This paper was converted on www.awesomepapers.org from LaTeX by an anonymous user.
Want to know more? Visit the Converter page.

Characteristic classes in TMFTMF of level Γ1(3)\Gamma_{1}(3)

Gerd Laures Fakultät für Mathematik, Ruhr-Universität Bochum, NA1/66, D-44780 Bochum, Germany
Abstract.

Let TMF1(n)TMF_{1}(n) be the spectrum of topological modular forms equipped with a Γ1(n)\Gamma_{1}(n)-structure. We compute the K(2)K(2)-local TMF1(3)TMF_{1}(3)-cohomology of BStringB{\mathit{S}tring} and BSpinB{\mathit{S}pin}: both are power series rings freely generated by classes that we explicitly construct and which generalize the classical Pontryagin classes. As a first application of this computation, we show how to construct TMF(3n)TMF(3n)-cohomology classes from stable positive energy representations of the loop groups LSpinL{\mathit{S}pin}.

2000 Mathematics Subject Classification:
primary: 55N34, 55R40, secondary: 55P50, 22E66

1. Introduction and Statement of Results

Characteristic classes are cohomology classes which are naturally associated to principal GG-bundles on topological spaces. Examples include Stiefel-Whitney classes for O(n)O(n)-bundles or KOKO-Pontryagin classes for SO(n)SO(n)-bundles. If the cohomology theory is oriented with respect to GG-bordism, the characteristic classes can be evaluated on fundamental classes of manifolds with GG-structure and yield characteristic numbers. It is well known that oriented bordism is determined by Stiefel-Whitney numbers and rational Pontryagin numbers, whereas spin bordism is determined by Stiefel-Whitney numbers and KOKO-Pontryagin numbers.

In this paper we consider characteristic classes for spin and string bundles. Here, StringString is the 6-connected cover of the SpinSpin-group. The cohomology theory is TMF1(3)TMF_{1}(3), the cohomology of topological modular forms of level Γ1(3)\Gamma_{1}(3). It is possible that string bordism is determined by Stiefel-Whitney and TMFTMF-characteristic numbers for some level structures. At the prime 2, the connective version of TMFTMF should split off of the Thom spectrum MStringMString and there is some evidence that another summand is given by the 16th suspension of the connective cover of TMF0(3)TMF_{0}(3) (compare [MR09]). However, in order to get a map, a better understanding of the characteristic classes is necessary. (The relation between TMF1(3)TMF_{1}(3) and TMF0(3)TMF_{0}(3) is described in section 2.)

The homology groups EBU6E_{*}BU\langle 6\rangle for complex oriented theories EE have been described in [AHS01] in terms of cubical structures on the formal group of EE. Here, BU6BU\langle 6\rangle is the complex analogue of BStringB{\mathit{S}tring}. One may hope that the real case EBStringE_{*}B{\mathit{S}tring} admits a similar description in terms of “real” cubical structures. In [LO] the relation between cubical structures and the results of the current article has been further investigated.

The cohomology groups EBStringE^{*}B{\mathit{S}tring} for various complex oriented versions of TMFTMF have been considered in [Lau99]. When 22 is inverted these groups have been computed by the topological qq-expansion principle. The difficulty appears at the prime 22 since here a splitting result for the string bordism spectrum along the lines of Hovey-Ravenel [HR95] is missing. This paper deals with this remaining prime.

We first look at characteristic classes for spin bundles and show the following qq-expansion principle.

Theorem 1.1.

The diagram

TMF1(3)BSpin\textstyle{TMF_{1}(3)^{*}B{\mathit{S}pin}\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}λ\scriptstyle{\lambda}KTateBSpin\textstyle{K_{Tate}^{*}B{\mathit{S}pin}\ignorespaces\ignorespaces\ignorespaces\ignorespaces}ch\scriptstyle{ch}H(BSpin,TMF1(3)𝐐)\textstyle{H^{*}(B{\mathit{S}pin},TMF_{1}(3)^{*}_{\mathbf{Q}})\ignorespaces\ignorespaces\ignorespaces\ignorespaces}H(BSpin,KTate𝐐)\textstyle{H^{*}(B{\mathit{S}pin},{K_{Tate}}^{*}_{\mathbf{Q}})}

is a pullback.

The horizontal map is Miller’s elliptic character which corresponds to the evaluation at the Tate curve on the moduli stack of elliptic curves ([Mil89].) On coefficients this map is just the traditional q-expansion map for modular forms. The theory KTateK_{Tate} is the power series ring K[1/3]((q))K[1/3]((q)) of KK-theories as in [Lau99]. The right vertical map is the Chern character and the left vertical map is the Dold character, that is, the map to rational cohomology induced by the exponential of the formal group law (see ibid.)

The theorem determines the ring of TMF1(3)TMF_{1}(3)-characteristic classes for spin bundles as follows. An element of KTateBSpinK_{Tate}^{*}B{\mathit{S}pin} is a KTateK_{Tate}-characteristic class for spin bundles, that is, a formal series of virtual vector bundles which is naturally defined for spin bundles. If its Chern character is invariant under the appropriate Möbius transformations then it gives rise to a unique TMF1(3)TMF_{1}(3)-characteristic class. This property allows the construction of many natural classes. An application to representations of loop groups will be given later.

The proof of the theorem is given in Section 3. Locally at the prime 2 the cohomology theory TMF1(3)TMF_{1}(3) is a generalized Johnson-Wilson spectrum E(2)E(2) by Proposition 2.4. Hence it relates to the Morava KK-theory K(2)K(2) by a Bockstein spectral sequence. A splitting principle which is based on the Kitchloo-Laures computation [KL02] of the Morava K(2)K(2)-homology of BSpinB{\mathit{S}pin} forces the universal coefficient spectral sequence to collapse. Its E2E_{2}-term is studied via a chromatic resolution. Note that this part of the work does not involve the theory of elliptic curves but only the study of formal group laws of height at most 2. However, the statement of the theorem is a transchromatic property which finds a natural formulation in the qq-expansion of modular forms once the generalized E(2)E(2)-theory has been chosen to be TMF1(3)TMF_{1}(3). The pullback property is reduced via the universal coefficient isomorphisms to the classical q-expansion principle.

The theorem allows us to the construct Pontryagin classes for TMF1(3)TMF_{1}(3). The pullback diagram relates these to the classical Pontryagin classes of KK-theory and rational Pontryagin classes in a nice way.

Theorem 1.2.
  1. (i)

    There are unique classes piTMF1(3)4iBSpinp_{i}\in TMF_{1}(3)^{4i}B{\mathit{S}pin} with the following property: the formal series p(t)=1+p1t+p2t2p(t)=1+p_{1}t+p_{2}t^{2}\ldots is given by

    i=1m(1tρ(xix¯i))\prod_{i=1}^{m}(1-t\rho^{*}(x_{i}\overline{x}_{i}))

    when restricted to the classifying space of each maximal torus of Spin(2m)Spin(2m). Here, ρ\rho is the map to the maximal torus of SO(2m)SO(2m) and the xix_{i} (and x¯i\overline{x}_{i}) are the first TMF1(3)TMF_{1}(3)-Chern classes of the canonical line bundles LiL_{i} (resp. L¯i\overline{L}_{i}) over the classifying spaces of the tori.

  2. (ii)

    The classes pip_{i} freely generate the TMF1(3)TMF_{1}(3)-cohomology of BSpinB{\mathit{S}pin}, that is,

    TMF1(3)BSpinTMF1(3)[[p1,p2,]].TMF_{1}(3)^{*}B{\mathit{S}pin}\cong TMF_{1}(3)^{*}[\![p_{1},p_{2},\ldots]\!].

Next we turn to the calculation of the characteristic classes for BStringB{\mathit{S}tring}. Here it turns out that the qq-expansion principle only holds for odd primes. A string characteristic class may not even be determined by its characters. However, we have the following result.

Theorem 1.3.

Let TMF^1(3)\widehat{TMF}_{1}(3) be the K(2)K(2)-localization of TMF1(3)TMF_{1}(3) at the prime 2. Then there is an isomorphism of algebras

TMF^1(3)BStringTMF^1(3)[[r,p1,p2,]]\widehat{TMF}_{1}(3)^{*}B{\mathit{S}tring}\cong\widehat{TMF}_{1}(3)^{*}[\![r,p_{1},p_{2},\ldots]\!]

where p1,p2,p_{1},p_{2},\ldots are the Pontryagin classes coming from BSpinB{\mathit{S}pin} and rr restricts to a topological generator of degree 6 in the K(2)K(2)-cohomology of K(,3)K({\mathbb{Z}},3).

Theorems 1.2 and 1.3 solve the problem stated in [KLW04a]: they give explicit generators for the cohomology rings of BSpinB{\mathit{S}pin} and BStringB{\mathit{S}tring}. Their proofs are given in Section 4. Theorem 1.3 relies on a description of the Morava K(2)K(2)-homology of BStringB{\mathit{S}tring} given by Kitchloo-Laures-Wilson in [KLW04a, KLW04b]. We remark that the completed TMF1(3)TMF_{1}(3) can be described as a fixed point spectrum of Lubin-Tate theory (see [MR09].)

As a first consequence we give a proof of the stable version of a conjecture by Brylinski which was stated in [Bry90]. The precise formulation is given in the last section. For sufficiently large nn divisible by 3 as in Theorem 5.2, there is a group homomorphism

φ:(Pm)Γ(n)TMF(n)BString\varphi:(P_{m})_{\Gamma(n)}\longrightarrow{TMF}(n)^{*}B{\mathit{S}tring}

from a stable group of positive energy representations V of the free loop group LSpin{L}Spin to the TMFTMF-cohomology with level nn-structure. Here, the congruence group is larger than the one considered before since the character of the representation is only known to be invariant under the action of Γ(n)\Gamma(n) by a theorem of Kac and Wakimoto [KW88, Theorem A].

We will describe the map φ\varphi in terms of its elliptic character. Suppose PP is a StringString-principal bundle over XX. Let L~Spin\tilde{L}Spin be the universal central extension of LSpinL{\mathit{S}pin}. Then by [Wit88], LXLX carries a L~Spin\tilde{L}Spin-principal bundle L~P\tilde{L}P whose associated LSpinL{\mathit{S}pin}-bundle is LPLP. In particular, this holds for the universal StringString-bundle EStringE{\mathit{S}tring} over BStringB{\mathit{S}tring}. The elliptic character of φ\varphi gives the bundle

λφ(V)=(L~ESpin×L~SpinV)|BString\lambda\varphi(V)=(\tilde{L}ESpin\times_{\tilde{L}Spin}V)_{|B{\mathit{S}tring}}

(when VV is suitably normalized with a character of the rotation circle). In this formula, the right hand side is considered as a formal power series of virtual bundles by decomposing the bundle as a representation of the circle group which reparameterizes the loops. The evaluation of this class on the fundamental class of a string manifold is the formal index of the Dirac operator on LMLM with coefficient in the bundle associated to the representation VV (see [Bry90]).

Acknowledgements.

The author would like to thank Nitu Kitchloo, Martin Olbermann, Björn Schuster, Vesna Stojanoska, Neil Strickland and Steve Wilson for helpful discussions. He is also grateful to the referee for a very careful revision.

