Hostname: page-component-7bb8b95d7b-5mhkq Total loading time: 0 Render date: 2024-09-28T17:08:11.301Z Has data issue: false hasContentIssue false

Obstruction theory and the level n elliptic genus

Published online by Cambridge University Press:  03 August 2023

Andrew Senger*
Affiliation:
Department of Mathematics, Harvard University, 1 Oxford Street, Cambridge, MA 02138, USA [email protected]
Rights & Permissions [Opens in a new window]

Abstract

Given a height at most two Landweber exact $\mathbb {E}_\infty$-ring $E$ whose homotopy is concentrated in even degrees, we show that any complex orientation of $E$ which satisfies the Ando criterion admits a unique lift to an $\mathbb {E}_\infty$-complex orientation $\mathrm {MU} \to E$. As a consequence, we give a short proof that the level $n$ elliptic genus lifts uniquely to an $\mathbb {E}_\infty$-complex orientation $\mathrm {MU} \to \mathrm {tmf}_1 (n)$ for all $n\, {\geq}\, 2$.

Type
Research Article
Copyright
© 2023 The Author(s). The publishing rights in this article are licensed to Foundation Compositio Mathematica under an exclusive licence

1. Introduction

Complex-oriented ring spectra play a central role in the chromatic approach to homotopy theory. Given a homotopy associative ring spectrum $E$, recall that a complex orientation is a choice of class $u \in \widetilde {E}^2 (\mathbb {CP}^\infty )$ with the property that its restriction along $S^2 \cong \mathbb {CP}^1 \hookrightarrow \mathbb {CP}^\infty$ is the unit $1 \in E^0 (\ast ) \cong \widetilde {E}^{2} (S^2)$. A complex orientation determines an isomorphism of graded rings

\[ E^* (\mathbb{CP}^\infty) \cong E^* [\![u]\!]. \]

Complex orientations may also be described in terms of the complex cobordism spectrum $\mathrm {MU}$: complex orientations of $E$ are in natural bijection with maps of homotopy associative ring spectra $\mathrm {MU} \to E$. For more background on complex orientations, we refer the reader to [Reference AdamsAda74, Part II].

The complex cobordism spectrum $\mathrm {MU}$ admits much more structure than that of a homotopy associative ring spectrum: it is an $\mathbb {E}_\infty$-ring spectrum. When $E$ also admits the structure of an $\mathbb {E}_\infty$-ring spectrum, it is natural to ask whether a given complex orientation is induced by a map of $\mathbb {E}_\infty$-ring spectra

\[ \mathrm{MU} \to E. \]

We will call such maps $\mathbb {E}_\infty$-complex orientations. The $\mathbb {E}_\infty$-complex orientations of an $\mathbb {E}_\infty$-ring may be viewed as particularly canonical complex orientations.

Indeed, many of the most familiar (not necessarily complex) orientations admit lifts to $\mathbb {E}_\infty$-orientations. For example, the Atiyah–Bott–Shapiro orientations [Reference Atiyah, Bott and ShapiroABS64]

\[ \mathrm{MSpin} \to \mathrm{ko} \quad \text{and} \quad \mathrm{MSpin}^{\mathbb{C}} \to \mathrm{ku} \]

were refined to $\mathbb {E}_\infty$-orientations by Joachim [Reference JoachimJoa04], who gave an explicit geometric construction of such an $\mathbb {E}_\infty$-orientation. Indeed, one expects that any orientation of geometric origin may, with enough care, be refined to an $\mathbb {E}_\infty$-orientation. A more sophisticated example is the Ando–Hopkins–Rezk $\mathbb {E}_\infty$-String orientation of the connective $\mathbb {E}_\infty$-ring of topological modular forms $\mathrm {tmf}$ (see [Reference Ando, Hopkins and RezkAHR10])

\[ \mathrm{MString} \to \mathrm{tmf}, \]

which refines the Witten genus [Reference WittenWit87, Reference WittenWit88]. While it is expected that this $\mathbb {E}_\infty$-orientation has a geometric origin, and much work has gone into developing such a viewpoint (for a small sampling, see [Reference Stolz and TeichnerST04, Reference Stolz and TeichnerST11, Reference CostelloCos10, Reference CostelloCos11, Reference Douglas and HenriquesDH11, Reference Berwick-EvansBE21]), such a description has so far remained elusive. In Theorem 1.7, we will prove that the Hirzebruch level $n$ elliptic genera [Reference HirzebruchHir88], [Reference WittenWit88, § 5] for $n \geq 2$ may be lifted to $\mathbb {E}_\infty$-complex orientations

\[ \mathrm{MU} \to \mathrm{tmf}_1 (n). \]

1.1 The Ando criterion

An algebraic approximation of what it means for a complex orientation $\mathrm {MU} \to E$ to be $\mathbb {E}_\infty$ is given by the Ando criterion. It asks that the complex orientation be compatible in a suitable sense with the power operations in $E$. In many cases, the Ando criterion is equivalent to the property that the complex orientation $\mathrm {MU} \to E$ be a map of $\mathbb {H}_\infty$-ring spectra (see, e.g., [Reference Ando, Hopkins and StricklandAHS04, Proposition 6.1]).

Now let $E$ denote an $\mathbb {E}_\infty$, or more generally an $\mathbb {H}_\infty$-ring spectrum, with a fixed complex orientation $u$.

Notation 1.1 Given a complex vector bundle $V \to X$ of dimension $d$, we let $t_u (V) \in E^{2d} (\mathrm {Th}(V))$ denote the Thom class of $V$.

Fix a prime $p$, let $\rho$ denote the vector bundle over $\mathrm {BC}_p$ corresponding to the complex regular representation of $C_p$, and let $\gamma _1$ denote the tautological bundle over $\mathbb {CP}^\infty$. Let $I_{\rm tr} \subset E^* (\mathrm {BC}_p)$ denote the transfer ideal. Recall from [Reference Hopkins and LawsonHL18, § 7] that there are additive power operations

\[ \Psi_u : E^{2*} (\mathrm{Th}(\gamma_1)) \to E^{2p*} (\mathrm{Th}(\rho \boxtimes \gamma_1))/I_{\rm tr}. \]

Definition 1.2 We say that a complex orientation of $E$ satisfies the Ando criterion at the prime $p$ if

\[ \Psi_u (t_u (\gamma_1)) = t_u (\rho \boxtimes \gamma_1) \]

in $E^{2p} (\mathrm {Th}(\rho \boxtimes \gamma _1))/I_{\rm tr}$. We say that a complex orientation of $E$ satisfies the Ando criterion if it satisfies the Ando criterion for all primes $p$.

Remark 1.3 If $E$ is $p$-local, then $E^* (\mathrm {BC}_\ell ) / I_{\rm tr} = 0$ for all primes $\ell \neq p$. Therefore, a complex orientation of a $p$-local $\mathbb {E}_\infty$-ring satisfies the Ando criterion if and only if it satisfies the Ando criterion at $p$.

The complex cobordism spectrum $\mathrm {MU}$, equipped with the canonical complex orientation, satisfies the Ando criterion.Footnote 1 It follows that any $\mathbb {E}_\infty$-complex orientation, or more generally any $\mathbb {H}_\infty$-complex orientation, satisfies the Ando criterion.

1.2 Results

The first main theorem of this paper states that for many $\mathbb {E}_\infty$-rings of height at most two, this condition is sufficient, and that the resulting $\mathbb {E}_\infty$-complex orientations are determined up to homotopy by their underlying complex orientations.

Theorem 1.4 Let $E$ denote a height at most two Landweber exact $\mathbb {E}_\infty$-ring whose homotopy is concentrated in even degrees. Then any complex orientation $\mathrm {MU} \to E$ which satisfies the Ando criterion lifts uniquely up to homotopy to an $\mathbb {E}_\infty$-ring map $\mathrm {MU} \to E$.

In the above theorem, we say that a Landweber exact ring spectrum $E$ is of height at most $n$ if $v_n \in \pi _* (E) / (p, v_1, \ldots, v_{n-1})$ is a unit for all primes $p$.Footnote 2 As a corollary, we obtain the following result for height at most two Lubin–Tate theories.Footnote 3

Corollary 1.5 Let $k \subseteq \overline {\mathbb {F}}_p$ denote a field of characteristic $p > 0$ which is algebraic over the prime field $\mathbb {F}_p$, and let $\mathbb {G}$ denote a formal group of height at most two over $k$. Then any complex orientation of the associated $2$-periodic Morava $K$-theory $K(k,\mathbb {G})$ lifts uniquely up to homotopy to an $\mathbb {E}_\infty$-complex orientation of $E(k, \mathbb {G})$

\begin{align*} \mathrm{MU} \to E(k, \mathbb{G}). \end{align*}

Proof. This follows immediately from Theorem 1.4 and [Reference ZhuZhu20, Theorem 1.2], which implies that every complex orientation of $K(k,\mathbb {G})$ admits a unique lift to a complex orientation of $E(k,\mathbb {G})$ which satisfies the Ando criterion whenever $k$ is algebraic over $\mathbb {F}_p$.

Remark 1.6 Theorem 1.4 was inspired by recent work of Balderrama [Reference BalderramaBal21, Theorem 6.5.3]. Balderrama showed that every periodic complex orientation of a Lubin–Tate theory of height at most two satisfying an analogue of the Ando criterion lifts to an $\mathbb {E}_\infty$-orientation. He also showed that $\mathbb {E}_\infty$-refinements of periodic complex orientations of even periodic $K(1)$-local $\mathbb {E}_\infty$-rings exist whenever the Ando criterion is satisfied and are unique up to homotopy.

The key observation he made is the presence of evenness in the Goerss–Hopkins obstruction theory for (periodic) $\mathbb {E}_\infty$-complex orientations, which implies that the obstructions to existence and uniqueness appearing in his theorem vanish for formal reasons. Our results about existence of $\mathbb {E}_\infty$-complex orientations will be obtained by observing a similar evenness in the Hopkins–Lawson obstruction theory [Reference Hopkins and LawsonHL18].

In contrast to Theorem 1.4, Balderrama did not obtain any uniqueness results at height $2$ (see [Reference BalderramaBal21, Remark 6.5.4]). In Remark 3.6, we will prove that $\mathbb {E}_\infty$-refinements of periodic complex orientations of height $2$ Lubin–Tate theories are unique.

Our second main theorem uses Theorem 1.4 to give a simple proof of the following theorem.

Theorem 1.7 For $n \geq 2$, the Hirzebruch level $n$ elliptic genus lifts uniquely up to homotopy to a map of $\mathbb {E}_\infty$-rings

\[ \mathrm{MU} \to \mathrm{tmf}_1 (n). \]

While Theorem 1.4 does not apply directly to $\mathrm {tmf}_1 (n)$, it does apply to $\mathrm {TMF}_1 (n)$, and it is not hard to upgrade the resulting $\mathbb {E}_\infty$-complex orientation to one of $\mathrm {tmf}_1 (n)$. This reduces us to verifying the Ando criterion, which may be done following the strategy of Ando, Hopkins, and Strickland [Reference Ando, Hopkins and StricklandAHS04].

