London Mathematical Society Student Texts 33
A Primer of Algebraic D-modules
S. C. Coutinho IMPA, Rio de Janeiro
o
CAMBRIDGE
~ UNIVERSITY PRESS
Published by the Press Syndicate of the University of Camhrioge The Pitt Building. T11ImpingLOn Street, Cambridge CB2 lRP 40 West 20th Street, New York, NY 10011-4211, USA 10 Stamford Road, Oakleigh, Melbourne 3166, Australia © Cambridge University Press 1995 First published 1995 Printed in Great Britain at the University Press, Cambridge
A catalogue recordfor this book is available from the British library Library of Congress cataloging in publication dnfa
Coutinho. S.C. A primer of algebraic D-modules / S.C. Coutinho p cm. - (London Mathematical Society Student Texts; 33) Includes bibliographical references (p. ) ISBN 0-521-55119-6 (hardback). 0-521-55908-1 (paperback) 1. D-modules. 1. Title. II. Selles. QA614.3 .C68 1995 512'.4-dc20 95-6628 CIP ISBN 0521551196 hardback ISBN 0521 55908 1 paperback
For Sergio Montenegro
CONTENTS Preface Introduction 1. The Weyl algebra 2. Algebraic D-modules 3. The book; an overview 4. Pre-requisites
xi
1 3 5 6
Chapter 1. The Weyl algebra 1. Definition 2. Canonical form 3. Generators and relations 4. Exercises
12
Chapter 2. Ideal structure of the Weyl algebra. 1. The degree of an operator 2. Ideal structure 3. Positive characteristic 4. Exercises
14 16 17 18
Chapter 3. Rings of differential operators. 1. Definitions 2. The Weyl algebra 3. Exercises Chapter 4. Jacobian Conjecture. 1. Polynomial maps 2. Jacobian conjecture 3. Derivations 4. Automorphisms 5. Exercises Chapter 5. Modules over the Weyl algebra. 1. The polynomial ring 2. Twisting 3. Holomorphic functions 4. Exercises
8 9 10
20 22
24
26 28 30
32 34
36 38 40 41
viii
Contents
Chapter 6. Differential equations. 1. The D-module of an equation 2. Direct limit of modules 3. Microfunctions 4. Exercises Chapter 7. Graded and filtered modules. 1. Graded rings 2. Filtered rings 3. Associated graded algebra 4. Filtered modules 5. Induced filtration 6. Exer cises Chapter 8. Noetherian rings and modules. 1. Noetherian modules 2. Noetherian rings 3. Good filtrations 4. Exercises Chapter 9. Dimension and multiplicity. 1. The lIilbert polynomial 2. Dimension and multiplicity 3. Basic properties 4. Bernstein's inequality 5. Exercises Chapter 10. Holonomic modules. 1. Definition and examples 2. Basic properties 3. Further examples 4. Exercises
44 46 48 50
53 55 57
59
60 62
65
67 70 72
74 77
80 82 84
86 88 91 95
Chapter 11. Characteristic varieties. 1. The characteristic variety 2. Symplectic geometry 3. Non-holonomic irreducible modules 4. Exercises
100 104 106
Chapter 12. Tensor products. 1. Bimod ules 2. Tensor products 3. The universal property
108 109 111
97
Contents 4. Basic properties 5. Localization 6. Exercises
be
113 117 119
Chapter 13. External products. 1. External products of algebras 2. External products of modules 3. Graduations and filtrations 4. Dimensions and mUltiplicities 5. Exercises
121 122 124 127 128
Chapter 14. Inverse Image. 1. Change ofrings 2. Inverse images 3. Projections 4. Exercises
130 132 134 135
Chapter 15. Embeddings. 1. The standard embedding 2. Composition 3. Embeddings revisited 4. Exercises
138 139 142 144
Chapter 16. Direct images. 1. llight modules 2. Transposition 3. Left modules 4. Exercises
146 147 150. 153
Chapter 17. Kashiwara's theorem. 1. Embeddings 2. Kashiwara's theorem 3. Exercises
154 156 160
Chapter 18. Preservation of holonomy. 1. Inverse images 2. Direct images 3. Categories and functors 4. Exercises
162 165 166 170
Chapter 19. Stability of differential equations. 1. Asymptotic stability 2. Global upper bound 3. Global stability on the plane 4. Exercises
171 173
176 178
x
Contents
Chapter 20. Automatic proof of identities. 1. Holonomic functions 2. Hyperexponential functions 3. The method 4. Exercises
179 180 183 186
Coda
188
Appendix 1. Defining the action of a module using generators
191
Appendix 2. Local inversion theorem
194
References
197
Index
203
PREFACE
As its title says, this book is only a primer; in particular, you will learn very little 'grammar' from it. That is not surprising; to speak the language of algebraic D-modules fluently you must first learn some algebraic geometry and be familiar with derived categories. Both of these are beyond the bounds of an elementary textbook. But you can expect to know the answers to two basic questions by the time you finish the book: what are D-modules? and why D-modules? It is particularly easy to answer the latter, because D-module theory has
many interesting applications. Hardly any area of mathematics has been left untouched by this theory. Those that have been touched range from number theory to mathematical physics. I have tried to include some real applications, but they are not by any means the ones that have caused the greatest impact from the point of view of mathematics at large. To some, they may even seem a little eccentric. That reflects two facts. First, and most important, this is an elementary book. The most interesting applications (to singularity theory and representations of algebraic groups, for example) are way beyond the bounds of such a book. Second, among the applications that were elementary enough to be presented' here, I chose the ones that I like the most. The pre-requisites have been kept to a minimum. So the book should be accessible to final year undergraduates or first year post-graduates. But I have made no effort to write a book that is 'purely algebraic'. Such a book might be possible, but it would not be true. One of the attractions of the theory of D-modules is that it sits across the traditional division of mathematics into algebra, geometry and analysis. It would be a pity to lose that. So this is a book about mathematics, and in it you will find algebras and modules, differential equations and special functions, all in easy conviviality. The introduction contains a detailed description of the pre-requisites. While writing this book I often worried about the language. I am too fond of English not to shiver at tIl(' idrn of hadly
>lbu~illg
it.. But English
xii
Preface
is not my first language, and I am also well aware of my deficiencies. I can only hope that the many revisions have spared me from sharing the fate of the master of the brigantine in Conrad's Lord Jim whose 'flowing English seemed to be derived from a dictionary compiled by a lunatic'. The material in this hook is not original. I have only tried to present the foundations of D-module theory for the beginner that I was, when I started working on the subject ten years ago. In a sense this book is only a compilation: while writing it I have tapped many sources. I have truly plundered the literature for results and exercises to be included in the book. Since it would be very difficult to give references for all of these, I have limited my attributions to the main results and applications. The book grew out of notes for a basic course on algebraic D-modules taught at the Federal University of Rio de Janeiro. The constant questioning of the students greatly improved the exposition, and directed me to many interesting examples. Many people offered hints, provided references or explained some of the topics to me, particularly G. Meisters, C. Gutierrez, S. Toscano de Melo and A. Pacheco. M.B. Alves and D. Levcovitz read parts of the book and contributed many useful suggestions and corrections. D. Tranah and R. Astley, at Cambridge University Press, put up patiently with the naive questions of a novice and offered many helpful suggestions. Martin Holland read most of the book. His comments saved me from many mistakes while helping to make the exposition clearer; and his unflagging support and friendship kept me at work and made the book possible. To Andrea lowe whatever else was life. The work in this book was partially supported by a CNPq grant, and benefited from the excellent working conditions generously provided by the Institute of Pure and Applied Mathematics (IMPA) .
Rio de Janeiro, April 1995.
INTRODUCTION This introduction begins with an account of the history of the subject and its relations with other areas of mathematics. This is followed by an overview of the whole book, and a brief description of its chapters. 1. THE WEYL ALGEBRA.
The history of the Weyl algebra begins with the birth of quantum mechanics. The year was 1925. A number of people were trying to develop the principles of the mechanics that was to explain the behaviour of the atom. One of them was Werner Heisenberg. His idea was that this mechanics had to be based on quantities that could actually be observed. In the atomic model of Bohr, there was much talk about orbits; but these are only remotely connected with the things that are actually observed. In the words of Dirac: 'The things that are observed, or which are connected closely with the observed quantities, are all associated with two Bohr orbits and not with just one Bohr orbit: two instead of one. Now, what is the effect of that?' [Dirac 78]. To Heisenberg's initial dismay, the 'effect of that' was the introduction of noncommutative quantities. Let Dirac continue with the story: Suppose we consider all the quantities of a certain kind associated with two orbits, and we want to write them down. The natural way of writing down a set of quantities, each associated with two elements, is in a form like this:
(~
X
X
X
X
X
X
X
X
X
X
X
X
.
"
OJ
an array of quantities, like this, which one sets up in terms of rows and columns. One has the rows connected with one of the states, the columns connected with the other. Mathematicians call a set of quantities like this a matrix.
The sequence of events is a little more intricate than Dirac's comments will have us belil'VP. In 1925 matriC('H were not part of the toolkit of every
2
Introduction
physicist, as they now are. What Heisenberg originally introduced were quantum theoretical analogues of the classical Fourier series. These were supposed to describe the dynamical variables of the atomic systems; and they did not commute. Unable to go as far as he had hoped, Heisenberg decided to sum up his ideas in a paper that he presented to Max Born. Born was acquainted with matrices and was the first to realize that matrix theory offered the correct formalism for Heisenberg's ideas. According to Born's approach, the dynamical variables (velocity, position, momentum) should be represented by matrices in quantum theory. Denoting the position matrix by q and the momentum matrix by p, one may write the equation for a system with one degree of freedom in the form pq- qp=
m.
The work on matrix mechanics began with Born and his assistant P . Jordan, soon to be joined by Heisenberg himself. Early on they understood that the fundamental equation above could not be realized by finite matrices; the
matrices of quantum theory had to be infinite. Matrix mechanics was soon followed by other formalisms. First came E. Schrodinger's wave mechanics. In this approach everything begins with a partial differential equation; a more familiar object to physicists. There was also Dirac's formalism. He chose the relations among the dynamical variables as his starting point. The dynamical variables subject to those relations were the elements of what he called quantum algebra. From Dirac's point of view, one is interested in polynomial expressions in the dynamical variables momentum, denoted by p, and position, denoted by q.
It is assumed that the variables satisfy the (normalized) relation
pq - qp
= 1.
This is what we now call the first Weyl algebra. In particular,
he showed how one could use the relation between p and q to differentiate polynomial expressions with respect to p and to q. The Weyl algebras of higher index appear when one considers systems with several degrees of freedom. The relations among the operators in this case had been established as early as September 1925 by Heisenberg. Dirac's point of view received a masterly presentation in H. Weyl's pioneering book The theor'lJ of gmups
and
quant~'m mechanir:.~ ,
[Wuyl 50].
Introduction
3
From that point onwards the mathematicians were ready to take over. Of very special interest is the paper [Littlewood 33] of D. Littlewood. He begins by saying that although finite dimensional algebras had been intensively studied, the same was not true of algebras of infinite dimensions. One of the algebras he studies is Dirac's quantum algebra. Littlewood carefully constructs (infinite) matrices p and q for which the equation pq - qp
= 1 is
satisfied; see Exercise 1.4.10. In his paper Littlewood established many of the basic properties of the Weyl algebra. He showed that the elements of the algebra have a canonical form (Oh. 1 ,§2) and that the algebra is a domain (Ch. 2, §1). He also showed that the relation pq - qp = 1 is not compatible with any other relation. Or, as we would now say, the only proper ideal of this algebra is zero (Oh. 2, §2). The modern age in the theory of the Weyl algebra arrived when its connections with Lie algebras were realized. Suppose that n is a nilpotent Lie algebra over C. Let U(n) be its enveloping algebra. The quotient of U(n) by a primitive ideal is always isomorphic to the Weyl algebra; see [Dixmier 74, TMoreme 4.7.9]. In [Dixmier 63], the notation An was introduced for the algebra that corresponds to the physicist's system with n degrees of freedom. The name Weyl algebra was used by Dixmier as the title of [Dixmier 68] following, as he says, a suggestion of 1. Segal in [Segal 68]. The increasing interest in noncommutative noetherian rings that followed' A. Goldie's work and the intense development of the theory of enveloping algebras of Lie algebras contributed to keeping up the interest in the Weyl algebra. That is not the end of the story though: another theme was added in the seventies under the guise of D-module theory. 2, ALGEBRAIC D-MODULES.
Under the cryptic name of D-module hides a modest module over a ring of differential operators. The importance of the theory lies in its manifold applications, which span a vast territory. The representation theory of Lie algebras, differential equations, mathematical physics, singularity theory and even number theory have been influenced by D-modules. One of the roots of the theory is the idea of considering a differential equation as a module olfer a ring of differential operators discussed in Ch.
Int1'Oduction
4
6. This approach goes back at least to the sixties, when it was applied by B. Malgrange to equations with constant coefficients. A turning point was reached in 1971 with M. Kashiwara's thesis 'Algebraic study of systems of partial differential equations', where the same approach was systematically applied to equations with analytic coefficients. In this context, the theory is often called by the alternative name of algebraic analysis. At the same time, in the Soviet Union, I. N. Bernstein was developing the theory of modules over the Weyl algebra. His starting point, however, was entirely different. I.M. Gelfand had asked in the International Congress of Mathematicians of 1954 whether a certain function of a complex variable, which was known to be analytic in the half plane
~(z)
> 0, could be ex-
tended to a meromorphic function defined in the whole complex plane. The problem remained open until 1968, when M.F. Atiyah and, independently, Bernstein and I.S . Gelfand gave affirmative answers. Both proofs made use of Hironaka's resolution of singula1-ities, a very deep and difficult result. Four years later, Bernstein discovered a new proof of the same result that was very elementary. The key to the proof was a clever use of the Weyl algebra. In his papers, Bernstein introduced many of the concepts that we will study in this book. Of course the theory of D-modules is not restricted to the Weyl algebra. The theory has two branches: an analytic and an algebraic one ; depending on whether the base variety is analytic or algebraic. Highly sophisticated machinery is required in the study of general D-modules, and the most important results cannot be introduced without derived categories and sheaves. Perhaps the most spectacular result ofthe theory to date is the Riemann-
Hilbert correspondence obtained independently by M. Kashiwara and Z. Mebkhout in 1984. This is a result of noble parentage; it can be traced to Riemann's memoir on the hypergeometric function. Its genealogical tree includes the work of Fuchs on differential equations with regular singular points and Hilbert's 21st problem 0[1900. Very roughly speaking, the correspondence establishes an (anti-) equivalence between certain differential equations (described in terms of D-modules) and their solution spaces. Unfortunately, the correspo ndence requires deep res ults o f ca tegor y t.lleory au cl cannot b(~
Introduction
5
included in an elementary book. 3. THE BOOK: AN OVERVIEW. This book is about modules over the Weyl algebra, and the point of view is that of algebraic D-modules. It is also fair to say that this is a book about certain aspects of the representation theory of the Weyl algebra. So we will have a lot to say about irreducible modules and the dimension of modules, for example. The book can be divided into two parts. The first part, which goes up to Ch. 11, is concerned with invariants of modules over the Weyl algebra; the most important being the dimension and the multiplicity. The first two chapters deal with ring theoretic properties of the Weyl algebra itself. Most importantly, we show that it is a simple domain. Ch. 3 establishes the Weyl algebra as a member of the family of rings of differential operators. There is much talk of derivations in this chapter, and they are put to good use in the next chapter. Ch. 4 contains the first of our applications. It consists of using a conjecture about the Weyl algebra to derive the Jacobian conjecture.
With Ch. 5 we get into the realm of representation theory. The purpose of the chapter is to describe a few important examples of modules over the Weyl algebra. The relation of these modules with differential equations is' found in Ch. 6, which also includes an elementary description of the module of microfunctions.
These are used as generalized solutions of differential
equations. Chs. 7 to 9 form a sequence which culminates in the definition of the dimension and multiplicity of a module over the Weyl algebra and the study of their properties. Place of honour is given to Bernstein's inequality: the dimension of a module over the n-th Weyl algebra is an integer between nand
2n. The modules of minimal dimension form such a nice category that they have a special name: holonomic modules. The whole of Ch. 10 is dedicated to them. Ch. 11 requires a smattering of algebraic geometry; it deals with the relation between D-modules and symplectic geometry. The key concept is another invariant of a D-module, the charactedstic va1'iety. This allows us to give a geometrical int!'fpfdation to the Uim('ll.'lion previously ddined.
Introduction
6
The emphasis now shifts from invariants to operatiolls, which are the theme of the second part. These operations are geometrical constructions which use a polynomial map to produce new D-modules. Since their definition depends on the use of the tensor product, which may not be familiar to some readers, we have included a discussion of them in Ch. 12. Chs. 13 to 16 contain the definitions and examples of the three main operations that we apply to modules over the Weyl algebra: external products, inverse images and direct images. The results are served in homreopathic doses, since the calculations tend to be crammed with detail. Kashiwara's theorem is one of the pearls of the theory. A simple, rather meagre version is described in Ch. 17, but it will return in greater splendour in Ch. 18. In this chapter we also show that all the operations previously described map holonomic modules to holonomic modules. This is very mystifying since the dimension of a module is not preserved by some of these operations. Finally we return to applications in Chs. 19 and 20. The former is concerned with the global stability of ordinary differential equations. We will discuss conditions under which a system of differential equations with polynomial coefficients has a global stability point. The key lemma has a very neat D-module theoretic proof due to van den Essen. It is discussed in detail in §2. The proof of the stability theorem is sketched in §3. Ch. 20 is about the work of Zeilberger and his collaborators on the automatic computation of sums and integrals. That is, how can D-modules help a computer to calculate a definite integral? 4. PRE-REQUISITES . All the algebra required in the book can be found in [Cohn 84]. Like all general rules, this one has exceptions; these are Ch. 11 and Ch. 18, §3. We have already mentioned that one needs to know some algebraic geometry to follow Ch. 11. All the required results will be found in [Hartshorne 77, Ch. 1]. In §3 of Ch. 18 we rewrite the results of Chs. 14 to 17 using categories. The subjects of these sections are not used anywhere else in the book, except in some exercises. The book also requires a good knowledge of analysis to be properly appreciated. This is particularly true of examples and llpplieatiol1s. However, most,
Introduction
7
of the results we use are part of the standard undergraduate curriculum; and we will give several references to those which are not so well-known. Ch. 19, in particular, requires some results about ordinary differential equations. A good reference for our needs is [Arnold 81]. The book is linearly ordered and the order is almost, though not quite, total. There are two obvious exceptions: Ch. 11 and §3 of Ch. 18 depend on everything that comes before them, but are not used anywhere else in the book. The only other exceptions are Chs. 4 and 6. We will need Ch. 4, §1 later on, but that will be only in Ch. 14.
CHAPTER 1 THE WEYL ALGEBRA We will describe the main protagonist of this book, the Weyl algebra, in two different ways: as a ring of operators and in terms of generators and relations. 1. DEFINITION
In this section the Weyl algebra is introduced as a ring of operators on a vector space of infinite dimension. Let us fix some notation. Throughout this book, K denotes a field of characteristic zero and K[X] the ring of polynomials K [Xl, ... , Xn] in n conunuting indeterminates over K . The ring K[X] is a vector space of infinite dimension over K. Its algebra of linear operators is denoted by EndK(K[X]). Recall that the algebra operations in the endomorphism ring are the addition and composition of operators. The Weyl algebra will be defined as a subalgebra of EndK(K[X]) . Let Xl, ... , xn be the operators of K[X] which are defined on a polynomial
f
E K[X] by the formulae Xi(f) =
operators defined by 0;(1)
= o!lOXj.
Xi·
f. Similarly,
01, .. . ,
an
are the
These are linear operators of K [X ].
The n-th Weyl algebra An is the K-subalgebra of End](K[X]) generated by the operators
xt. ... ,xn and
0[, . .. , an. For the sake of consistency, we
write Ao = K. Note that for n 2: m, the action of the operators of Am on K[X] is welldefined. Thus Am is a sub algebra of An in a natural way. We sometimes write An(K) instead of An, if it is necessary to make explicit the field over which the algebra is defined. According to our definition, the elements of An are linear combinations over
K of monomials in the generators Xl , .. . , Xn, 01 , . . . , an. However, one has to be careful when representing the elements of An because this algebra is not commutative. This is easily checked, as follows . Consider the operator
a; ·Xi
and apply it to a polynomial f E K[X]. Using the rule for the differentiation of a product, we get a;· xi(f)
= xiof lax, + f.
Oi· Xi = Xi· 0, + 1
In other words,
1. The Weyl algebra
9
where 1 stands for the identity operator. It is better to rewrite this formula using commutators. If P, Q E An then their commutator is the operator [P, Q]
=
P . Q - Q. P. The formula above becomes [Bj, Xj]
=
1. Similar
calculations allow us to obtain formulae for the commutators of the other generators of An. These are summed up below:
[Bj,xJ ]
= 6jj '1,
[B j, Bj ] = [Xj, Xj] where 1 1 if i
=
~
i,j
~
= 0,
n. Recall that 6ij is the Kronecker delta symbol: it equals
j and zero otherwise. A final observation. We have denoted the
operator 'multiplication by x;' by the symbol Xj. From now on, we shall follow the standard convention and write
for both the variable and the
Xi
corresponding operator. This tends to make the notation less cluttered. For the same reason we shall dispense with the subscripts for the generators of
Ab and write them simply as x and B. 2. CANONICAL FORM
In this section we construct a basis for the Weyl algebra as a K-vector space. This basis is known as the canonical basis. If an element of An is written as a linear combination of this basis then we say that it is in.
canonical form. Of course, to compare two elements in canonical form it is enough to compare the coefficients of their linear combinations, and that is easily done. It is easier to describe the canonical basis if we use a multi-index notation .
A multi-index a is an element of Nn; say a
= (a"
.. . ,an), Now by xQ we
mean the monomial xr' . .. X~" . The degree of this monomial is the length
lal of the multi-index a, of multi-indices in
j\Jn
namely
10'1 = al + ... + an'
is itself a multi-index in
j\J2n,
Notice that a pair (0,(3) . so it makes sense to talk
of its length.
2.1 PROPOSITION. Tile set B vector space over K.
=
{xQ(]ii
0,(3 E
Nn} is a basis of An as a
The proof of this proposition uses a formula for the derivative of polynomials in terms of muiti-illllices tlla!. we state below. The factorial of a
1. The Weyl algebra
10
multi-index fi E Nn is defined by !3! terms of powers of the operators 2 .2 LEMMA. Let u
u,!3
a;.
=!3t! . .. fin!'
The formula is written in
The proof is left to the reader.
E Nn and assume that
lui 5 IfiI·
Then [)iJ(x")
=!3!
if
= !3, and zero otherwise.
PROOF OF THE PROPOSITION:
It is easy to see that the elements of B gen-
erate the Weyl algebra as a vector space. Consider a monomial on the generators of An. Using the relations of §l, one shows that if / E K[x), then
- f . a. = 8/ lax,. That allows us to bring all powers of x's to of all the a's. By doing that, the monomial automatically ends up A• . f
the left written
as a linear combination of the elements of B. Now to the uniqueness. Consider a finite linear combination of elements of B, say D
= E c (Jx a
n
(j3. We must show that if some c<>fJ is non-zero then
D ¥= O. But D is a linear operator of K[X]. Hence D¥-O if and only if there exists a polynomial/for which D(f)
¥= O.
Let u be a multi-index which satisfies C<>fJ = 0, for all indices fJ such that
We construct such an /. CM
lfil < lui.
¥=
0 for some index
Ct,
but
A straightforward calculation
using Lemma 2.2 shows that D(x") = u! En Cauxn . This is non-zero since at least one of the coefficients C nu is non-zero by the choice of u. Thus is the required polynomial.
f = x"
3. GENERATORS AND RELATIONS Another way to define the Weyl algebra is by generators and relations. More precisely, we may write the Weyl algebra as a quotient of a free algebra in 2n generators . The ideal that we factor out is generated by the relations calculated in §l. The free algebra K {Zl, ... , Z2n} in 2n generators is the set of all finite linear combinations of words in zl,.' " Z2n.
Multiplication of two mono-
mials is simple juxtaposition. We may define a surjective homomorphism
An by Au .
--+
1. The Weyl algebra 3.1 THEOREM.
11
¢ is an isomorphism.
PROOF: Exactly as in the proof of Proposition 2.1, we may use the relations between the classes
Zj
+J
to show that every element of K {Zb" . ,Z2n} / J
may be written as a linear combination of monomials of the form
By Proposition 2.1, the images of these monomials under
¢ form a
basis of
An as a vector space over K. In particular, the monomials must be linearly independent in K {Zl' ... , Z2n} / J. Hence ~ is an isomorphism of vector spaces and, a fortiori, an isomorphism of rings. We may apply this theorem to the construction of automorphisms of An. The corollary contains a very important example, which will be used in later chapters. 3.2 COROLLARY. Let m
< n be positive integers. Choose polynomials j; E
K[X], for 1 SiS n, as follows: if ism, then fi is a polynomial in the variables Xm+l, . .. , Xn; otherwise the formulae
fi = O. The map u : An ~ An defined by
U(Xi) = Xj + Ii,
is an automorphism of An. PROOF: Define a homomorphism
= Xi+ Ii
Choose i,:j such that 1 S i,:j ::; n. It is clear that
[
12
1. The Weyl algebra
The first commutator is O;j
+ 8/j 18x;; the second equals
Subtracting these two terms, and using the hypotheses on the !'s, we get that
= Ojj.
A similar calculation shows that
Thus
(7
= O.
= ~ of An.
A similar argument shows that the map r defined by the formulae
r(x;)
= Xi -
J;,
is an endomorphism of An. It is now easy to check that r is the inverse of 0'. Thus
0'
is an automorphism of An.
The automorphisms of An playa very important role in the theory, as we shall see. In particular, it is not known when an endomorphism of An is an automorphism. This question is related to the Jacobian conjecture, and will be discussed in Ch. 4, §4.
4. EXERCISES 4.1 Let / E K[X). Prove that [8;,/]
= 8/ 18x;
in An.
4.2 Find the canonical fOfm of the elements of A3 listed below.
(1) (2)
8i. xi &i. Xl . 8.! . X3 + X3 .81 . XI
(3) 81
, XI .
(4) 8r . xi
82
.
8.! . X 3
• X2
+ 8i . x~
4.3 Let R be a K -algebra and AI, A2 E K. Show that the commutator of two elements of R satisfies [a, AJ b1
+ A2b2] = Ada, bl ] + A2[a, b2)·
4.4 Let Vbe an infinite dimensional real vector space with basis {Un : n EN} . DenTlf' ~ ,'rl E: E"dlR Vhy
eC"7<) =
'II.,d l (n + 1)1/2 for all n , ""('110) = 0 a nd
1. The Weyl algebra l 2 Un- I n /
1/(Un) =
Show that R
~
13
for n > O. Let R be the algebra generated by ~ and 1/.
AI(IR).
4.5 Let R be a ring. Let a, b, e E R. Show that Jacobi's identity [a, [b,
ell +
[e, [a, b]] + [b, [e, all = 0 holds in R. 4.6 Show that if dEAn and
f, 9 E K[XJ, then
[[d,j ], gl
= [[d, gJ, fl·
Hint: Jacobi's identity. 4.7 Let S£(2, K) be the group of 2 x 2 matrices with coefficients in K and determinant 1. Let (aiJ') E S£(2 , K). Show that there exists an automorphism q
of
Al
which satisfies
q(8) = a H 8
+ a12X
and
q(x) = a218
4.8 Show that there exists an automorphism F : An F(x;)
~
+ a22X.
An which satisfies
= 8; and F(8;) = -Xi, for i = 1,2, .. . , n.
4.9 Let Mn(K) denote the algebra of n x n matrices with entries in K . If
A E Mn(K) show that the map t/J : Mn(K) ....... Mn+l (K) defined by
t/J(A)
=
(~ ~)
is an injective K-algebra homomorphism. So we may assume that Mn(K)
Mn+l(K). Show that Moo(K)
= U~l Mn(K)
~
is a K-algebra of infinite di-
mension. 4.10 Consider the following two matrices in Moo(K):
0 1 0 0 ... ) 0020... P = ( ~ 0 0 3 '. " :
and Q
=
(0 0 0 0 . . , ) 1 0 0 0 ... ~ 1 0 0 '. : : .
Show that the K-algebra of Moo(K) generated by P and Q is isomorphic to A1(K).
CHAPTER 2 IDEAL STRUCTURE OF THE WEYL ALGEBRA It is time to discuss the ideal structure of the Weyl algebra. We will introduce the degree of an operator and use it to show that the Weyl algebra is a domain whose only proper two-sided ideal is zero .
1. THE DEGREE OF AN OPERATOR. The degree of an operator of An, to be introduced in this section, behaves, in many ways, like the degree of a polynomial. The differences are accounted for by the noncommutativity of An. Let DEAn. The degree of D is the largest length of the multi-indices (a, (3) E N n
x N n for which x"'{jJ appears with non-zero coefficient in the
canonical form of D. It is denoted by deg(D). As with the degree of a polynomial, we use the convention that the zero polynomial has degree
An example will suffice: the degree of 2xl ~
+ xlx~8.J ~
is 6.
If D, D' E An are written in canonical form, then so is D
one concludes that deg(D
+ D') ::;
-00.
max{ deg(D) , deg(D')}.
+ D',
and
Note that if
deg(D) =f deg(D') then we have equality in the above formula. The formula deg(DD') = deg(D) + deg(D') also holds, but its proof is harder, because An is noncommutative. From now on we shall denote bye, the multi-index all of whose entries are zero, except the i-th entry, which is 1.
1.1 THEOREM . The degree satisfies the following properties; for D, D' E An:
(1) deg(D + D') ::; max{ deg(D), deg(D')}.
= deg(D) + deg(D') . deg[D, D'l ::; deg(D) + deg(D') -
(2) deg(DD') (3)
2.
PROOF: (1) has already been proved. We prove (2) and (3) at the same time, by induction on deg(D)
+ deg(D').
If either deg(D) or deg(D') is zero,
then the result is obvious. Suppose that deg(D), deg(D') ;:: 1 and that the formulae bold whenever deg(D)+deg(D') < k. Choose D, D' E An such that deg(D) + deg(D') = k . It follows from (1) that it is r.nollgh to prove (2) and
2. Ideal structure
15
(3) when D, D' are monomials. Suppose first that D
lal + 1131 = k.
= ff3
and D'
= x'" with
If 13; #- 0, then
By induction we have that
deg[&'-c"x"'] ~ and that deg[8;, x"']
~
lal -
lal + 1131- 3
1. Hence we may use the induction hypothesis
again to conclude that deg(8;[00-e"x"']) and deg([8 j ,x"']00-e,) are ~
1131- 2.
lal + 1131- 2.
Therefore deg[oo,x"'] ~
&'x'" Since deg(x"'{)f3) =
But
= [&',x"'] + x"'&'.
lal + 1m and deg[oo,x"'] deg(&'x"') = deg(x"'&')
Now let D
~
lal + 1m -
2, we conclude that
= lal + 1131·
= x"oo and D' = x"'8ry. If lal = 1131 = 0, the result is obvious.
Suppose that that is not the case. We have seen that ()f3x'" where P = [oo,x"'] has degree ~
lal + 1131- 2.
deg(DD/) = deg(xa+"'&,+ry)
+ deg(D /) -
2. Hence
= deg(D) + deg(D/). =
x u +"'ff3+'l + Q1> where deg(D) + deg(D') - 2. Similarly, we have that DID = x(1+'" 00+'1 +
The calculations above also show that D.D' ~
= x"'{)f3 + P,
Then
By the induction hypothesis deg(x U P8ry) ~ deg(D)
deg(Q1)
lal +
Q2, where deg(Q2) ~ deg(D) + deg(D ' ) - 2. Hence, [D, D/] = Q1 - Q2, and so deg[D, D/l ~ deg(D) + deg(D') - 2, which concludes the induction. As in the case of polynomial rings over a field, Theorem 1.1(2) may be used to prove the following result.
2. I deal structure
16
1.2 COROLLARY. The algebra An is a domain.
2. IDEAL STRUCTURE.
If you are used to commutative rings, you will find An very peculiar. Commutative rings have lots of two-sided ideals; not so An. A ring whose only proper ideal is zero is called simple. A commutative simple ring must be a field, but this is not true of noncommutative rings. The Weyl algebra is a simple ring, but it is very far from being even a division ring. 2.1 THEOREM. The algebra An is simple. PROOF: Let I be a non-zero two-sided ideal of An. Choose an element D =F 0 of smallest degree in I. If D has degree 0, it is a constant, and I = An. Assume that D has degree k
> 0 and let us aim at a contradiction.
Suppose that (0.,(3) is a multi-index of length k. If x"'f)f3 is a summand of D with non-zero coefficient and (3i =F 0, then [Xi, D) is non-zero and has degree k-1, by Theorem 1.1. Since I is a two-sided ideal of An, we have that
[Xi. D) E I. But this contradicts the minimality of D. Thus (3 Since k
> 0, we must have that
o.j
= (0, ... ,0).
=F 0, for some i = 1,2, ... , n. Hence
[8j , D] is a non-zero element of I of degree k - 1, and once again we have a contradiction. The kernel of an endomorphism of An is a two-sided ideal; thus we have the following corollary. 2.2 COROLLARY. Every endomorphism of' An is injective. Although An does not have any non-trivial two-sided ideals, it is not a division ring. In fact, the only elements of An that have an inverse are the constants. This is easy to see, as for polynomial rings, using the degree of an operator. If DEAn has an inverse, then there exists D' E An such that
DD'
= 1.
Calculating degrees, we obtain deg(D)
+ deg(D') = O.
Since the
degree of a non-zero operator is always non-negative, we must have deg(D}
=
O. Hence D is constant. Thus every non-constant operator generates a nontrivial left ideal of An. The Weyl algebra is not a left principal ideal ring either. For example, the left ideal generated by 0" {h ill A2 is not principal; sec Exorcise 4.1. However,
2. Ideal
17
struCtU1'e
every left ideal of An is generated by two elements. This is a very important result, due to J.T. Stafford. Unfortunately, its proof is very technical and beyond the scope of this book. It may be found in the original paper of Stafford [Stafford 78] and also in [Bjork 79, Ch. 1 ]. In fact, a similar result holds for left ideals over any ring of differential operators over a smooth variety, see [Coutinho and Holland 88].
3. POSITIVE CHARACTERISTIC. At the beginning of chapter 1 we made the hypothesis that the base field K over which the Weyl algebra is defined always has characteristic zero.
However, it is not immediately clear why one should make such a restriction. After all, the definition of the Weyl algebra makes perfect sense without any restriction on the characteristic of the field K. The problem lies deeper. First of all, the Weyl algebra in positive characteristic suffers from a split personality problem. This happens because the two ways by which we defined An in Ch. 1 give two non-isomorphic rings. Let us see what happens for the field Zp where p is a prime, in the one-variable case. Consider first the algebra of operators Rl generated by Zp[x J and its derivation B. Let us calculate BP(x k ). If k
< p,
then BP(x k )
since the coefficient is divisible by p. Thus BP
=
O. If k ~ p, then
= 0 as an operator on Zp[x].
We
conclude that the ring of operators Rl has nilpotent elements. In particular it is not a domain. Now consider the ring Rz generated over Zp by Zl, Zz subject to tIle relation
[Z2, Zl]
= 1.
This is a domain. However, it is not like the Weyl algebra in
characteristic zero in another respect: it is not a simple ring. For example, if I E Zp[ZlJ then [Z2.J]
= 81/ Bz 1. In particular,
zi commutes with every element of R2. Hence the zi in R2 is a two-sided ideal. In particular R2 is not
over Zp. It follows that left ideal generated by a simple ring,
2. Ideal structure
18
These superficial comments are enough to show that the positive characteristic case is very different from the characteristic zero case which is the subject of this book. The Weyl algebra in positive characteristic is studied in great detail in [Smith 86].
4 . EXERCISES.
4.1 Let n 2: 2. Show that the left ideal L of An generated by
8", ... ,8n is
not principal. Hint: Show that a generator would have degree 1, and obtain a contradiction from that. 4.2 Show that the left ideal of A3 generated by 81 , 82 , {h may also be generated by two elements. Hint: Choose D1 =
8" and D2 = ~+X1{h to be the generators, and calculate
(D.,D 2J. 4.3 Let K(X) be the rational function field in n variables. Let Bn(K) be the K-subalgebra of EndK(K(X)) generated by the elements of K(X) and the derivations
8", ... ,8n •
An element d E Bn(K) may be written in the form
d = La q,,8", where q" E K(X) . The order of d, denoted by ord(d), is the largest lal for which qa #- O. Show that if (/ E Bn(K) then: (1) ord(d + d') (2) ord(dd')
~
max{ord(d), ord(d')}.
= ord(d) + ord(d') .
4.4 Use the order defined above to show that:
(1) Bn(K) is a domain. (2) Bn(K) is a simple ring. 4.5 Show that B1 (K) admits a left division algorithm. That is, if a, b E
Bl (K), then there exist q, r E Bl (K) such that a = qb + rand ord( r) < ord(b). Use this to prove that every left ideal of B1(K) is principal. 4.6 If
n : : : 2 , is every left
ideal of IJ,, (K) principal '?
2. I deal structure
19
4.7 Let Al (2) be the 2-subalgebra of Al (Q) generated by x and 8. Show that Al (2) is not a simple ring. 4.8 Let dEAl (K), and write d L(d)
= gm(x) ::f 0 its
(1) The set Ln(J)
= L:;' gi(X )8;.
Call m the order of d and
leading term. If J is a left ideal of AI(K), show that:
= {L(d) : dE
J and d has order n} U {O} is an ideal of
K[x]. (2) If m S n, then Lm(J) ~ Ln(J).
(3) J is principal if and only if Ln(J)
= Ln+I(J)
for all n> 0 such that
Ln(J) ::f O. 4.9 Show that the left ideal AI8 2 + Al (x8 - 1) is not a cyclic ideal of Al (K). 4.10 Let 4J: A~
-+
Al be the map defined by 4J(a, b)
= a8+bx.
Show that 4J is
surjective and that its kernel is isomorphic to the left ideal Al & + Al (x8-1) . Conclude that this is a projective left ideal of AI(K). 4.11 Let H2 be the ring defined in §3. Show that:
(1) H2 is a domain. (2) The centre of H2 is 2p[zf. ~"J .
(3) H2 is finitely generated as a module over its centre.
CHAPTER 3 RINGS OF DIFFERENTIAL OPERATORS
In this chapter we show that the Weyl algebras are members of the family of rings of differential operators. These rings come up in many areas of mathematics: representation theory of Lie algebras, singularity theory and differential equations are some of them. 1. DEFINITIONS. Let R be a commutative K-algebra. The ring of differential operators of R is defined, inductively, as a subring of End](R). As in the case of the Weyl algebra, we will identify an element a E R with the operator of EndK(R) defined by the rule r
f---t
ar, for every r E R.
We now define, inductively, the order of an operator. An operator P E EndK(R) has order zero if (a, P]
defined operators of order
<
= 0, for every
a E R. Suppose we have
n. An operator P E EndK(R) has order n if
it does not have order less than n and [a, P] has order less than n for every a E R . Let Dn(R) denote the set of all operators of EndK (R) of order::::; n. It is easy to check, from the definitions, that Dn(R) is a K-vector space.
We may characterize the operators of order ::::; 1 in terms of well-known concepts. A derivation of the K-algebra R is a linear operator D of R which satisfies Leibniz's rule : D(ab)
= aD(b) + bD(a)
for every a, b E R.
Let Der]( R) denote the K-vector space of all derivations of R . Of course DerK(R) ~ EndK(R). If DE D erK(R) and a E R , we define a new deriva-
tion aD by (aD)(b)
= aD(b)
for every bE R. The vector space DerK(R)
is a left R-module for this action. Derivations are very important in commutative algebra and algebraic geometry as shown in [Matsumura 86] and [Hartshorne 77]. 1.1 LEMMA. The operators of order::::; 1 correspond to the elements of DerK(R)
+ R.
PROOF:
Let Q E DI(R) and put P = Q - Q(I). Note that P(I )
The elements of order zero are the elements of R .
= 0 and
that P hll..<; order::::; 1. Hence [P, a] hilS orner zero for every a E R. Thus for
3. Differential operators
21
every b E R, we have that [[P, a], b] = O. Writing the commutators explicitly, one obtains the equality
(Pa)b- (aP)b- b(Pa)
+ b(aP) = O.
= aP(b) + bP(a) baP(1). Since P(1) = 0, it follows that P is a derivation of R. But Q =
Applying this operator to 1 E R, we end up with P(ab)
P + Q(1)
E
DerK(R)
+ R, as
required. An easy calculation shows that if Q
has order zero then Q E R. The ring of differential operators D(R) of the K-algebra R is the set of all operators of EndK(R) of finite order, with the operations of sum and composition of operators. In other words, D(R) is the union of the K-vector spaces Dn(R) for n = 0, 1,2 . . . . For this definition to make sense, we must show that the sum and product of two operators of finite order has finite order. This is clear for the sum, but requires a proof for the multiplication. 1.2 PROPOSITION. Let P E Dn(R) and Q E Dm(R), then p. Q E Dn+m(R). PROOF:
The proof is by induction on m
+ n.
obvious. Suppose the result true whenever m
If m
+n =0
+n <
k. If m
the result is
+n =
k and
a E R, we have that
[PQ,aj = P[Q, a]
+ [P,a]Q.
