tqft


The main thing happening in my end of the world is that it’s relocated from Europe back to North America. I’m taking up a teaching postdoc position in the Mathematics and Computer Science department at Mount Allison University starting this month. However, amidst all the preparations and moving, I was also recently in Edinburgh, Scotland for a workshop on Higher Gauge Theory and Higher Quantization, where I gave a talk called 2-Group Symmetries on Moduli Spaces in Higher Gauge Theory. That’s what I’d like to write about this time.

Edinburgh is a beautiful city, though since the workshop was held at Heriot-Watt University, whose campus is outside the city itself, I only got to see it on the Saturday after the workshop ended. However, John Huerta and I spent a while walking around, and as it turned out, climbing a lot: first the Scott Monument, from which I took this photo down Princes Street:

10262171_10202760228751728_566218701861596938_n

And then up a rather large hill called Arthur’s Seat, in Holyrood Park next to the Scottish Parliament.

The workshop itself had an interesting mix of participants. Urs Schreiber gave the most mathematically sophisticated talk, and mine was also quite category-theory-minded. But there were also some fairly physics-minded talks that are interesting to me as well because they show the source of these ideas. In this first post, I’ll begin with my own, and continue with David Roberts’ talk on constructing an explicit string bundle. …

2-Group Symmetries of Moduli Spaces

My own talk, based on work with Roger Picken, boils down to a couple of observations about the notion of symmetry, and applies them to a discrete model in higher gauge theory. It’s the kind of model you might use if you wanted to do lattice gauge theory for a BF theory, or some other higher gauge theory. But the discretization is just a convenience to avoid having to deal with infinite dimensional spaces and other issues that don’t really bear on the central point.

Part of that point was described in a previous post: it has to do with finding a higher analog for the relationship between two views of symmetry: one is “global” (I found the physics-inclined part of the audience preferred “rigid”), to do with a group action on the entire space; the other is “local”, having to do with treating the points of the space as objects of a groupoid who show how points are related to each other. (Think of trying to describe the orbit structure of just the part of a group action that relates points in a little neighborhood on a manifold, say.)

In particular, we’re interested in the symmetries of the moduli space of connections (or, depending on the context, flat connections) on a space, so the symmetries are gauge transformations. Now, here already some of the physically-inclined audience objected that these symmetries should just be eliminated by taking the quotient space of the group action. This is based on the slogan that “only gauge-invariant quantities matter”. But this slogan has some caveats: in only applies to closed manifolds, for one. When there are boundaries, it isn’t true, and to describe the boundary we need something which acts as a representation of the symmetries. Urs Schreiber pointed out a well-known example: the Chern-Simons action, a functional on a certain space of connections, is not gauge-invariant. Indeed, the boundary terms that show up due to this not-invariance explain why there is a Wess-Zumino-Witt theory associated with the boundaries when the bulk is described by Chern-Simons.

Now, I’ve described a lot of the idea of this talk in the previous post linked above, but what’s new has to do with how this applies to moduli spaces that appear in higher gauge theory based on a 2-group \mathcal{G}. The points in these space are connections on a manifold M. In particular, since a 2-group is a group object in categories, the transformation groupoid (which captures global symmetries of the moduli space) will be a double category. It turns out there is another way of seeing this double category by local descriptions of the gauge transformations.

In particular, general gauge transformations in HGT are combinations of two special types, described geometrically by G-valued functions, or Lie(H)-valued 1-forms, where G is the group of objects of \mathcal{G}, and H is the group of morphisms based at 1_G. If we think of connections as functors from the fundamental 2-groupoid \Pi_2(M) into \mathcal{G}, these correspond to pseudonatural transformations between these functors. The main point is that there are also two special types of these, called “strict”, and “costrict”. The strict ones are just natural transformations, where the naturality square commutes strictly. The costrict ones, also called ICONs (for “identity component oplax natural transformations” – see the paper by Steve Lack linked from the nlab page above for an explanation of “costrictness”). They assign the identity morphism to each object, but the naturality square commutes only up to a specified 2-cell. Any pseudonatural transformation factors into a strict and costrict part.

The point is that taking these two types of transformation to be the horizontal and vertical morphisms of a double category, we get something that very naturally arises by the action of a big 2-group of symmetries on a category. We also find something which doesn’t happen in ordinary gauge theory: that only the strict gauge transformations arise from this global symmetry. The costrict ones must already be the morphisms in the category being acted on. This category plays the role of the moduli space in the normal 1-group situation. So moving to 2-groups reveals that in general we should distinguish between global/rigid symmetries of the moduli space, which are strict gauge transformations, and costrict ones, which do not arise from the global 2-group action and should be thought of as intrinsic to the moduli space.

String Bundles

David Roberts gave a rather interesting talk called “Constructing Explicit String Bundles”. There are some notes for this talk here. The point is simply to give an explicit construction of a particular 2-group bundle. There is a lot of general abstract theory about 2-bundles around, and a fair amount of work that manipulates physically-motivated descriptions of things that can presumably be modelled with 2-bundles. There has been less work on giving a mathematically rigorous description of specific, concrete 2-bundles.

This one is of interest because it’s based on the String 2-group. Details are behind that link, but roughly the classifying space of String(G) (a homotopy 2-type) is fibred over the classifying space for G (a 1-type). The exact map is determined by taking a pullback along a certain characteristic class (which is a map out of BG). Saying “the” string 2-group is a bit of a misnomer, by the way, since such a 2-group exists for every simply connected compact Lie group G. The group that’s involved here is a String(n), the string 2-group associated to Spin(n), the universal cover of the rotation group SO(n). This is the one that determines whether a given manifold can support a “string structure”. A string structure on M, therefore, is a lift of a spin structure, which determines whether one can have a spin bundle over M, hence consistently talk about a spin connection which gives parallel transport for spinor fields on M. The string structure determines if one can consistently talk about a string-bundle over M, and hence a 2-group connection giving parallel transport for strings.

In this particular example, the idea was to find, explicitly, a string bundle over Minkowski space – or its conformal compactification. In point of fact, this particular one is for $latek String(5)$, and is over 6-dimensional Minkowski space, whose compactification is M = S^5 \times S^1. This particular M is convenient because it’s possible to show abstractly that it has exactly one nontrivial class of string bundles, so exhibiting one gives a complete classification. The details of the construction are in the notes linked above. The technical details rely on the fact that we can coordinatize M nicely using the projective quaternionic plane, but conceptually it relies on the fact that S^5 \cong SU(3)/SU(2), and because of how the lifting works, this is also String(SU(3))/String(SU(2)). This quotient means there’s a string bundle String(SU(3)) \rightarrow S^5 whose fibre is String(SU(2)).

While this is only one string bundle, and not a particularly general situation, it’s nice to see that there’s a nice elegant presentation which gives such a bundle explicitly (by constructing cocycles valued in the crossed module associated to the string 2-group, which give its transition functions).

(Here endeth Part I of this discussion of the workshop in Edinburgh. Part II will talk about Urs Schreiber’s very nice introduction to Higher Geometric Quantization)

(This ends the first part of this update – the next will describe the physics-oriented talks, and the third will describe Urs Schreiber’s series on higher geometric quantization)

To continue from the previous post

Twisted Differential Cohomology

Ulrich Bunke gave a talk introducing differential cohomology theories, and Thomas Nikolaus gave one about a twisted version of such theories (unfortunately, perhaps in the wrong order). The idea here is that cohomology can give a classification of field theories, and if we don’t want the theories to be purely topological, we would need to refine this. A cohomology theory is a (contravariant) functorial way of assigning to any space X, which we take to be a manifold, a \mathbb{Z}-graded group: that is, a tower of groups of “cocycles”, one group for each n, with some coboundary maps linking them. (In some cases, the groups are also rings) For example, the group of differential forms, graded by degree.

Cohomology theories satisfy some axioms – for example, the Mayer-Vietoris sequence has to apply whenever you cut a manifold into parts. Differential cohomology relaxes one axiom, the requirement that cohomology be a homotopy invariant of X. Given a differential cohomology theory, one can impose equivalence relations on the differential cocycles to get a theory that does satisfy this axiom – so we say the finer theory is a “differential refinement” of the coarser. So, in particular, ordinary cohomology theories are classified by spectra (this is related to the Brown representability theorem), whereas the differential ones are represented by sheaves of spectra – where the constant sheaves represent the cohomology theories which happen to be homotopy invariants.

The “twisting” part of this story can be applied to either an ordinary cohomology theory, or a differential refinement of one (though this needs similarly refined “twisting” data). The idea is that, if R is a cohomology theory, it can be “twisted” over X by a map \tau: X \rightarrow Pic_R into the “Picard group” of R. This is the group of invertible R-modules (where an R-module means a module for the cohomology ring assigned to X) – essentially, tensoring with these modules is what defines the “twisting” of a cohomology element.

An example of all this is twisted differential K-theory. Here the groups are of isomorphism classes of certain vector bundles over X, and the twisting is particularly simple (the Picard group in the topological case is just \mathbb{Z}_2). The main result is that, while topological twists are classified by appropriate gerbes on X (for K-theory, U(1)-gerbes), the differential ones are classified by gerbes with connection.

Fusion Categories

Scott Morrison gave a talk about Classifying Fusion Categories, the point of which was just to collect together a bunch of results constructing particular examples. The talk opens with a quote by Rutherford: “All science is either physics or stamp collecting” – that is, either about systematizing data and finding simple principles which explain it, or about collecting lots of data. This talk was unabashed stamp-collecting, on the grounds that we just don’t have a lot of data to systematically understand yet – and for that very reason I won’t try to summarize all the results, but the slides are well worth a look-over. The point is that fusion categories are very useful in constructing TQFT’s, and there are several different constructions that begin “given a fusion category \mathcal{C}“… and yet there aren’t all that many examples, and very few large ones, known.

Scott also makes the analogy that fusion categories are “noncommutative finite groups” – which is a little confusing, since not all finite groups are commutative anyway – but the idea is that the symmetric fusion categories are exactly the representation categories of finite groups. So general fusion categories are a non-symmetric generalization of such groups. Since classifying finite groups turned out to be difficult, and involve a laundry-list of sporadic groups, it shouldn’t be too surprising that understanding fusion categories (which, for the symmetric case, include the representation categories of all these examples) should be correspondingly tricky. Since, as he points out, we don’t have very many non-symmetric examples beyond rank 12 (analogous to knowing only finite groups with at most 12 elements), it’s likely that we don’t have a very good understanding of these categories in general yet.

There were a couple of talks – one during the workshop by Sonia Natale, and one the previous week by Sebastian Burciu, whom I also had the chance to talk with that week – about “Equivariantization” of fusion categories, and some fairly detailed descriptions of what results. The two of them have a paper on this which gives more details, which I won’t summarize – but I will say a bit about the construction.

An “equivariantization” of a category C acted on by a group G is supposed to be a generalization of the notion of the set of fixed points for a group acting on a set.  The category C^G has objects which consist of an object x \in C which is fixed by the action of G, together with an isomorphism \mu_g : x \rightarrow x for each g \in G, satisfying a bunch of unsurprising conditions like being compatible with the group operation. The morphisms are maps in C between the objects, which form commuting squares for each g \in G. Their paper, and the talks, described how this works when C is a fusion category – namely, C^G is also a fusion category, and one can work out its fusion rules (i.e. monoidal structure). In some cases, it’s a “group theoretical” fusion category (it looks like Rep(H) for some group H) – or a weakened version of such a thing (it’s Morita equivalent to ).

A nice special case of this is if the group action happens to be trivial, so that every object of C is a fixed point. In this case, C^G is just the category of objects of C equipped with a G-action, and the intertwining maps between these. For example, if C = Vect, then C^G = Rep(G) (in particular, a “group-theoretical fusion category”). What’s more, this construction is functorial in G itself: given a subgroup H \subset G, we get an adjoint pair of functors between C^G and C^H, which in our special case are just the induced-representation and restricted-representation functors for that subgroup inclusion. That is, we have a Mackey functor here. These generalize, however, to any fusion category C, and to nontrivial actions of G on C. The point of their paper, then, is to give a good characterization of the categories that come out of these constructions.

Quantizing with Higher Categories

The last talk I’d like to describe was by Urs Schreiber, called Linear Homotopy Type Theory for Quantization. Urs has been giving evolving talks on this topic for some time, and it’s quite a big subject (see the long version of the notes above if there’s any doubt). However, I always try to get a handle on these talks, because it seems to be describing the most general framework that fits the general approach I use in my own work. This particular one borrows a lot from the language of logic (the “linear” in the title alludes to linear logic).

Basically, Urs’ motivation is to describe a good mathematical setting in which to construct field theories using ingredients familiar to the physics approach to “field theory”, namely… fields. (See the description of Kevin Walker’s talk.) Also, Lagrangian functionals – that is, the notion of a physical action. Constructing TQFT from modular tensor categories, for instance, is great, but the fields and the action seem to be hiding in this picture. There are many conceptual problems with field theories – like the mathematical meaning of path integrals, for instance. Part of the approach here is to find a good setting in which to locate the moduli spaces of fields (and the spaces in which path integrals are done). Then, one has to come up with a notion of quantization that makes sense in that context.

The first claim is that the category of such spaces should form a differentially cohesive infinity-topos which we’ll call \mathbb{H}. The “infinity” part means we allow morphisms between field configurations of all orders (2-morphisms, 3-morphisms, etc.). The “topos” part means that all sorts of reasonable constructions can be done – for example, pullbacks. The “differentially cohesive” part captures the sort of structure that ensures we can really treat these as spaces of the suitable kind: “cohesive” means that we have a notion of connected components around (it’s implemented by having a bunch of adjoint functors between spaces and points). The “differential” part is meant to allow for the sort of structures discussed above under “differential cohomology” – really, that we can capture geometric structure, as in gauge theories, and not just topological structure.

