Tải bản đầy đủ (.pdf) (209 trang)

multilinear algebra – d g northcott multilinear algebra – werner greub multilinear algebra – marvin marcus multilinear algebra and differential forms for beginners

Bạn đang xem bản rút gọn của tài liệu. Xem và tải ngay bản đầy đủ của tài liệu tại đây (2.26 MB, 209 trang )

<span class='text_page_counter'>(1)</span><div class='page_container' data-page=1></div>
<span class='text_page_counter'>(2)</span><div class='page_container' data-page=2></div>
<span class='text_page_counter'>(3)</span><div class='page_container' data-page=3></div>
<span class='text_page_counter'>(4)</span><div class='page_container' data-page=4>

<i><b>Multilinear algebra</b></i>



D. G. NORTHCOTT F.R.S.


<i><b>Formerly Town Trust Professor of Mathematics</b></i>
<i><b>at the University of Sheffield</b></i>


<i><b>The right of the</b></i>
<i><b>University of Cambridge</b></i>


<i><b>to print and sell</b></i>
<i><b>all manner of books</b></i>


<i><b>was granted by</b></i>
<i><b>Henry VIII in 1534.</b></i>
<i><b>The University has printed</b></i>
<i><b>and published continuously</b></i>


<i><b>since 1584.</b></i>


CAMBRIDGE UNIVERSITY PRESS


<i>Cambridge</i>


</div>
<span class='text_page_counter'>(5)</span><div class='page_container' data-page=5>

CAMBRIDGE UNIVERSITY PRESS


Cambridge, New York, Melbourne, Madrid, Cape Town, Singapore, Sao Paulo, Delhi
Cambridge University Press


The Edinburgh Building, Cambridge CB2 8RU, UK


Published in the United States of America by Cambridge University Press, New York
www. Cambridge. org



Information on this title: www.cambridge.org/9780521262699
© Cambridge University Press 1984


This publication is in copyright. Subject to statutory exception
and to the provisions of relevant collective licensing agreements,
no reproduction of any part may take place without the written
permission of Cambridge University Press.


First published 1984


This digitally printed version 2008


<i>A catalogue record for this publication is available from the British Library</i>
<i>Library of Congress Catalogue Card Number: 83-27210</i>


</div>
<span class='text_page_counter'>(6)</span><div class='page_container' data-page=6>

Contents



<i>Preface ix</i>


<b>1 Multilinear mappings 1</b>


General remarks 1
.1 Multilinear mappings 1
.2 The tensor notation 4
.3 Tensor powers of a module 6
1.4 Alternating multilinear mappings 6
.5 Symmetric multilinear mappings 10
i.6 Comments and exercises 13
.7 Solutions to selected exercises 15



<b>2 Some properties of tensor products 19</b>


General remarks 19
2.1 Basic isomorphisms 19
2.2 Tensor products of homomorphisms 22
2.3 Tensor products and direct sums 25
2.4 Additional structure 28
2.5 Covariant extension 29
2.6 Comments and exercises 31
2.7 Solutions to selected exercises 37


<b>3 Associative algebras 42</b>


</div>
<span class='text_page_counter'>(7)</span><div class='page_container' data-page=7>

<i>vi Contents</i>


3.8 Comments and exercises 58
3.9 Solutions to selected exercises 65


<b>4 The tensor algebra of a module 69</b>


General remarks 69
4.1 The tensor algebra 69
4.2 Functorial properties 72
4.3 The tensor algebra of a free module 74
4.4 Covariant extension of a tensor algebra 75
4.5 Derivations and skew derivations on a tensor algebra 76
4.6 Comments and exercises 78
4.7 Solutions to selected exercises 80



<b>5 The exterior algebra of a module 84</b>


General remarks 84
5.1 The exterior algebra 84
5.2 Functorial properties 87
5.3 The exterior algebra of a free module 89
5.4 The exterior algebra of a direct sum 93
5.5 Covariant extension of an exterior algebra 95
5.6 Skew derivations on an exterior algebra 96
5.7 Pfaffians 100
5.8 Comments and exercises 105
5.9 Solutions to selected exercises 111


<b>6 The symmetric algebra of a module 117</b>


General remarks 117
6.1 The symmetric algebra 118
6.2 Functorial properties 120
6.3 The symmetric algebra of a free module 121
6.4 The symmetric algebra of a direct sum 121
6.5 Covariant extension of a symmetric algebra 122
6.6 Derivations on a symmetric algebra 123
6.7 Differential operators 124
6.8 Comments and exercises 126


<b>7 Coalgebras and Hopf algebras 130</b>


</div>
<span class='text_page_counter'>(8)</span><div class='page_container' data-page=8>

<i>Contents vii</i>


7.9 Tensor products of Hopf algebras 156


<i>7.10 E(M) as a (modified) Hopf algebra 158</i>
7.11 The Grassmann algebra of a module 160
<i>7.12 S(M) as a Hopf algebra 163</i>
7.13 Comments and exercises 166
7.14 Solutions to selected exercises 169


<b>8 Graded duality 175</b>


General remarks 175
8.1 Modules of linear forms 175
8.2 The graded dual of a graded module 178
8.3 Graded duals of algebras and coalgebras 184
8.4 Graded duals of Hopf algebras 188
8.5 Comments and exercises 191
8.6 Solutions to selected exercises 194


</div>
<span class='text_page_counter'>(9)</span><div class='page_container' data-page=9></div>
<span class='text_page_counter'>(10)</span><div class='page_container' data-page=10>

Preface



This account of Multilinear Algebra has developed out of lectures which I
gave at the University of Sheffield during the session 1981/2. In its present
form it is designed for advanced undergraduates and those about to
commence postgraduate studies. At this general level the only special
prerequisite for reading the whole book is a familiarity with the notion of a
module (over a commutative ring) and with such concepts as submodule,
factor module and homomorphism.


Multilinear Algebra arises out of Linear Algebra and like its antecedent is
a subject which has applications in a great many different fields. Indeed,
there are so many reasons why mathematicians may need some knowledge
of its concepts and results that any selection of applications is likely to


disappoint as many readers as it satisfies. Furthermore, such a selection
tends to upset the balance of the subject as well as adding substantially to
the required background knowledge. It is my impression that young
mathematicians often acquire their knowledge of Multilinear Algebra in a
rather haphazard and fragmentary fashion. Here I have attempted to weld
the most commonly used fragments together and to fill out the result so as
to obtain a theory with an easily recognizable structure.


The book begins with the study of multilinear mappings and the tensor,
exterior and symmetric powers of a module. Next, the tensor powers are
fitted together to produce the tensor algebra of a module, and a similar
procedure yields the exterior and symmetric algebras. Multilinear
mappings and the three algebras just mentioned form the most widely used
parts of the subject and, in this account, occupy the first six chapters.
However, at this point we are at the threshold of a richer theory, and it is
Chapter 7 that provides the climax of the book.


</div>
<span class='text_page_counter'>(11)</span><div class='page_container' data-page=11>

<i>x Preface</i>


<i>known as a coalgebra. Now it sometimes happens that, on the same</i>
underlying set, there exist simultaneously both an algebra-structure and a
coalgebra-structure. When this happens, and provided that the two
<i>structures interact suitably, the result is called a Hopf algebra. It turns out</i>
that exterior and symmetric algebras are better regarded as Hopf algebras.


This approach confers further benefits. By considering linear forms on a
coalgebra it is always possible to construct an associated algebra; and, since
exterior and symmetric algebras have a coalgebra-structure, this
construction may be applied to them. The result in the first case is the
algebra of differential forms (the Grassmann algebra) and in the second case


it is the algebra of differential operators.


<i>The final chapter deals with graded duality. From every graded module</i>
we can construct another graded module known as its graded dual. If the
components of the original graded module are free and of finite rank, then
this process, when applied twice, yields a double dual that is a copy of the
graded module with which we started. For similarly restricted graded
algebras, coalgebras and Hopf algebras this technique gives rise to a full
<i>duality; algebras become coalgebras and vice versa; and Hopf algebras</i>
continue to be Hopf algebras.


Each chapter has, towards its end, a section with the title 'Comments and
exercises'. The comments serve to amplify the main theory and to draw
attention to points that require special attention; the exercises give the
reader an opportunity to test his or her understanding of the text and a
chance to become acquainted with additional results. Some exercises are
marked with an asterisk. Usually these exercises are selected on the grounds
of being particularly interesting or more than averagely difficult; sometimes
they contain results that are used later. Where an asterisk is attached to an
exercise a solution is provided in the following section. However, to prevent
gaps occurring in the argument, a result contained in an exercise is not used
later unless a solution has been supplied.


Once the guide-lines for the book had been settled, I found that the
subject unfolded very much under its own momentum. Where I had to
<i>consult other sources, I found C. Chevalley's Fundamental Concepts of</i>
<i>Algebra, even though it was written more than a quarter of a century ago,</i>
especially helpful. In particular, the account given here of Pfaffians follows
closely that given by Chevalley.



Finally I wish to record my thanks to Mrs E. Benson and Mrs J.
Williams of the Department of Pure Mathematics at Sheffield University.
Between them they typed the whole book; and their cheerful co-operation
enabled the exacting task of preparing it for the printers to proceed
smoothly and without a hitch.


</div>
<span class='text_page_counter'>(12)</span><div class='page_container' data-page=12>

Multilinear mappings



<b>General remarks</b>


<i>Throughout Chapter 1 the letter R will denote a commutative ring</i>
<i>which possesses an identity element. R is called trivial if its zero element and</i>
<i>its identity element are the same. Of course if R is trivial, then all its modules</i>
are null modules.


The standard notation for tensor products is introduced in Section (1.2)
and from there on we allow ourselves the freedom (in certain contexts) to
omit the suffix which indicates the ring over which the products are formed.
More precisely when (in Chapter 1) we are dealing with tensor products of
<i>K-modules, we sometimes use ® rather than the more explicit ®R. This is</i>


done solely to avoid typographical complications.


<b>1.1 Multilinear mappings</b>


<i>Let Mt, M</i>2<i>, . . . , Mp (p> 1) and M be JR-modules and let</i>


<i></>: Mi x M2 x • • • x Mp-+M (1.1.1)</i>


be a mapping of the cartesian product Mt<i> x M2 x • • • x Mp</i> into M. We use



<i>mi9</i> m2<i>, . . . , mp to denote typical elements of Mu</i> M2<i>, . . . , Mp</i> respectively


<i>and r to denote a typical element of R. The mapping cf> is called multilinear if</i>
<i>cj)(mu...9m/i + m",... ,mp)</i>


<i>= (t>(mu</i>..., m j , . . . , mp) + </>(m1?<i>..., m " , . . . , mp) (1.1.2)</i>


and


<i>(j){mu..., rmh . . . , mp) = r(j){mu . . . , mi9..., mp). (1.1.3)</i>


<i>(Here, of course, i is unrestricted provided it lies between 1 and p.) For</i>
<i>example, when p = 1 a multilinear mapping is the same as a homomorphism</i>
of K-modules.


</div>
<span class='text_page_counter'>(13)</span><div class='page_container' data-page=13>

<i>2 Multilinear mappings</i>


<i>multilinear mappings from it in the following way. Let h: M-+N be a</i>
<i>homomorphism of K-modules. Then h ° (j) is a multilinear mapping of</i>
<i>Mx</i> x M2<i> x • • • x Mp</i> into JV. This raises the question as to whether it is


<i>possible to choose M and (j) so that every multilinear mapping of Mx</i> x


M2 x • • • x Mp can be obtained in this way. More precisely we pose


<i><b>Problem 1. To choose M and (j) in such a way that given any multilinear</b></i>


<i>mapping</i>



<i>\I/:M1XM2X--XMP->N</i>


<i>there is exactly one homomorphism h: M —• N (of R-modules) such that</i>


<i><b>h °</b></i><b> </> = i^.</b>


<i>This will be referred to as the universal problem for the multilinear</i>
mappings of Mx x M2<i> x • • • x Mp.</i>


We begin by observing that if the pair (M, 0) solves the universal
<i>problem, then whenever we have homomorphisms ht: M —• N (i = 1,2) such</i>


<i>that hx° (f) = h2o (j), then necessarily h1</i> = /z2. Now suppose that (M, </>) and


(M', (/>') both solve our universal problem. In this situation there will exist
<i>unique R-homomorphisms X.M-+M' and A': M' —• M such that A ° 0 = </>'</i>
<i>and X' ° (f)' = (j). It follows that (A' ° >1) ° 0 = 0 or (k' ° X) ° (j) = i ° 0, where i is</i>
the identity mapping of M. The observation at the beginning of this
<i>paragraph now shows that X' ° X = i and similarly X ° A' is the identity</i>
<i>mapping of M'. But this means that X\M-+M' and X'.M'^M are inverse</i>
isomorphisms. Thus if we have two solutions of the universal problem, then
they will be copies of each other in a very precise sense. More informally we
<i>may say that if Problem 1 has a solution, then the solution is essentially</i>
<i>unique. Before we consider whether a solution always exists, we make some</i>
general observations about free modules.


From here on, until we come to the statement of Theorem 1, we shall
<i>assume that R is non-trivial. We recall that an ^-module which possesses a</i>
<i>linearly independent system of generators is called free and a linearly</i>
<i>independent system of generators of a free module is called a base. Now let</i>


<i>X be a set and consider homogeneous linear polynomials (with coefficients</i>
<i>in R) in the elements of X. These form a free K-module having the elements</i>
<i>of X as a base. This is known as the free module generated by X. (If X is</i>
empty, then the free module which it generates is the null module.) Any
<i>mapping of X into an K-module N has exactly one extension to an </i>
<i>R-homomorphism of this free module into N.</i>


<i>We are now ready to solve Problem 1. Let U(Ml, M</i>2<i>, . . . , Mp) be the</i>


free R-module generated by the cartesian product Mx x M2<i> x • • • x Mp. Of</i>


course, this has the set of sequences (mls m2<i>, . . . , mp) as a base. The elements</i>


</div>
<span class='text_page_counter'>(14)</span><div class='page_container' data-page=14>

<i>Multlinear mappings</i>


<i>" , . . . , mp)-(mu ..., m'h..., mp)</i>


- ( m1, . . . , m ;/<i>, . . . ,mp) (1.1.4)</i>


and


<i>(ml9..., rmh . . . , mp)-r(mu . . . , mi9..., mp) (1.1.5)</i>


generate a submodule F(M 1? M2, . . . , Mp) say. Put


M = ( 7 ( M1, M2, . . . , Mp) / K ( M1, M2, . . . , Mp) (1.1.6)


and define


0 : Mx x M2<i> x • • • x Mp-*M (1.1.7)</i>



so that 0(mx, m2, . . . , mp) is the natural image of (ml5 m2, . . . , mp),


<i>con-sidered as an element of U(Ml9</i> M2, . . . , Mp), in M. Since the elements of


<i>U(Ml9M2,..., M</i>p) described in (1.1.4) and (1.1.5) become zero in M, 0


satisfies (1.1.2) and (1.1.3) and therefore it is multilinear. It will now be
<i>shown that M and </> provide a solution to our universal problem.</i>


To this end suppose that


<i><b>ij/: M</b><b>l</b></i><b> x M2</b><i><b> x • • • xM</b>p—>N</i>


is multilinear. There is an .R-homomorphism


<i>in which (ml9m2,..., mp) is mapped into \j/(ml,m2,..., mp). Since \jj is</i>


multilinear, the homomorphism maps the elements (1.1.4) and (1.1.5)
into zero and therefore it vanishes on P/(M1,M2,... ,MP). Accordingly


<i>there is induced a homomorphism h:M—+N which satisfies</i>
/i(0(m1,m2,... ,mp)) = ^ ( m1, m2, . . . ,mp). Thus /i°(/> = ^. Finally if


<i>/i': M —• N is also a homomorphism such that h' ° (/> = ^, then /i and /i' have</i>
<i>the same effect on every element of the form 4>(ml,m2,..., m</i>p). However,


<i>these elements generate M as an /^-module and therefore h = h'. Thus (M, (/>)</i>
solves the universal problem. We sum up our results so far.



<i><b>Theorem 1. Let M</b></i>l 9 M2<i>, . . . , Mp {p > 1) be R-modules. Then the universal</i>


<i>problem for multilinear mappings of Mx</i> x M2<i> x • • • x Mp has a solution.</i>


<i>Furthermore the solution is essentially unique (in the sense explained</i>
<i>previously).</i>


<i><b>Corollary. Suppose that (M, c/>) solves the universal problem described above.</b></i>


<i>Then each element ofM can be expressed as a finite sum of elements of the</i>
<i>form (f)(mum2,... ,mp).</i>


</div>
<span class='text_page_counter'>(15)</span><div class='page_container' data-page=15>

<i>4 Multilinear mappings</i>


<i>directly on the fact that (M, (f>) meets the requirements of the universal</i>
problem. This is the method employed here.


Suppose then that M' is the R-submodule of M generated by elements of
<i>the form <l>(ml9</i> m2<i>, . . . , mp). Also let hx: M-+M/M' be the natural </i>


<i>homo-morphism and h2:M^>M/M' the null homomorphism. Then h1o(j) =</i>


<i>h2 ° <j> and therefore ht = h2. However, this implies that M = M'.</i>


<i>Let xeM = M'. Then x can be expressed in the form</i>
<i>x = rcj)(ml9m2,..., mp) + r'<j)(m\, m</i>2<i>, . . . , m'p) + • • *,</i>


where the sum is finite. However,


<i>r(/)(mum2,..., mp) = (f)(rmu</i> m2<i>, . . . , mp)</i>



and similarly in the case of the other terms. The corollary follows.


<b>1.2 The tensor notation</b>


<i>Once again Ml9</i> M2<i>, . . . , Mp, where p> 1, denote R-modules and</i>


<i>we continue to use ml9</i> m2<i>, . . . , mp</i> to denote typical elements of these


<i>modules, and r to denote a typical element of R. Let the pair (M, (f>) provide</i>
<i>a solution to the universal problem for multilinear mappings of Mx</i> x


M2<i> x • • • x Mp. It is customary to write</i>


<i>M = M, ®RM2®R-- ®RMp</i> (1.2.1)


<i>and to use m1</i> ® m2<i> ® • • • ® mp</i> to designate the element (^(m^ m2, . . . , mp)


<i>of M. When this notation is employed, Mx đR M2 đR ã ã ã ®R Mp</i> is called


<i>the tensor product of Ml9</i> M2<i>, . . . , Mp</i> over i^.


It will be recalled that the solution to the universal problem is unique
only to the extent that any two solutions are copies of each other. Thus we
<i>can have different models for the tensor product. However, if we have two</i>
such models, then they are isomorphic (as modules) under an isomorphism
<i>which matches the element represented by mx đ m2 đ ã ã ã đ mp</i> in the first


model with the similarly represented element in the second. On account of
this we usually do not need to specify which particular model we are using.



Next, because </> is multilinear, the relations
<i>ã ã ã đ mp</i>


<i>- - - đmp + m1 đ ã ã ã đ mf( đ ã ã • ®mp</i>


(1.2.2)
and


<i>m1®" - ®rmt®" - ® mp = r(ml ® • • • ® mx ® • • • ® mp) (1.2.3)</i>


both hold. Moreover the corollary to Theorem 1 shows that each element of
Mx<i> ®KM2 ®R"' ®RMp</i> is a finite sum of elements of the form


</div>
<span class='text_page_counter'>(16)</span><div class='page_container' data-page=16>

<i>The tensor notation 5</i>


solution to the universal problem for multilinear mappings may be
restated as


<i><b>Theorem 2. Given an R-module N and a multilinear mapping</b></i>


<i>i//:MlxM2X'"xMp^N</i>


<i>there exists a unique R-module homomorphism h, of Mx ®RM2®R" '</i>


<i>đR Mp into N, such that</i>


<i>h{m^ đ m2 đ ã • • ® mp) = \j/(m1</i>, m2<i>, . . . , mp)</i>


<i>for all ml9</i> m2<i>, . . . , mp.</i>



<i>We interrupt the main argument to observe that when p = 1 the universal</i>
problem can be solved by taking M to be M x<i> and <\> to be the identity</i>


<i>mapping. This confirms that for p = 1 the tensor product Mx đR M2</i> đ<i>R</i> * * ã


<i>đRMp</i> is just Mx as we should naturally expect.


The reader will have noticed that the full tensor notation is rather heavy.
To counteract this we shall often use a simplified version. Indeed, because in
<i>Chapter 1 we shall only be concerned with a single ring R, it will not cause</i>
<i>confusion if we use Mx</i> ® M2<i> đ ã ã ã đ Mp</i> in place of the typographically


more cumbersome but more explicit Mx<i> ®RM2 ®R" ' ®RMp. </i>


Neverthe-less, in the statement of theorems and other results likely to be referred to
later we shall restore the subscript which identifies the relevant ring.


So much for matters of notation. Before we leave this section we shall
seek to gain insight into the nature of tensor products by examining the
<i>result of forming the tensor product of a number of free modules.</i>


<i>Suppose then that, for 1 < i </?, M</i>f<i> is a free .R-module and that Bt</i> is a base


for Mf<i>. The first point to note is that any mapping ofBl x B2 x • • • x Bp</i> into


<i>an R-module N has precisely one extension to a multilinear mapping of</i>
<i>Mx</i> x M2<i> x • • • x Mp into N. We use this observation in the proof of our</i>


next theorem.



<i><b>Theorem 3. Let M, (i = 1, 2 , . . . , p) be a free R-module with a base B</b>t. Then</i>


Mt đ<i>RM2 đR- ã' đRMp is also a free R-module and it has the elements</i>


<i>bt ® b2 ® - • • ® bp, where bt e Bh as a base.</i>


<i>Proof Denote by M the free R-module generated by the set Bx x B2</i> x • • •


<i>x Bp. Thus the sequences (bl9 b2,..., bp) form a base for M. There is then a</i>


<i>mapping 0, of B1 x B2 x • • • x Bp into M, in which (f>(bu b2,..., bp) is the</i>


<i>base element (bl9 b2,..., bp). Now, as we noted above, the mapping has an</i>


extension (denoted by the same letter) to a multilinear mapping of
Mx x M2<i> x • • x Mp into M. Clearly all we need to do to complete the proof</i>


</div>
<span class='text_page_counter'>(17)</span><div class='page_container' data-page=17>

<i>6 Multilinear mappings</i>


Suppose then that we have a multilinear mapping
<i>\//: M1 xM2</i> x • • • xMp—•N.


<i>In these circumstances there exists an R-homomorphism h: M—•AT which</i>
<i>maps (bl</i> , fc2<i>,..., bp) into \j/{b1</i> , b2<i>, . . . , bp). Obviously ft ° </> = i/f. Moreover,</i>


<i>if h'\ M^>N is also a homomorphism satisfying /i'°0 = ^, then h and h!</i>
<i>agree on the base Bx x B2 x • • • x Bp</i> of M and therefore /i = /i'. The proof is


therefore complete.



<b>1.3 Tensor powers of a module</b>


<i>Let M be an ^-module and p > 1 an integer. Put</i>


<i>Tp(M) = M ®RM®R--®RM, (1.3.1)</i>


<i>where there are p factors. The K-module Tp(M) is called the p-th tensor</i>


<i>power of M. These powers will later form the components of a graded</i>
<i>algebra known as the tensor algebra of M. For the moment we note that</i>
<i>7; (M) = M. Also, if M is a free K-module and B is a base of M, then Tp(M) is</i>


<i>also free and it has the elements b1 ®b2®- - - ®bp, where bt<b> e B, as a base.</b></i>


This follows from Theorem 3.


<b>1.4 Alternating multilinear mappings</b>


As in the last section M denotes an K-module. In this section we
<i>shall have occasion to consider products such as M x M x • • • x M and</i>
<i>M ® M ®- - - ® M. Whenever such a product occurs it is to be understood</i>
<i>that the number of factors is p, where p > 1.</i>


A multilinear mapping
<i>rj: M xM x ••• xM-+N</i>


<i>is called alternating if f/(m</i>1,m2,... ,mp) = 0 whenever the sequence


(m1? m2, . . . , mp<i>) contains a repetition. (To clarify the position when p= 1</i>



<i>we make it explicit that all linear mappings M-^N are regarded as</i>
alternating.)


<i>Suppose that n has this property. If l<i<j<p, then on expanding</i>
<i>n(mu..., mi+nip..., mf + rn^-,..., m</i>p) = 0,


<i>where the element m,- + m}</i><b> occurs in both the i-th andj-th positions, we find</b>


that


<i>n{mu..., mi9..., mp ..., m</i>p) + >y(m1<i>,..., mp . . . , mi9</i>..., mp) = 0.


<i>Thus if we interchange two terms in the sequence {ml</i>, m2, . . . , mp) the effect


<i>on n(m1, m</i>2, . . . , mp) is to multiply it by — 1. From this observation we at


</div>
<span class='text_page_counter'>(18)</span><div class='page_container' data-page=18>

<i>Alternating multilinear mappings 7</i>


<i><b>Lemma 1. Let rj.MxMx'xM—^N be an alternating multilinear</b></i>


<i>mapping and let (i</i>l 5 i2<i>, . . . , ip) be a permutation of ( 1 , 2 , . . . , p). Then</i>


<i>rj(mii9 mi2,..., mip) = ±ri(m1,m2,..., mp\</i>


<i>where the sign is plus if the permutation is even and minus if it is odd.</i>
Another useful observation is recorded in


<i><b>Lemma 2. Let ri'.MxMx — -xM—>N be a multilinear mapping and</b></i>



<i>suppose that ri(ml9 m2,..., mp) = 0 whenever mi = mi + 1for some i. Then rj is</i>


<i>an alternating mapping.</i>


<i>Proof The argument just before the statement of Lemma 1 shows that</i>
<i>rj(ml9m2, • . . , mp) changes sign if we interchange two adjacent terms in the</i>


sequence (m1? m2<i>, . . . , mp). Now suppose that (ml,m2,..., mp) contains a</i>


repetition. Then either two equal terms occur next to each other or this
situation can be brought about by a number of adjacent interchanges. It
<i>follows that rj(ml,m2,..., mp) = 0 so the lemma is proved.</i>


<i>Consider an alternating multilinear mapping rj, of M x M x • • • x M into</i>
<i>an K-module N, and suppose that h: N-+K is a homomorphism of </i>
<i>R-modules. Then h ° rj is an alternating multilinear mapping of M x M x • • •</i>


<i>x M into K. This observation leads us to pose the following universal</i>
problem.


<i><b>Problem 2. To choose N and Y\ in such a way that given any alternating</b></i>


<i>multilinear mapping</i>


<i>there exists exactly one homomorphism h: N-^K (of R-modules) such that</i>
<i>C = horj.</i>


<i>To solve this problem we consider the tensor power Tp(M) and denote by</i>


<i>JP(M) the submodule generated by all elements m</i>x<i> đ m2 đ ã ã ã đ mp,</i>



where ( m1, m2, . . . , mp) contains a repetition. (It is understood that


J1<i>(M) = 0.) Put N = Tp(M)/Jp(M) and let rj be the mapping of M x M x • • •</i>


<i>x M into N which takes {mx,m2,..., mp) into the natural image of m</i>x<i> ®</i>


<i>m2 ® • • • ® mp in N. Then rj is an alternating multilinear mapping. If now</i>


<i>£:MxMx-xM->K</i>


is also an alternating multilinear mapping, then, by Theorem 2, there is a
<i>homomorphism of M đ M đ ã ã ã đ M into K which takes m</i>x<i> đ m2</i> đ ã ã •


<i>® mp</i> into C(wi1? m2<i>, . . . , mp). This homomorphism vanishes on Jp(M) and</i>


<i>so it induces a homomorphism /z: AT —• X. It is clear that h ° >/ = £. If we have</i>
<i>a second homomorphism, say h': N-^K, and this satisfies h' °rj = £, then /i</i>
and ft' agree on the elements ^/(rnj, m2<i>, . . . , mp) and therefore they agree on a</i>


</div>
<span class='text_page_counter'>(19)</span><div class='page_container' data-page=19>

<i>8 Multilinear mappings</i>


It has now been shown that the universal problem for alternating
<i>multilinear mappings of M x M x • • • x M has a solution. The solution is</i>
<i>unique in the following sense. Suppose that (AT, rj) and (AT, rjf) both solve</i>
<i>Problem 2. Then there are inverse isomorphisms X: N —> N' and k': N'-+N</i>
<i>such that A°rj = rj' and X' °Y\' = r\. The situation is, in fact, almost identical</i>
with that encountered in dealing with uniqueness in the case of Problem 1.


<i>Let us suppose that (N, rj) solves the universal problem for alternating</i>


<i>multilinear mappings of M x M x • • • x M. Put</i>


<i>EP(M) = N (1.4.1)</i>


and


<i>mx /\m2 A • • • Amp = rj(m1,m2,... ,m</i>p). (1.4.2)


Then, because fy is multilinear, we have
<i>mx</i> A • • • A (mj + m") A • • • A<i> mp</i>


= mx A • • • A<i> m[</i> A • • • A<i> mp + m1</i> A •


and


<i>ml</i> A • • • Arm; A<i> • • • Amp = r(m1</i> A • • •


But */ is also alternating. Consequently mx AWI2 A
<i>(ml,m2,..., mp) contains a repetition and, by Lemma 1,</i>


<i>miy Ami2</i> A • • • A ^ = ±mx Am2 A • • • Amp, (1.4.5)


<i>where the plus sign is to be used if (il9 i2,..., ip) is an even permutation of</i>


(1,2,...,/?) and the minus sign if the permutation is odd.


<i>The module Ep(M) is called the p-th exterior power of M. As we have seen</i>


<i>it is unique in much the same sense that Tp(M) is unique. Since when p = 1</i>



we can solve Problem 2 by means of M and its identity mapping, we have
<b>Am;'</b>


<b>A • * •</b>


<b>A • • •</b>
<b>A • • *</b>


<b>A mp</b>


<b>A mp</b>


<b>A m</b>


<b>).</b>


= 0


<i><b>„ (1.4.3)</b></i>


(1.4.4)
whenever


The defining property of the exterior power is restated in the next
theorem.


<i><b>Theorem 4. Given an R-module K and an alternating multilinear mapping</b></i>


<i>C:MxMx-xM-+K</i>



<i>there exists a unique R-homomorphism h: Ep(M)-+K such that</i>


<i>for all ml,m2,... ,mpin M.</i>


<i><b>Corollary. Each element of E</b>P(M) is a finite sum of elements of the form</i>


<b>wij A</b><i><b> m</b><b>2 A</b></i><b> • • • A</b><i><b> m</b><b>p</b><b>.</b></i>


</div>
<span class='text_page_counter'>(20)</span><div class='page_container' data-page=20>

<i>Alternating multilinear mappings 9</i>


<i><b>The mapping of MxMx- xM into E</b><b>p</b><b>(M) which takes</b></i>


<i><b>(m</b><b>l</b><b>,m</b><b>2</b><b>,.. •, m</b><b>p</b><b>) into m</b><b>l</b></i><b> A m2</b><i><b> A • • • A m</b><b>p</b></i><b> is multilinear and so, by Theorem</b>


<b>2, there is induced a homomorphism TP(M)—>£P(M). This is surjective</b>
<i><b>because the image of m</b><b>x</b><b> ® m</b><b>2</b></i><b> ® ' * * ® wp</b><i><b> is m</b><b>x</b></i><b> A m2 A • • • A mp. We refer to</b>
<i>Tp(M)—>Ep<b>(M) as the canonical homomorphism of the tensor power onto</b></i>


<i><b>the exterior power. Note that when p = 1 the canonical homomorphism is</b></i>
<i><b>the identity mapping of M if we make the identifications T</b><b>1</b><b>(M) = M =</b></i>


<i>EX{M).</i>


<i><b>As before we let J</b><b>P</b><b>(M) denote the #-submodule of T</b><b>p</b><b>(M) generated by all</b></i>


<i><b>products m</b><b>x</b><b> ® m</b><b>2</b><b> đ ã * * đ m</b><b>p</b></i><b> in which there is a repeated factor.</b>


<i><b>Theorem 5. The canonical homomorphism T</b>p(M) —• Ep(M) is surjective and</i>


<i>its kernel is Jp(M). Moreover Jp(M) is generated, as an R-module, by all</i>



<i>products mx</i> ® m2<i> ® • * • ® mp, where mi = mi + 1 for some i.</i>


<i><b>Proof. Let J'</b><b>P</b><b>(M) be the submodule of T</b><b>p</b><b>(M) generated by all m</b><b>x</b></i><b> ®</b>


<i><b>m</b><b>2</b><b> ® * * * ® m</b><b>p</b><b> with m</b><b>t</b><b> = m</b><b>i + 1</b><b> for some i, and let J"</b><b>P</b><b>{M) be the kernel of the</b></i>


<i><b>canonical homomorphism. Then J'</b><b>P</b><b>(M) ^J</b><b>P</b><b>(M)^ J"</b><b>P</b><b>(M). Next, by Lemma</b></i>


<b>2, the mapping</b>


<i>6: M x M x • • • xM-+Tp{M)/J'p(M),</i>


<i><b>in which 9{m</b><b>x</b><b>, m</b></i><b>2</b><i><b>, . . . , m</b><b>p</b><b>) = m</b><b>1</b></i><b> ® m2</b><i><b> ® • • • ® m</b><b>p</b><b> + J'</b><b>p</b><b>(M), is multilinear</b></i>


<b>and alternating and therefore, by Theorem 4, there is a homomorphism</b>


<i>X: Ep(M)-+ Tp(M)/J'p<b>(M) such that</b></i>


<i>X(ml A m2 A • • • A mp) = m</i>x<i> đ m2 đ ã ã ã ® mp 4- J'p(M).</i>


<b>But if A is combined with the canonical homomorphism Tp(M)—>£p(M),</b>
<b>then the result is the natural homomorphism</b>


<i>Tp(M)^Tp(M)/J'p(M).</i>


<i><b>Thus the natural homomorphism vanishes on J"</b><b>P</b><b>{M) and therefore</b></i>


<b>j ; ( M ) c jp</b><i><b>( M ) . Accordingly J'</b><b>p</b><b>(M) = J</b><b>p</b><b>(M) = J"</b><b>p</b><b>(M) and the theorem is</b></i>



<b>proved.</b>


<b>It is useful to know the structure of the exterior powers of a free module.</b>
<i><b>Let us assume therefore that M is a free jR-module and that B is one of its</b></i>
<i><b>bases. On the set formed by all sequences (b</b><b>l9</b><b> b</b><b>2</b><b>,...</b><b>9</b><b> b</b><b>p</b><b>) of p distinct</b></i>


<i><b>elements of B we introduce the equivalence relation in which (b</b><b>t</b></i><b>, ft2</b><i><b>,..., b</b><b>p</b><b>)</b></i>


<i><b>and (b'</b><b>l9</b><b> b</b><b>29</b><b>..., b'</b><b>p</b><b>) are regarded as equivalent if each is a permutation of the</b></i>


<b>other. From each equivalence class we now select a single representative. In</b>
<b>the next theorem the set formed by these representatives is denoted by</b>


<i><b>I</b><b>P</b><b>(B).</b></i>


<i><b>Theorem 6. Let M be a free R-module having the set B as a base. Then</b></i>


<i>Ep(M) is a free R-module having the elements bx A b2 A • • • A bp as a base,</i>


</div>
<span class='text_page_counter'>(21)</span><div class='page_container' data-page=21>

<i>10 Multilinear mappings</i>


<i>Proof. Let N be the free K-module generated by Ip(B). Define a mapping</i>


<i>n.BxBx ••• xB-^N</i>


<i>as follows. W h e n (bl9 b2,..., bp) c o n t a i n s a r e p e t i t i o n n{b1, b2,..., bp) is t o</i>


<i>b e z e r o . W h e n , h o w e v e r , bi,b2,-..9bp</i> a r e all different t h e r e is a u n i q u e


<i>p e r m u t a t i o n (b[, b'2,..., b'p\ of (bi9 b2,..., bp), s u c h t h a t (b\, b'2,..., b'p)</i>



<i>belongs to Ip(B); in this case we put</i>


where the plus sign is taken if the permutation is even and the minus sign if it
<i>is odd. The mapping n has a unique extension (denoted by the same letter)</i>
to a multilinear mapping of M x M x • • • x M into AT, and the construction
ensures that the extension is alternating as well as multilinear.


<i>We claim that (N,rj) solves Problem 2. Clearly once this has been</i>
established the theorem will follow. Suppose then that


<i>£:MxMx'xM-+K</i>


<i>is an alternating multilinear mapping. Since N is free, there is a </i>
<i>homo-m o r p h i s homo-m h : N — > K s u c h t h a t h ( b1, b2, . . . , bp) = C ( bl 9b2, . . . , bp)</i>


<i>whenever (bl9b2,...9bp) is in Ip(B). Thus h°n and ( are alternating</i>


<i>multilinear mappings which agree on Ip(B) and therefore on B x B x • • •</i>


<i>x B as well. If follows that h°rj = C. Moreover it is evident that h is the only</i>
<i>homomorphism of N into K with this property. The proof is therefore</i>
complete.


<b>1.5 Symmetric multilinear mappings</b>


<i>Once again M denotes an ^-module and all products M xM x • • •</i>
<i>x M and M ® M ® ' " ® M are understood to have p (p > 1) factors.</i>


<i>Let N be an /^-module. A multilinear mapping</i>


<i>9:MxMx-"xM-^N</i>


<i>is called symmetric if</i>


<i>0(ml9m2, • • •, mp) = 0(m</i>lV mlV<i> . . . , mip\ (1.5.1)</i>


<i>whenever mx,m2,...,mp belong to M and (il ,i2,..., ip) is a permutation of</i>


<i>( 1 , 2 , . . . , p). Clearly 6 is symmetric provided 6(ml9m2,..., mp) remains</i>


<i>unaltered whenever two adjacent terms are interchanged.</i>


<i>If 6 is a symmetric multilinear mapping and h: N—+K is a </i>
<i>homo-morphism of JR-modules, then h ° 6 is a symmetric multilinear mapping of</i>
<i>M x M x • • • x M into K. This inevitably prompts the next universal</i>
problem.


<i><b>Problem 3. To choose N and 6 so that given any symmetric multilinear</b></i>


</div>
<span class='text_page_counter'>(22)</span><div class='page_container' data-page=22>

<i>Symmetric multlinear mappings 11</i>


<i>co: M x M x • - x M —+K</i>


<i>there exists a unique R-homomorphism h: N-^K such that h°6 = oj.</i>
As Problem 3 can be treated in very much the same way as Problem 2
there will be no need to go into the same amount of detail.


<i>Denote by Hp(M) the submodule of Tp(M) generated by all differences</i>


w1đ m2<i>đ *** đmp miiđmi2đ ã ã ã ®m,-, (1.5.2)</i>



<i>where (il9i29..., ip) is a permutation of ( 1 , 2 , . . . , p). Put</i>


<i>N=Tp(M)/Hp(M)</i>


<i>and defined: MxMx- xM—•Nsothat 0(m</i>1,m2,... ,mp) is the natural


<i>image of mx ® m2 ® • • • ® mp in N. Then 6 is multilinear and symmetric.</i>


Moreover considerations, closely similar to those encountered in solving
<i>Problem 2, show that N and 6 solve the present problem. Of course the</i>
solution is unique in the same sense and for much the same reasons as apply
in the case of the other universal problems.


<i>Now let (JV, 6) be any solution to Problem 3. We put</i>


<i>SP(M) = N (1.5.3)</i>


and


<i>m1m2 . . . mp = 0(ml9</i> m2<i>, . . . , mp). (1.5.4)</i>


Then


<i>m±m2 . . . (wij + w i " ) . . . mpz=m^m2 . . . m[... mp-\-mim2 . . . m'l... mp</i>


(1.5.5)
and


<i>m1m2 . . . (rmt)... mp = r{mlm2 ...mi...mp) (1.5.6)</i>



<i>because 6 is multilinear, and</i>


<i>mlm2 . . . mp = miimi2... mip</i> (1.5.7)


<i>because it is symmetric. (Once again (il9</i> i2<i>, . . . , ip) denotes an arbitrary</i>


<i>permutation of ( 1 , 2 , . . . ,p).) The module Sp(M) is called the p-th symmetric</i>


<i>power of M. As in earlier instances, when p = 1 our universal problem is</i>
<i>solved by M and its identity mapping. Consequently S1(M) = M. The next</i>


theorem records the characteristic property of the general symmetric
power.


<i><b>Theorem 7. Given an R-module K and a symmetric multilinear mapping</b></i>


<i>there exists a unique R-homomorphism h: Sp(M)^>K such that</i>


<i>h(mlm2 ... mp) = w(ml,m2,..., mp)</i>


</div>
<span class='text_page_counter'>(23)</span><div class='page_container' data-page=23>

<i>12 Multilinear mappings</i>


<i><b>Corollary. Each element of S</b>P(M) is a finite sum of elements of the form</i>


<i>m1m2 . . . mp.</i>


As is to be expected a proof of this corollary can be obtained by making
minor modifications to the arguments used when establishing the corollary
to Theorem 1.



<i>We next observe that, by Theorem 2, there is a homomorphism Tp(M) —•</i>


<i>SP(M)9 of ^-modules, in which mx ® m2 ® ' ' * ® mp</i> is mapped into


<i>mlm2 . . . mp. This is known as the canonical homomorphism of Tp(M) into</i>


<i>Sp(M). Evidently it is surjective. If we identify T^M) and S^M) with M,</i>


<i>then, for p = 1, the canonical homomorphism is the identity mapping.</i>
<i>We recall that Hp(M) is the submodule of TP(M) generated by elements of</i>


the form (1.5.2).


<i><b>Theorem 8. The canonical homomorphism T</b>p(M) —• SP(M) is surjective and</i>


<i>its kernel is Hp(M). Furthermore Hp(M) is generated by all differences</i>


<i>mx đ ãã • ®mi®mi + 1 ® • •• ®mp</i>


<i>-ml®--' ® m</i>x + 1 đ m < đ ããã ® mp,


<i>where the second term is obtained from the first by interchanging two</i>
<i>adjacent factors.</i>


Here the demonstration parallels the proof of Theorem 5. We leave the
reader to make the necessary adjustments.


<i>Finally let us determine the structure of SP(M) when M is a free K-module</i>



<i>with a base B. To do this we consider all sequences (bl9 b29... 9 bp) of p</i>


<i>elements of B (on this occasion repetitions are allowed) and we regard the</i>
<i>sequences (bl9 bl9. . . 9 bp) and (bfl9 b'l9..., b'p) as equivalent if each can be</i>


obtained by reordering the other. From each equivalence class we now
<i>select a representative, and we denote by I*(B) the set consisting of these</i>
representatives.


<i><b>Theorem 9. Let Mbea free R-module with base B. Then S</b>p(M) is also a free</i>


<i>R-module and it has as a base the elements b1b2 ... bp9 where (bl9 b2,..., bp)</i>


<i>ranges over I*(B).</i>


<i>Remark. If the elements of B are regarded as commuting indeterminates,</i>
<i>then this result says that Sp(M) can be thought of as consisting of all</i>


<i>homogeneous polynomials of degree p, with coefficients in R9</i> in these


indeterminates.


<i>Proof. Let N be the free .R-module generated by I*(B) and define</i>


<i>d:BxBx ••• xB—>N</i>


<i>by putting 6(bl9 b2,..., bp) = {b'l9 b'2,..., b'p\ where (b'l9 b'l9..., b'p) is the</i>


</div>
<span class='text_page_counter'>(24)</span><div class='page_container' data-page=24>

<i>Comments and exercises 13</i>



<i>to a multilinear mapping of M x M x • • x M into N and it is evident that</i>
the multilinear mapping so obtained is symmetric. It now suffices to show
<i>that (N, 0) solves Problem 3.</i>


Suppose then that


<i>co'.MxMx- - - xM —>K</i>


<i>is multilinear and symmetric, and define the jR-homomorphism h: N-+K</i>
<i>so that, for (b1,b2,... ,bp) in I*(B), we have h(bl,b2,... ,bp) =</i>


<i>co(bl9 b2,..., bp). Then h ° 6 and co are symmetric multilinear mappings</i>


<i>which agree on I*(B). It follows that they must agree on B x B x • • x B,</i>
<i>and this in turn implies that h°6 = co. This completes the proof because it is</i>
clear that there is only one homomorphism M —• K which, when combined
<i>with 6, gives co.</i>


<b>1.6 Comments and exercises</b>


In this section we shall make some observations to complement
what has been said in the main text and we shall provide a number of
exercises. A few of these exercises are marked with an asterisk. Those that
are so marked are either particularly interesting, or difficult, or used in later
chapters. Solutions to the starred questions will be found in Section (1.7).


<i>Throughout Section (1.6) R is assumed to be non-trivial.</i>


<i>Since we are concerned with multilinear algebra it is fitting that the scope</i>
<i>of linear algebra (in the present context) should be made explicit. Often by</i>


linear algebra is meant the theory of vector spaces and linear
<i>trans-formations. When the term vector space is used it is frequently understood</i>
<i>that the scalars are taken from a field. We, however, will only require our</i>
<i>scalars to belong to an arbitrary commutative ring and when we make this</i>
change the terminology normally changes as well. What had previously
<i>been termed a vector space now becomes a module and what had been</i>
<i>described before as a linear transformation is now referred to as a</i>
<i>homomorphism. Thus, for us, linear algebra will mean the theory of modules</i>
(and their homomorphisms) over a commutative ring.


One of the key problems of linear algebra is to determine when a system
of homogeneous linear equations has a non-trivial solution. Let us pose this
problem in a very general form.


<i>Suppose that M is an ^-module and consider the equations</i>
<i>a11ml +al2rn2+ • • -+alqmq = 0</i>


<i>a21ml +a22m2 + • • • +a2 m =0</i>


: : : : : (1 6 1)


</div>
<span class='text_page_counter'>(25)</span><div class='page_container' data-page=25>

<i>14 Multilinear mappings</i>


<i>Here the au belong to R and we are interested in solving the equations in the</i>


<i>module M. Of course one solution is obtained by taking all the mt</i> to be zero.


<i>Any other solution is called non-trivial.</i>


The first exercise gives a necessary and sufficient condition for the


<i>existence of a non-trivial solution. This result is known as McCoy's</i>
<i>Theorem. In order to state it we first put</i>


<i><b>A</b></i> <i><b>=</b></i><b> 021 «22</b>


(1.6.2)


<i>so that A is the matrix of coefficients, and we denote by(Uv(A) the ideal that</i>


<i>is generated by the v x v minors of A. Then</i>
8 l o M ) 2 « i M ) 2 «2( i 4 ) 2 - - - ,


<i>where by M0(A)</i>we<i> mean the improper ideal R. Note that % (A) = 0 for v ></i>


<i>min(p, q) because, for such a v, there are n o v x v minors.</i>


<i><b>Exercise 1*. Show that the equations (1.6.1) have a non-trivial solution, in the</b></i>


<i>R-module M, if and only if $lq(A) annihilates a non-zero element of M.</i>


<i>Deduce that if M ^ O and p<q, then a non-trivial solution exists.</i>


<i>Exercise 1 can be used inter alia to establish some fundamental facts</i>
about free modules and their bases. We illustrate this by means of the next
exercises.


<i><b>Exercise 2. The R-module M can be generated by q elements. Show that any</b></i>


<i>q + l elements ofM are linearly dependent over R. Show also that if M is a</i>
<i>free module with a base of q elements, then (i) q is the smallest number of</i>


<i>elements which will generate M, and (ii) any q elements that generate M form</i>
<i>a base.</i>


<i><b>Exercise 3. Let Mbea free R-module and let B, B' be bases of M. Show that</b></i>


<i>either (i) B and B' are both infinite, or (ii) B and B' are both finite and contain</i>
<i>the same number of elements.</i>


Exercise 3 shows that the number of elements in a base of a free R-module
M is the same for all choices of the base. This number, which is called the
<i>rank of the free module, will be denoted by rank^(M). Thus mnkR(M) is</i>


either a non-negative integer or it is 'plus infinity'.
We now turn our attention to some other matters.


<i><b>Exercise 4. Let I</b><b>l9</b><b>I</b><b>2</b><b>,...,I</b><b>p</b> be ideals of R such that /</i>x + /2<i> + • • • +Ip = R,</i>


</div>
<span class='text_page_counter'>(26)</span><div class='page_container' data-page=26>

<i>x--Solutions to selected exercises 15</i>


<i>is a multilinear mapping of R-modules, then every element ofMl</i> x M2 x • • •


<i>x Mp is mapped into zero.</i>


<i>We recall that if M is an R-module and / is an ideal of R, then M/IM has a</i>
natural structure as an R/7-module as well as an R-module.


<i><b>Exercise 5*. Let M</b>X,M2,... ,Msbe R-modules and let I be an ideal ofR.</i>


<i>Show that the R/I-modules</i>



<i>(M1 ®RM2®R-- ®R MS)/I(M1 ®RM2®R-- ®R Ms)</i>


<i>and</i>


<i>(MJIMJ ®R/I (M2/IM2) ®R/I • • • ®R/I (MJIMS)</i>


<i>are isomorphic and exhibit an explicit isomorphism.</i>


In Chapter 5 we shall derive the basic facts about free modules, their
bases and their ranks by different means. The central fact in the alternative
approach is embodied in the next exercise.


<i><b>Exercise 6. Let M^Obea free R-module. Show that rank</b></i>R<i>(M) is the upper</i>


<i>bound of all integers p such that Ep(M)^0.</i>


<i><b>Exercise 7. Let M be a free R-module of rank n. Determine the ranks of the</b></i>


<i>free modules Tp(M), Ep(M) and Sp(M).</i>


The next exercise shows that it is possible for the tensor square of a
non-zero module to be non-zero.


<i><b>Exercise 8. Show that if Z denotes the ring of integers and Q the field of</b></i>


<i>rational numbers, then (Q/Z) ®</i>Z((Q/Z) = O.


<b>1.7 Solutions to selected exercises</b>


In this section we shall provide complete solutions to Exercises 1


and 5, and make observations about some of the other exercises.


<i><b>Exercise 1. Show that the equations (1.6.1) have a non-trivial solution, in the</b></i>


<i>R-module M, if and only if S&q (M) annihilates a non-zero element of M.</i>


<i>Deduce that if M ^ 0 and p<q, then a non-trivial solution exists.</i>


<i>Solution. First suppose that ml9</i> m2<i>, . . . , mq</i> satisfy the equations with at


<i>least o n e m, different from zero. We claim that <$lq(A)mi = 0fori=l,2,...,q.</i>


<i>Evidently in establishing this we may suppose that p>q.</i>
<i>From the first q equations, namely</i>


<i>flii»ii +al2rn2 + - - - +alqmq = 0</i>


<i>a21mx +a22m2 + • • • -\-a2qmq = 0</i>


</div>
<span class='text_page_counter'>(27)</span><div class='page_container' data-page=27>

<i>16 Multilinear mappings</i>


<i>we see that Dmt = Q (i= 1, 2 , . . . , q), where D is the q x q minor</i>


0 2 1


<i>In the same way we can show that any other q x q minor of A annihilates all</i>
<i>the mt. Accordingly ^Lq(A)mi = 0 and the necessity of the stated condition</i>


has been established.



Now suppose that 91^(^4) annihilates a non-zero element of M. If v is the
<i>smallest integer for which ^(A) annihilates a non-zero element, then</i>
<i>0 < v < q and 3^ (A)m = 0 for some m ^ 0 in M. If v = 1 the existence of a </i>
non-trivial solution is obvious. Otherwise there is a (v— 1) x (v— 1) minor that
does not annihilate m. Without loss of generality we can suppose that the
<i>(v— 1) x (v— 1) minor in question occurs in the top left-hand corner of A.</i>


In the determinant


<b>0 1</b> 01v


0 v - l , 2


<i>let cr be the cofactor of xr. Then c</i>v<i>m^0. Furthermore, for 1 <i<p,</i>
<i><b>-+a</b><b>iv</b><b>(c</b><b>v</b><b>m)</b></i>


0f2 ' ' * 0/v


<i><b>=</b><b> m</b></i><b> -</b><i><b> Ql2</b><b> '"</b><b> a u</b></i>
0 v - l , 2 ' ' ' 0v-l,v


<i>If 1 < i < v — 1 this is zero because the determinant has two equal rows; and</i>
<i>if i>v, then it is zero because Sllv(A)m = 0. Thus by taking</i>


wi1=c1<i>wi, m2 = c</i>


and


we obtain a non-trivial solution of the equations.



<i>Finally suppose that M ^ O and p<q. Then (Hq(A) = 0 and this certainly</i>


annihilates a non-zero element of M. Accordingly a non-trivial solution
exists.


Exercise 1 provides the key to Exercise 2.


<i><b>Exercise 5. Let M</b>1,M2,... 9Msbe R-modules and let I be an ideal of R.</i>


<i>Show that the R/I-modules</i>


</div>
<span class='text_page_counter'>(28)</span><div class='page_container' data-page=28>

<i>Solutions to selected exercises 17</i>


<i>and</i>


<i>đR/I (M2/IM2) đR/I ã ã ã đR/I (MJIMS)</i>


<i>are isomorphic and exhibit an explicit isomorphism.</i>


<i>Solution. For m, e Mj let m, denote its natural image in Mj/IMj. Now every</i>
R/7-module has an obvious structure as an R-module. On this
under-standing the mapping


<i>M,xM2X' -xM^iMJIM,) đR/I- ã đR/I(MS/IMS)</i>


<i>which takes (mum2,..., m</i>s) into mx<i> đ m2 đ * * ã đ ms</i> is a multilinear


mapping of R-modules. Accordingly there is induced an R-linear mapping
<i>M, đR ã ã ã đR Ms-+ (MJIM1) đR/Iã • • ®R/I (MJIMS)</i>



<i>and this, as it clearly vanishes on I(M1 đR- ã ã đRMS), gives rise to a</i>


mapping


<i>A: (Af i đ^- ã ã đRMs)/I(Ml đR-" đRMS)</i>


<i> (MJIM,) đR/I - • • ®R/I (MS/IMS).</i>


<i>In fact I is a homomorphism of ^//-modules and if mx đ ã ã ã đ ms</i> denotes


the natural image of mx<i>đ ã ã ã đms in {M1 ®R'" ®RMS)/</i>


<i>1(MX®R-" ®RMS\ then</i>


<i>Mmx ®- — ®ms) = ml®m2®" ' ®ms.</i>


<i>Now suppose that, for each ;, we have elements m, and m}, in Mp</i> such


<i>that rhj^m'j. Then</i>


<i>= ml ® • • • ® (mj — m'j) ® • • • ® ms</i>


<i>and this belongs to I(Ml®R" - ®RMS). Repeated applications of this</i>


observation show that
<i>mlđm2đ" ã đms</i>


<i>= m\đ m'2 đ - - đ mfs (mod I(MX đR ã ã ã ®R Ms)).</i>


It follows that there is a well-defined mapping


( M1/ / M1) x - - - x ( Ms/ / Ms)


-> (M! đn ã ã ã đ<i> R MM)/I(MX đR'"đR Ms)</i>


<i>in which (ml9</i> m2<i>, . . . , ms) goes into ml®m2®- - ®ms. This is a </i>


<i>multi-linear mapping of ^//-modules and so it gives rise to an </i>
R/I-homomorphism


<i>®R/I -' ã đR/I (MJIMS)</i>


</div>
<span class='text_page_counter'>(29)</span><div class='page_container' data-page=29>

<i>18 Multilinear mappings</i>


which satisfies /i(mx<i> đ ã ã ã đ ms) = (m</i>x đ m2 đ ã ã ã đ ms). Finally


<b>( j ) ^ !</b> <i><b>s</b></i> <i><b>) j (</b></i> <i><b>l s</b></i><b>) = m1 đ ã ã ã đ</b><i><b> m</b><b>s</b></i>


and


(A0<i>/^)^! đ ã ã • ® ms) = k(m1</i> ® • • • ®ms<i>) = m! đ ã ã ã đ ms.</i>


<i>It follows that k and ^ are inverse isomorphisms and now the solution is</i>
complete.


Our concluding observation has to do with the answers to the questions
<i>posed by Exercise 7. We record that, for a free /^-module M of rank n,</i>


</div>
<span class='text_page_counter'>(30)</span><div class='page_container' data-page=30>

<b>2</b>



Some properties of tensor products




<b>General remarks</b>


It was in Section (1.2) that we defined the tensor product of a finite
<i>number of modules over a ring R, but at that time we passed on rapidly to</i>
study the tensor, exterior and symmetric powers of a module. However, the
theory of tensor products is rich in results, some of which will be needed
later. Here we provide an account of the most fundamental properties. As in
<i>Chapter 1 we shall sometimes omit the subscript to the symbol ® if it is</i>
clear which is the ring over which the products are formed.


<i>Throughout Chapter 2 the letters R and S denote commutative rings each</i>
possessing an identity element.


