Category Theory
Daniel Naylor
Contents
1 Definitions and Examples
Definition 1.1 (Category).
A category
consists of:
-
(a)
a collection
of objects .
-
(b)
a collection
of morphisms .
-
(c)
two operations ,
from
to :
we write
for “
is a morphism and
and ”.
-
(d)
an operation from
to
sending
to .
-
(e)
a partial binary operation
on ,
such that
is defined if and only if ,
and in this case we have
and .
These are subject to the axioms:
Remark 1.2.
-
(a)
and
needn’t be sets. If they are, we call
a small category.
-
(b)
We could formalise the definition without mentioning objects, but we don’t.
-
(c)
means “first ,
then ”.
Example 1.3.
-
(a)
category of all sets and the functions between them. (Formally, a morphism of
is a pair
where
is a set-theoretic function, and
is its codomain.)
-
(b)
We have categories:
-
of groups and group homomorphisms
-
of rings and homomorphisms
-
of vector spaces over a field
-
and so on
-
(c)
We have categories
-
of topological spaces and continuous maps
-
of metric spaces and non-expansive maps (i.e.
such that )
-
of smooth manifolds and
maps
Also for
topological groups and continuous homomorphisms, etc...
-
(d)
We have a category with
the same objects as ,
but morphisms
are homotopy classes of continuous maps.
In general, given and
an equivalence relation
on
such that
|
and
|
we can form a quotient category .
-
(e)
The category has
the same objects as ,
but morphisms
are relations ,
with composition defined by
|
We can also define the category
of sets with partial functions.
-
(f)
For any category , the opposite
category has the same
objects and morphisms as
but
and
are interchanged and composition is reversed.
This yields a duality principle: if is a
true statement about categories, so is
obtained by reversing arrows in .
-
(g)
A (small) category with one object
is a monoid (a semigroup with an identity). In particular, a group is a
object
small category whose morphisms are all isomorphisms.
-
(h)
A groupoid is a category whose morphisms are all isomorphisms. For example, the fundamental groupoid
of a topological
space has points
of as objects, and
morphisms are homotopy
classes of paths from
to (c.f. the
fundamental group ).
-
(i)
A discrete category is one whose only morphisms are identities. If
is such that for any
pair of objects there is
at most one morphism
then becomes a reflexive,
transitive relation on .
We call such a
a preorder. In particular, a poset is a small preorder whose only isomorphisms are identities.
-
(j)
Given a field , the category
has natural numbers as
objects, and morphisms
are matrices,
with entries from ,
and composition is matrix multiplication.
Definition 1.4 (Functor).
Let
and be categories.
A functor consists
of mappings
and
such that:
We write
for the category of small categories and the functors between them.
Example 1.5.
-
(a)
We have forgetful functors ,
,
,
… or slightly more interestingly, ,
,
,
,
…
-
(b)
The construction of free groups is a functor :
given a set ,
is the group freely generated by ,
such that every mapping
where
has a group structure extends uniquely to a homomorphism .
Given ,
we define
to be the unique homomorphism extending .
If we also have ,
and
are both homomorphisms extending .
-
(c)
Given a set ,
we define
to be the set of subsets of .
Given ,
we define
by .
So
is a functor .
-
(d)
But we also have a functor
(or ):
and, for ,
is given by .
We use the term “contravariant functor ”
for a functor .
-
(e)
Given a vector space
over ,
we write
for the space of linear maps .
Given ,
we write
for the mapping .
This defines a functor .
-
(f)
The mapping ,
defines a functor .
-
(g)
A functor between monoids is a monoid homomorphism; a functor between posets is a monotone
map.
-
(h)
Given a group ,
a functor
is given by a set
equipped with a -action
,
i.e. a permutation representation of .
Similarly, a functor
is a -linear
representation of .
-
(i)
The fundamental group construction is a functor ,
where
is the category of topological spaces with basepoints, and morphisms being the continuous maps
which preserve the basepoints.
Definition 1.6 (Natural transformation).
Given categories
and
, and two functors
, a natural
transformation
assigns to each
a morphism
in , such
that for any
in , the
square
commutes (we call this square
the
naturality square for
at
). Given
as above,
and
, we
define
by
. We write
for the
category
of functors
and natural transformations between them.
Example 1.7.
-
(a)
Given a vector space ,
we have a linear map
sending
to the linear form
on .
These maps define a natural transformation .
-
(b)
There is a natural transformation ,
where is the free
group functor and is
the forgetful functor ,
whose value at
is the inclusion .
The naturality square commutes by
the definition of .
-
(c)
For any , we have a
mapping given by
. This is a natural
transformation
since
for any .
-
(d)
Given order-preserving maps
between posets, there exists a unique natural transformation
if and
only if
for all .
-
(e)
Given two group homomorphisms ,
a natural transformation
is given by
such that
for all , or
equivalently ,
i.e.
and
are conjugate homomorphisms. In particular, the group of natural transformations
is the centraliser
of the image of .
-
(f)
If and
are
-sets considered as
functors , a natural
transformation
is a -invariant
map, i.e.
such that
for all ,
.
-
(g)
The Hurewicz homomorphism links the homotopy and homology groups of a space
. Elements of
are homotopy classes of
basepoint-preserving maps .
If we think of
as ,
defines a
singular -cycle on
and homotopic maps differ
by an -boundary, so we
get a well-defined map .
is a homomorphism, and it’s
a natural transformation ,
where is the
forgetful functor .
We have isomorphisms of categories: e.g.
defined by ,
is its
own inverse.
But we have a weaker notion of equivalence of categories.
Lemma 1.8.
Assuming that:
Then is an
isomorphism in if
and only if is an
isomorphism in
for each .
Proof.
-
Obvious since composition in .
-
Suppose each
has an inverse .
Given in
, in
the diagram we have
.
□
Definition 1.9 (Equivalence of categories).
Let
and
be categories. An equivalence between
and
consists of functors
and
together with natural isomorphisms ,
.
We write
if there exists an equivalence between
and .
We say
is a categorical property if
|
Example 1.10.
-
(a)
The category of sets and partial
functions is equivalent to (the
category of pointed sets). We define
by and
if , with
if
and undefined
otherwise. Then define
by and
if ,
then
|
Then ;
,
but there is an isomorphism .
Note that .
-
(b)
We have an equivalence :
both functors are ,
and both isomorphisms are .
-
(c)
We have an equivalence :
we define
by ,
is the linear map
represented by
(with respect to standard
bases). TO define ,
choose a basis for each ,
and define ,
|
;
the choice of bases yields isomorphisms
for each ,
which form a natural transformation .
Definition 1.11 (Faithful / full / essentially surjective).
Let
be a
functor.
-
(a)
We say
is faithful if, given
and
in ,
,
,
.
-
(b)
We say
is full if, for every
in ,
there exists
in
with .
-
(c)
We say
is essentially surjective if, for any ,
there exists
with .
Note that if is full and faithful,
it’s essentially injective: given
in , the
unique
with is
an isomorphism.
We say is a full
subcategory if the inclusion
is a full functor.
Lemma 1.12.
Assuming that:
Proof.
-
Suppose give ,
and
as in Definition 1.9. Then
witnesses the fact that
is essentially surjective. If
satisfy ,
then ;
but ,
so .
Suppose given ;
then
satisfies
but
is faithful for the same reason as ,
so .
-
For each ,
chose
and an isomorphism .
Given ,
define
to be the unique morphism such that .
