A First Course in Category Theory (Agore)
A First Course in Category Theory (Agore)
Ana Agore
A First
Course in
Category
Theory
Universitext
Series Editors
Nathanaël Berestycki, Universität Wien, Vienna, Austria
Carles Casacuberta, Universitat de Barcelona, Barcelona, Spain
John Greenlees, University of Warwick, Coventry, UK
Angus MacIntyre, Queen Mary University of London, London, UK
Claude Sabbah, École Polytechnique, CNRS, Université Paris-Saclay, Palaiseau,
France
Endre Süli, University of Oxford, Oxford, UK
Universitext is a series of textbooks that presents material from a wide variety
of mathematical disciplines at master’s level and beyond. The books, often well
class-tested by their author, may have an informal, personal, or even experimental
approach to their subject matter. Some of the most successful and established books
in the series have evolved through several editions, always following the evolution
of teaching curricula, into very polished texts.
Thus as research topics trickle down into graduate-level teaching, first textbooks
written for new, cutting-edge courses may find their way into Universitext.
Ana Agore
This work was supported by Fonds Wetenschappelijk Onderzoek, Belgium and Romanian Ministry of
Education and Research, CNCS/CCCDI-UEFISCDI.
© The Editor(s) (if applicable) and The Author(s), under exclusive license to Springer Nature Switzerland
AG 2023
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
Categories were first considered in 1945 in a paper by S. Eilenberg and S. Mac Lane
[21] with the purpose of formalizing the concept of “natural transformation”, which
was informally used at that time in many papers from various fields, especially in
algebraic topology. The initial theory introduced in [21] developed rapidly, allowing
for several new mathematical disciplines to arise, as was the case, for example, with
homological algebra. Category theory is based on the idea that many mathematical
properties can be described using diagrams of arrows of different types. Working
in this very general setting allows for a better understanding of the common
constructions and patterns in mathematics and leads to a unified treatment of similar
concepts across different mathematical structures. An early and notable example can
be found in [19], where group cohomology, Lie algebra cohomology and associative
algebra cohomology are recast as derived functors in a suitable module category.
Over the years, category theory has become a universal language allowing
mathematicians to achieve important advancements by exchanging ideas and tech-
niques between seemingly unrelated domains. Using very abstract definitions
that capture the idea behind a certain concept in universal terms rather than its
isolated properties, purely categorical techniques have found their way into most
mathematical areas.
Nowadays category theory is an indispensable tool for doing research not only in
various areas of pure mathematics such as algebraic topology, homological algebra,
algebraic geometry and functional analysis, but also in theoretical computer science
(e.g., the development of algorithms, automata theory), physics (e.g., electrical
circuits), chemistry (e.g., chemical interactions), biology (e.g., biological systems)
and medicine (e.g., genetics). We refer to [52] for an approach to category theory
in the spirit of the applied sciences and to [29] for applications to the cognitive
sciences.
The purpose of this book is to provide students with no prior exposure to
categorical reasoning with an accessible source from which to learn the basic
material. The fundamentals of category theory are clearly and thoroughly covered
with the aim of leaving the reader able to confidently use categorical techniques as
well as to easily explore and understand more advanced topics.
vii
viii Preface
The book is based on my lecture notes from the graduate course on category
theory that I have taught at Vrije Universiteit Brussel. Additional fully worked
examples and complete proofs have been added with the purpose of making the
material suitable for self-study. Although the reader is expected to be at the advanced
undergraduate level, some background and full references are provided throughout
the book. The prerequisites include familiarity with group theory, rings, modules
and topological spaces, as well as a basic understanding of set theory. As opposed
to the standard category theory monographs by S. Mac Lane [35] and F. Borceux
[8–10], and the more recent ones [5, 34, 47, 48], which are more encyclopedic
in nature and oriented toward researchers rather than students, the present book
serves as a first introduction to the field. The excellent monographs [1, 2, 5, 8–
10, 12, 13, 22, 23, 25, 30, 34, 35, 46–48] have been used when preparing these notes
and have influenced the approach and the development of certain topics.
The first chapter introduces the fundamental concepts needed in the sequel.
Important notions such as (sub)categories, functors, natural transformations, repre-
sentable functors, which form the backbone of category theory, are well illustrated
by many familiar examples. A concise description of the duality principle, a crucial
reasoning process in category theory, is also presented. The first important result we
present is Yoneda’s lemma, which allows us to embed any (locally small) category
into a category of functors on that category. This generalizes the well-known group
theory result called Cayley’s theorem, stating that any group is isomorphic to a
subgroup of a symmetric group.
The second chapter treats the general theory of limits and colimits. Both are
very general concepts which arise in various forms in all fields of mathematics.
We introduce them gradually, starting with some special cases which might be
familiar to the reader such as: (co)products, (co)equalizers, pullbacks and pushouts.
A variety of detailed examples are included to illustrate the newly introduced
concepts. (Co)products and (co)equalizers are not only important special cases of
(co)limits but also generic in the sense that all (co)limits can be constructed out of
these two special cases. Certain types of functors are considered in connection to the
existence of (co)limits. The existence of (co)limits in several important categories
such as functor categories or comma categories is investigated in detail as well.
The third chapter deals with one of the most important notions in category
theory: adjoint functors. Several descriptions of adjoint functors are presented
and the theory is illustrated by a wide range of examples from various areas of
mathematics. Many important constructions in mathematics are shown to be part
of an adjunction, including for instance the classical free constructions present in
algebra, localizations in ring theory or Stone–Čech compactifications of topological
spaces. Important related concepts such as equivalence of categories, (co)reflective
subcategories or localization of categories are also investigated and well illustrated
by a plethora of detailed examples. Deeper connections with the concepts introduced
in the previous chapters are emphasized. For instance, (co)limits and representable
functors are equivalently described by means of adjoint functors. Going beyond
what is usually covered by an introductory text in category theory, the book ends
with a more advanced topic, the adjoint functor theorem. More precisely, two
Preface ix
variations of this celebrated theorem, namely Freyd’s Adjoint Functor Theorem and
the Special Adjoint Functor Theorem, are considered. They provide different kinds
of necessary and sufficient conditions for a functor to admit a left or a right adjoint.
I would like to take this opportunity to thank my coauthors as well as my
colleagues and students from Vrije Universiteit Brussel for everything I have learned
from them. My warmest thanks to Gigel Militaru for teaching me category theory
when I was a student and for the many wise suggestions he made after reading
a first draft of this book as well as to Alexandru Chirvasitu for the countless
illuminating discussions. I am very grateful to the Springer editors, especially
to Rémi Lodh, and to the anonymous referees for their comments and advice
which greatly improved the book’s presentation. During the preparation of this
manuscript, my work was supported at different stages by FWO (Fonds voor
Wetenschappelijk Onderzoek—Flanders) and a grant of Ministry of Research,
Innovation and Digitization, CNCS/CCCDI–UEFISCDI, project number PN-III-P4-
ID-PCE-2020-0458.
xi
xii Contents
References . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 277
Index . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 281
Frequently Used Notations
.∅ empty set
.N set of natural numbers
.Z ring of integers
.Q field of rational numbers
.R field of real numbers
.A\B set-theoretic difference
.ℵ0 cardinality of .N (and .Z)
.|X| cardinality of the set X
.⊆ inclusion
.f|X restriction of a map f to a subset X of its domain
.⊗ tensor product
.H G H is a normal subgroup of G
.Gab abelianization of the group G
.G(M) universal enveloping group of the monoid M
.kerf kernel of the morphism f
.Imf image of the morphism f
.S
−1 X localization of the ring (module) X by the multiplicative set S
.CX discrete category on the set X
PO(.X, ) category associated to the pre-ordered set .(X, )
Set category of sets
FinSet category of finite sets
Grp category of groups
SiGrp category of simple groups
Ab category of abelian groups
Div category of divisible groups
Rng category of rings
Ring category of unitary rings
c
.Ring category of commutative unitary rings
Field category of fields
R.M category of left R-modules
.MR category of right R-modules
xiii
xiv Frequently Used Notations
We start by setting very briefly the set theory model that will be assumed to hold
throughout. The main issue that arises is that most categories of interest have as
objects all sets, all groups, all topological spaces, etc. Therefore, a proper definition
of a category which includes the examples mentioned above is not possible in the
classical Zermelo–Fraenkel set theory. One way to get around this issue is by using
the von Neumann–Bernays–Gödel (NBG) set theory which introduces, in addition
to sets, the notion of a class to play the role of these “big sets” consisting of all sets,
all groups, etc. More precisely, the connection between sets and classes is given by
the so-called limitation of size axiom:
A class is a set if and only if it is not bijective with the class of all sets.
To conclude, we can use the word class to designate any collection of mathematical
objects; all sets are obviously classes.
The NBG axioms are in fact a conservative extension of the ZFC axioms.
Therefore, all statements about sets which can be proved in NBG hold in ZFC as
well. An important axiom included in NBG which will be used in many places in
the sequel is the following form of the axiom of choice:
We can choose an element from each of any class of nonempty sets.
Moreover, the following consequence of the NBG axioms will be intensively used
throughout: if A is a set and B is a subclass of A then B is a set. A detailed account
of the NBG set theory axioms can be found, for instance, in [51] or in the Appendix
of [46].
We start by providing the definition of a category and many examples which may
already be familiar to the reader.
Definition 1.2.1 A category .C consists of the following data:
(1) a class Ob .C whose elements A, B, .C, . . . are called objects;
(2) for every pair of objects A, B, a (possibly empty) set .HomC (A, B), whose
elements are called morphisms from A to B. An element .f ∈ HomC (A, B) will
be denoted by .f : A → B; A and B are called the domain and the codomain of
f , respectively;
(3) for every triple of objects A, B, C, a composition law
. (f, g) → g ◦ f ;
(4) for every object A, a morphism .1A ∈ HomC (A, A), called the identity on A
such that the following axioms hold:
(i) associative law: given morphisms .f ∈ HomC (A, B), .g ∈ HomC (B, C),
.h ∈ HomC (C, D) the following equality holds:
h ◦ (g ◦ f ) = (h ◦ g) ◦ f ;
.
(ii) identity law: given a morphism .f ∈ HomC (A, B) the following identities hold:
1B ◦ f = f = f ◦ 1A .
.
Examples 1.2.2
(1) Any set X can be made into a small category, called the discrete category on
X and denoted by .CX , as follows:
Ob CX = X
.
∅ if x = y
HomCX (x, y) = , for every x, y ∈ X.
{1x } if x = y
1.2 Categories: Definition and First Examples 3
Ob PO(X, ) = X
.
∅ if x y
HomPO(X, ) (x, y) = , for every x, y ∈ X,
{ux, y } if x y
Moreover, following the same idea, we can further generalize this example
to the level of classes. For instance, consider the class of all sets pre-ordered
by inclusion; we obtain a category denoted by .Set(⊆) which has the class of
all sets as objects and for all sets A, B we have
∅ if A B
HomSet(⊆) (A, B) =
. ,
{uA, B } if A ⊆ B
1A set X is called pre-ordered if is endowed with a binary relation . which is reflexive and
transitive.
2 Recall that the axiom of choice is assumed to hold.
4 1 Categories and Functors
Ob RelSet = Ob Set
.
Finally, the identity is defined as .1A = {(a, a) | a ∈ A}. RelSet is called the
category of relations.
(7) Grp is the category of groups, where Ob Grp is the class of all groups while
.HomGrp (A, B) is the set of all group homomorphisms from A to B. Similarly,
3 A non-trivial group is called simple if its only normal subgroups are the trivial group and the
group itself.
4 An abelian group .(G, +) is called divisible if for every positive integer n and every .g ∈ G, there
(A, B) → BA.
Remark 1.2.3 Notice that although we sometimes work with categories whose
objects are sets, morphisms in the sense of Definition 1.2.1 need not be functions.
This situation is best illustrated in Example 1.2.2, (6).
6S is called a multiplicative subset of the ring R if .1R ∈ S and for all s, .s ∈ S we have .ss ∈ S.
7 Given two pre-ordered sets .(X, X ) and .(Y, Y ), a map .f : X → Y is called order preserving
if .x X y implies .f (x) Y f (y).
8 A set X is called partially ordered if is endowed with a binary relation . which is reflexive,
Definition 1.2.4 Let .C, .C be two categories. We shall say that .C is a subcategory
of .C if the following conditions are satisfied:
(i) Ob .C ⊆ Ob .C, i.e., any object of .C is an object of .C;
(ii) .HomC (A, B) ⊆ HomC (A, B) for every A, .B ∈ Ob C ;
(iii) the composition of morphisms in .C is induced by the composition of
morphisms in .C;
(iv) the identity morphisms in .C are identity morphisms in .C.
Moreover, .C is said to be a full subcategory of .C if for every pair .(A, B) of objects
of .C we have
Examples 1.3.2
(1) In each of the categories Set, Grp, Ab, .R M monomorphisms coincide with
the injective homomorphisms, while in Top and KHaus monomorphisms
coincide with the injective continuous maps. We will only prove here that
monomorphisms in Set coincide with injective functions. Indeed, suppose
.f : A → B is an injective map and g, .h : C → A are such that .f ◦ h = f ◦ g.
Then, we have .f h(c) = f g(c) for any .c ∈ C and since f is injective we
get .h(c) = g(c) for any .c ∈ C, i.e., .g = h as desired.
Assume now that .f : A → B is a monomorphism and let a, .a ∈ A be such
that .f (a) = f (a ). We denote by .ia : {∗} → A, respectively .ia : {∗} → A,
the maps given by .ia (∗) = a, ia (∗) = a . This implies that .f ◦ ia = f ◦ ia
and since f is a monomorphism we obtain .ia = ia . Therefore .a = a and f is
indeed injective.
Note that the proof above can be carried over verbatim to the categories Top
and KHaus by simply considering on .{∗} the indiscrete topology.9
(2) Similarly, in each of the categories Set, Grp, Ab, .R M epimorphisms coincide
with the surjective homomorphisms, while in Top and KHaus epimorphisms
coincide with the surjective continuous functions. We prove here that epi-
morphisms in Set and Grp coincide with surjective functions and surjective
homomorphisms, respectively.
Consider first an epimorphism .f : A → B in Set. Define g, .h : B → {0, 1}
as follows:
1, if b ∈ f (A)
.g(b) = , h(b) = 1, for all b ∈ B.
0, if b ∈
/ f (A)
9 The topology consisting only of the set itself and the empty set is called the indiscrete topology.
8 1 Categories and Functors
as follows:
⎧
⎨ kx, if h = ky
.σ (h) = ky, if h = kx , k ∈ K.
⎩
h, if h ∈
/ Kx ∪ Ky
First, we show that .ψ ◦ f = ξ ◦ f . Indeed, for any .g ∈ G and .h ∈ H, we have
ψ(f (g))(h) = f (g)h and respectively .ξ(f (g))(h) = σ −1 ◦ ψ f (g) ◦ σ (h).
.
Finally, .h ∈
/ Kx ∪ Ky yields
ψ(y −1 )(x) = y −1 x,
.
i.e., .t1 = t2 , which shows that f is an epimorphism in Set. Note that the
argument above can be used verbatim for the categories Grp, Ab, .R M in order
to show that surjective morphisms are epimorphisms.
1.3 Special Objects and Morphisms in a Category 9
(3) In each of the categories Set, Grp, Ab, .R M, isomorphisms coincide with the
bijective homomorphisms. In Top, isomorphisms are exactly the homeomor-
phisms, i.e., continuous bijections whose inverses are also continuous.
(4) In the category Div of divisible groups, the quotient map .q : Q → Q/Z is
obviously not injective but it is a monomorphism. Indeed, let G be another
divisible group and f , .g : G → Q be two morphisms of groups such that .q◦f =
q ◦ g. Denoting .f − g by h we obtain .q ◦ h = 0. Now for any .x ∈ G we have
.q(h(x)) = 0 and thus .h(x) ∈ Z. Suppose there exists some .x0 ∈ G such
that .h(x0 ) = 0. We can assume without loss of generality that .h(x0 ) ∈ N\{0}.
Since we are working with divisible groups, we can find some .y0 ∈ G such that
.x0 = 2h(x0 )y0 . Applying h to the above equality we obtain
and therefore .f (1/z) = 1/f (z). Similarly we can prove that .g(1/z) = 1/g(z)
and since f and g coincide on .Z we get .f (1/z) = g(1/z). Now for any .z ∈ Z
we have
(8) A rather special situation occurs in KHaus, the subcategory of Top con-
sisting of all compact Hausdorff topological spaces. As opposed to the cat-
egory of topological spaces, in KHaus any bijective continuous map .f ∈
HomKHaus (K, H ) is automatically an isomorphism. Indeed, it will suffice to
show that the inverse map .f −1 : H → K is continuous too. To this end, we
need to show that images of closed sets of K under f are closed in H ([39,
Theorem 18.1]). Consider U to be a closed subset of K; as K is compact it
follows that U is compact as well ([39, Theorem 26.2]). Moreover, as the image
of a compact space under a continuous map is compact ([39, Theorem 26.5]) we
obtain .f (U ) compact. Now recall that compact subspaces of Hausdorff spaces
are closed ([39, Theorem 26.3]). Therefore .f (U ) is closed, as desired.
(9) In the category PO(.X, ) associated to a partially ordered set .(X, ), any
isomorphism is an identity morphism. Indeed suppose .f : x → y is an
isomorphism; this implies that .x y. If .g : y → x is the inverse of f then
we also have .y x. Due to the antisymmetry of . we obtain .x = y. Therefore
.f : x → x must be the identity on x. .
The category .Set(⊆) has no final object. Indeed, U being a final object in
Set(⊆) would imply that for all sets X there exists a unique morphism .uX, U .
.
It is well known that for any object in Set one can define the notion of a subset.
The corresponding concept in an arbitrary category is called a subobject:
Definition 1.3.11 Let .C be a category and .C ∈ Ob C. An equivalence class
of monomorphisms with codomain C is called a subobject of C, where two
monomorphisms .f ∈ HomC (A, C) and .g ∈ HomC (B, C) are equivalent if there
exists an isomorphism .u ∈ HomC (A, B) such that .g ◦ u = f .
Note that if the category .C is not small then the subobjects of a given object C
might form a class rather than a set.
Examples 1.3.12
(1) In the category Set, the class of subobjects of a set X is in bijection with the
power set of X. Indeed, if we denote by .SO(X) the class of all subobjects of X,
the map .ψX : P(X) → SO(X) defined as follows is bijective:
ψX (Y ) = i
. Y, Y ⊆ X,
Y = f, as desired.
have .iY ◦ u = f , which shows that .i
We are left to prove that .ψX is also injective. Consider now Y , .Z ⊆ X
Y = i
such that .i Z . Hence, there exists a bijective map .u : Y → Z such that
.iZ ◦ u = iY . This comes down to .u(y) = y for all .y ∈ Y and therefore .Y ⊆ Z.
11 Let .τ , .τ be two topologies on a set X. Then .τ2 is called finer than .τ1 (or .τ1 is called coarser
1 2
than .τ2 ) if .τ1 ⊆ τ2 .
1.3 Special Objects and Morphisms in a Category 13
.ψX (Z, U) = i
Z, (Z, U) ∈ T(X),
where
.iZ denotes the equivalence class of the inclusion monomorphism
.iZ : Z, U → (X, τ ). Note that .iZ is obviously continuous as .U contains .τZ ,
As the subspace topology .τY is the coarsest topology on Y for which the
inclusion map .iY is continuous, we have .τY ⊆ Uf . This shows that .(Y, Uf ) ∈
T(X). Furthermore, as .g : (Z, V) → (Y, Uf ) is a homeomorphism and
.f = iY ◦ g, we can conclude that .iY = f and therefore .ψX (Y, Uf ) = i
Y = f.
This shows that .ψX is also surjective, as desired.
(3) As opposed to Top, in KHaus subobjects of a given object K are in bijection
with the closed subsets of K. Indeed, if we denote by .SSc (K) and .SO(K) the
set of all closed subsets and respectively the class of all subobjects of K, the
map .ψK : SSc (K) → SO(K) defined as follows is a bijection:
ψK (Y ) = i
. Y, Y ∈ SSc (K),
where
.iY denotes the equivalence class of the inclusion monomorphism
.iY : Y, τY → (K, τ ) and .τY is the subspace topology on Y .
14 1 Categories and Functors
= {z ∈ Z | f (z) ∈ U } = f −1 (U ) ∈ V.
ψX (f ) = Ef ,
. for all epimorphisms f ∈ HomSet (X, Y ),
ψG (K) = πK ,
. for all normal subgroups K of G,
u ◦ πK = f.
. (1.2)
Categories for which the subobjects (resp. quotients) of any given object form a
set are particularly important.
Definition 1.3.15 A category .C is called well-powered if the subobjects of any
object form a set. Similarly, .C is called co-well-powered if the quotients of any
object form a set.
Example 1.3.16 The discussion in Examples 1.3.12 and 1.3.14 immediately
implies that Set and Grp are both well-powered and co-well-powered. Furthermore,
in light of Example 1.3.12, Top and KHaus are well-powered. .
In this section we provide several methods of constructing new categories. The first
one relies on formally reversing the direction of the morphisms in the given category
leading to what is called the dual category.
Definition 1.4.1 Given a category .C, the dual (or opposite) category of .C, denoted
by .Cop , is defined as follows:
(i) .Ob Cop = Ob C;
(ii) .HomCop (A, B) = HomC (B, A); in order to avoid any confusion we write
.f
op : A → B for the morphism of .Cop corresponding to the morphism
.f : B → A of .C;
is defined as follows:
g op ◦op f op = (f ◦ g)op , for all f op ∈ HomCop (A, B), g op ∈ HomCop (B, C);
.
1.4 Some Constructions of Categories 17
op
(iv) the identities are the same as in .C, i.e., .1C = 1C for all .C ∈ Ob C.
Examples 1.4.2
(1) Obviously .(Cop )op = C for any category .C.
(2) Let PO(.X, ) be the category associated to the pre-ordered set .(X, ). Then
.PO(X, )
op = PO(X, ), where . is the pre-order on X defined as follows:
there exists a morphism in .PO(X, )op from x to y if and only if there exists a
morphism in .PO(X, ) from y to x. .
The dual category introduced above suggests that we can assign a dual to any
categorical concept. More precisely, the dual of a certain concept will be obtained
by considering this concept in the dual category. To illustrate this, we highlight
several dual concepts we have encountered so far.
Proposition 1.4.3 Let .C be a category and .f ∈ HomC (A, B).
(1) f is a monomorphism in .C if and only if .f op is an epimorphism in .Cop .
(2) I is an initial object in .C if and only if I is a final object in .Cop .
(3) f is an isomorphism in .C if and only if .f op is a isomorphism in .Cop .
Proof
op op op op
(1) Let .h1 , .h2 ∈ HomCop (A, C) be such that .h1 ◦op f op = h2 ◦op f op . In other
words, we have .f ◦ h1 = f ◦ h2 and the desired conclusion follows easily.
(2) If I is an initial object in .C then .HomC (I, C) has exactly one element for all
.C ∈ Ob C. Since .HomCop (C, I ) = HomC (I, C), the conclusion follows.
(iv) the identity maps are given by the trivial paths on each object.
12 The smallness assumption on the graph G is needed in order for the paths between any two
1 2
. 1 2 3
.Ob G = {v1 , v2 , v3 },
HomG (v1 , v1 ) = {(v1 )}, HomG (v1 , v2 ) = {(v1 e1 v2 )},
HomG (v1 , v3 ) = {(v1 e1 v2 e2 v3 )},
HomG (v2 , v1 ) = ∅, HomG (v2 , v2 ) = {(v2 )}, HomG (v2 , v3 ) = {(v2 e2 v3 )},
HomG (v3 , v1 ) = ∅, HomG (v3 , v2 ) = ∅, HomG (v3 , v3 ) = {(v3 )}.
.
f
A B
g
h
.
C
f
A B
k g
C D
. h
20 1 Categories and Functors
f
A B
k g
C D
. h
.
g ◦ f ∼A,C g ◦ f ∼A,C g ◦ f ,
.
tion 1.4.11, (ii) shows that the composition law in .C/ ∼ is well-defined.
Examples 1.4.14
(1) Consider a group G and let .G be the associated category (as in Exam-
ple 1.2.2, (3)) whose unique object we denote by .∗. There is a bijection between
normal subgroups of G and congruence relations on .G. Furthermore, for a
normal subgroup N of G, the quotient category by the congruence relation
induced by N is the quotient group .G/N.
Indeed, suppose first that N is a normal subgroup of G and define the
following equivalence relation on .HomG (∗, ∗) = G:
tg(th)−1 = tgh−1 t −1 ∈ N,
. gt (ht)−1 = gtt −1 h−1 = gh−1 ∈ N.
hx(hy) −1 ∈ N∼ where the last term belongs to .N∼ due to condition (1) of
Proposition 1.4.12. We have proved that .N∼ is a normal subgroup of G, as
desired.
In order to conclude that there is a bijection between the normal subgroups
of G and the congruence relations on .G we are left to show that for all normal
subgroups N of G and all congruence relations .∼ on .G we have
N∼N = N
. and ∼N∼ = ∼ .
g = (xy −1 )h ∼ 1h = h, i.e., g ∼ h.
.
F (x, 0) = f (x),
. F (x, 1) = g(x), G(x, 0) = g(x), G(x, 1) = h(x).
1.4 Some Constructions of Categories 23
F (x, 0) = f (x),
. F (x, 1) = f (x).
We can now define .F : X × [0, 1] → Z by .F (x, t) = g F (x, t) . This yields
F (x, 0) = g F (x, 0) = g f (x) = (g ◦ f )(x),
.
F (x, 1) = g F (x, 1) = g f (x) = (g ◦ f )(x),
G(x, 0) = g(x),
. G(x, 1) = g (x).
Define .G : X × [0, 1] → Z by .G (x, t) = G f (x), t . This yields
.G (x, 0) = G f (x), 0 = g f (x) = (g ◦ f )(x),
G (x, 1) = G f (x), 1 = g f (x) = (g ◦ f )(x),
which shows that .g ◦ f ∼X,Z g ◦ f . We can now conclude that the homotopy
relation .∼ is a congruence on Top. The resulting quotient category .Top/ ∼ is
called the homotopy category and will be denoted by HTop. For a thorough
introduction to homotopy theory, one of the cornerstones of algebraic topology,
we refer the reader to [4]. .
13 The pasting lemma: Let .X = A B, where A and B are closed in X, and consider two
continuous maps .f : A → Y , .g : B → Y . If .f (x) = g(x) for all .x ∈ A B, then f and g
f (x), if x ∈ A
combine to give a continuous function .h : X → Y , defined by .h(x) = (see [39,
g(x), if x ∈ B
Theorem 18.3]).
24 1 Categories and Functors
1.5 Functors
Functors are structure preserving maps which will be used to relate different
categories in the way morphisms do with objects.
Definition 1.5.1 Let .C and .D be two categories. A covariant functor (respectively
contravariant functor) .F : C → D consists of the following data:
(1) a mapping .A → F (A) : Ob C → Ob D;
(2) for each pair of objects A, .B ∈ Ob C, a mapping
1C 1 1C 2 1C 3 1C 4
f g
: C1 C2 C3 C4
1D 1 1D 2 1D 3
h k
: D1 D2 D3
. k h
14 The image of a functor (or values of a functor as defined in [23]) .F : C → D consists of a class
.{F (C) | C ∈ Ob C} together with all sets .{F (f ) | f ∈ HomC (A, B)} for any A, .B ∈ Ob C.
1.5 Functors 25
Indeed, the morphisms h and k are contained in the image of F while their
composition .k ◦ h is not. However, if F is injective on objects then it can be easily
seen that its image is indeed a category.
Examples 1.5.3
(1) If .C is a subcategory of .C we can define the inclusion functor .I : C → C
which sends every object as well as every morphism to itself. If .C = C then I
is just the identity functor .1C on .C.
(2) Let .∼ be a congruence on a category .C and .C/ ∼ the corresponding quotient
category. Then, we can define a quotient functor .Π : C → C/ ∼ as follows
for all .C ∈ Ob C and .f ∈ HomC (A, B):
Π (C) = C,
. Π (f ) = f ,
where .f denotes the equivalence class of the morphism f of .C. For all .f ∈
HomC (A, B) and .g ∈ HomC (B, C) we have
(1.3)
Π (g ◦ f ) = g ◦ f = g ◦ f = Π (g) ◦ Π (f ),
.
(6) Let .C, .D be two categories and consider .C × D to be the product category as
defined in Definition 1.4.4. We can define two projection functors as follows:
pC : C × D → C, pC (C, D) = C, pC (f, g) = f,
.
pD : C × D → D, pD (C, D) = D, pD (f, g) = g
for all .(C, D) ∈ Ob C × D and .(f, g) ∈ HomC×D (C, D), (C , D ) .
(7) If I is a small discrete category, then a functor .F : I → C is uniquely defined
by a family of objects .(Ci )i∈I indexed by I . More precisely, such a functor is
completely determined by the images of each object .i ∈ Ob I , say .Ci . Indeed,
note that for all .i ∈ Ob I the image of the identity morphism .1i is forced to be
.1Ci . If I is the empty set, then there exists a unique functor .F : I → C, called
(10) Given a category .C and a fixed object .C ∈ Ob C, we can define two functors,
one of them being covariant and the other one contravariant, called the hom
1.5 Functors 27
functors. Indeed, define .HomC (C, −), .HomC (−, C) : C → Set as follows:
In certain cases, the hom sets .HomC (A, B) can inherit some extra structure
from the objects of .C, as can be seen in the following examples. This simple
observation is the main idea behind the concept of an enriched category. For
the precise definition and further details on enriched category theory we refer
the reader to [33].
(11) If X, .Y ∈ Ob Top then the set of continuous maps .HomTop (X, Y ) can be
endowed with the so-called compact-open topology,15 which will turn out to
be particularly important when dealing with adjoint functors in Chap. 3. We
check first that if .f ∈ HomTop (Y, Z) then .HomTop (X, f ) : HomTop (X, Y ) →
HomTop (X, Z) is a continuous map with respect to the compact-open topol-
ogy. Indeed, let .K ⊆ X be a compact subset, .U ⊆ Z an open sub-
set and .W (K, U ) a sub-basis open set of the compact-open topology on
.HomTop (X, Z). As .f
−1 (U ) is an open subset of Y , we have
HomTop (X, f )−1 W (K, U )
.
15 The compact-open topology on .HomTop (X, Y ) is the topology generated by the sub-basis
.W (K, U ) = {f ∈ HomTop (X, Y ) | f (K) ⊆ U }, where .K ⊆ X is compact and .U ⊆ Y is
open.
28 1 Categories and Functors
functor.
(15) By composing the dual space functor with itself we obtain a covariant functor
denoted by .(−)∗∗ : K M → K M and called the double dual space functor.
We only point out for further use that if .u : U → V is a linear map then
.u
∗∗ : U ∗∗ → V ∗∗ is defined by
(16) The cartesian product bifunctor .−×− : Set×Set → Set is defined as follows
for all X, .Y ∈ Ob Set and .f ∈ HomSet (A, C), .g ∈ HomSet (B, D):
(− × −)(A, B) = A × B;
.
(− × −)(f, g) = f × g : A × B → C × D,
where .(f × g)(a, b) = f (a), g(b) for all .a ∈ A, .b ∈ B.
(17) For any set X we can define the corresponding cartesian product functor .− ×
X : Set → Set as follows for all Y , .Z ∈ Ob Set and .f ∈ HomSet (Y, Z):
(− × X)(Y ) = Y × X;
.
(− × X)(f ) = f × 1X : Y × X → Z × X,
1.5 Functors 29
where .(f × 1X )(y, x) = f (y), x for all .y ∈ Y , .x ∈ X. Similarly we can
define the cartesian product functor .X × − : Set → Set.
Exactly as in the case of the hom functor, in certain cases, when the
given sets are endowed with some extra structures, the corresponding cartesian
product inherits this structure.
(18) Any .X ∈ Ob Top defines a functor .− × X : Top → Top, where for all .Y ∈
Ob Top we consider on .Y × X the product topology.16
(19) Similarly, any .G ∈ Ob Grp defines a functor .− × G : Grp → Grp, where for
all .H ∈ Ob Grp the group structure on .H × G is defined component-wise.
(20) Let K be a field and for simplicity denote the tensor product over K by .⊗
(i.e., .⊗ = ⊗K ). For any .X ∈ Ob K M we can define a functor .− ⊗ X : K M →
K M, called the tensor product functor, as follows:
(22) The so-called forgetful functors are functors which forget (some of) the
structure on objects of the domain category. For instance the categories in
Example 1.2.2, (7)–(13) allow for a forgetful functor to the category Set of
sets, which sends the objects of that category to the underlying set, and the
(23) For a group G we denote by .[G, G] its commutator subgroup; in other words
.[G, G] is the subgroup of .G generated by all elements of the form .xyx
−1 y −1 ,
ker(πH ◦ f ). Hence, the universal property of the quotient group .Gab yields
a unique group homomorphism .fab : Gab → Hab which makes the following
diagram commutative:
πG
G Gab
f ab
πH f
.
Hab (1.6)
iX
X X i.e., f iX iY f.
f
iY f
.
Y (1.7)
iM
M G(M) i.e., f iM iN f.
f
iN f
.
G(N) (1.8)
The functor .G : Mon → Grp defined below is called the universal enveloping
group functor:
(26) Given a monoid .(M, ·), we denote by .U (M) the set of all invertible elements
of M. Furthermore, if .f ∈ HomMon (M, N ), it can be easily seen that
.f|U (M) ⊆ U (N ). This gives rise to a functor .U : Mon → Grp defined as
follows:
(27) Similarly, given a ring .(R, +, ·) we denote by .U (R) the set of invertible
elements of .(R, ·). This yields a functor .U : Ring → Grp defined as follows:
Note that each .BilM, N (A) can be made into an abelian group as in (1.4), i.e.,
for all .α, .β ∈ BilM, N (A) and .m ∈ M, .n ∈ N define
(29) Let
−1R be a commutative ring with unity, S a multiplicative subset−1
of R, and
. S R, j the corresponding localization ring, where .j : R → S R is the
ring homomorphism defined by .j (r) = 1r , for all .r ∈ R. Furthermore, if
.M ∈ Ob R M, we denote by .(S
−1 M, ϕ ) the corresponding localization
M
module, where .S M ∈ Ob S −1 R M and .ϕM : M → S −1 M is the R-module
−1
M
M S 1M i.e., f f.
M N
f
N f
S 1N
. (1.9)
qX
X H (X) i.e., f iX iY f.
f
qY f
.
