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

On the slice spectral sequence for quotients of norms of Real bordism

Agnès Beaudry University of Colorado Boulder, Boulder, CO 80309 [email protected] Michael A. Hill University of California Los Angeles, Los Angeles, CA 90095 [email protected] Tyler Lawson University of Minnesota, Minneapolis, MN 55455 [email protected] XiaoLin Danny Shi University of Washington, Seattle, WA 98105 [email protected]  and  Mingcong Zeng Max Planck Institute for Mathematics, 53111 Bonn, Germany [email protected]
Abstract.

In this paper, we investigate equivariant quotients of the Real bordism spectrum’s multiplicative norm MU((C2n))MU^{(\!(C_{2^{n}})\!)} by permutation summands. These quotients are of interest because of their close relationship with higher real KK-theories. We introduce new techniques for computing the equivariant homotopy groups of such quotients.

As a new example, we examine the theories BP((C2n))m,mBP^{(\!(C_{2^{n}})\!)}\langle m,m\rangle. These spectra serve as natural equivariant generalizations of connective integral Morava KK-theories. We provide a complete computation of the aσa_{\sigma}-localized slice spectral sequence of iC2n1BP((C2n))m,mi^{*}_{C_{2^{n-1}}}BP^{(\!(C_{2^{n}})\!)}\langle m,m\rangle, where σ\sigma is the real sign representation of C2n1C_{2^{n-1}}. To achieve this computation, we establish a correspondence between this localized slice spectral sequence and the H𝔽2H\mathbb{F}_{2}-based Adams spectral sequence in the category of H𝔽2H𝔽2H\mathbb{F}_{2}\wedge H\mathbb{F}_{2}-modules. Furthermore, we provide a full computation of the aλa_{\lambda}-localized slice spectral sequence of the height-4 theory BP((C4))2,2BP^{(\!(C_{4})\!)}\langle 2,2\rangle. The C4C_{4}-slice spectral sequence can be entirely recovered from this computation.

1. Introduction

1.1. Motivation

Let E(k,Γ)E(k,\Gamma) be the Lubin–Tate spectrum associated to a formal group law Γ\Gamma of height hh over a finite field kk of characteristic 22. The Goerss–Hopkins–Miller theorem states that E(k,Γ)E(k,\Gamma) is a commutative ring spectrum and that there is an action of Aut(Γ)\mathrm{Aut}(\Gamma) on E(k,Γ)E(k,\Gamma) by commutative ring maps. Given a finite subgroup GG of Aut(Γ)\mathrm{Aut}(\Gamma), we can view E(k,Γ)E(k,\Gamma) as a GG-equivariant commutative ring spectrum via the cofree functor, and define a theory EO(k,Γ)(G)EO_{(k,\Gamma)}(G) by taking the fixed points under the action of GG:

EO(k,Γ)(G)E(k,Γ)hG.EO_{(k,\Gamma)}(G)\simeq E(k,\Gamma)^{hG}.

These are the higher real KK-theory spectra, so named because when the height is 11 and GG is C2C_{2}, these are a form of 22-completed real KK-theory.

Up to an étale extension, these spectra only depend on the height of Γ\Gamma and we will suppress (k,Γ)(k,\Gamma) from the notation by letting

Eh=E(k,Γ)andEOh(G)=EO(k,Γ)(G).E_{h}=E(k,\Gamma)\quad\text{and}\quad EO_{h}(G)=EO_{(k,\Gamma)}(G).

The spectra EOh(G)EO_{h}(G) play a central role in chromatic homotopy theory. Reasons of their importance include:

  1. (1)

    They detect interesting elements in the homotopy groups of spheres. For example, Hill–Hopkins–Ravenel’s work on manifolds of Kervaire invariant one [22] and Ravenel’s work [41] can be reinterpreted in terms of the Hurewicz images of EO4(C8)EO_{4}(C_{8}) and EOp1(Cp)EO_{p-1}(C_{p}). More recently, Li–Shi–Wang–Xu studied the Hurewicz image of EOh(C2)EO_{h}(C_{2}) [35].

  2. (2)

    They serve as fundamental building blocks for the K(h)K(h)-local sphere via the theory of finite resolutions. The theory of finite resolutions was developed by Goerss–Henn–Mahowald–Rezk [16] and expanded by Henn [20], followed by Bobkova–Goerss [10].

  3. (3)

    They are periodic theories that exhibit nice vanishing line properties [11, 29, 13]. This makes them more accessible to computations.

Historically, there have been few computations of homotopy groups of EOh(G)EO_{h}(G) for chromatic heights h>2h>2 at p=2p=2. At height h=1h=1, these computations are well understood via the relationship with complex and real KK-theory. At h=2h=2, computations are done using the close relationship of the higher real KK-theories with the spectrum tmftmf of topological modular forms and its analogues with level structures [12, 4, 9, 36, 27, 26]. However, at chromatic heights h>2h>2, such computations have been out of reach for a long time, in part due to the lack of nice geometric models for the higher real KK-theories such as koko and tmftmf.

