« In defense of drawing by coding » Catsters guide

Catsters guide

Posted on January 13, 2014
Tagged

Introduction

In an attempt to solidify and extend my knowledge of category theory, I have been working my way through the excellent series of category theory lectures posted on Youtube by Eugenia Cheng and Simon Willerton, aka the Catsters.

Edsko de Vries used to have a listing of the videos, but it is no longer available. After wresting a copy from a Google cache, I began working my way through the videos, but soon discovered that Edsko’s list was organized by subject, not topologically sorted. So I started making my own list, and have put it up here in the hopes that it may be useful to others. Suggestions, corrections, improvements, etc. are of course welcome!

As far as possible I have tried to arrange the order so that each video only depends on concepts from earlier ones. Along with each video you can also find my cryptic notes; I make no guarantee that they will be useful to anyone (even me!), but hopefully they will at least give you an idea of what is in each video. (For some of the earlier videos I didn’t take notes, so I have just copied the description from YouTube.)

I have a goal to watch two videos per week (at which rate it will take me about nine months to watch all of them); I will keep this list updated with new video links and notes as I go.

Terminal and Initial objects

Terminal and initial objects 1

http://youtu.be/yeQcmxM2e5I

Terminal and initial objects 2

http://youtu.be/9vhWpOVPlIE

Terminal and initial objects 3

http://youtu.be/yaPwKu5fHqI

Products and Coproducts

Products and coproducts 1

http://youtu.be/upCSDIO9pjc

Products and coproducts 2

http://youtu.be/BqRkULEhG40

Products and coproducts 3

http://youtu.be/TCQWHOUBwDE

Products and coproducts 4

http://youtu.be/n6nBgszToBE

Pullbacks and Pushouts

Pullbacks and pushouts 1

http://youtu.be/XGysPJvCXOc

Pullbacks and pushouts 2

http://youtu.be/LkkallToFQ0

Natural transformations

Natural transformations 1

http://www.youtube.com/watch?v=FZSUwqWjHCU

Natural transformations 2

http://youtu.be/XnrqHd39Cl0

Natural transformations 3

http://youtu.be/EG5xUYXHFeU

Natural transformations 3A

http://youtu.be/fsfzEz6qAGQ

Representable functors and the Yoneda Lemma

Representables and Yoneda 1

http://youtu.be/4QgjKUzyrhM

Representables and Yoneda 2

http://youtu.be/eaJmUUogb6g

Representables and Yoneda 3

http://youtu.be/TLMxHB19khE

Adjunctions (part 1)

Adjunctions 1

http://youtu.be/loOJxIOmShE

Adjunctions 2

http://youtu.be/JEJim3t-N9A

Adjunctions 4

http://youtu.be/nP5XQ6OBHHY

Monads

Monads 1

http://youtu.be/9fohXBj2UEI

Monads 2

http://youtu.be/Si6_oG7ZdK4

Monads 3

http://youtu.be/eBQnysX7oLI

Monads 3A

http://youtu.be/uYY5c1kkoIo

More on monoids as monad algebras of the list monad.

Monads 4

http://youtu.be/Cm-O_ZWEIGY

Monad algebras form a category (called \(\mathbf{Alg}\ T\)).

Adjunctions and monads

Adjunctions 3

http://youtu.be/2i_PpYsl8b8

Adjunctions 5

http://youtu.be/xqLgGB7Hv7g

“Every monad comes from an adjunction via its category of algebras.”

Last time we showed every adjunction gives rise to a monad. What about the converse?

Answer: yes. In fact, given a monad, there is an entire category of adjunctions which give rise to it, which always has initial and terminal objects: these are the constructions found by Kleisli and by Eilenberg-Moore, respectively. Intuitively, any other adjunction giving rise to the monad can be described by the morphisms between it and the Kleisli and Eilenberg-Moore constructions.

Let \((T : C \to C, \eta, \mu)\) be a monad.

Adjunctions 6

http://youtu.be/Ht1mQ97Zq2k

This time, initial solution to “does a monad give rise to any adjunctions”: Kleisli.

Adjunctions 7

http://youtu.be/D8g9xnVr0Lg

The adjunction that comes from the Kleisli category, giving rise to the original monad \(T\).

Again, let \((T : C \to C, \eta, \mu)\) be a monad. We will construct \(F_T \dashv G_T : C \to C_T\), where \(C_T\) is the Kleisli category defined in Adjunctions 6, with \(G_T F_T = T\).

Given a monad \(T\) on \(C\), we have a category of adjunctions \(\mathbf{Adj}(T)\) giving rise to \(T\) (morphisms are functors making everything commute). \(C_T\) is the initial object and \(C^T\) is terminal.

Question of monadicity: given an adjunction \(F \dashv G\), is \(D \cong C^T\)? If so, say “\(D\) is monadic over \(C\)”, i.e. everything in \(D\) can be expressed as monad algebras of \(C\). Or can say the adjunction is a “monadic adjunction”. Can also say that the right adjoint (forgetful functor \(G\)) “is monadic”. Monadic adjunctions are particularly nice/canonical.

