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

Orientations and Topological Modular Forms with Level Structure

Dylan Wilson111The author was supported by NSF grant DGE-1324585 while completing this work.
Abstract

Using the methods of Ando-Hopkins-Rezk, we describe the characteristic series arising from EE_{\infty}-genera valued in topological modular forms with level structure. We give examples of such series for tmf0(N)tmf_{0}(N) and show that the Ochanine genus comes from an EE_{\infty}-ring map. We also show that, away from 6, certain tmftmf orientations of MStringMString descend to orientations of MSpinMSpin.

Introduction

One of our jobs as topologists is to study invariants of manifolds. The most accessible of these are cobordism invariants, called genera, which are ring maps

ϕ:MSOR\phi:MSO_{*}\longrightarrow R

Here MSOMSO is the Thom spectrum with coefficients the ring of cobordism classes of oriented-manifolds. Precomposing with the forgetful map MUMSOMU_{*}\longrightarrow MSO_{*} this genus determines a formal group law over RR. If RR is a \mathbb{Q}-algebra, the genus is then determined by the logarithm of this formal group law which takes the form

logϕ(x)=i0ϕ(2i)2i+1x2i+1\log_{\phi}(x)=\sum_{i\geq 0}\frac{\phi(\mathbb{C}^{2i})}{2i+1}x^{2i+1}

or equivalently by the Hirzebruch characteristic series Kϕ(u)=ulogϕ1(u)K_{\phi}(u)=\dfrac{u}{\log_{\phi}^{-1}(u)}.

Example 0.1.

The LL-genus (i.e. the signature) with characteristic series

KSign(u)=exp(k22k+1(2k11)Bkkukk!)K_{\textup{Sign}}(u)=\exp\left(\sum_{k\geq 2}\frac{2^{k+1}(2^{k-1}-1)B_{k}}{k}\frac{u^{k}}{k!}\right)

where the Bernoulli numbes BkB_{k} are defined by the generating series k0Bkxk/k!=xex1\sum_{k\geq 0}B_{k}x^{k}/k!=\dfrac{x}{e^{x}-1}.

Example 0.2.

The A^\widehat{A}-genus, with characteristic series

KA^(u)=exp(k2Bkkukk!)K_{\widehat{A}}(u)=\exp\left(-\sum_{k\geq 2}\frac{B_{k}}{k}\frac{u^{k}}{k!}\right)

The first two genera have the useful property that they vanish on projective bundles associated to even-dimensional complex vector bundles. Such genera were classified in a beautiful theorem:

Theorem 0.3 (Ochanine).

The logarithm of a genus ϕ\phi that vanishes on projective bundles of even-dimensional complex vector bundles is always of the form

logϕ(x)=0xdu12δu2+ϵu4\textup{log}_{\phi}(x)=\int_{0}^{x}\frac{du}{\sqrt{1-2\delta u^{2}+\epsilon u^{4}}}

where δ,ϵR\delta,\epsilon\in R.

In particular, there is a universal such with target the ring [δ,ϵ]\mathbb{Q}[\delta,\epsilon] of level 2 modular forms over \mathbb{Q} called the Ochanine genus. The characteristic series was computed by Zagier [Zag88]:

KOch(u)=exp(k22G~kukk!)K_{\textup{Och}}(u)=\textup{exp}\left(\sum_{k\geq 2}2\widetilde{G}_{k}\frac{u^{k}}{k!}\right)

where the G~k\widetilde{G}_{k} are certain Eisenstein series of level 2.

Atiyah-Bott-Shapiro showed that the A^\widehat{A}-genus on Spin-manifolds arose from a finer, integral invariant on families, neatly expressed by a map of spectra

MSpinKO\textup{MSpin}\longrightarrow KO

and this motivated the search for a cohomology theory to be the target of the Ochanine genus. Such a theory was built by Landweber-Ravenel-Stong and dubbed elliptic cohomology. Around the same time, Witten introduced a genus on Spin-manifolds that took values in the ring of quasi-modular forms. He proved that when the Spin-manifold satisfies p1(TM)=0p_{1}(TM)=0 the genus is actually valued in modular forms. The characteristic series for this genus was also computed by Zagier [Zag88] in terms of the classical Eisenstein series GkG_{k}:

KWit(u)=exp(k22Gkukk!)K_{\textup{Wit}}(u)=\textup{exp}\left(\sum_{k\geq 2}2G_{k}\frac{u^{k}}{k!}\right)

The requirement on the first Pontryagin class comes from the fact that G2G_{2} is not a modular form.

In Hopkins’ 1994 ICM address he proposed that the Witten genus comes from a map of spectra into the then conjectural spectrum of topological modular forms

σ:MStringtmf\sigma:\textup{MString}\longrightarrow tmf

where MStringMString_{*} is the cobordism ring for Spin manifolds such that the generator p1/2H4(Bspin,)p_{1}/2\in H^{4}(Bspin,\mathbb{Z}) vanishes on their stable normal bundle.

Since then, work of Ando, Goerss, Hopkins, Mahowald, Rezk, Strickland, and others has culminated in a construction of this map of EE_{\infty}-ring spectra. Furthermore, Ando-Hopkins-Rezk [AHR10] determined all possible characteristic series of EE_{\infty}-genera on String manifolds valued in topological modular forms. Their method of proof sheds light on the appearance of the Bernoulli numbers and the Eisenstein series in the formulae for characteristic series above: it is crucial that both sequences can be pp-adically interpolated.

Their work settles the question of genera taking values in modular forms of level 1, but leaves open questions about genera valued in higher level modular forms, including the Ochanine genus and others constructed in work of Hirzebruch [Hir88].

Thanks to work of Hill and Lawson [HL13], we now have candidates for EE_{\infty}-rings that deserve to be called topological modular forms with level structure, and we can ask about constructing EE_{\infty}-maps out of Thom spectra

MGtmf(Γ)MG\longrightarrow tmf(\Gamma)

The purpose of this paper is to apply the machinery of [AHR10] and [ABG+08] to the case of topological modular forms with level structure and construct various genera. In order to state the main result we will recall a bit of notation. If ΓGL2(/N)\Gamma\subset GL_{2}(\mathbb{Z}/N) there is a notion of modular forms with level structure Γ\Gamma. When

Γ=Γ1(N):={(1b01):}\Gamma=\Gamma_{1}(N):=\left\{\begin{pmatrix}1&b\\ 0&1\end{pmatrix}:\right\}

then the group (/N)×(\mathbb{Z}/N)^{\times} acts on MF(Γ1(N))MF_{*}(\Gamma_{1}(N)) and there is a nice formula for the operation ψp\psi^{p} in terms of qq-expansions. Indeed, if gkMFk(Γ1(N))g_{k}\in MF_{k}(\Gamma_{1}(N)) has a qq-expansion given by gk(q)=anqng_{k}(q)=\sum a_{n}q^{n} and the action of p(/N)×p\in(\mathbb{Z}/N)^{\times} is given by

(gk)|p(q)=n0bnqn(g_{k})|_{\langle p\rangle}(q)=\sum_{n\geq 0}b_{n}q^{n}

then we have:

(ψpgk)(q)=pkn0bnqpn(\psi^{p}g_{k})(q)=p^{k}\sum_{n\geq 0}b_{n}q^{pn}

All of this will be reviewed in detail below, but for now we at least state the main result.

Theorem 0.4.

Let ΓGL2(/N)\Gamma\subset GL_{2}(\mathbb{Z}/N) be a level structure, and consider a sequence of modular forms {gk}k2MF(Γ)\{g_{k}\}_{k\geq 2}\in MF_{*}(\Gamma)\otimes\mathbb{Q}. Then

  1. 1.

    There exists an EE_{\infty}-ring map MString[1/N]tmf(Γ)\textup{MString}[1/N]\longrightarrow tmf(\Gamma) with associated characteristic series exp(2k4gkukk!)\exp(2\sum_{k\geq 4}g_{k}\frac{u^{k}}{k!}) if and only if

    1. (a)

      For each i2i\geq 2, g2i+1=0g_{2i+1}=0,

    2. (b)

      For every prime pNp\!\not|N and every unit λp×/{±1}\lambda\in\mathbb{Z}_{p}^{\times}/\{\pm 1\} the sequence of rational pp-adic modular forms {(1λk)(11pψp)gk}k4MFp,(Γ)\{(1-\lambda^{k})(1-\frac{1}{p}\psi^{p})g_{k}\}_{k\geq 4}\in MF_{p,*}(\Gamma)\otimes\mathbb{Q} satisfies the generalized Kummer congruences (2.24),

    3. (c)

      For every prime pNp\not|N, Tpgk=(1+pk1)gkT_{p}g_{k}=(1+p^{k-1})g_{k}, where TpT_{p} is the ppth Hecke operator,

    4. (d)

      We have the congruence gkGkg_{k}\equiv G_{k} mod MF(Γ,[1/N,ζN])MF_{*}(\Gamma,\mathbb{Z}[1/N,\zeta_{N}]), where GkG_{k} is the unnormalized Eisenstein series of weight kk.

  2. 2.

    Suppose that 2|N2|N. There exists an EE_{\infty}-ring map MSpin[1/N]tmf(Γ)\textup{MSpin}[1/N]\longrightarrow tmf(\Gamma) with associated characteristic series exp(2k2gkukk!)\exp(2\sum_{k\geq 2}g_{k}\frac{u^{k}}{k!}) if and only if the conditions (a)-(c) above are satisfied for k2k\geq 2 and we have the congruence

    gkG~k mod MF(Γ,[1/N,ζN])g_{k}\equiv\widetilde{G}_{k}\textup{ mod }MF_{*}(\Gamma,\mathbb{Z}[1/N,\zeta_{N}])

    where G~k\widetilde{G}_{k} is the level 2 modular form with qq-expansion G~k=Bk2k+n1qnd|n(1)n/ddk1\widetilde{G}_{k}=-\dfrac{B_{k}}{2k}+\sum_{n\geq 1}q^{n}\sum_{d|n}(-1)^{n/d}d^{k-1}.

In each instance the set of homotopy classes of EE_{\infty}-ring maps corresponding to a given characteristic series is a non-empty torsor for the group

[ΣKO2,LK(1)LK(2)tmf(Γ)2][\Sigma KO^{\wedge}_{2},L_{K(1)}L_{K(2)}tmf(\Gamma)^{\wedge}_{2}]

which has exponent at most 2. In particular, upon inverting 22, the homotopy class of an EE_{\infty}-genus valued in tmf(Γ)tmf(\Gamma) is determined by its characteristic series.

Corollary 0.5.

Up to homotopy, there is a unique EE_{\infty}-ring map MSpin[1/2]tmf0(2)\textup{MSpin}[1/2]\longrightarrow tmf_{0}(2) refining the Ochanine genus.

Remark 0.6.

This answers in the affirmative the Question 1.1 of [HL13].

A more careful analysis at the prime 22 reveals the following

Theorem 0.7.

Let MSwing denote the Thom spectrum associated to the fiber of the map

bspinw4Σ4H/2bspin\stackrel{{\scriptstyle w_{4}}}{{\longrightarrow}}\Sigma^{4}H\mathbb{Z}/2

Then Theorem 0.4 holds mutatis mutandis for EE_{\infty}-ring maps MSwing[1/N]tmf(Γ)\textup{MSwing}[1/N]\longrightarrow tmf(\Gamma) when 3|N3|N.

Remark 0.8.

It seems plausible that such orientations reproduce the real EE-theory orientations found in [KS04].

A curious corollary of the proof of the main theorem is the following

Theorem 0.9.

There is an EE_{\infty}-ring map

MSpintmf[1/6]\textup{MSpin}\longrightarrow tmf[1/6]
Remark 0.10.

No such map exists refining the Witten genus, even as homotopy commutative spectra. See Theorem 5.8 below.

We now turn to a more precise overview of our results and outline of the proofs. In §1 we give a quick introduction to topological modular forms and elliptic curves with level structure, and review the obstruction theory for EE_{\infty}-genera given in [ABG+08]. Our problem is to describe the space of dotted arrows

g\textstyle{g\ignorespaces\ignorespaces\ignorespaces\ignorespaces}gl1S\textstyle{gl_{1}S\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}gl1S/g\textstyle{gl_{1}S/g\ignorespaces\ignorespaces\ignorespaces\ignorespaces}gl1tmf(Γ)\textstyle{gl_{1}tmf(\Gamma)}

where gg is a delooping of either SpinSpin or StringString, gl1Rgl_{1}R denotes the spectrum of units for an EE_{\infty} ring, and the top line is a fiber sequence.

In §2 we analyze the substantially easier problem of parameterizing the space of dotted arrows

g\textstyle{g\ignorespaces\ignorespaces\ignorespaces\ignorespaces}gl1S\textstyle{gl_{1}S\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}gl1S/g\textstyle{gl_{1}S/g\ignorespaces\ignorespaces\ignorespaces\ignorespaces}LK(1)gl1tmf(Γ)p\textstyle{L_{K(1)}gl_{1}tmf(\Gamma)^{\wedge}_{p}}

where pp is a prime not dividing the level, and we have localized with respect to Morava KK-theory. Following Ando-Hopkins-Rezk, we use the solution to the Adams conjecture to reduce this to a computation of LK(1)tmf(Γ)0(bspin)L_{K(1)}tmf(\Gamma)^{0}(bspin) which we give in terms of measures on p×/{±1}\mathbb{Z}_{p}^{\times}/\{\pm 1\} valued in pp-adic modular forms of level Γ\Gamma. This sheds light on the appearance of Eisenstein series and Bernoulli numbers in the characteristic series of these genera. We have included an especially detailed description of this computation as it has not yet appeared elsewhere in the literature, even for the case of tmftmf.

In §3 we review Rezk’s logarithm, construct a topological lift of the Atkin operator, and describe the space of dotted arrows

g\textstyle{g\ignorespaces\ignorespaces\ignorespaces\ignorespaces}gl1S\textstyle{gl_{1}S\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}gl1S/g\textstyle{gl_{1}S/g\ignorespaces\ignorespaces\ignorespaces\ignorespaces}LK(1)K(2)gl1tmf(Γ)p\textstyle{L_{K(1)\vee K(2)}gl_{1}tmf(\Gamma)^{\wedge}_{p}}

as certain measures valued in pp-adic modular forms of level Γ\Gamma which are fixed by the Atkin operator. Up to now there is no significant difference from the case of tmftmf. When replacing the K(1)K(2)K(1)\vee K(2)-local case with the pp-complete case, however, our story diverges from that of [AHR10]. Indeed, the difference is measured by the fiber FF in the sequence

Fgl1tmf(Γ)pLK(1)K(2)gl1tmf(Γ)pF\longrightarrow gl_{1}tmf(\Gamma)^{\wedge}_{p}\longrightarrow L_{K(1)\vee K(2)}gl_{1}tmf(\Gamma)^{\wedge}_{p}

and the main obstruction to an orientation is the map

[g,F][g,gl1tmf(Γ)p][g,F]\longrightarrow[g,gl_{1}tmf(\Gamma)^{\wedge}_{p}]

When g=stringg=string we are in the case analyzed by [AHR10] and one can prove that the group [g,F][g,F] is zero. However, when g=sping=spin, this group never vanishes, so one must analyze the map itself. In §4 we investigate the spectrum FF and the above map. Finally, in §5 we prove the main theorem and build the examples described above. We expect the same methods to produce EE_{\infty}-ring maps

MUtmf1(N)\textup{MU}\longrightarrow tmf_{1}(N)

refining Hirzebruch’s level NN genera.

Acknowledgements. I would like to thank Tyler Lawson for asking me whether the Ochanine genus comes from an EE_{\infty}-ring map. I owe a huge intellectual debt to Matt Ando, Mike Hopkins, and Charles Rezk. They were each very helpful and graciously answered my questions about their paper. I would also like to thank Joel Specter and Frank Calegari for patiently explaining the theory of pp-adic modular forms. Finally, this paper would not exist without the help of my advisor, Paul Goerss, to whom I am very grateful.

1 Background on Orientations and Level Structures

1.1 Orientations

The spectrum analog of the group algebra functor, Σ+Ω\Sigma_{+}^{\infty}\Omega^{\infty}, participates in an adjunction ([ABG+14b, 5.2])

Σ+Ω:Spectra0E-Ring:gl1\Sigma_{+}^{\infty}\Omega^{\infty}:\textup{Spectra}_{\geq 0}\leftrightarrows E_{\infty}\textup{-Ring}:gl_{1}

from connective spectra to EE_{\infty}-rings. If XX is a space and RR an EE_{\infty}-ring, then

gl1(R)0(X)=R0(X)×gl_{1}(R)^{0}(X)=R^{0}(X)^{\times}

In the special case when X=SnX=S^{n}, n>0n>0, and we use reduced homology we get an identification of πn(gl1R)\pi_{n}(gl_{1}R) with the group

{1+x(πnR)×:xπnR}.\{1+x\in\left(\pi_{n}R\right)^{\times}:x\in\pi_{n}R\}.

We will make heavy use of the fact that the assignment 1+xx1+x\mapsto x gives an equivalence

πqgl1RπqR,q>0\pi_{q}gl_{1}R\cong\pi_{q}R,\quad q>0

though this is not induced by a map of spectra in general.

For a connective spectrum gg, we will occasionally denote the suspension Σg\Sigma g by bgbg, and their zeroth space by GG and BGBG respectively. (The only exceptions to this convention are the connective KK-theory spectra bobo and bubu whose zeroth spaces are BO×BO\times\mathbb{Z} and BU×BU\times\mathbb{Z}, respectively.)

Given a map

f:bgbgl1Sf:bg\longrightarrow bgl_{1}S

we can form the associated Thom spectrum, which can be defined as the homotopy pushout of EE_{\infty}-rings

Σ+BG\textstyle{\Sigma_{+}^{\infty}BG\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}Σ+BGL1S\textstyle{\Sigma_{+}^{\infty}BGL_{1}S\ignorespaces\ignorespaces\ignorespaces\ignorespaces}S\textstyle{S\ignorespaces\ignorespaces\ignorespaces\ignorespaces}MG\textstyle{MG}

One can show that the underlying spectrum agrees with the usual notion of Thom spectrum in the examples of interest (see, e.g., the comparison results in [ABG+14a, §3] for the connection with the Thom spectra constructed in Lewis’s thesis [LMSM86, Ch. IX]). An immediate corollary of this definition is the following:

Theorem 1.1 ([ABG+08, 4.3]).

Let RR be an EE_{\infty}-ring with unit map ι:SR\iota:S\longrightarrow R. Then there is a homotopy pullback diagram of spaces

MapCAlg(MG,R)\textstyle{\textup{Map}_{\textup{CAlg}}(MG,R)\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}MapSpectra(gl1S/g,gl1R)\textstyle{\textup{Map}_{\textup{Spectra}}(gl_{1}S/g,gl_{1}R)\ignorespaces\ignorespaces\ignorespaces\ignorespaces}{gl1ι}\textstyle{\{gl_{1}\iota\}\ignorespaces\ignorespaces\ignorespaces\ignorespaces}MapSpectra(gl1S,gl1R)\textstyle{\textup{Map}_{\textup{Spectra}}(gl_{1}S,gl_{1}R)}

and any choice of point in MapCAlg(MG,R)\textup{Map}_{\textup{CAlg}}(MG,R) gives an equivalence

MapSpectra(bg,gl1R)MapCAlg(MG,R)\textup{Map}_{\textup{Spectra}}(bg,gl_{1}R)\stackrel{{\scriptstyle\cong}}{{\longrightarrow}}\textup{Map}_{\textup{CAlg}}(MG,R)

In practice, we interpret this theorem as the statement that the space of EE_{\infty} RR-orientations of MGMG is the space of dotted arrows in the diagram

g\textstyle{g\ignorespaces\ignorespaces\ignorespaces\ignorespaces}gl1S\textstyle{gl_{1}S\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}gl1S/g\textstyle{gl_{1}S/g\ignorespaces\ignorespaces\ignorespaces\ignorespaces}gl1R\textstyle{gl_{1}R}

In particular, the obstruction to the existence of an orientation is a class in [g,gl1R][g,gl_{1}R]. We will have occasion to consider more general targets in this diagram, so we introduce a notation for this.

Definition 1.2.

Let ϵ:gl1SX\epsilon:gl_{1}S\longrightarrow X be a spectrum under gl1Sgl_{1}S, then we denote by Nulls(g,X)\textup{Nulls}(g,X) the homotopy pullback

Nulls(g,X)\textstyle{\textup{Nulls}(g,X)\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}MapSpectra(gl1S/g,X)\textstyle{\textup{Map}_{\textup{Spectra}}(gl_{1}S/g,X)\ignorespaces\ignorespaces\ignorespaces\ignorespaces}{ϵ}\textstyle{\{\epsilon\}\ignorespaces\ignorespaces\ignorespaces\ignorespaces}MapSpectra(gl1S,X)\textstyle{\textup{Map}_{\textup{Spectra}}(gl_{1}S,X)}

In particular, MapCAlg(MG,R)=Nulls(g,gl1R)\textup{Map}_{CAlg}(MG,R)=\textup{Nulls}(g,gl_{1}R).

By construction this gives a homotopy limit preserving functor

Nulls(g,):Spectragl1S/Spaces\textup{Nulls}(g,-):\textup{Spectra}_{gl_{1}S/}\longrightarrow\textup{Spaces}

For the remainder of the paper we will assume that we are in the geometric situation where the map bgbgl1Sbg\longrightarrow bgl_{1}S factors through the spectrum bsobso. In this setting we can recover Hirzebruch’s characteristic series when RR is a rational EE_{\infty}-ring. Indeed, in this case, we have a canonical orientation

MGMSOHSRMG\longrightarrow MSO\longrightarrow H\mathbb{Q}\stackrel{{\scriptstyle\cong}}{{\leftarrow}}S_{\mathbb{Q}}\longrightarrow R

Thus, by Theorem 1.1, any other element in π0MapCAlg(MG,R)\pi_{0}\textup{Map}_{\textup{CAlg}}(MG,R) is specified by an element in [bg,gl1R][bg,gl_{1}R]. However, since RR is rational, there is a logarithm

0:τ1(gl1R)τ1R\ell_{0}:\tau_{\geq 1}(gl_{1}R)\stackrel{{\scriptstyle\cong}}{{\longrightarrow}}\tau_{\geq 1}R

where τn\tau_{\geq n} denotes the truncation functor from spectra to spectra with homotopy concentrated in degrees at least nn. Since gg is connective we deduce that orientations are determined by difference classes in [bg,R][bg,R].

For n1n\geq 1 we denote by bo2n:=τ2nbobo\langle 2n\rangle:=\tau_{\geq 2n}bo the 2n2n-connective cover of bobo. We also observe the following conventions:

bso:=bo2,bspin:=bo4,bstring:=bo8Σ8bobso:=bo\langle 2\rangle,\quad bspin:=bo\langle 4\rangle,\quad bstring:=bo\langle 8\rangle\cong\Sigma^{8}bo

Recall that we have

πbu=[v],πbo=[η,x4,v]/(2η,ηx4,x424v,η3)\pi_{*}bu=\mathbb{Z}[v],\quad\pi_{*}bo=\mathbb{Z}[\eta,x_{4},v_{\mathbb{R}}]/(2\eta,\eta x_{4},x_{4}^{2}-4v_{\mathbb{R}},\eta^{3})

Complexification of vector bundles induces a map c:bobuc:bo\longrightarrow bu, and forgetting structure induces a map (which does not give a ring homomorphism on homotopy groups) r:bubor:bu\longrightarrow bo. On homotopy these are

cv=v,cx4=2v2andrv4k=2vk,rv4k+2=x4vk,rvodd=0c_{*}v_{\mathbb{R}}=v,\,\,c_{*}x_{4}=2v^{2}\quad\textup{and}\quad r_{*}v^{4}k=2v_{\mathbb{R}}^{k},\,\,r_{*}v^{4k+2}=x_{4}v_{\mathbb{R}}^{k},\,\,r_{*}v^{\textup{odd}}=0
Theorem 1.3 ([AHR10]).