More recently, the work of Hill–Hopkins–Ravenel [22] has made such computations more achievable. This is the approach we take in this paper. Specifically, we focus on the case of cyclic 2-groups, as a finite 22-subgroup of Aut(Γ)\mathrm{Aut}(\Gamma) at p=2p=2 is either a cyclic 22-group of order 2n2^{n} whenever h=2n1mh=2^{n-1}m for m1m\geq 1, or the quaternions when h=2mh=2m for mm odd [21]. Our restriction to the case of cyclic 2-groups allows us to use the equivariant slice filtration and related machinery developed in [22].

1.1.1. G=C2G=C_{2}

When G=C2G=C_{2}, there are two C2C_{2}-actions: one coming from the central subgroup C2C_{2} in Aut(Γ)\mathrm{Aut}(\Gamma) through “formal inversion” and the other coming from complex conjugation on the Real bordism spectrum MUMU_{{\mathbb{R}}} [15, 34]. Hahn and Shi [17] produced a Real orientation

MUEhMU_{{\mathbb{R}}}\longrightarrow E_{h}

from MUMU_{{\mathbb{R}}} to any Lubin–Tate spectrum at the prime 22. This Real orientation allows us to combine the two C2C_{2}-actions under one perspective, and to construct EhE_{h} as a localization of a quotient of MUMU_{{\mathbb{R}}}.

After localization at 22, MUMU_{{\mathbb{R}}} splits as a wedge of suspensions of the Real Brown–Peterson spectrum BPBP_{{\mathbb{R}}}. Let ρ2\rho_{2} be the regular representation. By work of Araki [2] and Landweber [34], we have

πρ2C2BP(2)[v¯1,v¯2,v¯3,]\pi_{*\rho_{2}}^{C_{2}}BP_{{\mathbb{R}}}\cong{\mathbb{Z}}_{(2)}[\bar{v}_{1},\bar{v}_{2},\bar{v}_{3},\ldots]

for generators v¯iπ(2i1)ρ2C2BP\bar{v}_{i}\in\pi_{(2^{i}-1)\rho_{2}}^{C_{2}}BP_{{\mathbb{R}}} whose underlying homotopy classes give generators viπ2(2i1)BPv_{i}\in\pi_{2(2^{i}-1)}BP for πBP\pi_{*}BP.

In this setup, we can refine two classical families of chromatic spectra to the C2C_{2}-equivariant world. They are both constructed as quotients of BPBP_{{\mathbb{R}}}:

  1. (1)

    The first family is the Real truncated Brown–Peterson spectrum

    BPh=BP/(v¯jj>h).BP_{{\mathbb{R}}}\langle h\rangle=BP_{{\mathbb{R}}}/(\bar{v}_{j}\mid j>h).

    The underlying non-equivariant spectrum is the classical truncated Brown–Peterson spectrum BPhBP\langle h\rangle. The K(h)K(h)-localization of BPhBP_{{\mathbb{R}}}\langle h\rangle gives, up to periodization, a model of Lubin–Tate theory EhE_{h} with its canonical C2C_{2}-action obtained through Goerss–Hopkins–Miller theory. These equivariant spectra and their v¯h\bar{v}_{h}-localizations were first studied by Hu–Kriz [32] and Kitchloo–Wilson [33].

  2. (2)

    The second family is the Real connective integral Morava KK-theory

    BPh,h:=BP/(v¯ii0,h),BP_{{\mathbb{R}}}\langle h,h\rangle:=BP_{{\mathbb{R}}}/(\bar{v}_{i}\mid i\neq 0,h),

    whose underlying spectra are the connective integral Morava KK-theories. After quotienting by 2 and periodization, we obtain the classical Morava KK-theories K(h)K(h).

1.1.2. Larger cyclic 22-groups

In this paper, we will study the C2nC_{2^{n}}-generalizations of the integral Morava KK-theories in great computational depth.

Let GG be a finite subgroup of Aut(Γ)\mathrm{Aut}(\Gamma) containing C2C_{2}. Since EhE_{h} is an equivariant commutative ring spectrum, the norm-forgetful adjunction produces a GG-equivariant orientation map

MU((G)):=NC2GMUEh.MU^{(\!(G)\!)}:=N_{C_{2}}^{G}MU_{{\mathbb{R}}}\longrightarrow E_{h}.

Since we are working 22-locally, we can substitute MUMU_{{\mathbb{R}}} with BPBP_{{\mathbb{R}}} using Quillen’s idempotent, thereby obtaining a map

BP((G)):=NC2GBPEh.BP^{(\!(G)\!)}:=N_{C_{2}}^{G}BP_{{\mathbb{R}}}\longrightarrow E_{h}.

This map allows us to regard (BP((G)))G\big{(}BP^{(\!(G)\!)}\big{)}^{G} as a global approximation for the theory EOh(G)EO_{h}(G).

