You are on page 1of 229

Texts and Readings in Mathematics 46

C.S. Seshadri

Introduction to the
Theory of Standard
Monomials
Second Edition
Texts and Readings in Mathematics

Volume 46

Advisory Editor
C.S. Seshadri, Chennai Mathematical Institute, Chennai

Managing Editor
Rajendra Bhatia, Indian Statistical Institute, New Delhi

Editor
Manindra Agrawal, Indian Institute of Technology Kanpur, Kanpur
V. Balaji, Chennai Mathematical Institute, Chennai
R.B. Bapat, Indian Statistical Institute, New Delhi
V.S. Borkar, Indian Institute of Technology Bombay, Mumbai
T.R. Ramadas, Chennai Mathematical Institute, Chennai
V. Srinivas, Tata Institute of Fundamental Research, Mumbai
The Texts and Readings in Mathematics series publishes high-quality textbooks,
research-level monographs, lecture notes and contributed volumes. Undergraduate
and graduate students of mathematics, research scholars, and teachers would find
this book series useful. The volumes are carefully written as teaching aids and
highlight characteristic features of the theory. The books in this series are
co-published with Hindustan Book Agency, New Delhi, India.

More information about this series at http://www.springer.com/series/15141


C.S. Seshadri

Introduction to the Theory


of Standard Monomials
Second Edition

123
C.S. Seshadri
Chennai Mathematical Institute
Chennai
India

This work is a co-publication with Hindustan Book Agency, New Delhi, licensed for sale in all
countries in electronic form only. Sold and distributed in print across the world by Hindustan
Book Agency, P-19 Green Park Extension, New Delhi 110016, India. ISBN: 978-93-80250-58-8
© Hindustan Book Agency 2015.

ISSN 2366-8725 (electronic)


Texts and Readings in Mathematics
ISBN 978-981-10-1813-8 (eBook)
DOI 10.1007/978-981-10-1813-8
Library of Congress Control Number: 2016944375

© Springer Science+Business Media Singapore 2016 and Hindustan Book Agency 2015
This work is subject to copyright. All rights are reserved by the Publishers, whether the whole or part
of the material is concerned, specifically the rights of translation, reprinting, reuse of illustrations,
recitation, broadcasting, reproduction on microfilms or in any other physical way, and transmission
or information storage and retrieval, electronic adaptation, computer software, or by similar or dissimilar
methodology now known or hereafter developed.
The use of general descriptive names, registered names, trademarks, service marks, etc. in this
publication does not imply, even in the absence of a specific statement, that such names are exempt from
the relevant protective laws and regulations and therefore free for general use.
The publishers, the authors and the editors are safe to assume that the advice and information in this
book are believed to be true and accurate at the date of publication. Neither the publishers nor the
authors or the editors give a warranty, express or implied, with respect to the material contained herein or
for any errors or omissions that may have been made.

This Springer imprint is published by Springer Nature


The registered company is Springer Science+Business Media Singapore Pte Ltd.
D E D I C A T I O N

I would like to dedicate this book to the memory of C. Musili


my former student, friend and collaborator in the subject of
these lectures. His sudden death in October, 2005, came as a
shock to his friends.
Contents

Preface to the second edition ix

Preface to the first edition xi

Introduction xiii

About the Author xv

1 Schubert Varieties in the Grassmannian 1


1.1 Plücker coordinates . . . . . . . . . . . . . . . . . . . . . . 1
1.2 Schubert varieties . . . . . . . . . . . . . . . . . . . . . . . 5
1.3 Standard monomials . . . . . . . . . . . . . . . . . . . . . 12
1.4 Some Applications . . . . . . . . . . . . . . . . . . . . . . 19
1.5 Degeneration of Schubert varieties . . . . . . . . . . . . . 32
1.6 Connection with determinantal varieties
and invariant theory . . . . . . . . . . . . . . . . . . . . . 38

2 Standard monomial theory on SLn (k)/Q 55


2.1 Some facts about G/Q . . . . . . . . . . . . . . . . . . . . 55
2.2 Young diagrams and standard monomials . . . . . . . . . 59
2.3 Linear independence of standard monomials . . . . . . . . 61
2.4 Some facts about the partial order on W/WQi . . . . . . 63
2.5 Preparation for the main theorem . . . . . . . . . . . . . 65
2.6 Main theorem . . . . . . . . . . . . . . . . . . . . . . . . . 70
2.7 Another proof for generation by standard monomials . . . 74

3 Applications 81
3.1 Singularities of Schubert varieties . . . . . . . . . . . . . . 81
3.2 Vanishing theorem . . . . . . . . . . . . . . . . . . . . . . 85
3.3 Character formula . . . . . . . . . . . . . . . . . . . . . . 89
3.4 Ideal theory of Schubert varieties . . . . . . . . . . . . . . 93
3.5 The variety of complexes . . . . . . . . . . . . . . . . . . . 97

vii
viii CONTENTS

4 Schubert varieties in G/Q 107


4.1 Some remarks on linear algebraic groups . . . . . . . . . . 107
4.2 Basic properties . . . . . . . . . . . . . . . . . . . . . . . . 110
4.3 Reduced decompositions . . . . . . . . . . . . . . . . . . . 118
4.4 The normalization map . . . . . . . . . . . . . . . . . . . 126
4.5 Chevalley’s multiplicity formula . . . . . . . . . . . . . . . 129
4.6 Deodhar’s Lemma . . . . . . . . . . . . . . . . . . . . . . 134

Appendix A. Cohen-Macaulay Properties 139

Appendix B. Normality of Schubert varieties 157

Appendix C. Standard Monomial Theory 165

Bibliography 213

Notation 217

Index 219

Symbols 221

Texts and Readings in Mathematics 223


Preface to the second edition
In the introduction to the first edition, there was a mention of conjec-
tures of a Standard Monomial Theory (SMT) for a general semi-simple
(simply-connected) algebraic group. These conjectures, due to Laksh-
mibai, appeared in a paper in the proceedings of a conference held at the
University of Hyderabad in 1989. This paper is added as Appendix C in
this second edition, keeping the same list of references as in the paper.
The conjectures were later proved by P. Littelmann by using ideas from
quantum groups and marked a significant progress in SMT.
Many typographical errors have been corrected and the Bibliography
has been revised.
I wish to thank my colleague Manoj Kummini for his meticulous
reading of the revised manuscript.

ix
Preface to the first edition
This book is a reproduction of the Brandeis Lectures Notes 4 with
corrections of typographical errors and bringing up to date some refer-
ences. I wish to thank my colleague K.V. Subrahmanyam for his help
in this regard.

xi
Introduction
These are the notes of a course of lectures that I gave at Brandeis
during the academic year 1983-84. The purpose of this course was to give
an introduction to the series of papers that I have been writing under the
title Geometry of G/P (see G/P- I to V) (or Standard monomial theory
- abbreviated SMT) in collaboration with V. Lakshmibai and C. Musili.
Hodge (see [Ho], [H-P]) studied Schubert varieties in a Grassmannian
by giving canonical bases of the homogeneous coordinate rings of these
varities (bases of consisting of what he called “Standard monomials” in
the Plücker coordinates, since these can be represented by “Standard
tableaux” in the sense of Young). The aim of these papers has been
to generalize this work of Hodge to the case of Schubert varieties in
flag varieties associated to any semi-simple, simply-connected, algebraic
group G. The goals in SMT are the following:

(i) To give a canonical basis of H 0 (G/P, L), where P is a maximal


parabolic subgroup of G and L is the ample generator of Pic G/P.
When the base field is of characteristic zero, this problem is equiv-
alent to giving canonical bases for the fundamental representations
of G.

(ii) To give a basis of H 0 (G/B, M ) and more generally of H 0 (X, M ),


formed out of “standard monomials” in the basis elements of (i),
where B is a Borel subgroup of G, M is a line bundle on the flag
variety G/B associated to a dominant character of B and X is a
Schubert subvariety of G/B. When the base field is of character-
istic zero and X = G/B, thanks to a theorem of Borel-Weil, this
problem is equivalent to giving a basis of any finite dimensional ir-
reducible G-module, formed by “standard monomials” in the basis
elements of (i).

These goals have been achieved in the papers Geometry of G/P - I to


V (see G/P-V for a self-contained account) when G is a classical group
(also for G = G2 , see [L]) and for certain classes of representations, even
if G is exceptional. There are also conjectures for an SMT for a general
semi-simple algebraic group.
In these lectures, we deal only with the case G = SLn . In this case,
the solution of question (i) above is immediate, namely the Plücker
coordinates give the required basis elements. In these notes one gets a
xiii
xiv INTRODUCTION

fair introduction to the methods and the material of our papers on SMT
but for the question (i).
A consequence of SMT is that it gives a canonical set of generators
for the ideal (or ideal sheaf) defining a Schubert variety in a flag vari-
ety. This leads to the determination of the singular locus of a Schubert
variety. We give here also many other applications of SMT like the van-
ishing theorems for the cohomology of line bundles on Schubert varieties
(associated to dominant characters of B), Cohen-Macaulay properties of
Schubert varieties, classical invariant theory etc. As things stand now,
for applications like the vanishing theorems, the methods in the papers
[M-R] and [R-R] go much farther and cover the case of an arbitrary
semi-simple algebraic group G.
When these lectures were given, because of the gap found in the work
of Demazure [D], one could not use the fact (as is done in [L-M-S1])
that Schubert varieties in flag varieties, associated to any semi-simple
algebraic group, are normal. In the main body of this course, this fact is
deduced as a consequence of SMT for the cases treated here (see G/P-
V for a more general situation). Just when the course ended, a general
proof of normality was found and this is given in the Appendix 2. Joseph
(see [J]) had just by that time justified the Demazure character formula
for large dominant weights (when the base field is of characteristic zero)
and it can be easily seen that this is equivalent to the normality of
Schubert varieties (when the base field is of characteristic zero). There is
now another proof of normality due to S. Ramanan and A. Ramanathan
(see [R-R] cited above).
These lectures were given during a pleasant stay at Brandeis and I
would like to take this opportunity to thank my colleagues there and
especially David Eisenbud who suggested that I give this type of course.
Lastly, it is a great pleasure to thank Peter Littlemann and Pradeep
Shukla who have done a temendous job in writing these notes.

C. S. Seshadri
1 September, 1984
About the Author

C.S. Seshadri, FRS, an eminent Indian mathematician, is director-emeritus of the


Chennai Mathematical Institute, India. He is known for his work in algebraic
geometry. The well-known “Seshadri constant” is named after him. His work with
M.S. Narasimhan on unitary vector bundles and the Narasimhan–Seshadri theorem
has influenced the field. His work on geometric invariant theory and on Schubert
varieties, in particular his introduction of standard monomial theory, is widely
recognized. A recipient of the Padma Bhushan in 2009, the third highest civilian
honor in India, he was elected Fellow of the Indian Academy of Sciences in 1971.
Professor Seshadri worked in the School of Mathematics at the Tata Institute of
Fundamental Research, Mumbai, during 1953–1984, starting as a research scholar
and rising to a senior professor. From 1984 to 1989, he worked in Institute of
Mathematical Sciences, Chennai, India. From 1989 to 2010, he worked as the
founding director of the Chennai Mathematical Institute.

xv
Chapter 1

Schubert Varieties in the Grassmannian

Throughout our basic goal will be to study Schubert varieties using the
standard monomial theory which we shall develop in various situations.
This chapter is devoted to the study of Schubert varieties in the Grass-
mannian, which is a particular case of a more general situation to be
treated in the subsequent chapters. We begin by briefly recalling in sec-
tion 1 the basic definitions associated with the Grassmannian. In section
2, we define Schubert varieties and give their basic properties. In section
3, we develop the standard monomial theory, essentially due to Hodge
[H-P], and apply it in section 4 to prove some results on Schubert vari-
eties. Section 5 contains a result on degeneration of Schubert varieties,
which connects the study of Schubert varieties to combinatorics. Fi-
nally in section 6, exploiting our knowledge of Schubert varieties, we
study determinantal varieties and prove two results of classical invariant
theory in the form given to them by Doubilet-Rota-Stein D-R-S] and De
Concini and Procesi D-P].
The material of sections 1-4 is essentially that of [Mu]. Section 5
gives a somewhat different exposition of the main “degeneration result”
of [D-E-P]. Section 6 follows [L-S-2].

1.1 Plücker coordinates

Let V be a vector space of dimension n ≥ 1 over an algebraically closed


field k. Fixing a basis {e1 , e2 , . . . , en } of V, we may identify the elements
of V with n × 1 matrices over k. This allows us to write the elements

© Springer Science+Business Media Singapore 2016 and Hindustan Book Agency 2015 1
C.S Seshadri, Introduction to the Theory of Standard Monomials,
Texts and Readings in Mathematics 46, DOI 10.1007/978-981-10-1813-8_1
2 1. Schubert Varieties in the Grassmannian

of V as column vectors with


 
  0  
1   0
   1   .. 
 0    
e1 =   , e2 =  0  , . . . , en =  . 
.
..
 .   ..   0 
 . 
0 1
0

For a fixed integer r with 1 ≤ r ≤ n, let I(r, n) denote the indexing set

{α = (α1 , α2 , . . . , αr ) | αi ∈ Z and 1 ≤ α1 < α2 < · · · < αr ≤ n} ,

and let {eα = e(α1 ,α2 ,...,αr ) , α ∈ I(r, n)}} be the basis of Λr V defined by

eα = eα1 ∧ eα2 ∧ . . . ∧ eαr

Let {xα | α ∈ I(r, n)} be the basis of the dual space (Λr V )∗ , which
is dual to the above basis. Then, the polynomial algebra k[xα ] is the
homogeneous coordinate ring of the projective space P(Λr V ) consisting
of one-dimensional linear subspaces of Λr V ; it is also the coordinate ring
of the affine space Λr V, which is a cone over P(Λr V ).

1.1.1 Grassmannian
We shall denote by Gr,n , the Grassmannian of r-dimensional linear sub-
spaces of V. We recall that it can be endowed with the structure of a
variety as follows. Let

Vr = V
| ⊕V ⊕
{z. . . ⊕ V}
r copies

= M (n, r), thesetof n × r matrices,

and

V r,0 = {(v1 , v2 , . . . , vr ) ∈ V r | v1 . . . , vr are linearly independent}



= M (n, r)0 , thesetof n × r matrices of rank r.

Clearly, the column vectors of a matrix in M (n, r)◦ generate an element


of Gr,n , and the column vectors of two matrices A, B in M (n, r)◦ gen-
erate the same element of Gr,n if and only if A = BC for some C in
GLr (k). Thus Gr,n can be identified with M (n, r)◦ modulo the equiva-
lence relation
1.1. Plücker coordinates 3


A ∼ B in M (n, r)◦ iff A = BC
(1.1)
for some C in GLr (k)
Let η : M (n, r)◦ → M (n, r)◦ / ∼ be the canonical map, and for α ∈
I(r, n), let
n o
Uα = A ∈ M (n, r)◦ | αth minor of A is the identity matrix ,

where by the αth minor of A we mean the r × r minor whose ith row
is the αth ◦
i row of A. Obviously, η : Uα → M (n, r) / ∼ is injective and
using the above identification of Gr,n , we have
[
Gr,n = η(Uα )
α∈I(r,n)

Further, it follows from the definition of Uα that there is a bijection


between each Uα and the affine r(n − r)-space. One can easily verify
that these bijections define compatible affine charts on Gr,n , which makes
Gr,n into a variety of dimension r(n − r).
We shall often identify η(Uα ) with Uα and shall regard Gr,n as a
union of Uα , α ∈ I(r, n).

1.1.2 Plücker embedding


We now realize Gr,n as a projective variety by embedding it in the pro-
jective space P(Λr V ). Consider the morphism π b : V r → Λr V defined
by π b(v1 , v2 , . . . , vr ) = v1 ∧ v2 ∧ · · · ∧ vr . Its restriction to V r,o induces
a morphism π : V r,o → P(Λr V ) which in turn induces a morphism
π̃ : Gr,n → P(Λr V ), since π is constant on the equivalence classes of the
relation (1.1)
First we show that Im π̃ is a closed subset of P(∧r ∨). To this end,
let
Aβ = {(zα ) ∈ P(∧r ∨) | zβ = 1} for β in I(r, n).
Since {Aβ , β ∈ I(r, n)} is an open covering of P(∧r ∨), it is enough to
show that for every β in I(r, n), Aβ ∩ Im π̃ is closed in Aβ . For simplicity,
we show this for β = (1, 2, . . . , r), the argument being the same for an
arbitrary β. In fact, it is easy to see that Aβ ∩ Im π̃(Uβ ) = π̃(Uβ ), which
implies that Aβ ∩ Im π̃ consists of elements (zα ) ∈ P ∈ (∧r ∨), where

zα = ± (Determinant of the αth minor of an element in Uβ ).


4 1. Schubert Varieties in the Grassmannian

Since Uβ consists of n × r matrices of the form


 
Ir Ir = the r × r identity matrix
,
A and A = [aij ] ∈ M (n − r, r),

it follows that the elements of Aβ ∩ Im π̃ are of the form

(. . . , aij , . . . , fλ (aij ), . . .),

where aij are arbitrary and fλ are polynomial functions in aij . This
shows that Aβ ∩ Im π̃ is closed in the affine space Aβ , as we can actu-
ally specify its defining equations. Thus Im π̃ is a projective variety in
P(∧r ∨).
Further, from the description of π̃(Uβ ), it also follows that π̃ : Uβ →
π̃(Uβ ) is an isomorphism of affine varieties and that π̃ −1 (π̃(Uβ )) = Uβ .
This shows that π̃ : Gr,n → Im π̃ is locally an isomorphism and in fact,
an isomorphism. Thus π̃ embeds Gr,n as a projective variety in P(∧r ∨).
It is called the Plücker embedding of the Grassmannian Gr,n and the
coordinates of its image are called the Plücker coordinates of Gr,n .
Hereafter, we shall not distinguish between the Grassmannian and
its image under the Plücker embedding and shall denote both by Gr,n .

1.1.3 Remark
1. Im π = Gr,n

b is the cone over the Grassmannian Gr,n .


2. Im π

3. GLr (k) acts freely on M (n, r)◦ by multiplication on the right and
its orbits are precisely the equivalence classes of the relation 1.1.
Thus Gr,n can also be viewed as the orbit space of M (n, r)◦ under
the action GLr (k).

4. π̂ is constant on the SLr (k)-orbits in M (n, r)◦ .

5. π : M (n, r)◦ → Gr,n defines a principal GLr (k)-bundle on Gr,n .

6. Gr,n can also be obtained as a quotient of GLn (k) or SLn (k) by a


“maximal parabolic subgroup” (see Chapter 2, Section 1).
1.2. Schubert varieties 5

1.1.4 The homogeneous coordinate ring of Gr,n :


For (v1 , v2 , . . . , vr ) in V r , let
n
X
vj = xij ei , 1 ≤ j ≤ r.
i=1

Then xij ’s are coordinate functions on the affine space V r ∼= M (n, r) and
the polynomial k-algebra k[xij ] is the coordinate ring of V r . Now the
morphism π̂ : V r → Λr V induces a k-algebra homomorphism between
the coordinate rings, namely, π̂ ∗ : k[xα ] → k[xij ] defined by

π̂ ∗ (xα ) = pα ,

where  
xα1 1 xα1 2 . . . xα1 r
 xα2 1 xα2 2 . . . xα2 r 
 
pα = ± det  .. .. .. ,
 . . . 
xαr 1 xαr 2 . . . xαr r
the determinant of the αth minor (with its sign) in the n × r matrix
[xij ]. Since ker π̂ ∗ is the ideal of Im π̂, the cone over Gr,n , it is the
homogeneous ideal of Gr,n in k[xα ]. Thus the homogeneous coordinate
ring of Gr,n can be identified with the k-subalgebra of k[xij ] generated
by pα , α ∈ I(r, n), and Gr,n has a natural scheme structure defined by
its homogeneous coordinate ring k[pα ], namely, Proj k[pα ]. In particular,
Gr,n is a closed integral subscheme of P(∧r V ).

1.1.5 Remark
(i) Note that pα is nothing but the αth Plücker coordinate of Gr,n ,
i.e., any point on Gr,n is of the form (pα ) for some [xij ] ∈ M (n, r).

(ii) We shall often denote the restriction of a Plücker coordinate pα to


a subvariety of Gr,n also by pα .

1.2 Schubert varieties

In this section, we describe a decomposition of the Grassmannian into


a finite disjoint union of affine spaces called the Schubert cells of the
Grassmannian. What is noteworthy about these cells is that they are
indexed by the set I(r, n) in such a way that the dimension of each cell
6 1. Schubert Varieties in the Grassmannian

and hence, the cell itself, is determined by its index α = (α1 , α2 , . . . , αr ).


Later in this section, we define Schubert varieties as closures of the
Schubert cells and describe some of their basic properties.
A flag in a vector space V is a chain {0} = V0 ⊂ V1 ⊂ · · · ⊂
Vm = V of subspaces of V with Vi 6= Vi+1 . A flag is called a full flag if
dim Vi /Vi−1 = 1 for each i, 1 ≤ i ≤ m.

1.2.1 Definition
Fix a full flag {0} = V0 ⊂ V1 ⊂ · · · ⊂ Vn = V in an n-dimensional vector
space V. We define the Schubert cell of Gr,n associated to the index α in
I(r, n) (or simply, the αth Schubert cell of Gr,n ) to be the set
C(α) = {W ∈ Gr,n | dim W ∩ Vj = i, if αi ≤ j < αi+1 ,
where 1 ≤ j ≤ n, 0 ≤ i ≤ r and α0 = 0} .
The above definition of a Schubert cell C(α) depends on the choice of
a full flag. However, given any two full flags V0 ⊂ V1 ⊂ · · · ⊂ Vn and
V0′ ⊂ V1′ ⊂ · · · ⊂ Vn′ = V in V, there exists an automorphism f of V
such that f (Vi ) = Vi′ , which shows that C(α) is well defined up to an
automorphism of V.
Thus, without any loss of generality, we may fix our full flag to be
V0 ⊂ V1 ⊂ · · · ⊂ Vn , where Vi = he1 , e2 , . . . , ei i is the subspace of
V generated by e1 , e2 , . . . , ei . As before, if we write the elements of V
as column vectors, an r-dimensional subspace W or V “corresponds”
to an n × r matrix in M (n, r)◦ , in the sense that the columns of this
matrix generate W. We intend to determine the shape of the matrices
corresponding to the subspaces W in C(α).
So, let W ∈ C(α). It follows from the definition of W that α1 is the
least integer such that dim W ∩ Vα1 = 1. This implies that W contains
a nonzero element of Vα1 , say
 
a11
 a21 
 
 .. 
 . 
 
 aα 1 
 1 
a1 = 


 0 
 
 0 
 
 . 
 .. 
0
1.2. Schubert varieties 7

Clearly, aα11 6= 0, for otherwise dim W ∩ Vi 6= 0 for some i < α1 ,


contrary to the definition of W. Further, since α2 is the least integer
with dim W ∩ Vα2 = 2 and W ∩ Vα1 ⊂ W ∩ Vα2 , W contains an element
of Vα1 − ha1 i, say  
a12
 a22 
 
 .. 
 . 
 
a2 =  
 aα22  ,
 0 
 
 . 
 .. 
0
where aα22 6= 0. Proceeding inductively in this manner, we find that w
contains r linearly independent vectors
 
a1i
 a2i 
 
 .. 
 . 
 
 aα i 
ai =  
 0  , 1 ≤ i ≤ r,
i

 
 0 
 
 . 
.
 . 
0

with aαi i 6= 0. Thus an n × r matrix whose columns generate a subspace


W in C(α) is of the form
 
a11 a12 · · · a1r
 a21 a22 · · · a2r 
 
 .. .. .. 
 . . . 
 
 aα 1 aα 2 · · · aα1 r 
 1 1 
 .. .. 
 0 . . 
 
 0 a · · · a  , where aα i 6= 0 for each i. (1.2)
 α 22 α 2r  i
 .. .
.. .
..  
 .
 
 0 0 · · · aαrr 
 
 0 0 ··· 0 
 
 . . . 
 . . .
. .
. 
0 0 ··· 0
8 1. Schubert Varieties in the Grassmannian

On the other hand, if E(α) is the set of n × r matrices of the above form,
then it is clear that any matrix in E(α) “corresponds” to an element of
C(α). Therefore, as in (1.1), C(α) is identified with E(α) modulo the
equivalence relation induced by the equivalence relation as in (1.1). Note
that C has to be upper triangular.
Now, let E ∗ (α) be the set of matrices in E(α), whose αth minor is
the identity matrix. Since αth minor of every matrix in E(α) is upper
triangular and nonsingular, for each A in E(α), there is a unique upper
triangular C in GLr (k) such that AC ∈ E ∗ (α). This shows that C(α)
can be identified with E ∗ (α). Furthermore, it clearly follows from the
definition of E ∗ (α) that E ∗ (α) is an affine space of dimension
r
X
αi − (dimension of upper triangular matrices in GL(r)))
i=1
r
X r(r + 1)
= αi − .
2
i=1

P
r
r(r+1)
Hence, C(α) is an affine space of dimension αi − 2 .
i=1
Finally, we observe that for any W in Gr,n ,
there is a least integer
α1 such that dim W ∩ Vα1 = 1; then there is a least integer α2 > α1
such that dim W ∩ Vα2 = 2, and so on. Thus each W in Gr,n determines
an element α = (α1 , α2 , . . . , αr ) in I(r, n) such that W ∈ C(α), which
shows that Gr,n = ∪ C(α) and the union is disjoint. We summarize
α∈I(r,n)
the observations of the above discussion in

1.2.2 Theorem (The cellular decomposition of Gr,n )


(i) For each α in I(r, n), the Schubert cell C(α) is an affine space of
Pr
dimension αi − r(r+1)
2 , which, in the Plücker embedding of Gr,n
i=1
can be realized as the image of E(α), the set of n × r matrices of
the form (1.2).

(ii) The Grassmannian Gr,n is a disjoint union of the Schubert cells


C(α), α ∈ I(r, n).

1.2.3 Corollary
Let αmin = (1, 2, . . . , r) and αmax = (n − r + 1, n − r + 2, . . . , n). Then
1.2. Schubert varieties 9

(i) C(αmin ) is a point.


S
(ii) C(αmax ) = and it is the only Schubert cell in the above de-
αmax
composition, which is an open subset of Gr,n . We call it the big
cell.

Proof. Obvious.

1.2.4 Lemma
Let B be the subgroup of upper triangular matrices of GLn (k). There
is an action of B on Gr,n whose orbits are precisely the Schubert cells.

Proof. The action of B on M (n, r)◦ by left multiplication induces


an action of B on M (n, r)◦ modulo the equivalence relation (1.1). This
defines an action of B on Gr,n . We have seen above that Gr,n is a disjoint
union of Schubert cells and each Schubert cell C(α) is identified with the
set E(α) (the set of matrices of the form (1.2)) modulo the equivalence
relation (1.1). Therefore, to prove the lemma, it suffices to show that
E(α) is an orbit under the action of B. In fact, it is easy to see that
B · E(α) ⊆ E(α) and if Eα is the n × r matrix whose αth minor is the
identity matrix and all other entries are zero, then BEα = E(α). Q.E.D

1.2.5 Definition
The closure of the Schubert cell C(α) in Gr,n is called the Schubert vari-
ety corresponding to the index α (or simply, the αth Schubert variety),
which we denote by X(α). As a subscheme of G(r, n), we take X(α) to be
endowed, by definition, with its canonical reduced subscheme structure.
Remark

(i) Since, the big cell C(αmax ) is open and Gr,n is irreducible,
X(αmax ) = Gr,n , i.e. Gr,n itself is a Schubert variety.
P
r
r(r+1)
(ii) In view of (1.2.2, (i)), dim X(α) = αi − 2 .
i=1

1.2.6 Proposition
The homogeneous ideal of a Schubert variety is a prime ideal.
Proof. For α in I(r, n), since C(α) is an irreducible variety, its closure
X(α) is irreducible. This together with the fact that X(α) is reduced,
10 1. Schubert Varieties in the Grassmannian

implies that X(α) is an integral scheme, which in turn implies that the
homogeneous ideal of X(α) is a prime ideal in k[xβ , β ∈ I(r, n)] (the
homogeneous coordinate ring of P(∧r ∨)). Q.E.D.
Let D(α) be the set of n×r matrices whose entries below the (αi , i)th
entry are zero, i.e., the elements of D(α) are of the form
 
∗ ∗ · · · · · · · · ∗
 · · · · · · · · · · · 
 
 · · · · · · · · · · · 
 
 · · · · · · · · · · · 
 
 ∗ ∗ · · · · · · · · ∗ 
 
 ∗ ∗ · · · · · · · · ∗  −α1 -th row
 
 0 ∗ · · · · · · · · ∗ 
 
 · · · · · · · · · · · 
 
 · · · · · · · · · · · 
 
 
 0 ∗ ∗ ∗ · · · · · · ∗ 
 
 0 ∗ ∗ ∗ · · · · · · ∗  −α2 -th row
 
 0 0 ∗ ∗ · · · · · · ∗ 
 
 · · · · · · · · · · · 
 
 · · · · · · · · · · · 
 
 · · · · · · · · · · · 
 
 0 0 · · · · · · · · ∗ 
 
 0 0 · · · · · · · · ∗  −αr -th row
 
 0 0 0 · · · · · · · 0 
 
 · · · · · · · · · · · 
 
 · · · · · · · · · · · 
0 0 0 · · · · · · · 0

1.2.7 Lemma
If D(α)◦ = D(α) ∩ M (n, r)◦ , then the closure of π(D(α)◦ ) is X(α).

Proof. First, we note that the set of matrices in D(α)◦ whose α-th
minor is nonsingular is precisely E(α), the set of n × r matrices of the
form (1.2). Since D(α)◦ is an irreducible (being open in D(α)) closed
subset of M (n, r)◦ and E(α) is open in D(α)◦ , we have

E(α) = D(α)◦ ,
where E(α) is the closure of E(α) in M (n, r)◦ . Now it follows that
π(D(α)◦ ) = π(E(α)) ⊂ π(E(α)) ⊂ π(D(α)◦ ),
1.2. Schubert varieties 11

and hence, π(D(α)◦ ) = π(E(α)) = X(α). Q.E.D.

1.2.8 Remark
It is also true that π(D(α)◦ ) = X(α), a statement stronger than the
above lemma, whose proof is easy but not quite elementary. We have
omitted the proof, as we will not have any occasion to use the stronger
statement in the sequel1 .

1.2.9 Definition
We define a partial order ≤ on I(r, n) by declaring α ≤ β if α1 ≤ β1 ,
α2 ≤ β2 , . . . , αk ≤ βk .

1.2.10 Proposition
(i) X(α) ⊆ X(β) if and only if α ≤ β.
(ii) pα |X(β) 6= 0 if and only if α ≤ β.

Proof. Let α ≤ β. Then clearly, D(α)◦ ⊆ D(β)◦ and hence, by (1.2.7),


X(α) ⊆ X(β). Further, we observe that D(α) contains the matrix Eα
whose α-th minor is the identity matrix and all other entries are zero.

Since pα (Eα ) = ±1, we have pα X(α) 6= 0, and therefore, pα X(β) 6=
0 if α ≤ β. This proves the “if” part of both (i) and (ii). To prove
the reverse implication in both the cases, first we note that
in view of
pα X(α) 6= 0, it is enough to show that α 6≤ β implies pα X(β) = 0. So
assume that α 6≤ β. Then, there is an integer m such that α1 ≤ β1 ,
α2 ≤ β2 , . . . , αm−1 ≤ βm−1 and αm > βm . Now if Aα is the α-th minor
of a matrix A in D(β)◦ , then it is easy to see that Aα is of the form
 
C ∗
A=
0 ∗
where C is the leading m × m minor of Aα ; moreover, since A ∈ D(β)◦
and αm > βm , the last row of C consists of zeros. Thus
pα π(A) = ± det Aα = 0,

which shows that pα π(D(β)◦ ) = 0. Hence, using (1.2.7), it follows that

pα X(β) = 0. Q.E.D.
1
the proof is given at the end of Chapter 1.
12 1. Schubert Varieties in the Grassmannian

1.3 Standard monomials

Standard monomials facilitate a concrete description of the homogeneous


coordinate ring of a Schubert variety. To be precise, we shall show that
the graded components of this ring have a basis consisting of standard
monomials. Subsequent to this, we describe the defining equations of
the Grassmannian Gr,n in P(∧r ∨). In fact, we shall see that the ideal
of Gr,n is generated by quadratic polynomials which can be explicitly
described.

1.3.1 Definitions
(i) A tableau in I(r, n) (or a tableau on Gr,n ) of length m is a se-
quence (α(1), α(2), . . . , α(m)) of elements in I(r, n). We say that
the tableau is standard, if α(1) ≥ α(2) ≥ · · · ≥ α(m).
(ii) A standard monomial on Gr,n of length m is a formal expression
of the form pα(1) pα(2) · · · pα(m) , where pα(i) are Plücker coordi-
nates and (α(1), α(2), . . . , α(m)), is a standard tableau. Thus two
standard monomials are distinct, if the corresponding standard
tableaus are distinct.
(iii) A standard monomial of length m on a Schubert variety X(α) is
a formal expression pα(1) pα(2) · · · pα(m) , with α ≥ α(1) ≥ α(2) ≥
· · · ≥ α(m), where pα(i) are Plücker coordinates restricted to X(α).
Note that considering the above formal expression as a product of
Plücker coordinates in the homogeneous coordinate ring R(α) of
X(α), a standard monomial on X(α) defines an element of R(α),
and abuse of language, we call this element also a standard mono-
mial on X(α) (also see remark (3.2.i) below).

1.3.2 Remark
(i) It is not a priori clear that distinct standard monomials on X(α)
(as formal expressions) define distinct elements of the homoge-
neous coordinate ring. However, this is in fact the case, as we shall
see in (3.3) below. Thus we can identify a standard monomial (as
a formal expression) on X(α) with the corresponding element in
the homogeneous coordinate ring of X(α).
(ii) Since the definition of a standard monomial depends only on the
indices α(i) in the partially ordered set I(r, n), we may define stan-
1.3. Standard monomials 13

dard monomials more generally in an arbitrary set of elements of


a ring which are indexed by a partially ordered set. In particular,
if S is a quotient ring of the polynomial ring S = k[xα , α ∈ I(r, n)]
and xα is the canonical image of xα in S, then we may define stan-
dard monomials in xα exactly in the same way as above. As before,
a standard monomial in xα defines an element of S, and if there
is no confusion, we call this element also a standard monomial in
xα .

1.3.3 Proposition
For every m ≥ 0, distinct standard monomials of length m on a Schu-
bert variety X(α) define linearly independent elements (over k) in the
homogeneous coordinate ring of X(α). (By abuse of language we say
that distinct standard monomials in the homogeneous coordinate ring
of X(α) are linearly independent.) In view of (1.2.5), the same holds, in
particular, for the standard monomials on Gr,n .

Proof. We apply induction on the dimension of X(α). Let dim X(α) =


P
0, i.e., αi − r(r+1)
2 = 0. Then α = (1, 2, . . . , r), and using (1.2.10,(ii)),
it follows that pα is the only Plücker coordinate which does not vanish
on X(α). Therefore, for any m ≥ 0, pm α is the only standard monomial
of length m on X(α). Thus the proposition is true when dim X(α) = 0.
Now suppose that dim X(α) > 0 and the proposition is true for all
Schubert varieties X(β) with dim X(β) < dim X(α). Let F1 , F2 , . . . , Fd
be distinct standard monomials of length m ≥ 1 on X(α) and suppose
that they satisfy a nontrivial dependence relation

d
X
ci Fi = 0. (1.3)
i=1

Clearly, there is no loss of generality in assuming that ci 6= 0 for all i.


Moreover, we may assume that at least one Fi does not start with pα .
For if all of them start with pα , we can cancel a power of pα in (1.3)
to achieve the desired situation, since the homogeneous coordinate ring
of X(α) is an integral domain (1.2.6) and pα 6= 0 on X(α). Thus there
is a standard monomial Fj such that Fj = pβ G, where α > β and G
is a standard monomial of length m − 1. Restricting (1.3) to X(β) we
get a nontrivial dependence relation among distinct standard monomials
14 1. Schubert Varieties in the Grassmannian

on X(β). Since dim X(β) < dim X(α), this contradicts the induction
hypothesis. Hence F1 , F2 , . . . , Fd must be linearly independent.
Q. E. D.
Before going further into the main results of this section we lay down
some preliminaries to be used in the sequel. Let I = (1, 2, . . . , n) and
let I r = |I × .{z
. . × I} . We know that the indeterminates xα in the homo-
r times
geneous coordinate ring of P(∧r ∨) are indexed by the set I(r, n) ⊂ I r .
However, for the sake of convenience, we extend our indexing set I r and
define xα for any α in I r as follows: For any α = (α1 , α2 , . . . , αr ) in
I r with αi ’s distinct, there is a permutation σ in Sr (the permutation
group on r symbols) such that ασ = (ασ1 , . . . , ασr ) is in the increasing
order, where ασi denotes the image of αi under σ. Now for any α in I r
we define

0, if αi ’s are not distinct
xα =
( sgn σ)xασ , if αi ’s are distinct
Thus xα make sense for all α in I r . A similar definition can be given for
the Plücker coordinates pα so that pα make sense for all α in I r .
Let M {µ1 , µ2 , . . . , µp } and N = {v1 , v2 , . . . , vq } be two sets. Then a
shuffle permutation or a shuffling of M and N is a partition of M ∪ N
into a subset of p elements and its complement. We denote by S(p, q)
the set of all shufflings of a set of cardinality p and a set of cardinality
q. It is easy to see that S(p, q) can be identified with Sp+q /Sp × Sq .

1.3.4 Lemma
Fix integers k and ℓ with 1 < k < ℓ < r. Then for α, β in I(r, n), we
have
X h
sgn (σ)p(α1 ,α2 ,...,αk ,ασk+1 ,...,ασr ) ×
σ∈S(r−k,ℓ)
i
p(β1σ ,β2σ ,...,βℓσ ,βℓ+1 ,...,βr ) = 0, (1.4)

where the summation runs over all the shufflings of {αk+1 , . . . , αr } and
{β1 , β2 , . . . , βℓ }.

1.3.4.1 Remark
Note that (α1 , . . . , αk , ασk+1 , . . . , ασr ) and (β1σ , . . . , βℓσ , βℓ+1 , . . . , βr ) may
not be in I(r, n). However, the sum does make sense, as we have defined
1.3. Standard monomials 15

xα for all α in I r .

Proof2 of the lemma. Let F denote the left hand side of (1.4). It is
clear that f is a multilinear form in (r − k)+ ℓ row vectors corresponding
to the indices αk+1 , . . . , αr , β1 , β2 , . . . , βℓ . We claim that F is also skew-
symmetric in these vectors, i.e., if a set of these row vectors contains two
identical rows, then F vanishes on it. In fact, if the indices corresponding
to the identical rows are either in {ασk+1 , . . . , ασr } or in {β1σ , . . . , βℓσ } for
some σ, then

pσ = sgn(σ)p(α1 ,α2 ,...,αk ,ασk+1 ,...,ασr ) p(β1σ ,...,βℓσ ,βℓ+1,...,βr )

vanishes on the given set of row vectors containing two identical rows.
On the other hand, if one of the indices is in {ασk+1 , . . . , ασr } and the
other in {β1σ , . . . , βℓσ } for some σ in S(r − k, ℓ), then it is clear that
by interchanging these two indices we get a term pλ of the sum (1.4),
whose value at the given set of row vectors cancels with that of pσ .
Thus F is a multilinear skew-symmetric form on (r − k) + ℓ vectors of
an r-dimensional vector space. Since (r − k) + ℓ > r, F must vanish
identically. Q.E.D.
Besides the partial order ≤, defined in (1.2.9), I(r, n) also carries the
lexicographic order which we denote by ≤ℓ to distinguish it from ≤ .
This in turn gives for every m > 0, a lexicographic order on I(r, n)m =
I(r, n) × . . . × I(r, n), the set of all tableau of length m. We denote this
| {z }
m times
lexicographic order also by ≤ℓ .

1.3.5 Lemma
Let α, β ∈ I(r, n) with α 6≤ β. Then, we can write
X
pα pβ = ±pα′ pβ ′
α′ ,β ′

with α′ < α so that (α′ , β ′ ) <ℓ (α, β),

Proof. Since α 6≤ β, there is an integer k such that α1 ≤ β1 , α2 , ≤


β2 , . . . , αk ≤ βk and αk+1 > βk+1 . Thus we have

β1 < β2 < · · · < βk+1 < αk+1 < αk+2 < · · · < αr .
2
This proof was suggested by Steinberg
16 1. Schubert Varieties in the Grassmannian

If σ is any nontrivial shuffling of {αk+1 , αk+2 , . . . , αr } and {β1 , β2 , . . . ,


βk+2 }, then it follows from the above inequalities that after rearrang-
ing in the increasing order {α1 , α2 , . . . , αk , ασk+1 , . . . , ασr } < α. Since the
previous lemma for ℓ = k + 1 gives
X h
sgn (σ)p(α1 ,α2 ,...,αk ,ασk+1 ,...,ασr ) ×
σ∈S(r−k,k+1)
i
p(β1σ ,β2σ ,...,βk+1
σ ,βk+2 ,...,βr ) = 0,
P
It is clear that we can write pα pβ as ±pα′ pβ ′ with α′ < α. Q.E.D.
α′ ,β ′

1.3.6 Proposition
Let Rm denote the m-th graded component of the homogeneous coor-
dinate ring R of Gr,n . Then, for every m ≥ 0, Rm is generated as a
vector-space by the standard monomials of length m on Gr,n .

Proof. Let (τi ) = (τ1 , τ2 , . . . , τm ) denote an element of I(r, n)m , where


τi ∈ I(r, n). There is a map Φ : I(r, n)m → Rm defined by Φ((τi )) =
pτ1 pτ2 . . . pτm . Clearly, it suffices to show that for each ((τi )) in I(r, n)m ,
Φ((τi )) is a linear combination of standard monomials in Rm . To show
this, we shall use induction with respect to the lexicographic order on
I(r, n)m .
The smallest element in I(r, n)m is (αmin , αmin , . . . , αmin ), where
αmin = (1, 2, . . . , r). We have Φ((αmin , αmin , . . . , αmin )) = pmαmin
, which
is a standard monomial. Now let F = pτ1 pτ2 . . . pτm be a non-standard
monomial in Rm . Then, there is a integer k such that τ1P≤ τ2 ≤ · · · ≤ τk
but τk 6≤ τk+1 . Using (3.5) we can write pτk pτk+1 = ±pα′ pβ ′ with
α′ ,β ′
α′ < τk . Substituting this expression in F we obtain
X X
F = ±pv1 pv2 . . . pvm = Φ((vi ))
(vi )∈I(r,n)m (vi )∈I(r,n)m

where each (vi ) <ℓ (τi ). Therefore, by induction, each Φ((vi )) is a linear
combination of standard monomials in Rm and hence, so is F. Q. E. D.

1.3.7 Theorem
Let Rα be the homogeneous coordinate ring of a Schubert variety X(α).
Then, for every m ≥ 0, its m-th graded component (Rα )m has a basis
1.3. Standard monomials 17

consisting of standard monomials of length m on X(α). In particular,


the same holds for the homogeneous coordinate ring R of Gr,n .

Proof. In view of the natural surjective homomorphism R → Rα map-


ping each Plücker coordinate pλ to pλ X(α) , the theorem immediately
follows from the preceding proposition and (3.3) Q. E. D.

1.3.8 Remark
Let S̄ = k[x̄α , α ∈ I(r, n)] be a quotient ring of the polynomial ring
S = k[xα , α ∈ I(r, n)]. As it make sense to talk about standard mono-
mials in xα (1.3.2,(ii)), it is natural to ask when in general, S̄ is generated
by standard monomials. The proofs of the above results suggest a suf-
ficient condition for this. The relation (1.3.4), satisfied by the Plücker
coordinates is a crucial ingredient in the proof of (1.3.6). We used it
to prove (1.3.5) which allowed us to use the lexicographic induction on
I(r, n)m in (1.3.6). After seeing these proofs, a moment’s reflection will
convince the reader that a quotient ring S̄ of S is generated by stan-
dard monomials, if x̄α satisfy relations of the form (1.3.4) or, in fact of
the form as in (1.3.5). We include this observation separately in (1.3.9)
below for further reference.

1.3.9 Proposition
Let S̄ = k[x̄α , α ∈ I(r, n)] be a quotient ring of S = k[xα , α ∈ I(r, n)]
by a homogeneous ideal P. Then each graded component S̄m of S̄ is
generated by standard monomials in x̄α , if P contains all the polynomials
of the form
X
(i) sgn(σ)x(α1 ,...,αk ,ασk+1 ,...,ασr ) x(β1σ ,...,βℓσ ,βℓ+1,...,βr ) , where α, β ∈
σ∈S(r−k,ℓ)
I(r, n) and k, ℓ are fixed integers with 1 < k < ℓ < r.

(ii) or, in fact, of the form as in Lemma (1.3.5).

Proof. See remark (1.3.8) above.

Now we are ready to describe the ideal of the Grassmannian Gr,n .


18 1. Schubert Varieties in the Grassmannian

1.3.10 Theorem
The homogeneous ideal J of Gr,n is generated by the homogeneous poly-
nomials of the form:
X
(∗) sgn(σ)x(α1 ,...,αk ,ασk+1 ,...,ασr ) x(β1σ ,...,βℓσ ,βℓ+1,...,βr ) ,
σ∈S(r−k,ℓ)

where k and ℓ are fixed integers with 1 < k < ℓ < r, and for every α, β in
I(r, n) the above sum runs over all the shufflings of {αk+1 , αk+2 , . . . , αr }
and {β1 , β2 , . . . , βℓ }.

Proof. Consider the homogeneous coordinate rings S = k[xα , α ∈


I(r, n)] and R = k[pα , α ∈ I(r, n)] of P(∧r ∨) and Gr,n respectively.
We have a natural homomorphism
Φ : S −→ R
xα 7−→ pα
whose kernel is the ideal J. If J ′ is the ideal generated by the polynomials
of the form (∗), then (1.3.4) implies that J ′ ⊂ ker Φ = J. Hence we have
a surjective homomorphism
Φ : S/J ′ −→ R
xα 7−→ pα .
Thus, in order to prove that J ′ = J, it is enough to prove that Φ is
injective. So let F be any nonzero element of S/J ′ . By (1.3.9) S/J ′ is
generated by standard monomials, and therefore F can be written as
a linear combination of distinct standard monomials. This shows that
Φ(F ) is a linear combination of distinct standard monomials on Gr,n .
Since standard monomials on Gr,n are linearly independent, it follows
that Φ(F ) 6= 0 and hence, J ′ = J. Q.E.D.

1.3.11 Remark
If the field k is of characteristic zero, then a set of generators for J
can be described in a simpler way, namely, J is then generated by the
polynomials of the form
X
x(α1 ,...,αs−1 ,βj ,αs+1 ,...,αr ) x(β1 ,...,βj−1 ,αs ,βj+1 ,...,βr ) ,
j=1

where α, β ∈ I(r, n) and s is a fixed integer with 1 ≤ s ≤ r. This is false


if the characteristic of k is nonzero (see [A]).
1.4. Some Applications 19

1.3.12 Proposition
Let Iβ denote the ideal of a Schubert variety X(β) in R = k[pα , α ∈
I(r, n)]. Then {pα | α 6≤ β} is a set of generators for Iβ , i.e., Iβ is
generated by the Plücker coordinates vanishing on X(β).

Proof. The proof is quite similar to that of (1.3.10). Let I ′ be the


ideal generated by the set {pα |α 6≤ β}, and let Rβ be the homogeneous
coordinate ring of X(β). Then we have a natural homomorphism

Φ : R/I ′ −→ Rβ

pα 7−→ pα X(β)

which is surjective. Since R is generated by standard monomials, so is


R/I ′ . Therefore, it follows exactly in the same way as in (1.3.10) that Φ
is injective and hence, I ′ = Iβ . Q.E.D.

1.3.13 Corollary
The homogeneous ideal of a Schubert variety X(β) in k[xα , α ∈ I(r, n)]
is J + Jβ , where J is the homogeneous ideal of the Grassmannian and
Jβ is the ideal generated by {xα | α 6≤ β}.

Proof. Obvious.

1.4 Some Applications

In this section we give some applications of the standard monomial the-


ory. To begin with we strengthen some of the results of the previous
section to the scheme theoretic unions of Schubert varieties and de-
scribe Pieri’s formula which gives the hyperplane section of a Schubert
variety. This strengthening is required in order to prove the cohomology
vanishing theorems for Schubert varieties. We conclude the section by
proving that Schubert varieties are arithmetically Cohen-Macaulay and
projectively normal.

1.4.1 Definition
We call a subset T of I(r, n) a left half space (LHS), if

α ∈ T, β ∈ I(r, n) with β ≤ α ⇒ β ∈ T.
20 1. Schubert Varieties in the Grassmannian

Similarly, we call T a right half space (RHS), if

α ∈ T, β ∈ I(r, n) with α ≤ β ⇒ β ∈ T.

1.4.2 Example
For every α in I(r, n), the set Tα = {λ ∈ I(r, n) | λ ≤ α} is a LHS and
its complement Tα′ = {λ ∈ I(r, n) | λ 6≤ α} is a RHS. In view of (1.2.10),
it is clear that

Tα = {λ ∈ I(r, n) | pλ X(α) 6= 0} and

Tα′ = {λ ∈ I(r, n)|pλ X(α) = 0}.

The following lemma is an immediate consequence of the above def-


inition.

1.4.3 Lemma
(i) T is a LHS if and only if its complement is a RHS.

(ii) If T1 , T2 are LHS’s (respectively RHS’s), then so are T1 ∪ T2 and


T1 ∩ T2 .

(iii) A LHS (respectively RHS) is characterized by its maximal (re-


spectively minimal) elements, i.e., if µ(1), µ(2), . . . , µ(m) are the
maximal elements of a LHS T, then
[
T = Tµ(i) .
1≤i≤m

A similar description of T can be given, if it is a RHS.

Proof. Immediate.

Let X1 , X2 , . . . , Xm be closed subschemes of a projective scheme Y ,


whose scheme structures are given by the homogeneous ideals IX1 , IX2 ,
. . . , IXm respectively in the homogeneous coordinate ring of Y. We re-
call that the scheme theoretic union (resp. intersection) of Xi ’s is by
definition, the subscheme of Y whose scheme structure is given by the
T P
m
homogeneous ideal 1≤i≤m IXi (respectively IXi ). In what follows,
i=1
1.4. Some Applications 21

by a union (resp. intersection) of schemes we shall mean the scheme


theoretic union (resp. intersection) unless otherwise mentioned.
For a subset T of I(r, n), let X(T ) denote the union of Schubert
varieties X(α), α ∈ T. Note that X(T ), being a union (scheme theoretic)
of reduced schemes, is reduced.

1.4.4 Proposition

The map T → X(T ) establishes a one to one correspondence between


the collection of LHS’s in I(r, n) and the collection of unions of Schubert
varieties in Gr,n .

Proof. To show that T → X(T ) is a bijection, we define a map P in


the opposite direction: if Y is a union of Schubert varieties, the we set

P (Y ) = {α ∈ I(r, n) | pα |Y 6= 0

which is clearly a LHS.



Since pα X(α) =6 0, for any LHS T we have pα X(T ) 6= 0 for α ∈ T.
This shows
that T ⊂ P (X(T )). On the other hand, if α ∈ P (X(T )),
then pα X(T ) 6= 0 and therefore pα X(β) 6= 0 for some β in T. Now from
(1.2.10) it follows that α ≤ β, and since T is a LHS, we have α ∈ T.
Thus P (X(T )) = T. Using similar arguments, we can also show that
X(P (Y )) = Y for a union of Schubert varieties Y. This proves that
T → X(T ) is a bijection. Q.E.D.
As mentioned earlier, in order to prove our vanishing theorems we
need to generalize the results of the previous section for a Schubert
variety to a union of Schubert varieties. The required generalizations,
as we shall see below, follow as easy consequences of the previous results.
For an LHS T in I(r, n), let R(T ) denote the homogeneous coordinate
ring of X(T ). If we denote the restriction of a Plücker coordinate pα to
X(T ) also by pα , then a standard monomial of length m on X(T )
(or a standard monomial in T ) is by definition, a formal expression
p = pα(1) , pα(2) . . . pα(m) with α(1) ≥ α(2) ≥ . . . ≥ α(m), where α(i) ∈ T,
1 ≤ i ≤ m; p defines an element of R(T ) and as usual, we call this
element also a standard monomial of length m on X(T ).
22 1. Schubert Varieties in the Grassmannian

1.4.5 Proposition
(i) For every m ≥ 0, distinct standard monomials of length m on
X(T ) are linearly independent elements (over k) of R(T )m (the
m-th graded component of R(T )).

(ii) For every m ≥ 0, R(T )m is generated by standard monomials.

Proof.

(i) Suppose that F1 , F2 , . . . Fd are distinct standard monomials in


R(T )m such that
Xd
ci Fi = 0, (1.5)
i=1

where ci ∈ k with at least one ci , say c1 , nonzero. Since F1 is


a standard monomial in T, there exists an α ∈ T such that the
restriction of F1 to X(α) (i.e. image of F1 under the canonical
map R(T ) → Rα ) is a standard monomial on X(α). But then the
restriction of (1.5) to X(α) gives a linear relation among distinct
standard monomials on X(α), a contradiction to (1.3.3). Hence,
F1 , F2 , . . . , Fd must be linearly independent.

(ii) Consider the canonical surjective map R → R(T ), where R is the


homogeneous coordinate ring of Gr,n . The assertion immediately
follows using (1.2.10, ii) and (1.3.6). Q.E.D

For a LHS T, let I(T ) denote the ideal in R = k[Pα , α ∈ I(r, n)]
generated by pα with α 6∈ T. In other words, I(T ) is the ideal generated
by the Plücker coordinates vanishing on X(T ). The following lemma
says that I(T ) is in fact the ideal of X(T ) in R.

1.4.6 Lemma
I(T ) = IX(T ) , i.e. the ideal (in R) of a union of Schubert varieties is
generated by the Plücker coordinates vanishing on it.

Proof. The lemma obviously follows, if we show that the natural map
Φ : R/I(T ) → R(T ) is an isomorphism. In fact, since Φ is surjective, we
have only to show that it is injective. So let F ∈ R/I(T ) with F 6= 0,
1.4. Some Applications 23

where F ∈ R and − denotes the natural map R → R/I(T ). Since R is


generated by standard monomials (see (3.6)), we can write
m
X
F = ci Fi ,
i=1

where ci ∈ k and F i’s are distinct standard monomials in R. Further,


since F 6= 0, we may assume that c1 6= 0 and F1 6∈ I(T ) for all i. This
implies that the index α of any pα appearing in Fi , 1 ≤ i ≤ m, lies in
T, which shows that Φ(F i ) is a standard monomial on X(T ) for all i.
P
m
Thus Φ(F ) = ci Φ(F i ) is a nontrivial linear combination of distinct
i=1
standard monomials on X(T ), which, in view of (1.4.5,(i)), must be
nonzero. This proves that Φ is injective. Q. E. D.

1.4.7 Corollary
Let T1 , T2 be LHS’s. Then

(i) I(T1 ∩ T2 ) = I(T1 ) + I(T2 )

(ii) I(T1 ∪ T2 ) = I(T1 ) ∩ I(T2 ).

Proof. (i) immediately follows from the definition of I(T ). To prove


(ii), first we note that if Iα is the ideal of X(α) in R, then for any LHS
T we have
IX(T ) = ∩α∈T Iα .
since X(T ) is the scheme theoretic union of X(α), α ∈ T. Therefore,
using (4.3) and (4.6) it follows that

I(T1 ∪ T2 ) = IX(T1 ∪T2 )


\
= Iα
α∈T1 ∪T2
 \   \ 
= Iα ∩ Iα
α∈T1 α∈T2
= IX(T1 ) ∩ IX(T2 )
= I(T1 ) ∩ I(T2 ).

Q.E.D.
24 1. Schubert Varieties in the Grassmannian

1.4.8 Corollary
The correspondence T 7→ X(T ) preserves unions and intersections, i.e.,
for any LHS’s T1 , T2 , we have

(i) X(T1 ∪ T2 ) = X(T1 ) ∪ X(T2 )

(ii) X(T1 ∩ T2 ) = X(T1 ) ∩ X(T2 )

Proof. As before, let IX denote the ideal in R defining a subscheme X


of Gr,n . Then (i) follows from the equality

IX(T1 ∪T2 ) = IX(T1 ) ∩ IX(T2 )

which we have proved in the previous corollary. As for (ii), we have

IX(T1 ∩T2 ) = I(T1 ∩ T2 )


= I(T1 ) + I(T2 )
= IX(T1 ) + IX(T2 )
= IX(T1 ) ∩ X(T2 )

which implies (ii). Q.E.D.


As in (1.4.2), for α ∈ I(r, n) let

Tα = {λ ∈ I(r, n) | λ ≤ α}

which is the LHS indexing the Plücker coordinates not vanishing on


X(α).

1.4.9 Lemma
Let α, β ∈ I(r, n). Then

(i) Tα ∩ Tβ = Tγ for some γ ∈ I(r, n).

(ii) Tα − {α} = ∪i Tµ(i) , where µ(i)’s are the maximal elements of


Tα − {α}.

(iii) If µ and µ′ are any two maximal elements of Tα − {α}, then X(µ)
is of codimension 1 and X(µ) ∩ X(µ′ ) is of codimension 2 in X(α).
1.4. Some Applications 25

Proof.

(i) If α = (α1 , α2 , . . . , αr ) and β = (β1 , β2 , . . . , βr ) then it is easy to


see that γ = (γ1 , γ2 , . . . , γr ), where γi = min(αi , βi ), 1 ≤ i ≤ r.

(ii) Since α is the maximal element of Tα , Tα − {α} is a LHS, and


hence (ii) follows from (1.4.3).

(iii) If µ = (µ1 , µ2 , . . . , µr ) is a maximal element of Tα − {α}, it differs


from α = (α1 , α2 , . . . , αr ) just at one place by 1. Hence, using the
dimension formula for a Schubert variety (1.2.5,ii), the assertions
in (iii) follow easily.

To prove the remaining assertion, first we note that for any α in


I(r, n), X(α) = X(Tα ). Now using (1.4.8) and (i) above we have

X(µ) ∩ X(µ′ ) = X(Tµ ) ∩ X(Tµ′ )


= X(Tµ ∩ Tµ′ )
= X(Tν )
= X(ν),

where νi = min(µi , µ′i ), 1 ≤ i ≤ r.


Clearly, ν = (ν1 , ν2 , . . . , νr ) differs from α = (α1 , α2 , . . . , αr ) by 1
at two places. Therefore, using the dimension formula (1.2.5,ii), the
assertion follows. Q.E.D.

1.4.10 Corollary
The intersection (scheme theoretic) of two Schubert varieties is again a
Schubert variety.

Proof. Since X(α) = X(Tα ) for any Schubert variety X(α), the corol-
lary immediately follows from (1.4.8,ii) and (i) of the above lemma.
Q.E.D.

1.4.11 Pieri’s formula


The hyperplane section (up to linear equivalence) of a Schubert variety
X(α) is the union Schubert varieties of codimension 1 in X(α), and so
26 1. Schubert Varieties in the Grassmannian

in particular, it is reduced. Explicitly, if H is the hyperplane in Gr,n


defined by pα = 0, then
[
H ∩ X(α) = X(Tα − {α}) = X(µ(i)) (scheme theoretically),
i

where µ(i)’s are the maximal elements of Tα − {α}.

