Skip to the Main Content

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

June 28, 2006

Seminar on 2-Vector Bundles and Elliptic Cohomology, V

Posted by Urs Schreiber

Part V of our seminar on elliptic cohomology and 2-vector bundles.

part topic based on I literature on elliptic cohom. introdution to 2-vector bundles II,III,IV ellipt. coh. as categorified K-theory Baas-Dundas-Rognes classes of 2-vector bundles V-VI elliptic cohom. and strings V introduction K-theory and 1dSQFT Stolz-Teichner Witten genus and 2dSCFT VI tmfspectrum material used in string th. applications (?) Kriz-Sati \array{ \mathbf{\text{part}} & \mathbf{\text{topic}} & \mathbf{\text{based on}} \\ \href{http://golem.ph.utexas.edu/string/archives/000737.html}{\text{I}} & \text{literature on elliptic cohom.} \\ & \text{introdution to 2-vector bundles} \\ \\ \href{}{\text{II}}, \href{}{\text{III}}, \href{}{\text{IV}} & \text{ellipt. coh. as categorified K-theory} & \href{http://arxiv.org/abs/math.AT/0306027}{\text{Baas-Dundas-Rognes}} \\ & \text{classes of 2-vector bundles} \\ \\ \text{V-VI} & \text{elliptic cohom. and strings} & \\ \\ \text{V} & \text{introduction} \\ & \text{K-theory and 1dSQFT} & \href{http://math.ucsd.edu/~teichner/Papers/Oxford.pdf}{\text{Stolz-Teichner}} \\ & \text{Witten genus and 2dSCFT} \\ \\ \text{VI} & \mathrm{tmf}-\text{spectrum} \\ \\ & \text{material used in string th. applications (?)} & \text{Kriz-Sati} }

Outline of Part V

\bullet Introduction.

\bullet Background information.

\;\;\;\;\;1) What is a genus?

\;\;\;\;\;2) What is a generalized cohomology theory?

\;\;\;\;\;3) What do we need to know about elliptic curves?

\;\;\;\;\;4) What is the elliptic genus?

\;\;\;\;\;5) What is elliptic cohomology?

\bullet 6) How is ellitpic cohomology realized geometrically?

a) Warmup: how is the landscape of superpoint theories equal to the K-theory spectrum?

b) How does one expect the landscape of superstring theories to be equal to the spectrum of elliptic cohomology?

Introduction.

Supersymmetric point particles have partition functions which compute the index of the supercharge (a Dirac operator). This index can be thought of as an element in the K-theory over a point.

Supersymmetric strings (heterotic strings, in particular) have partition functions which compute the Witten genus of a manifold. This Witten genus can be thought of as an element in elliptic cohomology over a point.

fundamental object partition function ring of values cohomology group law chrom. filt. (1)branes ? ? ordinary xy=x+y 0 point particles index K xy=x+y+xy 1 strings Wittengenus modular forms elliptic ell. curve 2. \array{ \mathbf{\text{fundamental object}} & \mathbf{\text{partition function}} & \mathbf{\text{ring of values}} & \mathbf{\text{cohomology}} & \mathbf{\text{group law}} & \mathbf{\text{chrom. filt.}} \\ (-1)-\text{branes} & ? & ? & \text{ordinary} & x\cdot y = x + y & 0 \\ \text{point particles} & index & \mathbb{Z} & K & x\cdot y = x + y + xy & 1 \\ \text{strings} & Witten genus & \text{modular forms} & elliptic & \text{ell. curve} & 2 } \,.

The supercharges whose indices are being computed here can be taken to define a background for the superpoint or the superstring. Hence an element of K-theory over a point corresponds to a connected component of the space of all superparticle backgrounds, in a sense.

The sense in which this is true has been made precise by Stephan Stolz and Peter Teichner, and elaborated on by Elke Markert in her thesis

E. Markert
Connective 1-dimensional euclidean field theories
(pdf)

Theorem. The space of 1-dimensional supersymmetric quantum field theories is homotopy equivalent to the K-theory spectrum.

This means in particular, that supersymmetric quantum mechanics knows much more than the K-theory over points - it knows all of K-theory. In string speak, it says that the landscape of superparticle theories is homotopy equivalent to the spectrum of K-theory.

Based on similar observations, Graeme Segal had suggested already twenty years ago, in the last section of

Graeme Segal
Elliptic Cohomology
Séminaire Bourbaki, no. 695
Astérisque 161-162 (1988),

that the space of all string theories is similarly related to elliptic cohomology.

The underlying strategy for making this precise was that of nn-transport (\to).

It is natural to encode the concept of propagation in (supersymmetric) quantum mechanics in terms of functors QM:1dCob Hilb (t) (Hexp(itΔ)H) \array{ QM : 1d\mathrm{Cob} &\to& \mathrm{Hilb} \\ (\bullet \overset{t}{\to} \bullet) &\mapsto& (H \overset{\exp(-i t \Delta)}{\to} H) } from 1-dimensional worldlines to Hilbert spaces.

Segal’s famous definition of 2-dimensional conformal field theory

Graeme Segal
The definition of conformal field theory
in U. Tillmann (ed.)
Topology, Geometry and Quantum Field Theory
Lond. Math. Soc. Lecture Note Series 308
Cambridge (2002)

accordingly says, roughly, that 2dCFT is a projective functor on 2-dimensional conformal cobordisms 2dCFT:2dConfCob Hilb, \array{ 2dCFT : 2d\mathrm{ConfCob} &\to& \mathrm{Hilb} } \,, and tries to relate the space of all these functors to elliptic cohomology.

Stolz and Teichner notice in

St. Stolz & P. Teichner
What is an elliptic object?
in U. Tillmann (ed.)
Topology, Geometry and Quantum Field Theory
Lond. Math. Soc. Lecture Note Series 308
Cambridge (2002)
(pdf)

that this idea has to be refined in two aspects.

1) It is necessary to concentrate on supersymmetric QFT. This naturally leads to the required grading and prevents the resulting spaces from being topologically trivial.

2) It is necessary to allow string bits. Otherwise excision does not hold, which, in string speak, means something like that otherwise instanton effects lead to nonlocalities.

The last point means, technically, that we consider not 1-functors on 2-cobordisms but 2-functors on 2-paths, of the form 2dSCFT:super conformal 2-paths2Hilb, 2d\mathrm{SCFT} : \text{super conformal 2-paths} \to 2\mathrm{Hilb} \,, where 2Hilb2\mathrm{Hilb} is something like a 2-category of 2-Hilbert spaces (\to).

The refined version of Segal’s conjecture would hence read

Conjecture. The space of all 2-functors 2dSCFT2d\mathrm{SCFT} is homotopy equivalent to the elliptic cohomomolgy spectrum tmf\mathrm{tmf} of topological modular forms.

I’ll explain what all this means in detail below.

Before closing this introduction, I’ll comment on the following question.


How is this related to our previous session on elliptic cohomology and 2-vector bundles? (\to)

The index of a Dirac operator takes values in the integers \mathbb{Z}. We should think of this integer as a decategorified virtual vector bundle over a point (kerD,cokerD) , \array{ (\mathrm{ker}D,\mathrm{coker}D) \\ \downarrow \\ \bullet } \,, whose fiber is the kernel minus the cokernel of the Dirac operator. indD =dimkerDdimcokerD =dimkerDcodimimD. \begin{aligned} \mathrm{ind} D &= \mathrm{dim}\,\mathrm{ker}D - \mathrm{dim}\,\mathrm{coker}D \\ &= \mathrm{dim}\,\mathrm{ker}D - \mathrm{codim}\, \mathrm{im}D \end{aligned} \,. So we see that here the integers which appear =K(pt)=Groth(Vect) \mathbb{Z} = K(\mathrm{pt}) = \mathrm{Groth}(\mathrm{Vect}) are really the decategorification of the category of vector spaces.

Now, the index of a Dirac operator on loop space in general takes values not on \mathbb{Z}, but in the polynomial ring [[q,q 1]]\mathbb{Z}[\![q, q^{-1}]\!] (more precisely, in the ring of integral modular forms in qq).

Again, the integer coefficients here are really K-classes of a point. In fact, the coefficients a ka_k in the index ind Witten(G)= ka kq k \mathrm{ind}_\text{Witten}(G) = \sum_{k \in \mathbb{Z}} a_k\, q^k of the heterotic string supercharge GG (a Dirac operator on loop space) are the dimension of infinitely many virtual vector spaces, one for each Fourier mode of the string ind Witten(G) = k(dimkerG| kdimcokerG| k)q k, \begin{aligned} \mathrm{ind}_\mathrm{Witten}(G) &= \sum_{k \in \mathbb{Z}} ( \mathrm{dim}\,\mathrm{ker}G|_k - \mathrm{dim}\,\mathrm{coker} G|_k ) \, q^k \end{aligned} \,, where G| kG|_k is the restriction of the supercharge to the kk-th eigenspace of the operator P:=LL¯ P := L - \bar L which generates rigid rotations of (the parametrization of) the string.

It follows, that this loop space index is not a virtual vector bundle over a point, but a virtual vector bundle over the integers. (kerG,cokerG) (kerG| 0,cokerG| 0) (kerG| 1,cokerG| 1) (kerG| 2,cokerG| 2) = {0} {1} {2} . \array{ (\mathrm{ker}G, \mathrm{coker} G)&& \cdots & (\mathrm{ker}G|_0, \mathrm{coker} G|_0) & (\mathrm{ker}G|_1, \mathrm{coker} G|_1) & (\mathrm{ker}G|_2, \mathrm{coker} G|_2) & \cdots \\ \downarrow &=& \cdots & \downarrow & \downarrow & \downarrow & \cdots \\ \mathbb{Z} && \cdots & \{0\} & \{1\} & \{2\} & \cdots } \,. Recall (\to, \to, \to) that we can interpret such a virtual vector bundle as a virtual 2-vector in the countably infinite dimensional Kapranov-Voevodsky 2-vector space Vect \mathrm{Vect}^\mathbb{Z}.


