∈ Y .<br />

Pro<strong>of</strong>. Let y be a k-point <strong>of</strong> Y , which is contained in the closure <strong>of</strong> Gx 0 . We may find a curve η in<br />

Gx 0 that has y in its closure. Take a curve ξ in G that dominates η under the morphism g ↦→ g · x 0 .<br />

Let p : C → G be the rational mapping from a smooth complete curve C, which represents ξ. By<br />

construction, there is a k-point c 0 <strong>of</strong> C such that lim c→c0 p(c) · x 0 = y.<br />

Let R = Spec k[[t]] be the spectrum <strong>of</strong> the formal power series ring in one variable t. As the<br />

completion <strong>of</strong> the local ring O C,c0<br />

<strong>of</strong> the curve C at c 0 is isomorphic to k[[t]], we have a rational<br />

mapping q : R → G such that lim t→0 q(t) · x 0 = y, where 0 is the closed point (t = 0) <strong>of</strong> R.<br />

By Theorem 8, we may find two morphisms α 1 , α 2 : R → G such that α 1 · q = 〈γ〉 · α 2 , where<br />

〈γ〉 : R → G is the the rational mapping given by taking the Laurent series expansion at 0 <strong>of</strong> a<br />

one-parameter subgroup γ <strong>of</strong> G.<br />

Let g i be the k-point α i (0) <strong>of</strong> G. The following limits exist and are equal in X by Lemma 8:<br />

g 1 · y = lim<br />

t→0<br />

[α 1 (t)] · lim<br />

t→0<br />

[q(t) · x 0 ] = lim<br />

t→0<br />

[〈γ〉(t) · α 2 (t) · x 0 ].<br />

Unfortunately, this limit does not always equal lim t→0 [γ(t) · g 2 · x 0 ]. However, the results <strong>of</strong> the<br />

following claim are sufficient to complete the pro<strong>of</strong>.<br />

Claim ([27]). 1. The limit lim t→0 [γ(t) · g 2 · x 0 ] exists in X.<br />

2. If X is a representation <strong>of</strong> G and y is the zero point 0, then lim t→0 [γ(t) · g 2 · x 0 ] = 0.<br />

First, we will show how the claim implies the theorem. By Lemma 7, we may find a G-<br />

equivariant morphism f : X → W , where W is a G-representation and Y = f −1 (0). By part 1 <strong>of</strong><br />

the claim, z := lim t→0 [γ(t)·g 2·x 0 ] exists in X. To show that z is in Y , it is enough to prove f(z) = 0,<br />

where f(z) = lim t→0 [γ(t) · g 2 · f(x 0 )]. As the above argument applies equally well for x ′ = f(x 0 ),<br />

y ′ = f(y) = 0 and X ′ = W , by part 2 <strong>of</strong> the claim, we must have f(z) = lim t→0 [γ(t) · g 2 · x ′ ] = 0.<br />

Therefore, lim t→0 [g −1<br />

2 γ(t)g 2 · x 0 ] = g −1<br />

2 · z exists and is contained in Y . Taking λ(t) = g −1<br />

2 γ(t)g 2,<br />

we have found the desired one-parameter subgroup.<br />

38


Now, to prove the claim, by Lemma 7, we may assume that X is a G-representation.<br />

As<br />

g 2 = α 2 (0), we may write α 2 (t) · x 0 = g 2 · x 0 + ε(t), where ε(t) involves only positive powers <strong>of</strong> t.<br />

Let ∑ i (g 2 · x 0 ) i + ε i (t) be the eigendecomposition <strong>of</strong> α 2 (t) · x 0 with respect to the G m -action via<br />

γ, so ε i (t) = 0 whenever i ≤ 0. Then,<br />

γ(t) · α 2 (t) · x 0 = ∑ i<br />

t i [(g 2 · x 0 ) i + ε i (t)].<br />

As the limit lim t→0 [γ(t)·α 2 (t)·x 0 ] exists, (g 2·x 0 ) i = 0 for all i < 0. Therefore, because γ(t)·g 2·x 0 =<br />

∑<br />

i≥0 ti (g 2 · x 0 ) i , part 1 <strong>of</strong> the claim is true and lim t→0 [γ(t) · g 2 · x 0 ] = (g 2 · x 0 ) 0 .<br />

For part 2 <strong>of</strong> the claim, if y is zero, then lim t→0 [γ(t)·α 2 (t)·x 0 ] = g 1 ·0 = 0. Hence, (g 2 ·x 0 ) i = 0<br />

if i ≤ 0. Therefore, lim t→0 [γ(t) · g 2 · x 0 ] = lim t→0<br />

∑i>0 ti (g 2 · x 0 ) i = 0, which proves the claim and<br />

hence the theorem.<br />

3.2 Classification <strong>of</strong> affine group embeddings<br />

The main objective <strong>of</strong> this chapter is the classification <strong>of</strong> all affine G-embeddings. This will be done<br />

in a sequence <strong>of</strong> three steps. First, we show that affine G-embeddings X are determined by the<br />

collection <strong>of</strong> one-parameter subgroups <strong>of</strong> G that have limit in the embedding. This is accomplished<br />

in the first subsection. Second, we study such sets <strong>of</strong> one-parameter subgroups <strong>of</strong> G and show that<br />

they are strongly convex lattice cones as in Definition 11. Finally, we show that every strongly<br />

convex lattice cone determines an affine G-embedding.<br />

3.2.1 Limits <strong>of</strong> one-parameter subgroups<br />

Our primary method for describing and thus classifying affine G-embeddings X is to make use<br />

<strong>of</strong> one-parameter subgroups <strong>of</strong> G. We are interested in the limits, when they exist, <strong>of</strong> the oneparameter<br />

subgroups <strong>of</strong> G in X. One-parameter subgroups and their limits have been employed in<br />

a number <strong>of</strong> applications, including the Hilbert-Mumford criterion <strong>of</strong> stability [32], the construction<br />

<strong>of</strong> the spherical building <strong>of</strong> the group G [32] and the Bialynicki-Birula decomposition <strong>of</strong> a smooth<br />

projective T -variety [1]. For our purposes, we will show that an affine G-embedding X is determined<br />

by the set <strong>of</strong> one-parameter subgroups γ <strong>of</strong> G such that lim t→0 γ(t)x 0 exists in X.<br />

39


Let X ∗ (G) denote the set <strong>of</strong> one-parameter subgroups <strong>of</strong> G. The group G acts on X ∗ (G) by<br />

conjugation, g • γ : t ↦→ gγ(t)g −1 . We will denote the trivial one-parameter subgroup t ↦→ e by ε.<br />

Each one-parameter subgroup γ ∈ X ∗ (G) determines a subgroup<br />

P (γ) = {g ∈ G : γ(t)gγ(t −1 ) ∈ G k[[t]] } (3.3)<br />

<strong>of</strong> G, which is parabolic if G is reductive [32]. In fact, every parabolic subgroup <strong>of</strong> a reductive<br />

group G is <strong>of</strong> the form P (γ) for some one-parameter subgroup γ <strong>of</strong> G [40]. We define an equivalence<br />

relation on the set <strong>of</strong> non-trivial one-parameter subgroups <strong>of</strong> G by<br />

γ 1 ∼ γ 2 if and only if γ 2 (t n 2<br />

) = gγ 1 (t n 1<br />

)g −1 (3.4)<br />

for positive integers n 1 , n 2 and an element g ∈ P (γ 1 ), for all t ∈ k × . Then the quotient (X ∗ (G) −<br />

{ε})/ ∼ is isomorphic to the spherical building <strong>of</strong> G [32], [48].<br />

Every parabolic subgroup P <strong>of</strong> G defines a subset ∆ P (G) = {γ ∈ X ∗ (G) : P (γ) ⊇ P } <strong>of</strong><br />

X ∗ (G). Clearly γ ∈ ∆ P (γ) (G) for all γ ∈ X ∗ (G), so X ∗ (G) = ⋃ ∆ P (G) where the union is over<br />

all parabolic subgroups <strong>of</strong> G. In the spherical building <strong>of</strong> G, the images <strong>of</strong> the sets ∆ P (G) are<br />

simplices and constitute a “triangulation” <strong>of</strong> the building so that it is the complex formed from<br />

these simplices [32].<br />

Given an affine G-embedding X with a choice <strong>of</strong> base point x 0 ∈ Ω, let<br />

Γ(X, x 0 ) = {γ ∈ X ∗ (G) : lim<br />

t→0<br />

γ(t)x 0 exists in X}. (3.5)<br />

Before we proceed, we make some immediate observations about such sets.<br />

Proposition 7. Let G be a connected reductive group. Suppose X is an affine G-embedding and<br />

x 0 ∈ X is a base point.<br />

1. If x ′ 0 = hx 0, then Γ(X, x ′ 0 ) = hΓ(X, x 0)h −1 .<br />

2. If γ ∈ Γ(X, x 0 ) and γ ≠ ε, then γ −1 ∉ Γ(X, x 0 ).<br />

3. If T is any torus <strong>of</strong> G, then T x 0<br />

∼ = T σ , where σ = Γ(X, x 0 ) ∩ X ∗ (T ) is a strongly convex<br />

40


ational polyhedral cone in X ∗ (T ) (Section 2.1.1, Definition 3).<br />

4. If γ ∈ Γ(X, x 0 ) and p ∈ P (γ), then pγ(t)p −1 ∈ Γ(X, x 0 ), and moreover<br />

Γ(X, x 0 ) = ⋃<br />

P • (Γ(X, x 0 ) ∩ ∆ P (G)) (3.6)<br />

P ⊂G<br />

where the union is taken over all parabolic subgroups P <strong>of</strong> G.<br />

5. The image <strong>of</strong> Γ(X, x 0 ) in the spherical building is convex ([32], Definition 2.10).<br />

Pro<strong>of</strong>. First, Γ(X, x 0 ) depends on the base point x 0 as follows. Suppose x ′ 0 ∈ Ω, so x′ 0 = h · x 0<br />

for some unique h ∈ G (because G → Ω is an isomorphism). Then Γ(X, x ′ 0 ) = hΓ(X, x 0)h −1 for<br />

if γ ∈ Γ(X, x 0 ) (that is, if lim t→0 γ(t)x 0 ∈ X), then lim t→0 (hγ(t)h −1 )x ′ 0 = lim t→0 hγ(t)h −1 hx 0 =<br />

lim t→0 hγ(t)x 0 = h lim t→0 γ(t)x 0 , which exists in X. Therefore hΓ(X, x 0 )h −1 ⊂ Γ(X, x ′ 0 ). By<br />

symmetry, since x 0 = h −1 x ′ 0 , h−1 Γ(X, x ′ 0 )h ⊂ Γ(X, x 0), so Γ(X, x ′ 0 ) ⊂ hΓ(X, x 0)h −1 . Hence,<br />

Γ(X, h · x 0 ) = h Γ(X, x 0 ) h −1 (3.7)<br />

for any h ∈ G.<br />

Second, as X is affine, if γ ∈ Γ(X, x 0 ) and γ is not the trivial one-parameter subgroup ε : t ↦→ e,<br />

then γ −1 ∉ Γ(X, x 0 ). Otherwise, if both lim t→0 γ(t)x 0 and lim t→0 γ −1 (t)x 0 exist in X, then the<br />

composition ψ x0 ◦ γ : G m → X extends to a morphism ˜γ : P 1 → X, which must therefore be<br />

constant, so γ = ε.<br />

Third, if T is any torus <strong>of</strong> G, then T x 0<br />

∼ = Tσ , where σ ⊂ X ∗ (T ) is the strongly convex lattice<br />

cone Γ(X, x 0 ) ∩ X ∗ (T ) by Theorem 3.<br />

Now suppose γ ∈ Γ(X, x 0 ) and p ∈ P (γ).<br />

Then p · γ · p −1 also belongs to Γ(X, x 0 ), for<br />

lim t→0 [(pγ(t)p −1 )x 0 ] = lim t→0 [p(γ(t)p −1 γ(t −1 ))γ(t)x 0 ] = p[lim t→0 γ(t)p −1 γ(t −1 )][lim t→0 γ(t)x 0 ]<br />

exists in X by Lemma 8 and the definition <strong>of</strong> P (γ). Therefore, it is clear that Γ(X, x 0 ) =<br />

⋃<br />

P ⊂G P • (Γ(X, x 0) ∩ ∆ P (G)), where the union is taken over all parabolic subgroups P <strong>of</strong> G.<br />

Lastly, if δ 1 , δ 2 ∈ (Γ(X, x 0 )−{ε})/ ∼, then there are one-parameter subgroups γ 1 , γ 2 ∈ Γ(X, x 0 )<br />

such that δ i = [γ i ] is the equivalence class <strong>of</strong> γ i . The one-parameter subgroups determine parabolic<br />

subgroups P (γ 1 ) and P (γ 2 ), whose intersection contains a maximal torus T <strong>of</strong> G. Then γ 1 and γ 2 are<br />

