Datasets:
Tasks:
Text Generation
Modalities:
Text
Sub-tasks:
language-modeling
Languages:
English
Size:
100K - 1M
License:
\chapter{Objects and morphisms} | |
\label{ch:cats} | |
I can't possibly hope to do category theory any justice in these few chapters; | |
thus I'll just give a very high-level overview of how many of the concepts we've | |
encountered so far can be re-cast into categorical terms. | |
So I'll say what a category is, give some examples, | |
then talk about a few things that categories can do. | |
For my examples, I'll be drawing from all the previous chapters; | |
feel free to skip over the examples corresponding to things you haven't seen. | |
If you're interested in category theory (like I was!), perhaps in | |
what surprising results are true for general categories, I strongly recommend \cite{ref:msci}. | |
\section{Motivation: isomorphisms} | |
From earlier chapters let's recall the definition of an \emph{isomorphism} of two objects: | |
\begin{itemize} | |
\ii Two groups $G$ and $H$ are isomorphic if there was a bijective homomorphism: | |
equivalently, we wanted homomorphisms $\phi : G \to H$ and $\psi : H \to G$ | |
which were mutual inverses, meaning $\phi \circ \psi = \id_H$ and $\psi \circ \phi = \id_G$. | |
\ii Two metric (or topological) spaces $X$ and $Y$ are isomorphic | |
if there is a continuous bijection $f : X \to Y$ such that $f\inv$ is also continuous. | |
\ii Two vector spaces $V$ and $W$ are isomorphic if there is a bijection $T : V \to W$ | |
which is a linear map. | |
Again, this can be re-cast as saying that $T$ and $T\inv$ are linear maps. | |
\ii Two rings $R$ and $S$ are isomorphic if there is a bijective ring homomorphism $\phi$; | |
again, we can re-cast this as two mutually inverse ring homomorphisms. | |
\end{itemize} | |
In each case we have some collections of objects and some maps, | |
and the isomorphisms can be viewed as just maps. | |
Let's use this to motivate the definition of a general \emph{category}. | |
\section{Categories, and examples thereof} | |
\prototype{$\catname{Grp}$ is possibly the most natural example.} | |
\begin{definition} | |
A \vocab{category} $\AA$ consists of: | |
\begin{itemize} | |
\ii A class of \vocab{objects}, denoted $\obj(\AA)$. | |
\ii For any two objects $A_1, A_2 \in \obj(\AA)$, | |
a class of \vocab{arrows} (also called \vocab{morphisms} or \vocab{maps}) between them. | |
We'll denote the set of these arrows by $\Hom_\AA(A_1, A_2)$. | |
\ii For any $A_1, A_2, A_3 \in \obj(\AA)$, | |
if $f \colon A_1 \to A_2$ is an arrow and $g \colon A_2 \to A_3$ is an arrow, | |
we can compose these arrows to get an arrow $g \circ f \colon A_1 \to A_3$. | |
We can represent this in a \vocab{commutative diagram} | |
\begin{center} | |
\begin{tikzcd} | |
A_1 \ar[r, "f"] \ar[rd, "h"'] & A_2 \ar[d, "g"] \\ | |
& A_3 | |
\end{tikzcd} | |
\end{center} | |
where $h = g \circ f$. | |
The composition operation $\circ$ is part of the data of $\AA$; | |
it must be associative. | |
In the diagram above we say that $h$ \vocab{factors} through $A_2$. | |
\ii Finally, every object $A \in \obj(\AA)$ has a special \vocab{identity arrow} $\id_A$; | |
you can guess what it does.\footnote{To be painfully explicit: | |
if $f \colon A' \to A$ is an arrow then $\id_A \circ f = f$; | |
similarly, if $g \colon A \to A'$ is an arrow then $g \circ \id_A = g$.} | |
\end{itemize} | |
\end{definition} | |
\begin{abuse} | |
From now on, by $A \in \AA$ we'll mean $A \in \obj(\AA)$. | |
\end{abuse} | |
\begin{abuse} | |
You can think of ``class'' as just ``set''. | |