2. Topological modular forms of level Γ1(3)\Gamma_{1}(3)

In this section we describe the spectrum of topological modular forms with level structures. Most of the results are well known and are collected for later reference. We start more generally than actually needed in order to put the main theorem in a broader picture.

Let {\mathcal{M}} denote the stack of smooth elliptic curves. A morphism f:Sf:S\longrightarrow{\mathcal{M}} determines an elliptic curve CfC_{f} over the base scheme SS (see Deligne and Rapoport [DR73].) The cotangent spaces of the identity section of each elliptic curve give rise to the line bundle ω\omega of invariant differentials on {\mathcal{M}}. A section of ωk\omega^{\otimes k} is called a modular form of weight kk. We write MkM_{k} for the group of modular forms of weight kk.

Let CC be an elliptic curve over a ring RR. Let C[n]C[n] denote the kernel of the self-map which multiplies by nn on CC. If nn is invertible in RR, then étale-locally, C[n]C[n] is of the form /n×/n{\mathbb{Z}}/n\times{\mathbb{Z}}/n. A choice of an isomorphism is called a Γ(n)\Gamma(n)-structure on CC. A monomorphism /nC[n]{\mathbb{Z}}/n\longrightarrow C[n] is a Γ1(n)\Gamma_{1}(n)-structure. It corresponds to a choice of a point of exact order nn. A choice of a subgroup scheme of C[n]C[n] which is isomorphic to /n{\mathbb{Z}}/n is called a Γ0(n)\Gamma_{0}(n)-structure. There are maps of associated moduli stacks with nn inverted

(3)

and the sections of the associated line bundles are called modular forms with the corresponding level structures. For example, a Γ1(n)\Gamma_{1}(n)-modular form ff of weight kk associates an element of the ring RR to each triple (C/R,ω,P)(C/R,\omega,P) where CC is an elliptic curve over R=R[1/n]R=R[1/n], ω\omega is a translation invariant nowhere vanishing differential and PP is a point of exact order nn. This association should only depend on the isomorphism class of the triple, it should be invariant under base change, and it should satisfy

f(C,aω,P)=akf(C,ω,P)f(C,a\omega,P)=a^{-k}f(C,\omega,P)

for all aR×a\in R^{\times}.

The case n=3n=3 can be made more explicit. Locally, for any such triple (C,ω,P)(C,\omega,P) the curve can uniquely be written in the form

(4) C:\displaystyle C: y2+a1xy+a3y=x3\displaystyle y^{2}+a_{1}xy+a_{3}y=x^{3}

in a way that PP is the origin (0,0)(0,0) and ω\omega has the standard form ω=dx/(2y+a1x+a3)\omega=dx/(2y+a_{1}x+a_{3}). A proof of this fact can be found in [MR09, 3.2]. It means that there is universal triple (C,ω,(0,0))(C,\omega,(0,0)) over the ring [1/3,a1,a3,Δ1]{\mathbb{Z}}[1/3,a_{1},a_{3},\Delta^{-1}] with the property that locally any other triple is obtained by base change. Hence a Γ1(3)\Gamma_{1}(3)-modular form is determined by its value on the universal triple and gives an element in this ring. Moreover, any element in the ring gives a modular form with level structure. We get

(5) MΓ1(3)\displaystyle{M_{\Gamma_{1}(3)}}_{*} \displaystyle\cong [1/3,a1,a3,Δ1].\displaystyle{\mathbb{Z}}[1/3,a_{1},a_{3},\Delta^{-1}].

One should mention that all moduli problems for Γ1(n)\Gamma_{1}(n) and Γ(n)\Gamma(n)-structures with n3n\geq 3 are representable this way but Γ0(n)\Gamma_{0}(n) is not representable. A good reference for these classical results on modular forms and level structures is the book [KM85].

There is a derived version of these concepts as follows.

Theorem 2.1.

[Goe10, HL, AHR] There is a sheaf 𝒪TMF{\mathcal{O}}_{TMF} of EE_{\infty}-ring spectra over {\mathcal{M}} in the étale topology. This sheaf satisfies:

  1. (i)

    The spectrum TMF=Γ𝒪TMFTMF=\Gamma{\mathcal{O}}_{TMF} only has 2 and 3 torsion in homotopy. Away from 6 it is concentrated in even degrees and we have an isomorphism

    π2kTMF[1/6]Mk[1/6].\pi_{2k}TMF[1/6]\cong M_{k}[1/6].
  2. (ii)

    There is an orientation map MStringTMFMString\longrightarrow TMF which induces the Witten genus in homotopy. In fact, its image coincides with the homotopy groups of a connective version of TMFTMF.

  3. (iii)

    The sequence (3) of moduli stacks gives a sequence of spectra

    TMF(n)\textstyle{TMF(n)}TMF1(n)\textstyle{TMF_{1}(n)\ignorespaces\ignorespaces\ignorespaces\ignorespaces}TMF0(n)\textstyle{TMF_{0}(n)\ignorespaces\ignorespaces\ignorespaces\ignorespaces}TMF[1/n]\textstyle{TMF[1/n]\ignorespaces\ignorespaces\ignorespaces\ignorespaces}

    and induced equivalences (with nn inverted)

    TMF[1/n]\displaystyle TMF[1/n] \displaystyle\cong TMF(n)hGl2(/n)\displaystyle TMF(n)^{h\,Gl_{2}({\mathbb{Z}}/n)}
    TMF0(n)\displaystyle TMF_{0}(n) \displaystyle\cong TMF1(n)hGl1(/n).\displaystyle TMF_{1}(n)^{h\,Gl_{1}({\mathbb{Z}}/n)}.
  4. (iv)

    For all étale f:Spec(R)f:Spec(R)\longrightarrow{\mathcal{M}} the spectrum E=Γf𝒪TMFE=\Gamma f^{*}{\mathcal{O}}_{TMF} is a complex orientable ring spectrum whose formal group E0BS1E^{0}BS^{1} is equipped with an isomorphism to the formal completion of CfC_{f}.

The spectrum TMF1(3)TMF_{1}(3) can be described much more elementarily if one is only interested in its associated cohomology theory. Since the moduli problem is representable, its homotopy coincides with the ring MΓ1(3)M_{\Gamma_{1}(3)} by property (iv) as we now explain. (The ring MΓ1(3)M_{\Gamma_{1}(3)} has been determined in (5).) When we choose a coordinate on the formal group of the curve (4) we obtain a formal group law. These are classified by a map from the Lazard ring LL, which coincides with the homotopy groups of complex bordism MUMU. For instance, the 2-typicalization of the standard coordinate has Hazewinkel generators (see [Lau04, Lemma 1]) at p=2p=2

(6) v1\displaystyle v_{1} =\displaystyle= a1\displaystyle a_{1}
(7) v2\displaystyle v_{2} =\displaystyle= a3.\displaystyle a_{3}.

The Hazewinkel generators form a regular sequence and hence satisfy the Landweber exactness conditions. Thus for finite complexes XX we have natural isomorphisms

(8) TMF1(3)X\displaystyle TMF_{1}(3)^{*}X \displaystyle\cong MUXMUMΓ1(3).\displaystyle MU^{*}X\otimes_{MU^{*}}M_{\Gamma_{1}(3)}.

(In the older literature for example [Bry90, Fra92, Bak94] this theory carried the names EllΓ1(3)Ell^{\Gamma_{1}(3)} or EΓ1(3)E^{\Gamma_{1}(3)}.)

Lemma 2.2.

The map

(2)[v1,v2±1,(v1327v2)1]\textstyle{{\mathbb{Z}}_{(2)}[v_{1},v_{2}^{\pm 1},(v_{1}^{3}-27v_{2})^{-1}]\ignorespaces\ignorespaces\ignorespaces\ignorespaces}TMF1(3)(2)\textstyle{{TMF_{1}(3)_{(2)}}_{*}}

is an isomorphism.

Proof.

The discriminant of the universal curve has the form

Δ=a33(a1327a3).\Delta=a_{3}^{3}(a_{1}^{3}-27a_{3}).

The result follows from (6) and (7). ∎

Let wkπMUw_{k}\in\pi_{*}MU denote any element which projects to a generator of the indecomposables in π2(pk1)BP\pi_{2(p^{k}-1)}BP. Recall from [Mat, Definition 3.9] that an MUMU-module spectrum MM is a form of BPnBP\langle n\rangle if there are wn+1,wn+2,w_{n+1},w_{n+2},\ldots such that

MBP/(wn+1,wn+2,).M\cong BP/(w_{n+1},w_{n+2},\ldots).

In the sequel, we will call these theories generalized BPnBP\langle n\rangle and will use the same notation.

Definition 2.3.

An MUMU-module spectrum MM is a generalized E(n)E(n) if there exist a generalized BPnBP\langle n\rangle and an element wπMUw\in\pi_{*}MU such that

BPn[w1]MBP\langle n\rangle[w^{-1}]\cong M

and the element ww coincides with a power of vnv_{n} modulo the ideal (p,v1,,vn1)(p,v_{1},\ldots,v_{n-1}).

Proposition 2.4.

Locally at the prime 2, the spectrum TMF1(3)TMF_{1}(3) is a generalized Johnson-Wilson theory E(2)E(2).

Proof.

Use [LN12] or [Mat, Example 3.11] to get a connective version of TMF1(3)TMF_{1}(3) which is a generalized BP2BP\langle 2\rangle. The claim then follows from Lemma 2.2. ∎

Theorem 1.1 involves the elliptic character map. This map originates from the Tate curve

y2+xy=x3+B(q)x+C(q)y^{2}+xy=x^{3}+B(q)x+C(q)

where

B(q)=148(E4(q)1) and C(q)=1496(E4(q)1)1864(E6(q)1)B(q)=\frac{-1}{48}(E_{4}(q)-1)\mbox{ and }\;C(q)=\frac{1}{496}(E_{4}(q)-1)-\frac{1}{864}(E_{6}(q)-1)

with the Eisenstein series E4E_{4} and E6E_{6}. The series BB and CC are integral power series in qq. The evaluation of an ordinary modular form on the Tate curve with its canonical differential corresponds to its qq-expansion. The formal group associated to the Tate curve is the multiplicative formal group. In order to get a Γ1(n)\Gamma_{1}(n)-structure, that is, a point of order nn, one can use the extension of scalars ((q))((q));qqn{\mathbb{Z}}((q))\rightarrow{\mathbb{Z}}((q));\;q\mapsto q^{n}. The resulting curve is usually denoted by Tate(qn)Tate(q^{n}) and its multiplicative reduction furnishes the Miller character map [Mil89][Lau99]

λ:TMF1(n)K[1/n]((q)).\lambda:TMF_{1}(n)\longrightarrow K[1/n]((q)).

In homotopy this map is the classical qq-expansion.

Lemma 2.5.

Locally at p=2p=2, the map

λX:TMF1(3)[a11]XK[1/3]((q))X\lambda_{*}X:{TMF_{1}(3)}[a_{1}^{-1}]_{*}X\longrightarrow K[1/3]((q))_{*}X

is a monomorphism for all XX.

Proof.