The definition of order implies that [Q, aj E Dm-l(R) and [P, a] E Dn- l(R). Thus, by the induction hypothesis P[Q, a]. [P, a]Q E Dn+m-l. Hence [PQ, a] belongs to Dn+m-l, as required. We end this section with the explicit calculation of DerI( (K[XI, ... , x n]), which will be needed in the next section.
1.3 PROPOSITION. Every derivation of K[X]
= K[x), ... ,Xn]
is ofthe form
E~ /;8, , for some /J, ... ,Jn E K[X]. PROOF:
Let D E DerK(K[X]) . Then D(xt) = kx~-) D(Xi), for i = 1, ... , n.
Hence n
(D -
E D(Xi)8i)(xr' ... x~n) = O. I
3. Differential operators
22
Since these monomials form a basis for K[X] we have that D = 2:~ D(x;)B;. Thus D
= l:~ liB;
with J;
= D(x;).
2 . THE WEYL ALGEBRA.
Our aim now is to show that the Weyl algebra is the ring of differential operators of the algebra of polynomials. For the proof of this result we need two lemmas. 2.1
Let P
LEMMA.
E
D(K[X]). If [P,x;] = 0 for every i = 1, ... , n, then
PEK[X]. PROOF: We will show that under this hypothesis, [P,f] = 0 for every f E K[X]. Then the result follows from Lemma 1.1. Since the commutator is additive, it is enough to prove that [P, fl Let
f
= x
O
,
for some
0
E
= 0 when f
n
N and assume that
[P, XO ]
0;
=f O. Then
= [P,Xi]X o- e, + Xi[p, Xt>-e' j.
By induction on the degree of the monomials, [P, x;] [P,X O ]
is a monomial in K[X].
= [P, xo-e'j = O.
Thus
= 0, as required.
The next lemma is formally equivalent to the fact that every polynomial vector field F
= (F1, ... ,Fn )
in Rn which satisfies BFj/Bx;
BFi/Bxj for all 1 ~ i,j :$ n, has a potential. Define Cr to be the set of operators in An which can be written in the form l:o f oBt> with 101 ~ T. A simple calculation =
shows that
= Cr+! n Dr(K[X]). By Proposition 1.3, we have that C1 = DeTK(K[X]) + K[X] Cr
and that Co
=
K[X]. We will use the convention that if k < n then Nk is embedded in N n as the set of n.tuples whose last n - k components are zero . 2.2
LEMMA.
Let H, ... , P n
E
Cr-
1
and assume that [Pi , Xj 1= [Pj, x;] when-
ever 1 ~ i,j ~ n. Then there exists Q E Cr such that P;
= [Q, Xi],
for
i= 1, ... ,no
PROOF: Suppose, by induction, that we have determined Q' E Cr such that
[Q',x;] = Pi for k
+1
:$ i :$ n; thus ([Q',X;],Xk] = [Pk,X;]. Write G
[Q',xkj- Pk. Then [G,x;J
= 0,
for k
+ 1 :$
i :$ n.
=
9. Differential operators
23
It follows from the identity
that if [&6,x n ]
=
0 then
{3n
= O.
Thus [e,x n ]
=0
implies that e can be
written as a linear combination of monomials ofthe form xa&6 with{3 E N n Since
= 0 for k + 1 :c:;
[e, Xi]
1
•
i:C:; n, we may apply this result several times
and conclude that
where fa E K[X]. Now write
Q"
= L (Ok + It 1f a aa+ek • aENk
However, Q' E Gr
~
Dr{K[X]) implies that
Since Pk also belongs to Gr- 1 , then so does hand, [Q", Xi] But [Q",Xk]
e.
Hence Q" E Gr . On the other
= 0 for k + 1 :c:; i:C:; n by construction.
Thus [Q' - Q" , Xi]
= Pi .
= e, and so
Hence, [Q' - Q" , Xi]
= Pi, for
k :c:; i:C:; n; and the induction is complete.
2.3 THEOREM. The ring of differential operators of K[X] is An{K). Besides
this, Dk{K[X]) = Gk . PROOF: It is enough to prove that Dk(K[X]) ~ Gk . Let P E D(K[X]) .
If P E Dl(K[X]) then by Lemma 1.1, P E Der]((K[X])
+ K [X] .
p E G1 by Proposition 1.3. Suppose, by induction, that Dk(K[X])
for k :c:; m - 1. Let P E Dm(K [X]) . Write Pi k ::; m - 1, it follows that Pi E Gm -
I•
=
Thus
=
Gk
[P, Xi]. Since Pi has order
But, for all 1 ~ i, j :c:; n ,
3. Differential operators
24
Thus by Lemma 2.2 there exists Q E Om such that [Q,Xi] Hence [Q - P, Xi]
= 0 in D(K[X)).
~
1
~
i ~ n.
Since this holds whenever 1 ~ i ~ n, we
conclude by Lemma 2.1 that Q- P E K[XJ
Dm(K[X])
= Pi,
= 00,
Therefore P E Om. Hence
Om, as we wanted to prove.
The ring of differential operators D(R) of a commutative domain R is not always generated by Rand DerK(R)j see Exercise 3.8. However this is true if the ring R is regular; for example, if it is the coordinate ring of a non-singular irreducible affine variety. For a proof of this see [McConnell and Robson 87, Ch.I5]. The ring of Exercise 3.8 is the coordinate ring of a singular curve, the cusp. 3. EXERCISES .
3.1 Let J be a right ideal of a ring R. The idealizer of J in R is the set
H(J) = {a E R : aJ E J}. Show that H(J) is the largest subring of R that contains J as a two-sided ideal. 3.2 Let J be an ideal of S = K[Xll'" ,xnJ. Let BE EndKS. (1) Show that the formula O(f) = B(!) defines a K -endomorphism of S/ J if and only if B(J) ~ J. In that case, show that 0 = 0 if and only if
9(S)
~
J.
(2) Show that if BEAn ~ End](S and B(J) ~ J, then '8 E D(SjJ), the ring of differential operators of the quotient S / J. Show also that the order of '8 cannot exceed the order of B. (3) Show that if B is a derivation of S then '8 is a derivation of S/ J. 3.3 Let J be an ideal of S
= K[xj, . .. ,x n].
Let P E An. Show that:
(1) If peS) ~ J then P E JAn. (2) If P(J) ~ J then P E n(JAn) . 3.4 Let J be an ideal of S
= K[Xl . . .. ,xnJ .
differential operators of the quotient ring
S/ J.
Let D(S/J) be the ring of Show that there is an injective
ring homomorphism from H(J An)/ JAn into D(S/ J). Hint: There is a homomorphism from n(JAn) to D(S/J) which maps B to '8 , in the notation of Excrc.ise a.2. Now apply Exercise 3.3.
3. Differential opemtors
25
f
be the image of f E S in
3.5 Let J be an ideal of S
= K[XI, ... ,Xn].
Let
the quotient ring SjJ. Suppose that D is a derivation of SjJ and choose 9j E S such that 9; = D(xj), for 1 $ i $ n.
(1) Show that if B = 2:~ 9j 8j, then
13 =
D, in the notation of Exercise
3.2. (2) Let Der.T(S) be the set of derivations DE Der(S) such that D(J) J.
~
Conclude, using Exercise 3.2, that there is an isomorphism of
vector spaces between Der.;(S)jJDer.l(S) and Der(SjJ). 3.6 Let R
= K[ £l, ts).
(1) Show that R is isomorphic to K[x, V)jJ, where J is the ideal of K[x, V) generated by (2) Let DI
=
'Il -
x3 .
+ 3x 28y and
2V8x
D2
=
3V8y - 2x8x. Use the previous exercise to show that the set of derivatives of K[x, yJj J is generated by DI and D2 as a module over K[x, VJ/J .
(3) Conclude that DerK(R) is an R-module generated by where 8 t
tat
and {l8 t ,
= djdt.
3.7 Let R be a commutative domain with field of fractions Q. Show that the set {P E D(Q) : peR)
~
R} is a subring of D(R) .
3.8 Let R = K[£l, ts]. Let 8
= djdt and
let BI be the K-algebra generated
by K(t) and 8; see Exercise 2.4.3. (1) Using Exercise 3.7, show that the following elements of BI belong to
D(R): 8 2 - 2t- 18,
ta2 -
8 and
aJ - 3t- 182 + 3t- 2 8.
(2) Show that these elements do not belong to the subring of BI generated by R and its derivations. (3) Conclude that D(R) is not generated by Rand DerK(R).
CHAPTER 4
JACOBIAN CONJECTURE The Jacobian conjecture was proposed by O.H. Keller in 1939. It asks whether a polynomial endomorphism of
enwhose Jacobian is constant must
be invertible. Despite its simple and reasonable statement, the conjecture has not been proved even in the two dimensional case. In this chapter we show that this conjecture would follow if one could prove that every endomorphism of the Weyl algebra is an automorphism. The chapter opens with a discussion of polynomial maps, which will play a central role in the second part of the book. We shall return to the Jacobian conjecture in Ch. 19. 1. POLYNOMIAL MAPS .
Let F : Kn
-t
K m be a map and p a point of Kn. We say that F
is polynomial if there exist Fl • . . . ,Fm E Klxl , " " xn] such that F(p)
=
(FI(p)" ", Fm(P» , A polynomial map is called an isomorphism or a polynomial isomorphism if it has an inverse which is also a polynomiall11ap. It is not always the case that a bijective polynomial map has an inverse which is also polynomial. For an example where this does not occur see Exercise 5.1. However, if K = C, every invertible polynomial map has a polynomial inverse. This is proved in [Bass, Connell and Wright 82; Theorem 2.1]. For the rest of the section we shall write X, Y for the spaces K n and Ktn; and K[X], K[l1 for the polynomial rings K[XI,'" ,x n] and K[Yl , ... • YmJ. A polynomial 9 E K[ 11 may be identified with the function of Y into
K which maps p E Y to g(p). Clearly if 9 = 0 as a polynomial, then it induces the zero function on Y. Since K is a field of characteristic zero, the converse is also true: a polynomial which induces the zero function on Y is identically zero. For a proof see Exercise 5.2. This identification is the key to the construction that follows. Suppose that F : X - t Y is a polynomial map. We may define a map , F~
; K [Y]
-t
K[X] ,
4. Jacobian coniecture by the formula FI(g)
= g. F,
27
where 9 E K[Y] . Note that the arrow gets
reversed as we go from F to Fl. The map F~ is called the comorphism of F. A routine calculation shows that FI is a homomorphism of polynomial rings. Let us calculate an example. Suppose that n
< m and that
F ;X
~
Y is
the map
Then the algebra homomorphism F~ ; K[Y] ~ K[X] maps a polynomial
9(1/1, ... , Ym) to 9(X}, . .. ,X n, 0, . .. ,0). We may turn the construction of the above paragraph inside out. Suppose that a ring homomorphism ¢; K[Y]
~
K[X) is given. Then we may use it
to construct a polynomial map from X to Y. Let Vi be an indeterminate in
K[n Then ¢(Vi) is a polynomial of K[X). Now let rPr. ; X
~
Ybe the map
whose coordinate functions are ¢(Vl), ... , ¢(Vm). These two constructions are each other's inverse, as the next results show.
1.1
THEOREM.
Let F : X ~ Y be a polynomial map, then (FI)~
Furthermore, if Z = Kr and G: Y
~
G· F; X ~ Z is a polynomial map and (G· F)I PROOF:
=
F.
Z is another polynomial map, then
= FI . Q1I.
Let Vi be an indeterminate in K[Y]. As a function Y
~
K we have
that Vi maps a point of Yonto its i-th coordinate. Hence ,
Ftt(Vi)
Vi' F = Fi·
=
Thus, the coordinate functions of (FI)I are Fl • ... ,Fm; which are the coordinate functions of F. Hence (Ftt)~
= F.
It is clear that G·F is a polynomial function. Let 9 E K[Z] Then
(G· F)~(9)
= K[ZI," " zr].
= 9' (G· F) .
Since the composition of maps is associative,
(G· F)tt(g)
= (g. G)· F = F~(ctt(g»,
as required. The converse of Theorem 1.1 is also true. We state it and leave the proof to the reader.
4. Jacobian conjecture
28 1.2 THEOREM.
then (~)~
= ifJ.
IfifJ: K[l1-+ K[X] is a homomorphism of polynomial rings,
Furthermore, if'IjJ: K[Z]
K[l1 is another homomorphism,
-+
then
The following result is an immediate consequence of Theorems 1.1 and 1.2.
1.3
COROLLARY.
A polynomial map F : X
-+
Y is an isomorphism if and
only if FI is an isomorphism. The Jacobian conjecture, discussed in the next section, proposes a simple criterion to determine whether a polynomial map is an isomorphism. 2.
Let F : K"
-+
JACOBIAN CONJECTURE
K" be a polynomial map. Denote by J(F) its jacobian
matrix. This is the matrix whose iJ' entry is 8F;/8xj. If F is an isomorphism, then it follows from the chain rule that J(F) is an invertible matrix at every point of K". In particular the determinant ,1F
= detJ(F)
is an invertible
polynomial, hence a constant. Now suppose that K
=
IR or C. If p E K" and ,1F(p) '" 0, then by the
inverse function theorem there exists a neighbourhood U of p such that F restricted to U is invertible. This leads to the following question. If,1F is non-zero everywhere on K", is there a function G : Kn
-+
Kn which is
the inverse of F? Two comments are in order. First, although F always has an inverse in the neighbourhood of every point (by the inverse function theorem) it is not clear whether these inverses can be 'glued ' together to produce an inverse in the whole of K". Secondly, even if the inverse exists, it may not be a polynomial map, as shown in Exercise 5.1. The Jacobian conjecture, first stated in [Keller 39J, is a refinement of the above question.
2.1
JACOBIAN CONJECTURE.
Let F : Kn
-+
Kn be a polynomial map. If
,1F = 1 on Kn then F has a polynomial inverse on the whole of Kn. Let us see what happens if n
= 1. In this case, we have a map
F : K -+ K
which is determined by a polynomial F(x) in one variable. The jacobian rm1trix is the derivative dFj d:c, and we a rc assuming t hat this is a constant.
4. Jacobian coniecture
29
Hence F is a linear map, and consequently it has an inverse. This proves that the Jacobian conjecture holds for n
=
1. The fact that an invertible
polynomial is linear is restricted to the one-dimensional case. For n
~
2,
an invertible polynomial map may have coordinate functions of any degree whatsoever; as shown by the examples of Exercise 5.3. Despite many attempts to settle it, the Jacobian conjecture remains open for every n
~
2. A number of results related to the Jacobian conjecture
have accumulated over the years; for a very nice survey of some of these results see [Bass, Connell and Wright 82]. For example, it is known that the conjecture is false over fields of positive characteristic, see Exercise 5.4. On the positive side, if all the coordinate functions of F have degree :5 2, then the conjecture is true; see Exercise 5.5. We now rephrase the Jacobian conjecture using comorphisms. Let X Kn and K[X]
= K[Xl, ... ,xn]'
=
The Jacobian Conjecture states that if flF =
1 on X then F~ is an invertible homomorphism of rings. Actually it is not
hard to see that if flF is constant, then FI is necessarily injective. We prove this in a little more generality. 2.2 LEMMA. Let F : X
o everywhere in X. PROOF:
----t
X be a polynomial map and suppose that flF::f
Then FI is injective.
Suppose that F~ is not injective, and choose the non-constant poly- .
nomial 9 E K[X] of smallest degree such that F~(g) Let gj
=
O. Then g(F)
=
0.
= og/OXj and
Hence, by the chain rule,
v(p). JF(p)
=
°
for every p EX. Since flF(p) = detJ F(p) ::f 0,
we conclude that yep)
=
0 for every p EX. Thus 9j(Fl, ... ,Fn )
=
0 for
1 :5 i :5 n. Since 9 is not constant, at least one of the 9j must be non-zero. But 9, has dpgree smaller than 9, a contradiction.
30
4. Jacobian conjecture
Denote by K[F!, . .. , Fn] the sub algebra of K[X] generated by the coordinate functions of F. This is the image of the homomorphism F~. If.tlF = 1 then F~ is injective by Lemma 2.2. Now assume that K[Fb ... , Fn]
= K[X]i
then FI is also surjective. Hence by Corollary 1.3, F itself is bijective. Thus the Jacobian conjecture may be rephrased as follows. 2.3 JACOBIAN CONJECTURE. Let F : Kn
Kn be a polynomial map and assume that.tlF = 1 in Kn. Then K[Fl , ..• ,Fn] = K[Xb ... ,xn]. --+
In §4 we show that if every endomorphism of the Weyl algebra is an automorphism then the Jacobian conjecture holds. This follows an idea of L. Vaserstein and V. Katz; see [Bass, Connell and Wright 82, p.297]. The next section is a digression on results about derivations that will be required in §4.
3. DERIVATIONS In this section we study some properties of derivations that will be required in the next section. Let D be a derivation of a K-algebra S. It follows from Leibniz's rule that the kernel of D is a sub ring of S, it is called the ring of
constants of D. The derivation D is locally nilpotent if given a E S, there exists kEN suc.h that Dk(a)
= O.
Note that the derivations 8 1 , ... ,8n are
locally nilpotent in K[Xl, ... xn], whilst x l 8 l is not. Let S be a ring and D a locally nilpotent derivation. Define a map ifJ :
S ---t SIx] by the rule ifJ(a) =
Lo 00
Dn( a )
- I - xn
n.
for every a E S. Note that ifJ(a) belongs to SIx] because D is locally nilpotent. It is easy to check that ifJ is a ring homomorphism which satisfies
We want to prove the following proposition from [Wright 81J.
4. Jacobian conjectu.re
31
PROPOSITION 3.1. Let S be a K -algebra and D 1 , ... , Dn be commuting
locally nilpotent derivations of S . Suppose that there exist tl" ' " tn E S such that Di(t;)
= oij .
Then
(1) S = R[t1 , • •• , tn], where R is the ring of constants with respect to Dl, .. . , Dn,
(2) tt, .. . , tn are algebraically independent over R , (3) D;
= a/at; for i = 1, ... , n.
The proposition is proved by induction. It is better to isolate the case
n = 1 in a lemma. LEMMA
3.2. Let S be a K-algebra and D a locally nilpotent derivation of
S . Suppose that for some t E S one has D( t)
= 1.
Then
(1) S = R[ 4, where R is the ring of constants of S, (2) t is algebraically independent over R , (3) D=d/dt. PROOF: Put S
= S/ St.
Let p : S --+ S[x] be the composition of ifJ defined
above and the projection S[x] --+ S[x]. We want to show that p is an
= x.
isomorphism. Note that p( t)
To prove that p is surjective it is enough to prove that its image contains
S. Let a E S .
Denote by ii its image in
exists n EN such that Dk(a)
S.
Since D is locally nilpotent, there '
= 0 for k > n.
Thus,
> 0 put ao = a and define = 1, ... ,n. It is easy to show, by
If n = 0, then pea) = a. If n
aj+l = OJ -
Dn- j(Oj)t!'-j/(n-j)!, for j
induction on
j, that Dk(aj)
Thus p(a n ) surj ective.
= 0 for k > n- j
= an'
and that
However, since 1 = 0, we have that p(a n )
= a.
Thus p is
4. Jacobian conjecture
32
Let us prove that p is injective. If not, then there exists a non-zero a E S such that pea) for some
al
= O.
Thus Dk(a) E tS, for every kEN. Hence a
E S. Since p(t)
= x, we have that p(al) = O. Thus
al
= al . t, E tS and
a = a2· f-, for some a2 E S. Continuing this way we conclude that t n divides ~ o. But this is impossible, unless a = O. Indeed,
a for all n
in the polynomial ring S[x]. Hence, if a¥-O we have that deg(
0, which is clearly impossible. Thus a
~
n for
= 0, as required.
We conclude that the homomorphism p : S ~ S[x) is an isomorphism. Since p . D
= dldx. p, we have that R = p-I(S). The result now follows if
we recall that p (t) = x. PROOF OF PROPOSITION 3.1: We proceed by induction on the number n of derivations. By Lemma 3.2, S = RdtJ], where RI is the ring of constants of D I . But tl is algebraically independent over RI and DI
=
dl dtl. Since
DI commutes with Di for i> 1, we have that Di(R I ) ~ R I . Thus, by the induction hypothesis, RI = R[~, ... t,..], and the proposition follows.
4. AUTOMORPHISMS We now return to the setup of the Jacobian conjecture. Let X
= Kn.
The
rational function field of K[X] will be denoted by K(X). Let F : X -) X be a polynomial map with coordinate functions Fh ... , Fn. Assume that
L1
= L1F ¥- 0
everywhere on X; a condition that is actually weaker than the one required by the Jacobian conjecture. Define a map D i : K(X) -) K(X) by
It is easy to check that Di is a K-linear map that satisfies Leibniz's rule. Hence Di is a derivation of K(X). Now let K[X,L1-1] be the K-subalgebra of K(X) of all rational functions whose denominator is a power of L1. Then Di restricts to a derivation of K[X, L1-I], since
4. Jacobian conjectu1'e
33
4.1 LEMMA. As derivations of K[X,L1- 1J the Di satisfy:
(1) D;(Fj)
= Cij .
(2) The Di commute pai.rwise.
PROOF: We prove (2); since (1) follows easily from properties of the determinant. Note first that ..1(0)
=1=
O. Thus ..1 is invertible as a power series
and K[X, L1- J ~ K[!XlJ. On the other hand, ..1 . Di is a derivation of 1
K[Xl, ... ,xnJ which can be extended to a derivation on the power series ring K[[XlJ
= K[[Xb '"
,xnlJi see Exercise 5.9. Since ..1 is invertible as a power series, then Di can also be extended to a derivation of K[[X]]. Put B
=
[Di' D j ]. We want to show that B = 0 on K[X, ..1- 1]. It is
enough to show that B
= 0 on the power series ring K[[XII . Since the
commutator of two derivations is a derivation (see Exercise 5.8), we have that B is a derivation of K[[XlJ. Moreover B(Fk)
= 0, for
B is zero in the subalgebra K[F}, . .. ,Fn). But F h
...
1 ::; k ::; ni and so
, Fn are algebraically
independent, by Lemma 2.2. Hence we may consider B as a derivation on the power series ring K[[F} , ... , Fnll. By (1), B is zero on K[[F}, ... , Fnll. For 1 ::;
i::;
n let ai
= Fj(O).
The jacobian matrices of (Fl - al,' .. ,Fn - an)
and F coincide. Since the latter is invertible in K[[x}, ... ,xnJ]' we conclude from the local inversion theorem (see Appendix 2) that
Thus B is zero on K[[x}, .. . ,xnll, as required. We now return to the Jacobian conjecture. The next theorem is of the type: a conjecture implies a conjecture! To simplify the proof we introduce the following notation. Let a E An. The map ada: An ada(b)
-t
An is defined by
= [a, bJ.
This is a K-linear map, but it is not a K-algebra homomorphism. 4 .2 THEOREM. Let F : Kn
-t
Kn be a polynomial map and assume that
L1F = 1 everywllere on Kn . If every endomolphism of An is an automorphism, tl/en tlJC .Jacobian conjecture llOlds.
4. Jacobian coniecture
34 PROOF: Since
ilF = 1, it follows from Lemma 4.1 that Dl, ... , Dn are
derivations of K[X] which satisfy
[D i, Fj]
Di(F;)
=
=
8ij and [D i , Dj ] = 0,
for 1 :::; i,i :::; n. By Ch.1 §3, there exists an endomorphism ifJ: An such that ¢;(Xi) = Fi and ¢;(8i) = D i , for 1 :::;
i:::; n. Note that for
Thus given bEAm there exists kEN such that (ada.?(b) ¢;(ada,(b»
~
An
bEAn,
= O. Since
= adD,¢;(b).
we have that (adDy(¢;(b» = O. Assuming that ifJ is an automorphism, we conclude that Di is locally nilpotent. It then follows by Proposition 3.1 that
K[F1, ... , Fn]
= K[X1, ' .. ,x,,], which is the Jacobian conjecture as stated in
2.3. Once again let us observe that it is not known whether every endomorphism of An is an automorphism. This conjecture first appeared in print as 'Probleme 11.1' in [Dixmier 68]. Note, however, that every endomorphism of An is injective, by Corollary 2.2.2. Thus to prove the conjecture it is enough to show that every endomorphism of An is surjective. Unfortunately this is not known even for
11,
=
1.
5. 5.1 Let F : R.2 ~
]R2
EXERCISES.
be the polynomial map defined by F(x, y)
Show that ilF ~ 1 in
]R2,
= (x3 +x, y).
but that it is not constant. Show that F has an
inverse but that it is not polynomial. 5.2 Let 9 E K[X]
= K[X1, ... ,x n]. Show that g(p) = 0 for every p E Kn if
and only if 9 = 0 as a polynomial. Hint: Proceed by induction on n. Write 9 in the form 'E~ giX~, where gi E K[X1" .. ,X n-1]. Suppose that there exists q E Kn-1 such that 9i(q) for some i. Then k
g(q,x n ) = L9j(Q)X:. o
¥- 0
4. Jacobian conjecture
35
is a polynomial in one variable with infinitely many roots. 5.3 Let d be a positive integer. Consider the polynomial map F : 1R.2 defined by F(x, y) everywhere in
]R2
=
«x - y)d
+y-
-+ ]R2
2x, (x - y)d - x). Show that 11F
=1
and that F has a polynomial inverse.
5.4 Let k be a field of characteristic p > O. Consider the map F: k n
-+
kn
defined by
Show that J(F) is the identity matrix but that F cannot be invertible. Hint:
Xl
+ xi is not an irreducible polynomial.
5.5 Let F :
en -+ en be a polynomial map.
Suppose that the coordinate
functions of F have degree at most 2. Show that if 11F f- 0 everywhere in
en, then F(p) - F(q)
p+q
= JF(-2-)(P- q).
Use this to prove that F must be injective. This is enough to prove the Jacobian conjecture for quadratic maps because by [Bass, Connell and Wright 82, Theorem 2.1J an injective polynomial map of en to itself must be bijective.
5.6 Which of the following derivations are locally nilpotent in K[xI, X2]?
(1) XI~
(2) Xl~
+ x2lh + lh
5.7 Check in detail that the Di defined in §4 are derivations of K(X). 5.8 Let R be a commutative ring and D, D' be derivations of R. Show that
[D, D'] is a derivation of R. 5.9 Let D be a derivation of K[XI,"" x n ] that is zero on the constants. Show that: (1) D can be extended to the power series ring K [[x I , . .. , xnlJ . (2) If 11 is a power series such that 11(0) f- 0 then 11- 1 . D is a derivation of the power series ring K[[xJ, ... ,:I;n]].
CHAPTER 5 MODULES OVER THE WEYL ALGEBRA This chapter collects a number of important examples of modules over the Weyl algebra. The prototype of all the examples we discuss here is the polynomial ring in n variables; and with it we shall begin. The reader is expected to be familiar with the basic notions of module theory, as explained in [Cohn 84, Ch.1O]. 1. THE POLYNOMIAL RING.
In Ch. 1, the Weyl algebra was constructed as a subring of an endomorphism ring. Writing K[X] for the polynomial ring K [Xl, ...
,X n ]
we have
that An(K) is a subring of EndKK[X] . One deduces from this that the polynomial ring is a left An-module. Thus the action of
Xi
on K[X] is by
straightforward multiplication; whilst 8i acts by differentiation with respect to
Xi.
This is a very important example, and we shall study it in some detail.
Let us first recall some basic definitions. Let R be a ring. An R-module is irreducible, or simple, if it has no proper submodules. Let M be a left R-module. An element
'U
E
M is a torsion element if amlR('U) is a non-
zero left ideal. If every element of M is torsion, then M is called a torsion
mod'Ule. 1.1 LEMMA. Let R be a ring and M an irreducible left R-module.
(1) Ifa
i u EM, then M
~
R/annR(u).
(2) If R is not a division ring, then M is a torsion module. PROOF:
Consider the map rp : R ---; M defined by rp(l) =
homomorphism of R- modules.
Since
'U
(1) that M
~
=
It is a
i a and M is irreducible, rp is
surjective. Thus (1) follows from the fact that ker tjJ Now suppose that annR( 'U)
'U.
0 for some
ai
= annR( 'U).
U
E
M. It follows from
R. Since M is irreducible, this can happen only if the left
ideals of R are trivial. But in this case R is a division ring, contradicting the hypothesis. Thus annRu # 0, which proves (2). Let us apply these results to the An-mod ule K[X).
37
5. Modules
1.2 PROPOSITION. K[X] is an irreducible, torsion An-module. Besides this, n
K[X] ~ Ani LAna;. 1
PROOF: First of all 1 is clearly a generator of K[X]. Now suppose that
f =!: 0
is a polynomial and consider the submodule An . f. Let x~ ... x~ be
a monomial of maximal possible degree among the monomials that appear in f with non-zero coefficients. Let a b~ its coefficient. Thus
otl ... o~· .f
il !. " in! . a is a non-zero constant in the submodule generated by An ' f
= K[X] . Thus K[X]
=
I . Hence
is irreducible. Since An is not a division ring, it
follows from Lemma 1.1(1) that K[X] is a torsion module. Now 1 is a non-zero element of K[X] which is annihilated by 81 , . . • , an. Hence the left ideal J generated by 01, . .. ,an is contained in annAn 1. Con-
f + Q, where implies that f = O.
versely, let P E annA.(I). Then P may be written in the form
Q E J and f Therefore P
E
= f . 1, which that J = annAn(l).
K[X] . Thus 0 = p. 1
= QE
J . We conclude
The isomorphism
now follows from Lemma 1.1(1). We may generalize this example as follows. Choose 91, . . " 9n E K[X] and consider the left ideal J of An generated by 81 - 91> ... , 8n - 9n. Every element of An is of the form
f + P,
for
f
E K[X] and
P E Jj see Exercise '
4.1. Hence the map IjJ : Ani J ---; K[X] defined by 1jJ(f
+ J) = I
is an
isomorphism of K-vector spaces. Although the action of the x's is preserved under this isomorphism, it is not an isomorphism of An-modules. Indeed, if
f
is a polynomial, then
0;(1 Thus,
8f + J) = -a + I . a;+ J Xi
=
of -a +f Xi
. 9; + J .
ljJ(odl +J)) = (Oi+9;)' I , where the right hand side is to be calculated
in K[X] with its natural action. The module Ani J is irreducible; the proof is similar to that of Proposition 1.2. Another module that is closely related to K [X ] is Ani vector space it is isomorphic to K[o]
= K[oll ' . . , 8n ],
I:~
A n' x ;. As a K-
the set of polynomials
in 01 , . '" 8 n . Using t his isomorphism, we may identify the action of An
5. Modules
38
directly on K(a]: the 8's act by multiplication, whilst
Xi
acting on 8j gives
-Oij·1. Apart from the obvious similarities, the modules K[8] and K[X] are
related in a deeper way that will be explained in the next section. The emphasis on irreducible modules in this chapter is justified by the fact that they play the role of building blocks in module theory. However, it is not true that every An-module is a direct sum of irreducible modules. The sense in which we say that (interesting) An-modules are built up frpm irreducible modules is explained in Ch.l0. 2. TWISTING. We begin with a general construction. Let R be a ring and M a left Rmodule. Suppose that u is an automorphism of R. We shall define a new left module M .. , as follows. As an abelian group, M" = M. The difference lies in the action of Ron M". Let a E Rand u EM, define a. u = u(a)u. A routine calculation shows that M .. is a left R-module. It is called the twisted
module of M by u. Not surprisingly, M .. inherits many of the properties of M. 2.1 PROPOSITION. Let R be a ring, M a left R-modu1e and u an automorphism of R. Then: (1) M" is irreducible ifand only jf Mis D.'reducib1e. (2) M" is a torsion module if and only if M is a torsion module. (3) If N is a submodu1e of M then (MIN)" ~ M"IN... (4) Let J be a left ideal of R. Set u(J) = {u(r) : r E J}. Then u(J) is a left ideal of An and (RIJ)" ~ R/u-I(J). PROOF: An R-module M is irreducible if and only if given any non-zero
u, v
E
M there exists a
as u-1(a) • u
o becomes
=
E
R such that au
=
v. This equation translates
v in M", which proves (1). Similarly, the equation au
u-1(a) •
U
=
OJ which proves (2).
=
Now (3) is an immediate
application of the first homomorphism theorem. To prove (4), note that u(J) is a left ideal, since u is an automorphism of An. Let
.p(I)
.p: R
= 1 + J.
~
(RIJ)" be the homomorphism of R-modules defined by
If b E R, then ¢(Ii) = /). ¢(l) = (1'(Ii) +.1.
5. Mod1J.les
39
Hence ¢i is surjective, and its kernel is q-l(J). Thus (4) also follows from the first homomorphism theorem. Let us apply this construction to An. An important example is the Fourier transform. Let F be the automorphism of An defined by F(x;) =
F(Oi)
= -Xi;
a; and
see Exercise 1.4.8. Let M be a left An-module. The twisted
module MF is called the Fourier transform of M. The reason for the name is clear, F transforms a differential operator with constant coefficients into a polynomial. 2.2 PROPOSITION. The Fourier transform of K[X] is K [o] . PROOF: It follows from Proposition 1.2 that K[X] L:~
~
Ani J, where J
=
An ' OJ. Since F-l(J) = L:~ An ' Xj we may apply Proposition 2.1(4) to
get the desired result. It follows from Propositions 2.1(1) and 2.2 that K[o] is irreducible. The
Fourier transform will reappear later on in this book. Let us consider other examples of twisting. For i = 1, ... ,n, let 9j E K[xil. Note that [OJ + 9i, OJ
gjl
= 0, for 1 ~
i,j ~ n. Hence there exists an automorphism
maps Xj to itself and
{)j
q
+
of An which
to OJ + 9j. A straightforward calculation shows that n
K!X)11 ~ AnIZ)An(oi - 9i)),
which is a particular case of the example considered in §l. Note that K[X]11 is irreducible, by Proposition 1.2 and Proposition 2.1(1). In fact K[X)" is irreducible for every automorphism
q
of An .
We may use the above construction to produce an infinite family of nonisomorphic irreducible left An-modules. For every positive integer r let the automorphism of An which satisfies a-.(xi)
= X;
and ur(O;) =
qr
be
a; - xi.
2.3 THEOREM. The modules K[X] 11, form an infinite family of pairwise non-
isomorphic irreducible modules over An. PROOF: Let r
<
t, and suppose that there exists an isomorphism, ¢i :
K[X ]I1' ~ K[X]a, . Since K[X]a, is irreducible, it is gellerated by 1. Thus ¢i is completely determined by the image of 1; say (PCl) = f "" O. Now the
40
5. Module3
equation ¢I(8;. 1)
= 8;. ¢I(I)
translates as the differential equation
8J = 8Xi
(x; - xDJ.
The left hand side of the equation has degree :::; deg J - 1. Since
r
< t,
+ t.
the right hand side has degree degJ
f ¥- 0 and
This is a contradiction, so
the theorem is proved. The construction of irreducible An-modules leads to many important questions and we shall return to it several times. 3. HOLOMORPHIC FUNCTIONS. Let U be an open subset of
= d/dz .
Then
z acts by multiplication, whilst 8 acts by differentiation on H( U). It is routine to check that the actions are well-defined and satisfy the required properties; see Appendix 1. As an Al (C)-module, H( U) is not irreducible. This is the same as saying that there are holomorphic functions which are not polynomials. On the other hand, a torsion element of H( U) is a holomorphic function which satisfies an ordinary differential equation with polynomial coefficients. These elements exist. For example, exp(z) is a solution of d/ dz1. However, H( U) is not a torsion module. This is a more interesting result that we now prove. In fact we show that the function exp(exp(z)) does not satisfy a polynomial differential equation. First a technical lemma. 3.1 LEMMA. Let h(z) be the hoiomol'pllic function exp(exp(z)). For every positive integer m there exists a polynomial Fm(x) E Clx] of degree m such
that
dmh/dz"" = Fm(eZ)h(z). PROOF: Since dh/dz
= e'· h(z), the result
holds for m= I with F1(z) = z.
Suppose, by induction, that the formula holds for m. Using the formula
dm+1h/dzm+l = d(dmh/dzm)/dz and the induction hypothesis, we have that
5. Modules
41
where F:" is the derivative of the polynomial Fm. Putting
we have the desired formula. Besides this, the degree of Fm+! is deg Fm + 1 = m + 1, by the induction hypothesis. In the proof of the next proposition we use the fact that the exponential function is not algebraic. A complex function! E 1i( U) is algebraic if there exists a non-zero polynomial G(x, y) E IC[x, y] such that G(z, fez)) all z E U. For a proof that
e
= 0,
for
is not algebraic, see Hardy's delightful book
The integration of functions of a single variable, (Hardy 28, eh.V, §16]. 3.2 PROPOSITION. The function h(z)
= exp(exp(z)) is not a tOl'sion element
of the Al (C)-module 1i(U). PROOF: Let P
= L.~!.(z)8i be an element of AI(C)'
We assume that!r ~ O.
Then P·h = 0 in 1i( U) is equivalent, by definition, to the differential equation
dih
r
'E,.!i(Z) dzi o Using Lemma 3.1 and the fact that h(z)
= O.
= exp(exp(z)) ~ 0 for every z
E C,
we obtain an equation of the form r
'E,.!i(z)Fi(e') = O. o This is equivalent to G(z, e)
= 0,
where G(x, y)
= L~ !;(x)F;(y)
is a poly-
nomial in two variables. Since Fi(Y) is a polynomial in y of degree i, the polynomial G(x, y) must be non-zero. This implies that tf is an algebraic function; a contradiction. In the next chapter, 1i( U) will play the role of module of solutions for differential equations represented in 'V-module language. 4. EXERCISES.
4.1 Let 91, ... ,gn be polynomials in K[X] . (1) Show, by induction on m, that Bin ma~ be written in the form D(B;9,)
+ f , where
DEAn aIld
f
E K[X].
5. Modules
42
(2) Conclude that every element of An can be put in the form Q where Q E L:~ An(8; - gi) and
+ f,
J E K[X] .
4.2 Show that if 9i E K[x,] then the module Ani L:~(An(8; - gi)) is irreducible. 4.3 A left R-module M is cyclic if M
= R· u,
for some u E M. Show that
an irreducible module is always cyclic. 4.4 Let R be a simple ring that is not a division ring. Suppose that M, M' are non-zero left torsion R-modules. Show that if M is cyclic and M' is irreducible, then M EEl M' is cyclic.
"# a E R such that au = O. Since i= O. Then u+ v generates M EEl M'.
Hint: Let u be a generator of M . Choose 0
R is simple aM'
i= O.
Let v E M' with av
4.5 Using Exercise 4.4, show that the direct sum of any finite number of irreducible An-modules is cyclic. 4.6 Let M, M' be left R-modules and let u be an automorphism of R . Show that (M EEl M')" ~ M" EEl M~. 4.7 Show that 'H.( U) is not a cyclic Al (C)-module. Hint: Let h E 'H.(U) be a generator. Show that h cannot be constant and that exp(h) .;. AICC)' h . 4.8 Let U be an open set of IR n and COO(U) be the real vector space of all functions of class Coo defined on U. Let
Xi
act by multiplication and 8; by
differentiation on COO( U). Show that this makes COO ( U) into a left An(lR)module. Is COO( U) an irreducible A n (IR)-module? Is it a torsion module? Is it cyclic? 4.9 Are sinCe') and cos(eZ ) torsion elements of the Al (C)-module 'H.(C)? 4.10 Are exp(sinz) and exp(cosz) torsion elements of the AI(C)-modulp. 'H.(C)? 4.11 Show that cos z and sin z are torsion elements of 'H.(C).
43
5. Modules 4.12 Let U be the open set C \ (-00,0]. If Q E JR, show that element of the At (C)-module 11.( U) .
ZO
is a torsion
CHAPTER 6
DIFFERENTIAL EQUATIONS We are now ready to justify the statement that the theory of D-modules offers an algebraic approach to linear differential equations. We begin by describing a system of differential equations and its polynomial solutions in module theoretic language. This leads to more general kinds of solutions: distributions, hyper functions and microfunctions. We end with a description of the module of micro functions in dimension 1.
1. THE D-MODULE OF AN EQUATION.
Let P be an operator in An. It may be represented in the form where a E N" and 9", E K[X l , ... ,x,,]
=
2:" 9",8"',
K[X]. This differential operator
gives rise to the equation
P(f) =
L 9",8",(1) = 0, ",
f may be a polynomial or, if K
JR, a Coo function on the variables x), ... , X n. More generally, if PI> . . . , Pm are differential operators in An, then we have a system of differential equations
where
=
(1.1) In this section we want to associate to this system a finitely generated left An-module in a canonical way. This will allow us to give a purely algebraic description of the polynomial solutions of (1.1). The An-module associated to the system of differential equations (1.1) is
Ani 2:;n AnPi . This definition is justified by the next theorem. A polynomial solution of (1.1) is a polynomial i
= 1, ... , m.
space.
f E K[X] which satisfies Pi(J) = 0, for
The set of all polynomial solutions of (1.1) forms a K-vector
45
6. Differential equations
1.2
THEOREM.
