Classically Semisimple Rings - Martin
Classically Semisimple Rings - Martin
Classically
Semisimple
Rings
A Perspective Through Modules
and Categories
Classically Semisimple Rings
Martin Mathieu
Classically Semisimple
Rings
A Perspective Through Modules
and Categories
Martin Mathieu
School of Mathematics and Physics
Queen’s University Belfast
Belfast, UK
© The Editor(s) (if applicable) and The Author(s), under exclusive license to Springer Nature Switzerland
AG 2022
This work is subject to copyright. All rights are solely and exclusively licensed by the Publisher, whether
the whole or part of the material is concerned, specifically the rights of translation, reprinting, reuse
of illustrations, recitation, broadcasting, reproduction on microfilms or in any other physical way, and
transmission or information storage and retrieval, electronic adaptation, computer software, or by similar
or dissimilar methodology now known or hereafter developed.
The use of general descriptive names, registered names, trademarks, service marks, etc. in this publication
does not imply, even in the absence of a specific statement, that such names are exempt from the relevant
protective laws and regulations and therefore free for general use.
The publisher, the authors, and the editors are safe to assume that the advice and information in this book
are believed to be true and accurate at the date of publication. Neither the publisher nor the authors or
the editors give a warranty, expressed or implied, with respect to the material contained herein or for any
errors or omissions that may have been made. The publisher remains neutral with regard to jurisdictional
claims in published maps and institutional affiliations.
This Springer imprint is published by the registered company Springer Nature Switzerland AG
The registered company address is: Gewerbestrasse 11, 6330 Cham, Switzerland
To my children,
who never cease
to amaze me.
Le savant n’étudie pas la nature
parce que cela est utile;
il l’étudie
parce qu’il y prend plaisir
et il y prend plaisir
parce qu’elle est belle.
This book is written for advanced undergraduate and beginning graduate students. It
tells the story of a classical classification problem, about rings that are semisimple in
the sense of Wedderburn and Artin. One of the first major achievements of modern
algebra, the Wedderburn–Artin theorem, paved the way for many like theories in the
first half of the twentieth century. We present this theory from a modern viewpoint
choosing an approach via modules. At the same time, the reader is allowed a
glimpse into category theory, which partly finds its origins in the theory of modules.
Category theory is highly abstract and therefore sometimes hard to digest in a first
helping. By interweaving the basic concepts with more concrete examples coming
from modules and rings, we aim to make the abstract ideas more accessible and, at
the same time, to present a classical beauty in an attractive modern setting.
ix
Contents
xi
xii Contents
5 Artinian Modules . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .. . . . . . . . . . . . . . . . . . . . 47
5.1 Finitely Cogenerated Modules . . . . . . . . . . . . . . . . . .. . . . . . . . . . . . . . . . . . . . 48
5.2 Commutative Artinian Rings . . . . . . . . . . . . . . . . . . . .. . . . . . . . . . . . . . . . . . . . 50
5.3 Artinian vs. Noetherian Modules . . . . . . . . . . . . . . .. . . . . . . . . . . . . . . . . . . . 52
5.4 Abelian Categories . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .. . . . . . . . . . . . . . . . . . . . 54
5.5 Exercises . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .. . . . . . . . . . . . . . . . . . . . 56
6 Simple and Semisimple Modules. . . . . . . . . . . . . . . . . . . . . .. . . . . . . . . . . . . . . . . . . . 59
6.1 Decomposition of Modules . . . . . . . . . . . . . . . . . . . . .. . . . . . . . . . . . . . . . . . . . 59
6.2 Projective and Injective Modules . . . . . . . . . . . . . . .. . . . . . . . . . . . . . . . . . . . 64
6.3 Projective and Injective Objects . . . . . . . . . . . . . . . .. . . . . . . . . . . . . . . . . . . . 68
6.4 Exercises . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .. . . . . . . . . . . . . . . . . . . . 70
7 The Artin–Wedderburn Theorem . . . . . . . . . . . . . . . . . . . .. . . . . . . . . . . . . . . . . . . . 73
7.1 The Structure of Semisimple Rings . . . . . . . . . . . . .. . . . . . . . . . . . . . . . . . . . 73
7.2 Maschke’s Theorem . . . . . . . . . . . . . . . . . . . . . . . . . . . . .. . . . . . . . . . . . . . . . . . . . 80
7.3 The Hopkins–Levitzki Theorem . . . . . . . . . . . . . . . .. . . . . . . . . . . . . . . . . . . . 81
7.4 Exercises . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .. . . . . . . . . . . . . . . . . . . . 89
8 Tensor Products of Modules . . . . . . . . . . . . . . . . . . . . . . . . . . .. . . . . . . . . . . . . . . . . . . . 93
8.1 Tensor Product of Modules . . . . . . . . . . . . . . . . . . . . .. . . . . . . . . . . . . . . . . . . . 93
8.2 Tensor Product of Algebras . . . . . . . . . . . . . . . . . . . . .. . . . . . . . . . . . . . . . . . . . 101
8.3 Adjoint Functors . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .. . . . . . . . . . . . . . . . . . . . 102
8.4 Exercises . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .. . . . . . . . . . . . . . . . . . . . 105
9 Exchange Modules and Exchange Rings . . . . . . . . . . . . .. . . . . . . . . . . . . . . . . . . . 109
9.1 Basic Properties of Exchange Modules . . . . . . . . .. . . . . . . . . . . . . . . . . . . . 110
9.2 Exchange Rings . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .. . . . . . . . . . . . . . . . . . . . 114
9.3 Commutative Exchange Rings . . . . . . . . . . . . . . . . . .. . . . . . . . . . . . . . . . . . . . 119
9.4 Exercises . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .. . . . . . . . . . . . . . . . . . . . 128
10 Semiprimitivity of Group Rings . . . . . . . . . . . . . . . . . . . . . .. . . . . . . . . . . . . . . . . . . . 131
10.1 Basic Properties . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .. . . . . . . . . . . . . . . . . . . . 131
10.2 Some Analytic Structure on C[G] . . . . . . . . . . . . . .. . . . . . . . . . . . . . . . . . . . 133
10.3 The Semiprimitivity Problem . . . . . . . . . . . . . . . . . . .. . . . . . . . . . . . . . . . . . . . 137
10.4 Exercises . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .. . . . . . . . . . . . . . . . . . . . 141
Bibliography . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .. . . . . . . . . . . . . . . . . . . . 145
Index of Symbols . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .. . . . . . . . . . . . . . . . . . . . 147
Index . . . . . . . . .. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .. . . . . . . . . . . . . . . . . . . . 149
Introduction
xiii
xiv Introduction
converse statement is in general false: the ring Z is the most obvious example.
The proof of this fundamental result is deferred to Sect. 7.3, but Exercise 5.5.8
already invites a proof in the commutative case. In Sect. 5.4 abelian categories are
introduced which are the most prominent generalisation of module categories, and
various typical properties and techniques are displayed.
The heart of the theory is Chap. 6, where simple and semisimple modules are
defined and studied. Finitely generated semisimple modules turn out to be both
Noetherian and Artinian (Theorem 6.2.4). The concept of a classically semisimple
ring is introduced, and these rings are characterised by the fact that their category of
(left) modules entirely consists of semisimple modules (Theorem 6.1.12). This, in
consequence, leads to other important classes of modules, projective and injective
modules (Sect. 6.2) and their generalisations to projective and injective objects in
arbitrary categories (Sect. 6.3). The description via the Hom-functors opens up the
connection to Homological Algebra which we do not pursue further here; see [21]
or [29] for a full development.
The Artin–Wedderburn theorem is the focus of Chap. 7. This was one of the
major achievements of the early theory of non-commutative rings and still is a cor-
nerstone of basic algebra. It forms the model for many structure theorems that were
obtained during the twentieth century. An important example of a semisimple ring
is the group ring of certain finite groups; this is Maschke’s theorem (Theorem 7.2.1)
discussed in Sect. 7.2. A full proof of the Hopkins–Levitzki theorem, Theorem 5.3.2,
is finally given in Sect. 7.3. This section also contains a discussion of the Jacobson
radical for general unital rings and the notion of a semiprimitive ring. Our proof
differs from the one commonly given, which relies on composition series, and uses
module techniques instead. We do not have anything to say about categories in this
chapter aside from Remark 7.1.13.
A construction of paramount importance is the tensor product of modules to
which Chap. 8 is devoted. The fundamental Hom-tensor adjointness relation is
established in Theorem 8.1.9 and further studied in Sect. 8.3 on adjoint functors
between general categories. The special situation of tensor products of algebras over
a field is the theme of Sect. 8.2.
An interesting generalisation of semisimple rings was introduced in 1972 by
Warfield, based on previous work on modules by Crawley and Jónsson. In order
to provide the reader with some more recent development in Ring Theory, we have
included Chap. 9 which gives an introduction to the theory of exchange modules
and exchange rings. More comprehensive treatments can be found in [13] and [16].
In Sect. 9.1, we compile a number of basic properties of exchange modules, and in
Sect. 9.2, exchange rings are introduced via a separation property using idempotents
(Definition 9.2.1). It is shown that a unital ring R is an exchange ring if and only
if the standard module R R is an exchange module (Theorem 9.2.6) which in turn is
equivalent to the fact that every projective module in the category R−mod has the
exchange property (Corollary 9.2.7). Commutative exchange rings are investigated
in great detail in Sect. 9.3 and the reader will discover a number of equivalent
properties characterising exchange rings (Theorem 9.3.10). We also make contact
with various notions of dimension of a commutative ring and with topology; e.g., it
xvi Introduction
In this chapter, the reader will become familiar with our notation and basic
assumptions on the one hand; on the other, they are invited to various instances
where the setting of Ring Theory in its strictest sense turns out to be too narrow,
or at least somewhat cumbersome, in order to provide us with enough tools for
studying certain properties of rings. We shall take this as part of our motivation to
introduce the more comprehensive setting of modules below.
While the first section will be more informal in that we do not provide every detail
of each argument, the second section will devoted to a few first steps in Category
Theory, where we will introduce the fundamental concepts and illustrate these with
a number of examples.
Definition 1.1 A ring R consists of a non-empty set together with two binary
operations “+” and “ · ”, which are usually referred to as the addition and the
multiplication in R, such that
hold.
The ring R is called unital if (R, ·) has an identity element, i.e., an element 1 such
that 1 · x = x = x · 1 for all x ∈ R. The ring R is called commutative if (R, ·) is
commutative.
Note that we will not automatically assume that every ring is unital but this
assumption will be made explicit whenever necessary; the theory of unital rings
runs more smoothly in general. We list a number of examples that are useful to
understand the general theory.
Examples 1.2
(i) Z, the ring of integers, is the most basic but still an interesting ring. Clearly it is
both unital and commutative. Any new concept introduced should be checked
for Z first.
(ii) Zn = Z × . . . × Z, for a natural number n, with coordinate-wise operations.
This gives another unital commutative ring which already has some different
features. E.g., it is no longer an integral domain.
(iii) Every field K— such as Q, R and C—provides us with an extremely well-
behaved commutative unital ring. To some extent, Ring Theory is about how
far one can relax the nice multiplicative properties in a field and still get useful
and interesting examples.
(iv) H, the quaternions, almost forms a field but for the non-commutativity of
multiplication. Thus, they yield the first interesting example of a non-trivial
division ring. (See, for example, [34, Sect. 5.2].)
(v) R[x], the ring of polynomials in the indeterminate x over a commutative unital
ring R. When iterated, this construction yields the ring R[x1 , . . . , xn ], n ∈ N of
polynomials in n indeterminates which is highly useful in Algebraic Geometry.
(vi) Mn (R), where R is a unital ring and n ∈ N, is the ring of n × n matrices with
entries in R. Even if R is commutative, this ring is only commutative in case
n = 1. In general, the matrix rings are a source of useful examples for many
ring theoretic properties. E.g., Mn (R) is simple (that is, has no non-trivial two-
sided ideals) if and only if R is simple. (See Proposition 7.1.1 in Chap. 7.)
(vii) Tn (R), for a unital ring R and n > 1, is the ring of strictly upper diagonal
matrices, that is, an n × n matrix (aij ) belongs to Tn (R) if aij = 0 for all i ≥ j
and aij ∈ R is arbitrary otherwise. This is a non-unital noncommutative ring.
In the above examples, the operations of addition and multiplication are the
canonical ones; hence there is no need to discuss these in further detail as they
are treated in every undergraduate textbook on ring theory such as [7] or [34], for
instance. The next two important examples may be less familiar; for that reason, we
spend a little more time on them. The first is fundamental for the development of
module theory.
with tk = k=gh rg sh . (Here, we write the group multiplicatively as it does
not have to be abelian.) If R is unital then G is canonically a subgroup of
R[G] via g → 1 · g, and if e ∈ G is the identity then 1 · e is the identity
in R[G]. If both R and G are commutative, so is R[G]. In the case when K is
a field, the group ring K[G] in addition is a K-vector space with basis G (and
the scalar multiplication simply defined via multiplication of the coefficients
in K), so it is a K-algebra.
In this way, methods from Ring Theory can be used to study groups. For
instance, one of the deep open and longstanding problems is whether a group
G is torsion free if and only if the group ring K[G] has no non-trivial divisors
of zero. See also Exercise 1.3.6. (Quite a lot on this problem can be found
in [31].) We shall discuss certain aspects of group rings in Chap. 10.
The aim of Ring Theory is to understand the structure and the properties of these
and many other classes of rings. An important tool to achieve this aim is the concept
of a module which allows us to look at a given ring from various perspectives.
(r + s) · m = r · m + s · m, r · (m + n) = r · m + r · n (r, s ∈ R, m, n ∈ M).
In the case when R is unital and 1 · m = m for each m ∈ M, the left module R M is
called unital (in the literature, unitary is used as well).
4 1 Motivation from Ring Theory
Definition 1.4 Let R and S be rings, and let M be an abelian group. We say that
M is an R-S-bimodule, written as R MS , if M is a left R-module as well as a right
S-module and the compatibility condition r · (m · s) = (r · m) · s holds for all r ∈ R,
s ∈ S and m ∈ M.
If one does not want to specify which kind of (one-sided) module one speaks of,
one says “a module M over the ring R”. We next come to the important concept of
a substructure.
Notation In order to save writing, we shall denote the collection of all left R-
modules over a fixed ring R by R-Mod. Analogous meaning is given to the symbols
Mod-R and R-Mod-S (where S is another ring). If R (and S) is unital, this
notation will automatically refer to the collection of all unital modules.
Since the module maps are of paramount importance in the theory, we also give the
sets comprising all of them special symbols.
1.1 Basics on Modules 5
Rather than delving further into the general theory of modules, let us now pause
to ask the question “Why study modules at all?” The following instances of the
appearance of modules shall provide us with motivation to develop a comprehensive
theory as well as with some first examples.
Let us indicate the argument for the sake of illustration. If R is such a ring
then
there is an injective ring homomorphism ϕ from R into the direct product α Rα
of integral domains Rα (with a possibly large index set of α’s) such that Rα =
πα ◦ ϕ(R) for each α, where πα is the canonical projection onto Rα . The key to
this result is that, for every x ∈ R \ {0}, one can find a proper prime ideal Px of R
with the property x ∈ / Px . The integral domain Rα is then defined by R/Px (so that
α = x and the index set is all of R !). See also Exercise 1.3.11.
The above method of taking appropriate homomorphic images and then embed
the ring into a product of those is already very much in the spirit of module theory:
we will see that quotient rings have a canonical module structure right away.
6 1 Motivation from Ring Theory
r · (a + L) = ra + L (r, a ∈ R).
In this way, R/L turns into a left R-module which is unital if R is unital. Note also
that L ∈ R-Mod too!
We shall see that these two types of modules are very popular.
Every unital ring can be embedded into an endomorphism ring; thus, the latter may
be considered as “the mother of all rings”. Using the language of modules, this
process, reviewed below, becomes very natural.
Let R be a unital ring. We now formalise the construction of the standard R-
module R R indicated in the previous section. The abelian group (R, +) becomes a
unital left R-module in a natural way via
r ∈ R, s ∈ R : r · s = rs.
λ : R → End((R, +)), r → λr ,
1.1 Basics on Modules 7
r ∈ R, g ∈ G : r · g = π(r)(g).
πM : R → End(M), πM (r)(m) = r · m (r ∈ R, m ∈ M)
The above discussion shows that modules appear in many guises; therefore it
should be fruitful to have a well-developed theory at hand. To complete this section,
we record some (further) basic examples of modules.
Examples 1.8
It is now time to start introducing the language of categories. Category Theory itself
is nowadays a vast and highly developed area in Pure Mathematics with manifold
applications in other fields. We will only be able to touch upon the most basic
concepts and features; nevertheless this shall allow us a new view on the interplay
between rings and modules.
S. Eilenberg and S. MacLane established the foundations in 1945 in their article
“General theory of natural equivalences”, see [15]. Following their conviction that
“It should be observed first that the whole concept of a category is essentially an auxiliary
one; our basic concepts are essentially those of a functor and of a natural transformation.”
1.2 Categories of Modules 9
we will very soon include functors (as well as natural transformations) in our
discussion; see Sect. 3.2.
In the following, we will have to deal with “very large collections of sets” which
will be called classes. The reader familiar with Russell’s Paradox will know that
we cannot perform arbitrary operations with sets without creating set-theoretic
problems. The solution to this situation is a careful distinction between actual
sets and more flexible classes; this however needs an axiomatic approach to Set
Theory which would go far beyond the scope of this book. Theories which extend
the Zermelo–Fraenkel set theory together with the Axiom of Choice and allow a
rigorous treatment are the von Neumann–Bernays–Gödel axioms or the Morse–
Kelley set theory. An alternative is to work within Grothendieck universes. The
interested reader will find answers to their questions in [30], for example. The main
difference between sets and classes is that any class that is an element of another
class is a set.
(b) (identities)
The category C is called concrete if the elements of ob(C) have underlying sets and
the morphisms of mor(C) are mappings between these underlying sets. (A more
10 1 Motivation from Ring Theory
precise definition will be given in 3.2.2.) The category C is called small if ob(C) is
a set.
Let us have a look at some of the categories the reader may have encountered.
S the category of all sets where the objects are sets and the morphisms are
mappings between sets.
Top the category of all topological spaces; the objects are topological spaces
and the morphisms are continuous mappings between them.
Comp the category of compact Hausdorff spaces; the morphisms are again the
continuous mappings.
Gr the category of groups with group homomorphisms as morphisms.
AGr the category of abelian groups with group homomorphisms as mor-
phisms.
Ring the category of rings with ring homomorphisms as morphisms.
Ring1 the category of unital rings with unital ring homomorphisms as mor-
phisms.
R-Mod the category of (unital) left R-modules over a given (unital) ring R where
the morphisms are the left R-module maps.
Ban∞ the category of complex Banach spaces with bounded linear operators
as the morphisms.
Ban1 the category of complex Banach spaces with linear contractions as the
morphisms.
Lat the category of lattices and lattice homomorphisms as the morphisms.
HTop the category of all topological spaces and homotopy classes of continu-
ous mappings as the morphisms.
All of the above categories but for the last one are concrete categories; the objects
have an underlying set structure and the morphisms arise from mappings between
these sets. Moreover, the composition law is provided by the composition of
mappings. In HTop, this is not the case.
The reader will have no difficulties in defining the categories Mod-R and
R-Mod-S for themselves.
Although many of the above categories look alike on the surface, they have
rather different properties, as we shall see when moving on in the later chap-
1.2 Categories of Modules 11
ters. For example, for two objects R M, R N ∈ R-Mod the set of morphisms
Mor(R M, R N ) = HomR (R M, R N ) has a group structure (see Exercise 1.3.2)
whereas this is not the case for X, Y ∈ ob(Top): there is no canonical addition
for continuous mappings between X and Y .