Note that the qq-expansion of a1a_{1} starts with 1 and hence is invertible in the target. We will use Equation (8) and replace MUMU with BPBP. Every comodule is the inductive limit of its finitely generated subcomodules. Hence, we may assume that BPXBP_{*}X has a finite Landweber filtration (Fk)(F_{k}) with subsequent quotients of the form BP/ItBP_{*}/I_{t} with ItI_{t} the invariant prime ideal (p,v1,,vt1)(p,v_{1},\ldots,v_{t-1}). By the qq-expansion principle the character map λ\lambda_{*} is injective and injective mod p. Hence it is injective when tensored with each of the quotients BP/ItBP_{*}/I_{t}. The claim follows from the obvious inductive argument. ∎

Remark 2.6.

It would be interesting to set up character maps for finer structures like Γ0(n)\Gamma_{0}(n). We will come back to this question in a subsequent work.

3. The TMF1(3)TMF_{1}(3)-cohomology of BSpinB{\mathit{S}pin}

In this section we will show the universal coefficient isomorphism for the TMF1(3)TMF_{1}(3)-cohomology of the space BSpinB{\mathit{S}pin} and the pullback diagram of Theorem 1.1. The main ingredient is the Morava K(2)K(2)-homology which has been computed by Kitchloo-Laures in [KL02]. We can use this result to obtain information about the E(k,n)E(k,n)-cohomology by methods of Ravenel-Wilson-Yagita. A chromatic argument enables us to compute the universal coefficients spectral sequence for TMF1(3)BSpinTMF_{1}(3)^{*}B{\mathit{S}pin}. We will show that it collapses at the E2E_{2}-term and obtain Theorem 1.1 from the classical qq-expansion principle. Some of the results of this section apply to other situations and hence are formulated more generally than actually needed.

From now on we fix a prime pp. In a first step we do not want to deal with lim1\lim^{1}-questions and hence work with pp-completed spectra. Let E(n)E(n) be a pp-completed generalized Johnson-Wilson spectrum. Let IkI_{k} be the invariant prime ideal (p,v1,,vk1)(p,v_{1},\ldots,v_{k-1}) and let E(k,n)E(k,n) be the spectrum

E(k,n)=E(n)/Ik.E(k,n)=E(n)/I_{k}.

By definition we have E(n,n)K(n)E(n,n)\cong K(n), E(0,n)=E(n)E(0,n)=E(n) and there are cofibre sequences

(11) E(k1,n)vk1E(k1,n)E(k,n).\displaystyle\lx@xy@svg{\hbox{\raise 2.5pt\hbox{\kern 25.24069pt\hbox{\ignorespaces\ignorespaces\ignorespaces\hbox{\vtop{\kern 0.0pt\offinterlineskip\halign{\entry@#!@&&\entry@@#!@\cr&&\crcr}}}\ignorespaces{\hbox{\kern-25.24069pt\raise 0.0pt\hbox{\hbox{\kern 0.0pt\raise 0.0pt\hbox{\hbox{\kern 3.0pt\raise-2.5pt\hbox{$\textstyle{E(k-1,n)\ignorespaces\ignorespaces\ignorespaces\ignorespaces}$}}}}}}}\ignorespaces\ignorespaces\ignorespaces\ignorespaces{}{\hbox{\lx@xy@droprule}}\ignorespaces\ignorespaces\ignorespaces{\hbox{\kern 26.4473pt\raise 5.0375pt\hbox{{}\hbox{\kern 0.0pt\raise 0.0pt\hbox{\hbox{\kern 3.0pt\hbox{\hbox{\kern 0.0pt\raise-0.9764pt\hbox{$\scriptstyle{v_{k-1}}$}}}\kern 3.0pt}}}}}}\ignorespaces{\hbox{\kern 48.04062pt\raise 0.0pt\hbox{\hbox{\kern 0.0pt\raise 0.0pt\hbox{\lx@xy@tip{1}\lx@xy@tip{-1}}}}}}{\hbox{\lx@xy@droprule}}{\hbox{\lx@xy@droprule}}{\hbox{\kern 48.04062pt\raise 0.0pt\hbox{\hbox{\kern 0.0pt\raise 0.0pt\hbox{\hbox{\kern 3.0pt\raise-2.5pt\hbox{$\textstyle{E(k-1,n)\ignorespaces\ignorespaces\ignorespaces\ignorespaces}$}}}}}}}\ignorespaces\ignorespaces\ignorespaces\ignorespaces{}{\hbox{\lx@xy@droprule}}\ignorespaces{\hbox{\kern 121.32193pt\raise 0.0pt\hbox{\hbox{\kern 0.0pt\raise 0.0pt\hbox{\lx@xy@tip{1}\lx@xy@tip{-1}}}}}}{\hbox{\lx@xy@droprule}}{\hbox{\lx@xy@droprule}}{\hbox{\kern 121.32193pt\raise 0.0pt\hbox{\hbox{\kern 0.0pt\raise 0.0pt\hbox{\hbox{\kern 3.0pt\raise-2.5pt\hbox{$\textstyle{E(k,n)}$}}}}}}}\ignorespaces}}}}\ignorespaces.
Remark 3.1.

Recall from Hovey and Sadofsky [HS99a, Theorem 3.4] that for all generalized K(n)K(n), there is a faithfully flat extension of its coefficient ring over which its formal group law becomes strictly isomorphic to the Honda formal group law. This allows us to carry over results from the classical to the generalized K(n)K(n).

In the following, for a fixed 0<kn0<k\leq n, let XX be a space with even E(k,n)E(k,n)-cohomology. The exact sequence induced by (11)

(12)
0\textstyle{0\ignorespaces\ignorespaces\ignorespaces\ignorespaces}E(k1,n)evX\textstyle{E(k-1,n)^{ev}X\ignorespaces\ignorespaces\ignorespaces\ignorespaces}vk1\scriptstyle{v_{k-1}}E(k1,n)evX\textstyle{E(k-1,n)^{ev}X\ignorespaces\ignorespaces\ignorespaces\ignorespaces}E(k,n)evX\textstyle{E(k,n)^{ev}X}E(k1,n)oddX\textstyle{E(k-1,n)^{odd}X\ignorespaces\ignorespaces\ignorespaces\ignorespaces}vk1\scriptstyle{v_{k-1}}E(k1,n)oddX\textstyle{E(k-1,n)^{odd}X\ignorespaces\ignorespaces\ignorespaces\ignorespaces}0\textstyle{0}

tells us that each element in E(k1,n)oddXE(k-1,n)^{odd}X is infinitely divisible by vk1v_{k-1}. The following result applies:

Theorem 3.2.

If xx is infinitely divisible by vkv_{k} in E(k,n)XE(k,n)^{*}X, then it is zero.

Proof.

The proof for [RWY98, Corollary 4.11] verbatim carries over to the generalized E(k,n)E(k,n). In more detail, one first observes that xx is infinitely divisible when restricted to each finite subcomplex of XX. Hence one can assume that XX is finite. Then for P(k)=BP/IkP(k)=BP/I_{k} one tensors a Landweber filtration of P(k)XP(k)^{*}X with the generalized E(k,n)E(k,n)^{*} to reduce the claim to modules of the form E(k,n)/(vk,vs)E(k,n)^{*}/(v_{k},\ldots v_{s}) for some s<ns<n. Here, xx has to be zero. ∎

Corollary 3.3.

If XX is a space with even Morava K(n)K(n)-cohomology then E(k,n)XE(k,n)^{*}X is even for all kk and the exact sequences (12) are short exact.

Proof.

By an inductive argument we may assume that all elements of E(k1,n)oddXE(k-1,n)^{odd}X are infinitely divisible by vk1v_{k-1}. Then the claim follows from Theorem 3.2. ∎

Next we look at the universal coefficient spectral sequence for E=E(n)E=E(n)

ExtEs,t(EX,E)Es+tX\operatorname{Ext}_{E_{*}}^{s,t}(E_{*}X,E_{*})\Longrightarrow E^{s+t}X
Lemma 3.4.

The global dimension of E=E(n)E_{*}=E(n)_{*} in the category of graded modules equals nn.

Proof.

The corresponding result is well known in the ungraded setting. Its graded version is harder to find in the literature but the proof given in [Eis95, 19.5] carries over: Let kk_{*} be the graded field 𝔽p[vn±]{\mathbb{F}}_{p}[v_{n}^{\pm}]. Since (p,v1,,vn1)(p,v_{1},\ldots,v_{n-1}) is a regular sequence the Koszul complex provides a free graded resolution of length nn. This implies the vanishing of Tori+1E(k,M)\operatorname{Tor}^{E_{*}}_{i+1}(k_{*},M_{*}) for all ini\geq n and for all MM_{*}. Next let F=(Fn,φn)F=(F_{n*},\varphi_{n}) be a graded minimal free resolution of of a finitely generated module MM_{*}. (The minimality condition means that for each nn a basis of FnF_{n} maps to a minimal set of generators of kerφn1\ker\varphi_{n-1}.) Since all differentials in kFk_{*}\otimes F are 0 we have

Tori+1E(k,M)kEFi+1,.\operatorname{Tor}^{E_{*}}_{i+1}(k_{*},M_{*})\cong k_{*}\otimes_{E_{*}}F_{i+1,*}.

This vanishes iff Fi+1,F_{i+1,*} vanishes because the resolution is free. Thus we have shown the claim for finitely generated modules. The general result follows from the graded version of Auslander’s Theorem [Eis95, 19.1]. ∎

Lemma 3.5.

Let EE be E(2)E(2) and let E/pE/p^{\infty} be the cofibre of Ep1EE\rightarrow p^{-1}E. Suppose ExtE0,t+1(EX,E/p)\operatorname{Ext}^{0,t+1}_{E_{*}}(E_{*}X,E_{*}/p^{\infty}) vanishes for all even tt. Then for even tt there is an isomorphism

ExtE2,t(EX,E)(E/p)t+1X.\operatorname{Ext}^{2,t}_{E_{*}}(E_{*}X,E_{*})\cong(E/p^{\infty})^{t+1}X.
Proof.

Suppose FF is a ring theory with rational coefficients. Then Serre’s result and the comparison theorem for homology theories imply that the FF_{*}-linear extension of the FF-Hurewicz map

Fπst(X)\textstyle{F_{*}\otimes\pi_{*}^{st}(X)\ignorespaces\ignorespaces\ignorespaces\ignorespaces}F(X)\textstyle{F_{*}(X)}

is an isomorphism. It follows that p1EXp^{-1}E_{*}X is a free p1Ep^{-1}E_{*} module. Hence

ExtEs(EX,p1E)Extp1Es(p1EX,p1E)\operatorname{Ext}^{s}_{E_{*}}(E_{*}X,p^{-1}E_{*})\cong\operatorname{Ext}^{s}_{p^{-1}E_{*}}(p^{-1}E_{*}X,p^{-1}E_{*})

vanishes for all s>0s>0, and the short exact sequence

E\textstyle{E_{*}\,\,\ignorespaces\ignorespaces\ignorespaces\ignorespaces}p1E\textstyle{p^{-1}E_{*}\ignorespaces\ignorespaces\ignorespaces\ignorespaces}E/p\textstyle{E_{*}/p^{\infty}}

induces isomorphisms

ExtEs+1(EX,E)ExtEs(EX,E/p).\operatorname{Ext}^{s+1}_{E_{*}}(E_{*}X,E_{*})\cong\operatorname{Ext}^{s}_{E_{*}}(E_{*}X,E_{*}/p^{\infty}).

