Skip to the Main Content

Note:These pages make extensive use of the latest XHTML and CSS Standards. They ought to look great in any standards-compliant modern browser. Unfortunately, they will probably look horrible in older browsers, like Netscape 4.x and IE 4.x. Moreover, many posts use MathML, which is, currently only supported in Mozilla. My best suggestion (and you will thank me when surfing an ever-increasing number of sites on the web which have been crafted to use the new standards) is to upgrade to the latest version of your browser. If that's not possible, consider moving to the Standards-compliant and open-source Mozilla browser.

May 24, 2007

Cohomology and Computation (Week 24)

Posted by John Baez

In the 1950s, Cartan, Eilenberg, Mac Lane and others systematically studied the cohomology of many different algebraic gadgets — groups, associative algebras, Lie algebras, and so on. It was later realized that underlying all these different cohomology theories there’s a marvelous unifying idea: the bar construction. In this week’s seminar on Cohomology and Computation, we began trying to understand what makes the bar construction tick:

  • Week 24 (May 17) - The bar construction. Why do adjoint functors give simplicial objects? First, Δ alg\Delta_{alg} is the free monoidal category on a monoid object — or "the walking monoid", for short. Second, adjoint functors give certain monoids, called "monads".

    Supplementary reading:

Last week’s notes are here; next week’s notes are here.

Why is Δ alg\Delta_{alg}, the free monoidal category on a monoid, called the “walking monoid”?

This is one of Jim Dolan’s jokes. Supppose some guy has really big bushy eyebrows, so that when you see him walking down the street you first notice his eyebrows and only later realize there’s a person attached. Then you might call him a "walking pair of eyebrows". Such a guy is basically just a minimal life support system for his eyebrows!

In math, we see the same phenomenon in things like “the free group on 2 elements”. This is a group whose sole purpose in life is to have 2 elements. To pick out 2 elements in any group GG, we just need a homomorphism from the free group on 2 elements to GG. Similarly, the monoidal category Δ alg\Delta_{alg} is a monoidal category whose sole purpose in life is to contain a monoid object.

I explained some variants of this idea back in week173 and week174 — namely, the walking equivalence, the walking adjoint equivalence, the walking adjunction, and the walking ambidextrous adjunction. Aaron Lauda discusses the walking adjunction and the eyebrows joke on page 20 of his paper Frobenius algebras and ambidextrous adjunctions. For more of the history of this joke, and how Jim sometimes feels uncomfortable when I tell his jokes in public — whoops! — read the interview by Bruce Bartlett.

Posted at May 24, 2007 12:43 AM UTC

TrackBack URL for this Entry:   https://golem.ph.utexas.edu/cgi-bin/MT-3.0/dxy-tb.fcgi/1285

9 Comments & 1 Trackback

Re: the Walking Monoid

There’s another nice fact about Δ\Delta which fits with the picture of “the walking monoid”, but one degree up in categorical dimension.

First, Δ\Delta, being the category of finite ordinals and order-preserving maps, is a 2-category: the 2-cells are instances fgf \leq g of the pointwise order between order-preserving maps. In fact, Δ\Delta with its ordinal sum ++ forms a monoidal 2-category.

Next, taking into account the 2-category structure, there are some adjoint relationships between the 1-cells. Let u:01u: 0 \to 1 and m:21m: 2 \to 1 be the unique maps to the terminal object 1; these of course are the unit and multiplication on the monoid 1. Then, there is a string of adjunctions

(1+u:12)(m:21)(u+1:12).(1 + u: 1 \to 2) \dashv (m: 2 \to 1) \dashv (u + 1: 1 \to 2).

This is very easy to see; for instance, the counit 2-cell m(u+1)id 1m(u + 1) \leq id_1 is an equality, and the unit 2-cell id 2(u+1)mid_2 \leq (u + 1)m is also clear, because (u+1)m(u + 1)m is the function taking both elements of 2 = {0, 1} to the maximal element 1.