<b>2.1 Basic isomorphisms</b>


<i>Let Mu</i> Af 2<i>, . . . , Mp and Nl9 N2,..., Nq</i> be K-modules. In what


<i>follows ml9m29 - - •, mp denote elements of Ml9 M2, • • •, Mp</i> respectively


<i>and nl9n2,... ,nq denote elements of Nl9N2,... 9Nq.</i>


<i><b>Theorem 1. There is an R-isomorphism</b></i>


<i>Mi đR'' - đRMp <g)RN1 đR-" đRNq</i>


<i>ô ( M ! ®R''' ®R Mp) ®R (N1 ®R'"®RNq)</i>


<i>in which mx ® • • • ® mp ® nx ® • • • ® nq is matched with (m1 ® • • • ® mp) ®</i>



<i>{nl®'"®nq).</i>


<i>Proof. By Theorem 2 of Chapter 1, the multilinear mapping</i>
Mx<i> x • • • x Mp x Nl x • • • x Nq</i>


<i>where (mi9..., mp, nl9..., nq) goes to (m</i>x<i> đ ã ã ã đ mp) ®(nl®--® nq),</i>


induces a homomorphism


</div>
<span class='text_page_counter'>(31)</span><div class='page_container' data-page=31>

<i>20 Some properties of tensor products</i>
<i>f:Mlđ--đMpđN1đ-'đNq</i>


<i>-->(Ml</i>


đ-in which


= (mx<i> đ ã ã ã đ mp) đ (ô! đ ã ã ã đ nq).</i>


We now seek to reverse this homomorphism. Suppose for the moment
<i>that we keep nl9n29...,nq fixed. Another application of Chapter 1,</i>


Theorem 2 yields a homomorphism


Mx đ M2 đ ã ã ã đ Mpã Mt<i> đ ã ã ã đ Mp đ Nx đ • • • ® Nq</i>


<i>in which m</i>

<i>l</i>

<i>®m</i>

<i>2</i>

<i>®' — ®m</i>

<i>p</i>

<i> is mapped into m</i>

<i>l</i>

<i>®--®m</i>

<i>p</i>

<i>®n</i>

<i>l</i>

<i></i>



<i>®--® nq. Consequently if in M</i>x ® M2<i> ® • • • ® Mp</i> we have a relation
<i>m1 ® m2 ® " ' ® mp + m\ ® mf2 đ " ã đ mfp</i>



then


<i>ml đ - - - ® mp ® n1 ® • • • ® nq</i> + mi


<i>is zero in Mx đ ã ã ã đ Mp đ Nx đ ã ã ã đ Nq. Of course a similar </i>


<i>observa-tion applies if the roles of the Mt and Nj are interchanged.</i>


<i>Let £ eM</i>

<i>x</i>

<i> ® M</i>

<i>2</i>

<i> ® • • • ® M</i>

<i>p</i>

<i> and rjeN</i>

<i>l</i>

<i>®N</i>

<i>2</i>

<i>®--®N</i>

<i>q</i>

<i>. We know</i>



<i>that £, can be represented as a finite sum of elements of the form</i>


<i>m</i>

<i>1</i>

<i>®m</i>

<i>2</i>

<i>®'"®m</i>

<i>p</i>

<i>. Suppose that</i>



f = Z mx ® m2 ® • • • ® mp = ^ ^ ® //2<i> ® • • • ® \xp</i>


are two such representations of ^ and that


<i>= Y<</i> vi ® V<i>2 ® ' * * ® vq</i>


<i>are two similar representations of rj. The remarks in the last paragraph</i>
show that


Consequently


<i>Y£jm1®'"®mp®nl®'"®nq</i> (2.1.1)


</div>
<span class='text_page_counter'>(32)</span><div class='page_container' data-page=32>

<i>Basic isomorphisms 21</i>


<i>that takes (£, n) into the element (2.1.1). This mapping, because it is bilinear,</i>


induces a homomorphism


<i>- • Ml ® • • • ® Mp</i>


<i>in which (m! ® • • • g m ^ g ) ^ ! ® • • • ®nq) becomes mx</i> ® • • •


Wi®"'"®wfl<i>. The theorem follows because g r ° / and f°g are clearly</i>


identity mappings.


<i><b>Corollary. There is an isomorphism</b></i>


<i>(Ml®RM2)®RM3*M1®R(M2®RM3)</i>


<i>of R-modules in which (mx ® m2) ® m3 corresponds to mx</i> ® (m2 ® w3).


Proo/ The theorem provides isomorphisms


(M! ® M2) ® M3 « MX ® M2 ® M3 ôAf! đ (M2 đ M3)


and the corollary follows by combining them.


<i><b>Theorem 2. Let (i</b>l9 i2,..., ip) be a permutation of ( 1 , 2 , . . . , p). Then there</i>


<i>is an isomorphism</i>


<i>Mx ®RM2®R" ®RMp*Mix ®RMh ®R-" ®RMip</i>


<i>of R-modules which associates m1®m2®'-®mp with mix®mi2®'"</i>



<i>®mip.</i>


<i>Proof The multilinear mapping</i>


<i>M1xM2X" xMp-^Mti ®Mh ®" - ®Mip</i>


<i>taking (ml9 m2,..., mp) into mix đ mi2 đ ã ã ã đ mt</i> induces a


homo-morphism


/ : Mj đ M2<i> đ ã ã ã ® Mp-+Mti ®Mh<b>®'"® M</b><b>ip</b><b>,</b></i>


where /(mx<i> ® m2 ® • • • ® mp) = mix ® mi2 ® • • • ® mi</i> . For similar reasons


there is a homomorphism


<i>with g(mti đ mi2 đ ã ã ã đ mt ) = m1®m2®- - ® mp. Evidently g ° f and</i>


<i>f°g are identity mappings and therefore / is an isomorphism. The theorem</i>
is therefore proved.


<i>The ring R is itself an K-module. As the next theorem shows, this</i>
particular /^-module plays a very special role in the theory of tensor
products.


<i><b>Theorem 3. Let M be an R-module. Then there is an isomorphism</b></i>


<i>R®RM&M (of R-modules) in which r®mis matched with rm. There is a</i>


</div>
<span class='text_page_counter'>(33)</span><div class='page_container' data-page=33>

<i>22 Some properties of tensor products</i>



<i>Proof Theorem 2 of Chapter 1 shows that there is a homomorphism</i>
<i>/ : R®RM-^M with f(r ® m) = rm. The mapping g: M —> R ® R M given</i>


<i>by g(m)=l®m is certainly a homomorphism. But r ® m = l ® r m and</i>
<i>from this it follows that g ° / and / ° g are identity mappings. Consequently</i>
/ is an isomorphism and the theorem is proved.


<b>2.2 Tensor products of homomorphisms</b>


<i>Let Ml9</i> M2<i>, . . . , Mp</i> and M i , M2<i>, . . . , M'p</i> be ^-modules, and let


<i><b>fi'.M^Ml ( i = l , 2 p) (2.2.1)</b></i>


be given homomorphisms. The mapping


<i>Mi x M</i>2 x • • • xMn—>Mi ® M'2 ® • • • ® M'


<i>in which the image of (ml9</i> m2<i>, . . . , mp) is fiintx) đ /</i>2(m2<i>) đ ã ã • ® fp(mp) is</i>


multilinear and therefore it gives rise to a homomorphism
MXđ M2<i> đ ã ã ã đMp-+M'1đM'2 đ ã • • ®M'p</i>


<i>of /^-modules. This homomorphism, which is denoted by fx đ f2</i> đ ã ã ã


<i>đ fp, satisfies</i>


<i>(/l đ fl ® ' ' ' ® fp) (</i>ml ®<i> m2 ® ' ' ' ® mp)</i>


=



/ i (m<i>i ) ® fi(mi) ® *' * ®/p(^</i>P). (2.2.2)


Note that if each /j is surjective, then / i ® /2<i> ® • • • ® fp</i> is surjective as well.


Note too that if MJ = M£<i> for i = 1 , 2 , . . . , p and /j is the identity mapping of</i>


<i>Mi9 then f1®f2®'"®fp is the identity mapping of M1®M2</i>


<i>®--®MP.</i>


Once again suppose that homomorphisms (2.2.1) are given and that, in
addition to these, we have further homomorphisms


<i>gt: M'l - * Mi (i = 1 , 2 , . . . , p). (2.2.3)</i>


It then follows, from (2.2.2), that


<i>(/i ® fi ® *' * ® fp) ° (#i ® 02 ® * " ® 9P)</i>


<i>= (fi°9i) ® (fi°9i) ®''' ®(fp°9P)' (2.2.4)</i>


<i>We can deduce from this that when each f is an isomorphism, then</i>
<i>fi ® fi ®''' ® fp is</i> als° a n isomorphism. For in this situation we can form
<i>both fi® f2®' " ® fp and f{~</i>x<i> đ f2 1 đ ã ã ã đ fp~1, and when these are</i>


combined, in either order, (2.2.4) shows that we obtain an identity mapping.
<i>Consequently not only is fx ® f2 ® • • • ® fp</i> an isomorphism, but we also


have



</div>
<span class='text_page_counter'>(34)</span><div class='page_container' data-page=34>

<i>Tensor products of homomorphisms 23</i>


<i>summarized succinctly by the statement that the p-Md tensor product is a</i>
<i>covariant functor of p variables.)</i>


There is an additional item of notation which it is convenient to record
<i>here. Suppose that fi,...<)fi_l9fi + l,...,fp</i> are the identity mappings of


M1 ?. . . ,M,-_1,Ml- + 1<i>, . . . ,Mp respectively, and that f'.M^M'i is a</i>


<i>general homomorphism. Under these conditions fl ® f2 đ''' đ fp</i> is often


written as Mx<i> đ ã • • ® f ® • • • ® Mp. Thus</i>


(Mx<i> đ ã ã ã đ f{</i> đ ã ã • ® Mp) (mx<i> ® • • • ® mi ® • • • ® mp)</i>


<i>= mx ® • • • ®fi(mi) ® • • • ® m</i>p. (2.2.6)


<i>We recall that if/, # are homomorphisms of an ^-module M into an </i>
<i>R-module M', and if r e R, then we can form new homomorphisms / + g and rf</i>
<i>of M into M'. Indeed to be explicit ( / + g)(m) = f(m) + g(m) and (rf)(m) =</i>
<i>rf(m). In this connection we note that</i>


<i>P p</i> (2.2.7)


and


<i>/ i ® *'' ® rft đ ' ' ã đ /</i>p<i> = r(/i đ *' ã đ fi đ ' *' ® fP), (2.2.8)</i>



where the notation is self-explanatory.


After these preliminaries we turn our attention to the behaviour of tensor
products in relation to exact sequences of the form


<i>M'l -^-> Mt —• M| —> 0 (2.2.9)</i>


<i>where i = 1 , 2 , . . . , p. Denote by N(</i> the image of Mx<i> đ ã ã ã đ gt đ ã • • ® Mp</i>


in Mx<i> ® M2 ® * * • ® Mp. Thus N{</i> is generated by the elements of the form


<i>mt</i> đ ã ã ã đ0,-(mj')đ ã ã • ®mp and, in particular, it is contained in the


kernel of /x đ/2 đ ã ã ã đ/p.


<i><b>Theorem 4. Suppose that for each i (1 <i<p) the sequence (2.2.9) is exact.</b></i>


<i>Then the homomorphism</i>


<i>->M'1đRM'2đR--đRMp</i>


<i>is surjective and its kernel is N1-\-N2-\- ã • • +iV</i>p<i>, where Nt is the image of</i>


<i>Mi ® • • • ® gt ® - - • ® Mp in Mt đR M2 đR ã ã ã đ* Mp.</i>


<i>Proof. For 1 < i <p let mj e M\ and suppose that mt and /Xj- belong to M,- and</i>


<i>are such that fi(mi) = fi(ni) = m'i. Then there exists m'l eM'l such that</i>


<i>m, — fii = gi(m'l) and so we see that</i>



mx<i> đ ã ã ã đ mt đ • • • ® mp — mx ® • •</i>


</div>
<span class='text_page_counter'>(35)</span><div class='page_container' data-page=35>

<i>24 Some properties of tensor products</i>


<i>m2 ® * * * ®mp — fi1 ® ^2®"</i>


<i>is in N1 + N2</i> + * • • + Np<i> and therefore the natural image of m1 đ m2 đ ã * *</i>


<i>đmp</i><b> in</b>


<i>(Mi đ M2 đ ã ã ã đ MP)/(NX</i> + AT2<i> + • • • + Np)</i>


depends only on (mi, m'2<i>,..., m'p). It is now easily checked that there is a</i>


multilinear mapping


<i>M\ x M</i>2 x • • • x M'p


which is such that (mi, m2, . . . , m^) is mapped into the image in question. In


this way we arrive at a homomorphism


<i>M\ ® M'2 ® • - • ® M'p</i>


<i>~^(Ml®M2®" -®MP)/(N1</i> +JV2 + • • • + N , ) (2.2.10)


<i>that takes m\®m'2®- - ® m'p into the natural image of ml ® m2 ® • • •</i>


<i>® mp</i> in the factor module.



Consider the homomorphisms


Mx<i> đ - ã ã đ Mp-ã Mi đ ã • • ® M'p</i>


<i>where the first is fx®f2®' - ®fp</i> and the second is (2.2.10). Their


<i>combined effect is to reproduce the natural mapping of M1 đ M2 đ ã ã •</i>


<i><b>® M</b><b>p</b><b> onto (M</b>l®M2®-"® Mp)l{Nl +N2 + • • • + Np<b>) so the kernel of</b></i>


<i>f\® fi®" ' ® fP must be contained in Nx</i> + N2<i> + • • • + Np. But we also</i>


<i>have the opposite inclusion because, as was noted earlier, each N(</i> is


contained in the kernel. The proof is now complete.


<i><b>Corollary. Let M"</b>t<b> —• M</b>fl —• M'fl<b> —• 0 be an exact sequence of R-modules.</b></i>


<i>Then the induced sequence</i>


<i>Mi đR' ' ' đRM'1 đR- ã ã ®RMp</i>


- * Mx ®K<i> • • • ®R Mfl ®R</i> • • • ®K Mp


<i><b>M; ®</b><b>R</b><b>-"®</b><b>R</b><b> M</b><b>P</b><b>->O</b></i>


<i>is also exact.</i>


<i>Proof. The corollary is derived from the theorem by taking all the fi9</i> with



<i>the exception of ffl, to be identity mappings. This ensures that N( = 0 for all</i>


</div>
<span class='text_page_counter'>(36)</span><div class='page_container' data-page=36>

<i>Tensor products and direct sums 25</i>


<b>2.3 Tensor products and direct sums</b>


<i>Let {Nt}ieI be a family of submodules of an ^-module N. Then N is</i>


<i>the direct sum of these submodules if each n e N has a unique representation</i>
of the form


<i><b>n = Y</b></i>

<i><b>J</b></i>

<i><b>n</b></i>

<i><b>h</b></i>

<b> (2.3.1)</b>



<b>16/</b>


<i>where n( e Nt</i> and only finitely many summands are non-zero. When this is


the case we shall usually write


AT = 5 > , (As.). (2.3.2)
<b>16/</b>


<i>However, when N is the direct sum of a finite number of submodules, say of</i>
<i>Nl9 N2,..., Ns, we shall use</i>


<i>N = Nl®N2@---®Ns</i> (2.3.3)


as an alternative.



<i>Suppose that we have the situation envisaged in (2.3.2). Then for each i e /</i>
<i>we have an inclusion mapping o^A^—»iV and a projection mapping</i>
<i>nt: N —> Nt. (For n e N the latter picks out from the representation (2.3.1) the</i>


<i>summand indexed by i.) Both a{ and n{</i> are #-homomorphisms. They have


the properties listed below:


<i>(A) 7ij ° (Tt is a null homomorphism if i ^j and it is the identity mapping of</i>


<i>Ntifi=j;</i>


<i>(B) for each neN, 7c,(n) is non-zero for only finitely many different</i>
<i>values of i;</i>


<i>(C) for each neN, Y.</i>


<i>Let us now make a completely fresh start. Suppose that N is an </i>
<i>R-module, and that {Nt}ieI</i> is a family of ^-modules which, however, are no


<i>longer assumed to be submodules ofN. Suppose too that, for each iel, we</i>
<i>are given homomorphisms ậ Nt—>N9</i> 7rf<i>: N—>Nt</i> and that these satisfy


conditions (A), (B) and (C).


<i>Since nt °Gt is the identity mapping of Ni9 o{ is an injection and n{</i> is a


<i>surjection. In particular o{ maps Nt</i> isomorphically onto <rf(Nf). Next from


(B) and (C) it follows that



<b>16/</b>


<i>and now, by using (A), we can deduce that N is the direct sum of its</i>
submodules {0",(Afi)}/e/ in the sense used at the beginning of this section.


</div>
<span class='text_page_counter'>(37)</span><div class='page_container' data-page=37>

<i>26 Some properties of tensor products</i>


<i><b>complete representation of N as a direct sum. Furthermore we shall continue</b></i>
to write


<b>ie/</b>


<i>in the general case, and N = N1®N2(B"'@NS</i> in the case where the


<i>family {Nt}iel reduces to {Nu</i> AT2,...,<i> N</i>


<i>s}-Now suppose that M1, M</i>2<i>, . . . , Mp</i> are R-modules and that each is given


a (generalized) direct sum decomposition, say


M . = L<i> Mi» (d.s.),</i>


where the complete representation is provided by homomorphisms


and


<i><b>Theorem 5. Suppose that for each \i (1 ^fi^p) we have</b></i>


<i><b>M, = I M^ (d.s.),</b></i>



<i>w/zere the details of the complete representation are as described above. Then</i>
<i>Ml®RM2 ®R-' ®RMp</i>


<i><b>= Z ^ . ^ M ^ . - ^ M ^ (d.s.),</b></i>



<b>( » i , . . . , * p )</b>


<i>vv/i^r^ the typical homomorphisms in the complete representation are oix</i> ®


<i>oh ®''' ® oip and nti ® ni2</i> ® • • • ® 7ilp.


Proo/. Put / = /! x /2<i> x • • • xlp, N = MX</i> ® M2<i> ® • • • ®Mp and, for i =</i>
<i>(iui2i...Jp) in / , set</i>


<b>a n</b>


<i><b>d 7ii = ^i</b><b>l</b><b> ®</b><b>n</b><b>i</b><b>2</b><b> ®' " ®</b><b>n</b><b>i</b><b>p</b><b>- Then a^Ni—^N,</b></i>


<i>Ui'.N—^Ni and to prove the theorem it will suffice to show that the</i>
conditions which were listed earlier as (A), (B), (C) all hold. That (A) holds is
an immediate consequence of (2.2.4). Indeed (B) and (C) are equally obvious
<i>as soon as it is realized that we need only verify them when n has the form</i>
<i>mx ® m2 ®''' ® mp.</i>


<i><b>Corollary. Let F be a free R-module with a base {^,}</b></i>/6/<i> and let N be an</i>


<i>arbitrary R-module. Then each element of F ®RN has a unique </i>


<i>representa-tion in the form £ (£,. ® n{\ where the nt belong to N and only finitely many of</i>



</div>
<span class='text_page_counter'>(38)</span><div class='page_container' data-page=38>

<i>Tensor products and direct sums 27</i>


<i>Proof. We have</i>


<i>iel</i>


<i>and N is also a direct sum with N itself as the only summand. Accordingly,</i>
by the theorem,


and so, to establish the corollary, we need only show that for each x in
<i>R£i ® N there is a unique rc e N such that x = £</i>f ® n.


<i>By Theorem 3, there is an isomorphism N&R®N in which n</i>
<i>corresponds to 1 ®n, and of course there is an isomorphism R®N&</i>
<i>R£t ® N which matches 1 ® n with ^ (g) rc. Together these provide an</i>
<i>isomorphism N^R^t</i> ® JV with n corresponding to £f ® rc. Consequently,


<i>as n ranges over N, £t ® n ranges over R£( (x) N giving each element once</i>


and once only.


We recall that a module which is a direct summand of a free module is
<i>called a projective module.</i>


<i><b>Theorem 6. Let 0 —• M" —• M —> M' —• 0 fee ^n exact sequence of R-modules</b></i>


<i>and let P be a projective R-module. Then the induced sequences</i>


<i>0^>P®RM"^P(g)RM-+P®RM'->0</i>



<i>and</i>


<i>0 - > M " ®RP-+M ®RP-+Mf ®RP-+0</i>


<i>are exact.</i>


<i>Proof Of course we need only consider the first sequence. Let / : M" —• M</i>
be an injective homomorphism. By Theorem 4 Cor., it will suffice to show
<i>that P ® / is also injective. We begin with the case where P is free.</i>


<i>Suppose that F is a free /^-module and that {£i}ieI</i> is one of its bases. Let


<i>x G F ® M". By Theorem 5 Cor., x = £ (f, ® m'/), where the mj' are in M"</i>
<i>and only finitely many of them are non-zero. Assume that (F ® / ) ( x ) = 0.</i>
Then


Zôiđ/ô)) = 0


and therefore /(mj') = 0 for all f G /, again by the corollary to Theorem 5. But
/ is an injection. Consequently all the m" are zero and hence x = 0. This
<i>proves that F ® / is an injection.</i>


</div>
<span class='text_page_counter'>(39)</span><div class='page_container' data-page=39>

<i>28 Some properties of tensor products</i>


<i>injection. But we already know that F ® / is an injection so</i>
<i>(F®f)°((j®M") = (T(g)f is an injection as well. Since cr®/=</i>
<i>(a ® M) ° (P ® / ) , this implies that P ® / is an injection, which is what we</i>
were aiming to prove.



<b>2.4 Additional structure</b>


<i>In Section (2.4) S, as well as R, denotes a commutative ring with an</i>
identity element. If M is both an R-module and an S-module, then we shall
<i>say it is an (R, S)-module provided (i) the sum of two elements of M is the</i>
<i>same in both structures, and (ii) s(rm) = r(sm) whenever reR, seS and</i>
<i>msM. Let M and M' be (R, 5)-modules. A mapping / : M —• M' is called an</i>
<i>(R, Syhomomorphism if it is a homomorphism of K-modules and also a</i>
<i>homomorphism of S-modules. A bijective (R, S)-homomorphism is, of</i>
<i>course, referred to as an (R, S)-isomorphism.</i>


Suppose that M x, . . . , M;i _<i> x</i>, M;| +<i> x,..., Mp</i> are JR-modules and that M;1


<i><b>is an (R, S)-module. For s e S define f</b>s</i>: M/4 —> M;< by /s(m/4) = sm/r Then /s is


<i>an R-homomorphism and therefore Ml®'-'®fs®"'(g)Mp9 where fs</i>


occurs in the /i-th position, is an endomorphism of the K-module Mx đ ã ã •


® M^ ® • • • (g) Mp. For x in Mx đ ã ã ã đ M/4<i> đ ã • • ® Mp</i> define sx by


5X = (MX ® • • • ® /s ® • • • ® Mp)(x). (2.4.1)


<i>Then, because fs + a = fs+fa and fsa = fs°f(T for s,aeS, it is easily verified</i>


that Mx ® • • • ® M/4<i> ® • • • ® Mp</i> is an S-module. Indeed we can go further


<i>and say that it is actually an (R, 5)-module with</i>


<i>s(m1 ® • • • ® m,, ® • • • ® mp) = m</i>x<i> đ ã ã ã đ sm^ đ • • • ® mp. (2.4.2)</i>



<i>Finally, if gt: M</i>f<i>—>Mj is an R-homomorphism for i= 1 , . . . , fi— 1, fi +1,</i>


. . . , p and ^;1: M/f —• MJ4 is an (K, S)-homomorphism, then ^1® *


-® 0/< -® ' * * -® ^p is itself an (jR, S)-homomorphism.


<i>After these preliminaries let A be an K-module, C an S-module, and B an</i>
<i>(R, 5)-module. The considerations set out in the earlier paragraphs show</i>
<i>that A ®RB is an (R, S)-module and therefore (A ®RB) ®</i>S<i>C is an (R, </i>


<i>S)-module. Similar considerations show that A ®R(B®SC) is an (R, S)-module</i>


<i>as well. The next theorem, which asserts that (A<g)RB)®sC and</i>


A®/?(B®S<i>C) are virtually identical bimodules, is known as the associative</i>


<i>law for tensor products.</i>


<i><b>Theorem 7. Let A be an R-module, B an (R, S)-module, and C an S-module.</b></i>


<i>Then there is an (R, S)-isomorphism</i>


<i>(A ®RB)®SC&A ®R(B ®</i>SC)


</div>
<span class='text_page_counter'>(40)</span><div class='page_container' data-page=40>

<i>Covariant extension 29</i>


<i>Proof. We borrow an idea from the proof of Theorem 1. Select an element c</i>
<i>from C and for the moment let it be kept fixed. Consider the mapping</i>



<i>where (a, b) is mapped into a®(b®c). If we regard A,B and</i>
<i>A ®R (B ®s C) simply as K-modules, then the mapping is bilinear and</i>


therefore it induces a homomorphism
<i>A ®R B ã A đR (B đs C)</i>


<i>in which a đ b goes into a ® (b ® c). It follows that if</i>
<i>a ®b + a' ®b' + a" ® b" -\ = 0</i>


<i>in A ®R B, then</i>


<i>a® (b®c) + af ®(b' ®c) + a" ®(b" ®c) + '-=0</i>


<i>inA®R(B®sC).</i>


<i>Let x belong to A ®R B. Then x can be represented in the form</i>


<i>x = ax ®bx +a2 ®b2 + — - +an ®bn</i>


and the remarks of the last paragraph show that


flx<i> ® (h ®c) + a2 ®{b2®c)+--+an® (bn®c) (2.4.3)</i>


<i>depends only on x and c and is independent of the choice of the</i>
<i>representation of x. Consequently there is a well-defined mapping of</i>
<i>(A ®RB)xC into A ®R (B ®s C) in which (x, c) is mapped into the element</i>


<i>(2.4.3). But if we regard A ®RB, C and A ®R (B ®s C) as 5-modules, then</i>


this is a bilinear mapping and therefore it induces an 5-homomorphism


<i>X: {A ®RB) ®SC-^A ®R (B ®SC)</i>


with


<i>k({a ®b)®c) = a®(b®c). (2.4.4)</i>


<i>It is now clear, from (2.4.4), that X is actually an {R, S)-homomorphism.</i>
<i>Entirely similar considerations show that there is an (R, </i>
S)-homomorphism


<i>fi: A đR (B đs C)-ã {A đR B) ®s C</i>


<i>for which n(a ® (b ®c)) = (a ®b) ®c. Since /i°X and X° \i are identity</i>
<i>mappings, k and \i are inverse isomorphisms. The proof is therefore</i>
complete.


<b>2.5 Covariant extension</b>


<i>Once again R and S denote commutative rings, only now we</i>
assume that a ring-homomorphism


<i>oj.R^S (2.5.1)</i>


</div>
<span class='text_page_counter'>(41)</span><div class='page_container' data-page=41>

<i>S-30 Some properties of tensor products</i>


<i>module U can be regarded as an K-module by defining rw, where reR and</i>
<i>ue U, to be the same as co(r)u. In fact this turns U into an (R, S)-module. It</i>
follows, as was shown in the last section, that if M is an K-module, then
<i>M ®R U is an (R, S)-module. Note that it is also true of M ®R U that</i>



<i>multiplication of an element by r is the same as multiplication by co(r).</i>
<i>A special case arises if we take U to be S itself. Thus if M is an /^-module,</i>
<i>then M ®R S is an (R, S)-module. Actually it is the S-module structure that</i>


<i>concerns us more because R operates through S in the simple manner just</i>
<i>explained. When M ®R S is considered as an S-module we call it the</i>


<i>covariant extension of M by means of the ring-homomorphism (2.5.1).</i>
Covariant extensions are important because they cover such basic
<i>constructions as localization and the formation of fractions and</i>
<i>polynomials. By combining these constructions under a general heading we</i>
can derive some of their most important properties while still keeping the
details comparatively simple.


<i><b>Theorem 8. Let Mbea free R-module with {£,-}</b> ieI as a base. If now the ring S</i>


<i>is non-trivial, then the covariant extension M ®R S is a free S-module and it</i>


<i>has {£i ® l }</i>/ 6 /<i> as a base.</i>


<i>Proof. Let x e M ®R S. By Theorem 5 Cor., there is a unique family {st} iel9</i>


<i>of elements of S, with only finitely many st non-zero and x = £ (^ ® s,). The</i>


<i>theorem follows because £t ®^i = si{£>i</i> ® 1).


<i><b>Corollary. The covariant extension of a protective R-module is a protective</b></i>


<i>S-module.</i>



<i>Proof Suppose that F is a free ^-module and that P, Q are #-submodules</i>
<i>with F = P ®Q. By Theorem 5,</i>


<i>F ®RS = (P ®RS) 0 (Q ®R</i>5), (2.5.2)


<i>this being, in the first instance, a direct sum of/^-modules. However, F ®R S,</i>


<i>P ®RS and Q ®R S are all of them S-modules, and now (2.5.2) is seen to be a</i>


<i>direct sum of S-modules. But it has just been shown that F ®R S is S-free. It</i>


<i>therefore follows that P ®RS and Q ®RS are 5-projective.</i>


Our final result in this chapter shows that tensor products and covariant
extensions commute.


<i><b>Theorem 9. Let M</b>1,M29.-.,Mp be R-modules. Then there is an </i>


<i>isomor-phism</i>


<i><b>(M, ®</b></i>

<i><b>R</b></i>

<i><b> S) ®</b></i>

<i><b>s</b></i>

<b> (M</b>

<b>2</b>

<i><b> đ</b></i>

<i><b>R</b></i>

<i><b> S) đ</b></i>

<i><b>s</b></i>

<i><b> - - ã đ</b></i>

<i><b>s</b></i>

<i><b> (M</b></i>

<i><b>p</b></i>

<i><b> đ</b></i>

<i><b>R</b></i>

<i><b> S)</b></i>



<i><b>*(M! ®</b></i>

<i><b>R</b></i>

<i><b>M</b></i>

<i><b>2</b></i>

<i><b> ®</b></i>

<i><b>R</b></i>

<i><b>-- ®</b></i>

<i><b>R</b></i>

<i><b>M</b></i>

<i><b>p</b></i>

<i><b>) ®</b></i>

<i><b>R</b></i>

<i><b>S</b></i>



<i>of S-modules in which (m</i>x<i> ® sx</i>) ® (m2<i> ®s2)®"'® (mp ® sp) corresponds</i>


</div>
<span class='text_page_counter'>(42)</span><div class='page_container' data-page=42>

<i>Comments and exercises 31</i>


<i>Proof. We begin with the case p = 2. By Theorem 7, there is an (R, </i>
S)-isomorphism



(Mx<i> ®R S) ®s</i> (Af 2<i> ®R S)«M</i>x ®* (5 ®s (M2 ®* 5)) (2.5.3)


which matches (mx ® sx) ® (m2 ® s2) with mx ® (sx ® (m2 ® s2)). Next,


Theorem 3 provides an S-isomorphism


<i>S ®s</i>(M2<i> ®RS)*M2®RS (2.5.4)</i>


<i>which pairs sx ® (m</i>2 ® s2) with m2<i> ® s ^ . Indeed this is also an (R, </i>


<i>isomorphism. Thus (2.5.3) and (2.5.4) together provide an (R, </i>
S)-isomorphism


<i>(M, ®RS) ®S(M2 ®RS)*M1 ®R (M2 ®RS)</i>


<i>in which (mx ® s</i>x) ® (m2 ® 52<i>) goes into m1 ® (m</i>2<i> ®sls2). Finally, the</i>


corollary to Theorem 1 provides an K-isomorphism


Mx ®^ (Af 2<i> ®R S)ô(Af! đ* M</i>2<i>) đR</i> 5, (2.5.5)


<i>where mlđ(m2đs) is associated with (m1®m2)®s. The case p = 2</i>


<i>follows as soon as it is noted that (2.5.5) is an S-isomorphism as well as an </i>
R-isomorphism.


<i>When p = 1 the theorem becomes a tautology. The general result follows</i>
<i>by induction if we make use of Theorem 1 and the case where p = 2.</i>



<b>2.6 Comments and exercises</b>


This section will be used to clarify certain aspects of the theory of
tensor products by means of comments and exercises. As in the last chapter
solutions are provided to the more interesting and difficult exercises, and
the exercises which come into this category are marked by an asterisk.


<i>Let Af' be a submodule of an K-module M and N' a submodule of an </i>
<i>R-module N. The purpose of the first exercise is to show that usually Af' ® Nf</i>
<i>cannot be regarded as a submodule of Af ® N. More precisely, the inclusion</i>
<i>mappings M-+M and N'^N induce a homomorphism</i>


<i>M' ®N'—>M ®N, (2.6.1)</i>


and now the point being made is covered by


<i><b>Exercise 1. Show by means of an example that the homomorphism</b></i>
<i>Af' đ N''ã Af ® N of (2.6.1) need not be an injection.</i>


<i>(Hint. Take R = Z, Af = Af' = Z/2Z, N = Z, and AT = 2Z. As always Z</i>
denotes the ring of integers.)


By contrast Exercise 2 embodies an important constructive result which
partially offsets the purely negative assertion of Exercise 1.


</div>
<span class='text_page_counter'>(43)</span><div class='page_container' data-page=43>

<i>32 Some properties of tensor products</i>


<i>yi» ^2? • • • •> y$ belong to an R-module N. Suppose thatxl ® yx + x2 ® y2</i> + ' ' '


<i>+ xs® ys = Oin M ® N. Show that there exist finitely generated submodules</i>



<i>M' and AT, of M and N respectively, such that</i>


<i>(i) the x( are in M' and the yt in AT,</i>


<i>and</i>


<i>(ii) x1®y1+x2®y2 + --+xs®ys = 0 in M'</i>


<i><b>Exercise 3*. Let I be an ideal of R and let M be an R-module. Establish an</b></i>


<i>isomorphism M ®R<b> (R/I)&M/IM of R-modules.</b></i>


The next exercise provides a version of a very useful result known as
<i>Nakayamas Lemma. It appears here because it helps to solve Exercise 5.</i>


<i><b>Exercise 4*. Let M be a finitely generated R-module. Show that if M = JM</b></i>


<i>for every maximal ideal J, then M = 0.</i>


<i>The condition that M be finitely generated cannot be dropped. This may</i>
<i>be seen by taking R to be Z and letting M be the field of rational numbers</i>
considered as a Z-module.


<i><b>Exercise 5*. Let M be an R-module. Show that the following two statements</b></i>


<i>are equivalent:</i>


<i>(i) for a finitely generated R-module N we have M ® N = 0 only when</i>
<i>N = 0;</i>



<i>(ii) for every maximal ideal J, of R, JM is different from M.</i>
We next consider a matter that concerns tensor products of
<i>homo-morphisms. To this end let / : M —• N and / ' : M' —• N' be homomorphisms</i>
of K-modules. Then, as we saw in Section (2.2), these induce a
homomorphism


<i>f®f: M®M'^N ®N'.</i>


<i>Now the homomorphisms of M into N can be added and they can be</i>
<i>multiplied by elements of R to produce new homomorphisms; in fact the</i>
<i>homomorphisms of M into N form an .R-module. This module is denoted</i>
by HomR<i>(M, N) and this is sometimes simplified to Hom(M, N) if there is</i>


<i>no uncertainty as to the ring of scalars. Since f ® f belongs to</i>
<i>Hom(M ® M', N ® AT'), we have a mapping</i>


<i>Hom(M, N) x Hom(M', AT)-> Hom(M ®M',N ® N')</i>


<i>in which (/, / ' ) is taken into / ® / ' , and this mapping is bilinear (see (2.2.7)</i>
and (2.2.8)). It therefore induces a homomorphism


<i>Hom(M, N) ® Hom(M', AT) - ã Hom(M đM\NđN') (2.6.2)</i>
of .R-modules.


</div>
<span class='text_page_counter'>(44)</span><div class='page_container' data-page=44>

<i>Comments and exercises 33</i>


<i>could denote either an element of Hom(M, N) ® Hom(M', N') or an</i>
<i>element of Hom(M ® M ' , N ® AT'); indeed what the mapping (2.6.2)</i>
accomplishes is to change the former interpretation into the latter. In


practice this double meaning does not cause problems. Usually / ® / '
<i>means the homomorphism of M ® M' into N ® JV' and in any event the</i>
element of doubt can be removed by examining the context. The next
exercise has been included to emphasize the difference between the two
meanings of / ® / ' .


<i><b>Exercise 6. Let R be the ring Z/4Z and let I = 2Z/4Z so that I is an ideal of</b></i>


<i>R. Put M = M' = R/I and N = N' = R. Show that in this case the</i>
<i>homomorphism</i>


<i>Hom(M, N) ® Hom(M', N') - ã Hom(M đM\N đ AT)</i>
<i>of (2.6.2) is neither an injection nor a surjection.</i>


We know, from Theorem 4 Cor., that tensor products are right exact.
<i>Thus if 0 —• E —• E —> E" —• 0 is an exact sequence of K-modules and M is an</i>
arbitrary ^-module, then the derived sequence


M ® F - > M ® £ - » M ® £ " - > 0


is exact. However, for some choices of M it can happen that the five-term
sequence


0 - > M ® F - » M ® £ - > M ® F ' - > 0


is not exact. This observation leads on to the definition of an importance
class of /^-modules.


<i><b>Definition. The R-module M is said to be 'flat' or 'R-flaf if whenever 0—•</b></i>



<i>E' —-• E —• E" —> 0 is an exact sequence of R-modules the resulting sequences</i>


0 - ^ M ® F - ^ M ® £ - ^ M ® £ '/- ^ 0 (2.6.3)
<i>and</i>


0 - » £ ' ® M — £ ® M - > F ' ® M - > 0 (2.6.4)
<i>are. exact.</i>


<i>The reader should note that two sequences have been mentioned in the</i>
interests of symmetry. It is very easy to show, using Theorem 2, that if either
one of (2.6.3) and (2.6.4) is exact, then so too is the other. An alternative (but
<i>equivalent) way of defining flat jR-modules is the following: M is flat if and</i>
<i>only if M ® / and f ® M are injections whenever f is an injective</i>
<i>homomorphism. Of course, if either / ® M or M ® / is an injection, then the</i>
other is an injection as well. This again is a consequence of Theorem 2.


</div>
<span class='text_page_counter'>(45)</span><div class='page_container' data-page=45>

<i>34 Some properties of tensor products</i>


<i><b>Exercise 7. Show that if the ideal I contains a non-zero divisor and I is</b></i>


<i>different from R, then R/I is a non-flat R-module.</i>


The next two exercises concern straightforward properties of flat
modules.


<i><b>Exercise 8. Show that an arbitrary direct sum of flat modules is a flat </b></i>


<i>R-module.</i>


<i><b>Exercise 9. Show that if M</b>1,M2,...,Mq are flat R-modules, then</i>



<i>Ml®M2®'"®Mqisa flat R-module.</i>


A deeper result is contained in


<i><b>Exercise 10*. Let M be an R-module with the property that every finitely</b></i>


<i>generated submodule is contained in a flat submodule. Show that M itself is a</i>
<i>flat R-module.</i>


It is now a simple matter to give an example of a flat module which is not
projective. Such an example is provided by


<i><b>Exercise 11 *. Show that the field Q of rational numbers, when considered as</b></i>


<i>a Z-module, is fiat but not projective.</i>


The remaining comments on Chapter 2 all have to do with covariant
extension. We already know from Theorem 8 and its corollary that
projective modules stay projective and free modules stay free when they
undergo such an extension. As the next exercise shows, a similar
observa-tion applies to flat modules.


<i><b>Exercise 12. Let R^>Sbea homomorphism of commutative rings and let M</b></i>


<i>be a flat R-module. Show that M ®R S is a flat S-module.</i>


The main text makes a brief mention of different ways in which covariant
extensions can arise, but so far we have not examined in detail any special
cases. This omission will now be put right.



<i>The simplest situation occurs when we have an ideal / of R and we use the</i>
<i>natural ring-homomorphism R—+R/I. For an .R-module M the </i>
<i>corre-sponding covariant extension is M ®R (R/I) and this we know is naturally</i>


<i>isomorphic to the J?//-module M/IM (see Exercise 3). The reader will find it</i>
an instructive exercise to check that the isomorphism described in Chapter


1, Exercise 5 can be regarded as a special case of Theorem 9.


<i>Next let Xl9 X2,..., Xn be indeterminates. Then R is a subring of the</i>


<i>polynomial ring R[X1, X2,..., Xn~\ and covariant extension by means of</i>


<i>the inclusion homomorphism is referred to as adjunction of the </i>
<i>indeter-minates Xl9 X2,..., Xn. This can be looked at in a more down-to-earth</i>


</div>
<span class='text_page_counter'>(46)</span><div class='page_container' data-page=46>

<i>Comments and exercises 35</i>


Let M be an K-module and consider formal polynomials
Z<i> mvlv2...vHX\lXV22</i> ' ' ' ^n"'


<i>where the coefficients mVlv2...Vn</i> are taken from M. It is clear that we can add


such polynomials and that in this way we obtain an abelian group. Indeed if
we denote the group by M[Xl 9<i> X2,. • •, Xn~\, then M\_Xl9 X2,..., Xn~\ can</i>


<i>be regarded as an R[XU X2,..., X</i>n]-module in which


<i>(Here, of course, r e R and m e M.) It is now a straightforward exercise to</i>


verify that there is an isomorphism


<i>M[X19 X29...9Xn-]vM ®R R[X19 X29..., XJ</i>


<i>of R[Xl9 X29...9 Xj-modules in which mX\'X\2... Xvn» is matched with</i>


<i>m ® XyX^2... XI*. Note that R[Xx, Xl9..., Xn~\ is a free and therefore flat</i>


^-module.


The final example is taken from the general theory of fractions. First we
<i>recall how fractions are formed from the ring R.</i>


<i>Let Z be a non-empty multiplicatively closed subset of R, that is to say Z</i>
<i>has the property that if ax, a2,..., an (n >0) belong to it, then axa2 ... an</i> is


in Z as well. Note that in allowing n = 0 as a possibility we are in fact
<i>assuming that the identity element of R is a member of the multiplicatively</i>
closed set.


<i>We now consider formal fractions r/a, where the numerator r is in R and</i>
<i>the denominator a is in Z. Two such fractions, say rl/al and r2/a2, are</i>


<i>regarded as the same and we write rxjax =r2/a2 if (and only if) aa2rx</i> =


<i>aaxr2</i> for some <jeZ; in other words we introduce an equivalence relation


but without using a special notation for the equivalence classes. It is then
possible to define addition and multiplication of fractions (compatible with
the definition of equality) so that



<i>r r' a'r-\-ar'</i>
<i>a a' aa'</i>
and


<i>r r' rr'</i>
<i>a a' aa''</i>


after which it is readily checked that the fractions form a commutative ring.
<i>This ring is known as the ring of fractions of R with respect to Z and in what</i>
follows it will be denoted by Z-1(jR). Note that the identity element of
Z "1<i>^ ) is 1/1, and that the fraction r/a is the zero element of Z "</i>1^ ) if and
<i>only if a'r = 0 for some a' e Z; also Z ~1 (R) is a trivial ring only when 0 is a</i>
member of Z.


</div>
<span class='text_page_counter'>(47)</span><div class='page_container' data-page=47>

<i>ring-36 Some properties of tensor products</i>


homomorphism


<i><b>K—• JL \K) {Z.O.J)</b></i>


<i>in which r ofR is mapped into the fraction r/1. Let us examine what happens</i>
<i>when we use (2.6.5) to pass from modules over R to modules over Z "* (R).</i>
<i>To this end suppose that M is an .R-module and let xeM ®</i>/?Z~1(.R).


<i>Then x = mx ® Xx + m2</i> ® A2<i> + • * * + ms</i> ® As, where mf- e M and A,- G Z "<i>1 (R).</i>


<i>Now we can write the fractions Xt</i> with a common denominator in Z, say


^. = r./cr, and then



<b>V r*</b>


<i>Thus each element of M ®R</i> Z "<i>l (R) can be written in the form m ® (1/cr),</i>


<i>where m e M and a e Z .</i>


<i><b>Exercise 13*. Le£ m belong to the R-module M and let o belong to the</b></i>


<i>multiplicatively closed subset Z of R. Show that m®(l/<j) = 0 in</i>
<i>M ®</i>R Z ~<i>l<b> (R) if and only if o'm = 0 for some o' e Z.</b></i>


The construction which produced the ring Z~X(JR)<i> from R can also be</i>
applied directly to an .R-module M. Thus we may consider fractions of the
<i>form m/cr, where meM and oeZ; and naturally we treat m1/Gl and m1JG2</i>


<i>as the same if oo2ml =aalm2</i> for some <reZ. These new fractions may be


<i>'added', the sum of m/a and m'/cr' being given by</i>
<i>m rri o'm + om!</i>


<i>o o' GO'</i>


In this way we obtain an abelian group which is frequently denoted by
Z ~x (M). In fact we can go further and regard Z "<i>x</i> (M) as a Z "x (K)-module
with


<i>rm</i>


Finally, using Exercise 13, we can readily verify that there is an


iso-morphism


of Z "x<i> (R)-modules in which mJG of Z ~1 (M) corresponds t o r n ® (l/o). We</i>
leave to the reader the task of filling in the details.


<i><b>Exercise 14*. Let I, be a multiplicatively closed subset of R and letHL~</b>l(R)</i>
<i>be regarded as an R-module by means of the canonical ring-homomorphism</i>
<i>R—•Z"</i>1<i>^). Show that Z</i>- 1<i>( R ) is a flat R-module.</i>


</div>
<span class='text_page_counter'>(48)</span><div class='page_container' data-page=48>

<i>Solutions to selected exercises 37</i>


<i>Suppose that co:R-^S and X:S—+T are homomorphisms of </i>
<i>com-mutative rings. By Theorem 7, if M is an ^-module we have an (R, </i>
<i>S)-isomorphism (M ®RS) ®s T^M ®R (S ®s T) in which (with a </i>


<i>self-explanatory notation) the element (m®s)®t corresponds to m ® (s ® t).</i>
On the other hand, Theorem 3 provides us with an isomorphism
<i>S ®ST^T. Together these yield an isomorphism</i>


<i><b>(M đ</b><b>R</b><b> S) đ</b><b>s</b><b> T ô M đ</b><b>R</b><b> T (2.6.6)</b></i>


<i>in which (m ® 5) ® t is matched with m (x) sr. To begin with (2.6.6) is an</i>
isomorphism of K-modules but, because of the way it operates, we can go
further and say that it is actually an isomorphism of T-modules.
<i>Con-sequently if we extend M by means of to and then extend the result by means</i>
<i>of X the total effect is essentially the same as if we had extended M by means</i>
<i>of X ° co. This is what was meant by saying that covariant extension is a</i>
transitive operation.


<b>2.7 Solutions to selected exercises</b>



<i><b>Exercise 2. Let x</b></i>l 5x2, . . . , xs<i> be elements of an R-module M and</i>


<i>^i? yi) - - - •> ys belong to an R-module N. Suppose that xx ® yt +x2 ® y^ + ' ' '</i>


<i>+ xs ® ys = 0 in M ® N. Show that there exist finitely generated submodules</i>


<i>M' and Nf, of M and N respectively, such that</i>


<i>(i) the xt are in M' and the yt are in N',</i>


<i>and</i>


(ii) xx<i> (x)^! +x2</i>


<i>Solution. Let U(M, N) be the free K-module generated by M x N and let</i>
<i>V(M9 N) be the submodule of U(M, N) generated by elements having one</i>


or other of the following forms:
<i>{mx</i> +m2<i>, n)- (m</i>1?<i> n)-(m2, n),</i>


<i>(m, nx +n2)- (m, n</i>x) - (m, w2),


<i>(rm, n) — r(m, n),</i>


<i>(m,rn)-r(m,n),</i>


where the notation is self-explanatory. Now in the course of solving
<i>Problem 1 of Chapter 1 it was shown that M ®N=U(M, N)/V(M9 N);</i>



<i>moreover in this context m ® n is the natural image of the basis element</i>
<i><b>(m, n) in U(M, N)/V(M, N).</b></i>


<i>Next, because xl đ y</i>x + ã ã • + xs<i> ® ys = 0 in M ®N, we see that</i>


(*i> .Vi)+ ' *' +US<i>» ^s) belongs to V(M, N), that is to say</i>


</div>
<span class='text_page_counter'>(49)</span><div class='page_container' data-page=49>

<i>38 Some properties of tensor products</i>


<i>where rt e R and ^ is an element of U(M, N) having one or other of the four</i>


<i>forms listed above. Note that (2.7.1) means simply that for each (m, n) in</i>
<i>M x N the sum of the coefficients of all the occurrences of (m, n) in the </i>
left-hand side equals the sum of the coefficients of its occurrences in the
<i>right-hand side. But only finitely many basis elements (m, n) are actually present</i>
<i>in (2.7.1) and therefore we can choose finitely generated submodules M' and</i>
<i>AT', of M and N respectively, so that if (m, n) appears on either side of (2.7.1),</i>
<i>then (m, n)eM' x N'. This ensures that x</i>l 5 x2, . . . , xs are in M' and that


yi» )^2» • • • j ys a r e *n<i> N'; it also ensures that (xi,yl)+- - + (x</i>s<i>, ys) is in</i>


<i>V(M\ AT) as well as in U(M', AT). But M' ®N'= U(M\ N')/V(M\ AT) and</i>
therefore


<b>+x2</b>


<i>in M' ® AT' as required.</i>


<i><b>Exercise 3. Let I be an ideal of R and let M be an R-module. Establish an</b></i>



<i>isomorphism M ®R (R/I) % M/IM of R-modules.</i>


<i>Solution. Let (/>: R —• R/I be the natural homomorphism. Because tensor</i>
products are right exact, the exact sequence


<i>0 —>I —>R —</i>


gives rise to the exact sequence


<i>But M (x) R is isomorphic to M under an isomorphism which matches</i>
<i>m®r with rm (see Theorem 3). It follows that there is a surjective</i>
<i>homomorphism M^>M ® (R/I) which maps m into m ® c/>(l), and </i>
<i>further-more the kernel of this homomorphism is the image of / ® M in M under</i>
the result of combining the homomorphisms


<i>I ®M —>R®M ^-+M.</i>


<i>Accordingly the kernel of the homomorphism M-+M ® (R/I) is IM and</i>
<i>thus there is induced an isomorphism M/IM ^M ® (R/I).</i>


<i><b>Exercise 4. Let M be a finitely generated R-module. Show that if M = JM</b></i>


<i>for every maximal ideal J, then M = 0.</i>


<i>Solution. Let m</i>l9 m2<i>, . . . , mn</i> generate M as an .R-module and let J be a


<i>maximal ideal. Then, since m{ belongs to M = JM, we have</i>


<i>mi = ailml +ai2m2 + • • • +ainmn</i>



or


</div>
<span class='text_page_counter'>(50)</span><div class='page_container' data-page=50>

<i>Solutions to selected exercises 39</i>


<i>where the atj are in J. Consequently if D is the determinant</i>


<b>21 0 2 2 -1<sub> • "</sub></b><i><b> a</b><b><sub>2n</sub></b></i>


<i>then Dmt = 0 for i = l,2, . . . , n . It follows that D belongs to the ideal,</i>


<i>Ann^(M) say, formed by the elements of R that annihilate M. But D =</i>
<i>a + (—1)", where aeJ, and therefore D$J. This shows that AnnR(M)^J,</i>


i.e. it shows that Ann^(M) is not contained in any maximal ideal.
<i>Accordingly AnnR(M) = R and now we see that M = 0.</i>


<i><b>Exercise 5. Let M be an R-module. Show that the following two statements</b></i>


<i>are equivalent:</i>


<i>(i) for a finitely generated R-module N we have M ® N = 0 only when</i>
<i>N = 0;</i>


<i>(ii) for every maximal ideal J, of R, JM is different from M.</i>
<i>Solution. Assume (i) and let J be a maximal ideal. Then, since R/J is singly</i>
<i>generated and non-zero, we have M ®(R/J)^0. But, by Exercise 3,</i>
<i>M ® (R/J) is isomorphic to M/JM and so JM # M. Accordingly (i) implies</i>
(ii).


<i>Assume (ii) and let iV be a finitely generated K-module such that</i>


<i>M ®RN = 0. If now J is a maximal ideal, then (Chapter 1, Exercise 5)</i>


<i>(M/JM) ®R/J(N/JN)zz(M ®RN)/J(M ®RN) = 0. But R/J is a field and</i>


<i>M/JM and N/JN are vector spaces over R/J and therefore they are free</i>
<i>K/J-modules. Furthermore, by hypothesis M/JM ^ 0. It follows (Chapter 1,</i>
<i>Theorem 3) that N/JN = 0. We now know that N = JN for every maximal</i>
<i>ideal J and so we may conclude, by virtue of Exercise 4, that N = 0. This</i>
completes the solution.


<i><b>Exercise 10. Let M be an R-module with the property that every finitely</b></i>


<i>generated submodule is contained in a flat submodule. Show that M itself is a</i>
<i>flat R-module.</i>


<i>Solution. Let / : U—> V be an injective homomorphism of K-modules. It</i>
<i>will be enough to show that f ®M: U ®M—>V ®M is an injection as</i>
<i>well. To this end let £ = ul®m1+u2®rn2+- - +us®ms, where uteU</i>


<i>and mteM, belong to the kernel of f®M. Then / ( w</i>1) ® m1+


-+/(w<i>s) ®ms = 0\nV ®M.lt now follows, from Exercise 2, that there exists</i>


<i>a finitely generated submodule V* (of V) containing f(ux), f(u2),..., f(us)</i>


<i>and a finitely generated submodule M* (of M) containing m</i>l9 ra2<i>,..., ms</i>


</div>
<span class='text_page_counter'>(51)</span><div class='page_container' data-page=51>

<i>40 Some properties of tensor products</i>


<i>/(Mi) ® m1 +f{u2) ® m2</i> + • • •



<i>in V* ® M * . Choose a flat submodule M' of M so that M*^M'. The</i>
<i>obvious homomorphism K* đ M* ã V đ M' then shows that</i>
<i>f(u1)®ml + -'+f(us)®ma = 0 in V ®M'.</i>


<i>Let ux</i> ® mx + w2<i> đ m2 + ã ã ã + us</i> ® ms<i> = x in £/ ® M' and consider the</i>


natural commutative diagram


<i>(7 đ M' ã V đ M'</i>


(/ ® M • K ® M.


<i>The image of x in F ® M' is zero, and / ® M' is an injection because / is an</i>
<i>injection and M' is flat. It follows that x = 0. But £ is the image of x under the</i>
homomorphism 1 / ® M ' — • L ^ M so £ = 0 a s well.


<i><b>Exercise 11. Show that the field Q of rational numbers, when considered as a</b></i>


<i>Z-module, is flat but not projective.</i>


<i>Solution. Let N be a finitely generated Z-submodule of Q. Then there exists</i>
<i>k e Z, k 7^0 such that N is contained in the Z-submodule of Q generated by</i>
l//c. But the Z-submodule generated by 1/fe is isomorphic to Z and therefore
<i>N is isomorphic to an ideal of Z. It follows that AT is a free Z-module so a</i>
<i>fortiori it is Z-flat. Exercise 10 now shows that Q is a flat Z-module.</i>


We next show that Q is not a submodule of a free Z-module. (This is more
than enough to establish that Q is not projective.) Assume the contrary and
<i>let Q be a submodule of a free Z-module F with a base {£,.} ieI. Then we have</i>



<i>a relation 1 = £ mt£h where mteZ and only finitely many mt's are non-zero.</i>


<i>Choose j e / so that m</i>;<i><b> ^ 0 and then choose p e Z, /? # 0 so that p does not</b></i>


divide m^. Now


<i>where the n,- are integers, and it follows from this that pn^ = m, and therefore</i>
<i>p divides my</i> This is the desired contradiction.


<i><b>Exercise 13. Let m belong to the R-module M and let a belong to the</b></i>


<i>multiplicatively closed subset Z of R. Show that m®(l/o-) = 0 in</i>
<i>M ®</i>K<i> E ~ * CR) if and only if a'm = 0 for some a' e S.</i>


<i>Solution. First suppose that m ® [I/a) = 0 in M ®R</i> 1 ~x (K). By Exercise 2,


</div>
<span class='text_page_counter'>(52)</span><div class='page_container' data-page=52>

<i>Solutions to selected exercises 41</i>


<i>respectively, such that m ® (I/a) = 0 in M' ® L. Choose ax e Z so that L c</i>


<i>Ril/aaJ; it then follows that m ® (1/(7) = 0 in M (^Ril/aậ</i>
Consider the exact sequence


<i>4> 1</i>


<i>0 —>I —>R —>R >0</i>


<b>(7(7 x</b>



<i>of /^-modules, where (f)(r) = r/aa1. First we note that / consists of all the</i>


<i>elements r of R such that c'r = 0 for some a ' e l Next there is induced an</i>
exact sequence


<i>M ® / >M ®R >M ®R >0</i>


<b>(7(7 x</b>


<i>and, moreover, m ® al</i> belongs to the kernel of M ® 0, that is to say it


<i>belongs to the image of M ® / in M ® R. The isomorphism M ® R % M in</i>
<i>which m ® r is matched with rm now shows that o^m belongs to IM. We can</i>
therefore find ( j2<i>e l s o that a2Gim = 0. But (72(7! e S so m is annihilated by</i>


an element of Z.


<i>Conversely suppose that a'm = 0 for some a' e Z. Then in M ®^ Z "1 (R)</i>
we have


<i>1 a' 1</i>
<i>m ® — = m ® — = a'm ® — = 0</i>


<i><b>a aa aa</b></i>


as required.


<i>Exercise 14. Let Jibe a multiplicatively closed subset of R and letT~l (R) be</i>
<i>regarded as an R-module by means of the canonical ring-homomorphism</i>
K — • Z "1<i>^ ) . Show that Z "</i>1<i>^ ) is a flat R-module.</i>



<i>Solution. Let / : M' —» M be an injective homomorphism of /^-modules. It</i>
is sufficient to show that / ® Z "<i>l (R) is also an injection. To this end let x in</i>
<i>W ® Z " * (R) belong to the kernel of / ® Z ~</i>x<i> (R\ and let x = m' ® (I/a),</i>
<i>where m! e M' and a e Z. (Every element of M' ®^ Z ~</i>x (K) can be written in
<i>this form.) Then f(m') ® (l/(7) = 0 and therefore, by Exercise 13, af(m') = 0</i>
<i>for some a' e Z. Accordingly f(a'm') — 0 and therefore a'm' = 0 because / is</i>
an injection. Finally


<i>x = m'</i>


and the solution is complete.


<b>1</b>


<i><b>a</b></i> <i><b>m' (</b></i>


<i><b>a</b></i>


<i><b>aa'</b></i> <i><b>a</b></i><b>'m'(</b>


<b>1</b>
<b>*> —</b>


</div>
<span class='text_page_counter'>(53)</span><div class='page_container' data-page=53>

<b>3</b>



Associative algebras



<b>General remarks</b>



Our main objective is to discuss the important algebras that can be
derived from a module, and so in this, the last of the introductory chapters,
we shall study those general aspects of the theory of Algebras that are
relevant to our goal. At the outset it needs to be stressed that we shall be
concerned solely with algebras that are associative and possess an identity
element. Accordingly, from now on, the term algebra will only be used in
this restricted sense.


<i>As before R and S denote commutative rings, each with an identity</i>
element, and a ring-homomorphism between them is required to take
identity element into identity element. Later, when we come to consider
homomorphisms of algebras, these too will be required to preserve identity
elements.