Functoriality follows from uniqueness, and naturality of .
We define
to be the unique morphism such that .
is an isomorphism, and naturality squares for
are mapped by
to naturality squares for ,
so they commute. □
Definition 1.13 (Skeleton).
By a skeleton of a category ,
we mean a full subcategory containing just one object from each isomorphism class.
We say
is skeletal if it’s a skeleton of itself.
However, working with skeletal categories involves heavy use of the axiom of choice.
Definition 1.14 (Monomorphism / epimorphism).
Let
be a morphism in a category .
We say
is a monomorphism (or monic) if, given ,
.
We say
is an epimorphism (or epic) if it’s a monomorphism in .
We write
to indicate that
is monic, and
to indicate that it’s epic.
We say
is balanced if every arrow which is monic and epic is an isomorphism.
We will call a monic morphism
split if it has a left inverse (and similarly we may define the notion of split epic).
Example 1.15.
-
(a)
In ,
monic
injective (
obvious; for
consider morphisms ).
Also, epic
surjective (
obvious; for
consider morphisms ).
-
(b)
In ,
monic
injective (for
consider homomorphisms ),
and epic
surjective (but
is quite non-trivial – it uses free products with amalgamation).
-
(c)
In ,
monic
injective, but epic does not imply surjective (for example, consider ).
-
(d)
In ,
monic
injective and epic
surjective (as in )
but
isn’t balanced.
-
(e)
In preorder, all morphisms are monic and epic, so a preorder is balanced if and only if it’s an
equivalence relation.
2 The Yoneda Lemma
Definition 2.1 (Locally small).
We say a category
is locally small if, for any two objects
and ,
the morphisms
in
are parameterised by a set .
If is an object of a
locally small category ,
we have a functor
sending to
and a morphism
to the mapping
(this is functorial
since composition in
is associative).
Dually, we have .
Lemma 2.2 (Yoneda).
Assuming that:
Then
Proof.
-
(i)
Given ,
we define .
Given , we
define
by . This is
natural in
since is a
functor: given
we have
|
For any ,
.
For any ,
for all .
So .
-
(ii)
Later. Seeing examples of usage of (i) is interesting first. □
Corollary 2.3.
Assuming that:
Proof.
Substitute
for
in Lemma 2.2(i): we have a bijection from
to the collection of natural transformations .
For a given ,
the natural transformation
sends
to ,
so this is functorial by associativity of composition .
Similarly, we have a full and faithful functor
sending
to .
We call this the Yoneda embedding: it allows us to regard any locally small category
as a full subcategory of a -valued
functor category. □
Compare with Cayley’s Theorem in group theory (every group is isomorphic to a subgroup of a
permutation group) and ‘Dedekind’s Theorem’ (every poset is isomorphic to a sub-poset of a power
set).
Definition 2.4 (Representable).
We say a functor
is representable if it’s isomorphic to a
for some .
By a representation of ,
we mean a pair
where
is such that
is an isomorphism. We call
a universal element of .
Corollary 2.5.
Suppose
and are both representations
of . Then there is a
unique isomorphism
such that .
Proof.
is equivalent to saying that
commutes,
so
must be the unique isomorphism, whose image under
Yoneda is
.
□
Proof of Yoneda(ii).
Suppose for the moment that
is
small, so that
is
locally small. Given
two
functors : the
first sends an object
to
, and
a morphism
to the diagonal of
The
second is the composite
|
where is a Yoneda
embedding. Then
and
define a natural isomorphism between these two.
In elementary terms, this says that if ,
and is its image under
the diagonal, then
is the composite
|
This makes sense without the assumption that
is small, and it’s true since the composite maps
|
Example 2.6.
-
(a)
The forgetful functor
is represented by ,
is represented by ,
is represented by .
-
(b)
The functor
is represented by .
This is the bijection between subsets of
and functions ,
and it’s natural. But
is not representable, since
isn’t a singleton.
-
(c)
The functor
sending
to the set of open subsets of ,
and
to
is representable by the Sierpinski space
with
open but
not open. This works since continuous maps
are the characteristic functions of open subsets of .
-
(d)
The functor
isn’t representable, but its composite with
is represented by .
-
(e)
For a group
considered as a -object
category, the unique representable functor
is the Cayley representation:
acting on itself by multiplication.
-
(f)
Given two objects
in a locally small category ,
we have a functor
sending
to .
If this functor is representable, we call the representing object a categorical product
and write
for the universal element. Its defining property is that given any pair ,
there is a unique isomorphism
such that
and .
Dually, we have the notion of coproduct
with coprojections ,
.
-
(g)
Given a parallel pair
in a locally small category ,
we have a functor
sending
to
and defined on morphisms in the same way as .
A representation of this functor is called an equaliser of :
it consists of
satisfying ,
and such that any
with
factors uniquely as .
Note that
is monic; we call a monomorphism regular if it occurs as an equaliser.
Dually, we have the notions of coequaliser and regular epi.
In , products are just
cartesian products (also in ,
,
, …). coproducts
in are disjoint
unions . In
, coproducts are
free products .
In , the equaliser of
is the inclusion of
and the coequaliser
of is the quotient of
by the smallest equivalence
relation containing .
Note that in ,
all monomorphisms and all epimorphisms are regular, but in
, a monomorphism
is regular if and only
if is topologised as
a subspace of . An
epimorphism is regular if
and only if is topologised
as a quotient of .
Note that if
is both regular monic and regular epic, then it’s an isomorphism since the pair
of which its equaliser
must satisfy .
Warning.
The following terminology is not standard. These are usually (both!) referred to as
“generating”, but to avoid confusion, in this course we will refer to them with separate names.
Definition 2.7 (Separating / generating family).
Let
be a family of objects of
a locally small category .
-
(a)
We say
is a
separating family if the
functors ,
are jointly
faithful, i.e. given a parallel pair
,
the equations
for all
with
imply
.
-
(b)
We say
is a detecting family if the
jointly reflect isomorphisms, i.e. given ,
if every
with
factors uniquely through ,
then
is an isomorphism.
If , we
call a
separator or a detector.
Proof.
-
(i)
Suppose
is a detecting family, and suppose
satisfy the hypothesis of Definition 2.7(a). Let
of :
then any
with
factors uniquely through ,
so
is an isomorphism, so .
-
(ii)
Suppose
is separating, and
satisfies the hypothesis of Definition 2.7(b). If
satisfy ,
then any
with
satisfies ,
since both are factorisations of
through .
So ;
hence
is monic.
Similarly, if
satisfy ,
then any
satisfies ,
since it factors through ,
so
and hence
is epic. Since
is balanced,
is an isomorphism. □
Example 2.9.
-
(a)
In ,
is a separator and a detector, since
is isomorphic to the identity functor. Also,
is a coseparator and a codetector, since it represents .
-
(b)
In
(respectively ),
(respectively )
is a separator and a detector, since it represents the forgetful functor.
But
has no coseparator or codetector set: given any set
of groups, there is a simple group
with
for all ,
so the only homomorphisms
with
are trivial.
-
(c)
For any small category ,
the set
is separating and detecting in .
This uses Yoneda and Lemma 1.8 (for the detecting case).
-
(d)
In ,
is a separator since it represents .
But
has no detecting set of objects: given a set
of spaces, choose
for all ,
and let
and
be a set of .
Give
the discrete topology and for ,
we set the closed sets be
plus all the subsets of .