Let RR be a rational EE_{\infty}-ring and n4n\leq 4. There is a natural isomorphism

π0MapCAlg(MO2n,R)knπ2kR\pi_{0}\textup{Map}_{\textup{CAlg}}(MO\langle 2n\rangle,R)\stackrel{{\scriptstyle\cong}}{{\longrightarrow}}\prod_{k\geq n}\pi_{2k}R

where the kkth component tkt_{k} is given by

S2kvkbu2nbo2nδτ1gl1Rτ1RS^{2k}\stackrel{{\scriptstyle v^{k}}}{{\longrightarrow}}bu\langle 2n\rangle\longrightarrow bo\langle 2n\rangle\stackrel{{\scriptstyle\delta}}{{\longrightarrow}}\tau_{\geq 1}gl_{1}R\cong\tau_{\geq 1}R

(which is zero if kk is odd). The Hirzebruch characteristic series is then K(u)=exp(kntkukk!)K(u)=\exp\left(\sum_{k\geq n}t_{k}\dfrac{u^{k}}{k!}\right).

We now return to the case where RR is an arbitrary EE_{\infty}-ring. The localization RRR\longrightarrow R_{\mathbb{Q}} gives a map

π0MapCAlg(MO2n,R)knπ2kR\pi_{0}\textup{Map}_{\textup{CAlg}}(MO\langle 2n\rangle,R)\longrightarrow\prod_{k\geq n}\pi_{2k}R\otimes\mathbb{Q}

It turns out ([AHR10, 5.14]) that the elements {(tk/2)}:={bk}π2kR\{(t_{k}/2)\}:=\{b_{k}\}\in\pi_{2k}R\otimes\mathbb{Q} are independent of the orientation modulo \mathbb{Z}, which motivates the following.

Definition 1.4.

We denote the image of the above map postcomposed with division by 2 as Char(Spin,R)\textup{Char}(\textup{Spin},R) when n=2n=2. This is the set of characteristic series of EE_{\infty} Spin genera valued in RR under the correspondence {bk}exp(2bkuk/k!)\{b_{k}\}\mapsto\exp\left(2\sum b_{k}u^{k}/k!\right). We define Char(String,R)\textup{Char}(\textup{String},R) analogously for n=4n=4.

1.2 Level Structures and Modular Forms

This section exists mainly to establish notation and provide references for precise definitions and results. Intimate knowledge of algebraic stacks is not necessary to understand the main results of the paper, nor their proofs. Our main reference is [Con07], but most of the results and definitions were originally written down in [KM85] and [DR73].

A generalized elliptic curve [Con07, 2.1.4] over a base scheme SS is a separated, flat map π:ES\pi:E\longrightarrow S such that all geometric fibers are either elliptic curves or Néron polygons [DR73, II.1.1] together with a morphism +:Esm×SEE+:E^{\textup{sm}}\times_{S}E\longrightarrow E and a section eEsm(S)e\in E^{\textup{sm}}(S) in the smooth locus. We require that this data gives EsmE^{\textup{sm}} the structure of a commutative group scheme with identity ee and defines an action of EsmE^{\textup{sm}} on EE that acts by rotations on the graph of irreducible components on each singular fiber.

A Γ(N)\Gamma(N)-structure [Con07, 2.4.2] on a generalized elliptic curve is an ordered pair (P,Q)(P,Q) with P,QEsm[N](S)P,Q\in E^{\textup{sm}}[N](S) such that

  1. 1.

    the rank N2N^{2} Cartier divisor

    D=i,j/N[iP+jQ]D=\sum_{i,j\in\mathbb{Z}/N\mathbb{Z}}[iP+jQ]

    is a subgroup scheme killed by NN

  2. 2.

    DD meets all irreducible components of all geometric fibers of EE.

In the case where E=EsmE=E^{\textup{sm}} and NN is invertible in 𝒪S\mathcal{O}_{S}, this is equivalent to the data of an isomorphism of constant group schemes (/N)2E[N](\mathbb{Z}/N\mathbb{Z})^{2}\cong E[N] (after finite étale base change).

Remark 1.5.

When doing homotopy theory, we will almost always be interested in the case when NN is invertible, since we must invert NN to get EE_{\infty}-versions of these moduli problems. However, to define Hecke operators (i.e. power operations) as in §3.1-3.2, we pass through moduli problems defined over bad primes. It is possible to avoid these more arithmetic moduli entirely through ad-hoc constructions, but the author feels the present approach is more conceptual.

Theorem 1.6 (Katz-Mazur, Deligne-Rappaport, Conrad).

The moduli problem of generalized elliptic curves with Γ(N)\Gamma(N)-structure is representable by a proper, flat Deligne-Mumford stack Γ(N)\mathscr{M}_{\Gamma(N)} which is CM of relative dimension 1 over \mathbb{Z}. Moreover, the restriction to [1/N]\mathbb{Z}[1/N] is smooth.

For each N1N\geq 1, there is a finite, faithfully flat map [Con07, 4.1.1]

Γ(N)Γ(1):=ell\mathscr{M}_{\Gamma(N)}\longrightarrow\mathscr{M}_{\Gamma(1)}:=\mathscr{M}_{\textup{ell}}

which forgets the data of PP and QQ and collapses Néron nn-gon fibers to a nodal cubic fibers.

There is an action of GL2(/N)GL_{2}(\mathbb{Z}/N\mathbb{Z}) on the universal generalized elliptic curve with Γ(N)\Gamma(N)-structure Γ(N)Γ(N)\mathscr{E}_{\Gamma(N)}\longrightarrow\mathscr{M}_{\Gamma(N)} given by modifying the points (P,Q)(P,Q). Unless otherwise specified, from now on Γ\Gamma will denote one of the following subgroups:

Γ1(N):={(10)},Γ0(N):={(0)}\Gamma_{1}(N):=\left\{\begin{pmatrix}1&*\\ 0&*\end{pmatrix}\right\},\quad\quad\Gamma_{0}(N):=\left\{\begin{pmatrix}*&*\\ 0&*\end{pmatrix}\right\}

When NN is square-free, we denote by Γ\mathscr{M}_{\Gamma} the normalization of Γ(1)\mathscr{M}_{\Gamma(1)} in the quotient stack Γ(N)[1/N]/Γ\mathscr{M}^{\circ}_{\Gamma(N)}[1/N]/\Gamma. Here Γ(N)\mathscr{M}^{\circ}_{\Gamma(N)} denotes the open complement of the nonsingular locus of the universal generalized elliptic curve. When NN is not square-free, one can still produce algebraic stacks by giving a modular description and proving representability, but we will not need these here.

These stacks classify generalized elliptic curves together with a point of ‘exact order NN’ in the sense of Drinfeld (resp. a subgroup scheme of order NN) which must satisfy conditions on the singular fibers [Con07, 4.1.5]. For smooth elliptic curves over a base where NN acts invertibly, one can remove the quotation marks. While Γ1(N)\mathscr{M}_{\Gamma_{1}(N)} is a Deligne-Mumford stack, the same is not true in general for Γ0(N)\mathscr{M}_{\Gamma_{0}(N)}, though it is so if NN is square-free or inverted [Con07, 3.1.7].

Let Γ(N)Γ(N)\mathscr{E}_{\Gamma(N)}\longrightarrow\mathscr{M}_{\Gamma(N)} denote the universal generalized elliptic curve. Denote by ωΓ(N)\omega_{\Gamma(N)} (or just ω\omega if there is no confusion) the relative dualizing sheaf for the map Γ(N)Γ(N)\mathscr{E}_{\Gamma(N)}\longrightarrow\mathscr{M}_{\Gamma(N)}. This turns out to be a line bundle, and can be equivalently defined as the dual of the Lie algebra of Γ(N)sm\mathscr{E}_{\Gamma(N)}^{\textup{sm}}, i.e. e(ΩΓ(N)/Γ(N)1)e^{*}(\Omega^{1}_{\mathscr{E}_{\Gamma(N)}/\mathscr{M}_{\Gamma(N)}}).

Definition 1.7.

The group of modular forms of level Γ\Gamma and weight kk over a ring RR is defined as

Mk(Γ,R):=H0(ΓR,ωk)M_{k}(\Gamma,R):=H^{0}(\mathscr{M}_{\Gamma}\otimes R,\omega^{\otimes k})

By flat base change, if SS is a flat RR-algebra, we have

Mk(Γ,S)=Mk(Γ,R)RSM_{k}(\Gamma,S)=M_{k}(\Gamma,R)\otimes_{R}S

In particular, Mk(Γ,)Mk(Γ,)M_{k}(\Gamma,\mathbb{Z})\otimes\mathbb{C}\cong M_{k}(\Gamma,\mathbb{C}) and one can check that this latter group recovers the classical group of holomorphic modular forms.

The Fourier expansion of a holomorphic modular form can be recovered algebraically using the theory of Tate curves. First, one may define an elliptic curve of the form 𝔾m/q\mathbb{G}_{m}/q^{\mathbb{Z}} over ((q))\mathbb{Z}((q)). This has a unique extension to a generalized elliptic curve over [[q]]\mathbb{Z}[[q]] with special fiber a nodal curve. This is called the Tate curve, Tate(q)\textup{Tate}(q), and comes with a canonical isomorphism of formal groups

𝔾^mTate(q)q=0\widehat{\mathbb{G}}_{m}\cong\textup{Tate}(q)^{\wedge}_{q=0}

For each N1N\geq 1, there is a unique generalized elliptic curve TateN\textup{Tate}_{N} over [[q1/N]]\mathbb{Z}[[q^{1/N}]] restricting to Tate1[[q]]((q1/N))\textup{Tate}_{1}\otimes_{\mathbb{Z}[[q]]}\mathbb{Z}((q^{1/N})) over ((q1/N))\mathbb{Z}((q^{1/N})) and having nn-gon fiber at q1/N=0q^{1/N}=0. This curve comes with a canonical Γ1(N)\Gamma_{1}(N) structure (corresponding to q1/Nq^{1/N}) classified by a map

Spec([[q1/N]])Γ1(N)\textup{Spec}(\mathbb{Z}[[q^{1/N}]])\longrightarrow\mathscr{M}_{\Gamma_{1}(N)}

Evaluation of a modular form at this curve is called the evaluation at the cusp, and, using the canonical trivialization of ω\omega on the Tate curve, corresponds to a power series in qq. This agrees with the Fourier expansion over \mathbb{C}.

The map to the fpqcfpqc algebraic stack of formal groups

ΓFG\mathscr{M}_{\Gamma}\longrightarrow\mathscr{M}_{FG}

classifying the formal group associated to the universal generalized elliptic curve is flat and so defines a diagram of weakly even periodic Landweber exact cohomology theories parameterized by flat maps

Spec(R)Γ\textup{Spec}(R)\longrightarrow\mathscr{M}_{\Gamma}
Theorem 1.8 (Goerss-Hopkins-Miller, Hill-Lawson).

There is a sheaf of EE_{\infty}-rings 𝒪top\mathcal{O}^{\textup{top}} on the étale site of Γ[1/N]\mathscr{M}_{\Gamma}[1/N] rigidifying the given diagram of Landweber exact cohomology theories.

We adopt the following notations:

TMF(Γ):=𝒪top(Γ[1/N]),Tmf(Γ):=𝒪top(Γ[1/N]),tmf(Γ):=τ0Tmf(Γ)TMF(\Gamma):=\mathcal{O}^{\textup{top}}(\mathscr{M}^{\circ}_{\Gamma}[1/N]),\quad Tmf(\Gamma):=\mathcal{O}^{\textup{top}}(\mathscr{M}_{\Gamma}[1/N]),\quad tmf(\Gamma):=\tau_{\geq 0}Tmf(\Gamma)
Remark 1.9.

The definition of tmf(Γ)tmf(\Gamma) as a connective cover is known to be unsatisfactory when the genus of the modular curve Γ\mathscr{M}_{\Gamma} is bigger than one. This doesn’t affect the construction of our genera, but instead suggests that, in these cases, the genera should have more severe restrictions on their image.

2 The K(1)K(1)-local Case

Fix a level Γ{Γ0(N),Γ1(N)}\Gamma\in\{\Gamma_{0}(N),\Gamma_{1}(N)\} and a prime pp not dividing NN. The goal of this section is to prove the following theorem. All the undefined terms and notation will be explained below.

Theorem 2.1.

The image of the map

π0Nulls(spin,LK(1)gl1tmf(Γ)p)k2MFp,k(Γ)pp\pi_{0}\textup{Nulls}(spin,L_{K(1)}gl_{1}tmf(\Gamma)^{\wedge}_{p})\longrightarrow\prod_{k\geq 2}MF_{p,k}(\Gamma)\otimes_{\mathbb{Z}_{p}}\mathbb{Q}_{p}

is an isomorphism onto the set of sequences {gk}k2\{g_{k}\}_{k\geq 2} of pp-adic modular forms satisfying the following conditions:

  1. 1.

    For all k1k\geq 1, g2k+1=0g_{2k+1}=0,

  2. 2.

    For all cp×c\in\mathbb{Z}_{p}^{\times}, the sequence {(1ck)gk(p)}\{(1-c^{k})g^{(p)}_{k}\} extends to a measure μc\mu_{c} on p×/{±1}\mathbb{Z}^{\times}_{p}/\{\pm 1\} valued in V(Γ)V_{\infty}(\Gamma),

  3. 3.

    For all cp×c\in\mathbb{Z}_{p}^{\times}, we have p×/{±1}1𝑑μc=12plog(cp1)\int_{\mathbb{Z}^{\times}_{p}/\{\pm 1\}}1\,d\mu_{c}=\frac{1}{2p}\log(c^{p-1}).

Remark 2.2.

As a corollary of 4.1, below, we have Nulls(spin,LK(1)gl1tmf(Γ)p)Nulls(spin,gl1LK(1)tmf(Γ)p)\textup{Nulls}(spin,L_{K(1)}gl_{1}tmf(\Gamma)^{\wedge}_{p})\cong\textup{Nulls}(spin,gl_{1}L_{K(1)}tmf(\Gamma)^{\wedge}_{p}) so this is actually a description of K(1)K(1)-local orientations.

In order to prove this theorem we need a good understanding of [gl1S/spin,LK(1)gl1tmf][gl_{1}S/spin,L_{K(1)}gl_{1}tmf]. It turns out that, as a consequence of the Adams conjecture and the computation of the K(1)K(1)-local sphere by Adams and Mahowald, we have

LK(1)(gl1/spin)KOpL_{K(1)}(gl_{1}/spin)\cong KO_{p}

Together with Rezk’s logarithm, this reduces much of the work to understanding [KOp,LK(1)tmf(Γ)][KO_{p},L_{K(1)}tmf(\Gamma)]. We will compute this in §2.2 using the K(1)K(1)-local Adams-Novikov spectral sequences, after collecting the necessary preliminaries on pp-adic modular forms in §2.1. Finally, we will translate our computation into the language of measures and complete the proof following [AHR10].

2.1 Complements on pp-adic modular forms

For this section we fix an integer N1N\geq 1, Γ{Γ(N),Γ1(N),Γ0(N)}\Gamma\in\{\Gamma(N),\Gamma_{1}(N),\Gamma_{0}(N)\}, and a prime pp not dividing NN. The Hasse invariant, AA, is a level 1 modular form that takes the value 11 on the mod pp Tate curve. Thus, from the point of view of qq-expansions, AA behaves like a unit mod pp. It is essentially for this reason that the theory of pp-adic modular forms requires that we invert a lift of (some power) of the Hasse invariant. This corresponds to restricting attention to those elliptic curves with height 1 formal group.

Recall that we have a map of fpqcfpqc-stacks

ΓFG\mathscr{M}_{\Gamma}\longrightarrow\mathscr{M}_{FG}

classifying the formal group of the universal generalized elliptic curve. Let FGnFG(p)\mathscr{M}_{FG}^{\geq n}\subset\mathscr{M}_{FG}\otimes\mathbb{Z}_{(p)} denote the regularly embedded, closed substack classifying formal groups of height at least nn, and denote by FG<n\mathscr{M}^{<n}_{FG} the open complement.

We define the ordinary locus of (Γ)p\left(\mathscr{M}_{\Gamma}\right)^{\wedge}_{p} by the pullback of formal stacks

Γord\textstyle{\mathscr{M}^{\textup{ord}}_{\Gamma}\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}(Γ)p\textstyle{\left(\mathscr{M}_{\Gamma}\right)^{\wedge}_{p}\ignorespaces\ignorespaces\ignorespaces\ignorespaces}FG<2\textstyle{\mathscr{M}^{<2}_{FG}\ignorespaces\ignorespaces\ignorespaces\ignorespaces}(FG)p\textstyle{\left(\mathscr{M}_{FG}\right)^{\wedge}_{p}}

This classifies elliptic curves over complete, local Noetherian algebras with residue field of characteristic pp whose formal group is a deformation of a height 1 formal group. It is the open complement of the supersingular locus (the pullback of FG2\mathscr{M}^{\geq 2}_{FG}).

Definition 2.3.

Let RR be a pp-complete p\mathbb{Z}_{p}-algebra. We define the ring of pp-adic modular forms of level Γ\Gamma and weight kk over RR by

MFp,k(Γ,R):=H0(Γord^R,ωk)MF_{p,k}(\Gamma,R):=H^{0}(\mathscr{M}^{\textup{ord}}_{\Gamma}\widehat{\otimes}R,\omega^{\otimes k})

Concretely, a pp-adic modular form over RR is a rule ff which assigns to each triple (R,E,α)(R^{\prime},E,\alpha) consisting of a pp-complete RR-algebra, a generalized ordinary elliptic curve EE over RR^{\prime}, and a level Γ\Gamma structure α\alpha, an element f(R,E,α)H0(E,ωk)f(R^{\prime},E,\alpha)\in H^{0}(E,\omega^{\otimes k}). Furthermore, this assignment only depends on the isomorphism class of (E,α)(E,\alpha) and commutes with extension of scalars.

By Lubin-Tate theory, FG<2\mathscr{M}^{<2}_{FG} admits a pro-Galois cover by Spf(p)\textup{Spf}(\mathbb{Z}_{p}),

Spf(p)FG<2\textup{Spf}(\mathbb{Z}_{p})\longrightarrow\mathscr{M}^{<2}_{FG}

classifying the multiplicative group over p\mathbb{Z}_{p}. In particular, this is a torsor for the group p×\mathbb{Z}^{\times}_{p}, which means that the shearing map

Spf(C(p×,p))×Spf(p)Spf(p)Spf(p)×FG<2Spf(p)\textup{Spf}(C(\mathbb{Z}_{p}^{\times},\mathbb{Z}_{p}))\times_{\textup{Spf}(\mathbb{Z}_{p})}\textup{Spf}(\mathbb{Z}_{p})\longrightarrow\textup{Spf}(\mathbb{Z}_{p})\times_{\mathscr{M}^{<2}_{FG}}\textup{Spf}(\mathbb{Z}_{p})

is an isomorphism. The subgroups 1+pnpp×1+p^{n}\mathbb{Z}_{p}\subset\mathbb{Z}^{\times}_{p} give intermediate covers

FG(pn)FG<2\mathscr{M}_{FG}(p^{n})\longrightarrow\mathscr{M}^{<2}_{FG}

which are Galois with Galois group (/pn)×(\mathbb{Z}/p^{n})^{\times}. Pulling back this tower gives a diagram

Γtriv\textstyle{\mathscr{M}^{\textup{triv}}_{\Gamma}\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}Spf(p)\textstyle{\textup{Spf}(\mathbb{Z}_{p})\ignorespaces\ignorespaces\ignorespaces\ignorespaces}\textstyle{\vdots\ignorespaces\ignorespaces\ignorespaces\ignorespaces}\textstyle{\vdots\ignorespaces\ignorespaces\ignorespaces\ignorespaces}Γ(pn)\textstyle{\mathscr{M}_{\Gamma}(p^{n})\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}FG(pn)\textstyle{\mathscr{M}_{FG}(p^{n})\ignorespaces\ignorespaces\ignorespaces\ignorespaces}\textstyle{\vdots\ignorespaces\ignorespaces\ignorespaces\ignorespaces}\textstyle{\vdots\ignorespaces\ignorespaces\ignorespaces\ignorespaces}Γord\textstyle{\mathscr{M}^{\textup{ord}}_{\Gamma}\ignorespaces\ignorespaces\ignorespaces\ignorespaces}FG<2\textstyle{\mathscr{M}^{<2}_{FG}}

The left hand tower is known as the Igusa tower. The formal stack Γ(pn)\mathscr{M}_{\Gamma}(p^{n}) classifies triples (E,α,η)(E,\alpha,\eta) where (E,α)(E,\alpha) is as before and η:μpnE[pn]\eta:\mu_{p^{n}}\hookrightarrow E[p^{n}] is an injection of pp-divisible groups. The formal stack Γtriv\mathscr{M}^{\textup{triv}}_{\Gamma} parameterizes triples (E,α,η)(E,\alpha,\eta) where η:𝔾^mE^\eta:\widehat{\mathbb{G}}_{m}\stackrel{{\scriptstyle\cong}}{{\longrightarrow}}\widehat{E} is a trivialization of the formal group. We record a few facts about this tower.

Proposition 2.4.

The stack Γ(pn)\mathscr{M}_{\Gamma}(p^{n}) is formally affine for pp odd and n1n\geq 1 and p=2p=2 for n2n\geq 2. We denote the corresponding pp-complete rings by Vn(Γ)V_{n}(\Gamma). In particular Γtriv\mathscr{M}^{\textup{triv}}_{\Gamma} is formally affine with coordinate ring V(Γ)V_{\infty}(\Gamma) which is flat over p\mathbb{Z}_{p}.

Proof.

A proof of representability at level 1 (which implies representability for higher levels) can be found in [Beh14, 5.2]. Flatness follows from the same result for Γord\mathscr{M}^{\textup{ord}}_{\Gamma}. ∎

Remark 2.5.

Of course, for sufficiently large NN and small Γ\Gamma, we get representability from the start.