In this case, we take \mathbb{H} to have objects which are spectral-valued infinity-stacks on manifolds. This may be unfamiliar, but the main point is that it’s a kind of generalization of a space. Now, the sort of situation where quantization makes sense is: we have a space (i.e. \mathbb{H}-object) of field configurations to start, then a space of paths (this is WHERE “path-integrals” are defined), and a space of field configurations in the final system where we observe the result. There are maps from the space of paths to identify starting and ending points. That is, we have a span:

A \leftarrow X \rightarrow B

Now, in fact, these may all lie over some manifold, such as B^n(U(1)), the classifying space for U(1) (n-1)-gerbes. That is, we don’t just have these “spaces”, but these spaces equipped with one of those pieces of cohomological twisting data discussed up above. That enters the quantization like an action (it’s WHAT you integrate in a path integral).

Aside: To continue the parallel, quantization is playing the role of a cohomology theory, and the action is the twist. I really need to come back and complete an old post about motives, because there’s a close analogy here. If quantization is a cohomology theory, it should come by factoring through a universal one. In the world of motives, where “space” now means something like “scheme”, the target of this universal cohomology theory is a mild variation on just the category of spans I just alluded to. Then all others come from some functor out of it.

Then the issue is what quantization looks like on this sort of scenario. The Atiyah-Singer viewpoint on TQFT isn’t completely lost here: quantization should be a functor into some monoidal category. This target needs properties which allow it to capture the basic “quantum” phenomena of superposition (i.e. some additivity property), and interference (some actual linearity over \mathbb{C}). The target category Urs talked about was the category of E_{\infty}-rings. The point is that these are just algebras that live in the world of spectra, which is where our spaces already lived. The appropriate target will depend on exactly what \mathbb{H} is.

But what Urs did do was give a characterization of what the target category should be LIKE for a certain construction to work. It’s a “pull-push” construction: see the link way above on Mackey functors – restriction and induction of representations are an example . It’s what he calls a “(2-monoidal, Beck-Chevalley) Linear Homotopy-Type Theory”. Essentially, this is a list of conditions which ensure that, for the two morphisms in the span above, we have a “pull” operation for some and left and right adjoints to it (which need to be related in a nice way – the jargon here is that we must be in a Wirthmuller context), satisfying some nice relations, and that everything is functorial.

The intuition is that if we have some way of getting a “linear gadget” out of one of our configuration spaces of fields (analogous to constructing a space of functions when we do canonical quantization over, let’s say, a symplectic manifold), then we should be able to lift it (the “pull” operation) to the space of paths. Then the “push” part of the operation is where the “path integral” part comes in: many paths might contribute to the value of a function (or functor, or whatever it may be) at the end-point of those paths, because there are many ways to get from A to B, and all of them contribute in a linear way.

So, if this all seems rather abstract, that’s because the point of it is to characterize very generally what has to be available for the ideas that appear in physics notions of path-integral quantization to make sense. Many of the particulars – spectra, E_{\infty}-rings, infinity-stacks, and so on – which showed up in the example are in a sense just placeholders for anything with the right formal properties. So at the same time as it moves into seemingly very abstract terrain, this approach is also supposed to get out of the toy-model realm of TQFT, and really address the trouble in rigorously defining what’s meant by some of the standard practice of physics in field theory by analyzing the logical structure of what this practice is really saying. If it turns out to involve some unexpected math – well, given the underlying issues, it would have been more surprising if it didn’t.

It’s not clear to me how far along this road this program gets us, as far as dealing with questions an actual physicist would like to ask (for the most part, if the standard practice works as an algorithm to produce results, physicists seldom need to ask what it means in rigorous math language), but it does seem like an interesting question.

So I spent a few weeks at the Erwin Schrodinger Institute in Vienna, doing a short residence as part of the program “Modern Trends in Topological Quantum Field Theory” leading up to a workshop this week. There were quite a few interesting talks – some on topics that I’ve written about elsewhere in this blog, so I’ll gloss over those. For example, Catherine Meusburger spoke about the project with Barrett and Schaumann to give a diagrammatic language for Gray categories with duals – I’ve written about John Barrett’s talks on this elsewhere. Similarly, I’ve written about Chris Schommer-Pries’ talks about fully-extended TQFT’s and the cobordism hypothesis for structured cobordisms . I’d like to just describe some of the other highlights that connect nicely to themes I find interesting. In Part 1 of this post, the more topological themes…

TQFTs with Boundary

On the first day, Kevin Walker gave a talk called “Premodular TQFTs” which was quite interesting. The key idea here is that a fairly big class of different constructions of 3D TQFT’s turn out to actually be aspects of one 4D TQFT, which comes about by a construction based on the 3D construction of Crane-Yetter-Kauffman.  The term “premodular” refers to the fact that 3D TQFT’s can be related to modular tensor categories. “Tensor” includes several concepts, like being abelian, having vector spaces of morphisms, a monoidal structure that gets along with these – typical examples being the categories of vector spaces, or of representations of some fixed group. “Modular” means that there is a braiding, and that a certain string diagram (which looks like two linked rings) built using the braiding can be represented as an invertible matrix. These will show up as a special case of the “premodular” theory.

The basic idea is to use an approach that is based on local fields (which respects the physics-land concept of what “field theory” means), avoids the path integral approach (which is hard to make rigorous), and can be shown to connect back to the Atyiah-Singer approach in which a TQFT is a kind of functor out of a cobordism category.

That is, given a manifold X we must be able to find the fields on X, called F(X). For example, F(X) could be the maps into a classifying space BG, for a gauge theory, or a category of diagrams on X with labels in some appropriate sort of category. Then one has some relations which say when given fields are the same. For each manifold Y, this defines a vector space of linear combinations of fields, modulo relations, called A(Y;c), where c \in F(\partial Y). The dual space of A(Y;c) is called Z(Y;c) – in keeping with the principle that quantum states are functionals that we can evaluate on “classical” fields.

Walker’s talk develops, from this starting point, a view that includes a whole range of theories – the Dijkgraaf-Witten model (fields are maps to BG); diagrams in a semisimple 1-category (“Euler characteristic theory”), in a pivotal 2-category (a Turaev-Viro model), or a premodular 3-category (a “Crane-Yetter model”), among others. In particular, some familiar theories appear as living on 3D boundaries to a 4D manifold, where such a  premodular theory is defined. The talk goes on to describe a kind of “theory with defects”, where two different theories live on different parts of a manifold (this is a common theme to a number of the talks), and in particular it describes a bimodule which gives a Morita equivalence between two sorts of theory – one based on graphs labelled in representations of a group G, and the other based on G-connections. The bimodule is, effectively, a kind of “Fourier transform” which relates dimension-k structures on one side to codimension-k structures on the other: a line labelled by a G-representation on one side gets acted upon by G-holonomies for a hypersurface on the other side.

On a related note Alessandro Valentino gave a talk called “Boundary Conditions for 3d TQFT and module categories” This related to a couple of papers with Jurgen Fuchs and Christoph Schweigert. The basic idea starts with the fact that one can build (3,2,1)-dimensional TQFT’s from modular tensor categories \mathcal{C}, getting a Reshitikhin-Turaev type theory which assigns \mathcal{C} to the circle. The modular tensor structure tells you what gets assigned to higher-dimensional cobordisms. (This is a higher-categorical analog of the fact that a (2,1)-dimensional TQFT is determined by a Frobenius algebra). Then the motivating question is: how can we extend this theory all the way down to a point (i.e. have it assign something to a point, so that \mathcal{C} is somehow composed of naturally occurring morphisms).

So the question is: if we know what \mathcal{C} is, what does that tell us about the “colours” that could be assigned to a boundary. There’s a fairly elegant way to take on this question by looking at what’s assigned to Wilson lines, the observables that matter in defining RT-type theories, when the line where we’re observing gets pushed onto the boundary. (See around p14 of the first paper linked above). The colours on lines inside the manifold could be objects of \mathcal{C}, and fusing them illustrates the monoidal structure of \mathcal{C}. Then the question is what kind of category can be attached to a boundary and be consistent with this.This should be functorial with respect to fusing two lines (i.e. doing this before or after projecting to the boundary should be the same).

They don’t completely characterize the situation, but they give some reasonable arguments which suggest that the result is that the boundary category, a braided monoidal category, ought to be the Drinfel’d centre of something. This is actually a stronger constraint for categories than groups (any commutative group is the centre of something – namely itself – but this isn’t true for monoidal categories).

2-Knots

Joost Slingerland gave a talk called “Local Representations of the Loop Braid Group”, which was quite nice. The Loop Braid Group was introduced by the late Xiao-Song Lin (whom I had the pleasure to know at UCR) as an interesting generalization of the braid group B_n. B_n is the “motion group” of isomorphism classes of motions of n particles in a plane: in such a motion, we let the particles move around arbitrarily, before ending up occupying the same points occupied initially. (In the “pure braid group”, each individual point must end up where it started – in the braid group, they can swap places). Up to diffeomorphism, this keeps track of how they move around each other – not just how they exchange places, but which one crosses in front of which, etc. The loop braid group does the same for loops embedded in 3D space. Now, if the loops always stay far away from each other, one possibility is that a motion amounts to a permutation in which the loops switch places: two paths through 3D space (or 4D spacetime) can always be untangled. On the other hand, loops can pass THROUGH each other, as seen at the beginning of this video:

This is analogous to two points braiding in 2D space (i.e. strands twisting around each other in 3D spacetime), although in fact these “slide moves” form a group which is different from just the pure braid group – but PB_n fits inside them. In particular, the slide moves satisfy some of the same relations as the braid group – the Yang-Baxter equations.

The final thing that can happen is that loops might move, “flip over”, and return to their original position with reversed orientation. So the loop braid group can be broken down as LB_n = Slide_n \rtimes (\mathbb{Z}_2)^n \rtimes S_n. Every loop braid could be “closed up” to a 4D knotted surface, though not every knotted surface would be of this form. For one thing, our loops have a trivial embedding in 3D space here – to get every possible knotted surface, we’d need to have knots and links sliding around, braiding through each other, merging and splitting, etc. Knotted surfaces are much more complex than knotted circles, just as the topology of embedded circles is more complex than that of embedded points.

The talk described some work on the “local representations” of LB_n: representations on spaces where each loop is attached some k-dimensional vector space V (this is the “local dimension”), so that the motions of n loops gets represented on V^{\otimes n} (a tensor product of n copies of V). This is already rather complex, but is much easier than looking for arbitrary representations of LB_n on any old vector space (“nonlocal” representations, if you like). Now, in particular, for local dimension 2, this boils down to some simple matrices which can be worked out – the slide moves are either represented by some permutation matrices, or some tensor products of rotation matrices, or a few other cases which can all be classified.

Toward the end, Dror Bar-Natan also gave a talk that touched on knotted surfaces, called “A Partial Reduction of BF Theory to Combinatorics“. The mention of BF theory – a kind of higher gauge theory that can be described locally in terms of a 1-form and a 2-form on a manifold – is basically to set up some discussion of knotted surfaces (the combinatorics it reduces to). The point is that, like many field theories, BF theory amplitudes can be calculated using a sum over certain Feynman diagrams – but these ones are diagrams that lie partly in certain knotted surfaces. (See the rather remarkable handout in the link above for lots of pictures). This is sort of analogous to how some gauge theories in 3D boil down to knot invariants – for knots that live on the boundary of a region cut out of the 3-manifold. This is similar, for a knotted surface in a 4-manifold.

The “combinatorics” boils down to showing some diagram presentations of these knotted surfaces – particularly, a special type called a “ribbon knot”, which is a certain kind of knotted sphere. The combinatorics show that these special knotted surfaces all correspond to ordinary knotted circles in 3D (in the handout, you’ll see the Gauss diagram for a knot – a picture which shows which points along a line cross over or under each other in a presentation of the knot – used to construct a corresponding ribbon knot). But do check out the handout for some pictures which show several different ways of presenting 2-knots.

(…To be continued in Part 2…)

Well, it’s been a while, but it’s now a new semester here in Hamburg, and I wanted to go back and look at some of what we talked about in last semester’s research seminar. This semester, Susama Agarwala and I are sharing the teaching in a topics class on “Category Theory for Geometry“, in which I’ll be talking about categories of sheaves, and building up the technology for Susama to talk about Voevodsky’s theory of motives (enough to give a starting point to read something like this).

As for last semester’s seminar, one of the two main threads, the one which Alessandro Valentino and I helped to organize, was a look at some of the material needed to approach Jacob Lurie’s paper on the classification of topological quantum field theories. The idea was for the research seminar to present the basic tools that are used in that paper to a larger audience, mostly of graduate students – enough to give a fairly precise statement, and develop the tools needed to follow the proof. (By the way, for a nice and lengthier discussion by Chris Schommer-Pries about this subject, which includes more details on much of what’s in this post, check out this video.)

So: the key result is a slightly generalized form of the Cobordism Hypothesis.

Cobordism Hypothesis

The sort of theory which the paper classifies are those which “extend down to a point”. So what does this mean? A topological field theory can be seen as a sort of “quantum field theory up to homotopy”, which abstract away any geometric information about the underlying space where the fields live – their local degrees of freedom.  We do this by looking only at the classes of fields up to the diffeomorphism symmetries of the space.  The local, geometric, information gets thrown away by taking this quotient of the space of solutions.

In spite of reducing the space of fields this way, we want to capture the intuition that the theory is still somehow “local”, in that we can cut up spaces into parts and make sense of the theory on those parts separately, and determine what it does on a larger space by gluing pieces together, rather than somehow having to take account of the entire space at once, indissolubly. This reasoning should apply to the highest-dimensional space, but also to boundaries, and to any figures we draw on boundaries when cutting them up in turn.

Carrying this on to the logical end point, this means that a topological quantum field theory in the fully extended sense should assign some sort of data to every geometric entity from a zero-dimensional point up to an n-dimensional cobordism.  This is all expressed by saying it’s an n-functor:

Z : Bord^{fr}_n(n) \rightarrow nAlg.

