You are on page 1of 516

CAMBRIDGE STUDIES IN ADVANCED MATHEMATICS 136

Editorial Board
B . B O L L O B Á S , W . F U L T O N , A . K A T O K , F . K I R W A N ,
P. SARNAK, B. SIMON, B. TOTARO

Local Cohomology

This Second Edition of a successful graduate text provides a careful and detailed
algebraic introduction to Grothendieck’s local cohomology theory, including in
multi-graded situations, and provides many illustrations of the theory in commutative
algebra and in the geometry of quasi-affine and quasi-projective varieties. Topics
covered include Serre’s Affineness Criterion, the Lichtenbaum–Hartshorne Vanishing
Theorem, Grothendieck’s Finiteness Theorem and Faltings’ Annihilator Theorem,
local duality and canonical modules, the Fulton–Hansen Connectedness Theorem for
projective varieties, and connections between local cohomology and both reductions of
ideals and sheaf cohomology.
The book is designed for graduate students who have some experience of basic
commutative algebra and homological algebra, and also for experts in commutative
algebra and algebraic geometry. Over 300 exercises are interspersed among the text;
these range in difficulty from routine to challenging, and hints are provided for some
of the more difficult ones.

M. P. Brodmann is Emeritus Professor in the Institute of Mathematics at the


University of Zürich.

R. Y. Sharp is Emeritus Professor of Pure Mathematics at the University of Sheffield.


CAMBRIDGE STUDIES IN ADVANCED MATHEMATICS

Editorial Board:
B. Bollobás, W. Fulton, A. Katok, F. Kirwan, P. Sarnak, B. Simon, B. Totaro

All the titles listed below can be obtained from good booksellers or from Cambridge University Press.
For a complete series listing visit: http://www.cambridge.org/mathematics.

Already published
93 D. Applebaum Lévy processes and stochastic calculus (1st Edition)
94 B. Conrad Modular forms and the Ramanujan conjecture
95 M. Schechter An introduction to nonlinear analysis
96 R. Carter Lie algebras of finite and affine type
97 H. L. Montgomery & R. C. Vaughan Multiplicative number theory, I
98 I. Chavel Riemannian geometry (2nd Edition)
99 D. Goldfeld Automorphic forms and L-functions for the group GL(n,R)
100 M. B. Marcus & J. Rosen Markov processes, Gaussian processes, and local times
101 P. Gille & T. Szamuely Central simple algebras and Galois cohomology
102 J. Bertoin Random fragmentation and coagulation processes
103 E. Frenkel Langlands correspondence for loop groups
104 A. Ambrosetti & A. Malchiodi Nonlinear analysis and semilinear elliptic problems
105 T. Tao & V. H. Vu Additive combinatorics
106 E. B. Davies Linear operators and their spectra
107 K. Kodaira Complex analysis
108 T. Ceccherini-Silberstein, F. Scarabotti & F. Tolli Harmonic analysis on finite groups
109 H. Geiges An introduction to contact topology
110 J. Faraut Analysis on Lie groups: An introduction
111 E. Park Complex topological K-theory
112 D. W. Stroock Partial differential equations for probabilists
113 A. Kirillov, Jr An introduction to Lie groups and Lie algebras
114 F. Gesztesy et al. Soliton equations and their algebro-geometric solutions, II
115 E. de Faria & W. de Melo Mathematical tools for one-dimensional dynamics
116 D. Applebaum Lévy processes and stochastic calculus (2nd Edition)
117 T. Szamuely Galois groups and fundamental groups
118 G. W. Anderson, A. Guionnet & O. Zeitouni An introduction to random matrices
119 C. Perez-Garcia & W. H. Schikhof Locally convex spaces over non-Archimedean valued fields
120 P. K. Friz & N. B. Victoir Multidimensional stochastic processes as rough paths
121 T. Ceccherini-Silberstein, F. Scarabotti & F. Tolli Representation theory of the symmetric groups
122 S. Kalikow & R. McCutcheon An outline of ergodic theory
123 G. F. Lawler & V. Limic Random walk: A modern introduction
124 K. Lux & H. Pahlings Representations of groups
125 K. S. Kedlaya p-adic differential equations
126 R. Beals & R. Wong Special functions
127 E. de Faria & W. de Melo Mathematical aspects of quantum field theory
128 A. Terras Zeta functions of graphs
129 D. Goldfeld & J. Hundley Automorphic representations and L-functions for the general linear group, I
130 D. Goldfeld & J. Hundley Automorphic representations and L-functions for the general linear group, II
131 D. A. Craven The theory of fusion systems
132 J. Väänänen Models and games
133 G. Malle & D. Testerman Linear algebraic groups and finite groups of Lie type
134 P. Li Geometric analysis
135 F. Maggi Sets of finite perimeter and geometric variational problems
136 M. P. Brodmann & R. Y. Sharp Local cohomology (2nd Edition)
137 C. Muscalu & W. Schlag Classical and multilinear harmonic analysis, I
138 C. Muscalu & W. Schlag Classical and multilinear harmonic analysis, II
139 B. Helffer Spectral theory and its applications
Local Cohomology
An Algebraic Introduction with
Geometric Applications
second edition

M. P. BRODMANN
Universität Zürich

R. Y. SHARP
University of Sheffield
cambridge university press
Cambridge, New York, Melbourne, Madrid, Cape Town,
Singapore, São Paulo, Delhi, Mexico City
Cambridge University Press
The Edinburgh Building, Cambridge CB2 8RU, UK
Published in the United States of America by Cambridge University Press, New York

www.cambridge.org
Information on this title: www.cambridge.org/9780521513630


C Cambridge University Press 1998, 2013

This publication is in copyright. Subject to statutory exception


and to the provisions of relevant collective licensing agreements,
no reproduction of any part may take place without the written
permission of Cambridge University Press.

First published 1998


Second Edition 2013

Printed and bound in the United Kingdom by the MPG Books Group

A catalogue record for this publication is available from the British Library

ISBN 978-0-521-51363-0 Hardback

Cambridge University Press has no responsibility for the persistence or


accuracy of URLs for external or third-party internet websites referred to
in this publication, and does not guarantee that any content on such
websites is, or will remain, accurate or appropriate.
To Alice
from the second author
Contents

Preface to the First Edition page xi


Preface to the Second Edition xvii
Notation and conventions xxi
1 The local cohomology functors 1
1.1 Torsion functors 1
1.2 Local cohomology modules 3
1.3 Connected sequences of functors 10
2 Torsion modules and ideal transforms 16
2.1 Torsion modules 17
2.2 Ideal transforms and generalized ideal transforms 21
2.3 Geometrical significance 39
3 The Mayer–Vietoris sequence 47
3.1 Comparison of systems of ideals 48
3.2 Construction of the sequence 51
3.3 Arithmetic rank 55
3.4 Direct limits 59
4 Change of rings 65
4.1 Some acyclic modules 66
4.2 The Independence Theorem 70
4.3 The Flat Base Change Theorem 74
5 Other approaches 81
5.1 Use of C̆ech complexes 82
5.2 Use of Koszul complexes 94
5.3 Local cohomology in prime characteristic 101
6 Fundamental vanishing theorems 106
6.1 Grothendieck’s Vanishing Theorem 107
viii Contents

6.2 Connections with grade 112


6.3 Exactness of ideal transforms 117
6.4 An Affineness Criterion due to Serre 122
6.5 Applications to local algebra in prime characteristic 127
7 Artinian local cohomology modules 135
7.1 Artinian modules 135
7.2 Secondary representation 139
7.3 The Non-vanishing Theorem again 143
8 The Lichtenbaum–Hartshorne Theorem 147
8.1 Preparatory lemmas 148
8.2 The main theorem 156
9 The Annihilator and Finiteness Theorems 164
9.1 Finiteness dimensions 164
9.2 Adjusted depths 168
9.3 The first inequality 171
9.4 The second inequality 176
9.5 The main theorems 183
9.6 Extensions 188
10 Matlis duality 193
10.1 Indecomposable injective modules 193
10.2 Matlis duality 199
11 Local duality 211
11.1 Minimal injective resolutions 212
11.2 Local Duality Theorems 216
12 Canonical modules 223
12.1 Definition and basic properties 224
12.2 The endomorphism ring 238
12.3 S2 -ifications 245
13 Foundations in the graded case 251
13.1 Basic multi-graded commutative algebra 253
13.2 *Injective modules 257
13.3 The *restriction property 261
13.4 The reconciliation 271
13.5 Some examples and applications 274
14 Graded versions of basic theorems 285
14.1 Fundamental theorems 286
14.2 *Indecomposable *injective modules 295
14.3 A graded version of the Annihilator Theorem 302
Contents ix

14.4 Graded local duality 309


14.5 *Canonical modules 313
15 Links with projective varieties 331
15.1 Affine algebraic cones 331
15.2 Projective varieties 336
16 Castelnuovo regularity 346
16.1 Finitely generated components 346
16.2 The basics of Castelnuovo regularity 351
16.3 Degrees of generators 358
17 Hilbert polynomials 364
17.1 The characteristic function 366
17.2 The significance of reg2 373
17.3 Bounds on reg2 in terms of Hilbert coefficients 378
17.4 Bounds on reg1 and reg0 383
18 Applications to reductions of ideals 388
18.1 Reductions and integral closures 388
18.2 The analytic spread 393
18.3 Links with Castelnuovo regularity 397
19 Connectivity in algebraic varieties 405
19.1 The connectedness dimension 406
19.2 Complete local rings and connectivity 410
19.3 Some local dimensions 416
19.4 Connectivity of affine algebraic cones 422
19.5 Connectivity of projective varieties 424
19.6 Connectivity of intersections 426
19.7 The projective spectrum and connectedness 432
20 Links with sheaf cohomology 438
20.1 The Deligne Isomorphism 439
20.2 The Graded Deligne Isomorphism 452
20.3 Links with sheaf theory 455
20.4 Applications to projective schemes 465
20.5 Locally free sheaves 476
References 480
Index 485
Preface to the First Edition

One can take the view that local cohomology is an algebraic child of geomet-
ric parents. J.-P. Serre’s fundamental paper ‘Faisceaux algébriques cohérents’
[77] represents a cornerstone of the development of cohomology as a tool in
algebraic geometry: it foreshadowed many crucial ideas of modern sheaf coho-
mology. Serre’s paper, published in 1955, also has many hints of themes which
are central in local cohomology theory, and yet it was not until 1967 that the
publication of R. Hartshorne’s ‘Local cohomology’ Lecture Notes [25] (on A.
Grothendieck’s 1961 Harvard University seminar) confirmed the effectiveness
of local cohomology as a tool in local algebra.
Since the appearance of the Grothendieck–Hartshorne notes, local cohomol-
ogy has become indispensable for many mathematicians working in the theory
of commutative Noetherian rings. But the Grothendieck–Hartshorne notes cer-
tainly take a geometric viewpoint at the outset: they begin with the cohomology
groups of a topological space X with coefficients in an Abelian sheaf on X and
supports in a locally closed subspace.
In the light of this, we feel that there is a need for an algebraic introduc-
tion to Grothendieck’s local cohomology theory, and this book is intended to
meet that need. Our book is designed primarily for graduate students who have
some experience of basic commutative algebra and homological algebra; for
definiteness, we have assumed that our readers are familiar with many of the
basic sections of H. Matsumura’s [50] and J. J. Rotman’s [71]. Our approach
is based on the fundamental ‘δ-functor’ techniques of homological algebra pi-
oneered by Grothendieck, although we shall use the ‘connected sequence’ ter-
minology of Rotman (see [71, pp. 212–214]).
However, we have not overlooked the geometric roots of the subject or the
significance of the ideas for modern algebraic geometry. Indeed, the book
presents several detailed examples designed to illustrate the geometrical sig-
nificance of aspects of local cohomology; we have chosen examples which
xii Preface to the First Edition

require only basic ideas from algebraic geometry. In this spirit, there is one
particular example, which we refer to as ‘Hartshorne’s Example’, to which we
return several times in order to illustrate various points.
The geometric aspects are, in fact, nearer the surface of our treatment than
might initially be realised, because we make much use of ideal transforms and
their universal properties, but it is only in the final chapter that we expose
the fundamental links, expressed by means of the Deligne Correspondence,
between the ideal transform functors and their right derived functors on the
one hand, and section functors of sheaves and sheaf cohomology on the other.
We define the local cohomology functors to be the right derived functors
of the appropriate torsion functor, although we establish in the first chapter
that one can also construct local cohomology modules as direct limits of ‘Ext’
modules; we also present alternative constructions of local cohomology mod-
ules, one via cohomology of C̆ech complexes, and the other via direct limits of
homology modules of Koszul complexes, in Chapter 5. (In fact, we do not use
this Koszul complex approach very much at all in this book.)
Chapters 2, 3 and 4 include fundamental ideas concerning ideal transforms
and their universal properties, the Mayer–Vietoris Sequence for local cohomol-
ogy and the Independence and Flat Base Change Theorems: we regard all of
these as technical cornerstones of the subject, and we certainly use them over
and over again.
The main purpose of Chapters 6 and 7 is the presentation of some of Groth-
endieck’s important vanishing theorems for local cohomology, which relate
such vanishing to the concepts of dimension and grade. This work is mainly
‘algebraic’ in nature. In Chapter 8, we present another vanishing theorem for
local cohomology modules, namely the local Lichtenbaum–Hartshorne Van-
ishing Theorem: this has an ‘analytic’ flavour, in the sense that it is intimately
related with ‘formal’ methods and techniques, that is, with passage to comple-
tions of local rings and with the structure theory for complete local rings. The
Lichtenbaum–Hartshorne Theorem has important geometric applications: for
example, we show in Chapter 19 how it can be used to obtain major results
about the connectivity of algebraic varieties.
Grothendieck’s Finiteness Theorem and G. Faltings’ Annihilator Theorem
for local cohomology are the main subjects of Chapter 9. These two theorems
also have major geometric applications, including, for example, in Macaulay-
fication of schemes. They also have significance for the theory of generalized
Cohen–Macaulay modules and Buchsbaum modules, two concepts which fea-
ture briefly in the exercises in Chapter 9.
We have delayed the introduction of duality (until Chapters 10 and 11) be-
cause quite a lot can be achieved without it, and because, for our discussion
Preface to the First Edition xiii

of duality, we have had to assume (on account of limitations of space) that the
reader is familiar with the Matlis–Gabriel decomposition theory for injective
modules over a commutative Noetherian ring (although we have reviewed that
theory and provided some detailed proofs). We have not explicitly used dual-
izing complexes or derived categories, as it seems to us that such technicalities
could daunt youthful readers and are not essential for a presentation of the
main ideas. After the introduction of local duality in Chapter 11, we show how
this duality can be used to derive some results established earlier in the book
by different means.
The many recent research papers involving local cohomology of graded
rings illustrate the importance of this aspect, and we have made some effort
to develop the fundamentals of local cohomology in the graded case carefully
in Chapters 12 and 13: various representations of local cohomology modules
obtained in the earlier chapters inherit natural gradings when the ring, module
and ideal concerned are all graded, and it seems to us that it is important to
know that there is really only one sensible way of grading local cohomology
modules in such circumstances. Our main aim in Chapter 12 has been to ad-
dress this point. In Chapter 13, ‘graded frills’ are added to basic results proved
earlier in the book.
The short Chapter 14 establishes some links between graded local cohomol-
ogy and projective varieties; it has been included to provide a little geometric
insight, and in order to motivate the work on Castelnuovo–Mumford regularity
in Chapters 15–17, and the connections between ideal transforms and section
functors of sheaves presented in Chapter 20.
In Chapter 15, we study the graded local cohomology of a homogeneous
positively graded commutative Noetherian ring R with respect to the irrelevant
ideal. One of the most important invariants in this context is Castelnuovo–
Mumford regularity. This concept has, in addition to fundamental significance
in projective algebraic geometry, connections with the degrees of generators of
a finitely generated graded R-module M : it turns out that M can be generated
by homogeneous elements of degrees not exceeding reg(M ), the Castelnuovo–
Mumford regularity of M . In turn, this leads on to connections with the theory
of syzygies of finitely generated graded modules over polynomial rings over a
field.
In certain circumstances, including when M is the vanishing ideal IPr (V )
of a projective variety V ⊂ Pr , the above-mentioned reg(M ) coincides with
reg2 (M ), the Castelnuovo–Mumford regularity of M at and above level 2.
In Chapters 16 and 17, we present bounds for the invariant reg2 (M ). Chapter
16 contains a priori bounds which apply whenever the underlying homogen-
eous positively graded commutative Noetherian ring R has Artinian 0-th
xiv Preface to the First Edition

component. Chapter 17 is more specialized, and contains bounds expressed


in terms of coefficients of Hilbert polynomials; our development of this theory
includes a presentation of basic ideas concerning cohomological Hilbert poly-
nomials. The motivation for our work in Chapter 17 comes from D. Mumford’s
classical work [54]: Mumford established the existence of bounds of the type
we present, but, in the spirit of this book, we have added some precision.
One could view Chapters 18 and 19 as propaganda for the effectiveness of
local cohomology as a tool in algebra and geometry. Chapter 18 presents some
applications of Castelnuovo–Mumford regularities to reductions of ideals. This
is a fast developing area, and we have not attempted to give an encyclopaedic
account; instead, we have tried to present the basic ideas and a few recent
results to whet the reader’s appetite. The highlight of Chapter 18 is a theorem
of L. T. Hoa; the statement of this theorem is satisfyingly simple, and makes no
mention of local cohomology, and yet Hoa’s proof, which we present towards
the end of the chapter, makes significant use of graded local cohomology.
Chapter 18 is a good advertisement for local cohomology as a ‘hidden tool’,
and Chapter 19 continues this theme, although here the applications (to the
connectivity of algebraic varieties) are more geometrical in nature. The only
appearances of local cohomology in Chapter 19 are in just two proofs, where a
few central ideas (such as the Mayer–Vietoris Sequence and the Lichtenbaum–
Hartshorne Vanishing Theorem) are used in crucial ways. No hypothesis or
conclusion of any result in the chapter makes any mention of local cohomol-
ogy, and yet we are able to show how the two results whose proofs use local
cohomology can be developed into a theory which leads to proofs of major re-
sults involving connectivity, such as Grothendieck’s Connectedness Theorem,
the Bertini–Gothendieck Connectivity Theorem, the Connectedness Theorem
for Projective Varieties due to W. Barth, to W. Fulton and J. Hansen, and to G.
Faltings, and a ring-theoretic version of Zariski’s Main Theorem. This chapter
is certainly a good advertisement for the power of local cohomology as a tool
in algebraic geometry!
Finally, in Chapter 20, we bring the subject ‘home to its roots’, so to speak,
by presenting links between local cohomology and the cohomology of quasi-
coherent sheaves over certain Noetherian schemes. (Chapter 20 is the only one
for which we have assumed that the reader has some basic knowledge about
schemes and sheaves.)
Some parts of our presentation are fairly leisurely: this is deliberate, and has
been done with graduate students in mind, because we found several prepara-
tory topics where either we knew of no suitable text-book account, or we
felt we had something to add to the existing accounts. Examples are the treat-
ments of Matlis duality in Chapter 10, of *canonical modules in Chapter 13, of
Preface to the First Edition xv

reductions of ideals in Chapter 18, and of connectedness dimensions in Chap-


ter 19; also, our presentation in Chapter 5 of some links between Koszul com-
plexes and local cohomology is deliberately slow.
Our philosophy throughout has been to try to give a careful and accessible
presentation of basic ideas and some important results, illustrating the ideas
with examples, to bring the reader to a level of expertise where he or she can
approach with some confidence recent research papers in local cohomology.
To help with this, the book contains a large number of exercises, and we have
supplied hints for many of the more difficult ones.
We have tried out parts of the book, especially the earlier chapters, on some
of our own graduate students, and their comments have influenced the final
version. We are particularly grateful to Claudia Albertini, Carlo Matteotti,
Francesco Mordasini, Henrike Petzl and Massoud Tousi for acting as ‘guinea
pigs’, so to speak. We should also like to express our gratitude to Peter Gabriel,
John Greenlees, Martin Holland and Josef Rung for continual interest and
encouragement, and to the Schweizerischer Nationalfonds zur Förderung der
wissenschaftlichen Forschung, the Forschungsrat des Instituts für Mathematik
der Universität Zürich, and the University of Sheffield Research Fund, for fi-
nancial support to enable several visits for intense collaboration on the book to
take place. Both authors would like to thank Alice Sharp: Markus Brodmann
thanks her for kind hospitality during pleasant visits to Sheffield for discus-
sions on the book; and Rodney Sharp thanks her for much sympathetic support
through the years during which this book was being written (as well as for
many things which have nothing to do with local cohomology). Finally, we
are very grateful to David Tranah and Roger Astley of Cambridge University
Press for their continual encouragement and assistance over many years, and,
not least, for their cooperation over our request that the blue stripe on the cover
of the book should match the blue of the Zürich trams!

Markus Brodmann Rodney Sharp


Zürich Sheffield
April 1997
Preface to the Second Edition

In the fifteen years since we completed the First Edition of this book, we have
had opportunity to reflect on how we could change it in order to enhance its
usefulness to the graduate students at whom it is aimed. As a result, this Sec-
ond Edition shows substantial differences from the First. The main ones are
described as follows.
One of the more dramatic changes is the introduction of a complete new
chapter, Chapter 12, devoted to the study of canonical modules. The treatment
of canonical modules in the First Edition was brief and restricted to the case
where the underlying ring is Cohen–Macaulay; we assumed that the reader was
familiar with the treatment in this case by W. Bruns and J. Herzog in their book
on Cohen–Macaulay rings (see [7]). In our new Chapter 12, we present some
of the basic work of Y. Aoyama (see [1] and [2]) and follow M. Hochster and
C. Huneke [39] in defining a canonical module over a (not necessarily Cohen–
Macaulay) local ring (R, m) to be a finitely generated R-module whose Matlis
dim R
dual is isomorphic to the ‘top’ local cohomology module Hm (R). Thus
this topic is intimately related to local cohomology.
Canonical modules have connections with the theory of S2 -ifications (here,
the ‘S2 ’ refers to Serre’s condition), and we realised that the development of
S2 -ifications can be facilitated by generalizations of arguments we had used to
study ideal transforms in §2.2 of the First Edition. For this reason, instead of
dealing just with ideal transforms based on the sequence of powers of a fixed
ideal, we treat, in §2.2 of this Second Edition, a generalization based on a set
B of ideals such that, whenever b, c ∈ B, there exists d ∈ B with d ⊆ bc.
This represents a significant change to Chapter 2.
Another major change concerns our treatment of graded local cohomology.
Our new Chapters 13 and 14 treat local cohomology in the situation where the
rings, ideals and modules concerned are graded by Zn , where n is a positive
integer. In the First Edition we dealt only with the case where n = 1; in the
xviii Preface to the Second Edition

years since that edition was published, there have been more and more uses
of local cohomology in multi-graded situations. It was not difficult for us to
adapt the treatment of Z-graded local cohomology from the First Edition to
the multi-graded case. The main point of our Chapter 13 is to show that, even
though there appear to be various possible approaches, there is really only
one sensible way of grading local cohomology. Chapter 14 adds ‘(Zn -)graded
frills’ to basic results proved earlier in the book. We illustrate this work with
some calculations over polynomial rings and Stanley–Reisner rings. Chapters
13 and 14 are generalizations of Chapters 12 and 13 from the First Edition;
however, we found it desirable to present some fundamental results from S.
Goto’s and K.-i. Watanabe’s paper [22] about Zn -graded rings and modules;
those results in the particular case when n = 1 are more readily available.
The last two decades have seen a surge in the use of local cohomology as a
tool in ‘characteristic p’ commutative algebra, that is, the study of commutative
Noetherian rings of prime characteristic p. The key to this is the fact that, for
an ideal a of such a ring R, and any integer i ≥ 0, the i-th local cohomology
module Hai (R) of R itself with respect to a has a so-called ‘Frobenius action’.
In the new §5.3, we explain why this Frobenius action exists, and in the new
§6.5, we use it to present Hochster’s proof of his Monomial Conjecture (in
characteristic p), and to give some examples of how local cohomology can be
used as an effective tool in tight closure theory.
Another new section is §20.5 about locally free sheaves; here we prove
Serre’s Cohomological Criterion for Local Freeness, Horrocks’ Splitting Cri-
terion and Grothendieck’s Splitting Theorem. We have also expanded §20.4
with an additional application to projective schemes: we now include a result
of Serre about the global generation of twisted coherent sheaves.
In order to include all this new material, we have had to omit some items
that were included in the First Edition but which, we now consider, no longer
command sufficiently compelling reasons for inclusion. The main topics that
fall under this heading are the old §11.3 containing some applications of local
duality (one can take the view that the new Chapter 12, on canonical modules,
represents a major application of local duality), and the a priori bounds of di-
agonal type on Castelnuovo–Mumford regularity at and above level 2 that were
treated in Chapter 16 of the First Edition (Chapter 17 has been reorganized to
smooth over the omission, and expanded by the addition of further bounding
results which follow from our generalized version of Mumford’s bound on
regularity at and above level 2).
There are also many minor changes, designed to improve the presentation or
the usefulness of the book. For example, the treatment of Faltings’ Annihilator
Theorem in Chapter 9 now applies to two arbitrary ideals a and b, whereas in
Preface to the Second Edition xix

the First Edition we treated only the case where b ⊆ a; the syzygetic character-
ization of Castelnuovo–Mumford regularity is given a full proof in the Second
Edition; and the graded Deligne Isomorphism in §20.2 is presented here for a
multi-graded situation.
We should also point out that two comments made about the First Edition
in its Preface do not apply to this Second Edition. Firstly, the example studied
in 2.3.7, 3.3.5, 4.3.7, . . . is not called ‘Hartshorne’s Example’ in this Second
Edition (but we do cite Hartshorne’s paper [28] when we first consider this
example); secondly, there are a few more appearances of local cohomology in
Chapter 19 than there were in the First Edition (because we followed a sug-
gestion of M. Varbaro that, in some formulas, the arithmetic rank of an ideal a
could be replaced by the cohomological dimension of a). Nevertheless, Chap-
ter 19 still contains several exciting examples of situations which represent
‘hidden applications’ of local cohomology, in the sense that significant results
that do not mention local cohomology in either their hypotheses or their con-
clusions have proofs in this book that depend on local cohomology. There are
other examples of such ‘hidden applications’ in §6.5 and in Chapter 18.
We would like to add, to the list of people thanked in the Preface to the First
Edition, several more of our students, namely Roberto Boldini, Stefan Fuma-
soli, Simon Kurmann, Nicole Nossem and Fred Rohrer, who all contributed
to this Second Edition, either by providing constructive criticism of the First
Edition, or by trying out drafts of changed or new sections that we planned to
include in the Second Edition. We are grateful to them all.
We thank the Schweizerischer Nationalfonds zur Förderung der wissensch-
aftlichen Forschung, the Forschungsrat des Instituts für Mathematik der Uni-
versität Zürich, and the Department of Pure Mathematics of the University of
Sheffield, for financial support for visits for collaboration on this Second Edi-
tion. We are also particularly grateful to the Scientific Council of the Centre
International de Rencontres Mathématiques (CIRM) at Luminy, Marseille, for
their award to us of a two-week ‘research in pairs’ in Spring 2011 that enabled
us, in the excellent environment for mathematical research at CIRM, to pro-
duce a complete draft. We again thank Alice Sharp for her continued support
and encouragement for our project. It is also a pleasure for us to record our
gratitude to Roger Astley and Clare Dennison of Cambridge University Press
for their encouragement and support.

Markus Brodmann Rodney Sharp


Zürich Sheffield
April 2012
Notation and conventions

All rings considered in this book will have identity elements.


Throughout the book, R will always denote a non-trivial commutative Noe-
therian ring, and a will denote an ideal of R. We shall only assume that R
has additional properties (such as being local) when these are explicitly stated;
however, the phrase ‘(R, m) is a local ring’ will mean that R is a commutative
Noetherian quasi-local ring with unique maximal ideal m.
For an ideal c of R, we denote Supp(R/c) = {p ∈ Spec(R) : p ⊇ c} by
Var(c), and refer to this as the variety of c.
By a multiplicatively closed subset of R, we shall mean a subset of R which
is closed under multiplication and contains 1. It should be noted (and this com-
ment is relevant for the final chapter) that, if S is a non-empty subset of R
which is closed under multiplication, then, even if S does not contain 1, we
can form the commutative ring S −1 R and, for an R-module M , the S −1 R-
module S −1 M . In fact, S −1 R ∼ = (S ∪ {1})−1 R, and, in S −1 R, the element
sr/s, for r ∈ R and s ∈ S, is independent of the choice of such s; similar
comments apply to S −1 M .
The symbol Z will always denote the ring of integers; in addition, N (respec-
tively N0 ) will always denote the set of positive (respectively non-negative)
integers. The field of rational (respectively real, complex) numbers will be de-
noted by Q (respectively R, C).
The category of all modules and homomorphisms over a commutative ring
R will be denoted by C(R ). When R is G-graded, where G is a finitely
generated, torsion-free Abelian group, the category of all graded R -modules
and homogeneous R -homomorphisms will be denoted by *C(R ) (or *C G (R )
when it is desirable to indicate the grading group G).
The symbol ⊆ will stand for ‘is a subset of’; the symbol ⊂ will be reserved
to denote strict inclusion. Thus, for sets A, B, the expression A ⊂ B means
that A ⊆ B and A = B.
xxii Notation and conventions

The identity mapping on a set A will be denoted by IdA . If f : A → C is


a mapping from the set A to the set C, and S ⊆ A, then f S : S → C will
denote the restriction of f to S. Thus f S (s) = f (s) for all s ∈ S.
Some of the exercises in the book are needed for the main development
later in the book, and these exercises are marked with a ‘’; however, exercises
which are used later in the book but only in other exercises have not been
marked with a ‘’.
1

The local cohomology functors

The main objective of this chapter is to introduce the a-torsion functor Γa


(throughout the book, a always denotes an ideal in a (non-trivial) commutative
Noetherian ring R) and its right derived functors Hai (i ≥ 0), referred to as the
local cohomology functors with respect to a. We shall see that Γa is naturally
equivalent to the functor lim HomR (R/an , • ) and, indeed, that Hai is natu-
−→
n∈N
rally equivalent to the functor lim ExtiR (R/an , • ) for each i ≥ 0; moreover,
−→
n∈N
as Γa turns out to be left exact, the functors Γa and Ha0 are naturally equivalent.
This chapter also serves notice that our approach is based on fundamental
techniques of homological commutative algebra, such as ones based on con-
nected sequences of functors (see [71, pp. 212–214]): readers familiar with
such ideas, and with the local cohomology functors, might like to just glance
through this chapter and to move rapidly on to Chapter 2.

1.1 Torsion functors



1.1.1 Definition. For each R-module M , set Γa (M ) = n∈N (0 :M an ),
the set of elements of M which are annihilated by some power of a. Note that
Γa (M ) is a submodule of M . For a homomorphism f : M −→ N of R-
modules, we have f (Γa (M )) ⊆ Γa (N ), and so there is a mapping Γa (f ) :
Γa (M ) −→ Γa (N ) which agrees with f on each element of Γa (M ).
It is clear that, if g : M → N and h : N → L are further homomorphisms
of R-modules and r ∈ R, then Γa (h ◦ f ) = Γa (h) ◦ Γa (f ), Γa (f + g) =
Γa (f ) + Γa (g), Γa (rf ) = rΓa (f ) and Γa (IdM ) = IdΓa (M ) . Thus, with these
assignments, Γa becomes a covariant, R-linear functor from C(R) to itself.
(We say that a functor T : C(R) −→ C(R) is R-linear precisely when it is
2 The local cohomology functors

additive and T (rf ) = rT (f ) for all r ∈ R and all homomorphisms f of


R-modules.) We call Γa the a-torsion functor.

1.1.2 Exercise. Let b be a second ideal of R. Show that

Γa (Γb (M )) = Γa+b (M )

for each R-module M .

1.1.3 Exercise. Let b be a second ideal of R. Show that Γa = Γb if and only


√ √
if a = b.

(The notation , attached to some exercises, is explained in the section of


‘Notation and conventions’ following the Preface to the Second Edition.)

1.1.4 Exercise. Suppose that the ideal b of R is a reduction of a; that is,


b ⊆ a and there exists s ∈ N such that bas = as+1 . Show that Γa = Γb .

1.1.5 Exercise. For a prime number p, find ΓpZ (Q/Z).

1.1.6 Lemma. The a-torsion functor Γa : C(R) −→ C(R) is left exact.


f g
Proof. Let 0 −→ L −→ M −→ N −→ 0 be an exact sequence of R-
modules and R-homomorphisms. We must show that
Γa (f ) Γa (g)
0 - Γa (L) - Γa (M ) - Γa (N )

is still exact. It is clear that Γa (f ) is a monomorphism and it follows immedi-


ately from 1.1.1 that Γa (g) ◦ Γa (f ) = 0, so that

Im(Γa (f )) ⊆ Ker(Γa (g)).

To prove the reverse inclusion, let m ∈ Ker(Γa (g)). Thus m ∈ Γa (M ), so that


there exists n ∈ N such that an m = 0, and g(m) = 0. Now there exists l ∈ L
such that f (l) = m, and our proof will be complete if we show that l ∈ Γa (L).
To achieve this, note that, for each r ∈ an , we have f (rl) = rf (l) = rm = 0,
so that rl = 0 because f is a monomorphism. Hence an l = 0.

The result of Lemma 1.1.6 will become transparent to many readers once
we have covered a little more theory, and related the a-torsion functor Γa to a
functor defined in terms of direct limits of ‘Hom’ modules. However, before
we proceed in that direction, we are going to introduce, at this early stage, the
fundamental definition of the local cohomology modules of an R-module M
with respect to a.
1.2 Local cohomology modules 3

1.2 Local cohomology modules


1.2.1 Definitions. For i ∈ N0 , the i-th right derived functor of Γa is denoted
by Hai and will be referred to as the i-th local cohomology functor with respect
to a.
For an R-module M , we shall refer to Hai (M ), that is, the result of applying
the functor Hai to M , as the i-th local cohomology module of M with respect
to a, and to Γa (M ) as the a-torsion submodule of M . We shall say that M is
a-torsion-free precisely when Γa (M ) = 0, and that M is a-torsion precisely
when Γa (M ) = M , that is, if and only if each element of M is annihilated by
some power of a.
It is probably appropriate for us to stress the implications of the above def-
inition at this point, and list some basic properties of the local cohomology
modules.
1.2.2 Properties of local cohomology modules. Let M be an arbitrary R-
module.
(i) To calculate Hai (M ), one proceeds as follows. Take an injective reso-
lution
d−1 d0 di
I • : 0 −→ I 0 −→ I 1 −→ · · · −→ I i −→ I i+1 −→ · · ·
of M , so that there is an R-homomorphism α : M −→ I 0 such that the
sequence
α d0 di
0 −→ M −→ I 0 −→ I 1 −→ · · · −→ I i −→ I i+1 −→ · · ·
is exact. Apply the functor Γa to the complex I • to obtain
Γa (d0 ) Γa (di )
0 - Γa (I 0 ) - ··· - Γa (I i ) - Γa (I i+1 ) - ···

and take the i-th cohomology module of this complex; the result,
Ker(Γa (di ))/ Im(Γa (di−1 )),
which, by a standard fact of homological algebra, is independent (up to R-
isomorphism) of the choice of injective resolution I • of M , is Hai (M ).
(ii) Since Γa is covariant and R-linear, it is automatic that each local co-
homology functor Hai (i ∈ N0 ) is again covariant and R-linear.
(iii) Since Γa is left exact, Ha0 is naturally equivalent to Γa . Thus, loosely,
we can use this natural equivalence to identify these two functors.
(iv) The reader should be aware of the long exact sequence of local coho-
mology modules which results from a short exact sequence of R-modules and
R-homomorphisms, and so we spell out the details here.
4 The local cohomology functors
f g
Let 0 −→ L −→ M −→ N −→ 0 be an exact sequence of R-modules and
R-homomorphisms. Then, for each i ∈ N0 , there is a connecting homomor-
phism Hai (N ) → Hai+1 (L), and these connecting homomorphisms make the
resulting long sequence
Ha0 (f ) Ha0 (g)
0 - H 0 (L) - H 0 (M ) - H 0 (N )
a a a
Ha1 (f ) Ha1 (g)
- H 1 (L) - H 1 (M ) - H 1 (N )
a a a

- ··· ···
Hai (f ) Hai (g)
- H i (L) - H i (M ) - H i (N )
a a a

- H i+1 (L) - ···


a

exact. The reader should also be aware of the ‘natural’ or ‘functorial’ properties
of these long exact sequences: if
f g
0 - L - M - N - 0

λ μ ν

? ? ?
f g
0 - L - M - N - 0

is a commutative diagram of R-modules and R-homomorphisms with exact


rows, then, for each i ∈ N0 , we not only have a commutative diagram
Hai (f ) Hai (g)
Hai (L) - H i (M ) - H i (N )
a a

Hai (λ) Hai (μ) Hai (ν)

? ? ?
Hai (f  ) Hai (g  )
Hai (L ) - H i (M  ) - H i (N  )
a a

(simply because Hai is a functor!), but we also have a commutative diagram

Hai (N ) - H i+1 (L)


a

Hai (ν) Hai+1 (λ)

? ?
Hai (N  ) -H i+1 (L )
a

in which the horizontal maps are the appropriate connecting homomorphisms.


1.2 Local cohomology modules 5

The following remark will be used frequently in applications. It is an easy


consequence of Exercise 1.1.3 and the definition of local cohomology functors
in 1.2.1.
√ √
1.2.3 Remark. Let b be a second ideal of R such that a = b. Then
Hai = Hbi for all i ∈ N0 , so that Hai (M ) = Hbi (M ) for each R-module M
and all i ∈ N0 .
The next four exercises might help the reader to consolidate the properties of
local cohomology modules listed in 1.2.2. The first three of these exercises (for
which non-trivial results from commutative algebra about injective dimension
over the relevant rings are very helpful) give a tiny foretaste of results about
the vanishing of local cohomology modules which are central to the subject,
and which will feature prominently later in the book.
1.2.4 Exercise. Show that, for every Abelian group (that is, Z-module) G
and for every a ∈ Z, we have HZa
i
(G) = 0 for all i ≥ 2.
1.2.5 Exercise. Suppose that (R, m) is a regular local ring of dimension d.
Show that, for each R-module M , we have Hai (M ) = 0 for all i > d.
1.2.6 Exercise. Suppose that (R, m) is a Gorenstein local ring (see, for ex-
ample, Matsumura [50, p. 142]) of dimension d. Show that, for each finitely
generated R-module M of finite projective dimension, we have Hai (M ) = 0
for all i > d. (Here is a hint: use the fact [50, Theorem 18.1] that the injective
dimension of R as an R-module is d, and then use induction on the projective
dimension of M .)
The next exercise investigates the behaviour of local cohomology modules
under fraction formation: its results show that, speaking loosely, the local co-
homology functors ‘commute’ with fraction formation. This is a fundamental
fact in the subject; however, we shall actually derive it as an immediate conse-
quence of a more general result in Chapter 4 concerning the behaviour of local
cohomology under flat base change (and we shall not make use of it until after
Chapter 4). Nevertheless, even at this early stage, its proof should not present
much difficulty for a reader familiar with the fact (proved in 10.1.14) that, if I
is an injective R-module and S is a multiplicatively closed subset of R, then
S −1 I is an injective S −1 R-module.
1.2.7 Exercise. Let M be an R-module and let S be a multiplicatively closed
subset of R. Show that S −1 (Γa (M )) = ΓaS −1 R (S −1 M ), and that, for all
i ∈ N0 , there is an isomorphism of S −1 R-modules
S −1 (Hai (M )) ∼ i
= HaS −1 R (S
−1
M ).
6 The local cohomology functors

It is now time for us to relate the a-torsion functor Γa to a functor defined in


terms of direct limits of ‘Hom’ modules. Fundamental to the discussion is the
natural isomorphism, for an R-module M and n ∈ N,

=
φ := φan ,M : HomR (R/an , M ) −→ (0 :M an )

for which φ(f ) = f (1 + an ) for all f ∈ HomR (R/an , M ). In fact, we are go-
ing to put the various φan ,M (n ∈ N) together to obtain a natural isomorphism

=
lim HomR (R/an , M ) −→ Γa (M ), but before we do this it might be helpful
−→
n∈N
to the reader if we give some general considerations about functors and direct
limits, as the principles involved will be used numerous times in this book.

1.2.8 Remarks. Let (Λ, ≤) be a (non-empty) directed partially ordered set,


and suppose that we are given an inverse system of R-modules (Wα )α∈Λ over
Λ, with constituent R-homomorphisms hα β : Wα → Wβ (for each (α, β) ∈
Λ × Λ with α ≥ β). Let T : C(R) × C(R) → C(R) be an R-linear functor
of two variables which is contravariant in the first variable and covariant in the
second. (A functor U : C(R) ×C(R) −→ C(R) is said to be R-linear precisely
when it is additive and U (rf, g) = rU (f, g) = U (f, rg) for all r ∈ R and all
homomorphisms f, g of R-modules.) We show now how these data give rise
to a covariant, R-linear functor

lim T (Wα , • ) : C(R) −→ C(R).


−→
α∈Λ

Let M, N be R-modules and let f : M −→ N be an R-homomorphism.


β : Wα −→ Wβ induces an
For α, β ∈ Λ with α ≥ β, the homomorphism hα
R-homomorphism

β , M ) : T (Wβ , M ) −→ T (Wα , M ),
T (hα

and the fact that T is a functor ensures that the T (hα


β , M ) turn the family
(T (Wα , M ))α∈Λ into a direct system of R-modules and R-homomorphisms
over Λ. We may therefore form lim T (Wα , M ). Moreover, again for α, β ∈ Λ
−→
α∈Λ
with α ≥ β, we have a commutative diagram
T (hα
β ,M )
T (Wβ , M ) - T (Wα , M )

T (Wβ ,f ) T (Wα ,f )

? T (hα
?
β ,N )
T (Wβ , N ) - T (Wα , N ) ;
1.2 Local cohomology modules 7

therefore the T (Wα , f ) (α ∈ Λ) constitute a morphism of direct systems and


so induce an R-homomorphism

lim T (Wα , f ) : lim T (Wα , M ) −→ lim T (Wα , N ).


−→ −→ −→
α∈Λ α∈Λ α∈Λ

It is now straightforward to check that, in this way, lim T (Wα , • ) becomes


−→
α∈Λ
a covariant, R-linear functor from C(R) to itself. Observe that, since passage
to direct limits preserves exactness, if T is left exact, then so too is this new
functor.

1.2.9 Examples. Here we present some examples that are central for our
subject.

(i) Probably the most important examples for us of the ideas of 1.2.8 con-
cern the case where we take for Λ the set N of positive integers with its
usual ordering and the inverse system (R/an )n∈N of R-modules under
the natural homomorphisms hnm : R/an → R/am (for n, m ∈ N with
n ≥ m) (in such circumstances, an ⊆ am , of course). In this way, we
obtain covariant, R-linear functors

lim HomR (R/an , • ) and lim ExtiR (R/an , • ) (i ∈ N0 )


−→ −→
n∈N n∈N

from C(R) to itself. Of course, the natural equivalence between the left
exact functors HomR and Ext0R leads to a natural equivalence between
the left exact functors

lim HomR (R/an , • ) and lim Ext0R (R/an , • )


−→ −→
n∈N n∈N

which we shall use without further comment.


(ii) Very similar considerations, this time based on the inclusion maps an →
am (for n, m ∈ N with n ≥ m), lead to functors (which are again
covariant and R-linear)

lim HomR (an , • ) and lim ExtiR (an , • ) (i ∈ N0 )


−→ −→
n∈N n∈N

from C(R) to itself, and a natural equivalence between the left exact
functors

lim HomR (an , • ) and lim Ext0R (an , • ).


−→ −→
n∈N n∈N

These functors will be considered in detail in Chapter 2.


8 The local cohomology functors

It will be convenient for us to consider situations slightly more general than


that studied in 1.2.9(i) above.

1.2.10 Definition and Example. Let (Λ, ≤) be a (non-empty) directed par-


tially ordered set. By an inverse family of ideals (of R) over Λ, we mean a fam-
ily (bα )α∈Λ of ideals of R such that, whenever (α, β) ∈ Λ × Λ with α ≥ β,
we have bα ⊆ bβ .
For example, if

b1 ⊇ b2 ⊇ · · · ⊇ bn ⊇ bn+1 ⊇ · · ·

is a descending chain of ideals of R, then (bn )n∈N is an inverse family of ideals


over N (with its usual ordering). In particular, the family (an )n∈N is an inverse
family of ideals over N.
Let (bα )α∈Λ be an inverse family of ideals of R over Λ. Then the natural
R-homomorphisms hα β : R/bα → R/bβ (for α, β ∈ Λ with α ≥ β) turn
(R/bα )α∈Λ into an inverse system over Λ, and so we can apply the ideas of
1.2.8 to produce covariant, R-linear functors

lim HomR (R/bα , • ) and lim ExtiR (R/bα , • ) (i ∈ N0 )


−→ −→
α∈Λ α∈Λ

(from C(R) to itself), the first two of which are left exact and naturally equiv-
alent.

1.2.11 Theorem. Let B = (bα )α∈Λ be an inverse family of ideals of R over


Λ, as in 1.2.10.

(i) There is a covariant, R-linear functor ΓB : C(R) → C(R) which is such


that, for an R-module M ,

ΓB (M ) = (0 :M bα ),
α∈Λ

and, for a homomorphism f : M −→ N of R-modules, ΓB (f ) :


ΓB (M ) −→ ΓB (N ) is just the restriction of f to ΓB (M ).
(ii) There is a natural equivalence

φ (= φB ) : lim HomR (R/bα , • ) −→ ΓB
=
−→
α∈Λ

(of functors from C(R) to itself) which is such that, for an R-module
M and α ∈ Λ, the image under φM of the natural image of an h ∈
HomR (R/bα , M ) is h(1 + bα ). Consequently, ΓB is left exact.
1.2 Local cohomology modules 9

(iii) In particular, there is a natural equivalence



=
φ0 (= φ0a ) : lim HomR (R/an , • ) −→ Γa
−→
n∈N

which is such that, for an R-module M and n ∈ N, the image under φ0M
of the natural image of an h ∈ HomR (R/an , M ) is h(1 + an ).
Proof. (i) This can be proved by straightforward modification of the ideas of
1.1.1, and so will be left to the reader.
(ii) Let f : M −→ N be a homomorphism of R-modules. For each α ∈ Λ,
let φbα ,M : HomR (R/bα , M ) −→ (0 :M bα ) be the R-isomorphism for
which φbα ,M (h) = h(1 + bα ) for all h ∈ HomR (R/bα , M ). Let α, β ∈ Λ
with α ≥ β, and let hα β : R/bα → R/bβ be as in 1.2.10. Since the diagram
φbβ ,M
HomR (R/bβ , M ) ∼
- (0 :M bβ )
=

HomR (hα
β ,M )

? φbα ,M
?
HomR (R/bα , M ) ∼
- (0 :M bα )
=

(in which the right-hand vertical map is inclusion) commutes, it follows that
there is indeed an R-isomorphism
∼ 
φM : lim HomR (R/bα , M ) −→ ΓB (M ) =
=
(0 :M bα )
−→
α∈Λ α∈Λ

as described in the statement of the theorem. It is easy to check that the diagram
φM
lim HomR (R/bα , M ) ∼
- ΓB (M )
−→ =
α∈Λ

lim HomR (R/bα ,f ) ΓB (f )


−→
α∈Λ
? ?
φN
lim HomR (R/bα , N ) ∼
- ΓB (N )
−→ =
α∈Λ

commutes, and the final claim is then immediate from 1.2.10.


(iii) This is immediate from (ii), since when we apply (ii) to the family of
ideals B := (an )n∈N , the functor ΓB of (i) is just the a-torsion functor Γa .
We commented earlier that it would in time become transparent that Γa is
left exact: we had 1.2.11 in mind when we made that comment.
1.2.12 Exercise. Provide a proof for part (i) of 1.2.11.
10 The local cohomology functors

1.3 Connected sequences of functors


In this section, we are going to use the concepts of ‘connected sequence of
functors’ and ‘strongly connected sequence of functors’. These are explained
on p. 212 of Rotman’s book [71]. For the reader’s convenience, we recall here
relevant definitions in the case of negative connected sequences, as we shall be
particularly concerned with this case.
1.3.1 Definition. Let R be a commutative ring.
A sequence (T i )i∈N0 of covariant functors from C(R) to C(R ) is said to
be a negative connected sequence (respectively, a negative strongly connected
sequence) if the following conditions are satisfied.
f g
(i) Whenever 0 −→ L −→ M −→ N −→ 0 is an exact sequence in C(R),
there are defined connecting R -homomorphisms
T i (N ) −→ T i+1 (L) for all i ∈ N0
such that the long sequence
T 0 (f ) T 0 (g)
0 - T 0 (L) - T 0 (M ) - T 0 (N )
1 1
- T 1 (L) T (f )
- T 1 (M ) T (g)
- T 1 (N )
- ··· ···
T i (f ) T i (g)
- T i (L) - T i (M ) - T i (N )
- T i+1 (L) - ···

is a complex (respectively, is exact).


(ii) Whenever

0 - L - M - N - 0

λ μ ν

? ? ?
0 - L - M - N - 0

is a commutative diagram of R-modules and R-homomorphisms with


exact rows, then there is induced, by λ, μ and ν, a chain map of the long
complex of (i) for the top row into the corresponding long complex for
the bottom row.
It might help if we remind the reader of the convention regarding the rais-
ing and lowering of indices in a situation such as that of 1.3.1, under which
1.3 Connected sequences of functors 11

T i would be written as T−i : with this convention, (T i )i≥0 can be written as


(Tj )j≤0 .
We also point out that, if T : C(R) → C(R ) is an additive covariant functor,
such as Γa , then its sequence of right derived functors (Ri T )i∈N0 is a nega-
tive strongly connected sequence of covariant functors from C(R) to C(R );
furthermore, if T is left exact, then R0 T is naturally equivalent to T . We shall
be concerned so often with left exact, additive, covariant functors that it will
considerably simplify the exposition if we adopt now the following convention
which will be in force for the rest of the book.
1.3.2 Convention. Whenever R is a commutative ring and T : C(R) −→
C(R ) is a covariant, additive functor which is left exact, then we shall identify
T with its 0-th right derived functor R0 T in the natural way. Likewise, we
shall identify Ext0R with HomR in the natural way.
1.3.3 Definition. Let R be a commutative ring, and let (T i )i∈N0 , (U i )i∈N0
be negative connected sequences of covariant functors from C(R) to C(R ). A
homomorphism Ψ : (T i )i∈N0 −→ (U i )i∈N0 of connected sequences is a family
(ψ i )i∈N0 where, for each i ∈ N0 , ψ i : T i → U i is a natural transformation of
functors, and which is such that the following condition is satisfied: whenever
0 −→ L −→ M −→ N −→ 0 is an exact sequence of R-modules and
R-homomorphisms, then, for each i ∈ N0 , the diagram

T i (N ) - T i+1 (L)

i i+1
ψN ψL

? ?
i
U (N ) - U i+1 (L)

(in which the horizontal maps are the appropriate connecting homomorphisms
arising from the connected sequences) commutes.
A homomorphism Ψ = (ψ i )i∈N0 : (T i )i∈N0 −→ (U i )i∈N0 of connected
sequences is said to be an isomorphism (of connected sequences) precisely
when ψ i : T i → U i is a natural equivalence of functors for each i ∈ N0 .
We hope the reader is sufficiently adept at techniques similar to those on
pp. 212–214 of [71] to find the following exercise straightforward; if not, he
or she might like to study Theorem 10 (and its Corollary) of Section 6.5 of
Northcott [60], which together provide a solution.
1.3.4 Exercise. Let R be a commutative ring, and let (T i )i∈N0 , (U i )i∈N0 be
two negative connected sequences of covariant functors from C(R) to C(R ).
12 The local cohomology functors

(i) Let ψ 0 : T 0 → U 0 be a natural transformation of functors. Assume that


(a) the sequence (T i )i∈N0 is strongly connected, and
(b) T i (I) = 0 for all i ∈ N and all injective R-modules I.
Show that there exist uniquely determined natural transformations
ψi : T i → U i (i ∈ N)
such that (ψ i )i∈N0 : (T i )i∈N0 −→ (U i )i∈N0 is a homomorphism of
connected sequences.
(ii) Let ψ : T 0 → U 0 be a natural equivalence of functors. Assume that
(a) the sequence (T i )i∈N0 is strongly connected,
(b) the sequence (U i )i∈N0 is strongly connected, and
(c) T i (I) = U i (I) = 0 for all i ∈ N and all injective R-modules I.
By part (i), there is a unique homomorphism of connected sequences
Ψ := (ψ i )i∈N0 : (T i )i∈N0 −→ (U i )i∈N0 for which ψ 0 = ψ. Show that
Ψ is actually an isomorphism of connected sequences.

We shall not state explicitly the analogues of 1.3.1, 1.3.3 and 1.3.4 for pos-
itive connected sequences, but we warn the reader now that we shall use such
analogues in Chapters 11 and 12.
The following consequence of 1.3.4(ii) essentially provides a characteriza-
tion of the right derived functors of a left exact, additive, covariant functor
from C(R) to C(R ), where R is a commutative ring.
1.3.5 Theorem. Let R be a commutative ring, and let T be a left exact,
additive, covariant functor from C(R) to C(R ). Let (T i )i∈N0 be a negative
strongly connected sequence of covariant functors from C(R) to C(R ) such

=
that there exists a natural equivalence ψ : T 0 −→ T and such that T i (I) = 0
for all i ∈ N and all injective R-modules I.
Then there is a unique isomorphism of connected sequences

=
Ψ = (ψ i )i∈N0 : (T i )i∈N0 −→ (Ri T )i∈N0
(of functors from C(R) to C(R )) such that ψ 0 = ψ. (Of course, we are em-
ploying Convention 1.3.2.)
The next exercise strengthens Exercise 1.2.7.
1.3.6 Exercise. Let S be a multiplicatively closed subset of R. Show that
 −1 i   i 
S (Ha ( • )) i∈N
0
and HaS −1 R (S −1 ( • )) i∈N 0

are isomorphic connected sequences of functors (from C(R) to C(S −1 R)).


1.3 Connected sequences of functors 13

1.3.7 Remarks. Let B = (bα )α∈Λ be an inverse family of ideals of R over


Λ, as in 1.2.10.
Let us temporarily write U i := lim ExtiR (R/bα , • ) for i ∈ N0 . These
−→
α∈Λ
functors were introduced in 1.2.10. We are going to show now how they fit
together into a negative strongly connected sequence of functors (from C(R)
to itself).
First of all, whenever 0 −→ L −→ M −→ N −→ 0 is an exact sequence
of R-modules and R-homomorphisms, there are induced, for each α ∈ Λ,
connecting homomorphisms

ExtiR (R/bα , N ) −→ Exti+1


R (R/bα , L) (i ∈ N0 )

which make the induced long sequence

0 - HomR (R/bα , L) - HomR (R/bα , M ) - HomR (R/bα , N )


- Ext1 (R/bα , L) - Ext1 (R/bα , M ) - Ext1 (R/bα , N )
R R R

- ··· ···
- Exti (R/bα , L) - Exti (R/bα , M ) - Exti (R/bα , N )
R R R

- Exti+1 (R/bα , L) - ···


R

exact. Moreover, these connecting homomorphisms are such that, for α, β ∈ Λ


with α ≥ β, the diagram

ExtiR (R/bβ , N ) - Exti+1 (R/bβ , L)


R

ExtiR (hα
β ,N ) Exti+1 α
R (hβ ,L)

? ?
ExtiR (R/bα , N ) - Exti+1 (R/bα , L)
R

(in which the horizontal maps are the appropriate connecting homomorphisms
and hαβ : R/bα → R/bβ is the natural homomorphism) commutes for each
i ∈ N0 . It follows that these diagrams induce ‘connecting’ R-homomorphisms

U i (N ) = lim ExtiR (R/bα , N ) −→ U i+1 (L) = lim Exti+1


R (R/bα , L)
−→ −→
α∈Λ α∈Λ

(for i ∈ N0 ); moreover, the fact that passage to direct limits preserves exactness
14 The local cohomology functors

ensures that the resulting long sequence

0 - U 0 (L) - U 0 (M ) - U 0 (N )

- U 1 (L) - U 1 (M ) - U 1 (N )
- ··· ···
- U i (L) - U i (M ) - U i (N )
- U i+1 (L) - ···

is exact. Next, standard properties of the extension functors ensure that, when-
ever

0 - L - M - N - 0

λ μ ν

? ? ?
0 - L - M - N - 0

is a commutative diagram of R-modules and R-homomorphisms with exact


rows, then, for all α ∈ Λ, the diagram

ExtiR (R/bα , N ) - Exti+1 (R/bα , L)


R

ExtiR (R/bα ,ν) Exti+1


R (R/bα ,λ)

? ?
ExtiR (R/bα , N  ) - Exti+1 (R/bα , L )
R

(in which the horizontal maps are the appropriate connecting homomorphisms)
commutes for each i ∈ N0 . It therefore follows that the diagram

lim ExtiR (R/bα , N ) - lim Exti+1 (R/bα , L)


−→ −→ R
α∈Λ α∈Λ

lim ExtiR (R/bα ,ν) lim Exti+1


R (R/bα ,λ)
−→ −→
α∈Λ α∈Λ
? ?
lim Ext i
(R/b α , N ) - lim Exti+1 (R/bα , L )
−→ R −→ R
α∈Λ α∈Λ

(in which the horizontal maps are again the appropriate connecting homomor-
phisms) commutes for all i ∈ N0 .
1.3 Connected sequences of functors 15
 
We have thus made lim ExtiR (R/bα , • ) into a negative strongly
−→
α∈Λ i∈N0
connected sequence of covariant functors from C(R) to C(R). Since we have
lim ExtiR (R/bα , I) = 0 for all i ∈ N whenever I is an injective R-module,
−→
α∈Λ
it now follows from 1.3.5 that there is a unique isomorphism of connected
sequences
 

 = ψ
Ψ i i
: lim ExtR (R/bα , • )
=
−→ Ri ΓB
i∈N0 −→ i∈N0
α∈Λ i∈N0

for which ψ0 is the natural equivalence φB of 1.2.11(ii); furthermore, both
these connected sequences are isomorphic to the negative (strongly) connected
sequence of functors formed by the right derived functors of
lim HomR (R/bα , • ).
−→
α∈Λ

A special case of 1.3.7 describes local cohomology modules as direct limits


of Ext modules. As this description is of crucial importance for our subject,
we state it separately.
1.3.8 Theorem. There is a unique isomorphism of connected sequences (of
functors from C(R) to C(R))
 

=
Φa = φia : lim ExtiR (R/an , • ) −→ Hai
i∈N0 −→ i∈N0
n∈N i∈N0

=
which extends the natural equivalence φ0a : lim HomR (R/an , • ) −→ Γa of
−→
n∈N
1.2.11(iii). Consequently, for each R-module M and each i ∈ N0 ,
Hai (M ) ∼
= lim ExtiR (R/an , M ).
−→
n∈N

1.3.9 Exercise. Let M be an R-module, not necessarily finitely generated.


Let a1 , . . . , an be an M -sequence (of elements of R) (see [50, p. 123]). Let
a1 ∈ R. Show that
(i) if a1 , a2 , . . . , an is also an M -sequence, then so too is a1 a1 , a2 , . . . , an ;
(ii) if h1 , . . . , hn ∈ N, then ah1 1 , . . . , ahnn is also an M -sequence; and
(iii) if a1 , . . . , an all belong to a, then ExtiR (R/a, M ) = Hai (M ) = 0 for all
i = 0, . . . , n − 1. (This theme will be pursued in Chapter 6.)
2

Torsion modules and ideal transforms

The first section of this chapter contains the essence of a useful reduction tech-
nique in the study of local cohomology modules of finitely generated modules.
The main points are these: if M is a finitely generated R-module, then it turns
out that the R-module M := M/Γa (M ) is a-torsion-free, and that a contains
a non-zerodivisor r on M ; moreover, for i > 0, the local cohomology modules
Hai (M ) and Hai (M ) are isomorphic, so that the study of these higher local co-
homology modules of M with respect to a can be reduced to the study of the
corresponding local cohomology modules of M ; the advantage of this is that
r
the exact sequence 0 −→ M −→ M −→ M /rM −→ 0 provides a route to
further progress. There are several places later in the book where this strategy
is used.
In the second section of this chapter, we develop the basic theory of the
functor Da := lim HomR (an , • ) which was mentioned in 1.2.9(ii). For an R-
−→
n∈N
module M , the module Da (M ) is called the a-transform of M , and we plan to
show that such transforms provide a powerful algebraic tool. The use of these
ideal transforms is an important part of our approach to local cohomology; we
show that they have a certain universal property, and that universal property
will help us with many technical details later in the book. In §2.2, we shall
actually develop the theory of the generalized ideal transform functor with re-
spect to what we call a system of ideals of R; one example of such a system
of ideals is the family (an )n∈N of powers of a, and the generalized ideal trans-
form functor with respect to this system is just the a-transform functor. Our
motivation for working in this generality is our wish to apply generalized ideal
transforms to the theory of S2 -ifications in Chapter 12.
Towards the end of the chapter, we show that, in certain cases, ideal trans-
forms have geometrical significance: we describe the ring of regular functions
on a non-empty open subset of an affine variety V over an algebraically closed
2.1 Torsion modules 17

field as an ideal transform of O(V ), the ring of regular functions on V . Actu-


ally, this is only a brief foretaste of what is to come at the end of the book, in
Chapter 20, where we explore the relationship between ideal transforms and
sheaf cohomology.

2.1 Torsion modules


We begin with results concerning a-torsion-free modules and a-torsion mod-
ules. Part (i) of our first lemma is related to Exercise 1.3.9(iv).

2.1.1 Lemma. Let M be an R-module.

(i) If a contains a non-zerodivisor on M , then M is a-torsion-free, that is,


Γa (M ) = 0.
(ii) Assume now that M is finitely generated. Then M is a-torsion-free if and
only if a contains a non-zerodivisor on M .

Proof. (i) Let r ∈ a be a non-zerodivisor on M , and let m ∈ Γa (M ). This


means that there exists n ∈ N with an m = 0. Thus rn m = 0, from which we
deduce that m = 0.
(ii) One implication follows from (i), and so we assume that a consists en-

tirely of zerodivisors on M . Then a ⊆ p∈Ass M p by [81, Corollary 9.36],
and, since M is finitely generated, Ass M is finite. Hence, by the Prime Avoid-
ance Theorem [81, 3.61], a ⊆ p for some p ∈ Ass M . Since M has a sub-
module whose annihilator is exactly p, it follows that (0 :M a) = 0, so that
Γa (M ) = 0. This completes the proof.

2.1.2 Lemma. For an R-module M , the module M/Γa (M ) is a-torsion-free.

Proof. Let m ∈ M be such that the element m + Γa (M ) of M/Γa (M ) is


annihilated by an , where n ∈ N. Our aim is to show that m + Γa (M ) = 0,
that is, that m ∈ Γa (M ).
Now an m ⊆ Γa (M ). Since an m is a finitely generated submodule of
Γa (M ), and each element of an m is annihilated by some power of a, it follows
that there exists t ∈ N such that at an m = 0. Therefore m ∈ (0 :M an+t ) ⊆
Γa (M ).

2.1.3 Remarks. The following points should be noted.

(i) If M is an a-torsion R-module, that is, if M = Γa (M ), then all sub-


modules of M and all R-homomorphic images of M are also a-torsion.
18 Torsion modules and ideal transforms

(ii) Consequently, for each R-module L and each i ∈ N0 , the i-th local
cohomology module Hai (L) is an a-torsion R-module. To see this, let
I • : 0 −→ I 0 −→ I 1 −→ · · · −→ I i −→ I i+1 −→ · · ·
be an injective resolution of L; use this in conjunction with 1.2.2(i) to
see that Hai (L) is a homomorphic image of a submodule of the a-torsion
module Γa (I i ); and then appeal to (i).
Our next aim is to show that, if M is an a-torsion R-module, then Hai (M ) =
0 for all i ∈ N. We approach this by first considering the effect of Γa on an
injective R-module.
2.1.4 Proposition. Let I be an injective R-module. Then Γa (I) is also an
injective R-module.
Proof. Let b be an ideal of R, and let h : b −→ Γa (I) be a homomorphism
of R-modules. By the Baer Criterion (see [71, Theorem 3.20]), it is enough for
us to show that there exists m ∈ Γa (I) such that h(r) = rm for all r ∈ b.
Since I is injective, there exists w ∈ I such that h(r) = rw for all r ∈ b.
Since R is Noetherian, h(b) is a finitely generated submodule of Γa (I), and
so there exists t ∈ N such that at h(b) = 0. Now h(b) is a submodule of
the finitely generated R-module Rw, and so, by the Artin–Rees Lemma [50,
Theorem 8.5], there exists c ∈ N such that, for all integers n ≥ c,
an (Rw) ∩ h(b) = an−c (ac (Rw) ∩ h(b)).
Hence at+c (Rw) ∩ h(b) ⊆ at h(b) = 0. Consequently, we can define an R-
homomorphism h̃ : at+c + b −→ Γa (I) for which h̃(s + r) = rw for all s ∈
at+c and r ∈ b: this follows because, if r1 , r2 ∈ b are such that r1 − r2 ∈ at+c ,
then r1 w − r2 w = (r1 − r2 )w ∈ at+c (Rw) ∩ h(b) = 0.
Now use again the fact that I is R-injective to see that there exists m ∈ I
such that h̃(r) = rm for all r ∈ at+c + b. Since h̃ extends h, the proof will be
complete if we show that m ∈ Γa (I).
To achieve this, just note that, for all s ∈ at+c ,
sm = h̃(s) = h̃(s + 0) = 0w = 0,
by definition of h̃. This completes the proof.
2.1.5 Corollary. Let I be an injective R-module. Then the canonical exact
sequence 0 −→ Γa (I) −→ I −→ I/Γa (I) −→ 0 splits.
Proof. This is immediate from the fact, established in 2.1.4, that Γa (I) is
injective.
2.1 Torsion modules 19

2.1.6 Corollary. Let M be an a-torsion R-module. Then there exists an in-


jective resolution of M in which each term is an a-torsion R-module.
Proof. First note that, if N is an arbitrary R-module, then there exists an
injective R-module I and an R-monomorphism h : N → I. Application of
the left exact functor Γa yields a monomorphism Γa (h) : Γa (N ) → Γa (I),
and Γa (I) is injective by 2.1.4.
If we apply the above paragraph to the a-torsion R-module M , we see that
M can be embedded in an a-torsion injective R-module I 0 . Suppose, induc-
tively, that n ∈ N0 and we have constructed an exact sequence
n−1

0 - M - I0 - ··· - I n−1 d - I n

of R-modules and R-homomorphisms in which I 0 , . . . , I n−1 , I n are all a-


torsion injective R-modules. Let C := Coker dn−1 , and note that, by 2.1.3(i),
C is an a-torsion module because I n is. Apply the first paragraph of this proof
to C to deduce that there is an a-torsion injective R-module I n+1 and an R-
monomorphism g : C → I n+1 . Let dn : I n → I n+1 be the composition of
the natural epimorphism from I n to C and g.
This completes the inductive step, and the proof.
2.1.7 Corollary. Let M and N be R-modules such that M is a-torsion. Then
(i) Hai (M ) = 0 for all i > 0;
(ii) Hai (Γa (N )) = 0 for all i > 0; and
(iii) the natural map π : N → N/Γa (N ) induces isomorphisms Hai (π) :

=
Hai (N ) −→ Hai (N/Γa (N )) for all i > 0.
Proof. (i) It was remarked in 1.2.2(i) that we can use any injective resolution
of M to calculate (up to isomorphism) the Hai (M ): by 2.1.6, there is an injec-
tive resolution of M in which each term is an a-torsion R-module, and use of
this shows that Hai (M ) = 0 for all i > 0.
(ii) This is immediate from (i) because Γa (N ) is an a-torsion R-module.
(iii) This is immediate from (ii) on use of the long exact sequence of local
cohomology modules induced by the short exact sequence
π
0 −→ Γa (N ) −→ N −→ N/Γa (N ) −→ 0.
2.1.8 Exercise. Let M be an a-torsion R-module, and let
d−1 d0 di
I • : 0 −→ I 0 −→ I 1 −→ · · · −→ I i −→ I i+1 −→ · · ·
be an injective resolution of M . Show that Γa (I • ), that is, the complex
Γa (d−1 ) Γa (di )
0 - Γa (I 0 ) - ··· - Γa (I i ) - Γa (I i+1 ) -· · · ,
20 Torsion modules and ideal transforms

is also an injective resolution of M .

2.1.9 Exercise. Let b be a second ideal of R, and let M be a b-torsion R-


i
module. Prove that Ha+b (M ) ∼
= Hai (M ) for all i ∈ N0 .
In Chapter 1, we indicated that we shall, at times, find it convenient to con-
sider an inverse family B = (bα )α∈Λ of ideals of R over a (non-empty) di-
rected partially ordered set Λ, as in 1.2.10. We were able to produce, for such
a B, a functor ΓB in 1.2.11, and the right derived functors Ri ΓB (i ∈ N0 ) of
ΓB are generalizations of the local cohomology functors Hai (i ∈ N0 ).
However, we cannot expect the theory of ΓB to imitate local cohomology
theory completely unless we impose additional conditions on B. For example,
the analogue for ΓB of 2.1.4 is not true in general. However, this difficulty
does not occur for the concept introduced in the next definition.

2.1.10 Definition. Let (Λ, ≤) be a (non-empty) directed partially ordered set.


A system of ideals (of R) over Λ is an inverse family B = (bα )α∈Λ of ideals
of R over Λ in the sense of 1.2.10 with the additional property that, for all
α, γ ∈ Λ, there exists δ ∈ Λ such that bδ ⊆ bα bγ . (It is clear that the δ in this
condition can be chosen so that δ ≥ α and δ ≥ γ, since (Λ, ≤) is a directed
set and, by 1.2.10, whenever (μ, ν) ∈ Λ × Λ with μ ≥ ν, we have bμ ⊆ bν .)
For such a system of ideals B, we shall denote Ri ΓB by HB i
(for all i ∈
N0 ). The reader should keep in mind that (a )n∈N is a fundamental example of
n

a system A of ideals (over N); however, we shall continue to use the notation
Hai of 1.2.1 (rather than HA
i
).

2.1.11 Examples. Here are some further examples of systems of ideals.

(i) Let A be a non-empty set of ideals of R. Then the set of all products
of finite families of ideals taken from A forms a system of ideals in an
obvious way.
In particular, if A is a non-empty multiplicatively closed set of ideals
of R, then A itself forms a system of ideals.
(ii) Let n ∈ N0 . The height, ht b, of a proper ideal b of R is defined in [50,
p. 31] and [81, 15.6]. Interpret the height of the improper ideal R of R as
∞. Then {b : b is an ideal of R and ht b ≥ n} forms a system of ideals
of R.

2.1.12 Exercise. Let B be a system of ideals over Λ in the sense of 2.1.10.


Let α, β ∈ Λ with α ≥ β. Show that, for each R-module M ,

Γbβ (M ) ⊆ Γbα (M ) ⊆ ΓB (M ) = Γbδ (M ).
δ∈Λ
2.2 Ideal transforms and generalized ideal transforms 21

2.1.13 Exercise. Let B be a system of ideals over Λ in the sense of 2.1.10.


Let M be an R-module. We shall say that M is B-torsion-free precisely when
ΓB (M ) = 0, and that M is B-torsion precisely when M = ΓB (M ).
(i) Show that M/ΓB (M ) is B-torsion-free.
(ii) Show that, if I is an injective R-module, then ΓB (I) is also an injective
R-module. Deduce that, if N is a B-torsion R-module, then there exists
an injective resolution of N in which each term is a B-torsion R-module,
i
and conclude that HB (N ) = 0 for all i > 0.
(iii) Prove that the natural epimorphism π : M → M/ΓB (M ) induces iso-

=
morphisms HB i
(π) : HBi
(M ) −→ HB i
(M/ΓB (M )) for all i > 0.
2.1.14 Exercise. Let M be an R-module. Show that the sets Ass(Γa (M ))
and Ass(M/Γa (M )) are disjoint, and that
Ass M = Ass(Γa (M )) ∪ Ass(M/Γa (M )).

2.2 Ideal transforms and generalized ideal transforms


The principal object of study in this section will be the ideal transform of an
R-module M with respect to a. This is defined as follows.
2.2.1 Definitions. In 1.2.9(ii), we constructed covariant, R-linear functors
Da := lim HomR (an , • ) and lim ExtiR (an , • ) (i ∈ N0 )
−→ −→
n∈N n∈N

from C(R) to itself. We shall refer to Da as the a-transform functor; note that,
by 1.2.8, this functor is left exact.
For an R-module M , we call Da (M ) = lim HomR (an , M ) the ideal trans-
−→
n∈N
form of M with respect to a, or the a-transform of M .
However, instead of working with the powers of a single ideal a, we are
going to work in this section in the more general framework of a system of
ideals. This generality is motivated by an application that will be presented in
Chapter 12. The reader should keep in mind, throughout this section, that a
basic example of a system of ideals is (an )n∈N .
2.2.2 Notation. Throughout this section, (Λ, ≤) will denote a (non-empty)
directed partially ordered set, and B = (bα )α∈Λ will denote a system of ideals
of R over Λ in the sense of 2.1.10.
The functors ΓB and lim HomR (R/bα , • ) were introduced in 1.2.11(i) and
−→
α∈Λ
22 Torsion modules and ideal transforms

1.2.10; for i ∈ N0 , the functor lim ExtiR (R/bα , • ) was discussed in 1.3.7,
−→
α∈Λ
where it was shown to be naturally equivalent to Ri ΓB , the i-th right derived
functor of ΓB ; in 2.1.10, we agreed to denote the R-linear functor Ri ΓB by
i
HB ; we shall refer to this as the i-th generalized local cohomology functor
with respect to B (and we shall use natural extensions of this terminology).
We denote by
 
 i  ∼
=
i
ΦB = φB i∈N0 : lim ExtR (R/bα , • ) −→ HB i
−→ i∈N0
α∈Λ i∈N0

the unique isomorphism of connected sequences for which φ0B is the natural
equivalence φB of 1.2.11(ii): see 1.3.7.

2.2.3 Definitions. The functors

DB := lim HomR (bα , • ) and lim ExtiR (bα , • ) (i ∈ N0 )


−→ −→
α∈Λ α∈Λ

(from C(R) to itself) can be defined using the ideas of 1.2.8 in conjunction
with the inclusion maps bα → bβ (for α, β ∈ Λ with α ≥ β). We shall refer to
DB as the B-transform functor; note that, by 1.2.8, this functor is left exact.
For an R-module M , we call DB (M ) = lim HomR (bα , M ) the general-
−→
α∈Λ
ized ideal transform of M with respect to B, or, alternatively, the B-transform
of M .

2.2.4 Exercise. For i ∈ N0 , we use Ri DB to denote the i-th right derived


functor of DB .  
Modify the ideas of 1.3.7 to show that lim ExtiR (bα , • ) is a nega-
−→
α∈Λ i∈N0
tive strongly connected sequence of functors from C(R) to itself. Use 1.3.5 to
show that there is a unique isomorphism of connected sequences (of functors
from C(R) to itself)
 

=
i
Ψ B = ψB : R i DB −→ lim ExtiR (bα , • )
i∈N0 i∈N0 −→
α∈Λ i∈N0

which extends the identity natural equivalence from DB to itself.

2.2.5 Exercise. Let M be an R-module. For each α ∈ Λ, let


παM : HomR (bα , M ) → DB (M )

be the natural homomorphism.


2.2 Ideal transforms and generalized ideal transforms 23

(i) Let α, β ∈ Λ, and let f ∈ HomR (bα , R) and g ∈ HomR (bβ , R). Since
B is a system of ideals of R, there exists δ ∈ Λ such that bδ ⊆ bα bβ .
Observe that f bδ , the restriction of f to bδ , maps bδ into bβ . Show that
g ◦ (f bδ ) = f ◦ (g bδ ).
(ii) Show that there is a binary operation ∗ on DB (R) which is such that, for
f ∈ HomR (bα , R) and g ∈ HomR (bβ , R),

παR (f ) ∗ πβR (g) = πδR (g ◦ (f bδ ))

for any choice of δ ∈ Λ with bδ ⊆ bα bβ ; show further that DB (R) is


a commutative ring with identity with respect to its R-module addition
and ∗ as multiplication.
(iii) Show that DB (M ) has the structure of a DB (R)-module such that, for
α, β ∈ Λ and for f ∈ HomR (bα , R) and h ∈ HomR (bβ , M ),
 
παR (f ) πβM (h) = πδM (h ◦ (f bδ ))

for any choice of δ ∈ Λ with bδ ⊆ bα bβ .


(iv) Show that DB is an additive, left exact, covariant functor from C(R) to
C(DB (R)). Thus all the Ri DB (i ∈ N0 ) can be considered as additive
functors from C(R) to C(DB (R)).

2.2.6 Theorem. Denote the identity functor on the category C(R) by Id.

(i) There are natural transformations of functors (from C(R) to itself)

ξ (= ξB ) : ΓB −→ Id, η (= ηB ) : Id −→ DB

ζ 0 (= ζB
0
) : DB −→ HB
1

such that, for each R-module M ,

(a) ξM : ΓB (M ) −→ M is the inclusion map,


(b) for each g ∈ M , ηM (g) is the natural image in DB (M ) of the
homomorphism fα,g ∈ HomR (bα , M ) given by fα,g (r) = rg for
all r ∈ bα (for any α ∈ Λ), and
(c) the sequence
ξM ηM ζ0
0 −→ ΓB (M ) −→ M −→ DB (M ) −→
M 1
HB (M ) −→ 0

is exact.
(ii) Let i ∈ N, and M be an R-module. For each α ∈ Λ, the connecting
24 Torsion modules and ideal transforms
i
homomorphism βα,M : ExtiR (bα , M ) −→ Exti+1R (R/bα , M ) is an iso-
morphism, and passage to the direct limit yields an R-isomorphism

=
i
βM : lim ExtiR (bα , M ) −→ lim Exti+1
R (R/bα , M ).
−→ −→
α∈Λ α∈Λ

=
i
Define γM : Ri DB (M ) −→ HB
i+1
(M ) by

B,M ◦ βM ◦ ψB,M ,
:= φi+1
i i i
γM

where φi+1 i
B and ψB are the natural equivalences of 2.2.2 and 2.2.4 re-
spectively. Then, as M varies through the category C(R), the γM
i
con-

=
stitute a natural equivalence of functors γ : R DB −→ HB .
i i i+1

(iii) For each i ∈ N, set ζ i (= ζB


i
) := (−1)i γ i . Then
 j   j+1
ζ j∈N : Rj DB j∈N −→ HB
0 0 j∈N0

is the unique homomorphism of connected sequences which extends the


natural transformation ζ 0 : DB −→ HB
1
of part (i).
Note. When B = (an )n∈N , we shall write ξa , ηa , ζa0 instead of ξB , ηB , ζB
0
.
Proof. (i), (ii) Let α, δ ∈ Λ with α ≥ δ; let jδα : bα −→ bδ be the inclusion
δ : R/bα −→ R/bδ be the natural epimorphism. Also, let M, N
map; and let hα
be R-modules and let f : M −→ N be an R-homomorphism.
The commutative diagram

0 - bα - R - R/bα - 0

jδα hα
δ

? ?
0 - bδ - R - R/bδ - 0

(in which the rows are the canonical exact sequences) induces a chain map of
the long exact sequence of Ext•R ( • , M ) modules induced by the top row to
that induced by the bottom row. Since R is a projective R-module, and since
HomR (R, M ) is naturally isomorphic to M , we therefore obtain a commuta-
tive diagram

0 - HomR (R/bδ , M ) - M - HomR (bδ , M ) - Ext1R (R/bδ , M ) - 0

? ? ?
0 - HomR (R/bα , M ) - M - HomR (bα , M ) - Ext1R (R/bα , M ) - 0
2.2 Ideal transforms and generalized ideal transforms 25

(in which the rows are exact), and, for each i ∈ N, a commutative diagram
i
βδ,M
ExtiR (bδ , M ) ∼
- Exti+1 (R/bδ , M )
= R

ExtiR (jδα ,M ) Exti+1 α


R (hδ ,M )

? i ?
βα,M
ExtiR (bα , M ) - Exti+1 .

= R (R/bα , M )

Now pass to the direct limits, bearing in mind the exactness-preserving proper-
ties of this process, and use the natural equivalences φ0B , φ1B of 2.2.2 to obtain
an exact sequence of R-modules and R-homomorphisms
ξM ηM ζ0
0 −→ ΓB (M ) −→ M −→ DB (M ) −→
M 1
HB (M ) −→ 0

(where ξM and ηM are as described in (a) and (b) of the statement of the
theorem) and, for each i ∈ N, an isomorphism

=
i
βM : lim ExtiR (bα , M ) −→ lim Exti+1
R (R/bα , M ).
−→ −→
α∈Λ α∈Λ

Moreover, since the diagram

0 - HomR (R/bα , M ) - M - HomR (bα , M ) - Ext1R (R/bα , M ) - 0

? ? ? ?
0 - Hom R (R/b α , N ) - N - Hom (b
R α , N ) - ExtR (R/bα , N ) - 0
1

and, for each i ∈ N, the diagram


i
βα,M
ExtiR (bα , M ) - Exti+1

= R (R/bα , M )

? i ?
βα,N
ExtiR (bα , N ) ∼
- Exti+1 (R/bα , N )
= R

(in all of which, all the unmarked vertical maps are induced by f : M −→
N ), are all commutative, it follows that, as M varies through C(R), the ξM ,
0 i
ηM , ζM and βM constitute natural transformations of functors ξ, η, ζ 0 and β i
respectively. This completes the proof of parts (i) and (ii).
 
(iii) Of course Rj DB j∈N is a negative strongly connected sequence of
0
26 Torsion modules and ideal transforms
j+1
covariant functors from C(R) to itself; also, HB is a negative con-
j∈N0
nected sequence of covariant functors from C(R) to itself. It is immediate
from 1.3.4(i) that there is a unique homomorphism of the first of these con-
nected sequences to the second which extends the natural transformation ζ 0 :
DB −→
1
j j
 B of part (i). The fact that this unique homomorphism is actually
H
(−1) γ j∈N0 follows from Rotman [71, Theorem 11.24], which shows that,
for each j ∈ N0 , each exact sequence 0 −→ L −→ M −→ N −→ 0 of
R-modules and R-homomorphisms and each α ∈ Λ, the diagram
ExtjR (bα , N ) - Extj+1 (bα , L)
R

? ?
Extj+1 - Extj+2 (R/bα , L) ,
R (R/bα , N ) R

in which all the homomorphisms are the obvious connecting homomorphisms,


is anticommutative.

2.2.7 Remark. Let the situation be as in 2.2.6. It is immediate from the exact
sequence
ξM ηM ζ0
0 −→ ΓB (M ) −→ M −→ DB (M ) −→
M 1
HB (M ) −→ 0

of 2.2.6(i)(c) that ηM : M −→ DB (M ) is an isomorphism if and only if


1
ΓB (M ) = HB (M ) = 0.

2.2.8 Corollary. Let M be an R-module, not necessarily finitely generated.


Assume that a contains an M -sequence of length 2. Then

ηM : M −→ Da (M )

is an isomorphism.

Proof. By Exercise 1.3.9(iv), we have Γa (M ) = Ha1 (M ) = 0, and so the


claim is immediate from 2.2.7.

2.2.9 Remark. It follows from 2.2.6 that, for each R-module M , there is an
R-monomorphism θM : M/ΓB (M ) −→ DB (M ), induced by ηM , such that
the sequence
θ ζ0
0 −→ M/ΓB (M ) −→
M
DB (M ) −→
M 1
HB (M ) −→ 0

is exact. As θM is induced by ηM , a precise formula for it can be extracted


from the statement of 2.2.6. Note also that, as M varies through C(R), the θM
2.2 Ideal transforms and generalized ideal transforms 27

constitute a natural transformation of functors since, whenever f : M −→ N


is a homomorphism of R-modules, the diagram
θM
M/ΓB (M ) - DB (M )

f∗ DB (f )

? θN
?
N/ΓB (N ) - DB (N )

(in which f ∗ denotes the homomorphism induced by f ) commutes.


This remark can be particularly helpful when used in conjunction with the
fact (see 2.1.13(iii)) that the canonical epimorphism π : M → M/ΓB (M )

=
induces isomorphisms HB i i
(π) : HB (M ) −→ HB i
(M/ΓB (M )) for all i > 0.
We use this idea in the proof of 2.2.10 below.
The exact sequence
ξM ηM ζ0
0 −→ ΓB (M ) −→ M −→ DB (M ) −→
M 1
HB (M ) −→ 0
of Theorem 2.2.6(i)(c), and the particular case
ξM ηM ζ0
0 −→ Γa (M ) −→ M −→ Da (M ) −→
M
Ha1 (M ) −→ 0
of it, are fundamental. In the next corollary, we present some important appli-
cations of such sequences.
2.2.10 Corollary. Let M be an R-module; we use the notation of 2.2.6 and
2.2.9. Let π : M −→ M/ΓB (M ) be the canonical epimorphism. Then the
following hold:
(i) DB (ΓB (M )) = 0;
(ii) DB (π) : DB (M ) −→ DB (M/ΓB (M )) is an isomorphism;
(iii) DB (ηM ) = ηDB (M ) : DB (M ) −→ DB (DB (M )) is an isomorphism;
1
(iv) ΓB (DB (M )) = 0 = HB (DB (M ));
(v) HB (ηM ) : HB (M ) −→ HB
i i i
(DB (M )) is an isomorphism for every
i > 1.
Proof. (i) Since ΓB (M ) is a B-torsion R-module, it is enough, in order to
prove this part, to show that, if N is a B-torsion R-module, then DB (N ) = 0.
Now, for such an N , we have HB 1
(N ) = 0 by 2.1.13(ii), and ξN : ΓB (N ) →
N is the identity map. Hence the exact sequence of 2.2.6(i)(c) for N reduces
to

=
0 −→ ΓB (N ) −→ N −→ DB (N ) −→ 0,
28 Torsion modules and ideal transforms

and so DB (N ) = 0, as required.
(ii) By 2.2.3, the functor DB is left exact. Therefore the canonical exact
π
sequence 0 −→ ΓB (M ) −→ M −→ M/ΓB (M ) −→ 0 induces an exact
sequence

0 - DB (ΓB (M ))- DB (M ) - DB (M/ΓB (M )) - R1 DB (ΓB (M )).


DB (π)

Now DB (ΓB (M )) = 0 by (i), while R1 DB (ΓB (M )) ∼ = HB 2


(ΓB (M )) by
2
2.2.6(ii). Since HB (ΓB (M )) = 0 by 2.1.13(ii), it follows that DB (π) is an
isomorphism.
(iii) It is left as an interesting exercise on direct limits for the reader to show
that DB (ηM ) = ηDB (M ) . We show that DB (ηM ) is an isomorphism.
Since ηM = θM ◦ π (where θM is as defined in 2.2.9) and we have al-
ready shown in (ii) that DB (π) is an isomorphism, it is enough for us to show
that DB (θM ) is an isomorphism. Since DB is left exact (by 2.2.3), the exact
sequence
θ ζ0
0 −→ M/ΓB (M ) −→
M
DB (M ) −→
M 1
HB (M ) −→ 0

of 2.2.9 yields a further exact sequence


0
DB (θM )
0 - DB (M/ΓB (M )) - DB (DB (M )) - DB (H 1 (M )).
DB (ζM )
B

1 1
But HB (M ) is B-torsion, and so DB (HB (M )) = 0 by (i). Hence DB (θM )
is an isomorphism, as required.
(iv) This is now immediate from (iii) and 2.2.7.
(v) We again use the fact that ηM = θM ◦ π. We already know, from
i
2.1.13(iii), that HB (π) is an isomorphism for all i ∈ N. It is therefore enough,
i
in order to complete the proof, to show that HB (θM ) is an isomorphism for all
i > 1.
However, for each i > 1, the exact sequence
θ ζ0
0 −→ M/ΓB (M ) −→
M
DB (M ) −→
M 1
HB (M ) −→ 0

of 2.2.9 induces an exact sequence

··· - H i−1 (H 1 (M ))
B B

i i 0
- H i (M/ΓB (M )) HB
- H i (DB (M ))
(θM ) HB (ζM
- H i (H 1 (M ))
)
B B B B

- ··· ,
2.2 Ideal transforms and generalized ideal transforms 29
1
and since HB (M ) is a B-torsion R-module, it follows from 2.1.13(ii) that
i−1 1 i 1 i
HB (HB (M )) = HB (HB (M )) = 0; hence HB (θM ) is an isomorphism, as
required.
2.2.11 Exercise. Complete the proof of 2.2.10(iii). In other words, show
that, in the notation of 2.2.10, DB (ηM ) = ηDB (M ) .
2.2.12 Exercise. Recall that, by Exercise 2.2.5, the B-transform DB (R) of
R has the structure of a commutative ring. Show that ηR : R −→ DB (R) is a
ring homomorphism.
In order to exploit our results on the generalized ideal transform DB (M ),
we are going to obtain a description of DB (M ) in terms of objects which are
perhaps more familiar. This work will, of course, apply to the ideal transform
Da (M ). Towards the end of the section, we shall obtain a particularly simple
description of Da (M ) in the case in which a is principal. One approach to
these results uses the fact that, for an R-module M , the homomorphism ηM :
M −→ DB (M ) can be viewed as the solution to a universal problem. Our
next proposition, which uses ideas similar to ones used by K. Suominen in
[85, §1], provides the key to this approach. The influence of the ideas of P.
Gabriel [21] should also be acknowledged.
2.2.13 Proposition. (See R. Y. Sharp and M. Tousi [82, Lemma 1.3].) Let
e : M −→ M  be a homomorphism of R-modules such that Ker e and
Coker e are both B-torsion. Let ψ : M −→ K be a further homomorphism of
R-modules.
(i) The map DB (e) : DB (M ) −→ DB (M  ) is an isomorphism.
(ii) There is a unique R-homomorphism ψ  : M  −→ DB (K) such that the
diagram
e
M - M

ψ ψ
? ηK
?
K - DB (K)

commutes. In fact, ψ  = DB (ψ) ◦ DB (e)−1 ◦ ηM  .


(iii) If ψ and ηM  : M  −→ DB (M  ) are both isomorphisms, then the ho-
momorphism ψ  of part (ii) is also an isomorphism.
Proof. (i) We shall use the exact sequences
τ λ
0 −→ Ker e −→ M −→ Im e −→ 0
30 Torsion modules and ideal transforms

and
ρ
0 −→ Im e −→ M  −→ Coker e −→ 0,
σ

in which the maps are the obvious homomorphisms. Now e = ρ ◦ λ; it is there-


fore enough for us to show that DB (ρ) and DB (λ) are both isomorphisms.
The first of the above exact sequences induces an exact sequence

0 - DB (Ker e) - DB (M ) DB (λ)
DB (τ )
- DB (Im e) - R1 DB (Ker e).

However, R1 DB (Ker e) = ∼ H 2 (Ker e) by 2.2.6(ii). By hypothesis, Ker e is


B
2
B-torsion. Hence DB (Ker e) = HB (Ker e) = 0, by 2.2.10(i) and 2.1.13(ii).
Therefore DB (λ) is an isomorphism.
ρ σ
Next, from the exact sequence 0 −→ Im e −→ M  −→ Coker e −→ 0 we
obtain an induced exact sequence

0 - DB (Im e) DB (ρ)
- DB (M  ) DB (σ)
- DB (Coker e).

However, by hypothesis, Coker e is B-torsion, and so DB (Coker e) = 0, by


2.2.10(i). Hence DB (ρ) is an isomorphism.
(ii) For this part of the proof, it will be convenient for us to write K  :=
DB (K) and h := ηK : K −→ DB (K) = K  . Application of the natural
transformation η : Id −→ DB to the modules and homomorphisms in the
diagram
e
M - M

?
K
h
- K

yields a commutative diagram


ηM
M - DB (M )
@ DB (ψ)
@ D (e)
e@ =@
∼ B
ψ @
R @
R
M  ηM  - DB (M  )
? ηK ?
K - DB (K)
@ @ D (h)
h@ =@
∼ B

@
R ηK  @
R 
K ∼
- DB (K ) .
=
2.2 Ideal transforms and generalized ideal transforms 31

It should be noted that DB (e) is an isomorphism, by part (i), that DB (K) =


K  , and that, by 2.2.10(iii),
DB (h) = DB (ηK ) = ηDB (K) = ηK 
is also an isomorphism.
Thus, if there were an R-homomorphism ψ  : M  −→ K  such that ψ  ◦ e =
h◦ψ, then it would satisfy DB (ψ  )◦DB (e) = DB (h)◦DB (ψ) and we would
have to have (since η : Id −→ DB is a natural transformation)
ψ  = DB (h)−1 ◦ ηK  ◦ ψ  = DB (h)−1 ◦ DB (ψ  ) ◦ ηM 
= DB (ψ) ◦ DB (e)−1 ◦ ηM  .
On the other hand, one can easily verify by means of an elementary diagram
chase that DB (ψ) ◦ DB (e)−1 ◦ ηM  ◦ e = h ◦ ψ.
(iii) This is immediate from part (ii), since, if ψ is an isomorphism, then so
too is DB (ψ).
2.2.14 Remark. Let h : M −→ N be a homomorphism of R-modules.
Recall from 2.2.6(i) that ηB : Id −→ DB is a natural transformation of
functors and that Ker ηM and Coker ηM are both B-torsion. It therefore fol-
lows from 2.2.13 that DB (h) : DB (M ) −→ DB (N ) must be the unique
R-homomorphism from DB (M ) to DB (N ) which makes the diagram
ηM
M - DB (M )

? ηN
?
N - DB (N )

commute.
2.2.15 Corollary. Let e : M −→ M  be a homomorphism of R-modules
such that Ker e and Coker e are both B-torsion.
(i) The map DB (e) : DB (M ) −→ DB (M  ) is an isomorphism.
(ii) There is a unique R-homomorphism ψ  : M  → DB (M ) such that the
diagram
e
M - M

@
@
ηM
ψ
@
@R ?
DB (M )
32 Torsion modules and ideal transforms

commutes. In fact, ψ  = DB (e)−1 ◦ ηM  .


(iii) The map ψ  of (ii) is an isomorphism if and only if ηM  is an iso-
morphism, and, by 2.2.7, this is the case if and only if ΓB (M  ) =
HB1
(M  ) = 0.

Proof. Use 2.2.13 with IdM : M −→ M in the rôle of ψ : M −→ K.

We are now going to examine the special case of 2.2.15 in which M = R.


Recall from 2.2.5 that DB (R) is a commutative ring, and from 2.2.12 that
ηR : R −→ DB (R) is a ring homomorphism. We shall meet several exam-
ples of the situation of 2.2.15 in which e : R −→ R is a homomorphism of
commutative rings such that, when R is regarded as an R-module by means
of e, both Ker e and Coker e are B-torsion: 2.2.17 below shows that, in these
circumstances, the R-homomorphism ψ  : R −→ DB (R) given by 2.2.15 is
actually a ring homomorphism too.

2.2.16 Proposition. Let R be a ring (with identity, but not necessarily com-
mutative), and let e : R −→ R be a ring homomorphism such that Im e is con-
tained in the centre of R and, when R is regarded as a left R-module by means
of e, both Ker e and Coker e are B-torsion. Assume also that ΓB (R ) = 0.
Then the ring R is commutative.

Proof. By 2.2.15, there is a unique R-homomorphism ψ  : R −→ DB (R)


such that the diagram
e
R - R

@
ηR
@ ψ
@
R ?
@
DB (R)

commutes. Let r1 , r2 ∈ R . Since Coker e is B-torsion, there exist α, δ ∈ Λ


such that bα (respectively bδ ) annihilates the natural image in Coker e of r1
(respectively r2 ). Let b1 ∈ bα and b2 ∈ bδ ; then there exist r1 , r2 ∈ R such
that bi ri = e(ri ), that is e(bi )ri = e(ri ), for i = 1, 2. Therefore, since Im e is
contained in the centre of R , we have

e(b1 )e(b2 )r1 r2 = e(b1 )r1 e(b2 )r2 = e(r1 )e(r2 )
= e(r2 )e(r1 ) = e(b2 )r2 e(b1 )r1 = e(b1 )e(b2 )r2 r1 .

Therefore b1 b2 (r1 r2 −r2 r1 ) = 0. Hence the element r1 r2 −r2 r1 is annihilated
by bα bδ . But B is a system of ideals, so that there exists μ ∈ Λ such that
2.2 Ideal transforms and generalized ideal transforms 33

bμ ⊆ bα bδ . Since ΓB (R ) = 0, we deduce that r1 r2 − r2 r1 = 0 and R is


commutative.

2.2.17 Proposition. Let R be a commutative ring (with identity), and let


e : R −→ R be a ring homomorphism for which the R-modules Ker e and
Coker e are B-torsion.
Then the unique R-homomorphism ψ  : R −→ DB (R) such that the dia-
gram
e
R - R
@
ηR
@ ψ
@
R ?
@
DB (R)

commutes (the existence of which follows from 2.2.15) is a ring homomor-


phism, and therefore an R-algebra homomorphism.

Proof. Let r1 , r2 ∈ R . Since Coker e is B-torsion, there exist α, δ ∈ Λ


such that bα (respectively bδ ) annihilates the natural image in Coker e of r1
(respectively r2 ). Let b1 ∈ bα and b2 ∈ bδ ; then there exist r1 , r2 ∈ R such
that bi ri = e(ri ), that is e(bi )ri = e(ri ), for i = 1, 2.
Note also that, in the commutative ring DB (R), we have

b1 b2 ψ  (r1 )ψ  (r2 ) = ψ  (b1 r1 )ψ  (b2 r2 ) = ψ  (e(r1 ))ψ  (e(r2 ))


= ηR (r1 )ηR (r2 ) = ηR (r1 r2 ) = ψ  (e(r1 r2 ))
= ψ  (e(r1 )e(r2 )) = ψ  (b1 r1 b2 r2 ) = b1 b2 ψ  (r1 r2 ),

so that b1 b2 (ψ  (r1 )ψ  (r2 ) − ψ  (r1 r2 )) = 0. Hence the element

ψ  (r1 )ψ  (r2 ) − ψ  (r1 r2 ) ∈ DB (R)

is annihilated by bα bδ . But there exists μ ∈ Λ such that bμ ⊆ bα bδ . Since


DB (R) is B-torsion-free by 2.2.10(iv), we have ψ  (r1 r2 ) = ψ  (r1 )ψ  (r2 ).
Also ψ  (1R ) = ψ  (e(1R )) = ηR (1R ) = 1DB (R) . Therefore ψ  is a ring
homomorphism.

2.2.18 Corollary. Let M be an R-module, and let S be a multiplicatively


closed subset of R which consists entirely of non-zerodivisors on M , and which
is such that S ∩ bα = ∅ for all α ∈ Λ. Then there is a unique R-isomorphism


ψM : (M :S −1 M bα ) −→ DB (M )
α∈Λ
34 Torsion modules and ideal transforms

for which the diagram


⊆ 
M - α∈Λ (M :S −1 M bα )
@
@
ηM

ψM
@
@
R ?
DB (M )

commutes. In the special case in which M = R (and S consists of non-



zerodivisors on R), the map ψR is actually a ring isomorphism.

Note. Of course, since S consists entirely of non-zerodivisors on M , the


canonical R-homomorphism M → S −1 M is injective; we are using this to
identify M as an R-submodule of S −1 M .

Proof. Set M  := α∈Λ (M :S −1 M bα ), and let e : M −→ M  denote the
inclusion homomorphism. Since Coker e = ΓB (S −1 M/M ) is B-torsion, it is

immediate from 2.2.15 that there is a unique R-homomorphism ψM : M  −→
DB (M ) which makes the above diagram commute; it also follows that, in

order to show that ψM is an isomorphism, it is sufficient for us to show that
 1 
ΓB (M ) = HB (M ) = 0. This we do.
We show first that HBi
(S −1 M ) = 0 for each i ∈ N0 . Let y ∈ HB i
(S −1 M ).
Then there exists α ∈ Λ such that bα y = 0. By hypothesis, there exists s ∈
bα ∩ S, so that sy = 0. Since the functor HB i
is R-linear, multiplication by
i −1
s on HB (S M ) must provide an automorphism; therefore y = 0. Therefore
HBi
(S −1 M ) = 0, as claimed. Hence ΓB (M  ) = 0. It now follows from the
exact sequence

0 −→ M  −→ S −1 M −→ S −1 M/M  −→ 0
1
that HB (M  ) ∼
= ΓB (S −1 M/M  ). However,
S −1 M/M  ∼
= (S −1 M/M )/(M  /M ) = (S −1 M/M )/ΓB (S −1 M/M ),
1
and this is B-torsion-free by 2.1.13(i). Hence HB (M  ) = 0. The final claim
follows from 2.2.17.

Of course, all our work so far in this section applies to the particular system
of ideals (an )n∈N , and, indeed, that is a very important example of such a
system. As we have now presented, in this section, enough of the theory of
generalized ideal transforms for our later needs, we are going to concentrate,
for the remainder of this section, on the ordinary a-transform functor Da .
In the case when a is principal, there is a result similar to 2.2.18, but under
2.2 Ideal transforms and generalized ideal transforms 35

weaker hypotheses, which has important consequences for our work in Chapter
3. We present this result next. Recall that, for an R-module M and a ∈ R,
the notation Ma denotes the module of fractions of M with respect to the
multiplicatively closed subset ai : i ∈ N0 .

2.2.19 Theorem. Let a ∈ R. There is a natural equivalence of functors

ω  : DRa = lim HomR (Ran , • ) −→ ( • )a


−→
n∈N

(from C(R) to C(R)) such that, for an R-module M , and an f ∈ DRa (M )



represented by ft ∈ HomR (Rat , M ) (for some t ∈ N), we have ωM (f ) =
t t
ft (a )/a .

Proof. Let M be an R-module. It is immediate from 2.2.15 that there is a


unique R-isomorphism νM : Ma −→ DRa (M ) such that the diagram

M - Ma
@
ηRa,M
@ ∼
= νM
@
@R ?
DRa (M )

(in which the horizontal homomorphism is the natural one) commutes: note
that HRai
(Ma ) = 0 for all i ∈ N0 because multiplication by a provides an
 −1
automorphism on all these local cohomology modules. Define ωM := νM . It
is straightforward to use the commutativity of the above diagram to show that

ωM satisfies the formula given in the statement of the theorem. Furthermore,
it is easy to use that formula to show that, as M varies through the category

C(R), the ωM constitute a natural equivalence of functors.

2.2.20 Remark. With the notation of 2.2.19, consider, for an R-module M ,


the fundamental exact sequence of 2.2.6(i)(c) in the particular case in which
B = (an )n∈N and a = Ra: we have

ξM ηM ζ0
0 −→ ΓRa (M ) −→ M −→ DRa (M ) −→
M 1
HRa (M ) −→ 0.

Note that ωM ◦ ηM : M −→ Ma is just the canonical R-homomorphism τM .
 −1
Set σM := ζM 0
◦ (ωM ) ; then there is an exact sequence of R-modules and
R-homomorphisms
ξM τ σ
0 −→ ΓRa (M ) −→ M −→
M
Ma −→
M 1
HRa (M ) −→ 0
36 Torsion modules and ideal transforms

such that, as M varies through C(R), the τM and σM constitute natural trans-
formations of functors
τ : Id −→ ( • )a , σ : ( • )a −→ HRa
1
.
2.2.21 Corollary. Let M be an R-module and let a ∈ R.
(i) The kernel of the natural homomorphism τM : M → Ma is precisely
ΓRa (M ), and so, in view of 2.2.20, M/ΓRa (M ) can be identified as a
submodule of Ma . With this identification,
1
HRa (M ) ∼
= Ma /(M/ΓRa (M )).
(ii) For all i ∈ N with i > 1, we have HRa
i
(M ) = 0.
Proof. (i) This is immediate from 2.2.20.
(ii) Let i ∈ N with i > 1. Now HRa i
(M ) ∼
= Ri−1 DRa (M ) by 2.2.6(ii).
However, by 2.2.19, the functor DRa is naturally equivalent to ( • )a ; as the
latter functor is exact and i − 1 > 0, it follows that Ri−1 DRa (N ) = 0 for all
i
R-modules N . Hence HRa (M ) = 0.
The above corollary will prove useful in Chapter 3, as it provides a basis for
an argument which uses induction on the number of generators of a, and which
relies on the Mayer–Vietoris Sequence for local cohomology for the inductive
argument. This Mayer–Vietoris Sequence forms a major part of the subject
matter of Chapter 3.
2.2.22 Exercise. Use 2.2.18 to obtain an alternative proof of Corollary 2.2.8.
More precisely, let M be an R-module, not necessarily finitely generated,
and assume that a contains an M -sequence x, y of length 2. Use 2.2.18 (with
the choice S := {xi : i ∈ N0 }) and 1.3.9(ii) to show that the map ηM : M −→
Da (M ) of 2.2.6(i) is an isomorphism.
2.2.23 Exercise. Let b be a second ideal of R.

(i) Suppose that a ⊆ b. Show that there is a unique natural transformation
of functors (from C(R) to itself) αb,a : Db −→ Da such that, for each
R-module M , the diagram
ηb,M
M - Db (M )

@
@
ηa,M
αb,a,M
@
@R ?
Da (M )

commutes.
2.2 Ideal transforms and generalized ideal transforms 37
√ √
(ii) Deduce that, if a = b, then Da and Db are naturally equivalent.

It will be convenient, in our discussion of geometric examples in the next


section, for us to have available a result which shows that, in a certain sense, the
ideal transform is ‘independent of the base ring’. To be more precise, consider a
second commutative Noetherian ring R and a ring homomorphism f : R −→
R ; let M  be an R -module. At times when we wish to be absolutely precise,
we shall use M  R to indicate that we are regarding M  as an R-module by
means of f . Note that R can be regarded as a functor from C(R ) to C(R). We
can form the ideal transform DaR (M  ) of M  with respect to the extension
aR of a to R via f , and then regard this as an R-module by means of f : this
is, then, the R-module DaR (M  ) R . Alternatively, we can regard M  as the
R-module M  R , and form Da (M  R ). Our next result will show, among
other things, that there is an R-isomorphism

DaR (M  ) = DaR (M  ) −→ Da (M  =: Da (M  ),
=
R R)

so that, speaking loosely, it does not matter whether we calculate these ideal
transforms over R or R . The advantage of this in practice is that there is some-
times an obvious choice of an ‘R ’ over which the calculations are easy.

2.2.24 Theorem. Let R be a second commutative Noetherian ring and let


f : R −→ R be a ring homomorphism. We use the notation introduced above.
There is a natural equivalence of functors

ε : DaR ( • ) R −→ Da ( • R)

(from C(R ) to C(R)) which is such that, for each R -module M  , the diagram
ηaR ,M 
M - DaR (M  )

εM 

ηa,M 
?
M - Da (M  )

commutes.

Proof. Let M  be an R -module. By 2.2.6(i)(c), the R -homomorphism

ηaR ,M  : M  −→ DaR (M  )

has kernel and cokernel which are, respectively, isomorphic to ΓaR (M  ) and
1 
HaR  (M ). Hence ηaR ,M  R has kernel and cokernel which are a-torsion. It
38 Torsion modules and ideal transforms

therefore follows from 2.2.15 that there is a unique R-homomorphism εM  :


DaR (M  ) −→ Da (M  ) such that the diagram
ηaR ,M 
M - DaR (M  )

εM 

ηa,M 
?
M - Da (M  )

commutes. In fact, it is easy to use the uniqueness aspect of Proposition 2.2.13


to show that, as M  varies through the category C(R ), the εM  constitute a
natural transformation of functors, and so it remains only to show that each
εM  is an isomorphism.
The fact that M  is an (R, R )-bimodule means that Da (M  ) inherits a nat-
ural structure as an R -module (such that Da (r IdM  ), for r ∈ R , provides
multiplication by r ). Furthermore, since ε is a natural transformation of func-
tors, we have

Da (r IdM  ) ◦ εM  = εM  ◦ DaR (r IdM  ) = εM  ◦ (r IdDaR (M  ) )

for all r ∈ R , so that εM  is an R -homomorphism. Likewise, ηa,M  becomes


an R -homomorphism.
Another use of 2.2.6(i)(c) shows that ηa,M  : M  −→ Da (M  ) has kernel
and cokernel which are a-torsion, so that, when we consider ηa,M  as an R -
homomorphism, its kernel and cokernel are aR -torsion. It therefore follows
from 2.2.15 that there is a unique R -homomorphism λM  : Da (M  ) −→
DaR (M  ) such that the diagram
ηa,M 
M - Da (M  )

λM 

ηaR ,M 
?
M  - DaR (M  )

commutes. The uniqueness aspect of 2.2.15, together with the facts that εM 
and λM  are both R- and R -homomorphisms, now yields that

λM  ◦ εM  = IdDaR (M  ) and εM  ◦ λM  = IdDa (M  ) ,

so that εM  is an isomorphism.

2.2.25 Exercise. Let the situation be as in 2.2.24, and consider the natural
2.3 Geometrical significance 39

equivalence ε of that theorem. Show that ε−1 


R ◦ Da (f ) : Da (R) −→ DaR (R )
is the unique ring homomorphism which makes the diagram
ηa,R
R - Da (R)

? ηaR ,R
?
R - DaR (R )

commute. (You might find it helpful to adapt the argument used in the proof of
2.2.17.)

2.2.26 Exercise. Let the situation be as in 2.2.24. Show that there is a natural
equivalence of functors
1
HaR ( • ) R −→ Ha1 ( • R)

(from C(R ) to C(R)).

The result of Exercise 2.2.26 is a particular case of a more general (and very
important) result concerning ‘independence of the base ring’ which will be
established in Chapter 4.

2.3 Geometrical significance


We are now going to show that the ideal transform Da (R) has an important
geometrical significance in the case when R is the ring of regular functions on
an (irreducible) affine algebraic variety over an algebraically closed field. Sev-
eral examples in this book will be concerned with affine algebraic geometry
over the field of complex numbers C, and it is convenient for us to introduce
some notation which will be consistently used in connection with such exam-
ples.

2.3.1 Notation. Let K be an algebraically closed field. For n ∈ N, we shall


use An (K) to denote affine n-space over K, that is, K n endowed with the
Zariski topology; we shall use An to denote complex affine n-space An (C).
All unexplained mentions of topological notions, including ‘open’ and ‘closed’
subsets, in connection with affine spaces will refer to the Zariski topology.
By an affine variety over K we shall mean an irreducible closed subset of
A (K) (with the induced topology), and by a quasi-affine variety over K we
n
40 Torsion modules and ideal transforms

shall mean a non-empty open subset of an affine variety over K (again with
the induced topology).
For a quasi-affine variety U over K, we shall use O(U ) to denote the ring
of regular functions on U and K(U ) to denote the function field of U . Thus,
when U is affine, O(U ) is just the coordinate ring of U ; this is actually an
integral domain (since U is irreducible), and its field of fractions is just K(U ).
We shall regard the polynomial ring K[X1 , . . . , Xn ] as the coordinate ring
O(An (K)) of An (K) in the obvious way (although we shall tend to use X, Y
instead of X1 , X2 in the case when n = 2).
For f1 , . . . , ft ∈ K[X1 , . . . , Xn ], we shall use VAn (K) (f1 , . . . , ft ) to denote
the affine algebraic set

{p ∈ An (K) : f1 (p) = · · · = ft (p) = 0}

corresponding to the ideal (f1 , . . . , ft ) of K[X1 , . . . , Xn ]. If V denotes this


affine algebraic set, then, for 1 ≤ i ≤ n, the restriction of the coordinate
function Xi of An (K) to V will be denoted by Xi V or xi .
For a quasi-affine variety W over K and a function f ∈ O(W ), we shall
use UW (f ) to denote the open subset {p ∈ W : f (p) = 0} of W .

We are now going to show that the ring of regular functions on a quasi-affine
variety can be expressed in terms of an ideal transform. Before doing so, we
remind the reader of the following point. For a non-empty open subset U of an
affine variety V over the algebraically closed field K, the ring O(U ) can be
identified in a natural way as a subring of K(V ), and, when this identification
is made, the restriction homomorphism

U : O(V ) −→ O(U )

is just the inclusion map.

2.3.2 Theorem. Let V be an affine variety over the algebraically closed field
K. Let b be a non-zero ideal of O(V ), let V (b) denote the closed subset of V
determined by b, and let U be the open subset V \ V (b) of V .

=
There is a unique O(V )-isomorphism νV,b : O(U ) −→ Db (O(V )) for
which the diagram

O(V )
U
- O(U )

@
@
ηO(V )

= νV,b
@
@
R ?
Db (O(V ))
2.3 Geometrical significance 41

commutes. Furthermore, νV,b is a ring isomorphism.


Proof. In view of 2.2.18, it is enough for us to show that the submodules

O(U ) and n∈N (O(V ) :K(V ) bn ) of K(V ) are equal.
Since the ring O(V ) is Noetherian, b is finitely generated, by h1 , . . . , ht ,
say. We can, and do, assume that h1 , . . . , ht are all non-zero. Note that U =
t
i=1 UV (hi ), and that O(UV (hi )) = O(V )hi when these two rings are iden-
tified with subrings of K(V ) in the natural ways.
Let f ∈ O(U ). Choose i ∈ N with 1 ≤ i ≤ t. Since f UV (hi ) ∈
O(UV (hi )) = O(V )hi , it follows that there exists ni ∈ N such that hni i f ∈
O(V ). We deduce that there exists n ∈ N such that bn f ∈ O(V ). Hence

O(U ) ⊆ (O(V ) :K(V ) bn ).
n∈N

Now let f ∈ n∈N (O(V ) :K(V ) bn ). Thus there exists n ∈ N such that
gi := hni f ∈ O(V ) for all i = 1, . . . , t. Let p ∈ U , and let m(p) be the
maximal ideal of O(V ) corresponding to p. Now p ∈ UV (hi ) for some i with
1 ≤ i ≤ t, and since hi (p) = 0, we have hi ∈ O(V ) \ m(p) and
gi
f = n ∈ O(V )m(p) = OV, p ,
hi
the local ring of V at p. It follows that f ∈ p∈U OV, p = O(U ). Hence

O(U ) ⊇ (O(V ) :K(V ) bn ),
n∈N

and the proof is complete.


2.3.3 Remark. It follows from 2.3.2 that, in the notation of that proposition,
the map ηO(V ) : O(V ) −→ Db (O(V )) is an epimorphism if and only if the
restriction map U : O(V ) −→ O(U ) is surjective, that is, if and only if every
regular function on U can be extended to a regular function on V . However,
by 2.2.6(i)(c), ηO(V ) fails to be an epimorphism if and only if Hb1 (O(V )) = 0.
Thus we can, in a sense, regard non-zero elements of the local cohomology
module Hb1 (O(V )) as obstructions to the extension of regular functions on U
to regular functions on V . We shall exploit this observation later in the chapter
in our discussion of the geometric significance of ideal transforms and local
cohomology modules in particular examples.
The next exercise establishes a certain ‘naturality’ property of the isomor-
phisms given by Theorem 2.3.2.
2.3.4 Exercise. Let β : W −→ V be a morphism of affine varieties over the
algebraically closed field K. Let b be a non-zero ideal of O(V ), and assume
42 Torsion modules and ideal transforms

that the extension bO(W ) of b under the induced K-algebra homomorphism


β ∗ : O(V ) −→ O(W ) is also non-zero.
We use the notation of Theorem 2.3.2. Since β −1 (V (b)) = V (bO(W )), the
restriction of β provides a morphism β : W \ V (bO(W )) −→ V \ V (b) of
quasi-affine varieties.
Use Exercise 2.2.25 to show that, with the notation of that exercise, the
diagram

β
O(V \ V (b)) - O(W \ V (bO(W )))

νV,b ∼
= ∼
= νW,bO(W )

? ε−1 ◦Db (β ∗ )
?
Db (O(V ))
O(W )
- DbO(W ) (O(W ))

of ring homomorphisms commutes.

2.3.5 Exercise. Let ι : A2 \ {(0, 0)} → A2 denote the inclusion morphism


of varieties. Prove that the induced C-algebra homomorphism

ι∗ : O(A2 ) → O(A2 \ {(0, 0)})

is an isomorphism, and deduce that the quasi-affine variety A2 \ {(0, 0)} is not
affine.

2.3.6 Exercise. Let K be a field, and let q be a proper ideal of height 2 in


the ring of polynomials K[X, Y ] in the indeterminates X, Y . Set R := K + q,
a subring of K[X, Y ]. Observe that q is a maximal ideal of R, and that the
simple R-module R/q is isomorphic to K, where the R-module structure on
K is such that f a = 0 for all f ∈ q and all a ∈ K.

(i) Show that the vector space dimension t := dimK (K[X, Y ]/q) is finite,
and deduce that there is a monic polynomial in K[X] ∩ q of degree not
exceeding t: let pX be such a polynomial of smallest possible degree.
Similarly, let pY be a monic polynomial in K[Y ] ∩ q of smallest possible
degree. Show that K[X, Y ] is a finitely generated R-module.
(ii) Let c1 , . . . , ch generate q (as an ideal of K[X, Y ]), let u := deg pX and
v := deg pY . Set

S := {ci X j Y l : 1 ≤ i ≤ h, 0 ≤ j < u, 0 ≤ l < v} ∪ {pX , pY } ⊂ R.

Show that R = K[S], so that R is a finitely generated K-algebra, and


therefore Noetherian.
2.3 Geometrical significance 43

(iii) Show that there is a unique R-isomorphism ψ  : K[X, Y ] −→ Dq (R)


such that the diagram

R - K[X, Y ]

@
@
ηR
ψ
@
@
R ?
Dq (R)

commutes, and that ψ  is a ring isomorphism. Deduce that there is an


exact sequence 0 −→ R −→ K[X, Y ] −→ Hq1 (R) −→ 0 in C(R), and
that dimK (Hq1 (R)) = t − 1.
(iv) In the special case in which q := XK[X, Y ] + Y (Y − 1)K[X, Y ], show
that R = K[X, XY, Y (Y − 1), Y 2 (Y − 1)], that q is the maximal ideal
(X, XY, Y (Y − 1), Y 2 (Y − 1)) of R, and that Hq1 (R) ∼
= K.
(v) In the special case in which q := XK[X, Y ] + Y 2 K[X, Y ], show that
R = K[X, Y 2 , XY, Y 3 ], that q is the maximal ideal (X, Y 2 , XY, Y 3 )
of R, and that Hq1 (R) ∼= K.
We are now in a position to present one of the geometric examples which
was promised earlier.

2.3.7 Example. (See R. Hartshorne [28, 3.4.2].) With the notation of 2.3.1,
let V be the affine algebraic set in A4 given by

V := VA4 (X1 X4 − X2 X3 , X12 X3 + X1 X2 − X22 , X33 + X3 X4 − X42 ).

It is easy to check that the morphism of varieties α : A2 → A4 for which


α((c, d)) = (c, cd, d(d − 1), d2 (d − 1)) for all (c, d) ∈ A2 satisfies Im α ⊆ V .
In fact, Im α = V , because the map β : V → A2 defined by



⎨(c1 , c2 /c1 ) if c1 = 0,
β((c1 , c2 , c3 , c4 )) = (c1 , c4 /c3 ) if c3 = 0,


⎩(0, 0) if c = c = 0
1 3

(for all (c1 , c2 , c3 , c4 ) ∈ V ) satisfies α ◦ β = IdV , the identity map on V .


Since A2 is irreducible, it follows that V is irreducible, too. We propose
to study the coordinate ring, R := O(V ), of this affine variety. Note that the
C-algebra homomorphism

α∗ : O(V ) −→ O(A2 ) = C[X, Y ]

induced by α is injective (because α is surjective). Also, if ι : V → A4 denotes


44 Torsion modules and ideal transforms

the inclusion morphism of varieties, then ι∗ : C[X1 , X2 , X3 , X4 ] → O(V ) is


the natural surjective C-algebra homomorphism given by restriction. Further-
more, Im α∗ = C[X, XY, Y (Y − 1), Y 2 (Y − 1)], since
α∗ (ι∗ (X1 )) = X, α∗ (ι∗ (X2 )) = XY,

α∗ (ι∗ (X3 )) = Y (Y − 1), α∗ (ι∗ (X4 )) = Y 2 (Y − 1).


It follows from 2.3.6(iv) that, if we use m to denote the maximal ideal of
R corresponding to the point (0, 0, 0, 0) of V , then Hm 1
(R) ∼= C. Note that
the open subset V \ V (m) is just V \ {(0, 0, 0, 0)}. By 2.3.2 and 2.3.3, there
must be a regular function on V \ {(0, 0, 0, 0)} that cannot be extended to a
regular function on V . The reader might find it interesting for us to find such a
function.
Observe that
V \ {(0, 0, 0, 0)} = UV (x1 ) ∪ UV (x3 )
and that, on UV (x1 )∩UV (x3 ), the functions x2 /x1 and x4 /x3 are both defined
and are equal. Thus β2 : V \ {(0, 0, 0, 0)} → C defined by

c2 /c1 if c1 = 0,
β2 ((c1 , c2 , c3 , c4 )) =
c4 /c3 if c3 = 0
(for all (c1 , c2 , c3 , c4 ) ∈ V \ {(0, 0, 0, 0)}) is a regular function on the set
V \{(0, 0, 0, 0)}. (In fact, the map from V \{(0, 0, 0, 0)} to A2 \{(0, 0), (0, 1)}
given by (c1 , c2 , c3 , c4 ) → (c1 , β2 ((c1 , c2 , c3 , c4 ))) is actually an isomorphism
of (quasi-affine) varieties, because it is inverse to α : A2 \ {(0, 0), (0, 1)} −→
V \ {(0, 0, 0, 0)}.)
Suppose that β2 can be extended to a regular function β2 : V → C, and
look for a contradiction. Now for v ∈ UV (x3 ), we have β2 (v) = x4 (v)/x3 (v).
Since x4 (v)2 = x3 (v)3 + x3 (v)x4 (v), it follows that
x4 (v)2 x4 (v)
β2 (v)2 − β2 (v) = −
x3 (v)2 x3 (v)
3
x3 (v) + x3 (v)x4 (v) x4 (v)
= − = x3 (v)
x3 (v)2 x3 (v)
for all v ∈ UV (x3 ). Since UV (x3 ) is a dense open subset of V , we deduce that
β22 −β2 = x3 . Hence β2 ((0, 0, 0, 0)) = ε, where ε = 0 or 1. Now the map β  :
V −→ A2 \ {(0, 1 − ε)} given by (c1 , c2 , c3 , c4 ) −→ (c1 , β2 ((c1 , c2 , c3 , c4 )))
is a morphism of varieties. In fact, it is an isomorphism of varieties, because
α : A2 \ {(0, 1 − ε)} −→ V is an inverse for it. This shows that the quasi-
affine variety A2 \ {(0, 1 − ε)} is affine.
2.3 Geometrical significance 45

However, Exercise 2.3.5 shows that the quasi-affine variety A2 \ {(0, 0)} is
not affine, and a similar argument will show that A2 \ {(0, 1)} is not affine. We
have therefore arrived at a contradiction, and this shows that
β2 : V \ {(0, 0, 0, 0)} → C
is a regular function which cannot be extended to a regular function on V .
2.3.8 Exercise. With the notation of 2.3.1, let V be the affine algebraic set in
A4 given by
V := VA4 (X12 X2 − X32 , X23 − X42 , X2 X3 − X1 X4 , X1 X22 − X3 X4 ).
(i) Show that the morphism of varieties α : A2 → A4 for which
α((c, d)) = (c, d2 , cd, d3 ) for all (c, d) ∈ A2
is injective and that its image is equal to V . Deduce that V is irreducible,
and so is an affine variety.
(ii) Show that
α : A2 \ {(0, 0)} −→ V \ {(0, 0, 0, 0)}
is an isomorphism of (quasi-affine) varieties, with inverse
β : V \ {(0, 0, 0, 0)} −→ A2 \ {(0, 0)}
given by

(c1 , c3 /c1 ) if c1 = 0,
β((c1 , c2 , c3 , c4 )) =
(c1 , c4 /c2 ) if c2 = 0
(for all (c1 , c2 , c3 , c4 ) ∈ V \ {(0, 0, 0, 0)}).
(iii) Let m denote the maximal ideal of O(V ) corresponding to the point
(0, 0, 0, 0) of V . Show that Hm 1
(O(V )) ∼= C.
(iv) Show that the function β2 : V \ {(0, 0, 0, 0)} → C defined by

c3 /c1 if c1 = 0,
β2 ((c1 , c2 , c3 , c4 )) =
c4 /c2 if c2 = 0
(for all (c1 , c2 , c3 , c4 ) ∈ V \ {(0, 0, 0, 0)}) is a regular function on the
set V \ {(0, 0, 0, 0)} that cannot be extended to a regular function on V .
Although Example 2.3.7 and Exercise 2.3.8 look very similar from an alge-
braic point of view, there are substantial geometric differences between them.
This is illustrated by the next two exercises. By the metric topology on an
affine variety V ⊆ An we mean the (subspace) topology induced on V by the
topology defined on Cn by the standard distance metric of analysis.
46 Torsion modules and ideal transforms

2.3.9 Exercise. Let the situation and notation be as in 2.3.7.


Prove that there does not exist a mapping β2 : V −→ C which extends
β2 and which is continuous for the metric topology. (Here is a hint: consider
points on the path s : {t ∈ R : 0 ≤ t ≤ 1} → V given by s(t) = α((0, t)) =
(0, 0, t(t − 1), t2 (t − 1)) (for 0 ≤ t ≤ 1) to show that, in the metric topology,
limv→(0,0,0,0) β2 (v) does not exist.)
2.3.10 Exercise. Let the situation and notation be as in 2.3.8.
Let β  : V −→ A2 be the map which extends
β : V \ {(0, 0, 0, 0)} −→ A2 \ {(0, 0)}
and is such that β  ((0, 0, 0, 0)) = (0, 0). Show that β  is actually continuous
for the metric topology, and deduce that α : A2 −→ V is a homeomorphism
with respect to the metric topology. (Again we offer a hint: show that β2 (v)2 =
x2 (v) for all v ∈ UV (x1 ) ∪ UV (x2 ) = V \ {(0, 0, 0, 0)}.)
2.3.11 Exercise. Use the notation of 2.3.1.
(i) Let 0 = g ∈ C[X, Y ] = O(A2 ). Let U := UA2 (g), a quasi-affine variety
over C, and let p ∈ U . Show that each regular function f : U \ {p} → C
can be extended to a regular function on U . (Do not forget that U is
affine!)
(ii) Let f ∈ K(A2 ). There is a maximum open subset U  of A2 on which f
is defined: the poles of f are precisely the points of A2 \ U  . Use (i) to
show that there does not exist an isolated pole of f , that is, a pole q of f
for which there exists an open subset U  of A2 such that q ∈ U  but q
is the only pole of f in U  .
3

The Mayer–Vietoris sequence

Any reader with a basic grounding in algebraic topology will recall the impor-
tant rôle that the Mayer–Vietoris sequence can play in that subject. There is an
analogue of the Mayer–Vietoris sequence in local cohomology theory, and it
can play a foundational rôle in this subject. It is our intention in this chapter to
present the basic theory of the Mayer–Vietoris sequence in local cohomology,
and to prepare for several uses of the idea during the subsequent development.
The Mayer–Vietoris sequence involves two ideals, and so throughout this
chapter, b will denote a second ideal of R (in addition to a). Let M be an R-
module. The Mayer–Vietoris sequence provides, among other things, a long
exact sequence

0 - H 0 (M ) - H 0 (M ) ⊕ H 0 (M ) - H 0 (M )
a+b a b a∩b

- H 1 (M ) - H 1 (M ) ⊕ H 1 (M ) - H 1 (M )
a+b a b a∩b

- ··· ···
- H i (M ) - H i (M ) ⊕ H i (M ) - H i (M )
a+b a b a∩b

- H i+1 (M ) - ···
a+b

of local cohomology modules. Its potential for use in arguments that employ
induction on the number of elements in a generating set for an ideal c of R can
be explained as follows. Suppose that c is generated by n elements c1 , . . . , cn ,
where n > 1. Set a = Rc1 +· · ·+Rcn−1 and b = Rcn , so that c = a+b. Each
of a and b can be generated by fewer than n elements, but at first sight it seems
that the ideal a ∩ b, which also appears in the Mayer–Vietoris sequence, could
√ √
present difficulties. However, (a ∩ b) = (ab), and so Γa∩b = Γab by 1.1.3;
hence Ha∩bi
= Habi
for all i ∈ N0 (see 1.2.3). Moreover, in our situation,
ab = (Rc1 + · · · + Rcn−1 )Rcn = Rc1 cn + · · · + Rcn−1 cn
48 The Mayer–Vietoris sequence

can be generated by n − 1 elements. Thus a, b and ab can all be generated by


fewer than n elements, and an appropriate inductive hypothesis would apply to
all of them. In addition, we have already obtained (in 2.2.21) a certain amount
of information about the local cohomology functors HRa i
(i ∈ N0 ) with respect
to a principal ideal Ra of R: this can provide a basis for an inductive argument.

3.1 Comparison of systems of ideals


i
The result that Ha∩b = Hab i
for all i ∈ N0 can be viewed as a particular ex-
ample of a more general phenomenon which concerns a situation where two
systems of ideals are ‘comparable’ in a sense made precise in the proposi-
tion below. This comparison result will not only be used to obtain the Mayer–
Vietoris sequence; it will also provide an important ingredient, which is the
subject of Exercise 3.1.4, in our proof of the local Lichtenbaum–Hartshorne
Vanishing Theorem in Chapter 8.

3.1.1 Proposition. Let (Λ, ≤) and (Π, ≤) be (non-empty) directed partially


ordered sets, and let B = (bα )α∈Λ be an inverse family of ideals of R over Λ,
as in 1.2.10; let C = (cβ )β∈Π be an inverse family of ideals of R over Π.
Assume that, for all α ∈ Λ, there exists β ∈ Π such that cβ ⊆ bα , and, for
all β  ∈ Π, there exists α ∈ Λ such that bα ⊆ cβ  . Then

(i) ΓB = ΓC ;
(ii) the negative strongly connected sequences of covariant functors
   
lim ExtiR (R/bα , • ) and lim ExtiR (R/cβ , • )
−→ −→
α∈Λ i∈N0 β∈Π i∈N0

are isomorphic; and


(iii) B is a system of ideals of R over Λ (in the sense of 2.1.10) if and only if
C is a system of ideals of R over Π.

Note. The functor ΓB was defined in 1.2.11, while in 1.3.7 it was explained
that the negative strongly connected sequence (Ri ΓB )i∈N0 of its right derived
functors is isomorphic to
 
lim ExtiR (R/bα , • ) ;
−→
α∈Λ i∈N0

of course, similar comments apply to ΓC .


3.1 Comparison of systems of ideals 49

Proof. Part (i) is clear; hence Ri ΓB = Ri ΓC for all i ∈ N0 , and the claim in
(ii) is an obvious consequence of the above note.
(iii) Suppose that B is a system of ideals over Λ, and let β1 , β2 ∈ Π. By
assumption, there exist α1 , α2 ∈ Λ such that bα1 ⊆ cβ1 and bα2 ⊆ cβ2 . Since
B is a system of ideals, there exists α3 ∈ Λ such that bα3 ⊆ bα1 bα2 . By
assumption, there exists β3 ∈ Π such that cβ3 ⊆ bα3 . Hence

cβ3 ⊆ bα3 ⊆ bα1 bα2 ⊆ cβ1 cβ2 .

Hence C is a system of ideals over Π.


In view of the symmetry of our hypotheses on B and C, the proof is com-
plete.

3.1.2 Example. Consider the descending chain B = (an + bn )n∈N of ideals


of R. Since (a + b)2n−1 ⊆ an + bn ⊆ (a + b)n for all n ∈ N, it follows from
3.1.1 that B is actually a system of ideals of R over N; also, it follows from
1.3.7 and 3.1.1 that the (negative strongly) connected sequences of covariant
functors (from C(R) to C(R))
 
i
Ha+b , lim ExtiR (R/(an + bn ), • ) and i
HB
i∈N0 −→ i∈N0
n∈N i∈N0

are all isomorphic. In particular, for each i ∈ N0 , the functors Ha+b


i
and
i n n
lim ExtR (R/(a + b ), • ) are naturally equivalent.
−→
n∈N

3.1.3 Exercise. Let (R, m) be a complete local ring, and let B = (bn )n∈N
be a descending chain of m-primary ideals of R such that n∈N bn = 0.
Show that B is a system of ideals of R over N, and that the (negative
strongly) connected sequences of covariant functors (from C(R) to C(R))
 
i
Hm , lim ExtiR (R/bn , • ) and i
HB
i∈N0 −→ i∈N0
n∈N i∈N0

are all isomorphic. (If you have no idea where to start with this, we suggest
that you look up Chevalley’s Theorem in, for example, [59, §5.2, Theorem 1]
or [89, Chapter VIII, §5, Theorem 13].)

3.1.4 Exercise. Let (R, m) be a complete local domain, and let p be a prime
ideal of R of dimension 1, that is, such that dim R/p = 1.

(i) Use Chevalley’s Theorem ([59, §5.2, Theorem 1] or [89, Chapter VIII,
§5, Theorem 13]) to prove that, for each n ∈ N, there exists t ∈ N such
that the t-th symbolic power p(t) of p is contained in mn .
50 The Mayer–Vietoris sequence

(ii) Deduce that (p(n) )n∈N is a system of ideals of R over N, and that the
(negative strongly) connected sequences of covariant functors
 
Hpi and lim ExtiR (R/p(n) , • )
i∈N0 −→
n∈N i∈N0

(from C(R) to C(R)) are isomorphic.

3.1.5 Corollary. The descending chain (an ∩ bn )n∈N is a system of ideals


of R (over N), and the (negative strongly) connected sequences of covariant
functors (from C(R) to C(R))
 
lim ExtiR (R/(an ∩ bn ), • ) and i
Ha∩b
−→ i∈N0
n∈N i∈N0

are isomorphic. In particular, for each i ∈ N0 , the functors Ha∩b


i i
= Hab and
lim ExtR (R/(a ∩ b ), • ) are naturally equivalent.
i n n
−→
n∈N

Proof. As we mentioned at the beginning of this chapter, it follows from 1.2.3


i
that Ha∩b = Habi
for all i ∈ N0 . Since (a ∩ b)n ⊆ an ∩ bn for all n ∈ N, it
is enough, in view of 3.1.1, for us to show that, for each n ∈ N, there exists
q(n) ∈ N such that aq(n) ∩ bq(n) ⊆ (a ∩ b)n . We use the Artin–Rees Lemma
[50, Theorem 8.5] to achieve this.
Fix n ∈ N. By the Artin–Rees Lemma, there is c ∈ N such that am ∩ bn =
a (a ∩ bn ) for all integers m > c. Hence
m−c c

an+c ∩ bn+c ⊆ an+c ∩ bn = an (ac ∩ bn ) ⊆ an bn ⊆ (a ∩ b)n .

The proof is therefore complete.

3.1.6 Exercise. For each h ∈ N, let a[h] denote the ideal of R generated by
all the h-th powers of elements of a. Show that, for each i ∈ N0 , the functor
Hai is naturally equivalent to lim ExtiR (R/a[h] , • ).
−→
h∈N
(A variation on this is the result that, if R contains a subfield of character-
istic p > 0, then, for each i ∈ N0 , the functor Hai is naturally equivalent to
e
lim ExtiR (R/a[p ] , • ). There are situations where this observation can be very
−→
e∈N
useful: see, for example, C. Peskine and L. Szpiro [66, Chapitre III, Proposi-
tion 1.8(3)], and C. L. Huneke and R. Y. Sharp [43, p. 770].)

3.1.7 Exercise. Suppose that the t elements a1 , . . . , at generate a. Now Nt


is a directed partially ordered set with respect to the ordering ≤ defined by, for
3.2 Construction of the sequence 51

(u1 , . . . , ut ), (v1 , . . . , vt ) ∈ Nt ,
(u1 , . . . , ut ) ≤ (v1 , . . . , vt ) if and only if uj ≤ vj for all j = 1, . . . , t.
Show that, for each i ∈ N0 , the functor Hai is naturally equivalent to
lim ExtiR (R/(au1 1 , . . . , aut t ), • ).
−→
(u1 ,...,ut )∈Nt

3.2 Construction of the sequence


Our first lemma in this section provides a fundamental tool for the construction
of the Mayer–Vietoris sequence.
3.2.1 Lemma. Let N1 , N2 be submodules of the R-module M . The sequence
of R-modules and R-homomorphisms
α β
0 −→ M/(N1 ∩ N2 ) −→ M/N1 ⊕ M/N2 −→ M/(N1 + N2 ) −→ 0,
in which α(m + N1 ∩ N2 ) = (m + N1 , m + N2 ) for all m ∈ M and
β((x + N1 , y + N2 )) = x − y + (N1 + N2 ) for all x, y ∈ M,
is exact.
Proof. It is clear that α is injective, that β is surjective and that β ◦ α = 0.
Let x, y ∈ M be such that (x + N1 , y + N2 ) ∈ Ker β. Then x − y = n1 + n2
for some n1 ∈ N1 , n2 ∈ N2 , so that x − n1 = y + n2 and
(x + N1 , y + N2 ) = (x − n1 + N1 , y + n2 + N2 ) ∈ Im α.
Briefly, the general strategy for our construction of the Mayer–Vietoris
sequence is to write down, for each n ∈ N, the long exact sequence of ‘Ext’
modules which results from application of the functor HomR ( • , M ) to the
exact sequence
0 −→ R/(an ∩ bn ) −→ R/an ⊕ R/bn −→ R/(an + bn ) −→ 0
resulting from 3.2.1, pass to direct limits, and then appeal to 3.1.2 and 3.1.5
to convert the result into information about local cohomology modules. There
is, however, one minor technical point that needs attention first, and that is the
identification of Hai ( • ) ⊕ Hbi ( • ) with the functor
lim ExtiR (R/an ⊕ R/bn , • )
−→
n∈N

(obtained using the ideas of 1.2.8 in an obvious way). Although perhaps a little
tedious, this is not particularly difficult.
52 The Mayer–Vietoris sequence

3.2.2 Exercise. In this exercise, we use, for R-modules L and M , the sym-
bolism
 - L  - L⊕M  - M  -0
0

to denote simultaneously the two split exact sequences associated with the di-
rect sum: thus we consider just the arrows pointing to the right to obtain one
of these split exact sequences, while the arrows pointing to the left provide the
other.
Let T : C(R)×C(R) → C(R) be an R-linear functor of two variables which
is contravariant in the first variable and covariant in the second. (For example,
T could be HomR or ExtiR for i ∈ N0 .)
For n, m ∈ N with n ≥ m, let

hnm : R/an −→ R/am n


and km : R/bn −→ R/bm

denote the natural homomorphisms, and consider the commutative diagrams


 - R/an  - R/an ⊕ R/bn  - R/bn  - 0
0

hn m ⊕km
hn n n
m km

? ? ?
 - R/am  - R/am ⊕ R/bm  - R/bm  - 0 ,
0

in which the rows represent the canonical split exact sequences. Apply the
contravariant, additive functor T ( • , M ), and pass to direct limits to obtain
split exact sequences
- lim T (R/bn , M )  - lim T (R/an ⊕ R/bn , M )
0 −→ −→
n∈N n∈N
 - lim T (R/an , M )  -
−→
0.
n∈N

Deduce that there is a natural equivalence between the functors

lim T (R/an ⊕ R/bn , • ) and lim T (R/an , • ) ⊕ lim T (R/bn , • )


−→ −→ −→
n∈N n∈N n∈N

(from C(R) to itself).


Deduce from 1.3.8 that the functors

lim ExtiR (R/an ⊕ R/bn , • ) and Hai ( • ) ⊕ Hbi ( • )


−→
n∈N

are naturally equivalent for each i ∈ N0 .


3.2 Construction of the sequence 53

3.2.3 Theorem: the Mayer–Vietoris sequence. For each R-module M ,


there is a long exact sequence (called the Mayer–Vietoris sequence for M
with respect to a and b)

0 - H 0 (M ) - H 0 (M ) ⊕ H 0 (M ) - H 0 (M )
a+b a b a∩b

- H 1 (M ) - H 1 (M ) ⊕ H 1 (M ) - H 1 (M )
a+b a b a∩b

- ··· ···
- H i (M ) - H i (M ) ⊕ H i (M ) - H i (M )
a+b a b a∩b

- H i+1 (M ) - ···
a+b

such that, whenever f : M −→ N is a homomorphism of R-modules, the


diagram

i
Ha+b (M ) - H i (M ) ⊕ H i (M ) - H i (M ) - H i+1 (M )
a b a∩b a+b

i i+1
Ha+b (f ) Hai (f )⊕Hb
i
(f ) i
Ha∩b (f ) Ha+b (f )
? ? ? ?
i
Ha+b (N ) - H i (N ) ⊕ H i (N ) - H i (N ) - H i+1 (N )
a b a∩b a+b

commutes for all i ∈ N0 .

Proof. Let n, m ∈ N with n ≥ m, and let

hnm : R/an −→ R/am n


and km : R/bn −→ R/bm

denote the natural homomorphisms. The diagram

0 - R/(an ∩ bn ) - R/an ⊕ R/bn - R/(an + bn ) -0

m ⊕km
hn n

? ? ?
0 - R/(am ∩ bm ) - R/am ⊕ R/bm - R/(am + bm ) - 0,

in which the upper row is the exact sequence resulting from application of 3.2.1
to the submodules an and bn of R, the lower row is the corresponding exact
sequence for m instead of n, and the two outer vertical homomorphisms are the
natural ones, commutes. Therefore, application of the functor HomR ( • , M )
54 The Mayer–Vietoris sequence

to this yields a long exact sequence

0 - HomR (R/(am + bm ), M ) - HomR (R/am ⊕ R/bm , M )


- HomR (R/(am ∩ bm ), M ) - Ext1 (R/(am + bm ), M )
R

- ···
- ExtiR (R/(am + bm ), M ) - Exti (R/am ⊕ R/bm , M )
R

- ExtiR (R/(am ∩ bm ), M ) - Exti+1 (R/(am + bm ), M )


R

- ···

and a chain map (induced by the vertical homomorphisms in the last commu-
tative diagram) of this long exact sequence into the corresponding one with m
replaced by n. Also, a homomorphism f : M → N of R-modules induces a
chain map of the above displayed long exact sequence into the corresponding
sequence with M replaced by N . Now pass to direct limits: it follows that there
is a long exact sequence

0 - lim HomR (R/(an + bn ), M ) - lim HomR (R/an ⊕ R/bn , M )


−→ −→
n∈N n∈N
- lim HomR (R/(an ∩ bn ), M ) - lim Ext1 (R/(an + bn ), M )
−→ −→ R
n∈N n∈N
- ···
- lim Exti (R/(an + bn ), M ) - lim Exti (R/an ⊕ R/bn , M )
−→ R −→ R
n∈N n∈N
- lim ExtiR (R/(an ∩ bn ), M ) - lim Exti+1 (R/(an + bn ), M )
−→ −→R
n∈N n∈N
- ···

and that f : M → N induces a chain map of this long exact sequence into the
corresponding sequence with M replaced by N . The result is now an immedi-
ate consequence of the natural equivalences of functors between
lim ExtiR (R/(an + bn ), • ) and i
Ha+b ,
−→
n∈N

lim ExtiR (R/an ⊕ R/bn , • ) and Hai ( • ) ⊕ Hbi ( • )


−→
n∈N

and
lim ExtiR (R/(an ∩ bn ), • ) and i
Ha∩b
−→
n∈N
3.3 Arithmetic rank 55

(for i ∈ N0 ) established in 3.1.2, 3.2.2 and 3.1.5 respectively.


3.2.4 Exercise. Let M be an R-module. Show that the sequence of R-modu-
les and R-homomorphisms
α
0 −→ HomR (a + b, M ) −→ HomR (a, M ) ⊕ HomR (b, M )
β
−→ HomR (a ∩ b, M )
in which α(f ) = (f a , f b ) for all f ∈ HomR (a + b, M ), and β((g, h)) =
g a∩b − h a∩b for all (g, h) ∈ HomR (a, M ) ⊕ HomR (b, M ), is exact.
Show also that, if M is an injective R-module, then β is an epimorphism.
3.2.5 Exercise. Prove, perhaps with the aid of Exercise 3.2.4, that, for each
R-module M , there is a long exact sequence

0 - Da+b (M ) - Da (M ) ⊕ Db (M ) - Da∩b (M )
- H 2 (M ) - H 2 (M ) ⊕ H 2 (M ) - H 2 (M )
a+b a b a∩b

- ··· ···
- H i (M ) - H i (M ) ⊕ H i (M ) - H i (M )
a+b a b a∩b

- H i+1 (M ) - ···
a+b

such that each homomorphism f : M → N of R-modules induces a chain map


of the above long exact sequence into the corresponding long exact sequence
for N .
As we have already mentioned, we shall make substantial use of the Mayer–
Vietoris sequence in this book. We present in the next two sections two illus-
trations of its use in order to give some idea of its potential to any reader who,
wondering why so much effort has been expended in setting up the Mayer–
Vietoris sequence, is impatient to see some applications.

3.3 Arithmetic rank


Our first application of the Mayer–Vietoris sequence will be in the proof of
a result which relates the number of elements required to generate the ideal
a (and, more precisely, its so-called ‘arithmetic rank’) to the vanishing of the
local cohomology functors Hai . This vanishing result provides a powerful tool
for applications of local cohomology to algebraic geometry, and, in particular,
will play a crucial rôle in our presentation of applications to connectivity in
Chapter 19.
56 The Mayer–Vietoris sequence

3.3.1 Theorem. Suppose that a can be generated by t elements. Then, for


every R-module M , we have Hai (M ) = 0 for all i > t.
Proof. This proof is an example of the use of induction outlined in the intro-
duction to this chapter.
When t = 0, we have a = 0 and Γa = Γ0R is the identity functor, so that
i
H0R (M ) = 0 for all i > 0. The result is therefore true when t = 0, and it was
proved in 2.2.21(ii) in the special case in which t = 1.
Now suppose, inductively, that t > 1 and the result has been proved for
ideals that can be generated by fewer than t elements. Suppose that a is gener-
ated by t elements a1 , . . . , at . Set b = Ra1 + · · · + Rat−1 and c = Rat , so
that a = b + c. By the inductive assumption, Hbi (M ) = 0 for all i > t − 1 and
Hci (M ) = 0 for all i > 1. By the Mayer–Vietoris sequence 3.2.3, we have, for
an arbitrary i > t, an exact sequence
i−1
Hb∩c (M ) −→ Hai (M ) −→ Hbi (M ) ⊕ Hci (M ).
i−1 i−1
Now Hb∩c (M ) = Hbc (M ) by 3.1.5, and, since
bc = (Ra1 + · · · + Rat−1 )Rat = Ra1 at + · · · + Rat−1 at
can be generated by t − 1 elements, it follows from the inductive assumption
i−1
that Hbc (M ) = 0. Since Hbi (M ) ⊕ Hci (M ) = 0 also, we have Hai (M ) = 0.
This completes the inductive step.
We now combine 3.3.1 with 1.2.3 to obtain a result which is often used
in geometric applications. To formulate this result, let us recall the following
definition from basic commutative algebra.
3.3.2 Definition. The arithmetic rank of a, denoted by ara(a), is the least
number of elements of R required to generate an ideal which has the same
radical as a. Thus ara(a) is equal to the integer
√ √
min {n ∈ N0 : ∃ b1 , . . . , bn ∈ R with (Rb1 + · · · + Rbn ) = a} .
Note that ara(0R) = 0.
The next corollary is now immediate from 3.3.1 and 1.2.3.
3.3.3 Corollary. For every R-module M , we have Hai (M ) = 0 for all i >
ara(a). 
Corollary 3.3.3 leads naturally to the following definition.
3.3.4 Definition. The cohomological dimension of a, denoted cohd(a), is
defined as the greatest integer i for which there exists an R-module M with
Hai (M ) = 0 if any such integers exist, and −∞ otherwise (for example
3.3 Arithmetic rank 57

when a = R). It follows from 3.3.3 that this definition makes sense, and that
cohd(a) ≤ ara a.

Theorem 3.3.1 can be used to obtain information about the number of el-
ements required to generate a specified ideal; also, in geometric situations,
Corollary 3.3.3 can be used to obtain information about the number of ‘equa-
tions’ needed to define an algebraic variety. The following example illustrates
this.

3.3.5 Example. This example concerns the affine variety V in A4 studied in


2.3.7 and given by

V := VA4 (X1 X4 − X2 X3 , X12 X3 + X1 X2 − X22 , X33 + X3 X4 − X42 ).

By 2.3.7, V = (c, cd, d(d − 1), d2 (d − 1)) ∈ A4 : c, d ∈ C . Thus the lines


L, L in A4 given by L := VA4 (X2 , X3 , X4 ) = (c, 0, 0, 0) ∈ A4 : c ∈ C
and L := VA4 (X1 − X2 , X3 , X4 ) = (c, c, 0, 0) ∈ A4 : c ∈ C are both
contained in V : our aim here is to use 3.3.3 to show that the subvariety L of V
cannot be ‘defined by one equation’, that is, there does not exist a polynomial
f ∈ C[X1 , X2 , X3 , X4 ] such that

L = {p ∈ V : f (x1 , x2 , x3 , x4 )(p) = 0} .

Our argument uses the morphism of varieties α : A2 → V of 2.3.7 for which


α((c, d)) = (c, cd, d(d − 1), d2 (d − 1)) for all (c, d) ∈ A2 ; recall that α :
A2 \ {(0, 0), (0, 1)} −→ V \ {(0, 0, 0, 0)} is an isomorphism of (quasi-affine)
varieties. Set L := (c, 0) ∈ A2 : c ∈ C and L := (c, 1) ∈ A2 : c ∈ C .
Note that α(L) = L and α(L ) = L .
Write O(A2 ) = C[X, Y ]; note that L = VA2 (Y ) and that (0, 1) ∈ L. It
therefore follows from 2.3.11 that the restriction map

O(A2 \ L) −→ O(A2 \ ({(0, 1)} ∪ L))

is surjective. Now L \ {(0, 1)} ⊆ A2 \ ({(0, 1)} ∪ L): our immediate aim is
to show that the restriction map O(A2 \ ({(0, 1)} ∪ L)) −→ O(L \ {(0, 1)})
is not surjective.
Set x := X L and consider the regular function γ : L \ {(0, 1)} → C
defined by γ(v) = x(v)−1 for all v ∈ L \ {(0, 1)} = UL (x). If γ could be
extended to a regular function on A2 \ ({(0, 1)} ∪ L), then it would follow
from the consequence of 2.3.11 noted in the previous paragraph that γ could
be extended to a regular function γ  : A2 \ L → C. Let γ  be the restriction
of γ  to L , a subset of A2 \ L. It then follows that, for all v ∈ UL (x), we
have γ  (v)x(v) = 1. However, UL (x) is a dense open subset of L , and so
58 The Mayer–Vietoris sequence

γ  (v)x(v) = 1 for all v ∈ L . As this implies that 0 = 1, we have obtained a


contradiction! Thus the restriction map

O(A2 \ ({(0, 1)} ∪ L)) −→ O(L \ {(0, 1)})

is not surjective, as claimed.