The reason we can't use the word ``set'' is | |
because of some paradoxical issues with | |
collections which are too large; | |
Cantor's Paradox says there is no set of all sets. | |
So referring to these by ``class'' is a way of sidestepping these issues. | |
Now and forever I'll be sloppy and assume all my categories | |
are \vocab{locally small}, meaning that $\Hom_{\AA} (A_1, A_2)$ | |
is a set for any $A_1, A_2 \in \AA$. | |
So elements of $\AA$ may not form a set, | |
but the set of morphisms between | |
two \emph{given} objects will always assumed to be a set. | |
\end{abuse} | |
Let's formalize the motivation we began with. | |
\begin{example} | |
[Basic examples of categories] | |
\listhack | |
\label{example:basic_categories} | |
\begin{enumerate}[(a)] | |
\ii There is a category of groups $\catname{Grp}$. The data is | |
\begin{itemize} | |
\ii The objects of $\catname{Grp}$ are the groups. | |
\ii The arrows of $\catname{Grp}$ are the homomorphisms between these groups. | |
\ii The composition $\circ$ in $\catname{Grp}$ is function composition. | |
\end{itemize} | |
\ii In the same way we can conceive a category $\catname{CRing}$ of (commutative) rings. | |
\ii Similarly, there is a category $\catname{Top}$ of topological spaces, | |
whose arrows are the continuous maps. | |
\ii There is a category $\catname{Top}_\ast$ of topological spaces with a \emph{distinguished basepoint}; | |
that is, a pair $(X, x_0)$ where $x_0 \in X$. | |
Arrows are continuous maps $f : X \to Y$ with $f(x_0) = y_0$. | |
\ii Similarly, there is a category $\catname{Vect}_k$ of | |
vector spaces (possibly infinite-dimensional) over a field $k$, | |
whose arrows are the linear maps. | |
There is even a category $\catname{FDVect}_k$ of | |
\emph{finite-dimensional} vector spaces. | |
\ii We have a category $\catname{Set}$ of sets, | |
where the arrows are \emph{any} maps. | |
\end{enumerate} | |
\end{example} | |
And of course, we can now define what an isomorphism is! | |
\begin{definition} | |
An arrow $A_1 \taking{f} A_2$ is an \vocab{isomorphism} | |
if there exists $A_2 \taking{g} A_1$ such that $f \circ g = \id_{A_2}$ | |
and $g \circ f = \id_{A_1}$. | |
In that case we say $A_1$ and $A_2$ are \vocab{isomorphic}, hence $A_1 \cong A_2$. | |
\end{definition} | |
\begin{remark} | |
Note that in $\catname{Set}$, $X \cong Y | |
\iff \left\lvert X \right\rvert = \left\lvert Y \right\rvert$. | |
\end{remark} | |
\begin{ques} | |
Check that every object in a category is isomorphic to itself. | |
(This is offensively easy.) | |
\end{ques} | |
More importantly, this definition should strike you as a little impressive. | |
We're able to define whether two groups (rings, spaces, etc.) are isomorphic | |
solely by the functions between the objects. | |
Indeed, one of the key themes in category theory (and even algebra) is that | |
\begin{moral} | |
One can learn about objects by the functions between them. | |
Category theory takes this to the extreme by \emph{only} looking at arrows, | |
and ignoring what the objects themselves are. | |
\end{moral} | |
But there are some trickier interesting examples of categories. | |
\begin{example} | |
[Posets are categories] | |
Let $\mathcal P$ be a partially ordered set. | |
We can construct a category $P$ for it as follows: | |
\begin{itemize} | |
\ii The objects of $P$ are going to be the elements of $\mathcal P$. | |
\ii The arrows of $P$ are defined as follows: | |
\begin{itemize} | |
\ii For every object $p \in P$, we add an identity arrow $\id_p$, and | |