The identity
is continuous, but not a homeomorphism, but its restriction to any subset of
is a homeomorphism, so
can’t detect the fact that
isn’t an isomorphism.
-
(e)
Let
be the category whose objects are the ordinals, with identities plus two morphisms
whenever
with composition defined by .
Then
is a detector for :
it can tell that
aren’t isomorphisms since neither factors through the other, and if
it can tell that
aren’t isomorphisms since
doesn’t factor through either.
But
has no separating set: if
is any set of ordinals, choose
for all
and then
can’t separate .
By definition, the functors
preserve monomorphisms, but they don’t always preserve epimorphisms.
Definition 2.10 (Projective).
We say an object
in a locally
small category is
projective if
preserves epimorphisms, i.e. if given
there
exists
with
. Dually,
is
injective if it’s
projective in
.
If satisfies this condition
for all in some class
of epimorphisms, we
call it -projective.
In , we consider the class of
pointwise epimorphisms, i.e. those
such that is
surjective for all .
Proof.
Immediate from Yoneda; given
with
pointwise
epic,
is
for
some
,
so
.
□
“ has
enough pointwise projectives”:
Proposition 2.12.
Assuming that:
Proof.
Set
where the disjoint union is over all pairs
with
and .
A morphism
is uniquely determined by a family of morphisms .
. Hence
is pointwise projective, since all the
are. But we have
whose -th
component is
and this is pointwise epic since any
appears as .
□
3 Adjunctions
Definition 3.1 (Adjunction, D. Kan 1958).
Let
and
be categories. An adjunction between
and
consists of functors
and
together with, for each
and ,
a bijection between morphisms
in
and morphisms
in ,
which is natural in
and .
(If
and
are locally small, this means that
and
are naturally isomorphic functors .)
We say
is left adjoint to ,
or
is right adjoint to ,
and we write .
Example 3.2.
-
(a)
The free functor
is left adjoint to the forgetful functor .
By definition, homomorphisms
correspond to functions ;
naturality in
was built into the definition of
in Example 1.5(b) and naturality in
is immediate.
-
(b)
The forgetful functor
has a left adjoint ,
which equips a set
with its discrete topology since any function
is continuous as a map .
also has a right adjoint
given by the ‘ìndiscrete’ topology.
-
(c)
The functor
has a left adjoint
given by discrete categories, and a right adjoint :
is the category with objects
and morphisms
for each .
also has a left adjoint :
is the set of connected components of ,
i.e. the quotient of
by the smallest equivalence relation which identifies
with
for all .
-
(d)
Given a set ,
we can regard
as a functor .
It has a right adjoint, namely .
Given
we can regard it as a function
by .
We call a category
cartesian closed if it has binary products as defined in Example 2.6(f) and each
has a right adjoint .
For example,
is cartesian clsosed, with
taken to be the .
-
(e)
Let
be the -element
monoid with
(and identity ).
We have a functor
sending
to
and a functor
sending
to .
We have :
since any
takes values in
and any
is determined by its restriction to
since .
However, note that this is not an equivalence of categories.
-
(f)
Let
be the category with one object and one morphism (which must the identity on the only object). A
left adjoint for the unique functor
picks out an initial object of ,
i.e. an object such that there is a unique
for each .
Dually, a right adjoint for
‘is’ a terminal object of
(a terminal object is an initial object in ).
Again, the example of
shows that these two can coincide.
-
(g)
Suppose given
in .
We have order-preserving mappings
and ,
and
since .
-
(h)
Suppose given a relation .
We define
and
by
These are contravariant functors and .
We say
and
are adjoint on the right.
-
(i)
is self-adjoint on the
right, since functions
and functions both
correspond to relations .
-
(j)
is self-adjoint on the
right, since linear maps
and both correspond
to bilinear maps .
Theorem 3.3.
Assuming that:
-
-
for
, let
be the
category whose
objects are pairs
where
and
, and whose
morphisms
are
morphisms
making
commute.
Proof.
First suppose .
For each , let
be the morphism
corresponding to .
Then is an
initial object of :
given any ,
the diagram
commutes if
and only if
corresponds to
under the
adjunction, by
naturality of the
adjunction bijection.
So there’s a unique morphism
in .
Conversely, suppose given in initial object
in for each
. We make
into a
function :
given ,
is the unique
morphism in
. Functoriality comes
from uniqueness: given ,
and
are both
morphisms in
. The adjunction
bijection sends to the
unique morphism
in , with inverse
sending to
. This is natural
in since
is a natural
transformation
and natural in
since is
functorial. □
Corollary 3.4.
Assuming that:
Then there is a canonical natural isomorphism .
Proof.
and
are both initial in ,
so there’s a unique isomorphism
between them.
is natural: given ,
and
are both morphisms
in ,
so they’re equal. □
As a result of this, we will often talk about “the” left adjoint of a functor (when it exists), because we don’t
usually care about which one in the isomorphism class we use.
Lemma 3.5.
Assuming that:
Then .
Proof.
Given ,
,
we have bijections between morphisms ,
morphisms ,
and morphisms
which are both natural in
and ,
.
□
Corollary 3.6.
Assuming that:
Then the square of
left adjoints commutes up to natural isomorphism.
Proof.
By Lemma 3.5, both ways round are left adjoint to ,
so by Corollary 3.4 they’re isomorphic. □
We saw in Theorem 3.3 that an adjoint
gives rise to a natural transformation , called
the unit of the adjunction. Dually, we have ,
the counit of .
Theorem 3.7.
Assuming that:
Proof.
Suppose .
We defined
in the proof of Theorem 3.3, and
is defined dually. Since
corresponds to ,
the composite
corresponds to .
But by definition
corresponds to .
The other identity is dual.
Conversely, suppose given
and
satisfying the triangular identities. Given ,
we define .
Dually, given ,
we define .
Then ,
and dually .
Naturality of
and
follows from naturality of
and .
□
In Definition 1.9, we had natural isomorphisms
and . These look like the unit
and counit of an adjunction :
do they satisfy the triangular identities? No, but we can always change them:
Proposition 3.8.
Assuming that:
Then there exist isomorphisms
and satisfying the triangular
identities. In particular, .
Proof.
We define
and take to
be the composite
|
Note that ,
since
commutes by
naturality of
,
and
is
monic.
Similarly,
.
To verify the triangular identities, consider
which commutes
by
naturality of
.
For the second triangular identity, we have
Hence by
Theorem
3.7 we have
.
But
and
also satisfy the triangular
identities for and
adjunction .
□
Lemma 3.9.
Assuming that:
Then
Proof.
-
(1)
Given
in ,
corresponds to
under the adjunction. So
epic if and only if
acts injectively on morphisms with domain
and specified codomain. Hence
epic for all
if and only if
is faithful.
-
(2)
Similarly,
full and faithful if and only if for all
and
composition with
is a bijection .
This happens if and only if
is an isomorphism for all .
□
Example 3.11.
-
(a)
is reflective in :
the left adjoint to the inclusion sends
to
where
is the subgroup generated by commutators. Any homomorphism
with
abelian factors uniquely through the quotient map .
-
(b)
Recall that a group
is torsion if all elements have finite order, and torsion free if its only element of finite order is
.
In an abelian group ,
the torsion leements form a subgroup ,
and
is right adjoint to the inclusion ,
since any homomorphism
whose
is torsion takes values in .
Similarly,
defines a left adjoint to the inclusion .