Finally when dealing with tensor products of modules we shall
sometimes omit from the symbol (g) the suffix which indicates the
underlying ring. In this chapter whenever the suffix is left out the ring in
<i>question is always R.</i>


<b>3.1 Basic definitions</b>


<i>Let A be a (not necessarily commutative) ring with an identity</i>
<i>element lA, and at the same time let it be a module over the commutative</i>


<i>ring R. We suppose that the sum of two elements of A is the same whether</i>
<i>we use the ring or the module structure. In these circumstances A is called</i>
<i>an R-algebra provided that</i>


<i>r{a1a2) = (ra1)a2=a1(ra2) (3.1.1)</i>



<i>whenever au a2eA and reR. This condition is equivalent to</i>


<i>(r1a1)(r2a2) = (rlr2)(ala2l (3.1.2)</i>


where now r1?<i> r2 denote elements of R.</i>


</div>
<span class='text_page_counter'>(54)</span><div class='page_container' data-page=54>

<i>Basic definitions 43</i>


<i>Let A be an K-algebra. The mapping</i>


<i>cj):R-+A, (3.1.3)</i>


<i>defined by 0(r) = r\A</i>, is both a ring-homomorphism and a homomorphism


<i>of R-modules. Furthermore, since (j)(r)a = ra = a(j)(r), (j){R) is contained in</i>
<i>the centre of A We shall call (3.1.3) the structural homomorphism of the </i>
JR-algebra.


This last concept provides an alternative way of looking at K-algebras.
<i>Suppose that A is a ring with identity element and assume that we are given</i>
<i>a ring-homomorphism cp: R—*A which preserves identity elements and</i>
<i>maps R into the centre of A. If we turn A into an ^-module by putting</i>
<i>ra = (t>(r)a, then A becomes an R-algebra with (f> as its structural </i>
<i>homo-morphism. For example, we see that R itself is an ^-algebra with the</i>
identity mapping as the structural homomorphism.


An important example of an K-algebra arises in the following way. Let M
<i>and N be R-modules. The R-linear mappings of M into N can be added and</i>
<i>they can be multiplied by elements of R. In fact they form the R-module</i>
<i>which is denoted by Hom^ (M, N). When N = M these homomorphisms are</i>


<i>the endomorphisms of M and in this case we use the notation End^(M)</i>
rather than HomR<i> (M,M). Now if /, g belong to EndR(M), then so does</i>


<i>f°g. Indeed it is easy to check that End</i>R (M) becomes a ring with identity if


<i>the product of / and g is taken to be / ° g. But for r e R we have</i>


<i>so that in fact End^(M) is an K-algebra. We call EndR(M) the </i>


<i>endo-morphism algebra of M. The identity element of EndR{M) is the identity</i>


<i>mapping of M and in the structural homomorphism R—>EndR(M) an</i>


<i>element r, of R, is mapped into the corresponding homothety that is the</i>
<i>mapping M —• M in which meM goes into rm.</i>


<i>We shall now introduce some further terminology. Let A and B be </i>
<i>#-algebras. A mapping f:A—+B is called an algebra-homomorphism, or a</i>
<i>homomorphism of R-algebras, if it is both a homomorphism of rings</i>
(preserving identity elements) and a homomorphism of R-modules. For
<i>example the structural homomorphism R—+A is a homomorphism of </i>
<i>R-algebras. Note that if (/>: R —• A and \j/: R—+B are the structural </i>
<i>homo-morphisms of A and B, then a mapping f:A—+B is an </i>
<i>algebra-homomorphism if and only if it is a algebra-homomorphism of rings and / ° c/> = \\t.</i>
<i>Naturally a bijective homomorphism is called an </i>
<i>algebra-isomorphism.</i>


</div>
<span class='text_page_counter'>(55)</span><div class='page_container' data-page=55>

<i>44 Associative algebras</i>


<i>intersection of any family of subalgebras of A is again a subalgebra. Hence if</i>


<i>U is any subset of A, there will be a smallest subalgebra, B say, containing U.</i>
<i>We call B the subalgebra generated by U. As an K-module B is spanned by</i>
<i>all products ulu1... un, where ut</i>e (7 and n > 0 . Here when n = 0we have to


<i>do with the empty product and this is interpreted as being the identity</i>
<i>element of A. Of course the inclusion mapping into A of one of its</i>
subalgebras is an algebra-homomorphism.


<i>Finally suppose that / is a two-sided ideal of the ^R-algebra A. Then / is</i>
<i>necessarily an K-submodule. Accordingly A/I is both a ring and an </i>
<i>R-module. Indeed A/I is an jR-algebra. Note that the natural mapping of A</i>
<i>onto the factor algebra A/I is a homomorphism of K-algebras.</i>


<b>3.2 Tensor products of algebras</b>


<i>LetAuA2,. • •,;4</i>P<i>be R-algebras. Then Ax đRA2 đR- ã ã đRApis</i>


certainly an /^-module. Indeed, as will be established shortly, it has a
natural structure as an K-algebra.


Consider the multilinear mapping


<i>AlxA2X'-xApxA1xA2x-- x Ap—>AX ®A2®- - ®Ap</i>


<i>in which (al9 a2,..., ap, a\,af2,..., a'p) is mapped into a±a\ ® a2a!2 đ ã ã ã</i>


<i>đapa'p. This (Chapter 1, Theorem 2) induces a homomorphism</i>


<i>Al®A2®'"®Ap®Al®A2®"-®Ap</i>



<i>-+Al®A2®--®Ap</i> (3.2.1)


<i>of ^-modules. Moreover (Chapter 2, Theorem 1) we have an </i>
R-isomorphism


<i>® Ap)</i>


<i>p</i> <i>® - ' ® A</i> <i>p</i> (3.2.2)


<i>in which, with a self-explanatory notation, (a1đ--- đap)đ(a'lđ ã • •</i>


<i>®a'p) corresponds to ax ® • • • ®ap®a\ ® • • • ®a'p. If therefore we</i>


combine (3.2.1) and (3.2.2) the result is a homomorphism


<i>(Ax đ - ã ã ® Ap) ® (Ax ®'' - ® Ap)-+ Ax ® A2 đ ã ã ã đ Ap</i> (3.2.3)


<i>in which (ax ® • • • ® ap) ® (a[ ® • • • ® a'p) is carried over into axa\ ®</i>


<i>a2d2 ®"-®apa'p.</i>


We now define a mapping


<i>/*: {Ax ® - • • ® Ap) x {A^ ® - • • ® Ap)</i>


<i>-+Al®A2®--®Ap</i> (3.2.4)


<i>by requiring jx(x, x% where x and x' belong to Ax đ A2 đ ã ã ã đ Ap, to be</i>


</div>
<span class='text_page_counter'>(56)</span><div class='page_container' data-page=56>

<i>Tensor products of algebras 45</i>



<i>= ala'l ® a2a'2 đ ã ã ã đ apa'p. (3.2.5)</i>


<i>It follows that if x = al®-- ®ap, x' = a\®'' * ®a'p, and xn = a'{®-'</i>


<i>®afp, then</i>


<i>rivix, x'\ x") = fi(x, ii{x\ x")). (3.2.6)</i>


<i>But \i is bilinear. Consequently (3.2.6) holds when x, x' and x" are any three</i>
elements of Xx ® A2<i> ® * * • ® Ap.</i>


<i>We are now ready to turn Ax đ A2 đ ã ã ã đ Ap</i> into an ^-algebra. To do


<i>this we define the product of two of its elements, say x and x\ to be fi(x, x'). It</i>
then follows from (3.2.6) that multiplication is associative; and the
<i>bilinearity of \i ensures that it is distributive with respect to addition. Let et</i>


<i>be the identity element of At. By (3.2.5) we have</i>


<i>'" ®ep,a1®-- ®ap)</i>


and therefore


<i>for all x in Ax đ A2 đ ã * ã đ Ap. Thus A^ ® A2 ®' * * ® Ap</i> is a ring with


identity. Finally for r e R we have
<i>fi(rx, x') = rfi(x, xf) = fi(x, rx')</i>


<i>and therefore A±® A2®- - - ® Ap\s &n /^-algebra.</i>



We summarize our conclusions so far.


<i><b>Theorem 1. Let A</b><b>l9</b><b> A</b><b>2</b><b>,.. ã, A</b>p be R-algebras. Then</i>


<i>AlđRA2 đR'-đRAp</i>


<i>is an R-algebra, where the R-module structure is the usual one and the product</i>
<i>of alđa2đ- - ã đap and a\ đ a!2 đ ã ã ã đ a'p is axa\ đa2a'2 ® • • • ®apdp.</i>


It will now be shown that the results of Section (2.1) yield certain
isomorphisms between algebras. We can deal with these fairly rapidly.
<i><b>Theorem 2. Let A</b>u<b> A</b><b>2</b><b>,...,A</b><b>P</b> and BUB2, ...,Bq be R-algebras. Then</i>


<i>there is an isomorphism</i>


<i>A1đR'"đRApđBlđR"-đRBq</i>


<i>ô ( At đRã ã ã đR Ap) đR (Bx đR'"đRBq)</i>


<i>of R-algebras in which al®---®ap®bl®--'®bq is associated with</i>


<i>(a1®--®ap)®(b1®'-®bq).</i>


<i>Proof By Theorem 1 of Chapter 2, there is an isomorphism /, of R-modules,</i>
<i>that maps Axđ ã ã ã đApđB1đ ã ã ã đBq onto (Axđ ã ã ã đAp)đ(B1đ</i>


</div>
<span class='text_page_counter'>(57)</span><div class='page_container' data-page=57>

<i>46 Associative algebras</i>


<i>f(a1đ'-đapđblđ--đbq)</i>



<i>= K đ ã ã ã đ ap) đ (b</i>x đ • •


x


<i>b'q. Then, by Theorem 1,</i>


<i>Let x = ax</i> ® • • • ® ap ® bx<i> ® • • • ® bq and x' =</i>


<i>xx' = ala'l</i> ® •


and therefore


<i>/(**') = ( a ^ i đ ã ã • ® apa'p) ® ( b ^ ; ® • • • ® b</i>gb;)


<i>But / is R-linear. Hence if y and / a r e any two elements of Ax</i> đ ã ã ã đ 4pđ


<i># i ® • • • ® Bq, then f(yy') = f(y)f(y')- The theorem follows because / is</i>


bijective.


Before we leave this result we note that the argument used to deduce the
corollary to Theorem 1 of Chapter 2 now shows that there is an
isomorphism


<i>(Al®RA2)®RA3*A1®R(A2®RA3) (3.2.7)</i>


<i>of K-algebras in which (al ® a2) ® a3 corresponds to al ® (a2 ® a3).</i>


The next theorem is derived, in much the same way, from Theorem 2 of


Chapter 2. The details are left to the reader.


<i><b>Theorem 3. Let A</b>l9A2,..., Ap be R-algebras and let (i</i>l5 i2, . . . , ip) be a


<i>permutation of ( 1 , 2 , . . . , p). Then there is an isomorphism</i>


<i>Al®RA2®R--®R Ap&Atl ®R Ah ®R"- ®R Aip</i>


<i>of R-algebras in which ax đ a2 đ ã ã ã đ ap corresponds to aix đ aiz</i> đ ã ã ã


<b>v</b>



It will be recalled that # itself is an K-algebra.


<i><b>Theorem 4. L^f A be an R-algebra. Then there is an isomorphism</b></i>


<i>R ®RA ^A of R-algebras which matches r ®a with ra. There is a similar</i>
<i>algebra-isomorphism A ®RRzzA, where a ®r is matched with ra.</i>


<i>Proof We need only consider the first assertion. By Theorem 3 of Chapter</i>
<i>2, there is an isomorphism f:R®RA—+A of K-modules for which</i>


<i>f(r®a) = ra. That / is compatible with multiplication is clear because</i>
<i>(r1a1)(r2a2)=(r1r2)(a1a2).</i>


We next examine tensor products in relation to homomorphisms of
<i>algebras. To this end let A1,A2,..., Ap and Bx, B2,..., Bp</i> be .R-algebras


<i>and suppose that we are given algebra-homomorphisms f: At—>Bi for i =</i>



1,2,...,/?. From Section (2.2) we know that
<i>/ i ® fi ® ' " ' đ fP- Ax đR A2 đR ã ã ã ®R Ap</i>


</div>
<span class='text_page_counter'>(58)</span><div class='page_container' data-page=58>

<i>Graded algebras 47</i>


<i>is a homomorphism of R-modules. We claim that in the present instance</i>
<i>it is actually a homomorphism of R-algebras. For it is clear that fx</i> ®


<i>fi ® ' ' * ® fp</i> takes identity element into identity element. Now suppose


<i>that x and xf belong to A1 đ A2 đ ã ã ã đ Ap. It suffices to show that the</i>


<i>image of xx' is the product of the separate images of x and x'. Indeed it is</i>
<i>enough to prove that this is so when x and x' have the special forms</i>
<i>al đ a2 đ ã * * đ ap and a\ đ a'2 đ ã • • ® a'p. But in these circumstances</i>


what we wish to prove is obvious.


<b>3.3 Graded algebras</b>


<i>Let A be an R-algebra and {An}nel a family of R-submodules of A</i>


<i>indexed by the integers. The family is said to constitute a grading on A</i>
provided that


<i>(i) A=^An</i> (As.),


<i>neZ</i>
and



<i>(ii) areAr and aseAs together imply that araseAr+s.</i>


<i>An K-algebra with a grading is known as a graded R-algebra. The elements</i>
<i>of An are said to be homogeneous of degree n. Thus condition (ii) says that the</i>


product of two homogeneous elements is again homogeneous, and the
degree of the product is the sum of the degrees of the two factors.


<i>We shall frequently meet with situations where An=0 for all n<0. A</i>


<i>grading with this additional property will be called a non-negative grading.</i>
<i>Suppose that A is graded by {An</i>}<i> neI. Then an element a e A has a unique</i>


representation in the form


<i>where an e An</i> and only finitely many terms on the right-hand side are


<i>non-zero. We call the summands in the representation the homogeneous</i>
<i>components of a.</i>


<i><b>Theorem 5. Let {A</b>n}neI be a grading on the R-algebra A. Then the identity</i>


<i>element \A belongs to Ao.</i>


<i>Proof. Let e be the homogeneous component of lA</i> of degree zero and let


<i>aneAn. By comparing the components of degree n in the equations \Aan =</i>


<i>an = anlA we see that san = an = ane. But every element of A is a sum of</i>



<i>homogeneous elements and therefore ex = x = xe for all xeA. Accordingly</i>
<i>lA=s and the theorem is proved.</i>


<i><b>Corollary. A</b>o is an R-subalgebra of A.</i>


</div>
<span class='text_page_counter'>(59)</span><div class='page_container' data-page=59>

<i>R-48 Associative algebras</i>


algebra, by its homogeneous elements of degree one. The relevant facts are
recorded in the next lemma.


<i><b>Lemma 1. Let {A</b>n}nel be a grading on the R-algebra A, and suppose that A</i>


<i>is generated, as an R-algebra, by Av Then the grading is non-negative, and,</i>


<i>for p>\, each element of Ap is a sum of products of p elements of Ax.</i>


<i>Furthermore Ao is generated, as an R-module, by the identity element 1A and</i>


<i>therefore Ao is contained in the centre of A.</i>


<i>Proof All the assertions become obvious as soon as it is noted that (i) A is</i>
<i>generated, as an R-module, by products of elements of Ax</i> (including the


<i>empty product), and (ii) such a product is homogeneous of degree p, where p</i>
is the number of factors.


<i>We need some further terminology. Let A and B be graded jR-algebras. A</i>
<i>mapping f:A—>B is called a homomorphism of graded algebras if it is a</i>
homomorphism of /^-algebras which preserves degrees, i.e. if it satisfies
<i>f(An)^Bn for every n. (Here {An}neI and {Bn}neZ</i> are the respective



<i>gradings.) By an isomorphism of graded algebras we naturally mean a</i>
<i>bijective homomorphism of graded algebras. If f:A—>B is such an</i>
<i>isomorphism, then, for each n, f maps An isomorphically onto Bn.</i>


<i>Now let if be a two-sided ideal of a graded R-algebra A.</i>


<i><b>Definition. The two-sided ideal H is called 'homogeneous' if whenever aeH</b></i>


<i>all the homogeneous components of a belong to H as well.</i>


<i><b>Lemma 2. Let H be a two-sided ideal of a graded R-algebra A. Then H is</b></i>


<i>homogeneous if and only if it can be generated by homogeneous elements.</i>


<i>Proof. If H is homogeneous, then it is certainly generated by the</i>
<i>homogeneous components of the various members of H. Thus H has a</i>
homogeneous system of generators.


Conversely suppose that Q is a set of homogeneous elements and that Q
<i>generates H. Then each element of H is a sum of elements of the form axb,</i>
<i>where a, be A and x e Q. Indeed we may suppose that a and b are themselves</i>
<i>homogeneous. Thus if a e H, then a is a sum of homogeneous elements of H</i>
<i>and therefore the homogeneous components of a belong to H. This</i>
completes the proof.


<i>Now suppose that H is a homogeneous two-sided ideal of the graded </i>
<i>R-algebra A. Put</i>


<i><b>H</b><b>n</b> = HnAn. (3.3.1)</i>



Then


</div>
<span class='text_page_counter'>(60)</span><div class='page_container' data-page=60>

<i>Graded algebras 49</i>


<i>We already know that AjH — B say is an K-algebra. Suppose that we put</i>


<i>BH = (AH+H)/H. (3.3.3)</i>


<i>Then Bn is an R-submodule of B and</i>


<i>We claim that the sum (3.3.4) is direct. To see this suppose that £„ bn</i> = 0,


where foneBn and there are only finitely many non-zero summands. For


<i>each n e / w e choose an e An so that bn = an</i> + //, taking care to arrange that


an<i> = 0 whenever bn</i> = 0. Then £„ an<i> belongs to / / and therefore, because H is</i>


<i>homogeneous, all the an are in H. But this means that all the bn</i> are zero.


Accordingly our claim is established.


<i>Clearly the product of an element of Br and an element of Bs</i> is an element


<i>of Br+S. Hence the R-algebra A/H is graded by its submodules</i>


<i>{(An +H)/H}neI and the natural mapping A —+A/H is a homomorphism of</i>


<i>graded algebras. This mapping induces, for each n, a homomorphism of An</i>



<i>onto Bn whose kernel is An <^H = Hn. Thus we have isomorphisms</i>


<i>AJHH*(Am + H)/H (3.3.5)</i>


of R-modules and these can be combined to give an isomorphism


<i>A/H*ZAJHn</i> (d.s.). (3.3.6)


<i>nsZ</i>


In the first instance (3.3.6) is just an isomorphism of /^-modules, but of
course there is a unique way to turn the right-hand side into a graded
JR-algebra so that (3.3.6) becomes an isomorphism of graded JR-algebras.


<i>We next investigate the tensor product of A(1\ A{2\..., Aip\ where each</i>
<i>A{i)</i> is a graded R-algebra. By Theorem 5 of Chapter 2,


<i>Ail) ® Ai2) ® • • • ® Aip) = £ A^ đ A\2) đ ã ã ã đ A\p)</i> (d.s.),
where the direct sum is taken over all sequences (ils<i> i2,..., ip) of p integers.</i>


Put


<i>A = A{1)</i> đ A(2)<i> đ ã • • ® Aip)</i> (3.3.7)
<i>and for each sequence / = (il9 i2,..., i</i>p) let


<i>A^A^QAW® - - - ®A\P\ (3.3.8)</i>


<i>Then (to recapitulate) A is an R-algebra and</i>



<i>A = Z*i (d-s.) (3.3.9)</i>
/


<i>this being a direct sum of/^-modules. Furthermore if J= (JiJ2> • • • JP) is a</i>


</div>
<span class='text_page_counter'>(61)</span><div class='page_container' data-page=61>

<i>50 Associative algebras</i>


It will be convenient to put


<i>\I\ = ii+i2 + ~' + ip (3.3.10)</i>


and


<i><b>?</b></i>

<i><b> R</b></i>

<i><b>"- ®</b></i>

<i><b>R</b></i>

<i><b>A\</b></i>

<i><b>P</b></i>

<i><b>K (3.3.11)</b></i>



<i>\l\ = n \l\ = n</i>


Then as a consequence of (3.3.9) we have


<i>A=YaAn</i> (d.s.) (3.3.12)


<i>nel</i>


<i>and it is clear that the product of an element of An and element of Ak</i> belongs


<i>toAn+k.</i>


<i>Let us sum up. It has now been shown that if A(1\ A{2\..., A{p)</i> are
<i>graded K-algebras, then A{1) đRAi2) đR- ã ã đRA{p) is also a graded </i>



<i>R-algebra, where the module of homogeneous elements of degree n is</i>


We call {An}weZ<i> the total grading on the tensor product. Evidently if the</i>


<i>gradings on A(1\ A{2\ . . . , Aip)</i> are non-negative, then the total grading on
<i>A{i) ® A{2) ® • • • (x) Aip)</i> is non-negative as well.


We next review some of our basic isomorphisms from the standpoint of
<i>the theory of graded algebras. Suppose then that A{1\ A{2\ . . . , A(p)</i> and
£(1)<i>, B{2\ . . . , Biq)</i> are graded K-algebras. It will be recalled that Theorem 2
provides an explicit isomorphism


<i>^(A(1) ®R-" ®RA(P)) ®R</i> (£(1)<i> ®R- • • ®RB™) (3.3.13)</i>


<i>of ordinary, that is ungraded, /^-algebras. However, on this occasion each</i>
side of (3.3.13) is a graded algebra and it is clear (from the way the
isomorphism operates and the way the gradings are defined) that the
degrees of homogeneous elements are preserved. Accordingly what we have
<i>is an isomorphism of graded algebras.</i>


Let us consider Theorem 3 in a similar way. To this end suppose that
<i>(A*i> M2> • • •» Vp) i</i>s a<i> permutation of ( 1 , 2 , . . . , p). Then the isomorphism</i>


<i>A(1)đRAi2)đR'"đRAip)</i>


ôi40<l)<i> đRAUl2) đR'" đRAUlo)</i> (3.3.14)


that we get from Theorem 3 preserves degrees and so it too is an
isomorphism of graded algebras.



<i>In the case of Theorem 4 a little more explanation is needed. If B is an </i>
<i>R-algebra we can obtain a grading on B by putting B0 = B and Bn = 0</i>


</div>
<span class='text_page_counter'>(62)</span><div class='page_container' data-page=62>

<i>A modified graded tensor product 51</i>


graded K-algebra. Then Theorem 4 provides algebra-isomorphisms


<i>R®RA*A (3.3.15)</i>


and


<i>A®RR*A. (3.3.16)</i>


<i>It is readily checked that these are isomorphisms of graded algebras</i>
<i>provided that we endow R with the trivial grading.</i>


Our final observation in this section concerns homomorphisms
<i>ft: A{i)-+B{i\ where i = 1 , 2 , . . . , p, of graded algebras. It was observed in</i>


Section (3.2) that / i ® /2 ® * * * ® /p is a homomorphism


<i>A{1) đRAi2) đR> ã ã đR</i>A(p)<i> B{1) ®RBi2) ®R-" ®RB(P)</i>


of K-algebras. However, in the present situation it has the additional
<i>property of preserving degrees and therefore, this time, / i đ f2 đ ã ã * ® fp</i> is


a homomorphism of graded algebras.


<b>3.4 A modified graded tensor product</b>



<i>Let A{1\ A{2\ . . . , Aip)</i> be graded ^-algebras. We have shown that
<i>A{1) ® A(2) ® • • • ® Aip) = A say has a natural structure as a graded </i>
<i>K-algebra. We propose to modify this structure. The outcome will be that A</i>
will remain a graded ^-algebra with the same R-module structure, the same
grading and the same identity element. Only the definition of the product of
<i>two elements of A will be changed.</i>


<i>Let / = (I'I, i2, - -, ip) be a sequence of p integers and put</i>


<i>as we did before. Then A is the direct sum of the A{ so that each as A has a</i>


unique representation in the form


fl = Ia,, (3-4.1)


<b>/</b>


<i>where aleAl</i> and only finitely many of the summands are non-zero.


<i>Next let J = {ji,j2i • - • JP) be a second sequence of p integers. The</i>


<i>multiplication already introduced on A provides a bilinear mapping</i>


<b>M/J:<sub> ^./</sub> x</b><i><b><sub> Aj >A</sub></b></i>
<i><b>I+J</b><b>,</b></i>


<i>w h e r e / + J = (ix -\-j1, i2 +j2, . . . , ip</i> + ; ' ) . P u t


(3.4.2)



JV(J,J)=£y.



<i><b>r>s</b></i>
and


</div>
<span class='text_page_counter'>(63)</span><div class='page_container' data-page=63>

<i>52 Associative algebras</i>


Suppose that


<i><b>I J</b></i>


<i>are elements of A. Then the modified product of a and a' is defined to be</i>


5 > ( / , J W a , , a ; ) . (3.4.4)
<i><b>IJ</b></i>


<i>Thus if ait</i> đ ai2 đ ã ã ã đ aip<i> belongs to Aj and a^ đ fl}</i>2 đ ã ã • ® a}p belongs


<i>to Xj, then, whereas their ordinary product is aixa'h</i> đ ôl2ô}2 đ * ã * đ af<i> a) ,</i>


their modified product is


e(/, J ) ^ ^ ® al2fl}2<i> ® • • • ® aipa'jp. (3.4.5)</i>


Clearly modified multiplication is distributive with respect to addition
<i>on both sides, and e{1) đ e(2) đ ã ã ã đ e(p\ where e(r)</i> is the identity element of
<i>A(r\ is still neutral for the new multiplication. Also the property that</i>
corresponds to (3.1.1) continues to hold. Hence in order to check that we
still have an R-algebra it suffices to verify that modified multiplication is
associative.



<i>L e t I = (il9i2>->>ip)> J = UiJ2>- • • JP)</i> a n d<i> K = (kl,k2,... ,kp) b e</i>


<i>sequences of integers and suppose that xeAj, yeAj and zeAK. It will</i>


<i>suffice to verify the associative law in the case of x, y and z. Indeed we may</i>
<i>go further and assume that x = aix đ ã ã ã ® at , y = a'jx ® • • • ® a) and z =</i>


<i>0*, ® ''' ®a'k • ^</i>u t n o w5 in yie w °f (3.4.5) what we wish to establish will


follow if we show that


<i>9 K).</i>


<i>However, this is clear because e(/ + J, K) = s(I, K)e(J, K) and e(/, J + K) =</i>
<i>e(/, J)£(/, K).</i>


<i>The new K-algebra will be called the modified tensor product of</i>
<i>A{1\ A{2\ . . . , A{p\ It will be denoted by</i>


<i>Ail)</i> ®K<i> A{2) ®R'"®R A{p)</i> (3.4.6)


<i>in order to distinguish it from A{1) ® Ai2) ® • • • ® Aip). Note that the total</i>
<i>grading on A{1) đ A{2) đ ã ã ã đ Aip)</i> also serves as a grading on
<i>A{1)</i> đ A(2)<i> đ ã ã • ® Aip\ Finally we recall that the two tensor products have</i>
<i>the same identity element; and that for x e At, y e A3</i> the modified product of


<i>x and y is e(/, J) times the ordinary product.</i>


The new tensor product is important in the theory of exterior algebras. In


the remainder of this section we shall prepare the way for its application.
<i><b>Theorem 6. Let A</b>{1\ A(2\ . . . , Aip) and B(1<b>\ B</b><b>(2</b><b>\ . . . , B</b><b>(q)</b> be graded </i>
<i>R-algebras. Then</i>


</div>
<span class='text_page_counter'>(64)</span><div class='page_container' data-page=64>

<i>A modified graded tensor product 53</i>


<i>and</i>


<i>(A^ ®R'--®R A^) ®R</i> (B(1)<i> ®R--®RB™) (3.4.8)</i>


<i>are isomorphic graded R-algebras under an isomorphism which</i>
<i>matches a{1)đ ã ã ã đaip)đb{1)đ • • • ®biq) with (a{1)® • • • ®aip))®</i>


<i>Proof. We already know that there is an isomorphism / , of graded</i>
<i>algebras, which maps A{1)đ ã ã ã đAip)đB{1)đ • • • ®Biq)</i> onto
<i>(Ail) ® • • • ® A{p)) đ (B{1) đ" ã đ Biq)) and matches elements in the way</i>
described. It is therefore enough to show that / behaves properly with
respect to modified products.


Suppose then that / = (/1,f2,... ,/p<i>), V=(vl9v2,...9vp), J =</i>


<i>O'i»7*2» • • • Jq)</i> a n (i W = (wls w2<i>, . . . , wq) are sequences of integers. We put</i>


<i>IJ= (il9..., ip,jl9... Jq) and define VW similarly. We also put \J\=Ji +</i>


<i>J2 + '"+Jq</i> a n d<i> \V\=Vi+V2 + '-+Vp.</i>


<i>Let x = aiiđ--đaipđbjiđ---đbjq and y = ocViđ ã • • ®ocVp®</i>


<i>PWl®'-'® PWq belong "to A^ ®---® A\p) ®Bfi)®---® Bf and</i>



<i>A™ ®'"®A[p)p® B{^1 ®--®Biq)q</i> respectively. The modified product of


<i>x and y is</i>


<i>e(/J, VW)aticcVi đ ã ã ã đ aipocVp </i> <i>q</i> <i>q</i>


and the image of this under / is
<i>e(IJ,VW)(aiixVi®---®aip(xVp)</i>


<i>®(bjxvi®'"®bjqpwq). (3.4.9)</i>


Thus the theorem will follow if we can show that (3.4.9) is the
<i>p r o d u c t o f (ati ®-- ®aip)®(bji ®-- ®bjq) a n d (a</i>y<i>.(x) ã ã ã đ<xVp)đ</i>


<i>(PWl đ- " đPWq) in (3.4.8). However, this latter product is</i>


<i>where { is the product of ati đ- ã ã đ a{ and ocVi đ • • • ® ocVp</i> in


<i>A{1) ® A{2) ® • - - ® A{p\ and n is the product of bh đ - ã ã đ bjq</i> and


<i>PWlđ" ' đ Pw in B{1) ® B(2) ®" ' ® B{q). Accordingly</i>


is just


<i>Vx®-- '®aipav)</i>


and therefore the proof will be complete if we show that
<i>e(/J, VW) = (-l)lJme(I, VW, W).</i>



</div>
<span class='text_page_counter'>(65)</span><div class='page_container' data-page=65>

<i>54 Associative algebras</i>


We make one final observation concerning modified products. Suppose
that


<i>ft:A®^B® (* = l , 2 , . . . , p )</i>


are homomorphisms of graded algebras. Then of course /x<i> đ f2 đ ã ã * đ fp</i>


is also a homomorphism


of graded algebras. But modified and unmodified tensor products have the
<i>same ^-module structure. Hence the same mapping fx ® f2 đ * ã ã đ fp</i> is a


homomorphism


<i>đRB{2) đR'" ®RBip)</i> (3.4.10)


<i>of K-modules. This certainly preserves degrees and identity elements. We</i>
<i>claim that in fact (3.4.10) is a homomorphism of graded algebras. To see this</i>
<i>let x and y belong to A(1) ® A(2) ® • • • ® Aip). We have to show that the</i>
image of their modified product is the modified product of their images. For
<i>this we may assume that x = aix đ at2 đ ã ã ã đ at and y = aji đ ocJ2</i> đ ã ã •


<i>® (Xj belong to A^ ® • • • ® A\p) and Af^ đ ã ã ã đ A^p)</i> respectively, where
we have made use of our usual notation. However, in this case the desired
result follows from the formula for the modified product (see (3.4.5)).


<b>3.5 Anticommutative algebras</b>



<i>Let {An} neZ be a grading on an K-algebra A. We say that the graded</i>


<i>algebra is anticommutative provided</i>


anam = ( - i P amaB (3.5.1)


<i>whenever am</i> and an<i> are homogeneous elements of degrees m and n</i>


respectively, and provided also that


<i>a2m = 0 (3.5.2)</i>


<i>whenever the degree of am</i> is odd. Thus we see that in an anticommutative


algebra the square of a homogeneous element of degree one is zero. As the
next lemma shows there is an important special situation where the
converse holds.


<i><b>Lemma 3. Let the graded R-algebra A be generated (as an R-algebra) by its</b></i>


<i>module AY of homogeneous elements of degree one. If now a</i>2<i> = 0 for every</i>


<i>a^eA^ then A is anticommutative.</i>


<i>Proof From Lemma 1 we see that An = 0 for all n<0 and also that Ao</i> is


</div>
<span class='text_page_counter'>(66)</span><div class='page_container' data-page=66>

<i>Anticommutative algebras 55</i>


<i>= 0 and y2 = 0 it follows that</i>



<i>0. (3.5.3)</i>


<i>Now suppose that u = xxx2</i> . . . xm<i>, ^ = ^1^2 • • • >*> where every xt</i> and ^


<i>belongs to Av</i> Then, by (3.5.3),


<i>vu = yly2...ynx1x2...xm</i>


= ( - i r x ! X2 . . . xmy ^2<i> . . . yn = (- l)mnuv.</i>


<i>But every element of Am (respectively An) is a sum of elements such as u</i>


<i>(respectively v) so the formula (3.5.1) holds quite generally.</i>


<i>From here on we suppose that m> 1 and is odd. Let am belong to Am.</i>


<i>Then am = ux + u2 + * • • + us, where each ut is a product of m elements of A x</i>,


and w? = 0 as is clear from (3.5.3) and the second hypothesis of the lemma.
Again, by what has just been proved,


<i>ujui = (-l)m2uiUj=-uiuj</i>


and now it follows that a^ = 0. Thus the lemma is proved.


<i><b>Theorem 7. The modified tensor product of a finite number of </b></i>


<i>anticommuta-tive R-algebras is itself an anticommutaanticommuta-tive R-algebra.</i>


<i>Proof Let A and B be anticommutative .R-algebras. If we can show that</i>


<i>A ® B is anticommutative, then the general result will follow by virtue of</i>
Theorem 6.


<i>Let x = ar®bs,y = ocp®Pa, where ar, ap are homogeneous elements of A</i>


<i>and bs, fia are homogeneous elements of B, and where the degree of each</i>


<i>element is indicated by its suffix. If x * y denotes the modified product of x</i>
<i>and y, then</i>


On the other hand


so that


<i>But in A ® B the element x has degree r + 5 = m say whereas y has degree</i>
<i>p + o = n say. Consequently y * x = (— \)mnx * y.</i>


<i>Now suppose that m is odd. Then one of r and s is odd and therefore the</i>
<i>modified square of x namely</i>


<i>{-\)sr + r2+s2a2r®b2s</i>


is zero.


</div>
<span class='text_page_counter'>(67)</span><div class='page_container' data-page=67>

<i>56 Associative algebras</i>


<b>3.6 Covariant extension of an algebra</b>


The process of covariant extension as applied to modules was
explained in Section (2.5). Here we shall examine it in relation to algebras.


<i>Let R, 5 be commutative rings and assume that we are given a </i>
ring-homomorphism


<i>a):R-+S (3.6.1)</i>


<i>that preserves identity elements. Assume further that A is an R-algebra.</i>
<i>From Section (2.5) we know that A ® R</i> 5 is an 5-module. On the other hand


<i>5 is an R-algebra (with w as its structural homomorphism) so that A ®R</i> 5 is


an K-algebra and in particular it is a ring. If we examine the 5-module
<i>structure and the ring structure on A ®R</i> 5 we find that they interact in such


<i>a way that we have an 5-algebra. Thus the covariant extension of the </i>
<i>R-algebra A is the S-R-algebra A ®R</i> 5. Note that the structural homomorphism


<i>5ã A đRS maps s into lA đs.</i>


<i>Now suppose that {An}nel is a grading on A. By Theorem 5 of Chapter 2,</i>


<i>A®RS=YJ(An®RS) (As.). (3.6.2)</i>


<i>nel</i>


Of course according to the theorem just quoted the right-hand side is to be
<i>understood as a direct sum of ^-modules. However, A ®R 5 and An ®R</i> 5


are 5-modules so that (3.6.2) is also a direct sum of 5-modules. Moreover the
<i>product of x in Am®RS and y in An®RS belongs to Am+n®RS.</i>



<i>Accordingly the 5-algebra A ®RS is graded by {An ®RS}neZ.</i>


<i>Let us turn our attention to the tensor product, over R, of R-algebras</i>
<i>Al9A2,..., Ap. (For the moment these algebras need not be graded.) By</i>


Theorem 9 of Chapter 2, there is an isomorphism
<i>(Ax ®R 5) ®s {A2 ®RS)®S--' ®s (Ap ®R</i> 5)


<i>*{Al®RA2®R'-®R Ap) ®R</i> 5 (3.6.3)


<i>of 5-modules in which (ax ® s</i>t<i>) ® (a2 ® s2) ® • • * ® (ap ® sp) corresponds</i>


<i>to (al ® a2 đ ã ã * đ ap) đ s1s2 . . . sp. This isomorphism is compatible with</i>


multiplication and therefore (3.6.3) is actually an isomorphism of
<i>5-algebras. Thus, for algebras, covariant extension commutes with the</i>
<i>formation of tensor products.</i>


<i>Finally suppose that each of the K-algebras Al9 A2, . . . , Ap</i> is graded.


Then the two sides of (3.6.3) are graded 5-algebras and it is easy to see that
<i>the isomorphism preserves degrees. Hence when Al9A2,. • . , Ap are graded</i>


<i>R-algebras (3.6.3) is an isomorphism of graded S-algebras.</i>


<b>3.7 Derivations and skew derivations</b>


</div>
<span class='text_page_counter'>(68)</span><div class='page_container' data-page=68>

<i>Derivations and skew derivations 57</i>


<i>A mapping h: A —• A is said to be of degree k if it raises the degree of each</i>


homogeneous element by this amount.


<i>By a derivation on A we shall understand an #-homomorphism</i>
<i>D: A-+A, of degree — 1, which satisfies</i>


<i>D(aP) = (Doc)P + (x(DI3) (3.7.1)</i>


<i>for all a, j8 in A. Note that if D is a derivation, then</i>


<i><b>n</b></i>


D(axa2 . . . a j = £ ax. . . ai_1(Z)ai)ai + 1. . . an. (3.7.2)


<b>i = l</b>


<i>Suppose that £>, D' are derivations on A and that reR.lt is easily checked</i>
<i>that the endomorphisms D + D' and rD of A are themselves derivations. In</i>
<i>fact the derivations on A form an tf-module. Of course D°Df</i> is an
<i>endomorphism of A9</i> but (usually) it is not a derivation.


<i><b>Lemma 4. Suppose that the graded algebra A is generated (as an </b></i>


<i>R-algebra) by its homogeneous elements of degree one, and let D, D' be</i>
<i>derivations on A. Then D' ° D = D° D'.If D and D' agree on all homogeneous</i>
<i>elements of degree one, then D = D'.</i>


<i>Proof Let a</i>1? a2, . . . , an<i>, where n>2, belong to Av Then D<xt and D'OL{</i>


<i>belong to Ao and are therefore in the centre of A. Also (Df ° D)OL{</i> = 0 because



<i>As = 0 for all s < 0 . Next, from (3.7.2) we have</i>


<i>(Ly°D)(<x1ai2...aH)</i>


7) + D(aj)D'(al.))a1... af. . . a , . . . aw,


where the symbol A indicates that the term over which it is placed is to be
omitted. It follows that


<b>... aJ</b>



<i>so that Dr°D = D°D'as stated. The final assertion of the lemma also follows</i>
from (3.7.2).


Derivations play an important role in the theory of commutative
algebras. For anticommutative algebras the concept needs to be modified.
<i>By a skew derivation on A we shall mean an .R-homomorphism A: A —> A,</i>
of degree — 1, which satisfies


A(a/J) = (Aa)j8 + ( - l)ma(A£) (3.7.3)
<i>for all a e Am and ft e A. The skew derivations on A also form an R-module.</i>


<i><b>Lemma 5. Let A be a skew derivation on the graded R-algebra A and let o^,</b></i>
a2<i>, . . . , 0Ln be homogeneous elements of degree one. Then</i>


</div>
<span class='text_page_counter'>(69)</span><div class='page_container' data-page=69>

<i>58 Associative algebras</i>


<i>Proof. Since</i>


<b>. an) = (A(a1a2 ... a ^ ^ K + t - l)"</b>



the lemma follows by induction.


<i><b>Lemma 6. Let the graded R-algebra A be generated (as an R-algebra) by its</b></i>


<i>homogeneous elements of degree one, and let A, A' be skew derivations on A.</i>
<i>Then A ° A = 0 and A' ° A + A ° A' = 0. / / A and A' coincide on the module of</i>
<i>homogeneous elements of degree one, then A = A'.</i>


<i>Proof Let a</i>l5 a2, . . . , an be homogeneous elements of degree one. Then Aa^


<i>belongs to the centre of A. Two applications of (3.7.4) now show that</i>
(A ° A)(axa2 . . . an) = 0 whence A ° A = 0. This proves the first assertion and


the second follows by expanding (A + A') ° (A + A') = 0. The final assertion is
clear from (3.7.4).


Derivations and skew derivations both behave well in relation to the
<i>process of covariant extension. To be precise let co: R—+S be a </i>
<i>homo-morphism of commutative rings and let D (respectively A) be a derivation</i>
<i>(respectively skew derivation) on a graded K-algebra A. Then D ®S</i>
<i>(respectively A®S) is a derivation (respectively skew derivation) on the</i>
<i>graded S-algebra A ®R S.</i>


<b>3.8 Comments and exercises</b>


Perhaps the first comment to be made concerning the notion of an
algebra is that it subsumes more than appears at first sight. Thus if Q is any
<i>ring with an identity element, then Q is an algebra over the integers. Here</i>
<i>the structural homomorphism Z—>Q maps the integer n into nln.</i>



Additional comments, supplemented by exercises, are to be found below,
and, as in Chapters 1 and 2, the exercises for which solutions are provided
are marked with an asterisk.


<i>We begin by noting that if Xl9 X2,..., Xn</i> are indeterminates, then the


polynomial ring K[Xl 9<i> X2,..., Xn~] is a commutative .R-algebra.</i>


<i><b>Exercise 1. Show that an R-algebra A can be generated (as an algebra) by a</b></i>


<i>single element if and only if it is a homomorphic image of the algebra R[X~],</i>
<i>where X is an indeterminate.</i>


Observe that this exercise shows that an algebra which can be generated
by a single element is necessarily commutative.


</div>
<span class='text_page_counter'>(70)</span><div class='page_container' data-page=70>

<i>Comments and exercises 59</i>


This said, let Sx and E2<i> be two multiplicatively closed subsets of R. Then</i>


S f1^ ) and Z^"1^) are ^-algebras so we can form their tensor product


<i>^HR)®RZ;HR)</i> (3-8.1)


and this too will be a commutative R-algebra. On the other hand, the set of
all products<i> CXG2, where (T1enL1</i> and <r2eZ2, is again a multiplicatively


<i>closed subset of R. Let this new multiplicatively closed subset be denoted by</i>
I1Z2 a nd let us form the K-algebra



<i>(L^r^n (3.8.2)</i>



The next exercise shows that (3.8.1) and (3.8.2) are virtually identical.


<i><b>Exercise 2*. Let Z</b></i><b>x</b><i><b> and Z</b></i><b>2</b><i> be multiplicatively closed subsets of R and put</i>
Z = Z!Z2<i>. Establish an isomorphism</i>


<b>sr'w®*^</b>

<b>1</b>

<b>^)^"</b>

<b>1</b>

<b>^)</b>



<i>of R-algebras.</i>


The existence of a grading on an R-algebra has some interesting
<i>consequences. Let A be a graded K-algebra and let {An}neZ</i> be its grading.


<i>First we make the rather trivial observation that if a belongs to the centre of</i>
<i>A, then all its homogeneous components belong to the centre as well. To see</i>
this let


a = - - - + a _1+ a0 + a1+ a2 +


---be the decomposition of a into its homogeneous components, where the
<i>suffix indicates the degree of an individual component. If now am<b> e A</b>m, then</i>


<i>from otam = am<x we obtain <xnam = am(xn</i> by comparing components of degree


<i>m + n. But now it follows that (xnx = x(xn for all xeA, i.e. a</i>w belongs to the


<i>centre of A.</i>



<i><b>Exercise 3*. Let A be a non-negatively graded R-algebra and let a belong to</b></i>


<i>the centre of A. If OL2 = a show that a is homogeneous of degree zero.</i>
<i>Of great importance is the notion of the free algebra generated by a given</i>
<i>set. To explain what this is let X be an arbitrary non-empty set. We consider</i>
<i>formal products x^x2 . . . xn (n >0) of elements of A' or, to use an alternative</i>


<i>terminology, words formed by using the elements of X as letters. (Note that,</i>
<i>because the possibility n = 0 is allowed, we have included the empty word.)</i>
Next we form the free .R-module which has the set of words as a base, and
afterwards turn the free module into an .R-algebra by defining the product
of two of its elements by means of the self-explanatory formula


</div>
<span class='text_page_counter'>(71)</span><div class='page_container' data-page=71>

<i>60 Associative algebras</i>


<i>Let A be this algebra. For n > 0 let An</i> be the K-submodule generated by


<i>all words with n letters, and put Aq = 0 when q<0. It is easy to check that</i>


<i>{As}seI is a non-negative grading on A. Note that A0 = R, that Ax</i> is the free


<i>R-module generated by X, and that this module generates the algebra. In</i>
the next chapter where with each module we shall associate an algebra
<i>called its tensor algebra, we shall see that A is the tensor algebra of the free</i>
<i>module generated by X.</i>


<i>Another interesting example arises in the following way. Let M be an </i>
<i>R-module and this time consider the set A formed by all 2 x 2 'matrices'</i>


<i>where reR and meM. If addition and multiplication on A are now</i>


defined by


and


<i>[</i>

<i>r m~\ Yr' m'~\ Yr + r' m + m'~|</i>
<i>0 r \ 0 r' 0 r + r'</i>


<i>r m l f r ' rri~\ Yrr' rm' + r'm~\</i>

<b>0 r l o r'J</b>

<b> =</b>

<b> L0</b>

<i><b><sub>rr' J</sub></b></i>



<i>respectively, then it is easily checked that A is a commutative ring with</i>


<i>as its identity element. Furthermore, if for peR we put</i>


<i>[</i>

<i>r m l Ypr pm~\</i>


<b>0 rj</b>

<b> =</b>

<i><b> [o pr\</b></i>



<i>then A is turned into a commutative K-algebra.</i>
<i>Now let Ao</i> consist of all matrices of the form


<i>let Al</i> consist of all matrices


<i><b>0 ml</b></i>


<b>0 Oj'</b>



<i>and for n ^ 0 , 1 put An = 0. Then {Aq}qeZ is a family of R-submodules of A.</i>


<i><b>Exercise 4. With the above notation, show that {A</b>q</i>}<i> q€l is an algebra-grading</i>



<i>for the R-algebra formed by all matrices of the form</i>


<b>[;:]•</b>



</div>
<span class='text_page_counter'>(72)</span><div class='page_container' data-page=72>

<i>Comments and exercises 61</i>


<i>Let A(l\ A{2\..., Aip)</i> be graded K-algebras. We have discussed two
graded algebras that can be obtained from them by using tensor products.
<i>These algebras are the ordinary tensor product A{1) đ A{2) đ ã ã ã ® Aip)</i>
<i>and the modified tensor product A{1) ® A{2) ® • • • ® Aip\ The latter will</i>
<i>also be known as the twisted tensor product.</i>


In establishing (3.3.13) we showed that the operation of combining
<i>graded algebras by means of the ordinary tensor product is associative and</i>
the discussion of (3.3.14) shows that it is commutative as well. Furthermore,
<i>from (3.3.15) and (3.3.16), we see that when R is endowed with the trivial</i>
grading it acts like an identity in this context.


For modified, that is twisted, tensor products the facts are not so
transparent. We have indeed shown (Theorem 6) that the associative law
continues to be satisfied, but so far we have not examined whether modified
tensor multiplication is commutative. This will be our next concern.


<i>Let A and B be graded ^-algebras and {An}neI and {Bn}n€l</i> their


<i>respective gradings. There is an isomorphism Am ® Bn % Bn ® Am</i> in which


<i>am ® bn is matched with bn ® am. (Here, of course, am e Am and bn e Bn.) It</i>



follows that there is an isomorphism


<i>Tmn:Am®Bn^-+Bn®Am</i> (3.8.3)


in which


<i>Tmn(am®bn) = (-\rnbn®am. (3.8.4)</i>


<i>But A®B = Y,Am®Bn (d.s.) and B®A = Y,Bn®Am</i> (d.s.) so we can


<i>combine the various Tmn</i> to obtain an isomorphism


<i>T:A®B-^-+B®A (3.8.5)</i>


of K-modules which is such that


<i>T(am®bn) = (-irnbn®am</i> (3.8.6)


<i>whenever am e Am and bneBn.T is known as the twisting isomorphism. Now</i>


<i>A ® B and A ® B are identical as ^-modules and the same is true of B ® A</i>
<i>and B ® A. Consequently T provides a bijection</i>


<i>T:A®B-+B®A. (3.8.7)</i>


<i>Exercise 5*. Let A and B be graded R-algebras. Show that the twisting</i>
<i>isomorphism T:A®B^+B®A produces an isomorphism</i>


<i>T:A®B - ^</i>



</div>
<span class='text_page_counter'>(73)</span><div class='page_container' data-page=73>

<i>62 Associative algebras</i>


The next exercise extends the result embodied in Exercise 5 to twisted
tensor products with an arbitrary number of factors.


<i><b>Exercise 6*. Let A</b>{1\ A(2\..., A{p) be graded R-algebras and let</i>
(s1? s2<i>, . . . , sp) be a permutation of ( 1 , 2 , . . . , p). Show that</i>


<i>A(l) ® A{2) ® •• • ® A{p)</i>
<i>and</i>


<i>A(Sl) Q A(s2) ® . . . g) A(sP)</i>


<i>are isomorphic graded algebras.</i>


<i>Let A be an R-algebra. The mapping A x A —> A in which (a, a) is taken</i>
<i>into aoc is certainly bilinear and so it induces an ^-linear mapping</i>
<i>fi: A đ A ã A in which n(a ® a) = aa. If A is a commutative /^-algebra, then</i>
<i>fi: A ®A—+Aisa homomorphism of algebras, and if A is a commutative</i>
<i>graded algebra, then \i is a homomorphism of graded algebras. The next</i>
exercise deals with the corresponding result for anticommutative algebras.
<i>Note that \i can also be regarded as an K-linear mapping of A ® A into A.</i>


<i><b>Exercise 7*. Let A be an anticommutative R-algebra. Show that the mapping</b></i>


<i>fi\ A ® A-^Ain which fi(a ® a) = ace is a homomorphism of graded algebras.</i>
Ordinary and modified tensor products of graded algebras have so many
features in common that the idea of finding an approach that will deal with
both types simultaneously has obvious attractions, and in fact something
can be done on these lines as will now be explained.



<i>Let A and B be graded /^-algebras with {An}neI and {Bn}neI</i> their


<i>respective gradings. Suppose that An = 0 whenever n is an odd integer. Then</i>


<i>A ® B and A ® B not only coincide as modules, but ordinary and modified</i>
<i>multiplication are the same. Consequently A ® B and A ® B are identical</i>
algebras and we may write


<i><b>A®B = A®B. (3.8.8)</b></i>


<i>Likewise (with the same assumption on A) we have</i>


<i>B®A=B®A. (3.8.9)</i>
<i>For example, (3.8.8) and (3.8.9) hold if the grading on A is trivial. In</i>
<i>particular, if R is endowed with the trivial grading, then, for any graded </i>
<i>K-algebra £, R®B = R®B and B ® R = B ® R and therefore we have</i>
isomorphisms


<i>R®B*B (3.8.10)</i>
and


<i>B®R*B (3.8.11)</i>


</div>
<span class='text_page_counter'>(74)</span><div class='page_container' data-page=74>

<i>Comments and exercises 63</i>


<i>Suppose now that C is a graded K-algebra and let Cn</i> denote its


<i>submodule of homogeneous elements of degree n. Put</i>
<i>fO if n is odd,</i>



<b>c:=</b>

<i><sub>Z</sub><sub>nj2</sub><sub> if n is even.</sub></i>


<i>Then C = J ] Q (d.s.) and indeed {C'n}neI</i> is an algebra-grading on C.


<i>Accordingly we have a new graded R-algebra, C say. The connection</i>
<i>between C and C is described by saying that C is obtained from C by</i>
<i>doubling degrees.</i>


<i><b>Exercise 8. Let C and D be graded R-algebras. Show that</b></i>


<i>C ®D' = (C® D)\</i>


<i>where C", D' and (C ® D)' are obtained from C, D and C ®D respectively by</i>
<i>doubling degrees.</i>


Let us retain, for the moment, the notation of Exercise 8. Then, by (3.8.8),
<i>C ® D' = C ® Df</i> and therefore


<i>C ®D' = {C®D)f<b>. (3.8.12)</b></i>


Now the operation of doubling degrees is easily reversed. Consequently
(3.8.12) provides a means whereby an ordinary tensor product can be
expressed in terms of a modified tensor product and this, on occasions,
offers some advantages.


<i>We next observe that the notions of derivation and skew derivation, as</i>
<i>defined in Section (3.7), can be considerably extended. For instance, let A be</i>
<i>a graded .R-algebra. Then by a generalized derivation of degree i we</i>
<i>understand an .R-linear mapping D: A—>A of degree i such that</i>



<i>for all a, oc in A. Thus the derivations considered in the main text are</i>
precisely the generalized derivations of degree — 1.


<i><b>Exercise 9. Let D and D' be generalized derivations, on the graded R-algebra</b></i>


<i>A, of degrees i and j respectively. Show that D' ° D — D° D' is a generalized</i>
<i>derivation of degree i +j.</i>


As is to be expected, generalizing the notion of a skew derivation is a
more complicated matter. It is helpful to make some preparatory
observations.


<i>Once again let {An</i>}<i> nsI be a grading on an K-algebra A. For each n e Z we</i>


<i>can define an automorphism Jn, of An, by putting Jn(x)= (— l)nx for all</i>


<i>xeAn; and then these automorphisms can be combined to give an</i>


automorphism


</div>
<span class='text_page_counter'>(75)</span><div class='page_container' data-page=75>

<i>64 Associative algebras</i>


<i>of the J?-module A which satisfies</i>


<i>J(an) = (-l)nan</i><b> (3.8.14)</b>


<i>for all an in An. Actually J is more than a module-automorphism as the next</i>


exercise shows.



<i><b>Exercise 10. Show that J, as defined in (3.8.13), is an automorphism of the </b></i>


<i>R-algebra A.</i>


<i>Indeed this is not the end of the matter. Not only is J an automorphism of</i>
<i>the algebra, but J2 is the identity automorphism, so that, in other words, J is</i>
<i>an involution of the graded algebra. It is known as the main involution of A.</i>
<i>Note that if i e Z, then J1 is the identity automorphism if i is even whereas it</i>
<i>is J itself if i is odd. Note too that if ameAm, then</i>


<i>Ji(am) = (-l)i*nam. (3.8.15)</i>


<i>After these preliminaries let A: A—-• A be an R-linear mapping of degree</i>
— 1. Then A is a skew derivation (as defined in Section (3.7)) if and only if


A(aa) = (Aa)a + J(a)A(a) (3.8.16)
<i>for all a, a in A.</i>


We are now ready to generalize the notion of a skew derivation.
<i>Specifically a mapping A: A —• A will be called a generalized skew derivation</i>
<i>of degree i if it is R-linear of degree i and</i>


A(aa) = (Aa)a + Jl(a)A(a) (3.8.17)
<i>for a, a in A. Consequently the skew derivations which occur in the main</i>
text are none other than the generalized skew derivations of degree — 1.
<i><b>Exercise 11*. Let A and A' be generalized skew derivations, of degrees i and]</b></i>


<i>respectively, on the graded R-algebra A. Show that</i>



A'°A-(-l)"A°A'


<i>is a generalized skew derivation of degree i +j. Show also that if the integer i is</i>
<i>odd, then A ° A is a generalized skew derivation of degree 2i.</i>


</div>
<span class='text_page_counter'>(76)</span><div class='page_container' data-page=76>

<i>Solutions to selected exercises 65</i>


<b>3.9 Solutions to selected exercises</b>


<i><b>Exercise 2. Let E</b></i><b>x</b><i><b> and E</b></i><b>2</b><i> be multiplicatively closed subsets of R and put</i>
Z = Z1Z2<i>. Establish an isomorphism</i>


<i>of R-algebras.</i>


<i>Solution. In what follows au a\ denote elements of E</i>x; <r2,<i> G'i elements of</i>


E2<i>; and r, r', p, p' elements of R.</i>


<i>If r/a1=r'/(T'1</i> in Z f1<i>^ ) and p/(J2 = p'laf2</i> in Z ^1^ ) , then it is easily


verified that
<i>rp r'p'</i>


in E "1^ ) . Hence there is a well-defined mapping


<i>in which (r/a1,p/a2) is mapped into rpJGxo2, and a straightforward check</i>


shows that the mapping is bilinear. Accordingly there is an K-linear
mapping



<i>where ^{r/a1 ® p/a2) =</i>


<i>We claim that <\> is an isomorphism of R-algebras. Indeed it is obvious that</i>
<i><j) is surjective and that it preserves identity elements. Next, if M is any </i>
<i>R-module, then we know that each element of M ®R</i> Z ^<i>l (R) can be written in</i>


the form m (g) (l/a2<i>), where meM. Consequently every element of</i>


E j "1^ ) (x^E^T1<i>^) can be expressed in the form (r/a^ (x) (p/a2). Let x =</i>


<i>(rfci) ® (p/(T2) and x' = (r'la\) ® (p'l&2). Then</i>


<i>(j){xx') = (t){{rr'lG1G\) ® {pp'l<J2o'2)}</i>


<i>= rr'pp'/(T1 G\O2O'2</i>
<i><b>= (rp/<J</b><b>1</b><b>(T</b><b>2</b><b>)(r'p'/<T'</b><b>1</b><b>(T'</b><b>2</b><b>)</b></i>


<i><b>= 4>(x)ct)(x')</b></i>



so that 0 is a homomorphism of .R-algebras.


