<<

C2-EQUIVARIANT TOPOLOGICAL MODULAR FORMS

DEXTER CHUA

Abstract. We compute the groups of the C2 fixed points of equi- variant topological modular forms at the prime 2 using the descent spectral sequence. We then show that as a TMF-module, it is isomorphic to the tensor product of TMF with an explicit finite cell complex.

Contents 1. Introduction 1 2. Equivariant elliptic 5 3. The E2 page of the DSS 9 4. Differentials in the DSS 13 5. Identification of the last factor 25 6. Further questions 29 Appendix A. Connective C2-equivariant tmf 30 Appendix B. Sage script 35 References 38

1. Introduction Topological K-theory is one of the first examples of generalized cohomology theories. It admits a natural equivariant analogue — for a G-space X, the group 0 KOG(X) is the Grothendieck group of G-equivariant vector bundles over X. In 0 particular, KOG(∗) = Rep(G) is the representation ring of G. As in the case of non-equivariant K-theory, this extends to a G-equivariant cohomology theory KOG, and is represented by a genuine G-spectrum. We shall call this G-spectrum KO, omitting the subscript, as we prefer to think of this as a global equivariant spectrum — one defined for all compact Lie groups. The G-fixed points of this, written KOBG, is a spectrum analogue of the representation ring, BG 0 BG −n with π0KO = KOG(∗) = Rep(G) (more generally, πnKO = KOG (∗)). These fixed point spectra are readily computable as KO-modules. For example, KOBC2 = KO ∨ KO, KOBC3 = KO ∨ KU. arXiv:2101.06841v2 [math.AT] 1 Apr 2021

This corresponds to the fact that C2 has two real characters, while C3 has a real character plus a complex conjugate pair. If one insists, one can write KU = KO∧Cη, providing an arguably more explicit description of KOBC3 as a KO-module. In general, KOBG decomposes as a direct sum of copies of KO, KU and KSp, with the factors determined by the representation theory of G [Seg68, p.133–134]. From the chromatic point of view, the natural object to study after K-theory is elliptic cohomology, or its universal version, topological modular forms. Equivariant elliptic cohomology, in various incarnations, has been of interest to many people, include geometric representation theorist and quantum field theorists. Most recently, in [GM20], Gepner and Meier constructed integral equivariant elliptic cohomology and topological modular forms for compact abelian Lie groups, following the outline 1 2 DEXTER CHUA in [Lur09] and the groundwork in [Lur18a; Lur18b; Lur19]. The introduction in [GM20] provides a nice overview of the relevant history, whose efforts we shall not attempt to reproduce. The spectra TMFBCn can be constructed as follows: in [Lur18b], Lurie constructed the universal (derived) oriented1 elliptic curve, which we shall denote p: E → M. Equivariant TMF is then constructed with the property that 1 BCn BS TMF = Γ(E[n]; OE[n]), TMF = Γ(E; OE ), where E[n] is the n-torsion points of the elliptic curve. This is to be compared to the homotopy fixed points (with trivial group action), where E is replaced by the formal group Eˆ. We are interested in explicit descriptions of these spectra as TMF-modules. Much work was done by Gepner–Meier themselves: in [GM20, Theorem 1.1], they computed 1 TMFBS = TMF ⊕ ΣTMF. This corresponds to the fact that the coherent cohomology of a (classical) elliptic curve is concentrated in degrees 0 and 1 by Serre duality. As for finite groups, [GM20, Example 9.4] argues that if ` - |G| or ` > 3, then BG TMF` splits as sums of shifts of TMF1(3), TMF1(2) and TMF. Further, TMF1(3) and TMF1(2) can themselves be described as the smash product of TMF with an 8- and 3-cell complex respectively (see [Mat16, 4] for details). Thus, we have BG an explicit description of TMF` as a TMF`-module. This leaves us with the case where ` = 2, 3 and ` | |G|. In this paper, we compute TMFBC2 at the prime 2. Theorem 1.1. There is a (non-canonical) isomorphism of 2-completed TMF- modules TMFBC2 =∼ TMF ⊕ TMF ⊕ TMF ∧ DL, −8 −4 −2 −1 where DL is the spectrum S ∪ν S ∪η S ∪2 S , as depicted in Figure 1. This space DL is so named because its dual L is a split summand of the spectrum 0 L0 defined in [BR19, Definition 2.3]; in fact, L0 = L ⊕ S .

−1 2 −2 η −4

ν

−8

Figure 1. Cell diagram of DL

Remark. Despite being a 4-cell complex, the TMF-module TMF ∧ DL is really a rank-2 TMF-module. After base change to the flat cover TMF1(3) = BPh2i, all of 2, η, ν vanish, and the (−1)-cell is attached to the (−8)-cell via v2. Since v2 is invertible in TMF1(3), these two cells kill each other off, and we are left with two free cells in degrees −2 and −4.

1“oriented” refers to complex orientation of the associated cohomology theory C2-EQUIVARIANT TOPOLOGICAL MODULAR FORMS 3

Remark. While the theorem is stated for TMF, the same result holds for all elliptic cohomology theories. Indeed, by [MM15, Theorem 7.2], taking global sections gives an equivalence of ∞-categories ∼ Γ: QCoh(M) → ModTMF . Thus, as quasi-coherent sheaves, we have ∼ p∗OE[2] = OM ⊕ OM ⊕ OM ⊗ DL. By the universality of M, the same must hold for all other elliptic cohomology theories. 1.1. Outline of proof. To prove the theorem, we begin by computing the homotopy groups of TMFBC2 . As in the case of TMF, there is a descent spectral sequence BC2 computing π∗TMF whose E2 page is the coherent cohomology of the 2-torsion points of the (classical) universal elliptic curve. BC2 Upon computing the E2 page for TMF , one immediately observes that there are two copies of TMF’s E2 page as direct summands (as one would expect from the answer). We can identify these copies as follows:

(1) Applying Γ to the map p: E[2] → M induces 1: Γ(M; OM) → Γ(E[2]; OE[2]). This is split by the identity section. (2) Since TMF is a genuine C2-equivariant cohomology theory, we get a norm ∞ BC2 map TMFhC2 = TMF ∧ RP+ → TMF . Restricting to the bottom cell ∞ BC2 of RP+ gives us a map tr: TMF → TMF , which we call the transfer. We will explore these further in Section 2.2. To simplify the calculation, we can quotient out these factors, and rephrase our original theorem as Theorem 1.2. There is an isomorphism TMFBC2 ≡ TMFBC2 /(1, tr) ' TMF ∧ DL.

This is proven by computing the homotopy groups of TMFBC2 via its descent spectral sequence, which is now reasonably sparse, followed by an obstruction theory argument. This implies the original theorem via the observation Lemma 1.3. Any cofiber sequence of TMF-modules TMF ⊕ TMF → ? → TMF ∧ DL splits. Proof. We have to show that

[TMF ∧ DL, ΣTMF ⊕ ΣTMF]TMF = 0.

This is equivalent to showing that π−1TMF ∧ L = 0. This follows immediately by running the long exact sequences building TMF ∧ L from its cells, since π−2TMF = π−3TMF = π−5TMF = π−9TMF = 0.  Remark. At first I only computed the homotopy groups of TMFBC2 . The above identification was discovered when I, for somewhat independent reasons, looked into the homotopy groups of TMF ∧ L, and observed that they looked almost the same as that of TMFBC2 . It is, however, TMF ∧ DL that shows up above; there is a cofiber sequence TMF ∧ L → TMF ∧ DL → KO, which induces a short exact sequence in homotopy groups. Thus, the homotopy groups of TMF ∧ DL and TMF ∧ L differ by a single copy of π∗KO, which is hard to notice after inverting ∆. On the other hand, the Adams and Adams–Novikov filtrations of the classes differ, which makes them easy to distinguish in practice. 4 DEXTER CHUA

BC2 Remark. As part of the proof, we compute the homotopy groups π∗TMF . To describe the group explicitly, under the decomposition, it remains to specify BC2 π∗TMF ∧ DL = π∗TMF . This group is given by the direct sum of the ko-like parts, namely M 8+24k 16+24k π∗Σ ko ⊕ π∗Σ ko, k∈Z and what is depicted in Figures 11 and 13. In these figures, each dot is a copy of Z/2, and the greyed out classes are ones that do not survive the spectral sequence (that is, the homotopy groups are given by the black dots). This part is 192-periodic via ∆8-multiplication. 1.2. Overview. In Section 2 we provide relevant background on equivariant elliptic cohomology. Building upon the results in [GM20], we construct Cn-equivariant elliptic cohomology as a functor Spop → QCoh(E[n]), which gives us the transfer Cn map tr. We then provide an explicit description of the descent spectral sequence for quasi-coherent sheaves over M. In Section 3, we compute the Hopf algebroid presenting E[2] and subsequently

BC2 the E2 page of the descent spectral sequence for TMF using the 2-Bockstein spectral sequence. Unfortunately, the coaction involves division in a fairly complex ring, and cocycle manipulations throughout the paper are performed with the aid of sage. The relevant sage code is included in Appendix B. In Section 4, we compute the differentials in the descent spectral sequence. The BC2 key input here is the fact that there is a norm map TMFhC2 → TMF whose composite all the way down to TMFhC2 is well-understood in terms of stunted projective spaces. This provides us with a few permanent classes, which combined with the TMF-module structure lets us compute all the differentials. Our calculations will make heavy use of synthetic spectra [Pst18], whose relation to the Adams spectral sequence is laid out in [BHS19, Theorem 9.19]. In Section 5, we conclude the story by constructing a map TMF∧DL → TMFBC2 via obstruction theory and showing that it is an isomorphism. Finally, in Appendix A, we use entirely different methods to study properties of a hypothetical connective version of C2-equivariant TMF, which we call tmfC2 (we put the C2 subscript since we do not purport to describe a global equivariant tmf). −1 C2 C2 We shall show that under reasonable assumptions, ∆ (tmfC2 ) is dual to TMF in the category of TMF-modules.

1.3. Conventions. – All categories are ∞-categories. – Unless otherwise specified, we work in the category of TMF-modules, and all maps are TMF-module maps. Further, we implicitly complete at the prime 2. – Our charts follow the same conventions as, say, [Bau08]. In each bidegree, a solid round dot denotes a copy of Z/2. More generally, n concentric circles denotes a copy of Z/2n. A white square denotes Z. A line of slope 1 denotes 1 h1 multiplication and a line of slope 3 denotes h2 multiplication. An arrow with a negative slope denotes a differential. Dashed lines denote hidden extensions. In particular, a dashed vertical line is a hidden 2-extension. We use Adams grading, so that the horizontal axis is t − s and vertical axis is s. – All synthetic spectra will be based on BP . We choose our grading con- ventions so that π (νX/τ) = Exts,t (E ,E X), i.e. π shows up at t−s,s E∗E ∗ ∗ x,y coordinates (t − s, s) = (x, y) in an Adams chart. This is not the grading convention used by [Pst18] and [BHS19]. C2-EQUIVARIANT TOPOLOGICAL MODULAR FORMS 5

1.4. Acknowledgements. I would like to thank Robert Burklund for helpful discus- sions on various homotopy-theoretic calculations, especially regarding the application of synthetic spectra in Section 4. Further, I benefited from many helpful discussions with Sanath Devalapurkar, Jeremy Hahn, and Lennart Meier regarding equivariant TMF and equivariant in general. Robert and Lennart also pro- vided helpful comments on an earlier draft. Finally, the paper would not have been possible without the support of my advisor, Michael Hopkins, who suggested the problem and provided useful guidance and suggestions throughout. The author was partially supported by NSF grants DMS-1803766 and DMS- 1810917 through his advisor.