Next, α(A2 \ ({(0, 1)} ∪ L)) = V \ L and

α(L \ {(0, 1)}) = L \ {(0, 0, 0, 0)} .

There is therefore a commutative diagram

O(V \ L) - O(L \ {(0, 0, 0, 0)})


= ∼
=

? ?
O(A2 \ ({(0, 1)} ∪ L)) - O(L \ {(0, 1)})

in which the vertical isomorphisms are induced by (the appropriate restrictions


of) α. Therefore the map σ : O(V \ L) −→ O(L \ {(0, 0, 0, 0)}) given by
restriction is not surjective.
Let ι : L −→ V be the inclusion morphism of varieties, and let c denote
the vanishing ideal {f ∈ O(V ) : f (v) = 0 for all v ∈ L} of L in O(V ). With
the notation of 2.3.2 and 2.3.4, we have

V (cO(L )) = ι−1 (V (c)) = ι−1 (L) = L ∩ L = {(0, 0, 0, 0)} ⊂ L ,

and so cO(L ) = 0. Note also that the restriction map σ of the preceding
paragraph is just ι ∗ : O(V \ V (c)) −→ O(L \ V (cO(L ))). Therefore ι ∗
is not surjective, and so it follows from 2.3.4 that Dc (ι∗ ) : Dc (O(V )) −→
Dc (O(L )) is not surjective. Hence, if b := Ker(ι∗ ), we have R1 Dc (b) = 0,
so that Hc2 (b) = 0 by 2.2.6(ii). It therefore follows from 3.3.3 that ara(c) ≥ 2.
Thus the subvariety L of V cannot be ‘defined by one equation’.

3.3.6 Exercise. Show that, over the ring R[X1 , X2 , X3 , X4 , X5 , X6 ] of pol-


ynomials in six indeterminates with coefficients in R,
4
H(X 1 ,X2 ,X3 )(X4 ,X5 ,X6 )
(R[X1 , X2 , X3 , X4 , X5 , X6 ]) = 0.

3.3.7 Exercise. Show that, over the polynomial ring R[X1 , X2 , X3 ],


3
H(X 2 ,X X +X 3 ,X 4 ) (R[X1 , X2 , X3 ]) = 0.
1 2
1 2 2
3.4 Direct limits 59

3.4 Direct limits


Our second illustration of the use of the Mayer–Vietoris sequence is in our
proof that the local cohomology functors ‘commute with the formation of di-
rect limits’. This is an important property of local cohomology, which plays
a significant rôle in Grothendieck’s development in [25]. It should be noted
that our approach below avoids the use of an injective resolution of a direct
system of R-modules and R-homomorphisms in the category formed by such
direct systems. It is possible to approach this theory in a different way, based
on the interchange of the order of direct limits; however, we have chosen the
approach below because we consider it to be more illuminating.

3.4.1 Terminology. Let (Λ, ≤) be a (non-empty) directed partially ordered


set, and let (Wα )α∈Λ be a direct system of R-modules over Λ, with constituent
R-homomorphisms hα β : Wβ → Wα (for each (α, β) ∈ Λ × Λ with α ≥ β).
Set W∞ := lim Wα , and let hα : Wα → W∞ be the canonical map (for each
−→
α∈Λ
α ∈ Λ).
Let T : C(R) → C(R) be a covariant functor. It is immediate from the
definition of functor that the T (hα
β ) turn the family (T (Wα ))α∈Λ into a direct
system of R-modules and R-homomorphisms over Λ. Also, whenever α, β ∈
Λ with α ≥ β, the commutative diagram


β
- Wα

@
@ hα

@
@R ?
W∞

induces the commutative diagram


T (hα
β)
T (Wβ ) - T (Wα )
@
@ T (hα )
@
T (hβ )

@
R ?
T (W∞ ) ,

and so there is induced an R-homomorphism


 
ωT : lim T (Wα ) −→ T lim Wα = T (W∞ ).
−→ −→
α∈Λ α∈Λ
60 The Mayer–Vietoris sequence

If, for all choices of such a directed set Λ and such a direct system (Wα )α∈Λ
over Λ, the map ωT is an isomorphism, then we shall say that T commutes with
the formation of direct limits or, more loosely, T commutes with direct limits.

In the next exercise, we again use the notation Ma , for an R-module M


and a ∈ R, to denote the module of fractions of M with respect to the multi-
plicatively closed subset ai : i ∈ N0 . In fact, we consider the functor ( • )a :
C(R) −→ C(R).

3.4.2 Exercise. Let a ∈ R. Show that the functor ( • )a : C(R) −→ C(R)


commutes with direct limits.

The next exercise gives another example, again needed later in the book, of
a functor which commutes with direct limits.

3.4.3 Exercise. Let (Λ, ≤) be a (non-empty) directed partially ordered set,


and let (Wα )α∈Λ be a direct system of R-modules over Λ, with constituent
R-homomorphisms hα β : Wβ → Wα (for each α, β ∈ Λ with α ≥ β). Set
W∞ := lim Wα .
−→
α∈Λ
 
(i) Show that lim ExtiR ( • , Wα ) can be made into a negative stron-
−→
α∈Λ i∈N0
gly connected sequence of contravariant functors from C(R) to C(R) in
such a way that the natural homomorphisms (of 3.4.1) give rise to a
homomorphism
 
 
Ψ = (ψ i )i∈N0 : lim ExtiR ( • , Wα ) −→ ExtiR ( • , W∞ ) i∈N
−→ 0
α∈Λ i∈N0

of connected sequences.
0
(ii) Prove that ψR is an isomorphism, and deduce that ψF0 is an isomorphism
whenever F is a finitely generated free R-module.
0
(iii) Deduce that ψM is an isomorphism whenever M is a finitely generated
R-module.
(iv) Prove, by induction on i, that, for each i ∈ N0 , the homomorphism ψM i

is an isomorphism whenever M is a finitely generated R-module.


(v) Deduce that, whenever M is a finitely generated R-module, the functor
ExtiR (M, • ) commutes with direct limits, for each i ∈ N0 .

Our main aim during the remainder of this chapter is to show that the local
cohomology functors Hai (i ∈ N0 ) commute with direct limits; we show first
that the a-torsion functor commutes with direct limits.
3.4 Direct limits 61

3.4.4 Proposition. The a-torsion functor Γa commutes with direct limits.


Proof. We use the notation of 3.4.1, so that (Λ, ≤) is a (non-empty) directed
partially ordered set, and (Wα )α∈Λ is a direct system of R-modules over Λ,
with constituent R-homomorphisms hα β : Wβ → Wα (for each (α, β) ∈ Λ×Λ
with α ≥ β). We must show that the R-homomorphism
 
ωΓa : lim Γa (Wα ) −→ Γa lim Wα
−→ −→
α∈Λ α∈Λ

is an isomorphism. We write F∞ := lim Γa (Wα ) and W∞ := lim Wα , and,


−→ −→
α∈Λ α∈Λ
for α ∈ Λ, we use fα : Γa (Wα ) → F∞ to denote the canonical map.
A typical element of Ker ωΓa can be expressed as fα (wα ) for some α ∈ Λ
and wα ∈ Γa (Wα ), where Γa (hα )(wα ) = 0 in Γa (W∞ ). Thus hα (wα ) =
0 in W∞ , and so there exists γ ∈ Λ with γ ≥ α such that hγα (wα ) = 0.
Since wα ∈ Γa (Wα ), this means that Γa (hγα )(wα ) = 0 in Γa (Wγ ), and so
fα (wα ) = fγ (Γa (hγα )(wα )) = 0. Hence ωΓa is injective.
We now show that ωΓa is surjective. Let y ∈ Γa (W∞ ). There exists α ∈ Λ
and wα ∈ Wα such that y = hα (wα ). We know that y is annihilated by a
power of a: let j ∈ N be such that aj y = 0. Let aj be generated by r1 , . . . , rn .
For each i = 1, . . . , n, we have hα (ri wα ) = 0, so that there exists βi ∈ Λ with
βi ≥ α such that hβαi (ri wα ) = 0. Since Λ is directed, there exists γ ∈ Λ such
that γ ≥ βi for all i = 1, . . . , n. We now have
ri hγα (wα ) = ri hγβi (hβαi (wα )) = hγβi (hβαi (ri wα )) = 0
for all i = 1, . . . , n, so that aj hγα (wα ) = 0 and hγα (wα ) ∈ Γa (Wγ ). But
y = hα (wα ) = hγ (hγα (wα )) = Γa (hγ )(hγα (wα )) ∈ Im(Γa (hγ )).
Set zγ := hγα (wα ), an element of Γa (Wγ ). Now fγ (zγ ) ∈ F∞ , and
ωΓa (fγ (zγ )) = Γa (hγ )(zγ ) = y.
Hence ωΓa is surjective and the proof is complete.
3.4.5 Exercise. Let T : C(R) → C(R) be a covariant, additive functor
which commutes with direct limits. Let (Lθ )θ∈Ω be a non-empty family of

R-modules; for each φ ∈ Ω, let qφ : Lφ −→ θ∈Ω Lθ be the canonical
injection.  
Prove that T θ∈Ω Lθ is the direct sum of its family of submodules
(Im T (qθ ))θ∈Ω , so that
 
  
T Lθ = ∼
Im T (qθ ) = T (Lθ ).
θ∈Ω θ∈Ω θ∈Ω
62 The Mayer–Vietoris sequence

We shall describe this result by the statement that T commutes with direct
sums.

To show that the i-th right derived functors of Γa (for i ∈ N) also commute
with direct limits, we plan to argue by induction on the number of elements
needed to generate a, and to use the Mayer–Vietoris sequence to complete the
inductive step. The following lemma will be helpful.

3.4.6 Lemma. Let Q, S, T, U, V be covariant functors (from C(R) to C(R))


and let σ : Q −→ S, τ : S −→ T, μ : T −→ U, ν : U −→ V be natural
transformations of functors such that, for each R-module M , the sequence
σ τ μM ν
Q(M ) −→
M
S(M ) −→
M
T (M ) −→ U (M ) −→
M
V (M )

is exact. Suppose that Q, S, U, V all commute with direct limits. Then T also
commutes with direct limits.

Proof. We again use the notation of 3.4.1. For each α, β ∈ Λ with α ≥ β,


there is a commutative diagram
σWβ τWβ μWβ νWβ
Q(Wβ ) - S(Wβ ) - T (Wβ ) - U (Wβ ) - V (Wβ )

Q(hα
β) S(hα
β) T (hα
β) U (hα
β) V (hα
β)

? ? ? ? ?
Q(Wα ) - S(Wα ) - T (Wα ) - U (Wα ) - V (Wα )
σWα τWα μWα νWα

with exact rows, and so there is induced an exact sequence


σ
∞ ∞ τ ∞ μ∞ ν
Q∞ −→ S∞ −→ T∞ −→ U∞ −→ V∞ ,

where Q∞ := lim Q(Wα ), etc., and σ∞ := lim σWα , etc. It is easy to see
−→ −→
α∈Λ α∈Λ
from the definition of ωQ , . . . , ωV in 3.4.1 that the diagram

σ∞ τ∞ μ∞ ν∞
Q∞ - S∞ - T∞ - U∞ - V∞

ωQ ωS ωT ωU ωV

? ? ? ? ?
Q(W∞ ) - S(W∞ ) - T (W∞ ) - U (W∞ ) - V (W∞ )
σW∞ τW∞ μW∞ νW ∞

commutes. Since this diagram has exact rows and ωQ , ωS , ωU and ωV are iso-
morphisms by hypothesis, it follows from the Five Lemma that ωT is an iso-
morphism too.
3.4 Direct limits 63

3.4.7 Remark. It is clear that the identity functor Id : C(R) → C(R) and the
zero functor 0 : C(R) → C(R) (for which 0(M ) = 0 for all R-modules M )
commute with direct limits.

3.4.8 Corollary. Suppose that, for some i ∈ N0 , the functors Hai and Hbi
commute with direct limits. Then Hai ( • ) ⊕ Hbi ( • ) commutes with direct limits.

Proof. This is an easy consequence of 3.4.6 and 3.4.7: for each R-module
M , consider the canonical split exact sequences
 - H i (M )  -H i (M ) ⊕ H i (M ) - H i (M )  -
0 a a b b 0.

3.4.9 Proposition. Let a ∈ R. Then the local cohomology functor HRa


1
com-
mutes with direct limits.

Proof. By 2.2.20, there are natural transformations of functors

τ : Id −→ ( • )a and σ : ( • )a −→ HRa
1

such that, for all R-modules M , the sequence


τ σ
M −→
M
Ma −→
M 1
HRa (M ) −→ 0 −→ 0

is exact. We propose to use 3.4.6, and with this in mind we observe that, by
3.4.2, the functor ( • )a : C(R) −→ C(R) commutes with direct limits. (This
will be obvious to any reader familiar with the facts that ( • )a is naturally
equivalent to ( • ) ⊗R Ra : C(R) −→ C(R) and that, for an arbitrary R-module
N , the functor ( • ) ⊗R N : C(R) −→ C(R) commutes with direct limits.
Alternatively, it is easy to prove directly.)
The claim now follows immediately from 3.4.6 and 3.4.7.

We are now able to prove that the local cohomology functors Hai (i ∈ N0 )
all commute with direct limits.

3.4.10 Theorem. For all i ∈ N0 , the local cohomology functor Hai commutes
with direct limits.

Proof. This proof is another example of the use of induction outlined in the
introduction to this chapter.
We suppose that a can be generated by t elements and proceed by induction
on t. When t = 0, we have a = 0 and Γa is the identity functor, so that Hai = 0
for all i ∈ N. The result is therefore immediate from 3.4.7 in this case.
We proved in 3.4.4 that Γa commutes with direct limits. When t = 1, it
follows from 3.3.1 that Hai = 0 for all i > 1, and from 3.4.9 that Ha1 commutes
with direct limits. The result is therefore also proved in the case when t = 1.
64 The Mayer–Vietoris sequence

Now suppose, inductively, that t > 1 and the result has been proved for ide-
als that can be generated by fewer than t elements. Suppose that a is generated
by t elements a1 , . . . , at . Set b = Ra1 + · · · + Rat−1 and c = Rat , so that
a = b + c. Note that
bc = (Ra1 + · · · + Rat−1 )Rat = Ra1 at + · · · + Rat−1 at
can be generated by t − 1 elements, and that Hb∩c i
= Hbci
for all i ∈ N0 by
3.1.5.
It therefore follows from the inductive hypothesis and what we have already
proved that the local cohomology functors Hbi (i ∈ N0 ), Hci (i ∈ N0 ) and
i
Hb∩c (i ∈ N0 ) all commute with direct limits. Hence, by 3.4.8, all the functors
Hai ( • ) ⊕ Hbi ( • ) (i ∈ N0 ) commute with direct limits.
It now follows from 3.4.6 and the Mayer–Vietoris sequence (we use here
i
the full statement of 3.2.3) that the functors Hb+c (i ∈ N0 ) all commute with
direct limits. Since a = b + c, the inductive step is complete.
3.4.11 Corollary. The a-transform functor Da commutes with direct limits.
Consequently, by 3.4.5, the functor Da commutes with direct sums (in the
sense of Exercise 3.4.5).
Proof. By 2.2.6(i), there are natural transformations of functors
ξ : Γa −→ Id, η : Id −→ Da , ζ 0 : Da −→ Ha1
such that, for each R-module M , the sequence
ξM ηM ζ0
Γa (M ) −→ M −→ Da (M ) −→
M
Ha1 (M ) −→ 0
is exact. Since Γa and Ha1 commute with direct limits (by 3.4.4 and 3.4.10
respectively), the claim follows from 3.4.6.
3.4.12 Exercise. Suppose that, for some i ∈ N0 , we have Hai (R) = 0. Show
that Hai (P ) = 0 for every projective R-module P .
3.4.13 Exercise. Let K  denote the full ring of fractions of R; that is, K  :=
S −1 R, where S is the set of all non-zerodivisors of R. Show that, if, for some
integer i ∈ N0 , we have Hai (R) = 0, then Hai (K  ) = 0 too.
Thus, in particular, if R is an integral domain with field of fractions K and
Ha (R) = 0 for some i ∈ N0 , then Hai (K) = 0.
i
4

Change of rings

The main results of this chapter concern a homomorphism of commutative


Noetherian rings f : R −→ R . More precisely, we shall prove two funda-
mental comparison results for local cohomology modules in this context. The
first of these, which we shall call the ‘Independence Theorem’, compares, for
an R -module M  and an i ∈ N0 , the local cohomology modules Hai (M  )
i  
and HaR  (M ): to form the first of these, we consider M as an R-module
by restriction of scalars using f ; also, aR denotes the extension of a to R


under f . Our second main result, which we shall refer to as the ‘Flat Base
Change Theorem’, compares the local cohomology modules Hai (M ) ⊗R R

 (M ⊗R R ) for i ∈ N0 and an arbitrary R-module M under the
i
and HaR
additional assumption that the ring homomorphism f is flat.
Our main results rely on the fact that certain modules are acyclic with respect
to torsion functors. We say that an R-module A is Γa -acyclic precisely when
Hai (A) = 0 for all i > 0. As was explained in 1.2.2, the most basic method
for calculation, for an R-module M and an i ∈ N0 , of Hai (M ) is to take an
injective resolution I • of M , apply Γa to I • to obtain the complex Γa (I • ),
and take the i-th cohomology module of this complex: we have Hai (M ) =
H i (Γa (I • )). However, it is an easy exercise to show that a resolution of M by
Γa -acyclic R-modules will serve this purpose just as well.
Of course, injective R-modules are Γa -acyclic, but in general the class of
Γa -acyclic R-modules is larger than the class of injective R-modules. Let f :
R −→ R be a homomorphism of commutative Noetherian rings. We shall
show that, if I  is an injective R -module, then, when viewed as an R-module
by restriction of scalars, I  is Γa -acyclic. We shall also show that, if the ring
homomorphism f is flat, then, for each injective R-module I, the R -module
I ⊗R R is ΓaR -acyclic. These results pave the way for the proofs later in
the chapter of the Independence Theorem (4.2.1) and the Flat Base Change
Theorem (4.3.2).
66 Change of rings

4.1 Some acyclic modules


For completeness, we begin with the formal definition.
4.1.1 Definition. We say that an R-module A is Γa -acyclic precisely when
Hai (A) = 0 for all i > 0. If B is a system of ideals (of R) over Λ as in 2.1.10,
i
then we say that A is ΓB -acyclic if and only if HB (A) = 0 for all i > 0.
Our presentation does not depend on the following exercise, but the reader
should be aware of the result it contains.
4.1.2 Exercise. Show that, given an R-module M , its local cohomology mo-
dules with respect to a can be calculated by means of a resolution of M by
Γa -acyclic modules as follows. Let
d−1 d0 di
A• : 0 −→ A0 −→ A1 −→ · · · −→ Ai −→ Ai+1 −→ · · ·
be a Γa -acyclic resolution of M , so that A0 , A1 , . . . , Ai , . . . are all Γa -acyclic
R-modules and there is an R-homomorphism α : M → A0 such that the
sequence
α d0 di
0 −→ M −→ A0 −→ A1 −→ · · · −→ Ai −→ Ai+1 −→ · · ·
is exact. Show that
Hai (M ) ∼
= Ker(Γa (di ))/ Im(Γa (di−1 )) for all i ∈ N0 .
We shall have two applications of our first proposition in this section.
4.1.3 Proposition. Let M be an R-module, and let C be a set of ideals of R
such that
(a) C is closed under the formation of finite sums and products,
(b) 0R ∈ C, and
(c) each ideal in C is the sum of finitely many principal ideals which belong
to C.
Assume that Hc1 (M ) = 0 for all c ∈ C. Then M is Γc -acyclic for all c ∈ C.
Remark. Of course, if M is Γc -acyclic for all c ∈ C, then it is automatic that
Hc1 (M ) = 0 for all c ∈ C. Thus this theorem could be phrased as an ‘if and
only if’ result.
Proof. Since Γ0R is the identity functor on C(R), it is clear that M is Γ0R -
acyclic. Also, it was proved in 2.2.21(ii) that, for a ∈ R, we have HRa
i
(M ) = 0
for all i ∈ N with i > 1; hence M is Γc -acyclic whenever c is a principal ideal
in C.
4.1 Some acyclic modules 67

Now suppose, inductively, that t > 1 and we have proved that, whenever

c ∈ C can be expressed as a sum of fewer than t principal ideals which all
belong to C, then M is Γc -acyclic. This is certainly the case when t = 2. Let
c = Rc1 + · · · + Rct , where c1 , . . . , ct ∈ R and Rc1 , . . . , Rct ∈ C. We shall
use the Mayer–Vietoris sequence to show that M is also Γc -acyclic.
Set a = Rc1 + · · · + Rct−1 and b = Rct , so that c = a + b. Note that b ∈ C,
and, by hypothesis (a) on C, each of a, ab, Rc1 ct , . . . , Rct−1 ct belongs to C.
Since b is principal, and a and
ab = (Rc1 + · · · + Rct−1 )Rct = Rc1 ct + · · · + Rct−1 ct
are both expressible as sums of t − 1 principal ideals belonging to C, it follows
from 2.2.21(ii) and the inductive hypothesis that M is Γa -acyclic, Γb -acyclic
and Γab -acyclic.
By the Mayer–Vietoris sequence 3.2.3, we have, for an arbitrary i > 1,
an exact sequence Ha∩b i−1
(M ) −→ Hci (M ) −→ Hai (M ) ⊕ Hbi (M ). Now
i−1 i−1
Ha∩b (M ) = Hab (M ) by 3.1.5, and so, by what we have already established,
i−1
Ha∩b (M ) = Hai (M ) = Hbi (M ) = 0. Thus Hci (M ) = 0 for all i > 1. Since
Hc1 (M ) = 0 by hypothesis (because c ∈ C), this completes the inductive step.
Since each ideal in C can be expressed as the sum of finitely many principal
ideals which belong to C, the proof is complete.
The two important applications of Proposition 4.1.3 which we have in mind
concern a homomorphism of commutative Noetherian rings, and it is appro-
priate for us to clarify notation at this point.
4.1.4 Notation. Throughout this chapter, R will denote a second commuta-
tive Noetherian ring and f : R −→ R will denote a ring homomorphism. As
in Chapter 2, for an ideal b of R, we shall use bR to denote the extension of b
to R under f .
For an R-module M , we shall use ZdvR (M ) (or Zdv(M ) if there is no am-
biguity about the underlying ring involved) to denote the set of elements of R
which are zerodivisors on M . Thus, for example, if M  is an R -module, then
R \ ZdvR (M  ) denotes the set of elements of R which are non-zerodivisors
on M  when the latter is regarded as an R-module by means of f .
4.1.5 Lemma. Let M  be a finitely generated R -module that is aR -torsion-
free. Then a ⊆ ZdvR (M  ); in other words, a contains a non-zerodivisor on
M  when the latter is regarded as an R-module by means of f .
Proof. Suppose that a ∩ (R \ ZdvR (M  )) = ∅ and look for a contradiction.

Then f (a) ⊆ ZdvR (M  ) = P∈AssR M  P. Since M  is a finitely generated
R -module, it follows from the Prime Avoidance Theorem in the form given in
68 Change of rings

[81, 3.61] that f (a) ⊆ P for some P ∈ AssR M  . Since there exists m ∈ M 
with (0 :R m ) = P, we deduce that 0 = m ∈ (0 :M  aR ). This is a
contradiction.

We are now ready for the proof of one of the main results of this section.

4.1.6 Theorem. Let I  be an injective R -module. Then I  , when viewed as


an R-module by means of f , is Γa -acyclic.

Proof. The strategy of this proof is to show first that Ha1 (I  ) = 0, and then
appeal to Proposition 4.1.3.
By 2.2.24, there is an R-isomorphism εI  : DaR (I  ) −→ Da (I  ) such that
the diagram
ηaR ,I 
I - DaR (I  )


= εI 

ηa,I 
?
I - Da (I  )

1 
commutes. Since HaR  (I ) = 0, it follows from 2.2.6(i)(c) that ηaR ,I  is sur-

jective. Hence ηa,I  is surjective, and so Ha1 (I  ) = 0, again by 2.2.6(i)(c).


It follows that Ha1 (I  ) = 0 for an arbitrary ideal a of R. We can now apply
4.1.3 (with the set of all ideals of R as the set C) to deduce that I  is Γb -acyclic
for every ideal b of R.

We now move on to consider the situation in which the ring homomorphism


f : R → R is flat, that is, such that R , when viewed as an R-module via f , is
flat. We need two preliminary results for this situation; for the first, we appeal
to [50] for a proof.

4.1.7 Lemma. (See [50, Theorem 7.11].) Assume that the ring homomor-
phism f : R → R is flat. Let M be an R-module. Then there is a natural
transformation of functors

μ : HomR ( • , M ) ⊗R R −→ HomR (( • ) ⊗R R , M ⊗R R )

(from C(R) to C(R )) which is such that

(i) μN (g ⊗ r ) = r (g ⊗ IdR ) for each R-module N , each r ∈ R and


each g ∈ HomR (N, M ), and
(ii) μN is an isomorphism if N is a finitely generated R-module. 
4.1 Some acyclic modules 69

4.1.8 Proposition. Assume that the ring homomorphism f : R → R is flat;


let I be an injective R-module. Then the natural R -homomorphism
β : I ⊗R R −→ HomR (aR , I ⊗R R )
(for which β(m )(r ) = r m for all r ∈ aR and all m ∈ I ⊗R R ) is
surjective.
Proof. Since f is flat, the natural R -epimorphism γ : a ⊗R R −→ aR (for
t t
which γ( i=1 ai ⊗ ri ) = i=1 f (ai )ri for all t ∈ N, a1 , . . . , at ∈ a and
r1 , . . . , rt ∈ R ) is an isomorphism (see [50, Theorem 7.7]).
Since a is a finitely generated R-module, the R -homomorphism
μa : HomR (a, I) ⊗R R −→ HomR (a ⊗R R , I ⊗R R )
of 4.1.7 is an isomorphism. Since I is R-injective, the R-homomorphism σ :
I → HomR (a, I) (for which σ(m)(a) = am for all a ∈ a and all m ∈ I) is
surjective, so that, since tensor product is right exact,
σ ⊗ IdR : I ⊗R R −→ HomR (a, I) ⊗R R
is an R -epimorphism. It follows that the composition
HomR (γ −1 , IdI⊗R R ) ◦ μa ◦ (σ ⊗ IdR ) : I ⊗R R −→ HomR (aR , I ⊗R R )
is surjective. Since this composition is just β, the proof is complete.
We are now in a position to prove a theorem which has some similarities to
Theorem 4.1.6.
4.1.9 Theorem. Assume that the ring homomorphism f : R → R is flat,
and let I be an injective R-module. Then I ⊗R R is ΓaR -acyclic.
Proof. We shall employ a strategy similar to that which we used in the proof

 (I ⊗R R ) = 0, and then we shall appeal
1
of 4.1.6: we shall show first that HaR
to Proposition 4.1.3.
For each n ∈ N, the natural R -homomorphism
I ⊗R R −→ HomR (an R , I ⊗R R )
of 4.1.8 is surjective, so that, since an R = (aR )n , the R -homomorphism
ηI⊗R R : I ⊗R R −→ DaR (I ⊗R R ) = lim HomR ((aR )n , I ⊗R R )
−→
n∈N

 (I ⊗R R ) = 0, by 2.2.6(i)(c).
1
of 2.2.6(i) is surjective. Hence HaR
We can now apply 4.1.3 (with the set C taken as the set of all ideals of R
which are extended from ideals of R) to deduce that I ⊗R R is ΓbR -acyclic
for every ideal b of R.
70 Change of rings

Theorems 4.1.6 and 4.1.9 pave the way for us to present, in the next two
sections, the two main results of this chapter.

4.2 The Independence Theorem


In this section, we prove the Independence Theorem, which shows, loosely
speaking, that local cohomology is ‘independent of the base ring’. We have
already seen results of a similar type for ideal transforms in 2.2.24, and for first
local cohomology modules in Exercise 2.2.26. Let us again use R : C(R ) →
C(R) to denote the functor obtained from restriction of scalars (using f ): thus,
if M  is an R -module and i ∈ N0 , we can form the R-modules HaR i 
 (M ) R

and Ha (M R ). For the first, we form the i-th local cohomology module of M 
i 

with respect to aR and then regard the resulting R -module as an R-module
by means of f ; for the second, we first consider M  as an R-module via f , and
then take the i-th local cohomology module with respect to a. Theorem 4.2.1
below shows, among other things, that there is an R-isomorphism
  ∼
−→ Hai (M  =: Hai (M  ),
i i =
HaR  (M ) = HaR (M ) R R)

so that, speaking loosely, it does not matter whether we calculate these local
cohomology modules over R or R .

4.2.1 Independence Theorem. The functors ΓaR ( • ) R and Γa ( • R) (from


C(R ) to C(R)) are the same, and there is a unique isomorphism
 i  = 
∼ 
Λ = (λi )i∈N0 : HaR ( • ) R
i∈N0
−→ Hai ( • R ) i∈N0

of negative (strongly) connected sequences of covariant functors (from C(R )


to C(R)) such that λ0 is the identity natural equivalence. In particular, for each
i ∈ N0 , the functors HaR
i i
 ( • ) R and Ha ( • R ) are naturally equivalent.

Proof. Since R : C(R ) → C(R) is covariant and exact, it is clear that


i
(HaR ( • ) R )i∈N0 and (Hai ( • R ))i∈N0

are negative strongly connected sequences of covariant functors from C(R ) to


C(R).
Now an R = (aR )n for all n ∈ N, and so ΓaR ( • ) R and Γa ( • R ) are
the same functor. Furthermore, whenever I  is an injective R -module, it is,

 (I ) = 0 for all i ∈ N, while it follows from
i
of course, automatic that HaR
i 
4.1.6 that Ha (I ) = 0 for all i ∈ N. The result now follows immediately on
application of 1.3.4(ii).
4.2 The Independence Theorem 71

Remark. Some readers might prefer to approach the final claim of 4.2.1 in
the following way. Observe first that ΓaR ( • ) R and Γa ( • R ) are the same
functor. Let M  be an R -module, and let

I • : 0 −→ I 0 −→ I 1 −→ · · · −→ I i −→ I i+1 −→ · · ·

be an injective resolution of M  over R . By 1.2.2, the i-th cohomology module


of the complex ΓaR (I • ) is isomorphic to HaRi 
 (M ) (for i ∈ N0 ).

By 4.1.6, the complex

I • R : 0 −→ I 0 R −→ I 1 R −→ · · · −→ I i R −→ I i+1 R −→ · · ·

is a resolution of M  R by Γa -acyclic R-modules, and so, by 4.1.2, the i-th


cohomology module of the complex Γa (I • R ) is isomorphic to Hai (M  R )
(for i ∈ N0 ). Since ΓaR ( • ) R = Γa ( • R ), this gives a quick and relatively
transparent proof that HaRi  ∼ i
 (M ) R = Ha (M

R ) for i ∈ N0 . Note, however,
that our use of 1.3.4 in the above proof of 4.2.1 led rather rapidly to additional
information.

4.2.2 Example. Suppose that the R-module M is annihilated by the ideal b


of R. Then, since M is b-torsion, it is immediate from Exercise 2.1.9 that
Hai (M ) ∼
= Ha+bi
(M ) for all i ∈ N0 . However, the reader might find the
following alternative approach, which uses the Independence Theorem 4.2.1,
illuminating.
Since M is annihilated by b, we can regard M as a module over R/b in a
natural way. Let i ∈ N0 . By the Independence Theorem 4.2.1, there is an R-
isomorphism Hai (M ) ∼ i
= Ha(R/b) (M ). But a(R/b) = (a + b)/b, and another
use of the Independence Theorem 4.2.1 provides us with an R-isomorphism
i
H(a+b)/b (M ) ∼ i
= Ha+b (M ), so that Hai (M ) ∼ i
= Ha+b (M ), as claimed.

4.2.3 Remark. In 3.3.4 we defined the cohomological dimension cohd(a)


of a. It is immediate from the Independence Theorem 4.2.1 that cohd(aR ) ≤
cohd(a).

4.2.4 Exercise. Let n ∈ N and consider the ring R[X1 , . . . , Xn ] of polyno-


mials over R.

1
(i) Show that the R[X1 ]-module H(X 1)
(R[X1 ]) is not finitely generated,
and therefore non-zero. (Here is a hint: you might find 2.2.21 helpful.)
(ii) Use 1.3.9(iv) to show that, when n > 1,
n−1
H(X 1 ,...,Xn )
(R[X1 , . . . , Xn ]) = 0.
72 Change of rings

(iii) The ring homomorphism ν : R[X1 , . . . , Xn ] → R[X1 , . . . , Xn−1 ]


(again for n > 1) obtained by evaluation at X1 , . . . , Xn−1 , 0 allows
us to regard R[X1 , . . . , Xn−1 ] as an R[X1 , . . . , Xn ]-module. Show that,
when this is done,
n
H(X 1 ,...,Xn )
(R[X1 , . . . , Xn−1 ]) = 0.

(iv) Assume again that n > 1. Deduce from the exact sequence
X
0 −→ R[X1 , . . . , Xn ] −→
n
R[X1 , . . . , Xn ]
ν
−→ R[X1 , . . . , Xn−1 ] −→ 0
n
of R[X1 , . . . , Xn ]-modules that H(X 1 ,...,Xn )
(R[X1 , . . . , Xn ]) has a sub-
n−1
module isomorphic to H(X1 ,...,Xn−1 ) (R[X1 , . . . , Xn−1 ]) (the latter be-
ing regarded as an R[X1 , . . . , Xn ]-module by means of ν), and that
n n
H(X 1 ,...,Xn )
(R[X1 , . . . , Xn ]) = Xn H(X 1 ,...,Xn )
(R[X1 , . . . , Xn ]).

(v) Prove by induction that, for all n ∈ N, the R[X1 , . . . , Xn ]-module


n
H(X 1 ,...,Xn )
(R[X1 , . . . , Xn ]) is not finitely generated, and so is non-
zero.

4.2.5 Exercise. Use 4.2.4 to show that, over the ring R[X1 , X2 , X3 , X4 ] of
polynomials in four indeterminates with coefficients in R,
3
H(X 1 ,X2 )∩(X3 ,X4 )
(R[X1 , X2 , X3 , X4 ]) = 0.

Deduce that the affine algebraic set W in A4 defined by

W := VA4 (X1 , X2 ) ∪ VA4 (X3 , X4 )

cannot be ‘defined by two equations’, that is, there do not exist two polynomi-
als f, g ∈ C[X1 , X2 , X3 , X4 ] such that W = p ∈ A4 : f (p) = g(p) = 0 .

We are now going to use the Independence Theorem to prove a proposi-


tion which presents a non-vanishing result for certain local cohomology mod-
ules. We shall use this proposition in Chapter 6 to prove a more general non-
vanishing result. The strategy of the proof of the proposition has some similar-
ities with the strategy used in Exercise 4.2.4.

4.2.6 Proposition. Suppose that (R, m) is a regular local ring of dimension


d
d > 0. Then the R-module Hm (R) is not finitely generated, and therefore
non-zero.
4.2 The Independence Theorem 73

Proof. We argue by induction on d. When d = 1, the maximal ideal m is


generated by one element, π say. Let K denote the field of fractions of R.
Apply 2.2.18 (in the case where B = (Rπ n )n∈N , M = R and S is the set of

non-zero elements of R), and note that n∈N (R :K π n ) = K: it follows from
this and 2.2.6(i)(c) that there is an exact sequence
0 −→ R −→ K −→ HRπ
1
(R) −→ 0.
Since R is finitely generated as an R-module but K is not, we deduce that
1
HRπ (R) is not finitely generated.
Now suppose, inductively, that d > 1 and that the result has been proved
for regular local rings of smaller (positive) dimensions. Let u1 , . . . , ud be d
elements of m which generate this maximal ideal. Let R := R/ud R, a regular
local ring of dimension d − 1 with maximal ideal m := m/ud R. By the induc-
d−1
tive hypothesis, the R-module Hm (R) is not finitely generated. Hence, by
d−1
the Independence Theorem 4.2.1, the R-module Hm (R/ud R) is not finitely
generated. The exact sequence
u
0 −→ R −→
d
R −→ R/ud R −→ 0
induces an exact sequence Hm d−1
(R) −→ Hm d−1
(R/ud R) −→ Hm d
(R). But
d−1 d
Hm (R) = 0 by 1.3.9(iv) since u1 , . . . , ud is an R-sequence. Hence Hm (R)
d−1
has a submodule isomorphic to Hm (R/ud R); since the latter is not finitely
d
generated, neither is Hm (R).
4.2.7 Exercise. Assume that R contains a field K as a subring. Suppose
that a can be generated by t elements a1 , . . . , at . The ring homomorphism
f : K[X1 , . . . , Xt ] −→ R obtained by evaluation at a1 , . . . , at allows us to
regard each R-module as a module over K[X1 , . . . , Xt ].
Show that, for an R-module M , there is a K[X1 , . . . , Xt ]-isomorphism
i
H(X 1 ,...,Xt )
(M ) ∼
= Hai (M ) for each i ∈ N0 .
Use the fact that the ring K[X1 , . . . , Xt ] has global dimension t to show that
Hai (M ) = 0 for all i > t. Compare this with Theorem 3.3.1.
4.2.8 Exercise. (In this exercise, there is no assumption about R beyond
the standard ones.) Suppose that a can be generated by t elements. Use an
appropriate ring homomorphism Z[X1 , . . . , Xt ] −→ R in conjunction with
the Independence Theorem 4.2.1 and the fact that the global dimension of
Z[X1 , . . . , Xt ] is t + 1 to show that Hai (M ) = 0 for all i > t + 1. Com-
pare this with Theorem 3.3.1.
The following refinement of the Independence Theorem 4.2.1 is occasion-
ally useful.
74 Change of rings

4.2.9 Exercise. Let the situation be as in the Independence Theorem 4.2.1.


Let M  be an R -module. Show that, for each i ∈ N0 , the R-module Hai (M  )
actually has a natural structure as an R -module under which

r x = Hai (r IdM  )(x) for all r ∈ R and x ∈ Hai (M  ).

Show further that, if f  : M  −→ N  is a homomorphism of R -modules, then


Hai (f  ) : Hai (M  ) −→ Hai (N  ) is an R -homomorphism.
Deduce that (Hai (• R ))i∈N0 is a negative strongly connected sequence of co-
variant functors from C(R ) to itself, and, as such, is isomorphic to (HaR
i
 )i∈N0 .

4.3 The Flat Base Change Theorem


The second fundamental result about local cohomology that we propose to
derive from the work in §4.1 is the Flat Base Change Theorem. This result is
a consequence of Theorem 4.1.9 and concerns the situation in which f : R →
R is flat: we are going to show that, in this situation, speaking loosely, the
formation of local cohomology ‘commutes’ with ‘extension of the base ring
from R to R ’. More precisely, we shall show that, for each R-module M and

each i ∈ N0 , there is an R -isomorphism HaR  
=
 (M ⊗R R ) −→ Ha (M )⊗R R .
i i

Although our proof of this result has some similarities with our proof of 4.2.1
above, a preparatory lemma will be helpful in this case.

4.3.1 Lemma. Assume that the ring homomorphism f : R → R is flat.


There is a natural equivalence of functors (from C(R) to C(R ))

ρ0 : Γa ( • ) ⊗R R −→ ΓaR (( • ) ⊗R R )

which is such that, for each R-module M , we have ρ0M (m ⊗ r ) = m ⊗ r for


all m ∈ Γa (M ) and all r ∈ R .

Proof. Let M be an R-module. Since f is flat, the inclusion map induces an


R -monomorphism Γa (M ) ⊗R R → M ⊗R R whose image is contained
in ΓaR (M ⊗R R ). It follows easily that there is a natural transformation of
functors ρ0 : Γa ( • ) ⊗R R −→ ΓaR (( • ) ⊗R R ) which is such that, for each
R-module M , ρ0M is monomorphic and satisfies
t t
ρ0M i=1 mi ⊗ ri = i=1 mi ⊗ ri

for all t ∈ N, m1 , . . . , mt ∈ Γa (M ) and r1 , . . . , rt ∈ R . It remains to show


that ρ0M is surjective for each R-module M .
4.3 The Flat Base Change Theorem 75

We can deduce this from 4.1.7 as follows. Let n ∈ N. Since R/an is a


finitely generated R-module, 4.1.7 provides us with an R -isomorphism

μR/an : HomR (R/an , M ) ⊗R R −→ HomR ((R/an ) ⊗R R , M ⊗R R )
=

such that μR/an (h ⊗ r ) = r (h ⊗ IdR ) for all h ∈ HomR (R/an , M ) and all
r ∈ R . We can incorporate μ := μR/an into the composition

(0 :M an ) ⊗R R −→ HomR (R/an , M ) ⊗R R
=

μ
−→ HomR ((R/an ) ⊗R R , M ⊗R R )

−→ HomR (R /an R , M ⊗R R )
=


−→ (0 :M ⊗R R (aR )n )
=

of R -isomorphisms, in which the other three isomorphisms are the obvious


natural ones. The reader can easily check that this composition ε is such that
t t
ε( i=1 mi ⊗ ri ) = i=1 mi ⊗ ri (for t ∈ N, m1 , . . . , mt ∈ (0 :M an ) and
r1 , . . . , rt ∈ R ). It is now clear that ρ0M is surjective.

4.3.2 Flat Base Change Theorem. Assume that the ring homomorphism f :
R −→ R is flat. There is a unique isomorphism

(ρi )i∈N0 : (Hai ( • ) ⊗R R )i∈N0 −→ (HaR 
=
 (( • ) ⊗R R ))i∈N0
i

of negative (strongly) connected sequences of covariant functors (from C(R)


to C(R )) which extends the natural equivalence ρ0 of 4.3.1. In particular, for
each i ∈ N0 , the functors Hai ( • ) ⊗R R and HaR
i 
 (( • ) ⊗R R ) are naturally

equivalent.

Proof. As in the proof of 4.2.1, we use 1.3.4. Since ( • ) ⊗R R : C(R) −→


C(R ) is exact, it is clear that the two sequences given in the statement of
4.3.2 are indeed negative strongly connected sequences of covariant functors
from C(R) to C(R ). Furthermore, whenever I is an injective R-module, it
is, of course, automatic that Hai (I) = 0 for all i ∈ N, while it follows from

 (I ⊗R R ) = 0 for all i ∈ N. The result is now an immediate
i
4.1.9 that HaR
consequence of 4.3.1 and 1.3.4(ii).

Remark. Again, some readers might prefer the following direct approach to
the final claim of 4.3.2. Let M be an R-module, and let

I • : 0 −→ I 0 −→ I 1 −→ · · · −→ I i −→ I i+1 −→ · · ·

be an injective resolution of M over R. For i ∈ N0 , the i-th cohomology


76 Change of rings

module of the complex Γa (I • ) is, of course, Hai (M ). Since ( • ) ⊗R R :


C(R) −→ C(R ) is additive and exact, we have an R -isomorphism

Hai (M ) ⊗R R = H i (Γa (I • )) ⊗R R ∼
= H i (Γa (I • ) ⊗R R ).

In view of the natural equivalence of 4.3.1, there is an R -isomorphism

H i (Γa (I • ) ⊗R R ) ∼
= H i (ΓaR (I • ⊗R R )).

But, by 4.1.9, the complex I • ⊗R R is a ΓaR -acyclic resolution of the R -


module M ⊗R R , and so, by 4.1.2, there is an R -isomorphism

H i (ΓaR (I • ⊗R R )) ∼ i
= HaR 
 (M ⊗R R ).

This gives us, fairly directly, an R -isomorphism

Hai (M ) ⊗R R ∼ i
= HaR 
 (M ⊗R R ).

Note again, however, that our use of 1.3.4 in the above proof of 4.3.2 led very
quickly to additional information.
If S is a multiplicatively closed subset of R, then, as is very well known, the
natural ring homomorphism R → S −1 R is flat. Thus we can obtain the result
of Exercise 1.3.6 as a special case of the Flat Base Change Theorem 4.3.2.

4.3.3 Corollary. Let S be a multiplicatively closed subset of R. Then

S −1 (Hai ( • )) and i
HaS −1 R (S
−1
( • ))
i∈N0 i∈N0

are isomorphic connected sequences of functors (from C(R) to C(S −1 R)).


In particular, for every i ∈ N0 and every R-module M , there is an
S −1 R-isomorphism S −1 (Hai (M )) ∼ i
= HaS −1 R (S
−1
M ). 

4.3.4 Exercise. Suppose that R is regular; that is, Rp is a regular local ring
for every p ∈ Spec(R).
Suppose that a is proper and has height t > 0. Show that the R-module
Hat (R) is not finitely generated, and therefore non-zero.

The first part of the next exercise is for those readers who have reached this
point without needing the fact that, for an arbitrary R-module N , the functor
( • ) ⊗R N : C(R) −→ C(R) commutes with direct limits!

4.3.5 Exercise. Let N be an R-module.

(i) Show that the functor ( • ) ⊗R N : C(R) −→ C(R) commutes with direct
limits (see 3.4.1).
4.3 The Flat Base Change Theorem 77

(ii) Assume that the ring homomorphism f : R → R is flat. Show that there
is a natural equivalence of functors
ε : Da ( • ) ⊗R R −→ DaR (( • ) ⊗R R )
(from C(R) to C(R )) which is such that, for each R-module M , the
diagram
ηa,M ⊗R R
M ⊗R R  - Da (M ) ⊗R R

εM

ηaR ,M ⊗ 
?
M ⊗R R 
RR
- DaR (M ⊗R R )

commutes. (The notations ηa,M and ηaR ,M ⊗R R are explained in the


note following the statement of 2.2.6.)
(iii) Deduce that, for a multiplicatively closed subset S of R, the functors
S −1 (Da ( • )) and DaS −1 R (S −1 ( • ))
(from C(R) to C(S −1 R)) are naturally equivalent.
We shall end this chapter with a geometric example and exercise; the fol-
lowing remark will be useful.
4.3.6 Remark. Let r1 , r2 ∈ R. Then
{p ∈ Spec(R) : p ⊇ r1 r2 R + a}
   
= p ∈ Spec(R) : p ⊇ (r1 R + a) ∩ (r2 R + a) ,
  
and so (r1 r2 R + a) = (r1 R + a) ∩ (r2 R + a).
4.3.7 Example. This example concerns the affine variety V in A4 studied in
2.3.7 and 3.3.5 and given by
V := VA4 (X1 X4 − X2 X3 , X12 X3 + X1 X2 − X22 , X33 + X3 X4 − X42 ).
In 3.3.5 we showed that the subvariety
L := VA4 (X2 , X3 , X4 ) = (c, 0, 0, 0) ∈ A4 : c ∈ C
of V (a line) cannot be ‘defined by one equation’; here, we shall show that V
itself cannot be ‘defined by two equations’.
Write, for convenience, R := C[X1 , X2 , X3 , X4 ], and let c be the ideal of
R given by
c = (X1 X4 − X2 X3 , X12 X3 + X1 X2 − X22 , X33 + X3 X4 − X42 ).
78 Change of rings

We wish to show that ara(c) > 2, and we shall achieve this by showing that
Hc3 (R) = 0 and then appealing to 3.3.3. Let R := C[[X1 , X2 , X3 , X4 ]]; since

the natural ring homomorphism R → R is flat, it follows from the Flat Base
Change Theorem 4.3.2 that Hc3 (R) ⊗R R ∼  and so it will be enough,
= Hc3R (R),
in order to achieve our aim, for us to show that H 3 (R) = 0. This is what we

cR
shall do.
 be given by
Let u ∈ R

(−1)n−1 2(2n − 2)! n


u := 1 + 2X3 − 2X32 + 4X33 − 10X34 + · · · + X3 + · · · .
n!(n − 1)!

 (by [81, 1.43]),


In fact, u2 = 1 + 4X3 ; note that, since the latter is a unit in R
so too is u. Write

f1 := X1 X4 −X2 X3 , f2 := X12 X3 +X1 X2 −X22 , f3 := X33 +X3 X4 −X42 ,

 = (f1 , f2 , f3 )R.
so that cR  Now

4f2 = 4X12 X3 + X12 − (4X22 − 4X1 X2 + X12 )


= X12 u2 − (2X2 − X1 )2
= (X1 u + 2X2 − X1 )(X1 u − 2X2 + X1 )
= (X1 (u − 1) + 2X2 )(X1 (u + 1) − 2X2 ).

A similar calculation shows that

4f3 = 4X33 + X32 − (4X42 − 4X3 X4 + X32 )


= (X3 (u − 1) + 2X4 )(X3 (u + 1) − 2X4 ).

Set
g1 := X1 (u − 1) + 2X2 , g2 := X1 (u + 1) − 2X2 ,

h1 := X3 (u − 1) + 2X4 , h2 := X3 (u + 1) − 2X4 .

Note that

h1 g2 − h2 g1 = X1 (u + 1)2X4 − X3 (u − 1)2X2
− 2X2 X3 (u + 1) + 2X4 X1 (u − 1)
= 4u(X1 X4 − X2 X3 ) = 4uf1 ,

so that f1 = u−1 (h1 g2 − h2 g1 )/4.


4.3 The Flat Base Change Theorem 79

Therefore, on use of 4.3.6, we have


  
cR = (f1 , f2 , f3 )R
 = (f1 , g1 g2 , h1 h2 )R

 
 ∩ (f1 , g2 , h1 h2 )R
= (f1 , g1 , h1 h2 )R 
√ √ √ √
= d11 ∩ d12 ∩ d21 ∩ d22

where dij = (f1 , gi , hj )R  for i, j = 1, 2. We now use the equation f1 =


−1
u (h1 g2 − h2 g1 )/4 to see that d11 = (f1 , g1 , h1 )R  = (g1 , h1 )R
 and d22 =
 
(f1 , g2 , h2 )R = (g2 , h2 )R; the same equation and 4.3.6 show that


d12 = (f1 , g1 , h2 )R 
  
= (h1 g2 , g1 , h2 )R = (h1 , g1 , h2 )R
 ∩ (g2 , g1 , h2 )R


and
  

d21 =  = (h2 , g2 , h1 )R
(f1 , g2 , h1 )R  ∩ (g1 , g2 , h1 )R.


Now put all this information together to see that


  
 = (g1 , h1 )R
cR  ∩ (g2 , h2 )R.


 the four elements g1 , g2 , h1 , h2 gen-


Next, note that, since u is a unit of R,
 of the 4-dimensional regular local
erate the maximal ideal (X1 , X2 , X3 , X4 )R
 and so, by [81, 15.38] for example, p := (g1 , h1 )R
ring R,  and q := (g2 , h2 )R

are prime ideals of R. Thus we have the interesting situation where, although
√ 
c is, by 2.3.7, a prime ideal of R, the ideal cR  of R is the intersection of
two distinct prime ideals of height 2.
Note that p + q = m  the maximal ideal of R.
 := (X1 , X2 , X3 , X4 )R, 

Since X1 , X2 , X3 , X4 is an R-sequence in m, it follows from 1.3.9(iv) that
3
Hp+q  = 0. However, Proposition 4.2.6 shows that H 4 (R)
(R)  = 0. By
p+q
i  i 
3.3.1, we have Hp (R) ⊕ Hq (R) = 0 for all i > 2. It therefore follows from
4.3.2, 1.2.3 and the Mayer–Vietoris sequence 3.2.3 that
∼
Hc3 (R) ⊗R R  = Hp∩q
= Hc3R (R) 3  ∼
(R) 4 
= Hm
 (R) = 0.

Hence Hc3 (R) = 0, and so ara(c) > 2 by 3.3.3. Therefore V cannot be ‘de-
fined by two equations’.

4.3.8 Exercise. Let K be a field of characteristic 0, and let R denote the ring
K[X1 , X2 , X3 , X4 ]; let c be the ideal of R given by

c = (X12 + X13 − X22 , X32 + X1 X32 − X42 , X2 X3 − X1 X4 ).


80 Change of rings
 := K[[X1 , X2 , X3 , X4 ]] and let u ∈ R
Let R  be given by

1 1 1 5 (−1)n−1 2(2n − 2)! n


u := 1+ X1 − X12 + X13 − X14 +· · ·+ X1 +· · · ,
2 8 16 128 4n n!(n − 1)!
so that u2 = 1 + X1 . Let z1 := X1 u − X2 , z2 := X3 u − X4 , z3 := X1 u + X2
and z4 := X3 u + X4 .

(i) Show that z1 , z2 , z3 , z4 generate the maximal ideal of R.

(ii) Show that, in R,
X2 X3 − X1 X4 = X1 z2 − X3 z1 = X3 z3 − X1 z4
and 2u(X2 X3 − X1 X4 ) = z2 z3 − z1 z4 . Deduce that

cR  = (z1 , z2 )R
 ∩ (z3 , z4 )R.


(iii) Prove that Hc3 (R) = 0.


4.3.9 Exercise. With the notation of 2.3.1, let V be the affine algebraic set in
A4 given by
V := VA4 (X12 + X13 − X22 , X32 + X1 X32 − X42 , X2 X3 − X1 X4 ).
(i) By considering the morphism of varieties α : A2 → A4 for which
α((c, d)) = (c2 − 1, c3 − c, d, cd) for all (c, d) ∈ A2 and the mapping
β : V \ {(0, 0, 0, 0)} −→ A2 \ {(1, 0), (−1, 0)}
given by

(c2 /c1 , c3 ) if c1 = 0,
β((c1 , c2 , c3 , c4 )) =
(c4 /c3 , c3 ) if c3 = 0
(for all (c1 , c2 , c3 , c4 ) ∈ V \ {(0, 0, 0, 0)}), show that V is irreducible,
and so is an affine variety.
(ii) Deduce from Exercise 4.3.8 above that V cannot be ‘defined by two
equations’.
5

Other approaches

Although we have now developed enough of the basic algebraic theory of lo-
cal cohomology so that we could, if we wished, start right away with serious
calculations with local cohomology modules, there are two other approaches
to the construction of local cohomology modules which are useful, and pop-
ular with many workers in the subject. One approach uses cohomology of
C̆ech complexes, and the other uses direct limits of homology modules of
Koszul complexes. Links between local cohomology and Koszul complexes
and C̆ech cohomology are described in A. Grothendieck’s foundational lecture
notes [25, §2]; related ideas are present in J.-P. Serre’s fundamental paper [77,
§61]. Among other texts which discuss links between local cohomology and
the C̆ech complex or Koszul complexes are those by W. Bruns and J. Herzog
[7, §3.5], D. Eisenbud [10, Appendix 4], M. Herrmann, S. Ikeda and U. Or-
banz [32, §35], P. Roberts [70, Chapter 3, §2], J. R. Strooker [83, §4.3] and J.
Stückrad and W. Vogel [84, Chapter 0, §1.3].
We shall make very little use in this book of the descriptions of local coho-
mology modules as direct limits of homology modules of Koszul complexes.
However, we will use the approach to local cohomology via cohomology of
C̆ech complexes, and so we present the basic ideas of this approach in this
chapter. As this work leads naturally to the connection between local coho-
mology and direct limits of homology modules of Koszul complexes, we also
present some aspects of that connection.
The C̆ech complex approach is particularly useful for calculations in Han (R)
when a can be generated by n elements. We shall illustrate this in §5.3 in the
case where R has prime characteristic p. Then each local cohomology module
Hai (R) (i ∈ N0 ) of R itself carries a natural ‘Frobenius action’, that is, an
Abelian group homomorphism F : Hai (R) −→ Hai (R) such that F (rm) =
rp F (m) for all m ∈ Hai (R) and r ∈ R. The Frobenius action on Han (R),
82 Other approaches

where a can be generated by n elements, can be described very simply using


the C̆ech complex.
d
Later in the book, we shall often study aspects of Hm (R) where (R, m) is
local and of dimension d. Let a1 , . . . , ad be a system of parameters for R, and

set q := Ra1 + · · · + Rad . Since q = m, we have Hm d
(R) = Hqd (R), and
so, in the case when R has prime characteristic p, the above-mentioned C̆ech
complex approach facilitates calculations with the natural Frobenius action on
Hmd
(R). There is an illustration of this idea in §6.5.

5.1 Use of C̆ech complexes


Throughout this chapter, a1 , . . . , an (where n > 0) will denote n elements
which generate a, and M will denote an arbitrary R-module. Our first task
is to define the (extended) C̆ech complex C(M )• of M with respect to the
sequence a1 , . . . , an . This has the form
d0 dn−1
0 −→ C(M )0 −→ C(M )1 −→ · · · −→ C(M )n−1 −→ C(M )n −→ 0.
The following lemma will be helpful. Recall again that, for a ∈ R, the nota-
tions Ra and Ma denote, respectively, the ring and module of fractions of R
and M with respect to the multiplicatively closed subset ai : i ∈ N0 of R.
5.1.1 Lemma. Let a, b ∈ R. There is an isomorphism of R-modules
μ : Mab −→ (Ma )b
for which μ(m/(ab)i ) = (m/ai )/bi for all m ∈ M and all i ∈ N0 .
Proof. Suppose that m, y ∈ M and i, j ∈ N0 are such that m/(ab)i =
y/(ab)j in Mab . Then there exists k ∈ N0 such that
 
(ab)k (ab)j m − (ab)i y = 0.
Thus, in Ma , we have bk+j m/ai = bk+i y/aj , so that (m/ai )/bi = (y/aj )/bj
in (Ma )b . Therefore there is indeed a mapping μ : Mab −→ (Ma )b given by
the formula in the statement of the lemma. It is now a very easy exercise to
check that μ is an R-isomorphism.
5.1.2 Exercise. Complete the proof of Lemma 5.1.1.
5.1.3 Remark. In situations such as that of 5.1.1, we shall use μ to iden-
tify the R-modules Mab and (Ma )b without further comment. Thus, when we
speak of the natural R-homomorphism ω : Ma −→ Mab , we shall be referring
to the natural R-homomorphism from Ma to (Ma )b and employing the above
5.1 Use of C̆ech complexes 83

identification, so that ω(m/ai ) = bi m/(ab)i for all m ∈ M and i ∈ N0 . These


ideas are employed in the construction of the C̆ech complex of M with respect
to a1 , . . . , an in 5.1.5 below.
5.1.4 Notation. For k ∈ N with 1 ≤ k ≤ n, we shall write
I(k, n) := (i(1), . . . , i(k)) ∈ Nk : 1 ≤ i(1) < i(2) < . . . < i(k) ≤ n ,
the set of all strictly increasing sequences of length k of positive integers taken
from the set {1, . . . , n}. For i ∈ I(k, n), we shall, for 1 ≤ j ≤ k, denote the
j-th component of i by i(j), so that i = (i(1), . . . , i(k)).
Now suppose that k < n, and s ∈ N with 1 ≤ s ≤ k+1. Let j ∈ I(k+1, n).
Then by
 . . . , j(k + 1))
j s or (j(1), . . . , j(s),
we mean the sequence (j(1), . . . , j(s − 1), j(s + 1), . . . , j(k + 1)) of I(k, n)
obtained by omitting the s-th component of j.
It is perhaps worth pointing out here that, if t ∈ N with 1 ≤ t < s, then
(j s)t = (j t )
s−1
.
Again under the assumption that k < n, let i ∈ I(k, n). By the n-comple-
ment of i we mean the sequence j ∈ I(n − k, n) such that
{1, . . . , n} = {i(1), . . . , i(k), j(1), . . . , j(n − k)} .
5.1.5 Proposition and Definition. Define a sequence C(M )• of R-modules
and R-homomorphisms
d0 dn−1
0 −→ C(M )0 −→ C(M )1 −→ · · · −→ C(M )n−1 −→ C(M )n −→ 0
as follows:
(a) C(M )0 := M ;
(b) for k = 1, . . . , n, and with the notation of 5.1.4,

C(M )k = Mai(1) ...ai(k) ;
i∈I(k,n)

(c) d0 : C(M )0 −→ C(M )1 is to be such that, for each h = 1, . . . , n,


the composition of d0 followed by the canonical projection from
C(M )1 to Mah is just the natural map from M to Mah ; and
(d) for k = 1, . . . , n − 1, i ∈ I(k, n) and j ∈ I(k + 1, n), the compo-
sition
dk
Mai(1) ...ai(k) −→ C(M )k −→ C(M )k+1 −→ Maj(1) ...aj(k+1)
(in which the first and third maps are the canonical injection and
84 Other approaches

canonical projection respectively) is to be the natural map from


Mai(1) ...ai(k) to Mai(1) ...ai(k) aj(s) multiplied by (−1)s−1 if i = j s
for an s with 1 ≤ s ≤ k + 1, and is to be 0 otherwise.

Then C(M )• is a complex, called the (extended) C̆ech complex of M with


respect to a1 , . . . , an . (Henceforth, we shall omit the word ‘extended’.)
We denote C(R)• by
d0 di
C • : 0 −→ C 0 −→ C 1 −→ · · · −→ C i −→ C i+1 −→ · · · −→ C n −→ 0.

Proof. We have only to prove that dk+1 ◦ dk = 0 for all k = 0, . . . , n − 2.


Let m ∈ M . To show that d1 ◦ d0 = 0, it is enough to show that, for each
i ∈ I(2, n), the component of d1 ◦ d0 (m) in the direct summand Mai(1) ai(2) of
C(M )2 is 0. The only contributions to this component that could conceivably
be non-zero must come ‘through’ the direct summands Mai(1) and Mai(2) of
C(M )1 . It follows that the component of d1 ◦ d0 (m) in the direct summand
Mai(1) ai(2) of C(M )2 is (−1)(m/1) + (m/1) = 0. Hence d1 ◦ d0 = 0.
Now consider the case where 1 ≤ k ≤ n−2. In order to show dk+1 ◦dk = 0,
it is enough to show that, for each i ∈ I(k, n), the restriction of dk+1 ◦ dk to
the direct summand Mai(1) ...ai(k) of C(M )k is zero, and this is what we shall
do. So let m ∈ M, v ∈ N0 and j ∈ I(k + 2, n): we calculate the component
of
 
m
d k+1
◦d k
(ai(1) . . . ai(k) )v

in the direct summand Maj(1) ...aj(k+2) of C(M )k+2 . This component will be
(conceivably) non-zero only if i = (j s)t for some integers s, t with 1 ≤ t <
s ≤ k + 2, and, when this is so, will (in view of the penultimate paragraph of
5.1.4) be
(−1)t−1 (−1)s−2 avj(t) avj(s) m (−1)s−1 (−1)t−1 avj(s) avj(t) m
+ ,
(aj(1) . . . aj(k+2) )v (aj(1) . . . aj(k+2) )v

which is zero. Hence dk+1 ◦ dk = 0, and the proof is complete.

5.1.6 Example. The reader might find it helpful if we write down explicitly
the C̆ech complex C • = C(R)• of R with respect to a1 , . . . , an when n has a
fairly small value: when n = 3, the complex is
d0 d1 d2
0 → R −→ Ra1 ⊕ Ra2 ⊕ Ra3 −→ Ra2 a3 ⊕ Ra1 a3 ⊕ Ra1 a2 −→ Ra1 a2 a3 → 0

where the di (i = 0, 1, 2) are described as follows. For r, r1 , r2 , r3 ∈ R and


5.1 Use of C̆ech complexes 85

v 1 , v 2 , v 3 ∈ N0 ,
r r r
d0 (r) = , , ,
1 1 1
 
r1 r2 r3
d1 , ,
av11 av22 av33
 v3 
a2 r3 av32 r2 av13 r3 av31 r1 av12 r2 av21 r1
= − , − , −
(a2 a3 )v3 (a2 a3 )v2 (a1 a3 )v3 (a1 a3 )v1 (a1 a2 )v2 (a1 a2 )v1
and
 
r1 r2 r3
d2 , ,
(a2 a3 )v1 (a1 a3 )v2 (a1 a2 )v3
av11 r1 av22 r2 av33 r3
= v
− v
+ .
(a1 a2 a3 ) 1 (a1 a2 a3 ) 2 (a1 a2 a3 )v3
5.1.7 Exercise. Suppose that n > 1 and M is Ran -torsion. Show that the
C̆ech complex of M with respect to a1 , . . . , an is the C̆ech complex of M with
respect to a1 , . . . , an−1 .
5.1.8 Exercise. Show that a homomorphism f : M → N of R-modules
induces a chain map of complexes C(f )• : C(M )• −→ C(N )• such that
C(f )0 : C(M )0 → C(N )0 is just f : M → N .
Show further that, with these assignments, C( • )• becomes a functor from
the category C(R) to the category of complexes of R-modules (and R-homom-
orphisms) and chain maps of such complexes.
We remind the reader that we use the notation
d0 di
C • : 0 −→ C 0 −→ C 1 −→ · · · −→ C i −→ C i+1 −→ · · · −→ C n −→ 0
for the C̆ech complex of R itself with respect to a1 , . . . , an . We interpret C i as
0 for i ∈ Z \ {0, 1, . . . , n}, of course.
It should be clear to the reader that H i (( • ) ⊗R C • ) is, for each i ∈ N0 , a
covariant R-linear functor from C(R) to itself.
5.1.9 Lemma. The sequence (H i (( • ) ⊗R C • ))i∈N0 is a negative strongly
connected sequence of covariant functors from C(R) to itself.
Proof. Let
α β
0 - L - M - N - 0

λ μ ν

? ? ?
α β
0 - L - M - N - 0
86 Other approaches

be a commutative diagram of R-modules and R-homomorphisms with exact


rows. Now for each i ∈ N, the R-module C i is flat, since it is a direct sum of
finitely many modules of fractions of the form S −1 R for suitable choices of
the multiplicatively closed subset S of R. It follows that there is a commutative
diagram
α⊗C • β⊗C •
0 - L ⊗R C • - M ⊗R C • - N ⊗R C • - 0

λ⊗C • μ⊗C • ν⊗C •

? ? ?
α ⊗C • β  ⊗C •
0 - L ⊗ R C • - M  ⊗R C • - N  ⊗R C • - 0

of complexes of R-modules and chain maps of such complexes such that, for
each i ∈ N0 , the sequence
α⊗C i β⊗C i
0 - L ⊗R C i - M ⊗R C i - N ⊗R C i - 0

is exact, and a similar property holds for the lower row. Thus the above commu-
tative diagram of complexes gives rise to a long exact sequence of cohomology
modules of the complexes in the top row, a similar long exact sequence for the
bottom row, and a chain map of the first long exact sequence into the second.
The claim follows from this.

The technique used in the above proof will solve the following exercise.

5.1.10 Exercise. Show that (H i (C( • )• ))i∈N0 is a negative strongly con-


nected sequence of covariant functors from C(R) to itself.

5.1.11 Exercise. Use the natural isomorphisms



=
M ⊗R Rai(1) ...ai(k) −→ Mai(1) ...ai(k) (for i ∈ I(k, n))

(where 1 ≤ k ≤ n) to produce an isomorphism of complexes



=

ωM : M ⊗R C • −→ C(M )• .

Show further that, as M varies through C(R), the ωM



constitute a natural
equivalence of functors ω • : ( • ) ⊗R C • −→ C( • )• (from C(R) to the
category of all complexes of R-modules (and R-homomorphisms) and chain
maps of such complexes).
Show also that ω • induces an isomorphism

=
(H i (ω • ))i∈N0 : (H i (( • ) ⊗R C • ))i∈N0 −→ (H i (C( • )• ))i∈N0
5.1 Use of C̆ech complexes 87

of negative strongly connected sequences of covariant functors (from C(R) to


C(R)).

5.1.12 Exercise. In this and the next few exercises, we shall be concerned
with relationships between C̆ech complexes of M with respect to different se-
quences of elements of R, and, for such discussions, it will help if we use a
slightly more complicated notation. Thus, in this and a few subsequent situa-
tions, we shall use C(a1 , . . . , an ; M )• to denote the C̆ech complex of M with
respect to a1 , . . . , an .

(i) Suppose that n ≥ 2, and let m ∈ N be such that 1 ≤ m < n. Let


b1 , . . . , bn denote the sequence

a1 , . . . , am−1 , am+1 , am , am+2 , . . . , an

obtained from the sequence a1 , . . . , an by interchange of the m-th and


(m + 1)-th terms.
Let k ∈ N with 1 ≤ k ≤ n. Let i ∈ I(k, n).
If neither m nor m + 1 appears as a term in i, then

Mai(1) ...ai(k) = Mbi(1) ...bi(k) :

define ψ(i) : Mai(1) ...ai(k) −→ Mbi(1) ...bi(k) to be the identity mapping


in this case.
If m + 1 does not appear as a term in i but m = i(s), then

Mai(1) ...ai(k) = Mbi(1) ...bi(s−1) bi(s)+1 bi(s+1) ...bi(k) :

define ψ(i) : Mai(1) ...ai(k) −→ Mbi(1) ...bi(s−1) bi(s)+1 bi(s+1) ...bi(k) to be


the identity mapping in this case.
If m does not appear as a term in i but m + 1 = i(s), then

Mai(1) ...ai(k) = Mbi(1) ...bi(s−1) bi(s)−1 bi(s+1) ...bi(k) :

define ψ(i) : Mai(1) ...ai(k) −→ Mbi(1) ...bi(s−1) bi(s)−1 bi(s+1) ...bi(k) to be


the identity mapping in this case.
If m and m + 1 both appear in i, with, say, m = i(s) (so that m + 1 =
i(s) + 1 = i(s + 1)), then Mai(1) ...ai(k) = Mbi(1) ...bi(k) : in this case,
define ψ(i) : Mai(1) ...ai(k) −→ Mbi(1) ...bi(k) to be the identity mapping
multiplied by −1.
Let

ψk = ψ(i) : C(a1 , . . . , an ; M )k −→ C(b1 , . . . , bn ; M )k .
i∈I(k,n)
88 Other approaches

Also, let ψ 0 : M −→ M be the identity mapping. Show that


 
Ψ = ψ k 0≤k≤n : C(a1 , . . . , an ; M )• −→ C(b1 , . . . , bn ; M )•

is an isomorphism of complexes.
(ii) Let σ be a permutation of the set {1, . . . , n}. Show that there is an iso-
morphism of complexes

=
C(a1 , . . . , an ; M )• −→ C(aσ(1) , . . . , aσ(n) ; M )• .

5.1.13 Definition. Let m ∈ Z, and let


dk dk+1
Q• : · · · - Qk Q•
- Qk+1 Q• - Qk+2 - ···

be a complex of R-modules and R-homomorphisms. We define {m}Q• , the


result of shifting Q• by m places, as follows: ({m}Q• )k = Qm+k for all k ∈
Z, and the k-th ‘differentiation’ homomorphism dk{m}Q• in {m}Q• is given
by dk{m}Q• = dm+k
Q• : Qm+k −→ Qm+k+1 . Thus changing Q• into {1}Q•
(respectively {−1}Q• ) amounts to shifting it one place to the left (respectively
right).

5.1.14 Exercise. In this exercise, we use the extended notation for C̆ech
complexes introduced in Exercise 5.1.12.
Let b ∈ R. Show that there is a sequence of complexes (of R-modules and
R-homomorphisms) and chain maps

0 −→ {−1}C(a1 , . . . , an ; Mb )• −→ C(a1 , . . . , an , b; M )•
−→ C(a1 , . . . , an ; M )• −→ 0

which is such that, for each i ∈ N0 , the sequence

0 −→ ({−1}C(a1 , . . . , an ; Mb )• )i −→ C(a1 , . . . , an , b; M )i
−→ C(a1 , . . . , an ; M )i −→ 0

is exact.

5.1.15 Exercise. In this exercise, we use the extended notation for C̆ech com-
plexes introduced in Exercise 5.1.12.
Let a ∈ R. Show that H i (C(a; M )• ) ∼
= HRai
(M ) for all i ∈ N0 .

5.1.16 Exercise. Show that all the cohomology modules of the C̆ech com-
plex C(M )• are a-torsion. (Here are some hints: try induction on the number
of elements needed to generate a, in conjunction with 5.1.14 and 5.1.12.)
5.1 Use of C̆ech complexes 89

5.1.17 Exercise. In this exercise, we again use the extended notation for C̆ech
complexes introduced in Exercise 5.1.12.
Suppose that b1 , . . . , bm also generate a. Prove that

H i (C(a1 , . . . , an ; M )• ) ∼
= H i (C(b1 , . . . , bm ; M )• ) for all i ∈ N0 ,

that is, the cohomology modules of the C̆ech complex C(M )• are, up to R-
isomorphism, independent of the choice of sequence of generators for a. (We
suggest that you use 5.1.14 to compare

C(a1 , . . . , an , b1 ; M )• and C(a1 , . . . , an ; M )• .

Also, you may find 5.1.16 and 5.1.12 helpful.)

Our first major aim in this chapter is to show that, for each i ∈ N0 , the local
cohomology module Hai (M ) is isomorphic to H i (C(M )• ), the i-th cohomol-
ogy module of the C̆ech complex C(M )• . The next remark is a first step.

5.1.18 Remark. In the C̆ech complex


d0
C(M )• : 0 −→ M −→ Ma1 ⊕ Ma2 ⊕ · · · ⊕ Man −→ . . . ,

we have

H 0 (C(M )• ) = Ker d0
= {m ∈ M : for each i = 1, . . . , n, there exists

hi ∈ N with ahi i m = 0
= Γa (M ).

It will probably come as no surprise to the reader to learn that we intend


to exploit the above Remark 5.1.18 by use of negative strongly connected
sequences of functors and 1.3.5.

5.1.19 Remark. It is immediate from 5.1.18 and 5.1.11 that there is an R-



=
isomorphism γM 0
: Γa (M ) −→ H 0 (M ⊗R C • ) for which γM 0
(m) = m ⊗ 1
for all m ∈ Γa (M ), and it is then clear that, as M varies through C(R), the
γM0
constitute a natural equivalence of functors γ 0 : Γa −→ H 0 (( • ) ⊗R C • )
from C(R) to itself.

5.1.20 Theorem. (Recall that a1 , . . . , an (where n > 0) denote n elements


which generate a, and C • denotes the C̆ech complex of R with respect to
a1 , . . . , an .) There is a unique isomorphism

=
(δ i )i∈N0 : (H i (C( • )• ))i∈N0 −→ (Hai )i∈N0
90 Other approaches

of negative (strongly) connected sequences of covariant functors (from C(R)


to C(R)) which extends the identity natural equivalence on Γa .
Consequently, there is a unique isomorphism

=
(γ i )i∈N0 : (Hai )i∈N0 −→ (H i (( • ) ⊗R C • ))i∈N0
of negative strongly connected sequences of covariant functors (from C(R) to
C(R)) which extends the natural equivalence γ 0 of 5.1.19.
Proof. Recall from 5.1.18 that H 0 (C( • )• ) = Γa . We shall be able to use
Theorem 1.3.5 to prove the first part provided we can show that H i (C(I)• ) =
0 for all i ∈ N whenever I is an injective R-module. We shall achieve this
by induction on n, and we shall use the extended notation for C̆ech complexes
introduced in Exercise 5.1.12.
τI
When n = 1, the C̆ech complex C(a1 ; I)• is just 0 −→ I −→ Ia1 −→ 0,
where τI is the natural map. But, by 2.2.20, Coker τI ∼ = HRa1
1
(I), and the
latter is zero because I is injective. Thus H i (C(I)• ) = 0 for all i ∈ N. We
note also that Ker τI = ΓRa1 (I), which is injective by 2.1.4. Hence the exact
τI
sequence 0 −→ ΓRa1 (I) −→ I −→ Ia1 −→ 0 of 2.2.20 splits, and Ia1 is an
injective R-module.
Now suppose, inductively, that n > 1 and the result has been proved for
smaller values of n. By 5.1.14, there is a sequence of complexes (of R-modules
and R-homomorphisms) and chain maps
0 −→ {−1}C(a1 , . . . , an−1 ; Ian )• −→ C(a1 , . . . , an ; I)•
−→ C(a1 , . . . , an−1 ; I)• −→ 0
which is such that, for each i ∈ N0 , the sequence
0 −→ ({−1}C(a1 , . . . , an−1 ; Ian )• )i −→ C(a1 , . . . , an ; I)i
−→ C(a1 , . . . , an−1 ; I)i −→ 0
is exact. This sequence of complexes therefore induces a long exact sequence
of cohomology modules. Since Ian is an injective R-module (by the imme-
diately preceding paragraph of this proof), we can deduce from our inductive
hypothesis that
H i ({−1}C(a1 , . . . , an−1 ; Ian )• ) = 0 for all i ≥ 2
and
H i (C(a1 , . . . , an−1 ; I)• ) = 0 for all i ≥ 1.
Furthermore, an easy check shows that the connecting homomorphism
H 0 (C(a1 , . . . , an−1 ; I)• ) −→ H 1 ({−1}C(a1 , . . . , an−1 ; Ian )• )
5.1 Use of C̆ech complexes 91

induced by the above sequence of complexes is just the map


ΓRa1 +···+Ran−1 (I) −→ ΓRa1 +···+Ran−1 (Ian )
induced by the natural homomorphism I → Ian , and this is surjective since
the canonical exact sequence
0 −→ ΓRan (I) −→ I −→ Ian −→ 0
of 2.2.20 splits. It follows from the long exact sequence of cohomology mod-
ules that H i (C(a1 , . . . , an ; I)• ) = 0 for all i ≥ 1. This completes the induc-
tive step. We can now use Theorem 1.3.5 to complete the proof of the first
part.
For the second part, we recall the isomorphism

=
(H i (ω • ))i∈N0 : (H i (( • ) ⊗R C • ))i∈N0 −→ (H i (C( • )• ))i∈N0
of negative strongly connected sequences of 5.1.11, and deduce that
 i • −1  i −1 ∼
=
(H (ω ))i∈N0 ◦ (δ )i∈N0 : (Hai )i∈N0 −→ (H i (( • ) ⊗R C • ))i∈N0
is an isomorphism of connected sequences; moreover, (H 0 (ω • ))−1 ◦ (δ 0 )−1 is
just the natural equivalence γ 0 of 5.1.19, as is easy to check. The uniqueness
in the second part follows from 1.3.4.
5.1.21 Remark. The reader should note that it is immediate from Theorem
5.1.20 that, when a can be generated by n elements, Hai (N ) = 0 for all R-
modules N whenever i > n: we proved this result by means of the Mayer–
Vietoris sequence in Theorem 3.3.1. This is one example of a situation where
the use of a different approach to the calculation of local cohomology modules
can provide a simpler proof and additional insight. The next exercise provides
another example.
5.1.22 Exercise. Prove that Han (R) = 0 if and only if there exists k ∈ N
such that, for every t ∈ N, it is the case that
(a1 . . . an )t ∈ Rat+k
1 + · · · + Rat+k
n .

Deduce that, for h ∈ N, in the ring R[X1 , . . . , Xh ] of polynomials over R, we


h
have H(X 1 ,...,Xh )
(R[X1 , . . . , Xh ]) = 0.
Compare this approach with that of Exercise 4.2.4.
5.1.23 Proposition. Let K 1 (M ) := Ker d1 , where d1 : C(M )1 −→ C(M )2
is the first ‘differentiation’ map in the C̆ech complex C(M )• of M with respect
to the sequence a1 , . . . , an . (It is clear from 5.1.8 that K 1 ( • ) can easily be
made into a functor from C(R) to itself.)
92 Other approaches

There is a natural equivalence of functors ε̃ : K 1 ( • ) −→ Da which is such


that, for each R-module M , the diagram
d0
M - K 1 (M )

ε̃M

ηa,M
?
M - Da (M )

commutes.

Proof. By 5.1.16, the cohomology modules H i (C(M )• ) are all a-torsion,


and so both the kernel and cokernel of d0 : M −→ K 1 (M ) are a-torsion. It
therefore follows from 2.2.15 that there is a unique R-homomorphism ε̃M :
K 1 (M ) −→ Da (M ) such that the diagram

d0
M - K 1 (M )

ε̃M

ηa,M
?
M - Da (M )

commutes, and that ε̃M = Da (d0 )−1 ◦ ηa,K 1 (M ) . This formula and 2.2.6(i)(c)
show that ε̃M is injective, since Γa (K 1 (M )) ⊆ Γa (C 1 (M )) = 0. Further-
more, it is easy to use the uniqueness aspect of 2.2.13 to deduce that, as M
varies through the category C(R), the ε̃M constitute a natural transformation
of functors.
Let t ∈ N, and let h ∈ HomR (at , M ). It is straightforward to check that the
element
 
h(at1 ) h(at2 ) h(atn )
, , . . . , ∈ C(M )1
at1 at2 atn

actually belongs to K 1 (M ). (Note that ati h(atj ) = h(ati atj ) = atj h(ati ) for
integers i, j with 1 ≤ i, j ≤ n.) Hence there is an R-homomorphism γt,M :
HomR (at , M ) → K 1 (M ) for which
 
h(at1 ) h(at )
γt,M (h) = t , . . . , tn for all h ∈ HomR (at , M ).
a1 an
5.1 Use of C̆ech complexes 93

Also, it is straightforward to check that, for t, u ∈ N with u ≥ t, the diagram

HomR (at , M )
@
HomR (jtu ,M ) @ γt,M
@
? R
@
HomR (au , M ) - K 1 (M )
γu,M

(in which jtu : au → at denotes the inclusion map) commutes, and so the
γt,M (t ∈ N) induce an R-homomorphism γM : Da (M ) −→ K 1 (M ) for
which γM ◦ ηa,M = d0 . Note that ε̃M ◦ γM = IdDa (M ) by the uniqueness
aspect of 2.2.15. Hence ε̃M is surjective, and so is an isomorphism.

5.1.24 Exercise. Consider the special case of 5.1.23 in which M = R. Note


n
that C 1 = C(R)1 = i=1 Rai has a natural structure as a (commutative
Noetherian) ring (with identity), and that d0 : R → C 1 is a ring homomor-
phism. Recall also from 2.2.5 that Da (R) has a structure as a commutative
ring with identity.
Show that K 1 (R) is a subring of C 1 , and deduce from 2.2.17 that ε̃R :
K (R) −→ Da (R) (where ε̃R is as in 5.1.23) is actually a ring isomorphism.
1

The next exercise concerns the ring K 1 (R) in a geometrical situation.

5.1.25 Exercise. Let V be an affine variety over the algebraically closed field
K. Consider C 1 and K 1 (R) (of 5.1.24) in the special case in which R =
O(V ), and assume that the ideal a of O(V ) is non-zero (and generated by
a1 , . . . , an , all of which are assumed to be non-zero). Let U be the open subset
V \ {p ∈ V : a1 (p) = · · · = an (p) = 0} of V .
For each j = 1, . . . , n, let Uj denote the open subset {p ∈ V : aj (p) = 0}
of U , identify the subring O(V )aj of K(V ) with the ring O(Uj ) of regular
functions on Uj in the natural way, and let ιj : O(U ) −→ O(Uj ) be the
restriction homomorphism.
n
Show that the map λ : O(U ) −→ j=1 O(Uj ) for which

λ(g) = (ι1 (g), . . . , ιn (g)) for all g ∈ O(U )

is an injective ring homomorphism with image K 1 (O(V )). In this way, we



obtain a ring isomorphism λ : O(U ) −→ K 1 (O(V )). Show that, with the
=
94 Other approaches

notation of 5.1.24, the diagram


O(V )
U
- O(U )
@
ηa,O(V )
@ ∼
= ε̃O(V ) ◦λ
@
@
R ?
Da (O(V ))

commutes, and deduce from 2.3.2 that ε̃O(V ) ◦ λ is the νV,a of that theorem.

5.2 Use of Koszul complexes


We revert to the general situation and remind the reader that, throughout this
chapter, a1 , . . . , an (where n > 0) denote n elements which generate a, and
M denotes an arbitrary R-module.
There is yet another method of calculation of local cohomology modules:
this describes them as direct limits of homology modules of Koszul complexes.
We present this description here, because it can be derived quickly from our
work so far in this chapter on the C̆ech complex. We need to specify our nota-
tion for the Koszul complexes that we shall use.
5.2.1 Notation. For all u ∈ N, we denote by K(au )• (or K(au1 , . . . , aun )• )
the usual Koszul complex of R with respect to au1 , . . . , aun . Thus, if F denotes
the free R-module Rn and, for each i = 1, . . . , n, the element
(0, . . . , 0, 1, 0, . . . , 0) ∈ Rn
which has i-th component 1 and all its other components 0 is denoted by ei ,
then K(au )• has the form
d(au )k
0 - K(au )n - ··· - K(au )k - K(au )k−1 -

··· - K(au )0 - 0,

where

k 
k
K(au )k = F = (Rn ) for k = 0, . . . , n

(so that K(au )0 = R) and, for k = 1, . . . , n and i ∈ I(k, n) (with the notation
of 5.1.4),
k
d(au )k (ei(1) ∧ . . . ∧ ei(k) ) = (−1)h−1 aui(h) ei(1) ∧ . . . ∧ e
i(h) ∧ . . . ∧ ei(k) ,
h=1
5.2 Use of Koszul complexes 95

where the ‘e


i(h) ’ indicates that ei(h) is omitted.
Of course, we set K(au )k = 0 for all k ∈ Z \ {0, 1, . . . , n}.
5.2.2 Lemma. Let u, v ∈ N with u ≤ v. There is a chain map
(ψuv )• = ((ψuv )k )k∈Z : K(au )• −→ K(av )•
of complexes of R-modules and R-homomorphisms such that (ψuv )n is the
!n
identity mapping of F , such that (ψuv )0 is the endomorphism of R given
by multiplication by (a1 . . . an )v−u , and such that, for k = 1, . . . , n − 1 and
i ∈ I(k, n),
(ψuv )k (ei(1) ∧ . . . ∧ ei(k) ) = (aj(1) . . . aj(n−k) )v−u ei(1) ∧ . . . ∧ ei(k) ,
where j ∈ I(n − k, n) is the n-complement of i (see 5.1.4).
Proof. We must check that d(av )k ◦ (ψuv )k = (ψuv )k−1 ◦ d(au )k for each
k = 1, . . . , n. We leave this to the reader in the case in which k = n. For
1 ≤ k < n and i ∈ I(k, n), j ∈ I(n − k, n) as in the statement of the lemma,
we have

(d(av )k ◦ (ψuv )k )(ei(1) ∧ . . . ∧ ei(k) )


k
= (−1)h−1 avi(h) (aj(1) . . . aj(n−k) )v−u ei(1) ∧ . . . ∧ e
i(h) ∧ . . . ∧ ei(k) ,
h=1

and

((ψuv )k−1 ◦ d(au )k )(ei(1) ∧ . . . ∧ ei(k) )

 k

= (ψuv )k−1 (−1)h−1 aui(h) ei(1) ∧ . . . ∧ e
i(h) ∧ . . . ∧ ei(k)
h=1
k
= (−1)h−1 (aj(1) . . . aj(n−k) )v−u avi(h) ei(1) ∧ . . . ∧ e
i(h) ∧ . . . ∧ ei(k)
h=1
= (d(av )k ◦ (ψuv )k )(ei(1) ∧ . . . ∧ ei(k) ).
The result follows.
5.2.3 Exercise. Complete the proof of 5.2.2. In other words, show that, with
the notation of the lemma, d(av )n ◦ (ψuv )n = (ψuv )n−1 ◦ d(au )n .
5.2.4 Remark. It is clear that, in the notation of 5.2.2, for u, v, w ∈ N with
u ≤ v ≤ w, we have (ψvw )• ◦ (ψuv )• = (ψuw )• , and that (ψuu )• is the identity
chain map of the complex K(au )• to itself. Thus the (ψuv )• turn the family
96 Other approaches

(K(au )• )u∈N into a direct system of complexes of R-modules and chain maps.
Our immediate aim is to show that the direct limit complex, which we denote
by K(a∞ )• , is isomorphic to a shift of the C̆ech complex C • of 5.1.5.

5.2.5 Theorem. There is an isomorphism of complexes



K(a∞ )• = lim K(au )• −→ {n}C •
=
−→
u∈N

between the direct limit complex of Koszul complexes described in 5.2.4 and
the shift {n}C • of the C̆ech complex of R with respect to a1 , . . . , an .

Proof. Let u ∈ N. We first define a chain map of complexes

(gu )• : K(au )• −→ {n}C • .

Define (gu )n : K(au )n −→ ({n}C • )n = C −n+n = C 0 = R by requiring


that

(gu )n (re1 ∧ . . . ∧ en ) = (−1)1+2+···+(n−1) r for all r ∈ R.

(We point out now, to help the reader discern a pattern in what follows, that
n−1 n
i=1 i = −n + i=1 i.) Now let k ∈ {1, . . . , n − 1} and let i be a typical
element of I(k, n) (with the notation of 5.1.4). Define an R-homomorphism

(gu )k : K(au )k −→ ({n}C • )k = C n−k

by requiring that, for i ∈ I(k, n) as above,

(−1)i(1)+···+i(k)−k
(gu )k (ei(1) ∧ . . . ∧ ei(k) ) =
(aj(1) . . . aj(n−k) )u

in the direct summand Raj(1) ...aj(n−k) of C n−k , where j ∈ I(n − k, n) is the


n-complement of i. Lastly, define an R-homomorphism (gu )0 : K(au )0 =
R −→ ({n}C • )0 = C n = Ra1 ...an by
r
(gu )0 (r) = for all r ∈ R.
(a1 . . . an )u
In order to show that ((gu )k )0≤k≤n gives rise to a chain map of complexes
(gu )• , we must show that (with the notation of 5.1.5 concerning the C̆ech com-
plex C • of R)

(gu )k−1 ◦ d(au )k = dn−k ◦ (gu )k for all k = 1, . . . , n.

We leave this to the reader in the case when k = n, and deal here with the
5.2 Use of Koszul complexes 97

case when 1 ≤ k < n. Let i ∈ I(k, n), and let j ∈ I(n − k, n) be the
n-complement of i. Then

((gu )k−1 ◦ d(au )k )(ei(1) ∧ . . . ∧ ei(k) )


 k 
= (gu )k−1 (−1)h−1 aui(h) ei(1) ∧ . . . ∧ e
i(h) ∧ . . . ∧ ei(k)
h=1
k
(−1)i(1)+···+i(h)+···+i(k)−(k−1) (−1)i(h)
= (−1)h−1 aui(h)
(aj(1) . . . aj(n−k) ai(h) )u
h=1
k
(−1)i(1)+···+i(k)−k aui(h)
= (−1)i(h)+h .
(aj(1) . . . aj(n−k) ai(h) )u
h=1

On the other hand,


 
(−1)i(1)+···+i(k)−k
(d n−k
◦ (gu )k )(ei(1) ∧ . . . ∧ ei(k) ) = d n−k
.
(aj(1) . . . aj(n−k) )u

We now refer back to 5.1.5: in order to evaluate the right-hand side of the above
equation, we need to know, for each h = 1, . . . , k, at which point i(h) should
be inserted in the sequence (j(1), . . . , j(n − k)) in order to make an increasing
sequence: if it should occupy the sh -th position in the new sequence, then

{1, 2, . . . , i(h) − 1} = {i(1), . . . , i(h − 1), j(1), . . . , j(sh − 1)} ,

so that i(h) − 1 = h − 1 + sh − 1. We thus see that

(dn−k ◦(gu )k )(ei(1) ∧ . . . ∧ ei(k) )


k
(−1)i(1)+···+i(k)−k aui(h)
= (−1)i(h)−h
(aj(1) . . . aj(n−k) ai(h) )u
h=1
= ((gu )k−1 ◦ d(au )k )(ei(1) ∧ . . . ∧ ei(k) ).

It follows that we do indeed obtain a chain map of complexes

(gu )• : K(au )• −→ {n}C • .

Next note that, for u, v ∈ N with u ≤ v and the chain map

(ψuv )• = ((ψuv )k )k∈Z : K(au )• −→ K(av )•


98 Other approaches

of complexes of 5.2.2, we have a commutative diagram


v
(ψu )•
K(au )• - K(av )•
@
@ (gv )•
(gu )•
@
@R ?
{n}C • .

It follows that there is induced a chain map of complexes


(g∞ )• = ((g∞ )k )k∈Z : K(a∞ )• = lim K(au )• −→ {n}C •
−→
u∈N

which is clearly such that (g∞ )k is surjective for all k ∈ Z. In order to show
that (g∞ )• is an isomorphism, it is enough to show that, for u ∈ N, k ∈ N0
with 0 ≤ k ≤ n and β ∈ Ker(gu )k , there exists v ∈ N with u ≤ v such that
β ∈ Ker(ψuv )k . This is very easy (and left to the reader) in the cases when
k = 0 and k = n, and so we suppose that 1 ≤ k < n. Let i ∈ I(k, n). Now
(gu )k maps the direct summand Rei(1) ∧ . . . ∧ ei(k) of K(au )k into the direct
summand Raj(1) ...aj(n−k) of C n−k = ({n}C • )k , where j ∈ I(n − k, n) is the
n-complement of i. In view of this and the fact that I(k, n) is a finite set, we
can assume that β has the form rei(1) ∧ . . . ∧ ei(k) for some r ∈ R. In this
case, the fact that β ∈ Ker(gu )k means that there exists w ∈ N0 such that
(aj(1) . . . aj(n−k) )w r = 0, and then (ψuu+w )k (rei(1) ∧ . . . ∧ ei(k) ) = 0.
It follows that (g∞ )• is an isomorphism of complexes.
5.2.6 Exercise. Complete the proof of Theorem 5.2.5.
Because the C̆ech complex C • can, by 5.1.20, be used to calculate the local
cohomology modules Hai (M ) (i ∈ N0 ), it is a consequence of Theorem 5.2.5
that these local cohomology modules can also be computed in terms of Koszul
complexes.
The result of the following exercise will be used in the proof of Theorem
5.2.9.
5.2.7 Exercise. Let N be an R-module. Let (Λ, ≤) be a (non-empty) di-
rected partially ordered set, and let ((Wα )• )α∈Λ be a direct system of com-
plexes of R-modules and chain maps over Λ, with constituent chain maps
β ) : (Wβ ) → (Wα ) (for each (α, β) ∈ Λ × Λ with α ≥ β). Set
(hα • • •

(W∞ )• := lim (Wα )• .


−→
α∈Λ

β ) (for α, β ∈ Λ with α ≥ β) turn the


Show that the chain maps N ⊗R (hα •
5.2 Use of Koszul complexes 99

family (N ⊗R (Wα )• )α∈Λ into a direct system of complexes of R-modules


and R-homomorphisms and chain maps over Λ, and show that there is an iso-
morphism of complexes
 
• ∼
lim (N ⊗R (Wα ) ) = N ⊗R (W∞ ) = N ⊗R lim (Wα ) .
• •
−→ −→
α∈Λ α∈Λ

5.2.8 Exercise. For each u ∈ N, let K(au , M )• denote M ⊗R K(au )• ,


the Koszul complex of M with respect to au1 , . . . , aun . Let α : M −→ N
be a homomorphism of R-modules. For v, w ∈ N with v ≤ w, there is a
commutative diagram
α⊗K(av )•
M ⊗R K(av )• - N ⊗R K(av )•

M ⊗(ψvw )• N ⊗(ψvw )•

? α⊗K(aw )•
?
M ⊗R K(aw )• - N ⊗R K(aw )•

of complexes and chain maps; use this to show that, for each j ∈ Z, the family
(Hj (K(au , M )• ))u∈N is a direct system, and that lim Hj (K(au , • )• ) is a
−→
u∈N
covariant R-linear functor from C(R) to itself.
Show further that
 
lim Hn−i (K(au , • )• )
−→
u∈N i∈N0

is a negative strongly connected sequence of functors from C(R) to itself.

5.2.9 Theorem. In the situation and with the notation of 5.2.8, there is a

=
natural equivalence of functors δ 0 : lim Hn (K(au , • )• ) −→ Γa from C(R) to
−→
u∈N
itself; furthermore, there is a unique isomorphism
 

=
(δ i )i∈N0 : lim Hn−i (K(au , • )• ) −→ Hai
−→ i∈N0
u∈N i∈N0

of negative (strongly) connected sequences of covariant functors (from C(R)


to C(R)) which extends δ 0 .
Consequently, for each i ∈ N0 and each R-module M ,

Hai (M ) ∼
= lim Hn−i (K(au , M )• ).
−→
u∈N
100 Other approaches

Proof. We shall use Theorem 1.3.5 again.


Let u ∈ N. Note that (with the notation of 5.2.1) the homomorphism

IdM ⊗d(au )n : M ⊗R K(au )n −→ M ⊗R K(au )n−1

has kernel {m ⊗ (e1 ∧ . . . ∧ en ) : m ∈ (0 :M (au1 , . . . , aun ))}, and that



Γa (M ) = (0 :M (au1 , . . . , aun )).
u∈N

Hence, for each u ∈ N, there is a monomorphism

δu,M : Ker(IdM ⊗d(au )n ) −→ Γa (M )

with image (0 :M (au1 , . . . , aun )); furthermore, for u, v ∈ N with u ≤ v, the


diagram

Ker(IdM ⊗d(au )n )
@
⊆ @ δu,M
@
? R
@
Ker(IdM ⊗d(av )n ) - Γa (M )
δv,M

commutes. It follows that there is induced an isomorphism



=
0
δM : lim Hn (K(au , M )• ) −→ Γa (M ),
−→
u∈N

and it is easy to check that, as M varies through C(R), the δM


0
constitute a

=
natural equivalence of functors δ : lim Hn (K(a , • )• ) −→ Γa from C(R) to
0 u
−→
u∈N
itself.
Let I be an injective R-module. The fact that passage to direct limits pre-
serves exactness ensures that
 
u ∼ u
lim Hj (K(a , I)• ) = Hj lim (K(a , I)• ) for all j ∈ Z.
−→ −→
u∈N u∈N

It now follows from Exercise 5.2.7 that there is an isomorphism of complexes

lim (K(au , I)• ) = lim (I ⊗R K(au )• ) ∼


= I ⊗R K(a∞ )• .
−→ −→
u∈N u∈N

But 5.2.5 shows that K(a∞ )• ∼


= {n}C • , and 5.1.20 shows that

H i (I ⊗R C • ) ∼
= Hai (I) for all i ∈ N0 .
5.3 Local cohomology in prime characteristic 101

Hence lim Hn−i (K(au , I)• ) = 0 for all i ∈ N.


−→
u∈N
We can now use Theorem 1.3.5 to complete the proof.

5.3 Local cohomology in prime characteristic


Since a can be generated by n elements, we must have Han+j (M ) = 0 for all
j ∈ N (by 3.3.3), so that interest focusses on Han (M ). In this section, we are
going to illustrate how the C̆ech complex approach of this chapter facilitates
calculation in Han (M ).
5.3.1 Remark. (Recall that a1 , . . . , an (where n > 0) denote n elements
which generate a.) By 5.1.20, the local cohomology module Han (M ) is iso-
morphic to Coker dn−1 , where

n
dn−1 : C(M )n−1 = Ma1 ...ai−1 ai+1 ...an −→ C(M )n = Ma1 ...an
i=1

is the (n − 1)-th homomorphism in the C̆ech complex of M with respect to


a1 , . . . , an . We use ‘[ ]’ to denote natural images of elements of Ma1 ...an
n
"in this cokernel. # Thus a typical element of Ha (M ) can be represented as
m/(a1 . . . an ) for some m ∈ M and i ∈ N0 . Note that, for u ∈ {1, . . . , n},
i

we have
$ %
aiu m
= 0 for all i ∈ N0 and m ∈ M.
(a1 . . . an )i
Thus [m /(a1 . . . an )i ] = 0 whenever m ∈ (ai1 , . . . , ain )M . It is important for
us to know exactly when an element of Coker dn−1 is zero, and this is covered
in the next lemma.
5.3.2 Lemma. Denote the product a1 . . . an by a. Let m, g ∈ M and i, j ∈
N0 . Then, with the notation of 5.3.1,
" # " #
(i) m/ai = g/aj if and only if there exists k ∈ N0 such that k ≥
n
max{i, j} and" ak−i m# − a k−j
g ∈ k
u=1 au M ;
(ii) in particular, m/ai = 0 if and only if there exists k ∈ N0 such that
n
k ≥ i and ak−i m ∈ u=1 aku M .
Proof. (i) Since
&m' & g ' $ aj m % $ a i g % $ a j m − a i g %
− j = i+j − i+j = ,
ai a a a ai+j
it is enough for us to prove (ii).
102 Other approaches
n
(ii) (⇐) There exist m1 , . . . , mn ∈ M such that ak−i m = u=1 aku mu .
Therefore
& m ' $ ak−i m % $ n ak m %
u=1 u u
= = =0
ai ak ak
 n 
because ak m /ak ∈ Im dn−1 .
" u i #u
u=1
(⇒) Since m/a = 0, we have m/ai ∈ Im dn−1 , so that there exist
j1 , . . . , jn ∈ N0 and m1 , . . . , mn ∈ M such that

m aj11 m1 ajnn mn
= + · · · + .
ai aj1 ajn
Let j := max{j1 , . . . , jn }; then there exist m1 , . . . , mn ∈ M such that
m/ai = (aj1 m1 + · · · + ajn mn )/aj , so that there is an h ∈ N0 such that

ah (aj m − ai (aj1 m1 + · · · + ajn mn )) = 0.

Take k := i + j + h to complete the proof.

5.3.3 Notation for the section. In addition to the standard notation for this
chapter, we are going to assume, for the remainder of this section, that R has
prime characteristic p. In these circumstances, the map f : R −→ R for which
f (r) = rp for all r ∈ R is a ring homomorphism (simply because the binomial
coefficient pi is an integer divisible by p for all i ∈ {1, . . . , p − 1}). We call
f the Frobenius homomorphism.
In this section, we shall use f to denote the functor obtained from restric-
tion of scalars using f , rather than the R of Chapter 4. We are making this
change in the interests of clarity: when the two rings concerned are the same,
the notation R could be confusing.
Thus R f denotes R considered as an R-module via f .
By a Frobenius action on the R-module M , we mean an Abelian group
homomorphism F : M −→ M such that F (rm) = rp F (m) for all m ∈ M
and r ∈ R. For example, the Frobenius homomorphism f : R −→ R is a
Frobenius action on R.
n
For an ideal b of R and n ∈ N0 , we shall denote by b[p ] the ideal of R
generated by all the pn -th powers of elements of b. This ideal is called the
pn -th Frobenius power of b. Observe that, if b can be generated by b1 , . . . , bt ,
n n n n n+1
then bp1 , . . . , bpt generate b[p ] , and that (b[p ] )[p] = b[p ] .

Among other things, we aim in this section to show that there is a natural
Frobenius action on each local cohomology module Hai (R) of R itself with
respect to a, and to give a detailed description of this action in the case where
i = n. (Recall our assumption that a can be generated by n elements.)
5.3 Local cohomology in prime characteristic 103

5.3.4 Theorem. (Recall that R has prime characteristic p, and that R f


denotes R considered as an R-module via f .) The Frobenius homomorphism
f : R −→ R f is a homomorphism of R-modules, and thus induces
R-homomorphisms Hai (f ) : Hai (R) −→ Hai (R f ) for all i ∈ N0 .
By the Independence Theorem 4.2.1, there is a unique isomorphism
= 
∼ 
Λ = (λi )i∈N0 : Hfi (a)R ( • ) f −→ Hai ( • f ) i∈N
i∈N0 0

of negative strongly connected sequences of covariant functors (from C(R) to


C(R)) such that λ0 is the identity natural equivalence.
For each i ∈ N0 , the map (λiR )−1 ◦ Hai (f ) is a Frobenius action on Hai (R).
√ √
Proof. Note that f (a)R = a[p] . Since a = a[p] , the local cohomol-
ogy functor with respect to a coincides with the local cohomology functor
with respect to a[p] . Thus (λiR )−1 is an R-isomorphism from Hai (R f ) to
Hfi (a)R (R) f = Hai [p] (R) f = Hai (R) f . Thus F := (λiR )−1 ◦ Hai (f ) :
Hai (R) −→ Hai (R) f is an R-homomorphism (and so certainly an Abelian
group homomorphism). Since F (rh) = f (r)F (h) = rp F (h) for all h ∈
Hai (R) and r ∈ R, we see that F is a Frobenius action on Hai (R).
5.3.5 Remark. It is important to note that the Frobenius actions defined in
5.3.4 on the Hai (R) (i ∈ N0 ) do not depend on any choice of generators for a.
However, our next task is to describe the Frobenius action on Han (R) given
by 5.3.4 in terms of our generators a1 , . . . , an for a.
5.3.6 Theorem. (Recall that R has prime characteristic p.) The natural
Frobenius action F on Han (R) of 5.3.4 is such that, with the notation of 5.3.1
and when we identify H n (C • ) with Han (R) by means of the isomorphism δR n

of 5.1.20,
$ % $ %
r rp
F = for all r ∈ R and k ∈ N0 .
(a1 . . . an )k (a1 . . . an )kp
Proof. We have F = (λnR )−1 ◦ Han (f ), and so the precise formula that we
must establish is that
$ % $ %
n −1 n −1 r rp
(δR ) ◦ (λR ) ◦ Ha (f ) ◦ δR
n n
= .
(a1 . . . an )k (a1 . . . an )kp
As in 5.3.4, let f : C(R) −→ C(R) denote the functor obtained from re-
striction of scalars using f . Since δ n is a natural equivalence of functors,
Han (f ) ◦ δR
n n
= δR f
◦ H n (C(f )• ); also
$ % $ %
r rp
H n (C(f )• ) = .
(a1 . . . an )k (a1 . . . an )k
104 Other approaches

It is therefore sufficient for us to show that


$ % $ %
n −1 n −1 m m
(δM ) ◦ (λM ) ◦ δM f n
=
(a1 . . . an )k (a1 . . . an )kp
for all m ∈ M and k ∈ N0 .
Let b ∈ R. It is straightforward to check that there is an R-isomorphism

=
νM,b : (M f )b −→ (Mb ) f for which νM,b (m/bj ) = m/bpj for all m ∈ M
and j ∈ N0 . Therefore, for each k ∈ N with 1 ≤ k ≤ n, we have an R-
isomorphism
 ∼
=
k
τM := νM,ai(1) ...ai(k) : C(M f )k −→ C(M )k f
i∈I(k,n)

(the notation I(k, n) was defined in 5.1.4). It is straightforward to check that


k
the τM (k ∈ {1, . . . , n}), together with the identity map on M f , constitute

=
an isomorphism τM •
: C(M f )• −→ C(M )• f of complexes of R-modules
and R-homomorphisms. Note that
n
τM (m/(a1 . . . an )k ) = m/(a1 . . . an )kp for all m ∈ M and k ∈ N0 .
As M varies through C(R), the τM constitute a natural equivalence of func-


=
tors τ • : C( • f )• −→ C( • )• f (from C(R) to the category of all complexes
of R-modules (and chain maps of such complexes)), and τ • induces an iso-
morphism

=
(H i (τ • ))i∈N0 : (H i (C( • f)

))i∈N0 −→ (H i (C( • )• ) f )i∈N0

of negative strongly connected sequences of covariant functors which extends


the identity natural equivalence on Γa ( • f ) = Γa ( • ) f .
We now use the isomorphism of connected sequences

=
(δ i )i∈N0 : (H i (C( • )• ))i∈N0 −→ (Hai )i∈N0
of 5.1.20 to produce further isomorphisms of connected sequences

=
(δ i• f
)i∈N0 : (H i (C( • f)

))i∈N0 −→ (Hai ( • f ))i∈N0

and

=
(δ i• f )i∈N0 : (H i (C( • )• ) f )i∈N0 −→ (Hai ( • ) f )i∈N0

which extend the identity natural equivalence on Γa ( • f ) = Γa ( • ) f . But


then
 −1  i −1
(δ i• f )i∈N0 ◦ (H i (τ • ))i∈N0 ◦ (δ • f )i∈N0 :
 i  = 
∼ 
Ha ( • ) f i∈N −→ Hai ( • f ) i∈N
0 0
5.3 Local cohomology in prime characteristic 105

is an isomorphism of connected sequences which extends the identity natural


equivalence on Γa ( • f ) = Γa ( • ) f . By the uniqueness aspect of the Inde-
pendence Theorem 4.2.1, this isomorphism must be the Λ = (λi )i∈N0 of 5.3.4.
Hence (λnM )−1 = (δM n
f ) ◦ H (τM ) ◦ (δM f )
n • n −1
. Therefore
$ % $ %
n −1 n −1 m m
(δM ) ◦ (λM ) ◦ δM f n n •
= H (τM )
(a1 . . . an )k (a1 . . . an )k
$ %
m
= ,
(a1 . . . an )kp
as required.
5.3.7 Exercise. For u, v ∈ N with u ≤ v, let hvu : R/(au1 , . . . , aun ) −→
R/(av1 , . . . , avn ) be the R-homomorphism induced by multiplication by av−u ,
where a := a1 . . . an . These homomorphisms turn the family
(R/(au1 , . . . , aun ))u∈N
into a direct system. For each u ∈ N, let
hu : R/(au1 , . . . , aun ) −→ lim R/(aw w
1 , . . . , an ) =: H
−→
w∈N

=
be the natural homomorphism. Show that there is an isomorphism α : H −→
Han (R).
Use the isomorphism α and the Frobenius action of 5.3.4 on Han (R) to put
a Frobenius action F  on H, and show that F  is given by the following rule:
F  (hu (r + (au1 , . . . , aun ))) = hpu (rp + (apu
1 , . . . , an )) for all u ∈ N, r ∈ R.
pu

We plan to exploit the Frobenius action described in 5.3.4 and 5.3.6, but the
applications we have in mind will have to be postponed until the final section of
the next chapter, by which point we shall have covered the Non-vanishing The-
orem and some interactions between local cohomology and regular sequences.
6

Fundamental vanishing theorems

There are many important results concerning the vanishing of local cohomol-
ogy modules. A few results of this type have already been presented earlier
in the book: for example, Exercise 1.3.9(iv) is concerned with the fact that,
if an R-module M is such that a contains an M -sequence of length n, then
Hai (M ) = 0 for all i < n; also, Theorem 3.3.1 shows that, if a can be gen-
erated by t elements, then, for every R-module M , we have Hai (M ) = 0 for
all i > t; and we strengthened the latter result in Corollary 3.3.3, where we
showed that, for every R-module M , we have Hai (M ) = 0 for all i > ara(a).
In this chapter, we shall provide a further result of this type: we shall prove
Grothendieck’s Vanishing Theorem, which states that, if the R-module L (is
non-zero and) has (Krull) dimension n, then Hai (L) = 0 for all i > n. We
shall also prove that, when (R, m) is a local ring and the non-zero, finitely
n
generated R-module M has dimension n, then Hm (M ) = 0, so that, in view
of Grothendieck’s Vanishing Theorem, n = dim M is the greatest integer i
i
for which Hm (M ) = 0. Also in this chapter, we shall explore in greater detail
the ideas of Exercise 1.3.9, and this investigation will lead to the result that
i
depth M is the least integer i for which Hm (M ) = 0. (Recall that depth M
is the common length of all maximal M -sequences.) It will thus follow that,
for such an M over such a local ring (R, m), it is only for integers i satisfying
depth M ≤ i ≤ dim M that it is possible that Hm i
(M ) could be non-zero,
while this local cohomology module is definitely non-zero if i is at either ex-
tremity of this range.
In §6.4, we shall exploit our earlier work to obtain a geometrical application:
we shall establish a special case of Serre’s Affineness Criterion (see [77, §46,
Corollaire 1]), concerning the following situation. Let V be an affine variety
over the algebraically closed field K. Let b be a non-zero ideal of O(V ), let
V (b) denote the closed subset of V determined by b, and let U be the open
subset V \ V (b) of V . Thus U is a quasi-affine variety. It is of fundamen-
6.1 Grothendieck’s Vanishing Theorem 107

tal importance in algebraic geometry to be able to determine whether such a


quasi-affine variety is itself affine. We shall show that U is affine if and only
if Hbi (O(V )) = 0 for all i ≥ 2. Thus this work is well suited to a chapter on
‘vanishing theorems’!
In §6.5, we shall present two applications of local cohomology to local al-
gebra. These applications are to results which have no mention of local coho-
mology in their statements, but which have proofs that make non-trivial use
of local cohomology. One concerns the Monomial Conjecture (of M. Hochster
[37, Conjecture 1]) that whenever (ai )ni=1 is a system of parameters for the n-
dimensional local ring R, then (a1 . . . an )k ∈ (ak+1
1 n ) for all k ∈ N0 .
, . . . , ak+1
We shall present a proof (due to Hochster) of this conjecture in the case where
R has prime characteristic p; our proof makes use of the Frobenius action on
n n
Hm (R) that was produced in 5.3.4 (and the fundamental fact that Hm (R) = 0).
The other application is also to local rings of characteristic p, and concerns
tight closure; it too uses a Frobenius action.

6.1 Grothendieck’s Vanishing Theorem


Our first main aim in this chapter is to present a proof of Grothendieck’s Van-
ishing Theorem. We preface this with a reminder about the dimension of an
R-module.

6.1.1 Reminder. Let M be a non-zero R-module. The (Krull) dimension,


dim M or dimR M , of M is the supremum of lengths of chains of prime ideals
in the support of M if this supremum exists, and ∞ otherwise. In the case when
M is finitely generated, this is equal to dim R/(0 : M ), the dimension of the
ring R/(0 : M ), but this need not be the case if M is not finitely generated.
We adopt the convention that the dimension of the zero R-module is −1.

6.1.2 Grothendieck’s Vanishing Theorem. Let M be an R-module. Then


Hai (M ) = 0 for all i > dim M .

Proof. Since, for each p ∈ Spec(R),

SuppRp (Mp ) = {qRp : q ∈ Supp M and q ⊆ p} ,

it follows from 4.3.3 that it is sufficient for us to prove this result under the
additional hypothesis that (R, m) is local. This is what we shall do.
When dim M = −1, there is nothing to prove, as then M = 0. The result
is also clear if a = R, as then Γa is the zero functor of 3.4.7. We therefore
suppose henceforth in this proof that M = 0 and a ⊆ m.
108 Fundamental vanishing theorems

We argue by induction on dim M . When dim M = 0, each non-zero ele-


ment g ∈ M is annihilated by a power of m (and therefore by a power of a),
because the ring R/(0 : g) is of dimension 0 and is therefore Artinian. Thus,
in this case, M is a-torsion, and so it follows from 2.1.7(i) that Hai (M ) = 0
for all i > 0 = dim M .
Now suppose, inductively, that dim M = n > 0, and the result has been
proved for all R-modules of dimensions smaller than n. Since, by 3.4.10, for
each i ∈ N0 , the local cohomology functor Hai commutes with direct limits,
and M can be viewed as the direct limit of its finitely generated submodules,
it is sufficient for us to prove that Hai (M  ) = 0 for all i > n whenever M  is
a finitely generated submodule of M . Since such an M  must have dimension
not exceeding n, we can therefore assume, in this inductive step, that M itself
is finitely generated. This we do.
By 2.1.7(iii), we have Hai (M ) ∼ = Hai (M/Γa (M )) for all i > 0. Also,
M/Γa (M ) has dimension not exceeding n, and is an a-torsion-free R-module,
by 2.1.2. In view of the inductive hypothesis, we can, and do, assume that M
is a (non-zero, finitely generated) a-torsion-free R-module.
We now use 2.1.1(ii) to deduce that a contains an element r which is a non-
zerodivisor on M . Let t, i ∈ N with i > n. The exact sequence
rt
0 −→ M −→ M −→ M/rt M −→ 0
(in which the second homomorphism is provided by multiplication by rt ) in-
rt
duces an exact sequence Hai−1 (M/rt M ) −→ Hai (M ) −→ Hai (M ) of lo-
cal cohomology modules. (The fact that the second homomorphism is again
provided by multiplication by rt follows from the fact that the functor Hai is
R-linear.)

Now dim(M/rt M ) < n since rt ∈ p∈Ass M p and every minimal member
of Supp M belongs to Ass M (see [81, Theorem 9.39]). Hence, by the induc-
tive hypothesis, Hai−1 (M/rt M ) = 0. Thus, for each t ∈ N, multiplication by
rt provides a monomorphism of Hai (M ) into itself. But r ∈ a and Hai (M )
is an a-torsion R-module, so that each element of it is annihilated by some
power of r. Therefore Hai (M ) = 0. This completes the inductive step, and the
proof.
The next theorem can be regarded as a companion to Grothendieck’s Van-
ishing Theorem, because it shows that, in some circumstances, this Vanishing
Theorem is best possible. The method of proof employed here uses the pow-
erful technique of reduction to the case where the local ring concerned is a
complete local domain. The argument is fairly sophisticated, as it relies on
ideas related to the structure theorems for complete local rings. Some readers
6.1 Grothendieck’s Vanishing Theorem 109

might like to be informed now that a different proof of the following theorem
is provided in Chapter 7. We preface the theorem with an elementary remark.
6.1.3 Remark. Let R be a second commutative Noetherian ring and let f :
R → R be a flat ring homomorphism. Let M be a non-zero, finitely generated
R-module, generated by m1 , . . . , mt ; set c = (0 :R M ).
There is an exact sequence
⊂ h
0 −→ c −→ R −→ Rm1 ⊕ · · · ⊕ Rmt
of R-modules and R-homomorphisms, where h(r) = (rm1 , . . . , rmt ) for all
r ∈ R. Since R is a flat R-module, the induced sequence
h⊗R R
0 - c ⊗R R  - R ⊗R R  - (Rm1 ⊕ · · · ⊕ Rmt ) ⊗R R

is again exact, and so it follows easily from the additivity of the tensor product
functor that the sequence
⊆ h
0 −→ cR −→ R −→ (Rm1 ⊗R R ) ⊕ · · · ⊕ (Rmt ⊗R R ),
where h (r ) = (m1 ⊗r , . . . , mt ⊗r ) for all r ∈ R , is exact. We thus deduce
that cR = (0 :R (M ⊗R R )).
It follows that, in particular, if (R, m) is local with (m-adic) completion R, 
and M is a non-zero, finitely generated R-module, then
 R
dimR M = dim R/c = dim R/c  = dim  (M ⊗R R).

R

6.1.4 The Non-vanishing Theorem. Assume that (R, m) is local, and let M
n
be a non-zero, finitely generated R-module of dimension n. Then Hm (M ) = 0.
 denote the (m-adic) completion of R. Since the natural ring ho-
Proof. Let R
momorphism R → R  is (faithfully) flat, it follows from the Flat Base Change

Theorem 4.3.2 that there is an R-isomorphism
∼
(M ) ⊗R R 
 (M ⊗R R),
n n
Hm = Hm R

 
 (M ⊗R R) = 0. Of course, mR
n
and so it is enough for us to show that Hm R
 
is the maximal ideal of the local ring R, and M ⊗R R is a non-zero, finitely

generated R-module; 
also 6.1.3 shows that this R-module has dimension n.
Consequently, we can, and do, assume henceforth in this proof that R is com-
plete.
Let p be a minimal member of Supp M for which dim R/p = n. Since
dim(pM ) < n + 1, it follows from Grothendieck’s Vanishing Theorem 6.1.2
that the natural epimorphism M → M/pM induces an epimorphism Hm n
(M )
−→ Hm (M/pM ), and so it is enough for us to show that Hm (M/pM ) = 0.
n n
110 Fundamental vanishing theorems
 
Now, since (0 :R M/pM ) = p + (0 :R M ) = p (by [81, 9.23], for
example), we can deduce that (0 :R M/pM ) = p. Hence dimR (M/pM ) = n,
and so, when we regard M/pM as an R/p-module in the natural way, it is
faithful, finitely generated, and of dimension n. By the Independence Theorem
4.2.1, we have Hm n
(M/pM ) ∼ = Hm/pn
(M/pM ) as R-modules. Since R/p is
a complete local domain, it follows that it is enough for us to prove the result
under the additional assumptions that R is a complete local domain and that
M is a faithful R-module.
At this point, we appeal to Cohen’s Structure Theorem for complete local
rings: by [50, Theorem 29.4], there exists a complete regular local subring
(R , m ) of R which is such that R is finitely generated as an R -module. Since
R is integral over R , it follows that m is the one and only prime ideal of R

which has contraction to R equal to m . Hence (m R) = m, and so it follows
from 1.1.3 and the Independence Theorem 4.2.1 that Hm n
(M ) = Hm n ∼
 R (M ) =
n  n
Hm (M ) as R -modules. It is thus sufficient for us to show that Hm (M ) = 0.
Obviously, M is faithful and finitely generated as an R -module. Moreover,
dim R = dim R = n (since R is integral over R ). We can therefore replace
R by R and thus assume that R is a complete regular local ring during the
remainder of this proof.
Next, let, for each R-module G,

τ (G) := {g ∈ G : there exists r ∈ R \ {0} such that rg = 0} ,

a submodule of G. Note that (0 : τ (M )) = 0, and so τ (M ) has dimension less


than n. Another use of Grothendieck’s Vanishing Theorem 6.1.2 shows that
the natural epimorphism M → M/τ (M ) induces an isomorphism Hm n
(M ) →
n n
Hm (M/τ (M )), and so it is sufficient for us to show that Hm (M/τ (M )) = 0.
Note that τ (M/τ (M )) = 0 and that M/τ (M ) still has dimension n (as its
annihilator is 0). Therefore we can, and do, assume that τ (M ) = 0.
Let K be the field of fractions of R, and note that, since τ (M ) = 0,
the natural R-homomorphism M → M ⊗R K is injective. Let t denote the
(torsion-free) rank of M as an R-module, that is, the vector space dimension
dimK (M ⊗R K). There exists r ∈ R \ {0} such that M , as an R-module, can
be embedded in the submodule

R 1r ⊕ · · · ⊕ R 1r (t copies) of K ⊕ · · · ⊕ K (t copies).

But R 1r ∼
= R, and so there is a free R-module F of rank t and an R-monomo-
rphism h : M → F ; also, since h ⊗R K : M ⊗R K → F ⊗R K must be
a K-monomorphism between t-dimensional vector spaces over K, and there-
fore an isomorphism, it follows that dim(Coker h) < n. It therefore follows
6.1 Grothendieck’s Vanishing Theorem 111

from Grothendieck’s Vanishing Theorem 6.1.2 that h induces an exact se-


quence Hm n
(M ) −→ Hm n
(F ) −→ 0. Now since R is a regular local ring, it
n
follows from 4.2.6 that Hm (R) = 0; finally, we deduce from the additivity of
n n n
the functor Hm that Hm (F ) = 0. Hence Hm (M ) = 0 and this completes the
proof.
6.1.5 Exercise. Suppose that M is a finitely generated R-module for which
M = aM . Show that there exists i ∈ N0 for which Hai (M ) = 0.
In 3.3.4, we introduced the cohomological dimension cohd(a) of a.
6.1.6 Lemma. If a is proper, then ht a ≤ cohd(a).
Proof. Let p be a minimal prime ideal of a such that ht p = ht a =: h. Then,
by 1.1.3 and 4.3.3, we have (Hah (R))p ∼ h
= HaR p
h
(Rp ) = HpR p
(Rp ), and this is
h
non-zero by the Non-vanishing Theorem 6.1.4. Therefore Ha (R) = 0, so that
ht a = h ≤ cohd(a).
6.1.7 Exercise. Assume that (R, m) is local, and let M be a non-zero, finite-
n
ly generated R-module of dimension n > 0. Show that Hm (M ) is not finitely
generated. (Here are some hints: use M/Γm (M ) to see that one can make the
additional assumption that M is an m-torsion-free R-module; use 2.1.1(ii) to
see that, then, m contains an element r which is a non-zerodivisor on M ; note
that, if Hmn
(M ) were finitely generated, then there would exist t ∈ N such that
t n
r Hm (M ) = 0; and consider the long exact sequence of local cohomology
rt
modules induced by the exact sequence 0 −→ M −→ M −→ M/rt M −→ 0
in order to obtain a contradiction.)
6.1.8 Exercise. Provide an example to show that the result of Theorem 6.1.4
is not always true if the hypothesis that M be finitely generated is omitted.
6.1.9 Exercise. Let T : C(R) −→ C(R) be an R-linear covariant functor.
For each R-module M , and each g ∈ M , let μg,M : R → M be the R-
homomorphism for which μg,M (r) = rg for all r ∈ R.
Show that there is a natural transformation of functors
θ : ( • ) ⊗R T (R) −→ T
(from C(R) to C(R)) which is such that, for each R-module M ,
θM (g ⊗ z) = T (μg,M )(z) for all g ∈ M and z ∈ T (R).
Show also that θF is an isomorphism whenever F is a finitely generated free
R-module, and deduce that, if T is right exact, then θM is an isomorphism
whenever M is a finitely generated R-module.
112 Fundamental vanishing theorems

6.1.10 Exercise. Let n be an integer such that Hai (M ) = 0 for all i > n
and for all R-modules M . (For example, this condition would be satisfied if
n ≥ ara(a) (by 3.3.3), or if n ≥ dim R (by 6.1.2). Use Exercise 6.1.9 to prove
that Han is naturally equivalent to ( • ) ⊗R Han (R).
We interpret the supremum of the empty set of integers as −∞.
6.1.11 Proposition. The cohomological dimension cohd(a) of a is equal to
sup{i ∈ N0 : Hai (R) = 0}.
Proof. The result is clear when a = R, and so we suppose that a is proper.
Denote by d the greatest integer i such that Hai (R) = 0. (Such an integer
exists, by 3.3.4 and 6.1.6.) Clearly d ≤ cohd(a) =: c. By 6.1.10, the functors
Hac and ( • ) ⊗R Hac (R) are naturally equivalent. Therefore Hac (R) = 0 and
c ≤ d.

6.2 Connections with grade


It is now time for us to explore in detail connections, already hinted at in Exer-
cise 1.3.9, between regular sequences and local cohomology. To set the scene,
and clarify precisely what background information the reader will require for
this work, we begin by recalling the definition of ‘M -sequence’ and quoting,
without proof, two results from Matsumura [50]; however, we shall use termi-
nology different from his.
6.2.1 Definition. (See [50, p. 123].) Let a1 , . . . , an ∈ R and let M be an R-
module. We say that the sequence a1 , . . . , an is a poor M -sequence precisely
when
(i) a1 is a non-zerodivisor on M and, for each i = 2, . . . , n, the element ai
i−1
is a non-zerodivisor on M/ j=1 aj M .
Furthermore, a1 , . . . , an is said to be an M -sequence, or an M -regular se-
quence, precisely when it is a poor M -sequence, that is, it satisfies condition
(i) above, and, in addition,
n
(ii) j=1 aj M = M .

6.2.2 Theorem. (See [50, Theorem 16.6].) Let M be a finitely generated R-


module such that aM = M ; let n ∈ N. Then the following statements are
equivalent:
(i) ExtiR (G, M ) = 0 for all i < n and each finitely generated R-module G
for which Supp G ⊆ Var(a);
6.2 Connections with grade 113

(ii) ExtiR (R/a, M ) = 0 for all i < n;


(iii) there is a finitely generated R-module G with Supp G = Var(a) for
which ExtiR (G, M ) = 0 for all i < n; and
(iv) there exists an M -sequence of length n contained in a. 

Ideas involved in Theorem 6.2.2 can be adapted to prove the result in the
following exercise, which is often useful.

6.2.3 Exercise. Let N be an R-module (note that it is not assumed that


N is finitely generated) such that there exists a poor N -sequence of length n
contained in a. Show that ExtiR (G, N ) = 0 for all i < n and each finitely
generated R-module G for which Supp G ⊆ Var(a).

Observe that, with the notation of 6.2.2, every M -sequence contained in a


(even the empty one!) can be extended to a maximal one, since otherwise, as
aM = M , there would exist an infinite sequence (bi )i∈N of elements of a such
that b1 , . . . , bn is an M -sequence for all n ∈ N, and this would lead to an
infinite strictly ascending chain

(b1 ) ⊂ (b1 , b2 ) ⊂ · · · ⊂ (b1 , . . . , bn ) ⊂ · · ·

of ideals of R. We incorporate this observation into the next theorem, which is


otherwise taken from Matsumura [50].

6.2.4 Theorem and Definition. (See [50, Theorem 16.7].) Let M be a finitely
generated R-module such that aM = M .

(i) There exists an M -sequence contained in a which cannot be extended


to a longer one by the addition of an extra term (such a sequence will
henceforth be referred to as a maximal M -sequence contained in a).
(ii) Every M -sequence contained in a can be extended to a maximal one.
(iii) All maximal M -sequences contained in a have the same length, namely
the least integer i such that ExtiR (R/a, M ) = 0.

We shall refer to the common length of all maximal M -sequences contained


in a as the M -grade of a, and we shall denote this non-negative integer by
gradeM a. 

6.2.5 Notes. The following points should be noted.

(i) If the ideal a is proper, then the R-grade of a is defined: we shall follow
Rees’s original terminology [68] and refer to this simply as the grade of
a, and we shall denote it by grade a.
114 Fundamental vanishing theorems

(ii) Suppose that (R, m) is local, and that M is a non-zero finitely generated
R-module. Then it follows from Nakayama’s Lemma that mM = M ,
and so gradeM m is defined: this is referred to as the depth of M , and
denoted by depth M or depthR M . Since every M -sequence must be
contained in m, we see that depth M is equal to the common length of
all maximal M -sequences.

6.2.6 Exercise. Let M be an R-module and let i ∈ N0 . Show that


 
Supp Hai (M ) ⊆ Supp M ∩ Var(a).

Deduce that, if M is finitely generated and aM = M , then Haj (M ) = 0 for


all j ∈ N0 .

6.2.7 Theorem. Let M be a finitely generated R-module such that aM = M .


Then gradeM a is the least integer i such that Hai (M ) = 0.

Proof. Let g := gradeM a. We use induction on g. When g = 0, every ele-


ment of a must be a zerodivisor on M , and so Γa (M ) = 0 by 2.1.1(ii). Now
suppose that g > 0 and that the result has been proved for each finitely gener-
ated R-module N with aN = N and gradeN a < g.
There exists a1 ∈ a such that a1 is a non-zerodivisor on M . Set M1 :=
M/a1 M , and observe that (aM1 = M1 and) gradeM1 a = g−1. Therefore, by
the inductive hypothesis, Hai (M1 ) = 0 for all i < g−1, while Hag−1 (M1 ) = 0.
a1
The exact sequence 0 −→ M −→ M −→ M1 −→ 0 induces, for each i ∈ N,
an exact sequence
1 a
Hai−1 (M ) −→ Hai−1 (M1 ) −→ Hai (M ) −→ Hai (M ).

This shows that, for i < g, the element a1 is a non-zerodivisor on Hai (M ),


so that, since this module is a-torsion, it must be zero. We therefore have an
exact sequence 0 −→ Hag−1 (M1 ) −→ Hag (M ), and since Hag−1 (M1 ) = 0, it
follows that Hag (M ) = 0.

6.2.8 Corollary. Assume that (R, m) is local, and let M be a non-zero, finite-
i
ly generated R-module. Then any integer i for which Hm (M ) = 0 must satisfy

depth M ≤ i ≤ dim M,
i
while for i at either extremity of this range we do have Hm (M ) = 0.

Proof. This is immediate from Grothendieck’s Vanishing Theorem 6.1.2, the


Non-vanishing Theorem 6.1.4 and the above 6.2.7, because M = mM by
Nakayama’s Lemma, and depth M = gradeM m.
6.2 Connections with grade 115

6.2.9 Corollary. Assume that (R, m) is local, and let M be a non-zero, finite-
i
ly generated R-module. Then there is exactly one integer i for which Hm (M )
= 0 if and only if depth M = dim M , that is, if and only if M is a Cohen–
Macaulay R-module (see [50, p. 134]).
In particular, if (R, m) is a regular local ring of dimension n, then n is the
unique integer i for which Hm i
(R) = 0. 
6.2.10 Remark. Let M be a finitely generated R-module such that aM =
M . It is immediate from Theorems 2.2.6(i)(c) and 6.2.7 that ηM : M −→
Da (M ) is an isomorphism if and only if gradeM a ≥ 2.
6.2.11 Exercise. Let T : C(R) −→ C(R) be an R-linear covariant functor
with the property that T (M ) is a-torsion for every R-module M .
(i) Show that, for each R-module M , the result Ri T (M ) of applying the
i-th right derived functor of T to M is again a-torsion.
(ii) Assume, in addition, that T is left exact and such that, for each finitely
generated R-module M , the ideal a contains a non-zerodivisor on M if
and only if T (M ) = 0. Let r ∈ N0 , and let M be a finitely generated
R-module such that aM = M . Prove that Ri T (M ) = 0 for all integers
i < r if and only if a contains an M -sequence of length r.
(iii) Use part (ii) applied to the functor HomR (R/a, • ) to reprove the equiv-
alence of statements (ii) and (iv) in Theorem 6.2.2.
(iv) Use part (ii) applied to the functor Γa to prove the result of 6.2.7.
Corollary 6.2.8 raises the following question: if J is any prescribed finite
non-empty set of non-negative integers, does there exist a local ring (R , m )

 (R ) = 0 if and only if i ∈ J? An affirmative
i
having the property that Hm
answer to this question was provided by I. G. Macdonald [46], and the next
two exercises sketch the essence of his argument.
6.2.12 Exercise. Let M be an R-module.
(i) Show that the Abelian group R⊕M is a commutative ring (with identity)
with respect to multiplication defined by
(r1 , m1 )(r2 , m2 ) = (r1 r2 , r1 m2 + r2 m1 )
for all (r1 , m1 ), (r2 , m2 ) ∈ R ⊕ M . This ring is called the trivial exten-
sion of R by M , and we shall denote it by R ∝ M .
(ii) Show that the nilradical of R ∝ M contains 0 × M , and show that
Spec(R ∝ M ) = {p × M : p ∈ Spec(R)}.
(iii) Show that the ring R ∝ M is Noetherian if and only if M is a finitely
generated R-module.
116 Fundamental vanishing theorems

(iv) Let S be a multiplicatively closed subset of R. Show that S × M is a


multiplicatively closed subset of R ∝ M and that there is a ring iso-

morphism φ : (S × M )−1 (R ∝ M ) −→ S −1 R ∝ S −1 M for which
=

φ((r, m)/(s, m )) = (r/s, (sm − rm )/s2 ) for all r ∈ R, s ∈ S and


m, m ∈ M . In particular, this will show that
(R ∝ M )(p×M ) ∼
= Rp ∝ Mp for all p ∈ Spec(R).
(v) Show that, if R is local and M is finitely generated, then R ∝ M is local
with maximal ideal m × M , and that dim(R ∝ M ) = dim R.
6.2.13 Exercise. Let h, n be integers such that 0 ≤ h ≤ n, and let J be
an arbitrary set of integers such that {h, n} ⊆ J ⊆ {i ∈ N0 : h ≤ i ≤ n}.
Let (R, m) be a regular local ring of dimension n, and let u1 , . . . , un be n
elements which generate the maximal ideal m. For each j = 0, . . . , n − 1, let
pj = Ruj+1 + · · · + Run . Set

M := R/pj .
j∈J\{n}

(i) Show that, for i ∈ N0 , we have Hm i


(M ) = 0 if and only if i ∈ J \ {n}.

(ii) Let R := R ∝ M , the trivial extension of R by M of 6.2.12 above. By
that exercise, R is a local ring, with maximal ideal m := m × M . The
ring homomorphism ψ : R ∝ M → R for which ψ((r, m)) = r for all
(r, m) ∈ R ∝ M enables us to regard both R and M as R -modules.
q ψ
Use the exact sequence 0 −→ M −→ R ∝ M −→ R −→ 0 of R -
modules (in which q is the canonical injection) to show that, for i ∈ N0 ,

 (R ) = 0 if and only if i ∈ J.
i
we have Hm
6.2.14 Exercise. Let M be an R-module, and let a1 , . . . , an ∈ R be such
that M = (a1 , . . . , an )M . Prove that a1 , . . . , an is an M -sequence if and only
if
i−1
H(a 1 ,...,ai )
(M ) = 0 for all i = 1, . . . , n.
(Here are some hints for the implication ‘⇐’. Use induction on n: for the
inductive step, on the assumption that n > 1 and the result has been proved for
smaller values of n, use Exercise 1.3.9(iv) (in conjunction with the hypotheses)
to help you prove that
i−2
H(a 1 ,...,ai )
(M/a1 M ) = 0 for all i = 2, . . . , n;
then use the Independence Theorem 4.2.1 to deduce that
i−2
H(a 2 ,...,ai )
(M/a1 M ) = 0 for all i = 2, . . . , n.)
6.3 Exactness of ideal transforms 117

6.2.15 Exercise. Let (R, m) be a regular local ring of dimension d ≥ 2, and


let u1 , . . . , ud be d elements which generate m. Let t ∈ N0 with t < d − 1,
and let p = Ru1 + · · · + Rut , a prime ideal of R. Let M denote the R-module
Rp /R.
(i) Show that pM = M .
(ii) Show that u1 , . . . , ut is a maximal M -sequence contained in m.
(iii) Show that the least integer i such that Hm i
(M ) = 0 is d − 1.
Contrast this with the situation for finitely generated R-modules!
6.2.16 Exercise. Let (R, m) be a regular local ring of dimension d > 2, let
K denote the quotient field of R, and let N := R ⊕ K/R.
Show that mN = N , that m consists entirely of zerodivisors on N , and that
ηN : N → Dm (N ) is an isomorphism. Compare this with the situation for
finitely generated R-modules described in 6.2.10.

6.3 Exactness of ideal transforms


This section prepares the ground for our presentation, in the next section, of
the promised special case of Serre’s Affineness Criterion. Central to this prepa-
ration are various necessary and sufficient conditions for the exactness of the
a-transform functor Da ; some of these were presented by P. Schenzel in [73].
6.3.1 Lemma. The following statements are equivalent:
(i) the a-transform functor Da is exact;
(ii) Hai (R) = 0 for all i ≥ 2;
(iii) Ha2 (M ) = 0 for each finitely generated R-module M ;
(iv) Ha2 (M ) = 0 for each R-module M .
Proof. (i) ⇒ (ii) By 2.2.6(ii), for each i ∈ N, the i-th right derived functor
Ri Da of Da is naturally equivalent to Hai+1 . Now exactness of Da implies
that Ri Da (R) = 0 for all i ∈ N; therefore Hai (R) = 0 for all i ≥ 2.
(ii) ⇒ (iii) Let ara(a) = t. By 3.3.3, for every R-module M , we have
Hai (M ) = 0 for all i > t. We now argue by descending induction on i. Suppose
that i ∈ N with i > 2, and we have proved that Hai (M  ) = 0 for each finitely
generated R-module M  ; let M be an arbitrary finitely generated R-module.
There exists an exact sequence
0 −→ N −→ F −→ M −→ 0
of finitely generated R-modules and R-homomorphisms in which F is free.
118 Fundamental vanishing theorems

This induces an exact sequence Hai−1 (F ) −→ Hai−1 (M ) −→ Hai (N ). Since


i − 1 ≥ 2, it follows from the additivity of the (i − 1)-th local cohomology
functor and condition (ii) that Hai−1 (F ) = 0; since Hai (N ) = 0 by our in-
ductive assumption, we can deduce that Hai−1 (M ) = 0. This completes the
inductive step. Hence statement (iii) is proved by descending induction.
(iii) ⇒ (iv) This is immediate from the fact (3.4.10) that Ha2 commutes with
direct limits, because each R-module can be viewed as the direct limit of its
finitely generated submodules.
(iv) ⇒ (i) By 2.2.6(ii), the first right derived functor R1 Da of Da is nat-
urally equivalent to Ha2 . It therefore follows from statement (iv) (and the left
exactness of Da ) that Da is exact.

6.3.2 Exercise. Assume that ht p = 1 for every minimal prime ideal p of a,


and that Rq is a UFD for each prime ideal q of R which contains a. Prove that
Da : C(R) → C(R) is exact.

6.3.3 Remark. Let R be a second commutative Noetherian ring and let f :


R → R be a ring homomorphism. Suppose that the a-transform functor Da is
exact. Then since the restriction functor R : C(R ) → C(R) is exact and, by
2.2.24, there is a natural equivalence of functors

ε : DaR ( • ) R −→ Da ( • R ),

it follows that the aR -transform functor DaR : C(R ) → C(R ) is also exact.

Recall from Exercise 2.2.5 that Da (R) has a natural structure as a commu-
tative ring with identity, and from Exercise 2.2.12 that ηR : R → Da (R) is a
ring homomorphism. Thus we can regard Da (R) as an R-algebra by means of
ηR . Our next proposition is concerned with this R-algebra structure.

6.3.4 Proposition. Suppose that aDa (R) = Da (R). Then Da (R) is a finitely
generated R-algebra.

Proof. Note that Γa (R) = R if and only if a is nilpotent, and that in this case
Γa is the identity functor, so that Da (R) = 0 by 2.2.6(i)(c) and the claim is
clear in this case. Thus we suppose henceforth in this proof that Γa (R) = R.
Set R := R/Γa (R). We mentioned just before the statement of the propo-
sition that ηR : R −→ Da (R) is a ring homomorphism. By 2.2.6(i)(c), we
have Ker(ηR ) = Γa (R), and so ηR induces an injective ring homomorphism
θR : R −→ Da (R). Moreover, as Coker ηR ∼ = Ha1 (R) by 2.2.6(i)(c), the
R-module Coker θR is a-torsion, where a := aR, the extension of a to R.
Therefore, by 2.2.17, the unique R-homomorphism ψ : Da (R) −→ Da (R)
6.3 Exactness of ideal transforms 119

such that the diagram

R
θR
- Da (R)

@
@
ηa,R
ψ
@
R ?
@
Da (R)

commutes is a ring homomorphism. Furthermore, by the Independence The-


orem 4.2.1, we have Hai (Da (R)) ∼ = Hai (Da (R)) for all i ∈ N0 , and so
i
Ha (Da (R)) = 0 for i = 0, 1 by 2.2.10(iv). Hence, in view of 2.2.15(iii),
the map ψ is a ring isomorphism. Therefore aDa (R) = Da (R), and, if we
can show that Da (R) is a finitely generated R-algebra, then it will follow that
Da (R) is a finitely generated R-algebra. Since Γa (R) = 0 (by 2.1.2), we
therefore assume henceforth in this proof that Γa (R) = 0.
By 2.1.1(ii), this means that a contains a non-zerodivisor s on R. Note that
ηR : R → Da (R) is injective, by 2.2.6(i)(c). We again use Rs to denote
the ring of fractions of R with respect to the multiplicatively closed subset

si : i ∈ N0 . By 2.2.18, the subring D := n∈N (R :Rs an ) of Rs satisfies
aD = D, and it will be sufficient for us to show that D is a finitely generated
R-algebra.
Let a1 , . . . , at be t elements which generate a. Since aD = D, there ex-
t
ist y1 , . . . , yt ∈ D such that 1 = i=1 ai yi . We aim to show that D =
R[y1 , . . . , yt ]. We achieve this by showing, by induction on n, that
(R :Rs an ) ⊆ R[y1 , . . . , yt ] for every n ∈ N0 .
This claim is clear for n = 0, and so we suppose that n > 0 and the claim has
been proved for smaller values of n. Let z ∈ (R :Rs an ). Note that
ai z ∈ (R :Rs an−1 ) ⊆ R[y1 , . . . , yt ] for all i = 1, . . . , t.
t t
Thus z = 1z = i=1 ai yi z = i=1 (ai z)yi ∈ R[y1 , . . . , yt ], and the induc-
tive step is complete.
Therefore D = R[y1 , . . . , yt ], as claimed, and Da (R) is a finitely generated
R-algebra.
In Lemma 6.3.1, we established several criteria for the exactness of the a-
transform functor Da . We are now in a position to prove a further such crite-
rion.
6.3.5 Proposition. The a-transform functor Da is exact if and only if
aDa (R) = Da (R).
120 Fundamental vanishing theorems

Proof. (⇒) Assume first that Da is exact. By 6.1.9, we have

Da (R/a) ∼
= (R/a) ⊗R Da (R) ∼
= Da (R)/aDa (R).

Since Da (R/a) = 0 by 2.2.10(i), we deduce that aDa (R) = Da (R).


(⇐) Assume now that aDa (R) = Da (R). By Lemma 6.3.1, it is enough
for us to show that Hai (R) = 0 for all i ≥ 2; by Corollary 2.2.10(v), we have
Hai (R) ∼= Hai (Da (R)) for all i ≥ 2, and therefore it is enough for us to show
that Ha (Da (R)) = 0 for all i ≥ 2.
i

By Proposition 6.3.4, the commutative R-algebra Da (R) is finitely gener-


ated, and therefore a Noetherian ring. Therefore, by the Independence Theorem
4.2.1,

Hai (Da (R)) ∼ i


= HaDa (R)
(Da (R)) for all i ∈ N0 .

However, the assumption that aDa (R) = Da (R) means that ΓaDa (R) is the
i
zero functor, and so HaDa (R)
(Da (R)) = 0 for all i ∈ N0 . This completes the
proof.

6.3.6 Corollary. Assume the a-transform functor Da is exact. Then ht p ≤ 1


for every minimal prime ideal p of a.

Proof. Suppose that p is a minimal prime ideal of a with ht p =: t ≥ 2, and


look for a contradiction. Since Da is exact, it follows from Lemma 6.3.1 that
Hat (R) = 0. Therefore, by 1.1.3 and 4.3.3,
t
HpR p
t
(Rp ) = HaR p
(Rp ) ∼
= (Hat (R))p = 0.

However, this contradicts Theorem 6.1.4, since dim Rp = t.

6.3.7 Exercise. Let c be a second ideal of R, and suppose that c ⊆ a. Show


that Dc (ηa,( • ) ) : Dc ( • ) −→ Dc (Da ( • )) is a natural equivalence of functors.
(Here is a hint: you might find the argument used in the proof of 2.2.10(iii)
helpful.)

6.3.8 Lemma. Let b be a second ideal of R and let M be an R-module. Then


the R-homomorphism

αa,ab,Db (M ) : Da (Db (M )) −→ Dab (Db (M ))

(which results from application of the natural transformation αa,ab of 2.2.23(i)


to the R-module Db (M )) is an isomorphism.
6.3 Exactness of ideal transforms 121

Proof. The diagram


ηa,Db (M )
Db (M ) - Da (Db (M ))

αa,ab,Db (M )

ηab,Db (M ) ?
Db (M ) - Dab (Db (M ))

commutes. By 2.2.10(iv), we have Hbi (Db (M )) = 0 for i = 0, 1. Hence, by


the Mayer–Vietoris sequence 3.2.3, for i = 0, 1, there is an exact sequence
Hai (Db (M )) −→ Hab
i
(Db (M )) −→ Ha+b
i+1
(Db (M )),
i
so that Hab (Db (M )) is a-torsion. Hence, by 2.2.6(i)(c) and 2.2.15, there is
a unique homomorphism θ : Dab (Db (M )) −→ Da (Db (M )) such that the
diagram
ηab,Db (M )
Db (M ) - Dab (Db (M ))

ηa,Db (M ) ?
Db (M ) - Da (Db (M ))

commutes. In fact, it is immediate from the uniqueness aspects of 2.2.15 that


θ ◦ αa,ab,Db (M ) = IdDa (Db (M )) and αa,ab,Db (M ) ◦ θ = IdDab (Db (M )) ,
and so αa,ab,Db (M ) is an isomorphism.
6.3.9 Exercise. Let b be a second ideal of R. Suppose that the functors Da
g h
and Db are both exact. Let L −→ M −→ N be an exact sequence of R-
modules and R-homomorphisms.

(i) Use 6.3.8 to show that the induced sequence


Dab (Db (g)) Dab (Db (h))
Dab (Db (L)) - Dab (Db (M )) - Dab (Db (N ))

is exact.
(ii) Use 6.3.7 to show that the induced sequence
Dab (g) Dab (h)
Dab (L) - Dab (M ) - Dab (N )

is exact, and conclude that the functor Dab is exact.


122 Fundamental vanishing theorems

6.4 An Affineness Criterion due to Serre


Although the main result of this section, 6.4.4, is, strictly speaking, only a
special case of Serre’s Affineness Criterion (see [77, §46, Corollaire 1]), we
shall nevertheless refer to it as ‘Serre’s Affineness Criterion’.

6.4.1 Notation. In our discussion of Serre’s Affineness Criterion, we shall


use the following notation. We shall use V to denote an affine variety over the
algebraically closed field K, and b will denote a non-zero ideal of O(V ); we
shall use U to denote the quasi-affine variety V \ V (b), where V (b) is the
closed subset of V determined by b.
Also, for an affine or quasi-affine variety W over K and a point q ∈ W , we
shall frequently denote the maximal ideal {f ∈ O(W ) : f (q) = 0} of O(W )
by IW (q). The local ring of W at q will be denoted by OW,q .

6.4.2 Reminders. Let the notation be as in 6.4.1.

(i) The quasi-affine variety U is said to be affine precisely when there exists

=
an affine variety W over K and an isomorphism of varieties U −→ W .
(ii) Let W be an affine variety over K. If α : U → W is a morphism of
varieties, we shall denote by α∗ : O(W ) → O(U ) the homomorphism
of K-algebras induced by α. Recall, from [30, Chapter I, Proposition
3.5] for example, that the correspondence α → α∗ provides a bijective
map from the set of all morphisms of varieties U → W to the set of all
K-algebra homomorphisms O(W ) → O(U ).
(iii) Let T be a closed subvariety of V , so that T is a closed, irreducible (and
so necessarily non-empty) subset of V and the ideal

p := {f ∈ O(V ) : f (t) = 0 for all t ∈ T }

of O(V ) is prime. Recall that codimV T , the codimension of T in V , is


given by

codimV T = dim V − dim T = dim O(V ) − dim O(T ) = htO(V ) p.

(iv) Recall that a non-empty closed subset C of V is said to be of pure codi-


mension r in V precisely when every irreducible component of C has
codimension r in V .
(v) Let u ∈ U . With the natural identifications, we have

O(V ) ⊆ O(U ) ⊆ OU,u = O(U )IU (u) ⊆ K(U ) = K(V ).

(vi) Finally, we recall that points on the quasi-affine variety U can be ‘sep-
arated by regular functions’. More precisely, let u1 , . . . , ur be r distinct
6.4 An Affineness Criterion due to Serre 123

points of U and let c1 , . . . , cr ∈ K. Then there is a function f ∈ O(U )


such that f (ui ) = ci for all i = 1, . . . , r. In particular, if u, u ∈ U are
such that IU (u) = IU (u ), then u = u .

6.4.3 Lemma. Let the notation be as in 6.4.1, and let W be an affine variety
over K. Let α : U → W be a surjective morphism of varieties for which α∗ :
O(W ) → O(U ) is an isomorphism of K-algebras. Then α is an isomorphism
of varieties.