The ring V(Γ)V_{\infty}(\Gamma) is called the ring of generalized pp-adic modular forms. This carries an action of p×\mathbb{Z}_{p}^{\times} by automorphisms of 𝔾m\mathbb{G}_{m}, as well as a ring map ψp:V(Γ)V(Γ)\psi^{p}:V_{\infty}(\Gamma)\longrightarrow V_{\infty}(\Gamma) which is a lift of Frobenius. Its effect on qq-expansions is via

ψp:anqnanqpn\psi^{p}:\sum a_{n}q^{n}\mapsto\sum a_{n}q^{pn}

We extend this to a map on V(Γ):=H0(Γtriv,ω)V_{\infty}(\Gamma)_{*}:=H^{0}(\mathscr{M}^{\textup{triv}}_{\Gamma},\omega^{\otimes\ast}) by asking that ψp(v)=pv\psi^{p}(v)=pv for the periodicity generator. For any generalized pp-adic modular form gkH0(Γtriv,ωk)g_{k}\in H^{0}(\mathscr{M}^{\textup{triv}}_{\Gamma},\omega^{\otimes k}) we set the following notation:

gk(p):=(11pψp)gkg_{k}^{(p)}:=\left(1-\frac{1}{p}\psi^{p}\right)g_{k}

so that, on qq-expansions, we have

gk(p)(q)=(anpk1an/p)qn, where gk=anqng_{k}^{(p)}(q)=\sum(a_{n}-p^{k-1}a_{n/p})q^{n},\textup{ where }g_{k}=\sum a_{n}q^{n}

We will also need to know that there is a formal scheme Spf(V(Γ)ss)\textup{Spf}(V_{\infty}(\Gamma)^{\textup{ss}}) which captures the data of the ordinary part of a formal neighborhood of the supersingular locus. We will construct this for full level NN structure, Γ=Γ(N)\Gamma=\Gamma(N), following [Beh14] and in general define

V(Γ)ss:=(V(Γ(N))ss)Γ.V_{\infty}(\Gamma)^{\textup{ss}}:=\left(V_{\infty}(\Gamma(N))^{\textup{ss}}\right)^{\Gamma}.

At level NN, a formal neighborhood of the supersingular locus is given by

ellss(Γ(N))=Spf(W(ki)[[u1]])\mathscr{M}^{\textup{ss}}_{\textup{ell}}(\Gamma(N))=\coprod\textup{Spf}(W(k_{i})[\![u_{1}]\!])

for some collection of finite fields kik_{i} depending on NN. We denote the global sections by ANA_{N} and define

BN:=An[u11]p,ellss(N)ord:=Spf(BN)B_{N}:=A_{n}[u_{1}^{-1}]^{\wedge}_{p},\quad\mathscr{M}^{\textup{ss}}_{\textup{ell}}(N)^{\textup{ord}}:=\textup{Spf}(B_{N})

Then Spf(V(Γ(N))ss)\textup{Spf}(V_{\infty}(\Gamma(N))^{\textup{ss}}) is defined by the formal pullback

Spf(V(Γ(N))ss)\textstyle{\textup{Spf}(V_{\infty}(\Gamma(N))^{\textup{ss}})\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}(p)\textstyle{\mathscr{M}(p^{\infty})\ignorespaces\ignorespaces\ignorespaces\ignorespaces}ellss(N)ord\textstyle{\mathscr{M}^{\textup{ss}}_{\textup{ell}}(N)^{\textup{ord}}\ignorespaces\ignorespaces\ignorespaces\ignorespaces}ellord\textstyle{\mathscr{M}^{\textup{ord}}_{\textup{ell}}}

2.2 K(1)K(1)-local topological modular forms

The intersection of the previous section with topology comes from the following proposition.

Proposition 2.6.

There are canonical isomorphisms

LK(1)tmf(Γ)𝒪top(Γord)L_{K(1)}tmf(\Gamma)\cong\mathcal{O}^{\textup{top}}(\mathscr{M}^{\textup{ord}}_{\Gamma})
LK(1)(KpLK(1)tmf(Γ))𝒪top(Γtriv)L_{K(1)}\left(K_{p}\wedge L_{K(1)}tmf(\Gamma)\right)\cong\mathcal{O}^{\textup{top}}(\mathscr{M}^{\textup{triv}}_{\Gamma})

and the latter induces an isomorphism of θ\theta-algebras

(Kp)(LK(1)tmf(Γ))(Kp)pV(Γ)(K_{p})^{\wedge}_{*}(L_{K(1)}tmf(\Gamma))\cong(K_{p})_{*}\otimes_{\mathbb{Z}_{p}}V_{\infty}(\Gamma)

where the structure on the right hand side is diagonal and the structure on V(Γ)V_{\infty}(\Gamma) comes from the action of p×\mathbb{Z}^{\times}_{p} and a lift of Frobenius. Similarly, we have

(Kp)(LK(1)LK(2)tmf(Γ))(Kp)pV(Γ)ss(K_{p})^{\wedge}_{*}(L_{K(1)}L_{K(2)}tmf(\Gamma))\cong(K_{p})_{*}\otimes_{\mathbb{Z}_{p}}V_{\infty}(\Gamma)^{\textup{ss}}
Proof.

In fact, with our current construction of tmf(Γ)tmf(\Gamma), this is part of the definition. However, this proposition can be recovered from any reasonable construction. Indeed, the first statement is a combination of the fact that the right hand side, being a homotopy limit of K(1)K(1)-local ring spectra, is K(1)K(1)-local, together with the fact that the Adams-Novikov spectral sequence (descent spectral sequence) terminates at a finite stage with a horizontal vanishing line. The second statement can be proved using a variant of the usual ‘stacky pullback lemma’ for formal stacks, once one identifies the given moduli of elliptic curves with the stack presented by the LL-complete Hopf algebroid (π(LK(1)(KpLK(1)tmf(Γ))),π(LK(1)(KpKpLK(1)tmf(Γ))))(\pi_{*}(L_{K(1)}(K_{p}\wedge L_{K(1)}tmf(\Gamma))),\pi_{*}(L_{K(1)}(K_{p}\wedge K_{p}\wedge L_{K(1)}tmf(\Gamma)))). For references at level 1, see [Beh14, 7.9, 8.6]. ∎

We will occasionally make use of the following notation:

tmf(Γ;pn):=𝒪top(Γ(pn))tmf(\Gamma;p^{n}):=\mathcal{O}^{\textup{top}}(\mathscr{M}_{\Gamma}(p^{n}))

When the corresponding moduli stack is representable this spectrum has an action of (/pn)×(\mathbb{Z}/p^{n})^{\times} by EE_{\infty}-ring maps and we have equivalences

tmf(Γ;pn)h(/pn)×LK(1)tmf(Γ)tmf(\Gamma;p^{n})^{h(\mathbb{Z}/p^{n})^{\times}}\cong L_{K(1)}tmf(\Gamma)
((Kp)tmf(Γ;pn))(/pn)×=(Kp)LK(1)tmf(Γ)((K_{p})^{\wedge}_{*}tmf(\Gamma;p^{n}))^{(\mathbb{Z}/p^{n})^{\times}}=(K_{p})^{\wedge}_{*}L_{K(1)}tmf(\Gamma)

Finally, we record two more computations.

Proposition 2.7.

For any prime pp, we have isomorphisms of θ\theta-algebras

(Kp)Kp(Kp)pC(p×,p)(K_{p})^{\wedge}_{*}K_{p}\cong(K^{\wedge}_{p})_{*}\otimes_{\mathbb{Z}_{p}}C(\mathbb{Z}^{\times}_{p},\mathbb{Z}_{p})
(Kp)KOp(Kp)pC(p×/{±1},p)(K_{p})^{\wedge}_{*}KO_{p}\cong(K^{\wedge}_{p})_{*}\otimes_{\mathbb{Z}_{p}}C(\mathbb{Z}^{\times}_{p}/\{\pm 1\},\mathbb{Z}_{p})

Where the θ\theta-algebra structures on C(A,p)C(A,\mathbb{Z}_{p}) come from the evident actions of p×\mathbb{Z}^{\times}_{p} together with a lift of Frobenius acting on p\mathbb{Z}_{p}.

Proof.

This goes back to Strickland. A useful reference is [GHMR05] or [BH14]. Alternatively, one may compute directly the associated (affine) moduli problem. ∎

Given a K(n)K(n)-local spectrum XX satisfying some mild conditions, the completed homology (En)X(E_{n})^{\wedge}_{*}X is an LL-complete comodule over the LL-complete Hopf algebroid ((En),(En)En)((E_{n})_{*},(E_{n})^{\wedge}_{*}E_{n}). In particular, this is so if (En)X(E_{n})^{\wedge}_{*}X is finitely-generated or pro-free. There is a standard method for computing spaces of maps between K(n)K(n)-local spectra.

Theorem 2.8 (Barthel-Heard).

Suppose that (En)X(E_{n})^{\wedge}_{*}X is pro-free and (En)Y(E_{n})^{\wedge}_{*}Y is either finitely generated, pro-free, or has bounded 𝔪\mathfrak{m}-torsion where 𝔪π0En\mathfrak{m}\subset\pi_{0}E_{n} is the maximal ideal. Then the E2E_{2}-term of the K(n)K(n)-local EnE_{n}-based Adams-Novikov spectral sequence is

E2s,t=Ext^(En)Ens,t((En)X,(En)Y)E_{2}^{s,t}=\widehat{Ext}^{s,t}_{(E_{n})^{\wedge}_{*}E_{n}}((E_{n})^{\wedge}_{*}X,(E_{n})^{\wedge}_{*}Y)

This spectral sequence is strongly convergent, with abutment πtsF(X,LK(n)Y)\pi_{t-s}F(X,L_{K(n)}Y).

We will recall the definition of Ext^\widehat{\textup{Ext}} below, but first we remark that, in the case n=1n=1, we can often interpret the s=0s=0 part in a much more pedestrian fashion. If MM is an LL-complete comodule which is pro-free, then it is a Morava module. That is, the coaction map

M(Kp)Kp^(Kp)M=C(p×,p)^pMM\longrightarrow(K_{p})^{\wedge}_{*}K_{p}\widehat{\otimes}_{(K_{p})_{*}}M=C(\mathbb{Z}^{\times}_{p},\mathbb{Z}_{p})\widehat{\otimes}_{\mathbb{Z}_{p}}M

defines a continuous semi-linear action of p×\mathbb{Z}^{\times}_{p} given explicitly by

p[[p×]]^pMp[[p×]]^pC(p×,p)^pMev1M\mathbb{Z}_{p}[[\mathbb{Z}_{p}^{\times}]]\widehat{\otimes}_{\mathbb{Z}_{p}}M\longrightarrow\mathbb{Z}_{p}[[\mathbb{Z}^{\times}_{p}]]\widehat{\otimes}_{\mathbb{Z}_{p}}C(\mathbb{Z}^{\times}_{p},\mathbb{Z}_{p})\widehat{\otimes}_{\mathbb{Z}_{p}}M\stackrel{{\scriptstyle ev\otimes 1}}{{\longrightarrow}}M

In this case we have a natural isomorphism

Homp×cts(M,N)HomComod^(Kp)Kp(M,N)\textup{Hom}^{\textup{cts}}_{\mathbb{Z}^{\times}_{p}}(M,N)\cong\textup{Hom}_{\widehat{\textup{Comod}}_{(K_{p})^{\wedge}_{*}K_{p}}}(M,N)

where the left hand side denotes continuous, p×\mathbb{Z}_{p}^{\times}-equivariant homomorphisms. At this point we may state our first main calculation.

Theorem 2.9.

The following edge maps are isomorphisms:

[KOp,LK(1)tmf(Γ)]Homp×cts(C(p×/{±1},p),V(Γ))[KO_{p},L_{K(1)}tmf(\Gamma)]\longrightarrow\textup{Hom}^{\textup{cts}}_{\mathbb{Z}_{p}^{\times}}(C(\mathbb{Z}^{\times}_{p}/\{\pm 1\},\mathbb{Z}_{p}),V_{\infty}(\Gamma))
[KOp,LK(1)LK(2)tmf(Γ)]Homp×cts(C(p×/{±1},p),V(Γ)ss)[KO_{p},L_{K(1)}L_{K(2)}tmf(\Gamma)]\longrightarrow\textup{Hom}^{\textup{cts}}_{\mathbb{Z}_{p}^{\times}}(C(\mathbb{Z}^{\times}_{p}/\{\pm 1\},\mathbb{Z}_{p}),V_{\infty}(\Gamma)^{\textup{ss}})
[LK(1)tmf(Γ),LK(1)tmf(Γ)]Homp×cts(V(Γ),V(Γ))[L_{K(1)}tmf(\Gamma),L_{K(1)}tmf(\Gamma)]\longrightarrow\textup{Hom}^{\textup{cts}}_{\mathbb{Z}_{p}^{\times}}(V_{\infty}(\Gamma),V_{\infty}(\Gamma))

We show here how to deduce this from the following theorem.

Theorem 2.10.

Let GG be a profinite group, and let RR be a complete, local Noetherian graded ring with maximal homogeneous ideal generated by a sequence of regular elements. If AA is a GG-torsor and MM is pro-free, then

Ext^Gs(M,A)=0,s>0\widehat{\textup{Ext}}^{s}_{G}(M,A)=0,\quad s>0

where this is computed in the category of LL-complete C(G,R)C(G,R)-comodules.

We defer the proof and a precise definition of torsor in this context until the next section. When pp is odd, the theorem is an immediate consequence of the following

Proposition 2.11.

Let p3p\geq 3. Then we have the following vanishing results for s>0s>0

Ext^p×s((Kp)KOp,(Kp)LK(1)tmf(Γ))=0\widehat{\textup{Ext}}^{s}_{\mathbb{Z}_{p}^{\times}}((K_{p})^{\wedge}_{*}KO_{p},(K_{p})^{\wedge}_{*}L_{K(1)}tmf(\Gamma))=0
Ext^p×s((Kp)KOp,(Kp)LK(1)LK(2)tmf(Γ))=0\widehat{\textup{Ext}}^{s}_{\mathbb{Z}_{p}^{\times}}((K_{p})^{\wedge}_{*}KO_{p},(K_{p})^{\wedge}_{*}L_{K(1)}L_{K(2)}tmf(\Gamma))=0
Ext^p×s((Kp)LK(1)tmf(Γ),(Kp)LK(1)tmf(Γ))=0\widehat{\textup{Ext}}^{s}_{\mathbb{Z}_{p}^{\times}}((K_{p})^{\wedge}_{*}L_{K(1)}tmf(\Gamma),(K_{p})^{\wedge}_{*}L_{K(1)}tmf(\Gamma))=0
Proof.

Let XX denote either (Kp)LK(1)tmf(Γ)(K_{p})^{\wedge}_{*}L_{K(1)}tmf(\Gamma) or (Kp)LK(1)LK(2)tmf(Γ)(K_{p})^{\wedge}_{*}L_{K(1)}L_{K(2)}tmf(\Gamma). Then it is enough, by Theorem 2.10, to show that XX is a summand of a p×\mathbb{Z}_{p}^{\times}-torsor. To do this, it is enough to realize XX as the GG-fixed points of a p×\mathbb{Z}_{p}^{\times}-torsor where the order of GG is prime to pp. This can always be done by adding level structure (c.f. [Beh14, Prop. 7.1, Prop. 8.6]). ∎

The case p=2p=2 is more delicate, in this case the relevant spectral sequence does not collapse. We first reduce the calculation of the E2E_{2} term to one in group cohomology.

Proposition 2.12.

For s>0s>0 and p=2p=2 we have vanishing as in (2.11) if we take Ext over the subgroup 1+421+4\mathbb{Z}_{2}.

Proof.

This follows from (2.10) and [Beh14, 5.2]. ∎

If 1-1 acts nontrivially on V(Γ)V_{\infty}(\Gamma) (e.g. for Γ1(N)\Gamma_{1}(N)), then, in fact, we could have taken the entire group 2×\mathbb{Z}_{2}^{\times} in the previous proposition. So the only remaining case is when 1-1 acts trivially (e.g. for Γ0(N)\Gamma_{0}(N)). We will compute the homotopy type of the mapping space in question by working our way up from KOKO (at the cusp), to KO((q))KO((q)) (a punctured neighborhood of the cusp), and finally to LK(1)tmf(Γ)L_{K(1)}tmf(\Gamma) or LK(1)LK(2)tmf(Γ)L_{K(1)}L_{K(2)}tmf(\Gamma).

In the next few propositions, we will denote by Er(X,Y)E_{r}(X,Y) (resp. Er(X)E_{r}(X)) the K(1)K(1)-local Adams-Novikov spectral sequence computing πF(X,Y)\pi_{*}F(X,Y) (resp. πX\pi_{*}X).

Proposition 2.13.

We have an isomorphism of spectral sequences

Er(KO2,KO2)=Er(KO2)^2Hom2×cts(C(2×/{±1},2),C(2×/{±1},2))E_{r}(KO_{2},KO_{2})=E_{r}(KO_{2})\widehat{\otimes}_{\mathbb{Z}_{2}}\textup{Hom}^{\textup{cts}}_{\mathbb{Z}_{2}^{\times}}(C(\mathbb{Z}_{2}^{\times}/\{\pm 1\},\mathbb{Z}_{2}),C(\mathbb{Z}_{2}^{\times}/\{\pm 1\},\mathbb{Z}_{2}))
Proof.

There is an action of KO2KO_{2} on F(KO2,KO2)F(KO_{2},KO_{2}), where the module structure map is adjoint to the composite

KO2KO2F(KO2,KO2)KO2F(KO2,KO2)evKO2KO_{2}\wedge KO_{2}\wedge F(KO_{2},KO_{2})\longrightarrow KO_{2}\wedge F(KO_{2},KO_{2})\stackrel{{\scriptstyle ev}}{{\longrightarrow}}KO_{2}

which induces a pairing on spectral sequences. The isomorphism claimed in the theorem certainly holds when r=2r=2, so we get an isomorphism of spectral sequences if we can show that the elements in bidegree (0,0)(0,0) are permanent cycles. To see this, recall from, e.g., [AHR10, 9.2], we have

[KO2,KO2]Homcts(C(2×/{±1},2),2)[KO_{2},KO_{2}]\cong\textup{Hom}^{\textup{cts}}(C(\mathbb{Z}_{2}^{\times}/\{\pm 1\},\mathbb{Z}_{2}),\mathbb{Z}_{2})

But we have a natural isomorphism

Hom2×cts(C(2×/{±1},2),C(p×/{±1},2))Homcts(C(2×/{±1},2),2)\textup{Hom}^{\textup{cts}}_{\mathbb{Z}_{2}^{\times}}(C(\mathbb{Z}_{2}^{\times}/\{\pm 1\},\mathbb{Z}_{2}),C(\mathbb{Z}_{p}^{\times}/\{\pm 1\},\mathbb{Z}_{2}))\cong\textup{Hom}^{\textup{cts}}(C(\mathbb{Z}_{2}^{\times}/\{\pm 1\},\mathbb{Z}_{2}),\mathbb{Z}_{2})

Indeed, given an equivariant homomorphism ϕ\phi from the left hand side, we note that its behavior is determined by knowledge of ϕ(x2k)\phi(x^{2k}) for k0k\geq 0. Since λx2k=λ2kx2k\lambda\ast x^{2k}=\lambda^{2k}x^{2k} for λ2×\lambda\in\mathbb{Z}_{2}^{\times}, ϕ(x2k)\phi(x^{2k}) must have the same property. The only such function is a scalar multiple of x2kx^{2k}, and so determined by an element in 2\mathbb{Z}_{2}. This completes the proof. ∎

Corollary 2.14.

The preceding result holds for Tate KK theories, KO2[[q]]KO_{2}[[q]] and KO2((q))KO_{2}(\!(q)\!). That is, we have isomorphisms of spectral sequences

Er(KO2,KO2[[q]])=Er(KO2)^2Hom2×cts(C(2×/{±1},2),C(2×/{±1},2[[q]]))E_{r}(KO_{2},KO_{2}[[q]])=E_{r}(KO_{2})\widehat{\otimes}_{\mathbb{Z}_{2}}\textup{Hom}^{\textup{cts}}_{\mathbb{Z}_{2}^{\times}}(C(\mathbb{Z}_{2}^{\times}/\{\pm 1\},\mathbb{Z}_{2}),C(\mathbb{Z}_{2}^{\times}/\{\pm 1\},\mathbb{Z}_{2}[[q]]))
Er(KO2,KO2((q)))=Er(KO2)^2Hom2×cts(C(2×/{±1},2),C(2×/{±1},2((q))))E_{r}(KO_{2},KO_{2}(\!(q)\!))=E_{r}(KO_{2})\widehat{\otimes}_{\mathbb{Z}_{2}}\textup{Hom}^{\textup{cts}}_{\mathbb{Z}_{2}^{\times}}(C(\mathbb{Z}_{2}^{\times}/\{\pm 1\},\mathbb{Z}_{2}),C(\mathbb{Z}_{2}^{\times}/\{\pm 1\},\mathbb{Z}_{2}(\!(q)\!)))
Proof.

There is no lim1\lim^{1} issue in the first case because of surjectivity of the maps in the inverse system. The second case follows from localization of the first. ∎

Proposition 2.15.

Suppose 1-1 acts trivially on V(Γ)V_{\infty}(\Gamma). Let XX be either LK(1)tmf(Γ)L_{K(1)}tmf(\Gamma) or LK(1)LK(2)tmf(Γ)L_{K(1)}L_{K(2)}tmf(\Gamma). Let MM denote the p\mathbb{Z}_{p}-module

Hom2×cts((K2)KO2,(K2)X)\textup{Hom}^{\textup{cts}}_{\mathbb{Z}_{2}^{\times}}((K_{2})^{\wedge}_{*}KO_{2},(K_{2})^{\wedge}_{*}X)

Then we have an isomorphism of spectral sequences

Er(KO2,X)Er(KO2)^2ME_{r}(KO_{2},X)\cong E_{r}(KO_{2})\widehat{\otimes}_{\mathbb{Z}_{2}}M

In particular,

π0F(KO2,X)=M,π1F(KO2,X)=M/2\pi_{0}F(KO_{2},X)=M,\quad\pi_{1}F(KO_{2},X)=M/2
Proof.

First notice that the isomorphism holds when r=2r=2. One way to see this is to compute the Ext group via the Cartan-Eilenberg spectral sequence

Hctss((/4)×,Ext^1+42m((K2)KO2,(K2)X))Ext^2×s+m((K2)KO2,(K2)X)H_{\textup{cts}}^{s}((\mathbb{Z}/4)^{\times},\widehat{\textup{Ext}}_{1+4\mathbb{Z}_{2}}^{m}((K_{2})^{\wedge}_{*}KO_{2},(K_{2})^{\wedge}_{*}X))\Rightarrow\widehat{\textup{Ext}}_{\mathbb{Z}_{2}^{\times}}^{s+m}((K_{2})^{\wedge}_{*}KO_{2},(K_{2})^{\wedge}_{*}X)

Since (K2)X(K_{2})^{\wedge}_{*}X is a torsor for 1+421+4\mathbb{Z}_{2}, this spectral sequence collapses to

Hctss((/4)×,(K2)^2M)H_{\textup{cts}}^{s}((\mathbb{Z}/4)^{\times},(K_{2})_{*}\widehat{\otimes}_{\mathbb{Z}_{2}}M)

and the action of (/4)×(\mathbb{Z}/4)^{\times} is trivial on MM so we get the desired isomorphism for r=2r=2.

For either choice of XX we have a restriction map

XKO2((q1/N))X\longrightarrow KO_{2}(\!(q^{1/N})\!)

which induces an injection

Hom2×cts(C(2×/{±1},2),(K2)X)Hom2×cts(C(2×/{±1},2),C(2×/{±1},2((q1/N))))\textup{Hom}^{\textup{cts}}_{\mathbb{Z}_{2}^{\times}}(C(\mathbb{Z}_{2}^{\times}/\{\pm 1\},\mathbb{Z}_{2}),(K_{2})^{\wedge}_{*}X)\longrightarrow\textup{Hom}^{\textup{cts}}_{\mathbb{Z}_{2}^{\times}}(C(\mathbb{Z}_{2}^{\times}/\{\pm 1\},\mathbb{Z}_{2}),C(\mathbb{Z}_{2}^{\times}/\{\pm 1\},\mathbb{Z}_{2}(\!(q^{1/N})\!)))