For instance, Comp is a full subcategory of Top and AGr is a full subcategory
of Gr. On the other hand, Ring1 is a non-full subcategory of Ring and Ban1 is a
non-full subcategory of Ban∞ . Note that Ring is not even a subcategory of AGr
as an abelian group can carry several non-isomorphic rings structures.
Let R-mod denote the category whose objects are the finitely generated left
modules over a fixed ring R (i.e., those modules that have a finite set of generators,
see Definition 2.3.1) together with the left R-module maps. Then R-mod is a full
subcategory of R-Mod.
Remark 1.12 Even if the objects in a category have an underlying set (and are
specified by some additional structure on these sets), the morphisms do not have
to be mappings on the underlying sets. An example is HTop. Another one arises
from the open subsets of a given topological space X: the category OX has as objects
all open subsets of X and, for U, V ∈ ob(OX ), Mor(U, V ) = {→} if U ⊆ V and
Mor(U, V ) = ∅ otherwise. The composition law is defined through the transitivity
of set inclusion.
1.3 Exercises
ker(f ) = {m ∈ R M | f (m) = 0}
Exercise 1.3.2 For two module maps f and g defined on a module M with values
in a module N over the ring R define their sum by (f + g)(m) = f (m) + g(m)
for all m ∈ M. Show that HomR (R M, R N ) and HomR-S (R MS , R NS ) (where both
M and N are R-S-bimodules and S is another ring) are abelian groups under this
addition.
Exercise 1.3.4 Let R be a ring without identity. Show that we can embed R
into a unital ring in the following way: equip the abelian group R × Z with the
multiplication
Verify that we indeed get a ring multiplication in this way, and that (0, 1) serves
as an identity. The ring thus obtained is denoted by R × and is called the minimal
unitisation of R or the Dorroh superring of R.
Show further that the monomorphism r → (r, 0) allows us to view R as an ideal
in R × . (However, if performed for a ring which already has an identity, this process
will destroy the original identity!)
Exercise 1.3.5 Suppose R is a unital ring and that R M is a left R-module. Then
R M0 = {m ∈ M | r · m = 0 ∀ r ∈ R} and R M1 = {m ∈ M | 1 · m = m}
Exercise 1.3.6 Let G be a group with non-trivial torsion; that is, there is g ∈ G
such that g n = e for some n ∈ N and g = e, where e is the neutral element in G.
Show that the group ring R[G] contains a non-trivial divisor of zero.
Exercise 1.3.8 Let H be a subgroup of a group G. Let a ∈ R[H ]. Use the previous
exercise to show that, if a is invertible in R[G], then it is invertible in R[H ], and
that, if a is a zero divisor in R[G], then it is a zero divisor in R[H ].
Exercise
1.3.9 Let
G be a group and R be a unital ring. Define the trace on R[G]
by tr g∈G rg g = r1 , the coefficient at the identity 1 ∈ G. Show that tr is a left
R-module map and that tr(ab) = tr(ba) for all a, b ∈ R[G].
Exercise 1.3.10 Let R be a commutative unital ring, and denote by nil(R) the set
of all nilpotent elements in R. Show that nil(R) is an ideal of R; it is called the
nil radical of R. Show further that the quotient ring R/nil(R) does not contain any
non-zero nilpotent element.
I2 ⊆ P . Using Zorn’s Lemma, show that the intersection of all prime ideals of R is
zero.
Exercise 1.3.12 Let (π, G) be a representation of the unital ring R (that is, π : R →
End(G) is a unital ring homomorphism from R into the endomorphism ring of the
abelian group G). Show that the action of R on G defined by r ·g = π(r)(g), r ∈ R,
g ∈ G turns G into a left R-module.
Exercise 1.3.14 Let R be a unital ring. Show that every monomorphism in R-Mod
is injective and every epimorphism in R-Mod is surjective. Does a similar result
hold in the full subcategory R-mod of R-Mod?
This chapter is devoted to various basic constructions that we can perform within
R-Mod. Once we have found some interesting modules, we want to be able to
obtain new ones from them. Special classes of modules will play a distinguished
role here. In the second part of this chapter we shall review these constructions from
the viewpoint of category theory and thus realise that seemingly similar categories
can behave rather differently.
a monomorphism if f is injective;
an epimorphism if f is surjective;
an isomorphism if f is bijective.
r ∈ R, m ∈ M : r · (m + N) = r · m + N.
r · n + N = r · m − r · (m − n) + N = r · n + N.
In this way we obtain the quotient module R M/R N ; if R is unital, this construction
gives the correct object in R-Mod. We also have the canonical epimorphism
πN : R M → R M/R N , πN (m) = m + N.
Clearly, ker(πN ) = R N.
Proof It is easily checked that the inverse image of a submodule under a module
map is a submodule of the domain. Therefore, for a submodule R L ≤ R M/R N ,
−1
R K = πN (R L) = {m ∈ M | πN (m) ∈ L} is a submodule of R M containing R N .
Conversely, let R K ≤ R M be such that R N ⊆ R K. Then πN (R K) is a submodule
of R M/R N (compare Exercise
1.3.1).
Evidently, R K ⊆ πN−1 πN (R K) . If m ∈ πN−1 πN (R K) then πN (m) ∈ πN (R K)
and so m differs from an element of R K by anelement in R N ⊆ R K. Consequently,
m ∈ R K and we obtain R K = πN−1 πN (R K) . On the other hand, we always have
πN πN−1 (R L) ⊆ R L for any R L ≤ R M/R N, and since πN is surjective, we get the
equality of the two sets.
Definition 2.2.2 Let R be a ring, and let R M ∈ R-Mod. For a non-empty subset
S ⊆ R M we let
be the annihilator of S in R.
2.4 Direct Sums and Products of Modules 17
Clearly, AnnR (S) is always a left ideal of R and AnnR (R M) is a two-sided ideal
of R. It goes without saying that all of the above can be defined analogously for
right modules and bimodules.
In this section, we shall take a look at various ways to generate submodules from
subsets of a given module.
Let R M ∈ R-Mod for some ring R. Given a family {R M i | i ∈ I } of
submodules R M i ≤ R M, their intersection is easily seen to be a submodule
of R M. Suppose now that S ⊆ R M is a non-empty subset. The intersection of
all submodules of R M that contain S—which is evidently the smallest submodule
of R M containing S—is called the submodule generated by S and will be denoted
by R S.
Suppose that R is unital. For m ∈ M, R {m} = {r · m | r ∈ R} = Rm and for
any ∅ = S ⊆ R M, we have
R S = Rs = rj · sj | rj ∈ R, sj ∈ S, J finite .
s∈S j ∈J
for all (mi ), (ni ) ∈ X and r ∈ R. It is easy to check that we indeed obtain
RMi
i∈I
a left R-module which is denoted by R M i and called the direct product of the
family {R M i | i ∈ I }. i∈I
18 2 Constructions with Modules
The submodule
R M i = (mi ) ∈ RMi | at most finitely many mi are non-zero
i∈I i∈I
i R Mi
f
pj
RN
fj
R Mj
i RMi
f
ιj
RN
fj
R Mj
2.5 Free Modules 19
Proof
(i) For n ∈ N, we set f (n)i = fi (n) for each i ∈ I . Then pi f = fi and it is easy
to verify that f is a left R-module map.
(ii) For m ∈ i R M i , there are only finitely many i’s with mi = 0. Hence we can
define f (m) = i fi (mi ) and obtain a left R-module map satisfying f ιj = fj
for all j ∈ I .
This section deals with those modules that resemble vector spaces most closely. In
a way, they turn out to be the most general modules.
Let X be a non-empty set, and let R be a unital ring. We consider the set
In this way, we obtain a left R-module called the free left R-module on X which is
denoted by RX (sometimes also denoted as R (X) ). Note that, if we put Mx = R R
for each x ∈ X, then RX is nothing but Mx .
x∈X
Clearly, analogous constructions give us free right modules and free bimodules,
respectively.
Free modules are distinguished in several ways.
Considering the elements of X inside the free module RX, we see that X is a
basis of it.
20 2 Constructions with Modules
Proposition 2.5.2 Let R be a unital ring. Let R M ∈ R-Mod with basis S. Then
R M is isomorphic to RS as left R-modules.
Remark Contrary to the situation of vector spaces over fields, the notion of the
‘length of a basis’ (and hence, the notion of ‘dimension’) is not well defined for free
modules. For example, let V = Q[x] and R = EndQ (Q V ). It can be shown that
∼ n ∼
R R = R R ⊕ R R and hence R R = R R for all n, m ∈ N. If R is commutative,
m
f
R X RM
Corollary 2.5.4 Every unital left module is a quotient of a free left module.
We shall see in the next chapter that the technique used in the proof of the above
corollary allows us to identify any homomorphic image of a module as a quotient
module.
2.6 Special Objects in a Category 21
f
F A
Note that we do not use a special symbol for the underlying set of an object in C.
For example, for every unital ring R, if X is a non-empty set then RX is a free
object in R-Mod, by Theorem 2.5.3. On the other hand, the only free object in
Z-mod (the finitely generated Z-modules) is the zero module.
Products and direct sums (called “coproducts”) are defined in a general category C
just as in R-Mod.
Let {Ai | i ∈ I } be a family of objects in a category C. A product of the family
{Ai | i ∈ I } consists of an object A ∈ ob(C) together with a family {pi | i ∈ I }
of morphisms in Mor(A, Ai ), i ∈ I , called projections, such that the following
diagram can be made commutative by a unique f ∈ Mor(C, A) for each C ∈ ob(C)
22 2 Constructions with Modules
and fi ∈ Mor(C, Ai ), i ∈ I :
A
f
C pi
fi
Ai
If such product exists, it is unique up to isomorphism and is denoted by i∈I Ai .
Dually, a coproduct of {Ai | i ∈ I } consists of an object A ∈ ob(C) together
with a family {ei | i ∈ I } of morphisms in Mor(Ai , A), i ∈ I , called injections, such
that the following diagram can be made commutative by a unique g ∈ Mor(A, B)
for each B ∈ ob(C), gi ∈ Mor(Ai , B), i ∈ I :
A
g
ei B
gi
Ai
Examples
Two special situations for product and coproduct need to be singled out. Suppose
I = ∅. The “empty” product, if it exists, is the unique object A ∈ ob(C) with the
property that for each C ∈ ob(C) there is a unique morphism C → A; that is,
MorC (C, A) has precisely one element. In this case, A is called a final object. The
“empty” coproduct is the unique object A ∈ ob(C), if it exists, satisfying: for each
C ∈ ob(C) there is a unique morphism A → C; that is, MorC (A, C) has precisely
one element. In this case, A is called an initial object.
For instance in S, ∅ is the initial object and {∅} is the final object. In R-Mod,
the zero module 0 serves as both the initial and the final object. In the category of
fields, however, there are no initial or final objects (because a field homomorphism
preserves the characteristic of a field).
An object in a category that is both initial and final is called a zero object As such
an object is unique (cf. Exercise 2.7.12) it is often denoted as 0.
Some categories such as R-Mod or Ban∞ have an additional structure on their
morphism sets: they are abelian groups. In generalisation of this, one calls a category
C additive if it has zero, finite coproducts and all morphism sets Mor(A, B),
A, B ∈ ob(C) carry the structure of an abelian group such that the composition
of morphisms is bilinear. In this case, finite products exist as well and agree with the
finite coproducts.
2.7 Exercises
Exercise 2.7.1 Let R be a unital ring with centre Z(R). Show that the mapping
a → La , where, for a ∈ R, La : R → R is given by La (x) = ax, x ∈ R provides
isomorphisms between the rings R and EndR (RR ), respectively, and between Z(R)
and EndR (R RR ), respectively.
Exercise 2.7.2 Let R be a ring, and let R M ∈ R-Mod. Show that the annihilator
of R M in R,
Exercise 2.7.5 Show that the direct product and the direct sum of a family of left
R-modules are uniquely determined up to isomorphism by the universal properties
stated in Proposition 2.4.1.
Exercise 2.7.6 Let R be a unital ring, and let R M ∈ R-Mod. Show that if R M is
finitely generated and free then it is isomorphic to R R n for some n ∈ N.
Exercise 2.7.7 Let R be a unital ring, and let R M, R N ∈ R-Mod. Suppose that
R N is free and let f : R M → R N be an R-epimorphism. Show that there exists
g ∈ HomR (R N , R M) such that f ◦ g = idN and that R M = ker(f ) ⊕ im(g) (that
is, R M = ker(f ) + im(g) and ker(f ) ∩ im(g) = 0).
R L ∩ (R K + R N) = R K + (R L ∩ R N)
Exercise 2.7.9 Let R M ∈ R-Mod for a unital ring R. Show the following two
statements.
M, we have Rm ∼
(i) For each m ∈ R = R/AnnR (m).
(ii) AnnR (R M) = f ∈HomR (R,M) ker f .
Just as in group theory and in ring theory, we need to be able to manipulate modules
freely under isomorphisms; this is done via the so-called isomorphism theorems
which are discussed in this short chapter. The second part of this chapter deals with
connections between categories, the functors, and with connections between the
functors, the natural transformations.
The basis for all the canonical isomorphism theorems is the following result.
f
RM RN
R M/ ker(f ) im(f )
f
We can now draw several nice consequences from the above theorem.
3.1.2 First Isomorphism Theorem Let R M, R N ∈ R-Mod for some ring R. For
every f ∈ HomR (R M, R N), we have R M/ ker(f ) ∼
= im(f ).
Proof The mapping fˆ in the canonical factorisation supplies the desired isomor-
phism.
(R N + R K)/R K ∼
= R N /(R N ∩ R K).
In this section we will discuss how ‘to move’ from one category to another; this
is done via ‘functions between categories’, the functors. We shall then see how to
move between functors; this is done via natural transformations.
such that
(a) ∀ A ∈ ob(C) : F(1A ) = 1F(A) ;
(b) ∀ A, B, C ∈ ob(C) ∀ f ∈ MorC (A, B), g ∈ MorC (B, C):
The functor F is called faithful if it is injective on morphism sets and it is called full
if it is surjective on morphism sets.
Examples 3.2.2
HomR (R M, −) : C −→ AGr
RN −→ HomR (R M, R N ) (3.2.1)
HomR (R N , R L) f −→ f∗ = HomR (R M, f )
F
F F
G G
G
covariant case
F
F F
G G
G
contravariant case
ηM : F(R M) −→ G(R M) (R M ∈ C)
3.2 Functors and Natural Transformations 31
3.3 Exercises
f g
0 RM RN RL 0 (3.3.2)
and is exact we speak of a short exact sequence. Note that we do not have to specify
the module maps at either end (they can only be the zero maps) and that exactness
f g
in this situation means that R M R N is injective, R N R L is
surjective and ker(g) = im(f ). By the first isomorphism theorem, it follows that
∼
R L = R N/im(f ) so one can think of R L as a quotient of R N . The degenerate case
when R L = 0 means that f itself is an isomorphism.
0 R M/( R N R L) R M/ R N R M/ R L R M/( R N + R L) 0.
(R N + R L)/(R N ∩ R L) ∼
= (R N + R L)/R N × (R N + R L)/R L.
3.3 Exercises 33
Exercise 3.3.5 In Exercise 1.3.3 we saw how to turn an S-module into an R-module
given a fixed ring homomorphism ρ : R → S between two rings R and S. Does this
procedure give us a covariant functor Fρ : S-Mod → R-Mod?
Exercise 3.3.6 Let R be a unital ring and let R L, R N be submodules of the right R-
module R M. We define [R L : R N ] = {a ∈ R | a · R N ⊆ R L}. Show the following
statements:
(i) [R L : R N ] = R if R N ⊆ R L.
(ii) [R L1 ∩ R L2 : R N ] = [R L1 : R N ] ∩ [R L2 : R N ] for any R L1 , R L2 ≤ R M.
(iii) [R L : R L + R N ] = [R L : R N ].
Exercise 3.3.7 Let R be a unital commutative ring and let R M ∈ R-Mod. Using
the same notation as in the previous exercise, show that the following two conditions
are equivalent:
Show further that, if I is an ideal of R, then the quotient module R R/R I has any of
the above two propeties if and only if I is a prime ideal.
Noetherian Modules
4
Theorem 4.1 Let R M ∈ R-Mod for a unital ring R. The following conditions are
equivalent.
Proof The implication (a) ⇒ (b) is trivial. Towards (b) ⇒ (c) let N be a non-
empty set of submodules of R M without a maximal element. Take R N 1 ∈ N; then
there is R N 2 ∈ N with R N 1 ⊂ R N 2 . Let R N 1 , . . . , R N k ∈ N be chosen such that
R N 1 ⊂ R N 2 ⊂ . . . ⊂ R N k . Since R N k is not maximal in N, there is R N k+1 ∈ N
strictly containing R N k . By induction, we obtain an infinite strictly ascending chain
of submodules of R M. Thus (b) fails.
(c) ⇒ (d) Let R N ≤ R M. Let N be the set of all finitely generated submodules
of R N . As these are submodules of R M, N has a maximal element R N 0 . If there is
n ∈ R N \ R N 0 then R N 0 +Rn is a finitely generated submodule of R N strictly larger
than R N 0 . As this contradicts the maximality of R N 0 in N, we find that R N = R N 0
is finitely generated.
Example 4.3 Let s and t be two symbols and let R = Zs, t | t 2 = ts = 0 be the
quotient of the free unital ring Zs, t by the stated relations. As every element of R
is a unique combination of words of the form s n and s n t, n ≥ 0, R can be written
as R = Z[s] ⊕ Z[s]t. The commutative ring Z[s] is Noetherian by Hilbert’s Basis
Theorem (see below). The ideal Z[s]t of R is considered as a Z[s]-module; therefore
the finitely generated Z[s]-module Z[s] ⊕ Z[s]t is Noetherian, by Corollary 4.1.4
below, in other words, R is left Noetherian.
In order to show that R is not right Noetherian, it suffices to verify that the
submodule Z[s]t ∈ Mod-R is not finitely generated. The monomials s n t, n ≥ 0
are Z-linearly independent; thus there cannot exist finitely many p1 , . . . , pk ∈ Z[s]t
such that an arbitrary element p = n≥0 mn s n t, where mn ∈ Z are non-zero for
k
at most finitely many n, can be written as p = j =1 mj pj as the latter would
4.1 Permanence Properties of Noetherian Modules 37
imply that Z[s]t is finitely generated as an abelian group. As a result, R is not right
Noetherian.
4.4 Hilbert’s Basis Theorem For every left Noetherian ring R, the polynomial
ring R[x1 , . . . , xn ] is left Noetherian.
Proof Since R[x, y] = R[x][y] it suffices to establish the result for polynomial
rings in one indeterminate and then proceed by induction. Suppose that R[x] is not
left Noetherian. Take a left ideal I in R[x] which is not finitely generated. Let f1 ∈ I
be a polynomial with minimal degree. Then I1 = R[x]f1 I . Take f2 ∈ I \ I1 of
minimal degree. Put I2 = R[x]f1 + R[x]f2 I . Since I is not finitely generated,
by induction we obtain an infinite strictly ascending chain (In )n∈N of left ideals
in R[x]. Let nk = deg fk , k ∈ N and let ank be the leading coefficient in fk (i.e.,
ank is the coefficient of x nk ). Observe that n1 ≤ n2 ≤ . . . ≤ nk ≤ nk+1 ≤ . . .. Let
Lk = Ran1 + . . . + Rank , k ∈ N. Clearly, (Lk )k∈N is an ascending chain of left
ideals of R. We claim that this sequence is strictly ascending (hence R is not left
Noetherian).