2. Equivariant elliptic cohomology 2.1. Elliptic cohomology. Let p: E → X be any (derived) oriented elliptic curve, and Cn be the cyclic group of order n. Lemma 2.1. There is an elliptic cohomology functor EllE : Spop → QCoh(E[n]) Cn Cn such that for any m | n, we have a natural identification EllE ((C /C ) ) = O . Cn n m + E[m] Moreover, if f : X0 → X is a morphism almost of finite presentation, then

∗ f ∗EllE (X) = Ellf E(X) ∈ QCoh(f ∗E[n]). Cn Cn If there is no risk of confusion, we omit the superscript E.

Proof. [GM20, Construction 5.4, Proposition 8.2] constructed the S1-equivariant version of this function — there is a symmetric monoidal functor op 1 EllS1 : SpS1 → QCoh(E), EllS1 ((S /Cm)+) = OE[m]. S1 The C case follows from this formally. Let Ind : Sp → Sp 1 be the induction n Cn Cn S 1 map, left adjoint to the restriction map. Then IndS ((C /C ) ) = (S1/C ) . Cn n m + m + 1 Since the restriction map is symmetric monoidal under the smash product, IndS Cn is oplax monoidal. Thus, the composite

1 IndS ∗ op Cn op EllS1 Ell : Sp Sp 1 QCoh(E) , Cn Cn S

0 is lax monoidal. Since S is a coalgebra in SpCn and every object in SpCn is naturally an S0-comodule, it follows that this functor canonically factors through the category of Ell∗ (S0) = O -modules in QCoh(E), which is equivalent to QCoh(E[n]).2 Cn E[n] Functoriality in X follows from functoriality in the S1 case as in [GM20, Propo- sition 5.6].  Remark. Unlike the case of S1, the map EllE : Spop → QCoh(E[n]) is in general Cn Cn not symmetric monoidal.

Corollary 2.2. There is a Cn-spectrum R such that for any Cn-spectrum X, we have X Cn (R ) = Γ(E[n], EllCn (X)).

2One has to check that the ring structure on O = Ell∗ (S0) that arises this way is the E[n] Cn standard ring structure, which follows from the construction of EllS1 . 6 DEXTER CHUA

We call this R the Cn-spectrum associated to the elliptic curve E → X. For example, when E → X is the universal elliptic curve, then R = TMF. This follows the argument of [GM20, Construction 8.3]. Proof. By the spectral Yoneda’s lemma [Lur12, Proposition 4.8.2.18], the Yoneda embedding Sp → FunR(Spop , Sp) is an equivalence. Thus, we have to show that Cn Cn the functor X 7→ Γ(E[n], Ell (X)) preserves limits as a functor Spop → Sp. Cn Cn op – By construction EllS1 : SpS1 → QCoh(E) preserves limits. S1 – Since Ind : Sp → Sp 1 is a left adjoint, it preserves colimits, hence its Cn Cn S op preserves limits. So Ell∗ : Spop → QCoh(E) preserves limits. Cn Cn – Since QCoh(E[n]) is the category of OE[n]-modules in QCoh(E), the for- getful functor QCoh(E[n]) → QCoh(E) creates limits. So Ell : Spop → Cn Cn QCoh(E[n]) preserves limits. – Finally, Γ: QCoh(E[n]) → Sp is a right adjoint and preserves limits.  We are interested in these global sections, which we can write as

Γ(E[n]; EllCn (X)) = Γ(X; p∗EllCn (X)). 0 By computing p∗EllCn (S ), this lets us understand the global sections in terms of quasi-coherent sheaves on X itself. This pushforward is fairly nice by virtue of Lemma 2.3. The map [n]: E → E is flat, hence so is p: E[n] → X. Proof. To check that [n]: E → E is flat, observe that the map on underlying ∗ (classical) stacks is flat. The condition that [n] πtOE = πtOE as sheaves on the ∗ underlying stack is automatic, since πtOE = p πtOX and p[n] = p. For the second part, we have a pullback square E[n] E

p [n] X E where the bottom map is the identity section, and flat morphisms are closed under pullbacks.  Corollary 2.4 ([GM20, Lemma 8.1]). The underlying stack of E[n] is the n-torsion points of the underlying stack of E. Proof. More generally, given a pullback of a flat morphism between non-connective spectral Deligne–Mumford stacks, it is also a pullback on the underlying classical stacks. To see this, since being flat and a pullback is local, we may assume that the stacks are in fact affine, in which case the result is clear.  2.2. The unit and transfer maps. There are two natural maps

1, tr: OX → p∗OE[n]. ∗ The map 1 is adjoint to the identity map p OX = OE[n] → OE[n], and is a map of OX -algebras. In particular, it is an OX -module homomorphism that sends 1 to 1. If X were affine, then this comes from taking the global sections of p: E[n] → X. This map is split by the identity section X → E[n]. The trace map tr comes from stable equivariant homotopy theory itself. To avoid double subscripts, set G = Cn. In the category of G-spectra, there are maps 0 G+ → S → G+ P ∞ whose composition is g∈G g. The first map comes from applying Σ+ to the map of unbased G-spaces G → ∗, whereas the second map is the Spanier–Whitehead dual P of the first map, using the self-duality of G+. Intuitively, it sends 1 7→ g∈G G. C2-EQUIVARIANT TOPOLOGICAL MODULAR FORMS 7

0 Now EllG(G+) = Ell{e}(S ) = OX , so we get maps of OE[n]-modules tr OX → OE[n] → OX whose composite is n (since G acts trivially on OX ). Applying p∗, we get a map tr: OX → p∗OE[n]. It will be useful to relate this to the norm map of Cn-spectra. Let R be the Cn-spectrum associated to E. Then unwrapping the definitions, we see that tr is the Cn-fixed points of the map 0 R ⊗ G+ −→ R ⊗ S , obtained by tensoring up the unique map G → ∗. Similarly, the norm map is induced by 0 R ⊗ EG+ −→ R ⊗ S , G using the Adams isomorphism (R ⊗ EG+) = RhG. Since G includes into EG, the trace map factors as Nm BG tr: R −→ RhG −→ R , where the left-hand map is the usual inclusion. Since G acts trivially on the underlying spectrum R, we have RhG = R ∧ BG+, and the left-hand map is the inclusion of the bottom cell of BG+. We now define p∗OE[n] by the following cofiber sequence in QCoh(X): 1⊕tr OX ⊕ OX −→ p∗OE[n] −→ p∗OE[n]. We then write BCn R = Γ(X; p∗OE[n]), BC R n = Γ(X; p∗OE[n]). In particular, when X = M and R = TMF, we have a cofiber sequence 1⊕tr TMF ⊕ TMF −→ TMFBCn −→ TMFBCn . In this paper, we are only interested in the case n = 2. 2.3. The descent spectral sequence. Our main computational tool is the descent spectral sequence, which we recall in this section. Let X be any non-connective spectral Deligne–Mumford stack and F a quasi- coherent sheaf on X. Let U → X be an étale cover of X. Then the sheaf condition tells us ∗ Γ(X; F) = Tot(Γ(U ×X · · · ×X U; π F)), where U ×X · · · ×X U is the Čech nerve of the cover, and π : U ×X · · · ×X U → X is the projection map. The descent spectral sequence is the Bousfield–Kan spectral sequence for the totalization, and the E2 page is given by the Čech cohomology s,t ˇ s E2 = H (Xcl; πtF) of the underlying classical stack Xcl with respect to the cover U. For us, we have X = M, and U = M1(3), the moduli stack of elliptic curves with a Γ1(3)-structure (i.e. a choice of 3-torsion point). By [MM15, Theorem 7.2], the map Γ: QCoh(M) → ModTMF is an equivalence of symmetric monoidal categories. Since i: M1(3) → M is affine, we know QCoh(M (3)) = Mod (QCoh(M)) 1 i∗OM1(3)

= ModTMF1(3)(ModTMF) = ModTMF1(3) . 8 DEXTER CHUA

Thus, we have ∗ Γ(M1(3) ×M × · · · ×M M1(3), π F) = TMF1(3) ∧TMF · · · ∧TMF Γ(M; F).

So the descent spectral sequence is also the TMF1(3)-based Adams spectral sequence in ModTMF. There is a well-known identification

Lemma 2.5. The TMF1(3)-based Adams spectral sequence in ModTMF is the same as the BP -based Adams–Novikov spectral sequence in spectra. We only use this result to apply the machinery of synthetic spectra to the descent spectral sequence; the morally correct approach would be to reproduce the theory of synthetic spectra inside ModTMF, but we’d rather not take that up.

Proof. Following [Mil81, Section 1], it suffices to show that any TMF1(3)-resolution of a TMF-module in ModTMF is also an BP -resolution in Sp. To do so, we have to show that every TMF1(3)-injective module in ModTMF is BP -injective in Sp, and every TMF1(3)-exact sequence in ModTMF is BP -exact in Sp.

(1) We have to show that TMF1(3) ⊗TMF X is BP -injective in Sp for any X ∈ ModTMF. Since TMF1(3) is complex orientable, there is a homotopy ring map MU → TMF1(3). Thus, TMF1(3) ⊗TMF X is a homotopy MU- module, hence MU-injective, hence BP -injective. (2) Since F (TMF, −) is right-adjoint to the forgetful functor ModTMF → Sp, by definition of exactness, it suffices to show that if X is BP -injective, then F (TMF,X) is TMF1(3)-injective. Again we may assume X = BP ∧ Y . Recall that TMF1(3) = TMF∧Z for some even spectrum Z. By evenness, BP is a retract of BP ∧ DZ. Thus, F (TMF,BP ∧ Y ) is a retract of F (TMF,BP ∧ DZ ∧ Y ) = F (TMF ∧ Z,BP ∧ Y ) = F (TMF1(3),BP ∧ Y ), which is a TMF1(3)-module, hence TMF1(3)-injective.  For convenience, set

A = π∗TMF1(3), Γ = π∗TMF1(3) ⊗TMF TMF1(3). Then (Γ,A) is a Hopf algebroid, and for any TMF-module N = Γ(M; F), we have s s ∗ ExtΓ(A, πt(TMF1(3) ⊗TMF N)) = ExtΓ(A, πt(i F)) ⇒ πt−sN. To perform calculations, it is of course necessary to identify (Γ,A) explicitly. From [MR08], we have −1 3 3 A ≡ π∗TMF1(3) = Z2[a1, a3, ∆ ], ∆ = a3(a1 − 27a3), |ai| = 2i, with associated elliptic curve 0 2 2 3 E : y z + a1xyz + a3yz = x . Spec Γ is the classifying scheme of two curves of the form E0 that are abstractly isomorphic, i.e. related by a coordinate transform. Consider the change of coordinates x 7→ x + r y 7→ y + sx + t In order to preserve the form of the equation, we need 2 0 = 3r − s − a1s 4 3 0 = s − 6st + a1s − 3a1t − 3a3s 6 2 5 2 2 4 2 3 3 0 = s − 27t + 3a1s − 9a1s t + 3a1s − 9a1st + a1s − 27a3t So we have Γ = A[s, t]/I, where I is the ideal generated by the relations above (we have eliminated r entirely). One checks that Γ is the free A-module on C2-EQUIVARIANT TOPOLOGICAL MODULAR FORMS 9

2 3 2 3 {1, s, s , s , t, st, s t, s t}, and these generators exhibits TMF1(3) ⊗TMF TMF1(3) as the sum of 8 suspended copies of TMF1(3). We can read off the structure maps of the Hopf algebroid to be

ηR(a1) = a1 + 2s

ηR(a3) = a3 + a1r + 2t ∆(s) = s ⊗ 1 + 1 ⊗ s ∆(r) = r ⊗ 1 + 1 ⊗ r ∆(t) = t ⊗ 1 + 1 ⊗ t + s ⊗ r. This Hopf algebroid (or rather, the connective version without inverting ∆) was studied in detail in [Bau08], whose computations and names we will use significantly.