This in turn gives an injection

E2(KO2,X)E2(KO2,KO2((q1/N)))E_{2}(KO_{2},X)\longrightarrow E_{2}(KO_{2},KO_{2}(\!(q^{1/N})\!))

The first nontrivial differential is a d3d_{3}, which must be zero on bidegree (0,0)(0,0) since this is true of the right hand side. This leaves nothing in the column ts=1t-s=-1, so that, in fact, the elements of bidegree (0,0)(0,0) are permanent cycles. The proposition follows. ∎

For the final function spectrum, we will need the following result, which is folklore at level 1.

Theorem 2.16.

Suppose 1-1 acts nontrivially on V(Γ)V_{\infty}(\Gamma). Then, when p=2p=2, LK(1)tmf(Γ)L_{K(1)}tmf(\Gamma) splits as a (completed) wedge of copies of KO2KO_{2} (with no suspensions.)

Proof.

Arguing as in the previous proposition, we have an injection

E2(LK(1)tmf(Γ))E2(KO2((q1/N)))E_{2}(L_{K(1)}tmf(\Gamma))\longrightarrow E_{2}(KO_{2}(\!(q^{1/N})\!))

The E2(KO2)E_{2}(KO_{2})-module structure on the right-hand side preserves the image, and we get

Er(LK(1)tmf(Γ))Er(KO2)^2V(Γ)2×E_{r}(L_{K(1)}tmf(\Gamma))\cong E_{r}(KO_{2})\widehat{\otimes}_{\mathbb{Z}_{2}}V_{\infty}(\Gamma)^{\mathbb{Z}_{2}^{\times}}

in particular, on homotopy we have

πLK(1)tmf(Γ)πKO2^2V(Γ)2×\pi_{*}L_{K(1)}tmf(\Gamma)\cong\pi_{*}KO_{2}\widehat{\otimes}_{\mathbb{Z}_{2}}V_{\infty}(\Gamma)^{\mathbb{Z}_{2}^{\times}}

so we just need a map from a wedge of KO2KO_{2}’s inducing this isomorphism. Define an equivariant homomorphism

ev1:C(2×/{±1},2)V(Γ)ev_{1}:C(\mathbb{Z}_{2}^{\times}/\{\pm 1\},\mathbb{Z}_{2})\longrightarrow V_{\infty}(\Gamma)

by ff(1)f\mapsto f(1). Then consider the composite

C(2×/{±1},2)^2V(Γ)2×ev11V(Γ)^2V(Γ)2×multV(Γ)C(\mathbb{Z}_{2}^{\times}/\{\pm 1\},\mathbb{Z}_{2})\widehat{\otimes}_{\mathbb{Z}_{2}}V_{\infty}(\Gamma)^{\mathbb{Z}_{2}^{\times}}\stackrel{{\scriptstyle ev_{1}\otimes 1}}{{\longrightarrow}}V_{\infty}(\Gamma)\widehat{\otimes}_{\mathbb{Z}_{2}}V_{\infty}(\Gamma)^{\mathbb{Z}_{2}^{\times}}\stackrel{{\scriptstyle mult}}{{\longrightarrow}}V_{\infty}(\Gamma)

By the part of Theorem 2.9 proved so far, we get a map

LK(1)(KO2(V(Γ)2×)+)LK(1)tmf(Γ)L_{K(1)}\left(KO_{2}\wedge(V_{\infty}(\Gamma)^{\mathbb{Z}_{2}^{\times}})_{+}\right)\longrightarrow L_{K(1)}tmf(\Gamma)

inducing the desired isomorphism on homotopy groups. ∎

Remark 2.17.

Despite the splitting, LK(1)tmf(Γ)L_{K(1)}tmf(\Gamma) is not a KO2KO^{\wedge}_{2}-algebra.

Corollary 2.18.

The edge map

[LK(1)tmf(Γ),LK(1)tmf(Γ)]Homp×cts(V(Γ),V(Γ))[L_{K(1)}tmf(\Gamma),L_{K(1)}tmf(\Gamma)]\longrightarrow\textup{Hom}^{\textup{cts}}_{\mathbb{Z}_{p}^{\times}}(V_{\infty}(\Gamma),V_{\infty}(\Gamma))

is an isomorphism.

Proof.

We have already seen the result for KO2KO_{2} as the source, so the result is immediate from the previous theorem by naturality of edge homomorphisms. ∎

2.3 Cohomology of torsors

We begin by briefly reviewing the definition of a torsor. Let GG be a profinite group and RR a complete, local Noetherian graded ring with a maximal homogeneous ideal generated by a sequence of regular elements.

Definition 2.19.

Let AA be a pro-free Morava module. Then AA is a GG-torsor if AA is a faithfully flat extension of AGA^{G} and the natural map

A^AGAC(G,A)A\widehat{\otimes}_{A^{G}}A\longrightarrow C(G,A)

given by (a,a)(ϕ(g)=ag(a))(a,a^{\prime})\mapsto(\phi(g)=ag(a^{\prime})) is an isomorphism of rings.

First we reduce Theorem 2.10 to a statement about group cohomology, as opposed to Ext groups.

Proposition 2.20.

Let MM and NN be pro-free Morava modules. Then we have a natural equivalence

Ext^C(G,R)(R,HomRcts(M,N))Ext^C(G,R)(M,N)\widehat{\textup{Ext}}^{*}_{C(G,R)}(R,\textup{Hom}^{\textup{cts}}_{R}(M,N))\cong\widehat{\textup{Ext}}^{*}_{C(G,R)}(M,N)
Proof.

The cochain complex computing the right hand side has terms

HomRcts(M,C(Gn,N))\textup{Hom}^{\textup{cts}}_{R}(M,C(G^{n},N))

Since MM is pro-free, the natural map to C(Gn,HomRcts(M,N))C(G^{n},\textup{Hom}^{\textup{cts}}_{R}(M,N)) is an isomorphism, and this computes the left hand side. ∎

The theorem is now a direct consequence of the following proposition applied to HomRcts(M,A)\textup{Hom}^{\textup{cts}}_{R}(M,A).

Proposition 2.21.

Let AA be a GG-torsor and MM a complete AA-module with compatible continuous GG-action. Then

Ext^C(G,R)s(R,M)=0,s>0\widehat{\textup{Ext}}^{s}_{C(G,R)}(R,M)=0,\quad s>0
Proof.

We recall the definition of this Ext group. Let

Ωn(N):=ΓRn+1RM\Omega^{n}(N):=\Gamma^{\boxtimes_{R}n+1}\boxtimes_{R}M

be the usual cobar resolution, where \boxtimes is the monoidal structure on LL-complete modules given by L0L_{0}\circ\otimes as in [BH14]. Then we have

Ext^Γs(R,M):=Hs(HomΓcts(R,Ω(M)))\widehat{\textup{Ext}}^{s}_{\Gamma}(R,M):=H^{s}\left(Hom^{\textup{cts}}_{\Gamma}(R,\Omega^{*}(M))\right)

But, by the coinduction adjunction, we have

HomΓcts(R,Ωn(N))=ΓRnRMHom^{\textup{cts}}_{\Gamma}(R,\Omega^{n}(N))=\Gamma^{\boxtimes_{R}n}\boxtimes_{R}M

as RR-modules. We want to show this complex is acyclic, and it is enough to do this after faithfully flat base change, i.e. we need only show that the complex with terms

ΓRnRMAGA\Gamma^{\boxtimes_{R}n}\boxtimes_{R}M\boxtimes_{A^{G}}A

is acyclic. But, by the torsor assumption, this complex is isomorphic to the Amitsur complex

MAGAMAG(AAGA)M\boxtimes_{A^{G}}A\longrightarrow M\boxtimes_{A^{G}}(A\boxtimes_{A^{G}}A)\longrightarrow\cdots

which is acyclic by faithfully flat descent. ∎

2.4 Reinterpretation via pp-adic measures

Now we’d like a better description of the group

Homp×cts(C(p×/{±1},p),V(Γ)).\textup{Hom}^{\textup{cts}}_{\mathbb{Z}_{p}^{\times}}(C(\mathbb{Z}^{\times}_{p}/\{\pm 1\},\mathbb{Z}_{p}),V_{\infty}(\Gamma)).

We first treat the non-equivariant case, following [Kat77]. Let XX be a compact, totally disconnected topological space and RR a pp-adically complete ring. We have

C(X,R)=C(X,p)^pRC(X,R)=C(X,\mathbb{Z}_{p})\widehat{\otimes}_{\mathbb{Z}_{p}}R

and a measure μ\mu on XX with values in RR is a (necessarily continuous) RR-linear map from C(X,R)C(X,R) to RR. That is

Meas(X,R):=HomR(C(X,R),R)=Homp(C(X,p),R)\textup{Meas}(X,R):=\textup{Hom}_{R}(C(X,R),R)=\textup{Hom}_{\mathbb{Z}_{p}}(C(X,\mathbb{Z}_{p}),R)

If UXU\subset X is compact and open then measures on UU are in bijection with measures on XX supported on UU. In the case where X=pX=\mathbb{Z}_{p} we have a very nice description of measures. The space C(p,p)C(\mathbb{Z}_{p},\mathbb{Z}_{p}) is pro-free with a very explicit basis.

Theorem 2.22 (Mahler).

For each integer k0k\geq 0 denote by (xk)\binom{x}{k} the function pp\mathbb{Z}_{p}\longrightarrow\mathbb{Z}_{p} given by

x{1k=0x(x1)(x(n1))k!k>0x\mapsto\begin{cases}1&k=0\\ \frac{x(x-1)\cdots(x-(n-1))}{k!}&k>0\end{cases}

Then C(p,p)C(\mathbb{Z}_{p},\mathbb{Z}_{p}) is pro-free with basis given by these functions.

Proof.

Let f:ppf:\mathbb{Z}_{p}\longrightarrow\mathbb{Z}_{p} be arbitrary. Then we can recover ff uniquely as

f(x)=k0(Δkf)(0)(xk)f(x)=\sum_{k\geq 0}(\Delta^{k}f)(0)\binom{x}{k}

where Δ(g)(x)=g(x+1)g(x)\Delta(g)(x)=g(x+1)-g(x) as long as this right hand side makes sense.

We need to know that the numbers (Δkf)(0)(\Delta^{k}f)(0) converge to zero pp-adically, and this is not obvious. Many proofs may be found in the literature: [Mah81] has four different proofs, and [Elk10] has a particularly short one.∎

Corollary 2.23.

Let RR be a pp-complete p\mathbb{Q}_{p}-algebra, then evaluation on the functions xkx^{k} for kmk\geq m gives a bijection

Meas(p,R)k0R\textup{Meas}(\mathbb{Z}_{p},R)\cong\prod_{k\geq 0}R
Proof.

The functions xkx^{k} for k0k\geq 0 form a basis of C(p,R)C(\mathbb{Z}_{p},R) in this case since k!(xk)=xk+O(xk1)k!\binom{x}{k}=x^{k}+O(x^{k-1}) and k!k! is a unit. ∎

Definition 2.24.

Let RR be a pp-adically complete, flat p\mathbb{Z}_{p}-algebra and XX a compact, totally disconnected subspace of p\mathbb{Z}_{p}. We say that a sequence {bk}R[1/p]\{b_{k}\}\in R[1/p] satisfies the generalized Kummer congruences (for XX) if, for every polynomial anznp[z]C(X,p)\sum a_{n}z^{n}\in\mathbb{Q}_{p}[z]\cap C(X,\mathbb{Z}_{p}) we have

anbnRR[1/p]\sum a_{n}b_{n}\in R\subset R[1/p]
Proposition 2.25.

Let RR be a pp-adically complete, flat p\mathbb{Z}_{p}-algebra. Then evaluation on xkx^{k} gives an injection

Meas(p,R)k0R[1/p]\textup{Meas}(\mathbb{Z}_{p},R)\hookrightarrow\prod_{k\geq 0}R[1/p]

onto the set of sequences {bk}\{b_{k}\} satisfying the generalized Kummer congruences.

Proof.

Injectivity follows from the previous corollary and flatness. To characterize the image, note that we really only need to check the polynomials (xk)\binom{x}{k}, and this is a necessary and sufficient condition by Mahler’s theorem. ∎

Corollary 2.26.

Let RR be as above and fix m0m\geq 0. Evaluation on x2kx^{2k} for 2km2k\geq m gives an injection

Meas(p×/{±1},R)kmR[1/p]\textup{Meas}(\mathbb{Z}^{\times}_{p}/\{\pm 1\},R)\hookrightarrow\prod_{k\geq m}R[1/p]

onto the set of sequences {bk}\{b_{k}\} such that b2k1=0b_{2k-1}=0 for all kk and satisfying the generalized Kummer congruences.

Proof.

This is immediate except for the added statement about only needing sufficiently high powers x2kx^{2k}. To see this, note that x(p1)pr1x^{(p-1)p^{r}}\rightarrow 1 uniformly as functions on p×/{±1}\mathbb{Z}^{\times}_{p}/\{\pm 1\}. Indeed, any pp-adic unit cc satisfies cp1±1 mod pc^{p-1}\equiv\pm 1\textup{ mod }p (where the minus only occurs when p=2p=2). It follows that c(p1)pr(±1)pr mod pr+1c^{(p-1)p^{r}}\equiv(\pm 1)^{p^{r}}\textup{ mod }p^{r+1} and taking limits completes the proof. ∎

We can now give a purely algebraic description of the group [KOp,LK(1)tmf(Γ)][KO_{p},L_{K(1)}tmf(\Gamma)].

Theorem 2.27.

Evaluation at the generator in π4kKOp\pi_{4k}KO_{p}, followed by division by 22 when kk is odd, gives an injection

[KOp,LK(1)tmf(Γ)]kmMFp,k(Γ,p)[KO_{p},L_{K(1)}tmf(\Gamma)]\longrightarrow\prod_{k\geq m}\textup{MF}_{p,k}(\Gamma,\mathbb{Q}_{p})

onto the set of sequences {gk}km\{g_{k}\}_{k\geq m} of pp-adic modular forms such that

  1. 1.

    For kk odd, gk=0g_{k}=0.

  2. 2.

    The sequence satisfies the generalized Kummer congruences.

Moreover, given f:KOpLK(1)tmf(Γ)f:KO_{p}\longrightarrow L_{K(1)}tmf(\Gamma), the effect on π0\pi_{0} is given by

π0f(1)=limrg(p1)pr\pi_{0}f(1)=\lim_{r}g_{(p-1)p^{r}}
Proof.

Recall that we have a natural inclusion

MFp,k(Γ,p)V(Γ,p)MF_{p,k}(\Gamma,\mathbb{Q}_{p})\hookrightarrow V_{\infty}(\Gamma,\mathbb{Q}_{p})

given by pulling back to a section of ωk\omega^{\otimes k} and then multiplying by a fixed unit in ω1\omega^{-1} coming from a (p1)(p-1)st root of the Hasse invariant. It suffices to show that the following diagram commutes:

[KOp,LK(1)tmf(Γ)]\textstyle{[KO_{p},L_{K(1)}tmf(\Gamma)]\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}12evv2k\scriptstyle{\frac{1}{2}ev_{v^{2k}}}\scriptstyle{\cong}kmMFp,k(Γ,p)\textstyle{\prod_{k\geq m}MF_{p,k}(\Gamma,\mathbb{Q}_{p})\ignorespaces\ignorespaces\ignorespaces\ignorespaces}Measp×(p×/{±1},V(Γ))\textstyle{\textup{Meas}_{\mathbb{Z}_{p}^{\times}}(\mathbb{Z}_{p}^{\times}/\{\pm 1\},V_{\infty}(\Gamma))\ignorespaces\ignorespaces\ignorespaces\ignorespaces}evx2k\scriptstyle{ev_{x^{2k}}}kmV(Γ,p)\textstyle{\prod_{k\geq m}V_{\infty}(\Gamma,\mathbb{Q}_{p})}

Here v2kπ4kKOpv^{2k}\in\pi_{4k}KO_{p} is the image of the similarly named element in πbu\pi_{*}bu under the forgetful map, and the subscript on measures indicates we restrict to p×\mathbb{Z}^{\times}_{p}-equivariant homomorphisms C(p×/{±1},p)V(Γ)C(\mathbb{Z}^{\times}_{p}/\{\pm 1\},\mathbb{Z}_{p})\longrightarrow V_{\infty}(\Gamma).

Assuming the commutativity of the diagram, we see that evaluation at v2kv^{2k} gives an injection into the prescribed set of sequences. On the other hand, given such a sequence, we automatically get a measure by the preceding corollary and we need only check it is equivariant. But we may check equivariance on the dense subspace of C(p×/{±1},p)C(\mathbb{Z}^{\times}_{p}/\{\pm 1\},\mathbb{Z}_{p}) consisting of the even polynomials. In this case equivariance is equivalent to the requirement that x2kx^{2k} maps to an element gg of V(Γ)V_{\infty}(\Gamma) which satisfies

λg=λ2kg,λp×\lambda\ast g=\lambda^{2k}g,\quad\lambda\in\mathbb{Z}^{\times}_{p}

Since pp-adic modular forms of weight 2k2k satisfy this requirement, we conclude that the measure is equivariant.

It remains to check the commutativity of the diagram. This follows from the commutativity of the diagram

[Kp,LK(1)tmf(Γ)]\textstyle{[K_{p},L_{K(1)}tmf(\Gamma)]\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}evv2k\scriptstyle{ev_{v^{2k}}}\scriptstyle{\cong}kmMFp,k(Γ,p)\textstyle{\prod_{k\geq m}MF_{p,k}(\Gamma,\mathbb{Q}_{p})\ignorespaces\ignorespaces\ignorespaces\ignorespaces}Measp×(p×,V(Γ))\textstyle{\textup{Meas}_{\mathbb{Z}_{p}^{\times}}(\mathbb{Z}_{p}^{\times},V_{\infty}(\Gamma))\ignorespaces\ignorespaces\ignorespaces\ignorespaces}evx2k\scriptstyle{ev_{x^{2k}}}kmV(Γ,p)\textstyle{\prod_{k\geq m}V_{\infty}(\Gamma,\mathbb{Q}_{p})}

by naturality and the observation that the map KOKKO\longrightarrow K sends v2kπKOpv^{2k}\in\pi_{*}KO_{p}, as we’ve been denoting it, to 2v2kπKp2v^{2k}\in\pi_{*}K_{p}. The commutativity of this simpler diagram follows from the same proof as in [AHR10, 9.5]. ∎

2.5 K(1)K(1)-local orientations

Before completing the proof of the main theorem of this section, we need to review the formula for Rezk’s logarithm in the K(1)K(1)-local case. Let RR be a K(1)K(1)-local EE_{\infty} ring spectrum and denote by ψp\psi^{p} the power operation on πR\pi_{*}R lifting the Frobenius.

Theorem 2.28 (Rezk).

There is a K(1)K(1)-local equivalence

1:gl1RR\ell_{1}:gl_{1}R\longrightarrow R

whose effect on homotopy groups in positive degrees is given by

1+x(11pψp)x1+x\mapsto\left(1-\frac{1}{p}\psi^{p}\right)x
Corollary 2.29.

The effect of 1\ell_{1} on π2kLK(1)gl1tmf(Γ)ppMp,k(Γ,p)\pi_{2k}L_{K(1)}gl_{1}tmf(\Gamma)\otimes_{\mathbb{Z}_{p}}\mathbb{Q}_{p}\cong M_{p,k}(\Gamma,\mathbb{Q}_{p}) for k1k\geq 1 is given by

gg(p)g\mapsto g^{(p)}

We will also need the following result

Theorem 2.30 (Ando-Hopkins-Rezk).

If cp×/{±1}c\in\mathbb{Z}_{p}^{\times}/\{\pm 1\} is a topological generator, we have a commutative diagram

LK(1)S\textstyle{L_{K(1)}S\ignorespaces\ignorespaces\ignorespaces\ignorespaces}ρ(c)\scriptstyle{\rho(c)}KOp\textstyle{KO_{p}\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}1ψc\scriptstyle{1-\psi^{c}}\scriptstyle{\cong}bc\scriptstyle{b_{c}}KOp\textstyle{KO_{p}\ignorespaces\ignorespaces\ignorespaces\ignorespaces}LK(1)gl1S\textstyle{L_{K(1)}gl_{1}S\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}1\scriptstyle{\ell_{1}}\scriptstyle{\cong}LK(1)gl1S/spin\textstyle{L_{K(1)}gl_{1}S/spin\ignorespaces\ignorespaces\ignorespaces\ignorespaces}KOp\textstyle{KO_{p}}

where ρ(c)1=12plog(cp1)\rho(c)^{-1}=\frac{1}{2p}\log(c^{p-1}).

Proof.

First we construct the map bcb_{c}. Recall that the (proven) Adams Conjecture states that the composite

BOp\textstyle{BO^{\wedge}_{p}\ignorespaces\ignorespaces\ignorespaces\ignorespaces}1ψc\scriptstyle{1-\psi^{c}}BOp\textstyle{BO^{\wedge}_{p}\ignorespaces\ignorespaces\ignorespaces\ignorespaces}J\scriptstyle{J}(BGL1S)p\textstyle{(BGL_{1}S)^{\wedge}_{p}}

is nullhomotopic. This remains true upon taking connective covers, so that we have a nullhomotopy for the composite

BSpinp\textstyle{BSpin^{\wedge}_{p}\ignorespaces\ignorespaces\ignorespaces\ignorespaces}1ψc\scriptstyle{1-\psi^{c}}BSpinp\textstyle{BSpin^{\wedge}_{p}\ignorespaces\ignorespaces\ignorespaces\ignorespaces}J\scriptstyle{J}(BGL1S)p\textstyle{(BGL_{1}S)^{\wedge}_{p}}

This produces a homotopy commutative diagram

fib(1ψc)\textstyle{\textup{fib}(1-\psi^{c})\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}Ac\scriptstyle{A_{c}}BSpinp\textstyle{BSpin^{\wedge}_{p}\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}Bc\scriptstyle{B_{c}}BSpinp\textstyle{BSpin^{\wedge}_{p}\ignorespaces\ignorespaces\ignorespaces\ignorespaces}(GL1S)p\textstyle{(GL_{1}S)^{\wedge}_{p}\ignorespaces\ignorespaces\ignorespaces\ignorespaces}(GL1S/Spin)p\textstyle{(GL_{1}S/Spin)^{\wedge}_{p}\ignorespaces\ignorespaces\ignorespaces\ignorespaces}BSpinp\textstyle{BSpin^{\wedge}_{p}\ignorespaces\ignorespaces\ignorespaces\ignorespaces}J\scriptstyle{J}(BGL1S)p\textstyle{(BGL_{1}S)^{\wedge}_{p}}

On the other hand, for any nn we have a factorization

SpectraLK(n)\scriptstyle{L_{K(n)}}Ω\scriptstyle{\Omega^{\infty}}SpectraK(n)\textstyle{\textup{Spectra}_{K(n)}}SpacesΦn\scriptstyle{\Phi_{n}}

where Φn\Phi_{n} is the Bousfield-Kuhn functor. So, using Φ1\Phi_{1}, we get a homotopy commutative diagram

LK(1)fib(1ψc)\textstyle{L_{K(1)}\textup{fib}(1-\psi^{c})\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}ac\scriptstyle{a_{c}}LK(1)bspin\textstyle{L_{K(1)}bspin\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}1ψc\scriptstyle{1-\psi^{c}}bc\scriptstyle{b_{c}}LK(1)bspin\textstyle{L_{K(1)}bspin\ignorespaces\ignorespaces\ignorespaces\ignorespaces}LK(1)gl1S\textstyle{L_{K(1)}gl_{1}S\ignorespaces\ignorespaces\ignorespaces\ignorespaces}LK(1)gl1S/spin\textstyle{L_{K(1)}gl_{1}S/spin\ignorespaces\ignorespaces\ignorespaces\ignorespaces}LK(1)bspin\textstyle{L_{K(1)}bspin}

We have a canonical equivalence

LK(1)bspinKOpL_{K(1)}bspin\cong KO_{p}

so our diagram becomes

jc\textstyle{j_{c}\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}ac\scriptstyle{a_{c}}KOp\textstyle{KO_{p}\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}1ψc\scriptstyle{1-\psi^{c}}bc\scriptstyle{b_{c}}LK(1)bspin\textstyle{L_{K(1)}bspin\ignorespaces\ignorespaces\ignorespaces\ignorespaces}LK(1)gl1S\textstyle{L_{K(1)}gl_{1}S\ignorespaces\ignorespaces\ignorespaces\ignorespaces}LK(1)gl1S/spin\textstyle{L_{K(1)}gl_{1}S/spin\ignorespaces\ignorespaces\ignorespaces\ignorespaces}KOp\textstyle{KO_{p}}

Since cc is a generator, the computation of the K(1)K(1)-local sphere shows that the map aca_{c} is an equivalence, whence so is bcb_{c}. Now define the map ρ(c):LK(1)SKOp\rho(c):L_{K(1)}S\longrightarrow KO_{p} to make the following diagram commute

LK(1)S\textstyle{L_{K(1)}S\ignorespaces\ignorespaces\ignorespaces\ignorespaces}ρ(c)\scriptstyle{\rho(c)}KOp\textstyle{KO_{p}\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}1ψc\scriptstyle{1-\psi^{c}}\scriptstyle{\cong}bc\scriptstyle{b_{c}}KOp\textstyle{KO_{p}\ignorespaces\ignorespaces\ignorespaces\ignorespaces}LK(1)gl1S\textstyle{L_{K(1)}gl_{1}S\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}1\scriptstyle{\ell_{1}}\scriptstyle{\cong}LK(1)gl1S/spin\textstyle{L_{K(1)}gl_{1}S/spin\ignorespaces\ignorespaces\ignorespaces\ignorespaces}KOp\textstyle{KO_{p}}

The description of ρ(c)\rho(c) as in the theorem statement can be found as [AHR10, 7.15]. ∎

Proof of Theorem 2.1.