Monads for which the multiplication is left adjoint to the unit are called Kock-Zöberlein monads; they were first pointed out in 1972 by Anders Kock. They often arise as “cocompletion monads” of one kind of another. A typical example: consider the ‘finite coproduct completion’ of a category CC, sometimes denoted Fam(C)Fam(C). The objects of Fam(C)Fam(C) are finite tuples (c 1,...,c n)(c_1, ..., c_n) of objects of CC, and a morphism

(c 1,...,c n)(d 1,...,d m)(c_1, ..., c_n) \to (d_1, ..., d_m)

consists of a function f:[n][m]f: [n] \to [m] (where [n] = {1, …, n}) together with an n-tuple of morphisms g j:c jd f(j)g_j: c_j \to d_{f(j)} of CC, composed in the obvious way. [If you know about categorical wreath products, Fam(C) can also be expressed as a wreath product FinCFin \int C.] Finite coproducts in Fam(C)Fam(C) are formed by concatenating tuples, and it is not hard to convince oneself that the obvious inclusion i:CFam(C)i: C \to Fam(C) has the universal property:

Any functor F:CDF: C \to D to a category with finite coproducts DD can be extended along ii to a coproduct-preserving functor Fam(C)DFam(C) \to D, uniquely so up to canonical isomorphism.

Now, if CC already has finite coproducts, then the counit, i.e., the coproduct-preserving functor s=sum:Fam(C)Cs = sum: Fam(C) \to C which extends the identity on CC, is left adjoint to the inclusion i:CFam(C)i: C \to Fam(C), i.e., the unit! The counit of sis \dashv i is an isomorphism, and the unit, applied to an object (c 1,...,c n)(c_1, ..., c_n) in Fam(C), is just the tuple of canonical injections into the coproduct ic i\sum_i c_i in CC:

(c 1,...,c n)( ic i).(c_1, ..., c_n) \to (\sum_i c_i).

In cases like this, where the counit of a bi-adjunction is left adjoint to the unit in this manner, we say that the bi-adjunction is KZ or Kock-Zöberlein.

Here is a more precise definition: a bi-adjunction (F:CB)(G:BC)(F: C \to B) \dashv (G: B \to C) between bicategories, with unit u:1 CGFu: 1_C \to G F and counit c:FG1 Bc: F G \to 1_B and triangulators

s:(Gc)(uG)1 Gs: (G c)(u G) \stackrel{\sim}{\to} 1_G

t:1 F(cF)(Fu),t: 1_F \stackrel{\sim}{\to} (c F)(F u),

is said to Kock-Zöberlein if ss is the counit of an adjunction GcuGG c \dashv u G (equivalently, if tt is the unit of an adjunction FucFF u \dashv c F; proving that equivalence is a fun exercise in bicategorical algebra). Similarly, there is a notion of Kock-Zöberlein (pseudo)monad (M,m,u)(M, m, u), where there is a string of adjunctions

MumuM.M u \dashv m \dashv u M.

If you’ve followed me this far, you will have no trouble guessing the punch line:

Theorem (Kock): The monoidal 2-category Δ\Delta is the “walking Kock-Zöberlein monoid.”

See the penultimate paper referenced here for details.

I think this result is pretty widely known within the categorical community, but a quick Google search on Kock-Zöberlein or KZ monads didn’t garner a huge number of hits, so I think the result probably bears repeating here. I haven’t thought much about what implications this result would have with respect to the bar construction.

Posted by: Todd Trimble on May 26, 2007 1:04 AM | Permalink | Reply to this

Re: the Walking Monoid

In my previous post, I should have added (I think it’s pretty cool, anyway) that in Δ\Delta, if you look at the n+1n+1 injections

d 0...d n:[n][n+1]d_0 \leq ...\leq d_n: [n] \to [n+1]

and the nn surjections

s 1...s n:[n+1][n]s_1 \leq ... \leq s_n: [n+1] \to [n]

then what you get is a big long string of adjunctions

d 0s 1d 1s 2...s nd nd_0 \dashv s_1 \dashv d_1 \dashv s_2 \dashv ... \dashv s_n \dashv d_n

as a result of the KZ property.

Posted by: Todd Trimble on May 26, 2007 1:41 AM | Permalink | Reply to this

Re: the Walking Monoid

That is cool. Thanks for explaining all this stuff about Kock-Zöberlein monads. I need to lock myself in the cellar and ponder them sometime.

I occasionally feel I understand them — thanks mainly to your attempts to explain them, but also to the prevalence of ‘weakly idempotent’ pseudomonads on CatCat having the flavor of ‘cocompletion processes’: I’ve seen lots of those in my time!

But, the relation of all this to the monoidal 2-category version of Δ\Delta seems mysterious to me, and phrases like “ss is the counit of an adjunction GcuGG c \dashv u G” make my mind freeze like a deer in the headlights.