41


equivalent to one-parameter subgroups γ 1 ′ , γ′ 2 ∈ X ∗(T ) and δ i = [γ i ′]. By part 4, γ′ 1 , γ′ 2 ∈ Γ(X, x 0)<br />

as well. Then γ 1 ′ , γ′ 2 ∈ Γ(X, x 0) ∩ X ∗ (T ), which is the strongly convex rational polyhedral cone<br />

associated to the toric variety T ⊂ X by part 3. As strongly convex rational polyhedral cones are<br />

convex, the line in X ∗ (T ) joining γ 1 and γ 2 is contained in Γ(X, x 0 ) ∩ X ∗ (T ), and hence the line<br />

in the spherical building joining δ 1 and δ 2 is contained in the image <strong>of</strong> Γ(X, x 0 ), so this image is<br />

semi-convex. It is convex by part 2, which implies no pair <strong>of</strong> antipodal points <strong>of</strong> the building can<br />

belong to the image <strong>of</strong> Γ(X, x 0 ).<br />

Consider the following examples.<br />

Example 5 (G viewed as an affine G-embedding). Consider the trivial G-embedding, G ⊂ G<br />

with base point e. Here Γ(G, e) = {γ ∈ X ∗ (G) : lim t→0 γ(t) exists in G} = {ε}, where ε denotes<br />

the trivial one-parameter subgroup, t ↦→ e, <strong>of</strong> G.<br />

Example 6 (One variety as two distinct embeddings). Let B = {( a −1 b<br />

0 a<br />

)<br />

: a, b ∈ k, a ≠ 0<br />

}<br />

,<br />

which is not a reductive group. Then B → X = A 2 via ( a −1 b<br />

0 a<br />

)<br />

↦→ (b, a) is a B-embedding. Here<br />

⎧ ⎛ ⎞ ⎫<br />

⎪⎨<br />

⎪⎬<br />

Γ(X, (0, 1)) =<br />

⎪⎩ γ(t) = ⎜<br />

⎝ t−n 0 ⎟<br />

⎠ : n ∈ N 0 .<br />

0 t n ⎪ ⎭<br />

However, B → Y = A 2 given by ( a −1 b<br />

0 a<br />

)<br />

↦→ (a −1 , a −1 b) is another B-embedding into A 2 , but<br />

⎧ ⎛<br />

⎞<br />

⎫<br />

⎪⎨<br />

⎪⎬<br />

Γ(Y, (1, 0)) =<br />

⎪⎩ γ(t) = ⎜<br />

⎝ t−n c(t n − t −n ) ⎟<br />

⎠ : c ∈ k, −n ∈ N 0 .<br />

0 t n ⎪ ⎭<br />

Therefore, isomorphism <strong>of</strong> embeddings is a finer relation than isomorphism <strong>of</strong> varieties. As a result,<br />

our notation Γ(X, x 0 ) may seem a bit clumsy as it does not indicate the manner in which B is<br />

embedded in X, but this will be clear from the context whenever it is used.<br />

Example 7 (Γ(X, x 0 ) for the group embedding associated to a toric variety). Suppose T is<br />

a maximal torus <strong>of</strong> G and T σ is an affine toric variety for T . Then G × T T σ is an affine embedding<br />

<strong>of</strong> G. Clearly σ ⊆ Γ(G × T T σ , [e, e]), as σ = Γ(G × T T σ , [e, e]) ∩ X ∗ (T ). For each parabolic<br />

subgroup P <strong>of</strong> G containing T , let ∆ P (T ) = {γ ∈ X ∗ (T ) : P (γ) ⊇ P }. Then Γ(G × T T σ , [e, e]) =<br />

42


⋃<br />

P ⊃T P • (σ ∩ ∆ P (T )) is a simplicial decomposition <strong>of</strong> Γ(G × T T σ , [e, e]) ⊂ X ∗ (G). We remark<br />

that this is a finite simplicial decomposition, for there are only finitely many parabolic subgroups<br />

<strong>of</strong> G containing T . (The Borel subgroups <strong>of</strong> G that contain T are all conjugate to a fixed one<br />

by elements <strong>of</strong> the Weyl group, W (T, G), which is finite, and the set <strong>of</strong> parabolic subgroups <strong>of</strong> G<br />

containing a given Borel subgroup B ⊃ T are indexed by subsets <strong>of</strong> the set <strong>of</strong> simple roots ∆(T, B)<br />

relative to B.)<br />

For instance, consider the GL 2 -embedding GL 2 × D 2<br />

A 2 . Here σ = {γ m,n (t) = ( t m 0<br />

0 t n )<br />

: m, n ∈<br />

N 0 } ⊂ X ∗ (D 2 ). There are only three parabolic subgroups <strong>of</strong> GL 2 containing D 2 , namely B + =<br />

{ ( )<br />

a b<br />

0 d }, B − = { ( )<br />

a 0<br />

c d } and GL2 itself. Then<br />

∆ B +(D 2 ) = {γ m,n ∈ X ∗ (D 2 ) : m ≥ n},<br />

∆ B −(D 2 ) = {γ m,n ∈ X ∗ (D 2 ) : m ≤ n},<br />

∆ GL2 (D 2 ) = {γ m,n ∈ X ∗ (D 2 ) : m = n},<br />

so σ ∩ ∆ B +(D 2 ) = {γ m,n : m ≥ n ≥ 0}, σ ∩ ∆ B −(D 2 ) = {γ m,n : n ≥ m ≥ 0} and σ ∩ ∆ GL2 (D 2 ) =<br />

{γ m,n : m = n ≥ 0}. Now<br />

B + • (σ ∩ ∆ B +(D 2 )) = {t ↦→<br />

(<br />

B − • (σ ∩ ∆ B −(D 2 )) = {t ↦→<br />

(<br />

)<br />

t m bt n (1−t m−n )<br />

0 t n<br />

)<br />

t m 0<br />

ct m (1−t n−m ) t n<br />

GL 2 • (σ ∩ ∆ GL2 (D 2 )) = {t ↦→ ( t m 0<br />

0 t n )<br />

: m = n ≥ 0},<br />

: m ≥ n ≥ 0},<br />

: n ≥ m ≥ 0},<br />

with limits<br />

⎧<br />

[( 1 b<br />

0 1<br />

[( 1 b<br />

0 1<br />

)<br />

, (0, 0)<br />

]<br />

)<br />

, (0, 1)<br />

]<br />

m > n > 0<br />

m > n = 0<br />

lim [(p • γ m,n(t)) · [e, (1, 1)]] =<br />

t→0<br />

⎪⎨<br />

⎪⎩<br />

[( 1 c 0<br />

1<br />

) , (0, 0)] n > m > 0<br />

[( 1 c 0<br />

1<br />

) , (1, 0)] n > m = 0<br />

[( 1 0 0<br />

1<br />

) , (0, 0)] m = n > 0<br />

[( 1 0 0<br />

1<br />

) , (1, 1)] m = n = 0<br />

43


and we see Γ(GL 2 × D 2<br />

A 2 , [e, e]) = ⋃ P ⊃D 2<br />

P • (σ ∩ ∆ P (D 2 )).<br />

Example 8 (The affine GL n -embedding M n ). Consider G = GL n and X = M n , the affine<br />

space <strong>of</strong> n × n matrices. Clearly M n is an affine GL n -embedding with the obvious left action. Any<br />

( t p 1 0<br />

one-parameter subgroup <strong>of</strong> G can be obtained from a one-parameter subgroup t ↦→ . .. ,<br />

0 t pn )<br />

for p 1 , . . . , p n ∈ Z, after suitable conjugation by some matrix A ∈ GL n . Such a one-parameter<br />

( t p 1 0<br />

)<br />

subgroup A . .. A −1 has a limit in M n if and only if p 1 , . . . , p n ≥ 0. Thus Γ(M n , I n ) =<br />

( 0 t pn<br />

t p 1 0<br />

{A . .. A −1 : A ∈ GL n and p 1 , . . . , p n ∈ N 0 }.<br />

0 t pn )<br />

Each γ ∈ X ∗ (G) may be viewed as a k((t))-point <strong>of</strong> G.<br />

In [30], a G-stable valuation v λ is<br />

associated to every λ ∈ G k((t)) in the following way.<br />

As λ is a k((t))-point <strong>of</strong> G, we obtain a<br />

dominant morphism<br />

G × Spec k((t)) 1×λ G × G µ G.<br />

This morphism induces an injection <strong>of</strong> fields i λ : k(G) → Frac(k(G) ⊗ k k((t))) → k(G)((t)). Then<br />

v t ◦ i λ : k(G) × → Z is a valuation <strong>of</strong> k(G), where v t : k(G)((t)) × → Z is the standard valuation<br />

associated to the order <strong>of</strong> t. We define v λ = 1 (v t ◦i λ ), where n λ ∈ Z is the largest positive number<br />

n λ<br />

such that (v t ◦ i λ )(k(G) × ) ⊂ n λ Z (except when λ = ε, in which case v ε (f) = 0 or ∞ as f(e) ≠ 0 or<br />

= 0, respectively). This is G-stable by left translations, i.e., v λ (s · f) = v λ (f) for all s ∈ G, since<br />

i λ is clearly equivariant and k(G)[[t]] is obviously stable for left translations by G in k(G)((t)). We<br />

include some <strong>of</strong> the properties <strong>of</strong> these valuations that are proven in [30] in the following lemma.<br />

Lemma 9 ([30]).<br />

1. Let γ be a one-parameter subgroup <strong>of</strong> G. For each f ∈ k(G), there is an<br />

open subset U ⊂ G, depending only on f, such that<br />

v γ (f) = inf<br />

s∈U v t(f(s · γ(t))) (3.8)<br />

2. Let γ 1 , γ 2 be one-parameter subgroups <strong>of</strong> G. Then v γ1 = v γ2 if and only if γ 1 ∼ γ 2 .<br />

Pro<strong>of</strong>. Part 1 is Lemma 4.11.1 in [30], where U = {s ∈ G : f(s) ≠ 0}. The second part is the result<br />

<strong>of</strong> Propositions 3.3 and 5.4 in [30].<br />

The sets <strong>of</strong> one-parameter subgroups Γ(X, x 0 ) described in (3.5) are significant for the following<br />

reasons.<br />

The first result, which will serve as our foundation for the classification theorem in<br />

44


Section 3.2.3, is a uniqueness theorem which shows that an affine G-embedding X with base point<br />

x 0 is determined by the set Γ(X, x 0 ). The second demonstrates that the prime divisors on the<br />

boundary <strong>of</strong> an affine G-embedding correspond to equivalence classes <strong>of</strong> edges <strong>of</strong> the set Γ(X, x 0 ).<br />

Theorem 10. Let G be a connected reductive group. If X is an affine G-embedding with base point<br />

x 0 , then X ∼ = Spec A Γ(X,x0 ), where<br />

A Γ(X,x0 ) := {f ∈ k[G] : v γ (f) ≥ 0 for all γ ∈ Γ(X, x 0 )}. (3.9)<br />

Pro<strong>of</strong>. The base point x 0 defines a morphism ψ x0 : g ↦→ g · x 0 from G to X. As both G and X are<br />

affine, ψ x0 corresponds to a homomorphism ψ ◦ x 0<br />

: k[X] → k[G], which is injective since the image<br />

<strong>of</strong> G is open in X and X is irreducible. The image <strong>of</strong> ψ ◦ x 0<br />

lies in the subalgebra A Γ(X,x0 ) since<br />

every γ ∈ Γ(X, x 0 ) extends to a morphism ˜γ : A 1 → X so that f(g · ˜γ(0)) exists, which implies that<br />

v γ (f) = inf s∈Gf v t (f(s · γ(t))) ≥ 0 for all f ∈ k[X]. We claim that k[X] ∼ = A Γ(X,x0 ). It suffices to<br />

show that every f ∈ A Γ(X,x0 ) extends to a regular function on X to prove that k[X] → A Γ(X,x0 ) is<br />

surjective and hence that k[X] ∼ = A Γ(X,x0 ).<br />

Suppose not and assume that f ∈ A Γ(X,x0 ) is not in the image <strong>of</strong> k[X]. Then f fails to extend<br />

to a regular function on X, but it is defined on Ω = Gx 0 by f(g · x 0 ) := f(g). Let P be the divisor<br />

<strong>of</strong> poles <strong>of</strong> f in X. Then P is closed, has codimension one in X, and P ⊆ ∂X. Let D be an<br />

irreducible component <strong>of</strong> ∂X and hence a closed subvariety <strong>of</strong> X. Since ∂X is G-stable and G is<br />

connected (and so is irreducible), D is a G-stable prime divisor since e G ∈ G fixes the generic point<br />

<strong>of</strong> the irreducible subvariety D. Therefore, Theorem 9 provides a one-parameter subgroup γ D <strong>of</strong> G<br />

with lim t→0 γ D (t)x 0 ∈ D, so γ D ∈ Γ(X, x 0 ). Yet f ∈ A Γ(X,x0 ) implies that v γD (f) ≥ 0, so f must<br />

be defined on the orbit G[lim t→0 γ D (t)x 0 ] ⊂ D. Therefore, P ∩ D is a closed subset <strong>of</strong> D not equal<br />

to D, since P does not contain G[lim t→0 γ D (t)x 0 ] ⊂ D as v γD (f) ≥ 0. Thus the codimension <strong>of</strong><br />

P in X, which is equal to the minimum <strong>of</strong> the codimensions <strong>of</strong> the P ∩ D as D ranges over the<br />

irreducible components <strong>of</strong> ∂X, is at least 2. This is a contradiction. Hence every f ∈ A Γ(X,x0 )<br />

extends to a regular function on X, so is in the image <strong>of</strong> ψx ◦ 0<br />

. Therefore, ψx ◦ 0<br />

: k[X] → A Γ(X,x0 ) is<br />

an isomorphism, so X ∼ = Spec A Γ(X,x0 ) as claimed. Furthermore, the selected base point x 0 ∈ X<br />

corresponds to the maximal ideal m x0 = (ψx ◦ 0<br />

) −1 (m e ∩A Γ(X,x0 )) <strong>of</strong> k[X] as ψx ◦ 0<br />

identifies f ∈ A Γ(X,x0 )<br />

45


with the unique extension <strong>of</strong> the function f(g · x 0 ) := f(g) to X.<br />

Corollary 3. Let X be an affine G-embedding with base point x 0 . If x ′ 0 = hx 0 is another base<br />

point, then<br />

A Γ(X,h·x0 ) = r h (A Γ(X,x0 )), (3.10)<br />

where r h denotes right translation by h in k[G], r h (f)(x) = f(xh).<br />

Pro<strong>of</strong>. Suppose X is an affine G-embedding.<br />

We have shown in (3.7) that the set Γ(X, x 0 ) is<br />

determined by X only up to conjugation, as any other base point is <strong>of</strong> the form h · x 0 for a unique<br />

element h ∈ G and Γ(X, h · x 0 ) = hΓ(X, x 0 )h −1 . We now show that A Γ(X,h·x0 ) = r h (A Γ(X,x0 )).<br />