-
(c)
Let
be the full subcategory of compact Hausdorff spaces.
is reflective in :
the left adjoint is the Stone-Čech compactification .
-
(d)
Let
be the full subcategory of sequential spaces, i.e. those in which all sequentially closed sets are closed.
The inclusion
has a right adjoint sending
to ,
the same set as
with all sequentially closed sets declared to be closed. The identity mapping
is (continuous, and) the counit of the adjunction.
-
(e)
The category
of preordered sets is reflective in :
the reflection sends
to
where
is the congruence identifying all paralell pairs in .
-
(f)
Given a topological space ,
the poset
of open subsets of
is coreflective in ,
since if
is open and
is arbitrary, we have
if and only if
(recall
denotes interior). Dually, the poset of closed subsets is reflective in .
4 Limits
Definition 4.1 (Diagram).
Let
be a category (almost always small, and often finite). By a diagram of shape
in a category ,
we mean a functor .
The objects ,
are called vertices of ,
and morphisms ,
are called edges of .
For example, if
is the category
a diagram of shape
is a commutative
square in .
If is
instead
then a diagram
of shape is
a not-necessarily-commutative square.
Definition 4.2 (Cone, limit).
Let
be a diagram. A cone over
consists of an object (its
apex) together with morphisms
for each
(the legs of the cone) such that
commutes
for each
in
.
A morphism of cones
is a morphism
such that for
all . We have a
category of cones
over ; a limit
for is a terminal
object of .
Dually, a colimit for
is an initial cone under .
If is the functor sending
to the constant diagram
with all vertices then a
cone over is a natural
transformation .
Also, is another
name for , defined
as in Theorem 3.3.
So by Theorem 3.3, has limits
for all diagrams of shape
if and only if
has a right adjoint.
Example 4.3.
-
(a)
Suppose .
If ,
then ,
so a limit for
is a terminal object.
-
(b)
If , a diagram
of shape
is a pair ,
and a cone over it is a span A
limit for it is a categorical coproduct
Dually, a colimit for it is a coproduct
-
(c)
If is a (small) discrete
category, a (co)limit for
is a (co)product
().
-
(d)
If is
, then a diagram
of shape is a
parallel pair .
A cone over it consists of satisfying
, or equivalently
of satisfying
. So a limit
for is an
equaliser for ,
as defined in Example 2.6(g).
-
(e)
If is
then a diagram
of shape
is a cospan A cone
over it has
legs, but if we omit the (redundant) middle one, it’s a span
completing the cospan to a commutative square.
A limit for
is called a
pullback for .
If
has binary products and equalisers, we can construct pullbacks by forming the equaliser
. Dually,
colimits of shape
are called pushouts.
-
(f)
If is the
-element with
, a diagram of
shape is an object
equipped with an idempotent
. A limit (respectively colimit)
for is the monic (respectively
epic) part of a splitting of .
Note that the functor
in Example 3.2(e) is ,
so this explains the coincidence of left and right adjoints.
-
(g)
Suppose
is the ordered set of natural numbers. A diagram of shape
is a
direct sequence
and a colimit for it is called a direct limit .
Dually, we have inverse sequences
and their limits are called inverse limits.
For example in topology, an infinite dimensional CW-complex
is the direct limit
of its -skeletons
. In algebra, the
ring of -adic
integers is the limit of the inverse sequence
|
in .
Proposition 4.4.
Assuming that:
Then
Proof.
-
(i) & (ii)
Let
be a diagram. Form the products
We have morphisms
defined by
and for
all . Let
be an equaliser
for . The
morphisms form
a cone over ,
since for any
we have
|
It is a limit: given any cone
over ,
the
form a cone over the discrete diagram with vertices
, so they induce
a unique .
Then since
the s form a
cone over ,
so factors
uniquely as , and
is the unique
factorisation of
through .
-
(iii)
If is a terminal
object of , then
we can construct
as the pullback of Then we can
construct
as .
To form an equaliser of ,
consider the pullback of
Any cone over this
has .
So a limit cone has the universal property of an equaliser for
.
□
Definition 4.5 (Limit preserving / reflecting / creating).
Let
be a
functor.
-
(a)
We say
preserves limits of shape
if, given
and a limit
cone
for it,
is a
limit for
.
-
(b)
-
(c)
We say
creates limits of shape
if, given
and a
limit cone
over
,
there exists a
cone over
whose image under
is
,
and any such
cone is a
limit in
.
We say a category
is complete if it has all small limits.
Corollary 4.6.
In each of the statements of Proposition 4.4, we may replace
‘ has’ by
either ‘
has and
preserves’ or ‘
has and
creates’.
Example 4.7.
-
(a)
The functor
creates all small limits: given a family of groups ,
there’s a unique structure on
making the projections into homomorphisms, and it’s a product in .
Similarly for equalisers. But
doesn’t preserve or reflect coproducts.
-
(b)
The forgetful functor
preserves small limits and colimits, but doesn’t reflect them.
-
(c)
The inclusion
reflects coproducts, but doesn’t preserve them.
A coproduct
in
is nonabelian if both
and
are nontrivial. So the only cones in
thot could map to coproduct cones in
are those where either
or
is trivial. But if
then
in either category.
-
(d)
If
is a reflective subcategory of ,
the inclusion
creates any limits which exist.
Given
and a limit cone
for it in ,
the morphisms
(where
is the left adjoint, and
is the unit) form a cone over ,
so they induce a unique .
Now
is
since it’s a factorisation of the limit through itself. So ,
i.e.
is a factorisation of
through itself, so .
So the
form a limit cone in ,
and hence in .
-
(e)
If
has limits of shape ,
so does
for any ,
and the forgetful functor
creates them (strictly).
Given ,
we can regard it as a functor .
For each ,
is a diagram of shape
in ,
so has a limit .
Given
in ,
the composites
form a cone over ,
so induce a unique .
Functoriality of
follows fro uniqueness, and this is the unique way of making
into a functor which lifts the
to a cone in .
The fact that it’s a limit cone is straightforward.
Lemma 4.9.
Assuming that:
Proof 1.
Suppose ,
and suppose
and have limits
of shape .
Then the diagram
commutes, and all the
functors in it have
right adjoints, so
commutes up to isomorphism by Corollary
3.6.
□
Proof 2.
Suppose given
and a limit cone
over it. Give a cone
over ,
the transposes
form a cone over
by naturality of the adjunction, so induce a unique
such that
for all .
Then
is the unique morphism satisfying
for all .
□
Lemma 4.10.
Assuming that:
Then for each
,
has
limits of
shape
and the
forgetful
functor
creates them.
Proof.
Suppose given ;
write
and let
be a limit for .
Since the edges of
are morphisms in ,
the
form a cone over ,
so there’s a unique
satisfying
for all .
So
is the unique lifting of
to an object of
which makes the
into morphisms
in .
The fact that these morphisms form a limit cone is straightforward. □
Can we represent an initial object as a limit?
Lemma 4.11.
Assuming that:
Proof.
First suppose
is initial. The unique morphisms ,
,
form a cone over ,
and it’s a limit cone since if
is any cone over ,
then
is its unique factorisation through the one with apex .
Conversely, suppose given a limit
for .
Then
is weakly initial (i.e. it admits morphisms to every object of );
and if
then .
In particular,
for all ,
so
is a factorisation of the limit cone through itself, so
and
is initial. □
The ‘primitive’ Adjoint Functor Theorem follows from Lemma 4.10, Lemma 4.11 and Theorem 3.3. But it
only applies to preorders (see Example Sheet).