Proof. Since X(α) = X(Tα ), the ideal of X(α) is I(Tα ) (1.4.6), and
therefore the ideal of H ∩X(α) is (pα )+I(Tα ) which is the ideal generated
by pα and the Plücker coordinates pλ with λ 6∈ Tα . This shows that the
ideal of H ∩X(α) is I(Tα −{α}). Now Pieri’s formula immediately follows
using (4.7.ii) and (4.9.ii). Q. E. D.

1.4.12 Remark
The formula gives an ideal-theoretic strengthening of the intersection
formula for [H] · [X(α)], say in the cohomology (or Chow) ring of Gr,n ,
where [H] and [X(α)] represent the cohomology classes defined by H
and X(α) respectively.
For a LHS T in I(r, n) and a nonnegative integer m, let S(T, m)
denote the number of standard monomials on X(T ) of length m. Note
that in view of (4.5), S(T, m) is nothing but dim R(T )m , where R(T )
is the homogeneous coordinate ring of X(T ).

1.4.13 Lemma
(i) Let T1 and T2 be LHS’s in I(r, n). Then for every nonnegative
integer m, we have

S(T1 ∪ T2 , m) = S(T1 , m) + S(T2 , m) − S(T1 ∩ T2 , m)

(ii) For α ∈ I(r, n) and any nonnegative integer m,

S(Tα , m) = S(Tα , m − 1) + S(Tα − {α}, m)

Proof.

(i) Obvious.
1.4. Some Applications 27

(ii) Among the standard monomials of length m in Tα , those beginning


with pα are S(Tα , m − 1) in number, where as those not beginning
with pα are S(T − {α}, m) in number. Q.E.D.

In the proof of the main theorem on the cohomology of a Schubert


variety, we require the following technical result which says that the as-
sertions of the main theorem would hold for a union of Schubert varieties,
provided it holds for every Schubert variety contained in the union.

1.4.14 Lemma
Let T be a LHS and let X = X(T ) be the corresponding union of
Schubert varieties. If OX (1) denotes the hyperplane line bundle on X
corresponding to the Plücker embedding X → P = P(∧r ∨). then

(i) H 0 (X, OX (m)) ∼


= R(T )m for every m ≥ 0,

(ii) H i (X, OX (m)) = 0 for i > 0 and m ≥ 0,


provided (i) and (ii) hold for every Schubert variety X(α) with
α ∈ T. Furthermore, if X is a part of the hyperplane section of a
Schubert variety (i.e. each irreducible component of X is a com-
ponent of the hyperplane section), then

(iii) H i (X, OX (m)) = 0 for i > 0 with i 6= dim X, and m < 0 provided
(iii) holds for every Schubert variety X(α) with α ∈ T.

Proof. We shall use induction on |T |, the cardinality of T. If X is a


Schubert variety (in particular, if |T | = 1), then the lemma is trivial. So
assume that |T | > 1 and that X is not a Schubert variety. Then clearly,
T has more than one maximal element and using (1.4.3), we can write
T = T1 ∪ T2 , where Ti ’s are LHS’s with Ti 6= T, i = 1, 2.
Now, let X1 = X(T1 ), X2 = X(T2 ), and consider the following exact
sequence of sheaves:

0 → OX → OX1 ⊕ OX2 → OX1 ∩X2 → 0


(f, g) 7→ f − g

Twisting this sequence by m ≥ 0 we obtain

0 → OX (m) → OX1 (m) ⊕ OX2 (m) → OX1 ∩X2 (m) → 0


28 1. Schubert Varieties in the Grassmannian

which in turn gives the long exact cohomology sequence

0 → H 0 (X, OX (m)) → H 0 (X1 , OX1 (m)) ⊕ H 0 (X2 , OX2 (m)) →


Φ
→ H 0 (X1 ∩ X2 , OX1 ∩X2 (m)) → · · · (1.6)

Since X1 ∩ X2 = X(T1 ) ∩ X(T2 ) = X(T1 ∩ T2 ), by induction, (i) holds for


X1 ∩ X2 . This implies that the natural map H 0 (P, OP (m)) → H 0 (X1 ∩
X2 , OX1 ∩X2 (m)) is surjective and hence, so is the map Φ in the sequence
(1.6).
Thus, writing h0 (∗, O∗ (m)) for dim H 0 (∗, O∗ (m)), (1.6) gives

h0 (X, OX (m)) = h0 (X1 , OX1 (m)) + h0 (X2 , OX2 (m))


−h0 (X1 ∩ X2 , OX1 ∩X2 (m)). (1.7)

Furthermore, since (i) holds for X1 , X2 and X1 ∩ X2 , we have

h0 (X1 ∩ X2 , OX1 ∩X2 (m)) = S(T1 ∩ T2 , m) and


0
h (Xi , OXi (m)) = S(Ti , m), i = 1, 2.

On the other hand, by (1.4.13), we have

S(T, m) = S(T1 , m) + S(T2 , m) − S(T1 ∩ T2 , m).

Hence, the equality (1.7) implies that

h0 (X, OX (m)) = S(T, m) which proves (i).

The assertion (ii) easily follows from the long exact sequence (1.6),
since Φ is surjective and (ii) holds for X1 , X2 and X1 ∩ X2 .
To prove (iii), first we note that we can choose T1 such that X1 =
X(T1 ) is an irreducible component of X. Now it follows from (1.4.9,iii)
that the irreducible components of X1 ∩ X2 are of codimension 1 in X1 .
Therefore dim X1 ∩ X2 = dim X − 1 and by (1.4.11), X1 ∩ X2 is a part
of the hyperplane section of X1 . Thus by induction (iii) holds for X1 , X2
and X1 ∩ X2 . Now, for 0 ≤ i < dim X, (iii) easily follows from the exact
sequence (1.6). Since, for i > dim X, (iii) is a consequence of the Serre-
Grothendieck vanishing theorem ([H], III, Theorem 2.7), we are done.
Q.E.D.
1.4. Some Applications 29

1.4.15 Theorem
Let X = X(α) be a Schubert variety and let Rα = ⊕m≥0 (Rα )m be
its homogeneous coordinate ring. If OX (1) denotes the hyperplane line
bundle on X corresponding to the Plücker embedding X → P = P(∧r ∨),
then
(i) H 0 (X, OX (m)) ∼
= (Rα )m for every m ≥ 0.
(ii) H i (X, OX (m)) = 0 for i > 0 and m ≥ 0.
(iii) H i (X, OX (m)) = 0 for i > 0 with i 6= dim X and all m < 0.

Proof. We shall use induction on dim X. If dim X = 0, the theorem is


trivial. So assume that dim X > 0. Let H be the hyperplane section of
X defined by the Plücker coordinate pα = 0. For every integer m, we
have the following exact sequence of line bundles:
0 → OX (m − 1) → OX (m) → OH (m) → 0,
which gives the cohomology exact sequence
Ψ
0 → H 0 (X, OX (m − 1)) → H 0 (X, OX (m)) → H 0 (H, OH (m))
→ H 1 (X, OX (m − 1)) → · · · → H i−1 (H, OH (m)) →
H i (X, OX (m − 1)) → H i (X, OX (m)) → · · · . (1.8)
Now by Pieri’s formula, H = X(Tα − {α}) which is the scheme theo-
retic union of Schubert varieties of codimension one in X(α). Therefore
by induction, the hypotheses of the preceding lemma are satisfied for H,
and consequently, (i), (ii), (iii) hold for H. But statement (i) for H im-
plies that for m ≥ 0, the natural map H 0 (P, OP (m)) → H 0 (H, OH (m))
Ψ
is surjective; hence, so is the map H 0 (X, OX (m)) → H 0 (H, OH (m)).
Thus, from the sequence (1.8), we obtain the exact sequence
O → H 0 (X, OX (m − 1)) → H 0 (X, OX (m)) → H 0 (H, OH (m)) → 0
(1.9)
which gives
h0 (X, OX (m)) = h0 (X, OX (m − 1)) + h0 (H, OH (m)), (1.10)
where h0 (∗, O∗ (−)) denotes dim H 0 (∗, O∗ (−)). Since (i) holds for H, we
have h0 (H, OH (m)) = S(Tα − {α}, m), and therefore, using (1.4.3,ii), we
obtain
h0 (H, OH (m)) = S(Tα , m) − S(Tα , m − 1).
30 1. Schubert Varieties in the Grassmannian

This together with the equality (1.10) implies

h0 (X, OX (m)) − S(Tα , m) = h0 (X, OX (m − 1)) − S(Tα , m − 1)

i.e., h0 (X, OX (m)) − dim(Rα )m = h0 (X, OX (m − 1)) − dim(Rα )m−1 .


Since, by Serre’s theorem, the left hand side is zero for sufficiently
large m, it follows by descending induction on m that h0 (X, OX (m)) =
dim(Rα )m for all m ≥ 0. This proves (i).
Since (ii) and (iii) hold for H and the map Ψ in the exact se-
quence (1.8) is surjective, it follows that O → H i (X, OX (m − 1)) →
H i (X, OX (m)) is exact for i > 0, m ≥ 0, and for i > 0, i 6= dim X,
m < 0. But H i (X, OX (m)) = 0 for i > 0 and sufficiently large m.
Hence, (ii) and (iii) hold for X. Q.E.D.

1.4.16 Theorem
The assertions (i) and (ii) of the above theorem hold for any union
of Schubert varieties. Moreover, if every irreducible component of the
union is a component of the hyperplane section of a Schubert variety,
then assertion (iii) also holds.

Proof. Immediate from (1.4.14) and (1.4.15).


In proving that Schubert varieties are arithmetically Cohen-Macaulay,
we shall require the following result which is a consequence of the local
cohomology theory.

1.4.17 Proposition
Let Y be a closed subscheme of a projective space P. Let R be the
homogeneous coordinate ring of Y and let Ŷ = Spec R, the cone over
Y. If A denotes the local ring at the vertex of Ŷ , then the following
conditions are equivalent:

(i) (a) The natural map H 0 (P, OP (m)) → H 0 (Y, OY (m)) is surjective
for all m ≥ 0.
(b) H i (Y, OY (m)) = 0 for 0 < i ≤ d − 2 and all m ∈ Z.

(ii) Depth A ≥ d.

(see [G], section 3.)


1.4. Some Applications 31

1.4.18 Lemma
Let Y be a closed subscheme of a projective space. The cone Ŷ over Y
is Cohen-Macaulay if and only if Ŷ is Cohen-Macaulay at its vertex.

Proof. Suppose that Ŷ is Cohen-Macaulay at its vertex (0). By [G-


D], 12.1.1, Ŷ is Cohen-Macaulay in a neighborhood U0 of (0). If p :
Ŷ − {(0)} → Y is the natural projection, then it is clear that U0 ∩
p−1 (y) ∪ {(0)} is an open dense subset of p−1 (y) ∪ {(0)} for all y in
Y, and hence, each fibre p−1 (y) meets U0 . But each fibre is simply a
Gm -orbit under the natural action of Gm on Ŷ − {(0)}, which implies
that Ŷ is Cohen-Macaulay at each point of p−1 (y) for all y in Y. This
proves that Ŷ is Cohen-Macaulay. Q.E.D.

1.4.19 Theorem
Schubert varieties are arithmetically Cohen-Macaulay.

Proof. Let X = X(α) be a Schubert variety. In view of the preceding


lemma, it suffices to show that the cone X̂ over X is Cohen-Macaulay
at the vertex. To this end, first we observe that (1.4.15,i) implies the
condition (a) of (1.4.17), namely, H 0 (P, OP (m)) → H 0 (X, OX (m)) is
surjective for all m ≥ 0. Now if A denotes the local ring of X̂ at the
vertex, then by (1.4.15) and (1.4.17), we have

Depth A ≥ dim X + 1.

But
Depth A ≤ dim A = dim X + 1.

Hence, A is Cohen-Macaulay. This proves the theorem.

Remark There is another proof of the fact that X̂ is Cohen-Macaulay at


the vertex, which does not use the local cohomology theory (see [Mu]).
Now we proceed to show that Schubert varieties are projectively
normal.

1.4.20 Proposition
The singular locus of a Schubert variety is a union of Schubert varieties.
32 1. Schubert Varieties in the Grassmannian

Proof. Let B be the subgroup of the upper triangular matrices of


GLn (k). It follows from (1.2.4) that the action of B on Gr,n induces an
action on every Schubert variety X(α) and consequently, Sing X(α) (the
singular locus of X(α)) is B-stable. Since, by (1.2.4), B-orbits in Gr,n
are the Schubert cells, Sing X(α) must be a union of Schubert cells.
Now the proposition follows from the fact that Sing X(α) is a closed
subset of X(α). Q.E.D.

1.4.21 Lemma
Schubert varieties are smooth in codimension one.
Proof. Let X = X(α) be a Schubert variety. Since, by (1.4.20), its
singular locus Sing X is a union of Schubert varieties, it suffices to show
that none of the codimension one Schubert varieties in X is contained
in Sing X.
By Pieri’s formula, the subscheme H of X defined by pα = 0 is
reduced. Also observe that the irreducible components of H are precisely
the codimension one Schubert varieties Xi , 1 ≤ i ≤ m, of X. Since H
is reduced, ∃ xi ∈ Xi , 1 ≤ i ≤ m, such that xi is a smooth point of
H. Hence, the maximal ideal of the local ring OXi ,xi is generated by
(d − 1) elements, where d = dim X. Now the maximal ideal of OX,xi
is generated by pα and a set of generators of OXi ,xi lifted to OX,xi ;
consequently, it has d generators. Hence, xi is smooth in X. This proves
that Xi 6⊆ Sing X, 1 ≤ i ≤ m. Q.E.D.

1.4.22 Theorem
Schubert varieties are projectively normal.
Proof. Let X(α) be a Schubert variety. Since it is Cohen-Macaulay
(1.4.19) and smooth in codimension one (1.4.21), it is normal (see [Ma],
Theorem 39). Further, (1.4.15,(i)) implies that the map H 0 (P, OP (m)) →
H 0 (X(α), OX(α) (m)) is surjective for all m ≥ 0, which together with
normality of X(α) implies that X(α) is projectively normal (see [H], I,
Exercise 5.14). Q.E.D.

1.5 Degeneration of Schubert varieties

Consider the polynomial algebra P = k[xβ , β ∈ Tα ], where α ∈ I(r, n)


and Tα = {β ∈ I(r, n) | β ≤ α} (a left half space).
1.5. Degeneration of Schubert varieties 33

1.5.1 Definition
For α ∈ I(r, n), let Iα be the ideal of P generated by the set of monomials

{xγ xδ | γ and δ are noncomparable elements of Tα } .

We define the discrete algebra associated to α (or to the Schubert variety


X(α)) to be the ring P/Iα and denote it by Dα .
The main result of this section is that Proj Dα is a degeneration of
the Schubert variety X(α), i.e., there exists a one parameter flat family
with generic fibre X(α) and special fibre Proj Dα . Before coming to this
result, we collect some preliminaries.
If γ and δ are noncomparable elements of Tα , then by (1.3.6), we
obtain a quadratic relation in Rα (the homogeneous coordinate ring of
X(α)) of the form
X
(∗) pγ pδ = aλ,µ pλ pµ , where aλ,µ ∈ k, aλµ 6= 0,
λ,µ

and the monomials pλ pµ in the sum are standard, i.e. λ ≥ µ. We shall


denote by Q the set of all such quadratic relations for various pairs of
noncomparable elements γ, δ in Tα .

1.5.2 Lemma
P
Given any quadratic relation pγ pδ = λ,µ aλµ pλ pµ in Q, we have γ < λ,
δ < λ for all λ appearing in the sum.

Proof. Fix a λ appearing in the sum and restrict the quadratic relation
to X(λ). Since pλ pµ are standard monomials, by (1.3.3), the restriction
of the right hand side and hence, of pλ pµ to X(λ) is nonzero. Therefore,
by (1.2.10), we must have γ ≤ λ, δ ≤ λ. However, since γ and δ are not
comparable, equalities cannot hold. Hence γ < λ, δ < λ. Q.E.D.

1.5.3 Lemma
Let N denote the set of positive integers. There exists
P a function θ :
Tα → N such that given any quadratic relation pγ pδ = aλµ pλ pµ in Q,
λ,µ
we have
θ(λ) + θ(µ) > θ(γ) + θ(δ)
for all λ, µ appearing in the sum.
34 1. Schubert Varieties in the Grassmannian

Proof. Since Tα is a finite partially ordered set, we can define an order


function ord:Tα → N such that β > β ′ implies ord (β) > ord (β ′ ). For β
in Tα , set
θ(β) = N ord(β) ,

where N is a positive integer. We claim that for a sufficiently large N, θ


is the required function. Indeed, using (1.5.2) we have ord (γ) < ord(λ)
and ord (δ) < ord (λ); therefore, choosing N sufficiently large,  we can
achieve the inequality 1+N ord µ−(ord) (λ) > N ord (γ) + N ord (δ) N −(ord) λ
for all quadratic relations in Q. Q.E.D.
As the last item of the preliminaries we recall the following well
known lemma from commutative algebra.

1.5.4 Lemma
Let M be a finitely generated module over the ring A = k[[t]] and let K
be the quotient field of A. Then M is a free module if and only if

dimk M ⊗A k ≤ dimK M ⊗A K.

Proof. If M is free, then obviously the inequality holds. Conversely,


suppose that the inequality holds. If dimk M ⊗A k = m, then by
Nakayama’s lemma, M can be generated by m elements and therefore,
we have an exact sequence

0 → N → Am → M → 0

of A-modules. Tensoring the sequence by K and using the inequality,


we conclude that N ⊗A K = 0. Since N is torsion free, N = 0 and hence,
M is isomorphic to Am . Q.E.D.

1.5.5 Theorem
There exists a projective scheme X̃ flat over Spec k[[t]] whose generic
fibre is X(α) and the special fibre is Proj Dα , where Dα is the discrete
algebra associated to α in I(r, n).
Proof. Let Rα be the homogeneous coordinate ring of X(α) and let Q
be the set of quadratic relation (∗) in Rα as defined before (1.5.2). For
1.5. Degeneration of Schubert varieties 35

P
each quadratic relation pγ pδ = aλµ pλ pµ in Q, consider the polynomial
λ,µ

X
Fγ,δ = xγ xδ − aλµ xλ xµ
λ,µ

in P = k[xβ , β ∈ Tα ], and let J be the ideal of P generated by all such


polynomials. If xβ denotes the canonical image of xβ in P/J, then it
is clear that P/J is generated by standard monomials in xβ , β ∈ Tα .
Moreover, we have a canonical surjective map Φ : P/J → Rα which
sends xβ to pβ . Since Rα has a basis consisting of standard monomials
in pβ , by the usual argument as in the proof of (1.3.10), it follows that
Φ is injective. Hence, P/J ∼ = Rα .
P a function θ : Tα → N such that for each
Now, using (1.5.3), choose
quadratic relation pγ pδ = aλµ pλ pµ in Q,
λ,µ

θ(λ) + θ(µ) − θ(γ) − θ(δ) > 0.

Further, let A = k[[t]] and corresponding to each quadratic relation in


Q as above, consider the polynomial
X
Fγ,δ,t = xγ xδ − aλµ tθ(λ)+θ(µ)−θ(γ)−θ(δ) xλ xµ
λ,µ

in PA = A[xβ , β ∈ Tα ]. Denoting by I the ideal generated by the poly-


nomials Fγ,δ,t in PA , we set Rα = PA /I. Let K be the quotient field of
A and let PK = K[xβ , β ∈ Tα ]. If J˜ is the ideal of PK generated by the
polynomials Fγ,δ and I˜ is the one generated by the polynomials Fγ,δ,t ,
then it is easy to see that PK /I˜ ∼= PK /J˜, the isomorphism being in-
duced by the K-automorphism of PK which sends xβ to t−θ(β) xβ . Thus
we have

Rα ⊗A K = PA /I ⊗A K

= PK /I˜
∼ PK /J˜
=

= P/J ⊗k K

= Rα ⊗ k K

and Rα ⊗A k ∼ = Rα /tRα ∼ = Dα . This shows that the generic fibre of Rα


is Rα and its special fibre is Dα . Now we show that Rα is flat over A.
36 1. Schubert Varieties in the Grassmannian

From the definition of Dα it is clear that Dα is generated by standard


monomials in the canonical images xβ of xβ , β ∈ Tα . Therefore, using
the above isomorphisms, it follows that for every m ≥ 0

(∗) dimk (Rα )m ⊗A k = dimk (Dα )m


≤ dimk (Rα )m
= dimK (Rα )m ⊗k K
= dimK (Rα )m ⊗A K,

where (Rα )m , (Dα )m , (Rα )m denote the m-th graded components of the
respective rings. Thus, in view of (1.5.4), each graded component of Rα
is a free A-module and hence, Rα is a flat A-algebra.
Now if we take X̃ to be Proj Rα , the theorem immediately follows.
Q.E.D.

1.5.6 Corollary
(i) Dα has a basis consisting of standard monomials in xβ , β ∈ Tα .

(ii) Dα is reduced.

(iii) If P = Proj P, where P = k[xβ , β ∈ Tα ], then the degree of X(α)


in the Plücker embedding is the same as the degree of Proj Dα in
P.

Proof.

(i) Since for each m ≥ o, (Rα )m os a free A-module, we have

dimk (Rα )m ⊗A k = dimK (Rα )m ⊗A K,

and therefore, the inequalities (∗) in the proof of the above theorem
imply that

(∗∗) dimk (Dα )m = dimk (Rα )m for each m ≥ 0.

since distinct standard monomials are linearly independent in Rα


(1.3.3) and Dα is generated by standard monomials. It follows
that Dα has a basis consisting of standard monomials.
1.5. Degeneration of Schubert varieties 37

P
(ii) Let f = ai fi , where ai ∈ k, ai 6= 0, and fi are standard monomi-
als in Dα . From the definition of Dα it is clear that any product of
standard monomials in Dα is either zero or a standard monomial.
Therefore, using (i), it follows that f m 6= 0 for every m ≥ 0. This
proves that Dα is reduced.

(iii) The equality (∗∗) in the proof of (i) implies that the Hilbert polyno-
mial of X(α) is the same as that of Dα , which proves the assertion.
Q.E.D.
Let T be a subset of I(r, n). Recall that a chain of length m in
T is a sequence Λ = (λ(0), λ(1), . . . , λ(m)) of elements of T such that
λ(i) > λ(i + 1), 0 ≤ i ≤ m − 1. We say that Λ is maximal, if it is not a
subchain of any chain in T different from Λ, i.e., if Λ is the only chain
in T which contains all the Λ(i), 0 ≤ i ≤ m.
For a chain Λ in Tα , α ∈ I(r, n), let L(Λ) denote the linear space in
P = Proj P defined by the equations xβ = 0, β ∈ Tα − Λ, where Tα − Λ
denotes the set of elements of Tα not appearing in Λ.

1.5.7 Proposition
If M (Tα ) denotes the set of maximal chains in Tα , then
[
Proj Dα = L(Λ) (scheme theoretically),
Λ∈M (Tα )

i.e., the linear spaces L(Λ), Λ ∈ M (α ) are the irreducible components of


Proj Dα .

Proof. The ideal of the scheme theoretic union on the right hand side
is \
I= IΛ .
Λ∈M (Tα )

where IΛ is the ideal of L(Λ) which is generated by the set {xβ | β ∈ Tα −


Λ}. We shall show that I is the ideal defining Proj Dα , i.e., I is generated
by the set of monomials S = {xγ xδ | γ, δ are noncomparable elements of
Tα }. In fact, if γ, δ are noncomparable in Tα , then both of them together
cannot belong to a maximal chain in Tα . This shows that xγ xδ ∈ IΛ for
all Λ ∈ M (Tα ) and hence, S ⊂ I. Now if F ∈ I, it is easy to see that
every monomial of F is in I. Thus to complete the proof, it suffices to
show that every monomial F in I is divisible by an element of S. In fact,
38 1. Schubert Varieties in the Grassmannian

since F ∈ IΛ for all Λ ∈ M (Tα ), it is clear that F is not a standard


monomial and hence, it is divisible by an element of S. Q.E.D.

1.5.8 Corollary
The degree of X(α) in the Plücker embedding is equal to the number of
maximal chains in Tα .

Proof. Immediate from (1.5.6,iii) and the above proposition.

1.5.9 Remark
It can be shown that the discrete algebra Dα is Cohen-Macaulay (see [D-
E-P], section 8). Since Cohen-Macaulay nature is preserved under flat
deformation, this gives another proof of the fact that Schubert varieties
are arithmetically Cohen-Macaulay. Thus, in general, one can transfer
all those properties of Proj Dα (resp. Dα ) to X(α) (resp. Rα ) which are
preserved under flat deformation. Furthermore since Dα is combinatorial
in nature, the study of Schubert varieties via degeneration to Proj Dα
leads to combinatorial situations. We have already seen in (5.8), how
the geometry of X(α) (deg X(α)) is reflected in the combinatorics of the
partially ordered set Tα . Another simple illustration of this connection is
that the dimension of X(α) is equal to the length of any maximal chain
in Tα , as one can easily verify.

1.6 Connection with determinantal varieties


and invariant theory

It turns out that a determinantal variety can be realized as an affine open


subset of a Schubert variety. This enables us to transfer the local results
on Schubert varieties to determinantal varieties, for instance, Cohen-
Macaulayness and normality. In addition to these results we prove in
this section two results of classical invariant theory in the form given to
them by Doubilet-Rota-Stein [D-R-S] and De Concini and Procesi [D-P].
We recall from section (1.1.1) that the Grassmannian Gr,n is covered
by the affine open sets Uα , α ∈ I(r, n) with each Uα being identified
with the set of n × r matrices of rank r whose α-th minor is the identity
matrix; moreover, in the Plücker embedding of Gr,n each Uα each Uα
is identified with the open subset defined by pα 6= 0. By (1.2.3), max U
α
1.6. Connection with determinantal varieties and invariant theory 39

where αmax = (n − r + 1, n − r + 2, . . . , n), is the only open set among


Uα ’s, which is a Schubert cell called the big cell. On the other hand, we
call the affine open set Uαmin , where αmin = (1, 2, . . . , r), the opposite big
cell.
Recall that the opposite
  big cell is identified with the set of n × r
Ir
matrices of the form , where Ir is the r×r identity matrix and A is
A
an (n−r)×r matrix. Thus, the opposite big cell is an affine (n−r)r-space  
min Ir
whose origin is the point cell C(α ) consisting of the point .
0
Since the point cell is contained in every Schubert variety (2.10,i), the
opposite big cell has nonempty intersection with every Schubert variety,
and the intersection is a closed subvariety of the affine r(n − r)-space.
Further, since the opposite big cell is defined by pαmin 6= 0 in Gr,n ,
its coordinate ring R(Uαmin ) is the homogeneous localization
 

k[pα , α ∈ I(r, n)]pαmin = k , α ∈ I(r, n)
pαmin

As a k-algebra R(Uαmin ) is generated by the regular functions pα /pαmin


which are nothing but the restrictions of the Plücker coordinates to
Uαmin . As usual, we denote them also by Pα and call them the Plücker
coordinates on Uαmin . The following lemma is immediate from (1.3.7)
and (1.3.12).

1.6.1 Lemma
(i) The coordinate ring R(Uαmin ) of Uαmin has a basis consisting of
standard monomials in the Plücker coordinates pα on Uαmin .

(ii) For a Schubert variety X(β), let W (β) = X(β) ∩ Uαmin . The ideal
of W (β) in R(Uαmin ) is generated by the Plücker coordinates pα
on Uαmin with α 6≤ β.

Before proceeding further, we fix some notations:

1. Given positive integers s and m with s ≤ m, I(s, m) will denote,


as usual, the set of sequences α = (α1 , α2 , . . . , αs ), where αi ∈ Z
and 1 ≤ α1 < α2 < · · · < αs ≤ m. Denoting a constant sequence
(c, c, . . . , c) by (c), we shall write α ± (c) for the sequence (α1 ±
c, α2 , ±c, . . . , αs ± c).
40 1. Schubert Varieties in the Grassmannian

2. Let A be an ℓ × m matrix and s, a positive integer with s ≤


min(ℓ, m). For α = (α1 , α2 , . . . , αs ) in I(s, ℓ) and β = (β1 , β2 , . . . , βs )
in I(s, m), we shall denote by pα,β (A) the determinant of the
s × s minor of A corresponding to αi -th rows and βj -th columns,
1 ≤ i, j ≤ s (i.e., the minor whose (i, j)th entry is the (αi , βj )th
entry of A). Thus for each pair (α, β) in I(ℓ, m) × I(s, m), pα,β is a
polynomial function on M (ℓ, m), the affine space of ℓ×m matrices.

3. If A is an n × r matrix of rank r identifying an element of the


Grassmannian Gr,n , then for α ∈ I(r, n), we write

pα (A) = ± det(α-th minor of A)

which is the α-th coordinate of A in the Plücker embedding of


Gr,n .

1.6.2 Lemma
 
Ir
Let A = be an element of the opposite big cell Uαmin . Let α ∈
A
I(r, n) with α =6 αmin and let s be the integer such that αs ≤ r and
αs+1 > r. If

λ = (αs+1 , αs+2 , . . . , αr ),
µ = λ − (r) = (αs+1 − r, . . . , αr − r), and
ν = (1, 2, . . . , α̂1 , α̂2 , . . . , α̂i , . . . , α̂s , , αs + 1, . . . , r)

(i.e., the complement of {α1 , α2 , . . . , αs } in {1, 2, . . . , r} arranged in in-


creasing order), then

pα (A) = ±pλ,ν (A) = ±pµ,ν (A).

Proof. It is clear that the rows of the α-th minor of A which belong
to Ir do not contribute any thing towards the value of pα (A) = ± det
(α-th minor of A) and pα (A) is in fact determined by those rows of the
α-th minor which belong to A. Using this fact, the lemma easily follows
from the definition of pλ,ν (A) and pµ,ν (A).
1.6. Connection with determinantal varieties and invariant theory 41

1.6.3 Lemma
Let J denote the anti-diagonal m × m matrix
 
1
 0 
 
 1 
 
 . 
 
 . 
 
 . 
 
 0 
1

For a positive integer s ≤ m, let α, β ∈ I(s, m) and let γ = (m + 1 −


αs , m + 1 − αs−1 , . . . , m + 1 − α1 ). If A is any m × m matrix and B = JA,
then

pγ,β (B) = ±pα,β (A).

Proof. It is easy to see that multiplication by J on the left reverses the


order of rows of A; in other words, the i-th row of A is the (m + 1 − i)th
row of B. Now it is clear that

pγ,β (B) = ±pα,β (A).

Q.E.D.

1.6.4 Definition
(i) Let s, t, m, ℓ be positive integers with s ≤ m and t ≤ ℓ. For α ∈
I(s, m) and β ∈ I(t, ℓ), we define α ≤ β, if s ≥ t and αi ≤ βi ,
1 ≤ i ≤ t.

(ii) Let si , ti , mi , ℓi(i = 1, 2) be positive integers with si ≤ mi and ti ≤


ℓi . For α1 , α2 ∈ I(s1 , m1) × I(s2 , m2 ) and β 1 , β 2 ∈ I(t1 , ℓ1 ) ×
I(t2 , ℓ2 ), we define α1 , α2 ≤ β 1 , β 2 , if α1 ≤ β 1 and α2 ≤ β 2 .

Given positive integers ℓ and m, let c = min(ℓ, m) and let


[
P (ℓ, m) = (I(t, ℓ) × I(t, m)).
1≤t≤c
42 1. Schubert Varieties in the Grassmannian

We have a partial order ≤ on P (ℓ, m) as defined above, and for each


(α, β) ∈ P (ℓ, m), we have a regular function pα,β on M (ℓ, m). Thus we
can define standard monomials in pα,β , (α, β) ∈ P (ℓ, m), in the usual
manner:

1.6.5 Definition
A standard monomial in pα,β , (α, β) ∈ P (ℓ, m) (or on M (ℓ, m)) is a
formal expression of the form

pα,β pγ,δ · · · pλ,µ ,

where (α, β) ≤ (γ, δ) ≤ · · · ≤ (λ, µ). The sequence ((α, β), (γ, δ), . . . ,
(λ, µ)) satisfying the above inequalities is called a a double standard
tableau in the sense of Doubilet-Rota-Stein (see [D-R-S]).
As before, a standard monomial on M (ℓ, m) gives an element of
the coordinate ring R(M (ℓ, m)) and following our usual abuse of lan-
guage we call this element also a standard monomial on M (ℓ, m) (or in
pα,β , (α, β) ∈ P (ℓ, m)).
For α ∈ I(r, n) with α 6= αmin , let s be the largest integer such that
αs ≤ r and let

α′ = (n + 1 − αr , n + 1 − αr−1 , . . . , n + 1 − αs+1 ),
α′′ = (1, 2, . . . , α̂1 , α̂2 , . . . , α̂s , αs + 1, . . . , r),

where ˆ denotes omission. We call the pair (α′ , α′′ ) the dual pair as-
sociated to α. Thus we have a map h : I(r, n) − {αmin } → P (n − r, r)
defined by h(α) = (α′ , α′′ ).

1.6.6 Lemma
The above defined map h is an order reversing bijection, i.e., α ≤ β ⇔
(β ′ , β ′′ ) ≥ (α′ , α′′ ). Moreover, if (λ, µ) ∈ P (n−r, r) (say, (λ, µ) ∈ I(t, n−
r)×I(t, r) for some t, 1 ≤ t ≤ min(n−r, r)), then the element α ∈ I(r, n)
such that h(α) = (λ, µ) is given as follows:

(α1 , α2 , . . . , αr−t ) = (1, 2, . . . , µ̂1 , µ̂2 , . . . , µ̂t , µt + 1, . . . , r)


= The complement of µ in (1, 2, . . . , r)
arranged in increasing order,
(αr−t+1 , . . . , αr ) = (n + 1 − λt , n + 1 − λt−1 , . . . n + 1 − λ1 ).
1.6. Connection with determinantal varieties and invariant theory 43

Proof. It is an easy consequence of the various definitions involved.

1.6.7 Proposition
There is an isomorphism from the affine space M (n − r, r) to the op-
posite big cell Uαmin such that the induced isomorphism between their
coordinate rings identifies a Plücker coordinate pα on Uαmin with the
regular function ±pα′ ,α′′ on M (n − r, r).

Proof. Let Φ : M (n − r, r) → Uαmin be the map defined by


 
Ir
α(A) = ,
JA

where Ir is the r × r identity matrix and J is the (n − r) × (n − r) matrix


 
1
 0 
 
 1 
 
 . 
 
 . 
 
 . 
 
 0 
1

Since J is an invertible matrix, Φ is an isomorphism. Let Φ∗ : R(Uαmin ) →


R(M (n − r, r)) be the induced isomorphism of the coordinate rings. If
pα is a Plücker coordinate on Uαmin and A ∈ M (n − r, r), then using
(1.6.2) and (1.6.3), it is easy to see that

Φ∗ (pα )(A) = pα (Φ(A)) = ±pα′ ,α′′ (A).

Hence,
Φ∗ (pα ) = ±pα′ ,α′′ .
Q.E.D.

1.6.8 Theorem
For positive integers ℓ and m, the coordinate ring R(M (ℓ, m)) of M (ℓ, m)
has a basis consisting of standard monomials in the regular functions pα,β
on M (ℓ, m).
44 1. Schubert Varieties in the Grassmannian

Proof. Let n = ℓ + m and r = m. By (1.6.7), we have an isomorphism


Φ∗ : R (Uαmin ) → R(M (ℓ, m)) such that Φ∗ (pα ) = ±pα′ ,α′′ . Moreover,
since α 7→ (α′ , α′′ ) is an order reversing bijection (1.6.6), Φ∗ establishes
a bijection between the set of standard monomials in the Plücker coor-
dinates pα on Uαmin and the set of standard monomials in the regular
functions pα,β on M (ℓ, m). Now the theorem immediately follows from
the fact that R(Uαmin ) has a basis consisting of standard monomials in
pα , (1.6.1,i). Q.E.D.

1.6.9 Definition
Consider the coordinate ring R(M (ℓ, m)) = k[xij , 1 ≤ i ≤ ℓ, 1 ≤ j ≤
m] of M (ℓ, m). For a fixed positive integer t ≤ min(ℓ, m), the ideal of
R(M (ℓ, m)) generated by determinants of the t × t minors of the matrix
[xij ] is called a determinantal ideal, and the affine variety defined by
such an ideal is called a determinantal variety which we shall denote by
Dt (ℓ, m).

1.6.10 Proposition
Let t be a positive integer with t ≤ min(n−r, r) and let αt = (1, 2, . . . , t).
If β is the element of I(r, n) whose associated dual pair is (αt , αt ) (see
(1.6.6)), then the affine variety W (β) = X(β) ∩ Uαmin is isomorphic to
the determinantal variety Dt+1 (n−r, r). Conversely, every determinantal
variety is isomorphic to X(β) ∩ Uαmin for some Schubert variety X(β)
in a Grassmannian.

Proof. By (1.6.7), we have an isomorphism Φ : M (n − r, r) → Uαmin


which induces a k-algebra isomorphism Φ∗ : R(Uαmin ) → R(M (n −
r, r)) such that Φ∗ (pα ) = ±pα′ ,α′ . Since the ideal of W (β) in R(Uαmin )
is generated by the Plücker coordinates pα with α 6≤ β (1.6.1,ii), it
follows from (1.6.6) that the ideal of Φ−1 (W (β)) in R(M (n − r, r)) is
generated by the polynomials pα′ ,α′′ with (α′ , α′′ ) 6≥ (αt , αt ). But using
the definition of the partial order (1.6.4), this simply means that the ideal
of Φ−1 (W (β)) is generated by the determinants of the s×s minors of the
matrix [xij ] with s ≥ t + 1 and hence, with s = t + 1. Thus Φ−1 (W (β)) =
Dt+1 (n − r, r) which proves the first assertion of the proposition.
Conversely, given a determinantal variety Dt (ℓ, m), we set n = ℓ + m
and r = m. For t = 1, Dt (ℓ, m) is isomorphic to the Schubert variety
X(αmin ) consisting of a single point. So assume that t > 1. Now if
1.6. Connection with determinantal varieties and invariant theory 45

β is the element of I(r, n) whose associated dual pair is (αt−1 , αt−1 ),


then it is clear from the first assertion that Dt (ℓ, m) is isomorphic to
X(β) ∩ Uαmin .

1.6.11 Theorem
(i) Determinantal varieties are integral (i.e., irreducible and reduced)
and normal.

(ii) dim Dt (ℓ, m) = (t − 1)(ℓ + m − t + 1)

(iii) The coordinate ring of Dt (ℓ, m) has a basis consisting of standard


monomials in the regular functions pλ,µ on Dt (ℓ, m) with #λ ≤ t,
where #λ denotes the number of elements in the sequence λ.

Proof.

(i) This follows from (1.6.10) and the fact that Schubert varieties are
integral and normal.

(ii) For t = 1, clearly the formula holds. So assume that t > 1. Set
n = ℓ + m, r = m, and let β be the element of I(r, n) whose
associated dual pair is (αt−1 , αt−1 ) where αt−1 = (1, 2, . . . , t − 1).
By (1.6.10), Dt (ℓ, m) is isomorphic to an affine open subset of X(β)
and hence, dim Dt (ℓ, m) = dim X(β). But by (1.6.6), we have

β = (t, t + 1, . . . , m, ℓ + m − t + 2, ℓ + m − t + 3, . . . , ℓ + m).

Therefore, using the dimension formula for a Schubert variety


(2.5,ii), we obtain
m
X t−2
X m(m + 1)
dim Dt (ℓ, m) = j+ (ℓ + m − j) −
2
j=t j=0
= (t − 1)(ℓ + m − t + 1).

(iii) To avoid the trivial case, we assume that t > 1. Let n, r and β
be as above. By (1.6.7), we have an isomorphism Φ : M (ℓ, m) →
Uαmin with the induced isomorphism Φ∗ : R(Uαmin ) → R(M (ℓ, m))
such that Φ∗ (pα ) = ±pα′ ,α′′ . Moreover, as we saw in the proof of
(1.6.10), Φ maps Dt (ℓ, m) onto the affine variety W (β) = X(β) ∩
46 1. Schubert Varieties in the Grassmannian

Uαmin . Since the ideal of W (β) in R(Uαmin ) is generated by the


Plücker coordinates pα with α 6≤ β (1.6.1,ii), it follows from (1.6.6)
that the ideal of Dt (ℓ, m) is generated by the polynomials pα′ ,α′′ in
R(M (ℓ, m)) with (α′ , α′′ ) 6≥ (αt−1 , αt−1 ). Now the assertion follows
using (1.6.8) and the definition of the partial order (1.6.4). Q.E.D.

1.6.12 Corollary
A determinantal ideal is a prime ideal.
Proof. Obvious.

Now we proceed to prove two results of classical invariant theory in


the form given to them by Doubilet-Rota-Stein [D-R-S] and De Concini
and Procesi [D-P]. In what follows we shall assume familiarity with the
basic definitions and results of geometric invariant theory, for which we
refer the reader to [N], Chapter 3 or [M-F] Chapter 1.

1.6.13 Theorem
Let X = M (r, n) and let A = k[xij , 1 ≤ i ≤ n, 1 ≤ j ≤ r] be its
coordinate ring. Consider the canonical action of G = SLr (k) on A
induced by the action of G on X by left multiplication. If Ĝr,n denote the
cone over the Grassmannian Gr,n , then Spec AG ∼ = Ĝr,n ; in other words
G
the ring of invariants A is generated as a k-algebra by the determinant
functions ±pα , α ∈ I(r, n).
Proof. Recall that R = k[pα , α ∈ I(r, n)] is the coordinate ring of
Ĝr,n (see (1.1.4)). Since each pα is a G-invariant, we have the inclusions
R ⊂ AG ⊂ A which give the following commutative diagram:

To prove the theorem, we shall show that Ψ is birational and bijec-


tive. Then, since Ĝr,n is normal (4.22), it would follow from the Zariski’s
main theorem that Ψ is an isomorphism.
Let X ◦ = M (r, n)◦ be the set of r × n matrices of rank r, which is
clearly an open subset of X. For every x ∈ X ◦ , the orbit O(x) under the
action of G is a closed subset of X isomorphic to G. Further, it can be
easily seen that the points of X ◦ are precisely the “stable points” of X
in the sense of Mumford ([M-F] or [N]), so that X − X ◦ coincides with
the set of “unstable” points, i.e., for every x ∈ X − X ◦ , O ∈ O(x) (the
1.6. Connection with determinantal varieties and invariant theory 47

π̂

Spec AG Ĝr,n
Ψ

(a)

closure of O(x) in X). Therefore, by geometric invariant theory (see [N],


§3), the quotient variety U = X ◦ /G is an open subset of Spec AG . Thus
we have the following commutative diagram:

Ψ
Spec AG Ĝr,n X◦

U U Gr,n − {(0)}
Gr,n − {(0)} Ψ
(b) (c)

where (0) denote the vertex of the cone Ĝr,n . Note that the map π̂ :
X → Ĝr,n induced by the inclusion R ⊂ A is nothing but the morphism
π̂ defined in (1.2), i.e., for B ∈ X, π̂(B) = (pα (B)). Therefore, π̂ : X ◦ →
Ĝr,n − {(0)} defines a principal G-bundle on Ĝr,n − {(0)}. Now using the
commutative diagram (c), it is easy to see that Ψ : U → Ĝr,n − {(0)} is
an isomorphism, which shows that Ψ : Spec AG → Ĝr,n is birational.
Since X − X ◦ consists of unstable points, its image under the map
X → Spec AG is a point (see [N], §3). However, since X → Spec AG
is surjective, Spec AG − U consists of one point and it is clear from the
diagram (a) that this point is mapped to (0) by Ψ. This proves that Ψ
is bijective and hence, the theorem. Q.E.D.
48 1. Schubert Varieties in the Grassmannian

Let V be a vector space of dimension n and let V ∗ denote the dual of


V. We have canonical actions of G = GL(V ) on V and V ∗ . Recall that
if we fix a basis {e1 , e2 , . . . , en } of V and if the action of g on V, g ∈ G,
is given by a matrix Mg with respect to this basis, then the action of
g on V ∗ is given by the matrix (t Mg )−1 with respect to the dual basis
e∗1 , e∗2 , . . . , e∗n of V ∗ . Let
∗ ∗
| ×V ×
X=V × . . . × V }∗ with n < min(ℓ, m)
{z. . . × V} × |V × V {z
ℓ copies m copies

and let R(X) be the coordinate ring of X. Denoting an element (x1 , x2 ,


. . . , xℓ , ξ1 , ξ2 , . . . , ξm ) of X by (x, ξ), we define a map Φ : X → M (ℓ, m)
by Φ((x, ξ)) = [hxi , ξj i], where hxi , ξj i = ξj (xi ). Consider the diagonal
action of G on X, namely

g · (x1 , . . . , xℓ, , ξ1 , . . . , ξm ) = (g · x1 , . . . , g · xℓ , g · ξ1 , . . . g · ξm ).

Since (x, ξ) → hxi , ξj i is a G-invariant function on X, Φ is a G-invariant


morphism.

1.6.14 Theorem
The morphism Φ : X → M (ℓ, m) maps X into the determinantal variety
Dn+1 (ℓ, m), and the induced homomorphism Φ∗ : R(Dn+1 (ℓ, m)) →
R(X) between the coordinate rings induces an isomorphism Φ∗ : R(Dn+1
(ℓ, m)) → R(X)G , i.e., the determinantal variety Dn+1 (ℓ, m) is a “good
quotient” of X by G in the sense of geometric invariant theory.

For the proof we need the following from geometric invariant theory:

1.6.15 Lemma
Let X be an affine space on which a reductive algebraic group G acts
linearly. Let AN be an N -dimensional affine space and let Ψ : X → AN
be a G-invariant graded morphism (i.e., Ψ(tx) = td Ψ(x) for x ∈ X and
t ∈ k). Let D be a closed subvariety of AN such that Ψ(X) ⊂ D. Then
D is a good quotient of X by G and Ψ : X → D is the canonical quotient
map, provided the following conditions are satisfied:
(i) For x in X ss (the set of semi-stable points in X), Ψ(X) 6= (0).
(ii) There is a G-stable open subset U of X ss such that G operates freely
1.6. Connection with determinantal varieties and invariant theory 49

on U and Ψ induces an immersion U/G → AN (i.e., an injective mor-


phism with the induced maps between tangent spaces being injective).
(iii) dim D = dim U/G and (iv) D is normal.
Proof of the lemma. Let R be the coordinate ring of X and let
Y = Spec RG . Since the canonical morphism π : X → Y is a good
quotient and Ψ : X → AN is G-invariant, there is a canonical morphism
ρ : Y → AN such that the following diagram is commutative:

Y AN
ρ

Let X1 = Proj R and Y1 = Proj RG . Since Ψ is graded, Ψ defines


a rational map Ψ1 : X1 → PN −1 and the hypothesis (i) implies that the
restriction of Ψ1 to X1ss is a morphism. Further, since the morphism
π1 : X1ss → Y1 (the restriction of the canonical map X1 → Y1 ) is a
good quotient (see [N], Theorem 3.4) and Ψ1 is G-invariant, we again
get a morphism ρ1 : Y1 → PN −1 such that the following diagram is
commutative:

X1ss

Ψ1

π1

Y1 PN −1
ρ1

We claim that ρ1 is a finite morphism. To prove this claim, it suffices


to show that the line bundle L = ρ∗1 (OPN−1 (1)) on Y1 is ample. In fact,
50 1. Schubert Varieties in the Grassmannian

if M = Ψ∗1 OPN−1 (1)), then for a given x ∈ X1ss we can find an invariant
section f of Mr for some r ≥ 1 such that f (x) 6= 0 and (X1ss )f =
{z ∈ X1 | f (z) 6= 0} is affine. This implies that given y = π1 (x) ∈ Y1 ,
there is a section h of Lr for some r ≥ 1 such that h(y) 6= 0 and
(Y1 )h = {x ∈ Y1 | h(z) 6= 0} is affine. But this means that L is ample
(see for instance [N], Lemma 3.20). Hence, the claim follows.
Now, if P is the coordinate ring of AN , then ρ : Y → AN and
ρ1 : Y1 → PN −1 are induced by a graded homomorphism P → (RG =
L
RG ). Since ρ1 is finite, there exists a positive integer m0 such that
Lm≥0 mG G
m≥m0 Rm is a finitely generated P -module. However, since ⊕m<m0 Rm
is finite dimensional, RG itself is a finitely generated P -module, which
shows that ρ : Y → AN is a finite morphism. Now the hypothesis (ii)
and (iv) imply that ρ induces a birational morphism ρ : Y → D. Since
Y and D are normal, it follows that ρ is an isomorphism. This proves
the lemma.

Proof of the theorem. We can suppose without loss of generality that


ℓ and m are sufficiently large and hence, that ℓ = m. For, let
X′ = V . . × V} × |V ∗ × .{z
| × .{z . . × V }∗ ,
ℓ′ copies m′ copies

where ℓ′ > ℓ and m′ > m. Then the projection map ρ : X ′ → X induces


the inclusion R(X)G = R(X ′ )G ∩ R(X) which proves our claim. So we
assume that ℓ = m.
Obviously, Φ maps X into the determinantal variety Dn+1 (m, m).
The theorem would immediately follow from the above lemma, if we
show that conditions (i)–(iv) of the lemma are satisfied. We check these
conditions one by one.
(i) Let (x, ξ) = (x1 , . . . , xm , ξ1 , . . . , ξm ) ∈ X ss . Let Wx be the sub-
space of V spanned by xi ’s and Wξ , the subspace of V ∗ spanned
by ξj ’s. Suppose that φ((x, ξ)) = 0 i.e. hxi , ξj i = 0 for all i, j.

Case (a): Wξ = 0, i.e., ξj = 0 for all j. Consider the one parameter


subgroup
   
 t 0 

 


  t  

   
 . .. 
Γ = It =   t 6= 0 of GL(V ).

   


  0  


 

t
1.6. Connection with determinantal varieties and invariant theory 51

Then It · (x, O) = (tx, O), so that It · (x, O) → (0) as t → 0. Thus


the origin (0) is in the closure of 0((x, ξ)) and consequently, (x, ξ)
is not semi-stable, a contradiction.

Case (b): Wξ 6= 0. Since the case Wx 6= 0 is similar to the


case (a), we may assume that Wx 6= 0. Since hxi , ξj i = 0 for
all i, j, we can choose a basis {e1 , . . . , en } of V such that Wξ =
Span of (e1 , e2 , . . . , er ), r < n, and Wξ ⊂ span of e∗r+1 , . . . , e∗n ,
where {e∗1 , . . . , e∗n } is the dual basis in V ∗ . Consider the 1-parameter
subgroup
   
 t 

 


  t 0  


     

  t  
  tIr 0
Γ = gt =  t −1  = t 6
= 0

   0 t−1 Im−r 


  . .  


  0 .  


 

t −1

of GL(V ). We have
gt · (x, ξ) = (tx, tξ) → 0 as t → 0.
Thus, due to the same reason as in case (a), the point (x, ξ) is
not semi-stable which leads to a contradiction. Hence, we have
φ((x, ξ)) 6= 0.
(ii) Let U = {(x, ξ) ∈ X | (x1 , x2 , . . . , xn ) and (ξ1 , . . . , ξn ) are linearly
independent }. Clearly, U is a G-stable open subset of X and G
operates freely on it. Let {e1 , e2 , . . . , en } be a basis of V and let
U ′ = {(x, ξ) | xi = ei , 1 ≤ i ≤ n}.
Then U ′ can be identified with U/G and we have a commutative
diagram

Thus, to verify (ii), we have only to show that the map Φ : U ′ →


M (m, m) and its differential dΦ are both injective. Indeed, if
(x, ξ), (x′ , ξ ′ ) are in U ′ such that Φ(((x, ξ)) = Φ((x′ , ξ ′ )) then
hei , ξj i = hei , ξ ′ i for all i, j
⇒ ξj = ξj′ for all j.
52 1. Schubert Varieties in the Grassmannian

U/G U′

Since {e1 , . . . , en } is a basis of V, the equality hxi , ξj i = hx′i , ξj i for


all i, j, implies that xi = x′j for all i. This shows that Φ : U ′ →
M (m, m) is injective. To prove that its differential is injective, we
merely note that the above argument remains valid for the points
over k[ε] (the ring of dual numbers), i.e., it remains valid, if we
replace k by k[ε] or in fact, by any k-algebra.

(iii) We have

dim U/G = dim U − dim G


= 2mn − n2
= dim Dn+1 (m, m), by (1.6.11,ii).

(iv) This follows from (1.6.11,i). Q.E.D.

1.6.16 Corollary
Let X and G be as above. Let fij denote the regular function (x, ξ) →
hxi , ξj i on X, 1 ≤ i ≤ ℓ, 1 ≤ j ≤ m, and let F denote the ℓ × m matrix
[fij ]. The ring of invariants R(X)G has a basis consisting of standard
monomials in the regular functions pλ,µ (F ) with #λ ≤ n+1, where #λ is
the number of elements in the sequence λ = (λ1 , λ2 , . . . , λt ) and pλ,µ (F )
is the determinant of the t × t minor defined by the rows corresponding
to λ1 , λ2 , . . . , λt and the columns corresponding to µ1 , µ2 , . . . , µt . (See
(1.6.4) and (1.6.5) for the definition of standard monomials in pλ,µ (F ).)

Proof. Immediate from (1.6.11,iii). Q.E.D.

Proof of Remark 1.2.8. We shall now indicate its proof. We have to


prove that π(D(α0 )) is closed. Let Br denote the upper triangular ma-
trices in GL(r). Then Br operates freely (on the right) on M (n, r)0 so
that M (n, r)0 is a principal fibre Br -bundle over M (n, r)0 /Br . Since
1.6. Connection with determinantal varieties and invariant theory 53

D(α)0 is a closed Br -stable subset in M (n, r)0 , its canonical image


Z in M (n, r)0 /Br is closed. Now M (n, r)0 /Br is a fibre space over
M (n, r)0 /GL(r) with fibre GL(r)/Br (a complete variety), the canon-
ical morphism M (n, r)0 |Br −→ M (n, r)0 /GL(r) is proper. Hence the
image of Z in M (n, r)0 /GL(r) = Gr,n is closed and this is precisely
π(D(α)0 ) i.e. π(D(α)0 ) is closed.
Chapter 2

Standard monomial theory on SLn(k)/Q

As in the previous chapter, we assume k to be an algebraically closed


field of arbitrary characteristic. Let Q be a parabolic subgroup of SLn (k)
and let X be a Schubert variety in the complete variety SLn (k)/Q. The
purpose of this chapter is to develop a standard monomial theory for
the Schubert varieties X in SLn (k)/Q, i.e., to give an explicit basis for
the space H 0 (X, La ) where La is a line bundle on SLn (k)/Q such that
a ≥ 0. (see (2.1.3)).
Section 1-6 follow the treatment of [M-S] (see also G/P −IV ), except
for the important fact that it avoids the usage of normality of Schubert
varieties and the normality of Schubert varieties is deduced as a con-
sequence. The proof given in section 7 is the one in G/P − V, when
specialized to the case of special linear group.

2.1 Some facts about G/Q

In the following let G′ denote GLn (k) and denote SLn (k) by G. A Borel
subgroup of G′ (resp. G) is a maximal element among its connected
solvable subgroups. Denote by B ′ (resp. B) the Borel subgroup of G′
(resp. G) consisting of upper triangular matrices. Recall that all Borel
subgroups of G′ and of G are conjugate.

2.1.1
The homogeneous space G/B(≃ G′ /B ′ ) is called the flag variety asso-
ciated to G. This terminology is justified as follows: Let {e1 , . . . , en } be
the canonical basis of V = kn and let F (V ) be the flag variety of V. G′
operates transitively on F (V ). Let F ∈ F (V ) be the flag
F := (0) ⊂ (e1 ) ⊂ (e1 , e2 ) ⊂ . . . ⊂ (e1 , . . . , en ) = V.

© Springer Science+Business Media Singapore 2016 and Hindustan Book Agency 2015 55
C.S Seshadri, Introduction to the Theory of Standard Monomials,
Texts and Readings in Mathematics 46, DOI 10.1007/978-981-10-1813-8_2
56 2. Standard monomial theory on SLn (k)/Q

The stabilizer of F in G′ is B ′ . The orbit map:

Φ : G′ → F (V )
g 7→ g · F

induces a bijective morphism

Ψ : G′ /B ′ → F (V ).

It is easy to see that this map is an isomorphism of varieties: moreover, Ψ


is G′ -equivariant with respect to the canonical left action of G′ on G′ /B ′ .
Since F (V ) is complete, it follows that G′ /B ′ (≃ G/B) is complete.

2.1.2
A closed subgroup Q of G is called a parabolic subgroup if there exists a
Borel subgroup B̃ of G such that B̃ ⊂ Q. The canonical morphism Φ :
G/B̃ → G/Q is surjective. Since G/B̃ (≃ G/B) is complete, we conclude
that G/Q is a complete variety. It is clear that Φ is G-equivariant with
respect to the canonical left actions of G on G/B̃ and G/Q. All maximal
parabolic subgroups of G containing B are of the form:

Pi := {g = (gℓ,m )1≤ℓ,m≤n ∈ SLn (k)|


gℓ,m = 0 for i + 1 ≤ ℓ ≤ n, 1 ≤ m ≤ i}

where 1 ≤ i ≤ n − 1.
It is easy to see that G/Pi is isomorphic to the Grassmannian Gi,n
for i = 1, . . . , n − 1.

2.1.3
We want to recall some facts:

(i) Let Q be any parabolic subgroup of G containing B. There exists


a subset {i1 , . . . , ir } of {1, . . . , n − 1} such that Q = Pi1 ∩ . . . ∩ Pir
not only set-theoretically, but scheme theoretically. In particular,
we know that B = P1 ∩ . . . ∩ Pn−1 . Denote Pi ∩ Pi+1 ∩ . . . ∩ Pn−1 by
πi π2
Qi . We have a canonical map: G/Qi −→ G/Pi × . . . × G/Pn−1 −→
P(∧i ∨) × . . . × P(∧n−1 ∨). The morphism π1 is a closed immersion
and π2 is given by the Plücker embeddings.
2.1. Some facts about G/Q 57

(ii) The Picard group Pic G/Pi is infinite cyclic and it is generated
by the very ample line bundle Li , where Li is the pull back of the
hyperplane line bundle on P(∧i ∨).

(iii) The map pm : G/Qi → G/Pm , i ≤ m ≤ n − 1, is a fibre bundle


with fibres Pm /Qi . By abuse of notation we denote the pull back
p∗m (Lm ) on G/Qi by Lm .

(iv) H 0 (G/Pm , Lm ) ≃ H 0 (G/Qi , Lm ) for i ≤ m ≤ n − 1.

(v) The Picard group Pic G/Qi of G/Qi is isomorphic to Zn−i and it
is generated by the line bundles Li , . . . , Ln−1 .
⊗a
(vi) We denote an element L⊗a i
i
⊗ . . . ⊗ Ln−1n−1 of Pic G/Qi by La(i) ,
where a(i) denotes the tuple (ai , . . . , an−1 ). We say a(i) ≥ 0 (resp.
a(i) > 0) if am ≥ 0 (resp. am > 0) for i ≤ m ≤ n − 1. The line
bundle La(i) is ample if and only if a(i) > 0.