Suppose on the contrary that ank+1 ∈ Lk = Ran1 +. . .+Rank for some k. Hence,
ank+1 = ki=1 ri ani for some ri ∈ R. Put
k
f (x) = fk+1 (x) − ri x nk+1 −ni fi (x).
i=1
n
n
ank+1 x nk+1 − ri x nk+1 −ni ani x ni = ank+1 − ri ani x nk+1 = 0.
i=1 i=1
This violates the choice of fk+1 being a polynomial in Ik+1 \ Ik of lowest degree.
As a result our initial assumption entails that R is not left Noetherian.
In this section, we shall study how Noetherian modules behave under various
canonical constructions, such as taking submodules or homomorphic images. The
key result is the following theorem in which we employ the concept of an exact
sequence of modules, see p. 32.
38 4 Noetherian Modules
f g
0 RM1 RM2 RM3 0
where m ≤ k and f (k) and g (k) are the restrictions of f and g, respectively to the
corresponding submodules. By construction, all rows are short exact sequences. As
(m) ⊆ N (k) it remains to prove the reverse inclusion.
RN R
Let x ∈ R N (k) . Since R N (k) (m)
3 = R N 3 there is y ∈ R N
(m) such that g (k) (x) =
(k)
g (y). Since ker(g ) = im(f ) there is y ∈ R N 1 such that x − y = f (k) (y ).
(k) (k) (k)
(k) (m) (m)
As R N 1 = R N 1 , y ∈ R N 1 . It follows that
0 RN RM R M/ R N 0,
Proof Using induction, it suffices to consider the case of two Noetherian modules
R L and R N . In this case, the statement follows directly from Theorem 4.1.1 applied
to the short exact sequence
0 RN RN RL RL 0.
Corollary 4.1.4 Every finitely generated unital left module over a left Noetherian
ring is Noetherian.
n
RR
n
−→ R M, (r1 , . . . , rn ) −→ ri · mi
i=1
We observe therefore that a unital ring R is left Noetherian if and only if every
module in R-mod is Noetherian.
Let C be a category with zero object, see Chap. 2, p. 23. We will denote this object
(which is unique up to isomorphism by Exercise 2.7.12) by 0. Let B, C ∈ ob(C).
The unique morphisms f : 0 → B and g : C → 0 have the properties
for all A ∈ ob(C) and h ∈ MorC (B, A), k ∈ MorC (A, C).
The above properties are easily checked (see Exercise 4.3.8 below) and the above-
mentioned unique morphism is called a zero morphism. As a result, a category C
with zero object has zero morphisms between all its objects.
In the sequel, we will assume that C has a zero object 0 and denote the zero
morphisms unambiguously by 0, too.
Let f ∈ MorC (A, B) for some A, B ∈ ob(C).
0
h
g
i f
K A B
0 (4.2.1)
A B C
f p
g
h
0
D (4.2.2)
Since kernel and cokernel of a morphism f , if they exist, are unique, it suffices
to give their domain and codomain, respectively, a symbol; these are, respectively,
ker(f ) and cok(f ).
For instance, in R-Mod, the kernel of f ∈ HomR (R M, R N ) is the usual kernel
ker(f ) together with the canonical embedding ker(f ) → R M, and the cokernel of f
is the canonical projection π : R N → R N /im(f ). However, in R-mod, kernels
need not exist since a submodule of a finitely generated module may be not finitely
generated.
We are now in a position to introduce the concept of an exact sequence.
f g
A B C
h
ker(g) (4.2.3)
42 4 Noetherian Modules
f g
0 A B C 0
f g
0 A B C 0
in C the sequence
F(f ) F(g)
0 F(A) F(B) F(C)
F(f ) F(g)
F(A) F(B) F(C) 0
f g
0 RM RN RL 0
4.3 Exercises 43
in R-Mod and we need to show that (a) f∗ is a monomorphism, and (b) im(f∗ ) =
ker(g∗ ); then the sequence
f g
0 HomR ( R U , R M) HomR ( R U , R N ) HomR ( R U , R L)
will be exact.
If f h = 0 for some h ∈ HomR (R U , R M) then h = 0 as f is injective.
Therefore, f∗ is injective, thus a monomorphism. This proves (a).
In order to prove (b), take h ∈ im(f∗ ) ⊆ HomR (R U , R N) so that h = f k for
some k ∈ HomR (R U , R M). Then g∗ (h) = gh = gf k = 0 as im(f ) ⊆ ker(g). It
follows that h ∈ ker(g∗ ) and thus im(f∗ ) ⊆ ker(g∗ ). Now take h ∈ ker(g∗ ) so that
gh = 0. Then h(U ) ⊆ ker(g) = im(f ). As a result, for u ∈ U , there is m ∈ M
with h(u) = f (m) and, since f is injective, we can define k ∈ HomR (R U , R M) by
k(u) = m. It follows that h = f k ∈ im(f∗ ), which proves (b).
4.3 Exercises
Exercise 4.3.2 Prove that every left Noetherian unital ring R is von Neumann finite,
that is, for x, y ∈ R, xy = 1 implies that yx = 1.
Exercise 4.3.3 Let R be a commutative unital ring. Using Zorn’s lemma, show that
R is Noetherian if every prime ideal of R is finitely generated.
Exercise 4.3.6 Let R M ∈ R-Mod for a unital ring R. Show that the following two
conditions on R M are equivalent.
ιJ f
R Mj −→ R Mi −→ R M
j ∈J i∈I
Exercise 4.3.7 Let R M ∈ R-Mod for a unital ring R. Show that R M is finitely
generated if andonly if, for every family {R Ni | i ∈ I } of submodules R Ni of
M such that i∈I R Ni = R M, there exists a finite subset J ⊆ I such that
R
j ∈J R Nj = R M.
Exercise 4.3.8 Verify the properties of the zero morphism as stated in Sect. 4.2.1.
Exercise 4.3.9 Prove that the kernel and the cokernel of a morphism in a category
with zero are unique up to isomorphism, provided they exist.
Exercise 4.3.10 Let C be an exact category. Show that for every f ∈ mor(C) the
following identities hold:
Exercise 4.3.11 Let f, g ∈ mor(C) be “parallel arrows”, that is, both f and g
belong to MorC (A, B) for two objects A, B in C. An equalizer of (f, g) is a
morphism e ∈ MorC (C, A) such that f e = ge and for every h ∈ MorC (D, A)
with f h = gh there is a unique h ∈ MorC (D, C) such that eh = h. Show that in
R-Mod every equalizer is a kernel.
Exercise 4.3.12 Identify the kernel and the cokernel of a morphism f ∈ Ban1 .
2 f
RN RL
g
(iii) Prove that R M ×L R N has the universal property of a pullback; that is, given
another commutative diagram of module maps of the form
1
RK RM
2 f
RN RL
g
f
RL RM
RN RM T RN
(iii) Prove that R M ⊕T R N has the universal property of a pushout; that is, given
another commutative diagram of module maps of the form
f
RL RM
RN RK
The dual condition to “ACC” is “DCC”, the descending chain condition which we
shall discuss in the present chapter. The resulting modules are termed Artinian after
Emil Artin (1898–1962).
Historical Note Born in 1898 in Vienna (Austria), Emil Artin became one of the
leading mathematicians of the twentieth century. He made major contributions
to algebraic number theory, notably class field theory, abstract algebra and braid
theory. In 1927 he solved Hilbert’s 17th problem.
Artin’s university studies at Vienna were interrupted when he was drafted in June
1918 but he continued them in Leipzig from 1919 onward where he was awarded his
PhD in 1921. After a 1-year postdoctoral position at Göttingen during which period
he worked closely with Emmy Noether and Helmut Hasse, Artin moved to Hamburg
where he advanced to the rank of Privatdozent, having completed his Habilitation in
1923. In 1925 he accepted a position as associate professor at Hamburg University
and was promoted to full professor in 1926.
Because of his wife’s Jewish father and his own distaste for the Hitler regime,
Artin lost his position in Hamburg and was forced to emigrate to the US where he
got a professorship at Notre Dame University, Indiana in 1937. In 1938 Artin and
his family moved again, to Indiana University in Bloomington. Apart from various
visiting positions at Stanford, Ann Arbor, Boulder as well as in Japan, Artin spent
the years between 1946 and 1957 at Princeton, which had become the mecca of
mathematics during those years. Finally, in 1958, he returned to Hamburg where he
would remain until he died of a heart attack in December 1962.
A detailed account of Emil Artin’s life and work, written by J. J. O’Connor and
E. F. Robinson can be found at
https://mathshistory.st-andrews.ac.uk/Biographies/Artin/
Exercises 4.3.7 and 5.5.1 illustrate nicely the interplay between the concepts of
finitely generated and finitely cogenerated modules.
We now come to the promised characterisation of Artinian modules.
Theorem 5.1.3 Let R M ∈ R-Mod for a unital ring R. The following conditions
are equivalent.
(a) R M is Artinian.
(b) Each strictly descending chain R N 1 ⊃ R N 2 ⊃ . . . ⊃ R N k ⊃ . . . of submodules
R N i ≤ R M is finite.
5.1 Finitely Cogenerated Modules 49
i∈I R N i = 0 implies that N= 0 and therefore R N min = 0. By Exercise 5.5.1
below, R M is finitely cogenerated.
Every homomorphic image of R M, and hence every quotient of R M, inherits
property (c). Thus, these quotient modules are finitely cogenerated too, proving (d).
(d) ⇒ (a) Let R N 1 ⊇ R N 2 ⊇ . . . ⊇ R N k ⊇ . . . be a descending chain of
submodules of R M.Put R N = k∈N R N k . By hypothesis, R M/R N is finitely
cogenerated. Since k∈N R N k /R N = 0, condition (d) together
with Exercise 5.5.1
below imply that there is a finite subset J ⊆ N such that j ∈J R N j /R N = 0. Since
R N 1 /R N ⊇ R N 2 /R N ⊇ . . . ⊇ R N k /R N ⊇ . . . it follows that R N r /R N = 0 for
some r ∈ N, that is, R N r = R N . As a result, R N r+k = R N r for all k ∈ N wherefore
R M is Artinian.
The proof of the next result is analogous to the one of Theorem 4.1.1 and is therefore
left to the reader, see Exercise 5.5.2.
f g
0 RM1 RM2 RM3 0
Corollary 5.1.7 For every unital ring R, the following conditions are equivalent.
Z ⊃ 2Z ⊃ 22 Z ⊃ . . . ⊃ 2k Z ⊃ . . .
is infinite.
In this section we want to study some special properties of Artinian rings that are
commutative.
Proposition 5.2.2 Every finitely generated torsion module over a principal ideal
domain is Artinian.
5.2 Commutative Artinian Rings 51
Proof Let R be a principal ideal domain, and let R M ∈ R-mod with generators
m1 , . . . , mk . Suppose that R M is a torsion module. Then, by Exercise 2.7.9,
The reader may want to compare the above result with the fact that every finitely
generated module over a principal ideal domain is Noetherian.
Commutative Artinian rings are in a way the simplest rings after fields. To give
evidence to that fact, we record a few more of their nice properties (some of which
do extend appropriately to non-commutative Artinian rings).
are called the Jacobson radical and the nil radical of R, respectively.
By definition, both nil(R) and rad(R) are ideals of R and, since every maximal
ideal is a prime ideal—the quotient by it is a field, hence an integral domain—, we
have nil(R) ⊆ rad(R). In general, this inclusion will be strict.
Proof Since nil(R)n ⊇ nil(R)n+1 for all n ∈ N, DCC gives nil(R)n = nil(R)k for
some k ∈ N and all n ≥ k. Put I = nil(R)k . Suppose I = 0; then the set J of all
ideals J in R with I J = 0 is non-empty. Since R is Artinian, there is a minimal
element J0 in J. Take x ∈ J0 with xI = 0. As (x) ⊆ J0 we conclude that (x) = J0 .
Since (xI )I = xI 2 = xI = 0, an analogous argument yields xI = (x) = J0 . It
follows that x = xy for some y ∈ I from which we obtain
x = xy = xy 2 = . . . = xy n for all n ∈ N.
We already observed that the ring Z is Noetherian but not Artinian. Here is another
example of the same ilk.
Example 5.3.1 We shall use again the ring from Example 4.3. The ring
is left Noetherian (but not right Noetherian). Moreover, R is neither left nor right
Artinian since the module Z[s] ∼ = R/Z[s]t is not Artinian; it contains (Z[s]s k )k∈N
as an infinite strictly descending chain of submodules.
In the opposite direction, there is the following surprising result for rings (it does
not extend to modules, though).
We defer the proof until Chap. 7 but the reader can already try their hands at the
commutative case, see Exercise 5.5.8.
Of course, an analogous statement holds with “left” replaced by “right” in both
the assumption and the conclusion.
5.3 Artinian vs. Noetherian Modules 53
Proposition 5.3.3 Let R be a unital ring, and let R M ∈ R-Mod. For every f ∈
EndR (R M), the following holds.
ker(f n ) + im(f n ) = R M;
ker(f n ) ∩ im(f n ) = 0;
ker(f n ) ⊕ im(f n ) = R M.
The third statement in the above proposition is also known as “Fitting’s Lemma”. In
all statements, f n stands for the n-th iterate of the mapping f .
Proof
becomes stationary; that is, im(f n+k ) = im(f n ) for some n ∈ N and all k ∈ N.
Therefore, for given x ∈ R M, there is y ∈ R M such that f n (x) = f 2n (y);
consequently, x − f n (y) ∈ ker(f n ). It follows that x = x − f n (y) + f n (y) ∈
ker(f n ) + im(f n ), as claimed.
If ker(f ) = 0 then ker(f n ) = 0; thus im(f n ) = R M which implies
im(f ) = R M.
54 5 Artinian Modules
becomes stationary; that is, ker(f n+k ) = ker(f n ) for some n ∈ N and all
k ∈ N. In particular, ker(f 2n ) = ker(f n ). Thus, for x ∈ ker(f n ) ∩ im(f n ),
say x = f n (y), y ∈ R M, we have f 2n (y) = f n (x) = 0. As y ∈ ker(f 2n ) =
ker(f n ), we obtain x = 0 as claimed.
If im(f ) = R M then im(f n ) = R M; thus ker(f n ) = 0 which implies that
ker(f ) = 0.
(iii) is a direct consequence of (i) and (ii). The proof is complete.
Let C be a category which is both additive and exact; see p. 23 and 42, respectively.
Then C is called abelian. Abelian categories were introduced by A. Grothendieck
in 1957 as an abstract setting which captured many features of module categories.
In this section we aim to illustrate this fact by discussing various results familiar
from module theory that can be obtained without the use of elements in the objects.
Proposition 5.4.1 Let f ∈ MorC (A, B), where C is an abelian category. Then f
is an isomorphism if and only if it is both a monomorphism and an epimorphism.
Proof The “only if”-part follows directly from the definitions; see Definition 1.14.
Suppose f is a monomorphism. If fg = 0 for some g ∈ MorC (D, A) then g = 0.
Therefore g can be uniquely factored through 0 −→ A, that is, ker(f ) = 0. Note
that cok(ker(f )) = cok(0) = 1A ; thus we have a commutative diagram
1A
0 A A
f h
cok(f ) B ker(cok(f ))
with a unique morphism h ∈ MorC (A, ker(cok(f ))). Thus f = ker(cok(f )).
5.4 Abelian Categories 55
1A
0 A A
f h
0 B B
1B
The two properties f = ker(cok(f )) and f = cok(ker(f )), respectively now give
us the existence of unique morphisms k, k ∈ MorC (B, A), respectively making the
diagrams below commutative, compare (4.2.1) and (4.2.2):
5.4.3 Short Five Lemma Suppose we are given the following commutative dia-
gram in the abelian category C:
m e
0 A B C 0
f g h
0 A B C 0
m e
Suppose both rows are exact and the morphisms f and h are both monomorphisms
(epimorphisms). Then g is a monomorphism (an epimorphism).
The relation between abelian categories and module categories is in fact rather
intimate. The Freyd–Mitchell theorem states that for every small abelian category
there is an full exact embedding in R-Mod for a suitable ring R. See, e.g., Sect. 4.14
in [30].
There exist various alternative but equivalent ways to define an abelian category,
some even derive the abelian group structure on the Hom-sets from other basic
axioms; see [17, Chap. 2] and [30, Chap. 4]. A very readable account is given in
Chap. 7 of [29].
5.5 Exercises
Exercise 5.5.1 Let R M ∈ R-Mod for a unital ring R. Show that R M is finitely
if and only if, for every family {R Ni | i ∈ I } of submodules
cogenerated R Ni of R M
such that i∈I R Ni = 0, there exists a finite subset J ⊆ I such that j ∈J R Nj = 0.
Exercise 5.5.2 Write out the details of the proof of Theorem 5.1.4.
5.5 Exercises 57
Exercise 5.5.3 Let R be an integral domain, and let R M ∈ R-Mod. The torsion
submodule of R M is defined as
(i) tor(R M) ≤ R M;
(ii) tor(R M/tor(R M)) = 0;
(iii) for f ∈ HomR (R M, R N ), where R N ∈ R-Mod, we have f (tor(R M)) ⊆
tor(R N).
Exercise 5.5.4 In an Artinian commutative ring R, there are only finitely many
prime ideals.
Exercise 5.5.5 Let R be a commutative unital ring. Show that the definitions of the
nil radical of R given in Definition 5.2.3 and in Exercise 1.3.10 agree with each
other.
Exercise 5.5.6 Let K be a field and let V be a K-vector space. Show that the
following three conditions are equivalent.
The next exercise is the analogue of the First Isomorphism Theorem in module
theory (Theorem 3.1.2) in abelian categories.
58 5 Artinian Modules
i p
ker(f ) A cok(ker(f ))
p i
cok(f ) B ker(cok(f ))
for a given f ∈ MorC (A, B), use the definition of kernel and cokernel in
Sect. 4.2.1 in order to obtain unique morphisms g : A −→ ker(cok(f )) and
k : cok(ker(f )) −→ B, respectively with the properties f = i g and f = kp,
respectively. In the commutative diagram
i p
ker(f ) A cok(ker(f ))
g k
f h
p i
cok(f ) B ker(cok(f ))
Exercise 5.5.11 Let R be a left Noetherian ring. Show that R-mod is an abelian
category.
Exercise 5.5.13 Fill in the details in the following statement: The additive category
Ban∞ is not abelian since the inclusion 1 → c0 is both a monomorphism and an
epimorphism but not an isomorphism. (For the definition of the spaces, see p. 68.)
Simple and Semisimple Modules
6
In this chapter, which is at the very heart of our exposition, we shall apply module
theory to study the structure of a certain class of rings; these will be termed
(classically) semisimple. To this end, we will, once again, introduce some special
classes of modules and investigate their properties.
From this point on, we shall only consider unital rings (but still state this
explicitly) and unital modules. These were introduced in the Exercises in Chap. 1
where it was also shown that this is no restriction of the generality.
We turn our attention to the modules that are of the utmost interest to us in this book.
Examples 6.1.3
Our next result, though very easy to prove, is fundamental for the application of
simple modules in ring theory. It was first obtained by I. Schur in 1905 in a different
language (see Exercise 6.4.2 below).
6.1.4 Schur’s Lemma Let R be a unital ring, and let R M ∈ R-Mod be simple.
Then EndR (R M) is a division ring.
Proof Since ker(f ) ≤ R M for each f ∈ EndR (R M), either ker(f ) = 0 — in which
case f is injective — or ker(f ) = R M, that is, f = 0. Similarly, im(f ) ≤ R M
implies that im(f ) = R M — in which case f is surjective — or im(f ) = 0, that is
f = 0. It follows that every non-zero endomorphism of R M is bijective, and hence
EndR (R M) is a division ring.
EndR (R M) ∼
= EndR/I (R/I ) ∼
= R/I
We shall now study the relations between simple and semisimple modules. It
is evident from the definition that every simple module is semisimple. To obtain
information in the other direction, we first record the following useful observation.