Theorem 4.12 (General Adjoint Functor Theorem).
Assuming that:
Then has a
left
adjoint if and only if
preserves
small limits and satisfies the
solution-set condition: for every
, there’s
a set
of
objects of
which is collectively
weakly initial.
Proof.
-
preserves limits by Lemma 4.9, and
is a singleton solution-set for each .
-
By Lemma 4.10, the categories
are complete, and they’re locally small since
is.
So we need to show: if
is complete and locally small, and has a weakly initial set ,
then it has an initial object. First form ;
then
is weakly initial. Now form the limit of the diagram with vertices
and ,
with the morphisms
being all endomorphisms of .
Writing
for this,
is still weakly initial. Suppose given ;
let
be their equaliser. There exists some .
Now ,
but we also have ,
so .
But
is monic, so we get ,
so
is split epic, and hence .
□
Example 4.13.
-
(a)
Consider the forgetful functor .
has and
preserves all small limits by Example 4.7(a), and
is locally small. Given ,
any
factors through
where
is the subgroup generated by .
Also .
Let
be a set of this cardinality: considering all subsets ,
all group structures on
and all functions ,
we get a solution-set at .
-
(b)
Let
be the category of complete lattices (posets with all joins and all meets).
creates limits just like .
In 1965, A. Hales showed that there exist arbitrarily large complete lattices generated by 3 element
subsets, so the solution-set condition fails for .
Now also that
doesn’t have a coproduct for 3 copies of .
Definition 4.14 (Subobject).
By a subobject of ,
we mean a monomorphism .
We order subobjects by
if there exists
We
write
for this preorder.
We say is
well-powered if every
is equivalent to a small poset.
For example, is well-powered
since the inclusions form a
representative set of subobjects of .
It is well-copowered since isomorphism classes of epimorphisms
correspond to
equivalence relations on .
Lemma 4.15.
Assuming that:
Proof.
Suppose given
with .
Then ,
but is
monic so .
So and
are
both factorisations of
through the
pullback, and hence
.
□
Theorem 4.16 (Special Adjoint Functor Theorem).
Assuming that:
Proof.
-
is Lemma 4.9.
-
Let .
As in Theorem 4.12,
inherits completeness and locally smallness from :
it also inherits well-poweredness since subobjects of
in
are those
in
such that
factors through .
(Note that the forgetful functor
preserves monomorphisms by Remark 4.8). And if
is a coseparating set for ,
then
is a coseparating set for .
So we need to show: if
is complete, locally small and well-powered and has a coseparating set
, then
has an initial
object. First form ;
now consider the limit of the diagram
whose edges are a representative
set of subobjects of .
If is the apex of the
limit cone, the legs
of the limit cone are all monic by the argument of Lemma 4.15, and in particular
is monic, and it’s a
least subobject of .
If we had , their
equaliser would be
a subobject of
contained in , so
is an isomorphism,
and hence .
Given any form
the product
over all pairs
with and the
morphism
with for all
. Since the
are coseparating,
is monic. We
also have
defined by
for all .
Form the pullback
then
is monic by
Lemma 4.15, so
factors as and
hence we have .
So
is initial. □
Remark 4.18.
-
(a)
The construction in Theorem 4.16 is closely parallel to Čech’s original construction of .
Given a space, Čech constructs
and the map
defined by .
Then he takes
to be the closure of the image of ,
i.e. the smallest subobject of
in .
-
(b)
We could have constructed
using Theorem 4.12: to get a solution-set for
at an object
of ,
note that any continuous
factors as
where
is the closure of the image of ,
and then since
has a dense subspace of cardinality ,
we have .
5 Monads
Suppose we have ,
. How much of the adjunction can
we describe in terms of (supposing
we can’t know anything about ,
or know very little about it)?
We have:
From the triangular identities of Theorem 3.7, we obtain the commutative triangles:
and from
naturality of
we obtain
Definition 5.1 (Monad).
A monad on a category
is a
triple
where ,
and
and
satisfy the commutative diagrams
Example 5.2.
-
(a)
Let
be a monoid. The functor
has a monad structure:
is
and
sends
to .
The three diagrams ‘are’ the unit and associative laws in .
-
(b)
The functor
has a monad structure: the unit
is the mapping
(Example 1.7(c)) and the multiplication
sends a set of subsets of
to their union.
Does every monad come from an adjunction?
Answered by Eilenberg-Moore and by Kleisli (1965).
Note that the monad of Example 5.2(a) is induced by
and that of Example 5.2(b)
is induced by ,
where
is the category of complete semilattices (posets, with arbitrary joins). The free complete semilattice on
is
: every
extends
uniquely to
where .
An -set (respectively a complete
semilattice) is a set equipped
with a suitable mapping
(respectively ).
Definition 5.3 (Eilenberg-Moore algebra).
Let
be a monad on . By an
Eilenberg-Moore algebra for
we mean a pair
where
and
satisfies
A
homomorphism
is a morphism
satisfying
We write
for the
category
of
-algebras
and homomorphisms.
Proposition 5.4.
Assuming that:
Proof.
We define
(an algebra by (2) and (3)) and
(a homomorphism by naturality of ).
Clearly,
is functorial and .
We establish the adjunction using Theorem 3.7: its unit is ,
and the counit
is just
(a homomorphism ,
by (5), and natural by (6)).
The triangular identity
is just
(1), and
is
(4).
Finally, by
definition of . So the
adjunction induces .
□
Note: induces
, we can replace
by its full subcategory
on objects .
So in trying to construct , we may
assume is surjective (indeed,
bijective) on objects. The morphisms
in must correspond
to morphisms
in .
Definition 5.5 (Kleisli category).
Let
be a monad on .
The Kleisli category
is defined by ,
morphsims
in
are morphisms
in .
The identity
is ,
and the composite of
is .
For the unit and associative laws, consider the diagrams
Proposition 5.6.
Assuming that:
Proof.
We define
and .
preserves identities by definintion, and preserves composition by
We
define
,
and
.
preserves identities by
(1), and preserves composites by
We verify the
adjunction
using Theorem
3.7:
by
(1) so
and we take
as
unit of the
adjunction.
We define
to be .
To verify the naturality square
the lower composite
is
and the upper
one is
, which agree
since
(2) tells us that
.
The triangular identities become
and
Finally,
, so
induces the
monad .
□
Given a monad
on , we write
for the category whose
objects are adjunctions
inducing , and
morphisms
are functors
satisfying
and .
Proof.
Suppose given
in .
We define
by
(an algebra by one of the triangular identities for
and ,
and naturality of ),
(a homomorphism by naturality of ).
is functorial since
is,
is obvious, and .
So
is a morphism of .
Suppose
is another such: then we must have
where
is a natural transformation since
is a homomorphism
for all .
Also, since ,
we have
for all .
For any , we
have naturality squares
whose left edges are equal, and whose top edge is
split epic, so we obtain
for all
. So
.
We define
by and
.
preserves identities and
satisfies , by the first
triangular identity for
and .
preserves the
composite
by
Also
and
|
So is a morphism
of . Note that
is full and faithful,
since it sends to
its traspose across .
If is any
morphism of , we
must have for
all , and since
and the adjunctions
have the same unit,
must send the transpose
of to its
transpose across .
So .
□
has
coproducts if
does, but has few other limits or colimits. In contrast, we have:
Proposition 5.8.