<i>Finally let x = (r/(T1) ® (p/(T2) and suppose that </>(x) = 0. Then</i>


<i>rp/G1(T2 = 0 in E "1 (R) and therefore there exist G\ e E</i>x<i> and a'2<b> e E</b></i>2 such that


<i>(j'1G2rp = Q. But now</i>


<i>x = (G'lr/(Ti(j'1) ® (o'2plo2of2)</i>


<i>= (v\<r'2rplG1o\) ® (l/a2(J2)</i>



<i>= 0.</i>


</div>
<span class='text_page_counter'>(77)</span><div class='page_container' data-page=77>

<i>66 Associative algebras</i>


<i><b>Exercise 3. Let A be a non-negatively graded R-algebra and let a belong to</b></i>


<i>the centre of A. If a</i>2<i> = a show that a is homogeneous of degree zero.</i>


<i>Solution. Let a = a</i>0 + ax + a2 + * • *, where am is homogeneous of degree m,


and put jS = ax -f- a2 + * • • so that a = a0 + /?. Then a2) = a0 and both a0<i> and P</i>


<i>are in the centre of A. Now</i>


and from this we see that (1 — ao)j3 is equal to its own square. Accordingly


<i>((1 — (xo)P)m = (1 — (xo)p for all m > 1. But (1 — a</i>o)j8 is a sum of homogeneous


elements of strictly positive degrees and therefore we may conclude that
(l-ao<i>)j3 = 0, i.e. that p = (xop. Next</i>


<i>whence p2 = - p. It follows that pm = ( - l)</i>m + 1<i> p for all m > 1. However, this</i>
<i>implies that jS = 0 because p is a sum of homogeneous elements of strictly</i>
positive degrees. Thus a = a0 and the solution is complete.


<i><b>Exercise 5. Let A and B be graded R-algebras. Show that the twisting</b></i>


<i>isomorphism T: A®B^*B®A produces an isomorphism</i>



<i>T:A®B -^->B®A</i>


<i>of graded algebras.</i>


<i>Solution. Let x and y belong to A (g) B. Since T preserves degrees it will be</i>
<i>enough to show that T(x *y)= T(x) * T{y), where an asterisk has been used</i>
<i>to indicate a modified (as distinct from an ordinary) product. But T is an</i>
isomorphism of ^-modules. We may therefore confine our attention to the
<i>case where x = ar®bs</i> and y = ap(x)j8ff<i>; here ar9 ccp</i> denote homogeneous


<i>elements of A and bs, Pa homogeneous elements of B, the degrees of these</i>


<i>elements being indicated by their suffixes. On this understanding x*y =</i>
<i>( - l)sparocp ® bspa</i> and therefore


<i>p</i>


<i>On the other hand T(x) = ( - l)rsbs ® ar and T(y) = {- \)paPa ® ocp. </i>


Con-sequently


<i>= T(x * y)</i>


which is what we were seeking to prove.


</div>
<span class='text_page_counter'>(78)</span><div class='page_container' data-page=78>

<i>Solutions to selected exercises 67</i>


(3.9.1)
<i>and</i>



<i>A{Sl) đ A^ đ ãã ã đ A{SJ (3.9.2)</i>
<i>are isomorphic graded algebras.</i>


<i>Solution. Suppose that \<i<p. By repeated applications of Theorem 6</i>
together with the remarks made concerning (3.4.10), we obtain the
following isomorphisms of graded algebras:


<i>A® ® A{i + 1) ® Aii + 2) đ ã ã ã đ Aip)</i>


)} đ (A(i + 2)<i> đ ã ã ã đ Aip))</i>
đ (X(l + 2) đ ã ã ã


In a similar manner we obtain an isomorphism


<i>2)</i>


<i> đ ã ã ã đ Aip)</i>


<i>But we know from Exercise 5 that A{i) ® Aii + 1)</i> and v4(l + 1 )® A( 0 are
isomorphic graded algebras. It follows that the algebra (3.9.1) remains
unchanged up to an isomorphism of graded algebras if we interchange the
<i>adjacent factors A{i) and A(i + 1). However, the factors may be brought into</i>
whatever order we please by a succession of adjacent interchanges and with
this observation the solution is complete.


<i>The reader may like to refine this result by constructing an explicit</i>
isomorphism between the algebras (3.9.1) and (3.9.2).


<i><b>Exercise 7. Let A be an anticommutative R-algebra. Show that the mapping</b></i>



<i>fi: A đ A ã A in which \i{a ® a) = act is a homomorphism of graded algebras.</i>


<i>Solution. It is clear that fi is K-linear and that it preserves identities and</i>
<i>degrees. Let x and xf belong to A ® A and let x * xr</i> denote their product in
<i>A ® A. We need only prove that ii{x *x') = fi(x)fi(xf) and in doing this we</i>
<i>may suppose that x = ar</i> ® as<i>, x' = a'p ® a^, where ar, a'p, a</i>s<i>, ct!a</i> are


<i>homo-geneous elements of A whose degrees are indicated by their suffixes. Now</i>
<i>x * x' = (— \)spara'p ® (xsa'a</i> and therefore


<i>^(x * x') = ( - \)spara'pas(x'a = arasa'pcda</i>


<i>because A is anticommutative. Accordingly fx{x * x') = fi(x)jn(x') as required.</i>


<i><b>Exercise 11. Let A and A' be skew derivations, of degrees i andj respectively,</b></i>


</div>
<span class='text_page_counter'>(79)</span><div class='page_container' data-page=79>

<i>68 Associative algebras</i>


<i>is a generalized skew derivation of degree i +7. Show also that if the integer i is</i>
<i>odd, then A° A is a generalized skew derivation of degree 2i.</i>


<i>Solution. It is clear that A'°A and A ° A' are endomorphisms of degree i +7</i>
and therefore the same is true of A' ° A — (— 1)°A ° A'. Now let a, a belong to
<i>A. Then A(aa)=(Aa)a + J</i>J(a)(Aa) and therefore


<i>A'(A(aa)) = (A'(Afl))a + Jj(Aa) (A'a)</i>


Similarly


<i>A(A'(aa)) = (A(A'fl))a + f {A'a) (Aa)</i>



<i>+ A(J'"(a)) (A'a) + J* (Jj(a)) (A(A'a)).</i>


<i>But Jj(Ji(a)) = Ji(Jj(a)) = Ji+j(a) and therefore it will follow that A ' ° A </i>
<i>-(— 1)°A ° A' is a skew derivation of degree i +7 provided we show that</i>


<i>Jj(Aa) = (-l)ijA(Jj(a)) (3.9.3)</i>
and


<i>Ji(A'a)=(-l)ijA'(Ji(a)). (3.9.4))</i>


Indeed it is enough to prove the first of these relations for then the second
will follow by symmetry; and furthermore it will suffice to prove (3.9.3)
<i>when a is a homogeneous element.</i>


<i>Suppose therefore that aeAm. Then</i>


which is what we were aiming to prove.


<i>From here on we shall suppose that i is an odd integer and that a, a are</i>
<i>general elements of A. Then</i>


and so we shall have proved that A ° A is a skew derivation of degree 2i if we
show that


</div>
<span class='text_page_counter'>(80)</span><div class='page_container' data-page=80>

<b>4</b>



The tensor algebra of a module



<b>General remarks</b>



In Section (1.3) we defined the tensor powers of an /^-module M. In
this chapter we shall show how the different tensor powers can be fitted
together to produce an K-algebra. This algebra, which is known as the
<i>tensor algebra of M, contains M as a submodule. In fact the tensor algebra</i>
solves a certain universal problem concerned with algebras that contain a
homomorphic image of M. The tensor algebra is particularly important in
our context because it has, as homomorphic images, the exterior and
symmetric algebras which we shall study later.


<i>As in previous chapters R and S always denote commutative rings. All</i>
<i>algebras are understood to be associative, rings and algebras are assumed to</i>
have identity elements, and homomorphisms of rings and algebras are
required to preserve identity elements. Finally, we shall often omit the suffix
from the tensor symbol (x) when the underlying ring can easily be inferred
from the context.


<b>4.1 The tensor algebra</b>


<i>Let M be an ^-module. We shall define its tensor algebra in terms</i>
<i>of a certain universal problem. To this end suppose that A is an K-algebra</i>
<i>and that </>: M —• A is a homomorphism of K-modules. If now h: A —• B is a</i>
<i>homomorphism of K-algebras, then, of course, /z°c/>:M—•£ is an </i>
<i>R-module homomorphism of M into B. This observation leads us to pose the</i>
following universal problem.


<i><b>Problem. To choose A and (f> so that given any R-module homomorphism</b></i>


<i>\j/:M—+B {B is an R-algebra) there shall exist a unique homomorphism</i>
<i>h: A^»B, of R-algebras, such that h°(f) = il/.</i>



<i>It is clear that if (A, (/>) and (A\ </>') both solve this problem, then there will</i>


</div>
<span class='text_page_counter'>(81)</span><div class='page_container' data-page=81>

<i>70 The tensor algebra of a module</i>


<i>be inverse algebra-isomorphisms k. A —>A' and k'\A'—>A such that</i>
<i>X o (f) = (j)' and Xf °cf)f = (j). Thus the problem has (essentially) at most one</i>
solution.


<i><b>Theorem 1. The universal problem possesses a solution. If (A,(f)) is any</b></i>


<i>solution, then</i>


<i>(i) <f>\M—*A is an injection;</i>


<i>(ii) A is generated, as an R-algebra, by (j){M);</i>
<i>(iii) there is a grading {An}neI on A with Ax =4>(M);</i>


<i>(iv) for p>l, Ap is the p-th tensor power of M with ml đ m2</i> đ ã ã ã


<i>(v) the structural homomorphism R—>A maps R isomorphically onto</i>
<i>Ao.</i>


<i>Remarks. Before starting the proof it may be helpful if certain points are</i>
clarified. For instance it should be noted that, once (ii) has been established,
<i>Lemma 1 of Chapter 3 ensures that there is at most one grading on A that</i>
has the property described in (iii). To amplify (iv) we observe that the
<i>mapping of the p-fold product MxMx- - xM into Ap</i> which takes


(m1,m2,... ,mp<i>) into <t>(m1)<l)(m2).. • (j>(mp) is multilinear. The theorem</i>



<i>asserts that Ap</i> and this mapping together solve the universal problem for


multilinear mappings of M x M x • • • x M (see Problem 1 of Chapter 1).


<i>Proof. It suffices to exhibit one solution to the problem which satisfies</i>
<i>(i)-(v). To this end, for p > 1, put Ap = M đM đ ã • • ® M where there are p</i>


<i>factors. Thus Ax =M. We also put A0 = R and define an K-module A by</i>


<i><b>ZP</b></i>


<i>Our immediate concern will be to turn A into an R-algebra.</i>


<i>For p > 1 and q > 1, Theorem 1 of Chapter 2 provides an isomorphism</i>
<i>Ap(x) AqzzAp+q. Next, from Theorem 3 of the same chapter we obtain</i>


<i>isomorphisms Ao ® Aq&Aq and Ap® A0&Ap. (Note that when p = q = 0</i>


<i>the latter isomorphisms coincide.) Thus for every p > 0 and q > 0 we have an</i>
<i>explicit isomorphism Ap ® Aq % Ap+q</i> and hence there is a bilinear mapping


<i>Vpq:ApxAq-+Ap+q</i> (4.1.1)


<i>in which iipq(xp, yq) is the image of xp ® yq</i> under the isomorphism in


<i>question. Accordingly if m1,m2,..., mp and m\, m'</i>2<i>,..., m'q</i> belong to M,


then



<i>p®m'l®---®mq. (4.1.2)</i>


</div>
<span class='text_page_counter'>(82)</span><div class='page_container' data-page=82>

<i>The tensor algebra 71</i>


<i>yq) = ryq</i> (4.1.3)


and


<i>^po(xp,r) = rxp, (4.1.4)</i>


<i>where now xpeAp, yqeAq and p>0, q>0.</i>


<i>Suppose next that p>0, g > 0 , t>0 and let xp, yq, zt belong to Ap, Aq, At</i>


respectively. Using (4.1.2), (4.1.3) and (4.1.4) it is a straightforward matter to
verify that


<i><b>Hp+qAftpqiXp* y<)></b><b> z</b><b>t) = f</b><b>i</b><b>p,</b><b>q</b><b> + t(x</b><b>P</b><b>, t*</b><b>qt</b><b>(y</b><b>P</b><b>></b><b> z</b><b>t))- (4.1-5)</b></i>
<i>We are now ready to define multiplication on A. Let x,yeA. These</i>
<i>elements have unique representations in the form x = x0</i> + xx<i> + x2</i> + • • • and


<i>y = yo + y1+y2 + " ', where xn, yn belong to An</i> and the sums are essentially


<i>finite. The required product of x and y is then fi(x, y), where</i>


<i>Kx,y)= Z »pq(xP,yq).</i>


<i>p>0,q>0</i>


It is readily checked that multiplication is distributive (on both sides) with


respect to addition, and from (4.1.5) it follows that it is associative as well.
<i>Next IK belongs to Ao and fi(lR, y) = y, [i(x, lR) = x by virtue of (4.1.3) and</i>


<i>(4.1.4). Finally if reR, then</i>


<i>fi(rx9 y) = rfi(x, y) = /i(x, ry).</i>


<i>These observations taken together show that A is an (associative) </i>
<i>R-algebra, and that 1^, considered as an element of Ao, is its identity element.</i>


<i>For n < 0 p u t An = 0. Then {An}nel is a grading on A, and, since M = Al9</i>


<i>M is a submodule of A. Let (/>: M —• A be the inclusion mapping. Evidently</i>
<i>A and (/> satisfy conditions (i)-(v). The proof will therefore be complete if we</i>
<i>show that A and 4> solve the problem which was enunciated at the</i>
beginning of the section.


<i>Assume then that \I/:M—+B is an K-linear mapping of M into an</i>
<i>^-algebra B. For p>\ there is a multilinear mapping of the p-fold</i>
<i>product MxMx-'xM into B that takes (m</i>1,m2,... ,mp) into


<i>^ ( m i ) ^ ^ ) . . . i//(mp), and this will induce a homomorphism hp: Ap—+B,</i>


where


<i>hp{m1 ®m2 ®''' ®m</i>p) = ^ ( m1) ^ ( m2<i>) . . . i//(mp).</i>


<i>Since A0 = R, we may take h0: Ao—>B to be the structural homomorphism</i>


<i>R—>B. But A is the direct sum of all the An with n > 0. Consequently there is</i>



<i>an K-homomorphism h: A-^B which agrees with hn on An. Certainly h</i>


</div>
<span class='text_page_counter'>(83)</span><div class='page_container' data-page=83>

<i>72 The tensor algebra of a module</i>


<i>Finally for XGM = AX we have h(x) = \j/(x) and therefore h°(ft = \l/. But</i>


<i>M = Ai generates A as an R-algebra and so there can only be one </i>
<i>algebra-homomorphism (of A into B) which when combined with (ft gives \ft.</i>
Consequently the proof is complete.


These conclusions will now be cast into a different form. Suppose that
<i>(A, (ft) solves our problem. Put</i>


<i>T(M) = A (4.1.6)</i>


<i>and denote by {Tn(M)}neZ the grading on T(M) that results from Theorem</i>


<i>1. Since (ft maps M isomorphically onto Tt</i> (M) we can use this isomorphism


to identify the two modules. Thus


<i>TX(M) = M (4.1.7)</i>


<i>so that T(M) contains M as a submodule. Indeed by this device the notation</i>
<i>is greatly simplified. We know that, as an K-algebra, T(M) is generated by</i>
<i>Tl(M) = M so that the grading is non-negative. Also the structural</i>


<i>homomorphism maps R isomorphically onto T0(M). Hence, when it suits</i>



<i>us, we can identify T0(M) with R. Finally, if ® is used to denote</i>


<i>multiplication on T(M), then, for p > 1, Tp(M) is the p-th tensor power of M</i>


<i>as defined in Section (1.3). It is because of this last property that T(M) is</i>
<i>called the tensor algebra of M.</i>


Our next theorem merely restates, but now in the new terminology, that
the tensor algebra solves the universal problem with which we started.


<i><b>Theorem 2. Let T(M) be the tensor algebra of the R-module M, and let</b></i>


<i>\//: M—> B be an R-linear mapping of M into an R-algebra B. Then \\i has a</i>
<i>unique extension to an algebra-homomorphism of T(M) intoB.</i>


<b>4.2 Functorial properties</b>


Let /:M—>iV be a homomorphism of K-modules and let
<i>T(M), T(N) be the tensor algebras of M and N respectively. Since N is</i>
<i>a submodule of T(N), Theorem 2 shows that / has a unique extension</i>
<i>to a homomorphism T(f):T(M)^T(N) of K-algebras. Note that if</i>
<i>ml, m</i>2<i>, . . . , mp</i> belong to M, then


<i>TifUnti đ m2 đ ã ã ã đ mp) = f(m1)đf(m2) đ ã ã ã đf(mp). (4.2.1)</i>


<i>Clearly if N = M and / is the identity mapping of M, then T(f) is the</i>
<i>identity mapping of T(M).</i>


We next observe, as a consequence of (4.2.1), that a homogeneous
<i>element of degree p > 1 keeps its degree when T(f) is applied. Moreover</i>


<i>T(f) maps T0(M) isomorphically onto T0(N). Consequently T(f) is</i>


</div>
<span class='text_page_counter'>(84)</span><div class='page_container' data-page=84>

<i>Functorial properties 73</i>


<i>Ts(f):Ts(M)^Ts(N) (4.2.2)</i>


<i>of JR-modules. In fact, as is shown by (4.2.1), when p> 1 we have</i>


<i>Tp(f) = f®f®'--®f (p factors) (4.2.3)</i>


and to this we may add the observation that To<i>(/): T0(M)-+ T0(N) is the</i>


<i>identity mapping when we make the identifications T0(M) = R and</i>
<i><b>T</b><b>0</b><b>(N) = R.</b></i>


<i>Next suppose that we have #-homomorphisms f:M—>N and</i>
<i>g:K-^M. Evidently</i>


<i>T(f)oT(g)=T(fog) (4.2.4)</i>


<i>and hence, for each s e Z,</i>


<i><b>T</b><b>s</b><b>(f)°T</b><b>s</b><b>(g) = T</b><b>s</b><b>(f°g). (4.2.5)</b></i>


<i>Thus in the language of the Theory of Categories T(M) is a covariant</i>
<i>functor from /^-modules to graded K-algebras. Note that if / : M —• N is an</i>
<i>isomorphism, then T(f) is also an isomorphism and T ( /</i>- 1) is its inverse.
<i>This is because T{f~l)°T{f) and T{f)°T{f~l) are identity mappings.</i>


<i>Now let K be a submodule of the K-module M. Since T1(M) = M, K</i>



<i>consists of homogeneous elements of T(M) of degree one. Accordingly K</i>
<i>generates a homogeneous two-sided ideal of T(M).</i>


<i><b>Theorem 3. Let f\M^>N bea surjective homomorphism of R-modules and</b></i>


<i>let K be its kernel. Then T(f): T(M)-+T(N) is surjective and its kernel is the</i>
<i>two-sided ideal that K generates in T(M).</i>


<i>Proof. It is clear from (4.2.1) that T(f) is surjective. Let I(K) be the </i>
<i>two-sided ideal generated by K and put Ip(K) = I(K) n Tp(M). Then, for p > 1,</i>


<i>Ip(K) is the submodule of Tp(M) generated by all products mx</i> (x)


m2<i> đ ã * ã ® mp, where mi</i>, m2<i>, . . . , mp belong to M and at least one m{ is in K.</i>


<i>Since Tp(f) = f đ / đ ã ã • ® / , it follows (Chapter 2, Theorem 4) that Ip(K)</i>


<i>is the kernel of Tp(f). Consequently the kernel of T(f) is</i>


as required.


<i>We shall now give a second proof that the kernel of T(f) is I(K). This</i>
alternative proof has the advantage that it can be adapted to deal with
similar situations that arise in the study of other algebras.


<i>Let h: T(M)^T(M)/I(K) be the natural homomorphism of graded</i>
<i>algebras. Since T1(M) = M and 7</i>1<i>(K) = X, the submodule of T(M)/I{K)</i>


</div>
<span class='text_page_counter'>(85)</span><div class='page_container' data-page=85>

<i>74 The tensor algebra of a module</i>



<i>homomorphism N-+T(M)/I(K) and this, by Theorem 2, extends to an</i>
<i>homomorphism T(N)-+ T(M)/I(K). Let us combine the </i>
<i>algebra-homomorphism T(N)^> T(M)/I(K) with T(f): T(M)-> T(N). The result</i>
<i>is an algebra-homomorphism T(M)-+ T(M)/I(K) which maps m of 7i(M)</i>
<i>into h(m). But T</i>x<i> (M) generates T(M) as an R-algebra, so in combining the</i>


<i>two homomorphisms we have recovered h: T(M)—+ T(M)/I(K). It follows</i>
<i>that the kernel of T(f) is contained in the kernel of h, that is to say it is</i>
<i>contained in I(K). The proof is now complete because the opposite</i>
inclusion is trivial.


<b>4.3 The tensor algebra of a free module</b>


In Section (4.3) we shall assume that R is non-trivial. (This is to
avoid certain tiresome and unimportant complications.) Let M be a free
<i>R-module and B one of its bases. We wish to describe the structure of T(M).</i>
<i>We know that T0(M) is a free R-module having the identity element of</i>


<i>T(M) as a base. Also, by Theorem 3 of Chapter 1, if p > 1, then Tp(M) is a</i>


<i>free R-module and it has as a base the set formed by all products bx ®</i>


<i>b2®---®bp, where bt e B. Accordingly T(M) is itself a free R-module and</i>


<i>the totality of products b1 ® b2 ® • • • ® bn, where b(eB and n > 0 ,</i>


constitutes a base. To complete the description of the algebra all we have to
do is explain how two elements in our base are to be multiplied. But here
<i>there is no problem because the product of bx ® b2 ® •' * ® bn and b\ ®</i>



<i>b'2 ® ''' ® K is simply the basis element bx ® • • • ® bn ® b\ ® • • • ® b'x.</i>


<i>Thus T(M) is what is called the free algebra generated by the set B.</i>
<i>The next theorem gives another characterization of T(M) when M is free.</i>
<i><b>Theorem 4. Let M be a free R-module with a base B and let (ft: £ —> A be a</b></i>


<i>mapping of B into an R-algebra A. Then </> has a unique extension to a</i>
<i>homomorphism T(M)—*A of R-algebras.</i>


<i>Remark. This result shows that T(M) solves a certain universal problem</i>
<i>involving mappings of B into an R-algebra. We leave the reader to make</i>
this precise.


<i>Proof The mapping 0 : B—>A can be extended to an R-linear mapping</i>
<i>(j>: M^>A. Let h: T(M)—+A be a homomorphism of R-algebras. Then h</i>
<i>extends <\> if and only if it extends (/>. However, we know (Theorem 2) that</i>
<i>there is precisely one h with the latter property.</i>


<i><b>Theorem 5. Let P be a projective R-module. Then the tensor algebra T(P) is</b></i>


<i>projective as an R-module. Consequently (for all n) Tn(P), since it is a direct</i>


</div>
<span class='text_page_counter'>(86)</span><div class='page_container' data-page=86>

<i>Covariant extension of a tensor algebra 75</i>


<i>Proof. Choose an ^-module Q so that P ®Q = F (say) is a free K-module.</i>
<i>Let G\ P—>F be the inclusion mapping and n: F—>P the projection onto</i>
<i>the first summand. Then T(F) is a free K-module (because F is free) and the</i>
<i>mappings T(o)\ T(P)-+T(F) and T(n): T(F)-+T{P) are IMinear. But</i>
<i>n ° a is the identity mapping of P and therefore T(n) ° T(<T) is the identity</i>


<i>mapping of T(P). The following lemma allows us to conclude that T(P) is</i>
<i>isomorphic to a direct summand of T(F); and, since T(F) is free, this means</i>
<i>that T(P) is projective.</i>


<i><b>Lemma 1. Let u: N —>M and v: M—+N be homomorphisms of R-modules</b></i>


<i>such that v°u is the identity mapping of N. Then u is an injection, v is a</i>
<i>surjection, and M is the direct sum of u(N) and the kernel, Ker v, of v. In</i>
<i>particular N is isomorphic to the direct summand u(N) of M.</i>


<i>Proof It is obvious that u is an injection and v a surjection. Let meM. Since</i>


<i>m=(m — u(v{m))) + u(v(m))</i>


<i>and m — u(v(m)) is in Ker v, it follows that M = Ker v + u(N). Now let x e</i>
<i>u(N) nKeri;. To complete the proof it suffices to show that x = 0. But</i>
<i>x = u(n) for some neN and now</i>


<i>n = (v°u)(n) = v(x) = 0.</i>


<i>Accordingly x = 0 and the proof is complete.</i>


<b>4.4 Covariant extension of a tensor algebra</b>
Let M be an ^-module and


<i>co:R-+S (4.4.1)</i>


a homomorphism of commutative rings. Our aim is to study the effect of the
covariant extension associated with (4.4.1) on the tensor algebra of M. Since
here we shall be concerned with more than one commutative ring, we shall


<i>embellish the symbol for the tensor algebra by writing TR</i> (M) rather than


<i>just T(M). This will help us to avoid certain ambiguities.</i>


<i>We know that TR(M) is a graded K-algebra and therefore (see Section</i>


(3.6)) its covariant extension 7^ (M) ®R<i> S is an 5-algebra that is graded by</i>


<i>the family {Tn(M) ®R S}neI</i> of 5-submodules. The elements of degree one


<i>form the 5-module M ®RS and this module generates TR (M) ®R S as an </i>


5-algebra.


By Theorem 2 there is a homomorphism


<i>X:TS(M®RS)^TR(M)®RS (4.4.2)</i>


<i>of 5-algebras which extends the inclusion mapping of M ®R S in</i>


<i>TR</i> (M) ®<i> R S. (Here Ts (M ®R S) denotes the tensor algebra of the 5-module</i>


</div>
<span class='text_page_counter'>(87)</span><div class='page_container' data-page=87>

<i>76 The tensor algebra of a module</i>


<i>of graded S-algebras. However, as the following theorem shows, much more</i>
is true.


<i><b>Theorem 6. The mapping</b></i>


<i>X:TS(M®RS)-+TR(M)®RS</i>



<i>{described above) is an isomorphism of graded S-algebras.</i>


<i>Proof Any S-algebra, and in particular TS(M ®R S), can be regarded as an</i>


^-algebra by leaving the ring structure unchanged and using (4.4.1) to turn
<i>it into an K-module. Now the mapping M —> TS(M ®RS) which takes m</i>


<i>into m ® 1 is K-linear. Accordingly, by Theorem 2, there is a</i>
homomorphism


<i>O:TR(M)-+TS(M®RS)</i>


<i>of R-algebras such that 0{m) = m ® 1 for all m e M.</i>


<i>Consider the mapping TR(M) x Sã TS(M đRS) in which the image of</i>


<i>(x, s) is s6(x). This is a bilinear mapping of /^-modules, and therefore it</i>
induces a homomorphism


<i>fx:TR(M)®RS->Ts(M®RS)</i>


<i>of K-modules which is such that fi(x ® s) = s6(x) for all xeTR (M) and seS.</i>


<i>It is now a simple matter to verify that \i is actually a homomorphism of </i>
<i><b>S-algebras and that \iiyn ® 5) = m ® s for meM and seS.</b></i>


<i>Consider /x ° k and X ° fi. Each of these is a homomorphism of S-algebras</i>
<i>and each leaves M ®RS elementwise fixed. It follows from this (and the fact</i>



<i>that the S-algebras concerned are generated by M ®R S) that \i ° X and X ° \i</i>


<i>are identity mappings and hence that X is an isomorphism. The proof is now</i>
complete.


In view of Theorem 6 we may write


<i>Ts(M ®RS) =TR(M)®RS (4.4.3)</i>


and


<i>Tn (M ®R S) = Tn (M) ®R S, (4.4.4)</i>


and we may note, in passing, that (4.4.4) is a special case of Theorem 9 of
Chapter 2.


<b>4.5 Derivations and skew derivations on a tensor algebra</b>


<i>Let M be an ^-module. An /^-linear mapping of M into R is called</i>
<i>a linear form on M. If now D is a derivation on T(M), then, since it is of</i>
<i>degree — 1 , it induces an R-homomorphism Tl(M)—^T0(M). But</i>


<i>7; (M) = M and we know that we can identify T0(M) with R. Thus D extends</i>


</div>
<span class='text_page_counter'>(88)</span><div class='page_container' data-page=88>

<i>Derivations and skew derivations on a tensor algebra 11</i>


Similar considerations, this time based on Lemma 6 of Chapter 3, show
that each skew derivation extends a linear form and that distinct skew
derivations give rise to distinct linear forms. In the next two theorems it will
<i>be proved that T(M) is fully endowed with both derivations and skew</i>


derivations.


<i><b>Theorem 7. Let f be a linear form on the R-module M. Then there is exactly</b></i>


<i>one derivation on T(M) which extends f</i>


<i>Proof. Suppose that p>l. There is a multilinear mapping of the p-fold</i>
<i>product M x M x • • x M i n t o T</i>p_ ^ M ) in which (ml5 m2, . . . ,rap) becomes


<i><b>p</b></i>


<i>X / K M ^ i đ ã ã' đ fht đ ••' ® mp). (4.5.1)</i>
<i><b>i= 1</b></i>


(Here, as usual, the A<i> over m{</i> indicates that this factor is to be omitted.)


<i>Consequently there is induced an K-homomorphism Dp: 7p(M)ã</i>


<i>Tp_l(M\ where Dp(m1 đ m</i>2<i> ® * * • ® mp) is the element (4.5.1). Note that</i>
<i><b>D</b><b>x</b><b>=f.</b></i>


Next the homomorphisms Dl 9 Z)2, D3<i>, . . . give rise to an </i>


<i>R-endomorphism D: T(M)-+ T(M), of degree — 1, which agrees with Dp</i> on


<i>Tp(M). To complete the proof we show that D is a derivation.</i>


Let ml9 m2<i>, . . . , mp</i> and mi, m'2<i>,..., m'q belong to M and put x = ml®</i>


<i>m2 ® •' • ® mp, x' = m\®m'2®''' ® m'q. Then</i>



<i>D(x ® x') = Dinti ®--®mp®mf1®--®mq)</i>


<i><b>p</b></i>


<i><b>= (Dx) ® x' + x ® (Dx</b><b>f</b><b>). (4.5.2)</b></i>


Since this relation extends by linearity to any two elements of T(M), the
proof is complete.


<i><b>Theorem 8. Let f be a linear form on the R-module M. Then there is one and</b></i>


<i>only one skew derivation on T(M) which extends f</i>


<i>Proof This parallels the proof of Theorem 7. We first show that for each</i>
<i>p> 1, there is an K-homomorphism Ap: Tp(M)^> Tp_1(M), where</i>


<i><b>p</b></i>


Ap<i>(m! ® m2 ® - - - ® mp)= X (-l)i + lf(^i)(^i ®"-®rhi®'"® mp).</i>


</div>
<span class='text_page_counter'>(89)</span><div class='page_container' data-page=89>

<i>78 The tensor algebra of a module</i>


which extends /. Moreover, in place of (4.5.2) we obtain
A(mx đ ã ã ã ® mp<i> ® mi ® • • • ® m'q)</i>


<i>= A(m1 đ ã ã ã đ mp) đ (mi đ • • • ® m'q)</i>


<i>+ ( - l)p(mx ® • • • ® mp) ® A(mi ® • • • ® m'q).</i>



This shows that A is a skew derivation and ends the proof.


<b>4.6 Comments and exercises</b>


We make here a few comments on the theory of tensor algebras
and provide some exercises. As usual certain of these exercises have been
marked with an asterisk; for these particular exercises solutions are
provided in the next section.


It is clear that if an ^-module M is generated by a given set of elements,
<i>then the same set of elements will also generate T(M) as an K-algebra. It</i>
<i>follows that if M is a cyclic ^-module, then T(M) can be generated by a</i>
single element. Consequently (see Exercise 1 of Chapter 3) the tensor
algebra of a cyclic module is a homomorphic image of the polynomial ring
<i>&[X] and, in particular, it is commutative. The first exercise builds on this</i>
observation.


<i><b>Exercise 1*. Let M be an R-module with the property that every finitely</b></i>


<i>generated submodule is contained in a cyclic submodule. Show that its tensor</i>
<i>algebra is commutative.</i>


<i>For example, if R is an integral domain and F is its quotient field, then the</i>
<i>tensor algebra of F (considered as an .R-module) is commutative.</i>


<i>We remarked in Section (4.3) that if M is a free .R-module with a base B,</i>
<i>then T(M) is the free R-algebra generated by B. (The notion of the free</i>
algebra generated by a set was enlarged upon in Section (3.8).) The next
exercise is concerned with such an algebra in the special case where the
<i>ground ring R is an integral domain.</i>



<i><b>Exercise 2*. Let R be an integral domain and M a free R-module. Further let</b></i>


<i>x and y be non-zero elements of the tensor algebra T(M). Show that the</i>
<i>product of x and y in T(M) is not zero.</i>


<i>Once again let M be an K-module. In Section (3.8) we saw how a</i>
commutative, graded K-algebra can be formed from the matrices


<i>[</i>

<i>r m</i>


<i><b>o</b></i>



<i>where reR and meM. Let us denote this algebra by T(M) and let Tn{M)</i>


</div>
<span class='text_page_counter'>(90)</span><div class='page_container' data-page=90>

<i>Comments and exercises 79</i>


whereas F J M ) is made up of the matrices
<b>0 m l</b>


<b>o oj</b>



<i>Finally Tn(M) = 0 if n±0,1.</i>


Consider the mapping M—>F(M) in which m of M is mapping into


<i><b>O ml</b></i>


<b>o oj</b>




This mapping is R-linear and so it extends to a homomorphism


<i><b>X\T{M)^T{M) (4.6.1)</b></i>


<i>of /^-algebras. Evidently X preserves degrees and is surjective.</i>


<i><b>Exercise 3*. Let X\ T(M)—>r(M) be the homomorphism of graded </b></i>


<i>R-algebras described in (4.6.1). Determine its kernel and show that k is an</i>
<i>isomorphism if and only if M (g)RM = 0.</i>


<i><b>Exercise 4. Let R be an integral domain and F its quotient field. Use</b></i>


<i>Exercise 3 to give a simplified description of the tensor algebra of the </i>
<i>R-module F/R.</i>


<i>The next exercise is of general interest. Let I^R be an ideal of R and let</i>
<i>M be an ^-module. Then IT(M) is a homogeneous two-sided ideal of T(M)</i>
<i>and therefore T(M)/IT(M) is a graded K-algebra. But / annihilates this</i>
<i>algebra so that T(M)/IT(M) = A (say) can be regarded as a graded </i>
<i>R/I-algebra. Now the submodule Au of A, formed by the homogeneous</i>


<i>elements of degree one is the image of Tl(M) = M, and the kernel of</i>


<i>Tl(M)-^A1 is IM. Hence the natural mapping T(M)-+T{M)/IT(M)</i>


induces a homomorphism
<i>M/IM->T(M)/IT(M)</i>


of jR//-modules and this in turn induces a homomorphism



<i>0: TR/I(M/IM)-+T(M)/IT(M) (4.6.2)</i>


<i>of R/7-algebras. (Here by TR/I (M/IM) we mean the tensor algebra of M/IM</i>


considered as an K/7-module.)


<i><b>Exercise 5*. Show that the homomorphism</b></i>


<i><t>:TR/l(M/IM)^T(M)/IT(M)</i>


<i>of (4.6.2) is an isomorphism of graded R/l-algebras.</i>
This result enables us to make the identification


</div>
<span class='text_page_counter'>(91)</span><div class='page_container' data-page=91>

<i>80 The tensor algebra of a module</i>


Our final comments concern generalized ordinary and skew derivations.
<i>These were defined in Section (3.8). First suppose that D is a generalized</i>
<i>derivation of degree i on the tensor algebra of the /^-module M. Then D</i>
<i>induces an K-linear mapping of Tx</i>(M) = M into 7^ + x<i>(M). Next \Re T0(M)</i>


and


<i>so that D(lR) = 0. It follows that the mapping To{M)-+ Tt(M) induced by D</i>


<i>is a null mapping. Again, if p > 1 and m</i>1? m2<i>, . . . ,mp</i> belong to M, then it is


<i>easily seen (using induction on p) that</i>
<i><b>p</b></i>



<i>D(m1</i> ® m2<i> ® • • • ® mp) = £</i> mi ® ' * * ® ^mv ® ' ' ' ® "V (4.6.4)
<b>v = 1</b>


<i>It follows that D is fully determined by its degree and its effect on</i>
<i>Tl(M) = M.</i>


<i>Next assume that A is a generalized skew derivation on T(M) of degree i.</i>
<i>Everything that was said about D in the last paragraph applies to A except</i>
that (4.6.4) has to be replaced by


A(mx (x) m2<i> đ ã ã ã đ mp)</i>


= £ ( - l )( v + 1 ) lm1® - - - ® A mv® - - - ® m;, . (4.6.5)
<b>v = l</b>


The next exercise generalizes Theorems 7 and 8.


<i><b>Exercise 6*. Let M be an R-module, let i be a given integer, and let</b></i>


<i>f: M —+Ti + l(M) be an R-linear mapping. Show that there is exactly one</i>


<i>generalized derivation of degree i and exactly one generalized skew derivation</i>
<i>of degree i, on T{M), that extend f</i>


<b>4.7 Solutions to selected exercises</b>


<i><b>Exercise 1. Let M be an R-module with the property that every finitely</b></i>


<i>generated submodule is contained in a cyclic submodule. Show that its tensor</i>
<i>algebra is commutative.</i>



<i>Solution. Let £, £' belong to T(M). We wish to prove that f f = £'£ and this</i>
will follow with full generality if we can establish it on the assumption that
<i>£ = mx</i> đ m2<i> đ ã ã * đ mp, Ê' = m\ đ m</i>2<i> đ ã ã ã đ m'q, where the mi and m) are</i>


<i>in M. Now, by hypothesis, there exists a cyclic submodule N, of M,</i>
<i>containing all the mt and the m). Let (/>: T(iV)^T(M) be the </i>


<i>algebra-homomorphism that extends the inclusion mapping of N into M. Then</i>
<i>there exist x and x' in T(N) such that (j)(x) = ^ and cj)(x') = £)' and therefore</i>


</div>
<span class='text_page_counter'>(92)</span><div class='page_container' data-page=92>

<i>Solutions to selected exercises 81</i>


<i><b>Exercise 2. Let R be an integral domain and M a free R-module. Further let</b></i>


<i>x and y be non-zero elements of the tensor algebra T(M). Show that the</i>
<i>product of x and y in T(M) is not zero.</i>


<i>Solution. We can write</i>


x2<i> + • • • +xh</i> (


<i>where xt e T{{M) and y^e 7}(M), and it will suffice to prove that the product</i>


<i>of xh and yk</i> is not zero. Equally well we may assume for the rest of the proof


<i>that x belongs to Th(M) and y to Tk(M). Before proceeding note that under</i>


the canonical isomorphism
<i>Th(M)®Tk(M)*Th+k(M)</i>



<i>the element of Th+k(M) that corresponds to x ® y is precisely the product of</i>


<i>x and y in T(M).</i>


<i>Put U=Th(M) and V= Tk{M). These are free K-modules because M is</i>


free. Let wl9 w2<i>, . . . , up be a base for U and vl9 v2,..., vq SL base for V. Then</i>


<i>(see Chapter 1, Theorem 3) the elements ut ® v} form a base for U ® V.</i>


<i>Next, we can write x = alul -\-a2u2 + • • • +apup and y = b1vl +b2v2</i> + • • •


<i>+ bqvq, where a</i>f, ^ belong to R, and then


<i>But x 7^ 0 and ^ ^ 0. Hence we can find i and j so that a</i>f # 0 and fc^ # 0 and


<i>then we have atbj^0 because R is an integral domain. However, the</i>


<i>products ur (x) vs are linearly independent over R and therefore it follows</i>


<i>that x ®y^0. This completes the solution.</i>


<i><b>Exercise 3. Let X: r(M)—>r(M) be rfte homomorphism of graded </b></i>


<i>R-algebras described in (4.6.1). Determine its kernel and show that X is an</i>
<i>isomorphism if and only if M ®RM = 0.</i>


<i>Solution. Let m</i>1? m2<i>, . . . , mp belong to M. Then X{m1 đ m2 đ ã ã ã (g) m</i>p) is



the product of the matrices


<b>Lo o j</b>

<b>( i = l , 2 p)</b>


<i>and this is zero if p>2. Thus A maps all of T</i>2(M), T3(M), T4(M),... into


zero.


<i>Let x belong to T(M) and let</i>
x = xo + x1 + x2 + • • •


<i>be its decomposition into homogeneous components. Then xoeR, x1eM</i>


</div>
<span class='text_page_counter'>(93)</span><div class='page_container' data-page=93>

<i>82 The tensor algebra of a module</i>


<i>x0</i>


It follows that


<i>K e r ^ = X Tn(M).</i>


<i><b>n>2</b></i>


<i>Of course, since k is surjective it is an isomorphism if and only if Ker k = 0.</i>
<i>Now if k is an isomorphism, then M ® M =T2(M) = 0. On the other</i>


hand, if M ® M = 0, then M ® M ® - - ® M = 0 provided there are at least
two factors (see Chapter 2, Theorem 1). Consequently if M ® M = 0, then
<i>Tn (M) = 0 for all n > 2 and therefore Ker k = 0. This completes the solution.</i>



<i>Exercise 5. Show that the homomorphism</i>
<i><t>:TR/I(M/IM)^T(M)/IT(M)</i>


<i>of (4.6.2) is an isomorphism of graded R/I-algebras.</i>


<i>Solution. It is clear that (/> preserves the degrees of homogeneous elements</i>
so that it is enough to construct an algebra-homomorphism, in the reverse
direction, which when combined with (/> (in either order) produces an
identity mapping.


<i>Let ft: T(M)—• T(M)/IT(M) be the natural homomorphism and for m in</i>
<i>M let m denote its natural image in M/1M. Next, for the moment, let us</i>
<i>regard the ^//-algebra TR/I(M/IM) as an K-algebra. Then the jR-linear</i>


<i>mapping M^»TR/I(M/IM) which takes m into m extends to a </i>


<i>homo-morphism T{M)^> TR/I(M/IM) of K-algebras. But this vanishes on IT(M)</i>


and so it induces a homomorphism


This is a homomorphism of K-algebras and of ^//-algebras and it satisfies
<i>il/(h(m)) = rh for all m in M. Now (p(m) = h(m) by construction so that</i>
<i>\jj{(j)(m)) = rh and 0(^(/z(m))) = /z(m). But the elements m generate the </i>
<i>R/I-algebra TR/I(M/IM) and, because M generates T(M), the elements h(m)</i>


<i>generate the K/7-algebra T(M)/IT(M). Thus \jj°(f) and 0 ° ^ must be</i>
identity mappings and with this the solution is complete.


<i>Exercise 6. Let M be an R-module, let i be a given integer, and let f: M —•</i>
<i>Ti + 1(M) be an R-linear mapping. Show that there is exactly one generalized</i>



<i>derivation of degree i and exactly one generalized skew derivation of degree i,</i>
<i>on T(M), that extend f</i>


</div>
<span class='text_page_counter'>(94)</span><div class='page_container' data-page=94>

<i>Solutions to selected exercises 83</i>


<i>Suppose that p > 1. The mapping of the p-fold product MxMx - - xM</i>
<i>into Ti+p(M) which takes (m</i>l9 m2<i>, . . . , mp) into</i>


is multilinear and therefore it induces an R-linear mapping
<i>Ap:Tp(M)-^Tp+i(M)</i>


which is such that


<i><b>p</b></i>


<i>A</i>

<i>p</i>

<i>(m</i>

<i>l</i>

<i>®m</i>

<i>2</i>

<i>®--®m</i>

<i>p</i>

<i>) = £ ( - l )</i>

0

^

1

<i>^ đ ã ã ã ® f(m</i>

<i>ll</i>

<i>) ® • • • ® m</i>

p

.



<b>A* = i</b>


Next we define Ao<i>: T0(M)-+ T((M) to be the null homomorphism. Then</i>


Ao, A1, A2, . . . can be combined to give an R-linear mapping


<i>A: T(M)-+T(M) of degree i which agrees with A</i>n<i> on Tn(M) for all n > 0 .</i>


<i>Since Ax=f, A extends /. Now to show that A is a generalized skew</i>


<i>derivation we have to check that, for x and y in T(M), we have</i>
<i>A(x ® y) = (Ax) ® y + Jl (x) ® {Ay),</i>



<i>where J is the main involution (see Section (3.8)). However, it suffices to do</i>
<i>the checking when x and y are homogeneous elements whose degrees are</i>
strictly positive.


<i>Suppose then that xeTp(M) and yeTq(M), where p> 1 and q> 1. We</i>


have to show that


<i>A(x ®y) = (Ax) ® y + ( - l)ipx ® (Ay).</i>


<i>But now we can specialize still further and suppose that x = ml ® m2 ® * * *</i>


<i>® mp and y = m\ đ m'2 đ ã ã ã ® m'q, where mi and m) are elements of M. On</i>


this understanding we have


<i>A(x®y)= £</i>


<i><b>n=i</b></i>


<i>(m'v)</i>


<i>= (Ax) ® y + ( - l)</i>ip<i>x ® (Ay)</i>


</div>
<span class='text_page_counter'>(95)</span><div class='page_container' data-page=95>

<b>5</b>



The exterior algebra of a module



<b>General remarks</b>



The exterior algebra is one of the most interesting and useful of the
algebras that can be derived from a module. As we shall see it is an
anticommutative algebra having intimate connections with the theory of
determinants. Our aim in this chapter will be to establish all its main
properties; and in so doing we shall follow a pattern which can be used
again, with only small modifications, to study a related algebra, namely the
symmetric algebra of a module.


<i>As usual R and S are reserved to denote commutative rings with an</i>
identity element, and we again allow ourselves the freedom to omit the
suffix from the tensor symbol when there is no uncertainty concerning the
ground ring. Algebras are understood to be associative and to have an
identity element; and homomorphisms of rings and algebras are required to
<i>preserve identity elements. Finally T(M) denotes the tensor algebra of an</i>
<i>^-module M.</i>


<b>5.1 The exterior algebra</b>


<i>Let M be an /^-module. We propose to define its exterior algebra,</i>
and, as in the case of the tensor algebra, we shall do this by means of a
universal problem. However, because we are now in a position to make use
of the properties of tensor algebras, the details on this occasion will be much
simpler.


<i>Suppose that 0 : M —> A is an jR-linear mapping of M into an ^-algebra</i>
<i>A, and suppose that (c/>(m))</i>2<i>=0 for all meM. If now h:A^>B is a</i>
<i>homomorphism of K-algebras, then h ° (/> is an K-linear mapping of M into</i>
<i>B with the property that, for every meM, the square of (h °4>)(m) is zero.</i>


<i><b>Problem. To choose A and (j) so that given any R-linear mapping \//: M —• J9</b></i>



</div>
<span class='text_page_counter'>(96)</span><div class='page_container' data-page=96>

<i>The exterior algebra 85</i>


<i>(B is an R-algebra) such that (i/^(m))</i>2<i><sub> = 0 for all meM, there shall exist a</sub></i>
<i>unique homomorphism h: A-^B, of R-algebras, such that / j ° ^ = f</i>


<i>Evidently the problem has at most one solution. More precisely, if (A, (j>)</i>
<i>and (A\ (/>') both solve the problem, then there exist inverse isomorphisms</i>
<i>k: A—• A'and A': A' -^A,oiK-algebras, such that A ° </> = (/>' and A'°<£' = 0.</i>
<i>The next theorem shows inter alia that there is always a solution.</i>


<i><b>Theorem 1. The above universal problem possesses a solution. If {A, (/>) is</b></i>


<i>any solution, then</i>


<i>(i) (/>: M —+ A is an injection;</i>


<i>(ii) A is generated, as an R-algebra, by (j)(M);</i>


<i>(iii) there is a grading {An}nel on A such that Ax</i> = (/>(M);


<i>(iv) for each p>l,Apisthe p-th exterior power of M where, for m1,m2,</i>


<i>. . . ,mpin M, we have m1 r\m2 A • • • Amp = (j)(ml)(t)(m2) • • • 0(</i>WP)»*


<i>(v) the structural homomorphism R^>A maps R isomorphically onto</i>
<i>Ao.</i>


<i>Remark. As in the case of the tensor algebra, condition (ii) ensures that</i>
there can only be one grading with the property described in (iii).



<i>Proof. It is sufficient to construct one solution of the universal problem that</i>
<i>has all the five properties. To this end let T(M) be the tensor algebra of M</i>
<i>and let m belong to M = Tx (M). Then m®m belongs to T2(M) and therefore</i>


<i>such elements generate a homogeneous two-sided ideal, J(M) say, in T(M).</i>
<i>Let Js(M) = J(M)nTs(M). Then J</i>o<i>(M) = 0 and J1(M) = O. Furthermore</i>


Jn(M) = O f o r a l l n < 0 .


<i>Since J(M) is homogeneous, the factor algebra T(M)/J(M) = A (say) has</i>
a grading {,4S}S6Z<i> in which As is the image of TS(M). Bearing in mind that</i>


Tx<i> (M) = M, we let </>: M —• A be the homomorphism induced by the natural</i>


<i>mapping T(M)^> T(M)/J(M). Then A and 0 satisfy (i) because J1(M) = 0.</i>


<i>Since Tx (M) generates T(M) as an algebra, (/>(M) generates A as an algebra</i>


<i>so that (ii) holds as well. Evidently (iii) is satisfied. Again T0(M) is mapped</i>


<i>isomorphically onto Ao</i> and therefore the requirement (v) is also met.


<i>Now suppose that p > 1. The natural mapping of T(M) onto A induces a</i>
<i>surjective homomorphism of Tp(M) onto Ap</i> and the kernel of this


<i>homomorphism is Jp(M). But, by the definition of J(M), Jp(M) is the</i>


<i>submodule of Tp(M) generated by all products ml®m2®-'-®mp</i> with



<i>mi = mi + 1 for some i. That condition (iv) holds therefore follows from</i>


Theorem 5 of Chapter 1.


</div>
<span class='text_page_counter'>(97)</span><div class='page_container' data-page=97>

<i>86 The exterior algebra of a module</i>


<i>Finally suppose that \j/: M —• B is an K-linear mapping (of M into an </i>
<i>R-algebra B) which is such that (i//(m))2 = 0. Then (Chapter 4, Theorem 2), \j/</i>
<i>extends to an algebra-homomorphism T(M)—+ B which, it is clear, vanishes</i>
<i>on J(M). Accordingly there is induced a homomorphism h: A—+B of </i>
<i>R-algebras which satisfies (ft°(/>)(m) = i/f(m) for all meM. This shows that</i>
<i>h°(j) = \l/ and now (ii) ensures that there is only one algebra-homomorphism</i>
with this property. Thus the proof is complete.


<i>We next introduce some convenient terminology. Let (A, (j>) solve our</i>
universal problem. Put


<i>E(M) = A (5.1.1)</i>


<i>and let {En</i> (M)}<i> neI be the grading referred to in Theorem 1. Since <\> maps M</i>


<i>isomorphically onto E1(M) we can use this fact to make the identification</i>


<i>E1(M) = M. (5.1.2)</i>


<i>In this way M becomes a submodule of E{M) and, as we know, it generates</i>
<i>E(M) as an K-algebra. The grading on E(M) is, of course, non-negative and</i>
<i>the structural homomorphism R—>E(M) maps R isomorphically onto</i>
<i>EQ(M). It follows that we may make the further identification</i>



<i>E0(M) = R (5.1.3)</i>


when it is convenient to do so.


<i>The symbol A will be used for multiplication on E(M). This secures that,</i>
<i>for p > 1, Ep(M) is the p-th exterior power of M according to the definition</i>


given in Section (1.4). Note that


mAm = 0 (5.1.4)
<i>for all min M = E1(M).</i>


<i>The algebra E(M) is called the exterior algebra of the module M. The</i>
following theorem simply records the fact that it solves the universal
problem with which we started.


<i><b>Theorem 2. Let E(M) be the exterior algebra of the R-module M and let</b></i>


<i>\jj\ M—+B be an R-linear mapping, of M into an R-algebra B, such that</i>
<i>(\jj(m))2 = 0 for all meM. Then \j/ has a unique extension to a homomorphism</i>
<i>E(M)^B of R-algebras.</i>


</div>
<span class='text_page_counter'>(98)</span><div class='page_container' data-page=98>

<i>Functorial properties 87</i>


<i>a homogeneous element is preserved and therefore T(M)—+E(M) is a</i>
<i>homomorphism of graded algebras. In particular for each p e Z there is</i>
induced a homomorphism


<i>Tp(M)-+Ep(M) (5.1.5)</i>



<i>of K-modules. Of course for p = 0 this is an isomorphism and for p = 1 it is</i>
<i>the identity mapping. Suppose next that p > 1 and let ml</i>, m2<i>, . . . , mp</i> belong


<i>to M. Then in (5.1.5) m1 ®m2 ®''' ®mp is mapped into ml /\m2 A • • • /\mp</i>


and therefore (5.1.5) is the canonical homomorphism of the p-th tensor
power onto the /?-th exterior power according to the definition given in
Section (1.4).


<i><b>Theorem 3. The exterior algebra E(M) of the R-module M is </b></i>


<i>anticommuta-tive.</i>


<i>Proof. Since m A m = 0 for all m e M, the theorem follows from Lemma 3 of</i>
<i>Chapter 3 and the properties of E(M) that have already been noted.</i>


<b>5.2 Functorial properties</b>


<i><b>Let / : M —• N be a homomorphism of K-modules. We can regard</b></i>
<i>N as a submodule of the exterior algebra E(N) and then f(m) A f(m) = 0 for</i>
<i>all meM. It follows (Theorem 2) that / has a unique extension to a</i>
homomorphism


<i>E(f):E(M)-+E(N) (5.2.1)</i>


<i>of K-algebras. This is such that if m1,m2,. • •, mp</i> belong to M, then


£(/)(»»! Am2<i> A - - - Amp) = f(ml)Af(m2)A"- Af(mp). (5.2.2)</i>


<i>Moreover, since M consists of the elements of degree one in E(M),</i>


<i>the homomorphism E(f) preserves degrees. Thus E(f): E(M)-^E(N)</i>
<i>is a homomorphism of graded algebras and therefore, for each peZ, it</i>
induces a homomorphism


<i>Ep(f):Ep(M)^Ep(N) (5.2.3)</i>


<i>of K-modules. Of course E0(f) is an isomorphism and, when p>l,</i>


<i>EP(f)(mi Am2 A'"</i> A^P)= =/ ( ^ i ) A / ( m2<i>) A - - - Af(mp). (5.2.4)</i>


<i>Naturally when N = M and / is its identity mapping, E(f) is the identity</i>
<i>mapping of E(M).</i>


Next suppose that in addition to / : M — • N we are given a second
<i>homomorphism g: K-+M of K-modules. Then</i>


<i>E(f)oE(g) = E(fog) (5.2.5)</i>


and also


</div>
<span class='text_page_counter'>(99)</span><div class='page_container' data-page=99>

<i>88 The exterior algebra of a module</i>


<i>for all peZ.ln particular the exterior algebra provides a second example of</i>
a covariant functor from K-modules to graded R-algebras. It follows, as in
<i>all such situations, that if / is an isomorphism of modules, then E(f) is an</i>
<i>isomorphism of algebras and E(f~l) is its inverse.</i>


At this point it is convenient to mention a notation which is very
commonly used in the theory of exterior algebras and of which the reader
<i>should be aware. This is the wedge notation which we have already</i>


employed to describe multiplication. When used more extensively we put
<i>f\M = E{M\ (5.2.7)</i>


<i>/\pM = Ep{M\ (5.2.8)</i>


<b>A /</b>

<b> =</b>

<b> £ ( A (5.2.9)</b>



and


<i>/\pf = Ep(f). (5.2.10)</i>


<i>Accordingly (5.2.5) and (5.2.6) become (/\f)°(f\g) = f\(f°g) and</i>
(AP / ) ° (AP<i> 9) ~ f\P ( / ° d) respectively. However, since we aim to develop</i>
the theories of several algebras along parallel lines, we shall keep to a
uniform notation. This will help a little when making comparisons.
<i><b>Theorem 4. Let f:M-+Nbea surjective homomorphism of R-modules and</b></i>


<i>let K be its kernel. Then E(f): E(M) —• E(N) is a surjective homomorphism of</i>
<i>graded algebras whose kernel is the two-sided ideal which K generates in</i>
<i>E(M).</i>


<i>Remark. Since M is a submodule of E(M) so too is K. It is therefore</i>
<i>meaningful to speak of the two-sided ideal which K generates in E(M).</i>
<i>Proof The corresponding result for tensor algebras is Theorem 3 of</i>
Chapter 4. Two proofs were given of that result and, of these, the second one
can be readily adapted. The details are left to the reader.