Let M be the An-module associated with the system (1.1).
The vector space of polynomial solutions of the system (1.1) is isomorphic to HomA.(M, K[X)). PROOF:
Let j
E
K[X] be a polynomial solution of (1.1).
Consider the
homomorphism of : An ~ K[X] which maps 1 E An to j. If Q E J
=
2:~ AnP;, then
Q(J)
= O.
Hence uf defines a homomorphism Uf : M
-----t
K[X].
Thus to a polynomial solution j of (1.1) we associate uf E HomA.(M, K[X)). Furthermore, the map j
;----+
uf is a linear transformation. Indeed, if j, f} E
K[X] and A E K, we have that
Of+>.g(P + J) = P(J
+ Af}) =
(uf
+ AUg)(P)
by the linearity of the operator P. The inverse map may be explicitly defined. If 7' E HomA.(M, K[X]), then it maps the class of 1 in M to a polynomial h. An easy calculation shows that h is a solution of (1.1) and that 7'
= Uh.
Hence the rule 7'
;----+
7'(1)'
defines the inverse linear transformation. Be warned that HomA.(M, K[X]) is neither an An- module nor even a K[X]-module; it is only a K-vector space. Worse still: it may be a vector space of infinite dimension; see Exercise 4.1. Although we restricted ourselves to polynomial solutions of the system (1.1) this is not really necessary. Suppose that H, ... , Pm E An(1R). The set of COO functions defined on an open set U of Rn, denoted by COO( U), is a left An(R)-module; see Exercise 5.4.8. Proceeding as above one shows that the COO-solutions of (1.1) correspond to homomorphisms in HomA.(M,COO(U). These examples inspire us to make the following definition. Let S be a
.
left An-module; and let M be a finitely generated left An-module. We will call HOUlA..(M, S) the solution space of M in S. Note that we have taken
6. Differential equations
46
care not to require that S be finitely generated. For example,
coo (lRn) is not
a finitely generated An (IR)-module; see Exercise 4.4. On the other hand, a system of differential equations will always correspond to a finitely generated module. The module associated to (1.1) is even cyclic. The advantage of a definition of this sort is that it allows us to introduce generalized solutions of differential equations in a natural way. All one has to do is choose an appropriate An-module S. This includes solutions in terms of distributions, hyperfunctions and microfunctions. These generalized solutions may be necessary, as some differential equations do not have solutions in terms of ordinary functions. Here is an example in one variable. Consider the operator x in Al (C). The differential equation xf non-zero solution even if we require only that
= 0 does
not have a
f be continuous. However, this
equation has a solution in terms of distributions, the famous Dirac 6-function. In the next sections we construct the module of micro functions in one variable in an algebraic way and express the Dirac 6 as a microfunction. It is then easy to check that x.6 = 0 . 2.
DIRECT LIMIT OF MODULES.
The construction of the module of microfunctions makes use of direct limits. This has independent interest and we begin by discussing it in some detail. Let 'I be a set with a relation $ . We say that 'I is pre-ordered if $ is reflexive and transitive. A pre-ordered set 'I is directed if, given i,j E 'I, there exists k E 'I such that k $ i and k $ j. Directed sets will play the role of index sets for our construction. Let R be a ring and 'I be a directed set. Suppose that to every i E I we associate a left R-module Mi. We say that {Mi : i E 'I} is a directed
family of left R-modules if given any i, j E 'I, satisfying i $ j, there exists a homomorphism of R-modules 71'j; :
and if i
~
Mj
---t
Mi ,
j ~ k, then 71'j i ' 7rkj
=
7rki·
47
6. Differential equations
An example will make things clearer. Let D(e) be the open disk of centre
o and
radius e in C. Let H(e) = H(D(e» be the set of all holomorphic
functions defined in D(e). Recall that H(e) is a left AI-module; see Ch. 5, §3. Take IR to be the index set. If e ~ € in IR, then H(€)
~
H(e). Hence we
may take
'Tre. : H(e')
---!
H(e)
to be the restriction of a holomorphic function in H(€) to D(e). This gives us a directed family of AI-modules. We return to the general construction. Let {Mi : i E I} be a directed family of left R-modules. Denote by U the disjoint union of the modules Mi; it may be identified with the set of all pairs (u, i), where u E Mi. We define an equivalence relation in U as follows: (u, i), (v,j) E U are equivalent if and only if there exists k E I such that k ~ i, k ~ j and
'Trik( u)
= 'Trjk( v).
The
direct limit of the family {Mi : i E I}, denoted by limMi' is the quotient ~
set of U by this equivalence relation. To simplify the notation, we will write (u, i) for the equivalence class in limM; as well as for its representative in U. ~
Let us apply this construction to the family {H(E) : e E IR}. Things are made a little simpler in this case because IR is completely ordered. An element of lim H(e) is represented by a pair (j, e) where ~
f
E H(e). Assuming that
e ~ E', we have that two pairs (j,e) and (g,€) are equal in ~H(e) when
g(z) = fez), for every z
-
E D(e). The elements of Ho = lim H(e) are called
germs of holomorphic functions at O.
We have not yet finished the construction, because we want to turn the direct limit into a module. Once again we return to the general situation.
-
Let (u, i), (v,j) E limM; and a E R. Choose k E I such that k ~ i and k :S: j. The sum (u, t)
+ (v, j) is the element
The product a' (1.£, i) is (au, i). We must show that these operations are independent of the various representatives of classes used to define them. Since most of this is routine, \
we check only that the sum is independent of the choice of k.
Suppose
48
6. Differential equations
that k, k' are both less than or equal to i and j . We want to show that S
= (lrik(U) + lrjk(v),k)
rEI, such that r
~
equals S' = (lrik'(U)
k and r
~
+ lrjk'(v),k') in limMi. Choose ~
k'. Thus
and both equallrir( u) + 7rjr( v). Hence S = S' = (lr;r( u) + 7rjr( v), r) in lim M;, ~
as required. The sum and product by scalar in lim M; are defined using the corre~
sponding operations in Mi. Thus the usual properties of the sum and scalar product in a module hold for lim Mi ' The details are left to the reader. ~
In the example lim H(f), the index set R is totally ordered. Thus (1, E) ~
+
(g, i) is (f + g, E") where i' = mini E, i}. The scalar product is P . (f, E) = (P(f),E) , for every P E AI(C) ' 3. MICROFUNCTIONS. The module of microfunctions is constructed as a direct limit. We must first define the family of AI-modules, the direct limit of which we will be calculating. As in §2, let D(E) be the open disk of C of centre 0 and radius E.
Let D'(E)
= D(E) \ O.
C : R e(z) < 109(E)}. = eZ . Note that 11' is surjective: if p < E is a positive real number, then pe iO = 7r(log(p) + if)) , and log(P) + if) E bee). The relative positions of these sets and maps are shown The universal cove1' of D'(E) is the set beE)
The projection
11'
= {z
E
of bee) on D'(E) is defined by 7r(z)
in the following diagram.
We make the set H (b(e)) of functions that are holomorphic in bee) into an Al (C)-module. Let hE H(b(e)) . The action of a polynomial on h is given by j. h formula 8 • h
= j(e")h(z).
= hi (z )e- z .
actions are well-defined .
The operator 8
f
E C[x)
= d/dx acts on h by the
A calculation (left to the reader) shows that these
6. Differential equations
49
3.1 PROPOSITION. The map
defined by 1I"*(h)(z)
h( 1I"(z» is an injective homomorphism of Al (C)-
modules. PROOF:
Since 11" is surjective, 11"* must be injective. Let us check that 11"* is a
homomorphism of Al (C)-modules. Suppose that hE 1t(D'(e». If f E then 1I"*(fh) = (fh)(e!) and
f. h(e!)
Clxl,
= f(e!)h(e Z ) are equal. Similarly,
is equal to 11"* (8· h). The proof that 11"* preserves the sum is an easy exercise Since D'(e) ~ D(e), we have that 1t(D(e)) is a submodule of 1t(D'(e»). Let M. denote the quotient module 1t(D(f)/1I"*(1t(D(e»)). If t ~
f,
then
D(t) ~ D(e). Thus 1t(D(e) ~ 1t(D(t». This induces a homomorphism of Al (C)-modules
Hence {M. : e E R} is a directed family of AJ(C)-modules, and we may take its direct limit. This limit, denoted by M, is called the module of. microfunctions. The canonical projection of1t(D(e» onto M. is compatible with the limit, and determines a homomorphism of AI-modules
can: 1t(D(e»
-+
M.
The function e- Z /27ri is holomorphic in D(e). But e- Z /27ri is also the image of 1/27riz under 11"*. However, 1/27riz is not holomorphic in D(e). Hence
can( e- Z /211"i) is a non-zero element of M, it is called the Dirac delta microfunction, and denoted by 8. As observed at the end of §1, the equation xh = 0 has no analytic (or COO) solution, but it is satisfied by the Dirac delta. This is easily checked. Note first that x8 = can(eZe- Z/211"i)!;= can(1/211"i).
6. Differential equations
50
But 1/27ri, being a constant, is holomorphic in D(e), hence zero in M . Thus
x6 = 0, as required. Another important example is the Heaviside microfunction, defined by Y
=
can(z/27ri}. Note that z/27ri is the image of log(z)/211'i under 7r".
Since 10g(z)/211'i is not holomorphic in D(€), the hyper function Y is nonzero. Moreover, Y is the integral of 0;
Microfunctions can be used to classify certain Armodules in terms of quiv-
ers, a combinatorial object. This is discussed in detail in [Brian<;on and Maisonobe 84] and [Malgrange 91]. 4. EXERCISES. 4.1 Show that the K-vector space of polynomial solutions in K[Xl,X2J of
xIB:! - x28t has infinite dimension. 4.2 Show that the set of polynomial solutions in Clx] of a differential operator
of Al (C) has finite dimension.
4.3 Let U be an open set of Rn. Using Baire's theorem and the fact that
COO ( U) is a metrizable vector space show that any basis of COO( U) as a real vector space is uncountable. Hint: Suppose the basis generated by
Vh.'"
{Vi; i E N}
is countable. Let li be the subspace
v,.. Show that \{ is a closed set of Coo (U) with empty
interior and that Uk 'Ii
= Coo(U).
Obtain a contradiction using Baire.
A proof that COO( U) is a metrizable vector space is found in [Rudin 91,
eh.lJ. 4.4 Show that COO( U) is not finitely generated as an An(R)-module. Hint; Suppose that it is generated by ft .... ,h, and show that xa[jJ. fi forms a countable set of generators for Coo (U) as a real vector space, for ex, (3 E Nn and 1
~
i ~ k. This contradicts Exercise 4.3.
Exercises 4.5 to 4.7 explain the const1'Uction of the module of hyperJunctions. This is ve?/I simila?' to the construction 0/ the micro/unctions presented in §3.
6. Differential equations
51
4.5 Let [} be an open interval of JR. A complex neighbourhood of
Q is
an
open set U of C that contains Q. Consider the set
u = {U: U is a
complex neighbourhood of a}.
Show that U is a directed set for the order ;2. 4.6 Let [} be an open interval of IR and U a complex neighbourhood of [}. (1) Show that ?t(U)
~
?t(U\Q) and that both are Al (lR)-modules, where
a polynomial acts by multiplication and 8 by differentiation. (2) Show that 11.( U\ Q) /?t( U) is a directed family of Al (R)-modules with respect to the directed set U. (3) Let B([})
= lim11.(U\ [})/?t(U) ---t
and hE ?tCU\ [}), for some U E U.
Denote by [h] the image of h in B(Q). Show that if h can be extended to a holomorphic function on U then [h]
= O.
(4) Using the notation of the previous item, show that if f E R[x] then
f . [h]
= [fh] and 8· [h] = [hi].
B(Q) is called the mod1.l.le of. hype7j1.l.nctions of [}. 4.7 Let
Q
= (0,1). The Heaviside hyperfunction is Y= [log(-z)/27T"z] and .
the Dirac hyperfunction 0 = [l/27T"iz). Show that 8· Y =
o.
4.8 Show that the submodule of the module of microfunctions M generated by 0 is isomorphic to C[a]. 4.9 Let 0' be the first derivative of the Dirac microfunction. Let Al (C)o' be the submodule of M generated by 0'. Show that
(1) A I (C)6'
= Al(C)6.
(2) Al(C)O' ~ AI(C)/J, where J is the left ideal of AI(C) generated by
x 2 and x8+2. Hint: If Q E AI(C) satisfies QO' annihilator of o. Write Q = Q2x2 C[8], and calculate q. 8.
=
0 then we have Q.
+ QIX + Qo, where Q2
aE
AI(C)X, the
E Al (C), Qo, QI E
52
6. Differential equations
4.10 Let 8m be the m-th derivative of the Dirac microfunction 8. Show that Al (C)8 rn is isomorphic to Al (C)/ J, where J is the left ideal of Al (C) generated by xm and x8 + m. Hint: Induction and Exercise 4.9.
, ./
CHAPTER 7 GRADED AND FILTERED MODULES Simple rings are very hard to study because most techniques in ring theory depend on the existence of two-sided ideals. In the case of the Weyl algebra, however, we have a way out. As we saw in Ch. 2, one may define a degree for the elements of the Weyl algebra. Using this degree, we construct a commutative ring, which works as a shadow of An. We may then draw an outline of what An really looks like . This is the best method we have for understanding the structure of An and of its modules. 1.
GRADED RINGS
An important feature of a polynomial ring is that it admits a degree function. We want to generalize and formalize what it means for an algebra to have a degree. This leads to the definition of graded rings. These rings find their justification in algebraic geometry, more precisely in projective algebraic geometry; for details see IHartshorne 77 , Ch. 1, §2] . For the sake of completeness, we define graded rings without assuming conunutativity. Let R be a K-algebra. We say that R is graded if there are K-vector . subspaces R i , i E N, such that
(1) R
= ffiiEN R i ,
(2) R i · Rj ~ Ri+j ' The Ri are called the homogeneous components of R , The elements of R ; are the homogeneous elements of degree i. If Ri
= 0 when i < 0 then we
say that the grading is positive. From now on all graded algebras will have a positive grading unless explicitly stated otherwise. The most important example of a graded algebra is the polynomial ring
K[xI, .. . ,xnJ. The monomials X~1
. • • x~·
with kl
+ ... + k n =
m form a
basis of the homogeneous component of degree m. Nonconunutative graded rings have been in evidence recently in the theory of quantum groups and noncommutative geometry. The quantum plane, for example, corresponds to the K-algebra generated by two elements x , y which satisfy the relation
54
7. Graded and filtered modules
xy
= >.yx,
for some>. E K; see Exercise 6.1. For more details see [Manin
88) .
The most important graded rings that come up in algebraic geometry are quotients of polynomial rings. They are constructed as follows. Let R be a graded K -algebra. EBi~(I n
A two-sided ideal I of R is a graded ideal if I
=
R;) . Thus a graded ideal is generated by homogeneous elements.
The converse also holds: an ideal generated by homogeneous elements must be graded. Now let S
= EBi?:O Si
be another graded K-algebra. A homomorphism of
K-algebras ¢> : R --. S is graded if ¢>(R i )
~
Sj. Thus a graded homomor-
phism is one that preserves the degree. The concepts of graded homomorphism and graded ideal are related, as shown in the next result. 1.1 PROPOSITION.
Let R = EBi?:O R; and S = EBi?:O Si be graded algebras
over K.
(1) The kernel of a graded homomol'phism of K-algebl'as ¢> : R --. S is a graded two-sided ideal of R. (2) If I is a graded two-sided ideal of R then R/ I is a graded K -algebra. PROOF:
Suppose first that ¢> is a graded homomorphism. Let a
= ao€B ' . ·€Ba.
be an element of the kernel of ¢>. Thus,
¢>(a)
= .p(ao) + .,. + .p(as ).
Since ¢> is graded, this sum is direct. Thus ai E ker(¢» for i
= 0, ... ,s.
This
proves (1). Conversely, let I
=
ker(¢» be a graded two-sided ideal. Then we may
decompose the quotient ring R/I into a direct sum of K-vector spaces, R/J ~ E9(R;/(I n Ri»' i?:O
If ai E Ri and Uj E Rj then (ai
+ I)(aj + I)
= a,aj
+I
corresponds to an
element of Ri+jj(I n R i+j ) under this isomorphism. Hence RjI is a graded ring, wllich proves (2).
55
7. Graded and filtered modules
This gives us a way to generate examples of graded rings. Let Ft, . . . , Fk be homogeneous polynomials in K[X]. The quotient K[X]/(Fl,'" ,Fk) is a graded ring. A graded algebra admits a special kind of module. Let R
= Eai2:0 Ri be
a graded K-algebra. A left R-module M is a graded module if there exist
K-vector spaces M i , for
(1) M
i;:::
0, such that
= Eai2:0 M i ,
(2) R i · Mj !; Mi+j . The Mi are the homogeneous components of degree i of M. Note that the definition of graded module depends on the graded structure chosen for the algebra R. We may also define graded submodules and graded module homomorphisms, mimicking the corresponding definitions for graded rings . Let R be a graded K-algebra and M, M' be graded left R-modules. A submodule N of M is a graded sub module if N
morphism B ; M
--+
= E9i2:0(N n Mi).
An R-module homo-
M' is graded if B(Mi) ~ M: . It follows that ker(B) is a
graded submodule and that the quotient module MIN is a graded R-module j the proof is similar to that of Proposition 1.1. This allows us to produce examples of graded modules. Let R
=
ffi i2:0
Ri
n
be a graded K-algebra and R be the free left R-module of rank n. This . module has a natural grading, its k-th homogeneous component is the vector space
L
(Ril ffi .. . ffi Ri.)'
it+·+i.=k
If L is a graded submodule of R" , then Rn I L is a finitely generated graded left module. The importance of graded rings is explained in the next section. 2. FILTERED RINGS.
In eh. 2 we introduced the degree of an operator in the Weyl algebra. However, this degree cannot be used to make An into a graded ring. The problem is that an element like lhxl ought to be homogeneous of degree 2, but it is equal to x\81 + 1, which is not homogeneous. To use this degree effectively we must generalize graded dugs, to get filtered rings.
7. Graded and filtered modules
56
Let R be a K-algebra. A family :F
=
{F;}i~O
of K-vector spaces is a
filtration of R if
(1) Fo
~
Fl
~
F2
~
...
~
R,
(2) R = U,~O Fi, (3) Fi · Fj ~ Fi+j' If an algebra has a filtration it is called a filtered algebra. It is convenient
to use the convention that Fj
= {OJ if j < O.
Let us show that every graded algebra is filtered . Suppose that G
EB,~O G; is a graded algebra. Consider the vector spaces Fk
Fk
~
= EB~ Gi .
=
Clearly
Fk+l and their union is the whole of G. Since Fk.Fm =
E9
GiGj
,
i+j9+m and GiGj ~ G i+j , we have that FkFm ~ Fk+m' Hence {Fkh~o is a filtration of G. On the other hand there are filtered algebras which do not have a natural grading. This happens to the Weyl algebra; which, however, has many different filtrations. The first filtration of An which we will discuss is the Bernstein filtration. It is the filtration defined using the degree of operators in An . Denote by
Bk the set of all operators of An of degree $ k. These are vector subspaces of An. Conditions (1) and (2) of a filtration are clearly satisfied by the B k , whilst (3) is a consequence of Theorem 2.1.1(2). Thus B = {BkhEN is a filtration of An. We also write B(An) and Bk(An) whenever necessary. The Bernstein filtration has a very special feature: each Bk is a vector space of finite dimension. A basis for Bk is determined by the monomiaL., :z;Ot{/J with
101 + 1.81
$ k. In particular, Bo
=K
and {I , Xl , · · . x n , 8 1 , . . . ,8n }
is a basis of B I . The fact that Bk has finite dimension will be fundamental in Ch. 9. Another important example of a filtration for An is the o1'der filtration, denoted by C. As in Ch. 3, §2, denote by Ok the vector space of all operators of order $ k in An. It is clear that properties (1) and (2) of a filtration hold for C, and (3) is a consequence of Proposition 3.1.2. Note that 0 0
= K[X]
is an infinite dimensional K-vector spacc. Despite this drawback, the order
7. Gmded and jilte1'ed modules
57
filtration has the advantage that, unlike the Bernstein filtration, it is welldefined for other rings of differential operators.
3. ASSOCIATED GRADED ALGEBRA. We may use a filtration of an algebra to construct a graded algebra. This is very useful because many properties of this graded algebra pass on to its parent filtered algebra, as we shall see in later chapters. Let R be a Kalgebra. Suppose that :F = {Fi}iEN is a filtration of R. As a first step in the construction of the graded algebra, we introduce the symbol map of order
k, which is the canonical projection of vector spaces
Thus for an operator d E F k, the symbol O'k(d) is non-zero if and only if d¢Fk-l '
Consider now the K -vector space
g?R
= EB(Fi/Fi - I ) . i~O
We want to make it into a graded ring. For that it is enough to define the multiplication of two homogeneous elements, and extend it by linearity. A homogeneous element of g?R is of the form O'k(a) for some a E Fk . Let O'm(b), for bE Fm be another homogeneous element , and define their product by
A straightforward verification shows that g? R with this multiplication is a graded K-algebra, with homogeneous components Fi! F i -
I.
This is called
the graded algebra of R associated with the jiltmtion :F. The definition of the multiplication in g? R hides some surprises. The best way to illustrate these is with an example, which turns out to be the most important applicat ion of this construction to be used in this book. Put Sn
= !J',-8 An.
7. Graded and filtered modules
58
3.1 THEOREM. The graded algebra Sn is isomorphic to the polynomial ring over K in 2n variables. PROOF: For i
= 1, . .. , n, let
Yi
= O"I(Xi)
and Yi+n
= 0"1 (8i ).
We break the
proof into several steps. First step: Sn is generated by YI, ... , 1/2n as a K -algebra. It is enough to prove this for the homogeneous elements of Sn. But a
homogeneous element of Sn is of the form O"k(d), for some element d of An of degree k. Now d is a linear combination of monomials xo.Ef3, with If lal
+ IfJl = k,
lal+lfJl
$ k.
then
Thus O",.(d) is a linear combination of monomials in YI , .. . , 1/2n of degree k, as we wanted to prove. Second step: Sn is a commutative ring. Since Sn is generated by YI, . .. , 1/2n, we need only show that these elements commute in Sn. For i 0"2(8iXi) . Since OiXi =
= 1, .. . , n, we have that Xi8;
+ 1, one has
YiYi+n
= 0"2(X;Oi) and Yi+nY; =
that
Thus YiYi+ n = Yi+nYi· It is even easier to check that Yi commutes with Vj when j ¥ i + n, since the corresponding elements in An commute. Let K[ZI , . . . ,Z2n] be the ring of polynomials in 2n variables. The previous two steps allow us to define a surjective ring homomorphism
by tP(z;}
= Vi·
Since the z's have degree 1 in K[ZI, . .. , Z2n] and the V's have
degree 1 in Sn, tP is a graded homomorphism of K-algebras. Third step:
tP is injective.
Let F E K[ZI, . . . , Z2n] , and suppose that tP(F)
= O.
Since tP is a graded
homomorphism, we may assume that F is a homogeneous polynomial. Let
7. Graded and filtered modules where
0'1
+ '" + an + f31 + ... + f3n = k.
59
Define an operator d of An by the
formula
Then O'k(d)
= F(Yl, ... , '112,,).
= 4J(F) = 0, then d E Bk - 1 • Hence d may be written as linear combination of monomials xoVO with 10'1 + 1f31 < k. By construction, d is also If O'k(d)
a linear combination of monomials of degree k. Hence, by Proposition 1.2.1, all the coefficients Ca/3 above are zero. Thus F is the polynomial zero and 4J is injective, as required. One may ask, at this point, what the algebra !J"c An associated to the filtration by order looks like. The answer is that it too is isomorphic to a polynomial ring in 2n variables, see Exercise 6.5. However, for the rest of the book we will deal almost exclusively with the Bernstein filtration. The advantages of the Bernstein filtration over C will only be apparent in Ch. 9. Till then most of the results that will be proved for the Bernstein filtration also hold for the filtration by order. Since the proofs are always very similar, we will leave the latter as exercises for the reader.
4. FILTERED MODULES.
To define a filtered module we must start with a filtered ring. For the sake of simplicity we shall give the definitions only for the Weyl algebra with the Bernstein filtration. Let M be a left An-module. A family r
= {ri};~O
of K-vector spaces of
M is a filtration of M if it satisfies
(1) To ~
n ~ ... ~ M,
(2) Ui~O r j = M, (3) BJj ~ ri+j. Although this is the standard definition of a module filtration, we shall require that the filtrations used in this book satisfy an additional condition. First note that (3), with i = 0, implies that each r; is a K-vector space. The fourth condition that a filtration must sahsfy is:
60
7. Graded and filtered modules (4) F. is a K-vector space of finite dimension.
The convention that
rj = {OJ if j < 0 remains in force.
Of course B is a filtration of An as an An-module. A more interesting example is the An-module K[X]. The vector spaces degree
~ i
r. of all polynomials of
form a filtration of K[X] for the Bernstein filtration B.
Following the pattern of §3, we may define the graded module associated with a filtered module. Let M be a left An-module M and let
r be a filtration
of M with respect to B. Define the symbol map 01 order k of the filtration
r
to be the canonical projection
Now put
grF M
= E9(r;jri - 1 ). ,~o
We will define an action of Sn on this vector space. If a E Fk and u E Fi let
Extending this formula by linearity we get an action of Sn on g~ M . We leave it to the reader to check that this action satisfies the required properties. The graded Sn-module g~ M is called the g1'aded module associated to the
filtration
r.
Let us return to a previous example. Let F be the filtration of K[X] with respect to B defined above. Then
ri / ri- 1 is isomorphic to the vector space of
all homogeneous polynomials of degree i. Hence g~ M is isomorphic to K[X ] as a vector space. However we saw in Theorem 3.1 that Sn is isomorphic to the polynomial ring in 2n variables Yl, ... , 'Y2n. We want to determine the action of the y's on a homogeneous polynomial be thought of as an element of Fr/FTYi . 1 =
xd·
For i
= n + 1, ... , 2n we
I.
For i
1
of 1egree r, which is to
=
1,\...,n,
is, by definition J.'r( 8,(1». But 8,(1) is homogeneous of degree Yi
·1 = O.
we have that
must be more careful. Note that Yi· ~
1
r-l. Hence
In particular anns. (g~ M) is the ideal generated by Yn+l, . .. , 'Y2n· 5. INDUCED FILTRATIONS.
Let M be a left An-module with a filtration
r with re!lpect to B.
Suppose
7. Gmded and filtered modules that N is a submodule of M. We may use
r
61
to construct filtrations for both
Nand MIN . These are called the filtrations induced by
r.
To get a filtration for N put r' = {N n ri};~O. The inclusion N S; M allows us to define injective linear maps:
These may be put together to produce a linear map:
a homomorphism of modules. Since the
r" be the subspace of MIN
defined by
A routine calculation shows that r" is a filtration of MIN. Note that
and let 7l'k :
HI H-l ---; r[ I rf-l
be the canonical projection. Putting these together we get a K-linear map,
One easily checks that this map is a surjective homomorphism of Sn-modules. 5.1 LEMMA. Let M be an An-module witl] a filtration
r
compatible with
B. The sequence of Sn-modules
o-+ gr-r' N
4>
-+
.. grr" MIN -+ 0 grr M -+
is exact. PROOF:
Note that the kernel of the map
7l'k
defined above is
62
7. Graded and filtered modules
Thus we have an exact sequence of vector spaces
The sequence of 8 n -modules in the lemma is obtained by adding these vector space sequences for k
~
O. Hence it is exact.
The exact sequence of Lemma 5.1 is very useful. Here is a typical application. Let d be an operator in An of degree r and put M
= AniAnd.
Take l3
to be the filtration of An as a left An-module. The induced filtration in And is l3~
= Bk_rd.
Thus,
Since Bkl B k - 1 is the homogeneous component of degree k of 8 m then
By Lemma 5.1, there is an exact sequence,
G.
= K{x, y}
6.1 Let F
EXERCISES
be the free algebra in two geperators, and I the two-
sided ideal of F generated by the relation xy - )..yt. Show that the quotient ring FI I is a graded ring. 6.2 Let R
= ei2:0 Ri be a graded ring and M
Let M(k);
I = ei2:0 M; a graded R-module.
= Mi-k.
(1) Show that M(k) = EBi2:0 M(k); is a graded R-module. (2) Show that if k > 0 then the identity map M --+ M(k) is an isomorphism of R-modulcs, but. not a graded isomorphism.
7. Graded and filtered modules
63
6.3 In this exercise we define a grading for Al that is not positive. Denote by x and 0 the generators of AI. Define G" = {d E Al : [x8,
dJ = kd}.
Let
K[xo] be the polynomial ring in the operator xo. Show that (1) G"
= K[xo]x" for k ~ 0,
(2) G" = K[x8]0-" for k ~ 0, (3) Al
= IDiEZ G" is a graded ring.
6.4 Let dEAn. Define the principal symbol of d by IT(d)
= 1T,,(d),
where k
is the degree of d and IT" denotes the symbol map of degree k relative to the
Bernstein filtration. Find the principal symbol of the following operators of Aa:
+ x~; x~8i + x~xgol~ + X2X48l + xf; oJ + x~ + xtx~ + 8:2x~ + X2'
(1) otx~
(2) (3)
6.5 Let 'Tk be the symbol map of order k with respect to the order filtration of An. Let
ei =
71 (OJ). Show that g.f An is isomorphic to the polynomial
ring K[X][~1>'" , en]. 6.6 Let R be a filtered K -algebra with a filtration F. Show that if g.,J R is . a domain, then so is R. 6.7 Let J be a left ideal of An. Define the symbol ideal of J to be the ideal gr(J)
= Lk~OIT,,(JnBk) of Sn. Let
M
=
AnjJ.
(1) Let 8' be the filtration of J induced by the Bernstein filtration. Show that g."B/ J
S;:
gr(J).
(2) Show that if 8" is the filtration of M induced by the Bernstein filtration, then g."B"M ~ Snlgr(J).
6.8 The definition of a filtered An-module for the 01'de1' filtration is analogous to that given in §4, except that (4) must be replaced by: M j is a finitely generated K[Xl-module, for i 2: O. Find a filtration of the An-module K[X] with respect to t.he order filt.ration and c:aklllate its associated graded module.
64
7. Gmded and filte1'ed modules
6.9 Let Vbe the K-vector subspace of An with basis {Xl, ... , X n , (h, ... , an}. Let Sp( V) be the symplectic group on V. (1) Show that an element u E Sp{ V) can be extended to an automorphism of An that preserves the Bernstein filtration. (2) Show that this automorphism induces an automorphism of Sn
=
g"J3 An. 6.10 Let (,\b'" ,'\2n) E K2n. Show that the formulae
U(Xi) = Xi - '\; and
U( 0;)
= Oi -
,\n+i
define an automorphism of An(K) that preserves the Bernstein filtration and induces the identity automorphism on Sn. 6.11 Let d E An(C) be an element of degree 2. Show that there exist an automorphism
Ot
of An(C) and constants '\1 •. .. , '\n, TI, . •.
, Tn
such that
n
DI(d)
=L
'\i(X~
+ a;) + Ti·
I
Hint; Use Exercises 6.9 and 6.10, and the fact that quadratic forms are diagonalizable.
CHAPTER 8 NOETHERIAN RINGS AND MODULES There is very little that one can say about a general ring and its modules. In practice an interesting structure theory will result either if the ring has a topology (which is compatible with its operations), or if it has finite dimension, or some generalization thereof. As an example of the former, we have the theory of C" -algebras. The latter class includes many important rings: algebras that are finite dimensional over a field, PI rings, artinian rings and noetherian rings. It is the last ones that we now study. In particular, we prove that the Weyl algebra is a noetherian ring.
1. NOETHERIAN MODULES.
In this book we shall be concerned almost exclusively with finitely generated modules. One easily checks that a homomorphic image of a finitely generated module is finitely generated. However a finitely generated module can have a submodule that is not itself finitely generated. An example is the polynomial ring in infinitely many variables K[xJ , X2,
...
J. Taken as a module
over itself this ring is a cyclic left module: it is generated by 1. However, the ideal generated by all the variables
Xl, X2, ...
every finite set of polynomials in K [Xl, X2,
•••
cannot be finitely generated:
J uses up only finitely many of
the variables. We get around this problem with a definition. A left R-module is called noetherian if all its sub modules are finitely generated. Examples are easy
to come by: vector spaces over K are noetherian K-modules. Every ideal of the polynomial ring in one variable K[xJ is a noetherian K[xJ-module. There are several equivalent ways to define noetherianness. We chose the most natural. Here are two more. l.1 THEOREM . Let M be a left R-module. The following conditions are
equivalent:
(1) M is nueth()j-jaIJ.
8. Noetherian modules
66
(2) For every infinite ascending chain NI there exists k
~
0 such that N;
= N,.
~
N2
~
. . . of SIlbmodules of M
for every i
~
k.
(3) Every set S of submodules of M has a submodule L which is not properly contained in any submodule of S.
Condition (2) is known as the ascending chain condition. This is probably the most common way in which noetherian modules are defined. Condition (3) is the maximal condition: the submodule L is called a maximal element of S. PROOF: Suppose that (1) holds. If NI ~ N2 ~ ... is an infinite ascending
chain of submodules of M, then Q =
Uj~l
N; is a submodule of M. Thus
Q is generated by finitely many elements, say UI,' ..• 'Ut. Hence there exists UI,· . .. • Ut E N k · Thus Q = Nk = N i , for every i ~ k, as
k ~ 0 such that
claimed in (2). Assume now that (2) holds. We prove (3) by reaching a contradiction. Suppose that S does not contain a maximal element. If NI
~
N2
~
... t:::; N,.
is any chain of elements of S, then we can make it longer. Since no element of S is maximal, there must exist Nk+l in S such that Nk
c
NHI . In this
way we construct a proper infinite ascending chain of submodules of M, thus contradicting (2). Finally, assume (3) and let N be a submodule of M . Let S be the set of all finitely generated submodules of N, and let L be a maximal element of S. Suppose LeN, and choose
U
E
N\L. Thus L+ Ru is finitely generated and
contains L properly: a contradiction. Hence L
= N,
is finitely generated.
In the next proposition we collect the basic properties of noetherian modules that are used in later sections. First, a technical lemma. Let M be a left module over a ring R. Let N, PI and P2 be submodules of M such that P2 t:::; Pl. IfN+PI = N+P2 and NnPI = NnP2 ,
1.2
LEMMA.
then PI = P2. PROOF: We need only show that PI t:::; P2 • Suppose that
Ul
E Pl' Since
67
8. Noetherian modules we have that
Ul
= x + 'lIo2, for x
In particular, x E P2 . Hence
Ul
E N and ~ E
P2. Thus
= x + ~ E P2 , as required.
1.3 PROPOSITION. Let M be a left module over a ring R, and let N be a submodule of M. (1) M is noetherian if and only if MIN and N are noetherian. (2) Let N' be another submodule of M and suppose that M
= N + N'.
If N ,N' are noetllerian, then so is M . PROOF:
It is clear that a submodule of a noetherian module is itself noethe-
rian. On the other hand, a submodule of MIN is of the form LIN for some submodule L of M which contains N. If M is noetherian, then L is finitely generated. Thus LIN is finitely generated, and MIN is noetherian. Conversely, suppose that N and MIN are noetherian. Let L1
~
L2
~
...
be an infinite ascending chain of submodules of M . Then
is an ascending chain of submodules of N . Since N is noetherian, this chain must stop. In other words, there exists s such that (L. n N) all i
~
= (L
j
n N) for
s. Similarly, the chain
stops; that is there exists r such that N t = max{s , r}, we have that N +L, Thus by Lemma 1.2, L,
= L, for
To prove (2), note that MIN
+ Lr = N + L j ,
for i
~
r. Taking
= N +Lj and L,nN = Ljn N , for
i ~ t.
i ~ tHence M is a noetherian ring. ~
N'/(N'nN} is noetherian by (1). Since
N is noetherian, we may use (1) again to conclude that M is noetherian.
We have pitifully few examples of noetherian modules, but this situation will be rectified in the next section. 2 . NOETHERIAN RINGS.
We say that a ring R is a left noetherian ring if R is noetherian as a left R-module . Exnlllpl('s are fi elds and any principal ideal domain, like
68
8. Noethe1'ian modules
Z or K[x]. Actually, the most important rings in algebraic geometry are noetherian, because the ring of polynomials in finitely many variables is noetherian. This was proved by D. Hilbert in 1890 and it is the keystone of commutative algebra. It is known as Hilbert's basis theorem. We shall now prove a modernized version of Hilbert's result from which the original theorem follows as an easy consequence. 2.1 THEOREM. Let R be
8
commutative noetherian ring. The polynomial
ring R[xJ is noetherian. PROOF: Suppose that R[xJ is not noetherian. Let I be an ideal of R[x] that is not finitely generated. We shall construct, inductively, an infinite ascending chain of ideals in R; thereby achieving a contradiction. Choose
h
E I of smallest possible degree. Assume, by induction, that
h,··· ,1k
have been chosen. Let fk+1 be the polynomial of smallest possible degree in
I \
ClI, ... ,fk) .
Since I is not finitely generated, this construction produces
an infinite sequence of polynomials aj
h, h, ... E I.
Let nj be the degree and
the leading coefficient of j;. Since R is noetherian, the ascending chain of ideals
must be stationary; say (al , "" ak) = (al, " " ak+l)' Hence ak+l for some bi E R. Put
= :L; bja;,
k
9
= fk+1
-
L
bjx",,+l-n'fi'
I
Since, by construction, the degrees satisfy
nl ~
is indeed a polynomial. Note that gEl , but 9 ¢
'»-2 ~ • •. it follows that 9
(h , ... ,fd . However 9 has
smaller degree than f k+1 , a contradiction. Thus I must be finitely generated . A simple induction using Theorem 2.1 is all that is needed to prove the original basis theorem of Hilbert. 2.2 COROLLARY. Let K be a field. The polynomial ring K[Xl,." ,xnJ is IJoethel"ian.
69
8. Noetherian modules
Although every ideal of K[XI,' .. ,X n ] is finitely generated, it is not true that there is an upper bound on the number of generators of ideals in this ring. For example, it is easy to construct ideals
h of K[x!, ,1;2) which cannot
be generated by less than k elements; see Exercise 4.2. We shall now use Hilbert's basis theorem to prove that An is a left noetherian ring. This will follow from a slightly more general theorem. Recall that
B is the Bernstein filtration of An and Sn
= grI' An.
2.3 THEOREM. Let M be a left An-module with a filtration
r
with respect
to the Bernstein Jiltl'ation B. If grY M is a noethel'ian Sn-module, then M is noetherian. PROOF: Let N be a submodule of M, and r' the filtration of N induced by
rj
see Ch.7, §5. Since g"I" N ~ g"I'M and the latter is noetherian, we
conclude that g"I" N is finitely generated. Since the generators of grY' N are finite in number , they have degree ~ m, for some integer m. We wish to show that N is generated by the elements in
r:n'
Suppose that it is not, and let k be the smallest integer for which there
exists v E k
>
r~
which cannot be generated by the elements of
m. Let J-tk be the symbol map of order k of
g"I" N, there exist
ai
E An and
E
'U;
J.Lk(V) =
r' .
r:,..
Clearly
By the hypothesis on
r;, such that
L• Uk-r.(ai)J.Lr,(Uj) 1
where rj
~
m, for i = 1, 2, ... , S . Hence J
V-
L: aiUj E r k_l , 1
which, by the minimality of k, may be written as an An-linear combination of elements in
r:n . Thus vitself may be written as an An-linear combination
of elements in r:". To finish the proof we must show that a finite number of elements of are enough to generate N. But
r:"
r:,.
is a subspace of the finite dimensional
vector space rln' Hence it has a finite basis, which will be a set of generators for N as An-modnlt-.
70
8. Noetherian modules
2.4 COROLLARY. An is a left noetllel'ian I'ing. PROOF: The graded ring Sn associated to the Bernstein filtration is a polynomial ring in 2n variables by Theorem 7.3.1. Hence, by Hilbert's basis theorem, Sn is noetherian. Thus An is left noetherian by Theorem 2.3. It has already been pointed out in Ch.2 that every left ideal of An can be
generated by two elements. Therefore, there is an upper bound on the number of generators of left ideals in An , which is not the case for a polynomial ring. However, this result is very hard to prove and is beyond our means in this book. There are good reasons why one ought to rejoice that An is a left noetherian ring. For example, it follows from Corollary 2.4 and the next proposition that every finitely generated An-module is noetherian.