Well, once we know what this means, we’ll know (in principle) what a TQFT is.  It’s less important, for the purposes of Lurie’s paper, what nAlg is than what Bord^){fr}_n(n) is.  The reason is that we want to classify these field theories (i.e. functors).  It will turn out that Bord_n(n) has the sort of structure that makes it easy to classify the functors out of it into any target n-category \mathcal{C}.  A guess about what kind of structure is actually there was expressed by Baez and Dolan as the Cobordism Hypothesis.  It’s been slightly rephrased from the original form to get a form which has a proof.  The version Lurie proves says:

The (\infty,n)-category Bord^{fr}_n(n) is equivalent to the free symmetric monoidal (\infty,n)-category generated by one fully-dualizable object.

The basic point is that, since Bord^{fr}_n(n) is a free structure, the classification means that the extended TQFT’s amount precisely to the choice of a fully-dualizable object of \mathcal{C} (which includes a choice of a bunch of morphisms exhibiting the “dualizability”). However, to make sense of this, we need to have a suitable idea of an (\infty,n)-category, and know what a fully dualizable object is. Let’s begin with the first.

(\infty,n)-Categories

In one sense, the Cobordism Hypothesis, which was originally made about n-categories at a time when these were only beginning to be defined, could be taken as a criterion for an acceptable definition. That is, it expressed an intuition which was important enough that any definition which wouldn’t allow one to prove the Cobordism Hypothesis in some form ought to be rejected. To really make it work, one had to bring in the “infinity” part of (\infty,n)-categories. The point here is that we are talking about category-like structures which have morphisms between objects, 2-morphisms between morphisms, and so on, with j-morphisms between j-1-morphisms for every possible degree. The inspiration for this comes from homotopy theory, where one has maps, homotopies of maps, homotopies of homotopies, etc.

Nowadays, there are several possible concrete models for (\infty,n)-categories (see this survey article by Julie Bergner for a summary of four of them). They are all equivalent definitions, in a suitable up-to-homotopy way, but for purposes of the proof, Lurie is taking the definition that an (\infty,n)-category is an n-fold complete Segal space. One theme that shows up in all the definitions is that of simplicial methods. (In our seminar, we started with a series of two talks introducing the notions of simplicial sets, simplicial objects in a category, and Kan complexes. If you don’t already know this, essentially everything we need is nicely explained in here.)

One of the underlying ideas is that a category C can be associated with a simplicial set, its nerve N(C)_{\bullet}, where the set N(C)_k of k-dimensional simplexes is just the set of composable k-tuples of morphisms in C. If C is a groupoid (everything is invertible), then the simplicial set is a Kan complex – it satisfies some filling conditions, which ensure that any morphism has an inverse. Not every Kan complex is the nerve of a groupoid, but one can think of them as weak versions of groupoids – \infty-groupoids, or (\infty,0)-categories – where the higher morphisms may not be completely trivial (as with a groupoid), but where at least they’re all invertible. This leads to another desirable feature in any definition of (\infty,n)-category, which is the Homotopy Hypothesis: that the (\infty,1)-category of (\infty,0)-categories, also called \infty-groupoids, should be equivalent (in the same weak sense) to a category of Hausdorff spaces with some other nice properties, which we call \mathbf{Top} for short. This is true of Kan complexes.

Thus, up to homotopy, specifying an \infty-groupoid is the same as specifying a space.

The data which defines a Segal space (which was however first explicitly defined by Charlez Rezk) is a simplicial space X_{\bullet}: for each n, there are spaces X_n, thought of as the space of composable n-tuples of morphisms. To keep things tame, we suppose that X_0, the space of objects, is discrete – that is, we have only a set of objects. Being a simplicial space means that the X_n come equipped with a collection of face maps d_i : X_n \rightarrow X_{n-1}, which we should think of as compositions: to get from an n-tuple to an (n-1)-tuple of morphisms, one can compose two morphisms together at any of (n-1) positions in the tuple.

One condition which a simplicial space has to satisfy to be a Segal space has to do with the “weakening” which makes a Segal space a weaker notion than just a category lies in the fact that the X_n cannot be arbitrary, but must be homotopy equivalent to the “actual” space of n-tuples, which is a strict pullback X_1 \times_{X_0} \dots \times_{X_0} X_1. That is, in a Segal space, the pullback which defines these tuples for a category is weakened to be a homotopy pullback. Combining this with the various face maps, we therefore get a weakened notion of composition: X_1 \times_{X_0} \dots \times_{X_0} X_1 \cong X_n \rightarrow X_1. Because we start by replacing the space of n-tuples with the homotopy-equivalent X_n, the composition rule will only satisfy all the relations which define composition (associativity, for instance) up to homotopy.

To be complete, the Segal space must have a notion of equivalence for X_{\bullet} which agrees with that for Kan complexes seen as \infty-groupoids. In particular, there is a sub-simplicial object Core(X_{\bullet}), which we understand to consist of the spaces of invertible k-morphisms. Since there should be nothing interesting happening above the top dimension, we ask that, for these spaces, the face and degeneracy maps are all homotopy equivalences: up to homotopy, the space of invertible higher morphisms has no new information.

Then, an n-fold complete Segal space is defined recursively, just as one might define n-categories (without the infinitely many layers of invertible morphisms “at the top”). In that case, we might say that a double category is just a category internal to \mathbf{Cat}: it has a category of objects, and a category of morphims, and the various maps and operations, such as composition, which make up the definition of a category are all defined as functors. That turns out to be the same as a structure with objects, horizontal and vertical morphisms, and square-shaped 2-cells. If we insist that the category of objects is discrete (i.e. really just a set, with no interesting morphisms), then the result amounts to a 2-category. Then we can define a 3-category to be a category internal to \mathbf{2Cat} (whose 2-category of objects is discrete), and so on. This approach really defines an n-fold category (see e.g. Chapter 5 of Cheng and Lauda to see a variation of this approach, due to Tamsamani and Simpson), but imposing the condition that the objects really amount to a set at each step gives exactly the usual intuition of a (strict!) n-category.

This is exactly the approach we take with n-fold complete Segal spaces, except that some degree of weakness is automatic. Since a C.S.S. is a simplicial object with some properties (we separately define objects of k-tuples of morphisms for every k, and all the various composition operations), the same recursive approach leads to a definition of an “n-fold complete Segal space” as simply a simplicial object in (n-1)-fold C.S.S.’s (with the same properties), such that the objects form a set. In principle, this gives a big class of “spaces of morphisms” one needs to define – one for every n-fold product of simplexes of any dimension – but all those requirements that any space of objects “is just a set” (i.e. is homotopy-equivalent to a discrete set of points) simplifies things a bit.

Cobordism Category as (\infty,n)-Category

So how should we think of cobordisms as forming an (\infty,n)-category? There are a few stages in making a precise definition, but the basic idea is simple enough. One starts with manifolds and cobordisms embedded in some fixed finite-dimensional vector space V \times \mathbb{R}^n, and then takes a limit over all V. In each V \times \mathbb{R}^n, the coordinates of the \mathbb{R}^n factor give n ways of cutting the cobordism into pieces, and gluing them back together defines composition in a different direction. Now, this won’t actually produce a complete Segal space: one has to take a certain kind of completion. But the idea is intuitive enough.

We want to define an n-fold C.S.S. of cobordisms (and cobordisms between cobordisms, and so on, up to n-morphisms). To start with, think of the case n=1: then the space of objects of Bord^{fr}_1(1) consists of all embeddings of a (d-1)-dimensional manifold into V. The space of k-simplexes (of k-tuples of morphisms) consists of all ways of cutting up a d-dimensional cobordism embedded in V \times \mathbb{R} by choosing t_0, \dots , t_{k-2}, where we think of the cobordism having been glued from two pieces, where at the slice V \times {t_i}, we have the object where the two pieces were composed. (One has to be careful to specify that the Morse function on the cobordisms, got by projection only \mathbb{R}, has its critical points away from the t_i – the generic case – to make sure that the objects where gluing happens are actual manifolds.)

Now, what about the higher morphisms of the (\infty,1)-category? The point is that one needs to have an \infty-groupoid – that is, a space! – of morphisms between two cobordisms M and N. To make sense of this, we just take the space Diff(M,N) of diffeomorphisms – not just as a set of morphisms, but including its topology as well. The higher morphisms, therefore, can be thought of precisely as paths, homotopies, homotopies between homotopies, and so on, in these spaces. So the essential difference between the 1-category of cobordisms and the (\infty,1)-category is that in the first case, morphisms are diffeomorphism classes of cobordisms, whereas in the latter, the higher morphisms are made precisely of the space of diffeomorphisms which we quotient out by in the first case.

Now, (\infty,n)-categories, can have non-invertible morphisms between morphisms all the way up to dimension n, after which everything is invertible. An n-fold C.S.S. does this by taking the definition of a complete Segal space and copying it inside (n-1)-fold C.S.S’s: that is, one has an (n-1)-fold Complete Segal Space of k-tuples of morphisms, for each k, they form a simplicial object, and so forth.

Now, if we want to build an (\infty,n)-category Bord^{fr}_n(n) of cobordisms, the idea is the same, except that we have a simplicial object, in a category of simplicial objects, and so on. However, the way to define this is essentially similar. To specify an n-fold C.S.S., we have to specify a whole collection of spaces associated to cobordisms equipped with embeddings into V \times \mathbb{R}^n. In particular, for each tuple (k_1,\dots,k_n), we have the space of such embeddings, such that for each i = 1 \dots n one has k_i special points t_{i,j} along the i^{th} coordinate axis. These are the ways of breaking down a given cobordism into a composite of k_i +1 pieces. Again, one has to make sure that these critical points of the Morse functions defined by the projections onto these coordinate axes avoid these special t_{i,j} which define the manifolds where gluing takes place. The composition maps which make these into a simplical object are quite natural – they just come by deleting special points.

Finally, we take a limit over all V (to get around limits to embeddings due to the dimension of V). So we know (at least abstractly) what the (\infty,n)-category of cobordisms should be. The cobordism hypothesis claims it is equivalent to one defined in a free, algebraically-flavoured way, namely as the free symmetric monoidal (\infty,n)-category on a fully-dualizable object. (That object is “the point” – which, up to the kind of homotopically-flavoured equivalence that matters here, is the only object when our highest-dimensional cobordisms have dimension n).

Dualizability

So what does that mean, a “fully dualizable object”?

First, to get the idea, let’s think of the 1-dimensional example.  Instead of “(\infty,n)-category”, we would like to just think of this as a statement about a category.  Then Bord^{fr}_1(1) is the 1-category of framed bordisms. For a manifold (or cobordism, which is a manifold with boundary), a framing is a trivialization of the tangent bundle.  That is, it amounts to a choice of isomorphism at each point between the tangent space there and the corresponding \mathbb{R}^n.  So the objects of Bord^{fr}_1(1) are collections of (signed) points, and the morphisms are equivalence classes of framed 1-dimensional cobordisms.  These amount to oriented 1-manifolds with boundary, where the points (objects) on the boundary are the source and target of the cobordism.

Now we want to classify what TQFT’s live on this category.  These are functors Z : Bord^{fr}_1(1).  We have two generating objects, + and -, the two signed points.  A TQFT must assign these objects vector spaces, which we’ll call V and W.  Collections of points get assigned tensor products of all the corresponding vector spaces, since the functor is monoidal, so knowing these two vector spaces determines what Z does to all objects.

What does Z do to morphisms?  Well, some generating morphsims of interest are cups and caps: these are lines which connect a positive to a negative point, but thought of as cobordisms taking two points to the empty set, and vice versa.  That is, we have an evaluation:This statement is what is generalized to say that n-dimensional TQFT’s are classified by “fully” dualizable objects.

ev: W \otimes V \rightarrow \mathbb{C}

and a coevaluation:

coev: \mathbb{C} \rightarrow V \otimes W

Now, since cobordisms are taken up to equivalence, which in particular includes topological deformations, we get a bunch of relations which these have to satisfy.  The essential one is the “zig-zag” identity, reflecting the fact that a bent line can be straightened out, and we have the same 1-morphism in Born^{fr}_1(1).  This implies that:

(ev \otimes id) \circ (id \otimes coev) : W \rightarrow W \otimes V \otimes W \rightarrow W

is the same as the identity.  This in turn means that the evaluation and coevaluation maps define a nondegenerate pairing between V and W.  The fact that this exists means two things.  First, W is the dual of V: W \cong V*.  Second, this only makes sense if both V and its dual are finite dimensional (since the evaluation will just be the trace map, which is not even defined on the identity if V is infinite dimensional).

On the other hand, once we know, V, this determines W \cong V* up to isomorphism, as well as the evaluation and coevaluation maps.  In fact, this turns out to be enough to specify Z entirely.  The classification then is: 1-D TQFT’s are classified by finite-dimensional vector spaces V.  Crucially, what made finiteness important is the existence of the dual V* and the (co)evaluation maps which express the duality.

In an (\infty,n)-category, to say that an object is “fully dualizable” means more that the object has a dual (which, itself, implies the existence of the morphisms ev and coev). It also means that ev and coev have duals themselves – or rather, since we’re talking about morphisms, “adjoints”. This in turn implies the existence of 2-morphisms which are the unit and counit of the adjunctions (the defining properties are essentially the same as those for morphisms which define a dual). In fact, every time we get a morphism of degree less than n in this process, “fully dualizable” means that it too must have a dual (i.e. an adjoint).

This does run out eventually, though, since we only require this goes up to dimension (n-1): the n-morphisms which this forces to exist (quite a few) aren’t required to have duals. This is good, because if they were, since all the higher morphisms available are invertible, this would mean that the dual n-morphisms would actually be weak inverses (that is, their composite is isomorphic to the identity)… But that would mean that the dual (n-1)-morphisms which forced them to exist would also be weak inverses (their composite would be weakly isomorphic to the identity)… and so on! In fact, if the property of “having duals” didn’t stop, then everything would be weakly invertible: we’d actually have a (weak) \infty-groupoid!