We now add a few remarks that are relevant to the study of the exterior
<i>algebra of a finitely generated module.</i>



<i>Let ex, e2,... ,en</i> belong to the K-module M and let elements w1,M2,...,Mr


of M satisfy relations


<i>U</i>


<i>s = elCls + e2C2s+ ' ' ' +enCns (</i>s== U 2, . . . , r). (5.2.11)
<i>(Here the coefficients ctj are in R and, for convenience, we have written them</i>


on the right-hand side of the module elements which they multiply.) For
<i>integers jl9j2, - • • Jr></i> a'l between 1 and n, we put


</div>
<span class='text_page_counter'>(100)</span><div class='page_container' data-page=100>

<i>The exterior algebra of a free module 89</i>


<i><b>Lemma 1. Let the situation be as described above. Then</b></i>


<i>ux A u2 A • • • A ur = Y, ieh A eh</i> A " ' '<i> A ej)Chh...Jr></i>


<i>where the sum is taken over all sequences (jl,j2, • • • Jr) which satisfy 1 <jt <</i>


<i>j2 < ''' <jr < n. In particular ul A U2 A • • • A ur = 0 if r>n.</i>


<i>Proof We have</i>


<i>ux A u2 A • • • A ur={elcl! + • • • + encnl) A (excl2 + • • • + encn2)</i>


<i>A - " A{elclr+---+encnr)</i>


<b>= E K A ^2</b><i><b> A • • • A e</b><b>ir</b><b>)c</b><b>ixl</b><b>c</b><b>il2</b></i><b> . . . cl>?</b>



(5.2.13)
<i>where the summation is taken over a// sequences (i1, /</i>2, • • • > 0 of r integers


<i>lying between 1 and n. However, eix Aeiz A • • • Aeir = 0 if the sequence</i>


<i>(il9 i2,..., ir) contains a repetition. Hence from now on we may assume that</i>


<i>r<n.</i>


<i>Suppose that 1 <j1 <j2 < • • • <jr<n and let (il9 i2,..., ir) be a </i>


<i>permuta-tion of (JiJ2,... Jr). Then, because e</i>p<i> A eq = — eq A ep, we have</i>


<i>eh</i> A ^ A " " A ei p= ± ^ A ^ A • • • A 6j r,


<i>where the plus (respectively minus) sign is to be taken if (iu i2,..., ir) is an</i>


<i>even (respectively odd) permutation of (jlj2, - • • Jr)- Accordingly the</i>


<i>permutations of {j1j2,... ,jr) contribute to (5.2.13) an amount equal to</i>


<i>(eh</i> A ^-2<i> A • • • A eJr)CJiJ2 jr. The lemma follows.</i>


As an application of the lemma we have


<i><b>Theorem 5. Let the R-module M be generated by e</b>l9 e2,..., en. Then</i>


<i>Er(M) = 0 if r>n. If l<r<n, then Er(M) is generated, as an R-module, by</i>


<i>the elements eji A eJ2 A • • • A ejr, where (JiJ2,... Jr) is a typical sequence of r</i>



<i>integers satisfying 1 <j1 <j2 < • • • <jr < n.</i>


<b>5.3 The exterior algebra of a free module</b>


<i>Throughout Section (5.3) we shall assume that the ring R is </i>
<i>non-trivial.</i>


<i>Let M be a free .R-module. If p > 1, then Ep(M) is free by Theorem 6 of</i>


<i>Chapter 1. Moreover E0(M) is free because it is isomorphic to R. But</i>


<i>£ ( M ) = £ En(M) (d.s.)</i>


<i><b>n>0</b></i>


<i>so that E(M) too is a free .R-module.</i>


<i><b>Theorem 6. Let P be a projective module. Then E(P), considered as an </b></i>


<i>R-module, is projective. Consequently, for all s, ES(P) (since it is a direct</i>


</div>
<span class='text_page_counter'>(101)</span><div class='page_container' data-page=101>

<i>90 The exterior algebra of a module</i>


<i>Proof. The corresponding result for tensor algebras is Theorem 5 of</i>
Chapter 4. The argument used there, with only trivial modifications, also
yields the desired result on this occasion. The details are left to the reader.
<i>Let M be a free R-module and let B be a base for M. If B is infinite, then, by</i>
<i>Theorem 6 of Chapter 1, all the exterior powers E0(M\ EX{M\ £</i>2( M ) , . . .



are non-zero free /^-modules.


<i>Now consider what happens when B consists of a finite number of</i>
elements, say w1? w2<i>, . . . , un, where n>0. By Theorem 5, Er(M) = 0 if r>n.</i>


On the other hand, if 1 < r < n , then Theorem 6 of Chapter 1 shows that
<i>Er(M) has a base consisting of the products uji</i> A<i> uJ2</i> A • • • AMJP, where


<i>UiJn • • • Jr) is</i> a<i> sequence of integers satisfying 1 <jl <j2< ''' <jr<n.</i>


<i>Thus Er(M) has a base consisting of (") elements and, in particular,</i>


<i>0. Of course Eo(M)^0 because it is isomorphic to R.</i>


<i><b>Theorem 7. Let M be a free R-module and B be a base of M. Then the</b></i>


<i>number of elements in B is the upper bound of all integers k such that</i>


This is an immediate consequence of our previous remarks.


<i><b>Corollary. Let B and B' be bases of the free module M. Then either B and B'</b></i>


<i>are both infinite, or they are both finite and contain the same number of</i>
<i>elements.</i>


<i>The number of elements in a base of a free K-module M is called the rank</i>
<i>of M and it will be denoted by rank</i>/?(M). Thus rankK(M) is either a


non-negative integer or else it is 'plus infinity'. If rankR(M)= oo, then



rankfl (£0<i>(M)) = 1 and rank^ (Ep(M)) = oo for all p > 1. On the other hand if</i>


rankR(M) = n, then


<i><b>mnk</b><b>R</b><b>(E</b><b>s</b><b>(M)) = (</b><b>n</b><b>) (5.3.1)</b></i>


for all s > 0 .


<i>Let F be a free K-module of rank n and let f:F—+F be an </i>
<i>JR-endomorphism of F. Then En(f) is an endomorphism of En(F) and, by</i>


<i>(5.3.1), En(F) has a base consisting of a single element. Consequently there</i>


<i>exists aeR such that En(F): En(F)—>En(F) is the homothety a and,</i>


<i>moreover, a is uniquely determined. This scalar a is called the determinant of</i>
/ and it is denoted by Det(/). Hence


/ ( * i ) A / ( X2) A • • • A / ( x J = Det(/)(x1 AX2<i> A • • • Ax„), (5.3.2)</i>


where xx, x2, . . . , xn<i> are any n elements of F.</i>


<i>Suppose now that e1, e2,... ,en is a base of F. Then</i>


</div>
<span class='text_page_counter'>(102)</span><div class='page_container' data-page=102>

<i>The exterior algebra of a free module</i> 91


where


<i><b>C =</b></i>



<b>C<sub>12</sub></b>


<i><b>C</b><b>22</b></i> <i><b>-In</b></i> <b><sub>(5.3.4)</sub></b>


is the matrix of / with respect to the given base. By Lemma 1,


<i>f(ex) Af(e2) A • • • A / ( e J = Det(C)(^ Ae2 A • • • Aen). (5.3.5)</i>


<i>Since ex A e2 A • • • A en is a base for En</i> (F) we see, from (5.3.2), that Det(/) =


Det(C). Thus the determinant of / equals the determinant of any of its
<i>representative matrices. Again if g: F—• F is also an .R-endomorphism of F,</i>
<i>then from En(g° f) = En(g)°En(f) it follows that Det(#°/) =</i>


Det(gf) Det(/). When this is expressed in terms of representative matrices
<i>for / and g, we obtain, of course, the usual formula for the determinant of</i>
the product of two square matrices.


<i>Once again letel9e29... ,enbea. base of F, let / be an #-endomorphism</i>


<i>of F, and let f(er) be given by (5.3.3). Suppose that 1 < p < n. In what follows</i>


<i>we shall use K = (kl9 k2,..., kp) (respectively J= (j 1J2,... Jp)) to denote a</i>


<i>sequence of p integers satisfying 1 </c</i>x<i> <k2 < • • • < kp< n (respectively 1 <</i>


<i>h</i> <A<<i>' ' ' <jp<n), and K' (respectively J') will stand for the sequence</i>


<i>obtained from (1, 2, 3 , . . . , n) by deleting the terms that belong to K</i>
(respectively J).



<i>Putxi = f(ei),xK = xk i Axki</i> A<i> • • • Axk ,andej = eji</i> A ^ A - -1 A ^ - . T h e n ,


by Lemma 1,


where


<i><b>C</b><b>JK</b><b> =</b></i>


<i><b>L</b></i>


<i><b>J</b><b>2</b><b>k</b><b>l</b></i> <i><b>^J</b><b>2</b><b>k</b><b>2</b></i>


Similarly


<i>where T= (tl912,..., tp) and T are to be understood in the same way as K</i>


<i>and K'. We now have</i>


<i>A er )CJK CrK'</i>


<i><b>J,T</b></i>


</div>
<span class='text_page_counter'>(103)</span><div class='page_container' data-page=103>

<i>92 The exterior algebra of a module</i>


<i>because e3 A er = 0 unless J=T. Next, from (5.3.5) we obtain</i>


<i>= £KDet(C)(el A £ ?</i>2<i>A - " Aen),</i>


<i>where eK</i> is the sign of (fel5...,/cp,/c'l9/c2,...) considered as a permutation of



(1,2, 3 , . . . ,n); that is to say e^ is —1 raised to the power (/cx — 1) +


<i>(k2 — 2) + - • • +(/c</i>p<i> — p). We also have e, AeJ' = 8J(e1 Ae2</i> A • • • A 4 If we


<i>now substitute in (5.3.6) and remember that ex A e2 A • • • A en</i> is actually a


<i>of En(F), we find that</i>


Thus


Det(C) = X £ ^ C ^ Cr r (5.3.7)


<i>where eJK = SjSK</i> is — 1 raised to the power


O'I<i> +h + - ' +Jp) + (ki + k2 + -- + fc</i>p).


<i>This expression for Det(C) is known as Laplace's expansion of the</i>
<i>determinant of the matrix C using the columns of the matrix that are</i>
indexed by fcl9 fc2<i>,..., kp.</i>


We give one more illustration of the connection between the theory of
exterior algebras and the theory of determinants. To this end consider free
<i>/^-modules U, V and W of finite rank. We select bases for these, namely</i>
<i>ul9</i> t*2<i>,..., um for (7, vuv2,..., vn for V and w</i>l5 w2<i>, . . . , wq</i> for W Let


<i>X\ I/—• F and JU: K—> VF be R-linear mappings and let their matrices, with</i>
<i>respect to the chosen bases, be A = ||an\\ and B=\\bkj\\ respectively. Thus A</i>


<i>is an n x m matrix, B is a q x M matrix,</i>



and


<b>k = i</b>


<i>Furthermore the matrix C of /i° A is given by C = BA.</i>


After these preliminaries let / = (il5 i2<i>, . . . , ip) and J=(JiJ2,... ,7</i>P) be


sequences of p integers, where 1 < i\ < z2<i> < • • • < ip < m and 1 <j1 <j2 < • • •</i>


</div>
<span class='text_page_counter'>(104)</span><div class='page_container' data-page=104>

<i>The exterior algebra of a direct sum</i> 93


where


In a similar manner we can show that


<i>where K = (kl9 k2,..., kp) is a sequence of integers satisfying l<fex<</i>


<i>k2< • • • <kp<q,wK = wki</i> A wfe2 A • • • Awk<i> , and the definition of B^) mimics</i>


<i>that of AW. On the other hand</i>


<i>because C = BA is the matrix of (j, ° L But the wK form a base for Ep(W) and</i>


£p(^ o ^) = £p(^) o j E p ( / l)< Consequently


(5.3.8)


<i>We mention briefly that if we fix p and regard the A$ as entries in a</i>


<i>matrix Aip\ then A</i>(p)<i> is known as the p-th exterior power of the matrix A.</i>
<i>The relation (5.3.8) then yields (BA)ip) = Bip)Aip\ which is the usual formula</i>
<i>for the p-th exterior power of the product of two matrices.</i>


<b>5.4 The exterior algebra of a direct sum</b>


<i>Let Ml,M2,.. • ,Mq</i> be ^-modules and let £(Mt-) denote the


<i>exterior algebra of Mt. Then, using the construction set out in Section (3.4),</i>


<i>we can form the modified tensor product £(M</i>X<i>) ® E(M2) ® ''' ® E(Mq).</i>


This, of course, is an K-algebra and we know (Chapter 3, Theorem 7) that it
is anticommutative.


Next there is an R-linear mapping


given by


<i>E(M2)</i> <i>E(Mq)</i>


2<i>, . . . ,mq)=</i> <i><sub>m</sub></i>


<i>t</i>


(5.4.1)


<b>(5.4.2)</b>


<i>(Naturally in 1 đ ã ã ã đ mt ® • • • ® 1 it is understood that mi</i> occurs in the



<i>i-th position.) Note that (t>(M1 đ ã • • ® Mq) generates E{MX) ® • • • ® E(Mq)</i>


</div>
<span class='text_page_counter'>(105)</span><div class='page_container' data-page=105>

<i>94 The exterior algebra of a module</i>


<i><b>Theorem 8. The modified tensor product</b></i>


£(MX<i>) ®RE(M2) ®R-- ®RE(Mq), (5.4.3)</i>


<i>together with the R-linear mapping 4> defined in (5.4.2), constitutes the</i>
<i>exterior algebra of the direct sum M</i>x<i> â M2 đ '' * © Mq. Furthermore the</i>


<i>grading which (5.4.3) possesses by virtue of being a modified tensor product is</i>
<i>the same as its grading as the exterior algebra of M</i>x<i> © M2 © * * * © Mq.</i>


<i>Proof For the moment let us leave on one side the assertions about</i>
gradings. For the rest, the isomorphisms


(Mx<i> â ã ã ã â Mq_l) â Mq^M1 ®'"®Mq.1®Mq</i>


and (Chapter 3, Theorem 6)


<i>®"-® E(Mq_l) ® E{Mq)</i>


show that the first part of the theorem will follow if we can establish it when
<i>q = 2.</i>


<i>Suppose then that U and V are ^-modules, put M=U ®V, and define</i>
<i>the K-linear mapping 0 : M —» E(U) ® E(V) by 0(w, v) = u ® 1 + 1 ® v. We</i>
<i>have to show that E(U) ® E(V) and 0 constitute the exterior algebra of M.</i>


<i>As a first step we note that, since (0(w, v))2 = 0 for all UGU and veV,</i>
<i>Theorem 2 shows that 0 extends to a homomorphism X\E(M)-+</i>
<i>E(U) ® E(V) of tf-algebras. Naturally X(u, v) = u ® 1 + 1 ® v.</i>


<i>Next the inclusion mappings U-+M and V-+M extend to </i>
<i>homo-morphisms h: E(U)^E(M) and k: E(V)^E(M) of graded K-algebras.</i>
<i>Let us consider the mapping E(U) x E(V)—> E(M) in which (x, y) becomes</i>
<i>h(x) Ak(y). This is a bilinear mapping of jR-modules. From this, and</i>
<i>because E(U) ® E(V) and E(U) ® E(V) coincide as K-modules, it follows</i>
that there exists an JR-linear mapping


<i>where A'(x ® y) = h(x) A k(y) and X\u ® 1 + 1 ® v) = (u, v) for u in U and v in</i>
<i>V.</i>


<i>Let x e Er(U), y e ES(V\ f e Ep(U) and n e Ea{V). By (3.4.5), the product of</i>


<i>x ® y and f ® n in E(U) ® E(V) is ( - l)sp(x A^)®(yArj) and the image of</i>
<i>this under X' is</i>


<i>But h and k preserve degrees and E(M) is anticommutative. Consequently</i>
the image in question is


<i>(h(x) A k(y)) A (fc({) A k(ri)) = X'(x®y)A X'{^ ® rj).</i>


</div>
<span class='text_page_counter'>(106)</span><div class='page_container' data-page=106>

<i>Covariant extension of an exterior algebra 95</i>


<i>Consider X' ° X. This is an algebra-homomorphism of E(M) into itself and</i>
<i>it induces the identity mapping on M. Consequently X' ° X is the identity</i>
<i>mapping of E(M). Next X'°X is the identity mapping of E(U)®E(V)</i>
<i>because it induces the identity mapping on (j)(M) and we know that, as an</i>


<i>algebra, E(U) ® E(V) is generated by 0(M). Hence X is an isomorphism of</i>
<i>algebras and therefore E(U) ® E(V), together with (f>, is the exterior algebra</i>
of M. As already observed, the first part of the theorem follows in full
generality.


<i>Finally, let us consider the two gradings on E(M1)®E(M2)®'"</i>


<i>® E(Mq) that are mentioned in the statement of the theorem. In both cases</i>


<i><l>{Ml © • • • © Mq) is the module of elements of degree one. But</i>


<i>(j)(M1 â ã ã ã â Mq) generates the algebra and therefore the two gradings</i>


must coincide.


<b>5.5 Covariant extension of an exterior algebra</b>


<i>Let M be an ^-module and a>: R—>S a homomorphism of</i>
commutative rings. Since we are here concerned with more than one
<i>commutative ring, we shall, for greater definiteness, use ER(M) to denote</i>


<i>the exterior algebra of M. Of course ER(M) is a graded /^-algebra, and</i>


<i>therefore (see Section (3.6)) ER(M) ®RS is an algebra graded by the </i>


<i>S-modules {En(M) ®R S}nel. In particular M ®R S is the module of elements</i>


<i>of degree one. It is easily verified that, as an S-algebra, ER(M) ®RS is</i>


<i>generated by M ®R S and that, in ER (M) ®R</i> 5, the square of an element of



M (g)<i>R S is zero. It follows, from Lemma 3 of Chapter 3, that ER</i> (M) <g)<i>R S is</i>


anticommutative.


<i>By Theorem 2, the inclusion mapping M đR Sã ER (M) ®R S extends to</i>


a homomorphism


<i>X:ES(M®RS)^ER(M)®RS</i>


<i>of 5-algebras. (Here ES(M ®RS) denotes the exterior algebra of the </i>


<i>S-module M ®R S.) Note that X is degree-preserving.</i>


<i><b>Theorem 9. The mapping</b></i>


<i>X:ES(M®RS)^ER(M)®RS,</i>


<i>described above, is an isomorphism of graded S-algebras.</i>


<i>Proof There is a close similarity between this result and Theorem 6 of</i>
Chapter 4. It will be found that it is a simple matter to adapt the arguments
used to prove the earlier result, so the details have been omitted.


<i>In view of Theorem 9 we may regard ER (M) ®R S as the exterior algebra</i>


<i>of the S-module M ®R S. We shall therefore write</i>


</div>
<span class='text_page_counter'>(107)</span><div class='page_container' data-page=107>

<i>96 The exterior algebra of a module</i>



<i>and, because k is an isomorphism of graded algebras, we can add to this</i>
<i>En(M®RS) = En(M)®RS. (5.5.2)</i>


<i>The latter relation holds for every integer n.</i>


<b>5.6 Skew derivations on an exterior algebra</b>


<i>The linear mappings of a module M into the ring R form the </i>
<i>R-module UomR (M, R). We put</i>


<i>M* = HomR(M,R). (5.6.1)</i>


<i>The members of M* are, of course, the linear forms on M. M* itself is known</i>
<i>as the dual of M.</i>


We recall that the structural homomorphism of the exterior algebra
<i>E(M) maps R isomorphically onto E0(M). This fact will be used throughout</i>


Section (5.6) to identify the two so that in what follows we have


<i>E0(M) = R. (5.6.2)</i>


<i>Now let A be a skew derivation on E(M). Then A induces an </i>
<i>R-homomorphism El{M)^>E0(M). But El(M) = M and E0(M) = R. </i>


Con-sequently A gives rise to a linear form on M. Furthermore, by Lemma 6 of
Chapter 3, different skew derivations determine different linear forms.


<i><b>Theorem 10. Let f be a linear form on the R-module M. Then there is one</b></i>



<i>and only one skew derivation on E(M) which extends f</i>


<i>Proof In view of what has already been said it will suffice to produce a skew</i>
<i>derivation that agrees with f on E1(M) = M. Suppose then that p > 1 is an</i>


<i>integer and consider the mapping of M x M x • • • x M (p factors) into</i>
<i>which takes (ml, m</i>2<i>, . . . , mp) into</i>


<i><b>p</b></i>


<i>£ (—\)l*lf(mi)ml A • • • Arhi A • • • /\mp.</i>


This mapping is multilinear and alternating. Consequently there is induced
<i>a homomorphism Ap: Ep(M)^>Ep__l(M) of K-modules with the property</i>


that


<i>A ^</i> A m2 A • • • A<i> mp)</i>


<i><b>t (5.6.3)</b></i>



<i>Note that when p = 1, Ap</i> coincides with /.


<i>Next, the various Ap (p = 1,2, 3,...) are the restrictions of a single </i>


R-endomorphism A: £(M)—>£(M) that has degree — 1 ; and now it only
remains for us to verify that A is a skew derivation.


</div>
<span class='text_page_counter'>(108)</span><div class='page_container' data-page=108>

<i>Skew derivations on an exterior algebra 97</i>



<i>A(x) A y + (— l)rx A A(y) and for this we may assume that r>\ and s>\.</i>
<i>Indeed it is enough to establish the relation in question when x = m1 A</i>


<i>m2 A • • • A mr and y = fix A \I2 A • • • A \IS, where the mt</i> and ^ are elements of


M. But then, by (5.6.3),
<b>r</b>


<i>A(x Ay)= £ (—l)l + 1f(mi)m1 A • • • A ^ A • • • Amr A / ^ A • • • A/HS</i>


<i><b>A • • • AfljA- ' ' Afi</b><b>s</b></i>


and with this the proof is complete.


Suppose now that / is a linear form on M. By Theorem 10, it has a unique
<i>extension, Ay say, to a skew derivation on E(M). We have seen that the skew</i>
<i>derivations on E(M) form an /^-module and it is evident that</i>


and


<i>Arf = rAf</i> (5.6.5)


for / i , /2<i>e M * and reR. Furthermore, by Lemma 6 of Chapter 3,</i>


A/° A/ = 0. (5.6.6)


Before we start to examine the consequences of these relations it will be
convenient to broaden the basis of the discussion. To this end we note first
that if



<i>yo:M*xM-+R</i>


<i>is defined by yo(f, m) = f(m), then y0 is bilinear and the mapping M^>R in</i>


<i>which m becomes yo(f, m) is just /.</i>


<i>Suppose now that U is an K-module and that we have a bilinear mapping</i>
<i>y:UxM-*R, (5.6.7)</i>


<i>i.e. suppose that we have a bilinear form on U x M . If we fix u in U, then</i>
<i>there is a linear form on M that maps m into y(u, m). This linear form will</i>
have a unique extension, Au<i> say, to a skew derivation on E(M). Thus</i>


<i>Au(m) = y(u,m) (5.6.8)</i>


<i>for all meM. Furthermore (5.6.4), (5.6.5) and (5.6.6) generalize to give</i>


AUl + M2 = AU i+AU 2, (5.6.9)


Aru = rAu, (5.6.10)


Au°Au = 0, (5.6.11)


</div>
<span class='text_page_counter'>(109)</span><div class='page_container' data-page=109>

<i>98 The exterior algebra of a module</i>


<i>U->EndR(E(M)) (5.6.12)</i>


<i>in which u has image A</i>M<i>. Note that if q > 1 and m1,m2,..., mq</i> are in M,



then, by (3.7.4) and (5.6.8),


<b>Au(m!</b> A m<i><b>2</b> A<b> • • •</b> A<b> m</b><b>q</b><b>)</b></i>
<i><b>q</b></i>


<i>= £ (—l)l + iy(u,mi)m1 A - • - Arhi A - • - Amq</i> (5.6.13)
<i><b>i= 1</b></i>


<i>for all u in U. But A</i>U°AU = O. Accordingly (5.6.12) extends to a


homo-morphism


<i>F: E(U)-+ End^ (E(M)) (5.6.14)</i>
<i>of K-algebras. Note that if ul9 u2,..., u belong to U, then</i>


and therefore


F ^<i> AU2 A - • •</i> AW/,) = AM I° AM 2° - •
< O


AV (5.6.15)


This shows that when F(w1<i> A U2 A • • • A up) operates on a homogeneous</i>


<i>element of E(M) it lowers its degree by p. We also see that for p>2</i>
F(w1<i> A u2 A • - • A up) = AUi ° T(u2 A u3 A - - - A up). (5.6.16)</i>


We now wish to make explicit the way in which F(w1<i> A U2 A • • • A up)</i>


<i>operates on E(M) and for this it is convenient to introduce some temporary</i>


notation.


<i>Suppose that \<p<q. We shall use T=(tl,t2,... ,tp) to denote a</i>


<i>sequence of p integers satisfying I<t1<t2<- - • <tp<q; and if</i>


<i>m1, m2,..., mq belong to M, then {mx Am2 A • • • Amq)T</i> will denote the


<i>result of striking out from m1 A m2 A • • • A mq the terms mh, mh,..., mt</i> .


Thus


<b>(mt Am2</b><i><b> A • • • Am</b><b>q</b><b>)</b><b>T</b></i>


<i><b>= m</b><b>x</b><b> A • - - Arh</b><b>ti</b><b> A - - - Am</b><b>t</b><b> A • • • Am</b><b>q</b><b>. (5.6.17)</b></i>


<i>(If q = p, then (m</i>x A m2<i> A • • • A mq</i>)<i> T</i> is understood to be the identity element


<i>of E(M).) Finally we put</i>


<i><b>y(u</b><b>l9</b><b>m</b><b>ti</b><b>) y(u</b><b>l9</b><b>m</b><b>t2</b><b>) ••• y(u</b><b>l9</b><b>m</b><b>t</b></i>


and


DT =


l ^ l = ^i


(5.6.18)



<b>' • + f n</b>


<i><b>Lemma 2. Suppose that \<p<q and let u</b>t,u2,... ,up belong to U and</i>


</div>
<span class='text_page_counter'>(110)</span><div class='page_container' data-page=110>

<i>Skew derivations on an exterior algebra</i> 99


<i><b>A u</b><b>2</b></i><b> A • • • A Mp))(mj A m2 A • • • A m9)</b>


= (-1)"X ( - D ' ^ r K Am2 A • • • Am,)r.


<i><b>T</b></i>


<i>Proof. We argue by induction on p. If u e U, then</i>


(5.6.19)


<i><b>Am</b><b>2</b></i><b> A • • •</b> <i><b>Am</b><b>2</b></i><b> A • • •</b>


A • • • A<i> m{ A</i> • • • A<i> mn</i>


<i>by (5.6.13). This shows that (5.6.19) holds when p= 1.</i>


<i>We now assume that p > 1 and that the relation in question has been</i>
established for all smaller values of the inductive variable. Then, by the
inductive hypothesis,


<i>{T(u2</i> A M3<i> A • • • A up))(m</i>x<i> A m2 A - - - Amq)</i>


= ( - Dp-1I ( - l ) "c |^ ( m1A m2A - - - A m , )K. (5.6.20)



<i><b>K</b></i>


<i>Here the sequence K = (/c</i>2<i>, k3,..., kp) is required to satisfy 1 < k2 < k3 < • • •</i>


<i><kp<q and</i>


<b>K<sub>~ :</sub></b>


<i>p</i> <i>2 y(up,mk3) ••• y(up,mkp)</i>


Next, if we operate with AMi on (5.6.20) and use (5.6.16) we find that


<i>(Ht/j AU2 A" ' A M p ) ) ^ Am</i>2<i> A ' ' ' Atnq)</i>


)i,). (5.6.21)


<i>We must now identify the coefficient of (m1 A m2 A • • • A mq)T</i> when we


<i>expand the right-hand side of (5.6.21). To this end, for 1 < v <p we put Kv =</i>


<i>( * ! , . . . , £ , , . . . , tp). Then when A</i>Mi<i>((m! A m2 A • • • A mq)K) is developed it</i>


provides the term


y(t/1,mfv)(-l)fv+v(m1 A m2<i> A - - - A mq)T.</i>


<i>We also have \KV \ + tv = \T\. Hence the coefficient of (m1 A m2 A • • • A mq)T</i>


in the expansion of the right-hand side of (5.6.21) is



<b>(-I)'"1</b><i><b> t (-D</b></i><b>m+vy(«i^rv)D5,v.</b>


</div>
<span class='text_page_counter'>(111)</span><div class='page_container' data-page=111>

<i>100 The exterior algebra of a module</i>


However, this follows on expanding the determinant (5.6.18) by means of its
first row.


<i><b>Corollary. Suppose that p > 1 . Let w</b></i>l5<i> u2,..., upbelongto U andmi,m2,...,</i>


<i>mp to M. Then</i>


<i>y(ul9m2) ••• y(ul9mp)</i>


<i>y(up9</i> mx<i>) y(up,m2) ••• y(up,mp)</i>


<i>Proof. If we take account of (5.6.15) this is simply what (5.6.19) becomes</i>
<i>when p = q.</i>


<b>5.7 Pfaffians</b>


Let M be an K-module and let


<i>y:MxM-+R (5.7.1)</i>


<i>be an alternating bilinear mapping o f M x M into the ring R (i.e. y is an</i>
<i>alternating bilinear form). Furthermore for meM let A</i>m denote the skew


<i>derivation on E(M) which satisfies</i>


Am(/*) = y(m, ju) (5.7.2)



<i>when pi belongs to M. (Note that we are continuing to identify E0(M) with</i>


<i>R.) Then if m</i>1? m2 e M and r e K, we have Ami +m2 = Ami + Am2 and Arm = rAm.


Next for m G M we define an endomorphism


<i><b>L</b><b>m</b><b>:E{M)-+E(M) (5.7.3)</b></i>


by


<i>Lm(x) = mAx. (5.7.4)</i>


<i>Note that Lmi +m2 = Lmi</i> + Lm2<i>, Lrm = rLm</i> and, whereas Am has degree - 1,


<i>the degree of Lm</i> is + 1 . We now put


Am = Lm + Am. (5.7.5)


Of course Am<i> is also an endomorphism of the R-module E(M).</i>


<i><b>Lemma 3. The mapping</b></i>


M->EndR(£(M)) (5.7.6)


<i>in which m of M becomes A</i>m<i> is R-linear. Furthermore A</i>m° Am<i> = 0 for all</i>


<i>meM.</i>


<i>Proof The first assertion is clear. Now let xeE(M). Then (Lm°Lm)(x) =</i>



</div>
<span class='text_page_counter'>(112)</span><div class='page_container' data-page=112>

<i>Pfaffians 101</i>


<i>But y(m, m) = 0 because y is an alternating bilinear form. Consequently</i>


and therefore Am°Lm + Lm° Am = 0. The lemma follows.


By Theorem 2, the mapping (5.7.6) extends to a homomorphism


£(M) -»End,, (£(M)) (5.7.7)


<i>of/^-algebras. Let us denote the image of an element x of E(M) by A</i>x. Then


Am i.m 2...m p=(Lm,+Am i)o(Lm 2 + Am2)o---o(Lmp + Amp) (5.7.8)


whenever ml5 m2, . . . , mp belong to M.


We next define


<i>Q:E(M)->E(M) (5.7.9)</i>


by


Q(x) = Ax(l£(M)). (5.7.10)


<i>Clearly Q is an endomorphism of the ^-module E(M) and Q(1</i>£(M)) = 1£(M).


<i>It follows that Q induces the identity mapping on E0(M). Furthermore if</i>


<i>and ueE(M), then</i>



and hence


Q(w)). (5.7.11)


<i><b>Theorem 11. Let Q be defined as above. Then for every skew derivation A on</b></i>


<i>E(M) we have Q ° A = A ° Q.</i>


<i>Proof. Suppose that n > 0. It will suffice to prove that if xeEn(M) and A is a</i>


<i>skew derivation, then Q(A(x)) = A(Q(x)). This will be established by</i>


<i>induction on n.</i>


First we note that if x e £0( M ) , then Q(A(x)) and A(Q(x)) are both zero.


<i>Thus all is well when n = 0.</i>


<i>From here on we assume that n > 1 and that the assertion in question has</i>
been established for all smaller non-negative values of the inductive
variable. Accordingly we have Q(A'(z)) = A'(O(z)) if A' is a skew derivation


<i>and zeEn_l(M).</i>


<i>Now let x GEn(M) and let A be a skew derivation. We wish to show that</i>


Q(A(x)) and A(Q(x)) are the same and for this step we may assume that x =


<i>niAy, where meM and yeEn_l(M). Now A(m) belongs to E0(M) = R.</i>


Consequently


Q(A(x)) = Q(A(m)j;-mAA(j;))


<i>= A(m)Q(y) - m A Q(A(y)) - (A</i>m


</div>
<span class='text_page_counter'>(113)</span><div class='page_container' data-page=113>

<i>102 The exterior algebra of a module</i>


= A(mAQ(y)+Am(Q(>;)))


<i>= A(m)fi(y) - m A A(Q(y)) + (A ° A J (Q(y)).</i>


But A(Q(y)) = Q(A(y)) by the inductive hypothesis and A ° Am = — Am ° A by


Lemma 6 of Chapter 3. Accordingly


= Q(A(x)).
The lemma follows.


We now introduce a second alternating bilinear form


<i>y:MxM-+R (5.7.12)</i>


by putting


<i>y(mum2)= -y(mi,m2) (5.7.13)</i>


and using this we define Am<i>, Ax</i> and Q just as we defined Aw, Ax and Q using


<i>y. By (5.7.13)</i>



Am= - Am. (5.7.14)


<i>Of course, like Q, Q is an endomorphism of E(M) and it commutes with all</i>
<i>skew derivations. Suppose that msM and ueE(M). Then, in view of</i>
(5.7.14) and (5.7.11), we have


Q(mAu) = mAQ(u)-Am(Q(u)). (5.7.15)


We also have, this time from (5.7.8),


A.l A.2 A...A m p=(Lm i-Am i)o(Lm 2-Am 2)o---o(Lm p-Am p) (5.7.16)


for ml5 m2<i>, . . . , mp</i> in M.


<i><b>Theorem 12. Q and Q ar^ inverse automorphisms of the R-module E(M).</b></i>


<i>Proof. We shall establish the following statement by induction on n: for all</i>
<i>n>0,ifxeEn(M), then Q(Q(x)) = x andQ(Q(x)) = x. This will clearly suffice.</i>


The statement is true when n = 0 because both Q and Q induce the
<i>identity mapping on E0{M). From here on we shall suppose that n > 1 and</i>


make the natural inductive hypothesis.


<i>Let xeEn(M). In showing that Q(Q(x)) = x we may suppose that x =</i>


<i>mAy, where meM and yeEn_1(M). But then, by (5.7.11) and (5.7.15),</i>


<i>Q(Q(x)) = Q{m A Q(y) + AJQ(y))}</i>



<i>= m A Q(Q(>;)) - A</i>m(Q(Q(>;))) + Q(Am(Q(>;)))


<i><b>= m Ay</b></i>


</div>
<span class='text_page_counter'>(114)</span><div class='page_container' data-page=114>

<i>Pfaffians 103</i>



<i>because Q(Sl(y)) = y by the inductive hypothesis and because Q commutes</i>
with Am. Since fi(Q(x)) = x for entirely similar reasons, the proof is


complete.


<i><b>Theorem 13. Let m</b>l, m</i>2<i>, . . . , mp (p > 1) belong to M. Then</i>


<i>y(mum2) ••• y(mump)</i>


<i>= co\ (5.7.17)</i>
<i>y(m2, mx) y(m2, m2) -m y(m2, mp)</i>


<i>y(mp, mx) y(mp, m2) • • • y(mp, mp)</i>


<i>where w is the component of degree zero of the result of applying</i>


(Lmi +Am,)° (Lm2 + Am2<i>)°- • - (Lm+\m)</i>


<i>to \E(MY V P *</i>s<i> °dd, then the determinant and co are both zero.</i>


<i>Remark. Theorem 13 is of interest not only because it shows that the</i>
determinant is a perfect square but also because it identifies its square root.
<i>Proof. For u in E(M) let no(u) denote its homogeneous component of</i>



<i>degree zero. Let x = ml Am2 A • • • Amp. Then the corollary to Lemma 2</i>


shows that the determinant in (5.7.17) is equal to


Put x = Q(x). By Theorem 12, x = Q(x) and therefore
Ax(x) = Ax(Q(x)) = Ax(Ax-(l£(M))


= (AX°AX-)(1£(M))
= AX A X-(1£ ( M )) = Q ( X A X ) .


But x = Q(x) is the result of applying


(Lm i-Am i)°(Lm 2-Am 2)o---°(Lm p-Am p)


to 1£(M)<i> (see (5.7.16)) so that x belongs to the ^-module spanned by all</i>


<i>products mix A mi2 A • • • A mih. However, x A mti A mi2 A • • • A mih = 0 if h >0</i>


<i>and thus we see that x AX = TZO(X)X. Consequently Ax(x) = n0(x)Q(x) and</i>


therefore the determinant in (5.7.17) is equal to ( - l)p ( p + 3)/27r0(x)7T0(Q(x)).


<i>On the other hand a> = no(Q(x)) so that it follows that the relation (5.7.17)</i>


will be established if we show that


<i>(_ l)P(r + Wn0(x) = n0(Q(x)). (5.7.18)</i>


Next, by (5.7.8),



Ax = (Lmi +AMi)o (L + Am, ) o . . .o ( L + A
<i><b>p</b></i>


</div>
<span class='text_page_counter'>(115)</span><div class='page_container' data-page=115>

<i>104 The exterior algebra of a module</i>


where Ap_2s,x<i> is an endomorphism of E(M) of degree p — 2s, and now we</i>


may conclude, from (5.7.16), that
<i><b>_ p</b></i>


Ax= X (-1)SAP
<b>-2s,x-s=0</b>


Accordingly


) =<sub> A</sub>


X( 1£ ( M )<i>) — YJ Ap_2</i>s,x(l£(M))
<b>s = 0</b>


and


<i>Thus if /? is odd, then both 7i</i>o(x) = 0 and 7io(Q(x)) = 0 and therefore, in this


<i>case, both the determinant and co are zero. On the other hand, ifp = 2k then</i>


= (-l)f c and therefore
But (~



as required.


We shall now extract from Theorem 13 some information which is
directly relevant to the theory of matrices. To this end let


<i>A =</i>


<b>«11 fl</b>


<b>12</b>


<b>C?2i ^22</b>


<b>fli</b>


a<i><b>PI</b></i>


<i>be a p x p matrix with entries in R. We call A an alternating matrix provided</i>
<i>that (i) au = 0 for all i, and (ii) ajt= —atj for all i and j .</i>


<i>Suppose then that A is an alternating matrix. We can construct a free </i>
<i>R-module M with a base el9e2,-..,ep of p elements and afterwards an</i>


<i>alternating bilinear form y: M xM^>R with y(ei,eJ) = aij. This enables us</i>


<i>to introduce the Pfaffian of A, which will be denoted by Pf(^4). In fact the</i>
Pfaffian is defined by


where 7r0<i>: E(M)-+E0(M) is the natural projection. It follows (Theorem 13)</i>



that


(5.7.19)
<i>and when p is odd P((A) is zero.</i>


</div>
<span class='text_page_counter'>(116)</span><div class='page_container' data-page=116>

<i>Comments and exercises</i> 105


<i>A. =</i>


<i>12</i>


<i>— a</i>


<i>12</i>


<b>"13 "14</b>


023 024


0 fl34


" 0 3 4 0


<i>In this case Pf (A) is the component of degree zero of</i>


<i>But Le</i> has degree + 1 and Ae has degree — 1. Consequently


=


023014-024013+012034-Thus the relation Det(A) = (Pf(/4))2 becomes



<i>- 0 i 2 0 a23</i>


- 0 1 3 - 0 2 3 0
- 0 1 4 - 0 2 4 - 0 3 4


as may be verified directly.


024
034
<b>0</b>


= (012034-013024+014023)2


<b>5.8 Comments and exercises</b>


The theory of exterior algebras is of particular interest because of
its connection with the theory of determinants, and indeed most of what will
be said here arises out of this connection. However, there are two comments
on the general theory that come to mind. These will be dealt with first.


As usual, the examples for which solutions are provided are marked with
<i>an asterisk. If M is an R-module, then E(M) - or ER</i> (M) when it is desired to


remind the reader of the ground ring in question - always denotes its
exterior algebra.


<i>First suppose that / is a proper ideal of R and that M is an R-module.</i>
<i>Then IE(M) is a homogeneous two-sided ideal of E(M) and therefore</i>
<i>E(M)/IE(M) inherits the structure of a graded K-algebra. But / annihilates</i>


<i>this algebra, so E(M)/IE(M) can also be regarded as an K/7-algebra.</i>


<i>Next, the natural mapping £(M)—• E(M)/IE(M) induces a </i>
<i>homo-morphism of El{M) = M into E(M)/IE(M) which vanishes on IM. Thus</i>


there is induced a homomorphism


<i>M/IM-+E(M)/IE(M). (5.8.1)</i>


</div>
<span class='text_page_counter'>(117)</span><div class='page_container' data-page=117>

<i>106 The exterior algebra of a module</i>


<i>>E{M)/IE(M)</i> (5.8.2)
of ^//-algebras.


<i><b>Exercise 1. Show that the homomorphism (5.8.2) is an isomorphism of graded</b></i>


<i>R/I-algebras.</i>


There is a solution to this exercise which is very similar to the solution
provided for Exercise 5 of Chapter 4. We shall therefore leave the reader to
make the minor adjustments that are required. Note that this result enables
us to write


<i>ER/l(M/IM) = ER(M)/IER(M).</i> <b>(5.8.3)</b>


Our next comment has to do with generalized skew derivations. We
<i>already know, from Theorem 10, that, for any R-module M, E(M) is fully</i>
endowed with skew derivations. The next exercise extends this result.


<i><b>Exercise 2. Let M be an R-module, let i be a given integer, and let f: M —•</b></i>



<i>Ei + l(M) be an R-linear mapping. Show that there is one and only one</i>


<i>generalized skew derivation of degree i, on £(M), that extends f</i>


This time we can adapt the solution to Exercise 6 of Chapter 4. The only
point to be noted is that if/? > 1, then the mapping of the p-fold product M x
<i>M x • • • x M into Ei + p(M) which takes (m</i>l9 m2 <i>, mp) into</i>


<i>is an alternating multilinear mapping. This is because E(M) is an</i>
anticommutative algebra (see Theorem 3).


Our next comments have to do with Laplace's expansion of a
deter-minant and we begin by recalling what has been established already.
Suppose that


<i><b>C =</b></i> <i>c22</i> <sub>(5.8.4)</sub>


<i>is an n x n matrix with entries in R and let p be an integer satisfying 1 < p < n.</i>
<i>In what follows J = (JiJ2, • • • Jp) respectively K = (kl9</i> / c2<i>, . . . , kp) will</i>


<i>denote a sequence of p integers with 1 <j\ < • • • <jp < n respectively 1 <</i>


<i>kx<- • <kp<n, and as before we put</i>


<i><b>C</b></i>


</div>
<span class='text_page_counter'>(118)</span><div class='page_container' data-page=118>

<i>Comments and exercises</i> 107


<i>Again J' respectively K' will be used to denote the increasing sequence that</i>


<i>is obtained by striking out from (1, 2 , . . . , « ) the terms that belong to J</i>
<i>respectively K. If now we put</i>


<i>\J\=h+J2 + '-+Jp> (5.8.6)</i>
then (5.3.7) can be rewritten as


<i>Det(C) = £ ( - iyJl + lKlCJKCfK>. (5.8.7)</i>


<i><b>j</b></i>


This, as we remarked at the time, is Laplace's expansion of the determinant
<i>of C using columns fc</i>l5<i> k2,..., kp. Now there is a companion formula,</i>


namely


Det(C) = Y (—1)'J| + 'K<i>'C C> ' (5 8 8)</i>
<i><b>K</b></i>


<i>which is Laplace's expansion using rows jl9j2,... Jp. This can be derived</i>


by applying (5.8.7) to the transpose of C.


The following observation will serve to introduce the next exercise.
Consider the 2 x 4 matrix


<i>X^ X2 X-$</i>


<i><b>y\ yi ^3</b></i>


<b>></b>



<b>x4</b>

<i><b>r</b></i>



<i><b>\yx</b></i>


<i>X2</i> X3


<b>=0.</b>


<i>X2</i> X4


<i><b>y</b><b>2</b><b> y</b><b>4</b></i>


There is an identity that connects its 2 x 2 subdeterminants which takes the
form


(5.8.9)


Using Laplace's expansion, (5.8.9) can be generalized as follows.


<i>Exercise 3*. Let A = | | a</i>l J<i>| | be an nx2n matrix with entries in R, let</i>
<i>/ = (i'i, i2,... ,in) be a typical sequence of n integers satisfying 1 < i i < • • •</i>


<i><in<2n, and let V be the sequence obtained from ( 1 , 2 , . . . , 2n) by deleting</i>


<i>the terms in I. If</i>


<i>D,= "</i>


<i>show that</i>



</div>
<span class='text_page_counter'>(119)</span><div class='page_container' data-page=119>

<i>108 The exterior algebra of a module</i>


<i>Show also that if k (where 1 <k<2n) is fixed and n is even, then</i>


' ) 0 / 0r = O. (5.8.11)


Here it is to be understood that the summation is over the different
sequences /, and that if/' = (i'i, i"2<i>,..., i'n), then by sgn(7, /') is meant the sign</i>


<i>of (il, i2,..., i'i, i</i>2<i>, . . . ) considered as a permutation of ( 1 , 2 , . . . , 2n). Of</i>


<i>course (5.8.9) is what (5.8.11) becomes when n = 2.</i>


We return to (5.8.8). Let S = (sl5 s2, . . . , sp<i>), where 1 <Sj < • • • <sp<n,</i>


<i>and let S' (the corresponding residual sequence) be (si, s</i>2<i>, . . . , s'n_p).</i>


<i>Further, let a n n x n matrix Q be formed as follows: the first p rows of Q are</i>
toberows;1 ?;2<i>>-.. Jp of the matrix C and the last n — p rows of Q are to be</i>


rows si, s2, . . . ,sj,_pofC. Evidently, if J = 0 ' i , ;2<i>, . . . Jp) and S are different,</i>


<i>then the determinant of Q is zero whereas if J = S, then</i>
Det(6) = sgn(J,J')Det(C)


and therefore


( - 1 )1 + 2 +<i>-+ p</i> Det(Q) = (~ l )| s | Det(C).
<i>On the other hand, for arbitrary J and S</i>



<i>as may be seen by employing Laplace's expansion on the first p rows of Q. It</i>
follows that


<i>Z</i>

<i>( nlsi + |Kir r</i>
<i>\—l</i>


<i>) ^JK^S'K'</i>
<i><b>K</b></i>


<i>is zero if J ^S and it has the value Det(C) if J = S. Hence if we put</i>
rKs = ( - l ) ' *l + | 5 |Q ' r (5.8.12)
then we obtain


t '

K A S

"lDet(C) if J = S.

(5>8

-

13)
The next exercise provides a relation that is similar to (5.8.13). To solve
the exercise we have merely to repeat the above argument, but this time
using Laplace's expansion in terms of columns rather than in terms of rows.


<i>Exercise 4. / / FKS is defined as in (5.8.12) show that</i>


<i>In the preceding discussion J, K and S denoted increasing sequences of p</i>
<i>integers all lying between 1 and n. The total number of such sequences is (</i>p).


Let us put these (p) sequences in some order - the particular way in which


</div>
<span class='text_page_counter'>(120)</span><div class='page_container' data-page=120>

<i><b>Comments and exercises 109</b></i>


<i>the entries in an (np) x (np) matrix Cip)</i> say. (This is the p-th exterior power of



<i>C.) Likewise the TJK can be regarded as the entries in another (J) x (np) matrix</i>


<i>Tip)</i> say. On this understanding (5.8.13) and (5.8.14) can be written in matrix
notation as


<i>C{p)r{p) = (Det(C))I = r{p)Cip\ (5.8.15)</i>
where / denotes the identity matrix with (J) rows and columns.


<i><b>Exercise 5*. Let C =11^-11 be an nxn matrix with entries in R and for</b></i>


<i>l<p<n let C{p) be the p-th exterior power of C. Show that</i>


<i>where by (r, s) is meant the binomial coefficient r \ / s \ (r — s)l.</i>


To be fair to the reader it should be remarked that the solution provided
for this exercise uses the fact that polynomials with integer coefficients form
a unique factorization domain and also the result that the general
determinant, considered as a polynomial in its entries, is irreducible.


We shall now interpret the result contained in Exercise 5 in terms of
<i>exterior algebras. To this end let F be a free K-module of rank n, and let</i>
<i>/ : F —• F be an endomorphism of F. Further, suppose that 1 < p < n. If now</i>
<i>el9</i> e2<i>, . . . , en is a base for F,f(er) = exclr + e2c2r+ • • • +encnr, where c^eR,</i>


and C=||co.||, then, of course, Det(/) = Det(C). Let J = 0'iJ2<i>> • • • JP\</i>


<i>where 1 <j^ < • • • <jp < n, and put e3 = eji A eJ2 A • • • A ejp. Then the e3</i> form


<i>a base for EP(F) and, as we remarked earlier,</i>



<i><b>J</b></i>


<i>But this means that, with respect to the base {ed} of Ep(F), the</i>


<i>endomorphism Ep{f) has matrix Cip). Consequently Det(£</i>p(/)) =


Det(C(p)). Exercise 5 now shows that


Det(£p(/)) = (Det(/))("~1'p"1 ) (5.8.16)


which is the relation we were seeking.


We end this section with some comments on Pfaffians. The simplest
alternating matrix of even order has the form


0 a"


<b>ro</b>


<b></b>



L-and the Pfaffian of this is a. At the end of Section (5.7) we calculated the
Pfaffian of the general 4 x 4 alternating matrix. The reader may like to verify
by the same method that the Pfaffian of


<b>0</b>


<i><b>-a</b><b>l2</b></i>


- ^ 1 4



-015


-016


012


<b>0</b>


- 0 2 3


- 0 2 4


- 0 2 5


- 0 2 6


013


023


<b>0</b>


- 0 3 4


- 0 3 5


- 0 3 6


014



024


034


<b>0</b>


- 0 4 5


- 0 4 6


015


025


035


045


<b>0</b>


- 0 5 6


</div>
<span class='text_page_counter'>(121)</span><div class='page_container' data-page=121>

<i>110 The exterior algebra of a module</i>


<b>is</b>


<i>+ a14(a23a56-a25a36+a26a35)</i>


<i>-a15(a23a46-a24a36+a26a34)</i>



<i>+ a16(a23a45-a24a35+a25a34).</i>


(5.8.17)


The general formula, of which this is a special case, will be derived later.
It will be noticed that the Pfaffian of an alternating matrix is a polynomial
<i>in its entries with integer coefficients. Let us make this more precise.</i>
<i>Consider the generic alternating matrix</i>


<i>0 xl2 x13 • - • xln</i>


<i>— x12 0 x23 - • • x2n</i>


<i>— x13 —x23 0 • • • x3n</i>


<i>— xln —x2n —x3n</i> • • • 0


<i>where the x(j (i<j) are n(n— l)/2 different indeterminants. We regard this</i>


matrix as having its entries in the polynomial ring Z[x1 2<i>, x13,..., x</i>n_l M]


<i>and then its Pfaffian is a polynomial, say Q(x12, x</i>1 3, . . . , xn_x J , with


<i>integer coefficients in the indeterminates in question. Now let R be an</i>
<i>arbitrary commutative ring and A = ||a</i>o<i>|| annxn alternating matrix with</i>


<i>entries in R. Then</i>


<i>with the same polynomial Q as before. This follows from the method given</i>
in Section (5.7) for calculating Pfaffians.



<i>For the next exercise we observe that if A and B are alternating matrices,</i>
then


is an alternating matrix as well.


<i><b>Exercise 6*. Let A and B be alternating matrices with entries in R. Show that</b></i>


<i>the Pfaffian of</i>


<i>~AA 0 1</i>

<b>o</b>

<i><b> B\</b></i>



<i>is the product of the Pfaffian of A and the Pfaffian of B.</i>


</div>
<span class='text_page_counter'>(122)</span><div class='page_container' data-page=122>

<i>Solutions to selected exercises 111</i>


<i>matrix, so that we can compare the Pfaffian of ATAA with that of A. The</i>
next exercise records the facts.


<i><b>E x e r c i s e 7*. Let A be an nxn alternating matrix and A an arbitrary nxn</b></i>


<i>matrix. Show that A</i>T<i>y4A is an alternating matrix and that</i>
Pf(ArylA)=(Det(A))Pf(A).


<i>Let Abeannxn(n>2)alternating matrix and suppose that 1 <i<j<n.</i>
If we strike out the i-th andj-th rows and also the i-th andj-th columns, then
the resulting matrix will also be alternating.


<i><b>Exercise 8*. Let A = || a</b>pq \\ beannxn(n>2) alternating matrix and, for 1 <</i>



<i>i <j < n, let %j be the matrix obtained from A by striking out the i-th andj-th</i>
<i>rows and columns. Show that</i>


<b>7=2</b>


This result gives a convenient way of evaluating Pfaffians. We
encoun-tered a special case of it in (5.8.17).


<i><b>Exercise 9. Let I</b>n be the identity matrix of order n. Show that the Pfaffian of</i>


<i>is equal to </i>


<b>(-5.9 Solutions to selected exercises</b>


<i><b>Exercise 3. Let A = \\a</b>ij\\ be an nx2n matrix with entries in R, let 1 =</i>


<i>(i1,i2,...,in) be a typical sequence of n integers satisfying l ^ i ^ " *</i>


<i><in<2n, and let I be the sequence obtained from ( 1 , 2 , . . . , 2n) by deleting</i>


<i>the terms of I. If</i>


<i><b>a</b><b><sub>2i</sub></b></i>
<i><b>n</b></i>


<i>show that</i>


Xsgn(/,/')D7/)r = 0.



<i>Show also that if k (where 1 <k<2n) is fixed and n is even, then</i>


</div>
<span class='text_page_counter'>(123)</span><div class='page_container' data-page=123>

<i>112 The exterior algebra of a module</i>


<i>Solution. Let C be the In x 2n matrix</i>


Then Det(C) = 0 because there are repeated rows. Hence if we use Laplace's
<i>expansion to evaluate the determinant (applying it to the first n rows) we</i>
find that


or


This is the first assertion. (Here we have used the fact that


which was noted previously.) Observe that
sgn(/, /') sgn(/\ /) = (—l)1/|-|-|/'l =
Consequently


<i>when n is even.</i>


<i>From here on we suppose that k (1 <k<2n) is a given integer which we</i>
<i>keep fixed, and we suppose also that n is even. Let /</i>l 5 J2, . . . , /s be the


<i>sequences / that contain k. Then those that remain are /'</i>l5 / '2<i>, . . . , I's</i> and we


have


<i>= t sgn(J,,/;)D</i>

/f

D

i;

<i>+ t sgn(r</i>

<i>v</i>

<i>r;)D</i>

<i>r</i>

<i>p</i>

<i>r;</i>


= 2 £ sgn(/

r

, /;)/),,D

;;


<b>f = 1</b>


because /'/ = /f<i> and sgn(/,, rt) = sgn(I't, It). Thus</i>


(5.9.1)


and now we have the problem of getting rid of the 2.
We proceed as follows. Let


<i>X =</i>


<b>*11</b><i><b> X</b></i>


<i><b>\2 ' ' '</b><b> X</b></i>


<i><b>\,2n</b></i>


<i><b>X</b></i>


<i><b>2\ * 2 2 ' ' '</b><b> X</b></i>


<i><b>2,2n</b></i>


<i><b>X</b></i>


<i><b>nl</b><b> X</b></i>


<i><b>n2</b></i> <i><b>X</b></i>



</div>
<span class='text_page_counter'>(124)</span><div class='page_container' data-page=124>

<i>Solutions to selected exercises</i> 113


<i>where the xtj are indeterminates, and consider X as a matrix with entries in</i>


Z [ xn, x1 2<i>, • • •, xn,in\- If</i> n o w A7<i> denotes the subdeterminant of order n that</i>


<i>comes from columns iu i2,..., in, then (5.9.1) shows that</i>


<i>kel</i>


but now, because 2 is not a zerodivisor in the polynomial ring, we may
conclude that


<b>Xsgn(/,/')A7Ar = 0.</b> (5.9.2)


Finally we observe that there is a ring-homomorphism of Z [ xn, x1 2<i>, - •.,</i>


<i>xn2n\ into R in which m (in Z) is mapped into m\R</i> and x0<i>- is mapped into atj.</i>


If we apply this homomorphism to (5.9.2) we obtain the required relation. A
convenient way to describe the last step is to say that


<i>kel</i>


<i>is obtained from (5.9.2) by specialization.</i>


<i><b>Exercise 5. Let 0=11^11 be an nxn matrix with entries in R and for</b></i>


<i>\<p<n let C</i>(p)



<i> be the p-th exterior power of C. Show that</i>


<i>where by (r, s) is meant the binomial coefficient r\/s\ (s — r)\.</i>
<i>Solution. Let</i>


<i>X =</i>


<i>X</i>


<i>21</i>


<i>X<sub>nl</sub> X<sub>n2</sub></i>


<i>2n</i>


<i>where the xtj are indeterminates and X is regarded as a matrix with entries</i>


in Z[x11? x1 2<i>, . . . , xnn~\. We shall show that</i>


Once this has been proved the desired result will follow by applying the
ring-homomorphism


<i>in which xtj</i> is mapped into ctJ.


<i>By (5.8.15), we can find a square matrix U (say), of the same size as Xip\</i>
<i>such that (i) the entries in U belong to Z [ x</i>n, x1 2, . . . , xn n<i>], and (ii) X{P)U is</i>


</div>
<span class='text_page_counter'>(125)</span><div class='page_container' data-page=125>

<i><b>114 The exterior algebra of a module</b></i>


<b>and therefore Det(Ar(p)</b><i><b>) is a factor, in the polynomial ring, of (Det(X))</b><b>in</b><b>></b><b>p</b><b>\</b></i>



<i><b>But Det(X) is irreducible in Z\_x</b>l 1?</i> x1 2<i>, . . . , xnn<b>~\ and + 1 are the only units</b></i>


<b>in this ring. We now see that</b>


<i>where k is an integer and s is either + 1 or — 1. By comparing degrees we</i>
find at once that


<i>k</i>


<i>(p-l)\(n-p)\'</i>


<i>To determine £ we give xl l9</i> x2 2, • • • ><i> xnn ^</i>e<i> value 1 and all the other xtj</i> the


<i>value zero. Then X specializes to the identity matrix /„ and we obtain e =</i>
Det(/ip))-<i> B u t I(np) i s i t s e l f a n</i> identity matrix so that e = + 1 .


<i><b>Exercise 6. Let A and B be alternating matrices with entries in R. Show that</b></i>


<i>the Pfaffian of</i>


<i>B_, (5-9.3)</i>


<i><b>is the product of the Pfaffian of A and the Pfaffian of B.</b></i>


<i>Solution. Let A have p rows and p columns and let B have q rows and q</i>
<i>columns. Put n = p-\-q. Next, let M be a free K-module with a base el9 e2,</i>


<i>..., en and let y: M xM—+R be the alternating bilinear form in which</i>



<i>\\y{eh</i> £y)|| is the matrix (5.9.3). Then, with the notation of Section (5.7), the


desired Pfaffian is the component of degree zero belonging to


Now


for suitable elements xl 9 x2<i>, . . •, xq in E(M). It is therefore enough to show</i>


that


<i>has 0 as its component of degree zero. But this is clear because if 1 < i <p,</i>
<i>then Ae(ep+j) = 0fovj= 1, 2 , . . . , q, whereas operating with Le</i> replaces an


<i>element of the form ep + 1 AXX<b> + • • • +e</b><b>p</b><b>+</b><b>q</b> Axq</i> by another element of the


same form.