I don’t think there’s any more you can do to help: I just need to do some calculations to see for myself how this stuff works!

if you look at the n+1n+1 injections

d 0...d n:[n][n+1]d_0 \leq ...\leq d_n: [n] \to [n+1]

and the nn surjections

s 1...s n:[n+1][n]s_1 \leq ... \leq s_n: [n+1] \to [n]

then what you get is a big long string of adjunctions

d 0s 1d 1s 2...s nd nd_0 \dashv s_1 \dashv d_1 \dashv s_2 \dashv ... \dashv s_n \dashv d_n

as a result of the KZ property.

Cool! I vaguely seem to remember results that say “any sufficiently long string of adjunctions has to be sort of degenerate in some way.” I’ve looked for such results, but I don’t remember what I found. This example constrains what theorems I could possibly be imagining.

Do you know any results along these lines?

Posted by: John Baez on May 26, 2007 2:52 AM | Permalink | Reply to this

Adjoint strings

I vaguely seem to remember results that say “any sufficiently long string of adjunctions has to be sort of degenerate in some way.” I’ve looked for such results, but I don’t remember what I found. This example constrains what theorems I could possibly be imagining. Do you know any results along these lines?

There’s an absolutely remarkable theorem of this flavor due to Richard Wood and Bob Rosebrugh; I don’t know if it’s close to what you were trying to remember. I think it goes something like this:

Start with the Yoneda embedding

y:CSet C op.y: C \to Set^{C^{op}}.

As we all know, yy preserves all limits which happen to exist in CC. But the condition that yy is a right adjoint: that turns out to be a very strong cocompleteness condition on CC. In that case, people sometimes say that CC is total (don’t know who made up that term).

Presheaf categories (and Grothendieck toposes generally) are total in this sense. In the presheaf case, the left adjoint mm to the Yoneda embedding

y:Set C opSet (Set C op) opy: Set^{C^{op}} \to Set^{(Set^{C^{op}})^{op}}

is (somewhat interestingly)

Set y C op:Set (Set C op) opSet C op.Set^{y_C^{op}}: Set^{(Set^{C^{op}})^{op}} \to Set^{C^{op}}.

(As an aside, the condition that the Yoneda embedding on CC have a left exact left adjoint is actually rather close to the condition that CC is a Grothendieck topos!).

Now, in the presheaf case Set C opSet^{C^{op}}, the left adjoint to Yoneda has itself a left adjoint

L y:Set C opSet (Set C op) opL_y: Set^{C^{op}} \to Set^{(Set^{C^{op}})^{op}} given essentially by taking left Kan extension along a Yoneda embedding. (I’m not sure about this, but the existence of such a string of left adjoints which starts from the Yoneda embedding y Cy_C and proceeds leftwards:

Lmy CL \dashv m \dashv y_C

might be equivalent to the condition that CC is a presheaf topos.)

Can the left Kan extension

L y:Set C opSet (Set C op) opL_y: Set^{C^{op}} \to Set^{(Set^{C^{op}})^{op}}

have a left adjoint? It can (but the noose is gradually tightening). For example, if CC is a small sup-lattice, then y:CSet C opy: C \to Set^{C^{op}} has a left adjoint ξ:Set C opC\xi: Set^{C^{op}} \to C, and the left adjoint to L yL_y would be a left Kan extension L ξL_{\xi} along ξ\xi (more or less).

“Okay, smarty-pants, can this left adjoint L ξL_{\xi} have a left adjoint?” Well, yes it can, but now we’re really getting to the end of the rope: if the sup-lattice CC in the previous paragraph were the terminal category 1, then ξ:Set1\xi: Set \to 1 has a left adjoint i:1Seti: 1 \to Set, namely the initial object (and the left adjoint to L ξL_{\xi} would then be a left Kan extension L iL_i). That’s the absolute limit to how far we can go.

So, then, the Rosebrugh-Wood results are:

  • Theorem: There exists a string of adjoints jklmy C,j \dashv k \dashv l \dashv m \dashv y_C, where y Cy_C is the Yoneda embedding of CC, if and only if CC is equivalent to SetSet.
  • Corollary: There is no such adjoint string ijklmy C.i \dashv j \dashv k \dashv l \dashv m \dashv y_C.

I see this result is downloadable here. (I saw Richard Wood give a seminar on this theorem at Macquarie; it was impressive.)

Posted by: Todd Trimble on May 26, 2007 4:32 AM | Permalink | Reply to this

Re: Cohomology and Computation (Week 24)

There was one other thing about the bar construction which I wanted to mention, since John told me that his course is ending soon and he might not get around to mentioning it himself.

