You are on page 1of 267

OX FO R D M AT H E M AT I C A L M O N O G R A P H S

Series Editors

J. M. BALL W. T. GOWERS
N. J. HITCHIN L. NIRENBERG
R. PENROSE A. WILES
OX FO R D M AT H E M AT I C A L M O N O G R A P H S

Hirschfeld: Finite projective spaces of three dimensions


Edmunds and Evans: Spectral theory and differential operators
Pressley and Segal: Loop groups, paperback
Evens: Cohomology of groups
Hoffman and Humphreys: Projective representations of the symmetric groups: Q-Functions
and Shifted Tableaux
Amberg, Franciosi, and Giovanni: Products of groups
Gurtin: Thermomechanics of evolving phase boundaries in the plane
Faraut and Koranyi: Analysis on symmetric cones
Shawyer and Watson: Borel’s methods of summability
Lancaster and Rodman: Algebraic Riccati equations
Thévenaz: G-algebras and modular representation theory
Baues: Homotopy type and homology
D’Eath: Black holes: gravitational interactions
Lowen: Approach spaces: the missing link in the topology–uniformity–metric triad
Cong: Topological dynamics of random dynamical systems
Donaldson and Kronheimer: The geometry of four-manifolds, paperback
Woodhouse: Geometric quantization, second edition, paperback
Hirschfeld: Projective geometries over finite fields, second edition
Evans and Kawahigashi: Quantum symmetries of operator algebras
Klingen: Arithmetical similarities: Prime decomposition and finite group theory
Matsuzaki and Taniguchi: Hyperbolic manifolds and Kleinian groups
Macdonald: Symmetric functions and Hall polynomials, second edition, paperback
Catto, Le Bris, and Lions: Mathematical theory of thermodynamic limits: Thomas-Fermi type
models
McDuff and Salamon: Introduction to symplectic topology, paperback
Holschneider: Wavelets: An analysis tool, paperback
Goldman: Complex hyperbolic geometry
Colbourn and Rosa: Triple systems
Kozlov, Maz’ya and Movchan: Asymptotic analysis of fields in multi-structures
Maugin: Nonlinear waves in elastic crystals
Dassios and Kleinman: Low frequency scattering
Ambrosio, Fusco and Pallara: Functions of bounded variation and free discontinuity problems
Slavyanov and Lay: Special functions: A unified theory based on singularities
Joyce: Compact manifolds with special holonomy
Carbone and Semmes: A graphic apology for symmetry and implicitness
Boos: Classical and modern methods in summability
Higson and Roe: Analytic K-homology
Semmes: Some novel types of fractal geometry
Iwaniec and Martin: Geometric function theory and nonlinear analysis
Johnson and Lapidus: The Feynman integral and Feynman ’s operational calculus, paperback
Lyons and Qian: System control and rough paths
Ranicki: Algebraic and geometric surgery
Ehrenpreis: The radon transform
Lennox and Robinson: The theory of infinite soluble groups
Ivanov: The Fourth Janko Group
Huybrechts: Fourier-Mukai transforms in algebraic geometry
Hida: Hilbert modular forms and Iwasawa theory
Boffi and Buchsbaum: Threading homology through algebra
Threading Homology Through
Algebra: Selected Patterns
GIANDOMENICO BOFFI
Università G. d’Annunzio

DAVID A. BUCHSBAUM
Department of Mathematics, Brandeis University

CLARENDON PRESS · OXFORD


2006
3
Great Clarendon Street, Oxford OX2 6DP
Oxford University Press is a department of the University of Oxford.
It furthers the University’s objective of excellence in research, scholarship,
and education by publishing worldwide in
Oxford New York
Auckland Cape Town Dar es Salaam Hong Kong Karachi
Kuala Lumpur Madrid Melbourne Mexico City Nairobi
New Delhi Shanghai Taipei Toronto
With offices in
Argentina Austria Brazil Chile Czech Republic France Greece
Guatemala Hungary Italy Japan Poland Portugal Singapore
South Korea Switzerland Thailand Turkey Ukraine Vietnam
Oxford is a registered trade mark of Oxford University Press
in the UK and in certain other countries
Published in the United States
by Oxford University Press Inc., New York

c Oxford University Press, 2006
The moral rights of the authors have been asserted
Database right Oxford University Press (maker)
First published 2006
All rights reserved. No part of this publication may be reproduced,
stored in a retrieval system, or transmitted, in any form or by any means,
without the prior permission in writing of Oxford University Press,
or as expressly permitted by law, or under terms agreed with the appropriate
reprographics rights organization. Enquiries concerning reproduction
outside the scope of the above should be sent to the Rights Department,
Oxford University Press, at the address above
You must not circulate this book in any other binding or cover
and you must impose the same condition on any acquirer
British Library Cataloguing in Publication Data
Data available
Library of Congress Cataloging in Publication Data
Data available
Typeset by Newgen Imaging Systems (P) Ltd., Chennai, India
Printed in Great Britain
on acid-free paper by
Biddles Ltd., King’s Lynn, Norfolk

ISBN 0–19–852499–4 978–0–19–852499–1

1 3 5 7 9 10 8 6 4 2
A coloro che amo

To Betty, wife and lifelong friend.


Though she can’t identify each tree, she shares with me the delight
of walking through the forest.
This page intentionally left blank
PREFACE

From a little before the middle of the twentieth century, homological methods
have been applied to various parts of algebra (e.g. Lie Algebras, Associative
Algebras, Groups [finite and infinite]). In 1956, the book by H. Cartan and
S. Eilenberg, Homological Algebra [33], achieved a number of very important
results: it gave rise to the new discipline, Homological Algebra, it unified the
existing applications, and it indicated several directions for future study. Since
then, the number of developments and applications has grown beyond counting,
and there has, in some instances, even been enough time to see various methods
threading their way through apparently disparate, unrelated branches of algebra.
What we aim for in this book is to take a few homological themes (Koszul
complexes and their variations, resolutions in general) and show how these affect
the perception of certain problems in selected parts of algebra, as well as their
success in solving a number of them. The expectation is that an educated reader
will see connections between areas that he had not seen before, and will learn
techniques that will help in further research in these areas.
What we include will be discussed shortly in some detail; what we leave out
deserves some mention here. This is not a compendium of homological algebra,
nor is it a text on commutative algebra, combinatorics, or representation the-
ory; although, it makes significant contact with all of these fields. We are not
attempting to provide an encyclopedic work. As a result, we leave out vast areas
of these subjects and only select those parts that offer a coherence from the
point of view we are presenting. Even on that score we can make no claim to
completeness.
Our Chapter I, called “Recollections and Perspectives,” reviews parts of Poly-
nomial Ring and Power Series Ring Theory, Linear Algebra, and Multilinear
Algebra, and ties these with ideas that the reader should be very familiar with.
As the title of the chapter suggests, this is not a compendium of “assumed
known” items, but a presentation from a certain perspective—mainly homolo-
gical. For example, almost everyone knows about divisibility and factoriality; we
give a criterion for factoriality that ties it immediately to a homological inter-
pretation (and one which found significant application in solving a long-open
question in regular local ring theory).
The next three chapters of this book pull together a group of classical results,
all coming from and generalizing the techniques associated with the Koszul com-
plex. Perhaps the major result in Chapter II, on local rings, is the homological
characterization of a regular local ring by means of its global dimension. Section
II.6 includes a proof of the factoriality of regular local rings which is much closer
to the original one, rather than the Kaplansky proof that is frequently quoted.
viii Preface

We have also included a section on multiplicity theory, mainly to carry through


the theme of the Koszul complex, and a section on the Homological Conjectures,
as they provide a good roadmap for still open problems as well as a historical
guide through much of what has been going on in the area this book is sketching.
Chapter III deals with a class of complexes developed with the following aim in
view: to associate a complex to an arbitrary finite presentation matrix of a mod-
ule (the Koszul complex does this for a cyclic module), and to have that complex
play the same role in the proof of the generalized Cohen–Macaulay Theorem that
the Koszul complex plays in the classical case. We have made an explicit connec-
tion, in terms of a chain homotopy, between an older, “fatter” class of complexes,
and a slimmer, more “svelte” class. We have also included a last section in which
we define a generalized multiplicity which has found interesting applications, of
late. Chapter IV applies some of the properties of these complexes to a system-
atic study of finite free resolutions, ending in a “syzygy-theoretic” proof of the
unique factorization theorem (or “factoriality”) in regular local rings.
The last three chapters and the Appendix not only focus on determinantal
ideals and characteristic-free representation theory, but also involve a good deal
of combinatorics. Chapter V employs the homological techniques developed in
the previous part in the study of a number of types of determinantal ideals,
namely Pfaffians and powers of Pfaffians. In Chapter VI we develop the basics
of a characteristic-free representation theory of the general linear group (which
has already made its appearance in earlier chapters). Because of the generality
aspired to, heavy use is made of letter-place methods, an idea used more by
combinatorialists than by commutative algebraists. As some of the proofs require
more detail than is probably helpful for those encountering this material for the
first time, we decided to place these details in a separate Appendix: Appendix A.
Much of the development of this chapter rests heavily on the notion of straight
tableaux introduced by B. Taylor. In Chapter VII we first present a number of
results that immediately follow from this more general theory. Then examples
are given to indicate what further use has been made of it, and in most cases
references are given to detailed proofs. It is in this part of the chapter that we see
the important influence of the work of A. Lascoux in characteristic zero. We give
some of the background to the Hashimoto example of the dependence of the Betti
numbers of determinantal ideals on characteristic. We deal with resolutions of
Weyl modules in general, and skew-hooks in particular, and we make connections
with intertwining numbers, Z-forms, and several other open problems.
The intended readership of this book ranges from third-year and above gradu-
ate students in mathematics, to the accomplished mathematician who may or
may not be in any of the fields touched on, but who would like to see what devel-
opments have taken place in these areas and perhaps launch himself into some
of the open problems suggested. Because of this assumption, we are allowing
ourselves to depend heavily on material that can be found in what we regard as
comprehensive and accessible texts, such as the textbook by D. Eisenbud. We
may at times, though, include a proof of a result here even if it does appear in
such a text, if we think that the method of proof is typical of many of that kind.
CONTENTS

I Recollections and Perspectives 1


I.1 Factorization 1
I.1.1 Factorization domains 1
I.1.2 Polynomial and power series rings 6
I.2 Linear algebra 8
I.2.1 Free modules 8
I.2.2 Projective modules 13
I.2.3 Projective resolutions 17
I.3 Multilinear algebra 21
I.3.1 R[X1 , . . . , Xt ] as a symmetric algebra 22
I.3.2 The divided power algebra 28
I.3.3 The exterior algebra 30
II Local Ring Theory 37
II.1 Koszul complexes 38
II.2 Local rings 43
II.3 Hilbert–Samuel polynomials 46
II.4 Codimension and finitistic global dimension 50
II.5 Regular local rings 54
II.6 Unique factorization 56
II.7 Multiplicity 59
II.8 Intersection multiplicity and the homological conjectures 64
III Generalized Koszul Complexes 69
III.1 A few standard complexes 69
III.1.1 The graded Koszul complex and its “derivatives” 70
III.1.2 Definitions of the hooks and their explicit bases 72
III.2 General setup 80
III.2.1 The fat complexes 82
III.2.2 Slimming down 83
III.3 Families of complexes 85
III.3.1 The “homothety homotopy” 88
III.3.2 Comparison of the fat and slim complexes 91
III.4 Depth-sensitivity of T(q; f ) 94
III.5 Another kind of multiplicity 99
IV Structure Theorems for Finite Free Resolutions 103
IV.1 Some criteria for exactness 104
IV.2 The first structure theorem 110
x Contents

IV.3 Proof of the first structure theorem 115


IV.3.1 Part (a) 115
IV.3.2 Part (b) 118
IV.4 The second structure theorem 119
V Exactness Criteria at Work 127
V.1 Pfaffian ideals 128
V.1.1 Pfaffians 128
V.1.2 Resolution of a certain pfaffian ideal 131
V.1.3 Algebra structures on resolutions 132
V.1.4 Proof of Part 2 of Theorem V.1.8 134
V.2 Powers of pfaffian ideals 136
V.2.1 Intrinsic description of the matrix X 137
V.2.2 Hooks again 138
V.2.3 Some representation theory 139
V.2.4 A counting argument 140
V.2.5 Description of the resolutions 143
V.2.6 Proof of Theorem V.2.4 145
VI Weyl and Schur Modules 149
VI.1 Shape matrices and tableaux 149
VI.1.1 Shape matrices 149
VI.1.2 Tableaux 153
VI.2 Weyl and Schur modules associated to shape matrices 154
VI.3 Letter-place algebra 156
VI.3.1 Positive places and the divided power algebra 156
VI.3.2 Negative places and the exterior algebra 159
VI.3.3 The symmetric algebra (or negative letters and places) 164
VI.3.4 Putting it all together 164
VI.4 Place polarization maps and Capelli identities 165
VI.5 Weyl and Schur maps revisited 167
VI.6 Some kernel elements of Weyl and Schur maps 169
VI.7 Tableaux, straightening, and the straight basis theorem 174
VI.7.1 Tableaux for Weyl and Schur modules 174
VI.7.2 Straightening tableaux 176
VI.7.3 Taylor-made tableaux, or a straight-filling algorithm 181
VI.7.4 Proof of linear independence of straight tableaux 183
VI.7.5 Modifications for Schur modules 186
VI.7.6 Duality 187
VI.8 Weyl–Schur complexes 187
VII Some Applications of Weyl and Schur Modules 193
VII.1 The fundamental exact sequence 193
VII.2 Direct sums and filtrations for skew-shapes 197
VII.3 Resolution of determinantal ideals 199
Contents xi

VII.3.1 The Lascoux resolutions 200


VII.3.2 The submaximal minors 202
VII.3.3 Z-forms 203
VII.4 Arithmetic considerations 206
VII.4.1 Intertwining numbers 206
VII.4.2 Z-forms again 208
VII.5 Resolutions revisited; the Hashimoto counterexample 209
VII.6 Resolutions of Weyl modules 211
VII.6.1 The bar complex 212
VII.6.2 The two-rowed case 215
VII.6.3 A three-rowed example 217
VII.6.4 Resolutions of skew-hooks 225
VII.6.5 Comparison with the Lascoux resolutions 227
A Appendix for Letter-Place Methods 237
A.1 Theorem VI.3.2, Part 1: the double standard tableaux generate 237
A.2 Theorem VI.3.2 Part 2: linear independence of double standard
tableaux 241
A.3 Modifications required for Theorems VI.3.3 and VI.3.4 244
A.4 Modifications required for Theorem VI.8.4 246
References 249
Index 253
This page intentionally left blank
I
RECOLLECTIONS AND PERSPECTIVES

This chapter is neither a collection of results which we assume to be known nor


the place to prove some results probably unknown to the reader, but needed in
the following. Although it resembles a little of both things, it is essentially a selec-
tion of topics, some elementary, some more advanced, which we feel are adequate,
or even necessary, to prepare the ground for the material of the chapters to come.
Since it is almost impossible to tell which “basic” material is truly universally
known, and which is not, we can only assure the reader that those terms in this
chapter which are unfamiliar can be easily found in the book by D. Eisenbud, [41].

I.1 Factorization
In this section, we deal with the basic topic of divisibility. In doing so, we review
a few properties of some rings, which are of importance to us. For more details,
we refer the reader to Reference [87].

I.1.1 Factorization domains


Let R be an integral domain, that is, a commutative ring (with 1) having no
zero divisors. Given a and b in R, we say that a is a divisor of b (written a | b)
if b = ac for some c in R. If a | b and b | a, then b = ua for some unit u, and
a and b are called associates. Being associate is an equivalence relation. a is a
proper divisor of b if a divides b, but is neither a unit, nor an associate of b.
In terms of ideals, a | b means (b) ⊆ (a), u being a unit is equivalent to
(u) = R, a and b associates says (a) = (b), and a properly divides b if and only
if (b) ⊂ (a).
Definition I.1.1 An element c ∈ R is called a greatest common divisor
(gcd) of a and b in R if c | a, c | b, and c is divisible by every d such that d | a
and d | b. An element c ∈ R is called a least common multiple (lcm) of
a and b in R if a | c, b | c, and c divides every d such that a | d and b | d.
Given a and b in R, gcd(a, b) may or may not exist. If it does, it is unique up
to associates. Similarly for lcm(a, b).
Remark I.1.2 If a, b ∈ R −{0}, and lcm(a, b) exists, then also gcd(a, b) exists,
and lcm(a, b) · gcd(a, b) = ab, up to units. If a, b ∈ R − {0}, and gcd(a, b)
exists, lcm(a, b) may not exist (cf. Example I.1.11 later on). However, if gcd(a, b)
2 Recollections and perspectives

exists for all choices of a and b in R − {0}, then lcm(a, b) exists for all choices
of a and b in R − {0} .

If 1 is a greatest common divisor for a and b, we will say that a and b are
coprime.
In terms of ideals, c being a common divisor of a and b means (a, b) ⊆ (c),
while c being a common multiple of a and b means (a) ∩ (b) ⊇ (c). The equality
(a, b) = (c) implies gcd(a, b) = c (but not conversely), while (a) ∩ (b) = (c) is
equivalent to lcm(a, b) = c.

Definition I.1.3 A non-zero, non-invertible element a ∈ R is called irredu-


cible if it does not have any proper divisors. A non-zero, non-invertible element
a ∈ R is called prime if, whenever a | bc, then either a | b or a | c.

In terms of ideals, a is irreducible if and only if (a) is maximal among the


proper principal ideals of R; a is prime if and only if (a) is a prime ideal. If a is
prime, then it is irreducible. But the converse does not hold.

Definition I.1.4 An integral domain R is called a factorization domain if


every non-zero, non-invertible a ∈ R can be expressed as a product of irreducible
elements. A factorization domain is called a unique factorization domain
(UFD) if every factorization into irreducibles is unique up to permutation of
the factors and multiplication of the factors by units.

In terms of ideals, an integral domain R is a factorization domain if and only if


there is no strictly ascending infinite chain of principal ideals in R. In particular,
every principal ideal domain (PID) is a factorization domain: given any ascending
chain of ideals, the union of these ideals is an ideal of the sequence.
In fact, a PID is always a UFD, by part (ii) of the following proposition.

Proposition I.1.5 Let R be a factorization domain. The following are


equivalent.
(i) R is a UFD.
(ii) lcm(a, b) exists for every choice of a, b in R.
(iii) Every irreducible element is prime.
Proof (i) ⇒ (ii) As in Z, one expresses a and b as products of powers (with
non-negative exponents) of suitable irreducibles (the same ones for a and b):
   max{si ,ti }
a = fisi and b = fiti , say. Then lcm(a, b) = fi .
(ii) ⇒ (iii) The gcd exists in R since the lcm does. If c = gcd(a, b), then
cd = gcd(ad, bd) for every d. For, given any two ideals a and b, d(a ∩ b) = da ∩ db;
hence dlcm(a, b) = lcm(ad, bd); using that adbd = gcd(ad, bd)lcm(ad, bd) and
ab = clcm(a, b), we are through.
Let an irreducible element c divide ab, and assume that c  b: we claim that
c | a. Since c is irreducible and c  b, b and c are coprime, that is, 1 = gcd(b, c).
Factorization 3

It follows that a = gcd(ab, ac); since c divides ab by assumption, c must divide


gcd(ab, ac), as claimed.
(iii) ⇒ (i) As in the ring of integers. 2
Corollary I.1.6 If R is a UFD, then hdR (R/(a, b)) ≤ 2 for all a and b in
R. (Here hd stands for homological dimension, sometimes called projective
dimension and denoted by pd.)
Proof If a = 0 = b, the quotient ring is R and the homological dimension is 0.
If a = 0 and b = 0, the quotient ring is R/(b) and we consider the exact
complex of R-modules:
b
0 → K → R → R → R/(b) → 0,
where K stands for the kernel of the map given by multiplication by b. Since R
is a domain, cb = 0 implies c = 0, and K = (0). Hence hdR (R/(b)) ≤ 1.
If both a and b are different from 0, we consider the following exact complex:
(a,b)
0 → K → R2 → R → R/(a, b) → 0,
where K stands for the kernel of the map given by the matrix (a, b). We want
to show that K is free over R, as this will give us our result on the homological
dimension of R/(a, b).
If (a) : b denotes the ideal {r ∈ R | rb ∈ (a)}, clearly K = (a) : b, because
rb ∈ (a) if and only if rb = sa for some (unique) s ∈ R, that is, (−s)a + rb = 0.
As R-modules, (a) : b ∼ = b((a) : b), and obviously b((a) : b) = (a) ∩ (b).
By the previous proposition, (a) ∩ (b) is a principal ideal; hence it is a rank 1
free R-module, and so is K. 2
Because we have not made significant use of homological dimension, we will
put off giving a formal definition of that term here; the reader will find it in the
next section (Definition I.2.25). The crucial fact that we needed in the proof of
the above corollary was just that K is free.
We will see in Chapter II that if R is a noetherian local ring, then R is a UFD
if and only if hdR (R/(a, b)) ≤ 2 for all a and b in R. This will lead to proving
that regular local rings are UFD.
Remark I.1.7 We have noticed in the proof of Proposition I.1.5 that if R is
a UFD, then c = gcd(a, b) implies cd = gcd(ad, bd) for every d. It follows that
gcd(a, b) = 1 and a | bc together imply a | c. In terms of ideals, this means that if
a and b are coprime in R, then b is a non-zero divisor in R/(a), although R/(a)
may no longer be an integral domain. Conversely, if b is a non-zero divisor in
R/(a), then gcd(a, b) = 1; for otherwise a/ gcd(a, b) would kill b in R/(a). This
set up will be generalized in Chapter II by the notion of M -sequence.
Remark I.1.8 The complex
(a,b)
0 → R2 → R → R/(a, b) → 0
4 Recollections and perspectives

is a truncation of the following Koszul complex (to be described in Chapter II):


 

−b 
a (a,b)
0→R → R2 → R → R/(a, b) → 0.
Does the latter complex coincide with the resolution of (a, b), a = 0 = b, described
in the proof of Corollary I.1.6? Recalling the identification K = (a) : b,
 
−b  
im = (s, t) ∈ R2 | s = −br, t = ar for some r ∈ R
a
corresponds to (a) ⊆ (a) : b and we are asking whether (a) = (a) : b. We claim
that equality holds if and only if gcd(a, b) = 1. For, by the previous remark
gcd(a, b) = 1 means that b is a non-zero divisor in R/(a), and so (a) : b vanishes
in R/(a).
When unique factorization of elements does not hold in our integral domain
R, we might relax the condition a bit and ask: what kind of ring allows for the
unique factorization of principal ideals into prime ideals? We do not know the
answer to that, but we can ask for a generalization of principal ideal domains,
namely what kind of ring, R, (besides a PID) may have unique factorization of
ideals into products of prime ideals?
Actually, there is a name for such a ring: Dedekind domain. It turns out
that this condition is equivalent to a combination of other properties, namely
that of being noetherian, normal (or integrally closed), and being of dimension
one. Some of these terms will be discussed in great detail in Chapter II, but we
briefly point out here that one of the many characterizations of a noetherian
ring is that every ideal is finitely generated (another one is that no infinite
strictly ascending chain of ideals can exist). This is certainly true of a PID, so
every PID is noetherian. To say that the dimension of a ring is equal to one turns
out to mean (see Section II.3) that every non-zero prime ideal is maximal, and
the observations immediately preceding Definition I.1.3 imply that in a PID all
prime ideals are indeed maximal. Finally, it is clear that every PID (in fact, every
UFD) is normal, so that we know every PID is a noetherian, normal domain of
dimension one. However, these three properties do not quite characterize a PID;
rather, there is the following theorem.
Theorem I.1.9 For an integral domain R, the following are equivalent.
(i) R is a noetherian normal domain of dimension 1.
(ii) Every proper ideal a of R can be expressed as a product of prime ideals,
in a unique way, up to permutations of the factors.
Proof Cf., for example, Reference [87, chapter 5, section 6, theorem 13, p. 275].
2
So we are led to make the following definition.
Factorization 5

Definition I.1.10 An integral domain is called a Dedekind domain if it


satisfies the equivalent conditions of Theorem I.1.9.
The ring of algebraic integers in any algebraic number √ field is always a Dede-
kind domain. Some very accessible ones are the rings Z+Z n with n a squarefree
element of Z − {0, 1} such that n is not congruent to √ 1 modulo 4 (this latter
condition ensuring that this is the ring of integers in Q( n)).
The family of Dedekind domains properly includes the family of principal ideal
domains, for Dedekind domains may have ideals which are not principal.
√ √
Example I.1.11 Let √ R = Z + Z −5, a = 1 + −5, b = 3. gcd(a, b) exists √ and
equals 1, for if s + t −5 divides both a and b, then its norm N (s + t −5) =
s2 + 5t2 must divide both 6 and 9, hence their gcd 3; but s2 + 5t2 | 3 forces t = 0
and s√= ±1. If a = (a, b) were a √principal ideal, gcd(a, b) = 1 would imply a = R.
But −5 ∈ / a, since otherwise −5 = αa + βb would give 5 = 6N (α) + 9N (β)
and 3 should√divide 5, a contradiction. Finally, notice that lcm(a, b) does not
exist: if s + t −5 were a lcm(a, b), s2 +√5t2 would be divisible by both 6 and 9,
hence by lcm(6, 9) = 18; moreover, s+t√ −5 should √ divide both 36 and 54,√ hence
gcd (36, 54) = 18, since both 6 = (1 + −5)(1 − −5) = 2 · 3 and 3 + 3 −5 are
common multiples; thus s2 + 5t2 = 18, which is impossible.
If one had simply wanted to prove that this √ ring is not
√ a PID, it would have
sufficed to point out that 6 = 2 × 3 = (1 + −5)(1 − −5), show that each of
these factors is irreducible, and conclude that this contradicts UFD, hence PID.
The rather longer discussion above, though, actually produces a non-principal
ideal.
While it may be slightly disappointing that there are Dedekind domains that
are not principal ideal domains, it is a well-known property of Dedekind domains
that all their ideals can be generated by at most two elements (cf., e.g. Refer-
ence [87, chapter 5, section 7, theorem 17, p. 279]). So at least we are not too
far off the mark.
If R is any commutative ring, the collection of prime ideals of R is called the
spectrum of R, written Spec(R). The set of maximal ideals of R is called the
maximal spectrum, and is denoted by Max(R).
By Theorem I.1.9, given a Dedekind domain R, Spec(R) = {0} ∪ Max(R),
Proposition I.1.12 If R is a Dedekind domain such that |Max(R)| < ∞, then
R is a PID.
Proof Let Max(R) = {m1 , m2 , . . . , mt }. For every i = 1, 2, . . . , t there exists
an element ai ∈ mi such that ai ∈ / m2i and ai ∈/ mj , j = i (by the Chinese
Remainder Theorem, Reference [87], chapter 5, section 7, theorem 17, p. 279).
Then (ai ) = mi , and the principality of all maximal ideals implies the principality
of every other ideal (by part (ii) of Theorem I.1.9). 2
If we localize a Dedekind domain at a non-zero prime m, Rm is still Dedekind,
since condition (i) of Theorem I.1.9 is preserved by localization. (We assume the
6 Recollections and perspectives

reader is familiar with the process of forming rings of quotients with respect to
a multiplicative subset. Essentially this is just the “fractions” having arbitrary
elements of the ring on top, and elements of the multiplicative subset as denom-
inators. All the bells and whistles of localization are explained in Reference [41],
section 2.1). In fact, Rm is a PID (by the last proposition), because 0 and mm
are its only prime ideals. If mm = (a) for some a ∈ Rm , then every other ideal of
Rm is of type (an ) for some positive n.
Notice that since Rp is a PID for every p ∈ Spec(R), unique factorization of
elements is locally true for every Dedekind domain.
Local Dedekind domains are known as discrete valuation rings.

I.1.2 Polynomial and power series rings


Given any commutative ring (with 1), say R, a (formal) power series in t
indeterminates over R, t ∈ N−{0}, is a function f : Nt → R. Power series can be
added and multiplied. Addition is simply addition of functions. Multiplication
is defined by (f g)(n1 , . . . , nt ) = mi +li =ni f (m1 , . . . , mt )g(l1 , . . . , lt ). The set of
all power series in t indeterminates over R turns out to be a commutative ring
(with 1) with respect to the indicated operations. The customary notation for
this ring is R[[X1 , . . . , Xt ]], for one identifies f : Nt → R with the formal sum

f (n1 , . . . , nt )X1n1 · · · Xtnt .

In particular, an X n stands for the function f : N →R such that f (n) = an
for every n ∈ N.
Clearly, R[[X1 , . . . , Xt ]] = (R[[X1 , . . . , Xt−1 ]])[[Xt ]].
Given R as above, a polynomial in t indeterminates over R is a power
series f : Nt → R which is zero almost everywhere. The corresponding
symbol f (n1 , . . . , nt )X1n1 · · · Xtnt is usually meant to be restricted to the
(finitely many) non-zero values f (n1 , . . . , nt ), thereby giving a finite formal sum.
Polynomials form a subring of the ring of power series, denoted by R[X1 , . . . , Xt ].
Clearly, R[X1 , . . . , Xt ] = (R[X1 , . . . , Xt−1 ])[Xt ].
Often one writes R[[X]] and R[X] instead of R[[X1 , . . . , Xt ]] and
R[X1 , . . . , Xt ], meaning that X = {X1 , . . . , Xt }.
The following proposition collects some properties valid when |X| = t = 1.

Proposition I.1.13 Let R be a commutative ring (with 1).


(i) f ∈ R[X] is invertible in R[X] if and only if a0 is invertible in
R and all
n
other coefficients are nilpotent in R (as usual, we assume f = i=0 ai X i ;
an element of a ring is nilpotent if some power of it is equal to 0).
(ii) f ∈ R[[X]] is invertiblein R[[X]] if and only if a0 is invertible in R (as

usual, we assume f = i=0 ai X i ).
(iii) R has no zero divisors if and only if R[X] has no zero divisors if and only
if R[[X]] has no zero divisors.
Factorization 7

Proof We only prove (ii), not because it is harder, but because


∞ we need it soon.
If f is invertible in R[[X]], there exists g ∈ R[[X]], g = i=0 bi X i say, such
that f g = 1. Hence
f g = a0 b0 + (a0 b1 + a1 b0 )X + (a0 b2 + a1 b1 + a2 b0 )X 2 + · · · = 1
forces a0 b0 = 1, and a0 is a unit in R.
Conversely, assume that a0 is a unit in R and look for some g as above, such
that f g = 1. The following equalities must be satisfied:
a0 b0 = 1, a0 b1 + a1 b0 = 0, a0 b2 + a1 b1 + a2 b0 = 0, ...
The invertibility of a0 allows us to solve these equations for b0 , b1 , b2 , . . . one after
the other, and we are done. 2
The last statement of Proposition I.1.13 hints at a general question: what
properties of R are inherited by R[X] and R[[X]]?
For instance, if R is a Euclidean domain (i.e. a domain where one has division
with remainder), the domain R[X] need not be Euclidean.
Proposition I.1.14 If R is noetherian, then both R[X] and R[[X]] are
noetherian.
Proof For R[X], this is the Hilbert basis theorem, (cf., for example, Refer-
ence [87], chapter 4, section 1, theorem 1, p. 201). For R[[X]], there is a proof
very much in the spirit of the proof of the Hilbert basis theorem, (cf., e.g.
Reference [87], chapter 7, section 1, theorem 4, p. 138). 2
Corollary I.1.15 If R is noetherian, then both R[X1 , . . . , Xt ] and
R[[X1 , . . . , Xt ]] are noetherian.
When R is a noetherian domain, both R[X1 , . . . , Xt ] and R[[X1 , . . . , Xt ]]
(being noetherian domains) are factorization domains: they cannot contain any
strictly ascending infinite chain of principal ideals. This remark leads to the fol-
lowing question: if R is a UFD, is it true that R[X1 , . . . , Xt ] and R[[X1 , . . . , Xt ]]
are UFD?
Unlike Proposition I.1.14, we cannot give a unique answer: we will prove in
a moment that R[X1 , . . . , Xt ] does inherit the property of being a UFD from R;
but R[[X1 , . . . , Xt ]] may not be a UFD. The first counterexample was given by
P. Samuel in 1961 (see [77]).
Theorem I.1.16 If R is a UFD, then R[X1 , . . . , Xt ] is a UFD.
Proof By induction on t, it suffices to show that R[X] is a UFD. Since we
already know that R[X] is a factorization domain, part (ii) of Proposition I.1.5
says that it is enough to prove that a lcm(f, g) exists for any two polynomials
f and g in R[X].
Let Q denote the field of quotients of R. Since Q is a field, Q[X] is a Euclidean
domain, hence a PID, hence a UFD. So a lcm(f, g) certainly exists in Q[X].
8 Recollections and perspectives

Call it h. Clearly, h can be expressed as h = c(h) · h , where h ∈ R[X] and has


coprime coefficients, while c(h) ∈ Q.
Write f = c(f ) · f  and g = c(g) · g  , where f  and g  are assumed to have
coprime coefficients. Since R is a UFD by hypothesis, a lcm(c(f ), c(g)) exists in R.
Call it c. Then c · h is a lcm(f, g) in R[X], as required. 2

Although a similar theorem does not hold for R[[X1 , . . . , Xt ]], we have the
following partial result.

Proposition I.1.17 If R is a field, then R[[X1 , . . . , Xt ]] is a UFD.

Proof We cannot reduce to the case t = 1. But since we already know that
R[[X1 , . . . , Xt ]] is a factorization domain, it suffices to show that every irreducible
element generates a (principal) prime ideal (cf. part (iii) of Proposition I.1.5).
This can be accomplished by induction on t, and using the statement of the
previous theorem, (cf., e.g. Reference [87], chapter 7, section 1, theorem 6, p. 148).
2

We give another property of K[[X1 , . . . , Xt ]], when K a field.

Proposition I.1.18 If K is a field, then K[[X1 , . . . , Xt ]] is a local ring with


maximal ideal m = (X1 , . . . , Xt ).

Proof The proof of part (ii) of Proposition I.1.13 works word for word in every
ring (R[[X1 , . . . , Xs−1 ]])[[Xs ]]. Since in our case R is a field, the non-units of
K[[X1 , . . . , Xt ]] are the elements with zero constant term. That is, (X1 , . . . , Xt )
consists of all the non-invertible elements of K[[X1 , . . . , Xt ]]. 2

When t = 1, K[[X]] is in fact a discrete valuation ring (= local Dedekind


domain), hence a PID (cf. Proposition I.1.12), for it is not hard to check that
every proper ideal in K[[X]] is a power of m = (X).

I.2 Linear algebra


In this section we deal with linear algebra over a commutative ring, not just
over a field. In doing so, we review some basics of homological algebra. For more
details, we refer the reader to References [15], [33], and [41].

I.2.1 Free modules


Let R be a commutative ring (with 1). An R-module is an immediate generaliz-
ation of a vector space. That is, if K is a field and V a vector space over K, we
know that V is an abelian group, K acts on V , and this action satisfies certain
conditions. One notices that the conditions in no way make use of the fact that
K is a field; thus we may replace K by the commutative ring R, write M for V ,
and get the definition for a module M over the ring R.
Linear algebra 9

The usual definitions of linearly independent subset, linearly dependent sub-


set, generators, submodule, submodule generated by a subset, that are used for
vector spaces apply mutatis mutandis to R-modules. The difference, as we will
see, lies in the fact that our base ring is not in general a field; thus such things
as the existence of a basis for every vector space do not hold true for modules
over arbitrary rings. (Recall that a basis of a module is a linearly independ-
ent subset which generates the module.) Yet the existence of maximal linearly
independent subsets of a module is proved in exactly the same way as is done
for vector spaces. It may be, however, that the empty set is a maximal lin-
early independent subset of a module, but the module is not necessarily the zero
module.
For example, the abelian group Z/(2), considered as a Z-module, has two
elements, but its maximal linearly independent set is the empty set. For {0} is
not independent and {1} is not independent because 2 · 1 = 0.
Thus, while for vector spaces we have the fact that a maximal independent
subset is a basis for (hence generates) the vector space, this is no longer the case
for general modules.

Definition I.2.1 An R-module M is called free if it has a basis.

We note immediately that the zero module is free (its basis is the empty
set). The following result shows that the basis of a free module can have any
cardinality.

Proposition I.2.2 Given any non-empty set I, there is a free R-module with
basis in one-to-one correspondence with I.

Proof Let M be the set {f : I → R | f is zero almost everywhere}. It is


an R-module with respect to the operations (f1 + f2 )(i) = f1 (i) + f2 (i) and
(rf )(i) = rf (i). Clearly M has an R-basis {fi }i∈I , where fi stands for the map
sending i to 1 and all other elements of I to 0. 2

Homomorphisms of R-modules, often called R-maps, are defined as in the


case of vector spaces.
The free module built in the proof of Proposition I.2.2 is canonically
R-isomorphic to ⊕i∈I Ri , where Ri is a copy of the R-module R for every i ∈ I.
The basis of ⊕i∈I Ri corresponding to {fi }i∈I in this isomorphism is called the
canonical basis of ⊕i∈I Ri . When I is finite, say |I| = t, we write Rt instead of
⊕i∈I Ri .

Remark I.2.3 A free R-module M having a finite basis B cannot have an


infinite basis B  . For every element of B can be expressed as a linear combination
of finitely many elements of B  , and if C is the (finite linearly independent) set
of all the elements of B  involved in the expressions of the elements of B, then
every element of M is generated by C, so that C = B  .
10 Recollections and perspectives

Proposition I.2.4 If B = {m1 , . . . , mt } and B  = {m1 , . . . , ms } are two finite


bases of the same free R-module M , then t = s.

Proof Write each mi (respectively, mj ) as an R-linear combination of the


elements of B  (respectively, B):


s 
t
mi = aij mj , mj = bji mi .
j=1 i=1

Call S the t × s matrix (aij ) and T the s × t matrix (bji ). Clearly, ST equals the
t × t identity matrix It , and T S = Is . If t = s, say t > s, consider the square
matrices of order t:
 
  T
S = (S | t − s zero columns) and T = .
t − s zero rows

Their product is still It , but det It = 1, while det(ST) = det S det T = 0:


a contradiction. Similarly for t < s, using T S = Is . 2

We remark that the definition of determinant is the same as for fields, and
that det S det T = det ST is purely formal and does not require that the base
ring be a field.

Definition I.2.5 A free R-module is called finite if all its bases have finitely
many elements. A finite free R-module has rank t if it has a basis consisting of
t elements (hence every basis consists of t elements).

Remark I.2.6 If a free R-module M happens to have a finite system of gen-


erators, then M must be a finite free R-module. Just argue as in Remark I.2.3,
calling B the system of generators and B  any basis.

Proposition I.2.7 Let F be a rank t free R-module with basis B =


{f1 , . . . , ft }. Let S be a t×t matrix with entries in R. Let C = {m1 , . . . , mt } ⊆ F
be defined by
   
m1 f1
 ..   . 
 .  = S  ..  .
mt ft

Then the following are equivalent.

(i) det S is a unit in R.


(ii) C is another basis of F .
(iii) C is a generating system of F .
Linear algebra 11

   
m1 f1
   
Proof (i) ⇒ (ii): S −1  ...  =  ...  shows that C generates F (S −1
mt ft
exists for det S invertible allows the use of the customary formula); independence
is easy.
(ii) ⇒ (iii): Trivial.
(iii) ⇒ (i): Call T the matrix expressing B in terms of C; then T S = It , and
det S is a unit in R. 2

Corollary I.2.8
(i) Every generating system {m1 , . . . , mt } of Rt is a basis.
(ii) Every generating system of Rt has at least t elements.
(iii) If ϕ : Rn → Rm is an R-epimorphism, then n ≥ m.

Proposition I.2.9 Let ϕ be an R-morphism from Rt to Rt , and let S be its


matrix with respect to the canonical basis of Rt . Then ϕ is injective if and only
if det S is a non-zero divisor in R.

Proof First part:det Sa non-zero


 divisor ⇒ ϕ injective.
 
x1 0 x1
     
If not, S  ...  =  ...  for some non-zero  ... , x1 = 0 say. Then
xt 0 xt
   
x1 0
   
S  ... | It2 · · · Itt  =  ... | S 2 · · · S t , where Iti stands for the i-th column
xt 0
of It , and similarly for S i . Thus det S · x1 = 0, a contradiction.
Second part: ϕ injective ⇒ det S a non-zero divisor.
It suffices to show that if det S is a zero divisor, then ϕ is not injective, that is,
the columns of S are not an independent system in Rt . If t = 1, this is obvious.
If t ≥ 2, the statement is a corollary of the following more general result.
Let m1 , . . . , ms be elements of Rt (t ≥ 2). If a ∈ R − {0} kills all the maximal
minors of the matrix (m1 · · · ms ), then {m1 , . . . , ms } is not an independent
system.
The proof is by induction on the number s of t-tuples.
If s = 1, then am1 = 0 with a = 0 prevents m1 from being independent.
We now assume that t ≥ s > 1. If a also kills all maximal minors of the matrix
(m2 · · · ms ), then {m2 , . . . , ms } is not an independent system (by induction
hypothesis) and a fortiori {m1 , . . . , ms } is not either.
If a does not kill all maximal minors of (m2 · · · ms ), let b be one of those
minors such that ab = 0; let us say that b is given by the last s − 1 rows of the
matrix (m2 · · · ms ). We now use the assumption that a kills all maximal minors
of (m1 · · · ms ).
12 Recollections and perspectives

Let T denote the t×t matrix ( It−s 0 | m1 · · · ms ). Since det T equals a maximal
minor of (m1 · · · ms ), a det T = 0. If T denotes the companion matrix of T ,
and Ti is the i-th column of T, then T T = det T · It implies:
 
1
 .. 
 . 0 
 
   1 
 
    
T T1 · · · Tt−s aTt−s+1 Tt−s+2 · · · Tt = det T   a ,

 1 
 
 .. 
 0 . 
1
with a in position (t − s + 1, t − s + 1). It follows (since a det T = 0):
 
0
 .. 
 . 
 
 0 
 ∗   
0 = T aTt−s+1 = T  
 ab  , with ab in row t − s + 1.
 ∗ 
 
 . 
 .. 

Hence 0 = abm1 + (∗m2 ) + · · · + (∗ms ), so that (since ab = 0) {m1 , . . . , ms }
cannot be an independent system.
Finally, we consider the case s > t, that is, (m1 · · · ms ) is a t × s matrix
with t < s. In that case, its maximal minors are of order t; if a kills all the
maximal minors, in particular it kills det(m1 · · · mt ). This implies (case s = t of
the induction hypothesis) that {m1 , . . . , mt } cannot be an independent system;
a fortiori, {m1 , . . . , ms } cannot be either.
This concludes the proof of the more general result on Rt , as well as of the
proposition. 2
Corollary I.2.10
(i) If ϕ : Rn → Rm is an R-monomorphism, then n ≤ m.
(ii) A non-zero ideal a of R is a finite free R-module if and only if it is
principal, and generated by a non-zero divisor.
ϕ
Proof (i) If n > m, then there would be a monomorphism Rn → Rm →
 
Rm ⊕ Rn−m = Rn associated with the n × n matrix S = (n−m zero S
rows) ,
where S is the matrix of ϕ with respect to the canonical bases. But det S = 0
contradicts the proposition.
(ii) If a is finite R-free, say a ∼
= Rt , the inclusion a → R gives a monomorphism
R → R, and t ≤ 1 by part (i). So a = Ra = (a) for some independent (that is,
t

non-zero divisor) a ∈ R. The converse is obvious. 2


Linear algebra 13

I.2.2 Projective modules


Every R-module M is the quotient of a free R-module F . For if {mi }i∈I is a
generating system of M (if necessary, the generating system may consist of all
the elements of M ), then we call F the free module of Proposition I.2.2. If {fi }i∈I
is the basis of F defined in that proposition, the R-epimorphism ϕ : F → M
sending fi to mi for every i does the job.
If N denotes the kernel of ϕ, then there exists another R-epimorphism ψ :
E → N with E an R-free module. And one gets the following exact complex:
ψ ϕ
(∗) E → F → M → 0,
which is called a free presentation of M (0 stands for the zero module and
M → 0 is the zero map).
We recall that complex means, wherever you have two consecutive arrows,
the image of the left arrow is included in the kernel of the right arrow. Exact
complex means that the inclusion is always an equality.
If |I| < ∞ (i.e. M is finitely generated), the above F is a finite free R-module,
but E need not be finite. Yet in some cases (for instance when R is noetherian),
E does have a finite basis, and M is said to be finitely presented. Then ψ can
be expressed by a matrix (relative to some fixed bases of E and F ), carrying
information on M = coker(ψ).
Let us go back to the R-epimorphism ϕ : F → M and consider the exact
complex (a short exact sequence is what it is generally called):
0 → ker(ϕ) → F → M → 0.
Does it imply that F ∼
= M ⊕ ker(ϕ)?
More generally, does an exact complex of R-modules
β
0 → M  → M → M  → 0
α
(∗∗)
imply that M ∼ = M  ⊕ M  ?
It is clear that the answer cannot be positive, in general (just think of the
exact complex of Z-modules
α β
0 → Z → Z → Z/(n) → 0,
where α is multiplication by the positive integer n and β is the canonical
projection).
Definition I.2.11 The exact complex (∗∗) is called split if it implies M ∼ =
M  ⊕ M  by means of an isomorphism ϕ : M → M  ⊕ M  such that ϕ−1 | M 
equals α and the composite
M → M  ⊕ M  → M 
ϕ pr2

equals β.
Conditions for being split are easily proven.
14 Recollections and perspectives

Proposition I.2.12 The following are equivalent for the exact sequence (∗∗).

(i) (∗∗) is split.


(ii) There exists an R-map γ : M → M  such that γ ◦ α = idM  .
(iii) There exists an R-map δ : M  → M such that β ◦ δ = idM  .
β
Sometimes one says that an exact complex M → M  → 0 is split, meaning
that there exists δ : M  → M such that β ◦ δ = idM  . Similarly for an exact
α
complex 0 → M  → M .
If the module M  of (∗∗) happens to be R-free, condition (iii) of
Proposition I.2.12 is automatically satisfied, and (∗∗) is split. For if {fi }i∈I is a
basis of M  , it suffices to choose one mi ∈ β −1 (fi ) for every i and define δ by
means of δ(fi ) = mi for every i.
The above is just an instance of an important property satisfied by free
modules.
π
Proposition I.2.13 Given any exact complex M → N → 0 of R-modules, and
any R-map ϕ : F → N with F free, there exists an R-map ψ : F → M (called a
lifting of ϕ) such that ϕ = π ◦ ψ.
Proof Fix a basis {fi }i∈I of F and choose one mi ∈ π −1 (ϕ(fi )) for every i.
Then define ψ by means of ψ(fi ) = mi for every i. 2
The proposition leads to a well-known generalization of free modules.
Definition I.2.14 An R-module P is called projective if whenever we are
π
given an exact complex M → N → 0 of R-modules and an R-map ϕ : P → N ,
there exists an R-map ψ : P → M such that ϕ = π ◦ ψ.
Proposition I.2.15 Given an R-module P , the following are equivalent.
(i) P is projective.
π
(ii) Every exact complex M → P → 0 is split.
(iii) P is a direct summand of a free R-module.
π
(iv) Every exact complex M → N → 0 induces an exact complex
π◦
HomR (P, M ) −→ HomR (P, N ) → 0,
where the left arrow sends f to π ◦ f .
Proof We just show (iii) ⇒ (iv). Let F = P ⊕ Q with F free. Then
π◦
HomR (F, M ) −→ HomR (F, N ) → 0
is exact, thanks to Proposition I.2.13, and we are done because
HomR (P ⊕ Q, X) = HomR (P, X) ⊕ HomR (Q, X)
for every R-module X. 2
Linear algebra 15

The family of projective R-modules does not always properly include the
family of free R-modules.

Lemma I.2.16 (Nakayama’s lemma) Let M be a finitely generated


R-module and a an ideal of R such that a ⊆ rad(R). Then
(i) if aM = M , then M = 0;
(ii) if {m1 , . . . , mt } ⊆ M induces a generating system {m1 , . . . , mt } of the
R/a-module M/aM , then {m1 , . . . , mt } is a generating system of M .
Proof Cf., for example, Reference [41, corollary 4.8, p. 124] 2

Proposition I.2.17 Let (R, m) be a local ring (not necessarily noetherian) and
let M = 0 be a finitely generated R-module. If M is R-projective, then M is
R-free.

Proof Part (i) of the lemma implies that M/mM (= M ⊗R R/m) is a non-zero
finitely generated R/m-vector space, thus it has a basis, say {m1 , . . . , mt }. By
part (ii) of the lemma, it follows that {m1 , . . . , mt } is a generating system of M ,
hence there exists an epimorphism ϕ : F → M with F a rank t free module.
Since M is projective, ϕ splits and F = M ⊕ N for a suitable N . It follows that
F/mF = M/mM ⊕ N/mN as R/m-vector spaces. But F/mF and M/mM have
both dimension t, so that N/mN must have dimension 0, that is, N = mN and
N = 0 (again part (i) of the lemma). 2

In the proof of Proposition I.2.17, we indicated that M/mM = M ⊗R R/m


without explaining the symbol, “⊗R ,” known as the tensor product. We will
assume that this operation is well known to the reader; if in doubt, a straight-
forward treatment may be found in Reference [41] sections 2.2 and A.2.2. We
also point out here that in Proposition I.2.18 we will write Mm = M ⊗R Rm ,
thereby indicating that the localization of a module is the same as taking the
tensor product of the module with the ring of quotients of the ring. These are
facts that we will use often in what follows.
• In connection with the use of tensor products in this book, we offer a caveat
to the reader. While we often try to indicate the ring over which a given tensor
product is taken, especially when a certain ambiguity exists, there are many times
that the base ring is omitted. We believe that when this occurs, the context is
such as to leave no question in the reader’s mind what the base ring is.
There are other cases when projective R-modules and free R-modules coin-
cide, for instance when R is a PID (see Corollary I.2.23 below) and when
R = K[X1 , . . . , Xn ], K a field (Quillen-Suslin Theorem).
A case in which the family of projective R-modules properly contains the
family of free R-modules is given by R a Dedekind domain: see Remark I.2.21
below.
16 Recollections and perspectives

The property of being projective is a local property, in the sense of the following
proposition.
Proposition I.2.18 Let M be a finitely presented R-module. M is projective
if and only if Mm is Rm -free for every m ∈ Max(R).
Proof The only if part follows from Proposition I.2.17, once we remark that
Mm = M ⊗R Rm R-projective implies Mm is Rm -projective (this comes from
Proposition I.2.15 (iii) and the fact that ⊗R commutes with ⊕).
The if part uses Proposition I.2.15 (iv), coupled with the remark that a map is
onto if and only if its localizations at all maximal ideals are onto. The assumption
that M is finitely presented ensures that
Rm ⊗R HomR (M, N ) ∼
= HomRm (Mm , Nm )
for every R-module N . 2
We now briefly turn our attention to a special class of rings.
Definition I.2.19 A nontrivial commutative ring (with 1) R is called hered-
itary if every ideal of R is a projective R-module.
Clearly every PID is hereditary, because of Corollary I.2.10 (ii). Another
example is given by Dedekind domains, due to the following result.
Proposition I.2.20 Every non-zero ideal a of a Dedekind domain R is a
projective R-module.
Proof Since R is noetherian, a is finitely presented as an R-module. Hence
by the last proposition, a is R-projective if and only if am is Rm -free for every
m ∈ Max(R).
Now given m ∈ Max(R), am ⊆ Rm satisfies am = (an ), where (a) is
the maximal ideal of Rm and n is a positive integer (recall the observa-
tions after Proposition I.1.12). Being a principal ideal, am is a rank 1 free
Rm -module. 2
Remark I.2.21 Dedekind domains are a more interesting example of hered-
itary rings than principal ideal domains, because in general not all ideals of
a Dedekind domain are free (if they were, they should all be principal, by
Remark I.2.6 and Corollary I.2.10 (ii)).
Theorem I.2.22 If R is hereditary, then every submodule of a free R-module
is R-isomorphic to a direct sum of ideals of R.
Proof Cf., for example, Reference [33], chapter I, theorem 5.3, p. 13. 2
Corollary I.2.23 If R is a PID, then every submodule of a free R-module is
free; in particular, every projective R-module P is free.
Corollary I.2.24 R is hereditary if and only if every submodule of a projective
R-module is projective.
Linear algebra 17

Proof If part: since R is R-projective (being R-free of rank 1), every


submodule of R (= every ideal of R) is projective.
Only if part: every R-projective P is a direct summand of some free R-module;
by the theorem, every submodule M of P is a direct sum of ideals, hence of pro-
jective modules (by definition of hereditary ring); but a direct sum of projective
modules is obviously projective. 2

I.2.3 Projective resolutions


We push the analysis of non-free modules a little further.
Given any R-module M , with a generating system {mi }i∈I , we have already
constructed a free presentation of M , that is, an exact complex
ϕ1 ϕ0
F1 → F0 → M → 0,
with each Fj a free R-module. Since ker(ϕ1 ), too, is a quotient of some free
R-module F2 , a longer exact complex is obtained:
ϕ2 ϕ1 ϕ0
F2 → F1 → F0 → M → 0.
Repeating the argument, one gets an a priori infinite exact complex
ϕn+1 ϕn ϕ2 ϕ1 ϕ0
(∗) · · · → Fn → · · · → F1 → F0 → M → 0
of free R-modules (except M ), which is called a free resolution of M .
Sometimes one says that the free resolution of M is just the exact complex
ϕn+1 ϕn ϕ2 ϕ1
· · · → Fn → · · · → F1 → F0
and that M = coker(ϕ1 ).
By the same token, we may also consider projective resolutions of a module,
M , namely, an exact sequence
dn+1 d d d
(∗∗) · · · → Pn →
n
· · · →2 P1 →1 P0 → M → 0
where each module Pi is projective. Since free modules are projective, and we
have already shown the existence of free resolutions, the existence of projective
resolutions is assured.
If for some non-negative integer n we have Fn (Pn ) = 0 and Fn+t (Pn+t ) = 0
for every t > 0, we say that M has a finite free (projective) resolution of
length n.
A famous theorem due to D. Hilbert states that if R = K[x1 , . . . , xt ], K a field,
then every finitely generated R-module has a finite free resolution of length less
than or equal to t. (Note that in the case considered by Hilbert, each of the
modules occurring in the resolution is finitely generated, since R is noetherian.)
We can now write down the formal definition of homological dimension that
we promised at the end of the proof of Corollary I.1.6.
18 Recollections and perspectives

Definition I.2.25 A module, M , has homological (free) dimension less


than or equal to n if it has a projective (free) resolution of length equal to n.
Otherwise it is said to have infinite homological (free) dimension.
Since every free module is projective, if the R-module M has a finite free
resolution of length n, then obviously its homological dimension hdR M cannot
be larger than n. However, it is quite possible for the free dimension of a module
to exceed its homological dimension, as the module R/(a, b) of Example I.1.11
easily shows.
Let M and N be two R-modules. Take a (possibly infinite) projective
resolution of M , such as (∗∗) above, and tensor it by N :
dn+1 ⊗1 d ⊗1 2 d ⊗1 1 d ⊗1
· · · −→ Pn ⊗R N −→
n
· · · −→ P1 ⊗R N −→ P0 ⊗R N → M ⊗R N → 0
(1 stands for the identity map on N ). We still have exactness at M ⊗R N (for
⊗R preserves surjectivity), but the truncated complex
dn+1 ⊗1 d ⊗12 d ⊗1 1 d ⊗1
· · · −→ Pn ⊗R N −→
n
· · · −→ P1 ⊗R N −→ P0 ⊗R N
may have nontrivial homology.
Definition I.2.26 For every n ≥ 0, the n-th torsion module, TorR
n (M, N ),
is defined as follows:

 coker(d1 ⊗ 1) if n = 0
TorR (M, N ) =
n
 ker(dn ⊗1)
im(dn+1 ⊗1) if n ≥ 1.

0 (M, N ) = M ⊗R N .)
(Notice that TorR
It is well known that the definition of TorR n (M, N ) does not depend on the
choice of the projective resolution of M . Furthermore, if one picks a projective
resolution of N and tensors it by M , the resulting truncated complex yields the
same homology modules. That is, TorR R
n (M, N ) = Torn (N, M ).
Clearly, Torn (M, N ) = 0 for every n ≥ 1, whenever either M or N is
R

projective.
As one might expect, Tor has some connection with the notion of torsion.
If N is an R-module, an element a ∈ N is said to be a torsion element
if there is a non-zero divisor r ∈ R such that ra = 0. The subset Ntor =
{a ∈ N | a is a torsion element} is clearly a submodule of N , and is called the
torsion submodule of N . N is called torsion-free if Ntor = 0; it is called a
torsion module if N = Ntor .
Example I.2.27 Let R be a commutative ring and r ∈ R a non-zero divisor.
Then
r
0 → R → R → R/(r) → 0
Linear algebra 19

is a projective (in fact free) resolution of R/(r), and for every R-module N , we
have that

1 (R/(r), N ) = {elements of N killed by r} .


TorR

1 (R/(r), N ) = 0 for every non-zero divisor r ∈ R, then N is a torsion-free


If TorR
R-module. In particular, N free implies N torsion-free (by the remark coming
immediately before this example).

It is well known that Tor “restores exactness” in the following sense. Given a
short exact sequence of R-modules:
0 → M  → M → M  → 0,
tensoring by N may destroy exactness on the left. But one can prove that there
is a long exact sequence:
...
 
→ TorR
n (M , N ) → Torn (M, N ) → Torn (M , N ) →
R R

...
 
→ TorR
1 (M , N ) → Tor1 (M, N ) → Tor1 (M , N ) →
R R

M  ⊗R N → M ⊗R N → M  ⊗R N → 0.
Proposition I.2.28 Let R be a PID and N a finitely generated R-module: N
is torsion free if and only if N is free.

Proof The if part is in the example above. Let us prove the only if part.
Thanks to Corollary I.2.23, Proposition I.2.18, and the fact that localization
preserves finite generation and torsion-freeness, we may assume that R is local,
hence a discrete valuation ring with maximal ideal (a), say.
If the R/(a)-vector space N ⊗R R/(a) has dimension t, by Lemma I.2.16 (ii)
we can find a rank t free R-module F such that
0→K→F →N →0
is exact (K simply stands for the kernel of the R-surjection F → N ). Tensoring
by R/(a), one gets the long exact sequence:

1 (R/(a), N ) → K ⊗R R/(a) → F ⊗R R/(a) → N ⊗R R/(a) → 0.


· · · → TorR
But the indicated Tor1 is zero (because N is torsion-free and the last example
applies), and the two vector spaces involving F and N have equal dimensions by
construction. Thus K ⊗R R/(a) = 0, that is, K/(a)K = 0 and by Lemma I.2.16
(i), K = 0 and F ∼ = N as wished. 2

In the last proposition, the assumption that N is finitely generated cannot be


removed: the Z-module Q is torsion-free, but not free.
20 Recollections and perspectives

Corollary I.2.29 Let R be a commutative ring, and N an R-module. Then


N/Ntor is torsion-free. Thus, if R is a PID and N is finitely generated, then N
is the direct sum of a free module and a torsion module.

Proof If r ∈ R is a non-zero divisor, and a ∈ N/Ntor is such that ra = 0, then


ra ∈ Ntor . Hence there exists a non-zero divisor s ∈ R such that 0 = s(ra) =
(sr)a. Since s, r are non-zero divisors, it follows a ∈ Ntor , that is, a = 0 which
proves that N/Ntor is torsion-free. But by the last proposition, if R is a PID,
N/Ntor is R-free if it is torsion-free.
Since N/Ntor is R-free, the short exact sequence
0 → Ntor → N → N/Ntor → 0
splits, and N = Ntor ⊕ N/Ntor . 2

Another example of “exactness restoration” is provided by Ext.


Let M and N be two R-modules. Taken a (possibly infinite) projective
resolution of M , such as (∗∗) above, the following complex is obtained:
1 ◦d 2 ◦d
0 → HomR (M, N ) −→ HomR (P0 , N ) −→ HomR (P1 , N ) −→ ··· ,
d
where ◦dn sends a map Pn−1 → N to the composite Pn −→ n
Pn−1 → N . We
still have exactness at HomR (M, N ), just because P0 projects onto M , but the
truncated complex
1◦d 2 ◦d
0 → HomR (P0 , N ) −→ HomR (P1 , N ) −→ ··· ,
may have nontrivial homology in dimension zero as well as elsewhere.

Definition I.2.30 For every n ≥ 0, the n-th extension module ExtnR (M, N )
is defined as follows:

 ker(◦d1 ) if n = 0
n
ExtR (M, N ) =
 ker(◦dn+1 )
im(◦dn ) if n ≥ 1.

(Notice that since Ext0R (M, N ) is the kernel of the map ◦d1 , and since
1 ◦d
0 → HomR (M, N ) → HomR (P0 , N ) −→ HomR (P1 , N )

is exact, we have Ext0R (M, N ) = HomR (M, N ).)

As in the case of Tor, one can check that the definition of ExtnR (M, N ) does
not depend on the choice of the projective resolution of M . But if one wants to
get the same homology modules starting with a resolution of N , it is necessary
to pick an injective resolution of N and apply HomR (M, ).
It is clear that ExtnR (M, N ) = 0 for every n ≥ 1, whenever M is projective.
Multilinear algebra 21

Moreover, one can prove that given a short exact sequence of R-modules:
β
0 → M  → M → M  → 0,
α

there exists a long exact sequence (note the reverse arrows):

HomR (M  , N ) ← HomR (M, N ) ← HomR (M  , N ) ← 0


← Ext1R (M  , N ) ← Ext1R (M, N ) ← Ext1R (M  , N ) ←
...
· · · ← ExtnR (M  , N ) ← ExtnR (M, N ) ← ExtnR (M  , N ) ←

By definition, an R-module L is called an extension of M by N if there exists


a short exact sequence

0 → N → L → M → 0.

If the sequence is split, the extension is trivial, because L = N ⊕ M .


As the reader may have guessed, these extensions may be turned into a
group which is isomorphic to Ext1R (M, N ), with the zero element being the split
extension. Thus Ext1R (M, N ) = 0 if and only if all extensions of M by N are
trivial.
If part: Any short exact sequence 0 → K → F → M → 0 with F free is split
by assumption; hence M is projective and ExtnR (M, N ) = 0 for every n ≥ 1, as
already observed.
Only if part: Apply HomR ( , N ) to every given short exact sequence
α β
0 → N → L → M → 0.

Since Ext1R (M, N ) = 0 by assumption, the long exact sequence of Ext yields the
surjectivity of
◦α
HomR (L, N ) → HomR (N, N ).

In particular, there exists ϕ : L → N such that ϕ ◦ α = idN , and we are done by


Proposition I.2.12 (ii).

I.3 Multilinear algebra


In this section, we deal with some algebras to be used extensively in the future,
namely symmetric, divided power, and exterior algebras. Our approach here is via
the structure of Hopf algebra, which involves the notions of algebra, coalgebra,
multiplication, comultiplication, and antipode map, among others. We do this in
order to make clear the relationship between the symmetric and divided power
algebras, and to bring out their many useful properites. For more details, we
refer the reader to References [15], [16], and [41].
22 Recollections and perspectives

I.3.1 R[X1 , . . . , Xt ] as a symmetric algebra


Let R be a commutative ring (with 1), and let S denote the polynomial ring
R[X1 , . . . , Xt ]. S is an R-module with respect to the usual addition of polyno-
mials and to their multiplication by elements of R. Moreover, multiplication of
general elements of S defines a map mS : S × S → S which is R-bilinear. Hence
one has an R-morphism S ⊗R S → S, still denoted by mS . If one further con-
siders the R-map uS : R → S defined by means of uS (1) = 1, the fact that
multiplication of polynomials is associative and has a neutral element implies
that S is an R-algebra, in the sense of the following definition.

Definition I.3.1 Given a commutative ring R (with 1), an R-algebra is an


R-module A endowed with two R-morphisms
mA : A ⊗R A → A (multiplication), uA : R → A (unit)
such that the following diagrams are commutative:
m⊗1
A⊗A⊗A → A⊗A

1⊗m↓ ↓m
m
A⊗A → A

1⊗u u⊗1
A⊗R → A⊗A R⊗A → A⊗A


=↓ ↓m ∼
=↓ ↓m

1 1
A → A A → A
(we have omitted subscripts, and 1 stands for the identity on A).

One should remark that an R-algebra A is always a ring, multiplication being


given by the composite map mA ◦ χ, where χ stands for the canonical bilinear
map A × A → A ⊗R A, and 1 being provided by uA (1R ). It is a commutative
ring if the following diagram commutes:
τ
A ⊗ A→A ⊗ A

m↓ ↓m

1
A → A
(here τ is the R-linear map sending a1 ⊗ a2 to a2 ⊗ a1 ).
By an ideal (right, left, two-sided) of the R-algebra A we always mean
an ideal (right, left, two-sided) of the ring A.
Multilinear algebra 23

Given two R-algebras A and B, an R-algebra homomorphism ϕ : A → B


is an R-map also compatible with the ring structures of A and B.
An important feature of a polynomial is its degree (total degree, that is). If for
every i ∈ N, Si stands for the R-linear span of all monomials of total degree i,
S decomposes (as an R-module) as the direct sum ⊕i∈N Si . Moreover, the product
of f ∈ Si and g ∈ Sj belongs to Si+j . This means that S is a graded algebra, in
the sense of the following definition.
Definition I.3.2 An R-algebra A is called graded if it decomposes (as an
R-module) as ⊕i∈N Ai , in such a way that Ai Aj ⊆ Ai+j for every i and j. The
(non-zero) elements of Ai are called homogeneous elements of degree i.
Let now a be an ideal of S. Clearly the ring S/a is again an R-algebra. But
S/a inherits the graduation of S if and only if a has a system of generators which
are homogeneous.
Remark I.3.3 Let A be any graded R-algebra, a a two-sided ideal of A. The
R-module A/a always inherits from A the structure of an R-algebra. But the
R-algebra A/a inherits the graduation of A if and only if a has a system of
generators which are homogeneous.
We now turn to a significant property of the R-algebra S. Let F be
the rank t free R-module corresponding to the set I = {x1 , . . . , xt } (recall
Proposition I.2.2). Denote by {fxi } the canonical basis of F and define the
R-map ϕ : F → S by means of ϕ(fxi ) = xi . Clearly, ϕ(F ) is a set of (commuting)
generators for the R-algebra S.
Proposition I.3.4 For every R-morphism ψ : F → A, with A an R-algebra,
such that the elements of ψ(F ) commute in A, there exists a unique R-algebra
morphism χ : S → A verifying ψ = χ ◦ ϕ.
Proof If we define
  
χ c(s1 , . . . , st )xs11 · · · xst t = c(s1 , . . . , st )ψ(fx1 )s1 · · · ψ(fxt )st ,

everything works. 2
The last proposition says that our polynomial ring S identifies with the
symmetric algebra S(F ), in the sense of the following definition.
Definition I.3.5 Let M be an R-module. We call a symmetric algebra
on M any pair (S(M ), ϕ) satisfying the following requirements: S(M ) is an
R-algebra, ϕ : M → S(M ) is an R-morphism, the elements of ϕ(M ) commute
in S(M ), and for every R-map ψ : M → A, with A an R-algebra, such that
the elements of ψ(M ) commute in A, there exixts a unique R-algebra morphism
χ : S(M ) → A verifying ψ = χ ◦ ϕ.
It is clear that, if symmetric algebras on M do exist, there is a unique
isomorphism from one to the other which is the identity on M .
24 Recollections and perspectives

Theorem I.3.6 Every M admits a symmetric algebra.

We sketch a proof of the theorem.


The idea is to build first an R-algebra T (M ) and an R-map σ : M → T (M )
with the following properties:
(1) for every R-map ψ : M → A, with A an R-algebra, there exists a unique
R-algebra morphism ρ : T (M ) → A verifying ψ = χ ◦ σ;
(2) the elements of σ(M ) generate T (M ).
Then one defines S(M ) as the quotient T (M )/a, where a is the two-sided ideal
generated by all the elements σ(m1 )σ(m2 ) − σ(m2 )σ(m1 ) with m1 and m2 in
M . The map ϕ is defined as the composite:
σ
M → T (M ) → T (M )/a.

Clearly, the elements of ϕ(M ) commute in S(M ). If ψ : M → A is given, such


that the elements of ψ(M ) commute in A, the map ρ : T (M ) → A verifying
ρ ◦ σ = ψ vanishes on a, and we call χ the map T (M )/a → A induced by ρ.
Finally, since T (M ) is generated by the elements of σ(M ), ϕ(M ) generates S(M )
and the uniqueness of χ follows.
A pair (T (M ), σ), necessarily unique, up to isomorphism, can simply be
obtained by taking the R-module ⊕i∈N M ⊗i and by imposing that every product
(m1 ⊗ · · · ⊗ mi )(m1 ⊗ · · · ⊗ mj ) be equal to m1 ⊗ · · · ⊗ mi ⊗ m1 ⊗ · · · ⊗ mj . (We
mean M ⊗0 = R.) The map σ identifies M with the summand M ⊗1 .
One should remark that T (M ) (called the tensor algebra on M ) is graded by
construction, and that the ideal a is generated by homogeneous elements. Hence
S(M ) inherits the graduation of T (M ). One usually writes S(M ) = ⊕i∈N Si (M ).
In particular, ϕ identifies M with S1 (M ), and the elements of S1 (M ) generate
S(M ). Thus S(M ) is commutative.
The construction of S(M ) is functorial [meaning that every R-map M1 → M2
induces an R-algebra morphism S(M1 ) → S(M2 )] and commutes with base
change [meaning that S(M ⊗R R ) = S(M ) ⊗R R for every ring homomorphism
R → R ]. These properties follow from those of the tensor algebra, that is,
of T (M ).
If M is R-free of finite rank, with basis {m1 , . . . , mt } say, then S(M ) is R-free,
and a basis for Si (M ) is given by all distinct elements ms11 · · · mst t such that
s1 + · · · + st = i (not surprisingly!).
Our polynomial ring S, being a symmetric algebra, is also a graded coalgebra,
and in fact a graded Hopf algebra.

Definition I.3.7 Given a commutative ring R (with 1), an R-coalgebra is


an R-module A endowed with two R-morphisms

cA : A → A ⊗R A (comultiplication), εA : A → R (counit)
Multilinear algebra 25

such that the following diagrams commute:


c⊗1
A⊗A⊗A ←− A ⊗ A

1⊗c↑ ↑c
c
A⊗A ←− A

1⊗ε ε⊗1
A ⊗ R ←− A ⊗ A R ⊗ A ←− A ⊗ A


=↑ ↑c ∼
=↑ ↑c

1 1
A ←− A A ←− A

(as before, we have omitted subscripts, and 1 stands for the identity on A).

It should be noticed that the diagrams occurring in the last definition are
obtained by reversing arrows in those describing algebra properties.
S(M ) is an R-coalgebra with respect to the following counit and
comultiplication.
ε is the identity on S0 (M ) = R and the zero map on Si (M ) for every i ≥ 1.
c is constructed from the diagonal mapping ∆ : m ∈ M → (m, m) ∈ M × M :
since ∆ is R-linear, a morphism S(M ) → S(M × M ) of R-algebras is induced;
but S(M × M ) ∼ = S(M ) ⊗R S(M ) as R-algebras, and c is defined to be the
composite S(M ) → S(M × M ) ∼ = S(M ) ⊗R S(M ).

Remark I.3.8 In the above, the following definition is understood: given two
R-algebras A and B, the R-algebra A ⊗R B is the R-module A ⊗R B with
multiplication defined by

(a1 ⊗ b1 )(a2 ⊗ b2 ) = a1 a2 ⊗ b1 b2
u ⊗u
and unit given by the composition RA −→ R ⊗R R A−→B A ⊗R B.

The fact, stated above, that S(M × M ) = ∼ S(M ) ⊗R S(M ) is easily seen if
one takes, instead of M × M the direct sum M ⊕ M . It is clear that M ⊕ M is
the degree one part of both S(M ⊕ M ) and of S(M ) ⊗R S(M ). The universality
property of the functor S(M ) does the rest.
As for an explicit description of c in the case of S(M ), since the isomorphism
S(M × M ) ∼ = S(M ) ⊗R S(M ) is induced by (m1 , m2 ) → m1 ⊗ 1 + 1 ⊗ m2 , and
c is a morphism of algebras, one gets

k 
c(m1 · · · mk ) = (mi ⊗ 1 + 1 ⊗ mi ) = (mi1 · · · mih ) ⊗ (mj1 · · · mjk−h ),
i=1
26 Recollections and perspectives

the sum ranging over all pairs of strictly increasing sequences i1 < · · · < ih and
j1 < · · · < jk−h such that
{i1 , . . . , ih } ∪ {j1 , . . . , jk−h } = {1, . . . , k}
(including the cases in which one of the two sequences is empty).
One should remark that S(M ) is a graded coalgebra, in the sense that all
maps are compatible with the given gradings.
Remark I.3.9 Since the comultiplication of S(M ) is obtained from the diag-
onal map M → M × M , it is customary to call diagonal map the
comultiplication of any coalgebra, replacing the letter c by ∆.
In order to explain what a (graded) Hopf algebra is, we need the definition
of the product of coalgebras. Let A and B be two given R-coalgebras. The
R-coalgebra A ⊗R B is the R-module A ⊗R B with counit εA⊗R B (a ⊗ b) =
εA (a)εB (b) and with diagonal map

∆A⊗R B (a ⊗ b) = (ai ⊗ bj ) ⊗ (ai ⊗ bj ),
i,j
 
where i ai ⊗ = ∆A (a) and j bj ⊗ bj = ∆B (b).
ai
We also need the following: given R-coalgebras A and B, an R-coalgebra
homorphism ϕ : A → B is an R-module map such that the diagrams
∆ ε
A −→
A
A⊗A A −→
A
R

ϕ↓ ↓ϕ⊗ϕ ϕ↓ ↓1
ε
B −→
B∆
B⊗B B −→
B
R

are commutative.
Definition I.3.10 A Hopf algebra over R is an R-module A, which is both
an algebra and a coalgebra over R, satisfying the following properties:
(1) the multiplication m and the unit u are coalgebra morphisms;
(2) the diagonal map ∆ and the counit ε are algebra morphisms;
(3) there exixts an R-module map sA : A → A (called the antipode) such that
the diagrams
1⊗s s⊗1
A ⊗ A −→ A ⊗ A A ⊗ A −→ A ⊗ A

∆↑ ↓m ∆↑ ↓m
u◦ε u◦ε
A −→ A A −→ A
are commutative.
Multilinear algebra 27

We leave as an exercise the verification that S(M ) is indeed a graded Hopf


algebra, that is, a Hopf algebra with all structure maps preserving the graded
structure. The antipode s is defined as 1 on all Si (M ) with i even, and as −1 on
all Si (M ) with i odd.
In fact, S(M ) is a commutative graded Hopf algebra, meaning that it is both
a commutative graded R-algebra and a cocommutative graded R-coalgebra.
A little bit of care is necessary here, for the general definition of (co-)
commutativity in the graded case does not suit the intuitive ideas suggested
by our polynomial ring.

Definition I.3.11 Let A = ⊕i∈N Ai and B = ⊕i∈N Bi be two graded R-modules.


Let τB,A : B ⊗R A → A ⊗R B be the map defined on homogeneous elements by
means of τB,A (b ⊗ a) = (−1)deg(b) deg(a) a ⊗ b.
If A has the structure of a graded R-algebra, we say that it is a commutative
graded R-algebra if the following diagram
τA,A
A ⊗ A −→ A ⊗ A

m↓ ↓m

1
A −→ A
is commutative.
If B has the structure of a graded R-coalgebra, we say it is a cocommutative
graded R-coalgebra if the following diagram
τB,B
B ⊗ B ←− B ⊗ B

∆↑ ↑∆

1
B ←− B
is commutative.

For these notions of commutativity and cocommutativity to suit a symmetric


algebra ⊗iεN Si (M ), one usually assumes that the elements of Si (M ) have degree
2i, not i.

Remark I.3.12 The last definition also has an impact on the construction
of the graded R-algebra A ⊗R B. The multiplication defined on A ⊗R B in
Remark I.3.8 is the composition
1A ⊗τ ⊗1B mA ⊗mB
A⊗B⊗A⊗B −→ A⊗A⊗B⊗B −→ A⊗B
(with τ (b, a)=(a, b)) and applies to non-graded algebras, as well as to graded
algebras concentrated in even degrees (such as the symmetric algebras). But in the
28 Recollections and perspectives

general graded case, it is replaced by the following, slightly different, composition


1A ⊗τB,A ⊗1B mA ⊗mB
A⊗B⊗A⊗B −→ A⊗A⊗B⊗B −→ A ⊗ B,
with τB,A as in the last definition. That is,
(a1 ⊗ b1 )(a2 ⊗ b2 ) = (−1)deg(b1 ) deg(a2 ) a1 a2 ⊗ b1 b2 .
In the future, unless otherwise stated, we always adopt for graded algebras
the specified graded versions of (co-)commutativity and tensor product. We also
assume that the elements of S(Mi ) have degree 2i.

I.3.2 The divided power algebra


We again let F be a rank t free R-module, and let a basis of F be {x1 , . . . , xt }.
We use E (rather than F ∗ ) to denote the R-module HomR (F, R), the dual of F .
Since ⊕ti=1 commutes with HomR , E is a rank t free R-module.  We indicate by
1 if i = j
{y1 , . . . , yt } the basis of E such that for every i, yi (xj ) = (the
0 otherwise
so-called dual basis of {x1 , . . . , xt }).
For every n ∈ N, the symmetric group Sn acts on the n-fold tensor product
Tn (E) = E ⊗R · · · ⊗R E by means of σ(e1 ⊗ · · · ⊗ en ) = eσ(1) ⊗ · · · ⊗ eσ(n) . The
elements of Tn (E) fixed by such an action are called symmetric tensors of order
n and form an R-submodule of Tn (E), denoted by Dn (E).
There is an R-map, sym : ⊕n∈N Tn (E) −→ ⊕n∈N Dn (E), called symmetriza-
tion, defined by means of

sym(z) = σ(z)
σ∈Sn

for every z ∈ Tn (E) and for every n ∈ N. If z ∈ Dn (E), then sym(z) = n!z.
Let D(E) stand for the graded R-module ⊕n∈N Dn (E).
Proposition I.3.13 D(E) is a graded R-algebra.
Proof
 Given z ∈ Di (E) and z  ∈ Dj (E), then zz  ∈ Di+j (E) is defined to be

σ∈Si,j σ(z ⊗ z ), where

Si,j = {σ ∈ Si+j | σ(1) < · · · < σ(i) and σ(i + 1) < · · · < σ(i + j)} .
The remainder of the proof is standard. 2
Remark I.3.14 If e1 , . . . , en belong to E, then their product e1 · · · en in D(E)
equals sym(e1 ⊗ · · · ⊗ en ).
Definition I.3.15 D(E) is called the divided power algebra on the free
R-module E.
If we assume that the elements of Dn (E) have degree 2n, then D(E) is
commutative as a graded R-algebra.
Multilinear algebra 29

The construction of D(E) is functorial and commutes with base change.