Proof. Let w ∈ W . Since α is surjective, there exists u ∈ U with α(u) =


w. Now, for f ∈ O(W ), we have f (w) = f (α(u)) = 0 if and only if
(α∗ (f ))(u) = 0. It follows that α∗ (IW (w)) = IU (u). This shows that IU (u)
is uniquely determined by w. By 6.4.2(vi), it follows that there is exactly one
u ∈ U for which α(u) = w. Hence α is bijective. It thus remains only for
us to show that its inverse α−1 : W → U is a morphism of varieties. For
this, let U  be a non-empty open subset of U , and let f ∈ O(U  ): it is enough
for us to show that (α−1 )−1 (U  ) = α(U  ) is an open subset of W and that
f ◦ (α−1 α(U  ) ) : α(U  ) −→ K is regular.
Before establishing these two points, we make one preparatory observation.
For each h ∈ O(U ), the regular function (α∗ )−1 (h) ∈ O(W ) has the property
that, for each w ∈ W ,
 ∗ −1    
(α ) (h) (w) = (α∗ )−1 (h) α(α−1 (w))
= α∗ ((α∗ )−1 (h))(α−1 (w)) = h(α−1 (w)).

We now turn our attention to U  . There is a non-zero ideal c of O(V )


such that U  = V \ V (c). Thus, with the notation of 2.3.1, we have U  =

g∈c UV (g). However, for each g ∈ c, we have

α(UV (g)) = α ({v ∈ V : g(v) = 0})


= α ({u ∈ U : (g U )(u) = 0})
−1
= {w ∈ W : (g U )(α (w)) = 0}
 
= {w ∈ W : (α∗ )−1 (g U ) (w) = 0}

by the preceding paragraph; since (α∗ )−1 (g U ) is a regular function on W , we



see that α(UV (g)) is an open subset of W . Hence α(U  ) = g∈c α(UV (g)) is
open too.
Next, let w ∈ α(U  ). Since f ∈ O(U  ), there exists an open subset U  of
U with α−1 (w) ∈ U  and regular functions h, k ∈ O(U ) such that k does


not vanish on U  and f (p) = h(p)/k(p) for all p ∈ U  . It follows from


the immediately preceding paragraph that α(U  ) is an open subset of W that
124 Fundamental vanishing theorems

contains
 ∗w,−1and from
 the paragraph before that that, for all q ∈ α(U  ), we
have (α ) (k) (q) = k(α−1 (q)) = 0 and
 ∗ −1 
 −1
 −1 h(α−1 (q)) (α ) (h) (q)
f ◦ (α α(U  ) ) (q) = f (α (q)) = = .
k(α−1 (q)) ((α∗ )−1 (k)) (q)
Since (α∗ )−1 (h) and (α∗ )−1 (k) are regular functions on W , it follows that
f ◦ (α−1 α(U  ) ) is a regular function on α(U  ), as required. This completes
the proof that α−1 : W → U is a morphism of varieties.

We are now ready to present Serre’s Affineness Criterion.

6.4.4 Serre’s Affineness Criterion. Let the notation be as in 6.4.1. Then the
following conditions are equivalent:

(i) U = V \ V (b) is affine;


(ii) Db : C(O(V )) −→ C(O(V )) is exact;
(iii) Hbi (O(V )) = 0 for all i ≥ 2;
(iv) Hb2 (M ) = 0 for each finitely generated O(V )-module M ;
(v) Hb2 (M ) = 0 for each O(V )-module M ;
(vi) bDb (O(V )) = Db (O(V )).

Proof. The equivalence of the last five conditions (ii) – (vi) was established in
6.3.1 and 6.3.5. It only remains for us to establish that these are also equivalent
to statement (i). Let ι : U → V denote the inclusion morphism of varieties.
(i) ⇒ (vi) Assume that U is affine, so that there is an affine variety W over

K and an isomorphism of varieties α : U −→ W . Set β := ι ◦ α−1 : W → V .
=

In view of 2.3.2, we therefore have a commutative diagram


ηO(V )
O(V ) - Db (O(V ))
@ 6
β ∗
ι∗
@ ∼
= ν=νV,b
@
? @
α∗
@
R
O(W ) - O(U )

=

of K-algebra homomorphisms. Let n be a maximal ideal of Db (O(V )). Then


−1 ∗−1 −1
ηO(V ) (n) = ι (ν (n)) = β ∗−1 (α∗−1 (ν −1 (n))).

Now α∗−1 (ν −1 (n)) is a maximal ideal of the ring O(W ), and so there exists
w ∈ W such that α∗−1 (ν −1 (n)) = IW (w). Hence
−1 ∗−1
ηO(V ) (n) = β (IW (w)) = IV (β(w)).
6.4 An Affineness Criterion due to Serre 125

However, β(w) = ι(α−1 (w)) ∈ U = V \ V (b), and so

−1
b ⊆ IV (β(w)) = ηO(V ) (n).

This is true for each maximal ideal n of Db (O(V )); therefore bDb (O(V )) =
Db (O(V )).
(vi) ⇒ (i) Assume that bDb (O(V )) = Db (O(V )). Then Db (O(V )) is a
finitely generated O(V )-algebra, by 6.3.4. Moreover Db (O(V )) is an integral
domain, and so there exists an affine variety W over K for which O(W ) =
Db (O(V )). It follows from Theorem 2.3.2 that there is an O(V )-isomorphism

=
ν : O(U ) −→ Db (O(V )) for which the diagram
ι∗
O(V ) - O(U )
@
ηO(V@

= ν
)
@
R
@ ?
Db (O(V )) = O(W )

commutes. By 6.4.2(ii), there is a morphism of varieties γ : U → W such


that γ ∗ = ν −1 , and there is a morphism of varieties β : W → V such that
β ∗ = ηO(V ) . Since (β ◦ γ)∗ = γ ∗ ◦ β ∗ = ι∗ , it also follows from 6.4.2(ii) that
β ◦ γ = ι. Our strategy is to use Lemma 6.4.3 to show that γ is an isomorphism
of varieties, and so our immediate aim is to show that γ is surjective.
Let w ∈ W . Note that β ∗−1 (IW (w)) = IV (β(w)). Since

(β ∗−1 (IW (w)))Db (O(V )) ⊆ IW (w) ⊂ Db (O(V )) = bDb (O(V )),

it follows that b ⊆ β ∗−1 (IW (w)) = IV (β(w)). Therefore β(w) ∈ U . Denote


β(w) by u. We aim to show that γ(u) = w.
Suppose that γ(u) = w, and look for a contradiction. By 6.4.2(vi), there
exists a function g ∈ O(W ) such that g(w) = 0 and g(γ(u)) = 0. Also, since
b ⊆ IV (u), there exists h ∈ b \ IV (u). Now Coker ηO(V ) ∼ = Hb1 (O(V )), by
2.2.6(i)(c), and so this cokernel is b-torsion. Therefore, there is n ∈ N such
that β ∗ (hn )g = ηO(V ) (hn )g = hn g ∈ ηO(V ) (O(V )) = β ∗ (O(V )). Thus
there exists k ∈ O(V ) such that β ∗ (hn )g = β ∗ (k). We shall now calculate
k(u) in two ways. First of all,

k(u) = k(β(w)) = (β ∗ (k))(w) = (β ∗ (hn )g)(w)


= ((β ∗ (h))(w))n g(w) = 0.
126 Fundamental vanishing theorems

On the other hand,

k(u) = k(ι(u)) = k(β ◦ γ(u)) = k(β(γ(u))) = (β ∗ (k))(γ(u))


= (β ∗ (hn )g)(γ(u)) = ((β ∗ (h))(γ(u))) g(γ(u))
n

n
= (h(β ◦ γ(u))) g(γ(u)) = (h(u))n g(γ(u));

this is non-zero by choice of h and g. This contradiction shows that γ(u) = w.


We have therefore proved that γ is surjective, and so we can now use Lemma
6.4.3 to deduce that γ is an isomorphism of varieties, so that U is affine.

6.4.5 Corollary. Let the notation be as in 6.4.1, and assume U = V \ V (b)


is affine. Then V \ U = V (b) is of pure codimension 1 in V .

Proof. By Serre’s Affineness Criterion 6.4.4, the functor Db is exact. There-


fore, by 6.3.6, ht p ≤ 1 for every minimal prime ideal p of b. But b = 0 and
O(V ) is an integral domain, so that ht p = 1 for every minimal prime ideal p
of b. Hence every irreducible component of V (b) has codimension 1 in V .

Our work on Serre’s Affineness Criterion raises the following questions.


We again use the notation of 6.4.1. First, if U = V \ V (b) is affine, so
that bDb (O(V )) = Db (O(V )), then it follows from 6.3.4 that Db (O(V ))
is a finitely generated O(V )-algebra: is the converse statement true, that is, if
Db (O(V )) is a finitely generated O(V )-algebra, is it necessarily the case that
U is affine? Second, is the converse of 6.4.5 true, that is, if V \ U = V (b) is
of pure codimension 1 in V , is it necessarily the case that U is affine? Another
examination of the example studied in 2.3.7 and 3.3.5 will provide us with
negative answers to both questions.

6.4.6 Example. We consider again the affine variety V in A4 studied in 2.3.7


and 3.3.5 and given by

V := VA4 (X1 X4 − X2 X3 , X12 X3 + X1 X2 − X22 , X33 + X3 X4 − X42 ).

As in 3.3.5, let L := VA4 (X2 , X3 , X4 ) = (c, 0, 0, 0) ∈ A4 : c ∈ C and


L := (c, 0) ∈ A2 : c ∈ C . Our argument uses the morphism of varieties
α : A2 → V of 2.3.7 for which α((c, d)) = (c, cd, d(d − 1), d2 (d − 1)) for
all (c, d) ∈ A2 . It was shown in 3.3.5 that the restriction of α provides an

=
isomorphism of (quasi-affine) varieties A2 \ ({(0, 1)} ∪ L) −→ V \ L. Now
L is of pure codimension 1 in V . (As L is actually a subvariety of V , this
statement is equivalent to the statement that codimV L = 1.) If we can show
that U := A2 \ ({(0, 1)} ∪ L) is not affine and that O(U ) is a finitely generated
C-algebra, then it will follow that both questions posed just after 6.4.5 have
6.5 Applications to local algebra in prime characteristic 127

negative answers. Let q := (0, 1) ∈ A2 . As A2 \ U = L ∪ {q} is not of pure


codimension 1 in A2 , it follows from 6.4.5 that U is not affine.
Next, L = VA2 (Y ), so that A2 \ L is affine, and, by 2.3.2 and 2.2.19, we can
identify O(A2 \ L) with the subring C[X, Y ]Y = C[X, Y, Y −1 ] of C(X, Y ).
Note that IA2 \L (q) = (X, Y − 1)C[X, Y, Y −1 ]. (Observe that the maximal
ideal (X, Y − 1) of C[X, Y ] does not meet Y i : i ∈ N0 .) Since IA2 \L (q)
contains a C[X, Y, Y −1 ]-sequence X, Y − 1 of length 2, it follows from 2.2.8
that
ηC[X,Y,Y −1 ] : C[X, Y, Y −1 ] −→ DIA2 \L (q) (C[X, Y, Y −1 ])

is an isomorphism. But, by 2.3.2, and as U = (A2 \ L) \ {q}, there is an


isomorphism

DIA2 \L (q) (C[X, Y, Y −1 ]) ∼


= O((A2 \ L) \ {q}) = O(U )

of O(A2 \ L)-algebras. Hence O(U ) is a finitely generated C-algebra. Also,


the fact that the map ηC[X,Y,Y −1 ] is an isomorphism shows that

(X, Y − 1)DIA2 \L (q) (C[X, Y, Y −1 ]) = DIA2 \L (q) (C[X, Y, Y −1 ]),

and so we see again (this time from Serre’s Affineness Criterion 6.4.4) that the
quasi-affine variety U is not affine.
Thus both questions posed just after 6.4.5 have negative answers.

6.4.7 Exercise. Let the notation be as in 6.4.1. Also, let b be a second non-
zero ideal of O(V ) and let U  denote the quasi-affine variety V \V (b ). Deduce
from Exercise 6.3.9 and Serre’s Affineness Criterion 6.4.4 that, if U and U  are
affine, then U ∩ U  is also affine.

6.5 Applications to local algebra in prime characteristic


In this section, we present some applications of local cohomology to the study
of algebra over local rings of prime characteristic. These applications concern
results that do not involve local cohomology in their statements, but which
have proofs that make non-trivial use of local cohomology.
In this work, we shall use some techniques that assist calculation with reg-
ular sequences, and our first exercises in the section are concerned with these.
The reader should recall the definition of poor M -sequence (where M is an
R-module) given in 6.2.1
128 Fundamental vanishing theorems

6.5.1 Exercise. Let M be an R-module, and let r1 , . . . , rn be a poor M -


sequence, where n ≥ 2. Show that r1 is a non-zerodivisor on
M/(r2 , . . . , rn )M .
6.5.2 Exercise. Let M be an R-module, let r2 , . . . , rn , a, b ∈ R, where
n ≥ 2, and suppose that a, r2 , . . . , rn is a poor M -sequence.
(i) Assume that
abm1 + r2 m2 + · · · + rn mn = am1 + r2 m2 + · · · + rn mn ,
where m1 , . . . , mn , m1 , . . . , mn ∈ M . Show that
m1 ∈ (b, r2 , . . . , rn )M.
(ii) Deduce that, if b, r2 , . . . , rn is also a poor M -sequence, then
ab, r2 , . . . , rn is a poor M -sequence.
6.5.3 Exercise. Let M be an R-module, and r1 , . . . , rn ∈ R, where n ∈ N.
(i) Let i ∈ {1, . . . , n}, and suppose that ri can be written as ri = ab, where
a, b ∈ R. Show that r1 , . . . , ri−1 , ri , ri+1 , . . . , rn is a poor M -sequence
if and only if
r1 , . . . , ri−1 , a, ri+1 , . . . , rn and r1 , . . . , ri−1 , b, ri+1 , . . . , rn
are poor M -sequences.
(ii) Let t1 , . . . , tn be arbitrary positive integers. Show that r1 , . . . , rn is a
poor M -sequence if and only if r1t1 , . . . , rntn is a poor M -sequence.
6.5.4 Notation for the section. Throughout the section, M will denote an
R-module, n will denote a positive integer, and Ln (R) will denote the set of
n × n lower triangular matrices with entries in R. For H ∈ Ln (R), we shall
use |H| to denote the determinant of H, that is, the product of the diagonal
entries of H. We shall use T to denote matrix transpose; displayed matrices
will be shown between rectangular brackets.
Let d1 , . . . , dn ∈ R. We shall use diag(d1 , . . . , dn ) to denote the diagonal
matrix in Ln (R) whose (i, i)-th entry is di (for each i = 1, . . . , n).
6.5.5 Remark. Suppose that x1 , . . . , xn , y1 , . . . , yn ∈ R and H = [hij ] ∈
T T
Ln (R) are such that [y1 · · · yn ] = H [x1 · · · xn ] . The fact that the adjoint
matrix Adj H satisfies (Adj H)H = |H|In ensures that
|H|(x1 , . . . , xn )M ⊆ (y1 , . . . , yn )M.
Multiplication by |H| therefore induces an R-homomorphism
M/(x1 , . . . , xn )M −→ M/(y1 , . . . , yn )M.
6.5 Applications to local algebra in prime characteristic 129

Let k ∈ {1, . . . , n}. Since H is lower triangular, the k × k submatrix Hk of


H obtained by deleting the (k +1)-th,. . ., n-th rows and columns of H satisfies
T T
[y1 · · · yk ] = Hk [x1 · · · xk ] . We deduce from the above paragraph that
|Hk |(x1 , . . . , xk )M ⊆ (y1 , . . . , yk )M , so that
|H|(x1 , . . . , xk )M ⊆ (y1 , . . . , yk )M
because |Hk | is a factor of |H|.
6.5.6 Theorem. (L. O’Carroll [64, Theorem 3.2]) Suppose that x1 , . . . , xn ,
y1 , . . . , yn ∈ R and H ∈ Ln (R) are such that
T T
(i) [y1 · · · yn ] = H [x1 · · · xn ] , and
(ii) y1 , . . . , yn is a poor M -sequence.
Then the R-homomorphism α : M/(x1 , . . . , xn )M −→ M/(y1 , . . . , yn )M
induced by multiplication by |H| is a monomorphism, and x1 , . . . , xn is also
a poor M -sequence.
Proof. We use induction on n. When n = 1, the result is immediate from
6.5.3(i). We therefore suppose that n > 1 and that the result has been proved
for sequences of length smaller than n. Let hij denote the (i, j)-th entry of H
(for all i, j = 1, . . . , n).
Now [x1 y2 · · · yn ] = H  [x1 x2 · · · xn ] , where H  = [hij ] ∈ Ln (R) is
T T

specified as follows:

1 if i = j = 1,
hij =
hij otherwise.
Let β : M/(x1 , x2 , . . . , xn )M −→ M/(x1 , y2 , . . . , yn )M be the R-homom-
orphism induced by multiplication by |H  |. Note also that
T T
[y1 y2 · · · yn ] = D [x1 y2 · · · yn ] where D := diag(h11 , 1, . . . , 1).
Let γ : M/(x1 , y2 , . . . , yn )M −→ M/(y1 , y2 , . . . , yn )M be the R-homom-
orphism induced by multiplication by |D| = h11 . Since |H| = |D||H  |, we
have γ ◦ β = α. In order to show that α is a monomorphism, it is therefore
sufficient for us to show that both γ and β are monomorphisms.
Since y1 = h11 x1 , we see by 6.5.3(i) that h11 , y2 , . . . , yn and x1 , y2 , . . . , yn
are poor M -sequences. Let m ∈ M be such that h11 m ∈ (y1 , y2 , . . . , yn )M ;
thus h11 m = y1 m1 + · · · + yn mn for some m1 , . . . , mn ∈ M . Since y1 =
h11 x1 , we obtain h11 (m − x1 m1 ) ∈ (y2 , . . . , yn )M . Since h11 , y2 , . . . , yn is
a poor M -sequence, we can use 6.5.1 to see that h11 is a non-zerodivisor on
M/(y2 , . . . , yn )M ; hence m − x1 m1 ∈ (y2 , . . . , yn )M . It follows that γ is a
monomorphism.
130 Fundamental vanishing theorems

We now turn our attention to β. Set R = R/x1 R and M = M/x1 M ;


for r ∈ R, denote the natural image of r in R by r. We noted in the last
paragraph that x1 , y2 , . . . , yn is a poor M -sequence. Therefore y2 , . . . , yn is
T T
a poor M -sequence in R. Moreover [y2 · · · yn ] = G [x2 · · · xn ] , where
G = [gij ] ∈ Ln−1 (R) is given by gij = hi+1,j+1 for all i, j ∈ {1, . . . , n − 1}.
Let β : M /(x2 , . . . , xn )M −→ M /(y2 , . . . , yn )M denote the R-homom-
orphism induced by multiplication by |G| = |H  |. By the inductive hypothesis,
β is a monomorphism. An easy calculation then shows that β is a monomor-
phism. Also, the facts that y1 = h11 x1 and y1 is a non-zerodivisor on M
ensure that x1 is a non-zerodivisor on M . The inductive hypothesis yields that
x2 , . . . , xn is a poor M -sequence, and so we can conclude that x1 , . . . , xn is a
poor M -sequence.
This completes the inductive step, and the proof.
6.5.7 Corollary. Suppose that a1 , . . . , an are n elements of R that generate
a, and let M be an R-module. Set a := a1 . . . an . Use the notation of 5.3.1
to denote natural images of elements of Ma in the n-th cohomology module of
the C̆ech complex of M with respect to a1 , . . . , an .
Suppose that a1 , . . . , an is a poor M -sequence, and that m ∈ M and i ∈ N
are such that [m/ai ] = 0. Then m ∈ (ai1 , . . . , ain )M .
n
Proof. By 5.3.2, there is k ∈ N0 such that k ≥ i and ak−i m ∈ u=1 aku M .
" i #
i T
" #T
1 , . . . , an ) a 1 · · · a n
Now diag(ak−i = ak1 · · · akn and ak1 , . . . , akn is a
k−i

poor M -sequence, by 6.5.3(ii). Since


| diag(ak−i k−i
1 , . . . , an )|m = a
k−i
m ∈ (ak1 , . . . , akn )M,
it follows from O’Carroll’s Theorem 6.5.6 that m ∈ (ai1 , . . . , ain )M .
6.5.8 Remark. Suppose that (R, m) is local with dim R = n > 0. Recall
that a system of parameters for R is a sequence (ri )ni=1 of n elements of m
such that the ideal (r1 , . . . , rn ) of R generated by the terms of the sequence is
m-primary. By a subsystem of parameters for R we mean a sequence (ri )ti=1 of
t elements of m, with t ≤ n, which can be extended to a system of parameters
for R by the addition of n − t extra terms.
Let (ai )ni=1 be a system of parameters for R.
When R is Cohen–Macaulay, so that a1 , . . . , an is an R-sequence, we must
have (a1 . . . an )k ∈ (ak+1 1 n ) for all k ∈ N0 , as we now show. If
, . . . , ak+1
(a1 . . . an ) ∈ (a1 , . . . , ak+1
k k+1
n ), then, since ak+1
1 , . . . , ak+1
n is an R-sequence
by 6.5.3(ii), we can use O’Carroll’s Theorem 6.5.6 in conjunction with the
equation
T " #T
diag(ak1 , . . . , akn ) [a1 · · · an ] = ak+11 · · · ak+1
n
6.5 Applications to local algebra in prime characteristic 131

to deduce that 1 ∈ (a1 , . . . , an ), a contradiction.


This observation leads to the following famous conjecture.
6.5.9 The Monomial Conjecture. (See M. Hochster [37, Conjecture 1].)
Suppose that (R, m) is local (but not necessarily Cohen–Macaulay) and that
dim R = n > 0. The conjecture that (a1 . . . an )k ∈ (ak+1
1 , . . . , ak+1
n ) for all
k ∈ N0 and all systems of parameters (ai )i=1 for R is known as the Monomial
n

Conjecture.
6.5.10 Theorem. (See M. Hochster [37, pp. 33–34].) Suppose that (R, m) is
local with dim R = n > 0, and has prime characteristic p. Then the conclu-
sion of the Monomial Conjecture is true in R.
Proof. Let (ai )ni=1 be a system of parameters for R. Suppose there exists
k ∈ N0 such that (a1 . . . an )k ∈ (ak+1
1 , . . . , ak+1
n ), and seek a contradiction.
n
Set a := a1 . . . an , and represent elements of Hm (R) using a1 , . . . , an and
a in the manner described in 5.3.1. Let F denote the Frobenius action on
Hmn
(R) of 5.3.4 and 5.3.6. Let y := [1/a] ∈ Hm n
(R); then the supposition
that a ∈ (a1 , . . . , an ) means that there exist r1 , . . . , rn ∈ R such that
k k+1 k+1

ak = ak+1
1 r1 + · · · + ak+1
n rn , so that, by the comments in 5.3.1,
$ % $ k % ( k+1 )
1 a a1 r1 + · · · + ak+1 n rn
y= = k+1 = = 0.
a a ak+1
n
But an arbitrary element z of Hm (R) can be expressed as z = [r/at ] for some
r ∈ R and t ∈ N0 . Choose h ∈ N such that ph ≥ t. On use of 5.3.6 we now
see that
( h ) $ %
&r' ap −t r 1
ph −t
= ap −t rF h (y) = 0.
h
z= t = p h = a r p h
a a a
n
We have therefore shown that Hm (R) = 0, contrary to the Non-vanishing
Theorem 6.1.4.
Hochster’s theorem above provides one example of an important result in
local algebra whose statement makes no mention of local cohomology but for
which local cohomology provides a proof. Below we present another such re-
sult, from the theory of tight closure. Some definitions are needed.
6.5.11 Definitions. Suppose that R has prime characteristic p. We use R◦
to denote the complement in R of the union of the minimal prime ideals of
R. An element r ∈ R belongs to the tight closure a∗ of a if and only if there
exists c ∈ R◦ such that crp ∈ a[p ] for all n  0. We say that a is tightly
n n

closed precisely when a∗ = a. The theory of tight closure was invented by M.


132 Fundamental vanishing theorems

Hochster and C. Huneke [38], and many applications have been found for the
theory: see [41] and [42], for example.
The next exercise establishes some basic properties of tight closure.
6.5.12 Exercise. Suppose that R has prime characteristic p.
(i) Show that the tight closure a∗ of a is an ideal of R.
(ii) Let b be a second ideal of R with a ⊆ b. Show that a∗ ⊆ b∗ .
(iii) Show that (a∗ )∗ = a∗ , so that tight closure really is a ‘closure operation’.
6.5.13 Exercise. Suppose that R has prime characteristic p, and let a, b be
ideals of R.
∗ ∗ ∗ ∗ ∗ ∗ ∗ ∗
(i) Show that (a + b)
√ = (a + b ) and (ab) √ = ∗(a b ) .

(ii) Show that 0 = 0,√ and conclude that 0 ⊆ c for every ideal c of R.
(iii) Let π : R −→ R/ 0 be the natural ring homomorphism. Show that
 √ ∗
a∗ = π −1 π(a)(R/ 0) .
(iv) Show that, if a is tightly closed, then so too is (a : b).
(v) Show that every maximal ideal of R is tightly closed.
6.5.14 Exercise. Let K be a field of prime characteristic p, and let R be the
subring of the ring of formal power series K[[X]] given by
∞
i=0 ai X ∈ K[[X]] : a1 = 0, a0 , a2 , a3 , . . . ∈ K .
i
R :=
In R, calculate (X 2 R)∗ , (X 3 R)∗ , (X 2 R)∗ ∩ (X 3 R)∗ , (X 2 R ∩ X 3 R)∗ ,
((X 2 R)∗ : (X 3 R)∗ ) and (X 2 R : X 3 R)∗ .
6.5.15 Proposition. Suppose that (R, m) is local, of prime characteristic p,
and Cohen–Macaulay with dim R = n > 0. Let (ai )ni=1 be a system of pa-
rameters for R, and suppose that the ideal q := (a1 , . . . , an ) is tightly closed.
Then the ideal (at11 , . . . , atnn ) is tightly closed for all t1 , . . . , tn ∈ N.
n
Proof. We argue by induction on t := i=1 ti . Suppose that t > n and that
the desired result has been proved for smaller values of t.
Without loss of generality, we can, and do, assume that t1 > 1. By our induc-
tive hypothesis, (a1t1 −1 , at22 , . . . , atnn ) is tightly closed. Let r ∈ (at11 , . . . , atnn )∗ .
Then r ∈ (a1t1 −1 , at22 , . . . , atnn )∗ = (a1t1 −1 , at22 , . . . , atnn ), and so there exist
r1 , . . . , rn ∈ R such that
r = r1 a1t1 −1 + r2 at22 + · · · + rn atnn .
Now r2 at22 + · · · + rn atnn ∈ (at11 , . . . , atnn ) ⊆ (at11 , . . . , atnn )∗ , and so
r1 at11 −1 = r − (r2 at22 + · · · + rn atnn ) ∈ (at11 , . . . , atnn )∗ .
6.5 Applications to local algebra in prime characteristic 133

Therefore, there exists c ∈ R◦ such that


 p j j
c r1 at11 −1
j
∈ at11 p , . . . , atnn p for all j  0.
j j
Now at11 p , . . . , atnn p (for j ∈ N) is an R-sequence, and the diagonal matrix
(t −1)pj
D := diag(a1 1 , 1, . . . , 1) ∈ Ln (R)
& j j
'
j T
& j j
'
j T
satisfies D ap1 at22 p · · · atnn p = at11 p at22 p · · · atnn p . We can there-
fore use O’Carroll’s Theorem 6.5.6 to deduce that
j j j j
cr1p ∈ ap1 , at22 p , . . . , atnn p for all j  0.

Therefore r1 ∈ (a1 , at22 , . . . , atnn )∗ . By the inductive hypothesis, the ideal


(a1 , at22 , . . . , atnn ) is tightly closed, so that r1 ∈ (a1 , at22 , . . . , atnn ). Therefore
r = r1 a1t1 −1 + r2 at22 + · · · + rn atnn ∈ (at11 , at22 , . . . , atnn ).
Thus (at11 , at22 , . . . , atnn ) is tightly closed. This completes the inductive step,
and the proof.
6.5.16 Theorem. (See R. Fedder and K.-i. Watanabe [17, Proposition 2.2].)
Suppose that (R, m) is local, of prime characteristic p, and Cohen–Macaulay
with dim R = n > 0. Suppose that there is one system of parameters (ai )ni=1
for R which generates a tightly closed ideal. Then the ideal of R generated by
each system of parameters for R is tightly closed, as is the ideal generated by
each subsystem of parameters for R.
Proof. First of all, by 6.5.15, the ideal (at11 , . . . , atnn ) is tightly closed for all
t1 , . . . , tn ∈ N. Let F denote the Frobenius action on Hm n
(R) of 5.3.4 and
5.3.6. We plan to prove that, if y ∈ Hm (R) is such that there exists c ∈ R◦
n

with cF j (y) = 0 for all j  0, then y = 0.


n n
Represent elements of Hm (R) = H(a 1 ,...,an )
(R) using a1 , . . . , an and a :=
a1 . . . an in the manner described in 5.3.1. Thus y ∈ Hm n
(R) can be written
as y = [r/a ] for some r ∈ R and t ∈ N0 . Suppose that there exists c ∈ R◦
t
j j
with cF j (y) = 0 for all j  0. By 5.3.6, this means that [crp /atp ] = 0
for all j  0. Since a1 , . . . , an is an R-sequence, we can deduce from 6.5.7
j
j
tpj t ∗
that crp ∈ (atp 1 , . . . , an ) for all j  0. Therefore r ∈ (a1 , . . . , an ) .
t

Since (a1 , . . . , an ) is tightly closed, we see that r ∈ (a1 , . . . , an ), so that


t t t t

y = [r/at ] = 0 by the comments in 5.3.1.


Let (bi )ni=1 be an arbitrary system of parameters for R. We show now that
the ideal (b1 , . . . , bn ) is tightly closed. Let s ∈ (b1 , . . . , bn )∗ , so that there
j
exists c ∈ R◦ such that c sp ∈ (bp1 , . . . , bpn ) for all j  0. Represent
j j
134 Fundamental vanishing theorems
n n
elements of Hm (R) = H(b 1 ,...,bn )
(R) using b1 , . . . , bn and b := b1 . . . bn in
the manner described in 5.3.1, and consider z := [s/b] ∈ Hm n
(R). At this
n
point, it is important to recall from 5.3.5 that the Frobenius action F on Hm (R)
does not depend on any choice of generators for any m-primary ideal of R.
Therefore, using 5.3.6 and 5.3.1 again, we can deduce that
( )
c sp
j
 j
c F (z) = = 0 for all j  0
bpj
j
because c sp ∈ (bp1 , . . . , bpn ) for all j  0. The claim proved in the im-
j j

mediately preceding paragraph therefore shows that z = [s/b] = 0, so that


s ∈ (b1 , . . . , bn ) by 6.5.7 because b1 , . . . , bn is an R-sequence. Therefore
(b1 , . . . , bn ) is tightly closed.
To complete the proof, let i ∈ {0, 1, . . . , n − 1} and set c := (b1 , . . . , bi );
we have to show that c is tightly closed. (Interpret c as 0 when i = 0.) By the
immediately preceding paragraph, the ideal (b1 , . . . , bi , bti+1 , . . . , btn ) is tightly
closed for all t ∈ N. Let v ∈ c∗ ; then v belongs to the tight closure of every
ideal of R that contains c (by 6.5.12(ii)). Hence
* *
v∈ (b1 , . . . , bi , bti+1 , . . . , btn )∗ = (b1 , . . . , bi , bti+1 , . . . , btn )
t∈N t∈N
*  * 
⊆ (b1 , . . . , bi ) + m t
= c + mt = c
t∈N t∈N

by Krull’s Intersection Theorem. Therefore c is tightly closed, and the proof is


complete.
6.5.17 Remark. Suppose that (R, m) is local and of prime characteristic.
We say that R is F -rational if and only if every proper ideal c of R which
can be generated by ht c elements is tightly closed. The Theorem 6.5.16 of
Fedder and Watanabe shows that a Cohen–Macaulay local ring of prime char-
acteristic is F -rational if one single system of parameters generates a tightly
closed ideal. It can be proved that, if R is excellent, or a homomorphic image
of a Cohen–Macaulay local ring, and is F -rational, then it must be Cohen–
Macaulay; however, that result is beyond the scope of this book. We have in-
cluded 6.5.16 because, firstly, it gives another example of a significant result
in local algebra whose statement makes no mention of local cohomology, but
for which local cohomology provides a proof, and, secondly, because it gives
some hints about the important rôle that local cohomology can play in tight
closure theory.
7

Artinian local cohomology modules

In this chapter, we shall show that certain local cohomology modules are
Artinian, that is, satisfy the descending chain condition on submodules, and
we shall use our results to provide a different proof of a theorem in Chapter 6.
Suppose, temporarily, that (R, m) is local. In the Non-vanishing Theorem
6.1.4, we proved that, if M is a non-zero, finitely generated R-module of di-
n
mension n, then Hm (M ) = 0. One consequence of our work in this chapter
is that we can give an alternative proof of this result, and, at the same time,
obtain more information than we deduced in Chapter 6. Our approach in this
n
chapter exploits the facts that the R-module Hm (M ) is actually Artinian, and
that, for Artinian modules over commutative rings, there is available a theory
of secondary representation that is, in several respects, dual to the theory of
primary decomposition of Noetherian modules over commutative rings.

7.1 Artinian modules


We begin with some arguments due to L. Melkersson ([51], [52]) which will
enable us to show that, for a non-zero, finitely generated module M of dimen-
sion n over the local ring (R, m), the n-th local cohomology module Han (M )
of M with respect to a is Artinian, and all the local cohomology modules
Hmi
(M ) (i ∈ N0 ) of M with respect to m are Artinian.
We revert to our standard hypotheses concerning R and a (although it is
worth pointing out that the first two results below (due to Melkersson) actually
hold under weaker hypotheses).

7.1.1 Lemma. (L. Melkersson [51, Lemma 2.1]). Let a ∈ R and let M be
an aR-torsion R-module. Suppose that N, N  are submodules of M such that
N  ⊆ N and ai (0 :N ai+1 ) = ai (0 :N  ai+1 ) for all i ∈ N0 . Then N = N  .
136 Artinian local cohomology modules

Proof. Since each element of N is annihilated by some power of a, it is


enough for us to show that (0 :N ai ) ⊆ N  for all i ∈ N. We prove this
by induction on i. By hypothesis, (0 :N a) = (0 :N  a) ⊆ N  , and so we
assume that, for i ∈ N, we have proved that (0 :N ai ) ⊆ N  .
Let z ∈ (0 :N ai+1 ). Then ai z ∈ ai (0 :N ai+1 ) = ai (0 :N  ai+1 ), and so
there exists z  ∈ (0 :N  ai+1 ) such that ai z = ai z  . Hence z − z  ∈ (0 :N ai ),
and so z − z  ∈ N  by the inductive assumption. Therefore z = (z − z  ) + z  ∈
N  , and the inductive step is complete.
7.1.2 Theorem. (L. Melkersson [51, Theorem 1.3]). Assume that M is an
a-torsion R-module for which (0 :M a) is Artinian. Then M is Artinian.
Proof. We suppose that a can be generated by t elements and proceed by
induction on t. When t = 0, we have a = 0 and (0 :M a) = M , so that there
is nothing to prove in this case.
Now suppose that t = 1 and a = Ra for a ∈ R. Let
L1 ⊇ L2 ⊇ · · · ⊇ Ln ⊇ Ln+1 ⊇ · · ·
be a descending chain of submodules of M . Observe that, for each i ∈ N0 and
each submodule L of M , we have ai (0 :L ai+1 ) ⊆ (0 :M a). In fact, for each
n ∈ N,
(0 :Ln a) ⊇ · · · ⊇ ai (0 :Ln ai+1 ) ⊇ ai+1 (0 :Ln ai+2 ) ⊇ · · ·
is a descending chain of submodules of the Artinian R-module (0 :M a), and
so is eventually stationary: let En denote its eventual stationary value, so that
there is kn ∈ N such that ai (0 :Ln ai+1 ) = En for all i ≥ kn .
Since
En = akn +kn+1 (0 :Ln akn +kn+1 +1 )
⊇ akn +kn+1 (0 :Ln+1 akn +kn+1 +1 ) = En+1 ,
we see that E1 ⊇ E2 ⊇ · · · ⊇ En ⊇ En+1 ⊇ · · · is a descending chain of
submodules of the Artinian R-module (0 :M a), and so there exists t ∈ N such
that En = Et for all n ≥ t. Now for all i, n ∈ N with n ≥ t and i ≥ kt , we
have
Et = akt (0 :Lt akt +1 ) ⊇ ai (0 :Lt ai+1 ) ⊇ ai (0 :Ln ai+1 ) ⊇ En = Et .
Thus ai (0 :Ln ai+1 ) = Et for all n ≥ t and i ≥ kt .
We plan to use Lemma 7.1.1. With this in mind, we consider, for each integer
i = 0, 1, 2, . . . , kt − 1, the descending chain
ai (0 :L1 ai+1 ) ⊇ · · · ⊇ ai (0 :Ln ai+1 ) ⊇ ai (0 :Ln+1 ai+1 ) ⊇ · · ·
7.1 Artinian modules 137

of submodules of the Artinian R-module (0 :M a): there exists u ∈ N with


u ≥ t such that ai (0 :Ln ai+1 ) = ai (0 :Lu ai+1 ) for all n ≥ u and all
i = 0, 1, 2, . . . , kt − 1.
We now have that, for each integer n ≥ u,

ai (0 :Ln ai+1 ) = ai (0 :Ln+1 ai+1 ) for all i ∈ N0 :

this equation is true for all i = 0, 1, 2, . . . , kt − 1 by the last paragraph and for
all i ≥ kt by the paragraph before that. Therefore, by Lemma 7.1.1, we have
Ln = Ln+1 for all n ≥ u . It follows that M is Artinian.
Now suppose, inductively, that t > 1 and the result has been proved for ide-
als that can be generated by fewer than t elements. Suppose that a is generated
by t elements a1 , . . . , at . Set b = Ra1 + · · · + Rat−1 and N = (0 :M b). Then
N is Rat -torsion and (0 :N at ) = (0 :M a). It therefore follows from what we
have already proved in the case where t = 1 that N = (0 :M b) is Artinian.
Since M is, of course, b-torsion and b can be generated by t − 1 elements, we
can apply the inductive hypothesis to deduce that M is Artinian. The inductive
step is complete.

We immediately exploit Theorem 7.1.2 to prove that certain local cohomol-


ogy modules are Artinian. The proofs presented below of the next two theo-
rems are due to L. Melkersson (see [52, Theorems 2.1 and 2.2]), although the
results themselves are somewhat older.

7.1.3 Theorem. Assume that (R, m) is local, and let M be a finitely gener-
i
ated R-module. Then the R-module Hm (M ) is Artinian for all i ∈ N0 .

Proof. We use induction on i. First, Γm (M ) is a finitely generated R-module


0
annihilated by a power of m; hence Hm (M ) has finite length.
Now suppose, inductively, that i > 0 and we have shown that Hm i−1
(M  ) is

Artinian for every finitely generated R-module M . Now
i
Hm (M ) ∼ i
= Hm (M/Γm (M ))

by 2.1.7(iii). Also, 2.1.2 shows that M/Γm (M ) is m-torsion-free. We therefore


assume in addition that M is an m-torsion-free R-module.
We now use 2.1.1(ii) to deduce that m contains an element r which is a
non-zerodivisor on M . The exact sequence
r
0 −→ M −→ M −→ M/rM −→ 0
r
induces an exact sequence Hm i−1
(M/rM ) −→ Hm i
(M ) −→ Hm i
(M ) of lo-
cal cohomology modules. Since M/rM is a finitely generated R-module, it
i−1
follows from the inductive hypothesis that Hm (M/rM ) is Artinian, so that,
138 Artinian local cohomology modules

in view of the above exact sequence, the R-module (0 :Hm i (M ) r) is Artinian.


i
Since Hm (M ) is m-torsion and therefore Rr-torsion, it follows from Theorem
i
7.1.2 that Hm (M ) is Artinian. The inductive step is complete.

The following exercise generalizes Theorem 7.1.3 to non-local situations.

7.1.4 Exercise. Assume that R/a is Artinian, and let M be a finitely gener-
ated R-module. Show that Hai (M ) is Artinian for all i ∈ N0 .

7.1.5 Exercise. Suppose that m is a maximal ideal of R, and that q is an


m-primary ideal of R. Let M be an R-module, and let i ∈ N0 .

(i) Let s ∈ R \ m. Show that multiplication by s provides an automorphism


of Hqi (M ), so that this local cohomology module has a natural structure
as an Rm -module.
(ii) Show that a subset of Hqi (M ) is an Rm -submodule if and only if it is an
R-submodule.
(iii) Deduce from (i), (ii) and Theorem 7.1.3 that, when M is finitely gener-
ated, Hqi (M ) is an Artinian R-module.

7.1.6 Theorem. Assume (R, m) is local, and let M be a non-zero, finitely


generated R-module of dimension n. Then the R-module Han (M ) is Artinian.

Proof. We use induction on n. If n = 0, then (0 : M ) = m, so that M is
annihilated by a power of m and so has finite length; therefore its submodule
Γa (M ) also has finite length.
Now suppose, inductively, that n > 0 and we have established the result
for (non-zero, finitely generated) R-modules of dimension smaller than n.
Now Han (M ) ∼ = Han (M/Γa (M )) by 2.1.7(iii). If dim(M/Γa (M )) < n, then
n
Ha (M/Γa (M )) = 0 by Grothendieck’s Vanishing Theorem 6.1.2, and there
is nothing to prove. Since M/Γa (M ) is a-torsion-free (by 2.1.2), we therefore
make the additional assumption that M is an a-torsion-free R-module.
The argument now proceeds like that used in the proof of Theorem 7.1.3
above. By 2.1.1(ii), the ideal a contains an element r which is a non-zerodivisor
r
on M . The exact sequence 0 −→ M −→ M −→ M/rM −→ 0 induces an
r
exact sequence Ha (M/rM ) −→ Han (M ) −→ Han (M ). Since r is not in
n−1

any minimal member of Supp M , we have dim(M/rM ) ≤ n − 1, so that,


by the inductive hypothesis (or Grothendieck’s Vanishing Theorem 6.1.2), the
R-module Han−1 (M/rM ) is Artinian. Therefore, in view of the above exact
sequence, (0 :Han (M ) r) is Artinian. Since Han (M ) is Rr-torsion, it follows
from Theorem 7.1.2 that Han (M ) is Artinian. The inductive step is complete.
7.2 Secondary representation 139

It is perhaps worth pointing out that, in our applications of Melkersson’s


Theorem 7.1.2 in the proofs of 7.1.3 and 7.1.6, we have only used the special
case of 7.1.2 in which a is principal. However, this is perhaps not surprising,
as most of our proof of 7.1.2 was devoted to the case when a is principal.
The following exercise generalizes Theorem 7.1.6 to non-local situations.
7.1.7 Exercise. Let M be a non-zero, finitely generated R-module of finite
dimension n. Show that the R-module Han (M ) is Artinian.

7.2 Secondary representation


We intend to exploit Theorem 7.1.3 to provide another proof of the Non-
vanishing Theorem 6.1.4, which states that, if M is a non-zero, finitely gener-
n
ated module of dimension n over the local ring (R, m), then Hm (M ) = 0. We
are going to use the theory of secondary representations of Artinian modules.
As this theory, although mentioned in Matsumura [50, Section 6, Appendix], is
not as well known as the theory of primary decomposition, we shall guide the
reader through the main points by means of a series of exercises. Although we
shall maintain our standard hypothesis that R is Noetherian, the reader might
be interested to learn that this condition is not strictly necessary for the devel-
opment of a worthwhile theory: see D. Kirby [44], I. G. Macdonald [45] or D.
G. Northcott [62].
7.2.1 Definitions and Exercise. Let S be an R-module. We say that S is
secondary precisely when S = 0 and, for each r ∈ R, either rS = S or
there exists n ∈ N such that rn S = 0. Show that, when this is the case,

p := (0 :R S) is a prime ideal of R: in these circumstances, we say that S is
a p-secondary R-module.
Show that a non-zero homomorphic image of a p-secondary R-module is
again p-secondary.
Show that, if S1 , . . . , Sn (where n ∈ N) are p-secondary submodules of an
n
R-module M , then so too is i=1 Si .
7.2.2 Definitions and Exercise. Let M be an R-module. A secondary rep-
resentation of M is an expression for M as a sum of finitely many secondary
submodules of M . Such a secondary representation
M = S1 + · · · + Sn with Si pi -secondary (1 ≤ i ≤ n)
of M is said to be minimal precisely when
(i) p1 , . . . , pn are n different prime ideals of R; and
140 Artinian local cohomology modules
n
(ii) for all j = 1, . . . , n, we have Sj ⊆ Si .
i=1
i=j

We say that M is a representable R-module precisely when it has a sec-


ondary representation. As the sum of the empty family of submodules of an
R-module is zero, we shall regard a zero R-module as representable.
Show that a representable R-module has a minimal secondary representa-
tion.
7.2.3 Exercise. The First Uniqueness Theorem. Let M be a representable
R-module and let
M = S1 + · · · + Sn with Si pi -secondary (1 ≤ i ≤ n)
and
M = S1 + · · · + Sn  with Si pi -secondary (1 ≤ i ≤ n )
be two minimal secondary representations of M . Prove that n = n and
{p1 , . . . , pn } = {p1 , . . . , pn } .
(Here is hint: show that, for p ∈ Spec(R), it is the case that p is one of
p1 , . . . , pn if and only if there is a homomorphic image of M which is p-
secondary.)
7.2.4 Definition. Let M be a representable R-module and let
M = S1 + · · · + Sn with Si pi -secondary (1 ≤ i ≤ n)
be a minimal secondary representation of M . Then the n-element set
{p1 , . . . , pn } ,
which is independent of the choice of minimal secondary representation of M
by 7.2.3, is called the set of attached prime ideals of M and denoted by Att M
or AttR M . The members of Att M are referred to as the attached prime ideals
or the attached primes of M .
7.2.5 Exercise. Suppose that the R-module M is representable, and let p ∈
Spec(R). Use the Noetherian property of R to show that p ∈ Att M if and
only if there is a homomorphic image of M which has annihilator equal to p.
7.2.6 Exercise. Let 0 −→ L −→ M −→ N −→ 0 be an exact sequence of
representable R-modules and R-homomorphisms. Prove that
Att N ⊆ Att M ⊆ Att L ∪ Att N.
7.2 Secondary representation 141

7.2.7 Exercise. The Second Uniqueness Theorem. Let M be a represent-


able R-module and let
M = S1 + · · · + Sn with Si pi -secondary (1 ≤ i ≤ n)
be a minimal secondary representation of M .
Suppose that pj is a minimal member of {p1 , . . . , pn } with respect to inclu-
sion. Prove that Sj = r∈R\pj rM .
In the light of the First Uniqueness Theorem 7.2.3, this means that, in a
minimal secondary representation of M , each secondary term corresponding
to a minimal member of Att M is uniquely determined by M and independent
of the choice of minimal secondary representation.
In order to make use of this theory of secondary representation, we shall
need the fact that every Artinian R-module is representable. This fact is the
subject of the next two exercises.
7.2.8 Definition and Exercise. We say that an R-module N is sum-irreduc-
ible precisely when it is non-zero and cannot be expressed as the sum of two
proper submodules of itself.
Prove that an Artinian sum-irreducible R-module is secondary.
7.2.9 Exercise. Let A be an Artinian R-module. Show that A can be ex-
pressed as a sum of finitely many sum-irreducible submodules, and deduce
from Exercise 7.2.8 above that A is representable.
We can therefore form the finite set Att A. Note that Att A = ∅ if and only
if A = 0.
Actually, the class of R-modules which possess secondary representations
is, in general, larger than the class of Artinian R-modules: this is illustrated by
the next exercise.
7.2.10 Exercise. Let E be an injective R-module.
(i) Suppose that Q is an R-module with the property that its zero submodule
is a p-primary submodule of Q. Prove that HomR (Q, E), if non-zero, is
p-secondary.
(ii) Let M be a finitely generated R-module. Prove that HomR (M, E) is
representable, and that AttR (HomR (M, E)) ⊆ AssR M .
(iii) Deduce that E is representable and that AttR E ⊆ Ass R = ass 0.
(iv) The injective R-module E is said to be an injective cogenerator for R
precisely when HomR (N, E) = 0 for every non-zero R-module N .
Prove that, when this is the case, AttR (HomR (M, E)) = AssR M for
every finitely generated R-module M .
142 Artinian local cohomology modules

7.2.11 Proposition. Let A be an Artinian R-module, and let r ∈ R. Then



(i) rA = A if and only if r ∈ R \ p∈Att A p; and

(ii) (0 : A) = p∈Att A p.

Proof. Clearly, we can assume that A = 0, since Att 0 = ∅. Let


A = S1 + · · · + Sn with Si pi -secondary (1 ≤ i ≤ n)
be a minimal secondary representation of M .

(i) Suppose that r ∈ R \ p∈Att A p; then rSi = Si for all i = 1, . . . , n,
and so rA = A. On the other hand, if r ∈ pj for some j with 1 ≤ j ≤ n, then
rh Sj = 0 for a sufficiently large integer h, and so
n
r h A = r h S1 + · · · + r h Sn ⊆ Si ⊂ A.
i=1
i=j
√ n √ n
(ii) To prove this, just note that (0 : A) = i=1 (0 : Si ) = i=1 pi .
Part (i) of 7.2.11 provides an Artinian analogue of the well-known fact that,
if N is a Noetherian R-module and r ∈ R, then r is a non-zerodivisor on N if
and only if r lies outside all the associated prime ideals of N .
There is more to the theory of secondary representation than the brief out-
line presented above: the interested reader should consult the references cited
earlier for this topic, especially [45]. We have presented little more than the
part of the theory we shall need to use. Its relevance for us lies in the fact that,
when (R, m) is local, and M is a finitely generated R-module, then, for each
i ∈ N0 , the local cohomology module Hm i
(M ) is, by 7.1.3, Artinian, and so we
i
can form the finite set of prime ideals Att(Hm (M )) and use the theory of sec-
ondary representation for these modules. In the subsequent work, we shall be
j
interested in whether, for particular j, the local cohomology module Hm (M )
is finitely generated (and so, since it is in any case Artinian, of finite length).
With this in mind, we present here one more result about secondary represen-
tation before going on to apply the theory to local cohomology modules.

7.2.12 Corollary. Let (R, m) be local and let A be an Artinian R-module.


Then A is finitely generated, and so of finite length, if and only if Att A ⊆ {m}.

Proof. (⇒) When A is of finite length, there exists h ∈ N such that mh A = 0,


so that A is either 0 or m-secondary.
(⇐) If Att A ⊆ {m}, then, by 7.2.11(ii), there exists h ∈ N such that
mh A = 0, and it then follows from, for example, [81, 7.30] that A has finite
length.
7.3 The Non-vanishing Theorem again 143

7.3 The Non-vanishing Theorem again


Our first lemma of this section is in preparation for the main theorem of this
chapter. Recall our convention that the dimension of the zero R-module is −1.

7.3.1 Lemma. Let (R, m) be local and let M be a non-zero, finitely generated
R-module of dimension n. Then the set

Σ := {N  : N  is a submodule of M and dim N  < n}

has a largest element with respect to inclusion, N say. Set G := M/N . Then

(i) dim G = n;
(ii) G has no non-zero submodule of dimension less than n;
(iii) Ass G = {p ∈ Ass M : dim R/p = n}; and
(iv) Hmn
(G) ∼
= Hmn
(M ).

Proof. Since M is a Noetherian R-module, the set Σ has a maximal member,


N say. Since the sum of any two members of Σ is again in Σ, it follows that N
contains every member of Σ, and so is the largest element of Σ.
(i) It follows from the canonical exact sequence

0 −→ N −→ M −→ G −→ 0

that Supp G ⊆ Supp M , and that any p ∈ Ass M with dim R/p = n must
belong to Supp G since it cannot belong to Supp N . Hence dim G = n and
{p ∈ Ass M : dim R/p = n} ⊆ Ass G.
(ii) Suppose that L is a submodule of M such that N ⊆ L ⊆ M and
dim(L/N ) < n. Consideration of the canonical exact sequence

0 −→ N −→ L −→ L/N −→ 0

shows that dim L < n; hence L ⊆ N and L/N = 0.


(iii) Let p ∈ Ass G. By (ii), dim R/p = n, so that, since Supp G ⊆
Supp M , we must have p ∈ Ass M . Thus

Ass G ⊆ {p ∈ Ass M : dim R/p = n} .

Since the reverse inclusion was established in our proof of (i) above, we have
completed the proof of (iii).
(iv) Since dim N < n, it follows from Grothendieck’s Vanishing Theorem
n n+1
6.1.2 that Hm (N ) = Hm (N ) = 0. The claim therefore follows from the
long exact sequence of local cohomology modules (with respect to m) that
results from the exact sequence 0 −→ N −→ M −→ G −→ 0.
144 Artinian local cohomology modules

The next theorem is the main result of this chapter. With it, we not only
offer an alternative proof of the result of the Non-vanishing Theorem 6.1.4 (as
promised just before 6.1.3), but we also provide more information than was
given in Theorem 6.1.4. (We should perhaps point out that our proof below in
7.3.2 does quote from Theorem 6.2.7 the fact that, if M is a non-zero, finitely
d
generated module of depth d over the local ring (R, m), then Hm (M ) = 0;
however, no use was made of 6.1.4 in our proof of 6.2.7, and, indeed, the
argument given in the proof of 6.2.7 is elementary.)

7.3.2 Theorem. Assume (R, m) is local, and let M be a non-zero, finitely


n
generated R-module of dimension n. Then Hm (M ) = 0 and
n
Att(Hm (M )) = {p ∈ Ass M : dim R/p = n} .

Proof. (This proof is due to I. G. Macdonald and R. Y. Sharp [47, Theorem


2.2].) Throughout this proof, we make tacit use of the fact, proved in Theorem
7.1.3, that, for each i ∈ N0 , the module Hm
i
(M ) is Artinian, so that, by 7.2.9,
i
it has a secondary representation and we can form the set Att(Hm (M )).
We use induction on n. When n = 0, the module M has finite length and so
is annihilated by some power of m. Hence Hm 0
(M ) ∼
= Γm (M ) = M = 0. By
7.2.9 and 7.2.12,
0
Att(Hm (M )) = Att M = {m} = Ass M = {p ∈ Ass M : dim R/p = 0} .

The result has been proved in this case.


Assume, inductively, that n > 0 and that the result has been proved for non-
zero, finitely generated R-modules of dimension n − 1. By 7.3.1, we can, after
factoring out, if necessary, the largest submodule of M of dimension smaller
than n, assume that M has no non-zero submodule of dimension smaller than
n. We shall make this assumption for the remainder of the inductive step, and
n
with this assumption our aim is to show that Att(Hm (M )) = Ass M .
Since n > 0, we have m ∈ Ass M , and so there exists r ∈ m which is a non-
n
zerodivisor on M . We suppose that Hm (M ) = 0, and look for a contradiction.
If n = 1, we have 1 ≤ gradeM m = depth M ≤ dim M = 1, so that
gradeM m = 1 and we have a contradiction to 6.2.7. Thus, in our search for a
contradiction, we can, and do, assume that n > 1.
Now, for each r ∈ m which is a non-zerodivisor on M , the module M/rM
(is non-zero and finitely generated and) has dim(M/rM ) = n − 1, and the
r
exact sequence 0 −→ M −→ M −→ M/rM −→ 0 induces a long exact
sequence of local cohomology modules
r
n−1
Hm (M ) −→ Hm
n−1
(M ) −→ Hm
n−1
(M/rM ) −→ 0
7.3 The Non-vanishing Theorem again 145

in view of our supposition that Hm n


(M ) = 0. Thus, for each r ∈ m which is
a non-zerodivisor on M , we have Hm n−1
(M )/rHmn−1
(M ) ∼
= Hmn−1
(M/rM ),
n−1
and this is non-zero by the inductive hypothesis. Therefore Hm (M ) = 0.
Our next step is to prove that m ∈ Att(Hm n−1
(M )). We suppose that m ∈
n−1
Att(Hm (M )) and look for a contradiction. Then, by the Prime Avoidance
Theorem,
  
m⊆ p∈Att(Hmn−1
(M )) p q∈Ass M q ,

so that, in the light of 7.2.11(i), there exists r1 ∈ m which is a non-zerodivisor


n−1 n−1
on M and such that Hm (M ) = r1 Hm (M ); this contradicts the fact that
n−1
Hm (M/r1 M ) = 0.
Thus m ∈ Att(Hm n−1
(M )): let p1 , . . . , pt be the remaining members of
n−1
Att(Hm (M )). Again by the Prime Avoidance Theorem, there exists
t  
r2 ∈ m \ i=1 pi q∈Ass M q .

Since r2 ∈ m and r2 is a non-zerodivisor on M , we again have


n−1
Hm n−1
(M )/r2 Hm (M ) ∼ n−1
= Hm (M/r2 M ),
n−1
and, by the inductive hypothesis, Hm (M/r2 M ) = 0 and
n−1
Att(Hm (M/r2 M )) ⊆ {p ∈ Ass(M/r2 M ) : dim R/p = n − 1} .

But, by 7.2.5,
n−1
Att(Hm n−1
(M )/r2 Hm (M )) ⊆ p ∈ Att(Hm
n−1
(M )) : r2 ∈ p ,

and m is the only member of the latter set. Since n > 1, we have obtained a
contradiction.
n
Thus we have proved that Hm (M ) = 0. To complete the inductive step,
since M now has no non-zero submodule of dimension smaller than n, it re-
n
mains for us to prove that Att(Hm (M )) = Ass M .
We know that gradeM m ≥ 1; also, for each r ∈ m which is a non-
zerodivisor on M , we have dim(M/rM ) = n − 1, so that Hm n
(M/rM ) = 0
by Grothendieck’s Vanishing Theorem 6.1.2, and the exact sequence of local
cohomology modules induced by the exact sequence
r
0 −→ M −→ M −→ M/rM −→ 0
n n
yields that Hm (M ) = rHm (M ). It therefore follows from 7.2.11(i) that
 
m \ p∈Ass M p ⊆ m \ q∈Att(Hm n (M )) q.

Let q ∈ Att(Hm
n
(M )): it follows from the above inclusion relation and the
146 Artinian local cohomology modules

Prime Avoidance Theorem that q ⊆ p for some p ∈ Ass M . Since Hm


n
is an
R-linear functor, it follows that
(0 : M ) ⊆ (0 : Hm
n
(M )) ⊆ q ⊆ p.
As n = dim R/(0 : M ) = dim R/p, it follows that q = p. Hence
n
Att(Hm (M )) ⊆ Ass M.
To establish the reverse inclusion, let p ∈ Ass M , so that dim R/p = n. By
the theory of primary decomposition, there exists a p-primary submodule Q of
M ; thus M/Q is a non-zero finitely generated R-module with Ass(M/Q) =
{p}. Note that M/Q cannot have any non-zero submodule of dimension less
than n (or else it would have an associated prime ideal other than p). Thus, by
the work in the preceding six paragraphs applied to M/Q rather than M , we
n
have Hm (M/Q) = 0 and
∅ = Att(Hm
n
(M/Q)) ⊆ Ass(M/Q) = {p} .
Hence Att(Hm n
(M/Q)) = {p}. Since dim Q < n + 1, Grothendieck’s Van-
n+1
ishing Theorem 6.1.2 tells us that Hm (Q) = 0; therefore, the canonical
exact sequence 0 −→ Q −→ M −→ M/Q −→ 0 induces an epimorphism
Hmn
(M ) −→ Hm n
(M/Q). It now follows from 7.2.5 that
{p} = Att(Hm
n
(M/Q)) ⊆ Att(Hm
n
(M )).
Hence Ass M ⊆ Att(Hm n n
(M )), and therefore Ass M = Att(Hm (M )). This
completes the inductive step.
We show next how to deduce the result of Exercise 6.1.7 very quickly from
Theorem 7.3.2.
7.3.3 Corollary. Assume (R, m) is local, and let M be a non-zero, finitely
n
generated R-module of dimension n > 0. Then Hm (M ) is not finitely gener-
ated.
n
Proof. By 7.1.3 and 7.3.2, the local cohomology module Hm (M ) is a non-
zero Artinian module which has a non-maximal attached prime ideal. Hence
n
Hm (M ) is not finitely generated, by 7.2.12.
7.3.4 Exercise. Let M be a finitely generated R-module for which the ideal
a + (0 : M ) is proper. Let p be a minimal prime ideal of a + (0 : M ), and let
t := dimRp Mp . Prove that Hat (M ) = 0, and that, if t > 0, then Hat (M ) is not
finitely generated.
8

The Lichtenbaum–Hartshorne Theorem

In this chapter, we take up again the main theme of Chapter 6, and establish
another vanishing theorem for local cohomology modules, namely the local
Lichtenbaum–Hartshorne Vanishing Theorem (see R. Hartshorne [29, Theo-
rem 3.1], and C. Peskine and L. Szpiro [66, chapitre III, théorème 3.1]). While
two important vanishing theorems in Chapter 6, namely Grothendieck’s Van-
ishing Theorem 6.1.2, and Theorem 6.2.7, which shows that, for a finitely
generated R-module M such that aM = M , we have Hai (M ) = 0 for all
i < gradeM a, can be regarded as ‘algebraic’ in nature, the Lichtenbaum–
Hartshorne Theorem is of ‘analytic’ nature, in the sense that it is intimately
related with ‘formal’ methods and techniques, that is, with passage to comple-
tions of local rings and with the structure theory for complete local rings.
Results of a ‘formal’ nature in algebraic geometry sometimes provide pow-
erful tools: the Lichtenbaum–Hartshorne Theorem is one such example, for we
shall see in Chapter 19 that it can play a crucial rôle in the study of connectivity
in algebraic varieties.
The local Lichtenbaum–Hartshorne Vanishing Theorem gives necessary and
sufficient conditions, over an n-dimensional local ring (R, m), for the vanish-
ing of n-th local cohomology modules with respect to a, where a is proper. The
sufficiency of these conditions is rather harder to prove than the necessity, and
so we shall just discuss the sufficiency in this introduction. Let R  denote the
n
completion of R. It turns out that Ha (M ) = 0 for every R-module M if, for
each P ∈ Spec(R)  satisfying dim R/P
 = n, we have dim R/(a  R  + P) > 0.
In the case when R is an n-dimensional complete local domain, the state-
ment simplifies: then, Han (M ) = 0 for every R-module M if dim R/a > 0.
The general statement can be deduced from this special case by means of stan-
dard reductions, and we now outline our strategy for proof of the special case.
This strategy is similar to that used by M. Brodmann and C. Huneke in [4].
Of course, the fact (3.4.10) that the local cohomology functor Hai commutes
148 The Lichtenbaum–Hartshorne Theorem

with direct limits means that it is enough to prove that Han (M  ) = 0 for each
finitely generated R-module M  in order to obtain the above result; in fact,
Lemma 8.1.7 below will show that it is enough to prove merely that Han (R) =
0. Then we shall use the Noetherian property of R to reduce to the case where a
is a prime ideal p with dim R/p = 1. As in our proof in Chapter 6 of Theorem
6.1.4, we again use ideas related to the structure theorems for complete local
rings to reduce to the case where R is a Gorenstein ring.
The arguments outlined in the last paragraph will reduce our proof to the
following: given a complete Gorenstein local domain R of dimension n and
p ∈ Spec(R) with dim R/p = 1, how can we show that Hpn (R) = 0? Our
approach to this will use the ideas concerning the symbolic powers of p of
Exercise 3.1.4 (for which we shall provide a solution!) to see that

Hpn (R) ∼
= lim ExtnR (R/p(j) , R);
−→
j∈N

we shall then use the facts that R is Gorenstein and depth R/p(j) > 0 to see
that ExtnR (R/p(j) , R) = 0 for all j ∈ N.

8.1 Preparatory lemmas


Our intention is to use Noetherian induction to reduce part of the proof of
the Lichtenbaum–Hartshorne Theorem to a case where (R is a complete local
domain and) a is a prime ideal p of R such that dim R/p = 1. To achieve this,
we wish to make our first application of a very useful proposition (8.1.2 below)
that relates local cohomology with respect to a + Rb, where b is an element of
R, to local cohomology with respect to a. Our proof of 8.1.2 uses the following
lemma about injective modules.

8.1.1 Lemma. Let b ∈ R, and let I be an injective R-module. Then the


sequence of R-modules and R-homomorphisms
ξI τ
0 −→ ΓRb (I) −→ I −→
I
Ib −→ 0,

in which ξI is the inclusion map and τI is the natural homomorphism, is


split exact. Consequently, when Ib is regarded as an R-module in the natu-
ral way, it is injective; furthermore, application of the additive functor Γa to
the above split exact sequence yields a further split exact sequence of injective
R-modules

0 −→ Γa+Rb (I) −→ Γa (I) −→ Γa (Ib ) −→ 0.


8.1 Preparatory lemmas 149
1
Proof. Since I is injective, HRb (I) = 0; therefore, the exactness of the first
sequence in the statement of the lemma is immediate from 2.2.20. By 2.1.4, the
R-module ΓRb (I) is injective, and so this exact sequence splits, and Ib , as R-
module, is injective. The second part is then immediate once one recalls from
1.1.2 that Γa (ΓRb (I)) = Γa+Rb (I), and from 2.1.4 that Γa (I  ) is injective
whenever I  is an injective R-module.
8.1.2 Proposition. Let b ∈ R and let f : M −→ N be a homomorphism of
R-modules.
(i) There is a long exact sequence of R-modules and R-homomorphisms
- H0 - H 0 (M ) - H 0 (Mb )
0 a+Rb (M ) a a

- H1 - H 1 (M ) - H 1 (Mb )
a+Rb (M ) a a

- ··· ···
- Hi - H i (M ) - H i (Mb )
a+Rb (M ) a a

- H i+1 (M ) - ···
a+Rb

such that the diagram


i
Ha+Rb (M ) - H i (M ) - H i (Mb ) - H i+1 (M )
a a a+Rb

i i+1
Ha+Rb (f ) Hai (f ) Hai (fb ) Ha+Rb (f )

? ? ? ?
i
Ha+Rb (N ) - H i (N ) - H i (Nb ) - H i+1 (N )
a a a+Rb

commutes for all i ∈ N0 .


(ii) For each i ∈ N0 , there is a commutative diagram

0 - H 1 (H i (M )) - H i+1 (M ) - ΓRb (H i+1 (M )) -0


Rb a a+Rb a

1 i+1
HRb (Hai (f )) Ha+Rb (f ) ΓRb (Hai+1 (f ))

? ? ?
0 - H 1 (H i (N )) - H i+1 (N ) - ΓRb (H i+1 (N )) -0
Rb a a+Rb a

with exact rows. The top row is referred to as the comparison exact sequence
for M .
Proof. Let
d−1 d0 di
I • : 0 −→ I 0 −→ I 1 −→ · · · −→ I i −→ I i+1 −→ · · ·
150 The Lichtenbaum–Hartshorne Theorem

be an injective resolution of M , so that there is an R-homomorphism α : M →


I 0 such that the sequence
α d0 di
0 −→ M −→ I 0 −→ I 1 −→ · · · −→ I i −→ I i+1 −→ · · ·
is exact. Similarly, let the exact sequence
β e0 ei
0 −→ N −→ J 0 −→ J 1 −→ · · · −→ J i −→ J i+1 −→ · · ·
provide an injective resolution for N . By the (dual of) the Comparison Theo-
rem [71, 6.9], there exists a chain map φ• = (φi )i∈N0 : I • −→ J • for which
the diagram
α
M - I0

f φ0
? ?
N
β
- J0

commutes.
It is immediate from Lemma 8.1.1 that there is a commutative diagram

0 - Γa+Rb (I • ) - Γa (I • ) - Γa ((I • )b ) -0

Γa+Rb (φ• ) Γa (φ• ) Γa ((φ• )b )

? ? ?
0 - Γa+Rb (J • ) - Γa (J • ) - Γa ((J • )b ) -0

of complexes of R-modules and chain maps of such complexes such that, for
each i ∈ N0 , the sequence 0 → Γa+Rb (I i ) → Γa (I i ) → Γa ((I i )b ) → 0 is
exact, and a similar property holds for the lower row. Thus the above commu-
tative diagram of complexes gives rise to a long exact sequence of cohomology
modules of the complexes in the top row, a similar long exact sequence for the
bottom row, and a chain map of the first long exact sequence into the second.
Since, by 8.1.1, (I • )b provides an injective resolution for Mb as R-module and
(J • )b provides an injective resolution for Nb as R-module, all the claims in (i)
follow.
(ii) For each R-module K, let τK : K −→ Kb be the natural map.
For an R-submodule L of I i , we use the Rb -monomorphism Lb −→ (I i )b
induced by inclusion to identify Lb as an Rb -submodule of (I i )b . With this
convention, it is routine to check that
Ker(Γa ((di )b )) = (Ker(Γa (di )))b , Im(Γa ((di−1 )b )) = (Im(Γa (di−1 )))b .
8.1 Preparatory lemmas 151

Hence
 
(Hai (M ))b = Ker(Γa (di ))/ Im(Γa (di−1 )) b

= (Ker(Γa (di )))b /(Im(Γa (di−1 )))b
= Ker(Γa ((di )b ))/ Im(Γa ((di−1 )b )) = Hai (Mb ).

This natural isomorphism (Hai (M ))b ∼ = Hai (Mb ), and the corresponding one
for N , enable us to deduce from (i) that there is a commutative diagram
τH i (M )
i
Ha+Rb (M ) - H i (M ) a
- (H i (M ))b - H i+1 (M )
a a a+Rb

i i+1
Ha+Rb (f ) Hai (f ) (Hai (f ))b Ha+Rb (f )

? ? τH i (N ) ? ?
i
Ha+Rb (N ) - H i (N ) a
- (H i (N ))b - H i+1 (N )
a a a+Rb

with exact rows. The result now follows from 2.2.20, which shows that

Ker τHai+1 (M ) = ΓRb (Hai+1 (M )),

that Ker τHai+1 (N ) = ΓRb (Hai+1 (N )) and that there is a commutative diagram
τH i (M )
Hai (M )
a
- (H i (M ))b - H 1 (H i (M )) - 0
a Rb a

1
Hai (f ) (Hai (f ))b HRb (Hai (f ))

? τH i (N ) ? ?
Hai (N )
a
- (H i (N ))b - H 1 (H i (N )) - 0
a Rb a

with exact rows.

8.1.3 Remark. In 3.3.4 we introduced cohd(a), the cohomological dimen-


sion of a. We deduce from 8.1.2(i) that, if b ∈ R, then

cohd(a + Rb) ≤ cohd(a) + 1.

8.1.4 Exercise. Let b be a second ideal of R and let i ∈ N0 . Show that the
i
restrictions of the functors Ha+b and Hai to the full subcategory of C(R) whose
objects are the b-torsion R-modules are naturally equivalent.

8.1.5 Exercise. Let f : M −→ N be a homomorphism of R-modules and


152 The Lichtenbaum–Hartshorne Theorem

let b ∈ R. Use 8.1.2(ii) and 8.1.4 to show that there is a commutative diagram
- H1 i - H i+1 (M ) - Γa+Rb (H i+1 (M )) - 0
0 a+Rb (Ha (M )) a+Rb a

1 i+1
Ha+Rb (Hai (f )) Ha+Rb (f ) Γa+Rb (Hai+1 (f ))

? ? ?
0 - H1 (H i
a (N ))
- H i+1 (N ) - Γa+Rb (H i+1 (N )) - 0
a+Rb a+Rb a

with exact rows.


We make use of Proposition 8.1.2 in the following lemma.
8.1.6 Lemma. Let (R, m) be a local integral domain of dimension n; suppose
that the set
B := {b : b is an ideal of R, Hbn (R) = 0 and dim R/b > 0}
is non-empty. Let p be a maximal member of B. (The fact that R is Noetherian
ensures that such a p exists.)
Then p is prime and dim R/p = 1.
Proof. Suppose that the statement ‘p is prime and dim R/p = 1’ is false.
This means that there exists q ∈ Spec(R) such that dim R/q = 1 and p ⊂ q.
Let b ∈ q \ p. Then, by 8.1.2(i), there is an exact sequence
n
Hp+Rb (R) −→ Hpn (R) −→ Hpn (Rb ).
By the Independence Theorem 4.2.1, there is an isomorphism of R-modules
Hpn (Rb ) ∼
= HpRn
b
(Rb ). Since dim Rb < n (as b ∈ m), we can deduce from
n
Grothendieck’s Vanishing Theorem 6.1.2 that HpR b
(Rb ) = 0. It therefore fol-
lows from the above exact sequence that Hp+Rb (R) = 0, and since p ⊂ p+Rb
n

and dim R/(p + Rb) > 0, this contradicts the maximality of p.


Our next lemma will come as no surprise to any reader who has solved Ex-
ercises 6.1.9 and 6.1.10. However, we provide a more direct argument, which
avoids use of 6.1.9.
8.1.7 Lemma. Assume that dim R = n and Han (R) = 0. Then Han (M ) = 0
for every R-module M .
Proof. Since Hai commutes with direct limits (by 3.4.10) and each R-module
is the direct limit of its finitely generated submodules, it is enough for us to
prove that Han (M ) = 0 whenever M is finitely generated. However, in that
case, there exists an exact sequence 0 −→ N −→ F −→ M −→ 0 of
finitely generated R-modules and R-homomorphisms in which F is free. This
8.1 Preparatory lemmas 153

induces an exact sequence Han (F ) −→ Han (M ) −→ Han+1 (N ). It follows


from the additivity of the local cohomology functor Han and the assumption
that Han (R) = 0 that Han (F ) = 0; since Han+1 (N ) = 0 by Grothendieck’s
Vanishing Theorem 6.1.2, we can deduce that Han (M ) = 0, as required.

8.1.8 Exercise. Assume that dim R = n and that a is proper, having minimal
prime ideals p1 , . . . , pt . Let M be an R-module for which Hpni (M ) = 0 for all
i = 1, . . . , t. Show that Han (M ) = 0.

In the course of our proof of the local Lichtenbaum–Hartshorne Vanishing


Theorem, we shall reduce to the case where R is a complete local domain
and then appeal to Cohen’s Structure Theorem for complete local rings: we
shall again quote [50, Theorem 29.4] to deduce that there exists a complete
regular local subring R of R which is such that R is finitely generated as an
R -module. The next lemma will be applied to this situation.

8.1.9 Lemma. Suppose that (R, m) is local, and that (R , m ) is a local sub-
ring of R which is such that R is finitely generated as an R -module. Let
p ∈ Spec(R) be such that dim R/p =  1. There exists b ∈ R such that, if
B denotes the subring R [b] of R, then (p ∩ B)R = p.

Proof. Of course, R is integral over R . By the Incomparability Theorem (see


[81, 13.33], for example), p must be a minimal prime ideal of (p ∩ R )R: let
p2 , . . . , pt be the other minimal prime ideals of (p ∩ R )R (there may be none).
t
By the Prime Avoidance Theorem, there exists b ∈ p \ i=2 pi . Then B :=
R [b], necessarily a Noetherian ring, is finitely generated as an R -module, and
so is integral over R .
To complete the proof, it is enough for us to show that p is the one and only
minimal prime ideal of (p ∩ B)R. It certainly is one, since R is integral over B
so that we can appeal to the Incomparability Theorem again, as we did at the
beginning of this proof. Let us suppose that q is another minimal prime ideal
of (p ∩ B)R, different from p.
Obviously q = m. Since R is integral over R , it follows that q ∩ R = m
(by [81, 13.31], for example). Since p ∩ R ⊆ p ∩ B ⊆ q, it follows that
p ∩ R ⊆ q ∩ R ⊂ m . But R/p is an integral extension of R /(p ∩ R ), and so
dim R /(p∩R ) = dim R/p = 1. We can therefore deduce that p∩R = q∩R .
Another use of the Incomparability Theorem as before shows that q must be
a minimal prime ideal of (p ∩ R )R, and so must be one of p2 , . . . , pt . But
b ∈ p ∩ R [b] ⊆ (p ∩ B)R ⊆ q, and this is a contradiction because b was
chosen outside p2 ∪ · · · ∪ pt .

In our application of Lemma 8.1.9 in which R is a complete regular local


154 The Lichtenbaum–Hartshorne Theorem

ring, it will be important for us to know that the ring B = R [b] is a Gorenstein
ring. This is the motivation behind the next lemma. In fact, the first part of this
lemma is actually a special case of [81, Proposition 13.40], but we include a
proof for the convenience of the reader.
8.1.10 Lemma. Let R be a subring of the integral domain S, and suppose
that R is integrally closed and S = R[b], where b ∈ S is integral over R. Let
K be the field of fractions of R. Then b is algebraic over K and its minimal
polynomial f over K belongs to R[X]. Also, the surjective ring homomor-
phism φ : R[X] → R[b] given by evaluation at b has kernel f R[X], so that φ
induces an isomorphism of R-algebras

=
φ : R[X]/f R[X] −→ R[b] = S.
Proof. Since b is integral over R, it is certainly algebraic over K. Let its
minimal polynomial over K be
f = X h + ah−1 X h−1 + · · · + a1 X + a0 ∈ K[X].
We aim to show that a0 , . . . , ah−1 ∈ R.
There exists a field extension L of the field of fractions of S such that
f splits into linear factors in L[X]: let s = s1 , . . . , sh ∈ L be such that
f = (X − s1 )(X − s2 ) . . . (X − sh ) in L[X]. Equate coefficients to see
that each of a0 , . . . , ah−1 can be written as a ‘homogeneous polynomial’ (in
fact, a ‘symmetric function’) in s1 , . . . , sh with coefficients ±1; in particular,
a0 , . . . , ah−1 ∈ R[s1 , . . . , sh ].
Since b is integral over R, there exists
g = X n + rn−1 X n−1 + · · · + r1 X + r0 ∈ R[X]
such that g(b) = 0. Therefore g ∈ Ker φ ⊆ f K[X], and so there exists
g1 ∈ K[X] such that g = f g1 in K[X]. For each i = 1, . . . , h, evaluate
both sides of this equation at si ∈ L to see that g(si ) = 0. Thus all the
si (i = 1, . . . , h) are integral over R, and so, by [81, 13.21] for example,
the ring R[s1 , . . . , sh ] is a finitely generated R-module; it therefore follows
that a0 , . . . , ah−1 are all integral over R, since they belong to R[s1 , . . . , sh ].
But a0 , . . . , ah−1 ∈ K and R is integrally closed; hence a0 , . . . , ah−1 ∈ R.
Finally, Ker φ = f K[X] ∩ R[X], and it is elementary to deduce from the
facts that f is monic and has all its coefficients in R that f K[X] ∩ R[X] =
f R[X].
8.1.11 Lemma. Suppose that (R, m) is local and of dimension n, and also
that R is a Gorenstein ring. Let M be a non-zero, finitely generated R-module.
Then depth M > 0 if and only if ExtnR (M, R) = 0.
8.1 Preparatory lemmas 155

Proof. Recall from [50, Theorem 18.1] (or from [3]) that R has injective
dimension n as a module over itself.
(⇒) There exists r ∈ m such that r is a non-zerodivisor on M . The exact
r
sequence 0 −→ M −→ M −→ M/rM −→ 0 induces an exact sequence
r
ExtnR (M/rM, R) −→ ExtnR (M, R) −→ ExtnR (M, R) −→ 0
since Extn+1
R (M/rM, R) = 0. Application of Nakayama’s Lemma to the
finitely generated R-module ExtnR (M, R) now shows that this module is zero.
(⇐) By [50, Theorem 18.1], we have ExtnR (R/m, R) = 0. It follows that
m ∈ Ass M , for otherwise there would be an exact sequence
0 −→ R/m −→ M −→ C −→ 0
of R-modules, and the induced exact sequence
ExtnR (M, R) −→ ExtnR (R/m, R) −→ Extn+1
R (C, R) = 0

would provide a contradiction.


The next exercise extends Lemma 8.1.11 in a precise manner.
8.1.12 Exercise. Suppose that (R, m) is a Gorenstein local ring of dimension
n. Let M be a non-zero, finitely generated R-module. Show that depth M is
n−i
the least integer i such that ExtR (M, R) = 0.
The next lemma provides the promised solution for Exercise 3.1.4.
8.1.13 Lemma. Let (R, m) be a complete local domain, and let p be a prime
ideal of R of dimension 1, that is, such that dim R/p = 1.
(i) For each n ∈ N, there exists t ∈ N such that p(t) ⊆ mn .
(ii) The family (p(n) )n∈N is a system of ideals of R over N in the sense
of 2.1.10, and the (negative strongly) connected sequences of covariant
functors (from C(R) to C(R))
 
Hpi and lim ExtiR (R/p(j) , • )
i∈N0 −→
j∈N i∈N0

are isomorphic.
Proof. Let ψ : R → Rp denote the natural homomorphism.
(i) Recall that p(n) := ψ −1 (pn Rp ) = ψ −1 ((pRp )n ). Hence
 
n∈N p
(n)
= ψ −1 n∈N (pRp )
n
= ψ −1 (0) = Ker ψ,
by Krull’s Intersection Theorem [50, Theroem 8.9]. Since R is a domain, it
follows that n∈N p(n) = 0.
156 The Lichtenbaum–Hartshorne Theorem

We can now use Chevalley’s Theorem ([59, Section 5.2, Theorem 1] or [89,
Chapter VIII, Section 5, Theorem 13]) to deduce that, for each n ∈ N, there
exists t ∈ N such that p(t) ⊆ mn .
(ii) Let n ∈ N. Note that, by [81, 5.47(i)] for example, p is the unique
minimal prime ideal of pn , and p(n) is the (uniquely determined) p-primary
term in any minimal primary decomposition of pn . Since dim R/p = 1, the
only possible associated prime ideal of pn in addition to p is m. Thus either
pn = p(n) , or there exists h ∈ N such that p(n) ∩ mh ⊆ pn . It now follows
from part (i) that there exists t ∈ N such that p(t) ⊆ pn .
Of course, pn ⊆ p(n) for all n ∈ N. Since it is clear that (p(n) )n∈N is an
inverse family of ideals of R over N in the sense of 1.2.10, it now follows from
Proposition 3.1.1(iii) that (p(n) )n∈N is actually a system of ideals of R over N,
and from part (ii) of the same proposition that the negative strongly connected
sequences of covariant functors
   
lim ExtiR (R/pj , • ) and lim ExtiR (R/p(j) , • )
−→ −→
j∈N i∈N0 j∈N i∈N0

are isomorphic. The proof can now be completed by an appeal to 1.3.8.

8.2 The main theorem


We are now in a position to put all our various lemmas of §8.1 together to
produce a proof of the main theorem of this chapter.

8.2.1 The Lichtenbaum–Hartshorne Vanishing Theorem. (See R. Hartsh-


orne [29, Theorem 3.1].) Suppose that (R, m) is local of dimension n, and also
that a is proper. Then the following statements are equivalent:

(i) Han (R) = 0;


 the completion of R,
(ii) for each (necessarily minimal) prime ideal P of R,
  
satisfying dim R/P = n, we have dim R/(aR + P) > 0.

Proof. (i) ⇒ (ii) Assume that Han (R) = 0 and that there exists a prime ideal
P of R such that dim R/P
  R
= n but dim R/(a  + P) = 0. Since the natural
ring homomorphism R → R  is flat, it follows from the Flat Base Change

Theorem 4.3.2 that there is an R-isomorphism H n (R) ⊗R R∼  and
= H n (R),
a 
aR
so 
HanR (R)
= 0.
Now mR  is the maximal ideal of the local ring R,
 and our assumptions
 
mean that (R/P, mR/P) is an n-dimensional local ring and (aR + P)/P is
8.2 The main theorem 157

an (mR/P)-primary ideal of this ring. It therefore follows from 1.1.3 and The-
orem 6.1.4 (or Theorem 7.3.2) that H(a n 
(R/P) = 0. We now deduce

R+P)/P
n 
from the Independence Theorem 4.2.1 that H (R/P) = 0. Therefore, by

aR
 = 0, and this is a contradiction.
8.1.7, we must have HanR (R)
 R
(ii) ⇒ (i) Assume that dim R/(a  + P) > 0 for each P ∈ Spec(R)  for
which dim R/P n
= n. We suppose also that Ha (R) = 0, and we again look
for a contradiction. We again use the Flat Base Change Theorem 4.3.2 to see

that there is an R-isomorphism Han (R) ⊗R R  ∼  from which we
= HanR (R),
deduce that H n (R)  = 0 (because the natural ring homomorphism R → R  is

aR
faithfully flat). We can therefore assume, in our search for a contradiction, that
R is complete: we make this assumption in what follows.
There is an ascending chain 0 = b0 ⊂ b1 ⊂ · · · ⊂ bh−1 ⊂ bh = R of
ideals of R such that, for each i = 1, . . . , h, there exists pi ∈ Spec(R) with
bi /bi−1 ∼ = R/pi . It follows from the half-exactness of Han that Han (R/pi ) = 0
for at least one i between 1 and h. The Independence Theorem 4.2.1 now en-
n
ables us to deduce that H(a+p i )/pi
(R/pi ) = 0. Note that, by Grothendieck’s
Vanishing Theorem 6.1.2, the complete local domain R/pi has dimension n.
Therefore, our assumptions imply that dim(R/pi )/((a + pi )/pi ) > 0. There-
fore we can, and do, assume henceforth in our search for a contradiction that
R is an n-dimensional complete local domain and the proper ideal a satisfies
Han (R) = 0 and dim R/a > 0. We shall show that such a situation is impossi-
ble.
By Lemma 8.1.6, there exists p ∈ Spec(R) such that Hpn (R) = 0 and
dim R/p = 1. We now appeal to Cohen’s Structure Theorem for complete
local rings: by [50, Theorem 29.4], there exists a complete regular local subring
(R , m ) of R which is such that R is finitely generated as an R -module. By
Lemma 8.1.9,
 there exists b ∈ R such that, if B denotes the subring R [b] of
R, then (p ∩ B)R = p. Note that, since R is integral over B, the latter is
also local and of dimension n, with unique maximal ideal m ∩ B. Note also
that, since B is a finitely generated R -module, it is a complete local ring.
Since R is integrally closed, it follows from Lemma 8.1.10 that there is a
monic polynomial f ∈ R [X] and an isomorphism of R -algebras

R [X]/f R [X] ∼
= R [b] = B.

By [50, Theorem 19.5], the ring R [X] is regular. As f is monic, it is a non-


zerodivisor in R [X]. We now deduce that B is a Gorenstein local ring as
follows. There exists exactly one maximal ideal Q of R [X] that contains f ;
158 The Lichtenbaum–Hartshorne Theorem

we have
B∼
= (R [X]/f R [X])Q/f R [X] ∼
= R [X]Q /(f /1)R [X]Q ;
and the latter
 is a Gorenstein local ring by [50, Exercise 18.1].n
Since (p ∩ B)R = p, we can deduce from 1.1.3 that H(p∩B)R (R) = 0,
n
and thence by the Independence Theorem 4.2.1 that H(p∩B) (R) = 0. There-
n
fore H(p∩B) (B) = 0 by Lemma 8.1.7. Let q := p ∩ B; observe that, since R
is integral over B, we have dim B/q = dim R/p = 1.
We have nearly arrived at a contradiction: B is a complete n-dimensional
Gorenstein local domain, q ∈ Spec(B) has dim B/q = 1, and Hqn (B) = 0.
This is, in fact, impossible, because, by Lemma 8.1.13(ii),
Hqn (B) ∼
= lim ExtnB (B/q(j) , B)
−→
j∈N

and the latter module is zero by Lemma 8.1.11, since depth B/q(j) > 0 for all
j ∈ N because each q(j) is a q-primary ideal.
8.2.2 Corollary. Suppose that (R, m) is local and has dimension n, and that
 R
dim R/(a  + P) > 0 for every prime ideal P of R
 such that dim R/P
 = n.
Then Ha (M ) = 0 for all i ≥ n and for every R-module M .
i

Proof. This follows from Grothendieck’s Vanishing Theorem 6.1.2, Lemma


8.1.7, and the Lichtenbaum–Hartshorne Vanishing Theorem 8.2.1.
8.2.3 Exercise. Suppose that (R, m) is local and that M is a non-zero finitely
generated R-module of dimension n. Assume that
 R
dim R/(a  + P) > 0
 such that dim R/P
for every prime ideal P of SuppR (M ⊗R R)  = n. Prove
that Ha (M ) = 0 for all i ≥ n.
i

8.2.4 Exercise. Suppose that (R, m) is local and that A is an Artinian R-


module. Show that, for each a ∈ A, there exists t ∈ N such that mt a = 0.
Deduce that A has a natural structure as a module over R, the completion of

R, that a subset of A is an R-submodule if and only if it is an R-submodule,
and that the map φ : A → A ⊗R R  defined by φ(a) = a ⊗ 1 for all a ∈ A is

an isomorphism of R-modules.
The next two exercises involve the theory of secondary representation and
attached primes of Artinian modules, discussed in §7.2.
8.2.5 Exercise. Let R denote a second commutative Noetherian ring and let
f : R → R be a ring homomorphism. Suppose that the R -module M  has
8.2 The main theorem 159

a secondary representation (7.2.2). Show that, when M  is viewed as an R-


module by means of f , it has a secondary representation as an R-module, and
that
AttR M  = f −1 (p ) : p ∈ AttR M  .

Is it possible for M  to have fewer attached prime ideals as an R-module than


it has as an R -module? Justify your response.

8.2.6 Exercise. Suppose that (R, m) is local and has dimension n, and also
that a is proper. By Theorem 7.1.6, the R-module Han (R) is Artinian.

(i) Use the Lichtenbaum–Hartshorne Vanishing Theorem 8.2.1 to prove that


 dim R/P
AttR (Han (R)) = {P ∩ R : P ∈ Spec(R),  = n,

 R
and dim R/(a  + P) = 0 .

(It is perhaps worth pointing out that the result of this exercise (which
is due to R. Y. Sharp [80, Corollary 3.5]) amounts, in effect, to a refine-
ment of the statement of the local Lichtenbaum–Hartshorne Vanishing
Theorem, because the set on the right-hand side of the above display is
empty if and only if Han (R) = 0.)
(ii) Deduce that, if n > 0, then Han (R), if non-zero, is not finitely generated.

8.2.7 Remark. Let n ∈ N and let K be a field. This comment is concerned


with the ring K[X1 , . . . , Xn ] of polynomials over K.
Each f ∈ K[X1 , . . . , Xn ] \ (X1 , . . . , Xn ) is a unit in K[[X1 , . . . , Xn ]], and
so there is a natural injective ring homomorphism

φ : K[X1 , . . . , Xn ](X1 ,...,Xn ) −→ K[[X1 , . . . , Xn ]].

It is not difficult to see that we can use φ to regard K[[X1 , . . . , Xn ]] as the


completion of the regular local ring K[X1 , . . . , Xn ](X1 ,...,Xn ) . Let m denote
the maximal ideal (X1 , . . . , Xn ) of K[X1 , . . . , Xn ].
It follows that, if b is an ideal of K[X1 , . . . , Xn ] contained in m, and we
use xi to denote the natural image of Xi in K[X1 , . . . , Xn ]/b (for each i =
1, . . . , n), then the injective ring homomorphism

K[x1 , . . . , xn ](x1 ,...,xn ) −→ K[[X1 , . . . , Xn ]]/bK[[X1 , . . . , Xn ]]

obtained by composing the natural ring isomorphism

K[x1 , . . . , xn ](x1 ,...,xn ) = (K[X1 , . . . , Xn ]/b)m/b



=
−→ K[X1 , . . . , Xn ]m /bK[X1 , . . . , Xn ]m
160 The Lichtenbaum–Hartshorne Theorem

with the injective ring homomorphism


K[X1 , . . . , Xn ]m /bK [X1 , . . . , Xn ]m
−→ K[[X1 , . . . , Xn ]]/bK[[X1 , . . . , Xn ]]
induced by φ provides the completion of K[x1 , . . . , xn ](x1 ,...,xn ) .
8.2.8 Exercise. Consider the special case of the situation of Remark 8.2.7 in
which n = 3 and b = (X1 X2 ). With these choices, let
R = K[x1 , x2 , x3 ](x1 ,x2 ,x3 ) = (K[X1 , X2 , X3 ]/(X1 X2 ))(X1 ,X2 ,X3 )/(X1 X2 ) .
Let p be the prime ideal (x1 /1, x3 /1) of R. Use the exact sequence
0 −→ R −→ R/(x1 /1) ⊕ R/(x2 /1) −→ R/(x1 /1, x2 /1) −→ 0
of Lemma 3.2.1 to show that Hp2 (R) is not finitely generated. Use the descrip-
 of R provided by 8.2.7 to find a prime ideal P of R
tion of the completion R 
  
such that dim R/P = 2 but dim R/(pR + P) = 0.
8.2.9 Exercise. Consider the special case of the situation of Remark 8.2.7
in which K has characteristic 0, n = 3 and b = (X22 − X12 − X13 ). Set
R := K[X1 , X2 , X3 ] and m := (X1 , X2 , X3 ). Let
c := (X1 + X2 − X2 X3 )R + ((X3 − 1)2 (X1 + 1) − 1)R .
(i) Show that b ∈ Spec(R ), that b ⊆ c, and also that cRm  
 ∈ Spec(Rm ).

(Here is a hint: do not forget that Rm is a regular local ring.)
(ii) Follow the procedure of 8.2.7, and set
R := (R /b)m /b and  = K[[X1 , X2 , X3 ]]/bK[[X1 , X2 , X3 ]];
R

regard R as the completion of R in the manner described in 8.2.7. It will be


 (for
convenient to use xi to denote the natural image of Xi in R /b and in R
i = 1, 2, 3). Show that R is a 2-dimensional local domain, and that
p := (x1 + x2 − x2 x3 )R + ((x3 − 1)2 (x1 + 1) − 1)R
is a prime ideal of R with dim R/p = 1.
 be given by
(iii) Let u ∈ R
1 1 1 5 4 (−1)n−1 2(2n − 2)! n
u := 1 + x1 − x21 + x31 − x1 + · · · + x1 + · · · ,
2 8 16 128 4n n!(n − 1)!
 and u2 = 1 + x1 . Show that
so that u is a unit of R
 = (x1 + x2 − x2 x3 )R
pR  + (x3 − 1 + u−1 )R,


and that this is a prime ideal of R.
8.2 The main theorem 161

(iv) Use the exact sequence


 −→ R/(x
0→R  2 + x1 u) −→ R/(x
 2 − x1 u) ⊕ R/(x  2 − x1 u, x2 + x1 u) → 0

 is not finitely generated, and deduce that


of Lemma 3.2.1 to show that Hp2R (R)
Hp2 (R) is not finitely generated.
(v) Find a prime ideal P of R such that dim R/P
 = 2 but
 R
dim R/(p  + P) = 0.

8.2.10 Definition and Exercise. Suppose that (R, m) is local. We say that R
 the completion of R, is an integral
is analytically irreducible precisely when R,
domain.
Note that an analytically irreducible local ring must itself be a domain.
Suppose that (R, m) is analytically irreducible, and that a is proper. Let
dim R = n. Show that Han (R) = 0 if and only if dim R/a > 0, and that,
when this is the case, Hai (M ) = 0 for all i ≥ n and for every R-module M .

8.2.11 Exercise. Suppose (R, m) is an analytically irreducible (see 8.2.10


above) local domain of dimension 2, and that dim R/a > 0. Show that the
a-transform functor Da is exact.

8.2.12 Exercise. Show by means of an example that, if the phrase ‘analyt-


ically irreducible’ is omitted from the hypotheses in Exercise 8.2.11 above,
then the resulting statement is no longer always true.

8.2.13 Exercise. For this exercise, we recommend that the reader refers to
the example studied in 2.3.7, 3.3.5 and 4.3.7, and applies the ideas of Remark
8.2.7 in the special case in which n = 3 and

b = (X1 X4 − X2 X3 , X12 X3 + X1 X2 − X22 , X33 + X3 X4 − X42 ).

Let S  denote the subring C[X, XY, Y (Y − 1), Y 2 (Y − 1)] of C[X, Y ].


Consider the ideals

n := (X, XY, Y (Y − 1), Y 2 (Y − 1)) and r := (XY, Y (Y − 1), Y 2 (Y − 1))

of S  .

(i) Show that rSq is a principal ideal of Sq for every maximal ideal q of S 
different from n.
(ii) Show that S := Sn is a local domain of dimension 2, and that p := rSn
is a prime ideal of S such that dim S/p = 1.
(iii) Use a calculation made at the end of 3.3.5 to show that Hp2 (S) = 0.
162 The Lichtenbaum–Hartshorne Theorem
+ of S provided by Remark
 M)
(iv) Use the description of the completion (S,
8.2.7, together with calculations made in 4.3.7, to see that S has exactly
two minimal prime ideals, P1 and P2 say, and that
+
P1 + P2 = M + for i = 1 or 2.
and Pi + pS = M
 is not finitely generated, and deduce that H 2 (S) is
Show that Hp2S(S) p
not finitely generated.

8.2.14 Exercise. Let R be the localization C[X, Y 2 , XY, Y 3 ](X,Y 2 ,XY,Y 3 )


of the subring C[X, Y 2 , XY, Y 3 ] of C[X, Y ]. Show that Ha2 (R) = 0 for every
ideal a of R which is not primary to the maximal ideal.

8.2.15 Definition and Exercise. Let V be a quasi-affine variety over an al-


gebraically closed field K, and let p ∈ V .
We say that V is analytically irreducible at p precisely when the completion

O V, p of the local ring OV, p of V at p is an integral domain, that is, if and
only if OV, p is analytically irreducible (see 8.2.10); otherwise, we say that V
is analytically reducible at p.

(i) Now let V be an affine surface over K (that is, an affine algebraic variety
over K of dimension 2), and let C ⊂ V be a (not necessarily irreducible)
curve (that is, a non-empty closed subset of pure codimension 1 in V (see
6.4.2(iv))) which avoids all the points of V at which V is analytically
reducible. Use Serre’s Affineness Criterion 6.4.4 to show that the quasi-
affine variety V \ C is, in fact, affine.
(ii) By considering the example studied in 2.3.7, 3.3.5, 4.3.7 and (especially)
6.4.6, show that the condition that C ‘avoids all the points of V at which
V is analytically reducible’ cannot be omitted from the hypotheses of
the result established in part (i) above.

8.2.16 Exercise. Let V be an affine surface over an algebraically closed field


K; let C ⊂ V be a (not necessarily irreducible) curve (see 8.2.15(i)). Let b be
the ideal of C in O(V ) and let p ∈ C. We say that C is fully branched in V
at p precisely when dim O 
V, p /(bOV, p + P) > 0 for all minimal primes P of
 
OV, p satisfying dim OV, p /P = 2.

(i) Show that C is fully branched in V at p if and only if


2
HbOV, p
(OV, p ) = 0.

(ii) Show that V \ C is affine if and only if C is fully branched in V at all its
points.
8.2 The main theorem 163

8.2.17 Exercise. This exercise is concerned with the situation of Exercise


8.2.9, and so we use the notation of that exercise. It was not necessary in 8.2.9
for the reader to know that the ideal c is actually prime, but, in fact, it is and
the first three parts of this exercise sketch a route to a proof of this fact.
(i) Note that K[X1 , X2 , X3 ] = K[X  1 , X2 , X
3 ], where X1 = X1 + 1 and
X3 = X3 − 1. Note also that X  1 , X2 , X
3 are algebraically independent
over K.
3 ]/(X
 1 , X2 , X  2X   and
(ii) Set A := K[X 3 1 − 1); consider A0 := K[X3 ]X 3

K[X2 , X3 ]  as subrings of K(X2 , X 3 ) in the natural ways. Show that


X3

A∼ 3 ]  = A0 [X2 ].
= K[X2 , X X3

(iii) Show that


 1 , X2 , X
K[X 3 ]/(X
1 −X2 X
3 − 1, X 32 X
1 − 1)
∼ 3 ]  /(X2 + X
= K[X2 , X  −3 ) ∼
 −1 − X = A0 ,
X3 3 3

and deduce that c is a prime ideal of K[X1 , X2 , X3 ].


(iv) Now assume that K is algebraically closed.
Let C := VA3 (K) (X1 + X2 − X2 X3 , (X3 − 1)2 (X1 + 1) − 1) and
V := VA3 (K) (X22 − X12 − X13 ). Show that C ⊆ V and (with or without
the aid of Exercise 8.2.16 above) that V \ C is not affine.
9

The Annihilator and Finiteness Theorems

There have been several examples earlier in this book of non-finitely gener-
ated local cohomology modules of finitely generated modules: for example,
in 6.1.7, and also in 7.3.3, we saw that, if (R, m) is local, and N is a non-
n
zero, finitely generated R-module of dimension n > 0, then Hm (N ) is not
0
finitely generated. Since Hm (N ), being isomorphic to a submodule of N , is
certainly finitely generated, it becomes of interest to identify the least integer
i
i for which Hm (N ) is not finitely generated. This integer is referred to as the
finiteness dimension fm (N ) of N relative to m. Our work in this chapter on
Grothendieck’s Finiteness Theorem will enable us to see that, in this situation,
and under mild restrictions on R,
 
fm (N ) = min depthRp Np + dim R/p : p ∈ Supp N \ {m} .

However, our approach will not restrict attention to the case where the ideal
with respect to which local cohomology is calculated is the maximal ideal of
a local ring; also, the approach we shall take will show that questions about
such finiteness dimensions are intimately related to questions about precisely
which ideals annihilate local cohomology modules. We shall, in fact, establish
Grothendieck’s Finiteness Theorem as a special case of Faltings’ Annihilator
Theorem, which is concerned with the following question: given a second ideal
b of R, and given a finitely generated R-module M , is there a greatest integer
i for which Hai (M ) is annihilated by some power of b, and, if so, what is it?

9.1 Finiteness dimensions


Our first two results will provide motivation for the formal introduction of the
concept of finiteness dimension. Let M be a finitely generated R-module. It is
9.1 Finiteness dimensions 165

clear that, if, for some i ∈ N, the local cohomology module Hai (M ) is finitely
generated, then au Hai (M ) = 0 for some u ∈ N, and so

a⊆ (0 : Hai (M )).

It is not quite so clear that there is a sort of converse to this result: we show
below, in our first proposition of this chapter, that, if, for some t ∈ N, it is

the case that a ⊆ (0 : Hai (M )) for all i < t, then it follows that Hai (M ) is
finitely generated for all i < t.

f g
9.1.1 Lemma. Let L −→ M −→ N be an exact sequence of R-modules and
√ √
R-homomorphisms, and suppose that a ⊆ (0 : L) and a ⊆ (0 : N ). Then

a ⊆ (0 : M ) also.

Proof. Let r ∈ a. Then, for some t ∈ N, we have rt g(M ) = 0, and so


rt M ⊆ Ker g = Im f . However, there exists u ∈ N such that ru L = 0, and so
ru (Im f ) = 0. Hence ru+t M = 0.

9.1.2 Proposition. Let M be a finitely generated R-module, and let t ∈ N.


Then the following statements are equivalent:

(i) Hai (M ) is finitely generated for all i < t;



(ii) a ⊆ (0 : Hai (M )) for all i < t.

Proof. (i) ⇒ (ii) As we remarked immediately before the statement of 9.1.1,


this implication is clear.
(ii) ⇒ (i) We use induction on t. When t = 1, there is nothing to prove,
since Ha0 (M ) = Γa (M ) is a submodule of M , and so is finitely generated. So
suppose that t > 1 and that the result has been proved for smaller values of t.
By this assumption, Hai (M ) is finitely generated for i = 0, 1, . . . , t − 2, and it
only remains for us to prove that Hat−1 (M ) is finitely generated.
It follows from 2.1.7(iii) that Hai (M ) ∼
= Hai (M/Γa (M )) for all i > 0.
Also, M/Γa (M ) is an a-torsion-free R-module, by 2.1.2. Hence we can, and
do, assume that M is an a-torsion-free R-module.
We now use 2.1.1(ii) to deduce that a contains an element r which is a non-

zerodivisor on M . Since a ⊆ (0 : Hat−1 (M )), there exists u ∈ N such that
ru Hat−1 (M ) = 0. The exact sequence

ru
0 −→ M −→ M −→ M/ru M −→ 0
166 The Annihilator and Finiteness Theorems

induces a long exact sequence


ru
0 - H 0 (M ) - H 0 (M ) - H 0 (M/ru M )
a a a
ru
- H 1 (M ) - H 1 (M ) - H 1 (M/ru M )
a a a

- ··· ···
ru
- H i (M ) - H i (M ) - H i (M/ru M )
a a a

- H i+1 (M ) - ··· .
a

It follows from this long exact sequence and 9.1.1 that



a ⊆ (0 : Hai (M/ru M )) for all i < t − 1,

so that, by the inductive hypothesis, Hai (M/ru M ) is finitely generated for all
i < t − 1. In particular, we see that Hat−2 (M/ru M ) is finitely generated;
since ru Hat−1 (M ) = 0, it follows from the above long exact sequence that
Hat−1 (M ) is a homomorphic image of Hat−2 (M/ru M ), and so is finitely gen-
erated. This completes the inductive step.

Proposition 9.1.2 provides some motivation for the following definition.


Here, and throughout the book, we adopt the convention that the infimum of
the empty set of integers is to be taken as ∞.

9.1.3 Definition. Let M be a finitely generated R-module. In the light of


Proposition 9.1.2, we define the finiteness dimension fa (M ) of M relative to
a by

fa (M ) = inf i ∈ N : Hai (M ) is not finitely generated



= inf i ∈ N : a ⊆ (0 : Hai (M )) .

Note that fa (M ) is either a positive integer or ∞, and that, since Ha0 (M ) is


finitely generated,

fa (M ) = inf i ∈ N0 : Hai (M ) is not finitely generated .

9.1.4 Exercise. For a finitely generated R-module M , show that fa (M ) > 1


if and only if the a-transform Da (M ) is finitely generated.

In the situation of 9.1.3, it is reasonable to regard the condition that a ⊆



(0 : Hai (M )) as asserting that Hai (M ) is ‘small’ in a sense, because if
this condition holds for all i less than some positive integer t, then Hai (M )
is finitely generated for all i < t (by 9.1.2). However, sometimes it is more
9.1 Finiteness dimensions 167

realistic to hope for a weaker condition than ‘a ⊆ (0 : Hai (M ))’: we intro-
duce a second ideal b of R, and, when b ⊆ a, think of Hai (M ) as being ‘small’

relative to b if b ⊆ (0 : Hai (M )).
9.1.5 Definition. Let M be a finitely generated R-module and let b be a
second ideal of R. We define the b-finiteness dimension fab (M ) of M relative
to a by

fab (M ) := inf i ∈ N0 : b ⊆ (0 : Hai (M )) .
Note that fab (M ) is either ∞ or a non-negative integer not exceeding dim M
√ √
and that faa (M ) = fa (M ) because a ⊆ (0 : Γa (M )). Note that, if b ⊆ a,

then b ⊆ (0 : Γa (M )), so that we can then write

fab (M ) := inf i ∈ N : b ⊆ (0 : Hai (M )) .

However, in general we shall not assume that b ⊆ a.
9.1.6 Exercise. Let the situation be as in 9.1.5, and let S be a multiplica-
tively closed subset of R. Show that
−1
fab (M ) ≤ fSS−1 ab (S −1 M ).
9.1.7 Exercise. Let R be a second commutative Noetherian ring and let
f : R → R be a ring homomorphism; assume that R , when regarded as an
R-module by means of f , is finitely generated.
Let M  be a finitely generated R -module and let b be a second ideal of R.
Prove that

fab (M  ) = faR
bR 
 (M ).

9.1.8 Lemma. Let b, c be further ideals of R, and let M be a finitely gener-


ated R-module.
(i) If b ⊆ c, then fab (M ) = fab (M/Γc (M )).
(ii) If fab (M ) > 0, then fab (M ) = fab (M/Γa (M )).
Proof. Suppose that there exists n ∈ N such that bn Γc (M ) = 0. Then
bn Haj (Γc (M )) = 0 for all j ∈ N0 . Now the exact sequence
0 −→ Γc (M ) −→ M −→ M/Γc (M ) −→ 0
induces, for each i ∈ N0 , an exact sequence
Hai (Γc (M )) −→ Hai (M ) −→ Hai (M/Γc (M )) −→ Hai+1 (Γc (M )),

and, since b ⊆ (0 : Haj (Γc (M ))) for all j ∈ N0 , it follows from 9.1.1 that
√ √
b ⊆ (0 : Hai (M )) if and only if b ⊆ (0 : Hai (M/Γc (M ))),
168 The Annihilator and Finiteness Theorems

so that fab (M ) = fab (M/Γc (M )).


(i) There exists n ∈ N such that cn Γc (M ) = 0. If b ⊆ c, then bn Γc (M ) = 0,
and the desired result follows from the above.
(ii) Assume that fab (M ) > 0. This means that bn Γa (M ) = 0 for some
n ∈ N0 . Now use the first paragraph of this proof with a taken as c.

9.1.9 Exercise. Let M be a finitely generated R-module for which M = aM .


Show that faR (M ) = gradeM a.

9.2 Adjusted depths


The b-finiteness dimension fab (M ) of 9.1.5 is one of the invariants used in
Faltings’ Annihilator Theorem. We now motivate the introduction of another
of the ingredients. This motivation concerns the situation of Exercise 7.3.4,
and we provide now a solution to the second part of that exercise.
Let M be a finitely generated R-module for which the ideal a + (0 : M )
is proper. Let p be a minimal prime of a + (0 : M ), and suppose that t :=
dimRp Mp > 0. Then, by 4.3.3, 4.2.2, and 1.1.3 used in conjunction with the
fact that p is a minimal prime of a + (0 : M ), we have

(Hat (M ))p ∼ t
= HaR p
(Mp ) ∼ t
= HaR p +(0:Rp Mp )
(Mp )
t t
= H(a+(0:M ))Rp (Mp ) = HpRp (Mp ).

By 7.3.3, this Rp -module is not finitely generated, since t := dimRp Mp > 0;


hence Hat (M ) is not finitely generated. We can therefore write fa (M ) ≤ t.
There exists a minimal prime q of (0 : M ) such that q ⊂ p and ht p/q = t.
Note that depthRq Mq = 0, that a ⊆ q, and that p is a minimal prime of a + q;
furthermore, we can write fa (M ) ≤ depthRq Mq + ht p/q.
Next suppose that r ∈ R is a non-zerodivisor on M and that there is a min-
imal prime p of a + (0 : M/rM ) such that t := dimRp (M/rM )p > 0.

Another application of Exercise 7.3.4 yields that Hat (M/rM ) is not finitely
generated. It therefore follows from the long exact sequence of local cohomol-
ogy modules induced by the exact sequence
r
0 −→ M −→ M −→ M/rM −→ 0
 
that either Hat (M ) or Hat +1 (M ) is not finitely generated. The reasoning in
the preceding paragraph applied to M/rM (rather than M ) shows that there
exists a minimal prime q of (0 : M/rM ) such that q ⊂ p and ht p /q = t .
Again, note that depthRq Mq = 1, that a ⊆ q , that p is a minimal prime of
9.2 Adjusted depths 169

a + q , and that
fa (M ) ≤ 1 + t = depthRq Mq + ht p /q .
It is hoped that, after this little discussion, the reader will not be too dis-
mayed to find that we now begin to consider the values of the expressions
depthRs Ms + ht(a + s)/s
for prime ideals s ∈ Supp M . (Here, ht R/s is to be interpreted as ∞.) Note
that, for such an s, we have ht(a + s)/s > 0 if and only if (a + s)/s is a
non-zero ideal of the integral domain R/s, that is, if and only if s ∈ Var(a).
We can think, loosely, of ht(a + s)/s as the ‘distance’ from s to Var(a) in
Spec(R), as it is, at least in the case when R is catenary (see [50, p. 31]), the
minimum length of a saturated ascending chain of prime ideals starting with s
(as its smallest term) and ending in Var(a).
9.2.1 Notation and Conventions. Let M be a finitely generated R-module.
For p ∈ Supp M we shall abbreviate depthRp Mp by depth Mp ; we shall
adopt the convention that the depth of a zero module over a local ring is ∞,
and accordingly, for p ∈ Spec(R) \ Supp M , we shall write depth Mp = ∞.
We interpret the height of the improper ideal R of R as ∞; accordingly, for
a proper ideal d of R, we write ht R/d = ∞.
9.2.2 Definitions. For a p ∈ Spec(R) and a finitely generated R-module M ,
we define the a-adjusted depth of M at p, denoted adja depth Mp , by
adja depth Mp := depth Mp + ht(a + p)/p.
Note that this is ∞ unless p ∈ Supp M and a + p ⊂ R, and then it is a
non-negative integer not exceeding dim M .
Let b be a second ideal of R. We define the b-minimum a-adjusted depth of
M , denoted by λba (M ), by
λba (M ) = inf {adja depth Mp : p ∈ Spec(R) \ Var(b)}
= inf {depth Mp + ht(a + p)/p : p ∈ Spec(R) \ Var(b)} .
Thus λba (M ) is either ∞ or a non-negative integer not exceeding dim M .
Faltings’ Annihilator Theorem, the main result of this chapter, asserts that,
under mild restrictions on R, the invariants λba (M ) and fab (M ) (where M is a
finitely generated R-module and b is a second ideal of R) are equal. Faltings’
original proof in [14] is rather different from the one we shall present below.
The special case of Faltings’ Annihilator Theorem in which b = a reduces
to Grothendieck’s Finiteness Theorem (see [26, Exposé VIII, Corollaire 2.3]),
170 The Annihilator and Finiteness Theorems

which is another fundamental result of local cohomology. This theorem pro-


vides information about the finiteness dimension fa (M ) of M relative to a
(again under mild restrictions on R).
Our approach to the proof of Faltings’ Annihilator Theorem begins with an
investigation of some of the properties of the invariant λba (M ) introduced in
9.2.2.

9.2.3 Remarks. Let b, c and d be further ideals of R such that c ⊆ b and


a ⊆ d. Let M be a finitely generated R-module. Then

b
(i) λba (M ) = λb√a (M ) = λ√a (M );
(ii) λba (M ) ≤ λbd (M ); and
(iii) λba (M ) ≤ λca (M ).

9.2.4 Lemma. Let b, c be further ideals of R, and let M be a finitely gener-


ated R-module.

(i) If b ⊆ c, then λba (M ) = λba (M/Γc (M )).


(ii) If λba (M ) > 0, then λba (M ) = λba (M/Γa (M )).

Proof. Suppose that, for each p ∈ Spec(R)\Var(b), we have (Γc (M ))p = 0.


Then (M/Γc (M ))p ∼
= Mp and

depth(M/Γc (M ))p + ht(a + p)/p = depth Mp + ht(a + p)/p,

so that λba (M ) = λba (M/Γc (M )).


(i) Suppose that b ⊆ c. Then, for each p ∈ Spec(R) \ Var(b) we have c ⊆ p,
so that (Γc (M ))p = 0. The claim therefore follows from the above paragraph.
(ii) Assume that λba (M ) > 0. Let p ∈ Spec(R) \ Var(b). We claim that
(Γa (M ))p = 0. This is clearly the case if a ⊆ p, so suppose that a ⊆ p.
If (Γa (M ))p were not zero, then there would be an associated prime ideal
q of Γa (M ) with q ⊆ p, and, necessarily, a ⊆ q. But then we would have
q ∈ Spec(R) \ Var(b) with adja depth Mq = depth Mq + ht(a + q)/q = 0,
contrary to the assumption that λba (M ) > 0. Thus (Γa (M ))p = 0 in all cases.
The desired result now follows from the first paragraph of this proof with a
taken as c.

9.2.5 Lemma. Let b be a second ideal of R, let M be a finitely generated


R-module, and let S be a multiplicatively closed subset of R. Then
−1
λba (M ) ≤ λSS −1 ba (S −1 M ).
9.3 The first inequality 171

Proof. Let P ∈ Spec(S −1 R) \ Var(S −1 b). Then there exists p ∈ Spec(R)


such that p ∩ S = ∅ and P = S −1 p. Now p ∈ Var(b), and so

λba (M ) ≤ depth Mp + ht(a + p)/p.

Now (S −1 M )P = (S −1 M )S −1 p ∼
= Mp as Rp -modules (when (S −1 M )S −1 p
is regarded as an Rp -module by means of the natural isomorphism Rp −→
(S −1 R)S −1 p ). Also

ht(a + p)/p ≤ ht(S −1 a + S −1 p)/S −1 p,

and so λba (M ) ≤ depth(S −1 M )P +ht(S −1 a+P)/P. The result follows.

The next lemma shows that λba (M ) is, in a certain sense, independent of the
base ring.

9.2.6 Lemma. Let b be a second ideal of R, let M be a finitely generated


R-module, and let c be an ideal of R such that c ⊆ (0 : M ). Then
(b+c)/c
λba (M ) = λ(a+c)/c (M ).

Proof. Let p ∈ Spec(R) \ Var(b). Either p ⊇ c or p ⊇ c. If p ⊇ c, then


Mp = 0 and depth Mp + ht(a + p)/p = ∞. This means that

λba (M ) = inf {depth Mp + ht(a + p )/p : p ∈ Var(c) \ Var(b)} .

However, if p ∈ Var(c) \ Var(b), then p/c ∈ Spec(R/c) \ Var((b + c)/c), and


it is an elementary exercise to check that depth Mp/c = depth Mp and

ht(((a + c)/c) + (p/c))/(p/c) = ht(a + p)/p.

The claim follows from these observations.

9.3 The first inequality


Our proof of Faltings’ Annihilator Theorem consists of demonstrations that,
with the notation of 9.1.5 and 9.2.2, we have fab (M ) ≤ λba (M ) and, under
mild restrictions on R, λba (M ) ≤ fab (M ). We start with the first inequality.
We prepare the ground with several lemmas.

9.3.1 Lemma. Let b be a second ideal of R, and let M be a finitely generated


R-module. Suppose that p ∈ Ass M \ Var(b) has ht(a + p)/p = 1. Then

b ⊆ (0 : Ha1 (M )).
172 The Annihilator and Finiteness Theorems

Proof. It suffices to show that a∩b ⊆ (0 : Ha1 (M )). Since ht(a+p)/p = 1,
we must have that a ⊆ p, so that p ∈ Var(a ∩ b). Therefore we may replace b
by a ∩ b; we therefore assume henceforth in this proof that b ⊆ a.
As p ∈ Ass M , there is an exact sequence
0 −→ R/p −→ M −→ N −→ 0
of R-modules and R-homomorphisms; this induces an exact sequence
Ha0 (N ) −→ Ha1 (R/p) −→ Ha1 (M ).

Since N is finitely generated, b ⊆ a ⊆ (0 : Ha0 (N )). Therefore, by 9.1.1,

it is enough for us to show that b ⊆ (0 : Ha1 (R/p)). By the Independence
Theorem 4.2.1, there is an R-isomorphism Ha1 (R/p) ∼ 1
= H(a+p)/p (R/p), and
so it is enough for us to show that
,
(b + p)/p ⊆ 1
0 : H(a+p)/p (R/p) .

It is therefore enough for us to prove that, if R is a domain, ht a = 1 and



0 = b ⊆ a, then b ⊆ (0 : Ha1 (R)). We shall achieve this by showing that, in
these circumstances, rHa1 (R) = 0 for every r ∈ a \ {0}.
Suppose that, on the contrary, rHa1 (R) = 0 for some r ∈ a \ {0}. Let q be
a minimal prime of a with ht q = 1. Let R := Rq and m := qRq , so that
(R , m ) is a 1-dimensional local domain. We deduce from 1.1.3 and 4.3.3 that
1
rHm  1 ∼ 1
 (R ) = rHaR (Rq ) = r(Ha (R))q = 0.
q

r ∈ m \ {0}. Since (R , m ) is a 1-dimensional local domain, we


Of course, √

have m = rR . Therefore, by 1.1.3 and 2.2.21(i),
1  1  ∼  
Hm  (R ) = HrR (R ) = Rr /R .

Thus rRr ⊆ R , so that r(1/r2 ) ∈ R . Hence r is a unit of R . This contradic-


tion completes the proof.
9.3.2 Lemma. Let b be a second ideal of R, and let M be a finitely gen-
erated R-module. Suppose that p ∈ SuppR M ∩ Var(a) \ Var(b) and let

t := gradeMp (aRp ). Then b ⊆ (0 : Hat (M )).
Proof. The hypotheses ensure that Mp = aRp Mp . In view of 4.3.3 and 6.2.7,
we have (Hat (M ))p ∼ = HaRt
(Mp ) = 0. By hypothesis, there exists b ∈ b \ p.
p √
It follows that (Hat (M ))b = 0, so that b ∈ (0 : Hat (M )).
9.3.3 Exercise. Show that, for a finitely generated R-module M for which
M = aM , we have λRa (M ) = gradeM a. (Here is a hint: observe that

gradeM a = inf{depth Mp : p ∈ Var(a)}.)


9.3 The first inequality 173

9.3.4 Lemma. Let M be a finitely generated R-module, and p, s ∈ Spec(R)


with p ⊆ s. Then
depth Ms ≤ depth Mp + ht s/p.

Proof. We use induction on h := ht s/p, there being nothing to prove when


h = 0.
Consider the case in which h = 1. Clearly, we can assume p ∈ Supp M .
Now Mp is Rp -isomorphic to the localization of Ms at the prime ideal pRs
(when that localization is regarded as an Rp -module by means of the natural

=
isomorphism Rp −→ (Rs )pRs ). In order to establish the desired result when
h = 1, it is therefore enough for us to show that, if (R, m) is local, M is a
finitely generated R-module, and p ∈ Supp M is such that dim R/p = 1, then
depth M ≤ depth Mp + 1. This we do.
Let gradeM p = t (see 6.2.4), and let a1 , . . . , at be a maximal M -sequence
contained in p. Then a1 /1, . . . , at /1 is an Mp -sequence contained in pRp . Let
t
N := M/ j=1 aj M . Then

depth N = depth M − t, depth Np = depth Mp − t

and gradeN p = 0. Thus we can achieve our aim by showing that depth N ≤
depth Np + 1. We can assume that depth N > 0, since otherwise there is
nothing to prove. This assumption means that m ∈ Ass N . But p consists

entirely of zerodivisors on N , and so p ⊆ p ∈Ass N p . It therefore follows
from the Prime Avoidance Theorem and the fact that dim R/p = 1 that p ∈
Ass N , and so depth N ≤ dim R/p = 1 by [50, Theorem 17.2].
The claim in the lemma has therefore been proved when h = 1; suppose
now that h > 1 and the claim has been proved for smaller values of h. Since
ht s/p = h, there exists q ∈ Spec(R) such that p ⊂ q ⊂ s, ht q/p = 1 and
ht s/q = h − 1. We can now use the inductive hypothesis and the (already
established) truth of the claim in the case when h = 1 to see that

depth Ms ≤ depth Mq + h − 1
≤ depth Mp + 1 + h − 1 = depth Mp + h.

This completes the inductive step, and the proof.

9.3.5 Lemma. Let M be a finitely generated R-module. Suppose that p, q ∈


Spec(R) are such that q is a minimal prime of a + p and ht q/p = ht(a + p)/p.
Then
adja depth Ms ≤ adja depth Mp

for all s ∈ Spec(R) with p ⊆ s ⊆ q.


174 The Annihilator and Finiteness Theorems

Proof. It is clear that q ⊇ a + s, and so ht q/s ≥ ht(a + s)/s. Hence, on use


of 9.3.4, we have
adja depth Ms = depth Ms + ht(a + s)/s
≤ depth Ms + ht q/s
≤ depth Mp + ht s/p + ht q/s
≤ depth Mp + ht q/p
= depth Mp + ht(a + p)/p = adja depth Mp ,
as required.
9.3.6 Lemma. Let s, q ∈ Spec(R) with s ⊂ q be such that ht q/s > 1. Let
a ∈ q \ s. Then there exists t ∈ Spec(R) with s ⊂ t ⊂ q such that a ∈ t.
Proof. Let R := R/s; for c ∈ R, denote the natural image of c in R by c.
Let the minimal primes of the proper principal ideal Ra be p1 /s, . . . , ph /s,
where p1 , . . . , ph ∈ Spec(R). Now ht pi /s = 1 for all i = 1, . . . , h, by Krull’s
Principal Ideal Theorem. Therefore, since ht q/s > 1, it follows from the
h
Prime Avoidance Theorem that q/s ⊆ i=1 pi /s. Hence, there exists b ∈ R
such that
h
b ∈ q/s \ pi /s.
i=1

Since Rb ⊆ q/s, there exists t ∈ Spec(R) with s ⊂ t ⊆ q such that t/s is a


minimal prime of Rb. Since ht q/s > 1 and ht t/s = 1, it follows that t ⊂ q.
h
Since b ∈ t \ i=1 pi , we can deduce that t is different from all of p1 , . . . , ph .
Hence a ∈ t.
9.3.7 Theorem. Let b be a second ideal of R, and let M be a finitely gener-
ated R-module. Then
fab (M ) ≤ λba (M ).
Proof. Set λ := λba (M ). If λ = ∞, then there is nothing to prove; we there-
fore suppose that λ is finite, and argue by induction on λ.
When λ = 0, there exists p ∈ Spec(R) \ Var(b) with
adja depth Mp = depth Mp + ht(a + p)/p = 0.
This means that depth Mp = 0 and p ∈ Var(a), so that
pRp ∈ AssRp Mp ∩ Var(aRp )

and gradeMp (aRp ) = 0. Therefore b ⊆ (0 : Ha0 (M )) by 9.3.2, so that
fab (M ) = 0.
9.3 The first inequality 175

Now suppose that λ > 0 and assume, inductively, that the desired inequality
has been proved for smaller values of λ. We can assume that fab (M ) > 0.
By 9.1.8(ii) and 9.2.4(ii), we have fab (M ) = fab (M/Γa (M )) and λba (M ) =
λba (M/Γa (M )). We may therefore replace M by M/Γa (M ), and, in view of
2.1.2, assume that M is a-torsion-free for the remainder of this proof.
Choose p ∈ Spec(R) \ Var(b) with

adja depth Mp = depth Mp + ht(a + p)/p = λ.

Assume first that p ∈ Var(a). Then t := gradeMp aRp ≤ depth Mp = λ, so


that fab (M ) ≤ t ≤ λ by 9.3.2.
Assume now that p ∈ Var(a). If λ = 1, then

adja depth Mp = depth Mp + ht(a + p)/p = 1.

These conditions mean that ht(a + p)/p = 1 and depth Mp = 0. Hence



p ∈ Ass M , and so it follows from Lemma 9.3.1 that b ⊆ (0 : Ha1 (M )).
Hence fab (M ) ≤ 1 = λ.
So we may assume that λ > 1 (and p ∈ Var(a)). Since p ∈ Var(a ∩ b),
there exists a ∈ a ∩ b \ p. Let q be a minimal prime of a + p such that ht q/p =
ht(a + p)/p. Then p belongs to the set

Σ := {s ∈ Spec(R) : p ⊆ s ⊆ q and a ∈ s } ;

let s be a maximal member of Σ. Now a ∈ a ∩ b ⊆ a ⊆ q, and so s ⊂ q.


Lemma 9.3.6 shows that ht q/s = 1. Note that the fact that a ∈ s ensures that
s ∈ Spec(R) \ Var(b).
We can now deduce from 9.3.5 and the definition of λ that

λ ≤ adja depth Ms ≤ adja depth Mp = λ.

Therefore adja depth Ms = λ. Since a ∈ a \ s, we have a ⊆ s. It follows


that q is a minimal prime of a + s such that ht q/s = ht(a + s)/s = 1. Thus
we can replace p by s and so make the additional assumption that ht q/p =
ht(a + p)/p = 1.
We now propose to localize at q. Note that pRq ∈ Spec(Rq ) \ Var(bRq ),
that qRq is a minimal prime of aRq + pRq , that

ht qRq /pRq = ht(aRq + pRq )/pRq = 1,

that Mq is a finitely generated aRq -torsion-free Rq -module, and that

adjaRq depth(Mq )pRq = depth(Mq )pRq + ht(aRq + pRq )/pRq


= depth Mp + 1 = λ.
176 The Annihilator and Finiteness Theorems
bR
We can therefore deduce from 9.2.5 that λaRqq (Mq ) = λ. Also, by 9.1.6, we
bR
have fab (M ) ≤ faRqq (Mq ). These considerations mean that it is enough for us
to prove the desired result under the additional assumption that (R, q) is local,
and we make this assumption in what follows.
Our next aim is to show that p contains a non-zerodivisor on M . Suppose
that this is not the case, so that p ⊆ s for some s ∈ Ass M . Since M is a-
torsion-free, a contains a non-zerodivisor on M , by 2.1.1(ii), so that, as q ⊇ a,
we see that s ⊂ q. As ht q/p = 1, it follows that p = s ∈ Ass M . This means
that depth Mp = 0, and λ = depth Mp + ht(a + p)/p = 1. This contradiction
shows that there exists r ∈ p which is a non-zerodivisor on M .
Now

λba (M/rM ) ≤ adja depth(M/rM )p = depth(M/rM )p + 1


= depth Mp − 1 + 1 < λ.

Therefore, by the induction hypothesis, fab (M/rM ) ≤ λba (M/rM ) < λ. Set

u := fab (M/rM ), so that b ⊆ (0 : Hau (M/rM )). But the exact sequence
r
0 −→ M −→ M −→ M/rM −→ 0

induces an exact sequence

Hau (M ) −→ Hau (M/rM ) −→ Hau+1 (M ),



and so it follows from 9.1.1 that b ⊆ (0 : Hai (M )) for i = u or i = u + 1.
Thus
fab (M ) ≤ u + 1 = fab (M/rM ) + 1 ≤ λ.

This completes the inductive step, and the proof.

9.4 The second inequality


We now embark on the more difficult part of our proof of Faltings’ Annihilator
Theorem, namely the proof that, when R is a homomorphic image of a regular
ring, then, with the notation of 9.1.5 and 9.2.2, we have λba (M ) ≤ fab (M ).
(Recall (see [50, p. 157]) that a commutative Noetherian ring is said to be
regular precisely when its localizations at all of its prime ideals are regular
local rings.)

9.4.1 Lemma. Let M be a finitely generated R-module, and let

p ∈ Spec(R) \ Supp M.
9.4 The second inequality 177

Then there exists s ∈ R \ p such that, for every ideal b of R, we have


sHbi (M ) = 0 for all i ∈ N0 .
Proof. Since M is finitely generated, there exists s ∈ (0 : M ) \ p, and this s
has the desired properties because the functors Hbi (i ∈ N0 ) are R-linear.
9.4.2 Lemma. Let M be a finitely generated R-module, and let p ∈ Spec(R)
be such that Mp is a non-zero free Rp -module. Then there exist t ∈ N and an
R-homomorphism π : Rt −→ M such that (Ker π)p = (Coker π)p = 0.
Proof. Set t := rankRp Mp . There exist m1 , . . . , mt ∈ M such that m1 /1,
. . . , mt /1 form a base for the free Rp -module Mp . Let π : Rt −→ M be the
t
R-homomorphism for which π((r1 , . . . , rt )) = i=1 ri mi for all (r1 , . . . , rt )
∈ Rt . Then the localization of π at p is an isomorphism, and so (Ker π)p =
(Coker π)p = 0.
9.4.3 Lemma. Let M be a finitely generated R-module, and let p ∈ Spec(R)
be such that Mp is a non-zero free Rp -module. Then there exists s ∈ R \p such
that, for every proper ideal b of R, we have
sHbi (M ) = 0 for all i < grade b.
Proof. By 9.4.2, there exist t ∈ N and an R-homomorphism π : Rt −→ M
such that (Ker π)p = (Coker π)p = 0. By 9.4.1, there exist u, v ∈ R \ p such
that, for every ideal b of R, we have
uHbi (Ker π) = vHbi (Coker π) = 0 for all i ∈ N0 .
Set s := uv. Let b be a proper ideal of R and let i ∈ N0 with i < grade b.
Now the exact sequence 0 −→ Ker π −→ Rt −→ Im π −→ 0 induces an
exact sequence Hbi (Rt ) −→ Hbi (Im π) −→ Hbi+1 (Ker π). Since i < grade b
we have Hbi (R) = 0, by 6.2.7, so that Hbi (Rt ) = 0 in view of the additivity of
the functor Hbi . By the immediately preceding paragraph, uHbi+1 (Ker π) = 0.
Therefore uHbi (Im π) = 0.
Next, the exact sequence 0 −→ Im π −→ M −→ Coker π −→ 0 induces
an exact sequence Hbi (Im π) −→ Hbi (M ) −→ Hbi (Coker π). As uHbi (Im π)
= 0 and vHbi (Coker π) = 0, it follows that sHbi (M ) = uvHbi (M ) = 0.
9.4.4 Conventions. Let M be an R-module. We shall denote the projective
dimension of M by proj dim M or, occasionally, by proj dimR M if it is es-
sential to specify the underlying ring concerned. In particular, the reader is
warned that, when S is a multiplicatively closed subset of R, we shall always
write proj dim S −1 M rather than proj dimS −1 R S −1 M . We adopt the con-
vention that a zero module has projective dimension −∞.
178 The Annihilator and Finiteness Theorems

9.4.5 Lemma. Let M be a finitely generated R-module, and let p ∈ Spec(R)


be such that proj dim Mp < ∞. Then there exists s ∈ R\p such that, for every
proper ideal b of R, we have
sHbi (M ) = 0 for all i < grade b − proj dim Mp .
Note. In the case when Mp = 0, one should interpret the ‘− − ∞’ in the
above statement as ‘∞’.
Proof. Set h := proj dim Mp . We use induction on h. If h = −∞, then
Mp = 0 and the result is clear from Lemma 9.4.1. When h = 0, the desired
result follows from Lemma 9.4.3, since then Mp is a non-zero free Rp -module.
We therefore assume, inductively, that h > 0 and the result has been proved
for smaller values of h. There is a non-zero, finitely generated free R-module
F and an exact sequence 0 −→ N −→ F −→ M −→ 0 of R-modules and
homomorphisms. Localization yields an exact sequence
0 −→ Np −→ Fp −→ Mp −→ 0.
Therefore proj dim Np = h − 1 (since h > 0), and so, by the inductive hy-
pothesis, there exists s ∈ R \ p such that, for every proper ideal b of R, we
have
sHbi (N ) = 0 for all i < grade b − h + 1.
Thus sHbi+1 (N ) = 0 for all i < grade b − h. Let b be a proper ideal of R and
let i ∈ N0 with i < grade b − h.
The exact sequence 0 −→ N −→ F −→ M −→ 0 induces a further
exact sequence Hbi (F ) −→ Hbi (M ) −→ Hbi+1 (N ). Since i < grade b we
have Hbi (F ) = 0, by 6.2.7 and the additivity of the functor Hbi . By the imme-
diately preceding paragraph, sHbi+1 (N ) = 0. Therefore sHbi (M ) = 0. This
completes the inductive step.
9.4.6 Lemma. Let M be a finitely generated R-module, and let p ∈ Spec(R)
be such that proj dim Mp < ∞. Then there exists s ∈ R \ p such that
proj dim Ms = proj dim Mp .
Proof. Set h := proj dim Mp . We use induction on h. If h = −∞, then
Mp = 0 and there exists s ∈ R \ p such that sM = 0. Hence Ms = 0 and
proj dim Ms = −∞ = h.
When h = 0, then, by Lemma 9.4.2, there exist t ∈ N and an R-homomor-
phism π : Rt −→ M such that (Ker π)p = (Coker π)p = 0. Then there exists
s ∈ R \ p such that s Ker π = s Coker π = 0. It follows that (Ker π)s =
(Coker π)s = 0 and πs : (Rt )s −→ Ms is an isomorphism. Since Ms = 0,
we have proj dim Ms = 0 = h.
9.4 The second inequality 179

We therefore assume, inductively, that h > 0 and the result has been proved
for smaller values of h. There is a non-zero, finitely generated free R-module
F and an exact sequence 0 −→ N −→ F −→ M −→ 0 of R-modules and
homomorphisms. Localization yields an exact sequence

0 −→ Np −→ Fp −→ Mp −→ 0.

Note that proj dim Np = h − 1, and so, by the inductive hypothesis, there
exists s ∈ R \ p such that proj dim Ns = proj dim Np = h − 1. But then the
exact sequence
0 −→ Ns −→ Fs −→ Ms −→ 0

shows that proj dim Ms ≤ h. Since p ∩ {sn : n ∈ N0 } = ∅, we see that Mp ∼ =


(Ms )pRs (when the latter is regarded as an Rp -module by means of the natural
isomorphism Rp −→ (Rs )pRs ). Hence h = proj dim Mp ≤ proj dim Ms ≤
h, so that proj dim Ms = proj dim Mp . This completes the inductive step.

9.4.7 Corollary. Let M be a finitely generated R-module. Then for each


t ∈ N0 ∪ {−∞}, the set

Ut (M ) := {p ∈ Spec(R) : proj dim Mp ≤ t}

is an open subset of Spec(R) (in the Zariski topology).

Proof. Let p ∈ Ut (M ), so that h := proj dim Mp ≤ t. By 9.4.6, there exists


s ∈ R \ p such that proj dim Ms = h. Therefore, for each q in the open neigh-
bourhood Spec(R) \ Var(sR) of p, we have proj dim Mq ≤ proj dim Ms ≤ t
(because Mq ∼ = (Ms )qRs when the latter is regarded as an Rq -module by
means of the natural isomorphism Rq −→ (Rs )qRs ), so that q ∈ Ut (M ).

9.4.8 Notation and Remarks. Let M be a finitely generated R-module. For


each t ∈ N0 ∪ {−∞}, let

Ut (M ) := {p ∈ Spec(R) : proj dim Mp ≤ t} ,

let Ct (M ) = Spec(R) \ Ut (M ), and let ct (M ) = p∈Ct (M ) p. Then

(i) Ut (M ) is an open, and so Ct (M ) is a closed, subset of Spec(R), for all


t ∈ N0 (by 9.4.7);
(ii) Spec(R) \ Supp M = U−∞ (M ) and

U−∞ (M ) ⊆ U0 (M ) ⊆ U1 (M ) ⊆ · · · ⊆ Ut (M ) ⊆ · · · ;

(iii) Supp M = C−∞ (M ) ⊇ C0 (M ) ⊇ C1 (M ) ⊇ · · · ⊇ Ct (M ) ⊇ · · · ;



(iv) (0 : M ) = c−∞ (M ) ⊆ c0 (M ) ⊆ c1 (M ) ⊆ · · · ⊆ ct (M ) ⊆ · · · ;
180 The Annihilator and Finiteness Theorems

(v) for each t ∈ N0 ∪ {−∞}, the ideal ct (M ) is radical (that is, is equal to
its own radical), and Var(ct (M )) = Ct (M ); and
(vi) if R is regular, so that, for all p ∈ Spec(R), the localization Rp has finite
global dimension and proj dim Mp is finite, then
 *
Ut (M ) = Spec(R) and Ct (M ) = ∅.
t∈N0 ∪{−∞} t∈N0 ∪{−∞}

9.4.9 Exercise. Let M be a finitely generated R-module and let S be a


multiplicatively closed subset of R. Show that, with the notation of 9.4.8,

ct (S −1 M ) = S −1 (ct (M )) for all t ∈ N0 ∪ {−∞} .

9.4.10 Proposition. Let M be a finitely generated R-module, and let t ∈


N0 ∪ {−∞}. We use the notation of 9.4.8. There exists n ∈ N such that, for
every proper ideal d of R, we have

ct (M )n Hdi (M ) = 0 for all i < grade d − t.

Note. In the case when t = −∞, one should interpret the ‘− − ∞’ in the
above statement as ‘∞’.

Proof. Let p ∈ Ut (M ), so that proj dim Mp ≤ t. By 9.4.5, there exists


sp ∈ R \ p such that, for every proper ideal d of R, we have sp Hdi (M ) = 0 for
all i < grade d − proj dim Mp , and therefore for all i < grade d − t.

Set g := p ∈Ut (M ) sp R, and observe that, for every proper ideal d of R,
we have gHdi (M ) = 0 for all i < grade d − t. Note that no prime ideal in
Ut (M ) contains g. The latter statement implies that Var(g) ⊆ Ct (M ), so that

ct (M ) ⊆ g. Hence there exists n ∈ N such that ct (M )n ⊆ g, and the result
follows from this.

9.4.11 Exercise. Assume that R is a homomorphic image of a regular (com-


mutative Noetherian) ring, and let M be a finitely generated R-module with
the property that Supp M has exactly one minimal member. Assume that c
is an ideal of R such that Mp is a Cohen–Macaulay Rp -module for all p ∈
Spec(R) \ Var(c). Prove that there exists n ∈ N such that

cn Hqi (M ) = 0 for all q ∈ Var(c) and all i < dimRq Mq .

9.4.12 Exercise. Let the situation be as in Proposition 9.4.10. Prove that there
exists n ∈ N such that, for every choice of flat ring homomorphism f : R →
R of commutative Noetherian rings and every choice of proper ideal B of
R , we have

ct (M )n HB
i
 (M ⊗R R ) = 0 for all i < grade B − t.
9.4 The second inequality 181

(Here is a hint: make appropriate modifications to the preparatory results, such


as 9.4.1, 9.4.3 and 9.4.5, that were used in our approach to the proof of Propo-
sition 9.4.10.)
9.4.13 Exercise. (The result of this exercise is due to C. L. Huneke.) As-
sume that R is a homomorphic image of a regular (commutative Noetherian)
ring for which there exists a non-zerodivisor c ∈ R such that the ring Rc is
Cohen–Macaulay. Assume that either (a) R is an integral domain, or (b) R is
an equidimensional (see [50, p. 250]) local ring.
Use Exercise 9.4.12 above to prove that there exists n ∈ N such that, for
all choices of r ∈ N0 and all choices of an ideal B of the polynomial ring
R[X1 , . . . , Xr ] (interpret this as R in the case when r = 0), we have
cn HB
i
 (R[X1 , . . . , Xr ]) = 0 for all i < ht B .
9.4.14 Lemma. Let (R, m) be a regular local ring of dimension d, let M
be a finitely generated R-module, and let b be an ideal of R. Then, with the
notation of 9.4.8, we have b ⊆ cd−λbm (M ) (M ).
Proof. Set t := d − λbm (M ). It is enough to prove that, for each p ∈ Ct (M ),
we must have b ⊆ p. Suppose that, for one such p, we have b ⊆ p, and look
for a contradiction. Then (p ∈ Supp M and)
∞ > adjm depth Mp = depth Mp + ht(m + p)/p ≥ λbm (M ).
Since R is a catenary domain, ht(m + p)/p = ht m/p = d − ht p. As Rp is a
regular local ring, it follows from the Auslander–Buchsbaum–Serre Theorem
that depth Mp + proj dim Mp = dim Rp = ht p. Therefore
ht p − proj dim Mp + d − ht p ≥ λbm (M ),
so that proj dim Mp ≤ t. This contradiction completes the proof.
We remind the reader that our present major aim is a proof of the fact that,
when R is a homomorphic image of a regular ring, then, with the notation of
9.1.5 and 9.2.2, we have λba (M ) ≤ fab (M ). In view of 9.1.7 and 9.2.6, it will
be enough to do this in the case when R itself is regular. In the light of this, the
next result already proves our desired inequality in a special case.
9.4.15 Proposition. Assume that R is regular, and that dim R/a = 0. Let b
be a second ideal of R, and let M be a finitely generated R-module. Then
λba (M ) ≤ fab (M ).
Proof. Since dim R/a = 0, we have Var(a) = ass a is a finite set of maximal
ideals of R: let its members be m1 , . . . , mh . Consider an integer j with 1 ≤
182 The Annihilator and Finiteness Theorems
bRm
j ≤ h. Set tj := ht mj − λmj Rjm (Mmj ). Since ht mj = grade mj Rmj , we
j
bRm
see that grade mj Rmj − tj = λmj Rjm (Mmj ). By 9.4.10, there exists nj ∈ N
j
such that
bRm
ctj (Mmj )nj Hm
i
j Rm
(Mmj ) = 0 for all i < λmj Rjm (Mmj ).
j j

By 9.4.14, we have bRmj ⊆ ctj (Mmj ).


Set n := max {n1 , . . . , nh }. We can use 1.1.3 and 4.3.3 to see that
i
Hm j Rm
i
(Mmj ) = HaR m
(Mmj ) ∼
= (Hai (M ))mj for all i ∈ N0 .
j j

Also
bRm bRm
λmj Rjm (Mmj ) = λaRmj (Mmj ) ≥ λba (M ),
j j

by 9.2.3(i) and 9.2.5. It therefore follows that


(bn Hai (M ))mj ∼
= (bRmj )n Hm
i
j Rmj
(Mmj ) = 0 for 1 ≤ j ≤ h, i < λba (M ).
Since a local cohomology module with respect to a is a-torsion, the sup-
port of such a local cohomology module must be contained in Var(a) =
{m1 , . . . , mh }. Hence
 
Supp bn Hai (M ) = ∅ for all i < λba (M ).
In view of the definition of fab (M ) (see 9.1.5), this completes the proof.
9.4.16 Theorem. Assume that R is a homomorphic image of a regular (com-
mutative Noetherian) ring. Let b be a second ideal of R, and let M be a finitely
generated R-module. Then
λba (M ) ≤ fab (M ).
Proof. In view of 9.1.7 and 9.2.6, we can, and do, assume that R itself is
regular.
We suppose that λba (M ) > fab (M ) and look for a contradiction. Let I(R)
denote the set of all ideals of R. Since R is Noetherian, we can, and do, assume
that a is a maximal member of the set
a ∈ I(R) : λba (M ) > fab (M ) .
Note that dim R/a > 0, by 9.4.15. Set λ := λba (M ), and let q be a minimal
prime of a such that ht q = ht a. By 9.4.14, 9.4.9, 9.2.5, 9.2.3(i) and 9.4.8(iv),
 
bRq ⊆ cht a−λbRq (M ) (Mq ) = cht a−λbRq (M ) (M ) ⊆ (cht a−λ (M ))q .
qRq q aRq q
q

Therefore, there exists u ∈ R \ q such that bRu ⊆ (cht a−λ (M ))Ru . Since
9.5 The main theorems 183

dim R/a > 0, there exists s ∈ Var(a) which is not a minimal prime of a: let
v ∈ s \ q and put s := uv. Note that (by 9.4.9 again)
a ⊂ a + Rs ⊂ R and bRs ⊆ (cht a−λ (M ))s = cht(aRs )−λ (Ms ).
Since ht aRs = grade aRs , it follows from 9.4.10 that there exists n ∈ N
such that
 n i
cht(aRs )−λ (Ms ) HaR s
(Ms ) = 0 for all i < λ.
Hence bn HaR
i
s
(Ms ) = 0 for all i < λ (when HaR i
s
(Ms ) is regarded as an
R-module in the natural way). Therefore, by the Independence Theorem 4.2.1,
we have bn Hai (Ms ) = 0 for all i < λ.
Since a ⊂ a + Rs, it follows from the ‘maximality’ assumption about a
made in the second paragraph of this proof that λba+Rs (M ) ≤ fa+Rs b
(M ).

Now λ = λa (M ) ≤ λa+Rs (M ), by 9.2.3(ii), and so there exists n ∈ N such
b b

that bn Ha+Rs
i
(M ) = 0 for all i < λ.
By 8.1.2(i), there is, for each i < λ, an exact sequence
i
Ha+Rs (M ) −→ Hai (M ) −→ Hai (Ms ).

It now follows from 9.1.1 that b ⊆ (0 : Hai (M )) for all i < λ. This shows
that λ = λba (M ) ≤ fab (M ), and this contradiction completes the proof.

9.5 The main theorems


We can now put together Theorems 9.3.7 and 9.4.16 to prove Faltings’ Anni-
hilator Theorem.
9.5.1 Faltings’ Annihilator Theorem. (See G. Faltings [14].) Assume that
R is a homomorphic image of a regular (commutative Noetherian) ring. Let b
be a second ideal of R, and let M be a finitely generated R-module. Then
λba (M ) = fab (M ).
Proof. This is now immediate from Theorems 9.3.7 and 9.4.16.
The special case of Faltings’ Annihilator Theorem in which a = b is Groth-
endieck’s Finiteness Theorem.
9.5.2 Grothendieck’s Finiteness Theorem. (See A. Grothendieck [26, Ex-
posé VIII, Corollaire 2.3].) Assume that R is a homomorphic image of a reg-
ular (commutative Noetherian) ring, and let M be a finitely generated R-
module. Then
λaa (M ) = fa (M ).
184 The Annihilator and Finiteness Theorems

In other words, there exists i ∈ N such that Hai (M ) is not finitely generated
if and only if Supp M ⊆ Var(a); moreover, when this is the case, the least
i ∈ N such that Hai (M ) is not finitely generated is equal to
min {depth Mp + ht(a + p)/p : p ∈ Supp M \ Var(a)} .
Proof. Put a = b in Faltings’ Annihilator Theorem 9.5.1, and use 9.1.2 for
the last part.
9.5.3 Example. Let K be a field, and let R = K[X, Y 2 , XY, Y 3 ], a sub-
ring of the ring of polynomials K[X, Y ]. Let m denote the maximal ideal
(X, Y 2 , XY, Y 3 ) of R.
In the ring of polynomials R[Z], let N = mR[Z] + ZR[Z]. We first cal-
culate the invariant λNN (R[Z]). Let P ∈ Spec(R[Z]) \ Var(N): we calculate
adjN depth R[Z]P . Two cases arise: if P = 0, then the fact that ht N = 3
implies that adjN depth R[Z]0 = 3; if P = 0, then the fact that R[Z]P is a
local domain which is not a field ensures that
adjN depth R[Z]P = depth R[Z]P + ht(N + P)/P ≥ 1 + 1 = 2.
Thus λNN (R[Z]) ≥ 2. To see that λN (R[Z]) is exactly 2, note that mR[Z] ∈
N

Spec(R[Z]) \ Var(N). Now ht N/mR[Z] = 1 since the ideal N/mR[Z] is


1
principal. Moreover, it follows from 2.3.6(v) that Hm (R) = 0; hence, since
R is a domain, we can deduce from 6.2.7 that grade m = 1. Now X is a
non-zerodivisor on R; the fact that m is a maximal ideal of R ensures that
m ∈ assR RX; and it is now easy to deduce that depth R[Z]mR[Z] = 1. Hence
adjN depth R[Z]mR[Z] = depth R[Z]mR[Z] + ht N/mR[Z] = 1 + 1 = 2.
Therefore λN
N (R[Z]) = 2.
Grothendieck’s Finiteness Theorem 9.5.2 thus tells us that the finiteness di-
mension fN (R[Z]) is 2. Let us show this directly.
Since Z, X is an R[Z]-sequence contained in N, it follows from 6.2.7 that
HN0
(R[Z]) = HN1
(R[Z]) = 0. Further, for each t ∈ N, the exact sequence
Zt
0 −→ R[Z] −→ R[Z] −→ R[Z]/Z t R[Z] −→ 0
induces an exact sequence
Zt
0 −→ HN
1
(R[Z]/Z t R[Z]) −→ HN
2
(R[Z]) −→ HN
2
(R[Z]);
2 (R[Z]) Z ) ∼
t 1
hence (0 :HN = HN (R[Z]/Z t R[Z]). But, by 1.1.3 and 4.2.2,
1
HN (R[Z]/Z t R[Z]) = HmR[Z]+Z
1 t
t R[Z] (R[Z]/Z R[Z])

∼ 1
= HmR[Z] (R[Z]/Z t R[Z]).
9.5 The main theorems 185

The Independence Theorem 4.2.1 shows that, as R-modules,


1
HmR[Z] (R[Z]/Z t R[Z]) ∼ 1
= Hm (R[Z]/Z t R[Z]) ∼ 1
= Hm (Rt );
this is isomorphic to the R-module K t , by 2.3.6(v). Therefore
2 (R[Z]) Z ) ⊂ (0 :H 2 (R[Z]) Z for all t ∈ N,
t t+1
(0 :HN N
)
2
and so HN (R[Z]) is not finitely generated. Thus we have shown directly that
the finiteness dimension fN (R[Z]) is 2.
9.5.4 Exercise. Assume that the local ring (R, m) is a homomorphic image
of a regular local ring, and let M be a finitely generated R-module.
(i) Show that
fm (M ) = inf {depth Mp + 1 : p ∈ Spec(R) and dim R/p = 1} .
(ii) Suppose that n := dim M > 0. Prove that fm (M ) = n if and only if
Mp is a Cohen–Macaulay Rp -module of dimension n − dim R/p for all
p ∈ Supp M \ {m}.
9.5.5 Exercise. Assume that (R, m) is a local domain which is a homomor-
phic image of a regular local ring; assume also that d := dim R > 0, and that
Rp is a Cohen–Macaulay ring for all p ∈ Spec(R) \ {m}.
i
Show that, if Hm (R) = 0 for all integers i such that 1 < i < d, then there
exists a Cohen–Macaulay subring of the quotient field of R which contains R
and is a finitely generated R-module. (Here is a hint: think of Dm (R).)
9.5.6 Exercise. Assume that (R, m) is a local domain which is a homomor-
phic image of a regular local ring, and that the ideal a is proper. Show that
fa (R) > 1 if and only if ht a = 1.
9.5.7 Exercise. Assume that (R, m) is local, and let M be a non-zero finitely
generated R-module of dimension n > 0. We say that M is a generalized
i
Cohen–Macaulay R-module precisely when Hm (M ) is finitely generated for
all i = n. (Such modules were called ‘quasi-Cohen–Macaulay modules’ by P.
Schenzel in [72, p. 238]; these modules were investigated by Schenzel, N. V.
Trung and N. T. Cuong in [76], and, since the publication of that paper, the
terminology ‘generalized Cohen–Macaulay module’ seems to have become
more commonplace than ‘quasi-Cohen–Macaulay module’.)
(i) Show that, if M is a generalized Cohen–Macaulay R-module, then we
have dim R/p = n for all p ∈ Ass M \ {m} and Mq is a Cohen–
Macaulay Rq -module for all q ∈ Supp M \ {m}.
(ii) Show conversely, that, if
186 The Annihilator and Finiteness Theorems

(a) R is a homomorphic image of a regular ring,


(b) dim R/p = n whenever p is a minimal member of Supp M , and
(c) Mq is a Cohen–Macaulay Rq -module for all q ∈ Supp M \ {m},
then M is a generalized Cohen–Macaulay R-module.
9.5.8 Exercise. Assume that (R, m) is local, and that the non-zero finitely
generated R-module M of dimension n > 0 is a generalized Cohen–Macaulay
R-module (see 9.5.7 above). Note that m ∈ Ass(M/Γm (M )) and Mp ∼ =
(M/Γm (M ))p for all p ∈ Spec(R) \ {m}.