Warning: Technical details of definitions may be imprecise. For instance, some necessary technical conditions on elliptic curves are missing. Don’t rely on what I write, but check the referenced literature.

Background information.

1) What is a genus?

See for instance

G. Segal, Elliptic Cohomology, or

Gerd Laures,
An introduction to Elliptic Cohomology and Conformal Field Theory
(pdf).

Some comments in the introduction of

Steven Rosenberg
Nonlocal Invariants in Index Theory
(pdf).

also prove useful.

Definition. An RR-valued genus μ\mu is a ring homomorphism μ:Ω SOR \mu : \Omega_{\mathrm{SO}} \to R from the bordisms ring (\to) Ω SO:=closed oriented manifoldscobordisms \Omega_{\mathrm{SO}} := \frac{\text{closed oriented manifolds}}{cobordisms} to RR. The product in Ω SO\Omega_\mathrm{SO} is the cartesian product of manifolds and the sum is the disjoint union of manifolds.

Examples.

\bullet The Euler characteristic Xχ(X)X \mapsto \chi(X) (which is the index of D=d+d *D = d + d^* under the grading by form degree) is close to being a genus, but is not cobordism invariant.

\bullet The signature genus is a genus. It is (as far as I understand) the index of D=d+d *D = d + d^*, but under the grading of differential forms as positive or negative, according to the signature of the bilinear form (v,w) Xvw. (v,w) \mapsto \int_X v \wedge w \,.

\bullet The A^\hat A-genus is the index of a Dirac operator coming from a spinor bundle (Atiyah-Singer).

\bullet The elliptic genus has been interpreted (non-rigorously) by Witten as the index of a Dirac operator (the heterotic’s string supercharge, in fact) on loop space.

Bottom line. Genera appear as indices of Dirac operators. As such, they are related to cohomologies of points.

2) What is a generalized cohomology theory?

See for instance

Jacob Lurie
A survey of elliptic cohomology
(pdf)
section 1.1 .

If you feel you need a more gentle introduction, try

John Baez
TWF 149
TWF 151 .


Let AA be any abelian group. Given a topological space XX, there are several equivalent ways to define the nnth singular cohomology group H n(X,A) H^n(X,A) with values in AA. A particularly nice one is to say that this group equals that of homotopy classes of maps from XX to the space K(A,n)K(A,n), which is defined to have all homotopy groups trivial except for the nnth one, which is isomorphic to AA: H n(X,A)[X,K(A,n)]. H^n(X,A) \simeq [X,K(A,n)] \,. For this reason, one says that K(A,n)K(A,n) represents ordinary cohomology.

Even more is true. It is a simple fact that the (based) loop space of K(A,n)K(A,n) is homotopy equivalent to K(A,n1)K(A,n-1) ΩK(A,n)K(A,n1). \Omega K(A,n) \simeq K(A,n-1) \,. A list E=(E n) nE = (E_n)_{n\in \mathbb{N}} of topological spaces such that each space is the loop space of the next one ΩE nE n1 \Omega E_n \simeq E_{n-1} is called a spectrum. The sequence E n:=K(A,n)E_n := K(A,n) is called the Eilenberg-MacLane spectrum, which represents ordinary cohomology.

Stated this way, the definition of generalized cohomology is obvious:

Definition. A generalized cohomology is a collection of functors H n:Top AbGrp X H n(X) \array{ H^n : \mathrm{Top} &\to& \mathrm{AbGrp} \\ X &\mapsto& H^n(X) } from topological spaces to abelian groups, which is represented by a spectrum E=(E n)E = (E_n): H n(X):=[X,E n]. H^n(X) := [X,E_n] \,.

Historically, things were found the other way around. First Eilenberg and Steenrod extracted a list of axioms, the Eilenberg-Steenrod axioms, which are satisfied by ordinary cohomology, and defined a generalized cohomology to be anything satisfying these axioms. Later, Brown proved his representability theorem, which says that every spectrum determines a generalized cohomology theory and every generalized cohomology comes from a spectrum (see Lurie, p.8).

The first interesting example of a generalized cohomology theory, right after ordinary cohomology, is K-theory XK n(X). X \mapsto K_n(X) \,. A well known result by Atiyah says that K 0K_0 is represented by the space Fred(H)\mathrm{Fred}(H) of Fredholm operators on some seperable Hilbert space HH K(X)[X,Fred(X)]. K(X) \simeq [X,\mathrm{Fred}(X)] \,.

(— to do: what is the entire spectrum? what happens for different flavors of K-theory? —)

Some generalized cohomologies have special properties. In particular, we are used to there being a graded ring structure on the cohomology groups (given by the cap product for ordinary cohomology, and by the tensor product of vector bundles for K-theory).

It is clear that in order for H (X)[X,E ]H^\bullet(X) \simeq[X,E_\bullet] to be a ring, the spaces E E_\bullet need to have a ring-like structure themselves. Since everything in sight is defined only up to homotopy, there is freedom in having this ring structure defined only up to higher coherent homotopy. There is one natural choice for how to deal with these higher coherencies (see Lurie, p. 9).

Definition. A spectrum EE which is a graded ring in this higher coherent sense is called a E E_\infty-ring spectrum , or simply an E E_\infty-ring. The “EE” is supposed to remind you of commutativity, while the “ {}_\infty” reminds us of the fact that everything holds only up to homotopy, which is coherent up to homotopy, which… and so on.

3) What do we need to know about elliptic curves?

If you don’t know anything about elliptic curves, the text

Charles Daney
Elliptic Curves and Elliptic Functions
(html)

is a rather nice brief introduction. Here, I won’t need anything that goes beyond the basic facts stated in that article. In fact, all that is really necessary in order to get a basic idea of elliptic cohomology are these

2 Facts.

1) An elliptic curve over a field kk is the collection of solutions in kxkk x k to an equation in kk of the form f(x,y)=Ax 3+Bx 2+Cx+Dy 2=0, f(x,y) = A x^3 + B x^2 + C x + D - y^2 = 0 \,, where A,B,C,DkA,B,C,D \in k are constants defining the elliptic curve. For the present purpose, the single most important property of an elliptic curve S={(x,y)k×k|f(x,y)=0}S = \{(x,y) \in k\times k | f(x,y) = 0\} is that it is naturally equipped with an abelian (algebraic) group structure S×SSS \times S \to S.

2) Elliptic curves over the complex numbers are precisely the same thing as Riemann surfaces of genus 1, hence the same as 1-loop string diagrams.

This makes it quite plausible that elliptic curves play a central role in a cohomology theory which assigns to a point in the space of 2dSCFTs the corresponding 1-loop partition function.

4) What is the elliptic genus?

See

G. Segal, Elliptic Cohomology

together with

Edward Witten
Elliptic Genera And Quantum Field Theory
Commun.Math.Phys. 109 525 (1987)
(archive)

and

Edward Witten
The Index Of The Dirac Operator In Loop Space.
Proc. of Conf. on Elliptic Curves and Modular Forms in Algebraic Topology
Princeton, 1986.
(archive).

I can also very much recommend to have a look at the trascript of a talk

Johannes Ebert
The partition function of CFT’s: Connection to modular forms.
(notes)

given in our winter school on elliptic objects (\to).


For our purposes, we are interested in 2-dimensional field theory with N=(1,0)N= (1,0)-superconformal symmetry, i.e. in the heterotic string. (We could also work with type II superstrings, but these would yield less cohomological information.)

There is a graded Hilbert space HH of states, which we can think of as something like the space of sections of a spinor bundle over free loop space.

Represented on HH are a Laplace-like operator Δ=L+L¯ \Delta = L + \bar L and an operator P=i(LL¯) P = i(L - \bar L) which generates rigid rotations of (the parameterization) of the loops.

The chiral combinations L =12(ΔiP) L¯ =12(Δ+iP) \begin{aligned} L &= \frac{1}{2}(\Delta -i P) \\ \bar L &= \frac{1}{2}(\Delta + i P) \end{aligned} are the 0-modes of two Virasoro algebras.

We say there is N=(n,m)N=(n,m) supersymmetry essentially (up to the fact that we are only considering 0-modes at the moment) if there are nn mutually (graded-)commuting odd-graded operators G (i)G^{(i)} such that L=(G (i)) 2 L = (G^{(i)})^2 and mm mutually (graded-)commuting odd-graded operators G¯ (i)\bar G^{(i)} such that L¯=(G¯ (i)) 2. \bar L = (\bar G^{(i)})^2 \,. Restricting to N=(1,0)N=(1,0) we have a single odd-graded operator GG with L=G 2L = G^2.

We are interested in something like the index of this GG.

For ordinary susy quantum mechanics, we compute the index by means of the partition function over a circle of length tt (which turns out to be independent of tt).

Here we compute the partition function over a torus of length tt and twist ss as ind WittenG =str(exp(tΔ)exp(sP)) =str(q Lq¯ L¯), \begin{aligned} \mathrm{ind}_\mathrm{Witten} G &= \mathrm{str} \left( \exp(-t \Delta)\exp(-s P) \right) \\ &= \mathrm{str} \left( q^L {\bar q}^{\bar L} \right) \,, \end{aligned} where q=exp((t+is))q = \exp(-(t+is)) and q¯=exp((tis))\bar q = \exp(-(t-is)).