\ii For any pair of distinct objects $p \le q$, we add a single arrow $p \to q$. | |
\end{itemize} | |
There are no other arrows. | |
\ii There's only one way to do the composition. What is it? | |
\end{itemize} | |
\end{example} | |
For example, for the poset $\mathcal P$ on four objects $\{a,b,c,d\}$ with $a \le b$ and $a \le c \le d$, we get: | |
\begin{center} | |
\begin{tikzpicture}[scale=3.5] | |
\SetVertexMath | |
\Vertices{square}{d,c,a,b} | |
\Edge[style={->}, label={$a \le b$}](a)(b) | |
\Edge[style={->}, label={$a \le c$}](a)(c) | |
\Edge[style={->}, label={$a \le d$}](a)(d) | |
\Edge[style={->}, label={$c \le d$}](c)(d) | |
\Loop[dist=8, dir=NO, label={$\id_a$}, labelstyle={above=1pt}](a) | |
\Loop[dist=8, dir=WE, label={$\id_b$}, labelstyle={left=1pt}](b) | |
\Loop[dist=8, dir=EA, label={$\id_c$}, labelstyle={right=1pt}](c) | |
\Loop[dist=8, dir=WE, label={$\id_d$}, labelstyle={left=1pt}](d) | |
\end{tikzpicture} | |
\end{center} | |
This illustrates the point that | |
\begin{moral} | |
The arrows of a category can be totally different from functions. | |
\end{moral} | |
In fact, in a way that can be made precise, the term ``concrete category'' refers | |
to one where the arrows really are ``structure-preserving maps between sets'', | |
like $\catname{Grp}$, $\catname{Top}$, or $\catname{CRing}$. | |
\begin{ques} | |
Check that no two distinct objects of a poset are isomorphic. | |
\end{ques} | |
Here's a second quite important example of a non-concrete category. | |
\begin{example} | |
[Important: groups are one-Object categories] | |
A group $G$ can be interpreted as a category $\mathcal G$ with one object $\ast$, | |
all of whose arrows are isomorphisms. | |
\begin{center} | |
\begin{tikzpicture}[scale=5.5] | |
\Vertex[x=0,y=0,L={$\ast$}]{a} | |
\Loop[dist=8, dir=NO, label={$1 = \id_a$}, labelstyle={above=1pt}](a) | |
\Loop[dist=7, dir=WE, label={$g_2$}, labelstyle={left=1pt}](a) | |
\Loop[dist=9, dir=SO, label={$g_3$}, labelstyle={below=1pt}](a) | |
\Loop[dist=8, dir=EA, label={$g_4$}, labelstyle={right=1pt}](a) | |
\end{tikzpicture} | |
\end{center} | |
As \cite{ref:msci} says: | |
\begin{quote} | |
The first time you meet the idea that a group is a kind of category, | |
it's tempting to dismiss it as a coincidence or a trick. | |
It's not: there's real content. | |
To see this, suppose your education had been shuffled and you took a course | |
on category theory before ever learning what a group was. | |
Someone comes to you and says: | |
``There are these structures called `groups', and the idea is this: | |
a group is what you get when you collect together all the symmetries | |
of a given thing.'' | |
``What do you mean by a `symmetry'?'' you ask. | |
``Well, a symmetry of an object $X$ is a way of transforming $X$ or mapping | |
$X$ into itself, in an invertible way.'' | |
``Oh,'' you reply, ``that's a special case of an idea I've met before. | |
A category is the structure formed by \emph{lots} of objects and mappings | |
between them -- not necessarily invertible. A group's just the very special case | |
where you've only got one object, and all the maps happen to be invertible.'' | |
\end{quote} | |
\end{example} | |
\begin{exercise} | |
Verify the above! | |
That is, show that the data of a one-object category with all isomorphisms | |
is the same as the data of a group. | |
\end{exercise} | |
Finally, here are some examples of categories you can make from other categories. | |
\begin{example} | |
[Deriving categories] | |
\listhack | |
\begin{enumerate}[(a)] | |