By 3.4 these groups vanish for s2s\geq 2 and so does ExtE0,t+1(EX,E/p)\operatorname{Ext}_{E_{*}}^{0,t+1}(E_{*}X,E_{*}/p^{\infty}) for even tt. Hence the claim follows from the universal coefficient spectral sequence

ExtEs,t(EX,E/p)(E/p)s+tX.\operatorname{Ext}_{E_{*}}^{s,t}(E_{*}X,E_{*}/p^{\infty})\Longrightarrow(E/p^{\infty})^{s+t}X.

We now specify to the case X=BSpinX=B{\mathit{S}pin} and prove a splitting principle using the following computation.

Theorem 3.6.

[KL02, 1.2] Let nn be 1 or 2. Let biK(n)2iBS1b_{i}\in K(n)_{2i}BS^{1} be the dual to the power c1i{c_{1}^{i}} of the first Chern class of the canonical line bundle. Denote its image under the map induced by the inclusion of the maximal torus

BS1\textstyle{BS^{1}\ignorespaces\ignorespaces\ignorespaces\ignorespaces}BSpin(3)\textstyle{B{\mathit{S}pin}(3)\ignorespaces\ignorespaces\ignorespaces\ignorespaces}BSpin\textstyle{B{\mathit{S}pin}}

by the same name. Then we have

K(n)BSpinπK(n)[b2n2,b2n4,b2n6,].K(n)_{*}B{\mathit{S}pin}\cong\pi_{*}K(n)[b_{2^{n}\cdot 2},b_{2^{n}\cdot 4},b_{2^{n}\cdot 6},\ldots].
Corollary 3.7.

Let T(m/2)T(\lfloor m/2\rfloor) be the maximal torus of Spin(m)Spin(m) and set BT=colimBT(m/2)BT^{\infty}=\mbox{co}\!\lim BT(\lfloor m/2\rfloor). Then the restriction map from K(n)BSpinK(n)^{*}B{\mathit{S}pin} to K(n)BTK(n)^{*}BT^{\infty} is injective for n2n\leq 2.

Proof.

It suffices to show that the dual map K(n)BTK(n)BSpinK(n)_{*}BT^{\infty}\longrightarrow K(n)_{*}B{\mathit{S}pin} is surjective. This is immediate from the theorem since each monomial bi1bi2bikb_{i_{1}}b_{i_{2}}\cdots b_{i_{k}} comes from the classifying space of the kk-dimensional torus. ∎

Corollary 3.8.

For E=E(k,n)E=E(k,n) with n2n\leq 2 the restriction map from EBSpinE^{*}B{\mathit{S}pin} to EBTE^{*}BT^{\infty} is injective.

Proof.

By descending induction on knk\leq n, by the previous corollary and by Corollary 3.3 we have a map of short exact sequences

E(k1,n)BSpin\textstyle{E(k-1,n)^{*}B{\mathit{S}pin}\,\,\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}vk1\scriptstyle{v_{k-1}}E(k1,n)BSpin\textstyle{E(k-1,n)^{*}B{\mathit{S}pin}\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}E(k,n)BSpin\textstyle{E(k,n)^{*}B{\mathit{S}pin}\ignorespaces\ignorespaces\ignorespaces\ignorespaces}E(k1,n)BT\textstyle{E(k-1,n)^{*}BT^{\infty}\,\,\ignorespaces\ignorespaces\ignorespaces\ignorespaces}vk1\scriptstyle{v_{k-1}}E(k1,n)BT\textstyle{E(k-1,n)^{*}BT^{\infty}\ignorespaces\ignorespaces\ignorespaces\ignorespaces}E(k,n)BT\textstyle{E(k,n)^{*}BT^{\infty}}

for which the last vertical map can be assumed to be injective. Hence, any element in E(k1,n)BSpinE(k-1,n)^{*}B{\mathit{S}pin} which restricts trivially to the torus is divisible by vk1v_{k-1} and the quotient again restricts trivially to the torus. Continuing this way, we see that it must be infinitely divisible by vk1v_{k-1} and thus has to vanish by Theorem 3.2. ∎

Theorem 3.9.

Let EE be E=E(1)E=E(1) or the 2-completed TMF1(3)TMF_{1}(3). Then the universal coefficient isomorphism

EBSpinHomE(EBSpin,E)E^{*}B{\mathit{S}pin}\cong\operatorname{Hom}_{E_{*}}(E_{*}B{\mathit{S}pin},E_{*})

holds.

Proof.

First note that we have the isomorphism

EBTHomE(EBT,E)E^{*}BT^{\infty}\cong\operatorname{Hom}_{E_{*}}(E_{*}BT^{\infty},E_{*})

because EBTcolimETkE_{*}BT^{\infty}\cong\operatorname*{colim}E_{*}T^{k} is free: a basis is given by arbitrary products of the form

βi1βi2βik\beta_{i_{1}}\otimes\beta_{i_{2}}\otimes\cdots\otimes\beta_{i_{k}}

where βi\beta_{i} is dual to c1ic_{1}^{i}.

For the space X=BSpinX=B{\mathit{S}pin} and for E=E(1)E=E(1) the universal coefficient spectral sequence degenerates to the short exact sequence

ExtE1,(E1X,E)\textstyle{\operatorname{Ext}^{1,*}_{E_{*}}(E_{*-1}X,E_{*})\,\,\ignorespaces\ignorespaces\ignorespaces\ignorespaces}EX\textstyle{E^{*}X\ignorespaces\ignorespaces\ignorespaces\ignorespaces}ExtE0,(EX,E)\textstyle{\operatorname{Ext}^{0,*}_{E_{*}}(E_{*}X,E_{*})}

by Lemma 3.4. Furthermore, the second map is injective because it factors through the restriction map into EBTExtE0,(EBT,E)E^{*}BT^{\infty}\cong\operatorname{Ext}^{0,*}_{E_{*}}(E_{*}BT^{\infty},E_{*}) and this map is injective by Corollary 3.8.

For the case E=TMF1(3)E=TMF_{1}(3) at p=2p=2, we already know from Corollary 3.3 that EXE^{*}X is concentrated in even degrees. Since EE_{*} is torsion free and p1EXp^{-1}E_{*}X is in even degrees, the group HomE(EX,E)\operatorname{Hom}_{E_{*}}(E_{*}X,E_{*}) must be even, too. Hence, we can restrict our attention to even degrees.

We would like to apply Lemma 3.5, and therefore we have to show that ExtE0,t+1(EX,E/p)\operatorname{Ext}^{0,t+1}_{E_{*}}(E_{*}X,E_{*}/p^{\infty}) vanishes for all even tt. By Lemma 2.5 or simply by the qq-expansion principle, the group E/pE_{*}/p^{\infty} injects into KTate/p{K_{Tate}}_{*}/p^{\infty} and so does the induced map on Ext0\operatorname{Ext}^{0}-groups. The isomorphisms at p=2p=2

ExtE0,t+1(EX,KTate/p)\displaystyle\operatorname{Ext}^{0,t+1}_{E_{*}}(E_{*}X,{K_{Tate}}_{*}/p^{\infty}) \displaystyle\cong ExtBP0,t+1(BPX,KTate/p)\displaystyle\operatorname{Ext}^{0,t+1}_{BP_{*}}(BP_{*}X,{K_{Tate}}_{*}/p^{\infty})
\displaystyle\cong ExtE(1)0,t+1(E(1)X,KTate/p)\displaystyle\operatorname{Ext}^{0,t+1}_{E(1)_{*}}(E(1)_{*}X,{K_{Tate}}_{*}/p^{\infty})

follow from the Landweber exactness of EE and KTateK_{Tate}. The latter group coincides with ExtE(1)1,t+1(E(1)X,KTate)\operatorname{Ext}^{1,t+1}_{E(1)_{*}}(E(1)_{*}X,{K_{Tate}}_{*}) which is a product of groups of the form ExtE(1)1,t+1(E(1)X,E(1))\operatorname{Ext}^{1,t+1}_{E(1)_{*}}(E(1)_{*}X,E(1)_{*}). Thus it is trivial.

The even dimensional Ext2\operatorname{Ext}^{2}-term of the universal coefficient spectral sequence has been identified with the odd part of (E/p)X({E/p^{\infty}})^{*}X in Lemma 3.5. Hence, for its vanishing it is enough to show the injectivity of the first map in the exact sequence

EX(p1E)X(E/p)X.\lx@xy@svg{\hbox{\raise 2.5pt\hbox{\kern 13.81248pt\hbox{\ignorespaces\ignorespaces\ignorespaces\hbox{\vtop{\kern 0.0pt\offinterlineskip\halign{\entry@#!@&&\entry@@#!@\cr&&\crcr}}}\ignorespaces{\hbox{\kern-13.81248pt\raise 0.0pt\hbox{\hbox{\kern 0.0pt\raise 0.0pt\hbox{\hbox{\kern 3.0pt\raise-2.5pt\hbox{$\textstyle{E^{*}X\ignorespaces\ignorespaces\ignorespaces\ignorespaces}$}}}}}}}\ignorespaces\ignorespaces\ignorespaces\ignorespaces{}{\hbox{\lx@xy@droprule}}\ignorespaces{\hbox{\kern 36.61241pt\raise 0.0pt\hbox{\hbox{\kern 0.0pt\raise 0.0pt\hbox{\lx@xy@tip{1}\lx@xy@tip{-1}}}}}}{\hbox{\lx@xy@droprule}}{\hbox{\lx@xy@droprule}}{\hbox{\kern 36.61241pt\raise 0.0pt\hbox{\hbox{\kern 0.0pt\raise 0.0pt\hbox{\hbox{\kern 3.0pt\raise-2.5pt\hbox{$\textstyle{(p^{-1}E)^{*}X\ignorespaces\ignorespaces\ignorespaces\ignorespaces}$}}}}}}}\ignorespaces\ignorespaces\ignorespaces\ignorespaces{}{\hbox{\lx@xy@droprule}}\ignorespaces{\hbox{\kern 107.02693pt\raise 0.0pt\hbox{\hbox{\kern 0.0pt\raise 0.0pt\hbox{\lx@xy@tip{1}\lx@xy@tip{-1}}}}}}{\hbox{\lx@xy@droprule}}{\hbox{\lx@xy@droprule}}{\hbox{\kern 107.02693pt\raise 0.0pt\hbox{\hbox{\kern 0.0pt\raise 0.0pt\hbox{\hbox{\kern 3.0pt\raise-2.5pt\hbox{$\textstyle{(E/p^{\infty})^{*}X}$}}}}}}}\ignorespaces}}}}\ignorespaces.