Since L=G 2L = G^2 is the square of an odd operator, the usual argument applies and we see that this partition function localizes on the kernel of LL =str(q¯ L¯| kerL). \cdots = \mathrm{str}({\bar q}^{\bar L}|_{\mathrm{ker}L}) \,. We can further simplify this by splitting the supertrace into the contributions coming from the eigenspaces Eig(P,k)=H kH\mathrm{Eig}(P,-k) = H^k \subset H of P=i(LL¯)P = i(L - \bar L) = k(q k)sdim(kerLH k). \cdots = \sum_{k \in \mathbb{Z}}(q^k) \mathrm{sdim}(\mathrm{ker}L \cap H^k) \,.

This power series in qq is the index of our loop space Dirac operator. (Had we used full N=(1,1)N=(1,1) supersymmetry the remaining sum would have localized itself on the kernel of L¯\bar L and collapsed to a mere integer.)

ind WittenG\mathrm{ind}_\mathrm{Witten} G is a weak integral modular form, which means that

\bullet it is a holomorphic function ff on the upper half plane

\bullet with the transformation property f(aτ+bcτ+d)=(cτ+d) kf(τ) f(\frac{a\tau + b}{c \tau + d}) = (c\tau + d)^k f(\tau) for all elements of SL 2()SL_2(\mathbb{Z}),

\bullet such that only finitely many terms for negative powers of kk are nonvanishing,

\bullet and such that all coefficients are integers.

All weak integral modular forms are combinations of

\bullet the discriminant Δ=q n=1 (1q n) 24 \Delta = q \prod_{n=1}^\infty (1-q^n)^{24} and

\bullet the two Eisenstein series c 4 = 1+240 k>0σ 3(k)q k c 6 = 1504 k>0σ 5(k)q k \array{ c_4 &=& 1 + 240 \sum_{k \gt 0} \sigma_3(k) q^k \\ c_6 &=& 1 - 504 \sum_{k \gt 0} \sigma_5(k) q^k } with σ r(k)= d|kdr\sigma_r(k) = \sum_{d|k}d r. In fact, the ring of weak integral modular forms, MF *MF_* is MF *=[c 4,c 6,Δ,Δ 1]/(c 4 3c 6 2(12) 3Δ). MF_* = \mathbb{Z}[c_4,c_6,\Delta,\Delta^{-1}]/(c_4^3 - c_6^2 - (12)^3\Delta) \,.

Bottom line. The index of an ordinary Dirac operator takes values in the integers, which is the KK-theory of a point. The index of a Dirac operator on loop space takes values in the ring MF *MF_* of weak integral modular forms. This lives in something like the elliptic cohomology of a point.

In fact, for this reason the generalized cohomology appearing here comes from an E E_\infty-spectrum called tmf\mathrm{tmf}, for “topological modular forms”. This is discussed in the next subsection.

5) What is elliptic cohomology?

See G. Segal, Elliptic Cohomology, or

Jacob Lurie
A survey of elliptic cohomology
(pdf)
section 1

Let HH be some multiplicative generalized cohomology (i.e. one coming from an E E_\infty-ring spectrum). It turns out (Lurie, p. 3) that the graded ring associated by HH to the space K(,2)P PU(H)BU(1) K(\mathbb{Z},2) \simeq \mathbb{C}P^\infty \simeq PU(H) \simeq BU(1) is of particular importance.

Since BU(1)BU(1) is the classifying space for U(1)U(1)-bundles, the ordinary cohomology group H (BU(1),)H^\bullet(BU(1),\mathbb{Z}) is the ring freely generated, over \mathbb{Z}, by the first Chern class c 1c_1 of the universal line bundle U(H)PU(H)U(H) \to PU(H).

It is a triviality, but of central importance for what follows, that under the tensor product of two line bundles L 1XL_1 \to X and L 2XL_2\to X, their Chern classes simply add up, in ordinary second cohomology c 1(L 1 XL 2)=c 1(L 1)+c 2(L 1). c_1(L_1 \otimes_X L_2) = c_1(L_1) + c_2(L_1) \,.

This particular property turns out to be modified for other generalized cohomology theories. They will have other group laws, different from the simple one on the right hand side of the above formula.

So let HH be any generalized cohomology theory. One finds that its value on BU(1)BU(1) is always a power series ring H (BU(1))=H (pt)[[t]] H^\bullet(BU(1)) = H^\bullet(\mathrm{pt})[\![t]\!] over a single generator tt of degree 2 with coefficients in the cohomology ring of the point. Recall that tt is hence itself a (class of a) map t:BU(1)E 2, t : BU(1) \to E_2 \,, if EE is the spectrum representing the generalized cohomology.

Given any complex line bundle on some space XX, classified by the class of the map ϕ:XBU(1)\phi : X \to BU(1), we can pull back the generator tt along this map to obtain an element of H (X)H^\bullet(X): ϕ *t:XϕBU(1)tE 2. \phi^*t : X \overset{\phi}{\to} BU(1) \overset{t}{\to} E_2 \,. Notice that for ordinary cohomology this is precisely the pullback of the universal first Chern class along ϕ\phi to XX, hence the Chern class c 1(L)=ϕ *tc_1(L) = \phi^* t of our complex line bundle on XX.

For generalized cohomologies ϕ *t\phi^*t is no longer the Chern class of our line bundle - so let’s call it a generalized Chern class.

The whole point of this exercise is that we can now (in principle) compute how these generalized Chern classes of complex line bundles behave under the tensor product operation of line bundles.

In general, they will, unlike the ordinary Chern class, not just add up.

For instance, for K-theory one finds the following group law c 1(L 1 XL 2)=c 1(L 1)+c 1(L 2)+c 1(L 1)c 1(L 2). c_1(L_1 \otimes_X L_2) = c_1(L_1) + c_1(L_2) + c_1(L_1) c_1(L_2) \,. Most generally, one finds c 1(L 1 XL 2)=f(c 1(L 1),c 2(L 2)), c_1\left(L_1 \otimes_X L_2\right) = f\left( c_1(L_1),\; c_2(L_2) \right) \,, where f(t 1,t 2)H (pt)[[t 1,t 2]] f(t_1,t_2) \in H^\bullet(\mathrm{pt})[\![t_1,t_2]\!] is any formal power series with coefficients in the cohomology ring of the point, satisfying the following three axioms

\bullet unit law: f(v,0)=f(0,v)=vf(v,0) = f(0,v) = v

\bullet commutativity: f(u,v)=f(v,u)f(u,v) = f(v,u)

\bullet associativity: f(u,f(v,w))=f(f(u,v),w)f(u,f(v,w)) = f(f(u,v),w).

Such a power series is called a commutative, 1-dimensional formal group law over the ring H (pt)H^\bullet(\mathrm{pt}).

One can go on and on about formal group laws, formal spectra, formal groups and what not, but the important point is the following.

Fact. (see Lurie, p. 7) Under suitable assumptions, over an algebraically closed field, there are only three types of 1-dimensional algebraic groups, corresponding to three types of formal group laws

0) the additive group law: xy=x+yx \cdot y = x + y

1) the multiplicative group law: xy=x+y+xyx \cdot y = x + y + xy

2) a group law defined by an elliptic curve (\to).

Hence we finally have the, now obvious

Definition. (roughly) An elliptic cohomology is any generalized cohomology whose formal group law (which governs the generalized Chern classes of tensor products of complex line bundles) is given by an elliptic curve.

So there are many elliptic cohomology theories. One can do something like taking the direct limit over all of these. The result is represented by an E E_\infty-spectrum called tmf\mathrm{tmf}. Its value over a point is the ring of weak integral modular forms, that arise as the partition function of heterotic strings.

6) How is ellitpic cohomology realized geometrically?

Above we have given the definition of elliptic cohomology. But we would also like to have a nice geometric realization of it. Like we can understand K-theory geometrically as being about classes of vector bundles.

In fact, we can alternatively understand the K-theory spectrum as the space of supersymmetric quantum mechanical systems. That’s the content of the following “warmup exercise”.

It suggests, that the tmf\mathrm{tmf}-spectrum for elliptic cohomologies is similarly related to the moduli space of 2-dimensional superconformal field theories.

6a) Warmup: how is the landscape of superpoint theories equal to the K-theory spectrum?

First some sources.

Apart from the the above cited texts by Stolz-Teichner-Markert, information on part a) can be found in a collection of lecture notes from a winter school on elliptic objects.

Sarah Massberg,
The definition of 1-dimensional EFT’s a la Stolz/Teichner.
(pdf),

Juan Wang
From Euclidean Field Theory to K-theory, I
pdf,

Benedikt Plitt
Fredholm Operators and K-theory
(pdf)

and

Moritz Wiethaupt
From Euclidean Field Theory to K-theory, II
(pdf).


Before saying anything about supersymmetric quantum mechanics, consider this standard example, which, in string speak is a point in the geometric phase of the superparticle landscape.

Example. (The Euler characteristic.)

Let XX be a compact Riemannian manifold with exterior bundle Λ (T *X)\Lambda^\bullet(T^*X). Let HH be the 2\mathbb{Z}_2-graded (actually even \mathbb{Z}-graded) Hilbert space of square integrable sections of this bundle with respect to the Hodge scalar product v|w= Xvw. \langle v | w \rangle = \int_X v \wedge \star w \,. From the exterior derivative d:HHd : H \to H and its adjoint d *:HHd^* : H \to H we construct the Dirac operator D:=d+d *. D := d + d^* \,. Time-independent euclidean quantum mechanics of a free (“RR-sector”) superparticle on XX could be described by taking D 2D^2 to be the corresponding Hamiltonian.