Suppose f ∈ r h (A Γ(X,x0 )). Then f = r h (f ′ ) for some f ′ ∈ A Γ(X,x0 ), which means that v γ (f ′ ) ≥ 0<br />

for all γ ∈ Γ(X, x 0 ). If γ ′ ∈ Γ(X, h · x 0 ) = hΓ(X, x 0 )h −1 , write γ ′ = h • γ for γ ∈ Γ(X, x 0 ) and<br />

consider<br />

v h•γ (f) = inf<br />

s∈U v t(f(s · hγ(t)h −1 )) = inf<br />

s∈U v t((r h f ′ )(s · hγ(t)h −1 )) = inf<br />

s∈U v t(f ′ (s · hγ(t)h −1 · h))<br />

= inf v t(f ′ (sh · γ(t))) = inf v t(f ′ (s ′ · γ(t))) = v γ (f ′ ) ≥ 0.<br />

s∈U s ′ ∈Uh<br />

Therefore, f ∈ A Γ(X,h·x0 ), so r h (A Γ(X,x0 )) ⊂ A Γ(X,h·x0 ). Likewise, r h −1(A Γ(X,h·x0 )) ⊂ A Γ(X,x0 ), so<br />

r h (A Γ(X,x0 )) = A Γ(X,h·x0 ) as claimed.<br />

Consider the following examples.<br />

Example 9 (Recovering G from Γ(G, e)). Recall that in Example 5 we saw Γ(G, e) = {ε} for<br />

any group G. Now A {ε} = {f ∈ k[G] : v ε (f) ≥ 0} = k[G], and G = Spec k[G] = Spec A {ε} . Note<br />

that Γ(G, e) = {ε} is G-stable for conjugation and that A {ε} is right-invariant as well. This will be<br />

true in general. See Section 3.3.2 for more details.<br />

Example 10 (Recovering M n from Γ(M n , I n )). Recall Example 8, the embedding <strong>of</strong> G =<br />

{ ( t p 1 0<br />

GL n in X = M n . Here Γ(M n , I n ) = A . .. A −1 : A ∈ GL n , p 1 , . . . , p n ∈ N 0<br />

}, whose<br />

0 t pn )<br />

associated algebra is A Γ(Mn,I n) = {f ∈ k[GL n ] : v γ (f)) ≥ 0 for all γ ∈ Γ(M n , I n )}. This is naturally<br />

isomorphic to k[X 11 , . . . , X nn ] ⊂ k[GL n ] = k[X 11 , . . . , X nn , det −1 ], and clearly Spec A Γ(Mn,I n) =<br />

M n . Again we see that Γ is GL n -stable for conjugation, and that the corresponding embedding M n<br />

is biequivariant.<br />

46


Therefore, the classification <strong>of</strong> affine G-embeddings is equivalent to the characterization <strong>of</strong> such<br />

subsets <strong>of</strong> X ∗ (G) that are obtained from affine G-embeddings. In order to classify admissible subsets<br />

Γ, we will explore the properties <strong>of</strong> the sets Γ(X, x 0 ) for an arbitrary affine G-embedding X with<br />

choice <strong>of</strong> base point x 0 in the next section. We record one such result here.<br />

Recall from Section 2.1.4 that if σ is a strongly convex rational polyhedral cone in X ∗ (T ) R , then<br />

σ(1) denotes the set <strong>of</strong> rays <strong>of</strong> σ, σ(1) = {τ < σ : dim τ = 1}. This is called the one-skeleton <strong>of</strong><br />

the cone. By Proposition 7, for each maximal torus T <strong>of</strong> G, Γ(X, x 0 ) ∩ X ∗ (T ) is a strongly convex<br />

rational polyhedral cone in X ∗ (T ). So we may define the one-skeleton <strong>of</strong> the set Γ(X, x 0 ) to be<br />

Γ 1 (X, x 0 ) = ⋃ T<br />

[Γ(X, x 0 ) ∩ X ∗ (T )](1), (3.11)<br />

which is the set <strong>of</strong> extremal rays <strong>of</strong> Γ(X, x 0 ).<br />

Proposition 8. There is a bijection between Γ 1 (X, x 0 )/ ∼ and the finite set <strong>of</strong> prime divisors <strong>of</strong><br />

X contained in ∂X.<br />

Pro<strong>of</strong>. Write X = Ω ∪ D 1 ∪ D 2 ∪ · · · ∪ D r , where D 1 , . . . , D r are the finitely many G-stable prime<br />

divisors <strong>of</strong> X in ∂X.<br />

Suppose that ρ ∈ Γ 1 (X, x 0 ). Then there is a maximal torus T <strong>of</strong> G such that ρ ∈ (Γ(X, x 0 ) ∩<br />

X ∗ (T ))(1) is a ray <strong>of</strong> the strongly convex rational polyhedral cone Γ(X, x 0 ) ∩ X ∗ (T ) corresponding<br />

to T ⊂ X. From the description <strong>of</strong> T -stable divisors in toric varieties in Section 2.1.4, the ray ρ<br />

corresponds to a prime divisor D ρ in T . Thus D ρ is an irreducible T -stable subvariety <strong>of</strong> T , so G · D ρ<br />

is an irreducible G-stable subvariety <strong>of</strong> X, as both G and D ρ are irreducible. Moreover, G · D ρ is<br />

contained in ∂X, so there is a G-stable prime divisor D i <strong>of</strong> X such that G · D ρ ⊆ D i ⊂ X. However,<br />

codim X (G · D ρ ) = dim X − dim G · D ρ = dim G − (dim G + dim D ρ − dim T ) = dim T − dim D ρ =<br />

codim T<br />

(D ρ ) = 1. Thus, as G · D ρ is irreducible and <strong>of</strong> codimension one in X, and G · D ρ ⊂ D i X,<br />

where D i is also irreducible and <strong>of</strong> codimension one, we conclude that G · D ρ = D i . Therefore,<br />

there is a map Γ 1 (X, x 0 ) → {D 1 , D 2 , . . . , D r } given by ρ ↦→ G · D ρ .<br />

Furthermore, this map respects the equivalence relation ∼ described in (3.4), for if γ ρ denotes the<br />

first lattice point in X ∗ (T ) along ρ and γ ρ1 ∼ γ ρ2 , then lim t→0 γ ρ1 (t)x 0 and lim t→0 γ ρ2 (t)x 0 belong to<br />

the same G-orbit in X, and hence to the same prime divisor D <strong>of</strong> X, as D is G-stable. Therefore,<br />

47


G · D ρ1 = D = G · D ρ2 , from the discussion above. Hence Γ 1 (X, x 0 )/ ∼ → {D 1 , D 2 , . . . , D r } is<br />

well-defined. To prove that it is a bijection, we use the following<br />

Lemma 10. Let ρ ∈ Γ 1 (X, x 0 ) and let γ ρ denote the first lattice point in X ∗ (G) along ρ. Then v γρ<br />

is the valuation <strong>of</strong> G · D ρ .<br />

Pro<strong>of</strong>. The ideal Γ(X, O(G · D ρ )) = {f ∈ k[X] : f = 0 on G · D ρ } is equal to k[X] ∩ m vγρ , where<br />

m vγρ = {f ∈ k(G) : v γρ (f) > 0} is the maximal ideal <strong>of</strong> the valuation ring O vγρ = {f ∈ k(G) :<br />

v γρ (f) ≥ 0}. For D ρ = T · z ρ , where z ρ = lim t→0 γ ρ (t)x 0 ∈ T , so that G · D ρ = G · z ρ as well. Thus,<br />

if f ∈ k[X] ∩ m vγρ then f(z ρ ) = f(lim t→0 γ ρ (t)x 0 ) = lim t→0 t vγρ (f) u = 0, where u ∈ k[X] ∩ O v × γρ<br />

,<br />

since v γρ (f) > 0. Therefore k[X] ∩ m vγρ<br />

⊂ Γ(X, O(G · D ρ )), where both are prime ideals in k[X]<br />

and the latter is <strong>of</strong> height one. Hence they are equal, so the Lemma is proven.<br />

We return to the pro<strong>of</strong> <strong>of</strong> the Proposition. Suppose ρ 1 , ρ 2 ∈ Γ 1 (X, x 0 ) such that γ ρ1 ≁ γ ρ2 .<br />

Then v γρ1 ≠ v γρ2 , so G · D ρ1 ≠ G · D ρ2 . Hence Γ 1 (X, x 0 )/ ∼ → {D 1 , D 2 , . . . , D r } is injective. It<br />

is also surjective, as any <strong>of</strong> the prime divisors D i <strong>of</strong> X are closed G-subvarieties, and thus contain<br />

the limit point <strong>of</strong> some one-parameter subgroup γ ∈ Γ(X, x 0 ) by Theorem 9. This γ is a oneparameter<br />

subgroup <strong>of</strong> some maximal torus T <strong>of</strong> G, and so T ∩ D i ≠ ∅ is a T -stable divisor <strong>of</strong> T .<br />

By Lemma 2, there is a prime divisor D ρ <strong>of</strong> T corresponding to a ray ρ ∈ Γ(X, x 0 ) ∩ X ∗ (T ) such<br />

that D ρ ⊂ T ∩ D i . Thus G · D ρ ⊂ D i , from which we conclude G · D ρ = D i as before. Therefore,<br />

Γ 1 (X, x 0 )/ ∼ → {D 1 , D 2 , . . . , D r } is a bijection.<br />

3.2.2 States and strongly convex lattice cones<br />

Our sets Γ(X, x 0 ) <strong>of</strong> one-parameter subgroups associated to an affine G-embedding are identical to<br />

sets arising in geometric invariant theory as studied by Mumford [32], Kempf [27] and Rousseau [39].<br />

In [27], Kempf studies affine G-schemes X and one-parameter subgroups γ <strong>of</strong> G with a specialization<br />

in X. In his paper, Kempf describes such sets in terms <strong>of</strong> bounded, admissible states as follows.<br />

Definition 9. A state Ξ is an assignment <strong>of</strong> a nonempty subset Ξ(R) ⊂ X ∗ (R) to each torus R <strong>of</strong><br />

G so that the image <strong>of</strong> Ξ(R 2 ) in X ∗ (R 1 ) under the restriction map X ∗ (R 2 ) → X ∗ (R 1 ) is equal to<br />

Ξ(R 1 ) whenever R 1 ⊂ R 2 are tori <strong>of</strong> G.<br />

48


For each k-point g <strong>of</strong> G, we have maps g ! : X ∗ (g −1 Rg) → X ∗ (R) defined by (g ! χ)(r) = χ(g −1 rg)<br />

for each torus R <strong>of</strong> G. We define the conjugate state g ∗ Ξ by the formula (g ∗ Ξ)(R) = g ! Ξ(g −1 Rg)<br />

for each torus R <strong>of</strong> G. With this notion in mind, we say that a state Ξ is bounded if, for each torus<br />

R <strong>of</strong> G, ⋃ g∈G k<br />

g ∗ Ξ(R) is a finite set <strong>of</strong> characters <strong>of</strong> R.<br />

Finally, any state defines a function µ(Ξ) on X ∗ (G) by<br />

µ(Ξ, γ) =<br />

min 〈χ, γ〉 (3.12)<br />

χ∈Ξ(γ(G m))<br />

called the numerical function <strong>of</strong> Ξ.<br />

We say Ξ is admissible if its numerical function satisfies<br />

µ(Ξ, γ) = µ(Ξ, p • γ) for all p ∈ P (γ), where P (γ) is the parabolic subgroup <strong>of</strong> G associated to γ.<br />

We will refer to bounded admissible states as Kempf states.<br />

Remark 3. Suppose H is a closed subgroup <strong>of</strong> a group G and that Ξ is a Kempf state for G. Then<br />

res G HΞ, which assigns to any torus R <strong>of</strong> H the set <strong>of</strong> characters Ξ(R) (for R is also a torus <strong>of</strong> G),<br />

is a Kempf state for H, as all <strong>of</strong> the compatibility conditions are clearly inherited from G.<br />

With these terms defined, we return to the situation <strong>of</strong> an affine G-scheme X. For each closed<br />

G-subscheme Y <strong>of</strong> X, define Γ Y (X, x 0 ) to be the set {γ ∈ Γ(X, x 0 ) : lim t→0 γ(t) · x 0 exists in Y }.<br />

Theorem 11 ([27], Lemma 3.3). Let x 0 be a k-point <strong>of</strong> an affine G-variety X. Let Y be a closed<br />

G-subvariety <strong>of</strong> X not containing x 0 . Then there are Kempf states Ξ X,x0 and Υ Y X,x 0<br />

such that<br />

1. Γ(X, x 0 ) = {γ ∈ X ∗ (G) : µ(Ξ X,x0 , γ) ≥ 0},<br />

2. Γ Y (X, x 0 ) = {γ ∈ Γ(X, x 0 ) : µ(Υ Y X,x 0<br />

, γ) > 0}.<br />

Pro<strong>of</strong>. By Lemma 7, there is a G-representation V and an equivariant closed embedding X ↩→ V .<br />

Identify X with its image in V . Since ψ x0 : G → X is an isomorphism onto the open orbit Ω ⊂ X,<br />

we may ensure x 0 is not zero in V . As X is a closed G-subvariety <strong>of</strong> V , Γ(X, x 0 ) = Γ(V, x 0 ). Hence,<br />

without loss <strong>of</strong> generality, we may assume that X = V .<br />

We define the state Ξ V,x0<br />

<strong>of</strong> x 0 in the representation V as follows. Let R be a torus <strong>of</strong> G. Let<br />

V = ⊕ χ∈X ∗ (R) V χ be the eigendecomposition <strong>of</strong> V with respect to the torus R and let proj Vχ<br />

(x 0 )<br />

be the projection <strong>of</strong> x 0 on the weight space V χ . Set<br />

Ξ V,x0 (R) = {χ ∈ X ∗ (R) : proj Vχ<br />

(x 0 ) ≠ 0}, (3.13)<br />

49


As x 0 is not zero, Ξ V,x0 (R) is a nonempty subset <strong>of</strong> X ∗ (R) for each torus R <strong>of</strong> G. Furthermore,<br />

if R 1 ⊂ R 2 is an inclusion <strong>of</strong> tori <strong>of</strong> G, then the eigendecomposition <strong>of</strong> x 0 with respect to R 2 is<br />

also the eigendecomposition with respect to R 1 whose weights are the restrictions <strong>of</strong> characters<br />

χ ∈ Ξ V,x0 (R 2 ) to R 1 , so the restriction property is true. Hence Ξ V,x0<br />

is a state.<br />

Let x 0 = ∑ v i be the eigendecomposition <strong>of</strong> x 0 with respect to a one-parameter subgroup γ <strong>of</strong><br />