Remark 1.8 During the writing process, [Reference AbsmeierAbs21] has also appeared, which follows the strategy of Ando, Hopkins, and Rezk [Reference Ando, Hopkins and RezkAHR10] to prove a similar result to Theorem 1.7 for $\mathrm {Tmf}_1 (n)$. Our method is rather different from that of Absmeier and completely avoids the consideration of $p$-adic Eisenstein measures.

1.3 Further questions

One of the key inputs in our proof of Theorem 1.4 is Theorem 2.11, which states that the Morava $K$-theory of certain finite groups is concentrated in even degrees. This is closely related to these groups being good in the sense of Hopkins, Kuhn, and Ravenel [Reference Hopkins, Kuhn and RavenelHKR00, § 7].

These groups come in a family, and to see that the Hopkins–Lawson obstruction theory is concentrated in even degrees one would like to show that the entire family has Morava $K$-theory concentrated in even degrees; see Remark 2.13.

Question 1.9 Is the Morava $K$-theory of the groups $\Gamma _k ^{(n)}$ of Definition 2.7 concentrated in even degrees for $k > 2$, at least for $n \gg 0$? The groups $\Gamma _k ^{(1)}$ are the extraspecial $p$-groups of type $p^{1+2k} _+$.

One could also ask about $C_2$-equivariant refinement of our results.

Question 1.10 The complex cobordism spectrum may be refined to a $C_2$-equivariant $\mathbb {E}_\infty$-ring $\mathrm {MU}_\mathbb {R}$, and $\mathrm {tmf}_1 (n)$ admits the natural structure of a $C_2$-equivariant $\mathbb {E}_\infty$-ring [Reference MeierMei21, Theorem 2.20]. Moreover, the Hirzebruch level $n$ elliptic genus admits a refinement to a map of homotopy $C_2$-ring spectra $\mathrm {MU}_\mathbb {R} \to \mathrm {tmf}_1 (n)$ (see [Reference MeierMei21, Theorem 3.5]). Is there a suitable $C_2$-equivariant analogue of Theorem 1.4 which may be used to prove a $C_2$-equivariant refinement of Theorem 1.7? See [Reference Hopkins and LawsonHL18, Remark 13] for a comment on a $C_2$-equivariant version of the Hopkins–Lawson obstruction theory.

On the other hand, it would be very interesting to study $\mathbb {E}_\infty$-complex orientations at heights $3$ and above. A natural choice of spectra to study would be Lubin–Tate spectra. Since the obstruction groups do not vanish for formal reasons at these heights, it seems likely that this will require an explicit analysis of the Goerss–Hopkins obstruction theory for $\mathbb {E}_\infty$-maps $\mathrm {MU} \to E(k,\mathbb {G})$. In particular, one would have to compute the $\mathrm {E}_2$-page.

Problem 1.11 Compute the Goerss–Hopkins obstruction groups for $\mathbb {E}_\infty$-maps $\mathrm {MU} \to E(k,\mathbb {G})$ for height $3$ and above Lubin–Tate theories $E(k,\mathbb {G})$.

2. Existence of $\mathbb {E}_\infty$-orientations

In this section, we will prove the half of Theorem 1.4 concerning the existence of $\mathbb {E}_\infty$-complex orientations. The main tool that we will utilize is an obstruction theory for $\mathbb {E}_\infty$-complex orientations studied by Hopkins and Lawson [Reference Hopkins and LawsonHL18]. Given this obstruction theory, the existence half of Theorem 1.4 reduces to the statement that certain obstruction groups vanish. Using work of Arone and Lesh [Reference Arone and LeshAL07], this can be further reduced to the evenness of the Morava $K$-theory of certain extraspecial $p$-groups and related groups, which we are able to extract from the literature.

2.1 Hopkins–Lawson obstruction theory

We begin by summarizing the main results of the Hopkins–Lawson obstruction theory [Reference Hopkins and LawsonHL18]. First, a definition.

Definition 2.1 Let $E$ denote a homotopy commutative ring spectrum. We let $\mathrm {Or}(E)$ denote the space of complex orientations of $E$, i.e. the fiber

\[ \mathrm{Or}(E) \to \mathrm{Map} (\Sigma^{\infty-2}\, \mathbb{CP}^\infty, E) \to \mathrm{Map}(\Sigma^{\infty-2}\, \mathbb{CP}^2, E) \simeq \mathrm{Map}(\mathbb{S}, E) \]

above the unit map $\mathbb {S} \to E$.

Theorem 2.2 [Reference Hopkins and LawsonHL18, Theorems 1 and 32]

There exists a diagram of $\mathbb {E}_\infty$-ring spectra

\[ \mathbb{S} \to \mathrm{MX}_1 \to \mathrm{MX}_2 \to \mathrm{MX}_3 \to \cdots \to \mathrm{MU} \]

such that the following hold:

  1. (i) the natural map $\varinjlim \mathrm {MX}_n \to \mathrm {MU}$ is an equivalence;

  2. (ii) the $\mathbb {E}_\infty$-ring $\mathrm {MX}_1$ is equipped with a natural complex orientation inducing an equivalence $\mathrm {Map}_{\mathbb {E}_\infty } (\mathrm {MX}_1, E) \xrightarrow {\sim } \mathrm {Or}(E)$ for each $\mathbb {E}_\infty$-ring $E$;

  3. (iii) given $m > 0$ and an $\mathbb {E}_\infty$-ring $E$, there is a pullback square

    where $F_m$ is a pointed space described in Recollection 2.6;
  4. (iv) the map $\mathrm {MX}_{m-1} \to \mathrm {MX}_m$ is a rational equivalence if $m > 1$, a $p$-local equivalence if $m$ is not a power of $p$, and a $K(n)$-local equivalence if $m > p^n$;

  5. (v) let $E$ denote an $\mathbb {E}_\infty$-ring such that $\pi _* E$ is $p$-local and $p$-torsion free. Then an $\mathbb {E}_\infty$-map $\mathrm {MX}_1 \to E$ extends to an $\mathbb {E}_\infty$-map $\mathrm {MX}_p \to E$ if and only if the corresponding complex orientation of $E$ satisfies the Ando criterion.

Using this theorem, we will reduce the proof of Theorem 1.4 to the following.

Lemma 2.3 Let $E$ denote a $p$-complete Landweber exact ring spectrum with homotopy concentrated in even degrees. Then $E^{2n} (F_p) \cong E^{2n+1} (F_{p^2}) \cong 0$ for all $n \in \mathbb {Z}$.

Remark 2.4 In fact, we only need that $E^{2n+1} (F_{p^2}) \cong 0$. However, we include the statement $E^{2n} (F_p) \cong 0$ since it is no harder for us to prove. This extra evenness implies uniqueness up to homotopy for $\mathbb {E}_\infty$-refinements of complex orientations of height at most one. However, we will prove uniqueness in a different way in § 3.

Question 2.5 Given a $p$-complete Landweber exact ring spectrum $E$ with homotopy concentrated in even degrees, is $E^{2*+k-1} (F_{p^k}) \cong 0$ for $k \geq 3$?

Proof Proof of existence in Theorem 1.4 assuming Lemma 2.3

We begin by reducing to the case where $E$ is $p$-complete for some prime $p$. We make use of the following fracture square (see [Reference BousfieldBou79, Proposition 2.9]).

Note that $E^{\wedge } _p$ is again even and Landweber exact; see Appendix A.

By Theorem 2.2, we have $\mathrm {Map}_{\mathbb {E}_\infty } (\mathrm {MU}, R) \simeq \mathrm {Or}(R)$ for a rational $\mathbb {E}_\infty$-ring $R$; it follows further that $\pi _1 \mathrm {Map}_{\mathbb {E}_\infty } (\mathrm {MU}, R) \cong \pi _1 \mathrm {Or}(R) \cong 0$ if $R$ has homotopy concentrated in even degrees. As a consequence, there are pullback squares of sets:

(1)

and

To lift a complex orientation of $E$ to an $\mathbb {E}_\infty$-complex orientation, it therefore suffices to lift the induced complex orientation of $E^{\wedge }_p$. We may therefore assume that $E$ is $p$-complete.

Let $E$ now denote an $p$-complete Landweber exact $\mathbb {E}_\infty$-ring with homotopy concentrated in even degrees. Using Theorem 2.2, we see that it suffices to show that

\[ \pi_0 \mathrm{Map}_{\mathbb{E}_\infty} (\mathrm{MX}_{p^2}, E) \to \pi_0 \mathrm{Map}_{\mathbb{E}_\infty} (\mathrm{MX}_p, E) \]

is surjective.

By Theorem 2.2, there is a fiber sequence

\[ \mathrm{Map}_{\mathbb{E}_\infty} (\mathrm{MX}_{p^2}, E) \to \mathrm{Map}_{\mathbb{E}_\infty} (\mathrm{MX}_p, E) \to \mathrm{Map}_* (F_{p^2}, \mathrm{Pic}(E)). \]

Now, there are equivalences

\[ \mathrm{Map}_* (F_m, \mathrm{Pic}(E)) \simeq \mathrm{Hom} (\Sigma^{\infty} F_m, \mathrm{pic}(E)) \simeq \mathrm{Hom} (\Sigma^{\infty} F_m, \Sigma E), \]

where in the second equivalence we have used the fact that $\Sigma ^\infty F_m$ is $(2m-1)$-connected by [Reference Hopkins and LawsonHL18, Corollary 4(5)]. It therefore follows from the above fiber sequence that it suffices to show that

\[ E^1 (\Sigma^{\infty} F_{p^2}) \cong 0. \]

Since $E$ is $p$-complete, this follows from Lemma 2.3.

2.2 Proof of Lemma 2.3

In the remainder of this section, we will prove Lemma 2.3. First, we must recall the definition of the spaces $F_m$.

Recollection 2.6 Let $L_m$ denote the nerve of the topologized poset of proper direct-sum decompositions of $\mathbb {C}^m$, and let $(L_m)^{\diamond }$ denote its unreduced suspension. The natural action of $U(m)$ on $\mathbb {C}^m$ endows $L_m$ and $(L_m)^{\diamond }$ with the structure of $U(m)$-spaces.

Furthermore, view $S^{2m}$ as a $U(m)$-space via its identification with the one-point compactification of $\mathbb {C}^m$, viewed as the fundamental representation of $U(m)$. Then $F_m$ is given by

\[ F_m \simeq ((L_m)^{\diamond} \wedge S^{2m})_{h U(m)}. \]

To prove Lemma 2.3, we will use a result of Arone–Lesh to reduce the study of the $E$-cohomology of $F_m$ to the $E$-cohomology of certain groups $\Gamma _k$, whose definition we now recall.