String diagrams

String diagrams 1

http://youtu.be/USYRDDZ9yEc

Way of notating natural transformations and functors. Poincare dual: 0D things (points, i.e. categories) become 2D (regions), 1D things (lines, i.e. functors) stay 1D, 2D things (cells, i.e. natural transformations) become 0D.

String diagrams should be read right-left and bottom-top.

Horizontal and vertical composition of NTs correspond to horizontal and vertical juxtaposition of string diagrams.

Can leave out vertical lines corresponding to identity functor.

String diagrams 2

http://youtu.be/JeGhNhgOTuk

Recall the interchange law, which says that vertical and horizontal composition of natural transformations commute. This guarantees that string diagrams are well-defined, since the diagram doesn’t specify which happens first.

Whiskering is represented in string diagrams by horizontally adjoining a straight vertical line.

String diagrams 3

http://youtu.be/pmvVE8AGAEA

Given an adjunction \(F \dashv G\), we have natural transformations \(\varepsilon : FG \to 1\) and \(\eta : 1 \to GF\), and two laws given by triangles. What do these look like as string diagrams? \(\varepsilon\) is a cap, \(\eta\) a cup, and the triangle laws look like pulling wiggly strings straight!

String diagrams 4

http://youtu.be/YNC5faXshAk

Monads in string diagrams. Draw \(\mu\), \(\eta\), and the monad laws as nice-looking string diagrams with nice topological intuition.

String diagrams 5

http://youtu.be/kiXjcqxVogE

Seeing how monads arise from adjunctions, using string diagrams.

Pipe cleaners

These are presented without any commentary or explanation that I can find. Each of the below videos just presents a 3D structure made out of pipe cleaners with no explanation. Maybe there is some other catsters video that presents a motivation or explanation for these; if I find it I will update the notes here. I can see that it might have something to do with string diagrams, and that you can make categories out of these sorts of topological structures (e.g. with gluing as composition) but otherwise I have no clue what this is about.

There is also:

This is a nice 5-minute presentation about Klein bottles, complete with pipe cleaner model. Though it seems to have little to do with category theory.

Also also:

This has nothing to do with either pipe cleaners or category theory, but it is midly amusing.

General Limits and Colimits

General limits and colimits 1

http://www.youtube.com/watch?v=g47V6qxKQNU

Defining limits in general, informally.

General limits and colimits 2

http://www.youtube.com/watch?v=SpCyaNi257w

Examples of limits.

General limits and colimits 3

http://www.youtube.com/watch?v=U3nzEUEnLKQ

General limits and colimits 4

http://youtu.be/7B4cawdLAPg

General limits and colimits 5

http://youtu.be/Ud_k4HFIogQ

General limits and colimits 6

http://youtu.be/9UOdrRF_pNc

Colimits using the same general formulation. “Just dualize everything”.

Slice and comma categories

Slice and comma categories 1

https://www.youtube.com/watch?v=f4jpvwwnq_s

Slice category. Given a category \(C\), fix an object \(X \in C\). Then we define the slice category \(C/X\) by

Coslice category, or “slice under” category \(X/C\) is the dual of \(C/X\), i.e. objects are pairs \((A,p)\) where \(p : X \to A\), etc.

Slice and comma categories 2

http://youtu.be/W6sG5uraex0

Comma categories are a generalization of slice categories. Fix a functor \(F : C \to D\) and an object \(X \in D\). Then we can form the comma category \(F \downarrow X\).

Of course we can dualize, \(X \downarrow F\) (“cocomma” sounds even stupider than “cocone”, perhaps).

Apparently comma categories give us nice ways to talk about adjunctions.

Let’s generalize even more! Fix the functor \(F\) but not the object \(X \in D\). Then we can form \(F \downarrow D\):

Can also dualize, \(D \downarrow F\).

An even further generalization! Start with two functors \(F : C \to D\), \(G : E \to D\). Form \(F \downarrow G\):

In fact, all of these constructions are universal and can be seen as limits/colimits from the right point of view. “Next time”. (?)

Coequalisers

Coequalisers 1

https://youtu.be/vYktRTtulek

Coequalisers are a colimit. Show up all over the place. Give us quotients and equivalence relations. Also tell us about monadicity (given an adjunction, is it a monadic one?).

Definition: a coequaliser is a colimit of a diagram consisting of two parallel arrows.

More specifically, given \(f,g : A \to B\), a coequaliser is an object \(C\) equipped with \(e : B \to C\) such that \(f;e = g;e\), with a universal property: given any other \(s : B \to V\) with \(f;s = g;s\), \(s\) factors uniquely through \(e\).

Coequalisers 2

https://www.youtube.com/watch?v=DMSPS6us__Y

Quotient groups as coequalisers. Consider a group \(G\) and a normal subgroup \(H \triangleleft G\). In the category of groups, consider two parallel maps \(H \to G\): the inclusion map \(\iota\), and the zero map \(0\) which sends everything to the identity element \(e \in G\). Claim: the coequaliser of these two maps is the quotient group \(G/H\), together with the quotient map \(G \to G/H\).