(i) Let r ∈ m be such that dim M/rM = n − 1. Show that r is a non-


zerodivisor on M/Γm (M ) and that, if n > 1, then M/rM is a general-
ized Cohen–Macaulay R-module.
(ii) Use part (i) and induction on n to show that every saturated chain of
prime ideals from a minimal member of Supp M (as smallest term) to
m (as largest term) has length n.
(iii) Deduce from part (ii) that the ring R/(0 : M ) is catenary.
9.5.9 Definitions. Assume that (R, m) is local, and let M be a non-zero
finitely generated R-module of dimension n > 0.

(i) By a system of parameters for M we mean a sequence (ri )ni=1 of n


n
elements of m such that M/ i=1 ri M has finite length. We say that
r1 , . . . , rn form a system of parameters for M precisely when (ri )ni=1
is a system of parameters for M . By a parameter for M , we mean a
member of a system of parameters for M .
n
(ii) Let (ri )ni=1 be a system of parameters for M ; let q := i=1 ri R. We say
that (ri )ni=1 is a standard system of parameters for M precisely when
k
qHm j
M/ i=1 ri M = 0 for all j, k ∈ N0 with j + k < n.

(See Stückrad and Vogel [84, p. 261].)


9.5.10 Exercise. Assume (R, m) is local, and let M be a non-zero finitely
generated R-module of dimension n > 1. Let (vi )ni=1 be a system of pa-
t
rameters for M , let t ∈ N with t < n, and let M  := M/ i=1 vi M . Let
r1 , . . . , rn−t ∈ m. Show that

(i) dim M  = n − t;
(ii) v1 , . . . , vt , r1 , . . . , rn−t form a system of parameters for M if and only
n−t
if (ri )i=1 is a system of parameters for M  ;
(iii) if (ri )i=1 is a system of parameters for M  and h ∈ N0 is such that
n−t
h t h
h < n − t, then M  / i=1 ri M  ∼ = M/( i=1 vi M + i=1 ri M );
9.5 The main theorems 187

(iv) if every system of parameters for M is standard, then every system of


parameters for M  is standard.

9.5.11 Exercise. Assume that (R, m) is local, and that M is a generalized


Cohen–Macaulay R-module of dimension n > 1. Let r be a parameter for M ,
and write M = M/Γm (M ). Show that

(i) (0 : Hm i
(M ))(0 : Hmi+1
(M )) ⊆ (0 : Hmi
(M /rM )) for all i = 1, . . . ,
n − 2;
(ii) (0 : Hm0 1
(M ))(0 : Hm (M/(0 :M r))) ⊆ (0 : Hm 0
(M/rM ));
(iii) (0 :M r) is m-torsion; and
(iv) (0 : Hm i i+1
(M ))(0 : Hm (M )) ⊆ (0 : Hm
i
(M/rM )) for all i = 0, 1, . . . ,
n − 2.

9.5.12 Exercise. Assume that (R, m) is local, and that M is a generalized


Cohen–Macaulay R-module of dimension n > 0. Let
 0
(n−1)  1
(n−1)  n−1
(n−1)
qM := 0 : Hm (M ) 0 0 : Hm (M ) 1 . . . 0 : Hm (M ) n−1 .

Prove that every system of parameters for M composed of elements in qM


is standard (see 9.5.9). (Here is a hint: use Exercises 9.5.11 and 9.5.8, together
with induction.)

9.5.13 Exercise. Assume (R, m) is local, and let M be a non-zero finitely


generated R-module of dimension n > 0. Assume that, for every system of
parameters (ri )ni=1 for M ,
k
mΓrk+1 R M/ i=1 ri M = 0 for all k ∈ N0 with k < n.

(i) Show that (0 :M r) = Γm (M ) for each parameter r for M .


j
(ii) Use part (i) and induction to prove that mHm (M ) = 0 for all j =
0, . . . , n − 1.

9.5.14 Definitions and Exercise. Assume that (R, m) is local, and let M be
a non-zero finitely generated R-module of dimension n > 0.
Let a1 , . . . , ah ∈ m. We say that a1 , . . . , ah is a weak M -sequence precisely
when
i−1 i−1
j=1 aj M :M ai = j=1 aj M :M m for all i = 1, . . . , h.

We say that M is a Buchsbaum R-module precisely when every sequence


r1 , . . . , rn forming a system of parameters for M is a weak M -sequence; we
say that R is a Buchsbaum ring if and only if it is a Buchsbaum module over
itself.
188 The Annihilator and Finiteness Theorems

Show that M is a Buchsbaum R-module if and only if, for every system of
parameters (ri )ni=1 for M ,
k
mΓrk+1 R M/ i=1 ri M = 0 for all k ∈ N0 with k < n.

There is an extensive theory of Buchsbaum rings and modules: see Stückrad


and Vogel [84]. The following final exercise in this section provides some al-
ternative characterizations of Buchsbaum modules.
9.5.15 Exercise. Assume (R, m) is local, and let M be a non-zero finitely
generated R-module of dimension n > 0.
Prove that the following conditions on M are equivalent:
(i) for every system of parameters (ri )ni=1 for M ,
k
mHm j
M/ i=1 ri M = 0 for all j, k ∈ N0 with j + k < n;

(ii) every system of parameters for M is standard (see 9.5.9);


(iii) M is a Buchsbaum R-module (see 9.5.14).
(It is perhaps appropriate to give some hints. For the implication ‘(ii) ⇒
(iii)’, note that m can be generated by parameters for M , and show that, if
r1 , . . . , rn form a system of parameters for M , then condition (ii) implies
k
that M/ i=1 ri M is a generalized Cohen–Macaulay R-module (see Exer-
cise 9.5.7) for all k = 0, . . . , n − 1. For the implication ‘(iii) ⇒ (i)’, use 9.5.14
and 9.5.13.)

9.6 Extensions
The remaining exercises in this chapter are directed at those readers who are
familiar with some fairly advanced concepts in commutative algebra, includ-
ing formal fibres of local rings, universally catenary rings, and L. J. Ratliff’s
Theorem [50, Theorem 31.7] that a local ring (R, m) is universally catenary if
and only if, for every p ∈ Spec(R) and for every minimal prime P of the ideal
pR of R,
 we have dim R/P = dim R/p.
So far in this chapter, we have established the main results, that is, the An-
nihilator and Finiteness Theorems, under the hypothesis that the underlying
ring is a homomorphic image of a regular ring. The remaining exercises in
this chapter, in conjunction with the Local-global Principle 9.6.2 for Finite-
ness Dimensions, show that the result of the Finiteness Theorem holds under
slightly weaker hypotheses, namely under the assumption that the underlying
ring is universally catenary and has the property that all the formal fibres of
9.6 Extensions 189

all its localizations are Cohen–Macaulay rings. We consider the Local-global


Principle, which is a consequence of the theorem of Faltings presented in 9.6.1
below, to be of considerable independent interest.

9.6.1 Theorem. (See G. Faltings [16, Satz 1].) Let M be a finitely generated
R-module, and let t ∈ N. Then the following statements are equivalent:

(i) Hai (M ) is finitely generated for all i < t;


i
(ii) HaR p
(Mp ) is a finitely generated Rp -module for all i < t and all p ∈
Spec(R).

Proof. (i) ⇒ (ii) Since 4.3.3 shows that HaR i


p
(Mp ) ∼= (Hai (M ))p for all
i ∈ N0 and all p ∈ Spec(R), this implication is clear.
(ii) ⇒ (i) We use induction on t. When t = 1, there is nothing to prove,
since Ha0 (M ) = Γa (M ) is a submodule of M , and so is finitely generated. So
suppose that t > 1 and that the result has been proved for smaller values of t.
By this assumption, Hai (M ) is finitely generated for i = 0, 1, . . . , t − 2, and it
only remains for us to prove that Hat−1 (M ) is finitely generated.
Set M := M/Γa (M ). Then, for each p ∈ Spec(R),

Mp ∼
= Mp /(Γa (M ))p = Mp /ΓaRp (Mp ),

and so it follows from 2.1.7(iii) that HaRi


p
(Mp ) ∼
= HaRi
p
(M p ) for all i ∈ N
and all p ∈ Spec(R). It follows that M also satisfies condition (ii) in the
statement of the theorem. Moreover, it would be sufficient for us to prove that
Hai (M ) is finitely generated for all i < t, since Hai (M ) ∼ = Hai (M ) for all
i ∈ N (by 2.1.7(iii)) and it is automatic that Ha (M ) is finitely generated. Now
0

M is an a-torsion-free R-module, by 2.1.2. Hence we can, and do, assume that


M is an a-torsion-free R-module.
We now use 2.1.1(ii) to deduce that a contains an element r which is a non-
zerodivisor on M . Let n ∈ N and let p ∈ Spec(R). The localization of the
rn
exact sequence 0 −→ M −→ M −→ M/rn M −→ 0 at p induces, for each
i ∈ N0 , an exact sequence
i−1
HaRp
(Mp ) −→ HaR
i−1
p
((M/rn M )p ) −→ HaR
i
p
(Mp ).

When i < t, the two outer modules in this last exact sequence are finitely
i−1
generated, and therefore so also is the middle one. Thus HaR p
((M/rn M )p ) is
a finitely generated Rp -module for all i < t and all p ∈ Spec(R). Therefore,
by the inductive hypothesis, Hai−1 (M/rn M ) is finitely generated for all i < t.
rn
Since the exact sequence 0 −→ M −→ M −→ M/rn M −→ 0 induces an
190 The Annihilator and Finiteness Theorems

exact sequence
rn
Hat−2 (M/rn M ) −→ Hat−1 (M ) −→ Hat−1 (M ),

it therefore follows that (0 :Hat−1 (M ) rn ) is finitely generated.


Set H := Hat−1 (M ), and Hn := (0 :H rn ) for all n ∈ N. Recall that our
aim is to prove that H is finitely generated; we have just proved that Hn is
finitely generated for all n ∈ N. Note that

H1 ⊆ H2 ⊆ · · · ⊆ Hn ⊆ Hn+1 ⊆ · · · ,

and that, since r ∈ a and H is an a-torsion R-module, H = n∈N Hn . We
shall achieve our aim by showing that there exists k ∈ N such that Hk = Hk+i
for all i ∈ N: it will then follow that H = Hk , and the latter module is finitely
generated.
For each n ∈ N, let Un denote the open subset Spec(R) \ Supp(Hn+1 /Hn )
of Spec(R). The next two stages in our proof aim to show that Spec(R) =

n∈N Un and

U1 ⊆ U2 ⊆ · · · ⊆ Ui ⊆ Ui+1 ⊆ · · · .

Let p ∈ Spec(R). Then Hp = (Hat−1 (M ))p = ∼ H t−1 (Mp ), and this is a


aRp
finitely generated, aRp -torsion, Rp -module. Therefore, there exists n ∈ N
such that Hp = (0 :Hp rn /1). Since

(0 :Hp rn /1) ⊆ (0 :Hp rn+1 /1) ⊆ Hp ,

it follows that (0 :Hp rn /1) = (0 :Hp rn+1 /1). But (0 :Hp ri /1) = (Hi )p
for all i ∈ N. Therefore (Hn+1 /Hn )p = 0, and so p ∈ Un . It follows that

Spec(R) = n∈N Un .
Next, let i ∈ N and p ∈ Ui . Then (0 :Hp ri /1) = (0 :Hp ri+1 /1), and so
 
(0 :Hp ri+2 /1) = (0 :Hp ri+1 /1) :Hp r/1
 
= (0 :Hp ri /1) :Hp r/1 = (0 :Hp ri+1 /1).

Thus (Hi+2 /Hi+1 )p = 0, and so p ∈ Ui+1 . Hence Ui ⊆ Ui+1 for all i ∈ N.


Thus the sets Ui (i ∈ N) form an ascending open cover of the quasi-compact
topological space Spec(R). Therefore there exists k ∈ N such that Uk =
Uk+i = Spec(R) for all i ∈ N. Thus

Supp(Hk+1 /Hk ) = Supp(Hk+i+1 /Hk+i ) = ∅ for all i ∈ N,

so that Hk = Hk+i for all i ∈ N. Thus H = Hk , and so is finitely generated.


This completes the inductive step.
9.6 Extensions 191

9.6.2 Local-global Principle for Finiteness Dimensions. Let M be a finitely


generated R-module. Then
fa (M ) = inf faRp (Mp ) : p ∈ Spec(R) .
Proof. This is now immediate from Theorem 9.6.1 above and the definition
of finiteness dimension in 9.1.3.
9.6.3 Exercise. Assume that (R, m) is local. Let b be a second ideal of R,
and let M be a finitely generated R-module. Prove that
 
 (M ⊗R R) = fa (M ).
fabRR b

9.6.4 Exercise. Assume that (R, m) is local, that M is a finitely generated


R-module, and that p ∈ Spec(R) \ Var(b), where b is a second ideal of R. Let
h := ht(a + p)/p.
(i) Prove that ht(aR + pR)/p
 R  = h.

(ii) Let Q ∈ Spec(R) be a minimal prime of aR  + pR with ht Q/pR
 = h,
 such that P ⊆ Q and ht Q/P = h.
and let P be a minimal prime of pR
Show that
(a) ht(aR  + P)/P ≤ h,
(b) P ∩ R = p and P ∈ Var(bR), 

(c) depth(M ⊗R R)P = depth Mp , and
(d) adj  depth(M ⊗R R)  P ≤ adja depth Mp .
aR
 
 (M ⊗R R) ≤ λa (M ).
R
(iii) Deduce that λbaR b

9.6.5 Exercise. Let the situation be as in Exercise 9.6.4 above, and assume in
addition that R is universally catenary and that all its formal fibres are Cohen–
Macaulay rings.
 \ Var(bR)
(i) Let Q ∈ Spec(R)  and set q := Q ∩ R. Show that q ∈
Spec(R) \ Var(b) and that
 Q .
adja depth Mq ≤ adjaR depth(M ⊗R R)
(You might find Ratliff’s Theorem [50, Theorem 31.7] helpful.)
(ii) Deduce from part (i) and Exercise 9.6.4(iii) above that
 
 (M ⊗R R) = λa (M ).
R
λbaR b

9.6.6 Exercise. Assume that (R, m) is a universally catenary local ring all
of whose formal fibres are Cohen–Macaulay rings. Prove that the conclusion
of the Annihilator Theorem holds over R. In other words, prove that, if M is
a finitely generated R-module and b is a second ideal of R, then λba (M ) =
fab (M ).
192 The Annihilator and Finiteness Theorems

9.6.7 Exercise. Assume that R is universally catenary and all the formal
fibres of all its localizations are Cohen–Macaulay rings. Deduce from Exercise
9.6.6 above and the Local-global Principle 9.6.2 that the conclusion of the
Finiteness Theorem holds over R. In other words, prove that, if M is a finitely
generated R-module, then λaa (M ) = fa (M ).
9.6.8 Exercise. Assume that (R, m) is a universally catenary local ring all of
whose formal fibres are Cohen–Macaulay rings. Let M be a non-zero finitely
generated R-module of dimension n > 0 such that
(a) dim R/p = n whenever p is a minimal member of Supp M , and
(b) Mq is a Cohen–Macaulay Rq -module for all q ∈ Supp M \ {m}.
Prove that M is a generalized Cohen–Macaulay R-module (see 9.5.7).
9.6.9 Exercise. (This exercise is related to Exercise 9.4.13.) Assume that R
is a universally catenary semi-local integral domain with the property that all
the formal fibres of all its localizations are Cohen–Macaulay rings. Assume
that there exists 0 = c ∈ R such that the ring Rc is Cohen–Macaulay. Prove
that there exists n ∈ N such that, for all choices of r ∈ N0 and all choices of
an ideal B of the polynomial ring R[X1 , . . . , Xr ], we have
c n HB
i
 (R[X1 , . . . , Xr ]) = 0 for all i < ht B .
10

Matlis duality

Prior to this point in the book, we have not made use of the decomposition
theory (due to E. Matlis [49]) for injective modules over our (Noetherian) ring
R. However, our work in the next Chapter 11 on local duality will involve use
of the structure of the terms in the minimal injective resolution of a Gorenstein
local ring, and so we can postpone no longer use of the decomposition theory
for injective modules. Our discussion of local duality in Chapter 11 will also
involve Matlis duality.
Our purpose in this chapter is to prepare the ground for Chapter 11 by re-
viewing, sometimes in detail, those parts of Matlis’s theories that we shall need
later in the book. Sometimes we simply refer to [50, Section 18] for proofs; in
other cases, we provide alternative proofs for the reader’s convenience. An
experienced reader who is familiar with this work of Matlis should omit this
chapter and progress straight to the discussion of local duality in Chapter 11:
for one thing, there is no local cohomology theory in this chapter! However,
graduate students might find this chapter helpful.

10.1 Indecomposable injective modules


10.1.1 Reminders. Let M be a submodule of the R-module L.

(i) We say that L is an essential extension of M precisely when B ∩ M = 0


for every non-zero submodule B of L.
(ii) We say that L is an injective envelope (or injective hull) of M precisely
when L is an injective R-module that is an essential extension of M .
(iii) (See [50, Theorem B4, p. 281].) The R-module M is injective if and
only if the only essential extension of M is M itself.
194 Matlis duality

(iv) If L is an injective envelope of M , and g : M −→ K is an R-monomor-


phism from M into an injective R-module K, then there is a homomor-
phism g  : L → K such that the diagram

0 -M -L

g
g
?
K

commutes. Since Ker g  ∩ M = Ker g = 0, it follows from the fact that


L is an essential extension of M that Ker g  = 0 and g  is a monomor-
phism. Hence L is isomorphic to a direct summand of K.
(v) (See [50, p. 281].) Each R-module has an injective envelope which is
uniquely determined up to isomorphism. In fact, if E and E  are both
injective envelopes of M (so that each of E and E  is an injective R-
module which contains M as a submodule, and each of E and E  is an
essential extension of its submodule M ), then there is an isomorphism
f : E → E  such that the diagram

M -E

@ ∼
@ = f
R ?
@
E

(in which the unnamed homomorphisms are the inclusion maps) com-
mutes.
(vi) We denote by E(M ) (or ER (M ) if it is necessary to specify the under-
lying ring) one choice of injective envelope of M .

10.1.2 Example and Warning. The reader should be warned that the iso-
morphism f : E → E  of 10.1.1(v) need not be uniquely determined; for this
reason, we cannot regard E( • ) as a functor from C(R) to itself.
To illustrate this point, consider the submodule
 r 
C2 (∞) := α ∈ Q/Z : α = n + Z for some r ∈ Z and n ∈ N0
2
of the Z-module Q/Z. Since C2 (∞) is a divisible Abelian group, it is an injec-
tive Z-module, by [71, Theorem 3.24] for example. Also C2 (∞) is an essential
extension of its Z-submodule C2 := Z( 12 + Z). Now ‘f = IdC2 (∞) ’ and ‘f =
10.1 Indecomposable injective modules 195

− IdC2 (∞) ’ are two different choices of automorphism f : C2 (∞) → C2 (∞)


for which the diagram

C2 - C2 (∞)

@
@ ∼
= f
@
@ ?
R
C2 (∞)

(in which the unnamed homomorphisms are the inclusion maps) commutes.

Injective envelopes play an essential rôle in the very satisfactory decompo-


sition theory for injective R-modules: for one thing, the fundamental ‘building
blocks’ on which the whole theory is based are described using the concept of
injective envelope.

10.1.3 Reminders. We shall assume that the reader is familiar with the fol-
lowing facts about injective R-modules.

(i) Let (Mι )ι∈Λ be a non-empty family of R-modules. It is immediate from



the definition of injective module that the direct product ι∈Λ Mι is an
injective R-module if and only if Mι is injective for all ι ∈ Λ. However,
since our ring R is Noetherian, it is also the case that the direct sum

ι∈Λ Mι is injective if and only if Mι is injective for all ι ∈ Λ: see [50,
Theorem 18.5(i)].
(ii) An R-module is said to be indecomposable precisely when it is non-zero
and cannot be written as the direct sum of two proper submodules. For
each p ∈ Spec(R), the injective R-module E(R/p) is indecomposable
(see [50, Theorem 18.4(i)]); moreover, each indecomposable injective
R-module is isomorphic to E(R/q) for some q ∈ Spec(R) (see [50,
Theorem 18.4(ii)]).
(iii) Let p ∈ Spec(R) and let r ∈ R \ p. Then multiplication by r provides an
automorphism of E(R/p) (see [50, Theorem 18.4(iii)]); moreover, each
element of E(R/p) is annihilated by some power of p, that is, E(R/p)
is p-torsion (see [50, Theorem 18.4(v)]).
(iv) Let p, q ∈ Spec(R). Then E(R/p) ∼ = E(R/q) if and only if p = q (see
[50, Theorem 18.4(iv)]).
(v) The results recounted in parts (ii) and (iv) above can be reformulated as
follows: there is a set of isomorphism classes of indecomposable injec-
tive R-modules, and there is a bijective correspondence between this set
196 Matlis duality

and Spec(R), under which a p ∈ Spec(R) corresponds to the isomor-


phism class of E(R/p).
10.1.4 Exercise. Let M1 , . . . , Mn be R-modules. Show that the map
M1 ⊕ · · · ⊕ Mn −→ E(M1 ) ⊕ · · · ⊕ E(Mn )
(obtained by taking the direct sum of the inclusion maps) provides an injective
n
envelope of i=1 Mi .
10.1.5 Exercise. Let M be a non-zero R-module. Show that the following
statements are equivalent:
(i) E(M ) is indecomposable;
(ii) E(M ) is an injective envelope of every non-zero submodule of itself;
(iii) the zero submodule of M cannot be expressed as the intersection of two
non-zero submodules of M .
10.1.6 Exercise. Let a be a proper ideal of R. Show that E(R/a) is inde-
composable if and only if a is irreducible.
Let q be a p-primary ideal of R. Prove that E(R/q) is isomorphic to a direct
sum of finitely many copies of E(R/p). What can you say about the number
of copies? (Here is a hint: recall that q can be expressed as an intersection
n
q = i=1 ji , where each ji (for 1 ≤ i ≤ n) is irreducible and irredundant in
the intersection. If you still find this exercise difficult, you might like to consult
[81, Exercise 8.30].)
10.1.7 Exercise. Let p ∈ Spec(R).
(i) Let 0 = x ∈ E(R/p). Show that (0 : x) is an irreducible ideal of R.

(ii) In 10.1.3(iii), we saw that E(R/p) = n∈N (0 :E(R/p) pn ). Show that

E(R/p) = n∈N (0 :E(R/p) p(n) ).
10.1.8 Exercise. Let I be a non-zero injective R-module. Show that I is a
direct sum of indecomposable injective submodules, perhaps by means of the
following intermediate steps.
(i) Apply Zorn’s Lemma to the set of all sets of indecomposable injective
submodules of I whose sum is direct, in order to find a maximal member
M of this set.

(ii) Let J := D∈M D. Suppose that J ⊂ I and seek a contradiction. Use
the injectivity of J to find a submodule K of I such that I = J ⊕ K.
(iii) Let q ∈ Ass K. Show that K has a submodule isomorphic to ER (R/q).
(iv) Use 10.1.3(ii) to find a contradiction to the maximality of M, and deduce
that J = I.
10.1 Indecomposable injective modules 197

In 10.1.3, we reviewed fundamental facts about indecomposable injective R-


modules. One of the reasons for the importance of these modules is provided
by the following.

10.1.9 Reminder. (See [50, Theorem 18.5(ii)].) By 10.1.8, each injective R-


module I is a direct sum of indecomposable injective submodules. Therefore,
by 10.1.3(ii), there is a family (pα )α∈Λ of prime ideals of R for which I ∼
=

α∈Λ E(R/p α ).

10.1.10 Remark. Let the situation be as in 10.1.9. Then p is equal to pα for


some α ∈ Λ if and only if p ∈ Ass I.

10.1.11 Exercise. Let I be an injective R-module, so that, by 10.1.9, there is



a family (pα )α∈Λ of prime ideals of R for which I ∼
= α∈Λ E(R/pα ). Prove
that

Γa (I) ∼
= E(R/pα ).
α∈Λ
a⊆pα

(This exercise provides another proof of the result of Proposition 2.1.4 that
Γa (I) is an injective R-module.)

In fact, the direct decompositions described in 10.1.9 have uniqueness prop-


erties. The route taken by Matsumura in [50, §18] to these uniqueness proper-
ties involves facts about the behaviour of indecomposable injective R-modules
under localization. We review this behaviour next.

10.1.12 Lemma. Let S be a multiplicatively closed subset of R, and let G be


an S −1 R-module. Then G is R-injective (that is, injective when viewed as an
R-module by means of the natural homomorphism R → S −1 R) if and only if
G is S −1 R-injective.

Proof. (⇒) Let H be an S −1 R-submodule of the S −1 R-module J, and let


h : H → G be an S −1 R-homomorphism. Then H is an R-submodule of
the R-module J, and h is an R-homomorphism. Since G is R-injective, there
exists an R-homomorphism j : J → G which extends h. It is easy to check
that j must be an S −1 R-homomorphism.
(⇐) Let M be an R-submodule of the R-module N , and let λ : M → G
be an R-homomorphism. As G is an S −1 R-module, the natural map ψ : G →
S −1 G is not only an R-isomorphism but also an S −1 R-isomorphism. Thus
S −1 M is an S −1 R-submodule of the S −1 R-module S −1 N , and the composi-
tion ψ −1 ◦ S −1 λ : S −1 M → G is an S −1 R-homomorphism.
Since G is S −1 R-injective, there is an S −1 R-homomorphism μ : S −1 N →
198 Matlis duality

G which extends ψ −1 ◦ S −1 λ. Then μ ◦ θ : N → G, where θ : N → S −1 N


is the natural map, is an R-homomorphism which extends λ.
10.1.13 Lemma. Let S be a multiplicatively closed subset of R, and let p ∈
Spec(R) be such that p ∩ S = ∅. By 10.1.3(iii), the indecomposable injective
R-module ER (R/p) has a natural structure as an S −1 R-module.
As S −1 R-module, ER (R/p) is isomorphic to ES −1 R (S −1 R/S −1 p).
Furthermore, ES −1 R (S −1 R/S −1 p), when considered as an R-module by
means of the natural homomorphism R → S −1 R, is isomorphic to ER (R/p).
Proof. It follows from Lemma 10.1.12 that the S −1 R-module ER (R/p) is
S −1 R-injective. Since an S −1 R-submodule of ER (R/p) is automatically an
R-submodule, it is immediate from 10.1.3(ii) that ER (R/p) is indecomposable
as S −1 R-module; we can therefore deduce from the same result that ER (R/p)
is S −1 R-isomorphic to ES −1 R (S −1 R/P) for some P ∈ Spec(S −1 R). By
10.1.3(iii),
P = α ∈ S −1 R : α IdER (R/p) is not an isomorphism .
Hence the contraction of P to R is just
r ∈ R : r IdER (R/p) is not an isomorphism = p.
Therefore P = S −1 p.
The final claim is now immediate.
10.1.14 Proposition. Let S be a multiplicatively closed subset of R.
(i) Let p ∈ Spec(R). Then

=0 if p ∩ S = ∅,
S −1 (ER (R/p))

= ES −1 R (S −1 R/S −1 p) if p ∩ S = ∅.

(ii) Let I be an injective R-module. Then S −1 I is both S −1 R-injective and


R-injective.
Proof. (i) Suppose that s ∈ p ∩ S. By 10.1.3(iii), each x ∈ E(R/p) is anni-
hilated by some power of s, and so S −1 (ER (R/p)) = 0.
If, on the other hand, p ∩ S = ∅, then, by 10.1.3(iii) again, E(R/p) has a
natural structure as an S −1 R-module, so that S −1 (E(R/p)) ∼ = E(R/p) both
−1
as R-modules and S R-modules. We can now use 10.1.13 to complete the
proof of part (i).
(ii) By 10.1.9, there is a family (pα )α∈Λ of prime ideals of R for which I ∼
=
 −1 ∼
 −1 −1
α∈Λ E(R/p α ). Then S I = α∈Λ S (E(R/p α )) (as S R-modules);
since, by (i), each S −1 (E(R/pα )) in the above display is S −1 R-injective, it
10.2 Matlis duality 199

follows from 10.1.3(i) that S −1 I is S −1 R-injective. An appeal to 10.1.12 now


completes the proof.
The behaviour of injective R-modules under fraction formation is used in
Matsumura’s proof of the following fundamental uniqueness property of the
direct sum decompositions described in 10.1.9.
10.1.15 Reminder. (See [50, Theorem 18.5(iii)].) Let I be an injective R-
module. By 10.1.9, there is a family (pα )α∈Λ of prime ideals of R for which

I∼= α∈Λ E(R/pα ).
Let p ∈ Spec(R), and let k(p) = Rp /pRp , the residue field of the local ring
Rp . Then the cardinality of the set {α ∈ Λ : pα = p} depends only on I and
p and not on the particular decomposition of I (as a direct sum of indecom-
posable injective submodules) chosen; in fact, this cardinality is equal to the
vector space dimension dimk(p) HomRp (k(p), Ip ).
10.1.16 Lemma. Suppose that the R-module M is annihilated by the ideal
b of R. We can regard M and (0 :ER (M ) b) as modules over R/b in natural
ways: when this is done, (0 :ER (M ) b) ∼
= ER/b (M ).
Proof. Let ι : (0 :ER (M ) b) → ER (M ) denote the inclusion map. Let
α
0 - G - H

?
(0 :ER (M ) b)

be a diagram of R/b-modules and R/b-homomorphisms in which the row


is exact. We can regard these modules as R-modules by means of the natural
homomorphism R → R/b, and then α, γ and ι◦γ become R-homomorphisms.
Since ER (M ) is an injective R-module, there exists an R-homomorphism β :
H −→ ER (M ) such that β ◦ α = ι ◦ γ. However, since bH = 0, it follows
that Im β ⊆ (0 :ER (M ) b). We can regard β as a map from H to (0 :ER (M ) b),
and, when we do that, β is an R/b-homomorphism such that β ◦ α = γ.
Thus (0 :ER (M ) b) is injective as an R/b-module. Obviously (0 :ER (M ) b)
is an essential extension of its R/b-submodule M .

10.2 Matlis duality


Although there is a treatment of Matlis duality in [50, Theorem 18.6], we
are going to present, for the reader’s convenience, a different approach to this
200 Matlis duality

theory. Our approach makes use of Melkersson’s Theorem [51, Theorem 1.3]:
we presented this theorem in 7.1.2, and applied it to local cohomology modules
in Chapter 7.
We begin by specifying some notation which we shall frequently use during
our discussion of Matlis duality.

10.2.1 Notation and Remarks. Suppose (R, m) is local. Set E = E(R/m),



the injective envelope of the simple R-module R/m. As usual, we shall use R
n
to denote the m-adic completion lim R/m of R. We shall use D to denote the
←−
n∈N
exact, contravariant, R-linear functor HomR ( • , E) from C(R) to itself. For
each R-module G, we shall refer to D(G) as the Matlis dual of G. Note that
D(R) is naturally isomorphic to E, and that D(E) = HomR (E, E) is just the
R-endomorphism ring of E considered as an R-module in the natural way.
For each R-module G, let

μG : G −→ DD(G) = HomR (HomR (G, E), E)

be the natural R-homomorphism for which (μG (x))(f ) = f (x) for all x ∈ G
and f ∈ HomR (G, E).
Note that, as G varies through C(R), the μG constitute a natural transforma-
tion μ from the identity functor to the functor DD.
If an R-module M has finite length, then we shall denote that length by
(M ).

10.2.2 Remarks. Suppose that (R, m) is local, and use the notation of 10.2.1.
Let G be an R-module.

(i) The R-homomorphism μG is injective, since if 0 = x ∈ G, there is an


R-homomorphism f  : Rx → R/m for which f  (rx) = r + m for all
r ∈ R, and the composition of f  and the inclusion map R/m → E can
be extended to an f ∈ HomR (G, E) for which f (x) = 0.
(ii) The annihilators of G and its Matlis dual D(G) are equal, because the
fact that D is an R-linear functor ensures that

(0 : G) ⊆ (0 : D(G)) ⊆ (0 : DD(G)),

while the injectivity of μG : G −→ DD(G) (proved in (i) above) en-


sures that (0 : DD(G)) ⊆ (0 : G).

First we analyse the case when (R, m) is an Artinian local ring. Recall that,
then, an R-module G is Artinian if and only if it is Noetherian, and this is the
case if and only if G is finitely generated.
10.2 Matlis duality 201

10.2.3 Proposition. Suppose that (R, m) is local and Artinian, and use the
notation of 10.2.1. Then

(i) D(R/m) ∼ = R/m (as R-modules);


(ii) for each finitely generated R-module G (that is, for each R-module G
of finite length), the Matlis dual D(G) is also finitely generated and
(D(G)) = (G);
(iii) E is finitely generated (and so Artinian), and (E) = (R);
(iv) for each finitely generated R-module G (that is, for each Artinian R-
module G), the map μG : G −→ DD(G) is an isomorphism; and
(v) for each f ∈ HomR (E, E), there is a unique rf ∈ R such that f (x) =
rf x for all x ∈ E.

Note. Condition (v) is equivalent to the statement that the homomorphism


θ : R −→ HomR (E, E) for which θ(r) = r IdE , for all r ∈ R, is an isomor-
phism.

Proof. (i) Let k denote the residue field of R. There is an isomorphism of


k-modules
D(R/m) = HomR (R/m, E) ∼ = (0 :E m).
But (0 :E m) ∼ = Ek (k) by 10.1.16; since every k-module, that is, every vector
space over k, is injective, Ek (k) = k. The claim follows from this.
(ii) Use induction on length: remember that D is an exact functor.
(iii) This is now immediate from (ii), since E ∼
= D(R).
(iv) Let G be a finitely generated R-module. Two uses of part (ii) show that
DD(G) is also finitely generated and has (DD(G)) = (G); since μG is
injective by 10.2.2(i), it follows that μG must be an isomorphism.
(v) By part (iv), the map μR : R → HomR (HomR (R, E), E) is an isomor-
phism. The claim therefore follows from the fact that the composition
μR ∼
=
θ : R −→ HomR (HomR (R, E), E) −→ HomR (E, E),

in which the second isomorphism is the obvious natural one, is such that θ(r),
for r ∈ R, is the endomorphism of E given by multiplication by r.

10.2.4 Exercise. Suppose that (R, m) is local and Artinian, and set E :=
E(R/m). Let L be a submodule of E. Show that L is faithful if and only if
L = E. (Here is a hint: apply the functor HomR ( • , E) to the canonical exact
sequence 0 −→ L −→ E −→ E/L −→ 0.)

Proposition 10.2.3 provides the main ingredients of Matlis duality for the
(very special!) case of an Artinian local ring. However, as we shall show below,
202 Matlis duality

this special case can be used to build up to the general case quickly. We proved
in 10.2.3(iii) that, if (R, m) is local and Artinian, then E(R/m) is Artinian.
We now use Melkersson’s Theorem [51, Theorem 1.3] (see 7.1.2) to obtain a
much more general statement.

10.2.5 Theorem. Let m be a maximal ideal of R. Then E(R/m) is an Ar-


tinian injective R-module.

Proof. By 10.1.16, we have (0 :ER (R/m) m) ∼ = ER/m (R/m). Since R/m


is a field, ER/m (R/m) = R/m. Hence (0 :ER (R/m) m) is an Artinian R-
module. Now ER (R/m) is m-torsion, by 10.1.3(iii), and so it follows from
Melkersson’s Theorem 7.1.2 that ER (R/m) is Artinian.

10.2.6 Definition. Let M be an R-module. The socle Soc(M ) of M is de-


fined to be the sum of all simple submodules of M .

10.2.7 Remark. Let A be a non-zero Artinian R-module. Then A is an essen-


tial extension of its own socle, because each non-zero submodule B of A must
contain a simple submodule (since a minimal member of the set of non-zero
submodules of B must be simple).

10.2.8 Corollary. Suppose that (R, m) is local, and set E := E(R/m). Let
M be an R-module. Then M is Artinian if and only if M is isomorphic to a
submodule of E t , the direct sum of t copies of E, for some t ∈ N.

Proof. (⇐) It is clear from 10.2.5 that, for every t ∈ N, every submodule of
E t is Artinian.
(⇒) Assume that M is Artinian. Clearly, we can assume that M = 0. Now
M is an essential extension of Soc(M ), by 10.2.7, and Soc(M ) is an Artinian
module annihilated by m. Thus Soc(M ) has a natural structure as a (finite-
dimensional) vector space over R/m, and so there exists t ∈ N such that
Soc(M ) ∼ = (R/m)t . Now compose this isomorphism with the direct sum of
the inclusion maps to obtain an R-monomorphism f : Soc(M ) → E t ; since
E t is injective (by 10.1.3(i)), this f can be extended to an R-homomorphism
f  : M → E t ; and f  must also be a monomorphism since M is an essential
extension of Soc(M ), by 10.2.7.

Key to the theory of Matlis duality are the facts that, when (R, m) is local,

E := E(R/m) has a natural structure as an R-module, and, for each f ∈
HomR (E, E), there is a unique rf ∈ R  such that f (x) = rf x for all x ∈ E.
We approach these facts next, and, in doing so, we touch on ideas mentioned
in Exercise 8.2.4.
10.2 Matlis duality 203

10.2.9 Remark. Suppose that (R, m) is local and that A is an Artinian R-


module.
Let a ∈ A. It is immediate from 10.2.8 and 10.1.3(iii) that there exists t ∈ N
such that mt a = 0. (In fact, one can prove this in a much more elementary
manner: if a = 0, then, since the R-module Ra is Artinian, R/(0 : a) is an
Artinian local ring, and so its maximal ideal is nilpotent.) Let, with an obvious
notation,

r = (rn + mn )n∈N = (rn + mn )n∈N ∈ lim R/mn = R,
←−
n∈N

so that rn − rn ∈ mn for all n ∈ N and rn+h − rn ∈ mn for all n, h ∈ N.


Then rt+h a = rt a = rt a = rt+h

a for all h ∈ N. It is straightforward to check

that A can be given the structure of an R-module such that, with the above

notation, ra = rt a. Note that, if we regard this R-module as an R-module by
means of the natural ring homomorphism R → R,  then we recover the original
R-module structure on A; note also that a subset of A is an R-submodule if

and only if it is an R-submodule.
10.2.10 Exercise. Suppose that (R, m) is local, and set E := E(R/m).
By 10.2.5, the R-module E is Artinian; therefore, by 10.2.9, it has a natural

structure as an R-module. 
Prove that there is an R-isomorphism E(R/m) ∼ =
 m),
E  (R/  where m 
 denotes the maximal ideal of R.
R

We are now able to prove the key result for Matlis duality that we mentioned
just before 10.2.9.
10.2.11 Theorem. Suppose that (R, m) is local, and set E := E(R/m).
By 10.2.5, the R-module E is Artinian; therefore, by 10.2.9, it has a natural

structure as an R-module.
The natural R-homomorphism θ : R  −→ HomR (E, E) for which θ(r) =

r IdE for all r ∈ R is an isomorphism. Thus, for each f ∈ HomR (E, E), there
is a unique rf ∈ R  such that f (x) = rf x for all x ∈ E.

Proof. For each n ∈ N, set En := (0 :E mn ). Let f ∈ HomR (E, E); let t ∈


N. Of course, R/m is annihilated by mt . By 10.1.16, there is an isomorphism
of R/mt -modules Et ∼ = ER/mt ((R/mt )/(m/mt )). Now f (Et ) ⊆ Et , and,
of course, R/mt is an Artinian local ring. Therefore, by 10.2.3(v), there exists
rt ∈ R such that f (e) = rt e for all e ∈ Et , and, moreover, if rt is any other
element of R such that f (e) = rt e for all e ∈ Et , then the uniqueness aspect
of 10.2.3(v) ensures that rt + mt = rt + mt , that is, rt − rt ∈ mt .
We can proceed as above for each t ∈ N, and so construct a uniquely deter-

mined sequence (rn +mn )n∈N ∈ n∈N R/mn with the property that, for every
204 Matlis duality

n ∈ N, we have f (e) = rn e for all e ∈ En . Furthermore, for n, h ∈ N, we


have En ⊆ En+h , and so, since f (e) = rn e = rn+h e for all e ∈ En , it follows
from the immediately preceding paragraph in this proof that rn − rn+h ∈ mn .
We have therefore found a uniquely determined sequence

(rn + mn )n∈N ∈ lim R/mn = R
←−
n∈N

such that, for every n ∈ N, we have f (e) = rn e for all e ∈ En . Since E =


  such that
∈ R
n∈N En by 10.1.3(iii), it follows that there is exactly one r
f (x) = rx for all x ∈ E.

We are now able to present the main aspects of Matlis duality over a com-
plete local ring.

10.2.12 Matlis Duality Theorem. Suppose that (R, m) is local and com-
plete, and use the notation of 10.2.1. (Thus E denotes the injective envelope
E(R/m) of the simple R-module and D := HomR ( • , E).)

(i) For each f ∈ HomR (E, E), there is a unique rf ∈ R such that f (x) =
rf x for all x ∈ E.
(ii) Whenever N is a finitely generated R-module, the natural homomor-
phism μN : N −→ DD(N ) is an isomorphism and D(N ) is Artinian.
(iii) Whenever A is an Artinian R-module, the natural homomorphism μA :
A −→ DD(A) is an isomorphism and D(A) is Noetherian.

Proof. (i) This is immediate from Theorem 10.2.11.


(ii) The composition
μR ∼
=
θ : R −→ HomR (HomR (R, E), E) −→ HomR (E, E),

in which the second map is the obvious natural isomorphism, is such that
θ(r) = r IdE for all r ∈ R. We have seen in part (i) that θ is an isomorphism;
therefore μR is an isomorphism.
The identity functor and DD are both additive, and the result of application
of an additive functor to a split short exact sequence is again a split short exact
sequence; also, μ is a natural transformation of functors. We can therefore
deduce, by induction on rank, that μF is an isomorphism whenever F is a
finitely generated free R-module.
Let N be an arbitrary finitely generated R-module. Then N can be included
in an exact sequence F1 −→ F0 −→ N −→ 0 in which F1 and F0 are finitely
generated free R-modules. Since the functor DD is additive and exact, and
10.2 Matlis duality 205

μ is a natural transformation of functors, the above exact sequence induces a


commutative diagram

F1 - F0 - N - 0 - 0


= μF1 ∼
= μF 0 μN

? ? ?
DD(F1 ) -DD(F0 ) - DD(N ) - 0 - 0

with exact rows. It therefore follows from the Five Lemma that μN is an iso-
morphism.
Finally, application of the contravariant, exact functor D to the exact se-
quence F0 −→ N −→ 0 shows that D(N ) is isomorphic to a submodule of
D(F0 ); since D(F0 ) is isomorphic to a direct sum of finitely many copies of
D(R) and D(R) ∼ = E, it follows from 10.2.8 that D(N ) is Artinian.
(iii) The composition
μE HomR (θ,IdE ) ∼
E - HomR (HomR (E, E), E) ∼
- HomR (R, E)
=
- E,
=

where θ is the isomorphism used in the proof of part (ii) above, is just the
identity map; therefore μE is an isomorphism. We can now use the additivity
of DD, and the natural transformation μ, to deduce, by induction on t, that
μE t is an isomorphism for all t ∈ N.
Let A be an arbitrary Artinian R-module. Two uses of Corollary 10.2.8 show
that there is an exact sequence 0 −→ A −→ E n0 −→ E n1 for suitable positive
integers n0 and n1 . We can now use the exactness of the functor DD, together
with the natural transformation μ, as we did in the above proof of part (ii), to
obtain from this exact sequence a commutative diagram with exact rows, and
another application of the Five Lemma will yield the desired conclusion that
μA is an isomorphism.
Finally, application of the contravariant, exact functor D to the exact se-
quence 0 −→ A −→ E n0 shows that D(A) is a homomorphic image of
D(E n0 ) ∼ = (D(E))n0 ; since θ : R → D(E) is an isomorphism, D(A) is
a homomorphic image of a finitely generated free R-module, and so is
Noetherian.

10.2.13 Exercise. Suppose that (R, m) is local and complete, and use the
notation of 10.2.1. Let G be an R-module of finite length, so that, by the Matlis
Duality Theorem 10.2.12, the Matlis dual D(G) also has finite length. Prove
that (D(G)) = (G).
206 Matlis duality

10.2.14 Remark. Suppose that (R, m) is local and complete, and use the
notation of 10.2.1.
The Matlis Duality Theorem 10.2.12 allows statements about Noetherian
R-modules to be translated into ‘dual’ statements about Artinian R-modules,
and vice versa. For example, it shows that every Noetherian R-module is iso-
morphic to the Matlis dual of an Artinian R-module, and that every Artinian
R-module is isomorphic to the Matlis dual of a Noetherian R-module.
To give a sample of this type of ‘translation’, let N be a Noetherian
R-module, so that D(N ) is an Artinian R-module: we can use the Matlis
Duality Theorem 10.2.12 to show quickly that Att D(N ) = Ass N , as fol-
lows. Let p ∈ Spec(R). Then p ∈ Ass N if and only if N has a submod-
ule with annihilator equal to p. Now D is an exact, contravariant functor and
DD(N ) ∼ = N ; also, by 10.2.2(ii), the annihilators of an R-module M and its
Matlis dual D(M ) are equal. Therefore N has a submodule with annihilator
equal to p if and only if D(N ) has a homomorphic image with annihilator
equal to p; and, by 7.2.5, this is the case if and only if p ∈ Att D(N ).

10.2.15 Exercise. Suppose that (R, m) is local and complete, and use the
notation of 10.2.1.
Let N be a Noetherian R-module and let A be an Artinian R-module; let
n ∈ N0 and h ∈ N.

(i) Prove that D(an N/an+h N ) ∼


= (0 :D(N ) an+h )/(0 :D(N ) an ).
 
(ii) Prove that D (0 :A an+h )/(0 :A an ) ∼= an D(A)/an+h D(A).
 
(iii) Prove that Att(0 :A a ) = Ass D(A)/ah D(A) .
h

So far, we have restricted our account of Matlis duality to situations where


the underlying local ring is complete. However, it is possible to obtain from
Theorem 10.2.11 a satisfactory partial result which is valid over any local ring.
We include this because we shall find a corollary of it useful in our applica-
tions of the local duality theory developed in Chapter 11. We shall need the
following technical lemma.

10.2.16 Lemma. Let M, I, J be R-modules.

(i) There exists a (unique) R-homomorphism

ξM,I,J : M ⊗R HomR (I, J) −→ HomR (HomR (M, I), J)

such that, for m ∈ M , f ∈ HomR (I, J) and g ∈ HomR (M, I), we


have (ξM,I,J (m ⊗ f )) (g) = f (g(m)). Furthermore, as M, I, J vary
10.2 Matlis duality 207

through the category C(R), the ξM,I,J constitute a natural transforma-


tion of functors

ξ • , • , • : ( • ) ⊗R HomR ( • , • ) −→ HomR (HomR ( • , • ), • )

(from C(R) × C(R) × C(R) to C(R)).


(ii) If J is injective, then ξM,I,J is an isomorphism whenever M is finitely
generated.

Proof. (i) This is straightforward and left to the reader.


(ii) It will be convenient to use V to denote ( • ) ⊗R HomR (I, J), to use W
to denote the functor HomR (HomR ( • , I), J), and to use ζ • to denote ξ • ,I,J .

= ζR ∼
=
The composition HomR (I, J) −→ V (R) −→ W (R) −→ HomR (I, J), in
which the first and last isomorphisms are the obvious natural ones, is just the
identity map; hence ζR = ξR,I,J is an isomorphism.
The functors V and W are both additive, and ζ is a natural transformation
of functors. We can therefore deduce, by induction on rank, that ζF is an iso-
morphism whenever F is a finitely generated free R-module.
Let M be an arbitrary finitely generated R-module. Then M can be included
in an exact sequence F1 −→ F0 −→ M −→ 0 in which F1 and F0 are finitely
generated free R-modules. Of course the functor V is right exact; since J
is injective, the functor W is also right exact; therefore, since ζ is a natural
transformation of functors, the above exact sequence induces a commutative
diagram

V (F1 ) - V (F0 ) - V (M ) - 0 - 0


= ζF1 ∼
= ζF0 ζM

? ? ?
W (F1 ) - W (F0 ) - W (M ) - 0 - 0

with exact rows. It thus follows from the Five Lemma that ζM is an isomor-
phism.

10.2.17 Exercise. Supply a proof for part (i) of Theorem 10.2.16.

10.2.18 Remarks. Suppose that (R, m) is local (but not necessarily com-
plete), and use the notation of 10.2.1. (Thus E denotes the injective envelope
E(R/m) of the simple R-module and D := HomR ( • , E).) Let m  denote the
maximal ideal of R.
208 Matlis duality

(i) By 10.2.9, an Artinian R-module A can be given a natural structure as



an R-module such that
 there exists r ∈ R with ra = ra,
(a) for each a ∈ A and each r ∈ R,
and

(b) if we regard this R-module as an R-module by means of the nat-
ural ring homomorphism R → R,  then we recover the original
R-module structure on A.

It follows that A is an Artinian R-module.
 
(ii) Let A be an Artinian R-module. Let a ∈ A . By 10.2.9, there exists
t ∈ N such that a is annihilated by m  Now let r ∈ R.
 t = mt R.  Then
t
there exists r ∈ R such that r − r ∈ m R. Therefore ra = ra , so


that, when A is regarded as an R-module by means of the natural ring


homomorphism R → R,  every R-submodule of A is automatically an

R-submodule, and therefore A is an Artinian R-module. If we then re-

gard the Artinian R-module A as an R-module using the method of

10.2.9, we recover the original R-module structure on A .
(iii) Thus, by parts (i) and (ii), the category of all Artinian R-modules and
all R-homomorphisms between them is equivalent to the category of all

Artinian R-modules 
and all R-homomorphisms between them; in fact,
we can regard these two categories as essentially the same.
(iv) Let ν be the natural transformation from the identity functor to the func-
tor ( • ) ⊗R HomR (E, E) (from C(R) to itself) given by the formula
νM (m) = m ⊗ IdE for each R-module M and each m ∈ M .
Let N be a finitely generated R-module. Then, since N is a homomor-
phic image of Rh for some h ∈ N, the Matlis dual D(N ) is isomorphic
to a submodule of D(Rh ); since D(Rh ) ∼ = D(R)h ∼ = E h , we deduce
from 10.2.8 that D(N ) is Artinian. Furthermore, it is straightforward to
check that the diagram
νN
N - N ⊗R HomR (E, E)
@
μN @ ξN,E,E
@
@
R ?
DD(N ) ,

where ξN,E,E is the isomorphism given by 10.2.16, commutes. Now, by


10.2.2(i), the natural map μN is injective; by 10.2.11, HomR (E, E) ∼
=R
under an isomorphism which maps IdE to 1R , and so, speaking loosely,
10.2 Matlis duality 209

we can regard the duality map μN : N −→ DD(N ) as the embedding


of N into its completion.
(v) Let A be an Artinian R-module. Part (i) shows that A can be regarded

as an Artinian R-module in a natural way; by 10.2.5, the R-module

E is Artinian, and so it too can be regarded as an Artinian R-module
in a natural way; by 10.2.10, there is an R-isomorphism E(R/m) ∼
 =
 m).
ER (R/  Also, D(A) = HomR (A, E) = HomR (A, E), by part (i)

(and 10.2.9), and so D(A) has a natural (unambiguous) R-module struc-
ture. It follows from these observations that, as R-module, D(A) ∼
 =
 m));
HomR (A, ER (R/  the Matlis Duality Theorem 10.2.12 shows that

this is a Noetherian R-module.

We now summarize some of the main points of 10.2.18 in a Partial Matlis


Duality Theorem.

10.2.19 Partial Matlis Duality Theorem. Suppose (R, m) is local (but not
necessarily complete), and use the notation of 10.2.1. (Thus E := E(R/m)
and D := HomR ( • , E).) By 10.2.5 and 10.2.9, the R-module E has a natural

structure as an R-module. 
 denote the maximal ideal of R.
Let m
 such
(i) By 10.2.11, for each f ∈ HomR (E, E), there is a unique rf ∈ R
that f (x) = rf x for all x ∈ E.
(ii) Whenever N is a finitely generated R-module, the natural homomor-
phism μN : N −→ DD(N ) is injective, D(N ) is Artinian, and there is
a commutative diagram

N - N⊗ R
R

@
μN
@ ∼
=
@
@
R ?
DD(N ) ,

in which the horizontal homomorphism is the canonical one.


(iii) Whenever A is an Artinian R-module, it has a natural structure as an

R-module, 
it is Artinian as such, and D(A) is a Noetherian R-module

R-isomorphic 
to the Matlis dual of A over R.

Proof. All the claims were explained in 10.2.18.

We show how the Partial Matlis Duality Theorem 10.2.19 can be used to
extend 10.2.14 to the case where the underlying local ring is not necessarily
complete.
210 Matlis duality

10.2.20 Corollary. Suppose that (R, m) is local (but not necessarily com-
plete), and use the notation of 10.2.1.
Let N be a Noetherian R-module, so that D(N ) is an Artinian R-module
by 10.2.19(ii), and we can form the set Att D(N ). Then Att D(N ) = Ass N .
Proof. Let p ∈ Ass N , so that N has a submodule B with annihilator equal to
p. Now D is an exact, contravariant functor; also, by 10.2.2(ii), the annihilators
of an R-module M and its Matlis dual D(M ) are equal. Therefore D(N ) has a
homomorphic image D(B) with annihilator equal to p, and so p ∈ Att D(N )
by 7.2.5.
Conversely, let p ∈ AttR D(N ), so that D(N ) has an R-homomorphic
image A with annihilator equal to p, by 7.2.5. Application of D and another
use of 10.2.2(ii) now show that DD(N ) has an R-submodule with annihilator
equal to p. By 10.2.19(ii), there is an R-isomorphism DD(N ) ∼ = N ⊗R R. 

Therefore there exists an R-submodule C of N ⊗R R with (0 :R C) = p.
Let C R be the R-submodule
 of N ⊗R R  generated by C. Note that C R  is a

finitely generated R-module, and that
 ∩ R = (0 :R C R)
(0 :R C R)  = (0 :R C) = p.

 such that P ∩ R = p. Of course,


It follows that there exists P ∈ AssR (C R)

P ∈ AssR (N ⊗R R). We can now use [50, Theorem 23.2] to deduce that
p ∈ AssR N .
10.2.21 Exercise. Suppose that (R, m) is local and complete, and use the
notation of 10.2.1. Assume that dim R > 0. Prove that there exists an R-
module G which is neither Noetherian nor Artinian but for which μG : G −→
DD(G) is an isomorphism.
11

Local duality

Suppose, temporarily, that (R, m) is local, and that M is a finitely generated R-


module. In Theorem 7.1.3, we showed that Hm i
(M ) is Artinian for all i ∈ N0 .
When R is complete, Matlis duality (see 10.2.12) provides a very satisfac-
tory correspondence between the category of Artinian R-modules and the cat-
egory of Noetherian R-modules, and so it is natural to ask, in this situation,
which Noetherian R-modules correspond to the local cohomology modules
Hm i
(M ) (i ∈ N0 ). Local duality provides a really useful answer to this ques-
tion.

In fact, the Local Duality Theorem 11.2.6 concerns the situation where R
is a (not necessarily complete) homomorphic image of a Gorenstein local ring
(R , m ) of dimension n . (Cohen’s Structure Theorem for complete local rings
(see [50, Theorem 29.4(ii)], for example) ensures that this hypothesis would be
satisfied if R were complete.) The Local Duality Theorem tells us that, if M
is a finitely generated R-module, then, for each i ∈ N0 , the local cohomology
i
module Hm (M ) is isomorphic to the Matlis dual of the finitely generated R-
n −i
module ExtR  (M, R ), and, as R is Gorenstein, quite a lot is known about
these ‘Ext’ modules. The Local Duality Theorem provides a fundamental tool
for the study of local cohomology modules with respect to the maximal ideal of
a local ring. Although it only applies to local rings which can be expressed as
homomorphic images of Gorenstein local rings, this is not a great restriction,
because the class of such local rings includes the local rings of points on affine
and quasi-affine varieties, and, as mentioned above, all complete local rings.

At the end of the chapter, some exercises are given to show that a local ring
which cannot be expressed as a homomorphic image of a Gorenstein local ring
need not have all the good properties flowing from the Local Duality Theorem.
212 Local duality

11.1 Minimal injective resolutions


The work in this chapter will involve knowledge of the structure of a mini-
mal injective resolution of a Gorenstein local ring; we begin by reviewing the
concept of minimal injective resolution.

11.1.1 Definition. Let M be an R-module. A minimal injective resolution of


M is an injective resolution
d0 di
I • : 0 −→ I 0 −→ I 1 −→ · · · −→ I i −→ I i+1 −→ · · ·

of M such that I n is an essential extension of Ker dn for every n ∈ N0 .

11.1.2 Exercise. Let M be an R-module. Prove that M has a minimal in-


jective resolution. (Start with the monomorphism M −→ E(M ), and, suc-
cessively, take the injective envelope of the cokernel of the last map you have
constructed.)

Suppose that, with the notation of 11.1.1, the injective resolution I • of M is


minimal. There is an augmentation R-homomorphism α : M → I 0 such that
the sequence
α d0 di
0 −→ M −→ I 0 −→ I 1 −→ · · · −→ I i −→ I i+1 −→ · · ·

is exact. Since I 0 is an essential extension of Ker d0 and Ker d0 = Im α ∼= M,


it follows that I 0 ∼
= E(M ), the injective envelope of M , and so is uniquely de-
termined up to isomorphism by M (see 10.1.1(v)). Actually, all the terms in a
minimal injective resolution of M are uniquely determined up to isomorphism:
this is the subject of the next exercise.

11.1.3 Exercise. Let M be an R-module. Let


d0 di
I • : 0 −→ I 0 −→ I 1 −→ · · · −→ I i −→ I i+1 −→ · · ·

and
δ0 δi
J • : 0 −→ J 0 −→ J 1 −→ · · · −→ J i −→ J i+1 −→ · · ·

be minimal injective resolutions of M , so that there are R-homomorphisms


α : M → I 0 and β : M → J 0 such that the sequences
α d0 di
0 −→ M −→ I 0 −→ I 1 −→ · · · −→ I i −→ I i+1 −→ · · ·

and
β δ0 δi
0 −→ M −→ J 0 −→ J 1 −→ · · · −→ J i −→ J i+1 −→ · · ·
11.1 Minimal injective resolutions 213

are both exact. Show that there is an isomorphism of complexes


 
φ• = φi i∈N0 : I • −→ J •

such that the diagram


α
M - I0

φ0
?
M
β
- J0

commutes.

11.1.4 Notation and Definition. Let M be an R-module. By 11.1.3, for each


i ∈ N0 , the i-th term in a minimal injective resolution of M is uniquely deter-
mined, up to isomorphism, by M , and is independent of the choice of minimal
injective resolution of M . We denote this i-th term by E i (M ), or by ERi
(M )
when it is desirable to emphasize the underlying ring.
Observe that E 0 (M ) ∼= E(M ), the injective envelope of M .
For each i ∈ N0 and each p ∈ Spec(R), we define the i-th Bass number
of M with respect to p as follows. By 10.1.9, there is a family (pα )α∈Λ of

prime ideals of R for which E i (M ) ∼= α∈Λ E(R/pα ). By 10.1.15 (see [50,
Theorem 18.5(iii)]), the cardinality of the set {α ∈ Λ : pα = p} depends only
on E i (M ) and p (and therefore only on M and p) and not on the particular
decomposition of E i (M ) (as a direct sum of indecomposable injective sub-
modules) chosen. This cardinality is denoted by μi (p, M ), and is referred to as
the i-th Bass number of M with respect to p. Symbolically, we write

E i (M ) ∼
= μi (p, M )E(R/p),
p∈Spec(R)

where ⊕μE denotes the direct sum of μ copies of E.

In the notation of 11.1.4, we can already give a description of μ0 (p, M ),


since it follows from 10.1.15 (and the fact that E 0 (M ) ∼
= E(M )) that

μ0 (p, M ) = dimk(p) HomRp (k(p), E(M )p ),

where k(p) = Rp /pRp , the residue field of the local ring Rp . In fact, a re-
finement of this, and analogues for the higher Bass numbers, are available: we
shall refer to Matsumura [50, Theorem 18.7] for these results, although we first
give, for the reader’s convenience, a slight refinement of one of Matsumura’s
lemmas.
214 Local duality

11.1.5 Lemma. (See [50, §18, Lemma 6].) Let f : L → M be a homomor-


phism of R-modules such that M is an essential extension of Im f . Let S be
a multiplicatively closed subset of R. Then S −1 M is an essential extension
of its submodule Im(S −1 f ) (where S −1 f : S −1 L → S −1 M is the S −1 R-
homomorphism induced by f ).
Proof. Suppose that x/s, where x ∈ M and s ∈ S, is a non-zero element
of S −1 M . Then S −1 (Rx) is a non-zero submodule of S −1 M , and so there
exists p ∈ Ass(Rx) such that p ∩ S = ∅. Also, there exists r ∈ R such that
(0 : rx) = p.
Since M is an essential extension of Im f , there exists r ∈ R such that
0 = r rx = f (y) for some y ∈ L. Now (0 : r rx) = p ⊆ R \ S, and so
r r x r rx f (y) y
0= = = = (S −1 f ) ∈ Im(S −1 f ).
1 s s s s
11.1.6 Corollary. Let M be an R-module, and let
d0 di
I • : 0 −→ I 0 −→ I 1 −→ · · · −→ I i −→ I i+1 −→ · · ·
be a minimal injective resolution of M . Let S be a multiplicatively closed
subset of R. Then
S −1 (di )
0 - S −1 (I 0 ) - ··· - S −1 (I i ) - S −1 (I i+1 ) - ···

is a minimal injective resolution of the S −1 R-module S −1 M .


Thus, for all i ∈ N0 ,
(i) S −1 (ERi
(M )) ∼
= ESi −1 R (S −1 M ) (as S −1 R-modules); and
−1 −1
(ii) μ (S p, S M ) = μi (p, M ) for all p ∈ Spec(R) with p ∩ S = ∅.
i

Proof. All the claims follow easily from 10.1.14 and 11.1.5.
11.1.7 Corollary. Let the situation and notation be as in 11.1.6. Then, for all
p ∈ Spec(R) and i ∈ N0 , the induced homomorphism HomRp (Rp /pRp , dip )
is zero.
Proof. In view of 11.1.6, we can, and do, assume that (R, m) is local and that
p = m, and then we have to show that HomR (R/m, di ) = 0. To achieve this,
it is sufficient for us to show that (0 :I i m) ⊆ Ker di . However, (0 :I i m) is
a direct sum of simple R-modules, and each simple R-submodule of I i has
non-trivial intersection with Ker di and so is contained in Ker di .
11.1.8 Theorem. (See [50, Theorem 18.7].) Let M be an R-module. For each
i ∈ N0 and each p ∈ Spec(R), the i-th Bass number μi (p, M ) is given by
 
μi (p, M ) = dimk(p) ExtiRp (k(p), Mp ) = dimk(p) ExtiR (R/p, M ) p ,
11.1 Minimal injective resolutions 215

where k(p) = Rp /pRp , the residue field of the local ring Rp . 

We have not explained the use of the word ‘minimal’ in the phrase ‘minimal
injective resolution’. We do this next.

11.1.9 Conventions. The injective dimension of an R-module M will be de-


noted by inj dim M or, occasionally, by inj dimR M if it is desirable to specify
the underlying ring concerned. We adopt the convention that a zero module has
injective dimension −∞.

11.1.10 Remark. Let M be an R-module, and let n ∈ N0 .

(i) If E n (M ) = 0, then E n+i (M ) = 0 for all i ∈ N, and, consequently,


inj dim M < n.
(ii) Conversely, if inj dim M < n, then, for all p ∈ Spec(R), we have
ExtnR (R/p, M ) = 0, so that μn (p, M ) = 0 by 11.1.8; it follows that
E n (M ) = 0.

11.1.11 Exercise. Let M be an R-module, and let


d0 di
I • : 0 −→ I 0 −→ I 1 −→ · · · −→ I i −→ I i+1 −→ · · ·

be a minimal injective resolution of M and


δ0 δi
J • : 0 −→ J 0 −→ J 1 −→ · · · −→ J i −→ J i+1 −→ · · ·

be an arbitrary injective resolution of M . Prove that there is a chain map of


complexes
 
φ• = φi i∈N0 : I • −→ J •

such that the diagram


α
M - I0

φ0
?
M
β
- J0

(in which the horizontal maps are the appropriate augmentation homomor-
phisms) commutes, and which is such that φi is a monomorphism for all
i ∈ N0 . Deduce that, for each i ∈ N0 , the i-th term J i has a direct summand
isomorphic to E i (M ).
216 Local duality

11.2 Local Duality Theorems


We shall again refer to Matsumura’s treatment of Gorenstein rings in [50, pp.
139–145]. We remind the reader of the following necessary and sufficient con-
dition, in terms of the Bass numbers μi (p, R), for R to be a Gorenstein ring.

11.2.1 Reminder. (See [50, Theorems 18.1 and 18.8].) The ring R is a
Gorenstein ring if and only if, for all i ∈ N0 and all p ∈ Spec(R),

i 0 if i = ht p,
μ (p, R) =
1 if i = ht p,

that is, if and only if E i (R) ∼
= E(R/p) for all i ∈ N0 .
p∈Spec(R)
ht p = i

11.2.2 Exercise. Suppose that (R, m) is a Gorenstein local ring of dimension


n, and let
d0 n−1

I• : 0 - I0 - I1 - ··· - I n−1 d - I n -0

be a minimal injective resolution of R. (We are using 11.2.1 when we ‘termi-


nate’ this resolution after the n-th term.) Set E := I n , so that E ∼
= E(R/m)
by 11.2.1. Let V := {p ∈ Spec(R) : ht p = n − 1 and p ⊇ a}, and, for each
p ∈ V , let c(p) denote the kernel of the natural ring homomorphism R −→ Rp .
   
(i) Prove that Im Γa (dn−1 ) = p∈V i∈N 0 :E p
(i)
.
(ii) Assume now that R is complete. Prove that

Han (R) ∼
= HomR p∈V c(p), E .

(The results of this exercise provided the key steps in a short proof of the
local Lichtenbaum–Hartshorne Vanishing Theorem (see 8.2.1) given by F. W.
Call and R. Y. Sharp in [8]: a reader who finds this exercise difficult is referred
to that paper for details.)

We now start our approach to the Local Duality Theorem.

11.2.3 Lemma. Let (R, m) be a Gorenstein local ring of dimension n. Then


Hmn
(R) ∼
= E(R/m), the injective envelope of the simple R-module R/m.

Proof. Let
d0 n−1

I• : 0 - I0 - I1 - ··· - I n−1 d - I n -0
11.2 Local Duality Theorems 217
n
be a minimal injective resolution of R. We calculate Hm (R) by working out

the n-th cohomology module of the complex Γm (I ).
By 11.2.1,


Ii = E(R/p) for all i ∈ N0 .
p∈Spec(R)
ht p = i

Let p ∈ Spec(R) with p = m. Then there exists r ∈ m \ p. Now, by 10.1.3(iii),


multiplication by r provides an automorphism of E(R/p), and so

Γm (E(R/p)) = 0.

Hence Γm (I i ) = 0 for all i ∈ N0 with i = n.


On the other hand, E(R/m) is m-torsion, again by 10.1.3(iii). Thus all the
terms other than the n-th of the complex Γm (I • ) are zero, while its n-th term
is isomorphic to E(R/m). Hence Hm n
(R) ∼
= E(R/m).

11.2.4 Exercise. This exercise is central to our treatment of local duality.

(i) Let T : C(R) −→ C(R) be a covariant R-linear functor, and let B be an


R-module.
Show that there is a natural transformation of functors

φ : T −→ HomR (HomR ( • , B), T (B))

(from C(R) to itself) which, for an R-module M , is such that

(φM (y))(f ) = T (f )(y) for all y ∈ T (M ) and all f ∈ HomR (M, B).

Show also that, if each endomorphism of B can be realized as multi-


plication by precisely one element of R, then φB is an isomorphism.
(ii) Now suppose that (R, m) is a Gorenstein local ring of dimension n; set
E := E(R/m) and D := HomR ( • , E).
Use part (i) and 11.2.3 to show that there is a natural transformation
of functors
n
φ 0 : Hm −→ D(HomR ( • , R))

(from C(R) to itself) which is such that φ0 R is an isomorphism.


Let M be a finitely generated R-module. Use Grothendieck’s Vanish-
ing Theorem 6.1.2, together with the fact that M can be included in an
exact sequence F1 −→ F0 −→ M −→ 0 in which F1 and F0 are finitely
generated free R-modules, to show that φ0 M is an isomorphism.

We can now present the Local Duality Theorem for a Gorenstein local ring.
218 Local duality

11.2.5 Local duality for a Gorenstein local ring. Let (R, m) be a local
Gorenstein ring of dimension n; set E := E(R/m) and D := HomR ( • , E).
By 11.2.4, there is a natural transformation of functors
n
φ0 : Hm −→ D(HomR ( • , R))

(from C(R) to itself) which is such that φ0 M is an isomorphism for every


finitely generated R-module M .
There is a unique extension of φ0 to a homomorphism
 n−i   
Φ := (φi )i∈N0 : Hm i∈N0
−→ D(Ext i
R ( • , R))
i∈N0

of (positive strongly) connected sequences of covariant functors from C(R) to


C(R). Furthermore, φi M is an isomorphism for all i ∈ N0 whenever M is a
finitely generated R-module.
In particular, for each finitely generated R-module M ,
n−i
Hm (M ) ∼
= D(ExtiR (M, R)) for all i ∈ Z.
n
Proof. By Grothendieck’s Vanishing Theorem  n−i6.1.2,
 the functor Hm is right
exact; it is an easy consequence of this that Hm i∈N0 is a positive strongly
connected sequence of covariant functors.
n−i
Since depth R = n, it follows from 6.2.8 and 3.4.10 that Hm (P ) = 0 for
all i ∈ N and all projective R-modules P ; also D(ExtR (P, R)) = 0 for all
i

i ∈ N and all projective R-modules P . It follows from the analogue of 1.3.4


for positive connected sequences that φ0 can be incorporated into a (uniquely
determined) homomorphism
 n−i   
Φ := (φi )i∈N0 : Hm i∈N0
−→ D(Ext i
R ( • , R))
i∈N0

of connected sequences. Furthermore, it is easy to prove by induction that, for


each i ∈ N, the homomorphism φi M is an isomorphism whenever M is a
finitely generated R-module: use the fact that such an M can be included in
an exact sequence 0 −→ K −→ F −→ M −→ 0 in which F is a finitely
generated free R-module.

11.2.6 Local Duality Theorem. Suppose that (R, m) is a local ring which
is a homomorphic image of a Gorenstein local ring (R , m ) of dimension n :
let f : R → R be a surjective ring homomorphism. Set E := E(R/m) and
D := HomR ( • , E).
An R-module M can be regarded as an R -module by means of f : then
M , and, for each j ∈ N0 , the modules ExtjR (M, R ) and D(ExtjR (M, R )),
inherit natural (R, R )-bimodule structures.
11.2 Local Duality Theorems 219

There is a homomorphism
 i n −i
Φ := (φi )i∈N0 : Hm i∈N0
−→ D(ExtR  ( • , R ))
i∈N0

of (negative strongly) connected sequences of covariant functors from C(R) to


C(R) which is such that φiM is an isomorphism for all i ∈ N0 whenever M is
a finitely generated R-module.
Consequently (since inj dimR R = n ), for each finitely generated R-
module M ,
i
Hm (M ) ∼ n −i
= D(ExtR 
n −i
(M, R )) = HomR (ExtR  (M, R ), E)

for all i ∈ Z.

Proof. Let E  := ER (R /m ), and let D := HomR ( • , E  ).


By 11.2.5, there is a homomorphism
n −i
 
Ψ := (ψi )i∈N0 : Hm  −→ D (ExtiR ( • , R )) i∈N0
i∈N0

of (positive strongly) connected sequences of covariant functors from C(R ) to


C(R ) which is such that ψi M  is an isomorphism for all i ∈ N0 whenever M 
is a finitely generated R -module. We can interpret
 i  n −i
(ψn −i )i∈N0 : Hm 
i∈N0
−→ D (ExtR  ( • , R ))
i∈N0

as a homomorphism of negative connected sequences.


Let R : C(R) → C(R ) denote the functor obtained from restriction of
scalars (using f ). If we precede each ψn −i by R , we can deduce from the
Independence Theorem 4.2.1 that there is a homomorphism
 i
(ψi )i∈N0 : Hm i∈N0
−→ D (ExtR n −i
 ( • , R ))
i∈N0

of (negative strongly) connected sequences of covariant functors from C(R) to


C(R ) which is such that ψM
i
is an isomorphism for all i ∈ N0 whenever M is
a finitely generated R-module.
Let c := Ker f . Let M be an arbitrary R-module. Then (0 :E  c), and, for
each j ∈ N0 , the modules ExtjR (M, R ) and D (ExtjR (M, R )) all inherit
j
natural R-module structures: for each of these modules, and for Hm (M ), we
   
have f (r )y = r y for all r ∈ R and y in the module. It follows that

D (ExtjR (M, R )) = HomR (ExtjR (M, R ), E  )


= HomR (ExtjR (M, R ), (0 :E  c))
= HomR (ExtjR (M, R ), (0 :E  c)).
220 Local duality

Thus, for each i ∈ N0 , we can regard


ψM
i i
: Hm n −i
(M ) −→ HomR (ExtR  (M, R ), (0 :E  c))
as an R-homomorphism. All that remains in order to complete the proof is to
note that, by 10.1.16, there are R-isomorphisms
(0 :E  c) ∼
= ER (R /m ) ∼
= ER (R/m) = E.
The following exercise will be useful in applications of the Local Duality
Theorem.
11.2.7 Exercise. Suppose that (R, m) is a local ring which is a homomor-
phic image of a local ring (R , m ): let f : R → R be a surjective ring ho-
momorphism. Let p ∈ Spec(R), and let p = f −1 (p), ∈ Spec(R ). Of course,
R /p ∼ = R/p, so that these rings have equal dimensions. Moreover, observe
that f induces a surjective ring homomorphism f  : Rp  −→ Rp which is such
that f  (r /s ) = f (r )/f (s ) for all r ∈ R and s ∈ R \ p .
Let M be a finitely generated R-module, and let j ∈ N0 . We can form
ExtjR (Mp , Rp  ), which has a natural Rp -module structure. We can also lo-
p

calize the R-module ExtjR (M, R ) at p. Show that

ExtjR (Mp , Rp  ) ∼


= ExtR (M, R )
j
as Rp -modules.
p p

The next exercise illustrates how the Local Duality Theorem can be used to
obtain some of the results of 7.3.2 and 7.3.3 in the case where the local ring
concerned is a homomorphic image of a Gorenstein local ring.
11.2.8 Exercise. Suppose that (R, m) is a local ring which is a homomorphic
image of a Gorenstein local ring. Let M be a finitely generated R-module. Let
p ∈ Spec(R), and let dim R/p = t.
(i) Let q ∈ Spec(R) be such that q ⊆ p, and let i ∈ Z. By 7.1.3, the Rp -
i i+t
module HpR p
(Mp ) is Artinian, and Hm (M ) is an Artinian R-module.
Show that
qRp ∈ AttRp HpR
i
p
(Mp ) if and only if q ∈ AttR (Hm
i+t
(M )).

(ii) Suppose p ∈ Ass M . Show that Hm t


(M ) = 0 and p ∈ Att(Hm t
(M )).
t
Show further that, if t > 0, then Hm (M ) is not finitely generated.
n
(iii) Deduce from (ii) that, if M is non-zero and of dimension n, then Hm (M )
= 0 and {p ∈ Ass M : dim R/p = n} ⊆ Att(Hm (M )); deduce also
n
n
that, if n > 0, then Hm (M ) is not finitely generated. Compare 7.3.2 and
7.3.3.
11.2 Local Duality Theorems 221

11.2.9 Exercise. Suppose that (R, m) is a local ring which is a homomorphic


image of a Gorenstein local ring. Let M be a finitely generated R-module,
j
and let j be an integer for which Hm (M ) = 0. Show that, for each p ∈
j
AttR (Hm (M )), we have dim R/p ≤ j.

Among the hypotheses for 11.2.8 and 11.2.9 was the assumption that the
underlying local ring is a homomorphic image of a Gorenstein local ring. Un-
fortunately, the results of those exercises are not all true for general local rings,
as can be seen from Exercise 11.2.13 below.
The next two exercises show that one of the implications in 11.2.8(i) is true
in general. These two exercises are directed at those readers who are expe-
rienced at working with the fibres of flat homomorphisms of commutative
Noetherian rings.

11.2.10 Exercise. Suppose that (R, m) is a local ring, that (R , m ) is a sec-
ond local ring, and that f : R −→ R is a flat local ring homomorphism
such that the extension mR of m to R under f is m -primary. Let i ∈ N0 ,
and let M be a non-zero, finitely generated R-module with the property that
i
Hm (M ) = 0.

(i) Use the Flat Base Change Theorem 4.3.2 to see that there is an R -
 ∼ 
 (M ⊗R R ) = Hm (M ) ⊗R R .
i i
isomorphism Hm
(ii) Let
i
Hm (M ) = S1 + · · · + Sh with Sj pj -secondary (1 ≤ j ≤ h)
i
be a minimal secondary representation of Hm (M ). For all j = 1, . . . , h,
let uj : Sj → Hm (M ) be the inclusion map, and let
i

Tj := (uj ⊗ IdR )(Sj ⊗R R ).


i
Show that Hm (M ) ⊗R R = T1 + · · · + Th and that

f −1 (P) : P ∈ AttR (Tj ) = {pj } for all j = 1, . . . , h.

(iii) Show that


i
AttR (Hm (M )) = f −1 (P) : P ∈ AttR (Hm
i 
 (M ⊗R R )) .

11.2.11 Exercise. Suppose that (R, m) is a local ring. Let M be a finitely


generated R-module. Let p ∈ Spec(R) and let dim R/p = t. Let i ∈ Z and
q ∈ Spec(R) be such that q ⊆ p and qRp ∈ AttRp HpR
i
p
(Mp ) . Prove that
q ∈ AttR (Hm i+t
(M )).
(Here are some hints: use Exercise 11.2.10, and Cohen’s Structure Theorem
222 Local duality

that a complete local ring is a homomorphic image of a regular local ring, in


conjunction with 11.2.8(i).)
11.2.12 Exercise. Use 11.2.11 to extend the result of 11.2.8(ii) to all local
rings.
Thus, in detail, suppose that (R, m) is a local ring. Let M be a non-zero,
finitely generated R-module, and let p ∈ Ass M . Suppose that dim R/p = t.
Prove that Hmt
(M ) = 0 and p ∈ Att(Hm t
(M )).
11.2.13 Exercise. This exercise shows that, if, in 11.2.9, the hypothesis that
the local ring (R, m) be a homomorphic image of a Gorenstein local ring is
dropped, then the corresponding statement is no longer always true. At the
same time, it shows that the result of 11.2.8(i) is not true for every local ring.
For a counterexample, suppose that (R, m) is a 2-dimensional local domain
whose completion (R,  m)
 possesses an embedded prime ideal P (associated
to its zero ideal). Such an example has been constructed by D. Ferrand and M.
Raynaud in [18]. Use 11.2.12 and 11.2.10 to prove that 0 ∈ Att(Hm 1
(R)), and
deduce that the result of 11.2.9 is not true for every local ring.
Show also that, for each p ∈ Spec(R) having dim R/p = 1,

0Rp ∈ AttRp HpR


0
p
(Rp ) .

Deduce that the result of 11.2.8(i) is not true for every local ring.
12

Canonical modules

Suppose that (R, m) is local of dimension n > 0. We have seen earlier in


the book some illustrations of the importance of the so-called ‘top’ local co-
n
homology module Hm (R) of R. (It is referred to as the ‘top’ local cohomol-
i
ogy module because Hm (R) = 0 for all i > n by Grothendieck’s Vanishing
n
Theorem 6.1.2, whereas Hm (R) = 0 by the Non-vanishing Theorem 6.1.4.)
n
Now Hm (R) is an Artinian R-module (by 7.1.3) and is not finitely generated;
commutative algebraists tend to be brought up to work with finitely generated
modules, and Artinian modules are perhaps a little less familiar. The philoso-
phy behind the concept of canonical module is the following: wouldn’t it be
n
nice if we could, in some sense, replace Hm (R) by some finitely generated
R-module that we could work with effectively to achieve our desired results?
The theory of canonical modules for Cohen–Macaulay local rings is devel-
oped by Bruns and Herzog in [7, Chapter 3]; their account is partly based on the
lecture notes of Herzog and E. Kunz [33]. However, here we are going to work
in the more general setting of an arbitrary local ring (R, m), and define a canon-
ical module for R to be a finitely generated R-module K whose Matlis dual
n
HomR (K, ER (R/m)) is isomorphic to Hm (R), where n := dim R. In the
special case where R is Cohen–Macaulay, this condition turns out to be equiva-
lent to Bruns’ and Herzog’s definition, namely that a canonical module for R is
a Cohen–Macaulay R-module K of dimension n for which inj dimR K < ∞
and μn (m, K) = 1. The more general approach that we shall take is based on
work of Y. Aoyama [1], [2]; the definition we use is that employed by Hochster
and Huneke in [39].
If R is a homomorphic image of a Gorenstein local ring R of dimension n
then the Local Duality Theorem 11.2.6 shows that the R-module

n −n
ExtR  (R, R )
is a canonical module for R. Thus the assumption that our local ring is a
224 Canonical modules

homomorphic image of a Gorenstein local ring again appears in hypotheses. In


fact, there is a result, proved independently by H.-B. Foxby [19] and I. Reiten
[69], to the effect that the Cohen–Macaulay local ring R admits a canonical
module if and only if R is a homomorphic image of a Gorenstein local ring.
It turns out that canonical modules are intimately related to Serre’s con-
dition S2 . Recall that a non-zero finitely generated R-module M (here R is
not assumed to be local) is said to satisfy Serre’s condition S2 if and only if
depthRp Mp ≥ min{2, dimRp Mp } for all p ∈ Supp M . When R is a nor-
mal domain, R itself satisfies S2 . It turns out that, if K is a canonical module
for the local ring R, then depth K ≥ min{2, dim K}, and this is the key to
the connection with the condition S2 . Furthermore, the endomorphism ring
HomR (K, K) is a semi-local commutative Noetherian ring that, under mild
conditions on R, acts as the so-called ‘S2 -ification’ of R. (We did not invent
that name!) There are links between this concept and the generalized ideal
transforms we studied in Chapter 2.
The concept of canonical module is of fundamental importance in the study
of Cohen–Macaulay local rings, and therefore we shall reconcile our approach
with that of Bruns and Herzog. In connection with the fact that a canonical
module over a Cohen–Macaulay local ring has finite injective dimension, it is
perhaps worth noting that the present second author proved in [79, Corollary
2.3], that, over a Cohen–Macaulay homomorphic image R of a Gorenstein
local ring, the finitely generated R-modules of finite injective dimension are
precisely those R-modules M that can be included in an exact sequence
0 −→ Gs −→ Gs−1 −→ · · · −→ G1 −→ G0 −→ M −→ 0
in which each of G0 , . . . , Gs is a direct sum of finitely many copies of the
canonical R-module K. However, the proof of that result is beyond the scope
of this book.

12.1 Definition and basic properties


Many of the results in this and the next section are due to Y. Aoyama [1], [2].
Here, we shall adopt the definition of canonical module used by M. Hochster
and C. Huneke in [39].
12.1.1 Notation. Throughout this section, we shall assume (R, m) is local,
 m)
and we shall use n to denote dim R and (R,  to denote the completion of R.
12.1.2 Definition. A canonical module for R is a finitely generated R-mod-
ule K such that HomR (K, ER (R/m)) ∼
= Hm n
(R).
12.1 Definition and basic properties 225

12.1.3 Remarks. We can make the following observations right away.


n
(i) Recall that the R-module Hm (R) is Artinian (by 7.1.3), and that the
Matlis dual of a finitely generated R-module is (by 10.2.19(ii)) also
Artinian.

(ii) Suppose that R has a canonical module K. There are R-isomorphisms

ER (R/  ∼
 m) 
= ER (R/m) ⊗R R
n  ∼ 
 (R) = Hm (R) ⊗R R (by the Flat
n
(by 8.2.4, 10.2.9 and 10.2.10) and Hm
Base Change Theorem 4.3.2). Therefore, by [50, Theorem 7.11], there

are R-isomorphisms
n  ∼ ∼ 
 (R) = Hm (R) ⊗R R = HomR (K, ER (R/m)) ⊗R R
n
Hm
∼  ER (R/m) ⊗R R)
= Hom  (K ⊗R R, 
R
∼  E  (R/
= HomR (K ⊗R R,  m)).

R

Therefore K ⊗R R  is a canonical module for R.



(iii) It is immediate from the Local Duality Theorem 11.2.6 that, if R is a
homomorphic image of an n -dimensional Gorenstein local ring R , then

R has a canonical module, namely ExtnR−n (R, R ).
(iv) In particular, if R itself is Gorenstein, then R is a canonical module for
itself.
(v) In particular, when R is complete, so that it is a homomorphic image
of a regular local ring by Cohen’s Structure Theorem, R has a canon-
ical module; in fact, in that case, it follows from the Matlis Duality
Theorem 10.2.12 that each canonical module for R is isomorphic to
n
HomR (Hm (R), ER (R/m)); thus, when R is complete, there is, up to
isomorphism, exactly one canonical module for R.
(vi) Suppose, in the case where the local ring (R, m) is not necessarily com-

plete, that K  is an R-module for which there is a R-isomorphism
∼
K  ⊗R R n 
= HomR (Hm  
 (R), ER
 (R/m)).

Then, by the Matlis Duality Theorem 10.2.12 and the faithful flatness of
R over R, it follows that K  is a finitely generated R-module such that
K  ⊗R R  is a canonical module for R. One can use [50, Theorem 7.11],
8.2.4, 10.2.10 and 4.3.2 again to see that HomR (K  , ER (R/m))⊗R R ∼
=
  
Hm (R) ⊗R R as R-modules. Since HomR (K , ER (R/m)) and Hm (R)
n n

are Artinian R-modules (by 10.2.19(ii) and 7.1.6 respectively), it follows


from 8.2.4 that there is an R-isomorphism HomR (K  , ER (R/m)) ∼ =
Hm n
(R). Thus K  is a canonical module for R.
226 Canonical modules

This and part (ii) reconcile the above Definition 12.1.2 with that used
by Herzog and Kunz in [33, Definition 5.6].
Our next aim is to extend 12.1.3(v) by showing that, if R has a canonical
module, then any two canonical modules for R are isomorphic.
12.1.4 Lemma. Let K and L be two finitely generated R-modules. Then
the set IsoR (K, L) of all R-isomorphisms from K to L is an open subset of
HomR (K, L) in the m-adic topology. In fact, if α ∈ IsoR (K, L), then
α + m HomR (K, L) ⊆ IsoR (K, L).
Proof. If K and L are not isomorphic, then IsoR (K, L) = ∅, an open subset

=
of HomR (K, L). So suppose that α : K −→ L is an isomorphism. We shall
show that α + m HomR (K, L) ⊆ IsoR (K, L), and this will be enough to
complete the proof.
So let β ∈ α + m HomR (K, L). Then there exist r1 , . . . , rt ∈ m and
t
λ1 , . . . , λt ∈ HomR (K, L) such that β = α + i=1 ri λi . To show that
β ∈ IsoR (K, L), it is enough for us to show that α−1 ◦ β is an automor-
phism of K, and, since K is finitely generated, it is therefore enough for us to
show that α−1 ◦ β is surjective (by [81, Exercise 7.2]). This we do. Now
t t
α−1 ◦ β = α−1 ◦ (α + i=1 ri λi ) = IdK + i=1 ri α−1 ◦ λi .
Let g ∈ K. Then
t
g = IdK (g) = α−1 ◦ β − i=1 ri α
−1
◦ λi (g) ∈ Im(α−1 ◦ β) + mK.

Therefore K = Im(α−1 ◦β)+mK, and so it follows from Nakayama’s Lemma


that α−1 ◦ β is surjective, as required.
12.1.5 Lemma. Let R and R be two commutative Noetherian rings and
let T : C(R ) −→ C(R ) be an exact additive functor which is faithful in
the sense that T (M ) = 0, for an R -module M , implies that M = 0. Let
f : M −→ N be an R -homomorphism.

(i) If T is covariant and T (f ) is a monomorphism (respectively, an epimor-


phism), then f is a monomorphism (respectively, an epimorphism).
(ii) If T is contravariant and T (f ) is a monomorphism (respectively, an epi-
morphism), then f is an epimorphism (respectively, a monomorphism).
Proof. These are all proved similarly, and so we just prove the first part of
(ii). Let C := Coker f . Apply T to the exact sequence
f
M −→ N −→ C −→ 0
12.1 Definition and basic properties 227
T (f )
to obtain an exact sequence 0 −→ T (C) −→ T (N ) −→ T (M ). Since T (f )
is a monomorphism, we must have T (C) = 0. Since T is faithful, C = 0, so
that f is an epimorphism.
12.1.6 Theorem. Any two canonical modules K and K  for R are isomor-
phic.
Proof. By 12.1.3(ii), both K ⊗R R  and K  ⊗R R
 are canonical modules for

 
R, and so, by 12.1.3(v), there is an R-isomorphism α : K ⊗R R −→
= 
K  ⊗R R.
 
Since K and K are finitely generated R-modules, HomR (K, K ) is finitely
generated, and so the canonical injective map

HomR (K, K  ) −→ HomR (K, K  ) ⊗R R
provides the completion of HomR (K, K  ). Now, by [50, Theorem 7.11], there

is an R-isomorphism

 −→ Hom  (K ⊗R R,
λ : HomR (K, K  ) ⊗R R
=  K  ⊗R R)

R

which is such that λ(f ⊗ r) = r(f ⊗ IdR ) for all f ∈ HomR (K, K  ) and

r ∈ R.
 there
Since the image of HomR (K, K  ) is dense in HomR (K, K  ) ⊗R R,

exists g ∈ HomR (K, K ) such that

g ⊗ 1 ∈ λ−1 (α) + m  .
 HomR (K, K  ) ⊗R R

Apply λ to deduce that g ⊗ IdR ∈ α + m  K  ⊗R R).


 HomR (K ⊗R R,  But,
by 12.1.4, the coset α + m   
 HomR (K ⊗R R, K ⊗R R) consists entirely of
 to K  ⊗R R,
isomorphisms; therefore g ⊗IdR is an isomorphism from K ⊗R R 
 
so that g : K −→ K is an isomorphism by 12.1.5, because R is faithfully flat
over R.
12.1.7 Remark. Suppose that there exists a canonical module for R. By
12.1.6, any two canonical modules for R are isomorphic. We shall denote by
ωR one choice of canonical module for R. We shall sometimes use the clause
‘ωR exists’ as an abbreviation for ‘there exists a canonical module for R’.
Note that, in particular, if R is a homomorphic image of an n -dimensional
Gorenstein local ring R , and also a homomorphic image of an n -dimensional
Gorenstein local ring R , then, in view of 12.1.3(iii), there is an isomorphism
n −n
(R, R ) ∼

ExtR  = ExtnR−n (R, R ) of R-modules.
12.1.8 Exercise. Let M be a non-zero finitely generated R-module, and sup-
pose that R is a homomorphic image of a Gorenstein local ring R . For i ∈ N0 ,
dim R −i
let K i (M ) denote the R-module ExtR  (M, R ).
228 Canonical modules

(i) Show that K i (M ) = 0 for i < depthR M and for i > dim M , and that
K i (M ) = 0 for i = depthR M and for i = dim M .
(ii) Show that, if R is complete, then, for a fixed i ∈ N0 , the R-module
K i (M ) is, up to R-isomorphism, independent of the choice of the local
Gorenstein ring R (of which R is a homomorphic image).
(iii) Prove the result of (ii) without the assumption that R is complete. In
other words, assume only that R is a homomorphic image of a local
Gorenstein ring R and prove that K i (M ) is, up to R-isomorphism, in-
dependent of the choice of R . (Here is a hint: use 12.1.4 in a manner
similar to the way it was used in 12.1.6.)

The module K i (M ), for i = dim M , is called the i-th module of deficiency


of M . In a sense, these modules give, if M is not Cohen–Macaulay, an indica-
tion of the extent of the failure of M to be Cohen–Macaulay. They have been
studied by P. Schenzel in [74] and [75].

We shall see during the chapter that the existence of a canonical module for
a local ring R imposes some restrictions on R.

12.1.9 Proposition. Suppose that dim R = n > 0 and that ωR exists. Let
a1 , . . . , an be a system of parameters for R. Then

(i) AssR ωR = {p ∈ Spec(R) : dim R/p = n}; and


(ii) a1 is a non-zerodivisor on ωR and, if n ≥ 2, then a1 , a2 is an
ωR -sequence.

Proof. Let D denote the Matlis duality functor HomR ( • , ER (R/m)).


(i) By definition, D(ωR ) ∼ n
= Hm (R); also,
n
Att(Hm (R)) = {p ∈ Spec(R) : dim R/p = n}
by 7.3.2; therefore, by 10.2.20,
Ass ωR = Att(D(ωR )) = {p ∈ Spec(R) : dim R/p = n} .

(ii) Since dim R/Ra1 = n − 1, it follows from part (i) that a1 does not
belong to any associated prime of ωR . Therefore a1 is an ωR -sequence.
Now suppose that n ≥ 2. Let α : R/(0 :R a1 ) −→ R be the R-monomorph-
ism induced by multiplication by a1 . The exact sequence
α
0 −→ R/(0 :R a1 ) −→ R −→ R/Ra1 −→ 0
yields an exact sequence
n
Hm (α)
n−1
Hm (R/a1 R) - Hm
n
(R/(0 :R a1 )) - Hm
n
(R) - 0
12.1 Definition and basic properties 229

because dim R/a1 R < n. Let π : R −→ R/(0 :R a1 ) be the natural epimor-


phism. The exact sequence
π
0 −→ (0 :R a1 ) −→ R −→ R/(0 :R a1 ) −→ 0

=
yields the isomorphism Hm n
(π) : Hm n
(R) −→ Hm n
(R/(0 :R a1 )) because
dim(0 :R a1 ) < n. It follows that there is an exact sequence
1 a
n−1
Hm (R/a1 R) −→ Hm
n
(R) −→ n
Hm (R) −→ 0.
1 a
But the exact sequence 0 −→ ωR −→ ωR −→ ωR /a1 ωR −→ 0 yields the
exact sequence
1 a
0 −→ D(ωR /a1 ωR ) −→ D(ωR ) −→ D(ωR ) −→ 0.

Since Hm n
(R) ∼= D(ωR ), it follows that D(ωR /a1 ωR ) is a homomorphic im-
n−1
age of Hm (R/a1 R).
The natural images of a2 , . . . , an in R/a1 R are a system of parameters for
that (n − 1)-dimensional local ring, and, by the Independence Theorem 4.2.1,
we have Hm n−1
(R/a1 R) ∼ n−1
= Hm/a 1R
(R/a1 R). Now it follows from the above
proof of part (i) that the natural image of a2 is not in any attached prime ideal of
n−1 n−1 n−1
Hm/a 1R
(R/a1 R), and therefore a2 Hm (R/a1 R) = Hm (R/a1 R). Since
n−1
D(ωR /a1 ωR ) is a homomorphic image of Hm (R/a1 R), it follows that

a2 D(ωR /a1 ωR ) = D(ωR /a1 ωR ).

Therefore, as D is a contravariant faithful exact R-linear additive functor from


C(R) to itself, we can use 12.1.5(ii) to see that a2 is a non-zerodivisor on
ωR /a1 ωR , so that a1 , a2 is an ωR -sequence.

12.1.10 Exercise. Let the situation and notation be as in 12.1.9, and let M
be a finitely generated R-module for which HomR (M, ωR ) = 0. Show that
a1 is a non-zerodivisor on HomR (M, ωR ) and, if n ≥ 2, then a1 , a2 is an
HomR (M, ωR )-sequence.

Our next aim is to identify, in the case where ωR exists, the annihilator of
dim R
ωR . By Matlis Duality, this is the same as the annihilator of Hm (R): see
10.2.2(ii).

12.1.11 Exercise. Let c be the intersection of the (uniquely determined)


primary components q of the zero ideal of R for which dim R/q = n, and let
*
g := p.
p∈Ass R
dim R/p<n
230 Canonical modules

(i) Show that

c = Γg (R)
= {r ∈ R : br = 0 for some ideal b ⊂ R with dim R/b < n}.

(ii) Let J (R) denote the set of all ideals of R. Set

L(R) = {b ∈ J (R) : dim b < n}


= {b ∈ J (R) : bRp = 0 ∀ p ∈ Spec(R) with dim R/p = n} .

Recall from 7.3.1 that there is a maximum member d of the set L(R).
Show that d = c.

12.1.12 Notation. Bearing in mind 12.1.11, we shall denote the maximum


member of the set

{b : b is an ideal of R with dim b < n}

by uR (0). (Note that we really are discussing dim b for an ideal b, as opposed
to dim R/b.)
We shall use some of the properties of uR (0) established in 12.1.11.

12.1.13 Lemma. A canonical module for R is annihilated by uR (0). More-


over, a finitely generated R-module K that is annihilated by uR (0) (so that
it can be regarded as an R/uR (0)-module in the natural way) is a canonical
module for R if and only if it is a canonical module for R/uR (0).

Proof. Since dim uR (0) < n, it follows from Grothendieck’s Vanishing


Theorem 6.1.2 and the exact sequence

0 −→ uR (0) −→ R −→ R/uR (0) −→ 0



=
that there is an isomorphism Hmn
(R) −→ Hm
n
(R/uR (0)). Therefore, if ωR is
a canonical module for R, then

HomR (ωR , ER (R/m)) ∼ n


= Hm (R) ∼ n
= Hm (R/uR (0)).
n
Since uR (0) annihilates Hm (R/uR (0)), it follows that uR (0) annihilates

HomR (ωR , ER (R/m)).

As an R-module and its Matlis dual have the same annihilator (by 10.2.2(ii)),
it follows that uR (0) annihilates ωR .
Now let K be a finitely generated R-module that is annihilated by uR (0),
and denote ER (R/m) by E. The second displayed isomorphism in the last
paragraph, together with the Independence Theorem 4.2.1, show that there is
12.1 Definition and basic properties 231
n
an R-isomorphism Hm (R) ∼ n
= Hm/uR (0)
(R/uR (0)). Now, in view of 10.1.16,
we have

HomR (K, E) = HomR (K, (0 :E uR (0))) = HomR/uR (0) (K, (0 :E uR (0)))


 

= HomR/u (0) K, ER/u (0) ((R/uR (0))/(m/uR (0))) .
R R

Therefore there is an R-isomorphism HomR (K, E) ∼ n


= Hm (R) if and only if
there is an R/uR (0)-isomorphism

HomR/uR (0) (K, ER/uR (0) ((R/uR (0))/(m/uR (0)))) ∼ n


= Hm/uR (0)
(R/uR (0));

thus K is a canonical module for R if and only if it is a canonical module for


R/uR (0).

12.1.14 Remark. Often in this chapter, including in the proof of the next
theorem, we shall want to work under the hypothesis that R is a homomorphic
image of an n -dimensional Gorenstein local ring (R , m ), so that there is
a surjective ring homomorphism g : R −→ R. We explain here why it is
possible to assume that n = n.
Let c = Ker g. Now ht c = n − n, and, since a Gorenstein ring is Cohen–
Macaulay, there exists an R -sequence r1 , . . . , rn  −n contained in c of length
n − n. Now R /(r1 , . . . , rn  −n ) is a Gorenstein local ring of dimension n (see
[50, Exercise 18.1]), and so we can assume that n = n.

12.1.15 Theorem. Suppose that ωR exists. Then


n
(0 :R ωR ) = (0 :R Hm (R)) = uR (0).
n
Proof. It follows from 10.2.2(ii) that (0 :R ωR ) = (0 :R Hm (R)), and we
proved in 12.1.13 that uR (0) ⊆ (0 :R ωR ). It is therefore sufficient for us to
show that (0 :R ωR ) ⊆ uR (0).
We first consider the case where R is a homomorphic image of an n -
dimensional Gorenstein local ring (R , m ), so that there is a surjective ring
homomorphism g : R −→ R. In view of 12.1.14, we assume that n =
n. By 12.1.3(iii), we have ωR ∼ = HomR (R, R ). By 12.1.9(i), Ass ωR =
{p ∈ Spec(R) : dim R/p = n}. Thus, in order to deal with this case, it is
enough for us to show that, if p ∈ Spec(R) has dim R/p = n, then the Rp -
module (ωR )p has zero annihilator, for that would show that (0 :R ωR )Rp =
(0 :Rp (ωR )p ) = 0. See 12.1.11 and 12.1.12.
To achieve this, let p := g −1 (p), ∈ Spec(R ), and use 11.2.7 to see that
there are Rp -isomorphisms

(ωR )p ∼
= (HomR (R, R ))p ∼
= HomRp  (Rp , Rp  ).
232 Canonical modules

However, Rp  is a 0-dimensional Gorenstein local ring, and so

Rp  ∼
= ERp  (Rp  /p Rp  ).
 
Therefore 0 :Rp  HomRp  (Rp , Rp  ) = (0 :Rp  Rp ) by 10.2.2(ii), so that
(0 :Rp (ωR )p ) = 0, as required.
We have therefore proved the theorem under the additional assumption that
R is a homomorphic image of a Gorenstein local ring. We now deal with
the general case. We use 12.1.3(ii) to see that ωR exists and there is an R- 
isomorphism ωR ⊗R R ∼ = ωR . Now R  is a homomorphic image of a regular
local ring by Cohen’s Structure Theorem, and so it follows from the first two
paragraphs of this proof that (0 :R ωR ⊗R R)  = u  (0). Note that
R

 ∩ R = u  (0) ∩ R,
(0 :R ωR ) = (0 :R ωR ⊗R R) R

and so our proof will be complete if we can show that it is not possible for

there to be a P ∈ assR 0 with dim R/P < n and dim R/(P ∩ R) = n. So we
suppose that such a P exists and seek a contradiction. Denote P ∩ R by p.
The inclusion homomorphism R −→ R  induces a flat local homomorphism
 
f : Rp −→ RP . Note that depthRP RP = 0, so that the depth of the fibre
ring of f over the maximal ideal of Rp is 0 (by [50, Theorem 23.3]). Since
p ∈ Ass ωR by 12.1.9(i), we have depthRp (ωR )p = 0, so that another use of
[50, Theorem 23.3] shows that
 
depthRP (ωR )p ⊗Rp R P = 0.

But (ωR )p ⊗Rp R P -isomorphic to (ωR ⊗R R)


P is R  P and ωR ⊗R R  ∼
= ωR .

Since dim R/P < n, there exists  a ∈ P that is a parameter for  By
R.
12.1.9(ii), the element  
a is an (ωR ⊗R R)-sequence, so that
 
depthRP (ωR )p ⊗Rp R P = depth  (ωR ⊗R R)  P ≥ 1.
RP

This is a contradiction, and the proof is complete.

We now remind the reader of Serre’s conditions Si (i ∈ N). Recall our con-
vention whereby the depth of the zero module over a local ring is interpreted
as ∞, and also our occasional abbreviation of depthRp Mp by depth Mp (for
a finitely generated R-module M and p ∈ Spec(R)).

12.1.16 Definition. Let R be a commutative Noetherian ring. Let k ∈ N


and let M be a faithful finitely generated R -module. We say that M sat-
isfies Serre’s condition Sk , or, more loosely, that M is Sk , precisely when
depth Mp ≥ min{k, ht p} for all p ∈ Spec(R ).
12.1 Definition and basic properties 233

More generally, we say that a non-zero finitely generated R -module N sat-


isfies Serre’s condition Sk precisely when it is Sk as an R /(0 :R N )-module,
that is, if and only if depthRp Np ≥ min{k, dimRp Np } for all p ∈ Supp N ,
or, equivalently (in view of our convention concerning the depth of zero mod-
ules), if and only if depthRp Np ≥ min{k, dimRp Np } for all p ∈ Spec(R ).
12.1.17 Exercise. Let R be a commutative Noetherian ring, and let N be a
non-zero finitely generated R -module. Show that N is S2 if and only if
(i) each associated prime of N is a minimal member of Supp N , and
(ii) for every a ∈ R which is a non-zerodivisor on N , each associated prime
of N/aN is a minimal member of SuppR (N/aN ).
It is important for us to identify the support of a canonical R-module, if one
exists.
12.1.18 Theorem. Suppose that ωR exists.
(i) We have
Supp ωR = {p ∈ Spec(R) : ht p + dim R/p = dim R};
also the R-modules ωR and HomR (ωR , ωR ) satisfy the condition S2 .
(ii) If R is a homomorphic image of a Gorenstein local ring, then, for each
p ∈ Supp ωR , the localization (ωR )p is a canonical module for Rp .
Note. Some readers may be surprised by the additional hypothesis in part (ii).
In fact, the statement is true without the hypothesis that R be a homomorphic
image of a Gorenstein local ring, but the proof is beyond the scope of this book.
Interested readers are referred to Aoyama [2, Corollary 4.3].
Proof. We deal first with the case where R is a homomorphic image of an
n -dimensional Gorenstein local ring (R , m ) via a surjective ring homomor-
phism g : R −→ R. In view of 12.1.14, we assume that n = n, so that we
can take ωR = HomR (R, R ).
Let p ∈ Supp ωR , so that there exists q ∈ Ass ωR with p ⊇ q. By 12.1.9,
dim R/q = n. Since R is catenary, ht p/q + dim R/p = dim R/q, so that
ht p + dim R/p = n.
Now let p ∈ Spec(R) be such that ht p+dim R/p = n. Let p := g −1 (p), ∈
Spec(R ), and use 11.2.7 to see that there are Rp -isomorphisms
(ωR )p ∼
= (HomR (R, R ))p ∼
= HomRp  (Rp , Rp  ).

We have dim R /p = dim R/p and ht p ≥ ht p. Therefore we must have


ht p + dim R /p = n and ht p = ht p. Therefore HomRp  (Rp , Rp  ) is a
234 Canonical modules

canonical module for Rp , so that there is an Rp -isomorphism (ωR )p ∼


= ωR p .
In particular, p ∈ Supp ωR . Also, it follows from 12.1.9 that depth ωRp ≥
min{2, dim Rp }, and from 12.1.10 that
depth HomRp (ωRp , ωRp ) ≥ min{2, dim Rp }.
We can therefore conclude that both ωR and HomR (ωR , ωR ) are S2 .
All parts of the theorem have now been proved in the case where R is a
homomorphic image of a Gorenstein local ring. We now deal with the general

case. We use 12.1.3(ii) to see that ωR exists and there is an R-isomorphism
 ∼ 
ωR ⊗R R = ωR . Now R is a homomorphic image of a regular local ring by
Cohen’s Structure Theorem, and so we know that the claims of part (i) hold true

for the R-canonical module ωR ⊗R R. In particular, the R-modules
 ωR ⊗ R R 
 
and HomR (ωR ⊗R R, ωR ⊗R R) are S2 . An argument similar to one in the proof
of [50, Theorem 23.9] shows that the R-modules ωR and HomR (ωR , ωR ) are
both S2 .
We now introduce some notation for a general prime ideal p of R. Let P ∈
Spec(R) be a minimal prime ideal of pR  for which dim R/P
  R
= dim R/p =
dim R/p. Then P ∩ R = p, and the inclusion homomorphism R −→ R 
induces a flat local homomorphism f : Rp −→ R P with the property that
its fibre ring over the maximal ideal of Rp is Artinian. Hence ht P = ht p, and
we have ht p + dim R/p = ht P + dim R/P  ≤ n. Note also that there is a

RP -isomorphism
(ωR ⊗R R)  ∼
 ⊗ R 
R P = (ωR ⊗R Rp ) ⊗Rp RP .

Since f is faithfully flat, it follows that p ∈ Supp ωR if and only if P ∈


SuppR (ωR ⊗R R). 
Now suppose that p ∈ Supp ωR . Then P ∈ SuppR (ωR ⊗R R),  so that
 
ht P + dim R/P = n because R is a homomorphic image of a Gorenstein
local ring. Therefore ht p + dim R/p = n.
Conversely, suppose that ht p + dim R/p = n. Then ht P + dim R/P =
 
n, so that P ∈ SuppR (ωR ⊗R R) because R is a homomorphic image of a
Gorenstein local ring; therefore p ∈ Supp ωR .
All parts of the theorem have now been proved.
12.1.19 Exercise. Suppose R is a homomorphic image of a Gorenstein local
ring and that p ∈ Supp ωR . Show that uRp (0) = uR (0)Rp .
The following variant of the Local Duality Theorem will be useful.
12.1.20 Theorem. Suppose that ωR exists. Set E := E(R/m), and let D :=
HomR ( • , E).
12.1 Definition and basic properties 235

(i) There is a natural transformation of functors


n
φ 0 : Hm −→ D(HomR ( • , ωR ))
which is such that φ0 M is an isomorphism whenever M is a finitely
generated R-module.
(ii) Now suppose that R is Cohen–Macaulay. Then there is a unique exten-
sion of φ0 to a homomorphism
 n−i   
Φ := (φi )i∈N0 : Hm i∈N0
−→ D(ExtiR ( • , ωR )) i∈N0
of (positive strongly) connected sequences of covariant functors from
C(R) to C(R). Furthermore, φi M is an isomorphism for all i ∈ N0
whenever M is a finitely generated R-module.
In particular, for each finitely generated R-module M ,
n−i
Hm (M ) ∼
= D(ExtiR (M, ωR )) for all i ∈ Z.
Proof. (i) By definition, Hm n
(R) =∼ D(ωR ). By 6.1.10, the functor H n is
m
naturally equivalent to ( • ) ⊗R Hm n n
(R). Therefore Hm is naturally equivalent
to ( • ) ⊗R HomR (ωR , E). We can now make use of 10.2.16 in order to arrive
at a natural transformation of functors
n
φ 0 : Hm −→ D(HomR ( • , ωR ))
which is such that φ0 M is an isomorphism whenever M is a finitely generated
R-module.
(ii) Now suppose that R is Cohen–Macaulay. Here, we reason as in the
proof of 11.2.5. Since depth R = n, it follows from 6.2.8 and 3.4.10 that
Hmn−i
(P ) = 0 for all i ∈ N and all projective R-modules P ; also
D(ExtiR (P, ωR )) = 0
for all i ∈ N and all projective R-modules P . It follows from the analogue
of 1.3.4 for positive connected sequences that φ0 can be incorporated into a
(uniquely determined) homomorphism
 n−i   
Φ := (φi )i∈N0 : Hm i∈N0
−→ D(ExtiR ( • , ωR )) i∈N0
of connected sequences, and it is easy to prove by induction that, for each
i ∈ N, the homomorphism φi M is an isomorphism whenever M is a finitely
generated R-module. For i < 0, we have Hm n−i
(M ) = ExtiR (M, ωR ) = 0 for
each R-module M .
12.1.21 Corollary. Suppose that the local ring R is Cohen–Macaulay, and
that ωR exists. Then ωR has finite injective dimension equal to n, and is a
Cohen–Macaulay R-module of dimension n.
236 Canonical modules

Proof. By 12.1.20 and the fact that ER (R/m) is an injective cogenerator


for R, we have ExtiR (M, ωR ) = 0 for all i > n and all finitely generated
R-modules M , so that, in particular, ExtiR (R/p, ωR ) = 0 for all i > n
and all p ∈ Spec(R). Thus all the Bass numbers μi (p, ωR ) are zero for all
i > n and all p ∈ Spec(R) (by 11.1.8). Also ExtnR (R/m, ωR ) = 0 because
D(ExtnR (R/m, ωR )) ∼ = Γm (R/m) = R/m = 0. Therefore ωR has finite in-
jective dimension equal to n.
It also follows similarly from 12.1.20 that
D(ExtiR (R/m, ωR )) ∼ n−i
= Hm (R/m) = 0 for all i ∈ {0, . . . , n − 1},
so that depth ωR = n and ωR is a Cohen–Macaulay R-module.
12.1.22 Corollary. Suppose that the local ring R is Cohen–Macaulay, and
that ωR exists. Then R is Gorenstein if and only if ωR ∼
= R, that is, if and only
if R is a canonical module for itself.
Proof. We observed in 12.1.3(iv) that a Gorenstein local ring is a canonical
module for itself. Conversely, suppose that R is a Cohen–Macaulay local ring
and that ωR exists and is isomorphic to R. Then inj dimR R < ∞, by 12.1.21,
and so R is Gorenstein.
12.1.23 Exercise. Suppose that the local ring R is Cohen–Macaulay and
that ωR exists. Show that μi (m, ωR ) = δi,ht m (the Kronecker delta) for all
i ∈ N0 . (Here is a hint: use the method of proof in 12.1.21.)
12.1.24 Exercise. Suppose that the local ring R is Cohen–Macaulay and
a homomorphic image of a Gorenstein local ring, so that R has a canonical
module ωR , by 12.1.3(iii). Show that, for all i ∈ N0 and all p ∈ Spec(R),

i 0 if i = ht p,
μ (p, ωR ) =
1 if i = ht p,
that is, μi (p, ωR ) is equal to the Kronecker delta δi,ht p . (Here is a hint: use
12.1.18 and 12.1.23.)
12.1.25 Exercise. Suppose that the local ring R is Cohen–Macaulay. Sup-
pose that K is a finitely generated R-module such that μi (m, K) = δi,ht m
for all i ∈ N0 . The purposes of this exercise are to show that K is actually a
canonical module for R, and to prove a result of I. Reiten [69] and H.-B. Foxby
[19] that R is a homomorphic image of a Gorenstein local ring. Recall that n
denotes dim R.
(i) Show that there exists an R-sequence r1 , . . . , rn which is also a K-
sequence.
12.1 Definition and basic properties 237

(ii) Denote the local ring R/(r1 , . . . , rn ) by (R, m) and use K to denote
K/(r1 , . . . , rn )K. Use [7, 3.1.16] to calculate the μjR (m, K) (j ∈ N0 ),
and conclude that there is an R-isomorphism K ∼ = ER (R/m). Deduce
that
(0 :R K/(r1 , . . . , rn )K) = (r1 , . . . , rn ).

(iii) By considering r1t , . . . , rnt for t ∈ N, show that (0 :R K) = 0.


(iv) Let R be the trivial extension of R by K, introduced in 6.2.12. It fol-
lows from that exercise that R is an n-dimensional local ring. Show
that (r1 , 0), . . . , (rn , 0) is an R -sequence, and deduce that R is Cohen–
Macaulay.
(v) Show that R / ((r1 , 0), . . . , (rn , 0)) is isomorphic to the trivial extension
of R by K, and calculate the socle of the latter ring. Deduce that R is
Gorenstein.
(vi) Conclude that R is a homomorphic image of a Gorenstein local ring, so
that ωR exists, by 12.1.3(iii). Prove that K ∼ = ωR . (Here is a hint: recall

from 12.1.3(iii) that ωR = HomR (R, R ).) 


12.1.26 Remark. Suppose that the local ring R is Cohen–Macaulay.


It follows from 12.1.3(iii), 12.1.23 and 12.1.25 that R has a canonical mod-
ule if and only if R is a homomorphic image of a Gorenstein local ring, and
that, when this is the case, an R-module K is a canonical module for R if and
only if it is finitely generated and μi (m, K) = δi,ht m for all i ∈ N0 .
In their book on Cohen–Macaulay rings [7], Bruns and Herzog (essentially)
defined a canonical module for a Cohen–Macaulay local ring (R, m) to be a
finitely generated R-module K for which μi (m, K) = δi,ht m for all i ∈ N0 .
Thus 12.1.23 and 12.1.25 reconcile their approach with the one we have taken
in this chapter.

12.1.27 Remark. Suppose (R, m) is a Cohen–Macaulay local ring. It follows


from 12.1.25, 12.1.26 and 12.1.24 that a finitely generated R-module C is a
canonical module for R if and only if μi (m, C) = δi,ht m (Kronecker delta)
for all i ∈ N0 , and that, when this is the case, R is a homomorphic image of a
Gorenstein local ring and Cp is a canonical module for Rp for all p ∈ Spec(R).

Bruns and Herzog also extended the definition of canonical module to the
case where the underlying ring is not necessarily local (but is Cohen–
Macaulay).

12.1.28 Definition. Let R be a (not necessarily local) Cohen–Macaulay


(commutative Noetherian) ring. Let C be a finitely generated R -module. We
238 Canonical modules

say that C is a canonical module for R precisely when Cm is a canonical



module for Rm for all maximal ideals m of R .
In that case, Cp is a canonical module for Rp for all p ∈ Spec(R ), by
12.1.27.
12.1.29 Remark. Let R be a (not necessarily local) Cohen–Macaulay (com-
mutative Noetherian) ring. Let C be a finitely generated R -module. It follows
from 12.1.27 that C is a canonical module for R if and only if μi (p, C) =
δi,ht p for all p ∈ Spec(R ) and all i ∈ N0 .
12.1.30 Proposition. Let R be a (not necessarily local) Cohen–Macaulay
(commutative Noetherian) ring, and assume that R has a canonical module
C. Then the trivial extension R ∝ C of R by C (see 6.2.12) is a Gorenstein
ring, so that R is a homomorphic image of a Gorenstein ring.
Proof. By 6.2.12(ii), a general prime ideal P of R ∝ C has the form p × C
for a prime ideal p of R ; also, 6.2.12(iv) shows that (R ∝ C)P ∼
= Rp ∝ Cp ,
which is Gorenstein by 12.1.25. Therefore R ∝ C is Gorenstein. The ring
homomorphism φ : R ∝ C −→ R for which φ((r, c)) = r for all r ∈ R and
c ∈ C is surjective.

12.2 The endomorphism ring


When (R, m) is local and ωR exists, it turns out that HomR (ωR , ωR ) has some
very interesting and useful properties, some of which are relevant to the theory
of S2 -ifications that we shall develop in the final section of this chapter. In this
section, we shall concentrate on the R-module structure of HomR (ωR , ωR );
we shall consider its ring structure later in the chapter.
12.2.1 Notation. Throughout this section also, we shall assume that (R, m)
is local, we shall use n to denote dim R, and we shall denote the completion
of R by (R, m).

In order to present the nice properties of HomR (ωR , ωR ), we are going to
concentrate first on the case where (R, m) is a Cohen–Macaulay local ring
(and ωR exists): by 12.1.15, we have (0 :R ωR ) = uR (0), and this is zero
when R is Cohen–Macaulay; in fact, we shall show, in that case, that each
R-endomorphism of ωR is given by multiplication by a uniquely determined
element of R.
12.2.2 Proposition. Suppose the local ring R is Cohen–Macaulay, and ωR
exists. Let a1 , . . . , aj ∈ m. Then a1 , . . . , aj is an R-sequence if and only if it
12.2 The endomorphism ring 239

is an ωR -sequence; moreover, when this is the case, ωR/(a1 ,...,aj ) exists and
ωR /(a1 , . . . , aj )ωR ∼
= ωR/(a1 ,...,aj ) .
Proof. Let D denote the Matlis duality functor HomR ( • , ER (R/m)). Note
that there is nothing to prove if n = 0, and so we suppose that n > 0.
We first deal with the case in which j = 1. Since R is Cohen–Macaulay,
Ass R = {p ∈ Spec(R) : dim R/p = n}, and this is equal to Ass ωR , by
12.1.9(i). Therefore a1 is a non-zerodivisor on R if and only if it is a non-
zerodivisor on ωR . Suppose that this is the case. Since grade m = n, we have
n−1
Hm (R) = 0, by 6.2.7. Therefore the exact sequence
1 a
0 −→ R −→ R −→ R/a1 R −→ 0

yields an exact sequence


1 a
0 −→ Hm
n−1
(R/a1 R) −→ Hm
n
(R) −→ n
Hm (R) −→ 0.
1a
Also, the exact sequence 0 −→ ωR −→ ωR −→ ωR /a1 ωR −→ 0 induces
a1
an exact sequence 0 −→ D(ωR /a1 ωR ) −→ D(ωR ) −→ D(ωR ) −→ 0. By
the definition of canonical module, we have D(ωR ) ∼
= mH n
(R). It follows that
∼ n−1
there is an R-isomorphism D(ωR /a1 ωR ) = Hm (R/a1 R).
Now Hm n−1
(R/a1 R) ∼ n−1
= Hm/a 1R
(R/a1 R) by the Independence Theorem
4.2.1. Also

D(ωR /a1 ωR ) = HomR (ωR /a1 ωR , ER (R/m))


= HomR (ωR /a1 ωR , (0 :ER (R/m) a1 )),

and 10.1.16 shows that there is an R/a1 R-isomorphism

(0 :ER (R/m) a1 ) ∼
= ER/a1 R ((R/a1 R)/(m/a1 R)).
We can therefore conclude that there is an R/a1 R-isomorphism

HomR/a1 R (ωR /a1 ωR , ER/a1 R ((R/a1 R)/(m/a1 R))) ∼ n−1


= Hm/a 1R
(R/a1 R).

Since the local ring R/a1 R has dimension n − 1, this shows that ωR /a1 ωR is
a canonical module for it.
To complete the proof, proceed by induction on j.

12.2.3 Lemma. Suppose that the local ring R is Cohen–Macaulay, and that
ωR exists. Let a ∈ m be a non-zerodivisor on R. Then the R-homomorphism

ξ : HomR (ωR , ωR ) −→ HomR (ωR /aωR , ωR /aωR )

for which ξ(g)(x + aωR ) = g(x) + aωR , for each endomorphism g of ωR and
each x ∈ ωR , is surjective, and Ker ξ = a HomR (ωR , ωR ).
240 Canonical modules

Proof. Denote ωR /aωR by ωR , and let π : ωR −→ ωR denote the natural


epimorphism.
Proposition 12.1.9(ii) shows that a is an ωR -sequence, and 12.2.2 shows that
n−1
depthR ωR = n. Therefore Hm (ωR ) = 0, and it follows from 12.1.20 that
1
ExtR (ωR , ωR ) = 0.
a
Therefore the exact sequence 0 −→ ωR −→ ωR −→ ωR −→ 0 yields the
exact sequence
a
0 −→ HomR (ωR , ωR ) −→ HomR (ωR , ωR ) −→ HomR (ωR , ωR ) −→ 0.
It also yields the exact sequence
a
0 −→ HomR (ωR , ωR ) −→ HomR (ωR , ωR ) −→ HomR (ωR , ωR ).
But a annihilates HomR (ωR , ωR ), and so the latter exact sequence shows that
Hom(π, ωR ) : HomR (ωR , ωR ) −→ HomR (ωR , ωR ) is an isomorphism. It is
straightforward to check that the diagram
ξ
HomR (ωR , ωR ) - HomR (ωR , ωR )

@
@ ∼
= Hom(π,ωR )
Hom(ωR ,π)
@
R
@ ?
HomR (ωR , ωR )

commutes. The sequence


a ξ
0 −→ HomR (ωR , ωR ) −→ HomR (ωR , ωR ) −→ HomR (ωR , ωR ) −→ 0
is therefore exact, and this proves all the claims of the lemma.
12.2.4 Notation. Suppose that ωR exists. We shall use
hR : R −→ HomR (ωR , ωR )
to denote the natural homomorphism for which hR (r) = r IdωR for all r ∈ R.
Note that hR is both an R-module homomorphism and a ring homomor-
phism, and that Im hR is contained in the centre of the ring HomR (ωR , ωR ).
We are going to be interested in the kernel and cokernel of hR , and, in partic-
ular, we shall show that h is an isomorphism when R is Cohen–Macaulay.
12.2.5 Remark. Suppose that ωR exists and that n > 0. By 12.1.11 and
12.1.15,
Ker hr = uR (0)
= {r ∈ R : br = 0 for some ideal b ⊂ R with dim R/b < n}.
12.2 The endomorphism ring 241

12.2.6 Theorem. Suppose that the local ring R is Cohen–Macaulay, and that
ωR exists. Then hR : R −→ HomR (ωR , ωR ) is an isomorphism.

Proof. We argue by induction on n. When n = 0, the ring R is an Artinian


local ring, and therefore complete; also Γm (R) = R; thus 12.1.3(v) shows that

ωR ∼ 0
= HomR (Hm (R), ER (R/m)) ∼
= HomR (R, ER (R/m)) ∼
= ER (R/m).
By 10.2.3(v), for each endomorphism f of ER (R/m), there is a unique rf ∈ R
such that f (x) = rf x for all x ∈ ER (R/m). It follows easily that, for our
endomorphism g of ωR , there is a unique rg ∈ R such that g(x) = rg x for all
x ∈ ωR .
Now suppose that n > 0 and assume, inductively, that the result has been
proved for all Cohen–Macaulay local rings of dimension n − 1 that possess
canonical modules. We already know that hR is injective, because Ker hR =
(0 :R ωR ) = uR (0) by 12.1.15, and this ideal is zero because R is a Cohen–
Macaulay local ring. So let C := Coker φR .
Let a ∈ m be a non-zerodivisor on R, and let R denote the (n − 1)-
dimensional Cohen–Macaulay local ring R/Ra. Now R has a canonical mod-
ule and we can identify ωR/Ra = ωR /aωR (by 12.2.2). The exact sequence
h
0 −→ R −→
R
HomR (ωR , ωR ) −→ C −→ 0 induces the exact sequence
hR ⊗R
R ⊗R R - HomR (ωR , ωR ) ⊗R R - C⊗ R - 0.
R

But it follows from 12.2.3 that there is an R-isomorphism



=
β : HomR (ωR , ωR ) ⊗R R −→ HomR (ωR /aωR , ωR /aωR )

such that β ◦ (hR ⊗ R) maps 1 ⊗ (r + Ra), for r ∈ R, to hR (r + Ra). Since


hR is an isomorphism by the inductive hypothesis, it follows that hR ⊗ R is
an isomorphism, so that C/aC = 0. Since C is finitely generated because ωR
is, we can use Nakayama’s Lemma to deduce that C = 0. Therefore hR is
surjective, and so is an isomorphism.
This completes the inductive step, and the proof.

12.2.7 Theorem. Suppose that ωR exists. Then

hR : R −→ HomR (ωR , ωR )

has Supp(Coker hR ) ⊆ {p ∈ Spec(R) : ht p ≥ 2} . Thus each element of


Coker hR has annihilator of height at least 2.

Proof. Let C := Coker hR . Let p ∈ Spec(R) have ht p ≤ 1; it is enough for


us to show that Cp = 0. Since Supp C ⊆ Supp(HomR (ωR , ωR )) ⊆ Supp ωR ,
242 Canonical modules

we can, and do, assume that p ∈ Supp ωR . Therefore ht p + dim R/p = n, by


12.1.18.
Now argue as in that part of the proof of 12.1.18 where the general case is
treated to find P ∈ Spec(R)  such that P ∩ R = p, dim R/P  = dim R/p
and ht P = ht p. Then the inclusion homomorphism R −→ R  induces a flat
local homomorphism f : Rp −→ R P , and so it is enough for us to show that

Cp ⊗Rp RP = 0. Hence it is enough for us to show that (C ⊗R R)⊗  R 
R P = 0.
Now, by 12.1.3(ii), we have ωR ⊗R R  ∼
= ωR . In view of the natural R-

isomorphism HomR (ωR , ωR ) ⊗R R  −→ Hom  (ωR ⊗R R,
=  ωR ⊗R R)  (see
R
[50, Theorem 7.11]), we can therefore assume for the remainder of this proof
that R is complete. Recall that p ∈ Supp ωR , ht p ≤ 1 and our aim is to show
that Cp = 0.
By 12.1.15 and 12.1.13, we have (0 :R ωR ) = uR (0) and ωR is a canonical
module for R/uR (0). Note that p ⊇ uR (0) and ht p = ht p/uR (0), in view of
12.1.15 and 12.1.18(i). Therefore, we can, and do, assume henceforth in this
proof not only that R is complete, but also that uR (0) = 0. Since ht p ≤ 1 and
uR (0) = 0, the localization Rp is Cohen–Macaulay. By 12.1.18(ii), we have
(ωR )p = ωRp . It therefore follows from 12.2.6 that hRp is an isomorphism.
As the standard Rp -isomorphism

=
ψ : (HomR (ωR , ωR ))p −→ HomRp ((ωR )p , (ωR )p )

is such that ψ ◦ (hR )p = hRp , it follows that Cp = 0. This completes the


proof.

The following proposition, due to Hartshorne and Grothendieck [28, Propo-


sition 2.1 and Remark 2.4.1] will be helpful in the determination of a neces-
sary and sufficient condition (in the case where ωR exists) for hR : R −→
HomR (ωR , ωR ) to be an isomorphism.
t+u
12.2.8 Proposition. Suppose (R, m) is local and S2 , and let 0 = i=1 qi be
a minimal primary decomposition of the zero ideal of R, where both t and u
are positive integers. Then
t t+u
(i) every minimal prime ideal of i=1 qi + j=t+1 qj has height 1;
(ii) if, in addition, R is catenary, then dim R/p = dim R for all p ∈ ass 0.
t t+u √
Proof. Write a := i=1 qi and b = i=t+1 qi , and let qi = pi for all
i = 1, . . . , t + u. Observe that a + b is a proper ideal.
(i) Let p be a minimal prime of a + b. There exists i ∈ {1, . . . , t} such that
p ⊇ pi , and there exists j ∈ {t + 1, . . . , t + u} such that p ⊇ pj . Localize at p:
12.2 The endomorphism ring 243

we obtain, in Rp , that
*
t *
t+u *
t+u
aRp = qi Rp , bRp = qj Rp and 0= qi Rp
i=1 j=t+1 i=1
pi ⊆p pj ⊆p pi ⊆p

are all minimal primary decompositions, that 0 = aRp ∩ bRp and that pRp is
the only associated prime ideal of aRp + bRp , and it is enough, in order for us
to complete the proof of this part, to show that ht pRp = 1. It is thus enough
for us to prove the claim under the additional assumption that p = m, and we
now assume this.
As R is S2 , all associated primes of 0 are minimal; since there is an i ∈
{1, . . . , t} such that m ⊇ pi , and there is a j ∈ {t + 1, . . . , t + u} such that
m ⊇ pj , we have ht m ≥ 1. Thus m ∈ ass 0, m ∈ ass a and m ∈ ass b; on
the other hand, m ∈ ass(a + b). Thus Γm (R) = Γm (R/a) = Γm (R/b) = 0,
whereas Γm (R/(a + b)) = 0. By 3.2.1, there is an exact sequence

0 −→ R −→ (R/a) ⊕ (R/b) −→ R/(a + b) −→ 0.

Application of local cohomology (with respect to m) to this sequence therefore


1
yields that Hm (R) = 0, so that depth R = 1. Since R is S2 , we must have
dim R = ht m = 1.
(ii) Now suppose, in addition, that R is catenary. Suppose that there ex-
ists p ∈ ass 0 such that dim R/p < n, and seek a contradiction. By this
assumption, we may suppose that the numbering, used above, of the primary
components of the zero ideal is such that dim R/pi = n for all i = 1, . . . , t
and dim R/pj < n for all j = t + 1, . . . , t + u. Then, with the above notation,
a + b will have a minimal prime ideal p, and by part (i), ht p = 1. Now p ⊇ pi
for some i ∈ {1, . . . , t} and p ⊇ pj for some j ∈ {t + 1, . . . , t + u}. Then, by
the catenarity, we have

n = dim R/pi = dim R/p + ht p/pi = dim R/p + 1


= dim R/p + ht p/pj = dim R/pj ,

and this is a contradiction.

12.2.9 Exercise. Suppose R is a homomorphic image of a Gorenstein local


ring, so that ωR exists. Show that hR : R −→ HomR (ωR , ωR ) is an isomor-
phism if and only if R satisfies the condition S2 .
Here are some hints for the implication ‘(⇐)’.

(a) Argue by induction on dim R. Use 12.2.6 to establish the claim in the
cases where dim R ≤ 2.
244 Canonical modules

(b) When dim R > 2, use 12.2.8(ii) and 12.2.5 to show that hR is monomo-
rphic, and consider the exact sequence
h
0 −→ R −→
R
HomR (ωR , ωR ) −→ Coker hR −→ 0.

12.2.10 Lemma. Suppose that R is a homomorphic image of a Gorenstein


local ring, and that uR (0) = 0. Then {p ∈ Spec(R) : Rp is S2 }, which we
refer to as the S2 -locus of R, is an open subset of Spec(R) in the Zariski
topology.

Proof. Consider hR : R −→ HomR (ωR , ωR ), and let C := Coker hR , a


finitely generated R-module. Since uR (0) = 0, we see that hR is monomorphic
and Supp ωR = Spec(R), by 12.1.15. Let p ∈ Spec(R). On use of the natu-
ral Rp -isomorphism between (HomR (ωR , ωR ))p and HomRp ((ωR )p , (ωR )p ),
and the fact that (ωR )p ∼
= ωRp (by 12.1.18(ii)), we see from 12.2.9 that Rp is
not S2 if and only if p ∈ Supp C.

The next exercise, which can be solved by use of 12.2.9, establishes a result
due to Y. Aoyama.

12.2.11 Exercise (Y. Aoyama [1, Proposition 2]). Suppose that ωR exists.
Show that hR : R −→ HomR (ωR , ωR ) is an isomorphism if and only if the
 satisfies the condition S2 .
completion R

12.2.12 Exercise. Suppose that ωR exists, that R is S2 and that all the formal
fibres of R are Cohen–Macaulay. Show that hR is an isomorphism.

12.2.13 Exercise. Suppose that R is a homomorphic image of a Gorenstein


local ring and that R is S2 . Show that, for all p ∈ Spec(R), the localization
(ωR )p is a canonical module for Rp .

The following exercise has important significance for algebraic geometry.

12.2.14 Exercise. Suppose that (R, m) is a Cohen–Macaulay local domain


of dimension n that is not Gorenstein but admits a canonical module ω := ωR .

(i) Show that ω is isomorphic to an ideal of R. (Here is a hint: consider


ω ⊗R Q, where Q is the quotient field of R.)
(ii) In the light of (i), regard ω as an ideal of R. Show that
(a) ht ω = 1 and ω is unmixed;
(b) R/ω is a Cohen–Macaulay ring of dimension n − 1; and
(c) R/ω is Gorenstein.
12.3 S2 -ifications 245

(Here are some more hints: for (ii)(a) and (ii)(b), apply local cohomology to
the exact sequence 0 −→ ω −→ R −→ R/ω −→ 0; for (c), use 12.1.20 to
see that the R/ω-module Ext1R (R/ω, ω) is a canonical module for R/ω, and
then apply the functor HomR ( • , ω) to the above-mentioned exact sequence
in order to deduce that R/ω is a canonical module for itself. If you still find
the exercise difficult, you might like to consult [7, Proposition 3.3.18].)

12.2.15 Exercise. Prove the following result of M. P. Murthy [55]: a Cohen–


Macaulay UFD that is a homomorphic image of a Gorenstein local ring must
itself be Gorenstein.

12.3 S2 -ifications
The purpose of this section is to relate the theory of canonical modules to the
concept of S2 -ification of a local ring discussed by Hochster and Huneke in
[39, Discussion (2.3)]. Our starting point is, however, a little more general, and
concerns the theory of generalized ideal transforms, discussed in §2.2.

12.3.1 Notation. In this section, we shall only assume that R is local when
that is explicitly stated. Throughout this section, (Λ, ≤) will denote a (non-
empty) directed partially ordered set, and B = (bα )α∈Λ will denote a system
of ideals of R over Λ in the sense of 2.1.10.
Later in the section, we shall take B to be a particular system of ideals
relevant to the condition S2 .
By a subsystem of B we shall mean a system of ideals C of R such that each
ideal in the family C is a member of B and C can be written as (bα )α∈Θ for
some directed subset Θ of the indexing set Λ.

12.3.2 Proposition. Let M be a finitely generated R-module whose support


0 1
is equal to the whole of Spec(R). Then HB (M ) = HB (M ) = 0 if and only
if Hbα (M ) = Hbα (M ) = 0 for all α ∈ Λ.
0 1

Proof. Since

Γbα (M ) ⊆ ΓB (M ) ⊆ Γbβ (M ),
β∈Λ

we see that HB 0
(M ) = 0 if and only if Hb0α (M ) = 0 for all α ∈ Λ. Suppose
that this is so.
Let S be the set of non-zerodivisors on M . Then, by 2.1.1(ii), S ∩ bα = ∅
for all α ∈ Λ. It follows from 2.2.6(i) that HB 1
(M ) = 0 if and only if ηB,M is
an isomorphism, and Hbα (M ) = 0 (for a given α ∈ Λ) if and only if ηbα ,M
1
246 Canonical modules

is an isomorphism. Moreover, by 2.2.18, the map ηB,M is an isomorphism


⊆ 
if and only if the inclusion monomorphism M −→ α∈Λ (M :S −1 M bα )
is an isomorphism, and, for an α ∈ Λ, the map ηbα ,M is an isomorphism if
⊆ 
and only if the inclusion monomorphism M −→ n∈N (M :S −1 M bnα ) is an
isomorphism. The proof can now be completed easily.
12.3.3 Reminders. We remind the reader of some properties of the natural
ring homomorphism ηB,R : R −→ DB (R) that were established in 2.2.6(i),
2.2.15, 2.2.16 and 2.2.17.

(i) Both the kernel and cokernel of ηB,R are B-torsion.


(ii) Suppose that R is a ring (with identity, but not necessarily commuta-
tive), and let e : R −→ R be a ring homomorphism such that Im e is
contained in the centre of R and, when R is regarded as a left R-module
by means of e, both Ker e and Coker e are B-torsion. If ΓB (R ) = 0,
then the ring R is commutative.
(iii) If now R is a commutative ring (with identity) and e : R −→ R is
a ring homomorphism such that the R-modules Ker e and Coker e are
B-torsion, then the unique R-homomorphism ψ  : R −→ DB (R) such
that the diagram
e
R - R
@
@
ηR
ψ
@
R ?
@
DB (R)

commutes is actually a ring homomorphism. The fact that this diagram


commutes can simply be recorded by the statement that ψ  is an R-
algebra homomorphism.
(iv) In the situation of part (iii), it follows from the formula for ψ  in 2.2.15
that ψ  is injective if and only if ηB,R is injective, and this is the case if
and only if ΓB (R ) = 0. Furthermore, ψ  is an R-algebra isomorphism
if and only if ηB,R is an isomorphism, and this is the case if and only if
ΓB (R ) = HB 1
(R ) = 0.
0
12.3.4 Remark. Let C be a subsystem of B, and suppose that both HB (R)
1
and HB (R) are C-torsion. Then it follows from 12.3.3 that there is a unique R-
algebra homomorphism DB (R) −→ DC (R); also the facts that Ker ηC,R ∼ =
HC0 (R) and Coker ηC,R ∼ = CH 1
(R) are C-torsion, and therefore B-torsion,
show that there is a unique R-algebra homomorphism e : DC (R) → DB (R).
12.3 S2 -ifications 247

The uniqueness aspects of these statements therefore mean that e is an R-


algebra isomorphism.
0 1
Similarly, if d is a member of B and HB (R) and HB (R) are d-torsion, then

=
there is a uniquely determined R-algebra isomorphism Dd (R) −→ DB (R).
12.3.5 Remark. Suppose DB (R) is a finitely generated R-module. Then it
1
follows from 2.2.6(i) that HB (R) is a finitely generated R-module; of course,
HB (R) is finitely generated. Since B is a system of ideals, there exists α ∈ Λ
0
0 1
such that bα annihilates both HB (R) and HB (R). Let C be a subsystem of
B such that bα belongs to the family C. Then it is immediate from 12.3.4 that
there are uniquely determined R-algebra isomorphisms

= ∼
=
Dbα (R) −→ DC (R) −→ DB (R).
12.3.6 Definition. A B-closure of R is a commutative R-algebra θ : R −→
A (with identity) such that
(i) the structural ring homomorphism θ makes A into a finitely generated
R-module;
(ii) both Ker θ and Coker θ are B-torsion; and
(iii) whenever e : R −→ R is a commutative R-algebra such that both Ker e
and Coker e are B-torsion, there is a unique R-algebra homomorphism
ψ  : R −→ A.
We are using the terminology ‘B-closure’ in 12.3.6 because the concept has
similarities with ‘la Z-clôture’ studied in Grothendieck [24, §5.9, §5.10].
12.3.7 Remarks. We can make the following comments about existence of a
B-closure of R.
(i) It is immediate from the definition that a B-closure of R, if it exists, is
uniquely determined up to R-algebra isomorphism.
(ii) Conditions (ii) and (iii) in Definition 12.3.6 require that a B-closure of
R has to be a solution to a certain universal problem. As ηB,R : R −→
DB (R) is a solution to that universal problem (by 12.3.3(iii)), it follows
that there is a B-closure of R if and only if DB (R) is a finitely generated
R-module, and then ηB,R : R −→ DB (R) is the B-closure of R.
(iii) If there exists a B-closure of R, then, by part (ii), the R-module DB (R)
is finitely generated, and so it is immediate from 12.3.5 that there is an
0 1
ideal a in the family B for which HB (R) and HB (R) are a-torsion, and
then, for any such a, and any subsystem C of B such that a belongs to
C, there are uniquely determined R-algebra isomorphisms

= ∼
=
Da (R) −→ DC (R) −→ DB (R).
248 Canonical modules

12.3.8 Notation. Throughout this section, H will denote the system of ideals
of R formed by the set of all ideals of R of height at least 2 (indexed by itself,
partially ordered by reverse inclusion). (Recall that we interpret the height of
the improper ideal R of R as ∞.)
We define the non-S2 locus of R to be {p ∈ Spec(R) : Rp is not S2 }. Also
throughout this section, S will denote the system of ideals of R formed by the
set of all ideals s such that Var(s) is contained in the non-S2 locus of R. Note
that an R-module M is S-torsion if and only if (0 :R m) ∈ S for all m ∈ M .
12.3.9 Definition. We define an S-closure of R, in the sense of 12.3.6, to be
an S2 -ification of R. It follows from 12.3.7(ii) that there is an S2 -ification of R
if and only if DS (R) is a finitely generated R-module, and then ηS,R : R −→
DS (R) provides the S2 -ification of R.
We shall reconcile this definition of S2 -ification with that made by Hochster
and Huneke in [39, Discussion (2.3)] in the case where R is local, uR (0) = 0
and ωR exists.
12.3.10 Theorem. Suppose that (R, m) is local, that uR (0) = 0 and that ωR
exists. Then

(i) S is a subsystem of H, and the ring homomorphism


hR : R −→ HomR (ωR , ωR )
of 12.2.4 has (kernel and) cokernel that are H-torsion;
(ii) the ring HomR (ωR , ωR ) is commutative;

(iii) there is a unique R-algebra isomorphism ψ  : HomR (ωR , ωR ) −→
=

DH (R);
1
(iv) DH (R) and HH (R) are finitely generated R-modules, and the ideal a :=
1
(0 :R HH (R)) belongs to the family H; also Var(a) is the non-S2 locus
of R;
(v) there are unique R-algebra isomorphisms

= ∼
= ∼
=
HomR (ωR , ωR ) −→ Da (R) −→ DS (R) −→ DH (R),
and each of these R-algebras provides the S2 -ification of R.
Proof. (i) Let a be an ideal in S, and suppose that a ⊆ p, where p is a
prime ideal of height less than 2. Then Rp is not S2 , and this is a contradic-
tion because uR (0) = 0. Now hR is injective (since its kernel is uR (0)), and
Coker hR is H-torsion by 12.2.7.
(ii) Note that Im hR is contained in the centre of H := HomR (ωR , ωR );
also ΓH (H) = 0, since otherwise there would be an associated prime of the
12.3 S2 -ifications 249

R-module H having height at least 2, and this is not possible because H is


a faithful R-module satisfying the condition S2 (by 12.1.18). It now follows
from 2.2.16 that H is a commutative ring.
(iii) It is now immediate from 2.2.17 that there is a unique R-algebra homo-
morphism ψ  : HomR (ωR , ωR ) =: H −→ DH (R). Since H is (faithful and)
S2 by 12.1.18, we see that Ha0 (H) = Ha1 (H) = 0 for all ideals a of R with
ht a ≥ 2. Hence HH 0
(H) = HH 1
(H) = 0 by 12.3.2. It therefore follows from

2.2.15 that ψ is an isomorphism.
(iv) Since H is a finitely generated R-module, DH (R) must also be finitely
1
generated, by part (iii). Therefore its epimorphic image HH (R) is finitely gen-
erated, and, since this module is H-torsion, its annihilator must be in H.
By 12.3.5, there is a uniquely determined R-algebra isomorphism

=
Da (R) −→ DH (R).

We show next that Var(a) is the non-S2 locus of R. Let p ∈ Spec(R). If


p ∈ Var(a), then there is an Rp -isomorphism Rp ∼ = Hp , so that Rp is S2
because H is. On the other hand, if p is a minimal member of Var(a), then

0 = (Ha1 (R))p ∼ 1
= HaR p
(Rp )

and depth Rp ≤ 1; since ht p ≥ 2, this means that Rp is not S2 .


(v) By part (iv), the ideal a lies in the family S. Also, S is a subsystem of
H, by part (i). We can now use part (iii) and 12.3.5 to complete the proof.

12.3.11 Corollary. Suppose (R, m) is local, ωR exists and uR (0) = 0. Let


S denote the set of all non-zerodivisors of R, so that R can be considered as
a subring of S −1 R, the total quotient ring of R. Then the S2 -ification of R

(which exists by 12.3.10) is given by the R-algebra A := b∈H (R :S −1 R b).
Furthermore,

(i) A is finitely generated and S2 as an R-module,


(ii) for all a ∈ A, we have ht(R :R a) ≥ 2, and
(iii) there are unique R-algebra isomorphisms
 ∼
=
b∈H (R :S −1 R b) −→ HomR (ωR , ωR )

= ∼
= ∼
=
−→ Da (R) −→ DS (R) −→ DH (R),
1
where a := (0 :R HH (R)).