3. The E2 page of the DSS 0 3.1. Computing the comodule. Let q : E → M1(3) be the canonical elliptic curve over M1(3), so that we have a pullback diagram

j E0 E q p

i M1(3) M. Then we have

BC2 ∗ BC2 TMF1(3) ⊗TMF TMF = Γ(i p∗OE[2]) = Γ(q∗OE0[2]) = TMF1(3) , and similarly with the bar version. BC2 In this section, we compute π∗TMF1(3) as a Γ-comodule, and then quotient BC2 out the image of 1 and tr. By Corollary 2.4, π∗TMF1(3) is given by (the global sections of) the classical scheme of 2-torsion points of E0. The naïve way to compute E0[2] is to write down the duplication formula for E0 and compute its kernel. However, the duplication formula is unwieldy. Instead, we write down the inversion map i: E0 → E0 and compute the equalizer with the identity map. The inversion map is induced by the map of projective spaces 2 2 P → P

[x : y : z] 7→ [x : −y − a1x − a3z : z] The equalizer of i with the identity is then cut out by the equations

x(2y + a1x + a3z) = 0

z(2y + a1x + a3z) = 0 Now observe that the 2-torsion points are contained in the affine chart y = 1. Indeed, if y = 0, then the equation defining E tells us x = 0. So the unique point on the curve when y = 0 is [0 : 0 : 1]. But this doesn’t satisfy the last equation above since a3 is invertible. Therefore, we work in the y = 1 chart. Following standard conventions, we redefine x z z = − , w = − . y y Then the 2-torsion points are cut out by the equations 3 2 z − w + a1zw + a3w = 0 2 2z − a1z − a3zw = 0 2 2w − a1zw − a3w = 0. 10 DEXTER CHUA

Adding the first and last equation gives z3 + w = 0. Eliminating w, we find that 0 2 4 E [2] = Spec A[z]/(2z − a1z + a3z ). In other words,

BC2 2 4 π∗TMF1(3) = A[z]/(2z − a1z + a3z ), |z| = −2.

Since a3 is invertible, this is a free A-module of rank 4. BC2 The Γ-coaction on π∗TMF1(3) comes directly from the construction of Γ itself; it is given by z − rz3 z 7→ . 1 − sz + tz3

BC2 3 Theorem 3.1. The map tr: TMF1(3) → TMF1(3) sends 1 to 2 − a1z + a3z . 3 In particular, by naturality, 2 − a1z + a3z is a permanent cocycle. The argument is similar to [Die72, Satz 4] (see also [HKR00, Remark 6.15]).

BC2 Proof. This map is a map of TMF1(3) -modules. Since z acts trivially on π∗TMF1(3), this means z tr 1 = 0. Since A[z] is a UFD, we know that tr 1 must be 3 a multiple of 2 − a1z + a3z . Moreover, since it is equal to 2 after modding out by z, the multiple must be 1.  BC 2 So after taking the cofiber by 1 and tr, we get π∗TMF1(3) 2 = A{z, z }, BC2 and the E2 page of the descent spectral sequence for π∗TMF is given by 2 ExtΓ(A, A{z, z }). 3.2. Computing the cohomology mod 2. While the coaction itself is fairly complicated, there is a major simplification after we reduce mod 2. By computer calculation (Appendix B), we find that Lemma 3.2. Let 2 2 b1 = a3z , b5 = a3z; |bi| = 2i. 2 Then A{z, z } = A{b1, b5}, and there is a short exact sequence of comodules

0 → A{b1}/2 → A{b1, b5}/2 → A{b5}/2 → 0, where both ends are cofree on the indicated generator, inducing a long exact sequence in Ext. 2 More precisely, the class b1 ∈ A{z, z }/2 is invariant, while 2 ψ(b5) − [1]b5 = [r ]b1. 2 Thus, the connecting map of the long exact sequence in Ext is given by b5 7→ [r ]b1. For reference, we display the cohomology of A/2 in Figure 2, as computed by [Bau08]. This chart is read as follows:

– Each dot represents a copy of F2; h1-multiplication and h3-multiplication are denoted by lines of slope 1 and 1/3 respectively. – [r2] represents the class x in bidegree (7, 1). This class is uniquely charac- 2 terized by the fact that a1 kills this class, coming from the cobar differential 2 2 2 d(a3) = [a1r ]. 4 4 −1 – The long dotted lines denote the extension h1 = a1∆ g. – The classes fading out continue a1-periodically, and each “period” consists of an infinite h1 tower. C2-EQUIVARIANT TOPOLOGICAL MODULAR FORMS 11

In Figure 3, we put two copies of this next to each other and draw the connecting differential. The resulting√ cohomology is in Figure 4. The hidden extensions follow from a “multiplication by ∆” operator, which we shall next explain.

7 6 5 4 g 3 x3 2 x2 1 x a1x 0 a1 ∆

0 2 4 6 8 10 12 14 16 18 20 22 24 26 28 30

Figure 2. Cohomology of A/2

3 2 1 x1,1 0 b1 b5

0 2 4 6 8 10 12 14 16 18 20 22 24 26 28 30

Figure 3. Connecting maps for ExtΓ(A, A{b1, b5}/2)

7 6 5 4 3 2 √ 1 x1,1 ∆x1,1 0 √ −1 b1 a1b1 ∆b1

0 2 4 6 8 10 12 14 16 18 20 22 24 26 28 30

Figure 4. ExtΓ(A, A{b1, b5}/2)

2 4 2 Originally, we have an action of A[z]/(2z − a1z + a3z ) on A{z, z }/2. Since 3 we have quotiented out by 2 and 2 − a1z + a3z , this reduces to an action by 3 A[z]/(2, a1z + a3z ). Further, since we are acting on z-multiples only, this reduces 2 to an action by A[z]/(2, a1 + a3z ). In this ring, we have 4 3 3 4 4 2 2 2 2 ∆ = a3 + a3a1 = a3 + a3a1z = (a3(1 + a1z)) . 12 DEXTER CHUA √ 2 2 One can check via sage that ∆ = a3(1 + a1z) is invariant in A[z]/(2, z + a3z ), 2 so acts on ExtΓ(A, A{z, z }√/2) (see again Appendix B). For example, the surviving 3 2 2 2 class in bidegree (14, 0) is ∆b1 = a3z − a3a1z. 3.3. 2-Bockstein spectral sequence. We now run the 2-Bockstein spectral se- quence, which we will find to degenerate on the E2 page. These Bockstein d1’s resemble the d3’s in the descent spectral sequence quite a bit. Thus, despite the fact that a lot of the differentials can be computed by writing down explicit cocycles, we try our best to argue them formally so that the same argument can be applied to the d3’s. Looking at the chart in Figure 4, it is not hard to see what to expect. All differentials have bidegree (−1, 1), and we know that nothing above the zero line survives, since (Γ,A) has no rational cohomology. Thus, for example, the class in bidegree (1, 1) must be hit by a differential from b1. The main work to do is to make sure nothing exotic happens with the highly a1-divisible classes coming from ∆ division. To begin, recall that in the 2-Bockstein for ExtΓ(A, A), we have

d1(a1) = h1. 2 Lemma 3.3. There are no non-zero classes of the form h1a on the E2 page. 2 Proof. If d1(h1a) 6= 0, then it doesn’t survive. Otherwise, consider d1(a). This must 2 2 be h1 torsion, so it is an h2 multiple. Then h1d1(a) = 0. So d1(h1a1a) = h1a.  The proof of the lemma is much more powerful than the conclusion itself. It tells 2 you about what sort of classes can kill h1x, and in certain bidegrees, it suffices to be h1 divisible. In general, we let xt−s,s denote a class in the corresponding bidegree that generates the bidegree after modding out by a1- and h1-multiples, if this makes sense. This class is well-defined up to a1- and h1-multiples. √ √ Lemma 3.4. d1(b1) = x1,1 and d1( ∆b1) = ∆x1,1.

Note that since x1,1 is only√ well-defined up to a1 multiples, this is equivalent to saying that√d1(b1) and d1( ∆)b1 are not a1 divisible. Since x1,1 is not well-defined, neither is ∆x1,1, and we are not claiming that there is a single choice of x1,1 for which both equations hold.

Proof. First observe that there is a choice of x1,1 that is permanent. Indeed, for 2 any choice of x1,1, the class d1(x1,1) must be h1-divisible, so it must be hit by a d1 from an a1-multiple by Lemma 3.3, which we can add to x1,1, so that it survives the d1. From the E2 page onwards, the target bidegree of the differential is 0 by Lemma 3.3 again. 3 4 Now h1x1,1 must be hit by a d1, and the source can only be h1b1 + O(a1), or else the differential would only hit highly a1-divisible elements. So d1(b1) must hit a version of x1,√1. The case of ∆b1 is analogous.  √ Lemma 3.5. d1(a1b1) = d1(a1 ∆b1) = 0.

This implies a1d1(b1) = h1b1. √ ± Proof. Note√ that b1 and ∆b1 generate the 0-line under a1 and ∆ , and d1(a1b1) and d1(a1 ∆b1) must be in the submodule generated by these and h1. We first show that the values of the differentials must be a1-divisible. Indeed, we cannot 2 have d1(a1b1) = h1b1 + O(a1), because applying d1 again would imply that 2 0 = h1x1,1 + O(a1), C2-EQUIVARIANT TOPOLOGICAL MODULAR FORMS 13 √ a contradiction. The argument for d1(a√1 ∆b1) is similar. Set ∆ = 1, and let x = a1b1, y = a1 ∆b1. Then we can write

d1(x) = h1(fx + gy), d1(y) = h1(hx + ky), for some f, g, h, k odd polynomials in a1. Applying d1 again gives us equations f 2 + f 0 + gh = 0 k2 + k0 + gh = 0 gk + gf + g0 = 0 hf + hk + h0 = 0 Adding the first two equations together tells us (f + k)2 = (f + k)0. Since squaring increases degree while differentiating decreases it, any polynomial that squares to its derivative must vanish. Thus f + k = 0. Then the last two equations tell us g0 = h0 = 0. But g and h are odd. So they must be 0. Then the first two terms tell 2 0 2 0 us f = f and k = k , so they must both be zero too.  √ Remark. We can in fact write down explicit lifts of a1b1 and a1 ∆b1, namely √ 3 a1b1 + 2a3z, a1 ∆b1 + 2a3z, 2 whose coboundary vanishes mod 2 . However, the proofs above will be used for d3’s in the descent spectral sequence too, and we cannot write down explicit cocycles for that.