Classifying TQFT

So finally, the point of the Cobordism Hypothesis is that a (fully extended) TQFT is a functor Z out of this nBord^{fr}_n(n) into some target (\infty,1)-category \mathcal{C}. There are various options, but whatever we pick, the functor must assign something in \mathcal{C} to the point, say Z(pt), and something to each of ev and coev, as well as all the higher morphisms which must exist. Then functoriality means that all these images have to again satisfy the properties which make Z(pt) a fully dualizable object. Furthermore, since nBord^{fr}_n(n) is the free gadget with all these properties on the single object pt, this is exactly what it means that Z is a functor. Saying that Z(pt) is fully dualizable, by implication, includes all the choices of morphisms like Z(ev) etc. which show it as fully dualizable. (Conceivably one could make the same object fully dualizable in more than one way – these would be different functors).

So an extended n-dimensional TQFT is exactly the choice of a fully dualizable object Z(pt) \in \mathcal{C}, for some (\infty,n)-category \mathcal{C}. This object is “what the TQFT assigns to a point”, but if we understand the structure of the object as a fully dualizable object, then we know what the TQFT assigns to any other manifold of any dimension up to n, the highest dimension in the theory. This is how this algebraic characterization of cobordisms helps to classify such theories.

Well, as promised in the previous post, I’d like to give a summary of some of what was discussed at the conference I attended (quite a while ago now, late last year) in Erlangen, Germany.  I was there also to visit Derek Wise, talking about a project we’ve been working on for some time.

(I’ve also significantly revised this paper about Extended TQFT since then, and it now includes some stuff which was the basis of my talk at Erlangen on cohomological twisting of the category Span(Gpd).  I’ll get to that in the next post.  Also coming up, I’ll be describing some new things I’ve given some talks about recently which relate the Baez-Dolan groupoidification program to Khovanov-Lauda categorification of algebras – at least in one example, hopefully in a way which will generalize nicely.)

In the meantime, there were a few themes at the conference which bear on the Extended TQFT project in various ways, so in this post I’ll describe some of them.  (This isn’t an exhaustive description of all the talks: just of a selection of illustrative ones.)


Categories with Structures

A few talks were mainly about facts regarding the sorts of categories which get used in field theory contexts.  One important type, for instance, are fusion categories is a monoidal category which is enriched in vector spaces, generated by simple objects, and some other properties: essentially, monoidal 2-vector spaces.  The basic example would be categories of representations (of groups, quantum groups, algebras, etc.), but fusion categories are an abstraction of (some of) their properties.  Many of the standard properties are described and proved in this paper by Etingof, Nikshych, and Ostrik, which also poses one of the basic conjectures, the “ENO Conjecture”, which was referred to repeatedly in various talks.  This is the guess that every fusion category can be given a “pivotal” structure: an isomorphism from Id to **.  It generalizes the theorem that there’s always such an isomorphism into ****.  More on this below.

Hendryk Pfeiffer talked about a combinatorial way to classify fusion categories in terms of certain graphs (see this paper here).  One way I understand this idea is to ask how much this sort of category really does generalize categories of representations, or actually comodules.  One starting point for this is the theorem that there’s a pair of functors between certain monoidal categories and weak Hopf algebras.  Specifically, the monoidal categories are (Cat \downarrow Vect)^{\otimes}, which consists of monoidal categories equipped with a forgetful functor into Vect.  Then from this one can get (via a coend), a weak Hopf algebra over the base field k(in the category WHA_k).  From a weak Hopf algebra H, one can get back such a category by taking all the modules of H.  These two processes form an adjunction: they’re not inverses, but we have maps between the two composites and the identity functors.

The new result Hendryk gave is that if we restrict our categories over Vect to be abelian, and the functors between them to be linear, faithful, and exact (that is, roughly, that we’re talking about concrete monoidal 2-vector spaces), then this adjunction is actually an equivalence: so essentially, all such categories C may as well be module categories for weak Hopf algebras.  Then he gave a characterization of these in terms of the “dimension graph” (in fact a quiver) for (C,M), where M is one of the monoidal generators of C.  The vertices of \mathcal{G} = \mathcal{G}_{(C,M)} are labelled by the irreducible representations v_i (i.e. set of generators of the category), and there’s a set of edges j \rightarrow l labelled by a basis of Hom(v_j, v_l \otimes M).  Then one can carry on and build a big graded algebra H[\mathcal{G}] whose m-graded part consists of length-m paths in \mathcal{G}.  Then the point is that the weak Hopf algebra of which C is (up to isomorphism) the module category will be a certain quotient of H[\mathcal{G}] (after imposing some natural relations in a systematic way).

The point, then, is that the sort of categories mostly used in this area can be taken to be representation categories, but in general only of these weak Hopf algebras: groups and ordinary algebras are special cases, but they show up naturally for certain kinds of field theory.

Tensor Categories and Field Theories

There were several talks about the relationship between tensor categories of various sorts and particular field theories.  The idea is that local field theories can be broken down in terms of some kind of n-category: n-dimensional regions get labelled by categories, (n-1)-D boundaries between regions, or “defects”, are labelled by functors between the categories (with the idea that this shows how two different kinds of field can couple together at the defect), and so on (I think the highest-dimension that was discussed explicitly involved 3-categories, so one has junctions between defects, and junctions between junctions, which get assigned some higher-morphism data).  Alteratively, there’s the dual picture where categories are assigned to points, functors to 1-manifolds, and so on.  (This is just PoincarĂ© duality in the case where the manifolds come with a decomposition into cells, which they often are if only for convenience).

Victor Ostrik gave a pair of talks giving an overview role tensor categories play in conformal field theory.  There’s too much material here to easily summarize, but the basics go like this: CFTs are field theories defined on cobordisms that have some conformal structure (i.e. notion of angles, but not distance), and on the algebraic side they are associated with vertex algebras (some useful discussion appears on mathoverflow, but in this context they can be understood as vector spaces equipped with exactly the algebraic operations needed to model cobordisms with some local holomorphic structure).

In particular, the irreducible representations of these VOA’s determine the “conformal blocks” of the theory, which tell us about possible correlations between observables (self-adjoint operators).  A VOA V is “rational” if the category Rep(V) is semisimple (i.e. generated as finite direct sums of these conformal blocks).  For good VOA’s, Rep(V) will be a modular tensor category (MTC), which is a fusion category with a duality, braiding, and some other strucutre (see this for more).   So describing these gives us a lot of information about what CFT’s are possible.

The full data of a rational CFT are given by a vertex algebra, and a module category M: that is, a fusion category is a sort of categorified ring, so it can act on M as an ring acts on a module.  It turns out that choosing an M is equivalent to finding a certain algebra (i.e. algebra object) \mathcal{L}, a “Lagrangian algebra” inside the centre of Rep(V).  The Drinfel’d centre Z(C) of a monoidal category C is a sort of free way to turn a monoidal category into a braided one: but concretely in this case it just looks like Rep(V) \otimes Rep(V)^{\ast}.  Knowing the isomorphism class \mathcal{L} determines a “modular invariant”.  It gets “physics” meaning from how it’s equipped with an algebra structure (which can happen in more than one way), but in any case \mathcal{L} has an underlying vector space, which becomes the Hilbert space of states for the conformal field theory, which the VOA acts on in the natural way.

Now, that was all conformal field theory.  Christopher Douglas described some work with Chris Schommer-Pries and Noah Snyder about fusion categories and structured topological field theories.  These are functors out of cobordism categories, the most important of which are n-categories, where the objects are points, morphisms are 1D cobordisms, and so on up to n-morphisms which are n-dimensional cobordisms.  To keep things under control, Chris Douglas talked about the case Bord_0^3, which is where n=3, and a “local” field theory is a 3-functor Bord_0^3 \rightarrow \mathcal{C} for some 3-category \mathcal{C}.  Now, the (Baez-Dolan) Cobordism Hypothesis, which was proved by Jacob Lurie, says that Bord_0^3 is, in a suitable sense, the free symmetric monoidal 3-category with duals.  What this amounts to is that a local field theory whose target 3-category is \mathcal{C} is “just” a dualizable object of \mathcal{C}.

The handy example which links this up to the above is when \mathcal{C} has objects which are tensor categories, morphisms which are bimodule categories (i.e. categories acted), 2-morphisms which are functors, and 3-morphisms which are natural transformations.  Then the issue is to classify what kind of tensor categories these objects can be.

The story is trickier if we’re talking about, not just topological cobordisms, but ones equipped with some kind of structure regulated by a structure group G(for instance, orientation by G=SO(n), spin structure by its universal cover G= Spin(n), and so on).  This means the cobordisms come equipped with a map into BG.  They take O(n) as the starting point, and then consider groups G with a map to O(n), and require that the map into BG is a lift of the map to BO(n).  Then one gets that a structured local field theory amounts to a dualizable objects of \mathcal{C} with a homotopy-fixed point for some G-action – and this describes what gets assigned to the point by such a field theory.  What they then show is a correspondence between G and classes of categories.  For instance, fusion categories are what one gets by imposing that the cobordisms be oriented.

Liang Kong talked about “Topological Orders and Tensor Categories”, which used the Levin-Wen models, from condensed matter phyiscs.  (Benjamin Balsam also gave a nice talk describing these models and showing how they’re equivalent to the Turaev-Viro and Kitaev models in appropriate cases.  Ingo Runkel gave a related talk about topological field theories with “domain walls”.).  Here, the idea of a “defect” (and topological order) can be understood very graphically: we imagine a 2-dimensional crystal lattice (of atoms, say), and the defect is a 1-dimensional place where the two lattices join together, with the internal symmetry of each breaking down at the boundary.  (For example, a square lattice glued where the edges on one side are offset and meet the squares on the other side in the middle of a face, as you typically see in a row of bricks – the slides linked above have some pictures).  The Levin-Wen models are built using a hexagonal lattice, starting with a tensor category with several properties: spherical (there are dualities satisfying some relations), fusion, and unitary: in fact, historically, these defining properties were rediscovered independently here as the requirement for there to be excitations on the boundary which satisfy physically-inspired consistency conditions.

These abstract the properties of a category of representations.  A generalization of this to “topological orders” in 3D or higher is an extended TFT in the sense mentioned just above: they have a target 3-category of tensor categories, bimodule categories, functors and natural transformations.  The tensor categories (say, \mathcal{C}, \mathcal{D}, etc.) get assigned to the bulk regions; to “domain walls” between different regions, namely defects between lattices, we assign bimodule categories (but, for instance, to a line within a region, we get \mathcal{C} understood as a \mathcal{C}-\mathcal{C}-bimodule); then to codimension 2 and 3 defects we attach functors and natural transformations.  The algebra for how these combine expresses the ways these topological defects can go together.  On a lattice, this is an abstraction of a spin network model, where typically we have just one tensor category \mathcal{C} applied to the whole bulk, namely the representations of a Lie group (say, a unitary group).  Then we do calculations by breaking down into bases: on codimension-1 faces, these are simple objects of \mathcal{C}; to vertices we assign a Hom space (and label by a basis for intertwiners in the special case); and so on.

Thomas Nickolaus spoke about the same kind of G-equivariant Dijkgraaf-Witten models as at our workshop in Lisbon, so I’ll refer you back to my earlier post on that.  However, speaking of equivariance and group actions:

Michael MĂĽger  spoke about “Orbifolds of Rational CFT’s and Braided Crossed G-Categories” (see this paper for details).  This starts with that correspondence between rational CFT’s (strictly, rational chiral CFT’s) and modular categories Rep(F).  (He takes F to be the name of the CFT).  Then we consider what happens if some finite group G acts on F (if we understand F as a functor, this is an action by natural transformations; if as an algebra, then ).  This produces an “orbifold theory” F^G (just like a finite group action on a manifold produces an orbifold), which is the “G-fixed subtheory” of F, by taking G-fixed points for every object, and is also a rational CFT.  But that means it corresponds to some other modular category Rep(F^G), so one would like to know what category this is.

A natural guess might be that it’s Rep(F)^G, where C^G is a “weak fixed-point” category that comes from a weak group action on a category C.  Objects of C^G are pairs (c,f_g) where c \in C and f_g : g(c) \rightarrow c is a specified isomorphism.  (This is a weak analog of S^G, the set of fixed points for a group acting on a set).  But this guess is wrong – indeed, it turns out these categories have the wrong dimension (which is defined because the modular category has a trace, which we can sum over generating objects).  Instead, the right answer, denoted by Rep(F^G) = G-Rep(F)^G, is the G-fixed part of some other category.  It’s a braided crossed G-category: one with a grading by G, and a G-action that gets along with it.  The identity-graded part of Rep(F^G) is just the original Rep(F).

State Sum Models

This ties in somewhat with at least some of the models in the previous section.  Some of these were somewhat introductory, since many of the people at the conference were coming from a different background.  So, for instance, to begin the workshop, John Barrett gave a talk about categories and quantum gravity, which started by outlining the historical background, and the development of state-sum models.  He gave a second talk where he began to relate this to diagrams in Gray-categories (something he also talked about here in Lisbon in February, which I wrote about then).  He finished up with some discussion of spherical categories (and in particular the fact that there is a Gray-category of spherical categories, with a bunch of duals in the suitable sense).  This relates back to the kind of structures Chris Douglas spoke about (described above, but chronologically right after John).  Likewise, Winston Fairbairn gave a talk about state sum models in 3D quantum gravity – the Ponzano Regge model and Turaev-Viro model being the focal point, describing how these work and how they’re constructed.  Part of the point is that one would like to see that these fit into the sort of framework described in the section above, which for PR and TV models makes sense, but for the fancier state-sum models in higher dimensions, this becomes more complicated.