Proof. Set H := HomR (ωR , ωR ). Since uR (0) = 0, it follows from 12.1.10


that S consists entirely of non-zerodivisors on H, and that S ∩ b = ∅ for
all b ∈ H. We can now use 2.2.18 to see that there is a unique R-algebra
250 Canonical modules

= ⊆
isomorphism A −→ DH (R). It therefore follows from 12.3.10 that R −→ A
provides the S2 -ification of R. All the other claims are now immediate from
12.3.10.
12.3.12 Remark. Suppose that R is local and that uR (0) = 0. Hochster and
Huneke [39, Discussion (2.3)] defined an S2 -ification of R to be a subring A
of the total quotient ring of R such that R ⊆ A , such that A , as R-module, is
finitely generated and S2 , and such that, for all a ∈ A , ht(R :R a) ≥ 2. If ωR
exists, then the S2 -ification A of R found in 12.3.11 is an S2 -ification in the
sense of Hochster and Huneke.
12.3.13 Remark. Suppose (R, m) is local and ωR exists. By 12.1.13, the
ideal uR (0) annihilates ωR and ωR is a canonical module for R/uR (0). It
therefore follows from 12.3.11 that the endomorphism ring HomR (ωR , ωR ) of
ωR is a commutative Noetherian semi-local ring.
12.3.14 Exercise (Goto [2, Example 3.3]). Let
R := K[[X, Y, Z, W ]]/(X, Y ) ∩ (Z, W ),
where K is a field and X, Y, Z, W are independent indeterminates. Show that
the R-homomorphism hR : R −→ HomR (ωR , ωR ) of 12.2.4 is not an iso-
morphism, and that the (commutative) ring HomR (ωR , ωR ) is not local.
12.3.15 Exercise. Suppose that (R, m) is local and ωR exists. Assume that R
 is an integral
is analytically irreducible; that is, assume that the completion R
domain. Show that HomR (ωR , ωR ) is also an analytically irreducible local
ring. (Here is a hint: you might find [50, Theorem 8.15] helpful.)
Note. In [39, Theorem (3.6)], Hochster and Huneke provide, in the case
where (R, m) is local, complete and equidimensional, several conditions equi-
valent to the statement that the semi-local ring HomR (ωR , ωR ) is local.
13

Foundations in the graded case

If our Noetherian ring R is (Z-)graded, and our ideal a is graded, then it is


natural to ask whether the local cohomology modules Hai (R) (i ∈ N0 ), and
Hai (M ) (i ∈ N0 ) for a graded R-module M , also carry structures as graded
R-modules. Some of the realizations of these local cohomology modules that
we have obtained earlier in the book suggest that they should. For instance, if
a1 , . . . , an (where n > 0) denote n homogeneous elements which generate a,
then the C̆ech complex C • (M ) of M with respect to a1 , . . . , an is composed
of graded R-modules and homogeneous homomorphisms, and so Hai (M ) (for
i ∈ N0 ), which, by Theorem 5.1.20, is isomorphic to H i (C(M )• ), inherits a
grading. But is this grading independent of the choice of homogeneous gener-
ators for a?
Additional hopeful evidence is provided by the isomorphism
Hai (M ) ∼
= lim ExtiR (R/an , M )
−→
n∈N

of 1.3.8. For each n ∈ N, since R/an is a finitely generated graded R-module,


ExtiR (R/an , M ) is actually the graded R-module
* ExtiR (R/an , M )
(see [7, pp. 32–33]) with its grading forgotten, and, for n, m ∈ N with n ≥ m,
the natural homomorphism hnm : R/an → R/am is homogeneous, so that the
induced homomorphism
* ExtiR (hnm , M ) : * ExtiR (R/am , M ) −→ * ExtiR (R/an , M )
is homogeneous; hence lim ExtiR (R/an , M ) is graded, and Hai (M ) inherits
−→
n∈N
a grading by virtue of the above isomorphism. But is this grading the same as
that which comes from the approach using the C̆ech complex described in the
preceding paragraph?
252 Foundations in the graded case

One could take another approach to local cohomology in this graded situa-
tion, an approach which, at first sight, seems substantially different from those
described in the preceding two paragraphs. Again suppose that our Noetherian
ring R is (Z-)graded, and our ideal a is graded. The category *C(R) of all
graded R-modules and homogeneous R-homomorphisms is an Abelian cate-
gory which has enough projective objects [7, p. 32] and enough injective ob-
jects [7, 3.6.2]: we can therefore carry out standard techniques of homological
algebra in this category. In particular, the a-torsion functor Γa can be viewed
as a (left exact, additive) functor from *C(R) to itself, and so we can form its
right derived functors *Hai (i ∈ N0 ) on that category. This will produce, for a
graded R-module M , graded local cohomology modules *Hai (M ) (i ∈ N0 ),
which are constructed by the following procedure. Since *C(R) has enough
injectives, we can construct an injective resolution of M in this category, that
is, we can construct an exact sequence

α d0 di
0 −→ M −→ E 0 −→ E 1 −→ · · · −→ E i −→ E i+1 −→ · · ·

in *C(R) in which the E i (i ∈ N0 ) are injective objects in that category, that


is, they are *injective graded R-modules in the terminology of [7, §3.6]; then
we apply the functor Γa to the complex

d0 di
0 −→ E 0 −→ E 1 −→ · · · −→ E i −→ E i+1 −→ · · · ;

the i-th cohomology module of the resulting complex is *Hai (M ) (for each
i ∈ N0 ). But if we forget the grading on *Hai (M ), is the resulting R-module
isomorphic to Hai (M ), and, if so, is the grading on Hai (M ) induced by this
isomorphism the same as that which comes from the approaches in the first
two paragraphs of this chapter?
Our main purpose in this chapter is to reconcile these various approaches.
We think it is desirable that it should be established that they all give (up to
isomorphism in *C(R)) the same object of *C(R). Thus we are going to show
that the questions posed at the ends of the first three paragraphs of this chapter
all have affirmative answers.
In fact, because many modern applications of local cohomology are to multi-
graded rings and modules, we are going to deal in this chapter with a Zn -
graded commutative Noetherian ring and Zn -graded modules over it (note that
n will always denote a positive integer in this context).
13.1 Basic multi-graded commutative algebra 253

13.1 Basic multi-graded commutative algebra


13.1.1 Notation and Terminology. Throughout this chapter, G will denote
a finitely generated, torsion-free Abelian group, written additively. (Thus G is
either 0 or isomorphic to Zn for some n ∈ N.)

When we write that R = g∈G Rg is a G-graded ring, it is to be understood
that the direct decomposition is as Z-module, that Rg Rg ⊆ Rg+g for all
g, g  ∈ G, and that all other ‘graded’ objects related to R (such as R-modules,
ideals of R) are graded by G. Thus, when R is G-graded as above and we
write that M is a graded R-module, it is to be understood that M has a direct

decomposition M = g∈G M as Z-module and that Rg Mg ⊆ Mg+g for
g
all g, g  ∈ G. The elements of g∈G Mg are called the homogeneous elements
of M , and an element of Mg \ {0}, where g  ∈ G, is said to have degree g  .
We shall use ‘deg’ as an abbreviation for degree.

Suppose that R = g∈G Rg is G-graded. An R-homomorphism f : M =
 
g∈G Mg −→ N = g∈G Ng between graded R-modules will be said to
be homogeneous precisely when f (Mg ) ⊆ Ng for all g ∈ G; we shall denote
by *C(R) (or by *C G (R) when it is desirable to specify the grading group G)
the category of all graded R-modules and homogeneous R-homomorphisms
between them.
For g0 ∈ G, we shall denote the g0 -th shift functor by ( • )(g0 ) : *C(R) −→

*C(R): thus, for a graded R-module M = g∈G Mg , we have (M (g0 ))g =
Mg+g0 for all g ∈ G; also, f (g0 ) (M (g0 ))g = f Mg+g0 for each morphism f
in *C(R) and all g ∈ G.
If S is a multiplicatively closed subset of R consisting of non-zero homoge-
neous elements, then it is routine to check that the ring S −1 R is also G-graded,
with g-th component, for g ∈ G, equal to the set of all elements of S −1 R that
can be expressed in the form r/s with s ∈ S, r a homogeneous element of R
and (r = 0 or) r = 0 and deg r − deg s = g. Similarly, for a graded R-module
M , the S −1 R-module S −1 M is also graded.

13.1.2 Remark. A total order ≺ on G is said to be compatible with addition


if, whenever g, g  , h ∈ G with g ≺ g  , we have g + h ≺ g  + h. Note that it
is always possible to put a total order compatible with addition on G: this is
trivial if G = 0, and otherwise G ∼= Zn for some n ∈ N, and, for example, the
lexicographical order on Z is compatible with addition. This lexicographical
n

order ≺ is defined as follows: for a = (a1 , . . . , an ) and b = (b1 , . . . , bn ) ∈


Zn , we set a ≺ b if and only if a = b and, if i is the least integer in {1, . . . , n}
for which ai = bi , then ai < bi .
254 Foundations in the graded case

13.1.3 Elementary Reminders. Assume that R is G-graded, and let M be a


graded R-module.

(i) We have 1R ∈ R0 ; furthermore, R0 is a Noetherian subring of R and


Rg is a finitely generated R0 -module, for each g ∈ G. (See [61, §2.11,
Theorem 21, and §4.3, Theorem 13].)
(ii) We shall say that a submodule of M is graded precisely when it can be
generated by homogeneous elements: see [61, §2.11, Proposition 28].
(iii) Recall from [61, §2.11, Proposition 29] that the sum and intersection of
the members of an arbitrary family of graded submodules of M are again
graded.
(iv) Let N be a graded submodule of M and assume that a is graded. Then
(N :R M ), aN and (N :M a) are all graded: see [61, §2.11, Propositions
30 and 31].
(v) Assume that a is graded. Then a is prime if and only if it is proper and,
whenever r and r are homogeneous elements of R \ a, then rr ∈ R \ a
too: see [61, §2.13, Lemma 13].
(vi) For an arbitrary ideal b of R, we denote by b* the (necessarily graded)
ideal generated by all homogeneous elements of b. Thus b* is the largest
graded ideal of R contained in b. By [61, §2.13, Proposition 33], if b is
prime, then so too is b*.
(vii) We denote by * Spec(R) (or * SpecG (R)) the set of all graded prime
ideals of R.
(viii) The radical of a graded ideal of R is again graded: see [61, §2.13, Propo-
sition 32].

13.1.4 Example. Let R0 be a commutative Noetherian ring, let n ∈ N, and


let R := R0 [X1 , . . . , Xn ], the ring of polynomials over R0 .

(i) We shall often consider R as Zn -graded, with grading given by

R(i1 ,...,in ) = R0 X1i1 . . . Xnin for all (i1 , . . . , in ) ∈ Zn .

Thus R(i1 ,...,in ) = 0 unless ij ≥ 0 for all j = 1, . . . , n. Observe that Xi


is homogeneous of degree (0, . . . , 0, 1, 0, . . . , 0) (where the ‘1’ is in the
i-th spot), for all i = 1, . . . , n.
(ii) Consider the special case in which R0 is a field. Then the graded ideals
of R are just the ideals that can be generated by monomials X1i1 . . . Xnin ,
where (i1 , . . . , in ) ∈ N0 n . The graded prime ideals of R are just the
ideals that can be generated by a subset of the set {X1 , . . . , Xn } of the
variables. It follows that there are only finitely many graded prime ideals
of R.
13.1 Basic multi-graded commutative algebra 255

(iii) Now consider the case where R0 = S is a G-graded ring, with grading

given by S = g∈G Sg . Then R = S[X1 , . . . , Xn ] is (G ⊕ Zn )-graded,
with (g, (i1 , . . . , in ))-th component equal to Sg X1i1 . . . Xnin for all g ∈
G and (i1 , . . . , in ) ∈ Zn .

13.1.5 Example. Let n ∈ N. A simplicial complex on the vertex set V :=


{1, . . . , n} is a set Δ of subsets of V which is closed under passage to subsets,
that is, whenever F ∈ Δ and F  ⊆ F , then F  ∈ Δ too.
Let R0 be a commutative Noetherian ring and let R := R0 [X1 , . . . , Xn ], the
ring of polynomials over R0 in indeterminates X1 , . . . , Xn , considered to be
Zn -graded as in 13.1.4. Let Δ be a simplicial complex on {1, . . . , n}. Let aΔ
be the ideal of R generated by all square-free monomials Xi1 . . . Xit such that
{i1 , . . . , it } ∈ Δ. Observe that, if F  is a subset of V that does not belong to
Δ, then no subset of V that contains F  can belong to Δ. Observe also that aΔ
is a graded ideal of R, as it is generated by homogeneous elements. Therefore
the ring R0 [Δ] := R/aΔ inherits a Zn -grading from R. The ring R0 [Δ] is
called the Stanley–Reisner ring of Δ with respect to R0 .

13.1.6 Lemma. Assume that R is G-graded, let M be a graded R-module,


and let p be a prime ideal of R.

(i) If p ∈ Supp M , then p* ∈ Supp M .


(ii) If p ∈ Ass M , then p is graded; in addition, p = (0 :R m) for some
homogeneous element m ∈ M .
(iii) In particular, if a is a graded ideal of R, then ass a consists of graded
prime ideals.

Proof. The special case of this lemma in which G = Z is proved in [7,


Lemma 1.5.6(b)]. The same proof with minor modifications works in this
G-graded case provided one puts a total order ≺ on G compatible with ad-
dition (see 13.1.2), and uses ≺ instead of the usual order < on Z.

13.1.7 Categorical Reminders. Assume that R is G-graded.

(i) The category *C(R) of all graded R-modules and homogeneous R-hom-
omorphisms is Abelian.
(ii) If f : M −→ N is a morphism in *C(R) (that is, if M and N are
graded R-modules and f is a homogeneous R-homomorphism), then
the ordinary kernel and image of f are graded submodules of M and N
respectively, and act as Ker f and Im f in *C(R).
(iii) A sequence M −→ N −→ P of objects and morphisms in *C(R) is
exact in that category if and only if it is exact in C(R).
256 Foundations in the graded case

(iv) Projective (respectively injective) objects in the category *C(R) will be


referred to as *projective (respectively *injective) graded R-modules.
The category *C(R) ‘has enough projectives’, for if M is a graded R-
module, then M is a homogeneous homomorphic image of a graded

R-module i∈I R(gi ), where (gi )i∈I is a family of elements of G, and

it is easy to see that i∈I R(gi ) is *projective. A graded R-module

of the form i∈I R(gi ) will be called *free. That *C(R) ‘has enough
injectives’ is proved in §13.2 below.

13.1.8 Exercise and Definitions. Assume that R is G-graded, and let M =


 
g∈G Mg and N = g∈G Ng be graded R-modules.

(i) Let g0 ∈ G. We say that an R-homomorphism f : M −→ N is


homogeneous of degree g0 precisely when f (Mg ) ⊆ Ng+g0 for all
g ∈ G. We denote by * HomR (M, N )g0 the set of all homogeneous
R-homomorphisms from M to N of degree g0 . Show that
(a) * HomR (M, N )g0 is an R0 -submodule of HomR (M, N ), and

(b) the sum g ∈G * HomR (M, N )g is direct.
(ii) We set

* HomR (M, N ) := * HomR (M, N )g = * HomR (M, N )g .
g  ∈G g  ∈G

Show that this is an R-submodule of HomR (M, N ), and that the above
direct decomposition turns * HomR (M, N ) into a graded R-module.
Deduce that * HomR ( • , • ) : *C(R) × *C(R) −→ *C(R) is a left
exact, additive functor.
(iii) Show that, if M is finitely generated, then HomR (M, N ) is actually
equal to * HomR (M, N ) with its grading forgotten.
(iv) Let i ∈ N0 . In order to avoid lengthy excursions into the homological
algebra of the category *C(R), we shall define * ExtiR ( • , N ) to be the
i-th right derived functor in *C(R) of * HomR ( • , N ).
Show that, if M is finitely generated, then ExtiR (M, N ) is actually
equal to * ExtiR (M, N ) with its grading forgotten.

13.1.9 Exercise. Suppose R is G-graded, and let M = g∈G Mg and N =

g∈G Ng be graded R-modules. Let (M ⊗R N )g , for a g ∈ G, be the Z-
submodule of M ⊗R N generated by all elements mg1 ⊗ng2 , where g1 , g2 ∈ G
are such that g1 + g2 = g, and mg1 ∈ Mg1 , ng1 ∈ Ng1 . It is clear that

M ⊗R N = g∈G (M ⊗R N )g ; the aim of this exercise is to show that this
sum is direct, and provides M ⊗R N with a structure as a graded R-module.
13.2 *Injective modules 257

(i) Let F be a *free R-module; thus F = i∈I R(gi ), where (gi )i∈I is a

family of elements of G. Show that the sum F ⊗R N = g∈G (F ⊗R N )g
is direct, and that the decomposition

F ⊗R N = (F ⊗R N )g
g∈G

provides a grading for F ⊗R N .


(ii) Consider an exact sequence F1 −→ F0 −→ M −→ 0 in the category
*C(R), where F1 and F0 are *free R-modules. Use part (i) to show that

the sum M ⊗R N = g∈G (M ⊗R N )g is direct; deduce that the de-

composition M ⊗R N = g∈G (M ⊗R N )g provides a grading for
M ⊗R N .
(iii) Deduce that ⊗R can be considered as a functor
⊗R : *C(R) × *C(R) −→ *C(R).

13.2 *Injective modules


One of the main aims of this section is to show that, when R is G-graded,
the category *C(R) ‘has enough injectives’. Our route to this will involve the
concept of *injective envelope.
13.2.1 Definition. Assume that R is G-graded. Let M be a graded submodule
of the graded R-module L.

(i) We say that L is a *essential extension of M precisely when B ∩ M = 0


for every non-zero graded submodule B of L. Such a *essential exten-
sion of M is said to be proper if and only if it is not equal to M .
(ii) We say that L is a *injective envelope (or *injective hull) of M precisely
when L is a *injective R-module and also a *essential extension of M .
13.2.2 Lemma. Assume that R is G-graded. Let M be a graded submodule
of the graded R-module L such that L is a *essential extension of M . Then,
with the gradings forgotten, L is an essential extension of M .
Proof. Let 0 = x ∈ L. Write x as a sum of homogeneous elements x =
xg1 + · · · + xgr , where g1 , . . . , gr are r different members of G, and 0 = xgi
for all i = 1, . . . , r. We show by induction on r that there exists a homogeneous
element a ∈ R such that 0 = ax ∈ M ; this will suffice to prove the lemma.
This is clear when r = 1, since then x is itself homogeneous, and so Rx∩M =
0 because L is a *essential extension of M .
258 Foundations in the graded case

Thus we suppose that r > 1 and make the obvious inductive assumption.
There exists a homogeneous element b ∈ R such that 0 = bxgr ∈ M ; let
x := x − xgr = xg1 + · · · + xgr−1 . If bx = 0, then bx = bxgr is a non-zero
element of M , as wanted; otherwise, 0 = bx , and bx has fewer than r non-
zero homogeneous components; therefore, by the inductive hypothesis, there
is a homogeneous element c ∈ R such that 0 = cbx ∈ M . Then, since b and
c are both homogeneous, 0 = cbx = cbx + cbxgr ∈ M .
13.2.3 Proposition. Assume that R is G-graded, and let M be a graded
R-module. Then M is *injective if and only if M has no proper *essential
extension.
Proof. (⇒) Assume that M is *injective, and let ι : M −→ N be the in-
clusion homomorphism from M into a *essential extension N of M . (Thus
N is graded and ι is homogeneous.) Since M is *injective, there exists a ho-
mogeneous R-homomorphism ϕ : N −→ M such that ϕ ◦ ι = IdM , and

N = M Ker ϕ. Since N is a *essential extension of M , we must have
Ker ϕ = 0, so that N = M .
(⇐) This can be proved just as in the ungraded case: see [7, Proposition
3.2.2].
In the next Theorem 13.2.4, we shall show that each graded R-module M
(where R is G-graded) has a *injective envelope; in particular, this will show
that M can be embedded, by means of a homogeneous R-homomorphism,
into a *injective R-module, and so prove that the category *C(R) has enough
injectives. The proof of 13.2.4 is modelled on Bruns’ and Herzog’s proof in [7,
Theorem 3.6.2] of the particular case in which G = Z.
13.2.4 Theorem. Assume that R is G-graded, and let M be a graded R-
module.
(i) The graded R-module M has a *injective envelope, which, with its grad-
ing forgotten, is an R-submodule of E(M ), the ordinary injective enve-
lope of M .
(ii) Between any two *injective envelopes of M there is a homogeneous
isomorphism which restricts to the identity map on M . We denote by
*E(M ) or *ER (M ) one choice of *injective envelope of M .
(iii) Considered as (ungraded) R-module, *ER (M ) is an essential extension
of M .
Proof. (i) Consider M as an ungraded R-module, and embed M into an un-
graded injective R-module I; for example, we could take I to be E(M ). Let S
denote the set of all R-submodules N of I that carry a grading with respect to
13.2 *Injective modules 259

which M is a graded submodule of N and N is a *essential extension of M .


There is a partial order ≤ on S defined as follows: for N1 , N2 ∈ S, we declare
that N1 ≤ N2 if and only if N1 is a graded submodule of N2 (with the spec-
ified gradings). By Zorn’s Lemma, S has a maximal member E; we note that
E is a graded R-module, a *essential extension of M , and an R-submodule
of I.
Suppose that E is not *injective; it then follows from 13.2.3 that E has a
proper *essential extension E ⊂ E  . By 13.2.2, when the gradings are for-
gotten, E  is an essential extension of E. As I is injective, there exists an
R-homomorphism (possibly not homogeneous) ψ : E  −→ I that extends the

inclusion homomorphism E −→ I. Since Ker ψ ∩ E = 0, and E  is an es-
sential extension of E, we see that Ker ψ = 0 and ψ is a monomorphism. We
can now use the grading on E  and the monomorphism ψ to put a grading on
Im ψ with respect to which E is a homogeneous submodule of Im ψ. Since
E ⊂ E  was a proper *essential extension, it follows that Im ψ is a proper
*essential extension of E, and so a proper *essential extension of M , and we
have a contradiction to the maximality of E in S. This contradiction shows
that E is *injective.
(ii) Suppose that M ⊆ E1 and M ⊆ E2 are two *injective envelopes of M .
Since E2 is *injective, there is a homogeneous R-homomorphism β : E1 −→
E2 whose restriction to M is just the identity map on M . Since Ker β∩M = 0,
we can deduce that Ker β = 0 because M ⊆ E1 is a *essential extension;
therefore β is monomorphic, and Im β is a *injective graded submodule of
E2 , and so has no proper *essential extension, by 13.2.3. But E2 is a *essential
extension of M , and therefore of Im β; hence Im β = E2 , and β : E1 −→ E2
is a homogeneous isomorphism.
(iii) This is immediate from 13.2.2, because *ER (M ) is a *essential exten-
sion of M .

Now that we know that the category *C(R) has enough injectives, we can
make progress with multi-graded local cohomology. We shall build on the
above introduction of the concept of *injective envelope, but not until §14.2.

13.2.5 Lemma. Assume that R is G-graded, and let I, M be graded R-


modules with I *injective.

(i) We have * ExtiR (M, I) = 0 for all i ∈ N.


(ii) If M is finitely generated, then ExtiR (M, I) = 0 for all i ∈ N.

Proof. (i) There is a *projective graded R-module F and a homogeneous


R-epimorphism ψ : F −→ M ; let K := Ker ψ. Since F is *projective,
260 Foundations in the graded case

* ExtiR (F, I) = 0 for all i ∈ N. The exact sequence


ψ
0 −→ K −→ F −→ M −→ 0

therefore induces an exact sequence

0 −→ * HomR (M, I) −→ * HomR (F, I) −→ * HomR (K, I)


−→ * Ext1R (M, I) −→ 0

(in the category *C(R)) and homogeneous isomorphisms

* ExtiR (K, I) ∼
= * Exti+1
R (M, I) for all i ∈ N.

The induced homomorphism * HomR (F, I) −→ * HomR (K, I) is surjec-


tive because I is *injective. Hence * Ext1R (M, I) = 0. But M was an arbitrary
graded R-module; therefore * Ext1R (K, I) = 0, and so * Ext2R (M, I) = 0.
Use induction to complete the proof of (i).
(ii) This is now immediate, since if, for finitely generated M , we forget the
grading on * ExtiR (M, I), then, by 13.1.8(iv), we obtain the ordinary ‘Ext’
module ExtiR (M, I).

Lemma 13.2.5 already enables us to prove the following result, which will
play a key rôle in this chapter.

13.2.6 Proposition. Assume that R is G-graded. Let I be a *injective graded


R-module. Let (Λ, ≤) be a (non-empty) directed partially ordered set, and let
B = (bα )α∈Λ be a system of ideals of R over Λ (as in 2.1.10), with the
property that all ideals in the system are graded. Then I is ΓB -acyclic.
In particular, if the ideal a is graded, then I is Γa -acyclic.

Proof. For each α ∈ Λ, the graded R-module R/bα is finitely generated, and
so ExtiR (R/bα , I) = 0 for all i ∈ N, by 13.2.5. Hence, by 1.3.7,
i
HB (I) ∼
= lim ExtiR (R/bα , I) = 0 for all i ∈ N.
−→
α∈Λ

The following exercise establishes a G-graded analogue of the well-known


Baer Criterion (see [71, Theorem 3.20]) for a module to be injective.

13.2.7 Exercise. Assume that R is G-graded, and let I be a graded R-


module. Show that I is *injective if and only if, for each graded ideal b of
R, each g0 ∈ G and each homogeneous homomorphism f : b −→ I of degree
g0 , there exists a homogeneous homomorphism f  : R −→ I of degree g0
such that f  b = f .
13.3 The *restriction property 261

We made use of the (ungraded) Baer Criterion in our proof of Proposition


2.1.4, and the G-graded version in the above exercise can be used in a similar
way to establish a G-graded version of that proposition. This is addressed in
the next exercise.

13.2.8 Exercise. Assume that R is G-graded and that the ideal a is graded.
Let I be a *injective graded R-module. Show that Γa (I) is *injective.
Now let (Λ, ≤) be a (non-empty) directed partially ordered set, and let
B = (bα )α∈Λ be a system of graded ideals of R over Λ. Show that ΓB (I)
is *injective.

13.2.9 Exercise. Assume that R is G-graded; let I be a graded R-module.


Show that the following conditions are equivalent:

(i) I is *injective;
(ii) * Ext1R (M, I) = 0 for all graded R-modules M ;
(iii) Ext1R (L, I) = 0 for all finitely generated graded R-modules L.

13.3 The *restriction property


13.3.1 Hypotheses for the section. We shall assume, throughout §13.3, that

R = g∈G Rg is G-graded. (Recall that we always assume that R is Noethe-
rian.) Also throughout this section, we shall let G be another finitely gener-

ated torsion-free Abelian group, and we shall let R = g ∈G Rg  be a G -
graded commutative ring. It should be noted that we are not assuming that R
is Noetherian; we shall assume that the reader is familiar with the elementary
properties of graded R -modules expounded in [61, §2.11 and §2.13].

13.3.2 Definition. A sequence (T i )i∈N0 of covariant functors from *C G (R)



to *C G (R ) is said to be a negative connected sequence (respectively, a nega-
tive strongly connected sequence) if the following conditions are satisfied.

(i) Whenever
f g
0 −→ L −→ M −→ N −→ 0

is an exact sequence in *C G (R), there are defined connecting homoge-


neous R -homomorphisms

T i (N ) −→ T i+1 (L) for all i ∈ N0


262 Foundations in the graded case

(in *C G (R )) such that the long sequence
T 0 (f ) T 0 (g)
0 - T 0 (L) - T 0 (M ) - T 0 (N )
1 1
- T 1 (L) T (f )
- T 1 (M ) T (g)
- T 1 (N )

- ··· ···
T i (f ) T i (g)
- T i (L) - T i (M ) - T i (N )

- T i+1 (L) - ···

is a complex (respectively, is exact).


(ii) Whenever

0 - L - M - N - 0

λ μ ν

? ? ?
0 - L - M - N - 0

is a commutative diagram of graded R-modules and homogeneous R-


homomorphisms with exact rows, then there is induced, by λ, μ and ν, a
chain map of the long complex of (i) for the top row into the correspond-
ing long complex for the bottom row.
13.3.3 Examples. Let M be a graded R-module.

(i) Let T : *C G (R) −→ *C G (R ) be an additive covariant functor. Since
the category *C G (R) has enough injectives (by 13.2.4), we can carry out
a standard procedure of homological algebra in that category and form
the right derived functors Ri T (i ∈ N0 ) of T . In more detail, for the
graded R-module M , we can construct an exact sequence
α d0 di
0 −→ M −→ E 0 −→ E 1 −→ · · · −→ E i −→ E i+1 −→ · · ·
in *C G (R) in which the E i (i ∈ N0 ) are *injective graded R-modules;
then we apply the functor T to the complex
d0 di
0 −→ E 0 −→ E 1 −→ · · · −→ E i −→ E i+1 −→ · · · ;
the i-th cohomology module of the resulting complex is the graded R -
module Ri T (M ) (for each i ∈ N0 ).
It should be noted that (Ri T )i∈N0 is a negative strongly connected

sequence of covariant functors from *C G (R) to *C G (R ); furthermore,
if T is left exact, then R0 T is naturally equivalent to T .
13.3 The *restriction property 263

(ii) When a is graded, the a-torsion functor Γa can be viewed as a (left ex-
act, additive) functor from *C G (R) to itself, and so we can form the
connected sequence (*Hai )i∈N0 of its right derived functors on that cat-
egory. For each i ∈ N0 , this leads to a graded local cohomology module
*Hai (M ): one of the aims of this chapter is to show that, if we forget
the grading on *Hai (M ), then the resulting R-module is isomorphic to
Hai (M ).
(iii) Another example is provided by a system of graded ideals of R. Let
(Λ, ≤) be a (non-empty) directed partially ordered set, and let B =
(bα )α∈Λ be a system of graded ideals of R over Λ. The B-torsion func-
tor ΓB can be viewed as a (left exact, additive) functor from *C G (R)
i
to itself, and so we can form the connected sequence (*HB )i∈N0 of its
right derived functors on that category. For each i ∈ N0 , this leads to
i
a graded generalized local cohomology module *HB (M ): in this chap-
i
ter, we shall show that, if we forget the grading on *HB (M ), then the
i
resulting R-module is isomorphic to HB (M ).

13.3.4 Definition. Let (T i )i∈N0 and (U i )i∈N0 be two negative connected se-

quences of covariant functors from *C G (R) to *C G (R ). A homomorphism
Ψ : (T i )i∈N0 −→ (U i )i∈N0 of connected sequences is a family (ψ i )i∈N0
where, for each i ∈ N0 , ψ i : T i −→ U i is a natural transformation of
functors, and which is such that the following condition is satisfied: when-
ever 0 −→ L −→ M −→ N −→ 0 is an exact sequence of graded R-modules
and homogeneous R-homomorphisms, then, for each i ∈ N0 , the diagram

T i (N ) - T i+1 (L)

i i+1
ψN ψL

? ?
U i (N ) - U i+1 (L)

(in which the horizontal maps are the appropriate connecting homomorphisms
arising from the connected sequences) commutes.
Such a homomorphism Ψ = (ψ i )i∈N0 : (T i )i∈N0 −→ (U i )i∈N0 of con-
nected sequences is said to be an isomorphism (of connected sequences) pre-
cisely when ψ i : T i −→ U i is a natural equivalence of functors for each
i ∈ N0 .
   
13.3.5 Exercise. Let T i i∈N0 and U i i∈N0 be negative connected sequen-

ces of covariant functors from *C G (R) to *C G (R ).
264 Foundations in the graded case

(i) Let ψ 0 : T 0 −→ U 0 be a natural transformation of functors. Assume


that
(a) the sequence (T i )i∈N0 is strongly connected, and
(b) T i (I) = 0 for all i ∈ N and *injective graded R-modules I.
Show that there are uniquely determined natural transformations ψ i :
T −→ U i (i ∈ N) such that (ψ i )i∈N0 : (T i )i∈N0 −→ (U i )i∈N0 is a
i

homomorphism of connected sequences.


(ii) Now let ψ 0 : T 0 −→ U 0 be a natural equivalence of functors. Assume
that
(a) the sequence (T i )i∈N0 is strongly connected,
(b) the sequence (U i )i∈N0 is strongly connected, and
(c) for all i ∈ N and *injective graded R-modules I, we have

T i (I) = U i (I) = 0.

Show that there exist uniquely determined natural equivalences ψ i :


T −→ U i (i ∈ N) such that (ψ i )i∈N0 : (T i )i∈N0 −→ (U i )i∈N0 is an
i

isomorphism of connected sequences.

13.3.6 Definition. Let T : C(R) −→ C(R ) be a covariant functor. If

(i) whenever M is a graded R-module, the R -module T (M ) is graded, and


(ii) the gradings in (i) are such that, whenever f : M −→ N is a homo-
geneous homomorphism of graded R-modules, then T (f ) : T (M ) −→
T (N ) is homogeneous,

then we say that T has the *restriction property (with respect to the gradings
specified in (i)).
Of course, the restriction of T to *C G (R) is a functor T : *C G (R) −→
C(R ). (We should, strictly speaking, denote T by T *C G (R) , but we shall
use the shorter notation in the interests of simplicity.) Note that T has the
*restriction property if and only if T can be viewed as a functor from *C G (R)

to *C G (R ).

13.3.7 Definition. Let T, U : C(R) −→ C(R ) be covariant functors. Let


α : T −→ U be a natural transformation of functors (from C(R) to C(R )).
Of course, the restriction of α to *C G (R) provides a natural transformation
α : T −→ U of functors from *C G (R) to C(R ).
Now assume that both T and U have the *restriction property. We say that α
has the *restriction property precisely when, for every graded R-module M ,
the map αM : T (M ) −→ U (M ) is homogeneous.
13.3 The *restriction property 265

Note that this is the case if and only if α can be viewed as a natural trans-

formation α : T −→ U of functors from *C G (R) to *C G (R ).

13.3.8 Remark. Observe that, if, in the notation of 13.3.7, α : T −→ U is


a natural equivalence of functors which has the *restriction property, then the
inverse natural equivalence α−1 : U −→ T also has the *restriction property.
Note also that, if V : C(R) −→ C(R ) is a third covariant functor having
the *restriction property, and both α and a natural transformation β : U −→ V
have the *restriction property, then the composition β ◦ α : T −→ V again has
the *restriction property.

13.3.9 Definition. Let (T i )i∈N0 be a negative strongly connected sequence of


covariant functors from C(R) to C(R ). We can regard (T i )i∈N0 as a negative
strongly connected sequence of covariant functors from *C G (R) to C(R ).
We shall say that (T i )i∈N0 has the *restriction property precisely when

(i) T i has the *restriction property for all i ∈ N0 ; and


(ii) whenever 0 −→ L −→ M −→ N −→ 0 is an exact sequence in the
category *C G (R), the connecting homomorphisms T i (N ) −→ T i+1 (L)
(i ∈ N0 ) (which exist by virtue of the fact that (T i )i∈N0 is a connected
sequence) are all homogeneous.

Note that this is the case if and only if (T i )i∈N0 can be viewed as a negative

strongly connected sequence of covariant functors from *C G (R) to *C G (R ).

13.3.10 Definition. Let (T i )i∈N0 and (U i )i∈N0 be negative strongly connec-


ted sequences of covariant functors from C(R) to C(R ); assume that both
these sequences have the *restriction property of 13.3.9. Let Ψ := (ψ i )i∈N0 :
(T i )i∈N0 −→ (U i )i∈N0 be a homomorphism of connected sequences. We say
that Ψ has the *restriction property if and only if, for all i ∈ N0 , the natural
transformation ψ i has the *restriction property of 13.3.7.

13.3.11 Proposition. Let V, M be graded R-modules with V finitely gen-


erated. By 13.1.8(iv), for each i ∈ N0 , ExtiR (V, M ) is actually the graded
R-module * ExtiR (V, M ) with its grading forgotten; hence each ExtiR (V, M )
has a natural structure as a graded R-module, and these structures are such
that, if V  is a second finitely generated graded R-module and h : V −→ V 
is a homogeneous homomorphism, then ExtiR (h, M ) : ExtiR (V  , M ) −→
ExtiR (V, M ) is also homogeneous.
Moreover, with respect to these
 graded R-module
 structures, the negative
strongly connected sequence ExtiR (V, • ) i∈N0 of covariant functors (from
C(R) to itself) has the *restriction property (see 13.3.9).
266 Foundations in the graded case

Proof. Only the claim in the second paragraph requires proof. Let F• be a
*free resolution of V in *C(R) by *free graded R-modules of finite rank.
An exact sequence 0 −→ L −→ M  −→ N  −→ 0 in C(R) induces a
sequence

0 −→ HomR (F• , L ) −→ HomR (F• , M  ) −→ HomR (F• , N  ) −→ 0

of complexes of R-modules and chain maps of such complexes such that, for
each i ∈ N0 , the sequence

0 −→ HomR (Fi , L ) −→ HomR (Fi , M  ) −→ HomR (Fi , N  ) −→ 0

(where Fi denotes the i-th term of F• ) is exact. It is straightforward to use


the long exact sequences
 of cohomology modules
 induced by such sequences
of complexes to turn H i (HomR (F• , • )) i∈N0 into a negative strongly con-
nected sequence of covariant functors from C(R) to itself, and a standard ap-
plication of Exercise
 1.3.4(ii) then
 shows that this negative connected sequence
is isomorphic to ExtiR (V, • ) i∈N .
0
In particular, when we apply this to a morphism f : M −→ M  in *C(R),
and to an exact sequence 0 −→ L −→ M −→ N −→ 0 of graded R-
modules and homogeneous R-homomorphisms, there result, for each i ∈ N0 ,
commutative diagrams

H i (HomR (F• , M )) - H i (HomR (F• , M  ))


= ∼
=

? ?
ExtiR (V,f )
ExtiR (V, M ) - Exti (V, M  )
R

and
H i (HomR (F• , N )) - H i+1 (HomR (F• , L))


= ∼
=

? ?
ExtiR (V, N ) - Exti+1 (V, L) .
R

In these circumstances, HomR (F• , M ), HomR (F• , M  ), HomR (F• , N ) and


HomR (F• , L) are complexes of graded R-modules and homogeneous R-hom-
omorphisms, so that all their cohomology modules are graded, and it is easy
to check that the top horizontal homomorphisms in the above two commu-
tative diagrams are homogeneous. Furthermore, the graded module structures
13.3 The *restriction property 267

on ExtiR (V, M ), ExtiR (V, M  ), ExtiR (V, N ) and Exti+1


R (V, L) induced by the
vertical isomorphisms in the diagrams are precisely the graded module struc-
tures referred to in the first paragraph of the statement of the proposition. All
the claims follow easily from these observations.
13.3.12 Remark. Let (Ω, ≤) be a (non-empty) directed partially ordered set,
and let (Wω )ω∈Ω be a direct system of graded R-modules and homogeneous
R-homomorphisms over Ω, with constituent R-homomorphisms hω ν : Wν →
Wω (for each (ω, ν) ∈ Ω × Ω with ω ≥ ν). We can forget the gradings and
calculate the ordinary direct limit
W∞ := lim Wω ,
−→
ω∈Ω

in C(R): let hω : Wω −→ W∞ be the canonical map (for each ω ∈ Ω). It


should be noted that W∞ inherits a natural grading, for which all the hω (ω ∈
Ω) are homogeneous, and that W∞ actually acts as the direct limit of the direct
system (Wω )ω∈Ω in the category *C(R).
13.3.13 Examples. The following examples will be important for us.
(i) Let (Λ, ≤) be a (non-empty) directed partially ordered set, and let B =
(bα )α∈Λ be a system of graded ideals of R over Λ. For α, β ∈ Λ with
α ≥ β, the natural homomorphism hα β : R/bα −→ R/bβ is homo-
geneous. It follows from 13.3.11 and 13.3.12 that the negative strongly
connected sequence of covariant functors
 
lim ExtiR (R/bα , • )
−→
α∈Λ i∈N0

from C(R) to itself has the *restriction property.


(ii) In particular, when the ideal a is graded, the negative strongly connected
sequence of covariant functors
 
lim ExtiR (R/an , • )
−→
n∈N i∈N0

from C(R) to itself has the *restriction property.


(iii) Assume that a is graded, and let a1 , . . . , an (where n > 0) denote n
homogeneous elements which generate a. Let C • denote the C̆ech com-
plex of R with respect to a1 , . . . , an , as in 5.1.5. Then it is straightfor-
ward to check that the negative strongly connected sequence of covariant
functors (H i (( • ) ⊗R C • ))i∈N0 from C(R) to itself has the *restriction
property.
268 Foundations in the graded case

13.3.14 Exercise. Let (Λ, ≤) be a (non-empty) directed partially ordered


set, and let B = (bα )α∈Λ be a system of graded ideals of R over Λ. Show
that the functor DB of 2.2.3 has the *restriction property, and that the natural
transformation ηB of 2.2.6(i) has the *restriction property.
Deduce that, when the ideal a is graded, the functor Da of 2.2.1 has the
*restriction property, and the natural transformation ηa of 2.2.6(i) has the *rest-
riction property.

13.3.15 Theorem. Let (T i )i∈N0 be a negative strongly connected sequence of


covariant additive functors from C(R) to C(R ). Suppose that, for each graded
R-module M , a grading is given on T 0 (M ), and that, with respect to these
gradings, T 0 has the *restriction property. Suppose further that T i (I) = 0 for
all i ∈ N whenever I is a *injective graded R-module.
Then there is exactly one choice of gradings on the T i (M ) (i ∈ N, M a
graded R-module) with respect to which (T i )i∈N0 has the *restriction prop-
erty.

Furthermore, if we denote by *T 0 : *C G (R) −→ *C G (R ) the (neces-
sarily left exact) functor induced by T 0 (see 13.3.6), then there is a unique

=
isomorphism (ψ i )i∈N0 : (T i )i∈N0 −→ (Ri (*T 0 ))i∈N0 of negative connected

sequences of covariant functors from *C G (R) to *C G (R ) for which ψ 0 :
T 0 = *T 0 −→ R0 (*T 0 ) is the canonical natural equivalence.

Proof. We first show, by induction on t ∈ N0 , that, if (T i )i∈N0 has the


*restriction property, then the gradings on the R -modules T t (M ) (M a graded
R-module) are uniquely determined by the gradings on the R -modules T 0 (L)
(L a graded R-module). This statement is certainly true when t = 0, and so we
suppose now that t > 0 and the statement is true for smaller values of t.
Let M be an arbitrary graded R-module. Since the category *C G (R) has
enough injective objects, there is an exact sequence
g
0 −→ M −→ I −→ N −→ 0

in *C G (R) with I a *injective graded R-module. The hypotheses and the as-
sumption that (T i )i∈N0 has the *restriction property therefore lead to an exact
sequence
T t−1 (g)
T t−1 (I) - T t−1 (N ) - T t (M ) - 0

of graded R -modules and homogeneous homomorphisms. Since, by the in-


ductive hypothesis, the grading on T t−1 (N ) is uniquely determined by our
assumptions, it follows that the grading on T t (M ) is similarly uniquely deter-
mined.
13.3 The *restriction property 269

Thus there is at most one choice of gradings on the T i (M ) (i ∈ N, M a


graded R-module) with respect to which (T i )i∈N0 has the *restriction property.
We still have to show that there is one such choice.
Since (T i )i∈N0 is a negative strongly connected sequence of covariant func-
tors from C(R) to C(R ), it is automatic that T 0 is left exact. Therefore *T 0 :

*C G (R) −→ *C G (R ) is left exact (and additive), (Ri (*T 0 ))i∈N0 is a nega-

tive strongly connected sequence of functors from *C G (R) to *C G (R ), and
there is a canonical natural equivalence ψ 0 : T 0 = *T 0 −→ R0 (*T 0 ). Note
0
that ψM is homogeneous for each graded R-module M .
Now forget, just temporarily, the grading on R : then (Ri (*T 0 ))i∈N0 and
(T i )i∈N0 are negative strongly connected sequences of covariant functors
from *C G (R) to C(R ) and ψ 0 : T 0 −→ R0 (*T 0 ) is a natural equivalence
(of functors from *C G (R) to C(R )). Moreover, T i (I) = Ri (*T 0 )(I) = 0 for
all i ∈ N and all *injective graded R-modules I.
At this point, regard, again temporarily, R as trivially graded, that is, graded
by 0: we can use 13.3.5(ii) to see that there exist uniquely determined nat-
ural equivalences ψ i : T i −→ Ri (*T 0 ) (i ∈ N) such that (ψ i )i∈N0 :
(T i )i∈N0 −→ (Ri (*T 0 ))i∈N0 is an isomorphism of connected sequences
(of functors from *C G (R) to C(R ) = *C 0 (R )).
Now remember the original G -gradings on R and its graded modules: for
each graded R-module M and each i ∈ N, the R -module Ri (*T 0 )(M ) is
graded. For such i and M , we can define a grading on T i (M ) in such a way
that the R -isomorphism

=
i
ψM : T i (M ) −→ Ri (*T 0 )(M )

is homogeneous. With these gradings in place, all the remaining claims of the
theorem follow routinely.

13.3.16 Corollary. Let T be a left exact additive covariant functor from C(R)
to C(R ). Suppose that, for each graded R-module M , a grading is given on
T (M ), and that, with respect to these gradings, T has the *restriction property.
Suppose further that each *injective graded R-module I is T -acyclic, that is,
Ri T (I) = 0 for all i ∈ N.
Then there is exactly one choice of gradings on the Ri T (M ) (i ∈ N, M a
graded R-module) with respect to which (Ri T )i∈N0 has the *restriction prop-
erty.

Furthermore, if we denote by *T : *C G (R) −→ *C G (R ) the functor in-
duced by T , then there is a unique isomorphism

=
(ψ i )i∈N0 : ((Ri T ) )i∈N0 −→ (Ri (*T ))i∈N0
270 Foundations in the graded case

of negative connected sequences of functors from *C G (R) to *C G (R ) for
which ψ 0 : T = *T −→ *T is the identity. 
13.3.17 Theorem. Let
(T i )i∈N0 and (U i )i∈N0
be negative strongly connected sequences of covariant additive functors from
C(R) to C(R ) which have the *restriction property of 13.3.9, and suppose
that T i (I) = 0 for all i ∈ N whenever I is a *injective graded R-module.
Let Ψ := (ψ i )i∈N0 : (T i )i∈N0 −→ (U i )i∈N0 be a homomorphism of con-
nected sequences. Suppose that ψ 0 has the *restriction property of 13.3.7.
Then it is automatic that ψ i has the *restriction property for all i ∈ N0 , so that
Ψ has the *restriction property of 13.3.10.
Furthermore, (ψ i )i∈N0 : (T i )i∈N0 −→ (U i )i∈N0 is the unique extension
of the natural transformation ψ 0 : T 0 −→ U 0 (of functors from *C G (R)

to *C G (R )) to a homomorphism of connected sequences of functors from

*C G (R) to *C G (R ).
Proof. We prove by induction on t that ψ t has the *restriction property for
all t ∈ N0 . We know that ψ 0 has the *restriction property, and so we suppose
now that t > 0 and that ψ i has the *restriction property for all i < t.
Let M be an arbitrary graded R-module. Since the category *C G (R) has
enough injective objects, there is an exact sequence
g
0 −→ M −→ I −→ N −→ 0
in *C G (R) with I a *injective graded R-module. The hypotheses therefore
lead to a commutative diagram
T t−1 (g)
T t−1 (I) - T t−1 (N ) - T t (M ) - 0

ψIt−1 t−1
ψN t
ψM

? ? ?
U t−1 (g)
U t−1 (I) - U t−1 (N ) - U t (M ) - U t (I)

of graded R -modules and homogeneous homomorphisms with exact rows. It


follows from the inductive hypothesis, and the fact that (T i )i∈N0 and (U i )i∈N0
have the *restriction property, that all the homomorphisms in the above dia-
t
gram, with the possible exception of ψM , are homogeneous. It is immediate
t
from the fact that the top row is exact that ψM must be homogeneous too.
The claim in the final paragraph is immediate from 13.3.5, 13.3.7 and 13.3.9.
13.4 The reconciliation 271

13.4 The reconciliation


13.4.1 Hypotheses for the section. We shall assume, throughout §13.4, that

R = g∈G Rg is G-graded, and that the ideal a is graded.
Also, we shall assume that (Λ, ≤) is a (non-empty) directed partially ordered
set, and that B = (bα )α∈Λ is a system of graded ideals of R over Λ.
13.4.2 Theorem. There is a unique choice of gradings on the HB i
(M ) (i ∈
N, M a graded R-module) with respect to which (HB )i∈N0 has the *restriction
i

property; furthermore, when these gradings are imposed, there is a unique


isomorphism

(φi )i∈N0 : (HB
=
i
)i∈N0 −→ (*HB
i
)i∈N0
of connected sequences of covariant functors from *C(R) to itself for which φ0
is the identity.
Proof. By 13.2.6, each *injective graded R-module I is ΓB -acyclic; also, it
is clear that ΓB : C(R) −→ C(R) has the *restriction property. The result
therefore follows from 13.3.16.
We record separately a very important special case of 13.4.2.
13.4.3 Corollary. There is a unique choice of gradings on the Hai (M ) (i ∈ N,
M a graded R-module) with respect to which (Hai )i∈N0 has the *restriction
property; furthermore, when these gradings are imposed, there is a unique
isomorphism

(ψi )i∈N : (H i )i∈N −→ (*H i )i∈N
=
0 a 0 a 0

of connected sequences of covariant functors from *C(R) to itself for which


ψ0 is the identity. 
13.4.4 Remark. For a graded R-module M , we do, of course, grade the
HBi
(M ) (i ∈ N) and the Hai (M ) (i ∈ N) using the unique gradings for which
the conclusions of 13.4.2 and 13.4.3 are satisfied.
Thus, during the rest of the book, whenever our current assumptions about
gradings (that R is G-graded, and a and all the ideals in B are graded) are in
force, we shall, without further ado, utilize the gradings on the Hai (M ) and
HBi
(M ) whenever M is a graded R-module; for each g ∈ G, we shall denote
the g-th component of the graded R-module Hai (M ) (respectively HB i
(M ))
i i
by Ha (M )g (respectively HB (M )g ). (In many of the examples that we shall
consider, G will actually be the additive group Z of integers.) We shall also
make much use of the fact that, whenever 0 −→ L −→ M −→ N −→ 0 is an
exact sequence of graded R-modules and homogeneous homomorphisms, not
272 Foundations in the graded case

only are all the terms in the induced long exact sequence of local cohomology
modules with respect to a (respectively generalized local cohomology mod-
ules with respect to B) graded, but all the homomorphisms in this sequence,
including the connecting homomorphisms, are homogeneous. These are really
crucial facts about graded local cohomology.

In the introduction to this chapter, we posed three questions about various


possible approaches to the construction of gradings on local cohomology mod-
ules. The next theorem will enable us to answer those questions.

13.4.5 Theorem. Let (T i )i∈N0 be a negative strongly connected sequence of


covariant additive functors from C(R) to itself. Suppose that

=
Ω = (ω i )i∈N0 : (T i )i∈N0 −→ (HB
i
)i∈N0

is an isomorphism of connected sequences (from C(R) to itself). Suppose that,


for each graded R-module M , a grading is given on T 0 (M ), and that, with
respect to these gradings, T 0 has the *restriction property. Suppose also that
ω 0 has the *restriction property.
Then there is exactly one choice of gradings on the T i (M ) (i ∈ N, M a
graded R-module) with respect to which (T i )i∈N0 has the *restriction prop-
erty; with these gradings (and those of 13.4.4) imposed, Ω has the *restriction
property of 13.3.10.

Proof. By 13.2.6, we have T i (I) = HB i


(I) = 0 for all i ∈ N whenever I
is a *injective graded R-module. Therefore, by 13.3.15, there is exactly one
choice of gradings on the T i (M ) (i ∈ N, M a graded R-module) with respect
to which (T i )i∈N0 has the *restriction property.
i
By 13.4.4, the connected sequence (HB )i∈N0 has the *restriction property.
The final claim therefore follows from 13.3.17.

13.4.6 Remarks. Here, we finally answer the three questions posed in the
introduction to this chapter.

(i) We apply Theorem 13.4.5 to the isomorphism


 
 i  ∼
=
i
ΦB = φB i∈N0 : lim ExtR (R/bα , • ) −→ HB
i
−→ i∈N0
α∈Λ i∈N0

of negative strongly connected sequences of functors (from C(R) to it-


self) of 2.2.2. It is clear from the definition of φ0B in 1.2.11(ii) that it has
the *restriction property. Hence Theorem 13.4.5 can be applied: there is
exactly one choice of gradings (and that was described in 13.3.13(i)) on
13.4 The reconciliation 273

the lim ExtiR (R/bα , M ) (i ∈ N, M a graded R-module) with respect


−→
α∈Λ
to which  
lim ExtiR (R/bα , • )
−→
α∈Λ i∈N0

has the *restriction property; with respect to these gradings, ΦB has the
*restriction property.
(ii) The following special case of part (i) is worthy of separate mention. We
can apply Theorem 13.4.5 to the isomorphism
 

=
Φa = φia : lim ExtiR (R/an , • ) −→ Hai
i∈N0 −→ i∈N0
n∈N i∈N0

of negative strongly connected sequences of functors (from C(R) to it-


self) of 1.3.8. It is clear from the definition of φ0a in 1.2.11(iii) that it has
the *restriction property. Hence Theorem 13.4.5 can be applied: there is
exactly one choice of gradings (and that was described in 13.3.13(ii)) on
the lim ExtiR (R/an , M ) (i ∈ N, M a graded R-module) with respect to
−→
n∈N
which  
lim ExtiR (R/an , • )
−→
n∈N i∈N0

has the *restriction property; with these gradings, Φa has the *restriction
property.
(iii) Let a1 , . . . , an (where n > 0) denote n homogeneous elements which
generate a. Let C • denote the C̆ech complex of R with respect to the
elements a1 , . . . , an , as in 5.1.5. It is easy to check that Theorem 13.4.5
can be applied to the isomorphism

((γ i )−1 )i∈N0 : (H i (( • ) ⊗R C • ))i∈N0 −→ (Hai )i∈N0
=

of negative strongly connected sequences of functors (from C(R) to it-


self) of 5.1.20. We conclude that there is exactly one choice of gradings
(and that was described in 13.3.13(iii)) on the H i (M ⊗R C • ) (i ∈ N,
M a graded R-module) with respect to which (H i (( • ) ⊗R C • ))i∈N0
has the *restriction property; with these gradings, ((γ i )−1 )i∈N0 has the
*restriction property.
(iv) In short, if we use the isomorphism of connected sequences
 

=
Φa = φia : lim ExtiR (R/an , • ) −→ Hai
i∈N0 −→ i∈N0
n∈N i∈N0
274 Foundations in the graded case

of 1.3.8 to define gradings on the Hai (M ) (i ∈ N) (for a graded R-


module M ), or if we use the isomorphism of connected sequences

((γ i )−1 )i∈N0 : (H i (( • ) ⊗R C • ))i∈N0 −→ (Hai )i∈N0
=

of 5.1.20, with any choice of homogeneous generators for a, to define


gradings on the Hai (M ) (i ∈ N), the resulting gradings are always
the same, and are precisely those with respect to which (Hai )i∈N0 has
the *restriction property: see 13.4.3. (Note also that it would make no
difference if we used C(M )• , the C̆ech complex of M with respect to
a1 , . . . , an , instead of M ⊗R C • : the constituent isomorphisms in the iso-
morphism of complexes of 5.1.11 are all homogeneous.) Furthermore,
as, when these gradings are imposed, there is a unique isomorphism

(ψi )i∈N0 : (Hai )i∈N0 −→ (*Hai )i∈N0 of connected sequences of co-
=

variant functors from *C(R) to itself for which ψ0 is the identity (see
again 13.4.3), all the questions posed in the introduction to this chapter
have been answered affirmatively!

13.5 Some examples and applications


In our discussion of examples, we shall sometimes wish to change gradings
and use a ‘less fine’ or ‘coarser’ grading than one that occurs naturally. In this
connection, the notation and terminology introduced in the following definition
will be helpful.
13.5.1 Definition. Let φ : G −→ H be a homomorphism of finitely gen-

erated torsion-free Abelian groups, and suppose that R = g∈G Rg is G-
graded.

For each h ∈ H, let Rhφ := g∈φ−1 ({h}) Rg . Then
⎛ ⎞
 φ  
Rφ := Rh = ⎝ Rg ⎠
h∈H h∈H g∈φ−1 ({h})

provides an H-grading on R, and we denote R by Rφ when considering it as


an H-graded ring in this manner.
 φ
Furthermore, for each G-graded R-module M = g∈G Mg , let Mh :=
 φ
 φ φ
g∈φ−1 ({h}) Mg and M := h∈H Mh ; then M is an H-graded Rφ -
module. Also, if f : M −→ N is a G-homogeneous homomorphism of G-
graded R-modules, then the same map f becomes an H-homogeneous homo-
morphism of H-graded Rφ -modules f φ : M φ −→ N φ .
13.5 Some examples and applications 275

In this way, ( • )φ becomes an exact additive covariant functor from *C G (R)


to *C H (R). We refer to it as the φ-coarsening functor, and we shall consider the
H-gradings constructed in this way as a coarsening of the given G-gradings.
The following lemma will help us to exploit coarsenings of graded local
cohomology modules.
13.5.2 Lemma. Let the situation be as in 13.5.1, so that R is G-graded and
φ : G −→ H is a homomorphism of finitely generated torsion-free Abelian
groups. Assume that a is graded. Then

Hai ( • )φ and Hai φ (( • )φ )


i∈N0 i∈N0

are isomorphic connected sequences of functors from *C G (R) to *C H (R).


Consequently, for each G-graded R-module M and each i ∈ N0 , there is
an H-homogeneous isomorphism Hai (M )φ ∼ = Hai φ (M φ ).
Proof. Because ( • )φ is an exact additive covariant functor from *C G (R) to
*C H (R), one easily sees that Hai ( • )φ and Hai φ (( • )φ ) are both
i∈N0 i∈N0
negative strongly connected sequences of covariant functors from *C G (R) to
*C H (R). Now Γa ( • )φ and Γaφ (( • )φ ) are the same functor. Also, whenever I
is a *injective G-graded R-module, we can see that Hai (I)φ = 0 = Hai φ (I φ )
for all i > 0: the first of these claims is immediate, whereas the second follows
from 13.4.3 because, without the H-grading, Hai φ (I φ ) ∼= Hai (I), and I is Γa -
acyclic (by 13.2.6). The result therefore follows from 13.3.5.
An illustration of the idea of coarsening occurs with our first example in
this section; in this example, we show that the C̆ech complex approach can
quickly lead to important information about graded local cohomology over a
polynomial ring.

13.5.3 Example. Let S = g∈G Sg be a G-graded commutative Noetherian
ring, let n ∈ N, and let R := S[X1 , . . . , Xn ], the ring of polynomials over S,
graded by G ⊕ Zn as in 13.1.4(iii). The reader should note the special case in
which S = S0 is trivially graded, in which case the grading on R is (essen-
tially) as described in 13.1.4(i), with S0 playing the rôle of R0 .
We propose to use the C̆ech complex of R with respect to the homogeneous
elements X1 , . . . , Xn ,
d0 di dn−1
C • : 0 −→ C 0 −→ C 1 −→ · · · −→ C i −→ C i+1 −→ · · · −→ C n −→ 0,
in conjunction with 13.4.6 above, to calculate the graded R-module
n n
H(X 1 ,...,Xn )
(S[X1 , . . . , Xn ]) = H(X 1 ,...,Xn )
(R).
276 Foundations in the graded case

There is a homogeneous R-isomorphism between this local cohomology mod-


ule and
n n−1  
RX1 ...Xn / t=1 d RX1 ...Xt−1 Xt+1 ...Xn .
 
Now RX1 ...Xn is a free S-module with X1i1 . . . Xnin (i ,...,i )∈Zn as a base;
n−1
 1 n 
for each t = 1, . . . , n, its R-submodule
 i1 d RX 1 ...X t−1 t+1 ...Xn
X is again
free as S-module, with base X1 . . . Xnin (i1 ,...,in )∈Zn , it ≥0 . Use −N to de-
note the set {n ∈ Z : n < 0}. It follows that the graded R-module
n n−1  
RX1 ...Xn / t=1 d RX1 ...Xt−1 Xt+1 ...Xn

can be considered as a free S-module with base


 i1 
X1 . . . Xnin (i ,...,i )∈(−N)n
1 n

and R-module structure such that, for (i1 , . . . , in ) ∈ (−N)n and t ∈ N with
1 ≤ t ≤ n,
 it−1 it +1 it+1
i1 in X1i1 . . . Xt−1 Xt Xt+1 . . . Xnin if it < −1,
Xt (X1 . . . Xn ) =
0 if it = −1.

The (G ⊕ Zn )-grading is such that, for g ∈ G and sg ∈ Sg \ {0},

deg(sg X1i1 . . . Xnin ) = (g, (i1 , . . . , in )) for all (i1 , . . . , in ) ∈ (−N)n .

We refer to this graded R-module as the module of inverse polynomials in


X1 , . . . , Xn over S, and denote it by S[X1− , . . . , Xn− ]. To summarize, the
C̆ech complex approach to the calculation of graded local cohomology mod-
ules quickly yields a (G ⊕ Zn )-homogeneous S[X1 , . . . , Xn ]-isomorphism
n
H(X 1 ,...,Xn )
(S[X1 , . . . , Xn ]) ∼
= S[X1− , . . . , Xn− ].

We could also regard R = S[X1 , . . . , Xn ] as (G ⊕ Z)-graded, with



R(g,m) = R(g,(i1 ,...,in )) for all (g, m) ∈ G ⊕ Z.
(i1 ,...,in )∈N0 n
i1 +···+in =m

However, with this grading, our polynomial ring is just the result Rφ of apply-
ing the φ-coarsening functor of 13.5.1 to R, where φ : G ⊕ Zn −→ G ⊕ Z is
the Abelian group homomorphism for which

φ((g, (i1 , . . . , in ))) = (g, i1 + · · · + in ) for all (g, (i1 , . . . , in )) ∈ G ⊕ Zn .


13.5 Some examples and applications 277

Thus, in view of 13.5.2, there are (G ⊕ Z)-homogeneous R-isomorphisms


φ
n φ ∼ n
H(X 1 ,...,Xn )
φ (S[X1 , . . . , Xn ] ) = H(X 1 ,...,Xn )
(S[X1 , . . . , Xn ])
 φ

= S[X1− , . . . , Xn− ] ,

where, in the right-hand module, for (i1 , . . . , in ) ∈ (−N)n and sg ∈ Sg \ {0}


(where g ∈ G), the element sg X1i1 . . . Xnin has degree (g, i1 + · · · + in ).
The reader should note the special case of the above in which S is trivially
graded, so that S = S0 = R0 : when R is considered to be Zn -graded as
described in 13.1.4(i), the isomorphism
n
H(X 1 ,...,Xn )
(R0 [X1 , . . . , Xn ]) ∼
= R0 [X1− , . . . , Xn− ]

is Zn -homogeneous and, in the module of inverse polynomials,

deg(X1i1 . . . Xnin ) = (i1 , . . . , in ) for all (i1 , . . . , in ) ∈ (−N)n ;

when R is considered to be Z-graded, with deg Xi = 1 for all i = 1, . . . , n,


then the isomorphism
n
H(X 1 ,...,Xn )
(R0 [X1 , . . . , Xn ]) ∼
= R0 [X1− , . . . , Xn− ]

is Z-homogeneous and, in the module of inverse polynomials, X1i1 . . . Xnin has


degree i1 + · · · + in for all (i1 , . . . , in ) ∈ (−N)n .

13.5.4 Exercise. Assume that R is G-graded and that the ideal a is graded.
Let (Λ, ≤) be a (non-empty) directed partially ordered set, and B = (bα )α∈Λ
be a system of graded ideals of R over Λ. By 13.3.14, the functor DB of 2.2.3
has the *restriction property, and the natural transformation ηB of 2.2.6(i) has
the *restriction property. In particular, Da has the *restriction property and ηa
has the *restriction property.

0
(i) Show that the natural transformation ζB of 2.2.6 also has the *restriction
property, so that, for each graded R-module M , all the homomorphisms
in the exact sequence
ξBM ηBM 0

0 - ΓB (M ) -M - DB (M ) ζBM
- H 1 (M ) -0
B

are homogeneous. Note that, as a special case, this exercise shows that
ζa0 (see 2.2.6(i)) has the *restriction property.
(ii) Let e : M −→ M  be a homogeneous homomorphism of graded R-
modules such that Ker e and Coker e are both B-torsion. Show that the
278 Foundations in the graded case

unique R-homomorphism ψ  : M  → DB (M ) for which the diagram


e
M - M

@
@
ηBM
ψ
@
R ?
@
DB (M )

commutes (see 2.2.15) is homogeneous.


The special case of this result in which B is taken to be the system formed
by the powers of a should be noted.
13.5.5 Exercise. Assume that R is G-graded and let a be a homogeneous
element of R. Show that the natural equivalence of functors
ω  : DRa = lim HomR (Ran , • ) −→ ( • )a
−→
n∈N

of 2.2.19 has the *restriction property, and deduce that, for a graded R-module
M , the isomorphism
1
HRa (M ) ∼
= Ma /(M/ΓRa (M ))
of 2.2.21(i) is homogeneous.
13.5.6 Exercise. Assume that R is G-graded and that the ideal a is graded.
Let a1 , . . . , an (where n > 0) denote n homogeneous elements which generate
a. Let u ∈ N and let K(au )• denote the Koszul complex of R with respect to
au1 , . . . , aun , as in 5.2.1.
(i) Show that K(au )• has the structure of a complex of graded R-modules
and homogeneous R-homomorphisms in which e1 ∧ . . . ∧ en has degree
n
0, the element 1 ∈ K(au )0 has degree −u i=1 deg ai , and, for k ∈
{1, . . . , n} with k < n and i ∈ I(k, n) (the notation is as in 5.1.4), the
n−k
degree of ei(1) ∧. . .∧ei(k) is −u h=1 deg aj(h) , where j ∈ I(n−k, n)
is the n-complement of i (see 5.1.4).
(ii) Let
 

=
(δ i )i∈N0 : lim Hn−i (K(au , • )• ) −→ Hai
−→ i∈N0
u∈N i∈N0

be the isomorphism of connected sequences


  C(R) to
of functors (from

itself) of Theorem 5.2.9. Show that lim Hn−i (K(au , • )• ) has


−→
u∈N i∈N0
13.5 Some examples and applications 279

the *restriction property, and that (δ i )i∈N0 has the *restriction property,
so that, for each graded R-module M , the isomorphism

=
i
δM : lim Hn−i (K(au , M )• ) −→ Hai (M )
−→
u∈N

is homogeneous.

13.5.7 Exercise. Assume that R is G-graded and that the ideal a is graded.
It follows from 13.3.14 that Da has the *restriction property.

(i) Use 13.3.11 and 13.3.12 to show that the negative strongly connected
sequence of covariant functors
 
lim ExtiR (an , • )
−→
n∈N i∈N0

from C(R) to itself has the *restriction property.


(ii) Recall from 2.2.4 that there is a unique isomorphism of connected se-
quences (of functors from C(R) to itself)
 

=
Ψa = ψai : R i Da −→ lim ExtiR (an , • )
i∈N0 i∈N0 −→
n∈N i∈N0

which extends the identity natural equivalence from Da to itself.


Let M be a graded R-module. We define structures as graded R-
modules on the Ri Da (M ) (i ∈ N) so that the R-isomorphism ψai M is
homogeneous for all i ∈ N. A consequence of this definition is that the
connected sequence (Ri Da )i∈N0 from C(R) to itself has the *restriction
property.
Denote by *Da : *C(R) −→ *C(R) the functor induced by Da .
Show that the three strongly connected sequences of covariant functors
(from *C(R) to itself)
 
 i   i 
R (*Da ) i∈N , (R Da ) i∈N and i n
lim ExtR (a , • )
0 0 −→
n∈N i∈N0

are isomorphic.

=
(iii) Let i ∈ N. Show that the natural equivalence γ i : Ri Da −→ Hai+1 of
2.2.6(ii) has the *restriction property, so that, for each graded R-module
M , there is a homogeneous isomorphism Ri Da (M ) ∼ = Hai+1 (M ) of
graded R-modules.
280 Foundations in the graded case

13.5.8 Exercise. Generalize 13.5.7 to systems of graded ideals.


In detail, assume that R is G-graded, let (Λ, ≤) be a (non-empty) directed
partially ordered set, and let B = (bα )α∈Λ be a system of graded ideals of R
over Λ. By 13.3.14, the functor DB of 2.2.3 has the *restriction property.

(i) Recall from 2.2.4 that


 
lim ExtiR (bα , • )
−→
α∈Λ i∈N0

is a negative strongly connected sequence of covariant functors from


C(R) to itself; show that it has the *restriction property.
(ii) Recall from 2.2.4 that there is a unique isomorphism of connected se-
quences (of functors from C(R) to itself)
 

=
i
Ψ B = ψB : R i DB −→ lim ExtiR (bα , • )
i∈N0 i∈N0 −→
α∈Λ i∈N0

which extends the identity natural equivalence from DB to itself. De-


duce that the first connected sequence here has the *restriction property.
Denote by *DB : *C(R) −→ *C(R) the functor induced by DB .
Show that the three strongly connected sequences of covariant functors
(from *C(R) to itself)
 
 i   i 
R (*DB ) i∈N0 , (R DB ) i∈N0 and i
lim ExtR (bα , • )
−→
α∈Λ i∈N0

are isomorphic.
(iii) Let i ∈ N. Show that there is a natural equivalence

=
i
γB : Ri DB −→ HB
i+1

that has the *restriction property.

Up to this point, this chapter has been fairly technical. In contrast, we end
the chapter with some concrete illustrations of some of the ideas developed so
far: the next exercise introduces the important concept of Veronesean subring,
and this idea is involved in Example 13.5.12 and Exercise 13.5.13.

13.5.9 Exercise: Veronesean subrings and functors. Assume that R =



n∈Z Rn is Z-graded
and that the ideal a is graded. Let r ∈ N and s ∈ Z be
fixed. Define R(r) := n∈Z Rrn : then R(r) is a subring of R, and is itself a Z-
graded (commutative Noetherian) ring with grading given by (R(r) )n = Rrn
13.5 Some examples and applications 281

for all n ∈ Z. We refer to R(r) , with this grading, as the r-th Veronesean
subring of R.
 
Let M = n∈Z Mn and L = n∈Z Ln be general graded R-modules, and
let f : M −→ L be a homogeneous homomorphism, with n-th component

fn : Mn −→ Ln for all n ∈ Z. We define M (r,s) := n∈Z Mrn+s , an R(r) -
submodule of M R(r) ; in fact, M (r,s) is a graded R(r) -module with grading
given by (M (r,s) )n = Mrn+s for all n ∈ Z. We refer to M (r,s) as the (r, s)-
(r,0)
th Veronesean submodule of M R(r) . Note that M (r,s) = (M (s)) . Also,
we denote by f (r,s)
:M (r,s)
−→ L (r,s)
the homogeneous homomorphism of
graded R(r) -modules for which (f (r,s) )n = frn+s : Mrn+s −→ Lrn+s for all
n ∈ Z. With these assignments, ( • )(r,s) : *C(R) −→ *C(R(r) ) becomes an
exact additive covariant functor, which we refer to as the (r, s)-th Veronesean
functor.
Note that, since a is graded, a(r) := a(r,0) is a graded ideal of R(r) .

(i) Show that there is an isomorphism of R(r) -modules



r−1

=
M (r,s+i) −→ M R(r) .
i=0
√ √
(ii) Show that a(r) R = a.
(iii) Show that Γa(r) (M (r,s) ) = (Γa (M ))(r,s) .
(iv) Let b be a graded ideal of R(r) , and let I be a *injective graded R-
module. Let j ∈ N. Show that there is an isomorphism of R(r) -modules

r−1

=
Hbj (I (r,s+i) ) −→ HbR
j
(I),
i=0

and deduce that the R(r) -module I (r,s) is Γb -acyclic.


(v) Show that there is a unique isomorphism
 
Φ = φi i∈N0 : Hai (r) (( • )(r,s) ) −→ (Hai ( • ))(r,s)
i∈N0 i∈N0

of negative connected sequences of covariant functors from *C(R) to


*C(R(r) ) for which φ0 is the identity natural equivalence.
(vi) Use 2.2.15 to show that there is a natural equivalence of functors

=
(Da ( • ))(r,s) −→ Da(r) (( • )(r,s) )

from *C(R) to *C(R(r) ) .

It is natural to ask whether one can generalize the concept of Veronesean


subring to multi-graded situations. This is the subject of the next exercise.
282 Foundations in the graded case

13.5.10 Exercise. Let n ∈ N and assume that G = Zn and R is G-graded;


let a be a graded ideal of R. Let G be a subgroup of finite index t in G, so
that rank G = n. Let g0 := 0G , g1 , . . . , gt−1 be representatives of the distinct

cosets of G in G. Let M = g∈G Mg be a (G-)graded R-module.
 
(i) Define RG := g ∈G Rg , and show that this is a Noetherian subring

of R; the decomposition of its definition provides RG with a grading by

G . (Here is a hint to help you show that RG is Noetherian: note that
 
RG is a direct summand of R as RG -module; consider an ascending

chain of ideals of RG , extend the ideals to R and then contract back to

RG .)
 
(ii) Let k ∈ {0, 1, . . . , t−1}. Set M G ,gk := g ∈G Mg +gk , and show that

this is an RG -submodule of M which is G -graded (by the decomposi-
tion given in its definition). Show further that there is an isomorphism of
 t−1  ∼
=
RG -modules j=0 M G ,gj −→ M RG .
  √  √
(iii) Let aG := aG ,0G . Show that aG R = a.
(iv) Let k ∈ {0, 1, . . . , t − 1}. Generalize arguments from 13.5.9 to produce
an isomorphism
   
Θ = θi i∈N0 : Hai G (( • )G ,gk ) −→ (Hai ( • ))G ,gk
j∈N0 i∈N0

of negative connected sequences of covariant functors from *C G (R) to


 
*C G (RG ) for which θ0 is the identity natural equivalence.
(v) Again for k ∈ {0, 1, . . . , t − 1}, generalize arguments from 13.5.9 to
produce a natural equivalence of functors
 ∼
= 
(Da ( • ))G ,gk −→ DaG (( • )G ,gk )
 
from *C G (R) to *C G (RG ).

13.5.11 Definition. When R = ⊕n∈Z Rn is Z-graded, we shall say that R is


positively graded precisely when Rn = 0 for all n < 0.
More generally, when R = ⊕g∈Zn Rg is Zn -graded for some positive integer
n, we shall say that R is positively graded if and only if Rg = 0 for all g =
(g1 , . . . , gn ) ∈ Zn \ N0 n .

13.5.12 Example. Let K be a field, and consider the ring K[X, Y ] of poly-
nomials over K in two indeterminates X and Y to be Z-graded so that K is
the component of degree 0 and deg X = deg Y = 1. (See 13.5.3.) Let d ∈ N
with d ≥ 3, and let A(d) be the subring of K[X, Y ] given by

A(d) := K[X d , X d−1 Y, XY d−1 , Y d ].


13.5 Some examples and applications 283

This is a subring of the d-th Veronesean subring K[X, Y ](d) of K[X, Y ], de-
scribed in 13.5.9; in fact, A(d) inherits a Z-grading from K[X, Y ](d) .
In this example, it will be convenient, when considering a positively graded

commutative Noetherian ring R = 
n∈N0 Rn , to denote the graded ideal
  
n∈N Rn by R+ .
Note that A(d)+ is the unique graded maximal ideal of A(d) . We shall now
illustrate some of the ideas of this chapter by showing that the ideal transform
DA(d)+ (A(d) ) can be naturally identified with K[X, Y ](d) , and then exploiting
1
this fact to obtain information about HA (d)+
(A(d) ).
Let φ : A(d) −→ K[X, Y ](d) denote the inclusion homomorphism. If m :=
X Y j with i, j ∈ N0 , i + j ≡ 0 (mod d) and i + j > 0, then we can write m =
i

X du Y dv X r Y d−r for some u, v, r ∈ N0 with 0 ≤ r ≤ d. If r = 0, 1, d − 1 or


d, then m ∈ A(d) . Now suppose that 2 ≤ r ≤ d − 2. Then
Y d(r−1) m = X du Y dv (XY d−1 )r ∈ A(d) ,
X d(d−r−1) m = X du Y dv (X d−1 Y )d−r ∈ A(d) ,
(X d−1 Y )r m = X du Y dv X dr Y d ∈ A(d) ,
(XY d−1 )d−r m = X du Y dv X d Y d(d−r) ∈ A(d) .
Hence Ker φ and Coker φ are both A(d)+ -torsion. Note that
 
A(d)+ K[X, Y ](d) = (X d , X d−1 Y, XY d−1 , Y d )K[X, Y ](d)
= (K[X, Y ](d) )+ = (K[X, Y ]+ )(d) .
Therefore, by 13.5.9 and the Independence Theorem 4.2.1, we have, for each
i ∈ N0 ,
i
HA (d)+
(K[X, Y ](d) ) ∼ i
= HA (d)+ K[X,Y ]
(d) (K[X, Y ]
(d)
)
i (d)
= H(K[X,Y ]+ )(d) (K[X, Y ] )
∼ i
= (H(X,Y (d,0)
) (K[X, Y ])) ,
and this is zero for i = 0, 1.
We can now use 2.2.15 to see that there is a unique A(d) -isomorphism φ :
K[X, Y ](d) −→ DA(d)+ (A(d) ) such that the diagram
φ
A(d) - K[X, Y ](d)

@
@
ηA(d)

= φ
@
@
R ?
DA(d)+ (A(d) )
284 Foundations in the graded case

commutes, and we can use 13.5.4(ii) to see that φ is homogeneous. So, by


13.5.4(i), there is a homogeneous isomorphism Coker φ ∼ = HA1
(d)+
(A(d) ) of
A(d) -modules. For each n ∈ N0 , let An (respectively Bn ) denote the n-th com-
ponent of the graded A(d) -module A(d) (respectively K[X, Y ](d) ). We now
compare An and Bn .
First, when 1 ≤ n < d − 2 the monomial X d−2 Y nd−d+2 cannot be ex-
pressed as a product of n factors taken from {X d , X d−1 Y, XY d−1 , Y d }, and
so An ⊂ Bn .
Next, we consider the case where n ≥ d − 2. Of course, X dn ∈ An ; we
therefore consider a monomial of the form X rd+s Y (n−r)d−s , where 0 ≤ r ≤
n−1 and 0 ≤ s ≤ d−1. We claim that either (a) r+s ≤ n or (b) r+s ≥ d−1:
if this were not the case, then we should have
n≤r+s−1 and r + s ≤ d − 2,
which would imply that n ≤ d − 3, a contradiction. In case (a), we have
X rd+s Y (n−r)d−s = (X d )r (XY d−1 )s (Y d )n−r−s ∈ An ,
while in case (b) we have
X rd+s Y (n−r)d−s = (X d )r+s+1−d (X d−1 Y )d−s (Y d )n−r−1 ∈ An .
Thus An = Bn in this case.
1
It follows that the n-th component of HA (d)+
(A(d) ) is non-zero if and only
if 1 ≤ n < d − 2.
13.5.13 Exercise. Let K be a field and let R := K[X1 , . . . , Xn ], the ring of
polynomials over K in n indeterminates (where n ∈ N), Z-graded so that Xi
has degree 1, for all i = 1, . . . , n. Let r ∈ N, and consider the r-th Veronesean
subring R(r) of R, as in 13.5.9. Thus R(r) is the K-subspace of R generated
by
{X1v1 . . . Xnvn : v1 , . . . , vn ∈ N0 , v1 + · · · + vn ≡ 0 (mod r)} .
(r)
Let R+ denote the unique graded maximal ideal of R(r) . Show that
i
HR (r) (R
(r)
) = 0 for all i ∈ N0 \ {n},
+

and deduce from [7, 1.5.8 and 1.5.9] that R(r) is Cohen–Macaulay.
14

Graded versions of basic theorems

We have now laid the foundations of multi-graded local cohomology theory


in Chapter 13, where the gradings on R and R-modules M are by a finitely
generated, torsion-free Abelian group G. Indeed, in the case where R is G-
graded and the ideal a is G-graded, we now know that, for a G-graded R-
module M , there is a natural way in which to define G-gradings on the local
cohomology modules Hai (M ) (i ∈ N0 ); furthermore, whenever f : M −→ N
is a morphism in *C G (R), then Hai (f ) is a homogeneous homomorphism for
all i ∈ N0 ; also, whenever 0 −→ L −→ M −→ N −→ 0 is an exact sequence
in the ‘G-graded’ category *C G (R), then all the homomorphisms, including
the connecting homomorphisms, in the induced long exact sequence of local
cohomology modules (with respect to a) are G-homogeneous.
This chapter is concerned with refinements available in this multi-graded
case of such fundamental results as the Independence Theorem 4.2.1, the Flat
Base Change Theorem 4.3.2, Faltings’ Annihilator Theorem 9.5.1, Grothendi-
eck’s Finiteness Theorem 9.5.2 and the Local Duality Theorem 11.2.6, and of
the theory of canonical modules developed in Chapter 12. However, although
it is true that part of this chapter is a retracing of steps through earlier chapters,
revisiting many of the highlights in order to ‘add graded frills’, we have felt
it necessary to include quite a bit of the underlying algebra of multi-graded
commutative Noetherian rings.
For example, there is a multi-graded analogue of Matlis’s decomposition
theory for injective modules over a commutative Noetherian ring, and this
multi-graded analogue has both strong similarities to, and fascinating links
with, the ungraded theory. We present the multi-graded version in §14.2, and
make much use of it in our treatment of graded local duality in §14.4 and
*canonical modules in §14.5.
In §14.3, we present some results due to S. Goto and K.-i. Watanabe [22,
§1.2] that also concern the algebra of a commutative Noetherian ring R that is
286 Graded versions of basic theorems

graded by a finitely generated torsion-free Abelian group. Let M be a finitely


generated graded R-module, and let p ∈ Supp M . We showed in 13.1.6(i)
that p* ∈ Supp M . Our principal aims in §14.3 are to present Goto’s and
Watanabe’s results that
dim Mp = dim Mp∗ + ht p/p* and depth Mp = depth Mp∗ + ht p/p*,
and to apply them to prove graded versions of Faltings’ Annihilator Theorem
9.5.1 and Grothendieck’s Finiteness Theorem 9.5.2. However, we also use the
Goto–Watanabe results later in the chapter.
Limitations on space mean that almost all of our treatment in §14.5 of *cano-
nical modules is concerned with the case where R is a Cohen–Macaulay multi-
graded ring with a unique maximal graded proper ideal m. We define a *can-
onical module for R to be a finitely generated graded R-module C for which
there is a homogeneous isomorphism * HomR (C, *ER (R/m)) ∼ = Hm n
(R).
This definition is the obvious graded analogue of our definition of canonical
module in the ungraded local case. However, in the special case in which R
is Z-graded, Bruns and Herzog give an alternative definition in [7, Definition
3.6.8] which is not obviously equivalent to ours. The work involved in our rec-
onciliation of these two approaches (see 14.5.12) has contributed significantly
to the length of this chapter.
Readers who are only interested in the case of Z-graded rings rather than the
full multi-graded case should be able to pass quickly over some parts of this
chapter. For example, the special cases of the above-mentioned results of Goto
and Watanabe in which the grading group is Z are essentially covered by [7,
Theorems 1.5.8 and 1.5.9].

14.1 Fundamental theorems


14.1.1 Notation and Terminology. Throughout this chapter, G will denote a
finitely generated, torsion-free Abelian group, written additively, and we shall
assume that our commutative Noetherian ring R is G-graded, with grading
given by R = ⊕g∈G Rg . Occasionally, we shall consider particular cases in
which R is Z-graded, that is, in which G is taken to be Z.
We shall employ the notation, conventions and terminology concerning G-
graded rings and modules described in 13.1.1. In addition, when the ideal a
is graded, and M is a graded R-module, we use Hai (M )g to denote the g-th
component of the graded R-module Hai (M ) (for i ∈ N0 and g ∈ G).
A maximal member of the set of proper graded ideals of R is referred to as
a *maximal graded ideal of R. We shall say that R is *local precisely when it
14.1 Fundamental theorems 287

has exactly one *maximal graded ideal. The statement ‘(R, m) is *local’ is to
be interpreted as meaning that m is the unique *maximal graded ideal of the
G-graded ring R. For an ideal b of R, we shall denote by b* the graded ideal
generated by all homogeneous elements of b.
Let S be a multiplicatively closed subset of R consisting of non-zero ho-
mogeneous elements. We pointed out in 13.1.1 that the ring S −1 R is also G-
graded, and that, if M is a graded R-module, then the S −1 R-module S −1 M
is also graded.
If p is a prime ideal of R and we take S to be the set of all homogeneous
elements of R that lie outside p, then the resulting G-graded ring S −1 R (re-
spectively module S −1 M ) is called the homogeneous localization of R (re-
spectively M ) at p, and is denoted by R(p) (respectively M(p) ). This concept
should not be confused with a different concept for which Hartshorne, in [30,
p. 18], uses the notation ‘(p) ’. Note that R(p) is a *local ring, with pR(p) as its
unique *maximal graded ideal.
Some of our work will be particularly concerned with the case where G = Z.

Suppose that R is Z-graded (that is, that G = Z), and let M = n∈Z Mn be
a graded R-module. We define the end of M by

end(M ) := sup {n ∈ Z : Mn = 0}

if this supremum exists, and ∞ otherwise. (We adopt the convention that the
supremum of the empty set of integers is to be taken as −∞, and we interpret
−∞+t as −∞ for all t ∈ Z.) With analogous conventions, we similarly define
the beginning of M , denoted by beg(M ), to be inf {n ∈ Z : Mn = 0} if this
infimum exists, and −∞ otherwise.

14.1.2 Remark. Suppose that M is a graded R-module.

(i) Recall from 13.1.6(i) that, if p ∈ Supp M , then the graded prime p*
also belongs to Supp M . Hence if M has no graded prime ideal in its
support, then M = 0.
(ii) Consequently, if (R, m) is *local (see 14.1.1) and Mm = 0, then M = 0.

14.1.3 Definition and Remark. Let R = g∈G Rg be a commutative G-
graded ring, and let f : R −→ R be a ring homomorphism.
We say that f is homogeneous precisely when f (Rg ) ⊆ Rg for all g ∈ G.

Assume that this is the case. Let M  =  
g∈G Mg be a graded R -module.
Then the same direct sum decomposition provides the R-module M  R with
a structure as a graded R-module. In fact, in the terminology of 13.3.6, the
functor R : C G (R ) −→ C G (R) has the *restriction property.
Whenever we regard, in such circumstances, a graded R -module M  as a
288 Graded versions of basic theorems

graded R-module, it is to be understood that the same grading is used for the
two structures. Thus we can write (M  R )g = (Mg ) R0 for all g ∈ G.

14.1.4 Proposition. Assume that a is graded, and let R = g∈G Rg be a
second commutative Noetherian G-graded ring. Let f : R −→ R be a ring
homomorphism which is homogeneous (see 14.1.3).
The natural equivalence of functors ε : DaR ( • ) R −→ Da ( • R ) of 2.2.24
(from C(R ) to C(R)) has the *restriction property.
Proof. Since aR is a graded ideal of R , it follows from 13.3.14 and 14.1.3
that both DaR ( • ) R and Da ( • R ) have the *restriction property.
Let M  be a graded R -module. The homomorphism
ηaR ,M  R : M R −→ DaR (M  ) R

is homogeneous by 13.3.14, and has kernel and cokernel which are a-torsion
by 2.2.6(i)(c). The result therefore follows from 13.5.4(ii).
14.1.5 Exercise: Graded Mayer–Vietoris sequence. Assume that a is
graded; let b be a second graded ideal of R. Show that, for a graded R-module
M , all the homomorphisms in the Mayer–Vietoris sequence (see 3.2.3)

0 - H 0 (M ) - H 0 (M ) ⊕ H 0 (M ) - H 0 (M )
a+b a b a∩b

- H 1 (M ) - H 1 (M ) ⊕ H 1 (M ) - H 1 (M )
a+b a b a∩b

- ··· ···
- H i (M ) - H i (M ) ⊕ H i (M ) - H i (M )
a+b a b a∩b

- H i+1 (M ) - ···
a+b

are homogeneous.
14.1.6 Exercise. Assume that a is graded; let i ∈ N0 .
(i) Let (Λ, ≤) be a (non-empty) directed partially ordered set; let (Wα )α∈Λ
be a direct system of graded R-modules and homogeneous R-homomor-
phisms over Λ, with constituent R-homomorphisms hα β : Wβ → W α
(for each (α, β) ∈ Λ × Λ with α ≥ β). (See 13.3.12.)
Show that the isomorphism
 

=
lim Hai (Wα ) −→ Hai lim Wα
−→ −→
α∈Λ α∈Λ

given by Theorem 3.4.10 (see also 3.4.1) is homogeneous.


14.1 Fundamental theorems 289

(ii) Let (Lθ )θ∈Ω be a non-empty family of graded R-modules. Show that the
isomorphism
  ∼= 
Hai θ∈Ω Lθ −→
i
θ∈Ω Ha (Lθ )

given by (Theorem 3.4.10 and) Exercise 3.4.5 is homogeneous.


Next we use Theorem 13.3.15 to establish quickly G-graded versions of the
Independence Theorem 4.2.1 and the Flat Base Change Theorem 4.3.2.
14.1.7 Graded Independence Theorem. Assume that a is graded, and let

R = g∈G Rg be a second commutative Noetherian G-graded ring.
Let f : R −→ R be a ring homomorphism which is homogeneous (see
14.1.3).
(i) Both the negative (strongly) connected sequences of covariant functors
i
(HaR ( • ) R )i∈N0 and (Hai ( • R ))i∈N0

from C(R ) to C(R) have the *restriction property.


(ii) The isomorphism of connected sequences

=
Λ = (λi )i∈N0 : (HaR
i
( • ) R )i∈N0 −→ (Hai ( • R ))i∈N0

of 4.2.1 has the *restriction property. Consequently, for each i ∈ N0 and


each graded R -module M  , there is a homogeneous R-isomorphism
 ∼
= 
λiM  : HaR  (M ) −→ Ha (M ).
i i

Proof. (i) Since f is homogeneous, the restriction functor R : C(R ) −→


C(R) has the *restriction property of 13.3.6: see 14.1.3. Since, by 13.4.3 and
13.4.4, the connected sequence (Hai )i∈N0 (from C(R) to C(R)) has the *restric-
tion property, it is immediate that (Hai ( • R ))i∈N0 has the *restriction property.
Since the extension aR of a to R under f is a graded ideal, it is just as easy
i
to see that (HaR  ( • ) R )i∈N0 has the *restriction property.
0
(ii) Since λ is the identity natural equivalence from ΓaR ( • ) R = Γa ( • R )
to itself, λ0 has the *restriction property.
For each i ∈ N0 and each graded R -module M  , we can use the grad-
ing on Hai (M  R ) of (i) to define a grading on HaR i
 (M )

R in such a

 = 
way that the isomorphism λiM  : HaR i
 (M ) R −→ H i
a (M R ) is homo-

geneous: we recover the grading of part (i) on the ΓaR (M ) R . With respect
i
to these gradings, (HaR  ( • ) R )i∈N0 also has the *restriction property. Since

HaR (I ) R = 0 for all i ∈ N whenever I  is a *injective graded R -module


i 

(by 13.2.6), it follows from 13.3.15 that these gradings coincide with the natu-
ral ones used in part (i). All the remaining claims follow from this.
290 Graded versions of basic theorems

Below, we shall use a similar argument to establish a graded version of the


Flat Base Change Theorem 4.3.2. However, the reader might find the following
preparatory remark helpful.
 
14.1.8 Remark. Let M = g∈G Mg be a graded R-module. Let R =
 
g∈G Rg be a second commutative Noetherian G-graded ring, and let f :
R −→ R be a ring homomorphism which is homogeneous (see 14.1.3). Then
R R is a graded R-module (as is explained in 14.1.3), and therefore M ⊗R R
has a structure as a graded R-module. In fact, the direct decomposition

M ⊗R R = (M ⊗R R )g
g∈G

described in 13.1.9 actually provides M ⊗R R with a structure as a graded


R -module, and ( • ) ⊗R R : C(R) −→ C(R ) has the *restriction property.

14.1.9 Graded Flat Base Change Theorem. Assume that a is graded, and

let R = g∈G Rg be a second commutative Noetherian G-graded ring.
Let f : R −→ R be a ring homomorphism which is homogeneous (see
14.1.3) and flat.

(i) Both the negative (strongly) connected sequences of covariant functors

(Hai ( • ) ⊗R R )i∈N0 and i


(HaR 
 (( • ) ⊗R R ))i∈N0

from C(R) to C(R ) have the *restriction property.


(ii) The isomorphism of connected sequences

(ρi )i∈N0 : (Hai ( • ) ⊗R R )i∈N0 −→ (HaR 
=
 (( • ) ⊗R R ))i∈N0
i

of 4.3.2 has the *restriction property, so that, for each i ∈ N0 and each
graded R-module M , there is a homogeneous R -isomorphism

ρiM : Hai (M ) ⊗R R −→ HaR 
=
 (M ⊗R R ).
i

Proof. (i) Since f is homogeneous, the functor ( • ) ⊗R R : C(R) −→


C(R ) has the *restriction property: see 14.1.8. Since the extension aR of a
to R under f is a graded ideal, it follows from 13.4.3 and 13.4.4 that the
 
 )i∈N0 (from C(R ) to C(R )) has the *restriction
i
connected sequence (HaR

 (( • ) ⊗R R ))i∈N0 has the *restriction property. It is just
i
property. Hence (HaR

as easy to see that (Ha ( • ) ⊗R R )i∈N0 has the *restriction property.
i

(ii) Note that, by 4.3.1, the natural equivalence ρ0 has the *restriction prop-
erty.
For each i ∈ N0 and each graded R-module M , we can use the grading
 
 (M ⊗R R ) of (i) to define a grading on Ha (M ) ⊗R R in such a
i i
on HaR
14.1 Fundamental theorems 291

way that the isomorphism ρiM : Hai (M ) ⊗R R −→ HaR 
=
 (M ⊗R R ) is
i

homogeneous: we recover the grading of (i) on the Γa (M )⊗R R . With respect
to these gradings, (Hai ( • ) ⊗R R )i∈N0 also has the *restriction property. Since
Hai (I)⊗R R = 0 for all i ∈ N whenever I is a *injective graded R-module (by
13.2.6), it follows from 13.3.15 that these gradings coincide with the natural
ones used in part (i). All the remaining claims follow from this.

Next, we explore the behaviour of graded local cohomology with respect to


a shift functor.

14.1.10 Remarks. Let g0 ∈ G and j ∈ N0 . Let rg0 ∈ Rg0 be a homogeneous


element of degree g0 . Let L, M be graded R-modules.

(i) The graded R-modules

* HomR (L, M (g0 )) and (* HomR (L, M ))(g0 )

are equal. Since * ExtjR ( • , N ) is the j-th right derived functor in *C(R)
of * HomR ( • , N ), we see that the graded R-modules * ExtjR (L, M (g0 ))
and (* ExtjR (L, M ))(g0 ) are again equal (and not just isomorphic in
*C(R)). Hence the graded R-modules
 
lim ExtjR (R/an , M (g0 )) and lim ExtjR (R/an , M ) (g0 )
−→ −→
n∈N n∈N

are equal.
(ii) If we forget the gradings on M and M (g0 ), we obtain the same ungraded
R-module M . By 13.4.6(ii), the grading on Haj (M ) can be defined from
the grading of 13.3.13(ii) on lim ExtjR (R/an , M ) simply by requiring
−→
n∈N
that the isomorphism φja M be homogeneous. Similarly, we can obtain
the grading on Haj (M (g0 )) by using φja M to ‘lift across’ the grading of
13.3.13(ii) on lim ExtjR (R/an , M (g0 )). It therefore follows from part
−→
n∈N
(i) that the graded modules Haj (M (g0 )) and (Haj (M ))(g0 ) are equal.
(iii) The fact that Rj Da (M (g0 )) and (Rj Da (M ))(g0 ) are the same graded
R-module can be deduced in a similar way from 13.5.7(ii).

14.1.11 Exercise. Obtain a ‘graded’ version of Proposition 8.1.2.


In detail, assume that a is graded; let b be a homogeneous element of R. Let
f : M −→ N be a homogeneous homomorphism of graded R-modules.
(i) Show that there is a long exact sequence of graded R-modules and
292 Graded versions of basic theorems

homogeneous R-homomorphisms

- H0 - H 0 (M ) - H 0 (Mb )
0 a+Rb (M ) a a

- H1 - H 1 (M ) - H 1 (Mb )
a+Rb (M ) a a

- ··· ···
- Hi - H i (M ) - H i (Mb )
a+Rb (M ) a a

- H i+1 (M ) - ···
a+Rb

such that the diagram

i
Ha+Rb (M ) - H i (M ) - H i (Mb ) - H i+1 (M )
a a a+Rb

i i+1
Ha+Rb (f ) Hai (f ) Hai (fb ) Ha+Rb (f )

? ? ? ?
i
Ha+Rb (N ) - H i (N ) - H i (Nb ) - H i+1 (N )
a a a+Rb

commutes for all i ∈ N0 .


(ii) Let i ∈ N0 . Show that there is a commutative diagram

0 - H 1 (H i (M )) - H i+1 (M ) - ΓRb (H i+1 (M )) -0


Rb a a+Rb a

1 i+1
HRb (Hai (f )) Ha+Rb (f ) ΓRb (Hai+1 (f ))

? ? ?
0 - H 1 (H i (N )) - H i+1 (N ) - ΓRb (H i+1 (N )) -0
Rb a a+Rb a

(in the category *C(R)) with exact rows. The top row is referred to as the
comparison exact sequence for M .

14.1.12 Lemma. Assume that G = Zn and R = g∈N0 n Rg is positively
graded, and that the ideal a is generated by homogeneous elements of degree
0. Let a0 = a ∩ R0 . Let h ∈ G. We denote by ( • )h : *C(R) −→ C(R0 ) the
functor which assigns, to each graded R-module, and to each homogeneous
homomorphism of graded R-modules, the h-th component.

=
There is an isomorphism (Hai ( • )h )i∈N0 −→ (Hai 0 (( • )h ))i∈N0 of negative
strongly connected sequences of covariant functors from *C(R) to C(R0 ).

Proof. Observe that Γa ( • )h and Γa0 (( • )h ) are the same functor. Let I =

g∈Zn Ig be a *injective graded R-module. Impose the trivial Z -grading
n
14.1 Fundamental theorems 293

on the commutative Noetherian ring R0 . Since a = a0 R, it follows from the


Graded Independence Theorem 14.1.7 and Exercise 3.4.5 that, for all i ∈ N,
 ∼
0 = Hai (I) = Hai 0 R (I) ∼
= Hai 0 (I R0 ) = Hai 0 g∈G Ig =
i
g∈G Ha0 (Ig ).

Hence Ih is Γa0 -acyclic. The result is now an easy consequence of 13.3.5.



14.1.13 Example. Assume that that G = Zn and R = g∈N0 n Rg is posi-
tively graded. Let h ∈ Zn and let L be an R0 -module. We can define a graded
R-module hL such that, for all g ∈ G,

h L if g = h,
( L)g =
0 if g = h.

(These conditions necessitate that rg m = 0 for all m ∈ hL and all rg ∈ Rg


whenever g ∈ G \ {0}.)
Assume that the ideal a is graded, and let a0 = a ∩ R0 . Let i ∈ N0 . We show

how to use Lemma 14.1.12 to calculate Hai (hL). Let R+ = g∈N0 n \{0} Rg ;
observe that R+ annihilates hL and that a+R+ = a0 R +R+ . The argument of
Example 4.2.2 can be modified to our ‘Zn -graded’ situation to produce homo-
geneous R-isomorphisms Hai (hL) ∼ i
= Ha+R +
(hL) ∼
= Hai 0 R (hL). We can now
use 14.1.12 to deduce that, for g ∈ Z ,
n

i h ∼ i h ∼ i h Hai 0 (L) if g = h,
Ha ( L)g = Ha0 R ( L)g = Ha0 (( L)g ) =
0 if g = h.

Hence there is a homogeneous R-isomorphism hHai 0 (L) ∼


= Hai (hL).
We end this section with some results that concern the special case of Z-
graded rings, that is, the special case where G = Z. In this case, much basic
theory has been developed by Bruns and Herzog in [7, §1.5 and §3.6]. We now
prepare for a Z-graded analogue of (a special case of) the local Lichtenbaum–
Hartshorne Vanishing Theorem 8.2.1. Before we come to the theorem itself,
however, we provide two preparatory exercises.
14.1.14 Exercise. Assume that G = Z and that (R, m) is *local; assume
further that m is actually a maximal ideal of R. Let M be a graded R-module.
Use [7, 1.5.6 and 1.5.8] to show that dim M = dimRm Mm .

14.1.15 Exercise. Assume that G = Z, that R = j∈N0 Rj is positively
graded, and that the subring R0 is a local ring having maximal ideal m0 .
(i) Show that R is *local with unique *maximal ideal
m := m0 ⊕ R1 ⊕ R2 ⊕ · · · ⊕ Rn ⊕ · · · .
294 Graded versions of basic theorems

(ii) Show that, for all j, i ∈ N, the j-th component of mj+i is contained in
mi0 Rj .
(iii) Assume that R = R0 . By [7, 1.5.4], there exist non-zero homogeneous
elements y1 , . . . , yt of R of positive degrees such that

R = R0 [y1 , . . . , yt ].

Let d := max{deg yi : 1 ≤ i ≤ t}. Show that R(i−1)d+j ⊆ mi for all


i, j ∈ N.
(iv) Show that the multiplication in R induces a natural ring structure on the

Abelian group j∈N0 Rj . (The following hint can ease the checking of

the ring axioms. For a := (a0 , a1 , . . . , aj , . . .) ∈ j∈N0 Rj and h ∈ N0 ,
define
a≤h := (a0 , a1 , . . . , ah , 0, 0, . . .);

show that (a≤h b≤h )≤h = (ab)≤h for a, b ∈ j∈N0 Rj .)
(v) Assume in addition that R0 is complete. Prove that the inclusion map

R −→ j∈N0 Rj provides the m-adic completion of R.

14.1.16 Graded Lichtenbaum–Hartshorne Vanishing Theorem. Assume



that G = Z, and that R = j∈N0 Rj is positively graded and an integral do-
main; assume also that the subring R0 is a complete local ring having maximal
ideal m0 .
Assume that the ideal a is graded and proper, and that dim R/a > 0. Set
d := dim R. Then Had (R) = 0.

Proof. Let m := m0 ⊕ R1 ⊕ R2 ⊕ · · · ⊕ Rj ⊕ · · · . By 14.1.15, our graded


ring R is *local with unique *maximal ideal m. Let R +m denote the completion
of the local ring Rm . In view of the natural ring isomorphisms

=
R/mj −→ Rm /(mRm )j for j ∈ N,

R+m is isomorphic to the m-adic completion of R, which is an integral domain


by 14.1.15.
Note that dim Rm = d, by 14.1.14, since m is actually a maximal ideal of
R. Since a is graded, all its minimal primes are graded, and so are contained
in m. Since dim R/a > 0, it therefore follows that m is not a minimal prime
of a. Hence dim R +m /aR +m = dim (Rm /aRm ) > 0. We can therefore use
the local Lichtenbaum–Hartshorne Vanishing Theorem 8.2.1 to deduce that
d
HaR m
(Rm ) = 0. Thus (Had (R))m = 0, by 4.3.3. Therefore Had (R) = 0, by
14.1.2(ii).
14.2 *Indecomposable *injective modules 295

Two other results to which we would like to add ‘graded frills’ are Faltings’
Annihilator Theorem 9.5.1 and Grothendieck’s Finiteness Theorem 9.5.2. We
shall have to defer graded versions of these results until the end of §14.3, by
which point we shall have presented more results about the behaviour of G-
graded modules.

14.2 *Indecomposable *injective modules


One of our major aims in this chapter is to provide versions of local duality
which apply in ‘graded’ situations, including versions which involve ‘graded
canonical modules’. To prepare for this, we shall develop the ‘graded’ ana-
logue of the decomposition theory for injective modules over a commutative
Noetherian ring.
Recall that we are assuming that R is G-graded throughout this chapter. In
13.2.4, we showed that each graded R-module M has a *injective envelope,
and between any two *injective envelopes of M there is a homogeneous iso-
morphism which restricts to the identity map on M ; we agreed to denote by
*E(M ) or *ER (M ) one choice of *injective envelope of M ; and we proved
that *E(M ), with its grading forgotten, is an essential extension of M .

14.2.1 Exercise. Consider a non-empty family (Mλ )λ∈Λ of graded R-mod-


ules.

(i) Show that λ∈Λ Mλ is *injective if and only if Mλ is *injective for all
λ ∈ Λ. (Here is a hint: you might find the ‘graded Baer criterion’ 13.2.7
useful.)
 
(ii) Show that the obvious map λ∈Λ Mλ −→ λ∈Λ *E(Mλ ) provides

the *injective envelope of λ∈Λ Mλ .

14.2.2 Exercise. Let I be a graded submodule of the graded R-module M ,


and suppose that I is *injective. Show that I is a direct summand of M with
graded complement.

14.2.3 Definition. A graded R-module is said to be *indecomposable pre-


cisely when it is non-zero and cannot be written as the direct sum of two proper
graded submodules.

14.2.4 Proposition. Let p ∈ * Spec(R).

(i) The *injective graded R-module *E(R/p)(g) is *indecomposable for


each g ∈ G.
296 Graded versions of basic theorems

(ii) A non-zero *injective graded R-module I has a *indecomposable *injec-


tive graded submodule which must, by 14.2.2, be a direct summand (with
graded complement). In fact, for each q ∈ Ass I, there exists g0 ∈ G and
a homogeneous element m of I of degree g0 for which (0 :R m) = q,
and then I has a graded submodule that is homogeneously isomorphic
to *E(R/q)(−g0 ).
(iii) Each *indecomposable *injective graded R-module is isomorphic (in
the category *C(R)) to *E(R/q)(−g0 ) for some q ∈ * Spec(R) and
g0 ∈ G.
(iv) Let r be a homogeneous element of degree g in R\p. Then multiplication
by r provides a homogeneous automorphism of degree g of *E(R/p).
Also, each element of *E(R/p) is annihilated by some power of p.
(v) Let q ∈ * Spec(R). If *E(R/p) ∼ = *E(R/q)(−g) (in *C(R)) for some
g ∈ G, then p = q.

Proof. (i) Let m be a homogeneous generator of the graded submodule R/p


of *E(R/p). Suppose that L, N are non-zero graded submodules of *E(R/p)
such that *E(R/p) = L ⊕ N . Then L ∩ Rm = 0 and N ∩ Rm = 0; thus there
exist homogeneous elements a, b ∈ R such that 0 = am ∈ L and 0 = bm ∈
N . Since (0 :R m) = p, prime, we must have a, b ∈ R \ p, so that ab ∈ p.
Therefore 0 = abm ∈ L ∩ N , and this is a contradiction. Therefore *E(R/p)
is *indecomposable. It follows easily that *E(R/p)(g) is *indecomposable for
each g ∈ G.
(ii) Since I = 0, it must have an associated prime, q say, which must be
graded, by 13.1.6(ii); moreover, q = (0 :R m) for some homogeneous element
m of I. Let deg m = g0 . Thus there is a homogeneous isomorphism ϕ :

=
(R/q)(−g0 ) −→ Rm such that ϕ(1 + q) = m.
Since I is *injective, we can extend ϕ to a homogeneous homomorphism
ψ : *E((R/q)(−g0 )) = *E(R/q)(−g0 ) −→ I; as

Ker ψ ∩ (R/q)(g0 ) = 0,

it follows that ψ is monomorphic; therefore J := Im ψ is a *injective graded


submodule of I and there is a homogeneous isomorphism

J∼
= *E(R/q)(−g0 ).

It follows from (i) that J is *indecomposable.


Since J is *injective, there is a homogeneous R-homomorphism ξ : I −→ J
that extends the identity map IdJ on J; then Ker ξ is a graded submodule of I
and J ⊕ Ker ξ = I. Thus J is a direct summand of I with graded complement.
14.2 *Indecomposable *injective modules 297

(iii) Apply (ii) to a *indecomposable *injective graded R-module I, and the


desired conclusion is immediate.
(iv) Multiplication by r provides a homogeneous R-homomorphism μr :
*E(R/p) −→ *E(R/p)(g); since Ker μr ∩(R/p) = 0, the map μr is monom-
orphic. Therefore Im μr is a *injective graded submodule of the *indecompos-
able *injective graded R-module *E(R/p)(g), and it follows from 14.2.2 that
μr must be surjective.
For the second claim, it is enough for us to show that an arbitrary non-zero
homogeneous element m ∈ *E(R/p) is annihilated by some power of p. Let
q ∈ Ass Rm, and recall from 13.1.6(ii) that q is graded. Since Rm ∩ (R/p) =
0, we see that q ∈ Ass(R/p) = {p}, so that q = p. Therefore (0 :R Rm) is
p-primary.
(v) Let h(R) denote the set of non-zero homogeneous elements of R, and
let g  ∈ G. By (iv), the set of homogeneous elements r of R for which multi-
plication by r provides an automorphism (of some degree) of *E(R/p)(−g  )
is precisely h(R) \ p. The desired conclusion is now immediate.
The next lemma can be proved by making straightforward modifications to
the proof of the corresponding ‘ungraded’ result in 10.1.12.
14.2.5 Lemma. Let S be a multiplicatively closed subset of homogeneous
elements of R, and let M be a (G-)graded S −1 R-module. Then M is *injective
over R if and only if it is *injective over S −1 R.
14.2.6 Lemma. Let S be a multiplicatively closed subset of homogeneous
elements of R, and let p ∈ Spec(R) be such that p ∩ S = ∅. By 14.2.4(iv),
the *indecomposable *injective graded R-module *ER (R/p) has a natural
structure as a G-graded S −1 R-module.
In the category *C(S −1 R), we have
*ER (R/p) ∼
= *ES −1 R (S −1 R/S −1 p).
Furthermore, *ES −1 R (S −1 R/S −1 p), when considered as a G-graded R-
module by means of the natural homomorphism R −→ S −1 R, is homoge-
neously isomorphic to *ER (R/p).
Proof. By Lemma 14.2.5, the graded S −1 R-module *ER (R/p) is *injective
over S −1 R. Since a graded S −1 R-submodule of *ER (R/p) is automatically a
graded R-submodule, it is immediate from 14.2.4(i) that *ER (R/p) is *indec-
omposable as S −1 R-module.
There is a homogeneous generator m of degree 0 of the graded R-submodule
R/p of *ER (R/p). When the latter module is considered as a graded S −1 R-
module, (0 :S −1 R m) = S −1 p, and m still has degree 0. It therefore follows
298 Graded versions of basic theorems

from 14.2.4(ii) that there is a homogeneous S −1 R-isomorphism *ER (R/p) ∼


=
*ES −1 R (S −1 R/S −1 p).
The final claim is now immediate.
14.2.7 Exercise. Suppose that our G-graded ring R is *simple, that is, R
has exactly two graded ideals, namely 0 and R. Use the ‘graded Baer criterion’
13.2.7 to show that every graded R-module is *injective.
14.2.8 Exercise. Assume that a is graded, and that p ∈ * Spec(R) is such
that a ⊆ p. Show that the graded R/a-module (0 :*ER (R/p) a) is homoge-
neously isomorphic to *ER/a ((R/a)/(p/a)).
14.2.9 Exercise. Assume that (R, m) is *local. Show that there are homo-
geneous isomorphisms
* HomR (R/m, *E(R/m)) ∼
= (0 :*E(R/m) m) ∼
= *ER/m (R/m) = R/m.
14.2.10 Exercise. Show that each *injective graded R-module I is a direct
sum of *indecomposable *injective graded submodules. (Here is a hint: adapt
the argument in the proof of 10.1.8 to our G-graded situation, that is, apply
Zorn’s Lemma to the set of all sets of *indecomposable *injective graded sub-
modules of I whose sum is direct.)
14.2.11 Definition and Exercise. Let M be a graded R-module. A minimal
*injective resolution of M is a *injective resolution
d0 di
I • : 0 −→ I 0 −→ I 1 −→ · · · −→ I i −→ I i+1 −→ · · ·
of M (in the category *C(R)) such that I i is a *essential extension of Ker di
for every i ∈ N0 .
Show that M has a minimal *injective resolution, and that the i-th term in
such a resolution is uniquely determined, up to isomorphism in *C(R), by M .
We denote this i-th term by *E i (M ), or by *ER
i
(M ).
With the notation of 14.2.11, and for i ∈ N0 , it follows from 14.2.4(iii) and
14.2.10 that there is a family (pα )α∈Λ of graded prime ideals of R and a family
(gα )α∈Λ of elements of G for which there is a homogeneous isomorphism
= 

*E i (M ) −→ α∈Λ *E(R/pα )(−gα ).
Of course, we would like, as in the analogous ungraded situation (see 11.1.4)
to be able to show that, for p ∈ * Spec(R), the cardinality of the set
{α ∈ Λ : pα = p}
depends only on *E i (M ) and p (and therefore only on i, M and p) and not on
14.2 *Indecomposable *injective modules 299

the particular decomposition of *E i (M ) (as a direct sum of *indecomposable


*injective submodules) chosen. This is indeed the case, and, interestingly, the
cardinality in question turns out to be equal to the ordinary Bass number
μi (p, M ) of 11.1.4. We are, in the spirit of this book, going to establish this;
however, instead of trying to imitate the standard ‘ungraded’ argument, we
shall employ Theorem 13.2.4(iii).

14.2.12 Proposition. Let M be a graded R-module. By 14.2.10 and 14.2.4,


there is a family (pα )α∈Λ of graded prime ideals of R and a family (gα )α∈Λ

=
of elements of G for which there is a homogeneous isomorphism *E(M ) −→

α∈Λ *E(R/pα )(−gα ).
Let p ∈ * Spec(R). Then the cardinality of the set {α ∈ Λ : pα = p} is
equal to the ordinary Bass number μ0 (p, M ) (see 11.1.4), and so depends only
on M and p and not on the particular decomposition of *E(M ) (as a direct
sum of *indecomposable *injective submodules) chosen.

Proof. If (Nα )α∈Λ is a family of R-modules, then, by 14.2.1(ii) (applied in


the case where R is considered to be trivially G-graded, that is, R0 = R and
Rg = 0 for all g ∈ G \ {0}),
  
E ∼
α∈Λ Nα = α∈Λ E(Nα ).

Now consider the given G-grading on R. For a graded R-module U and


g ∈ G, it follows from 13.2.4(iii) that E(*E(U )(g)) ∼
= E(U ). Hence, on use
of this and the preceding paragraph, we see that
 
E(M ) ∼ = E(*E(M )) ∼ =E α∈Λ *E(R/pα )(gα )
∼  
= E(*E(R/pα )(gα )) ∼
α∈Λ = E(R/pα ),
α∈Λ

from which the claim is clear.

We want to extend the result of 14.2.12 to the ‘higher’ terms in minimal


*injective resolutions. The following exercise concerns an important point in
the theory of Bass numbers for ungraded situations.

14.2.13 Exercise. Let R be a commutative Noetherian ring, let M be an


R -module, and let
d0 di
0 −→ E 0 (M ) −→ E 1 (M ) −→ · · · −→ E i (M ) −→ E i+1 (M ) −→ · · ·

be the minimal injective resolution of M , so that there is an augmentation R -


homomorphism α : M → E 0 (M ) such that the sequence
α di
0 −→ M −→ E 0 (M ) −→ · · · −→ E i (M ) −→ E i+1 (M ) −→ · · ·
300 Graded versions of basic theorems

is exact.
Deduce from 11.1.7 that, for each p ∈ Spec(R ), the induced homomor-
phism HomRp (Rp /pRp , αp ) is an isomorphism.
It will also be convenient if we record now some further consequences of
Lemma 13.2.4(iii).
14.2.14 Remarks. Let M be a graded R-module. Let
e0 ei
0 −→ *E 0 (M ) −→ *E 1 (M ) −→ · · · −→ *E i (M ) −→ *E i+1 (M ) −→ · · ·
be the minimal *injective resolution of M , with associated (necessarily ho-
mogeneous) augmentation homomorphism β : M −→ *E 0 (M ). Also, let
α : M −→ E 0 (M ) provide the injective envelope of M .

(i) Since E 0 (M ) is injective, there is a commutative diagram

M
β
- *E 0 (M )

φ0

α
?
M - E 0 (M )

of R-modules and R-homomorphisms. Now *E 0 (M ) is an essential ex-


tension of Im β, by 13.2.4(iii), and so, since Ker φ0 ∩ Im β = 0, it
follows that φ0 is actually a monomorphism.
(ii) Let p ∈ Spec(R) and set k(p) := Rp /pRp . Now
HomRp (k(p), φ0p ) ◦ HomRp (k(p), βp ) = HomRp (k(p), αp ),
which is an isomorphism by 14.2.13; however, the left-exactness of the
‘Hom’ functor ensures that HomRp (k(p), φ0p ) and HomRp (k(p), βp ) are
both monomorphisms, and so it follows that they are both isomorphisms.
Our need for a proof of the following lemma is explained by our ‘first vari-
able only’ approach to the functors * ExtiR ( • , N ) (i ∈ N0 ): see 13.1.8.
14.2.15 Lemma. Let L, M be graded R-modules with M finitely generated.
Let
f0 fi
I • : 0 −→ I 0 −→ I 1 −→ · · · −→ I i −→ I i+1 −→ · · ·
be a *injective resolution of L (in the category *C(R)). Then, for each i ∈ N0 ,
there is a homogeneous isomorphism
* ExtiR (M, L) ∼
= H i (* HomR (M, I • )).
14.2 *Indecomposable *injective modules 301

Proof. For each i ∈ N0 , set K i := Ker f i . Note that there is a homogeneous


isomorphism K 0 ∼ = L. Suppose that j ∈ N with j > 1. For all i = 1, . . . , j −1,
there is an exact sequence

0 −→ K i−1 −→ I i−1 −→ K i −→ 0

of graded R-modules and homogeneous homomorphisms. Since M is a finitely


generated graded R-module, it follows from (13.2.5 and) 13.3.11 that there are
homogeneous isomorphisms

* ExtjR (M, L) ∼ j ∼ * Extj−1 (M, K 1 ) =


= * ExtR (M, K 0 ) = R
∼ ···
∼ * Ext (M, K
= 2 j−2
)∼ 1
= * Ext (M, K j−1
).
R R

This means that it is now enough for us to prove the claim in the statement of
the lemma in the special cases in which i = 0 and i = 1. However, the claims
in these two cases follow easily from the exact sequences

0 −→ * HomR (M, L) −→ * HomR (M, I 0 ) −→ * HomR (M, K 1 )


−→ * Ext1R (M, L) −→ 0

and 0 −→ * HomR (M, K 1 ) −→ * HomR (M, I 1 ) −→ * HomR (M, K 2 ) of


graded R-modules and homogeneous homomorphisms which can also be ob-
tained from 13.2.5 and 13.3.11.

14.2.16 Theorem. Let M be a graded R-module. Let j ∈ N0 , and denote the


j-th term in the minimal *injective resolution of M by *E j (M ), as in 14.2.11.
By 14.2.10 and 14.2.4(iii), there is a family (pα )α∈Λ of graded prime ideals of
R and a family (gα )α∈Λ of elements of G for which there is a homogeneous
isomorphism

=

*E j (M ) −→ *E(R/pα )(−gα ).
α∈Λ

Let p ∈ * Spec(R). Then the cardinality of the set {α ∈ Λ : pα = p} is


equal to the ordinary Bass number μj (p, M ).

Proof. The claim was proved in 14.2.12 in the special case in which j = 0.
We consider next the case in which j = 1.
Use the notation introduced in 14.2.14, and set K i = Ker ei for each i ∈ N0 .
Also, let
d0 di
0 −→ E 0 (M ) −→ E 1 (M ) −→ · · · −→ E i (M ) −→ E i+1 (M ) −→ · · ·

be the (ordinary, ungraded) minimal injective resolution of M (with associated


302 Graded versions of basic theorems

augmentation homomorphism α : M −→ E 0 (M )). For each i ∈ N0 , set


C i := Ker di . Consider the commutative diagram

0 - M - *E 0 (M ) - K1 - 0

φ0 φ0
? ?
0 - M - E 0 (M ) - C1 - 0

with exact rows, in which φ0 is the homomorphism induced by φ0 . Localize


this diagram at p, and then apply the functor HomRp (k(p), • ), noting that
Ext1R (R/p, *E 0 (M )) = 0 by 13.2.5, and that HomRp (k(p), φ0p ) is an isomor-
phism by 14.2.14(ii). The Five Lemma shows that

=
HomRp (k(p), φ0 p ) : HomRp (k(p), (K 1 )p ) −→ HomRp (k(p), (C 1 )p )
is an isomorphism. Now *E 1 (M ) = *E(K 1 ), and so, by 14.2.12, the cardinal-
ity of the set {α ∈ Λ : pα = p} is equal to μ0 (p, K 1 ); we have just shown that
this is μ0 (p, C 1 ), which is equal to μ1 (p, M ) by the theory of Bass numbers in
the ungraded case.
Now suppose that j > 1. By 14.2.15, there is a homogeneous isomorphism
* Ext1R (R/p, K j−1 ) ∼ j
= * ExtR (R/p, M ), so that μ1 (p, K j−1 ) = μj (p, M ).
However,
⊆ dj−1
0 −→ K j−1 −→ *E j−1 (M ) −→ *E j (M )
is the start of the minimal *injective resolution of K j−1 , and so it follows from
what we have already proved in the case in which j = 1 that the cardinality of
the set {α ∈ Λ : pα = p} is equal to μ1 (p, K j−1 ), which we now know to be
equal to μj (p, M ).

14.3 A graded version of the Annihilator Theorem


The purpose of this section is to present graded versions of Faltings’ Annihi-
lator Theorem 9.5.1 and Grothendieck’s Finiteness Theorem 9.5.2. To prepare
for this, we are going to present some results due to S. Goto and K.-i. Watanabe
[22]. Recall that we are assuming that R is G-graded throughout this chapter.
The key points are that, if M is a non-zero finitely generated graded R-module,
then, for each p ∈ Supp M , we have p∗ ∈ Supp M and
dim Mp = dim Mp∗ + ht p/p∗ and depth Mp = depth Mp∗ + ht p/p∗ .
14.3 A graded version of the Annihilator Theorem 303

We begin with a description of the structure of *simple G-graded commuta-


tive Noetherian rings.

14.3.1 Example. Let K be a field and let G be a subgroup of G. Then the


group ring K[G ] is a G-graded commutative ring with

Kg if g ∈ G ,
K[G ]g =
0 if g ∈ G .

Let e1 , . . . , eh be a free base for G as Abelian group, and, for each i =


1, . . . , h, use Ti to denote the element ei = 1ei in K[G ]. Then T1 , . . . , Th
are algebraically independent over K, and

K[G ] ∼
= K[T1 , . . . , Th , T1−1 , . . . , Th−1 ]

under an isomorphism which maps ei to Ti (for i ∈ {1, . . . , h}). Thus K[G ] is


a (Noetherian) regular unique factorization domain. Note also that every non-
zero homogeneous element of K[G ] is a unit, and so K[G ] is a *simple G-
graded commutative Noetherian ring. Our first aim in this section is to establish
the converse statement.

14.3.2 Theorem. (See S. Goto and K.-i. Watanabe [22, Theorem 1.1.4].)
(Recall that R is G-graded.) The following statements are equivalent.

(i) There is a field K and a subgroup G of G such that R is homogeneously


isomorphic to K[G ], where the latter is G-graded in the manner de-
scribed in 14.3.1.
(ii) The G-graded ring R is *simple.
(iii) Every graded R-module is *injective and *free and R is non-trivial.

Proof. (i) ⇒ (ii) This was proved in 14.3.1.


(ii) ⇒ (iii) Assume that R is *simple. This means that the only graded ideals
of R are 0 and R itself. Let M be a graded R-module. It is immediate from the
‘graded Baer criterion’ 13.2.7 that M is *injective. Therefore, by 14.2.10 and
14.2.4(iii), there is a family (pα )α∈Λ of graded prime ideals of R and a family
(gα )α∈Λ of elements of G for which there is a homogeneous isomorphism
= 

M −→ α∈Λ *E(R/pα )(−gα ). However, the only proper graded ideal of R
is 0, and so, since every graded R-module is *injective, we have *E(R/pα ) =
R for all α ∈ Λ. Therefore M is *free.
(iii) ⇒ (i) Assume that every graded R-module is *injective and *free. Let
a be a proper graded ideal of R. By assumption, the graded R-module R/a is
free, and so has zero annihilator. Therefore a = 0. Thus R is *simple.
304 Graded versions of basic theorems

Let g ∈ G. Each non-zero homogeneous element ug ∈ Rg must be in-


vertible in R, with u−1 g ∈ R−g . In particular, each non-zero element of R0 is
invertible in R0 , so that R0 is a field, K say, and, for each g  ∈ G for which
Rg = 0, the R0 -module Rg , that is, the K-vector space Rg , is generated by
any one of its non-zero elements, for if u, v ∈ Rg \{0}, then vu−1 ∈ R0 = K,
so that v ∈ Ku. Also in particular, the product of two non-zero homogeneous
elements of R is again non-zero. Let G := {g ∈ G : Rg = 0}, a subgroup
of G. Let e1 , . . . , eh be a free base for G as Abelian group, and, for each
i = 1, . . . , h, let Ti be a non-zero element of Rei .
Consider K[G ] as G-graded in the manner described in 14.3.1. There is
a homogeneous ring homomorphism φ : K[G ] −→ R which acts as the
identity on K = R0 and maps ei to Ti (for each i = 1, . . . , h). The choice of
G ensures that φ is surjective, and since Ker φ is a proper graded ideal of the
*simple G-graded ring K[G ], we must have that φ is an isomorphism.

14.3.3 Remark. It follows from 14.3.1 and 14.3.2 that, if R is *simple, then
R is a regular unique factorization domain with dim R ≤ rank G.

14.3.4 Corollary. Each *maximal ideal of R is *prime.

Proof. Let m be a *maximal ideal of R. Then the G-graded ring R/m has
exactly two graded ideals, and so is *simple. Therefore R/m is an integral
domain, by 14.3.3.

14.3.5 Proposition. Assume that the G-graded ring R is an integral domain,


and let p ∈ Spec(R). Then ht p = ht p* + ht p/p*.

Proof. Recall that p* ∈ Spec(R). It is clear that ht p ≥ ht p* + ht p/p*; we


shall prove the opposite inequality by induction on h := ht p*. When h = 0,
we have p* = 0 (because R is an integral domain), so that ht p = ht p/p* and
the desired inequality is clear.
So suppose, inductively, that h > 0 and that the result has been proved in
situations where ht p* < h. Denote ht p by n. Let 0 = r be a homogeneous
element of R contained in p*. Then dim Rp /rRp = n − 1 because R is an
integral domain, and so there exists a chain of prime ideals p0 ⊂ p1 ⊂ · · · ⊂
pn−1 = p of R with r ∈ p0 . Now p0 , being a minimal prime ideal of the
graded ideal Rr, must be graded, and so p0 ⊆ p*. In fact, (p/p0 )* = p*/p0 .
Note that ht p/p0 = n − 1 and ht(p/p0 )* ≤ h − 1.
Therefore, if we apply the inductive hypothesis to the prime ideal p/p0 in
the G-graded integral domain R/p0 , we see that

ht p/p0 = ht(p/p0 )* + ht(p/p0 )/(p/p0 )* = ht(p/p0 )* + ht p/p*,


14.3 A graded version of the Annihilator Theorem 305

so that n − 1 ≤ h − 1 + ht p/p*. Therefore ht p ≤ ht p* + ht p/p*, and the


inductive step is complete.
The next theorem, again due to Goto and Watanabe, is a strengthening of
14.3.5.
14.3.6 Theorem. (See S. Goto and K.-i. Watanabe [22, Proposition 1.2.2].)
(The reader is reminded that R is G-graded.) Let p ∈ Supp M where M is a
graded R-module. Recall from 13.1.6(i) that p* ∈ Supp M . We have
dim Mp = dim Mp∗ + ht p/p*.
Proof. For q ∈ Supp M , we shall denote dim Mq by htM q. It is clear that
htM p ≥ htM p*+ht p/p*; we shall prove the opposite inequality by induction
on h := htM p*.
First suppose that h = 0. After homogeneous localization at p, we can, and
do, assume that (R, p*) is *local. Since h = 0, it follows that p* is the one and
only graded prime ideal in Supp M ; therefore Ass M = {p*}. Hence each
q ∈ Supp M such that q ⊆ p must have q* = p*, and it is clear from this that
htM p = ht p/p*.
Now suppose, inductively, that h > 0 and the desired result has been proved
for smaller values of h. We can again assume that (R, p*) is *local. Denote
htM p by n and ht p/p* by d. Since n ≥ h + d, we have n > d.
There exists a chain of prime ideals p0 ⊂ p1 ⊂ · · · ⊂ pn = p in Supp M .
Set q := pn−d ; then htM q = n − d and ht p/q = d. If q ⊆ p*, then n − d =
htM q ≤ htM p* = h and n ≤ d + h, as required. So we suppose that q ⊆ p*,
so that q is not graded because (R, p*) is *local. We do have q* ⊆ p* and
q* ∈ Supp M . Let t := ht q/q* and u := ht p*/q*. Note that t > 0 because q
is not graded.
By 14.3.5 applied to the G-graded integral domain R/q*, we have
ht p/q* = ht(p/q*)* + ht(p/q*)/(p/q*)* = ht p*/q* + ht p/p* = u + d.
But ht p/q* ≥ ht p/q+ht q/q* = d+t. Therefore u ≥ t > 0. This means that
q* is strictly contained in p*, so that htM q* < h. The inductive hypothesis
therefore yields that n − d = htM q ≤ htM q* + ht q/q* = htM q* + t. Now
htM q* + u = htM q* + ht p*/q* ≤ htM p* = h. Therefore
n − d ≤ htM q* + t ≤ h − u + t,
so that
n ≤ d + h + (t − u) ≤ d + h
since u ≥ t. This completes the inductive step.
306 Graded versions of basic theorems

In addition to Theorem 14.3.6, we would like to have available the compan-


ion result, also due to Goto and Watanabe, that, if M is a finitely generated
graded R-module and p ∈ Supp M , then depth Mp = depth Mp∗ + ht p/p*.
We shall obtain this as a corollary of the following result about Bass numbers.

14.3.7 Theorem. (See S. Goto and K.-i. Watanabe [22, Theorem 1.2.3].)
(Recall that R is G-graded.) Let M be a graded R-module. Let p ∈ Spec(R)
and let d := ht p/p*. Then

i 0 if 0 ≤ i < d,
μ (p, M ) =
μ (p*, M ) if i ≥ d.
i−d

Proof. By homogeneous localization at p, and use of the ‘invariance’ of Bass


numbers under fraction formation (see [50, p.150]), we reduce to the case
where (R, p*) is *local. Then R/p* is a *simple G-graded ring, and so, by
14.3.2, every graded R/p*-module is *free. Also, d is the dimension of the
local ring Rp /p*Rp , and this is regular, by 14.3.3. Let u1 , . . . , ud ∈ p be such
that their natural images in Rp /p*Rp , which we denote by α1 , . . . , αd respec-
tively, generate the maximal ideal of that regular local ring. Thus

pRp = p*Rp + (α1 , . . . , αd )Rp .

For each j = 0, . . . , d, set Qj := p*Rp + (α1 , . . . , αj )Rp , and note that


this is a prime ideal of Rp . Therefore, if j < d there is an exact sequence of
(ungraded) Rp -modules and Rp -homomorphisms
αj+1
0 - Rp /Qj - Rp /Qj - Rp /Qj+1 - 0,

and this induces a long exact sequence

0 - HomRp (Rp /Qj+1 , Mp ) - HomRp (Rp /Qj , Mp )


αj+1
- HomRp (Rp /Qj , Mp ) - Ext1Rp (Rp /Qj+1 , Mp )
- ···
- ExtiR (Rp /Qj+1 , Mp ) - ExtiRp (Rp /Qj , Mp )
p
αj+1
- ExtiRp (Rp /Qj , Mp ) - Exti+1
Rp (Rp /Qj+1 , Mp )

- ···.

Note that Q0 = p*Rp . Let i ∈ N0 . Now ExtiR (R/p*, M ) is the R-module


underlying the graded R-module * ExtiR (R/p*, M ). The latter has a natural
structure as graded R/p*-module, and we observed above that every graded
14.3 A graded version of the Annihilator Theorem 307

R/p*-module is *free. Consequently, ExtiR (R/p*, M ) is a free R/p*-module;


therefore α1 is a non-zerodivisor on ExtiRp (Rp /Q0 , Mp ) and, in view of the
above long exact sequence (in the case where j = 0), we have
ExtiRp (Rp /Q1 , Mp ) ∼
= Exti−1 i−1
Rp (Rp /Q0 , Mp )/α1 ExtRp (Rp /Q0 , Mp )

(as Rp -modules) for all i ∈ Z. (When i is negative, the statement is obviously


true.) Our immediate aim is to show that
ExtiRp (Rp /Qj , Mp )
∼ i−j i−j
= ExtRp (Rp /Q0 , Mp )/(α1 , . . . , αj ) ExtRp (Rp /Q0 , Mp )
for all i ∈ Z for all j = 2, . . . , d.
So suppose, inductively, that j ∈ {1, . . . , d − 1} and we have shown that
ExtiRp (Rp /Qj , Mp )
∼ i−j i−j
= ExtRp (Rp /Q0 , Mp )/(α1 , . . . , αj ) ExtRp (Rp /Q0 , Mp )

(as Rp -modules) for all i ∈ Z. Since ExtkRp (Rp /Q0 , Mp ) is a free Rp /Q0 -
module for all k ∈ Z, it follows that ExtiRp (Rp /Qj , Mp ) is a free Rp /Qj -
module for all i ∈ Z. For each i ∈ Z it therefore follows that αj+1 is a non-
zerodivisor on Exti−1
Rp (Rp /Qj , Mp ), so that, in view of the above long exact
sequence,
ExtiRp (Rp /Qj+1 , Mp ) ∼
= Exti−1 i−1
Rp (Rp /Qj , Mp )/αj+1 ExtRp (Rp /Qj , Mp )

(as Rp -modules). We can now use the inductive hypothesis to deduce that
ExtiRp (Rp /Qj+1 , Mp )
∼ i−j−1 i−j−1
= ExtRp (Rp /Q0 , Mp )/(α1 , . . . , αj+1 ) ExtRp (Rp /Q0 , Mp )
(as Rp -modules) for all i ∈ Z. This completes the inductive step.
Since Qd = pRp , it therefore follows, by induction, that
ExtiRp (Rp /pRp , Mp )

= Exti−d (Rp /Q0 , Mp )/(α1 , . . . , αd ) Exti−d (Rp /Q0 , Mp )
Rp Rp

= Exti−d i−d
Rp (Rp /p*Rp , Mp )/(α1 , . . . , αd ) ExtRp (Rp /p*Rp , Mp )

(as Rp -modules) for all i ∈ Z.


For each prime ideal q of R, denote the field Rq /qRq by k(q). Let i ∈ Z.
Recall from [50, Theorem 18.7] that
 
μi (q, M ) = dimk(q) ExtiRq (k(q), Mq ) = dimk(q) ExtiR (R/q, M ) q .

(We interpret μj (q, M ) for negative j as zero.) Now Exti−d


R (R/p*, M ) is
308 Graded versions of basic theorems

a free R/p*-module, and it follows from that that Exti−d Rp (Rp /p*Rp , Mp )
is a free Rp /p*Rp -module of the same rank. At this point, recall also that
pRp = p*Rp + (α1 , . . . , αd )Rp . It therefore follows from the above inductive
argument that
 
μi−d (p*, M ) = dimk(p∗ ) Exti−d R (R/p*, M ) p∗

= rankR/p∗ Exti−d
R (R/p*, M )

= rankRp /p∗ Rp Exti−d


Rp (Rp /p*Rp , Mp )

= rankk(p) ExtiRp (Rp /pRp , Mp )


= dimk(p) ExtiRp (k(p), Mp ) = μi (p, M ).
In particular, μi (p, M ) = 0 for all i < d.
When discussing depths of localizations of a finitely generated R-module,
we shall follow the notation and conventions of 9.2.1.
14.3.8 Corollary. Let M be a finitely generated graded R-module. Let p ∈
Spec(R). Then
depth Mp = depth Mp∗ + ht p/p*.
Proof. Recall from from [50, Theorems 16.7 and 18.7] that, for q ∈ Spec(R),
depth Mq is the least integer i such that μi (q, M ) = 0 (if any such integers
exist, and ∞ otherwise). The claim is therefore immediate from 14.3.7.
14.3.9 Exercise. (Recall that R is G-graded.)

(i) Show that the following statements are equivalent:


(a) R is Cohen–Macaulay;
(b) R(p) is Cohen–Macaulay for all p ∈ * Spec(R);
(c) Rp is Cohen–Macaulay for all p ∈ * Spec(R).
(ii) Show that the following statements are equivalent:
(a) R is Gorenstein;
(b) R(p) is Gorenstein for all p ∈ * Spec(R);
(c) Rp is Gorenstein for all p ∈ * Spec(R).

We are now in a position to present G-graded versions of Faltings’ Anni-


hilator Theorem 9.5.1 and Grothendieck’s Finiteness Theorem 9.5.2. Loosely,
these graded versions say that, in the G-graded case, both theorems can be re-
formulated so that only graded prime ideals need be considered. As the Graded
Finiteness Theorem has important geometric significance in connection with
the cohomology of projective schemes, we provide a proof.
14.4 Graded local duality 309

14.3.10 Graded Annihilator and Finiteness Theorems. (Recall that R is


G-graded.) Assume that R is a homomorphic image of a regular (commutative
Noetherian) ring. Assume that the ideal a is graded; let b be a second graded
ideal of R. Let M be a finitely generated graded R-module. Then
fab (M ) = inf {adja depth Mp : p ∈ * Spec(R) \ Var(b)}
= inf {depth Mp + ht(a + p)/p : p ∈ * Spec(R) \ Var(b)} .
In particular,
fa (M ) = inf {depth Mp + ht(a + p)/p : p ∈ * Spec(R) \ Var(a)} .
Proof. By Faltings’ Annihilator Theorem 9.5.1,
fab (M ) = inf {depth Mp + ht(a + p)/p : p ∈ Spec(R) \ Var(b)} .
It is therefore sufficient for us to show that, for each non-graded prime ideal
p ∈ Spec(R) \ Var(b) for which depth Mp is finite, the graded ideal p* is
such that p* ∈ * Spec(R) \ Var(b) and
depth Mp∗ + ht(a + p*)/p* ≤ depth Mp + ht(a + p)/p.
This we do.
Of course p* is a graded prime ideal of R (see 13.1.3(vi)), and it is clear that
p* ⊇ b. Let d := ht p/p*; by 14.3.8, depth Mp = depth Mp∗ + d.
Let t := ht(a+p)/p and let q be a minimal prime of a+p such that ht q/p =
ht(a + p)/p. Since R is a homomorphic image of a regular ring, it is catenary,
and therefore ht q/p* = ht q/p + ht p/p* = t + d. Since a + p* ⊆ a + p ⊆ q,
it follows that ht(a + p*)/p* ≤ ht q/p* = t + d. Therefore
depth Mp∗ + ht(a + p*)/p* ≤ depth Mp − d + t + d
= depth Mp + ht(a + p)/p,
and this completes the proof.

14.4 Graded local duality


The purpose of this section is to develop a G-graded analogue of the Local
Duality Theorem 11.2.6. This will concern the situation where (R, m) is a
*local G-graded ring that can be expressed as a homomorphic image of a
Gorenstein *local G-graded (commutative Noetherian) ring by means of a ho-
mogeneous homomorphism. Many of the G-graded rings that occur in appli-
cations are finitely generated algebras over fields, and so we are only imposing
a mild restriction.
310 Graded versions of basic theorems

Recall that we are assuming that R is G-graded throughout this chapter.

14.4.1 Graded Local Duality Theorem. Assume that (R, m) is *local with
ht m = n. Assume also that there is a Gorenstein *local G-graded com-
mutative Noetherian ring (R , m ) and a surjective homogeneous ring ho-
momorphism f : R −→ R. Let ht m = n . Let *D denote the functor
* HomR ( • , *E(R/m)) from *C(R) to itself.
Let M be a graded R-module, let N  be a graded R -module, and let j ∈
N0 . Now M can be regarded as a graded R -module by means of f ; then,
* ExtjR (M, N  ) has a natural structure as graded R-module.
n  ∼   
There exists g ∈ G such that Hm  (R ) = *ER (R /m )(−g) in *C(R ). For

any such g, there is a homomorphism


 i n −i
Ψ := (ψ i )i∈N0 : Hm i∈N0
−→ *D(* ExtR  ( • , R (g)))
i∈N0

of negative connected sequences of covariant functors from *C(R) to *C(R)


i
which is such that ψM is a (necessarily homogeneous) isomorphism for all
i ∈ N0 whenever M is a finitely generated graded R-module.

Proof. We shall first deal with the special case where R = R (and f is
the identity map), so that (R, m) itself is a Gorenstein *local G-graded ring.
Since dim Rm = n , it follows from Grothendieck’s Vanishing Theorem 6.1.2
that (Hmj
(N ))m ∼ j
= HmRm (Nm ) = 0 for all j > n and all R-modules N .
j
Therefore, by 14.1.2(ii), we have Hm (M ) = 0 for all j > n and all graded
n
R-modules M .
 n−i Consequently, H m is a right exact functor from *C(R) to it-
self, and Hm i∈N0
is a positive strongly connected sequence of covariant
functors from *C(R) to *C(R).
Let g ∈ G. Note that the long complex resultingfrom application of the
positive connected sequence *D(* ExtiR ( • , R(g))) i∈N0 to a short exact se-
quence of graded R-modules and homogeneous homomorphisms

0 −→ L −→ M −→ N −→ 0
 
is exact. Thus *D(* ExtiR ( • , R(g))) i∈N0 is a positive strongly connected
sequence of covariant functors from *C(R) to itself.
n
Our next aim is to find a natural equivalence between the functors Hm and
*D(* HomR ( • , R(g))) (from *C(R) to itself) for a suitable g ∈ G. Since R
is Gorenstein, it follows from 14.2.16 that there is a family (gp )p∈∗ Spec(R) of
elements of G such that there is a homogeneous isomorphism

*ERi
(R) ∼= *ER (R/p)(−gp ) for each i ∈ N0 .
p∈∗ Spec(R)
ht p=i
14.4 Graded local duality 311
n
We can calculate the graded module Hm (R) by applying Γm to the minimal
*injective resolution of R and then taking cohomology; in view of 14.2.4(iv),
n
the result is that Hm (R) ∼
= *ER (R/m)(−gm ) in *C(R).
We can now modify the ideas of 11.2.4 to establish the existence of a natural
transformation
n
φ : Hm −→ * HomR (* HomR ( • , R), Hm
n
(R))

of functors from *C(R) to itself, for which, for each graded R-module M , we
have (φM (y))(f ) = Hm n
(f )(y) for all y ∈ Hmn
(M ) and f ∈ * HomR (M, R).
One easily checks that φR is an isomorphism.
Choose g ∈ G such that Hm n
(R) ∼= *ER (R/m)(−g) in *C(R). (For exam-
ple, we could take g := gm .) We obtain from φ a natural transformation
n
ψ0 : Hm −→ * HomR (* HomR ( • , R), *ER (R/m)(−g))
= * HomR (* HomR ( • , R(g)), *ER (R/m))

of functors from *C(R) to itself, for which ψ0 R is an isomorphism. Note that,


since R is Gorenstein, (Hm n−i
(R))m ∼
= HmRn−i
m
(Rm ) = 0 for all i > 0, so that
Hm (R) = 0 for all i > 0 by 14.1.2(ii). Also * ExtiR (P, R (g)) = 0 for
n−i

all i > 0 for all *projective graded R-modules P . It follows from the graded
version of the analogue of 1.3.4 for positive connected sequences that ψ0 can
be incorporated into a (uniquely determined) homomorphism
 n−i   
Φ := (φi )i∈N0 : Hm i∈N0
−→ *D(* ExtiR ( • , R(g))) i∈N0

of positive connected sequences of covariant functors from *C(R) to *C(R).


Furthermore, it is easy to prove by induction that, for each i ∈ N, the homo-
morphism φi M is an isomorphism whenever M is a finitely generated graded
R-module: use the fact that such an M can be included in an exact sequence
0 −→ K −→ F −→ M −→ 0 in *C(R) in which F is a finitely generated
*free R-module. Note that we can interpret
 i  
(φn−i )i∈N0 : Hm i∈N0
−→ *D(* ExtR n−i
( • , R(g))) i∈N0

as a homomorphism of negative connected sequences.


We have thus established the claims of the theorem in the case where R =

R . We now deal with the general case, where there is a Gorenstein *local G-
graded commutative Noetherian ring (R , m ) and a surjective homogeneous
ring homomorphism f : R −→ R. Let *D := * HomR ( • , *ER (R /m )).
Our work so far in this proof establishes the existence of a g ∈ G such that
n  ∼   
Hm  (R ) = *ER (R /m )(−g) in *C(R ) and, for such a g, the existence of a
312 Graded versions of basic theorems

homomorphism
 i  
Ψ := (ψ i )i∈N0 : Hm 
i∈N0
−→ *D (* ExtnR−i ( • , R (g)))
i∈N0

of negative connected sequences of covariant functors from *C(R ) to *C(R )




such that φi M is a (necessarily homogeneous) isomorphism whenever M is a


finitely generated graded R -module. To complete the proof, we use the argu-
ment in the proof of the Local Duality Theorem 11.2.6, modified for the G-
graded context. Important ingredients include the Graded Independence The-
orem 14.1.7, and the fact that the graded submodule (0 :*E  (R /m ) Ker f )
R
of *ER (R /m ) is, when viewed as a graded R-module, homogeneously iso-
morphic to *ER (R/m).

Note. It should be noted that the element g in the statement of the Graded
Local Duality Theorem 14.4.1 need not be uniquely determined. The interested
reader is referred to [6, Lemma 1.5].
In the situation of 14.4.1, the functor *D = * HomR ( • , *E(R/m)) would
seem to be an obvious graded analogue of the functor of 10.2.1 used to con-
struct Matlis duals over a local ring. However, there is another approach to *D
which is particularly useful when the 0-th component R0 of the *local ring R
is a field, and this approach is the subject of the next exercise. The exercise is
based on [7, Proposition 3.6.16], but our approach is a little different from that
of Bruns and Herzog; also, we are working in the G-graded context, whereas
Bruns’ and Herzog’s treatment is for Z-graded rings.

14.4.2 Exercise: Graded Matlis Duality. Suppose (R, m) is *local. Note


that R0 is local with maximal ideal m0 := m ∩ R0 ; let E0 := ER0 (R0 /m0 ).

(i) Let M = g∈G Mg be a graded R-module. Now R0 can be consid-
ered as a G-graded ring with trivial grading, and any R0 -module can
be considered as a graded R0 -module concentrated in degree 0. Also,
the grading on our graded R-module M provides a structure as graded
R0 -module on M .
Show that M ∨ := * HomR0 (M, E0 ) has a natural structure as a
graded R-module with g-th component (M ∨ )g = HomR0 (M−g , E0 )
for all g ∈ G. (Remember that * HomR0 (M, E0 ) is a submodule of
HomR0 (M, E0 ), and the latter has a natural structure as an R-module.)
Deduce that ( • )∨ is an exact, additive functor from *C(R) to itself.
(ii) Show that the functors ( • )∨ and * HomR ( • , R∨ ), that is, the functors
* HomR0 ( • , E0 ) and * HomR ( • , * HomR0 (R, E0 )), from *C(R) to
itself, are naturally equivalent. (Here is a hint: if we forget the gradings,
14.5 *Canonical modules 313

there is a natural equivalence


κ : HomR0 ( • , E0 ) −→ HomR ( • , HomR0 (R, E0 ));
for a graded R-module M , consider an appropriate restriction of κM .)
(iii) Let φ0 : R −→ E0 be the R0 -homomorphism whose restriction to R0
is the composition of the natural epimorphism R0 → R0 /m0 and the
inclusion R0 /m0 → E0 = ER0 (R0 /m0 ), and whose restriction to Rg
for all g ∈ G \ {0} is zero. Show that R∨ is a *essential extension of its
graded submodule Rφ0 .
(iv) Show that R∨ ∼ = *ER (R/m) (in the category *C(R)), and deduce that
the functors ( • )∨ and *D := * HomR ( • , *E(R/m)) (from *C(R) to
itself) are naturally equivalent.
(v) Show that, if the local ring R0 is complete, then, whenever M is a finitely
generated graded R-module, there is a homogeneous R-isomorphism
M∼ = (M ∨ )∨ =: M ∨∨ .
Thus, in the situation of Exercise 14.4.2, it is reasonable for us to regard
the functor ( • )∨ = * HomR0 ( • , ER0 (R0 /m0 )) as the ‘graded Matlis duality
functor’. In the particular case in which R0 is a field K (and this is the situation
in many practical applications of graded ring theory), ER0 (R0 /m0 ) = K, and,

for a graded R-module M = g∈G Mg , the grading of the ‘graded Matlis
dual’ is given by the attractively simple formula

M ∨ = * HomK (M, K) = HomK (M−g , K).
g∈G

14.5 *Canonical modules


Recall that we are assuming that R is G-graded throughout this chapter. Also,
recall from 12.1.2 that a canonical module for a local ring (R , m ) is a finitely
generated R -module K such that HomR (K, ER (R /m )) ∼ = Hm dim R
 (R ).
The obvious graded analogue is given in the following definition.
14.5.1 Definition. Suppose that (R, m) is *local; let ht m = n. A *canonical
module for R is a finitely generated graded R-module C for which there is a
homogeneous isomorphism
* HomR (C, *ER (R/m)) ∼ n
= Hm (R).
14.5.2 Example. Assume that (R, m) is *local with ht m = n, and that there
is a Gorenstein *local G-graded commutative Noetherian ring (R , m ) with
ht m = n and a surjective homogeneous ring homomorphism f : R −→ R.
314 Graded versions of basic theorems

It follows from the Graded Local Duality Theorem 14.4.1 that, for some
g ∈ G, there is a homogeneous isomorphism
n −n
* Hom(* ExtR  (R, R (g)), *ER (R/m)) ∼ n
= Hm (R).

n −n
Therefore * ExtR  (R, R (g)) is a *canonical module for R.
14.5.3 Proposition. Assume that (R, m) is *local with a *canonical module
C. Then Cm is a canonical module for Rm .
Proof. By 13.2.4(iii), the R-module *ER (R/m) (with its grading forgot-
ten) is an essential extension of R/m, and so there is a monomorphism α :
*ER (R/m) −→ ER (R/m). Denote *ER (R/m) by *E and ER (R/m) by E.
Let K := Coker α, and let k(m) denote the residue field of the local ring Rm .
When the grading on *E is forgotten, we have
Ext1Rm (k(m), (*E)m ) ∼
= (Ext1R (R/m, *E))m = 0,
α
by 13.2.5. The exact sequence 0 −→ *E −→ E −→ K −→ 0 therefore
induces an exact sequence
0 −→ HomRm (k(m), (*E)m ) −→ HomRm (k(m), Em )
−→ HomRm (k(m), Km ) −→ 0.
Since HomRm (k(m), Em ) is a 1-dimensional vector space over k(m) and
HomRm (k(m), (*E)m ) = 0,
we must have HomRm (k(m), Km ) = 0; therefore Km = 0, since Km is an
Artinian Rm -module.
Therefore αm is an isomorphism.
Since C is a *canonical module for R, there is a homogeneous R-isomorph-
ism * HomR (C, *ER (R/m)) ∼ = Hm n
(R), where n = ht m. Now forget the
gradings, localize at m, and use the isomorphism αm : there result isomor-
phisms of Rm -modules
HomRm (Cm , ERm (Rm /mRm )) ∼
= HomRm (Cm , (ER (R/m))m )

= HomR (Cm , (*ER (R/m))m )
m


= (HomR (C, *ER (R/m)))m

= (H n (R))m ∼
= H n (Rm ).
m mRm

Therefore Cm is a canonical module for Rm .