As John said, the bar construction is a wonderful gadget which can be used to compute cohomology for general algebraic theories (groups, algebras, Lie algebras, …), as well as do other things like constructing classifying bundles. Given any monad MM and an MM-algebra XX, the bar construction is an efficient machine for constructing a ‘free resolution’ of XX, to which one can then apply the machinery of derived functors to compute the cohomology of XX.

What John gave in his course was a very nice introduction, with lots of diagrams and pictures, which gave the precise conceptual details underlying the bar construction: a simplicial object whose nn-dimensional component is a free algebra MM nXM M^n X over the monad MM, with face operators which look like:


...      mMX
         -->     mX
         MmX     -->     a
... MMMX --> MMX     MX --> X
         -->     -->
         MMa     Ma

What I want to talk about here is the conceptual sense in which the bar construction is a resolution of XX, i.e., the acyclicity of the bar construction as a simplicial object. I also want to talk about the sense in which the bar construction on XX is a universal MM-free resolution of XX.

In order for this note to make sense, I’ll first have to recall some of what John said. A monad MM is a monoid in a monoidal category, and the story begins by considering the string diagram calculus for monoids in monoidal categories. As John showed in pictures, the string diagrams, modulo the equivalence relation imposed by the monoid equations, look exactly like pictures (cospans) of order-preserving functions between finite ordinals. This was summarized in the slogan, “the category Δ\Delta is the walking monoid”, i.e., Δ\Delta equipped with its terminal object [1] is precisely the initial strict monoidal category equipped with a monoid.

Or, we could say that Δ op\Delta^{op} is the “walking comonoid”. It’s nice to know by the way that Δ op\Delta^{op} is equivalent to the category of finite intervals (i.e., finite totally ordered sets with a top and bottom, and maps which preserve the order and top and bottom). This is something you can see very well if you stare at John’s pictures of planar 2d cobordisms, which are planar thickenings of string diagrams for monoids, and then let your attention flicker over to the complement (of different shading) and read it upside-down as a 2d cobordism. This idea and the planar 2d cobordisms also figure prominently in Aaron Lauda’s Frobenius algebras and ambidextrous adjunctions.

With that, the bar construction is easy to construct. Let (M,m:MMM,u:1M)(M, m: M M \to M, u: 1 \to M) be a monad on a category EE. One has an adjunction

(F:EE M)(U:E ME)(F: E \to E^M) \dashv (U: E^M \to E)

where E ME^M is the Eilenberg-Moore category of MM-algebras. This in turn gives a comonad FUF U acting on MM-algebras, that is to say, a comonoid in a monoidal category of endofunctors. Because Δ op\Delta^{op} is the walking comonoid, there is a unique monoidal functor (the bar construction)

Bar M:Δ op[E M,E M]Bar_M: \Delta^{op} \to [E^M, E^M]

which sends the comonoid [1] in Δ op\Delta^{op} to FUF U. By postcomposing this by the evaluation functor at an MM-algebra XX, one gets a simplicial MM-algebra

B M(X):Δ opE MB_M(X): \Delta^{op} \to E^M

and this is the bar construction applied to XX. This description appears in Mac Lane’s Categories for the Working Mathematician, which is where I first learned about it (see section VII.6 in the second edition).

As I say, the important point here is that this object is acyclic, in a very strong sense of the word which I need to explain. Part of the philosophy here is that the bar construction is a piece of pure equational logic, and we will correspondingly treat its acyclicity purely algebraically, not as a property involving an existential quantifier but as a structure, called a ‘resolution’ for lack of anything better. Roughly speaking, a resolution structure on a simplicial object YY with augmentation,


... Y[2] --> Y[1] --> Y[0],
         -->

is a contracting homotopy which realizes a homotopy equivalence between YY and the constant simplicial object at Y[0]Y[0], and moreover a contracting homotopy with some very good properties.

Recall that contracting homotopy hh on a simplicial object Y:Δ opEY: \Delta^{op} \to E is given by a collection of maps

h n:Y[n]Y[n+1]h_n: Y[n] \to Y[n+1]

satisfying some equations. Observe also that the map [n][n+1][n] \mapsto [n+1] is the object part of a comonad [1]+():Δ opΔ op[1] + (-): \Delta^{op} \to \Delta^{op}. Here is the key notion:

  • A resolution is a simplicial object YY together with a right-sided coalgebra structure h:YY([1]+())h: Y \to Y \circ ([1] + (-)) over the comonad [1]+()[1] + (-).