RL ⊕ RL = R L ⊕ (R N ∩ R L ) = R N ∩ (R L ⊕ R L ) = R N
im(f ) = f (R M) = f (R N) ⊕ f (R N ) = R L ⊕ f (R N )
and thus im(f ) is semisimple. The first isomorphism theorem (Theorem 3.1.2)
completes the argument.
With the above preparations at hand, we can now ‘completely reduce’ any
semisimple module into its simple constituents.
(a) RM is semisimple;
(b) RM is the direct sum of simple submodules;
(c) R M is the sum of simple submodules.
Proof
As R M = i∈I R M i , we may, and will, consider R M as a submodule of
i∈I R M i . Let pj : i∈I R M i → R Mj be the canonical projection onto the j th
component. Since R N = 0, pj (R N ) = 0 for at least one j ∈ I . Therefore, f =
pj |R N ∈ HomR (R N , R Mj ) is a non-zero homomorphism. Since R Mj is simple,
im(f ) = R Mj ; since R N is simple, ker(f ) = 0. It follows that f is an isomorphism.
which is simple. We have j ∈I R Nj = 0, hence the intersection over all maximal
submodules of R M is zero.
Conversely, let {R N i | i ∈ I } be the collection of all maximal submodules
j ∈ I , put R M i = R M/R N i which is a simple left R-module.
of R M. For each
Suppose that
i∈I R N i = 0. By Exercise 5.5.1, there is a finite subset J ⊆ I
such that j ∈J R Nj = 0. Consequently, the R-module map R M −→ j ∈J R Mj
is an embedding of R M into a semisimple module, hence R M is semisimple
(Proposition 6.1.6).
Our next immediate goal is to determine for which rings every (left) module
is semisimple. Clearly, the question for which ring every module is simple does
not make sense, since we can always form direct sums. The key to the answer to
the above question is the following concept for short exact sequences of modules.
(Compare with (3.3.2) on p. 32.)
f g
Definition 6.1.11 Let 0 RM 1 RM 2 RM 3 0 be a short exact
sequence in R-Mod. We say the sequence splits if im(f ) = ker(g) is a direct
summand in R M 2 .
Theorem 6.1.12 The following conditions are equivalent for a unital ring R.
Proof (a) ⇒ (b) Let RN ≤ RMand consider the short exact sequence
0 RN RM R M /R N 0. . By hypothesis, this sequence splits
and hence R N is a direct summand in R M.
(b) ⇒ (c) ⇒ (d) ⇒ (e) are obviously valid.
(b) ⇒ (a) If each submodule of a module in R-Mod is a direct summand so in
particular the image 0 RM 1 R M 2 in a given short exact sequence as in
Definition 6.1.11.
(e) ⇒ (b) Let R M ∈ R-Mod. For m ∈ M, the cyclic module Rm is
semisimple since it is a homomorphic image of the semisimple module R R
(m)
(Proposition 6.1.6). By Proposition 6.1.8, Rm is thus the sum j ∈Jm R Nj of
simple submodules R Nj(m) . As R M = m∈M Rm, it follows that R M is the sum of
simple submodules and therefore, by Proposition 6.1.8, is semisimple.
We can now introduce the class of rings this book is centred around.
64 6 Simple and Semisimple Modules
Definition 6.1.13 A unital ring R satisfying any, and hence every, condition in
Theorem 6.1.12 is called (classically) semisimple or completely reducible.
Remarks 6.1.14
In this section we continue our study of the nice properties that semisimple rings
enjoy; this will lead us to the Artin–Wedderburn theorem in the next chapter. To this
end, we introduce two extremely important classes of modules.
g
RM RP
f
g0
RN
6.2 Projective and Injective Modules 65
g
RM RI
f
g0
RN
(a) R is semisimple;
(b) every left R-module is projective;
(c) every left R-module is injective.
Proof (a) ⇒ (b) Let R P ∈ R-Mod be part of the following diagram in which f is
an epimorphism
RP
g0
RM RN 0
f
66 6 Simple and Semisimple Modules
RP
g
g0
0 ker(f ) RM RN 0
f
f
0 RN RM
g0
RI
f
0 RN RM R M/ im(f ) 0
f
g
g0
RI
f
(c) ⇒ (a) Let 0 RK RM RN 0 be a short exact sequence
in R-Mod. By hypothesis, R K is injective; thus, for g0 = idK , we find f ∈
HomR (R M, R K) with f f = g0 = idK . Therefore the sequence splits, and
Theorem 6.1.12 entails that R is semisimple.
(c) (a)
(e)
(d) (b)
The implications (c) ⇒ (a) and (d) ⇒ (b) follow from Theorems 4.1 and
5.1.3, respectively. In order to show (a) ⇒ (e), suppose R M = i∈I R M i is
a decomposition of R M into a direct sum of simple submodulesR M i . As R M
is finitely generated, there is a finite subset J ⊆ I such that j ∈J R Mj =
j ∈J R Mj = R M, by Exercise 4.3.7. Since R M i ∩ R Mj = 0 for i = j , it follows
that J = I so that (e) holds.
In order to show (b) ⇒ (e) suppose R M is finitely cogenerated. Let R M =
i∈I R M i be a decomposition of R M into a direct sum of simple submodules R M i .
Fix j ∈ J and put R Nj = i=j R M i . In this way, we obtain a family of submodules
of R M with the property j ∈I R N j = 0, as R Mj ⊆ R Nj . By Exercise 5.5.1, there
is a finite subset J ⊆ I such that j ∈J R Nj = 0. Let k ∈ I \ J ; then R M k ⊆ R Mj
for any j ∈ J by simplicity. Hence R M k ⊆ R Nj for all j ∈ J . It follows that
R M k = 0. As a result, I \ J = ∅, that is, I = J is finite and (e) holds.
68 6 Simple and Semisimple Modules
Finally suppose that R M = i∈I R M i for a finite set I . If all R M i are simple,
they are both Noetherian as well as Artinian and R M inherits these properties by
Corollaries 4.1.3 and 5.1.6, respectively. This shows (e) ⇒ (c) and (e) ⇒ (d).
From this theorem, we gain a lot of insight into the structure of a semisimple ring
R as the module R R is clearly finitely generated.
Corollary 6.2.5 Every semisimple unital ring R is both left Noetherian as well as
left Artinian. Moreover, R = ni=1 Li for a finite set {L1 , . . . , Ln } of minimal left
ideals of R.
It is clear that there is a right-handed version for each of the last few results.
However, we aim for a symmetric version of the above corollary which will be
achieved in the next chapter.
Sometimes, to be injective does not require any additional property; e.g., in the
category of vector spaces, every object is injective. On the other hand, in AGr, the
injective objects are precisely the divisible abelian groups and the full subcategory
of finite abelian groups has no projective objects at all. We shall focus on injective
objects in the following, trying to illustrate what they might be good for. Evidently,
projectivity is the dual concept to injectivity.
In some categories the requirement on an injective object according to the above
definition is too restrictive: e.g., in Ban∞ , the category of all Banach spaces,
the categorical definition would imply that C is not injective. To see this, let 1
denote the Banach space of all absolutely summable sequences with the norm
(ξn )n∈N 1 = ∞ n=1 |ξn | and let c0 denote the Banach space of all null sequences
(sequences converging to 0) with the norm (ξn )n∈N ∞ = sup n |ξn |. Define a
∞
contractive linear functional g0 : 1 → C by g0 ((ξn )n∈N ) = n=1 ξn . Upon
1
embedding into c0 canonically we see that g0 cannot be extended to a bounded
linear functional g : c0 → C. Since C is complemented in every non-zero Banach
6.3 Projective and Injective Objects 69
space, it follows that 0 is the only injective object according to the categorical
definition. (Use Exercise 6.4.5 to see this.) An analogous argument shows that 0
is the only projective one, too.
The solution to this problem is to restrict the class of admissible monomorphisms.
g
B I
f
g0
injectives (where M is simply the class of all monomorphisms). This follows from
the fact that HomZ (R, G) is injective in R-Mod for every divisible abelian group.
For the details, see [29], Sect. 2.4 or [21], Sect. I.8.
70 6 Simple and Semisimple Modules
is exact. It can be shown that, if C has enough M-injectives, then every object in C
has an injective resolution.
Let F : C → AGr be a left exact functor (compare Sect. 4.2.3), where C is an
abelian category with enough M-injectives. The right derived functors of F, Rn F,
n ∈ N0 , are defined as follows. For an object A in C, choose an injective resolution
as above and consider the complex
and denote by H n (F(I n )) = ker F(d (n) )/ imF(d (n−1) ), n ≥ 1 the nth cohomology
group of the complex. Then Rn F(A) = H n (F(I n )) for n ≥ 1 and R0 F(A) = F (A).
Derived functors are at the heart of Homological Algebra; for more information,
see, e.g., [21] or [29].
6.4 Exercises
Exercise 6.4.1 Let R be a unital ring, and let RM ∈ R-Mod. Show that the
following conditions on R M are equivalent.
(a) RM is simple;
(b) RM is cyclic and every non-zero element m ∈ R M generates R M;
(c) ∼
R M = R/I for some maximal left ideal I of R.
Exercise 6.4.3 Show that the index sets of any two decompositions of a non-
zero semisimple module into a direct sum of simple submodules have the same
cardinality.
6.4 Exercises 71
f g
0 RM1 RM2 RM3 0
be a short exact sequence in R-Mod. Show that the sequence splits if and only if
either of the following conditions holds.
Exercise 6.4.7 Let R P ∈ R-Mod for a unital ring R. Show that the following
conditions on R P are equivalent.
(a) R P is projective;
(b) R P is (isomorphic to) a direct summand of a free left R-module;
(c) every epimorphism onto R P splits.
Exercise 6.4.9 Show that a unital ring R which is simple as a left R-module is a
division ring.
72 6 Simple and Semisimple Modules
Exercise 6.4.10 Let R M ∈ R-Mod for a unital ring R and let n ∈ N. Given
f ∈ EndR (R M n ), let αk be the composition
i f pk
n n
0 RM RM RM RM 0
of the projection pk onto the kth factor with f with the injection i of the th
Show that f → (αk ) defines a ring isomorphism between EndR (R M )
summand. n
and Mn EndR (R M) .
Exercise 6.4.11 Show that, for every unital ring R and each n ∈ N, the centre
Z(Mn (R)) of Mn (R) is isomorphic to Z(R).
In this chapter, we shall obtain the full structure theorem for semisimple rings.
This result, due to J. H. M. Wedderburn (1907) for semisimple finite-dimensional
algebras and to E. Artin (1927) in the general case, enables us to determine
completely this class of rings from the more elementary class of division rings.
It is generally regarded as the first major result in the abstract structure theory of
rings. In Sect. 7.2 below, we will briefly discuss the semisimplicity of group rings
via Maschke’s Theorem 7.2.1. A detailed proof of the Hopkins–Levitzki theorem,
already stated in Theorem 5.3.2, is given in Sect. 7.3.
Proposition 7.1.1 Let R be a unital ring, and let n ∈ N. Every ideal J in Mn (R) is
of the form J = Mn (I ) for some ideal I in R. Thus, Mn (R) is simple if and only if
R is simple.
In the next result we collect a number of properties of matrix rings over division
rings.
Theorem 7.1.2 Let D be a division ring, and let R = Mn (D) for some n ∈ N.
Then
Proof Every division ring is simple; hence R is simple by Proposition 7.1.1. Let
V be the n-tuple column space D n , viewed as a right D-vector space. Then R =
Mn (D) acts on the left of V by matrix multiplication, so R V ∈ R-Mod is a faithful
unital left R-module. In fact, Mn (D) can be identified with EndD (VD ) in the usual
way by choosing a basis of VD . An argument analogous to the one in Examples 6.1.3
shows that R V is a simple R-module.
For each i ∈ {1, . . . , n}, let Li denote the left ideal of R consisting of those n×n-
matrices over D whose columns other than the ith are zero. Then R = L1 ⊕. . .⊕Ln
and, since R Li ∼= R V , we get R R ∼ = R V n . By Proposition 6.1.8, R R is semisimple
so the ring R is (left) semisimple. By Corollary 6.2.5, R is thus left Artinian as well
as left Noetherian, which completes the proof of (i).
The uniqueness of R V stated in (ii) now follows from Proposition 6.1.9.
In order to prove (iii), we define a mapping σ : D → EndR (R V ) by v · σ (d) =
v · d, v ∈ V , d ∈ D. Note that we write here endomorphisms on the right in order
to avoid the opposite ring D op ; see Remark 7.1.3 below. Clearly, σ is a unital ring
homomorphism and, since D acts faithfully on V , σ is injective. To show that σ is
⎛ ⎞ ⎛ ⎞
1 d
⎜0⎟ ⎜∗ ⎟
⎜ ⎟ ⎜ ⎟
surjective, let f ∈ EndR (R V ). Then ⎜ . ⎟ f = ⎜ . ⎟ for some d ∈ D and hence,
⎝ .. ⎠ ⎝ .. ⎠
0 ∗
⎛ ⎛ ⎞⎞ ⎛⎛ ⎞ ⎞
⎛ ⎞ ⎛ ⎞ 1 ⎛ ⎞ 1
a1 ⎜ 1 a 0 . . . 0 ⎜0⎟⎟ a 1 . . . ⎜⎜0⎟ ⎟
⎜ .. ⎟ ⎜⎜ . ⎟ ⎜ ⎟⎟ ⎜. ⎟ ⎜⎜ ⎟ ⎟
⎝ . ⎠ f = ⎜⎝ .. ⎠ ⎜ .. ⎟⎟ f = ⎝ .. ⎠ ⎜⎜ .. ⎟ f ⎟
an
⎝
an
0 ⎝ . ⎠⎠
an
0 ⎝⎝ . ⎠ ⎠
0 0
⎛ ⎞
⎛ ⎞ d ⎛ ⎞ ⎛ ⎞
a1 . . . ⎜ ⎟ a1 d a1
⎜ .. ⎟ ⎜ ∗⎟ ⎜ .. ⎟ ⎜ .. ⎟
=⎝. ⎠ ⎜ .. ⎟ = ⎝ . ⎠ = ⎝ . ⎠ σ (d)
an
0 ⎝.⎠
an d an
∗
so that f = σ (d).
7.1 The Structure of Semisimple Rings 75
Remarks 7.1.3
Historical Note Joseph Henry Maclagan Wedderburn was born in 1882 in Scotland
as the 10th of 14 children. He entered the University of Edinburgh at the age of
16 and was elected Fellow of the Royal Society of Edinburgh when he was only
21 years old. After further studies at Leipzig, Berlin and Chicago he returned to
Edinburgh in 1905 and was awarded his DSc and his PhD in 1908. Wedderburn
served in the First World War between 1914 and 1918 after which he became
an Associate Professor at Princeton in 1921. Among his (three) PhD students at
Princeton was Nathan Jacobson. He made important advances in the theory of rings,
algebras and matrix theory; among these is his celebrated result that every finite
division ring is a field which also has important consequences for finite projective
geometry. His 1907 paper “On hypercomplex numbers" contains the structure
theorem for finite-dimensional semisimple algebras. See also Artin’s paper [5].
Wedderburn died in Princeton in October 1948.
https://www-history.mcs.st-andrews.ac.uk/Biographies/Wedderburn.html
We shall now see that the above Example 7.1.4 already exhausts all possibilities.
Proof Let R be a semisimple unital ring. By Corollary 6.2.5, there are finitely many
minimal left ideals L1 , . . . , Ln in R such that
R = L1 ⊕ . . . ⊕ Ln .
76 7 The Artin–Wedderburn Theorem
Some of the Li ’s may be isomorphic to each other as left R-modules; from each
isomorphism class we pick one representative and denote this simple submodule of
R R by R Mj (so that R Mj =
∼
R M i for j = i). Suppose we have such isomorphism
classes. Then there exist n1 , . . . , n ∈ N such that
RR
∼ n n
= R M 11 ⊕ . . . ⊕ R M . (7.1.1)
EndR (R R) ∼
= EndR
ni
RMi
i=1
nj
ni
= HomR RMi , R Mj
i=1 j =1
∼
= HomR
ni
RMi ,
nj
R Mj
i=1 j =1
nj
ni
= HomR RMi , R Mj
i=1 j =1
nj
∼
= HomR ni
R M i , R Mj ,
i=1 j =1
where we used the universal properties of the direct sum and the direct product of
modules and the fact that these constructions coincides with each other if the index
set is finite, see Exercise 2.7.10.
We now apply a similar chain of arguments to compute the individual homomor-
phism groups.
n
ni nj
HomR R M ni i , R Mj j = HomR RMi , R Mj
ni nj
∼
= HomR R M i , R Mj
⎧ ni ni
⎨
= HomR RMi , RMi if i = j
⎩
0 if i = j,
7.1 The Structure of Semisimple Rings 77
where we used that HomR R M i , R Mj = 0 for i = j since both modules are simple
and non-isomorphic. Putting these two computations together we find
ni
EndR (R R) ∼
ni
= HomR RMi , RMi
i=1
= EndR (R M ni i )
i=1
∼
= Mni (Di ).
i=1
R∼
= Mn1 (D1 ) × . . . × Mn (D )
Remark 7.1.6 If we want to avoid writing endomorphisms on the right in the above
proof, we obtain instead
R∼
= EndR (R R)op ∼
op
= Mni (Di )op = Mni (Di ),
i=1 i=1
Corollary 7.1.7 A unital ring is left semisimple if and only if it is right semisimple.
Proof This follows from the fact that we have an analogous theory of semisimple
right modules and right semisimple rings leading to the same Artin–Wedderburn
Theorem which is a symmetric statement.
Proof While the “if”-part is immediate, the “only if”-part follows from Theo-
rem 7.1.5 together with the fact that Mn (D) is commutative if and only if n = 1
and D is commutative.
78 7 The Artin–Wedderburn Theorem
7.1.9 Uniqueness Theorem for Semisimple Rings Let R be a unital ring and
suppose
n m
R= Ri = Rj
i=1 j =1
Proof We will identify each Ri and Rj with a two-sided ideal in R in the canonical
way. Then Ri = RRi = Ri R and Rj = RRj = Rj R for all i, j . Therefore
m
Ri = Ri R = Ri Rj (1 ≤ i ≤ n).
j =1
Putting Theorems 7.1.5 and 7.1.9 together we arrive at a very satisfactory result
illustrating in a beautiful way what the structure theory of rings is about. This result
is at the very heart of our book.
Corollary 7.1.10 A unital ring R is semisimple if and only if it is the unique finite
direct product of simple Artinian rings R1 , . . . , Rn .
Note that we must insert “Artinian” in the sufficiency condition as not every simple
ring is semisimple; this will be illustrated in the example below.
Example 7.1.11 Let D be a division ring (for a concrete example, take D = Q). For
each n ∈ N, let Rn = M2n (D) which is a simple unital ring by Proposition 7.1.1.
We define unital ring monomorphisms Rn → Rn+1 by
!
A 0
M (D) A −→
2n ∈ M2n+1 (D).
0 A
7.1 The Structure of Semisimple Rings 79
Identifying
∞ Rn with a unital subring of Rn+1 in this way, we put R = M2∞ (D) =
n=1 Rn . (This is an instance of a “direct limit” of a directed system of rings.)
Clearly, R is a unital ring.
In order to show that R is simple, let I be a non-zero ideal of R. Then there is
some Rn such that I ∩ Rn = 0. As I ∩ Rn is an ideal of Rn and Rn is simple, we
obtain I ∩ Rn = Rn , that is, Rn ⊆ I . Since 1R = 1Rn ∈ Rn it follows that I = R.
We next show that R is not left Artinian (and hence cannot be semisimple, by
Corollary 6.2.5). For n ∈ N, let en denote the matrix unit e11 in Rn . Then en+1 =
en+1 en ∈ Rn+1 . For instance,
!