This is exactly as in [AHR10, 14.6]. For convenience, we recall the argument here. We are interested in maps α\alpha making the following diagram commute up to homotopy

gl1S\textstyle{gl_{1}S\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}gl1S/spin\textstyle{gl_{1}S/spin\ignorespaces\ignorespaces\ignorespaces\ignorespaces}α\scriptstyle{\alpha}LK(1)gl1tmf(Γ)p\textstyle{L_{K(1)}gl_{1}tmf(\Gamma)^{\wedge}_{p}}

(We are permitted to consider just homotopy classes here by the argument in [AHR10, 14.3]). Since the target is K(1)K(1)-local, we may use the previous theorem to replace this diagram with the following one

LK(1)S\textstyle{L_{K(1)}S\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}ρ(c)\scriptstyle{\rho(c)}KOp\textstyle{KO_{p}\ignorespaces\ignorespaces\ignorespaces\ignorespaces}1αbc\scriptstyle{\ell_{1}\alpha b_{c}}LK(1)tmf(Γ)p\textstyle{L_{K(1)}tmf(\Gamma)^{\wedge}_{p}}

By Theorem 2.9 a map KOpLK(1)tmf(Γ)pKO_{p}\longrightarrow L_{K(1)}tmf(\Gamma)^{\wedge}_{p} is determined by its rationalization so we must understand the composite

KOp\textstyle{KO_{p}\ignorespaces\ignorespaces\ignorespaces\ignorespaces}1αbc\scriptstyle{\ell_{1}\alpha b_{c}}LK(1)tmf(Γ)p\textstyle{L_{K(1)}tmf(\Gamma)^{\wedge}_{p}\ignorespaces\ignorespaces\ignorespaces\ignorespaces}LK(1)tmf(Γ)p\textstyle{L_{K(1)}tmf(\Gamma)^{\wedge}_{p}\otimes\mathbb{Q}}

Since gl1Sgl_{1}S\otimes\mathbb{Q} is contractible, Theorem 2.30 implies that we have a factorization

KOp\textstyle{KO_{p}\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}1αbc\scriptstyle{\ell_{1}\alpha b_{c}}1ψc\scriptstyle{1-\psi^{c}}KOp\textstyle{KO_{p}\ignorespaces\ignorespaces\ignorespaces\ignorespaces}1β\scriptstyle{\ell_{1}\beta}LK(1)tmf(Γ)p\textstyle{L_{K(1)}tmf(\Gamma)^{\wedge}_{p}\ignorespaces\ignorespaces\ignorespaces\ignorespaces}LK(1)tmf(Γ)p\textstyle{L_{K(1)}tmf(\Gamma)^{\wedge}_{p}\otimes\mathbb{Q}}

for some β:KOpLK(1)gl1tmf(Γ)p\beta:KO_{p}\longrightarrow L_{K(1)}gl_{1}tmf(\Gamma)^{\wedge}_{p}\otimes\mathbb{Q}. Thus, if tk(α)t_{k}(\alpha) are the moments of the measure determined by 1αbc\ell_{1}\alpha b_{c}, and bk(α):=βνkMFp,k(Γ)b_{k}(\alpha):=\beta_{*}\nu^{k}\in MF_{p,k}(\Gamma)\otimes\mathbb{Q}, then the diagram and the formula (2.29) implies

tk(α)=(1ck)bk(p)t_{k}(\alpha)=(1-c^{k})b_{k}^{(p)}

Thus, the existence of such a commutative square is determined by a sequence of pp-adic modular forms {gk}k4\{g_{k}\}_{k\geq 4} such that the sequence {(1ck)g(p)}\{(1-c^{k})g^{(p)}\} extends to a measure μc\mu_{c} on p×/{±1}\mathbb{Z}_{p}^{\times}/\{\pm 1\} valued in V(Γ)V_{\infty}(\Gamma). This square, in turn, makes the diagram

LK(1)S\textstyle{L_{K(1)}S\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}ρ(c)\scriptstyle{\rho(c)}KOp\textstyle{KO_{p}\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}1αbc\scriptstyle{\ell_{1}\alpha b_{c}}1ψc\scriptstyle{1-\psi^{c}}KOp\textstyle{KO_{p}\ignorespaces\ignorespaces\ignorespaces\ignorespaces}1β\scriptstyle{\ell_{1}\beta}LK(1)tmf(Γ)p\textstyle{L_{K(1)}tmf(\Gamma)^{\wedge}_{p}\ignorespaces\ignorespaces\ignorespaces\ignorespaces}LK(1)tmf(Γ)p\textstyle{L_{K(1)}tmf(\Gamma)^{\wedge}_{p}\otimes\mathbb{Q}}

commute if and only if the effect of 1αbc\ell_{1}\alpha b_{c} on π0\pi_{0} is ρ(c)1\rho(c)^{-1}. By [AHR10, 7.15] this happens if and only if p×/{±1}1𝑑μc=12plog(cp1)\int_{\mathbb{Z}_{p}^{\times}/\{\pm 1\}}1d\mu_{c}=\frac{1}{2p}\log(c^{p-1}), which completes the proof. ∎

3 Gluing and the Atkin operator

Crucial to the construction of these genera is an understanding of the bottom map in the fiber square

LK(1)K(2)gl1tmf(Γ)p\textstyle{L_{K(1)\vee K(2)}gl_{1}tmf(\Gamma)^{\wedge}_{p}\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}LK(2)gl1tmf(Γ)p\textstyle{L_{K(2)}gl_{1}tmf(\Gamma)^{\wedge}_{p}\ignorespaces\ignorespaces\ignorespaces\ignorespaces}LK(1)gl1tmf(Γ)p\textstyle{L_{K(1)}gl_{1}tmf(\Gamma)^{\wedge}_{p}\ignorespaces\ignorespaces\ignorespaces\ignorespaces}LK(1)LK(2)gl1tmf(Γ)p\textstyle{L_{K(1)}L_{K(2)}gl_{1}tmf(\Gamma)^{\wedge}_{p}}

Using a K(2)K(2)-local version of Rezk’s logarithm, and the K(1)K(1)-local version we’ve already met, we can replace this square with the following one

LK(1)K(2)gl1tmf(Γ)p\textstyle{L_{K(1)\vee K(2)}gl_{1}tmf(\Gamma)^{\wedge}_{p}\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}LK(2)tmf(Γ)p\textstyle{L_{K(2)}tmf(\Gamma)^{\wedge}_{p}\ignorespaces\ignorespaces\ignorespaces\ignorespaces}LK(1)tmf(Γ)p\textstyle{L_{K(1)}tmf(\Gamma)^{\wedge}_{p}\ignorespaces\ignorespaces\ignorespaces\ignorespaces}LK(1)LK(2)tmf(Γ)p\textstyle{L_{K(1)}L_{K(2)}tmf(\Gamma)^{\wedge}_{p}}

By naturality, the right hand vertical map is the usual one, but the bottom horizontal map is not. One of the key insights of Ando-Hopkins-Rezk is a computation of this map in terms of the Atkin operator on pp-adic modular forms. In §3.1 we review the necessary algebraic facts about this operator and its relationship to Hecke operators. In §3.2 we construct a topological lift of the Atkin operator on K(1)K(1)-local topological modular forms and show that it fits in a fiber square as above. Finally, in §3.3 we compute the connected components of Nulls(spin,LK(1)K(2)gl1tmf(Γ)p)\textup{Nulls}(spin,L_{K(1)\vee K(2)}gl_{1}tmf(\Gamma)^{\wedge}_{p}), which is the penultimate step in the program to understanding pp-complete orientations.

3.1 Hecke operators on modular forms

Fix Γ{Γ1(N),Γ0(N)}\Gamma\in\{\Gamma_{1}(N),\Gamma_{0}(N)\}, and a prime pp not dividing NN. A (Γ1(N);p)(\Gamma_{1}(N);p) level structure on a generalized elliptic curve E/SE_{/S} is a pair (P,C)(P,C) where PP is a Γ1(N)\Gamma_{1}(N)-level structure and CC is a finite locally free SS-subgroup scheme in EsmE^{\textup{sm}} that is cyclic of order pp subject to the condition that the effective Cartier divisor

j/N(jP+C)\sum_{j\in\mathbb{Z}/N\mathbb{Z}}(jP+C)

meets all irreducible components of all geometric fibers. A (Γ0(N);p)(\Gamma_{0}(N);p) level structure is a pair (G,C)(G,C) consisting of a cyclic subgroup GEsmG\subset E^{\textup{sm}} of order NN such that, fppffppf locally where GG admits a generator PP, the pair (P,C)(P,C) is a (Γ1(N);p)(\Gamma_{1}(N);p) level structure. These define stacks (Γ1(N);p)\mathscr{M}_{(\Gamma_{1}(N);p)} and (Γ0(N);p)\mathscr{M}_{(\Gamma_{0}(N);p)}. On the smooth loci, there are two canonical maps π10,π20:(Γ;p)0(Γ;p)0\pi_{1}^{0},\pi_{2}^{0}:\mathscr{M}^{0}_{(\Gamma;p)}\longrightarrow\mathscr{M}^{0}_{(\Gamma;p)} given for (Γ1(N);p)(\Gamma_{1}(N);p) by

π10(E;P,C)=(E,P),π20(E;P,C)=(E/C,Pmod C)\pi_{1}^{0}(E;P,C)=(E,P),\quad\pi_{2}^{0}(E;P,C)=(E/C,P\textup{mod }C)

and defined via flat descent for (Γ0(N);p)(\Gamma_{0}(N);p).

Theorem 3.1 (Conrad).

The correspondence (π10,π20)(\pi_{1}^{0},\pi_{2}^{0}) uniquely extends to a finite, flat correspondence, the Hecke correspondence,

(Γ;p)\textstyle{\mathscr{M}_{(\Gamma;p)}\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}π2\scriptstyle{\pi_{2}}π1\scriptstyle{\pi_{1}}Γ\textstyle{\mathscr{M}_{\Gamma}}Γ\textstyle{\mathscr{M}_{\Gamma}}

Likewise, the natural map on the smooth locus

(π20)ωΓω(Γ;p)(\pi_{2}^{0})^{*}\omega_{\Gamma}\longrightarrow\omega_{(\Gamma;p)}

uniquely extends to a map on all of (Γ;p)\mathscr{M}_{(\Gamma;p)}.

Completing at pp and restricting to the ordinary locus we get a correspondence

(Γ;p)ord\textstyle{\mathscr{M}^{\textup{ord}}_{(\Gamma;p)}\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}π1\scriptstyle{\pi_{1}}π2\scriptstyle{\pi_{2}}Γord\textstyle{\mathscr{M}^{\textup{ord}}_{\Gamma}}Γord\textstyle{\mathscr{M}^{\textup{ord}}_{\Gamma}}

which induces a map pTp:Mp,k(Γ)Mp,k(Γ)pT_{p}:M_{p,k}(\Gamma)\longrightarrow M_{p,k}(\Gamma) defined as the composite

H0(Γord,ωΓk)\textstyle{H^{0}(\mathscr{M}^{\textup{ord}}_{\Gamma},\omega_{\Gamma}^{\otimes k})\ignorespaces\ignorespaces\ignorespaces\ignorespaces}π2\scriptstyle{\pi_{2}^{*}}H0((Γ;p)ord,π2ωΓk)\textstyle{H^{0}(\mathscr{M}^{\textup{ord}}_{(\Gamma;p)},\pi_{2}^{*}\omega_{\Gamma}^{\otimes k})\ignorespaces\ignorespaces\ignorespaces\ignorespaces}H0((Γ;p)ord,ωk)=H0(Γord,(π1)π1ωk)\textstyle{H^{0}(\mathscr{M}^{\textup{ord}}_{(\Gamma;p)},\omega^{\otimes k})=H^{0}(\mathscr{M}^{\textup{ord}}_{\Gamma},(\pi_{1})_{*}\pi_{1}^{*}\omega^{\otimes k})\ignorespaces\ignorespaces\ignorespaces\ignorespaces}H0(Γord,ωk)\textstyle{H^{0}(\mathscr{M}^{\textup{ord}}_{\Gamma},\omega^{\otimes k})}

The image of this map lies in pMp,kpM_{p,k}, which is torsion free, so we have a well-defined operator

Tp:Mp,k(Γ)Mp,k(Γ)T_{p}:M_{p,k}(\Gamma)\longrightarrow M_{p,k}(\Gamma)

called the ppth Hecke operator. As it happens, (Γ;p)ord\mathscr{M}^{\textup{ord}}_{(\Gamma;p)} is the disjoint union of two copies of Γord\mathscr{M}^{\textup{ord}}_{\Gamma}. The correspondence then decomposes as a sum of two correspondences corresponding to the following finite, flat maps:

Frob:(E;P,C)(E(p),ϕ(P),ϕ(C))\textup{Frob}:(E;P,C)\mapsto(E^{(p)},\phi(P),\phi(C))
pV:(E(p);pϕ(P),ϕ(C))(E;P,C)\langle p\rangle V:(E^{(p)};p\phi(P),\phi(C))\mapsto(E;P,C)

Define the Atkin operator UpU_{p} as 1ptr(Frob)\frac{1}{p}\textup{tr}(\textup{Frob}) on cohomology, and similarly define ψp\psi^{p} as the trace of the other map divided by pp. We get

pTp=tr(Frob)+ψppT_{p}=\textup{tr}(\textup{Frob})+\psi^{p}

on Mp,k(Γ)M_{p,k}(\Gamma), which yields a pp-adic lift of the Eichler-Shimura relation:

Tp=Up+1pψpT_{p}=U_{p}+\frac{1}{p}\psi^{p}

By pulling back along the map from Γtriv\mathscr{M}^{\textup{triv}}_{\Gamma}, all of these operators are defined on V(Γ)V_{\infty}(\Gamma) as well, and the same pp-adic lift of the Eichler-Shimura relation holds.

The effect on qq-expansions of these operators is as follows for fMk(Γ0(N))f\in M_{k}(\Gamma_{0}(N)) with f(q)=anqnf(q)=\sum a_{n}q^{n}:

Upf(q)=anpqn,ψpf(q)=pkanqnpU_{p}f(q)=\sum a_{np}q^{n},\quad\psi^{p}f(q)=p^{k}\sum a_{n}q^{np}

and for level Γ1(N)\Gamma_{1}(N) there are similar formulae that depend on the nebentypus of the modular form. From these, or directly from the definition, one verifies that Upψp=pkU_{p}\psi^{p}=p^{k} for arbitrary level.

Remark 3.2.

The reader may be confused that ψp\psi^{p} seems to be acting simultaneously like the Frobenius and a multiple of the Verschiebung. The mystery is solved by noting that there are two distinct notions of a level pp structure on an elliptic curve. One is a morphism μpE[p]\mu_{p}\hookrightarrow E[p] (an Igusa level structure) and one is a morphism /pE[p]\mathbb{Z}/p\mathbb{Z}\longrightarrow E[p] (a classical level pp structure). These are ‘transposes’ of each other in the same way that the Frobenius and the Verschiebung are transposes of each other. See [Gro90, 3.12] and the remarks and references therein.

Remark 3.3.

In [AHR10], for level 1 modular forms, the authors use the notation VpV_{p} for what corresponds to our 1pkψp\frac{1}{p^{k}}\psi^{p} in weight kk.

3.2 Lifting the Atkin operator

In addition to Rezk’s logarithm for K(1)K(1)-local EE_{\infty} rings, he also constructed a K(n)K(n)-local version specifically for forms of Morava EE-theory. For our purposes, we require the following statement:

Theorem 3.4 (Rezk).

Let kk be a perfect field and GG a height nn formal group, denote by E(k,G)E(k,G) the associated Morava EE-theory. Then there is a K(n)K(n)-local equivalence

n:gl1E(k,G)E(k,G)\ell_{n}:gl_{1}E(k,G)\longrightarrow E(k,G)

Moreover, this is natural in the pair (k,G)(k,G).

At height 22, in the case E(k,C^)E(k,\widehat{C}) for CC a supersingular elliptic curve over kk, the effect on π2k\pi_{2k} is:

1+f(1Tp+pk1)f1+f\mapsto(1-T_{p}+p^{k-1})f

We would like to apply this map to the K(2)K(2)-localization of tmf(Γ)tmf(\Gamma). The relationship comes from the following

Proposition 3.5.

Let SSp(Γ)\textup{SS}_{p}(\Gamma) denote the set of suspersingular elliptic curves at the prime pp with level Γ\Gamma-structure. Then we have an equivalence

LK(2)tmf(Γ)(C,α)SSpE(kC,C^)hAut(C,α)L_{K(2)}tmf(\Gamma)\cong\prod_{(C,\alpha)\in\textup{SS}_{p}}E(k_{C},\widehat{C})^{h\textup{Aut}(C,\alpha)}
Proof.

On an étale map from an affine, this is the definition of LK(2)𝒪topL_{K(2)}\mathcal{O}^{\textup{top}} (there are no automorphisms left). So choose an affine cover, and then descent gives the desired homotopy fixed points. ∎

We can tie these two together to get a logarithm for topological modular forms.

Proposition 3.6.

There is a K(2)K(2)-local equivalence

gl1tmf(Γ)pLK(2)tmf(Γ)pgl_{1}tmf(\Gamma)^{\wedge}_{p}\longrightarrow L_{K(2)}tmf(\Gamma)^{\wedge}_{p}

such that the following diagram commutes for k1k\geq 1,

π2kgl1tmf(Γ)p\textstyle{\pi_{2k}gl_{1}tmf(\Gamma)^{\wedge}_{p}\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}π2kLK(2)tmf(Γ)p\textstyle{\pi_{2k}L_{K(2)}tmf(\Gamma)^{\wedge}_{p}\ignorespaces\ignorespaces\ignorespaces\ignorespaces}MFk(Γ)p\textstyle{MF_{k}(\Gamma)^{\wedge}_{p}\ignorespaces\ignorespaces\ignorespaces\ignorespaces}1Tp+pk1\scriptstyle{1-T_{p}+p^{k-1}}MFp,k(Γ)\textstyle{MF_{p,k}(\Gamma)}
Proof.

Recall that, by [Bou87, Kuh89] , we have

LK(n)gl1RLK(n)gl1LK(n)RL_{K(n)}gl_{1}R\cong L_{K(n)}gl_{1}L_{K(n)}R

so we may replace tmf(Γ)ptmf(\Gamma)^{\wedge}_{p} by its K(2)K(2)-localization in the source. By Goerss-Hopkins, there is an action of 𝔾n\mathbb{G}_{n} on each E(k,G)E(k,G) in the homotopy theory of EE_{\infty}-rings. Applying gl1gl_{1} and the natural transformation 2:gl1LK(2)\ell_{2}:gl_{1}\longrightarrow L_{K(2)} we get a map of diagrams. Now take the homotopy limits indicated in Proposition 3.5 and use that gl1gl_{1} commutes with homotopy limits, as it is a right adjoint. ∎

Using this equivalence we get the following diagram:

LK(1)K(2)gl1tmf(Γ)\textstyle{L_{K(1)\vee K(2)}gl_{1}tmf(\Gamma)\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}LK(2)gl1tmf(Γ)\textstyle{L_{K(2)}gl_{1}tmf(\Gamma)\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}2\scriptstyle{\ell_{2}}\scriptstyle{\cong}LK(2)tmf(Γ)\textstyle{L_{K(2)}tmf(\Gamma)\ignorespaces\ignorespaces\ignorespaces\ignorespaces}LK(1)gl1tmf(Γ)\textstyle{L_{K(1)}gl_{1}tmf(\Gamma)\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}1\scriptstyle{\ell_{1}}\scriptstyle{\cong}LK(1)LK(2)gl1tmf(Γ)\textstyle{L_{K(1)}L_{K(2)}gl_{1}tmf(\Gamma)\ignorespaces\ignorespaces\ignorespaces\ignorespaces}LK(1)2\scriptstyle{L_{K(1)}\ell_{2}}\scriptstyle{\cong}LK(1)LK(2)tmf(Γ)\textstyle{L_{K(1)}L_{K(2)}tmf(\Gamma)}LK(1)tmf(Γ)\textstyle{L_{K(1)}tmf(\Gamma)\ignorespaces\ignorespaces\ignorespaces\ignorespaces}

The remainder of the section will be devoted to describing this dotted arrow as in the statement below.

Theorem 3.7.