H (Y ) (1.10)
Now we can define a functor .H : Top → Haus, called the Hausdorff quotient
functor, as follows:
jR
R D(R) i.e., f jR jS f.
f
jS f
.
D(S) (1.11)
Now we can define a functor .D : Rng → Ring, called the Dorroh extension
functor, as follows:
(32) For any .u ∈ HomRingc (R, S) we can define a functor .Fu : S M →R M, called
restriction of scalars, as follows for all .M ∈ Ob S M and .f ∈ HomS M (M, N ):
Fu (f ) = f.
(33) Given a pre-ordered set .(X, X ), we denote by .(X, AX ) the Alexandroff (or
Alexandrov) topology20 on X with respect to the pre-order .X . Furthermore,
if .f : (X, X ) → (Y, Y ) is order preserving then .f : (X, AX ) →
(Y, AY ) is continuous. Indeed, let .U ⊆ Y be an open subset, .x ∈ f −1 (U )
and .x ∈ X such that .x X x . As f is order preserving, we have
.f (x) X f (x ) and since .f (x) ∈ U we obtain .f (x ) ∈ U . Therefore,
.x ∈ f
−1 (U ), which shows that .f −1 (U ) ⊆ X is an open set. This yields a
F (f ) = f.
(34) A contravariant functor .F : Top → Poset can be defined as follows for all
topological spaces .(X, τ ), .(Y, γ ) and continuous maps .f : (X, τ ) → (Y, γ ):
. F (X, τ ) = {U ⊆ X | U is open in τ };
F (f ) = f −1 ,
20 Given a pre-ordered set .(X, X ), the Alexandroff topology on X with respect to .X is defined
by considering a subset U of X to be open if .x X x and .x ∈ U imply .x ∈ U .
1.5 Functors 35
Proof We only show the first assertion. To this end, for any .B ∈ Ob B we
have .FA (1B ) = F (1A , 1B ) = F (1(A, B) ) = 1F (A, B) , where the last equality
holds because F is a functor. Thus .FA respects identities. Furthermore, for any
.f ∈ HomB (B, B ) and .g ∈ HomB (B , B ) we have
FA (g ◦ f ) = F (1A , g ◦ f ) = F (1A , g) ◦ (1A , f ) = F (1A , g) ◦ F (1A , f )
.
= FA (g) ◦ FA (f ).
We have proved that .FA respects compositions as well and the proof is now finished.
36 1 Categories and Functors
Examples 1.5.6
(1) For any object C in a category .C, the hom functor .HomC (C, −) is the right
associated functor of the Hom bifunctor with respect to C. Similarly, the
contravariant hom functor .HomC (−, C) is the left associated functor of the Hom
bifunctor with respect to C.
(2) For any set X, the cartesian product functors .X × − and .− × X are the right
and respectively the left associated functors with respect to X of the cartesian
bifunctor. .
We start by introducing the following notions, which are weaker than isomorphism
but very useful.
Definition 1.6.1 Let .F : C → D be a functor and for all A, .B ∈ Ob C consider the
following induced mapping:
FA, B : HomC (A, B) → HomD F (A), F (B) , f → F (f ).
. (1.12)
(1) The functor F is called faithful if the mappings .FA, B are injective for all A,
.B ∈ Ob C.
(2) The functor F is called full if the mappings .FA, B are surjective for all A, .B ∈
Ob C.
(3) The functor F is called fully faithful if the mappings .FA, B are bijective for all
A, .B ∈ Ob C.
(4) The functor F is called essentially surjective if each object .D of D is
isomorphic to an object of the form .F (C) for some .C ∈ Ob C.
Examples 1.6.2
(1) The inclusion functor is automatically faithful. If the subcategory is full then
the inclusion functor is also full.
(2) The inclusion functor .I : Ab → Grp is fully faithful.
(3) The quotient functor .Π : C → C/ ∼ is always full, where .∼ is a congruence on
the category .C and .C/ ∼ denotes the corresponding quotient category.
(4) The quotient functor .Π : Top → HTop is full but not faithful.
(5) The inclusion functor .I : Cgrp → C is faithful but not full unless .C itself is
a groupoid, where .Cgrp is the core groupoid of the category .C constructed in
Example 1.3.6, (2).
(6) Given .f ∈ HomRingc (A, B), the restriction of scalars functor .Ff : B M →
A M defined in Example 1.5.3, (32) is obviously faithful. Furthermore, if f is
an epimorphism in .Ringc then the corresponding restriction of scalars functor
is also full ([53, Proposition XI.1.2]). Indeed, let M, .N ∈ Ob B M and .u ∈
1.6 Isomorphisms of Categories 37
Note that throughout this example all module actions will be denoted by
juxtaposition and we will see B as an A-module via f , i.e., .ab = f (a)b for
all .a ∈ A and .b ∈ B.
Now let .m ∈ M and consider the map .v : B ⊗A B → N defined for all b,
.b ∈ B as follows:
v(b ⊗A b ) = b u(bm).
. (1.14)
.α(b) = b ⊗A 1B , β(b) = 1B ⊗A b.
This follows easily by noticing that since .Z is the initial object in Ring, the
set .HomRing (Z, Fp ) has only one element while the cardinality of the set
.HomGrp (Z2 , Zp−1 ) is .gcd(2, p − 1) = 2 (see [27]).
38 1 Categories and Functors
In order to show that .U is not faithful either, consider the polynomial ring
k[X] over a field k and recall that .U (k[X]) = k\{0}. Then, the following
.
UF2 [X], F2 [X] : HomRing (F2 [X], F2 [X]) → HomGrp ({1}, {1}),
.
where .{1} denotes the trivial group. Indeed, the set .HomGrp ({1}, {1}) obviously
has only one element while the cardinality of .HomRing (F2 [X], F2 [X]) is at
least two. .
Example 1.6.4 The categories FinSet, Grp, Ab, Rng, Ring, Top, .R M are all
concrete categories due to the existence of forgetful functors from any of the above
categories to Set, which are obviously faithful. .
In fact, we have a lot more examples of concrete categories, as can be seen from
the next result which, as noted in [6, Theorem 7.5.6], resembles Cayley’s theorem
from group theory.
Theorem 1.6.5 Any small category is concrete.
Proof For any small category .C we construct a faithful functor .F : C → Set as
follows. Given .C ∈ Ob C and .u ∈ HomC (C, C ) we define
.u1 = u2 , as desired.
Not every category admits a faithful functor to Set. The interested reader may
find such an example in [24], where it is shown that the homotopy category of
pointed spaces is not concrete.
Definition 1.6.6 A functor .F : C → D is called an isomorphism of categories if
there exists another functor .G : D → C such that .F ◦ G = 1D and .G ◦ F = 1C .
In this case we say that the categories .C and .D are isomorphic and G is called
the inverse of F . A contravariant isomorphism of categories is called an anti-
isomorphism of categories.
Let .F : C → D be a functor and for all A, .B ∈ Ob C consider the induced
mapping .FA, B defined in (1.12). If F is an isomorphism of categories with inverse
.G : D → C then each .FA, B is a bijective map with inverse given by .GF (A), F (B) ,
1.6 Isomorphisms of Categories 39
where
GF (A), F (B) : HomD F (A), F (B) → HomC (A, B), GF (A), F (B) (g) = G(g)
.
for all .g ∈ HomD F (A), F (B) .
In particular, an isomorphism of categories takes initial (resp. final) objects to
initial (resp. final) objects. Indeed, it follows from the above discussion that there is
a bijection between the sets .HomC (A, G(D)) and respectively .HomD (F (A), D);
hence, if A is an initial object in .C then .F (A) is an initial object in .D.
Examples 1.6.7
(1) The forgetful functor .F : Z M → Ab is an isomorphism of categories. Indeed,
the inverse of F is the functor .G : Ab → Z M defined by .G(M) = M, .G(u) =
u, where .M ∈ Ab has a left .Z-module structure as follows:
⎧
⎪
⎪
⎪
⎪ +m+
m · · · + m if t > 0
⎪
⎪
⎨ t times
.t · m = 0M if t = 0 , for every t ∈ Z, m ∈ M.
⎪
⎪
⎪ −m
⎪
⎪
⎪ − m
− · · · − m if t < 0
⎩
−t times
(2) Let R be a ring and denote by .R op the opposite ring.21 Then we have an
isomorphism of categories .F : R M → MR op given by
F (u) = u,
21 In .R op we have .(R op , +) = (R, +) and the multiplication is given by .r ·op r = r r, for all r,
.r ∈ (R op , +).
40 1 Categories and Functors
such that
Since F is full we can find .g ∈ HomC (B, A) such that .F (g) = h. Therefore,
the above identities come down to
HomC (I, C) ∼
. = HomD (F (I ), F (C)).
Examples 1.6.10
(1) The forgetful functor .U : Grp → Set reflects isomorphisms. Indeed, recall that
a group homomorphism is an isomorphism if and only if it is a bijection.
(2) The forgetful functor .U : Top → Set does not reflect isomorphisms, as can be
easily seen from Example 1.3.2, (7). .
above mentioned categories all injective maps are monomorphisms and respectively
all surjective maps are epimorphisms. However, the converse is not necessarily true,
as can be seen from Example 1.3.2, (4) and (5) respectively.
αC
F (C) G(C) i.e., α C F(f) G( f ) α C .
F( f ) G( f )
F (C ) G(C )
αC
. (1.15)
a morphism .ϕ∗ : → in .G2 (i.e., an element of the group .G2 ) which makes
the following diagram commute for all morphisms .t : ∗ → ∗ in .G1 (i.e., an
element of the group .G1 ):
follows:
ηV
V V
u u
W W
.
ηW
−1
αC−1 = αC
. for all C ∈ Ob C
. G
K
α
. G
called the whiskering of the natural transformation .α on the left by the functor
K.
Furthermore, if .α is a natural isomorphism then both .H α and .αK are natural
isomorphisms. Indeed, the first assertion follows by Proposition 1.6.9, (1) while
the second one is an easy consequence of .α itself being a natural isomorphism.
.
F H
. G K
. (1.18)
Therefore, we have .(β ∗ α)C = βG(C) ◦ H (αC ) = K(αC ) ◦ βF (C) for all .C ∈ Ob C.
Let .f ∈ HomC (C, C ); showing that .β ∗ α is a natural transformation comes down
to proving the commutativity of the following diagram:
. (1.19)
46 1 Categories and Functors
To this end, the naturality of .α and functoriality of H render the following diagram
commutative:
(1.20)
(1.21)
(1.20)
βG(C ) ◦ H (αC ) ◦ H F (f ) = βG(C ) ◦ H G(f ) ◦ H (αC )
.
(1.21)
= KG(f ) ◦ βG(C) ◦ H (αC ),
which proves that (1.19) holds and the proof is now finished.
(2) for all .B0 ∈ Ob B, the family of morphisms .(α B0 )A : F B0 (A) → GB0 (A) in .C,
indexed by .A ∈ Ob A, form a natural transformation .α B0 : F B0 → GB0 , where
.(α 0 )A = α(A,B0 ) .
B
that .α : F → G
Proof Assume first is a natural transformation; therefore, for all
(f, g) ∈ HomA×B (A, B), (A , B ) the following diagram is commutative:
.
(1.22)
. (1.23)
. (1.24)
G(u, v) ◦ α(X,Y ) = G (u, 1Y ) ◦ (1X , v) ◦ α(X,Y )
.
= GY (u) ◦ (α Y )X ◦ FX (v)
(1.24) Y
= α X ◦ F Y (u) ◦ FX (v)
= α(X ,Y ) ◦ F (u, 1Y ) ◦ F (1X , v)
= α(X ,Y ) ◦ F (u, v).
This shows that for all .(u, v) ∈ HomA×B (X, Y ), (X , Y ) the following diagram
is commutative:
Indeed,
for any .g ∈ HomGrp (Z, X) we have .αY ◦ HomGrp (Z, f )(g) = αY f ◦
g = (f ◦ g)(1) and .f ◦ αX (g) = f (g(1)).
(2) The forgetful functor .U : Top → Set is representable and the representing
object is any singleton topological space .{x0 } (with the discrete topology).
To this end, for any .X ∈ Ob Top and any .h ∈ HomTop ({x0 }, X) we define
.αX : HomTop ({x0 }, X) → U (X) = X by .αX (h) = h(x0 ) ∈ X. Each .αX is
obviously bijective. Furthermore, it can be easily seen that the above diagram is
commutative for any .f ∈ HomTop (X, Y ):
.
50 1 Categories and Functors
f ◦ αX (h) = f (h(x0 ))
.
αY ◦ HomTop ({x0 }, f )(h) = αY f ◦ h = (f ◦ h)(x0 ).
bijective map .τ : HomSet (A, {∗}) → P({∗}). This leads to a contradiction since
.|HomSet (A, {∗})| = 1 and .|P({∗})| = 2. Therefore, .P is not representable. .
Proof Suppose first that F is representable, i.e., there exists a natural isomorphism
.ϕ : HomC (A, −) → F for some .A ∈ Ob C. In particular we have a bijection of sets
.ϕA : HomC (A, A) → F (A) and we denote .ϕA (1A ) ∈ F (A) by a. We will show
.
1.7 Natural Transformations: Representable Functors 51
−1
i.e., .ϕB ◦ HomC (A, f ) = F (f ) ◦ ϕA . This yields .F (f ) = ϕB ◦ HomC (A, f ) ◦ ϕA
and we obtain
−1
F (f )(a) = ϕB ◦ HomC (A, f ) ◦ ϕA
. (a)
= ϕB ◦ HomC (A, f )(1A )
= ϕB f ◦ 1A
= ϕB (f ) = b.
Assume now that .(A, a) is a representing pair. Let .ψ : HomC (A, −) → F be the
natural isomorphism defined as follows for any .B ∈ Ob C:
The property assumed to be satisfied by .(A, a) implies that each such map .ψB is
bijective. The proof will be finished once we show that .ψ is a natural transformation,
i.e., for any .g ∈ HomC (B, C) the following diagram is commutative:
F (u)(g) = u ◦ g,
for any .G ∈ Ob Grp and any .u ∈ Hom Grp (G, G ), .g ∈ F (G). Then F is
representable and . A/H, π : A → A/H is the representing pair, where .π is
the canonical projection. Indeed, consider another pair .(G, f : A → G) with
52 1 Categories and Functors
π
A A/H
f
f
.
G
Since .H ⊆ kerf , the universal property of the quotient group .A/H yields a unique
f ∈ HomGrp (A/H, G) such that the above diagram is commutative, i.e., .f ◦π = f .
.
The last equality is equivalent to .F (f )(π ) = f and the desired conclusion now
follows from Proposition 1.7.7. .
The dual category allows us not only to define a dual notion for every concept but
also to state a dual result for any theorem. This new result requires no proof and
is obtained by reversing all morphisms and consequently the order of composition
in the given theorem. Indeed, if a given statement T is valid in any category then
the dual statement .T op is also valid in any category. This can be easily seen by
noticing that proving the statement .T op in a category .C is equivalent to proving the
statement T in the category .Cop , which is assumed to be valid. To illustrate this
duality principle we look at the following statement proved in an equivalent form in
Proposition 1.3.9:
When it exists, the initial object of a category is unique up to isomorphism.
By simply applying the duality principle we obtain the following dual statement:
When it exists, the final object of a category is unique up to isomorphism.
A certain care is required, however, when dealing with statements which involve
functors. More precisely, in the process of dualizing these statements, all categories
are replaced by their duals, and all morphisms are reversed, while functors .C → D
are not reversed but replaced by dual functors .Cop → Dop as defined below.
Proposition 1.8.1 Let .F : C → D be a functor. There exists a functor .F op : Cop →
Dop , called the dual functor, defined as follows for any C, .D ∈ Ob Cop and any
.f
op ∈ Hom op (C, D):
C
F op (C) = F (C),
. F op (f op ) = F (f )op .
(G ◦ F )op = Gop ◦ F op ,
. (1.25)
1.8 The Duality Principle 53
In light of the duality principle, any statement about covariant functors has a
correspondent for contravariant functors obtained by replacing the given functor
.F : C → D by .F ◦ OCop : C
op → D (or .O ◦ F : C → Dop ). For example, in order
D
to obtain the contravariant version of Proposition 1.7.7 we consider .G : C → Set
to be a contravariant functor. Then .G ◦ OCop : Cop → Set is a covariant functor
and Proposition 1.7.7 implies that .G ◦ OCop : Cop → Set is representable if and
only if there exists a pair .(A, a) with .A ∈ Ob Cop and .a ∈ G ◦ OCop (A)
satisfying the following property: for any other pair .(B, b) with .B ∈ Ob Cop and
.b ∈ G ◦ OCop (B) there exists a unique .f
op ∈ Hom op (A, B) such that
C
.G ◦ OCop (f )(a) = b. Since .Ob C = Ob C and .HomCop (A, B) = HomC (B, A)
op op
property: for any other pair .(B, b) with .B ∈ Ob C and .b ∈ G(B) there exists a
unique .f ∈ HomC (B, A) such that .G(f )(a) = b.
Finally, the dual of a natural transformation is defined as follows:
Proposition 1.8.3 Let F , .G : C → D be two functors and .ψ : F → G a natural
transformation. There exists a natural transformation .ψ op : Gop → F op , called the
opposite or dual natural transformation, defined as follows for all .C ∈ Ob C:
op op
. ψ C = ψC (1.26)
and any natural transformation between .Gop and .F op appears as the dual of some
natural transformation between F and G. Furthermore, .ψ is a natural isomorphism
if and only if .ψ op is a natural isomorphism.
Proof For any .f op ∈ HomCop (C, C ), the naturality of .ψ renders the following
diagram commutative:
. (1.27)
We will show that this implies the naturality of .ψ op . Indeed, (1.27) takes the
following equivalent forms:
ψC ◦ F (f ) = G(f ) ◦ ψC
.
op op
⇔ ψC ◦ F (f ) = G(f ) ◦ ψC
⇔ F (f )op ◦op (ψC )op = (ψC )op ◦op G(f )op
⇔ F op f op ◦op (ψC )op = (ψC )op ◦op Gop f op ,
The second claim follows by noticing that for any natural transformation we have
(ψ op )op = ψ.
.
Numerous examples of duality arguments will be used in proofs throughout the
book.
Comma Categories
Comma categories are constructed using two functors with the same codomain. We
will see many instances of comma categories at work later on when dealing with
(co)limits or adjoint functors.
Theorem 1.8.4 Any two functors .F : A → C and .G : B → C define a comma
category denoted by .(F ↓ G) as follows:
(1) the objects are triples .(A, f, B) consisting of two objects .A ∈ Ob A, .B ∈ Ob B
and a morphism .f ∈ HomC (F (A), G(B));
(2) a morphism in .(F ↓ G) from .(A, f, B) to .(A , f , B ) is a pair .(a, b), where
.a ∈ HomA (A, A ), .b ∈ HomB (B, B ) such that the following diagram is
commutative:
. (1.28)
(3) the composition law in .(F ↓ G) is that induced by the composition laws of .A
and .B, i.e.:
.(a, b) ◦ (a , b ) = (a ◦ a , b ◦ b );
Proof First note that for any .(A, f, B) ∈ Ob (F ↓ G), the following diagram is
trivially fulfilled:
.
56 1 Categories and Functors
Examples 1.8.5
(1) Let .A and .B be discrete categories with only one object, say .Ob A = {A} and
.Ob B = {B}. As we have seen in Example 1.5.3, (7), defining two functors
.C. Then, the comma category .(F ↓ G) is isomorphic to the discrete category on
the set .HomC (C, C ). Indeed, the objects of the corresponding comma category
.(F ↓ G) are triples .(A, f, B), where .f ∈ HomC (C, C ). The morphisms in
.(F ↓ G) between two objects .(A, f, B) and .(A, f , B) are pairs .(a, b) where
.a ∈ HomA (A, A) and .b ∈ HomB (B, B). Now since the only morphisms in
the only morphisms in .(F ↓ G) are the identities on each object. Therefore
the functor from .(F ↓ G) to the discrete category on .HomC (C, C ) sending
any object .(A, f, B) to f and any morphism .(1A , 1B ) to the identity on f is
obviously an isomorphism of categories.
(2) Let .A = 1 be a discrete category with only one object, say A, and .B = C =
Top. Furthermore, consider the functors .F : 1 → Top, .G : Top → Top defined
as follows:
F (A) = {},
. F (1A ) = 1{} , G = 1Top ,
1.8 The Duality Principle 57
where .{} denotes a singleton set regarded as a topological space in the obvious
way. Then, the comma category .(F ↓ G) is isomorphic to the category of
pointed topological spaces .Top . To this end, the objects of the corresponding
comma category .(F ↓ G) are triples .(A, f, X), where .X ∈ Ob Top and
.f ∈ HomTop ({}, X). The morphisms in .(F ↓ G) between two objects
.(A, f, X) and .(A, g, Y ) are pairs .(a, b), where .a ∈ HomA (A, A) and
.b ∈ HomTop (X, Y ) such that diagram (1.28) is commutative. Now since the
. (1.29)
In what follows we write down, for further use, two important special cases of
comma categories:
Corollary 1.8.6 Let .F : A → C and .G : B → C be two functors.
(1) If .A is the discrete category with only one object and .F : A → C is the constant
functor at .C0 ∈ Ob C then the comma category .(F ↓ G) = (C0 ↓ G) is
isomorphic to the category defined as follows:
(i) the objects are pairs .(f, B), where .B ∈ Ob B, .f ∈ HomC (C0 , G(B));
(ii) a morphism in .(C0 ↓ G) between two objects .(f, B) and .(f , B )
is a morphism .b ∈ HomB (B, B ) such that the following diagram is
58 1 Categories and Functors
commutative:
C0 i.e., h f f
f f
C C
. h
1.8 The Duality Principle 59
f f
.
C0
In other words, u is the unique morphism in . {} ↓ F between .(fa , A) and .(g, B),
as desired.
Conversely, let .(f, A) be the initial object of . {} ↓ F , where .A ∈ Ob C and
.f ∈ HomSet ({}, F (A)), and consider .a = f (). Then, .(A, a) is the representing
theinitial object
of . {} ↓ F , there exists a unique morphism .u : (f, A) → (fb , B)
in . {} ↓ F such that .F (u) ◦ fa = g. To conclude, there exists a unique morphism
.u ∈ HomSet (A, B) such that .F (u)(a) = b and therefore .(A, a) is the representing
(1.30)
1.8 The Duality Principle 61
This shows that (1.30) is commutative if and only if the following diagram is
commutative:
. (1.31)
Hence we have .(a, b) ∈ Hom (F ↓G) (A2 , f2 , B2 ), (A1 , f1 , B1 ) . In fact, all the
above shows that . bop , a op ∈ HomGop ↓F op (B1 , f1 , A1 ), (B2 , f2 , A2 if
op op
and only if .(a, b) ∈ Hom(F ↓G) (A2 , f2 , B2 ), (A1 , f1 , B1 ) .
We
opcan now define two functors .H : (F ↓ G)op → Gop ↓ F op and
.T : G ↓ F op → (F ↓ G)op as follows:
. H (A, f, B) = B, f op , A ,
(A, f, B) ∈ Ob (F ↓ G)op ,
H (a, b)op = bop , a op ,
(a, b)op ∈ Hom(F ↓G)op (A1 , f1 , B1 ), (A2 , f2 , B2 ) ,
T (B, f op , A) = (A, f, B),
(B, f op , A) ∈ Ob Gop ↓ F op ,
T (bop , a op ) = (a, b)op ,
(bop , a op ) ∈ HomGop ↓F op (B1 , f1 , A1 ), (B2 , f2 , A2 .
op op
62 1 Categories and Functors
The above discussion shows that both H and T are well-defined. We will show that
they are indeed functors. To start with, for all .A ∈ Ob A and .B ∈ Ob B we have
op op
H 1(A, f, B) = H (1A , 1B )op = 1B , 1A = 1B, f op , A ,
op
.
op op
T 1B, f op , A = T (1B , 1A ) = (1A , 1B )op = 1(A, f, B) .
op
op ∈ Hom
Consider now two morphisms .(a, b) (F ↓G)op (A 1 , f1 , B1 ), (A2 , f2 , B2 )
and .(c, d)op ∈ Hom(F ↓G)op (A2 , f2 , B2 ), (A3 , f3 , B3 ) . We have
op op
H (c, d)op ◦op (a, b)op = H (a, b) ◦ (c, d)
. = H a ◦ c, b ◦ d
= (b ◦ d)op , (a ◦ c)op = d op ◦op bop , cop ◦op a op = d op , cop ◦ bop , a op
= H (c, d)op ◦ H (a, b)op .
we have
T (r op , s op ) ◦ (uop , v op ) = T (r op ◦op uop ), (s op ◦op v op
.
= T (u ◦ r)op , (v ◦ s)op
op op
= v ◦ s, u ◦ r = (v, u) ◦ (s, r) = (s, r)op ◦op (v, u)op
= T (r op , s op ) ◦op T (uop , v op ) .
The proof is now finished as H and T are obviously inverses to each other.
As the name suggests, functor categories have functors as objects and natural
transformations as morphisms. The next result shows that this is indeed a category
under certain conditions.
Proposition 1.9.1 Let I and .C be two categories. If I is a small category then the
functors from I to .C and the natural transformations between them as morphisms
1.9 Functor Categories 63
form a category, called a functor category, which we denote by .Fun (I, C). If .C is
also small then .Fun (I, C) is small.
Proof The composition of natural transformations is given by the vertical composi-
tion as defined in Example 1.7.2, (5). This composition law is obviously associative
and the identity at each functor F is just the identity natural transformation defined
in Example 1.7.2, (4).
Note that . i∈Ob I HomC F (i), G(i)
is a union
of sets indexed by another set,
namely .Ob I . Hence . i∈Ob I HomC F (i), G(i) is a set as well. Finally, note that
for any two functors F , .G : I → C, a natural transformation .η : F → G is
determined by a map
η : Ob I →
. HomC F (i), G(i) .
i∈Ob I
η ∗ : CF → CH , η : D F → D H .
.
Indeed, for all .i ∈ Ob I , we define .ψi to be the unique morphism in .C from .C0 to
.F (i) (recall that .C0 is, in particular, an initial object in .C). We are left to show that
.ψ as defined above is a natural transformation. Given any .f ∈ HomI (i, j ), both .ψj
ϕ is the unique natural transformation which can be defined between the two
.
functors above as for each .i ∈ Ob I there exists a unique morphism between .F (i)
and .C0 .
Theorem 1.9.5 Let .B, .C be two categories with .B small. Then we have an
isomorphism of categories between .Fun(Bop , Cop ) and .Fun(B, C)op .
Proof Throughout this proof, for any morphism .α : F → G in .Fun(B, C)
(i.e., natural transformation) we denote by .α op : G → F the corresponding
opposite morphism in .Fun(B, C)op (Definition 1.4.1) while .α op : Gop → F op
stands for the opposite natural transformation as defined in Proposition 1.8.3.
op
Moreover, .◦op denotes the composition in .Fun(B, C)op and .◦C (resp. .◦C ) stands
for the composition in the category .C (resp. .Cop ). Finally, we use .• to denote
the composition in .Fun(Bop , Cop ) and the unadorned .◦ for the composition in
.Fun(B, C).
66 1 Categories and Functors
Define a functor .T : Fun(B, C)op → Fun Bop , Cop as follows:
T (F ) = F op ,
. T (α op ) = α op (1.32)
(β ◦ α)op = α op • β op .
. (1.33)
We obtain:
(1.33)
T α op ◦op β op = T (β ◦ α)op = (β ◦ α)op = α op • β op = T α op • T β op
.
and therefore T is indeed a functor. We are left to construct the inverse func-
tor of T . To this end, consider .S : Fun Bop , Cop → Fun(B, C)op defined
by
S(U ) = U op ,
. S(β) = (β op )op (1.34)
We aim to show that .S(β • α) = S(β) ◦op S(α). To start with, note
that for any .B ∈ Ob B we have .αB ∈ HomCop (U (B), V (B)) and therefore
op
.αB = α
B for some morphism .α B ∈ HomC (V (B), U (B)). Similarly, .βB =
op
β B for some morphism .β ∈ HomC (W (B), V (B)). This leads to the follow-
ing:
op op op op op
.(β • α)op B = (β • α)B = (βB ◦C αB )op = βB ◦C αB op
op op
= (αB ◦C βB )op = αB ◦C βB = (αB op )op ◦C (βB )op
op op
= αB ◦C βB = (α op )B ◦C (β op )B = α op ◦ β op B .
1.9 Functor Categories 67
Hence, we have:
(β • α)op = α op ◦ β op .
. (1.35)
as desired. Finally, we will show that T and S are inverses to each other. To this
end, given two functors F , .G : B → C and a natural transformation .α : F → G we
have
(1.32) (1.34)
.S T (F ) = S(F op ) = (F op )op = F,
(1.32) (1.34) op
S T (α op ) ) = S(α op ) = (α op )op = α op .
(1.34) (1.32)
T S(U ) = T (U op ) = (U op )op = U,
.
(1.34) (1.32)
T S(β) = T (β op )op = = (β op )op = β
.ψ
op : F op → Gop is an epimorphism in .Fun(I op , Cop ).
Ev(F, i) = F (i),
. Ev(η, f ) = G(f ) ◦ ηi .
. (1.38)
We have proved that .Ev preserves identities. Consider now .f ∈ HomI (i, j ), .g ∈
HomI (j, k) and .η : F → G, .ξ : G → H , natural transformations, where F , G,
.H : I → C. We obtain:
Ev (ξ, g) ◦ (η, f ) = Ev(ξ ◦ η, g ◦ f ) = H (g ◦ f ) ◦ (ξ ◦ η)i
.
where in the fourth equality we used the naturality of .ξ . Hence .Ev preserves
compositions and is indeed a functor.
Ev(F, i) = F (i)
. and Ev(η, 1i ) = G(1i ) ◦ ηi = ηi (1.39)
Another functor which will later turn out to be important is the diagonal functor
defined as follows:
Proposition 1.9.8 Let I be a small category and .C an arbitrary category. There
exists a functor . : C → Fun(I, C), called the diagonal functor, defined as follows
for all C, .D ∈ Ob C and all .f ∈ HomC (C, D):
.(C) = C , (f ) = η : C → D ,
.
70 1 Categories and Functors
comes down to
t
C D
1C 1D
C D
. t
which is obviously commutative. This shows that .η defined by .ηi = t for all .i ∈
Ob I is indeed a natural transformation and therefore . is well-defined.
Clearly, for any .C ∈ Ob C, we have .(1C ) = ξ , where .ξ : C → C is the
natural transformation defined by .ξi = 1C for all .i ∈ Ob I . Hence .ξ is the identity
natural transformation .1C and we have proved that . respects identities.
Consider now .f ∈ HomC (C, D) and .g ∈ HomC (D, E) and let .(f ) = μ,
.(g) = ν where .μ : C → D and .ν : D → E are natural transformations
Yoneda’s lemma is arguably one of the most important results contained in this
book. It allows us to embed any category into its category of presheaves by means of
a fully faithful functor and this approach opens the way to a plethora of applications.
Some important consequences of Yoneda’s lemma, which go beyond the scope of
this book, are mentioned at the end of the section.
Theorem 1.10.1 (Yoneda’s Lemma) Let .F : C → Set be a functor and .C ∈
Ob C. Then the natural transformations from .HomC (C, −) to F are in bijection
with the elements of the set .F (C) and the bijection is given for any natural
transformation .ϕ : HomC (C, −) → F as follows:
πC, F : Nat HomC (C, −), F → F (C), πC, F (ϕ) = ϕC (1C ) ∈ F (C).
. (1.40)
In particular, .Nat HomC (C, −), F is a set.
1.10 Yoneda’s Lemma 71
Proof Consider .τC, F : F (C) → Nat HomC (C, −), F , defined for every .x ∈
F (C) by:
τC, F (x) = hx where hx D (f ) = F (f )(x)
. (1.41)
for every .D ∈ Ob C and .f ∈ HomC (C, D). First we have to check that .hx is a
natural transformation. This comes down to proving that given a morphism .f ∈
HomC (A, B) the following diagram is commutative:
= F (f ◦ g)(x) = F (f ) ◦ F (g)(x)
= F (f )(hx )A (g).
Thus .hx is a natural transformation. The proof will be finished once we show that
.πC, F and .τC, F are inverses to each other. To start with, consider .x ∈ F (C). Then
we have
.(πC, F ◦ τC, F )(x) = πC, F τC, F (x) = πC, F hx
= hx C (1C ) = F (1C )(x)
= 1F (C) (x) = x.
.
72 1 Categories and Functors
(1.42)
= F (f ) ϕC (1C )) = ϕD (f ),
which implies .τC, F ◦ πC, F (ϕ) = ϕ and the proof is now complete.
It turns out that the bijections .πC, F defined in (1.40) form a natural transforma-
tion in the variable C. Furthermore, if .C is a small category, then the bijections .πC, F
form a natural transformation in the variable F as well. The precise statement is the
following:
Theorem 1.10.2 Let .C be a category, .C ∈ Ob C and .F : C → Set a functor.
(1) If .G : C → Set is the functor defined as follows for all .C ∈ Ob C and .f ∈
HomC (C, C ):
G(C) = Nat HomC (C, −), F ,
.
G(f ) : Nat HomC (C, −), F → Nat HomC (C , −), F ,
G(f )(ψ) = ψ ◦ HomC (f, −), ψ ∈ Nat HomC (C, −), F
Proof
(1) Let .f ∈ HomC (C, C ); we need to prove the commutativity of the following
diagram:
Indeed, for all .ϕ ∈ G(C) = Nat HomC (C, −), F we have
(1.40)
.F (f ) ◦ πC, F (ϕ) = F (f ) ϕC (1C ) = ϕC (f ) = ϕC HomC (f, C )(1C )
= ϕ ◦ HomC (f, −) C (1C ) = πC , F ϕ ◦ HomC (f, −) = πC , F ◦ G(f ) (ϕ),
To this end, let .ϕ ∈ H (F ) = Nat HomC (C, −), F ; we obtain
(1.40)
. EvC (ψ) ◦ πC, F (ϕ) = EvC (ψ) πC, F (ϕ) = EvC (ψ) ϕC (1C )
(1.39)
= ψC ϕC (1C ) = (ψ ◦ ϕ)C (1C ) = πC, F (ψ ◦ ϕ)
= πC, F H (ψ)(ϕ) = πC, F ◦ H (ψ) (ϕ),
as desired.
Remark 1.10.3 The contravariant version of Yoneda’s lemma can be easily
obtained by replacing the category .C with .Dop in Theorem 1.10.1 and noticing
that the functors .HomDop (C, −) : Dop → Set and .HomD (−, C) : Dop → Set
coincide.