Let’s see why. Suppose we have another group \(V\) with a group homomorphism \(\theta : G \to V\) such that \(\iota ; \theta = 0 ; \theta = 0\); that is, \(\theta(h) = e\) for all \(h \in H\). We must show there is a unique homomorphism \(G/H \to V\) which makes the diagram commute.

Notation: \(g \in G\) under the quotient map gets sent to \([g] = gH\) (\(g_1 \sim g_2\) iff \(g_2 g_1^{-1} \in H\)). For the homomorphism \(G/H \to V\), send \([g]\) to \(\theta(g)\). Note this is required to make things commute, which gives us uniqueness; we must check this is well-defined and a group homomorphism. If \(g_1 \sim g_2\) then \(g_2 g_1^{-1} \in H\). By definition, \(\theta(g_2 g_1^{-1}) = e\), and since \(\theta\) is a group homomorphism, \(\theta(g_2) = \theta(g_1)\). Hence it is well-defined, and must additionally be a group homomorphism since \([g_1] [g_2] = [g_1 g_2]\) and \(\theta\) is a group homomorphism.

Monoid objects

Monoid objects 1

https://www.youtube.com/watch?v=PH-OhkrXXvA

Idea: take the definition of monoids from \(\mathbf{Set}\), and “plunk it” into any other category with enough structure.

Now let’s reexpress this categorically in \(\mathbf{Set}\). Note we have been talking about elements of sets; we have to replace this with use of only objects and morphisms of \(\mathbf{Set}\).

Now we take the definition and port it to any monoidal category.

Monoid objects 2

https://www.youtube.com/watch?v=7Sf3Y4sesZE

Today: monoid object in the category of monoids is a commutative monoid.

Note first the category of monoids is itself monoidal under Cartesian product. That is, given two monoids \(M\) and \(N\), \(M \times N\) is also a monoid.

Now, what is a monoid object in \(\mathbf{Mon}\)?

\(\eta : 1 \to M\) is a monoid morphism so it has to send the single object of \(1\) to the unit of \(M\). Hence \(\eta\) is entirely constrained and uninteresting.

\(\mu : M \times M \to M\) has to be a monoid map. That is, \(\mu((a,b) \circ (c,d)) = \mu(a,b) \circ \mu(c,d)\), i.e. \(\mu(a \circ c, b \circ d) = \mu(a,b) \circ \mu(c,d)\). So \(\mu\) has to “commute” with \(\circ\). This is precisely the condition needed to apply Eckmann-Hilton.

Monoid object is also required to satisfy unital and associativity laws, but we can already deduce those from Eckmann-Hilton.

2-categories

2-categories 1

https://www.youtube.com/watch?v=k-RehY4tLdI

Generalization of categories: not just objects and morphisms, but also (2-)morphisms between the (1-)morphisms. Primordial example: categories, functors, and natural transformations.

Note: today, strict 2-categories, i.e. everything will hold on the nose rather than up to isomorphism. A bit immoral of us. [If we let things be a bit looser we get bicategories?]

Recall: a (small) category \(C\) is given by

equipped with

To make this into a 2-category, we take the set of morphisms and categorify it. That turns some of the above functions into functors. Thus, a \(2\)-category \(C\) is given by a set of objects along with

(Note: why not turn the set of objects into a category? That’s a good question. Turns out we would get something different.)

Let’s unravel this a bit. If \(C(x,y)\) is a category then the objects are morphisms (of \(C\)) \(x \to y\), and there can also be morphisms (of \(C(x,y)\)) between these morphisms: \(2\)-cells. \(2\)-cells can be composed (“vertical” composition).

We also have the composition functor \(C(y,z) \times C(x,y) \to C(x,z)\). On “objects” (which are \(1\)-cells in \(C\)) the action of this functor is just the usual composition of \(1\)-cells. On morphisms (i.e. \(2\)-cells), it gives us “horiztonal” composition.

Next time: how functoriality gives us the interchange law.

2-categories 2

http://youtu.be/DRGh-HESyag

Interchange in a 2-category comes from functoriality of the composition functor. The key is to remain calm.

The functor is \(C(y,z) \times C(x,y) \to C(x,z)\). On morphisms, it sends pairs of \(2\)-cells to a single \(2\)-cell, the horizontal composite. What does functoriality mean? It means if we have two (vertically!) composable pairs of \(2\)-cells; the functor on their composition (i.e doing vertical composition pointwise) is the same as applying the functor to each (i.e. first doing the horizontal compositions) and then composing (vertically).

Eckmann-Hilton

Eckmann-Hilton 1

https://www.youtube.com/watch?v=Rjdo-RWQVIY

NOTE: There seems to be no catsters video actually explaining what a “bicategory” is. According to the nlab it is a weaker version of a 2-category, where certain things are required to hold only up to coherent isomorphism rather than on the nose.