We will concentrate on the case G=C2nG=C_{2^{n}} and h=2n1mh=2^{n-1}m in this paper. To compute the homotopy groups πGBP((G))\pi_{\star}^{G}BP^{(\!(G)\!)}, Hill, Hopkins, and Ravenel invented the equivariant slice spectral sequence [22]. The current approach of using the slice spectral sequence to understand πGBP((G))\pi_{\star}^{G}BP^{(\!(G)\!)} is to pass to quotients of BP((G))BP^{(\!(G)\!)} that generalize BPhBP_{{\mathbb{R}}}\langle h\rangle. To define these quotients, note that as in [22], the C2C_{2}-equivariant homotopy groups of BP((G))BP^{(\!(G)\!)} in degrees an integer multiple of ρ2\rho_{2} are

πρ2C2BP((G))(2)[Gv¯1G,Gv¯2G,],\pi_{*\rho_{2}}^{C_{2}}BP^{(\!(G)\!)}\cong{\mathbb{Z}}_{(2)}[G\cdot\bar{v}_{1}^{G},G\cdot\bar{v}_{2}^{G},\ldots],

where v¯iGπ(2i1)ρ2C2BP((G))\bar{v}_{i}^{G}\in\pi_{(2^{i}-1)\rho_{2}}^{C_{2}}BP^{(\!(G)\!)}. Here, Gv¯iGG\cdot\bar{v}_{i}^{G} denotes a set of 2n12^{n-1} elements

{v¯iG,γv¯iG,,γ2n11v¯iG},\left\{\bar{v}_{i}^{G},\gamma\bar{v}_{i}^{G},\ldots,\gamma^{2^{n-1}-1}\bar{v}_{i}^{G}\right\},

where γ\gamma represents a generator of GG and the Weyl action of GG is made obvious by the notation except for γ2n1v¯iG=v¯iG\gamma^{2^{n-1}}\bar{v}_{i}^{G}=-\bar{v}_{i}^{G}. The method of twisted monoid rings [22, Section 2] then allows one to form quotients of BP((G))BP^{(\!(G)\!)} by collections of permutation summands of the form Gv¯iGG\cdot\bar{v}_{i}^{G}. These quotients are the main objects of study in this paper.

The generalizations of the Real truncated Brown–Peterson spectrum BPhBP_{{\mathbb{R}}}\langle h\rangle and the Real connective integral Morava KK-theory BPh,hBP_{{\mathbb{R}}}\langle h,h\rangle are the following quotients by permutation summands:

  1. (1)

    The quotient

    BP((G))m:=BP((G))/(Gv¯m+1G,Gv¯m+2G,)BP^{(\!(G)\!)}\langle m\rangle:=BP^{(\!(G)\!)}/(G\cdot\bar{v}_{m+1}^{G},G\cdot\bar{v}_{m+2}^{G},\ldots)

    generalizes the spectrum BPhBP_{{\mathbb{R}}}\langle h\rangle. These spectra were studied in [7], where it was shown that BP((G))mBP^{(\!(G)\!)}\langle m\rangle is of height 2n1m{\leq 2^{n-1}m} ([7, Theorem 7.5]). Up to periodization and K(h)K(h)-localization, the GG-fixed points of BP((G))mBP^{(\!(G)\!)}\langle m\rangle gives a model for EOh(G)EO_{h}(G). The theories BP((G))mBP^{(\!(G)\!)}\langle m\rangle also give a chromatic filtration of BP((G))BP^{(\!(G)\!)} via the tower

    BP((G))mBP((G))m1BP((G))1.\cdots\longrightarrow BP^{(\!(G)\!)}\langle m\rangle\longrightarrow BP^{(\!(G)\!)}\langle m-1\rangle\longrightarrow\cdots\longrightarrow BP^{(\!(G)\!)}\langle 1\rangle. (1.1)

    The slice spectral sequences for these quotients have been computed for BPmBP_{{\mathbb{R}}}\langle m\rangle (m1m\geq 1), BP((C4))1BP^{(\!(C_{4})\!)}\langle 1\rangle, and BP((C4))2BP^{(\!(C_{4})\!)}\langle 2\rangle [32, 26, 28].

  2. (2)

    The equivariant generalization of the connective integral Morava KK-theories are the quotients

    BP((G))m,m:=BP((G))m/(Gv¯1G,,Gv¯m1G).BP^{(\!(G)\!)}\langle m,m\rangle:=BP^{(\!(G)\!)}\langle m\rangle/(G\cdot\bar{v}_{1}^{G},\ldots,G\cdot\bar{v}_{m-1}^{G}).

Given BP((G))m,mBP^{(\!(G)\!)}\langle m,m\rangle, we can apply further quotienting and localization to form the GG-spectrum

KG(h):=NC2G(v¯mG)1BP((G))m,m/G(v¯mGγv¯mG).K_{G}(h):=N_{C_{2}}^{G}(\bar{v}_{m}^{G})^{-1}BP^{(\!(G)\!)}\langle m,m\rangle/G\cdot(\bar{v}_{m}^{G}-\gamma\bar{v}_{m}^{G}).

The underlying spectrum of KG(h)K_{G}(h) is the 2(2m1)2\cdot(2^{m}-1)-periodic Morava KK-theory of height h=2n1mh=2^{n-1}\cdot m, with coefficient ring

πeKG(h)=𝔽2[v¯m±].\pi_{*}^{e}K_{G}(h)=\mathbb{F}_{2}[\bar{v}_{m}^{\pm}].

The group GG acts trivially on πeKG(h)\pi_{*}^{e}K_{G}(h), but the action of GG on KG(h)K_{G}(h) is nontrivial and compatible with the stabilizer group GG-action on the height-hh Lubin–Tate theory.