<i><b>Exercise 7. Let A be an nxn alternating matrix and let A be an arbitrary</b></i>


</div>
<span class='text_page_counter'>(126)</span><div class='page_container' data-page=126>

<i>Solutions to selected exercises</i> 115


<i>Solution. Let X be the alternating matrix</i>


0 x1 2 x1 3
<b> n </b>


<b>v-12 ^ 2 3</b>
<b>— X13 ~ ^ 2 3 0</b>


<i>— xln —x2n —x3n</i>



<i>xln</i>
<i><b>X</b><b><sub>2n</sub></b></i>


and let


<i>Y =</i>


<i><b>y</b><b>nn</b></i>


<i>Here the xtj {i<j) and the y,</i>a<i>, are distinct indeterminates and X and Y are</i>


regarded as matrices with entries in the ring formed by adjoining all these
indeterminates to the ring of integers. Now


<i>(YTXY)T= YTXTY= - (YTXY)</i>


<i>and so (because we are in characteristic zero) YTXY is an alternating</i>
<i>matrix. Note that it will suffice to show that Pf(YTXY) = (Det(Y)) Pf(X)</i>
because the general result will then follow by specialization. But


= (Det(Y))2Det(X)
= ((Det(Y))Pf(X))2


<i>so either Pf(YTXY) = (Det(Y)) Pf(X) or else Pf(YTXY) =</i>


<i>— (Det(7)) Pf (X). However, we have only to specialize Y so that it becomes</i>
the identity matrix to see that it is the former of these which is correct.


<i><b>Exercise 8. Let A = \\a</b>pq \\ beannxn(n> 2) alternating matrix and, for 1 <</i>



<i>i <j < n, let %j be the matrix obtained from A by striking out the i-th andj-th</i>
<i>rows and columns. Show that</i>


<i>Solution. It is convenient to deal with the problem in its generic form, that is</i>
we consider the matrix


<i>X =</i>


0 x1 2


<i>~xl</i>


</div>
<span class='text_page_counter'>(127)</span><div class='page_container' data-page=127>

<i>116 The exterior algebra of a module</i>


<i>where the xtj (i<j) are indeterminates and X is regarded as having its</i>


<i>entries in Z[xl 2</i>, xx 3<i>, . . . , xn</i> _ x<i> „]. Let Etj</i> be the matrix obtained by striking


<i>out the f-th and j-ih rows and columns. Then it will suffice to prove that</i>


<i><b>j=2</b></i>


since all other cases will follow by specialization. Note that we can suppose
<i>that n is even for otherwise Pf(X) and the Pf(S</i>l7) will all be zero thus


making the assertion trivial.


<i>Suppose that 1 < r < n. The indeterminates x</i>A/1<i> (X < /i) with r e (A, fi) occur</i>



only in the r-th row and the r-th column. Each term of Det(X) is of degree 2
in these particular indeterminates, and therefore, since Det(X)= (Pf(X))2,
<i>each term of Pf (X) will contain exactly one x^t with k < \i and r e (A, JX). It</i>


follows that


<i>where Uj is a polynomial not involving any indeterminate xpq (p < q) with</i>


<i>either 1 or j in (p, q).</i>


<i>Suppose that 2 <j<n. Replace xpq (p < q) by zero if exactly one of 1 and;</i>


<i>is in (p, q) and replace xXj</i> by 1. (All the other indeterminates are to remain


<i>unaltered.) Then X becomes a new alternating matrix, Xj say, and Uj is</i>
<i>unchanged. We therefore have Uj = Pf(Xj).</i>


<i>Let A be the matrix obtained from the n x n identity matrix by taking its</i>
;-th column and moving it step by step to the left until it becomes the second


<i>column. Then Det(A) = ( - l)j</i> and
0 1 0 ••• 0
- 1 0 0 ••• 0


0 0


Consequently (see Exercises 6 and 7)
Pf(S1<;<i>.) = Pf(ATXjA) = (Det(A))</i>


<i>whence Uj=(—l)jPf(Elj). Accordingly</i>



<i><b>3=2</b></i>


</div>
<span class='text_page_counter'>(128)</span><div class='page_container' data-page=128>

<b>6</b>



The symmetric algebra of a module



<b>General remarks</b>


We have examined the origins and properties of the tensor and
<i>exterior algebras of a module. The symmetric algebra, which concerns us</i>
now, fits into the same general pattern and, as in the other cases, we shall
motivate its investigation by means of an appropriate universal problem.
However, from a different standpoint, one can say that the exterior algebra
<i>can be obtained from the tensor algebra by making it anticommutative.</i>
Viewed from this position, the symmetric algebra is the result of making the
<i>tensor algebra commutative.</i>


Because much of the theory of the symmetric algebra runs closely parallel
to that of the exterior algebra, we can frequently borrow proofs and adapt
them without difficulty, and when this is the case we shall restrict the
amount of detail that is given.


There is, however, one area in this account of symmetric algebras which is
not foreshadowed in Chapter 5. The symmetric algebra of a module can be
regarded as a generalization of a polynomial ring; and for a polynomial
ring the partial differentiations with respect to the indeterminates generate
a commutative algebra. When this observation is analysed it leads naturally
<i>to the algebra of differential operators and this is one of the topics we shall</i>
discuss. In the next chapter this particular algebra will be examined from a


more general standpoint and, as a result, we shall be able to complete the
analogy between exterior and symmetric algebras.


The general conventions that apply to this chapter are the same as those
set out in the General Remarks at the commencement of Chapter 5 with one
<i>variation. Where we have to consider another commutative ring besides R</i>
<i>we shall denote the second ring by R' rather than by S. This is because the</i>
<i>letter S is widely used as part of the standard notation for symmetric</i>
algebras and symmetric powers.


</div>
<span class='text_page_counter'>(129)</span><div class='page_container' data-page=129>

<i>118 The symmetric algebra of a module</i>


<b>6.1 The symmetric algebra</b>


<i>Let M be an .R-module, A an K-algebra, and suppose that we are</i>
<i>given an K-linear mapping (ft: M-^A which is such that</i>


<i><b>(ft(m)(ft(m</b></i>

<i><b>f</b></i>

<i><b>) = (ft(m')(ft(m) (6.1.1)</b></i>



<i>for all m and m' in M. A natural universal problem connected with this</i>
situation can be posed as follows.


<i><b>Problem. To choose A and (ft so that given any R-linear mapping \ft: M^>B</b></i>


<i>(B is an R-algebra) such that ^/(m)\ft{m') = \jj(m')\j/(m) for all m, m! in M, there</i>
<i>shall exist a unique homomorphism h:A-+B, of R-algebras, such that</i>
<i>h°(ft = \jj.</i>


This type of problem will, by now, have become familiar. Clearly the
<i>problem has at most one solution in the sense that if (A, (ft) and (A', (ft') both</i>


<i>meet the requirements, then there exist inverse isomorphisms k: A —> A' and</i>
<i>k': A —>A,of ^-algebras, such that k° (ft = (ftf and k' ° (ft' = (ft. As for the rest</i>
the essential facts are contained in the following theorem.


<i><b>Theorem 1. The universal problem just enunciated possesses a solution.</b></i>


<i>Furthermore if (A, (ft) is any solution, then</i>


<i>(i) (ft: M—+A is an injection;</i>


<i>(ii) (ft(M) generates A as an R-algebra;</i>


<i>(iii) there is a, necessarily unique, algebra-grading {An}n£Z on A with</i>


<i>A,=(ft{M);</i>


<i>(iv) for each p>\, Ap is the p-th symmetric power of M, with</i>


<i>mlm2 . . . mp = (j)(ml)(ft(m2)... (ft(mp) for elements m</i>1? m2<i>, . . . , mp</i>


<i>in M;</i>


<i>(v) the structural homomorphism R^>A maps R isomorphically onto</i>
<i>Ao.</i>


<i>Remark. This result should be compared with Theorem 1 of Chapter 5. As</i>
will be seen the proofs are very similar.


<i>Proof Because any two solutions of the universal problem are copies of</i>
<i>each other, it is sufficient to find one solution that satisfies all of (i)-(v).</i>



<i>Suppose that m, m! e M. Then m (g) m' — m! ® m is a homogeneous</i>
<i>element of degree two in the tensor algebra T(M), and therefore such</i>
<i>elements generate a homogeneous two-sided ideal, H(M) say, in T(M). Put</i>
<i>Hp(M) = H{M) n Tp{M) and A = T(M)/H(M). Of course A has a natural</i>


<i>structure as a graded .R-algebra; let {An}nel</i> be the grading which it inherits


<i>from T(M). Then, in the natural homomorphism T(M)—+A of graded</i>
<i>algebras, Tn(M) is mapped onto An. In particular, since M = TX(M), the</i>


</div>
<span class='text_page_counter'>(130)</span><div class='page_container' data-page=130>

<i>The symmetric algebra 119</i>


<i>that, because m®m' — m' ®m belongs to H(M), its image in A is zero and</i>
therefore $(m)0(m') = 0(m')0(m).


<i>It is a simple matter to check that A and 0 satisfy conditions (i), (ii), (iii)</i>
and (v) in the statement of the theorem. To see that (iv) also holds suppose
<i>that p>\. Then the surjective homomorphism Tp(M)-^Ap</i> induced by


<i>T(M)-+A has kernel Tp(M) nH(M) = Hp(M). Furthermore Hp(M) is the</i>


<i>submodule of Tp(M) generated by all elements of the form</i>


<i>mx (x) • • • (x) mt ® mi</i> +<i> x (x) • • • (x) mp</i>


<i>- m1 ® • • • <g) mi +! (x) m</i>t- <g) • •


<i>where the second term is the same as the first except that mt and mt +</i> x have



<i>been interchanged. That Ap</i> is the p-th symmetric power of M now follows


from Theorem 8 of Chapter 1. Thus (iv) has been verified as well.


<i>All that remains of the proof is for us to verify that (A, <f>) solves the</i>
<i>universal problem. Suppose then that \j/: M —> B is an K-linear mapping, of</i>
<i>M into an K-algebra B, and that \jj(m)\f/(m') = \jj(m')\lj(m) for all m, m' in M.</i>
<i>The extension of \\i to an algebra-homomorphism T{M)^>B vanishes on</i>
<i>H(M) and so induces a homomorphism h: A-^B of K-algebras. Clearly h</i>
<i>satisfies h ° (f) = \\J ; and, since </>(M) generates A as an algebra, it is the only</i>
<i>homomorphism of A into B to do so. This ends the proof.</i>


<i>We next introduce the standard notation and terminology. Let (A, 0) be</i>
a solution of our universal problem. We put


<i>S(M) = A (6.1.2)</i>


<i>and we employ {Sn(M)}nel</i> to denote the grading referred to in Theorem 1.


<i>Also we use the fact that <f> maps M isomorphically onto Al =Sl(M) to</i>


make the identification


<i>M = SX{M\ (6.1.3)</i>


<i>and then M generates S(M) as an R-algebra. Of course the grading on S(M)</i>
is non-negative.


<i>We shall use ordinary juxtaposition to denote multiplication in S(M) so</i>
<i>that, because <}){m)(j)(m') = (f){m')(j)(m), we have from (6.1.3)</i>



<i>mm'= m'm (6.1.4)</i>
<i>for m, m' in M. Next, when p > 1, Sp(M) is the p-th symmetric power of M so</i>


<i>naturally we call S(M) the symmetric algebra of M. Finally Theorem 1 tells</i>
<i>us that the structural homomorphism R^>S(M) maps R isomorphically</i>
<i>onto S0(M). It follows that we may make the identification S0{M) = R</i>


whenever it is convenient to do so.


<i><b>Theorem 2. Let S(M) be the symmetric algebra of M and let \jj: M —> B be an</b></i>


</div>
<span class='text_page_counter'>(131)</span><div class='page_container' data-page=131>

<i>120 The symmetric algebra of a module</i>


<i><b>for all m, m' in M, then i// has a unique extension to a homomorphism</b></i>


<i><b>S{M)-^B of R-algebras.</b></i>


<i>Of course Theorem 2 holds because of the connection between S{M) and</i>
the universal problem with which we started.


The proof of Theorem 1 reveals the relation of the symmetric algebra to
the tensor algebra. However, we shall now describe this relation in a less
encumbered context. To this end observe that the inclusion mapping
<i>M-+S(M) extends to a surjective homomorphism</i>


<i><b>T(M)->S(M) (6.1.5)</b></i>


<i>of g r a d e d R - a l g e b r a s , a n d s o for e a c h peZ t h e r e is i n d u c e d a surjective</i>
h o m o m o r p h i s m



<i><b>T</b><b>p</b><b>(M)^S</b><b>p</b><b>(M) (6.1.6)</b></i>


<i>of R-modules. If /?>1, then the mapping (6.1.6) takes ml®m2®''' ®mp</i>


<i>into m1m2... m</i>p and therefore it is none other than the canonical


homomorphism which we first met in Section (1.5). Thus (6.1.5) is effectively
a combination of canonical homomorphisms.


The next result is the counterpart of Theorem 3 of Chapter 5.
<i><b>Theorem 3. The symmetric algebra S(M) is commutative.</b></i>


<i>This follows from (6.1.4) and the fact that M generates S(M) as an </i>
R-algebra.


<b>6.2 Functorial properties</b>


The functorial properties of the symmetric algebra now follow in a
<i>straightforward manner. Thus if f:M—»iV is a homomorphism of </i>
<i>R-modules, then N = SX (N) by (6.1.3) and therefore (see Theorem 2) / can be</i>


extended to a homomorphism


<i>S(f):S(M)-+S(N) (6.2.1)</i>
<i>of R-algebras. But S(f) preserves the degrees of homogeneous elements.</i>
<i>Consequently for each integer p there is induced an K-linear mapping</i>


<i><b>S</b><b>p</b><b>(f):S</b><b>p</b><b>(M)-+S</b><b>p</b><b>(N). (6.2.2)</b></i>



<i>Note that Sl(f) = f and that if we make the identifications S0(M) = R =</i>


<i>S0{N), then S0(f) is the identity mapping of R. Note too that if p > 1 and</i>


<i>ml, ra</i>2<i>,..., mp</i> belong to M, then


<i><b>S</b><b>p</b><b>(f)(m</b><b>1</b><b>m</b><b>2</b><b> . . . m</b><b>p</b><b>) = f(m</b><b>l</b><b>)f(m</b><b>2</b><b>)... f(m</b><b>p</b><b>). (6.2.3)</b></i>


<i>Of course, if / is the identity mapping of M, then S{f) is the identity</i>
<i>mapping of S(M). Also if / : M —> N and g: N —• K are homomorphisms of</i>
K-modules, then


</div>
<span class='text_page_counter'>(132)</span><div class='page_container' data-page=132>

<i>The symmetric algebra of a direct sum 121</i>


<i>and hence, for all p e Z,</i>


<i>Sp(g°f) = Sp(g)oSp(f). (6.2.5)</i>


<i>Thus S(M) can be regarded as a covariant functor from ^-modules to</i>
<i>commutative graded ^-algebras. Naturally if / is an isomorphism of </i>
<i>R-modules, then S(f) is an isomorphism of graded ^-algebras and</i>


<i>1</i>


<i>). (6.2.6)</i>


<b>6.3 The symmetric algebra of a free module</b>


<i>Throughout Section (6.3) we shall assume that the ring R is </i>
non-trivial.



<i>Let M be a free K-module with a base B. If p>l, then (Chapter 1,</i>
<i>Theorem 9) Sp(M) is a free module and it has a base consisting of the</i>


<i>different products blb2... bp, where b</i>feB. We also know that two such


<i>products, say b1b2...bp and b\b'2...bp, are equal if and only if</i>


<i>(b'l9 b2,..., b'p) is a rearrangement of (fo</i>1?<i> b2,..., bp). Of course 5</i>0(M) is a


<i>free K-module having the identity element of S(M) as a base.</i>


<i>The above remarks show that S(M) itself, considered as an K-module, is</i>
<i>free and has as a base the set of all distinct products blb2 . . . bn9 where bt eB</i>


<i>and n > 0 ; moreover two products blb2... bn</i> and frifr'2<i> . . . b'q</i> are the same if


<i>and only if n = q and (b\, b'2,..., b'q) is a rearrangement of (b</i>l9<i> b2,..., bn).</i>


<i>Since the multiplicative structure of S(M) is determined by the way the basis</i>
elements are to be multiplied, it follows that


<i>S(M) = R[Bl (6.3.1)</i>


<i>Here # [ £ ] denotes the ordinary ring of polynomials, with coefficients in R,</i>
<i>in which the elements of B are treated as distinct commuting</i>
indeterminates.


<i><b>Theorem 4. Let P be a projective module. Then S(P), considered as an </b></i>



<i>R-module, is also projective. Hence, for each neZ, Sn(P) is a projective </i>


<i>R-module.</i>


<i>Proof This result resembles Theorem 5 of Chapter 4, and there should be</i>
no difficulty in adapting the proof of the earlier theorem to meet the needs of
the present one.


<b>6.4 The symmetric algebra of a direct sum</b>


<i><b>If A</b>l9 A2,..., Aq are commutative .R-algebras, then A1 ® A2</i> ® • • •


<i>® Aq is also commutative. Hence if Ml9</i> M2<i>, . . . , Mq</i> are ^-modules, then


<i>S(MX) đ S(M2) đ * ã' đ S(Mq) (6.4.1)</i>


</div>
<span class='text_page_counter'>(133)</span><div class='page_container' data-page=133>

<i>122 The symmetric algebra of a module</i>


the ordinary and not the modified tensor product.) The elements of degree
one in this algebra form the module


<i>M1®R®--®R-\-R®M2®--®R</i>


<i>+ '-+R<g>R®-'®Mq, (6.4.2)</i>


<i>where each term in the sum contains q factors; moreover this module</i>
<i>generates S(Af J ® S{M2) ®''' ® S{Mq) as an K-algebra.</i>


Consider the K-linear mapping



<i>đS(M2) đ ã ã ã đ S(Mq), (6.4.3)</i>


where


<i><b>q</b></i>


<b>(/>(m1,m2,... , mp) = Ê 1 đ ã ã ã ® mt- ® • • • ® 1 (6.4.4)</b>
<b>;= I</b>


<i>it being understood that in 1 đ ã ã ã đ mt đ ã ã ã đ 1 the factor mt</i> occurs in the


i-th position.


<i><b>Theorem 5. The tensor product</b></i>


<b>S(Af</b> 2<i>) ®RS(M2) ®R--- ®RS(Mq) (6.4.5)</i>


<i>together with the R-linear mapping</i>


<i><t>: Mx ®M2 ® • • • 0 Mq-^S(M1)®R" • ®RS(Mq)</i>


<i>constitute the symmetric algebra of M1đ M2 đ ã ã ã â Mq. Furthermore the</i>


<i>grading which (6.4.5) possesses by virtue of being a tensor product of graded</i>
<i>algebras is the same as its grading when considered as the symmetric algebra</i>
<i>of M1®M2®---®Mq.</i>


<i>Proof This result should be compared with Theorem 8 of Chapter 5 and</i>
once again it is a simple matter to adapt the proof of the corresponding
result for exterior algebras. For instance it is possible to reduce the general


<i>result to the case where q = 2. The only point which possibly deserves</i>
mention is that whereas in Chapter 5 we used the isomorphism


<i>đ E(M2) đ ã ã • ® E(Mq)</i>


this time all we need is the less sophisticated


<i><b>*S(M</b><b>l</b><b>)®S(M</b><b>2</b><b>)®--®S(M</b><b>q</b><b>)</b></i>


which, of course, is a special case of (3.3.13).


<b>6.5 Covariant extension of a symmetric algebra</b>


</div>
<span class='text_page_counter'>(134)</span><div class='page_container' data-page=134>

<i>Derivations on a symmetric algebra 123</i>


<i>denoted by SR(M). We note that SR(M) ®RRf is a commutative K'-algebra</i>


<i>and that it is graded by the family {Sn(M) ®R R'}neI</i> of K'-submodules. The


<i>homogeneous elements of degree one in SR(M) ®RRf</i> form the covariant


<i>extension M ®RR' of M, and M ®RR' generates SR(M) ®RR' as an </i>


R-algebra.


<i>Let SR> (M ®R R) denote the symmetric algebra of M ®R Rf</i> considered as


<i>an ^'-module. Then, because SR(M) ®RRf</i> is commutative, the inclusion


<i>mapping M ®RR'—>SR(M) ®RR' extends to a homomorphism</i>



<i>X: SAM ®RR')^SR(M) ®RR' (6.5.1)</i>


of .R'-algebras which evidently preserves degrees.


<i><b>Theorem 6. With the above notation</b></i>


<i>X\ SAM ®RR')^SR(M) ®RR'</i>


<i>is an isomorphism of graded R'-algebras.</i>


We leave to the reader the straightforward task of adapting the proof of
Theorem 6 of Chapter 4. Finally we observe that Theorem 6 allows us to
make the identifications


<i><b>S</b><b>R</b><b>. (M ®</b><b>R</b><b>R') = S</b><b>R</b><b> (M) ®</b><b>R</b><b> R' (6.5.2)</b></i>


<i>and (for all neZ)</i>


<i>RR'. (6.5.3)</i>


<b>6.6 Derivations on a symmetric algebra</b>


<i>Let M be an K-module and let us identify the submodule S0(M) of</i>


<i>the symmetric algebra with R. If now D is a derivation on 5(M), then D</i>
induces a homomorphism of Sx<i> (M) = M into S0(M) = R, so that in effect D</i>


extends a linear form on M.



<i><b>Theorem 7. Let f be a linear form on M. Then there is one and only one</b></i>


<i>derivation on S(M) that extends f</i>


<i>Proof Lemma 4 of Chapter 3 shows that at most one derivation extends /</i>
<i>Now suppose that p > 1. There is a symmetric multilinear mapping, of the </i>
<i>p-fold product M x M x • • • x M into Sp-l(M), in which (m</i>l9 m2<i>, . . . , mp)</i>


becomes


<i><b>p</b></i>


<i><b>£ f{m</b><b>i</b><b>)m</b><b>l</b><b>...m</b><b>i</b><b>...m</b><b>p</b><b>,</b></i>


the A<i> over mt</i> indicating, as usual, that this term is to be omitted. It follows


<i>that there exists an K-homomorphism Dp: Sp(M)-^Sp_1(M) that satisfies</i>
<i><b>p</b></i>


</div>
<span class='text_page_counter'>(135)</span><div class='page_container' data-page=135>

<i>124 The symmetric algebra of a module</i>


<i>Next, by combining DX,D2,D3,... we obtain an K-endomorphism</i>


<i>D: S(M)^S(M), of degree - 1 , that agrees with Dp on SP(M). A simple</i>


<i>verification now shows that D is a derivation. It extends / because D</i>x = / .


<b>6.7 Differential operators</b>


<i>The notion of a differential operator is one that we have not</i>


encountered before. It occurs naturally in connection with polynomial
rings and it can be extended readily to symmetric algebras. We shall
therefore investigate it here.


<i>Let M be an K-module and h SL non-negative integer. A differential</i>
<i>operator of degree h, on 5(M), is an K-linear mapping cj): S(M)—• S(M) of</i>
<i>degree — h with a certain additional property. To describe the property it</i>
will be useful to introduce some notation.


<i>Suppose that p >h and that mu</i> m2<i>, . . . , mp</i> belong to M. Let / = (j1? i2,


<i>. . . , ih) be a sequence of h integers with 1 <i</i>x<i> < i2 < • • • < ih <p, and let /' be</i>


<i>the increasing sequence obtained from ( 1 , 2 , . . . , p) by deleting the terms in</i>
/. Finally put


<i>mI = mimii...mih</i> (6.7.1)


<i>and define mr</i> similarly.


<i>On this understanding an #-endomorphism <p: S(M)—>5(M) of degree</i>
<i>— h is called a differential operator of degree h provided that</i>


<i>^ ( m ^ . . . mp) = YJ (t>(mi)^i' (6.7.2)</i>


<i>for all p >h and arbitrary elements m</i>1? m2<i>, . . . , mp in M. It is clear that the</i>


<i>differential operators of a given degree h form an K-submodule of</i>
<i>End*(S(M)). This submodule will be denoted by VhS(M). Of course,</i>



V0<i>S(M) consists of the endomorphisms of S(M) produced by multiplication</i>


<i>by the various elements of R. An explanation of the name 'differential</i>
operator' will be given in Section (6.8). Note that a differential operator of
<i>degree 1 is identical with a derivation as defined in Section (3.7).</i>


There is a second form of (6.7.2) to which it is convenient to draw
<i>attention. Suppose that xi9 x2,..., xq</i> are elements of M and that sl5 s2, . . . ,


<i>sq are non-negative integers with s1+s2 + m" +sq>h. If we replace</i>


<i>mlm2.. .mp</i> by xjxx22... x j , then (6.7.2) becomes


<i><b>= 1 (l</b></i>

<i><b>1</b></i>

<i><b>) • • • M*!</b></i>

<b>1</b>

<b>""</b>

<b>1</b>

<b>*?""</b>

<b>2</b>

<i><b> • • • K</b></i>

<i><b>q</b></i>

<i><b>~</b></i>

<i><b>aq</b></i>

<i><b>> (</b></i>

<b>6</b>

<b>-</b>

<b>7</b>

<b>-</b>

<b>3</b>

<b>)</b>



where the summation is over all sequences (a1? a2<i>, . . . , aq) of integers with</i>


</div>
<span class='text_page_counter'>(136)</span><div class='page_container' data-page=136>

<i>Differential operators 125</i>


and


- " - + a , = /i. (6.7.5)
<i>Note that we have allowed the possibility that some of sx</i>, s2<i>, . . . , sq</i> may be


zero and that this involves the use of appropriate conventions to cover this
situation.


<i><b>Lemma 1. Let h>0 and k>0 be integers, and let <f> and \jj be differential</b></i>


<i>operators, on S(M), of degrees h and k respectively. Then i// ° 4> = </> ° \f/ and</i>


<i>this is a differential operator of degree h + k.</i>


<i>Proof Suppose that p>h + k and that ra</i>x, m2, . . . , mp belong to M. In what


<i>follows / = (il9 i2,..., ih),J = (JiJ2> • • • Jk) and L = ( /</i>l 9<i>l2, . . . , lh+k) denote</i>


<i>strictly increasing sequences of integers, between 1 and p, whose lengths are</i>
<i>h, k and h + k respectively. We shall use (/, J) to denote the sequence</i>
<i>obtained from (il9 i2, . . . , jl 9 j2, . . . ) by arranging the different integers</i>


<i>present so that they form an increasing sequence. We defined / ' above. L is</i>
defined similarly.


<i>It is clear that \\t ° <j> and (j) ° \jj are endomorphisms of degree —(h + k), and</i>
from (6.7.2) we see that


<i>^. (6.7.6)</i>


<i><b>= L</b></i><b> J</b>


<i>This shows that \j/ ° <j> = <f> ° ij/. Again</i>


<i>and now we see that \// ° c/> is a differential operator.</i>


<i>A differential operator of degree h {h >0) induces a linear form on Sh(M)</i>


<i>and (6.7.2) shows that if we know h and the associated linear form on Sh(M),</i>


then the differential operator is fully determined. The next theorem shows
that the linear form may be prescribed arbitrarily.



<i><b>Theorem 8. Let h (h >0) be a given integer, let M be an R-module, and let f</b></i>


<i>be a linear form on Sh(M). Then there is exactly one differential operator of</i>


<i>degree h, on S(M), which extends f</i>


<i>Proof Suppose that p > h and consider the mapping of the p-fold product</i>
<i>M x M x • • • x M into Sp_h(M) in which (m</i>l5 m2<i>, . . . , mp) becomes</i>


(Here / = (il9<i> i2,..., ih) satisfies 1 <i1 < • • • <ih <p and ml and mr</i> have the


</div>
<span class='text_page_counter'>(137)</span><div class='page_container' data-page=137>

<i>126 The symmetric algebra of a module</i>


with the property that


<i>Thus far we have assumed that p > h. For 0 <p < h we take 4>p</i> to be the null


<i>homomorphism of Sp(M) into Sp-h(M).</i>


We can now combine 0O, (/>1?<i> (j>2,... so as to obtain an endomorphism</i>


<i>c/>:S(M)->S(M) of degree -h which agrees with <\>q on Sq(M). By</i>


<i>construction, (/> is a differential operator of degree h and it extends / Finally,</i>
we already know that there can be no other differential operator with these
properties.


<i><b>Corollary. Let h be a non-negative integer. Then the module W</b>hS(M) of</i>



<i>differential operators of degree h is isomorphic to the module of linear forms</i>
<i>on Sh{M) under an isomorphism which matches (j), in V</i>h<i>S(M), with its</i>


<i>restriction to Sh(M).</i>


<i>Of course WhS(M) is an K-submodule of End* (S(M)). Put</i>


V S ( M ) = £ V,S(M), (6.7.7)


<i><b>h>0</b></i>


where the sum is taken in EndK<i> (S(M)). It is easy to see that this sum is direct.</i>


Indeed in view of Lemma 1 we have


<i><b>Theorem 9. V5(M) is a commutative R-subalgebra of End</b>R(S(M)) and it is</i>


<i>non-negatively graded by its submodules {VhS(M)}h>0.</i>


<i><b>Definition. The graded, commutative algebra VS(M) is called the 'algebra of</b></i>


<i>differential operators' on S(M).</i>


In the next chapter we shall meet this algebra in a different guise.


<b>6.8 Comments and exercises</b>


We make some miscellaneous comments on the subject of
<i>symmetric algebras. As always, if M is an K-module, then S(M) - or SR(M) if</i>



we wish to be more explicit - denotes its symmetric algebra.


<i>Let us consider what happens to S(M) if we factor out an ideal of R.</i>
<i>Suppose then that / is an K-ideal. Then IS(M) is a homogeneous, two-sided</i>
<i>ideal of S(M) and hence S(M)/IS(M) is a graded algebra both with respect</i>
<i>to R and with respect to the ring R/I. The homomorphism of M = S</i>1(M)


<i>that is induced by the natural mapping S(M)—• S(M)/IS(M) vanishes on</i>
<i>IM and so there results an ^//-linear mapping</i>


<i><b>M/IM-*S(M)/IS(M). (6.8.1)</b></i>


</div>
<span class='text_page_counter'>(138)</span><div class='page_container' data-page=138>

<i>Comments and exercises 127</i>


to a homomorphism


<i>(j): SR/I (M/IM) - » S(M)/IS(M) (6.8.2)</i>


of ^//-algebras.


The next exercise should be compared with Exercise 1 of Chapter 5.


<i><b>Exercise 1. Show that the homomorphism (6.8.2) is an isomorphism of graded</b></i>


<i>R/I-algebras.</i>


This exercise shows that


<i>SR/I (M/IM) = SR (M)/ISR</i> (M). (6.8.3)



<i><b>Exercise 2. Let M be an R-module, let i be a given integer, and let f: M —•</b></i>


<i>Si + l(M) be an R-linear mapping. Show that there is one and only one</i>


<i><b>generalized derivation of degree i, on S(M), that extends f</b></i>


This, of course, is the counterpart of Exercise 2 of Chapter 5. It can be
solved in very much the same way.


We turn now to the consideration of differential operators and we begin
by indicating the origin of the name. First, however, it will be convenient to
introduce some additional notation.


Leta = (al9 a2, . . . , an)andj5 = (jS1, /?2<i>, •. •,/?„) be two sequences of n </i>


non-negative integers and let us put


and


<i><b>(P</b></i>

<sub>(6.8.5)</sub>


<i>We shall use a</? to mean that OL{ </?,- for f = 1 , 2 , . . . , n so that, on this</i>


<i><b>understanding, (?) = 0 whenever a $ /?. Lastly, a + /? and a — ft will denote the</b></i>
<i>sequences (a1</i> + / ?l 9. . . , an+/?n) and (oq — j 8l 9. . . , an — /?„) respectively.


<i>Now let M be a free K-module having Xx</i>, Z2<i>, . . . ,Xn</i> as a base. Then, by


<i>(6.3.1), S(Af) is the polynomial ring R[Xl9 X2,..., Xn~\ and, if p > 0 , 5</i>p(M)



<i>consists of all the homogeneous polynomials, i.e. forms, of degree p. It will</i>
<i>be convenient to denote the power-product XI1 X22... X^n by Xa. We can</i>


<i>then say that Sp{M) is a free K-module having the X* with |a| =p as a base.</i>


<i>Suppose next that </>: 5(M) —• S{M) is an /^-linear mapping of degree — h</i>
<i>(h>0). By (6.7.3) this will be a differential operator of degree h provided</i>
<i>that, whenever \/3\>h, we have</i>


</div>
<span class='text_page_counter'>(139)</span><div class='page_container' data-page=139>

<i>128 The symmetric algebra of a module</i>


<i>g\v\</i>
<i>dXy<sub>S dX</sub>y</i>


<i>22... dXyn»</i>


<i>is a differential operator of degree \y |. It is this fact which accounts for the</i>
terminology.


<i>Suppose now that/? >0. We know that Sp{M) has {X^} ^|</i>= pasabase. For


<i>\(x\=p= \p\, let D</i>(a)<i> be the linear form on Sp(M) which satisfies</i>


Then the D(a)<i> form a base for the linear forms on Sp(M). But, by Theorem 8,</i>


D(a)<i> has a unique extension to a differential operator of degree p on S(M); let</i>
<i>us use Di<x)</i> also to denote the differential operator. We now see not only that
<i>VPS(M) is a free K-module, but also that it has {D</i>(a)} N = p as a base. Note


<i>that for every sequence y = (y1,y2,- • •, y</i>n) of non-negative integers, we have



(0 ifa
This formula makes it clear that


<i>Exercise 3. With the above notation show that</i>


a


We turn now to other matters. Let (7 and M be K-modules and suppose
that we are given a bilinear form


<i>y:UxM-+R. (6.8.7)</i>


<i>For a fixed M, in £/, the mapping m\-+y(u,m) is a linear form on M and this, by</i>
Theorem 7, will have a unique extension, DM<i> say, to a derivation on S(M).</i>


Thus, by construction,


DB(m) = y(M,m) (6.8.8)


<i>for all raeM. Next, the mapping U^>EndR(S(M)) given by u\-^Du</i> is


<i>R-linear and, by Lemma 4 of Chapter 3, DUi° DU2 = DU2° DUi for all w</i>x, w2 m<i> U.</i>


It follows that there is a homomorphism


<i>r : S(U)^EndR{S(M)) (6.8.9)</i>


of K-algebras which has the property that



<i>T { ulu2. . . up) = DU x° DU 2° - - - ° DU p</i> (6.8.10)


<i>for uu u2,. • •, up in U. But, because a derivation is a differential operator of</i>


</div>
<span class='text_page_counter'>(140)</span><div class='page_container' data-page=140>

<i>Comments and exercises</i> 129


<i>(6.8.9) is a degree-preserving, algebra-homomorphism of S(U) into the algebra</i>
<i>of differential operators on S(M).</i>


Let us examine, in more detail, how this homomorphism operates. First
we recall that if


<i><b>C =</b></i>


<b>12</b>


<i><b>C</b><b><sub>n2</sub></b></i>


<i>• • * cm</i>


<i>is an n x n matrix, then its permanent, Per(C), is defined by</i>


<i><b>= 2 ^ C</b><b>ln{l)</b><b>C</b><b>2</b><b>n(2)' ' '</b><b> C</b><b>nn(n)></b></i>


<i>where n is a typical permutation of (1, 2 , . . . , n). The next exercise uses this</i>
<i>concept to describe the effect, on m1m2 ... mp, of the differential operator</i>


<i>associated with u1u2 ... up.</i>


<i><b>Exercise 4. Let u1</b>,u2,...,up belong to U and m1, m2,... ,mpto M. Show</i>



<i>that (DUi ° DU2 ° • • • ° Du )(m1m2 ... mp) is equal to the permanent of the</i>


<i>matrix</i>


<i>y ( u2, m2) ••• y ( u2, mp)</i>


</div>
<span class='text_page_counter'>(141)</span><div class='page_container' data-page=141>

<b>7</b>



Coalgebras and Hopf algebras



<b>General remarks</b>


<i>A coalgebra is a concept which is dual (in a sense that belongs to</i>
Category Theory) to that of an associative algebra, and consequently to
almost every result that we have concerning algebras there is a
corre-sponding result for coalgebras. Now it sometimes happens that an algebra
and a coalgebra are built on the same underlying set. When this occurs, and
provided the algebra and the coalgebra interact suitably, the result is called
<i>a Hopf algebra. Our concern with these matters stems from the fact that,</i>
<i>when M is an i?-module, both E(M) and S(M) are Hopf algebras of</i>
particular interest.


Whenever we have a coalgebra the linear forms on it can be considered as
<i>the elements of an algebra. The algebra which arises in this way from E(M)</i>
<i>is known as the Grassmann algebra of M; for S(M) the resulting algebra has</i>
very close connections with the algebra of differential operators which was
described in Chapter 6.


<i>Throughout this chapter we shall follow our usual practice of using R to</i>


denote a commutative ring with an identity element; and when the tensor
<i>symbol (x) is used it is understood that the underlying ring is always R</i>
unless there is an explicit statement indicating the contrary.


Finally, it happens to be convenient, before introducing the notion of a
coalgebra, to reformulate the definition of an ^-algebra. This reformulation
is carried out in Section (7.1).


<b>7.1 A fresh look at algebras</b>


<i>Let A be an associative K-algebra with an identity element. Then</i>
<i>inter alia A is an jR-module. Further, we have an K-linear mapping</i>
<i>\i\ A <g> A —> A which is such that /J.(X ® y) = xy and this, because </i>
<i>multiplica-tion on A is associative, makes</i>


</div>
<span class='text_page_counter'>(142)</span><div class='page_container' data-page=142>

<i>Afresh look at algebras 131</i>


<i>H®A</i>


<i>A® A® A • A ®A</i>


(7.1.1)


a commutative diagram.


<i>Now letn: R—>Abe the structural homomorphism of the algebra A. This</i>
<i>too is K-linear and, because rj(l) is the identity element of A, the diagram</i>


(7.1.2)



<i>R ® A • A® A < Al®R</i>


<i>also commutes. (Here R ® A ã A and A đ R ã A are the isomorphisms</i>
provided by Theorem 3 of Chapter 2.)


<i>At this point we make a fresh start. Suppose that A is an R-module and</i>
<i>that we are given K-linear mappings /i: A ® A —> A and n: R—+A which are</i>
<i>such that (7.1.1) and (7.1.2) are commutative diagrams. For x, y in A put</i>
<i>xy = fi(x ® y). Then it is easily verified that this definition of multiplication</i>
<i>turns A into an fl-algebra that has n(l) as its identity element and n: R —> A</i>
as its structural homomorphism.


When algebras are looked at in this way we shall say that the triple
<i>(A, \i, n) constitutes an associative ^-algebra with identity, and (in this</i>
<i>context) jx is called the multiplication mapping and n the unit mapping of the</i>
algebra. Let us now review the general aspects of the theory of algebras
from this new standpoint.


<i>Suppose that (A, fiA,nA) and (B, \iB, nB) are (associative) R-algebras and</i>


<i>let / be a mapping of A into B. It is readily checked that / is an </i>
algebra-homomorphism, in the sense of Section (3.1), if and only if the conditions
<i>(i) / is /Minear, (ii) fo^iA = (f®f)o^iB, and (iii) f°nA = nB</i> are all satisfied.


<i>Now assume that, for 1 <i<p, {At, nt, rjt) is an .R-algebra. Let</i>


<i>ã (zlj đ Ax) đ - - - đ (Ap ® Ap) (7.1.3)</i>


<i>be the isomorphism (of /^-modules) which matches (a đ ã ã ã đap) đ</i>



<i>(a\ đ '' - ® a'p) with [ai ® a\) ® • • • ® (ap ® a'p). Further, let</i>


<i>A(RP):R-+R ®R ®--- ®R, (7.1.4)</i>


</div>
<span class='text_page_counter'>(143)</span><div class='page_container' data-page=143>

<i>132 Coalgebras and Hopf algebras</i>


<i>Now, by Section (3.2), Ax ® A2 ®''' ®Ap</i> is an K-algebra. In fact its


multiplication and unit mappings are given by


<i>VAX ®-®AP=(V\ ®^2®-"®^P)°^P</i> (7.1.5)


and


<i>iAl®-®Ap=(rii®ri2®--'®rip)o&iRp)</i> (7.1.6)


respectively.


<i>We turn next to matters involving gradings. Suppose that (A, fi, rj) is an</i>
<i>K-algebra and let {An}neI be a family of #-submodules of A such that</i>


<i>A=YJAn</i> (d.s.), (7.1.7)


MeZ


<i>that is we assume that {An}neZ grades A as an R-module. Then this gives rise</i>


<i>to a grading {(A ®A)n}nel on the module A ®A, where</i>


<i>(A®A)n= X Ap®Aq</i> (d.s.). (7.1.8)



<i>(Of course (7.1.8) is the usual total grading on A ® A.) This said, if {An}neI</i>


<i>happens to be an algebra-grading, then the mappings fi: A ®A—+A and</i>
<i>rj: R —> A preserve degrees, it being understood that R is to be graded</i>
<i>trivially. Conversely, if \i: A ® A —• A and Y\ : R —* A are degree-preserving,</i>
<i>then {An}neI is an algebra-grading on A.</i>


<i>Suppose now that A{1\ A{2\ . . . , Aip) are graded K-algebras and let A(i)</i>
<i>have fit and rjt</i> as its multiplication and unit mappings. Further, let / =


<i>(iu i2,..., ip) and J = (jlJ2, - • • Jp) be sequences of p integers and put</i>


(7.1.9)


(7.1.10)
(see (3.4.2) and (3.4.3)). Then there is an isomorphism of


C41<i>* ® AV ®-"® A\f) ® (Aft ® Ag> đ'' ã đ Aff) (7.1.11)</i>
onto


<i>(Al? đ i4jj>) đ (A\? ®Af</i>

<i>2)</i>

<i>)®-"® (Ajf ® Aff) (7.1.12)</i>



in which, with a self-explanatory notation,


<i>K ®ai2®-"® aip) ® {a'h ®a'h®-'® a'jp)</i>


is mapped into


<i>e(I, J)(ati ® a'h) ® (ah ®a'h)®--® (aip ® a'jp).</i>



</div>
<span class='text_page_counter'>(144)</span><div class='page_container' data-page=144>

<i>Coalgebras 133</i>


(7.1.13)


<i>We recall, from Section (3.4), that the modified tensor product A(1)</i> đ
<i>A{2) đ ã ã ã đ Aip)</i> is a graded K-algebra. The multiplication and unit
mappings of this algebra are given by


AV>®- ® ^ = 0*i ® /*2 ® ã ' ' đ /*P) AP (7.1.14)


and


^( 1<i>> ® - ® ^ = (>7i ® ^2 ® • •' ® r\p<b>)° Ajf> (7.1.15)</b></i>


respectively, where Ajf} is the homomorphism previously encountered in
(7.1.4).


Now that we have described some of the basic features of algebras in
terms of mappings (rather than in terms of elements), we are ready to
introduce the dual theory which was mentioned briefly in the introduction
to this chapter.


<b>7.2 Coalgebras</b>


<i>Let A be an module and this time suppose that we are given </i>
<i>R-linear mappings A: A ã A đ A and e: A —• R. The triplet (A, A, e) is called</i>
<i>an R-coalgebra provided that the diagrams</i>


A



<i>A ã A đA</i>


<i>AđA (7.2.1)</i>


<i>AđA</i>


<i>A > A đAđA</i>


and


_<i> eđA</i>


<i>RđA < Ađ A ã A"đ R</i>


<b>(7.2.2)</b>


<i>are commutative. Here A ã A đ R maps a into a ® 1 and correspondingly</i>
<i>for A-+R® A.</i>


<i>If (A, A, e) is a coalgebra, then A: A ã A đ A is called the comultiplication</i>
<i>or diagonalization mapping of the coalgebra, and the fact that (7.2.1) is</i>
<i>commutative is described by saying that comultiplication is coassociative.</i>
<i>The mapping e: A—+R is known as the counit.</i>


</div>
<span class='text_page_counter'>(145)</span><div class='page_container' data-page=145>

<i>134 Coalgebras and Hopf algebras</i>


<i>take e: R —• R to be the identity mapping. Whenever we speak of R as being</i>
a coalgebra it is always this structure that we have in mind.



<i>Now suppose that (A, AA, sA) and (B, A</i>B<i>, eB) are coalgebras. A mapping</i>


<i>f:A—*B is called a homomorphism of coalgebras or a </i>
<i>coalgebra-homomorphism provided that (i) / is ^-linear, (ii) A</i>B° / = ( / ® / ) ° A ^ and


<i>(iii) eB ° / = eA. (We can describe conditions (ii) and (iii) by saying that / has</i>


to be compatible with comultiplication and it must preserve counits.) If a
homomorphism of coalgebras is also a bijection, then it is called an
<i>isomorphism of coalgebras and when this is the case the inverse mapping is</i>
<i>also a coalgebra-isomorphism. In the last paragraph we observed that R</i>
<i>itself is a coalgebra. Now if AR is the comultiplication of R and eR</i> is its


<i>counit, then sR ° BA = eA and AR°eA = (eA ®eA)° AA. (The latter follows from</i>


<i>the commutative property of (7.2.2).) Accordingly the counit mapping</i>
<i>sA: A—+R is a homomorphism of R-coalgebras.</i>


Finally, homomorphisms of coalgebras can be combined, that is to say
<i>that if / : A —> B and g: B —• C are coalgebra-homomorphisms, then g ° f is</i>
<i>a homomorphism of the coalgebra A into the coalgebra C.</i>


<b>7.3 Graded coalgebras</b>


<i>Let (A, A, e) be an /^-coalgebra and let {An}neZ be a grading on A</i>


considered, for the moment, simply as an R-module. Then there is induced
<i>on the module A ® A the usual total grading in which (A ® A)n</i> is given by


<i>(7.1.8). Let R be given the trivial grading.</i>



<i><b>Definition. We say that 'A is graded as a coalgebra by {A</b>n}ne^ or that</i>
<i>i</i>


<i>{An}neI is a coalgebra-grading on A' if A and e preserve the degrees of</i>


<i>homogeneous elements.</i>


<i>Thus, in the case of a graded coalgebra, the counit e maps all those</i>
<i>homogeneous elements of A whose degree is different from zero into the</i>
<i>zero element of R.</i>


Any coalgebra can be turned into a graded coalgebra by giving it the
trivial grading (i.e. all non-zero elements are treated as being homogeneous
<i>and of degree zero). When R itself is considered as a graded coalgebra, it is</i>
<i>always to be understood that the grading in question is the trivial one.</i>


<i>Now let A and B be graded coalgebras. By a homomorphism A-^B of</i>
<i>graded coalgebras is meant a coalgebra-homomorphism which preserves</i>
degrees. Naturally if such a mapping happens to be a bijection, then it is
<i>called an isomorphism of graded coalgebras in which case its inverse will also</i>
be an isomorphism of graded coalgebras.


</div>
<span class='text_page_counter'>(146)</span><div class='page_container' data-page=146>

<i>Tensor products of coalgebras 135</i>


<b>7.4 Tensor products of coalgebras</b>


<i>Let Al,A2,..., Ap</i> be K-modules. There is an isomorphism


<i>Vp:(Ai®A1)®---®(Ap®Ap)</i>



<i>-^-+(Al®--®Ap)®(Al®---®Ap) (7.4.1)</i>


<i>in which, with a self-explanatory notation, the element (ax đa\) đ ã ã ã</i>


<i>đ (ap đ a'p) of the first module is matched with the element</i>


<i>(ax ® - " đ ap) đ (a\ đ - - ã ® a'p) of the second. Thus Vp</i> is an isomorphism


which accomplishes a certain type of rearrangement. In like manner we
shall use


<i>Wp: {Ax đ A! đ A!) đ ã ã ã ® (Ap ® Ap ® Ap)</i>


- ^ (v4t<i> (x) • • • <g> Ap) ® (Ax ® - •• ® Ap) đ (Ax đ ã ã ã đ Ap)</i>


(7.4.2)
for the rearranging isomorphism which makes


<i>(ax ® a\ ® a'{) ® (a2 ® a'2 ®a"2)®'"® (ap ®a'p® a"p)</i>


and


<i>ap)® {a\ ®af2®-- ®a'p)</i>


<i>®'-® a"p)</i>


correspond.


<i>Now suppose that, for \<i<p, (A^A^Ei) is an /^-coalgebra (not</i>


necessarily graded) and put


<i>A=A1®A2®-"®Ap. (7.4.3)</i>


<i>At this stage A is just an R-module. We note that the mapping Aj đ</i>
<i>A2đm ã' đAP has domain Ax đ A2 đ- ã ã đAp</i> and codomain


<i>{Ax ®AX)® (A2 ®A2) ® • • - ® (Ap® Ap). Consequently, if we put</i>


<i>A = Vp o (Aj ® A</i>2<i> ® - - • ® Ap\ (7.4.4)</i>


<i>then A is an R-linear mapping of A into A® A.</i>


<i>Consider the product R ® R ® • • • ® R where the number of factors is p.</i>
There is a homomorphism


<i>$>: R ® R đ ã ã ã đ R -> R (7.4.5)</i>


which satisfies ^p )(r<i>i ® ri ®''' ® rP) — riri • • • rP></i> a n (l using this we may


define


<i>e:A-+R (7.4.6)</i>
by


<i>e = ixtf ° (et ® e2 ® • • • ® ep). (7.4.7)</i>


</div>
<span class='text_page_counter'>(147)</span><div class='page_container' data-page=147>

<i>136 Coalgebras and Hopf algebras</i>


<i><b>Theorem 1. For l<i<p let (A</b>h Ahet) be an R-coalgebra. Then with the</i>



<i>above notation {A, A, e) is also an R-coalgebra.</i>
<i>Proof. The mapping</i>


04! ® Ax<i>) o A</i>x<i> ® (A2</i> ® A2) ° A2 ® • • • ® (Ap ® Ap) ° Ap


takes X1® ^2<i> <$••• ® Ap into ( ^ ® ^ ® Ax) ® • • • ® (Ap</i> ® Ap<i> ® Ap).</i>


Consequently


<i>Wp°((Al®A1)°A1®--®(Ap®Ap)°Ap) (7.4.8)</i>


<i>maps A into /I ® A ® A. We claim that (7.4.8) is the same as (A ® A) ° A.</i>
In order to verify this claim we shall introduce some notation which is
rather cumbersome but which is fully explicit. Later, when dealing with
similar but sometimes more complicated situations, we shall simplify the
notation in order to prevent it becoming unwieldy. The following argument
will provide a model for expanding the simplified expressions if the reader
wishes to have more details.


<i>Suppose that al9 a2,..., ap belong to Al9A2,..., Ap</i> respectively and let


us write


and so on until finally we have


(The simplification that we shall use later consists in replacing these
<i>relations by AA.(ai) = Yl^i ®a'l for 1 = 1 , 2 , . . . , p. Thus there will be no</i>


<i>explicit references to the parameters k, ft,..., x used to index the terms in</i>


the various sums. The reader is to take their presence as understood.) It then
follows that


<i>(Ax đ A</i>2 đ ã ã • ® Ap<i>)(fl! ® a2 ® • • • ® ap)</i>


</div>
<span class='text_page_counter'>(148)</span><div class='page_container' data-page=148>

<i>Tensor products of coalgebras 137</i>


and therefore


<i>(A ® A) (A(fl! đ a2</i> đ ã ã ã


= Z (fli



On the other hand,


<i>((A! đ Aj) A</i>x<i> đ ã ã • ® (Ap</i> ® Ap) ° Ap) (fl!


from which it follows that


<b>>---®fZ</b>

<b>fl</b>

<b>p(</b>

<b>T</b>

<b>)®Vp(</b>

<b>T</b>

<b>))</b>



<i>applied to ax</i> ® a2 ® ' *' ®<i> aP</i> yields


<i>• • • ®a'p(T))</i>


This establishes our claim.
It has now been shown that


and similarly it can be proved that



<i>(A ® A) ° A = Wp</i>° ((Ax ® Ai) o Aj đ ã ã ã đ (Ap đ Ap) Ap).


But (Xf<i> ® At) ° A</i>f = (Af<i> ® At) ° A</i>f because Af is coassociative. Accordingly


<i>(A ® A) ° A = A ° (A ® A) and therefore A is coassociative as well.</i>
It remains for us to demonstrate that


<i>R®A < A ® A</i>


is a commutative diagram, and, as the two triangles can be treated similarly,
we shall confine our attention to the one on the left-hand side. In this case it
<i>will suffice to show that the effect, on ax ® a2 ® * * * ® ap, of applying in</i>


</div>
<span class='text_page_counter'>(149)</span><div class='page_container' data-page=149>

<i>138 Coalgebras and Hopf algebras</i>


A<i> r.đA</i>


<i>A ã A ®A >R®A -^—> A</i>


is to leave the element unchanged. Now, when these homomorphisms are
applied the outcome is


<i><b>/</b></i>


<i><b> ]e</b><b>p</b><b>(a'</b><b>p</b><b>(x))a"</b><b>p</b><b>(r)\</b></i>


But ^]/le1(ai(A))a'][(A) = a1 because the diagram


<i>-A,</i>



commutes, and similarly in the other cases. Accordingly the image of ax ®


<i>di ®''' ® dp</i> is itself and with this observation the proof is complete.


<i><b>Definition. The coalgebra (A,A,e)of Theorem 1 is called the tensor product</b></i>


<i>of the coalgebras AX,A2,. • • ,Ap and it is denoted by Ax ® A2 ® ' ' ' ® Ap.</i>


<i>Since R is both an algebra and a coalgebra, the p-fold product R đ</i>
<i>R đ ã ã ã đ R is also an algebra and a coalgebra. Now by (7.1.4) and (7.4.5)</i>
we have jR-linear mappings


<i><b>Atf: R-+R ® R ®- - ®R</b></i>
and


in fact Ajf}<i> is the unit mapping of the algebra R ® R ® • • • ® R whereas ^</i>
<i>is the counit mapping of the coalgebra R ® R ® • • • ® R. It follows that A{f</i>}
is an algebra-homomorphism and /xjf* is a coalgebra-homomorphism.
However, more is true as is shown by


<i><b>Lemma 1. Ajf* is a homomorphism of both R-algebras and R-coalgebras;</b></i>


<i>and the same is true of fi^\</i>


<i>Proof It is clear that ///> is an algebra-homomorphism so we need only</i>
show that Ajf<i>)</i> is a homomorphism of coalgebras. Evidently Ajf) preserves
counits so we are left with showing that the diagram


<i>R -> R đ R</i>



<i>Rđ'- đR > (R đ - ã - đ R) đ (R đ ã ã ã đ R)</i>


</div>
<span class='text_page_counter'>(150)</span><div class='page_container' data-page=150>

<i>Tensor products of coalgebras 139</i>


<i>element of R when it is transported to the bottom right-hand corner by the</i>
two available routes.


Now that tensor products of coalgebras have been defined it is natural to
ask whether our results on tensor products of algebras have valid
analogues. And in fact they do. However, partly because a coalgebra is a
relatively unfamiliar object, and partly because of the greater use of
mappings made in characterizing coalgebras, it is not an entirely trivial
matter to adapt the arguments originally given in Chapter 3. Consequently,
until the reader has had a chance to become familiar with the way in which
adjustments may be made, we shall expound the dual theory in a fairly
leisurely manner.


<i>First suppose that Ai9 A2, . . . , Ap and Bl9 B2,..., Bp</i> are jR-coalgebras


and let


<i>fi'.A^Bi (7.4.9)</i>


<i>be a coalgebra-homomorphism for i= 1 , 2 , . . . , p. Then certainly</i>
<i>/ i ® fi ®''' ® fp is a homomorphism of the module Al®A2®- - ®Ap</i>


<i>into the module Bl®B2®"'®Bp. However, as the following lemma</i>


shows, the homomorphism respects the coalgebra structures.



<i><b>Lemma 2. With the above assumptions</b></i>


<i>/i ® * * * ®fp'- Ax ®A2 ® • - - ®Ap-+B1 ®B2 ® • • • ®Bp</i>
<i>is a homomorphism of coalgebras.</i>


<i>Proof Put / = / i ®f2®'"®fp,A = Al®A2®-"®Ap,B = B1®</i>


<i>B2®-"® Bp, and let at belong to At (i = 1, 2 , . . . , p). If now, using the</i>


<i>abbreviated notation (see the proof of Theorem 1), we set AA,(ai) =</i>


<i>X a\ ® a", then</i>


<i>&A(al®a2®---® ap) = Ya {a\®--® a'p) ® (af[ ®'</i>


and hence


<i>(f®f)(AA(di®"-®ap))</i>


<i>Again, because f is a homomorphism of coalgebras,</i>
Af l |(/^)


and therefore


= Z (/ia<i>i ® ' *' ® fp<*p) ® (/ifl'i ® ' ' ' ® fPa"P</i>


</div>
<span class='text_page_counter'>(151)</span><div class='page_container' data-page=151>

<i>140 Coalgebras and Hopf algebras</i>


In the case of the counits we have



<i>(efl/)(fli đ ' ã' đap) = eBi(f1a1)eB2(f2a2)... eBp(fpap).</i>


<i>But &B. (ftai) ~ ÊAt (ai)- Consequently</i>


<i>(sBf)(a1 đ ã ã ã đ ap) = eAi(a1)eA2(a2)... e^fo,)</i>


<b>= *U(0i®"-®0p).</b>


<i>Thus sB° f = eA</i> and the proof is complete.


We turn next to the basic isomorphism theorems for coalgebras and
prove results which are the counterparts of Theorems 2, 3 and 4 of
Chapter 3.