Eckmann-Hilton argument. Originally used to show all higher homotopy groups are Abelian. We can use it for other things, e.g.

Idea: given a set with two unital binary operations, they are exactly the same, and commutative — as long as the operations interact in a certain coherent way.

Given a set with two binary operations \(\circ\) and \(\star\), such that

then \(\circ = \star\), and the operation is commutative.

Geometric intuition: \(\circ\) and \(\star\) could be vertical and horizontal composition of \(2\)-cells in a bicategory. Then distributivity is just the interchange law.

Proof: use the “Eckmann-Hilton clock”. See video for pictures. Given e.g. \(a \circ b\), “rotate” \(a\) and \(b\) around each other by inserting units and using interchange law.

In fact, it is not necessary to require that the two units are the same: it is implied by the interchange law. Left as an exercise.

Eckmann-Hilton 2

https://www.youtube.com/watch?v=wnRqo7UHa-k

This time, show the interchange law implies the units are the same and associativity.

Let \(v\) be the vertical unit and \(h\) the horizontal unit. Then \(((v \circ h) \star (h \circ v)) = h \star h = h\) but also by interchange law it is equal to \(((v \star h) \circ (h \star v)) = v \circ v = v\), hence \(h = v\).

\((a \circ 1) \star (b \circ c) = a \star (b \circ c)\); interchange gives \((a \star b) \circ (1 \star c) = (a \star b) \circ c\). Since the two operations have to be the same, this gives associativity.

Example. A (small) \(2\)-category with only one \(0\)-cell and only one \(1\)-cell is in fact a commutative monoid. Underlying set is set of \(2\)-cells. Operation is either \(\circ\) or \(\star\), which by Eckmann-Hilton are the same and commutative.

Bicategory case is a bit more complicated, since horizontal composition is not strictly unital. A bicategory with only one \(0\)-cell is a monoidal category. A bicategory with only one \(1\)-cell is a commutative monoid.

Distributive laws

Distributive laws 1

https://www.youtube.com/watch?v=mw4IhOLhDwY

Monads represent algebraic structure; a distributive law says when two algebraic structures interact with each other in a coherent way. Motivating example: multiplication and addition in a ring.

Let \(S\), \(T\) be monads on a category \(C\). A distributive law of \(S\) over \(T\) is a natural transformation \(\lambda : ST \to TS\), satisfying the “obvious” axioms: \(\lambda\) needs to interact properly with the monad structure of \(S\) and \(T\), that is:

Example: \(C = \mathbf{Set}\). \(S\) = free commutative monoid monad (“multiplication”), \(T\) = free abelian group monad (“addition”). Define \(\lambda_X : STX \to TSX\): \(TX\) is formal sums of elements of \(X\), like \((a + b + c)\); \(S\) constructs formal products. So we have to send things like \((a + b)(c + d)\) to formal sums of formal products, \(ac + bc + ad + bd\).

In fact we have constructed the free ring monad, \(TS\).

If we start with a monoid and consider the free group on its underlying elements, we can define a product using distributivity; so the free group on a monoid is a group. Formally, the free group monad lifts to the category of monoids (?).

Distributive laws 2

https://www.youtube.com/watch?v=TLgjH9Y8HOc

More abstract story behind our favorite example: combining a group and a monoid to get a ring.

Note: distributive law (at least in this example) is definitely non-invertible: you can turn a product of sums into a sum of products, but you can’t necessarily go in the other direction.

Main result: A distributive law \(\lambda : ST \to TS\) is equivalent to a lift of \(T\) to a monad \(T'\) on \(S\)-\(\mathbf{Alg}\). \(TS\) becomes a monad, and \(TS\)-\(\mathbf{Alg}\) is equivalent to \(T'\)-\(\mathbf{Alg}\).

When is \(TS\) a monad? We need \(\mu : TSTS \to TS\); can do this if we have \(\lambda : ST \to TS\), then use \(\mu_T \mu_S\). The laws for a distributive law ensure that this satisfies the monad laws.

Distributive law is equivalent to a lift of \(T\) to a monad on \(S\)-\(\mathbf{Alg}\)?

\(TS\)-\(\mathbf{Alg}\) is equivalent to \(T'\)-\(\mathbf{Alg}\)?

Distributive laws 3 (aka Monads 6)

https://www.youtube.com/watch?v=g-SCYArh5RY

Recall that a monad is a functor together with some natural transformations; we can see this as a construction in the \(2\)-category of categories, functors, and natural transformations. We can carry out the same construction in any \(2\)-category \(C\), giving monads in \(C\).

Let \(C\) be a \(2\)-category (e.g. \(\mathbf{Cat}\)). A monad in \(C\) is given by

satisfying the usual monad axioms.

In fact, we get an entire \(2\)-category of monads inside \(C\)!