G. Then γ(t)·x 0 = ∑ t i v i , where each v i ≠ 0 and i runs through a finite nonempty set I <strong>of</strong> integers,<br />

since X ∗ (G m ) ∼ = Z. Suppose R is a torus such that γ ∈ X ∗ (R). Then I = {〈χ, γ〉 : χ ∈ Ξ V,x0 (R)}<br />

by the definition <strong>of</strong> a state.<br />

As µ(Ξ V,x0 , γ) = min I, lim t→0 γ(t) · x 0 exists in V if and only if<br />

µ(Ξ V,x0 , γ) ≥ 0. Thus Γ(X, x 0 ) = Γ(V, x 0 ) = {γ ∈ X ∗ (G) : µ(Ξ V,x0 , γ) ≥ 0}.<br />

It may be shown that Ξ V,x0 is a bounded, admissible state. First <strong>of</strong> all, for any torus R and any<br />

vector v ∈ V , the set Ξ V,v (R) is contained in the set <strong>of</strong> R-weights <strong>of</strong> V , which is finite. Furthermore,<br />

observe that ∑ g · v χ is the eigendecomposition <strong>of</strong> g · x 0 with respect to R whenever ∑ v χ is the<br />

eigendecomposition <strong>of</strong> x 0 with respect to T = g −1 Rg, so<br />

Ξ V,g·x0 (R) = g ! (Ξ V,x0 (g −1 Rg)) = g ∗ Ξ V,x0 (R) (3.14)<br />

for all g ∈ G k . Thus, the union ⋃ g∈G k<br />

g ∗ Ξ V,x0 (R) is contained in the finite set <strong>of</strong> R-weights <strong>of</strong><br />

V , and hence is finite. Therefore, Ξ V,x0<br />

is bounded. Now let γ be any one-parameter subgroup<br />

<strong>of</strong> G. Let p be a k-point <strong>of</strong> P (γ) = {g ∈ G : γ(t)gγ(t −1 ) ∈ G k[[t]] }, the parabolic subgroup <strong>of</strong> G<br />

associated to γ. As P (γ) = P (p·γ ·p −1 ), it will be enough to prove that µ(Ξ V,x0 , γ) ≤ µ(Ξ V,x0 , p·γ ·<br />

p −1 ). We may think <strong>of</strong> µ(Ξ V,x0 , γ) as the largest integer n such that lim t→0 t −n γ(t)x 0 exists in V .<br />

Then lim t→0 t −n pγ(t)p −1 x 0 = p −1<br />

0 = lim t→0 t −n p(γ(t)p −1 γ(t −1 ))γ(t)x 0 = p[lim t→0 γ(t)p −1 γ(t −1 )] ·<br />

[lim t→0 t −n γ(t)x 0 ], which exists in V by Lemma 8. Therefore, µ(Ξ V,x0 , p · γ · p −1 ) is at least n, so<br />

Ξ V,x0<br />

is admissible, as claimed.<br />

To define Υ Y X,x 0<br />

, we use part 2 <strong>of</strong> Lemma 7, which says that there is an equivariant morphism<br />

f : X → W , for some G-representation W , such that Y = f −1 (0). Set Υ Y X,x 0<br />

equal to the<br />

state <strong>of</strong> the point f(x 0 ) in the vector space W , defined by (3.13) above. Therefore, Υ Y X,x 0<br />

is a<br />

bounded admissible state. Moreover, γ ∈ Γ(X, x 0 ) implies that γ ∈ Γ(W, f(x 0 )), so lim t→0 γ(t)x 0<br />

is contained in Y if and only if γ ∈ Γ(X, x 0 ) and lim t→0 γ(t)f(x 0 ) = 0 in W . Let f(x 0 ) = ∑ w i<br />

be the eigendecomposition <strong>of</strong> f(x 0 ) with respect to the action <strong>of</strong> γ, where the w i ≠ 0 and the sum<br />

50


is over a finite set <strong>of</strong> integers J. Then γ(t)f(x 0 ) = ∑ t i w i , so lim t→0 γ(t)f(x 0 ) = 0 if and only if<br />

i > 0 for all i ∈ J. But µ(Υ Y X,x 0<br />

, γ) = min J, so Γ Y (X, x 0 ) = {γ ∈ Γ(X, x 0 ) : µ(Υ Y X,x 0<br />

, γ) > 0}.<br />

Given a Kempf state Ξ, define the set<br />

Γ Ξ = {γ ∈ X ∗ (G) : µ(Ξ, γ) ≥ 0}. (3.15)<br />

Example 11 (The Kempf state <strong>of</strong> M n as a GL n -embedding). Consider the GL n -embedding<br />

into M n with base point I n . Here Ξ Mn,In<br />

is the state which assigns<br />

(<br />

to the maximal torus D n <strong>of</strong><br />

r1 0<br />

diagonal matrices in GL n the set {χ 11 , χ 22 , . . . , χ nn }, where χ jj<br />

. .. = r j for j = 1, 2, . . . , n.<br />

0 r n<br />

)<br />

This corresponds to the eigendecomposition E 11 + E 22 + · · · + E nn <strong>of</strong> I n relative to D n . The state<br />

Ξ Mn,I n<br />

is GL n -stable under the conjugation action (because M n is a biequivariant GL n -embedding,<br />

see section 3.3.2 for more details), so Ξ Mn,I n<br />

(gD n g −1 ) = g ! Ξ Mn,I n<br />

(D n ) = {g ! χ 11 , . . . , g ! χ nn } for all<br />

g ∈ GL n . The closed GL n -subvariety ∂M n = V (det) corresponds to the state Υ ∂Mn<br />

M n,I n<br />

to each torus the set {det}. Clearly<br />

that assigns<br />

Γ(M n , I n ) = {γ ∈ X ∗ (GL n ) : lim<br />

t→0<br />

γ(t)I n exists in M n }<br />

= {γ ∈ X ∗ (GL n ) : min 〈χ, γ〉 ≥ 0}<br />

χ∈Ξ Mn,In (γ(G m))<br />

= {γ ∈ X ∗ (GL n ) : µ(Ξ Mn,I n<br />

, γ) ≥ 0},<br />

Γ ∂Mn (M n , I n ) = {γ ∈ Γ(M n , I n ) : lim<br />

t→0<br />

γ(t)I n exists in ∂M n }<br />

= {γ ∈ Γ(M n , I n ) : 〈det, γ〉 > 0}<br />

= {γ ∈ Γ(M n , I n ) : µ(Υ ∂Mn<br />

M n,I n<br />

, γ) > 0}.<br />

Therefore, given an affine G-embedding X and a choice <strong>of</strong> base point x 0 ∈ Ω, we obtain a Kempf<br />

state Ξ X,x0 such that Γ(X, x 0 ) = {γ ∈ X ∗ (G) : µ(Ξ X,x0 , γ) ≥ 0}. For each closed G-subvariety Y <strong>of</strong><br />

X, we obtain another Kempf state Υ Y X,x 0<br />

such that Γ Y (X, x 0 ) = {γ ∈ Γ(X, x 0 ) : µ(Υ Y X,x 0<br />

, γ) > 0}.<br />

However, the Kempf state for (X, x 0 ) depends on the choice <strong>of</strong> G-representation V in Lemma 7.<br />

Thus, for our purposes, rather than using Kempf states, we shall employ monoid states.<br />

Definition 10. A monoid state is a state Ψ such that Ψ(R) is a monoid for each R and for which<br />

51


there is a Kempf substate Ξ <strong>of</strong> Ψ such that Ψ(R) = { ∑ n χ χ : χ ∈ Ξ(R), n χ ∈ N 0 }.<br />

Recall that a monoid is a non-empty set M together with a multiplication operation such that<br />

M is associative and has an identity element with respect to this multiplication.<br />

Observe that if Ψ is a monoid state, then each monoid Ψ(R) is finitely generated since Ξ(R)<br />

must be a finite set as Ξ is a bounded state. Furthermore, if Ξ ′ is another Kempf substate that<br />

generates Ψ, then, for all γ ∈ X ∗ (G), µ(Ξ, γ) ≥ 0 if and only if µ(Ξ ′ , γ) ≥ 0. Suppose µ(Ξ, γ) ≥ 0.<br />

Each χ ′ ∈ Ξ ′ (γ(G m )) ⊂ Ψ(γ(G m )) may be written in the form χ ′ = ∑ χ∈Ξ(γ(G m)) n χ ′ ,χχ, where all<br />

<strong>of</strong> the coefficients n χ ′ ,χ ∈ N 0 . Therefore, if µ(Ξ, γ) ≥ 0, then each 〈χ, γ〉 ≥ 0 so that µ(Ξ ′ , γ) =<br />

min χ ′〈χ ′ ∑<br />

, γ〉 = min χ ′<br />

χ n χ ′ ,χ〈χ, γ〉 ≥ 0 as well. Likewise, every χ ∈ Ξ(γ(G m )) may be written in<br />

the form χ = ∑ χ ′ m χ,χ ′χ′ , so if µ(Ξ ′ , γ) ≥ 0, then µ(Ξ, γ) = min χ 〈χ, γ〉 = min χ<br />

∑χ ′ m χ,χ ′〈χ′ , γ〉 ≥<br />

0 also. Hence, a monoid state Ψ uniquely determines a subset Ψ ∨ = Γ Ξ ⊂ X ∗ (G), independent <strong>of</strong><br />

the choice <strong>of</strong> Kempf substate Ξ, thought <strong>of</strong> as the dual set <strong>of</strong> the monoid state.<br />

the set<br />

Conversely, suppose Γ is a subset <strong>of</strong> X ∗ (G). Consider Γ ∨ , which assigns to each torus R <strong>of</strong> G<br />

Γ ∨ (R) = {χ ∈ X ∗ (R) : 〈χ, γ〉 ≥ 0 for all γ ∈ Γ ∩ X ∗ (R)}. (3.16)<br />

First, 0 ∈ Γ ∨ (R) for all R implies that Γ ∨ (R) is non-empty. Suppose R ⊂ S are tori <strong>of</strong> G. Then<br />

it is clear that res S R (Γ∨ (S)) ⊆ Γ ∨ (R) by the definition <strong>of</strong> Γ ∨ above. However, this need not be an<br />

equality for every Γ, so Γ ∨ is not necessarily a state.<br />

Example 12 (A non-convex set <strong>of</strong> one-parameter subgroups). Consider the subset Γ =<br />

{γ (1,0) : t ↦→ (t, 1), γ (0,1) : t ↦→ (1, t)} ⊂ X ∗ (G 2 m). Then Γ ∨ (G 2 m) = {χ ij ∈ X ∗ (G 2 m) : 〈χ ij , γ (1,0) 〉 =<br />

52


i ≥ 0, 〈χ ij , γ (0,1) 〉 = j ≥ 0} = {χ ij ∈ X ∗ (G 2 m) : i, j ≥ 0}:<br />

• • • • •<br />

<br />

<br />

<br />

<br />

<br />

• • γ (0,1) <br />

• •<br />

<br />

<br />

<br />

<br />

X ∗ (G 2 <br />

m) : • • ◦<br />

γ (1,0) •<br />

• X ∗ (R) • • •<br />

• • • • •<br />

Now consider the subtorus R = diag(G 2 m) = {(r, r) : r ∈ G m }.<br />

Clearly Γ ∩ X ∗ (R) = ∅, so<br />

Γ ∨ (R) = {χ ∈ X ∗ (R) : 〈χ, γ〉 ≥ 0 for all γ ∈ Γ ∩ X ∗ (R) = ∅} = X ∗ (R). Yet the restriction <strong>of</strong><br />

Γ ∨ (G 2 m) to X ∗ (R) is {χ j : j ≥ 0} X ∗ (R), where χ j (r, r) = r j .<br />

If Γ ∨ is a state, then it is canonical, which is an advantage over the Kempf state Ξ X,x0 associated<br />

to an affine G-embedding (X, x 0 ) by Theorem 11. Moreover, it is obvious that each Γ ∨ (R) is a<br />

monoid, but this does not ensure that Γ ∨ is a monoid state, according to our definition, which<br />

requires the existence <strong>of</strong> a Kempf substate Ξ <strong>of</strong> Γ ∨ that generates each Γ ∨ (R).<br />

However, by<br />

Theorem 11, if Γ = Γ(X, x 0 ) for some affine G-variety, then Ξ X,x0<br />

is such an underlying Kempf<br />

substate. Hence Γ(X, x 0 ) ∨ is a monoid state. Moreover, Theorem 11 ensures that Γ(X, x 0 ) = {γ ∈<br />

X ∗ (G) : µ(Ξ X,x0 , γ) ≥ 0} = (Γ(X, x 0 ) ∨ ) ∨ . Thus, the combined requirements that Γ ∨ be a monoid<br />

state and that Γ = (Γ ∨ ) ∨ eliminates unwanted sets Γ, which leads to the following definition:<br />

Definition 11. We call a subset Γ ⊂ X ∗ (G) a lattice cone <strong>of</strong> one-parameter subgroups <strong>of</strong> G if Γ is<br />

saturated with respect to the equivalence relation (3.4) <strong>of</strong> one-parameter subgroups (i.e., if γ 1 ∼ γ 2<br />

and γ 1 ∈ Γ, then γ 2 ∈ Γ) and the quotient Γ(1)/ ∼ <strong>of</strong> the one-skeleton <strong>of</strong> Γ (3.11) is a finite set. A<br />

lattice cone Γ is called a convex lattice cone if Γ ∨ is a monoid state and Γ = (Γ ∨ ) ∨ . Additionally,<br />

Γ is a strongly convex lattice cone if it is a convex lattice cone and γ, γ −1 ∈ Γ if and only if γ = ε<br />

is the trivial one-parameter subgroup <strong>of</strong> G.<br />

Our terminology is compatible with that <strong>of</strong> toric geometry. If Γ is a strongly convex lattice<br />

53


cone as above, then each Γ ∩ X ∗ (T ) is one in the sense <strong>of</strong> toric geometry as defined in Section 2.1.1,<br />

Definition 3. For if Γ is a strongly convex lattice cone, Γ = {γ ∈ X ∗ (G) : µ(Ξ, γ) ≥ 0} for a Kempf<br />

state Ξ which generates Γ ∨ . Thus each Γ ∩ X ∗ (T ) is the intersection <strong>of</strong> finitely many half-spaces,<br />