There is a map Up:LK(1)tmf(Γ)LK(1)tmf(Γ)U_{p}:L_{K(1)}tmf(\Gamma)\longrightarrow L_{K(1)}tmf(\Gamma) making the following two diagrams commute

π2kLK(1)tmf(Γ)p\textstyle{\pi_{2k}L_{K(1)}tmf(\Gamma)^{\wedge}_{p}\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}Up\scriptstyle{U_{p}}π2kLK(1)tmf(Γ)p\textstyle{\pi_{2k}L_{K(1)}tmf(\Gamma)^{\wedge}_{p}\ignorespaces\ignorespaces\ignorespaces\ignorespaces}MFp,k(Γ)\textstyle{MF_{p,k}(\Gamma)\ignorespaces\ignorespaces\ignorespaces\ignorespaces}Up\scriptstyle{U_{p}}MFp,k(Γ)\textstyle{MF_{p,k}(\Gamma)}
LK(1)K(2)gl1tmf(Γ)p\textstyle{L_{K(1)\vee K(2)}gl_{1}tmf(\Gamma)^{\wedge}_{p}\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}LK(2)tmf(Γ)p\textstyle{L_{K(2)}tmf(\Gamma)^{\wedge}_{p}\ignorespaces\ignorespaces\ignorespaces\ignorespaces}LK(1)tmf(Γ)p\textstyle{L_{K(1)}tmf(\Gamma)^{\wedge}_{p}\ignorespaces\ignorespaces\ignorespaces\ignorespaces}1Up\scriptstyle{1-U_{p}}LK(1)tmf(Γ)p\textstyle{L_{K(1)}tmf(\Gamma)^{\wedge}_{p}\ignorespaces\ignorespaces\ignorespaces\ignorespaces}LK(1)LK(2)tmf(Γ)p\textstyle{L_{K(1)}L_{K(2)}tmf(\Gamma)^{\wedge}_{p}}
Remark 3.8.

This theorem is claimed in [AHR10]. See also [Bak89].

Before turning to the proof, we note an interesting and as-yet unexplained corollary, which is the existence of a logarithm for pp-complete tmftmf (though it is no longer an equivalence).

Corollary 3.9.

There is a fiber square

LK(1)K(2)gl1tmf(Γ)p\textstyle{L_{K(1)\vee K(2)}gl_{1}tmf(\Gamma)^{\wedge}_{p}\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}tmf\scriptstyle{\ell_{tmf}}Tmf(Γ)p\textstyle{Tmf(\Gamma)^{\wedge}_{p}\ignorespaces\ignorespaces\ignorespaces\ignorespaces}LK(1)tmf(Γ)p\textstyle{L_{K(1)}tmf(\Gamma)^{\wedge}_{p}\ignorespaces\ignorespaces\ignorespaces\ignorespaces}1Up\scriptstyle{1-U_{p}}LK(1)tmf(Γ)p\textstyle{L_{K(1)}tmf(\Gamma)^{\wedge}_{p}}
Proof.

This follows from pasting homotopy pullback squares together with the fact that LK(1)K(2)tmf(Γ)p=Tmf(Γ)pL_{K(1)\vee K(2)}tmf(\Gamma)^{\wedge}_{p}=Tmf(\Gamma)^{\wedge}_{p}. ∎

First we construct the map.

Lemma 3.10.

There is a unique map

Up:LK(1)tmf(Γ)LK(1)tmf(Γ)U_{p}:L_{K(1)}tmf(\Gamma)\longrightarrow L_{K(1)}tmf(\Gamma)

whose effect upon applying (Kp)(K_{p})^{\wedge}_{*} is the Atkin operator for generalized pp-adic modular forms.

Proof.

This is immediate from Theorem 2.9 and the fact that UpU_{p} is p×\mathbb{Z}_{p}^{\times} equivariant. ∎

Lemma 3.11.

We have a commutative diagram

π2kLK(1)tmf(Γ)p\textstyle{\pi_{2k}L_{K(1)}tmf(\Gamma)^{\wedge}_{p}\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}Up\scriptstyle{U_{p}}π2kLK(1)tmf(Γ)p\textstyle{\pi_{2k}L_{K(1)}tmf(\Gamma)^{\wedge}_{p}\ignorespaces\ignorespaces\ignorespaces\ignorespaces}MFp,k(Γ)\textstyle{MF_{p,k}(\Gamma)\ignorespaces\ignorespaces\ignorespaces\ignorespaces}Up\scriptstyle{U_{p}}MFp,k(Γ)\textstyle{MF_{p,k}(\Gamma)}
Proof.

This follows from naturality of the edge homomorphism in the Adams-Novikov spectral sequence ∎

This proves half of Theorem 3.7. For the other half, we first need a corollary of Theorem 2.9.

Corollary 3.12.

Evaluation on π\pi_{*} gives an injection

[LK(1)tmf(Γ)p,LK(1)LK(2)tmf(Γ)p]Hom(πLK(1)tmf(Γ)p,πLK(1)LK(2)tmf(Γ)p)[L_{K(1)}tmf(\Gamma)^{\wedge}_{p},L_{K(1)}L_{K(2)}tmf(\Gamma)^{\wedge}_{p}]\longrightarrow\textup{Hom}(\pi_{*}L_{K(1)}tmf(\Gamma)^{\wedge}_{p}\otimes\mathbb{Q},\pi_{*}L_{K(1)}L_{K(2)}tmf(\Gamma)^{\wedge}_{p}\otimes\mathbb{Q})
Proof.

By (2.9), the left hand side is torsion-free. ∎

Proposition 3.13.

The diagram

LK(1)K(2)gl1tmf(Γ)p\textstyle{L_{K(1)\vee K(2)}gl_{1}tmf(\Gamma)^{\wedge}_{p}\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}LK(2)tmf(Γ)p\textstyle{L_{K(2)}tmf(\Gamma)^{\wedge}_{p}\ignorespaces\ignorespaces\ignorespaces\ignorespaces}LK(1)tmf(Γ)p\textstyle{L_{K(1)}tmf(\Gamma)^{\wedge}_{p}\ignorespaces\ignorespaces\ignorespaces\ignorespaces}1Up\scriptstyle{1-U_{p}}LK(1)tmf(Γ)p\textstyle{L_{K(1)}tmf(\Gamma)^{\wedge}_{p}\ignorespaces\ignorespaces\ignorespaces\ignorespaces}LK(1)LK(2)tmf(Γ)p\textstyle{L_{K(1)}L_{K(2)}tmf(\Gamma)^{\wedge}_{p}}

commutes.

Proof.

It suffices to check commutativity of the following diagram

LK(1)gl1tmf(Γ)\textstyle{L_{K(1)}gl_{1}tmf(\Gamma)\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}1\scriptstyle{\ell_{1}}LK(1)LK(2)gl1tmf(Γ)\textstyle{L_{K(1)}L_{K(2)}gl_{1}tmf(\Gamma)\ignorespaces\ignorespaces\ignorespaces\ignorespaces}LK(1)2\scriptstyle{L_{K(1)}\ell_{2}}LK(1)tmf(Γ)\textstyle{L_{K(1)}tmf(\Gamma)\ignorespaces\ignorespaces\ignorespaces\ignorespaces}1Up\scriptstyle{1-U_{p}}LK(1)LK(2)tmf(Γ)\textstyle{L_{K(1)}L_{K(2)}tmf(\Gamma)}

By Corollary 3.12 we need only check that the diagram

πLK(1)gl1tmf(Γ)\textstyle{\pi_{*}L_{K(1)}gl_{1}tmf(\Gamma)\otimes\mathbb{Q}\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}1\scriptstyle{\ell_{1}}LK(1)LK(2)gl1tmf(Γ)\textstyle{L_{K(1)}L_{K(2)}gl_{1}tmf(\Gamma)\otimes\mathbb{Q}\ignorespaces\ignorespaces\ignorespaces\ignorespaces}LK(1)2\scriptstyle{L_{K(1)}\ell_{2}}πLK(1)tmf(Γ)\textstyle{\pi_{*}L_{K(1)}tmf(\Gamma)\otimes\mathbb{Q}\ignorespaces\ignorespaces\ignorespaces\ignorespaces}1Up\scriptstyle{1-U_{p}}πLK(1)LK(2)tmf(Γ)\textstyle{\pi_{*}L_{K(1)}L_{K(2)}tmf(\Gamma)\otimes\mathbb{Q}}

but now this is a statement about modular forms, which follows from Rezk’s formulae for the logarithms and the pp-adic lift of the Eichler-Shimura relation. ∎

3.3 Spin orientations in the unscrewable case

We are now ready to solve the extension problem

spin\textstyle{spin\ignorespaces\ignorespaces\ignorespaces\ignorespaces}gl1S\textstyle{gl_{1}S\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}gl1S/spin\textstyle{gl_{1}S/spin\ignorespaces\ignorespaces\ignorespaces\ignorespaces}LK(1)K(2)gl1tmf(Γ)p\textstyle{L_{K(1)\vee K(2)}gl_{1}tmf(\Gamma)^{\wedge}_{p}}
Theorem 3.14.

If p2p\neq 2, the map

π0Nulls(spin,LK(1)K(2)gl1tmf(Γ)p)π0Nulls(spin,LK(1)gl1tmf(Γ)p)\pi_{0}\textup{Nulls}(spin,L_{K(1)\vee K(2)}gl_{1}tmf(\Gamma)^{\wedge}_{p})\longrightarrow\pi_{0}\textup{Nulls}(spin,L_{K(1)}gl_{1}tmf(\Gamma)^{\wedge}_{p})

is an isomorphism onto the set of sequences {gk}k2\{g_{k}\}_{k\geq 2} as in Theorem 2.1 satisfying the additional condition that Upgk(p)=gk(p)U_{p}g_{k}^{(p)}=g_{k}^{(p)}. If p=2p=2, the map

Char(Spin,LK(1)K(2)gl1tmf(Γ))Char(Spin,LK(1)tmf(Γ))\textup{Char}(Spin,L_{K(1)\vee K(2)}gl_{1}tmf(\Gamma))\longrightarrow\textup{Char}(Spin,L_{K(1)}tmf(\Gamma))

(cf. Definition 1.4) is an isomorphism onto the same set of sequences.

Proof.

We have a homotopy pullback square

Nulls(spin,LK(1)K(2)gl1tmf(Γ)p)\textstyle{\textup{Nulls}(spin,L_{K(1)\vee K(2)}gl_{1}tmf(\Gamma)^{\wedge}_{p})\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}Nulls(spin,LK(2)tmf(Γ)p)\textstyle{\textup{Nulls}(spin,L_{K(2)}tmf(\Gamma)^{\wedge}_{p})\ignorespaces\ignorespaces\ignorespaces\ignorespaces}Nulls(spin,LK(1)tmf(Γ)p)\textstyle{\textup{Nulls}(spin,L_{K(1)}tmf(\Gamma)^{\wedge}_{p})\ignorespaces\ignorespaces\ignorespaces\ignorespaces}Nulls(spin,LK(1)LK(2)tmf(Γ)p)\textstyle{\textup{Nulls}(spin,L_{K(1)}L_{K(2)}tmf(\Gamma)^{\wedge}_{p})}

The upper right hand square is contractible, so the image of the map

π0Nulls(spin,LK(1)K(2)gl1tmf(Γ)p)π0Nulls(spin,LK(1)tmf(Γ)p)\pi_{0}\textup{Nulls}(spin,L_{K(1)\vee K(2)}gl_{1}tmf(\Gamma)^{\wedge}_{p})\longrightarrow\pi_{0}\textup{Nulls}(spin,L_{K(1)}tmf(\Gamma)^{\wedge}_{p})

consists of those sequences {gk}k2\{g_{k}\}_{k\geq 2} as in Theorem 2.1 such that (1Up)gk=0(1-U_{p})g_{k}=0 (c.f. [AHR10, 13.7]). The fiber of this map over any point is a torsor for

[ΣKOp,LK(1)LK(2)tmf(Γ)p][\Sigma KO_{p},L_{K(1)}L_{K(2)}tmf(\Gamma)^{\wedge}_{p}]

This group vanishes for p2p\neq 2 proving the claim in this case. When p=2p=2 this group is all torsion, so the discrepancy vanishes on characteristic series. ∎

4 Parameterizing pp-complete Orientations

Now we’d like to solve the extension problem

g\textstyle{g\ignorespaces\ignorespaces\ignorespaces\ignorespaces}gl1S\textstyle{gl_{1}S\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}gl1S/g\textstyle{gl_{1}S/g\ignorespaces\ignorespaces\ignorespaces\ignorespaces}gl1tmf(Γ)p\textstyle{gl_{1}tmf(\Gamma)^{\wedge}_{p}}

where gg is either spinspin or stringstring. We have already solved this problem for LK(1)K(2)gl1tmf(Γ)L_{K(1)\vee K(2)}gl_{1}tmf(\Gamma), so the composite

ggl1Sgl1tmf(Γ)LK(1)K(2)gl1tmf(Γ)g\longrightarrow gl_{1}S\longrightarrow gl_{1}tmf(\Gamma)\longrightarrow L_{K(1)\vee K(2)}gl_{1}tmf(\Gamma)

is nullhomotopic. Let FF be defined by the fiber sequence

Fgl1tmf(Γ)LK(1)K(2)gl1tmf(Γ)F\longrightarrow gl_{1}tmf(\Gamma)\longrightarrow L_{K(1)\vee K(2)}gl_{1}tmf(\Gamma)

then there exists a dotted arrow making the following diagram commute:

g\textstyle{g\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}gl1S\textstyle{gl_{1}S\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}gl1S/g\textstyle{gl_{1}S/g}F\textstyle{F\ignorespaces\ignorespaces\ignorespaces\ignorespaces}gl1tmf(Γ)p\textstyle{gl_{1}tmf(\Gamma)^{\wedge}_{p}\ignorespaces\ignorespaces\ignorespaces\ignorespaces}LK(1)K(2)gl1tmf(Γ)\textstyle{L_{K(1)\vee K(2)}gl_{1}tmf(\Gamma)}

and the obstruction to the existence of an orientation of gg is determined by the map

[g,F][g,gl1tmf(Γ)p][g,F]\longrightarrow[g,gl_{1}tmf(\Gamma)^{\wedge}_{p}]

When g=stringg=string as in [AHR10] the group [string,F][string,F] vanishes so there is no obstruction. However, when g=sping=spin this group never vanishes (4.10) so we must understand this map. We begin in §4.1 by collecting some necessary preliminaries on Hida’s theory of ordinary modular forms. In §4.2 we review what the work of Ando-Hopkins-Rezk can tell us about the homotopy type of FF in general. In §4.3 we compute π3F\pi_{3}F at level 1 in terms of modular forms and show how this implies the existence of spinspin orientations of tmftmf away from 6. In §4.4 we specialize to level pp topological modular forms and parameterize spinspin orientations of tmf0(p)tmf_{0}(p)^{\wedge}_{\ell} when 2\ell\neq 2. Finally, in §4.5 we treat the case when =2\ell=2 at various prime levels.

4.1 Units and localization

Given an LnL_{n}-local EE_{\infty} ring spectrum RR, the spectrum of units gl1Rgl_{1}R is almost never LnL_{n}-local. The following theorem of Ando-Hopkins-Rezk measures the difference.

Theorem 4.1 (Ando-Hopkins-Rezk, Theorem 4.11).

Let RR be an LnL_{n}-local EE_{\infty} ring spectrum and let dn(R)d_{n}(R) denote the fiber

dn(R)gl1RLngl1Rd_{n}(R)\longrightarrow gl_{1}R\longrightarrow L_{n}gl_{1}R

Then πqdn(R)=0\pi_{q}d_{n}(R)=0 for qn+1q\geq n+1.

We would like to apply this to the spectrum FF, which is the fiber of LK(1)K(2)L_{K(1)\vee K(2)}-localization. The relationship between the two is given by the following two lemmas.

Lemma 4.2.

The map L2XLK(1)K(2)XL_{2}X\longrightarrow L_{K(1)\vee K(2)}X is pp-completion.

Proof.

We have fiber squares

L2\textstyle{L_{2}\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}LK(1)K(2)\textstyle{L_{K(1)\vee K(2)}\ignorespaces\ignorespaces\ignorespaces\ignorespaces}LK(0)\textstyle{L_{K(0)}\ignorespaces\ignorespaces\ignorespaces\ignorespaces}LK(0)LK(1)K(2)\textstyle{L_{K(0)}L_{K(1)\vee K(2)}}LK(1)K(2)\textstyle{L_{K(1)\vee K(2)}\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}LK(2)\textstyle{L_{K(2)}\ignorespaces\ignorespaces\ignorespaces\ignorespaces}LK(1)\textstyle{L_{K(1)}\ignorespaces\ignorespaces\ignorespaces\ignorespaces}LK(1)LK(2)\textstyle{L_{K(1)}L_{K(2)}}

Note that K(n)S/p=K(n)K(n)\wedge S/p=K(n), and thus all K(n)K(n)-local spectra are pp-complete. Since pp-complete spectra are closed under homotopy limits, the right hand square implies that LK(1)K(2)L_{K(1)\vee K(2)} is pp-complete. So it suffices to show that the top horizontal arrow of the left hand square is an equivalence, or equivalently that the bottom horizontal map is an equivalence. But this is clear because LS/pLK(0)=0L_{S/p}L_{K(0)}=0.∎

Lemma 4.3.

Let RR be a pp-complete and L2L_{2}-local EE_{\infty}-ring spectrum. Define FF as the fiber

Fgl1RLK(1)K(2)gl1RF\longrightarrow gl_{1}R\longrightarrow L_{K(1)\vee K(2)}gl_{1}R

Then πqF=0\pi_{q}F=0 for q4q\geq 4.

Proof.

Recall we have defined dn(R)d_{n}(R) via the fiber sequence

dn(R)gl1RLngl1Rd_{n}(R)\longrightarrow gl_{1}R\longrightarrow L_{n}gl_{1}R

and, by (4.1), we know that πqdn(R)=0\pi_{q}d_{n}(R)=0 for q3q\geq 3.

Let 0 denote a zero object, and consider the larger diagram

dn(R)\textstyle{d_{n}(R)\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}0\textstyle{0\ignorespaces\ignorespaces\ignorespaces\ignorespaces}F\textstyle{F\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}F\textstyle{F^{\prime}\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}0\textstyle{0\ignorespaces\ignorespaces\ignorespaces\ignorespaces}gl1R\textstyle{gl_{1}R\ignorespaces\ignorespaces\ignorespaces\ignorespaces}L2gl1R\textstyle{L_{2}gl_{1}R\ignorespaces\ignorespaces\ignorespaces\ignorespaces}LK(1)K(2)gl1R\textstyle{L_{K(1)\vee K(2)}gl_{1}R}

where all the squares are homotopy cartesian. Since πqdn(R)=0\pi_{q}d_{n}(R)=0 for q3q\geq 3 it suffices to show that πqF=0\pi_{q}F^{\prime}=0 for q4q\geq 4. By the previous lemma, we have that the map

L2LK(1)K(2)L_{2}\longrightarrow L_{K(1)\vee K(2)}

is pp-completion. Now, since

πqRπqgl1RπqL2gl1R\pi_{q}R\cong\pi_{q}gl_{1}R\cong\pi_{q}L_{2}gl_{1}R

for q4q\geq 4 and a spectrum is pp-complete if and only if its homotopy groups are Ext-pp-complete [Bou79, Prop. 2.5], we have that πqL2gl1R\pi_{q}L_{2}gl_{1}R is Ext-pp-complete for q4q\geq 4.

The homotopy groups of a pp-completion, by [Bou79, Prop. 2.5], are given by:

0\textstyle{0\ignorespaces\ignorespaces\ignorespaces\ignorespaces}Ext(/p,πqL2gl1R)\textstyle{\textup{Ext}(\mathbb{Z}/p^{\infty},\pi_{q}L_{2}gl_{1}R)\ignorespaces\ignorespaces\ignorespaces\ignorespaces}πqLK(1)K(2)gl1R\textstyle{\pi_{q}L_{K(1)\vee K(2)}gl_{1}R\ignorespaces\ignorespaces\ignorespaces\ignorespaces}Hom(/p,πq1L2gl1R)\textstyle{\textup{Hom}(\mathbb{Z}/p^{\infty},\pi_{q-1}L_{2}gl_{1}R)\ignorespaces\ignorespaces\ignorespaces\ignorespaces}0\textstyle{0}

By the previous remark, since Ext(/p,)\textup{Ext}(\mathbb{Z}/p^{\infty},-) is Ext-pp-completion, the first entry is just πqL2gl1R\pi_{q}L_{2}gl_{1}R for q4q\geq 4, and the last entry vanishes for q5q\geq 5. Thus the map

πqL2gl1RπqLK(1)K(2)gl1R\pi_{q}L_{2}gl_{1}R\longrightarrow\pi_{q}L_{K(1)\vee K(2)}gl_{1}R

is an equivalence for q5q\geq 5 and an inclusion for q=4q=4, whence πqF=0\pi_{q}F^{\prime}=0 for q4q\geq 4 which completes the proof. ∎

Remark 4.4.

The same proof gives analogous results comparing LnL_{n} and LK(1)K(n)L_{K(1)\vee\cdots\vee K(n)}.

4.2 Analysis of FF

In order to understand FF in the fiber sequence

F\textstyle{F\ignorespaces\ignorespaces\ignorespaces\ignorespaces}gl1tmf(Γ)\textstyle{gl_{1}tmf(\Gamma)\ignorespaces\ignorespaces\ignorespaces\ignorespaces}LK(1)K(2)gl1tmf(Γ)\textstyle{L_{K(1)\vee K(2)}gl_{1}tmf(\Gamma)}

we first need to understand something about LK(1)K(2)gl1tmf(Γ)L_{K(1)\vee K(2)}gl_{1}tmf(\Gamma). The crucial homotopy group is π4\pi_{4}. Recall that the Atkin operator was constructed in §3.1, and it’s effect on qq-expansions is

Up(anqn)=anpqnU_{p}(\sum a_{n}q^{n})=\sum a_{np}q^{n}
Proposition 4.5.

Let p3p\geq 3 be a prime not dividing the level NN. Then there is a pullback square

π4LK(1)K(2)gl1tmf(Γ)p\textstyle{\pi_{4}L_{K(1)\vee K(2)}gl_{1}tmf(\Gamma)^{\wedge}_{p}\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}MF2(Γ)p\textstyle{MF_{2}(\Gamma)^{\wedge}_{p}\ignorespaces\ignorespaces\ignorespaces\ignorespaces}MFp,2\textstyle{MF_{p,2}\ignorespaces\ignorespaces\ignorespaces\ignorespaces}1Up\scriptstyle{1-U_{p}}MFp,2\textstyle{MF_{p,2}}

and the vertical maps are injective.

Proof.

By (3.9) we have a homotopy fiber square

LK(1)K(2)gl1tmf(Γ)p\textstyle{L_{K(1)\vee K(2)}gl_{1}tmf(\Gamma)^{\wedge}_{p}\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}Tmf(Γ)p\textstyle{Tmf(\Gamma)^{\wedge}_{p}\ignorespaces\ignorespaces\ignorespaces\ignorespaces}LK(1)tmf(Γ)p\textstyle{L_{K(1)}tmf(\Gamma)^{\wedge}_{p}\ignorespaces\ignorespaces\ignorespaces\ignorespaces}1Up\scriptstyle{1-U_{p}}LK(1)tmf(Γ)p\textstyle{L_{K(1)}tmf(\Gamma)^{\wedge}_{p}}

The result follows from the observation that πoddLK(1)tmf(Γ)p=0\pi_{\textup{odd}}L_{K(1)}tmf(\Gamma)^{\wedge}_{p}=0 for p3p\geq 3. To see this, recall that we have an equivalence

LK(1)tmf(Γ)ptmf(Γ;p)h(/p)×L_{K(1)}tmf(\Gamma)^{\wedge}_{p}\longrightarrow tmf(\Gamma;p)^{h(\mathbb{Z}/p)^{\times}}

The homotopy fixed point spectral sequence collapses to the zero line, so it suffices to make the same observation about tmf(Γ;p)tmf(\Gamma;p). But Γ(p)\mathscr{M}_{\Gamma}(p) is representable by a formally affine scheme when p3p\geq 3, so the descent spectral sequence for tmf(Γ;p)tmf(\Gamma;p) collapses to the zero line and we have the result. ∎

Remark 4.6.