The partition function of this superparticle, for the worldline being a closed path of time length tt, would be the supertrace Z(t)=str H(exp(tD 2))=tr((1) Nexp(tD 2)), Z(t) = \mathrm{str}_H \left( \exp\left( - t\, D^2 \right) \right) = \mathrm{tr} \left( (-1)^N \exp\left( - t\, D^2 \right) \right) \,, where NN is the operator which measures the degree of a differential form. Due to “supersymmetry” (namely due to the fact that the (euclidean) time-translation operator D 2D^2 is the square of an odd-graded operator) all non-vanishing eigenvalues of D 2D^2 occur in pairs, one corresponding to an eigenspace in even, the other to one in odd degree.

Therefore, the trace reduces to that over 0-eigenstates, which are precisely the harmonic forms, and we find that the partition function is the Euler characteristic (\to) χ(X)\chi(X) of XX Z(t)= i(1) ib i=χ(X), Z(t) = \sum_i (-1)^i b_i = \chi(X) \,, where b ib_i are the Betti numbers (\to).

χ(X)\chi(X) is the index (\to) of DD, being the difference between the dimension dim +\mathrm{dim}_+ of the space of “bosonic” (forms of even degree) and dim \mathrm{dim}_-, that of “fermionic” 0-modes (forms of odd degree).

A sophisticated way of thinking about this index is as an element of the K-theory (\to) over a point, namely the class of a virtual vector bundle over a point, hence a virtual vector space. ( dim +, dim ) (\mathbb{R}^{\mathrm{dim}_+},\mathbb{R}^{\mathrm{dim}_-})


This is one of the standard examples of supersymmetric quantum mechanics. We want to formalize the general notion of supersymmetric quantum mechanics.

Definition. Denote by 1Cob1\mathrm{Cob} the monoidal category of oriented 1-dimensional Riemannian cobordisms and by Hilb\mathrm{Hilb} the category of real seperable Hilbert spaces with bounded operators between them. Then a euclidean 1-dimensional QFT, also known as euclidean quantum mechanics (or as statistical mechanics, for that matter), is a functor 1dQFT:1CobHilb, 1d\mathrm{QFT} : 1\mathrm{Cob} \to \mathrm{Hilb} \,, which respects tensor products, orientation involution and the adjunctions on 1Cob1\mathrm{Cob}.

Any such functor is determined by

a) a choice of Hilbert space (“space of states”) HH

b) a choice of, possibly unbounded, self-adjoint operator Δ\Delta, the “Hamiltonian” 1dQFT (H,Δ)(t)=Hexp(tΔ)H. 1d\mathrm{QFT}_{(H,\Delta)} ( \bullet \overset{t}{\to} \bullet) \;=\; H \overset{\exp(-t \Delta)}{\to} H \,.

Now we want to formulate supersymmetric quantum mechanics in an analogous manner. We will define a category 1sCob1\mathrm{sCob} of 1-dimensional super-Riemannian cobordisms and take Hilb\mathrm{Hilb} to be the category of 2\mathbb{Z}_2-graded seperable (real) Hilbert spaces. Once 1sCob1\mathrm{sCob} is available, we will make the obvious

Definition. A euclidean 1-dimensional super QFT, also known as euclidean supersymmetric quantum mechanics, is a functor 1dsQFT:1sCobHilb, 1d\mathrm{sQFT} : 1\mathrm{sCob} \to \mathrm{Hilb} \,, which respects tensor products, orientation involution and the adjunctions on 1sCob1\mathrm{sCob}.

Essentially, every such functor is determined by specifying a 2\mathbb{Z}_2-graded Hilbert space HH, and a self-adjoint odd-graded operator DD on HH, the supercharge, whose square plays the role of the Hamiltonian above.

Given that the space of Fredholm operators is a classifying space for K-theory (\to) we may expect that the space of all 1dsQFT1d\mathrm{sQFT}s is somehow related to K-theory. If we choose our definitions reasonably, then the theorem will say that both spaces are in fact homotopy equivalent.

In order to get there, we first need the category of 1-dimensional super-Riemannian cobordisms.

Definition. The category 1dsCob1d\mathrm{sCob} of D=1D=1, N=1N=1 super-Riemannian cobordisms has a single object, the superpoint = 0|1\bullet = \mathbb{R}^{0|1}, and a superspace of morphisms Hom(,) + 1|1. \mathrm{Hom}(\bullet,\bullet) \simeq \mathbb{R}_+^{1|1} \,. Composition of morphisms is defined in terms of SS-points (see below) as Hom(,)×Hom(,) Hom(,) ( + 1|1× + 1|1)(S) + 1|1(S) (t 1,θ 1,t 2,θ 2) (t 1+t 2+θ 1θ 2,θ 1+θ 2). \array{ \mathrm{Hom}(\bullet,\bullet) \times \mathrm{Hom}(\bullet,\bullet) &\to& \mathrm{Hom}(\bullet,\bullet) \\ (\mathbb{R}_+^{1|1} \times \mathbb{R}_+^{1|1})(S) &\to& \mathbb{R}_+^{1|1}(S) \\ (t_1,\theta_1,t_2,\theta_2) &\mapsto& (t_1+t_2+ \theta_1\theta_2, \, \theta_1 + \theta_2) } \,.

Here we are working with superspaces in terms of sheaves on the category of 2\mathbb{Z}_2-graded ringed spaces (\to).

Briefly, this means the following:

Definition. 1) For us, a representable supermanifold is a manifold XX equipped with a sheaf O XO_X of 2\mathbb{Z}_2-graded rings, equipped with a epimorphism onto the sheaf of continuous functions C (X)C^\infty(X) on XX.

2)In particular, we denote by n|m\mathbb{R}^{n|m} the supermanifold whose underlying manifold is n\mathbb{R}^n and whose sheaf of graded rings has as global sections the ring O n|m( n):=C (C)Λ m O_{\mathbb{R}^{n|m}(\mathbb{R}^n)} := C^\infty(C)\otimes \Lambda^\bullet \mathbb{R}^m of smooth functions with values in the exterior algebra generated by mm linear independent differential forms (“Grassmann coordinates”).

2a) We need in particular the super half-line + 1|1\mathbb{R}_+^{1|1}, whose underlying manifold is +=[0,)\mathbb{R}_+ = [0,\infty) and whose ring of global sections is C ( +)Λ C ( +)θC ( +). C^{\infty}(\mathbb{R}_+)\otimes \Lambda^\bullet \mathbb{R} \simeq C^{\infty}(\mathbb{R}_+) \;\oplus\; \langle \theta \rangle \otimes C^{\infty}(\mathbb{R}_+) \,.

3) The category of representable supermanifolds repSMfld\mathrm{repSMfld} is, in the obvious way, a subcategory of that of ringed spaces.

4) A general supermanifold is a sheaf of sets (thinking of presheaves suffices for our present purposes) on repSMfld\mathrm{repSMfld}, hence a contravariant functor SMfldSets. \mathrm{SMfld} \to \mathrm{Sets} \,. These functors form the category of general supermanifolds, SMfld\mathrm{SMfld}.

5) By the Yoneda embedding (\to), every representable supermanifold is a general supermanifold repSMfld SMfld (X,O x) Hom repSMfld(,(X,O X)) \array{ \mathrm{repSMfld} &\to& \mathrm{SMfld} \\ (X,O_x) &\mapsto& \mathrm{Hom}_{\mathrm{repSMfld}}(--,(X,O_X)) } which associates to each representable supermanifold S=(|S|,O S)S = (|S|,O_S) the set of SS-points X(S):=Hom repSMfld(S,X). X(S) := \mathrm{Hom}_{\mathrm{repSMfld}}(S,X) \,.

We reobtain the physicist’s way of thinking about supermanifolds in terms of Grassmann coordinates as follows:

Fact. The SS-points of n|m\mathbb{R}^{n|m} are in bijection with ordered tuples (t 1,t 2,,t n,θ 1,θ 2,,θ m) (t_1,t_2,\cdots, t_n,\; \theta_1,\theta_2,\cdots, \theta_m) of nn even and mm odd global sections of SS.

In particular, if SS is the superpoint, S= 0|1S = \mathbb{R}^{0|1}, then an SS-point of n|m\mathbb{R}^{n|m} is precisely one ordinary point in n\mathbb{R}^n together with precisely one 1-form on ( m) *(\mathbb{R}^{m})^*.

Remark. Stolz and Teichner in their work, and in particular Elke Markert in her thesis (based on ideas by H. Hohnhold), go through the trouble of deriving the above stated super-Hom\mathrm{Hom}-spaces of the category 1dsCob1d\mathrm{sCob}, including their composition law, from “first principles”, i.e. from just demanding the super-Riemann property and implementing it consistently in the arrow-theoretic setup. When the general-abstract-nonsense-dust has settled, what remains is the simple composition law stated above, known otherwise as ordinary super-translation.


As a nice example for these constructions, and also because it is central for the following constructions, consider the following

Observation. Every 2\mathbb{Z}_2-graded vector space V=V evV oddV = V^\mathrm{ev}\oplus V^\mathrm{odd} gives rise to a general supermanifold by defining its SS points as (S,O S)(VO S(S)) ev:=V evO S(S) evV oddO S(S) odd. (S,O_S) \mapsto (V \otimes O_S(S))^\mathrm{ev} := V^\mathrm{ev}\otimes O_S(S)^\mathrm{ev} \oplus V^\mathrm{odd}\otimes O_S(S)^\mathrm{odd} \,. (In fact, the supermanifold defined this way is representable (\to), but that need not concern us here.)