Limitations on space mean that we are not able, in this book, to develop the
theory of *canonical modules in generality similar to that of Chapter 12. As
14.5 *Canonical modules 315

a fairly short treatment is possible in the Cohen–Macaulay case, we content


ourselves with that.
14.5.4 Theorem. Assume that (R, m) is *local and Cohen–Macaulay and
that C is a finitely generated graded R-module such that μi (m, C) = δi,ht m
for all i ∈ N0 . Then Cp is a canonical module for Rp for all (graded or
ungraded) p ∈ Spec(R); that is, C is a canonical module for R in the sense of
12.1.28. Consequently, for all p ∈ Spec(R), we have μi (p, C) = δi,ht p for all
i ∈ N0 .
Proof. It follows from 12.1.27 that Cm is a canonical module for Rm . Let
p ∈ Spec(R), so that p* ∈ * Spec(R) and p* ⊆ m. It also follows from
12.1.27 that Cp∗ is a canonical module for Rp∗ , so that μi (p*, C) = δi,ht p∗
for all i ∈ N0 . It now follows from the Goto–Watanabe Theorems 14.3.6 and
14.3.7 that μi (p, C) = δi,ht p for all i ∈ N0 . Therefore Cp is a canonical
module for Rp , by 12.1.27 again.
14.5.5 Corollary. Assume that (R, m) is *local and Cohen–Macaulay and
that C is a *canonical module for R. Then C is a canonical module for R in
the sense of 12.1.28.
Proof. It follows from 14.5.3 that Cm is a canonical module for Rm . There-
fore, by 12.1.26, we have μi (m, C) = μi (mRm , Cm ) = δi,ht m for all i ∈ N0 .
Now 14.5.4 shows that C is a canonical module for R.
14.5.6 Lemma. Assume that R is Cohen–Macaulay and has a canonical
module C. We saw in 12.1.30 that the trivial extension R = R ∝ C of R by C
is a Gorenstein ring, so that R is a homomorphic image of a Gorenstein ring.
If C = ⊕g∈G Cg is graded, then R is G-graded in such a way that the
canonical homomorphism φ : R −→ R is homogeneous. If, in addition, R is
*local, then so too is R .
Proof. The decomposition R = ⊕g∈G (Rg ⊕ Cg ) provides a grading on R ,
and φ is homogeneous with respect to this grading and the original grading on
R. Furthermore, if m is the unique *maximal graded ideal of R, then a routine
check shows that m ⊕ C is the unique *maximal graded ideal of R .
The following exercise will be very helpful in our development of the theory
of *canonical modules.
14.5.7 Exercise. Let α : M −→ N be a homogeneous homomorphism of
graded R-modules.
(i) Show that, if αp : Mp −→ Np is an isomorphism for all p ∈ * Spec(R),
then α is an isomorphism.
316 Graded versions of basic theorems

(ii) Show that, if (R, m) is *local and αm : Mm −→ Nm is an isomorphism,


then α is an isomorphism.

When (R, m) is *local and Cohen–Macaulay, a *canonical module C for R


is a canonical module for R (by 14.5.5), and so satisfies μi (p, C) = δi,ht p for
all i ∈ N0 and p ∈ Spec(R). We are now going to study a general finitely gen-
erated graded R-module with the property that, when its grading is forgotten,
it is a canonical module for R.

14.5.8 Lemma. Assume (R, m) is *local and Cohen–Macaulay; set n :=


ht m. Suppose that C is a finitely generated graded R-module which is a
canonical module for R in the sense of 12.1.28, so that, for all p ∈ * Spec(R),
we have μi (p, C) = δi,ht p .

(i) There is a family (gp )p∈∗ Spec(R) of elements of G for which there exist
homogeneous isomorphisms

*E i (C) ∼
= *E(R/p)(−gp ) for all i ∈ N0 .
p∈∗ Spec(R)
ht p = i

(ii) For any family (gp )p∈∗ Spec(R) of elements of G as in part (i), there are
homogeneous isomorphisms
(a) * ExtiR (R/m, C) ∼
= 0 for i = n,
(b) * ExtR (R/m, C) ∼
n
= (R/m)(−gm ), and
(c) Hmn
(C) ∼= *E(R/m)(−gm ).

Proof. (i) This is immediate from Theorem 14.2.16.


(ii) Let E • denote the minimal *injective resolution of C.
For each p ∈ * Spec(R) \ {m}, there exists a homogeneous element rp ∈
m \ p, and so it follows from 14.2.4(iv) that * HomR (R/m, *E(R/p)) = 0
and Γm (*E(R/p)) = 0. Hence the complex * HomR (R/m, E • ) has all terms
other than its n-th equal to 0, while its n-th term is isomorphic (in *C(R))
to * HomR (R/m, *E(R/m)(−gm )). Parts (a) and (b) therefore follow from
14.2.15 and 14.2.9, while part (c) is a consequence of the fact (see 13.4.3) that
n
we can calculate the graded R-module Hm (C) by application of the functor

Γm to E .

Recall that, in 12.1.6, we proved that any two canonical modules for a local
ring are isomorphic. We are now going to address the analogous issue in the
Cohen–Macaulay G-graded *local case.

14.5.9 Theorem. Suppose (R, m) is Cohen–Macaulay and *local,


14.5 *Canonical modules 317

and let C, C  be graded R-modules that are canonical modules for R in the
sense of 12.1.28. Let n := ht m. Then

(i) the map β : R −→ * HomR (C, C) defined by β(r) = r IdC for all
r ∈ R is a homogeneous isomorphism; and

(ii) there exist g ∈ G and a homogeneous isomorphism φ : C −→ C  (g).
=

Proof. (i) For each p ∈ Spec(R), the Rp -module Cp is a canonical module


for Rp . We can now deduce from 12.2.6 that βp is an isomorphism for all
p ∈ Spec(R). Hence β is an isomorphism, and it is clearly homogeneous.
(ii) Let H denote the graded R-module * HomR (C, C  ). It follows from part
(i) above and 12.1.6 that Hp is a free Rp -module of rank 1, for all (graded or
ungraded) p ∈ Spec(R).
Let {φ1 , . . . , φt } be a generating set, consisting of t homogeneous elements,
for H, that is minimal in the sense that no proper subset of it also generates H.
Let deg φi = gi for i = 1, . . . , t.
t
There is a graded *free R-module F , of the form i=1 R(−gi ), and a ho-
mogeneous R-epimorphism ψ : F −→ H which maps the generator, ei say,
of degree gi in R(−gi ) to φi .
Let K := Ker ψ, a graded submodule of F . Our immediate aim is to
show that K ⊆ mF . Suppose that this is not the case; then there exists a
homogeneous element k ∈ K \ mF . Since k is homogeneous, we can write
t
k = i=1 ri ei , where, for i ∈ {1, . . . , t}, the element ri ∈ R is homogeneous
of degree deg k − gi . Then there exists j ∈ {1, . . . , t} such that rj ∈ m, so
that, since rj is homogeneous, it must be a unit of R. Application of ψ there-
t
fore yields that 0 = ψ(k) = i=1 ri φi , so that φj is an R-linear combination
of the other φi and we have a contradiction to the minimality.
Therefore K ⊆ mF , and it follows from this that the maps F/mF −→
H/mH and Fm /mRm Fm −→ Hm /mRm Hm induced by ψ are both isomor-
phisms. Therefore Fm /mRm Fm and Hm /mRm Hm have equal dimensions as
vector spaces over Rm /mRm .
Now Ext1R (H, K)p ∼ = Ext1Rp (Hp , Kp ) = 0 for all p ∈ Spec(R), and so
1
* ExtR (H, K) = 0. Therefore the exact sequence
⊆ ψ
0 −→ K −→ F −→ H −→ 0

splits in *C(R). It follows that there is an Rm -isomorphism

Fm /mRm Fm ∼
= Km /mRm Km ⊕ Hm /mRm Hm ,

and therefore our calculations above with vector space dimensions show that
318 Graded versions of basic theorems

Km /mRm Km = 0. Therefore Km = 0 by Nakayama’s Lemma, so that K = 0


by 14.1.2(ii).
Hence H is *free. Its rank must be 1, so that there exists a homogeneous
element φ ∈ H, of degree g say, which forms a base for H. Thus φ : C −→
C  (g) is a homogeneous R-homomorphism. For each p ∈ Spec(R), the Rp -
homomorphism φp : Cp −→ Cp generates HomRp (Cp , Cp ). By 12.1.6, there
is an Rp -isomorphism λ : Cp −→ Cp in HomRp (Cp , Cp ); therefore, there
exist r ∈ R and s ∈ R \ p such that λ = (r/s)φp . It follows that φp is
surjective, so that λ−1 ◦ φp : Cp −→ Cp is a surjective endomorphism, and
therefore an isomorphism. Consequently, φp is an isomorphism. As this is true
for each p ∈ Spec(R), it follows that φ : C −→ C  (g) is a homogeneous
isomorphism.
14.5.10 Theorem. Assume that (R, m) is Cohen–Macaulay and *local with
ht m = n, and admits a *canonical module C. Set *E := *E(R/m), and let
*D := * HomR ( • , *E).
There is a natural transformation
n
φ 0 : Hm −→ *D(* HomR ( • , C))
of functors from *C(R) to itself which is such that φ0 M is an isomorphism
whenever M is a finitely generated graded R-module.
There is a unique extension of φ0 to a homomorphism
 n−i   
Φ := (φi )i∈N0 : Hm i∈N0
−→ *D(* ExtiR ( • , C)) i∈N0
of (positive strongly) connected sequences of covariant functors from *C(R) to
*C(R). Furthermore, φi M is an isomorphism for all i ∈ N0 whenever M is a
finitely generated graded R-module.
In particular, for each finitely generated graded R-module M , there are
homogeneous isomorphisms
n−i
Hm (M ) ∼
= *D(* ExtiR (M, C)) for all i ∈ Z.
Proof. As in the proof of the Graded Local Duality Theorem 14.4.1, we can
j n
show that Hm (M ) = 0 for all j > n and all gradedR-modules M , that Hm
n−i
is a right exact functor from *C(R) to itself, and that Hm i∈N0
is a positive
strongly connected sequence of covariant functors from *C(R) to *C(R).
Next, the ideas of 6.1.9 and 6.1.10 can be modified to show that the func-
tors Hm n
and ( • ) ⊗R Hm n
(R), from *C(R) to itself, are naturally equivalent.
Since C is a *canonical module for R, there is a homogeneous isomorphism
* HomR (C, *E) ∼ = Hmn
(R); therefore, the functors
n
Hm and ( • ) ⊗R * HomR (C, *E),
14.5 *Canonical modules 319

from *C(R) to itself, are naturally equivalent.


Next, recall the natural transformation of functors
ξ • , • , • : ( • ) ⊗R HomR ( • , • ) −→ HomR (HomR ( • , • ), • )
(from C(R) × C(R) × C(R) to C(R)) of 10.2.16: it is such that, for R-modules
M , I and J, we have (ξM,I,J (m ⊗ f )) (g) = f (g(m)) for m ∈ M , f ∈
HomR (I, J) and g ∈ HomR (M, I). Take C for I and *E for J; then, when
M is graded, ξM,C,*E maps M ⊗R * HomR (C, *E) into
* HomR (* HomR (M, C), *E);
one can easily check that degrees are preserved. Set ψ0 M := ξM,C,*E for each
graded R-module M . Then
ψ0 : ( • ) ⊗R * HomR (C, *E) −→ * HomR (* HomR ( • , C), *E)
is a natural transformation of functors from *C(R) to itself. Moreover, one can
modify the argument in the proof of 10.2.16 to show that ψ0 M is an isomor-
phism whenever M is a finitely generated graded R-module: use of 13.2.5(i)
enables one to see that, whenever A −→ B −→ L is an exact sequence in the
category *C(R), then the induced sequence
* HomR (L, *E) −→ * HomR (B, *E) −→ * HomR (A, *E)
is again exact.
We can compose ψ0 with a natural equivalence from the second paragraph
of this proof to obtain a natural transformation
n
φ 0 : Hm −→ *D(* HomR ( • , C))
of functors (from *C(R) to itself) with the property that φ0 M is an isomor-
phism whenever M is a finitely generated graded R-module.
We now reason as in the proof of 14.4.1. Since R is Cohen–Macaulay,
(Hmn−i
(R))m ∼= HmR n−i
m
(Rm ) = 0 for all i > 0, so that Hm n−i
(R) = 0
i
for all i > 0 by 14.1.2(ii). Also * ExtR (P, C) = 0 for all i > 0 for all
*projective graded R-modules P . It follows from the graded version of the
analogue of 1.3.4 for positive connected sequences that φ0 can be incorporated
into a (uniquely determined) homomorphism
 n−i   
Φ := (φi )i∈N0 : Hm i∈N0
−→ *D(* ExtiR ( • , C)) i∈N0
of positive connected sequences of covariant functors from *C(R) to *C(R).
Furthermore, it is easy to prove by induction that, for each i ∈ N, the homo-
morphism φi M is an isomorphism whenever M is a finitely generated graded
R-module: use the fact that such an M can be included in an exact sequence
320 Graded versions of basic theorems

0 −→ K −→ F −→ M −→ 0 in *C(R) in which F is a finitely generated


*free R-module.
In order to complete the proof of the final claim, one should note that, in
view of 14.1.2(ii), the graded module * ExtjR (M, C) = 0 for all j > n for
each finitely generated graded R-module M , because (* ExtjR (M, C))m = 0
since inj dimRm Cm = n by 12.1.21.
14.5.11 Corollary. Assume that (R, m) is Cohen–Macaulay and *local with
ht m = n, and that it admits a *canonical module C. Then there is a homoge-
∼=
neous isomorphism Hm n
(C) −→ *E(R/m).
Proof. By 14.5.10, there is a homogeneous isomorphism
n
Hm (C) ∼
= * HomR (* HomR (C, C), *E(R/m)).
Since C is a canonical module for R (by 14.5.5), it follows from 14.5.9(i) that
there is a homogeneous isomorphism R ∼ = * HomR (C, C), and the desired
result follows.
In the case where (R, m) is a Cohen–Macaulay *local Z-graded ring, Bruns
and Herzog in [7, 3.6.8] gave a definition of *canonical module for R different
from ours. We are now in a position to reconcile these two approaches.
14.5.12 Corollary. Assume the G-graded ring (R, m) is *local and Cohen–
Macaulay; set ht m = n. Let C be a finitely generated graded R-module.
Then C is a *canonical module for R if and only if there are homogeneous
isomorphisms

0 for i = n,
* ExtiR (R/m, C) ∼
=
R/m for i = n.
Proof. (⇒) When C is a *canonical module for R, it follows from 14.5.10
that there are homogeneous isomorphisms
n−i
Hm (R/m) ∼
= * HomR (* ExtiR (R/m, C), *E(R/m)) for all i ∈ Z.
n−i
Since, for i = n, we have (Hm (R/m))m = 0, it follows from 14.1.2(ii) that
Hm (R/m) = 0, so that * ExtiR (R/m, C) = 0. Also, there is a homogeneous
n−i

isomorphism
* HomR (* ExtnR (R/m, C), *E(R/m)) ∼
= R/m.
Since the G-graded ring R/m is *simple, every graded R/m-module is *free
and *injective, by 14.3.2. Now (0 :*E(R/m) m) = R/m by 14.2.9, and so the
graded R/m-module * ExtnR (R/m, C) must be *free of rank 1. Thus
* ExtnR (R/m, C) ∼
= (R/m)(g)
14.5 *Canonical modules 321

in *C(R) for some g ∈ G, so that

R/m ∼
= * HomR (* ExtnR (R/m, C), *E(R/m))
= * HomR ((R/m)(g), *E(R/m)) ∼
∼ = (R/m)(−g)

in *C(R). Thus there is a homogeneous isomorphism (R/m)(−g) ∼ = R/m;


application of the shift functor ( • )(g) then shows that R/m ∼
= (R/m)(g), and
this part of the proof is complete.
(⇐) Suppose that there are homogeneous isomorphisms

i ∼ 0 for i = n,
* ExtR (R/m, C) =
R/m for i = n.

It follows from this that μi (m, C) = δi,ht m for all i ∈ N0 , so that C is a


canonical module for R by 14.5.4. Therefore, by 14.5.6, there is a Gorenstein
*local G-graded ring R and a homogeneous surjective ring homomorphism
R −→ R. It now follows from 14.5.2 that there is a *canonical module C  for
R. Since a *canonical module for R is automatically a canonical module for R
(by 14.5.5), we can now use 14.5.9(ii) to see that there exists g ∈ G for which
there is a homogeneous isomorphism C  ∼ = C(g).
It follows from the ‘(⇒)’ part of this proof that there is a homogeneous
isomorphism * ExtnR (R/m, C  ) ∼= R/m. We can now make use of 13.3.11 to
see that there are homogeneous isomorphisms

= * ExtnR (R/m, C  ) ∼
R/m ∼ = * ExtnR (R/m, C(g))
= * Extn (R/m, C)(g) ∼
R = (R/m)(g).

The homogeneous isomorphism R/m = ∼ R/m(g) leads to a homogeneous



isomorphism *E(R/m) = *E(R/m)(g), and application of the shift functor
( • )(−g) yields a homogeneous isomorphism *E(R/m)(−g) ∼ = *E(R/m).
By 14.5.10, there are homogeneous isomorphisms
n
Hm (R) ∼
= * HomR (* HomR (R, C  ), *E(R/m)) ∼
= * HomR (C  , *E(R/m)),

and use of our homogeneous isomorphisms obtained above then yield further
homogeneous isomorphisms
n
Hm (R) ∼
= * HomR (C(g), *E(R/m)) = * HomR (C, *E(R/m)(−g))

= * HomR (C, *E(R/m)).

(We have used 13.3.11 again.) Therefore C is a *canonical module for R.

14.5.13 Exercise. Assume that (R, m) is *local.


322 Graded versions of basic theorems

(i) Show that GR := {g ∈ G : R/m ∼= (R/m)(g) in *C(R)} is a subgroup


of G.
(ii) Suppose that R is Cohen–Macaulay and that C is a *canonical module
for R. Show that
GR = {g ∈ G : C(g) is a *canonical module for R}.
(iii) In the case where G = Zn for some n ∈ N and R is positively graded,
show that GR = 0.
(iv) Let H be an arbitrary subgroup of Z. Give an example of a (commutative
Noetherian) Z-graded *local ring S such that GS = H.
14.5.14 Theorem. Suppose that G = Zn , and that (R, m) is *local, Cohen–
Macaulay, and positively (Zn -)graded, and has a *canonical module. Then any
two *canonical modules for R are isomorphic in the graded category *C(R).
Proof. Let C, C  be *canonical modules for R. It follows from 14.5.5 that
C and C  are canonical modules for R in the sense of 12.1.28. Therefore, by

=
14.5.9(ii), there exist g ∈ G and a homogeneous isomorphism φ : C −→
 
C (g). Hence C (g) is a *canonical module for R, so that, since R is positively
(Zn -)graded, we must have g = 0 by 14.5.13.
14.5.15 Lemma. Assume that (R, m) is *local and Cohen–Macaulay with
ht m = n, and let C be a finitely generated graded R-module such that
μi (m, C) = δi,n for all i ∈ N0 , that is (by 12.1.25), such that Cm is canonical
for the Cohen–Macaulay local ring Rm .
Then there exists g ∈ G such that *E n (C) ∼ = *E(R/m)(−g) (in *C(R));
furthermore, for any such g, the shifted graded module C(g) is a *canonical
module for R.
Proof. It follows from 14.5.4 that C is a canonical module for R in the sense
of 12.1.28. It follows from Lemma 14.5.8(i) that there exists g ∈ G such
that *E n (C) ∼ = *E(R/m)(−g) (in *C(R)); furthermore, for any such g, parts
(ii)(a),(b) of the same lemma show that there are homogeneous isomorphisms

i ∼ 0 for i = n,
* ExtR (R/m, C) =
(R/m)(−g) for i = n.
The result therefore follows from 14.5.12 and 14.1.10(i).
14.5.16 Corollary. Suppose (R, m) is *local and Gorenstein. Then there ex-
ists g ∈ G such that R(g) is a *canonical module for R.
Proof. This is immediate from 14.5.15 because Rm is a canonical module for
Rm .
14.5 *Canonical modules 323

In the case where G = Zn and (R, m) is *local, Gorenstein and posi-


tively (Zn -)graded, it follows from 14.5.16 and 14.5.13 that there is a unique
g ∈ Zn such that R(g) is a *canonical module for R. Note that, since m =
 
(m ∩ R0 ) g∈N0 n \{0} Rg , the *simple R-module R/m is concentrated
in degree 0. By 14.5.11, there is a homogeneous isomorphism Hm n
(R(g)) =∼
n ∼
*E(R/m), so that H (R) = *E(R/m)(−g) in *C(R). This means that g
m
n
can be identified as the degree in which the *simple submodule of Hm (R) is
concentrated. This g is an important invariant that we shall discuss further (in
special cases) below; for the present we content ourselves with a calculation of
the invariant in the case of a polynomial ring over a field.

14.5.17 Example. Let K be a field and let R := K[X1 , . . . , Xn ], the ring


of polynomials over K in n indeterminates (where n ∈ N). Consider R to be
Zn -graded as in 13.1.4(ii) (so we are taking G to be Zn here).
Note that R is Gorenstein and *local, and is positively graded by Zn , with
unique *maximal ideal m := (X1 , . . . , Xn ). As was explained just above,
there is a unique g ∈ Zn such that R(g) is a *canonical module for R. We now
calculate g.
We use 14.5.11, 14.1.10(ii) and 13.5.3 to see that there are homogeneous
isomorphisms

*E(R/m) ∼ n
= Hm n
(R(g)) = (Hm (R))(g) ∼
= K[X1− , . . . , Xn− ](g).

Now the graded submodule R/m of *E(R/m) is generated by a homogeneous


element of degree 0 which has annihilator m; furthermore, the only homoge-
neous elements of K[X1− , . . . , Xn− ] which have annihilator m are the elements
αX1−1 . . . Xn−1 where α ∈ K \ {0}. It follows that g = (−1, . . . , −1).
As in 13.5.3, we can also regard K[X1 , . . . , Xn ] as Z-graded, where deg Xi
= 1 for all i = 1, . . . , n; then our polynomial ring is just the result Rφ of
applying the φ-coarsening functor of 13.5.1 to R, where φ : Zn −→ Z is the
Abelian group homomorphism for which

φ((i1 , . . . , in )) = i1 + · · · + in for all (i1 , . . . , in ) ∈ Zn .


φ ∼
Since Hmn n
φ (R ) = Hm (R)
φ
in *C Z (R) by 13.5.2, it follows that, when we
regard R as Z-graded in this way, the unique integer a for which R(a) is a
*canonical module for R is the degree in which the Z-graded *simple sub-
module of Hm n
(R)φ is concentrated, namely −n.

14.5.18 Example. Here we review graded local duality for a ring of polyno-

mials over a field K in n indeterminates X1 , . . . , Xn . Let R = g∈Zn Rg =
K[X1 , . . . , Xn ], considered to be Zn -graded as in 13.1.4, and let m denote
324 Graded versions of basic theorems

the unique *maximal ideal (X1 , . . . , Xn ) of R. We apply 14.4.1 to R: we


can take R = R and f : R −→ R to be the identity ring homomorphism;
also, R0 = K and m0 := m ∩ R0 = 0, so that E0 := ER0 (R0 /m0 ) is just
K. We saw in 14.5.17 that −1 = (−1, . . . , −1) is the unique g ∈ Zn such
that R(g) is a *canonical module for R, that is, the unique g ∈ Zn such that
Hmn
(R) ∼= *ER (R/m)(−g) in *C(R).
It therefore follows from 14.4.1 and 14.4.2 that graded local duality for this
R takes the following form: there is a homomorphism
 i  
Ψ := (ψ i )i∈N0 : Hm i∈N0
−→ * HomK (* ExtR n−i
( • , R(−1)), K) i∈N0
of (negative strongly) connected sequences of covariant functors from *C(R)
i
to itself which is such that ψM is a (homogeneous) isomorphism for all i ∈ N0
whenever M is a finitely generated graded R-module. In particular, for such i
and M , there are homogeneous isomorphisms
i
Hm (M ) ∼
= * HomK (* Extn−i
R (M, R(−1)), K)

and (since R0 = K is a complete local ring)


i
* HomK (Hm (M ), K) ∼ n−i
= * ExtR (M, R(−1)).
Similarly, when we regard R = K[X1 , . . . , Xn ] as Z-graded, where deg Xi
= 1 for all i = 1, . . . , n (so that m = (X1 , . . . , Xn ) is again the unique
*maximal ideal of R), graded local duality yields a homomorphism
 i  
Φ := (φi )i∈N0 : Hm i∈N
−→ * HomK (* ExtR n−i
( • , R(−n)), K) i∈N
0 0

Z
of (negative strongly) connected sequences of covariant functors from *C (R)
to itself which is such that φiM is a (homogeneous) isomorphism for all i ∈ N0
whenever M is a finitely generated (Z-)graded R-module.
14.5.19 Example. Let K and R = K[X1 , . . . , Xn ], considered to be Zn -
graded, be as in 14.5.17. We now describe the structure of the *indecomposable
*injective R-modules. Recall that each of these is homogeneously isomorphic
to a shift of *ER (R/p) for some graded prime ideal p of R, and that we calcu-
lated *ER (R/m) (where m = (X1 , . . . , Xn )) in 14.5.17.
It follows immediately from 14.2.6 and 14.2.7 that there are homogeneous
R-isomorphisms
*ER (R/0) ∼
= *ER((0)) (R((0)) /0) ∼
= K[X1 , . . . , Xn , X1−1 , . . . , Xn−1 ].
Now let t ∈ {1, . . . , n − 1} and p := (Xt+1 , . . . , Xn ). We now describe
*ER (R/p), and note that similar calculations (and shifts) will then provide a
complete description of all *indecomposable *injective R-modules.
14.5 *Canonical modules 325

Note that ht p = n − t. Since R is Gorenstein, μi (qR(p) , R(p) ) = δi,ht q


(Kronecker delta) for all i ∈ N0 and q ∈ * Spec(R) with q ⊆ p. Thus, if we use
n−t
the minimal *injective resolution to calculate HpR (p)
(R(p) ), the properties de-
scribed in 14.2.4(iv) yield a homogeneous R(p) -isomorphism H n−t (R(p) ) ∼ = pR(p)
*ER(p) (R(p) /pR(p) )(−g) for some g ∈ Zn . By 14.1.9, there is a homoge-
neous R(p) -isomorphism HpRn−t
(p)
(R(p) ) ∼
= (Hpn−t (R))(p) . Also, when viewed
as an R-module, *ER(p) (R(p) /pR(p) )(−g) is homogeneously isomorphic to
*ER (R/p)(−g). There is therefore a homogeneous R-isomorphism
(Hpn−t (R))(p) ∼
= *ER (R/p)(−g),
and so our strategy is to calculate the Zn -graded R-module Hpn−t (R) and then
homogenously localize it at p.
We consider R as K[X1 , . . . , Xt ][Xt+1 , . . . , Xn ] and use the calculations
in 13.5.3, with S taken to be the Zt -graded ring K[X1 , . . . , Xt ]. There results
a Zn -homogeneous R-isomorphism
Hpn−t (K[X1 , . . . , Xn ]) ∼ −
= K[X1 , . . . , Xt ][Xt+1 , . . . , Xn− ],
j
where deg(X1i1 . . . Xtit Xt+1 t+1
. . . Xnjn ) = (i, j) for all i = (i1 , . . . , it ) ∈ N0 t
and j = (jt+1 , . . . , jn ) ∈ (−N)n−t . Thus, after homogenous localization at p,
we see that a K-basis for *ER (R/p)(−g) is
j
X1i1 . . . Xtit Xt+1
t+1
. . . Xnjn ,
(i1 ,...,it )∈Zt ,(jt+1 ,...,jn )∈(−N)n−t
j
where deg(X1i1 . . . Xtit Xt+1 t+1
. . . Xnjn ) = (i, j) for all i = (i1 , . . . , it ) ∈ Zt
and j = (jt+1 , . . . , jn ) ∈ (−N)n−t , and that the Zn -graded R-module struc-
ture is such that
j
Xk (X1i1 . . . Xtit Xt+1
t+1
. . . Xnjn )


⎪ i1 ik +1
. . . Xtit Xt+1
jt+1
. . . Xnjn if 1 ≤ k ≤ t,
⎨X1 . . . Xk
j
= X1i1 . . . Xtit Xt+1 t+1
. . . Xkjk +1 . . . Xnjn if t + 1 ≤ k ≤ n, jk < −1,


⎩0 if t + 1 ≤ k ≤ n, jk = −1.
j
Furthermore, such an X1i1 . . . Xtit Xt+1
t+1
. . . Xnjn is annihilated by p if and only
if j = (−1, . . . , −1) ∈ Z , that is, if and only if it has the form
n−t

−1
X1i1 . . . Xtit Xt+1 . . . Xn−1 .
Note that multiplication by X1 provides a homogeneous automorphism of de-
gree (1, 0, . . . , 0) of *ER (R/p); similar comments apply to X2 , . . . , Xt . It
follows that there is a Zn -homogeneous isomorphism
*ER (R/p) ∼ −
= K[X1 , . . . , Xt ][Xt+1 , . . . , Xn− ]((k, (−1, . . . , −1)))
326 Graded versions of basic theorems

for any k ∈ Zt . A more combinatorial approach to this *indecomposable


*injective module is provided by Miller and Sturmfels in [53, Chapter 11].
We can use 14.2.8 and 14.5.19 to find the *indecomposable *injective mod-
ules over a homomorphic image of the polynomial ring of 14.5.19. In partic-
ular, this applies to the Stanley–Reisner rings (with respect to the field K) of
simplicial complexes on {1, . . . , n} that were introduced in 13.1.5.
14.5.20 Exercise. Let Δ be the simplicial complex on {1, 2, 3, 4} consisting
of all the subsets of {1, 2, 3} and all the subsets of {1, 4}. Let K be a field and
work in the polynomial ring K[X1 , X2 , X3 , X4 ].
(i) Show that aΔ = (X2 X4 , X3 X4 ).
(ii) Identify *EK[Δ] (K[Δ]/ ((X2 , X3 , X4 )/aΔ )) with a shift of a submod-
ule of K[X1 ][X2− , X3− , X4− ] via 14.2.8 and 14.5.19. Show that
B := {X1i1 X2i2 X3i3 X4i4 : i1 ∈ Z, i2 , i3 , i4 ∈ −N
and i2 = i3 = −1 or i4 = −1}
is a K-basis for *EK[Δ] (K[Δ]/ ((X2 , X3 , X4 )/aΔ )). What can you say
about the degrees of the elements of B?

14.5.21 Definition and Remarks. Suppose G = Z and (R, m) is Cohen–


Macaulay and *local, and is a positively (Z-)graded ring; assume also that R
has a *canonical module. By 14.5.14, this is uniquely determined up to homo-
geneous isomorphism: we denote by ωR one choice of *canonical module for
R. The a-invariant of R is defined to be
a(R) := − beg(ωR ) = − min{n ∈ Z : (ωR )n = 0}.
(See 14.1.1 for the definition of the beginning of a Z-graded R-module.)
(i) Note that, when R is Gorenstein, a(R) is the unique integer a for which
R(a) is a *canonical module for R. Thus the a-invariant of a polyno-
mial ring over a field, Z-graded so that each variable has degree 1, was
calculated in 14.5.17.
(ii) In the general case, we can use graded local duality, as described in
14.5.10, to see that (with the notation *D of that result) there are ho-
mogeneous isomorphisms
n
Hm (R) ∼
= *D(* HomR (R, ωR )) ∼
= *D(ωR ),
and so it follows from Graded Matlis Duality 14.4.2 that
a(R) = − beg(ωR ) = end(*D(ωR )) = end(Hm
n
(R)).
14.5 *Canonical modules 327

The following proposition gives some hints about ways in which graded
local duality and graded Matlis duality can be used in tandem.

14.5.22 Proposition. Suppose that (R, m) is Cohen–Macaulay and *local,


and has a *canonical module. Set n := ht m.
Then R is Gorenstein if and only if there is a homogeneous isomorphism
n (R) m) ∼
(0 :Hm = (R/m)(g) for some g ∈ G.

Proof. (⇒) There exists g ∈ G such that R(g) is a *canonical module for
R, by 14.5.16. By 14.5.5 and 14.5.8, there exist h ∈ G and a homogeneous
isomorphism Hm n
(R(g)) ∼
= *E(R/m)(−h). It follows that there are homoge-
neous isomorphisms Hm (R) ∼
n
= *E(R/m)(−h − g) and

n (R) m) ∼
(0 :Hm = (0 :*E(R/m)(−h−g) m) = (R/m)(−h − g).

(We have used 14.2.9 here.)


(⇐) Let C be a *canonical module for R, and let *D denote the functor
* HomR ( • , *E(R/m)) from *C(R) to itself. By definition, there is a homo-
geneous isomorphism *D(C) ∼ = Hmn
(R). There are therefore homogeneous
isomorphisms

n (R) m) ∼
n
(0 :Hm = * HomR (R/m, Hm (R))

= * HomR (R/m, * HomR (C, *E(R/m)))

= * HomR (R/m ⊗R C, *E(R/m))

= * HomR (C/mC, *E(R/m)).

Application of *D and use of 14.2.9 therefore yield further homogeneous iso-


morphisms

*D(*D(C/mC)) ∼ n (R) m)) ∼


= *D((0 :Hm = *D((R/m)(g)) ∼
= (R/m)(−g).

However, the canonical R-homomorphism C/mC −→ *D(*D(C/mC)) is


homogeneous and monomorphic. Since Cm /mRm Cm = 0 (because Cm is
canonical for Rm by 14.5.3), it follows that Cm is a cyclic Rm -module, so that
Rm is Gorenstein. Therefore Rp is Gorenstein for all p ∈ * Spec(R), so that
R is Gorenstein by 14.3.9(ii).

In §12.3, we discussed the concept of S2 -ification, and we showed that an


S2 -ification exists in a local ring R that has a faithful canonical module. We
are now going to guide the reader to some results about S2 -ifications in graded
situations.

14.5.23 Notation. Suppose that (R, m) is G-graded and *local, and that C
328 Graded versions of basic theorems

is a *canonical module for R. Recall that there is a minimal primary decom-


n
position 0 = i=1 qi for the zero ideal in which each term is graded. Sup-
pose that the qi are indexed so that ht m/qi = ht m for all i = 1, . . . , t
and ht m/qi < ht m for all i = t + 1, . . . , n. (Of course, t could be n.) Set
t
uR (0) = i=1 qi . Note that uR (0) is graded, so that uR (0) = 0 if and only if
uR (0)Rm = 0, and that uR (0)Rm = uRm (0) in the notation of 12.1.12.
Note also that, by 14.5.3, the localization Cm is a canonical module for
Rm . We have (0 :Rm Cm ) = uRm (0), by 12.1.15. Therefore the graded ideal
(0 :R C) is equal to uR (0).
As in 12.3.8, we shall denote by S the system of all ideals s of R such that
Var(s) is contained in the non-S2 locus of R. We shall guide the reader, in
Exercise 14.5.24 below, to the result that, when the *canonical R-module C is
S
faithful, there is an S2 -ification ηR : R −→ DS (R) in the sense of 12.3.9, and
this R-algebra is G-graded with homogeneous structural homomorphism.

14.5.24 Exercise. Let the situation and notation be as in 14.5.23, and denote
by *H the system of all graded ideals of R of height at least 2. Assume that
uR (0) = 0, so that uRm (0) = 0. Let hR : R −→ * HomR (C, C) =: H denote
the natural homogeneous R-homomorphism for which hR (r) = r IdC for all
r ∈ R.

(i) Show that there is an isomorphism of Rm -algebras



=
Hm −→ HomRm (Cm , Cm ),

and recall that HomRm (Cm , Cm ) is the endomorphism ring of the canon-
ical Rm -module Cm .
(ii) Show that the natural R-homomorphism H −→ Hm is injective. (Here
is a hint: note that an associated prime of the kernel of the specified
homomorphism would have to be an associated prime of H.)
(iii) Show that Ker hR = 0; use 12.2.7 to show that Coker hR is *H-torsion.
(iv) Use the Goto–Watanabe results 14.3.6 and 14.3.8, part (i) and 12.1.18(i)
to show that the R-module H is S2 (see 12.1.16).
(v) Show that Γ*H (H) = 0 and use part (ii) together with 12.3.10(ii) to
show that H is a commutative G-graded Noetherian ring.
(vi) Use 13.5.4(ii), 2.2.15, 2.2.17 and 12.3.2 to show that there is a unique
homogeneous R-algebra isomorphism

ψ  : H = * HomR (C, C) −→ D*H (R).
=

(vii) Use part (vi) to show that H*1H (R) is a finitely generated R-module; let
a be its annihilator, and note that a is graded. Show that
14.5 *Canonical modules 329

(a) ht a ≥ 2;
(b) there is a uniquely determined homogeneous R-algebra isomor-

=
phism Da (R) −→ D*H (R); and
(c) Var(a) is equal to the non-S2 locus of R, so that every S-torsion
R-module is a-torsion.
S
(viii) Deduce that R has an S2 -ification ηR : R −→ DS (R), that DS (R) is
S
a G-graded commutative Noetherian ring, that ηR is homogeneous, and
that there are unique homogeneous R-algebra isomorphisms

= ∼
= ∼
=
* HomR (C, C) −→ Da (R) −→ D*H (R) −→ DS (R).

We shall consider some examples of S2 -ifications in graded situations in


§15.2.
The result established in the following exercise is due to J. Herzog and E.
Kunz [34].

14.5.25 Exercise. Let a1 , . . . , ah ∈ N \ {1} satisfy GCD(a1 , . . . , ah ) = 1.


Let S := a1 N0 + · · · + ah N0 , the additive subsemigroup of N0 generated
by a1 , . . . , ah . Let K be a field and let R be the subring of the polynomial
ring K[X] given by R := K[X a1 , . . . , X ah ]. Of course, R is a 1-dimensional
Cohen–Macaulay ring; furthermore, it is positively Z-graded, by virtue of the
grading inherited from the usual Z-grading on K[X] in which deg X = 1, and,
with this grading, R is *local. We denote by R+ the unique graded maximal
ideal of R.

(i) Show that there exists c ∈ N such that n ∈ S for all n ∈ N with
n ≥ c. Thus L := N0 \ S is a non-empty finite set, which we refer to as
the set of non-members of S, and so has a greatest member, e say. The
semigroup S is said to be symmetric precisely when, for all integers n
with 0 ≤ n ≤ e, we have n ∈ S if and only if e − n ∈ S.
(ii) Show that R has a *canonical module.
(iii) Show that K[X]X = K[X, X −1 ] can be naturally identified with the
ideal transform DR+ (R) in the sense that there is a unique homogeneous
R-isomorphism φ : K[X, X −1 ] −→ DR+ (R) such that the diagram

R
φ
- K[X, X −1 ]
@
η
@ ∼
= φ
R
@
R
@ ?
DR+ (R) ,
330 Graded versions of basic theorems

in which φ is the inclusion map, commutes.


(iv) Use 13.5.4(i) to show that, for n ∈ Z,

1 1 if n < 0 or n ∈ L,
dimK (HR+ (R)n ) =
0 if n ∈ S,
1
and deduce that end(HR +
(R)) = e.
(v) Show that, if R is Gorenstein, then S is symmetric.
(vi) Now suppose that S is symmetric. Show that, if n ∈ L ∪ {−i : i ∈ N},
then s := e−n ∈ S and X s HR 1
+
(R)n = 0. Deduce that R is Gorenstein.
14.5.26 Exercise. Let K be a field and let R := K[X1 , . . . , Xn ], the ring of
polynomials over K in n indeterminates (where n ∈ N), graded by Z so that
R0 = K and deg Xi = 1 for all i = 1, . . . , n. Let r ∈ N, and consider the r-th
Veronesean subring R(r) of R, as in 13.5.9. In Exercise 13.5.13, the reader was
asked to show that R(r) is Cohen–Macaulay. Prove that R(r) is Gorenstein if
and only if n ≡ 0 (mod r).
14.5.27 Exercise. Assume that G = Z and that (R, m) is Cohen–Macaulay
and *local, and positively (Z-)graded; assume also that R has a *canonical
module. Let b ∈ R be a homogeneous element of positive degree d which is
a non-zerodivisor on R. Show that R/bR has a *canonical module and that
a(R/bR) = a(R) + d.
15

Links with projective varieties

One of the reasons for the interest in graded local cohomology is provided by
the numerous applications to projective algebraic geometry. This short chapter
is intended to provide a little geometric insight, with the aim of motivating the
work on Castelnuovo regularity in Chapter 16.
In 2.3.2, we saw that the ideal transform has a geometric meaning in certain
cases: if V is an affine variety over K, an algebraically closed field, b is a
non-zero ideal of O(V ), and U denotes the open subset of V determined by
b, then the ideal transform Db (O(V )) is isomorphic, as an O(V )-algebra, to
the ring of regular functions on U . One of our first aims for this chapter is the
establishment of a graded analogue of this result. This graded analogue applies
to irreducible affine algebraic cones.
Throughout this chapter, all graded rings and modules are to be understood
to be Z-graded.

15.1 Affine algebraic cones


15.1.1 Notation and Terminology. We shall employ the notation and termi-
nology concerning graded rings and modules described in 13.1.1, 13.1.3 and
14.1.1, but, in accordance with our convention for the whole of this chapter,
restricted to the special case in which G = Z. In addition, when R is graded
and the ideal a is graded, and M is a graded R-module, so that (see 13.4.3
and 13.4.4) the Hai (M ) (i ∈ N0 ) are all graded R-modules, we use Hai (M )n
to denote the n-th component of the graded module Hai (M ) (for i ∈ N0 and
n ∈ Z); also Da (M ) is graded (by 13.3.14), and we use Da (M )n to denote its
n-th component.
332 Links with projective varieties

Now assume that R = n∈N0 Rn is positively graded. We set
 
R+ := Rn = R n = 0 ⊕ R1 ⊕ R 2 ⊕ . . . ⊕ R n ⊕ . . . ,
n∈N n>0

the irrelevant ideal of R. Of course, R+ is graded, and so, for a graded R-


module M , we can define the R0 -modules HR i
+
(M )n (i ∈ N0 , n ∈ Z).

15.1.2 Geometric Notation and Reminders. Let K be an algebraically clo-


sed field, and let r ∈ N. We shall use the notation of 2.3.1, and we shall denote
the origin (0, . . . , 0) of Ar (K) simply by 0. For a subset C of a quasi-affine
variety W over K, we shall extend the notation of 6.4.1 to denote the vanishing
ideal
{f ∈ O(W ) : f (q) = 0 for all q ∈ C}

of C by IW (C).

(i) By a cone (with vertex 0) in Ar (K) we mean a set C ⊆ Ar (K) such that
0 ∈ C and, whenever q ∈ C, then λq ∈ C for all λ ∈ K. Such a cone
C is called an affine algebraic cone in Ar (K) if and only if it is also an
affine algebraic set; also C is said to be non-degenerate precisely when
C = {0}.
Since a graded ideal can be generated by homogeneous elements, it is
clear that, if b is a proper graded ideal of K[X1 , . . . , Xr ], then VAr (K) (b)
is an affine algebraic cone in Ar (K). It is easy to see that, conversely,
if C is an affine algebraic cone in Ar (K), then IAr (K) (C) is a proper
graded ideal. Thus the affine algebraic cones in Ar (K) are precisely the
algebraic sets in Ar (K) which have proper graded vanishing ideals.
(ii) Let C be an irreducible affine algebraic cone in Ar (K). Since IAr (K) (C)
is a graded prime ideal of K[X1 , . . . , Xr ], the ring O(C) of regular func-
tions on C inherits a grading from K[X1 , . . . , Xr ], in such a way that the
restriction homomorphism K[X1 , . . . , Xr ] −→ O(C) is homogeneous
(see 14.1.3). Of course, for each n ∈ N0 , we denote the n-th component
of O(C) by O(C)n . Note that O(C)0 can be identified with K. It is also
worth noting that O(C) is *local with unique *maximal graded ideal
O(C)+ ; furthermore, O(C)+ is actually a maximal ideal in this case.
(iii) With the notation of part (ii), let b be a non-zero graded ideal of O(C),
let VC (b) denote the closed subset of C determined by b, and let U
be the open subset C \ VC (b) of C. In fact, if c denotes the inverse
image of b in K[X1 , . . . , Xr ] under the restriction homomorphism, then
VC (b) = VAr (K) (c), and U is obtained from C by removal of another
affine algebraic cone (or the empty set).
15.1 Affine algebraic cones 333

Let n ∈ Z. A regular function f ∈ O(U ) is said to be homogeneous


of degree n precisely when, for each p ∈ U , there exists an open set
W ⊆ U with p ∈ W , an integer d ∈ N0 , and g ∈ O(C)d , h ∈ O(C)n+d
such that, for each q ∈ W , we have g(q) = 0 and f (q) = h(q)/g(q).
The set of all regular functions on U which are homogeneous of degree
n is denoted by O(U )n . The fact that this definition is not ambiguous in
the case when U = C is one consequence of the next proposition.
15.1.3 Proposition. Let the situation be as in 15.1.2(iii). Thus K is an al-
gebraically closed field, r ∈ N, C is an irreducible affine algebraic cone in
Ar (K), b is a non-zero graded ideal of O(C), and U = C \ VC (b).
The subsets O(U )n (n ∈ Z) defined in 15.1.2(iii) provide O(U ) with a
structure as a graded ring with respect to which the homomorphisms U and
νC,b in the commutative diagram

O(C)
U
- O(U )

@
@
ηO(C)

= νC,b
@
@R ?
Db (O(C))

of 2.3.2 are homogeneous. (It should be noted that, since b is graded, it fol-
lows from 13.3.14 that Db (O(C)) is a graded O(C)-module and that ηO(C)
is homogeneous.)
Proof. Let us abbreviate νC,b by ν and ηO(C) by η. Since ν is a ring isomor-
phism (by 2.3.2), it is enough for us to show that O(U )n = ν −1 (Db (O(C))n )
for each n ∈ Z, and this is what we shall do.
Let y ∈ Db (O(C))n . Since Coker η, being isomorphic to Hb1 (O(C)), is
b-torsion, there exists t ∈ N such that bt y ∈ Im η. Let p ∈ U . Since U =
C \ VC (b) and b is graded, there exists a homogeneous element g ∈ bt , of
degree d ∈ N0 say, such that g(p) = 0. Then W := U \ VC (gO(C)) is an
open subset of U which contains p.
As gy ∈ Im η, there is an element h ∈ O(C) with gy = η(h). As gy ∈
Db (O(C))n+d and as η is homogeneous and injective, we must have h ∈
O(C)n+d . Now
−1
g Uν (y) = ν −1 (η(g))ν −1 (y) = ν −1 (η(g)y) = ν −1 (gy)
= ν −1 (η(h)) = h U.

Therefore, for each q ∈ W , we have g(q) = 0 and ν −1 (y)(q) = h(q)/g(q).


Hence ν −1 (y) ∈ O(U )n . We have proved that O(U )n ⊇ ν −1 (Db (O(C))n ).
334 Links with projective varieties

Now let f ∈ O(U )n . Choose p ∈ U ; then there exists an open set W ⊆ U


with p ∈ W , an integer d ∈ N0 , and g ∈ O(C)d , h ∈ O(C)n+d such that,
for each q ∈ W , we have g(q) = 0 and f (q) = h(q)/g(q). This shows, in
particular, that g W .f W = h W . As W is a non-empty open subset of the
irreducible topological space U , it follows that g U .f = h U . Now apply ν:
we obtain
η(g)ν(f ) = ν(g U )ν(f ) = ν(g U .f ) = ν(h U) = η(h).
As η is homogeneous, η(g) ∈ Db (O(C))d and η(h) ∈ Db (O(C))n+d . As η
is injective and g = 0, we have η(g) = 0; also, as Db (O(C)) is a domain, it
follows that ν(f ) must be homogeneous of degree n. We have therefore shown
that O(U )n ⊆ ν −1 (Db (O(C))n ), and so the proof is complete.
Theorem 2.3.2, and its graded refinement 15.1.3, provide a link between lo-
cal cohomology and algebraic varieties. Towards the end of the book, we shall
encounter more general versions of these two results, because they are related
to the Deligne Isomorphism Theorem 20.1.14 (see 20.1.17) and its graded ver-
sion 20.2.7.
We explore now the special case of Proposition 15.1.3 in which dim C > 0
and the graded ideal b is the irrelevant ideal O(C)+ (see 15.1.1) of O(C): then

the open set U = C \ {0} is just the punctured cone C of the given irreducible
affine algebraic cone C.
15.1.4 Corollary. Consider the special case of 15.1.3 where the irreducible
affine algebraic cone C has dim C > 1 and b = O(C)+ . Then

U = C \ VC (O(C)+ ) = C \ {0} = C ,
the punctured cone of C; also
◦ ◦
(i) the restriction ring homomorphism C◦ : O(C) −→ O(C ) makes O(C )
into a finitely generated graded O(C)-module;
1
(ii) end(HO(C) +
(O(C))) < ∞ (see 14.1.1 for the definition of the end of a
graded module);
1
(iii) HO(C) +
(O(C))n is a finite-dimensional vector space over K, for all
n ∈ Z; and
1
(iv) HO(C) +
(O(C))n = 0 for all n ≤ 0.
Proof. Set R := O(C). Now (R, R+ ) is *local, by 15.1.2(ii), and a domain.
Hence, by 14.1.14, we have ht R+ = dim R = dim C > 1. Therefore, for all
p ∈ * Spec(R) \ Var(R+ ), we have p ⊂ R+ and
depth Rp + ht R+ /p ≥ 2 > 1.
15.1 Affine algebraic cones 335
1
It thus follows from the Graded Finiteness Theorem 14.3.10 that HR +
(R) is
finitely generated. By 2.2.6(i)(c), 13.5.4 and 15.1.3, there is an exact sequence
◦ ◦
R −→
C
O(C ) −→ HR
1
+
(R) −→ 0 of graded R-modules and homogeneous

R-homomorphisms. It is now immediate that C◦ makes O(C ) into a finitely
generated graded R-module.
1
Note also that, since HR +
(R) is finitely generated and R+ -torsion, we must
have HR+ (R)n = 0 for all n  0 (that is, for all n greater than some fixed
1

integer n0 ), and so end(HR 1


+
(R)) < ∞. We have therefore completed the
proofs of parts (i) and (ii), while part (iii) is now immediate from the facts that
R0 = K and, for each n ∈ Z, the component HR 1
+
(R)n is a finitely generated
R0 -module.
It remains for us to prove part (iv). Use C◦ to identify R as a subring of
◦ ◦
O(C ). It is enough for us to show that K = R0 = O(C)0 = O(C )0 and that

O(C )n = 0 for all n < 0, and this we do.

First, by part (i), the integral domain O(C )0 is an integral extension of the
◦ ◦
algebraically closed field K = R0 , and so R0 = O(C )0 . Second, since O(C )

is a finitely generated graded R-module and R is positively graded, O(C )n =

0 for all n  0. Hence, for each n < 0 and each γ ∈ O(C )n , there exists

t ∈ N such that γ t = 0, so that γ = 0 since O(C ) is a domain. This completes
the proof.

15.1.5 Exercise. Calculate O(C), O(C ) and HO(C) 1
+
(O(C)) for an irre-
ducible affine algebraic cone C of dimension 1 in A (K), where K is an al-
r

gebraically closed field and r ∈ N. (Here is a hint. Let c := (c1 , . . . , cr ) ∈


C \ {0}. Without loss of generality, one can assume that c1 = 0: with this
assumption, show that

IAr (K) (C) = (c2 X1 − c1 X2 , c3 X1 − c1 X3 , . . . , cr X1 − c1 Xr ).)



15.1.6 Exercise. Assume that R = n∈N0 Rn is positively graded and an
integral domain, and that the ideal a is graded and non-zero.

(i) Show that, if a ∩ R0 = 0, then Da (R)n = Ha1 (R)n = 0 for all n < 0.
(ii) Assume that the subring R0 is a homomorphic image of a regular ring,
and that ht a > 1. Show that ηa makes Da (R) into a finitely generated
R-module, and deduce that Da (R)n and Ha1 (R)n are finitely generated
R0 -modules for all n ∈ Z. Show further that Da (R)n = Ha1 (R)n = 0
for all n < 0.
(iii) Assume that R0 is an algebraically closed field K, and that ht a > 1.
336 Links with projective varieties

Using ηa to identify R as a subring of Da (R), show that Da (R)0 = K


and Ha1 (R)0 = 0.

15.1.7 Exercise. Let the situation and notation be as in 15.1.3, and assume
that ht b > 1. Use 15.1.6(ii),(iii) to show that O(U ) is a finitely generated
O(C)-module for which O(U )0 = K.

15.2 Projective varieties


In the situation of 15.1.3, we have seen (see 2.3.3) that we can regard non-zero
elements of the local cohomology module Hb1 (O(C)) as obstructions to the
extension of regular functions on U to regular functions on C; we have also

obtained, in 15.1.4, in the case when U is the punctured cone C , information
1
about some of the components HO(C) +
(O(C))n = Hb1 (O(C))n . In the light
of the connections between irreducible affine algebraic cones and projective
varieties (reviewed in 15.2.1 below), it would be reasonable for one to suspect
that there are links between graded local cohomology and projective varieties.
Such suspicions would be well founded, and we plan to expose some of the
links in this and subsequent chapters.

15.2.1 Reminders and Notation. Here we specify the notation and termi-
nology that we shall use for discussion of projective varieties. Let K be an
algebraically closed field, and let r ∈ N. We shall find it convenient to vary
slightly the notation of 2.3.1 and regard the polynomial ring K[X0 , . . . , Xr ] in
r + 1 indeterminates X0 , X1 , . . . , Xr as the coordinate ring O(Ar+1 (K)) of
affine (r + 1)-space Ar+1 (K) over K. As in 15.1.2, we shall denote the origin
(0, . . . , 0) of Ar+1 (K) simply by 0.
(i) For c := (c0 , . . . , cr ) ∈ Ar+1 (K) \ {0}, we use (c0 : · · · : cr ) to
denote the line {λc : λ ∈ K} in Ar+1 (K) through c and 0. We shall use
Pr (K) to denote projective r-space over K, that is, the set

(c0 : · · · : cr ) : (c0 , . . . , cr ) ∈ Ar+1 (K) \ {0}

of all lines through the origin in Ar+1 (K), endowed with the Zariski topology.
We remind the reader that the closed sets in this topology are precisely the
projective algebraic sets, that is, the sets of the form

VPr (K) (a) := {(c0 : · · · : cr ) ∈ Pr (K) : f (c0 , . . . , cr ) = 0 for all f ∈ a} ,

where a is a proper graded ideal of K[X0 , . . . , Xr ]. (When a = (X0 , . . . , Xr ),


the corresponding projective algebraic set is the empty one.) If f1 , . . . , ft are
15.2 Projective varieties 337

non-constant homogeneous polynomials in K[X0 , . . . , Xr ], then the projec-


tive algebraic set VPr (K) ((f1 , . . . , ft )) is denoted by VPr (K) (f1 , . . . , ft ). Of
course, every projective algebraic set in Pr (K) can be represented in this form.
We shall use Pr to denote complex projective r-space Pr (C). All unex-
plained mentions of topological notions, including ‘open’ and ‘closed’ subsets,
in connection with projective spaces will refer to the Zariski topology.
(ii) By the statement ‘V ⊆ Pr (K) is a projective variety’ we shall mean
that V is an irreducible closed subset of Pr (K) (with the induced topology),
and by the statement ‘W ⊆ Pr (K) is a quasi-projective variety’ we shall mean
that W is a non-empty open subset of a projective variety V ⊆ Pr (K) (again
with the induced topology). (It is wise for us to include the ‘⊆ Pr (K)’ in the
notation because (and some of our examples below will remind the reader of
this) isomorphic projective varieties, embedded in projective spaces in differ-
ent ways, can have non-isomorphic homogeneous coordinate rings! (See parts
(vi) and (viii) below.) This problem does not occur with affine varieties.)
By a variety we shall mean an affine, quasi-affine, projective, or quasi-
projective variety.
(iii) Let V be the closed subset of Pr (K) given by V = VPr (K) (a), where
a is a proper graded ideal of K[X0 , . . . , Xr ]. The affine cone Cone(V ) ⊆
Ar+1 (K) over V in Ar+1 (K) is defined by

Cone(V ) = (c0 , . . . , cr ) ∈ Ar+1 (K) \ {0} : (c0 : · · · : cr ) ∈ V ∪ {0}


= VAr+1 (K) (a),

an affine algebraic cone in Ar+1 (K).


On the other hand, for the affine algebraic cone C in Ar+1 (K) defined by
C = VAr+1 (K) (a ), where a is a proper graded ideal of K[X0 , . . . , Xr ], we
define the projectivization C + of C to be the closed subset

C + = {(c0 : · · · : cr ) ∈ Pr (K) : (c0 , . . . , cr ) ∈ C \ {0}} = VPr (K) (a )

of Pr (K). Thus V → Cone(V ) provides a bijective map from the set T :=


{V ⊆ Pr (K) : V is closed} of all closed subsets of Pr (K) to the set

C ⊆ Ar+1 (K) : C is an affine algebraic cone

of all affine algebraic cones in Ar+1 (K); the inverse map is given by C → C + .
Note that these two maps both preserve inclusion relations.
(iv) It follows from part (iii) and standard facts from affine algebraic geom-
etry that there is a bijective correspondence between the set T of all closed sub-
sets of Pr (K) and the set of all radical proper graded ideals of K[X0 , . . . , Xr ]
338 Links with projective varieties

under which a closed subset V ∈ T corresponds to


IAr+1 (K) (Cone(V )) = {f ∈ K[X0 , . . . , Xr ] : f (c0 , . . . , cr ) = 0 for all
(c0 , . . . , cr ) ∈ Cone(V )}
= {f ∈ K[X0 , . . . , Xr ]+ : f (c0 , . . . , cr ) = 0 for all
(c0 , . . . , cr ) ∈ Ar+1 (K) \ {0}
with (c0 : · · · : cr ) ∈ V } .
This is also denoted by IPr (K) (V ), and is called the vanishing ideal of V .
(v) Note that, if V1 , . . . , Vt are closed subsets in Pr (K), then
 t   t 
 
t * *t
Cone Vi = Cone(Vi ) and Cone Vi = Cone(Vi ).
i=1 i=1 i=1 i=1

Also, the minimal prime ideals of a proper graded ideal of K[X0 , . . . , Xr ] are
again graded (by 13.1.6(ii)), and so it follows that a closed subset V of Pr (K)
is irreducible if and only if Cone(V ) is irreducible and non-degenerate, and
that this is the case if and only if IAr+1 (K) (Cone(V )) is a graded prime ideal
of K[X0 , . . . , Xr ] properly contained in (X0 , . . . , Xr ).
Thus, in the bijective correspondence of part (iv), the projective varieties
in Pr (K) correspond to the graded prime ideals of K[X0 , . . . , Xr ] properly
contained in (X0 , . . . , Xr ).
(vi) Let V ⊆ Pr (K) be a projective variety. By parts (iv) and (v), the
ideal p := IAr+1 (K) (Cone(V )) is a graded prime ideal of K[X0 , . . . , Xr ] with
p ⊂ (X0 , . . . , Xr ). We refer to the positively graded *local ring
O(Cone(V )) = K[X0 , . . . , Xr ]/p = K[X0 , . . . , Xr ]/IAr+1 (K) (Cone(V ))
as the homogeneous coordinate ring of V ⊆ Pr (K).
(vii) Remember that the dimension dim V of a variety V is defined as the
maximum length l of a strictly descending chain C0 ⊃ C1 ⊃ · · · ⊃ Cl of
closed irreducible subsets of V , and that such a V is called a curve if and only if
dim V = 1 and a surface if and only if dim V = 2. Using the correspondence
described in part (iii) and the observation made in part (v), we obtain, for a
projective variety V  ⊆ Pr (K), that
dim V  = dim(Cone(V  )) − 1 = dim(O(Cone(V  ))) − 1.
(viii) We mentioned in (ii) above that isomorphic projective varieties, em-
bedded in projective spaces in different ways, can have non-isomorphic ho-
mogeneous coordinate rings. Properties of the homogeneous coordinate ring
O(Cone(V )) of a projective variety V ⊆ Pr (K) are referred to as arith-
metic properties of V . For example, we say that V ⊆ Pr (K) is arithmetically
15.2 Projective varieties 339

Cohen–Macaulay (respectively arithmetically Gorenstein, . . .) if and only if


O(Cone(V )) is a Cohen–Macaulay (respectively Gorenstein, . . .) ring. We
also define the arithmetic depth arithdepth V of V to be the grade of the
unique *maximal graded ideal of the homogeneous coordinate ring of V . Thus
arithdepth V = grade O(Cone(V ))+ .
15.2.2 Reminder and Exercise: Veronesean tranformations. Let K be
an algebraically closed field, and let r, d ∈ N. Let V ⊆ Pr (K) be a projective
variety, and let p denote (the graded prime) kernel IPr (K) (V ) of the restriction
homomorphism K[X0 , . . . , Xr ] −→ O(Cone(V )). Set
T = Tr(d) = (ν0 , . . . , νr ) ∈ N0 r+1 : ν0 + · · · + νr = d ,
 
a set with cardinality r+d
r . We shall use a family of (algebraically indepen-
dent) indeterminates (Yν )ν∈T indexed by T . Let
φ : K[Yν : ν ∈ T ] −→ K[X0 , . . . , Xr ]
be the K-algebra homomorphism for which φ(Y(ν0 ,...,νr ) ) = X0ν0 . . . Xrνr for
all (ν0 , . . . , νr ) ∈ T . Let θ be the composition
φ
K[Yν : ν ∈ T ] - K[X0 , . . . , Xr ] - O(Cone(V )),
Cone(V )

and let q := Ker θ, a graded prime ideal of K[Yν : ν ∈ T ].


(i) Show that, with the notation of 13.5.9, Im φ = K[X0 , . . . , Xr ](d) , the
d-th Veronesean subring of K[X0 , . . . , Xr ], that φ(q) = p(d) , and that
Im θ = O(Cone(V ))(d) .
Note that, when Veronesean subrings are given the grading described at the
beginning of 13.5.9, both φ : K[Yν : ν ∈ T ] −→ K[X0 , . . . , Xr ](d) and
θ : K[Yν : ν ∈ T ] −→ O(Cone(V ))(d)
are homogeneous ring homomorphisms.
r+d
If we consider K[Yν : ν ∈ T ] as O A( r ) (K) , then the graded prime
ideal q of this polynomial ring defines a projective variety
r+d
V (d) := V (r+d)−1 (q) ⊆ P( r )−1 (K),
P r (K)

called the d-th Veronesean of V . Note that there is a homogeneous isomor-



=
phism of graded K-algebras O(Cone(V (d) )) −→ O(Cone(V ))(d) .
The main aim of this exercise is to show that the varieties V and V (d)
are isomorphic. Some additional notation will be helpful. Given an element
r+d
β = (βν )ν∈T ∈ A( r ) (K) \ {0}, we use β = (βν )ν∈T to denote the line
340 Links with projective varieties
r+d
{λβ : λ ∈ K}, considered as an element of P( r )−1 (K). Also, if y =
(y0 , . . . , yr ) is an (r + 1)-tuple of elements of some commutative K-algebra
and ν = (ν0 , . . . , νr ) ∈ N0 r+1 , then we shall use y ν to denote y0ν0 . . . yrνr .
For each i = 0, . . . , r, let ei = (0, . . . , 0, 1, 0, . . . , 0) be the element of N0 r+1
whose only non-zero component is a ‘1’ in the i-th position. In order to sim-
plify notation, we shall also employ the Z-module structure on Zr+1 . For ex-
ample, with this notation, we can write that dei ∈ T for all i = 0, . . . , r.

(ii) Let s ∈ N and let ω (1) , . . . , ω (s) , μ(1) , . . . , μ(s) ∈ T be such that
s (i)
s
i=1 ω = i=1 μ(i) .

Show that Yω(1) . . . Yω(s) − Yμ(1) . . . Yμ(s) ∈ q.


(iii) Let ν = (ν0 , . . . , νr ) ∈ T . Deduce from part (ii) that, for all i, j, k ∈
{0, . . . , r}, we have
(a) Yek +(d−1)ei Yei +(d−1)ej − Ydei Yek +(d−1)ej ∈ q;
r
(b) α=0 (Yeα +(d−1)ei )να − Yν (Ydei )d−1 ∈ q; and
r
(c) Yνd − α=0 (Ydeα )να ∈ q.
(iv) Show that there is a morphism of varieties ϑ(d) : V −→ V (d) which is
such that

ϑ(d) ((α0 : · · · : αr )) = (αν )ν∈T for all α = (α0 : · · · : αr ) ∈ V.

(v) For i = 0, . . . , r, let Ui := V (d) \ V (r+d)−1 (Ydei ). Use part (iii)(c)


P r (K)
to show that U0 , . . . , Ur form an open covering of V (d) .
(vi) Show that there is a morphism of varieties ω : V (d) −→ Pr (K) which
is such that, for each i = 0, . . . , r and for all (βν )ν∈T ∈ Ui ,

ω((βν )ν∈T ) = (βe0 +(d−1)ei : · · · : βej +(d−1)ei : · · · : βer +(d−1)ei ).

(vii) Show that Im ω ⊆ V . (Here are some hints. Suppose i ∈ {0, . . . , r} and
(βν )ν∈T ∈ Ui . Let α := (βe0 +(d−1)ei , . . . , βer +(d−1)ei ). It is enough
to show that, for each homogeneous g ∈ p, we have g d (α) = 0 (as this
would imply that g(α) = 0). Now g d ∈ p(d) ; by part (i), we have p(d) =
φ(q), and so there exists a homogeneous f ∈ q such that φ(f ) = g d .
Now use part (iii)(b).)
(viii) Show that ω : V (d) −→ V and ϑ(d) : V −→ V (d) are inverse isomor-
phisms of varieties. The isomorphism ϑ(d) is called the d-th Veronesean
transformation of V or the d-th Veronesean map on V .

15.2.3 Example: rational normal curves. Consider the particular case of


Exercise 15.2.2 in which r = 1, V = P1 (K), but d is still arbitrary. The d-th
15.2 Projective varieties 341

Veronesean P1 (K)(d) of P1 (K) has the property that there is a homogeneous


isomorphism
O(P1 (K)(d) ) ∼
= O(P1 (K))(d) = K[X0 , X1 ](d)
= K[X0d , X0d−1 X1 , . . . , X0 X1d−1 , X1d ] ⊆ K[X0 , X1 ].
The Veronesean transformation ϑ(d) : P1 (K) −→ P1 (K)(d) is an isomor-
phism of varieties, given by ϑ(d) ((σ : τ )) = (σ d : σ d−1 τ : · · · : στ d−1 : τ d )
for all (σ : τ ) ∈ P1 (K). Thus P1 (K)(d) ⊆ Pd (K) is a curve which is iso-
morphic to the projective line P1 (K); it is called the rational normal curve
in projective d-space, and we shall denote it by N(d) ⊆ Pd (K). The curve
N(3) ⊆ P3 (K) is also called the twisted cubic.
The phenomenon mentioned in 15.2.1(ii),(viii) is illustrated by these ratio-
nal normal curves: it follows from 14.5.26 that, provided d ≥ 3, the rational
normal curve N(d) ⊆ Pd (K) is not arithmetically Gorenstein, whereas P1 (K),
to which it is isomorphic, is.
15.2.4 Proposition. Let K be an algebraically closed field, and let r, d ∈ N.
Let V ⊆ Pr (K) be a projective variety of positive dimension, and consider the
r+d
d-th Veronesean V (d) ⊆ P( r )−1 (K) of V . Set
1
e := end(HO(Cone(V ))+ (O(Cone(V )))).

Then

(d) >1 if d > e,
arithdepth V
=1 if d = e.

In particular, if the projective variety V ⊆ Pr (K) is a curve, then V (d) is


arithmetically Cohen–Macaulay if d > e, and is not arithmetically Cohen–
Macaulay if d = e.
Proof. Let R := O(Cone(V )). By 15.1.4(ii),(iv), we have e < ∞ and
HR1
+
(R)n = 0 for all n ≤ 0. Hence the (d, 0)-th Veronesean submodule
HR+ (R)(d,0) of HR
1 1
+
(R) R(d) (see 13.5.9) is zero if d > e and is non-zero if
d = e. By 13.5.9(v), there is a homogeneous R(d) -isomorphism

=
1
HR +
(R)(d,0) −→ H(R
1
+)
(d) (R
(d)
);

furthermore, (R+ )(d) = (R(d) )+ . Note also that H(R


0
(d) ) (R
+
(d)
) = 0 since
(d)
R(d) is a domain and R+ = 0. Since there is a homogeneous ring isomor-

=
phism O(Cone(V (d) )) −→ R(d) by 15.2.2, it follows that arithdepth V (d) >
1 if d > e and arithdepth V (d) = 1 if d = e.
342 Links with projective varieties

If V is a curve, then dim R = 2, so that, since R is an integral extension do-


main of R(d) , it follows that the *local graded ring O(Cone(V (d) )) is Cohen–
Macaulay if and only if arithdepth V (d) > 1.

15.2.5 Exercise. Let K be an algebraically closed field, let d ∈ N with


d ≥ 3, and let A(d) be the subring of the ring K[X, Y ] of polynomials over
K in two indeterminates X and Y described in 13.5.12 and given by A(d) :=
K[X d , X d−1 Y, XY d−1 , Y d ]. Recall from 13.5.12 that A(d) inherits a grading
from K[X, Y ](d) , the d-th Veronesean subring of K[X, Y ].
Let Y0 , Y1 , Y2 , Y3 be independent indeterminates over K and consider the
polynomial ring K[Y0 , Y1 , Y2 , Y3 ] as the coordinate ring O(A4 (K)). Let

ψ : K[Y0 , Y1 , Y2 , Y3 ] −→ A(d)

be the (surjective, homogeneous) K-algebra homomorphism for which

ψ(Y0 ) = X d , ψ(Y1 ) = X d−1 Y, ψ(Y2 ) = XY d−1 and ψ(Y3 ) = Y d .

Let p := Ker ψ; this is a graded prime ideal of K[Y0 , Y1 , Y2 , Y3 ] and so defines


a projective variety Σ(d) := VP3 (K) (p) ⊆ P3 (K).

(i) Show that there is a morphism ρ : P1 (K) −→ Σ(d) of varieties for


which ρ((σ : τ )) = (σ d : σ d−1 τ : στ d−1 : τ d ) for all (σ : τ ) ∈ P1 (K).
(ii) Show that Y2d − Y3d−1 Y0 ∈ p.
(iii) Set U := Σ(d) \ VP3 (K) ((Y0 )) and U  := Σ(d) \ VP3 (K) ((Y3 )). Show that
U and U  form an open covering of Σ(d) .
(iv) Show that there is a morphism μ : Σ(d) −→ P1 (K) of varieties for
which

(α : β) if (α : β : γ : δ) ∈ U,
μ((α : β : γ : δ)) =
(γ : δ) if (α : β : γ : δ) ∈ U  .

(v) Show that ρ and μ are inverse isomorphisms of varieties, so that Σ(d) is
again a curve isomorphic to P1 (K).
(vi) Let t ∈ N. Use 13.5.12 to show that, when d > 3, the curve Σ(d) ⊆
P3 (K) is not arithmetically Cohen–Macaulay, but that the t-th Verone-
t+3
sean (Σ(d) )(t) ⊆ P( 3 )−1 (K) is arithmetically Cohen–Macaulay pro-
d
vided that t ≥ d−2. Is (Σ )(d−3) ⊆ P(3)−1 (K) arithmetically Cohen–
(d)
Macaulay?

15.2.6 Remarks. Let the situation and notation be as in Exercise 15.2.5,


where the curve Σ(d) ⊆ P3 (K) was constructed, for each integer d ≥ 3. As
a variety, Σ(d) is isomorphic to the projective line P1 (K). Note that Σ(3) ⊆
15.2 Projective varieties 343

P3 (K) is just the twisted cubic N(3) ⊆ P3 (K) of 15.2.3. The curve Σ(4) ⊆
P3 (K) is called the twisted quartic or Macaulay’s curve.
Exercise 15.2.5(vi) shows that the arithmetic structure of Σ(d) ⊆ P3 (K),
which reflects the way in which the curve is embedded in P3 (K), is more
complicated for d > 3 then it is when d = 3. This is another illustration of the
phenomenon mentioned in 15.2.1(ii),(viii).

15.2.7 Exercise. Let K be a field, and let U, W, S, T be indeterminates over


K. In the 5-th Veronesean subring K[U, W, S, T ](5) of the polynomial ring
K[U, W, S, T ], set f1 := U 4 T, f2 := W S 4 and f3 := U 4 S − W T 4 . Consider
the four K-subalgebras A, B, C and R of K[U, W, S, T ](5) given by
A := K[U 4 S, U 4 T, W S 4 , W S 3 T, W S 2 T 2 , W ST 3 , W T 4 ],

B := K[U 4 S, U 4 T, W S 4 , W S 3 T, W ST 3 , W T 4 ] ⊆ A,

C := K[U 4 S, U 4 T, W S 4 , W S 2 T 2 , W ST 3 , W T 4 ] ⊆ A,

R := K[f1 , f2 , f3 ] = K[U 4 T, W S 4 , U 4 S − W T 4 ] ⊆ B ∩ C ⊆ A.
Observe that each of these subrings inherits a grading from K[U, W, S, T ](5)
that turns it into a homogeneous positively graded *local integral domain. It
follows from 14.5.2 and 14.5.24 that each of them has a *canonical module
and an S2 -ification.

(i) Show (by induction on n) that, for all n ∈ N0 , the set of monomials
{U 4i W n−i S k T 4n−3i−k : 0 ≤ i ≤ n, 0 ≤ k ≤ 4n − 3i}
is a K-basis of An and conclude that dimK An = (n + 1)(5n + 2)/2.
(ii) Show that
A2 = (f1 K + f2 K + f3 K)A1
and deduce that
An = (f1 K + f2 K + f3 K)An−1 for all n ≥ 2.
Deduce that the R-module A is generated by the five homogeneous ele-
ments
g1 := 1, g2 := U 4 S, g3 := W S 3 T, g4 := W S 2 T 2 , g5 := W ST 3 .
(iii) Show that dim A = 3 = dim R, that f1 , f2 , f3 are algebraically inde-
pendent over K, that f1 , f2 , f3 is an R-sequence, and that dimK Rn =
(n + 1)(n + 2)/2 for all n ∈ N0 .
344 Links with projective varieties

(iv) For each n ∈ N0 let


Mn := {f1ν1 f2ν2 f3ν3 : ν1 , ν2 , ν3 ∈ N0 , ν1 + ν2 + ν3 = n},
the set of all monomials of degree n in f1 , f2 , f3 , and let Mn gi , for
n ∈ N0 and i ∈ {1, . . . , 5}, denote {mn gi : mn ∈ Mn }. Use part (ii)
to show that, for each n ∈ N, the members of the set
Mn g1 ∪ Mn−1 g2 ∪ Mn−1 g3 ∪ Mn−1 g4 ∪ Mn−1 g5
span the K-space An , and use the formula in part (i) for dimK An to
deduce that the above-displayed set is actually a K-basis for An . Deduce
that the five homogeneous elements g1 , g2 , g3 , g4 , g5 form a base for the
R-module A, so that A is a *free graded R-module of rank 5.
(v) Deduce that f1 , f2 , f3 is an A-sequence and conclude that the ring A is
Cohen–Macaulay.
(vi) Show that A = B + W S 2 T 2 B and B+ A ⊆ B. Deduce that there is a
homogeneous B-isomorphism A/B ∼ = 1 K, where 1 K is as defined in
14.1.13.
(vii) Show that that there is a homogeneous B-isomorphism HB 1
+
(B) ∼
= 1 K;
deduce that grade B+ = 1 and that the non-S2 locus of B is {B+ } =
Var(B+ ).
(viii) Use 2.2.15(iii), 2.2.17 and 13.5.4(ii) to show that there is a homoge-
neous isomorphism of B-algebras A ∼ = DB+ (B), and that both these
B-algebras are isomorphic to the S2 -ification of B.

Let p be the ideal (U 4 S, U 4 T, W S 2 T 2 , W ST 3 , W T 4 ) of


C = K[U 4 S, U 4 T, W S 4 , W S 2 T 2 , W ST 3 , W T 4 ].

(ix) Show that A = C + W S 3 T C, and that pA ⊆ C.


(x) Show that, for all n ∈ N, we have (W S 4 )n W S 3 T ∈/ C.
(xi) Show that C/p = K[W S 4 + p], and that W S 4 + p is transcendental
over K (in the usual sense that the only polynomial f ∈ K[X] for which
f (W S 4 + p) = 0 is the zero polynomial). Deduce that p ∈ * Spec(C),
that dim C/p = 1 and that ht p = 2.
(xii) Show that the C-module A/C is generated by W S 3 T + C, and that the
annihilator of this element is p. Deduce that there is a homogeneous C-
isomorphism Hp1 (C) ∼ = (C/p)(−1). Deduce that grade p = 1 and that
the non-S2 locus of C is Var(p).
(xiii) Show that there is a homogeneous isomorphism of C-algebras A ∼ =
Dp (C), and that both these C-algebras are isomorphic to the S2 -ification
of C.
15.2 Projective varieties 345

15.2.8 Exercise. Let K be an algebraically closed field, and consider the


homogeneous, N0 -graded K-algebras A, B and C defined in 15.2.7. Let VA ⊂
P6 (K) be the irreducible projective variety whose homogeneous coordinate
ring satisfies O(Cone(VA )) = A; also let VB , VC ⊂ P5 (K) be the irreducible
projective varieties such that O(Cone(VB )) = B and O(Cone(VC )) = C
respectively.
(i) Show that
VA = {(αs : αt : βs4 : βs3 t : βs2 t2 : βst3 : βt4 ) ∈ P6 (K) :
(α, β), (s, t) ∈ K 2 \ {(0, 0)}},
so that VA is the disjoint union of the lines in P6 (K) joining the points
(s : t : 0 : 0 : 0 : 0 : 0) and (0 : 0 : s4 : s3 t : s2 t2 : st3 : t4 ),
where (s : t) runs through the projective line P1 (K). Readers with some
background in classical algebraic geometry might recognize VA as the
standard rational normal surface scroll S(1, 4) ⊂ P6 (K) (see [27, p.
94]).
(ii) Find descriptions for VB and VC similar to that for VA in part (a).
(iii) Let e4 = (0 : 0 : 0 : 0 : 1 : 0 : 0) ∈ P6 (K). Consider the projection
map (see [30, p. 22]) π4 : P6 (K) \ {e4 } −→ P5 (K) for which
π4 ((x0 : x1 : x2 : x3 : x4 : x5 : x6 )) = (x0 : x1 : x2 : x3 : x5 : x6 )
for all (x0 : x1 : x2 : x3 : x4 : x5 : x6 ) ∈ P6 (K) \ {e4 }. Show that
π4 (VA ) = VB .
(iv) Let e3 = (0 : 0 : 0 : 1 : 0 : 0 : 0) ∈ P6 (K). Consider the projection
map π3 : P6 (K) \ {e3 } −→ P5 (K) for which
π3 ((x0 : x1 : x2 : x3 : x4 : x5 : x6 )) = (x0 : x1 : x2 : x4 : x5 : x6 )
for all (x0 : x1 : x2 : x3 : x4 : x5 : x6 ) ∈ P6 (K) \ {e3 }. Show that
π3 (VA ) = VC .
Although the two varieties VB and VC are both obtained by projecting VA
from a point, the two projections turn out to be quite different, as can be seen
from the properties of their homogeneous coordinate rings developed in 15.2.7.

In the next Chapter 16, we shall discuss the situation where R = n∈N0 Rn
is positively graded; we shall study in considerable depth the components
HR i
+
(M )n (n ∈ Z) of the i-th (i ∈ N0 ) local cohomology module, with re-
spect to the irrelevant ideal R+ of R, of a finitely generated graded R-module
M . There is very strong motivation from projective algebraic geometry for
such study, and we hope that our discussions of projective varieties in this
chapter have given the reader some hints of this motivation.
16

Castelnuovo regularity

In Chapter 15, we have seen that, when (K is an algebraically closed field,


r ∈ N and) R is the homogeneous coordinate ring of a projective variety
V ⊆ Pr (K) of positive dimension, the end of the (necessarily graded) first
1
local cohomology module HR +
(R) is of interest: see 15.2.4. This is one moti-
vation for our work in this chapter, where we shall study, in the case when R =

n∈N0 Rn is positively graded, the ends of the local cohomology modules
i
HR +
(M ) for a finitely generated graded R-module M . Perhaps the most im-
portant invariant related to these ends is the so-called (Castelnuovo–Mumford)
regularity of M . This invariant is of great significance in algebraic geometry,
and, as we shall see in 16.3.7 and 16.3.8, it provides links between local co-
homology theory and the syzygies of finitely generated graded modules over a
polynomial ring over a field.
Throughout this chapter, all graded rings and modules are to be understood
to be Z-graded.

16.1 Finitely generated components


Our first goal in this chapter is to establish the basic facts that, in the notation of
the above introduction, for each i ∈ N0 , the R0 -module HR i
+
(M )n is finitely
generated for all n ∈ Z, and is zero for all sufficiently large n. These facts,
which generalize 15.1.4(iii),(ii), are the basis for much of the work in this and
the next chapter.

16.1.1 Notation and Terminology. We shall employ the notation and ter-
minology concerning graded rings and modules described in 13.1.1, 13.1.3,
14.1.1 and 15.1.1, but, in accordance with our convention for the whole of this
chapter, restricted to the special case in which G = Z. In particular, all polyno-
16.1 Finitely generated components 347

mial rings R0 [X1 , . . . , Xn ] (in n indeterminates X1 , . . . , Xn over a commuta-


tive Noetherian ring R0 ) considered in this chapter will be positively Z-graded
so that R0 is the component of degree 0 and deg Xi = 1 for all i = 1, . . . , n.

Recall [7, p. 29] that, when R = n∈N0 Rn is positively graded, we say
that R is homogeneous if and only if R is generated as an R0 -algebra by its
forms of degree 1, that is, if and only if R = R0 [R1 ].

Also, when R is merely positively graded, L = n∈Z Ln is a graded R-
module and t ∈ Z is fixed, we define

L≥t := Ln ,
n∈Z
n≥t

a graded submodule of L. For example, in 16.1.6 below, we shall be concerned


with the graded R-modules DR+ (M )≥t (for a graded R-module M ).
We shall need to use the following graded version of the Prime Avoidance
Theorem; for a proof, we refer the reader to [7].
16.1.2 Homogeneous Prime Avoidance Lemma. (See [7, 1.5.10].) Let R
be graded and assume the ideal a is graded and generated by elements of
positive degree. Let p1 , . . . , pn ∈ Spec(R) be such that, for all i = 1, . . . , n,
we have a ⊆ pi . Then there exists a homogeneous element in a\(p1 ∪· · ·∪pn ).

We can improve on the result of 16.1.2 when R0 is local with infinite residue
field and a is generated by elements of degree 1.

16.1.3 Lemma. (See [7, 1.5.12].) Assume that R = n∈Z Rn is graded,
that R0 is local with infinite residue field, and that the ideal a is graded and
generated by elements of degree 1. Let b1 , . . . , bn be ideals of R such that, for
all i = 1, . . . , n, we have a ⊆ bi . Then there exists a homogeneous element of
degree 1 in a \ (b1 ∪ · · · ∪ bn ).
Proof. Set a1 := a ∩ R1 , and let m0 denote the maximal ideal of R0 . The
hypotheses ensure that, for all i = 1, . . . , n,
(bi ∩ a1 + m0 a1 )/m0 a1 = a1 /m0 a1 .
n
Since R0 /m0 is infinite, there exists r1 ∈ a1 \ i=1 (bi ∩ a1 + m0 a1 ).

16.1.4 Lemma. Assume that R = n∈Z Rn is graded and that the ideal a
is graded and generated by elements of positive degree. Let M be a finitely
generated R-module such that aM = M and gradeM a > 0.
(i) There is a homogeneous element in a which is a non-zerodivisor on M .
348 Castelnuovo regularity

(ii) If, in addition, R0 is local with infinite residue field and a is generated by
elements of degree 1, then there exists a homogeneous element of degree
1 in a which is a non-zerodivisor on M .

Proof. Since a ⊆ p∈Ass M p, part (i) is immediate from 16.1.2, while part
(ii) follows from 16.1.3.

We are now ready to prove the basic finiteness and vanishing theorem which
was mentioned in the first paragraph of this section.

16.1.5 Theorem. Assume that R = n∈N0 Rn is positively graded; let M
be a finitely generated graded R-module.

(i) For all i ∈ N0 and all n ∈ Z, the R0 -module HR i


+
(M )n is finitely
generated.
(ii) There exists r ∈ Z such that HR
i
+
(M )n = 0 for all i ∈ N0 and all
n ≥ r.

Proof. Let i ∈ N0 . We prove, by induction on i, that HR i


+
(M )n is a finitely
generated R0 -module for all n ∈ Z, and is zero for all sufficiently large values
of n. This will prove not only part (i) but also part (ii), because, in view of
i
3.3.3, there can only be finitely many integers i for which HR +
(M ) = 0, since
i
HR+ (M ) = 0 for all i > ara(R+ ).
0
We consider first the case where i = 0. Since HR +
(M ) is a submodule
of M , it is finitely generated as an R-module, and so there exists u ∈ N
such that (R+ )u HR 0
+
(M ) = 0. Now (R+ )i HR 0
+
(M )/(R+ )i+1 HR 0
+
(M ) is
a Noetherian R/R+ -module, and so is a Noetherian R0 -module (for each
i = 0, . . . , u − 1). Therefore HR 0
+
(M ) is a Noetherian R0 -module. Hence
HR+ (M )n is a finitely generated R0 -module for all n ∈ Z, and only finitely
0
0
many of the HR +
(M )n can be non-zero.
Now suppose that i > 0 and our desired result has been proved for smaller
values of i (and for all choices of the finitely generated graded R-module M ).
Since ΓR+ (M ) is a graded submodule of M , it follows from 2.1.7(iii) that, for
all i ∈ N, there is a (homogeneous) isomorphism
i
HR +
(M ) ∼ i
= HR +
(M/ΓR+ (M )).

Thus, for the purpose of this inductive step, we can replace M by M/ΓR+ (M )
and so (in view of 2.1.2 and 2.1.1) assume that R+ contains a non-zerodivisor
on M . This we do.
We can assume that M = 0; we therefore assume that M = R+ M . Then,
by Lemma 16.1.4, there exists a homogeneous element r ∈ R+ which is a
non-zerodivisor on M . Let t denote the degree of r.
16.1 Finitely generated components 349
r
The exact sequence 0 −→ M −→ M (t) −→ (M/rM )(t) −→ 0 of graded
R-modules and homogeneous homomorphisms induces a long exact sequence
(in *C(R)) of local cohomology modules, from which we deduce (with the aid
of 14.1.10(ii)) an exact sequence
r
i−1
HR +
(M/rM )n+t −→ HR
i
+
(M )n −→ HR
i
+
(M )n+t
of R0 -modules for all n ∈ Z.
By the inductive hypothesis, there exists s ∈ Z such that
i−1
HR +
(M/rM )j = 0 for all j ≥ s.
r
Hence, for all n ≥ s − t, the sequence 0 −→ HR i
+
(M )n −→ HRi
+
(M )n+t is
exact; therefore, since HR+ (M ) is R+ -torsion and r ∈ R+ , we have
i

i
HR +
(M )n = 0 for all n ≥ s − t.
i−1
The inductive hypothesis also yields that HR +
(M/rM )j is a finitely gen-
erated R0 -module for all j ∈ Z. Fix n ∈ Z and let k ∈ N0 be such that
n + kt ≥ s − t, so that HR i
+
(M )n+kt = 0 by the last paragraph. Now, for each
j = 0, . . . , k − 1, there is an exact sequence
r
i−1
HR +
(M/rM )n+(j+1)t −→ HR
i
+
(M )n+jt −→ HR
i
+
(M )n+(j+1)t
i−1
of R0 -modules, and HR +
(M/rM )n+(j+1)t is finitely generated over R0 . We
i
can therefore deduce, successively, that HR +
(M )n+jt is finitely generated
over R0 for j = k − 1, k − 2, . . . , 1, 0; we thus conclude that HR i
+
(M )n
is finitely generated over R0 , and the inductive step is complete.

16.1.6 Corollary. Assume that R = n∈N0 Rn is positively graded; let

M = n∈Z M n be a finitely generated graded R-module. Then (with the
notation of 16.1.1)
(i) DR+ (M )≥t is a finitely generated R-module for all t ∈ Z;
(ii) DR+ (M )n is a finitely generated R0 -module for all n ∈ Z; and
(iii) for all sufficiently large n, the restriction to Mn of the map ηM : M −→
DR+ (M ) of 2.2.6(i)(c) provides an R0 -isomorphism
(ηM )n : Mn −→ DR+ (M )n .
Proof. These claims are all immediate consequences of Theorem 16.1.5 and
the exact sequence
ξM ηM ζM
0 −→ ΓR+ (M ) −→ M −→ DR+ (M ) −→ HR
1
+
(M ) −→ 0
of graded R-modules and homogeneous homomorphisms: see 13.5.4(i).
350 Castelnuovo regularity

Theorem 16.1.5 shows that, when R is positively graded, the theory of local
cohomology with respect to the irrelevant ideal of R is particularly satisfactory.
It is natural to ask whether local cohomology with respect to other homoge-
neous ideals exhibits similar properties. The next three exercises concern ex-
amples where the answer is clearly negative: the conclusions of these exercises
should be compared with those of Theorem 16.1.5.
16.1.7 Exercise. Let (R0 , πR0 ) be a discrete valuation ring, and let R denote
the polynomial ring R0 [X]. Let m := πR + XR, the unique maximal graded
ideal of R. Let M be a non-zero, torsion-free, finitely generated, graded R-
module.
1
(i) Use the Graded Finiteness Theorem 14.3.10 to show that Hm (M ) is
finitely generated.
2
(ii) Show that Hm (M/XM ) = 0.
(iii) Use 13.5.5 to see that there exists n0 ∈ Z for which Hm 1
(M/XM )n0 is
not finitely generated as R0 -module.
X
(iv) Use the exact sequence 0 → M −→ M (1) → (M/XM )(1) → 0 to
show that there exists n1 ∈ Z such that Hm 2
(M )n is not finitely gener-
ated for all n < n1 and there exists n2 ∈ Z such that Hm 2
(M )n = 0 for
all n > n2 .
16.1.8 Exercise. Let R0 be a commutative Noetherian ring, and let R denote
the polynomial ring R0 [X, Y ]. Use Exercise 13.5.5 to show that, for all n ∈ Z,
1
the R0 -module HXR (R)n is free but not finitely generated.
Compare this with the conclusions of Theorem 16.1.5.
16.1.9 Exercise. Let (R0 , πR0 ) be a discrete valuation ring as in Exercise
16.1.7, and let R denote the graded ring R0 [X, Y ], as in Exercise 16.1.8. Let
a := πR + XR. Use 16.1.8 and 14.1.11 to show that, for all n ∈ Z, the
R0 -module Ha2 (R)n is not finitely generated.

16.1.10 Exercise. Assume that R = n∈N0 Rn is positively graded and
such that (R0 , m0 ) is local. Set m = m0 R + R+ , the unique graded maximal
ideal of R. Let M be a finitely generated graded R-module such that d :=
dim(M/ΓR+ (M )) > 0.
(i) Show that there is a homogeneous element a of positive degree t which
is a non-zerodivisor on M := M/ΓR+ (M ).
d
(ii) Use Theorem 6.1.4 to show that Hm (M ) = 0, and part (i) to deduce that
there are infinitely many negative n ∈ Z such that Hmd
(M )n = 0.
(iii) Use 14.1.7 and 14.1.12 to show that, for each i ∈ N0 , there are only
finitely many integers n ∈ Z for which Hm i
(ΓR+ (M ))n = 0.
16.2 The basics of Castelnuovo regularity 351
d
(iv) Prove that Hm (M )n = 0 for infinitely many negative integers n.

16.2 The basics of Castelnuovo regularity


Assume that R is positively graded and homogeneous, and let M be a finitely
generated graded R-module. It is clear from 3.3.3 and 16.1.5 that there is an
integer r such that, for all s ∈ Z with s > r, we have HR i
+
(M )s−i = 0 for all
i ∈ Z (or, equivalently, for all i ∈ N0 ). The Castelnuovo regularity of M is the
infimum of the set of integers r with this property (interpreted as −∞ if this
infimum does not exist). This is a very important invariant of M , and a topic
which has featured in much recent research. In this section, we develop the
basic theory of this and some related, slightly more complicated, invariants.

16.2.1 Definitions. Assume that R is positively graded and homogeneous


(see 16.1.1). Let M be a finitely generated graded R-module, and let r ∈ Z
and l ∈ N0 .

(i) We follow A. Ooishi in [65, Definition 1] and say that M is r-regular


i
in the sense of Castelnuovo–Mumford if and only if HR +
(M )s−i = 0
for all i, s ∈ Z with s > r. In practice, the phrase ‘in the sense of
Castelnuovo–Mumford’ is usually omitted.

As pointed out just before this definition, there does exist an r ∈ Z such that
i
M is r-regular. Note that M is r-regular if and only if HR +
(M )s−i = 0 for
all i, s ∈ Z with s > r and i ≥ 0. This observation leads to the following more
general definition.

(ii) We say that M is r-regular at and above level l if and only if


i
HR +
(M )s−i = 0 for all i, s ∈ Z with s > r and i ≥ l.

Thus M is r-regular if and only if it is r-regular at and above level 0.

Motivation for Definition 16.2.1 comes from Theorem 16.2.5 below, which
is a version of a proposition of D. Mumford [54, p. 99], adapted to the con-
text of local cohomology. (See also Ooishi [65, Theorem 2].) However, before
presenting Theorem 16.2.5, we provide some technical comments which will
help us to make appropriate reductions in the proof of the theorem and other
results.

16.2.2 Remarks. Assume that R = n∈Z Rn is graded, and that the ideal a
is graded. Let R0 be a commutative Noetherian ring, and let f0 : R0 −→ R0
352 Castelnuovo regularity

be a flat ring homomorphism. Set R := R ⊗R0 R0 , and let f : R −→ R be


the natural ring homomorphism.
(i) Since R is a finitely generated R0 -algebra by [7, Theorem 1.5.5], it
follows that R is Noetherian. Furthermore, R has the structure of a

graded ring given by R = n∈Z Rn , where Rn is the natural image of
Rn ⊗R0 R0 in R for each n ∈ N. Thus the ring homomorphism f is
homogeneous (in the sense of 14.1.3), and flat.
(ii) It therefore follows from the Graded Flat Base Change Theorem 14.1.9
that the sequences (Hai ( • ) ⊗R R )i∈N0 and (HaR
i 
 (( • ) ⊗R R ))i∈N0

(restricted to the ‘graded’ category *C(R)) are isomorphic negative


connected sequences of covariant functors from *C(R) to *C(R ).

(iii) Let L = n∈Z Ln be a graded R-module. Then L ⊗R R is a graded
R -module: if we use the natural isomorphisms
∼ ∼
L ⊗R R = L ⊗R (R ⊗R0 R0 ) −→ (L ⊗R R) ⊗R0 R0 −→ L ⊗R0 R0
= =

to identify L ⊗R R with L ⊗R0 R0 =: L , then the grading on L is



given by L = n∈Z Ln , where Ln is the natural image of Ln ⊗R0 R0
in L .
(iv) It follows from part (ii), and part (iii) applied to the graded R-module
Hai (M ) (where M is a graded R-module), that, for all i ∈ N0 and all
n ∈ Z, there is a natural equivalence of functors (from *C(R) to C(R0 ))
Hai ( • )n ⊗R0 R0 −→ HaR
i 
 (( • ) ⊗R R )n .

(v) We can deduce from 4.3.5, 13.3.14 and the formula in 2.2.15(ii) that
there is a natural equivalence of functors
ε : Da ( • ) ⊗R R −→ DaR (( • ) ⊗R R )
from the ‘graded’ category *C(R) to *C(R ).
(vi) If we now use part (v), and part (iii) applied to the graded R-module
Da (M ) (where M is a graded R-module), we obtain, for each n ∈ Z, a
natural equivalence of functors (from *C(R) to C(R0 ))
Da ( • )n ⊗R0 R0 −→ DaR (( • ) ⊗R R )n .
We give two examples where the ideas of 16.2.2 can be used.

16.2.3 Example. Assume that R = n∈N0 Rn is positively graded and ho-
mogeneous (see 16.1.1), let M be a graded R-module, and let p0 ∈ Spec(R0 ).
We can apply the techniques (and notation) of 16.2.2 with the choices a =
R+ and R0 = (R0 )p0 , a flat R0 -algebra. The ring R = R ⊗R0 (R0 )p0 is
again positively graded and homogeneous, and M  = M ⊗R0 (R0 )p0 is a
16.2 The basics of Castelnuovo regularity 353

graded R -module. Also, R+ R = R+ 


, the irrelevant ideal of R . It follows
from 16.2.2(iv) that,for each i ∈ N0 and each n ∈ Z, there is an isomorphism
of (R0 )p0 -modules HR i
+
(M )n p0 ∼= HR i 
 (M )n . Finally, note that the 0-th
+
component of R is a local ring isomorphic to (R0 )p0 .

16.2.4 Example. Assume that R = n∈N0 Rn is positively graded and ho-
mogeneous, and such that (R0 , m0 ) is local; let M be a graded R-module.
We can apply the techniques (and notation) of 16.2.2 with the choices a =
R+ and R0 = R0 [X]m0 R0 [X] , the localization of the polynomial ring R0 [X]
at the prime ideal m0 R0 [X]. This R0 is a flat R0 -algebra, and
R = R ⊗R0 R0 [X]m0 R0 [X]
is positively graded and homogeneous, M  = M ⊗R0 R0 [X]m0 R0 [X] is a
graded R -module, and R+ R = R+ 
, the irrelevant ideal of R .
It follows from 16.2.2(iv) that, for each i ∈ N0 and each n ∈ Z, there is an
isomorphism of R0 [X]m0 R0 [X] -modules
i
HR +
(M )n ⊗R0 R0 [X]m0 R0 [X] ∼ i
= HR 
 (M )n ;
+

i
thus, since R0 [X]m0 R0 [X] is a faithfully flat R0 -algebra, HR +
(M )n = 0 if and
i 
only if HR (M )n = 0.
+
Finally, note that the 0-th component of R is a local ring isomorphic to
R0 [X]m0 R0 [X] , having infinite residue field.

16.2.5 Theorem. Assume that R = n∈N0 Rn is positively graded and ho-
mogeneous; let M be a finitely generated graded R-module. Let r ∈ Z and
l ∈ N: the reader should note that we are assuming that l is positive.
Assume that HR i
+
(M )r+1−i = 0 for all i ≥ l. Then M is r-regular at and
above level l (see 16.2.1(ii)).
Proof. It suffices to show that, for each p0 ∈ Spec(R0 ), the (R0 )p0 -module
(HR i
+
(M )s−i )p0 vanishes for all i ≥ l and all s > r.
We can apply Examples 16.2.3 and 16.2.4 in turn to see that it is enough for
us to establish the claim in the statement of the theorem under the additional
assumption that (R0 , m0 ) is a local ring with infinite residue field, and we shall
make this assumption in what follows.
It follows from [7, 1.5.4] that dim M is finite, and we are going to argue by
induction on dim M . When dim M = −1 there is nothing to prove, and when
dim M = 0 the result is an easy consequence of Grothendieck’s Vanishing
Theorem 6.1.2.
Now suppose that dim M > 0 and our desired result has been proved for
all finitely generated graded R-modules of smaller dimension. Since ΓR+ (M )
354 Castelnuovo regularity

is a graded submodule of M , it follows from 2.1.7(iii) that, for each i ∈ N,


there is a homogeneous isomorphism HR i
+
(M ) ∼
= HR i
+
(M/ΓR+ (M )). In
particular, since l is positive, HR+ (M/ΓR+ (M ))r+1−i = 0 for all i ≥ l. Now
i

we must have dim(M/ΓR+ (M )) ≤ dim M : if this inequality is strict, we can


use the inductive assumption to achieve our aim; otherwise, for the purpose
of this inductive step, we can replace M by M/ΓR+ (M ) and so (in view of
2.1.2 and 2.1.1) assume that M is R+ -torsion-free and that R+ contains a
non-zerodivisor on M . This we do.
Now M = R+ M (since otherwise M = 0), and so, by Lemma 16.1.4(ii),
there exists a homogeneous element a ∈ R1 (and so of degree 1) which is a
non-zerodivisor on M .
a
The exact sequence 0 −→ M −→ M (1) −→ (M/aM )(1) −→ 0 of graded
R-modules and homogeneous homomorphisms induces a long exact sequence
(in *C(R)) of local cohomology modules, from which we deduce (with the aid
of 14.1.10(ii)) an exact sequence
a
i
HR +
(M )n −→ HR
i
+
(M/aM )n −→ HR
i+1
+
(M )n−1 −→ HR
i+1
+
(M )n
of R0 -modules, for all i ∈ N and all n ∈ Z. Use of this with n = r+1−i shows
that HRi
+
(M/aM )r+1−i = 0 for all i ≥ l. Now dim(M/aM ) < dim M
(since a lies outside every minimal prime ideal of Supp M ), and so it follows
from the inductive hypothesis that M/aM is r-regular at and above level l.
Fix an integer i ≥ l. For each n ∈ Z, there is an exact sequence
a
i
HR +
(M )n−1−i −→ HR
i
+
(M )n−i −→ HR
i
+
(M/aM )n−i
i i
of R0 -modules. We know that HR +
(M )r+1−i = HR +
(M/aM )r+2−i =
0, and so it follows from the above exact sequence (with n = r + 2) that
i
HR +
(M )r+2−i = 0. We can now repeat this argument, using induction, to
deduce that HR i
+
(M )r+j−i = 0 for all j ∈ N. It follows that M is r-regular
at and above level l.
In the statement (and proof) of Theorem 16.2.5, we stressed that the integer
l is assumed to be positive. In fact, if l in that theorem is replaced by 0, then the
resulting statement is no longer always true: this is illustrated by the following
exercise.
16.2.6 Exercise. Let the situation and notation be as in Theorem 16.2.5. Find
a counterexample to the statement ‘if HR i
+
(M )r+1−i = 0 for all i ≥ 0, then
M is r-regular at and above level 0 (that is, M is r-regular)’ by considering
R := R0 [X1 , . . . , Xn ], the ring of polynomials in n (≥ 1) indeterminates over
a commutative Noetherian ring R0 , and M := tR0 ⊕ R for an appropriate
value of the integer t, where tR0 is as defined in 14.1.13.
16.2 The basics of Castelnuovo regularity 355

The result of Theorem 16.2.5 can be rephrased thus: for a fixed positive
integer l, if HRi
+
(M )r+1−i = 0 for all i ≥ l, then HR i
+
(M )s−i = 0 for all
i ≥ l and s > r. Some readers might find the interest in ‘reverse diagonal
vanishing’ surprising, and they might find the following diagram helpful. The
diagram concerns the special case of Theorem 16.2.5 in which R0 is Artinian,
so that, for all i, n ∈ Z, the R0 -module HR i
+
(M )n has finite length hiM (n).
i
In the diagram, hM (n) is plotted at the position (n, i) in the Oni co-ordinate
plane. Theorem 16.2.5 shows that, if in the diagram
i
6
.... 0 0 0 0 .... 0 0 0 .... 0 0 0 0 ....
.... 0 0 0 0 ....
0 0 0 .... 0 ....0 0 0
0 0 .. .... d
@
@ 0 . .. ....
@ ..
@ . . 0 ....
.
@ .. 0 0 ....
@
@ 0 0 0 ....
@ 0 0 . ....
@ ..
@ 0 .. . ....
r
@ . .. 0 ....
@ ..
h iM (n) @ . 0 0 ....
@ 0 0 0 ....
@
@ 0 0 0 ....
l−1 @
@
@
O @r -n
.... 0 0 0 0 .... 0 0 0 .... 0 0 0 0 ....
.... 0 0 0 0 .... 0 0 0 .... 0 0 0 0 ....

(in which d denotes dim M ) there is a line of zeros on the line i + n = r + 1


above the line i = l − 1, then there must be a similar line of zeros on the line
i + n = s above the line i = l − 1 for every integer s > r.

Some readers may wonder whether ‘vertical vanishing’ works just as well in
this context as ‘reverse diagonal vanishing’. It doesn’t: the following example
provides a counterexample to the statement ‘if HR i
+
(M )r+1 = 0 for all i ∈ N
356 Castelnuovo regularity

(where R is positively graded and homogeneous and M is a finitely generated


graded R-module), then HR i
+
(M )s = 0 for all i ∈ N and s > r’.

16.2.7 Example. We consider again the example studied in 13.5.12. Thus


K is a field, d ∈ N with d ≥ 3, and A(d) is the subring of the d-th Verone-
sean subring K[X, Y ](d) of the ring K[X, Y ] of polynomials over K in two
indeterminates X and Y given by A(d) := K[X d , X d−1 Y, XY d−1 , Y d ], with
grading inherited from K[X, Y ](d) .
i
Since A(d) is a 2-dimensional domain, we have HA (d)+
(A(d) ) = 0 if i = 0
or i > 2. By 13.5.12, for n ∈ Z, we have HA(d)+ (A(d) )n = 0 if and only if
1

1 ≤ n ≤ d − 3.
Let φ : A(d) −→ K[X, Y ](d) denote the inclusion homomorphism. We
showed in 13.5.12 that there is a homogeneous A(d) -isomorphism Coker φ ∼ =
1
HA (d)+
(A (d) ). It therefore follows from 2.1.7(i) that there is a homogeneous
A(d) -isomorphism H 2 A(d)+ (A(d) ) ∼
= H2
A(d)+ (K[X, Y ](d) ). Further, arguments
we provided in 13.5.12, together with 13.5.9 and the Graded Independence
Theorem 14.1.7, show that there are homogeneous A(d) -isomorphisms
2
HA (d)+
(K[X, Y ](d) ) ∼ 2
= HA (d)+ K[X,Y ]
(d) (K[X, Y ]
(d)
)
2 (d)
= H(K[X,Y ]+ )(d) (K[X, Y ] )
∼ 2
= (H(X,Y (d,0)
) (K[X, Y ])) .
2
It now follows from 13.5.3 and 13.5.9 that HA (d)+
(A(d) )n = 0 if and only if
n < 0.
We can now conclude, for r, s ∈ Z, that HA i
(d)+
(A(d) )s−i = 0 for all i ∈ Z
if and only if s > d − 2. Thus A(d) is r-regular if and only if r ≥ d − 2.
Notice also that HAi
(d)+
(A(d) )0 = 0 for all i ∈ N, whereas, when d > 3,
1
we have HA(d)+ (A(d) )1 = 0: thus this example shows that the statement ‘if
HR i
+
(M )r+1 = 0 for all i ∈ N (where R is positively graded and homoge-
i
neous and M is a finitely generated graded R-module), then HR +
(M )s = 0
for all i ∈ N and s > r’ is false.

16.2.8 Remarks. Assume that R is positively graded and homogeneous; let


M be a finitely generated graded R-module. Let r ∈ Z and let l be a positive
integer.

(i) It follows from Theorem 16.2.5 that M is r-regular at and above level l
if and only if HR i
+
(M )r+1−i = 0 for all i ≥ l.
(ii) However, M is r-regular at and above level 0 (that is, M is r-regular) if
and only if HR i
+
(M )r+1−i = 0 for all i ∈ Z and HR
0
+
(M )s = 0 for all
16.2 The basics of Castelnuovo regularity 357

s > r. This is precisely the condition used to define the condition ‘M is


r-regular’ by D. Eisenbud and S. Goto [11, p. 95].
We are now ready to define Castelnuovo regularity, and also some refine-
ments which have been studied by several authors, including U. Nagel [57],
Nagel and P. Schenzel [58, Definition 6.1] and L. T. Hoa and C. Miyazaki [36,
§2].

16.2.9 Definition. Assume that R = n∈N0 Rn is positively graded and
homogeneous, and let M be a finitely generated graded R-module. Let l ∈ N0 .
The end of a graded R-module was defined in 14.1.1.
We define the (Castelnuovo–Mumford) regularity reg(M ) of M by
 
reg(M ) := sup end(HR i
+
(M )) + i : i ∈ N0
 
= sup end(HR i
+
(M )) + i : 0 ≤ i ≤ dim M .

Also, the (Castelnuovo–Mumford)regularity regl (M ) of M atand above level


l is defined by regl (M ) := sup end(HR i
+
(M )) + i : i ≥ l . Thus the reg-
ularity reg0 (M ) of M at and above level 0 is the regularity reg(M ) of M .
Furthermore, regl (M ) = inf {r ∈ Z : M is r-regular at and above level l} .
i
Since there are only finitely many integers i for which HR +
(M ) = 0, it
follows from 16.1.5(ii) that reg(M ) is either an integer or −∞. Note also that
regl (M ) ≤ regl−1 (M ) ≤ · · · ≤ reg0 (M ) = reg(M ).
Observe also that, by 14.1.10(ii), regl (M (t)) = regl (M ) − t for all t ∈ Z.
16.2.10 Exercise. Let the situation be as in 16.2.9.

(i) Show that regl (M ) = −∞ if and only if HRi


+
(M ) = 0 for all i ≥ l.
(ii) Let N be an R+ -torsion graded submodule of M . Show that regl (M ) =
regl (M/N ) for all l > 0.
(iii) Let p be a minimal prime of R+ + (0 :R M ) and assume that l ≤
ht(p/(0 :R M )). Show that regl (M ) > −∞.
(iv) Assume that R0 is Artinian and local, and that M = 0. Show that
regl (M ) = −∞ if and only if l > ht ((R+ + (0 :R M ))/(0 :R M )).
16.2.11 Exercise. Let R0 be a commutative Noetherian ring, let n ∈ N,
and let R := R0 [X1 , . . . , Xn ], the ring of polynomials over R0 . Use Example
13.5.3 to show that reg(R) = 0.

16.2.12 Proposition. Assume that R = n∈N0 Rn is positively graded and

homogeneous, and let M = n∈Z Mn be a finitely generated graded R-
module. Then reg1 (M ) = −∞ if and only if Mn = 0 for all n  0.
358 Castelnuovo regularity

Proof. (⇐) Suppose n0 ∈ N is such that Mn = 0 for all n ∈ Z with |n| ≥ n0 .


Then (R+ )2n0 M = 0, so that M is R+ -torsion. Therefore, by 2.1.7(i), we have
HR i
+
(M ) = 0 for all i > 0, and so reg1 (M ) = −∞.
(⇒) Assume that reg1 (M ) = −∞. This means that HR i
+
(M ) = 0 for all
i
i > 0. It therefore follows from 2.1.7(iii) and 2.1.2 that HR+ (M/ΓR+ (M )) =
0 for all i ∈ N0 . Hence M/ΓR+ (M ) = R+ (M/ΓR+ (M )), by 6.2.7 and 6.2.4.
It follows from this that, since M/ΓR+ (M ) is finitely generated and graded,
we must have M/ΓR+ (M ) = 0. Hence M = ΓR+ (M ), and the desired result
follows from this and the fact that M is finitely generated.
16.2.13 Exercise. Let the situation be as in Proposition 16.2.12. Show that
reg(M ) = −∞ if and only if M = 0.

16.2.14 Exercise. Assume that R = n∈N0 Rn is positively graded and

homogeneous. Let M = n∈Z Mn be a finitely generated graded R-module,
let l, l ∈ N0 with l ≥ l , and let r ∈ Z.
It is clear that, if M is r-regular at and above level l , then M is r-regular at
and above level l. Give an example to show that the converse statement is not
always true.

16.2.15 Exercise. Assume that R = n∈N0 Rn is positively graded and
homogeneous, let l ∈ N0 , and let 0 −→ L −→ M −→ N −→ 0 be an
exact sequence of finitely generated graded R-modules and homogeneous ho-
momorphisms. Show that
(i) reg(L) ≤ max {reg(M ), reg(N ) + 1},
(ii) regl+1 (L) ≤ max regl+1 (M ), regl (N ) + 1 ,
(iii) regl (M ) ≤ max regl (L), regl (N ) , and
(iv) regl (N ) ≤ max regl+1 (L) − 1, regl (M ) .

16.3 Degrees of generators


The following theorem is one of the main reasons for our introduction of the
concept of regularity. Among its consequences are the links, hinted at in the
introduction to this chapter, between local cohomology theory and the syzygies
of finitely generated graded modules over a polynomial ring over a field.

16.3.1 Theorem. Assume that R = n∈N0 Rn is positively graded and ho-

mogeneous, and let M = n∈Z Mn be a non-zero finitely generated graded
R-module. Then M can be generated by homogeneous elements of degrees not
exceeding reg(M ).
16.3 Degrees of generators 359

Proof. Our strategy in this proof has some similarities to that for our proof of
Theorem 16.2.5 above, inasmuch as we apply 16.2.3 and 16.2.4 to reduce to
the case in which R0 is local with infinite residue field.

Let N be the graded submodule of M generated by n≤reg(M ) Mn : we
must show that M = N .
It suffices to show that Mp0 = Np0 for each p0 ∈ Spec(R0 ). We use 16.2.3:
consider the positively graded, homogeneous ring R = R ⊗R0 (R0 )p0 and
the graded R -module M  = M ⊗R0 (R0 )p0 , which is finitely generated. By
16.2.3, reg(M  ) ≤ reg(M ); also, under the canonical isomorphism between
M  and Mp0 , the R -submodule of M  generated by all homogeneous elements
of degrees not exceeding reg(M ) is mapped onto Np0 . It is therefore enough
for us to establish the claim in the statement of the theorem under the additional
assumption that R0 is local. We can then use 16.2.4 in a similar way to see
that it is enough for us to establish the claim under the additional assumption
that (R0 , m0 ) is a local ring with infinite residue field, and we shall make this
assumption in what follows.
Now d := dim M is finite, and reg(M ) = −∞ since M = 0 (by 16.2.13).
We argue by induction on d. Consider first the case where d = 0. Then
Ass M = {m0 ⊕ R+ }, so that there exists t ∈ N such that (R+ )t M = 0.
In this case, M = ΓR+ (M ) and so HR i
+
(M ) = 0 for all i ∈ N, by 2.1.7(ii).
Thus reg(M ) = end(M ), and it is obvious that M can be generated by homo-
geneous elements of degrees not exceeding end(M ).
Now suppose that dim M > 0 and our desired result has been proved for
all non-zero, finitely generated graded R-modules of smaller dimensions. Of
course, ΓR+ (M ), if non-zero, can be generated by homogeneous elements of
degrees not exceeding end(ΓR+ (M )); since end(ΓR+ (M )) ≤ reg(M ), it fol-
lows from the exact sequence
0 −→ ΓR+ (M ) −→ M −→ M/ΓR+ (M ) −→ 0
that it is enough for us to show that M/ΓR+ (M ) can be generated by homo-
geneous elements of degrees not exceeding reg(M ). Note that
reg(M/ΓR+ (M )) = reg1 (M ) ≤ reg(M ).
Now dim(M/ΓR+ (M )) ≤ dim M : if this inequality is strict, we can use the
inductive assumption to achieve our aim; otherwise, for the purpose of this
inductive step, we can replace M by M/ΓR+ (M ) and so (in view of 2.1.2
and 2.1.1) assume that M is R+ -torsion-free and that R+ contains a non-
zerodivisor on M . This we do.
Note that M = R+ M . Then, by Lemma 16.1.4(ii), there exists a homoge-
neous element a ∈ R1 (and so of degree 1) which is a non-zerodivisor on M .
360 Castelnuovo regularity

Apply 16.2.15(iv) to the exact sequence


a
0 −→ M (−1) −→ M −→ M/aM −→ 0
of graded R-modules and homogeneous homomorphisms to see that
reg(M/aM ) ≤ max reg1 (M (−1)) − 1, reg(M ) = reg(M ).
Now dim(M/aM ) < dim M , and so it follows from the inductive hypoth-
esis that M/aM can be generated by homogeneous elements of degrees not
exceeding reg(M/aM ). Hence M = N + aM . Now R is *local with unique
*maximal graded ideal m := m0 ⊕ R+ ; it follows from Nakayama’s Lemma
that Mm = Nm , so that M = N by 14.1.2(ii) applied to M/N .
16.3.2 Exercise. Let R := K[X1 , . . . , Xn ], the ring of polynomials in n
(≥ 1) indeterminates over a field K. Let a be a proper graded ideal of R such
0 n
that HR +
(R/a) = HR +
(R/a) = 0. Use 16.2.11, 16.2.15 and 16.3.1 to show
that 0 ≤ reg (R/a) = reg2 (a) − 1.
1

In the following definition and subsequent exercises, we consider a geomet-


ric significance of the Castelnuovo regularity.
16.3.3 Definition and Remark. Let K be an algebraically closed field, and
let r ∈ N. Let V ⊂ Pr (K) be a projective variety. Let R := K[X0 , . . . , Xr ],
and consider R as the coordinate ring O(Ar+1 (K)). Then the graded prime
ideal IPr (K) (V ) := IAr+1 (K) (Cone(V )) of R is non-zero: we define the Cast-
elnuovo–Mumford regularity reg(V ) of V by reg(V ) := reg(IPr (K) (V )).
0 1
Note that HR +
(R) = HR +
(R) = 0, and the homogeneous coordinate ring
O(Cone(V )) of V is an integral domain. It follows from these observations
and 16.3.2 that
reg(V ) = reg(IPr (K) (V )) = reg1 (IPr (K) (V )) = reg2 (IPr (K) (V ))
= reg1 (O(Cone(V ))) + 1 = reg(O(Cone(V ))) + 1.
Note that, by 16.3.1, the vanishing ideal IPr (K) (V ) can be generated by homo-
geneous polynomials of degrees not exceeding reg(V ).
16.3.4 Exercise. Let K be an algebraically closed field, and let r ∈ N. Let
V ⊂ Pr (K) be a projective variety such that dim V > 0. Let n ∈ N be such
that reg(V ) ≤ n + 1.
r+n
(i) Show that the n-th Veronesean V (n) ⊆ P( r )−1 (K) of V has

arithdepth V (n) > 1.


(ii) Show that reg(V (n) ) ≤ dim V + 2.
16.3 Degrees of generators 361

16.3.5 Exercise. Let K be an algebraically closed field, and let d ∈ N with


d > 1. Here we study again the rational normal curve in projective d-space
N(d) ⊆ Pd (K) of 15.2.3.
Consider the Veronesean subring K[X0 , X1 ](d) of the ring K[X0 , X1 ] of
polynomials over K in two indeterminates X0 , X1 , and the polynomial ring
K[Y0 , . . . , Yd ] in d + 1 indeterminates Y0 , . . . , Yd as O(Ad+1 (K)), the co-
ordinate ring of affine (d + 1)-space over K. Let φ : K[Y0 , . . . , Yd ] −→
K[X0 , X1 ](d) be the K-algebra homomorphism for which

φ(Y0 ) = X0d , φ(Y1 ) = X0d−1 X1 , . . . , φ(Yd−1 ) = X0 X1d−1 , φ(Yd ) = X1d .

Then Ker φ = IPd (K) (N(d) ).

(i) Calculate reg(N(d) ), and deduce that IPd (K) (N(d) ) can be generated by
quadratics.
(ii) Calculate the dimension of IPd (K) (N(d) )2 , the component of degree 2 of
IPd (K) (N(d) ), as a vector space over K.
(iii) Deduce that IPd (K) (N(d) ) can be generated by the 2 × 2 minors of the
matrix
$ %
Y0 Y1 Y2 . . . Yd−2 Yd−1
.
Y1 Y2 Y3 . . . Yd−1 Yd
16.3.6 Exercise. Use the notation of 15.2.5, but in the special case in which
d = 4. Thus K is an algebraically closed field, and A(4) is the subring

K[X 4 , X 3 Y, XY 3 , Y 4 ]

of the ring K[X, Y ] of polynomials over K in two indeterminates X and Y .


Also, ψ : K[Y0 , Y1 , Y2 , Y3 ] −→ A(4) is the K-algebra homomorphism for
which ψ(Y0 ) = X 4 , ψ(Y1 ) = X 3 Y , ψ(Y2 ) = XY 3 and ψ(Y3 ) = Y 4 , and
Ker ψ is the vanishing ideal of the projective variety Σ(4) ⊆ P3 (K).

(i) Calculate reg(Σ(4) ), and deduce that IP3 (K) (Σ(4) ) can be generated by
homogeneous polynomials of degrees not exceeding 3.
(ii) For i = 2, 3, calculate the dimension of IP3 (K) (Σ(4) )i , the component
of degree i of IP3 (K) (Σ(4) ), as a vector space over K. Find a quadratic
and three cubics which generate IP3 (K) (Σ(4) ).

The next result uses Theorem 16.3.1 to establish connections between the
regularity of a finitely generated graded module M over a polynomial ring
over a field and the syzygies of M : we hinted at these connections in the intro-
duction to this chapter. The result is presented by Eisenbud and Goto in [11,
p. 89], although they state that it is not new and that it has origins in ideas of
362 Castelnuovo regularity

Castelnuovo and Mumford. We refer the reader to [7, pp. 36–37] for funda-
mental facts concerning the minimal graded free resolution of M .
16.3.7 Theorem: syzygetic characterization of regularity. Let K be a field

and let R = j∈N0 Rj := K[X1 , . . . , Xn ], the ring of polynomials over K
in n indeterminates (where n ∈ N). Let M be a non-zero finitely generated
graded R-module having proj dim M = p, and let
fp f1
0 −→ Fp −→ Fp−1 −→ . . . −→ F1 −→ F0 −→ 0,
be the minimal graded free resolution of M . Thus there exist b0 , . . . , bp ∈ N
(j) (j)
and, for each j = 0, . . . , p, integers ai (i = 1, . . . , bj ) such that a1 ≥ · · · ≥
(j) b j (j)
abj and Fj = i=1 R(ai ). Then
 
(j)
reg(M ) = max −abj − j : j = 0, . . . , p .

Proof. The fi (i = 1, . . . , p) are homogeneous, and there is a homogeneous


R-epimorphism f0 : F0 −→ M such that Ker f0 = Im f1 .
In view of 16.2.11 and 16.2.9, we have
  
b 0 (0) (0) 
reg(F0 ) = reg i=1 R(ai ) = max reg R(ai ) : i = 0, . . . , b0
 
(0) (0)
= max −ai : i = 0, . . . , b0 = −ab0 .

We argue by induction on p. When p = 0, we have a homogeneous isomor-


phism M ∼ = F0 , and the claim follows from the above equations. So suppose,
inductively, that p > 0 and that the result has been proved for smaller values
of p. Set L := Ker f0 = Im f1 .
Now F0 , and therefore M , can be generated by homogeneous elements of
(0)
degrees not exceeding reg(F0 ) = −ab0 , by 16.3.1. By the minimality, M
cannot be generated by homogeneous elements all of whose degrees are less
(0) (0)
than −ab0 ; therefore reg(M ) ≥ −ab0 = reg(F0 ) by 16.3.1.

There is an exact sequence 0 −→ L −→ F0 −→ M −→ 0 in the category
*C(R). Application of 16.2.15(i) to this yields that
reg(L) ≤ max{reg(F0 ), reg(M ) + 1} = reg(M ) + 1,
so that reg(L) − 1 ≤ reg(M ); on the other hand, application of 16.2.15(iv) to
the same sequence yields that
(0)
reg(M ) ≤ max{reg1 (L) − 1, reg(F0 )} ≤ max{reg(L) − 1, −ab0 },
(0)
and so we see that reg(M ) = max{reg(L) − 1, −ab0 }.
fp f1
The exact sequence 0 −→ Fp −→ Fp−1 −→ . . . −→ F1 −→ L −→ 0
16.3 Degrees of generators 363

yields the minimal graded free resolution of L, so that proj dim L = p − 1.


Therefore, by the inductive hypothesis,
 
(j)
reg(L) = max −abj − (j − 1) : j = 1, . . . , p
 
(j)
= max −abj − j : j = 1, . . . , p + 1.

Therefore
 
(0)
reg(M ) = max reg(L) − 1, −ab0
   
(j) (0)
= max max −abj − j : j = 1, . . . , p , −ab0 ,

and the inductive step can be completed.


16.3.8 Example. Consider the situation of Theorem 16.3.7 above in the spe-
cial case in which M = R/b, where b is a proper graded ideal of R. Since
R/b can be generated by one homogeneous element of degree 0, and since
Ker fj ⊆ mFj for all j = 0, . . . , p, it follows that, in this case, b0 = 1,
(0) (j+1) (j) (j)
a1 = 0, and −a1 ≥ −a1 + 1 for all j = 0, . . . , p − 1. Hence −a1 ≥ j
for all j = 0, . . . , p.
(j)
On the other hand, by 16.3.7, we have −abj ≤ reg(M ) + j for all j =
(j)
0, . . . , p. Thus, for a given integer j such that 0 ≤ j ≤ p, the −ai (1 ≤ i ≤
bj ) are constrained to lie in the interval {r ∈ N0 : j ≤ r ≤ reg(M ) + j} of
length reg(M ).

16.3.9 Exercise. Assume that R = n∈N0 Rn is positively graded and
homogeneous, and let M be a non-zero finitely generated graded R-module.
Let a1 , . . . , ah be homogeneous elements of R with deg ai = ni > 0 for
i = 1, . . . , h, and suppose that a1 , . . . , ah is an M -sequence. Use Exercise
16.2.15 to show that
h h h
reg1 M/ i=1 ai M ≤ reg1 (M ) + i=1 ni − h ≤ reg M/ i=1 ai M ,

and that both inequalities in this display are equalities if gradeM (R+ ) > h.
17

Hilbert polynomials


Suppose that R = n∈N0 Rn is positively Z-graded

and homogeneous and
such that R0 is an Artinian ring, and let M = n∈Z Mn be a non-zero finitely
generated graded R-module of dimension d. By a classical theorem of Hilbert
(see [7, Theorem 4.1.3]), there is a polynomial PM ∈ Q[X] of degree d − 1,
and an n0 ∈ Z, such that PM (n) = R0 (Mn ) for all n ≥ n0 . The polynomial
PM is necessarily uniquely determined, and known as the Hilbert polynomial
of M . The classical ‘postulation’ problem asked for an explanation of the dif-
ference PM (n) − R0 (Mn ) for n < n0 . Serre solved this in [77] using sheaf
cohomology. In this chapter we present, via graded local cohomology, the ideas
behind Serre’s approach; the reader will find a translation into the language of
sheaf cohomology in 20.4.16.
We study, in §17.1 below, the so-called characteristic function χM : Z −→
Z of M and show that this is completely represented by a polynomial. In more
detail, it turns out that, for each i ∈ N0 and n ∈ Z, the n-th component
Ri DR+ (M )n of Ri DR+ (M ) has finite length as an R0 -module, and we de-
note this length by diM (n); it makes sense for us to define the characteristic
function χM : Z −→ Z of M by setting

χM (n) = (−1)i diM (n) for all n ∈ Z.


i∈N0

We shall see in 17.1.7 that there is a polynomial q ∈ Q[X] of degree d−1 such
that q(n) = χM (n) for all n ∈ Z. The fact that HR i
+
(M )n = 0 for all i ≥ 0
and n  0 (see 16.1.5) means that we can conclude that χM (n) = R0 (Mn )
for all n  0; this not only provides a proof of Hilbert’s Theorem mentioned
above (q turns out to be the Hilbert polynomial PM ), but also yields precise
information about the difference PM (n) − R0 (Mn ) for small n.
Also in §17.1, we use Hilbert’s Theorem and graded local duality to establish
the existence of ‘cohomological Hilbert polynomials’: for each i ∈ N0 , there
Hilbert polynomials 365

is a polynomial piM ∈ Q[X] of degree less than i such that


i
R0 (HR +
(M )n ) = piM (n) for all n  0.
In §17.2, we study the invariant reg2 (M ) (where M is as above). We show
1
that both DR+ (M ) and HR +
(M ) can be generated by homogeneous elements
of degrees not exceeding reg (M ) (provided reg2 (M ) is finite). This will en-
2

able us to prove (in 20.4.13) an important result on coherent sheaves due to


Serre [80]. This is a small hint about the significance of the invariant reg2 .
Furthermore, in the case where there is a homogeneous element a of R of de-
gree 1 that is a non-zerodivisor on M , we show that, as n increases beyond
reg2 (M/aM ) − 1, the integers R0 (HR 2
+
(M )n ) decrease strictly to 0 and then
remain at 0. This is a generalization of a result, actually formulated in terms of
sheaf cohomology, of Mumford [54, p. 99].
To appreciate fully the significance of the invariant reg2 , the reader will need
to become aware of the links between local cohomology and sheaf cohomol-
ogy, which will be treated in Chapter 20. In this chapter, we content ourselves
with the hints given above, and some elementary comments at the beginning
of §17.2 about the significance of reg2 in the framework of defining equations
of projective varieties.
In §17.3, we generalize an important result of D. Mumford [54, p. 101],
actually formulated in terms of sheaf cohomology, which shows that, for a
non-zero graded ideal b of the polynomial ring K[X1 , . . . , Xd ] over a field
K, the regularity reg2 (b) of b at and above level 2 is bounded in terms of the
Hilbert polynomial Pb of b. We generalize this in 17.3.6 and prove that, for a
polynomial ring R = R0 [X1 , . . . , Xd ] over an Artinian local ring R0 , for inte-
gers a1 , . . . , ar and a non-zero graded submodule M of the finitely generated
r
*free graded R-module i=1 R(ai ), the regularity reg2 (M ) is bounded in
terms of PM . The reader will find a sheaf-theoretic formulation of Mumford’s
Regularity Bound in 20.4.18.
In §17.4, we apply the main result of §17.3 to give upper bounds on the
invariants reg1 and reg0 in certain circumstances. We prove that the regular-
ity reg1 (M ) at and above level 1 is bounded in terms of PM and the number
and degrees of homogeneous generators of M . We also prove in the same sec-
tion that the regularity reg0 (M ) = reg(M ) is bounded in terms of PM , the
degrees of homogeneous generators of M and the so-called postulation num-
ber pstln(M ) of M , defined as the greatest integer n for which R0 (Mn ) =
PM (n) if any such integers exist, and −∞ otherwise.
Once again, throughout this chapter, all graded rings and modules are to
be understood to be Z-graded, and all polynomial rings R0 [X1 , . . . , Xd ] (over
a commutative Noetherian ring R0 ) are to be understood to be (positively)
366 Hilbert polynomials

Z-graded so that R0 is the component of degree 0 and deg Xi = 1 for all


i = 1, . . . , d.

17.1 The characteristic function


17.1.1 Reminders, Notation and Terminology. We shall use notation and
terminology employed by Bruns and Herzog in [7, §4.1]. Consider a function
f : Z −→ Z.

(i) We say that f is of polynomial type (of degree d) if and only if there
exists a polynomial p ∈ Q[X] (of degree d) such that f (n) = p(n) for
all n  0 (that is, for all n greater than some fixed constant integer n0 ).
(When this is the case, the polynomial p will be uniquely determined,
of course, since two rational polynomials which take the same values at
infinitely many integers must coincide. We adopt the convention that the
zero polynomial has degree −1.)
(ii) We define Δf : Z −→ Z by Δf (n) = f (n + 1) − f (n) for all n ∈ Z.
Thus Δ is a mapping from the set of functions from Z to Z to itself. For
n ∈ N, we use Δn to denote the result of repeating Δ n times; we also
write Δ0 f = f .
(iii) Let d ∈ N0 . Recall from [7, Lemma 4.1.2] that f is of polynomial type of
degree d if and only if there exists c ∈ Z with c = 0 such that Δd f (n) =
c for all n  0.
(iv) Recall that, for i ∈ N,
 
X +i (X + i)(X + i − 1) . . . (X + 1)
:= ∈ Q[X],
i 1.2.3. . . . .i
 
and that X+0 0 denotes the constant polynomial 1 ∈ Q[X].
(v) Let d ∈ N. Let P ∈ Q[X] be a non-zero polynomial of degree d − 1.
Recall from [7, Lemma 4.1.4] that P (n) ∈ Z for all n ∈ Z if and only if
there exist integers e0 , . . . , ed−1 such that
d−1  
i X +d−i−1
P (X) = (−1) ei .
i=0
d−i−1

When this is the case, we say that P is a numerical polynomial; also,


the integers e0 , . . . , ed−1 are uniquely determined by P , and we denote
them by e0 (P ), . . . , ed−1 (P ).
Also, given c := (c0 , . . . , cd−1 ) ∈ Zd , we shall denote by pc =
17.1 The characteristic function 367

p(c0 ,...,cd−1 ) the polynomial in Q[X] given by


d−1  
i X +d−i−1
pc (X) = (−1) ci .
i=0
d−i−1

Note that, when d > 1, we have


pc (X) − pc (X − 1)
d−1    
X +d−i−1 (X − 1) + d − i − 1
= i
(−1) ci −
i=0
d−i−1 d−i−1
d−2  
X +d−i−2
= (−1)i ci = p(c0 ,...,cd−2 ) (X).
i=0
d−i−2

(vi) We shall say that f is of reverse polynomial type (of degree d) if and only
if there exists a polynomial p ∈ Q[X] (of degree d) such that f (n) =
p(n) for all n  0 (that is, for all n less than some fixed constant integer
n0 ). Thus f is of reverse polynomial type of degree d if and only if the
function g : Z −→ Z defined by g(n) = f (−n) for all n ∈ Z is of
polynomial type of degree d in the sense of (i).
17.1.2 Exercise. Let 0 = P ∈ Q[X] have deg P = d − 1. Assume that
P takes integer values at d consecutive integers. Show that P (n) ∈ Z for all
n ∈ Z.
17.1.3 Exercise. Suppose that the function f : Z −→ Z is such that there
exists Q ∈ Q[X] for which Δf (n) = Q(n) for all n ∈ Z. By using, for
various values of the integer n0 , the fact that
n−1
f (n) = f (n0 ) + i=n0 Q(i) for all n ≥ n0 ,
show that there exists P ∈ Q[X] such that f (n) = P (n) for all n ∈ Z.

17.1.4 Further Notation and Terminology. Assume that R = n∈N0 Rn
is positively graded and homogeneous and such that R0 is an Artinian ring, and

let M = n∈Z Mn be a finitely generated graded R-module. It follows from
16.1.5(i) that the R0 -module HRi
+
(M )n has finite length, for all i ∈ N0 and all
n ∈ Z: we denote this length by hiM (n). Similarly, it follows from 16.1.6(ii)
that the R0 -module DR+ (M )n has finite length, for all n ∈ Z. Let i ∈ N.
For n ∈ Z, we denote the n-th component of Ri DR+ (M ) by Ri DR+ (M )n ;
i+1
by 13.5.7(iii), this R0 -module is isomorphic to HR +
(M )n , and so has finite
length.
We denote R0 (Ri DR+ (M )n ) by diM (n), for all i ∈ N0 and all n ∈ Z. Note
M (n) for i > 0, and that this is zero for i ≥ max {1, dim M }.
that diM (n) = hi+1
368 Hilbert polynomials

It therefore makes sense for us to define the characteristic function χM :



Z −→ Z of M by setting χM (n) = i∈N0 (−1)i diM (n) for all n ∈ Z. Note
that, for all n ∈ Z, we have d0M (n) = R0 (Mn ) + h1M (n) − h0M (n) by 13.5.4,
so that, for any integer d ≥ dim M ,
d d
χM (n) = (−1)i diM (n) = d0M (n) − (−1)i hiM (n)
i=0 i=2
d
= R0 (Mn ) − (−1)i hiM (n).
i=0

Thus it follows from 16.1.5(ii) that χM (n) = R0 (Mn ) ≥ 0 for all n  0.
Note also that, by 14.1.10(ii), we have χM (t) (n) = χM (t + n) for all t, n ∈ Z.

17.1.5 Exercise. Assume that R = n∈N0 Rn is positively graded and
homogeneous and such that R0 is an Artinian ring. Let

0 −→ L −→ M −→ N −→ 0

be an exact sequence of finitely generated graded R-modules and homoge-


neous homomorphisms. Show that, for each n ∈ Z, we have

χM (n) = χL (n) + χN (n).



17.1.6 Remark. Assume that R = n∈N0 Rn is positively graded and ho-
(1) (t)
mogeneous, and such that R0 is Artinian. Let m0 , . . . , m0 be the maxi-
mal ideals of R0 . Let M be a finitely generated graded R-module. For each
j = 1, . . . , t, set R(j) := R ⊗R0 (R0 )m(j) and M (j) := M ⊗R0 (R0 )m(j) .
0 0

(i) For each R0 -module N , the natural R0 -homomorphism ωN : N −→


t
j=1 Nm(j) (for which ωN (y) = (y/1, . . . , y/1) for all y ∈ N ) is an
0
isomorphism; therefore, when N is finitelygenerated, its length R0 (N )
t
satisfies R0 (N ) = j=1 (R0 ) (j) Nm(j) .
m0 0

(1) (t)
(ii) Since m 0 + R + , . . . , m0 + R + are the only *maximal graded ideals of
R and, for j = 1, . . . , t, the ring R(j) is isomorphic, as a graded ring, to
the homogeneous localization (see 14.1.1) R (j)  , it follows that
m0 +R+
 
dim M = max dimR(j) M (j) : j = 1, . . . , t .

(iii) Fix an integer j between 1 and t. Note that, by 16.2.3, the ring R(j) is
positively graded and homogeneous, and has 0-th component isomorphic
17.1 The characteristic function 369
(j)
to the Artinian local ring (R0 )m(j) ; moreover, R+ R(j) = R+ and there
0
is an (R0 )m(j) -isomorphism
0

i
HR (M )n ∼ i
= HR (j) (M
(j)
)n for each i ∈ N0 and each n ∈ Z.
+ (j)
m0 +

Also, by 16.2.2(vi), there is an (R0 )m(j) -isomorphism


0
 

DR+ (M )n m(j) = DR(j) (M (j) )n for each n ∈ Z.
0 +

(iv) It follows from parts (i) and (iii) that


 
regl (M ) = max regl (M (j) ) : j = 1, . . . , t for all l ∈ N0 ,
t
that hiM (n) = j=1 hiM (j) (n) for all i ∈ N0 and n ∈ Z, and that (with
the notation of 17.1.4)
t t
d0M (n) = d0M (j) (n) and χM (n) = χM (j) (n) for all n ∈ Z.
j=1 j=1

17.1.7 Theorem. Assume that R = n∈N0 Rn is positively graded and ho-
mogeneous, and such that R0 is Artinian; let M be a non-zero finitely gen-
erated graded R-module. Then there is a (necessarily uniquely determined)
polynomial PM ∈ Q[X] of degree dim M − 1 such that PM (n) = χM (n) for
all n ∈ Z.
Proof. It follows from 17.1.6 (and the fact that χM (n) ≥ 0 for all n  0)
that it is sufficient for us to prove this result under the additional assumption
that the Artinian ring R0 is local. We make this assumption in what follows,
and we let m0 be the maximal ideal of R0 .
We now use 16.2.4 and 16.2.2(vi) to reduce to the case where the residue
field of R0 is infinite. We therefore assume that R0 /m0 is infinite for the re-
mainder of this proof.
We now use induction on dim M . Consider first the case when dim M = 0.
Then Ass M = {m0 ⊕ R+ }, so that there exists t ∈ N such that (R+ )t M = 0.
In this case, M = ΓR+ (M ) and so HR i
+
(M ) = 0 for all i ∈ N, by 2.1.7(ii),
and DR+ (M ) = 0 by 2.2.10(i). Thus χM (n) = 0 for all n ∈ Z, and our claim
is proved in this case.
Now suppose that dim M > 0 and our desired result has been proved for
all non-zero, finitely generated graded R-modules of smaller dimension. Since
ΓR+ (M ) is a graded submodule of M , it follows from 2.1.7(iii) that there is a
(homogeneous) isomorphism
i
HR +
(M ) ∼ i
= HR +
(M/ΓR+ (M )) for each i ∈ N,
370 Hilbert polynomials

and from 13.3.14 and 2.2.10(ii) that there is a (homogeneous) isomorphism


DR+ (M ) ∼ = DR+ (M/ΓR+ (M )). Hence χM = χM/ΓR+ (M ) . Since ΓR+ (M )
is annihilated by some power of R+ , it follows that

Supp(ΓR+ (M )) ⊆ {m0 ⊕ R+ } ,

so that dim M = dim(M/ΓR+ (M )). Hence, for the purpose of this induc-
tive step, we can replace M by M/ΓR+ (M ) and so (in view of 2.1.2, 2.1.1
and 16.1.4(ii)) assume that M is R+ -torsion-free and that R1 contains a non-
zerodivisor a on M . This we do.
Application of 17.1.5 to the exact sequence
a
0 −→ M (−1) −→ M −→ M/aM −→ 0

of graded R-modules and homogeneous homomorphisms, together with the


observation that χM (−1) (n) = χM (n − 1) for all n ∈ Z, yields that

χM (n) − χM (n − 1) = χM/aM (n) for all n ∈ Z.

Now dim M/aM = dim M −1 (by 14.1.14), and so it follows from the induc-
tive hypothesis that there is a polynomial PM/aM ∈ Q[X] of degree dim M −2
such that PM/aM (n) = χM/aM (n) for all n ∈ Z. Hence, by 17.1.3, there ex-
ists a polynomial PM ∈ Q[X] such that PM (n) = χM (n) for all n ∈ Z.
Furthermore, if PM/aM = 0, then PM must have degree dim M − 1. How-
ever, if PM/aM = 0, then the above argument shows only that deg PM ≤ 0,
and we must show that χM (n) = 0 for some n ∈ Z in order to complete the
inductive step.
The assumption that PM/aM = 0, in conjunction with the inductive hypoth-
i
esis, implies that dim M = 1. Hence HR +
(M ) = 0 for all i > 1, and so
χM (n) = dM (n) for all n ∈ Z. However, the fact that ΓR+ (M ) = 0 ensures
0

that the map ηM : M −→ DR+ (M ) of 2.2.6(i) is a monomorphism, so that


DR+ (M ) = 0 and d0M (n) = 0 for some n ∈ Z.

17.1.8 Remark. Assume that R = n∈N0 Rn is positively graded and ho-

mogeneous, and such that R0 is Artinian; let M = n∈Z Mn be a non-zero
finitely generated graded R-module of dimension d.
It was pointed out in 17.1.4 that χM (n) = R0 (Mn ) for all n  0. Thus
Hilbert’s Theorem (see [7, Theorem 4.1.3]), that the function

R0 (M( • ) ) : Z −→ Z

is of polynomial type of degree d − 1, is a corollary of Theorem 17.1.7, and


the polynomial PM of that theorem is just the Hilbert polynomial of M of [7,
4.1.5].
17.1 The characteristic function 371

With the notation of 17.1.1(v), for each j = 0, . . . , d − 1, we set ej (M ) :=


ej (PM ) and refer to this as the j-th Hilbert coefficient of M . Note that, in the
terminology of [7, 4.1.5], the integer e0 (M ) is the multiplicity of M if d > 0.
To sum up, we can write
d−1  
X +d−i−1
PM (X) = (−1)i ei (M )
i=0
d−i−1

and
d−1  
n+d−i−1
χM (n) = i
(−1) ei (M ) for all n ∈ Z.
i=0
d−i−1

17.1.9 Exercise. Assume that R = n∈N0 Rn is positively graded and ho-
mogeneous, and such that R0 is Artinian; let M be a non-zero finitely gen-
erated graded R-module of dimension d. Suppose that a ∈ R1 is a non-
zerodivisor on M . Show that

(i) PM/aM (X) = PM (X) − PM (X − 1); and


(ii) if d > 1, then, with the notation of 17.1.1(v),

PM/aM (X) = p(e0 (M ),...,ed−2 (M )) (X),

and ei (M/aM ) = ei (M ) for all i = 0, . . . , d − 2.



17.1.10 Definition and Exercise. Assume that R = n∈N0 Rn is pos-
itively graded and homogeneous, and such that R0 is Artinian; let M be a
finitely generated graded R-module. The postulation number pstln(M ) of M
is defined by pstln(M ) = sup{n ∈ Z : R0 (Mn ) = PM (n)}. Thus pstln(M )
is an integer or −∞. Show that

(i) pstln(M ) ≤ reg(M ) ≤ max{reg1 (M ), pstln(M )}, and


(ii) if dim M ≤ 0, then pstln(M ) = reg(M ) = end(M ).

17.1.11 Theorem and Definition. Assume R = n∈N0 Rn is positively
graded and homogeneous, and such that R0 is Artinian; let M be a finitely
generated graded R-module. Let i ∈ N0 . Then the function hiM : Z −→ N0 is
of reverse polynomial type of degree less than i (in the sense of 17.1.1(vi)); in
other words, there is a polynomial piM ∈ Q[X] of degree less than i such that
i
R0 (HR +
(M )n ) = piM (n) for all n  0.

The function hiM is called the i-th cohomological Hilbert function of M ,


while the (uniquely determined) polynomial piM is called the i-th cohomologi-
cal Hilbert polynomial of M .
372 Hilbert polynomials

Proof. By 17.1.6, it is sufficient for us to prove the result under the additional
hypothesis that R0 is local. We therefore assume for the remainder of this
proof that R0 is an Artinian local
 ring, with maximal ideal m0 , say. With this
assumption, R is *local, and R+ = m0 + R+ =: m, the unique *maximal
ideal of R.
Since (R0 , m0 ) is a complete local ring, we can use Cohen’s Structure The-
orem for such rings (see [50, Theorem 29.4(ii)], for example) in conjunction
with standard facts about the structure of positively graded homogeneous com-
mutative Noetherian rings (see [7, Proposition 1.5.4]) to see that there is a
Gorenstein graded *local commutative Noetherian ring R and a surjective ho-
mogeneous ring homomorphism f : R −→ R. Let d := dim R .
We are going to use the Graded Local Duality Theorem 14.4.1. Let *D
denote the functor * HomR ( • , *E(R/m)) from *C(R) to itself. By 14.4.1,
there exists a ∈ Z such that there is a natural transformation

d −i
ψ i : HR
i
+
−→ *D(* ExtR  ( • , R (a )))
i
of covariant functors from *C(R) to *C(R) which is such that ψM  is a (nec-

essarily homogeneous) isomorphism for all i ∈ N0 whenever M  is a finitely


generated graded R-module.
d −i
Let N = * ExtR  (M, R (a )), a finitely generated graded R-module. We
now use Graded Matlis Duality 14.4.2. We use the notation ( • )∨ of 14.4.2 to
denote the functor * HomR0 ( • , ER0 (R0 /m0 )) from *C(R) to itself. It follows
from 14.4.2 that, for each n ∈ Z (and with an obvious notation), there are
R0 -isomorphisms
i
HR +
(M )n ∼
= (*D(N ))n ∼
= (N ∨ )n ∼
= (* HomR0 (N, ER0 (R0 /m0 )))n
= HomR0 (N−n , ER0 (R0 /m0 )),

so that hiM (n) = R0 (HomR0 (N−n , ER0 (R0 /m0 ))) = R0 (N−n ) by 10.2.13.
Thus, in order to prove that hiM is of reverse polynomial type of degree less
than i, it is now sufficient for us to prove that the function R0 (N( • ) ) : Z −→
N0 is of polynomial type of degree less than i; hence, by Hilbert’s Theorem
(see 17.1.8 above and [7, Theorem 4.1.3]), it is enough for us to prove that
dim N ≤ i.
d −i
To do this, let p ∈ SuppR N . Then ExtR 

(Mp , Rp  ) = 0, so that, since
p
inj dimR  Rp  = ht p , we must have ht p ≥ d −i. Therefore dim R /p ≤ i,
p
so that dimR N ≤ i.

17.1.12 Exercise. Let the notation be as in 15.2.7 and 15.2.8.

(i) Compute the Hilbert polynomial, all Hilbert coefficients, the postula-
17.2 The significance of reg2 373

tion number, all cohomological Hilbert functions and the Castelnuovo–


Mumford regularity for each of the three rings A, B and C.
(ii) Conclude that the vanishing ideal IP6 (K) (VA ) ⊂ K[X0 , . . . , X6 ] of VA
can be generated by homogeneous polynomials of degree 2, and that the
vanishing ideals IP5 (K) (VB ), IP5 (K) (VC ) ⊂ K[X0 , . . . , X5 ] of VB and
VC can be generated by homogeneous polynomials of degrees 2 and 3.
(You might find 16.3.2 and 16.3.3 helpful.)

17.1.13 Exercise. Let the situation be as in 17.1.11, and let M be a non-


zero finitely generated graded R-module of dimension d. Show that the d-th
cohomological Hilbert polynomial pdM of M has degree exactly d − 1.

17.1.14 Exercise. Assume that R = n∈N0 Rn is positively graded and ho-
mogeneous, and such that R0 is Artinian; let M be a non-zero finitely gener-
ated graded R-module of positive dimension. This exercise involves the finite-
ness dimension fR+ (M ) of M relative to R+ of 9.1.3, and, for r ∈ N and
s ∈ Z, the notation ( • )(r,s) : *C(R) −→ *C(R(r) ) of 13.5.9 for the (r, s)-th
Veronesean functor.

(i) Show that fR+ (M ) is finite.


(ii) Show that fR+ (M ) = max gradeM (r,s) ((R(r) )+ ) : r ∈ N, s ∈ Z .

17.2 The significance of reg2


In this section, we shall show that, if R is positively graded and homoge-
neous, M is a finitely generated graded R-module, and t ∈ Z is such that
t ≥ reg2 (M ), then DR+ (M ) and HR 1
+
(M ) can be generated by homoge-
neous elements of degrees not exceeding t. We shall also start to prepare the
ground for a result (to be presented in the next section) that bounds reg2 (M )
in certain circumstances.
As promised in the introduction to this chapter, we now give another hint
about the significance of the invariant reg2 . Let K be an algebraically closed
field, let r ∈ N and let V ⊂ Pr (K) be a projective variety. Let p denote
the vanishing ideal IPr (K) (V ) of V , so that p is the (non-zero) graded prime
ideal IAr+1 (K) (Cone(V )) of K[X0 , . . . , Xr ] = O(Ar+1 (K)). In 16.3.3, we
introduced the Castelnuovo–Mumford regularity reg(V ) of V , and pointed out
that reg(V ) = reg(p) = reg1 (p) = reg2 (p) and that p can be generated by
homogeneous polynomials of degrees not exceeding reg(V ). Since p can be
used to define V , in the sense that V = VPr (K) (p), the regularity reg2 (p) of p
at and above level 2 provides an upper bound on the degrees of homogeneous
374 Hilbert polynomials

polynomials needed to define V . This is another reason why one could be


interested in upper bounds on reg2 (p).

17.2.1 Proposition. Assume that R = n∈N0 Rn is positively graded and

homogeneous, and let M = n∈Z M n be a finitely generated graded R-
module.
If reg2 (M ) = −∞, then each of DR+ (M ) and HR 1
+
(M ) can be generated
by homogeneous elements of degrees not exceeding reg2 (M ).
If reg2 (M ) = −∞, then, for each t ∈ Z, both DR+ (M ) and HR 1
+
(M ) can
be generated by homogeneous elements of degrees not exceeding t.