We now give a reason for the name assigned to D(E).
It is clear that for every e ∈ E and for every n ∈ N, the n-fold tensor product
e ⊗ · · · ⊗ e belongs to Dn (E): we denote it by e(n) . Since sym(e(n) ) = n!e(n) , by
Remark I.3.14 it follows that n!e(n) equals the product in D(E) of n copies of
e (=e(1) ), that is, the n-th power of e. If n! happens to be invertible in R, then
e(n) coincides with the n-th power of e divided by n!
Definition I.3.16 For every e ∈ E and every n ∈ N, the element e(n) of
Dn (E) is called the n-th divided power of e.
Remark I.3.17 It is not hard to check that divided powers verify:
(i+j)! (i+j)
(1) e(i) e(j) = i!j! e
n (i) (n−i)
(2) (e1 +e2 )(n) = i=0 e1 e2 (no binomial coefficients in the summation!).
Recalling the basis {y1 , . . . , yt } of E, one shows that every Dn (E) is R-free
and has a basis given by all elements
(s1 ) (st )
y1 · · · yt with s1 + · · · + st = n.
D(E) is in fact a commutative graded Hopf algebra, with diagonal map
defined by:
  
(s ) (s ) (u ) (u ) (v ) (v )
∆ y1 1 · · · yt t = y1 1 · · · yt t ⊗ y 1 1 · · · y t t ,

where the sum ranges over all pairs (u1 , . . . , ut ) and (v1 , . . . , vt ) of elements of
Nt , such that ui + vi = si for every i in {1, . . . , t}. We leave all the details to the
reader.
The indicated basis of D(E) leads one to think that D(E) is a symmetric
algebra (a polynomial ring). This is indeed the case if R contains a copy of the
(s ) (s )
rationals. For then every basis element y1 1 · · · yt t can be replaced by
(s1 ) (st )
s1 ! · · · st !y1 · · · yt = y1s1 · · · ytst
and D(E) ∼ = R[y1 , . . . , yt ].
In general, however, D(E) is not a symmetric algebra. What is true, though,
is that D(E) is the graded dual of a symmetric algebra.
Remark I.3.18 It is well known that, given R-modules Mi , i ∈ N, and L,
 
 
HomR Mi , L ∼= HomR (Mi , L).
i∈N i∈N

In particular, the ordinary dual of a graded R-algebra A = ⊕i∈N Ai is i∈N A∗i .
Hence, in order to make sure that the dual of a graded algebra is still a graded
algebra, one replaces A∗ with another object, namely ⊕i∈N A∗i , which is called
the graded dual of A and is denoted by A∗gr .
30 Recollections and perspectives

Theorem I.3.19 D(E) is the graded dual of the symmetric algebra S(F ), and
S(F ) is the graded dual of the divided power algebra D(E).
If one uses the notation F ∗ in place of E, the above says that D(F ∗ ) = S(F )∗gr
and S(F ) = D(F ∗ )∗gr .
We sketch a proof of the theorem, by giving an idea of why D(E)∗gr equals
S(F ). Essentially, it all amounts to showing that the diagonal map of D(E)
induces on D(E)∗gr a graded R-algebra structure corresponding to that of S(F ).
Let us take f ∈ Di (E)∗ and g ∈ Dj (E)∗ , and use the diagonal map of
D(E) to define an element f g ∈ Di+j (E)∗ . Since f ∈ HomR (Di (E), R) and
g ∈ HomR (Dj (E), R), we may set f g equal to the composite map
∆ f ⊗g
Di+j (E) −→ Di (E) ⊗R Dj (E) −→ R ⊗R R −→ R,
an element of HomR (Di+j (E), R), as required. One checks that this multiplica-
tion gives D(E)∗gr (which equals S(F ) as an R-module) an R-algebra structure.
We claim that such a structure is the ordinary multiplication in S(F ).
Clearly, it suffices to prove that
 ∗  ∗  ∗
(s ) (s ) (r ) (r ) (s +r ) (s +r )
(∗) y1 1 · · · y t t y 1 1 · · · yt t = y1 1 1 · · · yt t t ,
 ∗
stands for the element of (Ds1 +···+st E)∗ = Ss1 +···+st F
(s1 ) (st )
where y1 · · · yt
(s ) (s )
dual to y1 1 · · · yt t (and due to correspond to xs11 · · · xst t ). But (∗) is obvious,
in view of the definition of ∆ in D(E) recorded above.

Let us take t = 2 and show that y1∗ (y1 y2 )∗ = (y1 y2 )∗ .


(2)
Example I.3.20
∆ f ⊗g
We consider D3 (E) −→ D1 (E) ⊗R D2 (E) −→ R ⊗RR −→ R, where f andg
are y1∗ and (y1 y2 )∗ , respectively. A basis for D3 (E) is y1 , y1 y2 , y1 y2 , y2 .
(3) (2) (2) (3)

(2)
We must prove that the composite map assigns 1 to y1 y2 and 0 to all other
basis elements.
The action of y1∗ ⊗(y1 y2 )∗ on ∆(y1 y2 ) = y1 ⊗y1 y2 +y2 ⊗y1 yields 1⊗1+0⊗0,
(2) (2)

hence 1.
The action of y1∗ ⊗(y1 y ∗ (2) (2) (2)
2 ) on ∆(y
(3)
) = y1 ⊗ y1 , on ∆(y1 y2 ) = y1 ⊗ y2 +
(3) (2)
y2 ⊗ y1 y2 , and on ∆ y2 = y2 ⊗ y 2 yields (respectively) 1 ⊗ 0, 1 ⊗ 0 + 0 ⊗ 1,
and 0 ⊗ 0; hence 0 all the times.
Concerning the example, one should notice that the product of y1 and y1 y2 in
(2)
D(E) equals 2y1 y2 (by part 1 of Remark I.3.17), while x1 times x1 x2 in S(F )
gives x21 x2 .

I.3.3 The exterior algebra


Definition I.3.21 Let M be an R-module. We call an exterior algebra on
M a pair (Λ(M ), ϕ) satisfying the following requirements: Λ(M ) is an R-algebra,
Multilinear algebra 31

ϕ : M → Λ(M ) is an R-morphism, every z ∈ ϕ(M ) gives z 2 = 0, and for every


R-map ψ : M → A, with A an R-algebra, such that every a ∈ ψ(M ) gives
a2 = 0 in A, there exixts a unique R-algebra morphism χ : Λ(M ) → A verifying
ψ = χ ◦ ϕ.
Clearly, if exterior algebras on M exist, there is a unique isomorphism from
one to the other which is the identity on M .
Theorem I.3.22 Every M admits an exterior algebra.
As in the case of S(M ), one takes the tensor algebra T (M ) and defines Λ(M )
as the quotient T (M )/b, where b is the two-sided ideal generated by all elements
m ⊗ m, with m ∈ M . The map ϕ is given by the composition:
M → T (M ) → T (M )/b.
Clearly, ϕ(m)2 = 0 for every m ∈ M . If ψ : M → A is given, such that every
a ∈ ψ(M ) gives a2 = 0 in A, then the map T (M ) → A extending ψ vanishes on
b, and induces the required χ : T (M )/b → A. Finally, since T (M ) is generated
by the elements of M , ϕ(M ) generates Λ(M ) and the uniqueness of χ follows.
One should remark that since b is homogeneous, Λ(M ) inherits the gradu-
ation of T (M ): Λ(M ) = ⊕i∈N Λi (M ) is the customary notation. In particular,
ϕ identifies M with Λ1 (M ), and the elements of Λ1 (M ) generate Λ(M ).
The construction of Λ(M ) is functorial and commutes with base change.
Multiplication in Λ(M ) is called exterior multiplication. The usual notation
for it is m1 ∧ m2 , and we stick to this convention.
Regarding the commutativity of Λ(M ), some considerations are in order. Since
m1 , m2 ∈ M implies
0 = (m1 + m2 ) ∧ (m1 + m2 )
= m1 ∧ m1 + m1 ∧ m2 + m2 ∧ m1 + m2 ∧ m2
= m 1 ∧ m2 + m 2 ∧ m 1 ,
we get m1 ∧ m2 = −m2 ∧ m1 . Hence
(m1 ∧ · · · ∧ mi ) ∧ (mi+1 ∧ · · · ∧ mi+j ) =(−1)ij (mi+1 ∧ · · · ∧ mi+j )
∧ (m1 ∧ · · · ∧ mi ).
But Λ(M ) generated by Λ1 (M ) then says
z ∧ w = (−1)deg(z) deg(w) w ∧ z
for every choice of homogeneous elements z and w.
We conclude that Λ(M ) is a commutative graded algebra in the sense of
Definition I.3.11.
In fact more is true. Λ(M ) is a commutative graded Hopf algebra. We do not
go into the details, but record here the diagonal map:

∆(m1 ∧ · · · ∧ mk ) = (−1)s (mi1 ∧ · · · ∧ mih ) ⊗ (mj1 ∧ · · · ∧ mjk−h ),
32 Recollections and perspectives

where s stands for the number of ordered pairs (u, v) verifying jv < iu , and
the sum ranges over all pairs of strictly increasing sequences i1 < · · · < ih and
j1 < · · · < jk−h such that
{i1 , . . . , ih } ∪ {j1 , . . . , jk−h } = {1, . . . , k}
(including the cases in which one of the two sequences is empty).
We now focus our attention on Λ(F ), where F is a finite free R-module,
and reobtain some results of Subsection I.2.1, namely Propositions I.2.4, I.2.7,
and I.2.9.

Lemma I.3.23 Let M be a finitely generated R-module and let {m1 , . . . , mt }


be a generating system. Then Λn (M ) = 0 for every n > t.

Proof For every n ≥ 1, Λn (M ) has a generating system given by all elements


of type mi1 ∧ · · · ∧ min , and if n > t, some mi occurs more than once. 2

Proposition I.3.24 Let F be a free R-modulet  with basis {f1 , . . . , ft }. Then


Λ(M ) is R-free with a basis of cardinality i=0 ti = 2t .

Proof Write F = F1 ⊕ · · · ⊕ Ft , where Fi = Rfi for every i = 1, . . . , t.


Since Fi has one generator, Lemma I.3.23 says that Λ(Fi ) = R ⊕ Fi ; hence
Λ(Fi ) is R-free with basis {1, fi }. It follows that Λ(F ) = Λ(F1 ) ⊗ · · · ⊗ Λ(Ft )
has a basis given by all elements m1 ∧ · · · ∧ mt , with mi ∈ {1, fi }. In par-
ticular, for every s ∈ {0, 1, . . . , t}, the elements m1 ∧ · · · ∧ mt with
 s factors
different from 1 are a basis, Bs , of Λs (F ), with cardinality st . Explicitly:
Bs = {fi1 ∧ · · · ∧ fis with i1 < · · · < is }. 2

Remark I.3.25 (Proposition I.2.4) If B = {f1 , . . . , ft } and B  = {f1 , . . . , fs }


are two finite bases of the same free R-module F , then t = s. For both t < s and
t > s give a contradiction. For instance, if t < s, then |B| = t implies Λs (F ) = 0
(by Lemma I.3.23),
 s while |B  | = s implies Λs (F ) = 0, since Λs (F ) has a basis
of cardinality s = 1 (by Proposition I.3.24).

In order to state the next lemma, we need the notion of an alternating n-linear
function on M . It is an n-linear function ψ : M × · · · × M → N , with N any
R-module, such that ψ(m1 , . . . , mn ) = 0 whenever two of the arguments coincide.
Clearly,
ϕ : M × · · · × M → Λn (M ), (m1 , . . . , mn ) → m1 ∧ · · · ∧ mn
is an alternating n-linear function. It is not hard to see that every ψ as above
factors through ϕ, that is, ψ = χ ◦ ϕ for some linear map χ : Λn (M ) → N .

Lemma I.3.26 Let F be a finite free R-module and {m1 , . . . , mt } ⊆ F . Then,


m1 ∧· · ·∧mt = 0 if and only if there exists a t-linear function ψ : F ×· · ·×F → R
such that ψ(m1 , . . . , mt ) = 0.
Multilinear algebra 33

Proof If part. Write ψ as the composite χ ◦ ϕ, where χ : Λt (F ) → R sends


m1 ∧ · · · ∧ mt to ψ(m1 , . . . , mt ). Then m1 ∧ · · · ∧ mt = 0 would give χ(m1 ∧ · · · ∧
mt ) = 0, a contradiction.
Only if part. If not, all elements of (Λt (F ))∗ would vanish on m1 ∧ · · · ∧ mt ,
that is, m1 ∧ · · · ∧ mt (as an element of (Λt (F ))∗∗ = Λt (F )) would vanish on
(Λt (F ))∗ , that is, m1 ∧ · · · ∧ mt = 0 in Λt (F ). 2
Proposition I.3.27 Let F be a finite free R-module and {m1 , . . . , mt } ⊆ F .
Then, m1 , . . . , mt are linearly dependent if and only if there exists a ∈ R − {0}
such that am1 ∧ · · · ∧ mt = 0.
Proof Only if part. If a1 m1 + · · · + at mt = 0 with not all coefficients equal
to zero, a1 = 0 say, then a1 m1 equals −a2 m2 − · · · − at mt and the properties of
exterior multiplication give a1 m1 ∧ m2 ∧ · · · ∧ mt = 0.
If part. By induction on t, case t = 1 being trivial. Assume am1 ∧· · ·∧mt = 0;
if also am2 ∧· · ·∧mt = 0, then m2 , . . . , mt are dependent by inductive hypothesis,
and a fortiori m1 , . . . , mt are dependent. If am2 ∧ · · · ∧ mt = 0, Lemma I.3.26
says that there is a (t − 1)-linear function ψ such that ψ(am2 , . . . , mt ) = b = 0.
Since m1 ∧ am2 ∧ · · · ∧ mt = 0, every t-linear function vanishes on
(m1 , am2 , m3 , . . . , mt )
(again by Lemma I.3.26). In particular, this happens for

t
(m1 , . . . , mt ) → i , . . . , mt )mi ,
(−1)i+1 ψ(m1 , . . . , m
i=1

where  means omitted. Thus ψ(am2 , . . . , mt )m1 = bm1 equals a linear


combination of m2 , . . . , mt , and m1 , . . . , mt turn out to be dependent. 2
Corollary I.3.28 Let F be a free R-module with basis {f1 , . . . , ft }, and let
{m1 , . . . , mt } ⊆ F . Then {m1 , . . . , mt } is a basis of F if and only if m1 ∧ · · · ∧
mt = af1 ∧ · · · ∧ ft with a invertible in R.
Proof Only if part. By Proposition I.3.24, {f1 , . . . , ft } a basis of F implies
f1 ∧ · · · ∧ ft a basis of Λt (F ) ∼
= R; hence
m1 ∧ · · · ∧ mt = af1 ∧ · · · ∧ ft
for some a in R. Similarly, {m1 , . . . , mt } a basis of F implies
f1 ∧ · · · ∧ ft = bm1 ∧ · · · ∧ mt
for some b in R. It follows
f1 ∧ · · · ∧ ft = baf1 ∧ · · · ∧ ft ,
hence ba = 1.
If part. Define χ : Λt (F ) → R by χ(f1 ∧ · · · ∧ ft ) = a−1 . Call ψ the composite
χ ◦ ϕ : F × · · · × F → R, where ϕ is as defined before Lemma I.3.26. Clearly,
34 Recollections and perspectives

χ(m1 ∧ · · · ∧ mt ) = aa−1 = 1; hence ψ(m1 , . . . , mt ) = 1. For every m ∈ F ,


m ∧ m1 ∧ · · · ∧ mt = 0 since Λt+1 (F ) = 0. By Lemma I.3.26, it must vanish on
(m, m1 , . . . , mt ) the (t + 1)-linear function

t
(m0 , m1 , . . . , mt ) → i , . . . , mt )mi .
(−1)i ψ(m0 , . . . , m
i=0

Hence

t
ψ(m1 , . . . , mt )m = i , . . . , mt )mi .
(−1)i+1 ψ(m, m1 , . . . , m
i=1

Since ψ(m1 , . . . , mt ) = 1, m turns out to be a linear combination of m1 , . . . , mt


and we have shown that {m1 , . . . , mt } is a generating system. It is in fact a basis,
thanks to Proposition I.3.27, because m1 ∧ · · · ∧ mt = 0 by hypothesis. 2
Corollary I.3.29 Let F be a free R-module with basis {f1 , . . . , ft }. Let ϕ : F →
N be an injective R-map. Then the induced R-map Λ(ϕ) : Λ(F ) → Λ(N ) is also
injective. (By definition, for every n ∈ N, Λ(ϕ) sends each m1 ∧· · ·∧mn ∈ Λn (F )
to ϕ(m1 ) ∧ · · · ∧ ϕ(mn ).)
Proof If not, let the kernel of Λ(ϕ) contain a linear combination of distinct
basis elements

ai1 ,...,is fi1 ∧ · · · ∧ fis ,

where every ai1 ,...,is is = 0. Choose a summand having minimum number


of factors in the exterior product, say the one corresponding to the indices
i1 < · · · < is0 . Call I the subset of {1, . . . , t} having empty intersection with
 
i1 , . . . , is0 , but containing all the other indices occurring in the sum. Say
I = {j1 , . . . , jk }, with j1 < · · · < jk . Ley fI denote the basis element fj1 ∧· · ·∧fjk .
Then
 
ai1 ,...,is fi1 ∧ · · · ∧ fis ∧ fI

simply equals the exterior product of the chosen summand and fI . Since it is
still an element of the two-sided ideal ker(Λ(ϕ)), if we apply Λ(ϕ) to it, we get:
0 = ai1 ,...,is ϕ(fi1 ) ∧ · · · ∧ ϕ(fis ) ∧ ϕ(fj1 ) ∧ · · · ∧ ϕ(fjk ).
0

By Proposition I.3.27, it follows that ϕ(fi1 ), . . . , ϕ(fis ), ϕ(fj1 ), . . . , ϕ(fjk ) are


0
linearly dependent: a contradiction, because fi1 , . . . , fis , fj1 , . . . , fjk are inde-
0
pendent by hypothesis, and the injective map ϕ preserves independence. 2
Remark I.3.30 (Proposition I.2.7)
1. Let F be a free R-module with basis {f1, . . . ,ft }. Let {m1, . . . ,mt } ⊆ F be
defined by means of (m1 , . . . , mt )T = S(f1 , . . . , ft )T , where S is a t × t matrix
Multilinear algebra 35

with entries in R, and T denotes transposition. Using the properties of exterior


multiplication, one checks that
m1 ∧ · · · ∧ mt = af1 ∧ · · · ∧ ft with a = det S.
Hence by Corollary I.3.28, {m1 , . . . , mt } is another basis of F if and only if det S
is invertible in R.
2. Let F be a free R-module with basis {f1 , . . . , ft }. Let {m1 , . . . , mt } ⊆ F be
a generating system. We claim that {m1 , . . . , mt } is in fact a basis. For
(f1 , . . . , ft )T = U (m1 , . . . , mt )T
for some appropriate t × t matrix U with entries in R. Hence
f1 ∧ · · · ∧ ft = det U m1 ∧ · · · ∧ mt
couples with
m1 ∧ · · · ∧ mt = af1 ∧ · · · ∧ ft ,
yielding a det U = 1. It follows that a is a unit, and {m1 , . . . , mt } is a basis by
Corollary I.3.28.
Remark I.3.31 (Proposition I.2.9) Let F be a free R-module with basis
{f1 , . . . , ft }. Let ϕ : F → F have matrix S with respect to the given basis. Con-
sider the induced map Λt (ϕ) : Λt (F ) → Λt (F ). This is a map R → R, identified
by the value assigned to 1, say a. Then
Λt (ϕ)(f1 ∧ · · · ∧ ft ) = af1 ∧ · · · ∧ ft .
In fact, a = det S (using the properties of exterior multiplication). Thus if ϕ
is injective by assumption, then Λt (ϕ) is injective as well (by Corollary I.3.30),
and det S cannot be a zero divisor.
Conversely, if det S is not a zero divisor, then
t ϕ must be injective. Otherwise,
ϕ(f ) = 0 for some non-zero f ∈ F implies ϕ( i=1 ai fi ) = 0 with some non-zero
t
ai , that is, i=1 ai ϕ(fi ) = 0. By Proposition I.3.27, there exists a ∈ R − {0}
such that aϕ(f1 ) ∧ · · · ∧ ϕ(ft ) = 0, that is, aΛt (F )(f1 ∧ · · · ∧ ft ) = 0, that is,
a det Sf1 ∧ · · · ∧ ft = 0. Thus a det S = 0, and det S is a zero divisor.
We call the reader’s attention to the fact that if F is R-free of finite rank t,
say, then
Λ(F ) = Λ0 (F ) ⊕ · · · ⊕ Λt (F )
implies
(Λ(F ))∗ = (Λ(F ))∗gr = (Λ0 (F ))∗ ⊕ · · · ⊕ (Λt (F ))∗
= Λ0 (F ∗ ) ⊕ · · · ⊕ Λt (F ∗ ) = Λ(F ∗ ),
where the penultimate equality is a consequence of the first part of the following
remark.
36 Recollections and perspectives

Remark I.3.32
1. Let F be R-free with basis {f1 , . . . , ft }. Let {g1 , . . . , gt } be the basis of F ∗
dual to {f1 , . . . , ft }. Then for every k ∈ {1, · · · t}, we can define an isomorphism
Λk (F ∗ ) → (Λk (F ))∗ by sending each basis element gi1 ∧· · ·∧gik to the application
. , ik ) ∈ HomR (Λk (F ), R) defined by means of ϕ(i1 , . . . , ik )(fj1 ∧ · · · ∧
ϕ(i1 , . .
1 if ih = jh ∀h
fjk ) = . The indicated isomorphism is independent of the choice
0 otherwise
of the basis {f1 , . . . , ft }.
2. Let F be a rank t free R-module. If a non-zero element of Λt (F ) is chosen,
for example, an isomorphism ψ0 : Λt (F ) → R is fixed, then further isomorphisms
ψk : Λt−k (F ) → Λk (F ∗ ) are obtained for k ∈ {1, . . . , t}. Namely, given z ∈
Λt−k (F ) and w ∈ Λk (F ), one sets ψk (z)(w) = ψ0 (z ∧ w). This means that an
isomorphism Λ(F ) ∼ = Λ(F ∗ ) also exists. However, it is not independent of the
choice of the basis in F (= the choice of the element of Λt (F )).
Also the minors of a matrix can be described in terms of exterior multiplic-
ation. For let E and F be finite free R-modules, with bases {e1 , . . . , es } and
{f1 , . . . , ft }, respectively. Let ψ ∈ HomR (E, F ) have t × s matrix S  with
  respect
to the indicated bases, and let Λk (ψ), 1 ≤ k ≤ min{t, s}, have kt × ks matrix
U relative to the corresponding bases of Λk (E) and Λk (F ). Then the entry of
U , which is located at the intersection of the row indexed by the basis element
fi1 ∧ · · · ∧ fik and the column indexed by the basis element ej1 ∧ · · · ∧ ejk , is
precisely the k × k minor of S with rows indexed by fi1 , . . . , fik and columns
indexed by ej1 , . . . , ejk . Such a minor is often denoted by (i1 , . . . , ik | j1 , . . . , jk ).
It is customary to denote by Ik (ψ) the ideal generated by the entries of U ,
that is, by the k × k minors of S. The ideal Ik (ψ) is independent of the choice
of bases because the isomorphism
HomR (Λk (E), Λk (F )) ∼
= (Λk (E))∗ ⊗R Λk (F ),
coupled with
(Λk (E))∗ ⊗R Λk (F ) ∼
= (Λk (E) ⊗R Λk (F ∗ ))∗ = HomR (Λk (E) ⊗R Λk (F ∗ ), R),
identifies Λk (ψ) with a map Λk (E) ⊗R Λk (F ∗ ) → R, whose image is pre-
cisely Ik (ψ).
This suggests that we define
Λk (ψ)
Ik (ψ) = im(Λk (E) ⊗R Λk (F ∗ ) −→ R) ∀k ≥ 0,
which, among other things, implies that the ideal is 0 for k > min{t, s} (no minors
exist), and that the 0 × 0 minors generate R (a 0 × 0 matrix has determinant 1).
II
LOCAL RING THEORY

The development of the first five sections of this chapter is guided by the desire to
establish a string of inequalities for local rings (always assumed to be noetherian
here) which lead to the result, in Section II.5, that a local ring is regular if and
only if its global dimension is finite. The development here is very close to that
used in the notes of Reference [19] and rests on the work found in References [7]
and [8].
In the first section we define and develop the properties of the Koszul complex,
which will play a fundamental role throughout the book. The next section, on
local rings, ends with a statement of a fundamental inequality whose proof rests
almost completely on these properties. This inequality states that the global
dimension of a local ring always exceeds the minimal number of generators of
its maximal ideal. While Section II.3 on Hilbert–Samuel polynomials does not
use the complex itself, it does rely on the notion of “regular sequence” for some
of its fundamental results, and this notion is intimately connected with that
tool. In that section we establish the fact that the dimension of any local ring
is always exceeded by the minimum number of generators of its maximal ideal
(Krull’s Principal Ideal Theorem). The results of Section II.4 on codimension
and finitistic global dimension are inextricably tied up with the Koszul complex;
there we establish the facts that the finitistic global dimension of a local ring
is equal to its codimension, and that its codimension is never greater than its
dimension. Section II.5 on regular local rings puts together all of the above to
achieve the result mentioned in our first paragraph. Another main result there
is the Cohen–Macaulay Theorem; its main new “homological” point is that the
“unmixedness” part of the classical theorem is sharpened by the observation that
the Koszul complex associated to the generators of the ideal is acyclic, which then
yields the result.
The sixth section of this chapter contains a full proof of the unique factoriz-
ation theorem for regular local rings. This theorem, too, was one of the prime
objectives of the application of homological methods to commutative ring theory.
In the penultimate section, we take up the notion of multiplicity, which is again
intimately tied up with the Koszul complex. While a detailed study of this the-
ory is beyond the scope of our book, it would be remiss to completely ignore it.
In this section, therefore, we give a definition of multiplicity and establish some
of its properties by means of the Koszul complex. Connected with this topic is
the notion of Cohen–Macaulay ring, and we will see that for such rings the two
38 Local ring theory

invariants of an ideal, its height and its depth, are equal. As the depth of an
ideal is a number that in a certain sense (that of relative primeness) measures
the “arithmetic” size of an ideal, while the height is a number that measures
its “geometric” size, their coincidence in the case of Cohen–Macaulay rings sug-
gests why the use of ideal-theoretic methods in algebraic geometry, when the
variety is Cohen–Macaulay, is so effective.
From multiplicity in Section II.7, we move to the Serre definition of intersection
multiplicity in Section II.8 and we close the chapter with a description of the so-
called homological conjectures. This book does not deal with these conjectures,
but we feel that no book that purports to present the interaction of local ring
theory and homology can omit at least a brief mention of these fundamental
questions and the work that has been done on them.

II.1 Koszul complexes


In the beginning, many mathematicians wondered how and how come homo-
logical methods could be related and applied to problems in local ring theory.
That homology is tied up naturally with algebra can be seen easily by looking
at the following example: Let R be a commutative ring (all rings are assumed to
have an identity element), and let x be an element of R. Then we may consider
the rather trivial complex
x
· · · → 0 → 0 → R → R → 0,
x
where R → R means multiplication by the element x. If we call this complex
K(x), we see that H0 (K(x)) = R/(x), and H1 (K(x)) = 0 : x = {r ∈ R| rx = 0}.
Thus in a sense H0 measures how divisible R is by x, while H1 tells us whether
or not x is a zero divisor. For an arbitrary module, M, we may also consider the
complex K(x) ⊗R M :
x
··· → 0 → 0 → M → M → 0
and observe that H0 (K(x)⊗R M ) = M/xM and H1 (K(x)⊗R M ) = 0 : x = {m ∈
M | xm = 0}. Again we may use these homology groups to tell us how divisible
the module M is by the element x, and whether or not x is a zero divisor for M.
Now let us suppose that we have two elements x, y in R. Then we have the
y
map K(x) → K(x) induced by multiplication by y:
x
···→0→0→R →R →0
↓y ↓y
x
···→0→0→R →R →0
and it is natural for homologists to consider the “mapping cone” of this map of
complexes:
f g
··· → 0 → 0 → R → R ⊕ R → R → 0
Koszul complexes 39

where g(r1 , r2 ) = r1 x + r2 y and f (r) = (−ry, rx). Calling this complex K(x, y),
we have H0 (K(x, y)) = R/(x, y); H2 (K(x, y)) = 0 : (x, y) = {r ∈ R| rx = 0 =
ry}. H1 (K(x, y)) = {(r1 , r2 )| r1 x+r2 y = 0}/{(−ry, rx)}. If we suppose that R is
an integral domain, and that neither x nor y is zero, we see that H2 (K(x, y)) = 0
and that {(r1 , r2 )| r1 x + r2 y = 0} is mapped isomorphically onto (x) : y = {r ∈
R| ry ∈ (x)} by the map sending (r1 , r2 ) to r2 . Under this map, the submodule
{(−ry, rx)} is sent onto the ideal (x) and thus H1 (K(x, y)) = (x) : y/(x). If R
were a unique factorization domain, then H1 (K(x, y)) would be zero if and only
if x and y were relatively prime. Therefore, we again see that a certain amount
of homological information is intimately connected with arithmetical questions.
We could proceed step by step in this way, building up complexes
K(x, y, z, . . .), but it is more efficient to consider the following general construc-
tion. Let f : A → R be a morphism from the R-module A to R. Then we
may define the Koszul complex, K(f ), by setting K(f )p = Λp A and defining
df : Λp A → Λp−1 A by
 ∧
df (a1 ∧ · · · ∧ ap ) = (−1)i+1 f (ai ) a1 ∧ · · · ∧ ai ∧ · · · ∧ ap ,

where ai means we omit ai . It is well known, and easy to check, that (df )2 = 0.
Also, if α ∈ Λp A and β ∈ Λq A, then

df (α ∧ β) = df (α) ∧ β + (−1)p α ∧ df (β).

To see how this subsumes the sort of thing we have been discussing, consider
a sequence, x1 , . . . , xs , of elements of our ring R. We can define a morphism
f : Rs → R, where Rs denotes the direct sum of R with itself s times, and f is
defined by f (1, . . . , 0) = x1 , . . . , f (0, 0, . . . , 1) = xs . In this case, we denote the
complex K(f ) by K(x1 , . . . , xs ). For s = 1 or s = 2, we obtain the complexes
K(x) and K(x, y) described above.
Before we return to the general case of a morphism f : A → R, let us review the
mapping cone construction for a map between complexes. First, a convention
about complexes. Since in most of our applications our complexes will be left
complexes, we will assume that a complex C is given by {Cn ; dn } with the
boundary map d = {dn } of degree −1. That is, for each n, dn : Cn → Cn−1 , and
we will usually write the complex:
d d d d d−1 d−2
· · · →3 C2 →2 C1 →1 C0 →0 C−1 → C−2 → · · · .

Thus, our complex K(f ) above has all its negative parts equal to zero, K(f )0 =
R, K(f )1 = A, and so on.

Definition II.1.1 Let F : C → D be a map of complexes. The mapping cone


of the map F, denoted by M(F ), is defined as follows

M (F )n = Dn ⊕ Cn−1 ;
40 Local ring theory

and the boundary map δn : M (F )n → M (F )n−1 is defined by


  
δn (xn , yn−1 ) = (dn (xn ) + (−1)n−1 Fn−1 (yn−1 ), dn−1 (yn−1 ) .

(Here we are writing the boundary map for the complex D as d = {dn }.)
Notice that, module-theoretically, the complex M(F ) is the direct sum of the
complexes C and D, except for one thing: the degree of the component coming
from C is off by one. If we denote by C the complex {C n ; dn }, where C n = Cn−1
and dn = dn−1 , then our mapping cone is, in each degree, the direct sum of C
and D.
Proposition II.1.2 Let F : C → D be a map of complexes. Then we have the
short exact sequence of complexes:
(∗) 0 → D → M(F ) → C → 0
and the long exact sequence of homology:
· · · → Hn+1 (C) → Hn (D) → Hn (M(F )) → Hn (C) → · · ·

where the maps Hn+1 (C) → Hn (D) are the maps in homology induced by ±F.
Proof The exactness of the short exact sequence (∗) is clear on the module-
theoretic level. What has to be checked is that the obvious injection and
projection maps are maps of complexes. But this is quite simple. Once the
exactness of (∗) is established, the long exact homology sequence is an imme-
diate consequence; what must be explained is the remark that the maps
Hn+1 (C) → Hn (D) are the maps in homology induced by ±F. From the defini-
tion of C, it is clear that Hn+1 (C) = Hn (C), so that Hn+1 (C) → Hn (D) is really
Hn (C) → Hn (D). But the definition of the boundary map of M(F ) involves
the map F with a sign. A careful diagram chase to describe the “connecting
homorphism” makes it clear that this morphism is induced by the map F in
homology, with the sign determined in each dimension according to the definition
of δn . 2
Returning to the general case of a morphism f : A → R, we state the funda-
mental properties of the complex K(f ) :
Proposition II.1.3 If a0 is an element of A, then f (a0 )Hp (K(f )) = 0 for all
p ≥ 0.
Proposition II.1.4 If x is any element of R, and f : A → R a morphism, we
have the map of complexes x : K(f ) → K(f ) :
f
· · · → Λ3 A → Λ2 A → A → R
↓x ↓x ↓x ↓x .
f
· · · → Λ3 A → Λ 2 A → A → R
Koszul complexes 41

The mapping cone of this map of complexes is denoted by K(f ) where f : A⊕R →
R is defined by f (a, r) = f (a) + rx.

Proof The proof of Proposition II.1.3 is obtained by writing a homotopy s :


Λp A → Λp+1 A defined by s(a1 ∧ · · · ∧ ap ) = a0 ∧ a1 ∧ · · · ∧ ap , and showing that
multiplication by f (a0 ) is chain homotopic, via s, to the zero morphism. That
is, one must show that

df s(a1 ∧ · · · ∧ ap ) + sdf (a1 ∧ · · · ∧ ap ) = f (a0 )(a1 ∧ · · · ∧ ap ).

for all p. Proposition II.1.4 is proved by observing that for any module A, Λp (A⊕
R) ∼= Λp A ⊕ Λp−1 A, and then checking that under this identification the map
df is the boundary map of the mapping cone. 2

From Proposition II.1.4 and Proposition II.1.2 we obtain the following.

Corollary II.1.5 If f : A → R is a morphism, x an element of R, and f :


A ⊕ R → R defined as in Proposition II.1.4, we have the exact sequence:
±x
· · · → Hp (K(f )) → Hp (K(f )) → Hp−1 (K(f )) −→ Hp−1 (K(f )) → · · ·
· · · → H0 (K(f )) → H0 (K(f )) → 0.

In particular, if x1 , . . . , xs+1 is a sequence of elements of R, we have the exact


sequence:

(E) · · · → Hp (K(x)) → Hp (K(x, xs+1 )) → Hp−1 (K(x))


±xs+1
−→ Hp−1 (K(x)) → Hp−1 (K(x, xs+1 )) → · · · ,

where x stands for the sequence x1 , . . . , xs . We also get, for an arbitrary R-


module M, the exact sequence:

(E(M )) · · · → Hp (M(x)) → Hp (M(x, xs+1 )) → Hp−1 (M(x))


±xs+1
−→ Hp−1 (M(x)) → Hp−1 (M(x, xs+1 )) → · · · ,

where M(x) is the complex K(x) ⊗R M (still called a Koszul complex).

Proof All but the last statement about the exactness of (E(M )) follows
immediately from the propositions cited. The exactness of (E(M )) is due
to the fact that the short exact sequence (∗) applied to the complexes
C = D =K(x1 , . . . , xs ) and K(x1 , . . . , xs+1 ) = M(F ), where F is multiplica-
tion by xs+1 , splits. Thus tensoring by M preserves exactness, and the long
homology sequence (E(M )) is the one associated to the short one. 2

From the remarks made about the meaning of H1 (K(x, y)), we are led to make
the following definition.
42 Local ring theory

Definition II.1.6 Let M be an R-module, and x1 , . . . , xs a sequence of


elements of R. Then x1 , . . . , xs is said to be an M -sequence if
(a) M/(x1 , . . . , xs )M = 0;
(b) for every i, with i = 1 . . . , s, the element xi is not a zero divisor for
M/(x1 , . . . , xi−1 )M.
For i = 1, this means that x1 is not a zero divisor for M.
Sometimes an M -sequence is called an M -regular sequence, or just regular
sequence when M = R.
Proposition II.1.7 If x1 , . . . , xs is an M -sequence, then Hp (M(x1 , . . . , xs )) =
0 for all p > 0.
The proof proceeds by induction on s, the case s = 1 being self-evident. To
go from s to s + 1, one merely uses the exact sequence (E(M )) to show that
x
Hp (M(x1 , . . . , xs )) = 0 for p > 1, and the fact that H0 (M(x1 , . . . , xs−1 )) →s
H0 (M(x1 , . . . , xs−1 )) is a monomorphism (because H0 (M(x1 , . . . , xs−1 )) equals
M/(x1 , . . . , xs−1 )M ) to show that H1 (M(x1 , . . . , xs )) = 0.
Remark II.1.8 It is not true in general that Hp (M(x1 , . . . , xs )) = 0 for
all p > 0 implies that x1 , . . . , xs is an M -sequence. For example, let R =
k[X, Y, Z]/Y (X − 1, Z), with k a field. If we write X and Z for the residue
classes of X and Z in R, we see easily that X, Z is an R-sequence, so that
Hp (K(X, Z)) = 0 for p > 0. Therefore we obviously have that Hp (K(Z, X)) = 0
for p > 0 (since for any ring R, any sequence of elements x1 , . . . , xs in R, and
any R-module M, we have M(x1 , . . . , xs ) ∼ = M(xπ(1) , . . . , xπ(s) ), where π is a
permutation of the set {1, . . . , s}). However, it is clear that Z is a zero divisor
in R, so that Z, X fails to be an R-sequence.
As a result of the above remark, it is natural to ask when we can prove the
converse of Proposition II.1.7. If we now use our assumptions that R is noeth-
erian and M finitely generated, we conclude that Hp (M(x1 , . . . , xs )) is a finitely
generated R-module for any sequence of elements x1 , . . . , xs in R. Moreover, if
xi is in the radical of R, we can then conclude that the map
∧ x ∧
Hp (M(x1 , . . . , xi , . . . , xs )) →i Hp (M(x1 , . . . , xi , . . . , xs ))

is an epimorphism if and only if Hp (M(x1 , . . . , xi , . . . , xs )) = 0. Using these facts
we can prove
Proposition II.1.9 Let R be a noetherian ring, M a finitely generated non-
zero R-module, and x1 , . . . , xs a sequence of elements in the radical of R. Then
(1) if Hp (M(x1 , . . . , xs )) = 0 for some p, we have Hp+k (M(x1 , . . . , xt )) = 0 for
all k ≥ 0 and all t, 1 ≤ t ≤ s;
(2) if H1 (M(x1 , . . . , xs )) = 0, then x1 , . . . , xs is an M -sequence.
Local rings 43

Thus, under these hypotheses, we have x1 , . . . , xs is an M -sequence if and only


if Hp (M(x1 , . . . , xs )) = 0 for all p > 0 (or for p = 1).
Proof The proof of part (1) proceeds by induction on s. For s = 1 and p = 1,
there is nothing to be said. If s = 1 and p = 0, this means that M/x1 M = 0.
But by Nakayama’s Lemma (cf. Lemma I.2.16), this implies that M = 0, and
the rest follows. Now assume that s > 1, and assume the statement true for all
smaller values. From the exactness (using (E(M ))) of
±x
Hp (M(x1 , . . . , xs−1 )) →s Hp (M(x1 , . . . , xs−1 )) → Hp (M(x1 , . . . , xs )),
we conclude that multiplication by xs is an epimorphism. But this says that
Hp (M(x1 , . . . , xs−1 )) = 0, and thus that Hp+k (M(x1 , . . . , xt )) = 0 for all k ≥ 0
and all 1 ≤ t ≤ s − 1. So it only remains to show that Hp+k (M(x1 , . . . , xs )) = 0
for all k > 0. But again using the exactness of (E(M )), this follows from the
vanishing of Hp+k (M(x1 , . . . , xs−1 )) for all k ≥ 0.
The proof of part (2) follows easily by induction on s, and from part (1). For if
H1 (M(x1 , . . . , xs )) = 0, then H1 (M(x1 , . . . , xs−1 )) = 0, and our induction hypo-
thesis tells us that x1 , . . . , xs−1 is an M -sequence. We again apply the exactness
of (E(M )) to conclude that
x
H1 (M(x1 , . . . , xs )) → H0 (M(x1 , . . . , xs−1 )) →s H0 (M(x1 , . . . , xs−1 ))
is exact. Since H0 (M(x1 , . . . , xs−1 )) is equal to M/(x1 , . . . , xs−1 )M, and since
H1 (M(x1 , . . . , xs )) = 0, we may conclude that xs is not a zero divisor on
M/(x1 , . . . , xs−1 )M. Thus we have established part 2. 2
Corollary II.1.10 The hypotheses being as above, we have that x1 , . . . , xs is an
M -sequence if and only if xπ(1) , . . . , xπ(s) is an M -sequence for every permutation
π of {1, . . . , s}.
Corollary II.1.11 If R is a noetherian ring, M a finitely generated R-module,
p a prime ideal of R such that Mp = M ⊗R Rp = 0 (i.e. p is in Supp(M )), and
x1 , . . . , xs a sequence of elements in p which is an M -sequence, then x1 , . . . , xs ,
considered as elements in Rp , is an Mp -sequence.

II.2 Local rings


We shall now assume that R is a local ring, with maximal ideal m. The residue
field R/m will be denoted by k; all R-modules will be assumed finitely generated.
In this case, the radical of R is m, and all proper ideals of R are contained in m.
If E is an R-module and I an ideal of R, we have E/IE = 0 if and only if E = 0
because of Nakayama’s Lemma. Furthermore, if e1 , . . . , et are elements of E such
that e1 , . . . , et generate E/mE as a k-vector space, then e1 , . . . , et generate E as
an R-module.
Remark II.2.1 Any generating set of an R-module E contains a minimal gen-
erating set of E; any two minimal generating sets of E have the same number of
44 Local ring theory

elements. This number is equal to the dimension of the vector space E/mE over
k, and is denoted by [E/mE : k]. Moreover, any subset of E linearly independent
modulo mE may be extended to a minimal generating set of E.
Lemma II.2.2 If E is an R-module, there exists a free module F and an
epimorphism g : F → E such that ker(g) is contained in mF.
Proof The proof depends on nothing deeper than the fact that an epimorphism
of a vector space onto another of the same dimension is an isomorphism. One
then chooses F to be free on the number of generators in a minimal generating
set of E. 2
Lemma II.2.3 If E is an R-module such that TorR
1 (k, E) = 0, then E is free.

Proof We use the above remark to obtain an exact sequence


0→K→F →E→0
with F free and K ⊂ mF. Tensoring this exact sequence with k = R/m, we get
the exact sequence

1 (k, E) → K/mK → F/mF → E/mE → 0.


TorR

Since the map F/mF → E/mE is an isomorphism, the map TorR 1 (k, E) →
K/mK is surjective. But here we can invoke the fact that TorR
1 (k, E) = 0 to
conclude that K/mK = 0 and, by Nakayama, K itself must be zero. 2
Lemma II.2.4 For every R-module E, hdR (E) ≤ hdR (k).
Lemma II.2.5 If R is a local ring, gldimR = hdR k.
Recall that the global dimension of R, gldimR, is defined to be sup hdR (M )
M
as M runs through all R-modules.
Lemmas II.2.4 and II.2.5 are easy consequences of Lemma II.2.3.
The main result we are heading for now is that [m/m2 : k] ≤ gldimR, that
is, that the global dimension of R is never less than the minimal number of
generators of the maximal ideal of R. The idea of the proof is the following.
We take a minimal generating set x1 , . . . , xn of the maximal ideal m, and form
the complex K(x1 , . . . , xn ). We then show that we can find a free resolution of k :
d d d
· · · → X3 →3 X2 →2 X1 →1 R → k → 0
such that di (Xi ) ⊂ mXi−1 and such that K(x1 , . . . , xn ) is a subcomplex of
this resolution. Tensoring this resolution with k makes the boundary maps
zero, so that TorR p (k, k) = Xp /mXp . Since K(x1 , . . . , xn ) is a subcomplex
of the resolution, and K(x1 , . . . , xn ) is non-zero in dimension n, we see that
TorRn (k, k) = 0. Thus hdR k ≥ n and we have the desired result using the facts
that gldimR = hdR k and n = [m/m2 : k].
Local rings 45

The complex K(x1 , . . . , xn ) is of the form:


δ 2 δ f
0 → Λn R n →
n
· · · → Λ2 R n → Rn → R.
We may therefore choose X1 of our resolution to be Rn with d1 : X1 → R
defined to be the morphism f. Assume now that X1 , . . . , Xp have been defined
such that
1. Xl = Xl ⊕ Λl Rn ;
2. the morphism dl : Xl → Xl−1 restricted to Λl Rn is δl ;
3. ker(dl ) ⊂ mXl ;
dp d d
4. Xp → · · · → X2 →2 X1 →1 R → k → 0 is exact.
If p < n, we will show that we can find a free module Xp+1 such that 1–4 are
true with l replaced by p + 1. Our conditions 1 and 2 tell us that the composition
δp+1 dp
Λp+1 Rn → Λp Rn → Xp → Xp−1 is zero. Thus δp+1 (Λp+1 Rn ) is contained in
Zp = ker(dp ). We will show that if {εi1 ∧ · · · ∧ εip+1 } are a basis of Λp+1 Rn , then
{δp+1 (εi1 ∧· ·
·∧εip+1 )} are linearly independent modulo mZp . That is, we want to
show that if γi1 ...ip+1 δp+1 (εi1 ∧ · · · ∧ εip+1
 ) ∈ mZp , then each γi1 ...ip+1 ∈ m. But
this is the same as showing that if γ = γi1 ...ip+1 εi1 ∧ · · · ∧ εip+1 in Λp+1 Rn is
such that δp+1 (γ) is in mZp , then γ is in mΛp+1 Rn . Suppose, then, that δp+1 (γ)
is in mZp . Since Zp ⊂ mXp , we have mZp ⊂ m2 Xp = m2 Λp Rn ⊕ m2 Xp . Since
δp+1 (γ) is in Λp Rn , we must have δp+1 (γ) ∈ m2 Λp Rn .
Now δp+1 : Λp+1 Rn → Λp Rn induces a map δ p+1 :Λp+1 Rn /mΛp+1 Rn →
mΛp Rn /m2 Λp Rn since δp+1 (Λp+1 Rn ) ⊂ mΛp Rn . If we can show that δ p+1 is a
monomorphism, we are done. For to say that δp+1 (γ) is in m2 Λp Rn is to say that
δ p+1 (γ) = 0 and thus that γ = 0 or γ ∈ mΛp+1 Rn . But Λp+1 Rn /mΛp+1 Rn =
R/m⊗R Λp+1 Rn , and mΛp Rn /m2 Λp Rn = m/m2 ⊗R Λp Rn and the map δ p+1 sends
 ∧
the basis element 1 ⊗ εi1 ∧ · · · ∧ εip+1 to (−1)j+1 xij ⊗ εi1 ∧ · · · ∧ εij ∧ · · · ∧ εip+1 .
Clearly this map is a monomorphism.
Since we have shown that {δp+1 (εi1 ∧ · · · ∧ εip+1 )} are linearly independent
modulo mZp , we know that these elements may be chosen as part of a minimal
generating set for Zp , that is, we have a minimal generating set {δp+1 (εi1 ∧ · · · ∧

εip+1 )} ∪ {z1, . . . , zq }. Letting Xp+1 be the free module on q generators, we have

Λp+1 Rn ⊕ Xp+1 = Xp+1 mapping onto Zp . The kernel of the map Xp+1 →
Zp → Xp is clearly in mXp+1 and
dp+1 d d
Xp+1 → Xp → · · · → X2 →2 X1 →1 R → k → 0
is exact. Thus we have completed the inductive step, and we have proved
Theorem II.2.6 If R is a local ring, then [m/m2 : k] ≤ gldimR.
Our next objective is to define the dimension (or Krull dimension) of a local
ring R, and show that dim R ≤ [m/m2 : k]. To do this, we shall use the method
46 Local ring theory

of Hilbert–Samuel polynomials which we introduce in a slightly more general


setting now.

II.3 Hilbert–Samuel polynomials


We let Z+ denote the set of positive integers, and let G be an abelian group.
Generally, G will be the group of integers or the Grothendieck group of some
category of modules.

Definition II.3.1 Let f : Z+ → G be a function. f is called a polynomial


 for all nn∈ Z , n
+
function if there are elements a0 , . . . , ad in G 
such that
d n
sufficiently large (i.e. n >> 0), we have f (n) = i=0 i ai where i denotes
the binomial coefficient n!/(i!(n − i)!). If f : Z+ → G is any function, we define
∆f : Z+ → G by ∆f (n) = f (n + 1) − f (n). ∆f is called the first difference
(function) of f. For any integer s > 1, we define ∆s f = ∆(∆s−1 f ).

Lemma II.3.2 A function f : Z+ → G is a polynomial function if and only if


∆f is a polynomial function.
d n
Proof If f is a polynomial function, we have f (n) = i=0 i ai for some
a0 , . . . , ad in G, and all n >> 0. Consequently, for all n >> 0 we have ∆f (n) =
f (n + 1) − f (n) is equal to
d 
  d  
 d 
    d  
n+1 n n+1 n n
ai − ai = − ai = ai .
i=0
i i=0
i i=0
i i i=1
i−1

Thus ∆f is a polynomial function.


d n
Conversely, if ∆f (n) = i=0 i bi for n >> 0, define g(n) = f (n) −
d  n 
i=0 i+1 bi . Then for n >> 0, ∆g(n) = 0 so that g(n) is a constant a0 . Thus,
d+1  
letting ai = bi−1 for i > 0, we have f (n) = i=0 ni ai for n >> 0. 2

We have proved Lemma II.3.2 in detail, since this definition of polynomial


function is slightly more general than the usual one. However, with Lemma II.3.2
established, we shall only state three facts we need and omit the proofs.

d   d  
Lemma II.3.3 If f (n) = i=0 ni ai = i=0 ni ai is a polynomial function,
then d = d and ai = ai for all i. Thus, the degree, d, of f is a well-defined
integer, and the coefficients ai of f are uniquely determined.

Now consider a commutative noetherian ring, R, a set of indeterminates,


{X1 , X2 , . . .}, and let A be a full abelian subcategory of the category of R-
modules. For each integer s = 0, 1, 2, . . . , let As be the category of finitely
Hilbert–Samuel polynomials 47


generated graded R[X1 , . . . , Xs ]-modules E = ν≥0 Eν such that
1. if s = 0, Eν is in A for all ν;
2. if s> 0, Eν is in A for  all ν and the graded R[X1 , . . . , Xs−1]-modules
 Xs   Xs 
ker ν≥0 Eν → ν≥0 Eν and coker ν≥0 Eν → ν≥0 Eν are in
As−1 .
Finally, let f0 be a function from the objects of A to an abelian group G which
factors through the Grothendieck group, K(A), of A, that is, f0 is additive with
respect to exact sequences in A. Then

Theorem II.3.4 Let E = ν≥0 Eν be an object in As and define fE : Z+ → G
by fE (ν) = f0 (Eν ). Then fE is a polynomial function of degree less than or equal
to s − 1.
An example of such a set-up occurs when R is a local ring, and A is the full
subcategory
 of R-modules of finite length. In that case, an R[X1 , . . . , Xs ]-module
E = ν≥0 Eν is in As if it is finitely generated, and if each Eν is an R-module
of finite length. We may then choose G to be the group, Z, of integers and f0 (E)
equals the length of E. This is of course the most usual example. To see that
it is not the only example, we may choose A to be the category of all finitely
generated R-modules, and G to be K(A) itself with f0 the usual map.
Corollary II.3.5 If R is a local ring, E a finitely generated R-module and q
an ideal of R containing some power mn of the maximal ideal m, then E/qν E
is an R-module of finite length, and the function χq (E) : Z+ → Z defined by
χq (E; ν) = length(E/qν E) is a polynomial function whose degree is less than or
equal to [q/mq : k].
To see why this is a corollary of Theorem II.3.4, we observe that we have exact
sequences
0 → qν E/qν+1 E → E/qν+1 E → E/qν E → 0
and therefore ∆χq (E; ν) equals the length of qν E/qν+1 E. The graded mod-
ule qν E/qν+1 E is a module over R[X1 , . . . , Xs ] where s = [q/mq : k], and
each q E/qν+1 E is an R-module of finite length. Thus ∆χq (E) is a polynomial
ν

function of degree ≤ s − 1, and hence χq (E) is what we claimed it to be.


Remark II.3.6 If q1 and q2 are two ideals containing some powers of the max-
imal ideal m, that is, if q1 ⊃ mn1 and q2 ⊃ mn2 , then the degrees of χq1 (E) and
χq2 (E) are equal. In particular, they are all equal to the degree of χm (E). This
is seen easily by observing that for any integer n, χmn (E) and χm (E) have the
same degree. Thus, if mn ⊂ q ⊂ m, we must have deg(χmn (E)) ≤ deg(χq (E)) ≤
deg(χm (E)), and hence equality.
Definition II.3.7 If R is a local ring, the dimension of R is the degree of the
polynomial function χm (R). The dimension of an R-module E is the degree
48 Local ring theory

of the polynomial function χm (E). If R is a noetherian ring, and p is a prime


ideal of R, then the height of p is the dimension of Rp .

We immediately obtain the following proposition.

Proposition II.3.8 Let R be a local ring, and let s be the smallest number of
elements required to generate an ideal q of R which contains some power of the
2
maximal ideal m. Then dim R ≤ s. In particular, dim R ≤ [m/m : k].

Although we will not be using this fact immediately, it is important to note


that for a local ring R, the following integers
(a) the dimension of R;
(b) the smallest number of elements required to generate an ideal containing a
power of the maximal ideal;
(c) the length of the longest chain of prime ideals in R, where the length of the
chain p0 ⊃ p1 ⊃ · · · ⊃ ph is h
are equal.
The proof that these three integers are equal is not completely trivial. If we
denote these integers by d, s and h, respectively, the procedure is to show that
d ≤ h ≤ s ≤ d. We point out that an ideal which contains a power of the
maximal ideal is called an ideal of definition. It is the same thing as being an
m-primary ideal. A set of elements, x1 , . . . , xd , with d = dimR, is called a system
of parameters if the ideal they generate, q, is an ideal of definition. The reader
is referred to D. Eisenbud’s book, [41], for those details about noetherian rings
which are mentioned here but not proved.
For various reasons, it is useful to be able to compare the dimension of a
module E over a local ring R with that of E/xE if x is an element of R. From
our point of view, it is extremely helpful to know what happens to dim E/xE
when x is not a zero divisor for E. To handle this situation, we quote (without
proof) the Artin–Rees Theorem:

Theorem II.3.9 Let R be a noetherian ring, M a finitely generated R-module,


M  a submodule of M, and I an ideal of R. Then there exists an integer h > 0
such that for all n ≥ h, we have

(I n M ) ∩ M  = I n−h (I h M ∩ M  ).

Corollary II.3.10 If R is a local ring and x is an element of the maximal


ideal m which is a non-zero divisor for an R-module E, there is an integer h > 0
such that for all ν ≥ h, mν E : x ⊂ mν−h E where mν E : x = {e ∈ E| xe ∈ mν E}.

The proof depends on the Artin–Rees Theorem and the easy observation that

x(mν E : x) = mν E ∩ (x)E.
Hilbert–Samuel polynomials 49

For then we have


x(mν E : x) = mν E ∩ (x)E = mν−h (mh E ∩ (x)E)
= xmν−h (mh E : x) ⊂ xmν−h E,
from which we conclude: mν E : x ⊂ mν−h E.
Proposition II.3.11 Let R be a local ring, E an R-module, and x an element
of m. Then dim E/xE ≥ dim E − 1. If x is not a zero divisor for E, we have
dim E/xE ≤ dim E − 1 and hence dim E/xE = dim E − 1.
Proof If we let E = E/xE, and m = m/(x), we are interested in the length of
E/mν E. But E/mν E = E/(mν , x)E, and we have an exact sequence:
0 → (mν , x)E/mν E → E/mν E → E/(mν , x)E → 0.
Thus length(E/(mν , x)E) = length(E/mν E) − length((mν , x)E/mν E). Since
(mν , x)E/mν E ∼= xE/mν E ∩ xE ∼ = xE/x(mν E : x), we have length(mν , x)
E/m E ≤ length(E/(m E : x)), with equality holding if x is not a zero divisor for
ν ν

E. Since x ∈ m, we know that mν−1 E ⊂ mν E : x so that length(E/(mν E : x)) ≤


length(E/mν−1 E). Thus
length(E/(mν , x)E) = length(E/mν E) − length((mν , x)E/mν E)
≥ length(E/mν E) − length((E/mν+1 E)),
so that dim E ≥ dim E − 1. If, however, x is not a zero divisor for E, we have
length((mν , x)E/mν E) = length(E/(mν E : x))
and by Corollary II.3.10, length(E/(mν E : x)) ≥ length(E/mν−h E) for suitable
fixed h and all ν ≥ h. Thus
length(E/(mν , x)E) = length(E/(mν E)) − length(mν , x)E/mν E
≤ length(E/(mν E)) − length(E/mν−h E),
which immediately implies that dim E ≤ dim E − 1. 2
In the above proof, we said we were interested in the length of E/mν E instead
of the length of E/mν E to emphasize the fact that dim E as an R-module is the
same as dim E as an R-module, where R = R/(x). This is mainly to ensure that
when E = R, dim R as an R-module is seen to be the dimension of the local ring
R. Of course, E/mν E ∼ = E/mν E.
As an immediate consequence of Proposition II.3.11 we have
Corollary II.3.12 Let R be a local ring, E an R-module, and x1 , . . . , xs an
E-sequence. Then
dim(E/(x1 , . . . , xs )E) = dim E − s.
In particular, s ≤ dim E.
50 Local ring theory

II.4 Codimension and finitistic global dimension


We will now let R be a noetherian ring (i.e., not necessarily local), E a finitely
generated R-module, and I an ideal in R such that E/IE = 0. Since E/IE = 0,
there is some prime ideal p in R such that E/IE ⊗R Rp = 0, and this prime
ideal p obviously contains I. From Corollary II.1.10, we see that if x1 , . . . , xs
is an E-sequence contained in I, then x1 , . . . , xs is also an E ⊗R Rp -sequence
in IRp and thus, by Corollary II.3.12, s ≤ dim E ⊗R Rp . Hence the number of
elements in an E-sequence contained in I is bounded, and any E-sequence in I
may be extended to a maximal E-sequence in I. We shall prove that any two
maximal E-sequences in I have the same number of elements, but first we give
a quick review of the notion of associated prime ideals. The Definitions II.4.1
and II.4.5, the Lemmas II.4.2 and II.4.3, Proposition II.4.4, Theorem II.4.6 and
Remark II.4.7 can all be found in Reference [41]. We simply write them here for
easy reference.
Definition II.4.1 If E is an R-module, a prime ideal p is said to be associ-
ated to E if there is some monomorphism R/p → E. The associator of E,
denoted by Ass(E), is the set of all primes associated to E.
Using the fact that R is noetherian, we have
Lemma II.4.2 If E is an R-module, then Ass(E) = ∅ if and only if E = 0.
Another useful lemma is the following:
Lemma II.4.3 Let
0 → E  → E → E  → 0
be an exact sequence of R-modules. Then
Ass(E) ⊂ Ass(E  ) ∪ Ass(E  ).
Proposition II.4.4 If R is a noetherian ring and E is a finitely generated
R-module, then Ass(E) is a finite set.
Definition II.4.5 If R is a noetherian ring and E a finitely generated R-
module, we define the height of E to be min{height p} where p runs through
all primes in Ass(E). If I is an ideal of R, we generally abuse notation and call
the height of R/I the height of the ideal I.
An important result relating the height of an ideal with the number of gen-
erators of the ideal, is the Krull Principal Ideal Theorem. We state this
separately as a theorem.
Theorem II.4.6 (Krull) If R is a noetherian ring, and I is an ideal generated
by elements x1 , . . . , xr , then any minimal prime of Ass(R/I) has height less than
or equal to r. In particular, height(I) ≤ r.
Codimension and finitistic global dimension 51

Remark II.4.7
(1) For any R-module E, the set of zero divisors of E is equal to the union of
all primes in Ass(E).
(2) Thus, if E is a finitely generated R-module (with R noetherian) and if I
is an ideal such that every element of I is a zero divisor of E, then there is a
non-zero element e of E such that Ie = 0. For if I is as above, then I ⊂ ∪p where
p runs through Ass(E) and hence I ⊂ p for some p ∈ Ass(E) (since Ass(E) is
finite). We therefore have the epimorphism R/I → R/p and a monomorphism
R/p → E. The image of 1 under the composite map R → R/I → R/p → E is
the desired element e ∈ E.
(3) If the ideal I in 2 above is a maximal ideal, then I ∈ Ass(E).
We are now ready to prove the main theorem of this section.
Theorem II.4.8 Let R be a noetherian ring, E an R-module, and I an ideal
of R generated by elements x1 , . . . , xn such that E/IE = 0. Let y1 , . . . , ys be a
maximal E-sequence in I. Then s+q = n, where q is the dimension of the highest
non-vanishing homology of the complex E(x1 , . . . , xn ). Furthermore,
Hq (E(x1 , . . . , xn )) ∼
= ((y1 , . . . , ys )E : I)/(y1 , . . . , ys )E.
Proof The proof proceeds by induction on s. When s = 0, it means that every
element of I is a zero divisor for E so that by 2 of Remark II.4.7, there is a non-
zero element e in E such that Ie = 0. Since Hn (E(x1 , . . . , xn )) = 0 : I = 0, we
see that q = n and Hn (E(x1 , . . . , xn )) = (y1 , . . . , ys )E : I/(y1 , . . . , ys )E = 0 : I.
When s > 0, we consider the exact sequence
y1
0→E→E→E→0
and get the long exact sequence
y1
Hq+1 (E(x1 , . . . , xn )) → Hq (E(x1 , . . . , xn )) → Hq (E(x1 , . . . , xn ))
→ Hq (E(x1 , . . . , xn )) → · · ·
where q is the dimension of the highest non-vanishing homology group of
E(x1 , . . . , xn ). Thus, Hq+1 (E(x1 , . . . , xn )) = 0, Hq (E(x1 , . . . , xn )) = 0, and
multiplication by y1 on Hq (E(x1 , . . . , xn )) and Hq−1 (E(x1 , . . . , xn )) is zero
since y1 is in I. Thus Hq (E(x1 , . . . , xn )) = 0 while Hq−1 (E(x1 , . . . , xn )) ∼ =
Hq (E(x1 , . . . , xn )). Noting that y2 , . . . , ys is a maximal E-sequence in I, and
using induction, we have (s−1) +q = n. Since, however, we have just shown that
q = q − 1, we have s + q = n. Finally, since Hq (E(x1 , . . . , xn )) ∼ = (y2 , . . . , ys )E :
I/(y2 , . . . , ys )E ∼= (y1 , . . . , ys )E : I/(y1 , . . . , ys )E, and since, in addition,
Hq (E(x1 , . . . , xn )) = Hq−1 (E(x1 , . . . , xn )) ∼ = Hq (E(x1 , . . . , xn )), we have

Hq (E(x1 , . . . , xn )) = (y1 , . . . , ys )E : I/(y1 , . . . , ys )E.
2
52 Local ring theory

Corollary II.4.9 If R, E and I are as in Theorem II.4.8, then any two


maximal E-sequences in I have the same length.
Definition II.4.10 The length of a maximal E-sequence in I is called the
I-depth of E, denoted by depth(I; E). If R is a local ring, the codimension
of a module, E, is defined as codim(E) = depth(m; E) where m is the maximal
ideal of R. If E = R, we simply write depth(I) for depth(I; R), and call it the
depth of I. Finally, we call the ideal I perfect if hdR (R/I) = depth(I).
Remark II.4.11 If I is any ideal of a noetherian ring R, then depth(I) ≤
height(I).
An important consequence of Theorem II.4.8 is
Theorem II.4.12 Let R be a local ring, and E an R-module such that
hdR E < ∞. Then
codim(R) = hdR E + codim(E).
Proof Let m = (x1 , . . . , xn ), where m is the maximal ideal of R. Denote by q the
dimension of the highest non-vanishing homology group of K(x1 , . . . , xn ), and by
qE the corresponding integer for the complex E(x1 , . . . , xn ). Since codim(R) =
n − q and codim(E) = n − qE , we want to show that qE − q = hdR E.
Now when hdR E = 0, E is free and clearly qE = q. Suppose that hdR E ≥ 1.
Then we have an exact sequence
0→L→F →E→0
with F a free module, and hdR E = 1 + hdR L. This gives us an exact sequence:
H1+qL (E(x1 , . . . , xn )) → HqL (L(x1 , . . . , xn )) → HqL (F(x1 , . . . , xn ))
→ HqL (E(x1 , . . . , xn )) → HqL −1 (L(x1 , . . . , xn )).
By induction on hdR E, we know that qL − q = hdR L so what must be shown
is that 1 + qL = qE . If qL − q > 0, then HqL (F(x1 , . . . , xn )) = 0 and thus
H1+qL (E(x1 , . . . , xn )) = 0, while clearly Ht+qL (E(x1 , . . . , xn )) = 0 for all t > 1.
Thus in this case, we would have 1 + qL = qE and we would be done. Our
problem, then, is to resolve the case when qL = q, that is, when hdR L = 0 or
L is free. In this case, we may assume that L and F have been chosen so that
L ⊂ mF.
If y1 , . . . , ys is a maximal R-sequence, it is also a maximal F - and L-sequence,
and Theorem II.4.8 tells us that
Hq (L(x1 , . . . , xn )) ∼
= (y1 , . . . , ys )L : m/(y1 , . . . , ys )L,
∼ (y1 , . . . , ys )F : m/(y1 , . . . , ys )F,
Hq (F(x1 , . . . , xn )) =
and the map Hq (L(x1 , . . . , xn )) → Hq (F(x1 , . . . , xn )) is the natural map of
(y1 , . . . , ys )L : m/(y1 , . . . , ys )L → (y1 , . . . , ys )F : m/(y1 , . . . , ys )F. If we show
that this map is not a monomorphism, we are done.
Codimension and finitistic global dimension 53

Since (y1 , . . . , ys ) : m (y1 , . . . , ys ), there is a z ∈/ (y1 , . . . , ys ) with zm ⊂


(y1 , . . . , ys ). Since L is free, zL is not included in (y1 , . . . , ys )L. But zL ⊂
zmF ⊂ (y1 , . . . , ys )F. Choosing an element w ∈ zL − (y1 , . . . , ys )L, we have
w ∈ (y1 , . . . , ys )L : m but w ∈ / (y1 , . . . , ys )L and thus w = 0 in (y1 , . . . , ys )L :
m/(y1 , . . . , ys )L. However, w is mapped to 0 in (y1 , . . . , ys )F : m/(y1 , . . . , ys )F so
that our map Hq (L(x1 , . . . , xn )) → Hq (F(x1 , . . . , xn )) is not a monomorphism,
and the proof is complete. 2
A particular consequence of Theorem II.4.12 is that if R is a local ring, and
E an R-module such that hdR E < ∞, then hdR E ≤ codimR.
Definition II.4.13 If R is a commutative ring, we define the finitistic global
dimension of R, written fgldimR, to be sup{hdR E} where E ranges over all
finitely generated R-modules of finite homological dimension.
Lemma II.4.14 If R is a local ring, E a finitely generated R-module, and
x ∈ m an element which is not a zero divisor for E, then hdR E/xE = 1+hdR E.
Proof Applying TorR (−, R/m) to the short exact sequence
x
0 → E → E → E/xE → 0,
it is easy to see that E and E/xE both have finite homological dimension,or
both have infinite homological dimension. In the latter case, the lemma is clearly
true. In the former, the long exact Tor sequence exhibits the equality. 2
It may be amusing to look at a slightly different proof of this result. One can
choose a minimal resolution, X, of E, and consider the map Fx : X → X given
by multiplication by x. (A resolution is called minimal if the image of Xp is
contained in mXp−1 for all p.) Since x ∈ m, it is easy to see that the mapping
cone of Fx is a minimal resolution of E/xE. Now a simple look at the mapping
cone gives the result.
This other proof can be looked on as an extension of the ideas behind the
use of the Koszul complex. It was inspired by a more special theorem found in
W. Gröbner’s work [46], and was one of the early results that indicated that
homological methods could be effectively applied to certain kinds of problems in
commutative algebra.
Proposition II.4.15 If R is a local ring, then fgldimR = codimR.
Proof We have already seen that fgldimR ≤ codimR. However, if y1 , . . . , ys is
an R-sequence with s = codimR, we have (by Lemma II.4.14)
hdR R/(y1 , . . . , ys ) = s = codimR
and hence the equality. 2
Putting together Theorem II.2.6, Proposition II.3.8, Corollary II.3.12, and
Proposition II.4.15, we obtain
54 Local ring theory

Theorem II.4.16 If R is a local ring, we have


2
fgldimR = codimR ≤ dim R ≤ [m/m : k] ≤ gldimR.

II.5 Regular local rings


We are now ready to apply what we have done to the study of some important
properties of regular local rings.
2
Definition II.5.1 A local ring R is regular if dim R = [m/m : k].
The following is a well-known property of regular local rings.
Proposition II.5.2 Let R be a regular local ring, and x1 , . . . , xn a minimal
generating set of the maximal ideal m. Then for each integer i ≥ 0, the ideal
(x1 , . . . , xi ) is a prime ideal. In particular, R is a domain and x1 , . . . , xn is a
(maximal) R-sequence.
Proof That the ideals (x1 , . . . , xi ) are prime for i ≥ 0, can be found in Refer-
ence [41], section 10.3. That x1 , . . . , xn is a maximal R-sequence is then an easy
consequence. 2
The prototypical example of a regular local ring is the formal power series ring
over a field: K[[X1 , . . . , Xn ]].
Our first main result about regular local rings is:
Theorem II.5.3 A local ring R is regular if and only if gldimR < ∞. If R is
regular, gldimR = dim R.

Proof If R is regular, Proposition II.5.2 tells us that hdR k = hdR R/


(x1 , . . . , xn ) equals n = dim R, and hence gldimR = n < ∞. Conversely,
if gldimR < ∞, then gldimR = fgldimR and by Theorem II.4.16, we have
2
dim R = [m/m : k]. 2
The utility of Theorem II.5.3 will be seen in the subsequent theorems. The
result that comes immediately below was actually the prime motivation for one
of the authors (prompted by a conversation with Emil Artin) to establish this
characterization of regular local rings.
Theorem II.5.4 If R is a regular local ring and p is a prime ideal of R, then
Rp is a regular local ring.
Proof gldimRp = hdRp Rp /pRp ≤ hdR R/p < ∞. Thus Rp is regular. 2
Lemma II.5.5 If R is a regular local ring, E a finitely generated R-module,
and p ∈ Ass(E), then hdR E ≥ height p.
Proof We have hdR E ≥ hdRp Ep . Since pRp is obviously in Ass(E), and pRp is
the maximal ideal of Rp , we have codimEp = 0 as an Rp -module. Thus hdRp Ep =
codimRp = dim Rp = heightp. 2
Regular local rings 55

A straightforward application of the Krull Principal Ideal Theorem


(Theorem II.4.6) yields the following.
Lemma II.5.6 If R is a noetherian ring, I an ideal of R, and x an ele-
ment of R such that x is not a zero divisor for R/I, and R/(I, x) = 0, then
height(I, x) ≥ 1 + heightI. In particular, if x1 , . . . , xs is an R-sequence, then
height(x1 , . . . , xs ) = s.
Lemma II.5.7 Let I be an ideal in a regular local ring R. Then depth(I; R) =
heightI.
Proof By localizing we know that depth(I; R) ≤ heightI. Suppose heightI = s
and let x1 , . . . , xt be a maximal R-sequence in I, that is, t = depth(I; R).
Since every element of I is a zero divisor for R/(x1 , . . . , xt ), we must have
I contained in some prime ideal p ∈ Ass(R/(x1 , . . . , xt )). But, by Lemma II.5.5,
if p ∈ Ass(R/(x1 , . . . , xt )) we have heightp ≤ hdR R/(x1 , . . . , xt ) = t. Thus, since
I ⊂ p implies heightI ≤ heightp, we have s ≤ t and thus s = t. 2
It is not hard to see the following result.
Lemma II.5.8 Let R be a noetherian ring, E a finitely generated R-module,
and x a non-zero divisor for E. If p is in Ass(E) and p is a prime ideal containing
(p, x), then there is a prime ideal p ∈ Ass(E/xE) such that p ⊃ p ⊃ p.
Definition II.5.9 We say a chain of prime ideals p0 ⊂ p1 ⊂ · · · ⊂ pn is
saturated if for each i there is no prime lying strictly between pi and pi+1 . We
say a ring satisfies the saturated chain condition for prime ideals (s.c.c.)
if any two saturated chains of primes between any two given primes p ⊂ p have
the same length.
Theorem II.5.10 Every regular local ring satisfies the saturated chain condi-
tion for prime ideals.
Proof It obviously suffices to show that if p ⊂ p is saturated, then heightp =
1 + heightp. Suppose, then, that s = heightp. Then there is an R-sequence
x1 , . . . , xs which is maximal in p. Clearly there is an element xs+1 in p − p such
that x1 , . . . , xs+1 is an R-sequence. Using Lemma II.5.8, there is a prime p ∈
Ass(R/(x1 , . . . , xs+1 )) such that p ⊃ p ⊃ p. Since heightp = s + 1 > heightp,
we cannot have p = p. Thus p = p and heightp = s+1 = 1+heightp. 2
Remark II.5.11
1. By constructing a local ring that does not satisfy the s.c.c., M. Nagata was
able to show that not every local ring is a factor ring of a regular local ring.
For clearly, a factor ring of a ring satisfying the s.c.c. must also satisfy the
s.c.c.
2. We see that if R is a regular local ring, then dim R/p+heightp = n = dim R.
3. It can be shown that if R satisfies the s.c.c., if I is an ideal of R, and x an
element of R, then height(I, x) ≤ 1 + heightI.
We now come to the Cohen–Macaulay Theorem.
56 Local ring theory

Theorem II.5.12 (Cohen–Macaulay) Let R be a regular local ring, and


I = (x1 , . . . , xs ) an ideal of height s. Then I is unmixed (i.e., every prime
p ∈ Ass(R/I) has height s), and x1 , . . . , xs is an R-sequence.

Proof The main point of the proof is to show that x1 , . . . , xs is an R-sequence,


for then the rest follows easily. To show that x1 , . . . , xs is an R-sequence, we
proceed by induction. Certainly when s = 1 we are done (since a regular local
ring is an integral domain). Assuming s > 1, we have s = height(x1 , . . . , xs ) ≤
1+ height(x1 , . . . , xs−1 ). But then s − 1 ≤ height(x1 , . . . , xs−1 ) ≤ s − 1, so by
induction we have x1 , . . . , xs−1 is an R-sequence. It is then trivial to show that
xs is not a zero divisor for R/(x1 , . . . , xs−1 ) (by using a height argument), and
so we have the result. 2

The definition of unmixed given in the statement of Theorem II.5.12 holds


for any noetherian ring, R.

II.6 Unique factorization


In this section we outline a proof of the fact that every regular local ring is a
unique factorization domain. In Section IV.2 we will give a second proof of this
theorem. The fact that regular local rings are factorial was another one of the
early major results about local rings obtained using homological techniques, or
at least the homological characterization of regularity.
In order to discuss unique factorization in regular local rings, we first establish
a few results that extend our discussion of Section I.1.

Proposition II.6.1 If R is a noetherian local ring, then R is a unique


factorization domain if and only if hdR R/(x, y) ≤ 2 for every x, y ∈ R.

Proof In Section I.1 we have proved the implication one way (Corollary I.1.6).
It remains to prove that if hdR R/(x, y) ≤ 2 for every x, y ∈ R, then R is a unique
factorization domain.
First we see from Corollary I.2.10, that R is a domain, since if we pick x = y =
0, we have the finiteness of the homological dimension of R/(x), which guarantees
that x is not a zero divisor. But then, as in the proof of Corollary I.1.6, we see
that the sequence

0 → (x) : y → R2 → R → R/(x, y) → 0

is exact, which tells us that hdR ((x) : y) = 0. Hence this ideal is projective,
indeed free, hence principal, and we are done. 2

An immediate corollary of this proposition is the following.

Corollary II.6.2 If R is a regular local ring of dimension less than or equal


to 2, then R is a unique factorization domain.
Unique factorization 57

Proof If R is regular of dimension less than or equal to 2, then its global


dimension is at most 2. Hence the condition of the preceding proposition is
satisfied, and R is a unique factorization domain. 2
Proposition II.6.3 Let R be a noetherian domain. Then the following are
equivalent:
(1) R is a unique factorization domain.
(2) Every minimal prime ideal of R is principal.
(3) Every unmixed ideal of height one is principal.
Proof (1) ⇒ (2): If R is a unique factorization domain, then every irreducible
element is prime. Let p be a minimal prime ideal, and let (x) be maximal among
the principal ideals contained in p. Clearly x is an irreducible element. But then
it is prime. Since (x) ⊆ p, and p is a minimal prime, the two prime ideals are
equal.
(2) ⇒ (1): We must show that irreducible elements are prime. If x is irreducible,
let (x) ⊆ p for some minimal prime p. We are assuming that p is principal, say
p = (y). From the irreducibility of x, we see that x must be a unit multiple of y,
and we are done.
(1) ⇒ (3): Since we are assuming factoriality, we know that all minimal primes
are principal. We first show that if q is p-primary for some minimal prime ideal p,
then q is principal. Since p is principal, we have p = (a). Suppose that q ⊆ (an ),
but q  (an+1 ). Set q = q : an . It is easy to see that q = q (an ). Now, if q = R,
then it must be contained in p. But q ⊆ p, implies that an+1 ∈ q, and this is a
contradiction. Hence q = R, and we see that q = (an ).
Since every unmixed ideal of height one is the intersection of primaries belong-
ing to minimal primes, every such ideal is the intersection of principal ideals. We
know from Section I.1 that in a unique factorization domain, the intersection of
principal ideals is principal, and so we have the result.
Clearly (3) implies (2), so we have proven the proposition. 2
We now begin a series of propositions to prove that every regular local ring
of dimension 3 is a unique factorization domain. Following that, we shall use a
proof of M. Nagata [67] to show that this is sufficient to conclude that all regular
local rings are factorial.
Proposition II.6.4 Let (R, m) be a local ring. Let x be in m and y in R such
that a = (x) : y satisfies the following conditions:
(a) hdR a ≤ 1, and
(b) x is not in ma.
Then a = (x) and x is not a zero divisor .
Proof Suppose x does not generate a. Since x is not in ma, there exist a1 , . . . , an
in a (n > 0) such that x, a1 , . . . , an is a minimal generating set for a. Let Rn+1
58 Local ring theory

denote the direct sum of n+1 copies of R, define f : Rn+1 → a by f (r0 , . . . , rn ) =


n
r0 x + i=1 ri ai , and let K = kerf . Since x, a1 , . . . , an is a minimal generating
set for a, we have that f is an epimorphism and K is contained in mRn+1 . From
the exact sequence 0 → K → Rn+1 → a → 0 and the fact that hdR a ≤ 1, we
have that K is R-free. Since a is not principal, we have that hdR a = 1 and thus
K = 0.
Let Ni = (ti0 , . . . , tin ) be a free basis for K over R (i = 1, . . . , m). Since
a1 is in a, we have that ya1 = −vx for some v in R. Let V =(v, y, 0, . . . , 0)
m
and T = (a1 , −x, 0, . . . , 0). Then V and T are in K. Let V = j=1 rj Nj and
m m
T =
m j=1 sj Nj . Now xV = −yT. Therefore we have that j=1 xrj Nj =
j=1 −ysj Nj . Since the {Nj } are a free basis for K over R, we have that
xrj = −ysj for all j = 1, . . . , m. But (x) : y = a. Therefore we have that
m each sj is
m
in a. Since T = (a1 , −x, 0, . . . , 0) = j=1 sj Nj , it follows that −x = j=1 sj tj1 .
Therefore x is in ma (since K is contained in mRn+1 ) which contradicts the fact
that x is not in ma. Thus a = (x). The fact that x is not a zero divisor follows
from the fact that hdR a ≤ 1. (Recall the proof of Proposition II.6.1) 2
Corollary II.6.5 Suppose p is a prime ideal in R such that dimRp = 1 and
hdR/p ≤ 2. Then p is a principal ideal.
Proof Since hdR R/p ≥ hdRp Rp /pRp = gldimRp , it follows that the gldimRp
is finite. Therefore Rp is a regular local ring of dimension 1. Let x1 , . . . , xt be
a minimal generating set for p. Then the x1 , . . . , xt , considered as elements in
Rp , generate pRp . Since Rp is a regular local ring of dimension 1, we have that
pRp = xj Rp for some j. Let (xj ) = q1 ∩ · · · ∩ ql be an irredundant, primary
decomposition for (xj ). From pRp = xj Rp it follows that one of the qi is p. Let
us say q1 = p. Then for y in (q2 ∩ · · · ∩ ql ) − p we have that (xj ) : y = p. Since
xj is not in mp and hdR p ≤ 1, it follows from the previous proposition that
p = (xj ). 2
Theorem II.6.6 Let R be a local domain of dimension ≤ 3 such that hdR R/p
is finite for all minimal primes p. Then R is a unique factorization domain.
Proof Since R is a noetherian domain, it follows from our Proposition II.6.3
that it suffices to show that each minimal prime ideal is principal in order to
show that R is a unique factorization domain. But by Corollary II.6.5, it will
follow that a minimal prime ideal p is principal if we can show that hdR R/p ≤ 2.
Since hdR R/p < ∞ we have that hdR R/p + codimR/p = codimR ≤ dimR.
But codimR/p ≥ 1 and dimR ≤ 3. Thus hdR R/p ≤ 2, which completes the
proof. 2
Since every module has finite homological dimension over a regular local ring,
we have established
Corollary II.6.7 Every regular local ring of dimension ≤3 is a unique factor-
ization domain.
Multiplicity 59

Theorem II.6.8 Every regular local ring is a unique factorization domain.