2.5 PROPOSITION. Let R be a left noetherian ring. Finitely generated left R-modules are noethel'ian. PROOF: Let M be a finitely generated left R-module. If M is generated by Ie elements then there exists a surjective homomorphism 4J : Rk
---t
M. Since
R is left noetherian, it follows from Proposition 1.3(2) that Rk is noetherian. Hence M is noetherian by 1.3(1). There is really nothing special about left modules in tills context, and all the results that we have stated so far hold for right modules with the obvious changes, Thus the ring An is also right noetherian, and we shall use this in later chapters, without further comment. 3. GOOD FILTRATIONS. Let M be a left An-module and
r
a filtration of M with respect to the
Bernstein filtration B. If g~ M is finitely generated, then it is noetherian by Proposition 2.5. Hence M is finitely generated, by Theorem 2.3. However , it is not always true that if M is finitely generated over An then 9~ M is finitely generated over Sn. When g~ M is finitely generated we say that is a good filtration of M.
r
It is nonetheless true that every finitely generated An-module admits a good filtration . Indeed, if !v! is gencratcd by
'ILl,"
" Us
then the filtration
8. Noetherian modules r of M defined by
n = L~ BkU;
71
is good. The graded module g~ M is
generated over Sn by the symbols of '1.£1> ••• , u•. We need an example ofa filtration that is not good. Before that, however, we shall establish an easy criterion to check whether a filtration is good. 3.1 PROPOSITION. Let M be a left An-module. A filtration
r
of M with
respect to B is good if and only if there exists ko such that ri+k = Bin, for all k ~ ko.
PROOF: Suppose that there exists ko such that The K-vector space
n. has a finite basis.
this basis generate g~ M. Thus
r iH =
BiH, for all k
r is good. '1.£11 • • • ,
be elements of M whose symbols generate g~ M. Assume that Uj E for j
Let k
~
ko .
The symbols of the elements in
Conversely, suppose that g".r M is finitely generated over Sn. Let rlcj-l,
~
u.
r", \
= 1,2, ... ,8, and that ko = max{kb ... , k.}. ko. We prove that ri+k
=
Birk by induction on i.
If i
=
0,
the result is obviously true. Suppose that the equality holds for i-I. Pick v E ri+k. Since g~ M is finitely generated, and denoting by J.Lk the symbol
of order k of r, one has
• J.Li+k(V) E LITk+i-k,(BHi-k)J.LIcj(Uj). j=1
Since B iH - ki
= Bi.Bk-kj, we conclude that
•
VEL B i .Bk - k,'I1:i
+ r iH- I .
j=l
Now Bk-k;'U.j E
n, for every j
and, by the induction hypothesis,
r k +i - 1 =
Bi-lH. Since B i - 1 ~ B i , then v E B in . Thus ri+k ~ Bin . The other inclusion is obvious; hence we have an equality, as desired. It is now easy to check whether a filtration is good or not. For example, let
n be the filtration of the left module An defined by nk = B 2k . Then we have that Bink = B i+2k is properly contained in B 2(i+lo) = niH. By Proposition 3.1, n is not a good filtration of An with respect to B. We end this section with a proposition that will allow us to compare two good filtratiuns.
8. Noetherian modules
72
3.2 PROPOSITION. Let M be a left An-module. Suppose that rand [} are two filtrations of M with respect to B.
(1) If r is good then there exists kJ such that rj ~ [}j+k•. (2) If rand [} are good then tllere exists k2 such that [}j-k, ~ rj ~ [}j+ko·
PROOF: By Proposition 3.1 there exists ko such that Birj
=
r i+j , for all
j ~ ko and i ~ O. Since rko is a finite dimensional vector space over K,
there exists kJ such that rko
~ [}k •.
Thus rj+ko
=
Bjrko is contained in
Bj.[}k. ~ [}j+k•. Hence rj ~ Ij+ko ~ [}j+kl
for all j
~
0, which proves (1). To prove (2) apply (1) twice, swapping
r
and [}. The full power of good filtrations will only be felt in the next chapter, where they will be used to define a dimension for finitely generated An-modules. 4. EXERCISES.
4.1 A K-algebra R is affine if there exist elements rJ, ... , r, E R so that the monomials
r;nl ... r;n.
form a K-basis for R.
(1) Show that if R is an affine commutative K-algebra then it is a homomorphic image of a polynomial ring in finitely many variables over
K. (2) Use (1) and Hilbert's basis theorem to conclude that a commutative affine K-algebra is noetherian. (3) Show that (2) is false if the commutative hypothesis is dropped. Hint: Construct an infinite ascending chain of left ideals for the free algebra
K{x, y}. 4.2 Let h be the ideal of the polynomial ring K[Xl,X2) generated by the monomials of degree k. In other words, the generators of h are of the form xi~ with i+j = k. Let M =
h.
(1) Show that M is a maximal ideal of K[Xl' X2], and that K[xJ, X2]/ M
K. (2) Show that h/Mh is a K-vector space of dimension k + l. (3) Show that h canllot be generated by less than k + 1 elements.
8:'
8. Noetherian modules
73
4.3 Show that the results of §3 remain true if we replace the Bernstein filtration B by the filtration by order. Hint: In some of the proofs it will be necessary to use Hilbert's basis theorem and Proposition 2.5. 4.4 Let M be a left An-module with a filtration that i
Ui E
r k., for
= 1, ... s,
r with respect to B.
Suppose
i = 1, ... ,s. Show that if grf M is generated by JJ k, (Ui), for
then M is generated by
UJ, . .. , U •.
4.5 The purpose of this exercise is to show that the converse of Exercise 4.4 is false. Let J be the left ideal of A3 generated by 01 and fh.
+ XlOa.
Recall
that by Exercise 7.6.7, if 8' is the filtration of J induced by the Bernstein filtration of A3 , then g".B'J ~ grJ. (1) Using Exercise 2.4.2, show that grJ is generated by Ul(Oi) for i
=
1,2,3. (2) Show that although 01 and fh.+XI0a generate J, their symbols do not generate grJ. 4.6 Let M be a left An-module with a good filtration annihilator of grf M is a homogeneous ideal of Sn.
r.
Show that the
CHAPTER 9 DIMENSION AND MULTIPLICITY Using the filtered and graded methods of the previous chapters we shall define a dimension for An-modules. This is a very useful invariant and it comes naturally associated with another invariant: the multiplicity. The first section contains a result in commutative algebra that is the key to the definition of dimension in §2.
1. THE HILBERT POLYNOMIAL.
Without further ado we state the main theorem of this section.
It is
expected that the reader will feel somewhat surprised by its content.
1.1
THEOREM.
Let M
=
ffii~O
M; be a finitely generated graded module
over the polynomial ring K[Xl, ... , xnJ. There exist a polynomial x( t) E Q[ ~ and a positive integer N such tllat
•
L dimK(M
i)
= Xes)
o
for every s ;::: N. The polynomial X(t) is known as the Hilbert polynomial of M. We need a technical lemma before we come to the proof of the theorem. The following definitions and notations make the proofs more bearable. A numerical polynomial is a polynomial p( t) of Q[ 4 such that p( n) E Z for all integers n »0. The difference function of a function f : C
LlJ(z)
= fez + 1) -
---t
C is defined by
J(z). Finally, set
G)
=t(t-l) ... (t-r+1)/r!
where 'r iR a positive illtogOl' and t is a variable. If r
= 0 we put
(~) = 1.
9. Dimension
75
1.2 LEMMA. The following are true:
(1) L1(~) = C~l)' (2) Let p( t) E Q[ t] be a numerical polynomial.
There exist integers
Co, . .. , Ck such that
In particular p( n) E Z, for all n E Z. (3) Let f : Z
--t
Z be a function. Suppose that there exists a numerical
polynomial q(t) E Q[t] such that L1f(n) = q(n), for all n» O. Then there exists a numerical polynomial p( t) E Q[ ~ such that f (n) = p( n),
for all n PROOF:
» O.
(1) follows from a straightforward calculation that we omit. Let
us prove (2) by induction on the degree of p( t). If p( t) has degree 0, then it equals an integer. Suppose that the result holds for all numerical polynomials of degree
~
k - 1. Let p( t) be a numerical polynomial of degree k. Since the
leading term of (~) is t r / r!, there exist rational numbers Co, •.. , Ck such that
pet)
= 2:~ Ck-iG)·
To complete the proof we must show that
Since L1p( t) has degree k - 1, it follows by induction that Thus p( t) -
Ck
Ci E
Z. By (1),
Ck-l,'" ,Co
E Z.
is integer valued. Since p( t) is a numerical polynomial, it
follows that for r» 0 both per) and per) -
Ck
are integers. Hence c/o E Z,
and we have proved (2). We now prove (3). By (2), the numerical polynomial q(t) may be written in the form
q(t) =
~ Ck-.(:) ,
with Co, ... , Ck E Z. Let k
pet)
= ~ Ck-i(i:
1)'
76
9. Dimension
By (1),
we
= q(t).
have that ..1p(t)
Hence..1(J - p)(n)
=
°for all n» O.
But this is possible only if (J - p)(n) equals a constant integer, to be called Ck+l,
for all n» 0. Thus f(n)
= p(n) + Ck+b for n» 0, as required.
PROOF OF THE THEOREM:
The proof is by induction on the number n of variables. If n
=
0, we
are reduced to the base field K. In this case the finitely generated module
M is a vector space of finite dimension. Hence it has only finitely many homogeneous components. Thus
L• dimJ( Mi = dimK M, a
for s
» 0, and we may choose X(t) = dimJ( M, a constant.
Suppose, by induction, that the theorem holds for any finitely generated graded module over K[xJ, .. . , Xn-l]' Let M
=
E9i~O
Mi be a finitely gen-
erated graded module over K[xI, ... , xn]. Define a function
f :Z
--t
Z
by 8
1(s)
=L
dimJ( Mi
-00
where Mi
= 0 if i < O.
Let ¢Ji : Mi
--t
M,+l be the linear map of vector
spaces defined by multiplication by x n' Set Qi = ker (
Now Q
= E9i~O Qi and
over K[Xl"" ,Xn].
L
= E9,~o Li
are finitely generated graded modules
They are respectively the kernel and cokernel of the
endomorphism of M defined by multiplication by X n • Moreover, the elements of Q and L are annihilated by
X n,
hence these are in fact finitely generated
graded K[Xl' ... ,xn_l]-modules. Thus, by induction, there exist polynomials Xl (t),
X2( t) E Q[ t] such that s
XI(S) =
L ()
•
dimJ( Qi and X2(S) = Ldiml( L j , 0
9. Dimension for s
77
» o.
On the other hand, the dimensions of the vector spaces in (1.3) are related by the formula dimK
Qi-l -
Adding these up for 0
:s;
dimK M i - 1 + dimK M; - dim]( Li i:S; sand s
Xl(S -1)
»
= O.
0, one obtains
+ <1f(s) -
X2(8)
= o.
Hence <1f (8) is a numerical polynomial for s »0. We may now conclude from Lemma l.2(3) that f(s) itself is a numerical polynomial for all s
»
0,
as required.
2. DIMENSION AND MULTIPLICITY. We are ready to define the dimension of a module over An. Note that the Bernstein filtration B is essential for the construction that follows. Let M be a finitely generated left An-module. Suppose that
r
is a good fil-
tration of M with respect to the Bernstein filtration B. Denote by X( t, r, M) the Hilbert polynomial of the graded module grI'M over the polynomial ring Sn. By Theorem 1.1, we have, for
t» 0,
t
X(t,r,M) = EdimK(r;jr;-l) o
(2.1)
= dim](rt ).
The last equality in (2.1) follows from the fact that dimK is additive over exact sequences of vector spaces. The dimension d(M) of M is the degree of X( t, r, M). Let ad(M) be the leading coefficient of X(t,
r, M).
The multiplicity of Mis m(M)
Both numbers are non-negative integers.
= d!ad(M)'
This is obvious for d(M), and
follows from Lemma 1.2(2) for m(M). The definitions of dimension and multiplicity apparently depend on the good filtration
r from which the Hilbert polynomial is calculated. Let us
show that any choice of good filtration will give the same result. Suppose that rand
n
are two good filtrations of M. By Proposition 8.:3.2, there exists
78
9. Dimension
Ie, such that
dimK
[}j+k.
{}j-k ~
rj
~ {}j+k.
In particular, dimK
We conclude from Theorem 1.1 that for :f ~
x(;i - k, [},M)
xC,;", r,M)
Since the behaviour of a polynomial at
00
~
[}j-k
»
< dimK Fj <
0
xC;' + k, [}, M).
is determined by its leading term,
it follows that X( t, [}, M) and X( t, r, M) have the same degrees and leading coefficients. Hence d(M) and m(M) are independent of the choice of good filtration for M. Let us consider a few examples. First let M be the left An-module An. The Bernstein filtration B is a good filtration of M and it is possible to calculate X(t,B,M) explicitly in this case. By (2.1) we must determine the dimension of B t . But the monomials x a Ef1 with
lal + 1.81
~ t form a basis
of B t as a K -vector space. So it is enough to count the elements of this basis. To do this we must count the non-negative solutions of the equation
+ ... + an + f31 + ... + f3n
t. It is an easy exercise in combinatorics to show that there are c~~n) such solutions. Hence X( t, B, M) = (~~n). As a polynomial in t it has degree 2n and leading coefficient 1/(2n)!. Thus d(An) = 2n and meAn) = 1. al
~
Another An-module that we know very well is K[X] eh. 7 we defined a filtration polynomials of degree
~ t.
r
of K[X] such that
= K[Xl,'"
rt
is the space of all
Since Bi contains the polynomials in
of degree i, we have that Birt
,x n]. In
Xl, ... ,
Xn
= r t+j. Hence r is a good filtration. It is easy
to show that dimK r t = (n~t). Thus
xC t, r, M) =
(n~t) is a polynomial of
degree nand leading coefficient lin!. Hence d(K[X])
= nand m(K[X]) = 1.
A large class of examples is obtained by twisting a module by an automorphism, as in Ch. 5, §2. Curiously, this will lead us to a different definition for the dimension. Recall that if u is an automorphism of An and M is a left
An-module, then the action of a E An 011
U
EM" is defined by a. u = u(a)u.
We begin with the Fourier transform of a module. In this case the automorphism :F is defined by :F(x;)
for 1
~
:F(B,)
= OJ and
F(o;)
=
-Xj,
i ~ n. This automorphism :F preserves the Bernstein filtration; thus
= B i . This is very convenicut.
9. Dimension
79
2.2 PROPOSITION. Let M be a finitely generated left An-module. Then M
and M:r have the same dimension and multiplicity. PROOF: Write
= BkrO
M. Then H by
ro for a K-vector space whose basis is a set of generators of is a good filtration for M. Now M:r is also generated
Fa, so we may construct a good filtration for M:r by setting [h = Bk • roo
Since :F preserves Bk, we have that ilk
= rio.
Hence M:r and M have the
same Hilbert polynomial, therefore also the same dimension and multiplicity. Things are a lot more complicated if the automorphism does not preserve the filtration. To get around the problem we must give a different definition for the dimension.
Start by choosing a finite number of elements which
generate An as a K-algebra and let Vbe the K-vector space generated by these elements and by 1. Put
Ua = K and Uk = V'. This is a filtration of An which satisfies dimK Uk < then Uk
= Bk
00.
Note that if V
= BJ,
is the Bernstein filtration of An.
Now let M be a finitely generated left An-module, with a good filtration
r
with respect to the Bernstein filtration. Without loss of generality we may assume that r k = BkrO, for k
,sCM, V)
~
O. Put ilk
= inf{v : dimK [l.
= UkrO ~
and
s" for s
»
O}.
2.3 PROPOSITION. Let V be a vector SIJace whose basis is a finite set of
generators fol' An. Then ,s(M, V)
= d(M).
PROOF: Since V is finite dimensional, there exists r E N such that V <; B r . Thus Uk <; Brio. Furthermore,
Hence dimK ilk ~ dim]( rrk.
In particular, 6(M, V) ~ 6(M, B J ).
The
= 6(M, BJ)'
But
opposite inequality is proved similarly. Thus 6(M, V)
dimJ( r. = Xes, M, r) for s » O. Since d(M) is the degree of the polynomial Xes, M, r), then SCM, B 1 ) = d(M), and the proof is complete. The result
w(~
want is a simple corollary of this proposition.
9. Dimension
80
2.4 COROLLARY. Let M be a finitely generated left An-module and u an
automorphism of An, then d(M,,) PROOF: Let V
= u(BJ ).
= d(M).
Since u is an automorphism, we have that V has
generators of An for its basis. Define Uk as above. Let
r be a good filtration
of M with respect to the Bernstein filtration, and assume that r k
= Bkra.
Now let
This a good filtration for M", hence by Proposition 2.3, d(M,,) = c(M", BJ)' But ilk = U(Bk)ra. Since U(Bk) = Uk, we have that ilk = Ukra and so c(M", Bd = c(M, V). By Proposition 2.3, the latter equals d(M). In ring theory 6(M, V) is called the Gelfand-Kirillov dimension of a module. It is discussed in detail, including many applications, in [Krause and Lenagan 85J and [McConnell and Robson 87, Ch. 8]. 3.
BASIC PROPERTIES.
Let M be a finitely generated left An-module and
ra good filtration of
M with respect to B. Let N be a submodule of M. Denote by r' and the filtrations induced by
r in N
r"
and MIN, respectively; see Ch. 7, §5. It
follows from Lemma 7.5.1 that we have an exact sequence of Sn-modules, namely
o - t grr'N - t grrM - t grr"MIN - t O. Since
r
is good,
gr M
is finitely generated. But Sn is a noetherian ring.
Hence gr' Nand grl! (MIN) are also finitely generated. Therefore both r' and
r"
are good filtrations.
On the other hand, since the sequence of vector spaces
is exact, we have that
Summing these terms for k = 0, 1, ... , s and assuming that s (3.1)
xCs, r', N) + xC·~,r", MIN) = xes, r, M).
»
0 one obtains
81
9. Dimension
3.2 THEOREM. Let M be a finitely generated left An-module and N a submodule of M. (1) d(M)
= max{d(N),d(MjN)}.
(2) If d(N)
= d(MjN)
then m(M)
= meN) + m(MjN).
PROOF: Since M is finitely generated it admits a good filtration
r.
Let r'
and r" be the good filtrations of Nand MjN induced by r. Thus (3.1) holds for an infinite number of values of s. Hence we have an equality of polynomials:
x(t, r', N)
+ X(t,r", MjN) = X(t, r, M).
Since d(M) corresponds to the degree of X( t, r, M), we have that d(M)
~
max{d(N), d(MjN)}.
But the leading coefficients of these polynomials are positive. Thus we must have equality in the formula above, and (1) is proved. Now, if d(MjN)
=
deN) then all these polynomials have the same de-
gree. Thus the leading term of x( t, r, M) is the sum of the leading terms of X(t,r',N) and X(t,r",MjN), and (2) follows immediately.
This theorem is useful in calculating the dimension of some modules. We saw in §2 that deAn)
= 2n.
We may use the theorem to calculate the dimen- .
sian and multiplicity of a free module of finite rank r: it has dimension 2n and multiplicity r. This follows from the following result. 3.3 COROLLARY. Let MJ. .. . , Mk be finitely generated left An-modules, and M = Ml EEl ..• EEl M k . (1) d(M) = max{d(M\) , ... , d(Mk)}.
(2) If d(M) = d(M;) for 1 ~ i ~ k, then m(M) = I:~ m(Mi ). PROOF: The prooffollows by induction if we apply Theorem 3.2 to the exact sequence
We may also usc the theorem to get an upper bound on the dimension of a finitely generated An-module.
9. Dimension
82
3.4 COROLLARY. Let M be a finitely generated An-module. Then d(M)
~
2n. PROOF:
Suppose that M is generated by r elements.
a surjective homomorphism
A~ ~
= max{ d(M), d(ker
that d(A~)
Then there exists
M. It follows from the theorem
Since d(A~)
= 2n by
Corollary 3.3, we
conclude that d(M) ~ 2n. This upper bound may be sharpened if the module is a quotient of An by a left ideal. 3.5 COROLLARY. Let I be a non-zero left ideal of An. Then d(Anl I)
~
2n-1. PROOF:
I
First consider the case of a cyclic left ideal. Let dEAn, and put
= And.
Then we have an exact sequence
where the map () is defined by 8(a) contradiction, that d(AnIAnd)
Since m(An)
= ad, for every a
E
An. Suppose, by
= 2n. Then by Theorem 3.2(2), we have that
= 1 and the multiplicity is a positive number, this equation is
impossible. Hence d(Anl And)
~
2n - l.
Now fur the general case. Let I be a non-zero left ideal of An and choose
o :f dEl.
Since And £; I, we have that Ani I is a quotient of Ani And. Since
the latter has dimension
~
2n - 1, so does Ani I, by Theorem 3.2(1).
4. BERNSTEIN'S INEQUALITY.
At the end of the previous section we obtained an upper bound on the dimension of a finitely generated An-module. In this section we establish a lower bound. We begin with a lemma from linear algebra. 4.1 LEMMA. Let M be a finitely generated left An-module wit}l filtration
with respect to B. Suppose that
ra :f O.
The K -linear transformation
r
9. Dimension
83
which maps a E B j to the linear transformation <Pa( u) = au is injective. PROOF:
The statement of the lemma is equivalent to ari
o =f a E Bj.
We will prove this by induction on i. If i
the statement above is equivalent to
ra =f O.
Suppose that if 0 =f bE Bi-l then br;_l of B j • If arj = 0, then a
~
monomial. Then
OjCX
then Bo
=K
and
This is true by hypothesis.
=f O. Let a be a non-zero element
K. Hence the canonical form of a has a term
cx o Ef3 with c E K \ {O} and O
= 0,
=f 0 whenever
101 + 1.81 >
O. Suppose that
OJ
=f 0 for this
-£iEf3 is a summand in the canonical form of [a, a;].
Thus [a, a;] is a non-zero element of B j -
[a, 8i ]r.- 1
~
1•
Since ari = 0, we conclude that
aair.- 1 .
= 0, which contradicts the induction hypothesis. We may reach a similar contradiction by assuming that .8; =f o.
But O;ri - 1
~
r i , hence [a, a;jr;_l
Thus the lemma is proved. 4.2 THEOREM.
If M is a finitely generated non-zero left An-module, then
d(M)? n. PROOF:
Choose a set of generators for M and let
r
be the good filtration
obtained by giving each of these generators degree zero, see Ch. 8, §3. Then
ra =f O.
Let x( t)
= x( t, r, M) be the corresponding Hilbert polynomial.
By the lemma, B j can be embedded in HomK(rj , r2i). In particular,
But HomK (r;, r 2i ) has dimension dimK F; . dimK
Hi.
Thus, assuming that
i» 0, we get that dimK Bi ~ X(i)x(2i).
On the other hand, dimJ( B; = c~~n) is a polynomial in i of degree 2n. Hence, as a polynomial in i, X( i)X(2i) must have degree? 2n. But the degree of X(i)X(2t) is 2d(M). Thus d(M) ? n, as required. This inequality was first proved by I. N.Bernstein in [Bernstein 72j; and is often called the Bernstein inequality. The proof above is due to A. Joseph. We have already seen that both bounds for the dimension are attained. For example, d(An)
=
211, and d(K[X]) = n. In fact. there exist An-modules of
9. Dimension
84
dimension k for every k in the interval n to 2n; see Exercise 5.3. The modules of minimal dimension are so important that Ch. 10 will be devoted to them.
5.
EXERCISES
5.1 Show that if M is a finitely generated torsion An-module, then d(M)
:s;
2n-1.
5.2 Show that if d is a non-zero operator of An then d(AnIAnd)
= 2n-1.
Hint: Let k be the degree of d. There is an exact sequence,
where 8" is the filtration of Ani And induced by B. Using this sequence show that the Hilbert polynomial of Ani And is (2~~t)
-
C~~~k), and that it has
degree exactly 2n - 1. 5.3 Let k be a positive integer such that 1 :s; k subalgebra of An generated by
Xl, ... ,X n
:s;
n. Denote by Rk the
and 01, ... , Ok. Let J
= AnOk+l +
... + AnOn. (1) Show that the complex vector space Bi+JIJ is isomorphic to BinRk. (2) Show that the Hilbert polynomial of Ani J with respect to the filtration induced by B is
(n:tt).
(3) Show that Ani J has dimension n + k. What is its multiplicity? 5.4 Let D be a (noncommutative) domain. We say that D satisfies the Ore condition if Dan Db =I 0, for any two non-zero elements a, bED. A division ring Q is a left quotient 1'ing of D if (1) D is a subring of Q;
(2) every non-zero element of D is invertible in Q; (3) every element of Q is of the form b-ia where a, bED and b =I O. Show that if a domain has a left quotient ring then it satisfies the Ore
COll-
dition. The converse is also true, but it is harder to prove. For details see [Cohn 79, Theorem 12.1.2].
9. Dimension
85
5.5 Show that An satisfies the Ore condition. Hint: Let a, b be non-zero elements of An. If Ana n Anb
= 0,
then An maps
injectively to An/Ana ffi An/ Anb. This contradicts Corollary 3.5. 5.6 Let dE An(C) be an operator of degree 2. Set M
= An/And.
(1) Show that there exists a submodule N of M such that d(M/N) (2) Show that M cannot be an irreducible module if n
~
= n.
2.
Hint: By Exercise 7.6.11, there exists an automorphism a of An(C) such that a(d)
=
2:=~ d; where di
=
~i(X~
+ af) + Ti
for ~i'
Ti E
C. Let J be
the left ideal generated by d1 , ... ,dn. Then An/ J is a homomorphic image of Ma-l
~
An/And and d(An/ J)
of An/J see Exercise 13.5.8.
= n.
For an easy method to find the dimension
CHAPTER 10 HOLONOMIC MODULES
The most important An-modules are the holonomic modules, also known among PDE theorists as maximally overdetermined systems. An An-module is holonomic if it has dimension n.
Ordinary differential equations with
polynomial coefficients correspond to holonomic modules. In this chapter we begin the study ofholonomic modules, which will be one of the central topics of the second half of the book.
1.
DEFINITION AND EXAMPLES.
A finitely generated left An-module is holonomic if it is zero, or if it has dimension n. Recall that by Bernstein's inequality this is the minimal possible dimension for a non-zero An-module. We already know an example of a holonomic An-module, viz. K[X]
= K[XI, .. . , x n ]. We also know that
An itself is not a holonomic module: it has dimension 2n. It is easy to construct holonomic modules if n ideal of AI' By Corollary 9.3.5, d(AdI) inequality, deAl I I)
= 1.
~ 1.
If I
= 1.
Let I
=I 0 be a left
=I AI, then, by Bernstein's
Hence AdI is a holonomic AI-module.
This is wonderful source of examples, which will pour forth with the help of the next proposition. 1.1 PROPOSITION. Let n be a positive integer.
(1) Submodules and quotients of llOlonomic An-modules are holonomic.
(2) Finite sums of lwlonomic An-modules al'e holonomic. PROOF:
(1) These follow from Bernstein's inequality. Let M he a left An-
module, and N a suhmodule of M. From Theorem 9.3.2, deN) d(MjN)
~
d(M). Since d(M)
~
d(M) and
= n, and using Bernstein's inequality, we
deduce that deN) = d(MIN) are also equal to n. Thus N and MIN are holonomic. Now (2) follows from Corollary 9.3.3 anu (1).
10. Holonomic modules
87
1.2 COROLLARY. Finitely gene1·ated torsion AI-modules are holonomic.
PROOF: Let M be a finitely generated torsion AI-module. Suppose that it is generated by UI, ... , U r • Since M is torsion, for each i exists 0
=I bi
E Al such that
biu;
= O. Hence
= 1,2, ... , r there
Al U; is a quotient of At! Al bi
which is a holonomic module. Thus each Al U; is holonomic. Since M is the sum of AIuj, for i= 1,2, ... , r, it is holonomic by Proposition 1.1(2).
=I
~
J are left ideals of AI, then the quotient I I J is a torsion AI-module. A good It is easy to construct torsion AI-modules. For example, if 0
I
exercise, which the reader may wish to try, is to prove that I I J is holonomic using only Proposition 1.1(1). Still on the subject of torsion modules, we have the following proposition.
1.3 PROPOSITION. Holonomic An-modules are torsion modules. PROOF: Let M be a holonomic left An-module. Suppose that 0 element of M. Consider the map
~
M, it follows that d(im
= n.
=I 0,
u is an
M defined by
Thus by Theorem 9.3.2,
2n = deAn) In particular ker ¢
~
=I
= d(ker ¢).
and u is a torsion element of M.
Putting 1.2 and 1.3 together, we conclude that for finitely generated A I modules, torsion and holollOmic are equivalent concepts. However, it is not true that all torsion An-modules are holonomic when n consider the module M
= Ani AnOn.
~
2. For example,
The Bernstein filtration induces in M
a filtration defined by
r i is isomorphic to the vector space generated by the monomials in XI, . .. , Xn and Bt, ... , On-I of degree::; i. We may use this to calculate the Note that
Hilbert polynomial of M explicitly, using combinatorics. It equals (2;!~~I). As a polynomial in i, this binomial number has degree 2n- 1; see Exercises
9.5.2 and 9.5.3. Hence d(M)
= 2n-1 > n, if n ~
2. On the other hand, it is
very easy to check that M is a torsion module: its elements are annihilated by powers of Dn·
10. Holonomic modules
88
2. BASIC PROPERTIES. Many interesting properties ofholonomic modules follow from the fact that they are artinian. A module M over a ring R is said to be arlinian if given a descending sequence Nl k such that Nk
=
Nj
,
~
N2
~
for every ;"
~
... of submodules of M, there exists k. In other words, every descending
chain of submodules of M is stationary. This is like the noetherian property. The following proposition collects some of the basic properties of artinian modules.
2.1 THEOREM. Let M be a left module over a ring R and let N be a submodule of M.
(1) M is artinian if and only if evel"J set S of submodules of M lJas an element (a minimal element) which does not contain any other element of S. (2) M is artinian if and only if N and MIN are artinian. (3) Let Nt be anotllel" submodule of M, and suppose that M = N + NI. If N, Nt are al·tinian, tllen so is M. The proof of (1) is like that of Theorem 8.1.1 and the proofs of (2) and (3) are like those in Proposition 8.1.3. The details are left to the reader. Let us now see why we are interested in artinian modules. 2.2 THEOREM. Holonomic modules are artinian. PROOF: Let M be a holonomic left An-module.
Suppose that M has a
descending chain of submodules
By Theorem 9.3.2 and Proposition 1.1(1), it follows that meN;)
= m(Ni+l)+
m(Ni/Ni+l). Putting these together, we get that r-1
m(M) =
L: m(N;jN +
i 1)
+ m(Nr )
~ r.
a Hence M cannot have a descending chain of more than r sub modules. In particular M cannot have an infinite descending chain.
10. Holonomic modules
89
Let us dispel any hopes that the reader may have formed, by saying that not all An-modules are artinian. For example, An is not artinian as a module over itself. It is easy to construct an infinite descending chain; take for instance
More examples of non-artinian An-modules are found in Exercises 4.2 and 4.3. A ring R that is artinian as a left R-module is called a left artinian
ring. The argument above shows that An is not a left artinian ring. Artinian rings have a very nice representation theory; see [Reiten 85]. We return, for a moment, to the general situation. Let M be a left nonzero module over a ring R. Assume that M is both artinian and noetherian. Suppose we have constructed a chain No <;;; Nl <;;; ..• <;;; Nk of submodules of M such that N;/Ni -
1
is irreducible. It follows from Theorem 2.1(1) that
there exists a submodule Nk+l of M such that Nk+llNk is a minimal nonzero submodule of MINk.
Since Nk+llNk is minimal in the set of non-
zero submodules of MINk, then it must be irreducible. Thus in the chain
No <;;; ... <;;; Nk <;;; Nk+l, the quotient of two adjacent terms is irreducible. This process cannot be continued forever, since M is also noetherian. Thus we get a chain
o = No <;;; ..• <;;; N r = M of submodules of M such that N;/ N i -
1
is irreducible for i
= 1, ... , r.
Such a
chain is called a composition series of M. It follows from the Artin-Schreier theorem [Cohn 84, 9.2 Theorem 2) that any two composition series of M have the same number of submodules. The number r is called the length of M. Since a holonomic module is noetherian and artinian, it must have a composition series. Hence its length is well-defined. It may be calculated using the argument in the proof of 2.2. 2.3
SCHOLIUM.
The lengtll of a holonomic module cannot exceed its multi-
plicity. 2.4 COROLLARY. A llOlonomic An-modules of multiplicity 1 is b-reducible. PROOF:
Let M be a left holonomic module of multiplicity 1 and suppose
that it has a lIoll-zero submodul0. N. Then Nand A1jN are also holollomic
10. Holonomic modules
90
= m(N) +
by Proposition 1.1. Hence by Theorem 9.3.2, we have that m(M)
m(MIN}. Since m(M) = 1 and N of 0, we have that MIN = O. Thus
= N and M must be irreducible.
M
It is surprisingly easy to calculate the number of generators of a module
that is artinian and noetherian. 2.5 THEOREM. Let R be a simple left noetherian ring and M a finitely
generated left R-module. If M is aJ:tinian but R is not artinian (as a left R-module), then M is a cyclic module. PROOF: The statement that we shall prove is even sharper: if M is generated by
Ub""
Ur
then there exist
a2, ... , a r
M. It is enough to prove the case r
E R such that
Ul
+ L:; aiU; generates
= 2. The general case follows by
induction. Suppose that M is generated by two elements, say u and v. Since M is artinian and noetherian, it must have finite length. We shall prove, by induction on the length of M, that there exists a E R such that M
=
A(u+ av). If M has length 1, then it is irreducible and there is nothing to prove. Suppose that the result holds for all R-modules of length < r, and that M has length r. In particular Rv has finite length, so it must have a non-zero irreducible submodule. This must be cyclic, hence generated by an element of the form
cv, for some e E R. Since M I Rev has smaller length than M, we may use the induction hypothesis to construct an element A E R such that
M
= R(u+ AV) + Rev.
It is better to rephrase this by saying that one may assume that M
= Ru+ Rv,
with Rv an irreducible module. Let > : R
-+
M be the homomorphism of modules defined by >(x) = xu.
If > is injective, then R ~ im> is a non-artillian submodule of M, which
contradicts Proposition 2.1. Hence ker > of O. Let d be a non-zero element of ker >. Since R is a simple ring, one has that RdR = R, and so RdR . v of O. In particular, there exists b E R, such that dbv of 0 . We claim that M is
10. Holonomic modules generated by u + bv. First note that d( u + bv)
= dbv is a
91 non-zero element
in Rv. Since Rv is irreducible, v is a multiple of dbv. Hence v, and therefore
u, are elements of R( u + bv), which must then equal M. 2.6
COROLLARY.
PROOF:
Holonomic modules are cyclic.
We know from Theorem 2.2 that a holonomic An-module must be
artinian. We have also seen that An is not artinian. The result follows from Theorem 2.5.
3.
FURTHER EXAMPLES.
In this section we construct a family of holonomic An-modules which is very important in applications. We begin with a technical lemma. Note that in this lemma we do not assume that the module is finitely generated: this will be proved as part of the lemma.
3.1
LEMMA.
Let M be a left An-module with a filtmtion
r
with respect to
the Bernstein filtration of An. Suppose t1lat tIl ere exist constants cI, c2 such
»0
that for j
dimK rj :::;
cdn/nJ + C2(;j + l)n-l.
Then M is a holonomic An-module whose multiplicity cannot exceed particular M is finitely generated. PROOF:
CI'
In
The proof breaks up naturally into two parts.
Part 1: Every finitely generated submodule of Mis holollomic of multiplicity :::; CI·
Let N be a finitely generated sub module of M. Thus N admits a good filtration, say [} . By Proposition 8.3.2, there exists a positive integer r such that
[}j ~
IJ+r n N. III particular, dimK
[}j :::;
dimK rj+r' Let X(t) be the
Hilbert polynomial of N for the filtration [} . Using the polynomial bound of the hypothesis, we conclude that, for very large J',
In particular the degree of X(t) cannot exceed n. Hence d(N) :::; n. From the Bernstein inequality we conclude that d( N)
= n equals the degree of X( t).
Going back to the polynomial inequality above, we obt.ain that. meN) $
Cl.
10. Holonomic modules
92
Part 2: M is finitely generated. Consider an ascending chain Nl
~
~
N2
'"
~
N r of finitely generated
sub modules of M. Each of these is holonomic. By Theorem 9.3.2, m(Ni )
m(Ni - 1 )
+ m(Ni/Ni-l), for
=
i = 1,2, ... , r. Adding them up one has
r
L m(Ni/Ni-l) + m(N
l)
= m(NT ) ~
Cl·
2
In particular, all ascending chains of finitely generated submodules of M have less than
Cl
steps. Thus M must be finitely generated. We may now apply
Part 1 to M itself, and conclude that it is holonomic of multiplicity
Let K(X)
~ Cj.
= K(Xl,"" xn) be the field of rational functions. We may
extend the left action of An on the polynomial ring K[X] to the field of rational functions. The
Xi
continue to act by multiplication. The action
of Oi on a rational function
f / 9 is defined by the rule for differentiation of
quotients, namely 8i(J /g) = (8;(J)g - f8;(g))/i.
A routine calculation shows that this action satisfies the required properties. Note that this module is not finitely generated. Suppose that a polynomial p is chosen in K[X]. Let K[X, p-l] denote the set of rational functions of the form
f /pr, where f
is a polynomial. Note
that a partial derivative of f / pr has denominator rfT. Hence these rational functions are preserved by partial differentiation and by multiplication by a polynomial. In other words, K[X,p-l] is a left An-submodule of K(X). 3.2 THEOREM. The An-module K[X, p-l] is holonomic and its multiplicity
is
~
(deg(p)
+ 1)n.
PROOF: To save on notation, put M
= K[X, p-l].
Suppose that p has degree
m. Set
r k = {f/l : deg(f) We first check, in detail, that
r
~ (m+ l)k}.
is a filtration for M.
10. Holonomic modules
93
Let IIrY' be an element of M and assume that I has degree s. Then IIrY' = I· P'1P'+k· But Ip' has degree s(m+ 1), which is ~ (m+ 1)(s + k). Hence
f IV<
E r.+k' It follows that
Next suppose that f tiplication by
Xi
II E n.
M is the union of all r k, for k ~ O.
I
Equivalently,
has degree ~ (m+l)k. Mul-
increases the degree of f by 1, thus Xj(J IIi)
rk+l' Differentiating
I frY'
with respect to
The numerator has degree ~ (m + l)k
Xi
= x;jplIi+ 1 E
we get
+ (m -
1), so that OJ(J IIi) E n+l.
This may be summed up as
Since B; = B1 we also have that Birk ~ ri+/:o Finally, the dimension of
rk
space of polynomials of degree
cannot exceed the dimension of the vector ~
rk . d111lK
~
rk is finite dimenr is a filtration of M, and shows that
(m + 1)k. Hence each
sional. This concludes the proof that
(m+ l)k n
+
n)
•
Since the two terms of highest degree in k of this binomial number are
it follows that
for very large values of k. By Lemma 3.1, M must be a holonomic module of multiplicity ~ (m + 1), as required. We may use the theory we have developed so far to construct the Bernstein polynomial of a differential operator with polynomial coefficients. As before, we start with a polynomial p E K[xl> . .. ,xnJ. Let s be a new variable. We are going to construct a holollomic module over the ring A,,(K(s», wherE'.
10. H olonomic modules
94
K (8) is the field of rational functions on s. The generator of this module will be denoted by p'. This is a formal symbol on which 8j acts by
It follows from this formula that An(K{s»P' is an An(K{s))-submodule of
K{s)[X,p-l]p'. Now define an automorphism t of K(s)[X,p-ljp' by the formula t(SiP')
= (s + l)ip. p'.
Note that this is An(K)-linear, but not
An(K{s))-linear, see Exercise 5.8.
3.3
THEOREM.
Let p E K[X]. There exist a polynomial B(s) E K[s] and a
differential operator D(s) in the polynomiall'ing An(K)[sj such that B(s)p' PROOF:
= D(s)pp'.
The An{K(s))-module K(s)[X,p-ljp' is holonomic. The proof is
like that of Theorem 3.2 and is left to the reader; see Exercise 4.6. Since An(K(s»P' is a submodule of K(S)[X,p-l]P', it must be holonomic. In particular An(K(s))P' has finite length by Scholium 2.3. Thus the descending sequence
An(K(s))p' ;2 An(K(s»p· p' ;2 An{K(s»p'l. p' ;2 '" must terminate. This means that there exists k > 0 such that
Applying t- k to both sides of this equation, we get p' E An(K(s»)p· p'. Clearing denominators, we conclude that there exists a polynomial B(s) E
K[s] such that B(s )p'
E
An(K)[s]p. p'
and the theorem is proved. The polynomial B(s) and the operator D(s) are not uniquely determined by p. However, the set of all the possible polynomials B(s) satisfying Theorem 3.3 is an ideal of K[s], as one can easily verify. The monic generator of this ideal is thus unique; it is called the Bernstein polynomial of p and denoted by bp(s).
10. Holonomic modules
95
The calculation of the Bernstein polynomial for a given p can be a very complicated affair. Here is a famous example that can actually be done by hand. Let p = x~
+ '" + x!. If we denote by
D the differential operator
8? + ... + o~ then D.]I+!
=
4(8
+ 1)(8 + n/2)]I.
Thus bp (8) = 4(8 + 1)(8 + n/2). Note that in this case D is independent of 8. The polynomial bp(s) was introduced by Bernstein in [Bernstein 71] as part of his solution of Gelfand's problem on the meromorphic extension of a certain analytic function. Bernstein's work is discussed in detail in [Krause and Lenagan 85, Ch. 8] and requires a basic knowledge of distributions. The Bernstein polynomial is also very important in singularity theory; see [Malgrange 76]. Building on work of Malgrange, Kashiwara showed that the roots of the Bernstein polynomial are strictly negative rational numbers; see [Kashiwara 76]. Kashiwara's approach works over any analytic manifold. For details see [Granger and Maisonobe 93, Ch. VI] and [Bjork 79, Ch. 6]. The explicit calculation of the Bernstein polynomial and its roots has been the subject of many papers; see for example [Yano 78] and [Cassou-Nogues 86].