It is not true that πtmf(Γ)p\pi_{*}tmf(\Gamma)^{\wedge}_{p} is concentrated in even degrees at large primes, in general. Indeed, when the modular curve has large genus, Serre duality forces contributions in odd, positive degrees for the homotopy of Tmf(Γ)Tmf(\Gamma) which then persist after taking the connective cover. Indeed, this observation is one of the reasons why the current definition of tmf(Γ)tmf(\Gamma) is undesirable when the modular curve has large genus.

Proposition 4.7.

The following diagram commutes

π4gl1tmf(Γ)p\textstyle{\pi_{4}gl_{1}tmf(\Gamma)^{\wedge}_{p}\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}\scriptstyle{\cong}π4LK(1)K(2)gl1tmf(Γ)p\textstyle{\pi_{4}L_{K(1)\vee K(2)}gl_{1}tmf(\Gamma)^{\wedge}_{p}\ignorespaces\ignorespaces\ignorespaces\ignorespaces}MF2(Γ)p\textstyle{MF_{2}(\Gamma)\otimes\mathbb{Z}_{p}\ignorespaces\ignorespaces\ignorespaces\ignorespaces}1ψp\scriptstyle{1-\psi^{p}}MFp,2(Γ)\textstyle{MF_{p,2}(\Gamma)}

Putting these two together we can often compute π3F\pi_{3}F in practice. For now, we only remark that it is torsion-free.

Corollary 4.8.

Let p5p\geq 5, then the group π3F\pi_{3}F is torsion free.

Proof.

By the above proposition and a short diagram chase, π3F\pi_{3}F injects into the cokernel of the map

1ψp:MF2(Γ)pMFp,2(Γ)1-\psi^{p}:MF_{2}(\Gamma)\otimes\mathbb{Z}_{p}\longrightarrow MF_{p,2}(\Gamma)

This is torsion-free since the reduction modulo pp of this map is injective and MFp,2(Γ)MF_{p,2}(\Gamma) is torsion-free. ∎

4.3 The Discrepancy Map and Weight 2 Modular Forms

We have a homotopy pullback diagram of spectra under gl1Sgl_{1}S

gl1tmf(Γ)p\textstyle{gl_{1}tmf(\Gamma)^{\wedge}_{p}\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}LK(1)K(2)gl1tmf(Γ)\textstyle{L_{K(1)\vee K(2)}gl_{1}tmf(\Gamma)\ignorespaces\ignorespaces\ignorespaces\ignorespaces}0\textstyle{0\ignorespaces\ignorespaces\ignorespaces\ignorespaces}ΣF\textstyle{\Sigma F}

where the structure map gl1SΣFgl_{1}S\longrightarrow\Sigma F is defined as the composite gl1SLK(1)K(2)gl1tmf(Γ)ΣFgl_{1}S\rightarrow L_{K(1)\vee K(2)}gl_{1}tmf(\Gamma)\rightarrow\Sigma F. Thus, we get a homotopy pullback diagram

Nulls(spin,gl1tmf(Γ)p)\textstyle{\textup{Nulls}(spin,gl_{1}tmf(\Gamma)^{\wedge}_{p})\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}Nulls(spin,LK(1)K(2)gl1tmf(Γ))\textstyle{\textup{Nulls}(spin,L_{K(1)\vee K(2)}gl_{1}tmf(\Gamma))\ignorespaces\ignorespaces\ignorespaces\ignorespaces}B\textstyle{\ast\ignorespaces\ignorespaces\ignorespaces\ignorespaces}ANulls(spin,ΣF)\textstyle{\textup{Nulls}(spin,\Sigma F)}

and we need to understand the maps A and B. By definition, the first map hits the orientation corresponding to the composite

gl1S/spin0ΣFgl_{1}S/spin\longrightarrow 0\longrightarrow\Sigma F

using this to trivialize the torsor we get the following tautological result.

Proposition 4.9.

Using the zero map gl1S/spinΣFgl_{1}S/spin\longrightarrow\Sigma F to trivialize the torsor Nulls(spin,ΣF)\textup{Nulls}(spin,\Sigma F), the image of the map A is the zero map

bspinΣFbspin\longrightarrow\Sigma F

To detect elements in Nulls(spin,ΣF)\textup{Nulls}(spin,\Sigma F) we need only check one homotopy group.

Lemma 4.10.

Evaluation at π4\pi_{4} gives an isomorphism

[bspin,ΣF]π4ΣF[bspin,\Sigma F]\cong\pi_{4}\Sigma F
Proof.

Since bspinbspin has no homotopy below π4\pi_{4}, we have [bspin,ΣF]=[bspin,τ4ΣF][bspin,\Sigma F]=[bspin,\tau_{\geq 4}\Sigma F]. Since ΣF\Sigma F has no homotopy above π4\pi_{4} this is [bspin,Hπ4ΣF][bspin,H\pi_{4}\Sigma F] and the result follows from Hurewicz. ∎

Let’s recall how to associate a sequence of modular forms to a nullhomotopy as in (3.14). Given an element in π0Nulls(spin,LK(1)K(2)gl1tmf(Γ))\pi_{0}\textup{Nulls}(spin,L_{K(1)\vee K(2)}gl_{1}tmf(\Gamma)) we have a map

gl1S/spinLK(1)K(2)gl1tmf(Γ)pgl_{1}S/spin\longrightarrow L_{K(1)\vee K(2)}gl_{1}tmf(\Gamma)^{\wedge}_{p}

Upon K(1)K(1)-localization, we get a composite

KOp\textstyle{KO_{p}\ignorespaces\ignorespaces\ignorespaces\ignorespaces}bc\scriptstyle{b_{c}}\scriptstyle{\cong}LK(1)gl1S/spin\textstyle{L_{K(1)}gl_{1}S/spin\ignorespaces\ignorespaces\ignorespaces\ignorespaces}LK(1)gl1tmf(Γ)p\textstyle{L_{K(1)}gl_{1}tmf(\Gamma)^{\wedge}_{p}\otimes\mathbb{Q}}

and the image of generators of π4kKOp\pi_{4k}KO_{p}, suitably divided by 2 as in (1.4), give us a sequence of modular forms of the shape (1c2k)(1ψp)g2k(1-c^{2k})(1-\psi^{p})g_{2k}. The sequence {g2k}\{g_{2k}\} is the associated sequence.

Lemma 4.11.

Let c0c\in\mathbb{Z}_{\geq 0} be a topological generator of p\mathbb{Z}_{p}. Let bc:KOpLK(1)gl1S/spinb_{c}:KO_{p}\cong L_{K(1)}gl_{1}S/spin be the equivalence constructed in (2.30). Then there exists an element xπ4gl1S/spinx\in\pi_{4}gl_{1}S/spin whose image under K(1)K(1)-localization is bc(x4)b_{c}(x_{4}), where we recall that x4π4KOpx_{4}\in\pi_{4}KO_{p} is a generator.

Proof.

Since π4S=0\pi_{4}S=0, we have an exact sequence

0\textstyle{0\ignorespaces\ignorespaces\ignorespaces\ignorespaces}π4gl1S/spin\textstyle{\pi_{4}gl_{1}S/spin\ignorespaces\ignorespaces\ignorespaces\ignorespaces}π4bspin\textstyle{\pi_{4}bspin\ignorespaces\ignorespaces\ignorespaces\ignorespaces}π3gl1S\textstyle{\pi_{3}gl_{1}S}

Let x4x_{4} denote the generator of π4bspin\pi_{4}bspin. Then, since π3S=/24\pi_{3}S=\mathbb{Z}/24, 24x424x_{4} is hit by an element xπ4gl1S/spinx\in\pi_{4}gl_{1}S/spin. This has the desired properties by inspection of the diagram (2.30). ∎

Theorem 4.12.

Suppose that π3F\pi_{3}F is torsion-free (e.g. if p5p\geq 5 by (4.8)). Identifying elements in

π0Nulls(spin,LK(1)K(2)gl1tmf(Γ))\pi_{0}\textup{Nulls}(spin,L_{K(1)\vee K(2)}gl_{1}tmf(\Gamma))

with certain sequences of pp-adic modular forms as reviewed above, the map B takes an element {gk}k2\{g_{k}\}_{k\geq 2} to the image of A if and only if g2g_{2} is in the image of the usual inclusion

MF2(Γ)pMFp,2(Γ)MF_{2}(\Gamma)\otimes\mathbb{Z}_{p}\longrightarrow MF_{p,2}(\Gamma)
Proof.

Consider the following diagram:

gl1S/spin\textstyle{gl_{1}S/spin\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}bspin\textstyle{bspin\ignorespaces\ignorespaces\ignorespaces\ignorespaces}LK(1)K(2)gl1tmf(Γ)p\textstyle{L_{K(1)\vee K(2)}gl_{1}tmf(\Gamma)^{\wedge}_{p}\ignorespaces\ignorespaces\ignorespaces\ignorespaces}ΣF\textstyle{\Sigma F}

By the previous lemma, there is an element xπ4gl1S/spinx\in\pi_{4}gl_{1}S/spin whose image under K(1)K(1)-localization is bc(x4)b_{c}(x_{4}). Let ϕ\phi denote the image of x4π4bspinx_{4}\in\pi_{4}bspin under the right hand vertical map, and let yy denote the image of xx under the left hand vertical map. Then commutativity of the diagram, together with the diagram of (2.30), gives the picture

x\textstyle{x\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}(1c2)x4\textstyle{(1-c^{2})x_{4}\ignorespaces\ignorespaces\ignorespaces\ignorespaces}y\textstyle{y\ignorespaces\ignorespaces\ignorespaces\ignorespaces}(1c2)ϕ\textstyle{(1-c^{2})\phi}

When p5p\geq 5, 1c21-c^{2} is a pp-adic unit. Indeed, if 1c201-c^{2}\equiv 0 mod pp, then cc would be ±1\pm 1 mod pp and so wouldn’t generate μp1/{±1}\mu_{p-1}/\{\pm 1\} when p5p\geq 5. Thus, when p5p\geq 5 or π3F\pi_{3}F is torsion-free, ϕ=0\phi=0 if and only if (1c2)ϕ=0(1-c^{2})\phi=0, and this happens if and only if yy maps to zero in π3F\pi_{3}F. Equivalently, we’re asking that yy be in the image of

π4gl1tmf(Γ)pπ4LK(1)K(2)gl1tmf(Γ)p\pi_{4}gl_{1}tmf(\Gamma)^{\wedge}_{p}\longrightarrow\pi_{4}L_{K(1)\vee K(2)}gl_{1}tmf(\Gamma)^{\wedge}_{p}

which we may check after composition with the injective map

π4LK(1)K(2)gl1tmf(Γ)pπ4LK(1)tmf(Γ)p.\pi_{4}L_{K(1)\vee K(2)}gl_{1}tmf(\Gamma)^{\wedge}_{p}\longrightarrow\pi_{4}L_{K(1)}tmf(\Gamma)^{\wedge}_{p}\otimes\mathbb{Q}.

By definition of xx, the element yy maps to the element 2(1c2)(1ψp)g22(1-c^{2})(1-\psi^{p})g_{2}, and by (4.7) this lies in the image of the map from π4gl1tmf(Γ)p\pi_{4}gl_{1}tmf(\Gamma)^{\wedge}_{p} if and only if g2g_{2} came from an element in MF2pMF_{2}\otimes\mathbb{Z}_{p}. ∎

4.4 The case p=3p=3

When p=3p=3 trouble may occur if π3F\pi_{3}F has 33-torsion. This doesn’t happen as soon as the moduli problem is representable.

Lemma 4.13.

If Γ[1/N]\mathscr{M}_{\Gamma}[1/N] is representable by a scheme, π3F\pi_{3}F is torsion-free. In particular, this is true for Γ=Γ0(2)\Gamma=\Gamma_{0}(2).

Recall we have an exact sequence

π4gl1tmf(Γ)π4LK(1)K(2)gl1tmf(Γ)3π3Fπ3gl1tmf(Γ)3\pi_{4}gl_{1}tmf(\Gamma)\longrightarrow\pi_{4}L_{K(1)\vee K(2)}gl_{1}tmf(\Gamma)^{\wedge}_{3}\longrightarrow\pi_{3}F\longrightarrow\pi_{3}gl_{1}tmf(\Gamma)^{\wedge}_{3}

Thus, it suffices to show that the first map has torsion-free cokernel and the last group is torsion-free. The former claim has the same proof as (4.8). The latter claim is the context of the next lemma, which completes the proof of the theorem.

Lemma 4.14.

When (Γ)3\left(\mathscr{M}_{\Gamma}\right)^{\wedge}_{3} is representable by a scheme, the group π3tmf(Γ)3\pi_{3}tmf(\Gamma)^{\wedge}_{3} is torsion-free.

Proof.

The only possible torsion is in π1\pi_{1}, see [HL13, 6.4]. ∎

4.5 The case p=2p=2

At the prime 22, torsion appears in even simple examples:

Lemma 4.15.

The torsion in π3F\pi_{3}F is at most a /2\mathbb{Z}/2.

Proof.

We have an exact sequence

π4gl1tmf0(3)π4LK(1)K(2)gl1tmf0(3)π3Fπ3gl1tmf0(3)\pi_{4}gl_{1}tmf_{0}(3)\longrightarrow\pi_{4}L_{K(1)\vee K(2)}gl_{1}tmf_{0}(3)\longrightarrow\pi_{3}F\longrightarrow\pi_{3}gl_{1}tmf_{0}(3)

and the last term is /2\mathbb{Z}/2 by [MR09]. The cokernel of the first map is also torsion-free by the argument as in the previous section, thus π3F\pi_{3}F can, at most, contain a /2\mathbb{Z}/2 as torsion. ∎

Corollary 4.16.

Let bswing denote the fiber of the map

bspinw4Σ4H/2bspin\stackrel{{\scriptstyle w_{4}}}{{\longrightarrow}}\Sigma^{4}H\mathbb{Z}/2

Then, for every prime p3p\neq 3, MSwing admits a tmf0(3)ptmf_{0}(3)^{\wedge}_{p}-orientation for every sequence {gk}k2MFp,k(Γ0(3))\{g_{k}\}_{k\geq 2}\in MF_{p,k}(\Gamma_{0}(3))\otimes\mathbb{Q} of modular forms satisfying the conditions of (3.14) and such that g2MF2(Γ0(3))g_{2}\in MF_{2}(\Gamma_{0}(3)).

Proof.

Combine the previous lemma with Theorem 4.12. ∎

We suspect a similar statement can be made for tmf0(5)tmf_{0}(5), though perhaps a larger power of 2 is necessary.

5 Building Orientations

We can now combine our results thus far to prove the main theorem.

Proof of Theorem 0.4.

Let MM be any of the Thom spectra as in the statement of the theorem. We have a homotopy pullback square

MapE(M,tmf(Γ))\textstyle{\textup{Map}_{E_{\infty}}(M,tmf(\Gamma))\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}pMapE(M,tmf(Γ)p)\textstyle{\prod_{p}\textup{Map}_{E_{\infty}}(M,tmf(\Gamma)^{\wedge}_{p})\ignorespaces\ignorespaces\ignorespaces\ignorespaces}MapE(M,tmf(Γ))\textstyle{\textup{Map}_{E_{\infty}}(M,tmf(\Gamma)_{\mathbb{Q}})\ignorespaces\ignorespaces\ignorespaces\ignorespaces}MapE(M,(ptmf(Γ)p))\textstyle{\textup{Map}_{E_{\infty}}\left(M,\left(\prod_{p}tmf(\Gamma)^{\wedge}_{p}\right)_{\mathbb{Q}}\right)}

and π1\pi_{1} of the lower right hand side is zero, so we get a pullback square on connected components. This mostly completes the proof, except that we are using the Eichler-Schimura relation and the fact that Upψp=pkU_{p}\psi^{p}=p^{k} to see that:

(1Up)(11pψp)=01+pk1Tp=0(1-U_{p})(1-\frac{1}{p}\psi^{p})=0\iff 1+p^{k-1}-T_{p}=0

We are free to replace the condition (2.1.3) on the total mass of the measures by the congruence condition (0.4(d)) once we have shown (§5.2) that there is at least one orientation whose characteristic series satisfies this congruence (c.f. [AHR10, 10.7]). ∎

We will use this theorem below to construct examples of genera valued in topological modular forms with level structure. Before we do, we must recall some preliminary algebraic results concerning Eisenstein series.

5.1 Eisenstein Series and the Eisenstein Measure

Let ΓΓ\mathscr{M}^{\infty}_{\Gamma}\subset\mathscr{M}_{\Gamma} denote the degenerate locus. This is a relative, effective Cartier divisor in Γ\mathscr{M}_{\Gamma} over \mathbb{Z}, and so corresponds to an exact sequence

0\textstyle{0\ignorespaces\ignorespaces\ignorespaces\ignorespaces}𝒪Γ(cusps)\textstyle{\mathcal{O}_{\mathscr{M}_{\Gamma}}(-\textup{cusps})\ignorespaces\ignorespaces\ignorespaces\ignorespaces}𝒪Γ\textstyle{\mathcal{O}_{\mathscr{M}_{\Gamma}}\ignorespaces\ignorespaces\ignorespaces\ignorespaces}𝒪Γ\textstyle{\mathcal{O}_{\mathscr{M}^{\infty}_{\Gamma}}\ignorespaces\ignorespaces\ignorespaces\ignorespaces}0\textstyle{0}

which gives an exact sequence

0\textstyle{0\ignorespaces\ignorespaces\ignorespaces\ignorespaces}ωΓk(cusps)\textstyle{\omega^{\otimes k}_{\mathscr{M}_{\Gamma}}(-\textup{cusps})\ignorespaces\ignorespaces\ignorespaces\ignorespaces}ωΓk\textstyle{\omega^{\otimes k}_{\mathscr{M}_{\Gamma}}\ignorespaces\ignorespaces\ignorespaces\ignorespaces}ωΓk\textstyle{\omega^{\otimes k}_{\mathscr{M}^{\infty}_{\Gamma}}\ignorespaces\ignorespaces\ignorespaces\ignorespaces}0\textstyle{0}

The Hecke correspondence preserves the divisor Γ\mathscr{M}^{\infty}_{\Gamma} so this sequence is Hecke equivariant.

Theorem 5.1.

Let II denote the image of the map H0(Γ,ωk)H0(Γ,ωk)H^{0}(\mathscr{M}_{\Gamma},\omega^{\otimes k})\longrightarrow H^{0}(\mathscr{M}^{\infty}_{\Gamma},\omega^{\otimes k}). When k2k\geq 2, there is a unique, Hecke-equivariant splitting of the exact sequence

0\textstyle{0\ignorespaces\ignorespaces\ignorespaces\ignorespaces}H0(Γ,ωΓk(cusps))\textstyle{H^{0}(\mathscr{M}_{\Gamma},\omega^{\otimes k}_{\mathscr{M}_{\Gamma}}(-\textup{cusps}))\ignorespaces\ignorespaces\ignorespaces\ignorespaces}H0(Γ,ωk)\textstyle{H^{0}(\mathscr{M}_{\Gamma},\omega^{\otimes k})\ignorespaces\ignorespaces\ignorespaces\ignorespaces}I\textstyle{I\ignorespaces\ignorespaces\ignorespaces\ignorespaces}0\textstyle{0}
Proof.

This is classical, see [Eme] for a discussion and further references. ∎

The image of the splitting guaranteed in (5.1) is called the Eisenstein subspace of Mk(Γ)M_{k}(\Gamma) and denotes Eisenk(Γ)\textup{Eisen}_{k}(\Gamma).

We have the following dimension formulae for distinct primes pp and \ell.

dim(Eisenk(Γ0(p)))={2k>21k=2\textup{dim}(\textup{Eisen}_{k}(\Gamma_{0}(p)))=\begin{cases}2&k>2\\ 1&k=2\end{cases}
dim(Eisenk(Γ0(p))={4k>23k=2\textup{dim}(\textup{Eisen}_{k}(\Gamma_{0}(p\ell))=\begin{cases}4&k>2\\ 3&k=2\end{cases}

An explicit basis is given as follows. The qq-expansion of the unnormalized Eisenstein series of weight kk and level 1, for k4k\geq 4 even is given by

Gk=Bk2k+n1(d|ndk1)qnG_{k}=\frac{B_{k}}{2k}+\sum_{n\geq 1}\left(\sum_{d|n}d^{k-1}\right)q^{n}

When k=2k=2, the formal power series still makes sense, but it is not a modular form. However, for any prime pp,

G2(q)pG2(qp)G_{2}(q)-pG_{2}(q^{p})

is the qq-expansion of a modular form of level Γ0(p)\Gamma_{0}(p).

Proposition 5.2.

For k>2k>2 even, the modular forms GkG_{k} and 1pkGk|ψp\frac{1}{p^{k}}G_{k}|_{\psi^{p}} form a basis for Eisenk(Γ0(p))\textup{Eisen}_{k}(\Gamma_{0}(p)), and the modular forms Gk,Gk(p),Gk(),Gk(p)()G_{k},G_{k}^{(p)},G_{k}^{(\ell)},G_{k}^{(p)(\ell)} form a basis for Eisenk(Γ0(p))\textup{Eisen}_{k}(\Gamma_{0}(p\ell)). When k=2k=2, we must remove G2G_{2} from the list.

Proof.

See, for example, [DS05, 4.5.2]. ∎

We now compute the action of UpU_{p} on the space of Eisenstein series.

Proposition 5.3.

Let k>2k>2. Using the basis {Gk(p),Gk1pkGk|ψp}\{G_{k}^{(p)},G_{k}-\frac{1}{p^{k}}G_{k}|_{\psi^{p}}\} for Eisenk(Γ0(p))\textup{Eisen}_{k}(\Gamma_{0}(p)) the action of UpU_{p} is via the matrix

(100pk1)\begin{pmatrix}1&0\\ 0&p^{k-1}\end{pmatrix}

The action of UpU_{p} on Eisenk(Γ0(p))\textup{Eisen}_{k}(\Gamma_{0}(p\ell)) preserves the two dimensional summands {Gk(p),Gk1pGk|ψp)}\{G_{k}^{(p)},G_{k}-\frac{1}{p}G_{k}|_{\psi^{p}})\} and {Gk(p)(),Gk()1pkGk()|ψp}\{G_{k}^{(p)(\ell)},G_{k}^{(\ell)}-\frac{1}{p^{k}}G_{k}^{(\ell)}|_{\psi^{p}}\} and acts on each via the same matrix:

(100pk1)\begin{pmatrix}1&0\\ 0&p^{k-1}\end{pmatrix}

In particular, the fixed points of UpU_{p} acting on Eisenk(Γ0(p))\textup{Eisen}_{k}(\Gamma_{0}(p\ell)) are spanned by {Gk(p),Gk(p)()}\{G_{k}^{(p)},G_{k}^{(p)(\ell)}\}.

Proof.

This is an exercise in the Eichler-Shimura relation and the relation Upψp=pkU_{p}\psi^{p}=p^{k}. ∎

Finally, we recall the existence of the Eisenstein measure.

Theorem 5.4 (Katz).