This follows from the observation that its composite with the restriction map to (p1E)BT(p^{-1}E)^{*}BT^{\infty} is injective by Corollary 3.8. In even degrees we obtain the short exact sequence

ExtE1,(E1X,E)EXExtE0,(EX,E).\lx@xy@svg{\hbox{\raise 2.5pt\hbox{\kern 44.38472pt\hbox{\ignorespaces\ignorespaces\ignorespaces\hbox{\vtop{\kern 0.0pt\offinterlineskip\halign{\entry@#!@&&\entry@@#!@\cr&&\crcr}}}\ignorespaces{\hbox{\kern-44.38472pt\raise 0.0pt\hbox{\hbox{\kern 0.0pt\raise 0.0pt\hbox{\hbox{\kern 3.0pt\raise-2.5pt\hbox{$\textstyle{\operatorname{Ext}_{E_{*}}^{1,*}(E_{*-1}X,E_{*})\,\,\ignorespaces\ignorespaces\ignorespaces\ignorespaces}$}}}}}}}\ignorespaces\ignorespaces\ignorespaces\ignorespaces{\hbox{\kern 44.38472pt\raise 0.0pt\hbox{\hbox{\kern 0.0pt\raise 0.0pt\hbox{\lx@xy@tip{1}\lx@xy@tip{-1}}}}}}{\hbox{\lx@xy@droprule}}\ignorespaces{\hbox{\kern 67.18465pt\raise 0.0pt\hbox{\hbox{\kern 0.0pt\raise 0.0pt\hbox{\lx@xy@tip{1}\lx@xy@tip{-1}}}}}}{\hbox{\lx@xy@droprule}}{\hbox{\lx@xy@droprule}}{\hbox{\kern 67.18465pt\raise 0.0pt\hbox{\hbox{\kern 0.0pt\raise 0.0pt\hbox{\hbox{\kern 3.0pt\raise-2.5pt\hbox{$\textstyle{E^{*}X\ignorespaces\ignorespaces\ignorespaces\ignorespaces}$}}}}}}}\ignorespaces\ignorespaces\ignorespaces\ignorespaces{}{\hbox{\lx@xy@droprule}}\ignorespaces{\hbox{\kern 117.60954pt\raise 0.0pt\hbox{\hbox{\kern 0.0pt\raise 0.0pt\hbox{\hbox{\kern-3.0pt\lower 0.0pt\hbox{\lx@xy@tip{1}\lx@xy@tip{-1}}}\lx@xy@tip{1}\lx@xy@tip{-1}}}}}}{\hbox{\lx@xy@droprule}}{\hbox{\lx@xy@droprule}}{\hbox{\kern 117.60954pt\raise 0.0pt\hbox{\hbox{\kern 0.0pt\raise 0.0pt\hbox{\hbox{\kern 3.0pt\raise-2.5pt\hbox{$\textstyle{\operatorname{Ext}_{E_{*}}^{0,*}(E_{*}X,E_{*})}$}}}}}}}\ignorespaces}}}}\ignorespaces.

Once more by Corollary 3.8 the kernel of the second map vanishes. ∎

Remark 3.10.

We do not know much about the generalized E(2)E(2)-homology of BSpinB{\mathit{S}pin}, not even if it is concentrated in even degrees. This is what makes the proof of the universal coefficient isomorphism difficult.

Proof of Theorem 1.1:.

We first show the pullback property for the 22-completed theories. Recall from [Lau99, Lemma 1.5] that there is a pullback of rings

TMF1(3)\textstyle{TMF_{1}(3)_{*}\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}K[1/3]((q))\textstyle{K[1/3]_{*}((q))\ignorespaces\ignorespaces\ignorespaces\ignorespaces}(TMF1(3))𝐐\textstyle{(TMF_{1}(3)_{*})_{\mathbf{Q}}\ignorespaces\ignorespaces\ignorespaces\ignorespaces}(K)𝐐((q))\textstyle{(K_{*})_{\mathbf{Q}}((q))}

and each corner is the coefficient ring of a Landweber exact theory. For each such theory EE, and for all spaces XX, we have the natural isomorphism

HomBP(BPX,E)HomE(EX,E).\operatorname{Hom}_{BP_{*}}(BP_{*}X,E_{*})\cong\operatorname{Hom}_{E_{*}}(E_{*}X,E_{*}).

Hence, when applying the left exact functor HomBP(BPBSpin,)\operatorname{Hom}_{BP_{*}}(BP_{*}B{\mathit{S}pin},-) to the diagram we still have a pullback. By Theorem 3.9 each corner satisfies the universal coefficient isomorphism and hence we get the desired pullback diagram.

It remains to show the integral result. For odd primes we still have the appropriate pullback diagram by [Lau99, Theorem 1.12]. Moreover, for all spectra YY we have an arithmetic pullback by [Bou79, Proposition 2.9]

(13) Y\textstyle{Y\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}pLS/pY\textstyle{\prod_{p}L_{S{\mathbb{Z}}/p}Y\ignorespaces\ignorespaces\ignorespaces\ignorespaces}LS𝐐Y\textstyle{L_{S{\mathbf{Q}}}Y\ignorespaces\ignorespaces\ignorespaces\ignorespaces}LS𝐐(pLS/pY)\textstyle{L_{S{\mathbf{Q}}}(\prod_{p}L_{S{\mathbb{Z}}/p}Y)}

where SGSG denotes the Moore spectrum. Mapping X=BSpinX=B{\mathit{S}pin} to the diagram for Y=TMF1(3)Y=TMF_{1}(3) yields a pullback of cohomology groups by the previous results. In this pullback diagram the lower right corner has the form

H(X,(pTMF1(3)p)𝐐)(pTMF1(3)p)𝐐X,H^{*}(X,\left(\prod_{p}{TMF_{1}(3)}_{p}^{\wedge}\right)^{*}_{\mathbf{Q}})\cong\left(\prod_{p}TMF_{1}(3)^{\wedge}_{p}\right)^{*}_{\mathbf{Q}}X,

which we can replace with

H(X,p(TMF1(3)p)𝐐)p(TMF1(3)p)𝐐X.H^{*}(X,\prod_{p}\left({TMF_{1}(3)}_{p}^{\wedge}\right)^{*}_{\mathbf{Q}})\cong\prod_{p}\left(TMF_{1}(3)^{\wedge}_{p}\right)^{*}_{\mathbf{Q}}X.

Then the enlarged diagram

TMF1(3)X\textstyle{TMF_{1}(3)^{*}X\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}p(TMF1(3)p)X\textstyle{\prod_{p}\left(TMF_{1}(3)^{\wedge}_{p}\right)^{*}X\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}pK^pX((q))\textstyle{\prod_{p}\hat{K}_{p}^{*}X((q))\ignorespaces\ignorespaces\ignorespaces\ignorespaces}TMF1(3)𝐐X\textstyle{TMF_{1}(3)^{*}_{\mathbf{Q}}X\ignorespaces\ignorespaces\ignorespaces\ignorespaces}p(TMF1(3)p)𝐐X\textstyle{\prod_{p}\left(TMF_{1}(3)^{\wedge}_{p}\right)^{*}_{\mathbf{Q}}X\ignorespaces\ignorespaces\ignorespaces\ignorespaces}p(K^p)𝐐X((q))\textstyle{\prod_{p}(\hat{K}_{p}^{*})_{\mathbf{Q}}X((q))}

is a pullback. Since the upper composite factors through KTateXK_{Tate}^{*}X the result follows. ∎

Corollary 3.11.

The integral universal coefficient isomorphism

TMF1(3)BSpin\displaystyle TMF_{1}(3)^{*}B{\mathit{S}pin} \displaystyle\cong HomTMF1(3)(TMF1(3)BSpin,TMF1(3))\displaystyle\operatorname{Hom}_{TMF_{1}(3)_{*}}(TMF_{1}(3)_{*}B{\mathit{S}pin},TMF_{1}(3)_{*})

holds.

Proof.

The right hand side satisfies the pullback property in the diagram of the main theorem. ∎

4. Pontryagin classes and the cohomology of BStringB{\mathit{S}tring}

In this section we construct explicit generators in the TMF1(3)TMF_{1}(3)-cohomology rings of BSpinB{\mathit{S}pin} and BStringB{\mathit{S}tring} with the help of Theorem 1.1. We start with a reminder of the KOKO-Pontryagin classes. We will see that for line bundles they are defined by the same formula as the Pontryagin classes in singular homology. Using Theorem 1.1 we then obtain a construction of TMF1(3)TMF_{1}(3)-Pontryagin classes.

Recall from [ABP66, Proposition 4.4(b)] that the restriction maps from KO0(BSO(n))KO^{0}(B{\mathit{S}O}(n)) to the maximal torus K0(BT(n/2))K^{0}(BT(\lfloor n/2\rfloor)) are monomorphisms with images the invariants of the Weyl groups. The KOKO-Pontryagin classes are defined by the preimage of the series

(14) i=1n/2(1t(xix¯i))K0(BT(n/2)).\displaystyle\prod_{i=1}^{\lfloor n/2\rfloor}(1-t(x_{i}\overline{x}_{i}))\in K^{0}(BT(\lfloor n/2\rfloor)).

Here, xi=1Lix_{i}=1-L_{i} are the first Chern classes of the canonical line bundles LiL_{i} over BT(n/2)BT(\lfloor n/2\rfloor) in KK-theory.

The Pontryagin classes freely (topologically) generate the ring KO0(BSO)KO^{0}(B{\mathit{S}O}). Since we do not know a reference for this fact we give a short argument: first note that in KK-theory we have for all x=1Lx=1-L the equality

x+x¯=xx¯.x+\bar{x}=x\bar{x}.

Hence a power series which is invariant under the map which interchanges xx and x¯\bar{x} can be written as a power series in xx¯x\bar{x}. We conclude that each class in KO0(BSO)KO^{0}(B{\mathit{S}O}) is a symmetric power series in xixi¯x_{i}\bar{x_{i}} when restricted to K0BTK^{0}BT for all tori TT. Thus it is a power series in the Pontryagin classes.

We also note that

KO0(BSpin)KO0(BSO).KO^{0}(B{\mathit{S}pin})\cong KO^{0}(B{\mathit{S}O}).

This fact follows from the arithmetic square (13) since the map from BSpinB{\mathit{S}pin} to BSOB{\mathit{S}O} is a K(1)K(1)-local and rational equivalence (see [KL02, Theorem 1.2(ii)]).

Proof of Theorem 1.2: .

The KK-Pontryagin classes freely generate the KTateK^{*}_{Tate}-algebra KTate(BSpin)K^{*}_{Tate}(B{\mathit{S}pin}). We also know that the classical Pontryagin classes in rational singular cohomology freely generate H(BSpin;TMF1(3)𝐐)H^{*}(B{\mathit{S}pin};TMF_{1}(3)^{*}_{{\mathbf{Q}}}) as a TMF1(3)𝐐TMF_{1}(3)^{*}_{{\mathbf{Q}}}-algebra. They are defined in the same way except that in Formula (14), the xix_{i} are the ordinary first Chern classes.

The multiplicative formal group law over KTate{K_{Tate}}_{*} is strictly isomorphic to the one coming from the Tate curve, that is, the qq-expansion of the curve (4) together with the standard coordinate. By [Mil89] there is a natural automorphism of KTateK_{Tate} which exchanges the corresponding two orientations. Hence, when we replace the xix_{i} in the formula for the KK-Pontryagin classes by the first Chern classes with respect to the new orientation we still have free generators. The same argument holds for rational singular homology because here all formal group laws are strictly isomorphic.