Definition 2.7 [Reference OliverOli94, Definition 1]

Let $\sigma _0, \ldots, \sigma _{k-1} \in \Sigma _{p^k}$ denote the permutations

\[ \sigma_r (i) = \begin{cases} i + p^r & \text{ if } i \equiv 1, \ldots, (p-1)p^r \mod p^{r+1}, \\ i-(p-1) p^r & \text{ if } i \equiv (p-1)p^r +1, \ldots, p^{r+1} \mod p^{r+1}. \end{cases} \]

We let $\Gamma _k \subset U(p^k)$ denote the subgroup generated by the permutation matrices corresponding to $\sigma _0, \ldots, \sigma _{k-1}$, the central $S^1$, and the diagonal matrices $A_0, \ldots, A_{k-1}$ given by

\[ (A_r)_{ii} = \zeta_p ^{\lfloor (i-1)/p^r \rfloor}, \]

where $\zeta _p$ is a primitive $p$th root of unity. Then $\Gamma _k$ lies in a central extension

\[ 1 \to S^1 \to \Gamma_k \to \mathbb{F}_p ^{2k} \to 1. \]

For each $n \geq 1$, there is a normal subgroup $\Gamma _k ^{(n)} \subset \Gamma _k$ which only contains the central $p^{n}$th roots of unity instead of the full $S^1$. Then there are central extensions

\[ 1 \to C_{p^n} \to \Gamma_k ^{(n)} \to \mathbb{F}_p^{2k} \to 1 \]

and exact sequences

\[ 1 \to \Gamma_k ^{(n)} \to \Gamma_k \to S^1 \to 1. \]

Remark 2.8 The groups $\Gamma _k ^{(1)}$ are examples of extraspecial $p$-groups, and in this language are commonly denoted by $p^{1+2k} _+$.

Proposition 2.9 [Reference Arone and LeshAL07, Propositions 9.6 and 10.3]

The $p$-completion of the spectrum

\[ \Sigma^\infty F_m \simeq \Sigma^{\infty} ((L_m)^{\diamond} \wedge S^{2m})_{hU(m)} \]

is null unless $m = p^k$, in which case it is a summand of the $p$-completion of

\[ \Sigma^{k} (S^{2p^k}\!)_{h\Gamma_k}, \]

where $\Gamma _k$ acts on $S^{2p^k}$ via the inclusion $\Gamma _k \subset U(p^k)$.

Now, $(S^{2p^k})_{h\Gamma _k}$ may also be described as the Thom spectrum associated to the composition

\[ B\Gamma_k \to \mathrm{BU} (p^k\!) \to \mathbb{Z} \times \mathrm{BU}, \]

from which it follows that

\[ E^* ((S^{2p^k}\!)_{h\Gamma_k}) \cong \widetilde{E}^{*-2p^k} (B\Gamma_k). \]

Lemma 2.3 therefore reduces to the following lemma.

Lemma 2.10 Let $E$ denote a $p$-local Landweber exact ring spectrum whose homotopy is concentrated in even degrees. Then, for $k \leq 2$, $E^* (B\Gamma _k\!)$ is concentrated in even degrees.

We will deduce Lemma 2.10 from the following Morava $K$-theory computations.

Theorem 2.11 For all $n \geq 0$, the following groups are concentrated in even degrees:

  1. (i) (Tezuka and Yagita [Reference Tezuka and YagitaTY89, Theorem 4.2]) $K(n)^* (B\Gamma _1 ^{(1)})$ at all primes $p$;

  2. (ii) (Schuster and Yagita [Reference Schuster and YagitaSY04, Theorem 5.4]) $K(n)^* (B\Gamma _2 ^{(1)})$ at the prime $2$;

  3. (iii) (Yagita [Reference YagitaYag05, Theorem 1.2]) $K(n)^* (B\Gamma _2 ^{(2)})$ at odd primes $p$.

A lemma of Strickland which builds on the work of Ravenel, Wilson, and Yagita [Reference Ravenel, Wilson and YagitaRWY98] allows us to transport this evenness from Morava $K$-theory to $E$-cohomology.

Corollary 2.12 Given a $p$-local Landweber exact ring spectrum $E$ whose homotopy is concentrated in even degrees, the following groups are concentrated in even degrees:

  1. (i) $E^* (B\Gamma _1 ^{(1)})$ at all primes $p$;

  2. (ii) $E^* (B\Gamma _2 ^{(1)})$ at the prime $2$;

  3. (iii) $E^* (B\Gamma _2 ^{(2)})$ at odd primes $p$.

Proof. Combine Theorem 2.11 with [Reference StricklandStr99, Lemma 8.25].Footnote 4$^,$Footnote 5

Proof of Lemma 2.10 The short exact sequence

\[ \Gamma_k ^{(n)} \to \Gamma_k \to S^1 \]

induces a fiber sequence

\[ B\Gamma_k ^{(n)} \to B\Gamma_k \to BS^1. \]

Using the associated Atiyah–Hirzebruch spectral sequence

\[ \mathrm{H}^* (BS^1; E^* (B\Gamma_k ^{(n)})) \Rightarrow E^* (B\Gamma_k), \]

we find that if $E^* (B\Gamma _k ^{(n)})$ is even for some $n$, then $E^* (B\Gamma _k)$ must be as well. Therefore, Lemma 2.10 follows from Corollary 2.12.

Remark 2.13 By the same arguments, to give a positive answer to Question 2.5 it suffices to show that $K(n)^* (B\Gamma _k ^{(i_k)})$ is concentrated in even degrees for a fixed $i_k$ not depending on $n$. This is closely related to the question of whether $\Gamma _k ^{(i_k)}$ is a good group in the sense of Hopkins, Kuhn, and Ravenel [Reference Hopkins, Kuhn and RavenelHKR00, § 7].

3. Uniqueness of $\mathbb {E}_\infty$-orientations

Our goal in this section is to prove the following result.

Theorem 3.1 Let $E$ denote an $L_2$-localFootnote 6 complex orientable $\mathbb {E}_\infty$-ring with the property that $K(1)_* E$ and $K(1)_* L_{K(2)} E$ are concentrated in even degrees at all primes $p$. Then each complex orientation of $E$ admits at most one refinement to an $\mathbb {E}_\infty$-complex orientation up to homotopy.

Example 3.2 Any Landweber exact ring spectrum $E$ of height at most two whose homotopy is concentrated in even degrees satisfies the hypotheses of Theorem 3.1. By Lemma A.1, $L_{K(2)} E$ is again Landweber exact and has homotopy concentrated in even degrees. It therefore suffices to show that $K(1)_* E$ is concentrated in even degrees. This is true because $K(1)_* E \cong K(1)_* \mathrm {MU} \otimes _{\pi _* \mathrm {MU}} \pi _* E$, and $K(1)_* \mathrm {MU}$ is concentrated in even degrees.

Combining Theorem 3.1 with Example 3.2, we obtain the uniqueness half of Theorem 1.4.

Example 3.3 The ring spectra $\mathrm {Tmf}_1 (n)$ satisfy the hypotheses of Theorem 3.1. This follows from [Reference WilsonWil15, Propositions 2.4 and 2.6], which imply that the $p$-complete complex $K$-theory of these spectra and their $K(2)$-localizations is $p$-torsionfree and concentrated in even degrees. The proof is exactly the same as the case without level structure, which is treated in [Reference Behrens, Douglas, Francis, Henriques and HillBeh14]. The key point is that the Igusa tower becomes formally affine at a finite stage.

Our proof of Theorem 3.1 will be based on the orientation theory of [Reference Ando, Blumberg, Gepner, Hopkins and RezkABGHR14] and the following lemma.

Lemma 3.4 Let $E$ denote an $\mathrm {MU}$-module with the property that $K(1)_* E$ is concentrated in even degrees. Then $[\mathrm {KU}_p, L_{K(1)} E]$ is torsionfree and $[\Sigma \mathrm {KU}_p, L_{K(1)} E] = 0$.

Definition 3.5 We say that a $(\mathrm {KU}_p)_*$-module is pro-free if it is the $p$-completion of a free module. Moreover, given a spectrum $X$, we write $\mathrm {KU}^{\vee }_* (X)$ for $\pi _* L_{K(1)} (\mathrm {KU} \otimes X)$.

Proof. Since $K(1)_* (\mathrm {KU}_p)$ is even, $\mathrm {KU}^{\vee } _* (\mathrm {KU}_p)$ is pro-free by [Reference Hovey and StricklandHSt99, Proposition 8.4(f)]. Therefore, by [Reference Barthel and HeardBH16, Proposition 1.14], there is an isomorphism

\[ \pi_* \mathrm{Hom} (\mathrm{KU}_p, L_{K(1)} (\mathrm{KU}_p \otimes E) \cong \mathrm{Hom}_{(\mathrm{KU}_p)_*} (\mathrm{KU}^{\vee}_* (\mathrm{KU}_p), \mathrm{KU}^{\vee}_* (E)). \]

By assumption, $K(1)_* (E)$ is even and, hence, $\mathrm {KU}^{\vee } _* (E)$ is even and pro-free by [Reference Hovey and StricklandHSt99, Proposition 8.4(f)]. In particular, it is torsionfree, so that $\mathrm {Hom}_{(\mathrm {KU}_p)_*} (\mathrm {KU}^{\vee } _* (\mathrm {KU}_p), \mathrm {KU}^{\vee } _* (E))$ is even and torsionfree.

The result then follows from the following facts.

  1. (i) Since $E$ is an $\mathrm {MU}$-module, the unit map $L_{K(1)} E \to L_{K(1)} (\mathrm {MU} \otimes E)$ admits a splitting given by the module structure map.

  2. (ii) There is a splitting of the map $L_{K(1)} \mathrm {MU} \to L_{K(1)} E(1)$ which is compatible with the unit map [Reference Hovey and SadofskyHSa99, Theorem 4.1].

  3. (iii) There is an equivalence of spectra $\mathrm {KU}_p \simeq \bigoplus _{i=0} ^{p-2} \Sigma ^{2i} L_{K(1)} E(1)$.

Proof of Theorem 3.1 Using the pullback square of sets (1), we may assume that $E$ is $p$-complete.

Recall the map $\Sigma ^{\infty } \mathbb {CP}^{\infty } \to \mathrm {bu}$ which is adjoint to the canonical map $\mathbb {CP}^\infty \to \mathrm {BU}$. By orientation theory [Reference Ando, Blumberg, Gepner, Hopkins and RezkABGHR14], we must show that

\[ [\mathrm{bu}, \mathrm{gl}_1 (E)] \to [\Sigma^{\infty} \mathbb{CP}^{\infty}, \mathrm{gl}_1 (E)] \]

is injective. Since $E$ is $p$-complete, $\mathrm {gl}_1 (E)$ agrees with $\mathrm {gl}_1 (E)^{\wedge } _p$ in degrees ${\geq }2$, so we may as well replace the former by the latter. Letting $F$ denote the fiber of the map $\mathrm {gl}_1 (E) \to L_2 \mathrm {gl}_1 (E)$, we find that there is a fiber sequence

\[ F^{\wedge} _p \to \mathrm{gl}_1 (E)^{\wedge}_p \to L_{K(1) \oplus K(2)} \mathrm{gl}_1 (E). \]

It therefore suffices to show that

\[ [\mathrm{bu}, F^{\wedge}_p] \to [\Sigma^{\infty} \mathbb{CP}^{\infty}, F^{\wedge}_p] \]

and

\[ [\mathrm{bu}, L_{K(1) \oplus K(2)} \mathrm{gl}_1 (E)] \to [\Sigma^{\infty} \mathbb{CP}^{\infty}, L_{K(1) \oplus K(2)} \mathrm{gl}_1 (E)] \]

are injective. The first is injective because $F^{\wedge } _p$ is $3$-coconnective by [Reference Ando, Hopkins and RezkAHR10, Theorem 4.11] and the cofiber of $\Sigma ^{\infty } \mathbb {CP}^{\infty } \to \mathrm {bu}$ is $4$-connective.