<i><b>Theorem 2. Let A</b><b>l9</b> A2,..., A</i>p<i><b> arcd B</b><b>1</b><b>,B</b><b>29</b><b>...,B</b><b>q</b> be R-coalgebras. Then</i>


<i>the canonical isomorphism</i>
<i>A1®--®Ap®B1®'-®Bq</i>


<i>of modules (see Chapter 2, Theorem 1) is actually an isomorphism of </i>
<i>R-coalgebras.</i>


<i>Proof Put A=AXđ ããã đAp,B = Bx đ '-đBq</i> and C =


<i>Axđ- ã - đApđB1đ" - đBq. In what follows a(</i> will denote an element


<i>of At and bj an element of By, furthermore we shall suppose that AAi(at) =</i>


X<i> a'i ®</i> af' a n (<i>^ ABj(bj) = Y, b'j ® fc}. Lastly we shall use</i>



to denote the canonical module-isomorphism. Of course, it will suffice to
<i>show that (j) is a homomorphism of coalgebras.</i>


To this end we first observe that the result of applying Ac to


<i>a1đ-" đapđb1đ"</i>


<i>' * ã đ a'p ® b\ ® • • • ® bq) ® {a'[ ® • • • ® d'p ® b\ ® • • • ® bq)</i>


<i>and therefore when {(j) ®(j>)° A</i>c is applied to the same element what we


obtain is


</div>
<span class='text_page_counter'>(152)</span><div class='page_container' data-page=152>

<i>Tensor products of coalgebras 141</i>


say, and we also have


Ay4<i>(a) = S (fli đ ' ã ã đa'p) đ (a'[ ®--'®a"p)</i>


and


<b>Afi</b><i><b>(fr)=£ (b\ ® • • • ® *>;) ® (//; ® • • • ® / g .</b></i>


<i>A c c o r d i n g l y AA<S)B°(f) a p p l i e d t o al®''-®ap®bl®-'-®bq</i> p r o d u c e s


<i>the element (7.4.10) and so we have established that ((/> ® 4>) ° A</i>c = A^ ^ ° </>.


<i>Finally when £A®B°(I) and e</i>c are made to act on ax ®- • • ® ap®


<i>^i ® * * * ® bq</i> the result in both cases is



<i>e^di)... ^</i>p(0p)£B l<i>(M . . . eBq(bq).</i>


<i>It follows that £A0BO(t) = ec</i> a nd now the proof is complete.


<i><b>Theorem 3. Let A</b>l9A2,... ,Apbe R-coalgebras and let (il9 i2,..., ip) be a</i>


<i>permutation of (1, 2 , . . . , p). Then the canonical isomorphism</i>


<i>A1®A2®---®Ap^Aii ®Ai2®--- ®Aip</i>


<i>of modules (see Chapter 2, Theorem 2) is in fact an isomorphism of </i>
<i>R-coalgebras.</i>


<i>Proof Let 0 : Ax đ A2 đ ã ã ã đ Ap-^Aix đ Ai2 đ ã ã ã đ Ai</i> be the


<i>module-isomorphism in question. For 1 <i <p let at belong to A{</i> and suppose that
A


^(ôi) = Zf lJđf li'- P u t<i> ^ = ^ 1 ® ^ 2 ®'"®Ap and B = Aix ®Ai2</i>


<i>®---® A{ . If now ((/) ® (j>) ° A^ and AB °<j) are applied to a1 đ a2 đ ã ã ã ® ap, then</i>


the result in both cases is


<i><b>£ K ® • •' ® a'</b><b>ip</b><b>) ® (a'l</b><b>x</b><b> ® • • • ® a'(</b><b>p</b><b>),</b></i>


<i>whereas if sB ° </> and sA</i> are made to operate on the same element, then the


<i>result, again in both cases, is eAi(a1 )eAi(a2)... eA (ap). The theorem follows.</i>



<i>We recall that R itself is a coalgebra.</i>


<i><b>Theorem 4. Let A be an R-coalgebra. Then the canonical </b></i>


<i>module-isomorphisms A ® R &A and R đ A ô A are module-isomorphisms of R-coalgebras.</i>


<i>Proof Let (/>: A đ R ã A be the module-isomorphism in which (f)(a ®r) =</i>
<i>ra, let a belong to A, and let A^ (a) = £ a' ® a''. Then</i>


so that (0 ® 0) ° A^<i> 0R = AA</i> ° 0. Again


and therefore<i> &AO4>=&A®R'</i> This shows that ,4 ® # ^ y 4 is a


coalgebra-isomorphism and the other coalgebra-isomorphism is dealt with similarly.


</div>
<span class='text_page_counter'>(153)</span><div class='page_container' data-page=153>

<i>142 Coalgebras and Hopf algebras</i>


<i>A =Ail) đ A(2) đ ã ã ã ® Aip)</i> (7.4.11)
<i>has a natural structure as a coalgebra. In what follows A^ and eA</i> will denote


the comultiplication and the counit of this coalgebra.


Let / = (il5<i> i2, • • •, ip) be a sequence of p integers, and (as in the case of</i>


graded algebras) put


<i>AI = A\1i)®A\22)®- -®A\P)</i> (7.4.12)


so that



<i>A=YtAI</i> (7.4.13)


this being a direct sum of /^-modules. If we now set


|/| = i1+ i2 + - " + ip (7.4.14)


and


<i>An= Z A!9</i> (7.4.15)


<i>where the sum is taken in A, then {An</i>} «6Z<i> is a grading on the module A. It is</i>


<i>clear that if R is graded trivially, then eA: A —• JR preserves the degrees of</i>


<i>homogeneous elements; and it follows from (7.4.4) that A^: A—+A ® A is</i>
<i>also degree-preserving provided that A ® A is given the usual total grading.</i>


<i>We can sum up these observations as follows. Let A{1\ A{2\ . . . , Aip)</i> be
<i>graded coalgebras. Then A{1) đ A{2) đ ã ã ã đ A{p) has the structure of a</i>
<i>coalgebra by virtue of Theorem 1. Furthermore the usual total grading on the</i>
<i>module A{i) ® Ai2) đ ã ã * đ A{p) is a coalgebra-grading.</i>


In the next section we shall discuss a way of modifying this structure. For
the present, however, we shall leave this refinement on one side and begin by
<i>observing that if A{1\ A{2\ . . . , A{p) and B{1\ B{2\ . . . , B{q)</i> are graded
coalgebras, then the isomorphism


<i>Aa) ® • • • ® A{p) ® B(1) ® • • • ® B{q)</i>



<i>% (Ail) ® • • • ® Aip)) đ (B</i>(1)<i> đ ã ã ã đ Biq)\ (7.4.16)</i>
<i>provided by Theorem 2, is an isomorphism of graded coalgebras; and that if</i>
<i>(il9 i2, • • •, ip) is a permutation of ( 1 , 2 , . . . , /?), then the isomorphism</i>


<i>Ail) đ A{2) đ ã ã • ® A{p) % A{il) ® A{il) ® • • • ® A{i^ (7.4.17)</i>
<i>of Theorem 3 is also an isomorphism of graded coalgebras. Yet again, if A is</i>
a graded coalgebra, then the coalgebra-isomorphisms


<i>A®R*A (7.4.18)</i>
and


<i>R®A^A (7.4.19)</i>


</div>
<span class='text_page_counter'>(154)</span><div class='page_container' data-page=154>

<i>Modified tensor products of coalgebras 143</i>


Finally, suppose that
<i>ft:A®^>B®</i>


<i>is a homomorphism of graded coalgebras for i = 1 , 2 , . . . , p. Then, because</i>
<i>ã ã ã đ Aip)->Bil) đ • • • ® Bip)</i>


/ i ® <i>® fp:</i>


<i>preserves degrees, it is a homomorphism of graded coalgebras on account of</i>
Lemma 2.


<b>7.5 Modified tensor products of coalgebras</b>


In this section we shall develop ideas that correspond to those
<i>described in Section (3.4). Suppose then that A(l\ A{2\ . . . , Aip)</i> are graded


^-coalgebras. We have already seen how the /^-module


<i>A=A(1)®Ai2)®'-'®Aip)</i> (7.5.1)
<i>can be turned into a graded coalgebra (A,AA,eA). We now propose to</i>


<i>modify the coalgebra structure on A. The new coalgebra will have the same</i>
/^-module structure, the same grading, and the same counit as before; only
the comultiplication will be changed and this in a comparatively simple
way. We are, as it were, putting a twist into (7.5.1).


<i>Let I = (il9 i2, • • •, ip) and J = (j1J2, • • • JP) be sequences of p integers,</i>


and, as in (7.1.9) and (7.1.10), put


<i><b>N(U)=Y</b><b>J</b><b>Us (7.5.2)</b></i>


<i><b>r>s</b></i>


and


<b>(7.5.3)</b>


<i>We shall use I + J to denote the sequence (^ +jl9 i2 +j2, • • •</i> ? *p+7P


<i>)-For given / and J, there is an isomorphism of</i>

<i><b>(A^ ® Atf) ® (4</b></i>

<b>2</b>


<b>2)</b>


<i><b> ® A%) ® • • • ® (Atf ® 4J?) (7.5.4)</b></i>




onto


in which


<i>K ® a'h) ® (ai2 ®a'h)®'"® (aip ® a'jp)</i>
is mapped into


<i>e(I, J)(ati ® ai 2® - - - ® aip) ® (a'h ® a 'h® - - - ® a ) )</i>


and these various isomorphisms can be combined to give an isomorphism
<i>Vp: (A(1) đ A{1)) đ ã ã ã đ (Aip) đ Aip))</i>


</div>
<span class='text_page_counter'>(155)</span><div class='page_container' data-page=155>

<i>144 Coalgebras and Hopf algebras</i>


<i>Let us compare Vp</i> with the untwisted isomorphism


<i>Vp</i>:


<i>which operates as in (7.4.1). If we denote by nu the projection of (A{1)</i> đ ã • •


<i>® A{p)) ® (A</i>(1)<i> ® • • • ® Aip)) onto ,4, ® i4</i>y, that is onto the module (7.5.5),


then we find that


<i>nu°Vp = e(I9J)7tuoVp. (7.5.7)</i>


<i>We next define an K-linear, degree-preserving mapping A: A^>A ® A</i>
by means of the formula



<i>A = Vp(AAn) đ A</i>i4(2> đ ã ã ã đ A^(P»). (7.5.8)


<i>Now, by Theorem 1, A = A{1)</i> ® y4(2)<i> ® • • • ® A{p)</i> is a coalgebra. Let us use
A and e to denote its comultiplication and counit mappings. Then, by
(7.4.4), (7.5.7) and (7.5.8),


<i>nIJ°A = e(I,J)nu°A. (7.5.9)</i>


<i><b>Lemma 3. The triple (A</b>{1) đ ã ã ã đ A{p), A, e) is a coalgebra and the usual</i>
<i>total grading on A(1) ® A{2) ® • • • ® A{p) grades the coalgebra.</i>


<i>Proof. It will suffice to prove the first statement and for this a little</i>
<i>temporary terminology will be useful. Let us say that an element a of A is</i>
<i>pure if it has the form a , , ® ^ ® - - - ® ^ (where ociveA{^) and when this is</i>


the case let us use {a} to denote the sequence (zl9<i> i2,..., ip). Pure elements</i>


<i>are, of course, homogeneous and every element of A is a finite sum of pure</i>
elements.


<i>The idea is to use the fact that (A, A, e) is already known to be a graded</i>
<i>coalgebra. Suppose then that a belongs to A and is pure. Then A(a) can be</i>
expressed in the form


A(a) = £ V ®a",


<i>where a', a" are pure and {a'} +{a"} = {a}. Note that nIJ(A((x)) = 0 unless</i>


J. Note, too, that



<i>Now suppose that /, J, K are three sequences each consisting of p</i>
<i>integers and let nIJK be the projection of A ® A ® A onto At ® A3 ® AK. It</i>


<i>is clear that if we identify A ® A ® A with A® (A® A) in the usual way,</i>
<i>then nlJK = nI ®nJK. Accordingly</i>


</div>
<span class='text_page_counter'>(156)</span><div class='page_container' data-page=156>

<i>Modified tensor products of coalgebras 145</i>


and


K , , K ° 0 4 ® A ) ° A ) ( a )


<i>= % e({*'}9 {*'})*,(*') ®nJJC(Aắ).</i>


<i>But nj,K(A(x") = e(J, K)nj,K(A(x") by (7.5.9) and therefore</i>


<i>(ni,j,K° 04 ® A)° A) (a)</i>


<i>= e(J, K) £ e({a'}, {a"})7i</i>7(a') ® 7i^(Aa")


<i>= e(J, K M / , J + X) X Ti; (a') ®</i> TC^ (Aa")


because 7c7<i> (a') = 0 unless {a'} = /, and nJK (Aa") = 0 unless {a"} =J + K. This</i>


shows that


<i>(ni,j,K ° (4 ® A) ° A) (a) = e(J, K)e(I, J + K)(nUiK ° (A ® A) ° A)(a)</i>


and in a similar manner we can show that



<i>(71^^° (A® A)o A)(oi) = e(I,J)e(I + J, K)(nu,Ko (A® A)° A)(a).</i>


However, it is easy to verify that


<i>e(I,J + K)e(J, K) = e(I, J)e{I + J, K)</i>


<i>and we know that (A ® A) ° A = (A ® A) ° A because A is a comultiplication.</i>
Accordingly


<i>KIJ,K ° (A ® A) ° A = nu x ° (A ® A) ° A</i>


<i>for all /, J, /C and therefore (A ® A) ° A = (A ® A) ° A.</i>
It remains to be shown that the diagram


(7.5.10)


<i>A đ^R < Ađ A ã /Tđ A</i>


commutes. Now the combined effect, on a, of the mappings


A<i> A®e</i>


<i>A >A ®A >A ®R -^—+A</i>


is to turn it into £ e({a'}, {a"})e(a")a'. But if 6(a")^0, then {a"} =
( 0 , 0 , . . . , 0) and therefore e({a'}, {a"})= 1. It follows that


and this is just a because the mappings


A<i> A®e</i>



<i>A > A ®A >A ®R -^—^ A</i>


</div>
<span class='text_page_counter'>(157)</span><div class='page_container' data-page=157>

<i>146 Coalgebras and Hopf algebras</i>


<i>The graded coalgebra (A{1) đ ã ã ã đ A{p\ A, e) will be called the modified</i>
<i>tensor product or the twisted tensor product of A(1\ A{2\... ,A</i>( p );andit will
<i>be denoted by A{1) ® A{2)</i> ® • • • ® v4(p) to distinguish it from the coalgebra
<i>A{ 1 ) ®A(2) ®--® Ai p ).</i>


We proceed to establish the basic properties of these modified tensor
products. Suppose first that


<i>f.:A(i)^B(i) ( i = i , 2 , . . . , p ) (7.5.11)</i>


<i>is a homomorphism of graded coalgebras. Then fx ® f2 ® • • • ® fp</i> is a


<i>homomorphism of the coalgebra A{1)</i> đ A(2)<i> đ ã ã ã đ A{p)</i> into the
<i>coalgebra B{1) đ B{2) đ ã ã ã đ Bip\ But changing to modified tensor</i>
products does not affect the underlying module structures. Consequently
/ i ® - " ® /p: ^( 1 )<i> ®'" ® A{ P )- > B{ 1 ) ® - - - ® Bi p ). (7.5.12)</i>


Note that (7.5.12) is JR-linear and preserves degrees.


<i><b>Lemma 4. / / f</b>r đ f2 đ ã ã • ® fp is considered as a mapping of A{1)</i> ®


<i>A{2) ® • • • ® Aip) into B{1) ® B{2) đ ã ã ã đ B{p\ then it is a homomorphism of</i>
<i>graded coalgebras.</i>


<i>Proof Consider the element ai{ ® a-l2 ® • • • ® a(, where aiveA^\ We can write</i>



<i>where aj, and aj; are homogeneous elements of AUl)</i> the sum of whose degrees
is i,,, and then


<i>Put A = A{1) ® • • • ® A{p\ B = B{1) ® • • • ® B{p\ f = /</i>x<i> ® f2 ® • • • ® fp</i> and


<i>set a' = ai ® • • • ® a'p, a" = a'[®- - ® a"p. If now we use {a1} to denote the</i>


<i>sequence formed by the degrees of a\, a'2, -.. •> a'p and we define {a"}</i>


similarly, then we find that


<i>AA(ati ®ah®- -®aip) = % e({a% {a"})ar</i>


Accordingly


But


<i>and {fa'} = {a'}, {fa"} = {a"} because ffl</i> is a homomorphism of the graded


</div>
<span class='text_page_counter'>(158)</span><div class='page_container' data-page=158>

<i>Modified tensor products of coalgebras 147</i>


<i>from which it follows that AB ° / = ( / ® / ) ° AA</i>. Thus / is compatible with


the appropriate comultiplications.


We still have to satisfy ourselves that / preserves counits. However, this
follows from the corresponding result for unmodified tensor products
because the mappings involved are exactly the same in both cases.



The next result corresponds to Theorem 6 of Chapter 3. In order to state
<i>it we suppose that A(1\ A{2\ . . . , A{p) and Ba\ B{2\ . . . , B{q) are graded </i>
R-coalgebras. Then, by Theorem 1 of Chapter 2, there is an ^-linear bijection


<i>®A{p) ®B(1) < Đ ã ã ã đBiq)</i>


<i>ã ã ã đ A(p)) đ (B</i>(1)<i> đ ã ã ã đ Biq)) (7.5.13)</i>


in which


<i>4>{ah đ--đaipđbhđ--đ bjq) = (ati đ ã ã ã đ aip) đ (bji đ ã ã ã đ bjq).</i>


<i><b>Theorem 5. The bijection (7.5.13) is an isomorphism</b></i>


<i>ã ã ã đ Aip) đ B(1) ® • • • ® Biq)</i>


<i>• • • ® A(p)) ® (B(1) ® • • • ® Biq))</i>


<i>of graded coalgebras.</i>


<i>Proof It is evident that </> preserves degrees and that it preserves counits.</i>
Consequently we need only show that it is compatible with the appropriate
comultiplication mappings. In doing this we shall put


<i>A=A{1)®Ai2)®---®Aip\</i>
<i>B = Bil) đB{2) đ"- đB(q\</i>
and


<i>C = Aa) đ ã ã • ® A{p) ® B{1) ® • • • ® Biq)</i>
in order to simplify the notation.



<i>Now suppose that at eA^ and b; EB\V\ Then</i>


<i>where a'fi, a"H are homogeneous elements of AUl)</i> the sum of whose degrees is


<i>itn and bfY, b" are homogeneous elements of B{v)</i> the sum of their degrees


being 7,,. Since


</div>
<span class='text_page_counter'>(159)</span><div class='page_container' data-page=159>

<i>148 Coalgebras and Hopf algebras</i>


it follows that


<i>= 1 *(Ki đ ã ã ã đ *>;}, {a\ ® • • • ® fc;})(ai ® • • • ® vq) ® (*; ® • • • ® &;),</i>


<i>where by {ai đ ã ã ã đ b'q) is meant the sequence formed by the degrees of a\,</i>


<i>. . . , a'p, b\,..., b'q</i> taken in order. Accordingly


= Z<i> e(Wi ® ' * * ®</i> b<i>i}> K i ® ' * * ® bq})(a' ® fc') ® (a" đ &"),</i>


where a' = ai đ ã ã ã ® ap, a" = a? ® • • • ® a,, &' = & i ® - - - ® ^ and ft" =


<i>Put / ' = {a\ ® • • • ® ^p}, J' = {b[ ® • • • ® fo^} and define J" and J"</i>
similarly. Also, if / ' = (i'i, r2, . . . , ip) and J' = ( /l 5/2<i>, . . . J'q) let us use / ' J ' to</i>


<i>denote the sequence ( i i , . . . , i'p9 fl9... Jq). We then have</i>


<i>((0 đ (j>)o Ac)(ati đ ã ã ã đ aip ® bh ® • • • ® bJq)</i>



<i>= X e(I'J\ I"J")(a' ® 6') ® {a" ® b"). (7.5.14)</i>
But, on the other hand,


<i>*AK ® *' * ® %) = ! e({af}, {a"}){a'®a") = Y e{I\ F)(af</i>
and


<i>e(J\ J"){V ® V)</i>


so therefore


<i>(A^ ® AB)((atl</i> đ ã ã ã đ afp<i>) đ (bh đ • • • ® ^ ) )</i>


<i>X ® (ft' ® W).</i>


<i>Now a', a" are homogeneous elements of A and b', b" are homogeneous</i>
<i>elements of B. Furthermore, if J' = (j'l9... J'q) and /" = ( i j , . . . , *"</i>p), then the


<i>degree of b' is | J' | =j\ +f2 + • • • + ; ; and that of a" is |/" | = i'[ + i'i + ã ã ã + f</i>p.


Accordingly


<i>đ ( ^ ® ' *' ® bjq))</i>


<i>r){-\^J^r'\af®b')®{an®b't). (7.5.15)</i>


However, it is easy to check that


<i>e(l\ I")e(J\ J"){-iyfl]I"l = e(rj\ I"J")</i>


and so, comparing (7.5.14) with (7.5.15), we see that ( ^ ® 0 ) ° AC and



<i>&A®BO(<sub>}</sub>><sub> produce the same result when applied to a^®- - ®a</sub></i>


<i>{ ®</i>


<i>bjt ® * * * ® bj. It follows that (0 ® (/>) ° A</i>c = A^^^ ° 0 and with this the


</div>
<span class='text_page_counter'>(160)</span><div class='page_container' data-page=160>

<i>Modified tensor products of coalgebras 149</i>


The final result in this section shows that the formation of modified
tensor products of coalgebras is a commutative operation. To prepare the
<i>way for proving this, first suppose that A and B are graded /^-modules with</i>
<i>{An}neZ the grading on A and {Bn}nEl the grading on B. We recall that the</i>


<i>twisting isomorphism (see (3.8.5))</i>


<i>T:A®B-^-+B®A (7.5.16)</i>


is an isomorphism of /^-modules in which


<i>®bn) = (-\rnbn®am. (7.5.17)</i>


<i>Here, of course, ameAm and bneBn.</i>


<i>Now assume that A and B are graded coalgebras. In this case we can</i>
<i>regard the twisting isomorphism as mapping the coalgebra A ®B onto the</i>
<i>coalgebra B ® A. The next theorem shows that, in this context, T is an</i>
isomorphism of coalgebras.


<i><b>Theorem 6. Let A and B be graded coalgebras. Then the R-linear bijection</b></i>



<i>T: A ® B—>B ® A, where T is the twisting isomorphism, is an isomorphism</i>
<i><b>of graded coalgebras.</b></i>


<i>Proof T certainly preserves degrees. Now suppose that ameAm and bneBn.</i>


Then


<i>because, since A and B are graded coalgebras, eB(bn)sA(am) is zero except</i>


<i><b>perhaps when m = n = 0. It follows that e</b>B(S)A(T(am ® bn)) = eA(8)B(am ® bn).</i>


<i>Accordingly sB<S)A°T=eA<S)B and we have shown that T preserves counits.</i>


We turn now to the question of compatibility with comultiplication. We
<i>can write A^ (am) — £ a' ® a", where a\ a" are homogeneous elements of A</i>


<i>the sum of whose degrees is m; likewise we have AB(bn) = YJ br ® b", where</i>


<i>this time b\ b" are homogeneous elements of B and the sum of their degrees</i>
<i>is n. Let us use {#'}, {a"), {b'}, {b") to denote the degrees of these various</i>
elements. Then


<i>*AQB(am ® bn) = £ ( - l)</i>{*'}{fl<i>V ® V) ® (a" ® b")</i>


and therefore


<i>(T®T)(AA®B(am®bn))</i>


</div>
<span class='text_page_counter'>(161)</span><div class='page_container' data-page=161>

<i>150 Coalgebras and Hopf algebras</i>



<i>= AB(^A(T(am®bn)).</i>


<i>This shows that (T ® T) ° AA^B = AB®A ° T and now the proof is complete.</i>


<b>7.6 Commutative and skew-commutative coalgebras</b>
<i>Let A and B be ^-modules, and let us use</i>


<i>U:A®B -^>B®A (7.6.1)</i>


<i>to denote the module-isomorphism in which U(a ® b) = b ®a. If now</i>
<i>(A, \i, n) is an K-algebra, then it is a commutative algebra if and only if</i>


<i>u</i>


<i>A®A >A®A</i>


(7.6.2)


is a commutative diagram.


<i>Suppose next that (A, ft, rj) is a graded algebra and let T: A ®A^+A ® A</i>
<i>be the twisting isomorphism (see (7.5.16)). We shall say that A is a </i>
<i>skew-commutative algebra provided that the diagram</i>


<i><b>T</b></i>


<i>A đA ã A đA</i>


(7.6.3)



commutes. This is equivalent to requiring that


« = (-ir«A (7.6.4)
<i>for homogeneous elements am, ocn of degrees m and n respectively. Note that</i>


<i>this is weaker than the requirement of being anticommutative; in fact a</i>
graded algebra is anticommutative if and only if (i) it is skew-commutative,
and (ii) the square of every homogeneous element of odd degree is zero.


These observations suggest the following definitions for coalgebras. A
<i>coalgebra (A, A, e) will be called commutative if the diagram</i>


<i>u</i>


<i>A ®A • A ®A</i>


<i><b>\ A (7.6.5)</b></i>



<i>A</i>


</div>
<span class='text_page_counter'>(162)</span><div class='page_container' data-page=162>

<i>Linear forms on a coalgebra 151</i>


<i>T</i>


<i>A®A >A®A</i>


(7.6.6)


<i>is a commutative diagram. Here, of course, T is the twisting isomorphism.</i>


With this definition we have reached the point where the parallelism
between algebras and coalgebras has been developed as far as we need.
Further results will be found in the exercises at the end of the chapter.


<b>7.7 Linear forms on a coalgebra</b>


It will now be shown that the linear forms on a coalgebra
constitute, in a natural way, an R-algebra. To see how this comes about,
<i>suppose that (A, A, e) is a given R-coalgebra and put</i>


<i>A* = HomR(A,R) (7.7.1)</i>


<i>so that A* is the module of linear forms on A. First we note that the counit e</i>
<i>belongs to A* and therefore there exists an ^-linear mapping</i>


<i>rj:R-+A* (7.7.2)</i>


<i>in which rj(r) = re. This mapping will provide the structural homomorphism</i>
of our algebra.


<i>Next let nR. R đ R ã R be the multiplication mapping of R (considered</i>


<i>as an algebra) and let / , g belong to A*. Then ixR ° ( / ® g) ° A also belongs to</i>


<i>A* and the mapping A* xA*—+A* in which (/, g) becomes fiR ° (f ® g) ° A</i>


is bilinear. It follows that there is a homomorphism


<i>fi:A*®A*-+A* (7.7.3)</i>



of jR-modules which is such that


(7.7.4)
<i><b>Theorem 7. Let (A, A, e) be an R-coalgebra. Then (with the above notation)</b></i>


<i>the triplet (A*, /x, rj) is an R-algebra.</i>


<i>Proof. Let / , g, h belong to A* and consider the mappings</i>


<i>fđgđh</i>


<i>A > Ađ Ađ A ã RđRđR > R, (7.7.5)</i>


<i>where A^A ®A ® A is (A ® A)° A = (A ® A)° A and R®R®R-^R</i>
<i>takes rx®r2® r3 into rx r2r3. The total mapping (7.7.5) can be factored into</i>


<i>R®h HR</i>


<i>A >A đA ã RđA >RđR >R</i>


</div>
<span class='text_page_counter'>(163)</span><div class='page_container' data-page=163>

<i>152 Coalgebras and Hopf algebras</i>


<i>But this is just ii(n(f ®g)® h). In a similar manner we can verify that the</i>
<i>complete mapping (7.7.5) is also the same as fi(f ® n(g ® h)). </i>
<i>Con-sequently, if we use \i to define multiplication on A*, then multiplication will</i>
be associative.


Next, by (7.7.4),


<i>and, because (A9A,s) is a coalgebra, (e®A)°A is the canonical </i>



<i>iso-morphism of A onto R ® A. It follows that fi(e ®f): A-+R takes a into</i>
<i>/ ( a ) and therefore fi(e ® / ) = / A similar argument shows that fi(f ® e) = /</i>
<i>as well. It is now apparent that A* is an associative K-algebra, with</i>
<i>multiplication ^ and identity element e. Furthermore rj:R —>A* is its</i>
structural homomorphism.


<i><b>Corollary. / / (A,A,e) is a commutative coalgebra, then (A*,/n,rj) is a</b></i>


<i>commutative algebra.</i>


<i>Proof. Let / , g belong to A*. If now a, a'eA, then the total mapping</i>


<i>u f®g nR</i>


<i>A® A > A ®A >R®R >R</i>


<i>takes a ®a' into f(a')g{a). Accordingly</i>
<i>v>R°(f ®g)°u=HR°(g® / ) .</i>


<i>But A=U ° A because {A, A, e) is commutative. Consequently</i>
<i>HR° (f ® g)° A = fiR° (g ® f)° A,</i>


<i>i.e. ji(f ®g) = fi{g ® f) which is what we were seeking to prove.</i>


<i>We next turn our attention to the case where (A, A, e) is endowed with a</i>
<i>coalgebra grading {An</i>}<i> neI. For a given integer k, the linear forms on A that</i>


<i>vanish on J]ni:kAn form a submodule of A*; moreover this submodule is</i>



<i>isomorphic to HomR(Ak, R) = A* under an isomorphism which maps the</i>


<i>relevant linear forms on A into their restrictions on Ak. We may therefore</i>


<i>regard A$ as an K-submodule of A*. On this understanding put</i>


9I=S^f (

7

-

7

-

6

)



<i>kel</i>


<i>so that 21 is a submodule of A*. Note that the sum (7.7.6) is direct. Usually</i>
<i>A* and 21 are different, but in the special case where only finitely many Ak are</i>


<i>non-zero we always have A* = SU.</i>


<i>It is clear that e e Ag because e preserves degrees. Assume now that feA*</i>
<i>and g e A*. Then f ® g, which of course maps A ® A into R ® R, vanishes</i>
<i>on As ® At unless both s = p and t = q. But A: A -ã A đ A preserves degrees</i>


</div>
<span class='text_page_counter'>(164)</span><div class='page_container' data-page=164>

<i>Hopf algebras 153</i>


<i>regarded as a graded subalgebra of the algebra (A, /x, rj) described in</i>
Theorem 7. More precisely we have proved


<i><b>Theorem 8. Let (A, A, e) be a graded R-coalgebra with {A</b>n}neZ as its</i>


<i>grading. Then (with the above notation) 21 is an R-algebra and it has {A*}neZ</i>


<i>as an algebra-grading. (Here A* is the dual of An and it is to be considered as a</i>



<i>submodule of the dual, A*, of A.)</i>


<i><b>Corollary. Let the graded R-coalgebra (A, A, e) be skew-commutative. Then</b></i>
<i>91 is a skew-commutative algebra.</i>


<i><b>Proof. Let / e A * and g e A*. It is easily checked that ifT: A ® A-^A ®A</b></i>
denotes the twisting isomorphism, then


<i>HR°(f®g)°T=(- l)pqjiiR° (g ® / ) .</i>


<i>But T ° A = A because (A, A, e) is skew-commutative, and now it follows that</i>


<b>7.8 Hopf algebras</b>


<i>Let A be an K-module and suppose that we are given R-linear</i>
mappings as follows:


<i>fi:A®A-+A, (7.8.1)</i>


<i>n:R-+A, (7.8.2)</i>


<i>A:A-+A®A, (7.8.3)</i>


<i>s:A->R. (7.8.4)</i>


<i>Then A and these various homomorphisms are said to constitute a Hopf</i>
<i>algebra provided that the four conditions</i>


<i>(i) (A, \i, n) is an R-algebra;</i>
<i>(ii) (A, A, e) is an R-coalgebra;</i>



<i>(iii) A: A-^A ® A and e: A^>R are homomorphisms of algebras;</i>
<i>(iv) [i\ A ® A-^A and rj: R-+A are homomorphisms of coalgebras;</i>
are all satisfied. When this is the case we shall speak of the Hopf algebra
<i>(rj, ft, A, A, e).</i>


The conditions (i)-(iy) are not independent; indeed the precise
connections between them are explained in the following lemma. Note that
<i>if (i) and (ii) are both satisfied, then A ® A is both an algebra and a</i>
coalgebra.


<i><b>Lemma 5. Suppose that (A,fi,rj) is an algebra and that (A,A,e) is a</b></i>


<i>coalgebra. Then</i>


</div>
<span class='text_page_counter'>(165)</span><div class='page_container' data-page=165>

<i>154 Coalgebras and Hopf algebras</i>


<i>(b) fi: A ® A—>A preserves counits if and only if e: A^>R is </i>
<i>com-patible with multiplication;</i>


<i>(c) n: R-+A is compatible with comultiplication if and only if A: A—+</i>
<i>A ® A preserves identities;</i>


<i>(d) n:R-+A preserves counits if and only if e:A-+R preserves</i>
<i>identities.</i>


<i>Proof (a) The mapping fi: A đ A ã A is compatible with comultiplication</i>
<i>if and only if A ° ^ = (/i ® fi)° AA(S>A. But</i>


<i>where U: A ® A -^ A ® A is the isomorphism in which U(a ® a') = a' ®a.</i>


Consequently our condition becomes


<i>A°fi = (/j.®n)°(A®U®A)°(A®A). (7.8.5)</i>


<i>On the other hand A: A—+ A®A is compatible with multiplication</i>
<i>precisely when A° fi = fiA(S)Ao</i> (A ® A) and this is equivalent to (7.8.5)


<i>because \iA (S)A — (\i ® \x) ° (A ® (7 ® A).</i>


<i>(b) For /i: A ® A^>A to preserve counits it is necessary and sufficient</i>
<i>that s ° n = eA (S)A, that is to say we require e ° \i = /nR</i> ° (e ® e). However, this is


<i>precisely the condition for a: A—>R to be compatible with multiplication.</i>
<i>(c) The structural homomorphism rj: R—+A is compatible with </i>
<i>comulti-plication if and only if (rj ®rj)°AR = A°rj, and this occurs when and only</i>


when A ( l J = l/ 1® l/ 4.


<i>(d) For rj: R^>A to preserve counits we require that eR = e°rj and this</i>


occurs precisely when £ ( 1 ^ ) = ^ .


We combine some of these observations in the following


<i><b>Corollary. Suppose that (A,fi,rj) is an algebra and that (A</b>9A,e) is a</i>


<i>coalgebra. Then fi: A ®A^>A and n.R-+A are homomorphisms of</i>
<i>coalgebras if and only if A: A ã A đ A and e: A ằ R are homomorphisms of</i>
<i>algebras.</i>



<i>Now suppose that (rjA, fiA,A,AA,sA) and (rjB, fiB,B, AB,eB) are Hopf</i>


<i>algebras. A mapping A^>B is called a homomorphism of Hopf algebras</i>
provided it is homomorphism both of algebras and of coalgebras. A
<i>bijective homomorphism of Hopf algebras is called an isomorphism of Hopf</i>
<i>algebras. Evidently if / : A —• B is such an isomorphism, then / "1: B —• A is</i>
<i>also an isomorphism of Hopf algebras. Again, if g: A —• B and h: B —• C are</i>
<i>homomorphisms of Hopf algebras, then so is their product h ° g.</i>


Another concept which we shall need is that of a commutative Hopf
<i>algebra. Quite simply, a Hopf algebra is said to be commutative if it is</i>
commutative both as an algebra and as a coalgebra.


</div>
<span class='text_page_counter'>(166)</span><div class='page_container' data-page=166>

<i>Hopf algebras 155</i>


<i>family of i^-submodules of A which grades A as an K-module. We then say</i>
<i>that {An}neI grades the Hopf algebra provided that it grades the algebra</i>


<i>(A, \i, rj) and the coalgebra {A, A, e). This will be the case if \i, rj, A, e all</i>
preserve the degrees of homogeneous elements. (Naturally, it is understood
<i>that A ® A has the total grading and that R is graded trivially.) Finally, if A</i>
<i>and B are graded Hopf algebras and f:A—>B is a degree-preserving</i>
<i>homomorphism (of Hopf algebras), then we say that / is a homomorphism of</i>
<i>graded Hopf algebras.</i>


In the presence of a grading there is an important way in which our
definition of a Hopf algebra can be modified. In order to explain how this
comes about we shall make a completely fresh start.


<i>Assume then that A is an /^-module and that K-linear mappings</i>


<i>jtz: A ® A—+ A, rj: R—>A, A: A-+A ®A, and e: A—+R are given. Suppose</i>
<i>also that A is graded, as an R-module, by {An}neZ. Should it happen that</i>


<i>this makes (A, fi, rj) a graded algebra and (A, A, e) a graded coalgebra, then</i>
<i>the twisted product A ® A will exist both as a graded algebra and as a</i>
graded coalgebra. Furthermore, in these circumstances we can regard A as
<i>a mapping of A into A ® A and \i as a mapping of A ® A into A. With this</i>
<i>in mind, we say that the complete system forms a modified Hopf algebra or a</i>
<i>twisted Hopf algebra p r o v i d e d t h a t A : A^>A ®A a n d s:A—>R a r e</i>
<i>homomorphisms of (graded) ^-algebras and \i\ A ® A-^A and rj: R-+A</i>
are homomorphisms of (graded) K-coalgebras. Thus to modify the
<i>definition of a Hopf algebra we introduce a grading and then replace A ® A</i>
<i>by A ® A.</i>


For the new situation the counterpart of Lemma 5 is


<i><b>Lemma 6. Let (A, fi, rj) be an algebra, let (A, A, e) be a coalgebra, and let</b></i>


<i>both the algebra and the coalgebra be graded by {An}n€l. Then</i>


<i>(a) fi: A ®A-^A is compatible with comultiplication if and only if</i>
<i>A: A-^ A ®A is compatible with multiplication;</i>


<i>(b) \i\ A ® A —• A preserves counits if and only if s:A^>R is</i>
<i>compatible with multiplication;</i>


<i>(c) n:R-^A is compatible with comultiplication if and only if</i>
<i>A: A-^A ® A preserves identities;</i>


<i>(d) n\R—*A preserves counits if and only if e:A^>R preserves</i>


<i>identities.</i>


</div>
<span class='text_page_counter'>(167)</span><div class='page_container' data-page=167>

<i>156 Coalgebras and Hopf algebras</i>


<i>where T: A ®A^*A ®A is the twisting isomorphism, so that what we</i>
require is


A o )U = ()U<i> ® fi) o (A ® T ® A) ° (A (g) A). (7.8.6)</i>


<i>On the other hand, for A: A —• A ® A to be compatible with multiplication</i>
<i>it is necessary and sufficient that A°fi = fiA^A° (A ® A) and this is</i>


<i>equivalent to (7.8.6) because fiA^A = (/x ® //) ° (A ® T ® A).</i>


<i><b>Corollary. Suppose that {A, ft, rj) is an algebra and (A, A, e) is a coalgebra,</b></i>


<i>and let both be graded by {An}neI. Then \i\ A ® A—>A and Y\:R^>A are</i>


<i>homomorphisms of (graded) coalgebras if and only if A: A-^A ®A and</i>
<i>e: A—>R are homomorphisms of (graded) algebras.</i>


For modified Hopf algebras, the definitions of homomorphism and
isomorphism are the same as for ordinary (i.e. unmodified) Hopf algebras,
except that such mappings are now understood to preserve degrees. A
<i>modified Hopf algebra is termed skew-commutative if its algebra and</i>
coalgebra components are both skew-commutative.


<b>7.9 Tensor products of Hopf algebras</b>


<i>For 1 <i <p let (rjh ni9 Ah</i> Af, ef<i>) be a Hopf algebra (over the ring R)</i>



and put


<i>A=A1®A2®"-®Ap. (7.9.1)</i>


<i>Then A has a natural structure both as an algebra and as a coalgebra.</i>
Indeed, by (7.1.5) and (7.1.6),


<i>° Ap 9</i> (7.9.2)


<i>pA^\ (7.9.3)</i>


and, by (7.4.4) and (7.4.7),


<i>A^ = Vp o (Ax</i> ® A2 ® • • • ® Ap), (7.9.4)


<i>*A = HRP) o(e1®e2®'"® ep). (7.9.5)</i>


<i>(The notation is explained in Sections (7.1) and (7.4).) Now, because At</i> is a


Hopf algebra, Ax đ A2 đ ã • * ® Ap<i> is an algebra-homomorphism of Ax ®</i>


<i>A2®'"®Ap into (A1®A1)®(A2®A2)®'"®(Ap®Ap), and it</i>


<i>follows, from Theorems 2 and 3 of Chapter 3, that Vp</i> is an


algebra-isomorphism of


<i>(Ax ®A1)®(A2®A2)®-"®(Ap® Ap)</i>



<i>onto - (A1đA2đ-"đ Ap) đ(A1đA2đ-"đ Ap). Accordingly</i>


<i>A^: A ã A đ A is a homomorphism of i^-algebras. Next p^ (see (7.4.5)) is a</i>
<i>homomorphism of R-algebras and, of course, £1®e2®- - ®ep</i> is an


<i>algebra-homomorphism of Ax đ A2 đ ã ã ã ® Ap into R ® R ® • • • ® JR.</i>


</div>
<span class='text_page_counter'>(168)</span><div class='page_container' data-page=168>

<i>Tensor products of Hopf algebras 157</i>


<i><b>Theorem 9. Let A</b>l,A1,...,Ap be Hopf algebras over R. Put A =Al ®</i>


<i>A2®" ' ® Ap and define fiA,rjA, AA,eA as above. Then (nA, fiA,A,AA,eA) is</i>


<i>a Hopf algebra.</i>


<i>Proof. The theorem follows at once from the fact that A^: A —• A ® A and</i>
<i>eA:A-+R are algebra-homomorphisms by applying the corollary to</i>


Lemma 5.


<i><b>Definition. The Hopf algebra (rj</b>A, [iA, A, AA, sA) of the theorem is called the</i>


<i>'tensor product' of the Hopf algebras Al9 A2,..., Ap and it is denoted by</i>


<i>A1®A2®--®Ap.</i>


The corresponding result for modified Hopf algebras presents minor
complications. The following lemma will help us to deal with these.


<i><b>Lemma 7. Let A and B be modified Hopf algebras. Then A ®B, which is</b></i>



<i>certainly both a graded algebra and a graded coalgebra, is in fact a modified</i>
<i>Hopf algebra.</i>


<i>Proof. The comultiplication A ^ ^ o f t h e coalgebra A ®B, can be regarded</i>
<i>as a mapping of A ® B into (A ® B) ® (A ® B). Furthermore, it is given by</i>


A ^B<i> = (A ® T® B)o (AA</i> ® AB), (7.9.6)


<i>where T: A ® B^B ® A is the twisting isomorphism. Next, (7.9.6) can be</i>
obtained by piecing together the mappings


<i>A®B -^—^(A ®A)®(B®B)xA ® (A ®B)®B</i>


(7.9.7)


where the unlabelled isomorphisms are the obvious module-isomorphisms
<i>based on Theorem 1 of Chapter 2. However, AA: A—+A ®A and A</i>B<i>: B —></i>


<i>B ® B are algebra-homomorphisms by hypothesis and T. A®B-^*B®A</i>
is an algebra-isomorphism by Exercise 5 of Chapter 3. It follows that all the
mappings in (7.9.7) are homomorphisms of algebras and therefore


<i>AA®B: A ®B-+(A ®B) ® (A ® B)</i>
is a homomorphism of algebras as well.


<i>We turn now to eA 0 B</i>. This is the result of combining the mappings
<b>(2)</b>


<i>eAđeB HR</i>



<i><b>AđB ã R đR ã R .</b></i>


<i>Since R ®R and R®R are identical as algebras, we see that</i>


<i>£</i>


<i>A®B: A ®B—+R\s also an algebra-homomorphism. Finally, to complete</i>


</div>
<span class='text_page_counter'>(169)</span><div class='page_container' data-page=169>

<i>158 Coalgebras and Hopf algebras</i>


<i><b>Theorem 10. Let A</b>{1\ A{2\ . . . , A{p) be modified Hopf algebras. Then the</i>
<i>algebra and coalgebra structures on A(1) ® A(2) ® • • • ® A(p) interact in such</i>
<i>a way as to make it a modified Hopf algebra.</i>


<i>Proof. We use induction on p. When p = 1 the statement is tautologous so</i>
<i>we shall assume that p > 1 and that the result in question has been proved in</i>
<i>the case of p - 1 modified Hopf algebras. This secures that A(1) đ A{2)</i> đ ã ã ã


<i>đ A{p~</i>1} is a modified Hopf algebra and therefore, by Lemma 7, the same is
true of


<i>{A{1) đ A{2) đ ã ã ã đ Ato-V) ®A{p).</i>


<i>Next, Theorem 1 of Chapter 2 provides a degree-preserving </i>
<i>module-isomorphism</i>


(7.9.8)
However, by Theorem 6 of Chapter 3, this is an isomorphism of graded
algebras, and (this time by Theorem 5 of the present chapter) it is an


isomorphism of graded coalgebras. Since the right-hand side of (7.9.8) is a
modified Hopf algebra, we may conclude that the same holds for the
left-hand side. The inductive step is now complete and the theorem follows.


<i><b>7.10 E(M) as a (modified) Hopf algebra</b></i>


<i>Let M be an ft-module. We already know that E(M) is an</i>
<i>anticommutative, graded K-algebra and that S(M) is a commutative,</i>
graded K-algebra. But this is far from being the whole story. It will be
<i>shown, in this section, that the structure on E(M) can be extended so that</i>
<i>E(M) becomes a modified, skew-commutative, Hopf algebra; and later we</i>
<i>shall establish that S(M) is a commutative, graded, Hopf algebra. Our</i>
<i>reason for dealing with E(M) first is because the details in this case are a</i>
<i>little more complicated. When we come to discuss S(M) in a similar</i>
manner, the reader will find that there is no difficulty in adapting the
treatment provided for exterior algebras.


Suppose then that we are given an K-module M. Let us form its exterior
<i>algebra E(M) and let us identify E0(M) with # . The multiplication mapping</i>


<i>fi: E(M) ® E{M)^>E(M) of the algebra satisfies fi(x®y) = x Ay and the</i>
<i>unit mapping rj: R—+ E(M) is simply the identity mapping of R onto E0(M).</i>


Now let


<i>e:E(M)->R (7.10.1)</i>


<i>be the projection of E(M) onto E0(M). This mapping will turn out to be the</i>


</div>
<span class='text_page_counter'>(170)</span><div class='page_container' data-page=170>

<i>E(M) as a (modified) Hopf algebra 159</i>



The definition of comultiplication requires some preparation. The
<i>modified tensor product E(M) ® E(M) of E(M) with itself is an </i>
anti-commutative R-algebra (see Chapter 3, Theorem 7). Consequently the
<i>mapping <t>: M ã E(M) đ E(M) in which </>(m) = m ® 1 + 1 ® m is not only</i>
<i>R-linear, but also satisfies ((j)(m))2 = 0 for all m in M. Consequently <\> can be</i>
<i>extended to a homomorphism</i>


A: £ ( M ) - > £ ( M ) ® £(M) (7.10.2)
<i>o/ R-algebras which is degree-preserving and which is such that</i>


<i>A(m) = m ® 1 + 1 ® m (7.10.3)</i>
<i>when meM. The homomorphism A is called the diagonalization mapping of</i>
<i>£(M). Of course, since £(M) ® E(M) and £(M) ® E(M) coincide as</i>
modules (but not as algebras), A can also be considered as an /^-linear
<i>mapping of E(M) into E(M) ® E(M).</i>


<i><b>Lemma 8. With the above notation the triple (E(M), A, e) is a coalgebra.</b></i>


<i>This is graded, as a coalgebra, by the exterior powers {En(M)}neZ of the</i>


<i>module M. As a graded coalgebra (E(M), A, e) is skew-commutative.</i>


<i>Proof To see that the triple constitutes a coalgebra it suffices to verify that</i>
the diagrams


<i>E(M) > E(M)®E(M)</i>


<i><b>E(M)®A (7.10.4)</b></i>



A®£(M) ^


<i>E(M) ® E(M) > E(M) ® E(M) ® E(M)</i>


and


<i>^E(M).</i>


(7.10.5)


£®£(M) y


<i> E(M)®e</i>


<i>R ® E(M) < E(M) ® E(M) > E(M) ® R</i>


are commutative. However, all the mappings in (7.10.4) and (7.10.5) are
<i>algebra-homomorphisms and we know that the K-algebra E(M) is generated</i>
<i>by E1(M) = M. Consequently, when verifying that the diagrams are</i>


commutative, we need only examine what happens to an element of M. This
reduces the verifications to trivialities.


<i>Next, the exterior powers of M grade E(M) as a module and it is obvious</i>
<i>that, with respect to this grading, A and e preserve degrees. Accordingly</i>
(£(M), A, s) is a graded coalgebra.


</div>
<span class='text_page_counter'>(171)</span><div class='page_container' data-page=171>

<i>160 Coalgebras and Hopfalgebras</i>


<i><b>T</b></i>



<i>E(M) ® E(M) —> E(M) ® E(M)</i>


<i>E(M)</i>


<i>where T is the twisting isomorphism, commutes. But, once again, all the</i>
mappings are homomorphisms of algebras so that we need only show that
<i>T(A(m)) = A(m) for each element m in M. This, however, is clear.</i>


<i><b>Theorem 11. Let M be an R-module and let the notation be as above. Then</b></i>


<i>(rj, \i, E(M), A, e) is a modified Hopf algebra whose grading consists of the</i>
<i>exterior powers of M. As a modified Hopf algebra, E(M) is </i>
<i>skew-commutative.</i>


<i>Proof. We know from Chapter 5 that (£(M), /i, rj) is a graded algebra and</i>
<i>we have just proved (Lemma 8) that (E(M\ A, e) is a graded coalgebra.</i>
<i>Furthermore, we noted earlier that A: E(M)-*E{M) đ E(M) and</i>
<i>e: E(M) ã R are homomorphisms of algebras. It therefore follows, from the</i>
<i>corollary to Lemma 6, that (rj, //, E(M), A, e) is a modified Hopf algebra.</i>
<i>This is skew-commutative because (E(M)9 \x, rj) is an anticommutative</i>


algebra and, by Lemma 8, (£(M), A, e) is a skew-commutative coalgebra.
Thus the theorem is proved.


<i><b>Theorem 12. Let f:E—*N be a homomorphism of R-modules. Then the</b></i>


<i>mapping E(f): E(M)—+E(N) defined in Section (5.2) is a homomorphism of</i>
<i>modified Hopf algebras.</i>



<i>Proof. We know, from Chapter 5, that E(f) is a homomorphism of graded</i>
algebras. We also know that


and e£(M)<i>: E(M)~>R are homomorphisms of algebras and that a similar</i>


<i>observation applies to AE[N) and eE(N). This makes it very easy to verify that</i>


<i>£ ( / ) : E(M)^>E(N) is a homomorphism of graded coalgebras. But this is</i>
all that is needed to complete the proof of the theorem. The details are left to
the reader.


<b>7.11 The Grassmann algebra of a module</b>


Let M be an R-module. Then, by the results of the last section,
<i>E(M) is a modified Hopf algebra, and so inter alia it is a graded coalgebra.</i>
We recall that its comultiplication mapping A satisfies


A(m) = m ® 1 + 1® m


</div>
<span class='text_page_counter'>(172)</span><div class='page_container' data-page=172>

<i>The Grassmann algebra of a module 161</i>


<i>E0(M). Furthermore, A: £ ( M ) - > £ ( M ) ® E(M) and e:E(M)-^R are</i>


homomorphisms of /^-algebras.


By Theorem 8 we can now obtain a graded R-algebra by considering the
<i>linear forms on E(M). It is this algebra that we are about to investigate.</i>
<i>Suppose that p> 1 is an integer. In what follows I = (il9</i> i2<i>, • • •, is) will</i>


<i>denote a sequence of s (0 < 5 <p) integers satisfying 1 < i1 < i2 < • • * < is <p,</i>



<i>and by /' we shall understand the sequence obtained from ( 1 , 2 , . . . , p) by</i>
deleting the terms that belong to /. Furthermore, if /' = (*i, i'2,...)> then


sgn(7,/') will denote the sign of ( i1 ?. . . , is, ii,i"2,...) considered as a


<i>permutation of ( 1 , 2 , . . . , p).</i>


<i>Now let ml9</i> m2<i>, . . . , mp belong to M. Put mI = mii A mi2</i> A • • • Amfj and


<i>define mr</i> similarly, it being understood that the natural conventions apply


<i>when s = 0 and s=p.</i>


<i><b>Lemma 9. With the above notation</b></i>


Am2<i> A • • • Amp) = YJ</i> sgn(J, 7')m7


<i>Proof. Since A: E(M)—>E(M) ®E(M) is a homomorphism of algebras,</i>
<i>A(ml Am2 A • • • A mp) is the product, in E(M) ® E(M), of A(ml)9A(m2),...,</i>


A(mp), that is to say it is the product of (mx<i> ® 1 -f 1 (g) mx</i>), (m2<i> ® 1 + 1 ® m2)</i>


and so on. Accordingly A(mx A m2 A • • • A mp) is the sum, taken over all


sequences / = (il9 i2, . . . , is), of the product in £(M) ® £(M) of the elements


1 ® mx, . . . , mfi<i> ® 1 , . . . , mis ® 1 , . . . , 1 ® mp.</i>


<i>But £(M) ® E(M) is an anticommutative algebra and here we are</i>


concerned with homogeneous elements of degree one. Consequently the
product in question is equal to sgn(/, /') times the product of


mfi<i> ® 1, mi2</i> ® 1 , . . . , 1 ® mi;<i>, 1 ® mi2,...</i>


<i>and therefore it has the value sgn(7,1')mI</i> ® m7'. The lemma follows.


We now put


<i>E(M)* = Hom</i>i?<i>(£(M), R) (7.11.1)</i>


<i>so that E(M)* consists of all the linear forms on E(M); we also put</i>
<i>En(M)* = HomR(En(M),R) (7.11.2)</i>


<i>and identify En(M)* with the module formed by the linear forms on E(M)</i>


<i>that vanish on Ek(M) for all k^n.</i>


<i>By Theorem 7, E{M)* is an JR-algebra whose identity element is the linear</i>
<i>form e: E(M) —> R which projects E{M) onto E0(M). For / and g in E(M)*</i>


<i>we shall use / A g to denote their product in this algebra. Naturally / A g is</i>
<i>also a linear form on E(M) and, by (7.7.4),</i>


</div>
<span class='text_page_counter'>(173)</span><div class='page_container' data-page=173>

<i>162 Coalgebras and Hopf algebras</i>


<i>(We recall that A is the comultiplication mapping of E(M) and that fiR</i> is the


<i>multiplication mapping of R.)</i>



<i><b>Lemma 10. Let m</b>l9</i> m2<i>, . . . , mp (p>0) belong to M and let f g belong to</i>


<i>£(M)*. Then (with the same notation as in Lemma 9)</i>


<i>Am2</i> A • • • Amp) = £ sgn(/, /')


<i>Remark. When p = 0 this is to be interpreted as asserting that ( / A g)(i) =</i>


<i>Proof. The desired result follows by combining (7.11.3) with Lemma 9.</i>


<i><b>Lemma 11. Let /</b></i>l 9/2, . •. , /p<i> ( p > l ) belong to M* = E1(M)* and let fx</i> A


<i>f2A"'Afpbe their product in the algebra £(M)*. Furthermore, let m</i>1? m2,


<i>. . . , mp belong to M. Then</i>


<b>( / i A /2 A - - - A /p) ( m ! A m2 A - - - )</b>


/p(m2)


(7.11.4)


<i>Proof. We use induction on /?. The assertion is obviously true when p = 1 so</i>
<i>we shall assume that p > 1 and that the lemma has been established for all</i>
<i>smaller values of the inductive variable. Put g = f2A"'AfP' Since</i>


<i>fxim^ Am</i>V2<i> A ••• Amv) = 0 except when h = l, it follows, by Lemma 10,</i>


that



<b>A m2 A - - - A mp)</b>


<i>/\mp)</i>
<b>( / i A /2A </b>


<i><b>-p</b></i>


<b>/2K)</b>


<i>because, by the inductive hypothesis, g(mt A • • • A mt</i> A • • • A mp) is equal to


the determinant obtained by striking out the first row and the i-th column
in (7.11.4).


Let us now put


(7.11.5)


</div>
<span class='text_page_counter'>(174)</span><div class='page_container' data-page=174>

<i>S(M) as a Hopf algebra 163</i>


(7.11.6)
<i>Then, by Theorem 8, G(M) is an K-subalgebra of E(M)* and it has</i>
<i>{Gn(M)}neZ</i> as an algebra-grading. Since


G0(M) = Hom^(£0(M),JR) = Homi?(^,JR), (7.11.7)


<i>G0(M) is a free R-module of rank one. Note that e is the identity element of</i>


<i>G(M) and that it forms a base for G0(M). We also have</i>



<i>Gl(M) = HomR(El(M)iR) = M*. (7.11.8)</i>


<i>Finally, we note that if M is a finitely generated R-module, then G(M) =</i>
<i>E(M)*. This is because finite generation ensures that En</i> (M) = 0 for all large


<i>values of n.</i>


<i><b>Definition. The graded R-algebra G{M) is called the 'Grassmann algebra' of</b></i>


<i>M.</i>


<i><b>Theorem 13. The Grassmann algebra G(M) is an anticommutative algebra.</b></i>


<i>Proof. By Lemma 8, E(M) is a skew-commutative coalgebra and therefore,</i>
<i>by Theorem 8 Cor., G(M) is a skew-commutative algebra. Now suppose</i>
<i>that feGp(M), where p is an odd integer. Then, to complete the proof, we</i>


<i>need only show that / A / = 0. Suppose therefore that mi,m2,... , m</i>2p


<i>belong to M. Since / A / belongs to G2p(M), the theorem will follow if we</i>


establish that


( / A / M m ! Am2 A - - - Am2p) = 0.


But, by Lemma 10,


( / A / ) ( mx Am2<i> A • • • Am2p) = YJ</i> sgn(/, //) / ( w/) / ( mr) ,


<i>where (because feGp(M)) we have only to sum over all sequences / =</i>



(il5<i> i2, • • •, ip) of length p with ! < / ! < • • • <ip<2p. (Here / ' is obtained by</i>


<i>deleting the terms in / from ( 1 , 2 , . . . , 2p).) But, for such an /, sgn(7, /') =</i>
<i>— sgn(/', /) because p is odd, and therefore</i>


<b>( / A / ) ( * * ! ! A m2 A ' " Am2 p) = 0</b>


as required.