What is a morphism of monads? A monad functor \((X_1, S_1, \eta_1, \mu_1) \to (X_2, S_2, \eta_2, \mu_2)\) (i.e. a \(1\)-cell in the \(2\)-category of monads in \(C\)) is given by

satisfying the axioms:

A monad transformation (i.e. a \(2\)-cell in the \(2\)-category of monads in \(C\)) is given by

Distributive laws 4

https://www.youtube.com/watch?v=FZeoHPRoBVk

Distributive laws, even more formally!

Consider the \(2\)-category of monads \(\mathbf{Mnd}(C)\) in an arbitrary \(2\)-category \(C\); monads in \(\mathbf{Mnd}(C)\) are distributive laws!

Recall that a monad in an arbitrary \(2\)-category is a \(0\)-cell equipped with an endo-\(1\)-cell and appropriate \(2\)-cells \(\eta\) and \(\mu\). In \(\mathbf{Mnd}(C)\):

Summarizing more concisely/informally, a monad in \(\mathbf{Mnd}(C)\) is

Consider the map \(C \mapsto \mathbf{Mnd}(C)\). This actually defines an endofunctor \(\mathbf{Mnd}(-)\) on \(2\)-\(\mathbf{Cat}\), the category of (strict) \(2\)-categories and (strict) \(2\)-functors. In fact, Street showed that \(\mathbf{Mnd}(-)\) is a monad! The “monad monad”.

The multiplication has type \(\mathbf{Mnd}(\mathbf{Mnd}(C)) \to \mathbf{Mnd}(C)\). Recall that objects in \(\mathbf{Mnd}(\mathbf{Mnd}(C))\) are a pair of monads \(S\),\(T\) plus a distributive law. In fact, the distributive law is precisely what is needed to make \(TS\) into a monad, which is the monad returned by the multiplication.

Group Objects and Hopf Algebras

Group Objects and Hopf Algebras 1

https://www.youtube.com/watch?v=p3kkm5dYH-w

Take the idea of a group and develop it categorically, first in the category of sets and then transport it into other categories (though it may not be completely obvious what properties of \(\mathbf{Set}\) we are using).

A group is of course a set \(G\) with an associative binary product, inverses, and an identity element. Let’s make this categorical: don’t want to talk about internal structure of \(G\) but just about \(G\) as an object in \(\mathbf{Set}\).

So a group is:

together with axioms expressed as commutative diagrams:

where \(\Delta : G \to G \times G\) is the diagonal map (note the fact that we are using \(\Delta\) is the most interesting part; see forthcoming lectures) and \(\varepsilon : G \to 1\) is the unique map to a terminal set.

Group Objects and Hopf Algebras 2

https://www.youtube.com/watch?v=kJ2X_U7X5WA

Note just \(\mu\) and \(\eta\) together with axioms (forgetting about \(\gamma\) and its axioms) is the definition of a monoidal category. Not surprising since a group is a monoid with inverses.

Recall \(\Delta : G \to G \times G\). We get that for free from the fact that the monoid we are using is really the categorical product; \(\Delta\) can be easily defined using the universal property of categorical product.

In fact, every set \(S\) is a comonoid in a unique way, since \(\times\) is a categorical product. That is, a comonoid on a set \(S\) is given by

And note we used \(\Delta\) and \(\varepsilon\) in the definition of a group, in particular in the axioms for \(\gamma\).

Group Objects and Hopf Algebras 3

https://www.youtube.com/watch?v=wAeHrtKMTHM

The definition given last time won’t work in general for any monoidal category, but it does work for any Cartesian category (that is, monoidal categories where the monoidal operation is categorical product). Examples of Cartesian categories, in which it therefore makes sense to have group objects, include:

Let’s see what a group object looks like in each of these examples.

What about non-Cartesian monoidal categories? Simplest example is \(\mathbf{Vect}\), category of (finite-dimensional) vector spaces with linear maps. Monoidal structure given by tensor product and complex numbers. Tensor product defined by

\(V \otimes W = \{ \sum_t \alpha_t (v_t \otimes w_t) \mid v_t \in V, w\_t \in W \} / [(\alpha_1 v_1 + \alpha_2 v_2) \otimes w \sim \alpha_1 (v_1 \otimes w) + \alpha_2(v_2 \otimes w) \text{ and symmetrically}]\)

Suppose \(\{v_i\}\) is a basis for \(V\) and \(\{w_j\}\) is a basis for \(W\), then \(\{v_i \otimes w_j\}\) is a basis for \(V \otimes W\).

The point is that \(\mathrm{dim}(V \otimes W) = \mathrm{dim}(V) \times \mathrm{dim}(W)\), but that’s different than \(\mathrm{dim}(V \times W) = \mathrm{dim}(V) + \mathrm{dim}(W)\), so \(\mathbf{Vect}\) is not Cartesian.

Group Objects and Hopf Algebras 4

https://www.youtube.com/watch?v=zZn9ZETVkF8

We still want to be able to define group objects in monoidal categories which are not Cartesian.