An important feature of BP((G))m,mBP^{(\!(G)\!)}\langle m,m\rangle is that its slice E2E_{2}-page contains significantly fewer classes compared to the slice E2E_{2}-page of BP((G))mBP^{(\!(G)\!)}\langle m\rangle. The lengths of its differentials are also more concentrated in certain ranges. These properties enhance the computational manageability of these theories, making them ideal as computable approximations for the equivariant truncated Brown–Peterson spectra BP((G))mBP^{(\!(G)\!)}\langle m\rangle and BP((G))BP^{(\!(G)\!)}.

To this end, the main objective of this paper is to investigate quotients of BP((G))BP^{(\!(G)\!)} by permutation summands, with a particular focus on the spectrum BP((G))m,mBP^{(\!(G)\!)}\langle m,m\rangle. By exploring the theoretical and computational properties of these equivariant integral Morava KK-theories, we seek to gain a deeper understanding of the overall structure of BP((G))mBP^{(\!(G)\!)}\langle m\rangle and the chromatic filtration tower (1.1).

1.2. Main results

We will now provide an outline of the paper and state our main results. Throughout the paper, the group G=C2nG=C_{2^{n}}.

Section 2

In the first section of this paper, we define various equivariant quotients by permutation summands and study their slice filtration. We begin by defining permutation summands for BP((G))BP^{(\!(G)\!)}, which are collections of elements of the form Gv¯jGG\cdot\bar{v}_{j}^{G} (see Definition 2.1). Our main result in this section is the following, which provides a simple description of the slice associated graded for quotients by permutation summands:

Theorem A (Theorem 2.5).

The slice associated graded of the quotient

BP((G))/(Gv¯jGjJ)BP^{(\!(G)\!)}/(G\cdot\bar{v}_{j}^{G}\mid j\in J)

where JJ is a subset of the natural numbers, is the generalized Eilenberg–Mac Lane spectrum

H¯[Gv¯iGiJ].H{\underline{{\mathbb{Z}}}}[G\cdot\bar{v}_{i}^{G}\mid i\not\in J].

Notably, Theorem A implies that the slice associated graded for BP((G))m,mBP^{(\!(G)\!)}\langle m,m\rangle is H¯[Gv¯mG]H{\underline{{\mathbb{Z}}}}[G\cdot\bar{v}_{m}^{G}]. We remark that our results do not depend on the specific choice of generators of the permutation summand: we can replace v¯m\bar{v}_{m} by any element s¯2m1\bar{s}_{2^{m}-1} in π(2m1)ρ2C2BP((G))\pi_{(2^{m}-1)\rho_{2}}^{C_{2}}BP^{(\!(G)\!)} that generates a permutation summand. To streamline notation, we write

GS¯={Gs¯2j1jJ}G\cdot\bar{S}=\{G\cdot\bar{s}_{2^{j}-1}\mid j\in J\}

for S¯={s¯2j1jJ}\bar{S}=\{\bar{s}_{2^{j}-1}\mid j\in J\}.

Section 3

In this section, we determine the chromatic heights of the generalized integral Morava KK-theory spectra BP((G))m,mBP^{(\!(G)\!)}\langle m,m\rangle.

Theorem B (Theorem 3.1).

The underlying spectrum of BP((G))m,mBP^{(\!(G)\!)}\langle m,m\rangle has non-trivial chromatic localizations at heights equal to rmrm, where 0r2n10\leq r\leq 2^{n-1}. That is,

  1. (1)

    for r=kmr=km where 0k2n10\leq k\leq 2^{n-1}, LK(r)ieBP((G))m,m≄L_{K(r)}i_{e}^{*}BP^{(\!(G)\!)}\langle m,m\rangle\not\simeq*, and

  2. (2)

    for all other r0r\geq 0, LK(r)ieBP((G))m,mL_{K(r)}i_{e}^{*}BP^{(\!(G)\!)}\langle m,m\rangle\simeq*.

In other words, the spectrum BP((G))m,mBP^{(\!(G)\!)}\langle m,m\rangle captures chromatic information at heights 0, mm, 2m2m, 3m3m, \ldots, 2n1m2^{n-1}m.

We note that Theorem B is similar to [7, Theorem 1.9], where the underlying spectrum of BP((G))mBP^{(\!(G)\!)}\langle m\rangle is shown to have nontrivial chromatic localizations at heights 2n1m\leq 2^{n-1}m. Furthermore, the GG-actions on BP((G))m,mBP^{(\!(G)\!)}\langle m,m\rangle and BP((G))mBP^{(\!(G)\!)}\langle m\rangle are compatible with the GG-action on E2n1mE_{2^{n-1}m} that is induced from the Morava stabilizer group.

Section 4

In this section, we introduce our main computational tool, the localized slice spectral sequence. This spectral sequence was developed in [39], and we summarize some of its key features here.

Suppose XX is a GG-spectrum and HH is a normal subgroup of GG. The localized slice spectral sequence is obtained by smashing the slice tower of XX with E~[H]\widetilde{E}\mathcal{F}[H], where [H]\mathcal{F}[H] is the family containing all proper subgroups of HH. By [39], the localized slice spectral sequence converges strongly to the GG-equivariant homotopy groups of E~[H]X\widetilde{E}\mathcal{F}[H]\wedge X, which is equal to the homotopy groups of (ΦH(X))G/H(\Phi^{H}(X))^{G/H} (see Theorem 4.5).