Proof We want to prove that if R is a regular local ring of dimension d ≥ 4,


then it is factorial (recall that “factorial” is the same as “unique factorization
domain”). We assume, by induction, that all regular local rings of dimension less
than d are factorial. Following M. Nagata [67], we proceed in two steps. In the
first, we take a minimal prime ideal p, write it (the way we did in Corollary II.6.5)
as p = (x) : y, and show that hdR R/(x, y) ≤ d − 1. We then show that, with the
inductive assumption, if elements x, y ∈ R are such that hdR R/(x, y) ≤ d − 1,
then the ideal (x) : y is principal.
We let the integer t be such that (x, y) : mt = (x, y) : mt+1 , and set J =
(x, y) : mt . Then J : m = J. We may therefore choose an element u ∈ m, u ∈ / m2 ,
such that J : u = J. Since R̄ = R/(u) is regular local of dimension d − 1, it is
factorial, and therefore hdR̄ R/(x, y, u) ≤ 2.
From the fact that codimR̄ R/(x, y, u) + hdR̄ R/(x, y, u) = d − 1, we see that
codimR̄ R/(x, y, u) ≥ d − 3 > 0. Since codimR̄ R/(x, y, u) = codimR R/(x, y, u) is
positive, we see that (x, y, u) : m = (x, y, u) so that (x, y, u) ⊇ J. (To see this,
note that (x, y) ⊆ (x, y, u) so that J = (x, y) : mt ⊆ (x, y, u) : mt = (x, y, u).)
But this tells us that (x, y) ⊆ J ⊆ (x, y, u) and, since J : u = J, it follows easily
that (x, y) = J. Therefore we have shown that (x, y) : m = (x, y) which implies
that hdR R/(x, y) ≤ d − 1. We have therefore completed the first step of our
proof.
We now want to show that hdR R/(x, y) ≤ d−1 implies that (x) : y is principal.
From the exact sequence used in the proof of Proposition II.6.1, we see that
hdR R/((x) : y) ≤ d − 2, which implies that we can find an element v ∈ m − m2
such that ((x) : y) : v = (x) : y and ((x) : y, v) : m = ((x) : y, v).
Let q be an associated prime of ((x) : y, v). Since q = m, the local ring Rq
has dimension less than d, so it is factorial. Thus, in Rq the ideal (x) : y is
principal, so the ideal ((x) : y, v)Rq is generated by two elements which form
a regular Rq -sequence. Consequently q is of height 2 (by the Cohen–Macaulay
Theorem), and we see then that ((x) : y, v)/(v) ⊆ R/(v) is unmixed of height 1
in the regular local ring R/(v) (which is of dimension d − 1). But this tells us
that ((x) : y, v)/(v) is a principal ideal, that is, ((x) : y, v)/(v) = (c, v)/(v) for
some c ∈ (x) : y. We therefore have, in the ring R, that (c, v) = ((x) : y, v) and
(c) ⊆ (x) : y ⊆ ((x) : y, v) = (c, v), from which it follows that (c) = (x) : y.
Recalling that our original ideal p = (x) : y, we see that we have proved that p
is a principal ideal. This completes the proof of our theorem. 2

II.7 Multiplicity
In Section II.3, we proved that if R is a local ring and q an ideal of defini-
tion of R (i.e., an ideal which contains a power of the maximal ideal), then for
any finitely generated R-module, E, the function χq (E) defined by χq (E; n) =
length(E/qn E) is a polynomial function. In fact, we defined the dimension of
E (Definition II.3.7) to be the common degree of this polynomial for any ideal
60 Local ring theory

of definition, q, and we saw that it was always less than or equal to the dimen-
sion, d, of the ring, R. In this section, we will take a closer look at the leading
coefficient of these polynomials or, more precisely, the coefficient of the term of
degree d.
First, let us expand a bit on what  we did in Section II.3 on polynomial
functions. Define the polynomials Xi in Q[X] by
 
X X(X − 1) · · · (X − i + 1)
= .
i i!
It is easy to show that these polynomials form a linear basis for the polynomial
ring, Q[X], over the rationals, Q. They also help to explain why the term “poly-
nomial” enters into the definition  of “polynomial function”
  in Section II.3: we
think of the binomial coefficient ni as the polynomial Xi evaluated at n.
If f (n) is an integer-valued polynomial function, then this means that there
are integers a0 , . . . , ad such that
   
n n
f (n) = a0 + a1 + · · · + ad
1 d
for all n sufficiently
d large
 or that, for sufficiently large n, the function, f , and
the polynomial, i=0 ai Xi , agree. Notice that while the term ad is an integer,
the coefficient of X d is ad /d!. We will use this observation in our definition of
multiplicity.
While we will not be making use of many of the convenient properties of the
polynomials, Xi , we point out two of them (without proof) that the reader may
explore if moved by curiosity.
   
(a) −X i = (−1)i X+i−1
i ;
X+Y  i X  Y 
(b) i = j=0 j i−j .
Employing these, a seemingly complicated identity such as
     
n+i−j−1 n n+i−1
(−1)j =− ,
j>0
i−j j i

becomes trivial to prove.


We return now to a definition of multiplicity.
Definition II.7.1 Let R be a local ring with maximal ideal, m, E a finitely gen-
erated R-module, and q an ideal of definition of R. We define the multiplicity
of q with respect to E, written eE (q), by
eE (q) = d! c,
where c is the coefficient of X d in the polynomial χq (E). We define the
multiplicity of R to be eR (m).
This multiplicity is also sometimes called the Samuel multiplicity.
Multiplicity 61

We see immediately that, since the degree of χq (E) ≤ d, we have eE (q) = 0


if and only if dim(E) < d. It is also evident, from the definition of the Hilbert–
Samuel polynomial, that eE (q) ≥ 0.
Our next goal is to tie up the multiplicity just defined with the Euler–Poincaré
characteristic of a certain Koszul complex. To do this, we first outline a proof
of the following theorem. The full proof can be found in Reference [10], p. 395.
In fact, the theorem proven in Reference [10] is a bit more general than the one
stated here; as we do not need the more general statement, we will just give the
more restrictive result.

Theorem II.7.2 Let E be a finitely generated module over the local ring, R,
and let q be an ideal of definition generated by x1 , . . . , xs . Then

χ(H(E(x1 , . . . , xs ))) = ∆s χq (E; n)

for n sufficiently large, where χ(H(E(x1 , . . . , xs ))) denotes the Euler–Poincaré


characteristic of the homology of the Koszul complex.

Before we outline a proof of this result, we point out that the number, s, must
be at least equal to d, the dimension of R (Section II.3). Therefore, the result is
most significant when s = d, that is, when x1 , . . . , xs is a system of parameters,
and dim(E) = d.

Proof (Outline) We will denote by C the complex E(x1 , . . . , xs ):

C : 0 → Cs → Cs−1 → · · · → C1 → C0 ,

where Ci = Λi Rs ⊗R E.
We now consider the graded ring, R, defined by

R = R/q ⊕ q/q2 ⊕ · · · = qn /qn+1 .
n≥0

If we let x̄1 , . . . , x̄s be the cosets of x1 , . . . , xs in q/q2 , we see that R is isomorphic


to the graded ring R/q [x̄1 , . . . , x̄s ], that is, it is the image of a polynomial ring
in s linear variables over R/q.
Our strategy is to relate our Koszul complex over R to another graded Koszul
complex over this graded ring, R. This requires a few more definitions and
constructions.
Given our finitely generated R-module,  E, we obtain a finitely generated
n n+1
graded R-module, E, by setting E = n≥0 q E/q E. Finally, given our
complex, C, we note that the complex

C : 0 → C s → C s−1 → · · · → C 1 → C 0

is none other than the (graded) Koszul complex E(x̄1 , . . . , x̄s ).


62 Local ring theory

Since E is a finitely generated R-module, we see that Hi (C) is finitely gener-


ated. However, this homology is annihilated by (x̄1 , . . . , x̄s ), so that it is a finitely
generated module over R/q, and hence of finite length (since R/q is). But Hi (C)
is graded; it is the direct sum
 (l)
Hi (C) = Hi (C).
l≥0

(l)
Therefore, for sufficiently large l, we must have all Hi (C) = 0. If we do this for
each i, we arrive at an integer l0 with the property that for all l ≥ l0 , we have
(l)
Hi (C) = 0 for all i. It is this fact that we exploit in order to obtain our result.
And we exploit it by writing explicitly what this means, namely, the following
sequence is exact for sufficiently large l:

(∗) 0 → ql Cs /ql+1 Cs → ql+1 Cs−1 /ql+2 Cs−1 → · · · → ql+s C0 /ql+s+1 C0 → 0.

The exactness of (∗) allows us to prove the exactness of the complex

C(l) : 0 → ql Cs → ql+1 Cs−1 → · · · → ql+s C0 → 0,

which in turn implies that Hi (C) = Hi (C/C(l) ) for all sufficiently large l. Thus
these two complexes have equal Euler–Poincaré characteristics. However, it is
clear that the terms of the complexes C/C(l) are of finite length over R, so that
we have

χ(H(C)) = χ(H(C/C(l) )) = χ(C/C(l) ),

and this is clearly seen to be equal to ∆s χq (E; l).


The argument that the exactness of (∗) implies the exactness of the complex
C(l) is found, in more generality, in Reference [10]. 2

From the above result follow a number of others of interest. We will list some
of them, and refer the reader to Reference [9], sections 4–6, and Reference [78]
for proofs and more detail.
This next result is an immediate corollary of the above theorem, and is the
aim of the foregoing discussion.

Theorem II.7.3 If R is a local ring of dimension d, q an ideal of defini-


tion generated by a system of parameters, x1 , . . . , xd , and E a finitely generated
R-module, then

χ(H(E(x1 , . . . , xd )) = eE (q).

From this, it is relatively straightforward to prove the following result ([9],


corollary 4.3).

Theorem II.7.4 If R is a local ring of dimension d, q an ideal of defini-


tion generated by a system of parameters, x1 , . . . , xd , and E a finitely generated
Multiplicity 63

R-module, then

d−1
eE (q) = length(E/qE)−length(qd−1 E : xd /qd−1 E) − e(ql−1 E:xl /ql−1 E) (q/ql )
l=1

where ql = (x1 , . . . , xl ), l = 0, . . . , d − 1.
This theorem shows us that the multiplicity, eE (q), always exceeds the length
of E/qE. Further investigation into what accounts for this difference led to the
study of Macaulay modules and rings, a notion that we now define.
Definition II.7.5 An R-module, E, is called a Cohen–Macaulay module
(C–M for short) if codimE = dimR. The ring, R, is Cohen–Macaulay (C–M
for short) if, when considered as a module over itself, it is a C–M module.
The following two results provide us with some insight into the relationship
between the property of being C–M, unmixedness, and systems of parameters.
These are theorem 5.6 and proposition 5.7 of Reference [9].
Theorem II.7.6 If E is a C–M module, and x1 , . . . , xs s ≥ 0 is a set of
elements in m such that height(E/(x1 , . . . , xs )E) = s, then the submodule (0)
in E/(x1 , . . . , xs )E is unmixed (that is, every prime ideal belonging to (0) in
E/(x1 , . . . , xs )E is of height s) and x1 , . . . , xs is an E-sequence if s ≥ 1.
If E is an R-module having the properties that dimE = dimR and for every set
of elements x1 , . . . , xs in m such that height(E/(x1 , . . . , xs )E) = s, the submodule
(0) in E/(x1 , . . . , xs )E is unmixed, then E is a C–M R-module.
Theorem II.7.7 Let E be an R-module (dimR > 0). Then the following
statements are equivalent:
(a) E is a C–M module.
(b) If x1 , . . . , xd is a system of parameters for R, then Hi (E(x1 , . . . , xd )) = 0
for all i > 0.
(c) If x1 , . . . , xd is a system of parameters for R, then x1 , . . . , xd is an E-
sequence.
(d) There exists a system of parameters for R which is an E-sequence.
Finally, we have the fact that the equality of length and multiplicity
characterizes the property of being C–M ([9], theorem 5.10).
Theorem II.7.8 Let R be a local ring (dimR > 0) with maximal ideal m, let
E be an R-module, and let x1 , . . . , xd be a system of parameters for R. Then the
following statements are equivalent:
(a) E is a C–M module.
(b) The system of parameters x1 , . . . , xd is an E-sequence.
(c) If (y1 , . . . , yd ) = (x1 , . . . , xd ), then codimE/(y1 , . . . , yd−1 )E > 0.
(d) If q = (x1 , . . . , xd ), then eE (q) = length(E/qE).
64 Local ring theory

Moreover, if E is a C–M module, then any system of parameters for R satisfies


conditions (b), (c), (d).

II.8 Intersection multiplicity and the homological conjectures


Certainly one of the most interesting threads intertwining homology and com-
mutative algebra is that of the so-called homological conjectures which, to a
large degree, emerged from the definition, proposed by J.-P. Serre, of intersection
multiplicity. Here is the definition [78].

Definition II.8.1 Let R be a regular local ring of dimension d, and p, q two


prime ideals of R such that R/p ⊗R R/q has finite length. Then Serre defines the
intersection multiplicity, χ (R/p, R/q), by


d
χ(R/p, R/q) = (−1)i length(TorR
i (R/p, R/q));
i=0

more generally, for finitely generated R-modules, M and N, such that M ⊗R N


have finite length, he defines


d
χ(M, N ) = (−1)i length(TorR
i (M, N )).
i=0

Without going into detail, suffice it to say that his choice of definition was
informed by the desire to be able to prove a “Bézout Theorem.”
The classical conjectures that immediately emerged are:
1. dim(M ) + dim(N ) ≤ dim(R).
2. (Non-negativity) χ(M, N ) ≥ 0.
3. χ(M, N ) > 0 if and only if dim(M ) + dim(N ) = dim(R).
Equivalently, one can write:
1. dim(M ) + dim(N ) ≤ dim(R).
2. (Vanishing) If dim(M ) + dim(N ) < dim(R), χ(M, N ) = 0.
3. (Positivity) If dim(M ) + dim(N ) = dim(R), χ(M, N ) > 0.
In 1985, P. Roberts and H. Gillet-C. Soulé, using K-theory methods, local
Chern classes [74, 75] and Adams operations on Grothendieck groups of com-
plexes [44], succeeded in proving Vanishing. (J.-P. Serre had already proven
that dim(M )+dim(N ) ≤ dim(R).) In 1997, O. Gabber proved it is always the
case that χ(M, N ) ≥ 0. (The best reference for this is the expository paper by
M. Hochster [52], and may be obtained on his web page. His references are to ref-
erences [11] and [36].) Therefore, what remains is to prove the Positivity. Almost
every method of attack on this problem is tied to the following fundamental fact.
The homological conjectures 65

Theorem II.8.2 Let x1 , . . . , xs be a sequence of elements of R, and I the ideal


they generate. Suppose that M/IM has finite length. Write K for the Koszul
complex associated to the sequence x1 , . . . , xs , and define

s
χ(K ⊗ M ) = (−1)i length (Hi (K ⊗ M )) .
i=0

Then

χ(K ⊗ M ) = eM (I),

where eM (I) is the Samuel multiplicity, that is, a slight generalization of the
multiplicity defined in the last section (Section II.7), and discussed in detail in
References [9] and [78].

Usually the proof requires the use of a spectral sequence argument, or else the
type of argument we used in the last section. However, a particular case that is
very suggestive is the following:
Consider, as usual, R, p and q, with length(R/p ⊗ R/q) < ∞. Suppose that q
is generated by a regular sequence x1 , . . . , xs . Then

dim(R/q) = dim(R) − s,

and therefore

dim(R/p) ≤ dim(R) − dim(R/q) = s

and

dim(R/p) + dim(R/q) = dim(R)

if and only if dim(R/p) = s. Since the sequence x1 , . . . , xs is regular, the Koszul


complex K is a free resolution of R/q. Therefore TorR
i (R/q, R/p) is the homology
Hi (K ⊗ R/p). From the above theorem, we see

χ(R/p, R/q) = eR/p (q).

Since the Samuel multiplicity is always non-negative, and positive if and only if
dim(R/p) = s, the conjectures are true in this special case.
In the equicharacteristic case, that is, in the case that both the local ring, R,
and its residue field, k, have the same characteristic, J.-P. Serre proved the mul-
tiplicity conjecture by means of a “reduction to the diagonal” method, one that
pretty much follows the classical discussions of intersection multiplicity in algeb-
raic geometry. If R is a formal power series ring over a field, K[[X1 , . . . , Xd ]], and
M and N R-modules such that M ⊗R N is of finite length, he introduces a new set
of indeterminates, Y1 , . . . , Yd , and considers N as a module over K[[Y1 , . . . , Yd ]].
He then defines a “complete” tensor product, M ⊗  K N over K, as a module over
66 Local ring theory

the ring K[[X1 , . . . , Xd , Y1 , . . . , Yd ]] and proves that


∼ K[[X,Y ]]  
TorRi (M, N ) = Tori M⊗  K N, K[[X, Y ]]/(X1 − Y1 , . . . , Xd − Yd ) .

Since X1 − Y1 , . . . , Xd − Yd form a regular sequence, we obtain in this way


the result for formal power series rings, and the conjectures for the general
equicharacteristic case may be reduced to the special case by means of com-
pletion and the Cohen structure theorems of complete local rings. The proof
of non-negativity due to Gabber uses a theorem of A. de Jong on the exist-
ence of “regular alterations” to reduce the conjectures to questions on regular
immersions into projective space on R.
Clearly all of these proofs are beyond the scope of this book. However, the
Serre conjectures have led to a large number of related conjectures: the so-called
homological conjectures. They have also led to an industry within commutative
algebra with branches that spread almost everywhere. For the reader who is
curious to see what these conjectures look like, we produce a list of them; for the
interested reader, we suggest a number of more detailed references: [42, 49, 52,
53, 74–76, 78].
A few paragraphs earlier, we introduced the term “equicharacteristic.”
Another term, “mixed characteristic,” comes up often in the context of the homo-
logical conjectures. This simply means that the local ring, R, has characteristic
zero, while its residue field, k, is of characteristic p = 0.
The Conjectures. Throughout, R is a local ring with maximal ideal, m, and
all modules are finitely generated unless otherwise indicated.
1. Zero divisor theorem. If M = 0 has finite homological dimension, and r ∈
R is a non-zero divisor for M, then r is a non-zero divisor for R. Current
status: true in all cases.
2. Bass’ question. If M = 0 has a finite injective resolution, then R is a
C–M ring. Current status: true in all cases.
3. Intersection theorem. If M ⊗R N = 0 is of finite length, then the Krull
dimension of N (i.e., dim(R/ann(N ))) ≤ hdR (M ). Current status: true
in all cases.
4. New intersection theorem. Let G : 0 → Gn → · · · → G0 → 0 be a finite
complex of free modules such that ⊕i Hi (G) is of finite length, but not zero.
Then the Krull dimension of R is less than or equal to n. Current status:
true in all cases.
5. New improved intersection theorem. Let G : 0 → Gn → · · · → G0 → 0
be a finite complex of free modules such that Hi (G) is of finite length for
i > 0 and H0 (G) possesses a minimal generator annihilated by a power of
the maximal ideal. Then dimR ≤ n. Current status: true for rings that
contain a field; open in mixed characteristic.
6. Direct summand conjecture. If R ⊆ S is a ring extension with R regular
(not necessarily local) and S is finitely generated as an R-module, then R
The homological conjectures 67

is a direct summand of S as an R-module. Current status: true for rings


that contain a field; open in mixed characteristic.
7. Canonical element conjecture. Let x1 . . . , xd be a system of parameters
for R, let G be a projective resolution of the residue field of R with G0 = R,
and let us denote by K the Koszul complex of R with respect to x1 . . . , xd .
Lift the identity R = K0 → G0 = R to a map of complexes. Then, whatever
the choice of system of parameters or lifting, the last map R = Kd → Gd is
not 0. Current status: true for rings that contain a field; open in mixed
characteristic.
8. Conjecture on existence of big balanced C–M modules. There is
an R-module W (not necessarily finitely generated) such that mW = W
and every system of parameters for R is a regular sequence on W. Current
status: true for rings that contain a field; open in mixed characteristic.
9. Conjecture on a direct summand being C–M. If R is a direct sum-
mand of a regular local ring, S, as an R-module, then R is C–M (R not
necessarily local). Current status: true for rings that contain a field; open
in mixed characteristic.
10. Conjecture on the vanishing of Tor maps. Let A ⊆ R → S be
morphisms, where R is not necessarily local, with A, S regular and R a
finitely generated A-module. Let W be an arbitrary A-module. Then the
map TorA i (W, R) → Tori (W, S) is zero for every i ≥ 1. Current status:
A

true for rings that contain a field; open in mixed characteristic.


11. The strong direct summand conjecture. Let A ⊆ R be a map of
complete local domains, and let q be a prime ideal of R of height 1 that lies
over xA, where A and A/xA are regular. Then xA is a direct summand of q
as an A-module. Current status: true for rings that contain a field; open
in mixed characteristic.
12. Conjecture on the existence of weakly functorial C–M algebras.
Let R → S be a local morphism of complete local domains. Then there
exist an R-algebra BR and an S-algebra BS such that BR is a big balanced
C–M algebra for R, BS equally for S, and a map BR → BS such that

BR → BS
↑ ↑
R → S

is commutative. Current status: true for rings that contain a field; open
in mixed characteristic.
13. Serre multiplicity conjectures. Suppose R regular of dimension d, and
M ⊗R N of finite length. Then we have

χ(M, N ) = (−1)i length(TorR
i (M, N )) = 0
68 Local ring theory

if dimM +dimN < d, and positive if the sum equals d. Current status:
true for rings that contain a field; open in mixed characteristic (question of
positivity open: non-negativity is always true).
14. The small C–M modules conjecture. If R is complete, then there exists
a finitely generated R-module, M = 0, such that some (or, equivalently,
every) system of parameters for R is a regular sequence for M . Current
status: true in dimensions ≤ 2; other cases, nothing known, not even if the
ring contains a field, and no matter what the characteristic.
The following is a diagram of implications:
13

14 1
 
8 → 7, 6, 5 → 4 →3
  
12 → 11, 10 2

9
III
GENERALIZED KOSZUL COMPLEXES

In the preceding chapter, we saw the fundamental role that the Koszul complex
played in the study of certain aspects of local ring theory. One of the main ideas
tied up with that complex is that of regular sequence, with the Cohen–Macaulay
Theorem one of the major results obtained through its use. There is another
theorem, known as the Generalized Cohen–Macaulay Theorem [35], which deals
with the heights of ideals generated by the minors of an m × n matrix. (More
precisely, the classical form of this theorem states that if R is a regular local ring,
then the ideal generated by the p × p minors of an m × n matrix with entries in
R, has height at most equal to (m − p + 1)(n − p + 1).) We can regard the “old”
Cohen–Macaulay Theorem, then, as having to do with the 1 × 1 minors of an
m × 1 matrix, and the Koszul complex as a complex that “approximates” the
resolution of a cyclic module (that is, a module generated by a single element).
With this point of view, it is natural to ask if one can associate a complex to an
arbitrary m×n matrix (m ≥ n) which generalizes the Koszul complex, and which
may be used to prove the generalized theorem mentioned above. This chapter
will focus on this question; in fact, we will develop a whole class of “generalized
Koszul complexes” that will also play a role in Chapter IV.
In our last section of this chapter we take a look at the connection between
leading coefficients of certain Hilbert–Samuel polynomials and the Euler–
Poincaré characteristics of our generalized Koszul complexes. It was natural
at the time that the papers [18, 27] were being written, to try to generalize
the notions of Hilbert–Samuel polynomials and multiplicity to the situation of
finitely generated modules rather than just cyclic modules, that is, to mod-
ules of the form coker(f : Rm → Rn ), m ≥ n, rather than those of the form
coker(f : Rm → R). In the last sections of Reference [27], these generalizations
were introduced, but the new multiplicity was not pursued further. However, in
the mid-1990s, a flurry of work on these notions began (see [43, 58–61]), which
includes more geometric approaches to the subject. (In the more recent literat-
ure, this multiplicity is often referred to as “Buchsbaum-Rim multiplicity.”) As
in Chapter II, we have to say that a detailed study of this topic would lead us
much too far beyond the scope of this book, so we merely indicate what the
original definition was and its relation to the generalized Koszul complex.

III.1 A few standard complexes


Before we start to develop generalized complexes, it is worthwhile to look a little
more closely at the Koszul complex and some of its immediate relatives. This
70 Generalized Koszul complexes

section will devote itself almost entirely to these considerations. The free modules
we deal with are always assumed to be nontrivial and finitely generated.

III.1.1 The graded Koszul complex and its “derivatives”


If F is a free R-module, with basis x1 , . . . , xn , we know from Chapter I that the
symmetric algebra, S(F ), is isomorphic to the polynomial ring R[x1 , . . . , xn ]. In
that ring, x1 , . . . , xn is a regular sequence, so that the Koszul complex associ-
ated to the ideal, J, generated by these elements is acyclic. (Lest there be any
confusion about the meaning of the expression acyclic complex, we will take
it to mean that the positive-dimensional homology groups of the complex are
all zero.) If we denote the ring S(F ) by S, and the free S-module S ⊗R F
∼ ∼ ∼
by F , then the terms of this Koszul complex are 0 → Λn F → Λn−1 F →

· · · → Λ1 F → S → 0. But all of these modules are graded S-modules; in fact,
∼ S ⊗R Λk F = 

Λk F = S ⊗ Λk F, and its grading is given by that of S
  p≥0 p

itself, namely, Λk F = Sp ⊗ Λk F. The boundary map of this complex, which
 p
maps p Sp ⊗ Λk F to p Sp ⊗ Λk−1 F, is easily seen to be the sum of the maps
Sp ⊗ Λk F → Sp+1 ⊗ Λk−1 F where each of these maps is the composition
1⊗∆ m⊗1
Sp (F ) ⊗ Λk F −→ Sp (F ) ⊗ F ⊗ Λk−1 F −→ Sp+1 (F ) ⊗ Λk−1 F,

∆ is the diagonal map from Λk F to F ⊗ Λk−1 F, and m is the multiplication map


from Sp (F ) ⊗ F to Sp+1 (F ). In fact, we see that the Koszul complex is the direct
sum of complexes of R-modules:

0 → Λq F → S1 ⊗ Λq−1 F → · · · → Sq−l ⊗ Λl F → · · · → Sq−1 ⊗ Λ1 F → Sq → 0

which are exact for all choices of q > 0. When q = 0, the complex is nothing
other than

0→R→0

and its homology in dimension zero is R itself. Since R = S/J, this is precisely the
statement that the original Koszul complex is acyclic, and its zero-dimensional
homology is S/J.

Definition III.1.1 For each free R-module, and each integer q ≥ 0, we define
Λq (F ) to be the complex

0 → Λq F → S1 ⊗ Λq−1 F → · · · → Sq−l ⊗ Λl F → · · · → Sq−1 ⊗ Λ1 F → Sq → 0.

• An important observation to make here (it will be used later in this chapter)
is that for q ≥ n = rankF , the length—or dimension—of this complex is equal
to n, and the last term on the left is Sq−n F ⊗ Λn F.
A few standard complexes 71

Definition III.1.2 If ϕ : G → F is a map of free R-modules, and q ≥ 0, we


define Λq (ϕ) to be the complex

0 → Λq G → S1 (F ) ⊗ Λq−1 G → · · · → Sq−l (F ) ⊗ Λl G
→ · · · → Sq−1 (F ) ⊗ Λ1 G → Sq (F ) → 0.

The complex Λq (F ) is the same as Λq (id), where id is the identity map on F.

The boundary maps in the complexes Λq (ϕ) are the evident analogs of those
in Λq (F ) :
1⊗∆ 1⊗ϕ⊗1
Sp (F ) ⊗ Λk G → Sp (F ) ⊗ G ⊗ Λk−1 G →
m⊗1
Sp (F ) ⊗ F ⊗ Λk−1 G → Sp+1 (F ) ⊗ Λk−1 G,

Of course, while the complexes Λq (F ) are exact for all q > 0, that is not
necessarily the case for arbitrary Λq (ϕ).
As we saw in Chapter I, the graded dual of the symmetric algebra is the
divided power algebra. Thus, if we were to take the linear dual of the complex
Λq (F ) above, we would obtain, for each q > 0, the exact sequence (exact because
Λq (F ) is split exact) Dq (F ) :

0 → Dq → Dq−1 ⊗ Λ1 F → · · · → Dq−l ⊗ Λl F → · · · → D1 ⊗ Λq−1 F → Λq F → 0.

For q = 0, we again just obtain the trivial complex

0 → R → 0.

This observation leads us to make the corresponding definition:

Definition III.1.3 If ϕ : G → F is a map of free R-modules, and q ≥ 0, we


define Dq (ϕ) to be the complex

0 → Dq (G) → Dq−1 (G) ⊗ Λ1 F → · · · → Dq−l (G) ⊗ Λl F


→ · · · → D1 (G) ⊗ Λq−1 F → Λq F → 0.

The complex Dq (F ) is the same as Dq (id), where id is the identity map on F .

Again, in the case of general ϕ we can make no assertion about exact-


ness or acyclicity. To be sure that there is no confusion about the boundary
maps in the complex Dq (ϕ), we point out that each boundary map, ∂pq , is the
composition
∆⊗1 1⊗ϕ⊗1
Dp (G) ⊗ Λk F → Dp−1 (G) ⊗ G ⊗ Λk F →
1⊗m
Dp−1 (G) ⊗ F ⊗ Λk F → Dp−1 (G) ⊗ Λk+1 F.
72 Generalized Koszul complexes

III.1.2 Definitions of the hooks and their explicit bases


Given the complexes Λq (F ) and Dq (F ) defined above, it is reasonable to ask
whether we can describe their cycles in some convenient way. That is, if we are
given a basis x1 , . . . , xn of F, can we in some succinct way describe a basis for
each of the modules of cycles of these complexes? For example, we know that
a basis of Dp (F ) may  be described as the set of all divided power monomials:
(α1 ) (α2 ) (αt ) 
xi1 xi2 · · · xit where 1 ≤ i1 < i2 < · · · < it ≤ n, αj = p, and αj > 0
for j = 1, . . . , t. (This is ofcourse equivalent to the description as the set of all
(α ) (α ) (α ) 
divided power monomials x1 1 x2 2 · · · xn n with αj = p and αj ≥ 0 for
j = 1, . . . , n, but it will be seen soon why we are using the first description.)
We also know that a basis for Λk F may be described as the set of all elements
{xj1 ∧ xj2 ∧ · · · ∧ xjk } where 1 ≤j1 < · · · < jk ≤ n. Thus we often describe the
(α ) (α ) (α )
basis of Dp (F )⊗ Λk F as the set xi1 1 xi2 2 · · · xit t ⊗ xj1 ∧ xj2 ∧ · · · ∧ xjk . A
convention among combinatorialists is to write these elements out in tableau
form in the following way:
 α1 α2 αt 
# $% & # $% & # $% &
 xi1 xi1 · · · xi1 xi2 xi2 · · · xi2 · · · xit xit · · · xit 
 x 
 1 j 
 x .
 j2 
 . 
 .. 
xjk
That is, the repeats in the row, when they occur, are to be read as the appropriate
divided power of the repeated element, rather than as simply the product of the
elements (hence our choice of the word “combinatorialists”) and the product of
these divided powers is the element of the divided power algebra that the total
row “represents”. Notice that when the tableau is written as above, it has the
property that the indices of the basis elements are weakly increasing in the rows
(in this case, just the top row is involved), while the column indices are strictly
increasing. So we see that we may describe the basis of Dp (F )⊗ Λk F as the set
of all tableaux of the form
 
xu1 xu2 · · · xup
 xj1 
 
 xj2 
T(u,j) =  
 .. 
 . 
xjk
with 1 ≤ u1 ≤ u2 ≤ · · · ≤ up ≤ n, 1 ≤ j1 < j2 < · · · < jk ≤ n, where
u = {u1 , u2 , . . . , up } and j = {j1 , j2 , . . . , jk }. The rest of this subsection will be
devoted to obtaining a description of a basis of the cycles of our complexes in a
tableau form that is equally suggestive.
The reader may ask how we can assume that there is a basis for the cycles;
after all, we have not directly established that the cycles are free modules. That
A few standard complexes 73

they are projective is clear, since the complexes Λq (F ) and Dq (F ) are split exact;
hence the cycles are direct summands of free modules and therefore projective.
To see that they are indeed free, let us observe that if R is any commutative
ring, and F a free R-module, then F = R ⊗Z F0 where F0 is a free Z-module (of
the same rank as F ). Since Λk F = R ⊗Z Λk F0 , Dp (F ) = R ⊗Z Dp (F0 ), Sp (F ) =
R ⊗Z Sp (F0 ), and since the complexes (over any ring) are split exact, we see
first of all that the cycles of Λq (F0 ) and Dq (F0 ) are free (because projective
abelian groups, being summands of free abelian groups, are torsion-free, hence
free by Proposition I.2.28), and then that the cycles of Λq (F ) and Dq (F ) are free,
because they are just the tensor product with R of their integral counterparts.
This property of a functor, that is, that is R-free and can be obtained by tensoring
its integral counterpart by the ring R, is called universal freeness. With the
freeness of these cycles established, how do we go about finding a “convenient”
basis for them?
To this end, we will use a standard procedure in homology: we will construct
a splitting homotopy for the complex Dq (F ), q > 0; this will pick out the basis
for the cycles that we are looking for. One proceeds almost identically for the
complexes Λq (F ); we will omit the discussion of the homotopy in that case, but
simply give a description of the resulting basis.
To construct the desired homotopy, we have to define maps
sqp : Dp (F ) ⊗R Λq−p F → Dp+1 (F ) ⊗R Λq−p−1 F for − 1 ≤ p ≤ q
such that
q
(i) ∂p+1 sqp + sqp−1 ∂pq = 1 for 0 ≤ p ≤ q;
(ii) sqp+1 sqp = 0 for 0 ≤ p ≤ q − 1.
Of course, the maps sq−1
and sqq
are the zero maps.
Since all of the modules we are considering are free, it suffices to define these
maps on the basis elements. For p > 0, we will use the tableau description of
basis elements that we discussed above, and define
 
xu1 xu2 · · · xup
 xj1 
 
 
sqp  xj2 
 .. 
 . 
xjq−p


 0 if u1 < j1 ;



  



 xj1 xu1 ··· xup
 xj2 
=  

  xj3 

   if u1 ≥ j1 .

  .. 

  . 


xjq−p
74 Generalized Koszul complexes

For p = 0, we define
 
xj1
 xj2 
 
sq0 (xj1 ∧ xj2 ∧ · · · ∧ xjq ) =  .. .
 . 
xjq
Now we have to check the conditions (i) and (ii) above. First of all, it is clear
that
 
xj1
 xj2 
 
∂1q sq0 (xj1 ∧ xj2 ∧ · · · ∧ xjq ) = ∂1q  .  = xj1 ∧ xj2 ∧ · · · ∧ xjq ,
 .. 
xjq
so (i) is verified for p = 0. For p > 0, we have
 
xu1 xu2 · · · xup
 xj1 
 
q q  xj2 
∂p+1 sp  
 .. 
 . 
xjq−p


 0 if u1 < j1



  



 xj1 xu1 · · · xup
 xj2 
=  

 q  xj3 

 ∂ p+1   if u1 ≥ j1 ,

  .. 

  . 


xjq−p
while
 
xu1 xu2 ··· xup
 xj1  m 
  
 xj2 
sqp−1 ∂pq 
q
 = sp−1 Th ,
 .. 
 .  h=1

xjq−p
where h runs, in order, through the m distinct elements of {u1, . . . , up } and
 
xu1 xu2 · · · x h · · · xup
 xh 
 
 xj1 
Th =  ,
 .. 
 . 
xjq−p
h indicates that the element xh is to be omitted.
where x
A few standard complexes 75

Now let us consider the two cases: u1 < j1 and u1 ≥ j1 . In the first case, we
must show that
 
xu1 xu2 · · · xup
m   xj1 
  
 
sqp−1 Th =  xj2 .
 .. 
h=1  . 
xjq−p
Since u1 ≤ · · · ≤ up , we see that except when h = u1 , the tableau Th always has
the property that u1 is less than all the subscripts in the first column, so that
sqp−1 (Th ) = 0. Thus, we see that
 
xu1 xu2 · · · xup
 xj1 
 
q q  xj2 
sp−1 ∂p  
 .. 
 . 
xjq−p
 
u1
x xu2 ··· xup
 xu1 
 
 xj1 
= sqp−1  .
 .. 
 . 
xjq−p
 
xu1 xu2 ··· xup
 xj1 
 
 xj2 
Since this latter term is equal to  , we are done.
 .. 
 . 
xjq−p
In the second case, we must calculate the sum
 
xj1 xu1 · · · xup
 xj2 
 
q  xj3 
∂p+1  
 .. 
 . 
xjq−p
 
xu1 xu2 ··· h
x ··· xup
 xh 

m  
 xj1 
+ sqp−1  .
 .. 
h=1  . 
xjq−p
76 Generalized Koszul complexes

Since we are now assuming that u1 ≥ j1 , we see that in each tableau Th for
which h = u1 , h > j1 so that

 
xu1 xu2 ··· h
x ··· xup
 xh 
 
 xj1 
sqp−1  
 .. 
 . 
xjq−p
 
xj1 xu1 ··· h
x ··· xup
 xh 
 
 xj2 
= − 
 .. 
 . 
xjq−p

for h = u1 , while for h = u1 we have

 
u1
x xu2 ··· xup
 xu1 
 
 xj1 
sqp−1  =0
 .. 
 . 
xjq−p

if u1 = j1 , while for u1 > j1 we have

 
u1
x xu2 ··· xup
 xu1 
 
 xj1 
sqp−1  
 .. 
 . 
xjq−p
 
xj1 u1
x xu2 ··· xup
 xu1 
 
 xj2 
= − .
 .. 
 . 
xjq−p
A few standard complexes 77

 
xj1 xu1 ··· xup
 xj2 
 
q  xj3 
As for the other summand, ∂p+1  , we see that
 .. 
 . 
xjq−p
it is equal to
 
xj1 xu1 ··· h
x ··· xup
 xh 
m   xj2


h=1   if u1 = j1 ;
 .. 
 . 
xjq−p
 
xu1 xu2 ··· xup
 xj1 
 
 xj2 
 
 .. 
 . 
xjq−p
 
xj1 xu1 ··· h
x ··· xup
 xh 
m  
 xj2 
+   if u1 > j1 .
 .. 
h=1  . 
xjq−p

Comparing terms (including signs), we see that we obtain our desired result. The
verification of condition (ii) is straightforward; in fact it is trivial.
With our homotopy in hand, how do we describe the bases we are looking for?
We claim that if we have a basis of Dp (F )⊗Λq−p F, and consider the subset {Tβ }
q
consisting of those elements for which sqp (Tβ ) = 0, then {∂p+1 sqp (Tβ )} generate
the p-cycles of Dq (F ). For if z ∈ Dp (F ) ⊗ Λ F is a p-cycle, write
q−p

 
z= cα Tα + d β Tβ
α β

where the Tα are the basis elements that map to zero, while the Tβ are those
already described. Since
q
z = ∂p+1 sqp (z) + sqp−1 ∂pq (z),
and z is a cycle, we have
 
 
q
z = ∂p+1 q
sqp (z) = ∂p+1 sqp  dβ Tβ  = q
dβ ∂p+1 sqp (Tβ ).
β β
78 Generalized Koszul complexes

q
Therefore, if the elements {∂p+1 sqp (Tβ )} are linearly independent, they form a
basis for the p-cycles.
That these elements are linearly independent is easy to see in the following
way. If we have a linear combination
 q
dβ ∂p+1 sqp (Tβ ) = 0,
β
 q
this means that the term z = β dβ sqp (Tβ ) ∈ ker(∂p+1 ). That is, z is a (p + 1)-
cycle, and as such we know z = ∂p+2 sp+1 (z). But since sqp+1 sqp = 0, this tells
q q

us that z = 0. However, we know that each tableau sqp (Tβ ) is a basis element
of Dp+1 (F ) ⊗ Λq−p−1 F, so that if z = 0, we must have that each dβ = 0. This
q
proves the linear independence we are after. (It also shows us that the map ∂p+1
restricted to the image of sp is one-to-one.) Now the set of tableaux {sp (Tβ )} is
q q

precisely the subset of the basis tableaux of Dp+1 (F ) ⊗ Λq−p−1 F of the form
 
xu1 xu2 · · · xup+1
 xj1 
 
 xj2 
T(u,j) =  
 .. 
 . 
xjq−p−1
with u1 ≤ u2 ≤ · · · ≤ up+1 and u1 < j1 < j2 < · · · < jq−p−1 . Since it is clear
that every element sqp (Tβ ) is of the form just described, it suffices to show that
every tableau T(u,j) of that form is sqp (Tβ ) for some tableau Tβ . But we see that
 
xu2 xu3 · · · xup+1
 xu1 
 
 
T(u,j) = sqp  xj1  , and so we are done.
 .. 
 . 
xjq−p−1
q
Now, though, we are after a good way to describe the elements ∂p+1 (T(u,j) )
 
xu1 xu2 · · · xup+1
 xj1 
 
 xj2 
for such tableaux T(u,j) =   . The convention
 .. 
 . 
xjq−p−1
generally adopted
 is to again use tableau notation,
 and to denote this boundary
xu1 xu2 · · · xup+1
 xj1 
 
 xj2 
by the tableau   , where we see that the rows are
 .. 
 . 
xjq−p−1
weakly increasing, and the column is strictly increasing. Such a tableau has
A few standard complexes 79

a special name; it is called a standard tableau. Notice, then, that this tableau
stands for the sum of a number of terms in Dp (F ) ⊗ Λq−p F obtained essentially
by diagonalizing the terms in the first row, and multiplying them with the term
in ΛF represented by the product of the terms in the first
 column starting with
x1 x1 x2 x2
xj1 . Thus, for example, the tableau  x3  represents the element
   x4 
x1 x2 x2 x1 x1 x2
 x1   x2 
 +  in D3 (F ) ⊗ Λ3 F.
 x3   x3 
x4 x4
The corresponding discussion for the complexes Λq (F ) considers basis ele-
ments of Sq−l ⊗ Λl F which are tableaux of the form
 
xu1 xu2 · · · xul
 xj1 
 
 xj2 
 ,
 .. 
 . 
xjq−l
where now we have u1 < u2 < · · · < ul and j1 ≤ j2 ≤ · · · ≤ jq−l , and no relation
assumed between u1 and j1 . This means that we are regarding the elements of
the first row as xu1 ∧ xu2 ∧ · · · ∧ xul ∈ Λl F, and xj1 · · · xjq−l as an element of
Sq−l (F ). One then shows that the l-cycles of Λq (F ) can be parametrized by the
set of tableaux
 
xu1 xu2 · · · xul+1
 xj1 
 
 xj2 
 
 .. 
 . 
xjq−l−1
with u1 < u2 < · · · < ul+1 , j1 ≤ j2 ≤ · · · ≤ jq−l−1 and u1 ≤ j1 . Again, to read
off what this element is, we diagonalize the element represented by the first row,
and multiply the first entry with the element of S(F ) represented by the product
of the terms in the first column starting with xj1 . (This procedure is a bit more
straightforward than for the first case as we have here no repeats in the top row.)
Definition III.1.4 The module of p-cycles of the complex Dq F is denoted by
K(p+1,1q−p−1 ) (F ); the module of p-cycles of the complex Λq (F ) is denoted by
L(p+1,1q−p−1 ) (F ).
Remark III.1.5 It should be noted immediately, and it will be used later in
this chapter, that for any free module, F, we have
K(ν+1,1q−1 ) F ∗ ∼
= (L(q,1ν ) F )∗ .
80 Generalized Koszul complexes

Notice that when p = 0, the basis of K(1,1q−1 ) (F ) is the set of tableaux




 x u1 


  
 xj1 
 
 xj2 
  with u1 < j1 < · · · < jk−1 . But this, as we would expect, is

 

.

  .. 


 

xjq−1
exactly what parametrizes the basis of Λq (F ). A similar observation applies to
the 0-cycles of Λq (F ).
The modules K(p+1,1q−p−1 ) (F ) are called hook Weyl modules; the mod-
ules L(p+1,1q−p−1 ) (F ) are called hook Schur modules. Later, when we make a
more systematic study of the representations of the general linear group (Chapter
VI), we will see that these modules are special instances of more general repres-
entations. At that later time, it will become clear why we use these particular
subscripts to designate these modules.

III.2 General setup


In order to construct Koszul-type complexes in a much more general form, we
first take a look at the familiar bar complex, which is even more classical
than the Koszul complex. It arises in the following way: If Λ is an algebra over
the commutative ring, R, and M a Λ-module, then one can form the following
complex:
··· → Λ ⊗ ··· ⊗ Λ ⊗ M → Λ ⊗ ··· ⊗ Λ ⊗ M → ··· → Λ ⊗ M → M
% &# $ % &# $
l l−1

where the map Λ ⊗ M → M is simply the action of Λ on M and, in general, the


map Λ ⊗ · · · ⊗ Λ ⊗ M → Λ ⊗ · · · ⊗ Λ ⊗ M is defined by
% &# $ % &# $
l l−1


l−1
λ1 ⊗ · · · ⊗ λl ⊗ m → (−1)l−j λ1 ⊗ · · · ⊗ λj λj+1 ⊗ · · · ⊗ λl ⊗ m
j=1

+ λ1 ⊗ · · · ⊗ λl−1 ⊗ λl m.
In the case of a graded algebra, Λ, one usually restricts the degrees of the λj
to be positive and, in that case, the complex is referred to as the normalized
bar complex (it is usually assumed that all the λj are homogeneous and that
the module, M, is graded also). We will always make this convention when there
is a grading. Observe, too, that in the graded case, we get a subcomplex by
stipulating that the degree of the λ1 should always be greater than or equal to
some fixed integer, s. Another useful observation is that we may always take as
our module, M, the algebra Λ itself. (It will always be assumed, unless specified
otherwise, that the tensor product is taken over the ground ring, R.)
If B is an arbitrary module over R, we may consider its dual, B ∗=HomR (B,R),
and the two exterior algebras: ΛB and ΛB ∗ . We have seen, in our original
General setup 81

discussion of the Koszul complex, that every element, β, of B ∗ induces an endo-


morphism, dβ (of degree −1) on ΛB. Thus we have a morphism, β → dβ of the
module B ∗ into the endomorphism ring of ΛB, or EndR (ΛB), with (dβ )2 = 0. By
the universal mapping property of the exterior algebra, this morphism extends
to a unique algebra map of ΛB ∗ into EndR (ΛB), and this establishes ΛB as a
ΛB ∗ -module.
Now suppose that we have a map f : A → B. This induces a map f ∗ : B ∗ → A∗
as well as an algebra map Λf ∗ : ΛB ∗ → ΛA∗ . Since, by the above remarks, ΛA
is a ΛA∗ -module, we immediately see that ΛA becomes a ΛB ∗ -module via the
map Λf ∗ . Consequently, we can form the bar complex, with ΛB ∗ as our ring Λ,
and with ΛA as our graded Λ-module:
· · · → ΛB ∗ ⊗ · · · ⊗ ΛB ∗ ⊗ ΛA → ΛB ∗ ⊗ · · · ⊗ ΛB ∗ ⊗ ΛA →
% &# $ % &# $
l l−1

· · · → ΛB ⊗ ΛB ⊗ ΛA → ΛB ∗ ⊗ ΛA → ΛA.
∗ ∗

We remarked above that if our ring Λ is graded, and our module M is a graded
Λ-module, then we can consider graded strands of the bar complex.
In particular, suppose that we are given two integers p ≥ 0 and s > 0. Then
instead of the complex we have just written, we could consider

 
··· → Λk1 B ∗ ⊗ · · · ⊗ Λkl B ∗ ⊗ Λp+k1 +···+kl A →
% &# $
k1 ≥s ki >0 l
 
··· → Λ B ∗ ⊗ Λk2 B ∗ ⊗ Λp+k1 +k2 A →
k1

k1 ≥s k2 >0

Λk B ∗ ⊗ Λp+k A → Λp A.
k≥s

Notice that if we choose s = 1, and sum for all p, we get the full (graded)
complex.
We can also do the same thing using ΛB as our ΛB ∗ -module and, given our
map Λf : ΛA → ΛB induced from f : A → B, it is not hard to see that we get
a commutative diagram
 
··· → Λk1 B ∗ ⊗ Λk2 B ∗ ⊗ Λp+|k| A → Λk B ∗ ⊗ Λp+k A → Λp A
k1 ≥s,k2 >0 k≥s

↓ ↓ ↓
 
∗ ∗ ∗
··· → Λ B ⊗Λ B ⊗Λ
k1 k2 p+|k|
B→ Λ B ⊗Λ
k p+k
B→Λ B
p

k1 ≥s,k2 >0 k≥s

where |k| denotes the sum of the indices ki and the vertical map is the map Λf
tensored with the identity on the appropriate sum of tensor products of copies
of ΛB ∗ .
82 Generalized Koszul complexes

III.2.1 The fat complexes


As in the case of Chapter II, when we see a map of complexes, we automatically
take the mapping cone to obtain yet another complex. But instead of doing
this in general, let us keep in mind that we are interested in studying the case
when A and B are free modules over a commutative ring, R, of ranks m and n
respectively. In that case, we see that if we choose our integers p and s above so
that p+s > n, then all the terms but Λp B in the bottom row of our commutative
map of complexes will vanish. In fact, we will take p + s = n + 1 to accomplish
this purpose. When one does this, we see that the mapping cone of our map of
complexes is simply the top row augmented by Λp B, that is, we have the complex
 
··· → Λk1 B ∗ ⊗ · · · ⊗ Λkl B ∗ ⊗ Λp+|k| A → · · ·
% &# $
k1 ≥s ki >0 l

→ Λ B ∗ ⊗ Λp+k A → Λp A → Λp B.
k

k≥s

It is worth remarking that the freeness of the modules A and B does not enter
into the above discussion; it is enough to assume that our module B is such that
Λn+1 B = 0 to achieve the  same result. It may also be amusing to think of the
fact that the composition k≥s Λk B ∗ ⊗ Λp+k A → Λp A → Λp B is zero as the
generalization of Cramer’s Rule from linear algebra. Although we will continue
to use the letters A and B, what we should really keep in the back of our minds
is that A = Rm , and B = Rn , with m ≥ n. So, for example, if m = n, our
complex is simply
· · · → 0 → 0 → Λp R n → Λ p R n ,
while for m = n + 1, we have
0 → Λn−p+1 Rn∗ ⊗ Λn+1 Rn+1 → Λp Rn+1 → Λp Rn .
Up to this point our complexes look fairly svelte, but as soon as the difference
m − n becomes 2 or larger, we get a great many terms that seem extraneous. We
will soon see how this problem is addressed in general, but for now, let us just
look at the case m = n + 3 to see how “fat” our complexes can get to be.
0 → Λn−p+1 Rn∗ ⊗ Rn∗ ⊗ Rn∗ ⊗ Λn+3 Rn+3
Λn−p+1 Rn∗ ⊗ Λ1 Rn∗ ⊗ Λn+2 Rn+3 ⊕
→ Λn−p+1 Rn∗ ⊗ Λ2 Rn∗ ⊗ Λn+3 Rn+3 ⊕
Λn−p+2 Rn∗ ⊗ Λ1 Rn∗ ⊗ Λn+3 Rn+3
Λn−p+1 Rn∗ ⊗ Λn+1 Rn+3 ⊕
→ Λn−p+2 Rn∗ ⊗ Λn+2 Rn+3 ⊕ → Λp Rn+3 → Λp Rn .
Λn−p+3 Rn∗ ⊗ Λn+3 Rn+3
Despite this obesity, the complex always has dimension m − n + 1, the number
suggested (at least in the case when p = n) by the generalized Cohen–Macaulay
General setup 83

Theorem. But it would be more efficient if we could reduce the fat while keeping
the correct dimension. To this end, let us take a closer look at what is going on
and what it is we want.

III.2.2 Slimming down


Ideally, what we want is a generically acyclic complex, which means that at
each step of our construction we want to cover the kernel of our preceding map.
So the first question we can ask is: Do the images of Λn−p+2 Rn∗ ⊗ Λn+2 Rn+3
and Λn−p+3 Rn∗ ⊗ Λn+3 Rn+3 in Λp Rn+3 above help in covering elements of the
kernel of the map Λp Rn+3 → Λp Rn that the map Λn−p+1 Rn∗ ⊗ Λn+1 Rn+3 →
Λp Rn+3 missed? And the answer is, No. For in general we have a commutative
diagram:

Λk+l B ∗ ⊗ Λt+l A
 
Λk B ∗ ⊗ Λl B ∗ ⊗ Λt+l A Λt−k A ,
 
Λk B ∗ ⊗ Λ t A

where the map Λk B ∗ ⊗ Λl B ∗ ⊗ Λt+l A → Λk+l B ∗ ⊗ Λt+l A is given by the multi-


plication in ΛB ∗ , and the map Λk B ∗ ⊗ Λl B ∗ ⊗ Λt+l A → Λk B ∗ ⊗ Λt A is given by
the action of ΛB ∗ on ΛA. The other two maps are given by the action of ΛB ∗
on ΛA. Since the first of the maps (the one given by multiplication in ΛB ∗ ) is
a surjection, we see that the image of Λk+l B ∗ ⊗ Λt+l A in Λt−k A is contained in
the image of Λk B ∗ ⊗ Λt A in Λt−k A.
This tells us that, from the point of view of covering the kernel of Λp Rn+3 →
p n
Λ R , we do not need the two extra terms above; we needed them in order
to use the bar complex construction to proceed; that is, we needed them in
order to catch terms coming from, say, Λn−p+1 Rn∗ ⊗ Λ1 Rn∗ ⊗ Λn+2 Rn+3 ,
under the boundary map of the bar complex. For under this map, a term
in Λn−p+1 Rn∗ ⊗ Λ1 Rn∗ ⊗ Λn+2 Rn+3 will have its image in the direct sum of
Λn−p+1 Rn∗ ⊗ Λn+1 Rn+3 and Λn−p+2 Rn∗ ⊗ Λn+2 Rn+3 . But the component of
the boundary that falls into Λn−p+2 Rn∗ ⊗ Λn+2 Rn+3 is there due to the mul-
tiplication in ΛRn∗ . Suppose that, instead of taking an arbitrary element of
Λn−p+1 Rn∗ ⊗ Λ1 Rn∗ ⊗ Λn+2 Rn+3 , we were to take an element of K ⊗ Λn Rn+3 ,
where for the moment we let K = K(2,1n−p ) (Rn∗ ) = ker(Λn−p+1 Rn∗ ⊗ Λ1 Rn∗ →
Λn−p+2 Rn∗ ), and apply our boundary map to it. In that case, it would end up in
Λn−p+1 Rn∗ ⊗ Λn+1 Rn+3 , we would not have to include the extra term in order
to “catch” the result of multiplying in ΛRn∗ , and our slimmed-down complex
would start out looking like this:

K(2,1n−p ) (Rn∗ ) ⊗ Λn+2 Rn+3 → Λn−p+1 Rn∗ ⊗ Λn+1 Rn+3 → Λp Rn+3 → Λp Rn .


84 Generalized Koszul complexes

We have observed that Λn−p+1 B ∗ may be described as K(1,1n−p ) (B ∗ ), so that


the complex above may be rewritten

K(2,1n−p ) (Rn∗ ) ⊗ Λn+2 Rn+3 → K(1,1n−p ) (Rn∗ ) ⊗ Λn+1 Rn+3


→ Λp Rn+3 → Λp Rn .

The natural question to ask is whether this slimming down can be continued
all along our fat complexes. That is, can we do what is suggested above, and
define a sleek complex associated to any map f : A → Rn and any integer
p≤n:
m
0 → K(m−n,1n−p ) (Rn∗ ) ⊗ ΛA → · · · → K(l,1n−p ) (Rn∗ ) ⊗ Λn+l A → · · ·
→ K(2,1n−p ) (Rn∗ ) ⊗ Λn+2 A → K(1,1n−p ) (Rn∗ ) ⊗ Λn+1 A → Λp A → Λp Rn ?

We notice that this “complex,” too, has dimension m−n+1, but it is considerably
slimmer than the ones we have considered so far.
We have placed quotation marks around the word complex because as yet we
have not defined the boundary maps and verified that they give us a complex.
To define the boundary map, we will focus on the definition of

K(l,1n−p ) (Rn∗ ) ⊗ Λn+l A → K(l−1,1n−p ) (Rn∗ ) ⊗ Λn+l−1 A.

We have the commutative diagram


Dl (Rn∗ ) ⊗ Λn−p Rn∗ ⊗ Λn+l A → Dl−1 (Rn∗ ) ⊗ Λn−p Rn∗ ⊗ Λn+l−1 A
n−p+l n−p+l−1
↓ ∂l ⊗1 ↓ ∂l−1 ⊗1
Dl−1 (R ) ⊗ Λ
n∗
R ⊗ Λ A → Dl−2 (R ) ⊗ Λn−p+1 Rn∗ ⊗ Λn+l−1 A
n−p+1 n∗ n+l n∗

where the vertical maps are the boundary maps of the complexes Dn−p+l Rn∗
and Dn−p+l−1 Rn∗ , each tensored with the indicated exterior powers of A. The
horizontal maps involve the diagonalization of D(Rn∗ ), and the action of Rn∗ on
ΛA. That is, the upper horizontal map is the composition

Dl (Rn∗ ) ⊗ Λn−p Rn∗ ⊗ Λn+l A → Dl−1 (Rn∗ ) ⊗ Rn∗ ⊗ Λn−p Rn∗ ⊗ Λn+l A

= Dl−1 (Rn∗ ) ⊗ Λn−p Rn∗ ⊗ Rn∗ ⊗ Λn+l A → Dl−1 (Rn∗ ) ⊗ Λn−p Rn∗ ⊗ Λn+l−1 A,
and the lower one is similarly defined. The commutativity of the diagram is easy
to verify.
Given that this diagram is commutative, it follows that the image of the
left-hand vertical map is carried into the image of the right-hand vertical
map. But these images are precisely the modules K(l,1n−p ) (Rn∗ ) ⊗ Λn+l A and
K(l−1,1n−p ) (Rn∗ ) ⊗ Λn+l−1 A that we are after. To check that the composition of
two successive such maps is zero is also very easy.
With this discussion, we are now ready to define the families of complexes that
we want to have at our disposal.
Families of complexes 85

III.3 Families of complexes


For the rest of this chapter, F and G will denote free R-modules of ranks m and
n, with m ≥ n.
Let f : F → G be a map of free modules over the commutative ring R, and
let q be a positive integer.

Definition III.3.1 We define the complex C(q; f ) as follows:



q
0 → Cm−n+1 → · · · → Ckq → · · · → Λn−q+s0 G∗ ⊗ Λs1 G∗ ⊗ Λn+|s| F
si ≥1

→ Λn−q+s G∗ ⊗ Λn+s F → Λq F → Λq G,
s≥1

where
C1q = Λq F
and

Ckq = Λn−q+s0 G∗ ⊗ Λs1 G∗ ⊗ · · · ⊗ Λsk−2 G∗ ⊗ Λn+|s| F, k ≥ 2,
si ≥1

|s| = si . The maps (except for Λq f : Λq F → Λq G) are the bar complex maps
associated to the action of the algebra ΛG∗ on ΛF.

As the signs of all our maps are quite crucial, we will make clear just what we
mean by “boundary map” in this context. Namely, if a0 ⊗a1 ⊗· · ·⊗ak−2 ⊗x ∈ Ckq
for k ≥ 3, then
∂(a0 ⊗ a1 ⊗ · · · ⊗ ak−2 ⊗ x)


k−3
= a0 ⊗ a1 ⊗ · · · ⊗ ak−2 (x) + (−1)k−i a0 ⊗ a1 ⊗ · · · ⊗ ai ∧ ai+1 ⊗ · · · ⊗ ak−2 ⊗ x.
i=0

Definition III.3.2 We define the complex T(q; f ) as follows:


q
0 → Tm−n+1 → · · · → Tkq → · · · → K(2,1n−q ) G∗ ⊗ Λn+2 F
→ Λn−q+1 G∗ ⊗ Λn+1 F → Λq F → Λq G,
where
T1q = Λq F
and
Tkq = K(k−1,1n−q ) G∗ ⊗ Λn+k−1 F, k ≥ 2.
86 Generalized Koszul complexes

The maps (except for Λq f : Λq F → Λq G and Λn−q+1 G∗ ⊗ Λn+1 F → Λq F ) are


given as the following compositions of maps:
K(k−1,1n−q ) G∗ ⊗ Λn+k−1 F → Λn−q+1 G∗ ⊗ Dk−2 G∗ ⊗ Λn+k−1 F
→ Λn−q+1 G∗ ⊗ Dk−3 G∗ ⊗ Λn+k−2 F.
From the discussion of the last section, we know that the image of this
composition lies in K(k−2,1n−q ) G∗ ⊗ Λn+k−2 F, so this defines our desired
boundary map.
We should point out, as we did in the earlier section, that the term Λn−q+1 G∗ ⊗
Λn+1 F is also of the general form: it is K(1,1n−q ) G∗ ⊗ Λn+1 F.
In Chapter II, where we introduced the Koszul complex, we saw that the
Koszul complex associated with a map g : F ⊕ R → R was the mapping cone
of the Koszul complex associated to the map f : F → R mapped into itself
by multiplication by the image of 1 under the map of R to R. The complexes
we introduced above share that property with the Koszul complex, although we
have to be a bit more careful in setting up our mapping cones in this case. To
this end, we make two more definitions.
Definition III.3.3 Let f : F → G be a map of free R-modules.
1. For every pair of integers (q, l), with 1 ≤ q ≤ n + 1 and l ≤ m, define the
complex C(q, l; f ) as follows:
(q,l) (q,l)
0 → Cm−n+q−l → · · · → Ck → · · ·

→ Λn−q+s1 G∗ ⊗ Λs2 G∗ ⊗ Λn−q+|s|+l F
si ≥1

→ Λn−q+s G∗ ⊗ Λn−q+s+l F → Λl F,
s≥1

where

Λn−q+s1 G∗ ⊗ Λs2 G∗ ⊗ · · · ⊗ Λsk ⊗ Λn−q+|s|+l F,
(q,l)
Ck = k ≥ 1,
si ≥1

and the maps are just those of the bar complex.


2. For every pair of integers (q, l), with 1 ≤ q ≤ n + 1 and l ≤ m, define the
complex T(q, l; f ) as follows:
→ · · · → K(2,1n−q ) G∗ ⊗ Λn−q+2+l F
(q,l) (q,l)
0 → Tm−n+q−l → · · · → Tk
→ K(1,1n−q ) G∗ ⊗ Λn−q+1+l F → Λl F
where
= K(k,1n−q ) G∗ ⊗ Λn−q+k+l F,
(q,l)
Tk
and the maps are essentially the ones we defined for the complexes T(q; f )
above.
Families of complexes 87

Note that our definition does not preclude the possibility that l be negative,
or that q be equal to n+1. In our immediate use of these complexes, these rather
bizarre possibilities will not show up. In the next section, however, we will make
use of this flexibility when we consider various proofs of “generic acyclicity”
(a term that will be defined in that section) of our complexes.
We now have the following proposition.

Proposition III.3.4 Let F and G be free R-modules. Let f : F → G and


γ : R → G be maps, and let g : F ⊕ R → G be the sum of the maps f and
γ. Then for all q > 0, γ induces maps γ1 : C(q, q − 1; f ) → C(q; f ) and γ2 :
T(q, q − 1; f ) → T(q; f ) such that the mapping cone of γ1 is C(q; g) and the
mapping cone of γ2 is T(q; g).

Proof Let y0 = γ(1). We define γ1,0 = γ2,0 : Λq−1 F → Λq G by

γ1,0 (x) = γ2,0 (x) = y0 ∧ Λq−1 f (x),

(q,q−1)
and γ1,k : Ck → Ckq by

γ1,k (b1 ⊗ · · · ⊗ bk ⊗ x) = b1 ⊗ · · · ⊗ bk−1 ⊗ y0 (bk )(x).

(q,q−1) (q,q−1)
To define γ2,k : Tk → Tkq we use the fact that Tk is the image of
n−q ∗ ∗ α n−q+1 ∗ ∗
Λ G ⊗ Dk G ⊗ Λ n+k−1
F →Λ G ⊗ Dk−1 G ⊗ Λ n+k−1
F, while Tkq is the
β
image of Λn−q G∗ ⊗ Dk−1 G∗ ⊗ Λn+k−1 F → Λn−q+1 G∗ ⊗ Dk−2 G∗ ⊗ Λn+k−1 F. We
then note that we have a map, which we will denote by ∂γ : Dl G∗ → Dl−1 G∗ ,
defined as the composition
∆⊗γ
Dl G∗ = Dl G∗ ⊗ R → Dl−1 G∗ ⊗ G∗ ⊗ G → Dl−1 G∗ ⊗ R = Dl−1 G∗ ,
1⊗ω

where ∆ : Dl G∗ → Dl−1 G∗ ⊗G∗ is the indicated diagonal map, and ω : G∗ ⊗G →


R is the map that sends an element b ⊗ a to a(b). Observe next that the following
diagram commutes:

∂γ
Λn−q G∗ ⊗ Dk G∗ ⊗ Λn+k−1 F → Λn−q G∗ ⊗ Dk−1 G∗ ⊗ Λn+k−1 F
↓α ↓β
∂γ
Λn−q+1 G∗ ⊗ Dk−1 G∗ ⊗ Λn+k−1 F → Λn−q+1 G∗ ⊗ Dk−2 G∗ ⊗ Λn+k−1 F.

Then the map ∂γ = 1 ⊗ ∂γ ⊗ 1 carries the image of α to the image of β, and this
is the map γ2,k that we are after.
As in the case of the classical Koszul complex, the fact that the mapping cones
of these maps are the appropriate complexes is not difficult to prove; again one
simply uses the identification of Λl (F ⊕ R) with Λl F ⊕ Λl−1 F. 2
88 Generalized Koszul complexes

Notice that when G = R, so that n = 1 and q = 1, the complex C(1; f ) is just


the Koszul complex (as is also T(1; f )), and the complex C(1, 0; f ) is also just
the Koszul complex associated to the map f : F → R.

III.3.1 The “homothety homotopy”


When we introduced the Koszul complex associated to a map f : F → R, we
showed that if a = f (x), for some x ∈ F, then multiplication by a on the Koszul
complex is homotopic to zero. Since multiplication by an element of the ring is
called a homothety, we can call the homotopy a “homothety homotopy.” What
we will show here is that if f : F → G is a map as above, then for each λ ∈ Λn F
and ξ ∈ Λn G∗ , the homothety ξ(λ) on the complex C(q; f ) is homotopic to zero.
This not only shows that the homology of C(q; f ) is annihilated by ξ(λ), but
that if A is any additive functor from the category of R-modules to itself which
preserves homotheties, the element ξ(λ) ∈ R also annihilates the homology of
A(C(q; f )) (since additive functors preserve chain homotopies).
We take ξ ∈ Λn G∗ and λ ∈ Λn F, and we want to show that multiplication in
the fat complex C(q; f ) by µ = ξ(λ) is homotopic to zero. We define

σ 0 : Λq G → Λ q F

by setting

σ0 (y) = y(ξ)(λ).

In order to show that ∂σ0 (y) = µy, we must prove some basic lemmas about the
operation of ΛG∗ on ΛG in general.

Lemma III.3.5 Let α ∈ Λl G∗ and u ∈ Λr G, v ∈ Λs G. Then


 
α(u ∧ v) = (−1)r|αi | αi (u) ∧ αi (v),

where we set ∆(α) = αi ⊗ αi , and |αi | denotes the degree of αi . (One says
that the operation of ΛG on ΛG satisfies the measuring identity or that ΛG∗

measures ΛG.)

Proof We will not give a detailed proof of this, but just an outline. For |α| = 1,
this is easy: it is just the statement (which we used in Chapter II) that the bound-
ary map of the Koszul complex is a derivation. We then proceed by induction
on the degree of α, and assume that α = α1 ∧ α2 where |α1 | = 1. One then uses
the fact that


∆(α1 ∧ α2 ) = ∆(α1 )∆(α2 ) = (α1 ⊗ 1 + 1 ⊗ α1 ) α2i ⊗ α2i
 

= α1 ∧ α2i ⊗ α2i + (−1)|α2i | α2i ⊗ α1 ∧ α2i
,
Families of complexes 89

together with the induction assumption that


  
α(u ∧ v) = α1 (−1)r|α2i | α2i (u) ∧ α2i

(v) .

The rest is simply careful comparison of signs. 2

An immediate corollary of this is the following useful fact.

Lemma III.3.6 Let u ∈ Λr G, v ∈ Λs G and α ∈ Λl G∗ . Then


 
u(α)(v) = (−1)|ui |(l−1) ui ∧ α(ui ∧ v).

Proof Again, we will not prove this; we just indicate that in this case, one can
use induction on the degree of u. 2

Using the formula above, one sees immediately that ∂σ0 (y) = µy, so our
homotopy is under way. That is, we have
 
∂σ0 (y) = (−1)|yi |(n−1) yi ∧ ξ(yi ∧ λ) = y ∧ ξ(λ) = µy,

because most of the summands disappear in the above sum (since the λ is now to
be considered as sitting inside Λn G, so that multiplication with yi is zero unless
degree of yi is zero).
To proceed with the definitionof the homotopy, we refine our notation a bit.
We have been writing ∆(x) = j xj ⊗ xj to indicate the total diagonal of x.
However, we occasionally want to specify the degrees
  of the terms that occur in
the sum, so to do that we will write ∆(x) = j l xjl ⊗ xjr−l to indicate the
degree l of the term xjl , and the degree r − l of the term xjr−l if the element x
is of degree r.
We now define

σ1 (x) = xjl (ξ) ⊗ xjq−l ∧ λ
j l<q

where x ∈ Λq F.
In order to see that this works, we prove the following two lemmas:

Lemma III.3.7 For x ∈ Λq F, ξ and λ as before, and for each integer l, we


have

 
q−l  
l+t 
xjl (ξ)(xjq−l ∧ λ) = (−1)t xjq−l−t ∧ xjl+t (ξ)(λ).
j j t=0
l
90 Generalized Koszul complexes

Proof By the measuring identity, we see that


 
xjl (ξ)(xjq−l ∧ λ) = ±ξit (xjq−l ) ∧ xjl (ξin−t

)(λ)
j j i,t

= ±xjq−l−t ∧ xjt (ξit )xjl (ξin−t

)(λ)
j i,t
 
 
= ±xjq−l−t ∧ xjl xjt (ξit )ξin−t

(λ)
j t i

= ±xjq−l−t ∧ xjl (xjt (ξ))(λ)
j t
  l + t
= ± xjq−l−t ∧ xjl+t (ξ))(λ).
j t
l
2
Lemma III.3.8 For x ∈ Λ F, and ξ and λ as before, we have
q

xjl (ξ)(xjq−l ∧ λ) = x ∧ ξ(λ) − x(ξ) ∧ λ.
j l<q

Proof As we saw above, for each l we have



q−l  
l+t 
xjl (ξ)(xjq−l ∧ λ) = (−1)t xjq−l−t ∧ xjl+t (ξ)(λ).
t=0
l
So this says that
    
 l β
xjl (ξ)(xjq−l ∧ λ) = (−1)β
(−1) x ∧ xjβ (ξ)(λ),
l jq−β
l<q β l

and the conclusion follows (because we know what happens to the alternating
sum of binomial coefficients). 2
Now it is easy to see that σ1 gives us what we need.
The next step is the “generic” one; that is, once we get this one, the others all
are of the same type. We define
 
σ2 (β ⊗ x) = β ⊗ xjn+s−q+l (ξ) ⊗ xjq−l ∧ λ
j l<q−s

for β ∈ Λn−q+s G∗ , x ∈ Λn+s F. There are two “tricks” to showing that this works:
one is to recognize that
β(xjn+s−q+l )(ξ) = β ∧ xjn+s−q+l (ξ)
because ξ is of degree n (we are applying the
corollary to the measuring formula).
Then when we compute the boundary of l<q−s β ⊗ xjn+s−q+l (ξ) ⊗ xjq−l ∧ λ,
Families of complexes 91

we can allow l = q − s, since the zero degree term in the middle cancels out in
the boundary. But then, since the degree of x is greater than n (since s ≥ 1),
the term x(ξ) = 0, so when we apply our lemma, we see that this definition of
σ2 works.
Now it is easy to see that for k ≥ 2, we may define
 
σk+1 (β0 ⊗ · · · ⊗ βk−1 ⊗ x) = β0 ⊗ · · · ⊗ βk−1 ⊗ xjn+|s|−q+l (ξ)
j l<q−|s|

⊗ xjq−l ∧λ

for β0 ∈ Λn−q+s0 G∗ , βi ∈ Λsi G∗ , for i ≥ 1, and x ∈ Λn+|s| F (|s| = i≥0 si ).
We therefore state (without proof) the following theorem.

Theorem III.3.9 Let λ ∈ Λn F, ξ ∈ Λn G∗ , and let µ = ξ(λ). Define σ :


C(q; f ) → C(q; f ) by

σ0 (y) = y(ξ)(λ);

σ1 (x) = xjl (ξ) ⊗ xjq−l ∧ λ;
j l<q

and
 
σk+1 (β0 ⊗ · · · ⊗ βk−1 ⊗ x) = β0 ⊗ · · · ⊗ βk−1 ⊗ xjn+|s|−q+l (ξ)
j l<q−|s|

⊗ xjq−l ∧ λ

for k ≥ 1. Then σ is a homothety homotopy for the homothety µ.