Recall: if we have a monoidal category \((C, \times, 1)\) where \(\times\) is the categorical product, then every object \(X \in C\) is a comonoid \((X, \Delta, \varepsilon)\) in a unique way, and every morphism is a comonoid map.

Notation: in \(\mathbf{Set}\), an object with an associative binary operation and an identity is called a monoid; in \(\mathbf{Vec}\) it’s called an algebra. So when we generalize to arbitrary categories sometimes “monoid” is used, sometimes “algebra”.

A Hopf algebra is a group object in a general monoidal (tensor) category. Details next time.

Group Objects and Hopf Algebras 5

https://www.youtube.com/watch?v=gmxZ_KCRZho

A Hopf algebra \(H\) in a (braided) monoidal category is as follows. We don’t get comonoid stuff for free any more so we have to put it in “by hand”.

(See video for string diagrams.) Note the monoid and comonoid also need to be “compatible”: this is where the braidedness comes in. In particular \(\mu\) and \(\eta\) need to be comonoid morphisms. So we need \(H \otimes H\) to be a coalgebra.

Lemma: suppose \(H\), \(K\) are comonoids. Then \(H \otimes K\) is a coalgebra if the category is braided: \(H \otimes K \to (H \otimes H) \otimes (K \otimes K)\) using comonoid structures on \(H\) and \(K\), and then using (associativity and) braiding we can flip inner \(H \otimes K\) around to get \((H \otimes K) \otimes (H \otimes K)\).

Can then write down what it means for \(\mu\) to be a coalgebra map aka comonoid morphism; left as an exercise (or the next video).

Group Objects and Hopf Algebras 6

https://www.youtube.com/watch?v=Gv1sRLOwVWA

String diagram showing comonoid \(\Delta\) for \(H \otimes K\).

\(\mu\) and \(\eta\) should be a comonoid morphism, i.e. must commute with \(\Delta\) (string diagram) and also with \(\varepsilon\) (another string diagram).

There seems to be some asymmetry: monoid + comonoid + monoid must be comonoid morphisms. But it’s the same to say that the comonoid must be monoid morphisms.

Ends

Ends 1

http://youtu.be/mxI9ba6Rexc

Given a functor \(T : C^{op}\times C \to D\), an end \(\int_{c \in C} T(c,c)\) is an object in \(D\) which is “limit-like” in some sense.

Ends are not as common as coends (and perhaps not as intuitive?). Two particular places where ends do show up:

Definition:

Note we write the object \(E\) using the intergral notation, \(\int_{c \in C} T(c,c)\) (the morphisms of the wedge are left implicit).

Ends 2

http://youtu.be/gyc86NFT0Sw

Simple example of an end: \(T = \mathrm{Hom} : C^{\text{op}} \times C \to \mathbf{Set}\). In this case a wedge \(x \stackrel{\bullet}{\to} T\) consists of:

That is, for every \(n \in x\) we have \(\omega_c(n) = n_c : c \to c\), such that \(f \circ n_c = c_{c'} \circ f\). i.e. the family \(n_c\) are the components of a natural transformation \(Id_C \to Id_C\).

Note this goes in the other direction too, that is, a wedge \(x \stackrel{\bullet}{\to} \mathrm{Hom}\) is precisely the same thing as a function \(x \to \mathrm{Nat}(Id_C, Id_C)\). Therefore, the universal such \(x\) is precisely this set of natural transformations. (Can be thought of as “set of symmetries” of a category. Also the Hochschild cohomology.)

Ends 3

http://youtu.be/TfSUxhCNZZ0

More examples. First, straightforward generalization: given functors \(F, G : C \to E\), form the bifunctor \(\mathrm{Hom}_E(F(-), G(-)) : C^{op}\times C \to \mathbf{Set}\). Then we can see that

\(\int_{c \in C} \mathrm{Hom}_E(F(c),G(c)) = \mathrm{Nat}(F,G)\).

(Proof is just a small generalization of the proof in Ends 2, left as exercise.) Useful in an enriched context, can use this end to construct an object of natural transformations instead of a set.

Another example, “baby Tannaka reconstruction” (see Tannaka duality and reconstruction theorem on nlab).

Proof (application of Yoneda):

Ends 4

http://youtu.be/3hOtm0paWXY

Combine some of the previous examples. Recall

What happens if we combine these two results? First, look at the end from last time:

Now look at the end of the bare hom-functor in the category of \(M\)-sets. i.e. \(\int_{M\text{-}\mathbf{Set}} \mathrm{Hom}_{M\text{-}\mathbf{Set}}(-,-) = ?\)

Adjunctions from morphisms

Adjunctions from morphisms 1

https://www.youtube.com/watch?v=SzzHjpRmrLU

General phenomenon: associate some category \(C(X)\) to an object \(X\). For example:

Question: if we have a morphism \(f : X \to Y\), how does that relate to the categories \(C(X)\) and \(C(Y)\) associated to \(X\) and \(Y\)?