\ii Given a category $\AA$, we can construct the \vocab{opposite category} | |
$\AA\op$, which is the same as $\AA$ but with all arrows reversed. | |
\ii Given categories $\AA$ and $\BB$, we can construct the \vocab{product category} $\AA \times \BB$ | |
as follows: the objects are pairs $(A, B)$ for $A \in \AA$ and $B \in \BB$, | |
and the arrows from $(A_1, B_1)$ to $(A_2, B_2)$ | |
are pairs \[ \left( A_1 \taking{f} A_2, B_1 \taking{g} B_2 \right). \] | |
What do you think the composition and identities are? | |
\end{enumerate} | |
\end{example} | |
\section{Special objects in categories} | |
\prototype{$\catname{Set}$ has initial object $\varnothing$ and final object $\{\ast\}$. An element of $S$ corresponds to a map $\{\ast\} \to S$.} | |
Certain objects in categories have special properties. | |
Here are a couple examples. | |
\begin{example} | |
[Initial object] | |
An \vocab{initial object} of $\AA$ is an object | |
$A_{\text{init}} \in \AA$ such that for any $A \in \AA$ (possibly $A = A_{\text{init}}$), | |
there is exactly one arrow from $A_{\text{init}}$ to $A$. | |
For example, | |
\begin{enumerate}[(a)] | |
\ii The initial object of $\catname{Set}$ is the empty set $\varnothing$. | |
\ii The initial object of $\catname{Grp}$ is the trivial group $\{1\}$. | |
\ii The initial object of $\catname{CRing}$ is the ring $\ZZ$ | |
(recall that ring homomorphisms $R \to S$ map $1_R$ to $1_S$). | |
\ii The initial object of $\catname{Top}$ is the empty space. | |
\ii The initial object of a partially ordered set is its smallest element, if one exists. | |
\end{enumerate} | |
\end{example} | |
We will usually refer to ``the'' initial object of a category, since: | |
\begin{exercise} | |
[Important!] | |
Show that any two initial objects $A_1$, $A_2$ of $\AA$ are \emph{uniquely isomorphic} | |
meaning there is a unique isomorphism between them. | |
\end{exercise} | |
\begin{remark} | |
In mathematics, we usually neither know nor care if two objects are actually equal | |
or whether they are isomorphic. | |
For example, there are many competing ways to define $\RR$, | |
but we still just refer to it as ``the'' real numbers. | |
Thus when we define categorical notions, we would like to check they are | |
unique up to isomorphism. | |
This is really clean in the language of categories, and definitions | |
often cause objects to be unique up to isomorphism for elegant reasons like the above. | |
\end{remark} | |
One can take the ``dual'' notion, a terminal object. | |
\begin{example} | |
[Terminal object] | |
A \vocab{terminal object} of $\AA$ is an object | |
$A_{\text{final}} \in \AA$ such that for any $A \in \AA$ (possibly $A = A_{\text{final}}$), | |
there is exactly one arrow from $A$ to $A_{\text{final}}$. | |
For example, | |
\begin{enumerate}[(a)] | |
\ii The terminal object of $\catname{Set}$ is the singleton set $\{\ast\}$. | |
(There are many singleton sets, of course, but \emph{as sets} they are all isomorphic!) | |
\ii The terminal object of $\catname{Grp}$ is the trivial group $\{1\}$. | |
\ii The terminal object of $\catname{CRing}$ is the zero ring $0$. | |
(Recall that ring homomorphisms $R \to S$ must map $1_R$ to $1_S$). | |
\ii The terminal object of $\catname{Top}$ is the single-point space. | |
\ii The terminal object of a partially ordered set is its maximal element, if one exists. | |
\end{enumerate} | |
\end{example} | |
Again, terminal objects are unique up to isomorphism. | |
The reader is invited to repeat the proof from the preceding exercise here. | |