74 1 Categories and Functors
follows:
φ
.h X (f ) = f ◦ φ, for all X ∈ Ob C and f ∈ HomC (C, X). (1.43)
follows:
δ
.t X (g) = δ ◦ g, for all X ∈ Ob C and g ∈ HomC (X, C). (1.44)
Proof
for every .X ∈ Ob C and .f ∈ HomC (C, X). Hence, any natural transformation
between .HomC (C, −) and .HomC (D, −) is of the form .hφ for a unique .φ ∈
HomC (D, C), as desired.
(1) Follows from the contravariant version of Yoneda’s lemma.
Proposition 1.10.5 Let .C be a category and C, .D ∈ Ob C. Then C and D are
isomorphic if and only if the functors .HomC (C, −) and .HomC (D, −) are naturally
isomorphic.
Proof Suppose first that C and D are isomorphic objects in .C and let .φ ∈
HomC (D, C) be an isomorphism. As proved in Corollary 1.10.4, (1) we have a
natural transformation
defined by .τ (φ) = hφ : HomC (C, −) → HomC (D, −),
where . h X (f ) = HomC (D, f )(φ) = f ◦ φ for all .X ∈ Ob C and .f ∈
φ
HomC (C, X). We will prove that . hφ X is a set bijection for every .X ∈ Ob C. To
φ
this end, we define . μ X : HomC (D, X) → HomC (C, X) by . μφ X (g) = g◦φ −1
for any .g ∈ HomC (D, X). We will see that . μφ X is the inverse of . hφ X . Indeed,
1.10 Yoneda’s Lemma 75
αD
Hom (C, D) Hom (D, D)
(C, u) (D, u)
⇔ u ◦ αD (v) = αC (u ◦ v)
⇔ u = αC (u ◦ v).
Evaluating the above diagram at .1C ∈ HomC (C, C) and using .αD (v) = 1D we
obtain
⇔ v ◦ αC (1C ) = αD (v ◦ 1C )
⇔ v ◦ u = 1D ,
76 1 Categories and Functors
22 Note that .HomC (−, C) : Cop → Set is a covariant functor and therefore it is an object of the
category of presheaves on .C.
1.11 Exercises 77
1.11 Exercises
θ
X Y
f g
X Y
. θ
f
A B h f m g,
g h
C D
. m
1.8 Prove that any group regarded as a one-object category is isomorphic to its
opposite. Is the assertion still true for monoids?
1.9 Give examples to show that:
(a) arbitrary functors do not necessarily preserve or reflect monomorphisms
and epimorphisms;
(b) functors which reflect isomorphisms are not necessarily fully faithful.
1.10 Construct an example of:
(a) a full functor which is not surjective on objects/morphisms;
(b) a faithful functor which is not injective on objects/morphisms.
1.11 Is there a functor F : Grp → Grp such that F (G) is equal to the center of G
for all groups G?
1.12 Is there a functor F : Top → Set such that F (X) is equal to the set of
connected components of X for all topological spaces X?
1.13 Let F : C → D and G : D → C be functors such that GF is naturally
isomorphic to 1C . Prove that F is faithful.
1.14 Let F : C → D and G : D → C be two functors.
(a) If GF is faithful then F is faithful.
(b) If GF is full and G is faithful then F is full.
1.15 Let I be a small category and F : C → D a fully faithful functor. If
F : Fun(I, C) → Fun(I, D) is the induced functor defined in (1.36), show
that:
(a) F is fully faithful;
(b) if G, H : I → C are two functors such that F (G) and F (H ) are naturally
isomorphic, then G and H are naturally isomorphic.
1.16 Let U : Mon → Set be the forgetful functor and define U × U : Mon → Set
as follows:
(U × U )(M) = U (M) × U (M) for all M ∈ Ob Mon;
(U × U )(f )(m, m ) = (f (m), f (m )) for all f ∈ HomMon (M, N ), m,
m ∈ U (M).
Prove that γ : U × U → U , defined by γM = mM for all M ∈ Ob Mon, is
a natural transformation, where mM denotes the multiplication of the monoid
M.
1.17 Prove the interchange law between vertical and horizontal composition of
natural transformations, i.e., (δ ∗ γ ) ◦ (β ∗ α) = (δ ◦ β) ∗ (γ ◦ α), where
80 1 Categories and Functors
F H
G I
. P Q
If F, G : PO(X,
(a) Describe the functors between these two categories.
) → PO(Y, ) are two such functors, describe Nat F, G .
(b) Show that Fun PO(X, ), PO(Y, ) is isomorphic to a category asso-
ciated to a pre-ordered set.
1.20 Let V be a finite-dimensional vector space. Show that the functors
HomK M (V , −), V ∗ ⊗ − : K M → K M, are naturally isomorphic.
1.21 Prove that for any category C there exists a functor F : C → Set which is not
representable.
1.22 Prove that the contravariant power-set functor P c : Set → Set is repre-
sentable.
1.23 Let G ∈ Ob Grp. Prove that the functor FG : Grp → Set defined as follows
for any H ∈ Ob Grp, f ∈ HomGrp (G, H ):
F (f )(g) = f ◦ g, g ∈ FG (H )
is representable.
1.24 An object I of a category C is called injective if for any u ∈ HomC (A, I )
and any monomorphism m ∈ HomC (A, B) there exists a morphism v ∈
HomC (B, I ) such that the following diagram is commutative:
.
1.11 Exercises 81
Limits and colimits are fundamental and unifying concepts in category theory. Many
seemingly unrelated constructions from different fields of mathematics such as the
free product of groups (with amalgamation), the tensor product of (co)algebras
or the direct sum of modules are in fact special instances of these very general
concepts.
there exists a unique morphism .f : P → i∈I Pi in .C such that the following
diagram commutes for all .j ∈ I :
P
fj
f
i I Pi Pj
.
pj
We say that .C is a category with (finite) products or that .C has (finite) products if
there exists a product in .C for any (finite)
family of objects. If I is a finite set, then
we also write .P1 × . . . × Pn instead of . i=1, n Pi .
Coproducts
are the dual notion of products; that is, the coproduct of a family
.Xi i∈I of objects of a category .C is defined to be the product of the same family of
objects in the dual category .C op . This comes down to the following:
Definition 2.1.2 Let I be a set and . Qi i∈I a family of objects in a category .C. A
coproduct of the family .(Qi )i∈I is a pair . i∈I Qi , (qi )i∈I , where
(1) . i∈I Qi ∈ Ob C;
(2) .qj : Qj → i∈I Qi are morphisms in .C for all .j ∈ I ,
and for any other pair . Q, (fi )i∈I , where
(1) .Q ∈ Ob C;
(2) .fj : Qj → Q are morphisms in .C for all .j ∈ I ,
there exists a unique morphism .f : i∈I Qi → Q in .C such that the following
diagram commutes for all .j ∈ I :
qj
Qj i I Qi
f
fj
.
Q
We say that .C is a category with (finite) coproducts or that .C has (finite) coproducts
if there is a coproduct in .C for any (finite) family of objects.
Proposition 2.1.3 When it exists, the (co)product of a family of objects is unique
up to isomorphism.
Proof We start by proving the uniqueness up to isomorphism
of the product. Let
C be a category and consider two products . P , (pi )i∈I , respectively . P , (pi )i∈I ,
.
in .C of the same family of objects . Pi i∈I . Since . P , (pi )i∈I is assumed to be
a product, there exists a unique .f ∈ HomC (P , P ) such that for any .j ∈ I we
2.1 (Co)products, (Co)equalizers, Pullbacks and Pushouts 85
have
p j ◦ f = pj .
. (2.1)
Similarly, as . P , (pi )i∈I is also a product of the same family of objects, we obtain
a unique .g ∈ HomC (P , P ) such that for any .j ∈ I we have
pj ◦ g = p j .
. (2.2)
P
pj
f
pj
1P P Pj
g
pj
.
P
By composing the equality in (2.2) with f on the right and using (2.1) we obtain
(2.1)
pj ◦ (g ◦ f ) = pj ◦ f = pj
. (2.3)
for any .j ∈ I . Applying Definition 2.1.1 to the pair . P , (pi )i∈I , seen both as
a product and as the other pair, yields a unique .h ∈ HomC (P , P ) such that
.pj ◦ h = pj for any .j ∈ I . By the uniqueness of h we must have .h = 1P .
Moreover, since by (2.3) the map .g ◦ f also fulfills the above identity we obtain
.g ◦ f = 1P . In the same manner we obtain .f ◦ g = 1
P and therefore P and .P are
isomorphic.
The assertion regarding coproducts follows
by applying
the duality principle.
Indeed, assume now that . Q, (qi )i∈I and . Q, (q i )i∈I are coproducts in .C of the
op op
same family of objects . Qi i∈I . Then . Q, (qi )i∈I and . Q, (q i )i∈I are both
products in .C op of the family of objects . Qi i∈I . According to the above proof, there
exists an isomorphism .f op ∈ HomC op (Q, Q). Therefore, we have an isomorphism
.f ∈ HomC (Q, Q).
Proposition 2.1.4 Let .C be a category such that any two objects admit a
(co)product. Then, any non-empty finite family of objects in .C admits a
(co)product.
Proof Let .{X1 , X2 , . . . , Xn } be a non-empty family of objects in .C. We use
induction on n to construct the product of this family. If .n = 1 then the family
.{X} has a product given by the pair .(X, 1X ). Indeed, if . Y, f ∈ HomC (Y, X)
86 2 Limits and Colimits
then the unique morphism which makes the following diagram commutative is
precisely f :
Y
f
f
X X
.
1X
Z i.e., i f f i.
fi
f
Y Xi
. i (2.4)
Moreover, as . X, (πX , πk+1 ) is the product of the family .{Y, Xk+1 }, there
exists a unique .g ∈ HomC (Z, X) such that the following diagram is commuta-
tive:
(2.5)
with these properties. Indeed, assume there exists a .g ∈ HomC (Z, X) such
that
πk+1 ◦ g = fk+1
. and πj ◦ πX ◦ g = fj , for all j = 1, 2, . . . , k.
In particular, we have .πj ◦ πX ◦ g = fj for all .j = 1, 2, . . . , k and by the
uniqueness of the morphism which makes (2.4) commutative we obtain .f = πX ◦g .
Putting everything together, we obtain .πk+1 ◦ g = fk+1 and .πX ◦ g = f .
As g is the unique morphism for which (2.5) holds it follows that .g = g , as
desired.
The claim concerning coproducts follows by duality. Indeed, assume that .C is a
category such that any two objects admit a coproduct. Then, in .C op any two objects
admit a product and, according to the above proof, any non-empty finite family of
objects in .C op admits a product. Therefore, any non-empty finite family of objects
in .C admits a coproduct, as desired.
Examples 2.1.5
(1) The product of any family .(Xi )i∈I of objects
in Set, where I is a set, is given
by the corresponding
cartesian product
. i∈I Xi together with the canonical
projections . πj : i∈I Xi → Xj j ∈I . More precisely, we have
. Xi = {(xi )i∈I | xi ∈ Xi }, πj : Xi → Xj ,
i∈I i∈I
πj (xi )i∈I = xj for all j ∈ I.
where .·i denotes the group multiplication in .Gi . It can now be easily checked
as in the previous example that this is indeed the product in Grp of the family
. Gi
i∈I
.
88 2 Limits and Colimits
(3) The product of any family of objects in Top is given by the cartesian product
of the underlying sets endowed with the product topology together with the
canonical
projections. Let I be a set, .(Xi )i∈I a family of topological spaces
and . i∈I Xi , (πi )i∈I the product of the underlying sets. Then, each .πj
is obviously
continuous as the product topology is the coarsest topology
on . i∈I Xi for which all projections .πj : i∈I Xi → Xj are continuous.
Assume now that .X ∈ Ob Top and .pj ∈ HomTop (X, Xj ), .j ∈ I , is
a family of continuous maps. It can be easily seen that the unique map
.f : X → i∈I Xi defined in Example 2.1.5, (1) is continuous too. Indeed,
recall that the product topology is generated by the sub-basis .S = j ∈I Sj ,
where .Sj = {πj−1 (Uj ) | Uj open in Xj } for all .j ∈ J , and therefore in
order to prove continuity of f it will suffice to show that the inverse image
of each sub-basis element of . i∈I Xi is open [39, p. 103]. To this end, for all
−1 π −1 (U ) = (π ◦ f )−1 (U ) = p −1 (U ) and since .p is
.j ∈ I we have .f j j j j j
j j
continuous, the latter term is an open set, as desired.
(4) Products in Haus and KHaus, the categories of Hausdorff and compact
Hausdorff spaces respectively, are constructed as in Top. Indeed, it will suffice
to show that given a family of Hausdorff (resp. compact Hausdorff) spaces
.(Xi )i∈I , where I is a set, the product of the underlying spaces together with
product of this family in Cat. It can be easily seen that for any small category
.D and anyfamily of functors .UCi : D → Ci there exists a unique functor
.F : D → i∈I Ci such that .pCi ◦ F = UCi for all .i ∈ I ; more precisely, F is
defined as follows for all .D ∈ Ob D and .f ∈ HomD (D, D ):
F (D) = UCi (D) i∈I ,
. F (f ) = UCi (f ) i∈I .
(7) The category SiGrp of simple groups does not admit products or coproducts.
Indeed,
suppose this category admits products and let H , K be simple groups.
Let . X, (p, q) be the product in SiGrp of H and K. In particular, X is a sim-
ple group and .p : X → H , .q : X → K are group homomorphisms. Consider
now the pair . H, (IdH , 0K ) where .IdH is the identity homomorphism on H
while .0K : H → K denotes the group homomorphism defined by .0K (h) = 1K
for all .h ∈ H . By Definition 2.1.1 there exists a unique homomorphism of
groups .f : H → X such that the following diagram is commutative:
(9) In Set, the coproduct of a family .(Xi )i∈I is just its disjoint union, i.e., the
union of the sets .Xi = Xi × {i}. Thus, the coproduct of the family .(Xi )i∈I
is the pair . i∈I X i , (qi ) i∈I , where . i∈I Xi = {(x, i) | i ∈ I, x ∈ Xi }
and .qj : Xj → i∈I X i , .qj (x) = (x, j ) for all .j ∈ I . Indeed, given
a setQ together with a collection of maps .fj : Xj → Q, .j ∈ I , define
.f : X i → Q by considering .f (x, j ) = fj (x) for all .(x, j ) ∈ Xj ⊂
i∈I
i∈I Xi . Since each .(x, j ) lies inside a unique copy of .Xj , the following map
is well-defined:
(10) The coproduct of any family . Xi , τi i∈I of objects in Top is given by
the disjoint union of the underlying sets . i∈I Xi constructed in the
previous example endowed with the finest topology .τ for which all
maps .qj are continuous. Consider now .Q ∈ Ob Top and a family
of continuous maps .fj : Xj → Q, .j ∈ I . We will show that the
unique map .f : i∈I Xi → Q defined in the previous example is
continuous. To this end, let .U ⊆ Q be an open set. Note that since .τ
is the finest topology for which all .qj are continuous, in order to show
that .f −1 (U ) is an open set it will suffice to prove that .qj−1 f −1 (U )
is open for all .j ∈ I . Indeed, we have .qj−1 f −1 (U ) = (f ◦
qj )−1 (U ) = fj−1 (U ) and since each .fj is continuous, the desired conclusion
follows.
(11) For certain categories, such as Grp, the coproducts are more complicated
than the products and the constructions do not rely on the ones performed
in Set. This is basically because unions do not usually preserve oper-
ations (for instance, the union of an arbitrary family of groups is not
necessarily a group). In group theory, the construction which gives the
coproducts is called
the free product of groups (see [50, Chapter 11, p.
388]). Let . Gi i∈I be a family of groups, where I is a set, and denote by
. ∗i∈I Gi , (jk : Gk → ∗i∈I Gi , )k∈I the corresponding free product. Consider
where .ai0 = a and .aj = 0 for all .j = i0 . Indeed, consider the pair
. H, (fi )i∈I , where H is an abelian group and .fj : Aj → H are group
homomorphisms
for all .j ∈ I . Then, the unique homomorphism of groups
.f : i∈I iA → H which makes the following diagram commutative for all
.j ∈ I :
is given by .f (ai )i∈I = i∈I fi (ai ).1 Suppose now that .g : i∈I Ai → H
is another group homomorphism such that .g ◦ qj = fj for all .j ∈ I .
Then .(f − g) ◦ qj is the zero map from .Aj to H and thus the image
of .qj is contained in .ker(f − g) for all .j ∈ I . Now observe that
any element in . i∈I Ai is a sum of finitely many elements of the
form .qj (aj ). Therefore, since .ker(f − g) is a subgroup in . Ai , we
i∈I
obtain that .ker(f − g) = i∈I Ai , i.e., .f = g and the proof is now
finished.
An analogous description of coproducts holds for the category .R M of
modules over a ring R.
(13) Let PO.(X, ) be the category corresponding to the pre-ordered set .(X, ).
Then, the coproduct of a family .(xi )i∈I , if it exists, is just its supremum.
.
(1) .E ∈ Ob C;
(2) .p ∈ HomC (E, X) such that .f ◦ p = g ◦ p,
1 Note that the sum in the right-hand side contains only finitely many non-zero terms.
92 2 Limits and Colimits
We say that .C is a category with equalizers or that .C has equalizers if any pair of
morphisms in .C with the same domain and codomain has an equalizer.
Next we introduce coequalizers, the dual notion of equalizers.
Definition 2.1.7 A coequalizer of the morphisms f , .g ∈ HomC (X, Y ) is a pair
(Q, q), where
.
(1) .Q ∈ Ob C;
(2) .q ∈ HomC (Y, Q) such that .q ◦ f = q ◦ g,
and for any other pair .(Q , q ), where
(1) .Q ∈ Ob C;
(2) .p ∈ HomC (E , X) such that .f ◦ p = g ◦ p ,
there exists a unique .v ∈ HomC (Q, Q ) which makes the following diagram
commute:
We say that .C is a category with coequalizers or that .C has coequalizers if any pair
of morphisms in .C with the same domain and codomain has a coequalizer.
At this point we only state the uniqueness up to isomorphism of (co)equalizers.
The proof relies on the same idea used in proving Proposition 2.1.3. Furthermore,
this uniqueness result will follow as a special case of Proposition 2.2.9.
Proposition 2.1.8 When it exists, the (co)equalizer of two morphisms is unique up
to isomorphism.
2.1 (Co)products, (Co)equalizers, Pullbacks and Pushouts 93
Now notice that both maps .h1 , .h2 : E → E fulfill the above equality. Due to the
uniqueness of u we obtain .u = h1 = h2 , as desired.
For the second part we use duality. Indeed, .(E, p) is the coequalizer of the pair
of morphisms f , .g : X → Y in category .C if and only if .(E, pop ) is the equalizer of
the pair of morphisms .f op , .g op : Y → X in the category .C op . The conclusion now
follows from the first part of the proof and Proposition 1.4.3, (1).
Examples 2.1.10
(1) In Set the equalizer of two functions f , .g : X → Y is the pair .(E, i), where
.E = {x ∈ X | f (x) = g(x)} and .i : E → X is the canonical inclusion. Indeed,
(3) Let f , .g ∈ HomTop (X, Y ) and endow the set E constructed in the first example
with the subspace topology.2 This implies that the inclusion .i : E → X is a
continuous map and therefore .(E, i) is the equalizer of the pair of morphisms
.(f, g) in Top.
(4) Equalizers in Haus and KHaus are constructed as in Top. Indeed, assume first
that f , .g ∈ HomHaus (X, Y ) and consider .E = {x ∈ X | f (x) = g(x)} endowed
with the subspace topology. We only need to show that E is a Hausdorff space.
To this end, let .x1 , .x2 ∈ E such that .x1 = x2 . As X is a Hausdorff space, we can
find two disjoint neighborhoods of .x1 and .x2 in X, say .U1 and .U2 respectively.
By definition of the subspace topology, the sets .U1 ∩ E and .U2 ∩ E are open
in E and therefore neighborhoods in E of .x1 and .x2 , respectively. Furthermore,
we have .(U1 ∩ E) ∩ (U2 ∩ E) = (U1 ∩ U2 ) ∩ E = ∅, which proves that E is
Hausdorff.
Consider now u, .v ∈ HomKHaus (K, H ). We have already proved that E
together with the subspace topology is Hausdorff; we are left to show that E
is compact as well. It will suffice to prove that E is closed in X as any closed
subspace of a compact space is compact as well [39, Theorem 26.2]. To this end,
let .x0 ∈ X − E. As .f (x0 ) = g(x0 ) and Y is Hausdorff we can find two disjoint
open sets U and V such that .f (x0 ) ∈ U and .g(x0 ) ∈ V . Then .T = f −1 (U )
and .W = g −1 (V ) are open sets in X by continuity of f and g. Consequently
.S = T ∩ W is also an open set in X and .x0 ∈ S. Furthermore, S and E are
imply .f (x) ∈ U ∩ V = ∅.
To summarize, we have proved that for each .x0 ∈ X − E there exists an open
set S such that .x0 ∈ S ⊂ X − E. This shows that .X − E is contained in the
interior3 of .X − E, which allows us to conclude that .X − E = Int(X − E).
Hence .X − E is an open subset of X and therefore E is a closed set, as desired.
(5) Let f , .g ∈ HomSet (X, Y ) be two functions. Consider .R = {(f (x), g(x)) | x ∈
let .∼R be the equivalence relation on Y generated by R. Then the
X} and 4
pair . Y /∼R , π is the coequalizer of the maps f and g in Set, where .Y /∼R is
the set of equivalence classes of Y with respect to .∼R , while .π : Y → Y /∼R ,
.π(y) = y, for all .y ∈ Y , is the canonical projection. To start with, for any
by .v(y) = q(y) for all .y ∈ Y /∼R . First we will show that v is well-defined.
To this end, let y, .y ∈ Y such that .y = y . Then .y ∼R y and this implies that
defined as the union of all open sets contained in U [39, p. 95]. Moreover, we have .Int(U ) ⊂ U .
4 The equivalence relation generated by a binary relation R on a set Y (regarded as a subset of
there exists some positive integer n and .y0 , .y1 , . . . , yn ∈ Y such that .y = y0 ,
.y = yn and for any .i = 1, 2 . . . , n we have either .yi ∼ yi+1 or .yi+1 ∼ yi .
Furthermore, note that if .yi ∼ yi+1 then .yi = f (x0 ), .yi+1 = g(x0 ) for some
.x0 ∈ X, which implies .q(yi ) = q ◦ f (x0 ) = q ◦ g(x0 ) = q(yi+1 ). Therefore, we
obtain .q(y0 ) = q(y1 ) = . . . = q(yn ), which leads to .q(y) = q(y ). This shows
that v is well-defined. Obviously we have .v ◦ π(y) = q(y) for all .y ∈ Y . If
.v : Y /∼R → Q such that .v ◦ π = q then we easily obtain .v (y) = v ◦ π(y) =
q(y) = v ◦ π(y) = v(y) for all .y ∈ Y . Thus .v = v , which completes the proof.
(6) Let f , .g ∈ HomTop (X, Y ) and let .π : Y → Y /∼R be the canonical projection
constructed in the previous example. Now we endow .Y /∼R with the quotient
topology5 with respect to .π and we will show that . Y /∼R , π is the coequalizer
of the pair .(f, g) in Top. In particular, .π : Y → Y /∼R is a continuous map.
Consider now .Q ∈ Ob Top and another continuous map .q : Y → Q such that
.q ◦ f = q ◦ g. We only need to prove that the unique map .v : Y /∼R → Q
5 Let X be a topological space, Y a set and .f : X → Y a surjective map. The quotient topology
with respect to f is the finest topology on Y such that f is continuous. In other words, .U ⊆ Y is
open if and only if .f −1 (U ) is open in X [39, p. 138].
6 See, for instance, [38, Examples 6.4.17, 6.4.19].
96 2 Limits and Colimits
in Top of .(f, g), there exists a unique morphism .t ∈ HomTop Y /∼R , Q such
that .t ◦ π = q . Now the universal property of the Hausdorff quotient (see [40])
yields a unique .v ∈ HomHaus (H, Q ) such that .v ◦ u = t. If we put everything
together we obtain .v ◦ q = v ◦ u ◦ π = t ◦ π = q . We are left to show that v
is the unique morphism with this property. Indeed, assume there exists another
.v ∈ HomHaus (H, Q ) such that .v ◦ q = q ; then we have .(v ◦ u) ◦ π = q and
since t is the unique morphism with this property it follows that .v ◦ u = t and
therefore .v = v .
(8) Let f , .g ∈ HomGrp (G, H ) and consider .H = {f (x)g(x)−1 | x ∈ X} ⊆ H .
If we denote by .N = H ⊆U H U the normal subgroup generated by .H then
.(H /N, π ) is the coequalizer of the pair of morphisms .(f, g) in the category
i.e., .(H /N, π ) is the coequalizer of the pair of morphisms .(f, g) in Grp.
(9) Let f , .g ∈ HomAb (A, B). Then the map .(f − g) : A → B defined by .(f −
g)(a) = f (a) − g(a) for all .a ∈ A is a morphism of groups and, therefore,
the set .N = {f (a) − g(a) | a ∈ A} is a subgroup of B. It can be easily
seen,
using the same arguments as in the previous example, that . B/N, π is the
coequalizer of the pair of morphisms .(f, g) in Ab. Coequalizers in .R M have a
similar description.
(10) Let G be a non-trivial group and .G the associated category as described in
Example 1.2.2, (3). If x, .y ∈ G such that .x = y then the pair of morphisms
.(x, y) does not admit a (co)equalizer in .G. Indeed, this follows easily by
noticing that there are no elements .z ∈ G such that .xz = yz or .zx = zy.
.
Definition 2.1.11 Let .C be a pointed category and .f ∈ HomC (A, B). The
(co)equalizer of the pair of morphisms .(f, 0A, B ) is called the (co)kernel of f .
Example 2.1.12 By applying the above definition to the pointed categories Grp,
Ab and .R M, we recover the familiar algebraic kernel of a morphism defined as the
preimage of the neutral element. .
We say that .C is a category with pullbacks or that .C has pullbacks if any pair of
morphisms in .C with the same codomain has a pullback.
Definition 2.1.14 Let .C be a category and .f ∈ HomC (A, B), .g ∈ HomC (A, C).
A pushout of .(f, g) is a triple .(P , f , g ), where
(1) .P ∈ Ob C;
(2) .f ∈ HomC (C, P ), .g ∈ HomC (B, P ) such that .g ◦ f = f ◦ g,
and for any other triple .(P , f , g ), where
(1) .P ∈ Ob C;
(2) .f ∈ HomC (C, P ), .g ∈ HomC (B, P ) such that .g ◦ f = f ◦ g,
there exists a unique . ∈ HomC (P , P ) such that .f = ◦ f and .g = ◦ g .
98 2 Limits and Colimits
We say that .C is a category with pushouts or that .C has pushouts if any pair of
morphisms in .C with the same domain has a pushout.
As in the case of (co)products and (co)equalizers, both pullbacks and pushouts
are unique up to isomorphism:
Proposition 2.1.15 When it exists, the pullback (resp. pushout) of two morphisms
is unique up to isomorphism.
Examples 2.1.16
(1) In Set the pullback of two morphisms .f : B → A, .g : C → A is given by the
triple .(B ×A C, π C , π B ), where .B ×A C = {(b, c) ∈ B × C | f (b) = g(c)}
and .π C : B ×A C → C, .π B : B ×A C → B are given by
π C (b, c) = c,
. π B (b, c) = b, for all (b, c) ∈ B ×A C.
π C ◦ (x) = f (x),
.
π B ◦ (x) = g (x).
below commutative:
(2) Let .f ∈ HomGrp (B, A), .g ∈ HomGrp (C, A) and consider the set .B ×A C
constructed in the previous example endowed with the group structure given
by the direct product. Then, the two projections .π C : B ×A C → C, .π B : B ×A
C → B defined in the previous example are group homomorphisms and .(B×A
C, π C , π B ) is the pullback of the pair of morphisms .(f, g). Indeed, let .P ∈
Ob Grp, .f ∈ HomGrp (P , C), .g ∈ HomGrp (P , B) such that .f ◦ g =
g◦f . The only thing left to check is that the map . : P → B ×A C defined in
the previous example is a group homomorphism. To this end, for all x, .y ∈ P
we have .(xy) = (g (xy), f (xy)) = (g (x), f (x))(g (y), f (y)) =
(x)(y). Furthermore, .(1P ) = (g (1P ), f (1P )) = (1B , 1C ) and the
desired conclusion follows.
(3) Let .f ∈ HomTop (B, A), .g ∈ HomTop (C, A) and consider the set .B ×A C
constructed in the first example. Furthermore, we see .B × C as a topological
space with respect to the product topology and endow .B ×A C ⊆ B × C
with the subspace topology. Then . B ×A C, π C , π B is the pullback in Top
of the pair of morphisms .(f, g). Indeed, note first that both .π B = πB ◦ i
and .π C = πC ◦ i are compositions of continuous maps and are therefore
continuous, where .i : B×A C → B×C is the inclusion map and .πB , .πC denote
the projection maps. Consider now .P ∈ Ob Top and .f ∈ HomTop (P , C),
.g ∈ HomTop (P , B) such that .f ◦ g = g ◦ f . We are left to show that the
map . : P → B ×A C defined in the first example is continuous. To this end,
recall that the topology on .B ×A C is generated by the sets .π −1 −1
B (U ), .π C (V )
−1 −1
for all open sets .U ⊆ B and .V ⊆ C. In other words, the sets .π B (U ), .π C (V ),
where .U ⊆ B and .V ⊆ C are open sets, form a sub-basis for the topology
on .B ×A C. Therefore, in order to show that . is a continuous map it will
suffice to show that the preimage of any of these sets through . is an open set.
Indeed, for all open sets .U ⊆ B, .V ⊆ C we have
−1 ◦ π −1
.
−1
B (U ) = (π B ◦ ) (U ) = g
−1
(U ),
−1 ◦ π −1 −1
C (V ) = (π C ◦ ) (V ) = f
−1
(V ).
The conclusion now follows easily as both .f and .g are continuous maps.
100 2 Limits and Colimits
(4) Assume R is a commutative ring and .R M is the category of left modules over
R. Let .f ∈ HomR M (B, A), .g ∈ HomR M (C, A) and consider the submodule
.B ×A C = {(b, c) ∈ B × C | f (b) = g(c)} of the direct product .B × C. Then
. B ×A C, π C , π B is the pullback in .R M of the pair of morphisms .(f, g),
Indeed, since both maps .f and .g are R-linear, for all r, .s ∈ R and x, .y ∈ P
we have
(rx + sy) = g (rx + sy), f (rx + sy)
.
= rg (x) + sg (y), rf (x) + sf (y)
= rg (x), rf (x) + sg (y), sf (y)
= r g (x), f (x) + s g (y), f (y)
= r(x) + s(y),
as desired.
(5) Let .f ∈ HomSet (A, B), .g ∈ Hom Set (A, C) and consider the disjoint union
of the sets B and C, denoted by .B C = {(b, 0) | b ∈ B} ∪ {(c, 1) | c ∈ C},
together with the corresponding inclusion maps .j0 : B → B C, .j1 : C →
B C defined by .j0 (b) = (b, 0), .j1 (c) = (c, 1) for all .b ∈ B, .c ∈ C. Define
.R = { (f (a), 0), (g(a), 1) | a ∈ A} ⊆ B C × B C and let .∼R be
the equivalence relation on .B C generated R. Then the pushout of the
by
pair .(f, g) is
given
by the quotient C /∼R together with the maps
set . B
.f : C → B C /∼R , .g : B → B C /∼R defined as follows:
f = π ◦ j1 ,
. g = π ◦ j0 ,
where .π : B C → B C /∼R is the canonical projection. Indeed, first
note that for all .a ∈ A we have .g ◦ f (a) = π(f (a), 0) = π(g(a), 1) =
f ◦ g(a). Consider now .P ∈ Ob Set and .f ∈ HomSet (C,
P ), .g ∈
HomSet (B, P ) such that .g ◦ f = f ◦ g and define .ψ : B C → P as
follows:
Then .a ∈ A we have .ψ f (a), 0 = g f (a) = f g(a) =
for all
1 . The universal
ψ g(a), property of the quotient set yields a unique map
. : B C /∼R → P such that . ◦ π = ψ. Furthermore, for all .b ∈ B and
.c ∈ C we have
(7) Let .f ∈ HomGrp (A, B), .g ∈ HomGrp (A, C) and let .B C be the
free product of the groups B and C together with the corresponding group
homomorphisms .iB : B → B C and .iC : C → B C. Furthermore, we let
.N = {iB (f (a))
−1 i (g(a)) | a ∈ A} and denote by N the normal subgroup of
C
.B C generated by .N . Then the quotient group .(B C)/N together with
that .N ⊆ N, we have
(f ◦ g)(a) = π iC (g(a)) = π iB (f (a)) = (g ◦ f )(a) for all a ∈ A.
.
(2.7)
We show first that .N ⊆ ker . Indeed, for all .a ∈ A we have
−1
−1
. iB (f (a)) iC (g(a)) = iB (f (a) iC (g(a)
= g (f (a))−1 f (g(a)) = 1.
Therefore, .N ⊆ ker and the universal property of the quotient groups yields
a unique group homomorphism . : (B C)/N → P such that . ◦ π = .
We obtain
◦ f = ◦ π ◦ iC = ◦ iC = f ,
.
◦ g = ◦ π ◦ iB = ◦ iB = g .
We are left to show that . is the unique group homomorphism with this
property. Assume . : (B C)/N → P is another group homomorphism
such that . ◦ f = f and . ◦ g = g . This leads to . ◦ π ◦ iC = f
and . ◦ π ◦ iB = g and since . is the unique group homomorphism which
makes diagram (2.7) commute, we obtain . ◦ π = . Now by the universal
property of the quotient groups, . is the unique group homomorphism with
this property and therefore . = .
(8) Assume R is a commutative ring and .R M is the category of left modules over
R. Let .f ∈ HomR M (A, B), .g ∈ HomR M (A, C) and consider the submodule
.S = {(f (a), −g(a)) | a ∈ A} of .B × C. Then the triple . (B × C)/S, f , g
χ : B × C → P ,
. χ (b, c) = g (b) + f (c),
We are left to prove the uniqueness of .. Let .Υ : (B × C)/S → P such that
.Υ ◦ g = g and .Υ ◦ f = f . To this end, for all .(b, c) ∈ B × C, we have
.Υ (b, c) = Υ (b, 0) + Υ (0, c)
= Υ ◦ g (b) + Υ ◦ f (c)
= g (b) + f (c)
= (b, c) .