When G=H=C2nG=H=C_{2^{n}}, the localized slice spectral sequence is the aσa_{\sigma}-localization of the slice spectral sequence, where σ\sigma the sign representation. If G=C2nG=C_{2^{n}} and H=C2iH=C_{2^{i}} for i<ni<n, then the localized slice spectral sequence is the aλi1a_{\lambda_{i-1}}-localization of the slice spectral sequence, where λi1\lambda_{i-1} is the representation that rotates the plane by an angle of π/2ni\pi/2^{n-i}.

For G=C2nG=C_{2^{n}}, the slice spectral sequence of XX is divided into different regions, separated by the lines through the origin of slopes (2i1)(2^{i}-1), 0in0\leq i\leq n. In [38], the authors proved the Slice Recovery Theorem (Theorem 4.6), which states that for XX a (1)(-1)-connected GG-spectrum, the map from the original slice spectral sequence of XX to the localized slice spectral sequence of E~[C2i]X\widetilde{E}\mathcal{F}[C_{2^{i}}]\wedge X induces an isomorphism between all the differentials on or above the line of slope (2i11)(2^{i-1}-1) for all 1in1\leq i\leq n. In other words, even though the localized slice spectral sequence only computes the geometric fixed points, its E2E_{2}-page and differentials captures all the corresponding information in the original slice spectral sequence (which computes the fixed points) within a range.

Section 5

In this section, we present the following computational result, using the tools discussed earlier.

Theorem C (Theorem 5.9 and Theorem 5.10).

Let JJ\subseteq\mathbb{N}, and let S¯={s¯2j1jJ}\bar{S}=\{\bar{s}_{2^{j}-1}\mid j\in J\} be a set of generators for permutation summands. The following hold:

  1. (1)

    The aσa_{\sigma}-localized slice spectral sequence of aσ1BP((G))/GS¯{a_{\sigma}^{-1}BP^{(\!(G)\!)}/G\cdot\bar{S}} has only nontrivial differentials of lengths (k)=2n(2k+11)+1\ell(k)=2^{n}(2^{k+1}-1)+1, where k+1Jk+1\notin J. Moreover, for a fixed kk, all the (k)\ell(k)-differentials are multiples of a nontrivial d(k)d_{\ell(k)}-differential on the class b2kb^{2^{k}}, where b=u2σ/aσ2b=u_{2\sigma}/a_{\sigma}^{2}.

  2. (2)

    The aσa_{\sigma}-localized slice spectral sequence of aσ1BP((G))/GS¯{a_{\sigma}^{-1}BP^{(\!(G)\!)}/G\cdot\bar{S}}, which converges to the homotopy groups of the GG-geometric fixed points of BP((G))/GS¯{BP^{(\!(G)\!)}/G\cdot\bar{S}}, completely determines all the differentials on or above the line of slope (2n11){(2^{n-1}-1)} in the slice spectral sequence of BP((G))/GS¯BP^{(\!(G)\!)}/G\cdot\bar{S}.

The computation in Theorem C is done using the Slice Differential Theorem of Hill–Hopkins–Ravenel [22]. The quotient map

BP((G))BP((G))/GS¯BP^{(\!(G)\!)}\longrightarrow BP^{(\!(G)\!)}/G\cdot\bar{S}

induces a map of the corresponding localized slice spectral sequences. The Slice Differential Theorem produces all the differentials in the aσa_{\sigma}-localized slice spectral sequence of aσ1BP((G))a_{\sigma}^{-1}BP^{(\!(G)\!)}. Using the module structure and naturality, we deduce all the differentials in the aσa_{\sigma}-localized slice spectral sequence of aσ1BP((G))/GS¯a_{\sigma}^{-1}BP^{(\!(G)\!)}/G\cdot\bar{S}.

When G=C2G=C_{2}, the aσa_{\sigma}-localized slice spectral sequence of aσ1BP/S¯a_{\sigma}^{-1}BP_{{\mathbb{R}}}/\bar{S} produces all the differentials in the slice spectral sequence of BP/S¯BP_{{\mathbb{R}}}/\bar{S}. This is explained in Corollary 5.12.

Theorem C is used to show that, in stark contrast to the non-equivariant setting, most of the quotients of BP((G))BP^{(\!(G)\!)} by permutation summands do not admit a ring structure even in the homotopy category.

Theorem D (Theorem 5.16).

Let JJ\subseteq\mathbb{N}, and let S¯={s¯2j1jJ}\bar{S}=\{\bar{s}_{2^{j}-1}\mid j\in J\} be a set of generators for permutation summands. If there is a kJk\in J such that (k+1)J(k+1)\notin J, then BP((G))/GS¯BP^{(\!(G)\!)}/G\cdot\bar{S} does not have a ring structure in the homotopy category.

Section 6

In this section, we analyze the next region of the slice spectral sequence of BP((G))m,mBP^{(\!(G)\!)}\langle m,m\rangle, namely the region between the lines of slopes (2n21)(2^{n-2}-1) and (2n11)(2^{n-1}-1). Let G=C2n1G^{\prime}=C_{2^{n-1}} be the index 2 subgroup of G=C2nG=C_{2^{n}}. To compute the differentials in this region, we examine the localized slice spectral sequence of

E~[G]BP((G))m,maλn21BP((G))m,m,\widetilde{E}\mathcal{F}[G^{\prime}]\wedge BP^{(\!(G)\!)}\langle m,m\rangle\simeq a_{\lambda_{n-2}}^{-1}BP^{(\!(G)\!)}\langle m,m\rangle,

which computes the homotopy groups of the G/GG/G^{\prime}-fixed points of the spectrum ΦG(BP((G))m,m)\Phi^{G^{\prime}}(BP^{(\!(G)\!)}\langle m,m\rangle).