However, we can illustrate more strongly the notion of duality to give a short proof. | |
\begin{ques} | |
Verify that terminal objects of $\AA$ are equivalent to initial objects of $\AA\op$. | |
Thus terminal objects of $\AA$ are unique up to isomorphism. | |
\end{ques} | |
In general, one can consider in this way the dual of \emph{any} categorical notion: | |
properties of $\AA$ can all be translated to dual properties of $\AA\op$ | |
(often by adding the prefix ``co'' in front). | |
One last neat construction: suppose we're working in a concrete category, | |
meaning (loosely) that the objects are ``sets with additional structure''. | |
Now suppose you're sick of maps and just want to think about elements of these sets. | |
Well, I won't let you do that since you're reading a category theory chapter, | |
but I will offer you some advice: | |
\begin{itemize} | |
\ii In $\catname{Set}$, arrows from $\{\ast\}$ to $S$ correspond to elements of $S$. | |
\ii In $\catname{Top}$, arrows from $\{\ast\}$ to $X$ correspond to points of $X$. | |
\ii In $\catname{Grp}$, arrows from $\ZZ$ to $G$ correspond to elements of $G$. | |
\ii In $\catname{CRing}$, arrows from $\ZZ[x]$ to $R$ correspond to elements of $R$. | |
\end{itemize} | |
and so on. | |
So in most concrete categories, you can think of elements as functions from special sets to the set in question. | |
In each of these cases we call the object in question a \vocab{free object}. | |
\section{Binary products} | |
\prototype{$X \times Y$ in most concrete categories is the set-theoretic product.} | |
The ``universal property'' is a way of describing objects in terms of maps | |
in such a way that it defines the object up to unique isomorphism | |
(much the same as the initial and terminal objects). | |
To show how this works in general, let me give a concrete example. | |
Suppose I'm in a category -- let's say $\catname{Set}$ for now. | |
I have two sets $X$ and $Y$, and I want to construct the Cartesian product $X \times Y$ as we know it. | |
The philosophy of category theory dictates that I should talk about maps only, | |
and avoid referring to anything about the sets themselves. | |
How might I do this? | |
Well, let's think about maps into $X \times Y$. | |
The key observation is that | |
\begin{moral} | |
A function $A \taking f X \times Y$ | |
amounts to a pair of functions $(A \taking g X, A \taking h Y)$. | |
\end{moral} | |
Put another way, there is a natural projection map $X \times Y \surjto X$ and $X \times Y \surjto Y$: | |
\begin{center} | |
\begin{tikzcd} | |
& X \\ | |
X \times Y \ar[ru, two heads, "\pi_X"] \ar[rd, two heads, "\pi_Y"'] & \\ | |
& Y | |
\end{tikzcd} | |
\end{center} | |
(We have to do this in terms of projection maps rather than elements, | |
because category theory forces us to talk about arrows.) | |
Now how do I add $A$ to this diagram? | |
The point is that there is a bijection between functions $A \taking f X \times Y$ | |
and pairs $(g,h)$ of functions. | |
Thus for every pair $A \taking g X$ and $A \taking h Y$ there is a \emph{unique} function | |
$A \taking f X \times Y$. | |
But $X \times Y$ is special in that it is ``universal'': | |
for any \emph{other} set $A$, if you give me functions $A \to X$ and $A \to Y$, I can use it | |
build a \emph{unique} function $A \to X \times Y$. | |
Picture: | |
\begin{center} | |
\begin{tikzcd} | |
&&& X \\ | |
A \ar[rrru, bend left, "g"'] \ar[rrrd, bend right, "h"] \ar[rr, dotted, "\exists! f"] && | |
X \times Y \ar[ru, two heads, "\pi_X"] \ar[rd, two heads, "\pi_Y"] & \\ | |