Remark The definitions and results above and in the following do not
depend on the special choice of the chain of parabolic subgroups: B =
Q1 ⊂ 6−
Q2 ⊂ 6−
... ⊂
6−
Qn−1 = Pn−1 . Let (i1 , . . . , in−1 ) be any permutation of
(1, . . . , n − 1). Let Qij denote Pij ∩ Pij+1 ∩ . . . ∩ Pin−1 . Thus the obvious
analogues hold for the chain of parabolic subgroups

B = Q i1 ⊂
6−
Q i2 ⊂
6−
... ⊂
6−
Qin−1 .

2.1.4
Let T be the subgroup of diagonal matrices in G. The subgroup T is
a maximal torus of G, i.e., T is a maximal connected diagonalizable
subgroup of G. Recall that all maximal tori of G are conjugate. It is
well known that T = CenG T = {g ∈ G | gtg −1 = t, ∀ t ∈ T }. Let
NorG T = {g ∈ G | gT g−1 ⊆ T }, the normalizer of T in G. The factor
group W = NorG T /T is called the Weyl group of G. It is easy to see
that W ≃ Sn , the group of permutations on n letters.

2.1.5
If ω ∈ W is an element of the Weyl group, then BωB is a well defined
double coset because T is contained in B. S
By the Bruhat decomposition
(see Chapter 4, 4.1.6) we know that G = · BωB is a mutually disjoint
ω∈W
58 2. Standard monomial theory on SLn (k)/Q

double coset decomposition. Let WQi denote NorQi T /T, the Weyl group
of Qi . We can consider WQi as a subgroup of W. Denote the coset Qi
in G/Qi by e(id) and denote by e(ω), ω ∈ W/WQi , the coset ωQi . Note
that the canonical map

W/WQi → W/WPi × . . . × W/WPn−1


Tn−1
is injective and we have: WQi = m=i WPm .

2.1.6
The points e(ω), ω ∈ W/WQi are fixed under the canonical left action
of T on G/Qi . We want to show that these are all the T -fixed points in
G/Qi . Let {e1 , . . . , en } be the canonical basis of V = kn . It is easy to
see that the T -fixed points in G/Pm ≃ Gm,n are the points represented
by ei1 ∧...∧eim , 1 ≤ i1 < . . . < im ≤ n, and that WPm ∼ = Sm × Sn−m ֒→
 
1...m...n
Sn . Let ω = be an element of W. Then ω mod Pm is
i1 . . . im . . . in
uniquely determined by arranging i1 , . . . , im in ascending order. Thus
we get a bijection between W/WPm and I(m, n), the indexing set of the
T -fixed points in G/Pm . Hence all the T -fixed points in G/Pm are of
form e(ω) with ω ∈ W/WPm . Since the morphism

πi : G/Qi → G/Pi × . . . × G/Pn−1

is G-equivariant, it is easy to see that all T -fixed points in G/Qi are of


the form e(ω), ω ∈ W/WQi .

2.1.7
The B-orbit B.e(ω), ω ∈ W/WQi , is an affine space (see 4.2.4 in the
later Chapter 4). Hence we have a Bruhat decomposition of G/Qi into
affine spaces:

G/Qi = B.e(ω).
ω∈W/WQi

Denote the orbit B.e(ω) by C(ω).

Definition The B-orbit C(ω) is called the Schubert cell corresponding


to ω ∈ W/WQi . Its closure X(ω) = C(ω) in G/Qi , endowed with the
2.2. Young diagrams and standard monomials 59

reduced scheme structure, is called the Schubert variety corresponding


to ω.
Let ω ∈ WQi and denote its canonical image in W/WPm , 1 ≤ m ≤
n−1, by ω m . It is obvious that the image of the Schubert variety X(ω) ⊂
G/Qi under the projection pm : G/Qi → G/Pm is the Schubert variety
X(ω m ) ⊂ G/Pm .

2.2 Young diagrams and standard monomials

2.2.1

We have seen (Chapter 1, (1.3.7)) that the standard monomials


 of length
0 ⊗m
m ≥ 0 in the Plücker coordinates form a basis of H G/Pj , Lj . Let
⊗a
 ⊗a

1 ≤ i1 , i2 ≤ n. If p ∈ H 0 (G/Pi , Li1 1 ) = H 0 G/Qi , Li1 1 and q ∈
i i

 ⊗ai
  ⊗ai

H 0 G/Pi , Li2 2 = H 0 G/Qi , Li2 1 , then the product p · q := p ⊗ q
 ⊗ai ⊗ai

is an element of H 0 G/Qi , Li1 1 ⊗ Li2 2 . We want to show that a
certain subset of these products
 form a basis for the multigraded ring
0
R = ⊕a(i)≥0 H G/Qi , L a(i) .

2.2.2

We have a canonical partial order on W/WQi . Let ω, τ ∈ W/WQi .

Definition ω ≤ τ ⇔ X(ω) ⊆ X(τ ).

Lemma Let i ≤ m ≤ n − 1 and denote by ω m and τ m the canonical


images of ω and τ in W/WPm . The following are equivalent:

(i) ω ≤ τ

(ii) ω m ≤ τ m for all m, i ≤ m ≤ n − 1.

Proof. See Chapter 4, (4.3.11).


60 2. Standard monomial theory on SLn (k)/Q

2.2.3 Definition
A Young diagram of type a(i) = (ai , . . . , an−1 ), where am ≥ 0, m =
i, . . . , n − 1, is an element

λ ∈ (W/WPi × . . . × W/WPi ) × . . . × W/WPn−1 × . . . × W/WPn−1
| {z } | {z }
ai copies an−1 copies

Let ω ∈ W/WQi and let X(ω) be the corresponding Schubert variety


in G/Qi . We say that

λ = λi,1 , λi,2 , . . . , λi,ai ; λi+1,1 , . . . , λn−1,an−1

is a Young diagram on X(ω) (or for short ω ≥ λ), if for every projection
pm : W/WQi → W/WPm , i ≤ m ≤ n − 1, we have pm (ω) = ω m ≥ λm,km ,
1 ≤ km ≤ am .

2.2.4
Let λ be Young
0 a(i)
 diagram of type a(i). We define the section pλ ∈
H G/Qi , L corresponding to λ to be the monomial:

pλ = pλi,1 · pλi,2 · · · pλn−1,an−1

of Plücker coordinates (see 1.1.4).



It is easy to see that: pλ X(ω) 6≡ 0 ⇔ ω ≥ λ.

Definition Let λ be a Young diagram of type a(i) and let X(ω), ω ∈


W/WQi , be a Schubert variety in G/Qi . The diagram λ is called standard
with respect to X(ω) if there exists an element

Φ = Φi,1 , Φi,2 , . . . , Φi,ai , Φi+1,1 , . . . , Φn−1,an−1 ,
Qn−1 am
in m=i (W/WQi ) such that

1) Φm,ℓm is a lift of λm,ℓm under G/Qi → G/Pm i ≤ m ≤ n − 1,


1 ≤ ℓ m ≤ am .

(2) ω ≥ Φi,1 ≥ Φi,2 ≥ · · · ≥ Φi,ai ≥ Φi+1,1 ≥ · · · ≥ Φn−1,an−1 in


W/WQi .
2.3. Linear independence of standard monomials 61

In this case Φ is called a defining chain for λ. The corresponding section


pλ is called a standard monomial on X(ω). We say that λ is standard
(resp. pλ ) is a standard monomial if it is standard on G/Q.
In the above definition there is an abuse of language for we consider
two standard monomials to be distinct if the corresponding Young dia-
grams are distinct. Note that a similar definition of standard monomials
(or standard tableau) could be given by fixing an arbitrary ordering of
the maximal parabolic subgroups of G containing B. The above defini-
tion corresponds to the ordering P1 , . . . , Pn−1 . The definition of standard
monomials depends upon the choice of an ordering of P1 , . . . , Pn−1 , but
all the considerations of this chapter go through for any ordering of
P1 , . . . , Pn−1 .

2.2.5 Remark
Let λ be a Young diagram of type a(i). For each λm,km write the cor-
responding element in I(m, n) (see (1.1.6)) as a column with increasing
entries. This gives us a one-to-one correspondence between the Young
diagrams of type a(i) in the above sense and the classical Young dia-
grams of shape a(i). In classical literature (see [W]), a Young diagram
is called “standard” if the rows have non-decreasing entries. Let us call
such diagrams weakly standard. It can be shown that standard diagrams
on G/B (in our sense) are precisely those which are weakly standard (see
G/P -IV). However, on a Schubert variety, being standard in our sense
is (in general) stronger than being weakly standard.

Example G = SL3 (k), Q = B, ω = (312) ∈ S3 = W. Then λ =


2 1
∈ W/WP1 × W/WP2 is weakly standard. Furthermore, ω ≥ λ,
3
but λ is not standard on X(ω). The possible lifts for λ11 are : Φ11 = (231)
or Φ11 = (213). Now (231) 6≤ (312). Hence the only possibility to build
1
up a defining chain is to take Φ11 = (213). The possible lifts for λ22 =
2
are Φ21 = (312) or Φ21 = (132). But in both cases we have Φ11 6≥ Φ21 .
Hence λ is not standard on X(ω).

2.3 Linear independence of standard monomials

Let Y = X(ω1 ) ∪ . . . ∪ X(ωr ) be a union of Schubert varieties in G/Qi . If


λ is a Young diagram of type a(i) and pλ is the corresponding monomial
62 2. Standard monomial theory on SLn (k)/Q

in the Plücker coordinates, then we say that λ (resp. pλ ) is standard


on Y if it is standard
 on some X(ωm ), 1 ≤ m ≤ r. Note that pλ ∈
H 0 G/Qi , La(i) . We want to show that the standard monomials are
linearly independent.

2.3.1
Let S(Y, a(i)) denote the set of all standard monomials on Y of type
a(i) ≥ 0 and let s(Y, a(i)) denote the cardinality of S(Y, a(i)).

Proposition If Y = X(ω1 )∪. . .∪X(ωr ) is a union of Schubert varieties


in G/Qi , then the set of standard monomials on Y is linearly indepen-
dent. In particular, the elements of S(Y, a(i)) are linearly independent
and

h0 Y, La(i) ≥ s(Y, a(i)).

Proof. We have already proved the proposition for the Grassmannian


(see Chap 1, (1.4.5)), so by induction on i we can assume that the
proposition holds for G/Qm , m ≥ i + 1. We perform induction on ai ,
where a(i) = (ai , . . . , an−1 ).

Case 1: ai = 0
Let η : G/Qi → G/Qi+1 denote the canonical map. Note that the cor-
responding map of global sections preserves standardness. Denote η(Y )
by Y ′ . Any standard monomial of type a(i) = (0, ai+1 , . . . , an−1 ) can be
identified with a standard monomial on Y ′ of type (ai+1 , . . . , an−1 ). The
claim follows from the induction hypothesis and the surjectivity of η.

Case 2: ai ≥ 1
Let pλ(k) , 1 ≤ k ≤ s, be a minimal set of linearly dependent standard
monomials of type a(i). We can suppose that λ(1) , . . . , λ(s) are distinct
(k)
as diagrams. Let φi,j be a defining chain for λ(k) . Since the pλ(k) are
(k)
standard, for each k there exists an ℓ such that ωℓ ≥ Φi,1 .

(1) (2) (s)


Case 2a: λi,1 = λi,1 = · · · = λi,1 = δ, say.
Every pλ(k) is by definition standard on a Schubert variety X(ωℓ ). Let
Z be the union of these irreducible components of Y. The monomials
pλ(k) , k = 1, . . . , s, are standard and linearly dependent on Z ⊆ Y. We
2.4. Some facts about the partial order on W/WQi 63

(k)
can write pλ(k) = pδ pµ(k) ; we get µ(k) from λ(k) by dropping λi,1 . Now
the monomials pµ(k) are of type a′ (i) = (ai − 1, ai+1 , . . . , an−1 ). They
are standard on Z, and pλ 6= 0 on every irreducible component of Z.
Furthermore, the monomials pµ(k) , k = 1, . . . , s, are again distinct as
diagrams. Since the pλ(k) are linearly dependent, the monomials pµ(k)
would be linearly dependent, which is a contradiction to our induction
hypothesis.
n o
(1) (s)
Case 2b: Let λ be a minimal element of the set λi,1 , . . . , λi,1 say,
(1) (s ) (s +1)
, . . . , λsi,1 .
λ = λi,1 = · · · = λi,10 and λ 6≥ λi,10
   (s 
(1)
Let Y0 be the union of Schubert varieties X Φi,1 ∪ . . . ∪ X Φi,1(0) in
G/Qi . Observe that Y0 ⊆ Y and that the monomials pλ(1) , . . . , pλ(s0 ) are
standard
  on Y0 . Let Pi : G/Qi → G/Pi be the projection map. Since
(k)
Pi X Φi,1 = X(λ) ֒→ G/Pi and pλ(r) ≡ 0 on the Schubert variety
i,1
X(λ) ֒→ G/Pi for r ≥ s0 + 1, we conclude pλ(r) Y0 ≡ 0 for r ≥ s0 + 1. Let
Pa
(k)
ak pλ(k) be a linear dependence relation on Y. By the minimality
k=1
of s we know that ak 6= 0 ∀ k = 1, . . . , s. If we restrict the relation to
Y0 , then we obtain a nontrivial linear dependence relation on Y0 for the
standard monomials pλ(1) , . . . pλ(s0 ) . Case 2a then applies. Q.E.D.

2.4 Some facts about the partial order on W/WQi

We want to recall some definitions and facts about the partial order
in W/WQi . A more general discussion about this and the proofs of the
Lemmas stated below can be found in Chapter 4.

2.4.1
We have already noted that W = Sn . The group W is generated by
the transpositions s1 = (1, 2), . . . , sn−1 = (n − 1, n). Let ω ∈ W and
ω = si1 . . . sir be a representation of ω. Call r the length of the rep-
resentation. The representation is called reduced if r is minimal, in
which case it is called the length ℓ(ω) of ω. Let ω ∈ W/WQi and let
ω ≡ si1 · · · · · sir mod WQi be a representation of ω. The representation
is called reduced if r is minimal, in which case r is called the length
ℓQi (ω) of ω. Any element τ ∈ W/WQi which has a reduced decompo-
sition τ ≡ siji · · · · · sijk mod WQi such that j1 < · · · < js is called a
64 2. Standard monomial theory on SLn (k)/Q

subword of ω.

2.4.2 Lemma
Let X(ω) and X(τ ), ω, τ ∈ W/WQi , be Schubert varieties in G/Qi . The
following are equivalent:
(i) X(ω) ⊆ X(τ )

(ii) ω is a subword of τ.

Proof. See Chapter 4, (4.3.8).

2.4.3 Lemma
Let m ≥ i and let η : G/Qm → G/Qi denote the canonical map. For
any ω ∈ W/WQi there exists a minimal representative τ min ∈ W/WQm
and a maximal representative τ max ∈ W/WQm such that
η
(i) the map X(τ min ) → X(ω) is birational and if ω ≡ s1 · · · · · · · sr
mod WQi is a reduced decomposition, then τ̃ min = s1 · · · · · sr is a
representative of τ min in W.
 
(ii) ℓQm τ̃ min · σ = ℓQm τ min + ℓQm (σ)∀ σ ∈ WQi .

(iii) ∀ ρ ∈ W/WQm such that η(X(ρ)) = X(ω), we have τ max ≥ ρ ≥


τ min in W/WQm and ρ ≡ τ̃ min · σ mod WQi for some σ ∈ WQi .

(iv) X(τ max ) = η −1 (X(ω)).

Proof. See Chapter 4, 4.3.10.

2.4.4
Let X(ω) be a Schubert variety in G/Qi and let λ be a standard Young
diagram of type a(i) on X(ω).

Lemma (Deodhar) There exists a unique maximal defining chain Φ+


and a unique minimal defining chain Φ− for λ such that if Ψ is any
defining chain for λ, then

Φ+ −
j,kj ≥ Ψj,kj ≥ Φj,kj , 1 ≤ j ≤ n, 1 ≤ kj ≤ aj .
2.5. Preparation for the main theorem 65

Proof. See Chapter 4, (4.6.3).

Remark The minimal defining chain is “independent of ω” in the sense


that if λ is standard on G/Qi , then λ is standard on X(ω) ⇔ ω > Φ̄i,1 .

2.5 Preparation for the main theorem

In this section we are going to prove some lemmas which we will need
for the proof (based on induction) of the main theorem (namely that
H 0 (X(ω), La(i) ) is generated by standard monomials).

2.5.1 Lemma
Let ω1 , . . . , ωt , τ1 , . . . , τs be elements of W/WQi and let a(i) ≥ 0. If
Y1 = ∪τj=1 X(ωj ) and Y2 = ∪sℓ=1 X(ωℓ ) are the corresponding unions of
Schubert varieties then

s (Y1 ∪ Y2 , a(i)) = s(Y1 , a(i)) + s(Y2 , a(i)]) − s(Y1 ∩ Y2 , a(i)).

Proof. We have only to show: if λ is a standard Young diagram on


Y1 and Y2 , then λ is standard on Y1 ∩ Y2 . It is enough to consider the
case Y1 = X(ω) and Y2 = X(τ ). Let Φ = (Φi,1 , Φi,1 , . . .) be a minimal
defining chain for λ. We have ω ≥ Φi,1 and τ ≥ Φi,1 . (Recall the remark
in 2.4.4.) Hence we know that X(Φi,1 ) ⊆ X(ω) ∩ X(τ ) and λ is standard
on Y1 ∩ Y2 . Q.E.D.

2.5.2
Denote by pi : G/Qi → G/Pi the canonical map. Recall that we denote
the image of ω ∈ W/WQi in W/Wpi by ω. To avoid ambiguity we denote
Pi (X(ω)) by XPi (ω). We have the commutative diagram:

X(ω) ֒→ G/Qi
↓ ↓
XPi (ω) ֒→ G/Pi

Assume that dim XPi (ω) > 0. Let pω be the Plücker coordinate on
G/Pi corresponding to ω ∈ W/WPi . We have seen (Chapter 1, (1.4.11))
that
HPi (ω) = XPi (ω) ∩ {pω = 0}
66 2. Standard monomial theory on SLn (k)/Q

is the union of Schubert varieties of codim 1 in XPi (ω). We can consider


{pω = 0} as a section of the line bundle Li on G/Qi . Denote by abuse
of notation the pull back of this section on G/Qi by {pω = 0}.

Definition Denote the proper intersection X(ω) ∩ {pω = 0} on G/Qi


by H(ω).
Note that H(ω)red is a union of Schubert varieties of codim. 1 in
X(ω).

2.5.3 Lemma
Let X(ω) be a Schubert variety in G/Qi and assume that dim XPi (ω) >
0. If a(i) = (ai , . . . , an−1 ) such that a(i) ≥ 0 and ai > 0, then

s(X(ω), a(i)) = s(X(ω), a′ (i)) + s(H(ω)red , a(i))

where
a′ (i) = (ai − 1, ai+1 , . . . , an−1 ).

Proof. It is clear that if pν is any standard monomial on X(ω) of type


a′ (i), then pω · pν is standard on X(ω) and is of type a(i). So we can
identify S(X(ω), a′ (i)) with the subset of monomials of S(X(ω), a(i))
which begin with pω .
It remains to show that those elements of S(X(ω), a(i)) which do not
begin with pω can be identified with the elements of S(H(ω)red , a(i)).
We can write H(ω)red = ∪j X(τ (j)), where the X(τ (j)) are Schubert
varieties of codim. 1 in X(ω). Denote by τ (j) the image of τ (j) in
W/WPi . It is obvious that τ (j) is a maximal representative of τ (j) in
W/WQi such that τ (j) ≤ ω. Let λ be any standard Young diagram of
type a(i) on X(ω) such that pλ does not begin with pω . We know that
ω > λi,1 . Recall that HPi (ω) := pi (H(ω)red ) is the union of all Schubert
varieties of codim. 1 in XPi (ω) (and hence all the Schubert varieties
properly contained in XPi (ω) are contained in HPi (ω)). Write HPi (ω) =
∪ℓ XPi (ξ(ℓ)), ξ(ℓ) ∈ W/WPi . For at least one ℓ we have ω > ξ(ℓ) ≥ λi,1 .
Hence λ is standard on X(τ (ℓ)), else pλ would vanish on HPi (ω) and
hence begin with p(ω). Q.E.D.
2.5. Preparation for the main theorem 67

2.5.4 Notations
(i) If ζ denotes (1, . . . , n − i), then Lζ is an ample line bundle on
| {z }
n−i
G/Qi .

(ii) Denote by St the set of all subschemes of G/Qi which are scheme
theoretic unions of Schubert varieties of dimension ≤ t. Note that
the elements of St are reduced schemes.

Lemma Let Y1 , Y2 be unions of Schubert varieties in G/Qi . Suppose


we have 
h0 Yi , Lm,ζ = s(Yi , mζ); i = 1, 2, m ≫ 0.
Then

(i) the scheme theoretic intersection Y1 ∩ Y2 is reduced.

(ii) a) h0 (Y1 ∪ Y2 , Lmζ ) = s(Y1 ∪ Y2 , mζ)


b) h0 (Y1 ∩ Y2 , Lmζ ) = s(Y1 ∩ Y2 , mζ) for m ≫ 0.

Proof. We have the exact sequence

0 → OY1 ∪Y2 → OY1 ⊕ OY2 → OY1 ∪Y2 → 0.

Denote Lmζ |Y by OY (m) for any union of Schubert varieties Y in G/Qi .


If we tensor the exact sequence by Lmζ , then we get:

0 → OY1 ∪Y2 (m) → OY1 (m) ⊕ OY2 (m) → OY1 ∩Y2 (m) → 0.

For m ≫ 0 we get (by Serre’s Theorem):

H 0 (Y1 ∩ Y2 , Lmζ ) = h0 (Y1 , Lmζ ) + h0 (Y2 , Lmζ ) − h0 (Y1 ∪ Y2 , Lmζ )

We know that s(Y1 ∪ Y2 , mζ) ≤ h0 (Y1 , ∪Y2 , Lmζ ). We obtain

h0 (Y1 ∩ Y2 , Lmζ ) ≤ s(Y1 , Lmζ ) + s(Y2 , mζ) − s(Y1 ∪ Y2 , mζ).

Thus we conclude by Lemma 2.5.1 that

(∗) h0 (Y1 , ∩Y2 , Lmζ ) ≤ s(Y1 ∩ Y2 , mζ) for m ≫ 0.


68 2. Standard monomial theory on SLn (k)/Q

Let Z denote Y1 ∩ Y2 . We claim that if Z is not reduced, then


h0 (Z, Lmζ )> h0 (Zred , Lmζ ) for n ≫ 0. The canonical map π : Zred → Z
gives us the exact sequence

0 → I → OZ → OZred → 0.

If we tensor by OZ (m), we obtain

0 → I(m) → Oz (m) → OZred (m) → 0.

Since the twisted sheaf I(m) is never trivial unless Z is reduced, we


obtain the claim by Serre’s theorem because then h0 (Z, I(m)) 6= 0 for
m ≫ 0. Since h0 (Zred , Lmζ ) ≥ s(Zred , mζ) we conclude that h0 (Y1 ∩
Y2 , Lmζ ) > h0 ((Y1 ∩ Y2 )red , mζ) for m ≫ 0 if Y1 ∩ Y2 is not reduced,
but this contradicts (∗). It follows that Y1 ∩ Y2 is reduced and formula
(ii)b holds. Formula (ii)(a) follows then from (ii)(b) and Lemma 2.5.1.
Q.E.D.

2.5.5 Lemma
Suppose that for all Schubert varieties X(ω) ∈ St (5.4 ii) we have

(∗) s(X(ω), a(i)) = h0 (X(ω), La(i) ) for all a(i) ≥ 0.

Then

(i) For any Y1 , Y2 ∈ St , Y1 ∩ Y2 is reduced.

(ii) (∗) holds for any Y ∈ St .

Proof. Using the previous lemma and induction on the number of


components, we see that (i) holds and that (ii) holds for a(i) = mζ,
m ≫ 0. Hence the only point to show is that (ii) holds for arbitrary
a(i). We proceed by induction on t and on the number of components
of Y. Assume that (ii) holds for Sd , d ≤ t − 1. Let Y1 denote X(ω(1)) ∪
. . . ∪ X(ω(r − 1)), let Y2 = X(ω(r)) and denote Y1 ∪ Y2 by Y. We can
suppose that (ii) holds for Y1 and Y2 . We have the exact sequence

0 → OY (a(i)) → OY1 (a(i)) ⊕ OY2 (a(i)) → OY1 ∩Y2 (a(i)) → 0.


2.5. Preparation for the main theorem 69

Since dim Y1 ∩ Y2 ≤ t − 1, we know that

h0 (Y1 ∩ Y2 , La(i) ) = s(Y1 ∩ Y2 , a(i)).

The corresponding sequence of global sections is hence exact and we get

h0 (Y1 , OY1 (a(i)) + h0 (Y2 , OY2 (a(i)))


= h0 (Y, OY (a(i))) + h0 (Y1 ∩ Y2 , OY1 ∩Y2 a(i))

By Lemma 2.5.1 we know that

s(Y1 , a(i)) + s(Y2 , a(i)) = s(Y, a(i)) + s(Y1 ∩ Y2 , a(i)).

Since we assumed that s(Yj , a(i)) = h0 (Yj , OYj (a(i))), j = 1, 2; we con-


clude h0 (Y, OY (a(i))) = s(Y, a(i)). Q.E.D.

Remark The same arguments give: If H j (X(ω), , La(i) ) = 0 for all


j > 0 and all Schubert varieties X(ω) in St , then H j (Y, La(i) ) = 0 for
j > 0 and all Y ∈ St .

2.5.6 Lemma
Let X(ω) be a Schubert variety in G/Qi . Denote by Φ : ^
X(ω) → X(ω)
the normalization map.

^ → X(ω) is bijective.
(i) The normalization map Φ : X(ω)

(ii) X(ω) is smooth in codim. 1; i.e. codim(singX(ω)) ≥ 2 in X(ω).

(iii) The intersection multiplicity of X(ω) and {pω = 0} considered


as subvarieties of G/Qi is 1 along the irreducible components of
H(ω)red .

Proof. See Chapter 4, (4.4), (4.5), (5.9).

Remark The normalization map Φ : X(ω)^ → X(ω) is an isomorphism


over an open subset U of X(ω) such that codim(X(ω)− U ) ≥ 2 in X(ω).
70 2. Standard monomial theory on SLn (k)/Q

2.5.7
If L is a line bundle on X(ω), then denote its pullback by φ on ^
X(ω)
^ ^
by L̃. Let H(ω) be the subscheme of X(ω) defined by the pull back of
^ is defined by only one equation.
{pω = 0}. Note that H(ω)

^ is reduced.
Proposition H(ω)

^ endowed with the


Proof. Let Zred be an irreducible component of H(ω)
reduced structure. The local ring of ^
X(ω) at the generic point of Zred is
isomorphic to the local ring of X(ω) at the generic point of φ(Z) = X(τ ),
where X(τ ) is a Schubert variety of codim. 1 in X(ω). (Note that Zred is
of codim. 1 in ^
X(ω) and the normalization map is an isomorphism over
an open subset U of X(ω) such that codim(X(ω) − U ) ≥ 2 in X(ω),
Remark 2.5.6) The local ring (A, M ) of X(ω) at the generic point of
X(τ ) is a discrete valuation ring (Lemma 2.5.6, (iii)). The fact that
the intersection multiplicity of X(ω) and {pω = 0} is one along the
irreducible components (Lemma 5.6, (iii)) is equivalent to saying that the
maximal ideal M of A is generated by pω . Again, since the normalization
map is an isomorphism outside a set of codimension one, the same is
^ Now the proposition follows from the following lemma.
true on X(ω).

2.5.8 Lemma
Let B be a noetherian normal ring and f ∈ B a non unit. Denote by
(f ) = ∩ri=1 qi the primary decomposition and let pi be the associated
prime ideal (of ht 1)Tto qi . Assume that the multiplicity of f along every
pi is 1. Then (f ) = ri=1 pi and B/f B is reduced.

(j)
Proof of the lemma. We have qi = pi , the j-th symbolic power of
pi . Since j is T
the multiplicity of f along pi and j = 1, we get qi = pi and
hence (f ) = ri=1 pi .

2.6 Main theorem

We will prove in this section that H 0 (Y, La(i) ), a(i) ≥ 0, is generated by


standard monomials for any union of Schubert varieties in G/Qi , and,
2.6. Main theorem 71

as a consequence, we will prove the normality of Schubert varieties in


G/Qi .

2.6.1 Main Theorem


(i) If Y = X(ω(1)) ∪ X(ω(2)) ∪ . . . ∪ X(ω(r)) is a union of Schubert
varieties in G/Qi , then h0 (Y, La(i) ) = s(Y, a(i)) for all a(i) ≥ 0.

(ii) All Schubert varieties X(ω) in G/Qi are normal.

Proof. By Lemma (2.5.5)(ii) it suffices to prove the theorem only for


Schubert varieties. Furthermore, to prove the theorem it suffices to show
: h0 (^ ^ denotes the normaliza-
X(ω), L̃a(i) ) = s(X(ω), a(i)), where X(ω)
tion of X(ω) and L̃a(i) denotes the pull back of La(i) via the normal-
ization map. For, in this case, the canonical map H 0 (X(ω), La(i) ) ֒→
H 0 (^
X(ω), L̃a(i) ) is an isomorphism and the normality follows from the
following lemma.

2.6.2 Lemma
Let p : X̃ → X be a finite morphism of projective varieties. Suppose
that there exists an ample line bundle L on X such that the canonical
map H 0 (X, Lm ) → H 0 (X̃, L̃m ) is an isomorphism for all m ≫ 0. Then
p is an isomorphism.

Proof of the Lemma. Consider the exact sequence

0 → OX → p∗ OX̃ → F → 0.

If we tensor the exact sequence by Lm , we get

0 → Lm → p∗ OX̃ ⊗ Lm → F ⊗ Lm → 0.

Since H 0 (X, Lm ) ≃ H 0 (X̃, L̃m ) ≃ H 0 (X, p ∗ OX̃ ⊗ Lm ) for m ≫ 0, we


get H 0 (X, F ⊗ Lm ) = 0 for m ≫ 0. But F is coherent and L is ample.
Hence we conclude that F = 0 and p is an isomorphism. Q.E.D.

2.6.3 Proof of the theorem.


The proof is by several induction steps. Let X(ω) be a Schubert variety
in G/Qi . The induction hypotheses are:
72 2. Standard monomial theory on SLn (k)/Q

1. If G = SLn (k), suppose that the theorem holds for all Schubert
varieties in G′ /Q′ , where G′ = SL2 (k), . . . , SLn−1 (k) and Q′ is
any parabolic subgroup of G′ . (The theorem holds obviously for
G = SL2 (k).)

2. The theorem holds for all Schubert varieties in G/Qm , m = i +


1, . . . , n − 1 (G/Qn−1 = G/Pn−1 is a Grassmann variety and hence
the theorem holds, see Chapter 1, (4.14) and (4.22).)

3. The theorem holds for Schubert varieties of dimension less than


dim X(ω).

4. Let a(i) = (ai , . . . , an−1 ) ≥ 0. We will use induction on ai .

Case 1 : ai = 0
Consider the map η : G/Qi → G/Qi+1 . The line bundle La(i) on G/Qi

is the pull back of the line bundle La (i) , a′ (i) = (ai+1 , . . . , an−1 ), on
G/Qi+1 via η. Denote η(X(ω)) by Y and denote by the same η the map
η : X(ω) → Y. One knows that on an open subset U of Y (big cell in Y ),
η −1 (U ) ≃ U × affine space. This implies in particular that the function
field of Y is algebraically closed in that of X(ω). Further Y is normal
and η is proper. Hence by Zariski’s connectedness theorem, the fibres of
η are connected. Moreover, since Y is normal. η∗ OX(ω) = OY , and since

η is proper, the induced map between the global sections H 0 (Y, La (i) ) →
H 0 (X(ω), La(i) ) is an isomorphism. Note that standardness is preserved
under this map and hence, by induction hypothesis 2 the theorem holds
if ai = 0.

Case 2 : ai ≥ 1
Denote by pi : G/Qi → G/Pi the projection map and denote pi (X(ω))
by XPi (ω).

Case 2a : dim XPi (ω) = 0


Note that in this case X(ω) ֒→ Pi /Qi ≃ SLn−i (k)/Bn−i , where Bn−i is
a Borel subgroup of SLn−i (k). Hence we can look at X(ω) as a Schubert
variety in a flag variety of a group of lower rank. Since standardness is
preserved under the corresponding map of global sections, we are done
by induction hypothesis.
2.6. Main theorem 73

Case 2b : dim XPi (ω) ≥ 1


Let Y1 , . . . , Yr be the irreducible components of H(ω)red and denote by
Ỹi , i = 1, . . . , r the irreducible components of ^
H(ω) (see 2.5.7). Since
the Yi , i = 1, . . . , r, are Schubert varieties of codim. 1 in X(ω), we know
by induction hypothesis 3 that h0 = s for unions and intersections of the
Yi and the Yi are normal. Hence we know that if Φ : H(ω) ^ → H(ω)red
denotes the canonical morphism, then the restrictions Φi : Ỹi → Yi are
isomorphisms for i = 1, . . . , r and Φ is a bijective map. We claim that
Φ is an isomorphism. This follows from the lemma below.

2.6.4 Lemma
Let p : Z̃ → Z be a bijective proper map of reduced schemes. If Z =
Z1 ∪ Z2 , Z̃ = Z̃1 ∪ Z̃2 such that Z1 , Z2 , Z̃1 , Z̃2 are closed subschemes,
Z1 ∩ Z2 is reduced and p induces isomorphisms p : Z̃1 → Zi , i = 1, 2,
then p is an isomorphism.

Proof of the lemma. We are reduced to the affine case. So let Zi =


Spec Ai , Z̃1 = Spec Ãi , Z1 ∪ Z2 = Spec A, Z̃1 ∪ Z̃2 = Spec Ã, Z1 ∩ Z2 =
Spec A12 and Z̃1 ∩ Z̃2 = Spec Ã12 . We have the following commutative
diagram of exact sequences:

0 A A1 ⊕ A2 A12 0

φ ζ

0 Ã Ã1 ⊕ Ã2 Ã12 0

The map Φ is injective since A is reduced and p is surjective. An


easy diagram chase shows that Φ is also surjective. Q.E.D.

2.6.5 Back to the proof of the theorem


^ = H(ω)red , we know that h0 = s for ^
Since H(ω) H(ω). Denote the ideal
e ^
sheaf of H(ω) by I (H(ω)). We have the exact sequence

^ →O
0 → I (H(ω)) ^ → O ^ → 0.
X(ω) H(ω)
74 2. Standard monomial theory on SLn (k)/Q

Tensoring by Lea(i) we get:

0 → O ^ (a′ (i)) → O ^ (a(i)) → O ^ (a(i)) → 0


X(ω) X(ω) H(ω)

where a′ (i) = (ai − 1, ai+1 , . . . , an−1 ). Since

^ La(i) ) = H 0 (H(ω)red , La(i) )


H 0 (H(ω),

is generated by standard monomials and these are induced by sections


on X(ω) (and a fortiori sections on ^
X(ω)), we conclude that the corre-
sponding sequence of global sections is exact:

^ La′ (i) ) → H 0 (^
0 → H 0 (H(ω), ^ La(i) ) → 0.
X(ω), La(i) ) → H 0 (H(ω),

^ La(i) ) = s(H(ω)red , a(i)) we get


Since h0 (H(ω),

s(H(ω)red , a(i)) = h0 (^
X(ω), Lea(i) ) − h0 (^

X(ω), Lea (i) ).
By induction hypothesis on ai , we know that

h0 (^

X(ω), Lea (i) ) = s(X(ω), a′ (i)) and hence
h0 (^
X(ω), Le a′ (i)
) = s(H(ω)red , a(k))s(X(ω), a′ (i))
= s(X(ω), a(i)) by Lemma 2.5.3.

Hence we conclude that

h0 (^
X(ω), Lea(i) ) = s(X(ω), a(i)) ∀ a(i) ≥ 0

which proves the theorem. Q.E.D.

2.7 Another proof for generation by standard monomials

We want to give another proof of theorem that H 0 (X(τ ), La(i) ), a(i) ≥ 0,


is generated by standard monomials. The proof is based on the same
induction procedure as the proof given before, but uses the methods
developed in Chapter 4. We will not need the fact that X(τ ) is smooth
in codim. 1 (proved by decreasing induction on dim X(τ ), Chapter
4, Cor.4.5). The proof for the generation will be purely by increasing
induction on dim X(τ ).
2.7. Another proof for generation by standard monomials 75

2.7.1
Let S be the set of simple roots of the root system Φ of G corresponding
to the choice of a Borel subgroup B (see Chapter 4, (1.3), (1.5)). If
α ∈ S, then denote the corresponding simple reflection by sα . We denote
by P (α) the minimal parabolic subgroup of G containing B and G−α
(see Chapter 4, (1.5)). Let L(α) be the semisimple subgroup of P (α)
isomorphic to SL2 (k) and containing Gα and G−α . Denote by B(α)
the Borel-subgroup L(α) ∩ B of L(α). If X(ω) and X(τ ) are Schubert
varieties in G/Qi such that ω < τ and ω = sα τ for some α ∈ S, then let
Z(ω, α) be the fibre bundle L(α) ×B(α) X(ω) (≃ P (α) ×B X(ω)) over
P1 with fibre X(ω) (see Chapter 4, 2.9).

Lemma The canonical map Ψ : Z(ω, α) → X(τ )((g, x) 7→ g · x)) is


surjective and the fibres are connected.

Proof. See Chapter 4, Cor. (2.11) and Lemma (4.3).

2.7.2 Theorem
(i) If Y = X(ω1 ) ∪ . . . ∪ X(ωr ) is a union of Schubert varieties in
G/Qi , then
h0 (Y, La(i) ) = s(Y, a(i)).

(ii) All Schubert varieties in G/Qi are normal.

(iii) If ω, τ ∈ W/WQi such that ω < τ and ω = sα τ for some α ∈ S,


then
H 0 (X(τ ), La(i) ) ≃ H 0 (Z(ω, α), Ψ∗ (La(i) ))

Proof. By Lemma (2.5.5) it suffices to prove the theorem for Schubert


varieties. Let X(τ ) be a Schubert variety in G/Qi . As in the earlier
proof, the induction hypotheses are:
The theorem holds for unions of Schubert varieties: (a) in G/Qm ,
i + 1 ≤ m ≤ n − 1; (b) in G′ /Q, with G′ of lower rank and (c) in G/Qi
of dimension less than dim X(τ ).
In the case when G = SL2 (k) or X(τ ) is a point or dim X(τ ) = 1,
there is nothing to be shown. We divide the induction procedure into
several steps.
76 2. Standard monomial theory on SLn (k)/Q

Step A: Let pi : G/Qi → G/Pi be the projection map. We denote


by XPi (τ ) ֒→ G/Pi the Schubert variety pi (X(τ )). Assume that XPi (τ )
is a point. If i = n − 1, then X(τ ) = XPi (τ ) and there is nothing to
be shown. If i < n − 1, then X(τ ) ֒→ Pi /Qi , and the result follows
by induction on rank G. Note that standardness is preserved under the
corresponding map of global sections. Hence by induction on the rank,
the theorem holds for X(τ ).
Step B Assume that XPi (τ ) is not a point. Choose XPi (ω) (
XPi (τ ) such that ω = sα τ for some α ∈ S. Let ω = sα τ in W/WQi . Note
that X(ω) ⊆ X(τ ) is of codim. 1. The fibre bundle Z(ω, α) is normal,
since the base is P1 and X(ω) is normal by induction hypothesis. Since
X(ω) is the preimage of a point under the map π : Z(ω, α) → P1 , we
know that the ideal sheaf I(X(ω)) of X(ω) on Z(ω, α) is π ∗ (OP1 (−1)).
Step C We want to recall the following general construction. Let F
be an object on X(ω), say a coherent sheaf, on which B acts (consistent
with the action of B on X(ω)). We denote by F # the “associated” object
on Z(ω, α) (i.e. F # = P (α) ×B F ). Let L = La(i) be a line bundle on
G/Qi such that a(i) ≥ 0. To avoid ambiguity, we denote by L(τ ) the
restriction of L to X(τ ). Now note that the map Ψ : Z(ω, α) → X(τ ) is
a P (α)-morphism. The line bundle L is a homogeneous line bundle on
G/Qi , i.e. G acts on L. Since X(τ ) is P (α)-stable (Chapter 4, Lemma
2.8), we know that P (α) acts on L(τ ) and hence on Ψ∗ (L(τ )). Let G0
denote the fibre of Ψ∗ (L(τ )) over the point e ∈ P (α)/B corresponding
to the coset B. We see that B acts on G0 . If G#0 denotes the associated
object over P (α)/B, we see easily that Ψ (L(τ )) = G#

0 . Since G0 =
L(ω), we get Ψ∗ (L(τ )) = L(ω)# .
Step D If XPi (ω) is a point, denote the divisor X(ω) in Z(ω, α)
by HZ . We have the following commutative diagram:

Ψ Z(ω, α)
X(τ )
π

Pi′ P1
pi

≃ Z(ω, α)
XPi (τ )
2.7. Another proof for generation by standard monomials 77

Since the line bundle π ∗ (OP1 (−1)) corresponds to the divisor X(ω)
in Z(ω, α), we get, by the commutativity of the diagram for the ideal
sheaf I(HZ ) of HZ : Ψ∗ (L−1
i ) = I(HZ ).
Step E Assume that XPi (ω) is not a point. Denote p−1 i (HPi (ω)) ∩
X(ω) by H(ω) (see 2.5.2). Since X(ω) is normal by induction hypothe-
sis, we know that the local rings of the generic points of the irreducible
components of H(ω) are discrete valuation rings. By Chevalley’s multi-
plicity formula (Chapter 4, Corollary (5.9)), we know that the intersec-
tion multiplicity of X(ω) and p∗i ({p(ω) = 0}), considered as subvarieties
of G/Qi is 1 along the irreducible components of H(ω). Hence the local
rings of the generic points of the irreducible components are generated
by p(ω), and H(ω) and H(ω)# are reduced. Denote by HZ the scheme
theoretic union H(ω)# ∪ X(ω) in Z(ω, α). Note that HZ is reduced. It
is obvious that I(H(ω)# ). I(X(ω)) (note that I(X(ω)) is locally prin-
cipal) is a subsheaf of the sheaf of ideals I(HZ ) of HZ . On the other
hand, given f ∈ I(HZ ) locally, we can write f = g.θ, where θ is a local
equation for X(ω). Since H(ω)# 6⊇ X(ω), g has to vanish on H(ω)# and
hence
I(HZ ) = I(H(ω)# .I(X(ω))
Since I(X(ω)) ∼ = p∗ (OP1 (−1)), we can write I(HZ ) = I(H(ω)# )
(−1). We have a canonical isomorphism I(H(ω)) ∼ = L−1
i on X(ω) given
by multiplication with p(ω). Since B acts nontrivially on p(ω), using
the corresponding fibre space construction, it is easy to see that we
get a “twist by OP1 (1),” i.e. we get an isomorphism I(H(ω)# ) ≃
(Li (ω))# (1) ≃ Ψ∗ (L(τ ))(1) (see C). Therefore we get

Ψ∗ (L−1
i ) = I(HZ ).

Note that set theoretically HZ = Ψ−1 (H(ω)). Both define the same line
bundle and hence Ψ∗ (H(τ )) = HZ .
Step F We want to show that Ψ∗ : H 0 (H(τ )red , La(i) ) → H 0 (HZ , Ψ∗
a(i)
(L )) is an isomorphism. Since we take the reduced structure on H(τ ),
we have only to show that Ψ∗ is surjective. Recall that if XPi (ω) is a
point, then HZ = X(ω) ֒→ Z(ω, τ ), and Ψ∗ is obviously an isomorphism
in this case. If XPi (ω) is not a point, then denote the irreducible com-
ponents of H(ω) by X(φi ), i = 1, . . . , r. The Schubert varieties X(φi )
are normal. The irreducible components of H(ω)# are the fibre spaces
X(φi )# = L(α) ×B(α) X(φi ), i = 1, . . . , r. Since ω = sα τ, we know that
sα φi ≤ ω in W/WQi and hence X(φi ) is P (α)-stable. The fibres of the
canonical map Ψi : L(α) ×B(α) X(φi ) → X(φi ) are connected (Lemma
78 2. Standard monomial theory on SLn (k)/Q

2.7.1) and φi is proper. Hence we deduce that Ψ∗ (OX(φi )# ) = OX(φi ).


Let HZ = ∪Si , Si irreducible, and let H(τ )red = ∪Ti , Ti irreducible, such
that Ψ(Si ) ⊆ Ti . If s ∈ H 0 (HZ , Ψ∗ (La(i) )), then for every i = 1, . . . , r + 1
there exists a ti ∈ H 0 (Ti , La(i) ) such that Ψ∗ (ti ) = s|Si . It remains to
show that the ti glue together to define a global section on H(τ )red .
Since we take the reduced structure on H(τ ), this is a purely set theo-
retical question. Let y be an element of H(τ )red . Choose a trivialization
of La(i) on a neighborhood U of y. Then Ψ∗ (La(i) ) is trivial on Ψ−1 (U ).
The “value” of s is constant on Ψ−1 (y) and the “value” of ti (y) is in-
dependent of the neighborhood of y. Hence the ti glue together to form
a global section t such that Ψ∗ (t) = s and Ψ∗ : H 0 (H(τ )red , La(i) ) →
H 0 (HZ , Ψ∗ (La(i) )) is an isomorphism.
Step G We will prove that H 0 (X(τ ), La(i) ), a(i) = (ai , . . . , an−1 )
is generated by standard monomials by induction on ai . Assume ai = 0.
Note that if i = n − 1, there is nothing to be shown. If i < n − 1,
then let η : G/Qi → G/Qi+1 be the projection map and denote the
Schubert variety η(X(τ )) by X ′ (τ ′ ). The line bundle La(i) is the pull
back of the line bundle La(i+1) on G/Qi+1 via η, where a′ (i + 1) =
(ai+1 , . . . , an−1 ). We know by induction that X ′ (τ ′ ) is normal. If X ′ (τ ′ )
is a point, then X(τ ) ֒→ Qi+1 /Qi , which is associated to “SL(k) in
lower rank” and hence we are done by induction. It X ′ (τ ′ ) is not a
point, choose X ′ (ω ′ ) ⊆ X ′ (τ ′ ) of codim. 1 such that ω ′ = sα .τ ′ , α ∈ S.
Let ω = sα τ in W/WQi . Note that X(ω) is of codim. 1 in X(τ ). We
have the following commutative diagram

Ψ
Z(ω, α) X(τ )

η′ η
Ψ′
Z ′ (ω ′ , α′ ) X ′ (τ ′ )

The fibre of Ψ′ , η and η ′ are connected, X ′ (τ ′ ) and Z ′ (ω ′ , α′ ) are nor-


mal by induction; Ψ′ , η and η ′ are proper and hence the induced maps
between the global sections are isomorphisms:
∗ ′
H 0 (Z(ω, α), Ψ∗ (La(i) ) = H 0 (Z ′ (ω ′ , α′ ), Ψ′ (La (i) )

= H 0 (X ′ (τ ), La (i+1) ) = H 0 (X(τ ), La(i) ).
2.7. Another proof for generation by standard monomials 79

Since standardness is preserved under η ∗ , we observe by induction that


H 0 (X(τ ), La(i) ) is generated by standard monomials.
Thus we have proved (i) and (ii) in the case ai = 0.
Step H Assume that ai > 0 and (i) and (iii) hold for all Lb(i) ∈
Pic G/Qi such that b(i) ≥ 0 and 0 ≤ bi < ai . We have an exact sequence

0 → I(HZ ) = Ψ∗ (L−1
i ) → OZ(ω,α) → OHZ → 0.

Denote by a′ (i) the tuple (ai − 1, ai+1 , . . . , an−1 ). If we tensor the


exact sequence by Ψ∗ (La(i) ) we get

0 → Ψ∗ (La (i) ) → Ψ∗ (La(i) ) → Ψ∗ (La(i) )|HZ → 0

Now by step F we know that

H 0 (HZ , Ψ∗ (La(i) )) ≃ H 0 (H(τ )red , La(i) ),

and by the induction hypothesis H 0 (H(τ )red , La(i) ) is generated by


standard monomials and these are induced by sections on X(τ ) (and a
fortiori on Z(ω, α)). Hence the corresponding sequence of global sections
is exact


0 → H 0 (Z(ω, α), Ψ∗ (La (i) )) → H 0 (Z(ω, α), Ψ∗ (La(i) )) →
→ H 0 (HZ , Ψ∗ (La(i) )) → 0.

By the induction hypothesis, we know that h0 (Z(ω, α), Ψ∗ (La (i) ) =

h0 (X(τ ), La (i) ) = s(X(τ ), a′ (i)) and h0 (H(τ )red , La(i) ) =
a(i)
s(H(τ )red , L ). Hence we get

h0 (Z(ω, α), Ψ∗ (La(i) ))



= h0 (Z(ω, α), Ψ∗ (La (i) )) + h0 (HZ , Ψ∗ (La(i) ))
= s(X(τ ), a′ (i)) + s(H(τ )red , a(i))
= s(X(τ ), a(i)) by Lemma 2.5.2.

We conclude that H 0 (X(τ ), La(i) ) = H 0 Z(ω, α), Ψ∗ (La(i) ) and H 0 (X(τ ), La(i) )
is generated by standard monomials.
Step I It remains to show that X(τ ) is normal. Denote by φ :
]
X(τ ) → X(τ ) the normalization map. The map Ψ : Z(ω, α) → X(τ )
factors through the normalization, so we get a map Ψ̃ : Z(ω, α) →
80 2. Standard monomial theory on SLn (k)/Q

]). Since the normalization map is bijective (Lemma 2.5.6(i)), we


X(τ
know that the fibres of Ψ̃ are connected. Now since Ψ̃ is proper and
]) is normal, it follows that Ψ̃∗ (OZ(ω,α) ) = O
X(τ ^ and the map
X(τ )
Ψ̃ : ]), φ∗ (La(i) ))
H 0 (X(τ → H 0 (Z(ω, α), Ψ∗ (La(i) ))
is an isomorphism for
any ample line bundle La(i) on X(τ ). Hence we have H 0 (X(τ ), La(i) ) =
]), φ∗ (La(i) )) for any ample line bundle La(i) on X(τ ) and thus
H 0 (X(τ
]) = X(τ ).
X(τ Q.E.D.
Chapter 3

Applications

We shall be concerned, in this chapter, with Schubert varieties in


SLn (k)/B, where B is a Borel subgroup containing a maximal torus
T in SLn (k). To begin with, in section 1, we compute the dimension of
the tangent space at the T -fixed points of a Schubert variety X(τ ); this
essentially describes the singular locus of X(τ ). In section 2, we gener-
alize the vanishing theorem (1, 4.17) to Schubert varieties in SLn (k)/B,
and in section 3, we derive a formula for the character of the T -module
H 0 (X(τ ), La ), where a ≥ 0. The aim of section 4 is to prove a result
(4.4) which is crucial for the “variety of complexes”(section 5). The
main result of section 5 is that the variety of complexes is an open sub-
set of a union of Schubert varieties in SLn (k)/Q (Q a parabolic subgroup
endowed with its canonical reduced scheme structure).
Section 1 is taken from [L-S1]. The proof in (3.2.7), (3.2.8) and
(3.2.9) is the one for Theorem 7.1, [M-S]. For section 3, see [S2] or [S3].
for sections 4 and 5, see again [M-S].

3.1 Singularities of Schubert varieties

Let G = SLn (k). Let T be a maximal torus in G, B a Borel subgroup


containing T and W the Weyl group of G. Recall the terminology in-
troduced in 2.1, Chap.2: Let B = P1 ∩ P2 ∩ . . . ∩ Pn−1 , where Pi are
the maximal parabolic subgroups of G containing B. If Li is the “hy-
perplane bundle” on the Grassmannian G/Pi (i.e. the ample generator
of Pic G/Pi ), then denoting its pullback to G/B also by Li we have
H 0 (G/Pi , Li ) ∼
= H 0 (G/B, Li ) = Vi (say), for 1 ≤ i ≤ n − 1. More-
over, if Wi denotes the Weyl group of Pi , then H 0 (G/Pi , Li ) has a basis
consisting of the Plücker coordinates pλ , λ ∈ W/Wi .

© Springer Science+Business Media Singapore 2016 and Hindustan Book Agency 2015 81
C.S Seshadri, Introduction to the Theory of Standard Monomials,
Texts and Readings in Mathematics 46, DOI 10.1007/978-981-10-1813-8_3
82 3. Applications

Also recall that we may identify Vi with the set of regular functions
f on G such that

(∗) f (gb) = Xi (b)f (g); b ∈ B, g ∈ G

for a suitable character Xi of B. Now if G denotes the Lie algebra of


G, then each Vi is also a G-module. Identifying each Y in G with the
corresponding right invariant vector field DY and using the above iden-
tification of Vi , it can be easily seen that DY f = Y f for every regular
function f on G as above.
Let X(τ ) be a Schubert variety in G/B. We intend to investigate
whether X(τ ) is smooth at the T -fixed points ew , w ∈ W. First we note
that

1. For λ, µ in W/Wi , pλ (eµ ) 6= 0 iff λ = µ.

2. Let Xα be a usual basis element (in the Chevelley basis) of G,


where α is a root. If Xα pα 6= 0, λ ∈ W/Wi , then Xα pλ = ±psαλ ,
where sα is the reflection corresponding to the root α.

The first remark is obvious, since {pλ | λ ∈ W/Wi } is a dual basis in


(∧i V )∗= H 0 (G/Pi , Li ) where V is an n-dimensional vector space. The
second remark is a consequence of SL2 theory, using the facts


(a) |hX, α∗ i| = 2(X,α)
(α,α) = 0 or 1, X being the weight of pλ .

(b) pλ is the lowest weight vector for the Borel subgroup B λ = λBλ−1 .

3.1.1 Theorem
If TX(τ ),e denotes the tangent space at the point e = eid of a Schubert
variety X(τ ) in G/B, then

dim TX(τ ),e = #{α ∈ Φ+ | τ ≥ sα }.

Proof. Consider the projection π : G → G/B and let X(τ ]) = π −1 (X(τ )).
To obtain the desired dimension of the tangent space, first we calculate
]). Using
the rank of the Jacobian matrix J(τ ) at the identity e in X(τ
0
the identification of H (G/P, Li ) as the set of regular functions on G
]) in G is defined by
satisfying (∗), it is easy to see that the ideal of X(τ
the regular functions on G corresponding to the Plücker coordinates pλ
3.1. Singularities of Schubert varieties 83

such that λ 6≤ τi , where τi is the natural projection of τ in W/Wi . We


denote these functions also by Pλ . Now if

B = {Xα | α ∈ Φ} ∪ {H1 , H2 , . . . , Hn−1 }

]) at e
is a Chevelley basis of G, then the Jacobian matrix Je (τ ) of X(τ
is of the form

Je (τ ) = [DY pλ ](e) = [Y pλ ](e)


 
Xα pν Xβ pν . . . H 1 pν H 2 pν . . .
 Xα pµ Xβ pν . . . H 1 pµ H 2 pµ . . . 
 
=  .. .. .. .. 
 . . . . 
Xα pδ Xβ pδ . . . H 1 pδ H 2 pδ . . . at e

where columns of the matrix are indexed by the Chevelley basis β and
the rows are indexed by λ ∈ W/Wi such that λ 6≤ τi . We claim that each
row in Je (τ ) has at most one nonzero element. For if Y = Hi , then

(Hi pλ )(e) = (cpλ )(e), where c is a constant.

However, since λ 6= id, the remark (1) above shows that (Hi pλ )(e) = 0,
1 ≤ i ≤ n − 1. On the other hand, if Y = Xα , α ∈ Φ and Xα pλ 6= 0,
then the remark (2) above implies that

Weight of Xα pλ = Weight of psαλ = Weight of pλ + α.

Therefore,

(Xα pλ )(e) 6= 0 ⇔ Weight of Xα pλ = Weight of pid Xα pλ 6= 0

Since weight of Xα pα 6= Weight of Xβ pλ for α 6= β, the claim follows.


We note also that since λ 6= id, i.e. pλ 6= a multiple of the lowest weight
vector, weight of Xα pλ 6= weight of pid , for α > 0. Thus we have

rank of Je (τ ) = Number of nonzero columns in Je (τ )


= # {α ∈ Φ | α > 0 and there is a λ ∈ W/Wi
1 ≤ i ≤ n − 1 such that X−α pλ = ±pid
and τi 6≤ λ} .

However, X−α pλ = ±psαλ = ±pid iff sα λ = id in W/Wi , i.e., iff λ = sα


in W/Wi . Hence
84 3. Applications

rank Je (τ ) = #{α > 0 | τi 6≥ sα in W/Wi } for some i


= #{α > 0 | τ 6≥ sα }
Now since the projection π : G → G/B is a smooth morphism, we have:
rank of Je (τ ) = Codimension of TX(τ
e ),e
= e ),e .
Codimension of TX(τ

This together with the fact dim G/B = #{α > 0} implies that
 
dim TX(τ ) , e = dim G/B − Codimension of TX(τe ),e

= dim G/B − rank of Je (τ )


= #{α > 0 | τ ≥ sα }.
Q.E.D.

3.1.2 Definition
For τ ∈ W, we define the length of τ, denoted by ℓ(τ ) to be the least
integer r such that τ = s1 s2 . . . sr , where si are simple reflections in W.
One knows that ℓ(τ ) = dim X(τ ) (see Chapter 4).

3.1.3 Corollary
X(τ ) is smooth iff ℓ(τ ) = #{α > 0 | τ ≥ sα }.

Proof. This immediately follows from the fact that X(τ ) is smooth iff
it is smooth at e. Q.E.D.

The above proof carries over to the case of any T -fixed point ew and
we obtain

3.1.4 Theorem
Let X(τ ) be a Schubert variety in G/B and let ew ∈ X(τ ). Let
Z(τ, w) = {α > 0 | τ ≥ wsα }.
If TX(τ ) , ew denotes the tangent space of X(τ ) at ew , then
dim TX(τ ),ew = #Z(τ, w).
3.2. Vanishing theorem 85

3.1.5 Corollary
X(τ ) is smooth at ew iff ℓ(τ ) = #Z(τ, w).

3.2 Vanishing theorem

Denote SLn (k) by G and let T be a maximal torus and let B be a


Borel subgroup of G containing T. It is known that if L = La is a line
bundle on G/B such that a ≥ 0 (i.e. a = (a1 , . . . , an−1 ) such that
ai ≥ 0, i = 1, . . . , n − 1), then H i (G/B, L) = 0 ∀ i > 0. This result is
due to G. Kempf (see [K]). In characteristic zero, this is classical and
can be deduced, for example, as a consequence of the Kodaira vanishing
theorem. We will now generalize this vanishing theorem to any Schubert
variety X(τ ) in G/B.

3.2.1 Theorem
If X is a union of Schubert varieties in G/B and L = La is a line bundle
on G/B such that a ≥ 0, then H i (X, L) = 0 ∀ i > 0.