Γ ∩ X ∗ (T ) = ⋂ χ∈Ξ(T ) {v ∈ X ∗(T ) : 〈χ, v〉 ≥ 0}, so it is a convex lattice cone in X ∗ (T ). The strong<br />

convexity follows as the same condition is required <strong>of</strong> Γ.<br />

Lemma 11. If X is an affine G-embedding with base point x 0 ∈ Ω, then the set Γ(X, x 0 ) = {γ ∈<br />

X ∗ (G) : lim t→0 γ(t)x 0 exists in X} is a strongly convex lattice cone.<br />

Pro<strong>of</strong>. Observe that if (X, x 0 ) is an affine G-embedding, then Γ(X, x 0 ) is a convex lattice cone by<br />

Proposition 8, Theorem 11 and our discussion above. Therefore, it is a strongly convex lattice cone,<br />

for we have shown that γ, γ −1 ∈ Γ(X, x 0 ) implies γ = ε and that ε ∈ Γ(X, x 0 ) in Proposition 7.<br />

Example 13 (The monoid state for Γ(G, e)). The monoid state corresponding to the group<br />

G as an affine G-embedding is equal to X ∗ , for Γ(G, e) = {ε} consists <strong>of</strong> only the trivial oneparameter<br />

subgroup and every character <strong>of</strong> every torus <strong>of</strong> G has non-negative value when paired<br />

with ε. A Kempf substate for X ∗ may be found by viewing G as a closed subgroup <strong>of</strong> some GL n ,<br />

which in turn is a closed subvariety <strong>of</strong> M n+1 defined by the equations det[(X ij ) n i,j=1 ]X n+1,n+1 = 1<br />

and X i,n+1 = X n+1,j = 0 for 1 ≤ i, j ≤ n. Notice that g ∈ GL n acts on M n+1 through matrix<br />

( {( ) }<br />

g 0<br />

g 0<br />

multiplication by<br />

0 det(g)<br />

), −1 under which GL n =<br />

0 det(g) −1 : g ∈ GL n is invariant. Then<br />

Ξ G,e = res GLn<br />

G<br />

ΞGLn M n+1 ,I n+1<br />

, which is a Kempf state by Remark 3, for Ξ Mn+1 ,I n+1<br />

is a Kempf state by<br />

the pro<strong>of</strong> <strong>of</strong> Theorem 11.<br />

The strong convexity condition implies that the dual monoid state Γ ∨ takes values Γ ∨ (R) that<br />

generate X ∗ (R) for all tori R. For if Γ is strongly convex, then (Γ ∩ X ∗ (R)) ∩ (−Γ ∩ X ∗ (R)) = {ε},<br />

so (Γ ∩ X ∗ (R)) ∨ + (−Γ ∩ X ∗ (R)) ∨ = [(Γ ∩ X ∗ (R)) ∩ (−Γ ∩ X ∗ (R))] ∨ = X ∗ (R). Therefore, Γ ∨ (R) =<br />

(Γ ∩ X ∗ (R)) ∨ generates X ∗ (R) as a group for all R.<br />

3.2.3 Classification theorem<br />

Let G be a connected reductive algebraic group defined over an algebraically closed field k <strong>of</strong><br />

characteristic zero. Let T be a maximal torus <strong>of</strong> G.<br />

54


Lemma 12. If γ 1 , γ 2 ∈ X ∗ (T ) ⊂ X ∗ (G), then<br />

(k[G] ∩ O vγ1 ) ∩ (k[G] ∩ O vγ2 ) ⊆ k[G] ∩ O vγ1 +γ 2<br />

, (3.17)<br />

where O vγ = {f ∈ k(G) : v γ (f) ≥ 0} is the valuation ring associated to v γ .<br />

Pro<strong>of</strong>. Recall that v γ is the valuation v t ◦ i γ <strong>of</strong> k(G), where i γ : k(G) → k(G)((t)) is the injection<br />

<strong>of</strong> fields filling the commutative diagram<br />

k[G]<br />

⊂<br />

k(G)<br />

µ ◦ <br />

k[G] ⊗ k[G]<br />

i γ<br />

id k[G] ⊗γ ◦ k[G] ⊗ k((t))<br />

⊂<br />

k(G)((t))<br />

and v t : k(G)((t)) × → Z is the standard valuation. Suppose γ 1 , γ 2 ∈ X ∗ (T ) ⊂ X ∗ (G) are oneparameter<br />

subgroups <strong>of</strong> G. Then γ 1 + γ 2 ∈ X ∗ (T ) is also a one-parameter subgroup <strong>of</strong> G contained<br />

in T . The valuations v γ1 , v γ2 and v γ1 +γ 2<br />

are obtained as above from the homomorphisms i γ1 , i γ2<br />

and i γ1 +γ 2<br />

, so it is enough to prove that i γ1 +γ 2<br />

(k[G] ∩ O vγ1 ∩ O vγ2 ) ⊂ k(G)[[t]] to prove the lemma.<br />

Yet<br />

k[G] ∩ O vγi<br />

k[G] ⊗ k[[t]]<br />

⊂<br />

k[G]<br />

µ ◦ k[G] ⊗ k[G]<br />

id k[G] ⊗γ ◦ i<br />

⊂<br />

k[G] ⊗ k((t))<br />

for i = 1, 2 and<br />

k[G]<br />

µ ◦ k[G] ⊗ k[G]<br />

id k[G] ⊗µ ◦ <br />

k[G] ⊗ k[G] ⊗ k[G]<br />

id k[G] ⊗(γ 1 +γ 2 ) ◦ <br />

id k[G] ⊗γ ◦ 1 ⊗γ◦ 2<br />

k[G] ⊗ k((t))<br />

id k[G] ⊗m 23<br />

k[G] ⊗ k((t)) ⊗ k((t))<br />

55


implies that<br />

i γ1 +γ 2<br />

(k[G] ∩ O vγ1 ∩ O vγ2 ) = (id k[G] ⊗ (γ 1 + γ 2 ) ◦ ) ◦ µ ◦ (k[G] ∩ O vγ1 ∩ O vγ2 )<br />

= (id k[G] ⊗ [m 23 ◦ (γ ◦ 1 ⊗ γ ◦ 2) ◦ µ ◦ ]) ◦ µ ◦ (k[G] ∩ O vγ1 ∩ O vγ2 )<br />

⊂ k[G] ⊗ k[[t]].<br />

Thus, (k[G] ∩ O vγ1 ) ∩ (k[G] ∩ O vγ2 ) ⊆ k[G] ∩ O vγ1 +γ 2<br />

as claimed.<br />

Theorem 12. Affine G-embeddings are classified, up to conjugation, by strongly convex lattice<br />

cones <strong>of</strong> one-parameter subgroups in X ∗ (G). Conjugation corresponds to the change <strong>of</strong> base point<br />

in the embedding.<br />

Pro<strong>of</strong>. Assume X is an affine G-embedding.<br />

The selection <strong>of</strong> a base point x 0 in X uniquely<br />

determines a set Γ(X, x 0 ) ⊂ X ∗ (G), which is a strongly convex lattice cone by Lemma 11. By<br />

Theorem 10, the set Γ(X, x 0 ) determines the affine G-embedding X and the selected base point x 0<br />

via an isomorphism ψ x0 : k[X] ∼ = A Γ(X,x0 ) such that m x0 = ψ −1<br />

x 0<br />

(m e ∩ A Γ(X,x0 )). By Corollary 3<br />

and formula (3.7), the selection <strong>of</strong> a different base point h · x 0 is equivalent to conjugating the<br />

cone Γ(X, h · x 0 ) = hΓ(X, x 0 )h −1 and a corresponding right translation <strong>of</strong> the algebra A Γ(X,h·x0 ) =<br />

r h (A Γ(X,x0 )).<br />

Thus each affine G-embedding determines a strongly convex lattice cone, modulo<br />

conjugation in X ∗ (G), which in turn recovers the variety up to isomorphism.<br />

Conversely, we prove that every strongly convex lattice cone Γ ⊂ X ∗ (G) determines an affine<br />

G-embedding X Γ such that Γ(X Γ , x 0 ) = Γ for some choice <strong>of</strong> base point x 0 ∈ X Γ . Assume Γ is a<br />

strongly convex lattice cone, so Γ ∨ is a monoid state generated by a Kempf substate Ξ. Define A Γ<br />

as in (3.9) to be the subalgebra <strong>of</strong> k[G] given by<br />

A Γ = {f ∈ k[G] : v γ (f) ≥ 0, for all γ ∈ Γ} = k[G] ∩ ⋂ γ∈Γ<br />

O vγ ,<br />

where O v is the valuation subring {f ∈ k(G) : v(f) ≥ 0} <strong>of</strong> k(G). Using this second description,<br />

since k[G] and all <strong>of</strong> the valuation rings O vγ<br />

are integrally closed in k(G), their intersection A Γ is<br />

an integrally closed domain. Furthermore, A Γ is left-invariant, since k[G] is and v γ (s · f) = v γ (f)<br />

for all f ∈ k(G) and s ∈ G implies that the valuation rings O vγ<br />

are G-stable as well. Hence A Γ is<br />

56


an integrally closed left-invariant subalgebra <strong>of</strong> k[G]. It only remains to prove that A Γ is finitely<br />

generated over k and that the variety X Γ = Spec A Γ contains G as an open orbit.<br />

Let γ 1 , . . . , γ m be a set <strong>of</strong> representatives for the equivalence classes in Γ(1)/ ∼, which is a finite<br />

set as Γ is a lattice cone. By Proposition 8 and Lemma 12,<br />

A Γ = k[G] ∩ ⋂ γ∈Γ<br />

O vγ = ⋂ γ∈Γ(k[G] ∩ O vγ ) =<br />

m⋂<br />

(k[G] ∩ O vγi )<br />

since any γ ∈ Γ belongs to some X ∗ (T ), in which case it is a sum <strong>of</strong> elements from (Γ ∩ X ∗ (T ))(1).<br />

Let f 1 , . . . , f n be a finite set <strong>of</strong> generators for the algebra k[G] over k, so k[G] = k[f 1 , . . . , f n ]. By<br />

Gordan’s Lemma, the set<br />

C = {(c l ) ∈ Z n : c 1 , . . . , c n ≥ 0,<br />

i=1<br />

n∑<br />

c j v γi (f j ) ≥ 0 for all i = 1, . . . , m}<br />

j=1<br />

is finitely generated. Thus the “monomials” f c 1<br />

1 f c 2<br />

2 · · · f n<br />

cn<br />

in the generators f 1 , . . . , f n <strong>of</strong> k[G]<br />

associated to the generators (c 1 , . . . , c n ) <strong>of</strong> C admit a finite set <strong>of</strong> generators <strong>of</strong> A Γ , as any element<br />

<strong>of</strong> A Γ is a sum <strong>of</strong> such monomials.<br />

Therefore X Γ = Spec A Γ is a normal affine G-variety with an open G-orbit, since A Γ ⊂ k[G]<br />

implies f : G → X Γ is dominant, so f(G) is an open orbit in X Γ ([40], Theorem 1.9.5). It only<br />

remains to show that f(G) is isomorphic to G as an orbit in X Γ . Consider, for each torus R <strong>of</strong><br />

G, the image <strong>of</strong> A Γ under the homomorphism k[G] → k[R]. By construction <strong>of</strong> A Γ = {f ∈ k[G] :<br />

v γ (f) ≥ 0 for all γ ∈ Γ} and the decomposition A Γ = ⊕ χ∈Γ ∨ (R) AR χ , where A R χ = {f ∈ A : f(xr) =<br />

χ(r)f(x) for all r ∈ R}, it is evident that the image <strong>of</strong> A Γ in k[R] is the k-monoid algebra k[Γ ∨ (R)].<br />

Hence we have the following commutative diagrams:<br />

⊂<br />

A Γ<br />

k[G]<br />

X Γ<br />

f<br />

G<br />

onto<br />

<br />

k[Γ ∨ (R)]<br />

⊂<br />

onto<br />

k[R] = k[X ∗ (R)]<br />

closed<br />

R Γ ∨ (R)<br />

open<br />

R<br />

closed<br />

Since Γ is strongly convex, Γ ∨ (R) is not contained in any hyperplane, so the monoid Γ ∨ (R) generates<br />

X ∗ (R) as a group. This implies that R is isomorphic to an open subset <strong>of</strong> the closed subvariety<br />

57


R Γ ∨ (R) = Spec k[Γ ∨ (R)] in X Γ , so that f(R) ∼ = R for every torus R <strong>of</strong> G. Thus f( ⋃ T T ) =<br />

⋃<br />

T f(T ) ∼ = ⋃ T T ⊂ X Γ. Yet ⋃ T<br />

T contains a dense open subset O <strong>of</strong> G ([40], Theorem 6.4.5 and<br />

Corollary 7.6.4) and f| ⋃ T T is an isomorphism, so O ∼ = f(O) is open in X Γ . Thus f : G → X Γ is a<br />

birational morphism [22]. As f(G) is an open orbit in X Γ containing an open subset birationally<br />

equivalent to G, f(G) is isomorphic to G and f : G → X Γ is an affine G-embedding.<br />

Finally, consider Γ(X Γ , f(e)) = {γ ∈ X ∗ (G) : lim t→0 γ(t)f(e) exists in X Γ } = {γ ∈ X ∗ (G) :<br />

γ ◦ (A Γ ) ⊂ k[t]}, that is, the dual map γ ◦ : A Γ ⊂ k[G] → k[t, t −1 ] factors through k[t]. Thus, if<br />

γ ∈ Γ, so v γ (f) ≥ 0 for all f ∈ A Γ , then γ ◦ (A Γ ) ⊂ k[t], and hence Γ ⊂ Γ(X Γ , f(e)). Now let<br />

γ ∈ Γ(X Γ , f(e)). Suppose γ is a one-parameter subgroup <strong>of</strong> a torus T <strong>of</strong> G. Then lim t→0 γ(t)f(e)<br />

exists in the toric variety T ⊂ X Γ , so the classification <strong>of</strong> toric varieties (Theorem 3) implies that<br />

γ ∈ Γ ∩ X ∗ (T ), and hence that Γ(X Γ , f(e)) ⊂ Γ.<br />

Therefore, X Γ = Spec A Γ is a normal affine<br />

G-embedding such that Γ(X Γ , f(e)) = Γ, which completes our pro<strong>of</strong>.<br />