(9) Let PO.(X, ) be the category corresponding to the pre-ordered set .(X, )
and a, b, .c ∈ X such that .a ≤ b and .a ≤ c. If it exists, the pushout of the
above maps is some element .p ∈ X satisfying:
104 2 Limits and Colimits
• .b ≤ p and .c ≤ p;
.
τ (x) = y, τ (y) = x,
.
ψx (x) = ψx (y) = x,
ψy (x) = ψy (y) = y.
The next two results highlight the different connections existing between
monomorphisms (resp. epimorphisms) and pullbacks (resp. pushouts) and will
be useful in the sequel.
Proposition 2.1.17 Let .C be a category.
(1) Let . X, u ∈ HomC (X, A), v ∈ HomC (X, C) be the pullback of the pair of
morphisms .f ∈ HomC (A, B) and .g ∈ HomC (C, B). If g is a monomorphism
then u is also a monomorphism.
(2) Let . Y, u ∈ HomC (B, Y ), v ∈ HomC (C, Y ) be the pushout of the pair of
morphisms .f ∈ HomC (A, B) and .g ∈ HomC (A, C). If g is an epimorphism
then u is also an epimorphism.
Proof .(1) Let .Y ∈ Ob C and assume there exists .α, .β ∈ HomC (Y, X) such that
u ◦ α = u ◦ β. In particular, this leads to .f ◦ u ◦ α = f ◦ u ◦ β and since .f ◦ u = g ◦ v
.
we also have .g ◦ v ◦ α = g ◦ v ◦ β.
.
2.2 (Co)limit of a Functor. (Co)complete Categories 105
This shows that .(A, 1A , 1A ) is the pullback of the pair of morphisms .(f, f ).
Conversely, assume that .(A, 1A , 1A ) is the pullback of the pair of morphisms
.(f, f ) and let u, .v ∈ HomC (P , A) such that .f ◦ u = f ◦ v. Hence, we have a
desired conclusion.
concepts of limit and colimit, which unify all the above. We start by introducing the
following:
Definition 2.2.1 Let .F : I → C be a functor.7 A cone on F consists of the
following:
(1) .C ∈ Ob C;
(2) for every .i ∈ Ob I , a morphism .si ∈ HomC (C, F (i)),
such that for any morphism .d ∈ HomD (i, j ), the following diagram is commuta-
tive:
The object C is called the vertex of the cone. If I is a small category and .C ∈ Ob C,
we denote by
.Cone(F, C) = { si ∈ HomC (C, F (i)) | C, (si )i∈Ob I is a cone on F }
i∈Ob I
The object C is called the vertex of the cocone. If I is a small category and .C ∈ Ob C,
we denote by
Cocone(F, C) = { ti ∈ HomC (F (i), C) i∈Ob I | C, (ti )i∈Ob I is a cocone on F }
.
. (2.8)
. (2.9)
108 2 Limits and Colimits
The cones on F together with morphisms between them as defined above form
a category denoted by .C(F ).
(2) A morphism between two cocones . C, (ti )i∈Ob I and . C, (ui )i∈Ob I on F is a
morphism .f ∈ HomC (C, C) such that the following diagram is commutative
for any .i ∈ Ob I :
pair .(f, 1∗ ) consisting of a morphism .f ∈ HomC (C, C ) and the identity morphism
.1∗ on .∗ such that the following diagram commutes:
where .1F denotes the identity natural transformation on the functor F . This leads to
the following equality between natural transformations: .α ◦ (f ) = α. Hence, for
all .i ∈ Ob I we
have .αi ◦ f = αi i.e.,
f is a morphism in .C(F ) between the cones
. C, (αi )i∈Ob I and . C , (α )i∈Ob I . We can now define two functors .U : C(F ) →
i
( ↓ TF ) and .V : ( ↓ TF ) → C(F ) as follows:
where .1F denotes the identity natural transformation on the functor F . This leads
to the following equality between natural transformations: . (f ) ◦ α = α . Hence,
for all .i ∈ Ob
I we have .f ◦ αi = αi , i.e., f is a morphism in .CO(F ) between
the cocones . C, (αi )i∈Ob I and . C , (αi )i∈Ob I . We can now define two functors
.U : CO(F ) → (TF ↓ ) and .V : (TF ↓ ) → CO(F ) as follows:
U C, (αi )i∈Ob I = (∗, α, C),
. U (f ) = (1∗ , f ),
V (∗, α, C) = C, (αi )i∈Ob i , V (1∗ , f ) = f.
110 2 Limits and Colimits
functor is just an object of .C. Furthermore, any morphism between two such
2.2 (Co)limit of a Functor. (Co)complete Categories 111
follows:
F (A1 ) = X,
. F (A2 ) = Y, F (u) = f, F (v) = g.
A
cone on F consists of an object .C ∈ Ob C and morphisms
.s ∈ HomC (C, X), t ∈ HomC (C, Y ) such that the following diagrams
are commutative:
functor .F : I → C as follows:
F (A1 ) = X,
. F (A2 ) = Y, F (A3 ) = Z,
F (u) = f ∈ HomC (Z, X), F (v) = g ∈ HomC (Z, Y ).
A
cocone on F consists of an object C together with
morphisms
.s ∈ HomC (X, C), t ∈ HomC (Y, C), l ∈ HomC (Z, C) such that the
112 2 Limits and Colimits
functor .F : J → C as follows:
F (B1 ) = X,
. F (B2 ) = Y, F (B3 ) = Z,
F (u) = f, F (v) = g.
It can be easily seen that the limit of the functor F defined above, if it exists, is
nothing but the pullback of the pair of morphisms .f : X → Z, .g : Y → Z in
.C. .
Example 2.2.11 The category .Set(⊆) defined in Example 1.2.2, (2) is obviously
not complete, as we have already established in Example 1.3.10, (6) that it has
no final objects (i.e., the empty functor from the empty category to .Set(⊆) does
not have a limit). However, .Set(⊆) is cocomplete. To this end, consider a functor
.H : I → Set(⊆) where I is a small category. Then . Z, (uH (i), Z )i∈Ob I is the
colimit of H , where .Z = i∈Ob I H (i). Recall that for all sets A, B such that
.A ⊆ B, we denote by .uA, B the unique morphism in .Set(⊆) from A to B. With this
in mind, it is straightforward
to see that . Z,
(uH (i), Z )i∈Ob I is a cocone on H .
Assume now that . W, (uH (i), W )i∈Ob I is another cocone on H . In particular,
as we have morphisms .uH (i), W : H (i) → W in .Set(⊆), it follows that .H (i) ⊆ W
for all .i ∈ Ob I . This implies that .Z = i∈Ob I H (i) ⊆ W and we have a unique
morphism in .Set(⊆) between Z and W , namely .uZ, W . Furthermore, for all .i ∈ Ob I
2.2 (Co)limit of a Functor. (Co)complete Categories 113
we have .uZ, W ◦ uH (i), Z = uH (i), W . This shows that .uZ, W is the unique morphism
in .Set(⊆) which makes the following diagram commutative for all .i ∈ Ob I :
Hence, . Z, (uH (i), Z )i∈Ob I is the colimit of H . .
Lemma 2.2.12 Let .G : I → C be a functor and let . L, i∈Ob I be
(pi )op a cone on G.
Then . L, (pi )i∈Ob I is the limit of G if and only if . L, (pi )i∈Ob I is the colimit
of the dual functor .Gop : I op → C op .
Proof Assume that . L, (pi )i∈Ob I is the limit of G and, furthermore, consider a
op
cocone . M, qi ∈ HomC op (Gop (i), M) i∈Ob I on .Gop . Lemma 2.2.3 implies
that . M, qi ∈ HomC (M, G(i)) i∈Ob I is a cone on G and since . L, (pi )i∈Ob I
is its limit, there exists a unique .f ∈ HomC (M, L) such that the following diagram
is commutative for all .i ∈ I :
. (2.10)
op op
In particular, this implies that we also have .f op ◦op pi = qi for all .i ∈ I , i.e., the
following diagram is commutative:
Moreover, the uniqueness of the morphism .f op which makes the above diagram
commutative follows from the uniqueness of the morphism
which makes
diagram
op
(2.10) commutative. Therefore, we can conclude that . L, (pi )i∈Ob I is the colimit
of .Gop .
As an easy consequence of Theorem 2.2.6 we obtain the following:
Corollary 2.2.13 Let I be a small category and .F : I → C a functor. Then:
(1) F has a limit of and only if the comma category .( ↓ TF ) has a final object;
114 2 Limits and Colimits
(2) F has a colimit if and only if the comma category .(TF ↓ ) has an initial
object.
We record here, for further use, the following useful results which generalize
Proposition 2.1.9:
Proposition 2.2.14 Let .F : I → C be a functor and .C ∈ Ob C.
(1) If . lim F, (pi )i∈Ob I is the limit of F and f , .g ∈ HomC (C, lim F ) such that
.pi ◦ f = pi ◦ g for all .i ∈ Ob I then .f = g.
(2) If . colim F, (qi )i∈Ob I is the colimit of .F : I → C and f , .g ∈
HomC (colim F, C) such that .f ◦ qi = g ◦ qi for all .i ∈ Ob I then .f = g.
Proof (1) To start with, note that . C, (pi ◦ f )i∈Ob I is a cone onthe functor F , i.e.,
.F (d) ◦ pi ◦ f = pj ◦ f for any .d ∈ HomI (i, j ). Indeed, since . lim F, (pi )i∈Ob I
is the limit of the functor F and in particular a cone on F , the following diagram is
commutative:
As both morphisms f , .g ∈ HomC (C, lim F ) render the above diagram commuta-
tive, the desired conclusion follows.
op
(2) Lemma 2.2.12 implies that . colim F, (qi )i∈Ob I is the limit of the functor
.F
op : I op → C op and moreover, we have .q ◦op f op = q op ◦op g op for all .i ∈ Ob I ,
op
i i
where .◦op denotes the composition in .C op . The first part of the proof implies .f op =
g op and therefore .f = g, as desired.
Lemma 2.2.15
Let F , .G : J → C be two functors,
where J is a small category,
and
denote by . L, (pj : L → F (j ))j ∈Ob J and . C, (qj : F (j ) → C)j ∈Ob J the limit
and colimit, respectively, of F . If .α : F → G is a natural transformation, then:
2.2 (Co)limit of a Functor. (Co)complete Categories 115
(1) . L, (αj ◦ pj : L → G(j ))j ∈Ob J is a cone on G. Furthermore,
if .α is a natural
isomorphism then . L, (αj ◦ pj : L → G(j ))j ∈Ob J is the limit of G.
(2) If .α is a natural isomorphism then . C, (qj ◦ αj−1 : G(j ) → C)j ∈Ob J is the
colimit of G.
Proof Given that in most of the previous proofs we have worked mainly with limits,
here we will prove the second
assertion concerning colimits.
(2) First we show that . C, (qj ◦ αj−1 : G(j ) → C)j ∈Ob J is a cocone on G.
To this end, consider .d ∈ HomJ (i, l); we need to prove the commutativity of the
following diagram:
. (2.11)
Since . C, (qj : F (j ) → C)j ∈Ob J is in particular a cocone on F , the following
diagram is commutative:
. (2.12)
. (2.13)
which proves that (2.11) holds and therefore . C, (qj ◦ αj−1 : G(j ) → C)j ∈Ob J is
indeed a cocone . C , (tj : G(j ) → C )j ∈Ob J
on G. Consider now another cocone
on G. Then . C , (tj ◦ αj : F (j ) → C )j ∈Ob J is a cocone on F . Indeed, for all
.d ∈ HomJ (i, l) we have
(2.13)
tl ◦ αl ◦ F (d) = tl ◦ G(d) ◦ αi = ti ◦ αi ,
.
where in the last equality we used the fact that . C , (tj : G(j ) → C )j ∈Ob J is a
cocone on G.
Now since . C, (qj : F (j ) → C)j ∈Ob J is the colimit of F , there exists a unique
.f ∈ HomC (C, C ) such that the following diagram is commutative for all .j ∈ Ob J :
Thus .f ∈ HomC (C, C ) is the unique morphism such that .f ◦ (qj ◦ αj−1 ) = tj ,
i.e., the unique morphism which makes the following diagram commutative for all
.j ∈ Ob J :
This proves that . C, (qj ◦ αj−1 : G(j ) → C)j ∈Ob J is the initial object in the
category of cocones on G, as desired.
.(1) To start with, note that .α
op : Gop → F op is also a natural isomorphism
−1
by Proposition 1.8.3. Furthermore, . α op : F op → Gop is again a natural
op
isomorphism (see Example 1.7.2, (6)). Lemma 2.2.12 implies that . L, (pj )j ∈Ob J
op
is the colimit of .F op and using part .(2) proved above we obtain that . L, (pj ◦op
op op op
αj )j ∈Ob J is the colimit of .Gop . Now since we have .pj ◦op αj = (αj ◦ pj )op ,
the conclusion follows by Lemma 2.2.12.
2.2 (Co)limit of a Functor. (Co)complete Categories 117
(Co)limit as a Functor
lim (F ) = lim F,
. lim (α) = α
for all functorsF, G : I → C and all natural transformations α : F → G,
where . lim F, (pi )i∈Ob I and . lim G, (si )i∈Ob I are the limits of F and G respec-
tively and .α ∈ HomC (lim F, lim G) is the unique morphism which makes the
following diagram commute for all .i ∈ Ob I :
Proof Before
going into the proof,
we point out that in light of Lemma 2.2.15, (1)
the pair . lim F, (αi ◦ pi )i∈Ob I is a cone on G. Therefore, the unique morphism .α ∈
HomC (lim F, lim G) which makes the diagram above commutative for all .i ∈ Ob I
exists by virtue of Definition 2.2.7.
Now let .F : I → C be a functor and .1F the identity natural transformation. Then,
.lim(1F ) is the unique morphism in .HomC (lim F, lim F ) such that the following
lim
As .1lim F makes the same diagram commutative we obtain .lim(1F ) = 1lim F . This
leads to .lim(1F ) = 1lim(F ) , as desired.
Consider now functors F , G, .H : I → C and natural transformations .α : F → G,
.β : G → H . The proof will be finished once we show that .lim(β ◦ α) = lim(β) ◦
lim(α). To this end, let . lim F, (pi )i∈Ob I , . lim G, (si )i∈Ob I , . lim H, (ti )i∈Ob I
be the limits of F , G and H , respectively. Recall that .β ◦ α, .β, .α are the unique
9 Note that defining the (co)limit functor requires an arbitrary choice of a limit for each functor.
i.e., ti ◦ β ◦ α = (β ◦ α)i ◦ pi , .
. (2.14)
si ◦ α = αi ◦ pi , . (2.15)
ti ◦ β = βi ◦ si . (2.16)
(2.16) (2.15)
ti ◦ β ◦ α = βi ◦ si ◦ α = βi ◦ αi ◦ pi = (β ◦ α)i ◦ pi for all i ∈ Ob I.
.
Now since .β ◦ α is the unique morphism for which (2.14) holds, we obtain .β ◦ α =
β ◦ α, and the proof is now complete.
For the sake of completeness we record below the dual result concerning colimits
and leave the proof to the reader:
Theorem 2.2.17 Let I be a small category and .C a cocomplete category. Then
colim : Fun(I, C) → C defined below is a functor:
.
colim(F ) = colim F,
. colim(α) = α
for all functorsF, G : I → C and all natural transformations α : F → G,
where . colim F, (qi )i∈Ob I and . colim G, (ti )i∈Ob I are the colimits of F and G
respectively and .α ∈ HomC (colim F, colim G) is the unique morphism which
makes the following diagram commute for all .i ∈ Ob I :
.
2.3 (Co)limit as a Representing Pair 119
In this section we show that the (co)limit of a functor arises as the representing pair
of a certain functor. Throughout, I is a small category and .F : I → C is a functor.
We can define two functors, .Cocone(F, −) : C → Set and .Cone(F, −) : C →
Set, as follows:
This comes down to . ti ◦f i∈Ob I = ui i∈Ob I . In other words, there exists a unique
.f ∈ HomC (C , C) such that .Cone(F, f ) (ti )i∈Ob I = (ui )i∈Ob I and it follows
by Corollary 1.8.2 that . C, ti ∈ HomC (C, F (i)) i∈Ob I is the representing pair
of the contravariant functor .Cone(F, −) : C → Set.
The converse follows by similar arguments and is left to the reader.
120 2 Limits and Colimits
Hom Cocone
. (2.17)
which shows that diagram (2.17) is indeed commutative. We are left to show
that each .αD is an isomorphism. To this end, recall that since . C, ti ∈
HomC (F (i), C) i∈Ob I is the colimit of F , for any . vi i∈Ob I ∈ Cocone(F, D)
there exists a unique .v ∈ HomC (C, D) such that the following diagram is
commutative for all .i ∈ Ob I :
. (2.18)
2.4 (Co)limits by (Co)equalizers and (Co)products 121
αD ◦ βD (vi )i∈Ob I = αD (v) = v ◦ ti i∈Ob I = vi i∈Ob I ,
.
βD ◦ αD (f ) = βD (f ◦ ti )i∈Ob I = f,
which shows that .βD is the inverse of .αD . Therefore, .α is a natural isomorphism.
This shows that the functor .Cocone(F, −) : C → Set is representable, as desired.
Suppose now that . C, ti ∈ HomC (F (i), C) i∈Ob I is the representing pair of
the functor .Cocone(F, −) : C → Set and consider .D ∈ Ob C and . vi i∈Ob I ∈
Cocone(F, D). Using Proposition 1.7.7, there exists a unique .f ∈ HomC (C, D)
such that
Cocone(F, f ) (ti )i∈Ob I = vi i∈Ob I .
.
This comes down to . f ◦ ti i∈Ob I = vi i∈Ob I . In other words, there exists a unique
.f ∈ HomC (C, D) such that the following diagram commutes for all .i ∈ Ob I :
Therefore, . C, (ti )i∈Ob I is the colimit of F .
(2) By Lemma 2.2.12, a cone . C, si ∈ HomC (C, F (i)) i∈Ob I on F is the
op
limit of F if and only if . C, si ∈ HomC op (F (i), C) i∈Ob I is the colimit of the
op
dual functor .F op . Using .1) it follows that . C, si ∈ HomC op (F (i), C) i∈Ob I
op
is the colimit of .F op if and only if . C, si ∈ HomC op (F (i), C) i∈Ob I is the
representing pair of the functor .Cocone(F op , −) : C op→ Set and by Lemma 2.3.1
this is equivalent to . C, si ∈ HomC (C, F (i)) i∈Ob I being the representing pair
of the functor .Cone(F, −) : C → Set.
(Co)products and (co)equalizers are perhaps the most important among the special
cases of (co)limits. This is due to the fact that all small (co)limits can be constructed
122 2 Limits and Colimits
We are left to prove that .Img ⊆ lim F . To this end, notice that .g(c) = sk (c) k∈Ob I
for all .c ∈ C. Moreover, since .F (f ) ◦ si = sj for any .f ∈ HomI (i, j ) we get
.F (f ) si (c) = sj (c) for all .c ∈ C. Thus .g(c) = sk (c) ∈ lim F for all
k∈Ob I
.c ∈ C. .
The previous example shows that the limit of a functor .F : I → Set, where
I is a small category, can be constructed as a subset of the product . i∈ObI F (i).
This suggests that equalizers are being used in order to construct the limit. The next
theorem shows that this method can be generalized to arbitrary categories allowing
for the construction of small limits out of products and equalizers.
Theorem 2.4.2 A category .C is (co)complete if and only if it has (co)products and
(co)equalizers.
2.4 (Co)limits by (Co)equalizers and (Co)products 123
Proof We will only prove the assertion regarding completeness and leave the (dual)
one about cocompleteness to the reader. Obviously, if a category is complete then
it has products and equalizers, as shown in Example 2.2.8, (2) and (3). Conversely,
assume .C is a category with products and equalizers and let .F : I → C be a functor,
where I is a small category. For any morphism f in I we will denote by .d(f )
the domain of f and by .c(f ) the codomain of f ; in other words we have .f ∈
HomI (d(f ), c(f )). We start by constructing
the products in .C of the families of
objects .(F (i))i∈ObI and . F (c(f )) f ∈Hom (d(f ), c(f )) , respectively:
I
. F (i), (ui )i∈Ob I , F (c(f )), (vc(f ) )f ∈HomI (d(f ), c(f )) .
i f
As . f F (c(f )), (vc(f ) )f ∈HomI (d(f ), c(f )) is the product in .C of the fam-
ily of objects . F (c(f )) f ∈Hom (d(f ), c(f )) , there exists a unique morphism
I
.α : i F (i) → f F (c(f )) in .C such that the following diagram is commutative
for all .g ∈ HomI (d(g), c(g)):
. (2.19)
Similarly, there exists a unique morphism .β : i F (i) → f F (c(f )) in .C such
that the following diagram is commutative for all .g ∈ HomI (d(g), c(g)):
(2.20)
124 2 Limits and Colimits
Now consider .(L, l) to be the equalizer in .C of the pair of morphisms .(α, β). The
complete picture is captured by the following diagram:
We willprove that . L, (pi = ui ◦ l)i∈Ob I is the limit of the functor F . First we
prove that . L, (pi )i∈Ob I is a cone on F . Indeed, if .r ∈ HomI (d(r), c(r)) we have
(2.20) (2.19)
F (r) ◦ pd(r) = F (r) ◦ ud(r) ◦ l = vc(r) ◦ β ◦ l = vc(r) ◦ α ◦ l = uc(r) ◦ l = pc(r) .
.
Moreover, consider another cone . M, (qi )i∈ObI on F . Since . i F (i), (ui )i∈Ob I
is the product in .C
of the family of objects .(F (i))i∈ObI , there exists a unique
morphism .q : M → i F (i) in .C such that for any .j ∈ Ob I we have
. (2.21)
(2.19) (2.21)
vc(r) ◦ α ◦ q = uc(r) ◦ q = qc(r) = F (r) ◦ qd(r)
.
(2.21) (2.20)
= F (r) ◦ ud(r) ◦ q = vc(r) ◦ β ◦ q ,
where in the third equality we used the fact that . M, (qi )i∈Ob I is a cone on F .
Therefore we have .vc(r) ◦ α ◦ q = vc(r) ◦ β ◦ q for any .r ∈ HomI (d(r), c(r))
and according to Proposition 2.2.14, (1) we obtain .α ◦ q = β ◦ q . Since .(L, l) is
the equalizer of the pair of morphisms .(α, β) in .C we obtain a unique morphism
.q : M → L such that
l ◦ q = q .
. (2.22)
2.4 (Co)limits by (Co)equalizers and (Co)products 125
It turns out that q is the unique morphism in .C which makes the following diagram
commute for all .j ∈ Ob I :
(2.22) (2.21)
Indeed, for any .j ∈ Ob I we have .pj ◦q = uj ◦l ◦ q = uj ◦q = qj . Finally,
we are left to prove the uniqueness of q. To this end, assume .q ∈ HomC (M, L)
is another morphism such that .pj ◦ q = qj for all .j ∈ Ob I . Hence, we obtain
.uj ◦ l ◦ q = qj for all .j ∈ Ob I and since .q is the unique morphism in .C which
(2.23)
colim F =
. F (i) ,
∼R
i∈I
where . i∈I F (i) denotes the quotient set by the equivalence relation .∼R and
∼R
.π : i∈Ob I F (i) → i∈I F (i) is the associated quotient function.
∼R
We start by proving that .(colim F, (ri )i∈Ob I ) is a cocone on F . To this end, let
.d ∈ HomI (i, j ); then, for any .x ∈ F (i) we have
(2.23)
rj ◦ F (d)(x) = π ◦ qj ◦ F (d)(x)=π ◦ (F (d)(x), j ) = π(x, i)=π ◦ qi (x)=ri (x),
.
Hence
.(colim F, (ri )i∈Ob I ) is indeed a cocone on F . Consider now another cocone
.C, (ti )i∈Ob I on F . Definition 2.1.2 yields a unique .ψ ∈ HomSet i∈I F (i), C
which renders the following diagram commutative for all .j ∈ Ob I :
. (2.24)
Let .(x, i), .(y, j ) ∈ i∈Ob I F (i) such that .(x, i) ∼R (y, j ). Thus we have either
.(x, i) R (y, j ) or .(y, j ) R (x, i). In the first case, there exists some .f ∈ HomI (i, j )
(2.24) (2.24)
ψ(x, i) = ψ ◦ qi (x) = ti (x) = tj ◦ F (f )(x) = tj (y) = ψ ◦ qj (y) = ψ(y, j )
.
where in the third equality we used the fact that . C, (ti )i∈Ob I is a cocone on F .
On the other hand, if .(y, j ) R (x, i) there exists some .g ∈ HomI (j, i) such that
.F (g)(y) = x. This leads to
(2.24) (2.24)
ψ(y, j ) = ψ ◦ qj (y) = tj (y) = ti ◦F (g)(y) = ti (x) = ψ ◦qi (x) = ψ(x, i),
.
where the third equality follows from the fact that . C, (ti )i∈Ob I is a cocone on F .
Putting all together, we proved that .(x, i) ∼R (y, j ) implies .ψ(x,
i) = ψ(y, j ).
Therefore, by the universal property of the quotient set . i∈I F (i) there exists
∼R
2.5 (Co)limit Preserving Functors 127
a unique map .ϕ : i∈I F (i) → C such that the following diagram commutes:
∼R
. (2.25)
Now it can be easily seen that .ϕ : i∈I F (i) → C is the unique morphism
∼R
which makes the following diagram commutative for all .i ∈ Ob I :
(2.25) (2.24)
ϕ ◦ ri = ϕ ◦ π ◦ qi = ψ ◦ qi = = ti .
.
.
. BilM, N ( Ai ), (BilM, N (pj ) : BilM, N ( Ai ) → BilM, N (Aj ))j ∈I
i∈I i∈I
is the product in Ab of the family BilM, N (Ai ) i∈I . Denote by
. BilM, N (Ai ), (πj : BilM, N (Ai ) → BilM, N (Aj ))j ∈I
i∈I i∈I
the product in Ab of the family BilM, N (Ai ) i∈I . Again by Example 2.1.5, (2) we
know that the underlying set of i∈I BilM, N (Ai ) is the cartesian product of the
s and π is the j -th projection. Now define ψ : Bil
Bil
M, N (Aj ) j
M, N ( i∈I Ai) →
i∈I Bil M, N (Ai ) by ψ(α) = p i ◦ α i∈I
for all α ∈ BilM, N ( i∈I A i ). It can be
easily seen that for all i ∈ I we have pi ◦ α ∈ BilM, N (Ai ), which shows that ψ
is well-defined. Furthermore, ψ is bijective. Indeed, if α, β ∈ BilM, N ( i∈I Ai )
such that ψ(α) = ψ(β), we obtain pi ◦ α = pi ◦ β for all i ∈ I . Now Propo-
(1) implies α = β, which shows that ψ is injective. Consider now
sition 2.2.14,
(ui )i∈I ∈ i∈I BilM, N (Ai ), where uj ∈ BilM, N (Aj ) for all j ∈ I . Then ψ(u) =
(ui )i∈I , where u ∈ BilM, N ( i∈I Ai ) is defined by u(m, n) = (ui (m, n))i∈I for
2.5 (Co)limit Preserving Functors 129
Indeed, for all α ∈ BilM, N ( i∈I Ai ) and j ∈ I , we have
(πj ◦ ψ)(α) = πj (pi ◦ α)i∈I = pj ◦ α = BilM, N (pj )(α).
.
while j : Q → BilM, N (A) denotes the inclusion. Now define ϕ : BilM, N (E) → Q
by ϕ(v) = i ◦ v for all v ∈ BilM, N (E). Consider w ∈ Q, i.e., w : M × N → A
is a bilinear map such that f ◦ w = g ◦ w. Then, for all m ∈ M, n ∈ N we
have f (w(m, n)) = g(w(m, n)), which implies w(m, n) ∈ E. If we denote by w
the map obtained from w by restricting its codomain to E, we have ϕ(w) = i ◦
w = w. Hence ϕ is surjective and is also trivially injective as i is a monomorphism
by Proposition 2.1.9. Moreover, ϕ is the unique group morphism which makes the
following diagram commutative:
.
130 2 Limits and Colimits
(2.26)
Indeed, as X, (si )i∈Ob I is a cone on G, the following diagram is commutative:
. (2.27)
Now it is straightforward to see that (2.26) holds true just by applying F to the
identity (2.27).
One of the most important examples of functors which preserve limits are the
hom functors.
Theorem 2.5.6 Let C be a category and C ∈ Ob C.
(1) The hom functor HomC (C, −) : C → Set preserves all existing small limits.
(2) The contravariant hom functor HomC (−, C) : C → Set maps existing small
colimits to small limits.
Hom
Hom
.
Therefore, for all m ∈ M we have HomC (C, G(f )) qi (m) = qj (m), which leads
to G(f ) ◦ qi (m) = qj (m). This implies that for each m ∈ M, C,(qi (m))i∈Ob I
is a cone on G, where qi (m) ∈ HomC (C, G(i)) for all i ∈ Ob I . As L, (pi )i∈Ob I
is the limit of G, it yields a unique morphism q(m) ∈ HomC (C, L) such that the
following diagram is commutative for all i ∈ Ob I :
Putting all this together we have defined a function q : M → HomC (C, L) (i.e.,
a morphism in Set) satisfying HomC (C, pi ) ◦ q = qi for any i ∈ Ob I , i.e., the
following diagram commutes:
Hom
Furthermore, the uniqueness of q with this property follows from that of the q(m)’s.
(2) Showing that the contravariant hom functor maps existing small colimits to
small limits follows the strategy used in the proof above and is left to the reader.
Definition 2.5.7 A functor F : C → D reflects (small) limits/colimits when for
every
functor G : I → C, where I is a (small) category, and every cone/cocone
L, (p i )i∈Ob I on G, if F (L), (F (pi ))i∈Ob I is the limit/colimit of F G, then
L, (pi )i∈Ob I is the limit/colimit of G.
132 2 Limits and Colimits
. (2.28)
Since F is fully faithful there exists a unique morphism f ∈ HomC (M, L) such
that F (f ) = f . Then (2.28) comes down to F (qi ) = F (pi ) ◦ F (f ) and since F
is faithful we obtain qi = pi ◦ f for all i ∈ Ob I , i.e., the following diagram is
commutative:
We are left to prove that f is the unique morphism which makes the above diagram
commutative. To this end, assume that g ∈ HomC (M, L) is another morphism such
that qi = pi ◦ g for all i ∈ Ob I . This implies F (pi ) ◦ F (g) = F (qi ) for all
i ∈ Ob I and since f is the unique morphism which makes diagram (2.28) commute
we obtain F (g) = f . Now recall that we also have F (f ) = f and since F is
faithful we arrive at g = f , as desired. Therefore L, (pi )i∈Ob I is a final object in
the category of cones on G, as desired.
The dual statement will be settled as usual by the duality principle. Indeed,
H : I → C be a functor, where
let I is a small category,
and consider a cocone
Q, (qi )i∈Ob I on H such that F (Q), (F (qi))i∈Ob I is the colimit of F ◦ H .
op
In particular, by Lemma 2.2.3, Q, (qi )i∈Ob I is a cone on H op . Lemma 2.2.12
op (1.25)
implies that F (Q), (F op (qi ))i∈Ob I is the limit of (F ◦ H )op = F op ◦ H op .
Since F op is obviously also fully faithful and by
the first part of
the proof any fully
op
faithful functor reflects limits, we obtain that Q, (qi )i∈Ob I is the limit of H op .
By applying Lemma 2.2.12 once more it follows that Q, (qi )i∈Ob I is the colimit
of H . This shows that F is colimit reflecting, as desired.
2.5 (Co)limit Preserving Functors 133
. (2.29)
Since F is a limit preserving functor then F (L), (F (pi ))i∈Ob I is also a limit of
F ◦ G. Exactly as in the proof of Proposition 2.1.3, one can show that there exists a
unique isomorphism g ∈ HomD (F (M), F (L)) such that the following diagram is
commutative for all i ∈ Ob I :
Hence F (f ) = g is an isomorphism
in D. Our assumption implies that f is an
isomorphism in C and thus M, (qi )i∈Ob I is also a limit of G, as desired.
let H : I →
For the dual statement, C be a functor,where I is a small category,
and consider a cocone Q, (qi )i∈Ob I on H such that F op(Q), (F (qi ))i∈Ob
I is the
op
colimit of F ◦ H . Lemma 2.2.12 implies that F (Q), (F (qi ))i∈Ob I is the limit
(1.25) op
of (F ◦ H )op = F op ◦ H op . Furthermore, by Lemma 2.2.3, Q, (qi )i∈Ob I is
a cone on H op while Lemma 2.5.2 implies that F op is limit preserving. As F op
is
obviously also
isomorphism reflecting, the first part of the proof
implies that
op
Q, (qi )i∈Ob I is the limit of H op . Now Lemma 2.2.12 shows that Q, (qi )i∈Ob I
is the colimit of H .
134 2 Limits and Colimits
Examples 2.5.10
(1) The forgetful functor U : Top → Set preserves products and equalizers (see
Examples 2.1.5 and 2.1.10). Therefore U preserves small limits by Proposi-
tion 2.5.3. Similar arguments show that the forgetful functor U : Grp → Set is
also limit preserving.
(2) The category Ab is complete and the inclusion functor I : Ab → Grp
preserves preserves products and equalizers (see Examples 2.1.5 and 2.1.10).
Therefore I preserves small limits by Proposition 2.5.3. Furthermore, I reflects
isomorphisms and, according to Proposition 2.5.9, I also reflects small limits.
.U (A, f, B) = A, U (a, b) = a,
V (A, f, B) = B, V (a, b) = b.
.
2.6 (Co)limits in Comma Categories 135
is a morphism in .(F ↓ G). This implies that .aij ∈ HomA U H (i), U H (j ) and
.bij ∈ HomB V H (i), V H (j ) such that the following diagram is commutative:
. (2.31)
. (2.32)
136 2 Limits and Colimits
The proof will be finished once we show the commutativity of the following
diagram:
Indeed, we have
(2.31)
= αH (j ) ◦ F (aij ) ◦ F (pi )
(2.32)
= αH (j ) ◦ F (pj ),
as desired.
Lemma 2.6.3 Let .H : I → (F ↓ G) be a functor, where I is a small category.