Proof. We can reduce the proof of the theorem to the case where
X := X(τ ) is just a Schubert variety, for if X = X(τ1 ) ∪ X(τ2 ) is a
union of Schubert varieties such that dim X(τ1 ) ≥ dim X(τ2 ), then the
scheme theoretic intersection X(τ1 ) ∩ X(τ2 ) is reduced and a union of
Schubert varieties of dimension less than dim X(τ1 ) (Lemma (2.5.4) and
Theorem (2.7.2)). The sequence
0 → OX → OX(τ1 ) ⊕ OX(τ2 ) → OX(τ1 )∩Xτ2 → 0
is exact. If we tensor the sequence by L, we get
0 → L → L|X(τ1 ) ⊕ L|X(τ2 ) → LX(τ1 )∩X(τ2 ) → 0.
The map H 0 (LX(τ1 ) )⊕ H 0 (LX(τ2 ) ) → H 0 (LX(τ1 )∩X(τ2 ) ) is surjective. By
induction on dim X(τ1 ) and the number of irreducible components of
X as in the case of the Grassmannian, the claim follows by the corre-
sponding long exact cohomology sequence. Now suppose X = X(τ ) is a
Schubert variety in G/B and dim X(τ ) ≥ 1. Assume that if X(ω) is a
Schubert variety of dimension less than dim X(τ ), then H i (X(ω), L) =
0, ∀ i > 0. Let α be a simple root such that ω = sα τ < τ. Denote by
Z(ω, α) the fibre bundle P (α) ×B X(ω), where P (α) denotes the mini-
mal parabolic subgroup of G generated by B and G−α (see Chapter 4
(4.2.9)). The proof of the theorem depends on the following lemmas.
86 3. Applications

3.2.2 Lemma
If Ψ : Z(ω, α) → X(τ )((g, x) 7→ g.x) is the canonical map and L = La ,
where a ≥ 0, then H i (Z(ω, α), Ψ∗ (L)) = 0, ∀ i > 0.

Proof of the Lemma. We know that H i (X(ω), L) = 0, ∀ i > 0.


By the main theorem of standard monomial theory (Chap. 2, Thm.
6.1) we know that the restriction map H 0 (G/B, L) → H 0 (X(ω), L)
is surjective. Denote Ψ∗ (L) by M and let E denote the G-module
H 0 (G/B, L). The fibre type of the fibration π : Z(ω, α) → P1 is X(ω).
Recall that L|X(ω) ∼ = M|X(ω) (Chapter 2, (7.2), C). Hence, if Y is a
fibre of the fibration π, then H i (Y, MY ) = 0, ∀ i > 0. We deduce that
H i (Z(ω, α), M) ≃ H i (P1 , π∗ (M)) ∀ i > 0. Hence it is sufficient to show
that H i (P1 , π∗ (M)) = 0, ∀ i > 0. Since π : Z(ω, α) → P1 is a locally
trivial fibration, it is easily seen that π∗ (M) is a vector bundle over
P1 whose fibres are isomorphic to H 0 (X(ω), L). Furthermore, π∗ (M) is
simply the associated vector bundle to the fibre space P (α) → P1 and
the B-module H 0 (X(ω), L). We have the exact sequence of B-modules:

0 → ker → (E = H 0 (G/B, L)) → H 0 (X(ω), L) → 0.

Denote by ker, H(X(ω), L), E the vector bundles over P1 associated to


the objects in this exact sequence. We get the following exact sequence
of vector bundles over P1 .

0 → ker → E → π∗ (M) → 0.

Since H 0 (G/B, L) is a P (α)-module, the vector bundle E is trivial over


P1 , and hence, H i (P1 , E) = 0 ∀ i > 0. Since dim P1 = 1, we know
that H 2 (P1 , ker) = 0 and H i (P1 , π∗ (M)) = 0 i ≥ 2. By the long exact
cohomology sequence we get:

. . . → H 1 (P1 , E) → H 1 (P1 , π∗ (M)) → H 2 (P1 , ker) → . . .

and hence, H i (P1 , π∗ (M)) = 0, ∀ i > 0. Q.E.D.

3.2.3 Lemma
If Ψ : Z(ω, α) → X(τ ) is the canonical map and F = Lb , b > 0, is an
ample line bundle on X(τ ), then

H q (Z(ω, α), Ψ∗ (F m )) = H 0 (X(τ ), (Rq Ψ∗ OZ(ω,α) ) ⊗ F m )


3.2. Vanishing theorem 87

for q ≥ 0 and m ≫ 0.

Proof of the Lemma. Since Ψ is proper, the sheaves Rq Ψ∗ OZ(ω,α) are


coherent. By Serre’s theorem we know that H p (X(τ ), (Rq Ψ∗ OZ(ω,α) ) ⊗
F m ) = 0 for p > 0, q ≥ 0 and m ≫ 0. The corresponding Leray spectral
sequence:

H p (X(τ ), Rq Ψ∗ (Ψ∗ (F m ))) = H p (X(τ ), Rq Ψ∗ OZ(ω,α) ⊗ F m )


⇒ H p+q (Z(ω, α), Ψ∗ (F m ))

degenerates and hence, for m ≫ 0, we get

H q (Z(ω, α), Ψ∗ (F m )) = H 0 (X(τ ), (Rq Ψ∗ OZ(ω,α) ⊗ F m ).

Q.E.D.

3.2.4 Lemma
If Ψ : Z(ω, α) → X(τ ) is the canonical map and L = La , a ≥ 0, then
Ψ∗ : H i (X(τ ), L) → H i (Z(ω, α), Ψ∗ (L)) is an isomorphism.

Proof of the Lemma. By Lemma (3.2.2) we know that H q (Z(ω, α),


Ψ∗ (F )) = 0 ∀ q > 0 and for every line bundle F = Lb , b ≥ 0, on G/B.
In particular we know that H q (Z(ω, α), Ψ∗ (F m )) = 0 ∀ q > 0, m ≫ 0.
Now assume that b > 0 (⇔ F is ample). By Lemma (3.2.3), we conclude
that H 0 (X(τ ), (Rq Ψ∗ OZ(ω,α) )⊗F m ) = 0, ∀ q > 0 and m ≫ 0. For q > 0
denote Rq Ψ∗ OZ(ω,α) by G. Since G is a coherent OX(τ ) -module and F
is ample, the fact that H 0 (X(τ ), G ⊗ F m ) = 0, for m ≫ 0 implies that
G = 0. We conclude that Rq Ψ∗ OZ(ω,α) = 0, for q > 0. Recall that the
map Ψ : Z(ω, α) → X(τ ) is birational (see later 4.4.2) and that X(τ ) is
normal (Chapter 2, Theorem 6.1). Hence we have Ψ∗ OZ(ω,α) = OX(τ ) .
It follows that Ψ∗ (Ψ∗ (L)) = L. Since Rq Ψ∗ (Ψ∗ L) ≃ Rq Ψ∗ OZ(ω,α) ⊗ L =
0, ∀ q > 0, we conclude that H i (X(τ ), L) ≃ H i (Z(ω, α), Ψ∗ (L)), ∀ i ≥
0. Q.E.D.

3.2.5
The theorem follows now from Lemma (3.2.2) and Lemma (3.2.4). Q.E.D.
88 3. Applications

3.2.6
We want to give a sketch of another proof of Theorem 3.2.1. The proof
is based on the usual induction procedure used in many of the proofs
before. The main problem here is to prove the theorem for the starting
point of the induction, i.e. for L = La , where a = (0, a1 , . . . , an−1 ).

3.2.7 Another proof of Theorem 2.1.


Let m ≥ 1 and assume that the vanishing theorem holds for unions
of Schubert varieties of dimension less than m. As in the proof before,
it is sufficient to prove the vanishing theorem for Schubert varieties of
dimension m. Let X(τ ) be a Schubert variety in G/B of dimension m
and let α be a simple root such that ω = τ sα < τ. Let S be the set of
all simple roots. Arrange the maximal parabolic subgroups p1 , . . . , pn−1
of G in such a way that P1 is the maximal parabolic subgroup of G
generated by B and the subgroups G−β , where β ∈ S−{α}. With respect
⊗a
to this order we can write L as the tensor product L⊗a
1
1
⊗ . . . ⊗ Ln−1n−1 of
the pull backs of the very ample line bundles Li on G/Pi (see Chapter
2, (1.3)).
A Assume ai = 0 : By the choice of α we know that X(τ ) is saturated
for the locally trivial P1 -fibration Φ : G/B → G/P (α) (recall that
P (α)/B ≃ P1 ). By Chevalley’s formula (Chapter 4, (5.9)), the
restriction of L1 to the typical fibre of this fibration is of degree 1
and the restriction of L−11 to the typical fibre is of degree −1. We
have the exact sequence
0 → L−1
1 → OX(τ ) → OH(τ ) → 0.

Tensoring the exact sequence L we get : 0 → L′ → L → L|H(τ ) → 0,


⊗a2 ⊗an−1
where L′ = L−1 i
1 ⊗L2 ⊗. . .⊗Ln−1 . If we can show that H (X(τ ), L ) =

0, ∀ i ≥ 0, then the vanishing of H i (X(τ ), L) for i ≥ 1 follows from the


corresponding long exact cohomology sequence, since H i (H(τ ), L) = 0 ∀
i ≥ 1 by induction on the dimension (recall that X(τ ) is normal (Chap.2,
Thm. (2.6.1) and hence H(τ ) is reduced (Chap 2, Proposition 2.5.7)
and is a union of Schubert subvarieties of X(τ ) of codim. 1). Denote
by Y the image of X(τ ) under the fibration Φ : G/B → G/P (α). Then
Φ : X(τ ) → Y is a locally trivial P1 -fibration such that the restriction of
L′ to the typical fibre P1 is of degree −1. The vanishing of H i (X(τ ), L′ )
for i ≥ 0 and therefore the vanishing of H i (X(τ ), L), for i ≥ 1 follows
from:
3.3. Character formula 89

3.2.8 Lemma
Let X and Y be normal varieties and let Φ : X → Y be a locally trivial
P1 -fibration. If L is a line bundle on X such that the restriction of L to
the typical fibre is of degree −1, then H i (X, L) = 0, ∀ i ≥ 0.

Proof of the Lemma. Let U ⊂ Y open and affine such that the
fibration is trivial. Since the restriction of L to the typical fibre is of
degree −1, the restriction of L to U × P1 (֒→ X) can be written as
p∗1 (OP1 (−1)) ⊗ p∗2 (M), where p1 : U × P1 → P1 and p2 : U × P →
U are the projection mapsP and M is a line bundle on U. Hence we
get H i (U × P1 , L) = H k (U, M) ⊗ H ℓ (P1 , OP1 (−1)) = 0 ∀ i > 0
k+ℓ=i
(Künneth formula). As an immediate consequence we conclude that
Rq φ∗ L = 0 ∀ q ≥ 0 and hence H i (X, L) ∼ = H i (Y, φ∗ L), ∀ i ≥ 0. Now
let U1 . . . Un be an affine open cover of Y such that the fibration is
trivial over Ui , i = 1, . . . , n. Since Y is integral, every finite intersection
of the U1 , . . . , Un is open, affine and the fibration is trivial over the
intersection. Hence the cohomology of φ∗ L restricted to one of these
intersections vanishes and we can compute H i (Y, φ∗ L) by computing the
Čech cohomology H i (ω, φ∗ L), where ω = (Ui )i=1,...,n . Since Rq φ∗ L = 0,
it follows that Ȟ i (ω, φ∗ L) = H i (Y, φ∗ L) = H i (X, L) = 0, ∀ i ≥ 0.
Q.E.D.

3.2.9
Now assume a1 ≥ 1 and suppose that H i (X(τ ), Lb ) = 0 ∀ i > 0 if
b = (b1 , . . . , bn−1 ) such that b1 < a1 . If we tensor the exact sequence
0 → L−1
1 → OX(τ ) → OH(τ ) → 0 by L, we get the exact sequence

0 → L′ → L → L|H(τ ) → 0,

where L′ denotes the line bundle L⊗a 1


1 −1
⊗ L⊗a
2
2
⊗ . . . × L⊗an −1
n−1 . Since
H(τ ) is a union of Schubert varieties of dimension less than dim X(τ )
and H i (X(τ ), L′ ) = 0 ∀ i ≥ 1, we conclude from the corresponding long
exact cohomology sequence that H i (X(τ ), L) = 0, ∀ i ≥ 1. Q.E.D.

3.3 Character formula

Let G denote SLn (k) and fix a maximal torus T in G. Denote by B a


Borel subgroup of G containing T and let S be the set of simple roots
90 3. Applications

corresponding to the choice of B. Recall that we denote by La , a =


⊗a
(a1 , . . . , an−1 ), the line bundle L⊗a
1
1
⊗ . . . ⊗ Ln−1n−1 on G/B. We want to
derive a formula for the character of the T -module H 0 (X(τ ), La ), where
X(τ ) is a Schubert variety in G/B and a ≥ 0 (i.e., a1 ≥ 0, . . . , an−1 ≥ 0).

3.3.1
Let λ be an element of the character group X(τ )(= X(B)) of T (see
Chapter 4, (1.1)) such that λ is a dominant weight (Chapter 4, (1.7)).

Definition Denote by Lλ the line bundle on G/B whose total space is


the quotient of G × k by the equivalence relation (g, x) ∼ (g.b, −λ(b)x),
where g ∈ G, b ∈ B and x ∈ k.
The G-module H 0 (G/B, Lλ ) is a finite dimensional rational G-
module. Note that if ̟1 , . . . , ̟n−1 are the fundamental weights (Chap-
ter 4, (1.7)) and λ = a1 ̟1 + · · · an−1 ̟n−1 , then Lλ = La , a = (a1 , . . . ,
an−1 ), in our usual notation. In the following we will use the notation
Lλ instead of La .

Remark If char k = 0, we know by the Borel-Weil theorem that


H 0 (G/B, Lλ ) is a finite dimensional irreducible rational G-module of
highest weight i(λ), where i denotes the Weyl involution (Chapter 4,
section 1); and every finite dimensional irreducible rational G-module
can be realized as the space of regular cross sections for some line bundle
Lλ , where λ is a dominant weight.

3.3.2
Denote by Z[Hom(T, GL1 (k))] the group ring of the multiplicative group
exp[Hom(T, GL1 (k))] = {exp λ | λ ∈ Hom(T, GL1 (k))}. If X(τ ) is a
Schubert variety and λ is a dominant weight, then let H 0 (X(τ ), Lλ ) =
⊕i (H 0 (X(τ ), Lλ ))χi be a decomposition of the T -module. H 0 (X(τ ), Lλ )
into weight spaces (Chapter 4, (4.1.1)).
P
Definition Denote by Char H 0 (X(τ ), Lλ ) = i exp(ni χi ),
ni = dim(H 0 (X(τ ), Lλ ))χi , the character of the T -module H 0 (X(τ ), Lλ ).
The character is an element of Z[Hom(T, GL1 (k))]. We want to deduce
a formula for Char H 0 (X(τ ), Lλ ) in terms of τ, λ and the root system φ
of G.
3.3. Character formula 91

3.3.3 Definition
Denote by Lsα , α ∈ S, the linear operator:

Lsα : Z[Hom(T, GL1 (k))] → Z[Hom(T, GL1 (k))]

defined by:
exp λ − exp (sα (λ))
Lsα (exp λ) =
1 − exp α
Let Msα be the operator defined by

Msα (exp λ) = exp ρ Lsα (λ − ρ).

where ρ is half of the sum of all positive roots.

3.3.4 Theorem
(Demazure Character Formula). If λ is a dominant weight and τ, ω ∈ W
such that ω < τ and ω = sα τ, α ∈ S, then Char H 0 (X(τ ), Lλ ) =
Msα (Char H 0 (X(ω), Lλ ).

Proof. Let P (α) denote the parabolic subgroup of G generated by G−α


and B. Let L(α) be the subgroup of P (α) which is isomorphic to SL2 (k)
and is generated by G−α and G. Denote by B(α) the Borel subgroup
B ∩ L(α) of L(α). Let Z(ω, α) be the filtration L(α) ×B X(ω) = (≃
P (α) ×B X(ω)) over P1 . We have the following canonical maps:
Ψ
Z(ω, α) → X(τ )
π↓
P1

We know that H 0 (X(τ ), Lλ ) = H 0 (Z(ω, α), Ψ∗ (Lλ )) (Chapter 2, The-


orem (7.2)), and H i (Z(ω, α), Ψ∗ (Lλ )) = 0 for i > 0 (Lemma 3.2.2).
We deduce that H 0 (X(τ ), Lλ ) ≃ H 0 (P1 , π∗ (Ψ∗ (Lλ ))). We know that
the fibres of π∗ (Ψ∗ (Lλ )) are vector spaces isomorphic to the B-module
W = H 0 (X(ω), Lλ ) and the vector bundle π∗ (ψ ∗ (Lλ )) on P1 is the vec-
tor bundle W = P (α) ×B W on P1 associated to the module W (see
proof of Lemma 2.2). If V is any vector bundle on P1 associated to a
B-module V, then define:

Char V = Char H 0 (P1 , V ) − Char H 1 (P1 , V ).


92 3. Applications

Since H1 (P1 , W ) = 0 (see proof of 3.2.2), we get: Char(X(τ ), Lλ ) =


Char H(P1 , W ). It remains to show that Char W = Msα (Char W ) (re-
call that W = H 0 (X(ω), Lλ )). The B-module W has a filtration of
B-submodules W = Wm ⊃ Wm−1 ⊃ . . . ⊃ W1 ⊃ W0 = (0) such that
Wi /Wi−1 , i ≥ 1, are 1-dimensional B-modules. It is easy to see that
if 0 → V1 → V → V2 → 0 is an exact sequence of B-modules, then
P
m
CharV = CharV 1 = CharV 2 . Hence we get CharW = CharWi /Wi−1 .
i=1
Thus it suffices to show that if V is any 1-dimensional B-module, then
CharV = Msα (Char V ). In this case we are reduced to check the claim
for G = SL2 (k). The action of T on V is given by an element θ of
Hom(T, GL1 (k)) and hence, we have Char V = exp θ. Now
exp(θ − ρ) − exp(sα (θ − ρ))
Msα (exp θ) = exp ρ.
1 − exp α
Since sα (θ − ρ) = θ − ρ − [hθ, αi − 1]α, we get:

exp(θ − ρ) − exp(θ − ρ − [hθ, αi − 1]α)


Msα (exp θ) = exp ρ
1 − exp α
1 − exp(−[hθ, αi − 1]α
= exp θ.
1 − exp α
a) If n = −hθ, αi ≥ 0, then Msα (exp θ) = exp θ + exp(θ + α) +
· · · + exp(θ + nα). The line bundle on P1 = SL2 (k)/B corre-
sponding to θ has degree n. The corresponding SL2 (k)-module of
global sections has θ as lowest weight. Since H 1 (P1 , V ) = 0, we
know by the SL2 (k)- representation theory that Char H 0 (P1 , V ) =
Msα (exp θ).
b) If hθ, αi = 1, the equality 0 = Msα (exp θ) = Char V trivially holds
since H 0 (P1 , V ) = 0.
c) If hθ, αi ≥ 2, then H 0 (P1 , V ) = 0 and H 1 (P1 , V ) is dual to
H 0 (P1 , V ∗ ⊗ K) where V ∗ is the dual space to V and K denotes
the canonical sheaf on P1 . The action of T on V ∗ is given by (−θ).
The degree of the line bundle V ∗ ⊗ K is non negative. Since the
character of a G-module is equal to the negative of the character
of its dual module, we get by a):

Char V = Char H 1 (P1 , V ) = Char (H 0 (P1 , V ∗ ⊗K))∗ = Msα (exp θ).


3.4. Ideal theory of Schubert varieties 93

So, Char V = Msα (Char V ) for any one-dimensional B-module V and


hence, Char V = Maα (Char V ) for any finite dimensional rational B-
module V. Thus it follows that Char W = Char H 0 (X(τ ), Lλ ), where
W = H 0 (X(ω), Lλ ), and hence, Char H 0 (X(τ ), Lλ ) = Msα (Char H 0
(X(ω), Lλ )). Q.E.D.

3.3.5 Corollary
Let τ ∈ W and let λ be a dominant weight. If τ = sα1 . . . . .sαr is a
reduced decomposition, then

Char H 0 (X(τ ), Lλ ) = Msα1 . · · · .Msαr (exp(−λ)).

(Note that the right hand side is independent of the reduced decom-
position.)

Proof. By the theorem it remains to show that Char H 0 (X(id), Lλ ) =


exp(−λ). Since B acts through the character (−λ) on the fibres of Lλ at
e(id) = X(id) (Recall the construction of Lλ described in (3.3.1)), the
assertion follows. Q.E.D.

Remark (Char k = 0) : If ω0 is the unique element of the Weyl


group such that ω0 (φ+ ) = φ− (see 4.1.3), then X(ω0 ) = G/B. Let
ω0 = sα1 . · · · .sαN (N = #(φ+ )) be a reduced decomposition of ω0 . If
λ is a dominant weight, then denote by Vλ the finite dimensional ir-
reducible rational G-module with highest weight λ. It follows from the
above corollary that Char H 0 (X(ω0 ), Lλ ) = Char Vi(λ) = Char (Vλ )∗ =
Msα1 . · · · .Msαr (exp(−λ)).

3.4 Ideal theory of Schubert varieties

Let G, B, T be as in the previous sections. Let P1 , P2 , . . . , Pn−1 be the


maximal parabolic subgroups of G containing B, and let Pmi be the
projective space in which G/Pi is embedded by the Plücker embedding.
Let Li denote the ample generator of Pic Pmi . Denoting the restriction
of Li to G/Pi also by Li , the homogeneous coordinate ring of G/Pi is

Ri = ⊕m≥0 H 0 (G/Pi , Lm
i ), 1 ≤ i ≤ n − 1.
The natural maps G/B → G/Pi , 1 ≤ i ≤ n − 1, induce the “diagonal
embedding” G/B → π1≤i≤n−1 G/Pi which in turn gives the embedding
94 3. Applications

G/B → π1≤i≤n−1 Pmi . As usual, we will denote the pull back of Li to


π1≤i≤n−1 Pmi and G/B via the natural projections also by Li .
Let Z denote the multi-projective space π1≤i≤n−1 Pmi and let S be
its multi-homogeneous coordinate ring, i.e.,

S = ⊕a≥0 H 0 (Z, La )
a
where a = (a1 , a2 , . . . , an−1 ), La = La11 ⊗ La22 ⊗ . . . ⊗ Ln−1
n−1
and a ≥ 0
means ai ≥ 0 for all i. In terms of coordinates we have

S = k[x10 , . . . , x1m1 ; . . . ; xi0 , . . . , ximi ; . . . ; . . .]

Let
A = Spec S = π1≤i≤n−1 Ami +1 .
where Ami +1 = Spec k[xi0 , . . . , ximi ],
If T ′ denotes the torus group {(t1 , . . . , tn−1 )|ti ∈ Gm }, then we have
a canonical action of T ′ on A, namely, multiplication by ti on the com-
ponent Ami +1 . Furthermore, if A◦ denotes the open subvariety of A
formed by the points x = (xi ), xi ∈ Ami +1 such that xi 6= 0 for all
i, then T ′ operates freely on A◦ and Z identifies with the orbit space
A◦ /T ′ .

3.4.1 Definition
Let X be a closed subscheme of the multiprojective space Z. The ideal
of X (in S), denoted I(X) is defined to be the ideal generated by all
f ∈ Sa , a ≥ 0, such that f vanishes on X (as a section of a line bundle
on Z). We call the scheme X̂ = Spec S/I(X) the multicone over X.
It is clear that the ideal I(X) of X is a multigraded ideal in S. On the
other hand, if J is a multigraded ideal in S, we can associate to J a closed
subscheme V (J) of Z as follows: The ideal J determines a T ′ -stable sheaf
of ideal J˜ on A. The restriction of J˜ to A◦ gives a sheaf of ideals on
A◦ /T = Z, which in turn defines a closed subscheme of V (J) of Z.
Alternatively and more concretely, V (J) can also be defined as follows:
Z can be covered by affine open subsets of the form U1 × U2 × . . . × Un−1 ,
where Ui is an affine open subset of Pm1 defined by xiki 6= 0, i.e.,
" #
xi0 ximi
Ui = Spec k i , · · · , i , 1 ≤ i ≤ n − 1.
xk i xk i
3.4. Ideal theory of Schubert varieties 95

We define V (J) locally in U1 ×U2 ×. . .×Un−1 to be given by the ideal


(in the coordinate ring of Ui × U2 × . . . × Un−1 ) generated by elements
of the form
F
, F ∈ J ∩ Sa .
π1≤i≤n−1 , (xiki )ai

3.4.2 Lemma
If J1 and J2 are multi-homogeneous ideals in S, then

(i) V (J1 + J2 ) = V (J1 ) ∩ V (J2 )

(ii) V (J1 ∩ J2 ) = V (J1 ) ∪ V (J2 ),

the intersection and union on the right hand side of (i) and (ii) being
scheme theoretic.

Proof. By definition, it suffices to verify the lemma locally in the affine


open sets of U1 × U2 × . . . × Un−1 . Q. E. D.
Consider the multigraded ring
M
R= H 0 (G/B, La ),
a≥0

a
where La = La11 ⊗ La22 ⊗ . . . ⊗ Ln−1n−1
. We know that the natural map
H 0 (Z, La ) → H 0 (G/B, La ) is surjective for all a ≥ 0 (for example, it
follows from (2.6.1)). Hence the natural map Φ : S → R is surjective
and its kernel is the ideal I(G/B) of G/B in S.
If J is a multigraded ideal of R, then V (φ−1 (J)) is a closed sub-
scheme of G/B, which we denote by V (J). Conversely, if X is a closed
subscheme of G/B, the ideal of X in R is the ideal generated by all
multi-homogeneous elements f ∈ R vanishing on X. This ideal is in fact
the image of I(X) under the natural map φ : S → R, which we also
denote by I(X).

3.4.3 Lemma
Let Ji be a homogeneous ideal of the homogeneous coordinate ring Ri
of G/Pi and let J i be the multi-homogeneous ideal of R generated by
the image of Ji under the map Ri → R induced by the natural map
96 3. Applications

η : G/B → G/Pi . If V (Ji ) denotes the closed subscheme of G/Pi defined


by the ideal Ji , then
V (J i ) = η −1 (V (J)) (scheme theoretically).

Proof. Restricting V (J i ) and V (Ji ) to suitable affine open subsets


of G/B and G/Pi respectively, we are reduced to the following affine
situation: η : Spec A → Spec Ai , Ji is an ideal of Ai and J i is the ideal
generated by the image of Ji under the map Ai → A. But in this case it
is clear that
V (J i ) = η −1 (V (J)) (scheme theoretically).
Q.E.D.
Let W be the Weyl group of G and Wi , 1 ≤ i ≤ n − 1, be the
Weyl group of Pi . Recall that a subset Ti of W/Wi is a left half space if
λ ∈ W/Wi , µ ∈ Ti and λ ≤ µ implies λ ∈ Ti .

3.4.4 Theorem
Let T = ∪1≤i≤n−1 Ti , where Ti is a left half space in W/Wi (some Ti could
be empty). If J is the ideal of R generated by the Plücker coordinates
P τ, τ 6∈ T, then the closed subscheme V (J) of G/B is reduced and is in
fact a union of Schubert varieties in G/B.

Proof. Let Ji (resp. J i ) be the ideal in Ri (resp. R) generated by the


Plücker coordinates pτ , τ 6∈ Ti . By (1.4.6), Ji is the ideal of the union of
Schubert varieties in G/Pi defined by the left half space Ti . Thus V (J)
is reduced and is a union of Schubert varieties. Hence, using (3.4.3),
V (J i ) = η −1 (V (J)) is reduced (since η : G/B → G/Pi is smooth) and
is a union of Schubert varieties in G/B, where η is the natural map
G/B → G/Pi . Now, since
V (J) = V (J 1 ) ∩ . . . ∩ V (J n−1 ).
the theorem follows from lemma 5.4 of Chapter 2. Q.E.D.

3.4.5 Remark
(i) If Q is any parabolic subgroup of G, then, analogous to the case
of G/B, we can define the multigraded ring of G/Q and the above
theorem has an obvious extension to this case.
3.5. The variety of complexes 97

(ii) The ideal theory of Schubert varieties in G/B does not extend to
the multigraded ring R in the same way as it does in the case of
G/P (P a maximal parabolic subgroup of G). For example, we
are faced with difficulties of the following nature:

(a) If X1 and X2 are Schubert varieities in G/B, then is it true


that
I(X1 ∩ X2 ) = I(X1 ) + I(X2 )?

(b) Let X(τ ) be a Schubert variety in G/B and let Xi (τ ) be the


image of X(τ ) in G/Pi , where Pi is a maximal parabolic sub-
group and τ is the image of τ in W/Wi . Let Rτ = R/I(X(τ )),
the multigraded ring of X(τ ). Let H(τ ) = X(τ ) ∩ {pτ = 0}
(scheme theoretically) and let Ir (H(τ )) be the ideal of H(τ )
in Rτ , i.e., the image of I(H(τ )) in Rτ . Then is it true that

(∗) Iτ (H(τ )) = pτ Rτ ?

(iii) Huneke and Lakshmibai [H-L] have shown that the multigraded
ring Rτ of a Schubert variety X(τ ) in G/B is Cohen-Macaulay.
This result can also be proved using the results of Chapter 2. The
main observation is that (∗) holds for a suitable maximal parabolic
subgroup Pi of G.

3.5 The variety of complexes

Let V1 , V2 , . . . , Vr+1 be finite dimensional vector spaces over k with dim


Vi = ni , 1 ≤ i ≤ r + 1, and let A denote the affine space ⊕1≤i≤r Hom (Vi ,
Vi+1 ). Clearly, the coordinate ring of A is the polynomial ring A =
k[Y (1) , Y (2) , . . . , Y (r) ], where each Y (i) is an ni+1 × ni matrix of inde-
terminates.

3.5.1 Definition
The variety of complexes C ⊂ A is defined to be the set of complexes of
the form
f1 f2 fr
V1 −→ V2 −→ · · · −→ Vr+1 , i.e.,

C = {(f1 , f2 , . . . , fr ) ∈ A | fi+1 ◦ fi = 0, 1 ≤ i ≤ r − 1}.


98 3. Applications

To be precise, C is the subscheme of A defined by the ideal I in A


generated by the quadratic forms, which are the entries of the matrix
products Y (i+1) Y (i) , 1 ≤ i ≤ r − 1.
In what follows, for the sake of simplicity of notation, we shall assume
that r = 2, i.e.,

A = Hom (V1 , V2 ) ⊕ Hom (V2 , V3 ) and C = {(f1 , f2 ) ∈ A | f2 ◦ f1 = 0}.

However, the results we prove are true in general for any variety of
complexes. Although generality is lost in restricting to the above special
case, the essential ideas of the proofs are preserved. Our proofs easily
generalize to the general case; for details, see [M-S].
Let n = n1 + n2 + n3 and G = SLn (k). Let Q be the following
parabolic subgroup of G :
  
 A1 ∗ ∗ 
A ∈ GLni (k),
Q =  0 A2 ∗  i
 1≤i≤3 
0 0 A3
Let Z be the unipotent radical of the parabolic subgroup of G op-
posite to Q, i.e.
  
 I1 0 0 
I is the ni × ni identity matrix,
Z =  ∗ I2 0  i
 1 ≤i≤3 
∗ ∗ I3
Recall the well known fact that the restriction of the natural mor-
phism G → G/Q to Z is an isomorphism of Z onto an affine open subset
of G/Q which is called the opposite big cell of G/Q. We identify Z with
its image in G/Q and denote the image also by Z.
Let P1 and P2 be the maximal parabolic subgroups of G of the shape:

  
 ∗ ∗ ∗ 
P1 =  
0 ∗ ∗ ∈ SLn (k)
 
0 ∗ ∗
  
 ∗ ∗ ∗ 
P2 =  ∗ ∗ ∗  ∈ SLn (k)
 
0 0 ∗
Clearly, Q = P1 ∩ P2 . If Pi , i = 1, 2, denotes the projective space
in which G/Pi is embedded by the Plücker embedding, then we have a
natural embedding of G/Q in P1 × P2 given by
3.5. The variety of complexes 99

G/Q → G/P1 × G/P2 → P1 × P2 .


Composing this with the natural map G → G/Q, we get a map
ε : G → P1 × P2 whose restriction to Z maps Z isomorphically onto its
image. Identifying Z and G/Q with their images in P1 × P2 it is easy
to see that

Z = G/Q ∩ (Z1 × Z2 ),
where Zi , i = 1, 2, is the opposite big cell in G/Pi (see (1, Sec. 6)). This
allows us to speak of a Plücker coordinate as a function on Z.
The following theorem is a consequence of a stronger result that will
be proved later (see (3.5.5)). However, this is given here as it makes
the connection between the variety of complexes and Schubert varieties
more intuitive.

3.5.2 Theorem
Let C be the variety of complexes in A = Hom (V1 , V2 ) ⊕ Hom (V2 , V3 )
with dim Vi = ni 1 ≤ i ≤ 3. If n = n1 +n2 +n3 , then C can be embedded
in G/Q (Q a parabolic subgroup of G = SLn (k)) in such a way that
its closure C (set theoretically) is a union of Schubert varieties in G/Q
which are in fact Q-stable.

Proof. Let Q and Z be as defined above. Let


  
 I1 0 0
Ii is the ni × ni identity matrix for 1 ≤ i ≤ 3, 
Y =  X1 I2 0 
 and Xj ∈ M (nj+1 , nj ), j = 1, 2, 
0 X2 I3

Identifying Y with the affine space A, we get an embedding of A


(and hence of C) into Z, under which the points of C correspond to
the matrices in Y with X2 X1 = 0. Since Z is embedded in G/Q, this
gives an embedding of C in G/Q, which, we shall show, has the desired
property. As usual, we shall denote the images of C and Y in G/Q by
the same letters.
Recall that Qred , the reductive part of Q is given by:
  
 A1 0 0 
A ∈ GL (k)
Qred =  0 A2 0  i ni
 1≤i≤3 
0 0 A3
100 3. Applications

The canonical left action of G on G/Q induces an action of Qred on


G/Q, and since
    
A1 0 0 I1 0 0 I1 0 0
 0 A2 0   X1 I2 0  =  A2 X1 A−1 1 I2 0 
−1
0 0 A3 ∗ X2 I3 ∗ A3 X2 A2 I3

A1 0 0
 0 A2 0 
0 0 A3
it follows that Qred acts on Z; moreover, the above computation also
shows that Qred acts on Y and C. Hence, Qred C ⊆ C, where C is the
closure of C in G/Q.
Let Qµ denote the unipotent part of Q, i.e.
  
 I1 ∗ ∗ 
Q µ =  0 I2 ∗  ∈ Q
 
0 0 I3
To show that C is a union of Schubert varieties it suffices to show
that if X ∈ C and g is in an open neighborhood of the identity in Qµ ,
then gX ∈ C. For then, it would follow that Qµ C ⊆ C which together
with Qred C ⊆ C implies that QC ⊆ C, i.e. C is Q-stable and hence, is
a union of Schubert varieties.
First, we claim that Y is stable under Q. In fact, let S be the subset
of G consisting of matrices of the form
 
∗ ∗ ∗
 ∗ ∗ ∗ 
0 ∗ ∗
It is easy to see that S is Q-stable under the left and right actions of
Q on S. This implies that the image of S in G/Q is a Schubert variety,
say X(τ ), and Y = Z ∩ X(τ ) which proves the claim.
Now let
   
I 1 U1 0 I1 0 0
g =  0 I2 U2  and X =  X1 I2 0 
0 0 I3 0 X2 I3
with X2 X1 = 0. We have
 
I1 + U1 X1 ∗ ∗
gX =  X1 (I2 + U2 X2 ) ∗ 
0 X2 I3
3.5. The variety of complexes 101

Since Y is Q-stable, we can find A1 , A2 , A3 , B1 , B2 such that

  
I1 + U1 X1 ∗ ∗ A1 B1 ∗
 X1 I2 + U2 X2 ∗   0 A2 B2 
0 X2 I3 0 0 A3
 
(I1 + U1 X1 )A1 ∗ ∗
=  X1 A1 X1 B1 + (I2 + U2 X2 )A2 ∗ 
0 X2 A2 X2 B2 + A3
is in Q, i.e.,
(i) (I1 + U1 X1 )A1 = I1

(ii) X1 B1 + (I2 + U2 X2 )A2 = I2

(iii) X2 B2 + A3 = I3 .
and the matrix blocks above the diagonal are zero. Thus to show that
gX ∈ C (in G/Q) it suffices to show that

(X2 A2 )(X1 A1 ) = 0

i.e.
X2 A2 X1 = 0.
Since in a suitable neighborhood of the identity in Qµ , I2 + U2 X2 is
invertible, from (ii) we have in this neighborhood

A2 = (I2 + U2 X2 )−1 (I2 − X1 B1 )


⇒ X2 A2 X1 = X2 (I2 + U2 X2 )−1 (I2 − X1 B1 )X1
Set X2′ = X2 (I2 + U2 X2 )−1 .

Then

X2 = X2′ (I2 + U2 X2 )
⇒ 0 = X2 X1 = X2′ (I1 + U2 X2 )X1
⇒ X2′ X1 = 0.

Thus we have :

X2 A2 X1 = X2′ (I2 − X1 B1 )X1


= 0.
102 3. Applications

Q.E.D.
Let W be the Weyl group of G and let Wi be the Weyl group of Pi ,
i = 1, 2. If m1 = n1 and m2 = n1 + n2 , then as usual, we can identify
W/Wi , i = 1, 2, with the set

I(mi , n) = {α = (α1 , α2 , . . . , αmi ) | 1 ≤ α1 < α2 < · · · < αmi ≤ n}.

3.5.3 Lemma
The ideal I defining the variety of complexes C (in the coordinate ring
of Z) is generated by the Plücker coordinates pα on Z, where α is of the
form:

(i) either α = (α1 , . . . , αmi ) ∈ I(m1 , n) such that αm1 > m2 , or

(ii) α = (α1 , . . . , αm2 ) ∈ I(m2 , n) such that αs 6= s at least for


one s with 1 ≤ s ≤ m1 , or equivalently, α is not of the form
(1, 2, . . . , m1 , αm1 , . . . , αm2 )

Proof. Recall that Z is the affine space consisting of matrices of the


form  
I1 0 0
[zij ] =  X1 I2 0  ,
∗ X2 I3
Ii being the ni × ni identity matrix, 1 ≤ i ≤ 3, and as a subvariety of Z,
the variety of complexes C is defined by the equations

(a) zij = 0, i > m2 , j ≤ m1

(b) X2 X1 = 0.

Thus we have to show that in the coordinate ring of Z, the ideal


generated by the Plücker coordinates pα , where α is given by (i) or (ii)
is the same as the ideal generated by the entries of X2 X1 (considered
as functions on Z) and the coordinate functions zij with i > m2 and
j ≤ m1 .
For U ∈ Z, let
   
I1 I1 0
U1 =  X1  and U2 =  X1 X2 
∗ ∗ X2
3.5. The variety of complexes 103

Note that the Plücker coordinates pα on Z with α ∈ I(m, n) (resp.


α ∈ I(m2 , n)) can be identified with the determinant functions on the
set of n × m1 (resp. n × m2 ) matrices of the form U1 (resp. U2 ), i.e., for
α ∈ I(m1 , n) (resp. α ∈ I(m2 , n)), we have

pα (U ) = ± (Determinant of the) αth minor of U1 (resp. U2 )).

With this observation, it is easy to see that each coordinate function


zij with i > m2 , j ≤ m1 appears as a ±pα for some α of the form (i),
and conversely, for every such α, pα is in the ideal generated by zij with
i > m2 , j ≤ m1 .
Thus to complete the proof, it suffices to show that modulo the
functions zij as in (a), the ideal generated by the entries of X2 X1 is the
same as the ideal generated by pα with α as in (ii); in other words, if Z
is the set of n × n matrices of the form
 
I1 0 0
U =  X1 I2 0  ,
0 X2 I3

then we must show that in the coordinate ring of Z, the ideal generated
by the entries of X2 X1 is the same as the ideal generated by the Plücker
coordinates pα on Z with α as in (ii). To this end, we note, as before,
that such a pα on Z is given by

pα (U ) = ± (Determinant of the) αth minor of U2 )

where  
I1 0
U 2 =  X1 I2 
0 X2
Since
   
I1 0   I1 0
 X1 I2  I1 0
= 0 I2  = U2′ say,
−X1 I2
0 X2 −X2 X1 X2

we have

pα (U ) = ± (Determinant of the) αth minor of U2′ ).

Now we recall from (1.6.6) that the map α 7→ (α′ , α′′ ) which sends
each α 6= αmin = (1, 2, . . . , m2 ) to its associated dual pair (α′ , α′′ ) sets
104 3. Applications

up a bijection between αth minors of U2′ , α 6= αmin , and the set of minors
of the matrix [−X2 X1 X2 ]. Moreover, by (1.6.7), this bijection is such
that the Plücker coordinate ±pα , α 6= αmin , is equal to the determinant
function pα′ ,α′′ , on [−X2 X1 X2 ]. Now if α is as in (ii), then it is clear
that the minor of −X2 X1 X2 corresponding to the pair (α′ , α′′ ) contains
a column whose entries are in the matrix X2 X1 . This shows that each
pα with α as in (ii) is in the ideal generated by the entries of X2 X1 .
Conversely, it is obvious that each entry of X2 X1 can be obtained as a
± Determinant of the αth minor of U2′ ) = ±pα (U ) for some α as in (ii).
This completes the proof of the lemma.

3.5.4 Definition

For a Schubert variety X in G/Q, the affine open subset X ∩ Z of X (or


the closed subset X ∩ Z of the opposite big cell Z of G/Q) is called the
opposite big cell of X.

3.5.5 Theorem

The variety of complexes C is reduced and its irreducible components


are the opposite big cells of suitable Schubert varieties in G/Q.

Proof. Let

T1 = {α = (α1 , . . . , αm1 ) ∈ I(m1 , n) | αm1 ≤ m2 },


T2 = {α = (α1 , . . . , αm2 ) ∈ I(m2 , n) | αi = i for 1 ≤ i ≤ m1 }

and let T = T1 ∪ T2 . It is easy to see that T1 and T2 are left half


spaces in I(m1 , n) and I(m2 , n) respectively. Now if J is the ideal in the
multigraded ring R of G/Q generated by the Plücker coordinates pα ,
α 6∈ T, then by (3.4.3) (also see (3.4.4,i)), the closed subscheme V (J) of
G/Q is reduced and is a union of Schubert varieties in G/Q. But, by the
preceding lemma, the restriction of V (J) to Z is precisely C. Hence, C is
reduced and its components are opposite big cells of Schubert varieties
in G/Q. Q.E.D.
As an immediate consequence of (2.6.1), (3.4.5.iii) and (3.5.3), we
have
3.5. The variety of complexes 105

3.5.6 Corollary
The irreducible components S of the variety of complexes C are normal
and Cohen-Macaulay. Further, the restrictions of standard monomials
(on Z) to C (resp. S) generate the coordinate ring of C (resp. S) and
the nonzero standard monomials on S form a basis for the coordinate
ring of S.
Chapter 4

Schubert varieties in G/Q

Let G be a semisimple, connected algebraic group over an algebraically


closed field k of arbitrary characteristic. Choose a maximal torus T in
G and let Q ⊃ T be a parabolic subgroup of G. Denote the Weyl group
of G by W and denote by WQ the Weyl group of Q with respect to T.
We are going to prove in this more general context the equivalence of
the partial orders on W/WQ stated in Chapter 2. We will show that
every Schubert variety X(ω) in G/Q is smooth in codim 1 and that
the normalization map π : X(ω)^ → X(ω) is bijective. We will also
prove Chevalley’s multiplicity formula. In the last section we will give
a proof of Deodhar’s lemma and establish the existence of the minimal
and the maximal defining chain for a standard Young diagram in the
case G = SLn (k) (Chapter 2, 4.4).
Sections 2 and 5 follow closely an unpublished manuscript of Cheval-
ley [C]. for section 6 see [L-M-S1] or [L-S3].

4.1 Some remarks on linear algebraic groups

In this section we want to recall some facts on algebraic groups and fix
some notations. We refer to [Hu] or [Bo] for the facts and notations
stated here without further comment.

4.1.1
Recall that a linear algebraic group T is called a torus if T is a connected
group isomorphic to a product of groups GL1 (k). The character group
X(T ) := Hom (T, GL1 (k)) is a free abelian group of rank equal to dim T.
Let V be a finite dimensional rational T -module. For χ ∈ X(T ) denote
by Vχ = {V ∈ V | t.v = χ(t).v, ∀t ∈ T }. The finitely many χ ∈ X(T )

© Springer Science+Business Media Singapore 2016 and Hindustan Book Agency 2015 107
C.S Seshadri, Introduction to the Theory of Standard Monomials,
Texts and Readings in Mathematics 46, DOI 10.1007/978-981-10-1813-8_4
108 4. Schubert varieties in G/Q

such that Vχ 6= (0) are called the weights of T in V and Vχ is called


the weight space corresponding to χ. The T -module V decomposes as a
direct sum into the weight spaces: V = ⊕χ∈X(T ) Vχ .

4.1.2
Let G be a semisimple, connected algebraic group. Fix a maximal torus
T of G. Denote the Lie algebra of G by Lie G. The torus T acts via the
adjoint representation on Lie G. The nonzero weights of this represen-
tation are called roots. Denote the set of roots by φ := (G, T ). It is well
known that dim Lie (G)α = 1, ∀ α ∈ φ. Furthermore, the weight space
corresponding to the trivial weight is Lie T. Hence, we get Lie G = Lie
T ⊕ (⊕α∈φ (Lie G)α ) Denote by W = NorG T /CentG T , the Weyl group
of G. The group W is finite and operates faithfully on T and on X(T ).
There exists a W -invariant scalar product (., .) on X(T )R := X(T )⊗Z R.
We identify φ canonically as a subset of X(T )R . The set φ is a reduced
root system in X(T )R . Denote by sα the reflection in X(T )R with re-
spect to the hyperplane orthogonal to α ∈ φ. Denote by W (φ) the
subgroup of GL(X(T )R ) generated by the reflections {sα | α ∈ φ}. The
canonical homomorphism ρ : W → GL(X(T )R ) induces an isomorphism
ρ : W → W (φ). Therefore we will identify these two groups in the fol-
lowing.

4.1.3
A subset S of φ is called a basis of φ if S = {α1 , . . . , αr } is a basis of
P r
X(T )R and every α ∈ φ has an expression α = ci αi such that the
i=1
ci are integers of like sign. Such a basis exists and the elements of a
fixed basis are called simple roots. The corresponding reflections sαi are
called simple reflections. The roots, which are non negative (resp. non
positive) linear combinations of the elements in S constitute the set φ+
(resp. φ− ) of positive (resp. negative) roots. We have a decomposition
φ = φ+ ∪ φ− and there exists a uniquely determined ω0 ∈ W such that
ω0 (φ+ ) = φ− . Note that ω02 = id. We write α > 0 (resp. α < 0) if
α ∈ φ+ (resp. α ∈ φ− ).

4.1.4 Lemma
Let α ∈ S and β > 0. The following are equivalent (1) sα (β) ∈ φ− (ii)
α = β.
4.1. Some remarks on linear algebraic groups 109

P
r
Proof. Let S = {α1 , . . . , αr } such that α = α1 , and let β = ci αi be
i=1
a representation of β with respect to S.

Xr
2(α, β) 2(α, β)
sα (β) = β − α= ci αi + (c1 − )α.
(α, β) (α, β)
i=2

If α 6= β, then ∃ at least one i ∈ {2, . . . , r} such that ci > 0. Since


all coefficients of sα (β) are of like sign, we get sα (β) ∈ φ+ , which leads
to a contradiction. Q.E.D.

4.1.5
Denote by Ga the affine line A1k with group law µ(x, y) = x + y. For
every α ∈ φ there exists a uniquely determined subgroup Gα in G such
that Gα ≃ Ga and Lie Gα = (Lie G)α . Let B be a Borel subgroup of G
such that B ⊃ T. Denote the unipotent radical of B by B u . There exists
a uniquely
Q determined basis S uof φ such that with respect to− S we have
Bu = Gα . Note that B/B ≃ T. The Borel subgroup B := ω0 Bω0
α∈φ+
(see 4.1.3) is called the opposite Borel subgroup. The unipotent radical
of B − is (B − )u = πα∈φ+ G−α and B ∩ B − = T.

4.1.6
Let B be a Borel subgroup of G such that B ⊃ T. If ω ∈ W, then BωB
is a well defined double coset, because T = CenG T (= Centralizer of T
in G) ⊂ B.

4.1.7 Theorem
Bruhat decomposition ([Hu]), Theorem 28). G = ∪ω∈W BωB, with
BωB = Bτ B if and only if ω = τ in W. Since this is a finite union,
there has to be a double coset in G such that BωB is dense in G. De-
note by (B − )u the unipotent radical of the opposite Borel subgroup
e · B) = ω0 ((B − )u · B), the
B − = ω0 Bω0 (see 4.1.5). Since Bω0 B = ω0 (B
following proposition shows that Bω0 B is an open dense subset of G.

Proposition ([Hu 1], Prop. 28.5). The product map (B − )u × B → G


is an isomorphism onto an open subset Ω of G (called the big cell).
110 4. Schubert varieties in G/Q

4.1.8
2(λ, α)
For λ ∈ X(T )R denote by hλ, αi. Let Λ(φ) be the set of all
(α, α)
elements λ ∈ X(T )R such that hλ, αi ∈ Z, again ∀ α ∈ φ. The elements
of Λ(φ) which form a lattice in X(T )R , called the weight lattice Λ(φ)
which is therefore W -stable. If S = {α1 , . . . , αr }, then denote by ̟i the
element in Λ(φ) such that h̟i , αi i = δij . The elements ̟1 , . . . , ̟r are
called fundamental weights. They span Λ(φ) over Z. Denote by Λr (φ)
(called the root lattice) the Z-span of φ in X(T )R . One can show that
Λr (φ) ⊆ X(T ) ⊆ Λ(φ). Call a pair (φ, Γ) a diagram. Here φ is a reduced
root system in some finite dimensional vector space over R and Γ is a
lattice between Λr (φ) and Λ(φ).

4.1.9 Theorem
([Hu 1], §§32, 33). The map G → (φ(G, T ), X(T )) induces a bijection
between isomorphism classes of semisimple algebraic groups over k and
isomorphism classes of diagrams.
If X(T ) = Λ(φ), then G is called simply connected (if k = C, then
G is a simply connected Lie group). The non negative integral linear
combinations of the fundamental weights are called dominant weights.
Let V be a finite dimensional vector space and σ : G → GL(V ) a rational
irreducible representation. If B is a Borel subgroup of G corresponding
to the basis S of φ, then ∃ a unique line D ⊆ V which is stable under B
and the weight λ(σ) of this line is dominant.

Theorem ([Hu 1], §31). The map σ 7→ λ(σ) induces a bijection of iso-
morphism classes of irreducible rational G-representations and dominant
weights belonging to X(T ).

4.2 Basic properties

In this section we are going to prove some basic properties of Schubert


varieties and we will develop the main tools we are going to use in
the later sections. Unless otherwise stated, throughout the following
sections, we will denote by G a semisimple connected algebraic group.
We fix a maximal torus T of G and a Borel subgroup B of G containing
T. Let Q be a parabolic subgroup of G containing B.
4.2. Basic properties 111

4.2.1
Let S be a basis of the root system φ of G such that B = T.πα>0 Gα .
Let W denote the Weyl group of G with respect to T. There exists a
uniquely determined subset SQ ⊂ S such that Q is generated by B and
P
t
πα∈R+ (Q) Gα , where R+ (Q) = {β ∈ φ+ | β = ni , αi , ni ≥ 0, αi ∈ SQ }.
i=1
Denote by WQ the Weyl group of Q with respect to T. We can consider
WQ as a subgroup of W. The simple reflections sα , α ∈ SQ , generate
WQ . If R(Q) = R+ (Q) ∪ (−R+ (Q)), then R(Q) is the root system of
the reductive part MQ of the Levi decomposition Q = MQ .UQ of Q.
The unipotent radical UQ of Q is generated by the subgroups {Gα | α ∈
φ+ \R+ (Q)}.

4.2.2
Let θ ∈ W and α > 0. If ψ ∈ WQ , then ψ(R(Q)) = R(Q) and hence
θ −1 (α) ∈ R(Q) ⇔ (θ.ψ)−1 (α) = ψ −1 (θ −1 (α)) ∈ R(Q). Now let β ∈ SQ
and γ > 0. We know by Lemma 4.1.4 that sβ (γ) < 0 implies that
β = γ. Hence, θ −1 (α) < 0 (resp. θ −1 (α) > (0) and θ −1 (α) 6∈ R(Q) ⇔
(θsβ )−1 (α) = sβ θ −1 (α) < 0 (resp. θsβ )−1 (α) > 0 and (θs−1 β )(α) 6∈
R(Q)). Now WQ is generated by the simple reflection {sβ | β ∈ SQ }.
Hence if θ, θ ′ ∈ W such that θ ≡ θ ′ mod WQ , then θ −1 (α) ∈ R(Q) ⇔
(θ 1 )−1 (α) ∈ R(Q); also θ −1 (α) 6∈ R(Q) and θ −1 (α) < 0 (resp. > 0) ⇒
(θ 1 )−1 (α) 6∈ R(Q) and (θ 1 )−1 (α) < 0 (resp. > 0).

Definition Let ω ∈ W/WQ and α > 0. We say that ω −1 (α) ∈ R(Q) if


θ −1 (α) ∈ R(Q) for some (hence every) representative θ ∈ W of ω. Note
that this condition is empty if Q = B. If ω −1 (α) 6∈ R(Q) we say that
ω −1 (α) < 0 (resp. ω −1 (α) > 0) if θ −1 (α) < 0 (resp. θ −1 (α) > 0) for
some (hence every) representative θ ∈ W of ω.

4.2.3
Let e(id) denote the coset of Q in G/Q and for ω ∈ W/WQ denote by
e(ω) the point ω.e(id). Recall that {e(ω) | ω ∈ W/WQ } is the set of
all T -fixed points in G/Q; for, T.g.Q = g.Q implies that g−1 T g ⊆ Q
and since ∃q ∈ Q such that (g.q)−1 T gq = T, we get g.q ∈ NorG T. Thus
g.Q = e(ω) for some ω ∈ W/WQ . Denote the isotropy group of G at e(ω)
by IG (ω).
112 4. Schubert varieties in G/Q

Lemma

(i) Let α ∈ φ and ω ∈ W/WQ . If Q = B, then Gα ⊆ IG (ω) ⇒


ω −1 (α) > 0. If Q 6= B, then Gα ⊆ IG (ω) ⇒ ω −1 (α) ∈ R(Q) or
ω −1 (α) 6∈ R(Q) and ω −1 (α) > 0.

(ii) If α ∈ φ and ω ∈ W/WQ such that ω −1 (α) 6∈ R(Q) then Gα ⊆


IG (ω) ⇒ G−α 6⊆ IG (ω).

Proof. This follows easily from the fact that IG (ω) = ωQω −1 . Q.E.D.

4.2.4
Let ω ∈ W/WQQ and let N (ω) = {α > 0 | ω −1 (α) 6∈ R(Q) and ω −1 (α) <
0}. Let Bωu = { α Gα | α ∈ N (ω)}. The set Bωu contains all subgroups
Gα of B such that Gα is not contained in IG (ω). Let C(ω) denote the
B-orbit B.e(ω) in G/Q and denote its closure B.e(ω) by X(ω). We en-
dow X(ω) with canonical structure of a reduced subscheme of G/Q.
Then C(ω) is called the Schubert cell corresponding to ω and X(ω) is
called the Schubert variety corresponding to ω. Let θ0 ∈ W denote the
unique element such that θ0 (φ+ ) = φ−1 and denote its canonical image
in W/WQ by ω0 .

Lemma

(i) If ω ∈ W/WQ , then the map ψ : Bωu → C(ω), (b 7→ b.e(ω)) is an


isomorphism. In particular dim Bωu = dim X(ω) = #N (ω) and
C(ω) is an affine variety.

(ii) X(ω0 ) = G/Q

(iii) dim X(sα ) = 1, ∀ α ∈ S − SQ .

Proof.

(i) By the definition of Bωu we have B.e(ω) = Bωu .e(ω). Now the
claim follows from the fact that the isotropy group scheme in B is
ωBω −1 ∩ B and Bωu is just the complement to the isotropy group
scheme.
4.2. Basic properties 113

(ii) By Proposition (4.1.7), we know that if B = Q, then B.e.(θ0 ) =


φ0 (B −1 .e(id)) is open in G/Q and hence, X(θ0 ) = G/B. If Q is
any parabolic subgroup of G containing B, then the projection map
η : G/B → G/Q is B-equivarient and hence, X(ω0 ) = η(X(θ0 )) =
η(G/B) = G/Q.

(iii) By Lemma 4.1.4, we know that if β ∈ φ+ , then sα (β) ∈ φ−1 ⇒


α = β. Since α 6∈ R(Q) we get by (i), dim X(sα ) = 1. Q.E.D.

Corollary The Bruhat decomposition G/Q = ∪ω∈W/WQ B.e(ω) is a “cel-


lular decomposition” in the usual topological sense (say when the base
field is C and we endow G/Q with the canonical “Euclidean topology”).

4.2.5

For every α ∈ φ+ there exists a uniquely determined subgroup L(α) in G


which is isomorphic to SL2 (k) or P SL2 (k) and contains the groups G−α
and Gα . Denote by B(±α), the Borel subgroup of L(α) containing G±α
respectively. Let T (α) be the maximal torus in L(α) contained in T. The
Weyl group of L(α) with respect to T (α) can be canonically embedded
into W. The image is the subgroup {e, sα } of W.

Lemma Let α ∈ φ and ω ∈ W/WQ such that ω −1 (α) 6∈ R(Q). The


orbit L(α).e(ω) is isomorphic to P−1 and the only T (α)-fixed points are
e(ω) and e(sα ω). Further, if ω −1 (α) < 0, then L(α).e(ω) = Gα .e(ω) ∪
e(sα ω) = Gα .e(ω), and if ω −1 (α) > 0, then L(α).e(ω) = G−α .e(ω) ∪
e(sα ω) = G−α .e(ω).

Proof. By Lemma (4.2.3) (ii) we know that either Gα or G−α is con-


tained in IG (ω). Hence, L(α).e(ω) = L(α)/B(α) ≃ P−1 or L(α).e(ω) ≃
L(α)/B(−α) ≃ P1 . Now T (α) has precisely two fixed points on P1 and
hence on L(α).e(ω). The orbit contains the two T -fixed points e(ω) and
e(sα ω). This proves the first part of the statement. Let ω −1 (α) < 0. We
have already seen that in this case L(α).e(ω) is isomorphic to the flag
variety L(α)/B(α). The decomposition L(α).e(ω) = Gα .e(ω) ∪ e(sα ω)
is just the corresponding cellular decomposition. The statement for the
case ω −1 (α) > 0 follows in the same way. Q.E.D.
114 4. Schubert varieties in G/Q

4.2.6 Lemma
(i) If α > 0 and ω ∈ W/WQ such that ω −1 (α) 6∈ R(Q) (if Q = B,
then R(Q) = ∅;), then:
ω −1 (α) < 0 ⇔ X(sα ω) ⊆ X(ω)
ω −1 (α) > 0 ⇔ X(ω) ⊆ X(sα ω).

(ii) If α > 0, then X(sα ω) = X(ω) ⇔ ω −1 (α) ∈ R(Q).