Proof. Since reg2 (M ) = reg2 (M/ΓR+ (M )) (by 16.2.10(ii)) and there are
homogeneous isomorphisms

= ∼
=
DR+ (M ) −→ DR+ (M/ΓR+ (M )) and HR
1
+
(M ) −→ HR
1
+
(M/ΓR+ (M ))

(by 13.3.14, 2.2.10(ii) and 2.1.7(iii)), we can replace M by M/ΓR+ (M ) and


so (in view of 2.1.2) assume that M is R+ -torsion-free. We shall make this
simplification.
Since M is finitely generated, there exists h ∈ Z such that Mn = 0 for all
n < h. If reg2 (M ) = −∞, choose t ∈ Z such that t ≤ min h, reg2 (M ) ;
if reg2 (M ) = −∞, let t be any integer such that t ≤ h. With the nota-
tion of 16.1.1, let N := DR+ (M )≥t . Note that N is a finitely generated
R-module, by 16.1.6(i). Also, by choice of t, the homogeneous monomor-
phism ηM : M −→ DR+ (M ) satisfies Im ηM ⊆ N ; thus N/ Im ηM ⊆
DR+ (M )/ Im ηM ∼ = HR 1
+
(M ), and so N/ Im ηM is finitely generated and
R+ -torsion. It therefore follows from 2.1.7(i) that there is a (homogeneous)

=
isomorphism HR i
+
(M ) −→ HR i
+
(N ) for each i > 1, so that reg2 (M ) =
2
reg (N ).

Let C = n∈Z Cn be the graded R-module DR+ (M )/N . Since
i
HR +
(DR+ (M )) = 0 for i = 0, 1

by 2.2.10(iv), it follows from the exact sequence

0 −→ N −→ DR+ (M ) −→ C −→ 0
0
(in the category *C(R)) that HR +
(N ) = 0 and there is a homogeneous iso-

=
morphism ΓR+ (C) −→ HR 1
+
(N ). Now ΓR+ (C) is a graded submodule of C,
and Cn = 0 for all n ≥ t. Therefore end(HR 1
+
(N )) + 1 ≤ t − 1 + 1 = t. In
the case where reg2 (M ) = −∞, it follows that

reg(N ) = reg2 (N ) = reg2 (M ) ≥ t,


17.2 The significance of reg2 375

whereas, in the case where reg2 (M ) = −∞, we have reg(N ) ≤ t.


It follows from 16.3.1 that N = DR+ (M )≥t , if non-zero, can be gener-
ated by homogeneous elements of degrees not exceeding max{t, reg2 (M )},
1
so that, since HR +
(M ) is a homomorphic image of DR+ (M ) by a homoge-
neous homomorphism, all the claims are proved.
17.2.2 Exercise. Let the notation and hypotheses be as in 17.2.1. Let t ∈ Z
with t ≥ reg2 (M ). Let p ∈ Proj(R) := * Spec(R) \ Var(R+ ). Recall that
• (p) denotes the homogeneous localization functor with respect to p.
Show that the natural homomorphism η := ηM : M −→ DR+ (M ) induces

=
a homogeneous isomorphism η(p) : M(p) −→ DR+ (M )(p) of graded R(p) -
modules.
Let φ(p) : DR+ (M ) −→ DR+ (M )(p) be the natural homogeneous ho-
momorphism of graded R-modules, and consider the composition β(p) =
−1
η(p) ◦ φ(p) : DR+ (M ) −→ M(p) . As usual, the t-th component of β(p) will be
denoted by (β(p) )t . Let S be a generating set for the R0 -module DR+ (M )t .
Use 17.2.1 and the fact that R is homogeneous to show that

M(p) = m∈S R(p) β(p) (m)

and (M (t)(p) )0 = (M(p) )t = m∈S (R(p) )0 (β(p) )t (m).
Note. Exercise 17.2.2 will be used in an application to sheaf cohomology in
Theorem 20.4.13; the aim of that theorem is to establish a (generalization of a)
result of Serre that certain sheaves are generated by their global sections.

17.2.3 Proposition. Assume that R = n∈N0 Rn is positively graded and
homogeneous, and let M be a finitely generated graded R-module. Suppose
that a ∈ R1 is a non-zerodivisor on M . Then
(i) for all integers m ≥ reg2 (M/aM ) − 1, the multiplication map
a
2
HR +
(M )m−1 −→ HR
2
+
(M )m
is surjective; and
(ii) for all integers m ≥ reg2 (M/aM ) such that HR
2
+
(M )m−1 = 0, the
a
2
multiplication map HR +
(M )m−1 −→ HR
2
+
(M )m is not injective.
Proof. (i) By 14.1.10(ii), the exact sequence
a
0 −→ M −→ M (1) −→ (M/aM )(1) −→ 0
of graded R-modules and homogeneous homomorphisms induces an exact se-
a
quence HR 2
+
(M )m−1 −→ HR2
+
(M )m −→ HR
2
+
(M/aM )m of R0 -modules,
for all m ∈ Z, and HR+ (M/aM )m = 0 when m + 1 ≥ reg2 (M/aM ).
2
376 Hilbert polynomials

(ii) For each m ∈ Z, let

β m : R 1 ⊗ R 0 HR
1
+
(M/aM )m −→ HR
1
+
(M/aM )m+1

be the R0 -homomorphism for which βm (r1 ⊗ z) = r1 z for all r1 ∈ R1 and


z ∈ HR 1
+
(M/aM )m , and let γm : R1 ⊗R0 HR 1
+
(M )m −→ HR 1
+
(M )m+1 be
defined similarly.
Let m0 ∈ Z be such that HR 1
+
(M/aM ) can be generated by homogeneous
elements of degrees not exceeding m0 : by Proposition 17.2.1, we can take
m0 = reg2 (M/aM ) if reg2 (M/aM ) = −∞, and we can take m0 to be an
arbitrary integer if reg2 (M/aM ) = −∞. Since R is homogeneous, it fol-
lows that, for all m ≥ m0 , each element y ∈ HR 1
(M/aM )m+1 can be ex-
t +

pressed in the form i=1 ai zi for suitable a1 , . . . , at ∈ R1 and z1 , . . . , zt ∈


HR 1
+
(M/aM )m . In other words, βm is surjective for all m ≥ m0 .
Let π : M −→ M/aM be the canonical epimorphism, and, for each m ∈ Z,
let αm : HR 1
+
(M )m −→ HR 1
+
(M/aM )m be the m-th component of the
1
homogeneous homomorphism HR +
(π). Observe that the diagram
γm
R1 ⊗R0 HR
1
(M )m - HR
1
(M )m+1
+ +

IdR1 ⊗αm αm+1

? βm
?
R1 ⊗R0 HR
1
(M/aM )m - HR
1
(M/aM )m+1
+ +

commutes. It follows that, if, for some m ≥ m0 , we know that αm is surjective,


then αm+1 is surjective too.
a
Consider an m ≥ m0 and suppose that HR 2
+
(M )m−1 −→ HR 2
+
(M )m is
injective. By 14.1.10(ii), the exact sequence
a
0 −→ M (−1) −→ M −→ M/aM −→ 0

induces an exact sequence


α a
1
HR +
(M )n −→
n 1
HR +
(M/aM )n −→ HR
2
+
(M )n−1 −→ HR
2
+
(M )n

of R0 -modules for all n ∈ Z. This exact sequence (in the case where n = m)
shows that αm is surjective, and so it follows, as explained in the immedi-
ately preceding paragraph, that αm+1 is surjective and αn is surjective for
a
all n ≥ m. It follows from the above exact sequence that HR 2
+
(M )n−1 −→
HR+ (M )n is injective for all n ≥ m. It therefore follows from 16.1.5(ii) that
2
2
HR +
(M )m−1 = 0.
17.2 The significance of reg2 377

We point out that, in the situation and with the notation of 17.2.3, and with
the additional assumption that R0 is Artinian, the proposition shows that, as n
increases beyond reg2 (M/aM ) − 1, the integers h2M (n) decrease strictly to 0
and then remain at 0.
The following corollary is a consequence of Propositions 17.2.1 and 17.2.3,
and it provides the key for some crucial arguments later in this chapter.

17.2.4 Corollary. Assume that R = n∈N0 Rn is positively graded and
homogeneous, and that R0 is Artinian; let M be a finitely generated graded
R-module. Suppose that a ∈ R1 is a non-zerodivisor on M .
Let n0 , q ∈ Z with q ≥ max n0 , reg2 (M/aM ) − 1 ; for each n ≥ n0 , set
n
s(n) := h2M (n0 ) + h2M/aM (m).
m=n0 +1

Then
(i) h2M (n) ≤ s(n) for all n ≥ n0 ;
(ii) s(n) = s(q) for all n ≥ q, and s(q) = s(q − 1) if q > n0 ; and
(iii) h2M (n) ≤ max {0, s(q) + q − n} for all n > q.
Proof. (i) By 14.1.10(ii), the exact sequence
a
0 −→ M −→ M (1) −→ (M/aM )(1) −→ 0
of graded R-modules and homogeneous homomorphisms induces, for each
m ∈ Z, an exact sequence
a
2
HR +
(M )m−1 −→ HR
2
+
(M )m −→ HR
2
+
(M/aM )m
of R0 -modules, from which we deduce that
h2M (m) = R0 (HR
2
+
(M )m ) ≤ R0 (HR
2
+
2
(M )m−1 ) + R0 (HR +
(M/aM )m )
= h2M (m − 1) + h2M/aM (m),
and the claim follows easily from this.
(ii) For m > q ≥ reg2 (M/aM ) − 1, we have
q ≥ end(HR
2
+
(M/aM )) + 1,
2 2
so that HR +
(M/aM )m = HR +
(M/aM )q = 0. Hence

h2M/aM (m) = h2M/aM (q) = 0.


Therefore s(n) = s(q) for all n ≥ q, and s(q) = s(q − 1) if q > n0 .
(iii) For m ≥ q + 1 ≥ reg2 (M/aM ), we have, by Proposition 17.2.3,
a
that the multiplication map HR2
+
(M )m−1 −→ HR 2
+
(M )m is surjective, and
378 Hilbert polynomials
2
is not injective unless HR +
(M )m−1 = 0; therefore, if h2M (m − 1) = 0, then
h2M (m) ≤ h2M (m − 1) − 1. Thus h2M (m) ≤ max 0, h2M (m − 1) − 1 for all
m ≥ q + 1. Repeated use of this then shows that

h2M (n) ≤ max 0, h2M (q) − (n − q) for all n ≥ q + 1,

and the claim follows from part (i).

17.3 Bounds on reg2 in terms of Hilbert coefficients



17.3.1 Definition. Assume that R = n∈N0 Rn is graded. Let D be a sub-
class of the class of all graded R-modules. By a numerical invariant for R-
modules in D (or on D) we mean an assignment μ which, to each R-module
M in D, assigns μ(M ) ∈ Z ∪ {−∞}, and which is such that μ(M ) = μ(N )
whenever M and N are modules in D for which there is a homogeneous iso-
morphism M ∼ = N . We say that such a numerical invariant μ is finite if and
only if μ takes only finite values.
Let μ1 , . . . , μs , ρ be numerical invariants for R-modules in D, such that
μ1 , . . . , μs are finite. We say that μ1 , . . . , μs form a bounding system for ρ
(for R-modules in D) (or on D) if and only if there is a function B : Zs −→
Z such that ρ(M ) ≤ B(μ1 (M ), . . . , μs (M )) for each R-module M in D.
Furthermore, we say that μ1 , . . . , μs form a minimal bounding system for ρ
(for R-modules in D) precisely when they form a bounding system for ρ on D
but no s − 1 of μ1 , . . . , μs form a bounding system for ρ on D.

The main result of this section is a generalization of a ring-theoretic formu-


lation of a classical result of D. Mumford on sheaves of ideals on projective
spaces. A consequence is that, over the polynomial ring S = K[X0 , . . . , Xr ]
where K is an algebraically closed field, e0 ( • ), . . . , er ( • ) form a bounding
system for reg2 for non-zero graded ideals of S.
The following exercise shows that, in general, for an Artinian local ring R0 ,
the Hilbert coefficients do not form a bounding system for reg2 on the class of
all finitely generated graded R0 [X1 , . . . , Xn ]-modules.

17.3.2 Exercise. Suppose that R = R0 [X1 , X2 ] is a polynomial ring in 2


indeterminates X1 , X2 over an Artinian ring R0 .
For each t ∈ N0 , set M (t) := R(t) ⊕ R(−t), and calculate e0 (M (t) ),
e1 (M (t) ), reg2 (M (t) ). Conclude that the numerical invariants e0 ( • ), e1 ( • )
do not form a bounding system for reg2 on the class of all finitely generated
graded R-modules.
17.3 Bounds on reg2 in terms of Hilbert coefficients 379

17.3.3 Notation and Remark. For (n, i) ∈ Z × N0 , we set


⎧ 
 + ⎨ n
n if n ≥ i,
:= i
i ⎩
0 if n < i.
This notation is helpful in the following situation. Suppose that

R = R0 [X1 , . . . , Xd ]

is a polynomial ring in d (> 0) indeterminates X1 , . . . , Xd over an Artinian


ring R0 , and let a ∈ Z. Then, for all n ∈ Z, the n-th component R(a)n of the
graded R-module R(a) has length as an R0 -module given by
 +
n+a+d−1
R0 (R(a)n ) = R0 (R0 ) .
d−1
17.3.4 Lemma. Suppose that R = R0 [X1 , . . . , Xd ] is a polynomial ring in
d > 1 indeterminates X1 , . . . , Xd over an Artinian ring R0 . Let r ∈ N and let
r
a1 , . . . , ar ∈ Z. Let M be a graded submodule of i=1 R(ai ), and let f be
an integer such that f ≥ reg2 (M/Xd M ). Then
r  +
f + ai + d − 3
reg2 (M ) ≤ R0 (R0 ) − χM (f − 2) + f.
i=1
d−1

Proof. Note that Xd is a non-zerodivisor on M , because M is a submodule


r
of G := i=1 R(ai ). Let i, n ∈ Z be such that i > 2 and n > f − i.
Since f ≥ reg2 (M/Xd M ) ≥ end HR i−1
+
(M/Xd M ) + i − 1, it follows that
i−1
HR+ (M/Xd M )n+1 = 0. Therefore, the exact sequence
X
0 −→ M (−1) −→
d
M −→ M/Xd M −→ 0

of graded R-modules and homogeneous homomorphisms induces an exact


sequence
X
0 −→ HR
i
+
(M )n −→
d i
HR +
(M )n+1
i
of R0 -modules. Hence HR +
(M )n = 0 for all i > 2 and n > f − i. Thus, by
Definition 17.1.4,

χM (f − 2) = d0M (f − 2) − h2M (f − 2),

so that h2M (f − 2) = d0M (f − 2) − χM (f − 2).


We are now going to use Corollary 17.2.4 with n0 = f − 2 and q = f − 1.
Observe that, with the notation of that result, h2M/Xd M (m) = 0 for all m >
n0 = f − 2, so that s(n) = h2M (f − 2) for all n ≥ n0 . It follows from
380 Hilbert polynomials

17.2.4 that h2M (n) ≤ max 0, h2M (f − 2) + f − 1 − n for all n ≥ f − 1.


Hence h2M (n) = 0 for all n > h2M (f − 2) + f − 2. As we know already that
hiM (n) = 0 for all i > 2 and n > f − i, it follows that

reg2 (M ) ≤ h2M (f − 2) + f = d0M (f − 2) − χM (f − 2) + f.

As gradeG R+ = d > 1, it follows from 6.2.7, 2.2.6(i)(c) and 13.3.14 that


ηR+ : G −→ DR+ (G) is a homogeneous isomorphism. It therefore follows
from the fact that the functor DR+ is left exact that there is a homogeneous
monomorphism DR+ (M ) −→ G; we thus deduce, on use of 17.3.3, that (with
an obvious notation)
r  +
f − 2 + ai + d − 1
d0M (f − 2) ≤ R0 (Gf −2 ) = R0 (R0 ) .
i=1
d−1

The claim now follows.

Our next theorem is the main result of this chapter. We need one preliminary
lemma.

17.3.5 Lemma. Suppose that R = R0 [X1 , . . . , Xd ] is a polynomial ring in


d (> 0) indeterminates X1 , . . . , Xd over an Artinian local ring (R0 , m0 ).
Let M be a non-zero finitely generated graded R-module, and suppose
that there exists a ∈ R1 which is a non-zerodivisor on M . Then there exist
X1 , . . . , Xd ∈ R1 such that Xd is a non-zerodivisor on M , the family (Xi )di=1
is algebraically independent over R0 , and R = R0 [X1 , . . . , Xd ].

Proof. We have a = b1 X1 + · · · + bd Xd for some b1 , . . . , bd ∈ R0 . Now m0


is nilpotent, and so there exists j ∈ {1, . . . , d} such that bj ∈ m0 . Hence bj is
a unit of R0 , and so, after multiplication by b−1 j , we can, and do, assume that
bj = 1. Let c := a − Xj ∈ R1 .
There is a homogeneous R0 -algebra homomorphism φ : R −→ R for which
φ(Xj ) = Xj + c = a and φ(Xi ) = Xi for all i = 1, . . . , d with i = j.
Similarly, there is a homogeneous R0 -algebra homomorphism ψ : R −→ R
for which ψ(Xj ) = Xj − c and ψ(Xi ) = Xi for all i = 1, . . . , d with i = j.
Since φ(c) = ψ(c) = c, it follows that φ and ψ are inverse isomorphisms. If
we now take Xj = a = Xj + c and Xi = Xi for all i = 1, . . . , d with i = j,
then (Xi )di=1 is algebraically independent over R0 and R = R0 [X1 , . . . , Xd ];
since Xj is a non-zerodivisor on M , we can now reorder the Xi (if necessary)
to complete the proof.

17.3.6 Theorem. Suppose R = R0 [X1 , . . . , Xd ] is a polynomial ring in d (>


0) indeterminates X1 , . . . , Xd over an Artinian local ring (R0 , m0 ).
17.3 Bounds on reg2 in terms of Hilbert coefficients 381

Let G be a non-zero finitely generated *free graded R-module. Then

e0 ( • ), . . . , ed−1 ( • )

form a bounding system for reg2 on the class of all non-zero graded submod-
ules of G.
r
i=1 R(ai ), where a := (a1 , . . . , ar ) ∈ Z , and set
r
Proof. Write G =
s := min{−ai : 1 ≤ i ≤ r}. We start by defining, for each integer h ≥ 2, a
(h)
numerical function Fa : Zh −→ Z. The definition will be made by induction
on h.
First, for (e0 , e1 ) ∈ Z2 , define (with the notation of 17.1.1(v))

Fa(2) (e0 , e1 ) = s + 1 − p(e0 ,e1 ) (s − 1).


(h−1)
Now suppose that h > 2 and that the function Fa has already been defined.
Then, for e := (e0 , . . . , eh−1 ) ∈ Zh and with e := (e0 , . . . , eh−2 ), we write
(h−1) 
f = Fa (e ) and set (again with the notation of 17.1.1(v))
r  +
f + ai + h − 3
Fa(h) (e) := R0 (R0 ) − pe (f − 2) + f.
i=1
h−1

Consider first the case where d = 1, and let M be a non-zero graded sub-
i
module of G. Then dim M = 1, so that HR +
(M ) = 0 for all i > 1 and
reg (M ) = −∞. We therefore assume henceforth in this proof that d > 1.
2

As in the proof of Theorem 17.1.7, we can use the ideas of Example 16.2.4 to
show that it is enough for us to prove the theorem under the additional assump-
tion that R0 /m0 is infinite, and we shall make this assumption in what follows.
We shall prove the result by induction on d. We shall make frequent use of the
fact that, as each associated prime p of a free R-module has dim R/p = d,
every non-zero submodule of a free R-module has dimension d.
Consider now the case where d = 2, and let M be a non-zero graded sub-
module of G. Then dim M/X2 M < 2, and so reg2 (M/X2 M ) = −∞. Use
Lemma 17.3.4 with the choice f = s+1: since f +ai +d−3 = s+ai +d−2 ≤
d − 2 < d − 1 for each i = 1, . . . , r, we obtain that
r  +
s + ai + d − 2
reg (M ) ≤ R0 (R0 )
2
− χM (s − 1) + s + 1
i=1
d−1
= s + 1 − PM (s − 1) = s + 1 − p(e0 (M ),e1 (M )) (s − 1)
= Fa(2) (e0 (M ), e1 (M )).

Therefore e0 ( • ), e1 ( • ) form a bounding system for reg2 on the class of all


non-zero graded submodules of G.
382 Hilbert polynomials

Now suppose d > 2 and we have proved, for R := R0 [X1 , . . . , Xd−1 ]


r
and for every non-zero graded submodule N of i=1 R(ai ), that reg2 (N )
(d−1)
≤ Fa (e0 (N ), . . . , ed−2 (N )).
Now let M be a non-zero graded submodule of G. As gradeG R+ = d > 1,
it follows from 6.2.7, 2.2.6(i)(c) and 13.3.14 that ηR+ : G −→ DR+ (G)
is a homogeneous isomorphism. It therefore follows from the fact that the
functor DR+ is left exact that there is a homogeneous monomorphism ε :
DR+ (M ) −→ G: let L := ε(DR+ (M )), a d-dimensional graded submodule

=
of G. Since there is a homogeneous isomorphism L −→ DR+ (M ), it fol-
lows from 13.3.14 and 2.2.10(iii) that there is a homogeneous isomorphism

=
DR+ (L) −→ DR+ (M ), and from 13.4.3, 13.4.4 and 2.2.10(v) that there are

=
homogeneous isomorphisms HR i
+
(L) −→ HR i
+
(M ) for all i > 1. Hence
χL = χM , so that PL = PM and ei (L) = ei (M ) for all i = 0, . . . , d − 1; also
reg2 (L) = reg2 (M ). Now HR 1
+
(L) ∼= HR1
+
(DR+ (M )) = 0 by 2.2.10(iv),
0 0
so that, since HR+ (G) = 0, we have HR+ (G/L) = 0. It is therefore suf-
ficient for us to complete this inductive step under the additional assumption
0
that HR +
(G/M ) = 0, and so we shall make this assumption in the remainder
of this proof. In view of 2.1.1 and 16.1.4(ii), this means that R1 contains a non-
zerodivisor on G/M , and it follows from Lemma 17.3.5 that we can assume
without loss of generality that Xd is a non-zerodivisor on G/M . This we do.
Set R := R0 [X1 , . . . , Xd−1 ], G := G/Xd G, and M := M/Xd M . We can
view G and M as R-modules by means of the inclusion homomorphism R −→
R, and the natural map M −→ G is an R-monomorphism because Xd is a non-
zerodivisor on G/M . Note also that there is a homogeneous R-isomorphism
= r

G −→ i=1 R(ai ). It therefore follows from the inductive hypothesis that

reg2 (M ) ≤ Fa(d−1) (e0 (M R ), . . . , ed−2 (M R )).

A straightforward adaptation of 4.2.2 to the graded case, with use of the


Graded Independence Theorem 14.1.7 instead of the Independence Theorem
4.2.1, will show that there are homogeneous R-isomorphisms

=
i
HR +
(M R) −→ HR
i
+
(M ) R for all i ∈ N0 ,

and a similar argument based on 14.1.4 will show that there is a homogeneous
R-isomorphism

=
DR+ (M R) −→ DR+ (M ) R.

It follows that reg2 (M R) = reg2 (M/Xd M ) and, in view of 17.1.9, that

ei (M R) = ei (M/Xd M ) = ei (M ) for all i = 0, . . . , d − 2.


17.4 Bounds on reg1 and reg0 383

Set e := (e0 (M ), . . . , ed−1 (M )) ∈ Zd and e := (e0 (M ), . . . , ed−2 (M )).


(d−1) 
Thus reg2 (M/Xd M ) ≤ Fa (e ). Recall that χM (n) = PM (n) = pe (n)
for all n ∈ Z. We can now deduce from Lemma 17.3.4 with the choice f =
(d−1) 
Fa (e ) that
r  +
f + ai + d − 3
reg2 (M ) ≤ R0 (R0 ) − χM (f − 2) + f
i=1
d−1
r  +
f + ai + d − 3
= R0 (R0 ) − pe (f − 2) + f
i=1
d−1
= Fa(d) (e0 (M ), . . . , ed−1 (M )).

This completes the inductive step.

Theorem 17.3.6 is based on work of D. Mumford: see [54, p. 101]. In


our notation (of 17.3.6), Mumford’s work is concerned with the case where
G = R; thus Mumford showed that e0 ( • ), . . . , ed−1 ( • ) form a bounding
system for reg2 on the class of all non-zero graded ideals of R. In fact, Mum-
ford showed that there exists a rational polynomial q in d indeterminates such
that, for every non-zero graded ideal b of R, reg2 (b) is bounded above by
q(e0 (b), . . . , ed−1 (b)).

17.4 Bounds on reg1 and reg0


In this section we show that the bounding result for reg2 obtained in the last
section leads to bounding results for reg1 and reg0 in which the Hilbert coeffi-
cients play an important rôle.

17.4.1 Theorem. Assume that R = n∈N0 Rn is positively graded and ho-
mogeneous, and such that R0 is Artinian and local, with maximal ideal m0 ;
assume also that R is generated as R0 -algebra by d homogeneous elements of
degree 1, where d > 1. Let r, t ∈ N and a := (a1 , . . . , ar ) ∈ Zr be fixed. Let
D denote the class of all graded R-modules M of dimension t which can be
r
written in the form M = j=1 Rmj with mj ∈ M−aj for all j ∈ {1, . . . , r}.
Then e0 ( • ), . . . , et−1 ( • ) form a bounding system for reg1 on D.

Proof. We begin by defining a function that we shall use to bound reg1 . Let
r
E denote the class of non-zero graded submodules of G = i=1 R(ai ). By
17.3.6, there is a function B : Zd −→ Z such that

reg2 (L) ≤ B(e0 (L), . . . , ed−1 (L))


384 Hilbert polynomials

for each R-module L in E. Let c := (c0 , . . . , ct−1 ) ∈ Zt and let pc be the


polynomial of degree t − 1 in Q[X] given by
t−1  
X +t−i−1
pc (X) = (−1)i ci .
i=0
t−i−1

(a)
(See 17.1.1(v).) Also, define pc ∈ Q[X] by
r  
X + aj + d − 1
p(a) (X) :=  R0 (R 0 ) − pc (X).
c
j=1
d−1

(a)
With the notation of 17.1.1(v), if deg(pc ) = d − 1, set

c ), . . . , ed−1 (pc )) − 1, −a1 , . . . , −ar },


Ca(d) (c) := max{B(e0 (p(a) (a)

(d) (a)
but let Ca (c) := max{−a1 , . . . , −ar } if deg(pc ) < d − 1.
The hypotheses ensure that there is a surjective homogeneous homomor-
phism φ : S := R0 [X1 , . . . , Xd ] −→ R of graded rings. On use of the Graded
Independence Theorem 14.1.7, we see that we can replace R by S; thus we
assume that R = R0 [X1 , . . . , Xd ]. Let M be a module in D; then there is an
exact sequence 0 −→ N −→ G −→ M −→ 0 in the category *C(R). Write
e := (e0 (M ), . . . , et−1 (M )). The Hilbert polynomials PN and PM satisfy
r  
X + aj + d − 1
PN (X) = PG (X) − PM (X) = R0 (R0 ) − pe (X)
j=1
d−1

= p(a)
e (X).

(a)
If N = 0, then deg(pe ) = −1 < d − 1, and
r r
reg1 (M ) = reg1 ( j=1 R(aj )) ≤ reg( j=1 R(aj ))
= max{−a1 , . . . , −ar } = Ca(d) (e).
(a)
Now consider the case where N = 0. Then dim N = d, and so deg(pe ) =
d − 1. Notice that N lies in the class E. Therefore, in view of 17.3.6 and
16.2.15(iv), we have
r
reg1 (M ) ≤ max{reg2 (N ) − 1, reg( j=1 R(aj ))}
r
≤ max{B(e0 (N ), . . . , ed−1 (N )) − 1, reg( j=1 R(aj ))}

e ), . . . , ed−1 (pe )) − 1, −a1 , . . . , −ar }


= max{B(e0 (p(a) (a)

= Ca(d) (e).
17.4 Bounds on reg1 and reg0 385

Observe that

ei (p(a)
e )
⎧ 
⎨ei r
if 0 ≤ i ≤ d − t − 1,
j=1 R(aj )
= r
⎩ei
j=1 R(aj ) − (−1) ei−(d−t) (M ) if d − t ≤ i ≤ d − 1,
d−t

r
and that each such ei j=1 R(aj ) depends only on R0 (R0 ), a1 , . . . , ar and
d. Thus e0 ( • ), . . . , et−1 ( • ) form a bounding system for reg1 on D.
 
17.4.2 Corollary. Let R = n∈N0 Rn be as in 17.4.1. Let W = n∈Z Wn
be a finitely generated graded R-module and let 0 = P ∈ Q[x]. Then there is
an integer G such that, for each homogeneous R-homomorphism f : W −→
M of finitely generated graded R-modules that is surjective in all large degrees
and is such that PM = P , we have reg1 (M ) ≤ G.
r
Proof Choose r ∈ N and a := (a1 , . . . , ar ) ∈ Zr such that W = j=1 Rwj
with wj ∈ W−aj for all j ∈ {1, . . . , r}. Let t := deg P + 1, and define
D as in 17.4.1, that is, as the class of all graded R-modules N of dimension
r
t which can be written in the form N = j=1 Ryj with yj ∈ N−aj for
all j ∈ {1, . . . , r}. By 17.4.1, there is a function C : Zt −→ Z such that
reg1 (N ) ≤ C(e0 (N ), . . . , et−1 (N )) for each R-module N in D.
Let f : W −→ M be a homomorphism as described in the statement,
r
and set W := W/ Ker f . Then W = j=1 Rw j with w j ∈ W −aj for all
j ∈ {1, . . . , r}. Moreover, f induces an exact sequence

0 −→ W −→ M −→ C −→ 0

(in the category *C(R)), where Cn = 0 for all except finitely many n ∈ Z. In
particular, this means that PW = PM = P and reg1 (C) = −∞. Therefore,
by 16.2.15(iii) and 17.4.1, we have

reg1 (M ) ≤ reg1 (W ) ≤ C(e0 (P ), . . . , edeg P (P )) =: G.

Note, in particular, that Corollary 17.4.2 tells us (with the notation of that
corollary) that, for all graded homomorphic images M , with specified Hilbert
polynomial P , of a fixed finitely generated graded R-module W , the invariant
reg1 is bounded in terms of P .
Next, we bound the regularity in terms of the Hilbert coefficients, the postu-
lation number and the degrees of generators.

17.4.3 Corollary. Let R = n∈N0 Rn be as in 17.4.1. Let g, π, t ∈ Z with
t > 0.
386 Hilbert polynomials

Let G denote the class of all finitely generated graded R-modules of dimen-
sion t that can be generated by homogeneous elements of degrees not exceed-
ing g and whose postulation number does not exceed π. Then the invariants
e0 ( • ), . . . , et−1 ( • ) form a bounding system for reg on G.

Proof Let g := max{g, π + 1}, and, for a graded module M ∈ G, consider



the graded submodule M := n≥g Mn of M . Observe that PM = PM = pe ,
where e := (e0 (M ), . . . , et−1 (M )). Observe also that R0 (M g ) = PM (g) =
pe (g), and that M is generated by pe (g) > 0 homogeneous elements of degree
g. This time write a := (−g, . . . , −g) ∈ Zpe (g) , and define D to be the class
of all graded R-modules N of dimension t that can be generated by pe (g)
homogeneous elements all of degree g. By 17.4.1, there is a function C  :
Zt −→ Z such that reg1 (M ) ≤ C  (e0 (M ), . . . , et−1 (M )) for each R-module
M in D .
Note that there is an exact sequence 0 −→ M −→ M −→ C −→ 0 (in the
category *C(R)) in which C is finitely generated and R+ -torsion. Therefore,
by 16.2.15(iii), reg1 (M ) ≤ reg1 (M ). Since M ∈ D , we have reg1 (M ) ≤
C  (e0 (M ), . . . , et−1 (M )). However, it follows from 17.1.10(i) that reg(M ) ≤
max{reg1 (M ), pstln(M )}. Putting these inequalities together, we have

reg(M ) ≤ max{C  (e0 (M ), . . . , et−1 (M )), π},

and this is enough to complete the proof.

17.4.4 Exercise. Let R = K[X1 , X2 ] be a polynomial ring in two indeter-


minates X1 , X2 over a field K. By the generating degree gendeg(M ) of a
non-zero finitely generated graded R-module M we mean the smallest integer
g such that M is generated by homogeneous elements of degrees not exceeding
g; thus
  
gendeg(M ) = inf g ∈ Z : M = R n≤g Mn .

For each h ∈ Z, let hK be the graded R-module (see 14.1.13) such that, for all
n ∈ Z,

h K if n = h,
( K)n =
0 if n = h.

(i) For each t ∈ N, let M (t) := X1t (R/X2 R)⊕ 0K ⊕ 1K ⊕· · ·⊕ t−1K, and
determine end(ΓR+ (M (t) ), end(HR 1
+
(M (t) ), PM (t) and pstln(M (t) ).
Conclude that the multiplicity e0 ( • ) and the postulation number pstln( • )
do not form a bounding system of invariants for either reg1 or reg on the
class of all non-zero finitely generated graded R-modules.
17.4 Bounds on reg1 and reg0 387

(ii) Use the graded R-modules N (t) := R ⊕ (R/R+ t


) (t ∈ N) to show that
e0 ( • ), e1 ( • ) and gendeg( • ) do not form a bounding system of invariants
for reg on the class D of all finitely generated graded R-modules of
dimension 2.
t−1 h h+1 
(iii) For each t ∈ N, let L(t) := h=0 ( K) R≥t , where R≥t is as
h h+1
defined in 16.1.1 and ( K) denotes the direct sum of h + 1 copies of
h
K. Use the L(t) (t ∈ N) to show that e0 ( • ), e1 ( • ) and pstln( • ) do not
form a bounding system of invariants for reg1 on the class D of part (ii).

17.4.5 Definition and Exercise. Assume that R = n∈N0 Rn is positively
graded and that the ideal a is graded. The saturation of a is defined as asat :=
 n sat
n∈N (a :R (R+ ) ). We say that a is saturated precisely when a = a . Show
that
(i) asat is a graded ideal of R that contains a;
(ii) asat /a = ΓR+ (R/a);
(iii) asat is the largest graded ideal of R which coincides with a in all large
degrees;
(iv) (asat )sat = asat ; and
1
(v) if ΓR+ (R) = HR +
(R) = 0, then there is a homogeneous isomorphism
sat ∼ 1
a /a = HR+ (a).
17.4.6 Exercise. Let R = R0 [X1 , . . . , Xd ] be the polynomial ring over the
commutative Noetherian ring R0 in d > 1 indeterminates. Assume that a is
graded.
1
Show that a is saturated if and only if HR +
(a) = 0, and that reg(a) =
2
reg (a) if a is saturated.
Now assume in addition that the base ring R0 is local and Artinian. For
P ∈ Q[T ], let AP := {b is a graded ideal of R : Pb = P }. Show that the
maximal members of AP are precisely the saturated ideals in AP . Show also
that the Hilbert coefficients e0 ( • ), . . . , ed−1 ( • ) form a bounding system for
reg on the class of all non-zero saturated graded ideals of R.
18

Applications to reductions of ideals

Graded local cohomology theory has played a substantial rôle in the study of
Rees rings and associated graded rings of proper ideals in local rings. We do
not have enough space in this book to include all we would like about the
applications of local cohomology to this area, and so we have decided to select
a small portion of the theory which gives some idea of the flavour. The part
we have chosen to present in this chapter concerns links between the theory of
reductions of ideals in local rings and the concept of Castelnuovo regularity,
discussed in Chapter 16. The highlight will be a theorem of L. T. Hoa which
asserts that, if b is a proper ideal in a local ring having infinite residue field,
then there exist t0 ∈ N and c ∈ N0 such that, for all t > t0 and every minimal
reduction a of bt , the reduction number ra (bt ) of bt with respect to a is equal
to c. This statement of Hoa’s Theorem is satisfyingly simple, and makes no
mention of local cohomology, and yet Hoa’s proof, which we present towards
the end of this chapter, makes significant use of graded local cohomology.
Throughout this chapter, all graded rings and modules are to be understood
to be Z-graded, and all polynomial rings R[X1 , . . . , Xt ] (and R[T ]) over R
are to be understood to be (positively) Z-graded so that each indeterminate has
degree 1 and deg a = 0 for all a ∈ R \ {0}.

18.1 Reductions and integral closures


Reductions of ideals of local rings were first considered by D. G. Northcott
and D. Rees in [63]. Nowadays, the concept is recognized as being of major
importance in commutative algebra. The original paper [63] of Northcott and
Rees was written under unnecessarily restrictive hypotheses, and so we begin
this chapter with a rapid development of the links between the two concepts of
18.1 Reductions and integral closures 389

reduction and integral closure of ideals over a general commutative Noetherian


ring. Throughout this chapter, b will denote a second ideal of R.

18.1.1 Definitions. (See D. G. Northcott and D. Rees [63].)

(i) We say that a is a reduction of b precisely when a ⊆ b and there exists


s ∈ N0 such that abs = bs+1 ; then the least such s is denoted by ra (b)
and called the reduction number of b with respect to a. Note that, if a is
a reduction of b, then am bj = bm+j for all m ∈ N and j ≥ ra (b).
(ii) We say that a is a minimal reduction of b if and only if a is a reduction
of b and there is no reduction c of b with c ⊂ a.
(iii) We say that r ∈ R is integrally dependent on a if and only if there exist
n ∈ N and c1 , . . . , cn ∈ R with ci ∈ ai for i = 1, . . . , n such that

rn + c1 rn−1 + · · · + cn−1 r + cn = 0.

18.1.2 Exercise. Assume that a is a reduction of b.


√ √
(i) Show that a = b.
(ii) Let c be a third ideal of R such that b is a reduction of c. Show that a is
a reduction of c.

18.1.3 Exercise. Show that, if a is a reduction of b and also a reduction of


another ideal b of R, then a is a reduction of b + b .

18.1.4 Notation and Exercise. Let {a1 , . . . , ah } be a generating set for a,


and let T be an indeterminate. We use R[aT, T −1 ] to denote the subring

R[a1 T, . . . , ah T, T −1 ]

of R[T, T −1 ] = R[T ]T , and refer to this as the extended Rees ring of a.


(Note that R[a1 T, . . . , ah T, T −1 ] is independent of the choice of finite gener-
ating set {a1 , . . . , ah } for a.) Note that R[aT, T −1 ] inherits a Z-grading from
R[T, T −1 ] = R[T ]T .
Let r ∈ R. Show that r is integrally dependent on a if and only if the element
rT of R[T, T −1 ] is integral over R[aT, T −1 ].

18.1.5 Proposition. Assume that a ⊆ b and let r ∈ R.

(i) The element r is integrally dependent on a if and only if a is a reduction


of a + Rr.
(ii) The ideal a is a reduction of b if and only if each element of b is inte-
grally dependent on a.
390 Applications to reductions of ideals

Proof. First we suppose that a is a reduction of b, so that there exists s ∈ N


such that am bs = bm+s for all m ∈ N. Hence, provided we interpret ai as R
when the integer i is negative, bn ⊆ an−s for all n ∈ N0 . Therefore, within the
ring R[T, T −1 ], we have R[bT, T −1 ] ⊆ T s R[aT, T −1 ], a finitely generated
module over the Noetherian ring R[aT, T −1 ]. It follows that R[bT, T −1 ] is
integral over its subring R[aT, T −1 ], and so, by 18.1.4, each element of b is
integrally dependent on a.
(i) It follows from the above paragraph that, if a is a reduction of a + Rr,
then r is integrally dependent on a. Conversely, if r is integrally dependent on
a, then there exist n ∈ N and c1 , . . . , cn ∈ R with ci ∈ ai for i = 1, . . . , n
such that rn + c1 rn−1 + · · · + cn−1 r + cn = 0. Then

(a + Rr)n = a(a + Rr)n−1 + Rrn = a(a + Rr)n−1

and a is a reduction of a + Rr.


(ii) Suppose each element of b is integrally dependent on a. Let {b1 , . . . , bt }
be a generating set for b. By part (i), for each i = 1, . . . , t, the ideal a is a
t
reduction of a + Rbi . Hence a is a reduction of a + i=1 Rbi = b, by 18.1.3.
The converse statement has already been proved in the first paragraph of this
proof.

18.1.6 Corollary and Definition. By 18.1.3, the set I of all ideals of R that
have a as a reduction has a unique maximal member, a say: a is the union of
the members of I. By 18.1.5, this ideal a is precisely the set of all elements of
R which are integrally dependent on a (and so the latter set is an ideal of R):
we refer to a as the integral closure of a. 

18.1.7 Exercise. Show that

(i) the integral closure a is not a reduction of any ideal of R which properly
contains it, and
(ii) a = b if and only if a and b are both reductions of a + b.

We now reproduce the classical argument of Northcott and Rees which


shows that, when R is local, every reduction of b contains a minimal reduction
of b.

18.1.8 Lemma. Assume that (R, m) is local and that a ⊆ b + am. Then
a ⊆ b.

Proof. Since b + a = b + am, we have (b + a)/b = m ((b + a)/b), and so


the result follows from Nakayama’s Lemma.
18.1 Reductions and integral closures 391

18.1.9 Lemma (Northcott and Rees [63, §2, Lemma 2]). Assume that (R, m)
is local and that a ⊆ b. Then a is a reduction of b if and only if a + bm is a
reduction of b.
Proof. It is clear that, if a is a reduction of b, then a + bm is a reduction of b.
Now assume that a + bm is a reduction of b. Then there exists n ∈ N such that
(a + bm)bn = bn+1 , that is, bn+1 = abn + bn+1 m. It therefore follows from
Lemma 18.1.8 that bn+1 = abn .
18.1.10 Lemma. Assume that (R, m) is local and that a is a reduction of b.
Let a1 , . . . , at ∈ a be such that their natural images in (a + bm)/bm form a
t
basis for this R/m-space, and set a = i=1 Rai . Then
(i) a is a reduction of b contained in a and a ∩ bm = a m; and
(ii) if a is a minimal reduction of b, then a = a and a ∩ bm = am.
Proof. (i) Note that a ⊆ a ⊆ b and a + bm = a + bm. Since a is a reduction
of b, it follows from Lemma 18.1.9 that a + bm = a + bm is a reduction of
b, so that a is a reduction of b by the same lemma. Also, if r1 , . . . , rt ∈ R
t
are such that i=1 ri ai ∈ bm, then, by choice of a1 , . . . , at , we must have
r1 , . . . , rt ∈ m.
(ii) Now suppose that a is a minimal reduction of b. Then a = a by part (i),
so that a ∩ bm = am, again by part (i).
18.1.11 Exercise. Assume (R, m) is local and that a is a minimal reduction
of b. Let d be an ideal of R such that a ⊆ d ⊆ b. Show that every minimal
generating set for a can be extended to a minimal generating set for d.
18.1.12 Theorem (Northcott and Rees [63, §2, Theorem 1]). Assume that
(R, m) is local and that a is a reduction of b. Then a contains a minimal
reduction of b.
Proof. Let Σ be the set of all ideals of R of the form d + bm, where d is a
reduction of b and is contained in a. Note that a + bm ∈ Σ. Since b/bm is a
finite-dimensional vector space over R/m, the set Σ has a minimal member,
and so there exists a reduction c of b contained in a such that c + bm is a
minimal member of Σ. Let c1 , . . . , ct ∈ c be such that their natural images in
t
(c + bm)/bm form a basis for this R/m-space, and set c = i=1 Rci . Now
c is a reduction of b contained in c , and c ∩ bm = cm by 18.1.10(i); we shall
prove that c is a minimal reduction of b.
Suppose that c0 is a reduction of b with c0 ⊆ c. Then c0 + bm ∈ Σ and
c0 + bm ⊆ c + bm. Hence, by the choice of c , we must have
c0 + bm = c + bm = c + bm.
392 Applications to reductions of ideals

It follows that c ⊆ c0 + bm; therefore c ⊆ c0 + (bm ∩ c) ⊆ c0 + cm by the


immediately preceding paragraph in this proof. It now follows from 18.1.8 that
c ⊆ c0 , so that c = c0 .

There is an important connection between minimal reductions of ideals in


local rings and analytical independence.

18.1.13 Definition. Assume that (R, m) is local. Then v1 , . . . , vt ∈ b are


said to be analytically independent in b if and only if, whenever h ∈ N and
f ∈ R[X1 , . . . , Xt ] (the ring of polynomials over R in t indeterminates) is a
homogeneous polynomial of degree h such that f (v1 , . . . , vt ) ∈ bh m, then all
the coefficients of f lie in m.
Note that, if v1 , . . . , vt ∈ b are analytically independent in b, and c :=
t
i=1 Rvi , then c ∩ b m = c m for all h ∈ N.
h h h

18.1.14 Lemma (Northcott and Rees [63, §4, Lemma 1]). Assume (R, m)
is local and b is proper, and let v1 , . . . , vt ∈ b be analytically independent
t
in b. Then v1 , . . . , vt form a minimal generating set for i=1 Rvi , and, fur-
t
thermore, if {w1 , . . . , wt } is a minimal generating set for i=1 Rvi , then
w1 , . . . , wt are analytically independent in b.
t t
Proof. Set c := i=1 Rvi . If r1 , . . . , rt ∈ R are such that i=1 ri vi ∈ cm,
then, since c ⊆ b, it follows that r1 , . . . , rt ∈ m. Hence v1 , . . . , vt form a
minimal generating set for c.
Next, let h ∈ N and f ∈ R[X1 , . . . , Xt ] be a homogeneous polynomial of
degree h such that f (w1 , . . . , wt ) ∈ bh m. There exist rij ∈ R (1 ≤ i, j ≤ t)
t
such that wi = j=1 rij vj for i = 1, . . . , t, and so
t t
f j=1 r1j vj , . . . , j=1 rtj vj ∈ bh m.

Since v1 , . . . , vt ∈ b are analytically independent in b, it follows that all the


coefficients of the homogeneous polynomial
t t
f j=1 r1j Xj , . . . , j=1 rtj Xj

lie in m.
Denote the natural image in k := R/m (respectively k[X1 , . . . , Xt ]) of r ∈
R (respectively q ∈ R[X1 , . . . , Xt ]) by r (respectively q). Now [rij ] is an
invertible t × t matrix over k, and so there exists a t × t matrix [sij ] over R
such that [rij ][sij ] = It , the t × t identity matrix. Since
t t
f j=1 r1j Xj , . . . , j=1 rtj Xj =0
18.2 The analytic spread 393

in k[X1 , . . . , Xt ], it follows that


t t t t
f j=1 r1j l=1 sjl Xl , . . . , j=1 rtj l=1 sjl Xl = 0,

and so f (X1 , . . . , Xt ) = 0. Thus all the coefficients of f lie in m.


18.1.15 Proposition (Northcott and Rees [63, §4, Lemma 2]). Assume that
(R, m) is local and that k := R/m is infinite. Suppose that b is proper and a
is a minimal reduction of b, and let v1 , . . . , vt form a minimal generating set
for a. Then v1 , . . . , vt are analytically independent in b.
Proof. Let h ∈ N and f ∈ R[X1 , . . . , Xt ] be a homogeneous polynomial of
degree h such that f (v1 , . . . , vt ) ∈ bh m. Suppose that the coefficient u of X1h
is a unit of R. Then v1h ∈ bh m + v2 ah−1 + · · · + vt ah−1 , and so
ah = Rv1h + (Rv2 + · · · + Rvt )ah−1 ⊆ bh m + (Rv2 + · · · + Rvt )ah−1 .
Set a := Rv2 + · · · + Rvt . Now there exists n ∈ N such that abn = bn+1 .
Hence bh+n = ah bn ⊆ bh+n m + a ah−1 bn = bh+n m + a bh+n−1 . Hence
bh+n ⊆ a bh+n−1 by 18.1.8, so that bh+n = a bh+n−1 and a is a reduction
of b. This contradicts the fact that a is a minimal reduction of b, since a ⊂ a.
Hence u ∈ m.
Next let [rij ] be a t × t matrix over R such that all the entries in the first
column, that is r11 , r21 , . . . , rt1 , are units of R and det[rij ] is also a unit.
Then there is a minimal generating set {w1 , . . . , wt } for a such that vi =
t
j=1 rij wj for all i = 1, . . . , t. Since
t t
f j=1 r1j wj , . . . , j=1 rtj wj ∈ bh m,

it follows from the first paragraph of this proof that f (r11 , . . . , rt1 ), which is
t t
the coefficient of X1h in the form f j=1 r1j Xj , . . . , j=1 rtj Xj , lies in
m. Thus, if we denote the natural image in k = R/m of r ∈ R by r, and the
natural image in k[X1 , . . . , Xt ] of q ∈ R[X1 , . . . , Xt ] by q, then
f (r11 , . . . , rt1 ) = 0.
Therefore f (α1 , . . . , αt ) = 0 for all choices of (α1 , . . . , αt ) ∈ (k \ {0})t ;
hence, since k is infinite, f = 0.

18.2 The analytic spread


We now consider the graded rings which will provide the framework for con-
nections between reductions of ideals and graded local cohomology.
394 Applications to reductions of ideals

18.2.1 Notation and Definition. Suppose that b is proper and a is a reduction


of b.
(i) By 18.1.6 and 18.1.4, the extended Rees ring R[bT, T −1 ] is integral over

its subring R[aT, T −1 ]. The associated graded ring i∈N0 bi /bi+1 will

be denoted by G(b) = i∈N0 G(b)i . Since there is a homogeneous iso-

morphism of graded rings R[bT, T −1 ]/T −1 R[bT, T −1 ] −→ G(b) (and
=

a similar one for G(a)), it follows that G(b) is integral over the natural
image of G(a).
(ii) Now suppose, in addition, that (R, m) is local. The extension mG(b) of
m to G(b) under the natural ring homomorphism is a graded ideal: in
 i i+1
fact, mG(b) = i∈N0 mb /b . The analytic spread spr(b) of b is
defined by
spr(b) := dim (G(b)/mG(b)) .
Note that G(b)/mG(b) is (homogeneously) isomorphic to the graded

ring i∈N0 bi /mbi , in which (ri + mbi )(rj + mbj ) = ri rj + mbi+j for
(i, j ∈ N0 and) ri ∈ bi , rj ∈ bj . Some authors refer to the graded ring
G(b)/mG(b) as the fibre cone of b.
(iii) Note that, since a is a reduction of b, it follows that G(b)/mG(b) is
integral over the natural image of G(a)/mG(a). Observe also that, if a
can be generated by elements which are analytically independent in b,
then ai ∩ bi m = ai m for all i ∈ N, so that the natural homogeneous ring
homomorphism G(a)/mG(a) −→ G(b)/mG(b) is injective.
18.2.2 Lemma. Suppose (R, m) is local and b is proper. Then spr(b) =
spr(bt ) for all t ∈ N.
Proof. By Definition 18.2.1(ii),
 
spr(b) := dim (G(b)/mG(b)) and spr(bt ) := dim G(bt )/mG(bt ) .
Now there are homogeneous ring isomorphisms

=
 ∼
=

G(b)/mG(b) −→ bi /mbi and G(bt )/mG(bt ) −→ bti /mbti .
i∈N0 i∈N0

It is clear from these that G(bt )/mG(bt ) is homogeneously isomorphic to the


t-th Veronesean subring (see 13.5.9) of G(b)/mG(b). The claim now follows
from the fact that a commutative Noetherian graded ring has the same dimen-
sion as its t-th Veronesean subring because the former is integral over the
latter.
The observations in 18.2.1(iii) have some very important consequences.
18.2 The analytic spread 395

18.2.3 Remark. Suppose that (R, m) is local, that a ⊆ b ⊂ R and that


a is generated by v1 , . . . , vt which are analytically independent in b. Set k :=
R/m. Then v1 , . . . , vt are, of course, analytically independent in a, and it is im-
mediate from Definition 18.1.13 that there is a homogeneous isomorphism φ :
= 

k[X1 , . . . , Xt ] −→ i i
i∈N0 a /ma of graded k-algebras (where X1 , . . . , Xt
are independent indeterminates) such that φ(Xi ) = vi +ma for all i = 1, . . . , t.

Since G(a)/mG(a) ∼ = i∈N0 ai /mai , we have dim(G(a)/mG(a)) = t.
18.2.4 Theorem (Northcott and Rees [63, §4, Theorems 1 and 2]). Assume
that (R, m) is local and that k := R/m is infinite. Suppose that b is proper and
that a is a reduction of b, and let t := dimk (a/ma), the number of elements in
each minimal generating set for a. Let {v1 , . . . , vt } be a minimal generating
set for a. Then

(i) spr(b) ≤ t;
(ii) a is a minimal reduction of b if and only if spr(b) = t;
(iii) hence, at least spr(b) elements are needed to generate a, and, if a can
be generated by spr(b) elements, then it is a minimal reduction of b;
(iv) a is a minimal reduction of b if and only if v1 , . . . , vt are analytically
independent in b.

Proof. Note that a contains a minimal reduction c of b, by Theorem 18.1.12;


let w1 , . . . , ws form a minimal generating set for c. Then w1 , . . . , ws are ana-
lytically independent in b by 18.1.15. By 18.2.1(iii), we can view the graded
ring G(b)/mG(b) as an integral extension ring of G(c)/mG(c), and so

dim(G(c)/mG(c)) = dim(G(b)/mG(b)) = spr(b).

However, it is immediate from 18.2.3 that dim(G(c)/mG(c)) = s, and so s =


spr(b).
(i) By 18.1.11, every minimal generating set for c can be extended to a min-
imal generating set for a. Hence spr(b) = s ≤ t.
(ii) If a is a minimal reduction of b, then a = c and t = s = spr(b).
Now suppose that t = spr(b). By 18.1.11, every minimal generating set for
c can be extended to a minimal generating set for a, and so spr(b) = s ≤ t =
spr(b). Therefore s = t and w1 , . . . , ws actually generate a. It follows that
a = c, and a is a minimal reduction of b.
(iii) This is a restatement of most of parts (i) and (ii).
(iv) If a is a minimal reduction of b, then v1 , . . . , vt are analytically inde-
pendent in b by 18.1.15. Conversely, if v1 , . . . , vt are analytically indepen-
dent in b, then t = dim(G(a)/mG(a)) by 18.2.3, and G(b)/mG(b) can be
viewed as an integral extension ring of G(a)/mG(a), by 18.2.1(iii); therefore
396 Applications to reductions of ideals

t = dim(G(b)/mG(b)) = spr(b), so that we can deduce from part (ii) that a


is a minimal reduction of b.

The next Lemma 18.2.5 is included to help the reader solve Exercise 18.2.6;
that exercise presents an important result of Northcott and Rees.

18.2.5 Lemma. Suppose that (R, m) is local and that v1 , . . . , vt are analyti-
cally independent in b. Then t ≤ spr(b).
t
Proof. Set k := R/m and c := i=1 Rvi . Since ci ∩bi m = ci m for all i ∈ N,
the natural homogeneous ring homomorphism G(c)/mG(c) −→ G(b)/mG(b)
is injective. Therefore, by 18.2.3, and with an obvious notation,
 
i+t−1
dimk (G(b)/mG(b))i ≥ dimk (G(c)/mG(c))i = for all i ∈ N0 .
t−1
Hence, with the notation of 17.1.8, deg PG(b)/mG(b) ≥ t − 1, and therefore
t ≤ spr(b) by 17.1.7.

18.2.6 Exercise (Northcott and Rees [63, §4, Theorem 3]). Assume (R, m) is
local and that k := R/m is infinite. Suppose that b is proper. Show that spr(b)
is the maximum number of elements of b which are analytically independent
in b.

18.2.7 Exercise. Assume (R, m) is local and let v1 , . . . , vt ∈ R. Recall from


[50, pp. 106–107] that v1 , . . . , vt are analytically independent if and only if,
whenever f ∈ R[X1 , . . . , Xt ] is a homogeneous polynomial such that

f (v1 , . . . , vt ) = 0,

then all the coefficients of f lie in m.


Show that v1 , . . . , vt are analytically independent if and only if they are
t
analytically independent in the ideal i=1 Rvi which they generate.

18.2.8 Exercise (Northcott and Rees [63, §4, Lemma 4, Theorems 4, 5, and
§6, Theorem 1]). Assume that (R, m) is local and that k := R/m is infinite.
Suppose that b is proper. We say that b is basic precisely when it has no reduc-
tion other than itself. Let v1 , . . . , vt form a minimal generating set for b.

(i) Show that ht b ≤ spr(b) ≤ dimk (b/mb) = t.


(ii) Show that the following statements are equivalent:
(a) b is basic;
(b) v1 , . . . , vt are analytically independent (see Exercise 18.2.7);
(c) spr(b) = dimk (b/mb).
18.3 Links with Castelnuovo regularity 397

(iii) Show that, if b can be generated by t elements and has height t, then
b is basic and the members of each minimal generating set for b are
analytically independent.
(iv) Deduce that the members of a system of parameters for R are analyti-
cally independent.
(v) Let q be an m-primary ideal of R. Show that spr(q) = dim R. (Here is
a hint: R (qn /qn+1 ) = R (R/qn+1 ) − R (R/qn ) for all n ∈ N, where
‘’ denotes length.) Deduce that each minimal reduction of q is an ideal
generated by a system of parameters of R.

18.3 Links with Castelnuovo regularity


Local cohomology has still not made an appearance in this chapter! We are
soon going to make, in the case where (R, m) is local, some calculations with

local cohomology over the associated graded ring G(b) = n∈N0 G(b)n of a
proper ideal b of R. We note that G(b) is a positively graded, homogeneous
commutative Noetherian ring: the theory in the next part of this chapter con-
cerns such rings and will eventually be applied to G(b).

18.3.1 Definition. Assume that R = n∈N0 Rn is positively graded and
homogeneous. We say that an ideal A of R is a *reduction of R+ if and only
if A is graded, can be generated by homogeneous elements of R of degree 1,
and is a reduction of R+ .

18.3.2 Lemma. Assume that R = n∈N0 Rn is positively graded and ho-

mogeneous, and let A = n∈N An be a graded ideal of R generated by homo-
geneous elements of degree 1. Then A is a *reduction of R+ if and only if there
exists m ∈ N0 such that Rm+1 = Am+1 ; when this is the case, Ri+1 = Ai+1
for all i ≥ m, and the least i ∈ N0 such that Ri+1 = Ai+1 is equal to the
reduction number rA (R+ ) (see 18.1.1), so that rA (R+ ) = end(R/A).
Proof. Suppose that R = R0 [x1 , . . . , xt ], where x1 , . . . , xt ∈ R1 , and that
A is generated by a1 , . . . , ah ∈ A1 . We use the notation R≥i , for i ∈ N, of
16.1.1.
Let i ∈ N0 . Then a typical element of Ri is a sum of finitely many elements
of the form r0 xj11 . . . xjtt , where r0 ∈ R0 and j1 , . . . , jt ∈ N0 are such that
t
k=1 jk = i, and a typical element of Ai+1 is a sum of finitely many elements
of the form al ri , where l ∈ {1, . . . , h} and ri ∈ Ri . It follows that, if Ri+1 =
Ai+1 , then Ri+2 = Ai+2 , that (R+ )i = R≥i , and that A(R+ )i = A≥i+1 . All
the claims now follow easily.
398 Applications to reductions of ideals

18.3.3 Remark. Assume that R = n∈N0 Rn is positively graded and that

(R0 , m0 ) is local; set k := R0 /m0 . Let A = n∈N An be a graded ideal of R
generated by homogeneous elements of degree 1. Then the minimum number
of homogeneous elements (of degree 1) needed to generate A is

dimk (A1 /m0 A1 ).

18.3.4 Lemma. Suppose that (R, m) is local, and that a ⊆ b ⊂ R. Let



G(a) denote the ideal of the associated graded ring G(b) = i∈N0 G(b)i of
b generated by {a + b2 : a ∈ a} ⊆ G(b)1 .
Then a is a reduction of b if and only if G(a) is a *reduction of G(b)+ .
Furthermore, when this is the case, ra (b) = rG(a) (G(b)+ ).

Proof. Let i ∈ N0 . Then the (i + 1)-th component G(a)i+1 of G(a) is given


by G(a)i+1 = (abi + bi+2 )/bi+2 . By 18.1.8 and the fact that b is proper,
abi = bi+1 if and only if abi + bi+2 = bi+1 , that is, if and only if G(a)i+1 =
G(b)i+1 . The claims therefore follow from 18.3.2.

18.3.5 Lemma (L. T. Hoa [35, Lemma 2.3]). Assume that (R, m) is local
and that k := R/m is infinite. Suppose that b is proper. Then spr(b) =
ara(G(b)+ ), and their common value is the greatest integer i such that
i
HG(b)+
(G(b)) = 0.

Proof. By 18.1.12 and 18.2.4, there exists a minimal reduction a of b, and this
can be generated by s := spr(b) elements. Therefore, by 18.3.4, and with the
notation   G(a) is a *reduction of G(b)+ . Hence, by 18.1.2(i),
of that Lemma,
we have G(a) = G(b)+ , and so ara(G(b)+ ) ≤ s. Let M be an arbitrary
G(b)-module. By 3.3.3, we have HG(b) i
+
(M ) = 0 for every i > ara(G(b)+ ),
and so, in particular, for every i > s.
Consequently, the epimorphism G(b) −→ G(b)/mG(b) induces an epimor-
s
phism HG(b) +
(G(b)) −→ HG(b)s
+
(G(b)/mG(b)), and so it is enough for us to
s
show that HG(b)+ (G(b)/mG(b)) = 0 in order to complete the proof. This we
do.
The extension of G(b)+ to the *local graded ring G(b)/mG(b) is the unique
*maximal ideal, and is, in fact, also maximal; it is therefore of height s, by
14.1.14. It therefore follows from Theorem 6.1.4 and the Graded Independence
s
Theorem 14.1.7 that HG(b) +
(G(b)/mG(b)) = 0.

18.3.6 Exercise. Let the situation be as in Lemma 18.3.5, and let a be a


minimal reduction of b (such an a certainly exists, by 18.1.12). Show that
the graded ideal G(a) of G(b) defined in 18.3.4 can be generated by spr(b)
homogeneous elements of degree 1, and not by fewer.
18.3 Links with Castelnuovo regularity 399

In order to make further progress, we wish to use the fact that, when R =

n∈N0 Rn is positively graded, homogeneous, and such that (R , m ) is local
0 0
with infinite residue field, and the non-zero graded ideal A = n∈N An of R
is a *reduction of R+ , it is possible to generate A by the members of an R+ -
filter-regular sequence of length dimR0 /m0 (A1 /m0 A1 ). We therefore explain
the concept of R+ -filter-regular sequence.

18.3.7 Definition. Assume that R is positively graded, and let M be a non-


zero finitely generated graded R-module. Let f1 , . . . , fh be a sequence of
homogeneous elements of R. We say that f1 , . . . , fh is an R+ -filter-regular
sequence with respect to M if and only if, for all i = 1, . . . , h, we have

fi ∈ P∈Ass(M/(f1 ,...,fi−1 )M )\Var(R+ ) P.

We collect some elementary properties of R+ -filter-regular sequences to-


gether in the next exercise.

18.3.8 Exercise. Assume that R is positively graded, and let M be a non-


zero finitely generated graded R-module and f1 , . . . , fh be a sequence of ho-
mogeneous elements of R. Show that the following statements are equivalent:

(i) f1 , . . . , fh is an R+ -filter-regular sequence with respect to M ;


(ii) for all P ∈ Spec(R) \ Var(R+ ), the sequence f1 /1, . . . , fh /1 of natural
images in RP is a poor MP -sequence;
(iii) ((f1 , . . . , fi−1 )M :M fi )/(f1 , . . . , fi−1 )M is R+ -torsion for all i =
1, . . . , h;
(iv) end ((f1 , . . . , fi−1 )M :M fi )/(f1 , . . . , fi−1 )M ) < ∞ for i = 1, . . . , h
(the end of a graded R-module was defined in 14.1.1);
(v) fi is a non-zerodivisor on M/(f1 , . . . , fi−1 )M for each i = 1, . . . , h,
where L, for an R-module L, denotes L/ΓR+ (L).

18.3.9 Exercise. Let the situation be as in 18.3.8, and assume that f1 , . . . , fh


is an R+ -filter-regular sequence with respect to M . Let deg fi = ni for i =
1, . . . , h. Show that
h h
reg1 M/ i=1 fi M ≤ reg1 (M ) + i=1 ni − h.

We now turn to the construction of R+ -filter-regular sequences.



18.3.10 Proposition. Assume that R = n∈N0 Rn is positively graded,
that (R0 , m0 ) is local with infinite residue field k := R0 /m0 , and that the

non-zero graded ideal A = n∈N An of R is a *reduction of R+ . Let t :=
t
dimk (A1 /m0 A1 ). Then there exist f1 , . . . , ft ∈ A1 such that A = i=1 Rfi
and f1 , . . . , ft is an R+ -filter-regular sequence with respect to R.
400 Applications to reductions of ideals

Proof. Suppose, inductively, that i ∈ N0 with i < t and that an R+ -filter-


regular sequence (with respect to R) f1 , . . . , fi of elements of A1 has been
constructed such that the natural images of f1 , . . . , fi in A1 /m0 A1 are linearly
independent in this k-space. This is certainly the√ case when i = 0. Set f :=
i
j=1 Rf j . Note that m 0 A + f = A and that A = R+ by 18.1.2(i). We
can use 16.1.3 to see that there exists
⎛ ⎛ ⎞⎞
 
fi+1 ∈ A1 \ ⎝(m0 A + f) ⎝ P ⎠⎠ .
P∈Ass(R/f)\Var(R+ )

Then the extended sequence f1 , . . . , fi , fi+1 is R+ -filter-regular (with respect


to R), and the choice of fi+1 ensures that the natural images of f1 , . . . , fi , fi+1
in A1 /m0 A1 are linearly independent. This completes the inductive step.
In this way, we can construct an R+ -filter-regular sequence (with respect
to R) f1 , . . . , ft of elements of A1 whose natural images in A1 /m0 A1 form
a basis for this k-space. Since A can be generated by elements of degree 1, it
follows that f1 , . . . , ft generate A.
We now show that R+ -filter-regular sequences lend themselves to satisfac-
tory calculations with various Castelnuovo regularities. We remind the reader
that, when R is positively graded and homogeneous, the regularity regl (M ) of
M at and above level l (for l ∈ N0 ) of a finitely generated graded R-module
M was defined in 16.2.9.

18.3.11 Proposition. Assume that R = n∈N0 Rn is positively graded and
homogeneous, let M be a non-zero finitely generated graded R-module, and
let f1 , . . . , fh ∈ R1 be an R+ -filter-regular sequence with respect to M . Then
regl (M ) ≤ regl−h (M/(f1 , . . . , fh )M ) for all l ≥ h and
regh (M ) ≤ reg(M/(f1 , . . . , fh )M ) ≤ reg(M ).

Proof. We prove this by induction on h, there being nothing to prove when


h = 0. Suppose, inductively, that h > 0, and that both statements in the claim
h−1
have been proved for smaller values of h. Set f := i=1 Rfi . By our inductive
hypothesis, regl (M ) ≤ regl−h+1 (M/fM ) for all l ≥ h−1 and regh−1 (M ) ≤
reg(M/fM ) ≤ reg(M ).
j
Since (fM :M fh )/fM is R+ -torsion, we have HR +
((fM :M fh )/fM ) = 0
for all j ∈ N, by 2.1.7(i); therefore, the canonical epimorphism π : M/fM −→
M/(fM :M fh ) induces homogeneous isomorphisms
j j ∼
=
j
HR +
(π) : HR +
(M/fM ) −→ HR +
(M/(fM :M fh )) for all j ∈ N.

Hence regl (M/fM ) = regl (M/(fM :M fh )) for all l ∈ N.


18.3 Links with Castelnuovo regularity 401

Multiplication by fh leads to an exact sequence

0 −→ (M/(fM :M fh ))(−1) −→ M/fM −→ M/(f + Rfh )M −→ 0

of graded R-modules and homogeneous homomorphisms. By 16.2.15(iv) and


the inductive hypothesis,

reg(M/(f + Rfh )M )
≤ max reg1 ((M/(fM :M fh ))(−1)) − 1, reg(M/fM )
= max reg1 (M/fM ), reg(M/fM ) = reg(M/fM )
≤ reg(M ).

We can also apply 16.2.15(ii) to deduce that, for l ≥ h,

regl−h+1 (M/fM )
= regl−h+1 ((M/(fM :M fh ))(−1)) − 1
≤ max regl−h+1 (M/fM ) − 1, regl−h (M/(f + Rfh )M ) ,

so that regl−h+1 (M/fM ) ≤ regl−h (M/(f + Rfh )M ). We can therefore use


the inductive hypothesis to see that

regl (M ) ≤ regl−h+1 (M/fM ) ≤ regl−h (M/(f + Rfh )M ) for all l ≥ h.

In particular, regh (M ) ≤ reg(M/(f + Rfh )M ), and so this completes the


inductive step.

18.3.12 Theorem. (See N. V. Trung [86, Proposition 3.2].) Assume (R, m)


is local and that k := R/m is infinite. Suppose that b is proper, and set s :=
spr(b). Let a be a minimal reduction of b. Then

regs (G(b)) ≤ ra (b) ≤ reg(G(b)).

Proof. By 18.3.4, and with the notation of that lemma, G(a) is a *reduction
of G(b)+ , so that G(b)/G(a) is G(b)+ -torsion and (by 18.3.2)

ra (b) = rG(a) (G(b)+ ) = end(G(b)/G(a)).


i
Note that, by 2.1.7(i), we have HG(b) +
(G(b)/G(a)) = 0 for all i ∈ N, and so

ra (b) = end(G(b)/G(a)) = reg(G(b)/G(a)).

Next, we observe from 18.3.6 that G(a) can be generated by s := spr(b)


homogeneous elements of degree 1, and not by fewer. Therefore, by 18.3.3
and 18.3.10, there exists in G(a) a G(b)+ -filter-regular sequence (with respect
402 Applications to reductions of ideals

to G(b)) f1 , . . . , fs of elements of degree 1 which generate this ideal. Conse-


quently, by Proposition 18.3.11, we have
regs (G(b)) ≤ reg(G(b)/G(a)) ≤ reg(G(b))
and this completes the proof because we have already shown that ra (b) =
reg(G(b)/G(a)).
We are now going to use Trung’s Theorem 18.3.12 to derive a theorem due
to L. T. Hoa, which shows that, in the situation of Trung’s Theorem, for t ∈ N
sufficiently large, the reduction number of bt with respect to a minimal reduc-
tion of bt is independent of the choice of minimal reduction and independent
of t. We need one preparatory result, also due to Hoa.
18.3.13 Proposition (L. T. Hoa [35, Lemma 2.4]). Assume that (R, m) is
local and that k := R/m is infinite. Suppose that b is proper, and assume that
s := spr(b) ≥ 1. For q ∈ Q, we use !q" to denote max{i ∈ Z : i ≤ q}; we
interpret !q" as −∞ when q = −∞. Let t ∈ N. Then
1 23
j j
(i) end HG(b t) (G(b t
)) ≤ end H G(b)+ (G(b)) t for all j ∈ N0 ;
+
1 23
s t s
(ii) end HG(b t ) (G(b ))
+
= end HG(b) +
(G(b)) t .
Proof. By 18.2.2 and 18.3.5, we have
j j
end HG(b t
t ) (G(b ))
+
= end HG(b)+
(G(b)) = −∞ for all j > s.
The inequality in part (i) is therefore certainly true when j > s.
Now consider the case where 0 ≤ j ≤ s, and let i ∈ N0 with i ≤ t. The
extension bi G(bt ) of bi to G(bt ) under the composition R → R/bt → G(bt )
of canonical ring homomorphisms is a graded ideal of G(bt ) with grading given
by

bi G(bt ) = btn+i /bt(n+1) .
n∈N0

Note that b G(b ) = 0. Our strategy is to consider the chain of submodules


t t

G(bt ) = b0 G(bt ) ⊇ · · · ⊇ br G(bt ) ⊇ br+1 G(bt ) ⊇ · · · ⊇ bt G(bt ) = 0,


and to obtain information about the ‘subquotients’ in the chain by use of
Veronesean functors (see 13.5.9).
There is a natural homogeneous surjective ring homomorphism θ : G(bt ) →
(G(b))(t) , and so any graded G(b)(t) -module can be regarded as a graded
G(bt )-module by means of θ. This applies, in particular, to

(G(b))(t,i) = btn+i /btn+i+1 .
n∈N0
18.3 Links with Castelnuovo regularity 403

It follows that, for i < t, there is an exact sequence


0 −→ bi+1 G(bt ) −→ bi G(bt ) −→ (G(b))(t,i) −→ 0
of graded G(bt )-modules and
1 homogeneous homomorphisms.
23
j
Let r ∈ Z with r > end HG(b) +
(G(b)) t . By 13.5.9(v) and the
Graded Independence Theorem 14.1.7, there are homogeneous G(bt )-isomor-
phisms
j ∼
= ∼
=
(t,i)
HG(bt ) ((G(b))
(t,i)
) −→ H j (t)
j
((G(b))(t,i) ) −→ HG(b) (G(b)) .
+ G(b)+ +

Hence
j
HG(bt ) ((G(b))
+
(t,i)
)r ∼ j
= HG(b)+ (G(b))rt+i = 0,
j
since rt + i > end(HG(b) +
(G(b))). Therefore the short exact sequence dis-
played in the last paragraph induces an R/bt -epimorphism
j j
HG(bt ) (b
+
i+1
G(bt ))r −→ HG(bt ) (b G(b ))r .
+
i t

Since bt G(bt ) = 0, we can therefore deduce by descending induction that


j j
t ) (b G(b ))r = 0 for i = t − 1, . . . , 1, 0. Hence HG(bt ) (G(b ))r = 0.
i t t
HG(b + 1 23 +
j j
Hence end HG(b t
t ) (G(b )) ≤ end HG(b) (G(b)) t .
+ +
1 23
s
Now consider the case where j = s. Let u := end HG(b) +
(G(b)) t .
Thus end(HG(b)s
+
(G(b))) = ut + i for some integer i with 0 ≤ i ≤ t − 1.
i
Recall from 18.3.5 that s is the greatest integer i such that HG(b) +
(G(b)) = 0.
s
It therefore follows from 16.2.5 that HG(b) +
(G(b)) n = 0 for all n ≤ ut + i,
and so, in particular, for n = ut. Recall also, from 18.2.2 and 18.3.5, that
s = ara(G(bt )+ ). It therefore follows from the exact sequence
0 −→ bG(bt ) −→ G(bt ) −→ (G(b))(t,0) −→ 0
of graded G(bt )-modules and homogeneous homomorphisms that there is an
exact sequence of R/bt -modules

t ) (G(b ))u −→ HG(b) (G(b))ut −→ 0.


s t s
HG(b + +

s
Hence HG(b t
t ) (G(b ))u = 0, and end
+
s
HG(b t
t ) (G(b ))
+
≥ u. This, in con-
junction with the result of part (i), completes the proof.
We are now ready to present Hoa’s proof of his theorem, mentioned in the
introduction to this chapter, about the asymptotic behaviour, with respect to
reduction numbers, of powers of a proper ideal of a local ring having infinite
residue field. We remark again that there is no mention of local cohomology in
404 Applications to reductions of ideals

the statement of the theorem, but the powerful tool of graded local cohomology
plays a major rôle in the proof.
18.3.14 Theorem (L. T. Hoa [35, Theorem 2.1]). Assume that (R, m) is local
and that R/m is infinite. Suppose that b is proper. Then there exist t0 ∈ N and
c ∈ N0 such that, for all t > t0 and every minimal reduction a of bt , we have
ra (bt ) = c.
Proof. Let s := spr(b). If s = 0, then b is nilpotent, and the result is obvious
because the only minimal reduction of the zero ideal of R is the zero ideal
itself. We therefore assume that s ≥ 1 for the remainder of this proof. Set
t0 := max{| end(HG(b) i
+
(G(b)))| : i ∈ N0 and HG(b)
i
+
(G(b)) = 0}. Note
that t0 ∈ N0 , by 18.3.5.
Suppose that t ∈ N with t > t0 . By Proposition 18.3.13, we have
i
end HG(b t
t ) (G(b ))
+
≤0 for i = 0, . . . , s − 1

and

⎨0 s
if end HG(b) (G(b)) ≥ 0,
s t +
end HG(bt ) (G(b )) =
+
⎩−1 if end H s
G(b)+ (G(b)) < 0.

i t
Of course, HG(b t ) (G(b )) = 0 for all i > s, by 18.2.2 and 18.3.5. Hence
+

⎨s s
if end HG(b) (G(b)) ≥ 0,
+
regs (G(bt )) = reg(G(bt )) =
⎩s − 1 if end H s
G(b)+ (G(b)) < 0.

The result now follows from Trung’s Theorem 18.3.12.


18.3.15 Exercise. Assume that (R, m) is local and that R/m is infinite; let
dim R = d. Suppose that q is an m-primary ideal of R such that G(q) is
Cohen–Macaulay. Show that each minimal reduction q of q has rq (q) =
d
end HG(q)+
(G(q)) + d.

As was mentioned in the introduction to this chapter, we have only been able
to present a small portion of the body of work linking graded local cohomology
and reductions of ideals. An interested reader might like to consult, in addition
to papers already cited in this chapter, [48] by T. Marley, [87] by N. V. Trung,
[31] by M. Herrmann, E. Hyry and T. Korb, [9] by C. D’Cruz, V. Kodiyalam
and J. K. Verma, [12] by J. Elias, and some of the papers cited by these authors.
19

Connectivity in algebraic varieties

The study of the topological connectivity of algebraic sets is a fundamental


subject in algebraic geometry. Local cohomology is a powerful tool in this
field. In this chapter we shall use this tool to prove some results on connectiv-
ity which are of basic significance. Our main result will be the Connectedness
Bound for Complete Local Rings, a refinement of Grothendieck’s Connect-
edness Theorem. We shall apply this result to projective varieties in order to
obtain a refined version of the Bertini–Gothendieck Connectivity Theorem.
Another central result of this chapter will be the Intersection Inequality for
Connectedness Dimensions of Affine Algebraic Cones. As an application it
will furnish a refined version of the Connectedness Theorem for Projective Va-
rieties due to W. Barth, to W. Fulton and J. Hansen, and to G. Faltings. The final
goal of the chapter will be a ring-theoretic version of Zariski’s Main Theorem
on the Connectivity of Fibres of Blowing-up.
The crucial appearances of local cohomology in this chapter are just in two
proofs, but the resulting far-reaching consequences in algebraic geometry il-
lustrate again the power of local cohomology as a tool in the subject. We shall
use little more from local cohomology than the Mayer–Vietoris sequence 3.2.3
and its graded version 14.1.5, the Lichtenbaum–Hartshorne Vanishing Theo-
rem 8.2.1 and the graded version 14.1.16, and the vanishing result of 3.3.3.
See the proofs of Proposition 19.2.8 and Lemma 19.7.2. The use of these tech-
niques in this context originally goes back to Hartshorne [29] and has been
pushed further by J. Rung (see [5]).
Throughout this chapter, all graded rings and modules are to be understood
to be Z-graded, and all polynomial rings K[X1 , . . . , Xd ] (over a field K) are
to be understood to be (positively) Z-graded so that K is the component of
degree 0 and deg Xi = 1 for all i = 1, . . . , d.
406 Connectivity in algebraic varieties

19.1 The connectedness dimension


To begin, we have to introduce a measure for the connectivity of an algebraic
set, or, more generally, of a Noetherian topological space. We start with some
reminders of topological concepts which are fundamental to our work in this
chapter.
19.1.1 Reminders. Here we are concerned with a general topological space.
(i) Recall that a non-empty topological space is said to be disconnected
precisely when it can be expressed as the disjoint union of two proper
open (or closed) subsets. Otherwise the space is said to be connected.
We adopt the convention whereby the empty set is considered to be dis-
connected.
(ii) Recall also that a topological space T is said to be quasi-compact pre-
cisely when every open covering of T has a finite subcovering, that
is, if and only if, whenever (Uα )α∈Λ is a family of open subsets of T

such that T = α∈Λ Uα , then there is a finite subset Φ of Λ such that

T = α∈Φ Uα .
(iii) Recall that a non-empty topological space T is said to be irreducible
precisely when T is not the union of two proper closed subsets, that is,
if and only if every pair of non-empty open subsets of T has non-empty
intersection.
19.1.2 Exercise. Let T be a non-empty topological space.
(i) Show that the following statements are equivalent:
(a) T is irreducible;
(b) every non-empty open subset of T is dense in T ;
(c) every non-empty open subset of T is connected.
(ii) Let S be an irreducible subset of T (that is, a subset of T which is an
irreducible space in the topology induced from T ), and let (Ci )1≤i≤n
be a finite covering of S by closed subsets of T (so that C1 , . . . , Cn are
n
closed subsets of T such that S ⊆ i=1 Ci ). Show that S ⊆ Ci for some
i with 1 ≤ i ≤ n.
(iii) Show that, if T is irreducible, then every non-empty open subset of T is
irreducible.
(iv) Let (Ui )1≤i≤n be a finite open covering of T , with Ui = ∅ for all i =
1, . . . , n. Prove that T is irreducible if and only if Ui is irreducible for
all i = 1, . . . , n and Ui ∩ Uj = ∅ for all i, j = 1, . . . , n.
(v) Let S be a subset of T . Show that S is irreducible if and only if its closure
S is irreducible.
19.1 The connectedness dimension 407

(vi) Let T  be a second topological space and let f : T → T  be a continuous


map. Show that, if T is irreducible, then so too is f (T ).

19.1.3 Exercise and Definition. Use Zorn’s Lemma to show that a non-
empty topological space T has maximal irreducible subsets.
The maximal irreducible subsets of T are called its irreducible components.
Show that the irreducible components of T are closed and that they cover
T ; show also that every irreducible subset of T is contained in an irreducible
component of T .

19.1.4 Exercise. Let K be an algebraically closed field, and let r ∈ N.

(i) Show that the irreducible components of an affine algebraic cone in


Ar (K) (see 15.1.2(i)) are again affine algebraic cones.
(ii) Let W be a non-empty closed subset of Pr (K); suppose that the distinct
irreducible components of W are W1 , . . . , Wn . Show that Cone(W )
(see 15.2.1(iii)) has Cone(W1 ), . . . , Cone(Wn ) as its (distinct) irreduci-
ble components.

19.1.5 Definition. Let T be a topological space. We say that T is a Noethe-


rian topological space precisely when it satisfies the following equivalent con-
ditions.

(i) Whenever (Ci )i∈N is a family of closed subsets of T such that

C1 ⊇ C2 ⊇ · · · ⊇ Ci ⊇ Ci+1 ⊇ · · · ,

then there exists k ∈ N such that Ck = Ck+i for all i ∈ N.


(ii) Every non-empty set of closed subsets of T contains a minimal element
with respect to inclusion.

19.1.6 Exercise. Let T be a topological space.

(i) Show that T is Noetherian if and only if every open subset of T is quasi-
compact.
(ii) Show that the spectrum of a commutative Noetherian ring, furnished
with its Zariski topology, is a Noetherian topological space.
(iii) Show that a quasi-affine variety over an algebraically closed field (see
2.3.1) is a Noetherian topological space.

19.1.7 Lemma. Let T be a non-empty Noetherian topological space. Then


408 Connectivity in algebraic varieties

T has only finitely many irreducible components. Also, if T1 , . . . , Tn are the


distinct irreducible components of T , then

n
Tj ⊆ Ti for all j = 1, . . . , n.
i=1
i=j

Proof. Suppose that T has infinitely many irreducible components. Let S


be the set of non-empty closed subsets of T which have infinitely many irre-
ducible components. Since T is Noetherian, S has a minimal member: let C
be one such. Then C itself cannot be irreducible, so that C can be written as
C = C1 ∪ C2 for some proper closed subsets C1 and C2 of C. Note that C1
and C2 are closed in T , and so, by the minimality of C, each of C1 and C2 has
only finitely many irreducible components. But, by 19.1.2(ii), each irreducible
subset of C must be contained in C1 or C2 , and so each irreducible component
of C must be an irreducible component of C1 or C2 (by 19.1.3). Hence there
can only be finitely many irreducible components of C, a contradiction.
The final claim follows easily from another use of 19.1.2(ii) since, by 19.1.3,
each Ti (1 ≤ i ≤ n) is closed in T .
19.1.8 Definition. Let T be a non-empty Noetherian topological space. The
dimension of T , denoted by dim T , is defined as the supremum of the lengths
n of all strictly descending chains Z0 ⊃ Z1 ⊃ · · · ⊃ Zn of closed irreducible
subsets of T if this supremum exists, and ∞ otherwise. Thus, by 19.1.3, dim T
is a non-negative integer or ∞.
The dimension of the empty space is defined to be −1.
Note that, for our Noetherian ring R, we have dim(Spec(R)) = dim R;
thus, in view of [56, Appendix, Example 1], the dimension of a Noetherian
topological space can be ∞.
19.1.9 Definition. Let T be a Noetherian topological space. The connected-
ness dimension c(T ) of T is defined to be the minimum of the dimensions of
those closed subsets Z of T for which T \ Z is disconnected. (Observe that
T \ T is certainly disconnected!) Thus
c(T ) := min {dim Z : Z ⊆ T, Z is closed and T \ Z is disconnected} .

For our Noetherian ring R, we write c(R) := c(Spec(R)).


19.1.10 Examples. Let T be a Noetherian topological space.

(i) Note that T is disconnected, that is, T \ ∅ is disconnected, if and only if


c(T ) = −1. We can thus conclude that c(T ) is negative if and only if T
is disconnected.
19.1 The connectedness dimension 409

(ii) It follows from part (i) that, if (R, m) is local, then c(R) ≥ 0.
(iii) Suppose that T is irreducible. Then if Z is a proper closed subset of T ,
it is impossible for T \ Z to be disconnected (by 19.1.2(i)); on the other
hand, T \ T is disconnected. Thus c(T ) = dim T in this case.
(iv) It follows from part (iii) that, if R is an integral domain, then c(R) =
dim R. Thus, in view of [56, Appendix, Example 1], the connectedness
dimension of a Noetherian topological space can be ∞.

19.1.11 Example. Assume that (R, m) is local, and has exactly two minimal
prime ideals, p and q. Then c(R) = 0 if and only if dim R/(p + q) = 0.

Proof. Note that the hypotheses ensure that dim R > 0.


(⇐) We have Spec(R) = Var(p) ∪ Var(q). Since m is the only prime ideal
of R which contains both p and q, it follows that

Var(p) ∩ (Spec(R) \ {m}) and Var(q) ∩ (Spec(R) \ {m})

are two non-empty disjoint closed subsets of Spec(R) \ {m} which cover this
space. Therefore, bearing in mind 19.1.10(ii), we see that c(R) = 0.
(⇒) Assume that c(R) = 0. Thus there is a closed subset Z of Spec(R) for
which dim Z = 0 and Spec(R) \ Z is disconnected. We must have Z = {m}.
Set T := Spec(R)\{m}. Thus there exist ideals a, b of R such that T ∩Var(a)
and T ∩ Var(b) are non-empty disjoint subsets of T which cover T .
Then p ⊇ a or p ⊇ b; also q ⊇ a or q ⊇ b. For the sake of argument, let us
assume that p ⊇ a.
Then q ⊇ a, since otherwise Spec(R) = Var(a) and T ∩ Var(b) = ∅.
Therefore q ⊇ b, and it follows that (T ∩ Var(p)) ∩ (T ∩ Var(q)) = ∅, so that
dim R/(p + q) = 0.

19.1.12 Exercise. Let f , g be non-constant and irreducible polynomials in


C[X, Y ] which are not associates of each other. Let T := VA2 (f g). Show that

−1 if VA2 (f ) ∩ VA2 (g) = ∅,
c(T ) =
0 otherwise.

19.1.13 Notation and Exercise. Assume that (R, m) is local. The topologi-
cal space Spec(R) \ {m}, with the topology induced from the Zariski topology
on Spec(R), is called the punctured spectrum of R, and denoted by Spec◦ (R).
◦ (R)) = c(R) − 1.
Show that c(Spec

19.1.14 Notation. For r ∈ N, denote by S(r) the set of all ordered pairs
(A, B) of non-empty subsets of {1, . . . , r} for which A ∪ B = {1, . . . , r}.
410 Connectivity in algebraic varieties

19.1.15 Lemma. Let T be a non-empty Noetherian topological space with


(distinct) irreducible
 components T1 , . . . , Tr . With the notation of 19.1.14,
 we
 
have c(T ) = min dim T
i∈A i ∩ j∈B jT : (A, B) ∈ S(r) .

Proof. We write c = c(T ) and m for the minimum that occurs on the right-
hand side of the equation in the above statement.
We first show that c ≥ m. To achieve this, let Z be a closed subset of T
with dim Z = c and such that T \ Z is disconnected. If Ti ⊆ Z for some
i ∈ {1, . . . , r}, then c = dim Z ≥ dim Ti ≥ m. Thus we can, and do, assume
that Ti ∩ (T \ Z) = ∅ for all i ∈ {1, . . . , r}. As T \ Z is disconnected we can
write T \ Z = U1 ∪ U2 , where U1 and U2 are non-empty open sets in T such
that U1 ∩ U2 = ∅. Set

A := {i ∈ {1, . . . , r} : Ti ∩ U1 = ∅} ,

B := {j ∈ {1, . . . , r} : Tj ∩ U2 = ∅} .

Then, the pair (A, B) ∈ S(r). Moreover, A and B are disjoint, as otherwise for
any index i ∈ A ∩ B the irreducible space Ti would contain the two non-empty
 
and disjoint open subsets Ti ∩ U1 and Ti ∩ U2 . Thus i∈A Ti ∩ j∈B Tj
has no point in common with U1 ∪ U2 , so that it is contained in Z and has
dimension not exceeding c. This proves that c ≥ m.
To prove the inequality m ≥ c, let (A, B) ∈ S(r) be a pair such that Z :=
 
i∈A Ti ∩ j∈B Tj is of dimension m. If r = 1, we have m = dim T ≥
c, as required. Assume therefore that r > 1. Then, by the minimality in the
definition of m, we can, and do, assume that A and B are disjoint. Then the
 
open sets U1 = T \ i∈A Ti and U2 = T \ j∈B Tj are non-empty. Note that
 
U1 ∩ U2 = ∅, since i∈A Ti ∪ j∈B Tj = T . Hence T \ Z = U1 ∪ U2
is disconnected. Consequently c ≤ dim Z = m.

19.1.16 Exercise. Let V be an affine variety over the algebraically closed


field K. Let b be an ideal of O(V ), and let V (b) denote the closed subset of V
determined by b. Show that c(V (b)) = c(O(V )/b).

19.2 Complete local rings and connectivity


We now introduce another invariant of Noetherian spaces.

19.2.1 Definition. Let T be a non-empty Noetherian topological space. The


subdimension sdim T of T is defined as the minimum of the dimensions of
19.2 Complete local rings and connectivity 411

the irreducible components of T . For our Noetherian ring R, we write sdim R


instead of sdim(Spec R).

Notice the following easy fact.

19.2.2 Lemma. Let T be a non-empty Noetherian topological space of finite


dimension. Then c(T ) ≤ sdim T . Moreover, equality holds here if and only if
T is irreducible.

Proof. Let T1 , . . . , Tr be the irreducible components of T . Then

sdim T = min {dim Ti : i = 1, . . . , r} ;

also, we have, by 19.1.7, for each j = 1, . . . , r, that


⎛ ⎞
⎜r

Tj ∩ ⎜ ⎟
⎝ Ti ⎠ ⊂ T j ,
i=1
i=j

so that, since T has finite dimension,


⎛ ⎛ ⎞⎞
⎜ ⎜ 
r
⎟⎟
dim ⎜ ⎜
⎝Tj ∩ ⎝ Ti ⎟ ⎟
⎠⎠ < dim Tj .
i=1
i=j

The claims now follow easily from Lemma 19.1.15.

19.2.3 Exercise. Calculate c(T ) and sdim T for the following choices of the
Noetherian topological space T :

(i) T := VA3 (X12 − X1 , X22 − X2 );


(ii) T := VA4 (X1 X3 , X1 X4 , X2 X3 , X2 X4 );
(iii) T := VA3 ((X1 − 1)(X12 + X22 + X32 − 1)).

19.2.4 Exercise. Assume that (R, m) is local and that dim R > 0. Show
◦ (R) of R (see 19.1.13)
that the subdimension of the punctured spectrum Spec

is given by sdim(Spec(R)) = sdim R − 1.

19.2.5 Remark. We shall frequently consider connectedness dimensions and


subdimensions of spectra of Noetherian rings. Therefore it will be helpful
to translate the previous lemmas into ring-theoretic terms. So, let p1 , . . . , pr
be the distinct minimal prime ideals of R. Then, again using the notation of
19.1.14, it follows from 19.1.15 that
 2 
c(R) = min dim R i∈A p i + j∈B p j : (A, B) ∈ S(r) .
412 Connectivity in algebraic varieties

Also, 19.2.2 shows that sdim R = min {dim R/pi : i = 1, . . . , r} ≥ c(R),


with equality if and only if r = 1.

19.2.6 Exercise. Suppose that (R, m) is local, catenary and S2 , and that R
has more than one minimal prime ideal; set d := dim R. Use Proposition
12.2.8 to show that c(R) = d − 1.

Our applications of local cohomology to connectedness dimensions will in-


volve use of the concepts of arithmetic rank and cohomological dimension of
an ideal, introduced in 3.3.2 and 3.3.4. Recall that cohd(a) ≤ ara(a). The
reader should be aware of the elementary properties of arithmetic rank de-
scribed in the following exercise.

19.2.7 Exercise. Assume that the ideal a is proper.

(i) Show that ht a ≤ ara(a).


(ii) Let R be a second commutative Noetherian ring and let f : R −→ R
be a ring homomorphism. Prove that ara(aR ) ≤ ara(a).

The next result will play a crucial rôle in our approach to connectivity. Its
proof uses most of the main ingredients from local cohomology theory that
we shall need in this chapter. The result relates, in certain circumstances, the
cohomological dimension cohd(a ∩ b) of the intersection of two ideals a and
b in a complete local domain R with the dimensions of R and R/(a + b).

19.2.8 Proposition. Assume that (R, m) is a complete local domain. Let b


be a second ideal of R, and assume that a and b are both proper and that
min {dim R/a, dim R/b} > dim R/(a + b). Then
cohd(a ∩ b) ≥ dim R − dim R/(a + b) − 1,
so that ara(a ∩ b) ≥ dim R − dim R/(a + b) − 1.

Proof. Set d := dim R and δ := dim R/(a + b). We proceed by induction


on δ. First, let δ = 0. Then, we have to show that cohd(a ∩ b) ≥ d − 1. By the
Mayer–Vietoris sequence 3.2.3, there is an exact sequence
d−1
Ha∩b (R) −→ Ha+b
d
(R) −→ Had (R) ⊕ Hbd (R).
It follows from the local Lichtenbaum–Hartshorne Vanishing Theorem 8.2.1
that Had (R) = Hbd (R) = 0. As δ = 0, the ideal a + b is m-primary, and so
d d−1
Ha+b (R) = 0 by 1.2.3 and 6.1.4. Altogether we obtain that Ha∩b (R) = 0,
and so cohd(a ∩ b) ≥ d − 1. This proves the claim when δ = 0.
So, let δ > 0 and make the obvious inductive assumption. Set cohd(a∩b) =:
r. As a + b is not m-primary we can find an element y ∈ m which lies outside
19.2 Complete local rings and connectivity 413

all the minimal prime ideals of a, b and a + b. We write a = a + Ry and


b = b + Ry and note that dim R/(a + b ) = δ − 1,

dim R/a = dim R/a − 1 > δ − 1 and dim R/b = dim R/b − 1 > δ − 1.

As
√  
a ∩ b = (a + Ry) ∩ (b + Ry) = (a ∩ b) + Ry,

we have cohd(a ∩ b ) ≤ cohd(a ∩ b) + 1 = r + 1 by 8.1.3. Therefore, by the


inductive hypothesis, we have r + 1 ≥ d − (δ − 1) − 1; hence r ≥ d − δ − 1.
This completes the inductive step, and the proof.

As an application of this we can now prove the following.

19.2.9 Lemma. Assume that (R, m) is a complete local ring. Let b be a


second ideal of R, and assume that a and b are both proper and that

min {dim R/a, dim R/b} > dim R/(a + b).

Then

dim R/(a + b) ≥ min {c(R), sdim R − 1} − cohd(a ∩ b),

so that dim R/(a + b) ≥ min {c(R), sdim R − 1} − ara(a ∩ b).

Proof. Set δ := dim R/(a + b). Let p1 , . . . , pn be the distinct minimal prime
ideals of R.
First, we treat the case where, for all i ∈ {1, . . . , n}, either

dim R/(a + pi ) ≤ δ or dim R/(b + pi ) ≤ δ.

After an appropriate reordering of the pi , there will be an s ∈ N0 such that


s ≤ n and dim R/(a + pi ) ≤ δ for 1 ≤ i ≤ s and dim R/(b + pj ) ≤ δ for
s + 1 ≤ j ≤ n. As max {dim R/(a + pk ) : 1 ≤ k ≤ n} = dim R/a > δ, we
see that s < n. As max {dim R/(b + pk ) : 1 ≤ k ≤ n} = dim R/b > δ, we
see that 1 ≤ s.
Now, let p be a minimal prime ideal of the ideal

c := (p1 ∩ · · · ∩ ps ) + (ps+1 ∩ · · · ∩ pn )

such that dim R/p = dim R/c. By 19.2.5, we have dim R/p ≥ c(R). More-
over we can choose indices i and j with 1 ≤ i ≤ s < j ≤ n and such that
pi , pj ⊆ p. It follows that

δ ≥ dim R/(a + pi ) ≥ dim R/(a + p)


414 Connectivity in algebraic varieties

and δ ≥ dim R/(b + pj ) ≥ dim R/(b + p); hence

δ ≥ dim R/((a + p) ∩ (b + p)) = dim R/((a ∩ b) + p).

As R/p is catenary (see [50, Theorem 29.4(ii)]), we can write

dim R/((a ∩ b) + p) = dim R/p − ht((a ∩ b) + p)/p.

As dim R/p ≥ c(R) and

ht((a ∩ b) + p)/p ≤ cohd(((a ∩ b) + p)/p) ≤ cohd(a ∩ b)

(by 6.1.6 and 4.2.3), we thus obtain that δ ≥ c(R) − cohd(a ∩ b).
Therefore, it remains for us to treat the case in which there exists i ∈
{1, . . . , n} such that

dim R/(a + pi ) > δ and dim R/(b + pi ) > δ.

As dim R/(a + b + pi ) =: δ  ≤ δ, it follows from 19.2.8 that

δ ≥ δ  ≥ dim R/pi − cohd (((a + pi ) ∩ (b + pi ))/pi ) − 1.

Observing that dim R/pi ≥ sdim R and (by 4.2.3)

cohd (((a + pi ) ∩ (b + pi ))/pi ) = cohd (((a ∩ b) + pi )/pi ) ≤ cohd(a ∩ b),

we thus deduce that δ ≥ sdim R − 1 − cohd(a ∩ b).

Note. In the First Edition of this book, only the statements involving arith-
metic rank appeared in the results corresponding to 19.2.8 and 19.2.9. We are
very grateful to M. Varbaro for pointing out to us that those statements can be
strengthened by replacement of ‘arithmetic rank’ by ‘cohomological dimen-
sion’. We have left to the interested reader the formulation of similar strength-
enings of some subsequent results in this chapter, such as 19.2.10, 19.2.11 and
19.2.12.
We are now in a position to prove the first main result of this chapter, namely
the Connectedness Bound for Complete Local Rings.

19.2.10 Connectedness Bound for Complete Local Rings. Suppose that


(R, m) is a complete local ring, and let a be proper. Then

c(R/a) ≥ min {c(R), sdim R − 1} − ara(a).



Proof. Without loss of generality we can, and do, assume that a = a. Let
p1 , . . . , pn be the distinct minimal prime ideals of a, and set c := c(R/a).
If n = 1, we have a = p1 and c = dim R/p1 by 19.2.5. Choose a minimal
19.2 Complete local rings and connectivity 415

prime ideal p of R with p ⊆ p1 , and observe that ht p1 /p ≤ cohd(a) by 4.2.3


and 6.1.6. Since R is catenary, we deduce that

c = dim R/p − ht p1 /p ≥ sdim R − cohd(a),

from which our claim follows (since cohd(a) ≤ ara(a)).


Consider now the case where n > 1. By 19.2.5, there exist two non-empty
subsets A, B of {1, . . . , n} for which A ∪ B = {1, . . . , n} and
2
c = dim R i∈A pi + j∈B pj ;

moreover, we can, and do, assume that A and B are disjoint. Set r := i∈A pi
and s := j∈B pj ; then dim R/r > c and dim R/s > c (by the final comment
of 19.2.5), and r ∩ s = a. We can now use 19.2.9 to complete the proof.

19.2.11 Corollary. Let (R, m) and a be as in 19.2.10. Then

c(R/a) ≥ c(R) − ara(a) − 1.

If R has more than one minimal prime ideal, then the inequality is strict.

Proof. By 19.2.5, we have c(R) ≤ sdim R, with strict inequality if R has


more than one minimal prime ideal. The claim therefore follows from 19.2.10.

As another application of 19.2.10 we now prove Grothendieck’s Connected-


ness Theorem.

19.2.12 Grothendieck’s Connectedness Theorem. (See [26, Exposé XIII,


Théorème 2.1].) Assume that (R, m) is a complete local ring, and let a be
proper. Let k ∈ N0 be such that c(R) ≥ k and sdim R ≥ k+1. Then c(R/a) ≥
k − ara(a).

Proof. By 19.2.10, we have

c(R/a) ≥ min {c(R), sdim R − 1} − ara(a) ≥ k − ara(a).

19.2.13 Remark. Let the situation be as in 19.2.12. In Grothendieck’s origi-


nal version of that result, connectivity and subdimension are considered on the
punctured spectrum Spec ◦ (R) of R (see 19.1.13). However, since

◦ (R)) = c(R) − 1
c(Spec and ◦ (R)) = sdim R − 1
sdim(Spec

(by 19.1.13 and 19.2.4), Grothendieck’s version can be recovered from ours.
416 Connectivity in algebraic varieties

19.3 Some local dimensions


Up to now, we have obtained a certain understanding of the connectivity in the
spectrum of a complete local Noetherian ring. In order to apply this knowledge
to the non-complete case, we prove the following lemma. The reader is warned
that our proof of part (iv) of the lemma uses L. J. Ratliff’s Theorem [67] (see
also [50, Theorem 31.7]) that a local ring (R, m) is universally catenary only
if, for every p ∈ Spec(R) and for every minimal prime P of the ideal pR  of
 
R, we have dim R/P = dim R/p.

19.3.1 Lemma. Assume that (R, m) is local. The following hold:



(i) c(R) ≥ c(R);
  for all minimal prime ideals p of R, then equality holds
(ii) if pR ∈ Spec(R)
in (i);
(iii) sdim R ≥ sdim R;  and
(iv) if R is universally catenary, then equality holds in (iii).

Proof. Let P be the set of minimal prime ideals of R and Q be the set of
 Note that, by [50, Theorem 7.3(i)] for example,
minimal prime ideals of R.
each p ∈ P is the contraction to R of some member of Q. Also, Q ∩ R ∈ P
for all Q ∈ Q (by [50, Theorem 15.1(ii)], for example).
(i) By 19.2.5, we can find two non-empty subsets P1 and P2 of P such that
P1 ∪ P2 = P and, if a1 := p∈P1 p and a2 := p∈P2 p, then

c := c(R) = dim R/(a1 + a2 ).

Let Qi := {Q ∈ Q : Q ∩ R ∈ Pi } for i = 1, 2; then Q1 and Q2 are non-


empty and such that Q1 ∪ Q2 = Q. Let Bi = Q∈Qi Q (i = 1, 2). Then, by
 ≤ dim R/(B
19.2.5, we have c(R)  
1 + B2 ). Since ai R ⊆ Bi for i = 1, 2, it
 ≤ dim R/(B
follows that c(R)    
1 + B2 ) ≤ dim R/(a1 R + a2 R) = c.
 ∈ Spec(R)
(ii) Suppose that pR  for all p ∈ P. By 19.2.5, we can find
two non-empty subsets Q3 and Q4 of Q such that Q3 ∪ Q4 = Q and, if
Bi = Q∈Qi Q for i = 3, 4, then   = dim R/(B
c := c(R)  3 + B4 ).
Set Pi := {Q ∩ R : Q ∈ Qi } and ai = p∈Pi p for i = 3, 4. Since each
p ∈ P is the contraction to R of some minimal prime ideal of R,  we have

P3 ∪ P4 = P. In particular, we have (Q ∩ R)R = Q for all Q ∈ Q, and so
 = Bi for i = 3, 4. Hence, by
we can use [50, Theorem 7.4(ii)] to see that ai R
19.2.5,

 
c = dim R/(B   
3 + B4 ) = dim R/(a3 R + a4 R) = dim R/(a3 + a4 ) ≥ c(R).

The claim follows from this and part (i).


19.3 Some local dimensions 417

(iii) Let p ∈ P be such that dim R/p = sdim R. Now there exists Q ∈ Q
such that Q ∩ R = p. Then
 ≤ dim R/Q
sdim R   R
≤ dim R/p  = dim R/p = sdim R.

(iv) Assume now that R is universally catenary. Let Q ∈ Q be such that


sdim R = dim R/Q.
 Now p := Q ∩ R ∈ P; by Ratliff’s Theorem [50,

Theorem 31.7], we have dim R/Q = dim R/p. Hence
 = dim R/Q
sdim R  = dim R/p ≥ sdim R.
The claim follows from this and part (iii).
19.3.2 Definitions. Let T be a Noetherian topological space, and let p ∈ T .
The local dimension of T at p, denoted by dimp T , is defined as the supre-
mum of the lengths of all strictly descending chains Z0 ⊃ Z1 ⊃ · · · ⊃ Zn
of closed irreducible subsets of T which all contain p if this supremum exists,
and ∞ otherwise. Thus, by 19.1.3, dimp T is a non-negative integer or ∞.
Note that, for p ∈ Spec(R), we have dimp Spec(R) = dim Rp .
Let V be an affine variety over the algebraically closed field K. Since, in
an integral domain R which is a finitely generated K-algebra, every maximal
ideal has height equal to dim R (see, for example, [81, 14.33]), it follows that
dimq V = dim OV,q = dim O(V ) = dim V for all q ∈ V .
If T1 , . . . , Tr are the irreducible components of T which contain p, we write
T p := T1 ∪ · · · ∪ Tr , and call this subspace T p of T the p-component of T .
The local connectedness dimension of T at p, denoted by cp (T ), is defined
to be the minimum of the local dimensions at p of those closed subsets Z of the
p-component T p of T which contain p and for which T p \ Z is disconnected.
Thus
cp (T ) := min {dimp Z : Z ⊆ T p , p ∈ Z, Z is closed
and T p \ Z is disconnected} .
The local subdimension sdimp T of T at p is defined as the minimum of the
local dimensions at p of the irreducible components of T which contain p. (Of
course, p does belong to at least one irreducible component of T , by 19.1.3.)
19.3.3 Exercise. Let T be a Noetherian topological space and let p ∈ T .
(i) Show that dimp Z = dimp (T p ∩ Z) and sdimp Z = sdimp (T p ∩ Z) for
each closed subset Z of T for which p ∈ Z.
(ii) Let T1 , . . . , Tr be the (distinct) irreducible components of T that contain
p. Show that, with the notation of 19.1.14,
   
cp (T ) = min dimp i∈A T i ∩ j∈B T j : (A, B) ∈ S(r) .
418 Connectivity in algebraic varieties

(Use the ideas of the proof of Theorem 19.1.15.)


The next exercise provides justification for the appearance of the word
‘local’ in the definitions in 19.3.2.
19.3.4 Exercise. Let T be a Noetherian topological space and let p ∈ T ; let
U be an open subset of T which contains p. Show that

(i) Up = Tp ∩ U;
(ii) dimp T = dimp U ;
(iii) sdimp T = sdimp U ; and
(iv) cp (T ) = cp (U ).

(Use Exercise 19.1.2(i),(iii),(v) to establish the existence of a bijection be-


tween the set of irreducible closed subsets of T containing p and the set of
irreducible closed subsets of U containing p; then use Exercise 19.3.3(ii).)
19.3.5 Exercise. Let p ∈ V := VA3 (X1 X2 X3 ). Calculate cp (V )

(i) when p = (0, 0, 0);


(ii) when exactly two of the co-ordinates of p are 0; and
(iii) in all other cases.
19.3.6 Exercise. Let p ∈ Spec(R). Show that cp (Spec(R)) = c(Rp ) and
sdimp (Spec(R)) = sdim(Rp ).
We wish to study the connectivity of varieties over an algebraically closed
field. We remind the reader about some elementary facts concerning such va-
rieties.
19.3.7 Reminders. Let K be an algebraically closed field, and let r ∈ N. Let
V ⊆ Pr (K) be a quasi-projective variety.

(i) Regard the polynomial ring K[X0 , X1 , . . . , Xr ] as the coordinate ring


O(Ar+1 (K)), and let i ∈ {0, . . . , r}. Let Ui Pr (K) denote the open
subset of Pr (K) given by
Ui Pr (K) = Pr (K) \ VPr (K) (Xi )
= {(c0 : · · · : ci : · · · : cr ) ∈ Pr (K) : ci = 0} .

=
There is an isomorphism of varieties σi : Ar (K) −→ Ui Pr (K) given
by σi ((a1 , . . . , ar )) = (a1 : · · · : ai : 1 : ai+1 : · · · : ar ) for all
r
(a1 , . . . , ar ) ∈ Ar (K). Note that Pr (K) = j=0 Uj Pr (K); it follows
that the quasi-projective variety V ⊆ Pr (K) has a finite covering by
open sets each of which is quasi-affine (in the sense that it is isomorphic
19.3 Some local dimensions 419

to a quasi-affine variety over K). As any quasi-affine variety has a finite


covering by open sets which are affine, we thus see that quasi-projective
varieties also have finite coverings by affine open sets.
The above isomorphisms show that affine varieties, and also quasi-
affine varieties, are quasi-projective: the reader should remember that,
for us, the word ‘variety’, as introduced in 15.2.1(ii), is synonymous with
‘quasi-projective variety’ (and does not mean the same as ‘(abstract) va-
riety’ in the sense of Hartshorne [30, p. 105]).
(ii) Let p ∈ V . By part (i), there is an open subset U of V such that p ∈ U
and U is an affine variety. We can, and do, identify the local ring OV,p
of p on V with OU,p = O(U )IU (p) (see 6.4.1). Let C be a closed subset
of V such that p ∈ C. The vanishing ideal IU (C ∩ U ) was defined in
15.1.2. Recall that the local vanishing ideal IV,p (C) of C at p is the
(radical) ideal of OV,p consisting of all germs of regular functions f ∈
O(W ) defined on some open neighbourhood W of p in V and such that
f (W ∩ C) = 0; hence

IV,p (C) = IU,p (C ∩ U ) = IU (C ∩ U )O(U )IU (p) = IU (C ∩ U )OU,p .

19.3.8 Exercise. Let K be an algebraically closed field.

(i) Let V be an affine variety over K. Let b be a proper ideal of O(V ), and
let V (b) denote the closed subset of V determined by b; let p ∈ V (b).
Show that cp (V (b)) = c(OV,p /bOV,p ) and

sdimp V (b) = sdim(OV,p /bOV,p )


= min {dim O(V )/q : q is a minimal prime of b
such that p ∈ V (q)} .

(ii) Let V  be a variety over K. Let W be a non-empty closed subset of V  ,


and let p ∈ W . Show that, with the notation of 19.3.7(ii), sdimp W =
sdim(OV  ,p /IV  ,p (W )) and cp (W ) = c(OV  ,p /IV  ,p (W )). (Use part
(i) and 19.3.4.)

19.3.9 Definition. Let V be a variety over the algebraically closed field K.


Let W be a non-empty closed subset of V , and let p ∈ W . We define the
formal connectedness dimension of W at p, denoted 
cp (W ), by


cp (W ) = c O
 
V,p /IV,p (W )OV,p = c ((OV,p /IV,p (W ))) .

Here again, IV,p (W ) is the local vanishing ideal of W at p: see 19.3.7(ii).


420 Connectivity in algebraic varieties

19.3.10 Remark. Let V , W and p be as in 19.3.9. Then it follows from


19.3.1(i) and 19.3.8(ii) that cp (W ) ≥ 
cp (W ). Moreover, if V is affine and b is
any ideal of O(V ) for which W = V (b), then  cp (W ) = c ((OV,p /bOV,p )).
19.3.11 Exercise. Let
V1 := VA2 (X 3 − Y 2 ) and V2 := VA2 (X 2 + X 3 − Y 2 );
also let
V := VA4 (X1 X4 − X2 X3 , X12 X3 + X1 X2 − X22 , X33 + X3 X4 − X42 ),
as in Example 2.3.7.
(i) Calculate c(0,0) (Vi ) and 
c(0,0) (Vi ) for i = 1, 2.
(ii) Show that c(0,0,0,0) (V ) = 0. (You might find 8.2.13(iv) and 19.1.11
helpful.)

19.3.12 Definition. Let V be a variety over the algebraically closed field K.


Let W and Z be closed subsets of V with Z ⊆ W , and let p ∈ Z. Note that
IV,p (W ) ⊆ IV,p (Z).
We define the (local) arithmetic rank of Z at p with respect to W , denoted
araW,p (Z), by
araW,p (Z) = ara (IV,p (Z)/IV,p (W )) .
(Here, IV,p (Z)/IV,p (W ) is considered as an ideal of OV,p /IV,p (W ).)
19.3.13 Definition. Let V be a variety over the algebraically closed field K.
Let W be a non-empty closed subset of V , and let p ∈ W . We extend the
terminology of 8.2.15 and say that W is analytically reducible at p precisely
when O 
V,p /IV,p (W )OV,p has more than one minimal prime ideal. Otherwise,
W is said to be analytically irreducible at p. Note that O  ∼
V,p /IV,p (W )OV,p =
(OV,p /IV,p (W )), and by [50, Theorem 32.2(i) and p. 259, Remark 1], W is
analytically irreducible at p if and only if (OV,p /IV,p (W )) is a domain.
19.3.14 Remark. Note that, in the situation of 19.3.9, 19.3.12 and 19.3.13,
the invariants 
cp (W ) and araW,p (Z), and the notion (for W ) of analytical re-
ducibility at p, do not depend on the ambient variety V , and, indeed, remain
unchanged if V , W and Z are replaced, respectively, by U , W ∩ U and Z ∩ U ,
where U is any open subset of V containing p.
We can now deduce the following from 19.2.10 and 19.2.11.
19.3.15 Proposition. Let V be a variety over the algebraically closed field
K. Let W and Z be closed subsets of V with Z ⊆ W , and let p ∈ Z. Then
19.3 Some local dimensions 421

cp (Z) ≥ min {
(i)  cp (W ), sdimp W − 1} − araW,p (Z);
cp (Z) ≥ 
(ii)  cp (W ) − araW,p (Z) − 1; and
(iii) if W is analytically reducible at p, then the inequality in part (ii) is
strict.

Proof. Let R := OV,p /IV,p (W ) and a := IV,p (Z)/IV,p (W ). Then


2
  and 
cp (W ) = c(R) cp (Z) = c O    
V,p IV,p (Z)OV,p = c(R/aR).

Next, araW,p (Z) = ara(a) (see 19.3.12), and so it follows from 19.2.7(ii) that
araW,p (Z) ≥ ara(aR). Also, sdimp W = sdim R = sdim R,  by 19.3.8 and
19.3.1(iv).
We can now apply the Connectedness Bound for Complete Local Rings
19.2.10 to prove part (i), and 19.2.11 to prove parts (ii) and (iii).

19.3.16 Exercise. Let V be the affine variety in A4 of Example 2.3.7 given


by

V := VA4 (X1 X4 − X2 X3 , X12 X3 + X1 X2 − X22 , X33 + X3 X4 − X42 ).

Use 19.3.11(ii) and 19.3.15 to show that araA4 ,(0,0,0,0) (V ) ≥ 3.

It seems natural to ask whether, if the formal connectedness dimensions in


the inequality of 19.3.15(ii) are replaced by the corresponding local connect-
edness dimensions, the resulting statement is still true. We shall now provide
an example which shows that this is not always the case.

19.3.17 Example. Let R be the subring of R := C[X1 , X2 , X3 ] = O(A3 )


given by R := C[X1 , X2 , X1 X3 , X2 X3 , X32 − 1, X3 (X32 − 1)].
Let Y1 , . . . , Y6 be independent indeterminates over C, and let

f : O(A6 ) = C[Y1 , Y2 , Y3 , Y4 , Y5 , Y6 ] −→ R

be the C-algebra homomorphism such that f (Y1 ) = X1 , f (Y2 ) = X2 , f (Y3 )


= X1 X3 , f (Y4 ) = X2 X3 , f (Y5 ) = X32 − 1 and f (Y6 ) = X3 (X32 − 1).
Then p := Ker f is a prime ideal of O(A6 ) (since R is an integral domain);
let V := VA6 (p) denote the affine variety determined by p, so that there is a
natural isomorphism of C-algebras O(V ) ∼ = R.
The inclusion mapping R → R = O(A3 ), which makes R integral over its
subring R, therefore gives rise to a finite morphism of varieties α : A3 → V
such that

α((c1 , c2 , c3 )) = (c1 , c2 , c1 c3 , c2 c3 , c23 − 1, c3 (c23 − 1))

for all (c1 , c2 , c3 ) ∈ A3 . Let p = (0, 0, 1), q = (0, 0, −1) ∈ A3 , and let
422 Connectivity in algebraic varieties

0 denote (0, 0, 0, 0, 0, 0) ∈ A6 . Now it is straightforward to check that α :


A3 \ {p, q} −→ V \ {0} is an isomorphism of (quasi-affine) varieties, with
inverse β : V \ {0} −→ A3 \ {p, q} given by



⎨(d1 , d2 , d3 /d1 ) if d1 = 0,
β((d1 , d2 , d3 , d4 , d5 , d6 )) = (d1 , d2 , d4 /d2 ) if d2 = 0,


⎩(d , d , d /d ) if d = 0
1 2 6 5 5

(for all (d1 , d2 , d3 , d4 , d5 , d6 ) ∈ V \ {0}).


Let W := V ∩ VA6 (Y5 ), a closed subset of V such that 0 ∈ W . Note that
araV,0 (W ) = ara(IV (W )OV,0 ) = 1.
Let E1 := VA3 (X3 − 1) and E2 := VA3 (X3 + 1). Since α is finite, it is a
closed map, and so α(E1 ) and α(E2 ) are closed subsets of W and we have
dim(α(Ei )) = dim Ei = 2 for i = 1, 2.
Since α−1 (W ) = VA3 (X32 − 1) = E1 ∪ E2 , it follows that W can be
expressed as W = α(E1 ) ∪ α(E2 ), where α(E1 ) and α(E2 ) are closed ir-
reducible subsets of dimension 2. Therefore α(E1 ) and α(E2 ) must be the
irreducible components of W . Hence, by 19.3.3(ii), we have

c0 (W ) = dim0 (α(E1 ) ∩ α(E2 )) = dim0 {0} = 0.

On the other hand, since V is irreducible, c0 (V ) = dim0 V = 3. We therefore


have the strict inequality

c0 (W ) = 0 < 1 = 3 − 1 − 1 = c0 (V ) − araV,0 (W ) − 1.

Thus, if the formal connectedness dimensions in the inequality of 19.3.15(ii)


are replaced by the corresponding local connectedness dimensions, the result-
ing statement is not always true.

19.4 Connectivity of affine algebraic cones


Connectedness dimensions of affine algebraic cones behave particularly satis-
factorily, and the next two lemmas provide the key to this good behaviour. We
shall use the notation of 15.1.2 for affine algebraic cones.

19.4.1 Lemma. Let K be an algebraically closed field, let r ∈ N, and let


C ⊆ Ar (K) be an affine algebraic cone. Then C is irreducible (that is, C is a
variety) if and only if C is analytically irreducible at 0.

Proof. (⇒) Assume that C is irreducible. Then, by 15.1.2(ii), the ring O(C)
19.4 Connectivity of affine algebraic cones 423

is a positively graded *local homogeneous domain with O(C)0 = K; fur-


thermore, O(C)+ is the unique *maximal graded ideal of O(C), and is ac-
tually maximal and equal to IC (0). Hence OC,0 ∼ = O(C)O(C)+ . Therefore,

by 14.1.15(iv),(v), the completion OC,0 is a domain, and so C is analytically
irreducible at 0.
(⇐) Assume that C is reducible, so that C has more than one irreducible
component. By 19.1.4(i), these irreducible components all contain 0. This
means that the local vanishing ideal IAr (K),0 (C) of C at 0 has more than one
minimal prime. Hence OC,0 is not a domain, and neither is its completion.
19.4.2 Lemma. Let V be an affine variety over the algebraically closed field
K, let W ⊆ V be a closed subset, let p ∈ W , and assume that all the irre-
ducible components of W which contain p are analytically irreducible at p.
Then 
cp (W ) = cp (W ).
Proof. Let b := IV (W ) and R := OV,p /bOV,p . By 19.3.8(ii) and 19.3.10,
it is enough for us to show that c(R) = c(R);  therefore, by 19.3.1(ii), it is
enough for us to show that, for each minimal prime ideal p of R, we have
pR ∈ Spec(R),
 and this is what we shall do. Now p = qOV,p /bOV,p for some
minimal prime ideal q of b such that q ⊆ IV (p). But then Z := V (q), the
closed subset of V determined by q, is an irreducible component of W which
contains p. By hypothesis, Z is analytically irreducible at p, and so, by 19.3.13,
the ring (OV,p /IV,p (Z)) = (OV,p /qOV,p ) is a domain. Hence (R/p) is a
domain, so that pR ∈ Spec(R).


The next lemma establishes a very useful fact about the connectedness di-
mensions of an affine algebraic cone C: the formal connectedness dimension
of C at the origin and the local connectedness dimension of C at the origin are
equal, and they are both equal to the connectedness dimension c(C).
19.4.3 Proposition. Let K be an algebraically closed field, let r ∈ N, and
let C ⊆ Ar (K) be an affine algebraic cone. As in 15.1.2, we use 0 to denote
the origin of Ar (K). Then
(i) dim C = dim0 C,
(ii) sdim C = sdim0 C, and
(iii) c(C) = c0 (C) = 
c0 (C).
Proof. By 19.1.4(i), all the irreducible components of C are themselves affine
algebraic cones in Ar (K), and so contain 0. Now the dimension of an ir-
reducible affine algebraic cone C  satisfies dim C  = dim0 C  (by 19.3.2).
The claims in parts (i) and (ii) now follow immediately, while the equality
c(C) = c0 (C) follows from these considerations, 19.1.15 and 19.3.3(ii).
424 Connectivity in algebraic varieties

Finally, Lemma 19.4.1 shows that all the irreducible components of C are ana-
lytically irreducible at 0, and so Lemma 19.4.2 shows that c0 (C) = 
c0 (C).
We can now deduce the following corollary from 19.3.15.
19.4.4 Corollary. Let K be an algebraically closed field, let r ∈ N, let
D, E ⊆ Ar (K) be affine algebraic cones such that E ⊆ D. Then
(i) c(E) ≥ min {c(D), sdim D − 1} − araD,0 (E);
(ii) c(E) ≥ c(D) − araD,0 (E) − 1; and
(iii) if D is reducible, then the inequality in part (ii) is strict.
Proof. In view of Proposition 19.4.3, statements (i) and (ii) follow immedi-
ately from the corresponding statements of 19.3.15 (used with V = Ar (K),
W = D and Z = E). Statement (iii) follows from 19.3.15(iii) and Lemma
19.4.1.

19.5 Connectivity of projective varieties


In view of the close relationship between affine algebraic cones and projective
algebraic sets (see 15.2.1(iii)), we can exploit 19.4.4 to study the connectivity
of closed sets in projective varieties. We intend to do this, but first we need a
few preliminaries.
19.5.1 Lemma. Let K be an algebraically closed field, let r ∈ N, let W ⊆
Pr (K) be a non-empty closed subset of Pr (K), and consider the affine cone
Cone(W ) ⊆ Ar+1 (K) over W , as in 15.2.1(iii). Then
(i) dim W = dim(Cone(W )) − 1 = dim0 (Cone(W )) − 1;
(ii) sdim W = sdim(Cone(W )) − 1 = sdim0 (Cone(W )) − 1; and
(iii) c(W ) = c(Cone(W )) − 1 = c0 (Cone(W )) − 1 = 
c0 (Cone(W )) − 1.
Proof. Let the distinct irreducible components of W be W1 , . . . , Wn . By
19.1.4(ii), Cone(W1 ), . . . , Cone(Wn ) are the irreducible components (again
distinct) of Cone(W ). Furthermore, it follows from 15.2.1(vii) that
dim Wi = dim(Cone(Wi )) − 1 for i = 1, . . . , n.
In view of 19.4.3, the claims in statements (i) and (ii) are now immediate.
(iii) With the notation of 19.1.14, let (A, B) ∈ S(r). By 15.2.1(v), we have
  
Cone i∈A Wi ∩ j∈B Wj
  
= i∈A Cone(Wi ) ∩ j∈B Cone(Wj ) .
19.5 Connectivity of projective varieties 425

Since Cone(Wi ) (i = 1, . . . , n) are the irreducible components of Cone(W ),


it now follows from part (i) and 19.1.15 that c(W ) = c(Cone(W ))−1. Finally,
we can use 19.4.3 to complete the proof.

19.5.2 Definition. Let K be an algebraically closed field and let r ∈ N. Let


W and Z be non-empty closed subsets of Pr (K) with Z ⊆ W . Note that
IAr+1 (K) (Cone(W )) ⊆ IAr+1 (K) (Cone(Z)), and that

IAr+1 (K) (Cone(Z))/IAr+1 (K) (Cone(W ))

is an ideal of O(Ar+1 (K))/IAr+1 (K) (Cone(W )).


We define the arithmetic rank of Z with respect to W , denoted araW (Z),
by
 
araW (Z) = ara IAr+1 (K) (Cone(Z))/IAr+1 (K) (Cone(W )) .

It should be noted that, by 19.2.7(ii) and 19.3.7(ii), we have

araW (Z) ≥ araCone(W ),0 (Cone(Z)).

We are now able to state and prove a form of the Bertini–Grothendieck


Connectivity Theorem.

19.5.3 The Bertini–Grothendieck Connectivity Theorem. (See [26, Expo-


sé XIII, Corollaire 2.3].) Let K be an algebraically closed field and let r ∈ N.
Let W and Z be non-empty closed subsets of Pr (K) with Z ⊆ W . Then

(i) c(Z) ≥ min {c(W ), sdim W − 1} − araW (Z);


(ii) c(Z) ≥ c(W ) − araW (Z) − 1; and
(iii) if W is reducible, then the inequality in part (ii) is strict.

Proof. By Lemma 19.5.1, we have sdim W = sdim(Cone(W )) − 1,

c(W ) = c(Cone(W )) − 1 and c(Z) = c(Cone(Z)) − 1.

Moreover, araW (Z) ≥ araCone(W ),0 (Cone(Z)) by 19.5.2, while 15.2.1(v)


shows that Cone(W ) is reducible if W is reducible. Therefore, all three state-
ments follow from the corresponding statements of Corollary 19.4.4.

Let K be an algebraically closed field and let r ∈ N. Recall that a hyper-


surface in Pr (K) is a closed set VPr (K) (f ) defined by a single homogeneous
polynomial f ∈ K[X0 , X1 , . . . , Xr ] of positive degree. We can now deduce
the following corollary from the Bertini–Grothendieck Connectivity Theorem
19.5.3.
426 Connectivity in algebraic varieties

19.5.4 Corollary. Let K be an algebraically closed field, let r ∈ N, let W


be a non-empty closed subset of Pr (K), and let H1 , . . . , Ht ⊆ Pr (K) be
hypersurfaces. Then

(i) c(W ∩ H1 ∩ · · · ∩ Ht ) ≥ min {c(W ), sdim W − 1} − t;


(ii) c(W ∩ H1 ∩ · · · ∩ Ht ) ≥ c(W ) − t − 1; and
(iii) if W is reducible, then the inequality in part (ii) is strict.

Proof. For i = 1, . . . , t, there is a homogeneous polynomial (of positive de-


gree) fi ∈ K[X0 , X1 , . . . , Xr ] such that Hi = VPr (K) (fi ). Set
 
Z := W ∩ H1 ∩ · · · ∩ Ht = VPr (K) IPr (K) (W ) + (f1 , . . . , ft ) ,

so that IPr (K) (Z) = IPr (K) (W ) + (f1, . . . , ft ). This equation shows that
araW (Z) = ara IPr (K) (Z)/IPr (K) (W ) ≤ t. The claims now follow from
application of 19.5.3.

19.5.5 Exercise. Let K be an algebraically closed field and let r ∈ N.

(i) Prove the ‘classical’ form of Bertini’s Connectivity Theorem, that, if


V ⊆ Pr (K) is a projective variety such that dim V > 1 and H ⊆ Pr (K)
is a hypersurface, then V ∩ H is connected.
(ii) Provide an example which shows that if the irreducibility of V is dropped
from the statement in part (i) above, then the resulting statement is no
longer always true.
(iii) Provide an example of an affine variety V ⊆ Ar (K) with dim V > 1
such that V ∩ H is disconnected for a hyperplane H ⊆ Ar (K).

19.6 Connectivity of intersections


Our next aim is the study of the connectivity of the intersection of two affine al-
gebraic cones. For this, we recall, in 19.6.1 and 19.6.2 below, some elementary
facts about products of affine algebraic sets.

19.6.1 Reminder and Remark. Let K be an algebraically closed field and


let r, s ∈ N. We consider polynomial rings K[X1 , . . . , Xr ] = O(Ar (K)),
K[Y1 , . . . , Ys ] = O(As (K)) and

K[X1 , . . . , Xr ; Y1 , . . . , Ys ] = O(Ar+s (K)).

Let a be an ideal of K[X1 , . . . , Xr ] and b be an ideal of K[Y1 , . . . , Ys ], and


set V := VAr (K) (a) and W := VAs (K) (b). Recall that the product of V and
19.6 Connectivity of intersections 427

W is just the Cartesian product V × W ⊆ Ar+s (K); it is an affine algebraic


set because

V ×W
= VAr+s (K) (aK[X1 , . . . , Xr ; Y1 , . . . , Ys ] + bK[X1 , . . . , Xr ; Y1 , . . . , Ys ]).

(i) Recall from Hartshorne [30, Chapter I, Exercise 3.15] that, when V ⊆
Ar (K) and W ⊆ As (K) are irreducible, then V × W ⊆ Ar+s (K) is
again irreducible and, moreover, dim(V × W ) = dim V + dim W .
(ii) Now suppose r = s. Let Δ(r) be the diagonal

{(c, c) : c ∈ Ar (K)} ⊆ A2r (K).

Note that Δ(r) = VA2r (K) (X1 − Y1 , . . . , Xr − Yr ), so that Δ(r) is ir-


reducible, and also that there is the diagonal isomorphism of varieties
δ (r) : Ar (K) −→ Δ(r) for which δ (r) (c) = (c, c) for all c ∈ Ar (K). If
V, W ⊆ Ar (K) are closed subsets of Ar (K), then

δ (r) (V ∩ W ) = (V × W ) ∩ Δ(r) ,

and δ (r) gives rise to a homeomorphism



δ (r) : V ∩ W −→ (V × W ) ∩ Δ(r) .

19.6.2 Exercise. Let K be an algebraically closed field, let r, s ∈ N, and


let V ⊆ Ar (K) and W ⊆ As (K) be non-empty closed sets. Consider their
product V × W ⊆ Ar+s (K), as in 19.6.1.

(i) Show that, if V and W are affine algebraic cones, then V × W is again
an affine algebraic cone.
(ii) Let V1 , . . . , Vp (respectively W1 , . . . , Wq ) be the distinct irreducible
components of V (respectively W ). Show that the products

Vi × Wj (i = 1, . . . , p, j = 1, . . . , q)

are the (distinct) irreducible components of V × W .


(iii) Show that dim(V × W ) = dim V + dim W and sdim(V × W ) =
sdim V + sdim W .

19.6.3 Lemma. Let K be an algebraically closed field, let r, s ∈ N, and let


V ⊆ Ar (K) and W ⊆ As (K) be non-empty closed sets. Then

(i) c(V × W ) ≥ min {sdim V + c(W ), sdim W + c(V )};


(ii) c(V × W ) ≥ c(V ) + c(W ); and
(iii) if V and W are both reducible, then the inequality in part (ii) is strict.
428 Connectivity in algebraic varieties

Proof. Let V1 , . . . , Vp be the distinct irreducible components of V and let


W1 , . . . , Wq be the distinct irreducible components of W . By 19.6.2(ii), the
products Vi × Wj (i = 1, . . . , p, j = 1, . . . , q) are the (distinct) irreducible
components of V × W . By 19.1.15, there are two non-empty subsets A, B ⊆
{1, . . . , p} × {1, . . . , q} such that A ∪ B = {1, . . . , p} × {1, . . . , q} and
 
c(V × W ) = dim (i,j)∈A Vi × Wj ∩ (k,l)∈B Vk × Wl .

The argument now splits into two cases. Suppose first that there exists i0 ∈
{1, . . . , p} such that there are indices j, l ∈ {1, . . . , q} for which (i0 , j) ∈ A
and (i0 , l) ∈ B. Then

A := {j ∈ N : (i0 , j) ∈ A} and B := {l ∈ N : (i0 , l) ∈ B}

are non-empty sets such that A ∪ B = {1, . . . , q}. Set


 
Z := j∈A Wj ∩ l∈B Wl .

By 19.1.15, dim Z ≥ c(W ). Let Y be an irreducible component of Z such


that dim Y = dim Z. Then there exist j ∈ A and l ∈ B such that Y ⊆ Wj
and Y ⊆ Wl . Therefore Vi0 × Y ⊆ (Vi0 × Wj ) ∩ (Vi0 × Wl ), and so (see
19.6.2(iii))

c(V × W ) ≥ dim ((Vi0 × Wj ) ∩ (Vi0 × Wl ))


≥ dim(Vi0 × Y ) = dim Vi0 + dim Y ≥ sdim V + c(W ).

We now deal with the remaining case, when there is no index i0 with the
properties described above. Choose i ∈ {1, . . . , p} such that (i, j  ) ∈ A for
some j  ∈ {1, . . . , q}. Then, for all j ∈ {1, . . . , q}, we must have (i, j) ∈ B,
so that (i, j) ∈ A. But there is also a pair (k, j0 ) ∈ B; thus (i, j0 ) ∈ A and
(k, j0 ) ∈ B, and we can use the argument of the previous paragraph, with the
rôles of V and W interchanged, to deduce that c(V × W ) ≥ sdim W + c(V ).
This proves statement (i).
Parts (ii) and (iii) are now immediate from part (i) and 19.2.2.

19.6.4 Exercise. Let V  := VA2 (XY ) and W  := VA2 (X) (with the notation
of 2.3.1).

(i) Calculate sdim(V  × V  ) and c(V  × V  ).


(ii) Calculate sdim(V  × W  ) and c(V  × W  ).
(iii) Show that it is possible for the inequality in 19.6.3(ii) to be an equality
when just one of V and W is reducible.
19.6 Connectivity of intersections 429

19.6.5 Proposition: the Intersection Inequality for the Connectedness


Dimensions of Affine Algebraic Cones. Let K be an algebraically closed
field, let r ∈ N, and let C, D ⊆ Ar (K) be affine algebraic cones.
(i) We have

c(C ∩ D)
≥ min {sdim C + sdim D − 1, sdim C + c(D), sdim D + c(C)} − r.
(ii) Furthermore,
c(C ∩ D) ≥ min {sdim C + c(D), sdim D + c(C)} − r − 1,
with strict inequality if C or D is reducible.
(iii) Consequently, c(C ∩ D) ≥ c(C) + c(D) − r − 1 + ε, where ε = 0, 1 or
2 according as none, one or both of C and D are reducible.
Proof. By 19.6.2(i), the product C × D ⊆ A2r (K) is an affine algebraic
cone, and by 19.6.2(iii) we have sdim(C × D) = sdim C + sdim D. Also, by
19.6.3, we have c(C × D) ≥ min {sdim C + c(D), sdim D + c(C)} .
As in 19.6.1, write K[X1 , . . . , Xr ; Y1 , . . . , Yr ] = O(A2r (K)), and note
that the diagonal Δ(r) ⊆ A2r (K) is an affine algebraic cone in A2r (K). Since
IA2r (K) (Δ(r) ) = (X1 − Y1 , . . . , Xr − Yr ), we have

IA2r (K) ((C × D) ∩ Δ(r) ) = IA2r (K) (C × D) + (X1 − Y1 , . . . , Xr − Yr ).
It therefore follows from 19.2.7(ii) that
araC×D,0 ((C × D) ∩ Δ(r) )
≤ ara IA2r (K) ((C × D) ∩ Δ(r) )/IA2r (K) (C × D)
≤ r.
If we now apply 19.4.4(i) to the two affine algebraic cones (C × D) ∩ Δ(r) ⊆
C × D in A2r (K), we obtain
c((C × D) ∩ Δ(r) )
≥ min {sdim C + sdim D − 1, sdim C + c(D), sdim D + c(C)} − r.
In view of the homeomorphism provided by 19.6.1(ii), we have
c((C × D) ∩ Δ(r) ) = c(C ∩ D),
and so the claim in (i) is proved.
If C or D is reducible, then so too is C × D (by 19.6.2(ii)); we can therefore
deduce part (ii) from 19.4.4(ii),(iii).
430 Connectivity in algebraic varieties

In view of 19.2.2, part (iii) follows from part (ii) if at least one of C, D
is irreducible, and so we deal now with the case where both C and D are
reducible. Then, by 19.2.2, we have sdim C ≥ c(C) + 1 and also sdim D ≥
c(D) + 1, so that
sdim C +c(D) ≥ c(C)+c(D)+1 and sdim D+c(C) ≥ c(C)+c(D)+1.
We use these inequalities in conjunction with part (ii) to see that
c(C ∩ D) ≥ min {sdim C + c(D), sdim D + c(C)} − r − 1 + 1
≥ c(C) + c(D) + 1 − r − 1 + 1 = c(C) + c(D) − r − 1 + 2.
We now apply Proposition 19.6.5 to affine cones over projective algebraic
sets to deduce the following theorem.
19.6.6 Theorem: the Intersection Inequality for the Connectedness
Dimensions of Projective Algebraic Sets. Let K be an algebraically closed
field, let r ∈ N, and let V, W ⊆ Pr (K) be non-empty closed sets.

(i) We have
c(V ∩ W ) + r
≥ min {sdim V + sdim W − 1, sdim V + c(W ), sdim W + c(V )} .

(ii) Furthermore,
c(V ∩ W ) + r ≥ min {sdim V + c(W ), sdim W + c(V )} − 1,
with strict inequality if V or W is reducible.
(iii) Consequently, c(V ∩ W ) + r ≥ c(V ) + c(W ) − 1 + ε, where ε = 0, 1
or 2 according as none, one or both of V and W are reducible.

Proof. Let C := Cone(V ) ⊆ Ar+1 (K) and D := Cone(W ) ⊆ Ar+1 (K)


be the affine cones in Ar+1 (K) over V and W respectively. Observe that
C ∩ D = Cone(V ) ∩ Cone(W ) = Cone(V ∩ W ). By Lemma 19.5.1, we
have sdim C = sdim V + 1, sdim D = sdim W + 1,
c(C) = c(V ) + 1, c(D) = c(W ) + 1, c(C ∩ D) = c(V ∩ W ) + 1.
Moreover, it follows from 15.2.1(v) that C is reducible if and only if V is,
and D is reducible if and only if W is. All three parts of the theorem now
follow from the corresponding parts of Proposition 19.6.5 applied to the affine
algebraic cones C, D ⊆ Ar+1 (K).
Part of Theorem 19.6.6(iii) amounts to the following formulation of the Con-
nectivity Theorem due to W. Fulton and J. Hansen and to G. Faltings.
19.6 Connectivity of intersections 431

19.6.7 Corollary: the Connectivity Theorem of Fulton–Hansen and


Faltings. (See [20] and [15].) Let K be an algebraically closed field, let r ∈ N,
and let V, W ⊆ Pr (K) be projective varieties. Then
c(V ∩ W ) ≥ dim V + dim W − r − 1.
Proof. By 19.1.10(iii), we have c(V ) = dim V and c(W ) = dim W , and so
the claim follows from 19.6.6(iii).
19.6.8 Corollary (W. Fulton and J. Hansen [20, Corollary 1]). Let K be an
algebraically closed field, let r ∈ N, and let V, W ⊆ Pr (K) be projective
varieties such that dim V + dim W > r. Then V ∩ W is connected.
Proof. By Corollary 19.6.7, we have c(V ∩ W ) ≥ 0, so that V ∩ W is con-
nected by 19.1.10(i).
19.6.9 Example. Interpret the polynomial ring C[X1 , X2 , X3 ] as O(A3 ), as
in 2.3.1. In A3 , we consider the two surfaces
◦ ◦
V := VA3 (X1 − X2 X3 ) and W := VA3 (X1 + X22 − X2 X3 − 1).
Then
◦ ◦
V ∩ W = VA3 (X1 − X2 X3 , X1 + X22 − X2 X3 − 1)
= VA3 (X1 − X2 X3 , X22 − 1)
= VA3 (X1 − X2 X3 , X2 + 1) ∪ VA3 (X1 − X2 X3 , X2 − 1)
= VA3 (X1 + X3 , X2 + 1) ∪ VA3 (X1 − X3 , X2 − 1),
◦ ◦ ◦
so that V ∩ W is the union of the two lines L1 = VA3 (X1 + X3 , X2 + 1) and

L2 = VA3 (X1 − X3 , X2 − 1) and is disconnected. On the other hand,
◦ ◦
dim V + dim W = 2 + 2 > 3
◦ ◦
and V and W are both irreducible (as their defining polynomials are). There-
fore, the analogue of 19.6.8 for affine varieties is not always true.
Now interpret the polynomial ring C[X0 , X1 , X2 , X3 ] as O(A4 ) and con-
sider the projective varieties V, W ⊆ P3 defined by
V := VP3 (X0 X1 − X2 X3 ) and W := VP3 (X0 X1 + X22 − X2 X3 − X02 ).
(Note that both X0 X1 −X2 X3 and X0 X1 +X22 −X2 X3 −X02 are irreducible
◦ ◦
polynomials. One can think of V and W as the projective closures of V and W

=
respectively with respect to the isomorphism of varieties σ0 : A3 −→ U0 P3
of 19.3.7(i): see [30, Chapter I, Exercise 2.9].) Since dim V = dim W = 2,
Corollary 19.6.8 tells us that V ∩ W is connected.
432 Connectivity in algebraic varieties

Indeed, we have

V ∩W
= VP3 (X0 X1 − X2 X3 , X0 X1 + X22 − X2 X3 − X02 )
= VP3 (X1 + X3 , X2 + X0 ) ∪ VP3 (X1 − X3 , X2 − X0 ) ∪ VP3 (X0 , X2 )
= L1 ∪ L2 ∪ L,

where L1 := VP3 (X1 + X3 , X2 + X0 ) and L2 := VP3 (X1 − X3 , X2 − X0 ) are


◦ ◦
the projective closures of L1 and L2 respectively with respect to σ0 , and L is
the ‘line at infinity’ VP3 (X0 , X2 ) which intersects L1 at p := (0 : 1 : 0 : −1)
and L2 at q := (0 : 1 : 0 : 1).
◦ ◦
19.6.10 Exercise. Let V := VA3 (X13 + X12 − X22 ), W := VA3 (X2 ), and

V := VP3 (X13 + X0 X12 − X0 X22 ), W := VP3 (X2 ).


◦ ◦
Determine V ∩ W and V ∩ W .

19.7 The projective spectrum and connectedness


Let us take stock. Among the results we have presented so far in this chapter are
the Connectedness Bound for Complete Local Rings 19.2.10, Grothendieck’s
Connectedness Theorem 19.2.12, the Bertini–Grothendieck Connectivity The-
orem 19.5.3, the Intersection Inequality for the Connectedness Dimensions of
Projective Algebraic Sets 19.6.6, and the Fulton–Hansen Connectivity Theo-
rem 19.6.7. All of these are important results about connectivity, but none of
them mentions local cohomology in its statement. However, our proofs above
of these results depend on one proposition, crucial for our approach, that does
use local cohomology, namely Proposition 19.2.8. The key argument in our
proof of that proposition concerned part of an (exact) Mayer–Vietoris sequence
d−1
Ha∩b (R) −→ Ha+b
d
(R) −→ Had (R) ⊕ Hbd (R),

where (R, m) is a d-dimensional complete local domain and a and b are non-
zero proper ideals of R whose sum is m-primary. An exact sequence like the
one displayed above was considered in [5, p. 484], and so we shall refer to
the above sequence as Rung’s display. We used it in conjunction with the lo-
cal Lichtenbaum–Hartshorne Vanishing Theorem 8.2.1 and the non-vanishing
result of 6.1.4. From these few arguments from local cohomology, we have
derived far-reaching geometric consequences which do not involve local coho-
mology in their statements!
19.7 The projective spectrum and connectedness 433

Our final part of this chapter is concerned with a graded analogue of Rung’s
display, which we shall use in conjunction with the Graded Lichtenbaum–
Hartshorne Vanishing Theorem 14.1.16 and the non-vanishing result of Ex-
ercise 16.1.10(ii) in order to prove a ring-theoretic version of Zariski’s Main
Theorem on the Connectivity of Fibres of Blowing-up. To prepare for this, we
remind the reader, in 19.7.1 below, of some facts concerning the projective
spectrum of a positively graded commutative Noetherian ring.

19.7.1 Reminder and Exercise. Assume that R = n∈N0 Rn is positively
graded.

(i) Show that, for any two prime ideals p, q ∈ * Spec(R) with p ⊂ q and
ht q/p > 1, there exists a prime ideal s ∈ * Spec(R) with p ⊂ s ⊂ q.
(Recall that ⊂ denotes strict inclusion.)

Recall that the projective spectrum of R, denoted by Proj(R), is the set


* Spec(R) \ Var(R+ ) of all graded prime ideals of R which do not contain the
irrelevant ideal R+ (see 15.1.1). The Zariski topology on Proj(R) is defined
as the topology induced by the Zariski topology on Spec(R). Note that, since,
for an ideal b of R, we have * Spec(R) ∩ Var(b) = * Spec(R) ∩ Var(b) where
b is the ideal of R generated by all the homogeneous components of all the
elements of b, it follows that the set of closed sets for the Zariski topology on
Proj(R) is {Proj(R) ∩ Var(c) : c is a graded ideal of R}.
Let π : Proj(R) −→ Spec(R0 ) be the natural map, defined by π(p) =
p ∩ R0 for all p ∈ Proj(R). In the case when (R0 , m0 ) is local, we refer to
π −1 (m0 ) as the special fibre of π.

(ii) Show that * Spec(R), with the topology induced by the Zariski topology
on Spec(R), is a Noetherian topological space.
(iii) Show that Proj(R) is a Noetherian topological space.
(iv) Show that the set of closed irreducible subsets of Proj(R) is

{Proj(R) ∩ Var(p) : p ∈ Proj(R)}.

(v) Let b be a graded ideal of R. Show that

Proj(R) ∩ Var(b) = Proj(R) ∩ Var(b ∩ R+ ),

and deduce that each closed set in Proj(R) can be defined by finitely
many homogeneous elements of positive degree in R.
(vi) Show that π −1 (Var(b0 )) = Proj(R) ∩ Var(b0 R) for each ideal b0 of
R0 , and deduce that π : Proj(R) −→ Spec(R0 ) is continuous.
434 Connectivity in algebraic varieties

(vii) Let b be a graded ideal of R. Show that


 
π(Proj(R) ∩ Var(b)) = Var n∈N (b :R (R+ )n ) ∩ R0 .
(Here are some hints: show that it is sufficient to prove that, for p ∈
Proj(R), we have π(Proj(R) ∩ Var(p)) = Var(p ∩ R0 ); then show that,

in the special case in which (R0 , m0 ) is local, R+ ⊆ m0 R + p when
p ∈ Proj(R).)
 
Deduce that π is closed and that π(Proj(R)) = Var R0 ∩ ΓR+ (R) .
(viii) Assume that (R0 , m0 ) is local. Show that, for a graded ideal b of R,

we have m0 ∈ π(Proj(R) ∩ Var(b)) if and only if R+ ⊆ m0 R + b.
Show that π −1 (m0 ) ∩ W = ∅ for every non-empty closed subset W of
Proj(R).
Deduce that π −1 (m0 ) is connected if and only if π −1 (Z) is connected
for each non-empty closed subset Z of Spec(R0 ).
Our next lemma uses the promised graded analogue of Rung’s display.

19.7.2 Lemma. Assume that R = n∈N0 Rn is positively graded and that
(R0 , m0 ) is local and complete. Then the following statements are equivalent:
(i) Proj(R) is connected;
(ii) the special fibre π −1 (m0 ) (under the natural map π : Proj(R) −→
Spec(R0 ) of 19.7.1) is connected.
Proof. (ii) ⇒ (i) This is immediate from Exercise 19.7.1(viii).
(i) ⇒ (ii) Since Proj(R) = ∅, we have π −1 (m0 ) = ∅, by 19.7.1(viii), and
R+ = 0. If dim R0 = 0, then π −1 (m0 ) = Proj(R). Therefore we can, and
do, assume that dim R0 > 0.
Suppose that π −1 (m0 ) = Proj(R) ∩ Var(m0 R) is disconnected; we shall
obtain a contradiction. Then there exist two non-empty closed subsets Z1 , Z2
of Proj(R) such that Z1 ∩ Z2 = ∅ and Z1 ∪ Z2 = π −1 (m0 ). Let T1 , . . . , Tr
be the distinct irreducible components of Proj(R), and set
A := {i ∈ {1, . . . , r} : Ti ∩Z1 = ∅}, B := {j ∈ {1, . . . , r} : Tj ∩Z2 = ∅}.
Clearly each of A and B is non-empty, since Z1 = ∅ = Z2 . By 19.7.1(viii),
it follows that, for each k ∈ {1, . . . , r}, we have Tk ∩ (Z1 ∪ Z2 ) = ∅. This
shows that A ∪ B = {1, . . . , r}.
We show next that A ∩ B = ∅. Suppose, on the contrary, that A ∩ B = ∅,
 
and seek a contradiction. Then W := i∈A Ti ∩ j∈B Tj would be non-
empty because Proj(R) is connected; also, we would have
π −1 (m0 ) ∩ W = (Z1 ∪ Z2 ) ∩ W = (Z1 ∩ W ) ∪ (Z2 ∩ W ) = ∅,
19.7 The projective spectrum and connectedness 435

contrary to 19.7.1(viii). Hence there exists k ∈ A ∩ B.