If . (A, f, B), (pi , qi ) i∈Ob I is a (co)cone on H then . A, (pi )i∈Ob I and
. B, (qi )i∈Ob I are (co)cones on .U H : I → A and .V H : I → B, respectively.
Proof We only prove the statement concerning cones. To this end, consider .t ∈
HomI (i, j ) and let .H (t) = (aij , bij ), where .(aij , bij ) ∈ Hom(F ↓G) (H (i), H (j )).
In particular, we have
The proof will be finished once we show that the following diagrams commute:
. (2.33)
. (2.34)
2.6 (Co)limits in Comma Categories 137
Now since . (A, f, B), (pi , qi ) i∈Ob I is a cone on H , we have .H (t)◦(pi , qi ) =
(pj , qj ), which componentwise comes down to (2.33) and (2.34).
We can now state and prove the main result of this section:
Theorem 2.6.4 Let .F : A → C and .G : B → C be two functors.
(1) If .A and .B are complete categories and G preserves small limits then .(F ↓ G)
is also complete and both forgetful functors .U : (F ↓ G) → A and .V : (F ↓
G) → B preserve small limits.
(2) If .A and .B are cocomplete categories and F preserves small colimits then .(F ↓
G) is also cocomplete and both forgetful functors .U : (F ↓ G) → A and
.V : (F ↓ G) → B preserve small colimits.
complete
and therefore
the functors .UH : I → A and .V H : I → B have limits,
say . L, (pi )i∈Ob I and . M, (qi )i∈Ob I respectively, where:
L ∈ Ob A,
. pi ∈ HomA (L, U H (i)),
M ∈ Ob B, qi ∈ HomB (M, V H (i)).
Using Lemma 2.6.2, we have that . F (L), (αH (i) ◦ F (pi ))i∈Ob I is a cone on
.GV H : I → C, where .αH (i) ◦ F (pi ) ∈ HomC (F (L), GV H (i)). Furthermore,
. G(M), (G(qi ))i∈Ob I is the limit of the functor .GV H : I → C as G is limit
preserving. Hence, there exists a unique morphism .h ∈ HomC (F (L), G(M)) which
makes the following diagram commutative for all .i ∈ Ob I :
(2.35)
.
138 2 Limits and Colimits
which proves that .(pi , qi ) : (L, h, M) → H (i) is in fact a morphism in the comma
category .(F ↓ G), for all .i ∈ Ob I .
We will show that . (L, h, M), (pi , qi ) i∈Ob I is the limit of the functor H .
We start by proving that . (L, h, M), (pi , qi ) i∈Ob I is a cone on H . To this end,
let .t ∈ HomI (i, j ) and let .H (t) = (aij , bij ), where .(aij , bij ) : H (i) → H (j )
is a morphism in .(F ↓ G). Now observe that the commutativity
of the following
diagram is trivially implied by the fact that . L, (pi )i∈Ob I and . M, (qi )i∈Ob I are
in particular cones on .U H and .V H , respectively:
Thus, . (L, h, M), (pi , qi ) i∈Ob I is indeed a cone on H . Consider now another
cone . (L, h, M), (pi , qi ) i∈Ob I on H . Since .(pi , qi ) : (L, h, M) → H (i) is a
morphism in .(F ↓ G) for all .i ∈ Ob I , the following diagram is commutative:
(2.36)
Furthermore, Lemma 2.6.3 implies that . L, (pi )i∈Ob I and
. M, (qi )i∈Ob I are
cones on .U H and .V H respectively. Since . L, (pi )i∈Ob I and . M, (qi )i∈Ob I
are the limits of .U H and .V H , respectively, there exist two unique morphisms
.u ∈ HomA (L, L) and .v ∈ HomB (M, M) such that the following diagrams are
. (2.37)
2.6 (Co)limits in Comma Categories 139
. (2.38)
We will show that .(u, v) is a morphism in .(F ↓ G) from . L, h, M to .(L, h, M),
i.e., the following diagram is commutative:
. (2.39)
As . G(M), (G(qi ))i∈Ob I is the limit of the functor .GV H : I → C, using
Proposition 2.2.14, (1) it will suffice to show that the following holds for all
.i ∈ Ob I :
Indeed, we have
(2.35)
G(qi ) ◦ h ◦ F (u) = αH (i) ◦ F (pi ) ◦ F (u) = αH (i) ◦ F (pi ◦ u)
.
(2.37)
= αH (i) ◦ F (pi )
(2.36)
= G(qi ) ◦ h
(2.38)
= G(qi ◦ v) ◦ h = G(qi ) ◦ G(v) ◦ h,
as desired. Moreover, in light of (2.37) and (2.38), .(u, v) is obviously the unique
morphism in .(F ↓ G) which makes the following diagram commute for all .i ∈
Ob I :
.
140 2 Limits and Colimits
which shows that . (L, h, M), (pi , qi ) i∈Ob I is indeed the limit of the functor
H . Finally, note that both functors U and V are obviously limit preserving.
.(2) We use the duality principle. As .A
op and .B op are complete categories and
.F
op : Aop → C op preserves limits (Lemma 2.5.2) we obtain, by applying .1), that
As an easy consequence of the previous result we have:
Corollary 2.6.5
(1) Let .B be a complete category and .G : B → C a functor which preserves small
limits. Then, for all .C0 ∈ Ob C, the category .(C0 ↓ G) is complete.
(2) Let .A be a cocomplete category and .F : A → C a functor which preserves
small colimits. Then, for all .C0 ∈ Ob C, the category .(F ↓ C0 ) is cocomplete.
Proof Recall from Corollary 1.8.6 that the category .(C0 ↓ G) is obtained as
a special case of the comma-category .(F ↓ G) by considering .A to be the
discrete category with one object while the functor F is the object .C0 of .C. Note
that the discrete category with one object is obviously complete and the desired
conclusion now follows from Theorem 2.6.4, (1). Similarly, .(2) follows from
Theorem 2.6.4, (2).
As the identity functor on any category preserves all small (co)limits, we obtain:
Corollary 2.6.6 Let .C be a category and .C0 ∈ Ob C.
(1) If .C is complete then the coslice category .(C0 ↓ C) is also complete.
(2) If .C is cocomplete then the slice category .(C ↓ C0 ) is also cocomplete.
Proof .(1) Follows by considering .B = C and .F = 1C in Corollary 2.6.5, (1). Simi-
larly, .(2) can be obtain by specializing .A = C and .F = 1C in Corollary 2.6.5, (2).
Functor categories form another class of categories which behave well with respect
to (co)limits. In what follows I and J are small categories and .C is an arbitrary
category. We start by introducing the following induced functors:
Lemma 2.7.1 Let .F : I → Fun(J, C) be a functor. Then, we have
2.7 (Co)limits in Functor Categories 141
Fj (i) = F (i)(j ),
. Fj (t) = F (t)j , (2.40)
where the second equality holds because F is a functor. This proves that .Fj is a
functor.
.(2) Let .t ∈ HomI (i, r). The proof will be finished once we show the
commutativity of the following diagram:
(2.41)
Note that .F (t) is a natural transformation between the functors .F (i), .F (r) : J → C.
Writing down the naturality of .F (t) for the morphism .f ∈ HomJ (j, s) yields the
following commutative diagram:
(2.42)
142 2 Limits and Colimits
pointwise.
Proof We start by proving the claim concerning limits. For each .j ∈ Ob J , let
j
. Lj , (p : Lj → Fj (i))i∈Ob I be the limit of the functor .Fj : I → C, where .Lj ∈
i
Ob C. We will construct
a functor
.L : J → C together with a family of natural
transformations . L → F (i) i∈Ob I which will form the limit of F . First we define
the functor L on objects by .L(j ) = Lj , for all .j ∈ Ob J . In order to define L on
morphisms we show first that for any .f ∈ HomJ (j, s), the pair
j
. Lj , (Ff )i ◦ pi : Lj → Fs (i) i∈Ob I
is a cone on .Fs : I → C. Indeed, we aim to prove that for any .t ∈ HomI (i, k) the
following diagram is commutative:
(2.43)
To this end, the naturality of .Ff applied to t yields the following commutative
diagram:
(2.44)
2.7 (Co)limits in Functor Categories 143
j
Furthermore, as . Lj , (pi : Lj → Fj (i))i∈Ob I is in particular a cone on .Fj , the
following diagram is commutative:
. (2.45)
j j (2.44) j
Fs (t) ◦ (Ff )i ◦ pi = F (t)s ◦ F (i)(f ) ◦ pi
. = F (k)(f ) ◦ F (t)j ◦ pi
(2.45) j j
= F (k)(f ) ◦ pk = (Ff )k ◦ pk
j
and we have proved that . Lj , (Ff )i ◦ pi : Lj → Fs (i) i∈Ob I is indeed a cone
on .Fs . As . Ls , (pis : Ls → Fs (i))i∈Ob I is the limit of .Fs , there exists a unique
morphism, denoted by .L(f ), which makes the following diagram commutative for
all .i ∈ Ob I :
. (2.46)
j
pis ◦ L(f ) = F (i)(f ) ◦ pi , .
. (2.47)
pil ◦ L(g) = F (i)(g) ◦ pis , . (2.48)
j
pil ◦ L(g ◦ f ) = F (i)(g ◦ f ) ◦ pi . (2.49)
144 2 Limits and Colimits
(2.48)
pil ◦ L(g) ◦ L(f ) = F (i)(g) ◦ pis ◦ L(f )
.
(2.47) j j
= F (i)(g) ◦ F (i)(f ) ◦ pi = F (i)(g ◦ f ) ◦ pi ,
where in the last equality we used the fact that .F (i) is a functor. This shows that
.L(g) ◦ L(f ) fulfills (2.49), which implies that .L(g) ◦ L(f ) = L(g ◦ f ). Hence
.L : J → C is a functor and therefore an object in .Fun(J, C). Next, for each .i ∈ Ob I ,
j
we define a natural transformation .pi : L → F (i) by .(pi )j = pi for all .j ∈ Ob J .
j
In order to prove that the family of morphisms . pi : L(j ) → F (i)(j ) j ∈Ob J indeed
form a natural transformation we need to show the commutativity of the following
diagram for all .f ∈ HomJ (j, s):
. (2.50)
j j
By the commutativity of diagram (2.45) we have .F (t)j ◦ pi = pk for all .j ∈ Ob J .
This implies that the natural transformations .F (t)◦pi and .pk are equal and therefore
diagram (2.50) is commutative, as desired.
Consider now another cone . H, qi : H → F (i) i∈Ob I on F , where .H : J →
2.7 (Co)limits in Functor Categories 145
(2.51)
Moreover, as . H, qi : H → F (i) i∈Ob I is a cone on F , the following diagram
is commutative for any .t ∈ HomI (i, k):
j j
Fj (t) ◦ qi = qk ,
. (2.52)
j
which implies that . H (j ), qi : H (j ) → Fj (i) i∈Ob I is a cone on .Fj . Recall
j
now that . L(j ), pi : L(j ) → Fj (i) i∈Ob I is the limit of .Fj , so there exists
a unique morphism .ξj ∈ HomC H (j ), L(j ) such that the following diagram is
commutative for all .i ∈ Ob I :
. (2.53)
The proof will be finished once we show that the morphisms .ξi , .i ∈ Ob I form
a natural transformation .ξ : H → L or, equivalently, that .ξ is a morphism in the
category .Fun(J, C). To this end, we are left to prove the commutativity of the
146 2 Limits and Colimits
. (2.54)
which shows that . L, pi : L → F (i) i∈Ob I is indeed the limit of F .
Furthermore, this allows us to embed any arbitrary small category .C into the
(co)complete category of presheaves on .C through the Yoneda embedding functor
.Y : C → Fun(C , Set) (see Definition 1.10.7). As the category of presheaves
op
2.7 (Co)limits in Functor Categories 147
Hence we have .1F (i0 ) ◦ u = g and .1F (i0 ) ◦ u = f , which shows that .f = g, as
desired.
(2) Follows easily by the duality principle and Lemma 1.9.6. Indeed, assume
.C is a category with pushouts and .ψ is an epimorphism in the functor category
By the first part of the proof, the last statement is equivalent to .(ψ op )i : F op (i) →
Gop (i) being a monomorphism in .C op for all .i ∈ Ob I . As .(ψ op )i = (ψi )op , this is
the same as .ψi : F (i) → G(i) being an epimorphism in .C for all .i ∈ Ob I .
2.8 Exercises
2.2 Let C be a category with a final object T and binary products. Prove that there
is a natural isomorphism between the identity functor 1C on C and the product
functor − × T : C → C.
2.3 Let C be a category, f , g ∈ HomC (A, B) and (E, e) the equalizer of the pair
(f, g). Then the following are equivalent:
a. f = g;
b. e is an epimorphism;
c. e is an isomorphism;
d. (A, 1A ) is the equalizer of (f, g).
2.4 Let C be a category and F : C → Set a functor. Prove that
a. if F is representable then it preserves monomorphisms;
b. if F is contravariant representable then it maps epimorphisms to monomor-
phisms.
2.5 Describe binary products in the following categories:
a. PO(P(X), ⊆), where X is a non-empty set and ⊆ denotes the inclusion of
sets;
b. PO(N, |), where | denotes the usual divisibility relation on N.
2.6 Consider the following diagram in an arbitrary category C:
h ◦ f = h ◦ g,
. h ◦ v = 1C , g ◦ u = 1B , f ◦ u = v ◦ h.
In this case, the pair (C, h) is called the split coequalizer of (f, g). Prove that
(C, h) is the coequalizer of the pair of morphisms (f, g). State and prove the
dual statement.
2.8 Exercises 149
2.7 Let F : C → D be a functor and suppose that the following diagram is a split
coequalizer in C:
Show that F (C), F (h) is the coequalizer of the pair of morphisms
F (f ), F (g) .10 State and prove the dual statement.
2.8 Let C be a complete category, f , g ∈ HomC (X, Y ) and let (X × Y, (pX , pY ))
denote the product in C of X and Y . Show that if (E, p) is the equalizer of
the pair (f, g) then (E, p, p) is the pullback of the pair (f , g), where f ,
g ∈ HomC (X, X × Y ) are the unique morphisms such that the following
hold:
pX ◦ f = 1X , pY ◦ f = f, pX ◦ g = 1X , pY ◦ g = g.
.
Show that the left-side square is a pullback if and only if the outer rectangle is
a pullback.
2.19 Let G1 and G2 be two groups with a common subgroup H and i : H → G1 ,
j : H → G2 the inclusion morphisms.
a. Show that G1 ∗H G2 , f , g is the pushout of (i, j ), where G1 ∗H G2
denotes the free product with amalgamated subgroup and f : G2 → G1 ∗H
G2 and g : G1 → G1 ∗H G2 are its corresponding group morphisms.
b. Describe the pushout of (i, j ) when H (resp. G1 ) is the trivial group.
2.20 Let C be a category with equalizers and F : C → D a functor which preserves
equalizers and reflects isomorphisms. Prove that F is faithful.
2.21 Let I and C be two arbitrary categories and assume that I is small and has an
initial object. Show that any functor F : I → C has a limit.
2.8 Exercises 151
2.22 Let (X, ) be a pre-ordered set and PO(X, ) the corresponding category.
Describe (co)limits in PO(X, ).
2.23 Let G be a group and G the corresponding category. Is G (co)complete?
2.24 Let PO(Z, ) be the category corresponding to the poset (Z, ), where is
the usual ordering on the integers. Decide if the identity functor Id : PO(Z, )
→ PO(Z, ) has a (co)limit.
2.25 Decide if the following categories are (co)complete: Grp, Ab, Top, Ring,
R M, Field. Describe (co)limits whenever they exist.
Chapter 3
Adjoint Functors
Adjoint functors were first defined by Kan ([31]) in the 50s, motivated by homo-
logical algebra ([19, 28]). Nowadays they are present in most fields of mathematics,
as will be shown in the forthcoming examples. The terminology was inspired by
adjoint operators, whose definition is somewhat similar to the correspondence in
Definition 3.1.1.
which is natural in both variables. In this case, we say that F is left adjoint to G or
equivalently that G is right adjoint to F and the notation .F G is used to designate
such a pair of adjoint functors.
Unpacking the above naturality assumption in the two variables comes down
to the following: for any .Y ∈ Ob D, θ−,Y is a natural isomorphism between
the (contravariant) functors .HomD (F (−), Y ) and .HomC (−, G(Y )) and for any
.X ∈ Ob C, θX,− is a natural isomorphism between the (covariant) functors
.HomD (F (X), −) and .HomC (X, G(−)). In particular, this amounts to the com-
© The Author(s), under exclusive license to Springer Nature Switzerland AG 2023 153
A. Agore, A First Course in Category Theory, Universitext,
https://doi.org/10.1007/978-3-031-42899-9_3
154 3 Adjoint Functors
HomD (Y, Y ):
. (3.1)
. (3.2)
Most of the categories we have considered so far are categories of sets endowed with
some extra structure (e.g., groups, rings, vector spaces, algebras, topological spaces
etc.) which allow for various forgetful functors: for example, from Grp to Set, from
.K M to Set, from .AlgK (or .Alg ) to .K M. It turns out that all the forgetful functors
c
K
mentioned above do have left adjoints and this phenomenon can be explained by
the existence of the so-called free objects; to be more precise, we have a free group
(vector space) on any set, a free algebra (i.e., the tensor algebra) on any vector
space and a free commutative algebra (i.e., the symmetric algebra) on any vector
space. These free objects, together with their universal property, will be the main
ingredients in the construction of the left adjoints for the aforementioned forgetful
3.2 Adjoints Via Free Objects 155
functors. We consider below the case of the forgetful functor from Grp to Set, but
the same strategy works in general.
Example 3.2.1 Let .U : Grp → Set be the forgetful functor. We will see that U has
a left adjoint .F : Set → Grp called the free group functor. More precisely, F is
constructed as follows:
• for any .X ∈ Ob Set, define .F (X) = F X, the free group on the set X;1
• given .f ∈ HomSet (X, Y ), define .F (f ) : F X → F Y by .F (f ) = f ,
where .f is obtained from the universal property of the free group F X, i.e.,
.f ∈ Hom
Grp (F X, F Y ) is the unique group homomorphism which makes the
following diagram commute:
. (3.3)
for any .v ∈ HomGrp (F X, G). The inverse of .θX,G , denoted by .ψX,G , is defined
as follows:
. (3.4)
1A group G containing X as a subset is called the free group on X if for every group .G and
every function .f : X → G , there exists a unique group homomorphism .ψ : G → G such that
.f = ψ ◦ iX , where .iX : X → G is the inclusion map ([49, Section 5.5]).
156 3 Adjoint Functors
Since v makes the above diagram commutative we get .ψX,G ◦ θX,G (v) = v. On the
other hand, if .u ∈ HomSet (X, U (G)), we have
where .u is the unique group homomorphism which makes diagram (3.4) commute.
Thus .u ◦ iX = u and we obtain .θX,G ◦ ψX,G (u) = u, as desired.
Finally we check that the isomorphism .θ is natural in both variables. First,
fix .G ∈ Ob Grp and consider .f ∈ HomSet (X , X). We need to prove the
commutativity of the following diagram:
= r ◦ iX ◦ f
(3.3)
= r ◦ f ◦ iX = r ◦ F (f ) ◦ iX
= θX , G (r ◦ F (f )) = θX ,G ◦ HomGrp (F (f ), G)(r).
3.3 Galois Connections 157
Finally, fix .X ∈ Ob Set and consider .g ∈ HomGrp (G, G ). We are left to prove
that the following diagram is commutative:
.
Another class of generic examples of adjoint functors can be obtained from pre-
ordered sets regarded as categories (see Example 1.2.2, (2)). The general context
is the following: .(X, ) and .(Y, ) are two pre-ordered sets and we consider the
corresponding induced categories .PO(X, ) and .PO(Y, ), respectively. More-
over, functors between such categories are nothing but order-preserving functions
between the underlying pre-ordered sets, as we have seen in Example 1.5.3, (35).
If .F : PO(X, ) → PO(Y, )op and .G : PO(Y, )op → PO(X, ) are two
functors then F is left adjoint to G if and only if for all .x ∈ X and .y ∈ Y we have
Indeed, recall that the hom sets in any category induced by a pre-ordered set have
at most one element. Therefore, condition (3.5) can be equivalently expressed as
a bijection between .HomPO(Y, )op (F (x), y) and .HomPO(X, ) (x, G(y)). This
bijection is trivially natural as we have at most one element in each hom set.
A pair of adjoint functors as above is called a Galois connection from .(X, )
to .(Y, ). Important examples of Galois connections can be found all across the
158 3 Adjoint Functors
mathematical landscape. For instance, these include the classical Galois correspon-
dence for field extensions as well as the correspondence between algebraic sets and
radical ideals in algebraic geometry.
An important example concerns the Galois correspondence for field extensions.
Example 3.3.1 Throughout this example, .K ⊆ L is a field extension and for any
field .S ⊆ L we define
Furthermore, let
A = {S ⊆ L subfield | K ⊆ S ⊆ L} and
.
B = {H ⊆ G | H subgroup in G}.
Then .(A, ⊆) and .(B, ⊆) are pre-ordered sets, where by a slight abuse of notation
we use “.⊆” to denote both inclusions. We can now construct a Galois connection
from .(A, ⊆) to .(B, ⊆) as follows for all .S ∈ A and .H ∈ B:
First note that if .S ∈ A then any .ψ in .Gal(L/S) fixes all elements of S and, in
particular, those of K, which shows that .Gal(L/S) is a subgroup of .Gal(L/K) = G;
thus .F (S) ∈ B. Similarly, if H is a subgroup of G, and all elements of G fix all
elements of K, it follows that .K ⊆ Fix(H ); we can now conclude that .G(H ) ∈
A.
Moreover, if .S ⊆ S and .ψ ∈ Gal(L/S ) then .ψ(x) = x for all .x ∈ S and, in
particular, the same holds for all .x ∈ S. This shows that .Gal(L/S ) ⊆ Gal(L/S)
and therefore F is order-preserving, i.e., a functor between the corresponding
categories. Now if .H ⊆ H and .l ∈ Fix(H ) then .τ (l) = l for all .τ ∈ H
and, in particular, the same holds for all .τ ∈ H . Therefore, .Fix(H ) ⊆ Fix(H )
and G is also order-preserving. We are left to show that (3.5) also holds. Indeed
for all .S ∈ A and .H ∈ B we have .G(H ) = Fix(H ) ⊇ S if and only if
.τ (y) = y for all .τ ∈ H and .y ∈ S if and only if .H ⊆ F (S), as desired.
.
First, note that the properties of a group action imply that the set
.h
x = x for all .h ∈ H and, in particular, for all .h ∈ H . Hence .G(H ) ⊆ G(H )
and G is a well-defined functor from .(B, ⊆)op to .(A, ⊆). Finally, the two functors
fulfill condition (3.5); indeed, given .Y ∈ A and .H ∈ B we have .F (Y ) ⊇ H if and
only if .h y = y for all .y ∈ Y and .h ∈ H if and only if .Y ⊆ G(H ). To conclude,
we have proved that the functors F and G form a Galois connection.
We start this section with more examples of adjoint functors, spanning various fields.
Examples 3.4.1
(1) For any non-empty set X, the functor .− × X : Set → Set has a right adjoint
given by .HomSet (X, −) : Set → Set. Indeed, for all .Y, Z ∈ Ob Set, define
θY, Z : HomSet (Y × X, Z) → HomSet Y, HomSet (X, Z) ,
.
Indeed, we have
.ψY, Z ◦ θY, Z (f )(y, x) = θY, Z (f )(y) (x) = f (y, x),
θY, Z ◦ ψY, Z (g)(y)(x) = ψY, Z (g)(y, x) = g(y) (x).
We are left to show the commutativity of diagrams (3.1) and (3.2). To start with,
let .f ∈ HomSet (Y , Y ); we need to show the commutativity of the following
diagram:
ψY, Z (g)(y ⊗ x) = g(y) (x),
for all .g ∈ HomK M Y, HomK M (X, Z) , y ∈ Y and .x ∈ X.
Showing that the maps defined above are indeed K-linear is straightforward
using the properties of the tensor product while proving the commutativity of
the diagrams (3.1) and (3.2) goes very much along the lines of the previous
example and is left to the reader.
(3) If X is a locally compact and Hausdorff topological space then the
functor .− × X : Top → Top (see Example 1.5.3, (18)) is left adjoint to
.Hom
Top (X, −) : Top → Top (see Example 1.5.3, (11)) as proved, for instance,
in [12, Chapter 5] in a more general setting. Recall that for any .Y ∈ Ob Top
we consider on .HomTop (X, Y ) the compact-open topology while .Y × X is
endowed with the product topology.
In the first example above we proved that for any .Y, Z ∈ Ob Set we have
a set bijection between .HomSet (Y × X, Z) and .HomSet Y, HomSet (X, Z) .
We will show that this bijection induces
a continuous bijective
map between
.Hom
Top (Y × X, Z) and .Hom
Top Y, Hom Top (X, Z) with respect to the
previously mentioned topologies. To this end, for all Y , .Z ∈ Ob Top define
θY, Z : HomTop (Y × X, Z) → HomTop Y, HomTop (X, Z) ,
.
−1
If .f W (K, V ) = ∅ then .f is obviously continuous, as desired. Assume now
−1 −1
that .f W (K, V ) = ∅ and consider .y ∈ f W (K, V ) . Therefore, we
have .f (y) ∈ W (K, V ) and we obtain .f (y)(k) = f (y, k) ∈ V for all .k ∈ K.
162 3 Adjoint Functors
{y} × K ⊆ Uy × Ty ⊆ f −1 (V ).
.
−1
This shows that .y ∈ Uy ⊆ f W (K, V ) and therefore we have
−1
f
. W (K, V ) = Uy .
−1
y∈f W (K, V )
−1
We can now conclude that .f W (K, V ) is an open set as a union of open
sets. To summarize, we proved that if .f : Y × X → Z is continuous then
.f = θY, Z (f ) : Y → Hom
Top (X, Z) is continuous as well.
Consider now .ψY, Z , the inverse of .θY, Z , given as follows:
ψY, Z. : HomTop Y, HomTop (X, Z) → HomTop (Y × X, Z),
ψY, Z (g)(y, x) = g(y) (x),
for all .g ∈ HomTop Y, HomTop (X, Z) , y ∈ Y and .x ∈ X.
We are left to show that .ψY, Z is well-defined, i.e., if .g : Y →
HomTop (X, Z) is continuous then .g = ψY, Z (g) : Y × X → Z is continuous
as well. We start by showing that the evaluation map .ev : HomTop (X, Z) ×
X → Z defined by .ev(g, x) = g(x) is continuous at every point3
.(g, x) ∈ Hom
Top (X, Z) × X. Indeed, if .V ⊆ Z is an open set such that
.ev(g, x) = g(x) ∈ V then the continuity of g implies that .g
−1 (V ) ⊆ X
−1
is an open subset such that .x ∈ g (V ). As X is Hausdorff and locally
compact4 we can find an open subset .U ⊆ X whose closure .U is compact
and .x ∈ U ⊆ U ⊆ g −1 (V ). Therefore, we have .g(x) ∈ g U ⊆ V , which
shows that .W (U , V ) × U ⊆ HomTop (X, Z) × X is an open subset such that
.(g, x) ∈ W (U , V ) × U and .ev W (U , V ), U ⊆ V . Hence .ev is a continuous
2 The tube lemma: Let A and B compact subspaces of X and Y , respectively, and let N be an open
set in .X × Y containing .A × B. Then, there exist open subsets U and V in X and Y , respectively,
such that .A × B ⊆ U × V ⊆ N (see [39, Lemma 26.8 and exercise 9 on page 171]).
3 Let .f : A → B be a map between two topological spaces. We say that f is continuous at a
point .x ∈ A if, for each neighborhood V of .f (x), there is a neighborhood U of x whose closure
.f (U ) ⊂ V . The map f is continuous if and only if is continuous at every point .x ∈ A ([39,
Theorem 18.1]).
4 Recall that if X is a Hausdorff space then X is locally compact if and only if given .x ∈ X and
(1.6)
= t ◦ πG ◦ f = t ◦ fab ◦ πG
= θG , A (t ◦ fab ) = θG , A ◦ HomAb (fab , A)(t),
164 3 Adjoint Functors
which shows that (3.1) is commutative. Consider now .g ∈ HomAb (A, A ) and
.G ∈ Ob Grp. We are left to show the commutativity of the following diagram:
for all .uop ∈ HomDop (X, F op (Y )) Indeed, for all .t op ∈ HomCop (Gop (X), Y ) and
.u
op ∈ Hom op (X, F op (Y )) we have
D
−1 op
θ X, Y ◦ ξX, Y (uop ) = θ X, Y θY, X (u)op = θY,
.
X θ Y, X (u) = uop ,
op −1 op
−1
ξX, Y ◦ θ X, Y (t op ) = ξX, Y θY, X (t) = θY, X θ Y, X (t) = t op .
We are left to show that .θ is natural in both variables. To this end, consider first
g op ∈ HomDop (D , D). The naturality of .θ in the first variable comes down to
.
Since we have already proved that each .θ D, C is invertible with inverse .ξD, C it will
suffice to show that the following holds:
(3.2) op
= G(g) ◦ θC, D (u)
op
= θC, D (u) ◦op G(g)op
op
= HomCop (Gop (g op ), C) θC, D (u)
Relying again on the fact that .θ D, C is invertible with inverse .ξD, C it will suffice to
show that the following holds:
op
= f op ◦op θC, D (u)
op
= θC, D (u) ◦ f
(3.1)
op
= θC , D (u ◦ F (f ))
= ξD, C (u ◦ F (f ))op
= ξD, C F (f )op ◦op uop
= ξD, C ◦ HomDop (D, F op (f op ))(uop ).
Hence (3.9) holds and this shows the naturality of .θ in the second variable.
3.4 More Examples and Properties of Adjoint Functors 167
i.e., HomA (A, GH (d)) ◦ θA, H (i) = θA, H (j ) ◦ HomB (F (A), H (d)).
. (3.10)
Moreover, since . A, (qi : A → GH (i))i∈Ob I is a cone on GH the following
diagram is commutative:
. (3.11)
168 3 Adjoint Functors
HomA (A, GH (d)) ◦ θA, H (i) (ri ) = θA, H (j ) ◦ HomB (F (A), H (d))(ri )
.
. (3.12)
Denote .θA,L (f ) ∈ HomA (A, G(L)) by g. We are left to prove that the following
diagram is commutative for all .i ∈ Ob I :
. (3.13)
Using again the naturality of the bijection .θ we obtain the following commutative
diagram for all .i ∈ Ob I :
(3.12)
for any .i ∈ Ob I . Therefore, we have .pi ◦ f = ri = pi ◦ f for all .i ∈ Ob I . By
Proposition 2.2.14, (1) this implies .f = f and consequently .g = g, as desired.
The second part of the theorem follows easily by duality. Indeed, if .F G
then Theorem 3.4.3 implies that we also have .Gop F op . According to the above
proof, .F op preserves all existing limits. Now using Lemma 2.5.2 we obtain that F
preserves colimits, as desired.
Theorem 3.4.4 can be very useful in ruling out the existence of left/right adjoints
for certain functors, as shown in the following examples:
Examples 3.4.5
(1) The forgetful functor .F : Ab → Set does not preserve coproducts. Therefore,
by Theorem 3.4.4 it does not have a right adjoint.
(2) Consider now the inclusion functor .I : Ring → Rng. As .Z is an initial object
in .Ring but not in .Rng we can conclude by Theorem 3.4.4 that it does not admit
a right adjoint.
(3) The forgetful functor .U : Field → Set does not have a left adjoint. Indeed,
if .F : Set → Field is a left adjoint to U then by Theorem 3.4.4, F needs
to preserve colimits. In particular, this would imply the existence of an initial
object in Field, which contradicts Example 1.3.10, (4). .
170 3 Adjoint Functors
Our next result gives an important equivalent description of adjoint functors in terms
of two natural transformations called the unit and the counit of the adjunction.
Theorem 3.5.1 Let .F : C → D and .G : D → C be two functors. Then F is left
adjoint to G if and only if there exist two natural transformations
.η : 1C → GF, ε : F G → 1D
In this case .η and .ε are called the unit and the counit of the adjunction, respectively.
Proof Suppose first that .F G and let .θ : HomD (F (−), −) → HomC (−, G(−))
be the corresponding natural isomorphism. For each .C ∈ Ob C and .D ∈ Ob D we
have the following bijective maps:
we obtain
yields
Finally, we move on to proving that .η and .ε are natural transformations. First we will
collect some compatibilities using the commutativity of the diagrams (3.1) and (3.2).
Setting .X = C, X = C and .Y = F (C) in (3.1) yields the following commutative
172 3 Adjoint Functors
From the commutativity of the above diagram applied to .1F (C) we get
HomC (f, GF (C)) ◦ θC, F (C) (1F (C) ) = θC , F (C) ◦ HomD (F (f ), F (C))(1F (C) )
.
⇔ HomC (f, GF (C)) ◦ ηC = θC , F (C) F (f )
i.e., ηC ◦ f = θC , F (C) F (f ) .
. (3.17)
Finally, we use the commutativity of the diagram (3.2) for .X = G(D ), Y = D and
.Y = D. It yields the following commutative diagram for all .g ∈ HomD (D , D):
We are now in a position to prove that .η and .ε are natural transformations. Indeed,
the naturality of .η comes down to proving the commutativity of the following
diagram for all .h ∈ HomC (C , C):
where in the second equality we used (3.18) for .g = F (h). Thus .η is a natural
transformation.
The naturality of .ε comes down to proving the commutativity of the following
diagram for all .t ∈ HomD (D , D):
174 3 Adjoint Functors
(3.19) −1 (3.20)
.εD ◦ F G(t) = θG(D ), D G(t) = g ◦ εD .
Note that the first equality follows by applying (3.19) for .f = G(t).
Assume now that there exist two natural transformations .η : 1C → GF and
.ε : F G → 1D such that (3.15) and (3.16) are fulfilled for any .C ∈ Ob C and
for any .u ∈ HomD (F (C), D) and .v ∈ HomC (C, G(D)). First we will prove that
θC, D and .ϕC, D are inverses to each other for any .C ∈ Ob C and .D ∈ Ob D. To start
.
with, we note for further use that the naturality of .η and .ε imply the commutativity
of the following diagrams for all .u ∈ HomD (F (C), D) and .v ∈ HomC (C, G(D)):
. (3.21)
. (3.22)
Now, we have
.θC, D ◦ ϕC, D (v) = θC, D εD ◦ F (v) = G εD ◦ F (v) ◦ ηC
= G(εD ) ◦ GF (v) ◦ ηC
3.5 The Unit and Counit of an Adjunction 175
(3.21)
= G(εD ) ◦ ηG(D) ◦ v
(3.16)
= v,
ϕC, D ◦ θC, D (u) = ϕC, D G(u) ◦ ηC = εD ◦ F G(u) ◦ ηC
= εD ◦ F G(u) ◦ F (ηC )
(3.22)
= u ◦ εF (C) ◦ F (ηC )
(3.15)
= u.