10
e1 in R1 = M2 (D) is ,
00
⎛ ⎞
1000
⎜0 0 0 0⎟
e1 in R2 = M4 (D) is ⎜ ⎟
⎝0 0 1 0⎠ ,
0000
⎛ ⎞
1000
⎜0 0 0 0⎟
e2 in R2 = M4 (D) is ⎜ ⎟
⎝0 0 0 0⎠ ,
0000
Suppose en ∈ Ren+1 for some n. Then en = aen+1 where a ∈ Rj = M2j (D) for
some j > n. The (2n +1, 2n +1)-entry of en in Rn+1 is 1, however the (2n +1, 2n +
1)-entry of aen+1 is 0, which is impossible. As a result, the descending chain cannot
become stationary and R is not left Artinian.
As we shall see now, the Artinian property is the one that turns a simple ring into
a semisimple one.
Proposition 7.1.12 For every simple unital ring R, the following conditions are
equivalent.
(a) R is semisimple;
(b) R is left Artinian;
(c) R = Mn (D) for some division ring D and n ∈ N.
Proof The implications (c) ⇒ (a) and (c) ⇒ (b) are provided by Theorem 7.1.2.
(a) ⇒ (b) is Corollary 6.2.5 and (a) ⇒ (c) follows from Theorem 7.1.5.
80 7 The Artin–Wedderburn Theorem
This result shows that a right Artinian simple ring must be left Artinian (and vice
versa) since a similar characterisation as in Proposition 7.1.12 is available. For this
reason, one simply speaks of ‘Artinian simple rings’.
Remark 7.1.13 For an approach to the Artin–Wedderburn theorem from the view-
point of abelian categories and Morita equivalence the reader is invited to consult
[30, Sect. 4.12].
In this short section we shall discuss an important class of semisimple rings that is
related to group representations. In Sect. 1.1.5 we indicated the connections between
finite-dimensional representations of a group G and modules over the group ring
K[G], where K is a field. Indeed, given a finite group G and a field K, there is
a one-to-one correspondence between K-linear representations of G and finitely
generated left K[G]-modules. It turns out that the relation between the order of G
and the characteristic of K determines the structure of the group ring.
i
Proof Set R = K[G] and let the sequence 0 RN R M in R-Mod be
exact. As N and M are K-vector spaces, we can define a linear splitting p0 : M →
N easily: simply extend a basis of N to a basis of M and define p0 by p0 (i(b)) = b
for all basis elements in N and p0 = 0 otherwise. Then p0 i = idN . We intend to
“upgrade” p0 to a splitting map p ∈ R-Mod.
To this end, suppose that |G| = n and define p : M → N by
1
p(m) = g · p0 (g −1 m) (m ∈ M).
n
g∈G
7.3 The Hopkins–Levitzki Theorem 81
1 1
p(hm) = g · p0 (g −1 hm) = h(h−1 g) · p0 ((h−1 g)−1 m)
n n
g∈G g∈G
1
=h· (h−1 g) · p0 ((h−1 g)−1 m)
n −1
h g∈G
= h · p(m)
1 1 1
p(i(m)) = g · p0 (i(g −1 m)) = g(g −1 m) = m = m.
n n n
g∈G g∈G g∈G
On the other hand, it is not too hard to show that, if the characteristic of K does
divide |G|, then K[G] cannot be semisimple; see, e.g., [7, Proposition 4.1.5].
Since for every algebraically closed field K, the endomorphism ring
EndK[G] (R M) of a simple left K[G]-module R M is isomorphic to K
(Exercise 7.4.17), a combination of Theorems 7.1.5 and 7.2.1 yields the following
result.
Corollary 7.2.2 Let G be a finite group. Then the group ring C[G] is isomorphic
to a finite direct sum of matrix rings over C.
Proposition 7.3.2 In every ring R, the following sets agree with each other:
1 − xy is invertible ⇐⇒ 1 − yx is invertible .
By symmetry, we only have to verify one implication, and the observation will
establish the equivalence of (iii) and (iv) in the proposition.
Suppose that (1 − xy)u = 1 = u(1 − yx) for some u ∈ R. Then
show that 1 − yx is invertible with inverse yux + 1. Clearly (v) implies (iii), so
assume (iii) holds. Then 1 − zxy is invertible for all z, x ∈ R whence, by the
observation, 1 − xyz is invertible, which was to show.
Suppose y ∈ R is not in the intersection of all maximal left ideals of R and let
L be a maximal left ideal such that y ∈ / L. Then L + Ry = R and there is x ∈ R
with 1 − xy ∈ L. It follows that 1 − xy cannot be (left) invertible as L = R which
shows that the set described in (iii) is contained in the intersection of all maximal
left ideals. On the other hand, if 1 − xy is not invertible for some x ∈ R then the left
7.3 The Hopkins–Levitzki Theorem 83
ideal R(1 − xy) is proper, thus contained in a maximal left ideal L (by a standard
application of Zorn’s lemma). It follows that y ∈ / L which establishes the reverse
inclusion.
The equality of the sets in (ii) and (iv) is shown analogously.
Corollary 7.3.6 For every ring R, the quotient ring R/rad(R) is semiprimitive.
Proof Let us denote R/rad(R) by R̂ and its elements by x̂ etc. Note that rad(R)
is a proper ideal as 1 ∈ / rad(R). Take y ∈ rad(R̂); by Proposition 7.3.2, 1 − x̂ ŷ
is invertible in R̂ for all x̂ ∈ R̂. Let u ∈ R be such that û(1 − x̂ ŷ) = 1 in R̂.
Then u(1 − xy) = 1 + r for some r ∈ rad(R). As 1 + r is invertible in R, by
Proposition 7.3.2, u(1 − xy) and hence 1 − xy is left invertible for all x ∈ R. A
similar argument shows that 1 − xy is right invertible, thus invertible for all x ∈ R
which implies that y ∈ rad(R). As a result, ŷ = 0.
By Corollary 7.3.3, every simple left R-module is a simple left R̂-module, where
R̂ = R/rad(R). Conversely, every simple left R̂-module is a simple left R-module
in a canonical way. (See Exercise 2.7.3.)
84 7 The Artin–Wedderburn Theorem
The reader may wonder why we introduced the Jacobson radical en route to a
proof of the Hopkins–Levitzki theorem. The reason will be revealed to them now.
Proposition 7.3.7 For every ring R, the following properties are equivalent.
(a) R is semisimple;
(b) R is left Artinian and semiprimitive.
Proof (a) ⇒ (b) Every semisimple ring is left Artinian, by Corollary 6.2.5, so we
only need to show that rad(R) = 0. Since rad(R) is a direct summand of R R, there
is a left ideal L ⊆ R such that L ⊕ rad(R) = R. Let e ∈ L, f ∈ rad(R) be such that
e + f = 1. Then e2 + f e = e and, as f e ∈ rad(R) ∩ L = 0, it follows that e2 = e.
Similarly, f 2 = f so that e and f are orthogonal idempotents in R. As e = 1 − f is
invertible, it follows that e = 1 and so f = 0. We conclude that rad(R) = Rf = 0
as claimed.
(b) ⇒ (a) We first observe the following: let L be a non-zero left ideal of R. As
R is left Artinian, L contains a minimal left ideal (use Theorem 5.1.3 (c)). Every
minimal left ideal N of R is a direct summand of R R. Indeed, since rad(R) = 0 and
N = 0 there exists a maximal left ideal K of R not containing N. Then N ∩ K = 0
and R R = N ⊕ K.
Now suppose that R is not semisimple. Take a minimal left ideal N1 in R and
write R R = N1 ⊕ K1 for a maximal left ideal K1 as above. As K1 = 0 there
exists a minimal left ideal N2 ⊆ K1 . As before, N2 is a direct summand of R R and
thus of K1 (use Exercise 2.7.8). We therefore can write K1 = N2 ⊕ K2 for a left
ideal K2 K1 . By induction, we obtain a strictly descending chain of left ideals
K1 K2 K3 . . . which contradicts the assumption that R is left Artinian.
The fact that every semisimple ring is semiprimitive can also be deduced from
the Artin–Wedderburn theorem together with Exercise 7.4.7 since simple rings are
trivially semiprimitive. Of course, the argument above uses a less heavy tool.
We have now taken a substantial step towards the Hopkins–Levitzki theorem. If
R is left Artinian with rad(R) = 0 then, by the above proposition, R is semisimple
and hence left Noetherian, by Corollary 6.2.5. In particular, for any left Artinian ring
R, R/rad(R) is left Noetherian, by Corollaries 5.1.5 and 7.3.6. But, in the statement
of Theorem 5.3.2, there is no assumption on the Jacobson radical so what seems to
be an obstruction will have to be removed. We shall next see how an additional tool
enables us to achieve this goal.
The idea is to use a short exact sequence of modules like
together with Theorem 4.1.1 to obtain that R R is Noetherian from the same
properties of the outer modules. However, a priori, we do not know that rad(R)
is Noetherian, or even just finitely generated (though it will be once we have shown
7.3 The Hopkins–Levitzki Theorem 85
that R is left Noetherian). The remedy is to work with a suitable power of the radical
instead.
Throughout the remainder of this section we will denote rad(R) by J and
R/rad(R) by R̂.
For some n ∈ N, we replace (7.3.1) by
0 J n− 1 R R/J n −1 0 (7.3.2)
and use that, if R is left Artinian, then R̂ is semisimple. The choice of n will become
clear immediately.
Call an ideal I of R nil if every element in I is nilpotent and call I nilpotent if
I n = 0 for some n ∈ N.
Lemma 7.3.8 For every ring R, the Jacobson radical J contains each nil ideal of
R. If R is left Artinian then J is nilpotent.
is not empty, thus contains a minimal element, say L0 (Theorem 5.1.3). Take y ∈ L0
such that Iy = 0. Since I Iy = 0 and Iy ⊆ L0 it follows that Iy = L0 . Therefore
y = xy for some x ∈ I which implies that (1 − x)y = 0. Since x ∈ J n ⊆ J , 1 − x
is invertible so that y = 0, a contradiction. We conclude that J n = I = 0.
The reader may want to compare the above proof with the argument in Propo-
sition 5.2.5; see also Exercise 7.4.8. We observe that, since every nilpotent ideal is
nil, the Jacobson radical of a left Artinian ring is the largest nil ideal. On the other
hand, it is the smallest ideal such that the corresponding quotient is semisimple, as
we shall observe now.
Proposition 7.3.9 In a left Artinian ring R, the Jacobson radical is the smallest
ideal of R whose corresponding quotient is semisimple.
Proof By Corollary 7.3.6 and Proposition 7.3.7, R/J is semisimple. Conversely, let
K ⊆ R be an ideal such that R/K is a semisimple left R-module, and let ρ : R →
R/K be the canonical epimorphism. As R/K is semisimple, there is a left ideal W
in R, containing K, such that ρ(J ) ⊕ ρ(W ) = R/K. Suppose that W = R. Then
there is a maximal left ideal L of R which contains W . By definition, J ⊆ L so
86 7 The Artin–Wedderburn Theorem
Note also that, if K is an ideal of the left Artinian ring R and J ⊆ K, then R/K
is semisimple, as a homomorphic image of R/J .
We need another auxiliary result which can be obtained, for instance, using the
convenient characterisation of the Jacobson radical contained in Proposition 7.3.2.
Lemma 7.3.10 Let I be an ideal of the ring R such that I ⊆ J . Then rad(R/I ) =
J /I .
The above result will allow us to control the quotient R/J n−1 in (7.3.2). In order
to determine the behaviour of J n−1 as an R-module, we first need a definition.
Definition 7.3.11 Let I be an ideal of the ring R. We put rad(I ) = rad(R) ∩ I and
call it the Jacobson radical of I .
Proof Let us abbreviate the set on the right-hand side of (7.3.3) by J . Note at first
that, by definition,
rad(I ) = J ∩ I = {L ∩ I | L maximal left ideal of R}
= {L ∩ I | L maximal left ideal of R and I L},
7.3 The Hopkins–Levitzki Theorem 87
R/L = (L + I )/L ∼
= I /(L ∩ I )
Combining this lemma with Proposition 6.1.10 we obtain the next result.
Proposition 7.3.14 For every ideal I of a left Artinian ring R, we have rad(I ) = J I
and I /rad(I ) is semisimple.
Proof We prove the second statement first. From Theorem 3.1.3 we obtain
I /rad(I ) = I /(J ∩ I ) ∼
= (I + J )/J −→ R/J (7.3.4)
which together with Propositions 7.3.7 and 6.1.6 yields that I /rad(I ) is semisimple.
Clearly, J I ⊆ J ∩ I . By Exercise 7.4.10, rad(I /J I ) = rad(I )/J I so, in order to
establish the reverse inclusion, it suffices to show that rad(I /J I ) vanishes. Set M =
I /J I ; this is an Artinian left R-module, by Corollary 5.1.5. We can regard M as a
left R/J -module (Exercise 2.7.3 and Corollary 7.3.3); since R/J is a semisimple
ring (Corollary 7.3.6 and Proposition 7.3.7) it follows that M is a semisimple R/J -
module (Theorem 6.1.12) and hence semisimple as an R-module and an R/J I -
module (Exercise 2.7.3 again). By Corollary 7.3.13 above, rad(I /J I ) = 0 which
was to prove.
88 7 The Artin–Wedderburn Theorem
Remark 7.3.15 Let I be an ideal in a left Artinian ring R. The embedding (as rings)
of I /rad(I ) in (7.3.4) enables us to consider I /rad(I ) as an ideal in R/J ; call this
Ī . Then
Proof of Theorem 5.3.2 Let R be a left Artinian ring and let n ∈ N be the least
integer k such that J k = 0 (Lemma 7.3.8). We prove that R is left Noetherian
by induction on n. Let n = 1; then J = 0 and the assertion follows from
Proposition 7.3.7 as every semisimple ring is left Noetherian (Corollary 6.2.5).
Now assume that n > 1 and, for every 1 ≤ k < n and every left Artinian
ring S with rad(S)k = 0, we have S is left Noetherian. We will apply the short
exact sequence (7.3.2) with S = R/J n−1 . By Corollary 5.1.5, S is left Artinian. By
n−1
Lemma 7.3.10, rad(S) = rad(R)/J n−1 and hence, rad(S)n−1 = J /J n−1 = 0.
Thus, by induction hypothesis, S is left Noetherian.
It remains to show that J n−1 is left Noetherian and then to apply Theorem 4.1.1
to complete the proof. Clearly, J n−1 is left Artinian and, by Proposition 7.3.14,
rad(J n−1 ) = J J n−1 = 0 and J n−1 is semisimple (both as a left R/J - and a left
R-module). By Theorem 6.2.4, J n−1 is left Noetherian, which completes the claim.
Remark 7.3.16 The above proof is inspired by the arguments in Sect. 15 of [2]
where the theory of radicals of modules is discussed in detail.
Remark 7.3.17 The most commonly found proof of the Hopkins–Levitzki theorem
uses composition series and relies on the Jordan–Hölder theorem. Let R be a ring.
A composition series of a left R-module M is a descending chain of submodules
M = M0 ⊃ M1 ⊃ M2 ⊃ . . . ⊃ M = 0
such that each quotient Mi−1 /Mi is simple. In this case, is the length of the
composition series and the simple modules are called the composition factors. The
Jordan–Hölder theorem states that any two composition series of a module have the
same length and the composition factors are, up to permutation and isomorphism,
uniquely determined. This allows one to introduce the length of a module as the
length of a composition series, if such exists. It can be shown that a module has
a finite composition series if and only if it is both Artinian and Noetherian. The
7.4 Exercises 89
composition series that one uses in a proof of the Hopkins–Levitzki theorem derive
from the chain of inclusions
R ⊃ J ⊃ J2 ⊃ ... ⊃ Jn = 0
7.4 Exercises
Exercise 7.4.2 Show that every non-zero Artinian module contains a simple sub-
module.
Exercise 7.4.3 For a ring R, define its socle soc(R) as the sum of all minimal left
ideals of R and if R has no minimal left ideals, put soc(R) = 0. Show that soc(R)
is an ideal of R.
Exercise 7.4.4 For a ring R, an essential left ideal L of R is defined by the property
that L ∩ J = 0 for every non-zero left ideal J of R. Using Zorn’s Lemma show that,
for every left ideal I of R there is a left ideal I of R such that I ⊕ I is an essential
left ideal.
Exercise 7.4.5 Let R be a ring with non-zero socle soc(R). Using Zorn’s Lemma
show that, for each x ∈ R \ soc(R), there is an essential left ideal L of R containing
soc(R) such that x ∈ / L. Use this to prove that soc(R) is equal to the intersection of
all essential left ideals of R.
Exercise 7.4.6 Show that a unital ring R is semiprimitive if and only if it has a
faithful semisimple left module.
90 7 The Artin–Wedderburn Theorem
Exercise 7.4.9 A unital ring R is called semiprime if, for every ideal I of R, the
condition I 2 = 0 entails I = 0. Prove the following analogue of Proposition 7.3.7:
R is semisimple if and only if R is left Artinian and semiprime.
Exercise 7.4.10 Follow the argument in the proof of Lemma 7.3.10 to obtain the
following extension of this lemma: Let I be an ideal in a unital ring R and let K be
an ideal of R such that K ⊆ rad(I ). Then rad(I /K) = rad(I )/K.
Exercise 7.4.12 Let R be a unital ring and let n ∈ N. Using Proposition 7.1.1, show
that rad(Mn (R)) = Mn (rad(R)).
Exercise 7.4.13 A unital ring R is called von Neumann regular if, for each r ∈ R,
there is s ∈ R such that rsr = r. Show that, if R M ∈ R-Mod is semisimple, then
EndR (R M) is von Neumann regular.
Exercise 7.4.14 Let R be a von Neumann regular ring. Show that every principal
ideal of R (that is, an ideal which is generated by one element) is generated by an
idempotent.
Exercise 7.4.15 Show that every von Neumann regular ring is semiprimitive. Let
C[0, 1] be the ring of all continuous real-valued functions on the unit interval [0, 1].
Prove that C[0, 1] is semiprimitive but not von Neumann regular.
Exercise 7.4.16 Let R be a reduced ring in which every prime ideal is maximal.
Show that R is von Neumann regular.
Exercise 7.4.17 An algebra over a field K is a ring A which at the same time is a
K-vector space such that λ(xy) = x(λy) for all λ ∈ K and x, y ∈ A. Suppose K is
7.4 Exercises 91
RMi
RL lim I R M i
−→
f ji
R Mj
Exercise 7.4.19 (Nakayama’s Lemma) Let I be a left ideal of the (unital) ring R.
Show that the following three conditions are equivalent.
(i) I ⊆ rad(R);
(ii) for every M ∈ R-mod, I · M = M ⇒ M = 0;
(iii) for all R N , R M ∈ R-Mod with N ⊆ M such that R M/R N is finitely
generated, N + I · M = M ⇒ N = M.
Tensor Products of Modules
8
This chapter is devoted to a more sophisticated construction with modules, the so-
called tensor product. It is extremely useful, though fairly abstract. Its main benefit
lies in the fact that it allows us to convert bilinear mappings into homomorphisms of
abelian groups. The relations between tensor products and homomorphism groups
is fundamental and will lead us to the concept of adjoint functor in the later part of
the chapter.
Here, we will be dealing with left and right modules over a unital ring R at the same
time.
M N T (M, N)
Theorem 8.1.4 For all MR ∈ Mod-R and R N ∈ R-Mod over a unital ring R
there is a tensor product which is unique up to isomorphism.
Proof We will prove the uniqueness first. Suppose that both (T , τ ) and (S, σ )
satisfy the universal property of (8.1.3). By applying it to σ and to τ , respectively
we obtain the following two commutative diagrams
M N T M N S
S T
We claim that the group homomorphisms σ̂ and τ̂ are inverses of each other so that
T and S are isomorphic. We have
σ̂ τ̂ σ = σ̂ τ = σ and τ̂ σ̂ τ = τ̂ σ = τ.