&&& Y | |
\end{tikzcd} | |
\end{center} | |
We can do this in any general category, defining a so-called product. | |
\begin{definition} | |
Let $X$ and $Y$ be objects in any category $\AA$. | |
The \vocab{product} consists of an object $X \times Y$ | |
and arrows $\pi_X$, $\pi_Y$ to $X$ and $Y$ (thought of as projection). | |
We require that for any object $A$ and arrows $A \taking g X$, $A \taking h Y$, there | |
is a \emph{unique} function $A \taking f X \times Y$ such that the above diagram commutes. | |
\end{definition} | |
\begin{abuse} | |
Strictly speaking, the product should consist of \emph{both} | |
the object $X \times Y$ | |
and the projection maps $\pi_X$ and $\pi_Y$. | |
However, if $\pi_X$ and $\pi_Y$ are understood, | |
then we often use $X \times Y$ to refer to the object, | |
and refer to it also as the product. | |
\label{abuse:object} | |
\end{abuse} | |
Products do not always exist; for example, | |
take a category with just two objects and no non-identity morphisms. | |
Nonetheless: | |
\begin{proposition}[Uniqueness of products] | |
When they exist, products are unique up to isomorphism: | |
given two products $P_1$ and $P_2$ of $X$ and $Y$ | |
there is an isomorphism between the two objects. | |
\end{proposition} | |
\begin{proof} | |
This is very similar to the proof that initial objects are unique up to unique isomorphism. | |
Consider two such objects $P_1$ and $P_2$, and the associated projection maps. | |
So, we have a diagram | |
\begin{center} | |
\begin{tikzcd} | |
& & X & & \\ | |
\\ | |
P_1 \ar[rrdd, "\pi_Y^1"', two heads] \ar[rruu, "\pi_X^1", two heads] \ar[rr, "f", two heads] | |
&& P_2 \ar[rr, "g", two heads] \ar[uu, "\pi_X^2"', two heads] \ar[dd, "\pi_Y^2", two heads] | |
&& P_1 \ar[lluu, "\pi_X^1"', two heads] \ar[lldd, "\pi_Y^1"', two heads] \\ | |
\\ | |
&& Y && | |
\end{tikzcd} | |
\end{center} | |
There are unique morphisms $f$ and $g$ between $P_1$ and $P_2$ that | |
make the entire diagram commute, according to the universal property. | |
On the other hand, look at $g \circ f$ and focus on just the outer square. | |
Observe that $g \circ f$ is a map which makes the outer square commute, | |
so by the universal property of $P_1$ it is the only one. | |
But $\id_{P_1}$ works as well. | |
Thus $\id_{P_1} = g \circ f$. | |
Similarly, $f \circ g = \id_{P_2}$ so $f$ and $g$ are isomorphisms. | |
\end{proof} | |
\begin{abuse} | |
Actually, this is not really the morally correct theorem; | |
since we've only showed the objects $P_1$ and $P_2$ are isomorphic | |
and have not made any assertion about the projection maps. | |
But I haven't (and won't) define isomorphism of the entire product, | |
and so in what follows if I say ``$P_1$ and $P_2$ are isomorphic'' | |
I really just mean the objects are isomorphic. | |
\end{abuse} | |
\begin{exercise} | |
In fact, show the products are unique up to \emph{unique} isomorphism: | |
the $f$ and $g$ above are the only isomorphisms between | |
the objects $P_1$ and $P_2$ respecting the projections. | |
\end{exercise} | |
The nice fact about this ``universal property'' mindset | |
is that we don't have to give explicit constructions; assuming existence, | |
the ``universal property'' allows us to bypass all this work by saying | |
``the object with these properties is unique up to unique isomorphism'', | |
thus we don't need to understand the internal workings of the object | |
to use its properties. | |
Of course, that's not to say we can't give concrete examples. | |
\begin{example} | |
[Examples of products] | |