Thus .θC, D and .ϕC, D are inverses to each other for any .C ∈ Ob C and .D ∈ Ob D.
We are left to prove that .θ is natural in both variables, i.e., diagrams (3.1) and (3.2)
are commutative. Indeed, let .f ∈ HomC (C , C) and .u ∈ HomD (F (C), D); we
have
HomC (f, G(D)) ◦ θC, D (u) = HomC (f, G(D)) ◦ G(u) ◦ ηC
.
= G(u) ◦ ηC ◦ f
= G(u) ◦ GF (f ) ◦ ηC
= θC , D u ◦ F (f )
= θC , D ◦ HomD (F (f ), D)(u),
where in the third equality we used the naturality of .η applied to f . Thus, (3.1)
holds.
Consider now .g ∈ HomD (D, D ) and .u ∈ HomD (F (C), D). Then:
= G(g ◦ u) ◦ ηC
= θC, D ◦ HomD (F (C), g)(u).
This proves that (3.2) also holds and the proof is now finished.
Examples 3.5.2
(1) Let .F : C → D be an isomorphism of categories with inverse .G : D → C. Then
.(F, G) and .(G, F ) are pairs of adjoint functors with unit and counit given
εZ : HomK M (X, Z) ⊗ X → Z, εZ (f ⊗ x)
.εY ⊗X ◦ (ηY ⊗ 1X )(y ⊗ x) = εY ⊗X ηY (y) ⊗ x
= ηY (y)(x) = y ⊗ x,
HomK M (X, εY ) ◦ ηHom (f )(x) = HomK M (X, εY )(f ⊗ x)
K M (X, Y )
= εY (f ⊗ x) = f (x),
• for any .X ∈ Ob Set, define .R(X) = RX, the free module generated by X;5
• given .f ∈ HomSet (X, Y ), define .R(f ) : RX → RY by .R(f ) = f , where
.f is the unique homomorphism of R-modules which makes the following
diagram commute:
. (3.23)
where .(RX, iX ) and .(RY, iY ) are the free R-modules generated by X and
Y , respectively. .R is the left adjoint of the forgetful functor .U : R M → Set.
Indeed, the unit .η : 1Set → U R is defined for all .X ∈ Ob Set by .ηX = iX ,
where .iX : X → RX is the map corresponding to the free R-module on X
while the counit .ε : RU → 1R M is defined for all .M ∈ Ob R M as the unique
homomorphism of R-modules .εM : R(U (M)) → M such that the following
diagram is commutative:
. (3.24)
First, note that (3.23) implies in particular that for all .f ∈ HomSet (X, Y ) the
following diagram is commutative:
(3.24) (3.23)
1RX ◦ iX = = εRX ◦ iU (RX) ◦ iX = εRX ◦ R(iX ) ◦ iX .
.
This shows that .1RX ◦ iX = εRX ◦ R(iX ) ◦ iX and in light of [45, Definition
1.8] we can conclude that .1RX = εRX ◦ R(iX ). Hence, (3.15) also holds.
Consider now .g ∈ HomR M (M, N ); the proof will be finished once we show
the commutativity of the following diagram:
By the same argument used in the above paragraph, it will suffice to show that
g ◦ εM ◦ iU (M) = εN ◦ RU (g) ◦ iU (M) . Indeed, we have
.
(3.23)
εN ◦ RU (g) ◦ iU (M) = εN ◦ iU (N ) ◦ U (g)
.
(3.24) (3.24)
= 1N ◦ g = g ◦ 1M = g ◦ εM ◦ iU (M) ,
178 3 Adjoint Functors
as desired.
.
We record here for further use the following slightly more general version of the
compatibility conditions between the unit and counit of an adjunction:
Lemma 3.5.3 Let .F : C → D and .G : D → C be two functors such that .F
G and consider the corresponding natural isomorphism .θ : HomD (F (−), −) →
HomC (−, G(−)). If .η and .ε are the unit and counit of this adjunction, then for all
.u ∈ HomD (F (C), D) and .v ∈ HomC (C, G(D)) we have
εD ◦ F θC, D (u) = u, .
. (3.25)
−1
G θC, D (v) ◦ ηC = v. (3.26)
From the commutativity of the above diagram applied to .εD ∈ HomD (F G(D), D)
we obtain
.HomC (θC, D (u), G(D)) ◦ θG(D), D (εD ) = θC, D ◦ HomD F θC, D (u) , D (εD )
⇔ θG(D), D (εD ) ◦ θC, D (u) = θC, D εD ◦ F θC, D (u)
⇔ 1G(D) ◦ θC, D (u) = θC, D εD ◦ F θC, D (u)
.
3.5 The Unit and Counit of an Adjunction 179
The commutativity of the above diagram applied to .1F (C) ∈ HomD (F (C),
F (C)) yields
−1
θC, D ◦ HomD (F (C), θC,
.
D (v))(1F (C) )
−1
= HomC C, G θC, D (v) ◦ θC, F (C) (1F (C) )
−1
⇔ v = G θC, D (v) ◦ ηC ,
where .θ denotes the natural isomorphism induced by the adjunction. Similarly, if .η,
ε, θ denote the unit, the counit and respectively the natural isomorphism induced
.
C ∈ Ob C and .D ∈ Ob D:
.
i.e., u ◦ ψD
C
(v) = ψD
C
(u ◦ v) for all v ∈ HomD (F1 (C), D).
(3.2)
−1
⇔ G2 (u) ◦ θC,
2
D
2
θC, D αD ◦ θC,
1
D (v) = αD ◦ θC,
1
D (u ◦ v)
⇔ G2 (u) ◦ αD ◦ θC,
1
D (v) = αD ◦ θC, D (u ◦ v)
1
(3.2)
⇔ G2 (u) ◦ αD ◦ θC,
1
D (v) = αD ◦ G1 (u) ◦ θC, D (v),
1
3.5 The Unit and Counit of an Adjunction 181
2
Given the bijectivity of each .θC, F1 (C ) it will suffice to prove that the following
holds:
F1 (C ) α C ◦ F2 (g) = θC, F1 (C ) F1 (g) ◦ α C .
2 2
θC,
.
(3.17)
= αF1 (C ) ◦ ηC 1
◦g
= αF1 (C ) ◦ θC1 , F1 (C ) 1F1 (C ) ◦ g
(3.29)
= θC2 , F1 (C ) (α C ) ◦ g
(3.1)
= θC,
2
F1 (C ) α C ◦ F2 (g) ,
where .η1 and .θ 1 denote the unit and respectively the natural isomorphism induced
by the adjunction .F1 G1 .
Recall that if I is a small category then any functor .F : C → D induces a
functor between the corresponding functor categories .F : Fun (I, C) → Fun (I, D)
as in (1.36). Our next result shows that any adjunction can be lifted to an adjunction
between the corresponding induced functors.
182 3 Adjoint Functors
ηH = ηH, ε K = εK.
.
In order to show that .η and .ε are indeed natural transformations, consider two
natural transformations .α : H → H and .β : K → K , where .H, H : I → C
and K, .K : I → D are functors. First, by the naturality of .η and .ε, the following
diagrams are commutative for all .i ∈ Ob I :
(3.30)
To summarize, for all .i ∈ Ob I we obtain
(1.36) (3.30)
. G F (α) ◦ ηH i
= GF (αi ) ◦ ηH (i) = ηH (i) ◦ αi = (ηH ◦ α)i ,
(1.36) (3.30)
ε K ◦ F G (β) i = εK (i) ◦ F G(βi ) = βi ◦ εK(i) = (β ◦ ε K )i ,
which shows that .η and .ε are natural transformations. The proof will be finished
once we show that .η and .ε fulfill (3.15) and (3.16). To start with, note that since .η
and .ε fulfill (3.15) and (3.16), in particular the following hold for all .i ∈ Ob I :
1F H (i) = εF H (i) ◦ F ηH (i) ,
. 1GK(i) = G εK(i) ◦ ηGK(i) .
which shows precisely that .η and .ε fulfill (3.15) and (3.16) and the proof is now
finished.
. (3.31)
(3) there exists a natural transformation .ε : F G → 1D such that for any morphism
.f ∈ HomD (F (C), D) there exists a unique morphism .g ∈ HomC (C, G(D))
. (3.32)
As the notation suggests, the natural transformations .η and .ε are precisely the unit
and the counit, respectively, of the adjunction .F G.
Proof We start by proving the equivalence between (1) and (2). Suppose first that
.F G and let .θ be the corresponding natural isomorphism. We define the natural
transformation .η : 1C → GF as in the proof of Theorem 3.5.1, namely by .ηC =
θC,F (C) (1F (C) ) for any .C ∈ Ob C. Let .f ∈ HomC (C, G(D)); we will prove that
−1
.g = θ
C,D (f ) ∈ HomD (F (C), D) is the unique morphism in .D which makes
184 3 Adjoint Functors
HomC (C, G(u)) ◦ θC, F (C) (1F (C) ) = θC, D ◦ HomD (F (C), u)(1F (C) ),
.
Thus, we have
(3.33) −1
G(g) ◦ ηC = θC,D (g) = θC,D ◦ θC,D
. (f ) = f,
−1
which shows that .g = θC,D (f ) ∈ HomD (F (C), D) makes diagram (3.31) com-
mutative. Assume now that there exists another morphism .g ∈ HomD (F (C), D)
such that .G(g ) ◦ ηC = f and let .f = θC,D (g ). Following the same steps as in
the argument above it can be easily seen that .G(g ) ◦ ηC = θC,D (g ) = f . Our
assumption now implies that .f = f and therefore, since .θC, D is a bijection, we
obtain .g = g .
Assume now that .2) holds, i.e., for any .f ∈ HomC (C, G(D)) there exists a
unique morphism .g ∈ HomD (F (C), D) such that (3.31) is fulfilled. Given .C ∈
Ob C and .D ∈ Ob D we define the following map:
for any .u ∈ HomD (F (C), D). Obviously, our assumption implies that .θC, D is a
set bijection for all .C ∈ Ob C and .D ∈ Ob D. The fact that .θ defined in (3.34) is
natural in both variables follows exactly as in the proof of Theorem 3.5.1.
Finally, we are left to show the equivalence between (1) and (3). Indeed, by
Theorem 3.4.3, .F G if and only if .Gop F op . By applying the equivalence
between (1) and (2) we obtain that .Gop F op if and only if there exists a
natural transformation .η : 1Dop → F op Gop with the property that for any .f op ∈
HomDop (D, F op (C)) there exists a unique .g op ∈ HomCop (Gop (D), C) such that
3.6 Another Characterisation of Adjoint Functors 185
F op (g op ) ◦op ηD = f op .
. (3.35)
the adjunction.
(1) If .g, g ∈ HomD (F (C), D) such that .G(g) ◦ ηC = G(g ) ◦ ηC then .g = g .
(2) If .h, h ∈ HomC (C, G(D)) such that .εD ◦ F (h) = εD ◦ F (h ) then .h = h .
Proof
(1) Follows trivially from Theorem 3.6.1, (2) by considering .f = G(g ) ◦ ηC . Then
both morphisms g and .g make diagram (3.31) commutative, which implies .g =
g . The second part follows in a similar manner by using Theorem 3.6.1, (3).
Examples 3.6.3
(1) The forgetful functor .U : Top → Set has both a left and a right adjoint. We
start by constructing the left adjoint functor .F : Set → Top which endows
each .X ∈ Ob Set with the discrete topology. We define a natural transformation
.η : 1
Set → U F by .ηX (x) = x for any .X ∈ Ob Set and .x ∈ X. Consider now
.f ∈ Hom
Set (X, U (Y )), where .Y ∈ Ob Top. According to Theorem 3.6.1, (2)
in order to prove that .F U we need to find a unique morphism .g ∈
HomTop (F (X), Y ) such that the following diagram commutes:
−1
GU by .ηX (x) = x for all .X ∈ Ob Top and .x ∈ X. Now since .ηX (∅) =
−1
∅ and .ηX (G(X)) = G(X) = X we obtain that each .ηX is continuous.
Consider now .f ∈ HomTop (X, G(Y )), where .Y ∈ Ob Set. We aim to find
a unique morphism .g ∈ HomSet (U (X), Y ) such that the following diagram is
commutative:
As before, we set .g = f .
(2) The Stone–Čech compactification functor .S : Top → KHaus defined in Exam-
ple 1.5.3, (24) is left adjoint to the inclusion functor .I : KHaus → Top. Indeed,
let .i : 1Top → I S be the natural transformation defined for all topological
spaces X by the continuous map .iX : X → S(X) associated with the Stone–
Čech compactification of X. If .f ∈ HomTop (X, Y ) then .S(f ) is defined
by (1.7) as the unique morphism in .KHaus such that .I S(f ) ◦ iX = iY ◦ f .
Therefore, the following diagram is commutative for all .f ∈ HomTop (X, Y ):
Using Theorem 3.6.1, (2) we can conclude that .S is left adjoint to I , as desired.
(3) The Grothendieck group functor .G : Mon → Grp defined in Exam-
ple 1.5.3, (25) is left adjoint to the inclusion functor .I : Grp → Mon. Indeed,
let .i : 1Mon → I G be the natural transformation defined for all monoids M
3.6 Another Characterisation of Adjoint Functors 187
Using Theorem 3.6.1, (2) we can conclude that G is left adjoint to I , as desired.
(4) The functor .U : Mon → Grp defined in Example 1.5.3, (26) which assigns to
each monoid its group of invertible elements, is right adjoint to the inclusion
functor .I : Grp → Mon. Indeed, let .ε : I U → 1Mon be the natural
transformation defined for all .M ∈ Ob Mon by .εM : U (M) → M, εM = iM ,
where .iM denotes the inclusion map. If .f ∈ HomMon (M, N ), then for all
.m ∈ U (M) we have
.(f ◦ iM )(m) = f (m) = iN ◦ f|U (M) (m) = iN ◦ I U(f ) (m).
Using Theorem 3.6.1, (2) we can now conclude that .L is left adjoint to .Fj , as
desired.
3.6 Another Characterisation of Adjoint Functors 189
This shows that q is indeed a natural transformation. Now recall that by the
universal property of the Hausdorff quotient, for any .f ∈ HomTop (X, I (Z)),
where .Z ∈ Ob Haus, there exists a unique .g ∈ HomHaus (H (X), Z) such that
the following diagram is commutative:
Using Theorem 3.6.1, (2) we can conclude that .H is left adjoint to I , as desired.
(7) The Dorroh extension functor .D : Rng → Ring defined in Example 1.5.3, (31)
is left adjoint to the inclusion functor .I : Ring → Rng. Indeed, let .j : 1Rng →
I D be the natural transformation defined for all rings R by the ring homo-
morphism .jR : R → D(R) associated with the Dorroh extension of R. If
.f ∈ Hom
Rng (R, S) then .D(f ) is defined by (1.11) as the unique morphism
in Ring such that .jS ◦ f = I D(f ) ◦ jR . Therefore, the following diagram is
commutative for all .f ∈ HomRng (R, S):
In particular, this shows that j is a natural transformation. Now recall that by the
universal property of the Dorroh extension, for any .f ∈ HomRng (R, I (T )),
190 3 Adjoint Functors
As another application of Theorem 3.6.1 we will show that, when they exist,
left/right adjoints are unique up to natural isomorphism.
Theorem 3.6.4 Any two left (right) adjoints of a given functor are naturally
isomorphic.
Proof Assume .F, F : C → D are both left adjoint functors of .G : D → C. Then
there exist natural transformations .η : 1C → GF and .η : 1C → GF satisfying
the conditions in Theorem 3.6.1, (2). Given .C ∈ Ob C, as .F G and .ηC ∈
HomC (C, GF (C)), there exists a unique morphism .γC ∈ HomD (F (C), F (C))
such that
G(γC ) ◦ ηC
. = ηC . (3.36)
G(γC ) ◦ ηC = ηC
.
. (3.37)
We will see that each .γC is an isomorphism with the inverse given precisely by
.γC . Indeed, using (3.36) and (3.37) we can easily see that .G(γC ◦γC )◦ηC = η and
C
since we obviously also have .G(1F (C) ) ◦ ηC = ηC it follows by Corollary 3.6.2, (1)
that .γC ◦ γC = 1F (C) . Similarly, one can prove that .γC ◦ γC = 1F (C) .
We are left to prove that .γ : F → F is a natural transformation, i.e., for any
.f ∈ HomC (C, C ) the following diagram is commutative:
.
3.6 Another Characterisation of Adjoint Functors 191
Using Corollary 3.6.2, (1) it is enough to prove that the following holds:
G(F (f ) ◦ γC ) ◦ ηC
. = G(γC ◦ F (f )) ◦ ηC
. (3.38)
To this end, we use the naturality of .η and respectively .η ; that is, the commutativity
of the following diagrams:
. (3.39)
. (3.40)
Then, we have
(3.36)
GF (f ) ◦ G(γC ) ◦ ηC
. = GF (f ) ◦ ηC
(3.39)
= ηC ◦ f
(3.36)
= G(γC ) ◦ ηC ◦f
(3.40)
= G(γC ) ◦ GF (f ) ◦ ηC
.
Therefore, (3.38) indeed holds. To summarize, we have proved that there exists a
natural isomorphism .γ : F → F and the proof is now finished.
Adjunctions can also be used to easily derive important properties of certain func-
torial constructions, as the following examples show. This includes, for instance, the
commutation of tensor products or localizations with direct sums of modules. All of
these are obtained by applying Theorem 3.4.4.
Example 3.6.5 Given a commutative ring R, for any .X ∈ ObR M and any family
(Mi )i∈I of R-modules we have the following isomorphisms of R-modules:
.
. ⊕i∈I Mi ⊗ X ⊕i∈I Mi ⊗ X ,
S −1 ⊕i∈I Mi ⊕i∈I S −1 Mi ,
192 3 Adjoint Functors
where .⊗ = ⊗R . Indeed, both statements are consequences of the fact that both
the tensor product functor .− ⊗ X : R M → R M and the localization functor
.L : R M →S −1 R M are left adjoints (see Example 3.4.1, (2) and Example 3.6.3, (5))
indeed a monomorphism.
Conversely, assume .ηC is a monomorphism for all .C ∈ Ob C and let
.f1 , f2 ∈ HomC (C , C) such that .F (f1 ) = F (f2 ). Using (3.15) we obtain
.εF (C) ◦ F (ηC ◦ f1 ) = εF (C) ◦ F (ηC ◦ f2 ). Now Corollary 3.6.2, (2) implies
Therefore, F is faithful.
The result concerning the functor G follows by duality. Indeed, by The-
orem 3.4.3 we have .Gop F op and moreover, Corollary 3.5.4 shows that
the unit of this adjunction is precisely .εop . Since G is faithful if and only if
.G
op is faithful, the desired conclusion follows the first part of the proof and
(3.15)
εF (C) ◦ F (ηC ◦ uC ) = εF (C) ◦ F (ηC ) ◦ F (uC ) = F (uC )
.
= εF (C) = εF (C) ◦ F 1GF (C)
Let .g ∈ HomD (F (C), F (C )). The naturality of .η applied to .vC renders the
following diagram commutative:
(3.41)
Therefore, for all .C ∈ Ob C, we obtain
(3.41) (3.16)
GF (vC ) ◦ ηGF (C) = ηC ◦ vC = 1GF (C) = G(εF (C) ) ◦ ηGF (C) .
.
Now Corollary 3.6.2, (1) implies .F (vC ) = εF (C) . Finally, the naturality of .ε
applied to g yields
(3.42)
Putting all this together we obtain
(3.15)
g = g ◦ 1F (C) = g ◦ εF (C) ◦ F (ηC )
.
(3.42)
= εF (C ) ◦ F G(g) ◦ F (ηC ) = F vC ◦ G(g) ◦ ηC ,
This section is devoted to a special kind of adjunction, namely those for which one
of the functors involved is an inclusion.
Definition 3.7.1 A full subcategory .A of .B is called reflective if the inclusion
functor .I : A → B admits a left adjoint, called a reflector. Dually, a full subcategory
.A of .B is called coreflective if the inclusion functor admits a right adjoint, called a
coreflector.
We have already encountered many examples of such subcategories:
Examples 3.7.2
(1) Ab is a reflective subcategory of Grp, as shown in Example 3.4.1, (4).
(2) Grp is both a reflective and a coreflective subcategory of Mon as shown in
Example 3.6.3, (3) and (4).
(3) KHaus is a reflective subcategory of Top. The Stone–Čech compactification
provides the reflector, as shown in Example 3.6.3, (2).
(4) Haus is a reflective subcategory of Top. The Hausdorff quotient provides the
reflector, as shown in Example 3.6.3, (6).
(5) Ring is not a reflective subcategory of Rng. Indeed, note that although the inclu-
sion functor .I : Ring → Rng has a left adjoint, as shown in Example 3.6.3, (7)
the category Ring is not a full subcategory of Rng. .
. (3.43)
. (3.44)
We will prove first that . R(L), (qj : R(L) → F (j ))j ∈Ob J is a cone on F .
Indeed, for all .d ∈ HomJ (j, l) we have
(3.44) (3.43) (3.44)
I F (d)◦qj ◦ηL =I F (d)◦I qj ◦ ηL = I F (d) ◦ pj = pl = I ql ◦ ηL .
.
. (3.45)
196 3 Adjoint Functors
(3.45) (3.44)
Thus, for any .j ∈ Ob J we have .pj ◦ f ◦ηL = I (qj )◦ηL = pj = pj ◦1L
and using Proposition 2.2.14, (1) we obtain
f ◦ ηL = 1L .
. (3.46)
ηL ◦ f = I (t).
. (3.47)
Moreover, we have
(3.47) (3.46)
I (t) ◦ ηL = ηL ◦ f ◦ ηL = ηL = I (1R(L) ) ◦ ηL .
.
Using again Corollary 3.6.2, (1) we get .t = 1R(L) and hence .ηL ◦ f = 1I R(L) ,
so .ηL is an isomorphism, as desired.
Consider now another
cone . L , (tj : L → F (j ))j ∈Ob J on F . Then
. I (L ), (I (tj )j ∈Ob J is a cone on I F . Therefore, there exists a unique mor-
phism .g ∈ HomB (I (L ), L) such that the following diagram is commutative
for all .j ∈ Ob J :
. (3.48)
I (qj ) ◦ ηL ◦ g = I (tj ).
. (3.49)
As .ηL ◦ g ∈ HomB (I (L ), I R(L)) and I is fully faithful, there exists a unique
morphism .h ∈ HomA (L , R(L)) such that .I (h) = ηL ◦g. Then (3.49) becomes
.I (qj ◦h) = I (tj ) and since I is fully faithful we get .qj ◦h = tj for all .j ∈ Ob J ,
3.7 (Co)reflective Subcategories 197
The proof will be finished once we show that h is the unique morphism which
makes the above diagram commutative. Indeed, suppose there exists an .h ∈
HomA (L , R(L)) such that .qj ◦ h = tj for all .j ∈ Ob J . Then we also have
.I (qj ) ◦ I (h) = I (tj ) and using (3.44) and respectively (3.48) we get
−1
pj ◦ η L
. ◦ I (h) = pj ◦ g
−1
for all .j ∈ Ob J . Proposition 2.2.14, (1) implies .ηL ◦ I (h) = g and thus
.I (h) = ηL ◦ g. Since h is the unique morphism such that .I (h) = ηL ◦ g, we
get .h = h. This shows that . R(L), (qj )j ∈Ob J is indeed the limit of F and
therefore .A is a complete category.
(2) Let .A be a coreflective subcategory of a cocomplete category .B and denote by
.C : B → A the right adjoint of the inclusion functor .I : A → B and by .ε the
op
where .qj is the unique morphism which makes the following diagram commu-
tative:
op
and . L, (pj ∈ HomBop (L, I op F op (j )))i∈Ob J is the limit of the functor
.I
op F op : J op → Bop . In other words, for all .j ∈ Ob J, q is the unique
j
morphism in .A such that .εL ◦ I (qj ) = pj . Now we can conclude by
198 3 Adjoint Functors
Lemma 2.2.12 that . C(L), (qj ∈ HomA (F (j ), C(L)))i∈Ob J is the colimit
of the functor F and therefore .A is cocomplete.
In light of Proposition 3.7.3 the next natural question we are led to consider con-
cerns the cocompleteness of reflective subcategories (and, dually, the completeness
of coreflective subcategories).
Proposition 3.7.4 Let .A be a subcategory of .B.
(1) If .A is a reflective subcategory of a cocomplete category .B then .A is also
cocomplete.
(2) If .A is a coreflective subcategory of a complete category .B then .A is also
complete.
Proof
(1) Let .I : A → B be the inclusion functor and .R : B → A the reflector. Let
.
is the colimit of the functor .RI F : J → A. By Lemma 3.6.6, (3) we know that
the counit .ε : RI → 1A of the adjunction .R I is a natural isomorphism.
Therefore, the natural transformation .εF : RI F → F defined by
op
where . D, (qj )j ∈Ob J is the colimit of the functor .I op F op : J op → Bop . Now
Lemma 2.2.12 implies that . C(D), (ηF−1(j ) ◦ C(qj ))j ∈Ob J is the limit of the
functor F .
When studying categories which are practically the same, the first notion we usually
encounter is that of an isomorphism of categories, as introduced in Definition 1.6.6.
However, this concept turns out to be too strict, as there are many examples
of categories with similar properties (such as completeness, cocompleteness etc.)
which are not isomorphic. To express that two categories share many of the same
properties, a more suitable notion than isomorphism is the following:
Definition 3.8.1 A functor .F : C → D is called an equivalence of categories
and the category .C is said to be equivalent to .D if there exists another functor
.G : D → C such that we have natural isomorphisms .GF ∼ = 1C and .F G ∼ = 1D .
A contravariant functor .F : C → D for which .F : Cop → D is an equivalence of
categories is called a duality of categories.
Example 3.8.2 Given a field K, the category .MatK defined in Example 1.2.2, (19)
is equivalent to the category of finite-dimensional K-vector spaces .K Mf d . Indeed,
the functor .F : MatK → K Mf d defined below is an equivalence of categories:
To this end, we choose a basis6 .BV for each finite dimensional vector space V
and we define a functor .G : K Mf d → MatK as follows:
have
GF (n) = G K n = dim K n = n = 1MatK (n).
.
1 on the i-th (resp. j -th) position and zeros elsewhere for all n.i = 1, 2, . . . , m
(resp. .j = 1, 2, . . . , n), we obtain .MA (ei ) = Aei = j =1 aj i fj , where
.A = akl . This proves that .UMA = A, i.e., .GF (A) = A, as desired.
k=1, n, l=1, m
Hence, we have proved that .GF = 1MatK , which shows that, in particular, GF is
naturally isomorphic to .1MatK .
We are left to show that F G is naturally isomorphic to .1 f d . Consider
KM
.η : 1 f d → F G defined for any vector space V by .ηV : V → K
dim(V ) , η (v) =
KM
V
[v], where we denote by .[v] the (column) coordinate vector of v with respect to the
chosen basis of V . We claim that .η is a natural isomorphism. To start with, each
.ηV is clearly a linear bijection. We are left to check the naturality condition. To
where .m = dim(V ), n = dim(W ). The proof will be finished once we show that
the following diagram is commutative:
. (3.50)
If we define .Uα = ukl k=1, n, l=1, m , then for any .v = m vi ti ∈ V we have
m m n ni=1 m
.α(v) = v
i=1 i α(t i ) = v
i=1 i u w
j =1 j i j = j =1 vi uj i wj .
i=1
This shows that the j -th component of the column vector .[α(v)] is . m i=1 vi uj i ,
for all .j = 1, 2, . . . , n. Moreover, a similar straightforward computation shows
that the j -th component of .Uα [v] is . m i=1 uj i vi for all .j = 1, 2, . . . , n. Putting
everything together we have
[α(v)] = Uα [v].
. (3.51)
Remark 3.8.3 The categories .MatK and .K M from the previous example are
fd
equivalent but not isomorphic. Indeed, this follows easily by noticing that .MatK is
a small category while .K Mf d has a class of objects.
Proposition 3.8.4 Let .A, B and .C be three categories. The following hold:
(1) any category is equivalent to itself;
(2) if .A is equivalent to .B then .B is equivalent to .A;
(3) if .A is equivalent to .B and .B is equivalent to .C then .A is equivalent to .C.
Proof
(1) Any category .A is equivalent to itself as the identity functor .1A : A → A is
obviously an equivalence of categories.
(2) Assume that the category .A is equivalent to .B and .F : A → B is the
equivalence functor. Then there exists another functor .G : B → A and two
natural isomorphisms .GF ∼ = 1A and .F G ∼ = 1B . This shows that G is also an
equivalence of categories and therefore .B is equivalent to .A.
(3) Assume that .A is equivalent to .B and .B is equivalent to .C. Then, we have two
pairs of functors and their corresponding natural isomorphisms
F : A → B, G : B → A,
. α : F G → 1B , β : GF → 1A ,
H : B → C, T : C → B, γ : H T → 1C , σ : T H → 1B .
.
202 3 Adjoint Functors
Note that the above natural transformations are in fact natural isomorphisms
since .αT and .σF are natural isomorphisms (Example 1.7.2, (7)) and all functors
preserve isomorphisms (Proposition 1.6.9, (1)). This gives rise to the following
natural isomorphisms:
(3.52)
(3.53)
From (3.52) and (3.53) we obtain .ηC ◦ h1 = ηC ◦ h2 and since .ηC is an
isomorphism we get .h1 = h2 , as desired. Similarly, using the naturality
of .ε it follows that G is faithful as well.
Consider now .C, C ∈ Ob C and .g ∈ HomD (F (C), F (C )). Now
define
−1
f = ηC
. ◦ G(g) ◦ ηC ∈ HomC (C, C ). (3.54)
−1 (3.52)−1 (3.54)
ηC
. ◦ GF (f ) ◦ ηC = f = ηC ◦ G(g) ◦ ηC .
. (3.55)
−1 −1
F G(g) = εD
. ◦ g ◦ εD and F G(g ) = εD
◦ g ◦ εD .
This yields
−1
F G(g ) ◦ G(g) = εD
.
◦ g ◦ g ◦ εD . (3.56)
(3.57)
By naturality of .ε applied to .F (f ) we have the following commuta-
tive diagram:
i.e., εF (C ) ◦ F GF (f ) = F (f ) ◦ εF (C)
.
⇔ F GF (f ) ◦ εF−1(C) = εF−1(C ) ◦ F (f )
⇔ F GF (f ) ◦ F (ηC ) = F (ηC ) ◦ F (f )
⇔ F GF (f ) ◦ ηC = F (ηC ◦ f ).
have
−1
Consider now .D ∈ Ob D and .εD : D → F G(D). From the naturality
−1
of .ε applied to the morphism .εD we obtain the following commutative
diagram:
(3) ⇒ (4) Assume now that .G : D → C is a functor such that .F G and the
.
.εF−1(C) = F (ηC ), −1
ηG(D) = G(εD ).
which shows that the compatibility conditions (3.15) and (3.16) are
fulfilled for .ε−1 and .η−1 .
.(4) ⇒ (3) Follows in the same fashion as .3) ⇒ 4). The proof is now finished.
As an application of the previous theorem we will highlight an equivalence of
categories involving ring localizations.
3.8 Equivalence of Categories 207
Example 3.8.6 Let R be a commutative ring with unity and .(S −1 R, j ) its local-
ization at the multiplicative set .S ⊂ R. We will show, using Theorem 3.8.5, (2) that
the category .S −1 R M of modules over the localization ring .S −1 R is equivalent to
the category .R MS−aut of modules over R on which S acts as automorphisms (see
Example 1.2.2, (14)).
Indeed, consider the restriction of scalars functor .Fj : S −1 R M →R M induced
by the ring homomorphism .j : R → S −1 R as defined in Example 1.5.3, (32). First
note that since .j : R → S −1 R is an epimorphism in .Ringc (see Example 1.3.2,
(5)), the corresponding restriction of scalars functor .Fj is fully faithful, as proved
in Example 1.6.2, (6).
Furthermore, one can easily show that S acts as automorphisms on .Fj (M), for
any .S −1 R-module M. Indeed, it is straightforward to see that for all .s ∈ S the
inverse of the multiplication map .μs : M → M, μs (m) = sm is given by the R-
linear homomorphism .μ 1 , where the juxtaposition denotes the R-module structure
s
on .Fj (M) = M. This proves that the image of the restriction of scalars functor .Fj
is contained in the category .R MS−aut .
Therefore, we have a fully faithful functor .Fj : S −1 R M → R MS−aut . We are left
to show that .Fj is essentially surjective as well. To this end, let .M ∈ Ob R MS−aut .
Then M admits an .S −1 R-module structure defined for all .r ∈ R, s ∈ S and .m ∈ M
as follows:
r
. m = rn, (3.58)
s
where the juxtaposition denotes the R-module structure on M and n is the unique
element of M such that .sn = m. Note that the existence and uniqueness of the
element n with this property is a consequence of M being an R-module on which
S acts as an automorphism. Moreover, if .M ∈ Ob S −1 R M with the .S −1 R-module
structure given in (3.58) then .Fj (M) has the R-module structure defined as follows
for all .r ∈ R and .m ∈ M:
r
j (r) m =
. m = rm,
1
i.e., it coincides with the initial R-module structure on M. This finishes the proof.
.
category.
As .D is a complete category, the functor .F H : J → D has a limit, say
. L, (pj : L → F H (j ))j ∈Ob J . Moreover, as G is right adjoint
to F , Theorem 3.4.4
implies that . G(L), (G(pj ) : G(L) → GF H (j ))j ∈Ob J is the limit of the functor
.GF H : J → C. Now .η
−1 : GF → 1
C (as defined in Example 1.7.2, (6))
−1
is a natural isomorphism and consequently .ηH : GF H → H (as defined in
Example
1.7.2, (7)) is also a natural isomorphism. Now
Lemma 2.2.15, (1) implies
−1
that . G(L), (ηH (j ) ◦ G(pj ) : G(L) → H (j ))j ∈Ob J is the limit of H and therefore
the category .C is complete.
The statement concerning cocompleteness follows similarly using Theo-
rem 3.8.5, (4).
Definition 3.8.8 A skeleton of a category .C is a full subcategory .C0 of .C such that
each object of .C is isomorphic to exactly one object of .C0 .