± With ∆ and g periodicity, this gives all d1’s. No classes are left in positive s so we are done. The resulting the E2 page of the descent spectral sequence of TMFBC2 has a fairly regular pattern, which we exhibit in Figure 5. The names are intentionally left off; they can be found in Figure 6.

7 6 5 4 3 2 1 0

0 2 4 6 8 10 12 14 16 18 20 22 24 26 28 30

Figure 5. E2 page of ANSS

4. Differentials in the DSS

BC2 We have now computed the E2 page of the descent spectral sequence of TMF . The goal of this section is to compute the differentials. The main difficulty in computing the descent spectral sequence differentials is translational invariance — the E2 page is ∆-invariant, but the E∞ page will only be ∆8-invariant. If we had a connective version, then the leftmost class must be permanent since there is nothing to hit. Since we do not, we need external means of determining that certain classes are permanent. Once we do so, we can use standard techniques in homotopy theory to compute the remaining differentials. 14 DEXTER CHUA

We begin by computing the d3’s, where most of the hard work lies in. We depict the end result in Figure 6 for reference.

7 6 5 gt 4 x12,4 gx4,0 gx8,0 2 3 h1t 2 √ √ h2 ∆t 1 x1,1 =t ∆t ∆t 0 x4,0 x8,0 x16,0

0 2 4 6 8 10 12 14 16 18 20 22 24 26 28 30

Figure 6. E3 page of the descent spectral sequence

To compute the d3’s, we have to√ show that x1,1 is permanent by explicitly constructing a homotopy class t, while ∆x1,1 supports a d3. The rest then follows√ 4 2 formally using η = 0. Along the way, we will find a hidden ν-extension from h2 ∆t to h1gt, which will be useful later on.

Theorem 4.1. There is a choice of x1,1 that survives and has order 2. We call this class t. In fact, all choices survive and have order 2, but we will only get to see this after computing the spectral sequence fully.

Proof. We define t to be the composition

∞ Nm BC2 BC2 t: ΣTMF ,→ TMF ∧ RP+ = TMFhC2 −→ TMF → TMF /(1, tr), where the first map is the inclusion of the 1-cell. We claim this this has ANSS 4 filtration 1 and is non-v1-divisible. Then it must be detected by a choice of x1,1. To do so, consider the composite

t t0 : ΣTMF → TMFBC2 /(1, tr) hC ∞ 2 → TMF 2 /(1, tr) = TMFRP+ /(1, tr) → TMFRP+ /(1, tr). It suffices to prove the same properties for t0. The key fact from equivariant homotopy theory we use is the following: let X be a spectrum with trivial C2 action. Then the cofiber of the composition Nm ∞ m n ∞ hC2 RP+ RP+ X ∧ RP+ → X ∧ RP+ = XhC2 → X = X → X n n is ΣX ∧ P−m−1, where P−m is the stunted projective space. n 0 1 Take n = 1, so that RP+ = S ∨ S . The cell diagram of ΣP−3,1 is given by

2 1 0 −1 −2 C2-EQUIVARIANT TOPOLOGICAL MODULAR FORMS 15 where as usual the attaching maps of degree 1, 2, 4 are 2, η, ν respectively. In this diagram, we think of each cell as a TMF-module cell, i.e. a copy of TMF. We can read off all the information we need from this diagram. We start with 2 TMFRP+ , which is the bottom three cells in the diagram. We first understand what happens when we mod out 1 and tr. Recall that 1 is the global sections of the projection map E[2] → M, which is split by the identity section M → E[2]. The global sections of the identity section is the inclusion of the fixed points TMFBC2 → TMFhC2 → TMF by construction of 2 equivariant TMF. That is, 1 is a section of the projection TMFRP+ → TMF onto the 0-cell. Thus, quotienting out by 1 kills off the 0-cell, and we are left with the bottom two cells. By construction, tr is the attaching map of the 1-cell. Thus, further quotienting 2 by tr adds the 1-cell, and TMFRP+ /(1, tr) is the question mark complex, i.e. the subcomplex consisting of the (−2)-, (−1)- and 1-cell. Finally, t0 is the attaching map of the 2-cell. It must factor through the bottom cell since π2TMF/η = 0, and the diagram tells us this map is ν on the bottom cell, as desired.  2 Corollary 4.2. There is a choice of x4,0 with d3(x4,0) = h1t. 4 4 For any choice of x8,0, d3(x8,0) = O(v1) (i.e. it is divisible by v1).

In fact, we will later see that d3(x8,0) = 0. Proof. We must have η4t = 0, which forces the first part. For the second, we only 2 4 have to be concerned by the possibility that d3(x8,0) = h2t + O(v1). However, this is prevented by g-multiplication, since gx8,0 is h1-divisible.  √ Our next goal is to show that ∆t is not permanent,√ and instead supports a d3. We can think of this as a d3 on the hypothetical ∆ (which, if existed, must support √ 2 a d3 since ∆ supports a non-2-divisible d5). The proof is somewhat roundabout. Since t is 2-torsion, we get a map ΣTMF/2 → TMFBC2 picking out t. The homotopy groups of tmf/2 up to the 20th stem are depicted in Figure 7. We name these classes as follows — if y ∈ π∗TMF is 2-torsion, we let y˜ ∈ π∗TMF/2 be the class that is y on the top cell. This is well-defined up to an element in the image of π∗TMF. In particular, we are interested in the following classes:

(1) κ ∈ π14TMF is well-defined, while κ¯ ∈ π20TMF is well-defined mod 2. ˜2 ∼ (2) ν ∈ π7TMF/2 = Z/2 is the unique non-zero element in this degree. (3) κ˜ ∈ π15TMF/2 is well-defined up to ηκ. Thus, νκ˜ is well-defined.

5 ηκ¯ 4 κ¯ 3 2 κ 1 ν˜2 κ˜ 0

0 2 4 6 8 10 12 14 16 18 20

Figure 7. E∞ page of the ANSS of TMF/2 16 DEXTER CHUA

Lemma 4.3. In π∗TMF/2, we have ηκ¯ = ν2κ˜ + κν˜2.

Proof. We start with the Adams spectral sequence for π∗tmf, which is depicted in Figure 8. This may be computed by the May spectral sequence or a computer. The only possible d2’s in this range are the ones we have drawn, and any of the differentials implies all others by the Leibniz rule. We can get these via the fact 4 that v1ν = 0, for example. From this, Moss’ convergence theorem [Mos70] tells us 2 ηκ¯ = hκ, 2, ν i ∈ π∗TMF with no indeterminacy. By definition, the right-hand side is given by the composite3

κ

2 ν2

Let ν2 : Σ6TMF/2 → TMF/2 be the map that first projects to the top TMF-cell, and then maps via ν˜2. Then ν2 −ν2 maps trivially to the top cell, so factors through the bottom cell. This is a valid choice of “ν2 on the bottom cell”. So ηκ¯ = (ν2 − ν2)˜κ = ν2κ˜ − ν2κ.˜ Finally, note that ν2κ˜ = ν˜2κ, and that everything is 2-torsion, so we can drop the signs. 

8 7 6 5 4 4 v1 κ κ¯ 3 c0 2 1 0

0 2 4 6 8 10 12 14 16 18 20

Figure 8. E2 page of Adams spectral sequence for tmf

3In this diagram, the spectrum in the middle is (a shift of) TMF/2 and the ones at the end are TMF. The map on the left is any map such that if you project onto the top cell of TMF/2, then the map is κ (“κ on the top cell”); the map on the right is any map such that the restriction to the bottom cell of TMF/2 is ν2 (“ν2 on the bottom cell”). The composite of any two such choices given an element in the Toda bracket, and vice versa. C2-EQUIVARIANT TOPOLOGICAL MODULAR FORMS 17 √ Corollary 4.4. h2 ∆t represents κt˜√, and in particular is permanent. Further, 2 there is a hidden ν extension from h2 ∆t to h1gt. Proof. The class t gives a map ΣTMF/2 → TMFBC2 . Thus, the previous lemma gives 2 ˜2 BC2 ν κt˜ = ηκt¯ + ν κt ∈ π∗TMF ˜2 BC2 We know that ν t ∈ π8TMF has very high ANSS filtration (at least 7) because there is nothing in lower degrees. So ν2κt˜ = ηκt¯ + higher filtration.

ηκt¯ is represented by h1gt, so κt˜ must√ be detected by a permanent class with filtration at most 4, which must be h2 ∆t (this can alternatively follow from calculating the products in Ext, but we spare ourselves the trouble). 

If z is a cocycle in the E2 page, we let [z] denote any element of π∗ that is represented by z. √ k 2 k Corollary 4.5. There√ is a hidden ν extension from √∆ h2 ∆t to ∆ h1gt for every k 2 k 2 k k. That is, if ∆ h2 ∆t is permanent, then ν[∆ h2 ∆t] is detected by ∆ h1gt. Proof. We work in BP -synthetic spectra, and identify TMFBC2 with its synthetic analogue νTMFBC2 . We can rephrase the previous result as ν2κt˜ = τ 2ηκt¯ . √ Suppose ∆kh2 ∆t is permanent. Let α ∈ π TMFBC2 be a class whose 2 √ 16+24k,2 BC2 k 2 BC2 3 image in TMF /τ is ∆ h2 ∆t. Consider its image in TMF /τ . Since ∆ 3 survives to the E5-page, we know that ∆ lifts to π24,0TMF√ /τ (uniquely, since there 4 is nothing else in the bidegree). Since κt˜ represents h2 ∆t, we can write

k BC2 3 ∼ α = ∆ νκt˜ ∈ π16+24k,2TMF /τ = Z/2. So we know that

k 2 2 k BC2 3 να = ∆ ν κt˜ = τ ∆ ηκt¯ ∈ π∗,∗TMF /τ . √ BC2 2 k 2 3 So in π∗,∗TMF , we know that να = τ [∆ h2 ∆t] + O(τ ).  √ Lemma 4.6. ∆t does not survive to the E4 page. √ 3 Note that if ∆t survived to the E∞ page, then ηκt¯ would be ν -divisible. 3 2 2 However, ν is η -divisible in π∗TMF/2, and there is no candidate for the η division of ηκt¯ . The proof runs this argument in synthetic spectra to get the stronger claim that it doesn’t survive to E4. Proof. We again work in synthetic spectra. √ BC2 3 Suppose ∆t survived√ to the E4 page. Then it lifts to a class in π13,1(TMF /τ ), which we shall call ∆t again. Then √ 3 2 BC2 3 ν ∆t = τ ηκt¯ 6= 0 ∈ π22,4TMF /τ . √ BC2 3 Now note that 2 ∆t = 0 ∈ π13,1TMF /τ , since it is true algebraically, and there are no τ multiples in the bidegree. So we get a map of synthetic spectra √ Σ13,1TMF/2 → TMFBC2 /τ 3 picking out ∆t. From Figure 2, we can read that ν3 = η2ν˜2 ∈ π TMF/2, noting that there 9,√3 √ are no τ-divisible classes in that bidegree. Thus, ν3 ∆t = η2ν˜2 ∆t. However, √ ˜2 BC2 3 2 4 ν ∆t ∈ π20,2TMF /τ = 0, which is a contradiction (τ h1 = 0, so the h1 towers cannot contribute).  4Since we do not know that Cτ n is a ring, we have to interpret multiplication as composition here. 18 DEXTER CHUA √ Corollary 4.7. d3( ∆t) = x12,4. √ 2 Proof. This is equivalent√ to saying that d3( ∆t) is not v1-divisible. If it were, then 2 Lemma 3.3 tells us d3( ∆t + O(v1)) = 0, which the previous argument forbids. 