3.3 Functoriality <strong>of</strong> our classification<br />

In this section, we discuss how the classification <strong>of</strong> Theorem 12 also describes equivariant morphisms<br />

between affine G-embeddings. In particular, we show that equivariant maps between affine G-<br />

embeddings correspond to inclusions <strong>of</strong> the associated strongly convex lattice cones and conversely<br />

in Section 3.3.1. After that, we characterize biequivariant affine G-embeddings in terms <strong>of</strong> their<br />

cones and, using this result, construct the “biequivariant resolution” <strong>of</strong> an affine G-embedding in<br />

Section 3.3.2. We remark here that Proposition 10 below may be seen as an affine version <strong>of</strong> Brion’s<br />

classification [5] <strong>of</strong> regular G-compactifications presented in Section 2.2.1.<br />

3.3.1 <strong>Equivariant</strong> morphisms between affine G-embeddings<br />

Suppose f : X → Y is an equivariant morphism between affine G-embeddings X and Y .<br />

By<br />

equivariance, if x 0 ∈ X is a base point for X, then y 0 = f(x 0 ) is a base point for Y . Moreover, if<br />

γ is a one-parameter subgroup <strong>of</strong> G such that lim t→0 γ(t)x 0 = x γ exists in X, then lim t→0 γ(t)y 0<br />

exists in Y and is equal to f(x γ ) since f is continuous and f(γ(t)x 0 ) = γ(t)f(x 0 ) = γ(t)y 0 for all<br />

t ≠ 0. Therefore, there is an inclusion Γ(X, x 0 ) ⊂ Γ(Y, f(x 0 )) whenever there exists an equivariant<br />

morphism f : X → Y <strong>of</strong> affine G-embeddings. Moreover, f is the morphism dual to the inclusion<br />

58


<strong>of</strong> the subalgebras A Γ(X,x0 ) ⊂ A Γ(Y,f(x0 )) in k[G], because f(g · x 0 ) = g · f(x 0 ) and Gx 0 = Ω X is<br />

open in X implies that f is uniquely determined by its value at x 0 .<br />

Remark 4. Suppose X and Y are G-embeddings, not necessarily affine.<br />

If f : X → Y is an<br />

equivariant morphism and if x 0 is a base point for X (i.e., the orbit Gx 0 is isomorphic to G as<br />

G-varieties), then f(x 0 ) will be a base point for Y and we can define the sets Γ(X, x 0 ) = {γ ∈<br />

X ∗ (G) : lim t→0 γ(t)x 0 exists in X} and Γ(Y, f(x 0 )) = {δ ∈ X ∗ (G) : lim t→0 δ(t)f(x 0 ) exists in Y }<br />

as usual, even if X and Y are not affine. By the same argument as above, the existence <strong>of</strong> the<br />

equivariant morphism f : X → Y implies that Γ(X, x 0 ) ⊂ Γ(Y, f(x 0 )).<br />

Conversely, suppose Γ(X, x 0 ) ⊂ Γ(Y, y 0 ) for two affine G-embeddings X and Y .<br />

By Theorem<br />

10, X = Spec A Γ(X,x0 ) and Y = Spec A Γ(Y,y0 ). The definition <strong>of</strong> A Γ = {f ∈ k[G] : v γ (f) ≥<br />

0 for all γ ∈ Γ} implies that A Γ(Y,y0 ) ⊆ A Γ(X,x0 ), so there is a corresponding equivariant morphism<br />

<strong>of</strong> G-embeddings X → Y sending x 0 ↦→ y 0 . However, by Corollary 3, the subalgebra <strong>of</strong> k[G] isomorphic<br />

to k[X Γ ] is only determined up to right translations, which correspond to conjugates <strong>of</strong><br />

the cone Γ. Thus,<br />

Proposition 9. Suppose X 1 , X 2 are affine G-embeddings and Γ 1 , Γ 2 are strongly convex lattice<br />

cones.<br />

1. If x 1 ∈ X 1 is a base point and f : X 1 → X 2 is an equivariant morphism <strong>of</strong> G-embeddings,<br />

then Γ(X 1 , x 1 ) ⊂ Γ(X 2 , f(x 1 )) and f is the morphism recovered from the inclusion:<br />

k[X 1 ]<br />

f ◦ <br />

k[X 2 ]<br />

∼= A<br />

ψx ◦ Γ(X1 ,x 1 )<br />

1<br />

⊂<br />

∼=<br />

A<br />

ψf(x ◦ Γ(X2 ,x 2 )<br />

1 )<br />

2. If there is an element h ∈ G such that Γ 1 ⊂ hΓ 2 h −1 , then there is an equivariant morphism<br />

X Γ1 → X Γ2 sending the base point x 1 ∈ X Γ1 to h · x 2 ∈ X Γ2 , where m xi = A Γi ∩ m e for<br />

i = 1, 2.<br />

Pro<strong>of</strong>. We have already proven part 1 <strong>of</strong> this Lemma in the discussion prior to Remark 4.<br />

59


To prove par 2, by Theorem 12, we know that Γ 1 , Γ 2 correspond to affine G-embeddings X 1 =<br />

Spec A Γ1<br />

and X 2 = Spec A Γ2 , respectively, such that Γ i = Γ(X i , x i ) for i = 1, 2, where x i is the<br />

base point <strong>of</strong> X i corresponding to the maximal ideal m e ∩ A Γi . Now suppose there is an element<br />

h ∈ G such that Γ 1 ⊂ hΓ 2 h −1 as in the statement <strong>of</strong> the lemma. We know that hΓ 2 h −1 =<br />

hΓ(X 2 , x 2 )h −1 = Γ(X, h · x 2 ) by formula (3.7). Hence we have Γ(X 1 , x 1 ) ⊂ Γ(X 2 , h · x 2 ), so<br />

that A Γ(X1 ,x 1 ) ⊃ A Γ(X2 ,h·x 2 ) and thus X 1 → X 2 exists, is equivariant, and maps x 1 ↦→ h · x 2 as<br />

claimed.<br />

Example 14 (The subcone <strong>of</strong> Γ(X, x 0 ) associated to a torus closure). If X is an affine<br />

G-embedding and T is a maximal torus <strong>of</strong> G whose closure in X is denoted T , then G × T T is an<br />

affine G-embedding and we have an equivariant morphism G × T T → X. Let Γ = Γ(X, x 0 ) and<br />

σ = Γ(X, x 0 ) ∩ X ∗ (T ). Then Γ(G × T T , [e, x 0 ]) = ⋃ P ⊃T P • (σ ∩ ∆ P (T )), where the union is taken<br />

over all parabolic subgroups <strong>of</strong> G containing T and ∆ P (T ) = {γ ∈ X ∗ (T ) : P (γ) ⊇ P }. There<br />

are only finitely many parabolic subgroups P containing T , as there are only a finite number <strong>of</strong><br />

Borel subgroups containing T (in one-to-one correspondence with the elements <strong>of</strong> the Weyl group<br />

W (T, G), [[40], Corollary 6.4.12]), and, for each Borel subgroup B, there are only finitely many<br />

parabolic subgroups P ⊃ B (indexed by subsets <strong>of</strong> the basis ∆ positive roots <strong>of</strong> T relative to B, [[40],<br />

Theorem 8.4.3]). In contrast, Γ = ⋃ Q⊂G Q•(Γ∩∆ Q(G)), where the union is taken over the set <strong>of</strong> all<br />

parabolic subgroups Q <strong>of</strong> G (there are infinitely many, as there are infinitely many Borel subgroups<br />

each corresponding to cosets in G/B, which is projective) and ∆ Q (G) = {γ ∈ X ∗ (G) : P (γ) ⊇ Q}.<br />

Clearly, for each P ⊃ T , σ ∩ ∆ P (T ) = (Γ ∩ X ∗ (T )) ∩ ∆ P (T ) = Γ ∩ ∆ P (T ) = Γ ∩ ∆ P (G) since<br />

∆ P (G) ⊂ X ∗ (T ′ ) for all maximal tori T ′ contained in P . Therefore, Γ(G × T T , [e, x 0 ]) is a finite<br />

union <strong>of</strong> the “parabolic components” <strong>of</strong> Γ(X, x 0 ), so Γ(G × T T , [e, x 0 ]) is a finite polysimplicial<br />

subcomplex <strong>of</strong> Γ(X, x 0 ), as the ∆ Q (G) provide a triangulation <strong>of</strong> X ∗ (G) [32].<br />

3.3.2 Biequivariant resolutions<br />

In this section, we show that every affine G-embedding X canonically determines a (G × G)-<br />

equivariant affine G-embedding X G together with a left-G-equivariant morphism X G → X, which<br />

we call the biequivariant resolution <strong>of</strong> X. As we will be working with varieties some <strong>of</strong> which only<br />

have a left action and others both a left and a right action, we will be careful to specify how G acts<br />

60


on varieties discussed in this section.<br />

Our first tool is the following lemma, which allows us to detect when an affine G-embedding<br />

also has a right-G-action X × G → X extending the multiplication in G.<br />

Proposition 10. An affine G-embedding X will have both a left and a right G-action, and thus be<br />

a biequivariant G-embedding, if and only if the associated strongly convex lattice cone Γ(X, x 0 ), for<br />

any choice <strong>of</strong> base point x 0 ∈ X, is G-stable for the conjugation action <strong>of</strong> G on X ∗ (G).<br />

Pro<strong>of</strong>. Suppose that X is a (G × G)-equivariant affine G-embedding and let x ∈ X be a base<br />

point. Then X is determined by the strongly convex lattice cone Γ(X, x) = {γ ∈ X ∗ (G) :<br />

lim t→0 γ(t)x exists in X}. Let h ∈ G and recall that hΓ(X, x)h −1 = Γ(X, h · x) by formula (3.7).<br />

Thus it is enough to show that γ ∈ Γ if and only if γ ∈ Γ(X, h·x). Assume lim t→0 γ(t)x exists in X.<br />

Then consider lim t→0 [γ(t) · hx] = lim t→0 [γ(t) · xh ′ ] = [lim t→0 γ(t) · x] · h ′ , for some h ′ ∈ G, and this<br />

limit exists in X by Remark 2, since there is a right G-action on X. Thus Γ(X, x) ⊂ hΓ(X, x)h −1 .<br />

Now assume that γ ′ ∈ Γ(X, h · x). Then, the same argument implies γ ′ ∈ Γ(X, x) using Remark 2.<br />

Therefore, for every h ∈ G, Γ(X, x) = hΓ(X, x)h −1 . Thus Γ(X, x) is G-stable for the conjugation<br />

action <strong>of</strong> G on X ∗ (G) for any choice <strong>of</strong> base point x ∈ X.<br />

Now assume that Γ is a strongly convex lattice cone which is G-stable for the conjugation action<br />

<strong>of</strong> G on X ∗ (G). Theorem 12 implies that Γ = Γ(X Γ , x 0 ) for some base point x 0 ∈ X Γ = Spec A Γ .<br />

Then, by Corollary 3, for any h ∈ G we have r h (A Γ(XΓ ,x 0 )) = A Γ(XΓ ,h·x 0 ) = A hΓ(XΓ ,x 0 )h −1 =<br />

A Γ . Hence A Γ is a left- and right-G-invariant subalgebra <strong>of</strong> k[G], so there is a right G-action on<br />

X Γ = Spec A Γ , which extends the multiplication <strong>of</strong> G by [40], Proposition 2.3.6. Thus X Γ is a<br />

(G × G)-equivariant affine G-embedding.<br />

Corollary 4. If X is a biequivariant affine G-embedding and T is any maximal torus <strong>of</strong> G, then<br />

the closure <strong>of</strong> T in X determines X completely.<br />

Pro<strong>of</strong>. Let X be a biequivariant affine G-embedding with lattice cone Γ = Γ(X, x 0 ) for some choice<br />

<strong>of</strong> base point in X. Let T be any maximal torus <strong>of</strong> G. Consider T ⊂ X, which is an affine toric<br />

variety for T . Let σ ⊂ X ∗ (T ) be the cone <strong>of</strong> one-parameter subgroups <strong>of</strong> T with specializations<br />

in T . Then σ = Γ ∩ X ∗ (T ), by definition <strong>of</strong> Γ. If T ′ is any other maximal torus <strong>of</strong> G, then<br />

T ′ = gT g −1 for some g ∈ G and Γ∩X ∗ (T ′ ) = g[g −1 (Γ∩X ∗ (T ′ ))g]g −1 = g[g −1 Γg∩g −1 X ∗ (T ′ )g]g −1 =<br />

61


g[Γ ∩ X ∗ (T )]g −1 = gσg −1 . Thus Γ = ⋃ g∈G gσg−1 is determined by the cone σ, so X is determined<br />

by its toric subvariety T = T σ .<br />

We note the similarity between our classification in the case <strong>of</strong> biequivariant affine G-embeddings<br />

and Brion’s classification <strong>of</strong> regular G-compactifications in Section 2.2.1. In [5], Brion classified<br />

regular compactifications <strong>of</strong> a group G by the W (T, G)-invariant fan associated to the closure <strong>of</strong><br />

any maximal torus T <strong>of</strong> G in the compactification, demonstrating that regular compactifications<br />

<strong>of</strong> G are completely determined by any <strong>of</strong> the associated toric subvarieties. Likewise, Corollary 4<br />

implies that if an affine G-embedding X is (G × G)-equivariant and T is a maximal torus <strong>of</strong> G,<br />

then the cone for T recovers X. Therefore, we may think <strong>of</strong> Proposition 10 and Corollary 4 as an<br />

affine version <strong>of</strong> Brion’s classification.<br />

Remark 5. Suppose X is a (G × G)-equivariant affine G-embedding and suppose that x 0 ∈ X is<br />

a base point. By Proposition 10 above, the set Γ(X, x 0 ) is stable under the conjugation action <strong>of</strong><br />

G. However, by formula (3.7), we conclude that Γ(X, x 0 ) = hΓ(X, x 0 )h −1 = Γ(X, h · x 0 ) for all<br />

h ∈ G. Therefore, the strongly convex lattice cone associated to a biequivariant affine G-embedding<br />

is independent <strong>of</strong> the choice <strong>of</strong> base point.<br />