Example 3.8.9 A skeleton of a given category .C always exists; indeed, it can be
constructed by choosing8 an object from each isomorphism class of objects in .C
and considering the full subcategory of .C with this objects class. .
H (D) = G(D), D ∈ Ob D0 , .
. (3.59)
H (f ) = f , f ∈ HomD0 (D1 , D2 ). (3.60)
. H (g) ◦ H (f ) = g ◦ f = H (g ◦ f ).
This shows that H is a functor. The proof will be finished once we show that H is
the inverse of F . Indeed, for all .C ∈ Ob C0 and .D ∈ Ob D0 we have
(3.59) (3.59)
.F H (D) = F G(D) = D, H F (C) = GF (C) = C.
Similarly, if we have .t ∈ HomC0 (C1 , C2 ) = HomC0 (GF (C1 ), GF (C2 )), then
(3.60)
.H F (t) = t and the proof is now finished.
In light of our previous result, loosely speaking, we can conclude that two
equivalent categories might differ only by the numbers of isomorphic copies of the
same object. Another important consequence is the following:
Corollary 3.8.12 The skeleton of a category is unique up to isomorphism.
Proof As proved in Proposition 3.8.4, (1) any category .C is trivially equivalent to
itself. Now any two skeletons of .C are isomorphic by Proposition 3.8.11.
Categories with a small skeleton have been mentioned in passing in Exam-
ple 1.2.2, (5). We discuss them here in more detail.
Definition 3.8.13 A category is called essentially small if its skeleton is a small
category.
A useful characterization of essentially small categories is the following:
Proposition 3.8.14 A category is essentially small if and only if it is equivalent to
a small category.
Proof Consider .C to be a category equivalent to a small category .D. If .C0 and .D0
denote the skeleton of .C and .D, respectively, then in particular .D0 is also a small
210 3 Adjoint Functors
category. Now Proposition 3.8.11 implies that .C0 and .D0 are isomorphic categories
and therefore .C0 is small. This shows that .C is essentially small.
Conversely, if .C is essentially small then its skeleton .C0 is a small category and
the conclusion follows from Corollary 3.8.10.
Examples 3.8.15
(1) The categories FinSet and .K Mf d are essentially small. Indeed, a skeleton of
FinSet is given by its full subcategory whose objects are the sets .n, for all
∅ if n = 0
.n ∈ N, where .n = . The latter category is obviously
{1, . . . , n} if n ∈ N\{0}
small.
Furthermore, Example 3.8.2 shows that the category .K Mf d is equivalent
to the small category .MatK defined in Example 1.2.2, (19). Hence, .K Mf d is
essentially small by virtue of Proposition 3.8.14.
(2) The categories Set, Grp, Ring, Top and .K M are not essentially small. We only
prove the assertion regarding the category Set and leave the others to the reader.
To this end, assume there exists a small skeleton .C of Set. Given that .Ob C is a
set we can consider .Ob C = {Xi | i ∈ I }, where I is a set and .Xi ∈ Ob Set for
all .i ∈ I . Now let .X = P( i∈I Xi ) be the power set of the coproduct of the
family of objects .(Xi )i∈I in Set. As .C is assumed to be the skeleton of Set, there
exists some .i0 ∈ I and an isomorphism in Set (i.e., a set bijection) between X
and .Xi0 . From Cantor’s theorem9 we have .| i∈I Xi | < |X|. Furthermore, as
. i∈I Xi is the union of the sets .Xi = Xi × {i} (see Example 2.1.5, (9)) we also
have .|Xi | ≤ | i∈I Xi | for all .i ∈ I . Putting all this together leads in particular
to .|Xi0 | < |X| and we have reached a contradiction as X was assumed to be
isomorphic to .Xi0 . Therefore, Set cannot have a small skeleton, as desired. .
We end this section with a generic example of a duality of categories. Let .C
and .D be two small categories. We denote by .FunL (C, D) the full subcategory
of .Fun(C, D) consisting of all functors from .C to .D which admit a left adjoint.
Similarly, .FunR (C, D) denotes the full subcategory of .Fun(C, D) consisting of all
functors from .C to .D which admit a right adjoint.
Theorem 3.8.16 For all small categories .C and .D, we have a duality of categories
between .FunR (C, D) and .FunL (D, C).
Proof We will define an equivalence of categories
9 Cantor’s theorem: For any set X we have .|X| < |P(X)| ([37, Theorem 2.21]).
3.8 Equivalence of Categories 211
identities.
Consider now functors .Gi : D → C and choose .Fi : C → D to be the
corresponding left adjoints, .i = 1, 2, 3. Moreover, denote by .θ i the natural
isomorphism induced by the adjunction .Fi Gi , i = 1, 2, 3. If .α : G1 → G2
and .β : G2 → G3 are natural transformations and .H (α) = α, H (β) = β we aim to
show that .H (β◦α) = β ◦op α. This comes down to showing that for each .C ∈ Ob C,
the morphism . β ◦op α C = α C ◦β C is the unique one such that the following holds:
D (h ◦ α C ◦ β C ) = βD ◦ αD ◦ θC, D (h)
3 1
θC,
. (3.61)
for all .h ∈ HomD (F1 (C), D). To this end, recall that .α C and .β C are the unique
morphisms such that
D (f ◦ α C ) = αD ◦ θC, D (f ), .
2 1
.θC, (3.62)
D (g ◦ β C ) = βD ◦ θC, D (g)
3 2
θC, (3.63)
for all .f ∈ HomD (F1 (C), D), g ∈ HomD (F2 (C), D). Putting all the above
together yields
(3.62) (3.63)
.βD ◦ αD ◦ θC,
1
D (h) = βD ◦ θC, D (h ◦ α C ) = θC, D (h ◦ α C ◦ β C )
2 3
defined by
where .Fi : C → D is the left adjoint of .Gi and denote by .θ i the natural isomorphism
corresponding to the adjunction .Fi Gi , i = 1, 2.
Assume .α, β : G1 → G2 are natural transformations such that .HG1 , G2 (α) =
HG1 , G2 (β), i.e., .α = β. This implies .θC, 2 (f ◦ α ) = θ 2 (f ◦ β ) and
D C C, D C
consequently we have
αD ◦ θC,
.
1
D (f ) = βD ◦ θC, D (f ),
1
(3.64)
αD = θG
.
2
1 (D), D
1
(εD ◦ μG1 (D) ) ∈ HomD (G1 (D), G2 (D)),
where .ε1 denotes the counit of the adjunction .F1 G1 . We show first that the
morphisms .αD , D ∈ Ob D, form a natural transformation .α : G1 → G2 , i.e., for
all .r ∈ HomD (D, D ) the following diagram is commutative:
(3.65)
To start with, note that .ε1 ◦ μG1 : F2 G1 → 1D is a natural transformation and
therefore the following diagram is commutative:
(3.66)
3.8 Equivalence of Categories 213
Therefore, we have
1
G2 (r) ◦ αD = G2 (r) ◦ θG
.
2
1 (D), D
εD ◦ μG1 (D)
(3.2)
= θG
2
1 (D), D
r ◦ εD ◦ μG1 (D)
1
(3.66) 1
= θG
2
1 (D), D
εD ◦ μG1 (D ) ◦ F2 G1 (r)
(3.1) 1
= θG
2
1 (D ), D
εD ◦ μG1 (D ) ◦ G1 (r)
= αD ◦ G1 (r),
which shows that (3.65) indeed holds and hence .α is a natural transformation. The
proof will be finished if we show that .HG1 , G2 (α) = μ or, equivalently, that the
following holds for all .f ∈ HomD (F1 (C), D):
D (f ◦ μC ) = αD ◦ θC, D (f ).
2 1
θC,
.
. (3.67)
1 1
i.e., F1 θC, D (f ) ◦ μC = μG1 (D) ◦ F2 θC, D (f ).
αD ◦ θC,
.
1
D (f ) = θG1 (D), D (εD ◦ μG1 (D) ) ◦ θC, D (f )
2 1 1
(3.1) 2
1
= θC, D εD 1
◦ μG1 (D) ◦ F2 θC, D (f )
(3.67) 2
= θC, D εD 1
◦ F1 θC,
1
D (f ) ◦ μC
(3.25)
= θC,
2
D (f ◦ μC ),
of the most important ones: the category of compact topological abelian groups
is dual to the category of abelian groups (Pontryagin duality); the category of
commutative unital .C ∗ -algebras is dual to the category of compact Hausdorff
topological spaces (Gelfand-Naimark duality); the category of compact and totally
disconnected topological spaces10 is dual to the category of Boolean algebras (Stone
duality). For further details we refer the reader to [14, 43].
3.9 Localization
is commutative:
. (3.68)
Theorem 3.9.2 Let .C be a category. Then there exists a localization of .C by any set
of morphisms S of .C.
Proof In order to construct the localization of .C by the set S we start by defining an
oriented graph .Γ as follows:
• the vertices of .Γ are the objects of .C;
10 A topological space that is compact and totally disconnected is called a Stone space.
3.9 Localization 215
• the edges of .Γ are the morphisms of .C (any morphism .f ∈ HomC (X, Y ) is seen
Ob CS = Ob C;
.
HomCS (X, Y ) = {
. γ | γ is a path in Γ from X to Y },11 for all .X, Y ∈ Ob C,
with the composition of morphisms in .CS induced by the concatenation of paths and
the identity maps given by the trivial paths. The functor .F : C → CS is defined as
follows:
F (X) = X, for all .X ∈ Ob C;
.
diagram (3.68) commutative, then we have .H (X) = G(X) and .H (f) = G(f ),
for all .X ∈ Ob Cs = Ob C and .f ∈ HomC (X, Y ); furthermore, in order for H to
be a functor and to respect compositions and identities, it should satisfy .H xs =
G(s)−1 , for all .s ∈ S.
We are left to prove that H is well-defined. To this end, consider two paths u
and v in .Γ such that . u =
v . Since the paths u and v are equivalent, we can turn
u into v after a finite number of elementary operations. Thus, it suffice to prove
that by applying H to each of these elementary operations we obtain equalities
H (g
. g ) ◦ H (f) = H (
◦ f ) = G(g ◦ f ) = G(g) ◦ G(f ) = H ( g ◦ f).
xs ◦
H (
. s) = H ( s) = G(s)−1 ◦ G(s) = 1G(X) = G(1X ) = H (1X ).
xs ) ◦ H (
G ◦ F = F.
. (3.69)
G ◦ F = F.
. (3.70)
(3.70) (3.69)
F = G ◦ F = G ◦ G ◦ F = (G ◦ G) ◦ F.
. (3.71)
Applying Definition 3.9.1 to the pair .(CS , F ), seen both as a localization and as the
other pair, yields a unique functor .H : CS → CS such that .H ◦ F = F . By the
uniqueness of H we must have .H = 1CS . Moreover, since by (3.71) the functor
3.9 Localization 217
. (3.72)
Similarly one can prove that .G ◦ G = 1C and therefore the categories .CS and .CS
S
are isomorphic, as desired. The proof is now finished.
One of the situations when the localization of a category can be described, up to
equivalence of categories, even without assuming the localizing class of morphisms
to be a set, is that of reflective subcategories.
Theorem 3.9.4 Let .I : A → B be a reflective subcategory inclusion with reflector
R : B → A and denote by S the class of all morphisms s of .B such that .R(s) is an
.
for all .B ∈ Ob B.
Define a category .BS as follows:
Ob BS = Ob B;
.
Recall that S is the class of all morphisms s of .B such that .R(s) is an isomorphism
and therefore .F (s) is obviously an isomorphism for any .s ∈ S.
Consider now another functor .G : B → D such that .G(s) is an isomorphism
for all .s ∈ S. We need to find a functor .H : BS → D which makes the following
218 3 Adjoint Functors
diagram commutative:
. (3.73)
Having in mind that .ηB ∈ S for any .B ∈ ObB, it can be easily seen that a functor
H which makes the above diagram commute has the following property for all .B ∈
ObB:
(3.15) −1
H (εR(B) ) = H R(ηB )−1 = H R(ηB )
.
−1 (3.73)
= H F (ηB ) = G(ηB )−1 . (3.74)
Furthermore, for any morphism .f ∈ HomBS (B, B ) = HomA (R(B), R(B )), the
naturality of .ε : RI → 1A renders the following diagram commutative:
(3.75)
Therefore, for any .f ∈ HomBS (B, B ) = HomA (R(B), R(B )) we have
(3.15)
.H (f ) = H f ◦ εR(B) ◦ R(ηB )
(3.75)
= H εR(B ) ◦ RI (f ) ◦ R(ηB )
= H εR(B ) ◦ H RI (f ) ◦ H R(ηB )
(3.74)
= G(ηB )−1 ◦ H F I (f ) ◦ H F (ηB )
(3.73) −1
= G(ηB ◦ GI (f ) ◦ G(ηB ).
The above discussion proves that H is the unique functor which might render dia-
gram (3.73) commutative. We are left to prove that indeed H makes diagram (3.73)
commute. To this end we will use the naturality of .η, i.e., the commutativity of the
above diagram for any .g ∈ HomB (B, B ):
. (3.76)
Next we show that the category .BS is equivalent to .A. Indeed, consider the
functor .T : A → BS defined as follows:
T (A) = I (A), for all .A ∈ Ob A;
.
. (3.77)
−1
u = εA ◦ v ◦ εA
. ∈ HomC (A, A ). (3.78)
We will prove that .RI (u) = v. Indeed, using again the naturality of .ε we obtain
−1 (3.77) (3.78) −1
εA ◦ RI (u) ◦ εA
. = u = εA ◦ v ◦ εA .
Since .εA and .εA are isomorphisms we get .RI (u) = v and we have proved that T is
full.
Moreover, for any .B ∈ Ob BS we have an isomorphism
R(ηB ) ∈ HomA (R(B), RI R(B)) = HomBS (B, I R(B)) = HomBS (B, T R(B)),
.
L(M) = S −1 M,
. L(f ) = f,
x f (x)
f: S −1 M → S −1 N, f = , for all x ∈ M, s ∈ S.
s s
Note that we see .S −1 M as an R-module via j and by [3, Proposition 12.1], the
multiplication map .μs : S −1 M → S −1 M is bijective for all .s ∈ S and therefore L
is well-defined.
We will show that L is left adjoint to the inclusion functor I . To this end, let
.ϕ : 1 → I L be the natural transformation defined for all R-modules M by
RM
the R-module homomorphism .ϕM : M → S −1 M associated with the localization
3.10 (Co)limits as Adjoint Functors 221
m f (m)
.f◦ ϕM (m) = f = = ϕN (f (m)) = ϕN ◦ f (m),
1 1
which shows that .ϕ is a natural transformation, as claimed.
Consider .u ∈ HomR M (M, I (N )), with .N ∈ Ob R MS−aut . Now let
.u : S
−1 M → N be defined for all .x ∈ M and .s ∈ S by .u x = y, where y
s
is the unique element of N such that .sy = u(x); note that since the multiplication
by s is a bijection on N, we have a unique such y. It is straightforward to see that
.u is a well-defined R-module homomorphism. Furthermore, .u : S
−1 M → N is
We start by recalling from Theorem 2.2.16 (resp. Theorem 2.2.17) that taking
(co)limits yields a functor. It turns out that in certain conditions this limit (resp.
colimit) functor has a left (resp. right) adjoint, namely the diagonal functor defined
in Proposition 1.9.8.
Theorem 3.10.1 Let I be a small category and .C an arbitrary category.
(1) The diagonal functor .Δ : C → Fun(I, C) has a right adjoint if and only if .C is
complete. In this case, the right adjoint is the limit functor .lim : Fun(I, C) →
C.
(2) The diagonal functor .Δ : C → Fun(I, C) has a left adjoint if and only
if .C is cocomplete. In this case, the left adjoint is the colimit functor
.colim : Fun(I, C) → C.
Proof
(1) Assume first that any functor .F : I → C, where I is a small category,
has a limit. We will define a bijective map .θ : HomFun(I, C) (Δ, −) →
HomC (−, lim), natural in both variables. To this end, for any .X ∈ Ob C,
222 3 Adjoint Functors
.F ∈ Ob Fun(I, C) and any natural transformation .α : Δ(X) → F , we define
.θX, F (α) = f , where .f : X → lim F is the unique morphism in .C which makes
. (3.79)
and . lim F, (pi : lim F → F (i))i∈Ob I denotes the limit of F .
First we prove that each map .θX, F is bijective. Indeed, consider two
natural transformations .α, .β ∈ HomFun(I, C) (Δ(X), F ) such that .θX, F (α) =
θX, F (β) = f . This implies that for all .i ∈ Ob I we have .pi ◦ f = αi and
.pi ◦ f = βi . Hence .αi = βi for all .i ∈ Ob I , which implies that the two natural
Indeed, recall that . lim F, (pi : lim F → F (i))i∈Ob I is in particular a cone on
F and therefore we have .F (u)◦pi = pj . This yields .F (u)◦αi = F (u)◦pi ◦f =
pj ◦ f = αj , which shows that the above diagram is indeed commutative.
Next we show that .θ is natural in both variables. First, consider .f ∈
HomC (X , X). We will prove the commutativity of the following diagram,
which ensures the naturality in the first variable:
i.e.,
. HomC (f, lim F ) ◦ θX, F = θX , F ◦ HomFun(I, C) (Δ(f ), F ).
(3.80)
3.10 (Co)limits as Adjoint Functors 223
. (3.81)
Hence, we are left to show that .θX , F α ◦ Δ(f ) = t ◦ f . Having in mind the
way .θ was defined, this comes down to proving that .t ◦ f makes the following
diagram commutative for all .i ∈ Ob I :
(3.81)
Indeed, we have .pi ◦ t ◦f = αi ◦f for all .i ∈ Ob I and this shows that (3.80)
holds.
Consider now two functors F , .F : I → C and .β ∈ HomFun(I, C) (F, F ),
i.e., .β : F →
F is a natural transformation. We denote by . lim F, (pi : lim F →
i.e.,
. HomC (X, lim β) ◦ θX, F = θX, F ◦ HomFun(I, C) (Δ(X), β).
(3.82)
224 3 Adjoint Functors
(3.83)
We are left to show that .lim β ◦ t = r. To this end, recall that .lim β ∈
HomC (lim F, lim F ) is the unique morphism in .C which makes the following
diagram commute for all .i ∈ Ob I :
. (3.84)
By Proposition 2.2.14, (1) we only need to show that .si ◦lim β ◦t = si ◦r for all
(3.84) (3.83) (3.83)
.i ∈ Ob I . Indeed, we have .si ◦ lim β ◦t = βi ◦pi ◦ t = βi ◦ γi = si ◦r,
as desired.
Assume now that the diagonal functor .Δ : C → Fun(I, C) has a right
adjoint, denoted by .R : Fun(I, C) → C. We will show that any func-
tor .F : I → C has a limit. To this end, let .ε : ΔR → 1Fun(I, C) and
.θ : HomFun(I, C) (Δ, −) → HomC (−, R) be the counit and respectively the
yields a unique morphism .g ∈ HomC (X, R(F )) such that the following
diagram is commutative:
. (3.85)
The above equality between the natural transformations .εF ◦ Δ(g) and .α comes
down to identities between the corresponding morphisms associated to each
.i ∈ Ob I . In light of Example 1.7.2, (5) we have .(εF )i ◦ g = αi for all .i ∈ Ob I
The proof will be finished once we show that g is the unique morphism in .C
which makes the above diagram commutative. To this end, assume that .h ∈
HomC (X, R(F )) such that .(εF )i ◦ h = αi for all .i ∈ Ob I . This leads to
.εF ◦Δ(h) = α and the uniqueness of the morphism which makes diagram (3.85)
. (3.86)
Therefore, for all .i ∈ Ob I we have . Δ(G(u)) i ◦ (qj )i = (qt )i , which comes down
to .G(u) ◦ (qj )i = (qt )i , i.e., the following diagram is commutative:
. (3.87)
. (3.88)
3.11 Freyd’s Adjoint Functor Theorem 227
The proof will be finished once we show that .g = (gi )i∈Ob I : U → Δ(L) is a
natural transformation. To this end, let .v ∈ HomI (i, s); we are left to prove the
commutativity of the following diagram:
. (3.89)
In light of Proposition 2.2.14, (1) it is enough to prove that for all .j ∈ Ob J we have
pj ◦ gi = pj ◦ gs ◦ U (v). Indeed, we have
.
(3.88) (3.88)
.pj ◦ gs ◦ U (v) = (qj )s ◦ U (v) = (qj )i = pj ◦ g i ,
where the second equality holds because .qj : U → Δ(G(j )) is a natural transfor-
mation.
Theorem 3.4.4 shows that right (left) adjoints preserve all existing small limits
(colimits). However, in general, small limit/colimit preservation alone does not
guarantee the existence of a left/right adjoint. Indeed, consider the unique functor
.T : Set(⊆) → 1, where .1 is the discrete category with one object and .Set(⊆)
is the category defined in Example 1.2.2, (2). Note that the category .Set(⊆) is
cocomplete by Example 2.2.11 and does not posses a final object as shown in
Example 1.3.10, (6). Therefore, T does not admit a right adjoint as can easily be
seen from Example 3.4.1, (5) while it trivially preserves small colimits.
In this section we will prove that limit/colimit preservation is part of a necessary
and sufficient condition which needs to be fulfilled by a functor in order to admit
a left/right adjoint. Let .G : D → C be a functor, .X ∈ Ob C and let .(X ↓ G) be
the comma category defined in Corollary 1.8.6, (1). We have an obvious forgetful
functor .U : (X ↓ G) → D defined for any .(f, Y ) ∈ Ob X ↓ G and any morphism
h in .(X ↓ G) as follows:
U (f, Y ) = Y, U (h) = h.
.
(2) The functor G admits a right adjoint if and only if for all .X ∈ Ob C the comma
category .(G ↓ X) has a final object.
Proof
(1) Suppose first that G has a left adjoint .F : C → D and let .θ : HomD (F (−), −)
→ HomC (−, G(−)) be the natural isomorphism corresponding to the adjunc-
tion .F G. Now consider .X ∈ Ob C and let .η : 1C → GF be the
unit of the adjunction. We will prove that . ηX , F (X) is the initial object
of the category .(X ↓ G). Let .(v, W ) be another object in .(X ↓ G),
i.e., .W ∈ Ob D and .v ∈ HomC (X, G(W )). To this end, we need to
find a unique morphism .f : ηX , F (X) → (v, W ) in .(X ↓ G), i.e.,
a morphism .f ∈ HomD (F (X), W ) such that the following diagram is
commutative:
. (3.90)
Recall from (the proof of) Theorem 3.5.1 that for all .u ∈ HomD (F (X), W )
−1
we have .G(u) ◦ ηX = θX, W (u). Now if we consider .f = θX, W (v) we
obtain
v = θX, W (f ) = G(f ) ◦ ηX ,
.
as desired. The uniqueness of f with this property follows from the bijectivity
of .θX, W .
Conversely, assume now that for each .X ∈ Ob C the comma category
.(X ↓ G) has an initial object, which we denote by .(uX , VX ), where
.VX ∈ Ob D and .uX ∈ HomC (X, G(VX )). Hence, for any .(f, Y ) ∈
Ob (X ↓ G) there exists a unique morphism .h : (uX , VX ) → (f, Y ) in
.(X ↓ G); in other words, for any .f ∈ HomC (X, G(Y )) there exists a
. (3.91)
GF (f ) ◦ uX = uX ◦ f.
. (3.92)
desired.
(2) Theorem 3.4.3 shows that .G : D → C admits a right adjoint if and only if
.G
op : Dop → Cop admits a left adjoint. We have already proved in .1) that
.G
op has a left adjoint if and only if for all .X ∈ Ob Cop the comma category
an isomorphism of categories between .(X ↓ Gop )op and .(G ↓ X). Now
using Proposition 1.4.3, (2) we obtain that .(X ↓ Gop ) has an initial object
if and only if its opposite category, namely .(G ↓ X), has a final object.
By putting all the above together we obtain that G has a right adjoint if
and only if for all .X ∈ Ob C the comma category .(G ↓ X) has a final
object.
Definition 3.11.2 Let .C be a category.
(1) A family .(Ki )i∈I of objects of .C, where I is a set, is called a weakly initial set if
j
for any .C ∈ Ob C there exists a morphism .tC ∈ HomC (Kj , C) for some .j ∈ I .
(2) Dually, a family .(Wi )i∈I of objects of .C, where I is a set, is called a weakly
final set if it is a weakly initial set in .Cop ; that is, if for any .C ∈ Ob C there
j
exists a morphism .lC ∈ HomC (C, Wj ) for some .j ∈ I .
Lemma 3.11.3 Let .C be a category. Then:
(1) if .C is complete then .C has an initial object if and only if .C has a weakly initial
set;
(2) if .C is cocomplete then .C has a final object if and only if .C has a weakly final
set.
Proof
(1) Assume first that .C has an initial object I ; then .{I } is obviously a weakly initial
set.
230 3 Adjoint Functors
Conversely, let .(Ki )i∈I be a weakly initial set. As .C is complete and I is a
set we can consider the product . P , (πi : P → Ki )i∈I of the family of objects
.(Ki )i∈I . Notice that for each .C ∈ Ob C there exists at least one morphism
We will prove that L is the initial object of the category .C. Indeed, for
any .C ∈ Ob C there exists at least one morphism in .HomC (L, C) given
by the composition . Suppose now that we have two
such morphisms f , .g ∈ HomC (L, C) and consider .(E, e : E → L) to
be the equalizer of .(f, g). Since .E ∈ Ob C there exists a morphism .uE ∈
HomC (P , E) given by the composition for some .j ∈ I .
Thus .q ◦ e ◦ uE ∈ HomC (P , P ) and since .(L, q : L → P ) is in particular a
cone on F , the following diagram is commutative:
f = f ◦ 1L = f ◦ e ◦ uE ◦ q = g ◦ e ◦ uE ◦ q = g ◦ 1L = g,
.
where in the third equality we used the fact that .(E, e : E → L) is the equalizer
of .(f, g). We have obtained .f = g and hence L is an initial object of .C.
(2) If .C is cocomplete then .Cop is complete and, as proved in 1), .Cop has an initial
object if and only if .Cop has a weakly initial set. Equivalently, .C has a final
object if and only if .C has a weakly final set.
We are now ready to state the main result of this section:
Theorem 3.11.4 (Freyd’s adjoint functor theorem) Let .G : D → C be a
functor.
(1) If .D is a complete category then G has a left adjoint if and only if G preserves
all small limits and for each .X ∈ Ob C the comma category .(X ↓ G) has a
weakly initial set.
3.11 Freyd’s Adjoint Functor Theorem 231
Proof
(1) Suppose G has a left adjoint F . Then G is a right adjoint to F and by The-
orem 3.4.4 it preserves limits.
Moreover, by (the proof of) Lemma 3.11.1, (1)
for any .X ∈ Ob C, the pair . ηX , F (X) is an initial object in .(X ↓ G), where
.η : 1C → GF is the unit of the adjunction .(F, G).
.(X ↓ G ) is a weakly final set in .(G ↓ X). Putting all of the above together
op
we can conclude that G has a right adjoint if and only if G preserves all small
colimits and for each .X ∈ Ob C the comma category .(G ↓ X) has a weakly
final set, as desired.
Theorem 3.11.4 can be stated in an equivalent form without the use of comma-
categories. To this end, we introduce the following:
Definition 3.11.5 Let .F : C → D be a functor and .D ∈ Ob D. Then:
(1) F satisfies the solution set condition with respect to D if there exists a set .UD
of objects of .C such that for any .C ∈ Ob C and any .f ∈ HomD (D, F (C)),
there exists an object .C ∈ UD and morphisms .u ∈ HomC (C , C), .g ∈
HomD (D, F (C )) such that the following diagram is commutative:
(2) F satisfies the cosolution set condition with respect to D if there exists a set .WD
of objects of .C such that for any .C ∈ Ob C and any .f ∈ HomD (F (C), D),
there exists an object .C ∈ WD and morphisms .u ∈ HomC (C, C ), .g ∈
232 3 Adjoint Functors
u ◦ qi = αi .
. (3.93)
Δ(u) ◦ β = α.
. (3.94)
Proof
(1) Note that G satisfies the solution set condition with respect to X if and only
if the comma category .(X ↓ G) has a weakly initial set and the conclusion
follows by Theorem 3.11.4, (1). Indeed, assume first that G satisfies the solution
set condition, let .X ∈ Ob C and let .UX be the corresponding set as in
Definition 3.11.5, (1). Then .{(h, Y ) | h ∈ HomC (X, G(Y )), Y ∈ UX } is a
weakly initial set in .(X ↓ G). Conversely, if .K = {(t, Z) | Z ∈ Ob D, t ∈
HomC (X, G(Z))} is a weakly initial set in .(X ↓ G) then .UX = {D ∈
Ob D | there exists t ∈ HomC (X, G(Z)) such that (t, Z) ∈ K} is a set
which fulfills the condition in Definition 3.11.5, (1) and therefore G satisfies
the solution set condition.
(2) Similarly, G satisfies the cosolution set condition with respect to X if and only
if the comma category .(G ↓ X) has a weakly final set and Theorem 3.11.4, (2)
leads to the desired conclusion.
We end this section with some applications of the adjoint functor theorem. As
we will see, it can be used to show the existence of various free objects (such as free
groups, free algebras, free modules etc.) without explicitly constructing them.
Examples 3.11.8
(1) Let .U : Grp → Set be the forgetful functor. We will show that U admits a
left adjoint by using Freyd’s adjoint functor theorem. To start with, .Grp is
complete as shown in Example 2.4.3, (1) and U preserves small limits by
Example 2.5.10, (1). In order to conclude that U admits a left adjoint we
need to show that it satisfies the solution set condition with respect to each
.X ∈ Ob Set. To this end, for a given .X ∈ Ob Set we consider the class
.λ = max.{ℵ0 , |X|}, where .ℵ0 denotes the cardinality of the set of all natural
numbers. First we show that .UX is in fact a set. Indeed, recall that there is
only a set of composition laws (in particular, of composition laws which are
group structures) on any given set. Thus, we can conclude that we have, up to
isomorphism, only a set of group structures on any set whose cardinality is at
most .λ. This shows that .UX , being a reunion of sets, is a set itself.
In order to show that .UX satisfies the conditions in Definition 3.11.5, (1)
we start by proving that if .(gx )x∈X is a set of elements of a group G then the
subgroup .GX of G generated by this set has cardinality at most .λ. This can be
easily seen by observing that the following map is surjective:
f:
. (X × Z)n → GX , f (x1 , e1 ), (x2 , e2 ), . . . (xn , en ) = gxe11 gxe22 . . . gxenn ,
n∈N
234 3 Adjoint Functors
where the domain of f is the coproduct of the family .{(X × Z)n }n∈N in Set. We
have
n
.|GX | (X × Z) = |(X × Z)n |
n∈N n∈N
ℵ0 , if |X| ℵ0
= max{ℵ0 , |X|} = λ.
|X|, if |X| ℵ0
Consider now .f ∈ HomSet (X, U (G)) for some .G ∈ Ob Grp and let .G
be the subgroup of G generated by the set .(f (x))x∈X . Then, according to the
above discussion, .G is a group of cardinality at most .λ and we can find a
group .H ∈ UX and a group isomorphism .t : H → G . Denote by .i : G → G
the inclusion map and consider .i ◦ t ∈ HomGrp (H, G) and .U (t −1 ) ◦ f ∈
HomSet (X, U (H )) such that the following diagram is commutative:
We can now conclude by Freyd’s adjoint functor theorem that .U : Grp → Set
has a left adjoint.
(2) Let R be a ring, .M ∈ Ob MR , .N ∈ Ob R M and .BilM, N : Ab → Ab
the functor defined in Example 1.5.3, (28). We will show, using Freyd’s
adjoint functor theorem, that .BilM, N admits a left adjoint. By Example 2.5.4,
.BilM, N preserves limits. We are left to show that .BilM, N satisfies the
solution set condition for each .A ∈ Ob Ab. To this end, for a given .A ∈ Ob Ab
we denote by .UA the class of all isomorphism classes of abelian groups of
cardinality less than or equal to .λ = max.{ℵ0 , |A| · |M × N |}. It follows easily,
as in the previous example, that .UA is in fact a set.
We will show that .UA satisfies the conditions in Definition 3.11.5, (1).
Indeed, let .B ∈ Ob Ab and .f ∈ HomAb (A, BilM, N (B)). For all .a ∈ A
let .Pa = {f (a)(m, n) | m ∈ M, n ∈ N} and let .B be the abelian subgroup
of B generated by the set .P = ∪a∈A Pa . We start by proving that .B has
cardinality at most .λ. This can be easily seen by observing that the following
map is surjective:
n
.f: (P × Z)n → B , f (x1 , k1 ), (x2 , k2 ), . . . (xn , kn ) = ki xi ,
n∈N i=1
3.11 Freyd’s Adjoint Functor Theorem 235
where the domain of f is the coproduct of the family .{(P × Z)n }n∈N in Set. We
have
n
n
.|B | (P × Z) = (∪a∈A Pa ) × Z
n∈N n∈N
ℵ0 , if |A| · |M × N| ℵ0
|A| · |M × N|, if |A| · |M × N| ℵ0
max{ℵ0 , |A| · |M × N|} = λ.
functor and let .Xi = F (i), for all .i ∈ Ob I . Consider .λ = | i∈Ob I Xj | and
let .UF be the class of isomorphism classes of groups of cardinality less than or
equal to .λ, where . i∈Ob I Xj is the coproduct in Set of the underlying sets of
the family of groups .(Xi )i∈I . As in the first example, .UF can be easily proved
to be a set. Now let .C ∈ Ob Grp and .ψ : F → ΔC be a natural transformation.
Consider H to be the subgroup of C generated by . j ∈Ob I ψj (Xj ), where
.ψj : Xj → C, .j ∈ Ob I , are the group morphisms corresponding to the
and therefore .UF is indeed a solution set. We can now conclude by Freyd’s
adjoint functor theorem that .Δ has a left adjoint. Furthermore, we can derive
the cocompleteness of .Grp from Theorem 3.10.1, (2). .
In this section we show that under certain conditions on the domain category, a given
functor admits a left (right) adjoint if and only if it preserves small limits (colimits).
We start with some preparations.
Definition 3.12.1 Let .C be a category.
(1) We say that .C admits a set of generators (or separators) if there exists a set
of objects .{Si | i ∈ I } in .C with the property that for any two morphisms u,
.v ∈ HomC (A, B) such that .u = v there exists a morphism .g ∈ HomC (Sj , A),
z , if y = y ,
h(y) =
.
z , if y =
y,
where Z is a set with at least two elements and .z = z . This leads to .h(u(x0 )) =
h(y ) = z = z = h(y ) = h(v(x
0 )) and the desired conclusion follows.