4.
EXERCISES.
4.1 Show that An/Anan is a torsion An-module.
Hint: Every class in An/AnOn contains a representative of the form 2:=~ djx~, where d j E An-I' Show that this element is annihilated by O~+I E An/Anon. 4.2 Let I be a non-zero left ideal of An. Show that I is not an artinian
An-module. 4.3 Show that An/Anon is not an artinian An-module. Hint: The modules
form an infinite descending chain in An/ Anan. 4.4 Let p E K[X] be a non-constant polynomial. Is the module K[X, p-I) irreducible?
10. Holonomic modules
96 4.5 For r ~ n, let P AnP
~
= 2:; Eft -
2:~+1
at.
Find a left ideal J of An such that
J and Anj J is holonomic.
4.6 Let p E K[X].
Show that K(s)[X,p-l]V is a holonomic An(K(s»-
module. Hint: Let m be the degree of p. The K-vector spaces,
rk =
{q. p-k . V
:
deg(q) :::; (m+ l)k} give rise to a filtration of K(s)[X,p-l]. What are their dimensions? 4.7 Let p be a polynomial in K[Xl, ... , xn]. Denote by An [s] . V the submodule of K[s, X, p-l] generated by
V over
the polynomial ring An[s]. This is
also an An-module, but in this case it is not clear whethe1' it is finitely
generated! Show that if p belongs to the ideal of K[X] generated by its partial derivatives, then An[s] . V is finitely generated as an An-module. Hint: Let D
= 2:~ -I;; a;.
Then D· f
= sf·
= An[s]V be as in Exercise 4.7. Let map defined by t(D(s). V) = D(s + l)p· V.
4.8 Let p and M
t:
M
-t
M be the
(1) Show that t is an endomorphism of M as an An-module but not as an An[s]-module. (2) Show that [t, s]
= t.
(3) Use (2) to show that MjtM is an An[s]-module, even though tis not
An[s]-linear. It is true that MjtM is a holonomic module, but this is ha1'de1' to prove; see [Granger and Maisonobe 93, eh. VI].
= exp(Alxl+" +Anx n). = {P E An(lR) : P . f = a}.
4.9 Let AI,'" ,An be real numbers. Setf(xl, ... ,Xn) (1) Find generators for the left ideal J
(2) Show that An(R)j J is a holonomic module over An(R).
CHAPTER 11 CHARACTERISTIC VARIETIES In this chapter we give a geometrical interpretation of the dimension of an An-module. In order to do this we will have to use a number of results of algebraic geometry and linear symplectic geometry. All the algebraic geometry that we need can be found in [Hartshorne 77, Ch. 1]. Throughout the chapter we assume that the base field is C. 1. THE CHARACTERISTIC VARIETY. Let An be the 'Ilrth complex Weyl algebra and let M be a finitely generated left An-module with a good filtration
r.
Then grf M is a finitely generated
module over the polynomial ring Sn. Let ann(M, r) stand for the annihilator of grf M in Sn. Then ann(M, r) is an ideal of Sn. It depends not only on
M, but also on the choice of the good filtration
r;
see Exercise 4.1. The
radical of ann(M, r) however is independent of the filtration. 1.1 LEMMA. Let [} be another good filtration of M. Then
rad(ann(M, r)) PROOF:
= rad(ann(M, [})).
The ideals ann(M, r) and ann(M, [}) are homogeneous ( see Exer-
cise 8.4.6), hence so are their radicals (Exercise 4.2). Choose a homogeneous element
f
of degree s in rad(ann(M,r)). There exists dEB., the compo-
nent of degree ~ s of the Bernstein filtration, such that
f = used).
Since f E rad(ann(M, r)), there exists mEN such that f'n E ann(M, r). Hence,
dTnr ~ r m.+i-l, for every i ~ O. j
Iterating q times we get that
(1.2) On the other h.and, by Proposition 8.3.2 there exists k ~ 0 such that
for all i ~ D. Together with (1.2) for q
= 2k + 1, this leads to
r e re 'i+",-,(2k+l)-k-l _ .'i+ln1(2k+l)-1'
t"l C d rn (2k+l) d m (2k+l) Hi _ i+k _
t"l
11. Characteristic varieties
98 Thus d,n(2k+l)[}i ~
[}i+m8(2k+l)-b
and so rad(ann(M, r))
~
for all i ~ O. Hence f'n(2k+l) E ann(M, [}),
rad(ann(M, .a)). The opposite inclusion follows
by swapping rand [}. The ideal I(M)
=
rad(ann(M, r)) is called the characteristic ideal of
M. It follows from Lemma 1.1 that it is independent of the good filtration
r used to calculate it.
In other words, I(M) is an invariant of M. Thus, so
is the affine variety Ch(M)
=
Z(I(M)) ~ C 2n
that it determines. This variety is called the characteristic variety of M. Since I(M) is a homogeneous ideal ( see Exercise 4.2), the variety Ch(M) is homogeneous. This means that if p E Ch(M) and
~ E
C, then
~p E
Ch(M).
Note that Ch(M) is a subvariety of C2n, since Sn is a polynomial ring in 2n variables! Here is a simple example. Let dEAn be an element of degree r and put
M
= An/And.
If B" is the filtration of M induced by the Bernstein filtration
of An, then
as we saw in Ch. 7, §5. Therefore, ann(M, B")
= SnUT( d)
and so Ch(M)
=
Z(uT(d)) is a hypersurface.
1.3 PROPOSITION. Let M be a finitely genel'ated left An-module and N a
submodule of M. Then Ch(M) PROOF: Let
and
r"
r
= Ch(N) U Ch(M/N).
be a good filtration of M. It induces good filtrations
r
in N and MIN respectively. By Lemma 7.5.1, there exists an exact
sequence of finitely generated graded Sn-modules,
Clearly
ann(M, r) ~ anneN, r) n ann(M/N, r")
11. Characteristic varieties
99
and so Ch(N)UCh(M/N) ~ Ch(M), by [Hartshorne 77, Ch. 1, Proposition
1.2J. The opposite inclusion follows from ann(N, r)ann(M / N, r") £; ann(M, r). Before we proceed with the study of the characteristic variety, let us review some basic facts about the dimension of an affine variety. Let J be an ideal of Sn
=
C[YI, ... , YlnJ and put V
=
Z(J). If p is a point of V then the
Zariski tangent space of Vat p is the linear subspace of C2n defined by the equations 2n
of
L-(p)Yi=O lOY;
for all F E J. This space is denoted by 7;,( V)j it is a complex vector subspace of C2n. The dimension of V equals inf{ dime 7;,( V) : p E V}. Actually one need not look at every point of V. dim( V)
= dime 1;,( V).
If p is a non-singula1' point of V, then
This is equivalent to the definition in terms of heights
of prime ideals, see [Hartshorne 77, Ch. 1, Exercise 5.lOJ. We may also define 1;,( V) in terms of linear forms on linear form defined on the vector Y
= (YI, ... , Y2n) 2n
dpJ(Y) Consider the map d p : Sn
=
- t «((!2n)*
((!2n.
Let dpf be the
= dpf.
Let f, 9 E Sn
by
8f
L "UYi (P)Yi. I
defined by dp(f)
and), E C. Then dp satisfies dp(f
+ ),g) = dpf + )'dpg, dp(fg) = J(p)dpg + g(p)dP/.
We may rephrase the definition of Tz,( V) using d p , as follows: u E Tz,( V) if and only if dpF( u)
= 0, for every F E I( V).
This may be improved, by restricting
the elements of I( V) to a finite set. Suppose that FI, ... , Fm generate I( V). It follows from the properties of d p stated above that u E
if
Tz,( V) if and only
11. Characteristic varieties
100
The characterization of 7;,( V) in terms of dp will be used in the next section. The following theorem is an immediate consequence of the fact that if N is a graded module over Sn then the degree of its Hilbert polynomial is
dimZ(anns.N)j see [Hartshorne 77, Ch. 1 Theorem 7.5].
1.4
THEOREM.
dim Ch(M)
Let M be a finitely generated left module over An. Then
= d(M).
It is now very easy to show that if d
=I
0 is an operator in An, then
Ani And has dimension 2n - 1. Suppose that d has degree r
> O.
seen, the characteristic variety of Ani And has equation a"r{ d)
As we have
= O.
This is a
hypersurface of (:2nj hence its dimension is 2n-1. Thus d(AnIAnd) = 2n-1, by Theorem 1.4. Compare this example with Exercise 9.5.2. We may also use the characteristic variety to achieve a geometrical interpretation of Bernstein's inequality. This depends on results of symplectic geometry that we summarize in the next section. 2. SYMPLECTIC GEOMETRY.
A symplectic structure on (:2n is determined by a non-degenerate skew-
symmetric form. The standa1'd symplectic structure on (:2n is expressed by means of the matrix
where In is the nx n identity matrix. Given vectors u, v E (:2n, the symplectic form is
The matrix of any non-degenerate skew-symmetric form can be written in the form above for a suitable choice of coordinates [Cohn 84, §8.6, Theorem
1]. If U is a subspace of (:2n, then its skew-orthogonal complement is
UJ.
= {v E (:2n : w( 1.1, v) = 0 for
all
1.1
E
U}.
Note that it is not true that (:2n is the direct sum of a subspace and its skeworthogonal complement. For example, since w( 1.1,
1.1)
= 0 for
every u E (:2n,
11. Characteristic varieties
101
then everyone-dimensional space is contained in its complement. A subspace that is contained in its skew-symmetric complement is called isotmpic. However, since w is non-degenerate, it is always true that U and UJ. have complementary dimensions. The proof is short enough to be included here. First, some definitions. For wE c2n, define a linear form
--t
C by
cPw( v) = w( w, v). Let !l> : c 2n --t (c 2n)* be the linear map which associates cPw to w E C2n. 2.1
PROPOSITION.
Let w be a non-degenerate skew-symmetric form in
and U a subspace of c
2
n.
c2
n
Then:
(1) !l> is an isomorphism. (2) If a linear form
(3) dim U + dim UJ. PROOF:
cPw restricts to zero on U, then = 271.
w E UJ..
Note that the kernel of !l> is zero, because w is non-degenerate.
Hence, dim !l>(C 2 n) Since dim(C2n)*
= 271 -
dim ker(!l»
= 2n.
= 2n, it follows that !l> is surjective. Thus !l> is an isomor-
phism of vector spaces, and (1) is proved. Now let !l>lu :
c2
n
--t
U" be the map defined by !l>1U(w)
= rPwlv, the
restriction of cPw to U. Since every linear form on U extends to a linear form . on c2n, we have that !l>lv is surjective by (1). Hence the image of !l>lu is U". Now (2) is clear, and it implies that the kernel of !l>1 u is UJ.. Hence, 2n = dimker!l>1 v + dim!l>1 u(c 2 n) = dim UJ.
+ dim U*,
which implies (3). The subspaces that are important, from our point of view, are the ones that contain their skew-complement. They are called co-isotropic or involutive. A hyperplane is always an involutive subspace. Indeed, if H is a hyperplane of C2n that is not involutive, then there exists v ¢ H such that w(v, u) for all u E H. Since C2n
= H + Cv,
= 0,
this contradicts the non-degeneracy of
w. If V is an affine variety of C2n, then we will say that it is i71volutive if the tangent space 7;,( V) <; c2n is invollltive at every Iloll-singular point p E v:
11. Characteristic varieties
102
In particular, a hypersurface will always be involutive. The dimension of an involutive variety satisfies an inequality similar to Bernstein's. 2.2 PROPOSITION. The dimension of an involutive variety ofC 2n is greater
than or equal to n. PROOF: Let V be an involutive variety of
c2
n
and p a non-singular point
of V. By definition, we have that 7;.( V) 1. ~ 7;.( V). Hence dim 7;.( V)1.
5
dim 7;.( V). Thus, by Proposition 2.1,
2n = dim 7;.( V)1. and so dim 7;.( V) ~
1t.
+ dim 7;.( V) :5 2 dim 7;.( V),
Therefore, dim( V) ~ n.
The involutive varieties of dimension n are called lagrangian. Note that
if V is lagrangian in
c2n,
then 7;.( V)1. C 7;.( V) and both subspaces have
dimension n. Hence 7;.( V)
=
7;.( V)1.. In particular, V is also isotropic.
Thus lagrangian varieties can also be characterized as varieties which are involutive (co-isotropic) and isotropic. Talk of an affine variety, and one immediately thinks of its defining ideal. How can one decide whether a variety is involutive by looking at its ideal? The answer lies with the Poisson bracket of Sn. Let I denote the inverse ofthe map !l> defined above. The Poisson bracket of I, 9 E Sn is
An explicit calculation using coordinates shows that
{/,g}(p)
~ 81 8g = L..,{-(P)· -(p) I
8Yn+i
8Yi
8g
81
-(p). -(p)}. 8Yn+i 8y,
From this formula it is easy to see that {f, g} is a polynomial in Sn and that the map
{I, .} : Sn -+ Sn is a derivation of Sn.
An ideal J of Sn is closed for
the Poisson bracket if {f,g} E J whenever I,g E J. We want to show that a variety is involutive if its ideal is closed for the Poisson bracket. The proof will make use of a technical fact about the tangent space that we isolate ill a lemma.
11. Chamcteristic varieties
103
2.3 LEMMA. Let V be an affine variety ofC2n and p a point of V. If 8 is a form on C2n whose restriction to 1;.( V) is zero, then there exists
J E I( V)
such that 8 = dpJ. PROOF: Let F I , ... ,Fm be the generators of I( V). Set Y
= (Y1. ... .1l2n).
Then 1;.( V) is the solution set of the equations
in C2n . But 8( y)
=
0 is satisfied by the vectors of 1;.( V), by hypothesis.
Thus
where ai E C. Let I
= L:~ aiFi E I( V).
A straightforward calculation shows
that 8 = dp/. 2.4 PROPOSITION. An affine variety V of c2n is involutive if and only if its
ideal I( V) is closed for the Poisson bl·acket. PROOF: Suppose that V is involutive. Let p be a non-singular point of V. If I E I(V), then dpl restricted to 1;.(V) is zero. By Proposition 2.1(2),
I dp/ E 1;.( V).l.. Since V is involutive, then Idpl E 1;.( V). Now if 9 E I( V), then dpg is zero on 1;.( V), hence
{I, g}(p) =
-dpg(I dP/) =
o.
Since this identity holds for all non-singular points p of V, we conclude that the polynomial
{I, g} is in I( V), which is then closed for the Poisson bracket.
Conversely, assume that I( V) is closed for the Poisson bracket. Choose WE 1;.( V).l.. By Lemma 2.3, there exists 9 E I( V) such that
cPw = dpg. But
dpg( u) = w(I dpg, u). Since w is non-degenerate, w = I dpg. If I E I( V) then
Hence wE Tp( V). Thus Tp( V).l.
~
1;.( V).
We have seen that hypersurfaces are involutive, since their tangent spaces are hyperplanes. This is very
ca~y
t.o prove using Proposition 2.4, since the
11. Chamcteristic varieties
104
ideal of a hypersurface is principal, and the Poisson bracket of a polynomial with itself is always zero. The relation between characteristic varieties and symplectic geometry is the subject ofthe next theorem. The first proofs ofthis result were analytical. There is now a purely algebraic proof due to O. Gabber [Gabber 81]. We shall not include the proof here as it would take us too far from our intended course. 2.5 THEOREM. Let M be a finitely genel'ated left An-module. Then Ch(M) is involutive with respect to the standard symplectic stl'uctw'e ofC 2n . Equiv-
alently, I(M) is closed f01' the Poisson bracket. It is not obvious at first sight why this theorem should be difficult to prove.
In fact, it is easy to prove that if r is a good filtration of M, then ann(M, r) is closed for the Poisson bracket, The theorem does not follow easily from this fact because it is not true that the radical of a closed ideal is closed! Here is an example. Let J be the ideal of 8 1 generated by y'f, wand YI Y2. It is closed for the Poisson bracket, but its radical contains YI and Y2, and {YI, Y2}
= l.
Thus rad(J) is not closed for the Poisson bracket. Theorem 2.5
states the very subtle fact that the radical of the annihilator of the graded module of an An-module is closed, even though radicals of closed ideals arc)
not in general closed. Putting together Theorem 2.5 and Proposition 2.2 we get Bernstein's inequality. 2.6 COROLLARY. Let M be a finitely genel'ated left An-module. Then
d(M)
= dim Ch(M)
~
n.
It also follows from Theorem 2.5 that a holonomic module has a homoge-
neous lagrangian characteristic variety. A very spectacular use of the involutivity of the characteristic variety is discussed in the next section. 3. NON-HOLONOMIC IRREDUCIBLE MODULES
It is a very curious fact that, until 1985, it was widely believed that every irreducible An-module had to be holonomic. There was no good reason for
11. Characteristic varieties
105
this, except a lack of examples. The truth carne to light in [Stafford 85]. Stafford showed that if n
> 2 and A2 •...• An E C are algebraically indepen-
dent over Q, then the operator n S
n
= 01 + (~: A;XIX;O; + X;) + I)Xi 2
generates a maximal left ideal of An. Let M ducible and d(M)
= 2n -
Oi)
2
= Ani Ans.
Then M is irre-
1 > n, since n > 2. The fact that Ans is maximal
is proved by a long calculation, which can also be found in [Krause and Lenagan 85, Theorem 8.7].
In 1988, J. Bernstein and V. Lunts found a different and more geometric way of constructing irreducible modules of dimension 2n-1 over An. As we have seen in §§1 and 2, the characteristic variety of an An-module is always homogeneous and involutive. We will say that a homogeneous involutive variety of
c 2n is
minimal if it does not contain a proper involutive homo-
geneous subvariety. For example, an irreducible lagrangian variety must be minimal, since varieties of dimension less than n cannot be involutive. The work of Bernstein and Lunts depends on the following result. 3.1 PROPOSITION. Let d be an opel'ator of degree k in An and suppose that:
(1) the symbol uk(d) is ineducible in Sni (2) the hypersurface Z(uk(d» is minimal. Then the left ideal And is maximal. In particular, the quotient Ani And is an irreducible module of dimension 2n - lover An. PROOF: Suppose that
J is a left ideal of An such that And C J. Since these
are submodules of An we may consider their graded modules with respect to the filtrations induced by B; we get
The corresponding varieties are
o~ Z(gr(J»
C Z(uk(d).
11. Characteristic varieties
106
Note that the last inclusion is proper, because ITj,(ll) is irreducible. But
gr(An/J) ~ 8 n/ gr(J), and so ann(An/J, B) = gr(J). Therefore, Z(gr(J») = Ch(An/ J) is involutive by Theorem 2.5. Since Z(Uk(d) is minimal, then Z(gr(J)) = 0. Hence gr(J) = 8 n and so J = An. Thus And is a maximal left ideal of An. To put this proposition to good use we have only to construct minimal hypersurfaces in
[:2n.
That these hypersurfaces exist is the heart of the work
of Bernstein and Lunts. In fact they show that most hypersurfaces in
[:2n
are minimal. To make this into a precise statement we need a definition. Let 8 n (k) be the homogeneous component of degree k of 8 n . This is a finite dimensional complex vector space; so it makes sense to talk about hypersurfaces in 8 n (k). We say that a property P holds generically in 8 n (k) if the set
{f
8 n (k) : P does not hold for J}
E
is contained in the union of countably many hypersurfaces of 8 n (k). 3.2 THEOREM. The property 'Z(f) is minimal' holds generically in 8 n (k), ~
whenever k
4 and n
~
2.
TIns result was proved for n = 2 in [Bernstein and Lunts 88] and later generalized to n
~
2 in [Lunts 89]. The proof of Theorem 3.2 uses a lot
of algebraic geometry and also some results on differential equations, one of which goes back to Poincare's thesis! The proficient reader will find more details in the original papers. The gist of the work of Bernstein and Lunts is that 'most' irreducible Anmodules are not holonomic: almost the exact opposite of what was believed before 1985. 4. EXERCISES 4.1 Let M
= Ad A1x.
filtration defined by
ann(M, [})
Let
[}k
r
be the filtration of M induced by Band [} the
= Bk' (8 + A1x).
= 8 1 y'f + 8 1 Yl Yl.
Show that ann(M, r)
= 81 Yl
and
Compute their radicals.
4.2 Show that if J is a homogeneous ideal of a graded algebra R then Tad(J) is also homogeneous.
11. Chamcte7'istic varieties 4.3 Let J be a left ideal of An and put M
= An/J.
107 Show that Ch(M)
=
Z(gr(J».
4.4 Is the union of two involutive varieties involutive? What about their intersection? 4.5 Show that if V is an involutive homogeneous variety of C2 n then its irreducible components are also homogeneous and involutive. 4.6 Let s be the operator of An defined in §3. Let M = Ani Ans. (1) Compute Ch(M). (2) Is it an irreducible variety of c2n? (3) Is it a minimal hypersurface? 4.7 Let J be a left ideal of An. Show that if gr(J) is a prime ideal of Sn and Z(gr(J» is minimal, then J is a maximal ideal of An.
4.8 When is a hypersurface of C 2 n lagrangian? Give an example of a lagrangian variety of c2n, when n ~ 2. 4.9 A holonomic left An-module M is regular if there exists a filtration
r
for
M such that anns.grf M is a radical ideal of Sn. Let N be a submodule of . a regular holonomic module M. Show that N and MIN are also regular. 4.10 Show that if M is a regular holonomic module whose characteristic variety is irreducible, then M is an irreducible module. Why is the regular hypothesis required? 4.11 Let J be an ideal of Sn. Show that J2 is always closed for the Poisson bracket.
CHAPTER 12 TENSOR PRODUCTS All the operations on An-modules to be defined in the next chapters make use of the tensor product, which we are about to study. The construction of the tensor product presented in §2 usually seems artificial on a first encounter. Fear not; it is the neat universal property of §3 that is most often used in the applications. The final two sections before the exercises contain a number of results that will be required later. 1. BIMODULES.
To discuss tensor products in sufficient generality it is necessary to introduce bimodules. Let Rand S be rings and let M be an abelian group. To qualify as an R-S-bimodule, the group M must be a left R-module and a right S-module, and the R-action and the S-action must be compatible in the sense that if r E R, s E Sand u E M, then
r(us) If S
=
=
(ru)s.
R then we simply say that M is an R-bimodule.
A few examples and counter-examples will make the definition clear. If R is a ring then Rk is an R-bimodule. More generally, if R is a subring of S, then Sk is an S-R bimodule. In particular this applies to Weyl algebras; for if m < n, then Am is a subring of An. Thus A~ is an An-Am-bimodule. On the other hand, K[x] is a left AI-module and a right K[x]-module, but it is not an A1-K[x]-bimodule because the two actions are not compatible. Indeed, suppose that
f
E
K[x] is acted on the left by 8 and on the right by
x. Then 8(fx) = x8(f)
but (8. f)x
+f
= x8(f). More generally, K[X] is not an An-K[X]-bimodule.
Let M be an R-S-bimodule. A subgroup N of M is a sub-bimodule of M if it is stable under both the R-action and the S-action on M. In this
case, the quotient group !vII N has a natural ::;tructure of R-S-bimodule.
12. Tensor products
109
The following example will often come up in applications. Consider subring of
An+!
in the usual way. The left ideal
of the An+l-An-bimodule
An+!'
An+1Xn+l
An
as a
is a sub-bimodule
Thus the quotient
is an An+l-An-bimodule. We may similarly define a homomorphism of R-S-bimodules as a homomorphism of the underlying abelian groups which preserves both the left and right module structures. 2. TENSOR PRODUCTS.
Let R, Sand T be rings. Let M be an R-S-bimodule and let N be an ST-bimodule. We will define the tensor product of M and N over S, denoted by M®sN. First consider the set M x N of all pairs (u, v), with u EM, v EN. Let A be the free abelian group whose basis is formed by the elements of M x N. The elements of A are formal (finite) sums of the form (2.1)
with aj E Z,
La;(u"v;)
Ui
EM,
Vi
E N. Note that in this sum the pairs are mere
symbols: the sum is not an element of the direct sum M ffi N. In fact, if we assume that different indices correspond to different pairs in (2.1), then the sum is zero if and only if each
aj
= O.
r(u, v)
If r E Rand t E T, put
= (ru, v),
(u, v) t = (u, vt), where u EM and
V
EN. These are well-defined actions that make an R-T-
bimodule of A. Now consider the subgroup B of A generated by elements of the following types:
(U+ oJ, v) - (u, v) - (oJ, v), (u,
1)
+ V') -
(u., v) - (u., d),
12. Tensor prod1.lcts
110
(U8, v) - (u,sv), where r E R, s E S, t E T, u, 'Ii EM, v, 'If E N. One immediately checks that B is a sub-bimodule of A. The tensor product M @s N is the quotient AlB. The image of (u, v) in M @s N is denoted by 1.1 @ v. Since A has the pairs (1.1, v) as basis, the elements of the form u@ v generate M @s N. However, it is not true that every element of the tensor product is of the form u@ v. The elements 1.I@vsatisfyrelations which are determined by the generators of B, as follows (1.1 + J) @ v = 1.1 @ V
u@(v+v')
+ 1.1' @ v,
= u@v+u@v',
which are a sort of distributivitYi and also
U8@V= 1.I@SV, which means that the tensor product is balanced; finally
r(u@ v)
= ru@ v,
(1.1@ v)t= 1.I@vt, are, respectively, the left R-action and right T-action on M @s N. Note that if R is a K-algebra then a left R-module may be seen as an R-Kbimodule. This trivial observation must be kept in mind in the applications. For example, if M is a left R-module and N a right R-module, then N @R M is a K-vector space. We may also calculate M@l(N, which is an R-bimodule.
On the other hand, if R is commutative then an R-module is automatically an R-bimodule. In particular if M and N are R-modules and R is commutativc, then M
@R
N is an R-module.
Let us end this section by calculating a tensor product using the definition above. This is a simple example which, at the same time, carries a warning: the tensor product can behave in very unexpected ways. Consider the Zbimodules Zp and Zq, and assume that p and q are co-prime. We wish to calculate Zp 0z Zq. We know tllflt it is generated by elemcnts of the form
12. Tensor products
111
m 0 n, where m E Zp, n E Zq. But since p and q are co-prime, there exist integers a, b such that ap + bq = 1. Thus m0 n= (ap+ bq)m0 n= bq· m0 n
= O. Since the product is = bm0 qn. But qn = 0, Zp 0z Zq = o.
because pm
balanced, we deduce from the above
that m0 n
and so we end up with m® n = O.
Therefore,
3. THE UNIVERSAL PROPERTY The first property of the tensor product that we shall study is also the key to proving all the other ones. To make its statement more succinct it is convenient to introduce some terminology. For the rest of this section let R, Sand T be rings; let M be an R-S-bimodule and N an S- T-bimodule. Suppose that L is an R- T-bimodule and that ¢ : M x N
--t
L is a map. We
say that ¢ is bilinear if
¢(ru + u, v)
= r¢(u, v) + ¢( u, v), ¢( u, vt+ d) = ¢( u, v)t + ¢( u, d),
for all u, 11 EM, v, 'If EN, r E R, t E T. The map ¢ is said to be balanced, if
¢(us, v)
= ¢(u,sv)
for all s E S, u E M, v E N. The canonical example of a bilinear and balanced map is the projection 71' :
defined by 71'(u, v)
M x N
--t
M 0s N
= u® v.
Let U be an R- T-bimodule, and suppose that there is a bilinear and balanced map
'T} :
M x N
--t
U. We say that
'T}
is a universal bilinear
balanced map if given al1Y R- T-bimodule L and a bilinear and balanced map ¢ : M x N - - t L, there exists a unique R- T-bimodule homomorphism
;p: U --> L such that I/J = 1>. '1/.
12. Tensor products
112
3.1 THEOREM. The map
71' ;
M x N
M ®s N is a universal bilinear
--+
balanced map.
PROOF: First of all, we may extend ¢ to get a map 'IjJ : A 'IjJ(L: ai( U;, Vi»
=L
--+
L. In detail,
ai¢( Uj, Vi).
i
Since ¢ is bilinear and balanced, it follows that 'IjJ is zero on every element of the submodule 13. Thus 'IjJ induces an R-T-bimodule map
M ®s N
= AI13 --+ L.
This is the map that we call ¢. Since the canonical projection of A onto
AI13 is induced by
¢
now show that
M ®s N
--+
the equation ¢
71',
= ¢ . 71' follows
immediately. We must
is unique. Note that any bilinear and balanced map 8 :
L for which ¢ = 8· 71' satisfies 8(u® v) = ¢(u, v), for any u EM,
v EN. But this equation is enough to characterize ¢; hence 8 = ¢.
It follows from this theorem that the universal property uniquely characterizes the tensor product. 3.2 COROLLARY. Let
1] :
M x N --+ U be a universal bilinear balanced
map. Then U ~ M ®s N as R-T-bimodules. PROOF: Since N
--+
71'
is a universal bilinear balanced map, there exists 1i ; M ®s
= 17' 71'. On the other hand, since "I is also universal, U --+ M ®s N such that 71' = 7i' • 1]. Thus 1] = (1i' 7i') . 7)
U such that
there exists 7i' :
1]
. Using again the universality of 1], this time with respect to itself, and its uniqueness, we get that 1i . 7i' =
~,
where
~
; U
--+
U is the identity map.
Similarly, 7i"1i is the identity on M ®s N. We must now consider what happens with maps when we take tensor products. Let M' be an R-S-bimodule and N' be an S- T-bimodule. Let
M
--+
M' and 'IjJ : N
the map 9; M x N
--+
--+
N' be two bimodule homomorphisms. Consider
M' ®s N' defined by 9(u,v) = ¢(u) ® 'IjJ(v). It
follows from the properties of the tensor product that this map is bilinear and balanced. Hence, by the universal property of bilinear balanced maps, there exists a bimodule homomorphism: (j: M ®s N --+ M' 0s N'.
12. Tensor products
113
This map is usually denoted by cP@ 'I/J . Note that
(cP@ 'IjJ)(u@ v)
= cP(u) @ 'I/J(V).
4. BASIC PROPERTIES. We now use the universal property to prove some of the basic properties of the tensor product of modules. We illustrate the method with some examples, and leave the others as exercises for the reader. 4.1 PROPOSITION. Let R, S be rings, and M be an R-S-bimodule. Then
R@RM
~ M~
M@sS.
PROOF: Let 8 be the map from R x M to M defined by 8(r, u)
=
ru. It
follows from the definition of an R-module that B is bilinear and balanced. By the universal property of the tensor product there exists a bimodule homomorphism 9 from R@RM to M. This map satisfies the equation 9(r@
u)
= ru.
It is clear that
9 is onto, we show that it
is injective. Since we are
tensoring up over R, we have that r @ u = 1 @ ru. Thus every element of
R@RM can be written in the form l@u, for some u E M. But 9(1 @u)
= u,
hence '8 must be injective. The proof of the other isomorphism is analogous and is left to the reader. The property that we now prove is the distributivity of the tensor product with respect to the direct sum. It will be used often in later chapters. 4.2 THEOREM. Let R, S, T be l'ings, and I be an index set. Suppose that
M is an R-S-bimodule and that Ni are S-T-bimodules, for every i E I. Then
PROOF: The elements of EBi Ni will be denoted by every i E I. Consider the map
8: M x ffiN i
--+
EBi Vi,
ffi(M@sN i )
where
Vi
E N; for
12. Tensor products
114 defined by 9(U,EBiVi)
= EBi(u0 Vi).
It is a bilinear and balanced map. By
the universal property of the tensor product, there exists a bimodule map
8:M which satisfies 8(UI8>
0s
(EB; Vi»
EB N;
-+
EB(M 0s N
i)
EB;(U0 Vi).
=
On the other hand, let /-Ij : Nj
->
EBi Ni be the natural embedding.
Thus
we have maps
By the universal property of direct sums [Cohn 64, p.311j, there exists a unique map
such that /-I (EBi(U 0 Vi»
= u0 (EBiVi). One now checks that
/-I and Bare
inverse to each other by a simple calculation on generators; hence both maps are isomorphisms. A similar result holds when the direct sum is the first entry of the tensor product. Thus the following corollary is an immediate consequence of Theorem 4.2. 4.3 COROLLARY. Let R be a commutative ring and II, I2 be index sets. Let
Fi be free R-modules with bases {v~ : a E Ii}, for i is free with basis {v~ 18> ~ :
Q'
E
= 1,2.
Then H 0R F2
Ibf3 E I 2 }. Furthermore, if FI and F2 have
finite bases, then
Let us now state a few more properties of the tensor product: the proofs follow the general pattern of 4.1 and 4.2. 4.4 PROPOSITION. Let Ri be rings and Mi be R;-R;+l-bimodules, for i 1,2,3. Then
=
12. Tensor products
l15
4.5 PROPOSITION. Let R be a commutative ring and let M and N be R-
modules. Then
We must also consider the effect of the tensor product on exact sequences. The first result holds in great generality.
4.6
THEOREM.
Let R, Sand T be rings and let rjJ
tI>
M' --; M --; Mil --; 0
be an exact sequence of S-T-bimodules. If B is an R-S-bimodule, then the
sequence B 0s M'
1®1/> ---+
B 0s M
l®tI> ---t
B 0s Mil --; 0
is exact. PROOF:
Since r/J is surjective, the elements of B 0s Mil can be written in the
form
where bi E Band
11; E
M. Since this is equal to
we conclude that 1 0 r/J is surjective. On the other hand, (1 0 r/J) . (1 01jJ) equals zero, since ker(r/J)
= im(1/I).
= (10 r/J. 1jJ)
Hence
im(10 1/1) <; ker(1 0 r/J) and the proof will be complete if we show that the opposite inclusion holds. Let N
= im(l 0
1/1). Then r/J induces a map, (): (B ('0s M)jN --; B 08 Mil
12. Tensor products
116 defined by B(b 0 u + N)
=
b 0 cP(u). Note that () is well-defined because
N ~ ker(l 0 cP). Now let 7T be the projection 7T: B 0s M One easily checks that (). 7T
(B0s M)/N.
-+
= 1 0 cPo
Suppose we have shown that () is an
isomorphism, then ker(l0
= ker«(}. 7T) =
= N = im(l 0 1/J), by definition.
ker(7T).
Thus the proof will be complete
if we show that () is an isomorphism. We do this by explicitly constructing an inverse. Define a map, {3: B x M"
by (3(b, v)
(B0s M)/N,
-+
= b0u+N, where ¢(u) = v.
Assume that ¢(u)
= ¢(u) = v, where
We must check that it is well-defined. u, 'Ii E M. Then
u- U E ker(¢)
= im(1/J).
b0u=.b0U
(modN)
= 0 and so
But we have that
and {3 is well-defined. On the other hand, {3 is clearly bilinear and S-balanced. By the universal property ofthe tensor product there exists a homomorphism of R-T-bimodules,
73 : B 08 M"
-+
(B 0s M)/N.
A straightforward calculation shows that () and {3 are inverse to each other.
Of course there is a version of Theorem 4.6 with B tensoring on the right, and we shall use it whenever necessary without any further comment. The same applies to the next result.
12. Tensor products 4.7 COROLLARY.
117
Let S be a K-algebra and R a subalgebra of S. Suppose
that I <;;; S is an S-R-bimodule and that M is a right S-module. Then
M 0s S/I
~
M/MI
S
->
as right R-modules. PROOF: Tensoring the sequence
I
/I
->
S/I
->
0 with Mover S, and
using Theorem 4.6, we obtain the exact sequence
M0sI But M 0s S
~
I®/I
--+
M0s S
->
M0S/I
->
O.
M by Proposition 4.1. The image of the composition of 10 B
with this isomorphism is MI. Note that this is a right R-submodule of M. Thus, M0sS/I~M/MI
as right R-modules. An important warning: injective maps need not be preserved by tensor products. Consequently, tensor products do not always preserve short exact sequences. An example is given in Exercise 6.6. See also Exercise 6.5.
5.
LOCALIZATION.
It is time for our first application of the tensor product to the Weyl algebras. In it we generalize a result proved in Ch.10, §3. Let p be a non-zero polynomial in K[X] and consider the subset K[X, p-I] of all rational functions whose denominator is a power of p. We have proved that this is a holonomic left An-module in Theorem 10.3.2. Let M be a K[X]-module and put
Since K[X] is commutative, this is a K[XJ-module. Note that every element I
of M[p-I] can be written in the form p-k 0 u, for some u E M and k 2: 0, but this form is not
uIliq\l(~.
This posps the question: when is
p-k
0
U
= O?
12. Tensor products
118
5.1 PROPOSITION. Let M be a K[X]-module. Suppose that u EM satisfies
10 u = 0 in M[p-l]. Then there exists an integer k ~ 0 such that '1u = o. PROOF:
There exist an index set I and a surjective map IJ
K[XjI -tM, where K[XJZ denotes the direct sum of copies of K[X] indexed by I: a free
K[X]-module. Let F be the kernel of this map. Tensoring by K[X, p-l] over K[X], we get an exact sequence, (5.2)
By Theorem 4.2,
It is easy to check that the image of the first map of (5.2) in K[X, p-l]I is
K[X, p-l]F, the submodule generated by F. Therefore
as K[X]-modules. Choose v E K[XjI such that fJ(v) (1 0 fJ)(l 0 v)
= u.
Thus
= 1 0 u = O.
Identifying 10 v with v in K[X,p-IV, we have that v E K[X,p-l]F. Hence there exists k ~ 0 such that
'Iv E F;
from which we conclude that
'1u = 0
inM. Now assume that M is a left An-module. Then M[p-l] is a K[X]-modulc, and we want to turn it into a left An-module. We do that by defining the action of 8; as follows
fJ; . (p-k 0 u)
= p-k-I ® (pB; . u -
kfJ;(p)u).
It is not immediately clear that this formula gives a well-defined action,
because an element of l\IJ[p-l) does not have a unique representation in the
12. Tensor products
119
form p-k @ u. However, applying Proposition 5.1 one can show that it is indeed well-defined; the details are left to the reader. It is also easy to check that the relations of An are preserved by this action. Hence M[p-l] is a left
An-module by Appendix 1. It is called the localization of M at p. The name comes from the way this construction is used in algebraic geometry.
If we also assume that M has a good filtration
r, then we can construct a
filtration {} for M[p-l]. If m is the degree of p, put
(5.3) To check that this is a filtration of M[p-l], proceed as in the proof of Theorem 10.3.2.
5.4 THEOREM. If M is a holonomic An-module, then so is M[p-lj. PROOF: We have that
Let x( t, M, r) be the Hilbert polynomial of r. Then for k
»
0,
dimK (}k ~ x(k(m+ 1), M, r) a polynomial of degree n. Hence M[p-l] is holonomic by Lemma 10.3.1.
6. EXERCISES
6.1 Let p, q E Z, and let m be their greatest common divisor. Show that
Zp@z Zq
~
Zm.
6.2 Prove the following isomorphisms of Z-modules:
(1)
Q@zQ~Q.
(2) Q @z (Q/Z)
~ (Q/Z) @z (Q/Z).
6.3 Let R be a ring. ShaWl that Rn @R R1n
~
6.4 Prove Propositions 4.4 and 4.5 in detail.
Rmn.
120
12. Tensor products
6.5 Let R be a K-algebra and let k
~
0 be an integer. Show that if N
--t
M
is an injective map of left R-modules, then so is
6.6 Let cP : Z2 by cP(l
--t
Z4 be the injective homomorphism of Z-modules defined
+ 2Z) = 2 + 4Z.
Show that:
(1) Z2 0z Z2 ~ Z2. (2) Z2 0z Z4 ~ Z2. (3) The map Z2
--t
Z2 induced by tensoring cP by Z2 over Z is identically
zero. Conclude that the tensor product by Z2 does not preserve injectivity. 6.7 Show that for any non-zero p E K[X], there exist An-module isomorphisms:
(1) K[X,p-l]0J\'[xj K[X] ~ K[X, P-l]i
(2) K[X, p-l]0I<[Xj K[BJ, . .. , on]
= O.
CHAPTER 13 EXTERNAL PRODUCTS In this chapter we discuss the simplest operation to be performed on modules over the Weyl algebra: the external product. We also discuss the computation of the dimensions of the external product from the dimension of its factors. In particular, we will show that the external product of holonomic modules is itself holonomic.
1. EXTERNAL PRODUCTS OF ALGEBRAS. We begin with a general definition. Let A,B be K -algebras. The tensor product A 0]( B is a K-vector space, on which we define a multiplication. For a, a' E A and b, 1J E B, let
(a 0 b)(a' 01J) = aa' 0 blJ. It is routine to check that A 0]( B with this product is a K-algebra. This is called the external product of A and B. Since we use this construction very often, it is convenient to have a special notation for it: A®B. If we apply the construction to polynomial rings or to the Weyl algebra' we do not get anything new: that is the secret of its power. The key result is the following theorem. 1.1 THEOREM. Let R be a K-algebl·a and let A and B be subalgebras of R.