A valuable input to computing this spectral sequence is its restriction to the group GG^{\prime}, which computes the underlying homotopy groups of ΦG(BP((G))m,m)\Phi^{G^{\prime}}(BP^{(\!(G)\!)}\langle m,m\rangle). Using the Mackey functor structure, we can then deduce information about the GG-equivariant spectral sequence from the simpler GG^{\prime}-equivariant spectral sequence.

Another important spectral sequence that comes into play is the H𝔽2H{\mathbb{F}}_{2}-based Adams spectral sequence in the category of AA-module spectra (as in Baker–Lazarev [3]), where

A:=H𝔽2H𝔽2.A:=H\mathbb{F}_{2}\wedge H\mathbb{F}_{2}.

Here, H𝔽2H{\mathbb{F}}_{2} is given an AA-module structure via the multiplication map AH𝔽2A\to H{\mathbb{F}}_{2}. We call this spectral sequence the relative Adams spectral sequence.

As non-equivariant spectra, there is an equivalence

ΦG(BP((G))m,m)A/(ξi,ζi:im)\Phi^{G^{\prime}}(BP^{(\!(G)\!)}\langle m,m\rangle)\simeq A/(\xi_{i},\zeta_{i}:i\neq m)

for ξi\xi_{i} and ζi\zeta_{i} the usual Milnor generators and their conjugates. In fact, there is an intimate connection between the more classical relative Adams spectral sequence and the GG^{\prime}-equivariant localized slice spectral sequence, which we establish in Section 6.2.

Theorem E (Theorem 6.7, Corollary 6.8, and Corollary 6.10).

After a reindexing of filtrations, the GG^{\prime}-equivariant localized slice spectral sequence of BP((G))m,m{BP^{(\!(G)\!)}\langle m,m\rangle} is isomorphic to the relative Adams spectral sequence of A/(ξi,ζi:im)A/(\xi_{i},\zeta_{i}:i\neq m).

In Section 6.3, we apply the techniques developed in [6] and use the correspondence established in Theorem E to obtain the following computational result:

Theorem F (Theorem 6.20 and Summary 6.21).

We determine all the differentials in the following two spectral sequences:

  1. (1)

    the relative Adams spectral sequence of A/(ξi,ζi:im)A/(\xi_{i},\zeta_{i}:i\neq m);

  2. (2)

    the GG^{\prime}-equivariant localized slice spectral sequence of BP((G))m,mBP^{(\!(G)\!)}\langle m,m\rangle.

Theorem F demonstrates the effectiveness of our methods and allows us to proceed to the next stage of our analysis, which is to compute BP((C4))2,2BP^{(\!(C_{4})\!)}\langle 2,2\rangle. As a concrete illustration of Theorem F, we present a detailed analysis of the C2C_{2}-equivariant localized slice spectral sequence of BP((C4))2,2BP^{(\!(C_{4})\!)}\langle 2,2\rangle in Section 6.4.

Section 7

In the final section of the paper, we provide a complete computation of the slice spectral sequence of BP((C4))2,2BP^{(\!(C_{4})\!)}\langle 2,2\rangle. This serves as a showcase of the strength of our methods, and also offers insight into higher differentials phenomena when applied to higher heights and bigger groups.

Theorem G.

We determine all the differentials in the aλa_{\lambda}-localized slice spectral sequence of BP((C4))2,2BP^{(\!(C_{4})\!)}\langle 2,2\rangle. The spectral sequence terminates after the E61E_{61}-page and has a vanishing line of slope (1)(-1) on the EE_{\infty}-page.

By the Slice Recovery Theorem (Theorem 4.6), this computation completely determines all the differentials in the slice spectral sequence of BP((C4))2,2BP^{(\!(C_{4})\!)}\langle 2,2\rangle by truncating away the region below the line of filtration s=0s=0. In particular, Theorem G shows that the slice spectral sequence of BP((C4))2,2BP^{(\!(C_{4})\!)}\langle 2,2\rangle terminates after the E61E_{61}-page and has a horizontal vanishing line of filtration 61.

According to Theorem B, the underlying spectrum of BP((C4))2,2BP^{(\!(C_{4})\!)}\langle 2,2\rangle is of heights 0, 2 and 4, and is equipped with a C4C_{4}-action that is compatible with the stabilizer group action on a height-4 Lubin–Tate theory. The computation in Theorem G is a height-4 computation of a spectrum that is closely related to BP((C4))2BP^{(\!(C_{4})\!)}\langle 2\rangle, studied in [28]. There is a map

SliceSS(BP((C4))2)SliceSS(BP((C4))2,2).\operatorname{SliceSS}(BP^{(\!(C_{4})\!)}\langle 2\rangle)\longrightarrow\operatorname{SliceSS}(BP^{(\!(C_{4})\!)}\langle 2,2\rangle).