To prove that the second is injective, we first note that the Bousfield–Kuhn functor [Reference BousfieldBou87, Reference KuhnKuh89] and the chromatic fracture square for $L_{K(1) \oplus K(2)} \mathrm {gl}_1 (E)$ imply that there is an exact sequence

\[ [\Sigma \mathrm{KU}_p, L_{K(1)} L_{K(2)} E] \to [\mathrm{bu}, L_{K(1) \oplus K(2)} E] \to [\mathrm{KU}_p, L_{K(1)} E]. \]

Applying Lemma 3.4, we learn that $[\mathrm {bu}, L_{K(2) \oplus K(1)} E]$ is torsionfree. It therefore injects into its rationalization, so that the result follows from the fact that $\Sigma ^{\infty } \mathbb {CP}^{\infty } \to \mathrm {bu}$ is a rational equivalence.

Remark 3.6 Balderrama has shown that periodic complex orientations of height $2$ Lubin–Tate theories $E(k,\mathbb {G})$ which satisfy a version of the Ando criterion admit lifts to periodic $\mathbb {E}_\infty$-complex orientations $\mathrm {MUP} \to E(k,\mathbb {G})$ (see [Reference BalderramaBal21, Theorem 6.5.3(3)]). In this remark, we prove that such lifts are unique up to homotopy.

We begin with a result which has been proven by Rezk at height $2$ (which is the case that we use) [Reference RezkRez13] and in general is an unpublished theorem of Hopkins and Lurie that has now appeared in work of Burklund, Schlank, and Yuan [Reference Burklund, Schlank and YuanBSY22, Theorem H]. Let $\overline {k}$ denote the algebraic closure of $k$. Then we have

\[ \pi_* \mathbb{G}_m (E(\overline{k}, \mathbb{G})) := \pi_* \mathrm{Map} (\mathbb{Z}, \mathrm{gl}_1 (E(\overline{k}, \mathbb{G}))) \cong \begin{cases} \overline{k}^{\times} & *=0 ,\\ \mathbb{Z}_p & *=3, \\ 0 & \text{otherwise}. \end{cases} \]

On $\pi _0$, the map $\mathbb {Z} \to \mathrm {gl}_1 (E(\overline {k}, \mathbb {G}))$ corresponding to $a \in \overline {k}^{\times }$ picks out the Teichmuller lift $[a]$.

Since $\mathrm {Gal}(k)$ has $p$-cohomological dimension at most $1$ (see [Reference Gille and SzamuelyGS06, Proposition 6.1.9]), it follows that

\[ \pi_0 \mathbb{G}_m (E(k, \mathbb{G})) \cong k^{\times}. \]

In particular, the map $\pi _0 \mathbb {G}_m (E(k,\mathbb {G})) \to \pi _0 \mathrm {gl}_1 (E(k,\mathbb {G}))$ is injective. Now, by orientation theory it suffices to show that

\[ [\mathrm{ku}, \mathrm{gl}_1 (E(k,\mathbb{G}))] \to [\Sigma^\infty _+ \mathbb{CP}^\infty, \mathrm{gl}_1 (E(k,\mathbb{G}))] \]

is injective. By what we have proven above about uniqueness of $\mathbb {E}_\infty$-complex orientations, it suffices to show that

\[ [\mathbb{Z}, \mathrm{gl}_1 (E(k,\mathbb{G}))] \to [\mathbb{S}^0, \mathrm{gl}_1 (E(k, \mathbb{G}))] \]

is injective, which is what we showed above.

As noted in [Reference BalderramaBal21, Remark 6.5.4], this implies that the Goerss–Hopkins obstruction group $\mathrm {Ext}^2_{\Delta } (\hat {Q}(E(k,\mathbb {G})_0 ^{\wedge } \mathrm {MUP}), \omega )$ is equal to $0$.

4. The level $n$ elliptic genus

Convention 4.1 In this section $n$ will denote an integer greater than or equal to $2$.

In this section, we will prove Theorem 1.7, which states that the Hirzebruch level $n$ elliptic genus lifts uniquely up to homotopy to an $\mathbb {E}_\infty$-complex orientation $\mathrm {MU} \to \mathrm {tmf}_1 (n)$.

We will begin by recalling some background material about $\mathbb {E}_\infty$-rings of topological modular forms with level-$\Gamma _1 (n)$ structures in § 4.1. In § 4.2, we recall from [Reference Ando, Hopkins and StricklandAHS01, Reference Ando, Hopkins and StricklandAHS04] how complex orientations may be described in terms of $\Theta ^1$-structures. We then describe the level $n$ elliptic genus in this language, following Meier [Reference MeierMei21, § 3]. In § 4.3, we show that a complex orientation for $\mathrm {TMF}_1 (n)$ satisfies the Ando criterion at $p$ if and only if its composition along a map $\mathrm {TMF}_1 (n) \to E(k, \mathbb {G})$ to a Lubin–Tate theory does.

In § 4.4, we recall from [Reference Ando, Hopkins and StricklandAHS04] how the Ando criterion for Lubin–Tate theories may be rephrased in terms of $\Theta ^1$-structures. We then prove that the Hirzebruch level $n$ genus satisfies the Ando criterion. As a consequence of Theorem 1.4, it lifts uniquely up to homotopy to an $\mathbb {E}_\infty$-ring map $\mathrm {MU} \to \mathrm {TMF}_1 (n)$. Finally, in § 4.5, we prove that this $\mathbb {E}_\infty$-ring map admits a unique up to homotopy lift to an $\mathbb {E}_\infty$-ring map $\mathrm {MU} \to \mathrm {tmf}_1 (n)$, completing the proof of Theorem 1.7.

4.1 Topological modular forms with level structures

In this section, we will recall some basic facts about the $\mathbb {E}_\infty$-rings $\mathrm {tmf}_1 (n)$, $\mathrm {Tmf}_1 (n)$ and $\mathrm {TMF}_1 (n)$ of topological modular forms with level-$\Gamma _1 (n)$ structure. We begin by recalling the algebraic background.

Definition 4.2 We let $\mathcal {M}_1 (n)$ denote the Deligne–Mumford moduli stack of elliptic curves with level $\Gamma _1 (n)$ structure over $\mathbb {Z}[ {1}/{n}]$. Concretely, given a scheme $S$ on which $n$ is invertible, we have

\[ \mathcal{M}_1 (n) (S) = \text{elliptic curves } E \text{ over } S \text{ with a point } P \in E[n] (S) \text{ of exact order }n. \]

Moreover, we write $\overline {\mathcal {M}_1} (n)$ for the Deligne–Rapoport moduli stack of generalized elliptic curves with level $\Gamma _1 (n)$ structure [Reference Deligne and RapoportDR73]. This is again a Deligne–Mumford stack over $\mathbb {Z}[ {1}/{n}]$, and $\mathcal {M}_1 (n) \subset \overline {\mathcal {M}_1} (n)$ is an open substack.

We denote the universal family of curves by $\pi : \mathcal {C} \to \overline {\mathcal {M}_1} (n)$, and write $\pi ^{\mathrm {sm}} : \mathcal {C}^{\mathrm {sm}} \to \overline {\mathcal {M}_1} (n)$ for the smooth locus. Then $\mathcal {C}^{\mathrm {sm}}$ admits a natural structure of a group scheme, and we write $\omega$ for the line bundle of invariant differentials.

Finally, we write $\mathcal {M} = \mathcal {M}_1 (1)$ and $\overline {\mathcal {M}} = \overline {\mathcal {M}_1} (1)$ for the moduli stacks of (generalized) elliptic curves without level structure.

Recollection 4.3 (Goerss, Hopkins, and Miller [Reference Douglas, Francis, Henriques and HillDFHH14, Chapter 12], Lurie [Reference LurieLur18b], and Hill and Lawson [Reference Hill and LawsonHL16])

There is a sheaf $\mathcal {O}^{\mathrm {top}}_{\overline {\mathcal {M}}}$ of $\mathbb {E}_\infty$-ring spectra on the Kummer log-étale site of $\overline {\mathcal {M}}$ with the following properties.

  1. (i) There is a natural isomorphism of sheaves $\pi _0 \mathcal {O}^{\mathrm {top}}_{\overline {\mathcal {M}}} \cong \mathcal {O}_{\overline {\mathcal {M}}}$.

  2. (ii) There are natural isomorphisms of quasicoherent sheaves $\pi _{2i} \mathcal {O}^{\mathrm {top}} _{\overline {\mathcal {M}}} \cong \omega ^{i}$ and $\pi _{2i+1} \mathcal {O}^{\mathrm {top}}_{\overline {\mathcal {M}}} \cong 0$.

  3. (iii) Write $\widehat {\pi ^{\mathrm {sm}}} : \widehat {\mathcal {C}^{\mathrm {sm}}} \to \overline {\mathcal {M}}$ for the completion of $\pi ^{\mathrm {sm}} : \mathcal {C}^{\mathrm {sm}} \to \overline {\mathcal {M}}$ along the zero section. There is a natural isomorphism of sheaves of rings $\pi _0 \mathrm {Hom} (\Sigma ^\infty _+ \mathbb {CP}^{\infty }, \mathcal {O}^{\mathrm {top}}) \cong (\widehat {\pi ^{\mathrm {sm}}})_* \mathcal {O}_{\widehat {\mathcal {C}^{\mathrm {sm}}}}$.

Since the natural morphisms $\overline {\mathcal {M}_1} (n) \to \overline {\mathcal {M}}$ are Kummer log-étale (we refer the reader to [Reference Hill and LawsonHL16] for more details), we may define $\mathbb {E}_\infty$-rings:

\[ \mathrm{TMF}_1 (n) := \Gamma(\mathcal{M}_1 (n), \mathcal{O}^{\mathrm{top}}_{\overline{\mathcal{M}}}) \]

and

\[ \mathrm{Tmf}_1 (n) := \Gamma(\overline{\mathcal{M}_1 (n)}, \mathcal{O}^{\mathrm{top}}_{\overline{\mathcal{M}}}). \]

By definition, there are spectral sequences

\[ \mathrm{H}^s (\mathcal{M}_1 (n), \omega^i) \Rightarrow \pi_{2i-s} \mathrm{TMF}_1 (n) \]

and

\[ \mathrm{H}^s (\overline{\mathcal{M}_1} (n), \omega^i) \Rightarrow \pi_{2i-s} \mathrm{Tmf}_1 (n). \]

Remark 4.4 Let $n \geq 2$. It follows from [Reference MeierMei22a, Proposition 2.4(4)] that $\mathrm {H}^s (\mathcal {M}_1 (n), \omega ^i) = 0$ for all $s > 0$, so that $\mathrm {TMF}_1 (n)$ has homotopy groups concentrated in even degrees and that there are natural isomorphisms

\[ \pi_{2i} (\mathrm{TMF}_1 (n)) \cong \Gamma(\mathcal{M}_1 (n), \omega^i). \]

Moreover, $\mathrm {TMF}_1 (n)$ is Landweber exact. Indeed, this is a consequence of flatness of $\mathcal {M}_1 (n)$ over $\mathbb {Z}[ {1}/{n}]$, the integrality of $\mathcal {M}_1 (n)_{\mathbb {F}_p}$ and the fact that the formal group of an elliptic curve is of height at most two.