Suppose that (1) R = AB,
(2) [A, B] = 0, (3) thel·e exist K-bases {a; : i E N} and {bJ, : i E N} of A and B, respectively, such that {aibj : i, i E N} is a K -basis of R. Then R
8;:
A®B.
PROOF: Define a map r/J ; ~ x B
->
R by r/J( a, b) == abo It is clearly K -bilinear
and K-balanced. Therefore, there exists a map rp; A 01( B sli(a @ b) = abo This is called the 'I1'w.1Uplic.ation map.
->
R such that
13. External products
122
Let a, a! E A and b, If E B. Then
!Ii«a 18) b)(a'
@
IJ» = !Ii(aa'
@
bY) = aa'bli.
On the other hand,
!Ii(a@ b)!li(a' Thus (2) implies that
iJj
@
Ii)
= aba'Y.
is a K-algebra homomorphism.
Since, from (1), every element of R is a linear combination of monomials
ab, with a E A and bE B, we conclude that !Ii is surjective. The injectivity is a consequence of (3) and Corollary 12.4.3. A carefully chosen notation makes the application of Theorem 1.1 to polynomial rings and the Weyl algebra almost tautological. Our choice is the following. Let K[X]
= K[Xb ... ,x n] and
K[Y]
= K[Yl, ... , Ym]
be polyno-
mial rings. Write K[X, Y] for the polynomial ring on the x's and y's. Then
An will be the Weyl algebra generated by the x's and ax's, and Am the Weyl algebra generated by the y's and ay's. Both are subalgebras of Am+n> the Weyl algebra generated by the x's, y's and their derivatives. We shall retain this notation for the rest of the chapter. 1.2 COROLLARY. The following isomorphisms are induced by the multiplication map:
(1) K[X]®K[Y] ~ K[X, Y]. (2) Am®An ~ Am+n' The isomorphisms defined in the corollary are so natural that we shall take them to be equalities. Thus, from now on, we shall use without further comment that Am®An = Am+n' In other words if a E An and bEAm, we will identify the monomial ab with a @ b. Similarly for polynomial rings. 2.
EXTERNAL PRODUCTS OF MODULES.
Once again we turn to the general situation. Let A,B be K -algebras. Suppose that M is a left A-module and that N is a left B-module. Then we may turn the K-vector space M
18)K
N into a left (A®B)-module. The
action of a 18) b E Ai§B on u@ v EM 18)K N is given by the formula (a@b)(u0 'u) = atL@bv.
13. External products
It is routine to check that M
(9K
123
N is a module for this action. We shall
write M®N for this (A®B)-module. 2.1 PROPOSITION. If M and N are finitely generated modules, then so is M®N. PROOF:
Suppose that M is generated by Ut, ... , Us over A and that N is
generated over B by 'Ut, •.• , Vt. The elements of M®N are K-linear combinations of elements of the form u (9 v, for u EM, v E N. But u and v =
Ei bj Vj.
= E~ aiUj
Thus, u(9 v
= ~)a; (9 bj)(Uj (9 Vj). iJ
Hence M®N is generated over A®B by Uj (9 Vj, for 1 SiS sand 1 S j S t. Now let M be a left Am-module and N a left An-module. The external product M0N is a left Am+ n-module, under the convention that Am+n
=
Am0An. This will be used so often, that even at the risk of being pedantic,
we write the action explicitly. Let u(9 v E M®N. A monomial of Am+n can be written, in only one way, in the form ab with a E Am and bEAn. Thus, (ab)(u(9 v)
= au (9 bv.
The situation for polynomial rings is entirely similar. The next lemma will be very useful in calculations. We make use of these conventions in its statement and proof. 2.2 LEMMA. Let I be a left ideal of Am and J a left ideal of An. Denote by Am+nI + Am+nJ the left ideal of Am+n genel'ated by the elements of I and J. Then
as Am+n-modules. PROOF:
Since every monomial in Am+n may be written, uniquely, in the
form ab, for a E Am, bEAn, there exists a K-linear map,
13. External products
124 given by 'IjJ(ab)
=
(a + I) (/9 (b+ J). An easy calculation shows that this is an
Am+n-module homomorphism. It is clearly surjective. On the other hand, a monomial ab with a E lor bE J, gives 'IjJ(ab)
=
O. Hence Am+nI + Am+nJ ~
ker'IjJ . By the universal property of the tensor product, there is a K -linear map,
defined by c/J((a+ 1) (/9 (b+ J»
= ab+ Am+nI + Am+nJ.
This is a surjective
map, and
Hence Am+nI + Am+nJ
= ker 'IjJ and
'IjJ is an isomorphism of Am+n-modules.
Although we have not used in the proof of 2.2 that c/J is a homomorphism of Am+n-modules, this is easily shown to be true.
2.3
COROLLARY.
Let I be a left Am-module. Then
is an isomOl'phism of Am+n-An-bimodules. PROOF:
It follows from Lemma 2.2 that there exists an isomorphism of left
Am+n-modules,
given by 'IjJ(ab + Am+nI) = (a since I
~
+ I) (/9 b,
where a E Am, bEAn. Note that
Am, the elements of I commute with the elements of An. Hence
Am+nl Am+nI is a right An-module. So is (Ami I)0An, with An acting on the right component of the tensor product. Let c E An. Then
tf;((ab)c + A m+n1)
= (a + 1) (/9 bc = tf;(ab + A'n+n1)c.
Therefore, 'IjJ is also a right An-module isomorphism.
3.
GRADUATIONS AND FILTRATIONS.
We will now construct a good filtration for the external product of modules over the Weyl algebra and calculate its associated graded module. Throughout this section, let M be a finitely generated left Am-module with good
13. External products
filtration reM)
= {r;(M) :
125
i E N} and let N be a finitely generated left
An-module with good filtration rCN). We will proceed through a series of lemmas. 3.1 LEMMA. The Bernstein filtration of Am+n satisfies
L
BtCA m+n) =
Bp(Am)Bq(An).
p+q=t PROOF:
First note that
Now every monomial in Bt(Am+n) may be written as a product ab, with a E
Am and bEAn. Assume that a, b have degrees p, q, respectively. Since ab has degree p+q, it follows that ab E Bp(Am)Bq(An). But an element of Bt(Am+n) is a sum of such monomials. Therefore, Bt(Am+n) 3.2 LEMMA.
= L.:p+q=tBp(Arn)Bq(An).
The K -vector spaces n(M0N)
=
L r;(M) \Of{ rj(N)
;+J=k
form a good filtration of M0N as an Am+n-module. Note that we are identifying ri(M) \OK Tj(N) with a subspace of M0N, in the obvious way. PROOF:
Since the summation in the definition of n(M0N) is finite, we have
by Corollary 12.4.3 that nCM0N) is a finite dimensional K-vector space. Now every element of M0N is a finite linear combination of elements of the form u\O v, with u EM, v E N. Assume that u E T;CM) and v E Tj(N). Then
u\O v E Ti(M) \0 Tj(N) ~ ri+j(M0N). Hence, M0N = Uk~O Tk(M0N). Finally, since
13. External products
126
we conclude, using Lemma 3.1, that B t(A",+n)rk(M'0N)
= n+t for t» o.
Thus n(M'0N) is a good filtration of M'0N. We will now calculate the graded module with respect to this good filtration. The ring Sn
= g".8 An is a polynomial ring by Theorem 7.3.1.
Using the
conventions of §1 we may identify Sm'0Sn with Sm+n' Denote the homogeneous component of degree t of Sn by Sn(t). By an argument similar to that used in Lemma 3.1, we have that Sn+",(t) = EBp-tq=tSp(Am)Sq(An). Finally, write gr,.(M)
= rk(M)/ n-l (M).
3.3 LEMMA. Thel'e is an isomOl'phism of K -vector spaces
grk(M'0N)
9!
EB gri(M) 0]( gTj(N). (+j=k
PROOF: Consider first the K-linear map,
defined by 1)ij(u0v)
= J1.i(U) 0J1.j (v), where we are denoting the symbol maps
of both M and N by J1.. It is easy to see that 1);j is surjective and that its kernel is the subspace
Putting these maps together, we get a linear map
i+j=k
i+j=k
Note that the canonical projection gives a map
E9 (r;(M) 0 rj(N» i+j=k But ker 1)ij ~ ri+j-l (M'0N), thus
--) n(M'0N).
127
13. External products factors through
n(M0N)/Fk_l(M0N) --;
EB (grj(M) ®grj(N». i+j=k
Since the former map is bijective, so is the latter, which gives the required isomorphism. Taking direct sums for k
~
0, and using Lemma 3.3, we conclude that
there exists an isomorphism of K-vector spaces,
B : gr(M0N) --; gr(M)09r(N). Since M0N is an Am+n-module, gr(M0N) is an Sm+n-module. On the other hand, since Sm+n
= Sm0Sn, it follows that gr(M)0gr(N) is an Sm+n-module.
3.4 THEOREM. The lineal' map B is an isomorphism of Sm+n-modules. PROOF: We have only to check that B is compatible with the action of Sm+n. Choose a monomial of Srn+n, and write it ill the form
f g, with f E Sm(P),
9 E Sn(q). There exist operators a, b E Am+no such that
f
= O"p(a) and 9 = O"q(b).
In particular, a E Am and bEAn. Now let i
+ ;j = k.
U
® v E rj(M) ® Tj(N), with
Then
(jg)J.'k(u0 v) = J.'k+p+q«ab)(u0 v». But (ab)(u0 v)
= au0 bv; therefore
The latter is mapped onto J.'i+p(au) 0 J.'j+q(bv) by 8. Since this may be rewritten as (fg)(p,j(u) ®/-lj(v», we have
which completes the proof. 4. DIMENSIONS
AND MULTIPLICITIES
(
We are now ready to calculate the dimension and multiplicity of an external product.
128
13. External products
4.1 THEOREM. Let M be a finitely generated left Am-module and N a finitely generated left An-module. Then
(1) d(M@N)
= d(M) + d(N),
(2) m(M@N) ~ m(M)m(N). PROOF: We retain the notation of §3. It follows from Lemma 3.3 that k
dimKr,.(M@N)
=L
L dimKgri(M)dimKgrj(N).
r=O i+j=r
Therefore, dim](n(M@N) ~
k
k
;=0
;"=0
L dim](gri(M) L dimKgrj(N).
From this inequality and Lemma 3.2 we have that
Assuming that k
»
0, and using the Hilbert polynomials of the corresponding
filtrations we get that X(k, M@N) ~ X(k, M)X(k, N) ~ X(2k, M@N).
Since this holds for all large values of k, both (1) and (2) immediately follow. The corollary is an easy consequence of the theorem.
4.2 COROLLARY. Let M be a llOlonomic Am-module and N a holonomic An-module. Then M@N is a holonomic Am+n-module.
5. EXERCISES 5.1 Let Bn be the ring of differential operators of the rational function field
K(X)
= K(xJ,""
xn). Is it true that Bn@B m ~ Bn+m as K-algebras?
5.2 Show that the multiplication map induces an isomorphism of Am+nmodules, K[X,
11
~ K[X)&K[11.
13. External p1'oducts
129
5.3 Let p E K[X] and q E K[ Y]. Show that there exists an isomorphism of Am+n-modules,
5.4 Let
(7, (7'
be automorphisms of Am and An, respectively. Let M be a left
Am-module and N be a left An-module. (1) Show that there exists a unique automorphism 8 of Am+n whose restriction to Am is
0'
and whose restriction to An is
0".
(2) Show that
5.5 Let M be a left Am-module and N a left An-module. Show that the Fourier Thansform satisfies
5.6 Let M be a holonomic Am-module and N a holonomic An-module. Show that if M and N have multiplicity 1, then M0N is a simple holonomic Am+n-module. 5.7 In this exercise we use the results of eh. 11. Let M be a left Am-module and N a left An-module. Show that Ch(M0N) 5.8 Let 1
sis n.
= Ch(M) x Ch(N).
Suppose that d; is an operator of An of degree
is a linear combination of monomials in
Xi
~
1 which
and 8;. Let J be the left ideal
generated by d), . .. ,dn. Show that Ani J is a holonomic module over An. Hint: An/J ~ (AdA 1d1 )0 ... 0(AJ/A 1 dn).
CHAPTER 14 INVERSE IMAGES In this chapter we show how to change the ring of scalars of a modulll,
over a Weyl algebra. This is very important and leads to the construction of many new modules. As we shall see later, holonomic modules are preserved under this construction. 1. CHANGE OF RINGS
We start with a general construction for rings and modules. Let R, S be rings, and let <jJ : R
-+
S be a ring homomorphism. If M is a left R-module,
then we may use ¢ to turn it into a left S-module. This is called changing the base ring. The key to the construction is that S may be considered as
a right R-module in the following way. Let r E R, s E S; the right action of R on S is defined by S*
r
= s<jJ(r).
The juxtaposition on the right hand side of the equation denotes multiplication in the ring S. With this in mind, we may consider S as an S-R-bimodule. We are now allowed to t!lke the tensor product S
@R
M which is a left S-module. Thus
starting with the left R-module M we have constructed the left S-module S
@R
M. If it is necessary to call attention to the homomorphism ¢, we will
write S@¢M. As an application, let us rephrase the twisting construction of Oh. 5, §2 in terms of change of rings. 1.1 LEMMA. Let R be a ring and u an automorphism of R. Let M be a left
R-module. Then
PROOF:
Let a E R. The left action on M" is given by a. u
= u(a)u, where
u EM,,; and the right action on R by s*a = su- (a), where s E R. Consider 1
the map rjJ:RxM-M"
14. Inverse images defined by
= a _ u.
Since
131
= a_ bu = (u(a)b)u
=
is also bilin-
ear. Thus, by the universal property of the tensor product, there exists a homomorphism
Since M"
=M
as abelian groups, the map
¢ is surjective.
Let v E R@,,-l M. There exists u EM such that v = 1 @ u. Hence
O=¢(v)=u, which implies that u
= O. Thus v = 0 and ¢ is injective. Therefore, ¢ is an
isomorphism of R- modules. We shall now apply the change of rings construction to polynomial maps. We often use the results of Ch. 4, §1 in the coming chapters, so you may wish to read that section again before you proceed. A good notation will be of great help: we shall abide by the conventions we are about to make right to the end of the book.
1.2
NOTATION.
Put X
=
Kn. The polynomial ring K[XI, ... ,xnl will be.
denoted by K [X]; and the Weyl algebra genel'ated by tile x's and
a" 's
by
An. The n-tuple (Xl, ... ,x n) will be denoted by X. Similar conventions will hold for Y
= Km
and Z
= K r,
with polynomial rings K[ 11 and K[Z] and
Weyl algebras Am and A r . The space Km+n equals the cartesian product of X and Y and will be denoted by X x Y. FOl' the polynomial l'ing in the x's and y's we shall write K[X,
11.
The (n+ m)-th Weyl algebra Am+n will
always denote the algebra genel'ated by K[X,
11, a", 's and Oy's.
Note that
with these conventions Am+n = Am0An. Although these conventions may cause some initial confusion, they payoff later on by making the formulae easier to digest than they would otherwise be. I
Let F: X - t Ybe a polynomial map. Its comorphism FI : K[11-> K[X] is an algebra homomorphism. If M is a K[ YJ-module, we may use FI to
14. Inve1'se images
132
construct the K[X]-module K[X]0[([y) M. It is called the inverse image of
M by the polynomial map F, and denoted by F* M. Let us consider some examples. Let F : X
--l
X be a polynomial isomorphism, with inverse C. The
comorphism Fd is an automorphism of K[X] with inverse Cd. Let M be a
K[X]-mod ule. By Lemma 1.1, its inverse image F* M is isomorphic to MG'. Consequently, if u E F* M and hE K[X], then hu = C~(h)u. The second example is the projection 7r : X x Y coordinate: 7r(X, y)
--l
Yonto the second
= Y The comorphism 7rd maps a polynomial p E K[ Y]
to itself, but considered as an element of K[X, Y]. Let M be a module over
K[n then 7r* M
= K(X, Y] 0[([11 M.
A monomial in K[X, Y] may be written in the form pq, with p E K[X] and q E K[Y]. If u EM then
pq0u= p@qu as elements of 7r* M. Using this identity we may construct an isomorphism, 7r* M ~ K[X]0M, of K[X, Y]-modules. 2. INVERSE IMAGES. Let F : X
Y
= Km,
--+
Y be a polynomial map and M a left Am-module. Since
it follows that M is in particular a K[ Y]-module. Thus we may
compute the inverse image of M by F defined in §l:
F*(M)
=
K[X]0[([y) M
is a module over K[X]. We want to make this module into an An-module. At first sight this construction may seem very puzzling. Why do change of rings as if we had only a module over a polynomial ring and then turn it back into a module over a Weyl algebra? Why not do the change of rings at the Weyl algebra level? First of all, one is allowed to do change of rings at the level ofWcyl
alg{~bras,
since t.he construction works for any ring. However, in
14. Inverse images
133
doing that one is limited by the fact that Weyl algebras are simple rings; and homomorphisms of simple rings are necessarily injective. A more satisfactory answer is that this is really a construction in algebraic geometry, that we are lifting to the realm of noncommutat ive rings; see [Hartshorne 77, Ch. 2, §5]. Anyhow, the efficacy of the construction is its best justification, as we shall see later on. We know how a polynomial of K[X] acts on F*(M). We now give the recipe for the action of 0"". Let
q@uE F*(M), where q E K[X] and u EM. Let F l , ... , Fm be the coordinate functions of F. The action of ox; is defined by m
(2.1)
ox;(q@ u)
= ox.(q) @ u+ L qox, (Fk) @ oYk u k=l
for i
= I, ... , n.
The x's and ox's generate An, and we know how they act on F*(M). According to Appendix I, these formulae will make an An-module of F*(M) if they are compatible with the relations satisfied by the generators. The relations are
[O"'i,Xi] = 8ij l, [Xi,Xj]
= [OX"OXi] = 0,
for 1 S i, j S n. Let us carefully check that the first of these relations is compatible with the actiolls of x's and ox's defined above. Let wE F*(M). We want to show that
Clearly it is enough to check this when w is of the form q@ u, for q E K[X] and u E M. Let us begin by calculating ox) (Xi(q applying (2.1) to Xjq @ u, which gives
@
u)). We do this by
m
(2.2)
8xj (Xjq)
@
u+
L xiq8
xi
(Fk) ® Oy, u.
14. Inverse images
134
Using Leibniz's rule and the fact that OXj(Xi) = 6i:J" we may rewrite (2.2) in the form
Note that the expression in brackets equals OXj(q 0 u). Thus we have shown that
But this is equivalent to
which is what we wanted to prove. The proof that the other relations are compatible with these actions is an exercise for the reader. We will calculate some concrete examples in the next section. 3. PROJECTIONS.
Let us return to the projection 7r : X x Y --; Yof §l. Let M be a left Am-module. We have already seen that, as a K[X, l1-module, the inverse image 7r* M is isomorphic to K[X]0M. However, K[X] is an An-module and
M is an Am-module. So K[X]0M has a natural Am+n-module structure. We want to show that this structure coincides with the one determined by (2.1).
It is enough to check how
ox; and oy,
act on K[X]0M. For this we return
to the definitions. The K[X, l1-module isomorphism between 7r*(M} and K[X]0M maps qaxO 0 U to x a 0 qou, where a E Nn and qa E K[n We must show that this is compatible with the action of the derivatives. Consider first the action of ow' as defined by (2.1), m
Ow (qaxO0 u) = Ow (qa x O) 0 u+ L(OW(Yk)qa x O0 o!J>u}. I
14. Inverse images
135
Using Leibniz's rule, this may be rewritten as
ay; (q"x" 18> u) = x" 18> ay; (q"u). Summing up: if we identify n*(M) with K[X]0M, then
a!h
acts only on M,
in the usual way. Consider now the action of ax. . From the definition, we have that m
ax;(q"x'" 18> u)
= ax;(q"x"') 18> u + E ax;(Yk)q"x'" 18> a
Zk
u.
1
Since az;(Yk)
= ox,(q,,) = 0,
for all k and all ex, we end up with
ax;(q"x'" 18> u) = q",ox;(x") 18> u. Thus, under the previous identification, Ox, acts only on K[X]. We may sum up our calculations as follows. If the module 7r*(M) is identified with K[X]0M then the x's and o,,'s act only on the first factor K[XJ, and the y's and ay's act only on the second factor M. Besides, the actions are the natural ones. We may use this description to calculate the dimension of an inverse image. 3.1 THEOREM. Let M be a finitely generated left Am-module and 7r the
projection defined above. Then:
(1) 7r* M is a finitely generated A'n+n-module. (2) d(7r*M)=n+d(M). (3) m(7r*M)Sm(M). PROOF: (1) follows from Proposition 13.2.1; whilst (2) and (3) are consequences of Theorem 13.4.1 since, as an An-module, K[X] has dimension n and multiplicity 1; see Ch. 9, §2. 3.2 COROLLARY. Let M be a holonomic Am-module. Then 7r* M is a holo-
nomic Am+n-module. 4. EXERCISES 4.1 Let R, S, T be rings and cp: R
-+
Sand 'IjJ: S
phisms. If M is a left R-n\'odu]e. show that
--+
T be ring homomor-
136
14. Inverse images
4.2 Let F : K
---t
K be the polynomial map defined by F(x)
= x2•
Let M
be the algebra Al considered as a left module over itself. Show that F* M is not finitely generated over AI. 4.3 Let u be an automorphism of An. Suppose that the restriction B of (T to K[X] is an automorphism of this ring. Put F
=
B~
for the polynomial map
determined by B. Let M be a left An-module. Is it true that
as An-modules?
4.4 Let M and N be left An-modules. In particular these are modules over K[X). Hence the tensor product M 0K[X] N is a well-defined K[X]-module.
Define the action of 8x ; on u0 v E M 0J([x] N by the formula
8x ;(u0 v)
= 8x ;u0 v+ u0 8x ;v.
(1) Proceeding as in §2, show that this action makes M
01qX]
N into an
An-module.
(2) Give an example of two finitely generated An-modules M and N such that M 0K[X] N is not finitely generated over An.
4.5 Let F : C
---t
C be the map defined by y = F(x) = x 7n , where m
~
2 is
an integer. Let 6 be the Dirac microfunction; see Ch. 6, §3. The purpose of this exercise is to show that the inverse image F*(AI «(;)6) is isomorphic to the Al (C)-module generated by
(1) Show that x 7n and x8"
6'n, the
+m
m-th derivative of 6.
annihilate 1 06.
(2) Show by induction that there exist non-zero complex numbers cpq such that
for q = 0, 1, ... , m. (3) Using (2) show that
1 ® 8;;+16
= 2-8~rH)m(1 ® 6). c1nn
14. Inverse images
137
(4) Conclude from (3) that 1 08 generates F"(A l (C)8). (5) By (1), (3) and Exercise 6.4.10, F"(Al(C)8) is a homomorphic image of Al (C)8 m . Since the latter is irreducible, we have the desired isomorphism.
In the language of microfunctions we have proved that the inverse image of 8 under F is 8m .
CHAPTER 15
EMBED DINGS In the previous chapter we defined the inverse image of a module and calculated a formula for its dimension and multiplicity under projections. In this chapter we turn to embeddings. As we shall see, the behaviour of the inverse image under embeddings is a lot less regular than under projections. 1. THE STANDARD EMBEDDING
In this chapter we shall retain the notation of 14.1.2. We begin by considering the embedding
£ :
X
~
X x Y defined by LeX)
= (X, 0), where 0
denotes the origin of Y. We will call £ the standard embedding. The comorphism £1 : K[X, J1
---t
K[X) is defined by LI(g(X, Y))
= g(X, 0).
It may be
used to make K[X] into a K[X, J1-modulej as such, K[X] is generated by 1, since
g(X, Y) ·1
= £~(g(X,
Y)) . 1 = g(X, 0).
The annihilator of 1 in K[X, J1 is the ideal generated by Yt, ... , Yrn' Denoting this ideal by (Y) we have that K[X] ~ K[X, J1/(Y) as K[X, J1-modules. ~
Now consider a left Am+n-module M. Since An
A m+n, the Am+n-module
M is also an An-module. The elements of An commute with the variables Yt, ... , Yrn, thus (Y)M is an An-submodule of M. Hence M / ( Y)M is an An-
module. Note however that it is not an Am+n-submodule.We want to show that
£7M
= K[X] @K[X,y) M
~
M/(Y)M
as An-modules.
If u EM, let
u denote its image in M / ( Y)M.
Define the map
¢: £*M---tM/(Y)M' by ¢(q@u) = qu. It is a homomorphism of K[X]-modules. We want to show that the action of the derivations OXI' ... , that
oz. is compatible with ¢.
8q ~ OXk. 8z ,(q@u)=-8@u+L..,[email protected]. Xi
k=l
Xi
We have
139
15. Embeddings and so
8q 8x ,.(q® u) = 8x; ® u+ q® 8Xi u. The right hand side is mapped by r/J onto
8q_ -8 u+q8zi u, Xi
which is equal to
8Xi (qu)
= 8x ,.(r/J(q ® u)).
Thus r/J is an isomorphism of An-modules. Summing up: ~*M ~ Mj(Y)M as An-modules. Since M/(y)M is a quotient of An-modules, the action of x's and 8's is the natural one. Let us specialize this construction to the standard embedding ~ : K
--l
K2,
which takes x to (x,O). Let M be the ring A2 itself, considered as a left A2module. Applying the above results, we have that
~* M
taken as a left AI-module. Note that although M
= A2 is a finitely generated
left A2-module, the module
L* M
is the quotient A2/ yA2
is not finitely generated. In fact,
~* M
is a
free AI-module whose basis is the image of the powers of 8y in A2/ yA 2. Since this basis is infinite,
~* M
is not a finitely generated left AI-module.
We conclude from this example that the inverse image under embed dings does not preserve the noetherian property. This is very important, since we have defined dimensions only for noetherian modules. In particular there can be no general formula for the dimension of an inverse image under an embedding. Despite this, holonomic modules are preserved by inverse images under embeddings. This mystifying fact will have to wait until Oh. 18 for an explanation. It will be necessary, for future purposes, to consider a more general kind
of embedding, defined as a composition of the standard embedding and a polynomial isomorphism. Before we do this we must study the behaviour of the inverse image under composition of polynomial maps. 2.
COMPOSITION.
Let us first consider what happens to the change of rings construction of Cll. 14, §1 under composition. Let R, Sand T be rings and
->
S
15. Embeddings
140
and 1/J : S
---+ T
be ring homomorphisms. Let M be a left R-module. We
may turn M into a T-module in two different ways. On the one hand, we may apply the change of rings construction twice to get T®", S ®¢ M. On the other hand, the ring T is a right R-module via the homomorphism 'l/JcP , therefore T0",¢ M is a T-module, where
t® ru = t'I/JcP(r) ® u for t E T, r E Rand u E M. Using Propositions 4.1 and 4.4 of
eh. 12, we
conclude that these T-modules are isomorphic; thus,
It is not difficult to give a direct proof of this isomorphism. The T-module
T 0", S 0¢ M is generated by elements of the form t 0 s ®
U,
where t E T,
s E Sand u E M. But
to s 0 u = t'I/J(s) ® 10 u. Define a T-module homomorphism
by O(t0 s ® u)
= t'IjJ(s) 0
u. Note that if r E R then the element
t ® 1 0 ru
= t1/JcfJ( r) ® 1 0
u
is mapped to t'IjJcP( r) 0 u. It is easy to construct an explicit inverse for 0; hence it is an isomorphism. We will now apply this to An-modules. Recall that by Theorem 4.1.1, if
F: X
---+
Yand G: Y ---+ Z are polynomial maps, then (GF)~
= FaQII.
Note
the change in the order of the maps. 2.1
THEOREM.
Let F : X
---+
Yand G: Y
---+
M be an Ar-module. TlJeIl F'G*M
~
(GF)*M
Z be polynomial maps and
15. Embeddings
141
as An-modules.
By definition,
PROOF:
(CF)* M = K[X]0K[Z] M.
In this formula K[X] is a right K[Z]-module. Recall that the action of
f
E K[Z] on hE K[X] is given by h(J . C· F). On the other hand, F*G*M
= K[X]0ICfY) K[t]0K[Z] M.
We have already seen that these two modules are isomorphic as K[X]modules. The isomorphism B is defined by B(h 0 g 0 u)
= h(g . F) 0
u.
We must check that B is compatible with the action of the derivatives
ax.
By
definition,
az,(h 0 g0
u)
=
ah 0 (g 0 u) + hL (aF a-: ~ 0 a!/j(g 0 u) ) In
j=1
X,
X,
a
where F I , ... , Fm are the coordinate fUllctions of F. Replacing w(g 0 u) by its formula, the term
becomes
Applying B to this expression:
(a9 L (aF 8x, 8YJ In
_J.
-.'
F ) 0 u)
j=1
m L r +L j=1 k=1
(
aF (aCk) F
(9' F) _ J ax,
-.'
aYJ
)
0 OZk U ,
which, by interchanging the summations and applying the chain rule, becomes
~(aFj ~ Ox . j=1
(a
I
g
~. F ) ® u )
,.1J
+~ ~ (g. F) k=1
(O(Ck'F) {). 0 8 Zk U) . X,
15, Embeddings
142
Substituting the last expression in a(8",;(h ® 9 ®
L:
8h,® u + h m-8 (8 F (g , F) -8 J 89 ( ,F8®)u) x, j=1 x, YJ
u», one obtains
+ h Lr (g , F) 8(G8" " F) x,
k=1
®
8'k U.
Another application of the chain rule and Leibniz's rule to the first two terms shows that this is equal to
and hence to 8:z;;{h(g· F) ® u). We have thus proved that
Thus () is an isomorphism of An-modules. Applying this theorem to polynomial isomorphisms we get the following corollary.
2.2 COROLLARY. Let F ; X --; X be a polynomial isomorphism and G its
inverse. If M is a left An-module, then F* G* M
~
M.
3. EMBED DINGS REVISITED.
In this section we study a more general type of embedding. Let F : X --; Y be a polynomial map. Define a new polynomial map j ; X j(X)
= (X, F(X».
-->
X x Y by
This is an injective map, as one easily checks.
We shall write j as a composition of two maps, as follows. Let
~
:X
--l
Y be the standard embedding of §1. Let G : X x Y --; X X Y be the polynomial map defined by G(X, Y) = (X, Y + F(X». Then G is bijective X
X
andi=G'L Now let M be a left Am+n-module. We will calculate j* M. By Theorem 2.1, it equals
~*G* M.
Let us compute G* M. First of all, by Corollary 1.3.2,
there is an automorphism 0' of Am+n which maps
Xi
and satisfies
O'(Yj) = Yj - Fj(X), n
0'(8,,><>
= 8",; + E (8Fj/8xi) 8'Jj' j-1
and 8'Jj to themselves
143
15. Embeddings
Note that u restricts to an automorphism of K[X,
11 which we will also call
u. This automorphism is the inverse of the comorphism
G. By Lemma
14.1.1, G*M~M(7
as a K[X, l1-module. Let 'IjJ stand for this isomorphism; it satisfies 'if;(h@u) u(h)u, for hE K[X,
11 and
=
u E M.
3.1 THEOREM. Let M be a left A'n+n-module. Then G* M
~
M(7 as A m + n -
modules.
PROOF: We know that the isomorphism holds for K[X, l1-modules. Let us investigate the behaviour of the action of 8",. under 'if;. By definition, 8",(h @ u) equals
Applying 'if; to this formula, we get n
(3.2)
u(8x ;(h)u+ u(h)8,,;u+
L u(h8F /8x.j)81J) u. j
Since u leaves the elements of K[X] unchanged, we have that u(Fj)
= Fj .
Thus (3.2) is equal to u(8",(h»u+ u(h)u(8,,,)u.
But [8"" h]
= 8x ,(h).
Hence
from which we deduce that 'if;(8z ;(h@ u» = u(8"Ju(h)u.
Thus
as required. One may similarly check that the action of 8!h' is compatible I with 'if;. The next corollary is a combination of Theorem 3.1 and Corollary 9.2.4.
15. Embeddings
144
3.3 COROLLARY. Let M be a E.nitely generated left Am+n-module. Then
G* M and M have the same dimension. We will put all this together in a theorem. 3.4. Let M be a left Am+n-module. Then
THEOREM
i*M~M"/(Y)M"
as An-modules. As we have seen, j*M
PROOF:
~
£*G*M. By Theorem 3.1, G*M
~
M" .
By §1,
as claimed. 4. EXERCISES
K
K2 be the standard embedding. Compute the inverse image
4.1 Let
£ :
under
of the following A2-rnodules.
£
---+
(1) A2/A2fh
(2) K[XbX2J
(3) A2/A2x2 (4) A2/A2X2~ (5) A21A2~ 4.2 Which of the inverse images of Exercise 4.1 are finitely generated over AI?
4.3 Let
£ :
X --; X x K be the standard embedding, and denote by y the
coordinate of K. Let An+! be the Weyl algebra generated by x's and ax's and by y and Oy. Suppose that the term of highest order of dEAn+! is 8~. Show that L*(An+I/An+Id) is a free module of rank k over An4.4 Keep the notations of Exercise 4.3. Suppose that I is a left ideal of An+1
which properly contains An+! d. Show that £*(An+!/1) is finitely generated over An.
145
15. Ernbeddings
4.5 Let F : X -+ X x X be the polynomial map defined by F(X)
= (X, X).
This is a generalized embedding in the sense of §3. Let M and N be left modules over An. Then M0N is a left module over A 2n = An0An. Show that
F"(M0M) S!! M
@K[X)
N.
The latter was defined in Exercise 14.4.4. 4.6 Show that if £ : X -+ X x Y is the standard embedding, then
where K[oy] is the left Am-module K[oy" ... ,oYm].
CHAPTER 16 DIRECT IMAGES In the previous chapter we saw that starting with an Am-module and a
polynomial map F : Kn
-t
Km we can construct an An-module, its inverse
image. We shall now study a similar construction, which uses F to associate an Am-module to an An-module, its direct image. Curiously, the direct image is easier to define for right modules; and it is with them that we start. Throughout this chapter, the conventions of 14.1.2 remain in force. 1. RIGHT MODULES Let us briefly recall the definition of the inverse image. Let F : X
-t
Y
be a polynomial map. Let M be a left Am-module. The inverse image of M under F is F" M = K[X]
@]([Y]
M. This is a K[X]-module. It becomes an
An-module with the OX; acting according to the formula
oZi(h@u)
oh
m
Xi
1
op.
= {) @u+ Eh{/- @o'JjU' X;
Let us rewrite this definition in a slightly different way. Since
Am@A .. M 9!
M, we have that
Writing DX-tY for F* Am
= K[X] @K[y] Am,
one has that
Note that D x -+ Y is a left An-module and a right Am-module. One easily checks, that these two structures are compatible and that it is in fact an An-Am-bimodule. The bimodule D
X-+Y
is the key to the direct image construction. Let N
be a right An-module. The tensor product
16. Direct images
147
is a right Am-module, which is called the direct image of N under the polynomial map F. Consider the projection
11' :
X
X
Y -> Y defined by 1I'(X, Y)
= Y.
From
Ch.14, §3, it follows that
As a left An-module, K[X] is isomorphic to Ani L~ An8"". Hence, by Corollary 13.2.3, n
11'* Am
~ Am+nl
L A m+n8"" 1
is an isomorphism of Am+n-Am-bimodules. Now if N is a right Arn+n-module, then by Corollary 12.4.7,
as right Am-modules.
2. TRANSPOSITION We will now see how one can turn right modules into left modules, and vice versa. That will allow us to do direct images for left modules. We present the construction for general algebras. Let R be a K -algebra. A tmnsposition of R is an isomorphism r : R
->
R
of the underlying K -vector space that satisfies
(1) r(ab) = r(b)r(a), (2) -r2 = r. Examples oftranspositions abound. The most familiar occurs in the matrix ring over a field. An example in the Weyl algebra is provided by the map r : An -> An defined by
where h E K[xj, ... , xnJ and Q E !"in. We will refer to it as the standard tran.sposition of A". If dEAn, we also use the notation d for T(d). T
148
16. Direct images
Let us go back to the general construction. Let R be a K -algebra and
7'
a
transposition of R. If N is a right R-module then we define a left R-module
Nt as follows. As an abelian group, Nt = N. If a E Rand
U
E Nt then the
left action of a on U is defined by aou = U7'( a). The left action 0 is called the
transposed action. Instead of checking in detail that Nt is a left R-module, let us prove just one typical property, namely
(ab)
0
u = a 0 (bo u)
where a, b E Rand u EN. The left hand side is, by definition, u7'(ab). Since 7'
is a transposition, it equals u7'(b)7'(a). But this is ao (bo u). The same construction can be used to turn left modules into right modules.
Indeed, let M be a left R-module. The right R-module Mt equals M as an abelian group. The right action of a E R on u E M is defined by uoa Note that since
-? =
7' ,
= 7'( a) u.
the two constructions above are inverse to each
other. If N is a right R-module, then (Nt)t = Nj and the same holds for left modules. It is very easy to transpose the action when the module is cyclic. Let
R be a K-algebra with a transposition
7'.
If J is a left ideal of R, define
Jt = {7'(a) : a E J}. This is a right ideal of R. 2.1 PROPOSITION. Let R be a K -algebra with a transposition
7'
and J a left
ideal of R. The right R-module (R/ J) t is isomorphic to R/ P. PROOF:
If a E R, denote by
defined by ¢(a)
= 7'(a).
a its image in (R/ J)t.
Consider the map
It is clearly a homomorphism of additive groups. If
a, b E R, then
¢(ab) which equals 7'(b)7'(a)
= ¢(a) 0 b.
= 7'(ab),
Thus ¢ is a homomorphism of right R-
modules. One can easily check that it is surjective; let us calculate its kernel. Suppose that a E R satisfies
16. Direct images
149
conclude that ker
R/Jt';E, (R/J)t. Let us illustrate the proposition with a concrete example of a module over the Weyl algebra. Let J
= Anxn' and let
T be the standard transposition of
An. Since T(Xn) = xn> it follows that P = xnAn. Thus
This example will surface again later on. Al! this can be extend to bimodules. Let RI, R2 be K-algebras with transpositions
Tl, T2,
respectively. Let M be an R 1 -R2 -bimodule. The transposed
bimodule Mt is obtained by applying the above constructions to both the left and the right actions on M. Thus Mt and M have the same underlying abelian groups and, if al E R 1 , a2 E R2 and u EM, then
These actions are compatible with each other, since the original actions were so. Thus Mt is an R 2-R I -bimodule. The transposition of bimodules over the Weyl algebra has its own peculiarities. Suppose that m::; n and that Tm and Tn are the standard transpositions· of Am and An, respectively. Then Am Tin'
~
An and Tn restricted to Am equals
This considerably simplifies the calculations. For instance, if J is a
left ideal of An whicll is a right Am-module, then it follows from these considerations and Proposition 2.1 that (An/J)t is isomorphic to An/ Jt as an Am-An-bimodule . The left ideal J
= Anxn is an example of this situation.
This also means that we may denote the standard transposition simply by T,
irrespective of the index of the Weyl algebra in question. For future reference, we must consider the behaviour of the transposition
under tensor product. 2.2 LEMMA. Let R 1 , R2 and R3 be K -algebras with transposition. Suppose that MI is an RI-R 2 -bimc;Jdule and M2 an R2-R3-bimodule. Then
150
16. Direct images
as Rt-Ra -bimodules. PROOF: Consider the map
defined by
= Ut ® 'UIJ.,
where
Uj
E Mi. We show that
and leave the proof that it is bilinear to the reader. If a E R2 and
~
is the
transposition of R 2 , then
By the definition of
Now, using that the tensor product M t ®R. M2 is balanced, we get that
Hence,
Thus
It is clear from the definition of ¢ that it is bijective, and the proof is com-
plete.
3. LEFT MODULES Using the results of §2, we will do direct images for left modules. Let F :X
~
Y be a polynomial map. Consider the An-Am-bimodule D X-;y . Using the standard transposition for Am and An defined in the previous
section, put
This is an Am-An-bimodule.
151
16. Di1'ect images
Let M be a left An-module. The direct image of M by F is defined by the formula
It is clearly an Am-module. This is very convenient, because to compute
direct images one has only to transpose the actions of D X~y ; and this is usually easy to do. Let us work out the calculations for the projection
7r :
X x Y
~
Y. From
§1, we have that
Since the standard transposition of Am+n maps 8x, to -8x" we conclude by Proposition 2.1 that n
Dy<-xx y 2::
Am+n/
L 8x,Am+n.
111.1
III!
I
I
Note that the y's commute with the 8x 's. In particular, L~ 8""A m+ n is a left
Am- submodule of Am+n' Thus Am+n/ L~ 8""A m+n is a left Am-module for the quotient action. The direct image under projections does not take finitely generated modules to finitely generated modules. Indeed, Am+n is cyclic as a module over itself, but Am+n/ L~ 8"" Am+n is not finitely generated over Am· It is a free Am-module; with a basis given by the monomials on the x's. Using this description of Dy<-xx y we can easily calculate the direct image of any module under a projection. If M is a left Am+n-module, then by Corollary 12.4.7, n
7r*
M ~ M/
L 8",;M. I
Another interesting example is provided by polynomial isomorphisms. Let
F : X ~ Ybe a polynomial map, and consider the polynomial isomorphism G : X x Y ~ X x Y defined 'by G(X, Y)
=
(X, Y + F(X)).