Corollary 4.8. d3 vanishes on any class in bidegree (8k, 0). 2 Proof. Pick x4,0 and x16,0 to be such that they generate the 0-line under v1 and ± 2 4 ∆ . Define x8,0 = v1x4,0. Observe that d3(x8,0) and d3(x16,0) are both v1-divisible, so the argument of Lemma 3.5 shows that the d3’s must both in fact be 0. The result follows. 

This concludes the calculation of the E3 page. The E5 page (with the ko-like patterns omitted) is shown in Figure 9. The differentials come from applying the Leibniz rule with d5(∆) = h2g. We then have d7s in Figure 10 that are forced by the hidden ν extensions (it is easy to check that there cannot be differentials from the ko-like classes since the possible targets are 2 non-g-torsion). The E9 page is then depicted in Figure 9, which is still ∆ -invariant. We will show that the greyed out classes do not survive, while the black ones do. Afterwards, all the remaining differentials are long differentials that kill off high κ¯ powers. These are shown in Figure 12, and the E∞ page is shown in Figure 13. In the rest of the section, we shall show that the long differentials that occur are indeed what we indicated. We then conclude the calculation using that ∆±8 is permanent.

We start with the observation that Lemma 4.9. There are no classes above the s = 24-line that survive. 6 Proof. Any permanent class above the line is divisible by κ¯ = 0.  The hardest part is to show that ∆2t is permanent. The difficulty here is again translational invariance. Our starting piece of knowledge is that t is permanent, and we want to somehow deduce that ∆2t is permanent too. However, we must not allow ourselves to repeat the argument, using that ∆2t is permanent to deduce that ∆4t is, because it is not. The key property we can make use of is the fact that the class t extends to a map ∞ from TMF ∧ RP . Our job would be easy if ∆2 of the bottom cell is permanent ∞ in TMF ∧ RP , but that’s not true. However, we can get by with the following version: Lemma 4.10. The class t: ΣTMF → TMFBC2 extends to a map from TMF ∧ L. Recall that L is the dual of DL, as in the statement of the main theorem (Theorem 1.1). Its cell diagram is depicted in Figure 14. Proof. First of all, it extends to ΣTMF/2 since 2t = 0. The obstruction to extending to the 4-cell is hη, 2, ti. Since t comes from restricting the norm map TMFhC2 → BC2 4 BC2 BC2 TMF , we know it extends to a map RP → TMF . Let y ∈ π∗TMF be 4 the image of the 3-cell. Then the cell structure of RP (Figure 14) tells us hη, 2, ti = 2y. But all possible images of y are 2-torsion. So hη, 2, ti = 0. Finally, the obstruction to extending to all of L is hν, η, 2, ti, which is defined since hν, η, 2i = 0 with no indeterminacy. However, the only possible class is a ν-multiple, hence is in the indeterminacy. So 0 ∈ hν, η, 2, ti, and we can extend to L. 

Remark. A posteriori, we expect such a map to exist. We know that TMFBC2 = TMF ∧ DL, and this is the map TMF ∧ L → TMF ∧ DL whose cofiber is KO. C2-EQUIVARIANT TOPOLOGICAL MODULAR FORMS 19 t 4 96 ∆ 88 80 t 3 72 ∆ 64 56 t 2 48 ∆ 40 32 t 24 ∆ 16 8 t 0 8 0 16

Figure 9. E5 page of the descent spectral sequence 20 DEXTER CHUA t 4 96 ∆ 88 80 72 64 56 t 2 48 ∆ 40 32 24 16 8 t 0 8 0 16

Figure 10. E7 page of the descent spectral sequence C2-EQUIVARIANT TOPOLOGICAL MODULAR FORMS 21 t 2 ∆ t 4 10 96 t w ∆ 19 w t 2 ∆ 9 w t 18 w t 2 88 ∆ 8 w t 17 w t 2 ∆ 7 w t 80 16 w t 2 ∆ 6 w t 15 w t 2 ∆ 72 5 w t 14 w t 2 ∆ 4 w t 13 w 64 t 2 ∆ 3 w t 12 w 56 t 11 w t 10 w t 2 48 ∆ t 9 w t 8 40 w t 7 w 32 t 6 w t 5 w 24 t 4 w 16 t 3 w 8 t 0 8 0 16

Figure 11. E9 page of the descent spectral sequence 22 DEXTER CHUA t 4 t ∆ t 37 6 18 w t 184 w 2 ∆ 8 ∆ w t 4 27 t w ∆ t 36 6 17 w t w 2 ∆ 7 ∆ w t 4 26 t 176 w ∆ t 35 6 16 w t w 2 ∆ 6 ∆ w t 4 25 t w ∆ t 34 6 15 w t w 2 ∆ 5 168 ∆ w t 4 24 t w ∆ t 33 6 14 w t w 2 ∆ 4 ∆ w t 4 23 t w ∆ t 32 160 6 13 w t w 2 ∆ 3 ∆ w t 4 22 t w ∆ 31 12 w t w 2 ∆ t 152 4 21 t w ∆ 30 11 w t w 2 ∆ t 4 20 t w ∆ t 29 6 10 w t 144 w 2 ∆ ∆ t 4 19 t w ∆ 28 9 w w t 2 ∆ t 4 18 t 136 w ∆ 27 8 w w t 2 ∆ t 4 17 t w ∆ 26 7 w w t 2 128 ∆ t 4 16 t w ∆ 25 6 w w t 2 ∆ t 4 15 t w ∆ 24 5 120 w w t 2 ∆ t 4 14 t w ∆ 23 4 w w t 2 ∆ t 112 4 13 t w ∆ 22 3 w w t 2 ∆ 12 t w 21 w t 104 2 ∆ 11 t w 20 w t 2 ∆ t 4 10 96 t w ∆ 19 w t 2 ∆ 9 w t 18 w t 2 88 ∆ 8 w 8 0 40 32 24 16

Figure 12. Remaining long differentials C2-EQUIVARIANT TOPOLOGICAL MODULAR FORMS 23 184 176 168 160 t 2 ∆ 152 21 w t 2 ∆ 20 w 144 t 2 ∆ 18 136 w t 2 ∆ 17 w t 2 128 ∆ 16 w t 2 ∆ 15 w 120 t 2 ∆ 14 w t 2 ∆ 112 13 w t 2 ∆ 12 t w 21 w t 104 2 ∆ 11 t w 20 w t 2 ∆ 10 96 w t 2 ∆ 9 w t 18 w t 2 88 ∆ 8 w 8 0 40 32 24 16

Figure 13. The E∞ page of the descent spectral sequence 24 DEXTER CHUA

w8

ν

w4 η

w2 2 w1 4 L RP

4 Figure 14. Cell diagrams of L and RP

Let wk be the k-cell of L. Theorem 4.11. In the Adams–Novikov spectral sequence of TMF ∧ L, the class 2 ∆ w1 survives and has order 2. Proof. It suffices to prove this for tmf ∧ L instead. We do not know of a way to compute the E2 page of the Adams–Novikov spectral sequence for tmf ∧ L, as the attaching maps are filtration 0 but non-injective in homology, so there is no long exact sequence. To remedy this problem, we use a modified Adams spectral sequence via the technology of synthetic spectra. First observe that 2: tmf → tmf is in fact injective in BP -homology, since BP ∧ tmf is non-torsion (see e.g. [Mat16, Corollary 5.2]). So ν(tmf/2) = (νtmf)/2. We next construct synthetic νtmf-modules Q,e Le by the cofiber sequences

[ηw2] Σ3,0tmf ν(Σtmf/2) Qe

[νw4] Σ7,0tmf Qe Le

The universal property of the cofiber gives us natural comparison maps Qe → ν(tmf ∧ Q) and Le → ν(tmf ∧ L). For example, the second map is obtained from the first via

[νw4] Σ7,0tmf Q˜ Le

ν([νw ]) Σ7,0tmf = ν(Σ7tmf) 4 ν(tmf ∧ Q) ν(tmf ∧ L). Here the top row is a cofiber sequence in the category of synthetic spectra, and the bottom row is ν applied to a cofiber sequence in the category of spectra. The only thing to check is that the left-hand square commutes, which is true since every map Sk,0 → νZ is uniquely of the form νf; there are no τ-torsion classes in this bidegree since these would have to be hit by a differential from below the 0-line. 2 Given this, it suffices to show that ∆ w1 survives in Le. To understand Le, we start with Σtmf/2, whose ANSS was computed by [Bea+21] and is shown in Figure 15 (with ko-like terms omitted as usual). 3,0 The synthetic cofiber sequence Σ tmf → νΣtmf/2 → Qe tells us the E2 page of the ANSS for Q˜ sits in a long exact sequence between that of Σ3,0tmf and νΣtmf/2. This is displayed in Figure 16, where the blue differentials are the connecting map. The crucial claim in this diagram is that there is a hidden extension η[νx4] = xw1 on the E2 page. Then since Le is obtained by killing [νx4], we are done. C2-EQUIVARIANT TOPOLOGICAL MODULAR FORMS 25

2 To see this hidden extension, note that x ∈ π∗tmf/2 detects ν on the top cell. If we quotient out the bottom cell in Q˜, then we can write the class of interest as

2 η[νx4] = ηhν, η, w2i = hη, ν, ηiw2 = ν w2, as desired. 2 Finally, it is straightforward to check that there are no classes above ∆ w1, so it must have order 2. 

Corollary 4.12. The class ∆2t in the descent spectral sequence of TMFBC2 is permanent and has order 2.