However, because the groups $\mathrm {H}^1 (\overline {\mathcal {M}_1}, \omega ^i)$ do not in general vanish, we do not have a similar theorem for $\mathrm {Tmf}_1 (n)$. Instead, we have the $\mathbb {E}_\infty$-ring spectrum $\mathrm {tmf}_1 (n)$ from [Reference MeierMei21].

Recollection 4.5 [Reference MeierMei21, Theorem 1.1]

There is an essentially unique connective $\mathbb {E}_\infty$-ring spectrum $\mathrm {tmf}_1 (n)$ whose homotopy groups are concentrated in even degrees and which is equipped with an $\mathbb {E}_\infty$-ring map

\[ \mathrm{tmf}_1 (n) \to \mathrm{Tmf}_1 (n) \]

such that the induced maps

\[ \pi_{2i} \mathrm{tmf}_1 (n) \to \pi_{2i} \mathrm{Tmf}_1 (n) \to \Gamma(\overline{\mathcal{M}_1} (n), \omega^{i}) \]

are isomorphisms.

Remark 4.6 There is a sequence of natural maps

\[ \mathrm{tmf}_1 (n) \to \mathrm{Tmf}_1 (n) \to \mathrm{TMF}_1 (n). \]

4.2 $\Theta ^1$-structures

In this section, we describe complex orientations in terms of $\Theta ^1$-structures and give a description of the Hirzebruch level $n$ genus in this language.

Suppose that we are given a base Deligne–Mumford stack $S$ and a formal group or generalized elliptic curve $G$ over $S$. We denote the structure map by $p : G \to S$ and the zero section by $0 : S \to G$. Given a line bundle $\mathcal {L}$ on $G$, we set

\[ \Theta^1 (\mathcal{L}) := p^* 0^* \mathcal{L} \otimes \mathcal{L}^{-1}. \]

There is a canonical trivialization $0^* \Theta ^1 (\mathcal {L}) \cong \mathcal {O}_{S}$.

Definition 4.7 A $\Theta ^1$-structure on a line bundle $\mathcal {L}$ over $G$ is a trivialization of $\Theta ^1 (\mathcal {L})$ which pulls back to the canonical trivialization of $0^* \Theta ^1 (\mathcal {L})$.

Equivalently, a $\Theta ^1$-structure on $\mathcal {L}$ is an isomorphism $p^* 0^* \mathcal {L} \cong \mathcal {L}$ which pulls back to the canonical isomorphism $0^* p^* 0^* \mathcal {L} \cong 0^* \mathcal {L}$.

Remark 4.8 It is clear from the definition that there is a natural bijection between $\Theta ^1$-structures on $\mathcal {L}$ and $\mathcal {L}^{-1}$.

Remark 4.9 Note that a line bundle $\mathcal {L}$ on $G$ admits a $\Theta ^1$-structure if and only if it is pulled back from a line bundle on $S$ via the structure map $p$.

When $G$ is a generalized elliptic curve, $\Theta ^1$-structures are unique when they exist.

Lemma 4.10 Suppose that $G$ is a generalized elliptic curve. Then the set of $\Theta ^1$-structures on a line bundle $\mathcal {L}$ over $G$ is either empty or consists of a single element.

Proof. The set of isomorphisms $p^* 0^* \mathcal {L} \cong \mathcal {L}$ is a torsor for $\mathcal {O}_G (G)^{\times }$, whereas the set of isomorphisms $0^* p^* 0^* \mathcal {L} \cong 0^* \mathcal {L}$ is a torsor for $\mathcal {O}_S (S)^{\times }$. It therefore suffices to show that the map $\mathcal {O}_G (G) \to \mathcal {O}_S (S)$ induced by pulling back along the zero section is an isomorphism. This follows from [Sta23, Tag 0E0L].

Now let $E$ denote an even weakly periodic homotopy commutative ring spectrum, and let $\mathbb {G}_E = \mathrm {Spf} E^0 (\mathbb {CP}^\infty )$ denote the associated formal group over $\text {Spec}\ \pi _0 E$. We denote by $\gamma _1$ the canonical line bundle over $\mathbb {CP}^\infty$ and let $\mathcal {O}_{\mathbb {G}_E} (1)$ denote the line bundle over $\mathbb {G}_E$ corresponding to

\[ E^0 (\mathrm{Th}(\gamma_1)) \cong \widetilde{E^0} (\mathbb{CP}^\infty) \cong \ker (\mathcal{O}_{\mathbb{G}_E} \xrightarrow{0^*} \pi_0 E). \]

Then we have the following theorem.

Proposition 4.11 [Reference Ando, Hopkins and StricklandAHS01, Theorem 2.48]

There is a natural bijection between complex orientations of $E$ and $\Theta ^1$-structures on $\mathcal {O}_{\mathbb {G}_E} (1)$.

Proof. Recall that a complex orientation of $E$ consists of an element of $\widetilde {E^2} (\mathbb {CP}^{\infty })$ which restricts to the unit along the map $\widetilde {E^2} (\mathbb {CP}^\infty ) \to \widetilde {E^2} (\mathbb {CP}^1) \cong \pi _0 E$. Since the map $E^0 (\mathbb {CP}^\infty ) \to \pi _0 E$ is an infinitesimal thickening, complex orientations of $E$ may be identified with $E^0 (\mathbb {CP}^\infty )$-module isomorphisms $E^0 (\mathbb {CP}^{\infty }) \cong \widetilde {E^2} (\mathbb {CP}^{\infty })$ that become equal to the canonical isomorphism $\widetilde {E^2} (\mathbb {CP}^1) \cong \pi _0 E$ after tensoring down along $E^0 (\mathbb {CP}^\infty ) \to \pi _0 E$.

By Remark 4.8, we may replace $\mathcal {O}_{\mathbb {G}_E} (1)$ with $\mathcal {O}_{\mathbb {G}_E} (1)^{-1}$ in the statement of the proposition. To prove the proposition, it therefore suffices to identify the global sections of $\Theta ^1 (\mathcal {O}_{\mathbb {G}_E} (1)^{-1})$ with $\widetilde {E^2} (\mathbb {CP}^{\infty })$ and pullback along the zero section with $\widetilde {E^2} (\mathbb {CP}^\infty ) \to \widetilde {E^2} (\mathbb {CP}^1) \cong \pi _0 E$. This is an immediate consequence of the definitions.

Even though $\mathrm {tmf}_1 (n)$ is not weakly even periodic, Meier has shown that its complex orientations may still be described in terms of $\Theta ^1$-structures.

Proposition 4.12 [Reference MeierMei21, Lemma 3.2]

There is a natural bijection between complex orientations of $\mathrm {tmf}_1 (n)$ and $\Theta ^1$-structures on $\mathcal {O}_{\widehat {\mathcal {C}}} (1)$ over $\widehat {C} \to \overline {\mathcal {M}_1} (n)$.

We now recall the treatment of Hirzebruch's level $n$ elliptic genus from [Reference MeierMei21, § 3]. The following proposition is a mild rephrasing of [Reference MeierMei21, Lemma 3.3].

Proposition 4.13 Let $P : \overline {\mathcal {M}_1} (n) \to \mathcal {C}^{\mathrm {sm}}$ denote the universal level $\Gamma _1 (n)$ structure.

  1. (i) The pullback of the line bundle $\mathcal {O}_{\mathcal {C}} ([0] - [P])$ on $\mathcal {C}$ to $\widehat {\mathcal {C}}$ is naturally isomorphic to $\mathcal {O}_{\widehat {\mathcal {C}}} (1)$.

  2. (ii) There is a degree $n$ étale cover $q : \mathcal {C}' \to \mathcal {C}$ of generalized elliptic curves so that $q^* \mathcal {O}_{\mathcal {C}} ([0] - [P])$ admits a (necessarily unique) $\Theta ^1$-structure.

Since the induced map $\widehat {q} : \widehat {\mathcal {C}'} \xrightarrow {\sim } \widehat {\mathcal {C}}$ is an isomorphism, we obtain a $\Theta ^1$-structure on $\mathcal {O}_{\widehat {C}} (1)$ and, hence, by Proposition 4.12 a complex orientation of $\mathrm {tmf}_1 (n)$. This is the complex orientation corresponding to the Hirzebruch level $n$ elliptic genus.

4.3 Reduction to Lubin–Tate theory

Our goal in this section is to prove Proposition 4.16. This proposition implies that to verify the Ando criterion for the Hirzebruch level $n$ elliptic genus, it suffices to verify the Ando criterion after composition with a map $\mathrm {TMF}_1 (n) \to E(k, \mathbb {G})$ to a Lubin–Tate theory. This will be useful to us because work of Ando, Hopkins, and Strickland [Reference Ando, Hopkins and StricklandAHS04] rephrases the Ando criterion for Lubin–Tate theories in terms of $\Theta ^1$-structures.

We begin by recalling some basic facts about Lubin–Tate theories.

Recollection 4.14 (Goerss and Hopkins [Reference Goerss and HopkinsGH04] and Lurie [Reference LurieLur18b])

Let $k$ denote a perfect field of characteristic $p > 0$ and let $\mathbb {G}$ denote a formal group of finite height $h$ over $k$. To the pair $(k, \mathbb {G})$ one may associate an $\mathbb {E}_\infty$-ring spectrum $E(k,\mathbb {G})$, known as the Lubin–Tate spectrum of $(k,\mathbb {G})$. The ring $\pi _0 E(k,\mathbb {G})$ is naturally isomorphic to the universal deformation ring of $(k,\mathbb {G})$, which is non-canonically isomorphic to $\mathbb {W}(k) [\![u_1, \ldots, u_{h-1}]\!]$. We let $\mathbb {G}^{\mathrm {univ}}$ denote the universal deformation of $\mathbb {G}$ over $\pi _0 E(k, \mathbb {G})$, and denote by $\omega _{\mathbb {G}^{\mathrm {univ}}}$ its module of invariant differentials. There are natural isomorphisms of $\pi _{2i} E(k,\mathbb {G})$ with $\omega ^{i} _{\mathbb {G}^{\mathrm {univ}}}$.

Construction 4.15 Let $k$ denote a perfect field of characteristic $p > 0$. Associated to a supersingular $k$-point $(E, \alpha \in E[n] (k))$ of $\mathcal {M}_1 (n)$, there is a map of $\mathbb {E}_\infty$-ring spectra $\mathrm {TMF}_1 (n) \to E(k, \widehat {E})$. Indeed, this follows from the description of $E(k, \widehat {E})$ as an oriented deformation ring [Reference LurieLur18b, § 6], Lurie's Serre–Tate theorem for strict abelian varieties [Reference LurieLur18a, § 7], and the universal property of $\mathrm {TMF}_1 (n)$.

Proposition 4.16 Let $k$ denote a perfect field of characteristic $p$ and let $(E, \alpha \in E[n](k))$ denote an object of $\mathcal {M}_1 (n)(k)$ with $E$ a supersingular elliptic curve. Then a complex orientation of $\mathrm {TMF}_1 (n)$ satisfies the Ando criterion at $p$ if and only if its composite with the canonical map $\mathrm {TMF}_1 (n) \to E(k, \widehat {E})$ does.