Proof.
(i) By Lemma (4.2.3)(ii) we know that the fact ω −1 (α) < 0 is equiv-
alent to saying that G−α fixes e(ω). It follows that L(α).e(ω) =
Gα .e(ω). Since Gα ⊆ B we know that Gα .e(ω) ⊆ X(ω) and hence
e(sα ω) ∈ X(ω). But this implies that X(xα ω) = B.e(sα ω) ⊆
X(ω). If ω −1 (α) > 0, then (sα ω)−1 (α) = ω −1 (−α) < 0 and hence
X(ω) = X(xα (xα ω)) ⊆ X(sα ω).
(ii) Let θ ∈ W be a representative of ω. If ω −1 (α) ∈ R(Q), then
θ −1 (α) = β ∈ R(Q). Now sβ = θ −1 sα θ ∈ WQ and hence θ ≡
sα θ mod WQ . It follows that sα ω ≡ ω mod WQ and X(sα ω) =
X(ω). On the other hand, if X(sα ω) = X(ω), then e(sα ω) = e(ω).
Hence, α ∈ N (ω) ⇔ α ∈ N (sα ω). But θ −1 (α) = −(sα ω)−1 (α).
This implies immediately that θ −1 (α) ∈ R(Q). Q.E.D.

4.2.7 Corollary
Let α ∈ S and ω ∈ W/WQ such that ω −1 (α) 6∈ R(Q). (Note: R(Q) = ∅
if Q = B.)
(i) ω −1 (α) > 0 ⇔ dim X(sα ω) = dim X(ω) + 1.
(ii) ω −1 (α) < 0 ⇔ dim X(sα ω) = dim X(ω) − 1.

Proof. By Lemma (4.2.4)(i) we know that dim X(ω) = #N (ω) (=


#{β > 0 | ω −1 (β) 6∈ R(Q) and ω −1 (β) < 0}), and dim X(sα ω) =
#N (sα ω). By Lemma (4.1.4) we know that sα permutes all positive roots
except α. If β ∈ N (ω) and β 6= α, then sα (β) ∈ N (sα , ω), and if γ ∈
N (sα , ω), γ 6= α, then sα (γ) ∈ N (ω). Hence, #N (ω) and #N (sα ω) differ
at most by 1. By Lemma (4.2.6) we know that #N (ω) = #N (sα ω) = 1
and that X(ω) ⊆ X(sα ω) in case (i), and X(sα , ω) ⊆ X(ω) in case (ii).
Q.E.D.
4.2. Basic properties 115

4.2.8
For α ∈ S denote by P (α) the minimal parabolic subgroup of G generated
by B and G−α .

Lemma If α ∈ S and ω ∈ W/WQ such that ω −1 (α) 6∈ R(Q) and


ω −1 (α) < 0, or ω −1 (α) ∈ R(Q), then P (α) · X(ω) ⊆ X(ω).

Proof. It suffices to show that for g ∈ G−α and b ∈ B u we have


(g.b)e(ω) ∈ X(ω). We know that B u = (UP (α) ).Gα (see (4.2.1)). Ev-
ery b ∈ B u can be decomposed into b = c1 .c2 , c1 ∈ UP (α) , c2 ∈ Gα .
We have to show that (g.c1 .c2 ).e(ω) = (gc1 g−1 )(gc2 )e(ω) ∈ X(ω). But
gc1 g−1 ∈ UP (α) ⊆ B, so it is enough to show that G−α .Gα .e(ω) ⊆ X(ω).
If ω −1 (α) ∈ R(Q), nothing is to be shown. Assume that ω −1 (α) 6∈ R(Q)
and ω −1 (α) < 0. Denote by L(α) the subgroup of G isomorphic to
SL2 (k) or P SL2 (k) which contains Gα and G−α . (L(α) is just the
semisimple part of P (α) ). Then L(α).e(ω) = Gα .e(ω) ∪ e(sα ω) (Lemma
(4.2.5)). By Lemma (4.2.6), we know that e(sα ω) ∈ X(ω) and hence
L(α).e(ω) ⊆ X(ω)). Q.E.D.

Remark Let Q = B. The proof actually shows that if ω ∈ W and


α ∈ S such that ω −1 (α) < 0, then P (α).e(ω) ⊆ B.e(ω) ∪ B.e(sα ω).
If ω −1 (α) > 0, then the same proof (using Lemma (4.2.5)) shows that
P (α).e(ω) ⊆ B.e(ω) ∪ Be(sα ω). Since sα ∈ WP (α) , we get sα .B.e(ω) ⊆
B.e(ω) ∪ B.e(sα ω)∀ ω ∈ W, α ∈ S; or, writing it as an inclusion of cosets
we get the following:

T (1) : sα Bω ⊆ BωB ∪ Bsα BωB.

Recall that the quadruple (G, B, NorG T, {sα | α ∈ S}) forms a Tits
system and T (1) is one of the axioms of Tits system (see e.g. [Hu], §29).
On the other hand, we could have derived Lemma (4.2.8) from T (1).

4.2.9
Let α ∈ S and denote by P (α) the corresponding minimal parabolic
subgroup of G and let L(α) be its semisimple part (see (4.2.5) and
(4.2.8)). Denote by B(α) the Borel subgroup of L(α) containing Gα .
Let ω ∈ W/WQ . We have a canonical action of B(α) on L(α) × X(ω)
given by:
116 4. Schubert varieties in G/Q

b.(g.x) := (gb−1 , b.x), ∀ b ∈ B.

Definition Denote by Z(ω, α) the bundle L(α) ×B(α) X(ω) with fibre
X(ω) associated to the principal fibration L(α) → L(α)/B(α) ≃ P−1 .

Remark We have a canonical action of B on P (α) × X(ω) given by


b.(g.x) := (gb−1 , b.x). The bundle P (α) ×B X(ω) with fibre X(ω) asso-
ciated to the principal fibration P (α) → P (α)/B ≃ P−1 is canonically
isomorphic to Z(ω, α).

4.2.10
Let α ∈ S and τ ∈ W/WQ such that τ −1 (α) 6∈ R(Q) and τ −1 (α) < 0.
Denote sα τ by ω. By Lemma (4.2.8), we have a canonical morphism
ρ : P (α) × X(ω) → X(τ ).

Lemma The canonical map ρ : P (α) × X(ω) → X(τ ) is surjective.

Proof. The map ρ is B-equivariant with respect to the natural left B-


action on P (α). Since sα e(ω) = e(τ ), we have e(τ ) ∈ Im ρ and hence the
image of ρ is dense. Now ρ factors through P (α) ×B X(ω) ≃ Z(ω, α) :

ρ
P (α) × T (ω) X(τ )

Z(ω, α)

The fibre bundle Z(ω, α) is a complete variety. Hence, Im ρ = Im ψ


is closed and dense in X(τ ) and the required surjectivity follows. Q.E.D.

4.2.11
Since the canonical map φ : L(α) × X(ω) → X(τ ) factors through
Z(ω, α), we get as an immediate consequence:
4.2. Basic properties 117

Corollary The canonical maps φ : L(α) × X(ω) → X(τ ) and ψ :


Z(ω, α) → X(τ ) are surjective.

Remark Note that dim Z(ω, α) = dim X(ω) + 1 = dim X(τ ).

4.2.12 Lemma
Let α ∈ S and τ ∈ W/WQ such that τ −1 (α) < 0 and τ −1 (α) 6∈ R(Q).
Denote sα τ by ω. If e(θ) ∈ X(τ ) such that e(θ) 6∈ X(ω), then L(α).e(θ)
meets X(ω) at the unique point e(sα θ).

Proof. The map φ : L(α) × X(ω) → X(θ) is surjective by Corollary


(4.2.11). Hence, ∃ x ∈ X(ω) such that e(θ) ∈ L(α).x and therefore
L(α).e(θ) ∩ X(ω) 6= ∅. The orbit L(α).e(θ) is isomorphic to P1 (Lemma
(4.2.5)) and is not contained in X(ω). So the intersection has to be
discrete. Since the intersection is T (α)-stable and T (α) has only two
fixed points on L(α).e(θ), we conclude that L(α).e(θ) ∩ X(ω) = e(sα θ).
Q.E.D.

4.2.13 Lemma
Let ω1 , ω2 ∈ W/WQ and α ∈ S such that ω2−1 (α) 6∈ R(Q) and ω2−1 (α) <
0. Assume that X(ω1 ) ⊆ X(ω2 ). If X(ω1 ) 6⊆ X(sα ω2 ), then X(sα ω1 ) ⊆
X(ω1 ) and X(sα ω1 ) ⊆ X(sα ω2 ), i.e. in the second case we get the
following diagram of inclusion

X(ω2 )

X(sα ω2 )

X(ω1 )

X(sα ω1 )

Proof. Assume X(ω1 ) 6⊆ X(sα ω2 ) (Otherwise there is nothing to be


shown). Hence, we know that e(ω1 ) 6∈ X(sα ω2 ). By Lemma (4.2.12) we
118 4. Schubert varieties in G/Q

know that e(sα ω1 ) ∈ X(sα ω2 ) and hence, X(sα ω1 ) ⊆ X(sα ω2 ). Since


X(ω1 ) 6⊆ X(sα ω2 ), we conclude that X(sα ω1 ) ⊆ X(ω1 ) (by Lemma
(4.2.6)). Q.E.D.

4.3 Reduced decompositions

Let (α1 , . . . , αk ) be any subset of S. In the following we denote the cor-


responding simple reflections sα1 by si , i = 1, . . . , k. Recall that for
ω ∈ W/WQ , the length ℓQ (ω) is defined to be the minimal r such that ω ≡
s1 . . . . .sr mod WQ , where α1 , . . . , αr ∈ S. Note that ℓQ (ω) is constant on
the coset of ω in W/WQ . A decomposition ω ≡ s1 . . . . .sr mod WQ such
that r = ℓQ (ω) is called reduced. Note that for every i = 1, . . . , r − 1.
(si , .si−1 . . . s1 ω) ≡ si+1 . . . . .sr mod WQ is a reduced decomposition for
si . . . s1 ω ∈ W/WQ . In this section we are going to derive certain connec-
tions between the combinatorial properties of reduced decompositions
and geometric properties of the corresponding Schubert varieties. In
particular, we are going to prove the equivalence of the various defini-
tions of the partial orders on W/WQ stated in Chapter 3.

4.3.1 Lemma
Let ω ∈ W/WQ and let ω ≡ s1 . . . . .sr mod WQ be a reduced decom-
position. Then ω −1 (α1 ), (s1 ω)−1 (α2 ), . . . (sr−1 . . . . .s1 ω)−1 (αr ) are not
elements of R(Q).

Proof. Recall that sj . . . . .sr ≡ sj−1 . . . . .s1 ω mod WQ and hence, β =


(sj . . . . .sr )−1 (αj ) ∈ R(Q) if and only if (sj−1 . . . . .s1 ω)−1 (αj ) ∈ R(Q).
If β ∈ R(Q), then sβ = sr . . . . .sj+1 .sj .sj+1 . . . . .sr ∈ WQ and hence,
sj . . . . .sr ≡ sj+1 . . . . .sr mod WQ . But then ω ≡ s1 . . . . .sj−1 .sj+1 . . . . .sr
mod WQ , which is a contradiction to our assumption that ℓQ (ω) = r.

4.3.2 Lemma
Let ω ∈ W/WQ and let ω ≡ s1 . . . . .sr mod WQ be a reduced decompo-
sition. Let θ1 = α1 , θ2 = s1 (α2 ), . . . , θr = sr−1 . . . . .s1 (αr ). If α > 0
such that ω −1 (α) 6∈ R(Q) and ω −1 (α) < 0, then α = θi for some
i ∈ {1, . . . , r}.

Proof. By section (4.2.2) we see that our hypothesis implies (s1 . . . . .sr )−1 (α) <
0 and (s1 . . . . .sr )−1 (α) 6∈ R(Q). Hence, there exists an i ∈ {1, . . . , sr }
4.3. Reduced decompositions 119

such that s1 (α) > 0, . . . , si−1 . . . . .s1 (α) > 0, but si . . . . .s1 (α) < 0.
Now Lemma (4.1.4) implies that αi = si−1 . . . . .s1 (α) and hence, α =
s1 . . . . .si−1 (αi ).

4.3.3 Lemma
Let ω ∈ W/WQ . Then dim X(ω) = ℓQ (ω).

Proof. We use induction on ℓQ (ω). If ℓQ (ω) = 0 or 1, nothing is to be


shown (see Lemma (4.2.1)). Assume now that ℓQ (ω) = r > 1. Let ω ≡
s1 . . . . .sr mod WQ be a reduced decomposition. Then ℓQ (s1 ω) = ℓQ (ω)−
1 and s1 ωs2 . . . . .sr mod WQ is a reduced decomposition. By induction
we know that ℓQ (s1 ω) = dim X(s1 ω). We want to show ω −1 (α1 ) < 0.
Assume ω −1 (α1 ) > 0. Then (s1 ω)−1 (α1 ) < 0 and by Lemma 4.3.2 we
know that α1 = s2 . . . . .si−1 (αi ) for some i = 2, . . . , r. Hence, s1 =
s2 . . . . .si−1 .si .si−1 . . . . .s2 and ω ≡ (s2 . . . . .si−1 ).(si .si−1 . . . . .s2 ).s2 . . . . .sr
mod WQ ; consequently, ω ≡ s2 . . . . .si−1 .si+1 . . . . .sr mod WQ . But this is
a contradiction to our assumption that ℓQ (ω) = r. Hence, we know that
ω −1 (α1 ) < 0 and ω −1 (α1 ) 6∈ R(Q) (by Lemma (4.3.1)). Now Corollary
(4.2.7) implies that dim X(ω) = dim X(s1 ω) + 1 = ℓQ (s1 ω) + 1 = ℓQ (ω).
Q.E.D.

4.3.4
We use the same notations as in Lemma (4.3.2).

Corollary Let ω ∈ W/WQ and ω = s1 . . . . .sr mod WQ be a reduced de-


composition. Then N (ω) = {θ1 , . . . , θr }. In particular, for i = 1, . . . , r.θi
are distinct and positive.

Proof. By Lemma (4.2.4)(i) and Lemma (4.3.3) we know that r =


ℓQ (ω) = dim X(ω) = #N (ω). By lemma (4.3.2) we know that ∀ α ∈
N (ω), α = θi for some i = 1, . . . , r. Q.E.D.

4.3.5 Corollary
Let ω ∈ W/WQ and let ω ≡ s1 . . . . .sr mod WQ be a decomposition. The
following are equivalent:

(i) The decomposition is reduced.


120 4. Schubert varieties in G/Q

(ii) ω −1 (α1 ) 6∈ R(Q), ω −1 (α1 ) < 0; (s1 ω)−1 (α2 ) 6∈ R(Q), (s1 ω)−1 (α2 ) <
0; . . . ; (sr−1 . . . . .s1 ω)(αr ) 6∈ R(Q), (sr−1 . . . . .s1 ω)−1 (αr ) < 0.

Proof. (i) ⇒ (ii) Let ζ ∈ W be a representation of ω. Then (s1 . . . . si ζ)−1


(αi+1 ) = ζ −1 (s1 . . . . .si (αi+1 )) = ζ −1 (θi ) and hence, (si . . . . .s1 )−1 (αi+1 ) 6∈
R(Q) and (si . . . . .s1 ω)−1 (αi+1 ) < 0 if and only if ω −1 (θi ) 6∈ R(Q) and
ω −1 (θi ) < 0. But this is true by Corollary (4.3.4).

(ii) ⇒ (i): Since (si . . . . .s1 ) ≡ si+1 . . . . .sr mod WQ , we know that
(si+1 . . . . , sr )−1 (αi+1 ) < 0. Hence, we get a chain of Schubert varieties
X(sr ) ⊆ X(sr−1 sr ) ⊆ . . . ⊆ X(s1 . . . . .sr ) (Lemma (4.2.6) such that
X(si−1 . . . . .sr ) is of codim 1 in X(si . . . . .sr ). Hence r = ℓQ (ω). Q.E.D.

4.3.6 Corollary
Let ω, τ ∈ W/WQ such that X(τ ) ⊆ X(ω). There exists ω1 , . . . , ωℓ ∈
W/WQ such that ω1 ≡ ω, . . . , ωℓ ≡ τ mod WQ and X(ωi ) ⊆ X(ωi−1 ) is
of codim. 1.

Proof. It is enough to show that ∃σ ∈ W such that X(τ ) ⊆ X(σ) ⊆


X(ω) and X(σ) is of codim. 1 in X(ω). We prove this by induction
on dim X(ω). In the case dim X(ω) = 0, 1, nothing is to be shown.
Let dim X(ω) = r > 1 and let ω ≡ s1 . . . . .sr mod WQ be a reduced
decomposition. Since ω −1 (α1 ) 6∈ R(Q) and ω −1 (α1 ) < 0, we know that
X(s1 ω) ⊆ X(ω) (Lemma (4.2.5)) and is of codim. 1 in X(ω). If X(τ ) ⊆
X(s1 ω), we are done. Assume that X(τ ) 6⊆ X(s1 ω). By Lemma (4.2.13),
we get the following diagram of inclusion:

X(ω)

X(s1 ω)

X(τ )

X(s1 τ )
4.3. Reduced decompositions 121

By induction we know that ∃ σ ∈ W/WQ such that X(s1 .τ ) ⊆


X(σ) ⊆ X(s1 ω) and X(σ) is of codim 1 in X(s1 ω). We want to show
that X(s1 σ) is of codim. 1 in X(ω). Assume that this is not the case.
Then X(s1 σ) ⊆ X(σ) and hence, X(σ) is stable under the minimal
parabolic subgroup P (α1 ) (Lemma (4.2.6) and Lemma (4.2.8)). But
then P (α1 ).X(s1 .τ ) = X(τ ) ⊆ X(σ1 ) ⊆ X(s1 ω), which contradicts as-
sumption that X(τ ) 6⊆ X(s1 ω). Hence, X(τ ) ⊆ X(s1 σ) and X(s1 σ) is
of codim 1 in X(ω). Q.E.D.

4.3.7 Lemma
If ω ∈ W/WQ and ω ≡ s1 . . . . .sr mod WQ is a decomposition, then there
exists a sequence 1 ≤ i1 < · · · < iℓ ≤ r such that ω ≡ si1 . . . . .siℓ mod WQ
is a reduced decomposition.

Proof. By Corollary 4.3.5 we know that if ω −1 (α1 ) 6∈ R(Q), . . . , (sr−1 . . . . .


s1 ω)−1 (αr ) 6∈ R(Q) and all of them are < 0, then the decomposition is
reduced. Now assume that there exists an i ∈ {1, . . . , r} such that
(si−1 . . . . .s1 ω)−1 (αi )) ∈ R(Q). Since s1 . . . . .sr ≡ si−1 mod WQ this im-
plies that θ = (si . . . . .sr )−1 (αi ) ∈ R(Q). But then sθ = sr . . . . .si+1 .si .si+1
. . . . .sr ∈ WQ and si . . . . .sr ≡ si+1 . . . . .sr mod WQ . It follows that ω ≡
s1 . . . . .si−1 .si+1 . . . . .sr mod WQ i.e., we can drop si in the decomposition.
Now assume that ω −1 (α1 ) < 0, (s1 ω)(α2 ) < 0, . . . , (si−2 . . . . .s1 ω)−1 (αi−1 ) <
0 but (si−1 . . . . .s1 ω)−1 (αi ) > 0 and none of them is an element of R(Q).
Since si−1 . . . . .s1 ω ≡ si . . . . .sr mod WQ , this implies that (si . . . . .sr )−1 (αi ) >
0 and hence, (si+1 . . . . .sr )−1 (αi ) < 0. Since αi > 0, there exists a
j ∈ {i + 1, . . . r} such that si+1 (αi ) > 0, . . . , sj−1 . . . . .si+1 (αi ) > 0 but
sj . . . . .sr (αi ) < 0. By Lemma (4.1.4) we know that αj = sj−1 . . . . .(αi )
and hence, sj = sj−1 . . . . .si+1 .si .si+1 . . . . .sj−1 . If we replace sj in the
decomposition of ω by this expression, then we get:

ω ≡ s1 . . . . .si−1 .si+1 . . . . .sj−1 .sj+1 . . . . .sr mod WQ .


By repeating the procedure if necessary, we get a reduced decompo-
sition ω = sij . . . . .siℓ mod WQ such that 1 ≤ i1 < · · · < iℓ ≤ r. Q.E.D.

4.3.8
Recall that we have defined a partial order on W/WQ : If ω1 , ω2 ∈
W/WQ , then we say ω1 ≤ ω2 if and only if X(ω1 ) ⊆ X(ω2 ). We defined
ω1 to be a subword of ω2 if there exists a reduced decomposition ω2 ≡
122 4. Schubert varieties in G/Q

s1 . . . . .sr mod WQ and a sequence q ≤ i1 < · · · < iℓ ≤ r such that


ω1 ≡ si1 . . . . .siℓ mod WQ is a reduced decomposition.

Lemma Let ω1 , ω2 ∈ W/WQ . The following are equivalent:

(i) ω1 ≤ ω2

(ii) ω1 is a subword of ω2 .

Proof. The proof is by induction on ℓQ (ω2 ). In the case ℓQ (ω2 ) =


0 or 1, nothing is to be shown. Let ℓQ (ω2 ) = r > 1 and let ω2 ≡
s1 . . . . .sr mod WQ be a reduced decomposition. Denote s1 .ω by τ. Then
X(τ ) ⊆ X(ω2 ) is of codim 1 in X(ω) and τ is a subword of ω.

(i) ⇒ (ii). First assume ω1 ≤ τ. By induction we know that ω1 is a


subword of τ (since X(ω1 ) ⊆ X(τ ) ⊆ X(ω2 )) and hence, ω1 is a subword
of ω2 . Now assume that X(ω1 ) 6⊆ X(τ ). By Lemma (4.2.13), we have
the following diagram of inclusion:

X(ω2 )

X(ω1 )

X(τ )

X(s1 ω1 )

By induction we know that s1 ω1 is a subword of τ. Hence, ∃i1 . . . . , iq ,


2 ≤ i1 < · · · < iq ≤ r such that s1 ω1 ≡ si1 . . . . .siq mod WQ is a reduced
decomposition. Thus ω1 ≡ s1 .si1 . . . . .siq mod WQ is a reduced decompo-
sition of ω1 and hence ω1 is a subword of ω2 .

(ii) ⇒ (i): First assume that ω1 is a subword of τ. By induction we


know that X(ω1 ) ⊆ X(τ ) ⊆ X(ω2 ) and hence, ω1 ≤ ω2 . Now assume
that ω1 is not a subword of τ. Then there exists a reduced decomposition
ω1 ≡ si .si2 . . . . .siℓ mod WQ such that 2 ≤ i1 < · · · < iℓ ≤ r. Hence s1 ω1 is
4.3. Reduced decompositions 123

a subword of τ and by induction we get X(s1 ω1 ) ⊆ X(τ ) ⊆ X(ω2 ). Since


ω2−1 (α2 ) ∈ R(Q) and ω −1 (α1 ) < 0, we know that P (α1 ).X(ω2 ) ⊆ X(ω2 )
(Lemma (4.2.8)). By Lemma (4.2.6), we know that e(ω1 ) ∈ L(α).e(s1 ω1 )
and hence, e(ω1 ) ∈ X(ω2 ). But this implies that X(ω1 ) ⊆ X(ω2 ) and
hence, ω1 ≤ ω2 . Q.E.D.

Remark

(i) The proof shows that if ω1 is a subword of ω2 , then for every


reduced decomposition ω2 = s1 . . . . .sr mod WQ , ∃ i1 , . . . , iℓ , 1 ≤
i1 < · · · < iℓ ≤ r such that ω1 ≡ si1 . . . . .siℓ mod WQ is a reduced
decomposition.

(ii) Let Q = B be the Borel subgroup of G. Let ω ∈ W and α ∈ S


such that ω −1 (α) < 0. Then X(sα ω) ⊆ X(ω) and is of codim. 1.
Let ω = s1 . . . . .sr be a reduced decomposition of ω and let sα ω =
si1 . . . . .sir−1 be a reduced decomposition of sα ω given by a se-
quence 1 ≤ i1 < · · · < ir−1 ≤ r. Then ω = sα .si1 . . . . .sir−1 is
a reduced decomposition of ω. Hence , there exists an integer j
such that 1 ≤ j ≤ r and sα .s1 . . . . .sj−1 = s1 . . . . sj−1 .sj . Thus the
pair (W, {sα | α ∈ S}) satisfies the “exchange condition.” This is
equivalent to saying that the pair (W, {sα | α ∈ S}) is a Coxeter
system. (See [Bou].)

4.3.9 Corollary
Let ω1 , ω2 ∈ W/WQ be such that X(ω1 ) is of codim. 1 in X(ω2 ). Then
ω1 = sβ ω2 mod WQ for some positive root β.

Proof. Let ω2 ≡ s1 . . . . .sr mod WQ be a reduced decomposition and let


ω1 ≡ si1 . . . . .sir−1 mod WQ be a reduced decomposition of ω1 given by
a subword of ω2 . If si1 6= s1 , then ω1 = s1 ω2 mod WQ and we are done.
If si1 = s1 , then there exists a smallest integer j ∈ {2, . . . , r} such that
si1 6= sj . By Corollary (4.3.4) we know that θj = s1 . . . . .sj−1 (αj ) is a
positive root. Now sθj ω2 ≡ s1 . . . . .sj−1 .sj+1 . . . . .sr mod WQ and hence
sθj ω2 ≡ ω1 mod WQ .

4.3.10
Let Q1 , Q2 be two parabolic subgroups of G such that Q1 ⊇ Q2 ⊇
B. Denote by η : G/Q2 → G/Q1 the canonical projection. To avoid
124 4. Schubert varieties in G/Q

ambiguity we denote the Schubert variety X(ω) ֒→ G/Qi corresponding


to ω ∈ W/WQ by XQi (ω).

Corollary For every ω ∈ W/WQ1 there exists a minimal representative


τ min ∈ W/WQ2 and a maximal representative τ max ∈ W/WQ2 such that

(i) the map XQ2 (τ min ) → XQ1 (ω) is birational and if ω ≡ s1 . . . . .sr mod WQ1
is a reduced decomposition then τ̃ min = s1 . . . . .sr is a representa-
tive of τ min in W.

(ii) ℓQ2 (τ̃ min .σ) = ℓQ2 (τ min ) + ℓQ2 (σ), ∀ σ ∈ WQ1 .

(iii) ∀ζ ∈ W/WQ2 such that η(XQ2 (ζ)) = XQ1 (ω) we have τ min ≤ ζ ≤
τ max in W/WQ2 and ζ ≡ τ̃ min .σ mod WQ for some σ ∈ WQ1 .

(iv) XQ2 (τ max ) = η −1 (XQ1 (ω)).

Proof.

(i) Let ω ≡ s1 . . . sr mod WQ1 be reduced. Denote s1 . . . sr by τ min .


Then τ min ≡ s1 . . . . .sr mod WQ2 is a reduced decomposition and
by Corollary (4.3.4) and (4.2.4) we know that N (ω) = N (τ min ).
Therefore, Bωu = Bτumin (See  (4.2.4)).min
The u
 maps ψ : Bω → CQ1 (ω)
′ u
and ψ : Bω → CQ2 τ min b 7→ be(τ ) are isomorphisms. Hence,
the map η|XQ2 (τ min ) : XQ2 (τ min ) → XQ (ω) induces an isomor-
phism of the open cells CQ2 (τ min ) and CQ1 (ω). This proves (i).

(ii) Let σ ∈ WQ1 and let σ ≡ t1 . . . . .tℓ mod WQ2 be a reduced decompo-
sition. Then τ̃ min .σ ≡ s1 . . . . .sr .t1 . . . . .tℓ mod WQ2 is a decomposi-
tion. By Lemma 4.3.7, we know that we get a reduced decomposi-
tion of τ̃ min .σ by dropping certain elements in that decomposition.
But since the decomposition ω ≡ s1 . . . . .sr mod WQ1 is reduced
and t1 . . . . .tℓ ∈ WQ1 , we see easily that the decomposition of τ min σ
has to be reduced and hence ℓQ2 (τ̃ min .σ) = ℓQ2 (τ min ) + ℓQ2 (σ).

(iii) Denote by σ0 the unique element of WQ1 such that σ0 (R+ (Q1 )) =
R− (Q1 ).

Let MQ1 denote the semisimple part of Q1 and denote by BQ1 the
Borel subgroup of MQ1 corresponding to R+ (Q). The Schubert
variety in MQ1 /BQ1 corresponding to σ0 is MQ1 /BQ1 and hence
4.3. Reduced decompositions 125

σ0 ≥ ∀ σ ∈ WQ1 , and in particular σ0 ≥ σ in W/WQ2 ∀ σ ∈ WQ1 .


Denote τ̃ min .σ0 mod WQ2 by τ max . Let XQ2 (ζ) be any Schubert
variety in G/Q2 such that η(XQ1 (ζ)) = XQ1 (ω) = XQ1 (τ min ).
Then ζ ≡ τ min mod WQ1 and hence, ∃ σ ′ ∈ WQ1 such that ζ ≡
τ min .σ ′ mod WQ1 . Now σ ′ ≤ σ0 in W/WQ2 . Hence, by (ii) we get
that τ̃ min .σ ′ ≤ τ max in W/WQ2 .

(iv) The preimage η −1 (XQ1 (ω)) consists of all Schubert varieties XQ2 (ζ)
such that η(XQ1 (ζ)) = XQ1 (ω). They are all contained in XQ2 (τ max )
and by construction η(XQ2 (τ max )) = XQ1 (ω).

4.3.11
Let S = {α1 , . . . , αr } be a basis of the root system Φ of G. Denote by Pi
the maximal parabolic subgroup of G such that SPi = S − {αi }. Denote
by Q the parabolic subgroup Pi ∩ Pi+1 ∩ . . . ∩ Pn of G. For τ ∈ W/WQ
denote by τj , j ≥ i, the canonical image of τ in W/WPj .

Lemma If τ, ω ∈ W/WQ , then ω ≤ τ ⇔ ωj ≤ τj ∀ j = 1, . . . , n.

Proof. We have only to prove the implication “⇐ .” We use induction


on ℓQ (τ ). If ℓQ (τ ) = 0 or 1, nothing is to be shown. Assume ℓQ (τ ) > 1.
Choose a simple reflection sα such that sα τ < τ. If ωj ≤ (sα τ )j , ∀ j =
1, . . . , n, then ω ≤ sα τ by induction and hence, ω < τ. Suppose ωj ≤
(sα τ )j for some j. This implies that (sα ω)j ≤ (sα τ )j and (sα τ ) < ωj for
all these j (Lemma (4.2.13)). Hence, sα ω < ω and (sα ω)j ≤ ωj ∀ j =
1, . . . , n. Since ωj ≤ (sα τ )j for the other j’s, it follows that (sα ω)j ≤
(sα τ )j ∀ j = i, . . . , n, and by induction sα ω ≤ sα τ. Since sα τ < τ, we
know that X(τ ) is P (α)-stable (Lemma (4.2.8)) and hence, e(ω) ∈ X(τ ).
Thus it follows that X(ω) ⊆ X(τ ) and hence, ω ≤ τ. Q.E.D.

4.3.12
We use the same notations as before. For j = i, . . . , n denote by pj the
canonical morphism pj : G/Q → G/Pj .

Corollary If τ ∈ W/WQ , then X(τ ) = ∩nj=1 p−1


j (pj (X(τ )) (set theoret-
ically).

Proof. Let Z denote ∩nj=1 p−1


j (pj (X(τ ))). Obviously we have X(τ ) ⊆ Z.
126 4. Schubert varieties in G/Q

Note that Z is a union of Schubert varieties. It suffices to show that if


X(ω) ⊆ Z, then X(ω) ⊆ X(τ ). Since X(τ ) ⊆ Z ⊆ p−1 j (pj (X(τ ))) ∀ j =
1, . . . , n, we know that pj (Z) = pj (X(τ )) 6 ∀j = 1, . . . , n. If X(ω) ⊆ Z,
then pj (X(ω)) ⊆ pj (Z) = pj (X(τ ))) 6 ∀j = 1, . . . , n and hence, X(ω) ⊆
X(τ ) by Lemma (4.3.11). Q.E.D.

4.4 The normalization map

We will use the same notations as in the sections before. Let τ ∈ W/WQ
and let X(τ ) be the corresponding Schubert variety in G/Q. In this
section we are going to prove that X(τ ) is smooth of codim. 1 and that
the normalization map π : X̃(τ ) → X(τ ) is bijective.

4.4.1
Let α ∈ S, τ ∈ W/WQ such that τ −1 (α) 6∈ R(Q) and τ −1 (α) < 0. Denote
by P (α) the minimal parabolic subgroup of G generated by G−α and
B, and let L(α)(≃ SL2 (k) or P SL2 (k)) denote its semisimple part. Let
B(α) be the Borel subgroup of L(α) containing Gα . Denote sα τ by ω.
Recall that we denoted the bundle L(α) ×B(α) X(ω) by Z(ω, α) and that
the canonical map Ψ : Z(ω, α) → X(τ ) is surjective (Corollary (4.2.11)).

4.4.2 Lemma
The map Ψ : Z(ω, α) → X(τ ) (b 7→ b.e(ω)) is birational.

Proof. 4 Recall that Bωu = {ΠGα | α > 0, ω −1 (α) 6∈ R(Q), ω −1 (α) <
0}. Now Bωu contains all subgroups Gα , which are not contained in the
stabilizer of e(ω). If (B −)uω = {ΠGα | α < 0, ω −1 (α) −1
 6∈ R(Q), ω (α) <
Q Q
0}, then (B − )uω .Bωu = ω α<0 Gα ω −1 . Now α<0 Gα .e(id)
α6∈R(Q) α6∈R(Q)

is open and dense in G/Q and hence, (B − )uω .Bωu .e(w) is open and dense
in G/Q. Furthermore, the map Ψ : (B − )uω .Bωu → G/Q, b 7−→ b.e(ω)
is injective and the corresponding tangent map is also injective. Now
τ −1 (α) 6∈ R(Q) and τ −1 (α) < 0. Hence, ω −1 (−α) 6∈ R(Q) and ω −1 (α) <
0. Since C(ω) ≃ Bωu (See Lemma 4.2.4) and G−α ⊆ (B − )uω , it follows
that Ψ | G−α × C(ω) → X(τ ) ⊆ G/Q is injective, the tangent map is
injective and the image is dense. Note that G−α ×C(ω) can be identified
with an open dense subset of Z(ω, α). Denote by U the open subset of
4.4. The normalization map 127

smooth points in X(τ ). Then Ψ | Ψ−1 (U ) ∩ G−α × C(ω) → X(τ ) is an


isomorphism onto an open subset of X(τ ). Q.E.D.

4.4.3 Lemma
Let α ∈ S and τ ∈ W/WQ such that τ −1 (α) 6∈ R(Q) and τ −1 (α) < 0.
Denote sα τ by ω and denote by Ψ : Z(ω, α) → X(τ ) the canonical map
as in Lemma 4.4.3, as in Lemma (4.2.10) and (4.2.11). Let e(θ) ∈ X(τ )
for some θ ∈ W/WQ .

(i) Ψ−1 e(θ)) is a point ⇔ L(α).e(θ) is not contained in X(ω).

(ii) Ψ−1 (e(θ)) ≃ P1 ⇔ L(α).e(θ) ⊆ X(ω).

(iii) Every fibre of Ψ is either a point of P1 (set theoretically). In


particular, it is connected.

Proof. Consider the natural map λ : L(α)×X(ω) → X(τ ). If x ∈ X(τ ),


then λ−1 (x) = {(g, y) | y ∈ L(α).x ∩ X(ω), g.y = x}. If L(α).e(θ) is not
contained in X(ω), then L(α).e(θ) ∩ X(ω) is discrete. Since the intersec-
tion is T (α)-stable, we know that L(α).e(θ) ∩ X(ω) ⊆ {e(θ) ∪ e(sα θ)}.
If L(α).e(θ) ∩ X(ω) = {e(θ) ∪ e(sα θ)} then X(θ) ⊆ X(ω) and X(sα θ) ⊆
X(ω) and hence, L(α).e(θ) ⊆ X(ω). So we can assume without loss of
generality that e(θ) 6∈ X(ω). We know that L(α).e(θ) ∩ X(ω) = e(sα θ)
(Lemma 4.2.12), sα θ < θ and (sα θ)−1 (α) 6∈ R(Q) and (sα θ)−1 (α) > 0.
But then ψ −1 (e(θ)) is isomorphic to the isotropy group I of L(α) at
e(sα θ). Now I ≃ B(α) (Lemma 4.2.5) and hence, ψ −1 (e(θ)) is a point
since L(α) ×B(α) X(ω) = Z(ω, α). If L(α).e(θ) is contained in X(ω),
then ψ −1 (e(θ)) = {(g −1 , gy) | y any fixed point on L(α).e(θ), g ∈ L(α)}.
Hence, ψ −1 (e(θ)) is a principal fibration over P1 with structure group
B(α) and hence we see that ψ −1 (e(θ)) is isomorphic to P1 . The proof of
(iii) follows from (i) and (ii) and the B-equivariance of ψ. Q.E.D.

4.4.4 Corollary
]) → X(τ ) is bijective.
The normalization map π : X(τ

Proof. Induction on dim X(τ ). If dim X(τ ) = 0, nothing is to be


shown. Let dim X(τ ) = r > 0. Choose a simple reflection sα such that
128 4. Schubert varieties in G/Q

τ −1 (α) 6∈ R(Q) and τ −1 (α) < 0. By induction hypothesis the normal-


ization map π : X̃(sα τ ) → X(sα τ ) is bijective. Denote sα τ by ω. Note
that the normalization map is B-equivariant. Denote by Z̃(ω, α) the
fibre space L(α) ×B(α) X̃(ω). Note that Z̃(ω, α) is normal. Consider the
commutative diagram

L(α) × X̃(ω) L(α) × X(ω) X(τ )

π̃
π

Z̃(ω, α) Z(ω, α)
η

We get a morphism π̃ : Z̃(ω, α) → X(τ ) such that the fibres of π̃ are


connected, since η is bijective and the fibres of π are connected (Lemma
(4.4.3)); moreover, Z(ω, α) is normal and π̃ is birational. Consider the
Stein factorization of π̃ (see [I], Theorem (2.26)):

e
Z(ω, α)

 
π̃ Spec π̃∗ (OZ̃(ω,α) )

X(τ )

Now Spec π̃∗ (ÕZ̃(ω,α) ) is a normal variety and the fibre of ζ at a point
x ∈ X(τ ) contains as many points as the connected components of the
fibre of π̃ over x. Since the fibres of π̃ are connected, ζ is bijective.
Since ζ factors through the normalization map, the normalization map
is bijective too. Q.E.D.
4.5. Chevalley’s multiplicity formula 129

4.4.5 Corollary

X(ω) is smooth in codim. 1.

Proof. We use decreasing induction on dim X(ω). If X(ω) = G/Q,


there is nothing to show. Assume the claim is true for all Schubert
varieties of dim > dim X(ω) = r. Note that the singular set is B-stable.
Since the orbit B.e(ω) is open in X(ω), e(ω) is a smooth point of X(ω).
So we have only to show that if e(θ) ∈ X(ω) such that X(θ) is of codim.
1 in X(ω), then e(θ) is a smooth point of X(ω). If θ ≡ sα ω mod WQ ,
where α is simple, then e(θ) = e(sα ω) ∈ L(α)e(ω) and hence, a smooth
point of X(ω). Suppose we cannot write θ in that way. Choose α ∈ S
such that θ < sα θ. (Since X(θ) is at least of codim. 1 in G/Q, we can
find such an α because otherwise X(θ) would be stable under all L(α), α
simple, and hence stable under G, which would imply that X(θ) = G/Q.)
We want to show that ω < sα ω. Assume this is not the case. Then
X(ω) is stable under L(α); in particular, e(sα θ) ∈ X(ω) and hence,
sα θ ≡ ω mod WQ because dim X(ω) = dim X(θ) + 1 = dim X(sα θ).
But this is a contradiction to our assumption that θ 6≡ sα ω mod WQ for
some α ∈ S. Denote sα ω by τ. Then X(sα θ) ⊆ X(τ ) and is of codim. 1
in X(τ ). Consider the map ψ : Z(α, ω) → X(τ ). Since e(sα , θ) 6∈ X(ω),
we know that ψ −1 (e(sα θ)) is a point (Lemma (4.4.3)). Now e(sα θ) is
a smooth point of X(τ ) by induction hypothesis. Let U be an open
neighborhood of e(sα θ) in X(τ ) such that u ∈ U is a smooth point of
X(τ ) and ψ −1 (u) is a point. By Zariski’s Main Theorem ([M] p.413)
ψ : ψ −1 (U ) → U is an isomorphism. Hence, ψ −1 (e(sα θ)) is a smooth
point of Z(ω, α) which is represented by sα .e(θ). Hence, e(θ) is a smooth
point of X(ω). Q.E.D.

4.5 Chevalley’s multiplicity formula

Denote by A the Chow ring of G/Q for rational equivalence. For ω ∈


W/WQ , denote by [X(ω)] the equivalence class of the Schubert variety
X(ω) in the Chow ring. Let X(τ ) be a Schubert variety of codim. 1 in
G/Q. We want to derive a formula for the intersection multiplicity of a
Schubert variety X(ω) and X(τ ) considered as subvarieties of G/Q along
the codim. 1 Schubert varieties in X(ω). We will prove the multiplicity
formula first for the case Q = B.
130 4. Schubert varieties in G/Q

4.5.1
Let ω0 ∈ W denote the uniquely determined element in W such that
ω0 (φ+ ) = φ− . We know that X(ω0 ) = G/B (Lemma (4.2.4)).

Lemma
(i) If X(ω) is of codim. 1 in G/B, then ω = sα ω0 for some α ∈ S.

(ii) If X(ω) is of codim. 1 in X(τ ), then ω = τ.sβ for some β > 0.

Proof.
(i) For α ∈ S, we have either sα ω > ω or sα ω < ω. If sα ω < ω, then
X(ω) is of codim. 1 in X(sα ω), so that sα ω = ω0 and we are done.
Assume that sα ω < ω, ∀ α ∈ S. This would imply that X(ω) is
stable under all the minimal parabolic subgroups P (α), α ∈ S.
But then X(ω) would be stable under G and hence, ω = ω0 which
contradicts our assumption that X(ω) is of codim. 1 in G/B.

(ii) Let τ = s1 , . . . , sr be a reduced decomposition of τ and let ω =


s1 , . . . , si−1 , ŝi .si+1 . . . . .sr be a reduced decomposition of ω given
by a subword of τ. If β denotes (sr . . . . .si+1 )(αi ) = (si .si−1 . . . . .sl τ )−1 (αi )
(note that β > 0 by Corollary 4.3.5), then

ω = τ (sr . . . . .si+1 )(si )(si+1 . . . . .sr )


= τ sβ .

Q.E.D.

4.5.2
Denote by B − the opposite Borel subgroup ω0 Bω0 . Let ω ∈ W and de-
note the closure of the orbit B − .e(ω) endowed with the canonical struc-
ture of a reduced subscheme of G/Q by Y (ω). Note that B − .e(ω) =
(ω0 Bω0 ).e(ω) = ω0 B.e(ω0 ω) and hence Y (ω) = ω0 X(ω0 ω).

Lemma If ω, τ ∈ W, then ω ≤ τ ⇔ ω0 ω ≥ ω0 τ and the codimension of


X(ω) in X(τ ) is the codimension of X(ω0 .τ ) in X(ω0 ω).

Proof. Since ω02 = id, it is sufficient to show “⇒.” By Corollary


(4.3.6) it is sufficient to prove the lemma for the case when X(ω) is of
4.5. Chevalley’s multiplicity formula 131

codimension one in X(τ ). By Corollary (4.3.9), there is a β > 0 such


that ω = sβ τ and hence, ω −1 (β) > 0. Since (ω0 ω)−1 (ω0 β) = ω −1 (β), we
know that (ω0 ω)−1 (ω0 (β)) > 0. By Lemma 4.2.5

L(ω0 (β)).e(ω0 ω) = G−ω0 (β) .e(ω0 ω) ∪ e(sω0 (β) .ω0 ω)


= G−ω0 (β) .e(ω0 ω) ∪ e(ω0 sβ ω)
= G−ω0 (β) .e(ω0 ω) ∪ e(ω0 τ )
= G−ω0 (β) .e(ω0 ω)

Since −ω0 (β) ∈ φ+ we know that e(ω0 τ ) ∈ B.e(ω0 ω) = X(ω0 ω). But
this implies that X(ω0 τ ) ⊆ X(ω0 ω) and ω0 ω ≥ ω0 τ. Since dim X(ω0 ω) =
#N (ω0 ω) = #{φ+ − N (ω)} = #{φ+ − N (τ )} + 1 = #N (ω0 τ ) + 1, it
follows that the codim of X(ω0 τ ) in X(ω0 ω) is 1. Q.E.D.

4.5.3 Corollary
X(ω) ⊆ X(τ ) ⇔ Y (ω) ⊇ Y (τ ) and the codimension of X(ω) in X(τ ) is
the codim of Y (τ ) in Y (ω).

Proof. This follows from Lemma (4.5.2) and the fact that Y (ω) =
ω0 X(ω0 ω). Q.E.D.

4.5.4 Lemma
X(ω) ∩ Y (ω) = e(ω) and the intersection is transversal.

Proof. Obviously we have e(ω) ∈ X(ω)∩Y (ω). Let Tx X denote the tan-
gent space to a scheme X at x ∈ X. Since Te(ω) G/B ≃ {⊕(Lie G)α | α ∈
φ and ω −1 (α) < 0}, Te(ω) X(ω) ≃ {⊕(Lie G)α | α > 0 and ω −1 (α) < 0}
and Te(ω) Y (ω) ≃ {⊕(Lie G)α | α < 0 and ω −1 (α) < 0}, we get :
Te(ω) G/B = Te(ω) X(ω)⊕Te(ω) Y (ω); the intersection is hence transversal
at e(ω) and e(ω) is an isolated point in the intersection. The intersec-
tion is T -stable and by Borel’s Fixed Point Theorem, every irreducible
component of the intersection contains a T -fixed point. Hence, it is suf-
ficient to show that e(ω) is the only T -fixed point in the intersection. If
e(θ) ∈ X(ω) ∩ Y (ω) is a T -fixed point, then Be(θ) = X(θ) ⊆ X(ω) and
B − e(θ) = Y (θ) ⊆ Y (ω). Hence, by Corollary (4.5.3), θ = ω. Q.E.D.
132 4. Schubert varieties in G/Q

4.5.5 Lemma
If ω 6= τ, but dim X(ω) = dim X(τ ), then X(ω) ∩ Y (τ ) = ∅.

Proof. Assume that X(ω) ∩ Y (τ ) 6= ∅. The intersection is T -stable and


hence, contains a T -fixed point. If e(θ) ∈ X(ω) ∩ Y (τ ), then X(θ) ⊆
X(ω) and Y (θ) ⊆ Y (τ ) and hence, X(τ ) ⊆ X(ω) (Corollary (4.5.3)).
Since both have the same dimension, we have τ = ω. Q.E.D.

4.5.6
Let X(ω) ⊂ X(τ ) be of codim. 1. By Corollary 3.9 we know that ω ≡
sβ τ for some β > 0 and hence, L(β).e(ω) = G−β e(ω) ⊆ X(τ ) (Lemma
4.2.5). Since Y (ω) ⊇ Y (τ ) (Lemma 4.5.2), we get L(β).e(ω) ⊆ Y (ω) and
hence, L(β).e(ω) ⊆ Y (ω) ∩ X(τ ). Denote L(β).e(ω) by P1 (ω, τ ). (Recall
that L(β).e(ω) ≃ P1 by Lemma (4.2.4).)

Lemma If X(ω) ⊂ X(τ ) is of codim. 1, then Y (ω) ∩ X(τ ) = P1 (ω, τ )


and the intersection is transversal.

Proof. The intersection Y (ω) ∩ X(τ ) is T -stable. If e(θ) ∈ Y (ω) ∩ X(τ )


is a T -fixed point, then θ ≥ ω and θ ≤ τ. Hence, either θ = ω or
θ = τ , because X(ω) is of codim. 1 in X(τ ). It follows that the
only T -fixed points in Y (ω) ∩ X(τ ) are e(ω) and e(τ ). We know by
Lemma (4.5.4) that Te(τ ) X(τ ) ⊕ Te(τ ) Y (τ ) = Te(τ ) G/B. Since Y (τ ) ⊆
Y (ω) is of codim 1, e(τ ) is a smooth point in Y (ω) (Corollary 4.4.5).
Hence, Te(τ ) Y (τ ) ⊆ Te(τ ) Y (ω) is of codim. 1. But this implies that
Te(τ ) X(τ )+Te(τ ) Y (ω) = Te(τ ) G/B and Te(τ ) X(τ )∩Te(τ ) Y (ω) has dim 1.
Hence, X(τ ) ∩ Y (ω) is transversal at e(ω) and the only irreducible com-
ponent of X(τ ) ∩ X(ω) containing e(τ ) is P1 (ω, τ ). The same arguments
show that the intersection X(ω0 ω) ∩ Y (ω0 τ ) is transversal at e(ω0 ω).
Since ω0 (X(ω0 ω)) = Y (ω), ω0 (Y (ω0 ω)) = X(τ ) and ω0 (e(ω0 ω)) = e(ω),
it follows that X(τ ) ∩ Y (ω) is transversal at e(ω). Hence, the only irre-
ducible component of X(τ ) ∩ Y (ω) containing e(ω) is P1 (ω, τ ). Hence,
X(τ ) ∩ Y (ω) = P1 (ω, τ ) and the intersection is transversal.

4.5.7
Let S = {α1 , . . . , αr } be a basis of the root system φ of G and de-
note by ̟1 , . . . , ̟r the corresponding fundamental weights (see 4.1.8).
4.5. Chevalley’s multiplicity formula 133

Let X(sk ω0 ), αk ∈ S, be a Schubert variety of codim. 1 in G/B


(Lemma 4.5.1) and let X(τ ) be a Schubert variety in G/B such that
X(τ ) is not contained in X(sk ω0 ). Since the set {[X(ω)] | ω ∈ W } gen-
erates the additive group of the Chow ring A of G/B, we know that :
P
t
[X(τ )].[X(sk ω0 )] = di [X(sβi τ )], where βi > 0, X(sβi τ ) is of codim.
i=1
1 in X(τ ) (Corollary 4.3.9) and di ∈ N (resp. [X(τ )].[X(sk ω0 )] =
Pt
dj [X(τ sγj )], where γj > 0, X(τ sγj ) is of codim. 1 in X(τ ) (Lemma
j=1
4.5.1) and dj ∈ N); the central dot denotes the multiplication in A. We
call di the intersection multiplicity of [X(sβi ω)] in [X(ω)].[X(sk ω0 )].

(̟k ),βi ) 2(̟k ,γj )
Theorem (Chevalley)): di = 2(τ(β i ,βi )
(resp. (γj ,γj ) )

Proof. Let X(ωi ) ⊆ X(τ ) be of codim. 1. We know by Lemma (4.5.4)


that the intersection X(ωi ) ∩ Y (ωi ) is transversal, X(ωi ) ∩ Y (ωi ) = e(ωi )
and X(ωi ) ∩ Y (ωj ) = ∅, if ωi 6= ωj and dim X(ωi ) = dim X(ωj ).
Hence, [X(τ )].[X(sk ω0 )].[Y (sβi τ )] = di . By Lemma (4.5.5), we know
that the intersection X(τ )∩ Y (sβi τ ) is transversal and X(τ )∩ Y (sβi τ ) =
P1 (sβi τ, τ ). Hence [X(τ )].[X(sk ω0 )].[Y (sβi τ )] = [X(sk ω0 )].[P1 (sβi , τ )] =
di . By Lemma (4.5.1), we know that there exists a γ > 0 such that sβi τ =
τ sγ . The cycles P1 (sβi τ, τ ) and τ −1 P1 (sβi τ, τ ) = P1 (sγ , e) are hence al-
gebraically equivalent. The multiplicity of the intersection X(sk ω0 ) ∩
P1 (sγ .e) is given by the degree of the line bundle on P1 (sγ .e) ≃ L(γ)/B(γ)
induced by the line bundle on G/B corresponding to X(sk ω0 ). The in-
duced line bundle corresponds to the restriction of the character
̟k to
k ),γ
L(γ). The degree of this line bundle is given by di = 2(̟ (γ,γ) . Since the

2(τ (̟k ),τ (γ)) 2(τ (̟k ),βi )
scalar product is W -invariant, we get di = (τ (γ),τ (γ)) = (βi ,βi ) .
Q.E.D.

4.5.8
Let Q ⊃ B be a parabolic subgroup of G. Denote the Weyl group of
Q by WQ . Let φ0 ∈ W be the element such that φ0 (φ+ ) = φ− . Denote
its canonical image in W/WQ by ω0 . Every Schubert variety of codim
1. in G/Q can be written as X(sk ω0 ), αk ∈ S (same arguments as
in Lemma 4.5.1). Let X(τ ) ⊆ G/Q be a Schubert variety such that
X(τ ) is not contained in X(sk ω0 ). As in the case Q = B, we know that
134 4. Schubert varieties in G/Q

P
t
[X(τ )].[X(sk ω0 )] = di [(X(sβi τ ))], where βi > 0, X(sβi τ ) ⊆ X(τ ) of
i=1
codim. 1 (Corollary (4.3.9)) and di ∈ N is called the intersection multi-
plicity. Denote by abuse of notation the minimal representative of τ in
W also by τ. Since X(sk φ0 ) → G/B is the preimage of X(sk ω0 ) ֒→ G/Q
under the projection G/B → G/Q, we get as an immediate consequence
of the projection formula ([S], V-29) the same formula as before:

(̟k ),βi )
Corollary di = 2(τ(β i ,βi )

4.5.9 Corollary
If G = SLn+1 (k), then di = 1, ∀ i = 1, . . . , n.

Proof. This follows immediately from the fact that every positive root
Pn
β ∈ φ+ can be written as β = ni αi such that ni = 0 or ni = 1 ∀ i =
i=1
1, . . . , n (see e.g. [HU2], . 64). Q.E.D.

Remark This proves statement (iii) of Lemma (4.5.6) in Chapter 2.

4.6 Deodhar’s Lemma

In this section we are going to prove Deodhar’s Lemma and derive as a


consequence the existence of the minimal and maximal defining chains
for a standard Young diagram on a Schubert variety X(ω) ֒→ G/Q.

4.6.1
Let Q1 , Q2 be two parabolic subgroups of G containing B such that Q1 ⊃
Q2 . Denote the set of minimal representatives of W/WQ1 in W/WQ2 (see
(4.3.10)) by Ω(Q1 , Q2 ).

Deodhar’s Lemma Let σ2 .σ2 ∈ Ω(Q1 , Q2 ) be such that σ1 ≥ σ2 .

(i) Let ω1 ∈ WQ1 . There exists a ω2 ∈ WQ1 such that σ1 ω1 ≥ σ2 ω2 in


W/WQ2 and ω2 is maximal for this property and unique modulo
WQ 2 .

(ii) (dual version) Let ω2 ∈ WQ1 . There exists a ω1 ∈ WQ1 such that
σ1 ω1 ≥ σ2 ω2 in W/WQ2 and ω1 is minimal for this property and
4.6. Deodhar’s Lemma 135

unique modulo WQ2 .

Proof. This is by induction on ℓQ1 (σ1 ). If ℓQ2 (σ1 ) = 0 set ω2 = ω1


(resp. ω1 = ω2 in the dual case). Let ℓQ2 (σ1 ) ≥ 1. Choose a simple
reflection sα such that sα σ1 < σ1 . Note that sα σ1 ∈ Ω(Q1 , Q2 ).

Case 1: sα σ2 < σ2 . Then sα σ1 ≥ sα σ2 (Lemma 4.2.13) and sα σ2 ∈


Ω(Q1 , Q2 ). For any ω ∈ Ω(Q1 , Q2 ), τ ∈ WQ1 the length function is
additive: ℓQ2 (ωτ ) = ℓQ2 (ω) + ℓQ2 (τ ). Hence by Lemma 4.2.13, σ1 τ ≥
σ2 τ ′ ⇔ sα σ1 τ ≥ sα σ2 τ ∀τ, τ ′ ∈ WQ1 . We know by induction that there
exists a ω2 ∈ WQ1 (resp. a ω1 in WQ1 in the dual case) such that
sα σ1 ω1 ≥ sα σ2 ω2 in W/WQ2 . Hence, σ1 ω1 ≥ σ2 ω2 in W/WQ2 and ω2
is maximal (resp. ω1 is minimal in the dual case) for this property and
unique modulo WQ2 .

Case 2: sα σ1 ≥ σ2 . Then σ2 ≤ sα σ1 (Lemma 4.2.13). We want to prove


∀τ, τ ′ ∈ WQ1 , sα σ1 τ ≥ σ2 τ ≥ σ2 τ ′ ⇔ σ1 τ ≥ σ2 τ ′ . Since σ1 , sα σ1 ∈
Ω(Q1 .Q2 ) and σ1 > sα σ1 , “ ⇒′′ follows immediately. Now σ2 ≤ sα σ1
and hence σ2 is a subword of sα σ1 (Lemma 4.3.8). Since σ2 ∈ Ω(Q1 , Q2 ),
“⇐” follows. Now by induction there exists a ω2 ∈ WQ1 (there exists
a ω1 ∈ WQ1 in the dual case) such that sα σ1 ω1 ≥ σ2 ω2 and hence
σ1 ω1 ≥ σ2 ω2 in W/WQ2 such that ω2 is maximal (resp. ω1 is minimal
in the dual case) for this property and unique modulo WQ2 . Q.E.D.

4.6.2
Let G = SLn+1 (k). Let S = {α1 , . . . , αn } be a basis of the root system φ
of G and let B be the corresponding Borel subgroup. Denote by Pi the
maximal parabolic subgroup of G such that SPi = S − {αi }. Denote by
Qi the parabolic subgroup Pi ∩ Pi+1 ∩ . . . ∩ Pn of G. Recall that a Young
diagram of type a(i) = (ai , ai+k , . . . , an ), aj ≥ 0, ∀ j = 1, . . . , n, is an
element θ ∈ (W/WPi × . . . × W/WPi ) × . . . × (W/WPn × . . . × W/WPn ).
| {z } | {z }
ai an
Let ω ∈ W/WQi and let X(ω) ֒→ G/Qi be the corresponding Schu-
bert variety. Denote by pj the canonical morphism W/WQi → W/WPj ,
j = i, . . . , n. Recall that θ = (θi,1 , . . . , θi,ai , θi+1,1 , . . . , θn,an ) is said to be
a Young diagram on X(ω) if pj (ω) ≥ θj,kj , ∀ j = 1, . . . , r, kj = 1, . . . , an .
Recall that we define a Young diagram to be standard Q with respect to
X(ω) if there exists a defining chain φ for θ, i.e., ∃ φ ∈ nj=1 (W/WQi )aj ,
136 4. Schubert varieties in G/Q

φ = (φi,1 , . . . , φi.ai , φi+1,1 , . . . , φn.an ) such that

1) φj,kj is a lift for θj,kj ∀ i ≤ j ≤ n, 1 ≤ kj ≤ an

2) ω ≥ φi,1 ≥ φi,2 ≥ · · · ≥ φi,ai ≥ φi+1,1 ≥ · · · ≥ φn,an . We say θ is


standard if θ is standard on G/Q.

4.6.3 Corollary
Let θ be a standard Young diagram of type a(i) on X(ω) ֒→ G/Qi .
There exists a unique maximal defining chain φ+ and a unique minimal
defining chain φ− for θ such that if ψ is any defining chain for θ, then
φ+ −
j,kj ≥ ψj,kj ≥ φj.kj , i ≤ j ≤ n, 1 ≤ kj ≤ aj .