Higher Gauge Theory

There wasn’t as much on this topic as at our own workshop in Lisbon (though I have more remarks on higher gauge theory in one post about it), but there were a few entries.  Roger Picken talked about some work with Joao Martins about a cubical formalism for parallel transport based on crossed modules, which consist of a group G and abelian group H, with a map \partial : H \rightarrow G and an action of G on H satisfying some axioms.  They can represent categorical groups, namely group objects in Cat (equivalently, categories internal to Grp), and are “higher” analogs of groups with a set of elements.  Roger’s talk was about how to understand holonomies and parallel transports in this context.  So, a “connection” lets on transport things with G-symmetries along paths, and with H-symmetries along surfaces.  It’s natural to describe this with squares whose edges are labelled by G-elements, and faces labelled by H-elements (which are the holonomies).  Then the “cubical approach” means that we can describe gauge transformations, and higher gauge transformations (which in one sense are the point of higher gauge theory) in just the same way: a gauge transformation which assigns H-values to edges and G-values to vertices can be drawn via the holonomies of a connection on a cube which extends the original square into 3D (so the edges become squares, and so get H-values, and so on).  The higher gauge transformations work in a similar way.  This cubical picture gives a good way to understand the algebra of how gauge transformations etc. work: so for instance, gauge transformations look like “conjugation” of a square by four other squares – namely, relating the front and back faces of a cube by means of the remaining faces.  Higher gauge transformations can be described by means of a 4D hypercube in an analogous way, and their algebraic properties have to do with the 2D faces of the hypercube.

Derek Wise gave a short talk outlining his recent paper with John Baez in which they show that it’s possible to construct a higher gauge theory based on the Poincare 2-group which turns out to have fields, and dynamics, which are equivalent to teleparallel gravity, a slightly unusal theory which nevertheless looks in practice just like General Relativity.  I discussed this in a previous post.

So next time I’ll talk about the new additions to my paper on ETQFT which were the basis of my talk, which illustrates a few of the themes above.

Now for a more sketchy bunch of summaries of some talks presented at the HGTQGR workshop.  I’ll organize this into a few themes which appeared repeatedly and which roughly line up with the topics in the title: in this post, variations on TQFT, plus 2-group and higher forms of gauge theory; in the next post, gerbes and cohomology, plus talks on discrete models of quantum gravity and suchlike physics.

TQFT and Variations

I start here for no better reason than the personal one that it lets me put my talk first, so I’m on familiar ground to start with, for which reason also I’ll probably give more details here than later on.  So: a TQFT is a linear representation of the category of cobordisms – that is, a (symmetric monoidal) functor nCob \rightarrow Vect, in the notation I mentioned in the first school post.  An Extended TQFT is a higher functor nCob_k \rightarrow k-Vect, representing a category of cobordisms with corners into a higher category of k-Vector spaces (for some definition of same).  The essential point of my talk is that there’s a universal construction that can be used to build one of these at k=2, which relies on some way of representing nCob_2 into Span(Gpd), whose objects are groupoids, and whose morphisms in Hom(A,B) are pairs of groupoid homomorphisms A \leftarrow X \rightarrow B.  The 2-morphisms have an analogous structure.  The point is that there’s a 2-functor \Lambda : Span(Gpd) \rightarrow 2Vect which is takes representations of groupoids, at the level of objects; for morphisms, there is a “pull-push” operation that just uses the restricted and induced representation functors to move a representation across a span; the non-trivial (but still universal) bit is the 2-morphism map, which uses the fact that the restriction and induction functors are bi-ajdoint, so there are units and counits to use.  A construction using gauge theory gives groupoids of connections and gauge transformations for each manifold or cobordism.  This recovers a form of the Dijkgraaf-Witten model.  In principle, though, any way of getting a groupoid (really, a stack) associated to a space functorially will give an ETQFT this way.  I finished up by suggesting what would need to be done to extend this up to higher codimension.  To go to codimension 3, one would assign an object (codimension-3 manifold) a 3-vector space which is a representation 2-category of 2-groupoids of connections valued in 2-groups, and so on.  There are some theorems about representations of n-groupoids which would need to be proved to make this work.

The fact that different constructions can give groupoids for spaces was used by the next speaker, Thomas Nicklaus, whose talk described another construction that uses the \Lambda I mentioned above.  This one produces “Equivariant Dijkgraaf-Witten Theory”.  The point is that one gets groupoids for spaces in a new way.  Before, we had, for a space M a groupoid \mathcal{A}_G(M) whose objects are G-connections (or, put another way, bundles-with-connection) and whose morphisms are gauge transformations.  Now we suppose that there’s some group J which acts weakly (i.e. an action defined up to isomorphism) on \mathcal{A}_G(M).  We think of this as describing “twisted bundles” over M.  This is described by a quotient stack \mathcal{A}_G // J (which, as a groupoid, gets some extra isomorphisms showing where two objects are related by the J-action).  So this gives a new map nCob \rightarrow Span(Gpd), and applying \Lambda gives a TQFT.  The generating objects for the resulting 2-vector space are “twisted sectors” of the equivariant DW model.  There was some more to the talk, including a description of how the DW model can be further mutated using a cocycle in the group cohomology of G, but I’ll let you look at the slides for that.

Next up was Jamie Vicary, who was talking about “(1,2,3)-TQFT”, which is another term for what I called “Extended” TQFT above, but specifying that the objects are 1-manifolds, the morphisms 2-manifolds, and the 2-morphisms are 3-manifolds.  He was talking about a theorem that identifies oriented TQFT’s of this sort with “anomaly-free modular tensor categories” – which is widely believed, but in fact harder than commonly thought.  It’s easy enough that such a TQFT Z corresponds to a MTC – it’s the category Z(S^1) assigned to the circle.  What’s harder is showing that the TQFT’s are equivalent functors iff the categories are equivalent.  This boils down, historically, to the difficulty of showing the category is rigid.  Jamie was talking about a project with Bruce Bartlett and Chris Schommer-Pries, whose presentation of the cobordism category (described in the school post) was the basis of their proof.

Part of it amounts to giving a description of the TQFT in terms of certain string diagrams.  Jamie kindly credited me with describing this point of view to him: that the codimension-2 manifolds in a TQFT can be thought of as “boundaries in space” – codimension-1 manifolds are either time-evolving boundaries, or else slices of space in which the boundaries live; top-dimension cobordisms are then time-evolving slices of space-with-boundary.  (This should be only a heuristic way of thinking – certainly a generic TQFT has no literal notion of “time-evolution”, though in that (2+1) quantum gravity can be seen as a TQFT, there’s at least one case where this picture could be taken literally.)  Then part of their proof involves showing that the cobordisms can be characterized by taking vector spaces on the source and target manifolds spanned by the generating objects, and finding the functors assigned to cobordisms in terms of sums over all “string diagrams” (particle worldlines, if you like) bounded by the evolving boundaries.  Jamie described this as a “topological path integral”.  Then one has to describe the string diagram calculus – ridigidy follows from the “yanking” rule, for instance, and this follows from Morse theory as in Chris’ presentation of the cobordism category.

There was a little more discussion about what the various properties (proved in a similar way) imply.  One is “cloaking” – the fact that a 2-morphism which “creates a handle” is invisible to the string diagrams in the sense that it introduces a sum over all diagrams with a string “looped” around the new handle, but this sum gives a result that’s equal to the original map (in any “pivotal” tensor category, as here).

Chronologically before all these, one of the first talks on such a topic was by Rafael Diaz, on Homological Quantum Field Theory, or HLQFT for short, which is a rather different sort of construction.  Remember that Homotopy QFT, as described in my summary of Tim Porter’s school sessions, is about linear representations of what I’ll for now call Cob(d,B), whose morphisms are d-dimensional cobordisms equipped with maps into a space B up to homotopy.  HLQFT instead considers cobordisms equipped with maps taken up to homology.

Specifically, there’s some space M, say a manifold, with some distinguished submanifolds (possibly boundary components; possibly just embedded submanifolds; possibly even all of M for a degenerate case).  Then we define Cob_d^M to have objects which are (d-1)-manifolds equipped with maps into M which land on the distinguished submanifolds (to make composition work nicely, we in fact assume they map to a single point).  Morphisms in Cob_d^M are trickier, and look like (N,\alpha, \xi): a cobordism N in this category is likewise equipped with a map \alpha from its boundary into M which recovers the maps on its objects.  That \xi is a homology class of maps from N to M, which agrees with \alpha.  This forms a monoidal category as with standard cobordisms.  Then HLQFT is about representations of this category.  One simple case Rafael described is the dimension-1 case, where objects are (ordered sets of) points equipped with maps that pick out chosen submanifolds of M, and morphisms are just braids equipped with homology classes of “paths” joining up the source and target submanifolds.  Then a representation might, e.g., describe how to evolve a homology class on the starting manifold to one on the target by transporting along such a path-up-to-homology.  In higher dimensions, the evolution is naturally more complicated.

A slightly looser fit to this section is the talk by Thomas Krajewski, “Quasi-Quantum Groups from Strings” (see this) – he was talking about how certain algebraic structures arise from “string worldsheets”, which are another way to describe cobordisms.  This does somewhat resemble the way an algebraic structure (Frobenius algebra) is related to a 2D TQFT, but here the string worldsheets are interacting with 3-form field, H (the curvature of that 2-form field B of string theory) and things needn’t be topological, so the result is somewhat different.

Part of the point is that quantizing such a thing gives a higher version of what happens for quantizing a moving particle in a gauge field.  In the particle case, one comes up with a line bundle (of which sections form the Hilbert space) and in the string case one comes up with a gerbe; for the particle, this involves associated 2-cocycle, and for the string a 3-cocycle; for the particle, one ends up producing a twisted group algebra, and for the string, this is where one gets a “quasi-quantum group”.  The algebraic structures, as in the TQFT situation, come from, for instance, the “pants” cobordism which gives a multiplication and a comultiplication (by giving maps H \otimes H \rightarrow H or the reverse, where H is the object assigned to a circle).

There is some machinery along the way which I won’t describe in detail, except that it involves a tricomplex of forms – the gradings being form degree, the degree of a cocycle for group cohomology, and the number of overlaps.  As observed before, gerbes and their higher versions have transition functions on higher numbers of overlapping local neighborhoods than mere bundles.  (See the paper above for more)

Higher Gauge Theory

The talks I’ll summarize here touch on various aspects of higher-categorical connections or 2-groups (though at least one I’ll put off until later).  The division between this and the section on gerbes is a little arbitrary, since of course they’re deeply connected, but I’m making some judgements about emphasis or P.O.V. here.

Apart from giving lectures in the school sessions, John Huerta also spoke on “Higher Supergroups for String Theory”, which brings “super” (i.e. \mathbb{Z}_2-graded) objects into higher gauge theory.  There are “super” versions of vector spaces and manifolds, which decompose into “even” and “odd” graded parts (a.k.a. “bosonic” and “fermionic” parts).  Thus there are “super” variants of Lie algebras and Lie groups, which are like the usual versions, except commutation properties have to take signs into account (e.g. a Lie superalgebra’s bracket is commutative if the product of the grades of two vectors is odd, anticommutative if it’s even).  Then there are Lie 2-algebras and 2-groups as well – categories internal to this setting.  The initial question has to do with whether one can integrate some Lie 2-algebra structures to Lie 2-group structures on a spacetime, which depends on the existence of some globally smooth cocycles.  The point is that when spacetime is of certain special dimensions, this can work, namely dimensions 3, 4, 6, and 10.  These are all 2 more than the real dimensions of the four real division algebras, \mathbb{R}, \mathbb{C}, \mathbb{H} and \mathbb{O}.  It’s in these dimensions that Lie 2-superalgebras can be integrated to Lie 2-supergroups.  The essential reason is that a certain cocycle condition will hold because of the properties of a form on the Clifford algebras that are associated to the division algebras.  (John has some related material here and here, though not about the 2-group case.)

Since we’re talking about higher versions of Lie groups/algebras, an important bunch of concepts to categorify are those in representation theory.  Derek Wise spoke on “2-Group Representations and Geometry”, based on work with Baez, Baratin and Freidel, most fully developed here, but summarized here.  The point is to describe the representation theory of Lie 2-groups, in particular geometrically.  They’re to be represented on (in general, infinite-dimensional) 2-vector spaces of some sort, which is chosen to be a category of measurable fields of Hilbert spaces on some measure space, which is called H^X (intended to resemble, but not exactly be the same as, Hilb^X, the space of “functors into Hilb from the space X, the way Kapranov-Voevodsky 2-vector spaces can be described as Vect^k).  The first work on this was by Crane and Sheppeard, and also Yetter.  One point is that for 2-groups, we have not only representations and intertwiners between them, but 2-intertwiners between these.  One can describe these geometrically – part of which is a choice of that measure space (X,\mu).

This done, we can say that a representation of a 2-group is a 2-functor \mathcal{G} \rightarrow H^X, where \mathcal{G} is seen as a one-object 2-category.  Thinking about this geometrically, if we concretely describe \mathcal{G} by the crossed module (G,H,\rhd,\partial), defines an action of G on X, and a map X \rightarrow H^* into the character group, which thereby becomes a G-equivariant bundle.  One consequence of this description is that it becomes possible to distinguish not only irreducible representations (bundles over a single orbit) and indecomposible ones (where the fibres are particularly simple homogeneous spaces), but an intermediate notion called “irretractible” (though it’s not clear how much this provides).  An intertwining operator between reps over X and Y can be described in terms of a bundle of Hilbert spaces – which is itself defined over the pullback of X and Y seen as G-bundles over H^*.  A 2-intertwiner is a fibre-wise map between two such things.  This geometric picture specializes in various ways for particular examples of 2-groups.  A physically interesting one, which Crane and Sheppeard, and expanded on in that paper of [BBFW] up above, deals with the PoincarĂ© 2-group, and where irreducible representations live over mass-shells in Minkowski space (or rather, the dual of H \cong \mathbb{R}^{3,1}).

Moving on from 2-group stuff, there were a few talks related to 3-groups and 3-groupoids.  There are some new complexities that enter here, because while (weak) 2-categories are all (bi)equivalent to strict 2-categories (where things like associativity and the interchange law for composing 2-cells hold exactly), this isn’t true for 3-categories.  The best strictification result is that any 3-category is (tri)equivalent to a Gray category – where all those properties hold exactly, except for the interchange law (\alpha \circ \beta) \cdot (\alpha ' \circ \beta ') = (\alpha \cdot \alpha ') \circ (\beta \circ \beta ') for horizontal and vertical compositions of 2-cells, which is replaced by an “interchanger” isomorphism with some coherence properties.  John Barrett gave an introduction to this idea and spoke about “Diagrams for Gray Categories”, describing how to represent morphisms, 2-morphisms, and 3-morphisms in terms of higher versions of “string” diagrams involving (piecewise linear) surfaces satisfying some properties.  He also carefully explained how to reduce the dimensions in order to make them both clearer and easier to draw.  Bjorn Gohla spoke on “Mapping Spaces for Gray Categories”, but since it was essentially a shorter version of a talk I’ve already posted about, I’ll leave that for now, except to point out that it linked to the talk by Joao Faria Martins, “3D Holonomy” (though see also this paper with Roger Picken).