Our proof of Proposition 4.16 rests on the following lemma.

Lemma 4.17 Let $E \to F$ denote a map of complex orientable homotopy commutative ring spectra. Suppose that $\pi _* E$ and $\pi _* F$ are $p$-torsionfree and that the induced map $\pi _* E / p \to \pi _* F / p$ is an injection. Then the map $E^* (\mathrm {BC}_p)/ I_{\rm tr} \to F^* (\mathrm {BC}_p)/ I_{\rm tr}$ is an injection.

Proof. Since $\pi _* E$ is $p$-torsionfree, a choice of complex orientation gives rise to an isomorphism (see [Reference QuillenQui71, Proposition 4.2]):

\[ E^* (\mathrm{BC}_p)/I_{\rm tr} \cong E^* [[t]] / \langle p \rangle (t) \]

where $\lvert t \rvert = 2$, $\langle p \rangle (t) = {[p](t)}/{t}$, and $[p] (t)$ is the $p$-series of the formal group on $E^*$. Moreover, since $\pi _* E$ is $p$-torsionfree, we may identify $t^{n-1} E^* [[t]] / (\langle p \rangle (t), t^n)$ with a shift of $E^* / p$. The analogous statements for $F$ also hold.

Taking the induced complex orientation of $F$, we identify $E^* (\mathrm {BC}_p) / I_{\rm tr} \to F^* (\mathrm {BC}_p) / I_{\rm tr}$ with the natural map $E^* [[t]] / \langle p \rangle (t) \to F^* [[t]] / \langle p \rangle (t)$. To show that this map is injective, it suffices to show that it is injection on the associated graded for the $t$-adic filtration. This identifies with a shift of the natural map $E^* / p \to F^* / p$ in each degree, which is an injection by hypothesis.

Proof of Proposition 4.16 It is clear that if a complex orientation of $\mathrm {TMF}_1 (n)$ satisfies the Ando criterion at $p$, so does the induced complex orientation of $E(k, \widehat {E})$. To prove the converse, it suffices to show that the induced map

\[ \mathrm{TMF}_1 (n)^* (\mathrm{BC}_p)/I_{\rm tr} \to E(k, \widehat{E})^* (\mathrm{BC}_p)/I_{\rm tr} \]

is an injection. We begin by reducing to the case $n \geq 5$. Any map $\text {Spec}\ k \to \mathcal {M}_1 (n)$ fits into a diagram

for $k'$ a finite separable extension of $k$. It follows that there is a diagram

so that it suffices to show that

\[ \mathrm{TMF}_1 (n)^* (\mathrm{BC}_p) / I_{\rm tr} \to \mathrm{TMF}_1 (n^2)^* (\mathrm{BC}_p) / I_{\rm tr} \]

and

\[ \mathrm{TMF}_1 (n^2)^* (\mathrm{BC}_p) / I_{\rm tr} \to E(k', \widehat{E}_{k'})^* (\mathrm{BC}_p) / I_{\rm tr} \]

are injective.

We may therefore assume that $n \geq 5$ if we can show that the first map is an injection. Both $\pi _* \mathrm {TMF}_1 (n)$ and $\pi _* \mathrm {TMF}_1 (n^2)$ are $p$-torsionfree, and

\[ \pi_{2i} \mathrm{TMF}_1 (n) / p \cong \Gamma(M_1 (n)_{\mathbb{F}_p}, \omega^{i}) \to \Gamma(M_1 (n^2)_{\mathbb{F}_p}, \omega^{i}) \cong \pi_{2i} \mathrm{TMF}_1 (n) / p \]

is an injection since $\mathcal {M}_1 (n^2)_{\mathbb {F}_p} \to \mathcal {M}_1 (n)_{\mathbb {F}_p}$ is a finite étale cover. Therefore, we may apply Lemma 4.17 to show that the first map above is an injection.

We may now assume that $n \geq 5$, so that $\mathcal {M}_1 (n)$ is represented by an affine scheme $\text {Spec}\ R_n$ (see [Reference MeierMei22a, Proposition 2.4(2)]). Let $\mathfrak {m} \subset R_n$ denote the kernel of the map $R_n \to k$. Then $R_n / \mathfrak {m}$ is a finite field (hence, perfect), and the pair $(E, P \in E[n](k))$ descends to $R_n/ \mathfrak {m}$. As a consequence, there is a factorization $\mathrm {TMF}_1 (n) \to E(R_n/ \mathfrak {m}, \widehat {E}) \to E(k, \widehat {E})$. It follows immediately from Lemma 4.17 that the induced map

\[ E(R_n/ \mathfrak{m}, \widehat{E})^* (\mathrm{BC}_p) / I_{\rm tr} \to E(k, \widehat{E})^* (\mathrm{BC}_p)/ I_{\rm tr} \]

is an injection, so that it suffices to show that

\[ \mathrm{TMF}_1 (n)^* (\mathrm{BC}_p) / I_{\rm tr} \to E(R_n/ \mathfrak{m}, \widehat{E})^* (\mathrm{BC}_p)/ I_{\rm tr} \]

is an injection. To apply Lemma 4.17, we need to show that $\pi _{*} \mathrm {TMF}_1 (n) / p \to \pi _{*} E(R_n / \mathfrak {m}, \widehat {E}) / p$ is an injection. By abuse of notation, we let $\omega$ denote the invertible $R_n$-module corresponding to the line bundle $\omega$ on $\mathcal {M}_1 (n)$. Then the above map can be identified with the $\mathfrak {m}$-adic completion map $(\omega ^{*/2} / p) \to (\omega ^{*/2} / p)^{\wedge } _\mathfrak {m}$. This is an injection by the Krull intersection theorem, since $\mathcal {M}_1 (n)_{\mathbb {F}_p}$ is an integral scheme.

4.4 The Ando criterion and $\Theta ^1$-structures

In this section, we recall from the work of Ando, Hopkins, and Strickland how the Ando criterion for Lubin–Tate theories may be rephrased in terms of $\Theta ^1$-structures. We refer the reader to [Reference Ando, Hopkins and StricklandAHS04] for proofs and further details. We then combine this rephrasing with Proposition 4.16 to prove that the Hirzebruch level $n$ elliptic genus satisfies the Ando criterion. Finally, we deduce from Theorem 1.4 that the Hirzebruch level $n$ elliptic genus lifts uniquely up to homotopy to an $\mathbb {E}_\infty$-ring map

\[ \mathrm{MU} \to \mathrm{TMF}_1 (n). \]

We begin by recalling from [Reference Ando, Hopkins and StricklandAHS04, § 14] how $\Theta ^1$-structures may be normed along isogenies.

Recollection 4.18 [Reference Ando, Hopkins and StricklandAHS04, § 14]

Suppose we are given an isogeny $G \to G'$ of formal groups or elliptic curves. Given a line bundle $\mathcal {L}$ over $G$, there is a line bundle $N(\mathcal {L})$ over $G'$, called the norm of $\mathcal {L}$. Moreover, given a $\Theta ^1$-structure $s$ on $\mathcal {L}$, there is an associated $\Theta ^1$-structure $N(s)$ on $N(\mathcal {L})$, called the norm of $s$.

Remark 4.19 Given an isogeny of formal groups $\mathbb {G} \to \widetilde {\mathbb {G}}$, there is a natural isomorphism of line bundles $N(\mathcal {O}_\mathbb {G} (1)) \cong \mathcal {O}_{\widetilde {\mathbb {G}}} (1)$.

Remark 4.20 Suppose that we are given an elliptic curve $E$ with a point $P$ of exact order $n$, i.e. a level $\Gamma _1 (n)$ structure, and an isogeny $E \to \widetilde {E}$ of degree $p$ coprime to $n$. Then the image $\widetilde {P}$ of $P$ in $\widetilde {E}$ is again a point of exact order $n$, and there is a natural isomorphism of line bundles $N (\mathcal {O}_E ([0] - [P])) \cong \mathcal {O}_{\widetilde {E}} ([0] - [\widetilde {P}])$.

Following [Reference Ando, Hopkins and StricklandAHS04], we may now use this language to rephrase the Ando criterion for Lubin–Tate theory. We begin with some setup.

Recollection 4.21 Let $k$ denote a perfect field of characteristic $p > 0$ and let $\mathbb {G}$ denote a formal group of finite height over $k$, so that we have an associated Lubin–Tate theory $E(k,\mathbb {G})$. Then there are two ring maps $i, \psi : \pi _0 E(k, \mathbb {G}) \to E(k,\mathbb {G})^0 (\mathrm {BC}_p) / I_{\rm tr}$. The first, $i$, is induced by the projection $\mathrm {BC}_p \to \ast$. The second $\psi$, is the total power operation. Over the ring $E(k,\mathbb {G})^0 (\mathrm {BC}_p) /I_{\rm tr}$, there is a degree $p$ isogeny

\[ i^* \mathbb{G}^{\mathrm{univ}} \to \psi^* \mathbb{G}^{\mathrm{univ}} \]

induced by the total power operation on $E(k, \mathbb {G})^0 (\mathbb {CP}^\infty )$.

Given a $\Theta ^1$-structure $s$ on $\mathcal {O}_{\mathbb {G}^{\mathrm {univ}}} (1)$, there are therefore two naturally induced $\Theta ^1$-structures on $\mathcal {O}_{\psi ^* \mathbb {G}^{\mathrm {univ}}} (1) \cong N(\mathcal {O}_{i^* \mathbb {G}^{\mathrm {univ}}} (1))$: the pullback $\psi ^* (s)$ and the norm $N(i^* (s))$.

Definition 4.22 We say that a $\Theta ^1$-structure $s$ on $\mathbb {G}^{\mathrm {univ}}$ satisfies the Ando criterion if $\psi ^* (s) = N(i^* (s))$.

It follows from [Reference Ando, Hopkins and StricklandAHS04, § 5] that this is compatible with our previous definition of the Ando criterion.

Proposition 4.23 A complex orientation of $E(k, \mathbb {G})$ satisfies the Ando criterion if and only if the associated $\Theta ^1$-structure on $\mathcal {O}_{\widetilde {\mathbb {G}}} (1)$ satisfies the Ando criterion.

We are now able to prove the main theorem of this section.

Theorem 4.24 The Hirzebruch level $n$ elliptic genus $\mathrm {MU} \to \mathrm {tmf}_1 (n) \to \mathrm {TMF}_1 (n)$ satisfies the Ando criterion. As a consequence of Theorem 1.4 and Remark 4.4, it lifts uniquely up to homotopy to an $\mathbb {E}_\infty$-complex orientation

\[ \mathrm{MU} \to \mathrm{TMF}_1 (n). \]

Proof. Choose, for each $p$ not dividing $n$, $(E, P \in E[n] (k)) \in \mathcal {M}_1 (n)(k)$ with $E$ supersingular and $k$ a perfect field of characteristic $p$. By Proposition 4.16, it suffices to show that the induced complex orientation of $E(k,\widehat {E})$ satisfies the Ando criterion.