Proof.
Case 1: Maximal defining chain
Let X be a defining chain for θ. If X(ω) = G/Qi , define φ+ i,1 to be the
+
maximal representative of θi,1 . We have φi,1 ≥ ψi,1 for any defining chain
ψ of θ and φ+ i,1 is unique. If X(ω) 6⊆ G/Qi , denote by σ ∈ W/WQi a
minimal representative of pi (ω) ∈ W/WPi and denote by σi,1 ∈ W/WQi
a minimal representative of θi,1 . Since ω ≥ Xi,1 , we know that σ ≥ σi,1 .
Thus ω ≡ σ.ζ mod WQi for some ζ ∈ WPi and by Deodhar’s Lemma
+ + +
there exists a ζi,1 ∈ WPi such that ω ≥ σi,1 ζi.1 and ζi,1 is maximal for
that choice and unique modulo WQ . Denote σi,1 ζi,1 by φi,1 . Then φ+
+ +
i,1 is
a lift of θi,1 , ω ≥ φi,1 , φi,1 ≥ ψi,1 for any defining chain ψ for θ and φ+
+ + +
i,1
is unique for these properties. Now assume that we have already con-
structed φ+ + +
i,1 , φi,1 , . . . , φj,k . First assume that k < aj . Denote by σj,k a
minimal representative of θj,k = pj (φ+ j,k ) = pj (Xj,k ) and let σj,k+1 denote
a minimal representative of θj,k+1 = pj (Xj,k+1 ). Since Xj,k ≥ Xj,k+1 ,
we know that σj,k ≥ σj,k+1 . Since φ+ +
j,k ≡ σj,k .ζj,k mod WQ there ex-
+
ists a Deodhar’s Lemma a ζj,k+1 ∈ WPi such that φ+ +
j,k ≥ σj,k+1 .ζj,k+1
+
and ζj,k+1 is maximal for this property and is unique modulo WQ . De-
+
note σj,k+1 .ζj,k+1 by φ+ + + +
j,k+1 . Then φj,k+1 is a lift of θj,k+1.φj,k ≥ φj,k+1 ,
+ +
φj,k+1 ≥ ψj,k+1 for any defining chain ψ and φj,k+1 is unique for this
property. Now assume k = aj . Denote by σ a minimal representa-
tive of Pj+1 (φ+ j,aj ) and let σj+1,1 denote a minimal representative of
θj+1,1 in W/WQ . Since φ+ j,aj ≥ ψj,aj ≥ Xj+1,1 , we know that σ ≥ σi+1 .
+
Hence by Deodhar’s Lemma we can find a ζj+1,1 ∈ WPj+1 such that
+ + +
φj,aj ≡ σ.ζ ≥ σj+1,1 .ζj+1 in W/WQ2 and ζj+1 is maximal for this prop-
4.6. Deodhar’s Lemma 137

erty and unique modulo WQ . Hence, φ+ +


j+1,1 = σj+1,1 .ζj+1 is a lift of
θj+1,1 φ+ + +
j,aj ≥ φj+1,1 , φj+1,1 ≥ ψj+1,1 for any defining chain ψ and φj+1,1
+

is unique for this property. Now let φ+ denote (φi,1 , φi,2 , . . . , φn,an ). By
construction φ+ is a defining chain for θ and φ+ is maximal.

Case 2: Minimal defining chain


Let X be a defining chain for θ. Define φ− n,an ∈ W/WQ to be a mini-
mal representative for θn,an . Assume that we have already constructed
φ− − −
j,k , φj,k+1 , . . . , φn,an . First assume that k > 1. Denote by σj,k−1 ∈
W/WQi the minimal representative of θj,k−1 and denote by σj,k the min-
imal representative of θj,k . Now Xj,k−1 ≥ Xj,k and hence σj,k−1 ≥ σj,k .
Since φ− j,k ≡ σj,k .ζ mod WQ for some ζ ∈ WPj we know by Deodhar’s

Lemma that ∃ ζj,k−1 ∈ WPj such that φ− − −
j,k−1 ≡ σj,k−1 ≥ φj,k and ζj,k−1
is minimal for this property and unique mod WQ . Hence, φ− j,k−1 is a lift
− −
of θj,k−1, θj,k−1 ≥ φj,k , φj,k−1 ≤ ψj,k−1 for any defining chain ψ for θ and
φ−j,k−1 is unique for these properties. Now assume k = 1, j 6= 1. Denote
by σj−1,aj −1 ∈ W/WQi the minimal representative of θj−1,aj −1 and by
σ ∈ W/WQi the minimal representative of Pj−1 (φ− j,1 ). Since Xj−1,aj−1 ≥
Xj,1 ≥ φj,1 , we know that σj−1,aj−1 ≥ σ. Now φ−

j,1 ≡ σ.ζ mod WQi for

some ζ ∈ WPj−1 and hence by Deodhar’s Lemma ∃ ζj−1,a j−1
∈ WQ i
− − − −
such that φj−1,aj−1 ≡ σj−1,aj−1 .ζj−1,aj−1 ≥ φj,1 and ζj−1,aj−1 is minimal
for this property and unique modulo WQi . Hence, φ− j−1,k−1 is a lift of
θj−1,aj−1 , φ− j−1,aj−1 , φ −
j−1,aj−1 ≥ φ−
φ−
j,1 j−1,aj−1 ≤ ψ j−1,k−1 for any defin-

ing chain ψ and φj−1,aj−1 is unique for these properties. Now let φ−
 
denote φ− , φ
i,1 i,2

, . . . , φ − −
n,an . By construction φ is a defining chain for
θ and φ− is minimal.

Remark Note that the minimal defining is “independent of ω” in the


sense that if λ is a standard diagram on G/Qi , then λ is standard on
X(ω) if and only if ω ≥ φ−i,1 .
Appendix A

Cohen-Macaulay Properties

By V. Lakshmibai

Part I.

Definition 1 A graded poset is a finite partially ordered set P with a


unique maximal and a unique minimal element (denoted 1̂, 0̂ respec-
tively) in which all maximal chains 1̂ = x0 > x1 > · · · > xr = 0̂ have
the same length r. This common length r is called the rank of P.

Definition 2 A graded poset P is said to be lexicographically shellable


(or L-shellable (cf.[B-W1] Definition 3.2 or CL-shellable (cf. [B-W2]
Definition 2.2)), if every maximal chain m : 1̂ = x0 > x1 > · · · > xr = 0̂
can be labelled, say λ(m) = (λ1 (m), λ2 (m), . . . λr (m)) where λi (m) are
elements of another partially ordered set in such a way that the following
two conditions hold:

(L-1) If two maximal chains m, m′ have the same first d-edges, then the
corresponding labels of the first d-edges are the same (here, we think of
the element λi (m) as being associated with the edge xi−1 → xi ).

(L-2) Given an interval [x, y], together with a path c from 1̂ to y


(we refer to this as a rooted interval), among all the maximal chains
in [x, y], there exists a unique chain whose label is increasing; further
this unique increasing label is lexicographically < the label of any other
chain (here, a label for a maximal chain in [x, y] is the one induced by
the maximal chain obtained by following c by the maximal chain from
y to x under consideration, followed by an arbitrary chain from x to 0̂;
e.g., the Bruhat order in S3 . Denoting the transposition (1,2) and (2,3)
by s1 , s2 respectively, we may write

( 3 2 1 ) = s1 s2 s1 .

We label the maximal chains as shown in the diagram below:

© Springer Science+Business Media Singapore 2016 and Hindustan Book Agency 2015 139
C.S Seshadri, Introduction to the Theory of Standard Monomials,
Texts and Readings in Mathematics 46, DOI 10.1007/978-981-10-1813-8
140 Appendix A

w = s1 s2 s1

s 2 s 1 = τ2 • • • φ2 = s1 s2

s 1 = τ1 • • φ1 = s2

τ0 = id
1 • 1 • •3 •3
• • • •
3 1
2 2
• • • •
3 •2 1 • • 2

Remark The same edge occurring in two different maximal chains


may have different indexing. For instances, in the above example the
edge τ1 • • τ0 has the index 3 when considered as an edge in
m : w > τ2 > τ1 > τ0 and has the index 1 when considered as an edge
mm′ : w > φ2 > τ1 > τ0 .

Lexicographic shellability for Bruhat orders

Let (W, S) be a Coxeter group. For J ⊆ S, let W J be the set


of minimal representatives of W/WJ (here WJ is the subgroup of W
generated by the simple reflections sα , α ∈ J). Recall (e.g. [L-M-S1])
that W J = {w ∈ W | w(α) > 0, α ∈ J}. For w, w′ ∈ W J , note that the
interval [w, w′ ] consists of {u ∈ W J | w ≤ u ≤ w′ }. Following [B-W1],
we shall describe a labeling of the maximal chains in [w, w′ ]. Let us fix
a reduced expression w′ = s1 s2 . . . sq . Let m : w′ = w0 > w1 > · · · >
wr = w be a maximal chain in [w, w′ ], where r = ℓ(w′ )− ℓ(w). Now since
w1 < w′ and ℓ(w1 ) = ℓ(w′ ) − 1, a reduced expression for w1 is obtained
by omitting a reflection si in the reduced expression s1 s2 . . . sq for w′
and the deleted reflection is uniquely determined (cf. [C], for example).
We set λ1 (m) = i. Proceeding thus, at each step, we label an edge by the
position of the reflection that is deleted in the chosen reduced expression
for w′ .

Theorem 1 (Björner-Wachs. (cf. [B-W1], Theorem 4.2). Notations


being as above, the interval [w, w′ ] in W J is lexicographically shellable
for the labelling of the maximal chains in [w, w′ ] as described above.
Cohen-Macaulay Properties 141

Definition 3 The order complex: Given a finite poset P, we can


associate a simplicial complex (called the order complex of P and denoted
by △(P )), by taking a q-simplex to be a chain (i.e., a totally ordered
subset of P ) of length q.
We now define “shelling” of a simplical complex △ using induction
on dim △, the definition being obvious when dim △ = 0.

Definition 4 (cf. [B-W2], Definition 4.1) Let △ be a pure d-dimensional


simplicial complex, i.e., all maximal faces are of dimension d. An ordering
 facets (i.e., the maximal faces) of △ is said to be
F1 , F2 , . . . , Ft of the
j−1
a shelling if Fj ∩ ∪i=1 Fi is a (d − 1)-dimensional complex having a
shelling which extends to  a shelling
 of ∂Fj (i.e., ∂Fj has a shelling in
j−1
which the facets of Fj ∩ ∪i=1 Fi come first). Here, for a simplex F, F
denotes {G | G ⊆ F }. We say △ is shellable if it admits a shelling.

Theorem 2 (cf. [B-W1], Theorem 3.3) If P is lexicographically shellable,


then the order complex △(P ) is shellable.

Proof. In fact if P is L-shellable, we consider the total ordering < of


the maximal chains in △(P ) given by

m1 < m2 < · · · < ms

where

λ(m1 ) < λ(m2 ) < · · · < λ(ms ).

One may proceed as in [B-W1] to prove that the above total ordering
gives a shelling for △(P ).

Consequences for Schubert varieties


Let G be a connected semisimple algebraic group, T a maximal torus in
G, B a Borel subgroup of G, B ⊃ T and Q a parabolic subgroup contain-
ing B. Let W be the Weyl group of G relative to T. Let WQ denote the
Weyl group of Q and W Q the set of minimal representatives of W/WQ .
For τ ∈ W Q , let X(τ )(= Bτ Q(mod Q)) be the Schubert variety in G/Q
associated to τ. As in Definition 4, we define below a “nice indexing”on
a union of Schubert varieties inductively, the definition being obvious in
the case of dimension 0.
142 Appendix A

Definition 5 Let Z be an union of Schubert varieties in G/Q. Further,


let Z be pure of dimension d. We say Z admits a nice indexing, if there
exists an indexing, say X1 , . . . , Xr of the components of Z in such a way
that
 
j−1
(1) Xj ∩ ∪i=1 Xi is pure of dimension d − 1, for 2 ≤ j ≤ r.
 
j−1
(2) Xj ∩ ∪i=1 Xi admits a nice indexing which extends to a nice
indexing of Y (= the union of the Schubert divisors in Xj ), 2 ≤
j ≤ r(i.e., Y 
admits a nice indexing in which the components of
j−1
Xj ∩ ∪i=1 Xi come first).

Theorem 3 Let τ ∈ W Q and let Z be the union of the Schubert


divisors in X(τ ). Then Z admits a nice indexing.

Proof. The result follows immediately from Theorem 1 and 2.

Theorem 4 Let τ ∈ W Q and let P be a parabolic subgroup of G


containing Q. Let π : G/Q → G/P be the canonical projection. Let
Y = ∪w∈Bτ X(w) where

Bτ = {w ∈ W Q | X(w) is a divisor in X(τ ) and π(X(w)) ⊂ π(X(τ ))}.

Then Y admits a nice indexing.

Proof. Let us write τ = τ0 θ0 , where τ0 ∈ W P and θ0 ∈ WP and


ℓ(τ ) = ℓ(τ0 ) + ℓ(θ0 ). Let τ0 = s1 . . . st and θ0 = st+1 . . . st+k be reduced
expressions for τ0 and θ0 so that τ = s1 . . . st st+1 . . . st+k gives a reduced
expression for τ. Working with this reduced expression for τ, we index
the maximal chains in [id, τ ]. Now if X(w) is a divisor in X(τ ), then
w ∈ Bτ if and only if w = s1 . . . ŝℓ . . . st+k for some 1 ≤ ℓ < t. Hence the
maximal chains in [id, τ ] which start with an edge τ • • w, w ∈ Bτ .
come earlier (in the lexicographic ordering) than those which start with
an edge τ • • w, w ∈ Bτ . Now the result follows from Theorems 2
and 3 (recall that the linear ordering of the maximal chains in the proof
of Theorem 2 is given by

m1 < m2 < · · · < ms


and
Cohen-Macaulay Properties 143

λ(m1 ) < λ(m2 ) < · · · < λ(ms )).

Cohen-Macaulay properties for Schubert varieties in SLn /Q

Theorem 5 Let τ ∈ W and let L(π) be any ample line bundle on


P

G/B, say π = ai ω i , where ω are fundamental weights, ℓ = n − 1
i=1
[) be the cone over X(τ ) defined by the projective
and ai > 0. Let X(τ
[)
embedding of X(τ ) associated to the ample line bundle L(π). Then X(τ
is Cohen-Macaulay.

Proof. In view of Serre-Grothendieck criterion (see [Mu], for example)


for Cohen-Macaulayness of X̂ at the vertex in terms of the vanishing of
H i (X, L), it is enough to show that

(∗) H i (X(τ ), L(−π)) = 0, 0 ≤ i < dim X(τ ), π > 0

(since H i (X(τ ), L(π)) = 0, i > 0, π ≥ 0 (by the main results)). Let


us fix a simple root α so that τ > τ sα . (Note that X(τ ) is stable for the
P1 fibration G/B → G/Pα , where Pα is the rank 1 parabolic associated
to α.). Let X(τ ) be the projection of X(τ ) under G/B → G/P where P
is the maximal parabolic subgroup associated to α, (i.e., P = Pα̂ ) and
let H(τ ) = X(τ ) ∩ (p(τ ) = 0). Then H(τ ) = ∪w∈Bτ X(w) (here Bτ is as
in Theorem 4). Hence we obtain the exact sequence

(†) 0 → L−1 → OX → OH → 0

where X = X(τ ) and H = H(τ ). We first claim

(∗∗) H i (H, L(−π)) = 0, 0 ≤ i < dim X1 .

In view of Theorem 4, H admits a nice indexing X1 , . . . , Xr of its


components. Hence using induction on r and on dim X(τ ), we may
r−1
assume that (∗) holds for Y := ∪i=1 Xi , Z := Xi and Y ∩ Z (note that
Y ∩ Z admits a nice indexing (cf. Definition 5)). The required claim
(∗∗) now follows by considering the exact sequence
144 Appendix A

0 → OH → OY ⊕ OZ → OY ∩Z → 0
and the cohomology exact sequence

→ H i−1 (Y ∩Z, L(−π)) → H i (H, L(−π)) → H i (Y, L(−π)⊕H i (Z, L(−π)) →

Now tensoring (†) by L(−π) and writing down the cohomology exact
sequence, we have

→ H i−1 (H, L(−π) → H i (X, L(−π ′ )) → H i (X, L(−π)) → H i (H, L(−π)) →

where π ′ = (a1 + 1, a2 , . . . , aℓ ) (here we take P = Pα̂1 , α1 , α2 , . . . , αℓ


being simple roots.) Now in view of (∗∗), we obtain

0 → H i (X, L(−π ′ )) → H i (X, L(−π)), 1 ≤ i < dim X.

When a1 = 1, H i (X, L(−π))= 0, for all i (since X is stable for the


P1 -fibration G/B → G/Pα ). Hence we obtain

H i (X, L(−π)) = 0, 1 ≤ i ≤ dim X − 1, π>0 (1)


is exact. When i = 0, we obtain

H 0 (X, L(−π ′ )) ≈ H 0 (X, L(−π))


(since H 0 (H, L(−π)) = 0).
Again, we have H 0 (X, L(−π)) = 0, when a1 = 1. Hence we obtain

H 0 (X, L(−π)) = 0, π > 0. (2)

From (1) and (2) we obtain

H i (X, L(−π)) = 0, 0 ≤ i < dim X, π > 0.


This completes the proof of Theorem 5.

Theorem 6 Let τ ∈ W. Then the ring R(τ ) = ⊕π≥0 H 0 (X, L(π)) is


Cohen-Macaulay.
Cohen-Macaulay Properties 145

Proof. (By induction on dim X(τ ) ). Let α be a simple root such that
τ > τ sα . Let P = Pα̂ , the maximal parabolic subgroups corresponding
to α. Let X(τ ) be the projection of X(τ ) under G/B → G/P and let
H(τ ) = X(τ ) ∩ {p(τ ) = 0}. We now claim that

(∗) I(H(τ )) = p(τ )R(τ ).


Let Q = ∩P ′ , {P ′ , maximal parabolic, P ′ 6= P }. Let f ∈ R(τ )a
where a = (a1 , . . . , aℓ ) and let f ≡ 0 on H(τ ). If a1 = 0 then f ≡ 0
on X(τ ), since “f comes from G/Q” (to be precise, if a1 = 0, then
H 0 (X(τ ), L(π)) ≈ H 0 (X(w), L(π)) with w = τ sα , since X(w) and X(τ )
have the same projections under G/B → G/Q; further, f ≡ 0 on X(w),
since X(w) is an irreducible component on H(τ )). On the other hand
if a1 6= 0, one can easily see that f ∈ p(τ )R(τ ) : for, we can write
f = F1 + F2 , where each Fi is a sum of standard monomials on X(τ ) of
type a and each monomial in F1 involves p(τ ) while each monomial in
F2 does not involves p(τ ). Hence we obtain F1 ∈ I(H(τ )) and therefore,
F2 (= f − F1 ) ∈ I(H(τ )). On the other hand, each monomials in F2
remains standard on H(τ ) (since none of them involves p(τ ). Hence
the linear independence of standard monomials implies F2 ≡ 0. Thus
f ∈ p(τ )R(τ ) as claimed.
Next we claim that the multigraded ring R(τ )/p(τ )R(τ ) is Cohen-
Macaulay. More generally, we claim the following: let Z be an union of
Schubert varieties and further let Z be of pure of dimension < dim X(τ ).
If Z admits a nice indexing (cf. Definition 5), then R(Z) is Cohen-
Macaulay. We prove this by induction on the number of components of
Z and on dim Z. Let X1 , . . . , Xr be a “nice” indexing of the components
r−1
of Z. Let Z = X ∪ Y, where X = ∪i=1 Xi and Y = Xr . Denoting the
ideals of X, Y, Z in G/B by I(X), I(Y ), I(Z) respectively and denoting
by R the multigraded ring of G/B, the required result that R/I(Z)
is Cohen-Macaulay follows from the following well-known lemma from
Commutative algebra.

Lemma Let R = ⊕n≥0 Rn be a Noetherian graded ring with R0 = k; let


I, J be two homogeneous ideals of R such that dim R/I = dim R/J =
1 + dim R/(I + J). If R/I, R/J, and R/I + J are Cohen-Macaulay, then
R/I ∩ J is Cohen-Macaulay.
Now taking I = I(X), J = I(Y ), we have: R/I is Cohen-Macaulay
of dimension = dim Z (by the induction hypothesis on r); R/J is Cohen-
Macaulay of dimension = dim Z (by induction hypothesis on dim X(τ ));
146 Appendix A

R/I + J, being the ring of X ∩ Y, is Cohen-Macaulay (by the induction


hypothesis on dim Z, for any union V of Schubert varieties which is pure
of dimension < dim Z and which admits a nice indexing, the multigraded
ring of V is Cohen-Macaulay). Hence from the above lemma we obtain
that R/I ∩ J, i.e., the multigraded ring of Z is Cohen-Macaulay. On
the other hand, the ring of Z is precisely R(τ )/p(τ )R(τ ) and p(τ ) being
a nonzero divisor in R(τ ), we obtain, therefore, that R(τ ) is Cohen-
Macaulay.
This completes the proof of Theorem 6.

Part II Cohen-Macaulay Properties for Schubert varieties in the Flag


variety (by “degeneration” method).

In this section, we use the techniques of [D-E-P] to prove the arith-


metic Cohen-Macaulayness for Schubert varieties in the flag variety.

A. Schubert varieties in the Grassmannian.

Let α1 , α2 , . . . , αn−1 be simple roots of SLn (k) and let P = Pα̂α so


that Gd,n = SLn (k)/P.
Let X = X(τ ) be a Schubert variety in Gd,n . Let R(τ ) be the homo-
geneous coordinate ring of X for the Plücker embedding.

X(τ ) → Gd,n → P(Λd V )

(where V = kn ). Let H(τ ) = {φ ∈ I(d, n) | φ ≤ τ } and let k{H(τ )} be


the Stanley-Reisner ring associated to the partially ordered set H(τ ),
namely,

k{H(τ )} = k[xα ]α∈H(τ ) /I

where I is the ideal generated by {xα xβ | α, β non-comparable}. We


shall now show that R(τ ) can be obtained from k{H(τ )} by successive
flat deformations and as a consequence we shall show that R(τ ) is Cohen-
Macaulay by showing that k{H(τ )} is Cohen-Macaulay.

Proposition 1 Let F be a monomial,P say F = pτ1 pτ2 . . . pτm ∈ R(τ )


such that F is not standard. Let F = a(λ) pλ1 pλ2 . . . pλm be the ex-
(λ)
pression for F as a sum of standard monomials. Then for each term
Cohen-Macaulay Properties 147

pλ1 pλ2 . . . pλm on the R.H.S., we have: (λ1 , λ2 , . . . , λm ) is lexicographi-


cally ≥ (σ(τ1 ), σ(τ2 ), . . . , σ(τm )), σ ∈ Sm , (the permutation group on m
symbols).

Proof. We first prove the results for a monomial of degree 2, say


X
pτ1 pτ2 = a(λ) pλ1 pλ2 .
(λ)

Now restricting the above relation to X(λ1 ), R.H.S. is a nonzero sum


of standard monomials on X(λ1 ). Hence pτ1 pτ2 | X(λ1 ) 6≡ 0, from which
it follows that λ1 ≥ τ1 and τ2 . In fact, λ1 > τ1 and τ2 (this follows
from weight considerations pτ1 pτ2 is a weight vector of weight equal to
−(τ1 (ω)+τ2 (ω)) = −(λ1 (ω)+λ2 (ω)), where ω is the fundamental weight
associated to P ; hence λ1 = τ1 would imply that λ2 = τ2 and this in
turn implies that pτ1 pτ2 is standard).
Now let F = pτ1 . . . pτ2 be of degree m. We associate formal weight
n(F ) to F as follows. Fix N, a sufficiently large integer. Set N (F ) =
(n(τ1 ), . . . , n(τm )) in the N -adic representation (where n(τ1 ) = codimen-
sion of X(τi ) in Gd,n ) i.e., n(F ) = N m−1 n(τ1 )+N m−2 n(τ2 )+· · ·+n(τm ).
Since F is not standard, there exist integers P r, s ≤ m such that pτr pτs
is not standard. Hence writing pτr pτs = b(β) Pβ1 Pβ2 as a sum of stan-
(β)
P
dard monomials, we obtain F = Fi , where n(Fi ) < n(F ) (since each
i
β1 > τr and
P τs ). Now writing each Fi as a sum of standard monomials,
say Fi = cγ pγ1 . . . pγm , we have (by induction on n(F )) that for each
(γ)
(γ), (γ1 , . . . , γm ) is lexicographically > m-tuple of the τj ’s present in
any Fi taken in any order. But now, the τ ’s present in Fi are the same
as those in F except for {τr , τs } which have been replaced by {β1 , β2 }
where for each (β) = {β1 , β2 }, we have : {β1 , β2 } is lexicographically
greater than {τr , τs } and {τs , τr }. From this the required result follows
(we should observe that when n(F ) is least, we have F = pm−2 τ pδ1 pδ2
where X(δ P i ) are of codimension 1 in X(τ ), i = 1, 2. Now we have
pδ1 pδ2 = c(λ) pλ1 pλ2 , a sum of standard monomials. The fact that
(λ)
λ1 > δ1 implies that λ1 is in fact = τ for every (λ) = (λ1 , λ2 ) and the
result is obviously true in this case.
This completes the proof of Proposition 1.
148 Appendix A

Reduction of R(τ ) to k{H(τ )} by successive flat deformations

Let J(R(τ )) = {α ∈ H(τ ) | pα appears on the R.H.S. of some


straightening relation } (here by a straightening relation we mean a re-
lation expressing a nonstandard monomial as a sum of standard mono-
mials).

Proposition 2 Let α be a maximal element in J(τ ) and let Iα be the


ideal generated by pα in R(τ ). Then for k ≥ 1, Iαk has a basis consisting
of standard monomials involving pkα .

Proof. Let F be a monomial of degree m belonging to Iαk , say F =


pkα pτ1 . . . pτr , where r + k = m. If F is not already standard, then writing
F as a sum of standard monomials, we have
X
F = phα pτ1 . . . pτ2 = a(β) pβ1 . . . pβm .
(β)

Now by Proposition 1 we have (β1 . . . β2 ) is lexicographically


≥ (α, α, . . . , τ1 , . . . , τr ) . Now the maximality assumption on α implies
| {z }
k times
that β1 = . . . = βk = α. From this the result follows immediately.

Proposition 3 Let R = R[It−1 , t], where R = R(τ ) and I = Iα . Then


R is a k[t]-algebra with generators {Pα t−1 , Pβ , β 6= α}.

Proof. Obvious.

Proposition 4 Let Rα = gr1 (R) = R/tR. Then the k-algebra Rα


has a set of algebra generators given by {pβ | β ∈ H(τ )} such that
the standard monomials in pβ ’s form a basis for Rα . Further J(Rα ) ⊆
J(R)−{α} and Proposition 1 holds in Rα (Here J(Rα ) = {λ ∈ H(τ ) | pα
occurs on the R.H.S. of some straightening relation in Rα }).

Proof. The proof follows from Propositions 2 and 3.

Remark 5 The algebra Rα is the special fiber (t = 0) of the flat


deformation R over Spec k[t] whose generic fiber (t invertible ) is R.
Hence R would be Cohen-Macaulay if Rα is.
Cohen-Macaulay Properties 149

Theorem 6 There exists a sequence of k-algebras R1 , . . . , Rt (with


R1 = R(τ )) such that

(1) for 1 ≤ s ≤ t − 1, there is a flat deformation Rs whose special fiber


is Rs+1 and general fiber is Rs .

(2) J(Rs+1 ) ⊂ J(Rs ), 1 ≤ s ≤ t − 1.

(3) J(Rt ) = ∅.

The proof follows by repeated application of Proposition 4 (see also


Remark 5).

Remark 7 The ring Rt is nothing but k{H(τ )}.

Theorem 8 The ring k{H(τ )} is Cohen-Macaulay.

Proof. The poset H(τ ) is lexicographically shellable (cf. Part I and also
[B-W1]). Hence the order complex △(H(τ )) is shellable. The shellabil-
ity of △(H(τ )) implies that k{H(τ )} is Cohen-Macaulay (one may use
Mayer-Vietoris sequence or [R]).

Theorem 9 The ring R(τ ) is Cohen-Macaulay.

Proof. This follows from Theorem 6, Remark 7 and Theorem 8.

B. Schubert varieties in the Flag variety SLn /B.


Let Q be a parabolic subgroup, say Q = ∩ri=1 Pi , where Pi are some
maximal parabolic subgroups, 1 ≤ i ≤ r, and let τ ∈ W Q and let
R(τ ) = ⊕L≥0 H 0 (X(τ ), L), where L is a positive line bundle on G/Q. In
this section we want to show that R(τ ) is Cohen-Macaulay by using a
similar kind of deformation argument as in (A).
Let R = R(τ ) and let J(R) = {α ∈ ∪1≤i≤r W/Wi | pα occurs on the
R.H.S. of a straightening relation of a nonstandard nonzero monomial
on X(τ )}.

Proposition 10 Let λ be a non-standard Young diagram on X(τ )


of type m = (m1 , . . . , mℓ ), ℓ = nP− 1, and let Fλ be the associated
element in H 0 (X(τ ), L). Let Fλ = aα Fα be the straightening relation
for Fα (so that R.H.S. is a sum of standard monomials on X(τ )). If
150 Appendix A

S = Sm1 ×. . .×Smℓ , where Smi is the permutation group on mi symbols,


then for each α, the Young diagram α is lexicographically ≥ λσ , for any
σ ∈ S.

Proof. The proof is similar to that of Proposition 1 (for details, see


[H-L].)

Proposition 11 Let α ∈ J(R) be such that if (β, α) is a standard


Young diagram, then β 6∈ J(R) (note that such a “maximal” α exists).
Let I be the ideal generated by pα . Then, for all k ≥ 1, I k has a basis
consisting of standard monomials on X(τ ) involving pkα .

Proof. The proof is similar to that of Proposition 2 (one uses Proposi-


tion 10; see [H-L] for details).

Proposition 12 Let R = R[It, t−1 ], R and I being as above. The


graded algebra grI (R) (= R/tR) has a set of algebra generators {pβ |
β ∈ W/Wi , 1 ≤ i ≤ ℓ and pβ |X(τ ) 6= 0} such that the standard (Young)
monomials on X(τ ) form a basis for grI R. Further

J(grI R) ⊆ J(R) − {α}.

Proof. The proof follows from Proposition 11.

Remark 13 The algebra Rα (= grI R) is the special fiber of the flat


deformation R over Spec k[t] whose special fiber is R.

Theorem 14 There exists a sequence of k-algebras R1 , . . . , Rt (with


R1 = R(τ )) such that

(1) for 1 < s < t − 1, there is a flat deformation Rs whose special fiber
is Rs+1 and general fiber is Rs

(2) J(Ss+1 ) 6⊆ J(Rs ), 1 < s < t − 1

(3) J(Rt ) = ∅.

Proof. The proof follows by repeated application of Proposition 12.


Cohen-Macaulay Properties 151

Remark 15 In view of Theorem 14, R(τ ) would be Cohen-Macaulay,


if Rt is. Unlike the case of the Grassmannian (cf. (A)), where Rt turned
out to be k{H(τ )}, we cannot realize Rt as the Stanley-Reisner ring of
any partially ordered set. Nevertheless, we prove the Cohen-Macaulay
property for Rt using the idea behind the “principal radical systems”
as developed in [E-H]. Hereafter, we shall denote Rt (cf. Theorem 14)
by Rdef (τ ) (meaning “R(τ ) deformed”) and for X(τ ) = G/Q we denote
Rdef (w0 ) by just R. It is clear that I(τ ), the ideal of Rdef (τ ) in R,
is generated by J(τ ), the set of all square-free standard monomials on
G/Q which are not standard X(τ ). Further, I(τ ) has a basis consisting
of monomials standard on G/B but not standard on X(τ ).

Proposition 16 Let H = { all square-free standard monomials on


G/Q} together with the partial order “m ≥ m′ if and only if m′ contains
m as a submonomial.” Then Rdef (τ ) is a Hodge algebra (cf. [D-E-P])
on H.

Proof. In the notations of P [D-E-P], we see that Rdef (τ ) is a Hodge


algebra over H governed by = I(τ ). In fact it is a “discrete algebra”
(i.e.,
P the right hand sides of the straightening relations for thePgenerators
of are zero (note that J(τ ) gives the set of generators of )).

Definition 17 (cf. [E]) A subset J ⊂ H is said to be an ideal of H if


α ∈ J and β ≤ α imply β ∈ J.

Notation If J is an ideal of H, we denote by J the ideal of R generated


by the monomials in J.

Remark 18 Note that J(τ ) is an ideal of H and that J(τ ) = I(τ )


(recall that J(τ ) = { square-free standard monomials on G/Q which are
not standard on X(τ )}.

Lemma 19 Let A be a Hodge algebra over H. If I and J are ideals of


H, then

(1) I + J = I ∪ J

(2) I ∩ J = I ∩ J

The proof is quite easy (for a proof, one may also refer to [E]).
152 Appendix A

Lemma 20 Let w1 , . . . , wk ∈ W Q . Then

[I(w1 ) ∩ . . . ∩ I(wk−1 )]+I(wk ) = [I(w1 ) + I(wk )]∩. . .∩[I(wk−1 ) + I(wk )] .

Proof. In view of Lemma 19, it is enough to show that

[J(w1 ) ∩ . . . ∩ J(wk−1 )]∪J(wk ) = [J(w1 ) ∪ J(wk )]∩. . .∩[J(wk−1 ) ∪ J(wk )]

which is simply De Morgan’s law.

Definition 21 For a union of Schubert varieties, say Z = ∪ri=1 X(τi ),


we define Rdef (Z) = R/I(Z), where I(Z) = ∩ri=1 I(τi ); we have I(Z) =
J(Z), where J(Z) = ∩ri=1 J(τi ). Further, I(Z) has a basis consisting of
monomials standard on G/Q but not on X(τi ), 1 ≤ i ≤ r.

Lemma 22 Let Z = ∩ri=1 X(τi ) be an intersection of Schubert varieties.


P
r
Then I(Z) = I(τi ).
i=1

Proof. Let Z = ∪sj=1 X(wj ) so that

I(Z) = ∩sj=1 I(wj ).

Pr
Now each I(τi ) ⊂ I(Z) (obviously), 1 ≤ i ≤ r. Hence I(τi ) ⊂
P i=1
I(Z). On the other hand, let f = at ft ∈ I(Z), where each ft is a
t
maximal standard on G/B but not on X(wj ), 1 ≤ j ≤ s (since I(Z) =
∩I(wj )).

Claim ft cannot be standard on all X(τi ), 1 ≤ i ≤ r. For, if ft were


standard on X(τi ), 1 ≤ i ≤ r, then let us consider the Young diagram
λ associated to ft . Let θ = (θ1 , θ2 , . . .) be the unique minimal defining
chain (see (2.2.2), (2.4.4)) for λ. Then θ1 ≤ τi for all i (since ft is
standard on X(τi ), 1 ≤ i ≤ r). Hence θ1 ≤ wj , for some j, 1 ≤ j ≤ s.
Hence ft is standard on X(wj ), a contradiction (since ft is not standard
on X(wj ), 1 ≤ j ≤ s). This proves the claim that ft cannot be standard
Pr
on all X(τi ), 1 ≤ i ≤ r. Thus ft ∈ I(τi ). This completes the proof of
i=1
Lemma 22.
Cohen-Macaulay Properties 153

Proposition 23 Let τ ∈ W Q , Q = ∩ri=1 Pi and let X(τ ) be the projec-


π
tion of X(τ ) under G/Q → G/P1 . Let Bτ = { Schubert divisors X(w)
in X(τ ) | π(X(w)) ⊂ X(τ )}. Then in Rdef (τ ), we have: (pτ ), the ideal
generated by pτ , is the ideal Kτ ∩w∈Bτ I(w), where Kτ is the ideal in
Rdef (τ ) generated by {pα | α ∈ W P1 , α ≤ τ }.

Proof. It is clear that (pτ ) ⊆ Kτ ∩w∈Bτ I(w), since (pτ ) ⊆ Kτ (obvi-


Ppτ vanishes on X(w), w P∈1 Bτ . Conversely, let F ∈ Kτ ∩ I(w),
ously) and
say F = ci pαi Fi , where αi ∈ W , αi ≤ τ and pαi Fi is standard on
i
X(τ ). Now we claim that :
(∗) If a monomial pα F is standard on X(τ ), α ∈ W P1 and α < τ then
pα F is standard on X(w) for some w ∈ Bτ .
Observe that (∗) obviously implies the required result. The claim (∗)
immediately follows from the following lemma.

Lemma 24 Let τ ∈ W Q . If X(θ) is a Schubert subvariety of X(τ )


with π(X(θ)) ⊂ π(X(τ )) where π : G/Q → G/P1 , then there exists a
Schubert divisor X(w), w ∈ Bτ such that X(θ) ⊆ X(w). (For a proof of
this lemma refer to [H-L].)
This completes the proof of Proposition 23.

Theorem 25 Let τ ∈ W Q . Then the ring R(τ ) = ⊕L H 0 (X(τ ), L),


where L is a positive line bundle on G/Q, is Cohen-Macaulay.

Proof. As already remarked, it suffices to show that Rdef (τ ) is Cohen-


Macaulay. We observe that pτ is a nonzero divisor in Rdef (τ ) (since
if F is any standard monomial on X(w), then pτ F is also a standard
monomial on X(τ )). Hence to prove the theorem, it is enough to show
that Rdef (τ )/(pτ ) is Cohen-Macaulay.
Let Q = ∩rt=1 Pit , 1 ≤ i1 < i2 < · · · < ir ≤ n (where for 1 ≤ d ≤ n,
Pd denotes the maximal parabolic subgroup given by
( !)
d
Pd = A ∈ SL(n) | A = M1 M2 .
n−d 0 M3
By Proposition 23, we have

(p(τ )) = I ∩ J
154 Appendix A

where I = ∩w∈Bτ I(w), J = Kτ , the ideal in Rdef (τ ) generated by {pα |


α ∈ W Pi1 .α ≤ τ } (Note that here X(τ ) is the projection of X(τ ) under
G/Q → G/Pi1 ). We now claim that

(†) Let w1 , . . . , ws ∈ Bτ and let Z = ∪si=1 X(wi ). Then Rdef (τ )/I(Z)


is Cohen-Macaulay of dimension equal to dim Rdef (τ ) − 1.

To prove the claim (†), we first observe that Z admits a nice index-
ing, say Z = ∪si=1 Xi (this follows essentially from the fact that such a
result holds for a similar union in the Grassmannian. For details see
[H-L]). Now we use induction on s and on dim X(τ ). When s = 1,
Rdef (Z) = Rdef (Xi ). Hence (by induction on dim X(τ )) Rdef (X1 ) is
Cohen-Macaulay of dimension = dim X1 + r (where, recall, r is given
by Q = ∩rt=1 Pit ). Now we may assume (by induction on dim X(τ ))
that the result is true for a union ∪λ X(λ), where the union runs over
some of the λ’s belonging to Bw , dim X(w) < dim X(τ ). Hence writing
s−1
Z = Y ∪X, where Y = ∪i=1 Xi and X = Xs , we have I(Z) = I(Y )∩I(X)
and I(Y ) + I(X) = I(Z ) (cf. Lemma 22) where Z ′ = Y ∩ X further

Rdef (τ )/I(Y ), Rdef (τ )/I(X), and Rdef (τ )/I(Z ′ ) are Cohen-Macaulay


of dimension d − 1, d − 1, d − 2, respectively (where d = dim Rdef (τ ) =
dim X(τ ) + r). Hence by Lemma 7 of Part I, we obtain that Rdef (Z) is
Cohen-Macaulay of dimension d − 1.
This completes the proof of claim (†) above.

Return to the proof of Theorem 25 If X(τ ) is saturated for the


P1 -fibration G/Q → G/Q′ (equivalently τ > τ sα , where α = αi1 , Q′ =
∩rt=2 Pit ), then we have that ∩w∈Bτ I(w) is in fact = P (pτ ) (for (pτ ) ⊆
∩I(w) (obviously) and if f ∈ ∩w∈Bτ I(w), writing f = ci Fi , a sum of
standard monomials on X(τ ), we notice that if f ∈ H 0 (X(τ ), La) where
a = (a1 , . . . , ar ), then a1 6= 0; for a1 = 0 would imply that f ∈ I(w),
since w = τ sα (observe that X(w) and X(τ ) have the same projections
on G/P ′ for every maximal parabolic P ′ = Pit , 2 ≤ t ≤ r). Now a1 6= 0
and f ∈ ∩w∈Bτ I(w) imply that f ∈ (pτ ) (cf. claim (∗) in the proof
of Proposition 23). Thus, in this case Rdef (τ )/(pτ ) = Rdef (Z), where
Z = ∩w∈Bτ X(w), and hence it is Cohen-Macaulay of dimension d − 1
(where d = dim Rdef (τ ) = dim X(τ ) + r).
If X(τ ) is not saturated for the P1 -fibration G/Q → G/Q′ , Q′ =

∩rt=2 Pit , then we obtain that τ ∈ W Q . In this case, we see easily that

each w belonging to Bτ in fact belongs to W Q , and further, w ∈ Bτ′
Cohen-Macaulay Properties 155

(where Bτ′ is defined in a similar way as Bτ , X(τ ) now being con-


sidered as a Schubert variety in G/Q′ ; for details see [H-L]). Hence
def
Rdef (τ )/(p(τ )) = R′ (τ )/∩ I ′ (w) (note that (p(τ )) = J∩∩
w∈Bτ I(w) =
w∈Bτ
∩w∈Bτ′ I ′ (w)) and hence by claim (†) above (and by induction on r)
def
R′ (τ )/∩w∈Bτ , I ′ (w) is Cohen-Macaulay of dimension = dim X(τ ) +

r − 1 (note that τ ∈ W Q implies dim X(τ ) (in G/Q′ ) = dim X(τ ) (in
G/Q)).
Thus in either case, Rdef (τ )/p(τ ), is Cohen-Macaulay of dimension
d−1 and hence Rdef (τ ) is Cohen-Macaulay of dimension d = dim X(τ )+
r (since pτ is not a zero divisor in Rdef (τ ), as remarked earlier). This
completes the proof of Theorem 25.

Remark 26 It should be remarked that degeneration techniques do not


yield a proof of the Cohen-Macaulayness of the ring S(τ ) = ⊕n≥0 H 0 (X(τ ), Ln ),
where L is a positive line bundle on G/Q (compare with Theorem 5 of
Part I).
Appendix B

Normality of Schubert varieties

See [S3]. Let G be a semisimple, simply connected algebraic group


defined over an algebraically closed field. Fix a maximal torus T and a
Borel subgroup B of G containing T. Denote by W the Weyl group of
G. Let φ be the root system of G. Let S be the set of simple roots and
let φ+ be the set of positive roots of φ with respect to the choice of B.

(2.1) Theorem Every Schubert variety X(ω), ω ∈ W, in G/B is


normal.

Proof.
A. The proof is by decreasing induction on the dimension of X(ω). If
ω = ω0 , the unique element of the Weyl group such that ω0 (φ+ ) =
φ− , then X(ω0 ) = G/B; hence X(ω0 ) is smooth and in particular,
normal (see Chapter 4 (4.1.3) and (4.2.4)).

B. Now assume that X(ω) 6= G/B. If α is a simple root such that


τ = sα ω and τ > ω, then by induction hypothesis, X(τ ) is normal.
Let P (α) be the minimal parabolic subgroup of G generated by
B and G−α . Denote by Z(ω, α) the locally trivial fibration π :
P (α) ×B X(ω) → P1 (∼ = P (α)/B) with fibre X(ω) and let ψ :
Z(ω, α) → X(τ ) be the canonical morphism. Recall that ψ is
birational (Chapter 4, 4.4.2).

C. Let η : X̃(ω) → X(ω) be the normalization map and denote by


Z̃(ω, α) the fibration π̃ : P (α) ×B X̃(ω) → P1 . The canonical
map ψ̃ : Z̃(ω, α) → X(τ ) is birational. We have the following
commutative diagram:

D. Let M be an ample line bundle on G/B. Recall that M is ho-


mogeneous, i.e. G acts on M. Denote by M(ω) the restriction
of M to X(ω) and denote by M̃(ω) the pull back of M(ω) to
X̃(ω). Note that M̃(ω) is also ample. As in Chapter 2, (7.2), C,
we see that: ψ ∗ (M(τ )) ≃ (M(ω))# and ψ̃ ∗ (M(τ )) ≃ (M̃(ω))# ,

© Springer Science+Business Media Singapore 2016 and Hindustan Book Agency 2015 157
C.S Seshadri, Introduction to the Theory of Standard Monomials,
Texts and Readings in Mathematics 46, DOI 10.1007/978-981-10-1813-8
158 Appendix B

ν ψ
a) Z̃(ω, α) Z(ω, α) X(τ )

ψ̃

ν
b) Z̃(ω, α) Z(ω, α)

π
π̃
P1

where (M(ω))# (resp. (M̃(ω))# ) denotes the associated line bun-


dle P (α) ×B M(ω) on Z(ω, α) (resp. P (α) ×B M̃(ω) on Ž(ω, α)).

E. Let L be the line bundle on G/B corresponding to ζ = 21 sum


of the positive roots (same construction as in Chapter 3, (3.1)).
We use the notation of Serre, i.e. we denote F ⊗ Lm by F (m)
for any coherent sheaf F on a Schubert variety. Now denote
H 0 (X(ω), OX(ω) (m)) by Vm and denote H 0 (X̃(ω), OX̃(ω) (m)) by
Ṽm . Let Vm and Ṽm be the vector bundles on P1 associated to the
principal B-fibration P (α) → P1 for the B-modules Vm and Ṽm
respectively. Since ψ and ψ̃ are birational and X(τ ) is normal,
we see that ψ∗ (OZ(ω,α) ) (resp. ψ̃∗ (OZ̃(ω,α) ) ≃ OX(τ ) and hence:
H 0 (Z(ω, α), ψ ∗ (L(τ )m )) ≃ H 0 (Z̃(ω, α), ψ̃ ∗ (L(τ )m ) ≃ H 0 (X(τ ),
OX(τ ) (m)). We observe that P (α) acts on π∗ (ψ ∗ (L(τ )m )) (resp.
π̃∗ (ψ̃ ∗ (L(τ )m ))) consistent with its action on P1 , from which we
conclude easily that these sheaves are locally free. On the other
hand, we have

H 0 (P1 , π∗ (ψ ∗ (L(τ )m ))) ≃ H 0 (Z(ω, α), ψ ∗ (L(τ )m ))


(resp. H 0 (P1 , π̃∗ (ψ̃ ∗ (L(τ )m ))) ≃ H 0 (Z̃(ω, α), ψ̃ ∗ (L(τ )m )).

By similar arguments as in D we see that π∗ (ψ ∗ (L(τ )m )) ≃ Vm and


π̃∗ (ψ̃ ∗ (L(τ )m )) ≃ Vm and hence H 0 (P1 , Ṽm ) ≃ H 0 (X(τ ), OX(τ ) (m))
for m ≥ 0.
Normality of Schubert varieties 159

F. Let C denote the coherent sheaf of OX(ω) -modules OX̃(ω) /OX(ω) .


If we tensor the exact sequence 0 → OX(ω) → OX̃(ω) → C → 0 by
Lm (ω), we obtain the exact sequence:

(∗) 0 → OX(ω) (m) → OX̃(ω) (m) → C(m) → 0.

(Recall that for any line bundle N on X(ω) we have: η∗ (η ∗ (N )) ≃


OX̃(ω) ⊗ N , where η : X̃(ω) → X(ω) denotes the normaliza-
tion map. Since η is affine, we know that H 0 (X(ω), OX(ω) (m) ≃
H 0 (X̃(ω), η ∗ (L(ω)m )) ≃ Ṽm . By Serre’s theorem, we can choose
an m0 ≥ 0 such that ∀ m ≥ m0 , H 1 (X(ω), OX(ω) (m) = 0 and
H 0 (X(τ ), OX(τ ) (m))
→ H 0 (X(ω), OX(ω) (m)) → 0 is exact. The sequence of global sec-
tions corresponding to the exact sequence (∗) is hence exact for
m ≥ m0 :

0 → Vm → Ṽm → H 0 (X(ω), C(m)) → 0, m ≥ m0 .

Denote H 0 (X(ω), C(m)) by Wm . Since the normalization map η :


X̃(ω) → X(ω) is a B-morphism, the B-action on the coherent
OX(ω) -module C(ω) is consistent with the B-action on X(ω). De-
note by wm the associated vector bundle on P1 . We get the exact
sequence of vector bundles on P1

0 → Vm → Ṽm → Wm → 0, m ≥ m0 .

Since the map H 0 (X(τ ), OX(τ ) (m)) → H 0 (X(ω), OX(ω) (m)) is


surjective for m ≥ m0 and H 0 (X(τ ), OX(τ ) (m))) is P (α)-module,
we obtain, as in the proof of (3.2), Chapter 3, that H 1 (P1 , Vm ) = 0
for m ≥ m0 . We know that H 0 (P1 , Vm ) ≃ H 0 (P1 , Ṽm ) (see E). So
we conclude from the exact sequence

0 → H 0 (P1 , Vm ) → H 0 (P1 , Ṽm ) → H 0 (P1 , Wm ) → 0

that

H 0 (P1 , Wm ) = 0 for m ≥ m0 .

G. Now note that C = 6 (0) ⇔ X(ω) is not normal. To prove that


C = (0), we will show that if C =
6 (0), then there exists a simple
160 Appendix B

root α such that sα ω > ω and if W = H 0 (X(ω), C(m)), then


H 0 (P1 , Wm ) 6= 0 for m ≫ 0, where P1 = P (α)/B and Wm is the
associated vector bundle. This is a contradiction to our result that
H 0 (P1 , Wm ) = 0 for m ≥ m0 (see F) and hence we obtain that
C = (0) and X(ω) is normal. We prove the statement above in a
more general context.

H. If Q is the parabolic subgroup of G generated by the minimal


parabolic subgroup P (α) such that α ∈ S and P (α) leaves X(ω)
stable, then Q leaves X(ω) stable and C(m), m ≥ 0 is a Q−OX(ω) -
module. The vanishing of C and hence the required normality of
X(ω) is a consequence of the following:

(2.2) Lemma. Let F be a coherent Q − OX(ω) -module and let Wm


denote H 0 (X(ω), F(m)). If F =6 (0), then there exists a simple root α
0 1
such that sα ω > ω and H (P , Wm ) 6= 0 for m ≫ 0, where Wm denotes
the associated vector bundle to the B-module Wm on P1 ≃ P (α)/B.

Proof of the Lemma.

A. Let J denote the annihilator Ann F of F in OX(ω) . Then OY =


OX(ω) /J is a Q-sheaf and Y is a Q-scheme. The reduced scheme
Yred is also a Q-scheme. Note that F is the canonical extension
of a QY sheaf of modules which we denote also by F. We express
this by saying that F “lives on Y ”. Let I be the sheaf of ideals
on OY such that OYred = OY /I. Note that I and all powers I n ,
n ≥ 0, of I are Q − OY -modules. There exists an integer n ≥ 1
such that I n .F = (0) and I n−1 F =
6 (0). If G = I n−1 F and Wm
′ =

H 0 (X(ω), G(m)), then we have the exact sequence:


0 → Wm = H 0 (X(ω), G(m)) → Wm = H 0 (X(ω), G(m)),

which gives the exact sequence of vector bundles on P1 : 0 →


′ → W 0 1 ′
Wm m and hence the exact sequence: 0 → H (P , Wm ) →
0 1 0 ′
H (P , Wm ). So it suffices to show that H (P, Wm ) 6= 0.
B. Since I.G = 0, we know that G is a sheaf of OYred -modules. Let
Y1 , . . . , Yr be the irreducible components of Yred . Since Q is con-
nected, the irreducible components are Q-stable. We can assume
Normality of Schubert varieties 161

that the support of G is not contained in any union of a proper


subset of the irreducible components of Yred , for otherwise we can
replace Yred by the union of the irreducible components. I1 be
the sheaf of ideals in OYred defining the closed subscheme Y1 of
Yred and let I2 be the sheaf of ideals defining the closed subscheme
Y2 ∪ . . . ∪ Yr of Yred . Denote by G1 the Q − OYred subsheaf I2 .G of
G (the sections of G1 are those of G which vanish on Y2 ∪ . . . ∪ Yr ).
Since I1 annihilates G1 , the support of G1 is contained in Y1 , and
hence G1 is a OY1 = OYred /I1 module. Note that G1 6= (0), for oth-
erwise G1 = I2 G = 0 and hence G would “live” on Y2 ∪ . . . ∪ Yr , a
contradiction to our assumption about G made in B. By a similar
argument as in A, it suffices to prove the lemma for G as a sheaf
of OY1 -modules.

C. Let G1′ be the torsion subsheaf of the Q−OY1 -module G1 . Note that
Y1 is irreducible and Q-stable. Hence Y1 is a Q-stable Schubert
subvariety X(θ) of X(ω). The sheaf G1′ is a Q − OY1 submodule
of G1 . It is obvious, that if G1′ 6= (0), then it suffices to prove the
lemma for G1′ . Now observe that the support of G1′ is a properly
contained closed subscheme Y1′ of Y1 , and hence it suffices to prove
the lemma for the restriction of G1′ to Y1′ . By repeating the pro-
cedure described in B, we get a Q − OY2 -module G2 , G2 6= (0),
such that Y2 is a Schubert variety properly contained in Y, and it
suffices to prove the lemma for G2 . If we repeat this procedure by
taking the torsion subsheaves G2′ , G3′ . . . , etc., such that G2′ 6= (0),
G3′ 6= (0), . . . , etc., we get a strictly decreasing chain of Schubert
varieties. Since there are only finitely many such varieties, there
exists an n ≥ 0 such that Gn 6= (0) and Gn is torsion free on Yn .
Hence it is sufficient to prove the lemma when F is a coherent
sheaf which “lives” on a Schubert variety X(θ) ⊆ X(ω) such that
F considered as a sheaf on X(θ) is torsion free.

D. Since F is a Q − OX(θ) -module we know that X(θ) is Q-stable. Let


α be a simple root such that ρ = sα θ > 0. Note that sα ω > ω, for
if P (α) leaves X(ω)-stable, then P (α) ⊂ Q and we get sα ω > ω,
since X(θ) is Q-stable. By Borel fixed point theorem, ∃ m0 ∈ N
and a non-zero section s ∈ H 0 (X(θ), F(m0 )) such that the line
through s is B-fixed. Now multiplication by s induces an inclusion
j : OX(θ) ֒→ F(m0 ). Denote the image of OX(θ) in F(m0 ) by G.
f 7→ f.s The submodule G of F(m0 ) is a B − OX(θ) -submodule.
162 Appendix B

Note that G need not be B-isomorphic to OX(θ) . The action of B


on G differs from that on OX(θ) by the character X of B, which
defines the action of B on the line through s. It suffices to prove the
lemma for G, i.e., if Wm ′ = H 0 (X(θ), G(m))(≃ H 0 (X(ω), G(m)),

then H 0 (P1 , Wm
′ ) 6= 0 for m ≫ 0. The action of B on G(m) (for

all m) differs by the same character X from the action of B on


OX(ω) (m).

E. Consider the birational canonical morphism ψ : X(ω, α) = P (α)×B


X(θ) → X(ρ) (Chap. 4, (4.4.2)) and the fibration π : X(θ, α) →
P1 of fibre type X(θ). Denote by G # (m) the associated line bun-
dle on Z(θ, α) to the B − OX(θ) -module G(m). Since the B-action
on OX(θ) (m) differs from the B-action on G(m) by the character
X, we see that G(m)# = (OX(θ) (m))# ⊗ π ∗ (N ), where N is the
line bundle on P1 ≃ P (α)/B associated to the character X of B.
We see as in (2.1) D that (OX(θ) (m))# = ψ ∗ (OX(ρ) (m)). Hence
we get π∗ (G(m)# ) = π∗ (OX(ρ) (m)) ⊗ N . Denote H 0 (X(θ), L(θ)m )
by Vm and let Vm be the corresponding vector bundle on P1 . We
know that π∗ (OX(ω) (m)# ) ≃ Vm (2.1, E) and hence π∗ (G(m)# ) =
Vm ⊗ N . If Wm denotes H 0 (X(θ), G(m)) (= H 0 (X(ω), G(m)),
then π∗ (G(m)# ) ≃ Wm ≃ Vm ⊗ N . Furthermore, we see that
H 0 (P1 , Wm ) ≃ H 0 (X(θ, α), G(m)# ). Thus to show that H 0 (P1 , Wm )
6= 0 for m ≫ 0, it is sufficient to show that

H 0 (Z(θ, α), ψ # (OX(ρ) )(m)) ⊗OZ(θ,α) π ∗ (N ))) 6= 0

for m ≫ 0. Now

H 0 (Z(θ, ρ), ψ ∗ (OX(θ) )(m) ⊗ π ∗ (N )))


= H 0 (X(ρ), ψ∗ (ψ ∗ (OX(ρ) (m) ⊗ π ∗ (N ))))
= H 0 (X(ρ), OX(ρ) (m) ⊗OX(ρ) ψ∗ (π ∗ (N ))).

Since ψ is birational and N is a line bundle, we see that ψ∗ (π ∗ (N )) 6=


(0). Hence by Serre’s theorem, we have that H 0 (X(ρ), ψ∗ (π ∗ (N ))⊗
OX(ρ) (m)) 6= 0 for m ≫ 0, and hence H 0 (P1 , Wm ) 6= 0 for m ≫ 0.
Q.E.D.

Remark V. Mehta and A. Ramanathan have proved in a recent paper


[M-R] that the sequence H 0 (X(τ ), OX(τ ) (m)) → H 0 (X(ω), OX(ω) (m)) →
Normality of Schubert varieties 163

0 is exact for m ≥ 0. This result and the normality of the Schubert vari-
eties imply the validity of Demazure’s character formula for all dominant
weights in arbitrary characteristic.
Appendix C

Standard Monomial Theory1


V. Lakshmibai2 and C.S. Seshadri

1 Introduction

Let Gr,n denote the Grassmannian of r-dimensional linear subspaces of


an n-dimensional vector space (say over C). We have a canonical imbed-
ding of Gr,n in a projective space, namely the one given by the Plücker
coordinates (called the Plücker imbedding). In the 1940’s Hodge gave a
nice basis of the homogeneous coordinate ring R of Gr,n , as well as its
Schubert subvarieties (cf. [11], [12]). He called them Standard monomi-
als, probably since they were monomials in the Plücker coordinates and
were indexed by “standard Young Tableaux”. Recall that


 Gr,n L
= SL(n)/P, P a maximal parabolic subgroup

 R = m≥0 H 0 (Gr,n , Lm ), L being the hyperplane line bundle for

the Plücker imbedding (à priori we have only Rm = H 0 (Gr,n , Lm ),



 m ≫ 0, Rm being the mth homogeneous component of R,

but this stronger assertion also holds).