M N S M N T
S T
8.1 Tensor Product of Modules 95
τ : M × N → T, τ (x, y) = π(x, y)
This shows that K = ker π ⊆ ker β̃ and so π(f ) = 0 implies β̃(f ) = 0 for each
f ∈ F . As a result, β̂ is well defined.
Finally, β̂ ◦ τ = β̂ ◦ π|M×N = β yields the universal property.
96 8 Tensor Products of Modules
Remarks 8.1.5
k
ni (xi , yi ) + K,
i=1
The tensor product is rather well behaved and compatible with other construc-
tions with modules. As an example, we mention the result below.
(i) (M ⊕ M ) ⊗R N ∼
= M ⊗R N ⊕ M ⊗R N;
(ii) M ⊗R N ∼
= N ⊗R M provided R is commutative;
(iii) (M ⊗R N) ⊗R L ∼
= M ⊗R (N ⊗R L) provided R is commutative.
Proof We leave the verification of (i) and (iii) to the reader; see the Exercises. Note
that, if R is commutative, the tensor product carries a natural R-module structure
and therefore, it makes sense to iterate the construction as in (iii).
For illustration, we prove assertion (ii). Suppose R is commutative so that we can
consider the modules M and N as modules on the left and on the right. We have the
following commutative diagram
M N N M
1 2
2
M R N N R M
1
The additive maps τ̂1 , τ̂2 are given by the universal property of the tensor product
and satisfy τ̂2 ◦ τ1 = τ2 ◦ ι and τ̂1 ◦ τ2 = τ1 ◦ ι−1 . Therefore
and
which gives τ̂2 ◦ τ̂1 = id|N⊗R M and τ̂1 ◦ τ̂2 = id|M⊗R N , respectively, where, again,
we applied the universal property. Consequently, M ⊗R N ∼ = N ⊗R M as abelian
groups.
Since, for all x ∈ M, y ∈ N,
γ : N → R ⊗R N, x → 1 ⊗ x.
Then, for each i ai ⊗ xi ∈ R ⊗R N,
γ β̂ ai ⊗ xi = γ ai · xi = 1 ⊗ ai · xi = ai ⊗ xi .
i i i i
Our next aim is to investigate the relation between the tensor product and
homomorphism groups. To this end, we first record the following special situation.
and
HomR (R ⊗R M, R N ) ∼
= HomR (R M, R N ) ∼
= HomR (R M, HomR (R, R N ))
for all R M, R N ∈ R-Mod. We shall now see that a similar isomorphism can be
established in general once we allow ourselves to work with bimodules.
HomS (S U ⊗R M, S N ) ∼
= HomR (R M, HomS (S U , S N)).
8.1 Tensor Product of Modules 99
fx (u1 + u2 ) = f ((u1 + u2 ) ⊗ x)
= f (u1 ⊗ x + u2 ⊗ x) = f (u1 ⊗ x) + f (u2 ⊗ x)
= fx (u1 ) + fx (u2 ),
fx (s · u) = f (s · u ⊗ x) = s · f (u ⊗ x)
= s · fx (u)
where the last equality comes from the definition of r · fx . Consequently, for x ∈ M
and r ∈ R, we have
and thus
βh : U × M → N, βh (u, x) = h(x)(u)
100 8 Tensor Products of Modules
In categorical language the above theorem states that the functors U ⊗R – and
HomS (U, –) are adjoint functors; compare Sect. 8.3 below.
We next introduce the tensor product of module maps.
f ⊗ g : M ⊗R N −→ M ⊗R N
In this section, K will denote a field and A, B, C will be unital K-algebras. That
is, A is a K-vector space and carries an associative multiplication (x, y) → xy
such that λ(xy) = x(λy) for all x, y ∈ A and the two distributivity laws hold.
Furthermore, there is a multiplicative identity which we will denote by 1.
Our aim is to endow the module tensor product A ⊗K B with an algebra structure
and study the properties of this algebra. To this end, we first amend the definition of
a K-balanced mapping in a natural way. A mapping β : A × B → C will be called
K-bilinear if, for all a, a1 , a2 ∈ A, b, b1 , b2 ∈ B and λ ∈ K,
Note that the only difference to the original definition of an R-balanced mapping
(p. 93) is in the last identity of the last line above which is possible as C is a vector
space too.
A slight modification of the construction of the module tensor product in
Theorem 8.1.4 yields a K-vector space A ⊗K B which has the universal property
with respect to K-bilinear mappings as defined above, that is, results in a linear
mapping β̂ from A ⊗K B into a K-vector space E.
A B A K B
There is a canonical way of turning this vector space tensor product into a K-
algebra.
Definition 8.2.1 Let A and B be unital algebras over the field K. The tensor
product algebra A ⊗K B is the vector space tensor product endowed with the unique
associative product satisfying (a ⊗ b)(c ⊗ d) = ac ⊗ bd for all a, c ∈ A and
b, d ∈ B.
It should by now be evident how to define the product alluded to above: we first
have to linearise the K-bilinear mapping
(A ×B)×(A ×B) −→ (A ×B)⊗K (A ×B), (a, b), (c, d) − → (a, b)⊗(c, d),
102 8 Tensor Products of Modules
Proposition 8.3.2 For every unital ring R, the left module R R is flat.
Corollary 8.3.3 Every projective left module over a unital ring is flat.
Recall the concept of a von Neumann regular ring R: for each r ∈ R there
is s ∈ R such that rsr = r (Exercise 7.4.13). It turns out that a unital ring R
is von Neumann regular if and only if every left R-module is flat; see, e.g., [8,
Theorem 15.22]. This requires a more detailed study of the finitely generated left
ideals in a von Neumann regular ring.
By definition, a module R N ∈ R-Mod is flat if and only if the functor – ⊗R N
is exact (compare Sect. 4.2.3). We shall now discuss the “Fundamental theorem of
tensor products” (Theorem 8.1.9) from the point of view of adjoint functors.
The naturality of η in the above definition can be made explicit as follows. For
all f ∈ MorC (A , A), g ∈ MorD (B, B ) and h ∈ MorD (F(A), B), we have
Examples 8.3.5
1. Let F : S → R-Mod be the ‘free functor’ (compare Sect. 2.5) and let
G : R-Mod → S be the forgetful functor. Then F is left adjoint to G and G
is right adjoint to F by Theorem 2.5.3. This example extends to any concrete
category which has free objects in the sense of Sect. 2.6.1.
2. Theorem 8.1.9 tells us (in essence) that, whenever R, S are unital rings and S UR ∈
S-Mod-R, then the functor S U ⊗R – is left adjoint to the functor HomS (S U , –).
3. In a similar vein, if R and S are unital rings and ρ : R → S is a unital ring
homomorphism then, considering S as a right R-module (see Exercise 1.3.3), the
‘extension-of-scalars’ functor S ⊗R – : R-Mod → S-Mod is left adjoint to the
‘restriction-of-scalars’ functor S-Mod → R-Mod.
4. Let R be a unital ring and let G be a (multiplicative) group. The construction of
the group ring (see Examples 1.2 (ix)) provides us with a functor R[–] from the
category of groups into the category of unital rings. It is left adjoint to the functor
in the opposite direction which associates to a unital ring R its group of units R ∗ .
5. We return to Example 3.2.4 from our present point of view. For a commu-
tative unital C* -algebra A, its Gelfand space (A) is homeomorphic to the
space of all multiplicative linear functionals on A equipped with the weak *-
104 8 Tensor Products of Modules
topology; thus the functor can also be viewed as the representable functor
MorAC∗ (−, C) : AC1∗ −→ Comp. In the other direction, the functor C
1
associates with a compact Hausdorff space X the morphism set MorComp (X, C).
Therefore we obtain the adjunction
where we have to take care of the fact the both functors are contravariant.
Adjoint functors are ubiquitous in Mathematics; for a detailed study see, for
instance, Sect. II.7 in [21]. They have pleasant properties; e.g., left adjoint functors
preserve colimits while right adjoint functors preserve limits. As an illustration
of the arguments, we prove a special case of the latter statement. A pullback in
a general category, if it exists, is defined by the universal property as stated in
Exercise 4.3.13.
Proof Let {Bi | i ∈ I } be a family of objects in D and let (B, pi )i∈I denote their
product (assuming it exists). To show that (G(B), G(pi ))i∈I is the product of the
family {G(Bi ) | i ∈ I } in C, let fi : A → G(Bi ), i ∈ I be morphisms in C. Let
η : F → G be an adjugant and let ξ be its inverse. For each i, ξ(fi ) : F(A) → Bi so
that there exists a unique morphism g : F(A) → B in D such that pi ◦ g = fi for
all i ∈ I . It follows that
and η(g) is the unique morphism with this property as every morphism f : A →
G(B) is of the form f = η(g ) for some g in mor(D).
To prove the second assertion,
f G(f )
Y A G(Y ) G(A)
g g G(g) G(g )
B X G(B) G(X )
f G(f )
g ◦ ρ = ξ(k). Now applying η yields G(f ) ◦ η(ρ) = h and G(g) ◦ η(ρ) = k and, as
for products, η(ρ) is the unique morphism satisfying these identities.
8.4 Exercises
Exercise 8.4.1 Let R be a commutative ring and let R M ∈ R-Mod. Show that M
becomes a right R-module in a natural way by defining x · r = r · x, x ∈ M, r ∈ R.
As a result, when considering modules over commutative rings, one usually does
not specify “left” or “right”.
Exercise 8.4.4 Let K be a field and let G, H be groups. Prove that K[G] ⊗K
K[H ] ∼
= K[G × H ] as K-algebras.
Exercise 8.4.6 Let K be a field and let A be a unital K-algebra. Show that, for all
n ∈ N, Mn (A) ∼
= Mn (K) ⊗K A.
Exercise 8.4.7 Let V and W be vector spaces over the field K. Their vector space
tensor product V ⊗ W is defined as the module tensor product V ⊗K W . Show that,
for every z ∈ V ⊗W , there exist n ∈ N, linearly independent vectors
v1 , . . . , vn ∈ V
and linearly independent vectors w1 , . . . , wn ∈ W such that z = ni=1 vi ⊗ wi .
Exercise 8.4.8 For a K-vector space V , let L(V ) denote the K-algebra of all linear
mappings from V to itself. Suppose V and W are K-vector spaces. Show that, for
each S ∈ L(V ), T ∈ L(W ), the function γS,T defined by V × W → V ⊗ W ,
γS,T (v, w) = Sv ⊗ T w for all v ∈ V , w ∈ W yields a linear mapping S,T : V ⊗
W → V ⊗ W such that S,T (v ⊗ w) = Sv ⊗ T w and from this, a bilinear mapping
γ : L(V ) × L(W ) −→ L(V ⊗ W ), γ (S, T ) = S,T S ∈ L(V ), T ∈ L(W )
is obtained. Use the universal property of the algebra tensor product to obtain a
unique homomorphism
such that ϕ(S ⊗ T ) = S,T for all S ∈ L(V ), T ∈ L(W ). Use the previous exercise
to show that ϕ is injective, and refer to Remark 8.1.5 to prove that ϕ is surjective
provided both V and W are finite dimensional.
Exercise 8.4.10 With the assumptions and notation as in Proposition 8.1.10 show
that f ⊗ g is surjective if both f and g are surjective. Does an analogous statement
hold for injectivity?
(f ◦ f ) ⊗ (g ◦ g) = (f ⊗ g ) ◦ (f ⊗ g).
Exercise 8.4.13 Writing Z2 for Z/2Z consider the exact sequence of Z-modules
f g
0 2 0
f id g id
0 2 2 2 2 0.
f g
0 R N1 RN2 RN3 0
Exercise 8.4.16 Suppose the functor G : D → C has a left adjoint. Show that G
preserves equalizers and hence kernels. (Compare Exercise 4.3.11.)
Definition 9.1 Let R be a unital ring. The module M ∈ R-Mod has the exchange
property if, for every module A ∈ R-Mod and every decomposition
A=M ⊕N = Ai (9.1.1)
i∈I
A=M ⊕ Ai . (9.1.2)
i∈I
If the above condition holds for every finite index set I , M is said to have the finite
exchange property.
© The Author(s), under exclusive license to Springer Nature Switzerland AG 2022 109
M. Mathieu, Classically Semisimple Rings,
https://doi.org/10.1007/978-3-031-14209-3_9
110 9 Exchange Modules and Exchange Rings
We first observe that, the submodules Ai ≤ Ai in Definition 9.1 above are in fact
direct summands of the Ai . This follows by applying Exercise 9.4.1 to the chain of
inclusions
Ai ⊆ Ai ⊆ Ai ⊕ M ⊕ Ai
j =i
which yields Ai = Ai ⊕ Ci with Ci = Ai ∩ M ⊕ j =i Ai for each i ∈ I .
For finitely generated modules, the finite exchange property already implies the
full exchange property.
Aj = M ⊕ Aj
j ∈J j ∈J
and consequently,
A= M ⊕N = Aj ⊕ Ai = M ⊕ Aj ⊕ Ai ,
j ∈J i∈I \J j ∈J i∈I \J
As it turns out it is in fact sufficient for a module M to have the finite exchange
property for two-element decompositions; see Theorem 9.1.5 below. Working
towards this statement, we establish the following results.
9.1 Basic Properties of Exchange Modules 111
Proposition 9.1.2 Let {Ai | i ∈ I } be a family in R-Mod and let Bi ≤ Ai for all
i ∈ I . Suppose A, M, N and L in R-Mod are such that
A=M ⊕N ⊕L= Ai ⊕ L,
i∈I
Then
A=M⊕ Bi ⊕ L.
i∈I
Proof The assumption on A/L provides us with A = M + i∈I Bi + L. In order
is a direct sum, take m ∈ M, x∈ L and finitely many bi ∈ Bi such
to show this
that m + i∈I bi + x = 0. Then m + L + i∈I (bi + L) = 0 in A/L wherefore,
by assumption, m ∈ L and each bi ∈ L. Since the direct sum decomposition of A
yields m = 0 and bi = 0 (as Bi ∩ L ⊆ Ai ∩ L), it follows that x = 0 which was to
prove.
We say that a module M ∈ R-Mod has the c-exchange property if the defining
property in Definition 9.1 holds for all index sets I of cardinality at most c.
A =M ⊕N ⊕L = Ai ⊕ L.
i∈I
A=M⊕ Bi ⊕ L.
i∈I
This result will now be used to show that the exchange property is well behaved
under direct sums.
n
Theorem 9.1.4 Let M = i=1 Mi be a finite direct sum of modules Mi ∈
R-Mod. Then M has the c-exchange property if and only if each Mi , 1 ≤ i ≤ n
has the c-exchange property.
112 9 Exchange Modules and Exchange Rings
Remark This result does not extend to infinite direct sums; an example can be found
in [13, Example 12.15].
Proof Clearly it suffices to deal with the case n = 2; thus assume that M = X ⊕ Y
for some X, Y ∈ R-Mod. Suppose that A = M ⊕ N = i∈I Ai , where I has
cardinality at most c. If Y has the c-exchange property then, for some Bi ≤ Ai ,
A=X⊕Y ⊕N =Y ⊕ Bi .
i∈I
If X has the c-exchange property too, then, by Corollary 9.1.3, A = X⊕Y ⊕ i∈I Ci
for some submodules Ci ≤ Bi . It follows that M has the c-exchange property.
Conversely suppose that M = X ⊕ Y has the c-exchange property. If
A=X ⊕N = Ai
i∈I
with X ∼
= X and |I | ≤ c then, setting B = A ⊕ Y , we have
B =M ⊕N =Y ⊕ Ai ,
i∈I
where M = X ⊕ Y ∼
= M. Fix j ∈ I . Since
B = M ⊕ N = Y ⊕ Aj ⊕ Ai
i=j
B =M ⊕C⊕ Bi .
i=j
B = X ⊕ Y ⊕ Bj ⊕ Bi = X ⊕ Y ⊕ Bi .
i=j i∈I
Theorem 9.1.5 A module which has the 2-exchange property has the finite
exchange property.
Proof Suppose M has the 2-exchange property; we shall show that M has the n-
exchange property for every n ∈ N. Let n > 2 and assume that M has the (n − 1)-
exchange property. Let N, A1 , . . . , An ∈ R-Mod be such that M ⊕ N = ni=1 Ai .
Setting A = ni=1 Ai and B = ni=2 Ai we have A = A1 ⊕ B. By the 2-exchange
property,
A = M ⊕ A1 ⊕ B ,
where
A1 = A1 ⊕ A1 , B = B ⊕ B and B = B ∩ (M ⊕ A1 ).
M ⊕ A1 = B ⊕ B where B = (M ⊕ A1 ) ∩ (B ⊕ A1 ).
As a result,
A = M ⊕ A1 ⊕ B = B ⊕ B ⊕ B
n
= B ⊕B = Ai ⊕ B ⊕ B
i=2
n
=M⊕ Ai
i=1
Corollary 9.1.6 Let M ∈ R-mod. If M has the 2-exchange property then it has
the exchange property.
Proof This follows immediately from Proposition 9.1.1 and Theorem 9.1.5.
Example 9.1.7 The module Z fails the 2-exchange property. Note at first that Z is
indecomposable as a module: if Z ∼ = X ⊕ Y with X, Y ∈ Z-Mod then X or Y must
be trivial. Now start with the identity
and suppose that Z had the 2-exchange property. Then there are submodules X ≤
Z(7, 3), Y ≤ Z(5, 2) such that Z ⊕ Z = Z(1, 0) ⊕ X ⊕ Y . It follows that Z ∼
= X ⊕Y
whence X = 0 or Y = 0. Therefore
which is impossible as neither {(1, 0), (7, 3)} nor {(1, 0), (5, 2)} form a Z-basis of
Z ⊕ Z.
Definition 9.2.1 A unital ring R is said to be an exchange ring if, for any pair of
elements a and b in R with a + b = 1, there exist idempotents e ∈ Ra and f ∈ Rb
such that e + f = 1.
In this section, we will show that R is an exchange ring if and only if the standard
left module R R has the exchange property and study some of the properties of
this type of ring. It appears that we have defined a property that depends on the
choice of the standard left module R R over its right-handed analogue RR ; however
Proposition 9.2.3 below shows that this is not the case.
We prepare this by the following lemma.
Lemma 9.2.2 The following conditions are equivalent for an element x in a unital
ring R.
(a) ∃ e = e2 ∈ R : e − x ∈ R(x − x 2 );
(b) ∃ e = e2 ∈ Rx, c ∈ R : (1 − e) − c(1 − x) ∈ rad(R);
9.2 Exchange Rings 115
1 − e − (1 − rx)(1 − x) = 0 ∈ rad(R).
Proposition 9.2.3 A unital ring R is an exchange ring if any, and hence every, of
the conditions in Lemma 9.2.2 holds for all x ∈ R. Moreover, this is equivalent to
the condition that, for any pair a and b in R with a + b = 1, there exist idempotents
e ∈ aR and f ∈ bR such that e + f = 1.
Proof Since it is evident that condition (d) in Lemma 9.2.2 is equivalent to the
condition that, for any pair of elements a and b in R with a + b = 1, there exist
idempotents e ∈ Ra and f ∈ Rb such that e + f = 1, the first assertion in the
proposition holds.
Now assume that the condition in the second assertion of the proposition is
satisfied; we show that this implies that R is an exchange ring, and the converse
will be true by symmetry. Take a, b ∈ R with a + b = 1 and pick e = e2 ∈ aR,
f = f 2 ∈ bR such that e + f = 1. Choose r ∈ Re, s ∈ Rf with the property
e = ar and f = bs. It follows that r = re = rar and rbs = rebs = ref = 0 and
similarly, s = sf s and sar = 0. Put
r = 1 − sb + rb and s = 1 − ra + sa.
ar = a(1 − sb) + arb = a(1 − sb) + (1 − bs)b = a(1 − bs) + b(1 − sb) = 1 − sb.