The point in Joao’s talk starts with the fact that we can describe holonomies for 3-connections on 3-bundles valued in Gray-groups (i.e. the maximally strict form of a general 3-group) in terms of Gray-functors hol: \Pi_3(M) \rightarrow \mathcal{G}.  Here, \Pi_3(M) is the fundamental 3-groupoid of M, which turns points, paths, homotopies of paths, and homotopies of homotopies into a Gray groupoid (modulo some technicalities about “thin” or “laminated”  homotopies) and \mathcal{G} is a gauge Gray-group.  Just as a 2-group can be represented by a crossed module, a Gray (3-)group can be represented by a “2-crossed module” (yes, the level shift in the terminology is occasionally confusing).  This is a chain of groups L \stackrel{\delta}{\rightarrow} E \stackrel{\partial}{\rightarrow} G, where G acts on the other groups, together with some structure maps (for instance, the Peiffer commutator for a crossed module becomes a lifting \{ ,\} : E \times E \rightarrow L) which all fit together nicely.  Then a tri-connection can be given locally by forms valued in the Lie algebras of these groups: (\omega , m ,\theta) in  \Omega^1 (M,\mathfrak{g} ) \times \Omega^2 (M,\mathfrak{e}) \times \Omega^3(M,\mathfrak{l}).  Relating the global description in terms of hol and local description in terms of (\omega, m, \theta) is a matter of integrating forms over paths, surfaces, or 3-volumes that give the various j-morphisms of \Pi_3(M).  This sort of construction of parallel transport as functor has been developed in detail by Waldorf and Schreiber (viz. these slides, or the full paper), some time ago, which is why, thematically, they’re the next two speakers I’ll summarize.

Konrad Waldorf spoke about “Abelian Gauge Theories on Loop Spaces and their Regression”.  (For more, see two papers by Konrad on this)  The point here is that there is a relation between two kinds of theories – string theory (with B-field) on a manifold M, and ordinary U(1) gauge theory on its loop space LM.  The relation between them goes by the name “regression” (passing from gauge theory on LM to string theory on M), or “transgression”, going the other way.  This amounts to showing an equivalence of categories between [principal U(1)-bundles with connection on LM] and [U(1)-gerbes with connection on M].  This nicely gives a way of seeing how gerbes “categorify” bundles, since passing to the loop space – whose points are maps S^1 \rightarrow M means a holonomy functor is now looking at objects (points in LM) which would be morphisms in the fundamental groupoid of M, and morphisms which are paths of loops (surfaces in M which trace out homotopies).  So things are shifted by one level.  Anyway, Konrad explained how this works in more detail, and how it should be interpreted as relating connections on loop space to the B-field in string theory.

Urs Schreiber kicked the whole categorification program up a notch by talking about \infty-Connections and their Chern-Simons Functionals .  So now we’re getting up into \infty-categories, and particularly \infty-toposes (see Jacob Lurie’s paper, or even book if so inclined to find out what these are), and in particular a “cohesive topos”, where derived geometry can be developed (Urs suggested people look here, where a bunch of background is collected). The point is that \infty-topoi are good for talking about homotopy theory.  We want a setting which allows all that structure, but also allows us to do differential geometry and derived geometry.  So there’s a “cohesive” \infty-topos called Smooth\infty Gpds, of “sheaves” (in the \infty-topos sense) of \infty-groupoids on smooth manifolds.  This setting is the minimal common generalization of homotopy theory and differential geometry.

This is about a higher analog of this setup: since there’s a smooth classifying space (in fact, a Lie groupoid) for G-bundles, BG, there’s also an equivalence between categories G-Bund of G-principal bundles, and SmoothGpd(X,BG) (of functors into BG).  Moreover, there’s a similar setup with BG_{conn} for bundles with connection.  This can be described topologically, or there’s also a “differential refinement” to talk about the smooth situation.  This equivalence lives within a category of (smooth) sheaves of groupoids.  For higher gauge theory, we want a higher version as in Smooth \infty Gpds described above.  Then we should get an equivalence – in this cohesive topos – of hom(X,B^n U(1)) and a category of U(1)(n-1)-gerbes.

Then the part about the  “Chern-Simons functionals” refers to the fact that CS theory for a manifold (which is a kind of TQFT) is built using an action functional that is found as an integral of the forms that describe some U(1)-connection over the manifold.  (Then one does a path-integral of this functional over all connections to find partition functions etc.)  So the idea is that for these higher U(1)-gerbes, whose classifying spaces we’ve just described, there should be corresponding functionals.  This is why, as Urs remarked in wrapping up, this whole picture has an explicit presentation in terms of forms.  Actually, in terms of Cech-cocycles (due to the fact we’re talking about gerbes), whose coefficients are taken in sheaves of complexes (this is the derived geometry part) of differential forms whose coefficients are in L_\infty-algebroids (the \infty-groupoid version of Lie algebras, since in general we’re talking about a theory with gauge \infty-groupoids now).

Whew!  Okay, that’s enough for this post.  Next time, wrapping up blogging the workshop, finally.

I’d like to continue describing the talks that made up the HGTQGR workshop, in particular the ones that took place during the school portion of the event.  I’ll save one “school” session, by Laurent Freidel, to discuss with the talks because it seems to more nearly belong there. This leaves five people who gave between two and four lectures each over a period of a few days, all intermingled. Here’s a very rough summary in the order of first appearance:

2D Extended TQFT

Chris Schommer-Pries gave the longest series of talks, about the classification of 2D extended TQFT’s.  A TQFT is a kind of topological invariant for manifolds, which has a sort of “locality” property, in that you can decompose the manifold, compute the invariant on the parts, and find the whole by gluing the pieces back together.  This is expressed by saying it’s a monoidal functor Z : (Cob_d, \sqcup) \rightarrow (Vect, \otimes), where the “locality” property is now functoriality property that composition is preserved.  The key thing here is the cobordism category Cob_d, which has objects (d-1)-dimensional manifolds, and morphisms d-dimensional cobordisms (manifolds with boundary, where the objects are components of the boundary).  Then a closed d-manifold is just a cobordism from $latex\emptyset$ to itself.

Making this into a category is actually a bit nontrivial: gluing bits of smooth manifolds, for instance, won’t necessarily give something smooth.  There are various ways of handling this, such as giving the boundaries “collars”, but Chris’ preferred method is to give boundaries (and, ultimately, corners, etc.) a”halation”.  This word originally means the halo of light around bright things you sometimes see in photos, but in this context, a halation for X is an equivalence class of embeddings into neighborhoods U \subset \mathbb{R}^d.  The equivalence class says two such embeddings into U and V are equivalent if there’s a compatible refinement into some common W that embeds into both U and V.  The idea is that a halation is a kind of d-dimensional “halo”, or the “germ of a d-manifold” around X.  Then gluing compatibly along (d-1)-boundaries with halations ensures that we get smooth d-manifolds.  (One can also extend this setup so that everything in sight is oriented, or has some other such structure on it.)

In any case, an extended TQFT will then mean an n-functor Z : (Bord_d,\sqcup) \rightarrow (\mathcal{C},\otimes), where (\mathcal{C},\otimes) is some symmetric monoidal n-category (which is supposed to be similar to Vect).  Its exact nature is less important than that of Bord_d, which has:

  • 0-Morphisms (i.e. Objects): 0-manifolds (collections of points)
  • 1-Morphisms: 1-dimensional cobordisms between 0-manifolds (curves)
  • 2-Morphisms: 2-dim cobordisms with corners between 1-Morphisms (surfaces with boundary)
  • d-Morphisms: d-dimensional cobordisms between (d-1)-Morphisms (n-manifolds with corners), up to isomorphism

(Note: the distinction between “Bord” and “Cobord” is basically a matter of when a given terminology came in.  “Cobordism” and “Bordism”, unfortunately, mean the same thing, except that “bordism” has become popular more recently, since the “co” makes it sound like it’s the opposite category of something else.  This is kind of regrettable, but that’s what happened.  Sorry.)

The crucial point, is that Chris wanted to classify all such things, and his approach to this is to give a presentation of Bord_d.  This is based on stuff in his thesis.  The basic idea is to use Morse theory, and its higher-dimensional generalization, Cerf theory.  The idea is that one can put a Morse function  on a cobordism (essentially, a well-behaved “time order” on points) and look at its critical points.  Classifying these tells us what the generators for the category of cobordisms must be: there need to be enough to capture all the most general sorts of critical points.

Cerf theory does something similar, but one dimension up: now we’re talking about “stratified” families of Morse functions.  Again one studies critical points, but, for instance, on a 2-dim surface, there can be 1- and 0-dimensional parts of the set of cricical points.  In general, this gets into the theory of higher-dimensional singularities, catastrophe theory, and so on.  Each extra dimension one adds means looking at how the sets of critical points in the previous dimension can change over “time” (i.e. within some stratified family of Cerf functions).  Where these changes themselves go through critical points, one needs new generators for the various j-morphisms of the cobordism category.  (See some examples of such “catastrophes”, such as folds, cusps, swallowtails, etc. linked from here, say.)  Showing what such singularities can be like in the “generic” situation, and indeed, even defining “generic” in a way that makes sense in any dimension, required some discussion of jet bundles.  These are generalizations of tangent bundles that capture higher derivatives the way tangent bundles capture first-derivatives.  The essential point is that one can find a way to decompose these into a direct sum of parts of various dimensions (capturing where various higher derivatives are zero, say), and these will eventually tell us the dimension of a set of critical points for a Cerf function.

Now, this gives a characterization of what cobordisms can be like – part of the work in the theorem is to show that this is sufficient: that is, given a diagram showing the critical points for some Morse/Cerf function, one needs to be able to find the appropriate generators and piece together the cobordism (possibly a closed manifold) that it came from.  Chris showed how this works – a slightly finicky process involving cutting a diagram of the singular points (with some extra labelling information) into parts, and using a graphical calculus to work out how pasting works – and showed an example reconstruction of a surface this way.  This amounts to a construction of an equivalence between an “abstract” cobordism category given in terms of generators (and relations) which come from Cerf theory, and the concrete one.  The theorem then says that there’s a correspondence between equivalence classes of 2D cobordisms, and certain planar diagrams, up to some local moves.  To show this properly required a digression through some theory of symmetric monoidal bicategories, and what the right notion of equivalence for them is.

This all done, the point is that Bord_d has a characterization in terms of a universal property, and so any ETQFT Z : Bord_d \rightarrow \mathcal{C} amounts to a certain kind of object in \mathcal{C} (corresponding to the image of the point – the generating object in Bord_d).  For instance, in the oriented situation this object needs to be “fully dualizable”: it should have a dual (the point with opposite orientation), and a whole bunch of maps that specify the duality: a cobordism from (+,-) to nothing (just the “U”-shaped curve), which has a dual – and some 2-D cobordisms which specify that duality, and so on.  Specifying all this dualizability structure amounts to giving the image of all the generators of cobordisms, and determines the functors Z, and vice versa.

This is a rapid summary of six hours of lectures, of course, so for more precise versions of these statements, you may want to look into Chris’ thesis as linked above.

Homotopy QFT and the Crossed Menagerie

The next series of lectures in the school was Tim Porter’s, about relations between Homotopy Quantum Field Theory (HQFT) and various sort of crossed gizmos.  HQFT is an idea introduced by Vladimir Turaev, (see his paper with Tim here, for an intro, though Turaev also now has a book on the subject).  It’s intended to deal with similar sorts of structures to TQFT, but with various sorts of extra structure.  This structure is related to the “Crossed Menagerie”, on which Tim has written an almost unbelievably extensive bunch of lecture notes, of which a special short version was made for this lecture series that’s a mere 350 pages long.

Anyway, the cobordism category Bord_d described above is replaced by one Tim called HCobord(d,B) (see above comment about “bord” and “cobord”, which mean the same thing).  Again, this has d-dimensional cobordisms as its morphisms and (d-1)-dimensional manifolds as its objects, but now everything in sight is equipped with a map into a space B – almost.  So an object is X \rightarrow B, and a morphism is a cobordism with a homotopy class of maps M \rightarrow B which are compatible with the ones at the boundaries.  Then just as a d-TQFT is a representation (i.e. a functor) of Cob_d into Vect, a (d,B)-HQFT is a representation of HCobord(d,B).

The motivating example here is when B = B(G), the classifying space of a group.  These spaces are fairly complicated when you describe them as built from gluing cells (in homotopy theory, one typically things of spaces as something like CW-complexes: a bunch of cells in various dimensions glued together with face maps etc.), but B(G) has the property that its fundamental group is G, and all other homotopy groups are trivial (ensuring this part is what makes the cellular decomposition description tricky).

The upshot is that there’s a correspondence between (homotopy classes of) maps Map(X ,B(G)) \simeq Hom(\pi(X),G) (this makes a good alternative definition of the classifying space, though one needs to ).  Since a map from the fundamental group into G amounts to a flat principal G-bundle, we can say that HCobord(d,B(G)) is a category of manifolds and cobordisms carrying such a bundle.  This gets us into gauge theory.

But we can go beyond and into higher gauge theory (and other sorts of structures) by picking other sorts of B.  To begin with, notice that the correspondence above implies that mapping into B(G) means that when we take maps up to homotopy, we can only detect the fundamental group of X, and not any higher homotopy groups.  We say we can only detect the “homotopy 1-type” of the space.  The “homotopy n-type” of a given space X is just the first n homotopy groups (\pi_1(X), \dots, \pi_n(X)).  Alternatively, an “n-type” is an equivalence class of spaces which all have the same such groups.  Or, again, an “n-type” is a particular representative of one of these classes where these are the only nonzero homotopy groups.