(2) In Grp, the group of integers . Z, + is a generator. To this end, let u, .v ∈
HomGrp (G, H ) such that .u = v and consider .g0 ∈ G such that .u(g0 ) = v(g0 ).
Now define .g : Z → G by .g(k) = g0k for all .k ∈ Z, where the group structure on
G is considered to be multiplicative. It can be easily seen that g is a morphism
of groups and moreover .u(g(1)) = v(g(1)), which leads to .u ◦ g = v ◦ g, as
desired.
On the other hand, Grp has no cogenerators. To this end, assume there
exists a cogenerator U in Grp and let S be a simple group whose cardinality is
larger than that of U .12 Let .IdS , .0S ∈ HomGrp (S, S), where .IdS denotes the
identity morphism on S while .0S is defined by .0S (s) = 1S for all .s ∈ S.
As S is a simple group we have .IdS = 0S and since U is assumed to be
a cogenerator in Grp, we have a group homomorphism .f : S → U such
that:
f ◦ IdS = f ◦ 0S .
. (3.95)
12 Such a group is, for instance, the projective special linear group .PSL(2, k), where k is the field
of rational functions over the complex numbers in .|U | variables, i.e., .k = C(Xu )u∈U (see [50,
Theorem 9.46]).
238 3 Adjoint Functors
13 A topological space X is called normal if for each pair U , V of disjoint closed subsets of X there
is a closed interval then there exists a continuous map .f : X → [a, b] such that .f (x) = a for every
.x ∈ U and .f (x) = b for every .x ∈ V ([39, Theorem 33.1]).
3.12 Special Adjoint Functor Theorem 239
. (3.96)
is a monomorphism.
(2) Dually, assume .C has coproducts, S is a set of generators and consider
the coproduct . Q, (qf )f ∈Hom (Xi , C) of the family S, where the coproduct
C
consists of as many copies of .Xi as there are morphisms in .HomC (Xi , C).
Then, for any .C ∈ Ob C, the unique morphism .ξC : P → C which makes the
following diagram commutative for all .i ∈ I and all .f ∈ HomC (Xi , C):
. (3.97)
is an epimorphism.
Proof
(1) Consider u, .v ∈ HomC (A, C) such that .γC ◦ u = γC ◦ v. This implies
.pf ◦ γC ◦ u = pf ◦ γC ◦ v for all .i ∈ I and all .f ∈ HomC (C, Xi ).
Proof
(1) If F admits a left adjoint then F is a right adjoint and it preserves limits by
Theorem 3.4.4.
Assume now that F preserves small limits and consider .D ∈ Ob D. In light
of Theorem 3.11.7, (1) it suffices to find a solution set for D. To this end,
consider a coseparating set .S = {Gi | i ∈ I } of .C and denote by
.P , pf i∈I
f ∈Hom (C, Gi )
C
the product of the family S, where the product consists of as many copies of .Gi
as there are elements in .HomC (C, Gi ), where C is a fixed object. Similarly, we
consider the product
.P , qf i∈I
f ∈Hom
D (D, F (Gi ))
of the same family S, but this time consisting of as many copies of .Gi as there
are morphisms in .HomD (D, F (Gi )).
Let .UD = {T | T is a subobject of P }. Note that .UD is in fact a set as .C is
well-powered.
Consider now .g ∈ HomD (D, F (C)). By Proposition 3.12.3, (1) the unique
morphism .αC ∈ HomC (C, P ) such that the following holds for all .i ∈ I and
all .f ∈ HomC (C, Gi )
. (3.98)
is a monomorphism.
As . P , pf i∈I is a product, there exists a unique morphism
f ∈Hom (C, Gi )
C
.βC : P → P such that the following diagram is commutative for all .i ∈ I and
. (3.99)
3.12 Special Adjoint Functor Theorem 241
. (3.100)
. (3.101)
Now
using again the fact that F is
limit preserving we obtain, in particular,
that . F (P ), F (pf ) i∈I is the product of the family .{F (Gi ) | i ∈
f ∈Hom (C, Gi )
C
I }, where the product consists of as many copies of .Gi as there are morphisms
in .HomC (C, Gi ). Therefore, for all .i ∈ I and all .f ∈ HomC (C, Gi ) we have
(3.98)
(3.101)
F (pf ) ◦ F (αC ) ◦ g = F (pf ◦ αC ) ◦ g = F (f ) ◦ g
. = F qF (f )◦g ◦ λ
(3.99)
= F pf ◦ βC ◦ λ = F (pf ) ◦ F (βC ◦ λ.
Proposition 2.2.14, (1) implies that .F (αC ) ◦ g = F (βC ◦ λ. As F is
limit preserving it follows that .(F (S), F (μ), F (γ )) is the pullback of the
pair of morphisms .(F (αC ), F (βC )) and we obtain a unique morphism .g ∈
HomD (D, F (S)) such that .F (μ) ◦ g = g and .F (γ ) ◦ g = λ. The complete
242 3 Adjoint Functors
To conclude, we have proved that for any .C ∈ Ob C and .g ∈ HomD (D, F (C))
there exists some .S ∈ UD together with morphisms .μ ∈ HomC (S, C) and
.g ∈ HomD (D, F (S)) such that .F (μ) ◦ g = g. The desired conclusion now
Examples 3.12.6
(1) Let .U : KHaus → Top be the forgetful functor. Recall that KHaus, the
category KHaus of compact Hausdorff spaces, is well-powered, as shown in
Example 1.3.16, and has a cogenerator by Example 3.12.2, (3). Furthermore,
KHaus has products (Example 2.1.5, (4)) and equalizers (Example 2.1.10, (4)),
which shows that is complete by Theorem 2.4.2. As both products and
equalizers are constructed as in Top we can conclude by the Special Adjoint
Functor Theorem that U has a left adjoint.
(2) Let .F : K → J be a functor between small categories and consider the induced
functor .F : Fun(J, Set) → Fun(K, Set) defined in (1.37)
commutative:
. (3.102)
244 3 Adjoint Functors
j
Indeed, recall from Theorem 2.7.2 that . H (j ), (qi : Gj (i) → H (j ))i∈Ob I is
the colimit of the induced functor .Gj : I → Set and in particular a cocone on
j
.Gj , where .q = (qi )j for all .i ∈ Ob I and .j ∈ Ob J . Therefore, the following
i
diagram is commutative:
. (3.103)
F (H )(k) = H (F (k)),
.
it can be easily seen that the commutativity of (3.103) implies the commutativity
of (3.102).
Consider now another cocone . Xk , (sik : (F ◦ G)k (i) → Xk )i∈Ob I on the
functor .(F ◦ G)k : I → Set. Hence, for all .u ∈ HomI (l, t) the following
diagram is commutative:
As .(F ◦ G)k (u) = G(u)F (k) and .(F ◦ G)k (i) = GF (k) (i), the commu-
tativity of the above diagram comes down to .G(u)F (k) ◦ slk = stk . There-
fore, . Xk , (sik : GF (k) (i) → Xk )i∈Ob I is a cocone on the induced functor
F (k)
.GF (k) : I → Set. Now recall that the pair . H (F (k)), (q : GF (k) (i) →
i
H (F (k)))i∈Ob I is the colimit of the functor .GF (k) . Thus, we have a unique
morphism .f ∈ HomSet (H (F (k)), Xk ) such that the following diagram is
3.13 Representable Functors Revisited 245
f ◦ F (qi )k = sik .
.
This shows that . F (H )(k), F (qi )k i∈Ob I is the colimit of the functor
.F ◦ G : I → Fun(K, Set), as desired. .
This section collects new representability criteria for certain classes of functors. The
first one refers to limit preserving functors.
Theorem 3.13.1 (Representability criterion) Let .C be a complete category and
F : C → Set a functor such that
.
once we show that h is the unique morphism such that .G(h) ◦ αXC (1XC ) = g.
Recall that .α is a natural transformation and therefore the following diagram is
commutative:
as desired. We are left to show that h is the unique morphism with this
property. Indeed, assume there exists an .h ∈ HomD (XC , D) such that .G(h) ◦
αXC (1XC ) = g. Using again the naturality of .α this time for the morphism .h
yields .G(h) ◦ αXC (1XC ) = αD (h). It follows that .αD (h) = g and since .αD is
bijective we obtain .h = h, which finishes the proof.
(2) The second part follows in a similar manner. Indeed, if G is right adjoint
to F then for all .C ∈ Ob C and .D ∈ Ob D we have a bijective map
.θC, D : HomD (F (C), D) → HomC (C, G(D)) which is natural in both
variables. In particular, naturality in the first variable implies that for all .D ∈
Ob D, we have a natural isomorphism between the functors .HomD (F (−), D)
and .HomC (−, G(D)). Hence the functor .HomD (F (−), D) : C → Set is
representable and its representing object is .G(D).
Conversely, as the functors .HomD (F (−), D) : C → Set are representable
for all .D ∈ Ob D, there exists .YD ∈ Ob C and a natural isomorphism
.β : HomC (−, YD ) → HomD (F (−), D). Then, it can be easily proved that
. YD , βYD (1YD ) is the final object of the category .(F ↓ D). The conclusion
is a bilinear map, such that for any other pair .(A, f ), where A is an abelian group
248 3 Adjoint Functors
This means precisely that .M ⊗R N is the tensor product of the R-modules M and
N (see [45, Definition 1.5]). .
It is worth to point out that the various adjoint functor theorems proved in this
chapter have many notable applications, most of them exceeding the purpose of
this introductory book. For instance, the following corollary to Freyd’s theorem
is presumedly “more widely known than the theorem itself ” as stated in [7]: any
functor between varieties of algebras which respects underlying sets has a left
adjoint ([7, Corollary 8.17]). For precise definitions and more details we refer the
reader to [7].
Furthermore, the adjoint functor theorems have been extended to various settings
allowing for important applications. We only mention here the case of triangulated
categories and an important consequence in algebraic geometry. A criterion for
a functor between triangulated categories to admit a right adjoint was proved by
building on a version of the representability theorem for triangulated categories.
More precisely, a triangulated functor between triangulated categories satisfying
certain technical conditions which commutes with arbitrary coproducts admits a
right adjoint. A notable application of the aforementioned criterion on the existence
of adjoints for triangulated functors is the Grothendieck duality theorem proved by
A. Neeman (see [42] for further details).
3.14 Exercises
3.1 Prove that the forgetful functor U : Field → Ring does not admit a right or a
left adjoint.
3.2 Decide if the forgetful functor U : A → Set admits a right adjoint, where A
is Grp, Ring or R M.
3.3 Let R be a commutative ring. Show that the forgetful functor U : R M → Ab
has both a left and a right adjoint.
3.4 If R is a commutative ring, show that the forgetful functor F : AlgR → R M
(forgetting the multiplicative structure) has a left adjoint.
3.5 Decide if the inclusion functor I : Ringc → Ring has a left or a right adjoint.
3.6 Let F : C → D and G : D → C be functors such that F G. Prove that F
preserves epimorphisms and G preserves monomorphisms.
3.14 Exercises 249
4.1 Chapter 1
1.4 (a) Let u denote the unique morphism in HomC (X, C). Then m ◦ u ∈
HomC (X, X) = {1X } and therefore we have
m ◦ u = 1X .
. (4.1)
(4.1)
Furthermore, we have m ◦ u ◦ m = m and since m is a monomorphism
we obtain u ◦ m = 1C .
(b) This claim follows by the duality principle; indeed, applying a) for the
dual category Cop yields the desired claim.
1.5 (b) Let u, v ∈ HomC (E, A) such that f ◦ u = f ◦ v. This implies g ◦ f ◦ u =
g ◦ f ◦ v and since g ◦ f is a monomorphism, we obtain u = v, as desired.
(c) Consider t, w ∈ HomC (C, D) such that t ◦ g = w ◦ g. This implies
t ◦ g ◦ f = w ◦ g ◦ f and since g ◦ f is an epimorphism, we obtain
t = w.
1.6 (a) Let f ∈ HomC (A, B) be a split monomorphism and denote by t its left
inverse. If g1 , g2 ∈ HomC (A , A) such that f ◦ g1 = f ◦ g2 , then by
composing on the left with t we obtain g1 = g2 , which shows that f is a
monomorphism.
For the converse consider the group morphism f : Z2 → Z4 defined by
f (0) = 0̂ and f (1) = 2̂, where x and x̂ denote the residue classes modulo
2 and 4, respectively. It can be easily seen that f is a monomorphism;
to this end, let g1 , g2 ∈ HomGrp (G, Z2 ) such that f ◦ g1 = f ◦ g2 . If
there exists some x0 ∈ G such that g1 (x0 ) = g2 (x0 ) then we can assume
that g1 (x0 ) = 0 and g2 (x0 ) = 1. This implies
without loss of generality
0̂ = f (0) = f g1 (x0 ) = f g2 (x0 ) = f (1) = 2̂, which is an obvious
contradiction. Therefore, f is a monomorphism. We are left to show that
f is not a split monomorphism. Indeed, assume that there exists a group
© The Author(s), under exclusive license to Springer Nature Switzerland AG 2023 251
A. Agore, A First Course in Category Theory, Universitext,
https://doi.org/10.1007/978-3-031-42899-9_4
252 4 Solutions to Selected Exercises
f
A B
1A 1B
A B
. f
f
A B h f u g,
g h
C D
. u
1A 1B
f
A B
1C 1D 1C 1D
u v
.
C D C D
(a) The functor F : B → C defined below is full while the morphism v and
the object C are not in its image:
F (A) = C, F (B) = D,
.
F (1A ) = 1C , F (1B ) = 1D , F (f ) = u.
1 The opposite of a group (G, ·) is another group denoted by (Gop , ·op ), where Gop = G and the
i
. 2 S3 S 3 /H
Since Z(Z2 ) = Z2 and Z(S3 ) = {1} it follows that the composition below is
an isomorphism
i ,
F (f ) : F (X) → F (Y ), F (f )(Xi ) = X
and since GF is full there exists some f ∈ HomC A, B such that
G(g) = GF (f ). As G is faithful we obtain g = F (f ), which shows
that F is full.
1.15 (a) Let G, G : I → C be two functors; we need to show that the following
mapping is bijective:
Nat(G, G ) → Nat(F G, F G ), α → F α,
.
and since F is fully faithful there exists a unique βi ∈ HomD (G(i), G (i))
that γi = F (βi ). We are left to show that the family of morphisms
such
βi i∈Ob I form a natural transformation between the functors G and G .
To this end, consider f ∈ HomI (i, j ). Since γ : F G → F G is a natural
transformation, the following diagram is commutative:
.
256 4 Solutions to Selected Exercises
. (4.2)
4.1 Chapter 1 257
1.20 Let {vi | i = 1, n} be the basis of V over the field K and denote by {vi }i=1, n
the dual basis of V ∗ . Define two natural transformations as follows:
n
β : HomK M (V , −) → V ∗ ⊗ −, βU (t) = vi∗ ⊗ t (vi ),
i=1
F (C) = A,
. F (f ) = 1A .
which shows that ψ respects identities. Furthermore, for any two natural
transformations η : F → G, ζ : G → H , where F , G, H : I → Fun(J, C)
are functors, and any (i, j ) ∈ Ob (I × J ), we have
ψ(ζ ◦ η)(i, j ) = ζ ◦ ηi j = (ζi ◦ ηi )j = (ζi )j ◦ (ηi )j
.
4.2 Chapter 2
2.3 (a) ⇒ (b) By assumption, (E, e) is the equalizer of (f, f ), so there exists
a unique w ∈ HomC (A, E) such that the following diagram is
commutative:
e f
E A B i.e., e 1A .
f
1A
. A (4.3)
1A f
E A B
f
. C
2.6 Let t ∈ HomC (B, C ) such that t ◦ f = t ◦ g. First we prove that t ◦ v makes
the following diagram commutative:
f h
A B C
g
.
C
= e ◦ F (g) ◦ F (u) = e ◦ F (g ◦ u) = e.
Furthermore, suppose there exists a w ∈ HomC (F (C), E) such that the above
diagram is commutative, i.e., w ◦F (h) = e. By composing this equality on the
right with F (v) and having in mind that h ◦ v = 1C , we obtain w = e ◦ F (v).
2.8 Since (E, p) is the equalizer in C of (f, g) we have
f ◦ p = g ◦ p.
. (4.4)
4.2 Chapter 2 261
Therefore, we have
pX ◦ f ◦ p = 1X ◦ p = pX ◦ g ◦ p,
.
(4.4)
pY ◦ f ◦ p = f ◦ p = g ◦ p = pY ◦ g ◦ p.
p f
E X Y
g
.
E
This shows that w is the unique morphism which makes the following diagram
commutative:
E X
p
p g
X X Y
. f
f
I A
g qA
B C
qB
.
D
1B β 1B
.
B B
u ◦ qA = α,
. u ◦ qB = 1B , . (4.5)
v ◦ qA = β, v ◦ qB = 1B . (4.6)
p
A B D
.
D (4.7)
We obtain
t ◦ u ◦ qA = t ◦ α = t ◦ β = t ◦ v ◦ qA ,
.
t ◦ u ◦ qB = t ◦ 1B = t ◦ v ◦ qB .
Therefore, we have (t ◦u) ◦qA = (t ◦v) ◦qA and (t ◦u) ◦qB = (t ◦v) ◦qB .
Now Proposition 2.2.14, (2) implies that t ◦ u = t ◦ v.
C B
p
t
B D
p
t
.
D
g ◦ f = f ◦ g.
. (4.8)
Since B × C, (qB , qC ) is the coproduct of B and C, there exists
a unique u ∈ HomC (B × C, P ) such that the following diagram is
264 4 Solutions to Selected Exercises
commutative:
qB qC
B / B ×C o C
EE
EE yy
EE u yyy
EE y
g E" |yyy f
P
i.e., u ◦ qB = g , .
. (4.9)
u ◦ qC = f . (4.10)
We obtain
(4.9) (4.8) (4.10)
u ◦ qB ◦ f = g ◦ f = f ◦ g = u ◦ qC ◦ g.
.
.
P (4.11)
Moreover, we obtain
(4.11) (4.10)
v ◦ q ◦ qC = u ◦ qC = f ,
.
(4.11) (4.9)
v ◦ q ◦ qB = u ◦ qB = g ,
f
A B
g q qB
g
C Q
q qC
f
.
P
4.2 Chapter 2 265
We are left to show that v is the unique morphism which makes the above
diagram commutative. Indeed, suppose there exists a t ∈ HomC (Q, P )
such that t ◦ q ◦ qC = f and t ◦ q ◦ qB = g . This yields
(4.9)
t ◦ q ◦ qB = g = u ◦ qB ,
.
(4.10)
t ◦ q ◦ qC = f = u ◦ qC .
1A f
A A B
g f
t g
t
.
B
coequalizer of the pair (u, v). Consider now the following commutative
square:
f
A B
g h
C D
. m
m ◦ g ◦ u = h ◦ f ◦ u = h ◦ f ◦ v = m ◦ g ◦ v,
.
f
C A B
t
g
.
C
g
xg
h x− 1
y− 1 x
y
.
.
268 4 Solutions to Selected Exercises
In other words, the morphism fi0 ∈ HomC (C, F (i0 )) makes the following
diagram commutative for all j ∈ Ob I :
The proof will be finished once we show that fi0 is the unique morphism with
this property. To this end, let g ∈ HomC (C, F (i0 ) such that pj ◦ g = fj for
all j ∈ Ob I . We obtain
as desired.
2.23 See Exercise 2.16.
4.3 Chapter 3
3.1 Assume U : Field → Ring has a left adjoint L : Ring → Field. As noted in
Example 1.3.10, (3) Z is the initial object of Ring and Theorem 3.4.4 implies
that L(Z) is the initial object of Field. This contradicts Example 1.3.10, (4);
therefore U does not admit a left adjoint.
Assume now that U : Field → Ring has a right adjoint R : Ring → Field.
As noted in Example 1.3.10, (3) the zero ring is the final object of Ring and
Theorem 3.4.4 implies that R({0}) is the final object of Field. Again, this
contradicts Example 1.3.10, (4); therefore U does not admit a right adjoint.
3.3 The left adjoint L : Ab → R M is the tensor functor defined as follows:
L(A) = R ⊗ A,
. L(f )(r ⊗ a) = r ⊗ f (a)
ηA : A → R ⊗ A,
. ηA (a) = 1R ⊗ a,
εM : R ⊗ M → M, εM (r ⊗ m) = rm.
It will suffice to show that (3.15) and (3.16) hold. To this end, for all r ⊗ a ∈
R ⊗ A we have
which shows that (3.16) also holds. Thus, in light of Theorem 3.5.1, we obtain
that L U , as desired.
Next, the right adjoint of U is the hom functor T : Ab → R M defined as
follows:
for all A ∈ Ob Ab, f ∈ HomAb (A, B) and g ∈ HomZ (R, A). Note that
HomZ (R, A) ∈ Ob R M with the left R-module structure given by (rf )(t) =
f (rt) for all r, t ∈ R and f ∈ HomZ (R, A). Again, we use Theorem 3.5.1
to show that U R. Indeed, consider the natural transformations η : 1R M →
T U and ε : U T → 1Ab defined as follows for all A, B ∈ Ob Ab, M ∈
Ob R M, g ∈ HomZ (R, A) and m ∈ M:
ηM : M → HomZ R, U (M) ,
. ηM (m) = ψm : R → U (M),
ψm (r) = rm,
εA : HomZ (R, A) → A, εA (g) = g(1R ).
It will suffice to show that (3.15) and (3.16) hold. To this end, for all m ∈ M
we have
.εU (M) ◦ U (ηM ) (m) = εU (M) (ψm ) = ψm (1R ) = 1R m = m = 1U (M) (m),
and therefore (3.15) holds. Furthermore, for all g ∈ HomZ (R, A) and r ∈ R
we have εA ◦ ψg (r) = εA (rg) = (rg)(1R ) = g(r1R ) = g(r). Hence, we
obtain
εA ◦ ψg = g.
. (4.12)
270 4 Solutions to Selected Exercises
i.e., T (εA ) ◦ ηT (A) = 1T (A) , which shows that (3.16) holds as well and we
obtain the desired adjunction.
3.8 The equivalence (a) ⇔ (c) follows easily from (3.15).
(b) ⇒ (d) For all C ∈ Ob C we have
(3.16) b)
1GF (C) = G(εF (C) ) ◦ ηGF (C) = G(εF (C) ) ◦ GF (ηC ).
.
(3.16) b)
= GF (ηC ) = ηGF (C) .
Now Corollary 3.6.2, (1) implies that for all C ∈ Ob C we have F (ηC ) ◦
εF (C) = 1F GF (C) , and therefore
(4.16)
= ηGF (C) = G(1F GF (C) ) ◦ ηGF (C) .
Now Corollary 3.6.2, (1) implies F (ηC ) ◦ εF (C) = 1F GF (C) . Using (3.15) we
obtain that F (ηC ) is indeed an isomorphism.
3.9 Suppose first that ηD ∈ HomD (D, H F (D)) is an epimorphism for any d ∈
Ob D and let f , g ∈ HomD (D , H G(D)) such that
.εD ◦ f = εD ◦ g. (4.17)
(4.18)
272 4 Solutions to Selected Exercises
(4.19)
(4.19) (4.18)
G (qi ) = βT (i) ◦ G(qi ) ◦ ψ = G (qi ) ◦ βlim T ◦ ψ.
.
Now since G (lim T ), (G (qi ))i∈Ob I is the limit of G T : I → C,
Proposition 2.2.14, (1) implies that 1G (lim T ) = βlim T ◦ ψ. Furthermore,
for all i ∈ Ob I we have
= G(qi ).
Next we use Theorem 3.6.1, (1) in order to show that F I . To this end,
note that for all pre-ordered
sets (P , ) we have an order-preserving map
πP : (P , ) → P , defined by πP (x) = x for all x ∈ P . Moreover,
π : 1PreOrd → I F defined for any pre-ordered set (P , ) by πP is a natural
transformation. Consider now two pre-ordered sets (P , ), (Q, ) and a
morphism f : (P , ) → Q, in PreOrd; the proof will be finished once
we show that there exists a unique morphism g : P , , → Q, in Poset
such that I (g) ◦ πP = f . Define g : P , , → Q, by g(x) = f (x)
for all x ∈ P . The only thing left to prove is that g is well-defined. Indeed,
if x = y then x y and y x and since f is order-preserving we obtain
f (x)f (y) and f (y)f (x). Now recall that is a partial order on Q and
the anti-symmetry implies f (x) = f (y), as desired.
3.14 We will show that, unless X is a singleton set, the cartesian product functor
X×− does not preserve products and, therefore, by virtue of Theorem 3.4.4, it
does not admit a left adjoint. Note that if X is a singleton set then the identity
functor is obviously the left adjoint of the corresponding cartesian product
functor.
X be a set suchthat |X| = 1. Let Y , Z ∈ Ob Set and consider
Hereafter, let
their product Y × Z, (p1 , p2 ) in Set, as constructed in Example 2.1.5, (1)
i.e., Y × Z is the cartesian product of the two sets while p1 : Y × Z → Y and
p2 : Y × Z → Z denote the projections
on the first and second
component,
respectively. Furthermore, let X × Y × X × Z, (π1 , π2 ) be the product of
X × Y and X × Z in Set, as constructed in Example 2.1.5, (1) where π1 : X ×
Y × X × Z → X × Y , π2 : X × Y × X × Z → X × Z denote the projections
on X × Y and X × Z, respectively. Assume now that the cartesian product
functor X × − preserves products. Then, in light of Proposition 2.1.3, there
exists a unique isomorphism f : X × Y × Z → X × Y × X × Z in Set (set
bijection) such that the following diagram is commutative:
X × Y× Z
1X × p 1 1X × p 2
X × Y f X × Z
π1 π2
.
X × Y× X × Z
274 4 Solutions to Selected Exercises
G(u) ◦ h = G(v) ◦ h.
. (4.20)
G(u ◦ w) ◦ ηS = G(v ◦ w) ◦ ηS .
. (4.21)
. (4.22)
εI ◦ F (v) = w.
. (4.23)
(4.22) (4.23)
εI ◦ F (f ) = w ◦ F (m) = εI ◦ F (v) ◦ F (m) = εI ◦ F (v ◦ m).
.
commutative:
This shows that G(I ) is an injective object in C. The second claim follows by
duality.
References
1. Adamek, J., Rosicky, J.: Locally Presentable and Accessible Categories. London Mathemat-
ical Society Lecture Note Series, vol. 189. Cambridge University Press, Cambridge (1994)
2. Adamek, J., Herrlich, H., Strecker, G.E.: Abstract and concrete categories. Pure and Applied
Mathematics. John Wiley & Sons, New York (1990)
3. Altman, A., Kleiman, S.: A Term of Commutative Algebra. Worldwide Center of Mathe-
matics, Cambridge (2013)
4. Arkowitz, M.: Introduction to Homotopy Theory. Universitext, Springer, New York (2011)
5. Awodey, S.: Category Theory. Oxford Logic Guides, vol. 52. Oxford University Press,
Oxford (2010)
6. Bergman, G.M.: An Invitation to General Algebra and Universal Constructions. Universi-
text. Springer, Cham (2015)
7. Bergman, G.M., Hausknecht, A.O.: Co-Groups and Co-Rings in Categories of Associative
Rings. Mathematical Surveys and Monographs, vol. 45. American Mathematical Society,
Providence (1996)
8. Borceux, F.: Handbook of Categorical Algebra. 1. Encyclopedia of Mathematics and Its
Applications, vol. 50. Cambridge University Press, Cambridge (1994)
9. Borceux, F.: Handbook of Categorical Algebra. 2. Encyclopedia of Mathematics and Its
Applications, vol. 51. Cambridge University Press, Cambridge (1994)
10. Borceux, F.: Handbook of Categorical Algebra. 3. Encyclopedia of Mathematics and Its
Applications, vol. 52. Cambridge University Press, Cambridge (1994)
11. Bosch, S.: Algebraic Geometry and Commutative Algebra. Universitext. Springer, London
(2022)
12. Bradley, T.-D., Bryson, T., Terilla, J.: Topology: A Categorical Approach. MIT Press,
Cambridge (2020)
13. Bucur, I., Deleanu, A.: Introduction to the Theory of Categories and Functors. Pure and
Applied Mathematics, vol. XIX. John Wiley & Sons, London-New York-Sydney (1968)
14. Burris, S., Sankappanavar, H.P.: A Course in Universal Algebra. Graduate Texts in Mathe-
matics, vol. 78. Springer-Verlag, New York-Berlin (1981)
15. Burton, D.M.: A First Course in Rings and Ideals. Addison-Wesley Series in Mathematics.
Addison-Wesley Publishing Company, Boston (1970)
16. Caenepeel, S., Militaru, G., Zhu, S.: Doi-Hopf modules, Yetter-Drinfel’d modules and
Frobenius type properties. Trans. Am. Math. Soc. 349, 4311–4342 (1997)
17. Caenepeel, S., Militaru, G., Zhu, S.: Frobenius and Separable Functors for Generalized
Module Categories and Nonlinear Equations. Lecture Notes in Mathematics, vol. 1787.
Springer-Verlag, Berlin (2002)
© The Author(s), under exclusive license to Springer Nature Switzerland AG 2023 277
A. Agore, A First Course in Category Theory, Universitext,
https://doi.org/10.1007/978-3-031-42899-9
278 References
48. Riehl, E.: Category Theory in Context. Aurora: Dover Modern Math Originals. Dover
Publications, Mineola (2016)
49. Rotman, J.J.: Advanced Modern Algebra. Prentice Hall, Upper Saddle River (2002)
50. Rotman, J.J.: An introduction to the Theory of Groups. Graduate Texts in Mathematics, vol.
148. Springer-Verlag, New York (1995)
51. Shulman, M.A.: Set Theory for Category Theory. https://arxiv.org/abs/0810.1279v2 (2008).
Accessed 15 Aug 2023
52. Spivak, D. I.: Category Theory for the Sciences. MIT Press, Cambridge (2014)
53. Stenström, B.: Rings of Quotients. Grundlehren der mathematischen Wissenschaften, vol.
217. Springer-Verlag, New York-Heidelberg (1975)
Index
A of finite sets, 3
Abelianization, 30 of groups, 4
functor, 30 of Hausdorff topological spaces, 5
Adjunction/adjoint functors, 153 of left R-modules, 5
Alexandroff topology, 34 of monoids, 4
Anti-isomorphism of categories, 38 of objects over C (slice category), 59
Associative law, 2 of objects under C (coslice category), 58
Axiom of choice, 1 of partially ordered sets, 5
of pointed topological spaces, 5
of pre-ordered sets, 5
B of presheaves, 63
Balanced category, 10 with (finite) products, 84
Bifunctor, 24 with pullbacks, 97
Bimorphism, 10 with pushouts, 98
of relations, 4
of right R-modules, 4
C of rings, 4
Cartesian product of sets, 3
bifunctor, 28 of simple groups, 4
functor, 29 of small categories, 35
Category, 2 of topological spaces, 5
of abelian groups, 4 of unitary rings, 4
of (commutative) algebras, 5 Cayley’s theorem, 76
of cocones, 108 Cocomplete category, 110
with coequalizers, 92 Cocone, 106
of commutative unitary rings, 4 Codomain of a morphism, 2
of compact Hausdorff topological spaces, 5 Coequalizer, 92
of cones, 108 Cogenerator (coseparator), 237
with (finite) coproducts, 84 Cokernel of a morphism, 97
defined by a pre-ordered set, 3 Colimit
of divisible groups, 4 functor, 110, 118
with equalizers, 92 preserving functor, 127
of fields, 4 reflecting functor, 131
with finite colimits, 110 Comma-category, 55
with finite limits, 110 Commutative diagram, 19
© The Author(s), under exclusive license to Springer Nature Switzerland AG 2023 281
A. Agore, A First Course in Category Theory, Universitext,
https://doi.org/10.1007/978-3-031-42899-9
282 Index
E H
Edges of a graph, 18 Hausdorff quotient functor, 33
Empty Hom
category, 3 bifunctor, 26
functor, 26 (contravariant) functor, 27
Epimorphism, 6 Homeomorphism, 9
Equalizer, 91 Homotopy, 22
Equivalence category, 23
of categories, 199 Horizontal composition of natural
relation generated by a binary relation, transformations (Godement
94 product), 44
Index 283
I O
Identity Objects of a category, 2
law, 2 Order preserving map, 5
morphism, 2
Image of a functor, 25
Inclusion functor, 25 P
Indiscrete topology, 7 Partially ordered set, 5
Initial object, 11 Pasting lemma, 23
Injective object, 80 Path (in a graph), 18
Interchange law between vertical and Pointed category, 11
horizontal composition of natural Pointwise composition of functors, 25
transformations, 80 Pontryagin duality, 214
Interior of a subset of a topological space, 94 Power set (contravariant) functor, 29
Inverse Precomposition functor, 68
of a functor, 38 Pre-ordered set, 3
natural transformation, 43 Presheaf, 63
Isomorphic Product, 83
categories, 38 category, 17
objects, 6 topology, 29
Isomorphism of categories, 38 Projection functor, 26
Projective object, 81
Pullback, 97
K Pushout, 97
Kernel of a morphism, 97
Q
L Quotient
Left adjoint, 153 category, 21
Left/right associated functor of a bifunctor, functor, 25
35 object, 14
Limit topology, 95
functor, 110, 117
preserving functor, 127
reflecting functor, 131 R
Limitation of size axiom, 1 R-bilinear map, 32
Localization Reflective subcategory, 194
of a category by a class of morphisms, 214 Reflector, 194
functor with respect to a multiplicative set, Regular
33 epimorphism, 150
Locally monomorphism, 150
compact topological space, 162 Representability criterion, 245
small category, 2 Representable functor, 48
Representing
object, 48
M pair of a functor, 50
Monomorphism, 6 Restriction of scalars functor, 34
Morphisms of a category, 2 Right adjoint, 153
Multiplicative subset of a ring, 5
S
N Set
Natural isomorphism, 42 of cogenerators (coseparators), 236
Natural transformation/isomorphism, 41 of generators (separators), 236
Neumann–Bernays–Gödel set theory, 1 Simple group, 4
284 Index
T Z
Tensor product functor, 29 Zermelo–Fraenkel set theory, 1
Trivial path in a graph, 18 Zero-morphism, 11
Tube lemma, 162 Zero object, 11
U
Unit of an adjunction, 170
Universal enveloping group functor, 31