116 9 Exchange Modules and Exchange Rings
e + f = r a + s B = (1 − sb + rb)a + (1 − ra + sa)b
= a − sba + rba + b − rab + sab
= a + b + s(ab − ba) + r(ba − ab) = a + b = 1
Proof This follows immediately from the definition or any of the conditions listed
in Lemma 9.2.2.
Let L be a subgroup of (R, +). One says that idempotents lift modulo L if,
whenever x ∈ R satisfies x − x 2 ∈ L, there is an idempotent e ∈ R such that
e − x ∈ L. Proposition 9.2.3 entails the following characterisation of exchange
rings.
Corollary 9.2.5 A unital ring R is an exchange ring if and only if idempotents lift
modulo every left ideal of R.
We will now make contact with the exchange property, which will justify the
choice of the terminology ‘exchange ring’.
Theorem 9.2.6 Let R be a unital ring. Then M ∈ R-Mod has the finite exchange
property if and only if EndR (M) is an exchange ring. In particular, R is an
exchange ring if and only if R R has the exchange property.
Proof By Proposition 9.2.3, R is an exchange ring if and only if its opposite ring
R op is an exchange ring. Since R op ∼
= EndR (R R), see Exercise 2.7.1, the statement
on R follows immediately from the main statement in the theorem together with
Corollary 9.1.6.
Put S = EndR (M) and suppose S is an exchange ring. In order to deduce that
M has the finite exchange property, it suffices to show that M has the 2-exchange
property (Theorem 9.1.5). Assume, thus, that
A = M ⊕ N = A1 ⊕ A2 in R-Mod. (9.2.1)
A = (t1 A ⊕ t2 A) ⊕ A1 ⊕ A2 .
ai ti p = ai qi si = ai qi ai qi p = ai qi pai qi p = si2 = si
x = px = s1 x + s2 x = a1 t1 px + a2 t2 px = a1 t1 x + a2 t2 x = 0.
ti aj tj = ti paj qj tj = qi si sj aj qj tj = δij tj
x − a1 x1 − a2 x2 = x1 − a1 x1 + x2 − a2 x2 + y ∈ A1 ⊕ A2 .
x = a(a − b )x + b x = aa x + (1 − a)b x = aa x + bb x
which implies that e + f = 1. This proves that EndR (M) is an exchange ring, in
view of Proposition 9.2.3.
Remark The concept of an exchange ring was originally introduced by Warfield but
the equivalent formulation via Definition 9.2.1 is due to Nicholson; the proof of
Theorem 9.2.6 presented here is the original one in [28].
Corollary 9.2.7 A unital ring R is an exchange ring if and only if every finitely
generated projective module in R-Mod has the exchange property.
Proof Only the “only if”-part needs proof. By Theorem 9.2.6, R R has the exchange
property and therefore, every finitely generated free R-module has the exchange
property, by Theorem 9.1.4. As every finitely generated projective module is a direct
summand of a finitely generated free module, the same theorem yields the claim.
Theorem 9.2.8 A unital ring R is an exchange ring if and only if the quotient ring
R/rad(R) is an exchange ring and idempotents lift modulo rad(R).
Proof The “only if”-statement follows from Corollaries 9.2.4 and 9.2.5. Now
suppose that R = R/rad(R) is an exchange ring and let x ∈ R. Write x for
x + rad(R). There are an idempotent a ∈ R x and an element c ∈ R such that
1 − a = c (1 − x ), by condition (b) in Lemma 9.2.2. Without loss of generality we
can assume that a ∈ Rx. If idempotents lift modulo rad(R), there is f = f 2 ∈ R
such that f = a . Set u = 1 − f + a, which is a unit in R. It follows that
e = u−1 f u = u−1 f a is an idempotent in Rx. Moreover e = f = a and thus
(1−e)−c(1−x) ∈ rad(R) which proves that R is an exchange ring by Lemma 9.2.2.
9.3 Commutative Exchange Rings 119
Examples 9.2.9
1. Every von Neumann regular ring R is an exchange ring (cf. Exercise 7.4.13).
Let x ∈ R and take y ∈ R with xyx = x. Then f = yx is an idempotent
in Rx. Put e = f + (1 − f )xf to obtain an idempotent e ∈ Rx such that
1 − e = (1 − f )(1 − xf ) ∈ R(1 − x). The statement follows from Lemma 9.2.2.
2. Call a unital ring R clean if every element in R is the sum of a unit and an
idempotent. Every clean ring is an exchange ring. For, take x ∈ R and write it as
x = u + f where u is a unit and f is an idempotent. Set e = u−1 (1 − f )u. Then
u(e − x) = (1 − f )u − u(u + f ) = u − f u − uf − u2 = x − x 2
and thus, e − x ∈ R(x − x 2 ) so that Lemma 9.2.2 and Proposition 9.2.3 yield the
claim.
Remark 9.2.10 Exchange rings have rather interesting properties. For instance, it
was shown by Ara et al. [4] that a unital C* -algebra is an exchange ring if and
only if it has real rank zero, a rather important structural property useful for the
classification of C* -algebras. Ara [3] also introduced the concept of non-unital
exchange rings proving, amongst others, that a (possibly non-unital) ring R is an
exchange ring if and only if it contains an ideal I such that both I and R/I are
exchange rings and idempotents lift modulo I . For these, and further comments, see
[13, 11.43].
characterisation since it entails that the ring CR (X) is abundant with idempotents:
for each clopen subset U ⊆ X, its characteristic function χU belongs to CR (X) and
is an idempotent.
The following is a standard characterisation of zero-dimensional compact Haus-
dorff spaces; see, e.g., [19, Theorem 16.17].
(a) X is zero-dimensional;
(b) for every pair of disjoint closed subsets A and B of X, there exist disjoint open
subsets U and V of X such that X = U ∪ V , A ⊆ U and B ⊆ V ;
(c) for every closed subset A of X and every open subset U of X containing A,
there exists a clopen subset V of X such that A ⊆ V ⊆ U .
The exchange property for CR (X) turns out to be the equivalent of zero-
dimensionality of X.
Both Canfell dimensions of CR (X) agree with the covering dimension of the
compact Hausdorff space X, which was shown in [11] and [12], respectively.
Consequently, by Proposition 9.3.2, CR (X) is an exchange ring if and only if
dim1 CR (X) = 0 which is equivalent to dim2 CR (X) = 0. These results partly extend
to arbitrary commutative rings.
Proof Let a, b ∈ R be such that (a) = (b). Then there are s, t ∈ R with a = bs and
b = at so that b(1 − st) = 0. Since R is an exchange ring, there is an idempotent
e ∈ R such that e ∈ (1 − st)R and 1 − e ∈ stR. For some v, u ∈ R we thus have
e = (1 − st)v and 1 − e = stu. Replacing v by ve and u by u(1 − e) if necessary,
we can assume that v = ve and u = u(1 − e). Note that be = b(1 − st)v = 0 so
that b = b(1 − e).
Let k = s(1 − e) + (1 − st)e. Then k is invertible with inverse tu + v:
k(tu + v) = s(1 − e) + (1 − st)e tu(1 − e) + ve
= stu(1 − e) + (1 − st)ve = 1 − e + e = 1.
In addition,
Remark 9.3.6 Every integral domain has first Canfell dimension zero. This follows
immediately from a = bs and b = at, hence b(1 − st) = 0 so that st = 1 (unless
b = 0, the trivial case). However, not every integral domain is an exchange ring;
e.g., Z is not (Example 9.1.7).
a1 r1 + · · · + am rm = 1
(i) R is a Gelfand ring if, for every pair of elements a and b in R with a + b = 1,
there exist r, s ∈ R such that (1 + ar)(1 + bs) = 0.
(ii) R is a pm-ring if every prime ideal of R is contained in a unique maximal ideal.
(iii) R is a tb-ring if, for every pair of distinct maximal ideals of R, there is an
idempotent belonging to exactly one of them.
Definition 9.3.9 For a commutative unital ring R, let Spec(R) denote the set of all
prime ideals of R endowed with the following topology: a basis for the open subsets
is given by sets of the form
U (a) = {P ∈ Spec(R) | a ∈
/ P },
a ∈ R. This is the prime ideal space or spectrum of R and the topology is referred
to as the hull-kernel topology or Zariski topology.
The prime ideal property implies that U (ab) = U (a) ∩ U (b) for all a, b ∈ R.
It is customary to denote the complement of U (a) by V (a), that is, V (a) =
{P ∈ Spec(R) | a ∈ P }. In general, Spec(R) has poor separation properties;
its closed points are the maximal ideals in R. The subset Max(R) of all maximal
ideals is a subspace of Spec(R) when equipped with the subspace topology, and is
called the maximal spectrum of R. The basic subsets in Max(R) are thus U (a) =
U (a) ∩ Max(R), a ∈ R, and their complements will likewise be denoted by V (a).
Nevertheless both Spec(R) and Max(R) are compact spaces. More details on the
prime spectrum can be found in [6] or [10, Chap. II].
The next result provides us with the relations between the various types of rings
defined above.
The proof of Theorem 9.3.10 will be split up into a series of propositions which
emphasise more clearly the interconnections between the different conditions. We
start by examining pm -rings; clearly the acronym stems from the assumption that,
for each prime ideal, there is a unique maximal ideal containing it.
124 9 Exchange Modules and Exchange Rings
Proposition 9.3.11 A commutative unital ring R is a pm-ring if and only if, for
every pair of distinct maximal ideals M and M of R, there exist a ∈
/ M and b ∈
/M
such that ab = 0. In this case, Max(R) is a Hausdorff space.
S = {ab | a ∈
/ M, b ∈
/ M }.
If the semigroup S does not contain contain 0, a standard argument using Zorn’s
Lemma yields the existence of a prime ideal P with P ∩ S = ∅. Hence P ⊆ M ∩ M
and R is not a pm -ring.
In order to show that Max(R) is Hausdorff whenever R is a pm -ring let M, M ∈
Max(R), M = M . Take a ∈ / M and b ∈/ M such that ab = 0. Then M ∈ U (a),
M ∈ U (b) and U (a) ∩ U (b) = U (ab) = ∅ which proves the claim.
Corollary 9.3.13 Let R be a commutative exchange ring. Then the space Max(R)
is a compact Hausdorff space.
9.3 Commutative Exchange Rings 125
The terminology “tb -ring” stands for “topologically boolean ring” and is moti-
vated by the fact that a compact zero-dimensional Hausdorff space is also called a
boolean space and the result below.
Proof Note first that Max(R) is a compact Hausdorff space, by Propositions 9.3.11
and 9.3.12. It therefore suffices to show that the topology on Max(R) has a basis
consisting of clopen subsets. In fact, these will be the elements of E(R). By
Exercise 9.4.9, each U (e), e an idempotent in R, is clopen and, whenever a closed
subset K ⊆ Max(R) is covered by a family {U (ei ) | i ∈ I } of subsets from E(R),
by compactness, a finite subfamily suffices to cover K and, by Exercise 9.4.9, their
union belongs to E(R), so K is in fact contained in one subset of the form U (e).
Let K ⊆ Max(R) be closed and let M ∈ Max(R) be in the complement of K.
For each N ∈ K, there is an idempotent eN ∈ / N which belongs to M. Then N ∈
U (eN ), M ∈ U (1 −eN ) and U (eN )∩U (1 −eN ) = ∅. Since K ⊆ N∈K U (eN ),
the above argument yields an idempotent e ∈ R such that K ⊆ U (e) and M ∈ /
U (e). It follows that E(R) is a basis for the Zariski topology on Max(R).
Remark 9.3.15 The argument in the above proof in fact establishes the implication
(d) ⇒ (e) in Theorem 9.3.10. Let us now assume that condition (e) holds. This
entails that E(R) is the collection of all clopen subsets of Max(R). Take a ∈ R;
then V (a) is a closed subset of Max(R) and hence contained in some U (e), e = e2
by the above argument. Likewise V (1 − a) ⊆ U (1 − e) and hence condition (f) in
Theorem 9.3.10 holds.
Assuming (f) we will now show that R must be an exchange ring. Take a ∈ R. By
hypothesis, there is an idempotent e ∈ R such that V (a) ⊆ U (e) and V (1 − a) ⊆
U (1 − e). Let M be a maximal ideal of R. Suppose a ∈ M; then e ∈ / M and thus
e−a ∈ / M. Suppose a ∈ / M; if e − a ∈ M then a + M = e + M = 1 + M since
R/M is a field and e + M is a non-zero idempotent. It follows that 1 − a ∈ M so
1−e ∈ / M wherefore e ∈ M, which is impossible. As a result, e−a is invertible in R
and therefore, a = e − a + e can be written as a sum of a unit and an idempotent, in
other words, R is a clean ring. It follows from Example 9.2.9.2 that R is an exchange
ring.
Proof Suppose R is a Gelfand ring. Let M and M be two distinct maximal ideals
in R. For any m ∈ M \ M , M + (m) = R and thus, m + rm = 1 for some m ∈ M
and r ∈ R. By assumption, there are s, t ∈ R such that (1 + sm )(1 + trm) = 0.
Since 1 +sm ∈ / M and 1 +trm ∈ / M Proposition 9.3.11 entails that R is a pm -ring.
Now take a, b ∈ R with a + b = 1 and suppose that R fails to be a Gelfand ring.
Then the semigroup
does not contain 0. The usual argument invoking Zorn’s Lemma yields the existence
of a prime ideal P with P ∩ S = ∅. Since P + (a) = R (otherwise 1 = x − ar
with x ∈ P , r ∈ R would force x = 1 + ar ∈ S which is impossible) there is
a maximal ideal M ⊂ R containing P + (a). Similarly, there is a maximal ideal
M ⊂ R containing P + (b). As M = M (for a + b = 1), Proposition 9.3.11
implies that R is not a pm -ring.
To complete the proof of Theorem 9.3.10, we need to show that one of the
conditions (b) or (c) in the theorem imply that R is an exchange ring. The strategy
to achieve this is to construct two suitable orthogonal ideals of R whose sum is R.
These will arise as intersections of maximal ideals, and since the intersection of
maximal ideals is not a maximal but only a prime ideal, we will have to work within
the prime spectrum for the first time.
Proof We aim to show that E(R) is a basis for the Zariski topology of Max(R); as
seen in Remark 9.3.15 above, this implies that R is an exchange ring. First assume
that the nil radical nil(R) is zero. Let K ⊆ Max(R) be a clopen subset. Set
"
n "
m
K= U (ai ) and Spec(R) \ K = U (bj ).
i=1 j =1
9.3 Commutative Exchange Rings 127
Equivalently,
m
n
K= V (bj ) and Spec(R) \ K = V (ai ).
j =1 i=1
"
n "
m
Spec(R) = U (ai ) ∪ U (bj )
i=1 j =1
Lemma 9.3.18 Let R be a commutative unital ring. Then idempotents lift modulo
the nil radical nil(R).
and hence
9.4 Exercises
Exercise 9.4.2 Use Theorem 9.1.4 to prove that, for modules M, N, L, K and T ∈
R-Mod, if M = L ⊕ K = N ⊕ T and T has the finite exchange property and
N ≤ L, then K has the finite exchange property.
Exercise 9.4.6 Let R be an exchange n ring. Show by induction that, for every finite
i=1 ai = 1, there exist orthogonal idempotents
family a1 , . . . , an in R such that
ei ∈ Rai , 1 ≤ i ≤ n such that ni=1 ei = 1.
Exercise 9.4.7 Use Theorems 9.1.4 and 9.2.6 to show the following. Let e be an
idempotent in a unital ring R. Then R is an exchange ring if and only if eRe and
(1 − e)R(1 − e) are exchange rings.
Exercise 9.4.8 Show directly from the definition that every commutative exchange
ring is a Gelfand ring.
Exercise 9.4.9 For a commutative unital ring R, let E(R) = {U (e) | e = e2 ∈ R};
compare p. 123. Show that E(R) is closed under finite intersections and unions as
well as complements.
Exercise 9.4.10 Let R be a commutative unital ring. Show directly from the
definitions that, if R is a tb -ring, then R is an exchange ring.
Exercise 9.4.11 Let R be a commutative unital ring. Show directly from the
definitions that, if R is a clean ring, then R is a tb -ring.
Exercise 9.4.12 Use Proposition 5.2.4 to show that every Artinian commutative
ring is a pm -ring.
9.4 Exercises 129
The starting point for this chapter is the discussion around Maschke’s theorem,
Theorem 7.2.1, which provides us with conditions under which the group ring K[G]
is semisimple, provided G is a finite group. For any field K, the elements of G form
a basis of the K-vector space K[G] and if the ring K[G] is semisimple, then it is
necessarily Artinian, hence finite dimensional (Corollary 6.2.5 and Exercise 5.5.6).
As a result, we cannot expect K[G] to be semisimple for an infinite group G.
The ‘next best’ property one can expect for infinite groups therefore is semiprimi-
tivity, see also Proposition 7.3.7. An important and still open problem in ring theory,
which was popularised by Kaplansky, asks for characterisations of those groups for
which K[G] is semiprimitive. (This was known as the “semisimplicity problem”
before the terminology changed.) The larger part of this chapter will be devoted to
a study of some of the known results. In our approach, we will use techniques from
functional analysis (developed in detail in Sect. 10.2) to obtain a positive answer
when K = C, the field of complex numbers. This is a nice illustration of how
different areas of Mathematics can work together.
Throughout this chapter, K will denote a field and groups will be written
multiplicatively.
We begin by establishing some basic properties of group rings. For the definition of
the operations, see (ix) in Examples 1.2.
© The Author(s), under exclusive license to Springer Nature Switzerland AG 2022 131
M. Mathieu, Classically Semisimple Rings,
https://doi.org/10.1007/978-3-031-14209-3_10
132 10 Semiprimitivity of Group Rings
Corollary 10.1.2 Let G be an infinite cyclic group. Then K[G] is a principal ideal
domain.
Proof By the above proposition, we already know that K[G] is an integral domain.
Let I be a non-zero ideal of K[G]. Let g ∈ G be a generator. Then every element
of K[G] is of the form g k f (g) for some k ∈ Z and a polynomial f (g) ∈ K[G].
Suppose g k f (g) ∈ I ; then f (g) = g −k g k f (g) ∈ I . Take a non-zero polynomial
t (g) ∈ I of minimal degree and divide f (g) by t (g). By the division algorithm
and the minimality assumption we conclude that t (g) divides f (g) and thus t (g)
generates the ideal I . Consequently, K[G] is a principal ideal domain.
Lemma 10.1.4 Let R be a unital subring of the unital ring S. Suppose, as left R-
modules, R is a direct summand of S. Then rad(S) ∩ R ⊆ rad(R).
Proof Suppose that R S = R R⊕R M for some left R-module R M. Take α ∈ rad(S)∩
R and fix β ∈ R. Since there is s ∈ S such that (1 − αβ)s = 1 = s(1 − αβ) and s
can be written (uniquely) as s = r + m with r ∈ R, m ∈ M, we have
1 = (1 − αβ)r + (1 − αβ)m.
We turn our attention to fields with finite characteristic. Suppose K is a field with
characteristic p > 0 and let A be an algebra over K. If A is commutative, the power
map α → α p is a homomorphism, as is easily checked. In the general case, we have
the following generalisation. We shall denote by [A, A] the subspace of A spanned
by all commutators [α, β] = αβ − βα.