:1
~
I I
16. Direct images
152
In Ch. 15, §3, we saw how to calculate inverse images under this map. In
particular, there exists an automorphism
D Xx Y-+Xx Y
IJ'
of Am+n such that
= G*(Am+n) ~ (A m+n)".
We wish to transpose this module in order to calculate direct images. The final result is very simple, but a technical lemma is required. 3.1 LEMMA. Let T be the standard transposition of Am+n' Then T' IJ" T = PROOF:
It is enough to check that the identity holds on the generators of
A m +n . First, of K[X,
IJ'.
11.
T
is the identity on K[X,
Thus
T· IJ"
T
11 and IJ' restricts to an automorphism
= IJ' in K[X, 11. On the other hand,
TIJ'T(8,,J = TIJ'(-8",J = -T(8",;
+
E 8F. 8x'.8y,.). m
1
I
Since 8Fj 18x; commutes with 8y,. we have that
TIJ'T(8,,;)
= IJ'(8y,);
Similarly, TIJ'T(8y,) 3.2
and the proof is complete.
Let M be a left Am+n-module, then
PROPOSITION.
PROOF:
= IJ'(8,,;).
Let us first find the transpose of (Am+n)u. Consider the map
'I/J: ((Am+n),,)t defined by 'I/J(u)
= T(U).
if a, b E A'n+n and
U
-+
A'n+n
It is a K-vector space isomorphism. Furthermore,
E ((Am+n),,)t then
a 0 uo b = T(b)
* uT(a) = IJ'(T(b»uT(a).
Therefore, 'I/;(aouob) = aT(u)TIJ'T(b). By Lemma 3.1, TIJ'T(b) = IJ'(b); and so
Hence D Xx y<-Xx y is A'n+n as a bimodule over itself, but with the right action twisted by
IJ'.
Therefore,
By Lemma 14.1.1 this is isomorphic to MIT-I.
16. Direct images
153
3.3 COROLLARY. Let M be a finitely generated left A'n+n-module. Then G*M and M have the same dimension.
PROOF; Follows from Proposition 3.2 and Corollary 9.2.4.
4. EXERCISES 4.1 Let R be a K-algebra and
TI, 7?!
TI (7?!)-1
is an automorphism of R.
4.2 Let
71":
two transpositions of R. Show that
K2 ~ K be the projection on the second coordinate. Compute
the direct image of the following A 2 -modules under
71".
(1) A2/A2~ (2) K[Xb X2] (3) AdA281 (4) A2/A2X2
(5) AdA2&i 4.3 Which of the direct images of Exercise 4.2 are finitely generated over AI? 4.4 Let F ; X
~
Y be a polynomial map and M a left An-module. Show
that
4.5 Let
~
;X
~
X x Y be the standard embedding and TJ : X x Y
~
X be
the projection on the first coordinate. Let M be a left A'n+n-module. Show that
~* M
is the Fourier Transform of 7].,M.
4.6 Let F : X G(X, y)
~
Y be a polynomial map. Define G : X x Y
~
X x Y by
= (X, Y + F(X)). Let M be a left Am+n-module. Show that G.G*M
= M = G*G.M.
CHAPTER 17 KASHIWARA'S THEOREM
In this chapter we deal with the direct image under an embedding. This will lead us to an important structure theorem for a whole class of Anmodules. The noncommutativity of the Weyl algebra will be an essential ingredient in these results. In fact the theorem fails to hold for polynomial rings, as we show at the end of §2. Throughout this chapter, the notation of 14.1.2 will be in force. 1. EMBED DINGS
Let ~ : X
-+
X x Ybe the standard embedding: L(X)
= (X,O), where 0 is
the origin of Y. It follows from Ch. 15, §1, that
where (Y) is the ideal of K[X,
11 generated by the y's.
One easily verifies that
this isomorphism preserves the right Am+n-module structure of L*(Am+n). Transposing this bimodule with the help of the standard transposition of
Am+n and using Proposition 16.2.1,
This is an Arn+n - An-bimodule. If N is a left An-module, then its direct image under
~
is, by definition, the module ~N = DxxY<-x @A.
N.
Let us consider the structure of DxxY<-x in greater detail. By Corollary 13.2.3 there exists an isomorphism of bimodules,
But Am/Arney) is isomorphic to K[BI/l'" . ,By.] Therefore, (1.1)
= K[ByJ
as an Am-module.
155
17. Kashiwara's theorem
Let
Q
E IIln. Recall that the action of YJ on 0° E K[oy] is given by
It is worth stressing that on the right hand side of (1.1), the y's and Oy'S act only on K[oy], whilst the x's and o.,'s act only on An; as in eh. 13, §2. Thus if N is a left An-module,
It is very easy to calculate with the direct image in this form; we give two examples. The next theorem should be compared with Theorem 14.3.1. 1.2 THEOREM. Let
t :
X
~
X x Y be the standa.rd embedding. If M is a
fimtely generated left An-module, then:
(1) /.oM is a finitely generated left Arn+n-module. (2) d(4M)
= m+ d(M).
(3) m(4M):5:: m(M). PROOF: (1) follows from Proposition 13.2.1. Since K[oy] is the Fourier transform of K[X], it must have dimension m and multiplicity 1 by Proposition 9.2.2. Thus (2) and (3) follow from Theorem 13.4.1.
The following corollary is an immediate consequence of the theorem. 1.3 COROLLARY. Let M be a holonomic An-module and
t
the standard
embedding. Then /.oM is a holonomic Am+n-module. We now turn to another application. 1.4 LEMMA. Let N be a left An-module. Every element of t*N is anmhilated by a power of Yj, where 1 :5:: j :5:: m PROOF: Without loss of generality assume that j
= 1.
Since
in K[oy], it follows that 00. is annihilated by yf, for any k ~ every element of L*N may be written in the form I
QI
+ 1.
Now
17. Kashiwam's theorem
156
where Ua E N and I £;; Nn. Choose an integer k such that k every
Q
E I. Then yf . f)
= 0, y~ .
for every
(L::
f)a
Q
~ Qt
+ I,
for
E I. Thus
® Ua)
= O.
aEI
We also deduce from the isomorphism
that DxxY<-x is free of infinite rank as a right An-module. This follows from the fact that K[f)yj is a K-vector space of infinite dimension.
The
monomials in the By's form a basis for DxxY<-x as a right An-module. A simple consequence of this fact and Theorem 12.4.2 is the following corollary. 1.5 COROLLARY. If N is a left An-module, then
~*N
=
0 if and only jf
N=O.
2. KASHIWARA'S THEOREM
In this section we study the direct image under the following special case of the standard embedding
~
:X
-t
X x K. Let y denote the coordinate of
K. The polynomial ring in the x's and y will be denoted by K[X, yj. Let M be a K[X, yj-module and H the hyperplane of equation y = 0 in X x K. The submodule of M of elements with support on H is defined by
rHM
= {u EM:
U
is annihilated by a power of y}.
Note that by Proposition 12.5.1, this is exactly the kernel of the map M
-t
M[y-lj. Although we have defined rHM for K[X, yj-modules, in most of our applications it will also be an An+l-module. 2.1 PROPOSITION. Let M be an An+l-module. Then rHM is a submodule
oEM.
111 and 11' are the powers that 11 +'" annihilates both UI and '11.2, hence
PROOF: Let Ul, '11.2 E rHM. Suppose that annihilate Ut, '11.2, respectively. Then their Rum.
1
157
17. Kashiwara's theorem
It is harder to show that rH M is closed under multiplication by elements
of A n+l . Of course it is enough to check that it is closed for multiplication by the generators of An+!. Assume that u E rHM is annihilated by that
if commutes with the x's, 8",'s and with
and so 8z ,u E rHM. Similarly, that 8yu E
r H M.
XiU,
if.
Note
y. Thus,
yu E rHM. Hence we need only check
Using the Weyl algebra relations,
which completes the proof of the proposition. Given an An+!-module M, put kerM (y) Clearly kerM(Y)
<;::;
= {u EM: yu = a}.
rHM. Although kerM(Y) is closed under addition, it is
not an A n+1- submodule of M. But, since the y's commute with the x's and 8.,'s, it is an An-submodule of M. We now make explicit the true relation between kerM y and rHM. It all begins with a lemma. To simplify the notation, let Mo
= kerM(Y).
2.2 LEMMA. Let M be a left A 71+1-module, and let H be the hyperplane
Y= O. Then (1) For each kEN, the map
defined by right multiplication by 8y is injective. (2) y(An+IMO) = An+1Mo. (3) An+lMO = Mo EX! 8yMo EX! ~Mo EX! •..•
PROO~:
Let u E Mo
~n~
k be a positive integer. We have tha1[y'0;] = 'u, thus y. 8;u = -k~-I u. Therefore,
-k{}~ 1 and that y anmlulates
(2.3)
158
17. Kashiwara's theorem
for every i ~ O. Suppose that 8~u = O. Applying (2.3) with i = k to this equation, we conclude that u
= O.
This proves (1).
Now, every element of An+! can be written in the form
where bj E An[Y]. Since Mo is a left An-module that is annihilated by y, the elements of An+l Mo can be written as
where
U; E
Mo. In other words, An+lMo = Mo
Thus (2) follows from ~u = -(k
+ 8yMo + lfyMo + .... + 1)-ly. (8~+lu).
Assume now that
if and using (2.3) we conclude that = O. Thus we have a sum with k - 1 terms. By induction on the number
where tto, ... ,Uk E Mo. Multiplying by Uk
of terms, we conclude that tto
= ... = Uk = O.
Hence the sum is direct, and
(3) is proved. We are now ready to state a preliminary version of Kashiwara's theorem. 2.4 THEOREM. Let M be a left An+!-module, let H be the hyperplane y = 0 and denote by ~ : X
~
X x K the standard embedding. The An+! -modules
4(kerM y) and rHM are isomorphic. PROOF: We have seen that Mo
= kerM y is an An-submodule of M.
from §1 that
We wish to show that this is isomorphic to rJIM.
It follows
17. Kashiwara's theorem Since Mo
~
159
rH M, we can use the universal property of the tensor product
to define a map
by
CPU 0
u) = Iu, where I is a polynomial in By. The image of
and its kernel is zero, by Lemma 2.2(3). Thus to prove that cP is surjective it is enough to show that rH M is contained in the An+l-submodule of M generated by Mo. Let u E rHM and assume that 1Iu
=
O. We want to show that u E
An+lMo. The proof is by induction on k. If k = 1, then
U
E Mo. Suppose
that the result holds for every element of Mo annihilated by
11- 1 •
From
yku = 0, we deduce that By(yku) = O. Since [By, yk] = kyk-I, :I'-I(ku+ yByu)
= O.
11- 1 (yu)
Thus, by the induction hypothesis, ku + yByu E An+IMO. But
=
1Iu = 0 and so yu E.A 71+I Mo. Since An+IMO is an An+l-submodule of M,
we have that ku + yByu - ByYu E An+IMo. But this expression is equal to (k - 1)u, because [By, y]
1. Thus u E
An+IMO, as required. Combining the fact that rH(M)
= An+lMO with Lemma 2.2(2) we get the
following result, which will be used in the next chapter.
2.5 COROLLARY. Let M be a left An+l-module and let H be the hyperplane
y = O. T11en yrH(M)
= rH(M).
We say that a K[X, y]-module M has
SUPP017,
on H if rHM
= M.
In
particular this definition may be applied to An+1-modules. The next result is a structure theorem {or An+l-modules with support on H. The proof is an immediate consequence of Theorem 2.4 and the above definition.
2.6 COROLLARY. Let M be an An+l-module with support on H.
Then
M ~ ~*(kerM V). These results hold in greater generality. In Ch. 18, §3 we
VI
fo
use the
language of category theory to bring forth the true nature of Kashiwara's theorem.
17. Kashiwara's theorem
160
We have taken care to point out that both rHM and kerM yare defined for K[X, y]-modules as well as for An+1-modules. In fact there is also a natural way to define direct images for modules over polynomial rings under polynomial maps; see Exercise 3.4. Applied to the standard embedding
t :
X -; X x K, the direct image t+M of a K[X]-module M equals M, as
a vector space. The action of
f (X, Y)
E K[X, yJ on u E
M is given by
fu= f(X,O)u. For this direct image, it is not true that if M is a K[X, yJ-module then
Here is a simple example. Consider the K[X, yJ-module M Since every element of M is annihilated by
= K[X, yl/(1/).
1/, it follows that rHM = M.
a straightforward calculation shows that kerM y
= yM.
But
Since t+(kerM y)
=
yM, as vector spaces, we have that L+(kerM y) of rHM.
3. EXERCISES
3.1 Let
K -; K2 be the standard embedding given by t(x)
t :
Compute the direct image of the following modules under
(x,O).
t.
(1) K[x] (2) Ai/Alas (3) AtlA l x8 (4) A I /A 1 (x 2 8+4x+l)
3.2 Let
t ;
X -; X x Y be the standard embedding. Let M be a left An-
module. Is it true that
3.3 Let
L :
X -; X x K be the standard embedding and let M be a left
An+l-module with support on the hyperplane H of equation y = 0. Show that kerM y is isomorphic, as a left An-module, to
161
17. Kashiwam's theorem
3.4 The purpose of this exercise is to define a direct image for modules over polynomial rings. Suppose that F : X -; Y is a polynomial map. The comorphism of F is a homomorphism of rings FI : K[ YJ -; K[X]. Let M be a K[X]-module. We want to construct a K[YJ-module F+M. The definition is as follows. As K-vector spaces, F+M
= M.
The action of 9 E K[ YJ on
u EM is defined by gu = FI(g)u. (1) Check that this action makes F+M into a K[YJ-module. (2) Compute L+M for the standard embedding L : X -; X x Y. (3) Let N be a K[X, YJ-module. Compute 7r+N for the projection
7r :
X x Y -; Yonto the second coordinate. (4) If M is finitely generated over K[X], is F+M necessarily finitely generated over K[YJ,? 3.5 Let M be a left An-module and
cf> :
M -; M[x~l] the natural embedding.
Show that ker cf> and cokercf> are supported on the hyperplane
Xn
= O.
3.6 Let M be an irreducible An-module. For a non-zero element u in M let J(u)
= {f E K[X]: fu= O}. Show that:
(1) If 0 =I- v E M then radJ(v)
= radJ(u).
(2) radJ(u) is a prime ideal of K[X]. (3) If M is supported on the hypersurface
Xn
tained in the ideal of K[X] generated by
= 0, then radJ( u) is con-
X n•
/
CHAPTER 18 PRESERVATION OF HOLONOMY We have seen in the previous chapters that holonomic modules are preserved by inverse images under projections and by direct images under embeddings. However, as we also saw, inverse images under embeddings and direct images under projections do not preserve the fact that a module is finitely generated. Fortunately, though, holonomic modules are preserved by all kinds of inverse and direct images. The proof of this result will use all the machinery that we have developed so far. It gives yet one more way to construct examples of holonomic modules. We retain the notations of 14.1.2. 1. INVERSE IMAGES
The key to the results in this chapter is a decomposition of polynomial maps in terrns of embeddings and projections. The idea goes back to A. Grothendieck. Let F : X
~
Y be a polynomial map. We may decompose F as a com-
position of three polynomial maps: a projection, an embedding and an isomorphism. The maps are the following. The projection is 7r : X x Y defined by 7r(X, y)
G(X, Y)
= (X,
=
Y The isomorphism is G : X x Y
Y+F(X)). Finally, the embedding is L: X
~
~
~
Y,
X x Y where
X x Y, defined
by L(X) = (X, 0). One can immediately check that F = 7r. G· L . Let us consider the effect of an inverse image under F on the holonomic Am-module M. By Theorem 15.2.1, F*M
~
i*(G*(7r*M))).
It follows from Corollary 14.3.2 and Corollary 15.3.3 that 0*( 7r* M) is holo-
nomic if M is holonomic. Hence F* M will be a holonomic module if we can prove that i* preserves holonomy. We shall now do this. We will begin with the standard embedding dinates of X x K will be denoted by hyperplane y = o.
i :
Xl, ... , X n ,
X
~
X x K. The coor-
Y and H will stand for the
18. Holonomy
163
1.1 LEMMA. Let M be a left An+l-module. Put M'
= M/rHM.
Then
~"M~~"M'.
PROOF: The inverse image
~. M
is isomorphic to M/yM. Similarly
~. M' ~
M' /yM'. Since M' is a quotient of M, there exists a surjective map of An-modules,
Let
U
---!
EM and suppose that
M' /yM'.
= O. Then
U
E yM + rHM. Since
yrHM = rH M by Corollary 17.2.5; we conclude that u E yM and that
non-zero element with support on H.
Then~· M
is a holonomic An-module
whose multiplicity cannot exceed the multiplicity of M. PROOF: Let
r
be a good filtration for M. Put
Note that we are using a filtration of M - an An+l-module - to construct a filtration of M/yM - an An-module. It is clear that Uj~O {}j
= M/yM. Since Bj(An)
~ Bj(An+l),
{}j ~ {}j+l
and that
we have that Bj(An)rj ~ ri+j.
Thus
Hence {}
= {{}j
:
j ~ O} is a filtration of the An-module M/yM. Note that
we do not know whether this filtration is good. By the third homomorphism theorem for vector spaces, {}j ~
IJ/(IJ n yM).
Therefore,
But yrj -
J ~
Ij
n yM,
and so
I
/
18. Holonomy
164
Since M does not have any element supported on H, the map [j-l
---t
yrj-l given by right multiplication by y is injective. Hence dimK(yr,'-I) =
dimK (I;-d and so (1.3) Now let x( t, M, r) be the Hilbert polynomial of M with respect to the filtration
r . For i »
0, it follows from (1.3) that
dimK
{}j ~
xU, M, r) - xU -
1, M,
r).
The polynomials x( t, M, r) and x( t - 1, M, r) have the same leading term, m(M)f'/n!, where m(M) stands for the multiplicity of M. Hence their difference has leading term m(M)tn-1/(n - I)!. Thus there exists CEQ such that . m(M)r- 1 dllnJ( {}j ~ ( ), n-l.
.
+ c(; + 1)
n-2
.
We conclude from Lemma 10.3.1 that M/yM is a holonomic module whose multiplicity cannot exceed m(M).
1.4 THEOREM. Let
L :
X
---t
X
X
Y be the standard embedding and let M
be a holonomic Am+n-module. Tllen PROOF: Recall that Y that m
= 1.
L* M
is a holonomic An-module.
= Km. We proceed by induction on m. Suppose first
= 0,
Denote by H the hyperplane Yl
and put M'
= M / rH M.
The module M' is a quotient of a holonomic module, hence is holonomic. Since M' does not contain any element supported on H, it follows from Lemma 1.2 that L* M
L* M'
is holonomic. But by Lemma 1.1,
hence
is also holonomic.
Now the standard embedding
L :
X
---t
X
X
position of two standard embeddings, namely L2 :
L* M ~ L* M',
X x
Kn-l ---t
X x Y. By Theorem 15.2.1,
The result follows by induction. Let us now put all these results together.
Y may be written as a comLl
:
X
---t
X x
Kn-l
and
18. Holonomy
1.5 THEOREM. Let F : X
-t
165
Y be a polynomial map. If M is a holonomic
Am-module, then F* M is a llOlonomic An-module.
2. DIRECT IMAGES. That holonomy is preserved under direct images is proved in a very similar way to the inverse image case. However we must first tidy up some loose ends. 2.1 THEOREM. Let F : X
-t
Yand G: Y
-t
Z be two polynomial maps.
Let M be a left An-module. Then
PROOF: It follows from Theorem 15.2.1 that (GF)*(Ar)
~
F*G*(Ar). In the
notation of Ch. 16, §1, this is equivalent to
Applying Lemma 16.2.2 to this isomorphism we get that
The theorem is an immediate consequence of this formula. Now, in the notation of §1, we have that the polynomial map F : X can be written as F
=
7r' G
.~.
-t
Y
Let M be a holonomic Am-module. By
Theorem 2.1,
We already know, from Corollaries 16.3.3 and 17.1.3, that G*(~*M) is holonomic. In order to show that F*M is holonomic, it is enough to show that '1r*
preserves holonomy.
2.2 THEOREM. Let 7r: X x Y
-t
Y be tIle projection 7r(X, y) = Y. Let M
be a holonomic Am+n-module. TIlen
'1r.. M
is a holonomic Am-module.
PROOF: We know from Ch. 16, §3, that n
7r.. M ~ Mj
L8
xi
I
M.
/
18. Holonorny
166
By Proposition 5.2.1 this is the Fourier transform of M I E~ x;M. Therefore the modules 7T*M and M!E~XiM have the same dimension by Proposition 9.2.2. However, M!E~XiM is isomorphic to L*M where L: Y -+ X x Yis the embedding L( Y)
= (0,
Y). By Theorem 1.4, L* M is holonomic; hence so
is 7T*M. Summing up, we have proved the fullowing theorem. 2.3 THEOREM. Let F : X -+ Y be a polynomial map. If M is a holonomic
An-module, then F*M is a holonomic Am-module. 3. CATEGORIES AND FUNCTORS. We now introduce direct and inverse images as functors and generalize our previous version of Kashiwara's theorem using categories. Fbr that we assume that the reader is familiar with the language of category theory. All that is required here can be found in [Cohn 79]. We shall be dealing with three categories. The category ofleft An-modules and module homomorphisms will be denoted by M gory of finitely generated left An-modules by
M'l
n
,
and the full subcate-
The full subcategory of
Mn whose objects are holollomic left An-modules will be denoted by 1{n. Since holonomic modules are finitely generated, it follows that 1{n is a full subcategory of Mj. Note that these are all abelian categories. Let F : X
-+
Y be a polynomial map. The inverse image F* is a functor,
If M is an object in M
m
then
Thus F* is a right exact functor by [Cohn 79, §4.3, Proposition 3]. L :
K
-+
(x,O). Recall that L*(A2)
=
A2!x2A2' Let
general, this functor is not exact. Fbr example, let standard embedding: i(X)
¢> ; A2
-+
=
In
A2 be the map defined by ¢>(d)
get a homomorphism of AI-modules,
= dx 2 .
K2 be the
Applying the functor
L*
we
18. Holonomy
167
But 1 +x2A2 is mapped to zero by L*(cfJ); hence it cannot be injective. Thus £*
is not an exact functor. An important special case occurs when the polynomial map is a projection.
Define 1T: X x Y
-t
Yby 1T(X, y)
=
Y. Then n
DX x y .... y2i!:. Arn+nlL::Arn+nB"" 1
is free as a right Am-module: the monomials in the x's form a basis for this module. Since free modules are fiat, we conclude that the functor
is exact. We have also seen that 1T* preserves noetherianness. Hence we may restrict it to a functor,
This last statement does not hold true for general polynomial maps: it is false for embeddings, for example; see eh. 15, §1. On the other hand, if F :X
-t
Y is any polynomial map, then by Theorem 1.5 the inverse image
F* restricts to a functor from 1{rn to 1{n. So the behaviour of the inverse
image functor in the holonomic category is up to expectations. For direct images we have a similar situation. Given a polynomial map F ;X
-t
Y, the direct image F* determines a functor,
Since F* is also defined by a tensor product, the direct image is right exact. It is not in general exact. Although the direct image does not always preserve
finitely generated modules, it preserves holonomic modules by Theorem 2.3. Let us turn to direct images under embed dings. Denoting the coordinates of X x K by
Xl, ... ,X n ,
y, let
L :
X
-t
X x K be the standard embedding
L(X) = (X, 0). Then
DXl'K<-X
2i!:.
An+t/An+lY
is free as a right An-module, the monomials in By form a basis. H~;nce the functor
/
168 is exact.
18. Holonomy
Since
DXxK<-X
is finitely generated as a left An+l-module, the
functor ~" restricts to a functor M'i
---t
M 7+1 .
Now let H be the hyperplane defined by the equation y
= O.
Let Mn+l(H)
be the full subcategory of Mn+l of modules with support on H. Thus, M is
= M.
an object in Mn+1(H) ifand only if rHM that if M is an object in Mn, then
~*M
In Lemma 17.1.4 we showed
has support on H. This leads to a
categorical version of Kashiwara's theorem. 3.1 THEOREM. Let ~ ; X
X x K be the embedding t(X) = (X, 0) and H
---t
the hyperplane of equation y = O. The dll'ect image functor, t. :
Mn
"-7 Mn+l(H),
is an equivalence of categories. PROOF: Consider the functor
1C : Mn+l
--t
M n
defined on a left An+1-module M by 1C(M)
= kerM H.
By Theorem 17.2.4
we have that
= rHM. then ~,,(1C(M» = M. L.,,(1C(M»
If M has support on H,
On the other hand, let N be any left An-module. Then L."N!?;iK[8y ]0N
by Ch. 17, §1. Since the only elements of K[8 y ] that are annihilated by y are the constants, we conclude that ker,.N H = K0N
as An-modules. Hence 1C(t.N)
!?;i
!?;i
N
N.
We leave it to the reader to check that the natural isomorphisms defined above have the expected behaviour on maps. This may be generalized to embeddings in higher dimension. Let X x Ybe the embedding ~(X)
of equations Yl
= (X,O).
~
:X
--t
We will identify Y with the subspace
= ... = Ym = O. Let M be an object in Mm+n. Write Hi for = O. We say that M has support on Y if it has support
the hyperplane Yi
on Hi' for every i = 1, ... ,m. The full subcategory of Am+n-modules with support on }'will be denoted by M"'+n(Y).
18. H olonomy
169
3.2 COROLLARY. Tlle functor
is an equivalence of categories. PROOF: Recall that Y = Kin. We proceed by induction on m. The case m =
= Kin-I.
1 has been proved in the theorem. Let W £1 :
X
-t
X
X
Wand
£2 :
X x W -t X
X
There are embeddings
Y such that
£
=
£2 . £1.
It follows
by induction that (£1)* and (£2)* are equivalencies of categories. Since, by Theorem 2.1,
we have that
£..
is an equivalence of categories.
Since the direct image under embeddings preserves finitely generated modules, we may replace the categories in Corollary 3.2 by the corresponding full subcategories of finitely generated modules. A similar statement holds for the holonomic categories, see Exercise 4.5. Kashiwara's theorem can be used to prove a structure theorem for Anmodules with support on the origin. 3.3 COROLLARY. If M is a finitely
genel'a~ed
on the origin, then tllere exists an integer r M
PROOF: Let
£ :
{O}
-+
~
~
left An-module with support 0 such that
(K[llI, ... , anD"
X be the standard embedding: £(0)
= O.
Note first
that it follows from the definitions that
Hence K[a!> ... ,an] has support on every hyperplane Hi, for 1 ~ i ~ n, by Lemma 17.1.4. Therefore it has support on HI n··· n Hn
= {O}.
On the other hand, if M has support on the origin then K(M) is a finitely generated module over Ao £..
= K.
Hence K.(M)
~
Kr, for some
(K.(M) = M, whilst
r ~
/) )
£ ..
(K") ~ (K[ujl""
UnlY.
O. But
170
18. Holonomy
Therefore, M
Q!!
(K[Bt, . .. , an))', as required.
4. EXERCISES 4.1 Let
M,N
be holonomic An-modules. Show that
M®K[X] N
is a holonomic
An-module. Hint: Exercise 15.4.5. 4.2 Let p E K[X] be a non-zero polynomial and let M be a holonomic Anmodule. Show that M[P-l] = K[X, p-l] ®K[X] M is a holonomic An-modJIle. 4.3 Give an example to show that the direct image functor is not exact under projections. 4.4 Let n
~
k
~
2n be positive integers. Use Kashiwara's theorem and
induction to construct an An-module of dimension k. 4.5 As in §3, let us identify Y with the linear subspace of equations YI
... =
Yin
=
0 in X x Y. Let 1·{'n+n( Y) be the category of holonomic left
Atn+n-modul~s
1{n.
with support on Y Show that this category is equivalent to
CHAPTER 19 STABILITY OF DIFFERENTIAL EQUATIONS In this chapter we investigate the global asymptotic stability of a system of ordinary differential equations on the plane. This is closely related to the Jacobian conjecture. Holonomic modules will make a special appearance in
§2.
1. ASYMPTOTIC STABILITY
We begin with some basic facts about the stability of singular points of systems of differential equations. Let G:]Rn
~]Rn
be a function of class cr for some r> 2, and assume that G(O)
= o.
Consider
the differential equation
X=
(1.1)
G(X).
By the uniqueness theorem [Arnold 81, Ch. 2, §8.3], the solution
w~th
= 0 is
r/I
= O.
We are interested in the behaviour of
neighbouring initial conditions. The singular point X
= 0 of
equation (1.1) is asymptotically stable if:
> 0, there exists 8 > 0 (depending only on E and not on t) such that, for every Po with lPol < 8, the solution r/I of (1.1) with initial condition r/I(O) = Po can be extended to the whole half line t> 0 and satisfies Ir/I( t)1 < E for every t> O. (2) There exists T) > 0 such that limt-.+oo r/I( t) = 0 for all solutions r/I of (1) Given
E
(1.1) which satisfy r/I(O) < "'. Condition (1) above means that if the solution is initially wit}:n a ball /
of radius {) around the origin then it will never leave a ball-6f radius Asymptotic stability is easy to determine for linear systems.
E.
172
19. Stability
THEOREM
1.2. Let A be an n x n matrix with entries in R. The origin is
an asymptotically stable singulru' point of X
=
A . X if and only if all the
eigenvalues of A have negative real part. Lyapunov showed that this can be extended to give a criterion to determine whether 0 is asymptotically stable in terms of the linearized system
X=
JG(O)· X. THEOREM
1.3. If the l'eal part of every eigenvalue of JG(O) is negative, then
o is an asymptotically stable point of (1.1). For a proof of this theorem see [Arnold 81, Ch. 3, Theorem 23.3]. We shall say that 0 is globally asymptotically stable if
T)
may be taken. to be
in (2) above. For a linear system, if 0 is asymptotically stable, then it is
00
globally asymptotically stable. Markus and Yamabe conjectured in [Markus and Yamabe 60] the following criterion for global stability. 1.4
CONJECTURE.
for each P E
X=
]Rn,
The origin is globally asymptotically stable for (1.1) if,
the origin is an asymptotically stable point of the system
JG(P) ·X.
It is shown in [Gutierrez 93] that the conjecture is true if n
banov 88] gives a counter-example for n the case n
= 2 when
~
= 2.
[Bara-
4. In this chapter we will study
G is polynomial, which was settled in [Meisters and
Olech 88]. Let us return to the hypothesis in Conjecture 1.4. For systems on the plane, the Jacobian JG(P) is a 2 x 2 matrix. By Theorem 1.2 the system
X = JG(P) . X
has the origin as an asymptotically stable point if and only
if the eigenvalues of JG(P) have negative real part. Note that since we are in the 2 x 2 case, this is equivalent to saying that the matrix JG(P) has positive determinant and negative trace. This suggests a definition. Let :F be the class of C1 maps F : R2 properties: (1) F(O) = 0; (2) trJ F(P) < 0 for all P E R2j (3) det.JF(P) > 0 for all P E R2.
---t
R2
which satisfy the following
173
19. Stability
For an example of a polynomial function in F which is not linear, see Exercise 4.5.3. We may now state the result of Meisters and Olech, [Meisters and Olech 88, Theorem 1].
1.5 THEOREM. Let F be a polynomial function in F then the ol'igin is a globally asymptotically stable point of the system
X = F(X).
One of the key lemmas in. the proof of this result has a purely D-module theoretic proof due to van den Essen, which we discuss in §2. Before we close this section, let us see how Theorem 1.5 can be applied to the Jacobian conjecture.
1.6 PROPOSITION. Suppose that the origin is a global asymptotically stable
point of X = F(X) for every polynomial map F E F. Then the polynomial maps in F are injective. PROOF: Suppose, by contradiction, that F is not injective. Then there exist points P),P2 E ]R2 such that F(PI )
X = H(X)
where H(X)
=
F(P2 ) = Q. Consider the system
= F(X + H) - Q.
Note that it has two distinct
critical points, one at the origin and one at P2
-
PI
i=
O. Thus the origin
cannot be globally asymptotically stable. However, J(H)(X)
= J F(X + PI),
and so H is a polynomial map in F, which contradicts the hypothesis. Since, by Theorem 1.5, the hypothesis in Proposition 1.6 is always satisfied by inaps in F, we conclude that these maps are always invertible. This is especially interesting since S. Pinchuk has recently given an example of a polynomial map on the plane whose determinant is everywhere positive, but which does not have an inverse, see [Pinchuk 94]. 2. GLOBAL UPPER BOUND In this section we prove one of the key lemmas used to settle Theorem 1.5. The proof we give, due to van den Essen, is purely D-module theoretic and works over any field of characteristic zero. In this section we will make free use of the results of Oh. 4, §4. Let F : K n -+ Kn be a polynomial map and denote by FI ,.
F nits coordinate functions. Let L\(x) = detJF(x). Throughout thh/'section we will assume that L1(x) i= 0 for every x E j(n. Note that since we are not ./
),
19. Stability
174
assuming that K is algebraically closed, this does not imply that Ll(x) is constant. Put d = degF
= ma:z:{degFi: 1 SiS n}.
Let 9 E K[X, Ll-I] and consider the derivations Di of K[X,Ll- 1] defined for i
= 1, ... , n by
as in eh. 4, §4. Note that n
Di = Ll- 1 Eaik8k
(2.1)
1
where aik is the ik cofactor. Hence, deg(aik) S (n- 1)d. By Lemma 4.4.1, these derivations satisfy
[Di,Fj] for 1 S i,j
= 8;j and
[Di,Dj ]
=0
S n.
We shall use the Di to define an An-module structure on K[X,Ll-I] as follows: Xi
-g = F i • q,
8,. - 9 = D,.(q), where q E K[X,Ll-lJ. A routine argument using Appendix 1 shows that • gives a well-defined action of An-module on K[X, Ll- 1 ]. We denote this module by M(F). LEMMA 2.2. As an An-module, M(F) is holonomic and its multiplicity can-
not exceed 2n(2OO + 1)". PROOF: The proof follows the argument of Theorem 10.3.2. For v E JIl, put
r" = {g. Ll- 2" We show that
{rv},,~o
E K[X, .1- 1] : deg(g)
is a filtration of M(F).
Let us show first that B , ' r" because Bi
= B;' 8j
S 2v(2OO + I)}.
~
r 1>+i.
It is enough to prove this for i = 1,
Let q = gLl-2v E rv. Using the cllain rule, we have that _
q = Di(g).1- 2V
+ (-2v)g.1-·(2v+l)D
j
(.1).
19. Stability
175
Substituting for Di the formula in (2.1) we get
8 i • q = ..1- 2(,,+1)
..1
(
~ aik8,.(g) -
2vg
~a
ik 8 k (..1»)
.
Since deg(..1) ~ nd and deg(aik) ~ (n-l)d we conclude that 8. q E similar argument shows that
E
Xi. q
If q E K[X,..1- I ) then q
where 9 E K[X) has degree sand r ~ O. Put v
g(Ll 2 ,,-r)..1-2" and since v deg(g..1 211Thus g..1-r E
r",
r
~ s
)
~
r"
= max{r,s}.
=
g..1-r
Thus q
=
s,
+ (2v -
r)nd ~ s
which proves that
On the other hand,
A
r v+l.
ur" = M(F).
Finally we show that
r,,+l.
+ 2vnd ~ 2v(2nd + 1).
{r,,},,~o is
a filtration of M(F).
is a K-vector space of dimension equal to that of
the subspace of K[X) of polynomials of degree ~ 2v(2nd + 1). Hence, dimK
r"
2"(2nd + l)n n I v n.
~
+ terms of smaller degree in
v.
By Lemma 10.3.1, the module M(F) is holonomic. We are now ready to prove the main theorem of this section. Its purpose is to give a global bound on the number of elements in the inverse image
F-1(P) of a point P E K". Let P
= (PI, ... ,Pn ). It
is easy to see that the
number of elements in F-I(P) equals the number of solutions of the system FI - PI
= ... = F" -
Pn
= O.
2.3 THEOREM. Let F : Kn
~
Kn be a polynomial map. If detJ(F) of 0
everywhere'in Kn, then tllere exists a positive integer b such that F-l(P) does not have more than b" points for every P E Kn. PROOF: Let P E Kn and consider the polynomial map F - P. Since J(F -
P)
= J(F),
we have that ..1
Kn. Put M(P)
= M(F -
= detJ(F -
P)
= detJ(F) of
P). Note that M(P)
= K[X,
0 everywhere in
..1- 1 )
for all P E K",
it is only the action of An on M(P) that depends on P. By Lemma 2.2, M(P) is holonomic and its multiplicity is 2n(2nd+ 1)" By Theorem 18.1.4, n
M(?)! ~(Fi - Pi)M(P)
(
"
\/'
= M(P)! ~ Xi. M(P) )
= b.
176
19. Stability
is a vector space over K of dimension
~
b. In particular, the classes of
1,xl,XI, ... ,xt in M(P)/E~XiM(P) must be linearly dependent. Thus there exists a polynomial gl (XI) E K[xd of degree ~ b and a positive integer r such that n
Ll r • g(xI) E ~)Fj - Pj)K[XJ.
(2.4) Finally, if Q
= (QI. ... , Qn)
E Kn satisfies F(Q)
=
P, then by (2.4) we
have that Ll(Q)" . g(QI) = O. Since Ll has no zeros on Kn, it follows that
g(QI)
= O.
Hence there are at most b possibilities for the first coordinate of
Q. Arguing similarly for the other coordinates we have that F-I(P) cannot have more than bn elements. Since b is independent of P the theorem follows. Theorem 2.3 and its proof are due to A. van den Essen [van den Essen 91J.
In the special case K
= lR this result follows from topological arguments; see
[Bochnak, Coste and Roy 87, Theorem 11.5.2J. 3. GLOBAL STABILITY ON THE PLANE
We may now conclude the story that began in §1, by proving Theorem 1.5 on the global asymptotical stability of polynomial systems on the plane. The main ingredients of the proof are Theorem 2.3 and the following result. THEOREM 3.1. Let F E:F. If there exist positive constants p and r such
that IF(X) I ~ p wIlenever
IXI
~
r
then tIle origin is a globally asymptotically stable point of tIle system X = F(X). Since the proof of this Theorem is purely analytic and not very
straigh~
forward, we will not give it here. It was first proved in [Olech 63J where it follows from an application of Green's theorem. See also [Casull, Libre and Sotomayor 91J. Let us show how Theorems 2.3 and 3.1 can be used to prove Theorem 1.5. PROOF OF THEOREM 1.5; Since FE :F it follows from Theorem 2.3 that
177
19. Stability
Let P E ]R2 be a point at which the maximum is attained: that is #F- 1 (P)
=
K. Let Ql, .. " QK be the elements of P-l(P). By the inverse function
theorem F is invertible in the neighbour hood of every point of]R2. Hence for 1~ i
~
K, it is possible to choose p > 0 and a neighbour1iC'od V; of Qi such
that
is a diffeomorphism, where Bp(P) is the open ball centred on P of radius p. By decreasing the value of p, if necessary, we may also assume that V;n'lf = 0 if i of;'. Let us prove that, under these hypotheses, (3.2) It is clear that the union of the V;'s is contained in F-1(Bp(P». We prove
the opposite inclusion. Suppose, by contradiction, that it does not hold. Thus there exists a point W not in \{ U '" U \f( such that F( W) E Bp(P). Since F( V;)
= Bp(P),
there are points Y; E V; such that
F(Y;)
= P(W)
for i = 1, ... , K. Note that if i of j then Y; Furthermore Y;
i'
i' lj,
because V; n
If
=
0.
W since W does not belong to the union \{ U ... U \f(.
Hence, {W,}], ... ,
lId ~ F-1(F( W).
+1
Thus F-l (F( W) cannot have less than K
elements, a contradiction.
Thus (3.2) holds. Now choose 1"
> 0 so large that B,,(O) contains \{ U··· U lX. For this
r'
and the previously chosen p we have that IF(X) - PI ~ p if IXI ~
(3.3)
r'.
= F(X +Ql)-P, It follows from (3.3) that G(X) satisfies the hypothesis of Theorem 3.1 for r = r' + IQll. Hence the Consider the translated function G(X)
system
X=
G(X) has the origin as a globally asymptotically stablr point.
Thus by Proposition 1.6 the map G is injective. But for i CeQ; - QI)
= FCQi) -
P
\
=0=
C(O).
= 1,. /;K,
178
19. Stability
Since G is injective, we must have that K = 1. This means that F- 1 (P) has at most one point for any P E ]R2. Thus (3.3) is satisfied by P
= O.
But
this is the hypothesis of Theorem 3.1, from which Theorem 1.5 immediately follows. 4. EXERCISES
4.1 Let A be a 2 x 2 matrix with real coefficients. Show that the origin is a globally asymptotically stable point of the system
X=
A . X if and only if
the real part of the eigenvalues of A are negative. 4.2 Let F E:F. Show that if F is globally invertible in ]R2 then the origin is
a globally asymptotically stable point of the system
X = F(X).
Hint: By the inverse function theorem there exist p, r > 0 such that F maps
Bp(O) into Br(O). Since F is globally one-to-one the points outside Bp(O) must be sent outside Br(O). But this is the hypothesis of Theorem 3.1. 4.3 Let F E ]R[x, V].