\listhack | |
\begin{enumerate}[(a)] | |
\ii In $\catname{Set}$, the product of two sets | |
$X$ and $Y$ is their Cartesian product $X \times Y$. | |
\ii In $\catname{Grp}$, the product of $G$, $H$ | |
is the group product $G \times H$. | |
\ii In $\catname{Vect}_k$, the product | |
of $V$ and $W$ is $V \oplus W$. | |
\ii In $\catname{CRing}$, the product | |
of $R$ and $S$ is appropriately the ring product $R \times S$. | |
\ii Let $\mathcal P$ be a poset interpreted as a category. | |
Then the product of two objects $x$ and $y$ | |
is the \vocab{greatest lower bound}; for example, | |
\begin{itemize} | |
\ii If the poset is $(\RR, \le)$ then it's $\min\{x,y\}$. | |
\ii If the poset is the subsets | |
of a finite set by inclusion, | |
then it's $x \cap y$. | |
\ii If the poset is the positive integers ordered by division, | |
then it's $\gcd(x,y)$. | |
\end{itemize} | |
\end{enumerate} | |
\end{example} | |
Of course, we can define products of more than just one object. | |
Consider a set of objects $(X_i)_{i \in I}$ in a category $\AA$. | |
We define a \vocab{cone} on the $X_i$ to be an object $A$ | |
with some ``projection'' maps to each $X_i$. | |
Then the \vocab{product} is a cone $P$ which is ``universal'' in the same sense as before: | |
given any other cone $A$ there is a unique map $A \to P$ making the diagram commute. | |
In short, a product is a ``universal cone''. | |
The picture of this is | |
\begin{center} | |
\begin{tikzcd} | |
&& A | |
\ar[dd, "\exists! f"] | |
\ar[llddd, two heads, bend right] | |
\ar[lddd, two heads, bend right] | |
\ar[rddd, two heads, bend left] | |
\ar[rrddd, two heads, bend left] | |
&& \\ | |
&&&& \\ | |
&& P | |
\ar[lld, two heads] | |
\ar[ld, two heads] | |
\ar[rd, two heads] | |
\ar[rrd, two heads] | |
&& \\ | |
X_1 & X_2 && X_3 & X_4 | |
\end{tikzcd} | |
\end{center} | |
See also \Cref{prob:associative_product}. | |
One can also do the dual construction to get a \vocab{coproduct}: | |
given $X$ and $Y$, it's the object $X+Y$ | |
together with maps $X \taking{\iota_X} X+Y$ and $Y \taking{\iota_Y} X+Y$ | |
(that's Greek iota, think inclusion) | |
such that for any object $A$ and maps $X \taking g A$, $Y \taking h A$ | |
there is a unique $f$ for which | |
\begin{center} | |
\begin{tikzcd} | |
X \ar[rd, "\iota_X"'] \ar[rrd, "g", bend left] \\ | |
& X+Y \ar[r, "\exists! f"] & A \\ | |
Y \ar[ru, "\iota_Y"] \ar[rru, "h"', bend right] | |
\end{tikzcd} | |
\end{center} | |
commutes. | |
We'll leave some of the concrete examples as an exercise this time, | |
for example: | |
\begin{exercise} | |
Describe the coproduct in $\catname{Set}$. | |
\end{exercise} | |
Predictable terminology: a coproduct is a universal \vocab{cocone}. | |
Spoiler alert later on: | |
this construction can be generalized vastly to so-called ``limits'', | |
and we'll do so later on. | |
\section{Monic and epic maps} | |
The notion of ``injective'' doesn't make sense | |
in an arbitrary category since arrows need not be functions. | |
The correct categorical notion is: | |
\begin{definition} | |
A map $X \taking f Y$ is \vocab{monic} | |
(or a monomorphism) if for any commutative diagram | |
\begin{center} | |
\begin{tikzcd} | |
A \ar[r, shift left, "g"] \ar[r, shift right, "h"'] & X \ar[r, "f"] & Y | |
\end{tikzcd} | |
\end{center} | |
we must have $g = h$. | |
In other words, $f \circ g = f \circ h \implies g = h$. | |
\end{definition} | |
\begin{ques} | |
Verify that in a \emph{concrete} category, injective $\implies$ monic. | |
\end{ques} | |
\begin{ques} | |