Lemma 10.1.5 Let K be a field with characteristic p > 0, and for given n ∈ N,
set q = pn . Let A be an algebra over K and let α1 , . . . , αm ∈ A. Then, for some
β ∈ [A, A], we have
q q
(α1 + · · · + αm )q = α1 + · · · + αm + β. (10.1.1)
q q
Proof We have (α1 + · · · + αm )q = α1 + · · · + αm + β, where β is the sum of
all words αi1 · · · αiq with at least two distinct indices. Let Cq be the cyclic group
of order q. Suppose ω1 = αi1 · · · αiq and ω2 = αij · · · αiq αi1 · · · αij−1 are cyclic
permutations of each other. Then,
ω1 − ω2 = (αi1 · · · αij−1 )(αij · · · αiq ) − (αij · · · αiq )(αi1 · · · αij−1 ) ∈ [A, A].
Hence, modulo [A, A], all cyclic permutations of a word are equal and the number
of formally distinct permutations of a word which occur in β as above is the size of
a non-trivial orbit in Cq , therefore divisible by p. This proves the assertion.
Definition 10.2.1 Let α = g∈G ag g, β= g∈G bg g belong to C[G]. We put
(α | β) = ag b̄g
g∈G
(where b̄ denotes the complex conjugate of b ∈ C) and call this the canonical inner
product on C[G].
It will be shown in Exercise 10.4.1 that the above indeed defines an inner product
on C[G] and that α2 = (α | α)1/2 therefore is a norm on C[G]. However, the
associated metric space is in general not complete. In other words, C[G], (· | ·) is
in general not a Hilbert space.
Beside the structure of an inner product space, C[G] also has the structure of a
normed algebra.
Definition 10.2.2 Let α = g∈G ag g ∈ C[G]. We put α1 = g∈G |ag |, where
|a| of course denotes the absolute value of the complex number a.
Proposition 10.2.3 The above definition of α1 turns C[G] into a normed space.
Moreover, for all α, β ∈ C[G], we have αβ1 ≤ α1 β1 and C[G] is a complex
normed algebra with identity 1.
As a result, C[G], · 1 is a normed algebra. Clearly, 1 is a multiplicative identity.
As all good things come in threes, we will introduce another norm on C[G].
Let H = 2 (G) be the Hilbert space obtained from C[G], (· | ·) , compare
Exercise 10.4.1. The algebra C[G] acts on H in a canonical way: for clarity
let us
denote the elements in H by ξ so that ξ ∈ C[G] can be written as ξ = g∈G xg g
with xg ∈ C. For such ξ and α ∈ C[G] we have αξ ∈ C[G] ⊆ H , that is, C[G]
acts as left multiplication operator. The above inequalities imply that this action is
continuous, so can be extended to all of H . Moreover, since αξ 2 ≤ α2 ξ 2 for
all ξ ∈ H , the action is injective (α1 = 0 ⇒ α = 0) and it makes sense to define
our third norm as
$ %
αop = sup αξ 2 | ξ ∈ H, ξ 2 = 1 .
10.2 Some Analytic Structure on C[G] 135
(αγ ∗ | β) = (α | βγ ) = (β ∗ α | γ ). (10.2)
(αγ ∗ | β) = tr αγ ∗ β ∗ = tr α(βγ )∗ = (α | βγ )
= tr (βγ )∗ α = tr γ ∗ β ∗ α = (β ∗ α | γ ),
so that tr is a bounded linear functional and thus can be extended from the
dense subspace C[G] to the whole algebra Cr∗ (G). We shall denote this extended
functional by τ .
Lemma 10.2.5 The functional τ is a normalised faithful trace on Cr∗ (G), that is,
τ (1) = 1, τ (ab) = τ (ba) for all a, b ∈ Cr∗ (G) and τ (a ∗ a) ≥ 0 with τ (a ∗ a) = 0 if
and only if a = 0.
τ (a ∗ a) = tr α ∗ α − τ (α ∗ α − a ∗ a) ≥ tr α ∗ α − a ∗ a − α ∗ αop > tr α ∗ α − ε ≥ 0.
The next lemma is a useful general result in C* -algebras and can be found in
many of the standard textbooks. We provide a proof for the convenience of the
reader.
that is, e and a commute and hence e and a −1 commute. We put p = ee∗ a −1 and
observe that p∗ = a −1 ee∗ = p (since a is selfadjoint, so is its inverse). Furthermore,
Theorem 10.2.7 For every group G, the group ring C[G] is von Neumann finite.
This theorem in fact holds for all groups rings K[G] where K has characteristic
zero. A result by Kaplansky from 1969, it can be derived from the above special
case, or by other means, see, e.g., [31, Corollary 2.1.9].
A primitive ring is one that has a faithful simple module; according to the convention
in this book, we focus on the left-handed version (compare Definition 7.3.4). As
primitive rings are somewhat scarce, one wants to study more general rings by
looking at all their simple modules (equivalently, irreducible representations) but for
this to work, the canonical obstruction, the Jacobson radical, has to vanish. These
are the semiprimitve rings (Definition 7.3.4).
When is the group ring K[G] semiprimitive? The answer to this question will
depend on properties of the field K and the group G and possibly on the interaction
of the two. We will discuss this question for infinite groups in this section.
The first case we investigate is the group ring with complex numbers as
coefficients, this will lead us to one of the first results on the above question,
originally obtained by Rickart in 1950. We present a modified version of his
original proof which uses functional analytic techniques which we already prepared
ourselves for in the previous section.
Theorem 10.3.1 For every group G, the group ring C[G] is semiprimitive.
hand, we will show that the only α ∈ rad(C[G]) that allows for the construction of
the above-mentioned F is α = 0, which will complete the argument that C[G] is
semiprimitive.
Let us prove the last assertion first. Take β ∈ rad(C[G]) \ {0} and put α =
ββ ∗ ∗
β2 2 , where we use the notation of Sect. 10.2. Then α = α ∈ rad(C[G]) since the
138 10 Semiprimitivity of Group Rings
tr ββ ∗ β22
tr α = = = 1.
β22 β22
m
Suppose tr α 2 ≥ 1 for some m ∈ N0 . Then
= tr α 2 (α 2 )∗ = α 2 22 ≥ |tr α 2 |2 ≥ 1;
m+1 m m m m
tr α 2
m
thus, by induction, tr α 2 ≥ 1 for all m ∈ N0 and therefore the sequence (tr α n )n∈N
alluded to above cannot tend to zero.
We shall now go about the construction of the function F . Since, for every z ∈ C,
1 − zα is invertible in C[G] (Proposition 7.3.2), we can define φ : C → C[G] by
φ(z) = (1 − zα)−1 . Suppose α = 0 and that |z| < α 1
. In this case the series
∞ −1
1
F (z) − F (w)
= tr (α φ(z)φ(w)),
z−w
F (z) − F (w)
lim = tr (α φ(z)2 ).
w→z z−w
There do exist purely algebraic proofs of Theorem 10.3.1 but it is instructive to see
how Analysis can be used to obtain an algebraic result.
10.3 The Semiprimitivity Problem 139
From Lemma 7.3.8 we know that rad(K[G]) contains every nil ideal of K[G].
This is a starting point to obtain another case where K[G] is semiprimitive.
Lemma 10.3.2 In every group ring, each algebraic element in the radical is
nilpotent.
α n (1 + c1 α + c2 α 2 + · · · + cr α r ) = 0
Lemma 10.3.3 Let G be a group and let K be a field with the property that |K| >
dimK K[G] (as a strict inequality of possibly infinite cardinal numbers). Then every
element in rad(K[G]) is algebraic.
From the two results above we see that, for very large coefficient fields, every
element in the radical is nilpotent; hence the radical itself is a nil ideal.
Thus the remaining task is to determine when K[G] has no non-zero nil ideals.
i
p-supp α = {g ∈ G | ag = 0 and g p = 1 for some i ∈ N}
& '
for some β ∈ K[G], K[G] . Since the trace vanishes on commutators this entails
that
q q
0 = tr α q = ag = ag .
g q =1 g q =1
g∈p-supp α ag = 0 as claimed.
Let p be a prime. A group G is called a p -group if, for all g ∈ G, g = 1 and all
i
i ∈ N, g p = 1. We now combine the last two theorems to obtain an answer to the
semiprimitivity problem for certain fields and certain groups.
Theorem 10.3.6 Let K be an uncountable field with characteristic p > 0 and let
G be a p -group. Then K[G] is semiprimitive.
Remark 10.3.7 The above result, due to Amitsur (1957), also holds for fields with
characteristic 0 and arbitrary groups. To this end, one has to extend the previous two
theorems to this setting and add quite a bit of additional terminology and techniques.
Since these would take us here too far afield, we refer to [31, Chap. 7] for the details.
Remark 10.3.8 The three famous ‘Kaplansky conjectures’ on the structure of group
rings are as follows. Let K be a field and G be a torsion-free group. Then
(U) K[G] has no non-trivial units, that is, every element in K[G]× is of the form
ag with a ∈ K ∗ and g ∈ G.
(Z) K[G] has no non-trivial zero divisors, that is, every zero divisor is equal to 0.
(I) K[G] has no non-trivial idempotents, that is, every idempotent in K[G] is either
equal to 1 or to 0.
The second implication is easy: every idempotent e ∈ K[G] \ {0, 1} yields non-
trivial zero divisors via e(1 − e) = e − e2 = 0. The first is a bit harder and
uses Connell’s result ([31, Theorem 4.2.10]) that K[G] is a prime ring under
the given assumptions. They were open for many decades and proven in a large
number of special cases. Only in 2021, conjecture (U) was settled in the negative by
Gardam [18].
Since these questions are related to other important problems, for example, the
Baum–Connes conjecture in operator algebras, there is still quite some activity in
the study of group rings.
10.4 Exercises
Exercise 10.4.2 Prove that the normed algebra C[G] as described in Proposi-
tion 10.2.3 is in general not complete. Its completion to a Banach algebra is
generally denoted by 1 (G) and called an “L1 -group algebra”. (You find the
definition of 1 (G) in any of the above-cited books.)
Exercise 10.4.3 LetA be anunital Banach algebra. Show that, for every b ∈ A with
b < 1, the series ∞n=0 b converges and has limit (1−b) −1 . (This series is called
the Neumann series after Carl Neumann. See, e.g., [1, Sect. 4.4].)
Show that this indeed defines a norm on the complex vector space B(H ) of all
bounded linear mappings on H . Show also that, if S, T ∈ B(H ), then ST op ≤
Sop T op and that T op = T ∗ op and T ∗ T op = T 2op , where the adjoint
T ∗ of T is given by the formula (T ξ | η) = (ξ | T ∗ η) for all ξ, η ∈ H . Also prove
that B(H ) with this norm is complete. (If you need help with this exercise, look into
[1, Sect. 2.14], [26, Chap. 3, Sect. 9] or [32, Sect. 3.2].)
Exercise 10.4.7 For a group G, let F(G) denote the set of all its finitely generated
subgroups. Let α ∈ K[G], for some field K. Show that α ∈ rad(K[G]) if and only
if α ∈ rad(K[H ]) for all H ∈ F(G).
Exercise 10.4.8 Use Proposition 10.1.3 together with the previous exercise to show
that, for a group G, K[G] is semiprimitive if K[H ] is semiprimitive for all H ∈
F(G), and that rad(K[G]) is a nil ideal if rad(K[H ]) is nil for all H ∈ F(G).
Exercise 10.4.9 Let R be a non-zero ring and let G be an infinite group. Define the
augmentation map : R[G] → R by (r) = r for all r ∈ R and (g) = 1 for all
10.4 Exercises 143
1. Allan, G.R.: Introduction to Banach Spaces and Algebras. Oxford Graduate Texts in
Mathematics, vol. 20. Oxford University Press, Oxford (2011)
2. Anderson, F.W., Fuller, K.R.: Rings and Categories of Modules. Graduate Texts in Mathe-
matics, vol. 13. Springer-Verlag, New York (1992)
3. Ara, P.: Extensions of exchange rings. J. Algebra 197, 409–423 (1997)
4. Ara, P., Goodearl, K.R., O’Meara, K.C., Pardo, E.: Separative cancellation for projective
modules over exchange rings. Isr. J. Math. 105, 105–137 (1998)
5. Artin, E.: The influence of J. H. M. Wedderburn on the development of modern algebra. Bull.
Am. Math. Soc. 56, 65–72 (1950)
6. Atiyah, M.F., Macdonald, I.G.: Introduction to Commutative Algebra. Addison-Wesley,
London (1969)
7. Beachy, J.A.: Introductory lectures on rings and modules. London Mathematical Society
Student Texts, vol. 47. Cambridge University Press, Cambridge (1999)
8. Blyth, T.S.: Module Theory: An Approach to Linear Algebra. Clarendon Press, Oxford (1990)
9. Blyth, T.S., Robinson, E.F.: Basic Linear Algebra. Springer Undergraduate Mathematics
Series. Springer-Verlag, London (1998)
10. Bourbaki, N.: Elements of Mathematics. Commutative Algebra. Springer-Verlag, Berlin
(1989)
11. Canfell, M.J.: Uniqueness of generators of principal ideals in rings of continuous functions.
Proc. Am. Math. Soc. 26, 571–573 (1970)
12. Canfell, M.J.: An algebraic characterization of dimension. Proc. Am. Math. Soc. 32, 619–620
(1972)
13. Clark, J., Lomp, C., Vanaja, N., Wisbauer, R.: Lifting Modules; Supplements and Projectivity
in Module Theory. Frontiers in Mathematics. Birkhäuser Verlag, Basel (2006)
14. Conway, J.B.: A Course in Functional Analysis. Springer-Verlag, New York (1985)
15. Eilenberg, S., MacLane, S.: General theory of natural equivalences. Trans. Am. Math. Soc.
58, 231–294 (1945)
16. Facchini, A.: Module Theory: Endomorphism Rings and Direct Sum Decompositions in
Some Classes of Modules. Progress in Mathematics, vol. 167. Birkhäuser Verlag, Basel
(1998)
17. Freyd, P.J.: Abelian Categories. An Introduction to the Theory of Functors. Harper & Row,
New York (1964)
18. Gardam, G.: A counterexample to the unit conjecture for group rings. Ann. Math. 194, 967–
979 (2021)
19. Gillman, L., Jerison, M.: Rings of continuous functions. Graduate Texts in Mathematics, vol.
43. Springer-Verlag, New York (1976)
20. Golan, J.S.: The Linear Algebra a Beginning Graduate Student Ought to Know. Springer-
Verlag, Dordrecht (2007)
© The Author(s), under exclusive license to Springer Nature Switzerland AG 2022 145
M. Mathieu, Classically Semisimple Rings,
https://doi.org/10.1007/978-3-031-14209-3
146 Bibliography
21. Hilton, P.J., Stammbach, U.: A Course in Homological Algebra. Graduate Texts in Mathe-
matics, vol. 4. Springer-Verlag, New York (1971)
22. Jacobson, N.: Lectures in Abstract Algebra, vol. II. D. Van Nostrand, Toronto (1953)
23. Knapp, A.W.: Basic Algebra. Birkhäuser, Boston (2006)
24. Lam, T.Y.: A first course in noncommutative rings. Graduate Texts in Mathematics, vol. 131.
Springer-Verlag, New York (1999)
25. MacLane, S.: Categories for the Working Mathematician. Graduate Texts in Mathematics,
2nd. edn., vol. 5. Springer-Verlag, New York (1978)
26. Mathieu, M.: Funktionalanalysis. Ein Arbeitsbuch. Spektrum Akademischer Verlag,
Heidelberg-Berlin (1998)
27. McGovern, W.Wm.: Neat rings. J. Pure Appl. Algebra 205, 243–265 (2006)
28. Nicholson, W.K.: Lifting idempotents and exchange rings. Trans. Am. Math. Soc. 229, 269–
278 (1977)
29. Osborne, M.S.: Basic Homological Algebra. Graduate Texts in Mathematics„ vol. 196.
Springer-Verlag, New York (2000)
30. Pareigis, B.: Categories and Functors. Academic Press, New York/London (1970)
31. Passman, D.S.: The Algebraic Structure of Group Rings. John Wiley & Sons, New York
(1977)
32. Pedersen, G.K.: Analysis Now. Graduate Texts in Mathematics, vol. 118. Springer-Verlag,
New York (1995)
33. Taylor, M.E.: Linear Algebra. Pure and Applied Undergraduate Texts, vol. 45. American
Mathematical Society, Rhode Island (2020)
34. Wallace, D.A.R.: Groups, Rings and Fields. Springer Undergraduate Mathematics Series.
Springer-Verlag, London (1998)
Index of Symbols
Symbols ob(C), 9
B(H ), 142 MorC (A, B), 10
C(X), 28 1 (G), 141
K-algebra, 101 1, 68
L(V ), 105 2 (G), 141
Mn (R), 2 AGr, 10
R[x], 2 Ban1 , 10
Tn (R), 2 Ban∞ , 10
Z(R), 23 Cop , 12
c0 , 68 Comp, 10
p -group, 140 CMetric, 107
R[G], 3 C, 9
R R, 6 Gr, 10
HomR (MR , NR ), 5 HTop, 10
HomR-S (R MS , R NS ), 5 Lat, 10
End(G), 2 Metric, 107
GLn (K), 7 Mod-R, 4, 10
Max(R), 123 Ring1 , 10
Spec(R), 123 Ring, 10
AC1∗ , 31 S, 10
R M i , 18 Top, 10
i∈I soc(R), 89
codom(f ), 10 A ⊗K B, 101
cok(f ), 41 B(M × N, G), 93
dim1 R, 121 CR (X), 119
dim2 R, 121 Cr∗ (G), 135
dom(f ), 10 (A), 31
C, 2 E(R), 123
H, 2 M ⊗R N, 94
Q, 2 R × , 13
R, 2 R-Mod-S, 4, 10
Z, 2 R-Mod, 4, 10, 11
HomR (R M, R N), 5 x ⊗ y, 94
im(f ), 12 p-support, 140
ker(f ), 12, 41 AnnR (S), 16
p-supp
α, 140
R S, 17
R M i , 17 πN , 16
i∈I
mor(C), 9 tor(R M), 57
RX, 19
© The Author(s), under exclusive license to Springer Nature Switzerland AG 2022 147
M. Mathieu, Classically Semisimple Rings,
https://doi.org/10.1007/978-3-031-14209-3
Index
A D
Algebra Descending chain condition (DCC), 48
normed, 134 Dimension
over a field, 91 first Canfell dimension, 121
Annihilator, 16 second Canfell dimension, 121
Artin, Emil, 47 zero-dimensional space, 119
Artin–Wedderburn Theorem, 75 Direct limit
Ascending chain condition (ACC), 35 of modules, 91
Direct product
of modules, 17
B Direct sum
Bimodule, 4 of modules, 18
Dorroh superring, 13
C
C*-algebra, 135, 142 E
commutative, 31 Endomorphism
reduced group, 135 of a group, 2
Category, 9 Epimorphism, 12, 15
abelian, 54 canonical epimorphism, 16
additive, 23 Equalizer
arrow in a, 9 of two morphisms, 44
codomain of a morphism, 10 Evaluation map, 30
concrete, 10, 27 Exact sequence, 41
domain of a morphism, 10 of modules, 32
dual category, 12 short exact sequence, 32, 42
exact, 42 split, 63
final object, 23
full subcategory, 11
initial object, 23 F
morphism in a, 9 Free object
object in a, 9 in a category, 21
reflective subcategory, 107 Functor, 27
small category, 10, 11 additive, 33
subcategory, 11 constant, 27
zero object, 23 contravariant, 27
Cohomology group, 70 covariant, 27
Composition series, 88 exact, 42
Coproduct faithful, 27
in a category, 22 forgetful, 27
© The Author(s), under exclusive license to Springer Nature Switzerland AG 2022 149
M. Mathieu, Classically Semisimple Rings,
https://doi.org/10.1007/978-3-031-14209-3
150 Index