It follows, that we can regard the category of 2\mathbb{Z}_2-graded Hilbert spaces as a category whose objects and Hom\mathrm{Hom}-spaces are in fact supermanifolds (albeit possibly infinite dimensional ones). In particular, we have

Observation. Given a 2\mathbb{Z}_2-graded Hilbert space, the vector space Hom(H,H)\mathrm{Hom}(H,H) of bounded graded operators on that Hilbert space is a supermanifold whose superpoints (i.e SS-points for S= 0|1S = \mathbb{R}^{0|1}) are pairs (A,B)(A,B) consisting of one even graded and one odd graded operator.

This, then, allows us to make precise what we mean by a functor 1dsQFT:1CobHilb, 1d\mathrm{sQFT} : 1\mathrm{Cob} \to \mathrm{Hilb} \,, namely a functor on categories enriched over supermanifolds. This simply means that on morphisms it is a map of supermanifolds 1dsQFT:Hom 1dscob(,) 1|1 Hom Hilb(H,H) 1|1(S) (End HS) ev (t,θ) A(t)+θB(t), \array{ 1d\mathrm{sQFT} : \mathrm{Hom}_{1d\mathrm{scob}}(\bullet,\bullet) \simeq \mathbb{R}^{1|1} &\to& \mathrm{Hom}_\mathrm{Hilb}(H,H) \\ \mathbb{R}^{1|1}(S) &\to& (\mathrm{End}_H\otimes S)^\mathrm{ev} \\ (t,\theta) &\mapsto& A(t) + \theta B(t) \,, } where A(t)A(t) is an even-graded and B(t)B(t) and odd-graded endomorphism of HH.

Being maps of supermanifolds, these functors form a supermanifold themselves. Hence it makes sense to talk about the topology of the space of 1-dimensional super QFTs.


With this supermachinery in hand, we can now state the theorem that we are after, together with the two lemmas that prove it.

Theorem.([Stolz/Teichner] (simplified version)) The space of N=1N=1 supersymmetric 1-dimensional euclidean QFTs is homotopy equivalent to KO 0\mathrm{KO}_0, the 0th space in the Ω\Omega-spectrum of real K-theory.

Remark.

i) For the purposes of this presentation, I have suppressed a certain additional “Clifford grading” that Stolz-Teichner put on the super cobordisms. If one includes this, one finds spaces of supersymmetric 1D QFTs for all grades nn\in \mathbb{N}, and that these are homotopy equivalent to the nn-th space of the Ω\Omega-spectrum of KO\mathrm{KO}-theory.

(Personally, I should remark that, while introducing this additional grading leads to the desired result, it is otherwise not naturally motivated and looks a little ad hoc to me. It is slightly better motivated in the 2-dimensional case (superstrings) that we are ultimately interested in, since there it is related to the Pfaffian line bundles coming from the string’s path integral.

In the end, however, it seem to me that we should figure out how this grading arises naturally. I think I heard rumours that it would be implied by simply using higher supersymmetry on the worldvolume. But typing this, I realize that I may or may not have dreamed this rumour.)

ii) With just slightly more effort one also gets complex K-theory. But I will not discuss that. See p. 38 in Stolz-Teichner.


The above theorem follows from two lemmas.


Lemma 1. ([Stolz-Teichner, prop. 3.2.6 & lemma 3.2.14])

a) 1-dimensional supersymmetric euclidean field theories are in bijection with maps of supermanifolds from + 1|1\mathbb{R}_+^{1|1} to the superspace of self-adjoint graded Hilbert-Schmidt operators [1dsCob,Hilb]Hom( 1|1,HS sa(H)), [1d\mathrm{sCob},\mathrm{Hilb}] \simeq \mathrm{Hom}( \mathbb{R}^{1|1},\; HS^\mathrm{sa}(H) ) \,, which respect the super semigroup structure on + 1|1\mathbb{R}_+^{1|1}.

b) The latter space is homotopy equivalent to that of graded C *C^*-algebra morphisms from continuous functions to compact operators Hom( 1|1,HS sa(H))Hom C *(C 0(),K(H)). \mathrm{Hom}( \mathbb{R}^{1|1},\; HS^\mathrm{sa}(H) ) \simeq \mathrm{Hom}_{C^*}(C_0(\mathbb{R}), K(H)) \,.

Explanation and sketch of proof:

The first statement expresses essentially just functoriality. Under composition, the morphisms of 1dsCob1d\mathrm{sCob} realize the supertranslation semigroup + 1|1\mathbb{R}_+^{1|1}, and our sQFT\mathrm{sQFT} has to respect that. In fact, the QFT functor is realized by the obvious susy generalization of the bosonic propagator: given any suitable odd-graded operator DD (the supercharge, or Dirac operator), the QFT functor acts on SS-points of supercobordisms as 1dsQFT : Hom 1dsCob(,) Hilb 1|1(S) (End(H))(S) (t,θ) exp(tD 2)+θDexp(tD 2). \array{ 1d\mathrm{sQFT} &:& \mathrm{Hom}_{1d\mathrm{sCob}}(\bullet,\bullet) &\to& \mathrm{Hilb} \\ && \mathbb{R}^{1|1}(S) &\to& (\mathrm{End}(H))(S) \\ && (t,\theta) &\mapsto& \exp(-t D^2) + \theta D \exp(-t D^2) \,. } Notice how the right hand side in the last line is indeed a combination of even graded operator times even sections of SS plus an odd graded operator times an odd graded section of SS, as it should be.

It is an elementary exercise in supersymmetry reasoning to check that this assignment is indeed a supergroup homomorphism.

The second statement boils down essentially to a slight refinement of this situation. We take the algebra C 0()C_0(\mathbb{R}) of continuous functions on the real line to be graded in the sense that even functions (satisfying f(x)=f(x)f(x) = f(-x)) are assigned even grade and odd functions are assigned odd grade. Regarded as a superalgebra this way, we can, as above, define a supergroup homomorphisms 1|1(S) (C 0())(S) (t,θ) exp(tx 2)+θxexp(tx 2). \array{ \mathbb{R}^{1|1}(S) &\to& (C_0(\mathbb{R}))(S) \\ (t,\theta) &\mapsto& \exp(-t x^2) + \theta x \exp(-t x^2) \,. } Notice how the right hand side is indeed an even function plus an odd function times an odd section.

In conclusion, this tells us that the space of supersymmetric quantum mechanics is homotopy equivalent to the space of algebra homomorphisms from continuous functions on the real line to compact operators. This space, however, is well known to be homotopy equivalent to the classifying space of (real) K-theory.


Lemma 3.([Higson-Guentner]) The space Hom C *(C 0(),K(H)) \mathrm{Hom}_{C^*}(C_0(\mathbb{R}), K(H)) is homotpy equivalent to the 0th space in the Ω\Omega-spectrum of real K-theory.


Now some examples.

The most accessible examples, like the one already given above, come from the geometric phase of the theory, where our superparticles are goverened by operators that act on Hilbert spaces of sections of spinor bundles over some manifolds (as opposed to being completely abstractly defined odd graded operators on some abstract Hilbert space).

Example. Given a vector bundle with spin structure, we can define the Thom class […]


6b) How does one expect the landscape of superstring theories to be equal to the spectrum of elliptic cohomology?

Unfinished notes.

The entire point of the above exercise of reformulating K-theory in terms of supersymmetric quantum mechanics is that this lends itself to the expected generalization from point particles to strings.

\to

Expected Definition. A 2-dimensional superconformal field theory is a 2-functor 2dSCFT:scP 22Hilb 2d\mathrm{SCFT} : scP_2 \to 2\mathrm{Hilb} from superconformal-2-paths to graded 2-Hilbert spaces.

[…]

The heterotic string partition function and the Witten genus.

[…]

Now examples.

As before for the superparticle, we are interested in examples from the geometric phase, where our supercharge comes from an operator that acts on sections of some stringy generalization of a vector bundle with spin structure. We expect there to be stringy analogs of the Thom class and the Euler class in this sense.

Definition. A 2-vector bundle with connection is a 2-functor from 2-paths to Mod C\mathrm{Mod}_C, for CC some monoidal category.

Example.([L2B]) (surface transport in abelian gerbe (line 2-bundle))

\bullet Points are sent to the category ModA \multiscripts{_A}{\mathrm{Mod}}{}, for AA morita equivalent to \mathbb{C}. Hence AA is an algebra of compact operators and we get a bundle of compact operator - a PU(H)\mathrm{PU}(H)-bundle.

\bullet Morphisms are sent to bimodules.

\bullet 2-Morphisms are sent to bimodule homomorphisms.

Example.(Stolz-Teichner) (surface transport in string gerbe (Ω^Spin\hat \Omega \mathrm{Spin}-2-bundle \to))

\bullet Points are sent to the category ModA \multiscripts{_A}{\mathrm{Mod}}{}, for AA a type III 1\text{III}_1 von Neumann algebra factor.

\bullet Morphisms are sent to bimodules.

\bullet 2-Morphisms are sent to bimodule homomorphisms.

[…]

Posted at June 28, 2006 5:07 PM UTC

TrackBack URL for this Entry:   https://golem.ph.utexas.edu/cgi-bin/MT-3.0/dxy-tb.fcgi/862

17 Comments & 9 Trackbacks

Splitting hairs

A minor notational quibble.