This will require some unpacking, but let’s first see how this works in the case of the bar resolution. Since Bar MBar_M preserves the monoidal and comonoid structures, we have a commutative diagram


Delta^{op} -----> [E^M, E^M]
   |       Bar_M      |
   |                  |
   | [1]+( )          | FUo( )
   V                  V
Delta^{op} -----> [E^M, E^M]
           Bar_M

and we augment this diagram with another:


Delta^{op} -----> [E^M, E^M] ---> [E^M, E]
   |       Bar_M       |    U o( )   |
   |                   |             |
   |[1]+( )      FUo( )|       UFo( )| 
   V                   V             V
Delta^{op} -----> [E^M, E^M] ---->[E^M, E]
           Bar_M            U o( )

The bar resolution is the horizontal composite UBar MUBar_M. Notice there is is a canonical right FUF U-coalgebra structure on UU, given by the coaction uU:UUFU=MUu U: U \to U F U = M U. This gives a right action of the comonad [1]+()[1]+ (-) on UBar MUBar_M, as we see from the diagram, therefore we obtain a resolution structure with components

h n=uM n:M nM n+1.h_n = u M^n: M^n \to M^{n+1}.

Now, what exactly is a resolution structure? I’d like to tease that out.

A first trivial observation is that a right-sided coalgebra for the comonad [1]+()[1]+(-) is the same as an ordinary (left-sided) coalgebra for the comonad on simplicial objects

E [1]+():E Δ opE Δ opE^{[1]+(-)}: E^{\Delta^{op}} \to E^{\Delta^{op}}

given by pulling back along [1]+()[1] + (-). I’ll call this pullback comonad PP, for short.

Second, to make things easier (or perhaps more recognizable), let’s pretend for a moment that E=SetE = Set. The pullback comonad PP then has a left adjoint, given by left Kan extension along [1]+()[1]+ ( ). What is that left Kan extension? On the representable objects, it takes the affine simplex hom(,[n])hom(-, [n]) to hom(,[n+1])hom(-, [n+1]), i.e, it takes the cone of the affine simplex. This cone construction extends to all simplicial sets (by left Kan extension). I’d like to call the Kan extension the ‘cone functor’ on simplicial sets, although we should be careful here: we are not coning a space or simplicial set to a point, because that operation does not preserve disjoint sums (therefore isn’t a left adjoint). What the left Kan extension actually does is cone a space to its set of path components, i.e., it takes the disjoint union of all the cones over all the path components. But as I say, I’m still going to call this left Kan extension the ‘cone functor’, and denote it CC.

Now, we are in a general situation where we have a comonad PP with a left adjoint CC. There is a general piece of abstract nonsense here, dating back to the original 1965 paper by Eilenberg and Moore, which says that the comonad structure on PP corresponds precisely to a monad structure on CC, in such a way that the category of PP-coalgebras is naturally equivalent to the category of CC-algebras. So, now we want to know: what is a CC-algebra?

Personally, I find it more intuitive to think about it topologically. There is an analogous cone monad C:TopTopC: Top \to Top acting on topological spaces, where CXC X is defined to be the pushout of the diagram

[0,1]×Xinj 0Xqπ 0(X)[0, 1] \times X \stackrel{inj_0}{\leftarrow} X \stackrel{q}{\to} \pi_0(X)

where inj 0(x):=(0,x)inj_0(x) := (0, x) and qq is the canonical projection. A map a:CXXa: C X \to X is thus given by a pair of maps

h:[0,1]×XXh: [0, 1] \times X \to X i:π 0(X)Xi: \pi_0(X) \to X

where ii picks out a basepoint for each path component, and (by the pushout commutativity) hh contracts each path component to its basepoint. This already means the pair (h,i)(h, i) is explicit witness to the fact that XX is acyclic (is homotopy equivalent to its discrete space of path components), but there’s more since we have not yet accounted for the monad structure on CC.

The monad structure on CC is induced from a monoid structure on [0,1][0, 1], and we will take the monoid multiplication on [0,1][0, 1] to be the ‘tropical’ multiplication (x,y)min(x,y)(x, y) \mapsto min(x, y). An algebra with respect to the monad CC is given by a pair (h,i)(h, i) such that the homotopy hh behaves as a ‘flow’ with respect to tropical multiplication: the following equations are satisfied:

h(s,h(t,x))=h(min(s,t),x)h(s, h(t, x)) = h(min(s, t), x) h(1,x)=xh(1, x) = x h(0,x)=iq(x).h(0, x) = iq(x).