Proof. We previously constructed a map TMF ∧ L → TMFBC2 where the bottom cell hits t. Applying ν to this, we get a map t: ν(TMF ∧ L) → νTMFBC2 where 2 the bottom cell hits τt. Now consider t(∆ w1). This is a permanent class, and since ∆2 ∈ TMF/τ 2, after modding out by τ 2, we know that it must hit τ∆2t. So 2 2 t(∆ w1) is detected by ∆ t.  For the rest of the section, let z = t or ∆2t. It remains to consider the “w-chains” starting from z. There is a partially defined multiplication-by-w operation on the 4 E9 page, where w increases stem by 5. To formally define this, define w to be equal to g, and manually define

3 5 6 2 w z =κz, ˜ w z = ∆h1z, w z = ∆h2z. Corollary 4.13. The w chain starting from z is permanent. Proof. The argument of Corollary 4.4 shows that w3z is permanent. Since w5 = 5 6 [h1∆] is permanent, we know that w z is also permanent. This leaves the w terms, before we can conclude by g = w4-periodicity. We observe that ν3z = 0, since z is in the image of tmf ∧ L and ν3z = 0 over there. So w5z detects hκ,¯ ν3, zi and is permanent. 

4 k k+19 Corollary 4.14. There is a differential d?(∆ w z) = w z for all k. Here the length of the differential depends on the value of k (mod 4), which can be read off the charts. The precise values are, however, unimportant.

Proof. This follows from κ¯6 = 0 and g-division, since these are the only classes that can hit them. 

5. Identification of the last factor To identify TMFBC2 =∼ TMF ∧ DL, we map DL in by obstruction theory, and show it is an isomorphism after base change to TMF1(3). To do so, we need to understand the TMF1(3)-homology of DL.

Lemma 5.1. We can choose classes y−8, y−4, y−2 ∈ π∗TMF1(3) ∧ DL such that y−k ∈ π−kTMF1(3) ∧ DL; y−8 is the bottom cell of DL; {y−4, y−2} generates π∗TMF1(3) ∧ DL as a free π∗TMF1(3)-module; and −1 2 y−8 = v2 (a1y−4 + 2y−2) + O(2 ), 2 d(y−4) ≡ ψ(y−4) − [1]y−4 = [r]y−8 + O(2 ).

The choice of y−4 and y−2 is pretty much arbitrary. Other choices will result in slightly different formulas. These are chosen to simplify the ensuing calculation. 26 DEXTER CHUA 1 50 xw 2 g 1 w 48 2 ∆ 46 44 42 40 38 36 34 32 30 28 1 26 gw 2 h 1 w 24 ∆ 22 20 18 16 14 12 10 1 8 xw 6 4 ] 2 w 1 h [ 2 1 w 0 1 9 8 7 6 5 4 3 2 1 0 11 10 −

Figure 15. Adams–Novikov spectral sequence for Cf2 C2-EQUIVARIANT TOPOLOGICAL MODULAR FORMS 27 1 50 xw 2 g 48 46 44 42 40 38 36 34 32 30 28 26 24 22 20 18 16 14 12 10 1 xw 8 4 w 2 h 6 4 4 w 2 1 w 0 1 9 8 7 6 5 4 3 2 1 0 11 10 −

Figure 16. Adams–Novikov spectral sequence for Qe 28 DEXTER CHUA

Proof. We carefully construct TMF ∧ DL in the category of TMF-modules. We start with the bottom cell and attach y−4 to kill [r]y−8. The class y−4 is only 2 well-defined up to multiples of y−8, which in this case is integral multiples of a1y−8. 2 The coboundary of a1y−8 is [12r]y−8, and so

ψ(y−4) = y−4 + [kr]y−8, where k ≡ 1 (mod 4). The next cell will kill of the cocycle 2 {h1y−4} = [s]y−4 − [k(a1r + t)]y−8−?([s]a1 − [12t])y−8. Here ? is either 1 or 0, noting that twice the class is a coboundary. There is exactly one 2 4 choice of ? for which this cocycle is permanent, since there is a d3([s]a1 − [12t]) = h1. On the other hand, we know that y−4 is not entirely well-defined, and we can absorb the term into y−4 (and redefine k). We choose y−4 so that ? = 1. We now set ψ(y−2) = y−2 + {h1y−4}. and then the class 3 2y−2 + a1y−4 − ka3y−8 − (a1 − 12a3)y−8. is a cocycle, which the top cell kills off. So we get the relation 3 2 2y−2 + a1y−4 − (a1 − 27a3)y−8 + O(2 ), as desired.  We now construct a map f : TMF ∧ DL → TMFBC2 . This is constructed via obstruction theory. The relevant homotopy groups are in the range [−8, 0], which are depicted in Figure 17. In this range, all the homotopy groups come from the ko-like patterns.

3 2 1 0

−12 −10 −8 −6 −4 −2 0 2 4 6

Figure 17. Homotopy groups of TMFBC2 with ko-like terms

In general, let z−k be the images of y−k under f (after base change to TMF1(3)). In constructing f, the first step is to pick the image of the bottom cell, i.e. the BC2 value of z−8. This lives in π−8(TMF ), which is the direct sum of infinitely many copies of Z. Choosing z−8 requires a bit of care, but once we have chosen it, we can always extend it to a full map f. Indeed, the obstructions are

νx−8,0, hη, ν, x−8,0i, h2, η, ν, x−8,0i, which all vanish because

BC2 BC2 BC2 π−5TMF = π−3TMF = π−2TMF = 0. While the extension always exists, it is not unique. Specifically, the extension to the second cell is not unique. This results in z−4 being well-defined up to a BC2 permanent class in π−4TMF , which is again infinitely many copies of Z. After extending to the second cell, there is a unique way to extend all the way to DL, BC2 BC2 because π−2TMF = π−1TMF = 0. C2-EQUIVARIANT TOPOLOGICAL MODULAR FORMS 29

4 We are now ready to choose z−8. We definitely don’t want this to be v1-divisible, 4 but this only defines z−8 up to v1-multiples. After some experimentation, we settle on the following choice: Lemma 5.2. There is a permanent cocycle of the form −1 2 −1 2 2 z−8 = v2 [a1(z − a3 a1z) + 2z] + O(2 )

This will be our choice of z−8. −1 This is a lift of ∆ a1x14,0. Proof. Computer calculation (see Appendix B) verifies that this is a cocycle mod 22. In the bidegree (−8, 0), there are no 2-Bockstein differentials and DSS differentials, n so any cocycle mod 2 would lift to a permanent cocycle. 

Given the relation between the y−k, we know there must be some g such that 2 −1 2 2 z−4 = z − a3 a1z + 2g + O(2 )

z−2 = z + a1g + O(2). The indeterminacy tells us 2g is well-defined up to a permanent class. Lemma 5.3. There is a choice of f such that g = 0 mod 2.

Proof. In TMF1(3) ∧ DL, we have a cobar differential 2 d(y−4) = [r]y−8 + O(2 ). So we find that 2 d(z−4) = [r]z−8 + O(2 ). On the other hand, by computer calculation, we find that (see Appendix B) 2 2 −1 2 d(z − a1a3 z) = [r]z−8 + O(2 ). Comparing the two, we must have d(2g) = O(22). So g is a cocycle mod 2. Since there are no 2-Bocksteins, g lifts to an actual cocycle g˜, and 2g˜ is permanent. So we can subtract 2g˜ from z−4 so that g is now 0 mod 2. 

Corollary 5.4. This choice of f is an equivalence, and hence TMFBC2 =∼ TMF ∧ DL.

Proof. It suffices to show that f is an equivalence after base change to TMF1(3). Moreover, since we 2-complete, we can further reduce mod 2. Both surjectivity and injectivity in π∗ are easy linear algebra. 

6. Further questions – Is there a “geometric” description of the piece TMF ∧ DL, similar to the representation-theoretic description of KOBG? Can such a description streamline the calculations in the paper? (e.g. avoid the need of computers and “explain” the formula for z−8) In particular, the equation −v3y−8 +a1y−4 +2y−2 in TMF1(3)∧DL looks 4 2 BC2 remarkably similar to the defining equation a3z −a1z +2z in TMF1(3) . – The initial calculation of the Hopf algebroid is made possible by the fact that the inversion map of a Weierstrass elliptic curve has a simple formula. Is there a similar trick available for C3? Or must we bite the bullet and compute [3] by hand? 30 DEXTER CHUA

Appendix A. Connective C2-equivariant tmf At the prime 2, one prized property of tmf is

H∗tmf = AA(2)F2, which lets us identify the E2-page of its Adams spectral sequence as ExtA(2)(F2, F2). It is natural to expect C2-equivariant tmf to have a similar properties, and there have been attempts to construct C2-equivariant tmf along these lines (e.g. [Ric17]). The goal of this section is not to construct C2-equivariant tmf, but to deduce properties of any such construction based on its homology.

Theorem A.1. Let tmfC2 ∈ SpC2 be a spectrum such that

C2 C (HF2)?tmfC2 = A A 2 (2)M2. Then C2 ∼ H∗(tmfC2 ) = H∗(tmf ⊕ tmf ⊕ tmf ⊗ L)

C2 as a Steenrod comodule. If (tmfC2 ) has a tmf-module structure, then we have an isomorphism of 2-completed tmf-modules.

−1 C2 C2 Note that it is L that appears here, not DL, so ∆ (tmfC2 ) and TMF are duals as TMF-modules (after completion).

Since all we know about this hypothetical tmfC2 is its homology, it will be cleaner C2 to work with a general C2-spectrum whose homology is A AC2 (n)M2 for some n, specializing to n = 2 only at the very end. Nevertheless, we shall shortly see that such a spectrum can only exist for n ≤ 2, as in the non-equivariant case.

Conventions. We work with RO(C2)-graded homotopy groups throughout. We 2 have an explicit identification RO(C2) = Z[σ]/(σ − 1), and we shall write p + qσ = (p + q, q) — the first degree is the total degree of the representation and the second degree is the number of σ’s, also called the weight. We use ? to denote an RO(C2)-grading and ∗ for an integer grading. We let ρ: S0 → Sσ be the natural inclusion of the representation spheres (more 0 commonly known as a). Then ρ ∈ π−1,−1(S ).

A.1. C2-equivariant homotopy theory. Let X be a C2 spectrum. The main C2 difficulty in analyzing H∗X is that taking categorical fixed points is not symmetric C2 C2 monoidal, so there is no a priori relation between HF2 ∧ X and (HF2 ∧ X) . C2 To understand X , we consider two other functors SpC2 → Sp that are symmetric monoidal, namely the geometric fixed points XΦC2 and the underlying spectrum ιX. The goal of this section is to understand the effects of these constructions on RO(C2)-graded homotopy groups. The homotopy groups of the categorical fixed points are easy to compute. Essen- tially by definition, we have C2 π∗X = π∗,0X. The underlying spectrum and geometric fixed points are obtained by performing certain constructions in SpC2 and then applying categorical fixed points. In the C2-equivariant world, the geometric fixed points admit a particularly −1 0 simple description, since ρ S is a model of EC˜ 2: Lemma A.2. The geometric fixed points is given by XΦC2 = (ρ−1X)C2 , and we can write its homotopy groups as

ΦC2 −1 π∗X = π∗,0(ρ X) = π?X/(ρ − 1), C2-EQUIVARIANT TOPOLOGICAL MODULAR FORMS 31 where in the latter formulation, an element of bidegree (s, w) in π?X gives an element ΦC2 of degree s − w in π∗X . The underlying spectrum is slightly more involved. It can be expressed as

C2 C2 ιX = ((C2)+ ⊗ X) = F ((C2)+,X) , since (C2)+ is self-dual. To compute its homotopy groups, we use the cofiber sequence 0 ρ σ (C2)+ → S → S . We can either apply F (−,X) or (−) ⊗ X to this, and get Lemma A.3. There are long exact sequences ρ res ρ · · · −→ πs+1,w+1X −→ πs,wX −→ πs,w(C2)+ ∧ X −→ πs,w+1X −→ · · · . and ρ tr ρ ··· −→ πs,w−1X −→ πs,w(C2)+ ∧ X −→ πs,wX −→ πs−1,w−1X −→ · · · .