These Pontryagin classes define elements in each corner of the pullback diagram of Theorem 1.1 and hence free generators of TMF1(3)BSpinTMF_{1}(3)^{*}B{\mathit{S}pin}. Moreover, these classes are determined by their restrictions to the maximal tori by Corollary 3.8 and are here given by the displayed formula. ∎

Next we consider the cohomology of BStringB{\mathit{S}tring}. The space BStringB{\mathit{S}tring} is defined as the homotopy fibre of the map BSpinK(,4)B{\mathit{S}pin}\rightarrow K({\mathbb{Z}},4) which kills the lowest homotopy group. In particular, we have a sequence of infinite loop spaces

(17) K(/2,2)K(,3)BStringBSpin.\displaystyle\lx@xy@svg{\hbox{\raise 2.5pt\hbox{\kern 25.38193pt\hbox{\ignorespaces\ignorespaces\ignorespaces\hbox{\vtop{\kern 0.0pt\offinterlineskip\halign{\entry@#!@&&\entry@@#!@\cr&&&\crcr}}}\ignorespaces{\hbox{\kern-25.38193pt\raise 0.0pt\hbox{\hbox{\kern 0.0pt\raise 0.0pt\hbox{\hbox{\kern 3.0pt\raise-2.5pt\hbox{$\textstyle{K({\mathbb{Z}}/2,2)\ignorespaces\ignorespaces\ignorespaces\ignorespaces}$}}}}}}}\ignorespaces\ignorespaces\ignorespaces\ignorespaces{}{\hbox{\lx@xy@droprule}}\ignorespaces{\hbox{\kern 48.18185pt\raise 0.0pt\hbox{\hbox{\kern 0.0pt\raise 0.0pt\hbox{\lx@xy@tip{1}\lx@xy@tip{-1}}}}}}{\hbox{\lx@xy@droprule}}{\hbox{\lx@xy@droprule}}{\hbox{\kern 48.18185pt\raise 0.0pt\hbox{\hbox{\kern 0.0pt\raise 0.0pt\hbox{\hbox{\kern 3.0pt\raise-2.5pt\hbox{$\textstyle{K({\mathbb{Z}},3)\ignorespaces\ignorespaces\ignorespaces\ignorespaces}$}}}}}}}\ignorespaces\ignorespaces\ignorespaces\ignorespaces{}{\hbox{\lx@xy@droprule}}\ignorespaces{\hbox{\kern 109.52344pt\raise 0.0pt\hbox{\hbox{\kern 0.0pt\raise 0.0pt\hbox{\lx@xy@tip{1}\lx@xy@tip{-1}}}}}}{\hbox{\lx@xy@droprule}}{\hbox{\lx@xy@droprule}}{\hbox{\kern 109.52344pt\raise 0.0pt\hbox{\hbox{\kern 0.0pt\raise 0.0pt\hbox{\hbox{\kern 3.0pt\raise-2.5pt\hbox{$\textstyle{B{\mathit{S}tring}\ignorespaces\ignorespaces\ignorespaces\ignorespaces}$}}}}}}}\ignorespaces\ignorespaces\ignorespaces\ignorespaces{}{\hbox{\lx@xy@droprule}}\ignorespaces{\hbox{\kern 175.00879pt\raise 0.0pt\hbox{\hbox{\kern 0.0pt\raise 0.0pt\hbox{\lx@xy@tip{1}\lx@xy@tip{-1}}}}}}{\hbox{\lx@xy@droprule}}{\hbox{\lx@xy@droprule}}{\hbox{\kern 175.00879pt\raise 0.0pt\hbox{\hbox{\kern 0.0pt\raise 0.0pt\hbox{\hbox{\kern 3.0pt\raise-2.5pt\hbox{$\textstyle{B{\mathit{S}pin}}$}}}}}}}\ignorespaces}}}}\ignorespaces.
Theorem 4.1 ([KLW04a] [KLW04b]).

The sequence (17) induces an exact sequence of Hopf algebras in Morava K(2)K(2)-homology at the prime 2

K(2)\textstyle{K(2)_{*}\ignorespaces\ignorespaces\ignorespaces\ignorespaces}K(2)K(/2,2)\textstyle{K(2)_{*}K({\mathbb{Z}}/2,2)\ignorespaces\ignorespaces\ignorespaces\ignorespaces}K(2)K(,3)\textstyle{K(2)_{*}K({\mathbb{Z}},3)\ignorespaces\ignorespaces\ignorespaces\ignorespaces}K(2)BString\textstyle{K(2)_{*}B{\mathit{S}tring}}K(2)BSpin\textstyle{K(2)_{*}B{\mathit{S}pin}\ignorespaces\ignorespaces\ignorespaces\ignorespaces}K(2).\textstyle{K(2)_{*}.}

Algebraically, there is short exact sequence of Hopf algebras

K(2)K(,3)\textstyle{K(2)_{*}K({\mathbb{Z}},3)\ignorespaces\ignorespaces\ignorespaces\ignorespaces}K(2)BString\textstyle{K(2)_{*}B{\mathit{S}tring}\ignorespaces\ignorespaces\ignorespaces\ignorespaces}K(2)BSpin\textstyle{K(2)_{*}B{\mathit{S}pin}}

which splits. (Here, the first map is not the one which is induced from the second map of (17).) In particular, the module K(2)BStringK(2)_{*}B{\mathit{S}tring} is concentrated in even dimensions.

We will use this information for the computation of the cohomology ring of BStringB{\mathit{S}tring} with respect to

E=LK(2)E(2)E=L_{K(2)}E(2)

for a generalized E(2)E(2). The coefficients of EE are given by E=(E(2))I2E_{*}=(E(2)_{*})^{\wedge}_{I_{2}}.

Proposition 4.2 ([HS99b] Proposition 2.5).

Suppose XX is a space with even Morava K(2)K(2)-homology. Then EXE^{*}X is the completion with respect to I2I_{2} of a free EE_{*}-module.

Proposition 4.3.

There is an isomorphism of algebras

EK(,3)E[[r]]E^{*}K({\mathbb{Z}},3)\cong E^{*}[\![r]\!]

with rr in degree 6.

Proof.

The K(2)K(2)-cohomology of K(,3)K({\mathbb{Z}},3) has been computed in [RW80] and [JW85] (see also Su [Su07] for its E(1,2)E(1,2)-cohomology.) It is topologically free on a generator of degree 6. Lift this generator to an EE-cohomology class. This is possible because the algebra EK(,3)E^{*}K({\mathbb{Z}},3) is concentrated in even degrees with the Morava K(2)K(2)-cohomology as its I2I_{2}-reduction. (This can be seen as before with the exact sequence (12).) Clearly, when restricted to a finite subcomplex of K(,3)K({\mathbb{Z}},3) every reduced class becomes nilpotent. Hence, we obtain an algebra map from E[[r]]E^{*}[\![r]\!] to EK(,3)E^{*}K({\mathbb{Z}},3). The result follows from Proposition 4.2 and the following version of Nakayama’s Lemma. ∎

Lemma 4.4.

Let RR be a graded ring with a unique maximal homogeneous ideal mm and let PP and QQ be pro-free RR modules. Then f:PQf:P\rightarrow Q is an isomorphism if and only if it is so modulo mm.

Proof.

Tensor the short exact sequences

mk/mk+1\textstyle{m^{k}/m^{k+1}\,\,\ignorespaces\ignorespaces\ignorespaces\ignorespaces}R/mk+1\textstyle{R/m^{k+1}\ignorespaces\ignorespaces\ignorespaces\ignorespaces}R/mk\textstyle{R/m^{k}}

with ff and use the fact that mk/mk+1m^{k}/m^{k+1} is a free module over R/mR/m. ∎

Proof of Theorem 1.3: .

By Theorem 4.1, we find a lift rr of the generator of K(2)K(,3)K(2)^{*}K({\mathbb{Z}},3) to K(2)BStringK(2)^{*}B{\mathit{S}tring}. Since K(2)BStringK(2)^{*}B{\mathit{S}tring} is concentrated in even degrees it is the quotient of EBStringE^{*}B{\mathit{S}tring} by the ideal I2I_{2}. Hence we can lift rr further to a class in EBStringE^{*}B{\mathit{S}tring} and obtain an algebra map

f:EBSpin^E[[r]]EBString.f:\lx@xy@svg{\hbox{\raise 2.5pt\hbox{\kern 38.71082pt\hbox{\ignorespaces\ignorespaces\ignorespaces\hbox{\vtop{\kern 0.0pt\offinterlineskip\halign{\entry@#!@&&\entry@@#!@\cr&\crcr}}}\ignorespaces{\hbox{\kern-38.71082pt\raise 0.0pt\hbox{\hbox{\kern 0.0pt\raise 0.0pt\hbox{\hbox{\kern 3.0pt\raise-2.5pt\hbox{$\textstyle{E^{*}B{\mathit{S}pin}\hat{\otimes}E^{*}[\![r]\!]\ignorespaces\ignorespaces\ignorespaces\ignorespaces}$}}}}}}}\ignorespaces\ignorespaces\ignorespaces\ignorespaces{}{\hbox{\lx@xy@droprule}}\ignorespaces{\hbox{\kern 61.51074pt\raise 0.0pt\hbox{\hbox{\kern 0.0pt\raise 0.0pt\hbox{\lx@xy@tip{1}\lx@xy@tip{-1}}}}}}{\hbox{\lx@xy@droprule}}{\hbox{\lx@xy@droprule}}{\hbox{\kern 61.51074pt\raise 0.0pt\hbox{\hbox{\kern 0.0pt\raise 0.0pt\hbox{\hbox{\kern 3.0pt\raise-2.5pt\hbox{$\textstyle{E^{*}B{\mathit{S}tring}}$}}}}}}}\ignorespaces}}}}\ignorespaces.

This map is an isomorphism by Proposition 4.2 and the previous lemma. ∎

Remark 4.5.

Let EE be the 2-complete TMF1(3)TMF_{1}(3). Then there is a pullback square of cohomology rings

EBString\textstyle{E^{*}B{\mathit{S}tring}\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}(LK(2)E)BString\textstyle{(L_{K(2)}E)^{*}B{\mathit{S}tring}\ignorespaces\ignorespaces\ignorespaces\ignorespaces}(LK(1)E)BString\textstyle{(L_{K(1)}E)^{*}B{\mathit{S}tring}\ignorespaces\ignorespaces\ignorespaces\ignorespaces}(LK(1)LK(2)E)BString\textstyle{(L_{K(1)}L_{K(2)}E)^{*}B{\mathit{S}tring}}

for which we have computed the three corners: K(1)K(1)-locally the map from BSpinB{\mathit{S}pin} to BStringB{\mathit{S}tring} is an equivalence because the K(1)K(1)-homology of K(,3)K({\mathbb{Z}},3) vanishes. Hence, the lower horizontal map is the inclusion

(LK(1)E)[[p1,p2,]]\textstyle{(L_{K(1)}E)^{*}[\![p_{1},p_{2},\ldots]\!]\ignorespaces\ignorespaces\ignorespaces\ignorespaces}(LK(1)LK(2)E)[[p1,p2,]]\textstyle{(L_{K(1)}L_{K(2)}E)^{*}[\![p_{1},p_{2},\ldots]\!]}

but for the right vertical map the image of the class rr is unclear. We will address this question in a subsequent work.