Compared to the slice spectral sequence of BP((C4))2BP^{(\!(C_{4})\!)}\langle 2\rangle, the slice spectral sequence of BP((C4))2,2BP^{(\!(C_{4})\!)}\langle 2,2\rangle has fewer classes on the E2E_{2}-page, and the lengths of differentials are concentrated in certain ranges.

The following features of the slice spectral sequence of BP((C4))2,2BP^{(\!(C_{4})\!)}\langle 2,2\rangle are essential to the computation in Theorem G:

  1. (1)

    Differentials on or above the line of slope 1 are determined by aσa_{\sigma}-localized slice spectral sequence, which is computed in Theorem C.

  2. (2)

    The shorter differentials (d31d_{\leq 31}) are all determined from the C2C_{2}-slice differentials in Theorem F and Mackey functor structures.

  3. (3)

    To determine the higher differentials, we identified two key classes, α=𝔡¯t¯28u24σa24λ\alpha=\bar{\mathfrak{d}}_{\bar{t}_{2}}^{8}u_{24\sigma}a_{24\lambda} at (48,48)(48,48) and b32=u32λ/a32λb^{32}=u_{32\lambda}/a_{32\lambda} at (64,64)(64,-64). Multiplication with respect to these classes gives rise to periodicity of differentials and a vanishing line of slope 1-1 (Theorem 7.20). These phenomena determine all the higher differentials (d>31d_{>31}).

We believe these features can be extended to higher heights and larger groups, leading to a global description of all slice spectral sequence computations of quotients of BP((G))BP^{(\!(G)\!)} (see Section 1.3).

1.3. Open questions and future directions

The relationship between equivariant and chromatic homotopy theory is an exciting landscape whose exploration has only just begun. The results we present here reveal new aspects of the connection between GG-spectra and K(h)K(h)-local phenomena. They also open questions and suggest conjectures. We end this introduction by highlighting a few.

GG-equivariant periodic Morava-KK theory

In this paper, our focus has mostly been on the theories BP((G))m,mBP^{(\!(G)\!)}\langle m,m\rangle. Recall the spectrum

KG(h):=NC2G(v¯mG)1BP((G))m,m/G(v¯mGγv¯mG)K_{G}(h):=N_{C_{2}}^{G}(\bar{v}_{m}^{G})^{-1}BP^{(\!(G)\!)}\langle m,m\rangle/G\cdot(\bar{v}_{m}^{G}-\gamma\bar{v}_{m}^{G})

defined at the end of Section 1.1. This spectrum is the GG-equivariant generalization of the height-hh Morava KK-theory. The following questions about KG(h)K_{G}(h) will provide further insights into the structure of the Morava stabilizer group and the behavior of quotients of BP((G))BP^{(\!(G)\!)}.

Question 1.1.

What is the slice filtration of KG(h)K_{G}(h) and differentials in the slice spectral sequence of KG(h)K_{G}(h)?

Note that since KG(h)K_{G}(h) is not a quotient by permutation summands, Theorem A does not directly apply. Nonetheless, we have determined the slice filtration of KG(h)K_{G}(h) when G=C2G=C_{2} and C4C_{4}, as well as the slice differentials for all hh when G=C2G=C_{2}, and for h=2h=2 when G=C4G=C_{4}.

Question 1.2.

Is it possible to build an equivariant chromatic fracture square with the theories BP((G))mBP^{(\!(G)\!)}\langle m\rangle, BP((G))m1BP^{(\!(G)\!)}\langle m-1\rangle, BP((G))m,mBP^{(\!(G)\!)}\langle m,m\rangle, and KG(h)K_{G}(h)? In particular, what is the relationship between BP((G))m,mBP^{(\!(G)\!)}\langle m,m\rangle, KG(h)K_{G}(h), and the fiber of the map BP((G))mBP((G))m1BP^{(\!(G)\!)}\langle m\rangle\to BP^{(\!(G)\!)}\langle m-1\rangle in the chromatic filtration of BP((G))BP^{(\!(G)\!)}?

Question 1.3.

What are the Hurewciz images of BP((G))m,mBP^{(\!(G)\!)}\langle m,m\rangle and KG(h)K_{G}(h), compared to that of BP((G))mBP^{(\!(G)\!)}\langle m\rangle?

The C2C_{2}-equivariant relative Adams spectral sequence

In Section 6, the correspondence established in Theorem E between the relative Adams spectral sequence and the C2C_{2}-localized slice spectral sequence played a crucial role in determining the C2C_{2}-slice differentials in the localized slice spectral sequence of aλ1BP((G))m,ma_{\lambda}^{-1}BP^{(\!(G)\!)}\langle m,m\rangle. In [19], Hahn and Wilson constructed a C2C_{2}-equivariant relative Adams spectral sequence, which can be utilized to compute the homotopy groups of NeC2H𝔽2N_{e}^{C_{2}}H\mathbb{F}_{2}-modules. Notably, the C2C_{2}-geometric fixed points of quotients of BP((C4))BP^{(\!(C_{4})\!)} equipped with the residue C4/C2C_{4}/C_{2}-action are NeC2H𝔽2N_{e}^{C_{2}}H\mathbb{F}_{2}-modules. Both the C2C_{2}-equivariant relative Adams spectral sequence and the C4C_{4}-localized slice spectral sequence can be used to compute the homotopy groups of such quotients.