We often get some sort of “pullback” functor \(f^* : C(X) \leftarrow C(Y)\). (Also often get some sort of monoidal structure on \(C(X)\) and \(C(Y)\), and \(f^*\) is often monoidal.)

We also get various “pushforwards” \(f_* : C(X) \to C(Y)\), right adjoint to \(f^*\). In some situation we also get a left adjoint to \(f^*\).

This is the beginning of the story of “Grothendieck’s 6 operations”. Lots of similar structure arises in all these different areas.

Adjunctions from morphisms 2

https://www.youtube.com/watch?v=jAQfNGEOass

Baby examples of some particular adjunctions (in generality, they show up in Grothendieck’s 6 operations, Frobenius reciprocity, …). Idea: start with (e.g.) sets; to each set associate a category; to each morphism between sets we will get functors between the categories.

Adjunctions from morphisms 3

https://www.youtube.com/watch?v=tWhB7E-HS8Y

(Foreshadowing: taking the disjoint union gives us another adjoint.)

Adjunctions from morphisms 4

http://youtu.be/Cf7hCiTspJc

Proof of the adjunction \(f^* \dashv f_*\). (Come up with your own mnemonic to remember which way around the adjunction goes; suggested: think of a “falling star”.)

Adjunctions from morphisms 5

https://www.youtube.com/watch?v=MSOGEtW39qM

Last time, we proved an adjunction \(f^* \dashv f_*\), i.e.

\(\mathrm{Hom}_X (f^* \hat F, \hat E) \cong \mathrm{Hom}_Y (\hat F, f_* \hat E).\)

In fact, we showed that both are isomorphic to

\(\prod_{x \in X} \mathrm{Hom}(\hat F (f(x)), \hat E(x)).\)

i.e. given some \(f : X \to Y\), for each \(x\) we get a map going the other way, from the fiber over \(f(x)\) to the fiber over \(x\). (See the video for a nice picture.) But we can imagine turning these maps around, giving

\(\prod_{x \in X} \mathrm{Hom}(\hat E(x), \hat F(f(x))).\)

Using the same trick as last time, this is equivalent to \(\prod_{y \in Y} \prod_{x \in f^{-1}(y)} \mathrm{Hom} (\hat E (x), \hat F (y))\), which is in turn equivalent to

\(\prod_{y \in Y} \mathrm{Hom} (\coprod_{x \in f^{-1}(y)} \hat E (x), \hat F (y))\)

(since \(\mathrm{Hom}(-,B)\) turns limits into colimits; concretely, note that \(A^X A^Y = A^{X + Y}\)).

This gives us a left adjoint \(f_! \dashv f^*\), defined by

\(\displaystyle (f_! \hat E)(y) = \coprod_{x \in f^{-1}(y)} \hat E(x).\)

Remark: note that if we view bundles over \(X\) as objects of the slice category \(\mathbf{Set}/X\), \(f_!\) is just composition.

Double Categories

Double Categories

https://www.youtube.com/watch?v=kiCZiSA2W3Q

Internal categories in \(\mathbf{Cat}\). Recall that an internal category in \(E\) is a pair of objects \(A\) (representing objects) and \(B\) (representing morphisms), and a pair of parallel arrows \(s,t : B \to A\) in \(E\) recording the source and target of each morphism, all suitably equipped with unit and composition.

If \(A\) and \(B\) are themselves categories, and \(s\) and \(t\) are functors, then \(B\) itself has sets of objects \(B_0\) and morphisms \(B_1\) with source and target functions, and the same for \(A\). Then the functors \(s\) and \(t\) have actions on morphisms and objects, so we get a square with two parallel arrows on each side.

What about composition? Note \(B\) and \(A\) already come equipped with composition, which together give us “vertical composition” of \(2\)-cells. Composition in the internal category gives horizontal composition of \(2\)-cells.

Note if all vertical \(1\)-cells are identities, this collapses to the usual idea of a \(2\)-category. (Or symmetrically, with horizontal \(1\)-cells being identities.)

Spans

Spans 1

https://www.youtube.com/watch?v=SQfUXOCMUhI

NOTE: There seems to be no catsters video actually explaining what a “bicategory” is. According to the nlab it is a weaker version of a 2-category, where certain things are required to hold only up to coherent isomorphism rather than on the nose.

Let \(E\) be a category with (chosen) pullbacks. \(\mathbf{Span}(E)\) is a bicategory with

Can check all the axioms etc.

Now, note monads can be constructed inside any bicategory, and are given by

It turns out that monads in \(\mathbf{Span}(E)\) are great! For example, monads in \(\mathbf{Span}(\mathbf{Set})\) are small categories. Next time we’ll see why.

Spans 2

https://www.youtube.com/watch?v=Jn5dZuebeXU

Monads in \(\mathbf{Span}(Set)\) are small categories. These notes make a lot more sense when you can look at the diagrams. Watch the video or work out the diagrams yourself.

We have

And of course there are some monad laws which amount to the category laws.