The point being that if we’re considering maps X \rightarrow B up to homotopy, we may only be detecting the n-type of X (and therefore may as well assume X is an n-type in the last sense when it’s convenient).  More precisely, there are “Postnikov functors” P_n(-) which take a space X and return the corresponding n-type.  This can be done by gluing in “patches” of higher dimensions to “fill in the holes” which are measured by the higher homotopy groups (in general, the result is infinite dimensional as a cell complex).  Thus, there are embeddings X \hookrightarrow P_n(X), which get along with the obvious chain

\dots \rightarrow P_{n+1}(X) \rightarrow P_n(X) \rightarrow P_{n-1}(X) \rightarrow \dots

There was a fairly nifty digression here explaining how this is a “coskeleton” of X, in that P_n is a right adjoint to the “n-skeleton” functor (which throws away cells above dimension n, not homotopy groups), so that S(Sk_n(M),X) \cong S(M,P_n(X)).  To really explain it properly, though I would have to really explain what that S is (it refers to maps in the category of simplicial sets, which are another nice model of spaces up to homotopy).  This digression would carry us away from higher gauge theory, which is where I’m going.

One thing to say is that if X is d-dimensional, then any HQFT is determined entirely by the d-type of B.  Any extra jazz going on in B‘s higher homotopy groups won’t be detected when we’re only mapping a d-dimensional space X into it.  So one might as well assume that B is just a d-type.

We want to say we can detect a homotopy n-type of a space if, for example, B = B(\mathcal{G}) where \mathcal{G} is an “n-group”.  A handy way to account for this is in terms of a “crossed complex”.  The first nontrivial example of this would be a crossed module, which consists of

  • Two groups, G and H with
  • A map \partial : H \rightarrow G and
  • An action of G on H by automorphisms, G \rhd H
  • all such that action looks as much like conjugation as possible:
    • \partial(g \rhd h) = g (\partial h) g^{-1} (so that \partial is G-equivariant)
    • \partial h \rhd h' = h h' h^{-1} (the “Peiffer identity”)

This definition looks a little funny, but it does characterize “2-groups” in the sense of categories internal to \mathbf{Groups} (the definition used elsewhere), by taking G to be the group of objects, and H the group of automorphisms of the identity of G.  In the description of John Huerta’s lectures, I’ll get back to how that works.

The immediate point is that there are a bunch of natural examples of crossed modules.  For instance: from normal subgroups, where \partial: H \subset G is inclusion and the action really is conjugation; from fibrations, using fundamental groups of base and fibre; from a canonical case where H = Aut(G)  and \partial = 1 takes everything to the identity; from modules, taking H to be a G-module as an abelian group and \partial = 1 again.  The first and last give the classical intuition of these guys: crossed modules are simultaneous generalizations of (a) normal subgroups of G, and (b) G-modules.

There are various other examples, but the relevant thing here is a theorem of MacLane and Whitehead, that crossed modules model all connected homotopy 2-types.  That is, there’s a correspondence between crossed modules up to isomorphism and 2-types.  Of course, groups model 1-types: any group is the fundmental group for a 1-type, and any 1-type is the classifying space for some group.  Likewise, any crossed module determines a 2-type, and vice versa.  So this theorem suggests why crossed modules might deserve to be called “2-groups” even if they didn’t naturally line up with the alternative definition.

To go up to 3-types and 4-types, the intuitive idea is: “do for crossed modules what we did for groups”.  That is, instead of a map of groups \partial : H \rightarrow G, we consider a map of crossed modules (which is given by a pair of maps between the groups in each) and so forth.  The resulting structure is a square diagram in \mathbf{Groups} with a bunch of actions.  Each of these maps is the \partial map for a crossed module.  (We can think of the normal subgroup situation: there are two normal subgroups H,K of G, and in each of them, the intersection H \cap K is normal, so it determines a crossed module).  This is a “crossed square”, and things like this correspond exactly to homotopy 3-types.  This works roughly as before, since there is a notion of a classifying space B(\mathcal{G}) where \mathcal{G} =   (G,H,\partial,\rhd), and similarly on for crossed n-cubes.   We can carry on in this way to define a “crossed n-cube”, which correspond to homotopy (n+1)-types.  The correspondence is a little bit more fiddly than it was for groups, but it still exists: any (n+1)-type is the classifying space for a crossed n-cube, and any such crossed n-cube has an (n+1)-type for its classifying space.

This correspondence is the point here.  As we said, when looking at HQFT’s from HCobord(d,B), we may as well assume that B is a d-type.  But then, it’s a classifying space for some crossed (d-1)-cube.  This is a sensible sort of B to use in an HQFT, and it ends up giving us a theory which is related to higher gauge theory: a map X \rightarrow B(\mathcal{G}) up to homotopy, where \mathcal{G} is a crossed n-cube will correspond to the structure of a flat (n+1)-bundle on X, and similarly for cobordisms.  HQFT’s let us look at the structure of this structured cobordism category by means of its linear representations.  Now, it may be that this crossed-cube point of view isn’t the best way to look at B, but it is there, and available.

To say more about this, I’ll have to talk more directly about higher gauge theory in its own terms – which I’ll do in part IIb, since this is already pretty long.

A more substantial post is upcoming, but I wanted to get out this announcement for a conference I’m helping to organise, along with Roger Picken, JoĂŁo Faria Martins, and Aleksandr Mikovic.  Its website: https://sites.google.com/site/hgtqgr/home has more details, and will have more as we finalise them, but here are some of them:

Workshop and School on Higher Gauge Theory, TQFT and Quantum Gravity

Lisbon, 10-13 February, 2011 (Workshop), 7-13 February, 2011 (School)

Description from the website:

Higher gauge theory is a fascinating generalization of ordinary abelian and non-abelian gauge theory, involving (at the first level) connection 2-forms, curvature 3-forms and parallel transport along surfaces. This ladder can be continued to connection forms of higher degree and transport along extended objects of the corresponding dimension. On the mathematical side, higher gauge theory is closely tied to higher algebraic structures, such as 2-categories, 2-groups etc., and higher geometrical structures, known as gerbes or n-gerbes with connection. Thus higher gauge theory is an example of the categorification phenomenon which has been very influential in mathematics recently.

There have been a number of suggestions that higher gauge theory could be related to (4D) quantum gravity, e.g. by Baez-Huerta (in the QG^2 Corfu school lectures), and Baez-Baratin-Freidel-Wise in the context of state-sums. A pivotal role is played by TQFTs in these approaches, in particular BF theories and variants thereof, as well as extended TQFTs, constructed from suitable geometric or algebraic data. Another route between higher gauge theory and quantum gravity is via string theory, where higher gauge theory provides a setting for n-form fields, worldsheets for strings and branes, and higher spin structures (i.e. string structures and generalizations, as studied e.g. by Sati-Schreiber-Stasheff). Moving away from point particles to higher-dimensional extended objects is a feature both of loop quantum gravity and string theory, so higher gauge theory should play an important role in both approaches, and may allow us to probe a deeper level of symmetry, going beyond normal gauge symmetry.

Thus the moment seems ripe to bring together a group of researchers who could shed some light on these issues. Apart from the courses and lectures given by the invited speakers, we plan to incorporate discussion sessions in the afternoon throughout the week, for students to ask questions and to stimulate dialogue between participants from different backgrounds.

Provisional list of speakers:

  • Paolo Aschieri (Alessandria)
  • Benjamin Bahr (Cambridge)
  • Aristide Baratin (Paris-Orsay)
  • John Barrett (Nottingham)
  • Rafael Diaz (Bogotá)
  • Bianca Dittrich (Potsdam)
  • Laurent Freidel (Perimeter)
  • John Huerta (California)
  • Branislav Jurco (Prague)
  • Thomas Krajewski (Marseille)
  • Tim Porter (Bangor)
  • Hisham Sati (Maryland)
  • Christopher Schommer-Pries (MIT)
  • Urs Schreiber (Utrecht)
  • Jamie Vicary (Oxford)
  • Konrad Waldorf (Regensburg)
  • Derek Wise (Erlangen)
  • Christoph Wockel (Hamburg)

The workshop portion will have talks by the speakers above (those who can make it), and any contributed talks.  The “school” portion is, roughly, aimed at graduate students in a field related to the topics, but not necessarily directly in them.  You don’t need to be a student to attend the school, of course, but they are the target audience.  The only course that has been officially announced so far will be given by Christopher Schommer-Pries, on TQFT.  We hope/expect to also have minicourses on Higher Gauge Theory, and Quantum Gravity as well, but details aren’t settled yet.

If you’re interested, the deadline to register is Jan 8 (hence the rush to announce).  Some funding is available for those who need it.

On a tangential note, let me point out John Baez’ most recent “This Week’s Finds”, which has an accessible but fairly in-depth discussion of climate modelling.  There have been many years of very loud public discussion of this which, for reasons of politics, seems to involve putting the “Mathematical models are inherently elitist gibberish” and “Science knows everything so shut up, moron” positions on display and letting viewer decide.  This is known in the journalism trade as “balance”.  Obviously, within the research community working on them, there’s a mountain of literature on what the models model, how detailed they are, how they work, etc., but it mostly goes over my head, so John’s post strikes a nice balance for me.

Like most computer simulation models, they’re basically discrete approximations to big systems of differential equations – but exactly which systems, how they’re developed, how accurately they model the real thing, and the relative merits of simple vs. complex models is the main point.  The use of Monte Carlo methods and Bayesian analysis to tune the various free parameters is a key part of the matter of how accurate they should be.  Anyway – check it out.

Meanwhile, the TQFT club at IST recently started up its series of seminars.  The first few speakers were Rui Carpentier, Anne-Laure Thiel, and Marco Mackaay.  Rui is faculty here at IST, and a former student of Roger Picken (his thesis was on a topic closely related to what he was talking about).  Anne-Laure is a post-doc here at IST, mainly working with Marco, who, however, is actually at the University of the Algarve in Faro, Portugal, and had to come up to Lisbon specially for the seminar.  Anne-Laure and Marco were both speaking mainly about some of the Soergel bimodule stuff which came up at the Oporto meeting on categorification, which I posted about previously, so I’ll go over that in a bit more detail here.

First, though, Rui Carpentier’s talk:

3-colourings of Cubic Graphs and Operators

All these talks involve algebraic representations of categories that can be represented by some graphical calculus, but in this case, one starts with a category whose morphisms are precisely graphs with loose ends.  (The objects are non-negative integers, or, if you like, finite sets of dots which act as the vertices of the loose ends).  The graphs are trivalent (except at the input and output vertices, which are 1-valent), hence “cubic graphs”.  This category is therefore called \mathbf{CG}, and it has a small number of generators, which happen to be quite similar to those which generate the category of 2D-cobordisms (one of the connections to TQFT), though the relations are slightly different.

Roughly, and without drawing the pictures: the generators are cup and cap (the shapes \cup and \cap), two different trivalent vertices (a Y, and the same upside-down), the swap (an X where the strands cross without a vertex), and the identity (just a vertical line).  There are a number of relations, including Reidemeister moves, on these generating pictures, which ensure that they’re enough to identify graphs up to isotopy of the pictures.

Then the point is to describe graphs using operators – that is, construct a representation F :\mathbf{CG} \rightarrow \mathbf{Vect}.   Given any such representation, these generators provide all the structure maps of a bialgebra – chiefly, unit, counit, multiplication and co-multiplication – and the relations imposed by isotopy make this work (though unlike some other situations, it’s neither commutative nor cocommutative).  The representation F he constructs is based on 3-colourings of the edges of the graphs.  At the object level, it assigns to a dot the 3-dimensional vector space V= span(e_1,e_2,e_3).  Being monoidal, F takes the object n to V^{\otimes n} – the tensor product of the spaces at each vertex.

The idea is that choosing a basis vector in this space amounts to picking a colouring of the incoming and outgoing edges.  For morphisms, we should note that the rule that says when a colouring is admissible is that all the edges incident to a given vertex must have different colours.  Then, given a morphism (graph) G : m \rightarrow n, we can describe the linear map F(G) most easily by saying that the component in the matrix, given an incoming and outgoing basis vector, just counts the number of admissible graphs that agree with the chosen colourings on the in-edges and out-edges.

There’s another functor, \hat{F}, which counts these graphs with a sign, which marks whether the graph contains an odd or an even number of crossings of differently-coloured edges – negative for odd, positive for even.  This  is the “Penrose evaluation” of the graph.

So these maps give the “operators” of the title, and the rest of the point is to use them to study graphs and their colourings.  One can, in this setup, rewrite some graphs as linear combinations of others – so-called “Skein relations” hold, for example, so that, after applying F, the composite of multiplication and comultiplication (taking two points to two points, through one cut-edge) is the same as the identity minus the swap.  This sort of thing appears in formal knot theory all the time, and is a key tool for recoupling in spin networks, and so on…

Given this “recoupling” idea, there are some important facts: first, any graph can be rewritten as a linear combination of planar graphs, and any planar graph with cycles can be reduced to a sum of planar graphs without cycles.  (Rui gave the example of decomposing a pentagonal cycle as a linear combination of four other graphs, three of which are disconnected).  So in fact any graph decomposes as a linear combination of forests (cycle-free graphs, the connected components of which are called “trees”, hence the name).  Another essential fact is that, due to the Euler characteristic of the plane, any planar graph can be split into two parts with at most five edges between them (the basis of the solution to the three utilities puzzle).  Then it so happens that the space of graphs connecting zero in-edges to five out-edges is a 6-dimensional space, \mathcal{V}^o_5, generated by just six forests (including one lonesome tree).

So one theorem which Rui told us about, which can be shown using the so-called Penrose relations (provable using the representations F and \hat{F}), is that there’s just one such graph (which he described in the particular basis above) that evaluates to zero when composed with some other graph.  The proof of this uses the Four Colour Theorem (3-colouring of graph edges being related to 4-colouring of planar regions); in fact, the two theorems are equivalent so if anyone can find an alternative proof of this one, the bonus is another proof of the FCT.