Assuming that:
Then
Proof.
-
(i)
Suppose given ;
write ,
and let
be a limit for .
The composites
form a cone over .
So they induce a unique .
And
is a -algebra
structure on ,
since the identities
and
follow from uniqueness of factorisations through limits and it’s the unique lifting of the limit cone
in
to a cone in .
The fact that it’s a limit cone is straightforward.
-
(ii)
If
creates colimits then it preserves them, but so does
since it’s a left adjoint, so
preserves them too.
Conversely, given
and a colimit cone
under ,
we need to know that
is a colimit cone to obtain
(and that
is a colimit to verify that
is a -algebra
structure). Otherwise, the argument is as before. □
Given ,
, how can we
tell when is
part of an equivalence?
Note: is an equivalence
if and only if
is essentially surjective.
We call (or
the functor )
monadic if
is part of an equivalence.
Lemma 5.9.
Assuming that:
-
-
for every
algebra
,
the pair
has a
coequaliser in
Proof.
Write for the coequaliser.
For any homomorphism
the two left hand squares in
commute, so
we get a unique
making the right hand square commute. As usual, uniqueness implies functoriality of
.
For any , morphisms
correspond to
morphisms
satisfying .
If is the
transpose of
across , then
transposes
to , whereas
transposes to
. But we can write
by the proof of
Theorem 3.7, so .
So if
and only if
commutes, which
happens if and only if
in
.
Naturality of the bijection follows from that of .
□
Note that since ,
we have by
Corollary 3.6, and
preserves coequalisers.
Definition 5.10 (Reflexive / split coequaliser diagram / -split).
Theorem 5.12 (Crude Monadicity Theorem).
Assuming that:
Proof.
-
(5.11, )
Necessity of
is obvious. For the other condition, it’s enough to show that
creates coequalisers
of -split
pairs. This is a re-run of Proposition 5.8(ii): if
are such that
is a split coequaliser diagram, the coequaliser is preserved by
and
by , so
acquires a unique
algebra structure
making a
homomorphism, and
is a coequaliser in .
-
(5.11 and 5.12)
Either set of hypotheses implies that
has the coequalisers needed for Lemma 5.9, so
has a left
adjoint .
So we need to show that the unit and counit of
are isomorphisms.
The unit is the
factorisation of
through the (-split)
coequaliser
of .
But either set of hypothesis implies that
preserves the
equaliser of ,
so this factorisation is an isomorphism.
The counit is the
factorisation of through
the coequaliser of .
The hypotheses of Theorem 5.11 imply that
is a coequaliser of this pair, so the counit is an isomorphism. Those of
Theorem 5.12 imply that the factorisation is mapped to an isomorphism by
,
so it’s an isomorphism. □
Remark 5.13.
-
(1)
Reflexive coequalisers are colimits of shape ,
where
is the category satisfying
,
and
.
-
(2)
All colimits can be constructed from coproducts and reflexive coequalisers. This was proved in Proposition 4.4:
the pair
appearing in that proof is coreflexive with common left inverse
defined
by for
all .
-
(3)
If is
reflexive, then in any commutative square we
have .
So a pushout for is a coequaliser
for .
-
(4)
In , or more generally in a
cartesian closed category, if
() are reflexive
coequalisers, then
is also a coequaliser. To see this, consider in
which all rows and columns are coequalisers. Then the lower right square is a pushout; but if
coequalises
, then is also
coequalises
and ,
so if factors through the top and left edges of the lower right square, and hence through
.
Example 5.14.
-
(a)
The forgetful functor
is monadic, and satisfies the hypotheses of Theorem 5.12. If
is a reflexive pair in ,
with coequaliser
in ,
then
is a coequaliser, so the multiplication
induces a binary operation ,
which is the unique group multiplication on
making
a homomorphism, and it makes
into a coequaliser in .
The same argument works for ,
,
,
,
….
It doesn’t work for categories like
or ,
but here we can use Theorem 5.11 provided the forgetful functor has a left adjoint.
-
(b)
Any reflection is monadic: this can be proved using Theorem 5.11. If
is a reflective
subcategory, and
is a pair in
for which there exists in
satisfying the equaitions of
Definition 5.10(b), then
since is
full, so
is in ,
but
is closed under splittings of idempotents by Example 4.7(d), so
belongs to it.
-
(c)
Consider the composite adjunction where
is the adjunction
of Example 3.11(b). The two factors are monadic, but the composite isn’t since free abelian groups are torsion free, so
and its category
of algebras is .
-
(d)
The contravariant power-set functor
is monadic, and satisfies the hypotheses of Theorem 5.12. Its left adjoint is
by
Example 3.2(i), and it reflects isomorphisms by Example 2.9(a).
Let be a coreflexive
equaliser diagram in .
Then
is
a pullback by Remark 5.13(c), so commutes. But
we also have
and
since
and
are injective, so is
a split coequaliser diagram.
-
(e)
The fogetful functor is
not monadic; the monad on
induced by is
so its category
of algebras is .
-
(f)
The composite adjunction
is monadic. We’ll prove this using Theorem 5.11: suppose given
in
and
a split coequaliser in
. The quotient topology
on is the unique
topology making into
a coequaliser in , and
it’s compact, so will
be a coequaliser in
provided
is Hausdorff. It is also the unique topology that could make
into a
morphism of .
But, given an equivalence relation
on a compact Hausdorff space ,
is Hausdorff
if and only if
is closed in .
In our case, if
(i.e. )
then
and
satisfy ,
and
.
Conversely, if we have
and as
above, then ,
so
where
is . But
is closed in
since it’s the
equaliser of .
So is compact,
so is
compact, so
is closed in .
Definition 5.15 (Monadic tower).
Let
be an adjunction where has reflexive
coequalisers. The monadic tower of
is the diagram
where
is the
monad
induced by
,
and
and
are as in Theorem
5.7
and Lemma
5.9, and
is
the
monad induced by
and so on.
We say has monadic
length if we reach
an equivalence after
steps.
6 Filtered Colimits
Definition 6.1 (Filtered).
We say a category
is filtered if every finite diagram
has a cone under it.
Lemma 6.2.
is filtered if and only if:
For preorders, we say directed instead of filtered.
Definition 6.3 (Has filtered colimits).
We say
has filtered colimits if every ,
where
is small and filtered, has a colimit.
Note that direct limits as in Example 4.3(g) are directed colimits.
Lemma 6.4.
Assuming that:
Proof.
By Proposition 4.4(i), enough to show
has all small coproducts.
Given a set-indexeud family
of objects, the finite coproducts ,
for
finite, form the vertices of a diagram of shape
whose edges are coprojections.
is directed, and a colimit for this diagram has the universal property of a coproduct .
□
Suppose given a ,
where has limits
of shape and
colimits of shape .
We can form
, by Example 4.7(e) these are
the vertices of a diagram ,
and we can form .
Similarly, the colimits form
a diagram of shape , and
we can form . We get an
induced morphism ; if this
is an isomorphism for all ,
we say colimits of shape
commute with limits of shape
in .
Equivalently, preserves
limits of shape ,
or preserves
colimits of shape .
In Remark 5.13(d) we saw that reflexive coequalisers commute with finite products in
.
Theorem 6.5.
Assuming that:
Proof.
-
Let
be a diagram with
finite. We have a diagram
defined by .
For each ,
is a singleton since every
is identified with
in the colimit, so
is a singleton.