By the Serre–Tate theorem [Reference KatzKat81, § 1], $\pi _0 E(k, \widehat {E})$ is the universal deformation ring of $E$. We let $E^{\mathrm {univ}}$ denote the universal deformation of $E$ over $\pi _0 E(k, \widehat {E})$, and let $P^\mathrm {univ} \in E^{\mathrm {univ}}[n](\pi _0 E(k,\mathbb {G}))$ denote the unique lift of $P$. The associated formal group $\widehat {E^{\mathrm {univ}}}$ is a universal deformation of $\widehat {E}$.

Applying Proposition 4.23, we must show that the $\Theta ^1$-structure $s$ on $\mathcal {O}_{\widehat {E^\mathrm {univ}}} (1)$ corresponding to the level $n$ elliptic genus satisfies the Ando criterion, i.e. that $\psi ^* (s) = N(i^* (s))$. By the definition of the level $n$ elliptic genus, there is a degree $n$ étale isogeny $q: (E^{\mathrm {univ}})' \to E^\mathrm {univ}$ and a $\Theta ^1$-structure $\overline {s}$ on $q^* \mathcal {O}_{E^{\mathrm {univ}}} ([0] - [P^{\mathrm {univ}}])$ which induces $s$. By the Serre–Tate theorem, the isogeny of formal groups

\[ i^* \widehat{E^{\mathrm{univ}}} \to \psi^* \widehat{E^{\mathrm{univ}}} \]

over $E(k, \widehat {E})^0 (\mathrm {BC}_p)/ I_{\rm tr}$ lifts to the following diagram of isogenies of elliptic curves.

Let $\overline {P^{\mathrm {univ}}}$ denote the image of $i^*(P^\mathrm {univ} )$ in $\psi ^* E^\mathrm {univ}$. From the above diagram, we obtain $\Theta ^1$-structures $\psi ^* (\overline {s})$ and $N(i^* (\overline {s}))$ on

\[ (q_\psi )^* \mathcal{O}_{\psi^* E^\mathrm{univ}} ([0]-[\overline{P^\mathrm{univ}}]) \cong N ((q_i )^* \mathcal{O}_{i^* E^\mathrm{univ}}([0]-[i^*(P^\mathrm{univ})])). \]

As these induce $\psi ^* (s)$ and $N( i^* (s))$, it suffices to show that $\psi ^* (\overline {s}) = N(i^* (\overline {s}))$. But this follows immediately from the uniqueness of $\Theta ^1$-structures over elliptic curves proven in Lemma 4.10.

4.5 Lift to $\mathrm {tmf}_1 (n)$

In this section, we will complete the proof of Theorem 1.7 by proving the following two lemmas.

Lemma 4.25 Suppose that we are given a complex orientation $\mathrm {MU} \to \mathrm {Tmf}_1 (n)$ with the property that the composite

\[ \mathrm{MU} \to \mathrm{Tmf}_1 (n) \to \mathrm{TMF}_1 (n) \]

lifts to an $\mathbb {E}_\infty$-ring map. Then this complex orientation lifts uniquely to an $\mathbb {E}_\infty$-complex orientation $\mathrm {MU} \to \mathrm {Tmf}_1 (n)$.

Lemma 4.26 Any $\mathbb {E}_\infty$-complex orientation of $\mathrm {Tmf}_1 (n)$ lifts uniquely to an $\mathbb {E}_\infty$-complex orientation of $\mathrm {tmf}_1 (n)$.

Proof of Lemma 4.25 The uniqueness will follow from Theorem 3.1 and Example 3.3 once we know that a complex orientation of $\mathrm {Tmf}_1 (n)$ is determined by the induced complex orientation of $\mathrm {TMF}_1 (n)$. This follows from the fact that the map $\pi _* \mathrm {Tmf}_1 (n) \to \pi _* \mathrm {TMF}_1 (n)$ is injective in even degrees by [Reference MeierMei22b, Proposition 2.5] and the descent spectral sequence.

It therefore suffices to show that the $\mathbb {E}_\infty$-map $\mathrm {MU} \to \mathrm {TMF}_1 (n)$ lifts to an $\mathbb {E}_\infty$-map $\mathrm {MU} \to \mathrm {Tmf}_1 (n)$. We begin with the pullback square of $\mathbb {E}_\infty$-rings coming from [Reference Hill and LawsonHL16]:

This square satisfies the following properties.

  1. (i) The $\mathbb {E}_\infty$-rings $K^{\mathrm {cusp}}_1 (n)$ and $\Delta ^{-1} K^{\mathrm {cusp}} _1 (n)$ are Landweber exact and of height at most one.

  2. (ii) The induced map $\pi _* K^{\mathrm {cusp}}_1 (n)/p \to \pi _* \Delta ^{-1} K^{\mathrm {cusp}} _1 (n)/p$ is injective for all $p$.

It follows that to construct an $\mathbb {E}_\infty$-lifting $\mathrm {MU} \to \mathrm {Tmf}_1 (n)$, it suffices to lift the composite

\[ \mathrm{MU} \to \mathrm{TMF}_1 (n) \to \Delta^{-1} K^{\mathrm{cusp}} _1 (n) \]

to an $\mathbb {E}_\infty$-map

\[ \mathrm{MU} \to K^{\mathrm{cusp}} _1 (n). \]

By the uniqueness in Theorem 1.4 and (i) above, it suffices to lift the complex orientation and verify that it satisfies the Ando criterion. However, we are given a lift of the complex orientation by assumption, and it follows from Lemma 4.17 and (ii) above that it satisfies the Ando criterion.

Proof of Lemma 4.26 For this, we use orientation theory [Reference Ando, Blumberg, Gepner, Hopkins and RezkABGHR14]. We have the sequence of maps

\[ \mathrm{bu} \xrightarrow{J} \mathrm{bgl}_1 (\mathbb{S}) \to \mathrm{bgl}_1 (\mathrm{tmf}_1 (n)) \to \mathrm{bgl}_1 (\mathrm{Tmf}_1 (n)), \]

and $\mathbb {E}_\infty$-complex orientations of $\mathrm {tmf}_1$ and $\mathrm {Tmf}_1 (n)$ correspond to nullhomotopies of the respective composites. Since $\mathrm {bu}$ is $2$-connective, we may as well replace all occurrences of $\mathrm {bgl}_1$ with $\mathrm {bsl}_1 := \tau _{\geq 2} \mathrm {bgl}_1$.

Now, it follows from the definition of $\mathrm {tmf}_1 (n)$ (see [Reference MeierMei21]) that the map $\mathrm {tmf}_1 (n) \to \tau _{\geq 0} \mathrm {Tmf}_1 (n)$ fits into the following pullback square.

As a consequence, $\mathrm {tmf}_1 (n) \to \mathrm {Tmf}_1 (n)$ induces an isomorphism on $\pi _k$ for $k \geq 2$. Since $\pi _1 \mathrm {tmf}_1 (n) = 0$, it follows that the composite

\[ \mathrm{bsl}_1 (\mathrm{tmf}_1 (n)) \to \mathrm{bsl}_1 (\mathrm{Tmf}_1 (n)) \to \tau_{\geq 3} \mathrm{bsl}_1 (\mathrm{Tmf}_1 (n)) \]

is an equivalence, so that there is a splitting

\[ \mathrm{bsl}_1 (\mathrm{Tmf}_1 (n)) \simeq \mathrm{bsl}_1 (\mathrm{tmf}_1 (n)) \oplus \Sigma^2 \pi_1 \mathrm{Tmf}_1 (n), \]

from which the lemma follows.

Proof of Theorem 1.7 Combine Theorem 4.24 with Lemmas 4.25 and 4.26.

Acknowledgements

The author would like to thank Robert Burklund and Jeremy Hahn for useful comments on a draft.

Conflicts of Interest

None.

Appendix A. $K(n)$-localizations of Landweber exact ring spectra

In this appendix, we prove the following lemma, which the author was unable to find a reference for in the literature.

Lemma A.1 Let $E$ denote a Landweber exact ring spectrum whose homotopy is concentrated in even degrees. Then $L_{K(n)} E$ is also a Landweber exact ring spectrum whose homotopy is concentrated in even degrees.

Before we proceed to the proof of Lemma A.1, we need a lemma from commutative algebra.

Lemma A.2 Let $R_*$ denote a graded commutative ring and let $x_1, \ldots, x_n \in R_*$ denote a regular sequence of homogeneous elements. Then the sequence $x_1, \ldots, x_n$ remains regular in the completion $(R_*) ^{\wedge } _{(x_1, \ldots x_n)}$.

Proof. Let $I = (x_1, \ldots x_n)$. We claim that there are short exact sequences

(A.1)\begin{equation} 0 \to R_* / I^{k-1} \xrightarrow{x_1} R_* / I^{k} \to R_* / (I^k + (x_1)) \to 0 \end{equation}

for all $k \geq 1$. Supposing this for the moment, we find by taking the limit that

\[ 0 \to (R_*)^{\wedge} _I \xrightarrow{x_1} (R_*)^{\wedge} _I \to (R_* / x_1)^{\wedge} _I \to 0 \]

is short exact. In particular, we find that $x_1$ is a regular element in $(R_*)^{\wedge } _I$ and that $(R_*)^{\wedge } _I / x_1 \cong (R_* / x_1)^{\wedge } _I$. Taking $R_* / x_1$ as our new ring and the image of $x_2, \ldots, x_n$ in $R_* / x_1$ as our exact sequence, we may conclude by induction on the length of our regular sequence.

It remains to establish (A.1). It is sufficient to prove that (A.1) is exact on the associated graded of the $I$-adic filtration where $x_1$ is considered as a map of $I$-adic filtration degree $1$. Since $x_1, \ldots, x_n$ is regular, this associated graded may be identified with the sequence

which is easily verified to be short exact.

Proof Proof of Lemma A.1

By abuse of notation, we let $v_i \in \pi _* E$ inductively denote an arbitrary lift of the class $v_i \in \pi _* E / (p, \ldots, v_{i-1})$. Given a positive integer $k$, let $I_k = (p, v_1, \ldots, v_{k-1})$ and let $I_0 = (0)$. It follows from [Reference Hovey and StricklandHSt99, Proposition 7.10] that $\pi _* L_{K(n)} E \cong (v_n ^{-1} \pi _* E)^{\wedge } _{I_n}$. In particular, $\pi _{2*-1} L_{K(n)} E = 0$. It is clear that the operation of inverting $v_n ^{-1}$ preserves Landweber exactness, so that it suffices to prove that completion with respect to $I_n$ does as well. However, this follows immediately from Lemma A.2.

Footnotes

During the course of this work, the author was supported by NSF Grants DGE-1745302 and DMS-2103236.

1 This follows from [Reference QuillenQui71] as explained in [Reference LawsonLaw18, Proposition 4.1.2]. Alternatively, it follows directly from (the proof of) [Reference Hopkins and LawsonHL18, Theorem 32].

2 See [Reference RavenelRav86, Appendix A2.2] for the $v_i$.

3 See Recollection 4.14 for a refresher on the Lubin–Tate theories $E(k,\mathbb {G})$. The associated $2$-periodic Morava $K$-theories $K(k,\mathbb {G})$ are obtained by cofibering out by the sequence $(p,u_1,\ldots, u_{n-1})$, or equivalently by the sequence $(p,v_1, \ldots, v_{n-1})$.