Finally, he gave a conjecture that, if true, would help recognize planar graphs just by the operators produced by the representation \hat{F} (at least it proposes a necessary condition).  This conjecture says that if a planar graph with five output edges (the maximum, remember) is written in the basis mentioned above, then the sum of the coefficients of the five disconnected trees is nonnegative.  (Thus, the connected tree doesn’t contribute to this measure).  This is still just a conjecture – Rui said that to date neither proof nor counterexample has been found.

Soergel Bimodules, Singular and Virtual Braids

As I mentioned up top, I previously posted a bit about work on Soergel bimodules when describing Catharina Stroppel’s talk at the meeting in Faro in July.  To recap: they are associated with categories of modules over rings – specifically, rings of certain classes of symmetric functions.  Even more specifically, given a partition \lambda of an integer n, there is a subgroup of the symmetric group S_{\lambda} \subset S_n which fixes the partition.  All such groups act on the ring of n-variable polynomial functions R =\mathbb{Q}[x_1, \dots, x_n], and the ones fixed by S_{\lambda} form the ring R^{\lambda}.

Now, these groups are all related to each other in a web of containments, hence so are the rings.  So the module categories R^{\lambda} are connected by various functors.  Given a containment R^{\lambda '} \subset R^{\lambda}, modules over R^{\lambda} can be restricted to ones over R^{\lambda '}, and modules over R^{\lambda '} can be induced up to ones over R^{\lambda}.  The restriction and induction functors can be represented as “tensor with a bimodule” (this is much the same classification as that for 2-linear maps which I’ve said a bunch about here, except that those must be free).  Applying induction functors repeatedly gives abitrarily large bimodules, but they are built as direct sums of simple parts.  Those simple parts, and any direct sums of them, are Soergel bimodules.  The point is that such bimodules describe morphisms.

So in the TQFT club, Marco Mackaay gave the first of a series of survey talks on this topic, and Anne-Laure Thiel gave a talk about the “Categorification of Singular Braid Monoids and Virtual Braid Groups”.  Since Marco’s talk was the first in a series of surveys, and a lot of what it surveyed was work described in my post on the Faro meeting, I’ll just mention that it dealt with the original motivation of a lot of this work in categorifying representation theory of Lie algebras (c.f. the discussion of the Khovanov-Lauda categorification of quantum groups in the previous post), and also got a bit into some of the different diagrammatic calculi created for that purpose, along the lines of the talks by Ben Webster and Geordie Williamson at that meeting.  Maybe when Marco has given more of these talks, I’ll return to this one here as well.

Now, the starting point of Anne-Laure’s talk was that the setup above lets one define a category with a presentation like that of the Hecke algebra (a quotient of the group algebra of the braid group), where exact relations become isomorphisms.  That is, we go from a category where morphisms are braids (up to isotopy and Reidemeister moves and so forth as usual) to a 2-category where the morphisms are bimodules, which happen to satisfy the same relations.  (The 2-morphisms, bimodule maps, are what allow relations to hold weakly…)

Specifically, the generators of the braid group are \sigma_i, the braids taking the i^{th} strand over the (i+1)^{st}.  The parallel thing is B_i = R \otimes_{R^{\sigma_i}} R, where here we’re talking about the subgroup generated by the transposition of i and i+1.  In the language of partitions, this corresponds to a \lambda with one part of size two, (i,i+1), and the rest of size one.  Now, since this bimodule is actually built from polynomials in R, it naturally has a grading – this corresponds to the degree of q, since the Hecke algebra involves a quotient giving q-deformed relations – so there is a degree-shift operation that categorifies multiplication by q.  This much is due to Soergel.

Anne-Laure’s talk was about extending this to talk about a categorification, first of the braid group in terms of complexes of these bimodules (due actually to Rouquier), then virtual and singular braids.  These, again, are basically creatures of formal knot theory (see link above).  They can be described by a presentation similar to that for braids – just as the braid group has a generators-and-relations presentation in terms of over-crossings of adjacent strands, these incorporate other kinds of crossings.  Singular braids allow a sort of “through” crossing, where the i^{th} strand goes neither over nor under the (i+1)^{st}.  Virtual braids (the braid variant on virtual knots) have a special type of marked crossing called the “virtual crossing”, drawn with a little circle around it.  These are included as new generators in describing the virtual braid group, and of course some new relations are added to show how they relate to the original generators – variations on the Reidemeister moves, for example.

To categorify this, Anne-Laure explained that these new generators can also be represented by bimodules, but these ones need to be twisted.  In particular, twisting the bimodule R by the action of a permutation \omega \in S_n gives R_{\omega}, which is the same as R as a left R-module, but is acted on by an element a \in R on the right through multiplication by \omega(a), so that b \cdot p \cdot a = bp(\omega(a)).  Then the new generators, beyond the B_i = R \otimes_{R^{\sigma_i}} R, are of the form R_{\omega} \otimes_{R^{\omega '}} R.  These then satsify the right relations for this to categorify the virtual braid group.

So I recently received word that this paper had been accepted for publication by Applied Categorical Structures. Since I’ll shortly be putting out another which uses its main construction to build Extended Topological Quantum Field Theories, it’s nice and appropriate to say something about that. But actually, just at the moment, I want to take a slightly different approach.

Toward the end of February, I went up to Waterloo to the Perimeter Institute, where my friend Derek Wise was visiting with Andy Randono – apparently they’re working on a project together that has something to do with Cartan Geometry, which is a subject that plays a big role in Derek’s thesis.

However, Derek was speaking in their seminar about Extended TQFT (his slides are now up on his website, and there’s also a video of the talk available). Actually, a lot of what he was talking about was work of mine, since we’re working on a project together to constructs ETQFT’s from Lie groups (most likely compact ones at first, since all the usual analytical problems with noncompact groups turn up here). However, I really enjoyed seeing Derek talk about it, because he has a sharper grasp than I do of how this subject appears to physicists, and the way he presented this stuff is very different from the way I usually talk about it (you can see me in the video trying to help deal with a question at the end from Rafael Sorkin and Laurent Freidel, and taking a while to correctly understand what it was, partly because of this jargon gap – I hope to get better).

So, for example, describing a TQFT in the Atiyah/Segal axiomatic formulation is fairly natural to someone who works with category theory, but Derek motivated it as a way of taking a “deeper look at the partition function” for a certain field theory. The idea is that a partition function Z for a quantum field theory associates a number to a space M, satisfying certain rules. It is usually described by some kind of integral. Typically in QFT, these are rather tricky integrals – a topological QFT has the nice feature that, since it has no local degrees of freedom, these integrals are much more tractable. Of course, this is a mathematically nice feature that comes at the expense of physical relevance, but such is life.

Anyway, the idea is that the partition function Z for an n-dimensional TQFT can be thought of as assigning, not just numbers to n-dimensional manifolds M, but something more which reduces to this in a special case. Specifically, Z assigns a Hilbert space to any codimension-1 submanifold of M, in a particular way which Derek passed over by saying it “satisfies some compatibility conditions”. For an audience of mathematicians, you can gloss over this just as quickly by saying the assignments are “functorial”, or even with more detail saying the conditions make Z a symmetric monoidal functor.

Part of the point is that these conditions are about as obvious on physical grounds as they are if you’re a category theorist. For example, the fact that composition is preserved by the functor Z can be interpreted physically as saying that the number Z(M) given by the partition function isn’t affected by how we chop up the manifold M to analyse it. The fact that Z is a monoidal functor ends up meaning that the “unit” for manifolds under unions (namely, the empty manifold with no points, which you can add to things without affecting them) gets assigned the Hilbert space \mathbb{C}, which is the unit for Hilbert spaces with respect to the tensor product \otimes. The fact that this is so means we can treat a manifold with no boundary as going from one (empty) boundary to another (empty) boundary – it therefore gets assigned a linear map from \mathbb{C} to \mathbb{C} – a number. Seeing how this linear map comes from composing pieces of the manifold is what “a deeper look at the partition function” means.

ETQFT does essentially the same thing, at one level deeper. The point is that a TQFT breaks apart a manifold by treating it as a series of pieces – manifolds with boundary, glued together at their boundaries. An ETQFT does the same to these pieces, treating them as composed of pieces – manifolds with corners – which are glued orthogonally to the gluing just mentioned. That is, there are two kinds of composition, so we’re in some sort of 2-category (bi-, or double- depending on how you formulate things). The essential point is that now, to manifolds without boundary, which are of codimension 1, we assign Hilbert spaces – and to top-dimensional manifolds WITH boundary, we assign maps of Hilbert spaces.

An ETQFT attempts to give a “deeper-still look at the partition function” by seeing how the Hilbert space arises from composition of pieces in this new direction, along boundaries of codimension 2. The way Derek describes this for physicists is to say that the ETQFT describes how that Hilbert space is “built from local data”, which he described in the usual physics language of path integrals. First of all, the conventional thing in physics is to take Z(\Sigma) for a (codimension-1) manifold \Sigma to be L^2(\mathcal{A}_0(\Sigma)/\mathcal{G}(\Sigma)) – the space of square-integrable functions on the quotient of the space \mathcal{A}_0(\Sigma) of flat G-connections on M by the action of the group of gauge transformations \mathcal{G}(\Sigma).

Given a manifold M with boundary components \Sigma and \Sigma ', the standard quantum field theory formalism to describe the map Z(M) : Z(\Sigma) \rightarrow Z(\Sigma ') given by a TQFT is to describe how it interacts with particular state-vectors in the Hilbert spaces for the source and target boundary components of M. So then:

\langle \psi | Z(M) | \phi \rangle = \int_{\mathcal{A}_0(M)/\mathcal{G}} \mathcal{D}A \overline{\psi(A|_{\Sigma '})} e^{i S([A])} \phi(A|_{\Sigma})

The point being, a flat connection A has some action on it, which depends only on its gauge equivalence class [A] (“the Lagrangian has gauge symmetry”), and it restricts to give flat connections on \Sigma and \Sigma ', on which the L^2-functions \psi and \phi act, to give something we can integrate. The measure \mathcal{D}[A] is a crucial entity here, and in general can be a real puzzle, but at least for discrete groups, it’s just a weighted counting measure which effectively gives us the groupoid cardinality of the quotient space. As for the action S, the simplest possible case just says the action of any flat connection is zero – hence this expression is just finding the (groupoid) cardinality, or more generally measuring the (stacky) volume, of the configuration space for flat connections. There are other possible actions, though.

Derek gives an explanation of how to interpret this in terms of the “pull-push” construction, which I’ve talked about elsewhere here, including in the above paper, so right now, I’ll just pass to the next layer of the ETQFT layer cake – codimension-2. Here, there is a similar formula, which also has an interpretation in terms of a “pull-push” construction, but which can be written as a categorified path integral.

So now the \Sigma has boundary, and connects “inner” codimension-2 boundary component B_1 to “outer” boundary component B_2. Then, say, B_1 gets assigned the category of all gauge-equivariant “bundles” of Hilbert spaces on \mathcal{A}_0(B_1), rather than the space of gauge-invariant functions. (Derek carefully avoided using the term “category”, to stay physically motivated – and the term “bundle” is accurate in the case of a discrete gauge group G, but in general one has to appeal to the theory of measurable fields of Hilbert spaces, since they needn’t be locally trivial). Then given particular Hilbert bundles \mathcal{H} and \mathcal{K} on the spaces \mathcal{A}_0(B_1) and \mathcal{A}_0(B_2) respectively, we can define what Z(\Sigma) is by:

\langle \mathcal{K} | Z(M) | \mathcal{H} \rangle = \int_{\mathcal{A}_0(M)/\mathcal{G}} \mathcal{D}A \mathcal{K}(A|_{B_2}) \otimes T_A \otimes \mathcal{H}(A|_{B_1})

The interpretation is much like the previous formula: now we’re direct-integrating Hilbert spaces, instead of integrating complex functions – and we get a Hilbert space instead of a complex number, but this is in some sense superficial. Something any physicist would notice right away (or anyone comparing this to the previous formula) is that the exponential of the action S([A]) seems to have gone missing, to be replaced by some Hilbert space T_A. If we’re using the trivial action S \cong 0, this is fine, but otherwise, how exactly S affects the direct integral would take some explaining. For now, let’s just say that we should think of S([A]) as being folded into either the inner product on T_A, or into the measure \mathcal{D}A: it shows up in its effect on the inner product on the Hilbert space that this direct integral produces.

Let me jump to the end of Derek’s talk here, to get at some conceptual aspect of what’s happening here. The axiomatic way of talking about ETQFT, namely Ruth Lawrence’s way, is to say we assign a 2-Hilbert space to the codimension-2 manifolds. But “2-Hilbert space” is an off-putting bit of jargon, so instead the suggestion is to replace it with “von Neumann algebra”.

The point is that 2-Hilbert spaces are thought (according to a paper by Baez, Baratin, Friedel and Wise) to be just categories of representations of vN algebras. Being a 2-Hilbert space means, for instance, that they’re additive (by direct sum), \mathbb{C}-linear (there is a vector space of intertwiners between any two representations), have duals, and so on. Moreover, they’re monoidal 2-Hilbert spaces, since there is a tensor product. Their idea is that the two ideas correspond exactly. In any case, the way the ETQFT construction in question works actually passes through a von Neumann algebra. This comes from the groupoid algebra that’s associated to a certain group action. Namely, the action of the gauge group on the space of flat G-connections on the manifold M.

Then the way we can look more closely at the “structure of the partition function” is by seeing the Hilbert space associated to a codimension-1 manifold as actually being a kind of morphism of von Neumann algebras. In particular, it’s a Hilbert bimodule, which is acted on by the source algebra (say A) on the left, and the target algebra (B) on the right. This is intimately connected with the stuff I was writing about recently about Morita equivalence, and so to the 2-Hilbert space view. In particular, a Hilbert bimodule H gives an adjoint pair of linear functors (or “2-linear maps”) between the representation categories of algebras.

So shortly I’ll make a post about some papers coming out, and get back to this point…

Next Page »