But elements of
are cones under
with apex ,
so if
is nonempty there must be such a cone for some .
-
Suppose given
where
is finite and
is filtered. In general, the colimitof
is the quotient of
by the smallest equivalence relation identifying
with
for all
in .
For filtered ,
this identifies
with
if and only if there exists
with ,
and moreover if
we may assume .
Now, given an element
of ,
we can write it as
where
is an equivalence class of elements .
If
in ,
then
and
represent the same element of
so by repeated use of Lemma 6.2(ii) we can choose representatives in
for some fixed ,
and by repeated use of Lemma 6.2(iii) we can assume that these representatives define an element of
.
This defines an element of
mapping to the given element of .
The proof of injectivity is similar: if two elements
of
have the same image in
we can choose representatives
in
and then find
so that each of the components
and
map to the same element of
under .
So
in .
□
Corollary 6.6.
Assuming that:
Then
Proof.
-
(i)
This is just like Example 5.14(a): Given a filtered diagram
and a colimit for
with apex ,
then
is the colimit of
for all ,
so each -ary
operation on the ’s
induces an -ary
operation on ,
and
also inherits all the equations defining ,
so there’s a unique lifting of the colimit cone under
to a colimit cone for .
-
(ii)
Follows from (i) and Theorem 6.5, since
also creates finite limits (and reflects isomorphisms). □
Similar results hold for categories such as .
Given a functor , the
category of elements of
is : its objects
are pairs
with and
morphisms are
morphisms
such that .
Proposition 6.8.
Assuming that:
Then the following are equivalent:
Proof.
-
(i) (ii)
By Lemma 4.10,
has finite limits so
is filtered.
-
(ii) (iii)
Consider the diagram
where
is the forgetful functor and
is the Yoneda embedding. A cone under this diagram (with apex ,
say) yields a family of morphisms
for each ,
subject to compatibility conditions which say that
for every
in ,
i.e. such that
is a natural transformation .
So the cone
has the universal property of a colimit for the diagram.
-
(iii) (i)
Functors of the form
preserve any limits which exist, so this follows from Theorem 6.5 plus the fact that
colimits in
are computed pointwise. □
Given a category with filtered
colimits, we say is finitary if it
preserves filtered colimits. If ,
then a finitary is determined
by its restriction to ,
since any set is the directed union of its finite subsets.
In fact the restriction functor
has a left adjoint (the left Kan extension functor) and the finitary functors are those in the image of this left
adjoint (up to isomorphism).
For a category
as in Example 5.14(a) or Corollary 6.6, the corresponding monad
on
is finitary.
From now on,
will denote the skeleton of the category of finite sets whose objects are the sets
.
For example, if is a
monad on , the full
subcategory of whose
objects are the sets
is a Lawvere theory.
Lemma 6.10.
Assuming that:
Proof.
Given a model ,
we have
for all .
Also, any morphism
induced by a morphism
in
is determined by its composites with the projections ,
so specifying
on morphisms is determined by its effect on morphisms with domain .
So, given a set ,
specifying a model
with is equivalent to
specifying operations
for each in
, subject
to
whenever is
the -th
coprojection, and
commutes whenever
commutes.
□
Note that the characterisation of -models
in any category with finite products. Note also that the equations of Lemma 6.10 allow us to reduce any compound
operation to a
single operation .
Theorem 6.11.
Assuming that:
Then the following are equivalent:
-
(i)
is
equivalent to a finitary algebraic
category in the sense of Example
5.14(a). (Recall that these
categories are those whose objects are sets with finitary operations satisfying some equations,
and morphisms are homomorphisms between them.)
-
(ii)
-
(iii)
for a finitary
monad
on
.
Proof.
-
(ii) (i)
Let
be the full subcategory of
on the free algebras ,
for .
Then
is a Lawvere theory, and for every object
of ,
the functor
restricted to
preserves finite products, so it’s a model of .
This defines a functor ;
but
for some finitary monad
on ,
so we get a functor
which is the identity on underlying sets.
In this situation,
is induced by a morphism of monads ,
i.e. a natural transformation
commuting with the units and multiplications. (Clearly, such a
induces a functor
sending
to ).
But we know
is bijective for all ,
since elements of the free algebras on
are just morphisms
in .
But both functors are finitary, so
is bijective for all ,
i.e. it’s an isomorphism of monads. □
For a general monad
on , this construction
produces a finitary monad
which is the coreflection of
in the category of finitary monads.
For example:
-
For ,
we obtain .
-
For ,
we obtain the trivial monad .
7 Regular Categories
Definition 7.1 (Image, cover).
We say a category
has images if, for every
in ,
there exists a least
in
through which
factors. We call
the image of ,
and we say
is a cover if its image is .
We write
to indicate that
is a cover.
Proof.
If is
strong epic, applying the definition to commutative squares of the form
shows
that
is
a
cover.
For the converse, a cover
is epic since it can’t factor through the equaliser of any
with
. To
verify the other condition, suppose given
then the
pullback of
along
is
monic by Lemma
4.15,
and
factors through it, so it’s
an isomorphism. So we get
by composing with the top edge of the
pullback square.
□
Here, if has images, image
facorisation defines a functor :
given
if we
form the image factorisations we get a
unique
making both squares commute.
Definition 7.3 (Regular category).
We say
is regular if it has finite limits and images, and image factorisations are stable under pullback, i.e. if
the left hand square above is a pullback then so are both right hand squares. (This is equivalent to
saying that covers are stable under pullback).
Proposition 7.5.
Assuming that:
Proof.
-
Regular epimorphism implies strong epimorphism by Exercise 21˙4.
-
Suppose is a
cover; let
be its kernel-pair, i.e. the pullback of Suppose
given with
; form the
image of
. We’ll show
is an isomorphism,
so that is a
factorisation of
through .
is a cover
since is, so we
need to prove
is monic.
Let such
that ;
form the pullback
factors
as , so
is a cover,
and
is a cover.
Now so
factors
through .
But and
is monic,
so ,
so ,
i.e. .
Also is
epic, so .
□
By a relation
in a category
with finite products, we mean an isomorphism class of subobjects
.
If has images, we define
the composite of
by forming the pullback
forming the
image of .
This is well-defined up to isomorphism and has the
as
-sided
identities.
Lemma 7.6.
Composition of relations in
is associative if and only if
is regular.
Proof.
-
Suppose given .
Consider the relations
Composing the right hand pair first, we get and
thus we get
Composing the left hand pair first, we begin by forming the pullback and we endup
with the image of ;
so
must be a cover.
-
Suppose given relations .
If we form the pullbacks then both
and
are the
image of .
□
We write for the category whose
objects are those of and whose
morphisms are relations. Note that
is just
as defined in Example 1.3(e).
We have a faithful functor which is
the identity on objects and sends
to (for faithfulness, see
Exercise 4.22(i)). We write
for .
Note that there’s an isomporphism which
is the identity on objects and sends
to ; we denote
this by , and
write
for .
Also, is enriched
over (provided
is locally small, i.e.
is well-powered),
i.e. each
has a partial order which is preserved by composition.
We say is left
adjoint to if
and
.
Proposition 7.7.
is
a left adjoint in if and
only if it is of the form .
Proof.
-
We show : the
composite is just
the kernel-pair
of , and
factors through
it. Also
is the image of so it
contains .
-
Conversely, suppose
has a right adjoint .
In forming ,
we take the pullback So the
image of
contains , so
factors as a cover followed
by a split epimorphism, so
is a cover.