<i><b>7.12 S(M) as a Hopf algebra</b></i>


<i>Let M be an K-module and let S(M) be its symmetric algebra. We</i>
<i>propose to show how the structure of S(M) as an algebra can be</i>
complemented so that the final result is a graded Hopf algebra. In what
<i>follows we identify S0(M) with R.</i>


<i>Let s: S(M)-^R be the projection of S(M) onto S0(M). This will be the</i>


</div>
<span class='text_page_counter'>(175)</span><div class='page_container' data-page=175>

<i>164 Coalgebras and Hopf algebras</i>


<i>To define the comultiplication or diagonalization mapping on S(M), we</i>
<i>first observe that there is an R-linear mapping M —> S(M) ® S{M) in which</i>
<i>m, in M, becomes m ® 1 + 1 ® m. However, S(M) ® 5(M) is a commutative</i>
algebra and therefore the /Minear mapping can be extended to a
homomorphism


A:S(M)->S(M)®S(M) (7.12.1)
of algebras. We record that, by construction,



A(m) = m ® l + l ® m (7.12.2)
for all m in M.


<i><b>Lemma 12. With the above notation, (S(M), A, e) is a coalgebra. It is graded,</b></i>


<i>as a coalgebra, by the symmetric powers {Sn(M)}neI of the module M. The</i>


<i>coalgebra (S(M), A, e) is commutative.</i>


<i>Proof. This lemma is the counterpart of Lemma 8 and it can be proved in</i>
very much the same way. The details will therefore be omitted.


<i>In the next theorem, \i and rj denote the multiplication and unit mappings</i>
of the symmetric algebra.


<i><b>Theorem 14. Let the notation be as explained above. Then (rj, JJL, S(M), A, s)</b></i>


<i>is a graded Hopf algebra, the grading being given by the symmetric powers</i>
<i>{Sn(M)}neI of the module M. As a Hopf algebra, S(M) is commutative.</i>


<i>Proof That S(M) is a Hopf algebra follows from the corollary to Lemma 5.</i>
It is obviously commutative.


<i><b>Theorem 15. Let f: M —• N be a homomorphism of R-modules. Then the</b></i>


<i>mapping S(f): S(M)-^S(N), which was defined in Section (6.2), is a</i>
<i>homomorphism of graded Hopf algebras.</i>


<i>Proof Theorem 15 corresponds to Theorem 12. In the case of Theorem 12</i>
we described the essential features of the demonstration. Similar


considera-tions apply here.


<i>In Section (7.11) we used the fact that E(M) is a graded coalgebra in order</i>
<i>to derive from it the Grassmann algebra of M. In like manner we can use the</i>
<i>graded coalgebra S(M) to derive what we might expect to be a new algebra.</i>
Let us see what the construction produces.


To this end we consider


<i>Un = Sn(M)* = HomR(Sn(MlR) (7.12.3)</i>


<i>as a submodule of S(M)* = HornR(S(M),R) and put</i>


H=IH, (7.12.4)


<i>nel</i>


</div>
<span class='text_page_counter'>(176)</span><div class='page_container' data-page=176>

<i>S(M) as a Hopf algebra 165</i>


by the family {9lw}neZ. Furthermore, the corollary to Theorem 7 shows that


<i>the algebra is commutative. For / , g in 91 we shall use fg to denote their</i>
product. Then, by (7.7.4),


<i>fg = HR°(f®g)°A, (7.12.5)</i>
<i>where A is the diagonalization mapping of S(M).</i>


Next, we recall that in Section (6.7) we defined an algebra V5(M) which
we named the algebra of differential operators. By construction it is a
commutative subalgebra of End^(5(M)) and, moreover, it is graded. The


<i>module of differential operators of degree h was denoted by V</i>h5(M).


<i>Suppose that </>eVhS(M), where h>0. Then </> is an K-endomorphism of</i>


<i>S(M) of degree —h and it induces a linear form, ^ say, on Sh(M). Also, by</i>


<i>Theorem 8 of Chapter 6, the mapping 0h-x/> is an isomorphism of VhS(M)</i>


<i>onto Sh(M)*. Accordingly we have an isomorphism</i>


^ > 91, (7.12.6)


<i>for each value of h and, by combining these isomorphisms, we arrive at a</i>
degree-preserving isomorphism


(7.12.7)


of ^-modules.


<i><b>Theorem 16. The module-isomorphism (7.12.7) is actually a </b></i>


<i>degree-preserving algebra-isomorphism of the algebra VS(M) of differential</i>
<i>operators onto the algebra 91.</i>


<i>Proof Suppose that 0eV</i>fcS(M) and i^eVk<i>S(M), where h>0 and /c>0.</i>


<i>Then </> ° \jj is a differential operator of degree h + k and its image in 91,,+k</i> is


<i>its restriction </>°^ to Sh+k(M). The theorem will follow if we show that</i>



<i>(j) ° \\t is the product of cj) and ^ in 91.</i>


To see this let ml5 m2<i>, . . . , rnh+k</i> belong to M. It will now suffice to prove


<i>that {(f)°\jj)(m1m2 . . . rnh+k) and (^{j/)(m1m2 .. .mh+k) are the same. In this</i>


connection we note that
<i>...mh+k)</i>


imi 2<i>. . . mih)il/(mjmJ2... mjk),</i>


<i>where (i1,i2,--4h)</i> anc* (/iJ2>""Jk) constitute a variable pair of strictly


increasing sequences of integers which, between them, include all the
<i>positive whole numbers from 1 to h + k (see (6.7.6)). Next, from (7.7.4), we see</i>
<i>that (j)ij/ = tiR ° (0 (g) ^ ) ° A, where in this context 0 and ij/ are regarded as</i>


</div>
<span class='text_page_counter'>(177)</span><div class='page_container' data-page=177>

<i>166 Coalgebras and Hopf algebras</i>


<i>S{M) ® S(M), of the elements</i>


mf<i> ® 1 + 1 ® mt (i = 1 , 2 , . . . , h</i>


<i>Of course, these elements are of degree one and S(M) ® 5(M) is a</i>
<i>commutative algebra; also </> is null on Sn(M) if n ^h, whereas \j/ is null on</i>


<i>Sn(M) if n=£k. From these observations we conclude that</i>


<i><b>((f ®\j/)o A)(</b><b>mi</b><b>m</b><b>2</b><b> ... m</b><b>h+k</b><b>)</b></i>



where the sum is taken over the same pairs of sequences as before. But
<i>extends </> and \j/ extends \j/. Consequently</i>


<i><b>= £ ^(m^m^ . . . m</b><b>ih</b><b>)\l>{m</b><b>h</b><b>m</b><b>h</b><b> . . . m</b><b>Jk</b><b>)</b></i>


and with this the proof is complete.


<i>Thus, to sum up, when we consider linear forms on the coalgebra S(M)</i>
we do not obtain an essentially new graded algebra; instead we recover up
<i>to isomorphism the algebra of differential operators, on S{M), that we</i>
encountered in Chapter 6.


<b>7.13 Comments and exercises</b>


It will be clear to the reader, and indeed it has been mentioned
already, that the analogy between the theory of algebras and the theory of
coalgebras can be developed further than was done in the main text. Thus
our first exercise corresponds to the fact that the tensor product of a number
of commutative algebras is itself commutative, and the second exercise is
suggested by Theorem 7 of Chapter 3. The two exercises can be solved in
very much the same way, so a solution will be provided only for the second
one. As on earlier occasions, those exercises for which a solution is given are
marked with an asterisk.


<i><b>Exercise 1. Show that if At, A</b>2, • . . , Ap are commutative R-coalgebras, then</i>


<i>Al đ A2 đ ã * • ® Ap is also a commutative R-coalgebra.</i>


<i><b>Exercise 2*. Show that if A</b>{1\ A{2\..., Aip) are skew-commutative </i>
<i>R-coalgebras, then A{1) ® A{2) ® • • • ® A{p) is also a skew-commutative R~</i>


<i>coalgebra.</i>


</div>
<span class='text_page_counter'>(178)</span><div class='page_container' data-page=178>

<i>Comments and exercises 167</i>


<i>modified tensor product remains unchanged to within an isomorphism of</i>
<i>graded coalgebras.</i>


<i>The formal difference between the definition of an anticommutative</i>
<i>algebra and that of a skew-commutative algebra is clear enough. The next</i>
exercise invites the reader to construct an example to show that the
distinction is a real one.


<i><b>Exercise 4*. Give an example of a skew-commutative algebra which is not</b></i>


<i>anticommutative.</i>


<i>Next we return to the original definition of a graded algebra. Suppose</i>
<i>then that A is an K-algebra and {An}neI is a family of R-submodules of A</i>


<i>that grades A as a module. By the definition given in Section (3.3), {An}neZ</i>


<i>grades A as an algebra if from ar e Ar and as e As it follows that aras</i> belongs


<i>to Ar+S. Thus we have an algebra-grading provided only that the</i>


<i>multiplication mapping ix: A® A-+A preserves degrees. However, by</i>
<i>Theorem 5 of Chapter 3, in these circumstances the identity element of A is</i>
<i>homogeneous of degree zero and so the unit mapping rj.R^A </i>
auto-matically preserves degrees as well. The following exercise shows that a
similar situation exists in the case of coalgebras.



<i><b>Exercise 5*. Let A be an R-coalgebra and let {A</b>n}neI be a family of </i>


<i>R-submodules of A such that</i>


<i>A=YJAn</i> (d.s.).
<i>neZ</i>


<i>Further, let the comultiplication mapping A: A —• A (x) A preserve the degrees</i>
<i>of homogeneous elements. Show that if R is given the trivial grading, then the</i>
<i>counit mapping a: A-+R also preserves degrees.</i>


We now add some comments concerning the results of Section (7.7).
<i>There we saw that if A is an R-coalgebra and A* = Hom</i>R<i> (A, R), then A* has</i>


a natural structure as an R-algebra. Let us examine the functorial aspects of
<i>the way the algebra A* depends on the coalgebra A.</i>


<i>First we note that each ^-linear mapping </>: A—+B, of A into an </i>
<i>R-module B, gives rise to a R-module-homomorphism</i>


<i></>*:B*-+A* (7A3A)</i>
in which <£*(/) = / °(/>.


<i><b>Exercise 6*. / / (j>: A—>B is a homomorphism of R-coalgebras, show that</b></i>


<i>cj)*: B*^>A* is a homomorphism of R-algebras.</i>


</div>
<span class='text_page_counter'>(179)</span><div class='page_container' data-page=179>

<i>168 Coalgebras and Hopf algebras</i>



<i>Let M be an ^-module. In Theorem 16 we saw that when S(M) is</i>
regarded as a graded coalgebra and we use Theorem 8 to derive from it a
graded algebra, what we obtain is isomorphic to the algebra of differential
<i>operators on S(M); in particular the application of Theorem 8 produces an</i>
<i>algebra that is isomorphic to a certain graded subalgebra of EndR(S(M)).</i>


<i>This suggests that the Grassmann algebra G(M) may be related to some</i>
<i>graded subalgebra of EndR(E(M)). As will now be shown this is indeed the</i>


case.


<i>Suppose that h > 0 and consider the K-endomorphisms of E(M) of degree</i>
<i>— h. Such an endomorphism induces a linear form on Eh(M). In what</i>


<i>follows p denotes an integer satisfying p > h, I = (i</i>x<i>, i2,..., ih) is a sequence</i>


<i>of integers with 1 <*! <i2<'' • <ih<p, and / ' denotes the increasing</i>


<i>sequence obtained from ( 1 , 2 , . . . , p) by deleting the terms that belong to /.</i>
<i>Finally, if m1,m2,..., mp</i> belong to M, then m7 will be used for the element


<i>mti A mi2</i><b> A • • • A/Mj and m</b>r is to be defined similarly.


<i>By a differential form on E(M) of degree h, we shall understand an </i>
<i>R-endomorphism cj) of degree — h that has the following property: for all p > h</i>
<i>and all mx, m</i>2, . . . , mp in M


^(m! Am2 A - - - Amp) = £sgn(/,/')</>(w/)"V- (7.13.2)


<i>As usual, if I = (il,i2<b>,...) and r = (i'</b>l9i2,.. .)> then sgn(/,/') stands for the</i>



sign of (il5<i> i2,..., i\, i'</i>2,...) considered as a permutation of ( 1 , 2 , . . . , p).


<i>Evidently the differential forms of degree h constitute an K-submodule of</i>
EndK<i> (£(M)); and if two differential forms of degree h induce the same linear</i>


<i>form on Eh(M), then they coincide.</i>


<i><b>Exercise 7*. Let (j) and if/ be differential forms on E(M) of degrees h>0and</b></i>


<i>k>0 respectively. Show that ij/ ° (/> and (f)o{j/ are differential forms of degree</i>
<i>h + k and that xj/ ° 0 = (— \)hk(j) ° \j/. Show also that if h is odd, then 0 ° 0 = 0.</i>


The next exercise should be compared with Theorem 8 of Chapter 6.


<i><b>Exercise 8*. Let M be an R-module and f a linear form on E</b>h(M), where</i>


<i>h>0. Show that, on E(M), there is exactly one differential form of degree h</i>
<i>that extends f</i>


<i>From Exercises 7 and 8, we see that there is a subalgebra of End# (E(M))</i>
that is non-negatively graded by the modules of differential forms. Let us
<i>call this the algebra of differential forms on E(M). By Exercise 7, the algebra</i>
is anticommutative.


<i>Let (/> be a differential form on E(M) and let the degree of <f> be h. Then <j></i>
<i>restricts to a linear form, </> say, on Eh(M). Note that 0 belongs to the</i>


</div>
<span class='text_page_counter'>(180)</span><div class='page_container' data-page=180>

<i><b>Solutions to selected exercises 169</b></i>



<i><b>Exercise 9*. Let <f> and i// be differential forms on E(M). Show that the linear</b></i>


<i>form \f/o(f> associated with i// ° cj) satisfies \\i ° </> = 0 A \j/, where (j> A \j/ is the</i>
<i>product of (j> and \j/ in G(M).</i>


<i>Let A be a ring with identity and let A be the ring that is opposite to A.</i>
<i>(Thus A and A coincide as sets and addition is the same for both of them;</i>
<i>however, the product of two elements of A is their product in A but with the</i>
<i>order of the factors reversed.) If A happens to be an fl-algebra, then A will</i>
<i>also be an K-algebra with the same R-module structure as A. Naturally we</i>


<i>refer to A as the opposite algebra to A. Finally, if {An}neZ grades A as an </i>


R-algebra, then the same family provides an algebra-grading on the opposite
algebra.


This terminology enables us to sum up our main conclusions as follows:


<i>the Grassmann algebra G(M) and the opposite of the algebra of differential</i>
<i>forms on E(M) are isomorphic graded algebras.</i>


<b>7.14 Solutions to selected exercises</b>


<i><b>Exercise 2. Show that if A</b>(1\ A{2\... ,Aip) are skew-commutative </i>
<i>R-coalgebras, then A(1) ® Ai2) ® * • • ®A(p) is also a skew-commutative </i>
<i>R-coalgebra.</i>


<i>Solution. The property of being skew-commutative is preserved under an</i>


isomorphism of graded coalgebras, and because of this it is enough to prove


<i>the result in question in the case of two coalgebras. We shall therefore</i>
<i>assume that A and B are skew-commutative coalgebras and prove that</i>


<i>A ® B is skew-commutative as well.</i>


<i>We use our usual notation. Let am belong to Am and bn to Bn. Further let</i>


<i>&A(</i>flm) = If l' ®<i> a"</i> a r id AB<i>(bn) = £ b' ® ft", where a', a" are homogeneous</i>
<i>elements of A and b\ b" homogeneous elements of B. Then</i>


<i>*A9B(am ® K) = l (~ l){a"m(a' ® b') ® (an ® b"),</i>


<i><b>where it is understood that {a</b><b>r</b><b>}, {a") denote the degrees of a! and a", and</b></i>


<i>similarly in the case of {b'} and {b"}.</i>


<i>Let TA:A®A^>A ® A, TB\B®B^+B®B and</i>
<i>T: (A ®B)®(A ® B) -^-> (A ® B) ® {A ® B)</i>


be twisting isomorphisms. Then


<i>(T°AA®B)(am®bn)</i>


</div>
<span class='text_page_counter'>(181)</span><div class='page_container' data-page=181>

<i>170 Coalgebras and Hopf algebras</i>


and now it follows that


<i>b") ® (a' ® V)</i>


<i>Accordingly T° AA0B = AA0B</i> and with this the solution is complete.



<i><b>Exercise 4. Give an example of a skew-commutative algebra which is not</b></i>


<i>anticommutative.</i>


<i>Solution. Let F be a field of characteristic two and let A be the free </i>
<i>F-algebra generated by a set {X,Y} of two elements. Then XY+YX</i>
<i>generates a homogeneous two-sided ideal, / say, in A. Let B denote the</i>
<i>graded F-algebra A/1 and denote by x and y the natural images of X and Y</i>
<i>in B.</i>


<i>The homogeneous elements of B that have degree one form the vector</i>
<i>space B^ = Fx + Fy, and this generates B as an algebra. By construction</i>
<i>xy + yx = 0 and now, because F has characteristic two, we have bb' + b'b = O</i>
<i>for all b, b' in Bv</i> That £ is skew-commutative is a simple consequence of


this.


<i>Finally X2$I and therefore x</i>2<i>=^0. This shows that B is not</i>
anticommutative.


<i><b>Exercise 5. Let A be an R-coalgebra and let {A</b>n}neI be a family of </i>


<i>R-submodules of A such that</i>


<i>Further, let the comultiplication mapping A: A —• A ® A preserve the degrees</i>
<i>of homogeneous elements. Show that if R is given the trivial grading, then the</i>
<i>counit mapping e: A-+R also preserves degrees.</i>


<i>Solution. Suppose that ameAm, where m / 0 . It is sufficient to show that</i>



<i>s(am) = 0.</i>


We can write A(flm) = ]T af<i> ® fii9 where a,-, pt</i> are homogeneous elements


<i>of A the sum of whose degrees is m. Now, because £ is the counit mapping,</i>
we have


flm = Zfi(«i)A = Ze(i8l)a< (7.14.1)


and therefore e(am) = ^e(ai)£(j?l).


<i>Let us use {a,}, respectively {Pt}, to denote the degree of a</i>f, respectively


</div>
<span class='text_page_counter'>(182)</span><div class='page_container' data-page=182>

<i>Solutions to selected exercises 111</i>


and therefore


£ e(af)£(ft) = 0. (7.14.2)


<b>/:{a,-}*0</b>


But, again from (7.14.1),


and so


<b>/:{«,} = 0</b>


If now we combine this with (7.14.2) we find that £ e(ai)e()8f) = 0, that is to



say a(am) = 0. This completes the solution.


<i><b>Exercise 6. / / </>: A-^B is a homomorphism of R-coalgebras, show that</b></i>
<i>(/>*: B*—>A* is a homomorphism of R-algebras.</i>


<i>Solution. With the usual notation we have (j)*(eB) = sB°(j) = sA because <j></i>


preserves counits. Consequently 0* preserves identity elements.


<i>Now suppose that hx and h2 belong to B*. By (7.7.4), their product in this</i>


<i>algebra is fiR ° (h1 ® h2) ° A</i>B and, moreover,


because $ is compatible with comultiplication. But


<i>= ^ ° ^ i ) ® <j)*(h2)°AA</i>


and this is just the product of (^(/i^ and 0*(/i2) in 4*- Thus 0 * is


compatible with multiplication and the solution is complete.


<i><b>Exercise 7. Let (j) and \j/ be differential forms on E(M) of degrees h>0 and</b></i>


<i>k>0 respectively. Show that i// ° (j) and (po\j/ are differential forms of degree</i>
<i>h + k and that \jj ° </> = ( - 1)^0 ° i//. Show also that if h is odd, then (j) ° (j) = 0.</i>


<i>Solution. Evidently \jj ° </> and 0 ° i/^ are endomorphisms of £(M) and each</i>
<i>has degree — (ft + /c). Now suppose that p>h + k and that ml9 m2,... ,mp</i>


<i>belong to M. In what follows I = (il9 i2,..., ih) and J=UiJ2,... ,7</i>fc) will



<i>denote a pair of sequences of integers, where 1 < ix < • • • < ih < p, 1 <7\ < • • •</i>


<i><jk<p and / and J have no term in common. We put mI = mii A mi2</i> A • • •


A m^ and define my<i> similarly. Further we use /', respectively J\ to denote the</i>


</div>
<span class='text_page_counter'>(183)</span><div class='page_container' data-page=183>

<i>172 Coalgebras and Hopf algebras</i>


By (7.13.2),


^(m! Am2 A - - - Amp) = ^ s g n ( / , / ' ) 0 ( m/)


/


<i>and now if we operate on this with \j/ we obtain</i>


<i><b>--- Am</b><b>p</b><b>)</b></i>


= £ sgn(/, /') sgn(J,


<i><b>u</b></i>


<i>Note that the sequence J followed by the sequence (IJ)' is a permutation of</i>
<i>/'. It is the sign of this permutation that is denoted by sgn(J, (IJ)').</i>


<i>Now /, J and (IJ)', taken in order, provide a permutation of ( 1 , 2 , . . . , p).</i>
<i>If sgn(/, J, (U)f) denotes the sign of this permutation, then</i>


<i>sgn(/, /') sgn(J, (/J)') = sgn(J, J, (IJ)')</i>


and therefore


<i>(if/ ° ^)(m! Am</i>2 A - - -<i> Amp)</i>


= £ sgn(/, J, (/jy^K^^K,), (7.14.3)


Similarly we find that


X K ^ (7.14.4)
<i>and from (7.14.3) and (7.14.4) it is clear that \\i ° 0 = ( - l)*</i>k0 °


<i><A-For the last step we assume that h is odd and we take \j/ = </> so that ft = /c.</i>
By (7.14.3),


(0 °0)(m1 Am2 A • • • Amp) =


<i>u</i>
= 0
because, under the present assumptions,


sgn(/, J, (JJ)') + sgn(J, /, (J/)') = 0.
Accordingly 0 ° (/> = 0 as required.


<i>Exercise 8. Let M be an R-module and f a linear form on Eh(M\ where</i>


<i>h>0. Show that, on E(M), there is exactly one differential form of degree h</i>
<i>that extends f</i>


</div>
<span class='text_page_counter'>(184)</span><div class='page_container' data-page=184>

<i>Solutions to selected exercises 173</i>



where the notation is the same as in (7.13.2). This mapping is clearly
multilinear.


<i>We claim that it is also an alternating mapping. For suppose that \<j<p</i>
<i>and that m}f</i> = m7- +<i> v If both j and j+1 are in 7,thenm</i>7 = O;andifneitherisin


/ then m7<i>' = 0. Thus in both these cases f(mI)mI=0.</i>


<i>The remaining /'s are those which contain just one of j and j +1. Such</i>
<i>sequences can be arranged in pairs, Ix and I2</i> say, where each member of a


<i>pair becomes the other member when j and j + 1 are interchanged. Now</i>
/(w/^m^ is equal to /(m/2)my2<i>, because mj = mj+l; and</i>


s g n ( /1, / i ) = - s g n ( /2, //2) .


Consequently


sgn(/1?<i> ry)f(mh)mVx</i> + sgn(/2<i>, I2)f(ml2)ml2 = 0</i>


and therefore we have


Xsgn(/,/')/(m7)mr = 0.


This establishes our claim (see Chapter 1, Lemma 2).


<i>It now follows that there is an R-linear mapping of Ep(M) into Ep_h(M)</i>


in which



<i>mx</i> A<i> m2</i> A •••


When /? = /* this is just / Accordingly if we combine these various
<i>homomorphisms into an endomorphism of E(M) it will extend / ;</i>
moreover, by construction, the endomorphism so obtained will be a
<i>differential form of degree h. This provides all that is needed because the</i>
part of the exercise that is concerned with uniqueness has been dealt with
already.


<i><b>Exercise 9. Let (j) and \\/ be differential forms on E(M). Show that the linear</b></i>


<i>form \\i ° (/> associated with \\i ° (j) satisfies i// °</> = (j> A \j/9 where <p A \j/ is the</i>


<i>product of (j) and ij/ in G(M).</i>


<i>Solution. Let the degrees of (/> and \jj be h > 0 and k > 0 respectively, and let</i>
<i>ml9</i> m2<i>, . . . , mh+k belong to M. By (7.13.2),</i>


<b>Am2 A</b>


and therefore


</div>
<span class='text_page_counter'>(185)</span><div class='page_container' data-page=185>

<i>174 Coalgebras and Hopf algebras</i>


<i>Here / = (ix, i2,..., ih) is an increasing sequence of integers between 1 and</i>


<i>h + k and I denotes the residual sequence.</i>


<i>We now consider (</> A\j/)(ml</i> Am2<i> A • • • Amh+k). This can be obtained</i>



<i>from Lemma 10; and, if we remember that <^ and \jj (considered as elements</i>
<i>of £(M)*) belong to Eh(M)* and Ek(M)* respectively, we find that the</i>


relation provided by the lemma is


! Am2<i> A • • • Amh+k) = YJ</i> s g n ( / , / ' ^ ( m ; ) ^ ^ ) ,


where / ranges over the same set of sequences as before. However, 0 (mz) =


</div>
<span class='text_page_counter'>(186)</span><div class='page_container' data-page=186>

<b>8</b>



Graded duality



<b>General remarks</b>


From a graded coalgebra it is possible, by considering linear forms
on the submodules which make up the grading, to derive a graded algebra.
In fact, this is the conclusion of Theorem 8 of Chapter 7. An attempt to
perform a similar construction starting instead with a general graded
algebra soon runs into difficulties; but when the algebra is suitably
restricted the process can be carried through successfully to produce a
graded coalgebra as the end-product.


In the present chapter these ideas are developed in detail and it is shown
that, for non-negatively graded algebras and coalgebras whose grading
modules are free modules of finite rank, one has a full duality theory. This
not only interchanges algebras and coalgebras, but also preserves ordinary
and modified Hopf algebras.


The first two sections of the chapter are used to establish the results on


<i>modules that provide the basis of this theory. As always, R denotes a</i>
commutative ring with an identity element, and the symbol ®, when used
<i><b>without a subscript, indicates a tensor product formed over R.</b></i>


<b>8.1 Modules of linear forms</b>


It was remarked in the introduction to this chapter that the duality
we shall be discussing arises from the study of linear forms on free modules
with finite bases. However, before we restrict our attention to such modules
it is convenient to make some observations concerning linear forms on
arbitrary modules.


<i>Suppose then that M and N are R-modules and put</i>


<i>M* = HomR(M,R), (8.1.1)</i>


<i>N* = HomR(N, R) so that M* and N* are the so-called duals of M and N</i>


</div>
<span class='text_page_counter'>(187)</span><div class='page_container' data-page=187>

<i>176 Graded duality</i>


<i>respectively. If now k: M—>N is K-linear, then the dual homomorphism</i>
<i>k*:N*->M* (8.1.2)</i>


is defined by


<i>k*((t)) = <j)°k (8.1.3)</i>


<i>for all <j> in N*. Note that if id</i>M denotes the identity mapping of M, then


(idM)* = idM*; (8.1.4)



<i>and that if we have a second homomorphism, \i\ N-^K say, then</i>


(^°/l)* = /l*0/**. (8.1.5)


<i>(In the language of Category Theory, the taking of duals is a contravariant</i>
<i>functor from ^-modules to R-modules.)</i>


<i>If we consider R as a module and form its dual R*, then this is isomorphic</i>
<i>to R itself; in fact we have an isomorphism</i>


<i>R**R (8.1.6)</i>


<i>in which / in R* is matched with / ( I ) in R. This will be called the canonical</i>
<i>isomorphism of R* onto R and, in the sequel, it will often be used to identify</i>
<i>R* with R.</i>


<i>An ^-module M is connected with its double dual M** by means of an </i>
R-linear mapping


<i>M^M** (8.1.7)</i>
<i>in which m in M becomes the homomorphism M*-^>R given by/i—>f(m).</i>
<i>The homomorphism M^>M** is natural in the sense that if k: M—>N is # </i>
-linear, then


<i>M >M**</i>


(8.1.8)


<i>N > N**</i>



is a commutative diagram.


We must now consider the duals of finite direct sums and of tensor
<i>products. The former are easily described for if Ml9</i> M2<i>, . . . , Mp are </i>


R-modules, then we have an isomorphism


( M1e M20 - - - 0 Mp) * ^ M1* © M f 0 - - - 0 M * ; (8.1.9)


<i>here / in (Mi</i> â M2 â ã ã * â Mp)* corresponds to (/l9 /2, . . . , /p), where /•


is the restriction of / to Mj(


For tensor products the situation is more complicated. Let us again
<i>assume that Mt</i>, M2<i>, . . . , Mp are /^-modules and let ft belong to Mf for i =</i>


<i>1,2,... ,p. If now</i>


</div>
<span class='text_page_counter'>(188)</span><div class='page_container' data-page=188>

<i>Modules of linear forms</i> 177


denotes p-fold multiplication (so that /ijf


<i>r</i>


<i>iri" -</i> rp)> then /4JP)<i> °{fi® f2®" ' ® fp</i>


M2<i> ®' • • ® Mp</i> and the mapping


Mx x M2 x • • • x Mp- > (Mx ® M2



<i>which takes (fl9</i> /2, . . . , /p) into ^jf}<i> ° (fi</i>


Accordingly there is an K-linear mapping
cằ: Mf đ Mf đ ã ã ã ® MJ - • (Mx


which is such that


<i><o(fi ®fi®" ' ® /</i>P) = 4<i>P) o(fi®f2</i>


and therefore, when mf<i> e M, for i = 1 , 2 , . . . , p,</i>


i ®<i> ri ® " ' ® rP) is equal to</i>


is a linear form on Mi đ


ã • • ® Mp)*


<i>f2®' " ® fp) is multilinear.</i>


M2 ® • • • ® Mp)* (8.1.11)


<i>"®fP) (8-1-12)</i>


<i>^fi(m1)f2(m2)...fp(mp). (8.1.13)</i>


<i>It will be noticed that there is an ambiguity surrounding the use of fx</i> ®


<i>f2®" ' ® fp- On the one hand it can denote an element of Mf đ MÊ đ ã • •</i>



® M* and on the other it can stand for a mapping of Mx đ M2<i> đ ã ã • ® Mp</i>


<i>into R ® R ® • • • ® R; indeed in (8.1.12) both interpretations occur in the</i>
same formula. The reader will therefore need to examine the context of each
occurrence to discover the appropriate meaning.


It will be seen shortly that when we restrict our attention to free modules
with finite bases the homomorphisms (8.1.7) and (8.1.11) become
isomorphisms. To simplify the language used to describe these matters, we
<i>shall employ the term finite free module to describe a free module of finite</i>
rank.


<i>Let M be such an K-module and let bx</i>, fc2<i>,..., bs</i> be one of its bases. Then


<i>for each integer i satisfying 1 <f < s there is a unique ft</i> in M* such that


<i>It is easily seen that fx</i>, /2<i>, . . . , fs</i> is a base of M* and therefore M* is a finite


free module of the same rank as M itself. The particular base (/1? /2<i>, . . . , fs)</i>


<i>is said to be dual to the base (bl9</i> fe2<i>,..., bs) of M.</i>


<i>Still supposing that M is a finite free module, let 6: M-+M** be the</i>
homomorphism (8.1.7). Then, using the same notation as in the last
<i>paragraph, we find that O(bi)(fj) = fj(bi) has the value 1 if i=j and is zero</i>


<i>otherwise. Thus d{bx), 6(b2),..., 6{bs) is the base of M** that is dual to the</i>


<i>base fl9f29...,fa</i> of M*.



<i><b>Lemma 1. Let Mbea finite free R-module. Then the natural homomorphism</b></i>


</div>
<span class='text_page_counter'>(189)</span><div class='page_container' data-page=189>

<i>178 Graded duality</i>


<i>This is clear because we have just seen that M^>M** takes a base of M</i>
into a base of M**. Of course, when M is a finite free module, Lemma 1 may
be used to identify M** with M.


Now suppose that Ml 5 M2<i>, . . . , Mp</i> are all of them finite free modules, let


<i>Bt be a base of Mi9 and let Bf be the base of Mf that is dual to Bt. In what</i>


<i>follows bt denotes an element of B{ and fj an element of Bf.</i>


By Theorem 3 of Chapter 1, the products bx đ fo2 đ ã ã • ® fopform a base


<i>of Ml</i> ® M 2 ® • • • ® Mp whereas the products / / ® /2 ® * *' ® /p


constitute a base of Mf đ MJ đ ã ã • ® M*. If now


<i>co: Mf ® Mf ® • • • ® M* -> (M1</i> ® M2 ® • • • ® Mp)*


is the homomorphism (8.1.11), then, by (8.1.13),
<i>*' * đ fp) (bi đ b2 đ ã ã ã ® bp)</i>


But this is zero unless


<i>in which case it has the value 1. It follows that {co(f{ ® f2 ® • • • ® fp)} is the</i>


<i>base of (Mx</i> ® M2 đ ã ã ã đ Mp<i>)* that is dual to the base {bx</i> đ fo2 đ ã ã ã



đ fop<i>} of Mx</i> đ M2 đ * ã ã đ Mp<i>. Accordingly co maps a base of Mf ®</i>


<i>Mf ® • • • ® M* into a base of (Mx</i> ® M2 ® • • • ® Mp)* and we have


proved


<i><b>Lemma 2. Let M</b></i>l 5 M2, . . . , Mp<i> foe y?m£e /ree R-modules. Then the natural</i>


<i>homomorphism</i>


Mf đ Mf đ ã ã ã đ M* -> (Mi đ M2<i> đ ã ã ã đ Mp)*</i>


<i>{see (8.1.11)) is an isomorphism.</i>


This lemma shows how, when M1, M2<i>, . . . ,MP</i> are finite free


mod-ules, it is possible to identify M f ® M f ® ã ã ã đ M * with
( M1đ M2đ - - - ® Mp) * .


<b>8.2 The graded dual of a graded module</b>


Finite free modules are, of course, very special. More general than
<i>these are graded modules with finite free components (the components of a</i>
graded module are the submodules which make up its grading), and the aim
of this section is to extend our results to this wider class. However, we shall
begin by considering arbitrary graded modules.


<i>Let M be an ^-module and {Mn}neZ</i> a grading on M. Then M* =



HomR (Mn<i>, R) can be regarded as a submodule of M* = Hom</i>R<i> (M, R); more</i>


<i>precisely, if we fix n, then the linear forms on M that vanish on Y*k*n Mk</i>


</div>
<span class='text_page_counter'>(190)</span><div class='page_container' data-page=190>

<i>The graded dual of a graded module 179</i>


<i>will be embedded in M* by means of this isomorphism. The reader will</i>
recall that we have already met this kind of situation in Section (7.7).


This said, we can form


<i>ZM: = Mf</i> (8.2.1)


<i><b>n</b></i>


say, where the summation is carried out in M*. It is easy to check that the
sum in (8.2.1) is direct. Accordingly M+ is graded by {M*}neZ.


<i><b>Definition. The graded module M</b></i>+<i> is called the 'graded dual' of the graded</i>
<i>module M.</i>


<i>Now suppose that M and N are graded /^-modules with {Mn}neZ</i> and


<i>{Nn}neZ</i> as their respective gradings; further, let


<i>X:M-^N (8.2.2)</i>


<i>be a degree-preserving ^-linear mapping, that is to say X is a </i>
<i>homo-morphism of graded modules. By (8.1.2), A* maps N* into M* and it is easy</i>
<i>to see that X*(Nk¥)^Mk¥ for all k. Accordingly there is induced a</i>



homomorphism


^ A T —M+ (8.2.3)


<i>of graded modules; moreover if Xk: Mk^>Nk</i> is the homomorphism in


<i>degree k induced by X, then the homomorphism AT£ —• Mjf induced by Xf</i> is
<i>X£. Evidently</i>


(idM)+ = idMt (8.2.4)


<i>and if \i\ N-+K is another homomorphism of graded modules, then</i>


<i>(fi°Xy = kf°fi\ (8.2.5)</i>


Thus the formation of graded duals is a contravariant functor from the
category of graded fl-modules (and their degree-preserving
<i>homo-mo rphisms) to itself. Naturally if X: M—>N is an isohomo-morphism of graded</i>
<i>modules, then so too is Xf: AT</i>+^M+ and


<i>(^)-l=(X-l)\ (8.2.6)</i>


<i>We next consider the homomorphism M—>M** of (8.1.7). This</i>
<i>associates with each element of M a linear form on M* and hence, by</i>
restriction, a linear form on M+. Thus we arrive at a homomorphism
<i>M—>(Mf)* and now an easy verification shows that the image of M</i>
belongs to (M+)+. Accordingly we have a homomorphism


Af-»M+ t (8.2.7)



</div>
<span class='text_page_counter'>(191)</span><div class='page_container' data-page=191>

<i>180 Graded duality</i>


homomorphism of graded modules, then


M >M++


t (8.2.8)


<i>N > Nn</i>


is a commutative diagram.


<i><b>Lemma 3. Let M be a graded R-module with finite free components. Then</b></i>


<i>the homomorphism M —• M</i>++<i> of (8.2.7) is an isomorphism of graded modules.</i>


This follows from Lemma 1 as soon as we consider components.


Now suppose that M(1), M( 2 ), . . . , M(p) are graded K-modules. We can


turn M( 1 )® M( 2 )® - - - ® M( p ) and M(1)+ đ M( 2 ) + đã ã ã ®M( p ) + into


graded modules by giving them the usual total gradings. Let /1e M( 1 ) t,


<i>f2 e M</i>(2)+ and so on. Then /4f}<i> ° (/i ® fi ® * * * ® fP\ where /^</i>p) denotes


p-fold multiplication, is a linear form on M( 1 ) ® M(2) ® • • • ® M(p) and


among the submodules Af jf* đ Mj2)<i> đ ã ã ã đ M\p\ of M</i>( 1 ) đ ã ã ã đ M(p),



there are only finitely many on which jijf} ° (/x ® /2<i> ® • • • ® fp) does not</i>


vanish. Consequently the linear form belongs to (M(1) ® M(2) ® • • •


® M(p))+. It follows that there is an K-linear mapping


ã ã ã đ M(p))+ (8.2.9)


in which /x<i> đ f2 đ ã ã * đ fp</i> (considered as an element of the left-hand side)


<i>is mapped into /#> (fx đ f2</i> đ ã ã ã đ /p). But M( 1 ) t đ M(2)+ đ ã ã • ® M(p)+


<i>is the direct sum of its submodules M^* ® M[</i>2)* ® • • • ® M[P)*, and


(Mj^ ® M|2) đ ã ã ã đ M|p))* is a submodule of the component of degree


*'i+1''2 + - ' -2+ ;p of ( A /( 1 )® M( 2 )® - - - ® M( P ))+. Furthermore (8.2.9)


induces the homomorphism


<i>ã ã ã đ M\P))* (8.2.10)</i>


which operates exactly as in (8.1.11). In particular we see that (8.2.9) is a
homomorphism of graded modules.


<i><b>Lemma 4. Let M</b></i>( 1 ), M( 2 ), . . . ,M( P )<i> be graded modules with finite free</i>
<i>components, and let all the gradings be non-negative. Then the homomorphism</i>


Md ) t (g M<b>(2)t g ) . . . g) M</b>( p ) +- > (M(1) đ M(2)<i> đ ã ã ã đ Mip)Y</i>



<i>(see (8.2.9)) is an isomorphism of graded modules.</i>


<i>Proof. Since the components are finite free modules all the </i>


homo-morphisms (8.2.10) are isohomo-morphisms; and, because the gradings are


</div>
<span class='text_page_counter'>(192)</span><div class='page_container' data-page=192>

<i>The graded dual of a graded module 181</i>


<i>direct sum of its submodules</i>


(M£> ® M<2)<i> ® • • • ® M\P))* (i! + i</i>2 + • • • + ip = fc)


(see (8.1.9)). The lemma follows.


We next consider some important situations involving homomorphisms
like (8.2.9). First suppose that we have homomorphisms


^:M('> — JV<'> (i = l , 2 p)


of graded modules. It is then easily verified that


đ JV(2)+<i> đ ã ãã đ N(p)f > (N(1)</i> đ iV(2)<i> đ ã ãã đ Nip)Y</i>


<b>Md ) t (g M(2)t (g . . . 0 M(P)t > (M(l) g) M(2) 0 .</b>


(8.2.11)


is a commutative diagram. Of course, it is understood that the horizontal
mappings are derived from (8.2.9).



By Lemma 4, if all the modules have non-negative gradings and finite free
components, then the horizontal mappings in (8.2.11) are isomorphisms.
Hence in these circumstances we may make the identifications


AT(1)+<i> ® Ni2)f đ ã ã ã đ Nip)f = (N{1)</i> đ JV(2)<i> đ ã ã ã đ Nip)Y</i>
and


Md ) t ^ M<i>(2)t 0 . . . 0 Mip)f = (M</i>(1) ® M(2)<i> ® • • • ® M(p))\</i>


The commutative property of the diagram then ensures that


4+i đ A2 đ ã • • ® A^ = (Ai ® A2 ® • • • ® /lp)f. (8.2.12)


Another situation which will concern us arises in the following way. Let
M(1), M(2)<i>, . . . , M{p)</i> and K(1), X(2)<i>, . . . , K{q)</i> be graded /^-modules. The
homomorphisms


Md ) t ^ M(2)t 0 . . . 0 M( p ) +-> (M(1) đ M(2) đ ã ã ã đ M(p))+


and


<b>Ê(Dt ^ X(2)t 0 . . . 0 X ^t- > (X^> 0 X(2) ® ' ' ' ® K(9))t</b>


induce a homomorphism


(Md)t 0 . . . 0 M( p ) + ) đ (X(1)t đ ã ã ã đ X(^)+)


- ã ( M( 1 )<i> đ ã ã • ® Mip)y ® ( x</i>( i )<i> ® • • • ® Kiq)y</i>



of graded modules. But, again from (8.2.9), we have a homomorphism


</div>
<span class='text_page_counter'>(193)</span><div class='page_container' data-page=193>

<i>182 Graded duality</i>


<i><b>® • • • ® M</b><b>ip)f</b><b>)</b></i>


(8.2.13)
This too is a homomorphism of graded modules.


Now consider the diagram


<i>® Kiq))f</i>


(8.2.14)
(Here the upper horizontal mapping comes from (8.2.9) and the lower one
from (8.2.13). The left vertical isomorphism comes from Theorem 1 of
Chapter 2 whereas the one on the right-hand side is the graded dual of the
isomorphism


<i>ã ã ã đ Mip)) đ (X</i>(1)<i> đ ã ã • ® Kiq))</i>


<i>• • • ® Mip)</i> ® X(1)<i> ® • • • ® Kiq)</i>


provided by the same result.) It is a straightforward matter to check that
<i>(8.2.14) is commutative. We shall describe this fact by saying that the taking</i>
<i>of graded duals is compatible with the associative property of tensor products</i>
as embodied in Theorem 1 of Chapter 2. Note that all the mappings
preserve degrees and that when the gradings are non-negative and all
components are finite free modules, the horizontal homomorphisms are
actually isomorphisms.



There is one more result connected with (8.2.9) that it is useful to place on
record. Let M(1), M( 2 )<i>, . . . , Mip)</i> be graded modules and let A(M(1), M(2),
. . . , M(p)) denote the isomorphism


(M(1)<i> ® • • • ® Mip)) ® (M</i>(1)<i> ® • • • ® Mip))</i>


<i>(Mip) ® Mip)) (8.2.15)</i>


of graded modules in which


<i>(mti đ mh đ ã ã ã đ mip) đ (m'h đ m'h đ ã ã ã đ m'jp)</i>


<i>is mapped into (mti đ m'h) đ (m</i>l2<i> đ m'h) đ ã • • ® (m</i>;<i> ® m)). (The notation</i>


<i>is the normal one, i.e. mix and m'h</i> belong to Mj^ and MJ*} respectively and


so on.) The isomorphism


<i>(Mip) ® Mip))</i>


</div>
<span class='text_page_counter'>(194)</span><div class='page_container' data-page=194>

<i>The graded dual of a graded module 183</i>


<i>that is inverse to (8.2.15) will be denoted by V(M{1\ M</i>( 2 ), . . . ,
It is now a simple matter to verify that the diagram


+


<i>) đ ã ã ã đ (Mip)f</i>



<b>ã • • ®</b><i><b> M</b><b>ip)</b><b>)</b></i>


(8.2.17)
is commutative. (Of course, the horizontal mappings are obtained by using
homomorphisms of the kind typified by (8.2.9).) Once again, all the
mappings preserve degrees.


Should it happen that M(1), M( 2 )<i>, . . . , M{p)</i> have non-negative gradings
and finite free components, tfren the horizontal mappings in (8.2.17) will be
isomorphisms so that in each case we can identify the domain with the
codomain. If this is done we find that


K(M(1)t, M( 2 ) +<i>, . . . , Mip)f) = A(M</i>(1), M( 2 )<i>, . . . , Mip))f</i> (8.2.18)
and now, by taking inverses, we can add the relation


A(M(1)+, M( 2 ) +<i>, . . . , Mip)f)= F(M</i>(1), M( 2 )<i>, . . . , Mip))\ (8.2.19)</i>
The isomorphisms (8.2.15) and (8.2.16) may be modified by introducing
signs just as we did in (7.1.3) and (7.5.6). Let A(M( 1 ),... ,M(P)) and
F ( M( 1 ), . . . , M(p)) be the results of making these changes. We then find that
when M(1), M(2)<i>, . . . , Mip)</i> have non-negative gradings and finite free
components we can replace (8.2.18) and (8.2.19) by


<i>V(M{1)\ M</i>( 2 ) +<i>, . . . , Mip)f) = A(M</i>(1), M( 2 )<i>, . . . , Mip))f</i> (8.2.20)
and


A(M(1)+, M( 2 ) +<i>, . . . , M(p)i)= F(M</i>(1), M( 2 ), . . . , M(p))+ (8.2.21)
respectively.


Our final comments in this section involve much simpler situations.
<i>Evidently if the /^-module K is graded trivially, then Kf is just K* with the</i>


trivial grading. In particular


<i>Rf = R**R, (8.2.22)</i>


<i>where the isomorphism R*zzR comes from (8.1.6). We shall use (8.2.22) to</i>
<i>identify Rf and R as graded modules.</i>


Now consider the homomorphisms


</div>
<span class='text_page_counter'>(195)</span><div class='page_container' data-page=195>

<i>184 Graded duality</i>


trivially graded module. Then


/ijf)+: £+<i>- > (R đ R đ ã ã ã đ R)f = Rf đ Rf</i> đ ã ã
and


Ajf> +<i>: Rf</i> đ K+<i> đ ã ã ã đ Rf = {R đ R đ ã ã ã đ R)</i>f<i> - ã R\</i>


<i>so that the identification Rf — R turns /4f</i>)+ into a mapping R—>R®
K ® - - - ® R and it turns Ajf)+ into a similar mapping in the reverse
direction. On this understanding we find that


A«jr = Aj? (8.2.23)


and


^ . (8.2.24)


<b>8.3 Graded duals of algebras and coalgebras</b>


In Section (7.7) we considered an algebra that was formed by linear


forms on a coalgebra, and now that we are about to discuss the graded duals
of algebras and coalgebras, it is convenient to point out the connections
between the present discussion and the ideas that were developed earlier. It
is for this reason that we begin with coalgebras.


<i>Suppose then that (A, A, e) is a non-negatively graded coalgebra whose</i>
components are finite free modules. Then A+<i>: (A ® A)f—>Af</i> and
ef: JR+—>y4f<i>. But, by Lemma 4, (A ® A)f can be identified with Af ® A^ and,</i>
<i>by (8.2.22), Rf can be identified with R. In this way we arrive at </i>
degree-preserving K-linear mappings


A1: ^ ® ^1" - ^1" (8.3.1)
and


e+:R —X+. (8.3.2)


<i><b>Theorem 1. Let (A, A, e) be a graded R-coalgebra with a non-negative</b></i>


<i>grading and finite free components. Then (A\ A</i>+, a+<i>) is a graded R-algebra.</i>
<i>Proof. Because (A\ A\ ef) will turn out to be a special case of a graded</i>
algebra that was encountered in the last chapter, we shall use an argument
that exhibits this relationship rather than one which is based directly on the
definition.


<i>Let f,geA\ By (8.2.9), when A^®Af is identified with (A®AY the</i>
<i>element / ® g of the former becomes fiR°(f ® g), where fiR</i> denotes the


<i>multiplication mapping of R. Thus for A</i>f<i>: Af ® Af—>Af</i> we have
A+( / ® 0 ) = / i *o( / ® 0 ) ° A . (8.3.3)
On the other hand



e+(l) = id^oe = e. (8.3.4)


</div>
<span class='text_page_counter'>(196)</span><div class='page_container' data-page=196>

<i>Graded duals of algebras and coalgebras 185</i>


natural structure as a graded /^-algebra and now (8.3.3) and (8.3.4) show
that Af and e+ are the multiplication and unit mappings of the previously
encountered algebra. This proves the theorem. It should be observed,
however, that Theorem 8 of Chapter 7 was proved under much more
general conditions.


<i><b>Definition. With the above notation the graded algebra (A\ A</b></i>+, ef<i>) is called</i>
<i>the 'graded duaV of the graded coalgebra (A, A, e).</i>


Note that the graded dual also has a non-negative grading and finite free
components.


<i><b>Theorem 2. Let A and B be graded coalgebras with non-negative gradings</b></i>


<i>and finite free components. Further, let X: A-+B be a homomorphism of</i>
<i>graded coalgebras. Then 1</i>+<i>: Bf—> Af is a homomorphism of graded algebras.</i>


This follows at once from Exercise 6 of Chapter 7. However, a direct
proof can be obtained by adapting the arguments, given below, to establish
the dual result (Theorem 4).


<i>We turn now to algebras. Suppose that (A, \i, rj) is a non-negatively</i>
graded algebra with finite free components. Then /x+ maps A+<i> into (A ® A)f</i>
and *7+<i> maps Af into R\ Consequently if we make the identifications</i>
<i>(A ®A)f = Af®Af and Rf = R, then we obtain K-linear mappings</i>



<i>fS:Af->A+®Af</i> (8.3.5)
and


<i>nf:A'-^R (8.3.6)</i>


which preserve the degrees of homogeneous elements.


<i><b>Theorem 3. Let (A, jj,,n) be a graded R-algebra, where the grading is </b></i>


<i>non-negative and has finite free components. Then (A\ / / , nf) is a graded </i>
<i>R-coalgebra.</i>


<i>Proof. Since \i ° {\i ® A) = pt ° {A ® fi) we have</i>


<i>(n°(n®A)y = (ti°(A®ii))f. (8.3.7)</i>


<i>Now, to be quite precise, \i ° {\i ® A) is the result of combining the</i>
homomorphisms


<i>A® A® A - ^ - > {A® A)® A > Ađ A ã A</i>


<i>so that (fi (/i đ A))f</i> is the total mapping


<i>A* > (A đ AY ã ((A đA)đ AY</i>


</div>
<span class='text_page_counter'>(197)</span><div class='page_container' data-page=197>

<i>186 Graded duality</i>


<i>Let us make the identifications (A ®AY = Af ® Af</i> and
<i>((A ® A) ® A)</i>+<i>= (A ® AY ®Af = (Af®Af)® A\</i>



Then the first mapping in (8.3.8) becomes the mapping (8.3.5) and, by
<i>(8.2.12), the second one is turned into / / ® A\ Accordingly (8.3.8) becomes</i>


<i>((A</i>


<i>and if next we identify {A ® A ® A)f with Af ®Af ® A\ then the final</i>
<i>mapping is the module-isomorphism (Af ® Af) ® Af-^Af ® Af ® Af</i> (see
<i>(8.2.14)). Thus (fi°(fi®A))\ considered as a mapping of Af into Af ®</i>
<i>Af ® A\ is simply (/j</i>f<i> ® Af) ° /z</i>+; likewise


<i>{li-{A ® /x))</i>f<i>: A*^>A* ®A^®Af</i>
<i>is none other than (Af</i> ® /x+) ° //. Accordingly


<i>so that fif: Af—>A^ ® Af</i> is coassociative.
Next, the total mapping


<i><b>A</b><b> -^-+</b><b> ^ ^</b></i>


<i>is just the identity mapping of A and therefore</i>


<i>A' — (A ® AY ~ (R ® AY - ^ A' (8.3.9)</i>


<i>is the identity mapping of A\ If now we put {A ® AY = A^ ® Af</i> and
<i>(R®AY = R*® Af = R®A*</i>


we find that (8.3.9) becomes


<i>{R®AY -^A\</i>



<i>Consider the final mapping in this sequence. If / e A\ then 1 ® /</i>
<i>(regarded as a member of (R ® AY) maps r®a into rf(a), hence the image</i>
<i>of 1 ® / in Af</i> is just / itself. Accordingly the total mapping


</div>
<span class='text_page_counter'>(198)</span><div class='page_container' data-page=198>

<i>Graded duals of algebras and coalgebras 187</i>


<i><b>Definition. The graded coalgebra (A\fi\rj</b>f) of Theorem 3 is called the</i>
<i>'graded duaV of the graded algebra (A, fi, n).</i>


Naturally the grading on the graded dual is non-negative and the
components are finite free modules.


<i><b>Theorem 4. Let A and B be graded R-algebras with non-negative gradings</b></i>


<i>and finite free components, and let X: A-+B be a homomorphism of graded</i>
<i>algebras. Then A</i>+: J5+<i>—+Af is a homomorphism of graded coalgebras.</i>
<i>Proof If we apply graded duality to the commutative diagram</i>


<i><b>PA</b></i>


<i>A®A >A</i>


<i><b>B®B >B</b></i>


the result is a new commutative diagram; and if we then make the
<i>identifications (A ® A)f = Af ® Af and (B ® B)</i>+ = B+<i> ® B\ the new </i>
diag-ram becomes


<i>Af</i>



B+<i> ® Bf < £</i>+


<i>where the horizontal mappings are comultiplications. Thus Xf</i> is compatible
with comultiplication. A similar exercise carried out on the commutative
diagram


<i>"B</i>


shows that 2+ also preserves counits. The proof is therefore complete.
<i>We know that R is both a graded K-algebra and a graded R-coalgebra;</i>
<i>indeed the results of this section are applicable to it. Consequently R* is also</i>
a graded algebra and a graded coalgebra.


<i><b>Lemma 5. The isomorphism R</b>f&Rof (8.2.22) is an isomorphism of graded</i>
<i>algebras and graded coalgebras.</i>


The verification is trivial if we use (8.2.23) and (8.2.24).


</div>
<span class='text_page_counter'>(199)</span><div class='page_container' data-page=199>

<i>188 Graded duality</i>


<i><b>Theorem 5. Let A be a graded algebra, respectively coalgebra, with a </b></i>


<i>non-negative grading and finite free components. Then the canonical isomorphism</i>
<i>A-^An of graded modules (see Lemma 3) is actually an isomorphism of</i>
<i>graded algebras, respectively coalgebras.</i>


<i>Proof We shall deal only with the case where A is a coalgebra; the other</i>
case can be dealt with similarly.


<i>Let A and e be the comultiplication and counit mappings of A. By (8.2.8),</i>


the diagram


<i>A®A >(A®A)n</i>


<i>is commutative. If we make the identifications (A ® A)n = (Af ® Af)f =</i>
<i>An ® An we find that A đ A ã (A đ A)n, considered as a homomorphism</i>
<i>of A ® A into An ® An, is the tensor product of the homomorphism</i>
<i>A—>An with itself. This shows that A-+Aff</i> is compatible with
comultiplication.


Next we consider the commutative diagram


<i>R >Rn</i>


and observe that when we put


<i>the mapping R —• Rn becomes the identity mapping of R. Accordingly</i>
<i>A —• An</i> preserves counits and with this the proof is complete.


<b>8.4 Graded duals of Hopf algebras</b>


We can extend the results of the last section to Hopf algebras, but
first of all it is necessary to see how tensor products of graded algebras and
coalgebras are effected by graded duality.


<i>We begin with algebras. Let A{1\ A(2\... ,Aip)</i> be algebras with
<i>non-negative gradings whose components are finite free modules. Then A{1)</i> đ
<i>A{2) đ ã ã ã ® A{p)</i> is a graded algebra of the same kind. We propose to
<i>compare (A(1) đ Ai2) đ ã ã ã ® Aip))f</i> and A(1)+ ® 4( 2 ) +<i> ® • • • ® Aip)f</i> as
<i>graded coalgebras.</i>



</div>
<span class='text_page_counter'>(200)</span><div class='page_container' data-page=200>

<i>Graded duals of Hopf algebras 189</i>


<i>® Aip)) ® U</i>( 1 )<i> đ ã ã ã (x) Aip))</i>


<i>- ^ (A(1) đ A{1)) ®-"® (A{p) ® Aip)) (8.4.1)</i>
of graded modules which operates as in (8.2.15) and let the inverse
isomorphism


<i>"-® (A{p) ® A{p))</i>


-^-> (A(1)<i> đ ã ã ã đ A(p)) đ (A(1) đ • • • ® Aip)) (8.4.2)</i>
<i>be denoted by V{A(1\ Ai2\ . . . , A(p)) (see (8.2.16)). If now fit and rjt</i> are the


<i>multiplication and unit mappings of A{i\ then the multiplication mapping</i>


<i>P \ . . . , Aip))</i>


and its unit mapping is


(f71®f/2®---®f/p)°Ajf).


At this point let us make the identifications


<i>ip)</i>


<i>Y = A</i>(1)+ đ A(2)+<i> đ ã ã ã ® Aip)\</i>
<i>(A{p) ® Aip))Y</i>


<i>® • • • ® (A(p)f ® Aip)f)</i>


and


(04( 1 )<i> đ ã ã ã đ Aip)) đ {A{1) đ"-đ Aip)))f</i>


<i>= {Ail)f đ-"đ Aip)f) đ {A(1)f đ ã • • ® Ai p ) f) .</i>


Then, using (8.2.18), we see that the comultiplication mapping of (,4(1)
<i>Ai2) ® "-® A(p))f is</i>


<i>= V(A{1)\ Ai2)\ . . . , Aip)f) o (fi\ ®fif2®'"® n\) (8.4.3)</i>


and that its counit mapping is


<i>P</i> (8.4.4)


(see (8.2.24)).


Now the right-hand sides of (8.4.3) and (8.4.4) are none other than the
<i>comultiplication and counit mappings of Ail)f ® • • • ® A{p)f. Consequently</i>
we have proved that the natural isomorphism


</div>

<!--links-->

×