More generally, monads in \(\mathbf{Span}(E)\) are categories internal to \(E\). I.e.

Multicategories

Multicategories 1

http://www.youtube.com/watch?v=D_pPNgGZYDs

Like categories, but morphisms have multiple objects as their source.

A (small) multicategory \(C\) is given by

Note that one can have a morphism with no inputs.

This can all be expressed nicely using the “free monoid monad” (i.e. list monad). Let \(T\) be the free monoid monad on \(\mathbf{Set}\), i.e. the list monad; that is, \(T\) sends each set \(S\) to the free monoid on \(S\) (i.e. lists of \(S\)).

Make a bicategory of \(T\)-spans. Just as monads in the category of spans were small categories, monads in the (bi)category of \(T\)-spans are multicategories.

\(T\)-span has:

Bicategory axioms follow from monad laws for \(T\). Next time: monads in this category are multicategories.

Multicategories 2

https://www.youtube.com/watch?v=WytjdlserwU

We’ve seen that monads in Span are categories.

We’ve seen a category of \(T\)-spans, spans with a \(T\) on the left. We’ll see that monads in \(T\)-\(\mathbf{Span}\) are multicategories.

Recall that \(T\) is the list monad.

A monad in \(T\)-\(\mathbf{Span}\) is:

Key point: we can actually do this with other monads \(T\)! And even on other categories with pullbacks, as long as \(T\) preserves pullbacks (and \(\eta\) and \(\mu\) commutative diagrams are pullbacks). This yields a notion of a \(T\)-multicategory. The source of each morphism is not just a list of objects but a \(T\)-structure of objects.

Metric Spaces and Enriched Categories

Metric Spaces and Enriched Categories 1

https://www.youtube.com/watch?v=be7rx29eMr4

Idea due to Lawvere. A metric \(d\) on a metric space satisfies:

Compare to the data for a category, written in a slightly funny way:

These look remarkably similar! In fact, they are both examples of enriched category. We’ll start with a normal category and show how to generalize it to an enriched category.

Let \(C\) be a category. We have:

Important thing to note: composition and identity are morphisms in \(\mathbf{Set}\). What properties of \(\mathbf{Set}\) have we used? Just a Cartesian product and the one-element set \(\{\star\}\). Right generalization is a monoidal category.

In particular, if \((V, \otimes, I)\) is a monoidal category, we can define categories enriched in \(V\). Definition becomes:

\(C\) is a \(V\)-category (category enriched in \(V\)):

e.g. pick \(V\) to be category of Abelian groups (yields “additive category”), or category of vector spaces (yields “linear categories”).

What if we take \(V\) to be a non-concrete category? e.g. take the poset of nonnegative real numbers under \(\geq\). Can make this monoidal by taking the usual \(+\), identity is \(0\). Then it turns out that categories enriched in this poset category are metric spaces!

Metric Spaces and Enriched Categories 2

https://www.youtube.com/watch?v=0p3iS3Nf-fs

Explains in more detail how categories enriched in \((\mathbb{R}^+, \geq)\) poset (with monoidal structure given by \(+\) and \(0\)) are metric spaces.

This is actually a generalized metric space. More general than a metric space in several ways:

Now we can study metric spaces categorically.

Given two \(V\)-categories, a \(V\)-functor \(\Phi : C \to D\) consists of

In the generalized metric setting, such a \(\Phi\) is a nonexpansive map.

Metric Spaces and Enriched Categories 3

https://www.youtube.com/watch?v=kMSt_Ci54BE

Enriched natural transformations. There are actually various ways to generalize. Today: simple-minded version. Apparently if we generalize the notion of a set of natural transformations we get a slightly better definition—this will be covered in future videos. [editor’s note: to my knowledge no such future videos exist.]

Standard definition of a natural transformation \(\theta : F \to G\) given functors \(F,G : C \to D\). Problem: we are supposed to have \(\theta_x \in \mathrm{Hom}(F x, G x)\), but in the enriched setting \(\mathrm{Hom}(F x, G x)\) may not be a set, but just some object. So what should it mean for \(\theta_x\) to be an “element” of it?

Simple way around this: let \((V, \otimes, 1)\) be a monoidal category. We can define a canonical functor \(\Gamma : V \to \mathrm{Set}\) (“generalized elements”) which takes an object \(v \in V\) to the Hom-set \(\mathrm{Hom}(1,v)\).

e.g. if \(V = \mathrm{Set}\), \(\Gamma\) is the identity. Another example I don’t understand involving category of complex vector spaces.

In the example we care about, \(V = \mathbb{R}^+\), and \(I = 0\). In this case \(\Gamma\) sends \(v\) to the Hom-set \(0 \geq v\), that is, \(\{\star\}\) if \(v = 0\) and the empty set otherwise.

So now we can say that \(\theta_x\) should be a generalized element of \(\mathrm{Hom}(F x, G x)\).

So, let \(F, G\) be \(V\)-functors. Define a \(V\)-natural transformation \(\theta : F \to G\) as