Show that the composition of two monic maps is monic. | |
\end{ques} | |
In most but not all situations, the converse is also true. | |
\begin{exercise} | |
Show that in $\catname{Set}$, $\catname{Grp}$, $\catname{CRing}$, | |
monic implies injective. (Take $A = \{\ast\}$, $A = \ZZ$, $A = \ZZ[x]$.) | |
\end{exercise} | |
More generally, as we said before there are many categories | |
with a ``free'' object that you can use to think of as elements. | |
An element of a set is a function $1 \to S$, | |
and element of a ring is a function $\ZZ[x] \to R$, et cetera. | |
In all these categories, | |
the definition of monic literally reads | |
``$f$ is injective on $\Hom_\AA(A, X)$''. | |
So in these categories, ``monic'' and ``injective'' coincide. | |
That said, here is the standard counterexample. | |
An additive abelian group $G = (G,+)$ is called \emph{divisible} | |
if for every $x \in G$ and $n \in \ZZ$ there exists $y \in G$ with $ny = x$. | |
Let $\catname{DivAbGrp}$ be the category of such groups. | |
\begin{exercise} | |
Show that the projection $\QQ \to \QQ/\ZZ$ is monic but not injective. | |
\end{exercise} | |
Of course, we can also take the dual notion. | |
\begin{definition} | |
A map $X \taking f Y$ is \vocab{epic} | |
(or an epimorphism) if for any commutative diagram | |
\begin{center} | |
\begin{tikzcd} | |
X \ar[r, "f"] & Y \ar[r, "g", shift left] \ar[r, "h"', shift right] & A | |
\end{tikzcd} | |
\end{center} | |
we must have $g = h$. | |
In other words, $g \circ f = h \circ f \implies g = h$. | |
\end{definition} | |
This is kind of like surjectivity, although it's a little farther than last time. | |
Note that in concrete categories, surjective $\implies$ epic. | |
\begin{exercise} | |
Show that in $\catname{Set}$, $\catname{Grp}$, $\catname{Ab}$, $\catname{Vect}_k$, $\catname{Top}$, | |
the notions of epic and surjective coincide. | |
(For $\catname{Set}$, take $A = \{0, 1\}$.) | |
\end{exercise} | |
However, there are more cases where it fails. | |
Most notably: | |
\begin{example} | |
[Epic but not surjective] | |
\listhack | |
\begin{enumerate}[(a)] | |
\ii In $\catname{CRing}$, for instance, the inclusion $\ZZ \injto \QQ$ is epic | |
(and not surjective).. | |
Indeed, if two homomorphisms $\QQ \to A$ agree on | |
every integer then they agree everywhere (why?), | |
\ii In the category of \emph{Hausdorff} topological spaces | |
(every two points have disjoint open neighborhoods), | |
in fact epic $\iff$ dense image (like $\QQ \injto \RR$). | |
\end{enumerate} | |
Thus failures arise when a function $f : X \to Y$ can be determined by just some of the points of $X$. | |
\end{example} | |
\section\problemhead | |
\begin{problem} | |
In the category $\catname{Vect}_k$ of $k$-vector spaces | |
(for a field $k$), | |
what are the initial and terminal objects? | |
\end{problem} | |
\begin{dproblem} | |
What is the coproduct $X+Y$ in the categories | |
$\catname{Set}$, $\catname{Vect}_k$, and a poset? | |
\end{dproblem} | |
\begin{problem} | |
In any category $\AA$ where all products exist, | |
show that \[ (X \times Y) \times Z \cong X \times (Y \times Z) \] | |
where $X$, $Y$, $Z$ are arbitrary objects. | |
(Here both sides refer to the objects, as in \Cref{abuse:object}.) | |
\label{prob:associative_product} | |
\end{problem} | |
\begin{problem} | |
\gim | |
Consider a category $\AA$ with a \vocab{zero object}, | |
meaning an object which is both initial and terminal. | |
Given objects $X$ and $Y$ in $A$, | |
prove that the projection $X \times Y \to X$ is epic. | |
\end{problem} | |