5. Applications to representations of loop groups

A principal Spin(d)Spin(d)-bundle PP over XX and a representation VV of Spin(d)Spin(d) give rise to a vector bundle over XX by associating VV to each fibre of PP. Hence PP and VV define an element in the KK-theory ring K0(X)K^{0}(X). If XX is a spin manifold and PP is the principal bundle associated to its tangent bundle, the pushforward of this KK-theory class to a point is the index of the Dirac operator twisted by VV (see [AS68]). The construction induces a ring map

(20) Rep(Spin(d))K0(X).\displaystyle\lx@xy@svg{\hbox{\raise 2.5pt\hbox{\kern 32.65259pt\hbox{\ignorespaces\ignorespaces\ignorespaces\hbox{\vtop{\kern 0.0pt\offinterlineskip\halign{\entry@#!@&&\entry@@#!@\cr&\crcr}}}\ignorespaces{\hbox{\kern-32.65259pt\raise 0.0pt\hbox{\hbox{\kern 0.0pt\raise 0.0pt\hbox{\hbox{\kern 3.0pt\raise-2.5pt\hbox{$\textstyle{Rep(Spin(d))\ignorespaces\ignorespaces\ignorespaces\ignorespaces}$}}}}}}}\ignorespaces\ignorespaces\ignorespaces\ignorespaces{}{\hbox{\lx@xy@droprule}}\ignorespaces{\hbox{\kern 55.45251pt\raise 0.0pt\hbox{\hbox{\kern 0.0pt\raise 0.0pt\hbox{\lx@xy@tip{1}\lx@xy@tip{-1}}}}}}{\hbox{\lx@xy@droprule}}{\hbox{\lx@xy@droprule}}{\hbox{\kern 55.45251pt\raise 0.0pt\hbox{\hbox{\kern 0.0pt\raise 0.0pt\hbox{\hbox{\kern 3.0pt\raise-2.5pt\hbox{$\textstyle{K^{0}(X)}$}}}}}}}\ignorespaces}}}}\ignorespaces.

It factors through the map from K0(BSpin(d))K^{0}(B{\mathit{S}pin}(d)) to K0(X)K^{0}(X) which classifies PP (see [AS69]).

In [Bry90] Brylinski conjectured a similar connection for loop space representations and the elliptic cohomology of the 7-connected cover of BSpin(d)B{\mathit{S}pin}(d) which is usually denoted by BString(d)B{\mathit{S}tring}(d). Recall from [PS86]) that a positive energy representation VV of the free loop space LSpin(d)L{\mathit{S}pin}(d) means that VV is a representation of the semi-direct product L~Spin(d)S1\tilde{L}Spin(d)\rtimes S^{1} where L~Spin(d)\tilde{L}Spin(d) is the universal central extension of LSpin(d)L{\mathit{S}pin}(d) and S1S^{1} acts on L~Spin(d)\tilde{L}Spin(d) by reparameterizing the loops. ‘Positive energy’ means that for the action of Rt=e2πitS1R_{t}=e^{2\pi it}\in S^{1} the vector space

Vn={vV:Rt(v)=e2πintv for all t}V_{n}=\{v\in V:\>R_{t}(v)=e^{2\pi int}v\mbox{ for all }t\}

is finite dimensional for all nn and vanishes for all n<n0n<n_{0} for some n0n_{0}. (Note that we can always multiply the rotation action with a character of S1S^{1} to get n0=0n_{0}=0.) When identifying the group Spin(d)Spin(d) with a subgroup of L~Spin(d)S1\tilde{L}Spin(d)\rtimes S^{1}, we obtain a formal power series

n[Vn]qnRep(Spin(d))((q))\sum_{n\in{\mathbb{Z}}}[V_{n}]q^{n}\in Rep(Spin(d))((q))

and hence via the map (20) a class in KTate0(X)K_{Tate}^{0}(X) for every Spin(d)Spin(d)-bundle PP over X. If PP comes from a string bundle then the loop space LXLX carries a L~Spin(d)\tilde{L}Spin(d)-principal bundle (see ibidem). Hence, the class in KTate0(X)K_{Tate}^{0}(X) associated to VV can be viewed as

(L~P×L~Spin(d)V)|X.(\tilde{L}P\times_{\tilde{L}Spin(d)}V)_{|X}.

In case XX is a string manifold and PP comes from its tangent bundle, the pushforward of this KTateK_{Tate}-class to a point is the formal index of the S1S^{1}-equivariant Dirac operator on the free loop space of XX with coefficient in the bundle associated to the representation VV (see [Wit88]). The fact that this index is a modular form of some level leads to the hope that the KTate(M)K_{Tate}(M)-cohomology class refines to a class in topological modular forms (c.[Bry90]). More precisely, let mm be an integer and let Pm(d)P_{m}(d) be the free abelian group generated by the isomorphism classes of irreducible positive energy representations of L~Spin(d)\tilde{L}Spin(d) of level mm (see [PS86, 9.3] for the meaning of ‘level’ ).

Conjecture 5.1.

There is an integer nn depending on dd and mm and an additive map

φd:Pm(d)TMF(n)0BString(d)\varphi_{d}:P_{m}(d)\longrightarrow TMF(n)^{0}B{\mathit{S}tring}(d)

whose elliptic character λ\lambda (defined as before in [Mil89]) coincides with the bundle

(21) λφd(V)\displaystyle\lambda\varphi_{d}(V) =\displaystyle= (L~ESpin(d)×L~Spin(d)V)|BString(d).\displaystyle(\tilde{L}E{\mathit{S}pin}(d)\times_{\tilde{L}Spin(d)}V)_{|B{\mathit{S}tring}(d)}.

Let n=24(m+g)n=24(m+g), with g=d2g=d-2 the Coxeter number. Theorem A in [KW88] states that the character JJ of a positive energy representation VV is a formal Jacobi modular form of level nn, weight w=0w=0 and index m2\frac{m}{2}. This means that it is invariant under the action of Γ(n)\Gamma(n) defined by

(22) (J[abcd])(z,τ)=(cτ+d)we2πimc(jzj2)/(cτ+d)J(zcτ+d,aτ+bcτ+d)\displaystyle\qquad\left(J\,\begin{bmatrix}a&b\\ c&d\end{bmatrix}\right)(z,\tau)=(c\tau+d)^{-w}e^{-2\pi imc(\sum_{j}z_{j}^{2})/(c\tau+d)}J(\frac{z}{c\tau+d},\frac{a\tau+b}{c\tau+d})

when suitably normalized with a character of the rotation group. (In this formula the Chern roots xix_{i} are replaced by 2πizi2\pi iz_{i} (c.[Bry90]).) Since for string bundles the first Pontryagin class p1=jzj2p_{1}=\sum_{j}z_{j}^{2} vanishes we have the following result.

Theorem 5.2.

[Bry90, Remark 3.10] Conjecture 5.1 holds rationally for n=24(m+g)n=24(m+g).

We will show a stable version of the integral conjecture. Let PmP_{m} be the inverse limit of all Pm(d)P_{m}(d)’s. For nn divisible by 3, let (Pm)Γ(n)(P_{m})_{\Gamma(n)} be the subgroup of PmP_{m} consisting of representations VV with character JJ invariant under the action (22) of Γ(n)\Gamma(n).

Theorem 5.3.

There is an additive map

φ:(Pm)Γ(n)TMF(n)^0BString\varphi:(P_{m})_{\Gamma(n)}\longrightarrow\widehat{TMF(n)}^{0}B{\mathit{S}tring}

whose elliptic character is the class described in Equation (21).

The proof will be given at the end of the section.

Lemma 5.4.

Let 3 be inverted in the following rings of modular forms and let nn be divisible by 3. Then the following ring extensions are flat:

MΓ1(3)\displaystyle M_{\Gamma_{1}(3)} \displaystyle\longrightarrow MΓ(3)\displaystyle M_{\Gamma(3)}
MΓ(3)\displaystyle M_{\Gamma(3)} \displaystyle\longrightarrow MΓ(n)\displaystyle M_{\Gamma(n)}
Proof.

This follows from [KM85, 5.5.1] since the moduli problem for Γ1(3)\Gamma_{1}(3) structures is representable. ∎

Lemma 5.5.

Let RR be coherent, MM be a finitely generated RR module and NN be a flat RR-module. Then we have the isomorphism

HomR(M,R)RNHomR(M,N).\operatorname{Hom}_{R}(M,R)\otimes_{R}N\cong\operatorname{Hom}_{R}(M,N).
Proof.

This result should be standard. Choose a finitely generated free presentation FF_{*} of MM. Since NN is flat the left hand side of the claim is the kernel of

HomR(F0,R)RNHomR(F1,R)RN.\operatorname{Hom}_{R}(F_{0},R)\otimes_{R}N\longrightarrow\operatorname{Hom}_{R}(F_{1},R)\otimes_{R}N.

By finiteness it coincides with the kernel of

HomR(F0,N)HomR(F1,N).\operatorname{Hom}_{R}(F_{0},N)\longrightarrow\operatorname{Hom}_{R}(F_{1},N).

which is the right hand side. ∎

Proposition 5.6.

The diagram

TMF(n)BSpin\textstyle{TMF(n)^{*}B{\mathit{S}pin}\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}λ\scriptstyle{\lambda}KTateBSpin\textstyle{K_{Tate}^{*}B{\mathit{S}pin}\ignorespaces\ignorespaces\ignorespaces\ignorespaces}H(BSpin,TMF(n)𝐐)\textstyle{H^{*}(B{\mathit{S}pin},TMF(n)^{*}_{\mathbf{Q}})\ignorespaces\ignorespaces\ignorespaces\ignorespaces}H(BSpin,KTate𝐐)\textstyle{H^{*}(B{\mathit{S}pin},{K_{Tate}}^{*}_{\mathbf{Q}})}

is a pullback for all nn divisible by 3.

Proof.

This follows as before from the universal coefficient isomorphism:

TMF(n)BSpin\displaystyle TMF(n)^{*}B{\mathit{S}pin} \displaystyle\cong TMF1(3)BSpin^MΓ1(3)MΓ(n)\displaystyle TMF_{1}(3)^{*}B{\mathit{S}pin}\hat{\otimes}_{M_{\Gamma_{1}(3)}}M_{\Gamma(n)}
\displaystyle\cong Hom(TMF1(3)BSpin,MΓ1(3))^MΓ1(3)MΓ(n)\displaystyle\operatorname{Hom}(TMF_{1}(3)_{*}B{\mathit{S}pin},M_{\Gamma_{1}(3)})\hat{{\otimes}}_{M_{\Gamma_{1}(3)}}M_{\Gamma(n)}
\displaystyle\cong Hom(TMF1(3)BSpin,MΓ(n)).\displaystyle\operatorname{Hom}(TMF_{1}(3)_{*}B{\mathit{S}pin},M_{\Gamma(n)}).