Hodge gave also a similar result for the flag variety SL(n)/B, where B is
the Borel subgroup of SL(n), and for certain (but not all) Schubert sub-
varieties of SL(n)/B. The aim of Standard Monomial Theory (written
briefly as SMT) which we have pursued in a series of papers (along with
Musili) has been to generalize the above work of Hodge in the context
of groups which are more general than SL(n).
Recall the Borel-Weil Theorem, namely, when G is a semi-simple,
simply-connected algebraic group (over C), every finite-dimensional ir-
reducible G-module is G-isomorphic to H 0 (G/B, L), where L is a line
bundle on G/B. Hence giving nice bases of finite-dimensional irreducible
G-modules is a part of SMT. Our results on SMT provide a very satis-
factory generalization of the work of Hodge when G is a classical group
(cf. [43], [33], [26], [27], [30]). Such a generalization has also been done
1
Proceedings of the Hyderabad Conference on Algebraic Groups, Published for
the National Board for Higher Mathematics, Manoj Prakashan, First Edition, 1991
May, ISBN No: 81-231-0090-6.
2
partially supported by NSF grant No. DMS-8701043

© Springer Science+Business Media Singapore 2016 and Hindustan Book Agency 2015 165
C.S Seshadri, Introduction to the Theory of Standard Monomials,
Texts and Readings in Mathematics 46, DOI 10.1007/978-981-10-1813-8
166 Appendix C

for the exceptional groups G2 and E6 (cf. [19], [28]). A beginning has
also been done for Kac-Moody groups (cf. [32], [20]). There is also a
general conjecture (due to Lakshmibai, cf. [32] and §4 below) which
seems extremely plausible since it holds in all the cases where SMT has
been established (including the Kac-Moody case).

2 First Basis Theorem

The main references for this section are §5, [27] and §3, [30].
Let G be a semi-simple, simply connected algebraic group over an al-
gebraically closed field k ( more generally we could work with a Cheval-
ley group over Z). We fix a maximal torus T and a Borel subgroup
B, B ⊃ T . Relative to T and B, we talk of roots, weights etc., and we
denote by W (G) or simply W (= N (T )/T, N (T ) normalizer of T ) the
Weyl group. We write ∆ (resp. ∆+ )= system of roots (resp. system of
positive roots), S = set of simple roots {α1 , . . . , αℓ }, ℓ = rank of G. We
denote by ( , ) a W -invariant scalar product on hom(T, Gm ) ⊗ Q and
write in the usual manner:
2(λ, α)
(λ, α∗ ) = , α∈∆
(α, α)

We denote by e(w) the canonical image of w ∈ W in G/B (we see


that it is well-defined). The B-orbits {Be(w)} are the Schubert cells in
G/B and their respective closures Be(w) = X(w), endowed with the
canonical reduced structures are the Schubert varieties in G/B. More
generally, if Q is a parabolic subgroup (Q ⊃ B), we can define Schubert
varieties {X(w)} in G/Q and they are indexed by w ∈ W/WQ , where
WQ is the Weyl group of Q. We have a partial order in W/WQ , namely

τ1 ≥ τ2 ⇔ X(τ1 ) ⊇ X(τ2 ), τi ∈ W/WQ

Let now P be a maximal parabolic subgroup of G. Let L denote the


ample generator of Pic G/P ≃ Z. Recall that P is associated canonically
to a fundamental weight ω and this association can be done in such
a manner that the G-module H 0 (G/P, L) is of highest weight i(ω) (i
being the Weyl involution) and consequently of lowest weight −ω. For
τ ∈ W/WP , denote by X(τ ) the element of the Chow ring Ch( (G/P ) of
G/P determined by the Schubert variety X(τ ). Recall that we have a
unique codimension one Schubert variety H in G/P and that [H] = [L]
(the image of an effective divisor associated to L in Ch(G/P )). It is also
Standard Monomial Theory 167

known that L is very ample (cf. Prop. 5.7, [27]), so that [H] could be
referred to as the “hyperplane class”. It can be shown that (cf. [43])
X
X(τ ) · [H] = di X(φi ), di > 0

where · denotes multiplication in Ch(G/P ) and the summation runs


over the set of all the Schubert subvarieties of X(τ ) of codimension one.
By a formula of Chevalley (cf. [3], [4]), we have

di = |(ω, α∗i )|, τ = φi sαi

sαi being the reflection with respect to αi ∈ ∆+ . We call di the multi-


plicity of X(φi ) in X(τ ). Recall that the fundamental weight ω is said
to be minuscule if |(ω, α∗ )| ≤ 1, ∀α ∈ ∆. It can be seen easily that ω is
minuscule if and only if di ≤ 1, ∀i, τ . For G = SL(n), every fundamen-
tal weight is minuscule. There are exceptional groups with no minuscule
fundamental weights. We say that the fundamental weight ω (resp. P )
is of classical type if
|(ω, α∗ )| ≤ 2, ∀α ∈ ∆
It can be seen easily that ω is of classical type if and only if di ≤ 2, ∀i, τ .
For a classical group, every fundamental weight is of classical type (the
reason for the terminology—classical type).

2.1 Definition Suppose that the maximal parabolic P is of classical


type. A pair of elements (τ, φ) in W/WP is called an admissible pair if
(i) either τ = φ, in which case (τ, φ) is called the trivial admissible
pair, or
(ii) there exist {τi }, 1 ≤ i ≤ s, τi ∈ W/WP , having the following
properties:
(a) τ = τ1 > τ2 > · · · > τs = φ
(b) X(τi ) is of codimension one in X(τi−1 ) and the multiplicity of
X(τi ) in X(τi−1 ) is 2. An admissible pair (τ, φ) on X(θ) is an admissible
pair (τ, φ) such that θ ≥ τ .

2.2 Theorem (First basis theorem for classical type): There exists a
canonically determined basis {p(τ, φ)} of H 0 (G/P, L), indexed by the
admissible pairs in W/WP , such that
(i) p(τ, φ) is a weight vector (under T ) of weight equal to − 12 (τ (ω) +
φ(ω))
(ii) the restriction of p(τ, φ) to a Schubert variety X(θ) is not iden-
tically zero if and only if θ ≥ τ .
168 Appendix C

(iii) ∀θ ∈ W, the elements {p(τ, φ)}, indexed by admissible pairs on


X(θ), θ ∈ W/Wp form a basis of H 0 (X(θ), L) (to be precise, we take
the restrictions of p(τ, φ) to X(θ) etc.)

2.3. Remark We write p(τ ) = p(τ, τ ). Note that the weight of p(τ ) =
−τ (ω), so that it is the τ translate τ p(id) of the lowest weight vector
p(id). Hence {p(τ )} is the set of all the extremal weight vectors.

Idea of proof of the First Basis Theorem (cf. [7], [30])


The above theorem implies that one has a formula for the character
of the T -module H 0 (X(τP
), L), namely
0
(i) Char H (X(τ ), L) = exp( 21 λ(−ω) + µ(−ω)). On the other hand,
one has the Demazure character formula for H 0 (X(τ ), L) which is de-
fined as follows: Let Z[N ] denote the group ring on the group N =
hom(T, Gm ) and with the usual notation, we define the operators:

Lsα , Msα : Z[N ] → Z[N ],

exp(λ) − exp(sα (λ))


Lsα (exp λ) = , λ ∈ N, α ∈ S
1 − exp(α)

Msα (exp(λ) = exp(ρ) · Lsα (exp(λ − ρ)),
ρ = half sum of positive roots
Writing si for sαi for notational convenience, if τ = s1 . . . sr is a reduced
expression, we define
Mτ = Ms1 . . . Msr .
Then the Demazure character formula for H 0 (X(τ ), L) is simply Mr (exp(−ω)).
The first step is to check that
(ii) Mr (exp(−ω)) = RHS of (i) above
Now (ii) is proved by induction on dim X(τ ) as follows: We can write
τ = sα φ (α simple and X(φ) of codimension one in X(τ ), X(φ) is called
a “moving divisor” in X(τ ), moved by α). We assume the truth of (ii)
for X(φ). Let Iλ denote the set of admissible pairs on X(λ), λ ∈ W/WP .
If δ = (λ, µ) is an admissible pair, we set
1
δ(x) = (λ(x) + µ(x)), x∈N
2
Standard Monomial Theory 169

We check the following:


(iii) Admissible pairs on X(τ ) consist of those which are admissible on
X(φ) together with those of the form

(sα λ, sα µ), (λ, µ) admissible pair on X(φ),
(a)
φ 6≥ sα λ, where r = −(δ(−ω), α∗ ) = 1

(sα λ, sα µ) and (sα λ, µ), (λ, µ) admissible pair on X(φ),
(b)
φ 6≥ sα λ, when r = −(δ(−ω), α∗ ) = 2

Partition Iφ into the following subsets:


I 0 = {δ ∈ Iφ | (δ(ω), α∗ ) = 0},
I − = {δ ∈ Iφ | (δ(ω), α∗ ) < 0}
I + = {δ ∈ Iφ | (δ(ω), α∗ ) > 0 and φ ≥ sα λ, δ = (λ, µ)}
I e = {δ ∈ Iφ | (δ(ω), α∗ ) > 0 and φ ≥ sα λ, δ = (λ, µ)}
Set F 0 = Σδ∈I 0 exp(δ(−ω)),
F − = Σδ∈I − exp(δ(−ω))
F + = Σδ∈I + exp(δ(ω)),
F e = Σδ∈I e exp(δ(−ω)).

We check the following:



Msα (F 0 ) = F 0 , Msα (F + + F − ) = F + + F − , so that
(iv)
Msα (F 0 + F + + F − ) = F 0 + F + + F −
and


 Msα (F e ) = F e + Σδ∈Iτ −Iφ exp(δ(−ω)),
 in fact, if δ = (λ, µ), δ ∈ I e ,




Msα (exp δ(−ω)) = exp δ(−ω) + exp δ1 (−ω) + exp δ2 (−ω),
(v)

 with δ1 = (sα λ, µ), δ2 = (sα λ, sα µ) if r = −(δ(−ω), α∗ ) = 2



 M (exp δ(−ω)) = exp δ(−ω) + exp η(−ω)
 sα
with η = (sα λ, sα µ) if r = −(δ(−ω), α∗ ) = 1
Now together with (iii) above, the claim (ii) above follows.

2.4 Remark The validity of Demazure character formula is not strictly


required. This would follow, à posteriori, as a consequence of SMT.

Let GQ (resp. GZ ) denote a split group scheme over Q (resp. over


Z) such that the base change of GZ by k (resp. by Q is G (resp. GQ
is G (resp. GQ ). Let U denote the Lie algebra of GQ . Let UZ (resp.
U Z+ , resp. UZ− ) be the Z subalgebra of U generated by Xαn /n!, α ∈ ∆
170 Appendix C

(resp. α ∈ ∆+ , resp. α ∈ ∆− ), where Xα denotes the element in the


Chevalley basis of Lie GQ corresponding to α. Let Uα (resp. Uα,Z denote
the Q-vector subspace (resp. Z-submodule) of U (resp. UZ ), generated
by Xαn /n!).
Let V denote the finite-dimensional (over Q) GQ -module of highest
weight ω. Fix a highest weight vector ǫ ∈ V . We set

VZ = U Z e

Then VZ is a GZ -module (called the “Weyl module” or what could be


called the “minimal Z-form” in V ) and we have V = VZ ⊗ Q. For
τ ∈ W/WP (which can be represented by a Z-valued point of GZ ), we
set
VZ (τ ) = UZ+ er , where eτ = τ e

(we check VZ (ω0 ) = VZ , ω0 = the element of maximal length in W ). Now


the above theorem is a consequence of its following dual version:

2.2′ Theorem There exists a basis {Q(τ, φ)} of VZ , indexed by admis-


sible pairs in W/WP , such that

(a) Q(τ, φ) is a weight vector of weight 21 (τ (ω) + φ(ω))

(b) if W (θ) denotes the Z-submodule of VZ generated by Q(τ, φ) such


that θ ≥ τ , then W (θ) = VZ (θ).

The basis elements p(τ, φ) of Theorem 2.2 are precisely the elements
P (τ, φ)⊗ 1 in the k-vector space V ∗ ⊗ k (V ∗ dual of VZ ), where {P (τ, φ)}
is the basis of V ∗ dual to {Q(τ, φ)}.
The proof Theorem 2.2′ is again done by induction on dim X(τ ),
taking τ = sα φ (X(φ) a moving divisor in X(τ )), and assuming Theorem
2.2′ for X(φ). Now guided by (ii) and its proof outlined above, we define
the “new basis” elements indexed by the admissible pairs on X(τ ), which
do not live on X(φ) as follows:
Let δ1 = (λ1 , µ1 ) be an admissible pair on X(φ) such that (δ1 (ω), α∗ ) >
0 and φ 6≥ sα λ. Then

(a) if (δ1 (ω), α∗ ) = 1, define

Q(λ, µ) = X−α Q(λ1 , µ1 ), (λ, µ) = (sα λ1 , sα µ1 )


Standard Monomial Theory 171

(b) if (δ1 (ω), α∗ ) = 2, define


2
X−α
Q(λ, µ) = Q(λ1 , µ1 ), (λ, µ) = (sα λ1 , sα µ1 )
2
Q(sα λ1 , µ1 ) = X−α Q(λ1 , µ1 )

Then one checks that the {Q(λ, µ)} which live on X(φ), together with
the new ones defined above, satisfy the properties of Theorem 2.2′ .

2.5 Remark Let (λ, µ) be an admissible pair with



 λ0 > λ1 > λ2 > · · · > λr = µ
X(λi ) of codimension one and multiplicity 2 in X(λi−1 ),

λi = sαi λi−1 , αi ∈ S

Then we have
Q(λ, µ) = X−α1 . . . X−αr Q(u)
It can be shown (cf. Remark 3.8, [30]) that Q(λ, µ) is independent of
the choice of the “chain” connecting λ and µ above. We see then that
the Q(λ, µ) are determined canonically, once the choice of the extremal
vectors {Q(µ)} has been made. We note that {Q(µ)} are determined
up to ± depending on the choice of the highest weight vector e ∈ V .

3 Main Theorems for Classical Type

The main references for this section are [27] and Theorem 9.6, [30].
We keep the notations as in §2.

3.1 Definition Let P be of classical type. A Young diagram of length


m in W/WP (or on G/P ) is an ordered sequence

(τ1 , φ1 ), . . . , (τm , φm )

of m admissible pairs in W/WP . We say that it is standard if

τ1 ≥ φ1 ≥ τ2 ≥ φ2 ≥ · · · ≥ τm ≥ φm

and that it is standard on the Schubert variety X(θ) in G/P , if moreover


θ ≥ τ1 . A standard monomial of length m on G/P (resp. X(θ)) is the
formal expression
p(τ1 , φ1 ) · · · p(τm , φm )
172 Appendix C

attached to a standard Young diagram of length m on G/P (resp. on


X(θ)) as above. This determines an element of H 0 (G/P, Lm ) (resp.
H 0 (X(θ), Lm )) and by abuse of language we call this element a standard
monomial of length m.
We have then the following:

3.2 Theorem Standard monomials of length m on X(θ) form a basis


of H 0 (X(θ), Lm ).

Let G be now a classical group of rank ℓ and P1 , . . . , Pℓ denote the


set of the maximal parabolic subgroups of of G(⊃ B), taken in some
order. We fix this order in the sequel.

3.3 Definition By a Young diagram of type (or multi-degree) m =


(m1 , . . . , mℓ ), mi ≥ 0, we mean a pair (τ, φ) defined as follows:

τ = τij , φ = φij , (τij , φij ) admissible pair in W/WP ,
1 ≤ j ≤ mi 1 ≤ i ≤ ℓ
We say that the Young diagram (τ, φ) is standard on the Schubert variety
X(θ) ⊂ G/B, θ ∈ W (written as θ ≥ (τ, φ)), if there is a pair (α, β) called
a defining pair for (τ, φ) such that:

(i) α = (αij ), β = (βij ); αij , βij in W

(ii) each αij (resp. βij ) is a lift in W of τij (resp. φij )

(iii) θ ≥ α11 ≥ β11 ≥ · · · > αij ≥ βij ≥ αi,j+1 ≥ βi,j+1 ≥ · · · ≥ αi,m ≥


βi,m ≥ αi+1,1 ≥ · · · ≥ αℓ,m ≥ βℓ,m (the notation ≥ refers to the
partial order in W ).

Associated to the Young diagram (τ, φ) we define the monomial p(τ, φ)


as follows: Y Y
p(τ, φ) = p(τij , φij )
1≤i≤ℓ 1≤j≤mi

If (τ, φ) is a standard Young diagram on X(θ), we say that p(τ, φ)


is a standard monomial of type m on X(θ). Let {Li }, 1 ≤ i ≤ ℓ, denote
respectively the ample generators of Pic G/Pi . Then p(τ, φ) defines an
mℓ
element of H 0 (G/P, Lm ) (or H 0 (X(θ), Lm ), where Lm = Lm
1 ⊗· · ·⊗Lℓ
1

(here by Li , we mean the pull-back of Li on G/B or its restriction to


X(θ)). We have then the following:
Standard Monomial Theory 173

3.4 Theorem Standard monomials of type m on X(θ) form a basis of


H 0 (X(θ), Lm ).

3.5 Remark More generally, we could work with G/Q and its Schubert
subvarieties, where Q is a parabolic subgroup such that if {Pi }, 1 ≤ i ≤
ℓ′ , denotes the set of all maximal parabolic subgroups containing Q then
Pi is of classical type. One has then an assertion similar to Theorem 3.4
above.
Idea of proof of Theorems 3.2 and 3.4
Theorem 3.2 is a particular case of Theorem 3.4 if we formulate
Theorem 3.4 in the more general form as in Remark 3.5, above. We
shall limit ourselves mainly to giving an idea of proof of Theorem 3.2.
This proof generalizes also for Theorem 3.4 but its technically more
complicated.

Linear independence of standard monomials (cf. §6, [27] and §5,


[30])
We have to prove the linear independence of standard monomials of
length m on a Schubert variety X(τ ) in G/P , where P is a maximal
parabolic subgroup of classical type. This is done by induction on dim
X(τ ). Suppose first that P is minuscule. Suppose then that we have a
relation on X(τ ) of the form:

(i) a1 F1 + a2 F2 + · · · + ar Fr = 0, ai ∈ k, ai 6= 0, 1 ≤ i ≤ r

where {Fi }, 1 ≤ i ≤ r, are distinct standard monomials of length m on


X(τ ). Then we should show that this leads to a contradiction. We have

F1 = p(λ1 )G1 , F2 = p(λ2 )G2 , . . . , Fr = p(λr )Gr

Suppose that p(λi ) = p(τ ) for all i, 1 ≤ i ≤ r. Then in (i), cancelling


p(τ ), we get a contradiction by an induction argument on the length m
(note {Gi } are standard monomials of length (m − 1) on X(τ )). We can
therefore suppose that there is a λi , say λ1 such that τ > λ1 , τ 6= λ1 .
Then restrict the relation (i) to X(λ1 ). Note that if λ1 6≥ λi , then
p(λi ) vanishes on X(λ1 ) (cf. (ii) of Theorem 2.1). Therefore, only terms
involving λi such that λ1 ≥ λi survive, yielding a similar to (i) on X(λ1 ).
Since dim X(λ1 ) < dim X(τ ), we get a contradiction to the induction
hypothesis.
In the general case when P is no longer minuscule, we see easily that
we have only to prove that a relation of the form:
174 Appendix C

a1 p(τ, λ1 ))G1 + · · · + ar p(τ, λr )Gr = 0, a1 6= 0, ai ∈ k
(ii)
{p(τ, λi )Gi }, 1 ≤ i ≤ r, are distinct standard monomials on X(τ ),
leads to a contradiction. For this we make use of the special quadratic
relations, namely:

(iii) p(τ, λ)2 = p(τ )p(λ) on X(τ )

To prove (iii), one observes that the zero set of p(τ ) on X(τ ) is precisely
the union H(τ ) of all the codimension one Schubert varieties of X(τ )
and that p(τ, λ)2 vanishes with bigger order than that of p(τ ) on the
irreducible components of H(τ ). Then by the normality of X(τ ), we see
that p(τ, λ)2 /p(τ ) is regular on X(τ ), and since its weight is equal to
that of p(λ), it follows that this is equal to p(λ) (all by Theorem 2.2).
In a similar manner, we get the relation:

p(τ, λ1 )p(τ, λ2 ) = p(τ )F, F 2 = p(λ1 )p(λ2 ),
(iv)
on X(τ ), F regular on X(τ )

Now we multiply the relation (ii) by p(τ, λ1 ), cancel p(τ ) (using (iii) and
(iv) above) and restrict to X(λ1 ). Then we get a relation saying that
distinct standard monomials on X(λ1 ) are linearly independent, which
leads to a contradiction.
One need not use the normality of X(τ ). This follows, à posteriori,
by a suitable induction argument.
One can define standard monomials on a union of Schubert varieties
in G/P and prove their linear independence.
To carry over the above argument of linear independence to the
“mixed case” i.e., for Schubert varieties or their unions in G/B (as in
Theorem 3.4 above), one uses the existence of the unique maximal (resp.
minimal) defining pair for a standard Young diagram, which follows from
certain lemmas due to V. Deodhar.

Generation by standard monomials (cf. §7, [27] and §9, [30])


We give an outline of proof for Theorem 3.2 so that we work with
Schubert varieties in G/P , where P is a maximal parabolic of classical
type.
Let X(τ ) be a Schubert variety in G/P . Let X(φ) be a “moving
divisor” in X(τ ) with τ = sα φ. The proof is again by induction on dim
X(τ ) so that in X(τ ) with τ = sα φ. The proof is again by induction on
dim X(τ ) so that we suppose that standard monomials of length m gen-
Standard Monomial Theory 175

erate H 0 (X(φ), Lm ) (and hence form a basis since linear independence


has already been proved).
For a Schubert variety X(λ) in G/P , let OX(λ) (m) denote the restric-
tion of Lm to X(λ) (as well as the corresponding sheaf of sections). Let
H(φ) be the closed subscheme of X(φ), defined by p(φ) = 0. Consider
the set
Mφ = {λ | (φ, λ) admissible pair on G/P }
and index it as {λi }, 1 ≤ i ≤ N , so that if λi < λj then i > j. Then
from the fact that the standard monomial theory holds on X(φ) (by the
induction hypothesis), we deduce easily a filtration of the following form
for the ideal sheaf of H(φ)red called the “reduced hyperplane section of
X(φ)”

 I0 = I(H(φ)) ⊂ I1 ⊂ I2 ⊂ · · · ⊂ IN = I(H(φ)red )
(i) Ij /Ij−1 ≃ OX(λ) (−1), j ≥ 1

I0 is (of course) OX(φ) (−1)

We see on the other hand that SMT is a consequence of filtrations


as in (i). Hence one would like to prove the analogue of filtrations (i)
for the ideal sheaf H(τ )red of the “reduced hyperplane section” of X(τ ).
This does not seem to be easy to prove directly. The crucial idea is
that one can deduce a filtration similar to (i) , for the ideal sheaf of
a reduced subscheme of the “Demazure one-step modification of X(τ )”
(cf. [4]), which projects onto H(τ )red ) and this filtration suffices to prove
generation by standard monomials of H(τ ). Intuitively, one could expect
this, since by the induction hypothesis one understands the situation on
X(φ) and the Demazure 1-step modification is locally a product of X(φ)
and something smooth (in fact one-dimensional). Kempf had used the
1-step modification of Demazure for a similar purpose (cf. [17]).
Let Bα = B ∩ SL(2, α), where SL(2, α) is the “SL(2)” associated
to the simple root α. Then X(τ ) is stable under the canonical action of
SL(2, α) on G/P . Set
Zφ = SL(2, α) ×Bα X(φ)
so that Zφ is the fibre space with fibre X(φ) associated to the principal
Bα -fibration
SL(2, α) −→ SL(2, α)/Bα ≃ P1
We have also a canonical morphism
ψ : SL(2, α) ×Bα X(φ) = Zφ −→ X(τ )
176 Appendix C

defined by (g, x) 7→ gx, g ∈ SL(2, α) and x ∈ X(φ). Now ψ is the “1-


step Demazure modification” mentioned above. It is checked easily that
ψ is birational. Now if we have a Bα -object M on X(φ) (e.g., a coherent
sheaf on X(φ) with Bα -action consistent with the canonical Bα -action
f on Zφ (namely
on X(φ)), we can associate to it canonically an object M
f = SL(2, α) ×Bα M ), called the canonical extension of M to Zφ .
M
We now consider the canonical extension to Zφ of the Bα -filtration
(i) above. We write this as:

(ii) I˜0 ⊂ I˜1 ⊂ I˜2 ⊂ · · · ⊂ I˜N on Zφ

Recall that IN = I(H(φ)red ) (ideal sheaf on Zφ ). We have a canonical


morphism Zφ −→ P1 which is a fibre space of fibre type X(φ). In par-
ticular, the fibre over the point e ∈ P1 ≃ SL(2, α)/Bα (corresponding
to the identity element in SL(2, α)) is ≈ X(φ) and we identify X(φ) as
a closed subscheme of Zφ in this manner. We see that

]
H(φ)red ∪ X(φ) = ∆

is a reduced subscheme of Zφ which projects onto the “reduced hy-


perplane section of X(τ )” (the reduced subscheme of X(τ ) defined by
p(τ ) = 0). We see easily that the following holds on Zφ :

I(∆) = I(H(φ)red ) ⊗ C

where C is the ideal sheaf of the subscheme X(φ) of Zφ . Now tensoring


(ii) by C, we obtain a filtration for the ideal sheaf of ∆ on Zφ :

(iii) K0 ⊂ K1 ⊂ · · · ⊂ KN (= I(∆)), Kj = I˜j ⊗ C

In the case when X(φ) is of multiplicity two in X(τ ), this filtration


has to be replaced by

(iii)′ M0 ⊂ K0 ⊂ · · · ⊂ KNi M0 = K0 ⊗ C

The filtrations (iii) and (iii)′ of the ideal sheaf of the reduced sub-
scheme ∆ of Zφ projecting on H(τ )red are the filtrations, mentioned
above, for deducing generations by standard monomials on X(τ ). To
give a very brief idea of this, let Zλj denote the closed subvariety of Zφ ,
defined by

Zλ = SL(2, α) ×Bα X(λj ); (φ, λj ) admissible pair


Standard Monomial Theory 177

One shows that the sheaf Kj /Kj−1 lives on Zλj ; in fact it is of the form:

ψ ∗ (L−1 )|Zλj up to a twist by a power of C (which is 0,1 or − 1)

We tensor (iii) (resp. (iii)′ ) by Lm (to be precise ψ ∗ (Lm )), split it into
short exact sequences and using the induction hypothesis (which imply
suitable vanishing theorems as well), we check that

h0 (Zφ , Lm ) = #{standard monomials of length m on X(τ )}

which à fortiori, implies that

h0 (X(τ ), Lm ) = #{standard monomials of lengthm on X(τ )}

4 The Exceptional Groups and the Kac-Moody Groups

Let M be a symmetrizable, generalized Cartan matrix. Let g (resp. G)


be the associated Kac-Moody Lie algebra (resp. Kac-Moody group). Let
W be the Weyl group. Let U be the universal enveloping algebra g and
UZ+ , the Z-subalgebra of U generated by Xαn /n!, α a simple root. Let
λbe a dominant, integral weight and Vλ the integrable, highest weight
module (over C) with highest weight λ. Let us fix a generator e for the
highest weight space (note that e is unique up to scalars). For τ ∈ W ,
let eτ = τ e, VZ (τ ) = UZ+ eτ .

A conjectural basis for VZ (τ )


VZ (τ ) has a basis Bτ given as follows:
(1) Iτ , the indexing set of Bτ , is given by

Iτ = {1 ≥ pit /qit > pit−1 /qit−1 > · · · > pi1 /qi1 ≥ 0, µit+1 < µit < · · · < µi1 }

such that

(a) there exists elements µi ∈ W, 0 ≤ i ≤ r + 1 with

τ ≥ µ0 = µi1 > µ1 > µ2 > · · · > µit+1 = µr+1

where for 1 ≤ ℓ ≤ t + 1, each µi1 = µm for some m, 0 ≤ m ≤ r + 1;


further, for 0 ≤ i ≤ r, X(µi+1 ) is a Schubert divisor in X(µi ), say

µi = sβi µi+1 , mi = |(µi (λ), β ∗ )|


178 Appendix C

(b) there exist positive integers ni , 0 ≤ i ≤ r such that

1 ≥ nr /mr ≥ · · · ≥ n0 /m0 ≥ 0

Further in (b) nℓ /mℓ > nℓ−1 /mℓ−1 if and only if ℓ ∈ {i1 , . . . , it }


and for such an ℓ, nℓ /mℓ = pℓ /qℓ .

(2) The Vectors in Bτ : Let Nθ,µ be an element of Iτ , where θ = µi1 and


(n ) (n ) (n )
µ = µii+1 . To Nθ,µ , we associate the vector X−β00 X−β11 · · · X−βrr eµ (here,
(n) n /n!). Note that the above vector
for a real root β, X−β stands for X−β
(if nonzero) is a weight vector of weight χ(Nθ,µ ), where
r
X
χ(Nθ,µ ) = µ(λ) − ni βi
i=0

4.1 Remark

(a) If t = 1 and pi1 = 0, then (Nθ,µ ) is simply µ.

(b) If t = 1 and pi1 /qi1 = 1, then (Nθ,µ ) is simply θ.


P
(c) χ(Nθ,µ ) = t+1
ℓ=1 (pi1 /qiℓ − piℓ−1 /qiℓ−1 )µiℓ (ω)

(here piℓ+1 /qiℓ+1 = 1 and pi0 /qi0 = 0).


In the sequel, we shall denote Nθ,µ as

Nθ,µ = [(µi1 , . . . , µiℓ+1 ); (pi1 /qi1 , . . . , pit /qit )]

and a chain as in (1) satisfying (a) and (b) shall be referred to as a


defining chain for Nθ,µ and shall be denoted as

[(µ0 , µ1 , . . . , µr+1 ); (n0 /m0 , n1 /m1 , . . . , nr /mr )]

Also, for a divisor X(φ) in X(τ ), m(φ, τ ) shall denote the multiplicity of
X(φ) in X(τ ).
The above conjecture has been verified to hold for the cases where a
standard monomial theory has been developed. In this section, we shall
briefly recall the results of [19], [20], [28], [32].

Exceptional Groups
Let us suppose that the fundamental weight ω is such that (ω, α∗ ) ≤ 3,
for α ∈ ∆+ . We have the following possibilities for Nθ,µ :
Standard Monomial Theory 179

(1) [(µi1 , µi2 ); (0)] or [(µi1 , µi2 ); (1)] (note that the corresponding vectors
are the extremal weight vectors eµi2 and eµi1 respectively).
 
(2) (µi1 , µi2 ); 12
 
(3) (µi1 , µi2 ); 31
 
(4) (µi1 , µi2 , µi3 ); 31 , 12
 
(5) (µi1 , µi2 , µi3 ); 13 , 23
 
(6) (µi1 , µi2 , µi3 ); 12 , 23
 
(7) (µi1 , µi2 , µi3 , µi4 ); 13 , 21 , 23

4.2 Definition For τ in W/WP , let

Iτ = {Nθ,µ , τ ≥ θ}

4.3 Remark It should be remarked that in the multiplicity 3 case, there


do exist chains of the form X(µ0 ) ⊃ X(µ1 ) ⊃ X(µ2 ) ⊃ X(µ3 ), where
X(µi ) is a divisor in X(µi−1 ), 1 ≤ i ≤ 3, such that m(µ1 , µ0 ) = 2 =
m(µ3 , µ2 ), m(µ2 , µ1 ) = 3 and such chains do not figure in Iτ since we
cannot choose ni , i = 0, 1, 2, such that 1n0 /2 ≥ n1 /3 ≥ n2 /2 > 0.

4.4 Theorem There exists a basis {Q(Nθ,µ )} for VZ (τ ), indexed by Iτ .

4.5 Definition Let {P (Nθ,µ )} be the basis of VZ (τ )∗ , Z-dual of VZ (τ ),


dual to {Q(Nθ,µ )}. For any field k let p(Nθ,µ ) = P (Nθ,µ ) ⊗ 1.

4.6 Theorem (1) p(Nθ,µ ))|X(τ ) 6≡ 0 if and only if τ ≥ θ.


(2) {p(Nθ,µ ), τ ≥ µ} is a basis for H 0 (X(τ ), L).

4.7 Definition A monomial p(Nθ1 ,µ1 )p(Nθ2 ,µ2 ) · · · p(Nθm ,µm ) is said to
be standard on X(τ ) if

τ ≥ θ 1 ≥ µ1 ≥ θ 2 ≥ · · · ≥ µm

4.8 Theorem The standard monomials on X(τ ) of degree m form a


basis of H 0 (X(τ ), Lm ).
180 Appendix C

The above results are proved in the same spirit as the theorems of
§2. We are required to write down more relations among the p(Nθ,ν )’s
to prove the linear independence of standard monomials is proved by
considering ψ : Z → X(τ ) (cf. §3) and chasing some exact sequences.

4.9 Definition Let τ ∈ W, L = ⊗ℓi=1 Lm i .


i
A monomial F =
Qℓ Qmi
i=1 j=1 p(Nθij ,µij ) of multidegree m = (m1 , . . . , mℓ ) is said to be
standard on X(τ ), if there exists a sequence {λij , δij , 1 ≤ j ≤ mi , 1 ≤
i ≤ ℓ}, where λij , δij ∈ W such that
(1) τ ≥ λ11 ≥ δ11 ≥ λ12 ≥ · · · ≥ λ1m1 ≥ δ1m1 ≥ λ21 ≥ · · · ≥ δℓmℓ

(2) πi (X(λij )) = X(θij ), πi (X(δij )) = X(µij ), under πi : G/B → G/Pi .

4.10 Theorem The standard monomials on X(τ ) of multidegree m form


a basis of H 0 (X(τ ), L).

4.11 Remark The conjecture (in the beginning of this section) has been
verified to hold for the multiplicity 4 case also (cf. [25]). And there seems
no apparent difficulty in verifying it for the cases where the multiplicity
is 5 or 6 and thus extending the theory to these cases. (The details will
appear elsewhere.) Thus, it could be said (in view of the conjecture)
that the standard monomial theory is complete for semisimple groups.

The group SL dn
Let H = SL(n, k), k being the base field which we assume to be alge-
braically closed. Let A = k[t, t−1 ], A+ = k[t], A− = k[t−1 ]. Let D be
the maximal torus in H consisting of all upper triangular matrices. Let
N be the normalizer of D in H. The projection π + : A+ → k sending
t to 0, induces a map π + : H(A+ ) → H. Let B = (π + )−1 (B), W =
N (A)/D, G = H(A). We have a cellular decomposition
[
(*) G= BwB
w∈W

Let g be the Kac-Moody Lie algebra corresponding to the matrix


 
2 −1 0 0 · · · 0 −1
−1 2 −1 0 · · · 0 0
 
 .. 
 . 
−1 0 0 ··· ··· −1 2
Standard Monomial Theory 181

With notation as in [15], let S = {α0 , αS


1 , . . . , αn−1 } be the set of simple
roots of g. For w ∈ W , let X(w) = Bτ B (modB) be the Schubert
variety in G/B (see [16] for generalities on the infinite-dimensional flag
variety G/B). Let us fix a fundamental weight ωi , 0 ≤ i ≤ n − 1 and let
P be the maximal parabolic subgroup of G obtained by “omitting αi ”.
Let W P be the set of minimal representatives of W/WP .

The case n = 2: In this case we have the following (cf. [32]).

1. The Schubert varieties in G/P are totally ordered.

2. In each dimension there is precisely one Schubert variety in G/P .

3. Let X(φ) be the Schubert divisor in X(τ ). Let dim X(τ ) = d.


Then m(φ, τ ) = d.

According to the conjecture (stated in the beginning of this section),


VZ (τ ) has a basis indexed by Iτ . Now in view of (3) above, a typical
element in Iτ looks like

Nθ,µ = [(µ0 , µ1 , . . . , µr+1 ); (n0 ≤ n1 ≤ · · · ≤ nr )]

where µ0 = θ, µr+1 = µ, nr ≤ dim X(µ) and X(µi ) is a divisor in


X(µi−1 ), 1 ≤ i ≤ r + 1. Such a basis is in fact constructed in [32]. To
make it more precise, let dim X(µ) = d, so that dim X(θ) = d + r + 1.
Let N = (n0 , n1 , . . . , nr ) where n0 ≤ n1 ≤ · · · ≤ nr ≤ d. Let
(n ) (n ) (n )
Q(θ, µ)N = X−β00 X−β10 · · · X−βrr eµ

Then it is shown (cf. [32], Theorem 2.15) that {Q(θ, µ)N , τ ≥ θ} is a Z


basis for VZ (τ ). Now defining {P (θ, µ)N , τ ≥ θ} as the basis of VZ (τ )∗ ,
dual to {Q(θ, µ)N , τ ≥ θ}, we obtain analogous results for Schubert
varieties in G/P and also in G/B.

The general case


As above, we fix a maximal parabolic subgroup P with associated fun-
damental weight ω. We now briefly describe the results of [20].

4.12 Definition (cf. [7]): Let A = (Λ1 , . . . , λr ) be a Young diagram


with λj boxes in the j th row. We say Λ is admissible if

(1) λ1 ≥ λ1 ≥ · · · ≥ λr
182 Appendix C

(2) Number of rows of same length is < n − 1.


Let ω = ωd , for some d, 0 ≤ d ≤ n − 1. We have associated a vector
QA in Vω , to an admissible Young diagram Λ. To describe QA , we first
fill in Λ with integers. We fill in the first row with αd , αd+1 , αd+2 , · · ·
(or just d, d + 1, d + 2, · · · (modulo n)). Then we go down each column
decreasing the value by 1 at a time (again modulo n).

4.13 Definition A corner box in Λ is one such that


(1) It is a box at the end of a certain row, say the kth row rk

(2) ℓ(τk ) > ℓ(τk+1 ).


Consider the right most corner box. It is filled with αj say. Let us
(b)
denote αj by just α. QA is defined as QA = X−α QΛ′ , where Λ′ is a
certain admissible Young diagram obtained from Λ by deleting b of the
corner α-boxes in Λ. (The rule as to which of the corner α-boxes are
to be deleted is described in [20].) Further, to Λ is associated a pair
of Weyl group elements ρ(Λ), δ(Λ) (cf. [20] ) and then it is shown (cf.
[20], Theorem 3.4) that {QA , ρ(Λ) ≤ τ } is a Z-basis for VZ (τ ). Let
Jτ = {Λ | τ ≥ ρ(Λ)}. We shall now exhibit a bijection between Jτ and
Iτ (where recall that Iτ = {Nθ,µ | τ ≥ θ}).
To make the proof simple, we shall assume n = 3 (the proof for the
case n = 2 will be deduced as a special case of that of the case n = 3).
We shall first gather some facts on the configuration of Schubert varieties
inG/P , G = SL d3 .

4.14 Lemma (cf. [20], Lemmas 1.6 and 1.7). Let τ ∈ W P . Let βi , 1 ≤
i ≤ s + 1 < n be consecutive simple roots. Let (τ (ω), βi∗ ) = xi , 1 ≤ i ≤
s + 1, where
(i) x1 + x2 + · · · + xi > 0, 1 ≤ i ≤ s

(ii) x1 + · · · + xs+1 = −m, for some m ≥ 1.


P
Let γ = s+1 i=1 βi and η = sγ τ . Then X(η) is a divisor in X(τ ) with
m(η, τ ) and conversely.

4.15 Lemma Let τ ∈ W P . Further, let (τ (ω), βi∗ ) < 0, i = 1, 2 and βi


simple. Let si denote the reflection with respect to βi . Let τi , 0 ≤ i ≤ 5
be respectively τ, s1 τ, s2 τ, s1 s2 τ, s2 s1 τ, s1 s2 s1 τ . Then X(τi ), 0 ≤ i ≤ 5
form a configuration similar to that of SL3 /B.
Standard Monomial Theory 183

The proof is obvious.

4.16 Definition With notation as in Lemma 4.15, we have the forma-


tion of a hexagon (as shown in the diagram). We shall refer to τ0 (resp.
τ5 ) as the head (resp. the tail) of the hexagon.

τ0

m m2
1

τ1 τ2

m 1 + m2
m 1+ m2

m1 m2
τ4 τ3

m2 m1

τ5

4.17 Lemma With notation as in Definition 4.16, let m(τ5 , τ3 ) = m1 ,


m(τ5 , τ4 ) = m2 . Then we have m(τ4 , τ2 ) = m(τ1 , τ0 ) = m1 , m(τ3 , τ1 ) =
m(τ2 , τ0 ) = m1 , m(τ4 , τ1 ) = m(τ3 , τ2 ) = m1 + m2 .

The proof is an easy verification.

4.18 Lemma With notation as in Lemma 4.17, let γ = S −{β1 , β2 }. Let


θi = sγ τi , i = 1, 2. Then X(τi ) is a divisor in X(θi ), m(τi , θ1 ) = m2 + 1
and m(τ2 , θ2 ) = m1 + 1. Further, θi , i = 1, 2 occur as heads of two
hexagons (adjacent to the hexagon having τ0 as the head).
184 Appendix C

τ0 θ2
θ1

m 2+ 1 m1 m2 m1 + 1

τ2
τ1

The proof is an easy verification.

4.19 Theorem There exists a bijection between

I = {Nθ,µ } and J = {A}

Proof. We shall define maps ξ : I → J, ψ : J → I such that ψ ◦ ξ = idI


and ξ ◦ ψ = idJ .
The following lemma (cf. [30]) will be repeatedly used in the discus-
sion below.

4.20 Lemma Let X(φ) be a moving divisor in X(τ ), say τ = sα φ, for


some α ∈ S. Let X(η) be any other divisor in X(τ ). Let η = sβ τ , for
some positive root β. Let δ = sα η (note that X(δ) is a divisor in X(η)).
Then m(δ, φ) = m(η, τ ).

The map ψ : J → I
Let Λ be an admissible Young diagram. We shall define ψ(A) inductively,
in such a way that if ψ(Λ) = Nθ,µ , then θ = ρ(Λ) (cf.[20], for definition
of ρ(Λ)). If Λ = ∅, then we define ψ(Λ) = the identity element in W .
(a)
Let then Λ 6= ∅. Let QΛ = X−α QΛ′ (cf. [20], §2). Let ρ(Λ′ ) = φ, so that
ρ(Λ) = sα φ. Let ψ(Λ′ ) = Nθ′ ,µ′ (by induction hypothesis). Note that
θ ′ = φ. Let

CΛ′ = [(λ0 , λ1 , . . . , λs+1 ); (n0 /m0 , n1 /m1 , . . . , ns /ms )]

be a defining chain for Nθ′ ,µ′ where λ0 = θ ′ (= φ), λs+1 = µ′ . Let θ = sα φ


and m(φ, θ) = m. We shall define Nθ,µ (by giving a defining chain).

I Let a = 1.
Standard Monomial Theory 185

We shall now define

CΛ = [(µ0 , µ1 , . . . , µr+1 ); (N0 /M0 , N1 /M1 , . . . , Nr /Mr )]

so that CΛ = becomes a defining chain for some Nθ,µ . If n0 /m0 ≥ 1/m,


then we set r = s + 1, µ0 = sα λ0 , N0 = 1, M0 = m, µt = λt−1 , 1 ≤ t ≤
r + 1, Nt = nt−1 , Mt = mt−1 , 1 ≤ t ≤ r. It is easily checked that CΛ
defines an Nθ,µ . Let then n0 /m0 < 1/m. This implies m < m0 . For
simplicity of notation, let us denote m0 by p. Let {β, γ} = S − {α}. We
now distinguish the following two cases:

Case 1 β0 is nonsimple (recall that λ0 = sβ λ1 ). This implies that θ is the


head of hexagon. Let (θ(ω), β ∗ ) < 0 (note that (θ(ω), α∗ ) is also < 0).
Then we have (φ(ω), β ∗ ) < 0 and (φ(ω), δ∗ ) > 0, δ = α, γ. Now β0 being
nonsimple, we have β0 = β + α or β + γ. Now (φ(ω), δ∗ ) = −(m − 1),
if δ = β + γ and hence we conclude that β0 = β + α (since m0 cannot
be m − 1 as m0 > m). Let us fix an i such that sα λj > λj , j ≤ i + 1
(such an i exists since sα λ0 = θ > λ0 = φ and sα λ1 > λ1 ; note that
β0 = β + α ⇒ (λ1 , α∗ ) > 0). The following facts are easily checked

(i) β0 , β1 , . . . , βi are alternatively nonsimple and simple

(ii) mj (= m(λj+1 , λj )) = p − j, 0 ≤ j ≤ i (recall that p = m(λ1 , λ0 ))

(iii) ni ≥ ni−1 ≥ · · · ≥ n0 (note that in view of (ii), the condition


ni /mi ≥ · · · ≥ n0 /m0 reduces to ni ≥ ni−1 ≥ · · · ≥ n0 ). Now
n0 /p(= n0 /m0 ) < 1/m ⇒ (n0 + 1)/(m + p) > n0 /p.

(1) If n1 /(p − 1) ≥ (n0 + 1)/(m + p), then we take r = s + 1, µt =


sα λt , t = 0, 1; N0 = n0 , M0 = m0 , N1 = n0 + 1, M1 = m + p, µt =
λt−1 , 2 ≤ t ≤ r + 1, Nt = nt−1 , Mt = mt−1 , 2 ≤ t ≤ r.
If n1 /(p − 1) < (n0 + 1)/(m + p), this then implies that (n0 − n1 +
1)/(m + 1) > n1 /(p − 1). We then have 1 > n1 − n0 and hence we obtain
n1 = n0 (since n1 ≥ n0 ).
(2) If n2 /(p − 2) ≥ 1/(m + 1), then we take r = s + 1, µt = sα λt , 0 ≤
t ≤ 2; Nt = nt , Mt = mt , t = 0, 1, N2 = 1, M = m + 1, µt = λt−1 , 3 ≤ t ≤
r + 1, N = nt−1 , Mt = mt−1 , 3 ≤ t ≤ r.
If n2 /(p−2) < 1/(m+1), this then implies that (n2 +1)/(m+p−1) >
(n2 )/(p − 2).
(3) If n3 /(p − 3) ≥ (n2 + 1)/(m + p − 1), then we take r = s + 1, µt =
sα λt , 0 ≤ t ≤ 3; Nt = nt , Mt = mt , 0 ≤ t ≤ 2, N3 = n2 + 1, M3 =
m + p − 1, µt = λt−1 , 4 ≤ t ≤ r + 1, Nt = nt−1 , Mt = mt−1 , 4 ≤ t ≤ r.
186 Appendix C

Let 
1 + nk−1 , if k is odd
xk =
1, if k is even
and 
m + p − q, if k = 2q + 1
yk =
m + q, if k = 2q
Then proceeding as above, we have the following two possibilities:

A. There exists a j ≤ i such that

(a) n2k+1 = n2k , 2k + 1 ≤ j

(b) nk /(p − k) < xk /yk , k < j

(c) nj /(p − j) ≥ xj /yj

(or)
B. (a) n2k+1 = n2k , 2k + 1 ≤ i
(b) nk /(p − k) < xk /yk , k ≤ i
If (A) holds, then we take r = s + 1, µt = sα λt , 0 ≤ t ≤ j; Nt =
nt , Mt = mt , 0 ≤ t < j, Nj = xj , Mj = yj , µt = λt−1 , j < t ≤
r + 1, Nt = nt−1 , Mt = mt−1 , j < t ≤ r.
If (B) holds, then the condition ni /(p − i) < xi /yi implies that
(xi ± ni )/(yi ± (p − i)) > ni /(p − i). Let

 xi + n i


 , if i is even
xi y i+p−i
=
yi′ 
 xi − n i
 , if i is odd
yi − (p − i)

If i = s, then we take r = s+1, µt = sα λt , 0 ≤ t ≤ s+1, Nt = nt , Mt =


mt ,
0 ≤ t ≤ s, µr+1 = λs+1 , Nr = x′ , Mr = y ′ . Let then i < s. We
consider the following two subcases:

Subcase 1(a) i is odd, say i = 2q − 1.


We have x′ = 1, y ′ = m+q. Further, λi and sα λi+1 occur as the heads
of two adjacent hexagons. If {β, γ} = S−{α}, then we have (λi (ω), δ∗ ) <
0, δ = β, γ. Also (sα µi+1 (ω), α∗ ) < 0 and (sα µi+1 (ω), δ∗ ) < 0 for
precisely one δ ∈ {β, γ}, say (sα µi+1 (ω), β ∗ ) < 0. Hence βi+1 = β, β + α
or β + γ.
Standard Monomial Theory 187

sα λ i+1
λi

α
γ
λi+ 1

β
If ni+1 /mi+1 ≥ 1/(m + q), then we take r = s + 1, µt = sα λt , 0 ≤
t ≤ i + 1; Nt = nt , Mt = mt , 0 ≤ t ≤ i, Ni+1 = 1, Mi+1 = m + q, µt =
λt−1 , i + 2 ≤ t ≤ r + 1, Nt = nt−1 , Mt = mt−1 , i + 2 ≤ t ≤ r. Let then
ni+1 /mi+1 < 1/(m + q). This implies that βi+1 = β or β + α (since for
βi+1 = β + γ, we have mi+1 = m + q − 1 and ni+1 /mi+1 > 1/(m + q)).
(a) Let βi+1 = β.
We have mi+1 = m + p − q. Now the relation ni+1 /mi+1 ≥ ni /mi
implies that

(∗) ni+1 > ni

(since in this case mi+1 = m + p − q > m − i = mi ). Also, ni+1 /mi+1 <


1/(m + q) ⇒

ni+1 − 1 1
(∗∗) <
p−i−1 m+q

Further, we have

ni ni+1 − 1
(∗∗∗) ≤
p−i p−i−1

(since ni+1 − 1 ≥ ni (cf. (∗) above)). Thus we obtain (from (∗∗) and
(∗∗∗))
ni ni+1 − 1 1
≤ <
p−i p−i−1 m+q
(1) Let i + 2 = s + 1 or i + 2 < s + 1 and ni+2 /mi+2 ≥ 1/(m + q). We
take r = s + 1, µt = sα λt , 0 ≤ t ≤ i + 1; Nt = nt , Mt = mt , 0 ≤ t ≤ i,
188 Appendix C

µi+2 = sα+β λi+2 , Ni+1 = ni+1 − 1, Mi+1 = p − i − 1, Ni+2 = 1, Mi+2 =


m + q, µt = λt−1 , i + 3 ≤ t ≤ r + 1, Nt = nt−1 , Mt = mt−1 , i + 3 ≤ t ≤ r.
(2) Let i + 2 < s + 1 and ni+2 /mi+2 < 1/(m + q). Now, βi+2 = α or
γ and hence in view of the condition that ni+2 /mi+2 < 1/(m + q), we
have in fact βi+2 = α (note that if βi+2 = γ, then mi+2 = m + q − 1
and 1/(m + q) < ni+2 /mi+2 ). Now βi+2 = α ⇒ mi+2 = p − i − 1; and
ni+2 /mi+2 < 1/(m + q) implies that

ni+2 + 1 1
(!) <
m+p−q m+q

Also ni+1 /mi+1 ≤ ni+2 /mi+2 < 1/(m + q) implies that

ni+1 − 1 ni+2 + 1
(!!) <
p−i−1 m+p−q

(for, letting m + q = b, p − i − 1 = c, ni+1 = x, ni+2 = y, we have


1/b > y/c ≥ x/(b + c) which implies that
y+1 (cx/b + c) + 1 cx + b + c x−1
> = 2
>
b+c b+c (b + c) c
(since x < (b + c)/c). We take
r = s + 1, µt = sα λt , 0 ≤ t < i + 1; Nt = nt , Mt = m, 0 ≤ t ≤ t,
µi+2 = sα+β λi+2 , µi+3 = sβ λi+3 , µt = λt−1 , i + 4 ≤ t ≤ r + 1,
Ni+1 = ni+1 − 1, Mi+1 = p − i − 1, Ni+2 = ni+2 + 1, Mi+2 = m + p − q,
Ni+3 = 1, Mi+3 = m + q, Nt = nt−1 , Mt = mt−1 , i + 4 ≤ t ≤ r.
We have (in view of (!) and (!!)) Ni+3 /Mi+3 ≥ Ni+2 /Mi+2 ≥ · · · ≥
N0 /M0 . Now Ni+4 /Mi+4 may not be ≥ Ni+3 /Mi+3 , but then, by the
above procedure we have replaced
[(sα λ0 , sα λ1 , . . . , sα λi+1 , λi+1 , . . . , λs+1 );
(N0 /M0 ≤ . . . ≤ Ni+1 /Mi+1 > Ni+2 /Mi+2 ≤ . . . ≤ Nr /Mr )]
where sα λi+1 is the head of a hexagon and Ni+1 /Mi+1 = 1/(m + q) by
[(sα λ0 , . . . , sα λi+1 , sα+β λi+2 , sβ λi+3 , λi+3 , . . . , λs+1 );
(N0 /M0 ≤ . . . ≤ Ni+3 /Mi+3 /Mi+3 )] where sβ λi+3 is the head of a
hexagon and Ni+3 /Mi+3 = 1/(m+q). Hence by repeating the above pro-
cedure as many times as required it is clear that we obtain the required
CΛ .
Standard Monomial Theory 189

(b) Let βi+1 = β + α.


We have mi+1 = p − i − 1. Now
ni + 1 1 ni + 1 ni+1 + 1
< ⇒ <
mi + 1 m+q mi + 1 m+p−q
(1) Let i + 2 = s + 1 or ni+2 < s + 1 and i + 2/mi+2 ≥ ni+1 + 1/m +
p − q. We take r = s + 1, µt = sα λt , 0 ≤ t ≤ i + 2; Nt = nt , Mt =
mt , 0 ≤ t ≤ i + 1, Ni+2 = ni+1 + 1, Mi+2 = m + p − q, µt = λt−1 , i + 3 ≤
t ≤ r + 1, Nt = nt−1 , Mt = mt−1 , i + 3 ≤ t ≤ r.
(2) Let i + 2 < s + 1 and ni+2 /mi+2 < ni+1 + 1/m + p − q. We
have βi+2 = β or γ. We claim: βi+2 = γ. If possible, let us assume that
βi+2 = β. This implies that mi+2 = m + q. We have

ni+1 + 1 ni+2 ni+1


($) > ≥
m+p−q m+q p−i−1

Now ni+1 + 1/m + p − q > ni+2 /m + q ⇒


ni+2 ni+1 +1−ni+2
($$) m+q < p−i−1

Now from ($) and ($$), we obtain


ni+1 ni+1 + 1 − ni+2 ni+1
< <
p−i−1 p−i−1 p−i−1
which is not possible. Hence our assumption that βi+2 = β is wrong and
the claim follows. Now βi+2 = γ implies that mi+2 = p − i − 2. We have
ni+1 ni+2 ni+1 + 1
< <
p−i−1 p−i−2 m+p+q
This implies ni+1 + 1 > ni+2 ≥ ni+1 and hence we obtain ni+2 = ni+1 .
Now the relations ni+2 = ni+1 and ni+2 /mi+2 < ni+1 + 1/m + p − q
imply that ni+2 /p − i − 2 < 1/(m + q + 1). We now consider
ni ni+1 ni+2 1
≤ ≤ <
p−i p−i−1 p−i−2 m+q+1
and take r = s + 1, µt = sα λt , 0 ≤ t ≤ i + 3; Nt = nt , Mt = mt , 0 ≤
t ≤ i + 2, Ni+3 = 1, Mi+3 = m + q + 1, µt = λt−1 , i + 4 ≤ t ≤ r + 1,
Nt = nt−1 , Mt = mt−1 , i + 4 ≤ t ≤ r.
As in (a) (the case βi+1 = β), we have replaced ([(sα λ0 , sα λ1 , . . . , sα λi+1 ,
λi+1 , . . . , λs+1 ); N0 /M0 ≤ · · · ≤ Ni+1 /Mi1 > Ni+2 /Mi+2 ≤ . . . ≤
190 Appendix C

Nr /Mr ] where sα λi+1 is the head of a hexagon and Ni+1 /Mi+1 = 1/(m+
q) by
  
N0 Ni+3
(sα λ0 , . . . , sα λi+3 , λi+3 , . . . , λ); ≤ ··· ≤ ...
M0 Mi+3

where sα λi+3 is the head of a hexagon and Ni+3 /Mi+3 = 1/(m + q + 1).
Hence by repeating the above procedure as many times as required, it
is clear that we obtain the required CΛ .

Subcase 1(b) i is even, say i = 2q.


We have x′ = ni + 1, y ′ = m + p − q. In this case, λi+1 occurs as the
head of a hexagon. Let {β, γ} = S − {α}. Then (λi+1 (ω), α∗ ) > 0 and
(λi+1 (ω), δ∗ ) < 0, δ = β, γ. Further, mi+1 = p − i − 1 or m + q.
If x′i /yi′ ≤ ni+1 /mi+1 , then we take r = s+1, µt = sα λt , 0 ≤ t ≤ i+1;
Nt = nt , Mt = mt , 0 ≤ t ≤ i, Ni+1 = ni + 1, Mi+1 = m + p − q, µt =
λt−1 , i + 2 ≤ t ≤ r + 1, Nt = nt−1 , Mt = mt−1 , i + 2 ≤ t ≤ r.
Let then x′i /yi′ > ni+1 /mi+1 . We claim mI+1 6= m + q. For, if
possible, let us assume that mi+1 = m + q. Now
 
x′i ni + 1 ni+1 ni + 1 − ni+1 ni+1
′ = > ⇒ >
yi m+p−q m+p p−i m+q

This together with the fact that ni+1 /mi+1 > ni /(p − 1) (note that
mi+1 = m + q) ⇒ ni + 1 − ni+1 > ni ⇒ ni+1 < 1, which is not
possible. Hence the assumption that mi+1 = m + q is wrong and the
claim follows. Now the claim implies that mi+1 = p − i − 1. Hence the
relation x′i /yi′ > ni+1 /mi+1 ⇒

ni+1 n + i + 1 − ni+1
(∗) <
p−i−1 m+q+1

Further, we have mi+1 = p − i − 1, mi = p − i and ni+1 /mi+1 ≥ ni /mi .


Hence we obtain

(∗∗) ni+1 ≥ ni

Now (∗) and (∗∗) ⇒ ni = ni+1 (since (∗) implies that ni + 1 − ni1 > 0).
Thus we obtain
ni ni+1 1
(∗∗∗) ≤ <
p−i p−i−1 m+q+1
Standard Monomial Theory 191

We take r = s + 1, µt = sα λt , 0 ≤ t ≤ i + 2; Nt = nt , Mt = mt , 0 ≤ t ≤
i + 1; Ni+2 = 1, Mi+2 = m + q + 1, µt = λt−1 , i + 3 ≤ t ≤ r + 1, Nt =
nt−1 , Mt = mt−1 , i + 3 ≤ t ≤ r.
We have N0 /M0 ≤ . . . ≤ Ni+2 /Mi+2 , Ni+3 /Mi+3 ≤ . . . ≤ Nr /Mr .
Further µi+2 is the head of a hexagon. Thus we are reduced t Subcase
1(a) (note that i + 1 is odd and Ni+2 = 1, Mi+2 = mm+q+1 ).

Case 2. β0 is simple, say β0 = β.