Now, in the pullback
and
are covers, but
the image of is
contained in
so .
But ,
so ,
,
and
. So
is monic, and hence
an isomorphism, so .
□
P. Freyd developed a theory of allegories which have the structure of categories of relations and axiomatised those
allegories for which
the subcategory
is regular.
In a regular category ,
we say a relation
is reflexive if ,
symmetric if , and
transitive if .
is an equivalence relation if it
has all three properties. For any
in , the
kernel-pair of
is an equivalence relation. We say
an equivalence relation is effective
if it occurs as a kernel-pair, and
is effective regular if all equivalence relations are effective.
is regular but not effective
regular: is a non-effective
equivalence relation on .
Note that an equivalence relation is idempotent in
, and if
is an allegory and
is a class of symmetric
idempotents in then
(as defined in Exercise
1.18) is an allegory; and if
is for a regular
category ,
then:
Proposition 7.8.
Assuming that:
Note that if
is effective regular, its equivalence relations are split idempotents in
: if
is the
kernel-pair of
then it splits as
(as we saw for
in Exercise 1.19).
In ,
is the power-set
of , the unit is
the mapping of
Example 1.7(c), and .
Note that (isomorphism classes of) subobjects of
are in bijection with morphisms .
C. J. Mikkelses showed that any topos has finite colimits; we’ll give Bob Paré’s proof, which is much
simpler.
Proposition 7.10.
Assuming that:
Proof.
We make the assignment
into a functor
and a functor :
given ,
corresponds
to the image of ,
and
corresponds to the pullback of
Given
corresponding
to
,
corresponds to the
image of
and similarly
given
, composing
with
corresponds to
pulling back along
.
Given a pullback square
in
,
commutes, since both ways correspond to the
image of the left vertical composite in
where
both squares are
pullbacks.
Now, as in Example 5.14(d), we have that if
is a coreflexive in ,
then
is a
split coequaliser
coequaliser in
.
Also,
is
self-adjoint on the right, and it
reflects isomorphisms by Exercise 7.17(v). The second assertion follows from
Proposition
5.8(i).
□
Definition 7.11 (Support, totally supported, capital).
-
(a)
By the
support of an object
in a
regular category, we mean the
image of
.
We say
is
well-supported if
is a
cover.
-
(b)
We say a
regular category
is
totally supported if every object is well-supported. We say
is
almost totally supported if every object is either well-supported or a strict
initial object, where we
cann an object
strict if every
is an isomorphism. (Given finite
limits, a strict object is
initial since for any
there exists
,
and the
equaliser of any pair
is
).
-
(c)
A representable functor
always preserves limits, so it’s a regular functor if and only if
is
cover-projective (c.f. Definition 2.10).
Lemma 7.12.
Assuming that:
Proof.
Since covers are stable under pullback, we need to show that every
is split epic. If ,
nothing to prove. If not, the projections
aren’t equal (since their coequaliser is ,
by Proposition 7.5). So there exists
not factoring through their equaliser, so there exists .
□
If is regular, the
full subcategory
of well-supported objects is closed under finite products since
is a pullback, and under
pullbacks of covers since if
then
and
have the same support.
We write for the category
obtained from by adjoining
a strict initial object :
this is regular and almost totally-supported and the functor
sending all
non-well-supported objects to
is regular (c.f. Exercise 5.19).
Lemma 7.13.
Assuming that:
Proof.
Recall from Exercise 7.17:
regular implies
regular for any ,
and for any
in
pullback along
defines a regular functor ,
which has a left adjoint
sending
to .
And
reflects isomorphisms if and only if
is a cover.
We’ll define
as
where
is easier to describe.
To satisfy the desired conclusion for a single well-supported object ,
enough to take ,
since
acquires a point
not factoring through
for any proper .
More generally, for any finite list
of well-supported objects, we can take .
We define a base to be a finite list
of distinct well-supported objecs of .
We preorder the set
of bases by
if
contains all the members of .
We write
for the product
and if
we write
for the product projection .
This makes
into a functor .
Hence the assignment ,
is ‘almost’ a functor .
We now define :
its objects are pairs
where
is a base and
is an object of .
Morphisms
are represented by pairs
where
is a base containing
and
and
in ,
subject to the relation which identifies
with
if
and the pullback of
to
is isomorphic to .
Clearly, each
sits inside
as a non-full subcategory; so in particular
is a subcategory of ,
is regular, and the inclusions
are isomorphism-reflecting regular functors.
Given a finite diagram in ,
we can choose
such that all edges of the diagram appear as morphisms in ,
and take the limit there, and this is a limit in .
Similarly for images.
Also, if a morphism
becomes an isomorphism in ,
its inverse must live
for some ,
hence
is an isomorphism .
We define :
the induced functor
is still isomorphism reflecting since
is almost totally-supported. □
Lemma 7.14.
Assuming that:
Proof.
Consider the sequence
where each
is obtained from
by the construction of Lemma 7.13.
We define
to be the pseudo-colimit of this sequence: objects are pairs
where ,
and morphisms
are represented by pairs
where
and
in ,
modulo the identification of
with
if
and .
The proof that
is regular, and that the embeddings
are isomorphism-reflecting regular functors, is as in Lemma 7.13.
Given any non-invertible monomorphism
in ,
it lives in
for some ,
so there exists
in
not factoring through .
But if
isn’t monic in ,
the legs
of its kernel-pair aren’t equal, so there exists
not factoring through their equation, so
are distinct but have the same composite with .
So
reflects monomorphisms and hence reflects isomorphisms. □
Theorem 7.15.
Assuming that:
Proof.
Let be a representative
set of subobjects of
in , and
for each
consider the composite
|
where the third factor is the functor of Lemma 7.14 and the fourth is represented by .
Given any non-invertible morphism
in ,
if
is the support of
then
remains non-invertible in
and its codomain is well-supported there, so it remains non-invertible in
and hence in .
So these functors collectively reflect isomorphisms. □
˙
Index
Eilenberg-Moore algebra
Lawvere theory
adjoint
adjunction
allegory
adjoint on the right
almost totally-supported
balanced
capital
category
categorical coproduct
categorical property
categorical coproduct
cartesian closed
coequaliser
colimit
complete
completeness
cone
contravariant
coproduct
cospan
counit
cover
cover-projective
creates
detector
detecting
diagram
directed
direct limit
direct sequence
edge
effective
effective regular
Eilenberg-Moore
(4)
(5)
(6)
epimorphism
epic
equaliser
equivalence
equivalent
equivalence relation
essentially injective
essentially surjective
faithful
filtered
full
functor
functorial
-split
has filtered colimits
image
initial
inverse limit
inverse sequence
initial object
Kleisli category
Kleisli
left adjoint
commute
limit
locally small
model
monad
monadic
monic
monomorphism
(1)
(2)
(3)
natural
natural isomorphism
naturality square
natural transformation
naturality
naturally isomorphic
injective
reflexive
preserve
preserves
product
projective
pullback
pushout
pointwise
right adjoint
reflect
reflects
reflective
reflection
reflexive
reflective subcategory
regular
regular
regular epi
relation
representation
represented
representable
separator
separating
strict initial object
skeletal
skeleton
small
span
split coequaliser
split
solution-set condition
solution-set
subobject
support
symmetric
terminal
terminal object
topos
transitive
totally-supported
unit
vertex
well-copowered
weakly
well-powered
well-supported
Yoneda embedding