III.3.2 Comparison of the fat and slim complexes


As we would much prefer to work with the slim complexes, we would like to
know that they, too, have homothety homotopies for every homothety µ of the
previous subsection. In fact, just as a matter of curiosity, it is worth knowing just
how the fat and slim complexes do compare. What we will show here is that for
every q > 0, there are maps αq : T(q; f ) → C(q; f ), θq : C(q; f ) → T(q; f ) such
that θq αq = 1, and αq θq is homotopic to the identity. Using these maps, we will
be able to transfer the homotopy we have on the fat complexes to the slim ones.
We will also have established the fact that these two families of complexes are
homotopically equivalent, so that our cavalier “slimming down” that we did was
justified. We will not present all the details of proof of these facts; the interested
reader is referred to Reference [12] for the full story.
We first define the map αq : T(q; f ) → C(q; f ). In dimensions 0 and 1, of
course, the map is the identity. For ease of notation, we will label these maps
q q
α−2 and α−1 respectively and, in general, we will denote by αkq the map that
q q
takes Tk+2 to Ck+2 .
92 Generalized Koszul complexes

Definition III.3.10 (Definition of the map αq ) For q = −2 and −1 we


define αkq to be the identity. For k ≥ 0, we define αkq as the composition
q
Tk+2 → Λn−q+1 G∗ ⊗ Dk G∗ ⊗ Λn+k+1 F →
Λn−q+1 G∗ ⊗ D1 G∗ ⊗ · · · ⊗ D1 G∗ ⊗ Λn+k+1 F,
% &# $
k

where the right arrow is the k-fold diagonalization of Dk G∗ . (We observe that
q
the latter term is a summand of Ck+2 .)
It is a relatively straightforward calculation to see that αq , thus defined, is a
map of complexes.
To define the maps θq : C(q; f ) → T(q; f ) requires a bit of extra preparation.
In dimensions 0 and 1 it is of course defined to be the identity and denoted by
q q
θ−2 and θ−1 and as in the definition of the map αq , we will denote by θkq the
q q
map that sends Ck+2 to Tk+2 . But now we introduce some notation to facilitate
the definition.
Notation
Assume that a fixed basis of G∗ is given, say y1 , . . . , yn . An element yj1 ∧ · · · ∧ yjl
will be written either as j1 ∧ · · · ∧ jl or j1 · · · jl or simply as J. In short, the index
on a basis element will be used to denote that element (as is the practice when
working with tableaux), and products of elements will be denoted by products of
their indices. For l = 1, we will usually write j or j1 instead of J. When working
with products in the divided power algebra, we will use tableau notation in order
to avoid confusion about whether juxtaposition means the usual product within
that algebra, or the divided power when there are repeats. For example, we will
(2)
write, for yu1 yu2 , with u1 < u2 , the tableau [u1 u1 u2 ].
We will use freely the standard basis, consisting of standard tableaux, for
“hook Weyl modules” (as described in Subsection III.1.2), and a typical basis
q
element of Tk+2 would be denoted by
 
j1 u1 · · · uk
 j2 
 
 .. ⊗x
 . 
jn−q+1
 
j1 u1 ··· uk
 j2 
 
where x is an element of Λn+k+1 F, and  ..  is a standard
 . 
jn−q+1
tableau which stands for the image of the element j2 · · · jn−q+1 ⊗ [j1 u1 · · ·
uk ] ∈ Λn−q G∗ ⊗Dk+1 G∗ in Λn−q+1 G∗ ⊗Dk G∗ under the map which diagonalizes
Dk+1 G∗ → G∗ ⊗ Dk G∗ , and then multiplies Λn−q G∗ into Λn−q+1 G∗ by using
the G∗ factor.
Families of complexes 93

Recall that “standard tableau” means that the indices are strictly increasing
in the column, and weakly increasing in the row. For reasons that will become
apparent later, we will make one more (unusual) convention, which we will use
in this section only, about our use of tableau notation in the case of rows: we will
assume that the tableau is zero if the top row is not weakly increasing as written.
(2)
Thus, in the case of yu1 yu2 with u1 > u2 , we would have to write [u2 u1 u1 ]
to represent it as a tableau.
We are now in a position to define our maps θkq for all k ≥ −2.
Definition III.3.11 (Definition of the map θq ) For k = −2 and −1, we
have already said that the map is to be the identity. For k = 0, and Y = J ⊗ x,
with J a basis element of Λn−q+s G∗ and x ∈ Λn+1 F, we define
θ0q (Y ) = (−1)(s−1)(n−q+1) js · · · jn−q+s ⊗ j1 · · · js−1 (x).
For q > 0, and Y = J ⊗ u1 ⊗ · · · ⊗ uk ⊗ x with all the ui basis elements of degree
one, and J still of degree n − q + s, we define
 
js uk · · · u1
 js+1 
 
θk (Y ) = (−1)(s−1)(n−q+1)+(s−1)k  .  ⊗ j1 · · · js−1 (x).
 .. 
jn−q+s
It is essential to remember here that the tableau is to be read as equal to zero if
the row is not standard. Assume that the map θlq has been defined on elements
Y  = J ⊗U1 ⊗· · ·⊗Ul ⊗x for Ui basis elements of arbitrary degree, and l < k, and
that θkq has been defined on elements Z = J ⊗U1 ⊗· · ·⊗Uk ⊗x with U1 , . . . , Ut basis
elements of arbitrary degree (we make the convention that U0 = J), Ut+2 , . . . , Uk
basis elements of degree 1, and Ut+1 basis element of degree st+1 ≤ r (r > 0).
We now let Z  = J ⊗ U1 ⊗ · · · ⊗ Ut+1 ⊗ · · · ⊗ Uk ⊗ x with the basis element
Ut+1 = v ∧ W, v of degree 1, all of the basis elements Ut+2 , . . . , Uk are of degree
1, and degree(W ) = r. Define
θkq (Z  ) = θkq (B + (−1)(k−t−1)r E)
where B = J ⊗ U1 ⊗ · · · ⊗ Ut v ⊗ W ⊗ Ut+2 ⊗ · · · ⊗ Uk ⊗ x, and E = J ⊗ U1 ⊗
· · · ⊗ Ut ⊗ v ⊗ Ut+2 ⊗ · · · ⊗ Uk ⊗ W (x).
Notice that in position t + 1, the elements B and E are of degree less than
or equal to r, while the terms of higher index are not affected in degree. From
this we see immediately (using a simple induction proof) that θkq (Z  ) = 0 unless
Ut+2 ≥ · · · ≥ Uk .
While it is trivial to show that the map αq is a map of complexes, it is not
trivial to prove that the same is true of θq . It is at this point that we seriously
invoke the suggestion that the reader look at Reference [12] for details. In any
event, it is quite straightforward to show that the composition θq αq is equal to
the identity on the complex T(q; f ).
94 Generalized Koszul complexes

This fact is actually enough for us to transfer the homothety homotopy from
the fat to the slim complexes. In fact, we have the following general lemma that
will give us the result we want.

Lemma III.3.12 Let X and Y be complexes, α : X → Y and θ : Y → X maps


of complexes such that θα = idX . Let s : Y → Y be a homotopy on Y which
makes the scalar µ homotopic to zero on Y. Then the map τ = θsα : X → X is
a homotopy on X carrying µ to zero.

Proof First we have to show that ∂X τ0 = µ. But ∂X τ0 = ∂X θ1 s0 α0 =


θ0 ∂Y s0 α0 = θ0 µα0 = µθ0 α0 = µ. For i > 0, we have to show that

∂X τi + τi−1 ∂X = µ.

But here again we have

∂X τi = ∂X θi+1 si αi = θi ∂Y si αi = θi (µ − si−1 ∂Y ) αi
= µ − θi si−1 ∂Y αi

while

τi−1 ∂X = θi si−1 αi−1 ∂X = θi si−1 ∂Y αi ,

and this does it. 2

With this lemma, we now know that our slim complexes carry the desired
homotopy.
Although, as we said, we will not give a proof of the following fact, we state
it for the sake of completeness.

Theorem III.3.13 The composition of maps, αq θq is chain homotopic to the


identity map on the complex C(q; f ). Thus, the complexes C(q; f ) and T(q; f )
are homotopically equivalent.

III.4 Depth-sensitivity of T(q; f )


In this section we prove that the complexes T(q; f ), like the Koszul complex, give
us a way of determining the depth of the ideal generated by the maximal minors
(that is, the minors of order n = rankG). Recall that we are always assuming
that m ≥ n, where m = rankF. The method of proof that we will use here is the
one that was used in Reference [27], Proposition 2.1.

Definition III.4.1 A sequence of additive functors {T i ; i ≥ 0} from an abelian


category, A, to another, B, is called a homology functor if
1. For every exact sequence E : 0 → A → B → C → 0 in A, and every i ≥ 0,
there is given a morphism ∂ i (E) : T i (C) → T i+1 (A) which is “natural”, that is,
Depth-sensitivity of T(q; f ) 95

if E : 0 → A → B  → C  → 0 is another exact sequence, and ϕ : E → E is a


map of complexes, then the diagram
∂ i (E)
T i (C) → T i+1 (A)
↓ ↓
∂ i (E )
T i (C  ) → T i+1 (A )
is commutative.
2. For each exact sequence E above, the sequence
∂ 0 (E)
0 → T 0 (A) → T 0 (B) → T 0 (C) → T 1 (A) → · · ·
∂ i (E)
→ T i (C) → T i+1 (A) → · · ·
is exact.
As one example, one that we will be dealing with almost immediately, we may
consider the complex T(q; f ), and define the functors
T i (A) = H i (HomR (T(q; f ), A)) .
Another one, equally relevant, is to define the functors
U i (A) = Hm−n+1−i ( T(q; f ) ⊗R A).
Because of the fact that T(q; f ) is a free complex, tensoring with or taking
Hom into a module, preserves exact sequences. Thus, to each exact sequence of
R-modules, we get a long exact sequence of homology (or cohomology) modules.
Hence these are examples of homology functors.
Theorem III.4.2 Let I be an ideal in the commutative, noetherian ring R, and
let {T i } be a homology functor from the category of finitely generated R-modules
into itself which satisfies the following conditions:
1. T i (α) = α for every homothety α ∈ R;
2. SuppT i (A) ⊆ Supp R/I;
3. T 0 (A) = 0 if and only if Supp R/I ∩ Ass(A) = 0, where Ass(A) is the set
of all associated primes of A.
Then for each R-module, A, such that A/IA = 0, we have that depth(I; A)
is equal to the smallest integer r for which T r (A) = 0. Furthermore, if T 0 (∗) =
HomR (M, ∗) for some fixed R-module M , then T d (A) = ExtdR (M, A), where
d = depth(I; A).
Proof We prove the first part of the statement by induction on depth(I; A)
(which we know to be a non-negative integer).
Since A/IA = 0, we know that depth(I; A) = 0 if and only if every element
of I is a zero divisor for A. But this means that depth(I; A) = 0 if and only if I
is contained in the union of the associated primes of A. Thus depth(I; A) = 0 if
96 Generalized Koszul complexes

and only if Supp R/I ∩ Ass(A) = 0, and this is if and only if T 0 (A) = 0. Hence
we have our result for depth(I; A) = 0.
Let c(A) be the smallest integer r such that T r (A) = 0, and let us assume that
depth(I; A) = d > 0. We want to show, as we have in the case depth(I; A) = 0,
that d = c(A). Let α1 , . . . , αd be a maximal regular A-sequence contained in I,
and consider the exact sequence
0 → A → A → A/α1 A → 0.
We know that depth(I; A/α1 A) = d − 1 so, by induction, we have that
c(A/α1 A) = d − 1. From the exactness of
T i (α1 )
· · · → T i−1 (A/α1 A) → T i (A) → T i (A) → T i (A/α1 A) → · · ·
it follows that T i (α1 ) is a monomorphism for all i ≤ c(A/α1 A). But if T i (α1 ) is
a monomorphism, then so is T i (α1h ) for all h. However, SuppT i (A) ⊆ Supp R/I,
so that the radical of the annihilator of T i (A) contains the radical of I. Since
α1 ∈ I, some power of α1 must annihilate T i (A), and since we are assuming that
α1h = T i (α1h ), we cannot have that T (α1h ) is a monomorphism unless T i (A) = 0.
From this we see that T i (A) = 0 for all i ≤ d − 1. To see that T d (A) = 0, we
simply observe that
0 → T d−1 (A/α1 A) → T d (A),
is exact, and that tells us what we want to know, that is, that c(A) = d. This
concludes the proof of the first statement of our theorem.
Before we prove the second part, notice that the functors ExtiR (R/I, ∗) and
ExtiR (R/radI, ∗) each satisfy the conditions of our theorem with respect to
either of the ideals, I and radI. Thus we see immediately that depth(I; A) =
depth(radI; A).
Now for the proof of the second statement: we assume that there is a fixed
R-module, M, such that T 0 (∗) = HomR (M, ∗). If we take any prime ideal p, and
use the fact that
Supp HomR (M, R/p) = Supp T 0 (R/p) ⊆ Supp R/I,
we see that Mp = 0 for all p not in Supp R/I, that is, Supp M ⊆ Supp R/I.
Therefore, both {T i (∗)} and {ExtiR (M, ∗)} are homology functors satisfying the
conditions of the theorem. Suppose that d = depth(I; A). If we let J = Ann(M )∩
Ann(T d (A)) ∩ I, and apply the fact that Supp R/I contains both Supp M and
Supp T d (A), we see that radJ = radI. Therefore depth(J; A) = depth(I; A) = d,
and we can find a maximal A-regular sequence α1 , . . . , αd in J. Now, by the
property of being homology functors, and from the fact that the αi annihilate
T d (A) and ExtdR (M, A), we have
T d (A) ∼
= T 0 (A/(α1 , . . . , αd )A) = HomR (M, A/(α1 , . . . , αd )A) ∼
= ExtdR (M, A),
and the proof of the theorem is complete. 2
Depth-sensitivity of T(q; f ) 97

In the examples we gave immediately after the definition of homology functor


(Definition III.4.1), we see that if we choose I to be the ideal I(f ) generated
by the minors of order n of the morphism f, then both satisfy the conditions
(1) and (2) of our theorem with respect to I(f ). This is where we make use of
the fact that we have homothety homotopies which show that the ideal I(f )
annihilates all of our homology in these complexes. In the case of the functors
T i , we do see that T 0 (A) = HomR (M, A), where M = cokerf. It is easy to
see that Supp M = SuppR/I(f ). For our homothety homotopy says that the
0-dimensional homology of our complex, namely M, is annihilated by all the
n × n minors of f, so I(f ) ⊆ ann(M ). On the other hand, if a ∈ ann(M ),
then an ∈ ann(coker(Λn f )) = I(f ). This says that a ∈ radI(f ), so we have
I(f ) ⊆ ann(M ) ⊆ radI(f ), and establishes the asserted equality of supports.
What about U 0 (A)? Can this also be expressed as Hom of some fixed module
into A? To see that the answer is yes, we first consider the case m > n + 1, and
recognize that U 0 (A) =
ker(K(m−n,1n−q ) G∗ ⊗ Λm F ⊗ A → K(m−n−1,1n−q ) G∗ ⊗ Λm−1 F ⊗ A)
if m > n + 1. Using the remark at the end of the subsection on hooks, we see
that
 ∗
K(m−n,1n−q ) G∗ ⊗ Λm F ∼
= L(n−q+1,1m−n−1 ) G ⊗ Λ0 F
and that
 ∗
K(m−n−1,1n−q ) G∗ ⊗ Λm−1 F ∼
= L(n−q+1,1m−n−2 ) G ⊗ Λ1 F .
Then using the fact that, for any finitely generated free module, H, and any
finitely generated module, A, H ∗ ⊗R A ∼ = HomR (H, A), we see that U 0 (A) is
equal to the kernel of the map
HomR (L(n−q+1,1m−n−1 ) G ⊗ Λ0 F, A) → HomR (L(n−q+1,1m−n−2 ) G ⊗ Λ1 F, A).
 
If we let N = coker L(n−q+1,1m−n−2 ) G ⊗ Λ1 F → L(n−q+1,1m−n−1 ) G ⊗ Λ0 F ,
then the kernel above is simply HomR (N, A). To see what map N is the cokernel
of, consider the commutative diagram:
Λt+1 G ⊗ Sν−2 G ⊗ F → Λt G ⊗ Sν−1 G ⊗ F → L(t,1ν−1 ) G ⊗ F → 0
↓ ↓ ↓ ψt,ν
Λt+1 G ⊗ Sν−1 G → Λ t G ⊗ Sν G → L(t,1ν ) G →0
↓ ↓ ↓
ϕt,ν
Λt+1 G ⊗ Sν−1 M → Λ t G ⊗ Sν M → Γt,ν →0
↓ ↓ ↓
0 0 0
where t = n − q + 1, ν = m − n − 1, M = coker(f ), and Γt,ν = coker(ϕt,ν ) =
coker(ψt,ν ). Clearly, N = Γn−q+1,m−n−1 . The horizontal maps in the first two
rows are those of the graded Koszul complex on G, while the vertical maps
98 Generalized Koszul complexes

between the first and second rows are the operation of F on SG through the
map f and, in the third column, we have the induced maps on the indicated
modules. We are using the fact that if M = coker(f ), for any map f : C → D,
then the sequence
Sν−1 D ⊗ C → Sν D → Sν M → 0
is exact for all ν.
Therefore we see that both T i and U i are homology functors satisfying the con-
ditions of our theorem, or we will as soon as we show that SuppN = Supp M =
Supp R/I(f ). Since M = 0 implies that N = 0, it suffices to show that M = 0
implies that N = 0. By localizing at primes, we may assume that R is local,
and then by reducing modulo the maximal ideal, we may assume that R is a
field. In that case, all the modules are finite-dimensional vector spaces, so we
may assume that G has a basis consisting of a basis of M together with a basis
for the image of f. It is easy, then, to construct an element of Λt G ⊗ Sν M which
is not in the kernel of ϕt−1,ν+1 . But, since the kernel of ϕt−1,ν+1 contains the
image of ϕt,ν , this gives us an element not in the image of ϕt,ν , and we are done.
When m = n, it is easy to see, using the same kinds of identifications used
above, that U 0 (A) ∼= HomR (Q, A), where Q = coker(Λn−q G∗ → Λn−q F ∗ ). Since
m = n, it is easy to prove that SuppQ = Supp R/I(f ). When m = n + 1, similar
identifications show that U 0 (A) = HomR (Q , A), where Q = coker(Λm−q F →
Λm−q G). Again, it is easy to see that the support of Q is equal to that of R/I(f ).
With these observations, we are able to apply our theorem to the functors T i
and U i . In fact, we will state a number of results without proof, as the details
of proof may be found in Reference [27]. (Although the proofs in Reference [27]
relate to the complexes C(q; f ), our discussion above is sufficient to see that they
apply equally to the complexes T(q; f ).)
Theorem III.4.3 Given our map f : F → G and an R-module M such that
M/(I(f ))M = 0, we have for each q with 1 ≤ q ≤ n, the following facts:
1. depth(I(f ); M ) = the smallest integer r for which H r (T(q; f ), M ) = 0, and
furthermore,
H d (T(q; f ), M ) = ExtdR (coker(Λq f ), M )
where d = depth(I(f ); M ).
2. m − n + 1 − depth(I(f ); M ) is equal to the largest integer r for which
Hr (T(q; f ), M ) = 0. Furthermore

ExtdR (coker(Λm−q f ), M ) if m = n or n + 1;
Hm−n+1−d (T(q; f ), M ) =

ExtdR (Γn−q+1,m−n−1 , M ) if m > n + 1.
Corollary III.4.4 supM depth(I(f ); M ) ≤ m − n + 1, where M runs through
all R-modules such that M/I(f )M = 0. In particular, depth(I(f ); R) ≤ m−n+1.
Another kind of multiplicity 99

We should note that there is a stronger result that can be proven, namely:
supp dim Rp ≤ m − n + 1, where p runs through all minimal primes containing
I(f ). The reader may find this in Reference [27], theorem 3.5.

Corollary III.4.5 Given our map f : F → G and an R-module M such that


M/I(f )M = 0, the following statements are equivalent:
1. For some q, 1 ≤ q ≤ n, Hr (T(q; f ), M ) = 0 for all r = 0.
2. For some q, 1 ≤ q ≤ n, H r (T(q; f ), M ) = 0 for all r = m − n + 1.
3. For all q, 1 ≤ q ≤ n, Hr (T(q; f ), M ) = 0 for all r = 0 and
H r (T(q; f ), M ) = 0 for all r = m − n + 1.
4. depth(I(f ); M ) = m − n + 1.
In particular, if coker(f ) = 0, T(q; f ) is a free resolution of coker(Λq f ) for some
q, 1 ≤ q ≤ n (or for all q, 1 ≤ q ≤ n), if and only if depth(I(f ); R) = m − n + 1.

Corollary III.4.6 Let f : F → G be a map with coker(f ) = 0. If


depth(I(f ); R) = m−n+1, then hdR (coker(Λq f )) = m−n+1 for all q, 1 ≤ q ≤ n.

We finish this subsection with the generalized Cohen–Macaulay Theorem due


to J. Eagon [37]. (We actually get the result not only for the ideal I(f ), but for
coker(Λq f ) for all q, 1 ≤ q ≤ n.)
Observe that if M is a Cohen–Macaulay module, then M is equidimensional,
that is, dim Rp is constant (and equal to dim M ) for all p in Ass(M ). We
also note that over a Cohen–Macaulay ring, equidimensionality is the same as
unmixedness, since dim Rp + dim R/p = dim R for all prime ideals p.

Lemma III.4.7 If M is a Cohen–Macaulay R-module, then ExtdR (N, M ) is


equidimensional for any module N such that N ⊗R M = 0, where d is equal to
depth(Ann(N ); M ).

Proof See Reference [27], lemma 2.8. 2

Theorem III.4.8 Let f : F → G be our usual map and assume that


depth(I(f ); R) = m − n + 1. If R is a Cohen–Macaulay ring, then coker(Λq f ) is
unmixed for all q, 1 ≤ q ≤ n.

III.5 Another kind of multiplicity


In Section II.7 we restricted ourselves to ideals of definition of our noetherian
local ring, R, although in Reference [9] the theory was developed for ideals, I,
of R and modules, E, such that the length of E/IE is finite. The analogous
restriction in this section would be to consider morphisms f : Rm → Rn such
that the length of coker(f ) is finite, and then consider, for any R-module, E,
the module coker(f ) ⊗R E. However, we shall adopt the more general setting of
Reference [9] (as this is the point of view taken in Reference [27]), and consider
100 Generalized Koszul complexes

morphisms f : Rm → Rn and modules, E, such that the length of coker(f ) ⊗R E


is finite.
Recall that if M is an R-module, we have associated to it the commutative
algebra, S(M ), that is, the symmetric algebra of M over R. If M = Rm , we
know that S(M ) = S(Rm ) is isomorphic to the polynomial ring in m variables:
R[X1 , . . . , Xm ]. Given the module, M , we can form the graded module, S(M )⊗R
M , over the graded ring, S(M ). What is more, we have the map of S(M ) ⊗R M
into S(M ) given by multiplication by M , that is, since M = S1 (M ), we have
for each ν the multiplication map of Sν (M ) ⊗R M into Sν+1 (M ). We will call
this map τM , and the corresponding Koszul complex we will denote by K(τM ).
When our module is Rm , we will write τm in place of τM , so that the Koszul
complex will be denoted by K(τm ). If N is a graded S(M )-module, the complex
K(τM ) ⊗S(M ) N will be denoted by K(τM ; N ).
Suppose now that we have a map f : Rm → Rn . This induces an algebra map
S(f ) : S(Rm ) → S(Rn ), and thus converts S(Rn ) into a graded S(Rm )-module
(or algebra). Therefore, if E is an R-module, the graded S(Rn )-module, S(Rn )⊗R
E, is a graded S(Rm )-module, and we have the complex K(τm ; S(Rn ) ⊗R E).
It is this complex that we will want to study in more detail, in relation to a
Hilbert–Samuel polynomial that we are about to introduce.
Let us suppose that in addition to our map f : Rm → Rn , we have an
R-module, E, with the property that length(coker(f ) ⊗R E) < ∞. Then it is
straightforward to see that length(coker(Sν (f )) ⊗R E) < ∞ for all ν > 0 (since
Supp coker(f ) = Supp coker(Sν (F )) for ν > 0). We can therefore define the func-
tion, Pf (E; ν) = length(coker(Sν (f ))⊗R E), which is a function from the positive
integers to themselves. One of the main results we have about this function is
the following (see Reference [27], theorem 3.1).

Theorem III.5.1 Let f : Rm → Rn be a map, and E an R-module such


that length(coker(f ) ⊗R E) < ∞. Then Pf (E; ν) is a polynomial function for
sufficiently large ν. Furthermore,

∆m Pf (E; ν) = (−1)m−q length(Hm−q (K(τm ; S(Rn ) ⊗R E)))
q

for all sufficiently large ν.

Proof The proof of this result has the same flavor as that of Theorem II.7.2.
The important observation that gets the proof rolling is that we have the exact
sequence of S(Rm )-modules

0 → S(Rm )E → S(Rn ) ⊗R E → coker(S(f )) ⊗R E → 0,

and while coker(S(f )) ⊗R E is in general not finitely generated over S(Rm ),


the module S(Rm )E is. Thus its Koszul homology is finitely generated, it is a
finitely generated graded module over R (since the homology is killed by the
Another kind of multiplicity 101

image of τm ), and therefore all of its graded components are zero from a certain
integer on.
The rest of the proof is easy to follow. 2
Definition III.5.2 For any polynomial function, ϕ, set
u(ϕ) = (degϕ)! × (the leading coefficient of ϕ).
We now have a sequence of results from Reference [27], section 3, which enable
us to define our generalized multiplicity and relate it to our Koszul complexes.
The first of these is Reference [27], theorems 3.3 and 3.4.
Theorem III.5.3 Let f : Rm → Rn be a map, and E and R-module such that
length(coker(f ) ⊗R E) < ∞. Then u(Pf (E; ν)) and degPf (E; ν) − n + 1 depend
only on the R-modules E and coker(f ). If R is a local ring, then degPf (E; ν) =
n − 1 + dimE.
We also have the following corollary (corollary 3.6 of Reference [27]).
Corollary III.5.4 Let R be a local ring, and f and E as above. Then m − n +
1 ≥ dimE, and hence m ≥ degPf (E; ν).
Definition III.5.5 Given a module, M , of finite length over a local ring, R,
choose an exact sequence (a presentation) Rm → Rn → M → 0. Then for each
R-module, E, the product (dimR + n − 1)!× (the coefficient of the term of degree
n−1+dimR in the polynomial Pf (E; ν)) is a non-negative integer which depends
only on M and E. We call this non-negative integer the multiplicity of M with
respect to E, and denote it by eE (M ).
In the case when E = R, we write eR (M ). When M = R/q with q an ideal of
definition, we retrieve our older definition of multiplicity (except that we have it
written now as eE (R/q) instead of eE (q)).
Definition III.5.6 In view of Corollary III.5.4, we call a map f : Rm → Rn a
parameter matrix for E if length(coker(f ) ⊗R E) < ∞ and m − n + 1 = dim E.
The following proposition sums up where we are to this point (Reference [27],
proposition 3.8).
Proposition III.5.7 Let f : Rm → Rn be a parameter matrix for E, and let
M = coker(f ). Then

1. eE (M ) = q (−1)m−q length(Hm−q (K(τm , (S(Rn ) ⊗R E)ν+q ))) for ν suffi-
ciently large.
2. eE (M ) ≥ 0 and eE (M ) = 0 if and only if dimE < dimR.
3. If 0 → E  → E → E  → 0 is an exact sequence, then eE (M ) = eE  (M ) +
eE  (M ).
In section 4 of Reference [27], there is a large amount of calculation all of which
is directed at proving the following theorem (see Reference [27], theorem 4.2).
102 Generalized Koszul complexes

Theorem III.5.8 Let R be a local ring, f : Rm → Rn a map of free R-modules,


and E an R-module such that length(coker(f ) ⊗R E) < ∞. Then we have
 
n−1
∆m Pf (E; ν) = χH(Λq f, E)
n−q
where H(Λq f, E) stands for the Euler–Poincaré characteristic of either the
complex C(q; f ) (as in the paper, [27]), or T(q; f ).
We remark that in Reference [27], the complexes T(q; f ) did not appear, as
they were not yet known at the time of writing that paper. However, since we
know that the two complexes are homotopic, their homologies are the same, as
are their Euler–Poincaré characteristics. The notation used in the statement of
the theorem is used not only because it is the notation of the statement cited,
but because it is “neutral,” that is, it can refer to either of the two complexes.
We do not give a proof here of the above theorem; one proof can be found
in the cited reference. Also, the works of D. Kirby and D. Rees [58, 59] give
algebraic treatments of this topic, and the paper [60] gives a geometric approach
to the subject.
IV
STRUCTURE THEOREMS FOR FINITE FREE
RESOLUTIONS

In the previous chapter, we saw a whole class of complexes associated with a given
map, f , of finite free modules. We proved that all of them were sensitive to the
depth of the ideal generated by the maximal minors of f , and gave a necessary
and sufficient condition for them to be finite free resolutions of the cokernels
of the exterior powers of f in terms of the depth of the ideal generated by the
maximal minors of f .
In this chapter, we deepen our understanding of what it means for a complex
of free modules to be a resolution, that is, we confront the problem: what does
it mean for a complex of free R-modules to be exact? Given that the underlying
notion of homology, whether it be topological or algebraic, is the exact sequence,
it seems reasonable to ask just what this notion means in everyday terms. Of
course, when we say “everyday terms” we mean in terms of the data that we are
given to hand: in the case of free resolutions, the matrices that represent the
boundary maps of free complexes. This, after all, is what we would look at if
we had a complex of vector spaces.
Of course, there is a vast difference between modules which have finite free
resolutions and those with infinite ones, so it seems reasonable to start with
the more familiar, that is, the ones whose resolutions are finite. Another reason
that compels us to study finite resolutions is that the most familiar rings are
regular: think of the power series rings and polynomial rings. (A commutative
noetherian ring is called regular if its global dimension is finite.) And if one is to
exploit the finiteness of a resolution, one natural starting point is to make use of
the “last” (left-most) matrix that occurs in it. So, from a general point of view,
addressing the question of the structure of finite free resolutions suggests that the
final matrix of the resolution be a jumping-off point. In addition to this purely
theoretical consideration, some problems, such as the Lifting Problem (which
will be explained in this chapter), very much indicated that the last matrix of
a finite free resolution was of significance.
We will see that the complexes studied in Chapter III help us to get some
insight into the minors of the matrices that occur in a resolution. At certain
crucial points, these complexes or their duals will play a role. We will also see
that while our methods carry us some distance toward our goal, there is still a
great deal more to be understood. But first we have to get down to the basics
104 Structure theorems for finite free resolutions

of exactness, and then we will discuss several interesting applications (including


another, better proof of the factoriality of a regular local ring), which in fact
triggered the investigation.
Unless otherwise stated, in this chapter we always assume that we are working
over noetherian rings.

IV.1 Some criteria for exactness


Let R be a nontrivial commutative ring (with 1). Consider the following complex
of finitely generated free R-modules
fn fk f1
F : 0 → Fn → Fn−1 → · · · → Fk → Fk−1 → · · · → F1 → F0 .
We discuss what conditions guarantee that F is exact, that is, that its
homology modules (which exist only in positive dimension) are all trivial.
Definition IV.1.1 Given a map ψ : E → F of finite free R-modules, define
the rank of ψ, written rank(ψ), to be the largest integer k such that the map
Λk ψ : Λk E → Λk F is not zero. If r denotes this rank, define I(ψ) to be the ideal
Ir (ψ) introduced after Remark I.3.32, namely, the ideal of “r × r minors” of ψ.
We see immediately that 0 ≤ r ≤ min{rankE, rankF } = s, since the map
Λ0 ψ : Λ0 E → Λ0 F is the identity map on R, and Λs+1 ψ : Λs+1 E → Λs+1 F is
clearly zero. We should also note that the ideal of 0 × 0 minors (i.e., I0 (ψ)) is
always equal to (1) = R. Another special case to mention is that of ψ = 0. In
this case, we have rank(ψ) = 0, and I(ψ) = R.
We have already met an instance of I(ψ) in Section III.4; ψ was the map
f : F → G with rankF = m ≥ n = rankG and rankf = n.
Remark IV.1.2 In general, the ideal I(ψ) does not localize. That is, if ψ :
F → G is a map of free modules over R, and we localize with respect to a
multiplicative subset, S, we may consider I(ψ) ⊗ Rs and I(ψ ⊗ RS ). In general,
these two ideals do not coincide. For instance, if e is a nontrivial idempotent
of R, m a maximal ideal containing e, and ψ : R → R the map sending 1 to
e, then I(ψ) = (e), (I(ψ))m = 0 but I(ψm ) = Rm . However, if I(ψ) contains a
non-zero divisor, then we have I(ψ)RS = I(ψ ⊗ RS ) for every multiplicatively
closed subset, S, of R.
Definition IV.1.3 A projective R-module, P , is said to have defined rank
if the free Rm -module Pm has the same rank for every maximal ideal m. This
common rank is defined to be the rank of P .
When R is a noetherian ring with no nontrivial idempotents, every projective
module has defined rank (cf., e.g. Reference [80],theorems 7.8 and 7.12).
Example IV.1.4 Let e be a nontrivial idempotent of R. Then R/(e) is a
projective R-module but its rank is not well defined.
Some criteria for exactness 105

Proof R/(e) is projective because the short exact sequence


inc
0 → (e) → R → R/(e) → 0
splits, by means of the map R → (e) sending 1 to e.
Suppose that m is a maximal ideal of R, and consider I(incm ). If e ∈ m,
then 1 − e is not in m, and every element of (e) is annihilated by 1 − e. Thus
im(incm ) = 0, and the rank of (R/(e))m is 1. If e ∈/ m, then (e) localized at m is
Rm , so that the rank of (R/(e))m is 0. Since each of the elements e and 1−e must
be contained in some maximal ideal, and since they cannot both be contained
in the same one, we see that R/(e) does not have well-defined rank. 2

Proposition IV.1.5 Given ψ : E → F as before, coker(ψ) is projective and


has well-defined rank if and only if I(ψ) = R.

Proof We first do the “if” part. Since I(ψ) = R contains a non-zero divisor,
rank(ψ) does not change when localizing. Hence we assume that R is local and
prove that coker(ψ) is free of rank equal to rankF − rank(ψ). If we call r the
rank of ψ, I(ψ) = R implies that some r × r minor of ψ is invertible. Hence we
may choose bases for E and F such that ψ has block matrix
 
AB
C D
with A equal to the r × r identity matrix, Ir . Elementary row and column oper-
ations show that B and C may be assumed to be zero. But then D must be
zero as well, otherwise the rank of the overall matrix would exceed r. It is now
obvious that the cokernel of
 
Ir 0
0 0
is free of rank equal to rankF − rank(ψ).
Now we do the “only if” part. Since coker(ψ) is projective, we get F ∼
=

coker(ψ) ⊕ im(ψ), E ∼
0
= ker(ψ) ⊕ im(ψ), and ψ = idim(ψ) ⊕ ker(ψ) → coker(ψ) .
Let m, n, t be the ranks of E, F , coker(ψ), respectively. If t = n, then ψm = 0
for every maximal ideal m, hence ψ = 0 and I(ψ) = R, as expected. If t < n,
then ψm always has the matrix
 
In−t 0
,
0 0
hence det In−t = 1 ∈ I(ψm ) and I(ψm ) = (I(ψ))m = Rm for every m. Again
R = I(ψ), as required. 2

The above proposition reaffirms that the rank of R/(e) (of Example IV.1.4)
ψ
is not well defined since it has a presentation R → R → R/(e) → 0 such that
106 Structure theorems for finite free resolutions

I(ψ) = R (just take ψ to be multiplication by e and observe that the nontriviality


of e means that I(ψ) = (e) is a proper ideal).
We point out that the above proposition still holds under the assumptions
that R is a noetherian ring with no nontrivial idempotents, and E and F are
projective. That is, we can weaken the hypothesis that the modules be free,
provided that we can still guarantee well-defined rank.
The following result generalizes a well-known property of vector spaces over a
field.
Proposition IV.1.6 Given a complex
ψ ϕ
C: E→F →G
of finite free R-modules such that I(ψ) = R = I(ϕ), C is exact if and only if
rankF = rankψ + rankϕ.
Proof As in the first part of the last proof, we may assume that R is local.
Thus a maximal minor of ψ (resp. of ϕ) is invertible, thanks to the fact that
I(ψ) = R (resp. I(ϕ) = R). Hence we may choose bases for E, F and G such
that ψ and ϕ have block matrices
   
Ir 0 00
and , respectively,
0 0 0 Is
where Ir is the r × r identity matrix, Is is the s × s identity matrix, and r =
rank(ψ), s = rank(ϕ). It is now obvious that C is exact if and only if r + s equals
rankF . 2
Theorem IV.1.7 The complex
fn fk f1
F : 0 → Fn → Fn−1 → · · · → Fk → Fk−1 → · · · → F1 → F0
given at the beginning of this section is exact if and only if
(1) rankFk = rankfk+1 + rankfk
(2) depthI(fk ) ≥ k
for all k = 1, . . . , n.
Notice that we make the convention that the ideal, (1), has infinite depth.
Proof We will start with the “only if” part. As usual, we want to use local-
ization to prove the result. In order for I(fk ) to localize, we prove that I(fk )
contains a non-zero divisor. By inverting all non-zero divisors, we may assume
that R is a semilocal ring (if every non-zero divisor is a unit, every maximal
ideal is in the associator of the ring, and the associator of a ring is a finite set).
Given any maximal ideal, m, codimRm = 0, since m is an associated prime of 0.
By Theorem II.4.12, (coker(fk ))m has homological dimension zero, hence is free.
It follows that the truncated complex
fn fk
0 → Fn → Fn−1 → · · · → Fk → Fk−1
Some criteria for exactness 107

is split exact after localization at m (split exact meaning exact with the further
property that im(fi+1 ) is a direct summand of Fi for every i), that the rank of
(coker(fk ))m is equal to

(−1)i−k+1 rankFi
i≥k−1

(a quantity independent of m), and that the projective R-module coker(fk ) has
well-defined rank. Therefore Proposition IV.1.5 applies and I(fk ) = R, as wished.
Now we can prove (1): invert all non-zero divisors, getting I(fk ) = R for all
k, and use Proposition IV.1.6.
As for the proof of (2), take a prime ideal p ⊇ I(fk ) such that codim(Rp ) =
depthI(fk ) (a prime associated to a maximal regular sequence of I(fk )). We
claim that codim(Rp ) ≥ k. To see this, localize F at p: (coker(fk ))p is not free
(by Proposition IV.1.5), hence hdR (coker(f1 ))p is at least k (we cannot shorten
Fp ). Again by Theorem II.4.12, it follows that codim(Rp ) ≥ k, as claimed.
This completes the proof of the “only if” part.
Now we prove the “if” part. We may assume that R is local, with maximal ideal
m (for exactness is a local property, and homology commutes with localization).
Let d indicate the depth of m. For every k > d, condition (2) implies I(fk ) = R.

In particular, coker(fd+1 ) is a free module by Proposition IV.1.5, say Fd+1 . If
 
fd stands for the map Fd → Fd induced by fd , F can be viewed as the result of
glueing together the following complexes:
fd+1
C : 0 → Fn → · · · → Fd+1 → Fd → Fd → 0
and
f fd−1
F : 0 → Fd →
d
Fd−1 → Fd−2 → · · · → F0 ,
both of which have to be proven to be exact.
The exactness of C comes immediately from Proposition IV.1.6.
The exactness of F (which satisfies the same conditions as F, but with length
d equal to the depth of m) comes by induction on dim R.
If dim R = 0, then d = 0, the complex is trivial and we are through. If
dim R > 0, since localization preserves all assumptions, we know by induction
that F turns out to be exact whenever localized at a non-maximal prime ideal
of R. Hence H(F ), the homology of F , is all supported in m, m is an associated
prime of H(F ), and depthH(F ) = 0. The conclusion now results from the
so-called Acyclicity Lemma due to C. Peskine and L. Szpiro, which we quote
below in a slightly weaker form than the original. 2

Proposition IV.1.8 (Acyclicity Lemma) Let


E : 0 → En → En−1 → · · · → Ek → Ek−1 → · · · → E1 → E0
be a complex of finite free modules over the local ring R. If there is some k > 0
such that Hk E = 0 and Hk+t E = 0 for all t > 0, then depthHk E ≥ 1.
108 Structure theorems for finite free resolutions

Proof Cf. Reference [69] (and also [23, lemma 3] and [41, lemma 20.11]).
2
We remark that both Proposition IV.1.6 and Theorem IV.1.7 above lend them-
selves to several generalizations (cf., e.g., [23]). In particular, the following holds
([23], corollary 2).
Proposition IV.1.9 Let M = 0 be an R-module, and let
ϕn ϕ1
A: 0 → Fn → Fn−1 → · · · → F1 → F0
be a complex of finitely generated projective R-modules of well-defined rank. Then
A ⊗R M is exact if and only if for all k = 1, . . . , n
(a) rank(ϕk+1 ⊗R M ) + rank(ϕk ⊗R M ) = rank Fk and
(b) I(ϕk ⊗R M ) contains an M -sequence of length k or I(ϕk ⊗R M ) = R.
We include this last result because it represents a curiosity in relation to an old
question in homological algebra, namely, the “rigidity problem”. Roughly put,
the rigidity question is this: If N and M are two modules of finite homological
dimension, and if TorR R
i (N, M ) = 0 for some given i, is it true that Torj (N, M ) =
0 for all j ≥ i? Positive results have been obtained on this problem when R is
regular local, but it is not true in general. Possibly the most recent results can
be found in Reference [54], where the authors examine the case when the ring
is a hypersurface, that is, R = S/(f ), where S is a regular local ring. In any
event, we mention this here because the above proposition seems to be made for
studying the rigidity problem; the resolution, A, above could be the resolution of
N (truncated at dimension i), in which case the Tor we are looking at is just the
homology of A ⊗R M . However, none of the methods of attack on this problem
has made use of this kind of result.
There is an important corollary of Theorem IV.1.7, which is often considered
as a second, independent, criterion of exactness for F. We have already remarked
that exactness is a local property. Hence F is exact if and only if F ⊗R Rp is
exact for every prime ideal p. The second criterion says that we may limit the
set of prime ideals, p, for which exactness of F ⊗R Rp must be checked.
Theorem IV.1.10 F is exact if and only if F ⊗R Rp is exact for all prime
ideals p such that depth(pRp ) < n.
Proof In order to prove the “if” part, it suffices to show that F satisfies condi-
tions (1) and (2) of Theorem IV.1.7. In fact, if (2) holds, (1) follows automatically,
since the ranks of the maps localize and (1) holds by assumption over every
Rp . Hence we just prove (2), by contradiction. If depthI(fk ) < k for some k
in {1, . . . , n}, then there exists a prime ideal p ⊇ I(fk ) associated to a max-
imal regular sequence contained in I(fk ). Hence rank((fk )p ) = rank(fk ) and
depth(pRp ) = depthI(fk ) < k. It follows that
I((fk )p ) = (I(fk ))p ⊆ pRp
Some criteria for exactness 109

and depthI((fk )p ) < k, contradicting what property (2) says about the exact
complex F ⊗R Rp . 2

We end this section with some properties enjoyed not just by F, but by a larger
family of exact complexes.
Let G denote a (possibly infinite) complex of finite free R-modules that is
exact:
gk g1
· · · → Gk → Gk−1 → · · · → G1 → G0 .

Suppose that depthI(gk ) ≥ 1 for every k ≥ 1. (By Theorem IV.1.7, this condition
is automatically satisfied when F is finite.)

Proposition IV.1.11 With the notation as above, we have:


(a) radI(gk ) ⊆ radI(gk+1 ) for all k ≥ 1.
(b) If rank(g1 ) = rankG0 and depthI(g1 ) = k, then

radI(g1 ) = · · · = radI(gk ).

Proof (a) For x ∈ radI(gk ), let S = {xi }. Then S ∩I(gk ) = ∅ and I(gk )S = RS .
Since by assumption I(gk ) contains a non-zero divisor, rank(gk ) = rank((gk )S )
and I((gk )S ) = (I(gk ))S . Hence coker(gk )S is projective, by Proposition IV.1.5.
Since G exact implies GS exact, the projectivity of coker(gk )S implies that of
coker(gk+1 )S . For the exactness of

0 → im(gk )S → (Gk−1 )S → coker(gk )S → 0

yields the projectivity of im(gk )S , and im(gk )S = coker(gk+1 )S because of the


exactness of

0 → im(gk+1 )S → (Gk )S → coker(gk+1 )S → 0.

Hence RS = I((gk+1 )S ) = (I(gk+1 ))S , again by Proposition IV.1.5, and x ∈


radI(gk+1 ).
(b) Thanks to (a), it suffices to show that radI(gk ) ⊆ radI(g1 ). Assume for
a contradiction that there exists some x ∈ radI(gk ) such that x ∈
/ radI(g1 ). By
localizing at S = {xi }, we may suppose that

I(gk ) = R  I(g1 ).

As I(g1 ) is a proper ideal, coker(g1 ) = 0 (as we have seen in the discussion


after Theorem III.4.2, I(ψ) and ann(coker(ψ)) have the same radical). Since for
every finitely generated R-module M , hdR M ≥ depth(ann(M )), we get

hdR (coker(g1 )) ≥ k,

due to the hypothesis depthI(g1 ) = k.


110 Structure theorems for finite free resolutions

Since I(gk ) = R, coker(gk ) is projective by Proposition IV.1.5. Hence im(gk ) =


ker(gk−1 ) is projective, and coker(g1 ) has a projective resolution stopping at
ker(gk−1 ). That is,
hdR (coker(g1 )) < k,
a contradiction. 2

Remark IV.1.12 The hypothesis depthI(gk ) ≥ 1 for every k ≥ 1 is necessary.


Consider for instance R = K[[X, Y ]]/(XY ), where K is any field and X, Y are
two indeterminates. Let X denote the image of X in R and take the complex
X
R −→ R −→ R/(X) −→ 0.
The kernel of multiplication by X is (Y ), where Y stands for the image of Y in
R, for gX = hXY in K[[X, Y ]] implies g = hY . Hence we have an infinite exact
complex
Y X Y X
· · · −→ R −→ R −→ R −→ R −→ R/(X) −→ 0,
where I(gk ) is either (X) or (Y ), so that the depth assumption fails to be verified.
But (X) and (Y ) are primes, so radI(gk ) ⊆ radI(gk+1 ) fails too.

IV.2 The first structure theorem


In this section we state and discuss the first structure theorem. A sketch of its
proof is in the next section.
Theorem IV.2.1 (First Structure Theorem) Let
fn fk f1
F : 0 → Fn → Fn−1 → · · · → Fk → Fk−1 → · · · → F1 → F0
be an exact sequence of oriented free R-modules, and let rk denote rankfk . Then,
(a) for all k ≥ 1, there exist unique maps ak : R → Λrk Fk−1 such that
1. an = Λrn fn
2. the following diagram is commutative
Λrk fk
Λ r k Fk −→ Λrk Fk−1
a∗k+1 ↓ ↑ ak
R = R
for each k < n
(b) radI(fk ) = radI(ak ) for all k ≥ 2.

Remark IV.2.2 (Oriented free modules) By definition, an oriented free


module is a finite free module F , of rank r say, together with the choice of
a generator of Λr F . As we have seen in Remark I.3.32, this amounts to fixing
(noncanonical) isomorphisms Λr−k F ∗ ∼= Λk F , k = 1, . . . , r.
The first structure theorem 111

Consider now the situation of Theorem IV.2.1. As F is exact, Theorem IV.1.7


says that rankFk = rk+1 + rk . As Fk is oriented, we have fixed isomorphisms
Λrk+1 Fk∗ ∼
= Λrk Fk . If k = n, that is, rn+1 = 0, the isomorphism is R ∼
= Λ r n Fn ,
and part 1. of (a) makes sense provided one thinks of an as the composite map
Λrn fn
R ∼= Λrn Fn → Λrn Fn−1 . When k < n, Λrk+1 Fk∗ ∼ = Λrk Fk makes part 2. of

(a) meaningful: one thinks of the dual map ak+1 : Λrk+1 Fk ∗ → R as the map
Λrk Fk → R.
Definition IV.2.3 The maps ak , k ≥ 1, are called multipliers.
We now present a series of applications of Theorem IV.2.1, the first one being
another (neater) proof of the fact that a regular local ring is factorial.
Theorem IV.2.4 Let R be a regular local ring. Then R is a unique factoriza-
tion domain.

Proof By Proposition II.6.1, it suffices to show that hdR (R/(x, y)) ≤ 2 for all
x and y in R.
By Theorem II.5.3, we know that hdR (R/(x, y)) is finite. Hence R/(x, y) has
a finite free resolution, say:
fn f2 (x,y)
0 → Fn → Fn−1 → · · · → F2 → R2 → R → R/(x, y) → 0.
Now, Theorem IV.2.1 says that the following diagram commutes
(x,y)
R2 −→ R

a2 ↓ ↑ a1
R = R,
hence the ideal ima∗2 can be generated by two elements, say I(a∗2 ) = (x , y  ),
and (x, y) = a1 (x , y  ). That is, (x, y) is equal to (x , y  ) times a non-zero divisor
(every element of a regular local ring is a non-zero divisor). If we prove that
hdR (x , y  ) = 1, it will follow hdR (x, y) = 1, hence hdR (R/(x, y)) ≤ 2 as required.
But Theorem IV.2.1 (b) says radI(f2 ) = radI(a2 ), which implies (recall
Theorem IV.1.7) depthI(a2 ) = depthI(f2 ) ≥ 2. So depthI(a∗2 ) ≥ 2. It easily
follows that x , y  is a regular sequence, and hdR (x , y  ) = 1 (the Koszul complex
is a resolution of R/(x , y  ), of length 2). 2
We now deal with some topics which in fact led to the first structure theorem.

Example IV.2.5 (Hilbert–Burch Theorem) Let a be an ideal which can


be generated by n + 1 elements of R and such that hdR R/a = 2. Then we have
(by Theorem IV.1.7):
f2 f1
0 → Rn → Rn+1 → R → R/a → 0
and we ask whether we can describe a in terms of the maps f2 and f1 .
112 Structure theorems for finite free resolutions

Theorem IV.2.1 says that the following diagram is commutative:


f1
Rn+1 −→ R
↓ a∗2 ↑ a1
R = R.
Moreover, Theorem IV.2.1 implicitly says that for all k ≥ 1, we have
(∗) I(fk ) = I(ak )I(ak+1 )
(this will be shown during the proof of the first structure theorem). As depthI(fk )
≥ 1 for all k ≥ 1 (Theorem IV.1.7), and depthI(fk ) ≥ 1 implies depthI(ak ) ≥ 1
for all k ≥ 2 (Theorem IV.2.1 (b)), we have from (∗) that depthI(a1 ) ≥ 1, that
is, depth(a1 ) ≥ 1 and a1 is a non-zero divisor.
Now, a2 = Λn f2 (Theorem IV.2.1 (a)); if a = (y1 , . . . , yn+1 ) and f2 = (cij ),
1 ≤ i ≤ n + 1, 1 ≤ j ≤ n, explicit computations show that yi = ±a1 ∆i , where
∆i is the maximal minor obtained by erasing the i-th row of (cij ).
In other words, every ideal of homological dimension 1 looks determinantal.
This result goes back to D. Hilbert [48, section 4], in a special case, and Refer-
ence [32] seems to contain the first proof of the general case, although the same
theorem cropped up independently, and with minor variations and motivations,
at about the same time.

The construction seen in the previous example is related to the “lifting prob-
lem” of Grothendieck, which was motivated by the Serre conjectures on the
positivity of intersection multiplicity. This lifting problem is known to be not
solvable in general (there are counterexamples due to M. Hochster [50]).
Statement of the General Lifting Problem: Given a local ring S and a
non-zero divisor, x, let R = S/(x). If M is an R-module, does there exist an
S-module M . such that x is a non-zero divisor on M . and M ./xM.∼ = M?
t . t
If M is free of finite type, say M = R , clearly M = S does the job.
Consider next any finitely generated M with a free resolution G, say:
gk g1
· · · → Gk → Gk−1 → · · · → G1 → G0 .
One can construct a sequence

 : ··· → G
G .k →
g k

G . g1 .
k−1 → · · · → G1 → G0

just by choosing matrices for the maps gk and lifting these matrices. Suppose we
 is a complex. We then
are lucky in our choices of the lifted matrices, so that G
get an exact sequence of complexes:
 →G
0→G  → G →0.
x

Clearly Hi (G) = 0 for positive i (since G is a resolution of M ). Furthermore,


 = 0 because multiplication by x is onto, x belongs to
for every i > 0, Hi (G)
The first structure theorem 113

the radical, and Nakayama’s lemma applies. Hence the higher homology vanishes
and the long exact homology sequence leaves us with the short exact sequence
 →
0 → H0 (G)
x  → H0 (G) → 0,
H0 (G)

implying that x is a non-zero divisor for H0 (G).  Since H0 (G) = M , setting


. 
M = H0 (G) solves the lifting problem.
The question then arises: when are we lucky enough to be able to choose our
lifted matrices in such a way that G be a complex?

If hdR (M ) = 1, G looks like 0 → G.1 → G.0 , certainly a complex.
If hdR (M ) = 2, the construction given in Example IV.2.5 comes in. Take:
f2
S n → S n+1
| |
f2 f1
0 → Rn → Rn+1 → R.
.i in f2 ; let yi = a1 ∆
Lift a1 to a1 and consider ∆ .i and set

f1 = (y1 , . . . , y
n+1 ).

It automatically follows that f1 ◦ f2 = 0, and we are through. Beyond this, that
is, if hdR (M ) > 2, anything can happen.
One would like to understand the arithmetic reasons why G  is not always
a complex. A possible approach is to look for generic pairs (A, B), namely, a
ring A and an exact sequence B of finitely generated free A-modules with the
following property: given any ring A and any exact sequence B of free A -
modules with the same ranks of those of B, there exists a map A → A such that
B = B ⊗A A . M. Hochster [51] has shown that the structure theorems of this
chapter are enough to get a generic pair for resolutions of length 2. But the struc-
ture of resolutions of length 3 is far more complicated (cf., e.g. References [71]
and [85]).
Remark IV.2.6 Theorem IV.2.1 can sometimes provide conditions for the
cokernel of a map to have finite homological dimension. We illustrate this in
the following example.
Let K be a field and S the polynomial ring K[Xij ], 1 ≤ i ≤ n, 1 ≤ j ≤ m,
with m ≥ n. Consider the generic map
(Xij )
g : S m −→ S n .
As shown in Section III.4, depthI(g) = m − n + 1 = hdR coker(g).
Also let Ip+1 (g) denote the ideal generated by the (p + 1)-minors of the matrix
(Xij ), and set R = S/Ip+1 (g). The generic map of rank p is the induced map
(X ij )
f : Rm → Rn ,
114 Structure theorems for finite free resolutions

where X ij denotes the image of Xij in R. (Notice that p must be less than n.)
We are going to prove that coker(f ) cannot have finite homological dimension.
If hdR coker(f ) were finite, then coker(f ) would have a finite free resolution
(by Quillen-Suslin); indeed a resolution by free R-graded modules, because R
inherits from S the structure of a graded ring (Ip+1 (g) is a graded ideal, g is a
homogeneous map) and coker(f ) is a graded R-module. Theorem IV.2.1 would
thus give the commutative diagram:
Λp f
Λp Rm −→ Λp Rn
↓ a∗2 ↑ a1
R = R.

In particular, every p-minor occurring in the first p columns of (X ij ) would


belong to the (proper) principal ideal of R generated by a∗2 (e1 ∧ · · · ∧ ep ), where
{e1 , . . . , em } denotes the canonical basis of Rm . As the p-minors of g have degree
p in S, while Ip+1 (g) is generated by elements of degree p+1, it would then follow
that all p-minors in the first p columns of (Xij ) would be contained in a proper
principal ideal of S. However these same p-minors of (Xij ) would generate the
ideal of p-minors of a generic n×p matrix, that is, an ideal of depth n−p+1 > 1:
a contradiction.
We end this section with an example showing that Theorem IV.2.1 may not
hold for infinite complexes.

Example IV.2.7 Let K be a field, X and Y indeterminates over K, and R =


K[[X, Y ]]/(Y 2 − X 3 ) (cusp at the origin). Write X and Y for the images of X
and Y in R. Then we have the following infinite resolution:
f4 f3 f2 (X,Y )
· · · → R2 → R2 → R2 → R → R/(X, Y ),
   
2 2
Y X Y X
where every fk , k ≥ 2, has matrix . As det = 0,
−X −Y −X −Y
rank(fk ) = 1 for all k ≥ 2.
If Theorem IV.2.1 held in this case, we should have commutative diagrams
f2 (X,Y )
R2 −→ R2 R2 −→ R
↓ a∗3 ↑ a2 and ↓ a∗2 ↑ a1 ,
R = R R = R
   
2
Y X X
with f2 = . As a1 = 1, a∗2 = (X, Y ) and a2 = . Hence
−X −Y Y
   
1 X
(a2 ◦ a∗3 ) = a2 (r) = ra2 (1) = r ,
0 Y
Proof of the first structure theorem 115

 
1
where r stands for the constant a∗3 . Thus
0
 
2    
Y X 1 Y
=
−X −Y 0 −X
   
X Y
would give r = , from which Y = rX, a contradiction.
Y −X

IV.3 Proof of the first structure theorem


IV.3.1 Part (a)
We first show how one defines an−1 starting from an (assume n ≥ 2). We have
seen in Section III.4 that given f : F → G with rankF = m ≥ n = rankG,
one can construct a family of depth-sensitive complexes T(q; f ). When q = n we
have:
ψm−n+1
T(n; f ) :0 → Dm−n G∗ ⊗ Λm F → Dm−n−1 G∗ ⊗ Λm−1 F →
ψk+1
· · · → Dk G∗ ⊗ Λn+k F → Dk−1 G∗ ⊗ Λn+k−1 F →

· · · → G∗ ⊗ Λn+1 F → Λn F → Λn G ∼
ψ2 ψ1
= R,
where ψ1 = Λn f and, for every k > 0, ψk+1 is the composite map

Dk G∗ ⊗ Λn+k F −→ Dk−1 G∗ ⊗ G∗ ⊗ Λn+k F −→ Dk−1 G∗ ⊗ Λn+k−1 F


∆⊗1 1⊗α

(α stands for the action of ΛG∗ on ΛF induced by f ∗ : G∗ → F ∗ ). Let us suppose


that f coincides with the dual of the left-most non-zero map of our complex F,
that is, f is fn∗ : Fn−1

→ Fn∗ . Then T(n; f ) looks like
∗ ψ3 ∗ ψ2
∗ Λrn f ∗
· · · → D2 Fn ⊗ Λrn +2 Fn−1 → Fn ⊗ Λrn +1 Fn−1 → Λrn Fn−1 → n Λrn Fn∗ ,
and the following truncation of its dual complex
Λrn fn ψ∗
0 → Λrn Fn → Λrn Fn−1 →2 Fn∗ ⊗ Λrn +1 Fn−1
must be exact, because depth(fn ) ≥ n ≥ 2 (Theorem IV.1.7). Notice that ψ2∗
is just multiplication by the trace of fn , that is, the element of Fn∗ ⊗ Fn−1
corresponding to fn via the canonical isomorphism
HomR (Fn , Fn−1 ) ∼
= Fn∗ ⊗ Fn−1 .
Now we make a general observation. Given
f
F → G → L → 0 exact,
we have
F ⊗ Λt−1 G → Λt G → Λt L → 0 exact,
116 Structure theorems for finite free resolutions

where the left-most arrow is the composite map


f ⊗1 m
F ⊗ Λt−1 G → G ⊗ Λt−1 G → Λt G.
It follows that if the composite F → G → H is zero, then also the composite
m◦(f ⊗1)
F ⊗ Λt−1 G → Λt G → Λt H
is zero, since the right-most map factors through Λt L:
Λt G  Λt L → Λt H.
Thanks to the previous observation, it is immediate to see that if βn is the
composite map
∗ ∗ f ∗ ⊗1
⊗ Λrn−1 −1 Fn−1

→ Fn∗ ⊗ Λrn−1 −1 Fn−1


Λrn−1 Fn−1 → Fn−1 n
,
then the following composite

∗ Λrn−1 fn−1 ∗ βn
Λrn−1 Fn−2 → Λrn−1 Fn−1 → Fn∗ ⊗ Λrn−1 −1 Fn−1

is zero.
Some calculations then show that the following diagram is commutative:

Λrn−1 fn−1 βn
∗ ∗
Λrn−1 Fn−2 → Λrn−1 Fn−1 → Fn∗ ⊗ Λrn−1 −1 Fn−1

 i1  1 ⊗ i2
Λrn fn ψ∗
0→ R∼
= Λ r n Fn → Λrn Fn−1 →2 Fn∗ ⊗ Λrn +1 Fn−1 ,
where i1 and i2 are the appropriate isomorphisms given by the orientations (recall
Remark IV.2.2).
It follows that there exists a unique map

Λrn−1 Fn−2 → Λ r n Fn ∼
= R,
whose dual (clearly verifying property 2) we define to be an−1 .
Suppose now that we have ak+1 and ak ; we want to define ak−1 (assume k ≥ 2).
As before, we see that the composite map

∗ Λrk−1 fk−1 ∗ βk
Λrk−1 Fk−2 −→ Λrk−1 Fk−1 −→ Fk∗ ⊗ Λrk−1 −1 Fk−1

is zero.
Consider the sequence
a γ
0→R→
k
Λrk Fk−1 → Fk∗ ⊗ Λrk +1 Fk−1 , (IV.3.1)
where γ is multiplication by the trace of fk , that is, the element of Fk∗ ⊗ Fk−1
corresponding to fk via the canonical isomorphism
HomR (Fk , Fk−1 ) ∼
= Fk∗ ⊗ Fk−1 .
Proof of the first structure theorem 117

Again we have a commutative diagram



Λrk−1 fk−1 βk
∗ ∗
Λrk−1 Fk−2 → Λrk−1 Fk−1 → Fk∗ ⊗ Λrk−1 −1 Fk−1

 
a γ
0→ R →
k
Λrk Fk−1 → Fk∗ ⊗ Λrk +1 Fk−1 ,
so that it suffices to show that the bottom row is exact, to be able to conclude
as we did in the case of an−1 .
In order to prove that (IV.3.1) is a complex, we first show that the composite
map
Λrk fk γ
Λrk Fk → Λrk Fk−1 → Fk∗ ⊗ Λrk +1 Fk−1 (IV.3.2)
is zero. Let {x1 , . . . , xt } be a basis of Fk and {ξ1 , . . . , ξt } the 
basis of dual to Fk∗
{x1 , . . . , xt }. Then the trace element of fk can be written as i ξi ⊗ fk (xi ), and
for every y1 ∧ · · · ∧ yrk ∈ Λrk Fk it turns out that
(γ ◦ Λrk fk )(y1 ∧ · · · ∧ yrk ) = γ(fk (y1 ) ∧ · · · ∧ fk (yrk ))

= ξi ⊗ fk (xi ) ∧ fk (y1 ) ∧ · · · ∧ fk (yrk )
i

= ξi ⊗ (Λrk +1 fk )(xi ∧ y1 ∧ · · · ∧ yrk ),
i
rk +1
which vanishes because Λ fk = 0.
Next we prove that I(fk ) = I(ak )I(ak+1 ) (this is the equality announced and
used in Example IV.2.5). As the inclusion ⊇ is straightforward, we just prove ⊆.
The three formulas

(Λrk fk )(xi1 ∧ · · · ∧ xirk ) = ±(j1 , . . . , jrk | i1 , . . . , irk )wj1 ∧ · · · ∧ wjrk
(where {w1 , . . . , ws } stands for a basis of Fk−1 and the coefficients occurring in
the sum are k × k minors of fk ),

ak+1 (1) = ci1 ···irk xi1 ∧ · · · ∧ xirk
and

ak (1) = dj1 ···jrk wj1 ∧ · · · ∧ wjrk ,
inserted into the diagram (commutative by hypothesis)
Λrk fk
Λrk Fk −→ Λrk Fk−1
a∗k+1 ↓ ↑ ak (IV.3.3)
R = R,
yield the following diagram
Λrk fk 
xi1 ∧ · · · ∧ xirk −→ ±(j1 , . . . , jrk | i1 , . . . , irk )wj1 ∧ · · · ∧ wjrk
↓a∗k+1 
a 
ci1 ···irk →
k
ci1 ···irk dj1 ···jrk wj1 ∧ · · · ∧ wjrk ,
and we are done.
118 Structure theorems for finite free resolutions

Since I(ak+1 ) ⊇ I(fk ) implies depthI(ak+1 ) ≥ 2, I(ak+1 ) contains a non-zero


divisor. Take the zero composite map (IV.3.2), where Λrk fk factors as indicated
by (IV.3.3). It follows that

0 = im(γ ◦ Λrk fk ) = I(ak+1 )im(γ ◦ ak ),

and since I(ak+1 ) contains a non-zero divisor and im(γ ◦ ak ) is included in a free
module, one necessarily has im(γ ◦ ak ) = 0, that is, the sequence (IV.3.1) is a
complex.
It remains to show that (IV.3.1) is exact. We resort to Theorem IV.1.10 and
localize at primes of depth at most 1. Since k ≥ 2 and depthI(fk ) ≥ 2, I(fk )
blows up after localization (i.e., it becomes the whole ring) and we may assume

R = I(fk ) = I(ak ) = I(ak+1 ) (IV.3.4)

(recall that I(fk ) = I(ak )I(ak+1 )). It is now obvious that ak is a monomorphism.
To show exactness at Λrk Fk−1 , it suffices to prove that

rank(ak ) + rank(γ) = rankΛrk Fk−1 (IV.3.5)

(by Theorem IV.1.7; the depth condition is satisfied because of (IV.3.4)). (IV.3.5)
is equivalent to

rank(γ) = rankΛrk Fk−1 − 1.

To verify this latter equality, it is enough to show that (IV.3.2) is exact and
use Theorem IV.1.7, since rank(Λrk fk ) = 1. But the exactness of (IV.3.2) is
now obtained by noting that I(fk ) = R implies coker(fk ) projective (Pro-
position IV.1.5), so that Fk = ker(fk ) ⊕ im(fk ), Fk−1 = im(fk ) ⊕ coker(fk )
and
0
fk = idim(fk ) ⊕ {ker(fk ) → coker(fk )}.

For the well-known property Λ(A⊕B) = ΛA⊗ΛB indicates that (IV.3.2) is made
up of strands involving ker(fk ), im(fk ) and coker(fk ), strands whose exactness
is easy to ascertain.
This completes the proof of Part (a) of Theorem IV.2.1.

IV.3.2 Part (b)


Since the equality I(fk ) = I(ak )I(ak+1 ) proved above implies that I(fk ) ⊆ I(ak ),
it suffices to show that I(ak ) ⊆ radI(fk ), for all k ≥ 2.
Suppose the contrary, namely, that I(ak )  radI(fk ). Then there exists a
minimal prime, p, associated to I(fk ) such that p  I(ak ). By localizing at p,
we may thus assume that R is local with maximal ideal m, that I(fk ) is primary
to m and that I(ak ) = R. In particular, I(fk ) = I(ak+1 ) (again by the fact that
I(fk ) = I(ak )I(ak+1 )). 
If we write ak (1) = dj1 ···jrk wj1 ∧ · · · ∧ wjrk as in the proof of Part (a),
then (dj1 ···jrk ) = I(ak ) = R. Without loss of generality, we may suppose that
The second structure theorem 119

d1···rk = 1 and let F denote the free submodule of Fk−1 generated by w1 , . . . , wrk .
We take the composite map
fk π
Fk → Fk−1 → F,
where π stands for the projection, and look at the complex
fn fk+1 π◦fk
0 → Fn → Fn−1 → · · · → Fk+1 → Fk → F. (IV.3.6)
By construction, rank(π ◦ fk ) = rankF = rk = rank(fk ) and I(π ◦ fk ) = I(fk ) =
I(ak+1 ), so that using Theorem IV.1.7 we obtain that (IV.3.6) is exact.
The exactness of (IV.3.6) shows that coker(π ◦ fk ), which is not equal to 0,
has finite homological dimension. Since I(fk ) = R (so that neither coker(fk ) nor
coker(π ◦ fk ) is projective),
hdR coker(π ◦ fk ) = hdR coker(f1 ) − k + 1;
hence k ≥ 2 implies hdR coker(π ◦ fk )  hdR coker(f1 ).
As coker(π ◦ fk ) has finite homological dimension, Theorem II.4.12 applies and
hdR coker(π ◦ fk )  codimR.
On the other hand, since rank(π ◦ fk ) = rankF , the discussion after Theorem
III.4.2 holds for π ◦ fk and
I(π ◦ fk ) ⊆ ann(coker(π ◦ fk )).
As I(π◦fk ) = I(fk ), this means that ann(coker(π◦fk )) contains an ideal primary
to m, codim(coker(π ◦ fk )) = 0 and (again by Theorem II.4.12)
hdR coker(π ◦ fk ) = codimR,
a contradiction.
This finishes the proof of Theorem IV.2.1.

IV.4 The second structure theorem


In Example IV.2.5 we considered an ideal a generated by n + 1 elements of R
and such that hdR R/a = 2. We took the resolution
f2 f1
0 → Rn → Rn+1 → R → R/a → 0
and were able to use the multipliers to describe f1 in terms of f2 (which had
positive implications for the lifting problem).
Suppose now that R is a local ring, that the ideal a is generated by three
elements, and hdR R/a = 3. Then R/a has a resolution
f3 f2 f1
0 → Rr → Rr+2 → R3 → R, (IV.4.1)
where rank(f1 ) = 1, rank(f2 ) = 2 and r denotes rank(f3 ). It turns out that the
multipliers are not enough to describe all of (IV.4.1) in terms of f3 . However
120 Structure theorems for finite free resolutions

there is something we can do in this case also, to get further information about
these maps.
Let C = (cij ) be the (r + 2) × r matrix of f3 . We will describe below an
(r + 2) × 3 matrix, D = (dis ), whose entries, along with those of C, we use to
describe the matrices of f2 and f1 . More precisely, given the (r + 2) × (r + 3)
matrix (C|D) = (cij |dis ), the description of matrices for f2 and f1 , up to signs,
and multiplied by non-zero divisors, goes as follows.
Description of the 3 × (r + 2) matrix for f2 : The (s, t)-coordinate is the
(r + 1) × (r + 1) minor of (C|D) obtained by erasing the row indexed by t and
the columns indexed by the two elements of the set of indices{r + 1, r + 2, r + 3}
which are not equal to r + s.
Description of the 1×3 matrix for f1 : The (1, s)-coordinate is the (r + 2)×
(r + 2) minor of (C|D) obtained by erasing the column indexed by r + s.
The question is: How does one get the matrix D?
We have seen in Section III.4 that given f : F → G with rankF = m ≥
n = rankG, one can construct a family of depth-sensitive complexes T(q; f ). Set
q = 1 and f equal to the map f3∗ : (Rr+2 )∗ → (Rr )∗ . Then T(1; f3∗ ) looks like:
ψ3 ψ2 f∗
T(1; f3∗ ) : 0 → Rr ⊗ Λr+2 (Rr+2 )∗ → Λr+1 (Rr+2 )∗ → (Rr+2 )∗ →
3
(Rr )∗ ,

where ψ3 is the action of Rr on Λr+2 (Rr+2 )∗ induced by the map f3 , and ψ2 is


given by the action of Λr (Rr )∗ on Λr+1 (Rr+2 )∗ (again induced by the map f3 ).
By depth-sensitivity, the dual of T(1; f3∗ ) then provides the following exact
sequence:
f3
0 → Rr → Rr+2 → Λr+1 (Rr+2 ) → (Rr )∗ ⊗ Λr+2 (Rr+2 ), (IV.4.2)

and we have the commutative diagram:


0 → Rr → Rr+2 → Λr+1 (Rr+2 ) → (Rr )∗ ⊗ Λr+2 (Rr+2 )
 
f∗ f∗
(R3 )∗ →
2
(Rr+2 )∗ 3
→ (Rr )∗ .

Because of the exactness of (IV.4.2), it is possible to find a (not necessarily


unique) vertical map which yields the following commutative diagram:
0 → Rr → Rr+2 → Λr+1 (Rr+2 ) → (Rr )∗ ⊗ Λr+2 (Rr+2 )
↑  
f∗ f∗
(R3 )∗ →
2
(Rr+2 )∗ 3
→ (Rr )∗ .
This vertical map is the map whose matrix we have called D.
It is a relatively straightforward calculation to show that the descriptions of
the matrices of f2 and f1 in terms of this D are valid.
We take our cue from the foregoing discussion, and show how maps, similar
to the map D above, may be constructed in general.
The second structure theorem 121

Let R be a commutative ring and


fn fk f1
F : 0 → Fn → Fn−1 → · · · → Fk → Fk−1 → · · · → F1 → F0 ,

n ≥ 3, be an exact complex, with each Fk an oriented free R-module and rk =


rank(fk ). The multipliers ak : R → Λrk Fk−1 of Theorem IV.2.1 give information
about the rk × rk minors. But in (IV.4.1) above, we were looking for information
about the submaximal minors: f3 had rank 2. Take therefore the composite map
1⊗a m
a.
n : Fn−1 = Fn−1 ⊗ R → Fn−1 ⊗ Λ Fn−1 → Λ
n rn rn +1
Fn−1 ,

and consider the following complex (where γ is multiplication by the trace of fn ):


fn 
a γ
0 → Fn → Fn−1 →
n
Λrn +1 Fn−1 → Fn∗ ⊗ Λrn +2 Fn−1 ,

which is exact for n ≥ 3 (n = depthI(fn )). Notice that a.


n is precisely the map
Rr+2 → Λr+1 (Rr+2 ) of (IV.4.2). Again make identifications and obtain
fn 
a γ
0 → Fn → Fn−1 −→
n
Λrn +1 Fn−1 → Fn∗ ⊗ Λrn +2 Fn−1
  
Λrn−1 −1 fn−1

Λrn−1 −1 Fn−2

−→ Λrn−1 −1 Fn−1

→ Fn∗ ⊗ Λrn−1 −2 Fn−1

,

where the bottom row is a complex (as in βn ◦ Λrn −1 fn−1 ∗


= 0 of Subsec-
tion IV.3.1). It follows that there exists a map Λrn−1 −1 Fn−2∗
→ Fn−1 whose
dual we call bn−1 (b∗n−1 is like the vertical map (R3 )∗ → Rr+2 found above: just
note that (R3 )∗ is isomorphic to Λ2 (R3 )).

How can we go from bn−1 : Fn−1 → Λrn−1 −1 Fn−2 to further maps bk : Fk∗ →
rk −1
Λ Fk−1 that satisfy similar conditions?
Let us see these conditions first. Suppose that we mimic case n − 1 : we take the
composite map
1⊗ak+1 m
a
k+1 : Fk = Fk ⊗ R → Fk ⊗ Λrk+1 Fk → Λrk+1 +1 Fk ,

and consider the following (where γ is multiplication by the trace of fk+1 ):


fk+1 a ∗ γ
Fk+1 → Fk → Λrk+1 +1 Fk → Fk+1 ⊗ Λrk+1 +2 Fk .
k+1
(IV.4.3)

If, and this is a big if, (IV.4.3) were exact, we could make the usual identifications
and get b∗k as the vertical arrow induced by the commutative box of the following
diagram:
fk+1 a ∗ γ
Fk+1 → −→ Λrk+1 +1 Fk → Fk+1 ⊗ Λrk+1 +2 Fk
k+1
Fk
↑   
Λrk −1 f ∗
Λrk −1 Fk−1

−→ k Λrk −1 Fk∗ → Fk+1

⊗ Λrk −2 Fk∗ .
122 Structure theorems for finite free resolutions

Hence the conditions would be: the diagrams



a
Λrk+1 +1 Fk∗ −→ Fk∗
k+1

 ↓ bk
Λrk −1 fk
Λrk −1 Fk −→ Λrk −1 Fk−1

are commutative (assume k ≥ 2).


We thus want to show the exactness of (IV.4.3). Take:
fn fk+1 a ∗ γ
0 → Fn → · · · → Fk −→ Λrk+1 +1 Fk → Fk+1 ⊗ Λrk+1 +2 Fk .
k+1
(IV.4.4)

If we prove that (IV.4.4) is exact, it will follow that (IV.4.3) is exact too (here
is a case in which it is easier to deal with a longer sequence). In order to show
the exactness of (IV.4.4), we can first prove that it is a complex and then apply
Theorem IV.1.7.
That γ ◦ ak+1 = 0 and a  k+1 ◦ fk+1 = 0 follows in the same way that im(γ ◦
ak+1 ) = 0 was proved in Subsection IV.3.1. The rank conditions can be checked
by localizing at S = {xi }, where x is a non-zero divisor in I(fk+1 ), and proceeding
as at the end of Subsection IV.3.1. As for the depth conditions, since k ≥ 2,
depthI(ft ) is all right for all t ≥ k; depthI(
ak+1 ) ≥ 2 we know from the properties
of ak+1 ; depthI(γ) ≥ 1 follows from Lemma IV.4.1 below.

Lemma IV.4.1 radI(γ) ⊇ radI(fk+1 ).

Proof Assume for a contradiction that there exists x ∈ / radI(γ) such that
xt ∈ I(fk+1 ) for some t. Localize at S = {xi }. As S ∩ I(fk+1 ) = ∅ = S ∩ I(γ),
I(fk+1 ) blows up and I(γ) does not. Hence a contradiction will follow, if one
proves that I(fk+1 ) = R implies I(γ) = R.
By Proposition IV.1.5, I(fk+1 ) = R implies that coker(fk+1 ) is projective;
thus we have Fk+1 = ker(fk+1 ) ⊕ im(fk+1 ), Fk = im(fk+1 ) ⊕ coker(fk+1 ) and
 
0
fk+1 = idim(fk+1 ) ⊕ ker(fk+1 ) → coker(fk+1 ) .

From the well known property: Λ(A ⊕ B) = ΛA ⊗ ΛB, it follows that the
right-most map of (IV.4.4) is made up of strands involving ker(fk ), im(fk )
and coker(fk ), strands whose careful study shows coker(γ ∗ ) = (coker(fk+1 ))∗ ⊗
Λrk+1 (im(fk+1 ))∗ . Hence coker(γ ∗ ) is projective and Proposition IV.1.5 yields
I(γ ∗ ) = R. 2

This whole preceding discussion amounts to a sketch of the proof of what is


called the second structure theorem.