If you want the elliptic genus to be power series in qq (and not in q¯\overline{q}), then you should say: ind WittenG¯ =str(q Lq¯ L¯) =str(q L| ker(L¯)) = kq ksdim(ker(L¯)H k) \begin{aligned} ind_{\text{Witten}} \overline{G} &= str\left(q^L\overline{q}^{\overline{L}}\right)\\ &=str\left(q^{L}|_{ker(\overline{L})}\right)\\ &=\sum_{k\in\mathbb{Z}}q^k sdim\left(ker(\overline{L})\cap H^k\right) \end{aligned}

That would match up with your subsequent text (and with common convention).

Posted by: Jacques Distler on July 1, 2006 5:59 PM | Permalink | PGP Sig | Reply to this

Re: Splitting hairs

A minor notational quibble.

Thanks. I’ll correct that.

Posted by: urs on July 2, 2006 12:57 PM | Permalink | Reply to this

Re: Seminar on 2-Vector Bundles and Elliptic Cohomology, V

It would be great if you could put this stuff on your blog. But if you don’t feel like it, that’s no big deal.

You wrote in private email:

Let’s see:

Let OO be the spectrum which represents the cohomology theory called “complex cobordisms” (or maybe spin cobordisms, see below).

I think the spectrum for complex cobordism theory is usually called MUMU, but it doesn’t really matter…

Let KUKU be the spectrum which represents K-theory.

Let pt be the point.

Then

[pt,O][pt,O] is the cobordism ring (whose sum is the disjoint union of spin manifolds, product is the cartesian product, all up to spin cobordisms )

Or complex cobordisms, if that’s what you’re doing.

[pt,KU][pt,KU] is the ring of integers.

There is a ring homomorphism

(1)[pt,O][pt,KU] [pt,O] \to [pt,KU]

given by composing with the unique morphism

(2)OKU O \to KU

of spectra. As far as I understand.

If what you’re calling O is the spectrum for complex cobordism theory, there’s a unique-up-to-homotopy map of spectra

(3)OKU O \to KU

preserving the complex orientation - a structure I explain here.

This is because complex cobordism is the universal complex oriented cobordism theory.

So, yeah, you’re basically right.

This morphism

(4)[pt,O][pt,KU] [pt,O] \to [pt,KU]

is, by definition of genus, a genus.

Okay, good.

I guess it must be the A^\hat A-genus?? At least if we take our cobordisms to be spin?

I’m getting mixed up about the relation between complex cobordism vs. spin cobordism. For any increasing sequence of groups G(n)G(n) sitting in O(n)O(n), we can take the limit of these groups and get a group GG. We define a “GG-cobordism” to be a cobordism whose stable normal bundle has been reduced to a GG-bundle. By this I mean that we take our cobordism and embed it in a high-dimensional Euclidean space, take its normal bundle (which is a vector bundle of some dimension nn), and reduce the structure group from O(n)O(n) to G(n)G(n). Then we take the limit as nn \to \infty… the limit of the normal bundles is called the “stable normal bundle” and it becomes a GG-bundle. I’m being rough here but I hope you get the idea.

Whenever we have this situation we get a spectrum for GG-cobordism theory called “MGMG”. This spectrum is built in a functorial way (the “Thom construction”) from GG.

So, the spectrum for complex cobordism theory is MUMU while that for spin cobordism theory is MSpinM\mathrm{Spin}. To understand their relation, the key thing we need is a homomorphism between UU and Spin\mathrm{Spin}! There must be one… which way does it go? More precisely, we need one between U(n)U(n) and Spin(2n)\mathrm{Spin}(2n) - because U(n)U(n) really acts not on R nR^n but R 2nR^{2n}.

I should know which way this homomorphism goes, but I’m blanking out! I believe you used to talk a lot about Kähler manifolds and spinors, so you should know this stuff.

Hmm - looking around I see this written by Peter Woit.

“…if MM is Kähler, the frame bundle can be chosen to be a U(n)U(n) bundle, but such an MM will often not have a spin structure…”

He seems to say there’s no good homomorphism U(n)Spin(2n)U(n) \to \mathrm{Spin}(2n). It’s possible that something weaker, like U(n)Spin(2n+k)U(n) \to \mathrm{Spin}(2n+k), would be good enough.

Anyway, I’m a bit confused, but I hope this helps you start pondering the relation between MUMU and MSpinM\mathrm{Spin}.

I guess Atiyah-Singer implies hence that we can compute the morphism

(5)[pt,O][pt,KU] [pt,O] \to [pt,KU]

by

- choosing, for every spin manifold XX (regarded as an element in [pt,O][pt,O], since that ring happens to be a ring of spin manifolds) a Dirac operator on some spinor bundle on XX.

This sounds right - but I bet we want not just any spinor bundle on XX, but one determined by the spin structure. This should be unique up to isomorphism, so there’s no real choice involved.

Of course we need to pick a Dirac operator (e.g. by picking a Riemannian metric), but the index theorem says the index is independent of that choice.

- computing the index of that operator and interpreting it as an element in [pt,KO][pt,KO] .

Please correct me if this sounds wrong.

This stuff sounds right, modulo the issue of “spin cobordism versus complex cobordism”.

Now, in my original statement, I tried to relate this to the observation that Stolz-Teichner-Markert base their work on, namely the theorem that the space KUKU is homotopy equivalent to the space of 1d susy field theories, which is essentially the space of “abstract Dirac operators”.

Hmm… you’re probably on the right track, but I’m worrying about various things…

I’ll think about it. I should learn Atiyah-Singer in more detail. Also K-theory. Also lots of other things…

Yes, but you seem to be doing great.

Looking around I found an old conversation on the web, which may help a bit.

Best, jb

From: Dev Sinha Subject: Re: Elliptic cohomology Date: Wed, 17 May 2000 14:07:03 -0400 Newsgroups: sci.math.research Summary: [missing]

Hi John,

I’m not an expert in this subject, and I won’t fully answer your question, but I think I have a couple things to say which might help.

First is to think of the Index Theorem (“the case down the chromatic ladder from elliptic cohom”), which applies to any smooth, compact manifold. While the theorem applies to any compact mfld with an elliptic operator, the cohomology theory which plays a vital role is K-theory, which is complex-oriented. The topological index is best defined via K-theory. The trick is that for any X, TX is naturally a complex manifold.

A second thing to note is that sometimes complex-oriented theories have “companions” which are not complex-oriented. The prototypical example is real K-theory, which is not complex-oriented but can be built from K-theory if one pays attention to the Z/2 action coming from conjugation. The versions of elliptic cohom constructed by Hopkins and Miller, which are now being called rings of topological modular forms and one of which is the proper target for the Witten genus, follow this pattern.

In a fairly unrelated matter (of which I am reminded because we are talking about orientations), I’d like to share a bit of knowledge which I learned embarrassingly recently: We all know that in order for a manifold to satisfy Poincare duality for integral coefficients, it must be oriented. What about PD for other cohomology theories - what conditions are needed? Algebraic topologists know well that what you need are Thom isomorphisms for the normal bundle of the manifold in question, but can one get more specific? Might usual orientability be enough? When the cohomology theory in question is real K-theory, there is a beautiful answer: there is PD for KO-theory if and only if the manifold in question is spin. According to a friend who studies elliptic operators on manifolds, the KO-theory orientation class is, in an appropriate setting, represented by the Dirac operator. - I know this is old hat to some, but I was happy to learn it recently.

-Dev

From: baez@math.ucr.edu (John Baez) Subject: Re: Elliptic cohomology Date: 18 May 2000 23:42:35 GMT Newsgroups: sci.math.research Summary: [missing]

In article Pine.GSO.4.10.10005171335390.5268-100000@newton.math.brown.edu, Dev Sinha wrote:

>I’m not an expert in this subject, and I won’t fully answer your question, >but I think I have a couple things to say which might help.

Your comments were very helpful, as were those of Mark Hovey, who assured me that after we invert the prime 2, oriented cobordism theory and spin cobordism theory become the same, while complex cobordism becomes *almost* the same. More precisely, we’ve got a spectrum MU for complex cobordism theory, and we’ve got a spectrum MSO for oriented cobordism, but localizing away from 2 MU becomes isomorphic to MSO wedged with the double suspension of MSO. What this means is that lots of stuff can be done either for oriented cobordism, spin cobordism, or complex cobordism, without it making much difference - as long as we don’t care about 2-torsion. Apparently the authors who were confusing the heck out of me by using oriented cobordism to study elliptic cohomology were actually trying to be nice, they were afraid I’d be scared of complex cobordism!

(For years I had an instinctive fight-flight reaction whenever anyone did something algebraic like “localization” to a spectrum, so I sympathize with anyone who feels that way, but I think that in the last couple of months I’ve overcome that instinct, so now I’m gonna talk like I’ve understood this stuff since birth.)

Posted by: John Baez on July 2, 2006 1:18 PM | Permalink | Reply to this

Re: Seminar on 2-Vector Bundles and Elliptic Cohomology, V

There is a natural map from U(n) to SO(2n), because every complex vector space is naturally oriented when thought of as a real vector space. There is no natural map from U(n) to Spin(2n). But if we take the universal cover of the map from U(n) to SO(2n), you get a map from

SU(n) to Spin(2n).

This induces a map of cobordism theories from MSU to MSpin. MSpin is the universal KO-orientable cobordism theory. That is, there is a natural genus MSpin —> KO and a manifold has a KO-orientation (I believe on its stable normal bundle) if and only if it is Spin.