Use Green's theorem to show that if det.J(F)
everywhere on ]R2 then F is a map of]R2 that preserves area.
=1
CHAPTER 20 AUTOMATIC PROOF OF IDENTITIES Special functions are back in fashion. Recent years have seen the development of new approaches to the theory and also of many applications of special function identities. Two important examples of the latter are de Branges' proof of Bieberbach's conjecture, [de Branges 85], and Apery's proof of the irrationality of «(3), [van der Poor ten 79). Holonomic modules can be used to find the differential equations satisfied by certain special functions and also to determine whether a given identity is true. In many cases this can be done automatically; that is, by a computer. In this chapter we give an introduction to the theoretical foundations of this approach, pioneered by Zeilberger and his collaborators. 1. HOLONOMIC FUNCTIONS.
We will assume throughout this chapter that the coefficient field is lR. Thus
An will stand for An(lR). A function is holonomic if it is the solution of a holonomic module. Let U be an open set of ]Rn. If f E COO ( U), set
It = {d E An(R) : d(f) = a}. Then f is holonomic if Ani II is a holonomic module. But Ani II is isomorphic to the sub mod ule Anf of Coo (U) generated by f. Thus f is a holonomic function if and only if An! is a holonomic module. A polynomial
f E lR[xJ, ... ,x n ) is a holonomic function. Indeed, if f has
total degree k, then II contains the monomials [J<> with
lal
= k
+ 1.
Thus
Ani II is holonomic, see Exercise 4.1. The next result is of great help in producing examples of holonomic functions. 1.1 PROPOSITION. J-et f,9 E COO(U) be holonomic functions. Then f
and f 9
al'e
+9
holonomic.
Since f and 9 are holonomic functions, th~ modules An! ".nd An9 are holonomic. Hence An/$An9 and its sub quotient AnCf +g) ar(.-tl~lonomic, by Proposition 10.1.1. Thus f + 9 i~ a holonomic function.
PROOF:
180
20. Automatic proof
Let M be the An-submodule of COO(U) generated by 8 (f)tJP(g) for all a,{3 E Nn. It follows from Leibniz's rule that An(fg) ~ M. Now, the Q
multiplication map defines a homomorphism of An-modules,
whose image is M. Since Anf
@R[X]
Ang is holonomic by Exercise 18.4.1, it
follows that AnU g) is holonomic by Proposition 10.1.1. The composition of holonomic functions, however, is not holonomic. As we have seen in eh. 5, §3, the function exp(exp(x) does not satisfy a differential equation with polynomial coefficients. Hence, although the exponential function is holollomic, the composite exp(exp(x» is not holonomic.
2. HYPEREXPONENTIAL FUNCTIONS. Let ai, a:/, ... be a sequence of real numbers. It is called a geometric sequence, if the sequence of ratios an+I! an is constant. A natural generalization is to assume that the quotients an+1 / an are a rational function of n. These sequences are called hype1'geomet·ric. In the realm of functions, the object that corresponds to a geometric sequence is the exponential function, which satisfies:
f' / f
is constant. This
suggests that a function f E Coo (U) should be called hyperexponential if
8;(1)/f is a rational function of Xl,' .. , Xn for i
= 1, ... ,n.
Note that if q
is a rational function in R(X), then exp(q) is hyperexponential. To produce more examples, one may use the next result.
2.1 PROPOSITION. TIle pl'Oduct of hyperexponential functions is hyperexponential.
PROOF: Let
f, 9 be
hyperexponential functions. By Leibniz's rule,
8M g) = 8;(1) + 8;(g). 19
f
9
The result follows from the fact that the sum of rational functions is a rational function.
20. Automatic prool
181
We will now show that hyperexponential functions are holonomic. This is easy in one variable. If 1 is hyperexponential, then 8(1)/1 are polynomials. Thus
= plq, where p, q
1 satisfies the differential equation
= o.
(q8 - p)(1)
Equivalently,l is a solution of the module AI/A 1 (q8 - p). Since q =I- 0 and we are in dimension 1, this module is holonomic. Thus The
1 is holonomic.
case is very much harder. Suppose that 1 is a hyperex-
~dimensional
ponential function of n variables. Then 8i(1) 11 = p;1 qj, where Pi, qi E JR[X]. Thus 1 satisfies the system of differential equations
(q i8i - Pi)(f) = 0 for 1 ~ i ~ n. Let J
= E~ An(qi 8i - Pi).
Then J ~ If·
Put M = Ani J and let q be the least common multiple of ql, ... , qn. Denote by M[q-l] the module JR[X, q-l] @R[x] M; see eh. 12, §5. Every element of M[q-l] can be written in the form q-k that the action of 8; on q-k
@ u
@
u, where u EM. Recall
E M[q-l] is defined by
2.2 THEOREM. If M[q-l] is holonomic, then
1 is holonomic.
PROOF: Consider the map
defined by cP(d+ If)
= 1@(d+If)'
By Proposition 12.5.1, the element d+If
is in the kernel of cP if and only if ld E If for some positive integer k. This happens if td(f)
= O.
Since q is a polynomial, we must have that d(f)
= O.
Hence d E If. Summing up: the homomorphism cP is injective. Now, since J
~
h,
Thus JR[X,q-l]
@
AnlIf is a homomorphic image of M[q-l], by Theorem
it follows that AnlIf is a homomorphic image of M.
12.4.6. Hence JR[X, q-l]
@
Ani If is holonomic over An. But we have seen
that AniIf is a submodule of JR[X, q-l] Hence f is a holonomic function.
@
AniIf, thus it is itself holonomic. )
Let us prove that M[q-l] is a holonomic A~-module. We b..gr~ with a technical lemma.
20. Automatic proof
182
2.3 LEMMA. Let s = max { deg(Pi) : 1 ~ i ~ n} of An of degree k, then
+ deg(q).
If d is an operator
ld == 9 (mod J) where 9 is a polynomial of degree PROOF:
~
ks.
The operator d is a finite linear combination of monomials xc>EI' and
q commutes with xC>. Thus we have only to prove the following statement:
jf f3 E Ilin and 1f31 ~
= k,
then lEI' == 9 (mod J) where 9 E JR[X] has degree
ks.
i
=
=
1, then EI'
=
8i for some 1, ... ,n. Note that since q is the least common multiple of ql, ... , qn, jt
We will proceed by induction on k. If k follows that
q/ qi is a polynomial.
Thus,
q8i == Pi(q/qi) and deg(Pi(q/qi» ~ deg(Pi)
+ deg(q/qi)
(mod J). ~
s. Assume that the result holds for k-l and let us prove that lEI' E JR[X]+J
when 1f31
=k~
2. Since 1f31 > 0, it follows that f3i
i' 0 for
some i
= 1, ... ,n.
Now
(2.4) But
l8;
= 8; .£/- k£/-18;(q).
By the induction hypothesis there exists 9 E JR[X] of degree
~
(k - l)s such
that qk-1El'-c, == g(modJ). Substituting in (2.4) we obtain
£/fJ3 == 8; . q. 9 - kg8;(q) But 8;· qg
= g(q8;) + 8;(qg).
(mod J).
Since q8; == Pi(q/qi) (mod J), we conclude
that
lfJ3 == gPi(q/q;)
+ 8;(gq) -
kg8;(q) (rrwdJ).
The right hand side is a polynomial of degree less than or equal to max{(k -1)8
+ s, (s(k - 1) + 8)
-1,8(k -1)
+ (8 - I)} ~ 8k
as required. The proof of the theorem follows the argument already used in Theorem 10.3.2 and Oh. 12, §5.
20. Automatic pmol 2.5
THEOREM.
PROOF:
Let m
183
The An-module M[q-lj is holonomic.
= deg(q)
and
r
be the good filtration of M induced by the
Bernstein filtration of An. Define
This is a filtration of M[q-l] as shown in Theorem 12.5.4. By Lemma 2.3,
where 9 is a polynomial of degree
5 s(m+ l)k. Thus, as a real vector space,
By Lemma 10.3.1, it follows that M[q-l] is holonomic. Putting together Theorems 2.2 and 2.5 we have the required result.
2.6
THEOREM.
A hypel'exponential function in n val'iables is llOlonomic.
A more general result can be found in [Takayama 92, Appendix]. In the next section we explain the theoretical foundations of an algorithm which calculates the differential equation satisfied by a given definite integral with parameters. As an example, we calculate the integral of a hyperexponential function. 3.
THE METHOD.
Denote by x, y the coordinate functions of R2. Let U = (a, b) x (-00, (0) and
I
(3.1)
be a function in Coo (U) which satisfies
lim xQ{/3 lex, y) = 0
y ......... ±oo
for all n, j3 E N2 This implies that if P E A 2 , then
J""
.. "" f'(J)dy < 00.
"
20. A utomatic proof
184 for every x E
Ca, b).
In particular this is the case for P
= 1.
Put
+00
R(x)
=
1
-00
lex, y)dy.
We will describe an algorithm which finds a differential equation satisfied by
R(x), when f is a given holonomic function. Let M = A2/If and
11" :
R Z -+ R be the projection on the first coordinate.
The direct image of M under
If we now assume that
11"
is
f is a holonomic function, then
M is a holonomic
A2-module. Thus 1I"*M is holonomic over Al by Theorem 18.2.2. Hence the kernel of the homomorphism
which maps dEAl to d + (If
+ oyAz)
must be non-zero. Thus there exists
a non-zero operator DEAl such that
where QI E If and Qz E A z. Since QI (I)
= 0, we have that
(3.2) Integrating the right hand side of (3.2) between
-00
and
00
and using the
fundamental theorem of calculus,
Thus, integrating the left hand side of (3.2) and using differentiation under the integral sign [Buck 56, Ch. 4, §4.4, Theorem 29], one gets
o=
I:
D(f)dy = DCR).
185
20. Automatic proof Summing up: R(x) satisfies the differential equation D(R) = O.
We will apply the method to an example, borrowed from [Almkvist and Zeilberger 90j. Let f(x) = exp(-(x/y)2 - y2) and R(x) be its integral between
-00
and
00.
The function
f
is hyperexponential, hence holonomic by
Theorem 2.6. It is easy to check that it satisfies condition (3.1). A simple calculation shows that
f
is a solution of the equations
(y 28",
+ 2x)(f) = 0, (y3 8 y + 2y4 - 2X2)(f) = O. Let L
= It + 8 y A 2 .
Since y38 y
= 8y ' ~ -
3y2, we have that
We are allowed to multiply this identity by ~ on the left, because L is an A 1suhmodule of A 2 • Doing this and using the identity y28",
== -2x (mod J),
three times, we get that
D = 68", . x
+ 88", . x 2 + 8x -
2~ . x 2
is in L. Now D factors as
(x8",
+ 3)(8", -
2)(8", + 2).
In this special case we may use the differential equation D(R)
0 to
calculate R. The general solution of this equation is of the form Cl
exp(2x)
+ C2exP( -2x) + C3h,
where h satisfies the differential equation
for some constant A
of O.
We must now decide, among the possible solutions, which one coincides with R(x). First of all, h cannot be bounded at O. But
186 thus
20. A utomatic proof C3
= O. Furthermore,
limx--->+oo
R(x)
= 0, hence Cl = O. We are left with
R(x) = C2exp(-2x), and so
Therefore, R(x)
= (.j7l'tl exp(-2x). This is to be taken as a mere illustra-
tion: no one would dream of calculating this particular integral in this way. For the easy solutions, see Exercise 4.5.
In the above example we found D from the equations for f by an ad hoc calculation. Takayama showed in [Takayama 92] that one may use Grabner bases to determine D. If f is hyperexponential, there is also an algorithm in [Almkvist and Zeilberger 90], based on Gosper's summation algorithm [Gosper 78]. Thus the calculation of D can be done automatically.
If one tries to extend the algorithm to find R(x) by solving D(R) = 0 like we did, the difficulties are of another order of magnitude. There are two main problems: (1) The integration D(R)
= O.
(2) Checking initial conditions. One way to get around (1) is to settle for an algorithm to certify an identity.
In other words, we have 'guessed' R(x) by whatever method, and we want to show that our guess is correct. That still leaves (2). Since the equation D(R)
= 0 is usually of order greater than 1, checking initial conditions can
be very tricky. However, hyperexponential functions are like hypergeometric sequences, and integrals like sums; and it turns out that these ideas can be used to certify identities of hypcrgeometric sums. In this case (2) does not present any problem and the algorithm is very effective. However, we are no longer dealing with modules over the Weyl algebra. The base ring in this case is a localization of the Weyl algebra. Many interesting applications of these algorithms are discussed in [Zeilberger 90], [Almkvist and Zeilberger 90] and [Cartier 92]. 4. EXERCISES
4.1 Let J k be the left ideal of An generated by the monomials 8 a with Show tbat AnI J k is a hojolllllllic A,,-mod1Lk.
lal
= k.
20. Automatic prool
187
Hint: There exists an exact sequence,
4.2 Show that the derivative of a hyperexponential function is hyperexponential. 4.3 Find an upper bound for the multiplicity of the module A 11,/ If' when
I
is a hyperexponential function. Hint: What is the bound on the multiplicity of M[q-l] in Theorem 2.5'1 4.4 Show that the function sin(xy) is holonomic, but not hyperexponential.
4.5 Let I(x, y) = exp( _(x/y)2_1l) and R(x) be the integral of I with respect
to y between
-00
and
00.
Calculate R(x) in two different ways, as follows:
(1) Using the substitution t= y- x/y in
J::' exp(-t2)dt= (.j7l')-1.
(2) Using differentiation under the integral sign to obtain the equation
R' = -2R and integrating it. 4.6 Let I(x,y) = exp(-xy2) and R(x) = J~oo/(x,y)dy. Find an operator
D E Al such that D(R) = O.
4.7 Let Ah" . , A, E 1R" and let X be the ~tuple (Xh . .. , x 71 ). Denote by Aj'X
the formal inner product of the two n-tuples. For Pb . .. ,p, E K[Xh . .. ,xnJ put
•
I (x) = L Pi (X) eXp(Aj . X). 1
Show that
I
is a holonomic function.
CODA Since this book is only a primer, it is convenient to give the interested reader directions for further study. The comments that follow are based on this author's experience and inevitably reflect his tastes. First of all, the theory of algebraic D-modules is itself a part of algebraic geometry. Thus we must start with an algebraic variety X, If we assume that X is affine, then its algebraic geometric properties are coded by the ring of polynomial functions on X (and its modules). This is a conunutative ring, called the ring of cOO1"dinates and denoted by O(X). The ring of differential operators D(X) is the ring of differential operators of O(X) as defined in eh. 3. If the variety is smooth (non-singular) then D(X) is a simple noetherian ring. To deal with general varieties it is necessary to introduce sheaves. The
structure sheafkeeps the same relation to a general variety as the coordinate ring does to an affine variety. From it we may derive the sheaf of 1'ings of diffe1'ential opemto·I's. If the variety is smooth, this is a coherent sheaf
of rings. The purpose of D-module theory is the study of the category of coherent sheaves of modules over the sheaf of rings of differential operators of an algebraic variety. It is plain that a good knowledge of algebraic geometry is essential to
make sense of these statements. The standard reference is the first three chapters of [Hartshorne 77J. One can also find the required sheaf theory in Serre's beautiful "Faisceaux algebriques coMrents", [Serre 55]. But a thorough grounding in classical algebraic geometry is necessary before one tackles this paper. Two good references for that are [Harris 92J and the first chapter of [Hartshorne 77J. With this background it is possible to extend most of the material covered in the book to general varieties, including a full version of Kashiwara's theorem. Unfortunately there is no introduction to algebraic D-modules assuming only a good knowledge of sheaf theory. In this respect its twin sister, the theory of analytic D-modules, is better served. Analytic D-modules have an ana1ytic (instead of algebraic) manifold for base space .. The two theories
Coda
189
are very close, but there are some differences. However, with a little care one should be able to use a book on analytic D-modules to learn about the algebraic theory. An excellent introduction is [Pham 79]. A more recent work in the same spirit is [Granger and Maisonobe 93]. See also [Bjork 79]. Since we have mentioned family relationships: D-module theory has another close relative in the theory of modules over rings of microdifferential operators; see [Schapira 85]. It is always good to have a substantial result to aim at when one is first
learning a theory. There is Kashiwara's theorem, of course. Another result that one could tackle is the theorem of Beilinson and Bernstein on the structure of modules over the ring of differential operators of projective space: [Beilinson and Bernstein 81]. This has the added advantage that the theorem was one of the main steps in the proof of a famous conjecture in representation theory, see [Kirwan 88]. The Beilinson-Bernstein theorem is proved in [Borel et a1. 87, eh. VII §8] and [Benoist 93]. The background so far described is not enough when one comes to study direct images and the structure theory of holonomic modules. For example, direct images can be easily generalized to varieties following the principles of eh. 16, but this definition will not produce a functor. The problem is that Theorem 18.2.1 will break down for general varieties. To get over this problem it is necessary to give up the categories of sheaves of D-modules and work instead with their derived categories. Derived categories belong to homological algebra. It turns out that homological algebra is, right from the beginning, one of the essential tools in the study of D-modules. In fact, to avoid using homological algebra we had to make great sacrifices in this book. One obvious point has to do with the duality properties of holonomic modules, which are defined using the functor Ext, see [Borel et a1. 87, eh. V] and [Milicic 86].
Unfortunately it is not possible to come to a deep understanding of Dmodules without derived categories. This is a very technical subject. A good starting pOint is [Iversen 86]. For a more detaile
190
Coda
Riemann-Hilbe7'i c01"respondence, one of the jewels of this theory,
A lot of background is required before one gets to the heart of D~module theory, which means a rather long preparation. Is it worth it? The answer must be yes: the ascent may be steep, but the view is truly breathtaking. Besides, the power of this machinery is such that applications are legion and contacts with other areas of mathematics the rule.
APPENDIX 1 DEFINING THE ACTION OF A MODULE USING GENERATORS In Ch. 1 we defined the Weyl algebra in two different ways. First as a subalgebra generated by certain endomorphisms of the polynomial ring. Second as the quotient of a free algebra by an ideal of relations. In both cases, an arbitrary element of the Weyl algebra is only determined a posteri01-i, as a linear combination of monomials on the generators. So when it comes to defining an action of a Weyl algebra on a vector space, it should be enough to say how the generators are to act on the vectors. This would be the case if the generators were not bound up by relations. It is the relations that dictate which actions are well-defined and which are not. This is a very simple matter, but it is usually not carefully discussed in elementary books. It plays such an important role in this book that it seems better to treat it in detail in this appendix. We begin with a very general setup. Let R be a K-algebra and let M be a left R-module. The action of R on M gives rise to a homomorphism of K-algebras
which maps an element a E R onto the endomorphism
tPa( u) = au,
tPa
of M defined by
for u EM. Given such a homomorphism we may easily recover
the action of R by the preceding formula. Suppose now that F is a free algebra and that there is a surjective homomorphism
71' :
F
~
R. The composition tP· 71' gives a map from F to EndK (M)
that makes M into an F -module. Besides K er( 71')M
= O.
This last equation
tells us that M is not only an F- module, but in fact an R-module. The recipe for checking whether a proposed action of generators is welldefined is inspired by the setup above. Assume now that the algebra R is generated by ai, ... ,ak' Let F be the free algebra ~n k generators Xl, ... ,Xk and let IVI be a K -vector space. Suppose that an action aiU has been postulated for eVE'ry U E M and i t.r1lly an actioIl of R on M?
=
1, ... ,k. How can we check whether this is
192
Appendix 1
First of all put
XiU
= ai'lL.
In this way we have defined an action of F (and
not just its generators) on M, because the generators of F do not satisfy any relations. Hence any arbitrary definition will give a good action. Thus we have a map 'IjJ: F
---t
EndI«M) such that 'IjJ,;;(U) = XjU= aiu,
F
---t
We need to know whether 'IjJ factors through
71'.
This will happen if K er7l' C;;
U
E M. There is also a map
R, given by
=
71' :
for every
71'(x;)
ai.
K er'IjJ. In other words, a well defined action of R on M has been defined if Ker7l"
M = O.
Let us apply this recipe to the Weyl algebra. Let M be a
K-vector
space.
To define an action of AnCK) on M we begin by prescribing values for and 8i u, for every
U
E M and i
XjU
= 1, ... , n. Since the ideal of relations is
generated by a finite number of elements, the next step is very easy. All we have to do is to make sure that
and
for every u E M. This is not usually easy to do, as the following example shows. It is borrowed from Ch. 19, §2. Let FJ , • •• ,Fn be polynomials in
K[XI, ... , xn]
and suppose that there exist
pairwise commuting derivations D J , ... ,Dn such that Di(Fj)
= Oij. This will
be the case if the jacobian determinant of F1 , .•• , Fn is 1; cf. Ch. 4, §4. We would like to define an action of An on K[X] by the formulae: X;.9
= Fi9
8;·9= D i (9) where 9 E K[X]. Is this action well defined? Let us check that one of the relations, say [8;,x J ] we have that
= o;J'
is satisfied on 9 E K[X]. Using the definitions,
Appendix 1
193
Since D; is a derivation, we may apply Leibniz's rule to get
The other relations can be similarly checked, and we conclude that the action can be extended to all elements of An. The most important application of this principle occurs in the definition of inverse image in Ch. 14.
APPENDIX 2 LOCAL INVERSION THEOREM The purpose of this appendix is to give a self-contained proof of the local inversion theorem, which is used in the proof of Lemma 4.4.1. Our exposition follows [Bourbaki 59]. LOCAL INVERSION THEOREM. Let F
= (FI' ... ,Fn)
be an n-tuple of power
series in K[[XI'" ., xnll. Assume that the jacobian determinant detJ F(O)
f. 0
and that F(O) = O. Then there exist GI, ... , Gn E K[[XI"" ,x n]], without constant terms, suell that
for i
= 1, ... , n.
The proof of the theorem follows from the following lemma. LEMMA. For 1 ~ i
~
m, let
be formal power series and denote by P( Y, X) the c01'l'esponding m-tuple.
Let Ll(Y,X)
= det
(8E) ~ YJ
. l$iJ$m
Suppose that P(O,O) = 0 and that Ll(O, X) is invertible in K[[XI, ... , xnlJ. Then there exists a unique m-tuple G = (GI , ... , Gm ) in K [[Xl, ... ,xnll such that
for i
= 1, ... , m.
PROOF: We may write each Pi as a power series in the y's with coefficients in K[[XlJ
= K[[Xl,' .. ,xnlJ.
Using multi-indices, we have that In
Pi = ail
+L I
aij Yj
+
L 1"1>2
air>Y'\
195
Appendix 2
where
atO, aij
and
aio
belong to K[[X]] and
all)
has no constant term. We
begin by showing that it is enough to prove the lemma in a special case. By hypothesis, the matrix
(aij)l:5i,j:<>m
is invertible over K[[X]]. Let us
denote its inverse by (h ij )):5i,j:5m' Put m
Qi
= .E hij?;. )
Then we have Qi
= -CIl) + Yi -
.E ciaya lal?:2
where caO, Cia E K[[XJl and
CIl)
has no constant term. It is enough to solve
the problem for the Q's, because Pi Suppose now that series Gb
... ,
= I:;n aijQj.
Gm have been obtained which satisfY
Thus,
(1)
Gi
= Cill + .E ciaGo. lal?:2
Let
Sik
be the sum of the homogeneous components of degree
Given a multi-index
Q
= (a), ... , am)
~
k of G ik .
we will write
Sf = Sft ... S~. Since the G's have no constant term, the component of degree k of G'''', for
lal 2:
2, coincides with the component of degree k of
Sf_I'
It follows
from (1) that S) is equal to the homogeneous component of degree 1 of Cill. Furthermore, for all k > 1 the homogeneous component of degree k of G; coincides with the corresponding component of CIl)
-+ L 1"12'2
CiaSk_1
(-
Appendix 2
196
and can thus be determined by recursion. Note that since the coefficients of G; are recursively determined using (1), it follows that G j itself is unique. Let us prove that the local inversion theorem follows from this lemma. PROOF OF THE LOCAL INVERSION THEOREM: Denote by P the power series in K[[Yl, ... , Yn, Xl,' ..
for 1 ::;
i::;
11.
Then P(O, 0) =
,
xnll
~tuple
of
whose components are
a and
..1(Y,X) = detJF(y), so that ..1(0, X)
= detJ F(O)
is a non-zero constant. Hence we may apply the
lemma, which asserts that there exist power series G1 , ... , Gn E K[[X]] such that
as claimed in the theorem.
REFERENCES
Almkvist, G. and Zeilberger, D. 1990 The method of differentiation under the integral sign, J. Symbolic Computation, 10, 571-591. Arnold, V.l. 1981 Ordinary differential equations, The MIT Press, Cambridge, Mass.London. Barabanov, N. E. 1988 On a problem of Kalman, Siberian Math. J.,29,333-341. Bass, H., Connell, E.H. and Wright, D. 1982 The Jacobian Conjecture: reduction of degree and formal expansion of the inverse, Bull. Amer. Math. Soc. {New Series), 7, 287-330. Beilinson, A. and Bernstein, J. 1981 Localisation de g-modules, C.R. Acad. Sci. Pm'is, 292, 15-18. Benoist, Y. 1993 D-module sur la variete des drapeaux, in Images directes et constructibiliU, ed. P. Maisonobe et C. Sabbah, Travaux en Cours, 46, Hermann, Paris. Bernstein, LN. 1971 Modules over a ring of differential operators: study of the fundamental solutions of equations with constant coefficients, Funct. Anal. Appl., 5, 89-101. 1972 The analytic continuation of generalized functions with respect to a parameter, Funct. Anal. Appl., 6, 273-285. Bernstein J. and Lunts, V. 1988 On non-holonomic irreducible D-modules, Invent. Math., 94,<223-243. Bjork, J.-E. 1979 Rings of d'ifferential opemto1's, North Holland Mathematics Library 21, Amsterdam.
198
References
Bochnak, J., Coste, M. and Roy, M.-F. 1987 Geometrie algebrique reelle, Ergebnisse der Math. U.L Grenzgebieten, Springer-Verlag, Berlin-Heidelberg-New York. Borel, A. et al. 1987 Algebmic D-modules, Perspectives in Mathematics 2, Academic Press, London-New York. Bourbaki, N. 1959 Algebre, chapitre 4: polynomes et fmctions rationelles, deuxieme edition, Actualites Scientifiques et Industrielles 1102, Hermann, Paris. Brianc;on, J. and P. Maisonobe 1984 Ideaux de germes d'operateurs differentiels a une variable, Ens. Math., 30, 7-38. Buck, R.C. 1956 Advanced calculus, McGraw-Hill Book Company, New York-TorontoLondon. Cartier, P. 1992 Demonstrations "automatique" d'identites et fonctions hypergeometri-
ques [d'apres D. Zeilberger], Seminaire Bourbaki 1991/92, Asterisque, 206, 41-91. Cassou-Nogues, P. 1986 Racines de polyn6mes de Bernstein, Ann. Inst. Fourier, Grenoble, 36, 1-30.
Cohn, P.M. 1979 Algebm, vol. 2, John Wiley and Sons, Chichester-New York-BrisbaneToronto-Singapore.
1984 Algebm, vol. 1, second edition, John Wiley and Sons, Chichester-New York-Bris bane-Toronto-Singapore. Coutinho, S.C. and Holland, M.P. 1988 Module structure of rings of differential operators, Pmc. Math. Soc., 57, 417-432.
London
de Branges, L. 1985 A proof of the Bieberbach conjecture, Acta Math., 154, 1:17-152.
References
199
Dirac, P.A.M. 1978 The development of quantum mechanics, in Dirac; Di1'ections in Physics, ed. H. Hora and J.R. Shepanski, John Wiley and Sons, New YorkLondon-Sidney-Toronto. Dixmier, J. 1963 Representations irreductibles des algebres de Lie nilpotentes, Anais Acad. Bras. Cienc., 35, 491-519. 1968 Sur les algebres de Weyl, Bull. Soc. Math. France, 96, 209-242. 1974 Algebres enveloppantes, Cahiers Scientifiques 37, Gauthier-Villars, Paris.
Gabber, O. 1981 The integrability of the cllaracteristic variety, Ame1'. J. Math., 103, 445-468.
Gasull, A., Llibre, J. and Sotomayor, J. 1991 Global asymptotic stability of differential equations in the plane, J. Diff. Eq., 91, 327-335. Gosper Jr, R.W. 1978 Decision procedure of indefinite summation, Proc. Nat. Acad. Sci. USA, 75, 40-42. Granger, M. and Maisonobe, P. 1993 A basic course on differential modules, in D-modules coherents et holonomes, ed. P. Maisonobe et C. Sabbah, Travaux en Cours 45, Hermann, Paris. Gutierrez, C. 1993 A solution to the bidimensional global asymptotic stability conjecture, Informes de Matematica IMPA, Serie A-097. Hardy, G.H. 1928 The integration of functions of a single variable, Cambridge Tracts in Mathematics and Mathematical Physics 2, Cambridge University Press, Lon~n. ( Harris, J. 1992 Algeb'I'wie uernnet'I'Y: a first 133, Sprillger-Verlag, New York.
COU1'se,
Graduate Texts in Mathematics
200
References
Hartshorne, R. 1977 Algebraic geometry, Graduate Texts in Mathematics 52, SpringerVerlag, New York-Heidelberg-Berlin. Iversen, B. 1986 Cohomology of sheaves, Springer-Verlag, Heidelberg. Kashiwara, M. 1976 B-functions and holonomic systems: rationality of roots of b-functions, Invent. Math., 38, 33-53. Keller, O.H. 1939 Ganze Cremona-Transformationen, Monatsh. fu?' Math. und Phys. (Leipzig und Wiell), 47, 299-306. Kirwan, F. 1988 An introduction to intersection homology theory, Pitman Research Notes in Mathematics 187, Longman Scientific and Technical, Harlow, Essex. Krause, G. and Lenagan, T.H. 1985 Growth of algebras and Gelfand-Kirillov dimension, Research Notes in Mathematics 116, Pitman, London. Littlewood, D.E. 1933 On the classification of algebras, Proc. 200-240.
London Math.
Soc., 35,
Lunts, V. 1989 Algebraic varieties preserved by generic flows, Duke Math. J., 58, 531-554. Malgrange, B. 1976 Le polynome de Bernstein d'ulle singularite isoiee, Lecture Notes in Math. 459, Springer-Verlag, 98-119. 1991 Equations diife?""entielles Ii coefficients polynomiaux, Progress in Mathematics 96, Birkhiiuser, Boston, Mass.-Basel-Berlin.
Manin, Yu. I. 1988 Quantum groups and non-commutative geometry, Publications du Centre de Redwfches MatitcmatiqulJs, Moutnbl.
References
201
Markus, L. and Yamabe, H. 1960 Global stability criteria for differential systems, Osaka Math. J., 12, 305-317. Matsumura, H. 1986 Commutative 7'ing theory, Cambridge Studies in Advanced mathematics 8, Cambridge University Press, Cambridge. McConnell, J.C. and Robson, J.C. 1987 Noncommutative noethe7'ian rings, Wiley Series in Pure and Applied Mathematics, John Wiley and Sons, Chichester-New York-Brisbane-TorontoSingapore. Meisters G. and Olech, C. 1988 Solution of the global asymptotic stability Jacobian conjecture for the polynomial case, Analyse Matfu!.matique et Applications, Gauthier-Villars, Paris. Milicic, D. 1986 Lectures on algebraic theory of D-modules, University of Utah. Olech, C. 1963 On the global stability of an autonomous system on the plane, Contributions to Differential equations, 1, 389-400. Pham, F. 1979 SingulariUs des systemes diffc,'entiels de Gauss-Manin, Progress in Mathematics 2, Birkhiiuser, Boston. Pinchuk, S. 1994 A counterexample to the strong real jacobian conjecture, Math. Z., 217, 1-4. Reiten, L 1985 An introduction to the representation theory of Artin algebras, Bull. London Math. Soc., 17,209-233. Rudin, W. 1991 Functional analysis, second edition, International Series in P.h-~ and Applied Mathematics, McGraw-Hill International Edition, Singapore. Schapira, P. 1985 l\IlicmdijJe'/"'eni1:al systems in the complex domain, Grundlehren der mathematisdwll Wissellsehaften 269, Sprinp;er-Verlag, TI0flin-Heidelberg.
202
References
Segal, I. 1968 Quantized differential forms, Topology, 7, 147-171. Serre, J .-P. 1955 Faisceaux algebriques coMrents, Ann. of Math., 61, 197-278. Smith, S.P. 1986 Differential operators on Ai and II"I in charp > 0, in Seminaire Dubreil-Malliavin 1985, ed. M.- P. Malliavin, Lecture Notes in Mathematics 1220, Springer-Verlag, Berlin- New York, 157-177. Stafford, J. T. 1978 Module structure of Weyl algebras, J. London Math. Soc., 18, 429442. 1985 Non-holonomic modules over Weyl algebras and enveloping algebras, Invent. Math., 79, 619-638.
Takayama, N. 1992 An approach to the zero recognition problem by Burchberger algorithm, J. Symbolic Computation, 14, 265-282. van den Essen, A. 1991 A note on Meisters and Olech's proof of the global asymptotic stability jacobian conjecture, Pacific J. Math., 151, 351-356. van der Poorten, A. 1979 A proof that Euler missed ... Apery's proof of the irrationality of «(3), Math. Intelligence1', 1, 195-203. Weyl, H. 1950 The theory of groups and quantum mechanics, Dover, New York. Wright, D. 1981 On the Jacobian Conjecture, Illinois J. Math., 25,423-440. Yano, T. 1978 On the theory of b-functions, Publ. RIMS, Kyoto Univ., 14, 111-202. Zeilberger, D. 1990 A holonomic systems approach to special functions identities, J. Compo App. Mathematics, 32, 321-368.
INDEX
abelian category, 166
characteristic variety, 98
algebraic analysis, 4
closed ideal, 102
algebraic function, 41
co-isotropic subspace, 101
artinian
commutator, 9
- module, 88
comorphism, 27
- ring, 89
complex neighbourhood, 51
Artin-Schreier theorem, 89
composition
ascending chain condition, 66
- of direct images, 165
associated graded
- of inverse images, 140
- algebra, 57
- series, 89
- module, 60 asymptotically stable point, 171
degree of operator, 14
asympototic stability, 171
derivation, 20
automorphisms of An, 11
difference function, 74 dimension
balanced map, 111
- of Gelfand-Kirillov, 80
Bernstein
- of An-module, 77
- filtration, 56
- of variety, 99
- inequality, 83
Dirac's 6, 46
- polynomial, 94
- as hyperfunction, 51
bilinear map, 111
- as microfunction, 49
bimodule, 108
direct image - as functor, 167
canonical
- of left modules, 151
- basis, 9
- of right modules, 146
- form, 9
- under embeddings, 154
category, 166 COO-function, 42, 45
- under isomorphisms, 152 ("
change of rings, 130
- under projections, 151 direct limit, 47
characteristic ideal, 98
directed set, 46
204
Index
directed family, 46
good filtration, 70
distributions, 44
Gosper's algorithm, 186
Dixmier conjecture, 34
graded
D-module,3
- algebra, 53, 57
- of differential equation, 44
- homomorphism, 54, 55
dynamical variables, 2
- ideal, 54 - module, 55
endomorphisms of An, 16
-submodule, 55
enveloping algebra, 3
grading, 53
external product
Green's theorem, 176, 178
- of algebras, 121 - of modules, 122
Heaviside - hyper function, 51
factorial of multi-index, 10
- microfunction, 50
filtered
Hilbert
- algebra, 56
- basis theorem, 68
- module, 59
- polynomial, 74
filtration
holomorphic function, 40
- of algebra, 56
holonomic
- of module, 59
- function, 179
Fourier
- module, 86
- series, 2
homogeneous
- transform, 39
- component, 53
functor, 166
- element, 53
Fundamental theorem
hyperexponential functions, 180
of Calculus, 184
hyperfunctions, 44, 51 hypergeometric function, 4
Gelfand-Kirillov dimension, 80
hypergeometric sequences, 180
generically defined property, 106 germs of holomorphic functions, 47
induced filtration,61
global asymptotic stability, 172
inverse function
globally asymptotically stable point, 172
theorem, 28 inverse image, 132
Index inverse image
205 matrix mechanics, 2
- as functor, 166
maximal
- of An-modules, 132
- condition, 66
- of Dirac's 6 137
- element, 66
- over polynomial rings, 132
metrizable space, 50
- under embeddings, 138
microfunctions, 44, 46, 49
- under isomorphisms, 143
microdifferential operators, 189
- under projections, 134
minimal
involutive
- element, 88
- subspace, 101
- involutive variety, 105
- variety, 101
module
irreducible module, 36
- with support, 159
isotropic subspace, 101
- of hyperfunctions, 51 - of microfunctions, 49
jacobian conjecture, 28, 33
multi-index, 9
- for comorphisms, 30
multiplication map, 121
- in dimension 1, 28
multiplicity, 77
- in positive characteristic, 35 jacobian determinant, 28
noetherian
jacobian matrix, 28
- module, 65 - ring, 67
Kashiwara's theorem, 158, 168
non-singular point, 99 numerical polynomial, 74
lagrangian varieties, 102 length
order
- of composition series, 89
- filtration, 56, 63
- of multi-index, 9
- in An, 22, 56
Lie algebra, 3
- in Bn(K), 18
localization, 119, 186
- of operator, 20
local inversion
Ore condition, 84
theorem, 33, 194 locally nilpotent derivation, 30
('
Poisson bracket, 102 polynomial algebra, 8
206 polynomial - isomorphism, 26
Index skew orthogonal complement, 100
- map, 26
skew symmetic form, 100
- ring, 8, 36
solution space, 45
- solution, 44
standard
positive grading, 53
--embedding, 138
pre-ordered set, 46
-symplectic structure, 100
preservation of holonomy
- transposition, 147
- under direct image, 165
structure sheaf, 188
- under inverse image, 162, 165
sub-bimodule, 108
principal symbol, 63
support - of element, 156
quantum
- of module, 159
- algebra, 2
symbol
- group, 53
- ideal, 63
- mechanics, 1
- map, 57, 60
- plane, 53
symplectic
quivers, 50
- geometry, 100 - group, 64
regular holonomic
- structure, 100
module, 107 resolution of singularities, 4
tensor product, 109
Riemann-Hilbert
transposed
correspondence, 4, 190
- action, 148
ring of constants, 30
- bimodule, 149
ring of differential
- module, 148
operators, 21 ring of coordinates, 188
transposition, 147 twisted module, 38 torsion
simple
- element, 36
- module, 36
- module, 36
- ring, 16
uniqueness theorem
singular point, 99
for differential equations, 171
Index
207
universal
Weyl algebra
- balanced map, 111
- as differential operators, 8
- cover, 48
- as quotient of free algebra, 10
- property of tensor product, 112
- in positive characteristic, 17
wave mechanics, 2
Zariski tangent space, 99
LONDON MA1HEMATICAL SOCIETY STUDENT lEXTS Managing editor: Dr C.M. Series, Mathematics Institute University of Warwick, Coventry CV4 7AL, United Kingdom
Introduction to combinators and A-calculus, 1.R. HINDLEY & J.P. SELDIN 2
Building models by games, WILFRID HODGES
3
Local fields, .r.W.S. CASSELS
4
An Introduction to twistor theory: Second edition, SA HUGGETI & KP. TOD
5
Introduction to general relativity, L.P. HUGHSTON & KP. TOD
6
Lectures on stochastic analysis: diffusion theory, DANIEL W. STROOCK
7
The theory of evolution and dynamical systems, J. HOFBAUER & K SIGMUND
8
Summing and nuclear norms in Banach space theory, GJ.O. JAMESON
9
Automorphisms of surfaces after Nielsen and Thurston, A. CASSON & S. BLEILER
10 Nonstandard analysiS and its applications, N. CUTLAND (ed) 11 Spacetime and singularities, G. NABER 12 Undergraduate algebraic geometry, MILES REID 13 An introduction to Hankel operators, J .R. PARTINGTON 14 Combinatorial group theory: a topological approach, DANIEL E. COHEN 15 Presentations of groups, D.L. JOHNSON 16 An introduction to noncom mutative Noetherian rings, KR. GOODEARL & R.B. WARFIELD, JR 17 Aspects of quantum field theory in curved spacetime, S.A. FULLING 18 Braids and coverings: selected topics, VAGN LUNDSGAARD HANSEN 19 Steps in commutative algebra, R.Y. SHARP 20 Communication theory, C.M. GOLDIE & RG.E. PINCH 21 Representations of finite groups ofUe type, FRAN<;:OIS DIGNE & JEAN MICHEL 22 Designs, graphs, codes, and their links, P.l. CAMERON & J.H. VAN LINT 23 Complex algebraic curves. FRANCES KIRWAN 24 Lectures on elllptic curves, .r.W.S. CASSELS 25 Hyperbollc geomeu'y, BIRGER IVERSEN 26 An Introduction to the theory of L-functions and Eisenstein series, H. HIDA 27 Hilbert Space: compact operators and the trace theorem, l.R. RETHERFORD 28 Potential Theory in the Complex Plane, T. RANSFORD 29 Undergraduate Commutative Algebra, M. REID 32 Lectures on Ue Groups and Lie Algebras, R CARTER, G. SEGAL & I. MACDONALD 33 A Primer of Algebraic D-lllodules, S.C. COUTINHO r
I