Theorem IV.4.2 (Second Structure Theorem) Let


fn fk f1
F : 0 → Fn → Fn−1 → · · · → Fk → Fk−1 → · · · → F1 → F0 ,
The second structure theorem 123

n ≥ 3, be an exact sequence of oriented free R-modules. Let rk denote rankfk ,


and let
.k : Fk−1 → Λrk +1 Fk−1
a
be the composite
1⊗a m
Fk−1 = Fk−1 ⊗ R →k Fk−1 ⊗ Λrk Fk−1 → Λrk +1 Fk−1 ,
where ak is the multiplier of Theorem IV.2.1. Then for every k = 2, . . . , n − 1,
there exists a map
bk : Fk∗ → Λrk −1 Fk−1
(not unique) such that the following diagram commutes:

a
Λrk+1 +1 Fk∗ −→ Fk∗
k+1

 ↓ bk (IV.4.5)
rk −1 Λrk −1 fk rk −1
Λ Fk −→ Λ Fk−1 .
Remark IV.4.3 A map bn : Fn∗ → Λrn −1 Fn−1 can be trivially added to the
statement of the second structure theorem. Simply define it as the composite
Λrn −1 fn
F∗ ∼
n = Λrn −1 Fn −→ Λrn −1 Fn−1 .
Remark IV.4.4 Although Theorems IV.2.1 and IV.4.2 give such good inform-
ation about the maximal and submaximal minors of every fk , it is very hard to
guarantee the existence of a sequence F with prescribed maps ak and bk .
We have just said in the previous remark that it is hard to recover F from the
maps ak and bk . As a matter of fact, given a concrete resolution F, it may not
be easy to find its maps bk either. This motivates the following.
Notice that the commutativity of the diagram in part (a) 2. in the statement
of Theorem IV.2.1 amounts to
ak (a∗k+1 (z(ε))) = Λrk fk (z) (IV.4.6)
for all z ∈ Λrk Fk . Here ε ∈ Λrk+1 +rk Fk∗ denotes the orientation of Fk , that is, the
chosen isomorphism in HomR (Λrk+1 +rk Fk , R) ∼ = Λrk+1 +rk Fk∗ (cf. Remark I.3.32),
rk+1 ∗
and z(ε) ∈ Λ Fk is the contraction of z on ε.
Notice that the commutativity of the diagram (IV.4.5) in the statement of
Theorem IV.4.2 amounts to
b∗k (x)(ak+1 (1)(ε)) = Λrk −1 fk∗ (x)
for all x ∈ Λrk −1 Fk−1

.
Notice also that it is possible to define maps c∗k : Λrk −1 Fk−1

→ Fk such that

fk (ck (x)) = x(ak (1)), because
fk−1 (x(ak (1))) = 0 for all x ∈ Λrk −1 Fk−1

, (IV.4.7)
124 Structure theorems for finite free resolutions

so that one wonders whether the relatively easier maps ck can replace the
maps bk .
[Proof of (IV.4.7). It suffices to show y · fk−1 (x(ak (1))) = 0 for some non-zero
divisor y ∈ R. Since we have already seen in the proof of Theorem IV.2.1, Part
(a), that I(a∗k+1 ) contains a non-zero divisor, we may choose y to be a∗k+1 (z(ε))
for some suitable z ∈ Λrk Fk . We then have
y · fk−1 (x(ak (1))) = fk−1 (x(ak (y))) = fk−1 (x(ak (a∗k+1 (z(ε))))),
that is (recall (IV.4.6)),
fk−1 (x(Λrk fk (z))),
which is zero, since x(Λrk fk (z)) ∈ Fk−1 belongs in fact to im(fk ).]
Proposition IV.4.5 For the maps bk of the second structure theorem, the
following are equivalent:
(1) b∗k (x)(ak+1 (1)(ε)) = Λrk −1 fk∗ (x) for all x ∈ Λrk −1 Fk−1

(2) fk (b∗k (x)) = x(ak (1)) for all x ∈ Λrk −1 Fk−1



.
Proposition IV.4.5 says that the maps ck are indeed another way of defining
the maps bk . It also coincides with case  = 1 of the following stronger result.
Theorem IV.4.6 Let
fn fk f1
F : 0 → Fn → Fn−1 → · · · → Fk → Fk−1 → · · · → F1 → F0
be an exact sequence, let the maps ak be the multipliers of F, and let
bk : Λ Fk∗ → Λrk − Fk−1
be a map. Then bk satisfies
rk − ∗
b∗
k (x)(ak+1 (1)(ε)) = Λ fk (x) for all x ∈ Λrk − Fk−1

(IV.4.8)
if and only if it satisfies
Λ fk (b∗
k (x)) = x(ak (1)) for all x ∈ Λrk − Fk−1

. (IV.4.9)
We do not give the proof, but refer the reader to Reference [20]. We point out
that at one point it was hoped maps bk might help in the study of lower order
minors, so it was encouraging to have description (IV.4.9), which is relatively
easier than (IV.4.8) But it was soon discovered that maps bk do not always
exist [17].
In a way, all of the foregoing chapter may be looked at as the first steps of
the “homological baby” coming to grips with the outside world of commutative
algebra. When one gets right down to the core, it seems obvious, almost naive,
to ask: What makes a complex exact? For vector spaces, the answer is so obvious
that it was never even asked; for arbitrary commutative ground rings, this prob-
lem had to be confronted. While the structure theorems give a promising start
The second structure theorem 125

to understanding some basic questions, we see that they still leave a great many
questions open. Even where the answers to some of our questions are negative,
the counterexamples do not yield the insight we need in order to fully understand
the phenomena that we want to study.
To point to an even more basic, and hence irksome, question, consider the fact
that a module, M , is completely determined by its presentation (which, for the
sake of retaining our sanity, we will assume to be finite): F1 → F0 → M → 0.
Presumably, then, it should be possible to read off all the information we might
want to have about M from the presentation matrix. For example, what property
does the matrix have to have in order that hdR M be finite? Notice that all the
(limited) information we have obtained from the structure theorems comes from
the assumption that the module is known to have finite homological dimension;
using that, we have managed to drag information about the matrix at the tail
of the resolution to the head (i.e., the presentation matrix). We know that the
answer cannot be given solely in terms of the minors of the matrix, though, as the
cokernel of the transpose matrix may have characteristics completely different
from the original.
So, to be uncompromisingly candid, our baby steps do not begin to address
some of the most elementary yet fundamental problems of commutative algebra.
While this may be taken by some to be disheartening, to say the least, we
can (and the authors do) take the view that there is a lot of growing yet to
do, and perhaps the recipe resides somewhere in the willingness to accept that
our accomplishments are as yet meager, but we should not lose sight of the
fundamental questions.
This page intentionally left blank
V
EXACTNESS CRITERIA AT WORK

Since the time that the structure theorems of the last chapter appeared, there
have been any number of applications and refinements of these results. What we
have tried to do is to select some area in which these theorems have played an
essential role, and in which there are still lurking some fascinating, only partially
explored areas of investigation.
We have chosen to look at resolutions of Gorenstein ideals of depth three
in a local ring, and their powers. (Recall [Definition II.4.10] that the depth
of an ideal, also called its grade, is the length of a maximal R-sequence in
I.) A Gorenstein ideal is a perfect ideal, I, of grade g, with the property
that TorR ∼
g (R/I, k) = R, where k is the residue field of the local ring, R. For
example, any complete intersection (i.e. an ideal generated by an R-sequence)
is a Gorenstein ideal, as the Koszul complex associated to any minimal set
of its generators is a minimal resolution and the last term of that complex is
precisely R.
We will look at Gorenstein ideals of depth three for a number of reasons. First,
in much the same spirit of the Hilbert–Burch Theorem (cf. Example IV.2.5), it is
possible to describe such ideals in terms of the minors of a certain kind of matrix,
namely an alternating one. We will show that all of these ideals are generated
by the so-called Pfaffians of order n − 1 of an alternating matrix of odd order,
n. Another reason is that for the first time in this book, the algebra structure
of such a resolution comes into play. It still is not completely clear just how
large a role the algebra structures of finite free resolutions must play in their
study, but these examples, as well as the vast amount of work on Poincaré series,
would indicate that such structure should not be ignored. The development in
the first section, where we deal exclusively with the Gorenstein ideal of grade
three follows to a large extent the treatment in Reference [25].
In Section V.2, we build on the previous one and construct resolutions for
all powers of the ideals in Section V.1. While this entails a bit of additional
technique, the results exhibit a phenomenon which seems to occur with some
frequency: a growth of homological dimension up to a critical point, and then
a constant value from a certain point on. This class of ideals also provides us
with the opportunity to illustrate the use of a counting method from represent-
ation theory to provide us with an idea of what the resolutions of such ideals
may look like. An elaboration of the material of Section V.2 can be found in
Reference [13].
128 Exactness criteria at work

V.1 Pfaffian ideals


Before we begin with our Gorenstein ideals of depth three, we will dispose easily
of those of lower depth, namely depths one and two.
If I is Gorenstein of depth one, then it has a resolution of the type

0 → R → R → R/I → 0.

Obviously, I is a principal ideal generated by a non-zero divisor.


If I is Gorenstein of depth two, it has a resolution of the type

0 → R → R2 → R → R/I → 0.

The condition that the last term of the minimal resolution be free of rank one
(the condition that we placed on Tor) does not provide much wiggle room for the
resolution. But now our Hilbert–Burch theorem tells us that, since the depth of
I is two, I is generated by a regular sequence of length two, and our resolution
is essentially the Koszul complex associated to those generators.
Thus, we see that Gorenstein ideals of depth one or two must be generated by
regular sequences. This is why the real investigation of Gorenstein ideals has to
begin with depth three. Of course, one might be led to suspect that Gorenstein
ideals are all complete intersections; what follows should disabuse the reader of
that idea.

V.1.1 Pfaffians
Let R be a noetherian commutative ring, and F a finitely generated free
R-module. A morphism f : F ∗ → F is called alternating if, with respect
to some (and therefore every) basis and dual basis of F and F ∗ , the matrix
W = (wij ) of f satisfies wij + wji = 0 whenever i = j, and wii = 0 for every i.
Using the isomorphism

HomR (F ∗ , F ) ∼
= (F ∗ )∗ ⊗ F ∼
= F ⊗ F,

f corresponds to an element ϕ ∈ F ⊗F . In fact ϕ ∈ Λ2 F , where Λ2 F is embedded


into F ⊗ F by means of the diagonal map.
Since ϕ is homogeneous of degree 2 in ΛF , there exist divided powers

ϕ(0) , ϕ(1) , ϕ(2) , . . .

kthe following sense. Let a ∈ ΛF be homogeneous of even degree, say a =


in
i=1 ai , where each ai is a product of degree 1 elements; then, for every p ≥ 0,
set

a(p) = ai1 ∧ · · · ∧ aip .
1≤i1 <···<ip ≤k
Pfaffian ideals 129

The sequence of elements a(0) , a(1) , a(2) , . . . (called a system of divided


powers) satisfies the following properties:
(1) a(0) = 1, a(1) = a, deg(a(p) ) = p deg(a)
  (p+q)
(2) a(p) a(q) = p+q
q a
(pq)! (pq)
(3) (a(p) )(q) = q!pq ! a .
A fuller description of these divided powers can be found in Reference [25]
(p. 482).
Lemma V.1.1 For every a ∈ Λt F ∗ , we have:

Λt f (a) = (−1)i ϕ(i) (a)(ϕ(t−i) ). (V.1.1)
i≥0

Proof Use induction on t and the formula in Lemma III.3.2 (cf. Reference [25],
pp. 458–460). 2
Corollary V.1.2 If rankF = 2k, then using the isomorphism
HomR (Λ2k F ∗ , Λ2k F ) ∼
= Λ2k F ⊗ Λ2k F,
the element Λ2k f corresponds to ϕ(k) ⊗ ϕ(k) .
Proof Set t = 2k in (V.1.1). Then the only non-zero summand is
 
(−1)k ϕ(k) (a) ϕ(k) ,
/ 0
since rankF = 2k. But ϕ(k) (a) = (−1)k ϕ(k) , a , where −, − stands for the
action of (ΛF ∗ )∗ on ΛF ∗ and we identify ϕ(k) ∈ Λ2k F with its image under the
canonical map
Λ2k F → Λ2k (F ∗ )∗ → (Λ2k F ∗ )∗ .
/ 0
It follows that Λ2k f (a) = ϕ(k) , a ϕ(k) for every generator a of Λ2k F ∗ . 2
Corollary V.1.2 has the following meaning. Choose an orientation of F , that
is, an isomorphism Λ2k F ∼ = R. Then Λ2k f corresponds to the determinant of the
appropriate matrix of f , and such a determinant is the square of the element
of R corresponding to ϕ(k) , called the pfaffian of f . We write Pf(f ), and also
Pf(W ), if W is a matrix associated with f .
If we take any alternating submatrix (of even order) W  of W , we may consider
Pf(W  ). It is the analog of the minor of a determinant. The following lemma
expresses a pfaffian of f of order 2m in terms of pfaffians of order 2m − 2 (rankF
is unspecified).
Lemma V.1.3 For every a ∈ Λ2m+1 F ∗ , we have:
a(ϕ(m+1) ) = ϕ(m) (a)(ϕ).
130 Exactness criteria at work

Proof We use induction on m and the formula in Lemma III.3.1(cf. Reference


[25], pp. 460–461). 2
The above implies, in particular, the explicit formula found in the following
remark, and which will be used later in this chapter.
Remark V.1.4 Let W = (wij ) be an alternating matrix of order m (m either
even or odd). For every fixed i0 ,

m
Pf(W ) = (−1)i+i0 +1 γii0 wii0 Pf(W )ii0 ,
i=1

where

 1 if i < i0
γii0 = 0 if i = i0

−1 if i > i0 ,
and Pf(W )h1 ,...,hr denotes the pfaffian of the alternating submatrix of W formed
by deleting rows and columns indexed by h1 , . . . , hr .
For every fixed i0 and j0 such that i0 = j0 ,

m
(−1)i+i0 +1 γii0 wij0 Pf(W )ii0 = 0.
i=1

For every fixed i0 ,



m
wii0 Ti (W ) = 0,
i=1

where Ti (W ) stands for (−1)i+1 Pf(W )i .


Given again any alternating map f : F ∗ → F , for every m ≥ 1 we can consider
the pfaffian ideal P f2m (f ) generated by all pfaffians of f of order 2m. Then
an interesting corollary of Lemma V.1.3 is the following.
Corollary V.1.5 Let f : F ∗ → F be an alternating map. For each integer
m ≥ 1,
P f2m+2 (f ) ⊆ P f2m (f ).
Another corollary relates pfaffian ideals and determinantal ideals.
Corollary V.1.6 Let f : F ∗ → F be an alternating map. For each integer
m ≥ 1, we have:
1. I2m (f ) ⊆ P f2m (f ) ⊆ radI2m (f )
2. I2m−1 (f ) ⊆ P f2m (f )
3. if rankf = 2m − 1, then I2m−1 (f ) is nilpotent; if rankF = 2m − 1, then
det f = 0.
Pfaffian ideals 131

Proof Recall formula (V.1.1). For every t, It (f ) is generated by the coordinates


of Λt f (a) for various a, while the coordinates of ϕ(i) and ϕ(t−i) involve pfaffians of
order 2i and 2(t−i), respectively. Hence if t is either 2m or 2m−1, every term on
the right-hand side is contained in P f2n (f ) for some n ≥ m. By Corollary V.1.5,
P f2n (f ) ⊆ P f2m (f ) and we get the second statement, as well as the first inclusion
of the first statement.
To finish the proof of statement 1, we simply remember that the square of
each order 2m pfaffian of f is a minor of order 2m.
Statements 1 and 2 imply the first part of statement 3. As for the second part,
again use formula (V.1.1): if rankF = 2m − 1, ϕ(i) = 0 for i ≥ m; but one of i
and 2m − 1 − i is larger than or equal to m. 2

V.1.2 Resolution of a certain pfaffian ideal


Let R be a noetherian commutative ring, and let n ≥ 3 be an odd integer; say,
n = 2k + 1. Let F be a free R-module of rank n, and let f : F ∗ → F be an
alternating map. Suppose that the pfaffian ideal P f2k (f ) has grade 3. We are
going to build a sequence
g∗ f g
E : 0 → R → F∗ → F → R
which is a resolution of R/P f2k (f ).
As usual, let ϕ be the element of Λ2 F corresponding to f , so that, by
Lemma V.1.1, f (a) = a(ϕ) for every a ∈ F ∗ . Let e be a generator of Λn F ∗
and define g : F → R by means of
g(b) = ϕ(k) (e)(b).
We claim that im(g) = P f2k (f ). For Lemma 3.3.2 implies that
ϕ(k) (e)(b) = b(e)(ϕ(k) ),

and as b ranges over a basis of F , b(e) ranges over a basis of Λ2k F . Hence im(g)
is the ideal generated by the coordinates of ϕ(k) , that is, im(g) = P f2k (f ).
We first check that E is a complex. Since f ∗ = −f , it is enough to show
gf = 0. Now, for a ∈ F ∗ ,
gf (a) = ϕ(k) (e)(f (a)) = ϕ(k) (e)(a(ϕ)) = (ϕ(k) (e) ∧ a)(ϕ)
= −(a ∧ ϕ(k) (e))(ϕ) = −a(ϕ(k) (e)(ϕ)).
But by Lemma V.1.3,
ϕ(k) (e)(ϕ) = e(ϕ(k+1) ) = 0,

since ϕ(k+1) ∈ Λn+1 F = 0.


In order to show that E is exact, we use Theorem IV.1.7. We need to check that
rankf + rankg = n, depthI(g) ≥ 3 and depthI(f ) ≥ 2 (note that I(g ∗ ) = I(g)).
132 Exactness criteria at work

By statement 1 of Corollary V.1.6, I1 (g) = P f2k (f ) has the same radical


as I2k (f ).
By statement 3 of Corollary V.1.6, rankf = 2k, and I2k (f ) has grade 3 so
rank g = 1 and I(g) = I1 (g) has grade 3 as well.
This completes the proof that E is exact.
Remark V.1.7 If, in addition, R is local, with maximal ideal m, and we
assume im(f ) ⊆ mF , then
P f2k (f ) = im(g) ⊆ mR and im(g ∗ ) ⊆ mF ∗ .
That is, E is a minimal free resolution of R/P f2k (f ).
We are now ready to state the main theorem of this section.
Theorem V.1.8 Let (R, m) be a noetherian local ring.
1. Let n ≥ 3 be an odd integer, and let F be a free R-module of rank n. Let
f : F ∗ → F be an alternating map whose image is contained in mF . Suppose
P fn−1 (f ) has grade 3. Then P fn−1 (f ) is a Gorenstein ideal, minimally
generated by n elements.
2. Every Gorenstein ideal of grade 3 arises as in Part 1.
Remark V.1.7 above indicates that we have completely proven Part 1 of the
theorem (the property of being Gorenstein comes from having R as the left-most
non-zero term in E).
The proof of Part 2 is given in the rest of this section.

V.1.3 Algebra structures on resolutions


Let R be a commutative ring, and let
d d d
P : · · · →3 P2 →2 P1 →1 P0 = R
be a projective resolution of a cyclic R-module R/a, where a = im(d1 ). It is
clear that P ⊗R P has the structure of a complex, with differential
d(f ⊗ g) = df ⊗ g + (−1)deg f f ⊗ dg,
and that any comparison map P ⊗R P → P covering the natural map
R/a ⊗R R/a → R/a
makes P into a homotopy-associative, homotopy-commutative differential graded
algebra. (The property of being “homotopy-X” means the property of being X
up to homotopy.) We claim that one can make P commutative, as well.
Let the symmetric square S2 (P) be the quotient (P ⊗R P)/M, where M is
the graded subcomplex of P ⊗R P generated by
 
f ⊗ g − (−1)deg f deg g g ⊗ f | f, g homogeneous elements of P .
Pfaffian ideals 133

Since d(M) ⊆ M, S2 (P) inherits the structure of a complex from P ⊗R P.


Furthermore, each component of S2 (P) is a projective R-module, because for
each k one has
 

S2 (P)k ∼
= Pi ⊗R Pj  + Tk ,
i+j=k, i<j

where

0 if k is odd
Tk = Λ2 Pk/2 if k is of the form 4n + 2

S2 Pk/2 if k is of the form 4n.
Since S2 (P) is isomorphic to P in degrees 0 and 1, the comparison theorem for
projective complexes says that there exists a map of complexes ϕ : S2 (P) → P
which extends the natural isomorphism R/a ⊗R R/a → R/a, and which is the
identity on the subcomplex R ⊗R P ⊆ S2 (P). If we set
f · g = ϕ(f ⊗ g),

where f , g are in P, and f ⊗ g is the image, in S2 (P), of f ⊗ g, then the con-


ditions the map satisfies are equivalent to the statement that · makes P into a
(not necessarily associative) commutative, differential graded algebra, with the
differential d, and with structure map R → P given by R = P0 .
The uniqueness up to homotopy of comparison maps implies that the above
algebra structure, induced by ϕ, is homotopy-associative. In fact, it is even asso-
ciative if Pi = 0 for i ≥ 4 (recall that in this section, we are concerned with
resolutions of length at most 3).
Proposition V.1.9 With notation as above, if Pi = 0 for i ≥ 4, then P is a
commutative, associative graded algebra.
Proof There is just one associativity to be checked:
(f · g) · h = f · (g · h),
with f, g, h in P1 . Since d3 is a monomorphism, it is enough to show:
d3 ((f · g) · h) = d3 (f · (g · h)).
But both sides coincide with:
d1 (f ) · g · h − f · d1 (g) · h + f · g · d1 (h),
for d is a derivation and d1 (f ), d1 (g), d1 (h) are in R. 2
We now specialize and suppose that R is a noetherian local ring, that a is a
Gorenstein ideal of R of grade g, and that
dg d d
F : 0 → Fg → Fg−1 → · · · →2 F1 →1 R
134 Exactness criteria at work

is a minimal free resolution of R/a, equipped with a commutative multiplication


F ⊗R F → F which we denote by a dot. Since a is Gorenstein, we can make the
identification Fg = R, so that, for each k ≤ g, the dot-multiplication map
Fk ⊗R Fg−k → Fg = R

induces a map sk : Fk → Fg−k .
Theorem V.1.10 For every k ≤ g, sk is an isomorphism.
Proof We claim that the following diagram is commutative up to sign:
dg d d
F : 0 → Fg → Fg−1 → · · · → F2 →2 F1 →1 R

↓ sg−1 ↓ s2 ↓ s1

d∗ d∗ d∗
F∗ : 0 → R →
1
F1∗ ∗
→ · · · → Fg−2 ∗
→ Fg−1 → Fg∗ .
g−1 g

For, if a ∈ Fk and b ∈ Fg−k+1 , then a · b ∈ Fg+1 = 0, so


0 = dg+1 (a · b) = dk (a) · b + (−1)k a · dg−k+1 (b),
or dk (a) · b = ±a · dg−k+1 (b). If, for α ∈ F∗ and b ∈ F, we write α, b ∈ R for
the value of the linear functional α at b, we thus have
sk−1 dk (a), b = dk (a) · b
= ±a · dg−k+1 (b)
= sk (a), dg−k+1 (b)
/ 0
= d∗g−k+1 (a), b .
Hence sk−1 dk = ±d∗g−k+1 sk , as required.
Now, since a is Gorenstein, both F and F∗ are minimal free resolutions of R/a
and it follows that any map which extends an isomorphism in degree 0 must be
an isomorphism. Thus each sk is an isomorphism. 2

V.1.4 Proof of Part 2 of Theorem V.1.8


Let a be a Gorenstein ideal of grade 3, minimally generated by n elements. Let
d d d
F : 0 → F3 →3 F2 →2 F1 →1 R
be a minimal free resolution of R/a, so that F1 ∼ = Rn . In the previous subsection
we saw that F has the structure of a commutative, associative, differential graded
algebra. By Theorem V.1.10, the multiplication of this algebra gives a perfect
pairing F2 ⊗R F1 → F3 . If we identify F3 with R, such a pairing yields an
identification F2 = F1∗ . We first show that, with respect to this identification, d2
is alternating. Let {ei } and {εi } be dual bases for F1 and F2 = F1∗ , respectively,
Pfaffian ideals 135

and write (wij ) for the matrix of d2 relative to these bases. Thus, writing −, −
for the pairing between F2 and F1 ,
wij = εi , d2 (εj ) = εi · d2 (εj ) ∈ R = F3 ,
where · is the product for the algebra structure on F. We must show that wij =
−wji and wii = 0. Applying d3 to the above equation, and using the formula for
the differentiation of a product, we get
d3 (wij ) = d2 (εi ) · d2 (εj ) + εi · d1 (d2 (εj ))
= d2 (εi ) · d2 (εj ).
Since · satisfies the graded commutative law and d2 (εi ) and d2 (εj ) have degree
1, we have
d3 (wij ) = −d3 (wji ), d3 (wii ) = 0.
Since d3 is a monomorphism, this shows that d2 is alternating.
We next show that d2 satisfies the other conditions of Part 1 of Theorem V.1.8.
By Theorem IV.1.7, rank d2 = n − 1 and In−1 (d2 ) has grade ≥ 2. By statement
3 of Corollary V.1.6, n − 1 is even, so n is odd. Since F is a minimal resolution,
im(d2 ) ⊆ mF1 . To see that grade P fn−1 (d2 ) ≥ 3, we first note that by statement
1 of Corollary V.1.6 , rad P fn−1 (d2 ) = rad In−1 (d2 ), so it is enough to show that
grade In−1 (d2 ) ≥ 3. Since In−1 (d2 ) = I(d2 ), and I(d1 ) = I1 (d1 ) = a has grade 3
by assumption, we are done by Part (a) of Proposition IV.1.11.
Having proved that d2 satisfies all the conditions of the map f in Part 1 of
Theorem V.1.8, proceeding as in Subsection V.2.2, we can construct an exact
sequence
g∗ g
E : 0 → R → F1∗ →2 F1 → R.
d

But under our identifications F3 = R and F2 = F1∗ , we have d3 = d∗1 . Since both
d∗1 and g ∗ are kernels of d2 , there is a unit u ∈ R such that the following diagram
is commutative:
g∗
R −→ F1∗
↓u  .

R −→

F 1
d1

Hence ug = d1 and
P fn−1 (d2 ) = im(g) = im(ug) = imd1 = a,
as required.
This completes the proof of Theorem V.1.8.
One should remark that, given a Gorenstein ideal, a, of grade 3 as above,
minimally generated by n = 2k + 1 elements, we can rewrite the resolution
g∗ g
E : 0 → R → F1∗ →2 F1 → R
d
136 Exactness criteria at work

as follows:
g f g
E : 0 → Λn F1 → Λn−1 F1 → F1 → R,
by means of the isomorphisms Λk F → Λn−k F ∗ such that a → a(e) [e = chosen
orientation of Λn F ∗ ]. Then the pairing on E , which gives the duality expressed
by Theorem V.1.10, is precisely the multiplication
F1 ⊗R Λn−1 F1 → Λn F1 , R ⊗R Λn F1 → Λn F1
in the exterior algebra.
Finally, we observe that there is a close analogy between E and the Koszul
complex (which E must be, if n = 3). Namely, we note that the following
diagram is commutative, where the first row is a truncation of the Koszul complex
associated with g : F1 → R, where γ : ΛF1 → ΛF1 is the map a → a ∧ ϕ(k−1) ,
and ϕ is the element of Λ2 F1 corresponding to f  .
g
Λ3 F1 → Λ2 F1 → F1 → R

↓γ ↓γ  

g f g
Λn F1 → Λn−1 F1 → F1 → R.

V.2 Powers of pfaffian ideals


Let R be a noetherian ring. Let n be a positive integer such that n = 2k + 1,
k ≥ 1, let Zij , 1 ≤ i < j ≤ n, be 12 n(n − 1) independent indeterminates, and call
X the n × n matrix (Xij ) such that

 Zij if i < j
Xij = −Zij if i > j

0 if i = j
(X is called a generic alternating matrix). Finally, let S stand for the polynomial
ring R[X] (=R[Zij ]).
It is a fact that the ideal P f2k (X) has grade 3 (cf. e.g. Reference [55],
corollary 2.5); hence P f2k (X) has a finite free S-resolution as described in Sub-
section V.1.2 above. In this section we describe a finite free S-resolution for each
of the ideals [P f2k (X)]m , m ≥ 2. Interest in such a class of ideals was originally
raised by the desire to compare its homological behavior with that of the family
studied in Reference [24]. Let R[Uij , Tk ], 1 ≤ i ≤ m, 1 ≤ j ≤ n, 1 ≤ k ≤ m, be
a polynomial ring in mn + m indeterminates, with m ≥ n. Let ϕ : Rm → Rn
be the map with matrix (Uij ), and let a : R → Rm be the map with matrix
(T1 , . . . , Tm ). Let J(ϕ, a) denote the ideal generated by the n × n minors of (Uij )
and by the entries of the product (T1 , . . . , Tm )(Uij ). It is proven in Reference [24]
that J(ϕ, a) has grade m and is perfect, i.e., its homological dimension equals its
grade. We will see below that, if n = 3 and m ≥ 3, [P f2k (X)]m is never perfect,
Powers of pfaffian ideals 137

for it always has grade 3 while its homological dimension depends on m (in a
very nice way).

V.2.1 Intrinsic description of the matrix X


Let F0 be a free R-module of rank n = 2k+1, and let A be the symmetric algebra:
A = S(Λ2 F0 ). Set F = A ⊗R F0 . We define a degree one A-map, f : F → F ∗ ,
in the following way. For every r ≥ 0, the map f , restricted to the component
Ar ⊗R F0 , is the composite
Ar ⊗R F0 −→ Ar ⊗R F0 ⊗R F0 ⊗R F0∗
1⊗1⊗CF0

−→ Ar ⊗R Λ2 F0 ⊗R F0∗ = Ar ⊗R A1 ⊗R F0∗ −→ Ar+1 ⊗R F0∗ ,


m⊗1
1⊗m⊗1

where m denotes multiplication in the appropriate algebras and CF0 stands for
the element of
F0 ⊗R F ∗ ∼
= HomR (F0 , F0 )
0

corresponding to the identity on F0 .


Choosing dual bases {e1 , . . . , en } and {ε1 , . . . , εn } for F0 and F0∗ , CF0 equals
n
j=1 ej ⊗ εj and the composite map:

F0 −→ F0 ⊗R F0 ⊗R F0∗ −→ Λ2 F0 ⊗R F0∗ ,
1⊗CF0 m⊗1

sends each ei to j (ei ∧ ej ) ⊗ εj . Hence (ei ∧ ej ) is the matrix associated to
f with respect to the bases of F and F ∗ induced by those of F0 and F0∗ (we
denote the induced bases with the same symbols used for the original ones). By
means of the identification ei ∧ ej → Xij , the matrix (ei ∧ ej ) coincides with the
generic alternating matrix X introduced at the beginning of this section, and
A = S(Λ2 F0 ) is isomorphic to S = R[X].
From now on, we use S to indicate both R[X] and S(Λ2 F0 ), dropping the
symbol A.
The morphism, f , is called the generic alternating map. Asseen in Section
1, f corresponds to an element α ∈ Λ2 F ∗ , namely, we have α = i<j Xij εi ∧ εj .
Moreover, the grade 3 ideal P f2k (X), which we denote by a, for short, has the
by now familiar resolution (a resolution of S/a, i.e.):
g∗ f g
E : 0 → S → F → F ∗ → S.
The morphism f is as above. As for g, choose an orientation e ∈ Λn F , that is,
an identification between Λ2k F and F ∗ . Then g can be identified with
ω2k : Λ2k F → S, a → β(α(k) ⊗ a),
where β stands for the natural pairing Λ2k F ∗ ⊗S Λ2k F → S given by the ΛF ∗ -
structure of ΛF .
Concretely, g can be viewed as sending each εi to Ti (X) = (−1)i+1 Pf(X)i .
Finally, g ∗ (1) = α(k) .
138 Exactness criteria at work

V.2.2 Hooks again


We already met in Subsection III.1.2 some “hook” GL(F )-representations, Lλ F ,
where λ stands for

(λ1 , 1, . . . , 1) = (λ1 , 1t ),
% &# $
t

or, more pictorially,


λ1 boxes
# $% &
···
|
|
λ= .. ,
t boxes  .
|
|

with λ1 and t nonnegative integers such that λ1 ≥ 1 if t > 0. As an S-module,


Lλ F is defined as the cokernel of the composite

δ : Λλ1 +1 F ⊗ St−1 F −→ Λλ1 F ⊗ F ⊗ St−1 F −→ Λλ1 F ⊗ St F,


∆⊗1 1⊗m

where ∆ (resp., m) is diagonalization (resp., multiplication) in the algebra ΛF


(resp., SF ).
Consider the basis of Λλ1 F ⊗St F induced by the basis {e1 , . . . , en } of F . Given
an element

e1 ∧ · · · ∧ eλ1 ⊗ ej1 · · · ejt

of such a basis, denote its image in L(λ1 ,1t ) F by the “tableau”

i1 ··· iλ1
j1

.. .
.

jt

As proved in Subsection III.1.2, the following statements hold:

• A basis of L(λ1 ,1t ) F is given by all tableaux such that the indices in the first
row are strictly increasing and those in the first column are weakly increasing
(“standard tableaux”).
• A tableau which is not standard is equal to a Z-linear combination of
standard tableaux (“straightening law”).
Powers of pfaffian ideals 139

Explicitly, the key step of the straightening law is:


i1 ··· iλ1 j1 i1 ··· ih ··· iλ1
j1 ih
j2 
λ1
= (−1)h−1 j2
.. h=1 ..
. .
jt jt

(j1 < i1 < · · · < iλ1 )

where ih means ih omitted.


Finally, suppose that F = F1 ⊕ F2 , with F1 generated by e1 , . . . , eh and F2
generated by eh+1 , . . . , en for some fixed h. One has an isomorphism of free
S-modules
 

L(λ1 ,1t ) F ∼
= (L(µ1 ,1u ) F1 ⊗ St−u F2 ⊗ Λλ1 −µ1 F2 ) ⊕ L(λ1 ,1t ) F2 ,
(µ1 ,1u )

where (µ1 , 1u ) ranges over all the hooks which are nested in (λ1 , 1t ). But as
GL(F1 ) × GL(F2 )-modules, such an isomorphism usually holds only up to a
filtration.

V.2.3 Some representation theory


It will be shown in Chapter VI that L(λ1 ,1t ) F is just an instance of a wider family
of GL(F )-representations, Lλ F , where λ = (λ1 , . . . , λq ) is a partition, that is,
either (0) or a weakly decreasing sequence of positive integers.
Lλ F is a free S-module and a GL(F )-representation. Moreover, the ΛF ∗ –
module structure of ΛF allows the construction of a natural isomorphism
Lλ F ⊗S Λn F ∗ ⊗S · · · ⊗S Λn F ∗ → Lλ∗ F ∗ ,
% &# $
q

where q is the number of non-zero parts of λ, and λ∗ is the partition (n − λq , n −


λq−1 , . . .) (cf., e.g., Reference [4], proposition 2.4.2).
Another key property of Lλ F is universality, namely, the construction of this
module is essentially defined over the integers and then carried over to any other
ring. In particular, if we are in the situation of Subsection V.2.1 above, we have
Lλ F = Lλ (F0 ⊗R S) ∼ = Lλ (F0 ) ⊗R S.
As we shall see in a little while, this has interesting implications if we assume
that our ground ring R is a field of characteristic zero. For then every finite
dimensional polynomial representation of GL(F0 ) is completely reducible, that
is, it decomposes as a direct sum of irreducibles, and a complete set of irreducible
representations of GL(F0 ) is
{Lλ F0 | λ is a partition}.
140 Exactness criteria at work

In particular, for every h ≥ 0,


Sh (Λ2 F0 ) ∼
= ⊕Lλ F0 ,
where λ = (λ1 , . . . , λq ) ranges over all partitions having even entries and such
that 2h = λ1 +· · ·+λq (cf., e.g. Reference [66], chapter 1; in particular, exercise 6
on p. 46).
Also every GL(F0 )-module of type Lλ F0 ⊗R Lµ F0 is a direct sum of GL(F0 )-
irreducibles:
Lλ F0 ⊗R Lµ F0 = ⊕γ(λ, µ; ν) Lν F0 ,
ν

and there is a combinatorial rule (known as the Littlewood-Richardson


rule) describing the indicated multiplicities γ(λ, µ; ν) (cf., e.g. Reference [66],
chapter 1; in particular, section 9).

V.2.4 A counting argument


This subsection contains some heuristic considerations leading to the construc-
tion of the resolutions of the powers am = [P f2k (X)]m . For simplicity, we restrict
to the case m = 2.
Having at hand the resolution E of a = P f2k (X), it is natural to assume that
the first map, ϑ1 , of a resolution of a2 must coincide with the second symmetric
power of the map g : F ∗ → S having im(g) = a. For every r ≥ 0, ϑ1 is the
composite
Sr ⊗R S2 (Λ2k F0 ) −→ Sr ⊗R S2k (Λ2 F0 ) → S2k+r (Λ2 F0 ),
1⊗S2 ((ω2k )0 ) m

where (ω2k )0 is the R-map Λ2k F0 → Sk (Λ2 F0 ) inducing ω2k over S. Since
S2 (Λ2k F0 ) ∼
= S2 F0∗ = L(2k+1,1,1) F0∗ ∼
= L(2k,2k) F0 , ϑ1 is in fact a morphism

Sr (Λ2k F0 ) ⊗R L(2k,2k) F0 −→ S2k+r (Λ2 F0 ).


We study the latter morphism, when r ranges in N, in order to get a clue about
ker(ϑ1 ). We do this by assuming that R is a characteristic zero field, for the given
modules are universal and we may hope that all the necessary information on our
complex is available in the characteristic zero case. If R is assumed to be a field of
characteristic zero, then the irreducible components of Sr (Λ2k F0 ) ⊗R L(2k,2k) F0
either are mapped onto the corresponding irreducibles of S2k+r (Λ2 F0 ), or must
occur in the kernel.
For r = 0, we have S0 ⊗R L(2k,2k) F0 → S2k (Λ2 F0 ), which is just the inclusion
of L(2k,2k) F0 into S2k (Λ2 F0 ) ∼
= ⊕Lλ F0 (each λ having even entries, adding up
to 4k). Hence there is no contribution to ker(ϑ1 ).
For r = 1, we have Λ2 F0 ⊗R L(2k,2k) F0 → S2k+1 (Λ2 F0 ). By the Littlewood–
Richardson rule, one checks that the domain is isomorphic to
L(2k+1,2k,1) F0 ⊕ L(2k,2k,2) F0
Powers of pfaffian ideals 141

(since F0 has rank k + 1). As S2k+1 (Λ2 F0 ) ∼


= ⊕Lλ F0 , each λ having even entries,
adding up to 4k + 2, L(2k+1,2k,1) F0 must occur in the kernel.
For r = 2, we thus have
Λ2 F0 ⊗R L(2k+1,2k,1) F0 → S2 (Λ2 F0 ) ⊗R L(2k,2k) F0 → S2k+2 (Λ2 F0 ).
Again by means of the Littlewood–Richardson rule, one checks that
Λ2 F0 ⊗R L(2k+1,2k,1) F0 ∼
= L(2k+1,2k+1,2) F0
⊕ L(2k+1,2k+1,1,1) F0
⊕ L(2k+1,2k,3) F0
⊕ L(2k+1,2k,2,1) F0
and that
S2 (Λ2 F0 ) ⊗R L(2k,2k) F0 ∼
= L(2k+1,2k,3) F0
⊕ L(2k,2k,4) F0
⊕ L(2k+1,2k+1,1,1) F0
⊕ L(2k+1,2k,2,1) F0
⊕ L(2k+2,2k,2,2) F0 .
Since S2k+2 (Λ2 F0 ) ∼
= ⊕Lλ F0 , each λ having even entries, adding up to 4k + 4,
it follows that L(2k+1,2k+1,2) F0 must occur in the complex, in degree 3.
For r = 3, however, we get
0 → Λ2 F0 ⊗R L(2k+1,2k+1,2) F0 → S2 (Λ2 F0 ) ⊗R L(2k+1,2k,1) F0
→ S3 (Λ2 F0 ) ⊗R L(2k,2k) F0 → S2k+3 (Λ2 F0 ),
that is, no new term is necessary in degree 4.
Thus, assuming R to be a characteristic zero field, one conjectures that a
resolution of a2 may look like this:
ϑ ϑ ϑ
0 → L(2k+1,2k+1,2) F →3 L(2k+1,2k,1) F →2 L(2k,2k) F →1 S.
Of course some extra terms could be necessary in a characteristic-free setting.
Yet we take the above as a reasonable candidate and start looking for possible
definitions of the boundary morphisms, about which we have no hints.
Note that our candidate can also be expressed in terms of F ∗ :
ϑ ϑ ϑ
0 → L(2k−1) F ∗ →3 L(2k,1) F ∗ →2 L(2k+1,1,1) F ∗ →1 S.

Λ2k−1 F ∗ S2 F ∗
Since the map f : F → F ∗ of E can easily be identified with

L(2k) F ∗ → L(2k+1,1) F ∗ , u → 
p(2k+1,1) (αl1 ∧ u ⊗ αl1 ),
l
142 Exactness criteria at work

 
where l αl1 ⊗ αl1 is the image of α given by the component Λ2 F ∗ → F ∗ ⊗S F ∗
of the diagonal map, and where p(2k+1,1) is the projection Λ2k+1 F ∗ ⊗S S1 F ∗ →
L(2k+1,1) F ∗ , one conjectures that ϑ2 and ϑ3 are induced by

Λ2k F ∗ ⊗S S1 F ∗ → Λ2k+1 F ∗ ⊗S S2 F ∗ , u ⊗ v → 
αl1 ∧ u ⊗ vαl1
l

and by

Λ2k−1 F ∗ → Λ2k F ∗ ⊗S S1 F ∗ , u → 
αl1 ∧ u ⊗ αl1 ,
l

respectively. One checks that ϑ2 and ϑ3 are actually well defined, by invoking
the following more general result, which holds for every ring R.
Lemma
  V.2.1 Let ϕ  be the map Λa F ∗ ⊗S Sb F ∗ → Λa+1 F ∗ ⊗S Sb+1 F ∗ , u⊗v →
α
l l1 ∧ u ⊗ vαl1 . Then ϕ induces a morphism
ϕ : L(a,1b ) F ∗ → L(a+1,1b+1 ) F ∗ .
Proof Recall the expression of a hook Schur module
 as a cokernel, and perform
calculations applying the diagonal map to α = i<j Xij εi ∧εj (cf. Reference [13],
p. 472). 2
It is not hard to see that our candidate is indeed a complex. That ϑ2 ◦ ϑ3 = 0
can be seen as follows. Let ϕ 2 and ϕ 3 be the maps inducing ϑ2 and ϑ3 ,
respectively. Then
 

2 (ϕ
ϕ 2 
3 (u)) = ϕ Xij (εi ∧ u ⊗ εj − εj ∧ u ⊗ εi )
i<j
 
= Xij Xpq (εp ∧ εi ∧ u ⊗ εq εj − εp ∧ εj ∧ u ⊗ εq εi
i<j p<q

− εq ∧ εi ∧ u ⊗ εp εj + εq ∧ εj ∧ u ⊗ εp εj )
and the double sum equals zero (think of what happens if one interchanges
j ↔ q and i ↔ p). We also have that ϑ1 ◦ ϑ2 = 0 because if εi stands for
ε1 ∧ · · · ∧ εi−1 ∧ εi+1 ∧ · · · ∧ εn , then

ϑ1 (ϑ2 (p(2k,1) (εi ⊗ εj ))) = Xpq ϑ1 (p(2k+1,1,1) (εp ∧ εi ⊗ εj εq ))
p =q

= Xiq (−1)i−1 Tj (X)Tq (X)
q
1 2

i−1
= (−1) Tj (X) Xiq Tq (X)
q

and the sum in brackets is zero by the last formula of Remark V.1.4.
This concludes our heuristic considerations.
Powers of pfaffian ideals 143

V.2.5 Description of the resolutions


We begin the formal construction of some complexes Cm , m ≥ 2; these will be
shown to be the required resolutions of the ideals am .
Definition V.2.2 Let f : F → F ∗ be the generic alternating map of
Subsection V.2.1, where rankF = n = 2k + 1 for some positive k.
1. If m ≥ 2k, then Cm is defined to be the sequence
ϕ ϕ
0 → L(1m−n+2 ) F ∗ → L(2,1m−n+2 ) F ∗ → · · ·
ϕ ϕ ψ
· · · → L(n−1,1m−1 ) F ∗ → L(n,1m ) F ∗ → S.
2. If m = 2r, 1 ≤ r ≤ k − 1, then Cm is defined to be the sequence
ϕ ϕ
0 → L(n−m) F ∗ → L(n−m+1,1) F ∗ → · · ·
ϕ ϕ ψ
· · · → L(n−1,1m−1 ) F ∗ → L(n,1m ) F ∗ → S.
3. If m = 2r + 1, 0 ≤ r ≤ k − 1, then Cm is defined to be the sequence
χ ϕ ϕ
0 → S → L(n−m) F ∗ → L(n−m+1,1) F ∗ → · · ·
ϕ ϕ ψ
· · · → L(n−1,1m−1 ) F ∗ → L(n,1m ) F ∗ → S.
In all cases, ϕ stands for the map induced by the appropriate

 : Λa F ∗ ⊗S Sb F ∗ → Λa+1 F ∗ ⊗S Sb+1 F ∗ ,
ϕ u ⊗ v → 
αl1 ∧ u ⊗ vαl1 ,
l

and ψ is the m-th symmetric power of the map g : F → S defined in Subsec-
tion V.2.1 (explicitly, ψ(εi1 · · · εim ) = Ti1 (X) · · · Tim (X)). As for χ in the third
case, it is defined by χ(1) = α(k−r) .
We point out that the morphisms ϕ of Cm are well defined, thanks to the
lemma of the previous subsection.
When m = 1, Cm gives back E, the resolution of the pfaffian ideal a.
In the special case n = 3, each Cm coincides with the well-known resolution
of the m-th power of an ideal generated by a regular sequence (having three
elements); one finds such a resolution in Reference [24]. It is interesting to notice
that the hook Schur modules L(λ1 ,1t ) F ∗ were introduced in that paper for the
first time.
Proposition V.2.3 Cm is a complex, for every m ≥ 1.
Proof One shows that every composition ϕ ◦ ϕ is zero exactly in the same way
we did for ϕ2 ◦ ϕ
3 at the end of the previous subsection.
One shows that every composition ψ ◦ ϕ is zero exactly in the same way we
did for ϑ1 ◦ ϑ2 at the end of the previous subsection.
144 Exactness criteria at work

It remains to show that if m = 2r + 1, 0 ≤ r ≤ k − 1, then ϕ ◦ χ = 0, that is,


ϕ(χ(1)) = ϕ(α(k−r) ) = 0. One writes down explicitly the divided power α(k−r) ,
applies ϕ, and checks vanishing, first, by taking into account the key step of
the straightening law recalled in Subsection V.2.2 and then by resorting to the
formulas of Remark V.1.4. (Cf. Reference [13], pp. 474–477.) 2

Theorem V.2.4 For every m ≥ 1, the complex Cm is a resolution of S/am .

The proof of this theorem is discussed in the following subsection. Here we


just make some comments.
• The resolution Cm is minimal, that is, for every i ≥ 1, the image of the
i-th morphism of Cm is included in m(Cm )i−1 , where m denotes the ideal
of S generated by the indeterminates occurring in the matrix X.
• Cm is also generic (or universal), in the sense that it is defined over Z[X]
and then carried over to every other R[X], R noetherian. In particular,
the Betti numbers of am (that is, the ranks of the modules (Cm )i ) are
independent of the characteristic of R.
• The existence of a generic minimal free resolution for S/am is remarkable,
since it is known that S/P f2p (X), 2 ≤ p ≤ k − 1, may not admit such a
resolution. More precisely, [62] contains a case in which the Betti numbers
of S/P f2p (X) depend on the characteristic of R. Thus the ideals P f2p (X)
behave pretty much in the same way as the determinantal ideals of a generic
n1 × n2 matrix (cf. [48]). More on this in Section VII.5.
• Cm shows that the homological dimension of S/am satisfies

n if m ≥ 2k
hdS/am = m + 1 if m = 2r, 1 ≤ r ≤ k − 1

m + 2 if m = 2r + 1, 0 ≤ r ≤ k − 1
(notice that m + 2 = n, when m = 2(k − 1) + 1). Since, as we know,
an ideal is called perfect if its depth and homological dimension coincide, it
follows from grade am = grade a = 3 that am is perfect in every characteristic
(“generically perfect”) when m = 1, 2 (no matter what the value of n) and
whenever n = 3 (in this case, P f2k (X) is generated by a regular sequence).
In all the other cases, am is not perfect. However, there is a regular pattern
for hd S/am , namely,
hd S/a2r+1 = hd S/a2r+2 = 2r + 3 if 0 ≤ r ≤ k − 2
hd S/am = 2k + 1 if m ≥ 2k − 1.
This regularity can be seen as a special case of a more general phenomenon
related to “almost alternating maps,” cf. Reference [63], section 5.
Another type of example can be found in References [5] and [6]. In those
papers, the homological dimensions of certain generic modules associated
to partitions are seen to depend on the size of the Durfee squares of those
Powers of pfaffian ideals 145

partitions. (A definition of this combinatorial invariant of a partition can be


found in Subsection VII.3.1.)

V.2.6 Proof of Theorem V.2.4


We proceed by induction on k = (n − 1)/2.
When k = 1, that is, n = 3, we have already observed (immediately before
Proposition V.2.3) that Cm resolves S/am . So we prove the statement for k ≥ 2,
assuming it true for all values less than k.
Since we already know that im ψ = am , it is enough to show that the positive
homology vanishes. We make use of Theorem IV.1.10. Let p ∈ SpecS be such
that

 n if m ≥ 2k
gradepSp < lengthCm = m + 1 if m = 2r, 1 ≤ r ≤ k − 1

m + 2 if m = 2r + 1, 0 ≤ r ≤ k − 1.

The ideal, m, generated in S by the indeterminates occurring in X, has grade


n(n − 1)/2 ≥ n. Thus grade mSp > grade pSp , no matter what m is, and one
of the variables Xij is invertible over Sp , say X12 . But then the exactness of
Cm ⊗S Sp follows, if we show that Cm is exact after localization at the powers
of X12 . The idea is that if X12 can be assumed invertible, one can find new dual
bases such that the corresponding matrix associated with the generic alternating
map is of type
 
0 1
 −1 0 ,
0

0 X

X  is a generic alternating matrix of order n = n − 2 = 2(k − 1) + 1, and the


inductive hypothesis applies.
More precisely, let R and S  be the localizations at the powers of X12
of R[X12 , X13 , X14 , . . . , X1n , X23 , X24 , . . . , X2n ] and S, respectively, and keep
denoting by F and F ∗ the corresponding objects over S  . As in Reference [13],
pp. 479–480 (also cf. [55]), there is an explicit way of choosing dual bases
{ε1 , . . . , εn } and {e1 , . . . , en } for F ∗ and F , resp., such that f ⊗S S  has the
matrix above, the entries of the alternating submatrix X  are algebraically inde-
pendent over R , S  = R [X  ], and aS  = a , where a is the ideal P f2(k−1) (X  )
of S  . Furthermore, ψ ⊗S S  sends ε1 and ε2 to 0, and sends εi (3 ≤ i ≤ n) to
X12 · Ti−2 (X  ). As for the element of Λ2 F ∗ associated with f ⊗S S  , it looks like

ε1 ∧ ε2 +  
Xij εi+2 ∧ εj+2
1≤i<j≤n

(n = 2k − 1, as above).
146 Exactness criteria at work

Accordingly, let us decompose F ∗ as H ∗ ⊕ G∗ , where H ∗ is generated by


ε3 , . . . , εn ,
and G∗ by ε1 , ε2 (so that ε1 ∧ ε2 can be denoted by αG∗ and

 
Xij εi+2 ∧ εj+2
1≤i<j≤n

by αH ∗ ). Recalling the end of Subsection V.2.2, we have


(Cm ⊗S S  )i = L(a,1b ) (H ∗ ⊕ G∗ )

= [ ⊕ (L(µ1 ,1u ) H ∗ ⊗ Sb−u G∗ ⊗ Λa−µ1 F2 )] ⊕ L(a,1b ) G∗ ,
(µ1 ,1u )

where a = n+1−i, b = m+1−i, and some summands may be 0, if rank H ∗ = n


is not large enough.
We intend to use the decompositions of the modules L(a,1b ) (H ∗ ⊕ G∗ ) in order
to filter the complex Cm ⊗S S  . Thus the exactness of Cm ⊗S S  will be reduced to
the exactness of the factors of the filtration. It is convenient to discuss separately
the four cases:
m ≥ 2k, m even
m ≥ 2k, m odd
m = 2r, 1≤r ≤k−1
m = 2r + 1, 0 ≤ r ≤ k − 1.
Here we deal with the first case only. The others are similar and can be found
in Reference [13] (pp. 487–490).
The H ∗ -content of a summand of L(a,1b ) (H ∗ ⊕ G∗ ) is defined as the number
µ1 + u, denoted by |H ∗ |. Hence L(a,1b ) (H ∗ ⊕ G∗ ) can be decomposed into the
direct sum of two modules: the one, M0 (a, 1b ), comprising all summands with
|H ∗ | even, and the other, M1 (a, 1b ), comprising all summands with |H ∗ | odd.
If a morphism ϕ⊗S S  is applied to L(a,1b ) (H ∗ ⊕G∗ ), then M0 (a, 1b ) is mapped
to M0 (a+1, 1b+1 ) and M1 (a, 1b ) is mapped to M1 (a+1, 1b+1 ). If ψ⊗S S  is applied
to L(a,1b ) (H ∗ ⊕ G∗ ), it means that a = n and b = m, whence µ1 = n ; recalling
the way ψ⊗S S  acts on the elements of {ε1 , . . . , εn }, it follows that ψ⊗S S  is zero
on all summands of L(a,1b ) (H ∗ ⊕G∗ ), except for L(µ1 ,1m ) H ∗ ⊗Λ2 G∗ ⊆ M1 (n, 1m )
(here we use the fact that m is even). Therefore Cm ⊗S S  is the direct sum of
two subcomplexes, M0 and M1 ; M0 is given by the terms M0 (a, 1b ); M1 is given
by the terms M1 (a, 1b ), together with S  . One shows that both of them are exact,
by filtering them separately.
As above, we will deal with only one case, that of M1 . The case of M0 is similar
and can be found in Reference [13] (pp. 485–486).
A filtration {Xt } of M1 is described as follows (recall that m ≥ 2k is assumed
to be even, say m = 2p).
For every fixed t ∈ {0, 1, . . . , p − 1}, in each L(a,1b ) (H ∗ ⊕ G∗ ) we assign to Xt
all the summands having
µ1 = n − j, u = 2t − j
Powers of pfaffian ideals 147

for some t ≤ t and some non-negative integer j. To Xp we assign S  and all the
summands of L(a,1b ) (H ∗ ⊕ G∗ ) having µ1 = n − j and u = 2t − j for some t ≤ p
and some non-negative integer j.
It is easy to see that Xp = M1 (cf. Reference [13], proposition at the end of
p. 481). Recalling the straightening law at the end of Subsection V.2.2, it is not
hard to check that each Xt , 0 ≤ t ≤ p, is indeed a complex (cf. Reference [13],
proposition on p. 482).
We now describe the factors Xt /Xt−1 , 0 ≤ t ≤ p − 1 (X−1 is set equal to 0),
and prove their exactness.
The modules occurring in Xt /Xt−1 are given by µ1 = n − j and u = 2t − j,
with j ranging between 0 and q, where

2t if 2t < n
q=
n − 1 if 2t > n .


It follows that a − µ1 = 3 − (i − j) and b − u = m + 1 − 2t − (i − j). Since


rank G∗ = 2 implies 0 ≤ a − µ1 ≤ 2, i − j can only be 1, 2, 3, whence
b − u = m − 2t, m − 2t − 1, m − 2t − 2,
respectively. Thus, no matter what j is, one finds terms of the following type:
A(j) = L(n −j,12t−j ) H ∗ ⊗ Sm−2t−2 G∗ ,
B(j) = L(n −j,12t−j ) H ∗ ⊗ Sm−2t−1 G∗ ⊗ Λ1 G∗ ,
C(j) = L(n −j,12t−j ) H ∗ ⊗ Sm−2t G∗ ⊗ Λ2 G∗ .
If one remembers the way αG∗ and αH ∗ operate, one checks that Xt /Xt−1 is the
total complex of the following double complex, Dt ,
0 0 0
↑ ↑ ↑
(0) (1)
ϕG∗ ϕG∗
0 → A(0) → B(0) → C(0) →0
(0) (1) (2)
↑ ϕH ∗ ↑ ϕH ∗ ↑ ϕH ∗
(0) (1)
ϕG∗ ϕG∗
0 → A(1) → B(1) → C(1) →0
(0) (1) (2)
↑ ϕH ∗ ↑ ϕH ∗ ↑ ϕH ∗
.. .. ..
. . .
(0) (1) (2)
↑ ϕH ∗ ↑ ϕH ∗ ↑ ϕH ∗
(0) (1)
ϕ ∗ ϕ ∗
0 → A(q − 1) →
G
B(q − 1) →
G
C(q − 1) → 0
(0) (1) (2)
↑ ϕH ∗ ↑ ϕH ∗ ↑ ϕH ∗
(0) (1)
ϕG∗ ϕG∗
0 → A(q) → B(q) → C(q) →0
↑ ↑ ↑
0 0 0
148 Exactness criteria at work

In the diagram, for each w = 0, 1, 2, the map


ϕH ∗ : L(µ1 ,1u ) H ∗ ⊗ Sv G∗ ⊗ Λw G∗ → L(µ1 +1,1u+1 ) H ∗ ⊗ Sv G∗ ⊗ Λw G∗
(w)

is defined by
p(µ1 ,1u ) (x ⊗ y) ⊗ z1 ⊗ z2

→ p(µ1 +1,1u+1 ) ((αH ∗ )l1 ∧ x ⊗ y(αH ∗ )l1 ) ⊗ z1 ⊗ z2 .
l

For each w = 0, 1, the map


ϕG∗ : L(µ1 ,1u ) H ∗ ⊗ Sv G∗ ⊗ Λw G∗ → L(µ1 ,1u ) H ∗ ⊗ Sv+1 G∗ ⊗ Λw+1 G∗
(w)

is defined by

x ⊗ y ⊗ z → (−1)µ1 x ⊗ y(αG∗ )l1 ⊗ (αG∗ )l1 ∧ z.
l

We remark that each line of Dt is a complex essentially because ϕ ◦ ϕ = 0.


The anticommutativity of the boxes is straightforward from the definitions,
(w)
particularly from the sign (−1)µ1 introduced in ϕG∗ .
Now, each row of Dt is isomorphic to a short exact sequence
0 → Sh G∗ → Sh+1 G∗ ⊗ Λ1 G∗ → Sh+2 G∗ ⊗ Λ2 G∗ → 0,
and the latter is isomorphic to the exact complex Λh+2 (G∗ ) of Subsection III.1.1,
since Λ2 G∗ ∼= S  . But the exactness of the rows of Dt implies that the total
complex of Dt , that is, Xt /Xt−1 , is exact as well, as anticipated.
We finally show that Xp /Xp−1 is exact, so that the whole M1 is exact.
The modules occurring in Xp /Xp−1 are S  and those associated with µ1 =
n − j and u = m − j, with j ranging between 0 and n − 1 (since m ≥ 2k gives


2p > n , and q = n − 1). Then a − µ1 = 3 − (i − j) and b − u = 1 − (i − j). Since


0 ≤ a − µ1 ≤ 2 and b − u ≥ 0, it follows that a − µ1 = 2 and b − u = 0. Hence
(recall Λ2 G∗ ∼
= S  ) Xp /Xp−1 is isomorphic to the complex
(2) (2)
ϕ ∗ ϕ ∗
Cm : 0 → L(1m−n +2 ) H ∗ →
H
L(2,1m−n +2 ) H ∗ →
H
···
(2) (2)
ϕ ∗ ϕ ∗
··· →
H
L(n −1,1m−1 ) H ∗ →
H
L(n ,1m ) H ∗ → S  ,

where ϕH ∗ is as before and L(n ,1m ) H ∗ → S  is the only non-zero component of


(2)

ψ ⊗S S  , namely (recall Λn H ∗ ∼

= S  ), the morphism defined by
(ε )b3 · · · (εn )bn → T3 (X  )b3 · · · Tn (X  )bn .
% 3 &# $
b3 +···+bn =m

But Cm
is a complex of the same type of Cm , relative to the generic alternating

matrix X of order 2(k − 1) + 1. Therefore it is exact by induction hypothesis.
This concludes our discussion of the exactness of the complexes Cm .
VI
WEYL AND SCHUR MODULES

In earlier chapters we introduced a very small dose of representation theory of


the general linear group when we discussed, in particular, the hook shapes. In
this chapter, we will deal more comprehensively with representations of that
group, over an arbitrary commutative ring, R. Our discussion will help place the
material in earlier chapters in proper perspective, and will move us forward to
some interesting observations relating to determinantal ideals and other matters.
In the classical theory, the fundamental shapes are the Ferrers diagrams cor-
responding to “partitions,” and the closely related “skew-partitions.” For our
purposes, we will have to consider a slightly larger class of shapes, corresponding
to the so-called almost skew-partitions. We should add that as of this writing,
the class of shapes that is being studied is far broader than this. However, to
study all of these would require quite a bit more of combinatorics than we pro-
pose to use here. We hope that the material we do cover, however, will help the
interested reader to get into this other area if he so desires.
We remind the reader that the details of the proofs of several of the the-
orems involving letter-place methods can be found in Appendix A following
Chapter VII.

VI.1 Shape matrices and tableaux


We will start this section with the definition of a shape matrix, and then will talk
about the partition or skew-partition, etc. associated with the matrix. However,
the reader should know that historically the sequence was the other way around:
the idea of partition and skew-partition, and Ferrers diagrams in general, arose
first; the idea of generalizing to shape matrices came afterwards. In practice, one
usually describes the shape by the diagram.
In the second subsection on tableaux, we will give a definition that generalizes
the one used in Chapter III. There, too, we introduce a quasi order on tableaux
that will be used in Appendix A.

VI.1.1 Shape matrices


Definition VI.1.1 A shape matrix is an infinite integral matrix A = (aij )
of finite support, with all the aij equal to zero or 1. (To say it has finite support
is to say that aij = 0 for only a finite number of indices i and j.) The last row
(column) of the shape matrix A is the last row (column) in which a non-zero
150 Weyl and Schur modules

term appears. Such a matrix is said to be row-convex (column-convex) if, in


each row (column), there are no zeroes lying between ones. (All the shapes that
we consider will be row- and column-convex.) The shape matrix B = (bij ) is a
subshape of A (written B ⊆ A) if bij ≤ aij for every i and j. The shape matrix,
A, is said to correspond to a partition if for all i, j, aij = 0 implies ai+1j = 0
and aij+1 = 0 implies aij = 0. It is said to correspond to a skew-partition or
skew-shape if A = B − C, where B and C correspond to partitions, and C ⊆ B.
It is said to be a bar shape if its only non-zero entries are in its last row, and
it is row-convex. Finally, it is said to correspond to an almost skew-partition
or almost skew-shape if A = B − C, where B corresponds to a skew-shape,
and C is a bar subshape matrix of B the index of whose last row coincides with
that of B, and whose first non-zero entry in that row occurs in the same place
as the first non-zero entry of B.
• Notice that, unless a given partition shape matrix is the zero matrix, a11 = 1.
We now illustrate each of these types of shape matrices. The typical partition
shape looks like this:
 
1 1 1 1 0 0···
 1 1 1 0 0 0··· 
 
 1 1 1 0 0 0··· 
 
(P )  1 1 0 0 0 0··· ,
 
 0 0 0 0 0 0··· 
 
.. .. .. .. .. ..
. . . . . .···
and is often represented by the Ferrers diagram:

The typical skew-shape looks like this:


 
0 0 1 1 1 0···
 0 0 1 1 0 0··· 
 
 0 1 1 0 0 0··· 
 
(S)  1 1 0 0 0 0··· ,
 
 0 0 0 0 0 0··· 
 
.. .. .. .. .. ..
. . . . . .···
and is often represented by the Ferrers diagram:

.
Shape matrices and tableaux 151

A bar shape looks like:


 
0 0 0 0 0 0···
 0 0 0 0 0 0··· 
 
 0 0 0 0 0 0··· 
 
(B)  0 1 1 1 0 0··· ,
 
 0 0 0 0 0 0··· 
 
.. .. .. .. .. ..
. . . . . .···
and an almost skew-shape looks like:
 
0 0 0 1 1 1 1 0···
 0 0 1 1 1 1 1 0··· 
 
 0 1 1 1 1 1 0 0··· 
 
(AS)  0 0 1 1 1 0 0 0··· ,
 
 0 0 0 0 0 0 0 0··· 
 
.. .. .. .. .. .. .. ..
. . . . . . . .···
and would be represented by the Ferrers diagram:

Note that the diagram ignores the fact that the left-most column of the shape
matrix consists completely of zeroes. Also notice that the empty diagram
corresponds to the zero matrix.
The shapes illustrated above are those that we will have most to do with, but
clearly we can associate to any shape matrix, A = (aij ), a Ferrers diagram, or
simply diagram: we just set up a grid equal to the effective size of the matrix
(say, s × t), and throw away the boxes whose entries are equal to zero. That is,
the (i, j)th box lies in the diagram if and only if aij = 1. If A is the shape matrix,
we will sometimes denote by (A) its corresponding diagram.
The partition shape has a uniquely associated partition, namely the sequence
λ = (λ1 , . . . , λs , . . .) where λi is the non-negative integer equal to the number
of ones in row i. Clearly, the sequence is decreasing: λ1 ≥ λ2 ≥ · · · ≥ λs ≥ · · · .
We say the length of λ is s if s is the smallest non-negative integer such that
λs+t = 0 for every positive integer t.
The skew-shape has two partitions uniquely associated to it, namely, λ =
(λ1 , . . . , λs , . . .) and µ = (µ1 , . . . , µt , . . .), with µi ≤ λi for all i, and such that
the length of µ is strictly less than that of λ. One then thinks of the shape as
the result of removing from the shape of λ the subshape corresponding to µ. In
fact, the notation most often used for a skew-shape is λ/µ. We say the length
of λ/µ is the length of λ. If one removes the condition that the length of µ be
152 Weyl and Schur modules

strictly less than that of λ, then we have other pairs of partitions (λ , µ ) that
will yield the same diagram. In that case, we still use the same notation, λ /µ
(but the length of λ /µ stays equal to the length of the previous λ).
Finally, we see that the almost skew-shape would be a skew-shape but for its
last row, which, rather than projecting beyond (or flush with) the penultimate
row, does not make it out that far to the left. In short, it would be a skew-
partition but for that inadequacy in the last row.
In our examples above, the partition λ associated to (P ) is (4, 3, 3, 2); the
pair of partitions associated to (S) are λ = (5, 4, 3, 2) and µ = (2, 2, 1). (For
convenience we have eliminated the zeroes to the right in our notation.) For (AS),
we might take the skew-partition to be (7, 7, 6, 5)/(3, 2, 1) with a bar having the
entries (1, 1) in the fourth row, or we might take (7, 7, 6, 5)/(3, 2, 1, 1), with a bar
having entries (0, 1) in the fourth row.
Another way to denote an almost skew-shape, which closely parallels the nota-
tion for a skew-shape is to first define an almost partition to be a sequence
(µ1 , . . . , µn ) such that µ1 ≥ · · · ≥ µn−1 and 0 ≤ µn ≤ µ1 . We then can denote
(not necessarily uniquely) an almost skew-shape by λ/µ, where λ is a partition
of length n, and µ is an almost partition having exactly n terms and satisfying
µi ≤ λi for all i. We can go one step further, and say that an almost partition,
µ, is of type n − (i + 1) if i is the largest integer less than n such that µn ≤ µi ,
and we say that the type of the almost skew-shape λ/µ is equal to the type of
µ. Clearly, the type is independent of the choice of λ and µ used to describe the
almost skew-shape. The choice of the pair (λ, µ) can be made canonical if, in the
case of type zero, we choose µn = 0, while for type greater than zero, we choose
µn−1 = 0. We define the length of the almost skew-shape to be the length of
the canonical partition, λ.
With this terminology, we see that an almost skew-shape of type 0 is a
skew-shape, and that for almost skew-shapes of length n, we can have types
0, 1, . . . , n− 2. In particular, almost skew-shapes of length 2 are necessarily skew-
shapes; for length 3, there are only skew-shapes and almost skew-shapes of type
1, and so on.
We spoke of shape matrices as infinite matrices in order not to have to specify
last row or column when we talked about subshapes. However, we see that a
shape matrix whose last row is row s and whose last column is column t, can
be thought of as an s × t-matrix; when we draw diagrams or shapes, we will
generally avoid the dots that we were forced to put into the illustrations above.
If we have two shape matrices A and B, with B ⊆ A, we will assume that they
are both s × t-matrices, simply by augmenting where necessary by zeroes.

Remark VI.1.2 Three immediate observations should be made here:

 is also
1. If A is a partition (or skew-partition) matrix, then its transpose, A,
a partition (or skew-partition) matrix.
Shape matrices and tableaux 153

2. If A is a partition matrix with associated partition λ, then we denote the



 by λ.
partition associated to A
3. We see that if A is a skew-partition matrix with associated partitions λ and
µ, then the matrix A  and µ
 has associated partitions λ .

The notation and terminology used above differ slightly from those used else-
where. For instance, in I. Macdonald’s book [66], the transpose is called the
“conjugate,” and is denoted by A ; in Reference [3], it is called the “dual,” but
denoted in the way it is above.

Definition VI.1.3 We introduce some standard terminology for shapes, and


partitions in particular:

1. The weight of a shape matrix A = (aij ) is aij , and is denoted by |A|.
 
2. If λ = (λ1 , . . . , λn ) and λ = (λ1 , . . . , λm ) are two partitions, we say λ ≥ λ
if either λ = λ or if for some i, λj = λj for all j < i, and λi > λi .

VI.1.2 Tableaux
Definition VI.1.4 Let A be a shape matrix and S a set. A tableau, T , of
shape A with values in S is a filling-in of the diagram (A) by elements of S.
We denote the tableau by the ordered pair T = ((A); τ ), where τ is the filling-in
of (A) by S.

We could have said that τ is a function from (A) to S, but for the fact that
we have not given a formal enough definition of “diagram” to do this. But if
one regards the diagram as a collection of cells, then τ would be a function with
domain (A). In most cases, we will simply refer to the tableau as T , with the set
S an understood ordered basis of a finitely generated free module. This is how
we saw tableaux used in earlier chapters.
Occasionally we will use the term row tableau; this is simply a tableau the
diagram of whose shape consists of one row.
For further work with tableaux, it will be convenient to have a quasi order on
tableaux with values in a totally ordered set. Suppose, then, that S is a totally
ordered set, say S = {s1 , . . . , sn } with s1 < · · · < sn , and suppose T is a tableau
with values in S. Define Tij to be the number of elements in {s1 . . . , si } that
appear in at least one of the first j rows of the diagram of T . Now suppose that
T  is another tableau.

Definition VI.1.5 We say that T  ≤ T if Tij ≥ Tij for every i and j. We say
that T  < T if T  ≤ T and for some i, j we have Tij > Tij .

To see that this is a quasi order and not an order, consider our set S with
three elements: S = {s1 , s2 , s3 }, with s1 < s2 < s3 , and consider the diagram
154 Weyl and Schur modules

corresponding to the partition λ = (4, 2, 1). Then the two tableaux

s1 s2 s2 s3
T = s2 s3
s3

and

s2 s3 s2 s1
T  = s3 s2
s3

are such that T ≤ T  and T  ≤ T , but T and T  are clearly not equal. However,
the “≤” relation is both reflexive and transitive, as can easily be checked.

VI.2 Weyl and Schur modules associated to shape matrices


To each finite free module, F, over a commutative ring, R, and each shape matrix
we will associate two maps, a Weyl map and a Schur map, whose images will be
called the Weyl and Schur modules of that shape. To do that, we first look at
some auxiliary ideas.
If a = (a1 , . . . , al ) is a sequence of non-negative integers, let α = a1 + · · · + al .

We define the maps δa : Dα F → Da1 F ⊗ · · · ⊗ Dal F and δa : Λα F → Λa1 F ⊗
· · · ⊗ Λ F to be the diagonalization maps of the indicated divided and exterior
al

powers of F into the indicated tensor products. We define the maps µa : Λa1 F ⊗

· · · ⊗ Λal F → Λα F and µa : Sa1 F ⊗ · · · ⊗ Sal F → Sα F to be the multiplication
maps from the indicated tensor products of exterior and symmetric powers to
the indicated exterior and symmetric powers.

Definition VI.2.1 (Weyl and Schur maps) Let F be a free module over the
commutative ring, R. For the s×t shape matrix A = (aij ), set ri = (ai1 , . . . , ait ),
t s
cj = (a1j , . . . , asj ), ρi = j=1 aij , γj = i=1 aij . The Weyl map associated
to A, ωA , is the map

ωA : Dρ1 F ⊗ · · · ⊗ Dρs F → Λγ1 F ⊗ · · · ⊗ Λγt F

defined as the composition


   
ωA = µc1 ⊗ · · · ⊗ µct θW δr 1 ⊗ · · · ⊗ δr s

where, since all the aij are equal to zero or one, we have identified Daij F with
Λaij F for all i, j; the map θW is the isomorphism comprising all of these iden-
tifications together with rearrangement of the factors. Pictorially what we have
Modules associated to shape matrices 155

is the following:
 

 Da11 F ⊗ · · · ⊗ Da1t F 


 


 ⊗ 

Dρ1 F ⊗ · · · ⊗ Dρs F → .
.
θW
−→
 . 

 


 ⊗ 

 
Das1 F ⊗ · · · ⊗ Dast F
 a 

 Λ 11 F ⊗ · · · ⊗ Λa1t F 


 


 ⊗ 

.. → Λγ1 F ⊗ · · · ⊗ Λγt F.
 . 

 


 ⊗ 

 as1 
Λ F ⊗ · · · ⊗ Λast F
The Schur map associated to A, σA , is the map
σA : Λρ1 F ⊗ · · · ⊗ Λρs F → Sγ1 F ⊗ · · · ⊗ Sγt F
defined as the composition
  
   

σA = µc1 ⊗ · · · ⊗ µct θS δr1 ⊗ · · · ⊗ δrs

where, since all the aij are zero or one, we have identified Λaij F with Saij F
for all i, j; the map θS is the isomorphism comprising all of these identifications
together with rearrangement of the factors. We can view the definition of the
Schur map “pictorially” in the same way we did the Weyl map.
Definition VI.2.2 (Weyl and Schur modules) Let F be a free R-
module, and A a shape matrix. We define the Weyl module of F associated
to A, denoted KA F , to be the image of ωA . We define the Schur module of
F associated to A, denoted LA F, to be the image of σA .
Remark VI.2.3 The following observations are easy to check and very
useful.
1. If A is a non-zero shape matrix with its initial column consisting only of
zeros, and B is the shape matrix with that initial column removed, it is clear
that the Weyl and Schur maps associated to A and B are the same. Hence,
we will generally assume that our shape matrices have at least one entry in
the first column equal to one.
2. If A is a shape matrix, and B is the shape matrix obtained from A by
a permutation of its rows (columns), then the associated Weyl and Schur
modules of these matrices are isomorphic.
With the definition of Weyl and Schur modules to hand, a natural question to
consider is whether these modules are free over the ground ring and, if so, how
can we describe a basis. For example, if we take a one-rowed partition λ, what
156 Weyl and Schur modules

are the Weyl and Schur modules associated to it? The Weyl module is the image
of the map Dλ F →ωλ Λ1 F ⊗ · · · ⊗ Λ1 F , where ωλ is the diagonalization map.
% &# $
λ
Clearly, then, the image is isomorphic to Dλ F itself, that is, Kλ F = Dλ F . In a
similar way we can show that Lλ F = Λλ F . In both of these cases, the modules
are clearly free R-modules, and we have a very concrete description of bases for
them. Not only do we have explicit descriptions of their bases, we even have a
description in terms of tableaux. In the case of Dλ F , a basis is parametrizable by
the set of all one-rowed tableaux: xi1 xi2 · · · xiλ where {x1 , . . . , xm } is an
ordered basis of the free module, F , and i1 ≤ · · · ≤ iλ . In the case of Λλ F , we have
that a basis is parametrizable by all one-rowed tableaux: xi1 xi2 · · · xiλ
where i1 < · · · < iλ .
It will be proven later that if λ/µ is an almost skew-partition, then Kλ/µ and
Lλ/µ are free, and bases for them can be parametrized by certain sets of tableaux
of shape λ/µ satisfying combinatorial conditions.

VI.3 Letter-place algebra


We have run up, and will run up, against tensor products of divided powers,
exterior powers and symmetric powers frequently. A tool that has proven helpful
in dealing with these kinds of terms is the so-called letter-place algebra. In this
section we will define this algebra, and develop some important combinatorial
properties of it. Our treatment will be a little less general than that given in
Reference [81]; the interested reader may go to that reference to see how multi-
signed alphabets are treated in a uniform and general way. We will deal with
the divided powers case in some detail, and then we will quickly treat the cases
of exterior powers and symmetric powers. Most of the proofs will be found in
Appendix A.

VI.3.1 Positive places and the divided power algebra


Usually we are given a fixed number, say n, of terms in the tensor product:
Dk1 (F ) ⊗ · · · ⊗ Dkn (F ), where F is a free module. We intuitively look at such
a product and know which is the first factor, the second, and so on. The idea
behind the letter-place approach is to clearly designate the places that the terms
in the product are actually in. As an example of what we mean, suppose that
x ∈ Dki (F ), and we want to write the element 1 ⊗ · · · ⊗ x ⊗ · · · ⊗ 1 in the tensor
% &# $
i
product above. The letter-place algebra will allow us to write this element as
(x|i(ki ) ). How this will help besides just shortening the amount we have to type
and the space it takes to type it will become evident as we develop and use this
approach.
Just as with the symmetric and exterior algebras, we have that D(F ⊕ G) =
D(F ) ⊗ D(G); it is, after all, the graded dual of the symmetric algebra. So,
Letter-place algebra 157

if we take D(F ⊗ Rn ) ∼
= D(F ⊕ · · · ⊕ F ), we see that D(F ⊗ Rn ) is equal to
% &# $
n
D(F ) ⊗ · · · ⊗ D(F ). This is natural with respect to the action of GL(F ), but
% &# $
n times
clearly not with respect to the action of GL(Rn ). In fact, we moved to the
notation Rn rather than G to indicate that we have made a choice of basis in
our free module, G. We can, though, use G in our preliminary discussion and,
assuming that the rank of this free module is n, still see that “in some way”,
D(F ⊗ G) ∼= D(F ) ⊗ · · · ⊗ D(F ). Now we want to introduce convenient notation
% &# $
n
to exhibit this isomorphism, as well as to get to the letter-place conventions.
To this end, let us suppose that G has the (ordered) basis, {y1 , . . . , yn } with
y1 < · · · < yn , and for any x ∈ D1 (F ), let us denote by (x|yi ) the element x ⊗ yi ,
(k)
and by (x(k) |yi ) the element corresponding in D(F ) ⊗ · · · ⊗ D(F ) to (x|yi )(k) ,
% &# $
n
that is, to the element in that n-fold tensor product of D(F ) having x(k) in the
ith factor.
• The picture to keep in mind is: (x|yi ) is the element 1 ⊗ · · · ⊗ x ⊗ · · · ⊗ 1.
% &# $
i

Now
k!(x ⊗ yi )(k) = (x ⊗ yi )k = (1 ⊗ · · · ⊗ x ⊗ · · · ⊗ 1)k
% &# $
i

= 1 ⊗ · · · ⊗ x ⊗ · · · ⊗ 1 = k! (1 ⊗ · · · ⊗ x(k) ⊗ · · · ⊗ 1),
k
% &# $ % &# $
i i
(k)
so that the above definition of (x(k) |yi ) makes sense.
Finally, if l = l1 + · · · + ln , and x ∈ Dl (F ), we set
(l )
 (l )
(x|y1 1 · · · yn(ln ) ) = (x(l1 )|y1 1 ) · · · (x(ln )|yn(ln ) ),

where x(l1 )⊗· · ·⊗x(ln ) indicates the image of the diagonal map into Dl1 (F )⊗
· · · ⊗ Dln (F ) applied to our element x.
Remark VI.3.1 The identities and conventions that we adopt for our discus-
sion are those that are clearly valid if one works over the ring of integers (as is
the case illustrated above, where we have cancelled k! because there is no torsion
over the integers). We will continue to do this in our treatment of the letter-
place algebra and all other structures that are transportable from Z to arbitrary
commutative base rings.
A simple illustration, just to fix our ideas, is this: Suppose x1 , x2 and x3 are
in D1 (F ), and consider the element
(2) (2)
(x1 x2 x3 |y1 y2 y3 ) ∈ D2 (F ) ⊗ D1 (F ) ⊗ D1 (F ).
158 Weyl and Schur modules

Then this element is equal to

(2) (2)
(x1 x2 |y1 )(x2 |y2 )(x3 |y3 ) + (x1 x2 |y1 )(x3 |y2 )(x2 |y3 )+
(2) (2) (2)
(x1 x3 |y1 )(x2 |y2 )(x2 |y3 ) + (x2 |y1 )(x1 |y2 )(x3 |y3 )+
(2) (2) (2)
(x2 |y1 )(x3 |y2 )(x1 |y3 ) + (x2 x3 |y1 )(x1 |y2 )(x2 |y3 )+
(2)
(x2 x3 |y1 )(x2 |y2 )(x1 |y3 ).

(a ) (a )
nthe symbol (w|y1 · · · yn ) equal to zero if the degree
1 n
(♠) We agree to set
of w is not equal to i=1 ai . The element w is supposed to be a homogeneous
element of D(F ).
The letter-place notation we have been using above lends itself very naturally
to writing tableaux. That is, suppose we wanted to write the product of the
(2) (2) (2)
above element, (x1 x2 x3 |y1 y2 y3 ) with, say, (x3 x1 |y1 y2 y3 ). As we saw above,
each of these terms is a sum of a number of addends, so that the notation we
have for each of these terms is already of considerable convenience. But now,
instead of using juxtaposition to denote the product of these two terms, let us
use “double tableau” notation, that is, let us write
 
(2) (2)
x1 x2 x3 y1 y2 y3
(2)
x3 x1 y1 y2 y3

for this product.