There is of course a map from MSU to MU.
Rationally, on homotopy groups, this map is the inclusion of the polynomial ring on the Chern classes c_2 and higher into the polynomial ring on all Chern classes. If we invert 2, it is still a monomorphism but it is not so simple. And if we don’t invert 2, there is a lot of 2-torsion in MSU that gets killed in going to MU.

Kahler manifolds are in fact SU-manifolds, not just U-manifolds.

Posted by: Mark Hovey on July 7, 2006 4:31 PM | Permalink | Reply to this

Re: Seminar on 2-Vector Bundles and Elliptic Cohomology, V

[…] there is a natural genus MSpinKO\mathrm{MSpin} \to KO […]

Does this genus have a name? Is it the A^\hat A-genus?

Posted by: urs on July 9, 2006 3:31 PM | Permalink | Reply to this

Re: Seminar on 2-Vector Bundles and Elliptic Cohomology, V

The genus MSpin –> KO is sometimes called the Atiyah genus or the Atiyah-Bott-Shapiro genus. In dimensions 4k it is the A-hat genus (when KO_4k is identified with the integers). In dimensions 8k+1 and 8k+2 it is a mod 2 invariant, so it is not the A-hat invariant. It is the index of the Dirac operator in a suitable sense that I can’t remember.
Mark

Posted by: Mark Hovey on July 12, 2006 2:36 PM | Permalink | Reply to this

SRSs

Another, less nitpicky, point. Something that hasn’t made an appearance yet, but which surely must, at some point, make an appearance in your discussion, as you go up in dimension (and start to lose us mere mortal in the forest of higher category theory).

Super-Riemann surfaces (the bordisms that become of interest, here) are not just (1|1)(1|1)-dimensional complex supermanifolds. They are (1|1)(1|1)-dimensional complex supermanifolds, equiped with an integrable odd distribution — a (0|1)(0|1)-dimensional subbundle, 𝒟T\mathcal{D}\subset T, of the holomorphic tangent bundle, such that 𝒟\mathcal{D} and {𝒟,𝒟}\{\mathcal{D},\mathcal{D}\} span TT.

This is much more specialized than a general (1|1)(1|1)-dimensional complex supermanifold.

Posted by: Jacques Distler on July 2, 2006 9:23 PM | Permalink | PGP Sig | Reply to this

Re: SRSs

This is much more specialized than a general (1∣1)-dimensional complex supermanifold.

We need a spin structure on our surfaces. I did mention some aspects of that in a previous entry. I should try to say more about this some day.

Posted by: urs on July 3, 2006 2:55 PM | Permalink | Reply to this

Re: SRSs

Urs wrote:

We need a spin structure on our surfaces.

Is that what Jacques was referring to when he referred to:

(1|1)-dimensional complex supermanifolds, equipped with an integrable odd distribution - a (0|1)-dimensional subbundle X of the holomorphic tangent bundle T, such that X and {X,X} span T.

Is this integrable odd distribution secretly the same thing as a spin structure?? If not, what is its role?

Posted by: John Baez on July 12, 2006 6:20 AM | Permalink | Reply to this

Re: SRSs

an integrable odd distribution - a (0|1)(0|1)-dimensional subbundle XX of the holomorphic tangent bundle TT, such that XX and {X,X}\{X,X\} span TT.

what is its role?

Jacques will correct me, but I think what he wrote is just a sophisticated way of talking about the holomorphic superderivative.

The point is, we want our surface not just to be super, but superconformal.

Ordinary conformal structures on 2d surfaces can be expressed by demanding that on double intersections the vector field z\partial_z transforms homogeneously.

This is superized by demanding that the holomorphic superdifferential

(1)D= θ+θ z D = \partial_\theta + \theta \partial_z

transforms homogeneously.

Due to D 2= zD^2 = \partial_z one finds that, locally, DD and {D,D}\{D,D\} span the holomorphic super tangent bundle.

I think what Jacques wrote is the global version of this statement.

It is a fact that N=1N=1 superconformal structure in 2D is the same as a spin structure on an ordinary Riemann surface.

This again is the same as a square root of the canonical line bundle, which is, I think, yet another way to describe the same structure.

See for instance the first few pages of

Holger Ninnemann, Deformations of Super Riemann Surfaces (pdf).

I think, but correct me if I am wrong, that we can think of the superconformal surface as the supermanifold obtained from a spinor bundle over a Riemann surface by the standard construction, where one takes the fibers to be odd graded.

Posted by: urs on July 12, 2006 3:13 PM | Permalink | Reply to this

Re: SRSs

That’s basically correct. The holomorphic superdifferential, DD is a local section of the (0|1)(0|1) line bundle I called XX. What I’m telling you is that those local sections patch together, so that D=f(z,θ)DD' = f(z,\theta) D on patch overlaps.

That is a restriction on the transition functions that makes your (1|1)(1|1) complex surface into something superconformal.

It is a worthwhile exercise to compute the most general transition functions (z,θ)=(z(z,θ),θ(z,θ))(z',\theta')= (z'(z,\theta),\theta'(z,\theta)) compatible with a superconformal structure. In general, they involve an even function, f(z,θ)f(z,\theta) and an odd function, α(z,θ)\alpha(z,\theta) (both of which may depend on parameters).

If there are no odd parameters (“supermoduli”), then f=f(z)f=f(z) and α=0\alpha=0 and your SRS is just an ordinary RS, with a choice of spin structure.

Posted by: Jacques Distler on July 12, 2006 10:53 PM | Permalink | PGP Sig | Reply to this

Re: SRSs

Bah.

I should just spell it out for you.

The most general transition functions for a (1|1)(1|1) complex surface take the form z =f(z)+β(z)θ θ =α(z)+h(z)θ\begin{aligned} z'&=f(z) + \beta(z) \theta\\ \theta' &= \alpha(z) + h(z)\theta \end{aligned} Demanding that (θ+θz)=g(z,θ)(θ+θz) \left(\frac{\partial}{\partial \theta} +\theta \frac{\partial}{\partial z}\right) = g(z,\theta) \left(\frac{\partial}{\partial \theta'} +\theta' \frac{\partial}{\partial z'}\right) yields h 2=fz,β=hα h^2 =\frac{\partial f}{\partial z},\quad \beta = - h \alpha and the transition function g(z,θ)=h(z)+θαz g(z,\theta) = h(z) + \theta \frac{\partial\alpha}{\partial z}

Note that in in passing from a general (1|1)(1|1) complex surface to a Super-Riemann surface, we’ve cut the number of independent components of the transition functions from 4 (2 even and 2 odd) to 2 (ff and α\alpha), together with a choice of square root (which went into defining hh).

Setting α0\alpha\equiv 0, the transition function for zz becomes that of an ordinary Riemann surface and the transition function for θ\theta becomes that of a spinor (a section of a square root of the canonical bundle).

Posted by: Jacques Distler on July 13, 2006 5:42 AM | Permalink | PGP Sig | Reply to this

Re: SRSs

Thanks for your reply!

Let’s see.

We want (on double intersections) the homogeneous transformation law

(1)D=gD D = g D'

hence

(2)(Dθ)θ+(Dz)z=gD. (D\theta')\frac{\partial}{\partial \theta'} + (D z') \frac{\partial}{\partial z'} = g D' \,.

For the first coefficient this implies

(3)g=Dθ g = D \theta'

which means

(4)g(z,θ)=h+θα, g(z,\theta) = h + \theta \alpha' \,,

as you said.

For the second coefficient we then get

(5)Dz=gθ, D z' = g \theta' \,,

which means that

(6)Dz=(Dθ)θ, D z' = (D\theta') \theta' \,,

or

(7)β+θf =(h+θα)(α+θh) =hα+θ(h 2αα). \begin{aligned} \beta + \theta f' &= (h + \theta \alpha') (\alpha + \theta h) \\ &= h\alpha + \theta(h^2 - \alpha\alpha') \,. \end{aligned}

This, finally, is equivalent to

(8) hα=β h 2=f+αα. \begin{aligned} & h\alpha = \beta \\ & h^2 = f' + \alpha\alpha' \end{aligned} \,.

I might have mixed up a sign.

Hm, so I have this extra αα\alpha\alpha' term in the equation for hh. The same term does appear in equation (3.14) in Ninnemann.

I should check that more closely, but I have to run now.

Another thing I need to better understand: how do N=(1,1)N=(1,1) SRSs differ from N=(1,0)N=(1,0) SRSs.

Also, setting α=0\alpha = 0 corresponds to the “split” case, which (Ninnemann, p. 269) corresponds to differentiable SRSs. So do we assume this or not?

Posted by: urs on July 13, 2006 12:18 PM | Permalink | Reply to this

Re: SRSs

Hm, so I have this extra αα\alpha \alpha' term in the equation for hh.

Whoops! Yes. I should not have specialized. That term is there in the general case.

Also, setting α=0\alpha=0 corresponds to the “split” case, which (Ninnemann, p. 269) corresponds to differentiable SRSs. So do we assume this or not?

I think you are definitely using the complex structure of your SRS. As a real supermanifold, you could always remove any odd components in the transition functions. In the real case, any supermanifold takes the form M^=ΠV\hat{M}=\Pi V, where VMV\to M is a vector bundle and Π\Pi is the operation that makes the fibers “odd”.

But you are interested in (super)conformal structures. That is, you want to use the complex structure of Σ\Sigma. Over the complexes, there are, in general, cohomological obstructions to splitness.

Posted by: Jacques Distler on July 13, 2006 1:47 PM | Permalink | PGP Sig | Reply to this

Re: SRSs

Over the complexes, there are, in general, cohomological obstructions to splitness.