[A note to experts: the topos of simplicial sets classifies the theory of intervals; in particular, geometric realization Set Δ opTopSet^{\Delta^{op}} \to Top is the model which takes the generic interval to the interval [0, 1]. The generic interval carries an intrinsic monoid operation ‘min’, and geometric realization thus maps it over to the monoid ([0,1],min)([0, 1], min); this is why we took the multiplication on the unit interval to be ‘tropical’.]

The cone monad on topological spaces is again left adjoint to a comonad PP on topological spaces, where PXP X is the pullback of the diagram

X [0,1]eval 0Xid|X|X^{[0, 1]} \stackrel{eval_0}{\to} X \stackrel{id}{\leftarrow} |X|

where |X||X| is the underlying set of XX equipped with the discrete topology, included in XX by the identity function, and a CC-algebra structure is again equivalent to a PP-coalgebra structure.

  • In summary: a resolution = a PP-coalgebra = a CC-algebra = a simplicial object YY equipped with a ‘homotopy flow’ which contracts YY to the discrete space of its path components Y[0]Y[0]. In other words, a resolution is a particularly nice kind of contracting homotopy witnessing acyclicity of YY.

Now for the main theorem:

  • The bar resolution UBar MUBar_M is a universal MM-algebra resolution for the monad MM.

This may require some amplification. As above, let B M(X)B_M(X) denote the result of applying the bar resolution functor to a particular MM-algebra XX, i.e., let B M(X)B_M(X) be the composite

Δ opBar M[E M,E M]eval XE MUX.\Delta^{op} \stackrel{Bar_M}{\to} [E^M, E^M] \stackrel{eval_X}{\to} E^M \stackrel{U}{\to} X.

Let YY be any MM-algebra resolution, i.e., a simplicial MM-algebra Δ opE M\Delta^{op} \to E^M equipped with a PP-coalgebra structure on the underlying simplicial object

UY:Δ opE.U Y: \Delta^{op} \to E.

There is a forgetful functor from the category of MM-algebra resolutions to MM-algebras, which remembers only the augmentation part Y[0]Y[0].

The universal property of B M(X)B_M(X) is that given an MM-algebra resolution YY and an MM-algebra map XY[0]X \to Y[0], there is a unique extension to a map of MM-algebra resolutions, B M(X)YB_M(X) \to Y. This is the familiar ‘acyclic models theorem’:

  • Theorem (Acyclic Models): The bar resolution B M()B_M(-) is left adjoint to the forgetful functor from MM-algebra resolutions to MM-algebras.

Sketch of proof: The unique simplicial extension f:B M(X)Yf: B_M(X) \to Y of f 0:XY[0]f_0: X \to Y[0] is constructed by induction on dimension. The map f n+1:M n+1XYf_{n+1}: M^{n+1} X \to Y is the unique MM-algebra map which extends the composite

M nXf nY[n]h nY[n+1].M^n X \stackrel{f_n}{\to} Y[n] \stackrel{h_n}{\to} Y[n+1].

It is immediate that ff preserves the contracting homotopies; all that remains to be checked is that ff is a simplicial map. This is based on a simple inductive argument; details may be found in my notes.

Posted by: Todd Trimble on May 31, 2007 8:55 PM | Permalink | Reply to this

Re: Cohomology and Computation (Week 24)

Shall we promote this elaboration to a guest post?

(By the way, it seems you forgot to add the hyperlink to your notes. If you send me or John or David an email with the address, we can insert it into your comment.)

Posted by: urs on May 31, 2007 9:02 PM | Permalink | Reply to this

Re: Cohomology and Computation (Week 24)

I’d be honored – thanks very much, Urs.

I’ll send you the link in a moment.

Posted by: Todd Trimble on May 31, 2007 9:20 PM | Permalink | Reply to this

Re: Cohomology and Computation (Week 24)

I’ll send you the link in a moment.

Okay. Fixed.

Posted by: urs on May 31, 2007 9:32 PM | Permalink | Reply to this

Re: Cohomology and Computation (Week 24)

Earlier I wrote:

This already means the pair (h,i) is explicit witness to the fact that X is acyclic (is homotopy equivalent to its discrete space of path components)

which is not quite right since we don’t yet have the equation h(1,x)=xh(1, x) = x. But in the scheme of things, this is a minor error.
Posted by: Todd Trimble on May 31, 2007 9:18 PM | Permalink | Reply to this
Read the post Cohomology and Computation (Week 25)
Weblog: The n-Category Café
Excerpt: Getting monads, comonads and simplicial objects from adjoint functors.
Tracked: June 6, 2007 5:23 PM

Post a New Comment