Suppose that we choose an ordered basis for F , say {x1 , . . . , xm } with x1 <
· · · < xm , and let us say that the elements xi above are among these basis
elements. Then the double tableau above does not change value if we write it as:
 
(2) (2)
x1 x2 x3 y1 y2 y3
(DT ) (2) .
x1 x3 y1 y2 y3

We point this out to indicate that we may always assume that our tableaux
are given in such a way that in each row, the elements are increasing. The
terminology for this is that the tableaux are row-standard (a notion that we
have already encountered in Chapter III). Recall that in Chapter III, we wrote
out the rows of the tableau repeating letters instead of using divided powers. This
helps to talk about the columns of a tableau; for instance, the tableau above has
two rows and four columns (the number refers to the arrays in the letters as well
as the places).
Usually, we call the basis of F the letters, while the basis of G is called the
places. A basic word of degree k is simply a basis element of Dk (F ), while
a word of degree k is a linear combination of basic words of degree k. Usually
Letter-place algebra 159

we will write a word as w, and we will write a general double tableau as


 
w1 1(a11 ) 2(a21 ) 3(a31 ) · · ·
 w2 1(a12 ) 2(a22 ) 3(a32 ) · · · 
(G)   ···

··· ··· ··· ··· 
wn 1(a1n ) 2(a2n ) 3(a3n ) · · ·
where αi = (a1i + a2i + a3i + · · · ) ≥ αj for 1 ≤ i < j ≤ n, and we have written i
for yi . We will continue to write i for yi as long as there is no danger of confusion.
Also, in most cases, the words wi will be basic words, in which case (since they
are basis elements of Dk (F )), they are increasing. Because of our convention (♠)
above, we see that we may assume that the degree of the element wi is equal to
αi . Note that our tableau is an element of Dk1 (F ) ⊗ · · · ⊗ Dkn (F ) when, for each
j = 1, . . . , n, l ajl = kj .
We will call a double tableau standard if the words wi are basic, the lengths
of the rows are decreasing (from the top), it is row-standard, and also column-
standard in the sense that when we have used repeat notation instead of divided
powers, the columns are strictly increasing from top to bottom. Our double
tableau (DT ) above is not a standard double tableau; if we replace the element
x1 in the second row by x2 , however, it will be standard.
Clearly there is a set of double tableaux that are a basis for Dk1 (F ) ⊗ · · · ⊗
Dkn (F ), namely:
 
w1 1(k1 )
 w2 2(k2 ) 
(W )   ···

··· 
wn n(kn )
where the wi run through the basis elements of Dki (F ). But these tableaux are
not in general standard. Even if it were the case that k1 ≥ · · · ≥ kn , so that
the “place” side of the tableau were standard, the “word” side of the tableau
would in general not be so. And if we had to reorder the rows so that they were
decreasing in length, we would upset standardness in the column of places.
What we do have is the following theorem.
Theorem VI.3.2 The set of standard double tableaux having the ith place
counted ki times is a basis for Dk1 (F ) ⊗ · · · ⊗ Dkn (F ).

The proof breaks up into two parts: the double tableaux of type (G), with
l ajl = kj , generate, and the number of such tableaux is equal to the number
of tableaux of type (W ) above, for fixed k1 , . . . , kn . The first part is given in
Appendix A, Section 1, and the second part in Appendix A, Section 2.

VI.3.2 Negative places and the exterior algebra


Now we take up the letter-place approach to the tensor product of a fixed number
of copies of ΛF for a fixed free module, F . As in the previous discussion, we
160 Weyl and Schur modules

use the fact that Λ(F ⊗ Rn ) ∼


= Λ(F ⊕ · · · ⊕ F ), which is, in turn, isomorphic
% &# $
n
to ΛF ⊗ · · · ⊗ ΛF . There are two natural ways to proceed with this discussion
% &# $
n
from a letter-place point of view: we could make the letters of F be positive and
the places of Rn negative, or vice versa. We will deal with the first case in some
detail, and indicate the necessary changes if we reverse sign.
Take the basis of Rn to be {1, . . . , n}, but this time we will treat them as
“negative” places (in fact, we have written them in bold face to distinguish them
from the “positive” places of the previous subsection). To make the meaning of
this clearer (if not altogether clear), we can think of the bases of our free modules
as “alphabets” from which we make “words” by stringing them together (as we
have been doing). But we can also think of the letters of our alphabet as being
“signed,” that is, either positive or negative. In the preceding discussion of tensor
products of divided powers, all of our letters and places were positive, so that
we can assign the number 0 to all of them (to indicate that they are positive).
However, in this case, we want to consider the basis elements of F as positive,
while those of Rn as negative. So, we assign the value 0 to the basis elements of F ,
and we assign the value 1 to the basis elements 1, . . . , n to indicate that they are
negative. In general, if you have signed alphabets A and B which are the bases
of A and B, respectively, then the element a ⊗ b ∈ A ⊗ B is assigned the value
|a ⊗ b| = |a| + |b| mod 2, where |x| stands for the sign of x. Of course, we will
write the element a ⊗ b as (a|b) when we adopt the “letter-place” language as we
did in the foregoing subsection.
As before, then, we will write the element (x|i) to stand for the element x ⊗ i ∈
Λ(F ⊗ Rn ), where x is a basis element of F . We think of this, under the identi-
fications made above, as the element 1 ⊗ · · · ⊗ x ⊗1 ⊗ · · · ⊗ 1 ∈ ΛF ⊗ · · · ⊗ ΛF .
% &# $ % &# $
i n
Since x has sign 0 and i has sign 1, the sign of (x|i) is 0 + 1 = 1. From the
identifications we have made, we see that (x|i)(y|i) = −(y|i)(x|i). This, and the
commutativity of multiplication in the case of divided powers is consistent with
the sign convention:
(a1 |b)(a2 |b) = (−1)|(a1 |b)||(a2 |b)| (a2 |b)(a1 |b).
Our object is to work toward the same sort of double tableau notation for this
tensor product that we had earlier. But before it was possible to take a positive
place, i, say, and consider the element i(2) as in (xy|i(2) ). In this case, since a
place i is negative, we see that i(2) = 0 (recall Remark VI.3.1), so we have to
define what we mean by the element (w|p1 ∧ · · · ∧ pk ) where w is an element
(word) of a basis of Dk F , and p1 , . . . , pk are distinct basis elements of Rn (so
that p1 ∧ · · · ∧ pk is plus or minus a basis element of Λk Rn ).
(k ) (k )
Suppose that w = a1 1 · · · al l , let k = k1 + · · · + kl and let b1 , . . . , bk be the
sequence a1 , . . . , a1 , . . . , al , . . . , al . Let Sk1 ,...,kl denote the Young subgroup of the
% &# $ % &# $
k1 kl
Letter-place algebra 161

symmetric group Sk consisting of those permutations that permute the first k1


elements of 1, . . . , k among themselves, the next k2 elements among themselves,
and so on. (This is a subgroup isomorphic to Sk1 ×· · ·×Skl consisting of k1 ! · · · kl !
elements.) Then we define

(F I) (w|p1 ∧ · · · ∧ pk ) = (bσ(1) |p1 ) · · · (bσ(k) |pk ),
σ

where σ runs through representatives of distinct cosets of Sk /Sk1 ,...,kl .


In the summation above we have written the product in our exterior algebras
as simple juxtaposition instead of using wedges. We do this to conserve a uniform
notation for multiplication in the letter-place algebra, in which (as we will see
later) letters and places may sometimes be positive and sometimes negative.
A few simple examples will make this completely clear.
• Consider (a(2) |p1 ∧ p2 ). We have

(a(2) |p1 ∧ p2 ) = (a|p1 )(a|p2 ).

• Consider (a1 a2 |p1 ∧ p2 ). We have

(a1 a2 |p1 ∧ p2 ) = (a1 |p1 )(a2 |p2 ) + (a2 |p1 )(a1 |p2 ).

• Consider (a(2) b(3) |p1 ∧ · · · ∧ p5 ). We have

(a(2) b(3) |p1 ∧ · · · ∧ p5 ) = (a|p1 )(a|p2 )(b|p3 )(b|p4 )(b|p5 )


+ (a|p1 )(b|p2 )(a|p3 )(b|p4 )(b|p5 )
+ (a|p1 )(b|p2 )(b|p3 )(a|p4 )(b|p5 )
+ (a|p1 )(b|p2 )(b|p3 )(b|p4 )(a|p5 )
+ (b|p1 )(a|p2 )(a|p3 )(b|p4 )(b|p5 )
+ (b|p1 )(a|p2 )(b|p3 )(a|p4 )(b|p5 )
+ (b|p1 )(a|p2 )(b|p3 )(b|p4 )(a|p5 )
+ (b|p1 )(b|p2 )(a|p3 )(a|p4 )(b|p5 )
+ (b|p1 )(b|p2 )(a|p3 )(b|p4 )(a|p5 )
+ (b|p1 )(b|p2 )(b|p3 )(a|p4 )(a|p5 ),
in other words, the 10 = 5!/2!3! terms that correspond to the ten distinct
cosets of S5 /S2,3 .
An important observation to make here, as we made before, when we use
(k ) (k )
tableau notation as above (i.e.: (w|p1 ∧ · · · ∧ pk )), where w = a1 1 · · · al l , the
word w is entered as our row-array b1 . . . bk .
With this notation taken care of, we can consider double tableaux as we did
earlier, but now the left side consists of words in the positive alphabet which is
162 Weyl and Schur modules

the basis of F , and the left side consists of words in the negative alphabet of
places, the basis {1, . . . , n} of Rn .
To see that our usual basis elements of Λk1 F ⊗ · · · ⊗ Λkn F can be expressed
as double tableaux, consider the following example.

• The element x2 ∧ x3 ∧ x5 ⊗ x1 ∧ x3 ⊗ x2 ∧ x4 ⊗ x3 ∧ x5 ∧ x6 ∈ Λ3 F ⊗ Λ2 F ⊗
Λ2 F ⊗ Λ3 F can be expressed as the double tableau:
 (2) 
x2 1 3
 (3) 
 x3 1 2 4 
 (2) 
 x 1 4 
 5 
 x1 2 
 
 x4 3 
x6 4

On the right hand side of the tableau we have omitted the wedge, and simply
spread the basis elements out along the row. We used the divided power
notation on the left hand of the column to simplify writing. Really, the top
row of the tableau above should look like:

(x2 x2 |1 3).

In our situation, we see that if we interchange rows of the tableau, we must


take sign into account. For example:
 (2)   (2) 
x2 1 3 x2 1 3
 (3)   (3) 
 x3 1 2 4   x3 1 2 4 
 (2)   (2) 
 x 1 4   
 5  = −  x5 1 4 
 x1 2   x1 2 
   
 x4 3   x6 4 
x6 4 x4 3

As in the previous case, we now have to define what we mean by a double


standard tableau. We will call a double tableau standard if it is standard in the
old sense on the left hand side of the vertical column, but on the right hand side,
has the property that it is strictly increasing in the rows, and weakly increasing
in the columns.
Notice that this definition implies that the shape of the tableau is that of a
partition. For instance, the tableau of the above example is very non-standard
for all the possible reasons: the shape is not that of a partition (i.e., the lengths
of the rows are not decreasing from top to bottom); on the left side the columns
are not strictly increasing; nor are they on the right side. However, both sides of
the tableau are row-standard: on the left side, the rows are weakly increasing,
while on the right they are strictly increasing. These are the requisite conditions
for row-standard when we are talking about positive and negative alphabets.
Letter-place algebra 163

An example of a standard double tableau would be the following:


 
(2)
x3 x4 1 2 3
 
E :  x(2)5 1 3 ,
x6 3

where we are again cheating a bit in that we are writing the left hand side in
terms of divided powers rather than as spread-out sequences. Just to make sure
that we understand what all the notation means, we remark that the tableau
above is equal to:

E = x3 ∧ x5 ⊗ x3 ⊗ x4 ∧ x5 ∧ x6
+ x3 ∧ x5 ⊗ x4 ⊗ x3 ∧ x5 ∧ x6
+ x4 ∧ x5 ⊗ x3 ⊗ x3 ∧ x5 ∧ x6 .

In the same way the double standard tableaux generate the tensor product of
divided powers, these double standard tableaux generate the tensor product of
exterior powers. We have the following theorem.

Theorem VI.3.3 The set of standard double tableaux having the ith place
counted ki times is a basis for Λk1 F ⊗ · · · ⊗ Λkn F .

A sketch of the proof of this theorem is in Section 3 of Appendix A.


It should be fairly clear that the discussion above could just as well have been
carried out if we had assumed that the alphabet for F were signed negatively, and
that for the places signed positively. In that case, we would have simply written
the basis elements of F in boldface, and those for Rn in ordinary typeface. There
are one or two differences that we would have to remark in this case. One is that
we would set (x1 |i)(x2 |i) = (x1 ∧ x2 |i(2) ). Another is that we would modify the
fundamental identity (F I) earlier in this section as follows. If w were the word
(k ) (k )
w = x1 ∧ · · · ∧ xk , and we had p = p1 1 · · · pl l with k = k1 + · · · + kl , then
setting {q1 , . . . , qk } equal to the sequence {p1 , . . . , p1 , . . . , pl , . . . , pl }, we define
% &# $ % &# $
k1 kl


(F I) (w|p) = (x1 |qσ(1) ) · · · (xk |qσ(k) ),
σ

where σ ranges over representatives of the cosets of the appropriate Young


subgroup.
However, if we prefer to keep the places in their original order (so that we could
write the right-hand side of the term as a product of divided powers, say), then
the order of the terms q1 , . . . , qk would be changed, and that would introduce a
164 Weyl and Schur modules

sign. For example, if we took (x1 ∧ x2 ∧ x3 |1(2) 2), we would get


(x1 ∧ x2 ∧ x3 |1(2) 2) = (x1 |1)(x2 |1)(x3 |2)
− (x1 |1)(x3 |1)(x2 |2)
+ (x2 |1)(x3 |1)(x1 |2).
For “standardness” of double tableaux we would have strictly increasing rows
in the letters, weakly increasing rows in the places; weakly increasing columns in
the letters, strictly increasing columns in the places. The proof that these double
standard tableaux form a basis is indicated in Section 3 of Appendix A.

VI.3.3 The symmetric algebra (or negative letters and places)


There is one last canonical algebra to consider, namely the tensor product of a
fixed number of copies of the symmetric algebra of F : S(F ) ⊗ · · · ⊗ S(F ). In
% &# $
n
this case, we consider the basis elements of both F and Rn negative (but we
will not bother to write these in boldface). Thus, an element (x|i) will have sign
0, and hence will always be positive (a not surprising development given that
we are treating the symmetric algebra). Consequently, we write (x1 |i)(x2 |i) =
(x2 |i)(x1 |i) and for w = l1 ∧ · · · ∧ lk , p = p1 ∧ · · · ∧ pk , we define

(F I) (w|p) = (x1 |pσ(1) ) · · · (xk |pσ(k) ).
σ

where now the σ runs over all permutations of {1, . . . , k}.


With this identity, we get the usual basis theorem for the n-fold tensor product
of the symmetric algebra:
Theorem VI.3.4 The set of standard double tableaux having the ith place
counted ki times is a basis for Sk1 F ⊗ · · · ⊗ Skn F .
The reader is again referred to Section 3 of Appendix A for a brief indication of
the proof.

VI.3.4 Putting it all together


We now want to put these various pieces together, and consider what happens
when we have “letter alphabets” and “place alphabets” that contain both posit-
ive and negative elements. To use more descriptive notation, we will let L and P
stand for the letter and place alphabets, respectively. Further, we will suppose
that L = L+ "L− and P = P + "P − , where the plus and minus superscripts indic-
ate the signs of the elements of these alphabets. If we now let L+ , L− , P + , P −
stand for the free modules generated by these alphabets (or bases), we may
consider what is called the letter-place superalgebra:
S(L|P) = Λ(L+ ⊗ P − ) ⊗ Λ(L− ⊗ P + ) ⊗ D(L+ ⊗ P + ) ⊗ S(L− ⊗ P − ).
Polarization maps and Capelli identities 165

The individual factors of the tensor product above have been described in
detail; the product of two terms from different components of the product
is simply the tensor product of these terms, while the product (l1 |p)(l2 |p) =
(−1)|(l1 |p)||(l2 |p)| (l2 |p)(l1 |p).

VI.4 Place polarization maps and Capelli identities


In Section VI.2 we defined the Weyl and Schur maps, which entailed a good
deal of diagonalization, identification and multiplication from a tensor product
of divided (exterior) powers to a tensor product of exterior (symmetric) powers.
We now know that these tensor products of various powers can be expressed in
letter-place terms, and we may ask if these complicated maps may be viewed in
a different way (hopefully, a simpler way) using the letter-place approach. The
answer, as was no doubt anticipated, is yes, and the method will be that of place
polarizations.
In this section, we will consider two types of maps, both of which are called
place polarizations: those from “positive places to positive places” and those
“from positive places to negative places”.
Definition VI.4.1 Let q ∈ P + , s ∈ P, s = q, and let (l|p) be a basis element
of S(L|P). Define the place polarization, ∂s,q , to be the unique derivation on
S(L|P) defined by
∂s,q (l|p) = δq,p · (l|s),
where δq,p is the Kronecker delta.
When we say that this map is a derivation on S(L|P), we mean that it has
the property
∂s,q {(l1 |p1 )(l2 |p2 )} = {∂s,q (l1 |p1 )}(l2 |p2 ) + (−1)|s||p1 | (l1 |p1 )∂s,q (l2 |p2 ).
2
A straightforward calculation shows that if s is a negative place, then ∂s,q = 0.
On the other hand, we can see easily that if s is positive, ∂s,q {(l1 |q)(l2 |q)} =
2

2{(l1 |s)(l2 |s)}, so that for q and s positive places, it makes sense to talk about
(k)
the higher divided powers of the place polarizations, ∂s,q , namely ∂s,q . In the
case of the divided square just discussed, for instance, we see that the equation
(2)
may be interpreted as ∂s,q (l1 l2 |q (2) ) = (l1 l2 |s(2) ). In general, then, we have
(k)
∂s,q (w|q (m) ) = (w|q (m−k) s(k) ),
where q and s are positive places.
One fundamental identity, which is easy to prove, is the following.
Fact 1 Let p, q, r be places with q and p positive, and consider the place
polarizations ∂r,q , ∂q,p and ∂r,p . Then
∂r,p = ∂r,q ∂q,p − ∂q,p ∂r,q .
166 Weyl and Schur modules

In short: ∂r,p is the commutator of ∂q,p and ∂r,q .

Proof It is clearly enough to consider these maps on a term of the form


(w|p(a) q (b) r(c) . . .). If r is negative, then c can be 0 or 1, while if r is positive,
there is no restriction on c. Let us suppose first that r is positive. We have

∂r,p (w|p(a) q (b) r(c) . . .) = (w|p(a−1) q (b) rr(c) . . .) = (c + 1)(w|p(a−1) q (b) r(c+1) . . .).

But we have

∂r,q ∂q,p (w|p(a) q (b) r(c) . . .) = (b + 1)∂r,q (w|p(a−1) q (b+1) r(c) . . .)


= (b + 1)(c + 1)(w|p(a−1) q (b) r(c+1) . . .),

and

∂q,p ∂r,q (w|p(a) q (b) r(c) . . .) = (c + 1)∂q,p (w|p(a) q (b−1) r(c+1) . . .)


= b(c + 1)(w|p(a−1) q (b) r(c+1) . . .).

Taking the difference, we get the desired result.


If we assume that r is negative, and c = 1, we get zero when we apply the maps.
On the other hand, if c = 0, the proof proceeds as it did for positive r. 2

We start with the case of positive-to-positive place polarizations.


Assume that our places p, q and r are all positive. Then as we know, we can
form the divided powers of all of the place polarizations involving these places,
and ask if there are identities associated to these that generalize the basic identity
proved above.

Proposition VI.4.2 (Capelli Identities) Let p, q, r be places with p, q and


r all positive, and consider the place polarizations ∂r,q , ∂q,p and ∂r,p . Then

(a) (b) (b−k) (a−k) (k)
(Cap) ∂r,q ∂q,p = ∂q,p ∂r,q ∂r,p ;
k≥0


(Cap ) (b) (a)
∂q,p ∂r,q = (a−k) (b−k) (k)
(−1)k ∂r,q ∂q,p ∂r,p .
k≥0

Proof For a = b = 1, this is just the commutation rule (for both Cap and
Cap ). One can now proceed, by direct calculation, or we can employ a standard
trick and prove this by induction, first on a for b = 1, and then for all a and
b. The standard trick is to behave as though we were working over the field of
rational numbers (which allows us to divide by integers), and then assert that
since our proposition works there, it works in general (see Remark VI.3.1).
We will therefore proceed to prove our proposition by induction. As we pointed
out, our proposition is true for a = b = 1, so we may let b = 1 and assume the
Weyl and Schur maps revisited 167

identity true for a − 1. We then see that


(a)
a∂r,q (a−1)
∂q,p = ∂r,q (a−1)
∂r,q ∂q,p = ∂r,q {∂q,p ∂r,q + ∂r,p }
(a−1) (a−1)
= ∂r,q ∂q,p ∂r,q + ∂r,q ∂r,p
= {∂q,p ∂r,q
(a−1) (a−2)
∂r,q + ∂r,q ∂r,p ∂r,q } + ∂r,q
(a−1)
∂r,p
(a)
= a∂q,p ∂r,q + (a − 1)∂r,q
(a−1) (a−1)
∂r,p + ∂r,q ∂r,p
(a)
= a{∂q,p ∂r,q (a−1)
+ ∂r,q ∂r,p }.
Dividing through by a, we get our identity (Cap) for b = 1. We have used,
(a) (b)
without explicitly mentioning it in doing the calculation, that ∂r,p and ∂r,q
commute, but this, we believe, is clear.
Now, assuming it true for all a and for b − 1, a similar inductive argument
proves the statement true for all a and b.
The proof of (Cap ) proceeds in the same way. 2
We next turn to positive-to-negative place polarizations.
In this case, we consider what happens if r is a negative place, with both p
and q still positive. If we look at the above proof, and keep in mind that higher
divided powers of positive-to-negative polarizations are zero, the identities (Cap)
and (Cap ) make sense only when a = 1, and our proof in the case a = 1 above,
is still valid. Hence we have the following proposition.
Proposition VI.4.3 (Capelli Identities) Let p, q, r be places with p, q pos-
itive, and r negative. Consider the place polarizations ∂r,q , ∂q,p and ∂r,p .
Then
(b) (b) (b−1)
(Cap)+ ∂r,q ∂q,p = ∂q,p ∂r,q + ∂q,p ∂r,p ,

(Cap )− (b)
∂q,p (b)
∂r,q = ∂r,q ∂q,p − ∂q,p
(b−1)
∂r,p .

VI.5 Weyl and Schur maps revisited


Recall the set-up for the definitions of the Weyl and Schur maps. We let F be
a finite free module over the commutative
n ring, R. For the n × m shape matrix
m
A = (aij ), set pi = j=1 aij , γj = i=1 aij . The Weyl map associated to A, ωA ,
is a map
ωA : Dp1 F ⊗ · · · ⊗ Dpn F → Λγ1 F ⊗ · · · ⊗ Λγm F
that we defined using many diagonalizations, identifications, and multiplications.
Similarly, we defined the Schur map
σA : Λp1 F ⊗ · · · ⊗ Λpn F → Sγ1 F ⊗ · · · ⊗ Sγm F.
We now maintain that these maps can be described using place polarizations;
in particular, positive-to-negative place polarizations.
168 Weyl and Schur modules

For the Weyl map, we are going to consider the basis, L+ , of F as a positive
letter alphabet (in the letter-place language), and our place alphabet P = P + "
P − , where P + = {1, . . . , n} and P − = {1, . . . , m}. For the Schur map, we are
going to regard the basis of F as a negatively signed letter alphabet, L− , and
our place alphabet the same as the above.
We next observe that S(L+ |P) = D(F ⊗ Rn ) ⊗ Λ(F ⊗ Rm ), which contains
the subalgebras S(L+ |P + ) = D(F ⊗ Rn ) and S(L+ |P − ) = Λ(F ⊗ Rm ). Our
discussion of the letter-place algebra tells us that D(F ⊗ Rn ) = DF ) ⊗ · · · ⊗ DF
% &# $
n
while Λ(F ⊗ Rm ) = ΛF ⊗ · · · ⊗ ΛF . A similar discussion applies to the algebra
% &# $
m
S(L− |P) = Λ(F ⊗ Rn ) ⊗ S(F ⊗ Rm ).
What we will show is that our Weyl (or Schur) maps are compositions of place
polarizations that take us from our desired domain to our desired target through
S(L+ |P) (or S(L− |P)).
Although we can carry out this project for arbitrary shapes, we will restrict
ourselves to the class of shapes we have already discussed in Section VI.1, namely
almost skew-shapes. Recall that an almost skew-shape can be represented as
λ/µ where λ is a partition and µ is an almost partition. In order to conform
to the notation used to describe the shape matrix, A, above, we will assume
that our partition λ has length n, and that λ1 − µn = m if µ is a partition,
and that λ1 − µn−1 = m if µ is not a partition. A quicker way to say this is
that λ1 − min(µn , µn−1 ) = m. As we have noted before, we may as well set
min(µn , µn−1 ) = 0.
Using this notation for our shapes, we see that the numbers pi and γj above
become:

pi = λi − µi ; γj = λ̃j − µ̃j ,

for i = 1, . . . , n and j = 1, . . . , m, where the tilde denotes the transpose shape


matrices of λ and µ.
For each i = 1, . . . , n, let

∆i = ∂λi ,i · · · ∂µi +1,i .

(Recall that we are assuming that min(µn , µn−1 ) = 0, so that m = λ1 .) Now


we set

∆λ/µ = ∆n · · · ∆1 .

We see that each ∆i is a composition of positive-to-negative place polarizations


from the positive place, i, to the negative places µi + 1 to λi . Hence the map
∆λ/µ is a composition of such place polarizations from 1, . . . , n to 1, . . . , m. We
see, therefore, that the image of ∆λ/µ is contained in that part of S(L+ |P) which
contains no positive places, namely in Λ(F ⊗ Rm ) or, what is the same thing, it
Some kernel elements of Weyl and Schur maps 169

is a map

∆λ/µ : DF ⊗ · · · ⊗ DF → ΛF ⊗ · · · ⊗ ΛF .
% &# $ % &# $
n m

If we restrict it to Dp1 F ⊗ · · · ⊗ Dpn F , it is immediate to see that we end in


Λγ1 F ⊗ · · · ⊗ Λγm F . It is laborious but straightforward to prove that this last
map is the same as the Weyl map ωA for A = λ/µ; we will sketch a procedure
for carrying out such an argument.
We know that a basis for Dp1 F ⊗ · · · ⊗ Dpn F consists of double tableaux
 
w1 1(p1 )
 w2 2(p2 ) 
(W )   ···
.
··· 
wn n(pn )
The result of applying ∆λ/µ to such a tableau yields the tableau
 
w1 µ1 + 1 · · · λ1
 w2 µ2 + 1 · · · λ2 
 .
 ··· ··· ··· ··· 
wn µn + 1 · · · λn
If one now reads this tableau as the element one obtains by diagonalizing wi over
the negative places µi + i, . . . , λi and multiplying, one sees that this is precisely
the definition of the map ωλ/µ .
The discussion of the Schur map is identical to this one, with the proviso
that we now consider the letters to be negative. However, we are still going
from positive places to negative ones, in exactly the same way, so that while the
domain and range of the Weyl and Schur maps are different, the expression of
them as composites of place polarizations is identical.

VI.6 Some kernel elements of Weyl and Schur maps


In this section, we will define some maps from the sum of tensor products of
divided powers (exterior powers) to the domain of the Weyl (Schur) map, and
show that the images are in the kernel of the Weyl (Schur) map. These maps are
what were called in Reference [3] the “box map”; here we will see that they are
expressible in terms of positive-to-positive place polarizations.
Consider our almost skew-shape λ/µ : λ = (λ1 , . . . , λn ), µ = (µ1 , . . . , µn ).
Remember that the shape is of type τ = n − (i + 1) if i is the largest integer
different from n such that µn ≤ µi . Thus, τ = 0 means that λ/µ is a skew-shape;
τ > 0 means that the bottom row of the diagram of λ/µ is indented on the left
from the penultimate row.
We will introduce some more notation that we will use uniformly when we
discuss these almost skew-shapes.
170 Weyl and Schur modules

Notation (almost skew-shapes)


We will set ti = µi − µi+1 for i = 1, . . . , n − 1. If τ = 0, this means that
µn ≤ µn−1 and tn−1 = µn−1 − µn = µn−1 . If τ > 0, this means that µn−1 − µn =
−µn < 0; moreover there is an i = n − 1 − τ such that µi+1 < µn ≤ µi ,
and we set s = µn − µi+1 . Finally, we denote our shape λ/µ by the notation
(p1 , . . . , pn ; t1 , . . . , tn−1 ).
With this notation, we see that the diagram of an almost skew-shape of type
τ = n − (i + 1) > 0 looks like this:
t1 p1
p2
.. .. .. ..
. . . .
ti pi
pi+1
.. .. .. ..
. . . .
tn−2 pn−2
pn−1
tn−2 + · · · + ti+1 + s pn
with 0 < s ≤ ti . Of course, tn−2 + · · · + ti+1 + s = µn − µn−1 = −tn−1 > 0.
We will now restrict ourselves to the Weyl case until the end of this section,
where we indicate how the results apply to the Schur case as well.
Assume that our shape (p1 , . . . , pn ; t1 , . . . , tn−1 ) is a skew-shape, that is,
assume that tn−1 ≥ 0.
For each i = 1, . . . , n − 1, and for each ki > 0, we consider the module Dp1 ⊗
· · · ⊗ Dpi +ti +ki ⊗ Dpi+1 −ti −ki ⊗ · · · ⊗ Dpn and the (positive-to-positive) place
polarization
(t +k )
i
∂i+1,i i
: Dp1 ⊗ · · · ⊗ Dpi +ti +ki ⊗ Dpi+1 −ti −ki ⊗ · · · ⊗ Dpn → Dp1 ⊗ · · · ⊗ Dpn .
Here, and from now on in most cases, we omit the underlying free module, F ,
from our notation.
Define λ/µ,i to be the map

λ/µ,i : Dp1 ⊗ · · · ⊗ Dpi +ti +ki ⊗ Dpi+1 −ti −ki ⊗ · · · ⊗ Dpn → Dp1 ⊗ · · · ⊗ Dpn ,
ki >0

i i (t +k )
which, on each summand, is equal to ∂i+1,i . Now define

Rel(λ/µ) = Dp1 ⊗ · · · ⊗ Dpi +ti +ki ⊗ Dpi+1 −ti −ki ⊗ · · · ⊗ Dpn ,
i ki

where the sum is taken over i = 1, . . . , n − 1, and all positive ki . And now define
λ/µ : Rel(λ/µ) → Dp1 ⊗ · · · ⊗ Dpn
to be the map which, for each i, is the map λ/µ,i .
Some kernel elements of Weyl and Schur maps 171

In short, λ/µ is the sum of many, many place polarizations.


We will often write Rel(p1 , . . . , pn ; t1 , . . . , tn−1 ) for Rel(λ/µ) when we want to
make the data for the shape more explicit. The reason for this elaborate notation
is that we will eventually show that the image of the map λ/µ is the kernel of
the Weyl map ∆λ/µ = ωλ/µ .
For an almost skew-shape of type τ > 0, the kernel of the Weyl map will
be given by relations of the kind above, plus τ additional kinds of terms. It is
evident from the definition of the map λ/µ above that the relations on the Weyl
map for a skew-shape involve shuffling between consecutive pairs of rows of the
shape. The additional terms that we must consider for the almost skew-shape of
type τ > 0 involve shuffling between the last row and those rows beyond which
it does not protrude (to the left), as well as the lowest row beyond which it does
protrude. In our diagram of the almost skew-shape of type τ > 0, this means
that we have to shuffle the last row with the rows from n − 2 up through the ith.
This makes n − (i + 1) = τ rows, and hence τ kinds of terms to describe these
shuffles.
We now formally describe these additional terms. For j = i + 1, . . . , n − 2,
define
tj
#λ/µ,j : Dp1 ⊗ · · · ⊗ Dpj +kj ⊗ Dpj+1 ⊗ · · · ⊗ Dpn−1 ⊗ Dpn −kj
kj =1

→ Dp1 ⊗ · · · ⊗ Dpn
(k )
to be the map which on each component is the place polarization ∂n,jj , and for
i = n − (τ + 1), define

s
#λ/µ,i : Dp1 ⊗ · · · ⊗ Dpi+ti−s+k ⊗ Dpi+1 ⊗ · · · ⊗ Dpn−1 ⊗ Dpn−(ti−s)−k
k=1

→ Dp1 ⊗ · · · ⊗ Dpn
to be, again, the map which on each component is the place polarization
(t −s+k)
∂n,ii .
We next define, for an almost skew-shape, λ/µ of type τ > 0, the overall
relations Rel(λ/µ) = Rel(p1 , . . . , pn ; t1 , . . . , tn−1 ) by:

Rel(p1 , . . . , pn ; t1 , . . . , tn−2 , 0)


n−2 
tj

Dp1 ⊗ · · · ⊗ Dpj +kj ⊗ Dpj+1 ⊗ · · · ⊗ Dpn−1 ⊗ Dpn −kj
j=i+1 kj =1


s
Dp1 ⊗ · · · ⊗ Dpi +ti −s+k ⊗ Dpi+1 ⊗ · · · ⊗ Dpn−1 ⊗ Dpn −(ti −s)−k ,
k=1
and the map
λ/µ : Rel(p1 , . . . , pn ; t1 , . . . , tn−1 ) → Dp1 ⊗ · · · ⊗ Dpn
172 Weyl and Schur modules

in the by now obvious way.


We point out that Rel(p1 , . . . , pn ; t1 , . . . , tn−2 , 0) are the relations of the skew-
shape obtained from the almost skew-shape by sliding the last row flush with
the penultimate one.
An easy way to think about the additional terms that we have to add to the
relations is this. Think of removing all the rows between the bottom one and the
jth. In that case, the bottom row is still pushed in from the jth one, and so push
it flush with it. The relations we write down are the terms that we would get
from the relations on that skew-shape, except that we interpolate the rows that
we removed. The last exceptional term, that is, the ith one (where τ = n − i − 1),
can also be thought of in the same way: we remove the intervening rows, and
observe that the last row is now protruding ti −s units beyond the jth. Thus, the
relations we write down are precisely those we would have from that skew-shape
after reinserting the excised rows.
Of course, one might ask why the summations in these terms do not go beyond
tj for the first type of term, or beyond s for the second. The fact is that one could
remove the upper limit of summation, as we shall see, but what we are trying
for is the smallest number of terms that we know we can get away with. To see
that the upper limit on the sums is not essential, we observe that an immediate
corollary of the Capelli identity (Cap ) is the following:

(v)
∂r,p = (−1)v ∂q,p ∂r,q −
(v) (v) (v−l) (l) (l)
(−1)l ∂r,p ∂r,q ∂q,p .
l>0

Thus, the image of a polarization of high order from one place to another far
down the line is already in the image of one of the same order from, say, adjacent
places, plus the images, by lower order polarizations, between the original places.
This is what permits us to limit the index of summation.
The main focus of this section is the proof of the following essential result.
Theorem VI.6.1 Let λ/µ be any almost skew-shape. Then the composition
λ/µ ∆λ/µ
Rel(λ/µ) −→ Dp1 ⊗ · · · ⊗ Dpn −→ Λγ1 ⊗ · · · ⊗ Λγm
is zero. That is, the image of λ/µ is contained in the kernel of the Weyl map.
Proof The proof depends heavily on the Capelli identity (Cap) involving
positive-to-negative polarizations.
First assume that our shape is a skew-shape, that is, that τ = 0. We want to
show that on each summand
Dp1 ⊗ · · · ⊗ Dpi +ti +ki ⊗ Dpi+1 −ti −ki ⊗ · · · ⊗ Dn ,
(t +k )
i i
the map ∆λ/µ ∂i+1,i is zero. Since ∆λ/µ is a composition of compositions of
positive-to-negative place polarizations (recall that ∆λ/µ = ∆n · · · ∆1 ), we can
(ti +ki )
commute ∂i+1,i past all the maps ∆1 , . . . , ∆i , and thus we have to study what
Some kernel elements of Weyl and Schur maps 173

(t +k )
the composition ∆n · · · ∆i+1 ∂i+1,i
i i
∆i · · · ∆1 looks like. We further dissect this
i i (t +k )
composition, and focus on the parts surrounding ∂i+1,i , namely
(t +k )
∂λi+1 ,i+1 · · · ∂µi+1 +1,i+1 ∂i+1,i
i i
∂λi ,i · · · ∂µi +1,i .
Using recursion and the identity (Cap), one sees that
(t +k )
∂λi+1 ,i+1 · · · ∂µi+1 +1,i+1 ∂i+1,i
i i

pi+1
 (t +k −α)

= ∂i+1,i
i i
∂λi+1 ,i+1 · · · ∂ˆβ1 ,i+1 · · · ∂ˆβα ,i+1 · · · ∂µi+1 +1,i+1 ∂β1 ,i · · · ∂βα ,i
α=0 βα <···<β1

where µi+1 < βα < · · · < β1 ≤ λi+1 , and the hat over the ∂ means that term is
to be omitted.
Next we observe that if α < ti , then each term
∂λi+1 ,i+1 · · · ∂ˆβ1 ,i+1 · · · ∂ˆβα ,i+1 · · · ∂µi+1 +1,i+1 ∂β1 ,i · · · ∂βα ,i
has at least pi+1 − ti polarizations from i + 1 to some negative place. But since
our original element has the place i + 1 only pi+1 − ti − ki times, our element is
carried to zero when α < ti . But if α ≥ ti , then the terms ∂β1 ,i · · · ∂βα ,i composed
with ∆i polarize the positive place i to α + pi ≥ ti + pi negative places between
µi+1 + 1 and λi . Since λi − (µi+1 + 1) = λi − µi + µi − µi+1 − 1 = pi + ti − 1, the
composition of these polarizations must take i to the same negative place twice,
which results in the zero map.
Now we assume that τ > 0, so that we must consider the extra summands
that comprise Rel(λ/µ). These are of two types:
D1 ⊗ · · · ⊗ Dpj +kj ⊗ Dpj+1 ⊗ · · · ⊗ Dpn−1 ⊗ Dpn −kj
and
D1 ⊗ · · · ⊗ Dpi +ti −s+k ⊗ Dpi+1 ⊗ · · · ⊗ Dpn−1 ⊗ Dpn −(ti −s)−k ,
To see what happens to terms of the first type under the composition, we again
(k )
see that the map λ/µ applied to them is simply ∂n,jj . We therefore just have
(k )
to look at ∆n ∂n,jj ∆n−1 . If we apply (Cap) as we did above, it is even easier to
see that the composition is zero, since the place j is sent to the negative places
µj + 1, . . . , λj , and µn > µj . For the term D1 ⊗ · · · ⊗ Dpi +ti −s+k ⊗ Dpi+1 ⊗
· · · ⊗ Dpn−1 ⊗ Dpn −(ti −s)−k , the same argument that was used above yields the
result. 2
Corollary VI.6.2 Let us define K̄λ/µ to be the cokernel of λ/µ . Then the
identity map on Dp1 ⊗ · · · ⊗ Dpn induces a map θλ/µ : K̄λ/µ → Kλ/µ .
Proof This follows immediately from the result above. 2
All of the above discussion carries over to the Schur map and Schur modules,
simply by replacing divided powers by exterior powers and exterior powers by
174 Weyl and Schur modules

symmetric powers. Or, if one wishes, one can simply replace the positive letter
alphabet by its negative counterpart. All the maps that we define are in terms
of the place alphabets, and these have not changed.

VI.7 Tableaux, straightening, and the straight basis theorem


The last theorem is a step toward giving us a presentation of our Weyl (Schur)
modules: since the image of ∆λ/µ is the Weyl module, Kλ/µ , it suggests that
perhaps the sequence
Rel(λ/µ) → Dp1 ⊗ · · · ⊗ Dpn → Kλ/µ → 0
is exact. At least we know it is a complex. In this section, we will prove a basis
theorem for our Weyl (Schur) modules, from which the exactness of the above
sequence will follow.

VI.7.1 Tableaux for Weyl and Schur modules


The Weyl module corresponding to the shape,

is the image of D4 ⊗ D3 under the map ∆2 ∆1 , where


∆1 = ∂5,1 ∂4,1 ∂3,1 ∂2,1 ;

∆2 = ∂3,2 ∂2,2 ∂1,2 .


Suppose {xi } is a basis for our free module, F (unspecified rank at this point),
(2)
and suppose we take the basis element of D4 ⊗ D3 : x2 x3 x4 ⊗ x1 x2 x4 . In our
double tableau notation for D4 ⊗ D3 , this would be written
 
x2 x2 x3 x4 1(4)
,
x1 x2 x4 2(3)
and its image under ∆λ/µ would be
 
x2 x2 x3 x4 2 3 4 5
.
x1 x2 x4 1 2 3
What we will do is write this element as
x2 x2 x3 x4
,
x1 x2 x4
namely as a tableau. This may cause some initial confusion as the element we
are representing by this tableau is in reality a sum of basis elements in Λ1 ⊗ Λ2 ⊗
Λ2 ⊗ Λ1 ⊗ Λ1 rather than simply a filling of a diagram. To be more meticulous,
Straightening and the straight basis theorem 175

we should really introduce some term such as Weyl-tableau to indicate that it


is more than just a filled diagram. However, it will be clear from the context of
our discussions, when we are using the term “tableau” in this extended sense,
and when we are using it in the strictly combinatorial or typographic sense.
This notation is not only the standard one used for these modules, but it is also
extremely efficient.
All of the above carries over mutatis mutandis for Schur modules: the divided
powers are replaced by exterior powers, and the exterior powers are replaced
by symmetric powers. In addition, the positive letters are replaced by negative
letters.
The next definitions of various kinds of standardness and straightness of
tableaux apply to tableaux of positive or negative letters; we will therefore
introduce a notation that will apply to both cases simultaneously.

Notation (signed inequalities)


If A is a multi-signed alphabet, we say that a <+ b if a < b or a and b are
positive, and a = b. We say that a <− b if a < b or a and b are negative, and
a = b.
In less formal language, a <+ b means, for example, that if a and b are both
positive, then a ≤ b. Otherwise, a < b.

With this notation we proceed with some definitions.

Definition VI.7.1 We say that a tableau (of any shape) is row-standard if in


each row it is <+ -increasing; we say it is column-standard if in each column it
is <− -increasing. We say it is standard if it is both row- and column-standard.

Definition VI.7.2 In a row-standard tableau, two elements aik , ajk (with


i < j) in the same column are said to form (or be) an inversion if they violate
column-standardness. That is, if aik is not <− ajk . The inversion is said to be
unflippable if there is an element in the tableau, bik−1 , immediately to the left
of aik , such that ajk <− bik−1 . Otherwise the inversion is called flippable. The
row-standard tableau is said to be straight if every inversion is unflippable.

Clearly, since a standard tableau has no inversions, a standard tableau is


necessarily straight. We do have a partial converse.

Proposition VI.7.3 If λ/µ is a skew-shape and T is a row-standard tableau


of that shape, then T is straight if and only if it is standard.

Proof It suffices to show that if the tableau is not standard, then it is not
straight. If T is not standard, it has an inversion, say aik is not <− ajk with
i < j. Let us suppose that this inversion is the left-most one that exists in the
tableau. If aik has no element immediately to the left of it, then there is no
way the inversion is unflippable, so T is not straight. If we do have an element
immediately to the left of aik , say bik−1 , then bik−1 lies above an element cjk−1 ,
176 Weyl and Schur modules

since our shape is a skew-shape. But since the k th column is the first one in
which we have an inversion, the column to the left of it is <− -increasing. Hence
bik−1 <− cjk−1 <+ ajk (the last inequality holding due to the row-standardness
of T ). Thus the inversion is flippable, and T is not straight. 2
From now on, we will focus our attention on Weyl modules; the appropriate
changes for Schur modules will mostly be left to the reader. In the Weyl case,
<+ -increasing means weakly increasing, while <− -increasing means strictly
increasing.

VI.7.2 Straightening tableaux


What we want to do now is show that our Weyl modules are generated by the
straight tableaux. There is a procedure that is called “straightening”, that is
used to do this. Before we get into the general details, we give two examples of
straightening.
Example VI.7.4 Let us take as our first example the tableau we have just
x2 x2 x3 x4
looked at, namely, . It fails to be standard because of an
x1 x2 x4
(2)
inversion in the second column. This element is the image of x2 x3 x4 ⊗x1 x2 x4 ∈
(3)
D4 ⊗ D3 . However, let us look at x1 x2 x3 x4 ⊗ x4 ∈ D6 ⊗ D1 ⊂ Rel((5, 3)/(1, 0)),
and at its image under (5,3)/(1,0) which, in this instance, means applying the
(2)
map ∂2,1 . We get
(2) (3) (3) (2)
x2 x3 x4 ⊗ x1 x2 x4 + x2 x4 ⊗ x1 x3 x4 + 2x2 x3 ⊗ x1 x4 +
(2) (2)
x1 x2 x3 x4 ⊗ x2 x4 + x1 x2 x4 ⊗ x2 x3 x4 +
(2) (2) (3) (2)
2x1 x2 x3 ⊗ x2 x4 + 2x1 x2 ⊗ x3 x4 .
Notice that the first summand is the term we started with. If we apply the Weyl
map to this sum, we get zero by the theorem of the last section. This means
that our original tableau is equal to the negative of the sum of the tableaux we
obtain by writing the images of all the remaining summands as tableaux. One
sees easily that all of these tableaux are standard, so that our original tableau is
a linear combination of standard tableaux.

Example VI.7.5 Another example to consider is the following tableau cor-


x3 x3 x3
responding to an almost skew-shape of type 1: x2 x4 x4 . Here we
x4
have an inversion that is flippable (in the second column, second and third
rows). In this case, we swoop up the third with the second row, and consider
(3) (3)
the element x3 ⊗ x2 x4 ⊗ 1 ∈ D3 ⊗ D4 ⊗ D0 ⊂ Rel((4, 3, 2)/1, 0, 1)). This
Straightening and the straight basis theorem 177

(3) (2) (3) (3)


maps (under ∂3,2 ) to the sum: x3 ⊗ x2 x4 + x3 ⊗ x4 ⊗ x2 , one of whose
summands is our original term. If we arrange these in tableaux, we have our
x3 x3 x3
original tableau plus another, namely x4 x4 x4 , which is not straight
x2
because of a flippable inversion in the second column, first and third row. We
will see soon that this tableau is “better” than the one we started with, but
we can go one step further to actually express our original tableau as a sum of
straight ones. For now we can swoop up the third and first rows together, getting
(3) (3)
−x2 x3 ⊗ x4 ⊗ 1 ∈ D4 ⊗ D3 ⊗ D0 ⊂ Rel((4, 3, 2)/(1, 0, 1)), and the image of
(3) (3) (2) (3)
this element under ∂3,1 is −x3 ⊗ x4 ⊗ x2 − x2 x3 ⊗ x4 ⊗ x3 . This shows that
x2 x3 x3
our original tableau is equal to the straight tableau: x4 x4 x4 .
x3

The main task is to formalize the argument underlying the procedure


illustrated above.
Let us assume, then, we have a tableau, T , which is not straight, which means
it has some flippable inversions. Of those, choose the ones furthest to the left,
and of those, choose the one: aik ≥ ajk ; i < j, for which i is maximal and j − i
is minimal. (Remember that in the case of skew-shapes, we would automatically
get that j = i + 1, but not so here.)
Since this inversion is flippable, we either have no element aik−1 , or we do and
aik−1 ≤ ajk .
Suppose first that aik−1 does not exist. Then if i < l < j, we must have aik < alk ,
for otherwise we should have a flippable inversion, and the minimality of j −i says
that this can not happen. But then alk > ajk , and by the maximality assumption
on i, this must be an unflippable inversion. This means that alk−1 exists, and
alk−1 > ajk . But our tableau is row-standard (we are always assuming that),
so since ajk−1 ≤ ajk (assuming ajk−1 exists), we have the inversion alk−1 >
ajk−1 . By the assumption on k, this inversion must be unflippable, so we have
alk−2 > ajk−1 , and so on. Following this pattern, we see that the only way to
avoid flippable inversions between rows l and j is for row l to protrude further
to the left than row j. This has to mean that j = n, and we have, for all l
with i < l < j, that µl < µn ≤ µi which means, in particular, that the almost
skew-shape is of type n − i − 1.
This was under the assumption that aik−1 does not exist, and that j > i + 1.
So now suppose that aik−1 does exist, but that we still have j > i + 1.
If for some l with i < l < j we had aik ≥ alk , we would have aik−1 > alk , since
the minimality of j − i forces this to be an unflippable inversion. But again, as
above (since our tableau is row-standard), this forces us to conclude that µl > µi ,
and since l < j ≤ n, this is impossible. So we see that if j − i > 1, we must
have, for all l with i < l < j, that aik < alk , which now implies that for all
178 Weyl and Schur modules

such l, alk > ajk , hence we have an inversion. Since (by maximality of i), this
inversion must be unflippable, we run through the usual argument to conclude
that µj > µl . Again we conclude that j = n, and the fact that our shape is of
type ≥ n − i − 1.
Let us observe too that if we have i < l < l < j, then we must have alk < al k .
Because if not, then (by maximality of i), the inversion has to be unflippable,
and by the usual leftward descent argument, we would conclude that µl < µl ,
which can not be since l < j ≤ n.
So finally, let us assume that j − i = 1. In that case, we have no extra inform-
ation except that if aik−1 exists, then aik−1 ≤ ai+1k . Furthermore, we have no
conclusion to draw about the type of the shape. We will, for the sake of simpli-
city of writing, refer to this flippable inversion of our non-straight tableau as the
extreme flippable inversion of T .
There are a few more observations we should make about these extreme flip-
pable inversions. As we have seen, we have aik−1 ≤ ajk if aik−1 exists. Let α
be the minimum value such that ajα exists and either aiα−1 does not exist, or
aiα−1 ≤ ajα . If row i does not protrude further to the left than row j or, what
is the same, µj ≤ µi , then α = µj + 1. If, on the other hand, µj > µi (in which
case we must have j = n), then we see that µn + 1 ≤ α ≤ k. This number α will
come into play when we describe our straightening algorithm. We should note
that if ajα−1 exists, then ajα−1 = ajα since we have ajα−1 < aiα−2 ≤ aiα−1
≤ ajα .
Our second example above suggests that we ask what happens if aik−1 = aik .
We saw in that example that we scooped up all of the second row along with
the third row in order to avoid inconvenient coefficients when we diagonalized.
We have to take this situation into consideration; it is a rather special case, but
merits special attention.
If we have aik−1 = aik , then we have aik−1 = aik = ajk (since aik−1 ≤ ajk ).
If ajk−1 exists, then we have aik−1 ≥ ajk−1 , so aik−2 must exist and be greater
than ajk−1 . Thus, in this case, our α = k, and we also see that we must have
j = n. (If ajk−1 does not exist, we certainly have α = k and j = n.)
This analysis of the type of configuration that can yield a non-straight tableau
is what informs our straightening procedure. But before we actually describe this
procedure, and maintain that applying it actually makes our tableaux “better”,
we have to say what “better” means.

Definition VI.7.6 Let T be a tableau. We define the column word of T ,


denoted uT , to be the word we obtain by writing down the elements of the columns
of T starting from the bottom of the left-most column of T , working up that
column, returning to the bottom of the next column, etc. We define the modified
column word of T , denoted wT , to be the word obtained from T reading the
columns in decreasing order. Finally, we define the reversed column word of
T , denoted wT , to be the word obtained from T reading the columns in increasing
order.
Straightening and the straight basis theorem 179

Just to be sure that there is no confusion about this definition, let us look at
x3 x3 x3
the tableau, x4 x4 x4 . The column word is uT = x4 x2 x4 x3 x4 x3 x3 ; the
x2
modified column word is wT = x4 x4 x3 x2 x4 x3 x3 , and the reversed column word
is wT = x4 x2 x3 x4 x3 x4 x4 .
Notice that the tableau above appeared in Example VI.7.5, but there it was
being treated as an element of Kλ/µ . Here, however, we are regarding the tableau
purely combinatorially.

Definition VI.7.7 Given two tableaux, T and T  , corresponding to the same


diagram, we say that T  < T if uT < uT  in the lexicographic ordering of words.

In Examples VI.7.4 and VI.7.5 above, the tableaux that were produced via
straightening were all less than the original tableaux we started with.
We are now set to show that the straight tableaux generate our Weyl module
or, what is the same, that every tableau can be written as a linear combination
(with integer coefficients) of straight tableaux. This is what we mean by the
straightening procedure referred to earlier. We proceed as we have done before:
we take a tableau, T , assume that it is not straight, and then show that it is
a linear combination of tableaux, T  , which are all smaller than T in the order
relation just defined. Since we are dealing with finite sets, this will do the job.
It will then remain to show that the straight tableaux are linearly independent,
and hence form a basis for the Weyl module. We will always assume that we
start with row-standard tableaux; when we read a tableau as an element of the
Weyl module, if it were not row-standard, making it so would not change the
element.
Let T be a (row-standard) non-straight tableau, and assume that its extreme
flippable inversion occurs in adjacent rows, i and i+1 < n, where n is the number
of rows in the tableau. Focusing our attention just on those two rows, we have
the situation:
t at+1 ··· ak−1 ak ··· aq
T = ,
b1 ··· bt+1 ··· bk−1 bk · · · bq 

with the inversion ak ≥ bk . (We are taking liberties here, calling just the two
rows T , but we think the meaning is clear. We have also changed the names of
the elements in the interests of clarity.)
We have drawn the diagram as though the cell immediately to the left of ak
exists; of course that need not be the case. But with our assumptions, t ≥ 0 and
we cannot have ak−1 = ak . We now let w ∈ Dk−1−t be the word at+1 · · · ak−1
(meaning that if there are repeats they are considered as divided powers rather
than simple powers), we let v ∈ Dq+1+ be the word b1 · · · bk · · · bk+ ak · · · aq ,
where bk = · · · = bk+ = bk++1 ( may be zero), and u ∈ Dq −k− the word
consisting of bk++1 · · · bq .
180 Weyl and Schur modules

From the Capelli identities, we see that the map


(k−1−t) (k+)
∂2,1 ∂3,2 : Dk−1−t ⊗ Dq+1+ ⊗ Dq −k− −→ D0 ⊗ Dq−t ⊗ Dq
applied to our element w ⊗ v ⊗ u gives us the element

k−1−t
(k+−h) (k−1−t−h) (h)
(−1)h ∂3,2 (∂2,1 ∂3,1 (w ⊗ v ⊗ u)),
h=0

and we observe that


(k−1−t−h) (h)
∂2,1 ∂3,1 (w ⊗ v ⊗ u) ∈ D0 ⊗ Dq−t+k+−h ⊗ Dq −k−+h ,
with q − t + k +  − h ≥ q +  + 1 > q. We thus see immediately that
(k−1−t) (k+)
∂2,1 ∂3,2 (w ⊗ v ⊗ u) is in the kernel of the Weyl map.
To see how this helps us, we must look at just what this term looks like. First
of all, we see that this term is a sum of tableaux, one of which is T itself. The
other tableaux that occur will all have w in the top row, u in the bottom row,
and the elements of v distributed around the top and bottom rows, as indicated
(k+)
by the polarization map ∂3,2 . To obtain a tableau other than T , we must either
have some term among the b1 , . . . , bk replaced by terms from among ak , . . . , aq ,
or all of those elements as they stand, but again some of the ai terms must come
down to the right of the bk . In either case, when one writes the column word
for such a tableau, one sees easily that it is greater in the lexicographic order
than the one for our original tableau, T . Since the sum of these tableaux is in
the kernel of the Weyl map, we see that T (considered as an element of the Weyl
module) is equal to the negative of the sum of the other tableaux, all of whose
column words are greater than that of T .
We now examine what happens when the extreme flippable inversion involves
the last row of the tableau. In this case, the two rows may or may not be adjacent,
but we will draw the picture as though they were, since all of our discussion (and
maps) involves just these two rows. We also may assume that the shape is not a
skew-shape but is an almost skew-shape of positive type, so that our top row of
the two projects to the left strictly beyond the bottom one. This means that we
are in the situation pictured thus:
a1 ··· as+1 ··· ak−1 ak ··· aq
T = ,
s bs+1 ··· bk−1 bk · · · bq 
with, again, the flippable inversion ak ≥ bk and ak−1 ≤ bk . (Because we are
assuming that s > 0, we know that ak−1 exists.) Let α be the index of the
column as described in our previous discussion, that is, such that either aα−1
does not exist, or aα−1 ≤ bα . We clearly have s + 1 ≤ α ≤ k. Because there are
no flippable inversions to the left of the kth column, we see that we have aj < bj
for α ≤ j ≤ k − 1. On the other hand, we have aj > bj for s + 1 ≤ j < α. Recall
that from our earlier discussion, bα−1 = bα (if bα−1 exists).
Straightening and the straight basis theorem 181

We now have to be a bit more careful about our definition of the word w: let w
be the word consisting of the elements in the top row to the left of ak−δ , where δ is
such that ak−δ−1 < ak−δ = · · · = ak . We let v be the word bα · · · bk+ ak−δ · · · aq ,
that is, we do not include the terms bs+1 · · · bα−1 . Finally, we let u be the
word bs+1 · · · bα−1 bk++1 · · · bq . We have w ⊗ v ⊗ u ∈ Dk−δ−1 ⊗ Dq−α+2+δ+ ⊗
(k−δ−1) (k++1−α)
Dq −s+α−(k++1) . Now let us consider ∂2,1 ∂3,2 (w ⊗ v ⊗ u). Again
using the Capelli identities, we see that

(k−δ−1) (k++1−α)

k+−α
(k++1−α−h) (k−δ−1−h) (h)
∂2,1 ∂3,2 = (−1)h ∂3,2 ∂2,1 ∂3,1
h=0
(α−−2−δ) (k−α++1)
+ (−1)k+−α+1 ∂2,1 ∂3,1 .
The important thing to notice here is that the terms under the summation
sign are all in the kernel of the Weyl map (since they all involve positive powers
of ∂32 ), so that under the Weyl map, the sum
(k−1−δ) (k++1−α) (α−−2−δ) (k−α++1)
∂2,1 ∂3,2 − (−1)k−α++1 ∂2,1 ∂3,1
applied to w ⊗ v ⊗ u gives zero. If one analyzes the tableaux that show up under
this sum, one sees that in addition to our tableau, T , one gets tableaux with w
in the top row and the terms of v spread between the top and bottom rows (as
in the previous situation), and other tableaux in which the word w gets spread
around the top and bottom rows. Because of the inequalities we observed at
the beginning of this discussion, we see that all the tableaux that appear, other
than T , have column words greater in the lexicographic ordering than that of T .
Hence, we have again “straightened” our original tableau, and we have shown
that the straight tableaux generate the Weyl module for an almost skew-shape.

VI.7.3 Taylor-made tableaux, or a straight-filling algorithm


Before we address the problem of proving that the set of straight tableaux is
linearly independent, we describe an algorithm, which we will call the Taylor
algorithm, that will produce straight tableaux of a given shape from certain
reverse column words. The algorithm was developed by B. Taylor, and, as with
most of these constructions involving straight tableaux, applies to the larger
class of row-convex shapes. We, however, will describe this procedure just for
our almost skew-shapes.
Let us denote by D the diagram of an almost skew-shape, with columns
c1 , . . . , cm . We make a few observations.
Fact 2 If T is a straight tableau of diagram D, then its reverse column word has
the property that its subwords corresponding to each column are strictly increas-
ing. Consequently, if we arrange the elements of the reverse column word as
elements of Λγ1 ⊗ · · · ⊗ Λγm (where γi is the number of boxes in ci ), we get a
basis element of this tensor product of exterior powers.
182 Weyl and Schur modules

Proof This is the same as saying that in each column, there are no repeats.
Suppose for some i < j we had aik = ajk . Then straightness implies that aik−1 >
ajk and this is clearly impossible. 2

Fact 3 If T is a straight tableau, and w = x11 ∧· · ·∧x1γ1 ⊗· · ·⊗xm1 ∧· · ·∧xmγm


is its reverse column word (written as a basis element of Λγ1 ⊗ · · · ⊗ Λγm ), then
in each column cj , the element xjl appears in the first box (from the top) not
occupied by xj1 . . . , xjl−1 which either has no neighbor to its immediate left, or
has one whose value is less than or equal to xjl .

Proof Certainly this is the case if l = γj . If l = γj , and if the assertion were not
true, then some xjs , with s > l, must be in that box. But xjs > xjl , so we have
an inversion. However, there is either no element to the left of the box occupied
by xjs , or there is and its value is less than or equal to xjl . Therefore, this is a
flippable inversion, contradicting the straightness of our tableau. 2

We stress again that all of what we have said makes sense if we think of the
tableau simply as a filled-in diagram. Whether the filled-in diagram is to be
regarded as an element of the Weyl module or not, depends upon the context in
which we are using it.
What we propose to do now is start with a basis element of Λγ1 ⊗· · ·⊗Λγm , and
associate to it, when possible, a filling of the diagram which is straight and whose
reverse column word is the given basis element we started with. When B. Taylor
introduced this algorithm[83], he called it the straight-filling algorithm, and
the resulting filled diagram a straight filling. In fact, we will see that if this
procedure does not produce a straight filling, then there is no straight tableau
with that given reverse column word.
Algorithm (Taylor Algorithm)
Let D be the diagram of an almost skew-shape with columns c1 , . . . , cm , and let
w = x11 ∧ · · · ∧ x1γ1 ⊗ · · · ⊗ xm1 ∧ · · · ∧ xmγm be a basis element of Λγ1 ⊗ · · · ⊗ Λγm .
Arrange x11 , . . . , x1γ1 in increasing order in column c1 . Next, place x21 in the
first box of c2 which either has no neighbor to its immediate left, or has such
a neighbor whose entry is less than or equal to x21 . If there is no such box in
c2 , the output of the algorithm is “no straight filling”. If there is such a box, fill
it with x21 . Assuming that we have placed x21 , . . . , x2j , we place x2j+1 in the
first empty box of c2 which either has no neighbor to its immediate left, or has
such a neighbor whose entry is less than or equal to x2j+1 . Again, if there is no
such box, our output is “no straight filling”; if there is, we fill it with x2j+1 . We
continue in this way with the remaining columns, obtaining an output of “no
straight filling”, or a filling which we shall call T (w).
Let us look at an example or two. As our shape, we will take one we have used
earlier, namely (4, 3, 2)/(1, 0, 1). This has three rows and four columns, with
γ1 = 1, γ2 = 3, γ3 = 2, γ4 = 1. The word w1 = x4 ⊗ x1 ∧ x2 ∧ x4 ⊗ x4 ∧ x5 ⊗ x6
Straightening and the straight basis theorem 183

x1 x4 x6
produces the filling x4 x4 x5 , while the word w2 = x4 ⊗ x1 ∧ x2 ∧ x5 ⊗
x2
x1 x4 x4
x4 ∧ x6 ⊗ x4 produces the filling x4 x5 x6 . On the other hand, words
x2
such as x4 ⊗ x1 ∧ x2 ∧ x4 ⊗ x5 ∧ x6 ⊗ x4 or x5 ⊗ x1 ∧ x2 ∧ x4 ⊗ x4 ∧ x6 ⊗ x4 produce
“no straight filling.”
We point out a few things about this algorithm. First of all, if we start with a
straight tableau, T , with reverse column word w (written as a basis element of
our tensor product of exterior powers), then T (w ) will clearly have the reverse
column word w , if T (w ) exists. But from our second fact above, T (w ) clearly
does exist and equals T . We therefore see that if two straight tableaux have the
same reverse column word (and hence the same modified column word), then
they are equal. Hence the straight tableaux are precisely those whose reverse
column words produce a successful outcome of the straight-filling algorithm.

VI.7.4 Proof of linear independence of straight tableaux


Remember that the Weyl map is a map

ωλ/µ : Dp1 ⊗ · · · ⊗ Dpn → Λγ1 ⊗ · · · ⊗ Λγm

associated to the almost skew-shape λ/µ. In the preceding subsection, we have


spoken about the basis elements of Λγ1 ⊗ · · · ⊗ Λγm in the “classical” (i.e., non-
letter-place) sense, and we have discovered that among those basis elements is a
subset, namely, that subset consisting of basis elements which produce a straight
tableau of shape λ/µ under the Taylor algorithm. This subbasis generates a free
submodule of Λγ1 ⊗ · · · ⊗ Λγm which we shall denote by Λ(λ/µ). By the same
token, our classical basis of Dp1 ⊗ · · · ⊗ Dpn contains a subset consisting of those
elements which yield a straight tableau when they are used to fill the diagram
of λ/µ. These form a basis of a free submodule of Dp1 ⊗ · · · ⊗ Dpn which we
will denote by D(λ/µ), and from the fact that the straight tableaux generate the
Weyl module, Kλ/µ , we know that the image of the Weyl map is equal to the
image of the composite map:
inc ωλ/µ
D(λ/µ) −→ Dp1 ⊗ · · · ⊗ Dpn −→ Λγ1 ⊗ · · · ⊗ Λγm ,

where the map labeled “inc” is the inclusion map. In addition to this inclusion
map, we also have the projection map proj:Λγ1 ⊗ · · · ⊗ Λγm → Λ(λ/µ). If we can
show that the composite map:
inc ωλ/µ proj
D(λ/µ) −→ Dp1 ⊗ · · · ⊗ Dpn −→ Λγ1 ⊗ · · · ⊗ Λγm −→ Λ(λ/µ)

is an isomorphism, then it will automatically follow that the straight tableaux


are linearly independent, and therefore are a basis for Kλ/µ .
184 Weyl and Schur modules

Since D(λ/µ) and Λ(λ/µ) are both free modules with selected bases, we will
show that the composite map above is an isomorphism by showing that the
matrix of this map, with respect to a suitable total order of the bases, is upper
triangular with ones on the diagonal. However, to make the picture of what we
are about to do clear, we first give an example.
We will take our favorite almost skew-shape: λ/µ = (4, 3, 2)/(1, 0, 1) whose
(2)
diagram is D = . Consider the element x1 x4 x6 ⊗ x4 x5 ⊗ x2 ∈ D3 ⊗

x1 x4 x6
D3 ⊗ D1 ; under our Weyl map it goes to the tableau T = x4 x4 x5 ,
x2
which is straight, so this element is a basis element of D(λ/µ). For convenience of
writing, let us agree to replace xi by its index, i. Also, instead of writing wedges
for the product in the exterior algebra, let us simply juxtapose, and instead
of writing the tensor product symbol, we will simply write a slash. Using this
shorthand, we see that in Λ1 ⊗ Λ3 ⊗ Λ2 ⊗ Λ1 , T is the sum of elements:
4/142/45/6 + 4/142/65/4 + 4/152/64/4 + 4/452/14/6 + 4/452/64/1
+ 5/142/64/4 + 4/642/15/4 + 4/642/45/1 + 4/652/14/4 + 5/642/14/4.
Of course, in calculating out all the terms that occur as summands of T , many
are zero, as any repeat in a column is automatically zero in the exterior powers.
What we are left with, then, is a sum of terms which, but for a sign, are basis
elements of Λ1 ⊗ Λ3 ⊗ Λ2 ⊗ Λ1 . If we rewrite these terms so that the sequences
between slashes are increasing, we can ask which of them yields a straight filling
under the Taylor algorithm. With very little difficulty one sees that there are
only two: the “original” filling, and the one that corresponds to 4/125/46/4,
1 4 4
whose associated filling is 4 5 6 .
2
We see, then, that the composite map applied to our element yields
4/142/45/6 + 4/152/64/4.
Let us notice that the second filling was constructed from the term we denoted
4/152/64/4, and the reversal of order in the third column was due to the Taylor
algorithm. If we write T and T  for the fillings corresponding to the two elements
above, and uT , uT  , their corresponding column words, we see immediately that
uT < uT  in the lexicographic order; thus T > T  . If, now, we let T  be the
straight filling of w (T  ) (which is pictured immediately above), then uT  < uT  ,
so T > T  .
This suggests the order we should put on the basis elements of Λ(λ/µ) and
D(λ/µ). Namely, we order the basis elements of Λ(λ/µ) according to the lexico-
graphic order on the column words of their straight fillings, and we order the
Straightening and the straight basis theorem 185

basis elements of D(λ/µ) according to lexicographic order of the column words


of their straight fillings. With this, we are able to prove the statements which
follow.

Proposition VI.7.8 The composite map proj ωλ/µ inc : D(λ/µ) → Λ(λ/µ) is
an isomorphism.

Proof The main argument in this proof is to show that with the ordering of the
basis elements that we have defined, the matrix of the map is upper triangular
with ones on the diagonal.
To this end, consider a basis element X = X1 ⊗ · · · ⊗ Xn ∈ D(λ/µ), and use
it to fill the diagram of λ/µ. Its image under the Weyl map is then the tableau
that we get using that filling (see example above). This tableau is straight, since
the hypothesis that X is in D(λ/µ) tells us precisely that, and the value of this
tableau in Λγ1 ⊗· · ·⊗Λγm is a sum of elements which are obtained by first diagon-
alizing each Xi into Λ1 ⊗ · · · Λ1 , and then multiplying the entries in each column.
% &# $
pi
To clarify this procedure, visualize the diagram as filled with the rearranged
terms; thus we obtain rows whose entries are permutations of the original row,
and the “columns” to which we refer are the columns of the diagram. (Again,
we recommend that you look at the example above). Because the multiplication
is occurring in an exterior algebra, any repeats will yield zero, so the summands
that remain will be those in which no repeats appear in any of the columns.
There are two facts that are fairly clear: the original filling is among the sum-
mands (the identity permutation on the rows is among the permutations of the
rows that we are considering); the column word of any of the fillings that occur,
other than the original one, is greater than that of the original one. This last can
be proven easily by observing that if there is any change in the first column, it
must be due to the substitution of one of the original entries by a larger one in
its row. A similar argument, column by column, proves the assertion.
The next thing we must do is apply the projection map to the surviving
summands; this now gives us the sum of those summands whose reverse column
word yields a straight filling. But it is easy to see that if T is any filling whose
reverse column word yields the straight filling, T  , then the column word of T  is
greater than that of T in the lexicographic order. For any change in any column
will be the result of moving a larger element in a column below a smaller one in
that same column (and one can now argue column by column as above).
As a result we see that the matrix of this map is indeed upper triangular, with
ones on the diagonal, and the proposition is proven. 2

We now have the straight basis theorem, as well as the presentation theorem
for almost skew-shapes.

Theorem VI.7.9 Let F be a free R-module, and λ/µ an almost skew-shape.


The following statements are true:
186 Weyl and Schur modules

1. The Weyl module, Kλ/µ (F ), is a free R-module with basis consisting of the
straight tableaux in a basis of F .
2. The map θλ/µ : K̄λ/µ (F ) → Kλ/µ (F ) is an isomorphism.
3. The functor Kλ/µ (F ), considered as a functor of F , is universally free.
Proof Our proposition above tells us that the straight tableaux are linearly
independent; in fact they are in one-one correspondence with the basis elements
of D(λ/µ) and of Λ(λ/µ). Since we already know that they generate Kλ/µ , they
form a basis of the Weyl module. This disposes of 1.
The map θλ/µ is clearly a surjection, so that the cosets of the basis elements
of D(λ/µ) generate K̄λ/µ . But it is clear that these elements are also linearly
independent, so they form a basis of K̄λ/µ . Hence the map θλ/µ must be an
isomorphism. This takes care of 2.
Finally, since K̄λ/µ is the cokernel of a natural map between two univers-
ally free functors, it must be universally free. But from 2, we conclude that
Kλ/µ is also universally free. (For a reminder of what universal freeness is, see
Subsection III.1.2.) 2

VI.7.5 Modifications for Schur modules


We have focused our attention in the preceding sections on Weyl modules; the
same results obtain for Schur modules, but some minor modifications have to
be made in the definitions and proofs. We will briefly discuss these changes, but
will not go into extensive detail.
First notice what row- and column-standard and straight mean for tableaux
of negative letters. Row-standard means strictly increasing in rows; column-
standard means weakly increasing in columns. Straight means the following:

• In a row-standard tableau, two elements aik , ajk (with i < j) in the same
column are said to form (or be) an inversion if they violate column-standardness.
That is, if aik > ajk . The inversion is said to be unflippable if there is an element
in the tableau, bik−1 , immediately to the left of aik , such that bik−1 ≥ ajk .
Otherwise the inversion is called flippable. The row-standard tableau is said to
be straight if every inversion is unflippable.
In a row-standard tableau of negative letters, the straightening is a bit easier
than is the case for positive letters, as there can be no repeated elements in
any given row. Hence, whereas in the straightening algorithm for positive letters
we had to scoop up elements in a row that repeated, here our straightening
procedure will not require that. Except for that, our straightening of tableaux
goes through as before, and we have the fact that the straight tableaux generate
the Schur module. To prove linear independence of the straight tableaux, we
must modify the Taylor algorithm.
Since the Schur map goes from tensor products of exterior powers to tensor
products of symmetric powers, and a basis of the tensor product of symmetric
Weyl–Schur complexes 187

powers is a tensor product of monomials, we have to show how to associate to


such a tensor product of monomials a straight tableau (or no tableau at all).
But clearly all we have to do is what we did with the positive letters, except for
insisting that if an element is put into a box which has a box to its immediate
left, then it has to be strictly greater than the occupant of that box. Otherwise,
the proofs go through mutatis mutandis.

VI.7.6 Duality
As anticipated in Remark III.1.5 for the special case of hooks, there is a duality
between some Weyl and Schur modules. It is well known, [3], that if A is the
shape matrix of a skew-shape, λ/µ, and A  is its transpose (again the shape matrix
∗ ∼
of a skew-shape), then KA (F ) = (LA(F ))∗ . The proof depends in part on the
fact that for such shapes, the modules in question are universally free. Now, if
A is the shape matrix of an almost skew-shape of positive type, its transpose
is no longer of the same kind. Therefore if we want an isomorphism like the
one stated, we would at least have to saturate the class of almost skew-shapes
with respect to transposition, and develop all of the preceding material for that
larger class of shapes. Since all of the shapes in that class would be row-convex,
and the straightening techniques and algorithms used in this chapter apply to
row-convex shapes, one could probably arrive at such a duality statement.

VI.8 Weyl–Schur complexes


In this section, we simultaneously extend the notions of Weyl and Schur modules
to a complex which we choose to call the Weyl–Schur (or W–S) complex. We
make this choice even though it was just called the Schur complex when it first
appeared [3]. If λ/µ is an almost skew-shape, the W–S complex associated with
it will be denoted by Kλ/µ (ϕ), to be consistent with the notation we have been
using throughout this chapter. Sometimes the notation, Lλ/µ (ϕ), is used as well.
We will consider a given map ϕ : G → F , where F and G are free modules, in
place of just a free module, F . We will define D(ϕ) to be the algebra D(G)⊗Λ(F )
endowed with an endomorphism, ∂ϕ : D(ϕ) → D(ϕ), defined as the composition
of the maps:
∆⊗id id⊗µ
D(G) ⊗ Λ(F ) −→ D(G) ⊗ D1 (G) ⊗ Λ(F ) −→ D(G) ⊗ Λ(F ),
where ∆ is the diagonal map of D(G) → D(G)⊗D1 (G), and µ is the composition:
ϕ⊗id m
D1 (G) ⊗ Λ(F ) −→ Λ1 (F ) ⊗ Λ(F ) −→ Λ(F ).
In this second composition, m stands for multiplication in the exterior algebra.
We set

Dk (ϕ) = Dα (G) ⊗ Λk−α (F ),
0≤α
188 Weyl and Schur modules

and note that our map ∂ϕ gives us a complex


0 → Dk (G) → Dk−1 (G) ⊗ Λ1 (F ) → · · · → D1 (G) ⊗ Λk−1 (F ) → Λk (F ) → 0.
The reader will recognize this to be one of the complexes we considered in
Chapter III.
We can actually say a bit more, namely that this endomorphism of D(ϕ)
makes D(ϕ) into a differential graded algebra; that is, with the graded algebra
structure of the tensor product of two algebras, the map ∂ϕ is a derivation. This
means that it satisfies the identity:
∂ϕ ((a1 ⊗ b1 )(a2 ⊗ b2 )) = ∂ϕ (a1 ⊗ b1 )(a2 ⊗ b2 ) + (−1)b1 (a1 ⊗ b1 )∂ϕ (a2 ⊗ b2 ),
where (−1)b1 means (−1) raised to the degree of the element b1 . The proof of
this is straightforward; it merely uses the fact that the diagonal map preserves
multiplication.
In a similar way, we can define Λ(ϕ) to be the algebra Λ(G) ⊗ S(F ) endowed
with an endomorphism, ∂ϕ : Λ(ϕ) → Λ(ϕ), defined as the composition of the
maps:
∆ ⊗id id⊗µ
Λ(G) ⊗ S(F ) −→ Λ(G) ⊗ Λ1 (G) ⊗ S(F ) −→ Λ(G) ⊗ S(F ),
where ∆ is the diagonal map of Λ(G) → Λ(G)⊗Λ1 (G), and µ is the composition:
ϕ⊗id m
Λ1 (G) ⊗ S(F ) −→ S1 (F ) ⊗ S(F ) −→ S(F ).
In this composition, too, m stands
 for multiplication in the symmetric algebra.
As above, we define Λk (ϕ) = 0≤α Λα (G) ⊗ Sk−α (F ), and note that our map
∂ϕ gives us a complex
0 → Λk (G) → Λk−1 (G) ⊗ S1 (F ) → · · · → Λ1 (G) ⊗ Sk−1 (F ) → Sk (F ) → 0.
This in turn makes Λ(ϕ) into a differential graded algebra.
If we want to apply letter-place methods in this situation as we did earlier, we
first establish the following proposition.
Proposition VI.8.1 Let ϕ1 : G1 → F1 and ϕ2 : G2 → F2 be two maps of
free modules. Then ϕ1 ⊕ ϕ2 : G1 ⊕ G2 → F1 ⊕ F2 is a map of free modules, and
D(ϕ1 ⊕ ϕ2 ) ∼
= D(ϕ1 ) ⊗ D(ϕ2 ). The corresponding statement for Λ(ϕ1 ⊕ ϕ2 ) is
also true.
Proof We will simply show that as graded modules they agree; the detailed
verification of the algebra and complex isomorphisms is straightforward, but
tedious.
We have

(∗) Dk (ϕ1 ⊕ ϕ2 ) = Dα (G1 ⊕ G2 ) ⊗ Λk−α (F1 ⊕ F2 )
α

= Dβ (G1 ) ⊗ Dα−β (G2 ) ⊗ Λγ (F1 ) ⊗ Λk−α−γ (F2 ),
α,β,γ
Weyl–Schur complexes 189

while

(∗∗) (D(ϕ1 ) ⊗ D(ϕ2 ))k = Dα (ϕ1 ) ⊗ Dk−α (ϕ2 )
α

= Dβ (G1 ) ⊗ Λα−β (F1 ) ⊗ Dγ (G2 ) ⊗ Λk−α−γ (F2 ).
α,β,γ

If we set γ = α − β, we see that the last sum in (∗∗) rearranges into the last sum
in (∗). 2

Corollary VI.8.2 If we denote by ϕ ⊗ Rn the direct sum ϕ ⊕ · · · ⊕ ϕ, we have


% &# $
n

D(ϕ ⊗ R ) = D(ϕ) ⊗ · · · ⊗ D(ϕ) .


n
% &# $
n

This clearly allows us to think of a basis of Rn as, say, positive places, and
to develop the letter-place context for this situation. In fact, we will think of a
basis of F , say F, as an ordered set of negative letters, and a basis of G, which
we will write G, as an ordered set of positive letters. For convenience (it is easy
to show that our choice makes no essential difference to our discussion), we shall
assume that every element of G precedes all the elements of F, and so we have
a total order on G " F.
We will use the customary double tableau notation, but since the “letters”
are basis elements of D(G) ⊗ Λ(F ), we should explain what we mean by an
expression such as (w ⊗ u|1(k1 ) 2(k2 ) · · · n(kn ) ), where w ∈ Dk (G), u ∈ Λl (F ) and
k + l = k1 + · · · + kn . To be consistent with our previous notation for letter-place
identities, we set

(w ⊗ u|1(k1 ) 2(k2 ) · · · n(kn ) ) = (w(s1 ) ⊗ u(t1 )|1(k1 ) ) · · · (w(sn ) ⊗ u(tn )|n(kn ) ),

where the sum ranges over all pairs si , ti such that si + ti = ki . We have written
w(si ), u(ti ) to indicate the components of the n-fold diagonalization of w and u
of degrees si , ti . Again, an example will serve to make this much clearer.