Hence, let X be a biequivariant affine G-embedding. By formula (3.14), if Ξ is a Kempf substate<br />

generating Γ(X, x) ∨ , then g ∗ Ξ is a Kempf substate that generates Γ(X, g · x) ∨ . Hence, if Γ is a G-<br />

stable strongly convex lattice cone whose monoid state is generated by the Kempf substate Ξ, then<br />

the Kempf state G ∗ Ξ : R ↦→ ⋃ g∈G g ∗Ξ(R) generates Γ ∨ as well and is G-invariant for conjugation.<br />

We use these observations to construct the biequivariant resolution <strong>of</strong> an affine G-embedding<br />

X. Suppose X is an affine G-embedding, x ∈ X is a base point, and Γ(X, x) is the corresponding<br />

strongly convex lattice cone. Then there is a unique maximal G-stable subset<br />

Γ G := ⋂<br />

hΓh −1 (3.18)<br />

<strong>of</strong> Γ. In fact,<br />

h∈G<br />

Lemma 13. If Γ is a strongly convex lattice cone in X ∗ (G), then Γ G = ⋂ h∈G hΓh−1 is a strongly<br />

convex lattice cone.<br />

62


Pro<strong>of</strong>. Let Γ be a strongly convex lattice cone.<br />

Thus Γ ∨ is a monoid state, Γ = (Γ ∨ ) ∨ , and<br />

γ, γ −1 ∈ Γ if and only if γ is trivial. As Γ ∨ is a monoid state, let Ξ be a Kempf substate generating<br />

Γ ∨ . Define Γ G = ⋂ h∈G hΓh−1 as in (3.18). We claim that G ∗ Ξ : R ↦→ ⋃ g∈G g ∗Ξ(R) is a Kempf<br />

state and that it generates (Γ G ) ∨ , in which case Γ G is a lattice cone.<br />

First, G ∗ Ξ is a state, for each g ∗ Ξ is a state implies that G ∗ Ξ(R) is nonempty for all R and<br />

that if S ⊃ R, then G ∗ Ξ(R) = ⋃ g∈G g ∗Ξ(R) = ⋃ [ ⋃ ]<br />

g∈G resS R [g ∗Ξ(S)] = res S R g∈G g ∗Ξ(S) =<br />

res S R [G ∗Ξ(S)]. Next, G ∗ Ξ is bounded, for if h ∈ G, then h ∗ [G ∗ Ξ](R) = ⋃ g∈G h ∗[g ∗ Ξ(R)] =<br />

⋃<br />

g∈G (hg) ∗Ξ(R) = G ∗ Ξ(R), so ⋃ h∈G h ∗[G ∗ Ξ](R) = G ∗ Ξ(R) = ⋃ g∈G g ∗Ξ(R) is a finite subset<br />

<strong>of</strong> X ∗ (R) for each R, since Ξ is a bounded state.<br />

function µ(G ∗ Ξ) is defined by<br />

Finally, G ∗ Ξ is admissible, for its numerical<br />

µ(G ∗ Ξ, γ) =<br />

min 〈χ, γ〉 = min<br />

χ∈G ∗Ξ(γ)<br />

χ∈ ⋃ g ∗Ξ(γ)<br />

〈χ, γ〉 = min<br />

min<br />

g∈G χ∈g ∗Ξ(γ)<br />

〈χ, γ〉 = min µ(g ∗Ξ, γ)<br />

g∈G<br />

and each g ∗ Ξ is admissible.<br />

That is, if γ is a one-parameter subgroup <strong>of</strong> G and p ∈ P (γ) belongs<br />

to the parabolic subgroup associated to γ, then µ(G ∗ Ξ, p • γ) = min g∈G µ(g ∗ Ξ, p • γ) =<br />

min g∈G µ(g ∗ Ξ, γ) = µ(G ∗ Ξ, γ). Thus G ∗ Ξ is a Kempf state.<br />

Now we show that each G ∗ Ξ(R) generates (Γ G ) ∨ (R), for all tori R <strong>of</strong> G. Suppose χ ∈ (Γ G ) ∨ (R)<br />

for some torus R <strong>of</strong> G. Now (Γ G ) ∨ (R) = ( ⋂ g∈G gΓg−1 ) ∨ (R) = ∑ g∈G (gΓg−1 ) ∨ (R), so χ is a sum <strong>of</strong><br />

elements from the monoids (gΓg −1 ) ∨ (R). As the sets g ∗ Ξ(R) generate the monoids (gΓg −1 ) ∨ (R),<br />

we may write χ = ∑ g∈G (∑ ξ∈g n ∗Ξ(R) ξξ) for non-negative integers n ξ , which is a finite sum since<br />

∑ ∑<br />

g∈G ξ∈g = ∑ ∗Ξ(R) ξ∈G and G ∗Ξ(R) ∗Ξ(R) is finite. Thus G ∗ Ξ(R) generates (Γ G ) ∨ (R) as a monoid<br />

for each R. Therefore, to prove that (Γ G ) ∨ is a monoid state, it suffices now to show that (Γ G ) ∨ is a<br />

state, for we have shown that G ∗ Ξ will be a suitable Kempf substate. Clearly each (Γ G ) ∨ (R) is nonempty,<br />

for it will contain at least the zero character. Assume R ⊂ S is an inclusion <strong>of</strong> tori <strong>of</strong> G. Then<br />

res S R [(ΓG ) ∨ (S)] ⊂ (Γ G ) ∨ (R) by definition <strong>of</strong> (Γ G ) ∨ as mentioned after formula (3.16). Now suppose<br />

that χ ∈ (Γ G ) ∨ (R). We may write χ = ∑ ξ∈G ∗Ξ(R) n ξξ for non-negative integers n ξ since (Γ G ) ∨ (R)<br />

is generated as a monoid by G ∗ Ξ(R). Yet G ∗ Ξ is a state, so each ξ ∈ G ∗ Ξ(R) is the restriction<br />

<strong>of</strong> some not-necessarily-unique ξ S ∈ G ∗ Ξ(S). Let χ S = ∑ ξ∈G ∗Ξ(R) n ξξ S ∈ (Γ G ) ∨ (S) and consider<br />

res S R (χS ) = ∑ ξ∈G ∗Ξ(R) n ξres S R (ξS ) = ∑ ξ∈G ∗Ξ(R) n ξξ = χ.<br />

Hence res S R [(ΓG ) ∨ (S)] = (Γ G ) ∨ (R),<br />

which proves that (Γ G ) ∨ is a state. Therefore, (Γ G ) ∨ is a monoid state and Γ G is a lattice cone in<br />

63


X ∗ (G).<br />

Furthermore, [(Γ G ) ∨ ] ∨ = {γ ∈ X ∗ (G) : µ(G ∗ Ξ, γ) ≥ 0} = {γ ∈ X ∗ (G) : µ(g ∗ Ξ, γ) ≥ 0 for all g ∈<br />

G} = ⋂ g∈G {γ ∈ X ∗(G) : µ(g ∗ Ξ, γ) ≥ 0} = ⋂ g∈G gΓg−1 = Γ G . Thus, Γ G is a convex lattice cone.<br />

Finally, as Γ G ⊂ Γ, it is clear that γ, γ −1 ∈ Γ G implies that γ = ε is the trivial one-parameter<br />

subgroup <strong>of</strong> G. Furthermore, ε ∈ hΓh −1 for all h ∈ G, so ε ∈ Γ G . Thus γ, γ −1 ∈ Γ G if and only if<br />

γ = ε. Therefore, Γ G is a strongly convex lattice cone in X ∗ (G).<br />

Definition 12. Let X be an affine G-embedding and select a base point x 0 ∈ X. Let Γ = Γ(X, x 0 )<br />

and define Γ G = ⋂ h∈G hΓh−1 , which is a strongly convex lattice cone.<br />

Then X G := Spec A Γ G<br />

is a (G × G)-equivariant affine G-embedding by Theorem 12 and Proposition 10, and there is a<br />

left-G-equivariant morphism β (X,x0 ) : X G → X corresponding to the inclusion A Γ(X,x0 ) ⊂ A Γ G<br />

<strong>of</strong><br />

subalgebras in k[G]. We call X G together with the left-G-equivariant morphism β (X,x0 ) : X G → X<br />

the biequivariant resolution <strong>of</strong> X.<br />

We make a few remarks about the biequivariant resolution <strong>of</strong> an affine G-embedding before we<br />

prove its universal property. First, if X is already a (G × G)-equivariant affine G-embedding, then<br />

Γ = Γ G , so X G = X. However, it is possible in some cases that X G = G even when X ≠ G.<br />

Example 15 (A trivial biequivariant resolution). Let D 2 denote the maximal torus <strong>of</strong> diagonal<br />

matrices in GL 2 . Let σ be the strongly convex lattice cone generated by γ 1,0 : t ↦→<br />

( t 0<br />

0 1 ) and let T σ be the corresponding toric variety for D 2 . Observe that T σ = Spec k[σ ∨ ] =<br />

Spec k[χ (1,0) , χ (0,1) , χ (0,−1) ] ∼ = A 1 × G m , whose base point is (1, 1) ∈ A 1 × G m . Consider the affine<br />

GL 2 -embedding X = GL 2 × D 2<br />

T σ . Then Γ(X, [I 2 , (1, 1)]) = ⋃ P ⊃D 2<br />

P • (σ ∩ ∆ P (D 2 )). There are<br />

only three parabolic subgroups <strong>of</strong> GL 2 containing D 2 : B + = {( )}<br />

a b<br />

0 d , B − = {( )}<br />

a 0<br />

c d , and GL2 .<br />

As we have shown in Example 7, ∆ B +(D 2 ) = {γ m,n ∈ X ∗ (D 2 ) : m ≥ n}, ∆ B −(D 2 ) = {γ m,n ∈<br />

X ∗ (D 2 ) : m ≤ n}, and ∆ GL2 (D 2 ) = {γ m,n ∈ X ∗ (D 2 ) : m = n}, where γ m,n is the one-parameter<br />

subgroup <strong>of</strong> D 2 , t ↦→ ( )<br />

t m 0<br />

0 t . n Therefore, Γ(X, [I2 , (1, 1)]) = B + • σ and Γ(X, [I 2 , (1, 1)]) GL 2<br />

=<br />

⋂<br />

g∈GL 2<br />

g(B + • σ)g −1 ⊆ B + • σ ∩ s(B + • σ)s −1 , where s = ( 0 1 1<br />

0 ) ∈ GL 2. Now a typical element <strong>of</strong><br />

B + • σ has the form ( )<br />

1 b<br />

0 1 (<br />

t n 0<br />

0 1 ) ( ) ( )<br />

1 −b<br />

0 1 =<br />

t n b(1−t n )<br />

for a non-negative integer n and an element<br />

0 1<br />

( ) ( )<br />

b ∈ k. An element <strong>of</strong> s(B + • σ)s −1 is <strong>of</strong> the form ( 0 1 1<br />

0 ) ( 0 1 1<br />

0 ) = 1 0<br />

b(1−t m ) t m for a<br />

t m b(1−t m )<br />

0 1<br />

non-negative integer m and an element b ∈ k. Thus B + • σ ∩ s(B + • σ)s −1 = {ε}, which implies<br />

64


that Γ(X, [I 2 , (1, 1)]) GL 2<br />

= {ε} and that the biequivariant resolution <strong>of</strong> X is β (X,[I2 ,(1,1)]) : G → X<br />

given by g ↦→ [g, (1, 1)].<br />

Second, while the cone Γ G is canonical, the morphism β (X,x0 ) : X G → X is only defined<br />

up to right translation, which corresponds to a different choice <strong>of</strong> base point x ∈ X as follows:<br />

r h (β (X,x0 )) = β (X,h·x0 ) : X G → X. This is true because the identification <strong>of</strong> k[X] with a subalgebra<br />

<strong>of</strong> k[G] is determined by the selection <strong>of</strong> a base point and changing the base point corresponds to<br />

right translation <strong>of</strong> the subalgebra in k[G] by Corollary 3. Then the morphism β (X,h·x0 ) : X G → X<br />

is defined by the inclusion <strong>of</strong> A Γ(X,h·x0 ) = r h (A Γ(X,x0 )) ⊂ A Γ G, from which it is clear that β (X,h·x0 ) =<br />

r h (β (X,x0 )).<br />

Proposition 11 (Universal Property <strong>of</strong> Biequivariant Resolutions). Suppose X is a left-<br />

G-equivariant affine G-embedding, x 0 ∈ X is a base point, Y is any (G × G)-equivariant affine<br />

G-embedding, and ϕ : Y → X is a left-G-equivariant morphism. Then there is a unique (G × G)-<br />

equivariant morphism ϕ (XG ,x 0 ) : Y → X G such that ϕ = β (X,x0 )◦ϕ (XG ,x 0 ). That is, the biequivariant<br />

resolution <strong>of</strong> X satisfies the following diagram:<br />

Y <br />

<br />

<br />

<br />

∀ ϕ<br />

X<br />

∃! ϕ (XG ,x 0 ) β (X,x0 )<br />

X G<br />

If x 1 = h · x 0 is another base point for X, then β (X,x1 ) = r h (β (X,x0 )) and ϕ (XG ,x 1 ) = r h −1(ϕ (XG ,x 0 )).<br />

Pro<strong>of</strong>. Let X be an affine G-embedding and identify k[X] with the left-invariant subalgebra A Γ(X,x0 )<br />

<strong>of</strong> k[G] by selecting a base point x 0 ∈ X. Suppose that Y is a (G × G)-equivariant affine G-<br />

embedding and that ϕ : Y → X is a left-G-equivariant morphism from Y to X. By equivariance,<br />

there is a unique element y 0 ∈ Y such that ϕ(y 0 ) = x 0 , since ϕ| ΩY<br />

: Ω Y → Ω X is an isomorphism<br />

with Ω Y<br />

∼ = ΩX ∼ = G. Then the cone Γ(Y, y0 ) is G-stable by Proposition 10 and is a subset <strong>of</strong><br />

Γ(X, x 0 ) by Proposition 9. Using y 0 , identify k[Y ] with the (G × G)-invariant subalgebra A Γ(Y,y0 )<br />