Let pp and \ell be distinct primes and fix cp×/{±1}c\in\mathbb{Z}_{p}^{\times}/\{\pm 1\}. Then there is a measure μc\mu_{c} on p×/{±1}\mathbb{Z}_{p}^{\times}/\{\pm 1\} valued in V(Γ)V_{\infty}(\Gamma) such that

p×/{±1}x2k𝑑μc=(1c2k)(11pψp)G2k()\int_{\mathbb{Z}_{p}^{\times}/\{\pm 1\}}x^{2k}d\mu_{c}=(1-c^{2k})(1-\frac{1}{p}\psi^{p})G^{(\ell)}_{2k}
Proof.

Take the measure constructed in [Kat77, 3.3.3, 3.4.1] with b=1b=1 and a=ca=c. The extra factor of 2 needed to get a measure on p×/{±1}\mathbb{Z}_{p}^{\times}/\{\pm 1\} follows as in [AHR10, 10.10]. ∎

5.2 Examples

Theorem 5.5.

There exist, up to homotopy, unique EE_{\infty}-ring maps

σOch,σWSig:MSpintmf0(2)\sigma_{\textup{Och}},\sigma_{\textup{WSig}}:\textup{MSpin}\longrightarrow tmf_{0}(2)

refining the Ochanine genus and Witten signature, respectively, as defined in (A.1). In either case, evaluating at the two different cusps of 0(2)\mathscr{M}_{0}(2) gives two different genera

MSpinKO[1/2]\textup{MSpin}\longrightarrow KO[1/2]

one of which is the A^\widehat{A}-genus and the other evaluates to Sign(M)/22d\textup{Sign}(M)/2^{2d} on an oriented manifold of dimension 4d4d.

Proof.

Uniqueness follows from Theorem 0.4 so we need only show existence. This, in turn, follows from the existence of the Eisenstein measure, and the results in the appendix (A.3). ∎

Remark 5.6.

Consider the following diagram:

MStringσ\scriptstyle{\sigma}tmf\textstyle{tmf\ignorespaces\ignorespaces\ignorespaces\ignorespaces}MSpinOchaninetmf0(2)\textstyle{tmf_{0}(2)}

The only map we know about on the right hand side comes from the forgetful map on the corresponding moduli. This map does not yield a commutative diagram, even rationally. Indeed, this follows immediately from the characteristic series calculations in (A.2). In order to get a commutative diagram, we would need to understand the functoriality of 𝒪top\mathcal{O}^{\textup{top}} with respect to isogenies of formal groups. As far as the author knows, this is not treated in the literature.

The previous result makes one wonder if the signature itself is an EE_{\infty}-ring map, a result that does not seem to appear in the literature, but is nevertheless an easy consequence of the work in [AHR10]. We record the result here.

Theorem 5.7.

There exists, up to homotopy, a unique EE_{\infty}-ring map

MSpinKO\textup{MSpin}\longrightarrow KO

refining the LL-genus.

Proof.

This follows already from the work in [AHR10] once one checks the hypotheses on the characteristic series, as in the appendix (A.3). ∎

Theorem 5.8.

There exist EE_{\infty}-ring maps

MSpintmf[1/6]\textup{MSpin}\longrightarrow tmf[1/6]

but none of these make the following diagram commute

MStringσ\scriptstyle{\sigma}tmf[1/6]\textstyle{tmf[1/6]}MSpin\scriptstyle{\not\exists}
Proof.

First we show that the given diagram cannot commute. Suppose it did, then we would have a commutative diagram

MStringσ\scriptstyle{\sigma}tmf[1/6]\textstyle{tmf[1/6]\ignorespaces\ignorespaces\ignorespaces\ignorespaces}MSpinA^\scriptstyle{\widehat{A}}KO5\textstyle{KO^{\wedge}_{5}}

Indeed, the composite

MStringtmf[1/6]KO5\textup{MString}\longrightarrow tmf[1/6]\longrightarrow KO^{\wedge}_{5}

has the same characteristic series as the restriction of the A^\widehat{A}-genus to String-manifolds, so it must be the A^\widehat{A}-genus by [AHR10, 7.12]. Moreover, the K(1)K(1)-localization of MString agrees with the K(1)K(1)-localization of MSpin, c.f. [Hov97, 2.3.1], so this determines the bottom map making the diagram commute. On the other hand, since π4tmf[1/6]=0\pi_{4}tmf[1/6]=0, this would imply that all 44-dimensional Spin-manifolds have trivial A^\widehat{A}-genus (after completing at 55). This is false, for example if KK is a K3 surface then A^(K)=2\widehat{A}(K)=2.

Now we prove existence of such a genus (there are many). It is enough to write down a suitable sequence of modular forms, {gk}k2\{g_{k}\}_{k\geq 2}. We will define

gk:=GkqkGkg_{k}:=G_{k}-q_{k}G_{k}

with qkq_{k}\in\mathbb{Z}. By Theorem 0.4 and [SN14, Def. 27, Thm 40], we are reduced to constructing a sequence {qk}k2\{q_{k}\}_{k\geq 2} of integers such that: (i) q2k+1=0q_{2k+1}=0 for k1k\geq 1, (ii) for all primes pp there is a measure μpMeas(p×/{±1},p)\mu_{p}\in\textup{Meas}(\mathbb{Z}_{p}^{\times}/\{\pm 1\},\mathbb{Z}_{p}) such that

q2k=p×/{±1}x2k𝑑μp,k1q_{2k}=\int_{\mathbb{Z}_{p}^{\times}/\{\pm 1\}}x^{2k}d\mu_{p},\quad k\geq 1

and

0=p×/{±1}x2k𝑑μp0=\int_{\mathbb{Z}_{p}^{\times}/\{\pm 1\}}x^{2k}d\mu_{p}

and (iii) q2=1q_{2}=1. By [SN14, Thm 31], there are uncountably many such sequences satisfying (i) and (ii) and we are free to specify the value of q2q_{2} however we like. ∎

The next natural set of examples are certain complex orientations MUtmf1(N)\textup{MU}\longrightarrow tmf_{1}(N) defined by Hirzebruch, which specialize to a version of the χy\chi_{y} genus at the ramified cusp. In these cases it is more difficult to check the corresponding integrality conditions on the characteristic series, so we leave this to a future paper.

Appendix A Formulae for characteristic series

It is not usually the case that the Hirzebruch characteristic series of a genus is given in the form exp(2tkxk/k!)\textup{exp}\left(2\sum t_{k}x^{k}/k!\right). For the convenience of the reader, we manipulate the formulae for the genera appearing in the paper into this form. We begin by fixing some notation and recalling the definitions in the literature.

Definition A.1.

The following define genera for MSOMSO_{*}.

  1. 1.

    LL-genus

    logSign(x)=n1x2n+12n+1\log_{\textup{Sign}}(x)=\sum_{n\geq 1}\frac{x^{2n+1}}{2n+1}
  2. 2.

    A^\widehat{A}-genus

    expA^(u)=2sinh(u/2)\exp_{\widehat{A}}(u)=2\sinh(u/2)
  3. 3.

    Ochanine genus

    logOch(x)=0xdt12δt2+ϵt4\log_{\textup{Och}}(x)=\int_{0}^{x}\frac{dt}{\sqrt{1-2\delta t^{2}+\epsilon t^{4}}}
  4. 4.

    Witten genus

    uexpWit(u)=u/2sinh(u/2)n=1(1qn)2(1qneu)(1qneu)\frac{u}{\exp_{\textup{Wit}}(u)}=\frac{u/2}{\sinh(u/2)}\prod_{n=1}^{\infty}\frac{(1-q^{n})^{2}}{(1-q^{n}e^{u})(1-q^{n}e^{-u})}
  5. 5.

    Witten signature

    uexpWSig(u)=u/2tanh(u/2)n=1(1+qneu1qneu1+qneu1qneu)/(1+qn1qn)2\frac{u}{\exp_{\textup{WSig}}(u)}=\frac{u/2}{\tanh(u/2)}\prod_{n=1}^{\infty}\left(\frac{1+q^{n}e^{u}}{1-q^{n}e^{u}}\cdot\frac{1+q^{n}e^{-u}}{1-q^{n}e^{-u}}\right)\bigg{/}\left(\frac{1+q^{n}}{1-q^{n}}\right)^{2}
Theorem A.2.

We have the following identities of formal power series (where we have some redundant factors of 2 we’ve added to put them in a more useable form for our purposes.)

uexpSign(u)=exp(2k22k+1(2k11)2kBkukk!)\displaystyle\frac{u}{\exp_{\textup{Sign}}(u)}=\exp\left(2\sum_{k\geq 2}\frac{2^{k+1}(2^{k-1}-1)}{2k}B_{k}\frac{u^{k}}{k!}\right) (1)
uexpA^(u)=exp(2k2Bk2kukk!)\displaystyle\frac{u}{\exp_{\widehat{A}}(u)}=\exp\left(2\sum_{k\geq 2}\frac{-B_{k}}{2k}\frac{u^{k}}{k!}\right) (2)
uexpWSig(u)=exp(2k22Gk(2)ukk!)\displaystyle\frac{u}{\exp_{\textup{WSig}}(u)}=\exp\left(2\sum_{k\geq 2}2G_{k}^{(2)}\frac{u^{k}}{k!}\right) (3)
uexpWit(u)=exp(2k2Gkukk!)\displaystyle\frac{u}{\exp_{\textup{Wit}}(u)}=\exp\left(2\sum_{k\geq 2}G_{k}\frac{u^{k}}{k!}\right) (4)
uexpOch(u)=exp(2k2G~kukk!)\displaystyle\frac{u}{\exp_{\textup{Och}}(u)}=\exp\left(2\sum_{k\geq 2}\widetilde{G}_{k}\frac{u^{k}}{k!}\right) (5)
Proof.

For (2) and (4) see [AHR10, 10.2, 10.9]. For (5) see [Zag88]. Zagier also states the result for the Witten’s signature, but for completeness we include the derivation of this here. The formula (1) for the characteristic series of the signature follows from evaluation at q=0q=0 and substitution of uu for u/2u/2, once we note that

logSign(x)=12(log(1+x)log(1x))\log_{\textup{Sign}}(x)=\frac{1}{2}\left(\log(1+x)-\log(1-x)\right)

and hence

expSign(u)=eueueu+eu=tanh(u)\exp_{\textup{Sign}}(u)=\frac{e^{u}-e^{-u}}{e^{u}+e^{-u}}=\tanh(u)

So we are left with formula (3). By the definition,

log(uexpWSig(u))=log(u/2tanh(u/2))+n1(log(1+qneu1qneu)+log(1+qneu1qneu)2log(1+qn1qn))\log\left(\frac{u}{\exp_{\textup{WSig}}(u)}\right)=\log\left(\frac{u/2}{\tanh(u/2)}\right)+\sum_{n\geq 1}\left(\log\left(\frac{1+q^{n}e^{u}}{1-q^{n}e^{u}}\right)+\log\left(\frac{1+q^{n}e^{-u}}{1-q^{n}e^{-u}}\right)-2\log\left(\frac{1+q^{n}}{1-q^{n}}\right)\right)

Using the Taylor series for log(1x)\log(1-x), we can rewrite this as

log(uexpWSig(u))\displaystyle\log\left(\frac{u}{\exp_{\textup{WSig}}(u)}\right) =\displaystyle= log(u/2tanh(u/2))+2n1d odd(eud+eud2)qndd\displaystyle\log\left(\frac{u/2}{\tanh(u/2)}\right)+2\sum_{n\geq 1}\sum_{d\textup{ odd}}(e^{ud}+e^{-ud}-2)\frac{q^{nd}}{d}
=\displaystyle= log(u/2tanh(u/2))+2n1d oddqndd(2+1k!k0dk(uk+(1)kuk))\displaystyle\log\left(\frac{u/2}{\tanh(u/2)}\right)+2\sum_{n\geq 1}\sum_{d\textup{ odd}}\frac{q^{nd}}{d}\left(-2+\frac{1}{k!}\sum_{k\geq 0}d^{k}(u^{k}+(-1)^{k}u^{k})\right)
=\displaystyle= log(u/2tanh(u/2))+2n1d oddqndd(2+2k!k evendkuk)\displaystyle\log\left(\frac{u/2}{\tanh(u/2)}\right)+2\sum_{n\geq 1}\sum_{d\textup{ odd}}\frac{q^{nd}}{d}\left(-2+\frac{2}{k!}\sum_{k\textup{ even}}d^{k}u^{k}\right)

When k=0k=0, we are left only with the constant (q=0q=0) term from cancellation. When k>0k>0 and even the coefficient of uk/k!u^{k}/k! is

2d odddk1n1qnd2\sum_{d\textup{ odd}}d^{k-1}\sum_{n\geq 1}q^{nd}

which is the same as

2n1qnd|n,d odddk1=Gk(2)(q)Gk(2)(0)2\sum_{n\geq 1}q^{n}\sum_{d|n,d\textup{ odd}}d^{k-1}=G_{k}^{(2)}(q)-G_{k}^{(2)}(0)

Thus it suffices to show that

log(u/2tanh(u/2))=2k22k22kBkukk!\log\left(\frac{u/2}{\tanh(u/2)}\right)=2\sum_{k\geq 2}\frac{2^{k}-2}{2k}B_{k}\frac{u^{k}}{k!}

Recall that tanh(x)=1cosh2(u)\tanh^{\prime}(x)=\frac{1}{\cosh^{2}(u)}, whence

dlog(u/2tanh(u/2))\displaystyle d\log\left(\frac{u/2}{\tanh(u/2)}\right) =\displaystyle= 1u12sinh(u)cosh(u)\displaystyle\frac{1}{u}-\frac{1}{2\sinh(u)\cosh(u)}
=\displaystyle= 1u2eueu\displaystyle\frac{1}{u}-\frac{2}{e^{u}-e^{-u}}
=\displaystyle= 1u2eue2u1\displaystyle\frac{1}{u}-2\frac{e^{u}}{e^{2u}-1}
=\displaystyle= 1u2eu+1e2u1+2e2u1\displaystyle\frac{1}{u}-2\frac{e^{u}+1}{e^{2u}-1}+\frac{2}{e^{2u}-1}
=\displaystyle= 1u2eu1+2e2u1\displaystyle\frac{1}{u}-\frac{2}{e^{u}-1}+\frac{2}{e^{2u}-1}

Now recall the exponential generating function for the Bernoulli numbers is u/(eu1)u/(e^{u}-1), so we get

dlog(u/2tanh(u/2))\displaystyle d\log\left(\frac{u/2}{\tanh(u/2)}\right) =\displaystyle= 1u2k0Bkkuk1(k1)!+2k02k1Bkkuk1(k1)!\displaystyle\frac{1}{u}-2\sum_{k\geq 0}\frac{B_{k}}{k}\frac{u^{k-1}}{(k-1)!}+2\sum_{k\geq 0}2^{k-1}\frac{B_{k}}{k}\frac{u^{k-1}}{(k-1)!}
=\displaystyle= 2k22k11kBkuk1(k1)!\displaystyle 2\sum_{k\geq 2}\frac{2^{k-1}-1}{k}B_{k}\frac{u^{k-1}}{(k-1)!}

(Here we’ve noted that the coefficient of 1/u1/u and the constant term are both zero from cancellation.) Integrating and multiplying by 22\frac{2}{2} gives the result as desired. ∎

Proposition A.3.

We have the following congruences

Bk2k2(2k11)Bk2kmod [1/2],k2\displaystyle-\frac{B_{k}}{2k}\equiv\frac{2(2^{k-1}-1)B_{k}}{2k}\quad\textup{mod }\mathbb{Z}[1/2],\quad k\geq 2 (6)
Bk2k2k+1(2k11)Bk2kmod ,k2\displaystyle-\frac{B_{k}}{2k}\equiv\frac{2^{k+1}(2^{k-1}-1)B_{k}}{2k}\quad\textup{mod }\mathbb{Z},\quad k\geq 2 (7)
GkG~k2Gk(2)mod [1/2],k2\displaystyle G_{k}\equiv\widetilde{G}_{k}\equiv 2G_{k}^{(2)}\quad\textup{mod }\mathbb{Z}[1/2],\quad k\geq 2 (8)
Proof.

Let m(k)m(k) denote the denominator of Bk2k\frac{B_{k}}{2k} for kk even. Then, by [Ada65, 2.7], for pp odd we have

νp(m(k))={0k0 mod (p1)νp(k)+1k0 mod (p1)\nu_{p}(m(k))=\begin{cases}0&k\not\equiv 0\textup{ mod }(p-1)\\ \nu_{p}(k)+1&k\equiv 0\textup{ mod }(p-1)\end{cases}

So for pp odd with (p1)|k(p-1)|k, write k=(p1)pνp(k)uk=(p-1)p^{\nu_{p}(k)}u. Then, since ϕ(pνp(k)+1)=(p1)pνp(k)\phi(p^{\nu_{p}(k)+1})=(p-1)p^{\nu_{p}(k)}, we have

2k=(2(p1)pνp(k))u1 mod pνp(k)+12^{k}=(2^{(p-1)p^{\nu_{p}(k)}})^{u}\equiv 1\textup{ mod }p^{\nu_{p}(k)+1}

thus

(2k2)+11+10 mod pνp(k)+1(2^{k}-2)+1\equiv-1+1\equiv 0\textup{ mod }p^{\nu_{p}(k)+1}

and we see that

2(2k11)+12kBk\frac{2(2^{k-1}-1)+1}{2k}B_{k}

has only powers of 22 in the denominator. The power of 2 that appears is, again by [Ada65, 2.7], 2m+22^{m+2} where k=2muk=2^{m}u. This divides 2k+12^{k+1} only when k+1m+2k+1\geq m+2, which happens as soon as k>0k>0. The final congruences follow from the first one by comparing constant terms on qq-expansions (the other coefficients are all integers). ∎

References

  • [ABG+08] M. Ando, A. J. Blumberg, D. J. Gepner, M. J. Hopkins, and C. Rezk. Units of ring spectra and Thom spectra. ArXiv e-prints, October 2008.
  • [ABG+14a] Matthew Ando, Andrew J. Blumberg, David Gepner, Michael J. Hopkins, and Charles Rezk. An \infty-categorical approach to RR-line bundles, RR-module Thom spectra, and twisted RR-homology. J. Topol., 7(3):869–893, 2014.
  • [ABG+14b] Matthew Ando, Andrew J. Blumberg, David Gepner, Michael J. Hopkins, and Charles Rezk. Units of ring spectra, orientations and Thom spectra via rigid infinite loop space theory. J. Topol., 7(4):1077–1117, 2014.
  • [Ada65] J. F. Adams. On the groups J(X)J(X). II. Topology, 3:137–171, 1965.
  • [AHR10] M. Ando, M.J. Hopkins, and C. Rezk. Multiplicative orientations of KO-theory and the spectrum of topological modular forms (v0.5). May 2010.
  • [Bak89] Andrew Baker. Elliptic cohomology, pp-adic modular forms and Atkin’s operator UpU_{p}. In Algebraic topology (Evanston, IL, 1988), volume 96 of Contemp. Math., pages 33–38. Amer. Math. Soc., Providence, RI, 1989.
  • [Beh14] Mark Behrens. Note on the construction of tmftmf. In Topological Modular forms, volume 201 of AMS Mathematical Surverys and Monographs, pages 131–188. AMS, 2014.
  • [BH14] T. Barthel and D. Heard. The $E_2$-term of the $K(n)$-local $E_n$-based Adams spectral sequence. ArXiv e-prints, October 2014.
  • [Bou79] A. K. Bousfield. The localization of spectra with respect to homology. Topology, 18(4):257–281, 1979.
  • [Bou87] A. K. Bousfield. Uniqueness of infinite deloopings for KK-theoretic spaces. Pacific J. Math., 129(1):1–31, 1987.
  • [Con07] Brian Conrad. Arithmetic moduli of generalized elliptic curves. J. Inst. Math. Jussieu, 6(2):209–278, 2007.
  • [DR73] P. Deligne and M. Rapoport. Les schémas de modules de courbes elliptiques. In Modular functions of one variable, II (Proc. Internat. Summer School, Univ. Antwerp, Antwerp, 1972), pages 143–316. Lecture Notes in Math., Vol. 349. Springer, Berlin, 1973.
  • [DS05] Fred Diamond and Jerry Shurman. A first course in modular forms, volume 228 of Graduate Texts in Mathematics. Springer-Verlag, New York, 2005.
  • [Elk10] N. Elkies. Mahler’s theorem on continuous pp-adic maps via generating functions. May 2010.
  • [Eme] Matthew Emerton. Eisenstein series as sections of line bundles on moduli spaces. MathOverflow. http://mathoverflow.net/q/34653 (version: 2010-08-05).
  • [GHMR05] P. Goerss, H.-W. Henn, M. Mahowald, and C. Rezk. A resolution of the K(2)K(2)-local sphere at the prime 3. Ann. of Math. (2), 162(2):777–822, 2005.
  • [Gro90] Benedict H. Gross. A tameness criterion for Galois representations associated to modular forms (mod pp). Duke Math. J., 61(2):445–517, 1990.
  • [Hir88] Friedrich Hirzebruch. Elliptic genera of level NN for complex manifolds. In Differential geometrical methods in theoretical physics (Como, 1987), volume 250 of NATO Adv. Sci. Inst. Ser. C Math. Phys. Sci., pages 37–63. Kluwer Acad. Publ., Dordrecht, 1988.
  • [HL13] M. Hill and T. Lawson. Topological modular forms with level structure. ArXiv e-prints, December 2013.
  • [Hov97] Mark A. Hovey. vnv_{n}-elements in ring spectra and applications to bordism theory. Duke Math. J., 88(2):327–356, 1997.
  • [Kat77] Nicholas M. Katz. The Eisenstein measure and pp-adic interpolation. Amer. J. Math., 99(2):238–311, 1977.
  • [KM85] Nicholas M. Katz and Barry Mazur. Arithmetic moduli of elliptic curves, volume 108 of Annals of Mathematics Studies. Princeton University Press, Princeton, NJ, 1985.
  • [KS04] I. Kriz and H. Sati. M Theory, Type IIA Superstrings, and Elliptic Cohomology. ArXiv High Energy Physics - Theory e-prints, April 2004.
  • [Kuh89] Nicholas J. Kuhn. Morava KK-theories and infinite loop spaces. In Algebraic topology (Arcata, CA, 1986), volume 1370 of Lecture Notes in Math., pages 243–257. Springer, Berlin, 1989.
  • [LMSM86] L. G. Lewis, Jr., J. P. May, M. Steinberger, and J. E. McClure. Equivariant stable homotopy theory, volume 1213 of Lecture Notes in Mathematics. Springer-Verlag, Berlin, 1986. With contributions by J. E. McClure.
  • [Mah81] Kurt Mahler. pp-adic numbers and their functions, volume 76 of Cambridge Tracts in Mathematics. Cambridge University Press, Cambridge-New York, second edition, 1981.
  • [MR09] Mark Mahowald and Charles Rezk. Topological modular forms of level 3. Pure Appl. Math. Q., 5(2, Special Issue: In honor of Friedrich Hirzebruch. Part 1):853–872, 2009.
  • [SN14] J. Sprang and N. Naumann. p-adic interpolation and multiplicative orientations of KO and tmf (with an appendix by Niko Naumann). ArXiv e-prints, September 2014.
  • [Zag88] Don Zagier. Note on the Landweber-Stong elliptic genus. In Elliptic curves and modular forms in algebraic topology (Princeton, NJ, 1986), volume 1326 of Lecture Notes in Math., pages 216–224. Springer, Berlin, 1988.