Then Zi ∩ Tk = ∅ for i = 1, 2 and so π −1 (m0 ) ∩ Tk is disconnected. By
19.7.1(iv), there exists p ∈ Proj(R) for which Tk = Proj(R) ∩ Var(p). Let

R = n∈N0 Rn = R/p, graded in the natural way, let m0 be the maximal
ideal of R0 , and let π : Proj(R) −→ Spec(R0 ) be the natural map. The
natural homeomorphism between Tk and Proj(R) maps π −1 (m0 ) ∩ Tk onto
π −1 (m0 ), and so this latter special fibre is disconnected. Hence, in our search
for a contradiction, we can, and do, assume that the graded ring R is a domain.
There exist graded ideals b, c of R such that Proj(R) ∩ Var(b) = ∅,
Proj(R) ∩ Var(c) = ∅, Proj(R) ∩ Var(b) ∩ Var(c) = ∅, and
(Proj(R) ∩ Var(b)) ∪ (Proj(R) ∩ Var(c)) = π −1 (m0 ).
Since Proj(R) ∩ Var(b) ⊆ π −1 (m0 ), we have
Proj(R) ∩ Var(b) = Proj(R) ∩ Var(b + m0 R),
and a similar comment applies to c; we therefore assume that b ⊇ m0 R and
c ⊇ m0 R. We can now deduce that
√ √ 
b + c = m0 R + R+ and b ∩ c = m0 R.
Set d := dim R. We can now use 1.2.3 in conjunction with the Graded Mayer–
Vietoris sequence 14.1.5 to see that there is an exact sequence of graded R-
modules and homogeneous homomorphisms
d−1
Hm 0R
(R) −→ Hm
d
0 R+R+
(R) −→ Hbd (R) ⊕ Hcd (R).
As Proj(R) ∩ Var(b) = ∅ and Proj(R) ∩ Var(c) = ∅, we have dim R/b > 0
and dim R/c > 0. Therefore, by the Graded Lichtenbaum–Hartshorne Vanish-
ing Theorem 14.1.16, we have Hbd (R) = Hcd (R) = 0; also, Lemma 14.1.12
d−1
shows that Hm 0R
(R)n ∼
= Hm d−1
0
(Rn ) for all n ∈ Z, and this is 0 for n < 0.
d
Hence Hm0 R+R+ (R)n = 0 for all n < 0, contrary to 16.1.10(ii).
19.7.3 Theorem: the Connectedness Criterion for the Special Fibre. As-

sume that R = n∈N0 Rn is positively graded and that (R0 , m0 ) is local. Let
R  := R ⊗R R
+0 denote the completion of R0 . As in 16.2.2, set R + ; after an ob-
  0 0 +0 ).
vious identification, we can consider R as a graded ring (Rn ⊗R R
n∈N0 0

Let π : Proj(R) −→ Spec(R0 ) be the natural map.


The following statements are equivalent:
 is connected;
(i) Proj(R)
(ii) the special fibre π −1 (m0 ) is connected;
(iii) for each non-empty closed subset Z of Spec(R0 ), the set π −1 (Z) is con-
nected.
436 Connectivity in algebraic varieties

Proof. The equivalence of (ii) and (iii) is the subject of part of Exercise
19.7.1(viii).
(i) ⇔ (ii) Let π  −→ Spec(R
 : Proj(R) +0 ) be the natural map. Let m+0 =
+ +
m0 R0 , the maximal ideal of R0 . Since there are natural homeomorphisms
π −1 (m0 ) ≈ Proj(R/m0 R) and π −1 (m
+0 ) ≈ Proj(R/ m  and since there
+0 R),

is a homogeneous ring isomorphism R/m0 R −→ R/m
=  
+0 R, there is a home-
omorphism π −1 (m0 ) ≈ π −1 (m
+0 ). We can therefore apply Lemma 19.7.2 to
complete the proof.

We are now going to apply the Connectedness Criterion for the Special Fibre
19.7.3 to the (ordinary) Rees ring of an ideal of a commutative Noetherian ring.

19.7.4 Notation. We use R(a) to denote the ordinary Rees ring n∈N0 an
of a. If, as in 18.1.4, we let {a1 , . . . , ah } be a generating set for a and T be an
indeterminate, then there is a homogeneous isomorphism of graded R-algebras

=
R(a) −→ R[a1 T, . . . , ah T ] =: R[aT ].
We remark here that R(a) is also called the blowing-up ring of a; this ter-
minology has its roots in the fact that Proj (R(a)) is the topological space
underlying the scheme obtained by blowing up Spec(R) with respect to a.

19.7.5 Corollary. Let p ∈ Spec(R) be such that R +p has only one minimal
prime ideal and the ideal aRp of Rp is not nilpotent. Let R(a) denote the
ordinary Rees ring of a (see 19.7.4), and let πa : Proj(R(a)) −→ Spec(R) be
the natural map. Then the fibre πa−1 (p) of p under πa is connected.

Proof. We can use localization at p to see that it is enough for us to prove


the claim under the assumption that (R, m) is local and p = m; we make
this assumption in what follows. Now R  is flat over R; also, R(a) can be

viewed as a subring of R(aR) and there is a homogeneous isomorphism of
∼  
R(a)-algebras R(a) ⊗R R  −→ =  Therefore Proj R(a) ⊗R R
R(aR).  and
 
Proj R(aR)  are homeomorphic. By Theorem 19.7.3, it is sufficient for us to
prove that Proj(R(aR)) is connected.
In order to prove this, we let p be the unique minimal prime ideal of R,
 given by P :=   
and P be the ideal of R(aR) n∈N0 (a R ∩ p); in fact P ∈
n


* Spec(R(aR)). Since  p = 0, there exists t ∈ N such that  p t = 0; hence
P = 0, so that P is the unique minimal prime of R(aR).
t 

It follows from the faithful flatness of R → R  that aR⊆ 0= p, and so
n  n
a R ∩ p ⊂ a R for all n ∈ N. Hence
 \ Var(R(aR)
P ∈ * Spec(R(aR)) 
 + ) = Proj(R(aR)).

 we also have Proj(R(aR))


As P is the unique minimal prime of R(aR),  =
19.7 The projective spectrum and connectedness 437

Proj(R(aR))  ∩ Var(P). Hence, by 19.7.1(iv), Proj(R(aR))


 is irreducible,
and so it is connected by 19.1.2(i).
We are now able to deduce from 19.7.5 the last main connectedness result
of this chapter.
19.7.6 Corollary: ring-theoretic version of Zariski’s Main Theorem on
the Connectivity of Fibres of Blowing-up. Assume that R is a domain and
that a = 0. Let p ∈ Spec(R) be such that R +p is also a domain. Let πa :
Proj(R(a)) −→ Spec(R) be the natural map. Then the fibre πa−1 (p) is con-
nected. 
19.7.7 Exercise. Assume that R is a domain and that a = 0. Define πa :
Proj(R(a)) −→ Spec(R) as in 19.7.5. Let Z ⊆ Spec(R) be a connected
+m is an integral domain for each maximal
closed subset of Spec(R) such that R
ideal m of R which belongs to Z. Prove that πa−1 (Z) is connected.
19.7.8 Exercise. Assume that (R, m) is a local domain and that a is non-zero

and proper. We use G(a) to denote the associated graded ring i∈N0 ai /ai+1 :
see 18.2.1.
 is an integral domain, then Proj(G(a)) is connected.
(i) Show that, if R
(ii) Show that Proj(G(m)) is connected if and only if Proj(R(mR))  is con-
nected.
19.7.9 Exercise. Let V denote the Cartesian curve VA2 (X 3 + X 2 − Y 2 ), and
W denote the cuspidal curve VA2 (X 3 − Y 2 ).
(i) Let (R, m) be the local ring OV,0 of the origin 0 on V . Show that R has
two minimal primes, and that Proj(G(m)) is a discrete topological space
with just two points.
+
(ii) Let (R , m ) be the local ring OW,0 of the origin 0 on W . Show that R
is a domain and that Proj(G(m )) is a singleton set.
20

Links with sheaf cohomology

In this last chapter we shall develop the links between local cohomology and
the cohomology of quasi-coherent sheaves over certain Noetherian schemes.
Here we shall assume for the first time that the reader has some basic know-
ledge about schemes and sheaves: our reference for these topics is Hartshorne’s
book [30]. The central idea in this chapter is to extend our earlier relations in
2.3.2 and 15.1.3 between ideal transforms and rings of regular functions on
varieties to quasi-coherent sheaves over certain Noetherian schemes. We shall
be very concerned with a generalization of the ‘Deligne Isomorphism’ (see [30,
Chapter III, Exercise 3.7, p. 217]) which links the group of sections (over an
open subset) of an induced sheaf on an affine scheme with an ideal transform.
More precisely, let M 6 denote the sheaf induced by an R-module M on the
affine scheme Spec(R), and let U = Spec(R)\Var(a), where a ⊂ R; then the
group of sections Γ(U, M 6) is isomorphic to the ideal transform Da (M ). We
shall use standard techniques involving negative strongly connected sequences
of functors to extend this Deligne Isomorphism, and our generalization of it,
to produce the Deligne Correspondence 20.3.11. This correspondence provides
connections between higher cohomology groups of induced sheaves on the one
hand, and local cohomology modules on the other.
We shall also examine the case when R is graded in some detail. Here the
central result for us is the Serre–Grothendieck Correspondence 20.3.15, which
we shall also derive from the Deligne Isomorphism by standard ‘connected se-
quence’ arguments. In this introduction, we mention only some consequences
of the Serre–Grothendieck Correspondence for projective schemes. Consider,

therefore, the special case in which R = n∈N0 Rn is positively Z-graded
and homogeneous; set T := Proj(R) = * Spec(R) \ Var(R+ ) and consider
the projective scheme (T, OT ) defined by R (see [30, Chapter II, §2, p. 76]).
If F is a coherent sheaf of OT -modules, there is a finitely generated graded
R-module N such that F is isomorphic to the sheaf of OT -modules associated
20.1 The Deligne Isomorphism 439

to N on Proj(R). The classical form of the Serre–Grothendieck Correspon-


dence yields, for each i ∈ N0 and n ∈ Z, an R0 -isomorphism between the i-th
cohomology group H i (T, F(n)) of the twisted sheaf F(n) and Ri DR+ (N )n ,
so that, for i > 0, we have H i (T, F(n)) ∼ = HR i+1
+
(N )n . These results enable
us to deduce quickly, from algebraic results about local cohomology estab-
lished earlier in the book, significant results about the cohomology of coherent
sheaves of OT -modules.
In the final two sections, we use this approach to present proofs of some
fundamental theorems, and extensions thereof, from projective algebraic ge-
ometry, including Serre’s Finiteness Theorem for the cohomology of coherent
sheaves over projective schemes (see 20.4.8), Serre’s Criterion for the global
generation of coherent sheaves over projective schemes (see 20.4.13), the ex-
istence of a Hilbert polynomial for a coherent sheaf over a projective scheme
over an Artinian base ring (see 20.4.16), Mumford’s Regularity Bound for
coherent sheaves of ideals over a projective space (see 20.4.18), the Severi–
Enriques–Zariski–Serre Vanishing Theorem for the cohomology of coherent
sheaves over projective schemes (see 20.4.23), Serre’s Cohomological Crite-
rion for local freeness of coherent sheaves over regular projective schemes
(see 20.5.6), Horrocks’ Splitting Criterion for coherent sheaves over projective
spaces (see 20.5.8), and Grothendieck’s Splitting Theorem for coherent locally
free sheaves over the projective line (see 20.5.9).

20.1 The Deligne Isomorphism


The basic result in this chapter is Theorem 20.1.14, a generalized version of
the Deligne Isomorphism. This result is not formulated explicitly in sheaf-
theoretic terms, but rather in terms of certain local families of fractions. We
now start to develop the notions of ‘S-topology’ and ‘S-local family of frac-
tions’ which we shall use in our formulation and proof of Theorem 20.1.14.

20.1.1 Notation and Terminology. Throughout this chapter, we shall use S


to denote a non-empty subset of R which is closed under multiplication. It
should be noted that we do not assume that 1 ∈ S.

(i) We denote by AS the set {(S  ) : ∅ = S  ⊆ S} of all ideals of R


generated by elements of S. As R is Noetherian, each ideal in AS can
be generated by finitely many elements of S. Note that, if (ai )i∈I is a

family of ideals in AS , then i∈I ai ∈ AS , and if b1 , . . . , br ∈ AS ,
r
then j=1 bj ∈ AS too.
440 Links with sheaf cohomology

(ii) We shall say that a subset T of Spec(R) is essential with respect to S


precisely when
(a) T ⊆ Spec(R) \ Var((S));
(b) there exists b ∈ AS such that T ⊆ Var(b); and

(c) c = p∈Var((S))∪(T ∩Var(c)) p for all c ∈ AS .
Observe that condition (a) is automatically satisfied if S contains a
unit, that condition (b) is automatically satisfied if S contains 0, and
that, for every subset T of Spec(R), we certainly have
√ *
c⊆ p for all c ∈ AS .
p∈Var((S))∪(T ∩Var(c))

Note also that, if S = {0}, then the empty set is the unique subset of
Spec(R) which is essential with respect to S. In general, if the empty set

is essential with respect to S, then c = (S) for all c ∈ AS , so that,
in particular,
 √ ≤ 1 if S contains a non-unit; if, in addition, 0 ∈ S,
ht(S)
then (S) = 0.
Note also that condition (a) implies that, for each p ∈ T , there exists
sp ∈ S \ p.
(iii) Let T be a subset of Spec(R) which is essential with respect to S. It is
an easy consequence of the last sentence in part (i) that
{T ∩ Var(b) : b ∈ AS }
is the set of closed sets in a topology on T : we refer to this topology
(S)
as the S-topology on T . We denote by UT the set of open sets in the
(S)
S-topology on T : thus UT = {T \ Var(b) : b ∈ AS }. For p ∈ T , we
(S) (S)
denote by UT,p the set {U ∈ UT : p ∈ U } of all open neighbourhoods
of p in the S-topology on T .
Since every non-empty set of closed subsets of T (in the S-topology)
has a minimal member with respect to inclusion, the S-topology makes
T into a Noetherian topological space, and hence (see 19.1.6(i)) every
open subset of T is quasi-compact.
We now state the assumptions that will be in force throughout this chapter.
20.1.2 Standard hypotheses. Throughout this chapter, S will denote a non-
empty subset of R which is closed under multiplication, AS will denote the set
{(S  ) : ∅ = S  ⊆ S} of all ideals of R generated by non-empty subsets of
S, T will denote a subset of Spec(R) which is essential with respect to S (see
20.1.1(ii)), a will denote an ideal in AS , and M will denote an R-module.
20.1.3 Examples. (The hypotheses of 20.1.2 apply.)
20.1 The Deligne Isomorphism 441

(i) If S = R, then AS is the set of all ideals of R, and T := Spec(R)


is essential with respect to S. The S-topology on T is just the ordinary
Zariski topology.
(ii) Assume that G is a finitely generated torsion-free Abelian group and
 
R = g∈G Rg is G-graded, and take S := g∈G Rg to be the set of
homogeneous elements of R. Then AS is the set of all graded ideals of
R, and T := * Spec(R) is essential with respect to S. It is easy to see
that the S-topology on T is again the ordinary Zariski topology.

20.1.4 Exercise. Assume that R = n∈N0 Rn is positively Z-graded, and

take S := n∈N Rn , the set of homogeneous elements of R of positive de-
grees, together with the zero element of R; then AS is the set of all graded
ideals of R which are contained in R+ .
Show that T := Proj(R) = * Spec(R) \ Var(R+ ) is essential with respect
to S, and that the S-topology on T is the Zariski topology.
The next exercise suggests how the notion of S-topology on T can be re-
garded as a generalization of the Zariski topology on an affine variety.
20.1.5 Exercise. Let V be an affine variety over the algebraically closed field
K. Take R := O(V ), and S := R; by 20.1.3(i), we know that T := Spec(R)
is essential with respect to S. Let max T denote the set of all maximal ideals
of R; note that max T is also essential with respect to S (since R is a Hilbert
ring), and that the S-topology on max T is the topology induced from the
Zariski topology on T .

Use the Nullstellensatz to show that there is a homeomorphism j : V −→
max T given by j(p) = IV (p) for all p ∈ V .
(S)
Show that the assignment U → U ∩max T defines a bijection between UT
(S)
and Umax T .
We next introduce the notion of ‘S-local family of fractions’: this general-
izes the concept of regular function on a quasi-affine variety.
20.1.6 Notation and Terminology. (The hypotheses of 20.1.2 apply.) Note

that, if T = ∅, then p∈T (S \ p) = ∅: a member of this set is called a family
of denominators in S.
(i) Let p ∈ T . Since S\p = ∅, we can form the commutative ring (S\p)−1 R
and the (S \ p)−1 R-module (S \ p)−1 M .

(ii) Let ∅ = U ∈ UT . The elements of p∈U (S \ p)−1 M are called fam-
(S)

ilies of fractions over U with numerators in M and denominators in S.



Such a family γ = (γp )p∈U ∈ p∈U (S \ p)−1 M is called local, or an
S-local family of fractions over U , if and only if, for each q ∈ U , there
442 Links with sheaf cohomology
(S)
exist W ∈ UT,q and (s, m) ∈ S × M such that, for each p ∈ U ∩ W , it
is the case that s ∈ S \ p and γp = m/s in (S \ p)−1 M .
6(U ).
The set of all S-local families of fractions over U is denoted by M
6
We set M (∅) = 0.

Note that p∈U (S \ p)−1 M has a natural structure as a module over
  )
the commutative ring p∈U (S \ p)−1 R. It is easy to check that R(U
 −1 6 
is a subring of p∈U (S \ p) R and that M (U ) is an R(U )-submodule

of p∈U (S \ p)−1 M .

Note that R(∅) 6(∅) = 0 is an R(∅)-
is a trivial ring; of course, M 
module.
(iii) Now let m ∈ M and let U be as in part (ii). Choose a family of de-

nominators (sp )p∈U ∈ p∈U (S \ p). For each p ∈ U , the fraction
sp m/sp ∈ (S \ p)−1 M is independent of the choice of denominator
sp ∈ S \ p. We can therefore define a family of fractions

  7
sp m
 :=
m ∈ (S \ p)−1 M,
sp p∈U p∈U

which does not depend on the choice of family of denominators (sp )p∈U .
Moreover, for q ∈ U , we have, for each p ∈ U \ Var((sq )), the relation
sp m/sp = sp sq m/sp sq = sq m/sq in (S \ p)−1 M . As U \ Var((sq )) ∈
(S)
UT,q , we thus see that the family of fractions m
 is local. We can therefore
U 6
define a map εM : M −→ M (U ) by εM (m) = mU
 = (sp m/sp )p∈U for
all m ∈ M . Of course, we define ε∅ : M −→ M 6(∅) = 0 to be the zero
M
homomorphism.

R : R −→ R(U ) is a ring homomorphism, and so turns
Note that εU
 ) into an R-algebra; furthermore, εU is an R-homomorphism. Sim-
R(U M
ilar comments apply to ε∅R and ε∅M .
(S)
(iv) Now let U, V ∈ UT with V ⊆ U . Suppose that V = ∅, and let γ =
(γp )p∈U ∈ M6(U ). Then it is clear that the restriction γ V := (γp )p∈V
6(V ). We therefore have a restriction map ρU V (= ρU V,M ) :
belongs to M
6(U ) −→ M
M 6(V ) for which ρU V (γ) = γ V for all γ ∈ M 6(U ). Of
6 6
course, we define ρU ∅ : M (U ) −→ M (∅) = 0 to be the zero map.
 ) −→ R(V
It is easy to see that ρU V,R : R(U  ) is a homomorphism of
6 6  )-homomor-
R-algebras and that ρU V,M : M (U ) −→ M (V ) is an R(U
20.1 The Deligne Isomorphism 443

phism. Note also that the diagram

M
@ V
εU
M @ε M
@
ρU V
@
R
6(U )
M - M 6(V )

commutes. Statements similar to those in this paragraph hold if V = ∅.

The next exercise shows that the notion of local family of fractions, as in-
troduced in 20.1.6, can be viewed as a generalization of the notion of regular
function on a quasi-affine variety.

20.1.7 Exercise. Let V be an affine variety over the algebraically closed field
K, and take R := O(V ), S := R and T := Spec(R), as in Exercise 20.1.5;

let j : V −→ max T be the homeomorphism introduced in that exercise.
Let b be a non-zero ideal of O(V ), and let U be the non-empty open subset
{p ∈ V : f (p) = 0 for some f ∈ b} of V . We have j(U ) = max T \ Var(b).

(i) Recall that, for p ∈ U , we have OV,p = O(V )IV (p) = O(V )j(p) =
(S \ j(p))−1 R, and that each f ∈ O(U ) can be viewed as an element
of the ring OV,p . Thus we can define a ring homomorphism max ιU :
O(U ) −→ R(j(U )) by the assignment f → (f )m∈j(U ) . Show that
max ιU is an isomorphism.
(ii) Now let U be the open subset T \Var(b) of T = Spec(R). Given p ∈ U ,
−1
there exists p ∈ U such that p ⊆ j(p), so that (S \j(p)) R = Rj(p) can
be viewed as a subring of Rp = (S \ p)−1 R. Thus we can define a ring
homomorphism ιU : O(U ) −→ R(  U
 ) by the assignment f → (f )  .
p∈U
Show that ιU is an isomorphism.

In 2.3.2, we described, in terms of ideal transforms, the ring of regular func-


tions on a non-empty open subset of an affine algebraic variety over an al-
gebraically closed field. The last exercise therefore shows that, in one special
case at least, a ring of local families of fractions can be described in terms of
ideal transforms. The next major aim for this chapter is a description of general
modules of local families of fractions in terms of ideal transforms, and we now
embark on the preparations for this result.

20.1.8 Lemma. (The standard hypotheses of 20.1.2 apply.) Set T \Var(a) =:


U ; assume that U = ∅. Then
444 Links with sheaf cohomology

(i) for each γ = (γp )p∈U ∈ M 6(U ), there exist r ∈ N, s1 , . . . , sr ∈ S ∩ a


r  
and m1 , . . . , mr ∈ M such that U = i=1 T \ Var((si )) and γp =
mi /si for all p ∈ T \ Var((si )) (for i = 1, . . . , r);
6(U )) = 0; and
(ii) Γa (M
6
(iii) Ker(εUM ) = Γa (M ), where εM : M −→ M (U ) is the homomorphism
U

defined in 20.1.6(iii).
Proof. (i) As U is quasi-compact (see 20.1.1(iii)), there exist n ∈ N, W1 , . . . ,
(S) n
Wn ∈ UT , t1 , . . . , tn ∈ S and l1 , . . . , ln ∈ M such that U = j=1 Wj and,
for each j = 1, . . . , n and each p ∈ Wj , we have tj ∈ p and γp = lj /tj .
(S)
Let j ∈ {1, . . . , n}. Since Wj ∈ UT , there exists bj ∈ AS such that
Wj = T \ Var(bj ). Now Wj ⊆ U = T \ Var(a), and so
Wj = Wj ∩ U = (T \ Var(bj )) ∩ (T \ Var(a)) = T \ Var(abj ).
We therefore can, and do, assume that bj ⊆ a. Suppose that bj is generated by
wj1 , . . . , wjnj ∈ S ∩ a. Set Wjk := T \ Var((wjk )) for all k = 1, . . . , nj ;
n j
then Wj = k=1 Wjk , and, for k ∈ {1, . . . , nj } and all p ∈ Wjk , we have
tj wjk ∈ S \ p (so that Wjk = T \ Var((wjk )) = T \ Var((tj wjk ))) and we
can write
lj wjk lj
γp = = in (S \ p)−1 M.
tj wjk tj
We have therefore only to relabel the pairs
(wjk tj , wjk lj ) ∈ (S ∩ a) × M (k = 1, . . . , nj , j = 1, . . . , n)
in order to complete the proof of part (i).
(ii) Let γ = (γp )p∈U ∈ Γa (M6(U )). There exists n ∈ N such that an γ = 0.
Consider a p ∈ U = T \ Var(an ). Since an ⊆ p and an can be generated by
elements of S, there exists sp ∈ (S ∩an )\p. Since sp γ = 0, we have sp γp = 0
and we see that, in (S \ p)−1 M ,
1
γp = (sp γp ) = 0.
sp
Hence γ = 0.
6
M : M −→ M (U ) is an R-homomorphism, we have
(iii) As εU
6
M (Γa (M )) ⊆ Γa (M (U )) = 0
εU
(we have used part (ii) here), and so Ker(εUM ) ⊇ Γa (M ).
Now let m ∈ Ker(εU M ). Choose a family of denominators
7
(sp )p∈U ∈ (S \ p).
p∈U
20.1 The Deligne Isomorphism 445

Then 0 = εU M (m) = (sp m/sp )p∈U , and so, for each p ∈ U , there exists
tp ∈ S \ p such that tp sp m = 0.

Let b = p∈U Rtp sp ; then b is an ideal belonging to AS and bm = 0. As,
for each p ∈ U , we have tp sp ∈ p, it follows that U ∩ Var(b) = ∅, so that
T ∩ Var(b) ⊆ Var(a) and Var((S)) ∪ (T ∩ Var(b)) ⊆ Var(a). Therefore,
since T is essential with respect to S,
√ * * √
b= p⊇ p = a ⊇ a.
p∈Var((S))∪(T ∩Var(b)) p∈Var(a)

Thus there exists h ∈ N such that ah ⊆ b; therefore ah m ⊆ bm = 0 and


m ∈ Γa (M ).

20.1.9 Lemma. (The standard hypotheses of 20.1.2 apply.) Set T \Var(a) =:


(S)
U , assume that U = ∅, and let W ∈ UT be such that U ⊆ W . Consider
6(W ) −→ M
the restriction homomorphism ρW U : M 6(U ) of 20.1.6(iv). Then
6
Ker(ρW U ) = Γa (M (W )).

Proof. By 20.1.8(ii), we have ρW U (Γa (M6(W ))) ⊆ Γa (M6(U )) = 0, so that


Ker(ρW U ) ⊇ Γa (M 6(W )).
There exists c ∈ AS such that W = T \ Var(c). Let γ = (γp )p∈W ∈
Ker(ρW U ). By Lemma 20.1.8(i), there exist r ∈ N, s1 , . . . , sr ∈ S ∩ c
r
and m1 , . . . , mr ∈ M such that W = i=1 T \ Var((si )) and, for each
i = 1, . . . , r and each p ∈ T \ Var((si )), we have γp = mi /si . Since
γ ∈ Ker(ρW U ), we have γp = 0 for all p ∈ U = T \ Var(a).
Let i ∈ {1, . . . , r}. Set
 
U  := (T \ Var(a)) ∩ T \ Var((si )) = T \ Var(si a).

Then mi /si = 0 in (S \ p)−1 M for all p ∈ U  . This means that εU
M (mi ) = 0,
so that there exists hi ∈ N such that (si a) mi = 0 by 20.1.8(iii). Let h :=
hi

max{hi : i = 1, . . . , r}.
Now let p ∈ W . There exists i ∈ {1, . . . , r} with p ∈ T \ Var((si )), and
then, for all d ∈ ah , we have, in (S \ p)−1 M ,

dmi dshi mi
dγp = = h+1 = 0.
si si
6(W )).
Hence ah γ = 0, and γ ∈ Γa (M

20.1.10 Lemma. (The standard hypotheses of 20.1.2 are in force.) Set U :=


T \ Var(a). The homomorphism εU 6
M : M −→ M (U ) of 20.1.6(iii) has a-
torsion cokernel.
446 Links with sheaf cohomology

Proof. We can assume that U = ∅.


Let γ = (γp )p∈U ∈ M 6(U ). By 20.1.8(i), there exist r ∈ N, s1 , . . . , sr
r  
∈ S ∩ a and m1 , . . . , mr ∈ M such that i=1 T \ Var((si )) = U and, for
each i = 1, . . . , r and each p ∈ T \ Var((si )), we have γp = mi /si .
Let i ∈ {1, . . . , r}, and let Ui := T \ Var((si )). Then, for each p ∈ Ui , we
have si γp = si mi /si in (S \ p)−1 M . But this means that

ρU Ui (si γ) = si ρU Ui (γ) = εU U
M (mi ) = ρU Ui (εM (mi )),
i

so that si γ − εU 6
M (mi ) ∈ Ker(ρU Ui ) = Γ(si ) (M (U )) by Lemma 20.1.9. Hence
there exists ni ∈ N such that si (si γ − εM (mi )) = 0. Define
ni U

n := max{ni + 1 : i = 1, . . . , r}.

Then, for all i = 1, . . . , r, we have

sni γ = sn−1−n
i
i ni
si si γ = sn−1−n
i
i ni U
si εM (mi ) = εU n−1
M (si mi ) ∈ ε U
M (M ).

As
   
r r 
r
T \ Var Rsni = T \ Var Rsi = Ui = U = T \ Var(a),
i=1 i=1 i=1
r
it follows that T ∩ Var ( i=1 Rsni ) ⊆ Var(a). Therefore, since T is essential
with respect to S,
 * * √
r n
i=1 Rsi = p⊇ p= a ⊇ a.
r
p∈Var((S))∪(T ∩Var( n
i=1 Rsi )) p∈Var(a)

r
Thus there exists h ∈ N such that ah ⊆ i=1 Rsni , and
r r
ah γ ⊆ ( i=1 Rsni ) γ = i=1 Rsni γ ⊆ εU M (M ).

20.1.11 Lemma. (The standard hypotheses of 20.1.2 apply.) Set T \ Var(a)


(S)
=: U , and let W ∈ UT be such that U ⊆ W . Consider the restriction homo-
morphism ρW U : M6(W ) −→ M6(U ) of 20.1.6(iv). The R-module Coker ρW U
is a-torsion.

Proof. Since εUM = ρW U ◦ εM by 20.1.6(iv), we see that Im εM ⊆ Im ρW U


W U
U
and Coker ρW U is a homomorphic image of Coker εM . The claim therefore
follows from Lemma 20.1.10.

20.1.12 Remark and Notation. With the hypotheses of 20.1.2, let h : M →


(S)
N be a homomorphism of R-modules, and let ∅ = U ∈ UT .
20.1 The Deligne Isomorphism 447

(i) Now h induces, for each p ∈ U , an (S \ p)−1 R-homomorphism

(S \ p)−1 h : (S \ p)−1 M −→ (S \ p)−1 N.



There is therefore induced a p∈U (S \ p)−1 R-homomorphism
7 7 7
(S \ p)−1 h : (S \ p)−1 M −→ (S \ p)−1 N.
p∈U p∈U p∈U

It is easy to see that the image under this map of an S-local family of
fractions over U with numerators in M is an S-local family of fractions
 )-homomorphism
over U with numerators in N . Thus h induces an R(U
 6 
h(U ) : M (U ) −→ N (U ) for which

h(U ) ((γp )p∈U ) = ((S \ p)−1 h(γp ))p∈U 6(U ).
for all (γp )p∈U ∈ M

Of course, we define  h(∅) : M6(∅) −→ N  (∅) to be the zero homomor-


phism.
8
(ii) It is clear that Id M (U ) = IdM (U ) , and that, if g : N −→ L is another

homomorphism of R-modules, then g ◦ h(U ) = g(U ) ◦  h(U ). Thus


• (U ) is a covariant functor from C(R) to C(R(U  )). It is straightforward
to check that this functor is R-linear (note that an R(U  )-module can be
regarded as an R-module by means of εU
R ).
Clearly, similar comments can be made when U is replaced by ∅.
(S)
(iii) Now consider a second open set V ∈ UT such that V ⊆ U . It is clear
that (even if V = ∅) the diagram

6(U )
M
h(U )
-N
 (U )

ρU V,M ρU V,N

? 
?
6
M (V )
h(V )
-N
 (V )

commutes, and so ρU V : • (U ) −→ • (V ) is a natural transformation of


 ))). Also, ρ∅∅ is a natural transformation.
functors (from C(R) to C(R(U
(iv) Finally, the diagram
h
M - N

εU
M εU
N

? 
?
6(U )
M
h(U )
- N
 (U )
448 Links with sheaf cohomology

also commutes, and so εU : Id −→ • (U ) is also a natural transformation


of functors (from C(R) to itself). In addition, ε∅ : Id −→ • (∅) is a
natural transformation.

20.1.13 Exercise. Show that, in the situation of 20.1.12, the functor • (U ) :


 )) is left exact.
C(R) −→ C(R(U

We are now ready to establish the central result of this chapter; this result is
a version of the Deligne Isomorphism.

20.1.14 The Deligne Isomorphism Theorem. (The standard hypotheses of


20.1.2 apply.) Set U := T \ Var(a); assume that U = ∅.


6(U ) −→
(i) There is a unique R-isomorphism νa,M : M
=
Da (M ) such that
the diagram
εU
M
M
-M
6(U )
@
@
ηa,M

= νa,M
@
R ?
@
Da (M )

U
commutes. (Occasionally, νa,M will be written as νa,M when it is impor-
tant to stress the dependence on the open set U .) Moreover, if h : M −→
N is a homomorphism of R-modules, then the diagram

6(U )
M
h(U )
- N
 (U )


= νa,M ∼
= νa,N

? Da (h)
?
Da (M ) - Da (N )

commutes, and so νa : • (U ) −→ Da is a natural equivalence of func-


tors (from C(R) to itself).

 ) −→
(ii) The map νa,R : R(U
=
Da (R) is an isomorphism of R-algebras.

Proof. By 20.1.8(iii) and 20.1.10, both the kernel and cokernel of εUM are
a-torsion. It is therefore immediate from 2.2.15(ii) that there is a unique
20.1 The Deligne Isomorphism 449
6(U ) −→ Da (M ) such that the diagram
R-homomorphism νa,M : M

εU
M
M
-M
6(U )
@
ηa,M
@ νa,M
@
R ?
@
Da (M )

commutes, and, in fact, it follows from the formula for νa,M provided by
2.2.15(ii) that this map is monomorphic, since we know from 20.1.8(ii) that
6(U )) = 0. Note also that it follows from 2.2.17 that νa,R is a homomor-
Γa ( M
phism of R-algebras, and so part (ii) will follow from part (i).
To show that νa,M is surjective, let y ∈ Da (M ). Then there exists n ∈ N
and h ∈ HomR (an , M ) such that y is the natural image of h in Da (M ). As
a is generated by elements of S, there exists, for each p ∈ U , an element
sp ∈ (a ∩ S) \ p. Note that

  7
h(snp )
δ := ∈ (S \ p)−1 M.
snp p∈U p∈U

Now, for q ∈ U , we have, for each p ∈ U \ Var((sq )), that, in (S \ p)−1 M ,

h(snp ) snq h(spn ) h(snq snp ) snp h(snq ) h(snq )


= = = = ,
snp snq snp snq snp snp snq snq

6(U ). Our
so that δ is an S-local family of fractions over U , that is, δ ∈ M
immediate aim is to show that νa,M (δ) = y.
For each r ∈ an , we have
     
h(snp ) h(rsnp ) snp h(r)
rδ = r = = = εU
M (h(r)).
snp p∈U
snp p∈U
snp p∈U

Hence rνa,M (δ) = νa,M (rδ) = νa,M (εU M (h(r))) = ηa,M (h(r)), and this
is just the natural image in Da (M ) of h ∈ HomR (an , M ), where h (r ) =
r h(r) = rh(r ) for all r ∈ an . We thus see that rνa,M (δ) = ry for all r ∈ an .
Hence νa,M (δ) − y ∈ Γa (Da (M )), which is zero by 2.2.10(iv). Thus νa,M is
surjective, and so is an isomorphism.
450 Links with sheaf cohomology

To prove the second part, we wish to show that, in the diagram


h
M - N
B@ εM U B@ εUN
B @@ R B @@ R
B 6 B- 
ηa,M B M (U )  N (U )
B
h(U )
B
B ηa,N
B
B
νa,M B νa,N

BBN B
? Da (h)
BN ?
Da (M ) - Da (N ) ,

the front square commutes. We know from the first part of this proof that
the two side triangles commute; furthermore, the top square commutes by
20.1.12(iv), while the sloping rectangle on the underside commutes because
ηa is a natural transformation. Therefore

νa,N ◦ 
h(U ) ◦ εU
M = νa,N ◦ εN ◦ h = ηa,N ◦ h
U

= Da (h) ◦ ηa,M = Da (h) ◦ νa,M ◦ εU


M.

However, by 2.2.13(ii), 20.1.8(iii) and 20.1.10, there is a unique R-homomor-


6(U ) → Da (N ) such that the diagram
phism h : M
εU
M
M
-M
6(U )

h h
? ηa,N
?
N - Da (N )

commutes, and so νa,N ◦ 


h(U ) = Da (h) ◦ νa,M , as required.

The observant reader might have noticed that, in the Deligne Isomorphism
Theorem 20.1.14, we did not consider the case where U = ∅, that is, where
T ⊆ Var(a). The next exercise shows that, in some circumstances, there is a
‘Deligne Isomorphism’ in the case where U = ∅.

20.1.15 Definition and Exercise. (The standard hypotheses of 20.1.2 ap-


ply.) We say that T is large with respect to S if and only if, whenever c ∈ AS
is such that T ⊆ Var(c), then Var(c) = Spec(R), that is, c is nilpotent. Sup-
pose that this is the case and that T ⊆ Var(a). Show that

(i) Ker(ε∅M ) = Γa (M ) (compare 20.1.8(iii));


(S)
(ii) if W ∈ U , then Ker(ρW ∅ ) = Γa (M 6(W )) (compare 20.1.9); and
T
20.1 The Deligne Isomorphism 451

(iii) there is a natural equivalence of functors νa∅ : • (∅) −→ Da (compare


20.1.14(i)).

In several subsequent results that depend on the Deligne Isomorphism The-


orem 20.1.14, the reader will find the hypothesis (about U := T \ Var(a)) that
‘U = ∅ (or U = ∅ and T is large with respect to S)’. The facts that the results
concerned are still valid under the alternative hypothesis in parentheses are in
most cases easy consequences of Exercise 20.1.15.

20.1.16 Exercise. Consider the special case of the situation of 20.1.12 in


which R = K[X, Y ] is the polynomial ring in two indeterminates over a field
K, and take S = R and T = Spec(R), so that T is essential with respect to
S (see 20.1.3(i)). Let U := T \ Var((X, Y )). Use the natural homomorphism
R → R/XR to show that the functor • (U ) : C(R) −→ C(R(U  )) is not exact.

The following exercise shows that Theorem 20.1.14 can indeed be viewed
as a natural generalization of the isomorphism established in 2.3.2.

20.1.17 Exercise. Consider again the situation of 20.1.5 and 20.1.7, so that V
is an affine variety over the algebraically closed field K, R := O(V ), S := R,
T := Spec(R), b is a non-zero ideal of O(V ), U is the open subset of V
determined by b, and U  is the open subset T \ Var(b) of T = Spec(R).

Let j : V −→ max T be the homeomorphism of 20.1.5, and consider the
ring isomorphisms ιU : O(U ) −→ R(  U ) and max ιU : O(U ) −→ R(j(U
 ))
 j(U )
of 20.1.7. Show that both νb,R ◦ ιU : O(U ) −→ Db (R) and νb,R ◦ max ιU :
U

O(U ) −→ Db (R) coincide with the isomorphism νV,b of 2.3.2.

20.1.18 Exercise. Consider again the situation and notation of the Deligne
Isomorphism Theorem 20.1.14.
 )-module via νa,R , the
(i) Show that, when Da (M ) is regarded as an R(U

6(U ) −→ Da (M ) is an R(U
map νa,M : M
=  )-isomorphism.
(ii) Let b ∈ AS be such that a ⊆ b, and set W := T \ Var(b). Show that the
diagram
ρW U
6(W )
M -M
6(U )


= νb,M ∼
= νa,M

? αb,a,M
?
Db (M ) - Da (M ) ,

in which αb,a,M is the natural map of 2.2.23, commutes.


452 Links with sheaf cohomology

20.1.19 Exercise. Consider once more the situation and notation of the
Deligne Isomorphism Theorem 20.1.14.
(i) Show that, if a contains a poor M -sequence of length 2, then the homo-
6
morphism εU M : M −→ M (U ) of 20.1.6(iii) is an isomorphism.
(ii) Deduce that, if S contains a unit of R, then εTM : M −→ M 6(T ) is an
isomorphism.
20.1.20 Exercise. Let V be an affine variety over the algebraically closed
field K, take R := O(V ), let (as in 20.1.2) S be a non-empty subset of R
which is closed under multiplication, and let T be a subset of Spec(R) which
is essential with respect to S. Let b be a non-zero ideal of AS , and let U =
T \ Var(b); assume that U = ∅.
Show that the functor • (U ) : C(R) −→ C(R(U  )) is exact if and only if
the open subset V \ V (b) of V determined by b (the notation is as in 6.4.1) is
affine.

20.2 The Graded Deligne Isomorphism


We now intend to ‘add graded frills’ to the Deligne Isomorphism Theorem
20.1.14: we shall call the refined version the ‘Graded Deligne Isomorphism
Theorem’. One can view this refinement process as analogous to the improve-
ment, in the case of rings of regular functions on affine varieties, in Theorem
2.3.2 afforded, in the Z-graded case, by Proposition 15.1.3. For this work, we
introduce the concept of ‘homogeneous S-local family of fractions’.
20.2.1 Hypotheses for the section. The standard hypotheses of 20.1.2 will
be in force throughout this section, and, in addition, we shall assume that G

is a finitely generated torsion-free Abelian group and that R = g∈G Rg

is G-graded, that M = g∈G Mg is a graded R-module, and that S consists
entirely of homogeneous elements (so that that all the ideals in AS are graded).
20.2.2 Remark and Notation. (The hypotheses of 20.2.1 apply.) Let ∅ =
(S)
U ∈ UT .
(i) For each p ∈ U , the non-empty set S \ p consists entirely of homo-
geneous elements, so that, in the light of 13.1.1, the ring (S \ p)−1 R
carries a natural G-grading, and the (S \ p)−1 R-module (S \ p)−1 M
is also naturally graded; note also that the natural ring homomorphism
R −→ (S \ p)−1 R is homogeneous in the sense of 14.1.3. For g ∈ G,
we use ((S \ p)−1 R)g and ((S \ p)−1 M )g to denote the components of
degree g of (S \ p)−1 R and (S \ p)−1 M respectively.
20.2 The Graded Deligne Isomorphism 453

(ii) Let g ∈ G. The elements of p∈U ((S \ p)−1 M )g are called families
of homogeneous fractions of degree g over U with numerators in M
6(U ) ∩  −1
and denominators in S. The set M p∈U ((S \ p) M )g of such
6
families which are also local will be denoted by M (U )g ; the members of
6(U )g are called homogeneous S-local families of fractions of degree
M
g over U . Of course, we set M 6(∅)g = 0.
(iii) Let g, h ∈ G. It is easy to check that R(U )0 is an R0 -subalgebra of
 6 
R(U ), that M (U )g is an R(U )0 -submodule of M 6(U ), and that σγ ∈
6  6
M (U )g+h for all σ ∈ R(U )g and all γ ∈ M (U )h . An important aim for
 )g (g ∈ G) provide a G-grading on R(U
us is to establish that the R(U  ),
and that the M 6(U )g (g ∈ G) provide a grading on the R(U  )-module
6(U ). Lemma 20.2.3 below provides a key for the establishment of this
M
aim.
 −1
(iv) Let γ = (γp )p∈U ∈ p∈U (S \ p) M , and let g ∈ G. Then, for
each p ∈ U , the element γp ∈ (S \ p)−1 M has g-th component, de-
noted by (γp )g , in ((S \ p)−1 M )g . We shall refer to the family γg :=

((γp )g )p∈U ∈ p∈U ((S \ p)−1 M )g as the g-th homogeneous part of
the family γ.
(S)
20.2.3 Lemma. (The hypotheses of 20.2.1 apply.) Let ∅ = U ∈ UT and let
6(U ). Then
γ = (γp )p∈U ∈ M

(i) γg := ((γp )g )p∈U ∈ M6(U )g for each g ∈ G;


(ii) the set {g ∈ G : γg = 0} is finite;

(iii) γ = g∈G γg ; and
(iv) γ = 0 if and only if γg = 0 for all g ∈ G.
Proof. Because U is quasi-compact (see 20.1.1(iii)), there exist r ∈ N, W1 ,
(S) r
. . . , Wr ∈ UT , t1 , . . . , tr ∈ S and l1 , . . . , lr ∈ M such that U = j=1 Wi
and, for each j = 1, . . . , r and each p ∈ Wj , we have tj ∈ p and γp = lj /tj .
For each j = 1, . . . , r, let hj ∈ G be such that tj ∈ Rhj (recall that S consists
of homogeneous elements of R), and let (lj )g denote the g-th homogeneous
component of lj (for all g ∈ G).
(i) It is obvious that, for each j = 1, . . . , r,
 
lj (lj )g+hj
(γp )g = = for all p ∈ Wj .
tj g tj

Hence the family γg = ((γp )g )p∈U is local, and so lies in M 6(U )g .


(ii) For each j = 1, . . . , r, let Hj denote {g ∈ G : (lj )g+hj = 0}, a finite
set, and observe that (γp )g = 0 for all g ∈ G \ Hj and all p ∈ Wj . Hence
γg = 0 for all g ∈ G \ (H1 ∪ · · · ∪ Hr ).
454 Links with sheaf cohomology

(iii) As γp = g∈G (γp )g for each p ∈ U , this is immediate from part (ii).
(iv) Just note that γ = 0 if and only if γp = 0 for all p ∈ U , and that this is
the case if and only if (γp )g = 0 for all p ∈ U and all g ∈ G.

20.2.4 Remark and Definition. (The hypotheses of 20.2.1 apply.) Let ∅ =


(S)  )g
U ∈ UT . It is now immediate from 20.2.2(iii) and 20.2.3 that R(U
g∈G
 ), and that M
provides a grading on the ring R(U 6(U )g 6(U )
provides M
g∈G
 )-module: we refer to these gradings as the
with the structure of a graded R(U
natural gradings, and any unexplained references to gradings on R(U  ) and
6
M (U ) should always be interpreted as references to these natural gradings.

Trivially, similar conclusions apply to R(∅) 6(∅).
and M
(S)
20.2.5 Remarks. (The hypotheses of 20.2.1 apply.) Let U ∈ UT .
6
M : M −→ M (U ) of 20.1.6(iii) is homoge-
(i) It is clear that the map εU
neous.
(S)
(ii) Let V ∈ UT with V ⊆ U . It is also clear that the restriction map
ρU V (= ρU V,M ) : M 6(U ) −→ M
6(V ) of 20.1.6(iv) is homogeneous.
(iii) Now let N be a second graded R-module and let h : M −→ N be
a homogeneous R-homomorphism. It is easy to check that the R(U  )-
 6 
homomorphism h(U ) : M (U ) −→ N (U ) of 20.1.12(i) is homoge-
neous.
(S)
20.2.6 Exercise. (The hypotheses of 20.2.1 apply.) Let U ∈ UT and g ∈
G. This exercise involves the g-th shift functor described in 13.1.1.

(i) Use the fact that (S \ p)−1 (M (g)) = ((S \ p)−1 M )(g) for all p ∈ U to

show that M (g)(U ) = M6(U )(g).
(S)
(ii) Let V ∈ UT with V ⊆ U . Show that ρU V,M (g) = ρU V,M (g), that is,

that the restriction homomorphism ρU V,M (g) : M 
(g)(U ) −→ M (g)(V )
6
is the g-th shift of the restriction homomorphism ρU V,M : M (U ) −→
6(V ).
M
(iii) Let N be a second graded R-module and let h : M −→ N be a homo-
8 )=
geneous R-homomorphism. Show that h(g)(U h(U )(g).

Our promised Graded Deligne Isomorphism Theorem can now be obtained


very quickly.

20.2.7 The Graded Deligne Isomorphism Theorem. (The hypotheses of


20.2.1 apply.) Define U := T \ Var(a); assume that U = ∅.

Then the Deligne Isomorphism νa,M (= νa,MU 6(U ) −→
) : M
=
Da (M ) of
20.3 Links with sheaf theory 455

20.1.14 is a homogeneous R-homomorphism (with respect to the natural grad-


6(U ) and the grading of 13.3.14 on Da (M )).
ing of 20.2.4 on M
6(U ) is graded by 20.2.4, and εU : M −→ M
Proof. Since a is graded, M 6(U )
M
is homogeneous by 20.2.5(i), this result is now immediate from 13.5.4(ii) and
the Deligne Isomorphism Theorem 20.1.14.
20.2.8 Remarks. (The hypotheses of 20.2.1 apply.) Set T \ Var(a) =: U .

(i) It is clear from 20.2.4 and 20.2.5(iii) that the functor • (U ) : C(R) −→
 )) has the *restriction property of 13.3.6, and (by 20.2.5(i)) that
C(R(U
the natural transformation εU : Id −→ • (U ) has the *restriction prop-
erty of 13.3.7.
(S)
(ii) Let V ∈ UT with V ⊆ U . By 20.2.5(ii), the natural transformation of
 ))) of 20.1.12(iii)
functors ρU V : • (U ) −→ • (V ) (from C(R) to C(R(U
has the *restriction property.
(iii) Assume now that U = ∅ (or U = ∅ and T is large with respect to
S). We see from 20.2.7 (or 20.1.15) that νa,M is a homogeneous R-
isomorphism; by 20.1.18(i), it is also an R(U )-isomorphism; since the
6
natural grading on M (U ) is a grading of this as an R(U )-module, it is
therefore automatic that the grading of 13.3.14 on Da (M ) is a grading
 )-module. Hence Da : C(R) −→ C(R(U
of Da (M ) as an R(U  )) has the
*restriction property, and νa : • (U ) −→ Da , when viewed as a natural
equivalence of functors from C(R) to C(R(U  )) (or, for that matter, from
C(R) to itself), has the *restriction property.

20.3 Links with sheaf theory


Most serious readers of this chapter will by now have realised that, in view
of the results we have obtained about local families of fractions, we have es-
sentially started to discuss sheaves, even if only implicitly. We now intend to
make the connection more explicit, and to formulate 20.1.14 and 20.2.7 in
sheaf-theoretic terms. Henceforth, we assume that the reader has some famil-
iarity with basic knowledge about schemes and sheaves, although we provide
numerous references to Hartshorne’s book [30].

20.3.1 Remarks and Exercise: sheaf-theoretic interpretations. (The stan-


dard hypotheses of 20.1.2 apply.)
 )
(i) It is not difficult to check that the family R(U , together with
(S)
U ∈UT
456 Links with sheaf cohomology
 ) −→ R(V
the restriction maps ρU V,R : R(U  ) (U, V ∈ U (S) with V ⊆
T
U ) defines a sheaf R  of R-algebras on the topological space T , so that
 is a ringed space (see [30, Chapter II, §2, p. 72]).
(T, R)
(ii) Similarly, the family M 6(U ) , together with the restriction maps
(S)
U ∈UT
ρU V,M : M 6(U ) −→ M 6(V ) (U, V ∈ U (S) with V ⊆ U ), defines a
T
sheaf M6 of R-modules
 on T (see [30, Chapter II, §5, p. 109]). We shall
6 the S-sheaf induced by M . We shall sometimes use
call this sheaf M
standard notation from sheaf theory (see [30, Chapter II, §1, p. 61]) and
denote M 6(U ) by Γ(U, M6): the elements of Γ(U, M6) are the sections of
the sheaf M6 over U .
(iii) Now let h : M −→ N be a homomorphism of R-modules. It fol-
lows from 20.1.12(i),(iii) that the homomorphisms  h(U ) : M6(U ) −→
 (S)  6 
N (U ) (U ∈ UT ) define a morphism h : M −→ N of sheaves of R- 
modules (see [30, Chapter II, §5, p. 109]), which we call the morphism
induced by h.
(iv) It is now clear that • is a functor from C(R) to the category S(R)  of

sheaves of R-modules.
(v) Let p ∈ T . The stalk R p := lim R(U  ) of the sheaf R  at p (see [30,
−→
(S)
U ∈UT ,p
Chapter II, §1, p. 62]) has a natural structure as an R-algebra. Also, the
6p := lim M
stalk M 6(U ) of M6 at p has a natural structure as an Rp -
−→
(S)
U ∈UT ,p
(S)
module. For U ∈ UT,p , we shall use ρU,p (= ρU,p,M ) : M 6(U ) −→ M 6p
to denote the natural map; thus ρU,p,R is a homomorphism of R-algebras
 )-homomorphism.
and ρU,p,M is an R(U
p
The universal property of direct limits leads to a map ψM :M6p −→
−1 (S)
(S \ p) M for which ψM (ρU,p,M (γ)) = γp for all U ∈ UT,p and γ =
p

(γq )q∈U ∈ M6(U ). Show that ψ p is an isomorphism of R-algebras, and


R
that, as M varies through C(R), the ψM p
constitute a natural equivalence
−1
ψ : • p −→ (S \ p) of functors from C(R) to itself.
p

Deduce that the functor • : C(R) −→ S(R)  of part (iv) is exact (see
[30, Chapter II, Exercise 1.2(c), p. 66]).

20.3.2 Exercise. (The standard hypotheses of 20.1.2 apply.) Let h : L −→


N be a homomorphism of R-modules. Assume that T = ∅.
6 is zero if and only if the R-module M is
(i) Show that the induced sheaf M
(S)-torsion.
(ii) Show that the induced morphism h:L  −→ N  of sheaves of R-modules

20.3 Links with sheaf theory 457

is injective (respectively surjective) if and only if Ker h (respectively


Coker h) is (S)-torsion.
20.3.3 Remarks. (The standard hypotheses of 20.1.2 are in force.) Let U :=
T \ Var(a); assume that U = ∅ (or U = ∅ and T is large with respect to S).
Here, among other things, we reformulate the Deligne Isomorphism Theorem
in the language of sheaves.
(i) The functor • (U ) : C(R) −→ C(R(U  )) of 20.1.12(ii) can be regarded
as the composition of the functor • : C(R) −→ S(R)  of 20.3.1(iv)
and the section functor Γ(U, • ) : S(R)  −→ C(R(U  )). Therefore, the
Deligne Isomorphism Theorem 20.1.14, with the refinement afforded by

=
Exercise 20.1.18(i), provides a natural equivalence νa : Γ(U, • ) −→ Da
of functors from C(R) to C(R(U  )). Of course, we can interpret both Da
and Γ(U, • ) as functors from C(R) to itself (strictly, we should then
write Γ(U, • ) R in the latter case, but we shall often omit the ‘ R ’ in
the interests of notational simplicity), so that νa can also be interpreted
as a natural equivalence of functors from C(R) to itself.
(ii) The above interpretation can be refined in the case of the Graded Deligne
Isomorphism Theorem 20.2.7, for which we assume, in addition, that R
is G-graded, where G is a finitely generated torsion-free Abelian group,
and that S consists entirely of homogeneous elements. Then, in view of
20.2.4, the sheaf R  becomes a sheaf of G-graded R-algebras, and the
functor • : C(R) −→ S(R)  of 20.3.1(iv) restricts to a functor from
 
*C(R) to the category *S(R) of sheaves of graded R-modules. We shall
again denote this functor by • , as we do not expect this to cause confu-
sion. It follows from 20.2.8 that Γ(U, • ), considered as a functor from
C(R) to either C(R) or C(R(U  )), has the *restriction property, and that,

=
also in the two cases, the natural equivalence νa : Γ(U, • ) −→ Da of
part (i) has the *restriction property.  
 )0
(iii) Note that, with the notation of part (ii), R(U (S) defines a sheaf
U ∈UT
   
R0 = R  of R0 -algebras, so that T, R 0 is also a ringed space. Sim-
0  
ilarly, for a graded R-module M and g ∈ G, M 6(U )g (S) defines a
U ∈UT
 
sheaf M 6g = M 6 of R 0 -modules.
g
(iv) Again with the notation of parts (ii) and (iii), for each p ∈ U , the stalk
6p inherits a structure as a G-graded R-module for which the natural
M
homomorphism ρU,p,M : M 6(U ) −→ M6p is homogeneous. It is straight-

forward
 to check  that there is a natural isomorphism of R0 -modules
M6g ∼ M6 p g for all g ∈ G. We shall use these isomorphisms as
p =
identifications.
458 Links with sheaf cohomology

A special case of 20.3.3(i) yields the ‘classical’ form of Deligne’s Isomor-


phism.

20.3.4 Example: Deligne’s Isomorphism for Affine Schemes. Take S = R


and T := Spec(R) (which is essential with respect to S by 20.1.3(i)). Then
 of 20.3.1(i) is just the structure sheaf OT of the affine
the induced sheaf R
scheme (T, OT ) defined by R (see [30, Chapter II, §2, p. 70]). Moreover, if
M is an R-module, then the induced sheaf M 6 of 20.3.1(ii) is just the sheaf
of OT -modules associated to M (see [30, Chapter II, §5, p. 110]). It follows
from [30, Chapter II, §5, p. 111, and Proposition 5.4, p. 113] that the induced
sheaves in the sense of 20.3.1(ii) are, up to isomorphism, precisely the quasi-
coherent sheaves of OT -modules, and that the sheaves which are induced by
finitely generated R-modules are, up to isomorphism, precisely the coherent
sheaves of OT -modules.
An arbitrary ideal b of R automatically belongs to AS in this case. Let U :=
T \ Var(b) = Spec(R) \ Var(b). In this special case, T is large with respect to

S, and the isomorphism νb,M : Γ(U, M 6) −→=
Db (M ) of 20.1.14 (or 20.1.15)
and 20.3.3(i) is just the classical isomorphism of Deligne (see [30, Chapter III,
Exercise 3.7, p. 217]).

20.3.5 Exercise. Let (T, OT ) be the affine scheme defined by R, and let
qcohT denote the category of all quasi-coherent sheaves of OT -modules. Let
M be an R-module. Use Exercise 20.1.19 to show that εTM : M −→ Γ(T, M 6)
is an isomorphism, and deduce that the functor • : C(R) −→ qcohT is an
equivalence of categories. (See [30, Chapter II, Corollary 5.5, p. 113].)

20.3.6 Exercise. (The standard hypotheses of 20.1.2 apply.) Let b be a sec-


ond ideal of AS . Set U := T \ Var(a) and Z := T ∩ Var(b). Set
6) := {γ ∈ Γ(U, M
ΓZ (U, M 6) : ρU,p,M (γ) = 0 for all p ∈ U \ Z}.

(See [30, Chapter II, Exercise 1.14, p. 67, and Exercise 1.20, p. 68].) Note that
ΓZ (U, • ) is a functor from C(R) to C(R), and also a functor from C(R) to
 )).
C(R(U
Now assume that U = ∅ (or U = ∅ and T is large with respect to S).
Use Lemma 20.1.10 to show that the restriction of the isomorphism νa,M :

Γ(U, M6) −→ =
Da (M ) of 20.1.14 and 20.3.3(i) provides an isomorphism

6) −→
νa,b,M : ΓZ (U, M
=
Γb (Da (M )),

and deduce that the functors ΓZ (U, • ) and Γb (Da ( • )), whether considered as
 )), are naturally equivalent.
functors from C(R) to itself or from C(R) to C(R(U
20.3 Links with sheaf theory 459

20.3.7 Example. Let (T, OT ) be the affine scheme defined by R. (We in-
terpret S as R here and T as Spec(R), as in 20.3.4, so that R as well as a
belongs to AS .) Set Z := Var(a) and U := T \ Var(a) = Spec(R) \ Z. If we
apply 20.3.6 to the open set T and the closed set Z, we obtain an isomorphism

νR,a,M : ΓZ (T, M 6) −→ =
Γa (DR (M )). Bear in mind that ηR,M : M −→
DR (M ) is an isomorphism. Also bear in mind 20.1.15. The diagram
ρT U
0 - Γ (T, M
6) - Γ(T, M
6) - Γ(U, M
6)
Z

μa,M ∼
= (εT −1 ∼
= νa,M ∼
=
M)

? ? ηa,M
?
0 - Γa (M ) - M - Da (M ) - H 1 (M ) - 0,
a

in which μa,M = Γa ((ηR,M )−1 ) ◦ νR,a,M , the second map in the top row is
the inclusion map, and the lower row comes from 2.2.6(i)(c), has exact rows
and commutes.
20.3.8 Exercise. Let (T, OT ) be the affine scheme defined by R, let Z be
a closed subset of T = Spec(R), and let U := Spec(R) \ Z. Let F be a
coherent sheaf of OT -modules (see Example 20.3.4). Show that the restriction
map ρT U : Γ(T, F) −→ Γ(U, F) associated with the sheaf F is injective if
and only if, for all p ∈ Z, the stalk Fp of F at p has positive depth. (Remember
that the depth of a zero module over a local ring is interpreted as ∞.) Show
further that ρT U is bijective if and only if depth Fp > 1 for all p ∈ Z.
The following exercise and reminders are in preparation for an extension of
the sheaf-theoretic version of Deligne’s Isomorphism, as presented in 20.3.3,
to higher cohomology. This work will lead to what we shall call the Deligne
Correspondence.
20.3.9 Exercise. (The standard hypotheses of 20.1.2 apply.)
(i) Let I be an injective R-module. Show that the S-sheaf induced by I is
flasque, that is (see [30, Chapter II, Exercise 1.16, p. 67]), for every pair
(S)
of open sets U, V ∈ UT with V ⊆ U , the restriction map ρU V,I :
Γ(U, I ) −→ Γ(V, I ) is surjective. (You might find 20.1.18(ii) helpful.)

(ii) Now assume in addition that R = g∈G Rg is G-graded, where G is a
finitely generated torsion-free Abelian group, and that S consists entirely
of homogeneous elements. Let J be a *injective graded R-module. Show
that the S-sheaf induced by J is flasque.
(iii) Let R, S and J be as in part (ii), and let g ∈ G. Thus J is a sheaf of

graded R-modules. Now (R)  0 is a sheaf of R0 -algebras and the g-th
460 Links with sheaf cohomology
 
homogeneous component J g of this sheaf carries a natural structure
   
 -modules: show that J is flasque.
as a sheaf of R 0 g

20.3.10 Reminders. (The standard hypotheses of 20.1.2 apply.) Set U :=


T \ Var(a). Recall [30, Chapter II, Exercise 1.8, p. 66] that the section functor
Γ(U, • ) (from the category of sheaves of Abelian groups on T to the category
C(Z) of Abelian groups) is left exact: for each i ∈ N0 , the i-th right derived
functor of Γ(U, • ) is denoted by H i (U, • ) and is referred to as the i-th sheaf
cohomology functor on U . See [30, Chapter III, §2, p. 207]. Thus H i (U, • ) is
again a functor from the category of sheaves of Abelian groups on T to C(Z);

however, if F is a sheaf of R-modules on T , then, for all i ∈ N0 , the Abelian
 )-module which is such that,
group H (U, F) carries a natural structure as R(U
i

 ) and all σ ∈ H i (U, F), we have λσ = H i (U, λ IdF )(σ). It


for all λ ∈ R(U  
then follows that H i (U, • ) i∈N0 is a negative strongly connected sequence of
covariant functors from S(R) to C(R(U
 )).
We noted in 20.3.1(v) that the functor • : C(R) −→ S(R)  of 20.3.1(iv)
is exact.
 If we follow
 this with the above sheaf cohomology functors, we find
that H i (U, • ) i∈N0 is a negative strongly connected sequence of covariant
functors from C(R) to C(R(U  )). Of course, for an R-module M , we can regard
 i
the R(U )-module H (U, M 6) as an R-module by restriction of scalars: this
device
 i leads to a negative strongly connected sequence of covariant functors
H (U, • ) R i∈N0 from C(R) to itself (although we shall normally drop the
‘ R ’ from the notation).
We are now ready to present the promised Deligne Correspondence. This
correspondence presents a fundamental connection between sheaf cohomology
and local cohomology.
20.3.11 The Deligne Correspondence Theorem. (The standard hypotheses
of 20.1.2 apply.) Set U := T \ Var(a); assume that U = ∅ (or U = ∅ and T
is large with respect to S).
There is a unique isomorphism
    = 
∼ 
Θ = θi i∈N0 : H i (U, • ) i∈N0 −→ Ri Da i∈N0
of negative strongly connected sequences of covariant functors from C(R) to

=
itself for which θ0 is the natural equivalence νa : Γ(U, • ) −→ Da of 20.3.3(i).
Consequently, by 2.2.6(iii), for each i ∈ N, the functors H i (U, • ) and Hai+1
from C(R) to itself are naturally equivalent.
Proof. Let I be an injective R-module. Of course, Ri Da (I) = 0 for all
 = 0 for
i ∈ N. By 20.3.9(i), the induced S-sheaf I is flasque, so that H i (U, I)
20.3 Links with sheaf theory 461

all i ∈ N, by [30, Chapter III, Proposition 2.5, p. 208]. The claim is therefore
immediate from 1.3.4(ii).
20.3.12 Remark. In the situation of 20.3.11, we can, on account  of 20.1.18,

consider Da as a functor from C(R) to C(R(U )), so that R Da i∈N is actu-
i
0
ally a negative strongly connected sequence of covariant functors from C(R) to

 )); furthermore, νa : Γ(U, • ) −→
C(R(U
=
Da is actually a natural equivalence
 )) (see 20.3.3(i)).
of functors from C(R) to C(R(U
One can therefore argue as in the above proof of 20.3.11 to see that there is a
    = 
∼ 
unique isomorphism δ i i∈N0 : H i (U, • ) i∈N0 −→ Ri Da i∈N0 of negative
strongly connected sequences of covariant functors from C(R) to C(R(U  )) for
which δ 0 is νa ; moreover, the uniqueness aspect of Theorem 20.3.11 shows
that δ i R = θi for all i ∈ N0 , so that, for each such i and each R-module N ,
the maps δNi
and θNi
coincide and θN i  )-isomorphism.
is an R(U
20.3.13 A Graded Version of the Deligne Correspondence. Here, we pre-
sent refinements of 20.3.11 which are available in the graded case, and so we

assume, in addition to the hypotheses of 20.3.11, that R = g∈G Rg is G-
graded, where G is a finitely generated torsion-free Abelian group, and that S
consists entirely of homogeneous elements. In particular, this means that a is
graded. We consider the isomorphism
    = 
∼ 
Θ = θi i∈N0 : H i (U, • ) i∈N0 −→ Ri Da i∈N0
of negative strongly connected sequences of covariant functors from C(R)
 )) of 20.3.11 and 20.3.12: recall that θ0 is the natural equivalence
to C(R(U

=
νa : Γ(U, • ) −→ Da of 20.3.3(i), which, by 20.3.3(ii), has the *restriction
property.
(i) Whenever I is a *injective graded R-module, we have
H i (U, I ) ∼
= Ri Da (I) ∼
= Hai+1 (I) = 0 for all i ∈ N,
by 20.3.11, 2.2.6(iii) and 13.2.6. Since Γ(U, • ) has the *restriction prop-
erty (see 20.2.8(i)), we can use Theorem 13.3.15 to deduce that there is
 )-modules H i (U, N 
exactly one choice of gradings on the R(U  i  ) (i ∈ N,
N a graded R-module) with respect to which H (U, • ) i∈N0 has the
*restriction property of 13.3.9. We shall refer to these gradings as the
natural gradings, and any unexplained gradings on these modules are
to be interpreted as these natural ones. Notice that they provide grad-
ings over R, and so must be the unique choice of gradings with respect
to which H i (U, • ) i∈N0 , considered as a negative strongly connected
sequence of functors from C(R) to itself, has the *restriction property.
462 Links with sheaf cohomology

(ii) Similarly, Da : C(R) −→ C(R(U  )) has the *restriction property (see


20.2.8(iii)), and there is exactly one choice of gradings on the R(U  )-

 i  R Da (N ) (i ∈ N, N a graded R-module) with respect to which


i
modules
R Da i∈N0 has the *restriction property. We again refer to these grad-
ings as the natural gradings; the uniqueness aspect of Theorem 13.3.15
means that these natural gradings, which work over both R(U  ) and R,
must be the ones found in 13.5.7(ii).
Now θ0 has the *restriction property. We can define gradings on the
R Da (N ) (i ∈ N, N a graded R-module) in such a way that all the θN
i i

(i ∈ N, N a graded R-module) are homogeneous. Since Θ is an iso-


morphismof connected
 sequences, it follows that, with respect to these
gradings, Ri Da i∈N0 has the *restriction property, and so these grad-
ings must be the natural gradings.
Thus, with respect to the natural gradings, for all i ∈ N0 , the natural

=
equivalence θi : H i (U, • ) −→ Ri Da of 20.3.11 has the *restriction
property.

=
(iii) Since, by 13.5.7(iii), for i ∈ N, the natural equivalence Ri Da → Hai+1
of 2.2.6(ii) has the *restriction property, it follows that the natural equiv-

=
alence H i (U, • ) −→ Hai+1 of the final paragraph of 20.3.11 also has the
*restriction property. This gives a satisfactory extension, to the graded
case, of our fundamental connection between sheaf cohomology and lo-
cal cohomology.

Next, we are going to study the cohomology of homogeneous components of


induced S-sheaves in the graded situation. This theme has great importance in
the study of sheaf cohomology over projective varieties and, more generally,
over projective schemes. Our aim is to produce a link between cohomology
of homogeneous components of induced S-sheaves and homogeneous com-
ponents of graded ideal transforms and local cohomology modules. We begin
with some preparations.
20.3.14 Remarks and Notation. Here, the hypotheses of 20.2.1 apply. Let
g ∈ G, and set U := T \ Var(a); assume that U = ∅ (or U = ∅ and T is large
with respect to S).
 is a sheaf of G-graded R-algebras, and we
(i) In 20.3.3(ii), we noted that R
introduced the functor • from *C(R) to the category *S(R)  of sheaves

of graded R-modules; this functor is exact, by 20.3.1(v).
 −→ S((R)
We shall use ( • )g : *S(R)  0 ) to denote the functor which

associates to each sheaf F of graded R-modules its g-th component Fg ,

which is a sheaf of (R)0 -modules.
20.3 Links with sheaf theory 463

It is a consequence of Exercise 20.2.6 that the functors (• )g and


8
(•(g))0 (from *C(R) to S((R) 0 )) are equal. (Here, ( • )(g) : *C(R) −→
*C(R) denotes the g-th shift functor.)
(ii) It follows from the final paragraph of part (i) that, in C((R(U ))0 ), we
have Γ(U, (M 6)g ) = Γ(U, M
(g))0 ) = Γ(U, (M 6)g (for all M ∈ *C(R)).

In alternative notation, (M 6)g (U ) = M
(g))0 (U ) = (M 6(U )g for all M ∈

=
*C(R). The natural equivalence νa : Γ(U, • ) −→ Da of 20.3.3 has the
*restriction property, by 20.3.3(ii). Hence, on taking g-th components,
we see that

=
νa,g : Γ(U, • )g −→ Da ( • )g
is a natural equivalence of functors from *C(R) to C((R(U ))0 ), and also
of functors from *C(R) to C(R0 ).
 
(iii) Since the functor •8  0 ) is exact, it follows that
(g) 0 from *C(R) to S((R)
   
H i U, •8 (g)0
can be considered as a negative strongly con-
i∈N0
 ))0 )
nected sequence of covariant functors from *C(R) to either C((R(U
or C(R0 ). We propose, in the Serre–Grothendieck
 Correspondence

Theorem 20.3.15 below, to compare this with Ri Da ( • )g i∈N , which
0
also can be considered as a negative strongly connected sequence of co-
 ))0 ) or C(R0 ).
variant functors from *C(R) to either C((R(U
20.3.15 The Serre–Grothendieck Correspondence Theorem. For this, we
adopt the hypotheses of 20.2.1. Let g ∈ G, and set U := T \ Var(a); assume
that U = ∅ (or U = ∅ and T is large with respect to S).
There is a unique isomorphism
      = 
∼ 
Ωg = ωgi i∈N : H i U, •8 (g) 0 −→ Ri Da ( • )g i∈N
0
0 i∈N0

of negative strongly connected sequences of covariant functors from *C(R) to


 ))0 ) for which ωg0 is the natural equivalence
C((R(U
    ∼
νa,g : Γ U, 
=
• (g)
0
−→ Da ( • )g
of functors of 20.3.14(ii).
Proof. Let I be a *injective graded R-module. Let i ∈ N. Then
Ri Da (I) ∼
= Hai+1 (I) = 0,
8  
by 2.2.6(ii) and 13.2.6. Also, by 20.3.14(i), we have I(g) 0
= I g , and, by
 8 
 0 -modules. Hence H i U, I(g)
20.3.9(iii), this is a flasque sheaf of (R) =0
0
by [30, Chapter III, Proposition 2.5, p. 208].
464 Links with sheaf cohomology
 
 ) with the trivial grading, we can use 13.3.5(ii) to
If we now endow R(U 0
complete the proof.

20.3.16 Remarks. Here we consider further the situation of, and use the
notation of, the Serre–Grothendieck Correspondence 20.3.15.

(i) Observe that Ωg can be regarded as an isomorphism of negative strongly


connected sequences of covariant functors from *C(R) to C(R0 ).
(ii) Let d ∈ G and rd ∈ Rd . For each graded R-module N , let μrd =
μrd ,N : N −→ N (d) be the homogeneous R-homomorphism given by
multiplication by rd . Set u = g + d. It is straightforward to check that
 i   
H U, μr d, •
(g) 0 i∈N0 :
 i   8       
H U, • (g) 0 i∈N −→ H i U,  • (u)
0 i∈N0
0

is a homomorphism of connected sequences of functors from *C(R) to


C(R0 ).
  
(iii) Let i ∈ N0 . One can check that j∈G H i U, (M  (j))0 has a structure
   
as a graded R-module such that rd zg = H i U, μ rd (g) 0 (zg ) for (d ∈
   
G and) rd ∈ Rd and zg ∈ H i U, M  (g) . It is also straightforward to
0
use the uniqueness aspect of 13.3.5(i) to show that, with respect to this
graded R-module structure, the R0 -isomorphism
      ∼ 
i
ωj,M : H i U, M  (j) 0 −→
=
Ri Da (M )j = Ri Da (M )
j∈G j∈G j∈G

given by the Serre–Grothendieck Correspondence 20.3.15 (and part (i))


is actually a homogeneous isomorphism of graded R-modules. For i =
0, this is just the homogeneous R-isomorphism
     
0
ωj,M = νa,M : Γ U, M  6) = M
(j) 0 = Γ(U, M 6(U )
j∈G j∈G

=

−→ Da (M )j = Da (M )
j∈G

of 20.3.3(ii).
(iv) Let i ∈ N. It follows from part (iii) and 13.5.7(iii) that there is a homoge-
     ∼
neous isomorphism j∈G H i U, M  =
(j) 0 −→ Hai+1 (M ) of graded
R-modules.

20.3.17 Exercise. Here the hypotheses of 20.1.2 and 20.2.1 apply. Consider
a p ∈ T.
20.4 Applications to projective schemes 465

(i) Show that the isomorphism of R-algebras ψR p


:Rp −→ (S \ p)−1 R of
20.3.1(v) is homogeneous, and that the natural equivalence

ψ p : • p −→ (S \ p)−1

of functors from C(R) to itself (again of 20.3.1(v)) has the *restriction


property.
(ii) Let U := T \ Var(a) and assume that p ∈ U . Show that there is a
commutative diagram with homogeneous maps
εU
6(U ) 
M
M
M

ρU,p,M φM
S

? p ?
ψM
6p
M ∼
- (S \ p)−1 M
=

in which φM
S is the canonical map and the homogeneous homomor-
phisms εU
M and ρU,p,M are defined by 20.1.6(iii) and 20.3.1(v) respec-
tively.

20.4 Applications to projective schemes


The Serre–Grothendieck Correspondence 20.3.15, as exploited in 20.3.16, can
be used in a very effective manner, in various situations, to derive results about
the cohomology of induced sheaves from purely algebraic results obtained ear-
lier in the book. We have neither the space nor the intention to present a com-
prehensive approach to sheaf cohomology in this book, and so we shall content
ourselves, in this section, with some applications of the Serre–Grothendieck
Correspondence to the particular case of projective schemes induced by homo-
geneous, positively Z-graded (Noetherian) rings. This situation is particularly
fertile in this context because the quasi-coherent sheaves are, up to isomor-
phism, just the sheaves induced by graded modules.
We hope that the illustrations which we present in this section will convince
the reader of the value of the Serre–Grothendieck Correspondence, and will
whet her or his appetite for exploration of applications of the correspondence
in other situations.

20.4.1 Hypotheses for the section. We shall assume throughout this section

that R = n∈N0 Rn is positively Z-graded and homogeneous; we shall take
466 Links with sheaf cohomology

S := n∈N Rn , the set of homogeneous elements of R of positive degrees to-
gether with 0 (so that AS is the set of all graded ideals of R which are contained
in R+ (see 20.1.4)); we shall take T := Proj(R) = * Spec(R) \ Var(R+ ) and
we shall assume that T = ∅; also, we shall assume that the R-module M is
graded. Note that our assumptions imply that T is large (see 20.1.15) with
respect to S.
Recall from 20.1.4 that the S-topology on Proj(R) is the Zariski topology.
For p ∈ Proj(R), the natural ring homomorphism (S \ p)−1 R −→ R(p) is an
isomorphism.

20.4.2 Remarks. (The hypotheses of 20.4.1 apply.)

(i) The sheaf (R) 0 of R0 -algebras of 20.3.9(iii) is just the structure sheaf
OT of the projective scheme (T, OT ) defined by R: see [30, Chapter II,
§2, p. 76]. Note that, for each open subset U of T , we have OT (U ) =
 ))0 , and, for p ∈ T , the stalk of OT at p is (R
(R(U  0 )p = ( R
p )0 (we
are using 20.3.3(iv) here), and so 20.3.17(ii) yields an isomorphism of

=
R0 -algebras (ψR p
)0 : OT,p −→ (R(p) )0 . The reader should notice that
our use of the notation R(p) is different from Hartshorne’s in [30, p. 18].
From now on, we shall identify OT,p with (R(p) )0 by means of the above
isomorphism.
(ii) The sheaf (M6)0 of (R)  0 -modules of 20.3.14(i) is just the sheaf of OT -
modules associated to M on Proj(R), as defined by Hartshorne in [30,
Chapter II, §5, p. 116]. Although Hartshorne’s notation for this sheaf
is commonly used, we shall continue to use the notation (M 6)0 , in an
attempt to avoid confusing readers.
Note that, for each open subset U of T and t ∈ Z, we have (M 6)t (U ) =
6
M (U )t . Moreover, by 20.3.3(iv) and 20.3.17(i), for each p ∈ T and

t ∈ Z, there is an isomorphism of R0 -modules (ψM p 6)t )p −→
)t : ((M
=

(M(p) )t . We may consider ((M 6)t )p as an (R(p) )0 -module via the iso-
p −1
morphism ((ψR )0 ) of part (i). It then follows easily from 20.3.17(ii)
p
that the above isomorphism (ψM )t is an isomorphism of modules over
(R(p) )0 = OT,p . Observe also, that by 20.3.14(i) we have (M  (t))0 =
6
(M )t for all t ∈ Z. Consequently, (ψ p
)0 = (ψ )t for all t ∈ Z.
p
M (t) M

20.4.3 Exercise. Let the situation be as in 20.4.2(ii), and let p ∈ T and


t ∈ Z. Note that (M(p) )t denotes the t-th component of M(p) , the homoge-
neous localization of M at p. Use the diagram of 20.3.17(ii), the fact that the
maps in the diagrams in 20.1.14(i) are homogeneous, and the identifications
   

M (t) 0 = M6 and (ψ p )0 = (ψ p )t of 20.4.2(ii), to show that there is a
t M (t) M
20.4 Applications to projective schemes 467

commutative diagram
  T
(νR )t
- DR+ (M )t
M(t) 0 (T )
+ ,M

=

ρ  (β(p) )t
T ,p,(M (t))0

 ?
  p  ?
(ψM )t

M (t) 0 p ∼
- M(p)
= t

in which β(p) : DR+ (M ) −→ M(p) is defined as in 17.2.2.

20.4.4 Theorem: the classical form of the Serre–Grothendieck


Correspondence. (The hypotheses of 20.4.1 apply.) It follows from the Serre–
Grothendieck Correspondence 20.3.15 and 20.3.16(iii) (with a = R+ ) that
there are homogeneous R-isomorphisms
  ∼
= ∼
=
j∈Z H (T, (M (j))0 ) −→ R DR+ (M ) −→ HR+ (M ) for all i ∈ N
i i i+1

   ∼
6(T ) −→
6(T )j = M =
and j∈Z Γ(T, (M (j))0 ) = j∈Z M DR+ (M ). 

We recall now some facts about sheaves over projective schemes which
mean that the classical form of the Serre–Grothendieck Correspondence 20.4.4
is a powerful tool for translation of algebraic results about local cohomology
into geometric results about sheaf cohomology.

20.4.5 Reminders. (The hypotheses of 20.4.1 apply.) Let t ∈ Z.

(i) By [30, Chapter II, Proposition 5.11(c), p. 116], the sheaf (M 6)0 is a
quasi-coherent sheaf of OT -modules; furthermore, if M is finitely gen-
6)0 is a coherent sheaf of OT -modules.
erated, then (M
(ii) The twisted sheaf (M6)0 (t), as defined by Hartshorne [30, Chapter II, §5,
p. 117] is naturally isomorphic (as a sheaf of OT -modules) to (M (t))0 ,
the sheaf associated to the shifted module M (t): see [30, Chapter II,
Proposition 5.12(b), p. 117]
(iii) Let F be a sheaf of OT -modules. Recall from [30, Chapter II, §5, p. 118]

that Γ∗ (T, F) := Γ∗ (F) := j∈Z Γ(T, F(j)) carries a natural struc-
ture as a graded R-module and is called the graded R-module associated
to F. We shall use the notation Γ∗ (T, F) instead of Hartshorne’s Γ∗ (F).
(iv) Let F be a quasi-coherent sheaf of OT -modules. Then, by [30, Chapter
II, Proposition 5.15, p. 119], there is a natural isomorphism

Γ∗
=
(T, F) −→ F
0
468 Links with sheaf cohomology

of sheaves of OT -modules. This shows, in particular, that each quasi-


coherent sheaf F of OT -modules is isomorphic to (N  )0 for some graded
R-module N . Also, in this situation, the induced R0 -isomorphism
 
Γ∗ (T, F) := Γ(T, F(j)) ∼ = Γ(T, (N (j))0 )
j∈Z j∈Z

(which arises in view of (ii) above) is a homogeneous R-isomorphism.


This is easily seen by comparing, by means of the natural isomorphism
of part (ii), the action of a homogeneous element rd ∈ Rd (where d ∈
N0 ) on Γ∗ (T, F) as described in [30, Chapter II, §5, p. 118] with the
 
action of rd on j∈Z Γ(T, (N (j))0 ) as described in 20.3.16(iii).
20.4.6 Exercise. (The hypotheses of 20.4.1 apply.) Let F be a coherent
sheaf of OT -modules. Then by 20.4.5(iv) we can write F ∼ 6)0 , where M
= (M
is a graded R-module. Our aim here is to show that we can choose M to be
finitely generated. We do this in several steps.
(i) Show that (N  )0 = 0 for each R+ -torsion graded R-module N .
(ii) Show that we can replace M by M≥0 (see 16.1.1) and hence assume that
beg(M ) ≥ 0. (You may find 20.3.14(i) helpful.)
(iii) Write R = R0 [f1 , . . . , fr ] with fi ∈ R1 \ {0}. Let i ∈ {1, . . . , r}.
Observe that, by [30, Chapter II, Proposition 2.5, pp. 76,77], the open
set Ui := T \ Var(fi R) is (empty or) affine with O(Ui ) = (Rfi )0 , a
Noetherian ring. Use 20.3.14(ii) to show that Γ(Ui , F) ∼
= Dfi R (M )0 ∼ =
(Mfi )0 . Conclude by [30, Chapter II, Corollary 5.5 p. 113] that the
(Rfi )0 -module (Mfi )0 is finitely generated. This is true for each i =
1, . . . , r.
(iv) Deduce that there is a finitely generated graded R-submodule P of M
such that M/P is R+ -torsion.
(v) Now use part (i) to show that F ∼ = (P)0 .
20.4.7 Exercise. (The hypotheses of 20.4.1 apply.)
(i) Let b ⊆ R be a graded ideal. Show (by use of 20.3.14(i)) that the inclu-
sion map b −→ R yields a monomorphism of sheaves of OT -modules
J := (b)0 −→ (R)  0 = OT . This means that, by definition, J is a sheaf
of ideals on T (see [30, p. 109]). Conclude that this sheaf is coherent.
(ii) Let J be a coherent sheaf of OT -modules which is a sheaf of ide-
als on T , so that, by definition, there is a monomorphism of sheaves
J −→ OT . Use the exactness of the twisting functors on the category
of sheaves of OT -modules, the left exactness of the section functor and
the Serre–Grothendieck Correspondence (see 20.3.16(iii)) to show that
20.4 Applications to projective schemes 469

d := Γ∗ (T, J ) is homogeneously R-isomorphic to a graded ideal c of


the graded R-algebra DR+ (R). Let b ⊆ R be the inverse image of c
under the natural homomorphism of graded rings ηR : R −→ DR+ (R).
Show that there are isomorphisms of sheaves of OT -modules ( b)0 ∼
=
∼  ∼
(c)0 = (d)0 = J , so that J is induced by a graded ideal b of R.
We can now prove the fundamental Finiteness Theorem of Serre concerning
the cohomology of coherent sheaves over projective schemes.
20.4.8 Serre’s Finiteness Theorem. (See [77, §66, théorème 1 and théorème
2(b)] and [30, Chapter III, Theorem 5.2, p. 228].) (The hypotheses of 20.4.1
apply.) Let F be a coherent sheaf of OT -modules. Then
(i) H i (T, F(j)) is a finitely generated R0 -module, for all i ∈ N0 and all
j ∈ Z; and
(ii) there exists r ∈ Z such that H i (T, F(j)) = 0 for all i ∈ N and all
j ≥ r.
Proof. By 20.4.6, there exists a finitely generated graded R-module N such
that F ∼
= (N  )0 ; furthermore, by 20.4.5(iv), there is a homogeneous isomor-
phism of graded R-modules

Γ∗ (T, F) ∼
= Γ(T, (N (j))0 ).
j∈Z

Also, by 20.4.5(ii), for each i ∈ N0 and j ∈ Z, there is an isomorphism of


R0 -modules H i (T, F(j)) ∼ 
= H i (T, (N (j))0 ).
Let j ∈ Z and i ∈ N. We can now use the classical form of the Serre–
Grothendieck Correspondence 20.4.4 to see that there are R0 -isomorphisms
Γ(T, F(j)) ∼ = DR+ (N )j and H i (T, F(j)) ∼ = HR i+1
+
(N )j . Since DR+ (N )j
i+1
is a finitely generated R0 -module by 16.1.6(ii), and HR +
(N )j is a finitely
generated R0 -module by 16.1.5(i), part (i) is now proved. Part (ii) follows from
16.1.5(ii).
20.4.9 Exercise. Consider the situation and use the notation of Serre’s Finite-
ness Theorem 20.4.8. Show that, for all r ∈ Z, the graded R-module

Γ∗ (T, F)≥r := Γ(T, F(j))
j∈Z, j≥r

is finitely generated.
The philosophy of the above proof of Theorem 20.4.8 suggests that a similar
approach to Castelnuovo regularity of coherent sheaves of OT -modules might
be profitable. This is indeed the case.
470 Links with sheaf cohomology

20.4.10 Definition. (The hypotheses of 20.4.1 apply.) Let F be a coherent


sheaf of OT -modules, and let r ∈ Z.
We say that F is r-regular in the sense of Castelnuovo–Mumford if and only
if H i (T, F(s − i)) = 0 for all i ∈ N and all s ∈ Z with s ≥ r. In practice, the
phrase ‘in the sense of Castelnuovo–Mumford’ is usually omitted.
We define the (Castelnuovo–Mumford) regularity reg(F) of F by

reg(F) = inf {r ∈ Z : F is r-regular} .

20.4.11 Remarks. In the situation, and with the notation, of 20.4.10, there
exists a finitely generated graded R-module N such that F ∼ = (N  )0 , by 20.4.6.
Also, by 20.4.5(ii) and the classical form of the Serre–Grothendieck Corre-
spondence 20.4.4, for each i ∈ N and each j ∈ Z, there exists an R0 -isomorph-
ism H i (T, F(j)) ∼= HR i+1
+
(N )j ; in particular,

H i (T, F(s − i)) ∼ i+1


= HR +
(N )s+1−(i+1) for all i ∈ N and s ∈ Z.

This enables us to use results from Chapter 16 to make deductions about the
regularity of coherent sheaves of OT -modules.

(i) By Theorem 16.2.5, if H i (T, F(r − i)) = 0 for all i ∈ N, then F is


r-regular.
(ii) In view of Definition 16.2.1(ii), we can say that F is r-regular if and
only if N is r-regular at and above level 2.
(iii) Hence reg(F) = reg2 (N ) (see Definition 16.2.9).

An important property that a sheaf of OT -modules might have is that of


being generated by global sections. We shall show that this property is closely
related to the concept of regularity.

20.4.12 Definition. Let the notation and hypotheses be as in 20.4.10. We say


that the coherent sheaf F is generated by its global sections precisely when,
for each p ∈ T , the stalk Fp of F at p is generated ‘by germs of global sections
of F’, that is, if and only if

Fp = γ∈Γ(T,F ) OT,p ρT,p,F (γ) for all p ∈ T,

where ρT,p,F : Γ(T, F) −→ Fp is the natural map.

20.4.13 Serre’s Criterion for Generation by Global Sections. (See [77,


§66, Théorème 2].) The hypotheses of 20.4.1 apply. Let F be a coherent sheaf
of OT -modules and let t ∈ Z with t ≥ reg(F). Then the twisted sheaf F(t) is
generated by its global sections.
20.4 Applications to projective schemes 471
 
Proof. By 20.4.6, we may write F = M 6 for some finitely generated
0
 
graded R-module M . By 20.4.5(ii), there is an isomorphism F(t) ∼
= M 
(t) . 0
Then, by 17.2.2 and with the notation of that exercise, and in view of the iden-
tification of 20.4.2(i), we have
 
(M(p) )t = m∈S (R(p) )0 (β(p) )t (m) = m∈S OT,p (β(p) )t (m)

for some set S ⊆ DR+ (M )t . Moreover, by 20.4.3 and 20.4.5(ii), we have a


commutative diagram
  T
(νR )t
- DR+ (M )t
Γ(T, (F(t))) 
M (t) 0 (T )
+ ,M

=

ρT ,p,F (t) (β(p) )t

?    p  ?
(ψM )t
(F(t))p 
M (t) ∼
- M(p) .
0 p = t
 T −1
Set T := (νR )
+ ,M t
(S) ⊆ Γ(T, (F(t))). Then, since (ψM
p p
(t) )0 = (ψM )t
is an OT,p -isomorphism (by 20.4.2(ii)), it follows that

(F(t))p = γ∈T OT,p ρT,p,F (t) (γ),

and this proves the claim.

The above theorem extends a result of Serre [77, §66, Théorème 2] which
states that, if R is the homogeneous coordinate ring of a projective variety and
F is a coherent sheaf of OT -modules, then F(t) is generated by its global
sections for all t  0.

20.4.14 Definition and Exercise. (The hypotheses of 20.4.1 apply.) Let F


be a non-zero coherent sheaf of OT -modules. Recall that the support of F,
denoted by Supp F, is the set {p ∈ T : Fp = 0} (see [30, Chapter II, Exercise
1.14, p. 67]); the dimension of F, denoted by dim F, is defined to be the
dimension of Supp F (see 19.1.8).
By 20.4.6, there exists a finitely generated graded R-module N such that
F∼ = (N )0 . Show that

(i) Supp F = Supp N ∩ Proj(R) (a closed subset of Proj(R)), and


(ii) when R0 is Artinian, dim F = dim N − 1 (so that dim F is finite) (you
might find 19.7.1(i),(iv) helpful here).

20.4.15 Notation. (The hypotheses of 20.4.1 apply.) Assume that R0 is


Artinian and let F be a coherent sheaf of OT -modules.
472 Links with sheaf cohomology

By Serre’s Finiteness Theorem 20.4.8, for each i ∈ N0 and n ∈ Z, the R0 -


module H i (T, F(j)) is finitely generated, and so has finite length: we denote
this length by hi (T, F(j)).

We show next how the Serre–Grothendieck Correspondence enables us to


produce quickly, for a coherent sheaf F of OT -modules as in 20.4.15, a Hilbert
polynomial of F and, for each i ∈ N0 , an i-th cohomological Hilbert polyno-
mial of F which play similar rôles to the corresponding polynomials in 17.1.8
and 17.1.11.

20.4.16 Theorem and Definitions. (The hypotheses of 20.4.1 are in force.)


Assume that R0 is Artinian and let F be a non-zero coherent sheaf of OT -
modules. Set d := dim F.
Now hi (T, F) = 0 for all i  0, and the Euler characteristic χ(F) of F is

defined by χ(F) := i∈N0 (−1)i hi (T, F) (see [30, Chapter III, Exercise 5.1,
p. 230]).
There is a (necessarily uniquely determined) polynomial PF ∈ Q[X] of
degree d = dim F such that PF (n) = χ(F(n)) for all n ∈ Z. This polynomial
PF is called the Hilbert polynomial of F.
With the notation of 17.1.1(v), for each j = 0, . . . , d, we set ej (F) :=
ej (PF ) and refer to this as the j-th Hilbert coefficient of F. Thus
d  
i X +d−i
PF (X) = (−1) ei (F) .
i=0
d−i
 
Moreover, if N is a finitely generated graded R-module with F = N 
0
(see 20.4.6), then we have dim N = d + 1, PF = PN , χF = χN and ei (F) =
ei (N ) for all i ∈ {0, . . . , d}.

Proof. By 20.4.6, there exists a finitely generated graded R-module N such


that F = (N  )0 ; by 20.4.14, dim F = dim N − 1, so that dim N = d + 1.
Let n ∈ Z and i ∈ N. We can now use 20.4.5(ii) and 20.3.15 to see that
there are R0 -isomorphisms Γ(T, F(n)) ∼ = DR+ (N )n and H i (T, F(n)) ∼ =
i ∼
R DR+ (N )n = HR+ (N )n .
i+1

Hence hi (T, F(n)) = 0 for all i > d, and, with the notation of 17.1.4, we
have χ(F(n)) = χN (n). The result therefore follows from 17.1.7.

20.4.17 Definitions and Exercise. (The hypotheses of 20.4.1 are in force.)


Assume that R0 is Artinian and let F be a non-zero coherent sheaf of OT -
modules. Let i ∈ N0 . The function hiF : Z −→ N0 defined by hiF (n) =
hi (T, F(n)) for all n ∈ Z is referred to as the i-th cohomological Hilbert
function of F.
20.4 Applications to projective schemes 473

Show that there is a polynomial piF ∈ Q[X] of degree at most i such that
hiF (n) = hi (T, F(n)) = piF (n) for all n  0. The (uniquely determined)
polynomial piF is called the i-th cohomological Hilbert polynomial of F.
Let d := dim F. Show that the leading term of pdF is

(−1)d e0 (F) d
X ,
d!
where e0 (F) is as defined in 20.4.16.

We come now to a basic bounding result, due to Mumford, for the regularity
of coherent sheaves of ideals.

20.4.18 Theorem: Mumford’s Regularity Bound [54, Theorem, p. 101].


Suppose that R = R0 [X0 , X1 , . . . , Xd ] is a polynomial ring in d + 1 indeter-
minates X0 , X1 , . . . , Xd over an Artinian local ring R0 , where d ∈ N. Regard

R = n∈N0 Rn as N0 -graded with deg Xi = 1 for all i = 0, . . . , d. As in
20.4.1, we set T = Proj(R). (Thus T is just PdR0 , projective d-space over R0 :
see [30, Chapter II, Example 2.5.1, p. 77].)
Then there is a function F : Zd+1 −→ Z such that, for each non-zero
coherent sheaf I of ideals on T , we have reg(I) ≤ F (e0 (I), . . . , ed (I)).

Proof. By 20.4.7,the  coherent sheaves of ideals on T are exactly the sheaves



of the form I = b 0 , where b is a graded ideal of R. If a = 0 is such a
graded   corresponding to the coherent sheaf of ideals J , then dim a =
 ideal,
dim  a 0 +1 = dim J +1 = d+1  ei (a) = ei (J ) for all i ∈ {0, . . . , d}
and
(see 20.4.16). Also, reg(J ) = reg  a 0 = reg2 (a) by 20.4.11(iii). The re-
sult therefore follows from 17.3.6.

So far in this chapter, the results which we have obtained about sheaf coho-
mology have not been obviously related to local properties of the underlying
sheaves. In our next sequence of results and exercises, we aim for the so-called
Severi–Enriques–Zariski–Serre Vanishing Theorem (see F. Severi [78], F. En-
riques [13], O. Zariski [88], and J.-P. Serre [77, §76, Théorème 4]), which
establishes a fundamental link between the local structure of a coherent sheaf
of OT -modules (conveyed by information about the depths of its stalks, for
example) and global properties of the sheaf (described by the vanishing of cer-
tain cohomology groups, for example). Our approach to this Vanishing Theo-
rem again makes use of the Serre–Grothendieck Correspondence, this time in
conjunction with the Graded Finiteness Theorem 14.3.10.

20.4.19 Exercise. (The hypotheses of 20.4.1 apply.) Assume that the graded
R-module M is finitely generated, and set F := (M 6)0 . Fix p = x ∈ T , and
474 Links with sheaf cohomology

let mT,x denote the maximal ideal (pR(p) )0 of the local ring OT,x , which we
identify with (R(p) )0 by means of the isomorphism of 20.4.2(i).

(i) Use 19.7.1(vii) to show that, if x is a closed point of T (so that p is a


maximal member of Proj(R) with respect to inclusion), then p ∩ R0 is
a maximal ideal of R0 and dim R/p = 1.
(ii) Since R is homogeneous, there exists t ∈ R1 \ p. Denote the ring of
fractions OT,x [X]X of the polynomial ring OT,x [X] by OT,x [X, X −1 ].
Show that there is a unique homogeneous isomorphism of OT,x -algebras

φt : OT,x [X, X −1 ] −→ R(p) for which φt (X) = t.
=

(iii) Identify the stalk Fx with (M(p) )0 by means of the isomorphism of


20.4.3. Show that, when M(p) is considered as an OT,x [X, X −1 ]-module
by means of the isomorphism φt of part (ii), there is a homogeneous iso-
morphism of graded OT,x [X, X −1 ]-modules

ψt,M : Fx ⊗OT ,x OT,x [X, X −1 ] −→ M(p)
=

for which

ψt,M (z ⊗ f ) = φt (f )z for all z ∈ (M(p) )0 and f ∈ OT,x [X, X −1 ].

(iv) Show that the isomorphism φt of part (ii) leads to a ring isomorphism

=
OT,x [X]mT ,x OT ,x [X] −→ Rp . Conclude that dim OT,x = ht p, that
depth OT,x = depth Rp , and that OT,x is a domain, respectively nor-
mal, Cohen–Macaulay, Gorenstein, regular, if and only if Rp has the
same property.
(v) Show that the isomorphism ψt,M of part (iii) gives rise to an isomor-

=
phism Fx ⊗OT ,x OT,x [X]mT ,x OT ,x [X] −→ Mp of OT,x [X]mT ,x OT ,x [X] -
modules. Conclude that depthOT ,x Fx = depthRp Mp and that Fx is
free of rank r over OT,x if and only if Mp is free of rank r over Rp .
(vi) Let q ∈ Spec(R) be such that q* ⊆ p. Show that Rq is a domain,
respectively normal, Cohen–Macaulay, Gorenstein, regular, if OT,x has
the same property.

20.4.20 Exercise. (The hypotheses of 20.4.1 are in force.) Let π : T =


Proj(R) −→ Spec(R0 ) be the natural map, defined by π(q) = q ∩ R0 for all
q ∈ T . Fix p = x ∈ T . In the graded ring R/p, let Σ(p) denote the set of non-
zero homogeneous elements of degree 0. Note that Σ(p)−1 (R/p) is positively
graded and homogeneous: set Tx := Proj(Σ(p)−1 (R/p)).

(i) Show that there is a homeomorphism Tx −→ π −1 ({π(x)}) ∩ {x},
where the ‘overline’ is used to indicate closure in T .
20.4 Applications to projective schemes 475

(ii) Show that dim Tx = ht(R+ + p)/p − 1.


20.4.21 Notation and Remark. (The hypotheses of 20.4.1 apply.) Let F be
a coherent sheaf of OT -modules. We set
 
δ(F) := inf depthOT ,x Fx + dim Tx : x ∈ T ,

where Tx , for x ∈ T , is as defined in 20.4.20.


By 20.4.6, there exists a finitely generated graded R-module N such that
F∼= (N )0 ; by 20.4.19(v), for x = p ∈ T , we have

depthOT ,x Fx = depthRp Np .
Therefore, by 20.4.20, with the notation of 9.2.2,
δ(F) = inf {depth Np + ht(R+ + p)/p − 1 : p ∈ * Spec(R) \ Var(R+ )}
 
= inf adjR+ depth Np − 1 : p ∈ * Spec(R) \ Var(R+ ) .

In the case when R is a homomorphic image of a regular (commutative Noethe-


rian) ring, we can deduce from the Graded Finiteness Theorem 14.3.10 that
δ(F) = fR+ (N ) − 1. This observation is the key to our proof of the Severi–
Enriques–Zariski–Serre Vanishing Theorem 20.4.23 below.
20.4.22 Exercise. (The hypotheses of 20.4.1 apply.) Let π : T → Spec(R0 )
be the natural map. Assume that R0 is Artinian, and let F be a coherent sheaf
of OT -modules.
(i) Show that π −1 ({π(x)}) ∩ {x} = {x} for all x ∈ T .
(ii) Use 9.3.5 to show that
 
δ(F) = inf depthOT ,x Fx : x is a closed point of T .

20.4.23 The Severi–Enriques–Zariski–Serre Vanishing Theorem. (See


Severi [78], Enriques [13], Zariski [88], and Serre [77, §76, Théorème 4]).
(The hypotheses of 20.4.1 apply.) Assume that R0 is a homomorphic image of
a regular (commutative Noetherian) ring. Let F be a coherent sheaf of OT -
modules, and let r ∈ N0 . Then δ(F) > r if and only if H i (T, F(j)) = 0 for
all i ≤ r and all j  0.
Proof. By 20.4.6, there is a finitely generated graded R-module N such that
F∼= (N )0 . By 20.4.4, there are homogeneous R-isomorphisms
  ∼
=
j∈Z H (T, (N (j))0 ) −→ HR+ (N ) for all i ∈ N
i i+1

  ∼
=
and j∈Z Γ(T, (N (j))0 ) −→ DR+ (N ). Also, it follows from 2.2.6(i)(c) that
1
DR+ (N ) is finitely generated if and only if HR +
(N ) is.
476 Links with sheaf cohomology

In view of [7, Proposition 1.5.4], the hypothesis on R0 ensures that R is a


homomorphic image of a regular (commutative Noetherian) ring. It therefore
follows from 20.4.21 that δ(F) = fR+ (N ) − 1, so that δ(F) > r if and only if
 
the graded R-module j∈ZH i (T, (N (j))0 ) is finitely generated for all i ≤ r.
The desired conclusions follow from these observations and Serre’s Finiteness
Theorem 20.4.8.
20.4.24 Exercise. (The hypotheses of 20.4.1 apply.) Assume that R0 is Ar-
tinian, and let F be a coherent sheaf of OT -modules. Recall that Ass F :=
{x ∈ T : depthOT ,x Fx = 0}. Show that the graded R-module Γ∗ (T, F) =

j∈Z Γ(T, F(j)) is finitely generated if and only if Ass F contains no closed
point of T .

20.5 Locally free sheaves


In this final section, we give some applications to locally free sheaves. These
correspond to vector bundles in algebraic geometry, and form an important
class of sheaves.
20.5.1 Hypotheses for the section. Throughout this section, the hypotheses
of 20.4.1 will be in force.
20.5.2 Definition. A coherent sheaf F of OT -modules is called locally free
if the stalk Fx is a free OT,x -module (of finite rank, as F is coherent) for all
x ∈ T . (See [30, Chapter II, §5, p. 109, and Exercise 5.7, p. 124].)
20.5.3 Exercise. Suppose that T is connected and OT,x is an integral do-
main for all x ∈ T . Show that T has a unique minimal member (with respect
to inclusion of prime ideals) and conclude that T is irreducible.
20.5.4 Reminder. Recall from [30, Chapter II, Exercise 3.8, p. 91] that T is
said to be normal if and only if OT,x is a normal integral domain for all x ∈ T .
Similarly, T is said to be regular if and only if the local ring OT,x is regular
for all x ∈ T . (See [30, Chapter II, Remark 6.11.1A, p. 142].)
Observe that, if T is regular, then it is normal. Note also that, by 20.4.19(iv),
Proj(K[X0 , . . . , Xr ]) is regular (where K is a field, r ∈ N, X0 , . . . , Xr are
indeterminates, and the polynomial ring is graded so that deg Xi = 1 for all
i = 0, . . . , r and K is the component of degree 0).
20.5.5 Exercise. Assume that R0 is a field and that T = Proj(R) is con-
nected. Set d := dim T .
Suppose that T is normal (see 20.5.4), and assume that d ≥ 2. Let F be a
20.5 Locally free sheaves 477

locally free coherent sheaf of OT -modules. Show that H 1 (T, F(j)) = 0 for all
j  0. (Compare this with Hartshorne’s presentation of the Enriques–Severi–
Zariski Lemma in [30, Chapter III, Corollary 7.8, p. 244].)

20.5.6 Theorem: Serre’s Cohomological Criterion for Local Freeness


[77, §75, Théorème 3]. Assume that R0 is a field, and that T = Proj(R) is
connected and regular (see 20.5.4). Let d := dim T . Let F be a coherent sheaf
of OT -modules. Then F is locally free if and only if H i (T, F(n)) = 0 for all
i < d and all n  0.

Proof. By 20.4.23, it is enough for us to show that F is locally free if and only
if δ(F) ≥ d. (See 20.4.21 for the definition of δ(F).) First of all observe that
T has a unique minimal member q (by 20.5.3). Since R0 is a field, Var(R+ ) =
{R+ }. Since q ∈ T , it follows that q is the unique minimal prime of R. Let
x = p ∈ T . Then we have q ⊆ p ⊆ p + R+ ∈ * Spec(R). So, on use of
20.4.19(iv) and 20.4.20, we obtain

dim OT,x + dim Tx = ht p + ht(R+ + p)/p − 1 = ht(R+ + p) − 1


= ht R+ − 1 = dim R − 1 = d

because R is a finitely generated R0 -algebra (and therefore catenary) and has


a unique minimal prime q. Thus dim OT,x + dim Tx = d for all x ∈ T .
Now assume that F is locally free. Then, for each x ∈ T ,

depthOT ,x Fx = depth OT,x = dim OT,x ,

as OT,x is regular, and therefore Cohen–Macaulay. It follows that

depthOT ,x Fx + dim Tx = dim OT,x + dim Tx = d for all x ∈ T,

so that δ(F) = d.
Conversely, suppose that δ(F) ≥ d. Then, for each x ∈ T , we have

depthOT ,x Fx + dim Tx ≥ d = dim OT,x + dim Tx ,

so that depthOT ,x Fx ≥ dim OT,x . As OT,x is regular, it follows from the


Auslander–Buchsbaum Theorem (see [50, Theorem 19.1], for example) that
the OT,x -module Fx is projective, and therefore free (by [50, Theorem 2.5]).

Finally, we consider locally free sheaves over projective d-space PdK =


Proj(K[X0 , . . . , Xd ]) over K, where K is an algebraically closed field.

20.5.7 Reminder and Exercise. (The hypotheses of 20.4.1 apply.)


478 Links with sheaf cohomology

(i) In the category S(OT ) of sheaves of OT -modules one may form direct
sums, and the functors ( • )0 (from *C(R) to S(OT )) and H i (T, • ) (i ∈
N0 ) (from S(OT ) to *C(R)) are additive. So, if r ∈ N, N1 , . . . , Nr are
graded R-modules and F1 , . . . , Fr are sheaves of OT -modules, we have
  r  6 
Nj ∼
r
j=1 =0
Nj and
j=1 0
r ∼ r H i (T, Fj )
j=1 Fj ) = for all i ∈ N0 .
i
H (T, j=1

(ii) Assume now that d ∈ N, that K is a field and R = K[X0 , . . . , Xd ]


is a polynomial ring, so that T = Proj(R) becomes projective d-space
PdK over K. A sheaf F of OT -modules is said to split (completely) (the
r
word ‘completely’ is often omitted) if and only if F ∼ = j=1 OT (aj )
for some r ∈ N and some integers a1 , . . . , ar . Show that, if this is the
case, then F is coherent and locally free, and if also the ai are numbered
so that a1 ≥ a2 ≥ · · · ≥ ar , then a := (a1 , . . . , ar ) ∈ Zr is uniquely
determined by F. In this situation, a = (a1 , . . . , ar ) is called the split-
ting type of F. (You might find it helpful to consider the cohomological
Hilbert functions n → h0 (T, F(n)) and n → h0 (T, OT (aj )(n)), in
conjunction with the Serre–Grothendieck Correspondence 20.4.4).
20.5.8 Theorem: Horrocks’ Splitting Criterion [40]. Let d ∈ N, let K be
a field, let T = PdK and let F be a non-zero coherent sheaf of OT -modules.
Then the following statements are equivalent:
(i) F splits;
(ii) H 0 (T, F(n)) = 0 for all n  0 and H i (T, F(n)) = 0 for all i ∈
{1, . . . , d − 1} and all n ∈ Z.

Proof. Here, T = Proj(R) where R is the polynomial ring K[X0 , . . . , Xd ]


with deg Xi = 1 for all i = 0, . . . , d.
r
(i) ⇒ (ii) Assume that F splits, so that F ∼ = j=1 OT (aj ) for some r ∈ N
and some integers a1 , . . . , ar . On use of 20.5.7(i), we obtain
r r  
H i (T, F(n)) ∼
= j=1 H i (T, OT (aj + n)) ∼ = j=1 H i T, R 0 (aj + n)
r  

= j=1 H i T, R(a j + n)0 for all i ∈ N0 and all n ∈ Z.
By the Serre–Grothendieck Correspondence 20.4.4,
 

H 0 T, R(a ∼
j + n)0 = DR+ (R)aj +n ,

and DR+ (R)aj +n ∼


= Raj +n as HR0
+
1
(R) = HR +
(R) = 0. Therefore
 

H 0 T, R(a j + n)0 = 0 for all n < −aj ,

and so H 0 (T, F(n)) = 0 for all n  0. Also, by Theorem 20.4.4 once again,
20.5 Locally free sheaves 479
 
for all i > 0 and all n ∈ Z, we have H i T, R(a  ∼ i+1
j + n)0 = HR+ (R)aj +n ,
and this vanishes for all i ∈ {1, . . . , d−1} as grade R+ = d+1. It follows that
H i (T, F(n)) = 0 for all i ∈ {1, . . . , d − 1} and all n ∈ Z. Hence statement
(ii) is true.
 ⇒ (i) Assume that statement (ii) is true. By 20.4.6, we can write F =
 (ii)
M6 for a finitely generated graded R-module M . Then it follows from the
0
Serre–Grothendieck Correspondence 20.4.4 that HR i
+
(M ) = 0 for all i ∈
{2, . . . , d} and DR+ (M )n = 0 for all n  0. Consider the exact sequence
ηM
0 −→ ΓR+ (M ) −→ M −→ DR+ (M ) −→ HR
1
+
(M ) −→ 0
of 13.5.4(i), in which all the homomorphisms are homogeneous. As the K-
vector space HR 1
+
(M )n has finite dimension for all n ∈ Z and vanishes for all
large n (by 16.1.5), it follows that the graded R-module DR+ (M ) is finitely
generated. Moreover, if we apply the exact functor ( • )0 to the above exact
1
sequence and observe that both ΓR+ (M ) and HR +
(M ) are R+ -torsion (so
    
that ΓR 1 (M ) = 0 (by 20.3.2(i))), we get an isomorphism
+ (M ) 0 = HR + 0
   
of sheaves F = M 6 ∼ 
DR+ (M ) 0 . Thus DR+ (M ) is finitely generated
0 =
and F ∼= (D i
R+ (M ))0 . Now by 2.2.10(iv),(v), we have HR+ (DR+ (M )) = 0
i ∼
for i = 0, 1 and H (DR (M )) = H (M ) for all i > 1, As H i (M ) = 0
i
R+ + R+ R+
for all i ∈ {2, . . . , d}, we get that HR
i
+
(DR+ (M )) = 0 for all i ∈ {0, . . . , d}.
 
Therefore grade R+ ≥ d + 1. As 0 = F = ∼ D R (M ) , we have
DR+ (M ) + 0
DR+ (M ) = 0, so that gradeDR (M ) R+ = d + 1. So, by Hilbert’s Syzygy
+
Theorem (see [7, Corollary 2.2.15]), the finitely generated graded R-module
r
DR+ (M ) is free, and so DR+ (M ) ∼ = j=1 R(aj ) (in *C(R)) for some r ∈ N
and some a1 , . . . , ar ∈ Z. Therefore, in view of 20.4.5(ii),
r r    ∼  r   
F∼ = j=1 R(aj ) ∼ = R(aj ) =
j=1 0
R (aj )
j=1 0
0
r
= j=1 OT (aj ).
Therefore F splits.

20.5.9 Corollary: Grothendieck’s Splitting Theorem [23, Théorème 2.1].


Let K be a field and let T = P1K be the projective line over K. Then each
non-zero coherent locally free sheaf F of OT -modules splits.
Proof. Here T = Proj(R) where R is the polynomial ring K[X0 , X1 ] with
deg X0 = deg X1 = 1.
As T is regular (see 20.5.4) of dimension 1 and connected, Serre’s Criterion
for Local Freeness 20.5.6 implies that H 0 (T, F(n)) = 0 for all n  0. We
can therefore apply 20.5.8 to complete the proof.
References

[1] Y. AOYAMA, ‘On the depth and the projective dimension of the canonical mod-
ule’, Japanese J. Math. 6 (1980) 61–66.
[2] Y. AOYAMA, ‘Some basic results on canonical modules’, J. Math. Kyoto Univ. 23
(1983) 85–94.
[3] H. BASS, ‘On the ubiquity of Gorenstein rings’, Math. Zeit. 82 (1963) 8–28.
[4] M. B RODMANN and C. H UNEKE, ‘A quick proof of the Hartshorne–Lichtenbaum
Vanishing Theorem’, Algebraic geometry and its applications (Springer, New
York, 1994), pp. 305–308.
[5] M. B RODMANN and J. RUNG, ‘Local cohomology and the connectedness dimen-
sion in algebraic varieties’, Comment. Math. Helvetici 61 (1986) 481–490.
[6] M. B RODMANN and R. Y. S HARP, ‘Supporting degrees of multi-graded local
cohomology modules’, J. Algebra 321 (2009) 450–482.
[7] W. B RUNS and J. H ERZOG, Cohen–Macaulay rings, Cambridge Studies in
Advanced Mathematics 39, Revised Edition (Cambridge University Press,
Cambridge, 1998).
[8] F. W. C ALL and R. Y. S HARP, ‘A short proof of the local Lichtenbaum–
Hartshorne Theorem on the vanishing of local cohomology’, Bull. London Math.
Soc. 18 (1986) 261–264.
[9] C. D’C RUZ, V. KODIYALAM and J. K . V ERMA, ‘Bounds on the a-invariant and
reduction numbers of ideals’, J. Algebra 274 (2004) 594–601.
[10] D. E ISENBUD, Commutative algebra with a view toward algebraic geometry,
Graduate Texts in Mathematics 150 (Springer, New York, 1994).
[11] D. E ISENBUD and S. G OTO, ‘Linear free resolutions and minimal multiplicity’,
J. Algebra 88 (1984) 89–133.
[12] J. E LIAS, ‘Depth of higher associated graded rings’, J. London Math. Soc. (2) 70
(2004) 41–58.
[13] F. E NRIQUES, Le superficie algebriche (Zanichelli, Bologna, 1949).
[14] G. FALTINGS, ‘Über die Annulatoren lokaler Kohomologiegruppen’, Archiv der
Math. 30 (1978) 473–476.
[15] G. FALTINGS, ‘Algebraisation of some formal vector bundles’, Annals of Math.
110 (1979) 501–514.
[16] G. FALTINGS, ‘Der Endlichkeitssatz in der lokalen Kohomologie’, Math. Annalen
255 (1981) 45–56.
References 481

[17] R. F EDDER and K. WATANABE, ‘A characterization of F -regularity in terms of


F -purity’, Commutative algebra: proceedings of a microprogram held June 15 –
July 2, 1987, Mathematical Sciences Research Institute Publications 15 (Springer,
New York, 1989), pp. 227–245.
[18] D. F ERRAND and M. R AYNAUD, ‘Fibres formelles d’un anneau local
Noethérien’, Ann. Sci. École Norm. Sup. 3 (1970) 295–311.
[19] H.-B. F OXBY, ‘Gorenstein modules and related modules’, Math. Scand. 31
(1972) 267–284.
[20] W. F ULTON and J. H ANSEN, ‘A connectedness theorem for projective varieties,
with applications to intersections and singularities of mappings’, Annals of Math.
110 (1979) 159–166.
[21] P. G ABRIEL, ‘Des catégories abéliennes’, Bull. Soc. Math. France 90 (1962) 323–
448.
[22] S. G OTO and K. WATANABE, ‘On graded rings, II (Zn -graded rings)’, Tokyo J.
Math. 1 (1978) 237–261.
[23] A. G ROTHENDIECK, ‘Sur la classification des fibrés holomorphes sur la sphère
de Riemann’, American J. Math. 79 (1957) 121–138.
[24] A. G ROTHENDIECK, ‘Éléments de géométrie algébrique IV: étude locale des
schémas et des morphismes de schémas’, Institut des Hautes Études Scientifiques
Publications Mathématiques 24 (1965) 5–231.
[25] A. G ROTHENDIECK, Local cohomology, Lecture Notes in Mathematics 41
(Springer, Berlin, 1967).
[26] A. G ROTHENDIECK, Cohomologie locale des faisceaux cohérents et théorèmes
de Lefschetz locaux et globaux (SGA 2), Séminaire de Géométrie Algébrique du
Bois-Marie 1962 (North-Holland, Amsterdam, 1968).
[27] J. H ARRIS, Algebraic geometry: a first course, Graduate Texts in Mathematics
133 (Springer, New York, 1992).
[28] R. H ARTSHORNE, ‘Complete intersections and connectedness’, American J.
Math. 84 (1962) 497–508.
[29] R. H ARTSHORNE, ‘Cohomological dimension of algebraic varieties’, Annals of
Math. 88 (1968) 403–450.
[30] R. H ARTSHORNE, Algebraic geometry, Graduate Texts in Mathematics 52
(Springer, New York, 1977).
[31] M. H ERRMANN, E. H YRY and T. KORB, ‘On a-invariant formulas’, J. Algebra
227 (2000) 254–267.
[32] M. H ERRMANN, S. I KEDA and U. O RBANZ, Equimultiplicity and blowing up
(Springer, Berlin, 1988).
[33] J. H ERZOG and E. K UNZ, Der kanonische Modul eines Cohen–Macaulay-Rings,
Lecture Notes in Mathematics 238 (Springer, Berlin, 1971).
[34] J. H ERZOG and E. K UNZ, Die Wertehalbgruppe eines lokalen Rings der Di-
mension 1, Sitzungsberichte der Heidelberger Akademie der Wissenschaften
Mathematisch-naturwissenschaftliche Klasse, Jahrgang 1971 (Springer, Berlin,
1971).
[35] L. T. H OA, ‘Reduction numbers and Rees algebras of powers of an ideal’, Proc.
American Math. Soc. 119 (1993) 415–422.
[36] L. T. H OA and C. M IYAZAKI, ‘Bounds on Castelnuovo–Mumford regularity for
generalized Cohen–Macaulay graded rings’, Math. Annalen 301 (1995) 587–598.
482 References

[37] M. H OCHSTER, ‘Contracted ideals from integral extensions of regular rings’,


Nagoya Math. J. 51 (1973) 25–43.
[38] M. H OCHSTER and C. H UNEKE, ‘Tight closure, invariant theory and the
Briançon–Skoda Theorem’, J. American Math. Soc. 3 (1990) 31–116.
[39] M. H OCHSTER and C. H UNEKE, ‘Indecomposable canonical modules and con-
nectedness’, Commutative algebra: syzygies, multiplicities, and birational alge-
bra (South Hadley, MA, 1992), Contemporary Mathematics 159 (American Math-
ematical Society, Providence, RI, 1994), pp. 197–208.
[40] G. H ORROCKS, ‘Vector bundles on the punctured spectrum of a local ring’, Proc.
London Math. Soc. (3) 14 (1964) 689–713.
[41] C. H UNEKE, Tight closure and its applications, Conference Board of the Math-
ematical Sciences Regional Conference Series in Mathematics 88 (American
Mathematical Society, Providence, RI, 1996).
[42] C. H UNEKE, ‘Tight closure, parameter ideals and geometry’, Six lectures on com-
mutative algebra (Bellaterra, 1996), Progress in Mathematics 166 (Birkhäuser,
Basel, 1998), pp. 187–239.
[43] C. L. H UNEKE and R. Y. S HARP, ‘Bass numbers of local cohomology modules’,
Transactions American Math. Soc. 339 (1993) 765–779.
[44] D. K IRBY, ‘Coprimary decomposition of Artinian modules’, J. London Math.
Soc. (2) 6 (1973) 571–576.
[45] I. G. M ACDONALD, ‘Secondary representation of modules over a commutative
ring’, Symposia Matematica 11 (Istituto Nazionale di alta Matematica, Roma,
1973) 23–43.
[46] I. G. M ACDONALD, ‘A note on local cohomology’, J. London Math. Soc. (2) 10
(1975) 263–264.
[47] I. G. M ACDONALD and R. Y. S HARP, ‘An elementary proof of the non-vanishing
of certain local cohomology modules’, Quart. J. Math. Oxford (2) 23 (1972) 197–
204.
[48] T. M ARLEY, ‘The reduction number of an ideal and the local cohomology of the
associated graded ring’, Proc. American Math. Soc. 117 (1993) 335–341.
[49] E. M ATLIS, ‘Injective modules over Noetherian rings’, Pacific J. Math. 8 (1958)
511–528.
[50] H. M ATSUMURA, Commutative ring theory, Cambridge Studies in Advanced
Mathematics 8 (Cambridge University Press, Cambridge, 1986).
[51] L. M ELKERSSON, ‘On asymptotic stability for sets of prime ideals connected
with the powers of an ideal’, Math. Proc. Cambridge Philos. Soc. 107 (1990)
267–271.
[52] L. M ELKERSSON, ‘Some applications of a criterion for artinianness of a module’,
J. Pure and Applied Algebra 101 (1995) 291–303.
[53] E. M ILLER and B. S TURMFELS, Combinatorial commutative algebra, Graduate
Texts in Mathematics 227 (Springer, New York, 2005).
[54] D. M UMFORD, Lectures on curves on an algebraic surface, Annals of Mathemat-
ics Studies 59 (Princeton University Press, Princeton, NJ, 1966).
[55] M. P. M URTHY, ‘A note on factorial rings’, Arch. Math. (Basel) 15 (1964) 418–
420.
[56] M. NAGATA, Local rings (Interscience, New York, 1962).
References 483

[57] U. NAGEL, ‘On Castelnuovo’s regularity and Hilbert functions’, Compositio


Math. 76 (1990) 265–275.
[58] U. NAGEL and P. S CHENZEL, ‘Cohomological annihilators and Castelnuovo–
Mumford regularity’, Commutative algebra: syzygies, multiplicities, and bira-
tional algebra (South Hadley, MA, 1992), Contemporary Mathematics 159 (Amer-
ican Mathematical Society, Providence, RI, 1994), pp. 307–328.
[59] D. G. N ORTHCOTT, Ideal theory, Cambridge Tracts in Mathematics and Mathe-
matical Physics 42 (Cambridge University Press, Cambridge, 1953).
[60] D. G. N ORTHCOTT, An introduction to homological algebra (Cambridge Univer-
sity Press, Cambridge, 1960).
[61] D. G. N ORTHCOTT, Lessons on rings, modules and multiplicities (Cambridge
University Press, Cambridge, 1968).
[62] D. G. N ORTHCOTT, ‘Generalized Koszul complexes and Artinian modules’,
Quart. J. Math. Oxford (2) 23 (1972) 289–297.
[63] D. G. N ORTHCOTT and D. R EES, ‘Reductions of ideals in local rings’, Proc.
Cambridge Philos. Soc. 50 (1954) 145–158.
[64] L. O’C ARROLL, ‘On the generalized fractions of Sharp and Zakeri’, J. London
Math. Soc. (2) 28 (1983) 417–427.
[65] A. O OISHI, ‘Castelnuovo’s regularity of graded rings and modules’, Hiroshima
Math. J. 12 (1982) 627–644.
[66] C. P ESKINE and L. S ZPIRO, ‘Dimension projective finie et cohomologie locale’,
Institut des Hautes Études Scientifiques Publications Mathématiques 42 (1973)
323–395.
[67] L. J. R ATLIFF , J R ., ‘Characterizations of catenary rings’, American J. Math. 93
(1971) 1070–1108.
[68] D. R EES, ‘The grade of an ideal or module’, Proc. Cambridge Philos. Soc. 53
(1957) 28–42.
[69] I. R EITEN, ‘The converse to a theorem of Sharp on Gorenstein modules’, Proc.
American Math. Soc. 32 (1972) 417–420.
[70] P. ROBERTS, Homological invariants of modules over commutative rings,
Séminaire de Mathématiques Supérieures (Les Presses de l’Université de
Montréal, Montréal, 1980).
[71] J. J. ROTMAN, An introduction to homological algebra (Academic Press,
Orlando, FL, 1979).
[72] P. S CHENZEL, ‘Einige Anwendungen der lokalen Dualität und verallgemeinerte
Cohen–Macaulay-Moduln’, Math. Nachr. 69 (1975) 227–242.
[73] P. S CHENZEL, ‘Flatness and ideal-transforms of finite type’, Commutative alge-
bra, Proceedings, Salvador 1988, Lecture Notes in Mathematics 1430 (Springer,
Berlin, 1990), pp. 88–97.
[74] P. S CHENZEL, ‘On the use of local cohomology in algebra and geometry’, Six
lectures on commutative algebra (Bellaterra, 1996), Progress in Mathematics 166
(Birkhäuser, Basel, 1998), pp. 241–292.
[75] P. S CHENZEL, ‘On birational Macaulayfications and Cohen–Macaulay canonical
modules’, J. Algebra 275 (2004) 751–770.
[76] P. S CHENZEL , N. V. T RUNG and N. T. C UONG, ‘Verallgemeinerte Cohen–
Macaulay-Moduln’, Math. Nachr. 85 (1978) 57–73.
484 References

[77] J.-P. S ERRE, ‘Faisceaux algébriques cohérents’, Annals of Math. 61 (1955) 197–
278.
[78] F. S EVERI, Serie, sistemi d’equivalenza e corrispondenze algebriche sulle varietà
algebriche (a cura di F. Conforto e di E. Martinelli, Roma, 1942).
[79] R. Y. S HARP, ‘Finitely generated modules of finite injective dimension over cer-
tain Cohen–Macaulay rings’, Proc. London Math. Soc. (3) 25 (1972) 303–328.
[80] R. Y. S HARP, ‘On the attached prime ideals of certain Artinian local cohomology
modules’, Proc. Edinburgh Math. Soc. (2) 24 (1981) 9–14.
[81] R. Y. S HARP, Steps in commutative algebra: Second edition, London Mathemat-
ical Society Student Texts 51 (Cambridge University Press, Cambridge, 2000).
[82] R. Y. S HARP and M. T OUSI, ‘A characterization of generalized Hughes com-
plexes’, Math. Proc. Cambridge Philos. Soc. 120 (1996) 71–85.
[83] J. R. S TROOKER, Homological questions in local algebra, London Mathematical
Society Lecture Notes 145 (Cambridge University Press, Cambridge, 1990).
[84] J. S T ÜCKRAD and W. VOGEL, Buchsbaum rings and applications (Springer,
Berlin, 1986).
[85] K. S UOMINEN, ‘Localization of sheaves and Cousin complexes’, Acta Mathemat-
ica 131 (1973) 27–41.
[86] N. V. T RUNG, ‘Reduction exponent and degree bound for the defining equations
of graded rings’, Proc. American Math. Soc. 101 (1987) 229–236.
[87] N. V. T RUNG, ‘The largest non-vanishing degree of graded local cohomology
modules’, J. Algebra 215 (1999) 481–499.
[88] O. Z ARISKI, ‘Complete linear systems on normal varieties and a generalization
of a lemma of Enriques–Severi’, Annals of Math. 55 (1952) 552–592.
[89] O. Z ARISKI and P. S AMUEL, Commutative algebra, Vol. II, Graduate Texts in
Mathematics 29 (Springer, Berlin, 1975).
Index

a-adjusted depth, see depth local ring, 108, 200, 201, 203, 241, 357,
b-minimum (λba (M )), see depth 369, 380, 387, 473
Abelian category, 255 module, 135–139, 141, 142, 144, 146, 158,
acyclic module, 65 159, 200–206, 208–210, 220
Γa -, 65–69, 71, 76, 260 ring, 355, 367–371, 373, 377–379, 471,
ΓB -, 66, 260 472, 475
affine associated graded ring (G(b)), 394, 437
algebraic cone, see cone a-torsion
algebraic set, 40 functor (Γa ), see functor
n-space (An (K)), 39 module, see module
complex (An ), 39 submodule (Γa (M )), 3
open set, 419, 452 a-torsion-free module, see module
scheme, 458, 459 a-transform, see transform
surface, 162 functor (Da ), see functor
variety, 39, 93, 122–124, 126, 127, 410, attached prime ideal, 140
423, 441, 443, 451, 452 attached primes (Att), 140–142, 144, 145,
a-invariant, 326 158, 159, 206, 210, 220–222
analytic spread (spr(b)), 394–398, 402
Baer Criterion, 18, 260, 261
analytically independent, 396
basic ideal, 396, 397
in b, 392, 393, 395, 396
Bass number (μi (p, M )), 213, 216, 236, 299,
analytically irreducible
301, 306–308, 316, 325
closed set, at p, 420
B-closure, see closure
local ring, see local ring
beginning of a graded module (beg(M )), see
analytically reducible
graded module
closed set, at p, 420
Bertini’s Connectivity Theorem, 426
Annihilator Theorem
Bertini–Grothendieck Connectivity Theorem,
of Faltings, see Faltings’
425, 432
Graded, see Faltings’
b-finiteness dimension relative to a (fab (M )),
arithmetic
see dimension
depth, 339, 341
bounding system, 378
properties, of a projective variety, 338
minimal, 378
arithmetic rank
B-torsion module, see module
of a closed subset, with respect to a larger
one (araW (Z)), 425 B-torsion-free module, see module
local, at p (araW,p (Z)), 420 B-transform, see transform
of an ideal (ara(a)), 56, 78, 398, 412–415 functor (DB ), see functor
arithmetically Buchsbaum
Cohen–Macaulay, 339, 341, 342 module, 187, 188
Gorenstein, 339, 341 ring, 187, 188
Artinian canonical module (ωR ), 224, 226–231, 233,
486 Index

234, 236–244, 248–250, 314–317, cohomological Hilbert


320–322, 328 function, see Hilbert
*canonical module, 313–330, 343 polynomial, see Hilbert
Cartesian curve, see curve comparison exact sequence, 149, 292
Castelnuovo(–Mumford) regularity complement
of a coherent sheaf of OT -modules, 470 n-, 83
of a graded module (reg(M )), 351–363, complete local domain, 110, 153, 155, 157,
373, 383–387, 400–402, 404 412
at and above level 1 (reg1 (M )), 359, complete local ring, 49, 157, 204, 206, 216,
360, 383–387 222, 228, 294, 372, 413, 414
at and above level 2 (reg2 (M )), 360, completion, 109, 158, 159, 161, 200, 203, 204,
373–384 208, 209, 222, 225, 227, 232, 238, 244,
at and above level l (regl (M )), 357–358, 250, 294, 435
369, 400–402, 404 complex
syzygetic characterization of, 362 C̆ech, 82–91, 94, 96, 98, 101, 130, 273–275
of a projective variety (reg(V )), 360, 373 affine n-space (An ), see affine
category Koszul, 94–99, 278
Abelian, see Abelian category projective r-space (Pr ), see projective
C(R), C(R ), xxi cone, 332
*C(R), *C G (R), *C(R ), *C G (R ), xxi, affine algebraic, 332–337, 407, 422–425,
252 427
of all quasi-coherent sheaves of non-degenerate, 332
OT -modules (qcohT ), 458 punctured, 334

of all sheaves of R-modules  456
(S(R)), connected sequence
 negative, 10, 261
of all sheaves of graded R-modules negative strongly, 10, 261
 457
(*S(R)), positive, 12
catenary local ring, see local ring connected sequences
catenary ring, 169, 181, 186, 242, 243, 414, homomorphism of, 11, 263
415 isomorphism of, 11, 263
C̆ech complex, see complex connected topological space, 406
characteristic function, see function Connectedness Bound for Complete Local
Chevalley’s Theorem, 49, 156 Rings, 414
closure Connectedness Criterion for the Special Fibre,
B-, 247 435
S-, 248 connectedness dimension (c(T )), 408–416,
coarsening, 275 429–431
coarser grading, 274 formal, at p (cp (W )), see formal
codimension connectedness dimension
of a closed subvariety, 122 local, at p (cp (T )), see local connectedness
pure, 122 dimension
Cohen’s Structure Theorem, 110, 153, 157, Connectedness Theorem of Grothendieck, see
221, 225, 232, 234, 372 Grothendieck’s
Cohen–Macaulay Connectivity Theorem
local ring, 130, 132–134, 231, 235–241, of Bertini, see Bertini’s Connectivity
244, 245, 308, 474 Theorem
*local graded ring, 315–323, 326, 329, 330 of Bertini–Grothendieck, see
module, 115, 180, 185, 192, 228, 235 Bertini–Grothendieck Connectivity
generalized, see generalized Theorem
Cohen–Macaulay module of Fulton–Hansen and Faltings, 431
quasi-, see quasi-Cohen–Macaulay coordinate ring, 40
module homogeneous, see homogeneous
ring, 181, 185, 189, 191, 192, 237, 238, curve, 338, 341
284, 308, 315, 404 Cartesian, 437
cohomological dimension of an ideal cuspidal, 437
(cohd(a)), 56, 71, 111, 112, 151, fully branched, at p, see fully branched
412–414 curve
Index 487

Macaulay’s, 343 Graded, see Grothendieck’s


rational normal, 341 First Uniqueness Theorem (for secondary
cuspidal curve, see curve representation), see secondary
Deligne Correspondence Theorem, 460 Flat Base Change Theorem, 75
Graded Version of, 461 Graded, 290
Deligne Isomorphism Theorem, 448, 451, 457 flat ring homomorphism, 68, 69, 75, 77, 78,
for Affine Schemes, 458 109, 180, 221, 232, 242, 290, 352
Graded, 454, 457 formal connectedness dimension, at p
depth, 114, 115, 154, 155, 169, 228, 232, 308, cp (W )), 419–421
(
459, 474 formal fibres, 188, 191, 192, 244
a-adjusted, 169 F -rational local ring, see local ring
b-minimum (λba (M )), 169 *free graded module, see module
diagonal, 427 Frobenius
isomorphism, 427 action, 102, 103, 105, 131, 133, 134
dimension homomorphism, 102, 103
b-finiteness relative to a (fab (M )), 167 power (of an ideal), 102
finiteness relative to a (fa (M )), 166, 373 full ring of fractions, 64
fully branched curve, at p, 162
injective, 215, 235
Fulton–Hansen Connectivity Theorem, see
of a coherent sheaf of OT -modules, 471
Connectivity Theorem
of a module, 107, 114, 115, 143, 144, 230,
function
305
characteristic, 364, 368–371
of a Noetherian topological space, 408
cohomological Hilbert, see Hilbert
of a variety, 338
of polynomial type, 366
projective, 177–179, 181, 362
of reverse polynomial type, 367
disconnected topological space, 406
functor
discrete valuation ring, 350
a-torsion (Γa ), 2
dual, see Matlis
a-transform (Da ), 21, 92, 117–120, 124,
duality, see Matlis or local duality
161, 268, 279
end of a graded module (end(M )), see graded commutes with direct limits, 64
module B-transform (DB ), 22, 268, 280
endomorphism ring of canonical module, faithful, 226
238–244, 248–250 φ-coarsening, 275
equidimensional local ring, see local ring generalized local cohomology (HB i ), 22
essential extension, 193, 194, 202, 212, 257 local cohomology, 3
essential subset (with respect to S), 440 commutes with direct limits, 63
*essential extension, 257–259 R-linear, 1, 6
Euler characteristic (χ(F)), 472 section, 457, 460, 468
extended Rees ring (R[aT, T −1 ]), 389, 394 sheaf cohomology, see sheaf
extension of an ideal, under f , 37, 67 shift, see shift
Faltings’ Veronesean, see Veronesean
Annihilator Theorem, 183, 191 generalized Cohen–Macaulay module,
Graded, 309 185–188, 192
Connectivity Theorem, see Connectivity generalized ideal transform, 22, 245
Theorem is independent of the base ring, 37
family of denominators, 441 right derived functors of, 22
family of fractions, 441 universal property of, 29, 32–34, 278
S-local, 441 generalized local cohomology functor (HB i ),
family of homogeneous fractions, 453 see functor
S-local, 453 generating degree, 386
fibre cone, 394 Gorenstein
filter-regular sequence, see R+ -filter-regular local ring, 5, 154–158, 212–221, 225, 227,
sequence 231–237, 243–245, 308, 474
finiteness dimension relative to a (fa (M )), *local graded ring, 309–311, 313, 321–323,
see dimension 326, 330, 372
Finiteness Theorem ring, 154, 216, 238, 308, 315
of Grothendieck, see Grothendieck’s Goto–Watanabe Theorems, 302–308, 328
488 Index

grade, 113, 180, 182–184, 239, 339 right derived functors of, 22
M -, 113 universal property of, 29, 32–34, 278
graded local cohomology, 251–330 indecomposable module, 195–199, 213
and the shift functor, 291 *indecomposable graded module, 295–302,
graded local duality, see local duality 324–326
graded module, 253–257 Independence Theorem, 70
associated to a sheaf, 467 Graded, 289
beginning of (beg(M )), 287 injective
*canonical, see *canonical module cogenerator, 141, 236
end of (end(M )), 287 dimension, see dimension
*indecomposable, see *indecomposable envelope (E( • )), 193–210, 212, 213, 216,
graded module 258, 300
*injective, see *injective hull, 193
graded ring, 253 module, 5, 12, 15, 18, 19, 21, 68–70, 75, 90,
associated, see associated graded ring 100, 141, 148, 193–199, 202, 207
*local, see *local graded ring resolution, 3, 18, 19, 21, 71, 75, 150
positively Z-graded, 282, 326, 332, 347, minimal, 212–217, 299, 301
388, 405 *injective
homogeneous, 347 envelope (*E( • )), 257–259, 295–302
positively Zn -graded, 282 graded module, 256–261, 289, 291, 292,
*simple, see *simple graded ring 295–303, 324–326
Grothendieck’s hull, 257
Connectedness Theorem, 415 resolution, 300, 325
Finiteness Theorem, 183, 192 minimal, 298–302
Graded, 309, 475 integral closure, of a (a), 390
Splitting Theorem, 479 integrally dependent, on a, 389–390
Vanishing Theorem, 107 Intersection Inequality for the Connectedness
group ring, 303 Dimensions
height (ht) of Affine Algebraic Cones, 429
of a proper ideal, 20, 42, 76, 79, 111, 118, of Projective Algebraic Sets, 430
120, 134, 185, 238, 248–250, 305–309, inverse family of ideals, 8, 13, 20, 48
328, 335, 397, 412 inverse polynomials, see module of
of the improper ideal, 20, 169, 248 irreducible
Hilbert component, 407
coefficients, 371, 378, 381–387, 472 ideal, 196
function topological space, 406
cohomological, 371, 373, 472 irrelevant ideal (R+ ), 332
polynomial, 370, 384, 385, 472 isomorphism of connected sequences, see
cohomological, 371, 373, 473 connected sequences
homogeneous Koszul complex, see complex
coordinate ring, 337–339, 345, 360 Krull dimension of a module, see dimension,
localization, 287 of a module
positively Z-graded ring, see graded ring large with respect to S, 450
regular function, 333 Lichtenbaum–Hartshorne Vanishing Theorem,
ring homomorphism, 287 156–159, 216, 294, 412
S-local family of fractions, see family of Graded, 294, 433, 435
homogeneous fractions local arithmetic rank, see arithmetic rank
Homogeneous Prime Avoidance Lemma, 347 local cohomology
homomorphism of connected sequences, see functor, see functor
connected sequences graded, see graded local cohomology
Horrocks’ Splitting Criterion, 478 module, see module
hypersurface, 425 local connectedness dimension, at p (cp (T )),
ideal transform, 21, 40, 93, 443, 448, 462 417–420
conditions for exactness of, 117–121, 124 local dimension, at p (dimp T ), 417
geometrical significance of, 41, 44, 45 local duality, 216–221, 225, 234
is independent of the base ring, 37 graded, 310–312, 318–320, 323, 326
Index 489

Local Duality Theorem, 218 graded, see graded module


Graded, 310, 372 indecomposable, see indecomposable
local ring module
analytically irreducible, 161, 162, 250 injective, see injective
Artinian, see Artinian local cohomology, 3
catenary, 242, 243 of deficiency, 228
Cohen–Macaulay, see Cohen–Macaulay of inverse polynomials, 276, 277, 323, 325
complete, see complete local ring p-secondary, 139
equidimensional, 181, 250 quasi-Cohen–Macaulay, see
F -rational, 134 quasi-Cohen–Macaulay module
Gorenstein, see Gorenstein representable, see representable module
of a quasi-affine variety, at q (OW,q ), 122 secondary, see secondary
regular, see regular local ring sum-irreducible, see sum-irreducible
universally catenary, see universally module
catenary Monomial Conjecture, 131
local subdimension, at p (sdimp T ), 417–421 morphism
local vanishing ideal, at p (IV,p (C)), 419–421 of sheaves, 456
Local-global Principle for Finiteness induced, 456
Dimensions, 189, 191 of varieties, 122, 123
locally free coherent sheaf, see sheaf M -regular sequence, see sequence
*local graded ring, 286 M -sequence, see sequence
Cohen–Macaulay, see Cohen–Macaulay maximal, see sequence
Gorenstein, see Gorenstein poor, see poor M -sequence
locus weak, see weak M -sequence
non-S2 , 248, 249, 344 multiplicatively closed subset, xxi
S2 -, 244 multiplicity, 371, 386
Macaulay’s curve, see curve Mumford’s Regularity Bound, 473
Matlis natural gradings, 454, 461, 462
dual, 200 n-complement, see complement
duality, 199–210, 225, 228, 229 negative
graded, 312, 372 connected sequence, see connected
Duality Theorem, 204 sequence
Partial, 209 strongly connected sequence, see connected
maximal M -sequence, see sequence sequence
*maximal graded ideal, 286 Noetherian topological space, 407–411, 417,
Mayer–Vietoris sequence, 51–56, 64, 67, 79, 418, 433, 440
121, 412 dimension of, see dimension
graded, 288, 435 Non-vanishing Theorem, 109, 144
metric topology, 46 non-S2 -locus, see locus
M -grade, see grade normal projective scheme, see projective
minimal graded free resolution, 362 numerical invariant, 378
module finite, 378
acyclic, see acyclic numerical polynomial, 366
Artinian, see Artinian
a-torsion, 3, 17–19, 37, 85, 92 open set, 179, 244
a-torsion-free, 3, 17, 67 affine, see affine
B-torsion, 21, 27, 29, 32, 33 quasi-affine, see quasi-affine
B-torsion-free, 21 parameter, 186
Buchsbaum, see Buchsbaum Partial Matlis Duality Theorem, see Matlis
canonical, see canonical module p-component, 417
*canonical, see *canonical module pole, 46
Cohen–Macaulay, see Cohen–Macaulay isolated, 46
faithful, 110, 201, 232, 249, 327, 328 polynomial ring, 40, 58, 71, 72, 91, 159, 181,
*free, 256, 257, 266, 303, 306, 307, 317, 192, 254, 275, 282, 284, 323, 324, 326,
318, 320, 344, 362, 381 329, 330, 336, 339, 342, 343, 347, 350,
generalized Cohen–Macaulay, see 353, 354, 357, 360, 362, 378–380, 386,
generalized Cohen–Macaulay module 387, 426, 451, 473, 476, 478, 479
490 Index

Veronesean subring of, see Veronesean *restriction property, 264–274, 277–280,


poor M -sequence, 112, 127–130 287–291, 455, 457, 461, 462, 465
positive connected sequence, see connected R+ -filter-regular sequence, 399–402
sequence ring homomorphism
postulation number, 371, 385–387 flat, see flat ring homomorphism
prime characteristic, 50, 102, 103, 127, homogeneous, see homogeneous
131–134 ringed space, 456
product of varieties, 427 R-linear functor, see functor
projective r-regular (in the sense of
algebraic set, 336 Castelnuovo–Mumford), 351, 354, 356,
dimension, see dimension 357, 470
r-space (Pr (K)), 336–343, 345 at and above level l, 351, 353, 354, 356, 358
complex (Pr ), 337 Rung’s display, 432, 434
scheme, 466–479 saturated ideal, 387
normal, 476 saturation of an ideal, 387
regular, 476, 477 scheme
spectrum (Proj(R)), 375, 433–437, 441, affine, see affine
466–479 projective, see projective
variety, 336–343, 345, 360, 426, 431 S-closure, see closure
*projective graded module, 256 Second Uniqueness Theorem (for secondary
projectivization, 337 representation), see secondary
punctured secondary
cone, see cone module, 139
spectrum, 409 representation, 139–142, 144, 158, 159
quasi-affine First Uniqueness Theorem for, 140
open set, 418 minimal, 139, 221
variety, 39, 40, 42, 44–46, 122, 127, 162, Second Uniqueness Theorem for, 141
407 section functor, see functor
analytically irreducible, at p, 162 sections of a sheaf, see sheaf
analytically reducible, at p, 162 semi-local ring, 192, 250
quasi-Cohen–Macaulay module, 185 sequence
quasi-compact topological space, 190, 406, M -, 112
440, 444, 453 maximal M -, 113
quasi-projective variety, 337, 418, 419 M -regular, 112
Serre’s
radical ideal, 180, 419 Affineness Criterion, 124
rational normal curve, see curve Cohomological Criterion for Local
rational normal surface scroll, 345 Freeness, 477, 479
Ratliff’s Theorem, 188, 191, 417 condition S2 , 224, 233, 234, 242–245, 249,
reduction, 2, 388–398 250
minimal, 389–391, 395, 398, 404 condition Si , 232
number, 389, 397, 402, 404 Criterion for Generation by Global
*reduction, 397–399 Sections, 470
Rees ring (R(a)), 436 Finiteness Theorem, 469, 476
extended, see extended Rees ring Serre–Grothendieck Correspondence, 463,
regular local ring, 5, 72, 73, 79, 110, 111, 464, 468–472, 478
115–117, 153, 154, 157, 159, 181, 185, classical form of, 467
222, 225, 232, 234, 306, 474 Severi–Enriques–Zariski–Serre Vanishing
regular projective scheme, see projective Theorem, 475
regular ring, 76, 176, 180–182, 186, 188, 303, sheaf
304, 309, 335, 475 associated to M (M , (M )0 ), 458, 466
regularity (Castelnuovo–Mumford), see coherent
Castelnuovo(–Mumford) regularity generated by global sections, 470
Regularity Bound of Mumford, see locally free, 476–479
Mumford’s Regularity Bound cohomology functor, 460
representable module, 140, 141 induced, 456
restriction map, 442 flasque, 459
Index 491

of ideals, 468 transform


of R-algebras, 456 a-, 21

of R-modules, 456 B-, 22, 29
sections of, 456 trivial extension, 115, 237, 238, 315
splits (completely), 478 twisted
twisted, 467 cubic, 341, 343
sheaves, morphism of, see morphism quartic, 343
shift sheaf, see sheaf
functor, 253, 291, 321, 322, 324, 326, 454 unique factorization domain (UFD), 118, 245,
of a complex, 88 303, 304
S2 -ification, 245, 248–250 universally catenary
in graded situations, 327–329, 343, 344 local ring, 188, 191, 192, 416, 417
*simple graded ring, 298, 303, 304, 306 ring, 188, 192
simplicial complex, 255, 326
vanishing ideal, 58, 332, 338, 360, 361, 373,
S-local family of fractions, 441–455
419
S2 -locus, see locus
local, at p (IV,p (C)), see local vanishing
socle, 202
ideal
special fibre, 433–436
Vanishing Theorem
splitting type, 478
Lichtenbaum–Hartshorne, see
stalk, 456, 457, 459, 466, 470, 473, 474, 476
Lichtenbaum–Hartshorne Vanishing
Stanley–Reisner ring, 255, 326
Theorem
S-topology, 440, 441, 466
of Grothendieck, see Grothendieck’s
structure sheaf of an affine scheme, 458
Severi–Enriques–Zariski–Serre, see
subdimension (sdim T ), 410–417, 429, 430
Severi–Enriques–Zariski–Serre
local, at p (sdimp T ), see local
Vanishing Theorem
subdimension
variety, 337
subsystem of parameters, 130, 133
affine, see affine
sum-irreducible module, 141
of an ideal (Var(a)), xxi
support of a coherent sheaf (Supp F ), 471
projective, see projective
surface, 338
arithmetically Cohen–Macaulay, see
affine, see affine
arithmetically
symbolic power (p(t) ), 49, 50, 155, 156
arithmetically Gorenstein, see
symmetric semigroup, 329, 330
arithmetically
system of graded ideals, 267, 271, 277, 280
quasi-affine, see quasi-affine
system of ideals, 20, 21, 23, 32, 48–50, 66,
quasi-projective, see quasi-projective
155, 156, 245–248
variety
subsystem of, 245, 246, 248
Veronesean, 339, 341, 342, 360
system of parameters, 130–133, 186–188, 228,
functor, 281, 373, 402
229, 397
map, 340
standard, 186–188
syzygetic characterization of regularity, see submodule (M (r,s) ), 281, 341
Castelnuovo(–Mumford) regularity subring (R(r) ), 281, 394
of a polynomial ring, 283, 284, 330, 339,
tight closure (a∗ ), 131–134 342, 343, 356, 361
tightly closed ideal, 131–134 transformation, 340, 341
topology
metric, see metric topology weak M -sequence, 187
S-, see S-topology Zariski topology, 39, 244, 336, 441
Zariski, see Zariski topology on Proj(R), 433
on Proj(R), see Zariski topology Zariski’s Main Theorem on the Connectivity
total order compatible with addition, 253, 255 of Fibres of Blowing-up, 437

You might also like