<strong>of</strong> k[G]. By Corollary 3, A Γ(Y,h·y0 ) = r h (A Γ(Y,y0 )) = A Γ(Y,y0 ), for all h ∈ G, so the algebra A Γ(Y,y0 ) is<br />

independent <strong>of</strong> the choice <strong>of</strong> base point. Hence it is the only subalgebra <strong>of</strong> k[G] which is isomorphic<br />

to k[Y ] as a (G × G)-algebra by Theorem 12.<br />

65


Clearly Γ(Y, y 0 ) = ⋂ h∈G hΓ(Y, y 0)h −1 ⊂ ⋂ h∈G hΓ(X, x 0)h −1 = Γ(X, x 0 ) G ⊂ Γ(X, x 0 ), so that<br />

A Γ(X,x0 ) ⊂ A Γ(X,x0 ) G<br />

⊂ A Γ(Y,y 0 ), as indicated in the diagram below. Moreover, as Γ(X, x 0 ) G is<br />

G-stable, applying Corollary 3 again shows that A Γ(X,x0 ) G<br />

is a uniquely determined subalgebra<br />

<strong>of</strong> k[G] which is independent <strong>of</strong> the choice <strong>of</strong> base point x 0 ∈ X made above. Thus we have the<br />

following diagram <strong>of</strong> algebras:<br />

k[Y ]<br />

ϕ ◦ <br />

k[X]<br />

k[G]<br />

⊂<br />

∼= A<br />

ψy ◦ Γ(Y,y0 )<br />

0<br />

⊂<br />

<br />

⊂<br />

⊂<br />

<br />

∼= A<br />

ψx ◦ Γ(X,x0 )<br />

0<br />

A Γ(X,x0 ) G = k[X G]<br />

The morphism ϕ (XG ,x 0 ) : Y → X G corresponds to the restriction <strong>of</strong> the homomorphism (ψ ◦ y 0<br />

) −1<br />

from k[X G ] = A Γ(X,x0 ) G → k[Y ], while β (X,x 0 ) : X G → X is dual to the composition <strong>of</strong> ψ ◦ x 0<br />

:<br />

k[X] → A Γ(X,x0 ) followed by the canonical inclusion A Γ(X,x0 ) ⊂ A Γ(X,x0 ) G. Then it is clear that ϕ<br />

factors as ϕ = β (X,x0 ) ◦ ϕ (XG ,x 0 ).<br />

Now suppose that x 1 = h · x 0 is another base point in X. Then h · y 0 is the unique element<br />

<strong>of</strong> Y such that ϕ(y) = x 1 , for ϕ(h · y 0 ) = h · ϕ(y 0 ) = h · x 0 . Then ψ ◦ x 1<br />

is an isomorphism from<br />

k[X] to A Γ(X,h·x0 ) = A hΓ(X,x0 )h −1 = r h(A Γ(X,x0 )) ⊂ k[G]. Yet Γ(Y, h · y 0 ) = Γ(Y, y 0 ) ⊂ Γ(X, x 0 ) G =<br />

Γ(X, h · x 0 ) G ⊂ Γ(X, h · x 0 ), so we have ϕ ◦ = (ψh·y ◦ 0<br />

) −1 ◦ ψh·x ◦ 0<br />

: k[X] → k[Y ]. As above, this factors<br />

through the inclusions A Γ(X,h·x0 ) ⊂ k[X G ] = A Γ(X,h·x0 ) G ⊂ A Γ(Y,h·y 0 ), giving morphisms ϕ (XG ,h·x 0 ) :<br />

Y → X G and β (X,h·x0 ) : X G → X corresponding to (ψh·y ◦ 0<br />

) −1 = [r h (ψy ◦ 0<br />

)] −1 = r h −1[(ψy ◦ 0<br />

) −1 ] and<br />

ψh·x ◦ 0<br />

= r h (ψx ◦ 0<br />

), respectively. Hence ϕ (XG ,h·x 0 ) = r h −1(ϕ (XG ,x 0 )) and β (X,h·x0 ) = r h (β (X,x0 )).<br />

66


References<br />

[1] A. Bialynicki-Birula, Some theorems on actions <strong>of</strong> algebraic groups, Ann. <strong>of</strong> Math. (2) 98<br />

(1973), 480–497.<br />

[2] F. Bien and M. Brion, Automorphisms and local rigidity <strong>of</strong> regular varieties, Compositio Mathematica<br />

104 (1996), 1–26.<br />

[3] E. Bifet, C. De Concini and C. Procesi, Cohomology <strong>of</strong> regular embeddings, Adv. Math. 82:1<br />

(1990), 1–34.<br />

[4] A. Borel, Linear <strong>Algebraic</strong> <strong>Groups</strong>, 2nd Edition, Graduate Texts in Mathematics 126, Springer-<br />

Verlag, 1991.<br />

[5] M. Brion, The behaviour at infinity <strong>of</strong> the Bruhat decomposition, Comm. Math. Helvetici 73<br />

(1998), 137–174.<br />

[6] M. Brion, Group completions via Hilbert schemes, J. <strong>Algebraic</strong> Geom. 12 (2003), no. 4, 605–<br />

626.<br />

[7] C. De Concini and C. Procesi, Complete symmetric varieties in Invariant Theory (Montecantini<br />

1982), 1–44, Lecture Notes in Math 996, Springer, 1983.<br />

[8] C. De Concini and C. Procesi, Complete symmetric varieties, II Intersection theory in <strong>Algebraic</strong><br />

<strong>Groups</strong> and Related Topics (Kyoto/Nagoya, 1983), 481–513, Adv. Stud. Pure Math. 6, North-<br />

Holland, 1985.<br />

[9] C. De Concini and C. Procesi, Cohomology and compactifications <strong>of</strong> algebraic groups, Duke<br />

Math. J. 53:3 (1986), 585–594.<br />

67


[10] D. Cox, The homogeneous coordinate ring <strong>of</strong> a toric variety, J. <strong>Algebraic</strong> Geom. 4:1 (1995),<br />

17–50.<br />

[11] D. Cox and S. Katz, Mirror Symmetry and <strong>Algebraic</strong> Geometry, Mathematical Surveys and<br />

Monographs 68, American Mathematical Society, 1999.<br />

[12] V.I. Danilov, The geometry <strong>of</strong> toric varieties, Uspekhi Mat. Nauk. 33:2 (1978), 85–134.<br />

[13] M. Demazure, Sous-groupes algébriques de rang maximum du groupe de Cremona, Ann. Sci.<br />

École Norm. Sup. (4) 3 (1970), 507–588.<br />

[14] G. Ewald, Combinatorial Convexity and <strong>Algebraic</strong> Geometry, Graduate Texts in Mathematics<br />

168, Springer-Verlag, 1996.<br />

[15] R. Fossum, The Divisor Class Group <strong>of</strong> a Krull Domain, Ergebnisse d. Math. 74, Springer-<br />

Verlag, 1973.<br />

[16] R. Fossum and B. Iversen, On Picard groups <strong>of</strong> algebraic fibre spaces, J. Pure Appl. Algebra<br />

3 (1973), 269–280.<br />

[17] W. Fulton, Introduction to Toric Varieties, Annals <strong>of</strong> Mathematics Studies 131, Princeton<br />

University Press, 1993.<br />

[18] R. Godement, Topologie Algébrique et Théorie des Faisceaux, Actualit’es Sci. Ind., no. 1252,<br />

Publ. Math. Univ. Strasbourg, no. 13, Hermann, 1958.<br />

[19] F. Grosshans, <strong>Algebraic</strong> homogeneous spaces and invariant theory, Lecture Notes in Math.<br />

1673, Springer-Verlag, 1997.<br />

[20] A. Grothendieck, Éléments de Géométrie Algébrique, I-IV, Inst. Hautes<br />

Études Sci. Publ.<br />

Math. 4,8,11,17,20,24,28,32 (1960–1967).<br />

[21] W. Haboush, Reductive groups are geometrically reductive, Ann. <strong>of</strong> Math. 102 (1975), 67–83.<br />

[22] R. Hartshorne, <strong>Algebraic</strong> Geometry, Graduate Texts in Mathematics 52, Springer-Verlag, 1977.<br />

[23] B. Iversen, A fixed point formula for action <strong>of</strong> tori on algebraic varieties, Inv. Math. 16 (1972),<br />

229–236.<br />

68


[24] B. Iversen, Cohomology <strong>of</strong> Sheaves, Lecture Notes Series 55, Aarhus Universitet, Matematisk<br />

Institut, 1984.<br />

[25] N. Iwahori and H. Matsumoto, On some Bruhat decomposition and the structure <strong>of</strong> Hecke ring<br />

<strong>of</strong> p-adic Chevalley groups, Publ. Math. IHES 25 (1965), 5–48.<br />

[26] J. Jantzen, Representations <strong>of</strong> <strong>Algebraic</strong> <strong>Groups</strong>, Pure and Applied Mathematics 131, Academic<br />

Press, 1987.<br />

[27] G. Kempf, Instability in invariant theory, Ann. <strong>of</strong> Math. (2) 108 (1978), no. 2, 299–316.<br />

[28] G. Kempf, F. Knudsen, D. Mumford, and B. Saint-Donat, Toroidal <strong>Embeddings</strong> I, Lecture<br />

Notes in Mathematics 339, Springer-Verlag, 1973.<br />

[29] F. Knop, The Luna-Vust theory <strong>of</strong> spherical embeddings, Proc. <strong>of</strong> the Hyderabad Conf. on<br />

<strong>Algebraic</strong> <strong>Groups</strong> (Hyderabad, 1989), 225–249, Manoj Prakashan, Madras, 1991.<br />

[30] D. Luna and Th. Vust, Plongements d’espaces homogènes, Comment. Math. Helv. 58 (1983),<br />

186–245.<br />

[31] H. Matsumura, Commutative Ring Theory, Cambridge Studies in Adv. Math. 8, Cambridge,<br />

1986.<br />

[32] D. Mumford, J. Fogarty and F. Kirwan, Geometric Invariant Theory, 3rd Enlarged Edition,<br />

Ergebnisse d. Math. Ser. 3 34, Springer-Verlag, 1994.<br />

[33] T. Oda, Convex Bodies and <strong>Algebraic</strong> Geometry: An introduction to the theory <strong>of</strong> toric varieties,<br />

Ergebnisse d. Math. Ser. 3 15, Springer-Verlag, 1988.<br />

[34] D.I. Panyushev, Complexity and rank <strong>of</strong> homogeneous spaces, Geometriae Dedicata 34 (1990),<br />

249–269.<br />

[35] V.L. Popov, Quasihomogeneous affine varieties <strong>of</strong> the group SL(2), Izv. Akad. Nauk SSSR<br />

Ser. Mat 37 (1973), 792–832.<br />

[36] V.L. Popov, <strong>Groups</strong>, Generators, Syzygies, and Orbits in Invariant Theory, Translations <strong>of</strong><br />

Mathematical Monographs 100, American Mathematical Society, 1992.<br />

69


[37] V.L. Popov and É.B. Vinberg, On a class <strong>of</strong> quasihomogeneous affine varieties, Izv. Akad.<br />

Nauk SSSR Ser. Mat 36 (1972), no. 4, 749–763.<br />

[38] V.L. Popov and É.B. Vinberg, Invariant theory in <strong>Algebraic</strong> Geometry, 4, 137–315, Itogi Nauki<br />

i Tekhniki, Akad. Nauk. SSSR, Vsesoyuz Inst. Nauchn. i Tekhn. Inform., 1989.<br />

[39] G. Rousseau, Immeubles sphériques et théorie des invariants, C.R. Acad. Sci. Paris Sér. A-B<br />

286 (1978), no. 5, A247–A250.<br />

[40] T.A. Springer, Linear <strong>Algebraic</strong> <strong>Groups</strong>, 2nd Edition, Progress in Mathematics 9, Birkhäuser,<br />

1998.<br />

[41] E. Strickland, A vanishing theorem for group compactifications, Math. Ann. 277:1 (1987),<br />

165–171.<br />

[42] H. Sumihiro, <strong>Equivariant</strong> completion, J. Math. Kyoto Univ. 14 (1974), no. 1, 1–28.<br />

[43] H. Sumihiro, <strong>Equivariant</strong> completion II, J. Math. Kyoto Univ. 15 (1975), no. 3, 573–605.<br />

[44] A. Tchoudjem, Cohomologie des fibrés en droites sur les compactifications des groupes<br />

réductifs, preprint math.AG/0303125, 37 pages.<br />

[45] D.A. Timashev, Classification <strong>of</strong> G-varieties <strong>of</strong> complexity 1, Izv. Ross. Akad. Nauk Ser. Mat.<br />

61 (1997), no. 2, 127–162; English translation: Izv. Math. 61 (1997), no. 2, 363–397.<br />

[46] D.A. Timashev, <strong>Equivariant</strong> compactifications <strong>of</strong> reductive groups, Mat. Sb. 194 (2003), no.<br />

4, 119–146.<br />

[47] J. Tits, On buildings and their applications, Proceedings <strong>of</strong> the International Congress <strong>of</strong><br />

Mathematicians, Vancouver, 1974<br />

[48] J. Tits, Buildings <strong>of</strong> Spherical Type and Finite BN-Pairs, Lecture Notes in Mathematics 386,<br />

Springer-Verlag, 1974.<br />

[49] É.B. Vinberg, Complexity <strong>of</strong> action <strong>of</strong> reductive groups, Funct. Anal. Appl. 20 (1986), 1–11.<br />

[50] O. Zariski and P. Samuel, Commutative Algebra, vol. 1 and 2, Graduate Texts in Mathematics<br />

28, 29, Springer-Verlag, 1975.<br />

70


Vita<br />

David Charles Murphy was born in Grand Rapids, Michigan on September 29, 1973. In 1996 he<br />

graduated Summa Cum Laude from Western Michigan University in Kalamazoo receiving a B.A.<br />

in Mathematics. He received his M.S. in Mathematics in 1998 from the University <strong>of</strong> Illinois at<br />

Urbana-Champaign. In 2004 he earned his Ph.D. in Mathematics from the University <strong>of</strong> Illinois at<br />

Urbana-Champaign. Throughout the course <strong>of</strong> his graduate studies, he was a teaching assistant and<br />

a VIGRE pre-doctoral fellow. During this time, he won the Department <strong>of</strong> Mathematics Teaching<br />

Assistant Instructional Award (2001), Delta Sigma Omicron Distinguished Teaching Award (2001),<br />

and the Irving Reiner Memorial Award in Algebra (2003).<br />

71

Hooray! Your file is uploaded and ready to be published.

Saved successfully!

Ooh no, something went wrong!