4 Note that we cannot apply [Reference StricklandStr99, Lemma 8.25] directly to $B\Gamma _k$, since it only applies to spaces with bounded above $\mathbb {Q}$-cohomology.

5 While it is assumed in [Reference StricklandStr99] that $E$ is even-periodic, this is not used in the proof of [Reference StricklandStr99, Lemma 8.25]. We also note that $E$ may be viewed as a summand of an even-periodic ring spectrum $E[x_2 ^{\pm 1}]$, so that we may conclude from [Reference StricklandStr99, Lemma 8.25] as stated.

6 Recall that $L_2$-localization refers to Bousfield localization with respect to height two Johnson–Wilson theory $E(2)$, or equivalently with respect to a height two Lubin–Tate theory: see [Reference RavenelRav84, Definition 5.1].

References

Absmeier, D., Strictly commutative complex orientations of $Tmf_1 (N)$, Preprint (2021), arXiv:2108.03079.Google Scholar
Adams, J. F., Stable homotopy and generalised homology, Chicago Lectures in Mathematics (University of Chicago Press, Chicago, IL, 1974).Google Scholar
Ando, M., Blumberg, A. J., Gepner, D., Hopkins, M. J. and Rezk, C., Units of ring spectra, orientations and Thom spectra via rigid infinite loop space theory, J. Topol. 7 (2014), 10771117.CrossRefGoogle Scholar
Ando, M., Hopkins, M. J. and Rezk, C., Multiplicative orientations of $KO$-theory and of the spectrum of topological modular forms, Preprint (2010), https://faculty.math.illinois.edu/~mando/papers/koandtmf.pdf.Google Scholar
Ando, M., Hopkins, M. J. and Strickland, N. P., Elliptic spectra, the Witten genus and the theorem of the cube, Invent. Math. 146 (2001), 595687.CrossRefGoogle Scholar
Ando, M., Hopkins, M. J. and Strickland, N. P., The sigma orientation is an $H_\infty$ map, Amer. J. Math. 126 (2004), 247334.CrossRefGoogle Scholar
Arone, G. and Lesh, K., Filtered spectra arising from permutative categories, J. Reine Angew. Math. 604 (2007), 73136.Google Scholar
Atiyah, M. F., Bott, R. and Shapiro, A., Clifford modules, Topology 3 (1964), 338.CrossRefGoogle Scholar
Balderrama, W., Algebraic theories of power operations, Preprint (2021), arXiv:2108.06802v2.Google Scholar
Barthel, T. and Heard, D., The $E_2$-term of the $K(n)$-local $E_n$-Adams spectral sequence, Topology Appl. 206 (2016), 190214.CrossRefGoogle Scholar
Behrens, M., Notes on the construction of tmf, in Topological modular forms, eds Douglas, C. L., Francis, J., Henriques, A. G. and Hill, M. A., Mathematical Surveys and Monographs, vol. 201 (American Mathematical Society, Providence, RI, 2014), 131188.Google Scholar
Berwick-Evans, D., Supersymmetric field theories and the elliptic index theorem with complex coefficients, Geom. Topol. 25 (2021), 22872384.CrossRefGoogle Scholar
Bousfield, A. K., The localization of spectra with respect to homology, Topology 18 (1979), 257281.CrossRefGoogle Scholar
Bousfield, A. K., Uniqueness of infinite deloopings for $K$-theoretic spaces, Pacific J. Math. 129 (1987), 131.CrossRefGoogle Scholar
Burklund, R., Schlank, T. M. and Yuan, A., The chromatic Nullstellensatz, Preprint (2022), arXiv:2207.09929v1.Google Scholar
Costello, K., A geometric construction of the Witten genus, I, in Proceedings of the International Congress of Mathematicians. Volume II (Hindustan Book Agency, New Delhi, 2010), 942959.Google Scholar
Costello, K., A geometric construction of the Witten genus, II, Preprint (2011), arXiv:1112.0816.Google Scholar
Deligne, P. and Rapoport, M., Les schémas de modules de courbes elliptiques, in Modular functions of one variable, II (Proc. Internat. Summer School, Univ. Antwerp, Antwerp, 1972), Lecture Notes in Mathematics, vol. 349 (Springer, Berlin, Heidelberg, 1973), 143316.Google Scholar
Douglas, C. L. and Henriques, A. G., Topological modular forms and conformal nets, in Mathematical foundations of quantum field theory and perturbative string theory, Proceedings of Symposia in Pure Mathematics, vol. 83 (American Mathematical Society, Providence, RI, 2011), 341354.CrossRefGoogle Scholar
Douglas, L., Francis, J., Henriques, A. G. and Hill, M. A. (eds), Topological modular forms, Mathematical Surveys and Monographs, vol. 201 (American Mathematical Society, Providence, RI, 2014).Google Scholar
Gille, P. and Szamuely, T., Central simple algebras and Galois cohomology, Cambridge Studies in Advanced Mathematics, vol. 101 (Cambridge University Press, Cambridge, 2006).CrossRefGoogle Scholar
Goerss, P. G. and Hopkins, M. J., Moduli spaces of commutative ring spectra, in Structured ring spectra, London Mathematical Society Lecture Note Series, vol. 315 (Cambridge University Press, Cambridge, 2004), 151200.CrossRefGoogle Scholar
Hill, M. and Lawson, T., Topological modular forms with level structure, Invent. Math. 203 (2016), 359416.CrossRefGoogle Scholar
Hirzebruch, F., Elliptic genera of level N for complex manifolds, in Differential geometrical methods in theoretical physics (Como, 1987), NATO Advanced Science Institutes Series C: Mathematical and Physical Sciences, vol. 250 (Kluwer Academic Publishers, Dordrecht, 1988), 3763.CrossRefGoogle Scholar
Hopkins, M. J., Kuhn, N. J. and Ravenel, D. C., Generalized group characters and complex oriented cohomology theories, J. Amer. Math. Soc. 13 (2000), 553594.CrossRefGoogle Scholar
Hopkins, M. J. and Lawson, T., Strictly commutative complex orientation theory, Math. Z. 290 (2018), 83101.CrossRefGoogle Scholar
Hovey, M. and Sadofsky, H., Invertible spectra in the $E(n)$-local stable homotopy category, J. Lond. Math. Soc. (2) 60 (1999), 284302.CrossRefGoogle Scholar
Hovey, M. and Strickland, N. P., Morava $K$-theories and localisation, Mem. Amer. Math. Soc. 139 (1999).Google Scholar
Joachim, M., Higher coherences for equivariant K-theory, in Structured ring spectra, London Mathematical Society Lecture Note Series, vol. 315 (Cambridge University Press, Cambridge, 2004), 87114.CrossRefGoogle Scholar
Katz, N., Serre-Tate local moduli, in Algebraic surfaces (Orsay, 1976–78), Lecture Notes in Mathematics, vol. 868 (Springer, Berlin-New York, 1981), 138202.Google Scholar
Kuhn, N. J., Morava K-theories and infinite loop spaces, in Algebraic topology (Arcata, CA, 1986), Lecture Notes in Mathematics, vol. 1370 (Springer, Berlin, 1989), 243257.CrossRefGoogle Scholar
Lawson, T., Secondary power operations and the Brown-Peterson spectrum at the prime 2, Ann. of Math. (2) 188 (2018), 513576.CrossRefGoogle Scholar
Lurie, J., Elliptic cohomology I, Preprint (2018), available online at http://www.math.ias.edu/~lurie/.Google Scholar
Lurie, J., Elliptic cohomology II: orientations, Preprint (2018), available online at http://www.math.ias.edu/~lurie/.Google Scholar
Meier, L., Connective models for topological modular forms of level $n$, Preprint (2021), arXiv:2104.12649v1.Google Scholar
Meier, L., Additive decompositions for rings of modular forms, Doc. Math. 27 (2022), 427488.CrossRefGoogle Scholar
Meier, L., Topological modular forms with level structure: decompositions and duality, Trans. Amer. Math. Soc. 375 (2022), 13051355.Google Scholar
Oliver, B., $p$-stubborn subgroups of classical compact Lie groups, J. Pure Appl. Algebra 92 (1994), 5578.CrossRefGoogle Scholar
Quillen, D., Elementary proofs of some results of cobordism theory using Steenrod operations, Adv. Math. 7 (1971), 2956.CrossRefGoogle Scholar
Ravenel, D. C., Localization with respect to certain periodic homology theories, Amer. J. Math. 106 (1984), 351414.CrossRefGoogle Scholar
Ravenel, D. C., Complex cobordism and stable homotopy groups of spheres, Pure and Applied Mathematics, vol. 121 (Academic Press, Orlando, FL, 1986).Google Scholar
Ravenel, D. C., Wilson, W. S. and Yagita, N., Brown-Peterson cohomology from Morava $K$-theory, K-Theory 15 (1998), 147199.CrossRefGoogle Scholar
Rezk, C., Power operations in Morava $E$-theory: structure and calculations (draft), Preprint (2013), available online at https://faculty.math.illinois.edu/~rezk/power-ops-ht-2.pdf.Google Scholar
Schuster, B. and Yagita, N., Morava $K$-theory of extraspecial 2-groups, Proc. Amer. Math. Soc. 132 (2004), 12291239.CrossRefGoogle Scholar
The Stacks Project Authors, Stacks project (2023), https://stacks.math.columbia.edu.Google Scholar
Stolz, S. and Teichner, P., What is an elliptic object?, in Topology, geometry and quantum field theory, London Mathematical Society Lecture Note Series, vol. 308 (Cambridge University Press, Cambridge, 2004), 247343.CrossRefGoogle Scholar
Stolz, S. and Teichner, P., Supersymmetric field theories and generalized cohomology, in Mathematical foundations of quantum field theory and perturbative string theory, Proceedings of Symposia in Pure Mathematics, vol. 83 (American Mathematical Society, Providence, RI, 2011), 279340.CrossRefGoogle Scholar
Strickland, N. P., Formal schemes and formal groups, in Homotopy invariant algebraic structures (Baltimore, MD, 1998), Contemporary Mathematics, vol. 239 (American Mathematical Society, Providence, RI, 1999), 263352.CrossRefGoogle Scholar
Tezuka, M. and Yagita, N., Cohomology of finite groups and Brown-Peterson cohomology, in Algebraic topology (Arcata, CA, 1986), Lecture Notes in Mathematics, vol. 1370 (Springer, Berlin, 1989), 396408.CrossRefGoogle Scholar
Wilson, D., Orientations and topological modular forms with level structure, Preprint (2015), arXiv:1507.05116v1.Google Scholar
Witten, E., Elliptic genera and quantum field theory, Comm. Math. Phys. 109 (1987), 525536.CrossRefGoogle Scholar
Witten, E., The index of the Dirac operator in loop space, in Elliptic curves and modular forms in algebraic topology (Princeton, NJ, 1986), Lecture Notes in Mathematics, vol. 1326 (Springer, Berlin, 1988), 161181.CrossRefGoogle Scholar
Yagita, N., Integral and BP cohomologies of extraspecial $p$-groups for odd primes, Kodai Math. J. 28 (2005), 130.CrossRefGoogle Scholar
Zhu, Y., Norm coherence for descent of level structures on formal deformations, J. Pure Appl. Algebra 224 (2020), 106382.CrossRefGoogle Scholar