Example VI.8.3 Consider (x(2) ⊗ y1 ∧ y2 |12(2) 3). This is equal to


(x ⊗ 1|1)(x ⊗ y1 |2(2) )(1 ⊗ y2 |3) − (x ⊗ 1|1)(x ⊗ y2 |2(2) )(1 ⊗ y1 |3) + (x ⊗ 1|1)
(1 ⊗ y1 ∧ y2 |2(2) )(x ⊗ 1|3) + (1 ⊗ y1 |1)(x(2) ⊗ 1|2(2) )(1 ⊗ y2 |3) − (1 ⊗ y2 |1)
(x(2) ⊗ 1|2(2) )(1 ⊗ y1 |3) + (1 ⊗ y1 |1)(x ⊗ y2 |2(2) )(x ⊗ 1|3) − (1 ⊗ y2 |1)
(x ⊗ y1 |2(2) )(x ⊗ 1|3).
We have been very meticulous about writing, say, 1 ⊗ y1 instead of just y1 . In
the future, we will generally drop the 1 and the tensor product sign when the
meaning is evident.
190 Weyl and Schur modules

Once we have this notation at our disposal, we are set to use our old double
tableau notation. If A stands for the multi-signed alphabet G " F, the notation
and definitions of Subsection VI.7.1 apply to our situation.
At this point, it must be fairly clear to the reader that the same “straightening”
arguments we have used before will work here to show that the double standard
tableaux generate the n-fold tensor product of D(ϕ). We refer the reader to
Section 4 of Appendix A to see how to prove linear independence in much the
same way as before. An illustrative example will also be found there. As a result
we have the theorem:
Theorem VI.8.4 Let ϕ : G → F be a morphism of free R-modules. Then the
double standard tableaux discussed above form a basis for D(ϕ ⊗ Rn ). The same
result obtains for Λ(ϕ ⊗ Rm ).
Clearly, one can now proceed with place polarizations as in Section VI.4.
Because the polarizations operate only on the places, all the structure attached to
the letters is preserved by them. In particular, the complex structure that we have
imposed on our algebra is preserved. As a result, given an almost skew-shape,
λ/µ, (or any shape for that matter), we can define the map
∆λ/µ (ϕ) : Dp1 (ϕ) ⊗ · · · ⊗ Dpn (ϕ) → Λγ1 (ϕ) ⊗ · · · ⊗ Λγm (ϕ)
as we did the Weyl and Schur maps, polarizing the positive places (basis of Rn )
to the negative places (basis of Rm ) according to the recipe dictated by the
diagram of λ/µ. (The pi and γj are read from the diagram of λ/µ in the usual
way.) This, by our observation above, is a map of complexes, which we call the
Weyl–Schur map (or W–S map); as a result, its image is also a complex.
Definition VI.8.5 The image of the map ∆λ/µ is the Weyl–Schur (W–S)
complex, Kλ/µ (ϕ), associated to the map ϕ and the shape λ/µ.
Remark VI.8.6 It is important to point out here that, since the map ∆λ/µ
does not depend on the map ϕ at all, the chains of the complex Kλ/µ (ϕ) depend
only on the shape λ/µ and the modules G and F .
With this definition, we can proceed to define the complex Rel(λ/µ) as we did
for Weyl and Schur modules, as well as the map of complexes, λ/µ , whose image
is in the kernel of ∆λ/µ (ϕ). We denote again, by K̄λ/µ (ϕ), the cokernel of the
map λ/µ and by θλ/µ the naturally induced map of complexes from K̄λ/µ (ϕ)
to Kλ/µ (ϕ).
Given our signed inequalities, namely, <+ and <− , our definitions of standard
and straight apply to the tableaux of shape λ/µ, filled with the elements of G "F,
and our analysis of flippable and unflippable inversions goes through without
change. Thus the straightening procedure we used to show that the straight
tableaux generate goes through using <+ and <− on the multi-signed alphabet
G " F. Similarly, the Taylor algorithm goes through without change as long as
we use the inequalities above. In fact, one can see that the “modifications” made
Weyl–Schur complexes 191

in discussing the Taylor algorithm for Schur modules rather than Weyl modules
were really unnecessary if we had used the <+ and <− inequalities throughout.
As a result, we can state the following theorem without further proof:
Theorem VI.8.7 Let ϕ : G → F be a morphism of free R-modules, and λ/µ
an almost skew-shape. The following statements are true:
1. The Weyl–Schur complex, Kλ/µ (ϕ), is a free R-complex with basis consist-
ing of the straight tableaux in a basis of F union a basis of G.
2. The map θλ/µ : K̄ λ/µ (ϕ) → Kλ/µ (ϕ) is an isomorphism.
This page intentionally left blank
VII
SOME APPLICATIONS OF WEYL AND SCHUR MODULES

In the previous chapter we developed the basic definitions of Weyl and Schur
modules, as well as Weyl–Schur complexes. In this chapter, we will give
indications of how we may apply some of these ideas.
In the first two sections, we deal with very basic material, and go into
some detail, as the results are fundamental for most of the applications of
characteristic-free representation theory that we will see. The remainder of the
chapter takes a more cavalier approach to the results considered; we include a few
historical notes and rely even more heavily than heretofore on recent reference
material.

VII.1 The fundamental exact sequence


Let ϕ : G → F be a morphism of free R-modules, and A a shape matrix. The first
thing we should do in this section is consider what our Weyl–Schur complexes
look like in two very special cases: (i) G = 0 and (ii) F = 0.
Remark VII.1.1 When G = 0, we see that our complex Dk (ϕ) reduces to
Λk F while the complex Λk (ϕ) reduces to Sk (F ). Since the Weyl–Schur map is
defined in terms of place polarizations in all of our cases, we see that the complex
KA (ϕ) simply is the Schur module LA (F ).
Similarly, when F = 0, we find that KA (ϕ) is just the Weyl module KA (F ). In
this case, we notice that the module does not lie in degree zero as in the former
case, but occurs in the highest degree term of our complex. This is simply because
we have chosen the convention of considering our complexes as left, rather than
right, complexes.
From Remark VI.8.6, we see that the chains of KA (ϕ) for an arbitrary ϕ are
the same as those for KA (ϕ ), where ϕ = ϕ1 ⊕ ϕ2 . with ϕ1 = G → 0 and
ϕ2 = 0 → F . We will return to the implications of this later, to get a description
of the chains when A is the shape matrix of a skew-shape.
We now consider an almost skew-shape, λ/µ = (p1 , . . . , pn ; t1 , . . . , tn−1 ), where
we use the notation introduced at the beginning of Section VI.6. (Remember
that the integer, tn−1 , may be non-negative, or negative; in the first case, we
have a skew-shape, while in the second, we have an almost skew-shape of pos-
itive type.) From this shape, let us construct the associated almost skew-shape,
λ /µ = (p1 , . . . , pn ; t1 , . . . , 1+tn−1 ). If we write λ/µ = (λ1 , . . . , λn )/(µ1 , . . . , µn ),
194 Some applications of Weyl and Schur modules

we can write λ /µ = (λ1 + 1, . . . , λn−1 + 1, λn )/(µ1 + 1, . . . , µn−1 + 1, µn ).


If our original shape is a skew-shape, our convention adopted in Chapter VI
would have us assume that µn = 0, and in that case our skew-shape λ /µ also
satisfies that convention. On the other hand, if our almost skew-shape we star-
ted with is of positive type, our convention would have us choose µn−1 = 0,
in which case our representation of λ /µ no longer satisfies that convention.
Of course, in this latter case, we could choose to present our new shape as
λ /µ = (λ1 , . . . , λn−1 , λn − 1)/(µ1 , . . . , µn−1 , µn − 1), since our assumption of
positive type implies that µn > 0.
For notational convenience, we denote by M the complex Dp1 ⊗R · · · ⊗R Dpn ,
where we have omitted the ϕ; that is, we have written Dpi for Dpi (ϕ). Then we
have the two exact sequences:
(E  ) Rel(λ /µ ) → M → Kλ /µ (ϕ) → 0,
(E) Rel(λ/µ) → M → Kλ/µ (ϕ) → 0,
and we observe that
(†) Rel(λ/µ) = Rel(λ /µ ) ⊕ N,
where
(‡) N = Dp1 ⊗R · · · ⊗R Dpi +ti −s+1 ⊗R Dpi+1 ⊗R · · · ⊗R Dpn−1 ⊗R Dpn −ti +s−1 ,
with s and i as described below. This last observation may require a bit of
explanation: we resort to the definitions of the terms Rel(λ/µ) and Rel(λ /µ )
found in Chapter VI, Section 6, which we will not reproduce here.
If our original shape, λ/µ, is a skew-shape, then so is λ /µ , and in that case
i = n − 1 and s = 0. There are no “extra” terms that have to be added to the
basic relations of a skew-shape, and it is easy to see that (†) and (‡) are true.
If λ/µ is an almost skew-shape of positive type, then there are a number
of possibilities, which we now examine. Recall that, in this case, our type τ
is equal to n − i − 1, where i is such that µi+1 < µn ≤ µi . (Here we will
adopt our convention that µn−1 = 0 and µn > 0.) If it turns out that we have
µi+1 < µn − 1 ≤ µi , then our almost skew-shape, λ /µ , is also of type τ , its “s”
is now s − 1, and again it is easy to see that (†) and (‡) are true.
The situation becomes a bit more complex when it turns out that µi+1 =
µn − 1. In this case, the “s” of our shape λ/µ is equal to 1, and the type of λ /µ
decreases. In fact, suppose that we take α to be the largest integer such that
µi+1 = · · · = µi+α , with α ≤ τ . Then the type of λ /µ is precisely τ − α ≥ 0. In
this case, we see that all the integers ti+1 , . . . , ti+α−1 are zero, and the “s” for
λ /µ is zero also. Now a quick examination of the terms comprising Rel(λ/µ)
and Rel(λ /µ ) will confirm the statements (†) and (‡).
From the exact sequences (E  ) and (E), we see that the identity map on M,
and the inclusion map of Rel(λ /µ ) into Rel(λ/µ), induce a surjection
Kλ /µ (ϕ) → Kλ/µ (ϕ).
The fundamental exact sequence 195

What we intend to prove is that the kernel of this surjection is the W-S complex,
Kλ /µ (ϕ), where
λ /µ = (p1 , . . . , pi−1 , pi + ti − s + 1, pi+1 , . . . , pn−1 , pn − ti + s − 1;
t1 , . . . , ti−2 , ti−1 + ti + s − 1, s − 1, ti+1 , . . . , tn−2 , −(tn−2 + · · · + ti )).
While the above description of λ /µ is very graphic, it is also very unwieldy.
Also, our insistence that in the case of a skew-shape we have µn = 0, but for
an almost skew-shape we have µn−1 = 0, makes for a certain amount of distinct
case analysis in our discussions. To avoid this, we will simply start with our
almost skew-shape, λ/µ, define our shape λ /µ to be (λ1 + 1, . . . , λn−1 + 1, λn )/
(µ1 + 1, . . . , µn−1 + 1, µn ), and then our shape λ /µ is simply:
(λ1 + 1, . . . , λn−1 + 1, λn )/(µ1 + 1, . . . , µi−1 + 1, µn , µi+1 + 1, . . . , µn−1 + 1, µi + 1).

• There is one caveat we must call attention to: if i = 1, then since


µ2 < µn ≤ µ1 , µ = (µn , µ2 + 1, . . . , µn−1 + 1, µ1 + 1) is not an almost partition
for µ1 + 1
µn . In this case, we set
λ = (λ1 + 1, . . . , λn−1 + 1, λn − (µ1 − µn + 1))
and
µ = (µn , µ2 + 1, . . . , µn−1 + 1, µn ).
All of this clearly does not affect any of our discussion above. However, in the
next part of this section we will be using the definition of the Weyl–Schur map
as a composition of place polarizations (positive to negative), and in Chapter VI
we set min(µn−1 , µn ) = 0 to define these maps. If, however, for an almost skew-
shape λ/µ we just set β = min(µn−1 , µn ), then we simply have to subtract β
from all the λ’s and µ’s that occur in the indices of the polarization maps.
With these modifications in mind, we can now state and prove the theorem
on fundamental exact sequences.
Theorem VII.1.2 Let λ/µ be an almost skew-shape, and let λ /µ and λ /µ
be as above. Then we have the exact sequence of complexes:
0 → Kλ /µ (ϕ) → Kλ /µ (ϕ) → Kλ/µ (ϕ) → 0,
where the map Kλ /µ (ϕ) → Kλ/µ (ϕ) is the one induced by the identity map
on the generators of these complexes, and the map Kλ /µ (ϕ) → Kλ /µ (ϕ) is
induced by the map of the generators of Kλ /µ (ϕ) into those of Kλ /µ (ϕ) via
(t −s+1)
the (positive to positive) place polarization ∂n,ii .
Proof That the map Kλ /µ (ϕ) → Kλ/µ (ϕ) is surjective, is trivial to see. It is
easy to see that the ranges of the Weyl–Schur maps ∆λ /µ and ∆λ /µ are the
196 Some applications of Weyl and Schur modules

same, say W. It is also clear that the domain of ∆λ /µ is N, where N is as
defined above. Therefore, if we can prove that the diagram:
(t −s+1)
∂n,ii
N / M

∆λ /µ ∆λ /µ


 
W / W
id

is commutative, where M too is as defined above, we will be done. But the


commutativity is easily demonstrated by application of the Capelli identities.
2
We point out here that the original proof of this theorem made use of a rather
complicated spectral sequence argument. This was due to the fact that at that
time, there was no general presentation theorem for almost skew-shapes; the
presentation theorem was originally proven only for skew-shapes. However, with
the advent of B. Taylor’s notion of straight tableaux, the presentation theorem
for almost skew-shapes was possible, and now we have the rather elementary
proof above of the theorem on fundamental exact sequences.
We can use this theorem, along with an induction assumption, to prove the
following result.
Theorem VII.1.3 Let ϕ : F → F be the identity map, and λ/µ an almost
skew-shape. Then, unless λ/µ = 0, in which case the homology is simply R, the
complex Kλ/µ (ϕ) is exact, that is, all the homology groups of the complex are
zero.
Proof We prove the nontrivial case by a double induction on the the number
of rows of the diagram, and the sum, σ, of the number of overlaps of the last
row of the diagram with each of the other rows. Note first that if λ/µ consists
of only one row, our result follows from the discussion immediately preceding
Definition III.1.3. Thus, we may assume our theorem true for shapes having at
most n − 1 rows, and prove it for those having n rows.
If σ = 0, then our complex Kλ/µ (ϕ) must be of the form Dλn /µn ⊗R Kλ /µ (ϕ),
where λ /µ is our shape λ/µ with its last row removed. But by our induction
hypothesis on the number of rows, we see that we have the tensor product of
two free exact complexes, and this is again exact. If we now start with an almost
skew-shape, λ/µ, with σ > 0, then we notice that the shapes λ /µ and λ /µ
of our Theorem VII.1.2 above both have smaller values of σ, and hence the
associated complexes are exact. It is now trivial to conclude that our original
complex is exact. 2
Theorem VII.1.2 deals only with what happens when one slides the last row of
the diagram one move to the left. There is, however, a special situation in which
Direct sums and filtrations for skew-shapes 197

one can move much more to the left (still by only one step). In fact, a similar,
but much easier argument than that used to prove Theorem VII.1.2 yields the
following result.

Theorem VII.1.4 Let α be the skew-shape represented by

(p1 , . . . , pk , pk+1 , . . . , pn ; t1 , . . . , tk−1 , pk+1 − 1, tk+1 , . . . , tn−1 ).

Then we have the exact sequence

0 → Kδ (ϕ) → Kβ (ϕ) ⊗ Kγ (ϕ) → Kα (ϕ) → 0 ,

where

β = (p1 , . . . , pk ; t1 , . . . , tk−1 ), γ = (pk+1 , . . . , pn ; tk+1 , . . . , tn−1 )

and

δ = (p1 , . . . , pk−1 , pk + pk+1 , pk+2 , . . . , pn ; t1 , . . . , tk−2 , tk−1 + pk+1 , tk+1 , . . . , tn−1 ).

VII.2 Direct sums and filtrations for skew-shapes


In this section we are going to consider just the class of skew-shapes. When we
discuss the W–S complex associated to a direct sum of maps and a skew-shape
λ/µ, we will take a brief look also at the kind of additional problem that arises
if the shape is assumed to be almost skew.
Even for skew-shapes we will not give full proofs, since a good reference is
available (Reference [3], section 5.1, where our W–S complexes are called Schur
complexes and denoted by Lλ/µ (ϕ)), and here it will be enough to just give an
indication of the main ideas.

Theorem VII.2.1 ([3], Theorem 5.1.13) For i = 1, 2, let ϕi : Gi → Fi be


two maps, and let ϕ = ϕ1 ⊕ ϕ2 . Then for any skew-shape λ/µ, the W–S complex,
Kλ/µ (ϕ), has a filtration by subcomplexes the terms of whose associated graded
complex are Kν/µ (ϕ1 ) ⊗R Kλ/ν (ϕ2 ), where ν runs over all partitions such that
µ ⊆ ν ⊆ λ.

Discussion. We will only discuss the considerations that lead to the definition
of the filtration.
First, we order the basis for G ⊕ F , with G = G1 ⊕ G2 and F = F1 ⊕ F2 ,
by making the basis elements of Gi precede those of Fi , and the basis elements
of G1 ⊕ F1 precede those of G2 ⊕ F2 . For notational convenience, let us denote
the basis of G1 ⊕ F1 by U, and the basis of G2 ⊕ F2 by V. Since for a skew-
shape, “straight” tableaux are the same as “standard” tableaux, we see that
row-standardness of a tableau for Kλ/µ (ϕ) places the elements of U before those
of V, while column-standardness requires that the “separators” between the first
198 Some applications of Weyl and Schur modules

and the second set of elements in each row together form a partition ν that
contains µ and is contained in λ. We may illustrate this situation as follows:
u v
(S) u v .
u v
This leads us to define the filtration on Kλ/µ (ϕ) as follows.
For each partition ν with µ ⊆ ν ⊆ λ, define
  
Mν (Dλ/µ (ϕ)) = im Dσ/µ (ϕ1 ) ⊗R Dλ/σ (ϕ2 ) → Dλ/µ (ϕ)
ν⊆σ⊆λ
  
Ṁν (Dλ/µ (ϕ)) = im Dσ/µ (ϕ1 ) ⊗R Dλ/σ (ϕ2 ) → Dλ/µ (ϕ)
ν⊂σ⊆λ
 
Mν (Kλ/µ (ϕ)) = ∆λ/µ (ϕ) Mν (Dλ/µ (ϕ))

 
Ṁν (Kλ/µ (ϕ)) = ∆λ/µ (ϕ) Ṁν (Dλ/µ (ϕ)) ,

where the top two maps are essentially the inclusion maps, and ∆λ/µ (ϕ) is the
usual map. The proof that the above is our desired filtration and that
Mν (Kλ/µ (ϕ)) / Ṁν (Kλ/µ (ϕ)) ∼
= Kν/µ (ϕ1 ) ⊗R Kλ/ν (ϕ2 )
is actually an isomorphism is found in Reference [3]; the reader should be warned
that there the bold-faced D is instead written as Λ.
Let us now look at what the situation would be if instead of a skew-shape we
had started with an almost skew-shape (still denoted by λ/µ). Keeping our same
ordering of the basis, and looking at a straight tableau, we see that as in the
previous case, our row-standardness has all the basis elements of U preceding
those of V. An analysis of straightness in the columns, however, tells us that the
line of demarcation in each row gives us, until we reach the penultimate row,
what looks like a subpartition of λ which contains µ. But straightness allows the
possibility that in the last row, we have some basis elements of U lying under
some elements of V in a higher row, provided that this higher row protrudes
beyond the bottom row. As a result, these demarcations delineate an almost
partition, rather than a partition. Again, it is useful to illustrate this situation
with the following diagram:
u v
u v
(AS) .
u v
u v
One approach to dealing with this could be to consider a larger class of shapes
than the almost skew-shapes. For instance, we could consider shapes λ/µ, with
Resolution of determinantal ideals 199

λ and µ both almost partitions of length n, such that µ ⊆ λ. It would then


have to be checked in detail that everything found in Chapter VI carries through
for these shapes, and that all shapes manufactured in the filtration are of the
same kind. In order to be able to apply to these shapes the same inductive
argument used in the proof of Theorem VII.1.3, we could require in addition
that τλ + τµ < n − 1, where τλ and τµ are the types of the almost partitions
λ and µ. However, as we will have no use in the rest of the book for a more
general form of Theorem VII.2.1, we leave this as an open question.
We next list three immediate corollaries of Theorem VII.2.1.
Corollary VII.2.2 Let ϕ : G → F be any map of free modules, and let µ ⊆ λ
be partitions. Denote by (Kλ/µ (ϕ))j the component in degree j of the complex
Kλ/µ (ϕ). There is a natural filtration on (Kλ/µ (ϕ))j whose associated graded
module is

Lν/µ (F ) ⊗R Kλ/ν (G),
µ⊆ν⊆λ, |λ|−|ν|=j

where |λ| stands for the sum, λi , and similarly for ν.
Proof Cf. Reference [3], corollary 5.1.14. 2
We remind the reader that a complex is acyclic if all the positive homology
modules vanish.
Corollary VII.2.3 Let ϕ : G → F be a split injection with (projective) coker-
nel E. Then Kλ/µ (ϕ) is acyclic and, when E is free, H0 (Kλ/µ (ϕ)) = Lλ/µ (E).
Proof Since it is sufficient to prove acyclicity locally, we may assume R is local,
that F = G ⊕ E, with E free, and that our map ϕ is the direct sum of ϕ1 = id :
G → G and ϕ2 : 0 → E. Hence by Theorem VII.2.1,  we see that
 up to filtration,
the homology of Kλ/µ (ϕ)) is the direct sum of H Kν/µ (ϕ1 ) ⊗R Lλ/ν (E). By
Theorem VII.1.3, it follows that all of these homology modules are zero except
when ν = µ, and this gives the acyclicity as well as the result that H0 (Kλ/µ (ϕ)) =
Lλ/µ (E). If E is free to begin with, the result that H0 (Kλ/µ (ϕ)) = Lλ/µ (E) is
also true globally. 2
By essentially the same method of proof one obtains the following result.
Corollary VII.2.4 If ϕ = ϕ1 ⊕ ϕ2 , where ϕ1 is an isomorphism, then the
complex Kλ/µ (ϕ) is homotopically equivalent to Kλ/µ (ϕ2 ).

VII.3 Resolution of determinantal ideals


When, in the early 1960s, the hunt was on for the generalized Koszul complex,
the hope was that it would be possible, and not too difficult, to find resolutions
of the ideals of p × p minors of the generic m × n matrix (m ≥ n). In fact, the
generalized Koszul complexes described in Chapter III do the job for the n × n,
200 Some applications of Weyl and Schur modules

or maximal, minors of the matrix. But if, using the notation of Chapter III, we
have the map f : A → B, with A = Rm , B = Rn , then we can see that the
cokernel of the map mq (f ) : Λq A ⊗R Λr B → Λq+r B has support equal to that of
the ideal generated by the minors of order n − r (where mq (f ) is the composition
Λq f ⊗R id
Λq A ⊗R Λr B −→ Λq B ⊗R Λr B → Λq+r B). Thus there was the hope that
one could find some way of resolving this cokernel for all r in the same way as
one did the case when r = 0, that is, the case of the maximal order minors. Alas,
the problem turned out to be far more complex than imagined, but fortunately
it did lead to a number of interesting developments, some of which we will try
to cover in the subsequent subsections.

VII.3.1 The Lascoux resolutions


We saw in Chapter III that the hooks had come into play in connection with the
generalized Koszul complexes, and it became clear to a number of those working
in this area that the terms of a resolution of an ideal generated by minors should
be representations of the appropriate general linear groups. Before we proceed
further, let us explain why this is so by examining in some detail just what we
really mean by “generic map.” (Also cf. Section V.2.)
Consider two free R-modules, F and G of ranks m and n. What do we mean
by the generic map from one to the other? For notational convenience, it will
be easier to talk about a “generic map” from F to G∗ , rather than from F
to G (obvious modifications can be made to make it work from F to G). Let
S = S(F ⊗R G) be the symmetric algebra associated to the free R-module F ⊗R G,
and let F = S ⊗R F, G∗ = S ⊗R G∗ . In order to define a map from F to G∗ , it is
enough to define a map from F to F ⊗R G ⊗R G∗ . This is so because S0 = R and
S1 = F ⊗R G. Therefore, if we have the map from F to F ⊗R G⊗R G∗ , this means
we have a map F → S1 ⊗R G∗ . Hence, to define the map Sν ⊗R F → Sν+1 ⊗R G∗ ,
we simply take the composite

Sν ⊗R F → Sν ⊗R S1 ⊗R G∗ → Sν+1 ⊗R G∗ ,

where the last map is just the multiplication in the symmetric algebra.
But the map F to F ⊗R G ⊗R G∗ simply sends an element x ∈ F to the
element x ⊗ cG ∈ F ⊗R G ⊗R G∗ , where cG ∈ G ⊗R G∗ is the element (as in
Chapter V) corresponding to the identity map in HomR (G, G) under the natural
identification of G⊗R G∗ with HomR (G, G). If we now choose bases {x1 , . . . , xm }
for F , {y1 , . . . , yn } for G, and set Xij = xi ⊗ yj ∈ S1 , it is easy to see that the
matrix corresponding to our map, using {1 ⊗ xi } and {1 ⊗ yj } as bases for F and
G∗ , is simply (Xij ). The upshot of all of this is that the “generic” map is defined
without recourse to a basis, so that it seemed natural to suppose that the modules
that would occur in a resolution of the ideal, Ip , generated by minors of order p
(which also can be defined without recourse to basis), should be independent of
a choice of basis, and hence should be representations of GL(F ) × GL(G).
Resolution of determinantal ideals 201

However, it was not until the mid-1970s that A. Lascoux found a way to expli-
citly use representation theory to describe the resolutions of these determinantal
ideals [64]. Working in characteristic zero, where every representation of the gen-
eral linear group splits into a direct sum of irreducible representations, he was
able to write down the GL(F ) × GL(G)-irreducible components of the repres-
entation modules Xkp that occur as the chains in dimension k of the resolutions
of the determinantal ideal Ip , the ideal of p × p minors of the generic m × n
matrix. Since the irreducible representations of GL(F ) × GL(G) are of the form
M ⊗R N , where M and N are irreducible representations of GL(F ) and GL(G),
resp., and the irreducible representations would be of the form Lλ F and Lµ G
for partitions λ, µ, the question to be answered is: for which pairs of partitions
λ, µ do the tensor products Lλ F ⊗R Lµ G appear as direct summands in Xkp ?
A. Lascoux determined what this family of pairs is by using a clever geometric
argument. First he “desingularized” the variety of Ip , by going up to the appropri-
ate Grassmannian; the ideal of this desingularized variety has a Koszul complex
as its resolution. Then, by Bott’s theorem on the vanishing of the cohomology of
homogeneous vector bundles ([14], theorem 4), the higher direct image functors
on the Koszul complex can be computed, and the resulting complex is the min-
imal resolution of Ip . While the desingularization technique holds in arbitrary
characteristic, Bott’s theorem is valid only in characteristic zero, which partially
explains why the Lascoux approach seems to be limited to that case. (We will
return to this point in a later section.) This geometric approach was originated
by G. Kempf in the 1970s [57]; a full and very clear discussion of this and similar
techniques can be found in J. Weyman’s book [86].
It should be pointed out here, for the sake of historical accuracy, that an
explicit description of the boundary maps of the Lascoux resolutions was not
given by A. Lascoux. Subsequent work by H. Nielsen [68] and P. Pragacz and
J. Weyman [70] was needed to clear this up.
To describe the partitions that do appear in the Lascoux resolutions, we
need some additional notation and terminology. We say that a partition, λ, has
Durfee square of size k if λi ≥ k for i = 1, . . . , k, and λk+1 ≤ k. We denote
the transpose partition of λ as usual (see Chapter VI), by λ,  so that we write
 = (λ
λ 1 , . . . , λ
t ) where λ
j = #{i| λi ≥ j}. It is clear that λ and λ
 have the same
Durfee square size. Finally, if λ and µ are partitions, we denote by λ + µ the
partition (λ1 + µ1 , . . . , λq + µq ).
We are now in a position to state (without proof) what the terms of the
Lascoux resolution are.

Theorem VII.3.1 Let Xkp denote the module of k-chains in the resolution of
Ip . Then

Xkp = Lλ+(p−1)k(λ) F ⊗R Lλ+(p−1)
 k(λ) G,

λ∈Λk
202 Some applications of Weyl and Schur modules

where k(λ) is the Durfee square size of λ, (a)k(λ) denotes the rectangular partition
(a, . . . , a), and Λk is the set of partitions, λ, such that |λ| = k.
% &# $
k(λ)

If one chooses p = 1, that is, the ideal generated


 by the “indeterminates”
Xij , one sees that the terms of the resolution are λ∈Λk Lλ F ⊗R Lλ G and, in
characteristic zero, this is the classical Cauchy decomposition of Λ(F ⊗R G).
What we see is that the terms for arbitrary p are the terms that one would get
for p = 1, modified by the adjunction of rectangular pieces of dimensions p − 1
by k(λ).

VII.3.2 The submaximal minors


While there are clearly a few technicalities connected to the preceding resolu-
tions whose clarification would be desired, it is nevertheless evident that this
work of A. Lascoux aroused a great deal of interest and hope. The hope resided
in the expectation that the Schur modules that he used were replicable in a
characteristic-free setting, and that, while Bott’s theorem is not true in positive
characteristic, its non-validity would not adversely affect the outcome of Lascoux’
construction, even if it obviated the use of the same geometric approach that he
had employed so effectively.
There had been some work already on reproducing Schur and Weyl modules
in arbitrary characteristic (see [34], and [84]), but the approach in Reference [3],
and presented in this book, proved to be most convenient for the purposes at
hand. One of the most, if not the most, important reason for this is that these
modules can be described by means of generators and relations, that is, an expli-
cit presentation can be given, and this makes it easier to compute them and with
them. In addition, as one saw in the preceding chapter, one can define these mod-
ules for a larger category of shapes than just partitions, and this allows for the
use of homological methods, as illustrated by the theorem on fundamental exact
sequences (Theorem VII.1.2). In any event, the terms of classical representation
theory were constructible in a characteristic-free setting, and one could begin
to carry out a “Lascouxification” of the problem of resolving the determinantal
ideals.
While the case p = 1 gave us the Koszul complex (or at least its terms), and
the case p = n gave us the Eagon–Northcott complex [40], both of which could be
defined in a characteristic-free way, it was not clear just how to produce a resol-
ution for the intermediate values of p, nor was it clear that the boundary map of
such a resolution, assuming it could be built out of nice representation modules,
would be “natural,” as is true in the extreme cases. Some simple examples suf-
ficed to show that a direct parroting of the Lascoux complexes over the integers
did not lead to acyclicity; there was torsion in the integral homology. But it was
possible to construct a characteristic-free resolution of the ideal of submaximal
minors (see [4]), that is, the minors of order n − 1 (recall that we are assum-
ing throughout that m ≥ n). In this case, though, for arithmetic considerations
anticipated due to the previously mentioned torsion, one had to choose Z-forms
Resolution of determinantal ideals 203

of the terms that appear in characteristic zero (the term Z-form will be explained
in the next subsection) in order to push the construction through.
As in the previous discussion, we will skip the details, but two observations
emerged from this project: one was that the boundary map which was explicitly
constructed was not “natural,” that is it was not equivariant with respect to
the action of the general linear groups involved; another was that it was nev-
ertheless possible to build the terms of the resolution in a way suggested by
the Lascoux approach (although many more terms might have to be involved—
more about this in a later section). An important additional observation was
that fairly delicate arithmetic considerations were involved in the construction
of these complexes, and this led to an independent study of the Z-forms which
we will discuss in the next subsection.

VII.3.3 Z-forms
Let F be a free abelian group of rank m, and let F be its extension to the
rationals, that is, F = Q ⊗Z F . We know that D2 F and S2 F are both GL(F )-
representations , where GL(F ) means the general linear group over the integers,
Z. Furthermore, D2 F = Q ⊗Z D2 F, S2 F = Q ⊗Z S2 F , and these are GL(F )-
representations. We know that there is a GL(F )-equivariant map from D2 F to
S2 F , namely the composition:
∆ m
D2 F −→ F ⊗Z F −→ S2 F,

where ∆ is the diagonal map, and m is the usual multiplication map. However,
this is not an isomorphism of the two integral representations. Nevertheless, the
corresponding map over the rationals is an isomorphism. We therefore say that
D2 F and S2 F are Z-forms of the same rational representation (in this case, S2 F ).
So, we are led to make the following definition.

Definition VII.3.2 Let F be a free abelian group. Two GL(F )-representations


are Z-forms of the same representation if, when tensored with the rationals,
Q, they are isomorphic GL(F )-representations.

In Subsection VII.3.2, we indicated that in the construction of the resolution of


the ideal of submaximal minors, one had to choose certain Z-forms of the terms
in the Lascoux resolution in order to get an integral complex which was acyclic.
These representations that arose were Z-forms of certain hooks, and they came
about in the following way.
We know that for all l > 0 the complexes

0 → Λl F → Λl−1 F ⊗Z F → · · · → Λl−t F ⊗Z St F → · · · → F ⊗Z Sl−1 F → Sl F → 0

are exact, where the boundary map is given by diagonalizing the exterior powers
and multiplying the symmetric powers. But what happens if we replace the
204 Some applications of Weyl and Schur modules

symmetric powers by divided powers, that is, if we consider the complex

0 → Λl F → Λl−1 F ⊗Z F → · · · → Λl−t F ⊗Z Dt F → · · · → F ⊗Z Dl−1 F → Dl F → 0

where we still diagonalize the exterior powers and multiply, this time, into the
divided powers (something that we have avoided doing so far throughout this
book)? As the reader may strongly suspect, this complex is no longer exact;
the surprising thing, however, is that, counting from the left, it is exact up to
the middle of the complex, that is, from t = 0 to t = [(l − 1)/2] where [x]
indicates the integral part of x ([4], proposition 2.22). As a result, the cycles
of this complex are Z-forms of the corresponding cycles of the complex above,
involving the symmetric powers, and these, as indicated in Chapter III, are just
the hooks.
Another, simpler, way to construct non-isomorphic Z-forms is the following:
Consider the short exact sequence

(†) 0 → Dk+2 → Dk+1 ⊗Z D1 → K(k+1,1) → 0

where K(k+1,1) is the Weyl module associated to the hook partition (k + 1, 1).
(We are leaving out the module F , as that is understood throughout.)
If we take an integer, t, and multiply Dk+2 by t, we get an induced exact
sequence and a commutative diagram:
0 → Dk+2 → Dk+1 ⊗Z D1 → K(k+1,1) → 0
↓t ↓ ↓
0 → Dk+2 → E(t; k + 1, 1) → K(k+1,1) → 0,

where E(t; k + 1, 1) stands for the cofiber product of Dk+2 and Dk+1 ⊗Z D1 .
Each of these modules is a Z-form of Dk+1 ⊗Z D1 , but for t1 and t2 , two
such are isomorphic if and only if t1 ≡ t2 mod k + 2 (see [2]). This says that
Ext1A (K(k+1,1) , Dk+2 ) ∼
= Z/(k + 2), where A is the Schur algebra of appropriate
degree (we will discuss the calculation of Ext in more detail in Subsection VII.4.1
on intertwining numbers).
Although we have not introduced the Schur algebra until now, it plays an
important role in representation theory as, classically, the Schur algebra of degree
k is the universal enveloping algebra for the homogeneous polynomial representa-
tions of degree d. For example, Dd and Λd are such representations of the general
linear group. As a result, when we speak of “equivariant maps” of representa-
tion modules, what we really mean are linear maps over the Schur algebra; the
Schur algebra is the underlying ring over which all of our representations are
modules.

Definition VII.3.3 Let F be a free R-module of rank n, and d a positive


integer. The symmetric group, Sd , acts on the left of F ⊗d = F ⊗R · · · ⊗R F by
% &# $
d
permuting factors. The Schur algebra of degree d is the algebra, AR (n, d), of
Resolution of determinantal ideals 205

endomorphisms of F ⊗d which commute with the action of Sd . That is,


AR (n, d) = EndSd (F ⊗d ).
We will usually just write A for AR (n, d) when the context is clear.
It can be seen without too much difficulty that this algebra is, as a mod-
ule, simply Dd (EndR (F )) [2]. (It is enough to recall that for any free module,
V , Dd (V ) is the submodule of V ⊗d invariant under the operation of the sym-
metric group Sd .) The description of the ring structure on A in this guise
is a bit less intuitive. It is based on the fact that for any free module, V ,
and any integer d, there is a map of Dd (V ) ⊗R Dd (V ) into Dd (V ⊗R V )
(the dual map of the projection Sd (V ⊗R V ) → Sd (V ) ⊗R Sd (V ) associated
with the filtration in Reference [3], Section 3.1). If V happens to be a ring,
as is the case when V = EndR (F ), we have a map of Dd (V ⊗R V ) into
Dd (V ) induced by the multiplication map V ⊗R V → V . The composite map
Dd (EndR (F )) ⊗R Dd (EndR (F )) → Dd (EndR (F ) ⊗R EndR (F )) → Dd (EndR (F ))
gives the multiplicative structure on the Schur algebra. All of this is discussed in
detail in Reference [2], as well as Reference [45]. In both of the indicated refer-
ences, it is actually proved that the Schur algebra, AR (n, d), is the direct sum of
the modules Da1 (F ) ⊗R · · · ⊗R Dan (F ), as (a1 , . . . , an ) runs through all integral
weights of rank d (that is, all n-tuples of non-negative integers adding up to d).
As a result, the modules Da1 (F ) ⊗R · · · ⊗R Dan (F ) are projective modules over
AR (n, d). (In Reference [2] it is shown that if we take the tensor product of more
than n factors, this tensor product need not be projective over AR (n, d).)
This, then, is the definition of the Schur algebra; when we talk about functors
like Hom or Ext for representation modules, the base ring will usually be the
Schur algebra, and we will write HomA or ExtA as one does for modules over
rings. The Schur algebra, by the way, is not commutative.
One more observation about the Schur algebra 3before we go on to other
applications. If M is an A-module, then since A ∼ = Da1 (F ) ⊗R · · · ⊗R Dan (F ),
and M = HomA (A, M ), we have a decomposition of M as the direct sum

M∼ = HomA (Da1 (F ) ⊗R · · · ⊗R Dan (F ), M ).

The modules HomA (Da1 (F ) ⊗R · · · ⊗R Dan (F ), M ) are no longer A-modules;


they are R-modules. However, these are known as weight submodules of M . In
terms of classical representation theory, these are defined to be the submodules
of M consisting of those elements m ∈ M with the property that if T is the
diagonal matrix:
 
t1 0 · · · 0
 0 t2 · · · 0 
 
T = . .. . . . ,
 .. . . .. 
0 0 · · · tn
then T m = ta1 1 ta2 2 · · · tann m.
206 Some applications of Weyl and Schur modules

It is not difficult to see that if M is either a Weyl or Schur module associated


with a skew-shape, then the weight submodule corresponding to the weight α =
(a1 , . . . , an ) is generated by all standard tableaux of content α. By a standard
tableau of content α, we mean a standard tableau in which the entries consist
of ai copies of the ith basis element for every i. Thus, for example, the weight
submodule of Λ2 corresponding to the weight (2, 0, . . . , 0) is zero (i.e., there are
no maps from D2 to Λ2 ), while the weight submodule of Λ2 corresponding to the
weight (1, 1, 0, . . . , 0) is the submodule generated by the element x1 ∧x2 , assuming
that we have chosen the basis {x1 , . . . , xn } for the underlying free module, F .
It is understood here that by “maps” we mean A-maps, where A = AR (n, 2) in
this case (since Λ2 is a representation of degree 2).

VII.4 Arithmetic considerations


We have already seen that the Weyl and Schur modules are universal, so that if
we define them over the integers and then tensor them with a commutative ring,
R, we get the corresponding module over the ring R. We also see that there is
some interest in looking at the Ext groups of certain Weyl and/or Schur modules.
The Universal Coefficient Theorem tells us how the Ext groups behave under ring
extension. The proof of this theorem is found in Reference [2], theorem 5.3.
Theorem VII.4.1 (Universal Coefficient Theorem) Let R be a commu-
tative hereditary ring, R → R a homomorphism of commutative rings, and A an
R-algebra. Let M and N be left A-modules which are free R-modules. Further-
more, assume that M has a resolution, P, over A by finitely generated projective
A-modules. Then there is a short exact sequence of R-modules
0 → R ⊗R ExtiA (M, N ) → ExtiA (M , N ) → TorR
1 (R, ExtA (M, N )) → 0
i+1

for each i ≥ 0 where we have set Y = R ⊗R Y , for Y = A, M and N .


In particular, then, if R = Z and R = Z/(p) for some prime p, we have the
exact sequence:
0 → Z/(p) ⊗Z ExtiA (M, N ) → ExtiA (M , N ) → TorZ1 (Z/(p), Exti+1
A (M, N )) → 0,

where A denotes the appropriate Schur algebra over Z, and A denotes Z/(p)⊗Z A.

VII.4.1 Intertwining numbers


The exact sequence immediately above is particularly applicable in modular rep-
resentation theory in relation to the problem of calculating intertwining numbers.
A standard question in modular representation theory is the following.
If we let λ = (λ1 , . . . , λq ) be a partition, and let µ = (µ1 , . . . , µq ) be the
partition obtained from λ by attaching d boxes from the bottom row of λ to
its top row, that is, µ1 = λ1 + d, µ2 = λ2 , . . . , µq = λq − d, then what is the
Arithmetic considerations 207

Z/(p)-dimension of the Z/(p)-vector space ExtiA (Kλ , Kµ ) where the notation is


as indicated above. These numbers are called intertwining numbers. From the
exact sequence above, we see that it suffices to calculate the integral Ext groups,
since the modular ones are simply the p-torsion part of one integral Ext plus the
reduction modulo p of another.
Conjectures about these numbers were first made by D. Kazhdan and G.
Lusztig [56] (in this case, the conjectures concerned coefficients of the so-called
Kazhdan–Lusztig polynomials), then by G. Lusztig [65] for modular representa-
tions (where the connection was made between the Kazhdan–Lusztig coefficients
and the intertwining numbers for modular representations). A fine presenta-
tion of these conjectures and results on their partial solution can be found in
Reference [79]. As is so often the case in conjectures about characteristic p, the
solutions that have been obtained relate to “sufficiently large primes.” One might
hope that the study of these numbers through the use of the integral represent-
ations would throw more light on whether one must really limit the size of the
primes and, if so, what that limit is due to.
To indicate how the universal coefficient theorem (UCT) can be used in just
a small special case, let us suppose that we want to evaluate the dimension of
Ext0A (Kλ , Kµ ), that is, of HomA (Kλ , Kµ ). The UCT tells us that we should first
look at HomA (Kλ , Kµ ). But HomA (Kλ , Kµ ) is a subgroup of HomZ (Kλ , Kµ ),
and since these Weyl modules are free abelian groups, their Hom over Z is free.
Thus HomA (Kλ , Kµ ) is a free abelian group, and has the same rank as its exten-
sion to the rationals. But over the rationals, the Weyl modules are irreducible, so
that there are no nontrivial AQ -maps between them. Thus, the Hom group over
the integers is also zero, and what is left for us to do is calculate Ext1A (Kλ , Kµ ).
For once we have that, we simply have to take its p-torsion to get our result.
To find the p-torsion of a finite abelian group is indeed fairly simple, so the
rub in the above discussion is how to find Ext1A (Kλ , Kµ ). The straightforward
answer of any halfway respectable homologist is, find an A-projective resolution,
P, of Kλ , Hom it into Kµ , and take the resulting homology in dimension one.
We saw in Subsection VII.3.3 that certain tensor products of divided powers
are A-projective, and resolutions by such modules have the advantage that one
can calculate (in principle) the result of taking Hom of such modules into our
Weyl modules. In Reference [2], it was proven that projective resolutions, P,
of this desired type exist (we will return to this in Section VII.6), but to actu-
ally calculate the homology of the Hom complex, we need explicit information
about the boundary maps of these resolutions as well as the explicit descrip-
tion of the terms. One simplification we meet in this calculation comes from
the fact that HomA (P, Kµ ) is a complex of free abelian groups, and that its
cohomology is torsion (since over the rationals the cohomology is zero). If we
write down the boundary maps in HomA (P, Kµ ) as integral matrices, elementary
arguments show us that the cohomology groups are determined by the invari-
ant factors of those matrices (the non-zero invariant factors). To calculate those
208 Some applications of Weyl and Schur modules

matrices, we clearly must know the maps in the projective resolution, P. We


will do one example to illustrate how such calculations may be done. The reader
is referred to References [2], [26], and [30] for more complete discussion of this
problem.
In Subsection VII.3.3, we made the assertion that the exact sequence (†) is
exact. This is a resolution of the degree k + 2 representation, K(k+1,1) ; but how
do we know that? In this case, it is fairly simple to see it, since we do have the
exactness of the complex

0 → Dk+2 → Dk+1 ⊗Z D1 → Dk ⊗Z Λ2 → · · · → D1 ⊗Z Λk+1 → Λk+2 → 0

and, in Chapter III, we saw that the hooks were the cycles in this complex.
(We will soon see other ways of arriving at this fact.) In any event, to compute
Ext1A (K(k+1,1) , Dk+2 ) (this, after all, is the situation for λ = (k + 1, 1), d = 1
and µ = (k + 2, 0) of our discussion), we simply look at

0 → HomA (Dk+1 ⊗Z D1 , Dk+2 ) → HomA (Dk+2 , Dk+2 ) → 0

and compute the invariant factors of the matrix of the map

HomA (Dk+1 ⊗Z D1 , Dk+2 ) → HomA (Dk+2 , Dk+2 ).

From our discussion of weight submodules, we see that both of these abelian
groups are free of rank one; the question is to identify bases for these groups and
make the map explicit in terms of these bases. A simple calculation shows that
the basis for HomA (Dk+2 , Dk+2 ) is the identity map, α, say, while the basis for
HomA (Dk+1 ⊗Z D1 , Dk+2 ) is the multiplication map, say β. The map from Dk+2
to Dk+1 ⊗Z D1 , call it δ, is just the diagonal map. Thus, the map β is carried
(k+2) (k+2)
to βδ, and we see that βδ(x1 ) = (k + 2)x1 = (k + 2)α. The matrix
in question, then, is the one-by-one matrix, (k + 2), and so we get the result
announced earlier, namely that Ext1A (K(k+1,1) , Dk+2 ) ∼ = Z/(k + 2).
While this calculation was quite simple, this is not the case in general. As of the
current writing, the only complete results that are known are for two- and three-
rowed partitions, and only for Ext1A . Again we refer the reader to References [2],
[26], and [30] for explicit calculations leading to the known results, but as we
see, there is a strong incentive to get our hands on an explicit description of
projective resolutions of Weyl modules. The reader may be interested to know
that a newer approach to calculating the invariant factors of the matrices that
we do encounter involves the use of letter-place methods, but this development
is still in its infancy (see [22], section 4).

VII.4.2 Z-forms again


The example that we did in the previous subsection shows that we have at least
k + 2 distinct Z-forms for Dk+1 ⊗Z D1 . Of course, over the rationals, the divided
Hashimoto counterexample 209

powers are the symmetric powers, and Dk+1 ⊗Z D1 splits up into the direct sum
of K(k+1,1) and Dk+2 . Over the integers, the exact sequence (†) not only does
not split, but that short exact sequence turns out to be the generator of Ext.
Clearly there is a great deal of arithmetic involved in our integral representations
of the general linear group over Z.
The following example will also illustrate how these Z-forms are directly con-
nected with resolutions of Weyl modules. In Section VII.6 we will see that a
projective resolution of K(k,2) is given by

Dk+1 ⊗Z D1
0 → Dk+2 → ⊕ → Dk ⊗Z D2 → 0,
Dk+2
where the map from each component of the split term is the appropriate diag-
onalization (possibly followed by multiplication), the map Dk+2 → Dk+1 ⊗Z D1
is the diagonal, and the map Dk+2 → Dk+2 is multiplication by 2. Thus the
cokernel of the left-hand map is the module we denoted by E(2; k + 1, 1) in
Subsection VII.3.3. We therefore have the short exact sequence
0 → E(2; k + 1, 1) → Dk ⊗Z D2 → K(k,2) → 0,
and over the rationals, the sequence would look the same, except that there we
could replace E(2; k +1, 1) by Dk+1 ⊗Z D1 , which gives the anticipated resolution
of the Weyl module in characteristic zero (see [64]). From the point of view of the
Grothendieck ring of representations, this tells us that while over the rationals
the class of K(k,2) may be represented as [Dk ⊗Z D2 ] − [Dk+1 ⊗Z D1 ], over the
integers it would be represented by [Dk ⊗Z D2 ] − [E(2; k + 1, 1)].
The thrust of the above discussions is that there are many reasons for trying
to find an explicit description of projective resolutions of Weyl modules.

VII.5 Resolutions revisited; the Hashimoto counterexample


When we put the material of the preceding two sections together, we see
that the construction of resolutions of determinantal ideals, particularly for
the minors of lower order, may involve arithmetic subtleties. In fact, they
arose very explicitly in the construction of the resolution of the submaximal
minors and, in subsequent work on the minors of order n − 2, that is the
subsubmaximal minors, the intricacies of the Z-form arithmetic became so oner-
ous that there was no attempt to push through the full construction in that
case. It was obviously the time to sit back and rethink the approaches to the
problem.
It had always been clear that the existence of a universal (that is, defined over
Z; also called generic) minimal free resolution of the ideal, Ip , of p × p minors
would imply that the Betti numbers of that ideal are independent of character-
istic (by Betti numbers we mean the dimensions over the ground field—Z/(p)
or Q in our case—of the vector spaces of chains of a minimal resolution). But since
210 Some applications of Weyl and Schur modules

these ideals were known to be perfect (i.e., their homological dimension is equal
to their depth) in every characteristic (see [38], [39]), the implications about
Betti numbers seemed perfectly (sic) natural. That is, while certain construc-
tions were universal, they lacked universal minimal free resolutions, but those
that were known “violators” were not perfect in all characteristics (such as the
projective plane, which is perfect in every characteristic but two [51], [72]). Some
attempts were made to prove this apparently true fact about the Betti numbers
by less cumbersome techniques and, in 1987, K. Kurano [62], proved that the
first Betti numbers of Ip were independent of the characteristic.
But in 1988, M. Hashimoto announced (in a preprint which later became Ref-
erence [47]) the surprising fact that for the generic 5 × 5 matrix, the second Betti
number of I2 in characteristic three is one greater than it is in characteristic zero.
(His paper [47] actually proved that the ideal Iq of the (q + 3) × (q + 3) matrix
has its second Betti number in characteristic three larger than in character-
istic zero.) Immediately following M. Hashimoto’s announcement, J. Roberts and
J. Weyman [73] published another proof that had the same geometric flavor as
the Lascoux construction of his resolutions of determinantal ideals (see also [86]).
We will not go into their proof in detail, but a few remarks are worth discuss-
ing. The first observation made by Roberts and Weyman is that the geometric
part of the Lascoux approach is characteristic-free. That is, the desingularization
that is used over the Grassmannian, and the fact that the ideal of this desin-
gularization is resolved by a Koszul complex, are valid over any ground ring. It
is in the application of Bott’s theorem, that is, the second step of the Lascoux
program, that the procedure breaks down (as one might suspect, since the Bott
theorem is not true in positive characteristic). But what Roberts and Weyman
do is identify the higher direct images that come up in the Bott theorem with
the homology of certain canonical subcomplexes of the Weyl–Schur complexes.
The homology of these complexes is directly related to the Betti numbers of the
determinantal ideals, and can be computed by means of representation theory
(at least theoretically).
In the Hashimoto example, starting with such facts as that the representation
Λ3 is contained in the representation K(2,1) in characteristic three, it is possible
to see that there is a linear two-cycle that cannot be covered by a degree one map.
Hence the second Betti number goes up in characteristic three. This observation
makes sense in the light of the specific construction of the resolutions involved;
we will forbear to go into those details here.
It suffices to say that this example of M. Hashimoto took the steam out
of the project of looking for universal minimal free resolutions of determin-
antal ideals. It did, however, open up a few new areas of investigation, and
did use in an essential way the characteristic-free approach to Weyl modules
and their resolutions. Finally, the relationship between Bott’s theorem and the
homology of certain subcomplexes of Weyl–Schur complexes that is evident in
the Roberts–Weyman approach to M. Hashimoto’s example should be studied
carefully.
Resolutions of Weyl modules 211

VII.6 Resolutions of Weyl modules


In the discussions of intertwining numbers and of Z-forms, as well as in the above
section, we have seen references to and examples of resolutions of Weyl modules.
As the reader may well imagine, the history of the search for a description of
these resolutions is a fairly long one. In this section, we will deal with some of
this history, as well as give a brief description of some of the latest results. A
good deal of what appears here can be found in Reference [30].
In the early 1980s, work was begun [1] on the problem of resolving Schur
modules in terms of direct sums of tensor products of exterior powers (or the
so-called fundamental representations). In those years, Schur modules and funda-
mental representations were the focus since the terms of the Lascoux resolutions
of determinantal ideals were expressed in terms of Schur modules, and the
Lascoux description in characteristic zero of the “resolutions” of Schur mod-
ules was given in terms of the fundamental representations. Therefore, in the
early references, most of the literature is written in those terms. It transpired
that as work progressed and difficulties became more numerous, the question
arose as to whether these resolutions really existed. To settle this question
without having to actually construct the resolutions, a natural isomorphism
Ω : HomA (Dλ (F ), Dµ (F )) → HomA (Λλ (F ), Λµ (F )) was defined in Reference [2]
which allowed one to dualize the resolutions of the Kλ (F ), given in terms of
sums of tensor products of divided powers, to get the resolutions of the Lλ (F ) in
terms of tensor products of exterior powers. (By Dλ we mean Dλ1 ⊗R · · ·⊗R Dλn ,
where λ = (λ1 , ..., λn ). By Λλ we mean Λλ1 ⊗R · · · ⊗R Λλn .) We will therefore
focus, as we have throughout the book, on Weyl modules and resolutions in
terms of tensor products of divided powers. We add here, as a purely philoso-
phical comment, that we have grown accustomed to regarding the Weyl modules
as “more fundamental” than the Schur modules because of the fact that the
tensor products of divided powers are projective, and the weight submodules of
representations are expressible in terms of them.
For two-rowed Weyl modules, using the fundamental exact sequence,

p+t+1 t+1 p t p
0→ → → → 0,
q−t−1 q q

and an induction argument on the number of overlaps between the two rows,
namely, the integer q − t as read from the above diagram, it was possible to give
a description of the desired type of resolution [1, 28] . This will be described in
Subsection VII.6.2.
The three-rowed case presented, of course, quite another aspect, for as we
know, the kernel of the analogous surjective map is an almost skew-shape about
which nothing was known at the time. In particular, a presentation of such shapes
was not even guessed at. To be precise, the analogue of the above sequence is
212 Some applications of Weyl and Schur modules

the following:
t1 p t1 p
0→ q+t2 +1 → t2 +1 q
t2 +1 r−t2 −1 r
t1 p
→ t2 q → 0.
r
and while one can try to do an induction on the integer r − t1 − t2 , the type
of shape has changed on us, that is, it has become an almost skew-shape rather
than remain a skew-shape, as it so conveniently did in the two-row situation.
Hence one must also try to resolve these “new” shapes.
The approach to this in Reference [2] was to resort to a fairly complicated
spectral sequence to prove a slightly weaker form of Theorem VII.1.2.
Despite the encouragement provided by the proven existence of resolutions
of the type desired, and the fundamental exact sequences which theoretically
provide a mechanism for an induction proof as in the two-rowed case, the methods
at hand still made the calculations almost impossible to deal with. As a result,
it was not until the 1990s that this problem was looked at again from the point
of view of letter-place methods, where the Weyl map and the presentation map
were seen to be place polarizations, and the Capelli identities could play a role
in simplifying and systematizing these calculations. And then the introduction
by B. Taylor of straight tableaux made it possible to develop our theory as we
have in this book, and to prove the theorem on fundamental exact sequences in
more general form.
One other tool that was essential in describing the resolutions was the bar
complex (see [29]). As the reader can see, the strategy in building the resolutions
lies in assuming that one knows what the terms look like for shapes that are
“less complex,” with “complexity” being measured by the number of overlaps
between rows. Theorem VII.1.2 allows us to reduce this complexity by presenting
our given shape in terms of shapes that are less complex (as we did in the proof
of Theorem VII.1.3). Assuming we have resolutions of these less complex shapes,
we have a map between these resolutions (this is where the projectivity of the
resolutions comes in), and the mapping cone of this map is then a resolution
of the shape we started with. By describing the resolutions in terms of the bar
complex, a complex which behaves well with respect to forming mapping cones,
this procedure of putting together resolutions by forming mapping cones becomes
manageable. In the rest of this section we will describe some of these techniques
in more detail.

VII.6.1 The bar complex


We made heavy use of the bar complex in Section III.2. It turns out that this
complex is really a special case of a whole class of complexes, which we call
differential bar complexes as in Reference [29].
Resolutions of Weyl modules 213

The main components of the construction of such complexes are the exterior
algebra over Z, Λ(S), on a set of free generators, S, called the separators, an
algebra, A, an A-module, M , and the free product of Λ(S) with A. The algebra
Λ(S) has a natural Z2 -grading: if m is a monomial in Λ(S), that is, a product of
generators, we set |m| = 0 if m is the product of an even number of generators,
and |m| = 1 if m is the product of an odd number of generators.

Definition VII.6.1 The free product of the algebra A and the algebra Λ(S)
will be called the bar algebra on the algebra A, with set of separators S,
and denoted by Bar(A; S).

The algebra Bar(A; S) inherits a Z2 -grading as follows. Every element of


Bar(A; S) is a linear combination of elements of the form
W = w1 m1 w2 m2 · · · wk mk (∗)
where the mi are non-zero monomials in Λ(S), and where the wi are elements
of A. (The monomial mk can, of course, be equal to a scalar, and w1 may be
the identity element of A.) We set |W | = 0 if |m1 m2 · · · mk | = 0 and |W | = 1
if |m1 m2 · · · mk | = 1. One extends this definition by linearity to a Z2 -grading
of the algebra Bar(A; S). Notice that Bar(A; S) is a two-sided A-module in a
natural way.
For every finite subset T of the set S of separators, the underlying mod-
ule of the algebra Bar(A; S) has a grading, which will be called the T -grading
of Bar(A; S), and which is defined as follows. The submodule Bar(A; S; T, i)
of T -degree i is spanned by all elements of the form (∗) where the integer
i equals the total number of occurrences of separators in the set T in the
sequence (m1 , m2 , . . . , mk ). Clearly, the submodule Bar(A; S; T, i) is a two-sided
A-module.
For every separator x, there exists a unique antiderivation, ∂x , of the algebra
Λ(S), such that ∂x (x) = 1 (where 1 is the identity of the exterior algebra Λ(S)),
and ∂x (y) = 0 for y in S not equal to x. We also have (∂x )2 = 0 and ∂x ∂y = −∂y ∂x
(see [29] and [30]).
The antiderivation ∂x uniquely extends to an antiderivation of the Z2 -graded
algebra Bar(A; S), again denoted by ∂x , defined as follows. If W is as in (∗), set
∂x (x) = 1 (where 1 is now the identity of the algebra Bar(A; S)), and
∂x (W ) = w1 ∂x (m1 )w2 m2 · · · wk mk
+ (−1)|m1 | w1 m1 w2 ∂x (m2 ) · · · wk mk + · · ·
k−1
|mi |
+ (−1) i=1 w1 m1 w2 m2 · · · wk ∂x (mk ).
The antiderivation ∂x is well defined. Again, we have (∂x )2 = 0 and ∂x ∂y =
−∂y ∂x .

Definition
 VII.6.2 If T is a non-empty finite subset of S, the operator ∂T
= x∈T ∂x , is called the T -boundary operator on Bar(A; S).
214 Some applications of Weyl and Schur modules

The boundary operator ∂T maps Bar(A; S; T, i + 1) into Bar(A; S; T, i), for


i = 0, 1, 2, ....
Now let M be a left A-module. If w is an element of A, we denote the action
of w on an element v ∈ M by w(v).
Definition VII.6.3 The free bar module of the A-module M , with set of
separators S, denoted by Bar(M, A; S), is the Bar(A; S)-module Bar(A; S)⊗A M.
Remark VII.6.4 The following two observations are easy, but worth under-
lining:
1. The module Bar(M, A; S) is spanned by all elements of the form
w1 m1 w2 m2 · · · wk mk ⊗ v.
If mk = 1, then
w1 m1 w2 m2 · · · wk ⊗ v = w1 m1 w2 m2 · · · mk−1 ⊗ wk (v),
since the tensor product is taken over A.
2. For each separator, x, there is a well-defined antiderivation on the free bar
module, Bar(M, A; S), again denoted by ∂x , defined as follows:
∂x (w1 m1 w2 m2 · · · wk mk ⊗ v) = ∂x (w1 m1 w2 m2 · · · wk mk ) ⊗ v.
It is worth noting from the equation above that, when mk = x, we get (setting
W = w1 m1 w2 m2 · · · wk x) :
∂x (W ⊗ v) = w1 ∂x (m1 )w2 m2 · · · wk x ⊗ v
+ (−1)|m1 | w1 m1 w2 ∂x (m2 ) · · · wk x ⊗ v + · · ·
k−1
|mi |
+ (−1) i=1 w1 m1 w2 m2 · · · mk−1 ⊗ wk (v).
That the antiderivation ∂x on Bar(M, A; S) is well defined, is clear. Again, we
have (∂x )2 = 0 and ∂x ∂y = −∂y ∂x . 
If T is a non-empty finite subset of S, the operator ∂T = x∈T ∂x , is called
the T -boundary operator on Bar(M, A; S). As with Bar(A; S; T, i), one can
also define Bar(M, A; S; T, i).
The boundary operator ∂T maps Bar(M, A; S; T, i + 1) into Bar(M, A; S; T, i),
for i = 0, 1, 2, . . . .
Example VII.6.5 Let S be a one-element set, containing the element x. Then
the module Bar(M, A; S, i) is spanned by all elements of the form
w1 xw2 x · · · wi x ⊗ v (∗∗)
and the derivation ∂x is computed as follows:
∂x (w1 xw2 x . . . wi x ⊗ v) = w1 w2 x...wi x ⊗ v
− w1 xw2 w3 x . . . wi x ⊗ v + · · ·
+ (−1)i−1 w1 xw2 x . . . x ⊗ wi (v).
Resolutions of Weyl modules 215

Thus, the free bar module on M with a single separator gives rise to the
classical bar complex, as one sees by replacing the symbol x by the symbol “|”.
What we have called Bar(M, A; S, i) are simply the chains in dimension i.

Example VII.6.6 Let A be a connected graded algebra with identity (where


connected means that the component in degree zero is the ground ring over
which we are taking the algebra). The submodule of the module of the example
above, spanned by elements of the form (∗∗) where w1 , w2 , ..., wi are all of pos-
itive degree, gives the classical normalized bar construction of A and M (see
Section III.2) with set of separators S = {x}.

Example VII.6.7 With A and M as above, assume that the degree of w1


is greater than t. One then obtains the (t+ )-submodule (and hence, also, the
(t+ )-complex) of the normalized bar construction. If in addition M is a graded
module, if a positive integer n is fixed, and if in (∗∗) it is assumed that deg(w1 )+
deg(w2 ) + · · · + deg(wi ) + deg(v) = n, one obtains a bar module.

Definition VII.6.8 The bar module which we obtain at the end of


Example VII.6.7 is called the (t+ )-graded strand of degree n.

VII.6.2 The two-rowed case


In this subsection, we will construct the resolution for the two-rowed Weyl mod-
ule much as it is described in Reference [28]. We do this for two reasons: to
illustrate the use of the classical bar complex in our context, and to describe a
contracting homotopy for the non-negative part of the resolution and a basis for
the syzygies.
Recall that the Weyl module associated to the skew-shape

t p
(A)
q

is the image of Dp ⊗R Dq under the Weyl map. The “box map” referred to at
the very beginning of Section VI.6, and denoted by λ/µ , was described there
as the sum of place polarizations,
 (k) 
∂2,1 : Dp+k ⊗R Dq−k → Dp ⊗R Dq .
k>t k>t

If we let Z2,1 stand for the generator of a divided power algebra in one free
(k)
generator, we see that Z2,1 acts on Dp+k ⊗R Dq−k and carries it to Dp ⊗R Dq .
Thus, we may take the (t+ )-graded strand of degree q of the normalized bar
complex of this algebra acting on Dp+k ⊗R Dq−k (where the degree of the
216 Some applications of Weyl and Schur modules

second factor determines the grading) to get a complex over the Weyl module:
 (t+k ) (k ) (k )
··· → Z2,1 1 xZ2,12 x · · · xZ2,1l+1 x ⊗R (Dt+p+|k| ⊗R Dq−t−|k| ) →
ki >0
 (t+k1 ) (k ) (k )
Z2,1 xZ2,12 x · · · xZ2,1l x ⊗R (Dt+p+|k| ⊗R Dq−t−|k| ) → · · ·
ki >0
 (t+k)
→ Z2,1 x ⊗R (Dt+p+k ⊗R Dq−t−k ) → Dp ⊗R Dq → 0,
k>0

where the symbol “x” is our separator variable as described above, and |k| stands
for the sum of the indices ki . Here, the boundary operator is ∂x or, what is the
same thing, is obtained by polarizing the variable x to the element 1. This, then,
describes a left complex over the Weyl module in terms of bar complexes and
letter-place algebra. We also know from the fact that the Weyl module is the
cokernel of the box map, that the zero-dimensional homology of this complex is
the Weyl module itself.
Now the question is: how do we show that this complex is an exact left complex
over the Weyl module? In other words, that it is in fact a resolution. One way,
is to produce a splitting contracting homotopy, which is what we will do here.
Another way is to use our fundamental exact sequences and a mapping cone
argument; we refer the reader to Reference [2] for this approach.
Definition VII.6.9 With our complex given as above, define the homotopy as
follows:
 (t+k)
s0 : Dp ⊗R Dq → Z2,1 x ⊗R Dt+p+k ⊗R Dq−t−k
k>0
 4 
w 44 1(p) 2(k)
sends the double standard tableau to zero if k ≤ t, and to
w 4 2(q−k)
 4 
(k) w 44 1(p+k)
Z2,1 x ⊗ if k > t. For higher dimensions (l > 0),
w 4 2(q−k)
 (t+k ) (k ) (k )
sl : ki >0 Z2,1 1 xZ2,12 x · · · xZ2,1l x ⊗R Dt+p+|k| ⊗R Dq−t−|k| →
 (t+k1 ) (k2 ) (kl+1 )
ki >0 Z2,1 xZ2,1 x · · · xZ2,1 x ⊗R Dt+p+|k| ⊗R Dq−t−|k|
 4 
(t+k1 ) (k2 ) (kl ) w 44 1(t+p+|k|) 2(m)
is defined by sending Z2,1 xZ2,1 x · · · xZ2,1 x ⊗ to
w 4 2(q−t−|k|−m)
 4 (t+p+|k|+m) 
(t+k ) (k ) (k ) (m) w 44 1
zero if m = 0, and to Z2,1 1 xZ2,12 x · · · xZ2,1l xZ2,1 x ⊗
w 4 2(q−t−|k|−m)
if m > 0.

The proofs of the following statements are in Reference [28].


Proposition VII.6.10 The collection of maps {sl }l≥0 provides a splitting
contracting homotopy for the complex above.
Resolutions of Weyl modules 217

Theorem VII.6.11 The complex above is a projective resolution of the Weyl


module associated to the shape A, over the Schur algebra of appropriate weight.

VII.6.3 A three-rowed example


In this subsection, we describe the resolution of the three-rowed skew-shape
having one triple overlap. A full understanding of this case will help understand
the difficulties still inherent in the general case, but will also anticipate the
procedure by which one discovers the terms (if not the boundary maps) of the
general resolution. For the sake of completeness, we will include the description
of the general case at the end of the subsection.
We start with the skew-shape
t1 p1
t2 p2 ,
p3

where we assume that the number of triple overlaps is ≤ 1. This means that
p3 ≤ t1 + t2 + 1. The important thing to notice about this class of shapes is that
our fundamental exact sequence is very special in that it does not entail any
almost skew-shapes. Namely, the exact sequence we are interested in reduces to:

0 → p1 , p2 + t2 + 1, p3 − t2 − 1; t1 + t2 + 1, −(t2 + 1))
→ (p1 , p2 , p3 ; t1 , t2 + 1) → (p1 , p2 , p3 ; t1 , t2 ) → 0. (VII.6.1)

(The notation we are using above is that used in Section VI.6.) But, since the
third row of (p1 , p2 + t2 + 1, p3 − t2 − 1; t1 + t2 + 1, −(t2 + 1)) does not overlap
with the top row, we see that the Weyl module associated to this shape is iso-
morphic to that associated to the skew-shape (p1 , p2 + t2 + 1, p3 − t2 − 1; t1 +
t2 + 1, 0). Hence we have the exact sequence of skew-shapes:

0 → (p1 , p2 + t2 + 1, p3 − t2 − 1; t1 + t2 + 1, 0)
→ (p1 , p2 , p3 ; t1 , t2 + 1) → (p1 , p2 , p3 ; t1 , t2 ) → 0. (VII.6.2)

We know that the Weyl module is presented by the box map:



k>0 Dp1 +t1 +k ⊗R Dp2 −t1 −k ⊗R Dp3
:  ⊕ → Dp1 ⊗R Dp2 ⊗R Dp3 ,
l>0 Dp1 ⊗R Dp2 +t2 +l ⊗R Dp3 −t2 −l

where the maps Dp1 +t1 +k ⊗R Dp2 −t1 −k ⊗R Dp3 → Dp1 ⊗R Dp2 ⊗R Dp3 are the
kth divided
 power of the place polarization from place 1 to place 2, and the
maps l>0 Dp1 ⊗R Dp2 +t2 +l ⊗R Dp3 −t2 −l → Dp1 ⊗R Dp2 ⊗R Dp3 are the lth
divided power of the place polarization from place 2 to place 3. Again taking our
cue from the two-rowed case, but recognizing that we now need two separators
instead of one, we introduce two generators, Z2,1 and Z3,2 , with their divided
218 Some applications of Weyl and Schur modules

powers, and we write, in place of the above,


 (t1 +k)
k>0 Z2,1 x ⊗R Dp1 +t1 +k ⊗R Dp2 −t1 −k ⊗R Dp3
⊕ → Dp1 ⊗R Dp2 ⊗R Dp3 ,
 (t2 +l)
l>0 Z3,2 y ⊗R Dp1 ⊗R Dp2 +t2 +l ⊗R Dp3 −t2 −l
where x and y stand for separator variables, and the boundary map is the sum
of polarizing x and y to 1. In short, we see that we have the makings of the
free bar module on the set, S, consisting of two separators x and y, the free
associative (non-commutative) algebra, A, generated by Z2,1 and Z3,2 , and their
divided powers, and the module, M, which is the appropriate direct sum of tensor
products of divided power modules Dp ⊗R Dq ⊗R Dr for suitable p, q and r, with
the action of Z2,1 and Z3,2 (and their powers) being that of the indicated place
polarizations. The boundary map that we use is then ∂x + ∂y . However, we do
not take the free bar module on these data, but the quotient after dividing out
by the following identities (and it is here that we have the first stirrings of the
Capelli identities):
Z32 x = x Z32 , (VII.6.3)
(a) (b)
 (b−k) (a−k) (k)
Z32 Z21 x = Z21 xZ32 ∂31 , (VII.6.4)
k<b

(k)
where the symbol ∂31 means the divided power of the usual place polarization,
and its presence on the right means that it is to be brought to operate on the
element of M that is to the right of the tensor product sign. For example, if
W = W  ⊗ m, with m ∈ M, then ∂31 W means W  ⊗ ∂31 (m). To illustrate: if
(k) (k)

(c)
we have W = Z21 x ⊗ m, then
(a) (b)
 (b−k) (a−k) (c) (k)
Z32 Z21 xW = Z21 xZ32 Z21 x ⊗ ∂31 (m).
k<b

Notice that this in turn becomes


 (b−k) (c−l) (a−k−l) (l) (k)
Z21 xZ21 xZ32 ∂31 ⊗ ∂31 (m)
k<b,l<c
 k + l (b−k) (c−l) (a−k−l) (k+l)
= Z21 xZ21 x ⊗ ∂32 ∂31 (m),
l
k<b,l<c

the last equality being due to the first observation of Remark VII.6.4 (as well as
to multiplication of divided powers).
Remark VII.6.12 The reader may wonder at the fact that we are restricting
the index of summation, k, to be less than b, and l < c. We do this because in the
applications, we want our divided powers of the Z21 terms to be positive (so that
we remain in a normalized complex).
Resolutions of Weyl modules 219

(a) (b)
Definition VII.6.13 If we denote by V the element Z32 Z21 xW, where W is
(c ) (c )
in the submodule spanned by all terms of the form Z211 x · · · xZ21q x ⊗ m, denote
 (b−k) (a−k) (k)
by V the term k<b Z21 xZ32 ∂31 W . The module generated by all terms of
the form V − V will be denoted by Cap(1, 2).
This last is an inductive definition that requires a little bit of explanation. If
 (b−k) (a−k) (k)
q = 0, then V is simply k<b Z21 x ⊗ ∂32 ∂31 (m) (where we are assuming
that W is just the element m contained in a tensor product of divided powers).
(c)
If W = Z21 x ⊗ m, then
 (b−k) (a−k) (c) (k)
V = Z21 xZ32 Z21 ⊗ ∂31 (m)
k<b
 (b−k) (c−l) (a−k−l) (l) (k)
= Z21 xZ21 x ⊗ ∂32 ∂31 ∂31 (m).
k<b l<c

In order to show that we still get a complex if we factor out by the submod-
ule, Cap(1, 2), generated by the relations of the type V − V we have just been
discussing, we must show that these relations are carried into Cap(1, 2) by the
boundary map. Since the boundary map, ∂, is just ∂x + ∂y , and since the only
separator in the terms of the relations is x, we need only show that the relations
are stable under ∂x . To this end we have the following proposition, whose proof
is in Reference [30], proposition 5.1.
Proposition VII.6.14 Let W be an element of the bar module of the form
(c ) (c ) (a) (b)
Z211 x · · · xZ21q x ⊗ m, and let V = Z32 Z21 xW. Denote by V the element
 (b−k) (a−k) (k)
k<b Z21 xZ32 ∂31 W . Then we have
  (a−b)
∂ V = ∂(V ) − Z32 W .

Consequently, if a < b, then ∂(V − V ) ∈ Cap(1, 2). More precisely, we have


(a) (b) (a) (b)
∂(V − V ) = Z32 ∂(Z21 xW ) − Z32 ∂(Z21 xW )
= ∂(V ) − ∂(V ),
and we have the necessary stability of the relations.
Definition VII.6.15 We define the module Mn ((p1 , p2 , p3 ; t1 , t2 )) , for n > 0,
to be the submodule of the quotient module described above, freely spanned by all
elements of the form
 (l1 ) (l2 ) (l ) (k ) (k )
l>0 Z3,2 yZ3,2 y · · ·4 yZ3,2p yZ2,11 xZ2,12 x · · ·
 
w 44 1(π) 2(σ1 ) 3(ρ1 )
· · · xZ2,1 x ⊗  w 44 2(σ2 ) 3(ρ2 ) 
(kq )
 4 (ρ3 )
w 3
220 Some applications of Weyl and Schur modules

where

l1 > t2 ; k1 > t1 + |l|;


li > 0; kj > 0; 0 < i ≤ p; 0 < j ≤ q
 
|l| = li ; |k| = kj ;
π = p1 + |k|;
σ1 + σ2 = p2 + |l| − |k|;
ρ1 + ρ2 + ρ3 = p3 − |l|;
p + q = n.

Further, we define

M0 ((p1 , p2 , p3 ; t1 , t2 )) = Dp1 ⊗R Dp2 ⊗R Dp3 ,

and we set

M−1 ((p1 , p2 , p3 ; t1 , t2 )) = (p1 , p2 , p3 ; t1 , t2 ).

As special cases, it should be remarked that when p = 0, we have the


terms of the resolution of (p1 , p2 ; t1 ) tensored with Dp3 . When q = 0, we have
Dp1 tensored with the terms of the resolution of (p2 , p3 ; t2 ). Notice that the
tableau, with the conditions on the π, σi and ρj , simply represents an element of
Dp1 +t1 +t2 +|l|+|k| ⊗R Dp2 −t1 −|k| ⊗R Dp3 −t2 −|l| . Notice too that the map ∂x + ∂y
maps Mn ((p1 , p2 , p3 ; t1 , t2 )) to Mn−1 ((p1 , p2 , p3 ; t1 , t2 )) for n > 0. To see that,
we observe that when we apply ∂x + ∂y , we are removing either an x or a y
between the Z1,j or immediately to the left of the tensor product. It is easy to
see that in the latter case, our term remains one of the type we allow. Also,
when we remove an x between two terms of type Z21 or a y between two of
type Z32 , we simply multiply the like terms, and we still have a term of the
type allowed. The only questionable term is one in which the y between a Z32
and a Z21 has been removed. It is precisely in this case that we invoke identity
(VII.6.4) above. Of course, we have to show that when we apply that identity,
the resulting summands are of the allowed type again. But notice that

(l ) (l ) (l ) (k ) (k )
Z3,21 yZ3,22 y · · · yZ3,2p Z2,11 xZ2,12 x
 4 
w 44 1(π) 2(σ1 ) 3(ρ1 )
· · · xZ2,1 x ⊗ w 44 2(σ2 ) 3(ρ2 ) 
(kq )
 4 (ρ3 )
w 3
Resolutions of Weyl modules 221

is simply


lp
(l ) (l ) (k −α) (l −α) (k )
Z3,21 yZ3,22 y · · · yZ2,11 xZ3,2p Z2,12 x
α=0
 4 (π−α) 
w 4 1 2(σ1 ) 3(α) 3(ρ1 )
4
· · · xZ2,1q x ⊗ w
(k ) 4 2(σ2 ) 3(ρ2 ) .
4
w 4 3(ρ3 )
p−1
If we set |l | = i=1 li , we see that the exponent on the first Z21 is greater than
t1 + |l |, so that condition is preserved. The remaining ones are easy to check;
we point out that it is here that we use the fact that the index of summation in
identity (VII.6.4) is constrained.