The proof in this case is the same as that of Subcase 1(a), (a). To
make it very precise, we take θ = sα µi+1 , φ = µi+1 in (a) of Subcase
1(a).
(a)
II Let a > 1. (recall, QΛ = X−α QΛ′ ). Let [(λ0 , λ1 , . . . , λs+1 ),
(a−1)
(n0 /m0 , . . . , ns /ms )] be a defining chain for ψ(X−α QΛ′ ) (by induc-
(a−1)
tion on a, we know ψ(X−α QΛ′ )). Let
 
nk+1 nk + 1 nk−1
Z = k, 0 ≤ k ≤ s | ≥ ≥
mk+1 mk mk−1
(for k = s (resp. 0), nk+1 /mk+1 (resp. nk−1 mk−1 ) is 1 (resp. 0)). Note
that Z is nonempty, since s ∈ Z. Let i be the smallest integer in Z. Set
  ′ 
n0 n′s
ψ(Λ) = (λ0 , λ1 , . . . , λs+1 ); ,..., ′
m′0 ms
where 
nk , k 6= 1
n′k =
ni + 1, k = 1
and
m′k = mk , ∀k
This completes the definition of the map ψ : J → I. We shall now define
The map ξ : → J
Let Nθ,µ ∈ I. Let Rθ,µ = { all defining chains for Nθ,µ }. Let us
write Rθ,µ = {R1 , . . . , Rk } for some k. Further let
  
ni0 nir
Ri = (µi0 , µi1 , . . . , µir+1 ); ,...,
mi0 mir
where µi0 = θ, µir+1 = µ. Let
  
′ ni0 nir
R = (µi1 , . . . , µir+1 ); ,...,
mi0 mir
192 Appendix C

By induction on r +1 (= codimension of X(µ) in X(θ)), we may suppose


that ξ(R′ ) has been defined. Let 0 denote the total ordering on J given
by the lexicographic ordering on {n(Λ) = (Λ1 Λ2 . . .), Λ ∈ J}, where, for
Λ ∈ J, Λk = # boxes in the kth row in Λ. Let Λ′ be the greatest (under
0) in {ξ(R′ ), 1 ≤ i ≤ k}, say Λ′ = ξ(Nρ,µ ). Let
  
n1 nr
(µ1 = ρ, µ2 , . . . , µr+1 = µ); ,...,
m1 mr

be a defining chain for Nρ,µ , so that


  
n0 nr
(µ0 = θ, µ1 , . . . , µr+1 = µ); ,...,
m0 mr

is a defining chain for Nθ,µ . We have θ(ω) ≤ χ(Nθ,µ ) ≤ χ(Nθ,µ ) where,


recall that χ(Nθ,µ ) = µ(ω) − Σni βi . Now θ(ω) and χ(Nρ,µ )(= χ(Λ′ ),
the weight of QΛ′ are both weights in Vω and hence χ(Nθ,µ ) is a weight
in Vω . Now the fact that χ(Nθ,µ ) = χ(Nρ,µ ) − n0 β0 is a weight in Vω
implies that (χ(Λ′ ), β ∗ ) ≥ n0 and hence we obtain that # blank corner
β0 -boxes in Λ′ ≥ n0 (recall that (χ(Λ′ ), β ∗ ) = # blank corner β0 -boxes
in Λ′ − # corner β0 -boxes in Λ′ (cf. [20])). Let Λ be the admissible
Young diagram obtained by filling in the first n0 blank corner β0 -boxes
in Λ′ with β0 . Set ξ(Nθ,µ ) = Λ.

Starting point of Induction


Suppose X(µ) is a divisor in X(θ), say θ = sβ µ for some real positive
p
root β. Let m(µ, θ) = m and Nθ,µ = {(θ, µ); ( m )}, If p = 0 (resp. m),
then ξ(Nθ,µ ) = µ (resp. θ ) (recall that for τ ∈ W, τ ∗ is the admissible
∗ ∗

Young diagram representing the extremal weight vector eτ of weight


τ (ω) (cf. [20]). Let then 0 < p < m. Now the fact that (µ(ω), β ∗ ) = m
implies that # blank corner β-boxes in µ∗ ≥ m and we set ξ(Nθ,µ ) = Λ,
where Λ is the admissible Young diagram obtained by filling in the first
p blank corner β-boxes in µ∗ with β. It can be easily checked that
ψ ◦ ξ = idI and ξ ◦ ψ = idJ . This completes the proof of Theorem 4.19.

4.21 Remark The proof of Theorem 4.19 includes ŜL2 as a particular


case. While defining ψ given Λ ∈ J, suppose QΛ = X−α(a) QΛ , and that
ψ(Λ′ ) = Nθ′ ,µ′ . If [(λ0 , λ1 , . . . , λs+1 ); (n0 /m0 , . . . , ns /ms )] is a defining
chain for Nθ′ ,µ′ (where λ0 = θ ′ , λs+1 = µ′ ) then [(sα λ0 , λ0 , λ1 , . . . , λs+1 );
(1/m, n0 /m0 , . . . , ns /ms )] defines ψ(Λ) for the case a = 1 (here
m = m(sα λ0 , λ0 )). The case a > 1 is treated in the same way (as
Standard Monomial Theory 193

in the proof of Theorem 4.19). The map ξ : I → J is defined in the


same way.
Having constructed a Z-basis {QΛ , Λ ∈ Jτ } for VZ (τ ), we define
{PΛ , Λ ∈ Jτ } as the basis of VZ (τ )∗ dual to {QΛ , Λ ∈ Jτ } and prove
analogous results for Schubert varieties in G/P and also in G/B. The
philosophy of the proof of these results is the same as in [30], namely, to
prove results for X = X(τ ), we fix a nice Schubert divisor Y in X. We
then construct a proper birational morphism ψ : Z → X, where Z is a
fiber space over P1 with fiber Y . By induction, we suppose the result
to be true on Y , prove the results for Z and then make the results “go
down” to X.

5 Applications

An important (and rather easy) consequence of SMT is that one obtains


explicit defining equations for the ideal of a Schubert variety, namely:

5.1 Theorem (cf. Theorem 10.3, [30]).


(i) Let X(τ ) be a Schubert variety in G/P , P a maximal parabolic sub-
group of classical type. Then the ideal sheaf of X(τ ) is generated by
{p(λ, µ)}, τ 6≥ λ.
(ii) Let X(τ ) be a Schubert variety in G/B, where G is a classical group
and let {X(τi )}, 1 ≤ i ≤ ℓ, be the images of X(τ ) in G/Pi , 1 ≤ i ≤ ℓ, by
the canonical morphisms G/B → G/Pi (Pi are the maximal parabolics
containing B). Then the ideal sheaf of X(τ ) in G/B is generated by
{p(λ, µ)}, of the form τi 6≥ λ, 1 ≤ i ≤ ℓ (here we take the pull-back to
G/B of p(λ, µ)).
Now (i) is an immediate consequence of Theorem 3.2. To prove (ii),
one observes that
\
X(τ ) = πi−1 (X(τi )) (set theoretically)

As a consequence of SMT, one sees that Schubert varieties “behave well”


with respect to scheme-theoretic unions and intersections, so that the
above equality holds in fact, scheme theoretically. Then (ii) follows from
(i).
The original motivation for SMT was to prove the Kodaira type of
vanishing theorems for line bundles on G/B and their restrictions to the
Schubert varieties. This includes questions like whether the singularities
of Schubert varieties are Cohen-Macaulay, etc. In many important cases
194 Appendix C

SMT gave such results for the first time (cf, [37], [8]). The proof is
always by induction on the dimension of the Schubert variety and the
reason why such a method could work is because SMT gives a hold
on the ideal of a Schubert variety, its hyperplane intersection, etc. At
present the best results in the direction (also valid for any semisimple
G) are proved by the methods of Frobenius splitting initiated by the
work of V. Mehta and A. Ramanathan (cf. [36]; see also [40], [41],
[42], [18]). Recall that Andersen and Haboush had earlier proved the
Kodaira type of vanishing theorems on G/B by Frobenius methods (cf.
[1], [13]). There is, however, an interesting relationship between Cohen-
Macaulayness and combinatorics, which SMT gives in many cases. We
shall now describe this briefly.

Degeneration of Schubert varieties


Let X(τ ) be a Schubert variety in the Grassmannian and I the par-
tially ordered set
I = {λ | τ ≥ λ}
Let {Yλ }, λ ∈ I, be a set of (distinct) indeterminates indexed by I and X0
the projective scheme defined by the following homogeneous equations:

Yλ Yµ = 0; λ, µ not comparable

As a consequence of SMT, it follows that X0 is a specialization of


X(τ ),i.e., there is a 1-parameter flat family {Xt }, t ∈ T (T smooth of di-
mension one), such that Xt0 ≈ X0 for a t0 ∈ T and for t 6= t0 , Xt ≈ X(τ ).
Further X0 is reduced. The scheme X0 is an object canonically associated
to the partially ordered set and can be shown to be Cohen-Macaulay.
This is a consequence of the fact that I is a“wonderful poset”. Then it
follows that X(τ ) is Cohen-Macaulay. All these results have been proved
by Deconcini, Eisenbud and Procesi (cf. [5]) by a critical analysis of the
earlier work of Musili (cf. [37]).
The above results were later extended to the case of Schubert vari-
eties in G/P , P a maximal parabolic subgroup of classical type in [8]
using the “shellability” of partial orders in Weyl groups (cf. [2]). These
results were then extended to the Case of Schubert varieties in SL(n)/B
(cf. [14]).
Similar results have not been proved for Schubert varieties in G/B,
where G is say a classical group.
Standard Monomial Theory 195

Singular locus of Schubert varieties (cf. [31], [21]).


This is probably the most interesting application of SMT, as at
present. Let X(τ ) be a Schubert variety in G/B, G being a classical
group. Let {p(λ, µ)} be the explicit set of generators for the ideal sheaf
of X(τ ) as in Theorem 5.1, above. Now if w ∈ W is such that τ ≥ w,
then e(w) ∈ X(τ ) and X(τ ) and X(τ ) is the union of the Schubert cells
Be(w). Now the rank of the Zariski tangent space to X(τ ) remains con-
stant on each one of these cells. Determination of the singular locus of
X(τ ) is a consequence of the determination of the Zariski tangent space
to X(τ ) at any point of X(τ ). The determination of dimension of the
Zariski tangent space to X(τ ) at one of its point is equivalent to the
determination of the rank of the Jacobian matrix of the above set of
generators of the ideal of X(τ ) in G/P . We determine this rank and it
suffices to do this at the point {e(w)}, τ ≥ w.
Let Jτ denote the matrix
 
··· ···
Jτ = (X−α p(λ, µ)) =  X−α p(λ, µ) X−β p(λ, µ) 
··· ···
where α, β, . . . ∈ ∆+ and X−α is the usual element of Lie G associated
to −α. Note that in each row, p(λ, µ) is kept fixed and in each column,
X−α is fixed. From the fact that the unipotent part U − of B − forms a
coordinate neighborhood of eB (= B.id) ∈ G/B and the fact that
Y
U− = Ga,−α (Ga, −α ∼ Ga )
α∈∆+

it is not difficult to guess and show that the rank of the Jacobian matrix
of {p(λ, µ)} at ew (with respect to the ambient smooth variety G/P ) is
the rank of the matrix Jτ (ew ). Say ew = eB . From the fact
p(λ, µ)(eB ) 6= 0 ⇔ p(λ, µ) = p(id)
and weight considerations, it follows that in each row of Jτ (eB ), there
is at most one non zero entry. Then we see that rank of Jτ (eB ) is the
number of its nonzero columns. These considerations hold for Jτ (ew )
and we get the following.

5.2 Theorem
(i) Rank of Jτ (eB ) = #R(τ ), where R(τ ) = {α | α ∈ ∆+ , X−α p(λ, µ) =
cp(id), c 6= 0, for some p(λ, µ) in the set of generators {p(λ, µ)} of the
ideal sheaf of X(τ ) on G/B, given above}
196 Appendix C

(ii) More generally, we have Rank Jτ (ew ) = #R(τ, w), where R(τ, w) =
{α | α ∈ w(∆+ ), X−α p(λ, µ)} = cp(w), c 6= 0 of the ideal sheaf of X((τ )
on G/B, given above }
(iii) dim TX(τ ),e(w) = N − #R(τ, w), where N = dim G/B = #∆+
and TX(τ ),e(w) is the Zariski tangent space of X(τ ) at ew .

5.3 Corollary Let G = SL(ℓ + 1). We have then

N − #R(τ, w) = #{α ∈ w(∆+ ) | τ ≥ sα w}

Hence
dim TX(τ ),e(w) = #{α ∈ w(∆+ ) | τ ≥ sα w}
In particular
dim TX(τ ),eB = #{α ∈ ∆+ | τ ≥ sα }

5.4 Remark Using Corollary 5.3, V. Deodhar has shown that X(τ ) ֒→
SL(ℓ + 1)/B is smooth at ew (τ ≥ w) if and only if it is rationally smooth
at ew , i.e., the Kazhdan-Lusztig polynomial P (τ, w) ≡ 1 (cf. [6]).
In [22], the vectors Q(λ, µ), (λ, µ) an admissible pair, have been ex-
plicitly computed for classical groups and these computations lead to
an explicit determination of the singular loci of Schubert varieties for
classical groups (cf. [21]). We shall briefly state below the results on the
singular loci of Schubert varieties for classical groups. For a maximal
parabolic subgroup Pd , 1 ≤ d ≤ ℓ, let WPd be the Weyl group of Pd and
W pd , the set of minimal representatives of W/WP in W of wWPd .

The involution σ
Let
A = {(a1 , . . . , ad ) | a1 < · · · < ad , ai ∈ Z}
We have a natural partial order ≥ in A, namely,

(∗) (a1 , . . . , ad ) ≥ (b1 , . . . , bd ), if ai ≥ bi , 1≤i≤d

This partial order among d-tuples will be used in the sequel in describ-
ing the Bruhat order in W Pd . Further, for any d-tuple (z1 , . . . , zd ) of
integers, we let
(z1 , . . . , zd ) ↑= (zi1 , zi2 , . . . , zid )
where j → ij is a permutation and zij ≤ zij+1 . Thus, (z1 , . . . , zd ) ↑ is the
d-tuple whose entries are obtained by arranging the entries (z1 , . . . , zd ) in
Standard Monomial Theory 197

increasing order. We shall denote the elements of the symmetric group


Sm , where m ∈ N, in the following way. Let σ ∈ Sm be such that
σ(i) = ci , 1≤i≤m
We shall denote σ by (c1 · · · cm ). Let k be the base field. For any positive
integer m, let {e, . . . , em } denote the standard basis of km .

I The sympletic
 group Sp(2n)
0 J
Let E = , where
−J 0
 
0 1
 · 
 

J = · 

 · 
1 0
Let (, ) be the skew symmetric bilinear form on k2n , represented by E,
with respect to {e1 , . . . , e2n }. Let

(1) G = Sp(2n) = {A ∈ SL(2n) | t AEA = E}

Let σ be the involution on SL(2n) defined by

(2) σ(A) = E(t A)−1 E −1 , A ∈ SL(2n)

We see that

(3) Sp(2n) = SL(2n)σ

In view of (3), we obtain an identification of W , the Weyl group of G,


with a subgroup of S2n (= the Weyl group of SL(2n)), namely

(4) W = {(α1 , . . . , a2n )} | ai = 2n + 1 − a2n+1−i , 1 ≤ i ≤ 2n}

See [33] for details.


The above identification (cf. [33]) of W , and straightforward calcu-
lations allow us to identify W Pd as
  
  (1) 1 ≤ a1 < a2 < · · · < ad ≤ 2n, 
(5)W Pd = (a1 , . . . , ad ) | (2) for 1 ≤ i ≤ 2n, if i ∈ {a1 , . . . , ad },
  
then 2n + 1 − i 6∈ {a1 , . . . , ad }
198 Appendix C

For w ∈ W , say w = (c1 , . . . , c2n ), we see easily that

(6) w(d) = (c1 , . . . , cd ) ↑

Under the above identification of W Pd , we have (cf. [39]), given two


elements (a1 , . . . , ad ), (b1 , . . . , bd ) in W Pd ,

(7) (a1 , . . . , ad )  (b1 , . . . , bd ) if and only if (a1 , . . . , ad ) ≥ (b1 , . . . , bd )

Thus, the Bruhat order in W Pd coincides with the natural order (cf. (*))
on d-tuples.

II The special orthogonal group SO(2n + 1)


Let  
0 1
 · 
 
E= · 

 · 
1 0 2n+1×2n+1

and let (, ) be the symmetric bilinear form on k2n+1 , represented by E,


with respect to {e1 , . . . , e2n+1 }. Let

(8) G = SO(2n + 1) = {A ∈ SL(2n + 1) | t AEA = E}

Let σ be the involution on SL(2n + 1) defined by

(9) σ(A) = E(t A)−1 E, A ∈ SL(2n + 1)

As in I, we have

(10) SO(2n + 1) = SL(2n + 1)σ

In view of (10), we obtain identification for the Weyl group W , and


also for W Pd similar to (4) and (5), namely

(11) W = {(a1 , . . . , a2n+1 ) ∈ S2n+1 | ai = 2n + 2 − ai′ , 1 ≤ i ≤ 2n + 1}

where i′ = 2n + 2 − i and

(12) W Pd =
Standard Monomial Theory 199
  

 
 (1) 1 ≤ a1 < a2 < · · · < ad ≤ 2n + 1, 

  
(2) ai 6= n + 1, 1 ≤ i ≤ d,
(a1 , . . . , ad ) |

 
 (3) for 1 ≤ i ≤ 2n + 1, if i ∈ {a1 , . . . , ad }, 

  
then 2n + 2 − i 6∈ {a1 , . . . , ad }

For w ∈ W , say w = (c1 , . . . , c2n+1 ), we have

(13) w(d) = (c1 , . . . , cd ) ↑

As in I, we have (cf. [39]) that the Bruhat order in W Pd coincides with


the natural order (cf. (*)) on d-tuples.

III The special orthogonal group SO(2n)


Let  
0 1
 · 
 
E= · 

 · 
1 0 2n×2n

and let (, ) be the symmetric bilinear form on k2n , represented by E,


with respect to {e1 , . . . , e2n }. Let

(14) G = SO(2n) = {A ∈ SL(2n) | t AEA = E}

Let σ be the involution on SL(2n) defined by

(15) σ(A) = E(t A)−1 E, A ∈ SL(2n)

We have

(16) SO(2n) = SL(2n)σ

As in I and II, we obtain, in view of (16), identifications (described


below) for W and W Pd . We have

(17)
   W =
(1) ai = 2n + 1 − ai′ , 1 ≤ i ≤ 2n,
(a1 , . . . , a2n ) ∈ S2n |
(2) #{i, 1 ≤ i ≤ n | ai > n} is even

where i′ = 2n + 1 − i. For 1 ≤ d ≤ n, let


200 Appendix C
  
  (1) 1 ≤ a1 < a2 < · · · < ad ≤ 2n, 
(18) Zd = (a1 , . . . , ad ) | (2) for 1 ≤ i ≤ 2n, if i ∈ {a1 , . . . , ad },
  
then 2n + 1 − i 6∈ {a1 , . . . , ad }

We have for d 6= n − 1

(19) W Pd = Zd

For d = n − 1, if w ∈ W Pd , then

(20) w ≡ wui (mod WP ), 0 ≤ i ≤ n, i 6= n − 1

where

 sαn if i = n,
(21) ui = Id if i = 0,

sαi sαi+1 . . . sαn−2 sαn if 1 ≤ i ≤ n − 2.

(Here Id denotes the identity element in W .) In particular, for w1 , w2 ∈


(n−1)
W , say w1 = (a1 , . . . , a2n ), w2 = (b1 , . . . , b2n ), we can have w1 =
(n−1)
w2 without (a1 , . . . , an−1 ) ↑ and (b1 , . . . , bn−1 ) ↑ being the same.
Thus W Pn−1 gets identified with a proper subset of Zn−1 (cf. definition
(18)). For w ∈ W , say w = (c1 . . . c2n ), we have

(22) w(d) = (c1 , . . . , cd ) ↑, d 6= n − 1

To describe w(n−1) , we let, for 1 ≤ i ≤ n, ı 6= n − 1

(i) (i) the (n − 1)-tuple given by the first (n − 1)


(23) (y1 , . . . , yn−1 ) =
entries in wui

and
(i) (i)
(24) Y = {(y1 , . . . , yn−1 ) ↑, 0 ≤ i ≤ n, i 6= n − 1}

We observe that Y is totally ordered under ≥ (cf. (*)). We have

(25) w(n−1) = the smallest (under ≥ ) element in Y

Unlike the cases of Sp(2n) (resp. SO(2n + 1)), the Bruhat order in W ,
the Weyl group of SO(2n), is not induced from the Bruhat order in S2n .
Standard Monomial Theory 201

Hence the Bruhat order in W Pd does not coincide with the natural order
on d-tuples (cf. (*)). We now describe the Bruhat order in W Pd .
For 1 ≤ i ≤ 2n, let |i| = min{i, i′ } (with i′ = 2n+1−i, as above). Un-
der the above identification, given two elements (a1 , . . . , ad ), (b1 , . . . , bd )
in W Pd , 1 ≤ d ≤ n, we have (cf. [39])

(a1 , . . . , ad )  (b1 , . . . , bd )

if and only if the following two conditions hold:


(A) (a1 , . . . , ad ) ≥ (b1 , . . . , bd ).
(B) Suppose for some r, 1 ≤ r ≤ d and some i, 0 ≤ i ≤ d − r,

(|ai+1 |, . . . , |ai+r |) ↑= (|bi+1 |, . . . , |bi+r |) ↑= {n + 1 − r, . . . , n}

Then
#{j, i + 1 ≤ j ≤ i + r | aj > n}
and
#{j, i + 1 ≤ j ≤ i + r | bj > n}
should both be odd or both even.

Singular Loci of Schubert Varieties for Classical Groups


Let w, τ ∈ W , with w  τ , (here  denotes the Bruhat order in W ).
Let α ∈ ∆+ and β = τ (α). Let T (w, τ ) be the Zariski tangent space to
X(w) at e(τ ).

5.5 Theorem Let G be a classical (assume char k 6= 2, 3, if G =


SO(m)). Then T (w, τ ) is spanned by {X−β | w(d)  τ (α,d) , 1 ≤ d ≤ n},
where τ (α,d) ∈ W Pd is given as follows:

τ (α,d) = (τ sα )(d) , if α = ǫj − ǫk , 1 ≤ j < k ≤ n

In the rest of the cases, τ (α,d) is given as follows:

The symplectic group Sp(2n)


Let τ = (a1 , . . . , a2n ). For 1 ≤ i ≤ 2n, let i′ = 2n + 1 − i and |i| =
min{i, i′ }.
(1) α = 2ǫj , 1 ≤ j ≤ n.

τ (α,d) = (τ sα )(d) , 1≤d≤n


202 Appendix C

(2) α = ǫj + ǫk , 1 ≤ j < k ≤ n.
Let s = min{|aj |, |ak |}, r = max{|aj |, |ak |}


(α,d) (τ sα )(d) if d < k
τ =
(a1 , . . . , âj , . . . , âk , . . . , ad , s′ , τ ) ↑, if k ≤ d ≤ n.

The orthogonal group SO(2n + 1)

Let τ = (a1 , . . . , a2n+1 ). For 1 ≤ i ≤ 2n + 1, let i′ = 2n + 2 − i and


|i| = min{i, i′ }
(1) α = ǫj + ǫk , 1 ≤ j < k ≤ n.
(a) If d < k or d = n, then

τ (α,d) = (τ sα )(d)

(b) If k ≤ d ≤ n − 1, then let s = min{|aj |, |ak |}, r = max{|aj |, |ak |}.


Define si , 0 ≤ i ≤ c(d), as the integers

r = s0 < s1 < s2 < . . . < sc(d) ≤ n

such that si 6∈ {|a1 |, . . . , |ad |}, i 6= 0. If precisely one of {aj , ak } is > n,


then
τ (α,d) = (a1 , . . . , âj , . . . , âk , . . . , ad , s′ , r) ↑
If aj , ak are either both > n or both ≤ n, then

τ (a,d) = (a1 , . . . , âj , . . . , âk , . . . , ad , s′c(d) , s′ ) ↑

(2) α = ǫj , 1 ≤ j ≤ n.
Define si , 0 ≤ i ≤ m(d), as the integers

s0 = |aj | < s1 < s2 < · · · < sm(d) ≤ n

such that si 6∈ {|a1 |, . . . , |ad |}, i 6= 0.


(
(τ sα )(d) , if d < j or d = n
τ (α,d) =
(a1 , . . . , âj , . . . , âk , . . . , ad , s′m(d) ) ↑, if j ≤ d ≤ n − 1
Standard Monomial Theory 203

The orthogonal group SO(2n)

Let τ = (a1 , . . . , a2n ). For 1 ≤ i ≤ 2n, let i′ = 2n + 1 − i and |i| =


min{i, i′ }. Let α = ǫj + ǫk , 1 ≤ j < k ≤ n.
(a) If d < k or n − 2, then
τ (a,d) = (τ sα )(d)
(b) If k ≤ d ≤ n − 2, then let s = min{|aj |, |ak |}, r = max{|aj |, |ak |}.
Define si , −ℓ(d) ≤ i ≤ c(d), as the integers
s < s−ℓ(d) < sℓ(d)+1 < · · · < s−1 < s0 = r < s1 < s2 < · · · < sc(d) ≤ n
such that si 6∈ {|a1 |, . . . , |ad |}, i 6= 0.
If precisely one of {aj , ak } is > n, then
τ (α, d) = (a1 , . . . , âj , . . . , âk , . . . , ad , s′ , r)| ↑
If aj , ak are either both > n or both ≤ n, then

(α,d) (a1 , . . . , âj , . . . , âk , . . . , ad , s′c(d)−1 , s′ ) ↑ if (ℓ(d), c(d)) =
6 (0, 0)
τ = ′ ′
(a1 , . . . , âj , . . . , âk , . . . , ad , r , s ) ↑ if (ℓ(d), c(d)) = (0, 0)

Further applications to classical groups


The varieties X(w) and Y (w)
Let G be Sp(2n) (resp. SO(2n+1), SO(2n)) and H be SL(2n) (resp.
SL(2n + 1), SL(2n)). Let B(H) be the Borel subgroup of H consisting
of all the upper triangular matrices. Let W (G) (resp. W (H)) be the
Weyl group of G (resp. H). As above, we have W (G) ֒→ W (H). For
w ∈ W (G), let ew (resp. fw ) denote the point of G/B (resp. H/B(H)),
where B = B(H)σ . Let X(w) (resp. Y (w)) be the associated Schubert
variety in G/B (resp. H/B(H)). A relative study of X(w) and Y (w)
has been carried out in [23], [24], [29] (note that Y (w) is σ-stable and
that X(w) ⊆ Y (w)σ ). It is shown loc. cit that
1. If G = Sp(2n), then X(w) = Y (w)σ scheme theoretically for all
w ∈ W (G).
2. If G = SO(2n+1), then X(w) = Y (w)σ , set theoretically and the w’s
for which X(w) = Y (w)σ scheme theoretically have been classified.
3. If G = SO(2n), there are w’s for which Y (w)σ is not even irreducible.
The w’s for which X(w) = Y (w)σ scheme theoretically (resp. set
theoretically) have been classified.
204 Appendix C

Generalization of Littlewood Richardson Rule (cf.[34])


Let µ be a dominant weight, say µ = Σℓi=1 ai ωi , ai ∈ Z+ . Let

T = (θ11 , δ11 , θ12 , δ12 , . . . , θ1a1 , θ21 , δ21 , . . . , θℓaℓ , δℓaℓ )

be a standard Young tableau of type a = (a1 , . . . , aℓ ) (cf. Definition


3.3). Let us write T simply as T = (τ1 , τ2 , . . . , τr ) where r = 2Σℓi=1 ai .
For 1 ≤ i ≤ r, set
i
X
χi (T ) = τk (ωt )
k=1

where t is given by τk = θtbt or δtbt , some bt , 1 ≤ bt ≤ at . Let


1
χ(T ) = χr (T )
2
5.6 Definition Let λ be an element of the weight lattice. T is called
λ-dominant if λ + χi (T ) is dominant for all 1 ≤ i ≤ r. Let

Iλ,µ = {T | T is λ-dominant}

5.7 Definition For a dominant weight θ, let Vθ denote the irreducible


G-module with highest weight θ.

5.8 Theorem(cf. [34]). Let char k = 0. Let λ, µ be two dominant


weights. Then M
Vλ ⊗ Vµ = Vλ+χ(T )
T ∈Iλ,µ

Decomposition under a Levi subgroup


Let Q be a parabolic subgroup with SQ as the associated subset of simple
roots. Let LQ or just L be the Levi part of Q.

5.9 Definition With notation as above, T is said to be L-dominant if

(χi (T ), α∗ ) ≥ 0, α ∈ SQ , 1≤i≤r

Let IL,µ = {T | T is L-dominant}

5.10 Definition For a weight λ such that (λ, α∗ ) ≥ 0, α ∈ SQ , let Uλ


denote the irreducible L-module with highest weight λ.
Standard Monomial Theory 205

5.11 Theorem (cf.[34]). Let char k = 0. Let µ be dominant. Then


M
Vµ = Uχ(T ) ( as L–modules)
T ∈IL,µ

Filtration (cf. [35])


Let L1 , L2 be two ample line bundles on G/Q. Using SMT, a nice fil-
tration for H 0 (G/B, L1 ) ⊗ H 0 (G/B, L2 ) by G-modules such that the
successive quotients are again of the form H 0 (G/B/L) has been con-
structed in [35].

Classical invariant theory


Let Gm,2m be the Grassmannian of m-dimensional linear spaces in a 2m
dimensional vector space. We have
 
∗ ∗
Gm,2m ≃ SL(2m)/P, P =
0 ∗

Let ψ be the canonical morphism SL(2m) → SL(2m)/P . Let Z denote


the subgroup of SL(2m):
 
Id 0
Z=
Y Id

Y ∈ M (m)–space of (m × m) matrices. Then we see that ψ maps Z


onto an open subset of Gm,2m , which is the opposite big cell (set of points
where the Plücker coordinate p(id) 6= 0 or equivalently the translate of
the big cell by the element w0 of maximal length in W ). Let

Ik = {(i1 , . . . , ik ) | 1 ≤ i1 < · · · < ik ≤ m}


S
Then we get a bijection W/WP ≃ 0≤k≤m (Ik × Ik ) (k = 0 corresponds
to the element of smallest length, namely (id) ∈ W/WP ). Let (i) =
(i1 , . . . , ik ) ∈ Ik and (j) = (j1 , . . . , jℓ ). We say that

(i) ≤ (j) if k ≥ ℓ and it ≤ jt for t ≤ ℓ

Let τ ∈ W/WP correspond to ((i), (j)) under the above bijection. Then
one gets a canonical identification

p(i),(j) = the function on M (n) defined by the minor
p(τ )|M (m) ↔
corresponding to the indices (i) × (j)
206 Appendix C

We check also that:

τ = ((i), (j)), τ ′ = ((i′ ), (j ′ ))

τ ≥ τ ′ ⇔ (i) ≤ (i′ ), (j) ≤ (j ′ )


Let Dt denote the determinantal subvariety of M (m), defined by the
vanishing of all the (n × n) minors of Mm with n ≥ (t + 1). Then one
checks that

 Dt = M (m) ∩ X(τ ), τ ↔ ((1, . . . , t), (1, . . . , t))
(here M (m) is identified with the opposite

big cell of Gm,2m as above)

Then as a consequence of SMT on Gm,2m , we deduce that the coordinate


ring of Dt has a basis consisting of the following standard monomials:


 p(i),(j) p(i′ ),(j ′ ) p(i′′ ),(j ′′ ) . . . ,

# of any (i), (j), (i′ ), (j ′ ), . . . etc. is ≤ t and
(I)

 (i) ≤ (i′ ) ≤ (i′′ ) ≤ . . .

(j) ≤ (j ′ ) ≤ (j ′′ ) ≤ . . .

Let now X stand for the affine space X = V × · · · × V × V ∗ × · · · × V ∗ ,


dim V = n, V ∗ dual of V . Let us take the canonical diagonal action of
GL(n) on X. Let R denote the coordinate ring of X. Write

x ∈ X, x = (x1 , . . . , xm ; y1 , . . . , ym ), xi ∈ V, yi ∈ V ∗

Consider the canonical morphism

ψ : X → Mm , defined by

x 7→ the matrix | < xi , yj > |


where <, > denotes the canonical bilinear form on V × V ∗ . Then we get
the following formulation of classical invariant theory for GL(n) (cf.[10]).

5.12 Theorem The morphism ψ factors through the determinantal va-


riety Dn ,i.e., we have a commutative diagram:
ψ
X −→Mm
ցφ ր
Dn
Standard Monomial Theory 207

Besides, the morphism φ : X → Dn is surjective and we have a canonical


identification (through φ) of the coordinate ring of Dn with the subring
RGL(n) of R formed by the GL(n) invariants so that by the preceding
discussions, we get a basis of RGL(n) formed of standard monomials as
in I above (with t = n).
Let Sp(2m) denote the sympletic subgroup of SL(2m), which we
take as the fixed point set of the involution σ on SL(2m):
 
0 1
     · 
0 J t −1 0 −J  
σ(A) = ( A) 
,J =  · 
J 0 J 0 
 · 
1 0

Then Q = Sp(2m) ∩ P (P the maximal parabolic of SL(2m) considered


above) is also a maximal parabolic subgroup of Sp(2m) and Sp(2m)/Q
is the sympletic Grassmannian formed by maximal (i.e., of dimension
m) isotropic subspaces of the 2m-dimensional linear space for the skew-
symmetric form associated to the above definition of the symplectic
group.
We see that the above subgroup Z of SL(2m) is σ-stable and the
fixed point subset Z σ of Z under σ, can be identified with the subset
  
I 0 t
, Y =J YJ
Y I

further Z σ can be identified with the opposite big cell of Sp(2m)/Q.


Setting Y = JX, we get

Y = J tY J ⇔ X = tX

i.e., Z σ can be identified with Sym M (m), the space of (m×m) symmet-
ric matrices. Also, we have an order reversing bijection of W (Sp(2m))/WQ
with

{τ ↔ ((i), (j)) ∈ W (SL(2m))/P | (#(i) ≤ m) such that (i) = (j)}

The crucial thing is that if τ = ((i), (i)), φ = ((j), (j)), then (τ, φ) is an
admissible pair in W (Sp(2m))/WQ if and only if

#(i) = #(j) & (i) ≤ (j)


208 Appendix C

One checks also that the restriction of p(τ, φ) to the opposite big cell
(namely, Sym M (m)) of Sp(2m)/Q can be canonically identified with
the polynomial function:
p(i),(j) : Sym M (m) −→ k; (i) ≤ (j)
Let Sym Dt = Dt ∩ Sym M (m); we call this a symmetric determinantal
variety. The element
τ = ((1, . . . , t); (1, . . . , t))
lies in W (Sp(2m)/WQ and if X(τ ) denotes the Schubert variety in
Sp(2m)/Q defined by τ , it can be shown that
X(τ ) ∩ Sym Mm = Sym Dk
i.e., Sym Dk is the opposite big cell of X(τ ). Then as a consequence
of Theorem 3.1, we see that the coordinate ring Sym Dt has a basis
consisting of standard monomials as follows:

p(i),(j) p(i′ ),(j ′ ) p(i′′ ),(j ′′ ) . . . ,
(II)
(i) ≤ (j) ≤ (i′ ) ≤ (j ′ ) ≤ . . . , #(i), #(i′ ), etc. ≤ t.

Let now X stand for the affine space


X = V × · · · × V (m times ), dim V = 2n
Consider the diagonal action of the orthogonal group O(2n) on X (base
field of char 6= 2). Let R be the coordinate ring of X. Consider the
canonical morphism
ψ : X −→ Sym Mm
defined by
x = (x1 , . . . , xm ), x 7→ the matrix (hxi , xj i)
where h, i denotes the scalar product on V defining O(2n). Then we
deduce the following formulation of classical invariant theory for O(2n)
due to DeConcini and Procesi (cf. [9]).

5.13 Theorem The morphism ψ factors through the symmetric deter-


minantal variety Sym D2n , i.e., we have a commutative diagram:
ψ
X −→ Sym Mm
ցφ ր
Sym D2n
Standard Monomial Theory 209

Further the morphism is surjective and we have a canonical identification


(through φ) of the coordinate ring of Sym D2n with the subring RO(2n) of
R formed by the O(2n)-invariants, so that by the preceding discussions
we get a basis of RO(2n) formed of standard monomials as in II above
(with t = 2n).
The above theorem has been deduced here as a consequence of SMT,
but as a matter of fact this was the motivation for the definition of
admissible pairs (cf. [33]).
A result similar to Theorem 5.13 above holds for the diagonal action
of Sp(2n) on V × · · · × V (m times), dim V = 2n, connecting the
invariants ring of under Sp(2n) with the coordinate rings of opposite
big cells of the orthogonal group in 2m variables.
For other applications see [38].

References
[1] H.H. Andersen: The Frobenius morphism on the cohomology of homoge-
neous vector bundles on G/B, Ann. Math. 112 (1980) 113–121.
[2] A.B. Bjorner: Shellable and Cohen-Macaulay partially ordered sets, Trans.
Amer. Math. Soc. 260 (1980) 159–183.
[3] C.C. Chevalley: Sur les décomposition celluaires des espaces G/B
(manuscript non publié circa) 1958.
[4] M. Demazure: Désingularization des variétiés de Schubert generaliseés, Ann.
Sci. École Norm. Sup. 7 (1974) 53–88.
[5] C. DeConcini, D. Eisenbud and C. Procesi: Hodge Algebras, Asterisque 91
1982.
[6] V.V. Deodhar: Local Poincaré duality and non-singularity of Schubert va-
rieties, Communications in Alg. 13 (6) (1985).
[7] E. Date, M. Jimbo, K. Juniba, T. Miwa and M. Okado, Paths, Maya dia-
ˆ C), Advanced Studies in Pure Math. 19
grams and representations of sℓ(r,
(1989) 149–191.
[8] C. DeConcini and V. Lakshmibai: Arithmetic Cohen-Macaulayness and
Arithmetic Normality for Schubert varieties, Amer. J. Math. 103 (1981)
835–850.
[9] C. DeConcini and C. Procesi: A characteristic-free approach to invariant
theory, Advances in Math. 21 (1976) 330-354.
[10] P. Doubilet, G. Rota and J. Stein, On the foundations of combinatorial
theory, Studies in Applied Mathematics, 53 (1974) 185–216.
210 Appendix C

[11] W.V.D. Hodge: Some enumerative results in the theory of forms, Proc.
Camb. Phil. Soc. 38 (1943) 22–30.
[12] W.V.D Hodge and C. Pedoe: Methods of Algebraic Geometry, Vol. II, Cam-
bridge University Press, 1952.
[13] W.J. Haboush: A short proof of Kempf’s vanishing theorem, Invent. Math.
56 (1980) 109–112.
[14] C. Huneke and V. Lakshmibai: Cohen-Macaulayness and Normality of mul-
ticones over Schubert varieties in SLn /B.
[15] V.G. Kac: Infinite dimensional Lie algebras, Progress in Math. 44 Birkhauser,
1983.
[16] M. Kashiwara: The flag manifold of Kac-Moody Lie algebras (preprint).
[17] G. Kempf: Linear systems on homogeneous spaces, Ann. Math. 103 (1976)
557-591.
[18] G. Kempf and A. Ramanathan: Multicones over Schubert varieties, Invent.
Math. 87 (1987) 353–363
[19] V. Lakshmibai: Standard Monomial Theory for G2 , J. Algebra, 98 (1986)
281–318.
c n , Progress in Math. 92,
[20] V. Lakshmibai: Standard Monomial Theory for SL
Birkhauser (1990) 197–217.
[21] V. Lakshmibai: Singular loci of Schubert varieties for classical groups, Bull.
A.M.S. 16 (1987) 83–90.
[22] V. Lakshmibai: Geometry of G/P -VI, J. Algebra, 108 (1987) 355–502.
[23] V. Lakshmibai: Geometry of G/P -VII, J. Algebra, 108 (1987) 403–434.
[24] V. Lakshmibai: Geometry of G/P -VIII, J. Algebra, 108 (1987)435–471.
[25] V. Lakshmibai: The groups F4 and E7 (in preparation).
[26] V. Lakshmibai, C. Musili and C.S. Seshadri: Geometry of G/P -III, Proc.
Ind. Acad. Sc. 87A (1978) 93–177.
[27] V. Lakshmibai, C. Musili and C.S. Seshadri: Geometry of G/P -IV, Proc.
Ind. Acad. Sci.88A (1979) 279–362.
[28] V. Lakshmibai and K.N. Rajeswari: Towards a Standard Monomial Theory
for Exceptional Groups, Contemporary Mathematics, 88 449–578.
[29] V. Lakshmibai and K.N. Rajeswari: Geometry of G/P -IX, J. Algebra, 130
(1990).
[30] V. Lakshmibai and C.S. Seshadri: Geometry of G/P -V, J.Algebra, 100
(1986) 462–557.
Standard Monomial Theory 211

[31] V. Lakshmibai and C.S. Seshadri: Singular locus of a Schubert variety, Bull.
A.M.S. II (2) (1984) 363–366.
c 2 Infi-
[32] V. Lakshmibai and C.S. Seshadri: Standard Monomial Theory for SL
nite Dimensional Lie Algebras and Groups, Advanced Series in Mathemat-
ical Physics, 7 178–234.
[33] V. Lakshmibai and C.S. Seshadri: Geometry of G/P -II, Proc. Ind. Acad.
Sci. 87 A (1978) 1–54.
[34] P. Littlemann: A Littlewood-Richardson rule for classical groups, Comptés
rendus (Série I) 306 (1988) 299–303.
[35] P. Littlemann’s manuscript on Filtrations (preprint).
[36] V.B. Mehta and A. Ramanathan: Frobenius splitting and cohomology van-
ishing for Schubert varieties, Ann. Math. 122 (1985) 27–40.
[37] C. Musili: Postulation formula for Schubert varieties, J. Indian Math. Soc.
36 (1972) 143–171.
[38] C. Musili: Applications of standard monomial theory, in these proceedings.
[39] R. Proctor: Classical Bruhat orders and lexicographic shallability, J. Alge-
bra, 77 (1982) 104–126.
[40] A. Ramanathan: Schubert varieties are arithmetically Cohen-Macaulay, In-
vent. Math. 80 (1985) 283–294.
[41] A. Ramanathan: Equations defining Schubert varieties and Frobenius split-
ting of diagonals, Publ. Math. I.H.E.S. 65 (1987) 61–90.
[42] S. Ramanan and A. Ramanthan: Projective normality of Flag varieties and
Schubert varieties, Invent. Math. 79 (1985) 217–224.
[43] C.S. Seshadri: Geometry of G/P -I, C.P. Ramanujam: A Tribute, 207 (Springer-
Verlag) published for Tata Institute, Bombay, (a978) 207–239.
Bibliography

[A] S. Abeasis, On the Plücker relations for the Grassmannian varieties, Adv.
in Math. 36, 1980 (277–282).
[Bo] A. Borel, Linear algebraic groups, Benjamin, New York, 1969.
[Bou] N. Bourbaki, Groupes et algèbres de Lie, Chapter V, VI, VII, Hermann,
Paris
[B-W1] A. Björner, M. Wachs, Bruhat order of classical groups and shellability,
Adv. in Math. 43 (1982).
[B-W2] A. Björner, M. Wachs, On lexicographically shellable posets, Trans. A.
M. S., Vol. 277, No.1, May 1983.
[C] C. Chevalley, Sur les décompositions cellulaires des espaces G/B, (unpub-
lished manuscript) 1958.
[D] M. Demazure, Désingularisation des variétés de Schubert generalisées,
Ann. Sc. E. N. S., t 7 (1974), 330–354.
[D-E-P] C. De Concini, D. Eisenbud, C. Procesi, Hodge algebras, Astérisque
91, 1982.
[D-P] C. De. Concini, C. Procesi, A characteristic-free approach to invariant
theory, Advances in Math. 21, 1976, 330–354.
[D-R-S] P. Doubilet, G. C. Rota, J. Stein, On the foundations of combinatorial
theory IX, Studies in Applied Math. 53, 1974, 185–216.
[E] D. Eisenbud, Introduction to algebras with straightening law, Proceedings
of the third Oklahoma Conference on Ring Theory, MacDonald, B., ed.
1980.
[E-H] J. Eagon, M. Hochster, Cohen-Macaulay rings, invariant theory and the
generic perfection of determinantal loci, Amer. J. Math., 93 (1971).
[G] A. Grothendieck, Local cohomology, Lecture Notes in Math. No.41,
Springer-Verlag, 1966.
[G-D] A. Grothendieck, J. Dieudonné, Eléments de Géométrie Algébrique,
EGA IV, Publ. Math. IHES, Nos. 28 (1966) and 32 (1967).

© Springer Science+Business Media Singapore 2016 and Hindustan Book Agency 2015 213
C.S Seshadri, Introduction to the Theory of Standard Monomials,
Texts and Readings in Mathematics 46, DOI 10.1007/978-981-10-1813-8
214 BIBLIOGRAPHY

[H] R. Hartshorne, Algebraic Geometry, Graduate Texts in Math. 52, Springer-


Verlag, 1977.
[Ho] W.V.D. Hodge, Some enumerative results in the theory of forms, Proc.
Camb. Phil. Soc., 39 (1943), 22–30.
[Hu1] J. E. Humphreys, Linear algebraic groups, Graduate Texts in Math. 21,
Springer-Verlag, 1975.
[Hu2] J. E. Humphreys, Introduction to Lie algebras and representation theory,
Graduate Texts in Math. 9, Springer-Verlag, 1972.
[H-P] W.V.D. Hodge, D. Pedoe, Methods of algebraic geometry, Cambridge
University Press, Vol. II, 1952.
[H-L] C. Huneke, V. Lakshmibai, Degeneracy of Schubert varieties, Contem-
porary Math., Vol 139, 1992, 181–235.
[I] S. Iitaka, Algebraic Geometry, Graduate Texts in Math. 76, Springer-Verlag,
1981.
[J] A. Joseph, On the Demazure character formula, Annales Scientifiques de
lÉcole Normale Sujérieure Sér. 4, 18, No.3 (1985), 389–419.
[K] G. Kempf, Linear systems on homogeneous spaces, Ann. Math. 103 (1976),
557–591.
[L] V. Lakshmibai, Standard monomial theory for G2 , J. of Algebra, 98 (1986),
281–318.
[L-S1] V. Lakshmibai, C. S. Seshadri, Singular locus of a Schubert variety, Bull.
Amer. Math. Soc., Vol.11, No.2, October, 1984.
[L-S2] V. Lakshmibai, C. S. Seshadri, Geometry of G/P -II (The work of De
Concini and the basic conjectures), Proc. Indian Acad. Sci., 87A, No.2,
1978, 1–54.
[L-S3] V. Lakshmibai, C. S. Seshadri, Geometry of G/P -V, J. of Algebra, 100,
No.2, (1986), 462–557.
[L-M-S1] V. Lakshmibai, C. Musili, C. S. Seshadri, Geometry of G/P -IV (Stan-
dard monomial theory for classical types), Proc. Indian Acad. Sci., 88A
(1979), 280–362.
[L-M-S2] V. Lakshmibai, C. Musili and C.S. Seshadri, Geometry of G/P-III,
Proc. Indian Acad. Sci. 87 (A) (1978), 93–177.
[Ma] H. Matsumura, Commutative Algebra, Benjamin, New York, 1970.
[M] D. Mumford, Introduction to algebraic geometry, Mimeographed notes,
Harvard University.
[Mu] C. Musili, Postulation formula for Schubert varieties, J. Indian Math.
Soc., 36, 1972, 143–171.
BIBLIOGRAPHY 215

[M-F] D. Mumford, J. Fogarty, Geometric invariant theory (Second edition),


Springer-Verlag, 1982.
[M-R] V. Mehta, A. Ramanathan, Frobenius splitting and cohomology vanish-
ing for Schubert varieties, Annals of Mathematics, 122 (1985), 27–40.
[M-S] C. Musili, C. S. Seshadri, Schubert varieties and the variety of complexes,
Arithmetic and Geometry: Papers dedicated to I. R. Shafarevich, Vol. 2,
Birkhäuser, Boston, 1983.
[N] P.E. Newstead, Introduction to moduli problems and orbit spaces, vol 51,
Tata Institute of Fundamental Research Lecture Notes on Mathematics and
Physics, 1978
[R] G. Reisner, Cohen-Macaulay quotients of polynomial rings, Adv. in Math.,
21 (1976).
[R-R] S. Ramanan, A. Ramanathan, Projective normality of flag varieties and
Schubert varieties, Inv. Math, 79, (1985), 217–224.
[S] J-P. Serre, Algèbre Locale-Multiplicites, Lecture notes in Mathematics 11,
Springer-Verlag, 1965.
[S1] C. S. Seshadri, Geometry of G/P -I (Standard monomial theory for minus-
cule P), C. P. Ramanujam: A Tribute, Springer-Verlag, Published for the
Tata Institute of Fundamental Research, 1978, 207–239.
[S2] C. S. Seshadri, Standard monomial theory and the work of Demazure, Ad-
vanced Studies in Pure Mathematics 1, 1983 (Proceedings of a symposium
on algebraic and analytic varieties), 355–384.
[S3] C. S. Seshadri, Line bundles on Schubert varieties, Bombay Colloquium
on vector bundles (1984), published by Tata Institute, 1987.
[W] H. Weyl, The classical groups, Princeton University Press, 1946.
Notation

GLn (k) = the group of n×n invert-


ible matrices with entries
in k, 4
Gm = the group k \ {0} under mul-
tiplication, 31
M (n, r) = the set all n×r matrices,
2
SLn (k) = the group of n × n ma-
trices of determinant 1, 4
Sr = the permutation group on r
symbols, 14
k = an algebraically closed field, 1

© Springer Science+Business Media Singapore 2016 and Hindustan Book Agency 2015 217
C.S Seshadri, Introduction to the Theory of Standard Monomials,
Texts and Readings in Mathematics 46, DOI 10.1007/978-981-10-1813-8
Index

αth minor of a matrix, 3 a Schubert variety in G/Q, 104


Order complex, 141
Big cell, 9, 109
Borel subgroup, 55, 109 Parabolic subgroup, 56
opposite, 109 Pieri’s formula, 25
Bruhat decomposition, 109 Plücker coordinate, 5
Plücker embedding, 3
Cellular decomposition of Gr,n , 8 Poset, 149
Chain, 37
maximal, 37 Reduced decomposition, 118
Character formula, 89 Reduced representation, 63
Character group, 107 Reduced root system, 108
Chevalley’s multiplicity formula, 129 Right half space, 20
Rooted interval, 139
Defining chain, 61 Root lattice, 110
maximal, 64 roots, 108
minimal, 64 negative, 108
Degeneration of Schubert varieties, positive, 108
32 simple, 108
Deodhar’s lemma, 134
Determinantal ideal, 44 Schubert cell, 5, 6, 58, 112
Determinantal variety, 44 Schubert variety, 9, 59, 112
Diagram, 110 Shelling, 141
Discrete algebra, 33 Shuffling, 14
Dual pair, 42 (or Shuffle permutation), 14
Simple reflection, 108
Facets, 141 Standard diagram, 61
Flag, 6 weakly, 61
full, 6 Standard Monomial, 61
Flag variety, 55 Standard monomial, 12, 21, 42, 63
Graded poset, 139
Tableau, 12
Grassmannian, 2
double standard, 42
Left half space, 19 standard, 12
Lexicographic shellability, 140 Tits system, 115
Torus, 107
Multicone, 94 maximal, 57

Nice indexing, 142 Variety of complexes, 97

Opposite big cell of, 39 Weights


G/Q, 98 dominant, 110
Gr,n , 38 fundamental, 110

© Springer Science+Business Media Singapore 2016 and Hindustan Book Agency 2015 219
C.S Seshadri, Introduction to the Theory of Standard Monomials,
Texts and Readings in Mathematics 46, DOI 10.1007/978-981-10-1813-8
220 INDEX

weights, 108
weight space, 108
Weyl group, 58

Young diagram, 60, 107, 135, 149


standard, 135
Symbols

(B − )u , 109 T (α), 113


(φ, Γ), 110 Tα , 20
B, 55 UQ , 111
B − , 109 Uαmax , 38
B ′ , 55 Uαmin , 39
B u , 109 V r,0 , 2
Bωu , 112 W , 57
C(α), 6 W (φ), 108
C(ω), 58 WQ , 111, 141
D(α), 10 WQi , 58
D(α)◦ , 10 X(T ), 21
Dt (ℓ, m), 44 X(T )R , 108
Dα , 33 X(α), 9
Gα , 109 X(ω), 58, 112
Gr,n , 2 XQi (ω), 124
H(ω), 66 Y (ω), 130
I(T ), 22 Z[Hom(T, GL1 (k))], 90
IG (ω), 111 Z(ω, α), 75, 116, 129
L(α), 75, 129 [X(ω)], 129
Lsα , 91 [x, y], 139
M (n, r)◦ , 2 Λ(φ), 110
Msα , 91 Λr (φ), 110
N (ω), 112 Ω(Q1 , Q2 ), 134
P (α), 75, 130 Φ, 75
Pi , 56 αt , 44
Pℓ,m , 42 αmax , 9
Pα̂ , 143 αmin , 9
Qµ , 100 α′ , α′′ , 42
Qi , 56 ℓ(ω), 63
Qred , 99 ℓ(τ ), 84
R(Q), 111 ℓQ (ω), 118
R(T ), 21 ℓQi (ω), 63
R+ (Q), 111 exp(Hom(T, GL1 (k))), 90
Rdef (Z), 152 0̂, 139
Rdef (τ ), 151 1̂, 139
S(Y, a(i)), 62 P1 (ω, τ ), 132
S, 75, 89, 109, 111 Lie G, 108
S(T, m), 26 ω0 , 108
S(p, q), 14 φ, 108
SQ , 111 φ+ , 108
SPi , 125 φ− , 108

© Springer Science+Business Media Singapore 2016 and Hindustan Book Agency 2015 221
C.S Seshadri, Introduction to the Theory of Standard Monomials,
Texts and Readings in Mathematics 46, DOI 10.1007/978-981-10-1813-8
222 SYMBOLS

τ max , 124
τ min , 124
ζ, 67
e(ω), 58, 111
e(id), 58, 111
k, 1
k{H(τ )}, 146
pα,β , 40
pα,beta , 42
pα , 5
sα , 75, 108
La(i) , 57
Lλ , 90

LHS, 19

RHS, 20
Texts and Readings in Mathematics

1. R. B. Bapat: Linear Algebra and Linear Models (Third Edition)


2. Rajendra Bhatia: Fourier Series (Second Edition)
3. C. Musili: Representations of Finite Groups
4. H. Helson: Linear Algebra (Second Edition)
5. D. Sarason: Complex Function Theory (Second Edition)
6. M. G. Nadkarni: Basic Ergodic Theory (Third Edition)
7. H. Helson: Harmonic Analysis (Second Edition)
8. K. Chandrasekharan: A Course on Integration Theory
9. K. Chandrasekharan: A Course on Topological Groups
10. R. Bhatia (ed.): Analysis, Geometry and Probability
11. K. R. Davidson: C* – Algebras by Example (Reprint)
12. M. Bhattacharjee et al.: Notes on Infinite Permutation Groups
13. V. S. Sunder: Functional Analysis — Spectral Theory
14. V. S. Varadarajan: Algebra in Ancient and Modern Times
15. M. G. Nadkarni: Spectral Theory of Dynamical Systems
16. A. Borel: Semisimple Groups and Riemannian Symmetric Spaces
17. M. Marcolli: Seiberg – Witten Gauge Theory
18. A. Bottcher and S. M. Grudsky: Toeplitz Matrices, Asymptotic
Linear Algebra and Functional Analysis
19. A. R. Rao and P. Bhimasankaram: Linear Algebra (Second Edition)
20. C. Musili: Algebraic Geometry for Beginners
21. A. R. Rajwade: Convex Polyhedra with Regularity Conditions
and Hilbert's Third Problem
22. S. Kumaresan: A Course in Differential Geometry and Lie Groups
23. Stef Tijs: Introduction to Game Theory
24. B. Sury: The Congruence Subgroup Problem
25. R. Bhatia (ed.): Connected at Infinity
26. K. Mukherjea: Differential Calculus in Normed Linear Spaces
(Second Edition)
27. Satya Deo: Algebraic Topology: A Primer (Corrected Reprint)
28. S. Kesavan: Nonlinear Functional Analysis: A First Course
29. S. Szabó: Topics in Factorization of Abelian Groups
30. S. Kumaresan and G. Santhanam: An Expedition to Geometry
31. D. Mumford: Lectures on Curves on an Algebraic Surface (Reprint)
32. J. W. Milnor and J. D. Stasheff: Characteristic Classes (Reprint)
33. K. R. Parthasarathy: Introduction to Probability and Measure
(Corrected Reprint)
34. Amiya Mukherjee: Topics in Differential Topology
35. K. R. Parthasarathy: Mathematical Foundations of Quantum
Mechanics
36. K. B. Athreya and S. N. Lahiri: Measure Theory
37. Terence Tao: Analysis I (Second Edition)
38. Terence Tao: Analysis II (Second Edition)

© Springer Science+Business Media Singapore 2016 and Hindustan Book Agency 2015 223
C.S Seshadri, Introduction to the Theory of Standard Monomials,
Texts and Readings in Mathematics 46, DOI 10.1007/978-981-10-1813-8
224 TEXTS AND READINGS IN MATHEMATICS

39. W. Decker and C. Lossen: Computing in Algebraic Geometry


40. A. Goswami and B. V. Rao: A Course in Applied Stochastic
Processes
41. K. B. Athreya and S. N. Lahiri: Probability Theory
42. A. R. Rajwade and A. K. Bhandari: Surprises and Counterexamples
in Real Function Theory
43. G. H. Golub and C. F. Van Loan: Matrix Computations (Reprint of the
Third Edition)
44. Rajendra Bhatia: Positive Definite Matrices
45. K. R. Parthasarathy: Coding Theorems of Classical and Quantum
Information Theory (Second Edition)
46. C. S. Seshadri: Introduction to the Theory of Standard Monomials
47. Alain Connes and Matilde Marcolli: Noncommutative Geometry,
Quantum Fields and Motives
48. Vivek S. Borkar: Stochastic Approximation: A Dynamical Systems
Viewpoint
49. B. J. Venkatachala: Inequalities: An Approach Through Problems
50. Rajendra Bhatia: Notes on Functional Analysis
51. A. Clebsch (ed.): Jacobi's Lectures on Dynamics
(Second Revised Edition)
52. S. Kesavan: Functional Analysis
53. V. Lakshmibai and Justin Brown: Flag Varieties: An Interplay of
Geometry, Combinatorics, and Representation Theory
54. S. Ramasubramanian: Lectures on Insurance Models
55. Sebastian M. Cioaba and M. Ram Murty: A First Course in Graph Theory and
Combinatorics
56. Bamdad R. Yahaghi: Iranian Mathematics Competitions, 1973-2007
57. Aloke Dey: Incomplete Block Designs
58 R. B. Bapat: Graphs and Matrices
59. Hermann Weyl: Algebraic Theory of Numbers (Reprint)
60. Carl Ludwig Siegel: Transcendental Numbers (Reprint)
61. Steven J. Miller and Ramin Takloo-Bighash: An Invitation to Number Theory
(Reprint)
62. John Milnor: Dynamics in One Complex Variable (Reprint)
63. R. P. Pakshirajan: Probability Theory: A Foundational Course
64. Sharad S. Sane: Combinatorial Techniques
65. Hermann Weyl: The Classical Groups: Their Invariants and Representations
(Reprint)
66. John Milnor: Morse Theory (Reprint)
67. R. Bhatia (ed.): Connected at Infinity II
68. Donald Passman: A Course in Ring Theory
69. Amiya Mukherjee: Atiyah-Singer Index Theorem: An Introduction
70. Fumio Hiai and Dénes Petz: Introduction to Matrix Analysis and Applications

You might also like