Moreover, res ◦ tr is 1 + σ on π∗ιX = π∗,0(C2)+ ∧ X, where σ is the generator of C2.

Corollary A.4. Suppose ρ is injective on π∗,0X and π∗,1X. Then

π∗ιX = (π?X/ρ)∗,0 with σ = −1. Proof. We first apply the first sequence with w = 0. By assumption, the last ρ is injective, hence has trivial kernel. So res is surjective with kernel given by ρ-multiples. Now consider the second sequence with w = 0. Since the second ρ is injective, it has trivial kernel, so tr has zero image. Hence 1 + σ = res ◦ tr = 0. 

A.2. The C2-equivariant Steenrod algebra. We recall the computation of the C2-equivariant Steenrod algebra. Lemma A.5 ([HK01, Proposition 6.2]).  γ  ≡ π H = [τ, ρ] ⊕ | k, ` ≥ 1 M2 ? F2 F2 ρkτ ` with the obvious ring structure. The degrees of the classes are |τ| = (0, −1), |ρ| = (−1, −1), |γ| = (1, 0). The terms involving γ are known as the “negative cone”. They will not play a role in the story (though we will have to verify that they indeed do not play a role). Lemma A.6 ([HK01, Theorem 6.41]).

C2 2 A ≡ π?(HF2 ∧ HF2) = M2[τ0, τ1, . . . , ξ1, ξ2 ...]/(τi = ρτi+1 +τξ ¯ i+1), where τ¯ = τ + ρτ0 and n+1 n n+1 n |τn| = (2 − 1, 2 − 1), |ξn| = (2 − 2, 2 − 1). We have

ηR(ρ) = ρ, ηR(τ) =τ ¯ = τ + ρτ0 and k X 2i X 2i ∆ξk = ξk−i ⊗ ξi, ∆τk = ξk−i ⊗ τi + τk ⊗ 1. i=0 32 DEXTER CHUA

C2 One observes that Corollary A.4 applies to M2 and A , so we can read off the homotopy groups of the underlying spectra of HF2 and HF2 ∧ HF2. This tells us ∼ ιHF2 = HF2 and π∗ι(HF2 ⊗ HF2) = A, as expected. The identification of the latter with A is canonical for the following unimaginative reason:

i Lemma A.7. Let ζ1, ζ2,... ∈ A be such that |ζi| = 2 − 1 and

X 2i ∆ζk = ζk−i ⊗ ζi.

If ζ1 6= 0, then ζi = ξi.

Proof. We prove this by induction on i. It is clear for i = 1, since there is a unique element of degree 1. Then if it is true for k < i, then ζi − ξi is primitive. But the 2k only primitive elements of A are of the form ξ1 , which is not in the bidegree. So this is zero, and ζi = ξi. 

A.3. The homology of geometric fixed points. Since (−)ΦC2 is symmetric ΦC2 monoidal, in order to deduce H∗X from (HF2)?X, it suffices to understand ΦC2 HF2 .

ΦC2 Lemma A.8. We have HF2 = HF2[τ] as an E1-ring, where |τ| = 1.

ΦC2 Truncation gives us an E∞ map HF2 → HF2, which we can understand as quotienting out τ.

ΦC2 Proof. The inclusion of the categorical fixed points gives HF2 the structure of an E∞-HF2-algebra. Since HF2[τ] is the free E1-ring on one generator over HF2, it ΦC2 suffices to observe that π∗HF2 = F2[τ] by Lemma A.2. 

Corollary A.9. If X ∈ SpC2 , then

ΦC2 ∼ H∗X = ((HF2)?X)/(ρ − 1, τ). If X is a homotopy ring, then this is an isomorphism of rings.

We are, of course, not only interested in the homology groups as groups, but as ΦC2 ΦC2 Steenrod modules. This involves understanding the comparison HF2 ∧ HF2 → ΦC2 HF2 ∧ HF2 given by squaring the projection HF2 → HF2.

ΦC2 ΦC2 Corollary A.10. The map π∗(HF2 ∧ HF2 ) → π∗(HF2 ∧ HF2) sends ξi 7→ ξi and kills τ0, τ.

2 C2 Proof. After setting ρ = 1, we can express τi+1 in terms of τi and ξi+1, so A /(ρ−1) is generated polynomially over F2[τ] by τ0, ξ1, ξ2,... with no relations. We know that ηL(τ) and ηR(τ) get mapped to zero, and these are τ and τ + τ0 respectively. So τ and τ0 get killed. ΦC2 As a spectrum, HF2 splits as a sum of HF2’s, we know that the map is surjective, and thus ξ1 must hit ξ1 since it is the only element left in the degree. Finally, the images of the ξi satisfy the conditions of Lemma A.7, so must be equal to ξi. 

ΦC2 Corollary A.11. If X ∈ SpC2 , then the coaction on H∗X is given by

C2 (HF2)?X/(ρ − 1) → A ⊗M2 (HF2)?X/(ρ − 1, τ ⊗ 1, τ0 ⊗ 1). C2-EQUIVARIANT TOPOLOGICAL MODULAR FORMS 33

C2 A.4. Spectra whose homology is A AC2 (n)M2. We now introduce the stan- dard quotient coalgebras of AC2 : Definition A.12. We define n−i+1 C2 C2 2 A (n) = A /(ξi , τn+1, τn+2,...).

C2 If n < 0, we set A (n) = M2. Then we have n C2 ¯2 ¯2 ¯ ¯ A AC2 (n)M2 = M2[ξ1 ,..., ξn, ξn+1, ξn+2,..., τ¯n+1, τ¯n+2,...]. As usual, the bar denotes the Hopf algebroid antipode.

Definition A.13. We let Yn ∈ SpC2 be any spectrum such that

C2 (HF2)?Yn = A AC2 (n)M2.

If n ≥ m, a map f : Yn → Ym is admissible if it induces the natural injection in homology. It is helpful to note that

Lemma A.14. There is a unique admissible map Yn → HF2.

Since we understand the situation of HF2 completely, we get to learn about Yn through comparison.

Proof. Since H?Yn is free over M2, the universal coefficients theorem tells us such maps are classified by M2-module maps H?Yn → M2, which in turn is in bijection C2 C2 with A -comodule maps H?Yn → A . 

Lemma A.15. H∗ιYn = AA(n)F2 with trivial C2 action as a Steenrod comodule, and admissible maps induce the natural inclusion.

This implies that Yn cannot exist for n > 2.

Proof. Since ι is symmetric monoidal, we have H∗ιYn = π∗ι(HF2 ∧ Yn). γ Observe that in (HF2)?Yn, the ρ-torsion term of smallest weight is ρτ with weight 2. So ρ is injective on (HF2)∗,0Yn and (HF2)∗,1Yn, and Corollary A.4 applies. This tells us H∗ιYn = ((HF2)?Yn/ρ)∗,0. It is then easy to check that the ring map ¯k 2i−1−1 k AA(n)F2 → H∗ιYn sending ξi to (τ τ¯i−1) is an isomorphism. Since the ring is 2-torsion, we have σ = −1 = 1. The comodule structure follows from a similar calculation.  We can similarly compute Lemma A.16. n ΦC2 ¯2 ¯2 ¯ H∗Yn = F2[¯τn+1, ξ1 ,..., ξn, ξn+1,...] ¯ with the coactions of the ξi as usual and

ψ(¯τn+1) = 1 ⊗ τ¯n+1. ¯2n An admissible map Yn → Ym acts as the “identity” on the ξi ’s and sends τ¯n+1 to 2n−m τ¯m+1 .

C2 Our goal is to understand the homology of the categorical fixed points Yn . To do so, we make use of the commutative diagram

C2 ΦC2 (Yn)hC2 Yn Yn Σ(Yn)hC2

(Yn)hC2 ιYn Yfn Σ(Yn)hC2 , 34 DEXTER CHUA

where Yfn is defined to be the cofiber of (Yn)hC2 → ιYn. C2 ΦC2 To compute H∗Yn , we have to understand the effect of Yn → Σ(Yn)hC2 on ΦC2 homology. This factors through Yfn, and we shall see that Yn → Yfn is injective in homology while Yfn → Σ(Yn)hC2 is surjective. The second map is easy to understand. A standard result tells us that

∞ ∞ H∗Yfn = AA(n)H∗ΣRP−1,H∗Yfn = AA(n)H∗ΣRP , ∞ ∞ and the map H∗Yfn → H∗Σ(Yn)hC2 is induced by the projection ΣRP−1 → ΣRP killing the bottom cell. ΦC2 As for the first map, in the case Y−1 = HF2, the map Y−1 → Yg−1 is an ΦC2 equivalence, and in particular injective in homology. Since H∗Yn and H∗Yfn inject ΦC2 into the n = −1 counterparts, we know that H∗Yn → H∗Yfn is injective as well. It remains to compute the image. At this moment it is slightly more convenient to consider cohomology. We follow some constructions from [Lin+80]. ∞ ∞ ∗ ∞ For any N, we have a map ΣRP−N → ΣRP−1 which presents H ΣRP−1 as ∗ ∞ ∗ ∗ ∞ ∗ a submodule of H ΣRP−N . Let Mn,N ⊆ H ΣRP−N be the A(n) -submodule ∗ ∗ ∗ generated under A(n) by elements of negative degree, and let Ln = Mn,N ∩ ∗ ∞ ∗ H ΣRP−1 for sufficiently large N (where it no longer depends on N). Let Gn = ∗ ∞ ∗ H ΣRP−1/Ln. We then have a short exact sequence of A(n)-comodules. ∞ 0 → Gn → H∗ΣRP−1 → Ln → 0.

ΦC2 Lemma A.17. There are no non-zero A-comodule maps H∗Yn → AA(n)Ln. Proof. By adjunction, this is equivalent to showing that there are no non-zero ΦC2 ∗ A(n)-comodule maps H∗Yn → Ln, or equivalently, no A(n) -module maps ∗ ∗ ΦC2 Ln → H Yn . Now note that in Ln all elements have positive degree, and are generated as an A(n)∗-module by elements of degree less than 2n − 1, since A(n)∗ is generated by elements of degree at most 2n. On the other hand, the elements of lowest degree n ΦC2 n ¯2 ∗ in H∗Yn are in degree 0 and 2 (given by 1 and ξ1 ). Thus, any A(n) -module ∗ ΦC2 map Ln → H Yn must kill the generators, hence must be zero. 

ΦC2 Corollary A.18. The map H∗Yn → H∗ΣYfn factors through AA(n)Gn.