Ok. One day I need to understand this in detail. But didn’t you say here that

An SRS, Σ^\hat \Sigma, is, perforce, a split supermanifold […]

?

One thing that might be noteworthy about the terms α\alpha and β\beta in the transition law

(1)z =f+βθ θ =α+hθ \begin{aligned} z' &= f + \beta \theta \\ \theta' &= \alpha + h\theta \end{aligned}

is that these can only appear if we regard the transition morphism

(2)U i 1|1g 1|1U j U_i \simeq \mathbb{C}^{1|1} \overset{g}{\to} \mathbb{C}^{1|1} \simeq U_j

between our patches as a morphism of generalized supermanifolds, in the sense of sheaves on the category of ordinary supermanifolds (\to).

Namely, a (holomorphic) morphism of mere ringed spaces

(3) 1|1 1|1 \mathbb{C}^{1|1} \to \mathbb{C}^{1|1}

is fixed by a ring homomorphism

(4)Γ( 1|1)Γ( 1|1) \Gamma(\mathbb{C}^{1|1}) \to \Gamma(\mathbb{C}^{1|1})

which is fixed by an equation of the form

(5)z =f(z) θ =h(z)θ. \begin{aligned} z' &= f(z) \\ \theta' &= h(z)\theta \,. \end{aligned}

There are no other odd sections α\alpha, β\beta on 1|1\mathbb{C}^{1|1} which would be linearly independent (over ordinary functions on \mathbb{C}) of θ\theta and θ¯\bar \theta.

Right?

So, it seems to me, the only way to interpret α\alpha and β\beta in the first place is in terms of maps of “SS-points”, i.e. as things determining a natural transformation between sheaves on the category of supermanifolds.

I’ll spell out in detail what I mean, at least for my own benefit.

Let’s regard some morphism

(6)S ϕ S˜ \array{ S \\ \;\;\downarrow \phi \\ \tilde S }

of ordinary supermanifolds (i.e. of 2\mathbb{Z}_2-graded ringed spaces).

Our supermanifold 1|1\mathbb{C}^{1|1} is represented by the contravariant functor, which maps this to the morphism of sets

(7){(z,θ)Γ ev(S)×Γ odd(S)} Φ {(z˜,θ˜)Γ ev(S˜)×Γ odd(S˜)} \array{ \left\{ (z,\theta) \in \Gamma^\mathrm{ev}(S) \times \Gamma^\mathrm{odd}(S) \right\} \\ \;\;\uparrow \Phi \\ \left\{ (\tilde z,\tilde \theta) \in \Gamma^\mathrm{ev}(\tilde S) \times \Gamma^\mathrm{odd}(\tilde S) \right\} }

obtained by acting with the ring homomorphism Φ\Phi that comes with ϕ\phi.

A morphism 1|1 1|1\mathbb{C}^{1|1} \to \mathbb{C}^{1|1} in the sense of generalized supermanifolds is a natural automorphism of this functor, coming from commuting diagrams of maps of sets of the form

(8){(z,θ)Γ ev(S)×Γ odd(S)} g S {(z,θ)Γ ev(S)×Γ odd(S)} Φ Φ {(z˜,θ˜)Γ ev(S˜)×Γ odd(S˜)} g S˜ {(z˜,θ˜)Γ ev(S)×Γ odd(S)}. \array{ \left\{ (z,\theta) \in \Gamma^\mathrm{ev}(S) \times \Gamma^\mathrm{odd}(S) \right\} &\overset{g_S}{\to}& \left\{ (z',\theta') \in \Gamma^\mathrm{ev}(S) \times \Gamma^\mathrm{odd}(S) \right\} \\ \;\;\uparrow \Phi && \;\;\uparrow \Phi \\ \left\{ (\tilde z,\tilde \theta) \in \Gamma^\mathrm{ev}(\tilde S) \times \Gamma^\mathrm{odd}(\tilde S) \right\} &\overset{g_{\tilde S}}{\to}& \left\{ (\tilde z',\tilde \theta') \in \Gamma^\mathrm{ev}(S) \times \Gamma^\mathrm{odd}(S) \right\} } \,.

Now we have more freedom. In particular, now g Sg_S may act by picking two fixed families of odd sections α(z)\alpha(z) and β(z)\beta(z) of SS (which may be higher dimensional and hence may contain lots of independent odd sections) and act as

(9)g S : {(z,θ)Γ ev(S)×Γ odd(S)} {(z,θ)Γ ev(S)×Γ odd(S)} (z,θ) (f(z)+β(z)θ,α(z)+h(z)θ), \array{ g_S &:& \left\{ (z,\theta) \in \Gamma^\mathrm{ev}(S) \times \Gamma^\mathrm{odd}(S) \right\} &{\to}& \left\{ (z',\theta') \in \Gamma^\mathrm{ev}(S) \times \Gamma^\mathrm{odd}(S) \right\} \\ && (z,\theta) &\mapsto& (f(z) + \beta(z) \theta, \alpha(z) + h(z)\theta) } \,,

if these choices of additional sections are compatible with the ring homomorphisms Φ\Phi as

(10)α =Φ(α˜) β =Φ(β˜). \begin{aligned} \alpha &= \Phi(\tilde \alpha) \\ \beta &= \Phi(\tilde \beta) \end{aligned} \,.

Ok, much ado about nothing. What I am trying to say is that, unless I am making a mistake, precisely in the split case we can get away with staying in the category of ringed spaces, while otherwise we need to be working in sheaves over ringed spaces.

I think.

Posted by: urs on July 13, 2006 6:16 PM | Permalink | Reply to this

Re: SRSs

You said:

But didn’t you say here that

An SRS, Σ^\hat{\Sigma}, is, perforce, a split supermanifold […]

and then:

There are no other odd sections α,β\alpha,\beta on 1|1\mathbb{C}^{1|1} which would be linearly independent (over ordinary functions on \mathbb{C}) of θ\theta and θ¯\overline{\theta}.

The answer to your first question is contained in your second.

The way to think about these things is to imagine that they always come in families 𝒳B\mathcal{X}\to B (with fiber Σ^\hat \Sigma) where BB is also a supermanifold. α(z)\alpha(z) and h(z)h(z) are functions also of the coordinates on BB. We’ll just assume that BB has as many odd coordinates as we need.

So I always think about families of supermanifolds. Even if the supermanifold itself doesn’t vary over the family, BB provides the nilpotents necessary to think about superfunctions (e.g., they allow you to make sense of the “ψ\psi” in Φ(x,θ)=ϕ(x)+θ αψ α(x)+θ αθ αF(x)\Phi(x,\theta) = \phi(x) +\theta_\alpha\psi^\alpha(x) +\theta_\alpha\theta^\alpha F(x).)

Now, if BB is just a point, then, as you quickly concluded, it supplies no nilpotents, and α(z)0\alpha(z)\equiv 0. So a single SRS (not part of a family) is necessarily split.

Posted by: Jacques Distler on July 14, 2006 7:25 AM | Permalink | PGP Sig | Reply to this

Re: SRSs

Great, thanks!

On p. 283 of his paper (\to) Holger Ninnemann promises to explain

[…] the necessity of considering families of super-Riemann surfaces in order to describe a ghost-free fermionic theory.

I feel like I should already know what he means. But apparently I don’t.

Posted by: urs on July 14, 2006 12:45 PM | Permalink | Reply to this
Read the post Freed, Moore, Segal on p-Form Gauge Theory, I
Weblog: The String Coffee Table
Excerpt: Some aspects of recent work by Freed, Moore and Segal on the quantum theory of abelian p-form fields and the grading of the Hilbert space by electric and magnetic flux modulo torsion.
Tracked: July 10, 2006 2:30 PM
Read the post Seminar on 2-Vector Bundles and Elliptic Cohomology, VI
Weblog: The String Coffee Table
Excerpt: More detials on elliptic curves, formal groups and "classical" elliptic cohomology.
Tracked: July 13, 2006 9:16 PM
Read the post Connes on Spectral Geometry of the Standard Model, I
Weblog: The n-Category Café
Excerpt: Connes is coming closer to the spectral triple encoding the standard model coupled to gravity. Part I: some background material.
Tracked: September 6, 2006 12:22 PM
Read the post Connes on Spectral Geometry of the Standard Model, II
Weblog: The n-Category Café
Excerpt: On the nature and implication of Connes' identification of the spectral triple of the standard model coupled to gravity.
Tracked: September 6, 2006 7:08 PM
Read the post Hopkins Lecture on TFT: Introduction and Outlook
Weblog: The n-Category Café
Excerpt: Introductory lecture by M. Hopkins on topological field theory.
Tracked: October 25, 2006 11:07 AM
Read the post Some Conferences
Weblog: The n-Category Café
Excerpt: A conference on bundles and gerbes, another one on topology, and comments on associated 2-vector bundles and String connections.
Tracked: April 19, 2007 8:56 PM
Read the post Workshop on Elliptic Cohomology in Hamburg
Weblog: The n-Category Café
Excerpt: A little workshop on elliptic cohomology, mainly the functorial QFT approach by Stolz and Teichner, in Hamburg, June 2007.
Tracked: May 9, 2007 10:01 PM
Read the post Extended Quantum Field Theory and Cohomology, I
Weblog: The n-Category Café
Excerpt: On understanding extended quantum field theory and generalized cohomology.
Tracked: June 8, 2007 2:18 PM
Read the post HIM Trimester on Geometry and Physics, Week 4
Weblog: The n-Category Café
Excerpt: Talk in Stanford on nonabelian differential cohomology.
Tracked: May 29, 2008 11:26 PM

Post a New Comment