Here, the first isomorphism holds for all finite spectra by Lemma 5.4. (The completion of the tensor product is then unnecessary). Taking inverse limits over all finite subspectra and using Corollary 3.11 we see that the isomorphism holds for BSpinB{\mathit{S}pin} for the completed tensor product. The second isomorphism follows again from Corollary 3.11. The last one is a consequence of Lemmas 5.4 and 5.5 when considering finite subspectra of BSpinB{\mathit{S}pin}. ∎

Proof of 5.3. .

Set E=TMF1(3)E=TMF_{1}(3). We first construct a specific invertible element SS in E0BStringE^{0}B{\mathit{S}tring} whose character is a Jacobi form of weight 0 and index 12\frac{1}{2}. Let x:MUEx:MU\rightarrow E be the orientation considered earlier. Its exponential is given in terms of the Weierstrass Φ\Phi-function by

Φ(τ,z)Φ(τ,ω)Φ(τ,zω)\frac{\Phi(\tau,z)\Phi(\tau,-\omega)}{\Phi(\tau,z-\omega)}

for the standard division point ω=2πi/3\omega=2\pi i/3 (see [HBJ92, 5.3, 6.4]). Let BU6BU\langle 6\rangle be the 5-connected cover of BUBU and MU6MU\langle 6\rangle be the corresponding Thom spectrum. Then EE admits two ring maps from MU6MU\langle 6\rangle: the one which factors through xx and the one which factors through the Witten orientation (Theorem 2.1(ii)). Using the Thom isomorphism we obtain a class

σxE0BU6\frac{\sigma}{x}\in E^{0}BU\langle 6\rangle

whose augmentation is 1. Let SS be the image of σx\frac{\sigma}{x} under the complexification map from BStringB{\mathit{S}tring} to BU6BU\langle 6\rangle. The character of SS is the function

Φ(τ,zω)Φ(τ,ω).\frac{\Phi(\tau,z-\omega)}{\Phi(\tau,-\omega)}.

One can check that this is a formal Jacobi function of weight 0 and index 12\frac{1}{2} either by direct calculation or one uses the fact that the Witten orientation σ\sigma comes from a formal Jacobi modular form of index 12\frac{1}{2} and xx is one of index 0 (see [Bry90, p.469]). Since σx\frac{\sigma}{x} has the inverse xσ\frac{x}{\sigma} the class SS is invertible.

Next, let VV be an irreducible positive energy representation in (Pm)Γ(n)(P_{m})_{\Gamma(n)} and let

V^=n[Vn]qn\hat{V}=\sum_{n}[V_{n}]q^{n}

be the the class in KTate0(BSpin)K_{Tate}^{0}(B{\mathit{S}pin}) constructed above. Without loss of generality let EE be LS/2TMF(n)L_{S{\mathbb{Z}}/2}TMF(n). Consider SS as a class in E0BStringE^{0}B{\mathit{S}tring}. Since the K(1)K(1)-homology of K(,3)K({\mathbb{Z}},3) vanishes (c.[RW80]) the 2-completed KTateK_{Tate}-cohomologies of BStringB{\mathit{S}tring} and BSpinB{\mathit{S}pin} coincide. Hence, we may consider the element

V^λ(S)mKTate0(BSpin).\hat{V}\lambda(S)^{-m}\in K_{Tate}^{0}(B{\mathit{S}pin}).

Since its character is a Jacobi form of index 0 and weight 0 we can apply Proposition 5.6 to obtain a unique class in E0BSpinE^{0}B{\mathit{S}pin} and hence in E0BStringE^{0}B{\mathit{S}tring}. Define φ[V]\varphi[V] as the product of this class with SmS^{m}.

References

  • [ABP66] D. W. Anderson, E. H. Brown, Jr., and F. P. Peterson, SU{\rm SU}-cobordism, KO{\rm KO}-characteristic numbers, and the Kervaire invariant, Ann. of Math. (2) 83 (1966), 54–67. MR 0189043 (32 #6470)
  • [AHR] M. Ando, M. Hopkins, and C. Rezk, Multiplicative orientations of ko-theory and of the spectrum of topological modular forms, preprint, 2014.
  • [AHS01] M. Ando, M. J. Hopkins, and N. P. Strickland, Elliptic spectra, the Witten genus and the theorem of the cube, Invent. Math. 146 (2001), no. 3, 595–687. MR 1869850 (2002g:55009)
  • [AS68] M. F. Atiyah and I. M. Singer, The index of elliptic operators. I, Ann. of Math. (2) 87 (1968), 484–530. MR 0236950 (38 #5243)
  • [AS69] M. F. Atiyah and G. B. Segal, Equivariant KK-theory and completion, J. Differential Geometry 3 (1969), 1–18. MR 0259946 (41 #4575)
  • [Bak94] Andrew Baker, Elliptic genera of level NN and elliptic cohomology, J. London Math. Soc. (2) 49 (1994), no. 3, 581–593. MR 1271552 (95e:55009)
  • [Bou79] A. K. Bousfield, The localization of spectra with respect to homology, Topology 18 (1979), no. 4, 257–281. MR 551009 (80m:55006)
  • [Bry90] Jean-Luc Brylinski, Representations of loop groups, Dirac operators on loop space, and modular forms, Topology 29 (1990), no. 4, 461–480. MR 1071369 (91j:58151)
  • [DR73] P. Deligne and M. Rapoport, Les schémas de modules de courbes elliptiques, Modular functions of one variable, II (Proc. Internat. Summer School, Univ. Antwerp, Antwerp, 1972), Springer, Berlin, 1973, pp. 143–316. Lecture Notes in Math., Vol. 349. MR 0337993 (49 #2762)
  • [Eis95] David Eisenbud, Commutative algebra, Graduate Texts in Mathematics, vol. 150, Springer-Verlag, New York, 1995, With a view toward algebraic geometry. MR 1322960 (97a:13001)
  • [Fra92] Jens Franke, On the construction of elliptic cohomology, Math. Nachr. 158 (1992), 43–65. MR 1235295 (94h:55007)
  • [Goe10] Paul G. Goerss, Topological modular forms [after Hopkins, Miller and Lurie], Astérisque (2010), no. 332, Exp. No. 1005, viii, 221–255, Séminaire Bourbaki. Volume 2008/2009. Exposés 997–1011. MR 2648680 (2011m:55003)
  • [HBJ92] Friedrich Hirzebruch, Thomas Berger, and Rainer Jung, Manifolds and modular forms, Aspects of Mathematics, E20, Friedr. Vieweg & Sohn, Braunschweig, 1992, With appendices by Nils-Peter Skoruppa and by Paul Baum. MR 1189136 (94d:57001)
  • [HL] Michael Hill and Tyler Lawson, Topological modular forms with level structure, arXiv:1312.7394 , 2013.
  • [HR95] Mark A. Hovey and Douglas C. Ravenel, The 77-connected cobordism ring at p=3p=3, Trans. Amer. Math. Soc. 347 (1995), no. 9, 3473–3502. MR 1297530 (95m:55008)
  • [HS99a] Mark Hovey and Hal Sadofsky, Invertible spectra in the E(n)E(n)-local stable homotopy category, J. London Math. Soc. (2) 60 (1999), no. 1, 284–302. MR 1722151 (2000h:55017)
  • [HS99b] Mark Hovey and Neil P. Strickland, Morava KK-theories and localisation, Mem. Amer. Math. Soc. 139 (1999), no. 666, viii+100. MR 1601906 (99b:55017)
  • [JW85] David Copeland Johnson and W. Stephen Wilson, The Brown-Peterson homology of elementary pp-groups, Amer. J. Math. 107 (1985), no. 2, 427–453. MR 784291 (86j:55008)
  • [KL02] Nitu Kitchloo and Gerd Laures, Real structures and Morava KK-theories, KK-Theory 25 (2002), no. 3, 201–214. MR 1909866 (2003i:55005)
  • [KLW04a] Nitu Kitchloo, Gerd Laures, and W. Stephen Wilson, The Morava KK-theory of spaces related to BOBO, Adv. Math. 189 (2004), no. 1, 192–236. MR 2093483 (2005k:55002)
  • [KLW04b] by same author, Splittings of bicommutative Hopf algebras, J. Pure Appl. Algebra 194 (2004), no. 1-2, 159–168. MR 2086079 (2005e:55019)
  • [KM85] Nicholas M. Katz and Barry Mazur, Arithmetic moduli of elliptic curves, Annals of Mathematics Studies, vol. 108, Princeton University Press, Princeton, NJ, 1985. MR 772569 (86i:11024)
  • [KW88] Victor G. Kac and Minoru Wakimoto, Modular and conformal invariance constraints in representation theory of affine algebras, Adv. in Math. 70 (1988), no. 2, 156–236. MR 954660 (89h:17036)
  • [Lau99] Gerd Laures, The topological qq-expansion principle, Topology 38 (1999), no. 2, 387–425. MR MR1660325 (2000c:55009)
  • [Lau04] by same author, K(1)K(1)-local topological modular forms, Invent. Math. 157 (2004), no. 2, 371–403. MR MR2076927 (2005h:55003)
  • [LN12] Tyler Lawson and Niko Naumann, Commutativity conditions for truncated Brown-Peterson spectra of height 2, J. Topol. 5 (2012), no. 1, 137–168. MR 2897051
  • [LO] Gerd Laures and Martin Olbermann, tmf0(3)tmf_{0}(3) characteristic classes of string bundles, arXiv:1403.7301, 2014.
  • [Mat] Akhil Mathew, The homology of tmftmf, arXiv:1305.6100, 2013.
  • [Mil89] Haynes Miller, The elliptic character and the Witten genus, Algebraic topology (Evanston, IL, 1988), Contemp. Math., vol. 96, Amer. Math. Soc., Providence, RI, 1989, pp. 281–289. MR 1022688 (90i:55005)
  • [MR09] Mark Mahowald and Charles Rezk, Topological modular forms of level 3, Pure Appl. Math. Q. 5 (2009), no. 2, Special Issue: In honor of Friedrich Hirzebruch. Part 1, 853–872. MR 2508904 (2010g:55010)
  • [PS86] Andrew Pressley and Graeme Segal, Loop groups, Oxford Mathematical Monographs, The Clarendon Press, Oxford University Press, New York, 1986, Oxford Science Publications. MR 900587 (88i:22049)
  • [RW80] Douglas C. Ravenel and W. Stephen Wilson, The Morava KK-theories of Eilenberg-Mac Lane spaces and the Conner-Floyd conjecture, Amer. J. Math. 102 (1980), no. 4, 691–748. MR 584466 (81i:55005)
  • [RWY98] Douglas C. Ravenel, W. Stephen Wilson, and Nobuaki Yagita, Brown-Peterson cohomology from Morava KK-theory, KK-Theory 15 (1998), no. 2, 147–199. MR 1648284 (2000d:55012)
  • [Su07] Hsin-hao Su, The E(1,2) cohomology of the Eilenberg-MacLane space K( Z ,3), ProQuest LLC, Ann Arbor, MI, 2007, Thesis (Ph.D.)–The Johns Hopkins University. MR 2709571
  • [Wit88] Edward Witten, The index of the Dirac operator in loop space, Elliptic curves and modular forms in algebraic topology (Princeton, NJ, 1986), Lecture Notes in Math., vol. 1326, Springer, Berlin, 1988, pp. 161–181. MR 970288