ΦC2 Corollary A.19. H∗Yn → AA(n)Gn is an isomorphism. Proof. We already know this is injective. By [Lin+80, Lemma 1.3] both sides are abstractly isomorphic as graded groups. Since they are finite dimensional in each degree, it must be an isomorphism. 

Since H∗Yfn → H∗Σ(Yn)hC2 is surjective in homology with kernel AA(n)F2, we learn that Corollary A.20. There is a short exact sequence of comodules

C2 0 → AA(n)Ln[−1] → H∗Yn → AA(n)F2 → 0.

These Ln can be explicitly computed.

Example A.21. L0[−1] = 0. L1[−1] = F2. L2[−1] = F2 ⊕ H∗L. Specializing to the n = 2 case, Lemma A.22. The short exact sequence splits non-canonically when n = 2. C2-EQUIVARIANT TOPOLOGICAL MODULAR FORMS 35

Proof. It is classified by an element in

1,0 ExtA (AA(2)F2, AA(2)L2[−1]) = ExtA(2)(AA(2)F2,L2[−1]). Thus, we have to equivalently show that any short exact sequence of A(2)∗-modules

∗ 0 → A//A(2) → Z → L2[−1] → 0

∗ splits. Let xi ∈ L2[−1] be the unique element in degree i, if exists. ∗ Let z0 ∈ Z be any lift of x0 ∈ L2[−1] and z1 the unique lift of x1. Then we can attempt to produce a splitting by mapping x0 to z0 and x1 to z1. Since the lowest non-zero degrees of A//A(2) are 0, 8, 12, the only obstruction is if we 4 2 1 4 4 2 1 cannot map x8 to Sq Sq Sq z1. This occurs if Sq Sq Sq Sq z1 =6 0. But 4 4 2 1 7 3 1 3 1 Sq Sq Sq Sq = Sq Sq Sq , and Sq Sq z1 has degree 5, where Z is trivial. So this cannot happen.  Corollary A.23. We have

C2 H∗Y2 = AA(2)(F2 ⊕ F2 ⊕ H∗L).

C2 Corollary A.24. If Y2 admits the structure of a tmf-module, then upon 2- completion, we have an isomorphism

C2 Y2 = tmf ⊕ tmf ⊕ tmf ∧ L.

Proof. The Adams spectral sequence has two h0-towers at 0, and this lets us map in two copies of tmf. After quotienting them out we are left with ExtA(2)(F2,H∗L), which is shown in Figure 18.

10

8

6

4

2

0

0 2 4 6 8 10 12 14 16 18 20 22 24

Figure 18. ExtA(2)(F2,H∗L)

Standard obstruction theory lets us map in L, which lifts to a map from tmf ∧ L as desired. 

Appendix B. Sage script This appendix contains the sage script used to perform the computer calculations. The actual computations are at the end of the script and the comments indicate the lemmas they prove. 36 DEXTER CHUA

2 4 # Define the ring Γ[z]/(2z − a1z + a3z ) # We first start with the generators S. = PolynomialRing(QQ)

# Some convenient constants. The convention is that\psi(a)= a2 rp=(sp^2 + a_1p* sp)/3 a_12p= a_1p+2* sp a_32p= a_3p+ a_1p* rp+2* tp a_32ip= a_32p^2 * (a_12p^3 - 27 * a_32p) a_3ip= a_3p^2 * (a_1p^3 - 27 * a_3p) dp= a_3p^3 * (a_1p^3 - 27 * a_3p)

s4=6* sp* tp- a_1p* sp^3 + 3 * a_1p* tp+3* a_3p* sp t2=(sp^6 + 3 * a_1p* sp^5 - 9 * a_1p* sp^2 * tp + 3 * a_1p^2 * sp^4 - 9 * a_1p^2 * sp* tp + a_1p^3 * sp^3 - 27 * a_3p* tp) / 27

relations=(sp^4 - s4, tp^2 - t2, 2 * zp- a_12p* zp^2 + a_32p* zp^4, dp - 1)

# Now we impose the relations T. =S.quo(relations)

a_12=T(a_12p) a_32=T(a_32p) a_32i=T(a_32ip) a_3i=T(a_3ip) r=T(rp)

# z2 is the coaction onz z2=(z-r*z^3) / (1 -s*z+t*z^3)

# Some sanity checks assert a_32i* a_32 == 1 # This is the conjugate of the relation we imposed. assert2* z2- a_1* z2^2 + a_3* z2^4 == 0

# Computes\psi(x)-1(x)x def diff(x): returnT(x(a_1, a_3,s,t, z2)-x(a_12, a_32,s,t,z))

# Substitute occurrences of‘old‘ with‘new‘ once def substitute(expr, old, new): quo, rem= expr.quo_rem(old) return quo* new+ rem

# Keep performing substitutions until you can’t def substitute_full(expr, old, new): quo=1 while quo != 0: quo, rem= expr.quo_rem(old) expr= quo* new+ rem return expr

# Substitutions used to simplify expressions SUBS=[ [zp^4, a_32ip * (-2 * zp+ a_12p* zp^2)], C2-EQUIVARIANT TOPOLOGICAL MODULAR FORMS 37

[tp^2, t2], # We check this substitution below [sp^4 * tp,-a_1p^3*a_3p*sp - 4*a_1p^2*a_3p*sp^2 - 3*a_1p*a_3p*sp^3 - a_1p^4*tp - 3*a_1p^3*sp*tp - 4*a_1p^2*sp^2*tp - 3*a_1p*sp^3*tp - 6*a_3p^2*sp + 3*a_1p*a_3p*tp - 3*a_3p*sp*tp], [sp^4, s4], [dp, 1], ]

# Check our substitutions for old, new inSUBS: assertT(old- new) == 0 v=QQ.valuation(2)

# This function takes an expression, writes it ina canonical form, # and then reduces mod 2^n def reduce(result,n): original= result result= result.lift()

4 2 # First impose a3z − a1z + 2z = 0 to get rid of all # higher multiples ofz result= substitute_full(result,*(SUBS[0]))

# In this case, quotienting asa module is the same as # quotienting asa ring. result= substitute(result, zp^3, a_32ip*(a_12p* zp - 2))

# Then we get rid of allt^2 multiples. result= substitute_full(result,*(SUBS[1]))

# Next we get rid of alls^4 multiples. The expression fors^4 # involvest, so we need to handles^4t separately to avoid # re-introducingt^2. while result.quo_rem(sp^4)[0] != 0: result= substitute_full(result,*(SUBS[2])) result= substitute(result,*(SUBS[3]))

# Set ∆ = 1 and drop constants result= substitute_full(result,*(SUBS[4])) result= result.quo_rem(zp)[0] * zp

# Finally, reduce mod 2^n final=0 for(c,f) in list(result): c=v.simplify(c, error=(n - 1), force=True) final +=c*f return final

# Check comodule coactions(Lemma 3.2) assert reduce(diff(a_3p^2*zp)-r^2*a_32*z^2, 1) == 0 assert reduce(diff(a_3p* zp^2), 1) == 0 assert diff(a_3p* zp^3 - a_1p* zp) == 0 √ # Check ∆ is invariant 38 DEXTER CHUA assert reduce(diff(a_3p^2 * (1 + a_1p* zp))*z, 1) == 0

# Check that z_{-8} exists(Lemma 5.2) z4= zp^2 - a_3ip* a_1p^2 * zp z8= a_3p^3 * (a_1p* z4+2* zp) assert reduce(diff(z8), 2) == 0

# Computea Massey product(Lemma 5.3) assert reduce(diff(z4)+3*r* z8(a_12, a_32,s,t,z), 2) == 0

References [Bau08] Tilman Bauer. “Computation of the homotopy of the spectrum tmf”. In: Groups, homotopy and configuration spaces. Vol. 13. Geom. Topol. Monogr. Geom. Topol. Publ., Coventry, 2008, pp. 11–40. [Bea+21] Agnès Beaudry et al. The topological modular forms of RP 2 and RP 2 ∧ CP 2. 2021. arXiv: 2103.10953 [math.AT]. [BHS19] Robert Burklund, Jeremy Hahn, and Andrew Senger. On the boundaries of highly connected, almost closed . 2019. arXiv: 1910.14116 [math.AT]. [BR19] Scott M. Bailey and Nicolas Ricka. “On the Tate spectrum of tmf at the prime 2”. In: Math. Z. 291.3-4 (2019), pp. 821–829. [Die72] Tammo tom Dieck. “Kobordismentheorie klassifizierender Räume und Transformationsgruppen”. In: Math. Z. 126 (1972), pp. 31–39. [GM20] David Gepner and Lennart Meier. On equivariant topological modular forms. 2020. arXiv: 2004.10254 [math.AT]. [HK01] Po Hu and Igor Kriz. “Real-oriented homotopy theory and an analogue of the Adams–Novikov spectral sequence”. In: Topology 40.2 (2001), pp. 317–399. [HKR00] Michael J. Hopkins, Nicholas J. Kuhn, and Douglas C. Ravenel. “Gener- alized group characters and complex oriented cohomology theories”. In: J. Amer. Math. Soc. 13.3 (2000), pp. 553–594. [Lin+80] W. H. Lin et al. “Calculation of Lin’s Ext groups”. In: Math. Proc. Cambridge Philos. Soc. 87.3 (1980), pp. 459–469. [Lur09] J. Lurie. “A Survey of Elliptic Cohomology”. In: Algebraic Topology: The Abel Symposium 2007. Ed. by Nils Baas et al. Berlin, Heidelberg: Springer Berlin Heidelberg, 2009, pp. 219–277. [Lur12] Jacob Lurie. Higher Algebra. eng. 2012. [Lur18a] Jacob Lurie. “Elliptic Cohomology I: Spectral Abelian Varieties”. eng. In: (2018). [Lur18b] Jacob Lurie. “Elliptic Cohomology II: Orientations”. eng. In: (2018). [Lur19] Jacob Lurie. “Elliptic Cohomology III: Tempered Cohomology”. eng. In: (2019). [Mat16] Akhil Mathew. “The homology of tmf”. eng. In: Homology, homotopy, and applications 18.2 (2016), pp. 1–29. [Mil81] Haynes R. Miller. “On relations between Adams spectral sequences, with an application to the stable homotopy of a moore space”. In: Journal of Pure and Applied Algebra 20.3 (1981), pp. 287–312. [MM15] Akhil Mathew and Lennart Meier. “Affineness and chromatic homotopy theory”. In: Journal of Topology 8.2 (May 2015), pp. 476–528. [Mos70] R. Michael F. Moss. “Secondary compositions and the Adams spectral sequence”. In: Math. Z. 115 (1970), pp. 283–310. [MR08] Mark Mahowald and Charles Rezk. Topological Modular Forms of Level 3. 2008. arXiv: 0812.2009 [math.AT]. REFERENCES 39

[Pst18] Piotr Pstrągowski. Synthetic spectra and the cellular motivic category. 2018. arXiv: 1803.01804 [math.AT]. [Ric17] Nicolas Ricka. Motivic modular forms from equivariant stable homotopy theory. 2017. arXiv: 1704.04547 [math.AT]. [Seg68] Graeme Segal. “Equivariant K-theory”. en. In: Publications Mathéma- tiques de l’IHÉS 34 (1968), pp. 129–151.