Theorem VII.6.16 ([30], theorem 5.2) Let (p1 , p2 , p3 ; t1 , t2 ) be a Weyl mod-


ule with p3 ≤ t1 + t2 + 1. The sequence of modules and maps

· · · → Mn ((p1 , p2 , p3 ; t1 , t2 )) → Mn−1 ((p1 , p2 , p3 ; t1 , t2 )) → · · ·


· · · → M1 ((p1 , p2 , p3 ; t1 , t2 )) → M0 ((p1 , p2 , p3 ; t1 , t2 ))

is a projective resolution of (p1 , p2 , p3 ; t1 , t2 ), where the map from M0 ((p1 , p2 , p3 ;


t1 , t2 )) to (p1 , p2 , p3 ; t1 , t2 ) is the Weyl map, and the arrows are the maps
described above, that is, the maps induced by ∂ = ∂x + ∂y . We will denote this
resolution by Res((p1 , p2 , p3 ; t1 , t2 )), and continue to denote the boundary maps
by ∂.

• The condition that p3 ≤ t1 + t2 + 1 is critical to this argument because


otherwise we would not be able to invoke the exactness of the sequence (VII.6.2).
To see how important this is the reader should look at the examples in
Reference [30], section 5.

Since the description of the terms of our complex, Definition VII.6.15, seems
a bit arbitrary and cumbersome, let us take a little time to give an alternate
description. To do this we first introduce some more notation. In general we
let Zn,m stand for polarizations from place m to place n, and we are going
to suppress the various separators that come into play (as a rule, there will
be a new separator introduced for each operator Zn,m ; in the case at hand,
it will be Z32 with separator y). For a fixed choice of n, m, ξ and l, we will
denote by
ξ
Zn,m  Z n,m
(l)

the homogeneous strand of the bar complex of total degree ξ + l with initial
ξ (0)
term of degree ≥ ξ. For example, Z32  Z 32 is just the complex concentrated
222 Some applications of Weyl and Schur modules

ξ(ξ) (2)
in dimension zero consisting of the element Z32 , while Z32  Z 32 is the
complex
(ξ+1)
Z32 yZ32
(ξ) (ξ+2)
0→ Z32 yZ32 yZ32 → ⊕ → Z32 → 0.
(ξ) (2)
Z32 yZ32

If we simply write

Z (l)
n,m ; l > 0,

this will mean the homogeneous strand of the normalized bar complex of degree
l. Now a typical term of Definition VII.6.15 arises in the following way. We choose
(l)
ξ = t2 +1, and any non-negative integer l. We then form the complex Z32 t2 +1
Z 32
and tensor it with

Res ((p1 , p2 + t2 + 1 + l; t1 + t2 + 1 + l)) ⊗R Dp3 −(t2 +1+l) .

If one looks at the terms of the outcome, one sees that these are the terms
described in Definition VII.6.15. So our complex, Res ((p1 , p2 , p3 ; t1 , t2 )), may
be described as

Res ((p1 , p2 ; t1 )) ⊗R Dp3 ⊕


 (t +1) (l)
Z322 Z 32 ⊗R Res((p1 , p2 +t2 +1+l; t1 +t2 +1+l))⊗R Dp3 −(t2 +1+l) .
l≥0

The foregoing example illustrates how one goes about constructing resolutions
of Weyl modules in an “easy” case. There are many more examples and tech-
niques given in the often-cited paper, [30], which are beyond the scope of this
monograph. However, in that paper, as well as at the end of this paragraph,
the reader will find an explicit description of all the terms of the resolutions
of Weyl modules associated to any almost skew-shape. A discussion of the
obstacles to describing the boundary maps explicitly is also in that article, as
are alternative approaches that may prove to be more effective in solving this
problem. Here, then, is the explicit, inductive description of the terms of the
resolutions.

Notation (for resolutions)


For an almost skew-shape, written (p1 , . . . , pm ; t1 , . . . , tm−1 ), we denote its
resolution by Res(p1 , . . . , pm ; t1 , . . . , tm−1 ).
We assume that we know Res(p1 , . . . , pm ; t1 , . . . , tm−1 ) for m < n, and we
describe Res(p1 , . . . , pn ; t1 , . . . , tn−1 ) by induction on n.
Resolutions of Weyl modules 223

Description of Terms of Resolutions


(1) For a skew-shape (p1 , . . . , pn ; t1 , . . . , tn−1 ), the resolution looks like this:

Res(p1 , . . . , pn ; t1 , . . . , tn−1 )
= Res(p1 , . . . , pn−1 ; t1 , . . . , tn−2 ) ⊗ D(pn )
 t +1 (l )
⊕ n−1
Zn,n−1  Z n,n−1
n−1

ln−1 ≥0

⊗ Res(p1 , . . . pn−2 , pn−1 + tn−1 ; t1 , . . . tn−3 , tn−2 + tn−1 ) ⊗ D(pn −tn−1 )


 t +1 (l ) t +1 (l )
⊕ n−1
(Zn,n−1  Z n,n−1
n−1
) ⊗ (Zn,n−2
n−2
 Z n,n−2
n−2
)
ln−j ≥0

⊗ Res(p1 , . . . pn−2 +tn−2 , pn−1 +tn−1 ; t1 , . . . tn−3 +tn−2 , tn−2 +tn−1 − tn−2 )

⊗ D(pn −tn−1 −tn−2 ) ⊕

..
.
 t +1 (l ) t +1 (l ) (l )
n−1
(Zn,n−1 Z n,n−1
n−1 n−2
)⊗(Zn,n−2  Z n,n−2
n−2
) ⊗ · · · ⊗(Zn,1
t1 +1
 Z n,1
1
)
ln−j ≥0

⊗ Res(p1 + t1 , . . . , pn−1 + tn−1 ; t1 + t2 − t1 , . . . , tn−2 + tn−1 − tn−2 )

⊗ D(pn −1≤j≤n−1 tj ) ,

where tj = tj + lj + 1 for j = 1, . . . , n − 1.
(2) For an almost skew-shape (p1 , . . . , pn ; t1 , . . . , tn−1 ) of positive type k, let
us denote by s the positive integer −(tn−1 + · · · + tn−k ), and by τ, the integer
(tn−1 + · · · + tn−k−1 ). With this notation, the terms of the resolution are

Res(p1 , . . . , pn ; t1 , . . . , tn−1 )
= Res(p1 , . . . , pn−1 ; t1 , . . . , tn−2 ) ⊗ D(pn )
 (ln−1 )
⊕ Z n,n−1 ⊗ Res(p1 , . . . pn−2 , pn−1 + ln−1 ; t1 , . . . tn−3 , tn−2 + ln−1 )
ln−1 >0

⊗ D(pn −ln−1 )
 (ln−2 ) (ln−1 )
⊕ (Z n,n−2 ) ⊗ (Z n,n−1 )
ln−2 >0,ln−1 ≥0
224 Some applications of Weyl and Schur modules

⊗ Res(p1 , . . . pn−2 +ln−2 , pn−1 +ln−1 ; t1 , . . . , tn−3 +ln−2 , tn−2 +ln−1 −ln−2 )

⊗ D(pn −ln−1 −ln−2 ) ⊕

..
.
 (l ) (l ) (l )
τ +1
(Zn,n−k−1  Z n,n−k−1
n−k−1
) ⊗ Z n,n−k
n−k
⊗ · · · ⊗ Z n,n−1
n−1
)
ln−j ≥0

⊗ Res(p1 , p2 , . . . , pn−k−1 +τ +1+ln−k−1 , pn−k +ln−k , . . . , pn−1 +ln−1 ; t1 ,

. . . , tn−k−2 +τ +1+ln−k−1 , s−1+ln−k −ln−k−1 , . . . , tn−2 +ln−1 −ln−2 )

⊗ D(pn −n−1 n−1


i=n−k−1 ln−i − j=n−k−1 tj −1)

 t +1 (l ) (l ) (l ) (l )
⊕ n−k−2
(Zn,n−k−2 Z n,n−k−2
n−k−2 τ+1
)⊗(Zn,n−k−1 Z n,n−k−1
n−k−1 n−k
)⊗Z n,n−k ⊗· · ·⊗Z n,n−1
n−1

ln−j ≥0

⊗ Res(p1 , p2 , . . . , pn−k−2 + tn−k−2 , pn−k−1 + τ + 1 + ln−k−1 , pn−k + ln−k ,

. . . , pn−1 +ln−1 ; t1, . . . tn−k−3 +tn−k−2 , tn−k−2 +τ +1+ln−k−1 −tn−k−2 ,

. . . , tn−2 + ln−1 − ln−2 )

⊗ D(pn −n−1 n−1 ⊕


i=n−k−2 ln−i − j=n−k−2 tj −2)

..
.
 (l ) t +1 (l ) (l )
t1+1
(Zn,1 Z n,1
1
)⊗· · ·⊗(Zn,n−k−2
n−k−2
Z n,n−k−2
n−k−2 τ+1
)⊗(Zn,n−k−1 Z n,n−k−1
n−k−1
)
ln−j ≥0

(l ) (l )
⊗ Z n,n−k
n−k
⊗· · ·⊗Z n,n−1
n−1

⊗ Res(p1 + t1 , . . . , pn−k−2 + tn−k−2 , pn−k−1 + τ + 1 + ln−k−1 , pn−k

+ ln−k , . . . , pn−1 + ln−1 ; t1 + t2 − t1 , . . . , tn−k−3 + tn−k−2 − tn−k−3 , tn−k−2

+ τ + 1 + ln−k−1 − tn−k−2 , . . . , tn−2 + ln−1 − ln−2 ) ⊗ D(pn −n−1 tj −k) ,


j=1

where tj is defined as above.


It must be understood that if the indicated shape is not legitimate, then that
term vanishes. This puts some constraints on the parameters over which we are
summing.
Resolutions of Weyl modules 225

VII.6.4 Resolutions of skew-hooks


There is a class of skew-shapes for which it is possible to completely describe
resolutions, namely the skew-hooks (see Reference [31] as well as Reference [21]).
These shapes are essentially “hooks strung together” like this:
k1

..
. l1

..
.
kt−1

.. .
. lt−1
kt

..
. lt

The one illustrated is a t-hook, and the indices ki , li are the lengths of the
horizontals (arms) and verticals (legs), respectively. The arms always include
the left-most box, but except for the top-most arm, exclude the right-most box.
The legs always include the bottom box, but exclude the top-most (as it is already
being counted as part of the arm). We insist that the ki always be positive so
that if, for instance, k1 = 1, this means that we are essentially starting with a
vertical piece, etc. We will denote the t-hook by (k1 , . . . , kt ; l1 , . . . , lt ). Note that
if lt = 0, the skew-hook ends with an arm; we have li > 0, however, if i < t.
Thus, the skew-hook represented by (1, 3; 2, 0) looks like this:

while the one represented by (1, 3; 1, 0) is:

In order to describe the resolution of K(k1 ,...,kt ;l1 ,...,lt ) , we first introduce some
general notation. Then we will apply this to our skew-hooks and indicate why
this yields the sought-for resolution.
Let n be a fixed positive integer, and S = {s1 , . . . , sp } be a strictly increasing
set of integers, with sp = n. Denote by Xn the complex consisting of terms, in
226 Some applications of Weyl and Schur modules

dimension n − α, of the form Di1 ⊗R · · · ⊗R Diα with i1 + · · · + iα = n, ij > 0,


and with the boundary map, δ, defined by

α
δ(Di1 ⊗R · · · ⊗R Diα ) = (−1)j+1 Di1 ⊗R · · · ⊗R ∆ (Dij ) ⊗R · · · ⊗R Diα ,
j=1

where ∆ is the “normalized” diagonal map. By “normalized” diagonal map


we mean that we never allow  the degrees to be zero. That is, the normalized
diagonal map will map Di to 0<β<i Dβ ⊗R Di−β . Observe that if i = 1, this
map is the zero map. So, for example, the boundary map applied to the element
a(2) ⊗ b(2) ⊗ c ∈ D2 ⊗R D2 ⊗R D1 is a ⊗ a ⊗ b(2) ⊗ c − a(2) ⊗ b ⊗ b ⊗ c. That
this is a complex is easy to verify; perhaps the easiest way to see it though is to
recognize that this is the dual of the normalized bar complex on the symmetric
algebra. In fact, this complex is the resolution of Λn (see [21]).
Given our set, S = {s1 , . . . , sp }, we define the subcomplex, ( S), as follows.
A term Di1 ⊗R · · · ⊗R Diα is in ( S) if for some β, i1 + · · · + iβ ∈
/ S. Thus, the
complex Xn / ( S) consists of those terms Di1 ⊗R · · ·⊗R Diα with i1 +· · ·+iβ ∈ S
for all β = 1, . . . , α. This means that our terms are of the form
Dsj1 ⊗R Dsj2 −sj1 ⊗R · · · ⊗R Dn−sjα−1 .
So the terms correspond to finding 1 ≤ sj1 < · · · < sjα−1 < n. If α > p, there
are no such terms. If α = p, there is only one term:
Ds1 ⊗R Ds2 −s1 ⊗R · · · ⊗R Dn−sp−1 .
It is this type of complex, Xn / ( S), that is going to provide the resolution of
our skew-shapes.
If our complex is going to be the resolution of some Weyl module, the rows of
the shape of that module determine the set S uniquely. In particular, if we want
such a complex to resolve a skew-hook, then we can describe explicitly what the
set S looks like. As before, we will denote the skew-hook by (k1 , . . . , kt ; l1 , . . . , lt )
and the total number of boxes in this count is simply |k| + |l|. In this notation,
the resolution should start off with
l1 −1 l2 −1
# $% & # $% &
Dk1 ⊗R D1 ⊗R · · · ⊗R D1 ⊗R Dk2 +1 ⊗R D1 ⊗R · · · ⊗R D1 ⊗R · · ·
lt
# $% &
⊗R Dkt +1 ⊗R D1 ⊗R · · · ⊗R D1 .
Hence, the set S that we are looking at is
{k1 , k1 + 1, . . . , k1 + l1 − 1, k1 + k2 + l1 ,
k1 + k2 + l1 + 1, . . . , k1 + k2 + l1 + l2 − 1,
S= k1 + k2 + k3 + l1 + l2 , . . . ,
k1 + · · · + kt−1 + l1 + · · · + lt−1 − 1,
|k| + |l| − lt , . . . , |k| + |l|}.
Resolutions of Weyl modules 227

Next we look at the following situation. Suppose we have the complex Xk , and
we have two sets of integers T ⊂ S, with both of them containing k. Then since
( S) ⊂ ( T ) , we clearly have Xk / ( S) mapping surjectively onto Xk / ( T ) .
Now we do not know what the kernel of this morphism is in general, but suppose
that the integer n is in S (n < k), and we let T = S − {n}. Then it is easy to see
that the kernel, K, of the morphism is spanned by the terms Di1 ⊗R · · · ⊗R Diα
of Xk / ( S) having the property that for some β, i1 + · · · + iβ = n. If we set
S = {s1 , . . . , sp , n + t1 , . . . , n + tq } , where sp = n, and tq = k − n = m, then K
is isomorphic to Xn / ( {s1 , . . . , sp }) ⊗R Xm / ( {t1 , . . . , tq }) . Moreover, there
is a map of Xk / ( T ) into Xn / ( {s1 , . . . , sp }) ⊗R Xm / ( {t1 , . . . , tq }) which
is fairly easy to define. Namely, suppose that Di1 ⊗R · · · ⊗R Diα is a term in
Xk / ( T ) . This means that for all β, we have i1 + · · · + iβ = n. Then there is
a β0 with i1 + · · · + iβ0 −1 = n − c < n, and i1 + · · · + iβ0 = n + d > n. We map
the term Di1 ⊗R · · · ⊗R Diα to the term
   
Di1 ⊗R · · · ⊗R Diβ0 −1 ⊗R Dc ⊗R Dd ⊗R Diβ0 +1 ⊗R · · · ⊗R Diα ,

by diagonalizing Diβ0 , which makes sense since iβ0 = c + d . In fact, the mapping
cone of this map (notice the shift of dimension by 1 here), is precisely Xk / ( S) .

Application to skew-hooks
If we now start with a skew-hook (k1 , . . . , kt ; l1 , . . . , lt ), we get by
Theorem VII.1.4 the exact sequence

0 → (k1 , . . . , kt−1 , kt + 1; l1 , . . . , lt−1 − 1, lt )


→ (k1 , . . . , kt−1 ; l1 , . . . , lt−1 − 1) ⊗R (kt + 1; lt )
→ (k1 , . . . , kt ; l1 , . . . , lt ) → 0.

We want to show, by induction on t and the number of rows, that the resolution
of (k1 , . . . , kt ; l1 , . . . , lt ) is our complex X|k|+|l| / ( S) where S is as above. By
induction, we know that the resolution of (k1 , . . . , kt−1 , kt + 1; l1 , . . . , lt−1 − 1, lt )
is X|k|+|l| / ( T ) , where T = S − {|k| + |l| − lt − 1} . On the other hand, we see
that the resolution of (k1 , . . . , kt−1 ; l1 , . . . , lt−1 − 1) ⊗R (kt + 1; lt ) is the tensor
product of two complexes which is precisely the kernel of the map of X|k|+|l|/( S)
onto X|k|+|l| / ( T ) . Our discussion in the preceding part of this subsection
completes the proof.

VII.6.5 Comparison with the Lascoux resolutions


We end this section, and in fact the whole chapter, with a very detailed compar-
ison of the characteristic-free resolution of the Weyl module, K(2,2,2) , with that
of A. Lascoux (whose maps are not put in evidence). We hope this will serve to
make clear what is behind the Lascoux resolutions, and how one can deal in a
“hands on” way with the characteristic-free case.
228 Some applications of Weyl and Schur modules

The terms of the Lascoux resolution of any skew-shape are read off from the
determinantal expansion of the Jacobi–Trudi matrix of the shape, with the pos-
ition of the terms of the complex determined by the length of the permutations
to which they correspond. In K-theory notation, the Jacobi–Trudi matrix of
the skew-shape λ/µ = (p1 . . . . , pn ; t1 , . . . , tn−1 ) is the n × n matrix

 
Dp1 Dp1 +t1 +1 Dp1 + t2 ··· Dp1 +t1
 Dp2 − t1 Dp2 Dp2 +t2 +1 ··· Dp2 +t2 
 
 Dp3 − t2 Dp3 −t2 −1 Dp3 ··· Dp3 +t3 
(J − T ) :  ,
 .. .. .. .. .. 
 . . . . . 
Dpn − tn−1 Dpn −t2 Dpn −t3 ··· Dpn

where ti = ti + · · · + tn−1 + (n − i), and 


ti = t1 − ti+1 .

VII.6.5.1 Lascoux and non-Lascoux resolutions for (2,2,2)


The Jacobi–Trudi matrix of the partition (2,2,2) is

 
D2 D3 D4
 D1 D2 D3  ,
D0 D1 D2

and the Lascoux resolution of the Weyl module associated to the partition (2, 2, 2)
looks like this:

D3 ⊗R D3 ⊗R D0 D3 ⊗R D1 ⊗R D2
0 → D4 ⊗R D2 ⊗R D0 → ⊕ → ⊕ → D2 ⊗RD2 ⊗RD2 → 0.
D4 ⊗R D1 ⊗R D1 D2 ⊗R D3 ⊗R D1

The correspondence between the terms of the resolution above, and permuta-
tions, is as follows:

D2 ⊗R D2 ⊗R D2 ←→ identity
D3 ⊗R D1 ⊗R D2 ←→ (12)
D2 ⊗R D3 ⊗R D1 ←→ (23)
D3 ⊗R D3 ⊗R D0 ←→ (123)
D4 ⊗R D1 ⊗R D1 ←→ (213)
D4 ⊗R D2 ⊗R D0 ←→ (13).
Resolutions of Weyl modules 229

By contrast, the terms of the characteristic-free resolution of the same Weyl


module, are:

X0 = D2 ⊗R D2 ⊗R D2 ;
(1) (2)
X1 = Z2,1 x ⊗R D3 ⊗R D1 ⊗R D2 ⊕ Z2,1 x ⊗R D4 ⊗R D0 ⊗R D2 ⊕
(1) (2)
Z3,2 y ⊗R D2 ⊗R D3 ⊗R D1 ⊕ Z3,2 y ⊗R D2 ⊗R D4 ⊗R D0 ;
(1) (1)
X2 = Z2,1 xZ2,1 x ⊗R D4 ⊗R D0 ⊗R D2 ⊕
(1) (2) (1) (3)
Z3,2 yZ2,1 x ⊗R D4 ⊗R D1 ⊗R D1 ⊕ Z3,2 yZ2,1 x ⊗R D5 ⊗R D0 ⊗R D1 ⊕
(2) (3) (2) (4)
Z3,2 yZ2,1 x ⊗R D5 ⊗R D1 ⊗R D0 ⊕ Z3,2 yZ2,1 x ⊗R D6 ⊗R D0 ⊗R D0 ⊕
(1) (1)
Z3,2 yZ3,2 y ⊗R D2 ⊗R D4 ⊗R D0 ⊕
(1) (1)
Z3,2 yZ3,1 z ⊗R D3 ⊗R D3 ⊗R D0 ;
(1) (2) (1)
X3 = Z3,2 yZ2,1 xZ2,1 x ⊗R D5 ⊗R D0 ⊗R D1 ⊕
(2) (3) (1)
Z3,2 yZ2,1 xZ2,1 x ⊗R D6 ⊗R D0 ⊗R D0 ⊕
(1) (1) (3)
Z3,2 yZ3,2 yZ2,1 x ⊗R D5 ⊗R D1 ⊗R D0 ⊕
(1) (1) (4)
Z3,2 yZ3,2 yZ2,1 x ⊗R D6 ⊗R D0 ⊗R D0 ⊕
(1) (1) (1)
Z3,2 yZ3,1 zZ2,1 x ⊗R D4 ⊗R D2 ⊗R D0 ⊕
(1) (1) (2)
Z3,2 yZ3,1 zZ2,1 x ⊗R D5 ⊗R D1 ⊗R D0 ⊕
(1) (1) (3)
Z3,2 yZ3,1 zZ2,1 x ⊗R D6 ⊗R D0 ⊗R D0 ;
(1) (1) (3) (1)
X4 = Z3,2 yZ3,2 yZ2,1 xZ2,1 x ⊗R D6 ⊗R D0 ⊗R D0 ⊕
(1) (1) (1) (1)
Z3,2 yZ3,1 zZ2,1 xZ2,1 x ⊗R D5 ⊗R D1 ⊗R D0 ⊕
(1) (1) (2) (1)
Z3,2 yZ3,1 zZ2,1 xZ2,1 x ⊗R D6 ⊗R D0 ⊗R D0 ⊕
(1) (1) (1) (2)
Z3,2 yZ3,1 zZ2,1 xZ2,1 x ⊗R D6 ⊗R D0 ⊗R D0 ;
(1) (1) (1) (1) (1)
X5 = Z3,2 yZ3,1 zZ2,1 xZ2,1 xZ2,1 x ⊗R D6 ⊗R D0 ⊗R D0 ,

where the subscripts on the X indicate the dimension in which these terms
(t)
appear. The symbols Za,b are the formal “polarization” operators defined in
preceding subsections, following Reference [30], and the letters x, y, z are the
separator variables also explained in that paper and above.
The boundary map for this complex is obtained by polarizing all the separator
(t)
variables to one. When the separator x disappears between a Za,b and elements
(t)
in the tensor product of divided powers, this means ∂a,b , or the place polarization
operator, applied to that tensor product. The only essentially new terms that
(1) (1)
we have here are the terms that involve Z3,2 yZ3,1 z, or more generally terms of
the form:

(1) (1) (k ) (k )
Z3,2 y Z3,1 z Z2,11 x · · · x Z2,1n−2 x ⊗R D2+1+|k| ⊗R D2+1−|k| ⊗R D0 ,

(of which, in this example, there are not that many). For these we have to use
identities of Capelli type, or some easy variants of them (Reference [30] and
230 Some applications of Weyl and Schur modules

Section VI.4). The boundary map on such a term sends it to


 
(1) (1) (k ) (k )
Z3,2 y Z3,1 z ∂ Z2,11 x · · · x Z2,1n−2 x ⊗R Dp+1+|k| ⊗R Dq+1−|k| ⊗R D0
 
(1) (1) (k ) (k )
± Z3,2 y Z3,1 Z2,11 x · · · x Z2,1n−2 x ⊗R Dp+1+|k| ⊗R Dq+1−|k| ⊗R D0
 
(1) (1) (k ) (k )
∓ Z3,2 Z3,1 z Z2,11 x · · · x Z2,1n−2 x ⊗R Dp+1+|k| ⊗R Dq+1−|k| ⊗R D0 ,

and we have to define the terms


(1) (k ) (k )
Z3,1 Z2,11 x · · · x Z2,1n−2 x ⊗R Dp+1+|k| ⊗R Dq+1−|k| ⊗R D0
and
 
(1) (1) (k ) (k )
Z3,2 Z3,1 z Z2,11 x · · · x Z2,1n−2 x ⊗R Dp+1+|k| ⊗R Dq+1−|k| ⊗R D0

with n ≥ 2.
If n = 2, we have
(1)
Z3,1 v = ∂3,1 (v),
while
(1) (1) (1) (2) (2)
Z3,2 Z3,1 z ⊗ v = −Z2,1 x ⊗ ∂3,2 (v)+Z3,2 y ⊗ ∂2,1 (v).
For n > 2, we have
(1) (k ) (k ) (k +1) (1) (k ) (k )
Z3,1 Z2,11 x · · · xZ2,1n−2 x ⊗ v = −Z2,11 xZ3,2 Z2,12 x · · · xZ2,1n−2 x ⊗ v
 
k1 + k2 (1) (k +k +1) (k )
+ Z3,2 yZ2,11 2 x · · · xZ2,1n−2 x ⊗ v,
k2 − 1
and
(1) (1) (k ) (k ) (1) (2) (k ) (k )
Z3,2 Z3,1 zZ2,11 x · · · xZ2,1n−2 x ⊗ v = −Z2,1 xZ3,2 Z2,11 x · · · xZ2,1n−2 x ⊗ v
(2) (k +1) (k )
+ (k1 − 1)Z3,2 yZ2,11 x · · · xZ2,1n−2 x ⊗ v.
With these identities in play, we can easily write down the boundary map for
this resolution.
VII.6.5.2 Reduction of non-Lascoux to Lascoux
It is of some interest to see how we can discard the “excess” terms of the
characteristic-free resolution to recover that of A. Lascoux. For example, we want
(2) (2)
to throw away the terms Z2,1 x⊗R D4 ⊗R D0 ⊗R D2 and Z3,2 y⊗R D2 ⊗R D4 ⊗R D0 .
(2)
If we look at the image of a term Z2,1 x ⊗ v, where v ∈ D4 ⊗ D0 ⊗ D2 , we
(2) (2)
see that it is ∂21 (v) = 1 2
2 ∂21 (v). So, the image of Z2,1 x ⊗ v is the same as
1 (1) (1)
the image of ⊗ 2 Z2,1 x ∂21 (v). Hence
in characteristic zero we can rig up the
boundary map taking this into account. Obviously the same kind of thing holds
Resolutions of Weyl modules 231

(2)
for the term Z3,2 y ⊗R D2 ⊗R D4 ⊗R D0 that is, its image is the same as that of
1 (1) (1)
2 Z3,2 x⊗ ∂32 (v). Thus, any term that is sent by the “full” boundary map into
one of the “redundant” terms above, should now be sent into the non-redundant
term instead. For example, the following terms should now be sent as indicated:
(1) (2) (1) (2) (1) (1)
Z3,2 yZ2,1 x ⊗ v → Z3,2 y ⊗ ∂2,1 (v) − Z2,1 x ⊗ ∂3,1 (v) − 12 Z2,1 x ⊗ ∂2,1 ∂3,2 (v)

and
(1) (1) (1) (1) (2) (1)
Z3,2 yZ3,1 z ⊗ v → Z3,2 y ⊗ ∂3,1 (v) − Z2,1 x ⊗ ∂3,2 (v) − 12 Z3,2 y ⊗ ∂3,2 ∂2,1 (v).
With this modification, the terms of the type
(1) (1)
Z2,1 xZ2,1 x ⊗R D4 ⊗R D0 ⊗R D2
and
(1) (1)
Z3,2 yZ3,2 y ⊗R D2 ⊗R D4 ⊗R D0
are automatically sent to zero. As a result, when we go to the last term that
counts in this complex, we see that under the boundary map in the big complex
we get
(1) (1) (1) (1) (1) (1) (2)
Z3,2 yZ3,1 z Z2,1 x ⊗ v → Z3,2 yZ3,1 z ⊗ ∂2,1 (v) + Z3,2 y Z2,1 x ⊗ ∂3,2 (v)
(1) (1) (2) (1) (1) (2)
− Z3,2 yZ3,2 y ⊗ ∂2,1 (v) − Z2,1 xZ2,1 x ⊗ ∂3,2 (v),
so that under the modified boundary, we can simply define the boundary map
on this term to be
(1) (1) (1) (1) (1) (1) (2)
Z3,2 yZ3,1 z Z2,1 x ⊗ v → Z3,2 yZ3,1 z ⊗ ∂2,1 (v) + Z3,2 y Z2,1 x ⊗ ∂3,2 (v).
Of course it remains to prove that in characteristic zero this is exact; we will
now indicate how to do this.
VII.6.5.3 Question of exactness
Here, we indicate more schematically how we modify the maps to take advantage
of divisibility in Q. Divide the terms of our big complex into the sum of those
of the Lascoux complex, and the others. That is, we look at the terms of the
complex as:
X0 = A0
X1 = A1 ⊕ B1
X2 = A2 ⊕ B2
X3 = A3 ⊕ B3
Xj = Bj for j = 4, 5,
where the Ai are the sums of the “Lascoux” terms, and the Bl are the sums of
the others.
232 Some applications of Weyl and Schur modules

Let σ1 be the map


B1 → A1
defined by
(2) 1 (1) (2) 1 (1)
Z2,1 x ⊗ v −→ Z x ⊗ ∂2,1 (v) and Z3,2 y ⊗ w −→ Z y ⊗ ∂3,2 (w)
2 2,1 2 3,2
where v ∈ D4 ⊗R D0 ⊗R D2 and w ∈ D2 ⊗R D4 ⊗R D0 . This situation we have
already discussed, but we should point out that the map σ1 satisfies the identity:
(C1 ) δA1 A0 σ1 = δB1 B0 ,
where by δA1 A0 we mean the component of the boundary of the large complex
that carries A1 to A0 , and similarly for δB1 B0 . Naturally, we will use notation
δAi+1 Ai , δAi+1 Bi etc. in the same way. We define
∂1 : A1 → A0
as
∂1 = δA1 A0 .
But now we are in position to define
∂2 : A2 → A1
by
∂2 = δA2 A1 + σ1 δA2 B1 .
We get immediately
Remark VII.6.17 The composition ∂1 ∂2 = 0.
Proof We have
∂1 ∂2 (a) = δA1 A0 {δA2 A1 (a) + σ1 δA2 B1 (a)}
= δA1 A0 δA2 A1 (a) + δA1 A0 σ1 δA2 B1 (a).
But since δA1 A0 σ1 = δB1 B0 , we see that
δA1 A0 σ1 δA2 B1 (a) = δB1 B0 δA2 B1 (a),
so
δA1 A0 δA2 A1 (a) + δA1 A0 σ1 δA2 B1 (a) = δA1 A0 δA2 A1 (a) + δB1 B0 δA2 B1 (a).
But this is zero since it is the boundary of the fat complex applied to a. 2
Before we can prove exactness at A1 , we have to define a map
σ2 : B2 → A2
Resolutions of Weyl modules 233

such that
(C2 ) δB2 A1 + σ1 δB2 B1 = {δA2 A1 + σ1 δA2 B1 }σ2 .
We define this map as follows:
(1) (1)
Z2,1 x Z2,1 x ⊗ v −→ 0
(1) (1)
Z3,2 y Z3,2 y ⊗ v −→ 0
(1) (3) (1) (2)
Z3,2 y Z2,1 x ⊗ v −→ 13 {Z3,2 y Z2,1 x ⊗ ∂2,1 (v)}
(2) (3) (1) (2) (1) (1) (2)
Z3,2 y Z2,1 x ⊗ v −→ 13 {Z3,2 yZ2,1 x ⊗ ∂3,1 (v)−Z3,2 yZ3,1 z ⊗ ∂2,1 (v)}
(2) (4) (1) (1) (3)
Z3,2 y Z2,1 x ⊗ v −→ − 12 {Z3,2 y Z3,1 z ⊗ ∂2,1 (v)}
where the element v is in the appropriate tensor product of divided powers as
given in the description of the large complex.
Remark VII.6.18 The map σ2 defined above satisfies the condition (C2 ).
Proof Trivial. 2
Remark VII.6.19 We have exactness at A1 .
Proof This is just a diagram chase. 2
Using σ2 we can now also define
∂3 : A3 → A2
by
∂3 = δA3 A2 + σ2 δA3 B2 .
Not too surprisingly, we have
Remark VII.6.20 ∂2 ∂3 = 0.
Of course, what we need now is a map σ3 : B3 → A3 similar to the maps σ
above, that is, satisfying
(C3 ) δB3 A2 + σ2 δB3 B2 = {δA3 A2 + σ2 δA3 B2 }σ3 .
We define such a σ3 as follows:

(1) (2) (1)


Z3,2 y Z2,1 x Z2,1 x ⊗ v −→ 0
(2) (3) (1) (1) (1) (1) (2)
Z3,2 y Z2,1 x Z2,1 x ⊗ v −→ 3 {Z3,2 y Z3,1 z Z2,1 x ⊗ ∂2,1 (v)}
1
(1) (1) (3) (1) (1) (1) (1)
Z3,2 y Z3,2 y Z2,1 x ⊗ v −→ − 13 {Z3,2 y Z3,1 z Z2,1 x ⊗ ∂2,1 (v)}
(1) (1) (4) (1) (1) (1) (2)
Z3,2 y Z3,2 y Z2,1 x ⊗ v −→ − 13 {Z3,2 y Z3,1 z Z2,1 x ⊗ ∂2,1 (v)}
(1) (1) (1) (1) (1) (1) (1)
Z3,2 y Z3,1 z Z2,1 x ⊗ v −→ 3 {Z3,2 y Z3,1 z Z2,1 x ⊗ ∂2,1 (v)}
1
(1) (1) (3)
Z3,2 y Z3,1 z Z2,1 x ⊗ v −→ 0
234 Some applications of Weyl and Schur modules

where the element v again is taken in the appropriate tensor product of divided
powers as described earlier.

Remark VII.6.21 The map σ3 defined above satisfies the condition (C3 ).

Proof Trivial. 2

We have yet to prove that we have exactness at A2 and that ∂3 is a


monomorphism. But first we explicitly define the boundary maps in the complex
∂ ∂ ∂
0 → A3 →3 A2 →2 A1 →1 A0 .

The map ∂1 is clear: it is just the operation of the indicated polarization


operators on the argument. The map ∂2 is defined as:
 
(1) (2) (1) (2) (1) (1) (1) (1) (1)
∂2 Z3,2 yZ2,1 x⊗v = Z3,2 y⊗∂2,1 (v)−Z2,1 x⊗∂3,1 (v)− 12 Z2,1 x⊗∂2,1 ∂3,2 (v);
 
(1) (1) (1) (1) (1) (1) (1) (1) (2)
∂2 Z3,2 yZ3,1 z⊗v = Z3,2 y⊗∂3,1 (v)− 12 Z3,2 y⊗∂3,2 ∂2,1 (v)+Z2,1 x⊗∂3,2 (v).

And finally, the map ∂3 is defined as:


 
(1) (1) (1) (1) (1) (1) (1) (2) (1)
∂3 Z3,2 yZ3,1 z Z2,1 x⊗v = Z3,2 yZ3,1 z⊗∂2,1 (v) + Z3,2 yZ2,1 x⊗∂3,2 (v).

Since one component of the map ∂3 is a diagonalization of D2 into D1 ⊗R D1 ,


it is clear that ∂3 is injective. Hence, the only remaining exactness to prove is at
A2 . To handle this, we suppose that we have V ∈ A1 with ∂2 (V ) = 0. We want
to produce an element W ∈ A3 such that ∂3 (W ) = V.
It is easy to see that we can add to V an element W ∈ B2 , such that under
the boundary of the characteristic-free complex, V + W goes to zero, while
σ2 (W ) = 0. Using the fact that this large complex is acyclic, we know that there
are elements α ∈ A3 , β ∈ B3 such that α + β goes to V + W under the boundary
map of that complex. It is then easy to show that α + σ3 (β) goes to V , and our
Lascoux “reduction” is exact.

VII.6.5.4 The Lascoux skeleton in general


We end this subsection with a description of an inductive method for finding the
“Lascoux skeleton” inside the characteristic-free resolution of any skew-shape.
That is, if we are given a term in the expansion of the Jacobi–Trudi determinant
(the determinant of the matrix denoted by (J−T) at the beginning of this sub-
section) corresponding to the permutation, σ ∈ Sn , we will show where, among
the terms of the characteristic-free resolution, this term will be found.
Recall that Lascoux [64] has described the terms and their placement in the
resolutions of Weyl modules (of skew-partitions) in terms of the lengths of the
permutations corresponding to the determinantal expansion of the Jacobi–Trudi
matrix for that Weyl module. If σ ∈ Sn is such that σ(n) = i, then σ can be
Resolutions of Weyl modules 235

written uniquely as a product:


σ = (n, n − 1)(n, n − 2) · · · (n, i)σ  ,
where (n, j) stands for the transposition on n and j, and σ  ∈ Sn−1 . (Notice
that the length of σ is n − i + length(σ  ).) One then goes to the part of the
characteristic-free resolution of λ/µ that involves the terms
t +1 (l ) t +1 (l ) (l )
n−1
(Zn,n−1  Z n,n−1
n−1 n−2
)⊗ (Zn,n−2  Z n,n−2
n−2
) ⊗ · · · ⊗ (Zn,i
ti +1
 Z n,ii ),
and then finds inside the corresponding Res the term corresponding to σ  . This
is the way of recovering the terms of the Lascoux resolution within the resolution
described at the end of Subsection VII.6.3.
This page intentionally left blank
APPENDIX A
APPENDIX FOR LETTER-PLACE METHODS

In this appendix, we furnish the proofs of the theorems stated in Sections VI.3
and VI.8 on letter-place methods. More precisely, Sections A.1 and A.2 give a
complete proof of Theorem VI.3.2 which involves straightening considerations
(Section A.1) as well as a combinatorial procedure known as the Robinson–
Schensted–Knuth correspondence (Section A.2). In Section A.3 we deal in a less
detailed way with the proofs of Theorems VI.3.3 and VI.3.4, mainly showing what
modifications are necessary to adapt the methods of proof of Theorem VI.3.1 to
the other situations. In Section A.4, we discuss the modifications necessary for
the proof of Theorem VI.8.4.

A.1 Theorem VI.3.2, Part 1: the double standard tableaux generate


To show the generation by double standard tableaux, we first of all can assume
that they are row standard all the time, since we can always row-standardize
them without changing their values. We can also assume that the lengths of the
rows are decreasing, since we can always arrange that without changing their
values. So we are talking about double tableaux of the form:
 
w1 1(a11 ) 2(a21 ) 3(a31 ) · · ·
 w2 1(a12 ) 2(a22 ) 3(a32 ) · · · 
(S)   ···

··· ··· ··· ··· 
wn 1(a1n ) 2(a2n ) 3(a3n ) · · ·
where αi = (a1i + a2i + a3i + · · · ) ≥ αj for 1 ≤ i < j ≤ n, and the words, when
written out, are increasing. Our tableaux of the form (W ):
 
w1 1(k1 )
 w2 2(k2 ) 
(W )   ···

··· 
wn n(kn )
are, by rearranging the order of the rows if necessary, of type (S) above. Since
they form a basis for Dk1 (F )⊗· · ·⊗Dkn (F ), it certainly suffices to prove that any
double tableau of type (S) is a linear combination of standard double tableaux.
Our procedure is to put a quasi order on these double tableaux. The object is
to show that we can express a double tableau that is not standard in the places
(or letters) as a linear combination of others which are “lower” in the quasi order.
238 Appendix for letter-place methods

Since the set of tableaux is finite, this will show that we can express any double
tableau as a linear combination of such that are standard in the places. We then
do the same for the letters. As this process cannot go on forever, each double
tableau must eventually be a linear combination of standard double tableaux;
that is, the standard ones generate.
At the cost of being ultra pedantic, we shall denote a double tableau, T , by a
triple: T = (λ; L, P ), to denote the diagram, the “letter,” and the “place” parts
of the tableau. We stress that the diagram is always to be that of a partition
in this case, that is, the lengths of the rows are decreasing from top to bottom.
In Subsection VI.1.1, we described a well-known partial order on partitions: we
said that λ > λ if the first row of λ (from the top) which differs in length
from the corresponding row of λ, is longer than that of λ. We also defined in
Subsection VI.1.2, a quasi order on single tableaux. Using these orderings, we
now define a quasi ordering of our double tableaux.
Definition A.1.1 Given tableaux, T = (λ; L, P ) and T  = (λ ; L , P  ), we
define (λ ; L , P  ) ≤ (λ; L, P ) if λ ≥ λ, Lij ≥ Lij for all i, j, and Pij ≥ Pij
for all i, j. We then say that (λ ; L , P  ) < (λ; L, P ) if (λ ; L , P  ) ≤ (λ; L, P ),
and either λ > λ or λ = λ and either Lij > Lij for some i, j, or Pij > Pij for
some i, j.
To give an example, and also to see how the general “straightening” proceeds,
consider an elementary, yet prototypical situation:
 
w1 1(a1 ) 2(b1 )
T = ,
w2 1(2) 2(b2 )
where w1 and w2 are words of degrees a1 + b1 and 2 + b2 respectively. (We are
now dropping the cumbersome designation of a tableau as a triple.)
Our problem in the “place” section of the tableau is that we have 1 in the
bottom row, so that no matter what the value of a1 , that section is not standard.
Recall that

T = w1 (a1 )w2 (2) ⊗ w1 (b1 )w2 (b2 )
where the letters in parentheses indicate the degrees of the wi under appropriate
diagonalization. (We will continue to use this notation to indicate that we have
diagonalized our terms in the indicated degrees.)
Consider now:
  (a1 +2) (b1 )

  w 1 w2 (2) 1 2
T = ,
w2 (b2 ) 2(b2 )
 
 w1 1(a1 +1) 2(b1 −1)
T = ,
w2 1 2(b2 +1)
 
w1 1(a1 +2) 2(b1 −2)
T  = .
w2 2(b2 +2)
Theorem VI.3.2, Part 1 239

Then it is straightforward to check that


 
 b2 + 2  
T = T − (b2 + 1)T − T .
2

• It is extremely important to observe at this point that the diagram for T  has
changed shape, but it is bigger than that of our original T . The shapes of T  and
T  are the same as our original, the letter part of the tableau has not changed
(that is, L = L = L), but we have both P  and P  which are different from
P . In both cases, though, we have decreased these tableaux (in the quasi order).
Notice that if we were to “straighten” our letter tableau similarly, the changes
would either lead to a bigger diagram, in which case we go down in the order,
or the places would remain unchanged, and the letter tableaux that intervened
would be lower in the quasi order.
• When checking this type of calculation, it is most often convenient to
assume that the words, wi , are simply divided powers. That is, we might assume
that w1 = x(a1 +b1 ) and that w2 = y (2+b2 ) . This makes keeping track of the
diagonalizations much easier.
The point of this exercise is to show that when we get rid of an instance of
non-standardness, the tableaux that emerge in the process are all less, in the
quasi-order we introduced, than the tableau we started with. In the case of T  ,
this is due to the fact that the diagram of T  is properly larger than that of T ;
in the other two cases, the “letter” side of the tableau has not changed, but the
“place” side has properly decreased. Of course, the tableau, T  , is manifestly
non-standard (the others may or may not be, depending on the relative values
of a1 , b1 and b2 ).
To see how a greater number of places affect the straightening procedure, we
will look at one more example. To this end, consider:
 
w1 1(a1 ) 2(b1 ) 3(c1 )
T = .
w2 1 2(b2 ) 3(c2 )
Then if we set:
 
 w1 w2 (1) 1(a1 +1) 2(b1 ) 3(c1 )
T = .
w2 (b2 + c2 ) 2(b2 ) 3(c2 )
 
 w1 1(a1 +1) 2(b1 −1) 3(c1 )
T = ,
w2 2(b2 +1) 3(c2 )
and
 
 w1 1(a1 +1) 2(b1 ) 3(c1 −1)
T = ,
w2 2(b2 ) 3(c2 +1)
we have
T = T  − (b2 + 1)T  − (c2 + 1)T  .
240 Appendix for letter-place methods

Again we see that we have eliminated the offending 1 in the bottom row, and
we have “replaced” T by the tableaux T  , T  and T  which are strictly lower in
our quasi-order than T .
As these two examples show, it is enough to work either on the letter or place
side of the tableau to push us further toward standardness. These examples are
prototypical in the sense that it is enough to work on straightening two adjacent
rows, for if we have a tableau that is not standard, this is because it is not
strictly increasing in the columns (since we start with row-standardness). But
then a violation of strict increase must also occur in two adjacent rows. This is
the reason that we can focus on two-rowed double tableaux such as:
 
w1 z11 z12 z13 · · · z1k
T =
w2 z21 z22 z23 · · · z2l
where we write zij to represent the integers from 1 to n, with possible repeats.
Suppose that the first violation of strict increase occurs in the ith column, that
is, z11 < z21 , . . . , z1i−1 < z2i−1 , but z1i ≥ z2i . (The ith column could be the first,
as it was in the above examples.) Next, suppose that z2i = · · · = z2t < z2t+1
(possibly t = i), and consider the sum of tableaux:
  w1 w2 (t) z11 · · · z1i−1 z21 · · · z2t z1i · · · z1k


T = .
w2 (l − t) z2t+1 · · · z2l
Now, for each subset J of U = {z21 , . . . , z2t , z1i , . . . , z1k }, of order 0 < s < t,
let J  be the complement of J in U , let ZJ  be the row tableau of that set, and ZJ
the corresponding row tableau of J. We let I stand for an arbitrary subset of U of
order precisely t, other than the subset {z21 , . . . , z2t } itself, I  its complementary
subset, and ZI  , ZI as for J. Set
  w1 w2 (t − s) z11 · · · z1i−1 ZJ 


TJ = ,
w2 (l − t + s) ZJ z2t+1 · · · z2l
and cJ equal plus or minus the product of binomial coefficients that would be
appropriate due to the multiplication in the divided power algebra, of ZJ with
z2t+1 · · · z2l (see the foregoing examples to get a more concrete picture of this
description). Finally, let
 w1 z11 · · · z1i−1 ZI 


TI = ,
w2 ZI z2t+1 · · · z2l
and cI equal plus or minus the corresponding product of binomial coefficients.
Then it is tedious, but straightforward to prove that
 
T = ±T  + cJ TJ + cI TI .
J I

The important thing to observe in this equation is that the diagrams of T 


and TJ are all bigger than that of T and so all these terms are strictly lower in
the quasi order. The terms TI all have the letter half of the tableau unchanged
Theorem VI.3.2 Part 2 241

from that of T . However, since our sets I must have t elements, and none can be
the set {z21 , . . . , z2t } itself, at least one of these elements must remain in the top
row, and one of the elements z1u > z2i must come down into the bottom row.
Thus, the resulting tableaux in this case are all less in the quasi order than the
original.
One may ask, when dealing with the tableaux other than the TI what we do
about keeping these tableaux within the class we are considering, namely, the
lengths of the rows decreasing. In order to keep to this recipe, we simply push
the top row of the two up as far as it has to go, and the bottom one down to
where it has to go to make it into a legitimate shape. But this still makes the
shape lower in our quasi order than the original.

A.2 Theorem VI.3.2 Part 2: linear independence of double standard


tableaux
In this section, we introduce the Robinson–Schensted–Knuth correspondence
(which will be written R–S–K in the future) to set up a one-to-one count between
the double standard tableaux, and the usual basis elements of Dk1 (F ) ⊗ · · · ⊗
Dkn (F ).
% ·&#
Let us agree, as we did in Chapter III, to write x(k) as x · · x$, when there is no
k
% ·&#
danger of confusion. That is, we write x · · x$ to mean the row tableau consisting
k
of k copies of x in a row. Then a basis element of Dk (F ) is the same as a non-
decreasing sequence of elements: xi1 ≤ xi2 ≤ · · · ≤ xik , and a basis element
of Dk1 (F ) ⊗ · · · ⊗ Dkn (F ) is a long string of such sequences. For instance, the
(2) (3)
basis element x1 x3 ⊗ x2 x3 ∈ D3 ⊗ D4 corresponds to the long sequence β =
x1 x1 x3 x2 x2 x2 x3 . Now R–S–K sets up a correspondence between such sequences
and pairs of tableaux. Before we describe (loosely) this correspondence in general,
let us look at the example at hand.
To β we want to associate two tableaux, L(β) and P (β); first we will describe
how we get L(β). Since x1 x1 x3 is increasing, we put these elements in a row
tableau: x1 x1 x3 . But now we hit up against x2 (the next term in our
sequence), and if we were to put that in as the next element of the row, we
would spoil row-standardness. So, we “bump” the x3 from the first row and
replace it by x2 while moving x3 to the second row of a now two-rowed tableau:
x1 x1 x2
. And now we see that the remaining three terms of the sequence,
x3
β, can all be placed in the first row without necessitating any bumping, so the
x1 x1 x2 x2 x2 x3
tableau we end up with is L(β) = . (Notice that
x3
this tableau is standard, an end result guaranteed by the nature of the bumping
process.)
242 Appendix for letter-place methods

To assign the next tableau, P (β), we will make a slight modification of the
usual R–S–K, and make use of the fact that we are looking very distinctly at
D3 ⊗ D4 , namely, a two-fold tensor product of divided powers of designated
degrees. This means that we have only two places to deal with, namely 1 and
2, and we want P (β) to be of the same shape as L(β), standard, and filled with
the places 1 and 2. Now the first part of the construction of our tableau involved
using the first three terms of the sequence, β, so that the first three boxes of the
first row should be filled in with the place, 1. The second row was next produced
by bumping, so we put the place 2 in that box of the second row. The next
three entries in the first row were produced by inserting entries from the second
factor, so we fill them in with 2, and the resulting tableau we get is P (β) =
1 1 1 2 2 2
.
2
On the other hand, given the pair of tableaux, L(β) and P (β), we can recon-
struct the sequence β. We look at the tableau, P (β), and peel off the highest
place from the highest row first, with its corresponding letters in L(β). This
tells us that in our second factor, we had x2 x2 x3 , and we still are left with the
x x1 x2 1 1 1
pair of tableaux: 1 and . This means that we still have a
x3 2
term from the second factor, and it was obtained by bumping from a single-rowed
tableau with three entries. The only way that bumping could have happened was
for x3 to have been in the upper row previously, and to have been bumped by x2
(for if an element of the first row had been bumped by x1 , that element would
have been x2 , and we would have had x2 in the second row). Therefore, the
full second factor must be x2 x2 x2 x3 , and our first factor is what is left, namely
x1 x1 x3 .
To explain the bumping procedure that leads to the construction of the
first tableau, let us consider the general situation of an ordered set, S =
{s1 , s2 , . . . , st }, and a sequence β = si1 si2 . . . sik formed from these elements.
To the element si1 we attach the tableau consisting of one box, and the one
entry, si1 . If si2 ≥ si1 , then we attach the one-rowed tableau consisting of si1
and si2 ; if si2 < si1 , we attach the two-rowed tableau having si2 in the top row,
and si1 in the second row. Suppose we have attached by this procedure a stand-
ard tableau, λ, to the first l elements of the sequence. We then take the element
sil+1 and try to add it to λ. If it exceeds (in the weak sense) every element in
the top row of λ, then we stick it onto the end of that row. If it does not, then
we look for the first element of the first row that is strictly greater than it, and
replace it by sil+1 . This leaves us with the exiled element of the first row, and
we take it and look at the second row. If it fits, we add it on to the second row,
otherwise bump as before and continue in this way. By this procedure, we arrive
at the tableau, L(β), associated to the sequence, β.
What if, one may ask, the first two rows of λ were of the same length and
the bumped element exceeded all the elements of the second row? Would not
sticking it onto the second row produce an inadmissible shape? But we have
Theorem VI.3.2 Part 2 243

assumed that the tableau, λ, is standard, and so we see that this situation cannot
arise.
To associate a “place” tableau to the sequence β, we have to be given a bit
more information along with the sequence (as when we saw in our example that
we were dealing with two factors—hence two places—and subsequences of fixed
length). So let us assume that our integer k = k1 + · · · + kn , and that our
sequence β has the property that the first k1 elements are increasing, the next
k2 are increasing, and so on. We want to assign to our β a standard place tableau
having n places. Notice that since the first k1 elements are increasing, they all
fit into a single-row tableau with k1 boxes. We record this by attaching a single-
rowed tableau with k1 boxes all filled in with the place 1. We then run through
the next k2 elements, keeping track of all the new boxes created in the L(β)
construction by labeling them with the place 2. Notice that if elements from the
second strand bump elements from the first row, the elements they bump grow
in size, so that these elements all go into the second row (they bump no more).
Also, once an element from this second strand gets placed in the first row, all
the others do. This means that the place tableau so far associated has 1’s in the
first row, and 2’s in the first and second rows. Obviously the number of 2’s in
the second row cannot exceed the number of 1’s in the first row (namely, k1 ).
We then proceed with the third strand placing 3’s, and so on.
Rather than spend more words on this description, a not too trivial example
may be in order. For simplicity, we will use numbers for letters as well as places;
we believe that this should cause no confusion.

Example A.2.1 Consider the element 223446 ⊗ 12335 ⊗ 3446 in D6 ⊗ D5 ⊗ D4 .


Following the recipes above, one ends with the following pair of tableaux (the
first one the “letter” tableau, the second the “place” tableau filled with only
three places):
1 2 2 3 3 3 4 4 6
2 3 4 4 5
6
and
1 1 1 1 1 1 3 3 3
2 2 2 2 2 .
3
The original “sequence” (we write it in quotes, since we have taken the liberty
to use the tensor product symbol to mark off where the substrands are to be
seen) can be reconstructed from this pair of tableaux as we indicated in the first
example. By removing all the entries labeled 3 in the first row, we see that the
third factor has to end with 446. But then we see that the 6 must have been
bumped into the third row by the first term of the third factor. If 6 had been
bumped from the second row, it would have had to be the 5 that did it, but if
244 Appendix for letter-place methods

the 5 had been the bumper, it would have already fit nicely onto the first row.
So, it must have been the 3 that bumped, which means that the third factor was
3446, and that 3 bumped into the two-rowed tableau having 122335 in the top
row and 23446 in the second. The corresponding place tableau now has only the
places 1 and 2 with all the entries labeled 1 in the top row and all those labeled
2 in the second. Now the previous tableau had six entries in the top row, and
only four in the second. The only way we could have arrived at the current stage
is if we had bumped a 5 into the tableau having 122346 in the top row, and 2344
in the second. And so forth. In this way we reconstruct our original sequence of
departure.

This procedure establishes a one-to-one correspondence between the usual


basis elements of Dk1 (F ) ⊗ · · · ⊗ Dkn (F ) and the double standard tableaux hav-
ing ki places i for i = 1, . . . , n. As a result we have a complete proof of the linear
independence of the double standard tableaux, and hence of Theorem VI.3.2.

A.3 Modifications required for Theorems VI.3.3 and VI.3.4


In Subsections VI.3.2 and VI.3.3, we discussed the various changes we have to
consider when we alter the signs of the letters or places. In this section, we will
sketch the proofs that underlie the theorems stated there.
We will look first at the case of positive letters and negative places.
In the same way that we proved that the double standard tableaux generated
the tensor product of divided powers, we prove that the appropriate double
standard tableaux generate the tensor product of exterior powers. That is, we
introduce a quasi-order on double tableaux, and then show that any non-standard
double tableau may be expressed as a linear combination of tableaux properly
lower in that order. It is worth pointing out that if there is a violation of weak
increase in the columns on the right hand side of a double tableau, the calculation
that we have to do to straighten this is a bit easier than in the case of positive
places, as there are no repeats in the rows (we may, as before, assume that we
are starting with a row-standard double tableau). We will omit the proof of this
fact as it is almost identical to the previous proof. The interested reader might
consult the paper by B. Taylor, [82], for a complete proof.
The proof of the linear independence of the standard double tableaux is slightly
different from that in the case of divided powers, so we will indicate the slight
variant of the R–S–K that must be used here. The essential difference is that
when we construct a letter tableau from a sequence of letters, we construct it
column-wise rather than row-wise, and if bumping is necessary, we start a new
column with the bumped element. We now insist, though, that the columns be
strictly increasing, so that we must bump if our new element does not strictly
exceed the largest element already in the column. The corresponding tableau of
places (this time negative) is built as we did earlier. Again, an example should
suffice to make this clear.
Theorems VI.3.3 and VI.3.4 245

Let us take the basis element x2 ∧ x3 ∧ x5 ⊗ x1 ∧ x3 ⊗ x2 ∧ x4 ⊗ x3 ∧ x5 ∧ x6


in Λ3 F ⊗ Λ2 F ⊗ Λ2 F ⊗ Λ3 F.
Because the first three terms are strictly increasing, after three steps we obtain
x2 1
the tableau: x3 , with the accompanying tableau of places: 1 .
x5 1
Now we go to our term in the second place (so that any boxes constructed
from these two terms get labeled 2 in the tableau of places we construct), and
bump in the term x1 . This has the effect of knocking out the x2 from the first
column, and replacing it by x1 , while we place x2 in the second column getting:
x1 x2 1 2
x3 , with accompanying tableau of places: 1 .
x5 1
Proceeding step by step, we end up with the pair of tableaux:
x1 x2 x2 1 2 3
x2 x3 x5 1 2 4
x3 x4 and 1 3 .
x5 4
x6 4
Putting these together, we get the standard double tableau:
 
x1 x2 x2 1 2 3
x2 x3 x5 1 2 4
 
x3 x4 1 3 .
 
x5 4 
x6 4

These arguments are sufficient to prove Theorem VI.3.3.

As for the case of negative letters and positive places, the proof that the
appropriate double standard tableaux generate is just a mild modification of the
above. The proof of linear independence involves a modified R–S–K, namely, in
this case, you build the letter tableau by rows, and again label the new boxes by
the place corresponding to the element that you bump in. To illustrate, we will
use the same element we used above, namely x2 ∧ x3 ∧ x5 ⊗ x1 ∧ x3 ⊗ x2 ∧ x4 ⊗
x3 ∧ x5 ∧ x6 in Λ3 F ⊗ Λ2 F ⊗ Λ2 F ⊗ Λ3 F. Our R–S–K applied to this yields the
double tableau:
 
x1 x2 x3 x5 x6 1 1 1 4 4
x2 x3 x4 2 2 3 .
x3 x5 3 4

The reader should be careful not to confuse this double standard tableau
with the double tableau that actually equals the element x2 ∧ x3 ∧ x5 ⊗ x1 ∧
x3 ⊗ x2 ∧ x4 ⊗ x3 ∧ x5 ∧ x6 in Λ3 F ⊗ Λ2 F ⊗ Λ2 F ⊗ Λ3 F. For this element
246 Appendix for letter-place methods

is simply
 
x2 x3 x5 1 1 1
 x1 x3 2 2 
 ,
 x2 x4 3 3 
x3 x5 x6 4 4 4

which is, of course, far from standard.

For the case of the symmetric algebra, the proof that the appropriate double
standard tableaux generate presents no new feature. As for their linear inde-
pendence, the R–S–K builds the letter part of the tableau by columns, and the
place part of the tableau in the usual way. To take an example, consider the
element x22 x3 x4 ⊗ x31 x2 ⊗ x2 x24 ∈ S4 F ⊗ S4 F ⊗ S3 F . The corresponding standard
double tableau is
 
x1 x2 1 2
 x1 x2 1 2 
 
 x1 x3 1 2 
 
 x2 x4 1 2  .
 
 x2 3 
 
 x4 3 
x4 3

This completes the proof of our theorems.

A.4 Modifications required for Theorem VI.8.4


We just have to show, in this section, how to modify the R–S–K for D(ϕ ⊗ Rn )
and Λ(ϕ ⊗ Rm ). As usual we discuss the case of D(ϕ ⊗ Rn ) in some detail and
point out that the same discussion obtains for Λ(ϕ ⊗ Rm ) where the basis of Rm
is considered to be negative.
If we fix our degrees, k1 , . . . , kn , we know that a basis for Dk1 (ϕ)⊗· · ·⊗Dkn (ϕ)
is of the form

(x11 · · · x1s1 ⊗ y11 ∧ · · · ∧ y1t1 ) ⊗ · · · ⊗ (xn1 · · · xnsn ⊗ yn1 ∧ · · · ∧ yntn )

where si + ti = ki and the xij , yij are elements in G, F, respectively such that for
all i, j, xij ≤ xij+1 , yij < yij+1 . We now must define the R–S–K that will assign
to such an element a double standard tableau. But this modification is simple:
we use the procedure that we used before, but this time we allow repeats in the
x entries, and no repeats in those of y. An example here will tell the whole story.
Theorem VI.8.4 247

Example A.4.1 Consider the element


(x2 x2 ⊗ y3 ∧ y4 ) ⊗ (x1 x2 ⊗ y2 ∧ y3 ∧ y4 ) ∈ D4 (ϕ) ⊗ D5 (ϕ).
From this, we build successively the tableaux:
 
  x1 x1 x2 y3 y4 1(4) x
x2 x2 y3 y4 1 (4) → →2
x2 2
   
x1 x2 x2 y4 1(4) y2 x1 x2 x2 y2 1(4) y3

 →
x2 y3 2(2) x2 y3 y4 2(3)
   
x1 x2 x2 y2 y3 1(4) 2 y4 x1 x2 x2 y2 y3 y4 1(4) 2(2)

 .
x2 y3 y4 2(3) x2 y3 y4 2(3)
Here, the overset arrows indicate we are bumping the element set over the
arrow into the tableau, and producing the tableau to the right of the arrow.
The reader can see that this process is essentially the same as that described
in earlier sections, and is reversible. This shows that the set of generating double
tableaux has the same cardinality as the module it is generating, and hence is a
basis.
This page intentionally left blank
REFERENCES

[1] K. Akin and D.A. Buchsbaum, Characteristic-free representation theory of the


general linear group, Adv. Math. 58 (1985), 149–200.
[2] —— and ——, Characteristic-free representation theory of the general linear group,
2. Homological considerations, Adv. Math. 72 (1988), 171–210.
[3] ——, —— and J. Weyman, Schur functors and Schur complexes, Adv. Math. 44
(1982), 207–278.
[4] ——, —— and ——, Resolutions of of determinantal ideals: The submaximal
minors, Adv. Math. 39 (1981), 1–10.
[5] M. Artale, Syzygies of a certain family of generically imperfect modules, J. Algebra
167 (1994), 233–257.
[6] ——, Perfection and representation of GL(N ), J. Algebra 168 (1994), 695–727.
[7] M. Auslander and D.A. Buchsbaum, Homological dimension in local rings,
Trans. Am. Math. Soc. 85 (1957), 390–405.
[8] —— and ——, Homological dimension in Noetherian rings, Trans. Am. Math. Soc.
88 (1958), 194–206.
[9] —— and ——, Codimension and multiplicity, Ann. Math. 68 (1958), 625–657.
[10] —— and ——, Corrections to codimension and multiplicity, Ann. Math. 70 (1959),
395–397.
[11] P. Berthelot, Altérations de variétés algébriques (d’après A. J. de Jong),
Astérisque 241 (1997), 273–311.
[12] G. Boffi and D.A. Buchsbaum, Homotopy equivalence of two families of
complexes, Trans. Am. Math. Soc. 356 (2004), 3077–3107.
[13] —— and R. Sánchez, On the resolutions of the powers of the pfaffian ideal, J.
Algebra 152 (1992), 463–491.
[14] R. Bott, Homogeneous vector bundles, Ann. Math. 66 (1957), 203–248.
[15] N. Bourbaki, Algebra (Part I), Addison-Wesley, Reading, MA (1973).
[16] ——, Algebra (Part II), Springer-Verlag, Berlin (1988).
[17] W. Bruns, The Buchsbaum-Eisenbud structure theorems and alternating syzygies,
Commun. Algebra 15 (1987), 873–925.
[18] D.A. Buchsbaum, A generalized Koszul complex, I, Trans. Am. Math Soc. 111
(1964), 183–196.
[19] ——, Lectures on Regular Local Rings, in Category Theory, Homology Theory and
their Applications, Lecture Notes in Mathematics 86, Springer Verlag, New York
(1969), pp. 13–32.
[20] ——, A remark on the structure maps for finite free resolutions, Commun. Algebra
15 (1987), 21–27.
[21] D.A. Buchsbaum, Letter-place methods and homotopy, in Mathematical Essays
in Honor of Gian-Carlo Rota, Progress in Mathematics 161 Birkhaüser (1998),
pp. 41–62.
250 References

[22] ——, A characteristic-free example of a Lascoux resolution, and letter-place


methods for intertwining numbers, Eur. J. Combin 25 (2004), 1169–1179.
[23] —— and D. Eisenbud, What makes a complex exact?, J. Algebra 25 (1973),
259–268.
[24] —— and ——, Generic free resolutions and a family of generically perfect ideals.
Adv. Math. 18 (1975), 245–301.
[25] —— and ——, Algebra structures for finite free resolutions, and some structure
theorems for ideals of codimension 3, Am. J. Math. 99 (1977), 447–485.
[26] —— and D. Flores de Chela, Intertwining Numbers; the Three-Rowed Case,
J. Algebra 183 (1996), 605–635.
[27] —— and D.S. Rim, A generalized Koszul complex, 2. Depth and Multiplicity,
Trans. Am. Math Soc. 111 (1964), 197–224.
[28] —— and G.-C. Rota, Projective resolutions of Weyl modules, Proc. Natl. Acad.
Sci. USA 90 (1993), 2448–2450.
[29] —— and ——, A new construction in homological algebra, Proc. Natl. Acad. Sci.
USA 91 (1994), 4115–4119.
[30] —— and ——, Approaches to resolution of Weyl modules, Adv. Appl. Math. 27
(2001), 82–191.
[31] —— and B.D. Taylor, Homotopies for resolutions of skew-hook shapes, Adv. Appl.
Math. 30 (2003), 26–43.
[32] L. Burch, On ideals of finite homological dimension in local rings, Proc. Cam. Phil.
Soc. 64 (1968), 941–946.
[33] H. Cartan and S. Eilenberg, Homological Algebra, Princeton University Press,
Princeton 1956.
[34] R.W. Carter and G. Lusztig, On the modular representations of the general
linear and symmetric groups, Math. Z. 136 (1974), 193–242.
[35] I.S. Cohen, Unmixed Ideals, Algebraic Geometry Conference Notes, mimeo-
graphed, University of Chicago, Chicago (1949).
[36] A.J. de Jong, Smoothness, semi-stability and alterations, Publ. Math. IHES 83
(1996), 51–93.
[37] J.A. Eagon, Ideals generated by the subdeterminants of a matrix, Ph.D. thesis,
University of Chicago, (1961).
[38] —— and M. Hochster, A class of perfect determinantal ideals, Bull. Am. Math.
Soc. 76 (1970), 1026–1029.
[39] —— and ——, Cohen-Macaulay rings, invariant theory, and the generic perfection
of determinantal loci, Am. J. Math. 93 (1971), 1020–1059.
[40] —— and D.G. Northcott, Ideals defined by matrices and a certain complex
associated with them, Proc. Royal Soc. A 269 (1962), 188–204.
[41] D. Eisenbud, Commutative Algebra with a View Toward Algebraic Geometry,
Graduate Texts in Mathematics No. 150, Springer-Verlag, New York (1995).
[42] W. Fulton, Intersection Theory, Springer-Verlag, Berlin (1984).
[43] T. Gaffney, Multiplicities and equisingularity of ICIS germs, Invent. Math. 123
(1996), 209–220.
[44] H. Gillet and C. Soulé, K-théorie et nullité des multiplicités d’intersection,
C. R. Acad. Sci. Paris Sér. I Math. 300 (1985), 71–74.
[45] J.A. Green, Polynomial Representations of Gln , Lecture Notes in Mathematics
830, Springer-Verlag, Berlin 1980.
References 251

[46] W. Gröbner, Algebraische Geometrie I, II, Bibliographisches Institut AG,


Mannheim 1968.
[47] M. Hashimoto, Determinantal ideals without minimal free resolutions, Nagoya
Math. J. 118 (1990), 203–216.
[48] D. Hilbert, Über die Theorie der algebraischen Formen, Math. Ann. 36 (1890),
473–534.
[49] M. Hochster, Topics in the Homological Theory of Modules Over Commutative
Rings, Regional Conference Series in Mathematics 24, AMS 1975.
[50] ——, An obstruction to lifting cyclic modules, Pacific J. Math. 61 (1975), 457–463.
[51] ——, Cohen-Macaulay rings, combinatorics, and simplicial complexes, Proceedings
of the Second Oklahoma Ring Theory Conference (March 1976), Marcel-Dekker,
New York (1977), pp. 171–223.
[52] ——, Non-negativity of Intersection Multiplicities in ramified regular local rings
following Gabber/De Jong/Berthelot. Expository Notes (Hochster home page), April
16, 1997.
[53] C. Huneke, Tight Closure and its Applications, Regional Conference Series in
Mathematics 88, AMS 1996.
[54] —— and R. Wiegand, Tensor products of modules and the rigidity of Tor, Math.
Ann. (Historical Archive) 299 (1994), 449–476.
[55] T. Józefiak and P. Pragacz, Ideals generated by pfaffians, J. Algebra 61 (1979),
189–198.
[56] D. Kazhdan and G. Lusztig, Representations of Coxeter groups and Hecke
algebras, Invent. Math. 53 (1979), 165–184.
[57] G.R. Kempf, Linear systems on homogeneous spaces, Ann. Math. 103 (1976),
557–591.
[58] D. Kirby and D. Rees, Multiplicities in graded rings. I. The general theory,
Contemp. Math. 159 (1994), 209–267.
[59] —— and ——, Multiplicities in graded rings. 2. Integral equivalence and the
Buchsbaum-Rim multiplicity, Math. Proc. Cam. Phil. Soc. 119 (1996), 425–445.
[60] S. Kleiman and A. Thorup, A geometric theory of the Buchsbaum-Rim
multiplicity, J. Algebra 167 (1994), 168–231.
[61] —— and ——, Mixed Buchsbaum-Rim multiplicities, Am. J. Math. 118 (1996),
529–569.
[62] K. Kurano, The first syzygies of determinantal ideals, J. Algebra 124 (1989),
414–436.
[63] A.R. Kustin and B. Ulrich, A family of complexes associated to an almost
alternating map, with applications to residual intersections, Mem. Am. Math. Soc.
95 (1992), no. 461, iv + 94 pp.
[64] A. Lascoux, Syzygies des variétés déterminantales. Adv. Math. 30 (1978), 202–237.
[65] G. Lusztig, Some problems in the representation theory of finite Chevalley groups,
Proc. Symp. Pure Math. 37 (1980), 313–317.
[66] I.G. Macdonald, Symmetric Functions and Hall Polynomials, Oxford University
Press (Clarendon), Oxford (1979).
[67] M. Nagata, A general theory of Algebraic Geometry over Dedekind Rings, 2, Am.
J. Math. 80 (1958), 382–420.
[68] H.A. Nielsen, Tensor functors of complexes, Aarhus University Preprint Series No.
15 (1978).
252 References

[69] C. Peskine and L. Szpiro, Dimension projective finie et cohomologie locale, Publ.
Math. IHES 42 (1973), 47–119.
[70] P. Pragacz and J. Weyman, On the construction of resolutions of determinantal
varieties: a survey, in Séminaire d’algèbre Paul Dubreil et Marie-Paule Malliavin,
37ème Année (Paris, 1985), Lecture Notes in Mathematics 1220, Springer-Verlag
(1986), pp. 73–92.
[71] —— and ——, On the generic free resolutions, J. Algebra 128 (1990), 1–44.
[72] G.A. Reisner, Cohen-Macaulay quotients of polynomial rings, Adv. Math. 21
(1976), 30–49.
[73] J. Roberts and J. Weyman, A short proof of a theorem of M. Hashimoto, J.
Algebra 134 (1990), 144–156.
[74] P. Roberts, The vanishing of intersection multiplicities of perfect complexes, Bull.
Am. Math. Soc. 13 (1985), 127–130.
[75] ——, Intersection multiplicities in commutative algebra, in Algebra colloquium
(Rennes, 1985), Univ. Rennes I, Rennes 1985, pp. 25–39.
[76] ——, Multiplicities and Chern Classes in Local Algebra, Cambridge Tracts in
Mathematics No. 133, Cambridge University Press, Cambridge (1998).
[77] P. Samuel, On unique factorization domains, Illinois J. Math. 5 (1961), 1–17.
[78] J.-P, Serre, Algèbre locale. Multiplicités. Cours au Collège de France, 1957–1958,
rédigé par Pierre Gabriel. Seconde édition, 1965. Lecture Notes in Mathematics 11,
Springer-Verlag, Berlin-New York (1965).
[79] W. Soergel, Conjectures de Lusztig, Astérisque 237 (1996), 75–85.
[80] R.G. Swan, Algebraic K-Theory, Lecture Notes in Mathematics 76, Springer-
Verlag, New York (1968).
[81] B.D. Taylor, Generalized straightening laws for products of determinants, Ph.D.
Thesis, MIT, (1997).
[82] ——, Compressed straight tableaux and a distributive lattice of representations, J.
Combin. Theory A 91 (2000), 598–621.
[83] ——, A straightening algorithm for row-convex tableaux, J. Algebra 236 (2001),
155–191.
[84] J. Towber, Two new functors from modules to algebras, J. Algebra 47 (1977),
80–109.
[85] J. Weyman, On the structure of free resolutions of length 3, J. Algebra 126 (1989),
1–33.
[86] ——, Cohomology of Vector Bundles and Syzygies, Cambridge Tracts in Mathem-
atics No. 149, Cambridge University Press, Cambridge, 2003.
[87] O. Zariski and P. Samuel, Commutative Algebra (Vols 1 and 2), Springer-Verlag,
Berlin (1979).
INDEX

acyclic complex, 70 exact complex, 13


Acyclicity Lemma, 107 Extn , 20
almost partition, 152 extension, 21
almost skew-partition (-shape), 150 exterior algebra, 30
alternating map, 128 extreme flippable inversion, 178
Artin–Rees Theorem, 48
associate, 1
associated prime ideal, 50 factorization domain, 2
associator of a module, 50 finite free module, 10
finitely presented, 13
finitistic global dimension, 53
bar algebra, 213 first difference function, 46
bar complex, 80 First Structure Theorem, 110
bar shape, 150 flippable inversion, 175
Betti numbers, 144, 209 free bar module, 214
free dimension, 18
free module, 9
canonical basis, 9 free presentation, 13
Capelli Identities, 166, 167 free resolution, 17
cocommutative graded R-coalgebra, 27
codimension of a module, 52
Cohen–Macaulay module, 63 generic alternating map, 137
Cohen–Macaulay ring, 63 generic (universal) resolution, 144
Cohen–Macaulay Theorem, 55 global dimension, 44
column word of a tableau, 178 Gorenstein ideal, 127
column-convex shape matrix, 150 grade of an ideal, 127
column-standard, 175 graded R-algebra, 23
commutative graded R-algebra, 27 greatest common divisor, 1
complex, 13
connected graded algebra, 215
content of a tableau, 206
coprime, 2 height of a module, 50
cyclic module, 69 height of a prime ideal, 48
hereditary ring, 16
Hilbert basis theorem, 7
Hilbert–Burch Theorem, 111
Dedekind domain, 4
homogeneous element, 23
depth of a module, 52 (ξ) (l)
diagonal map, 26 homogeneous strands Zn,m  Z n,m , 221
differential bar complex, 212 homological conjectures (list of), 66
dimension of a ring, 47 homological dimension, 3, 17
dimension of an R-module, 47 homology functor, 94
discrete valuation ring, 6 Hopf algebra, 26
divided power algebra, 28
divisor, 1
dual basis, 28 ideal of definition, 48
Durfee square, 144, 201 integral weight, 205
intersection multiplicity, 64
intertwining numbers, 207
equicharacteristic, 65 inversion, 175
Euclidean domain, 7 irreducible element, 2
254 Index

Jacobi–Trudi matrix, 228 projective resolution, 17


proper divisor, 1

Koszul complex, 39
Krull Principal Ideal Theorem, 50 Quillen-Suslin theorem, 15, 114

last row (column) of shape matrix, 149 R-algebra, 22


least common multiple, 1 R-coalgebra, 24
length of a partition, 151 R-maps, 9
length of a skew-shape, 151 rank of a free module, 10
length of almost skew-shape, 152 rank of a map, 104
letter-place superalgebra, 164 rank of a projective module, 104
letters, 158 regular local ring, 54
lifting problem, 112 regular sequence, 42
Rel(λ/µ) (almost skew-shape), 171
Rel(λ/µ) (skew-shape), 170
M-sequence, 42 reversed column word of a tableau, 178
mapping cone, 39 rigidity problem, 108
mapping cone construction, 39 Robinson–Schensted–Knuth (R–S–K), 241
maximal spectrum of R, 5 row tableau, 153
measuring identity, 88 row-convex shape matrix, 150
minimal resolution, 53, 144 row-standard tableau, 158, 175
mixed characteristic, 66
modified column word of a tableau, 178
multi-signed alphabet, 175 Samuel multiplicity, 60
multiplicity of q with respect to E, 60 saturated chain condition, 55
multiplicity of M with respect to E, 101 saturated chain of prime ideals, 55
multiplicity of R, 60 Schur algebra, 204
multipliers, 111 Schur complexes, 197
Schur map, 155, 169
Schur modules (general), 155
n-th extension module, 20 Schur modules (hooks), 80
n-th torsion module, 18 Second Structure Theorem, 122
Nakayama’s lemma, 15 separators, 213
noetherian, 3 shape matrix, 149
noetherian ring, 4 short exact sequence, 13
normalized bar complex, 80 signed inequalities, 175, 176, 186
skew-partition (-shape), 150
spectrum of R, 5
(k) split exact sequence, 13
operators Zi,j , 215 standard double tableau, 159, 162, 164
oriented free module, 110 standard tableau, 175
standard tableau (for hooks), 79
straight filling, 182
parameter matrix for a module, 101 straight tableau, 175
partition, 150 subshape, 150
perfect ideal, 52 symmetric algebra, 23
pfaffian, 129 symmetrization, 28
pfaffian ideal, 130 system of divided powers, 129
place polarization, 165–167 system of parameters, 48
places, 158
polynomial, 6
polynomial function, 46 T-boundary operator, 213, 214
power series (formal), 6 t+ -complex, 215
prime element, 2 t+ -graded strand, 215
projective dimension, 3 T-grading, 213
projective module, 14 t-hook, 225
Index 255

t+ -submodule, 215 unique factorization domain (UFD), 2, 59,


tableau with values in S, 153 111
Taylor Algorithm, 182 Universal Coefficient Theorem, 206
tensor algebra, 24 universal freeness, 73
tensor product, 15 unmixed ideal, 56
Torn , 18
torsion element, 18
torsion module, 18 weight of a shape, 153
torsion submodule, 18 weight submodule, 205
torsion-free, 18 Weyl map, 154, 169
transpose (dual) of a shape, 152 Weyl modules (general), 155
trivial extension, 21 Weyl modules (hooks), 80
type of almost skew-shape, 152, 170 Weyl–Schur complex, 190
Weyl–Schur map, 190

unflippable inversion, 175 Z-forms, 203

You might also like