Question 1.4.

Is there an equivariant analogue of the correspondence in Theorem E for BP((C4))m,mBP^{(\!(C_{4})\!)}\langle m,m\rangle and general quotients of BP((C4))BP^{(\!(C_{4})\!)}? In particular, can we establish a correspondence between differentials in the C2C_{2}-equivariant Adams spectral sequence and the C4C_{4}-localized slice spectral sequence?

Global structures in the slice spectral sequence

Understanding the equivariant homotopy groups of BP((C4))m,mBP^{(\!(C_{4})\!)}\langle m,m\rangle for all m1m\geq 1 would significantly deepen our knowledge of higher chromatic heights and provide valuable insight into the C4C_{4}-fixed points of height-(2m)(2m) Lubin–Tate theories. To facilitate these computations, it is important to establish certain general properties of the localized slice spectral sequences for these quotients.

Our computation of BP((C4))2,2BP^{(\!(C_{4})\!)}\langle 2,2\rangle in this paper leads us to believe that certain structures, such as vanishing lines and periodicity of differentials, should be present in the localized slice spectral sequences for all BP((C4))m,mBP^{(\!(C_{4})\!)}\langle m,m\rangle. Answering the following questions would significantly simplify the computation of localized slice spectral sequences for quotients of BP((G))BP^{(\!(G)\!)}.

Question 1.5.

Do vanishing lines of slope (1)(-1) always exist on the EE_{\infty}-pages of the localized slice spectral sequences for BP((G))m,mBP^{(\!(G)\!)}\langle m,m\rangle and other quotients of BP((G))BP^{(\!(G)\!)}?

The presence of such vanishing lines is closely linked to the existence of horizontal vanishing lines in the slice spectral sequence for quotients of BP((G))BP^{(\!(G)\!)}, a question raised in [13].

Question 1.6.

Do there exist analogues of the classes α\alpha and b32b^{32} that induce differential periodicity in the localized slice spectral sequence for BP((G))m,mBP^{(\!(G)\!)}\langle m,m\rangle and other quotients of BP((G))BP^{(\!(G)\!)}?

For BP((C4))1,1BP^{(\!(C_{4})\!)}\langle 1,1\rangle, the analogous classes are α1=𝔡¯t¯14u4σa4λ\alpha_{1}=\bar{\mathfrak{d}}_{\bar{t}_{1}}^{4}u_{4\sigma}a_{4\lambda} in bidegree (8,8)(8,8) and b1=b8=u8λ/a8λb_{1}=b^{8}=u_{8\lambda}/a_{8\lambda} in bidegree (16,16)(16,-16). In general, let

αm=𝔡t¯m2m+1u(2m1)2m+1σa(2m1)2m+1λ\alpha_{m}=\mathfrak{d}_{\bar{t}_{m}}^{2^{m+1}}u_{(2^{m}-1)2^{m+1}\sigma}a_{(2^{m}-1)2^{m+1}\lambda}

and

bm=b22m+1=u22m+1λ/a22m+1λ.b_{m}=b^{2^{2m+1}}=u_{2^{2m+1}\lambda}/a_{2^{2m+1}\lambda}.
Conjecture 1.7.

In the C4C_{4}-localized slice spectral sequence of BP((C4))m,mBP^{(\!(C_{4})\!)}\langle m,m\rangle, multiplications by the classes αm\alpha_{m} and bmb_{m} induce differential periodicity and a vanishing line of slope (1)(-1).

1.4. Acknowledgements

The authors would like to thank Mark Behrens, Christian Carrick, Mike Hopkins, Hana Jia Kong, Guchuan Li, Yutao Liu, Lennart Meier, Juan Moreno, Doug Ravenel, Vesna Stojanoska, Guozhen Wang, Zhouli Xu, and Guoqi Yan for helpful conversations. This material is based upon work supported by the National Science Foundation under Grant No. DMS-1906227 (first author), DMS-2105019 (second author) and DMS-2313842 (fourth author).

2. Quotient modules of MU((G))MU^{(\!(G)\!)}

2.1. Slices for some MU((G))MU^{(\!(G)\!)} modules

For a graded ring RR with augmentation ideal II, let QnRQ_{n}R denote the degree nn elements in I/I2I/I^{2}. One of the key computations in [22] was a convenient choice of algebra generators for the ρ2*\rho_{2}-graded homotopy groups of MU((G))MU^{(\!(G)\!)}. In particular, we have an isomorphism of graded [G]{\mathbb{Z}}[G]-modules

Qρ2(πρ2C2MU((G)))k1Σkρ2IndC2G(k),Q_{\ast\rho_{2}}\big{(}\pi_{\ast\rho_{2}}^{C_{2}}MU^{(\!(G)\!)}\big{)}\cong\bigoplus_{k\geq 1}\Sigma^{k\rho_{2}}\operatorname{Ind}_{C_{2}}^{G}(\mathbb{Z}_{-}^{\otimes k}),

where \mathbb{Z}_{-} is the integral sign representation.

Definition 2.1.

Let JJ\subseteq\mathbb{N} and

S¯={s¯jπjρ2C2MU((G))jJ}\bar{S}=\{\bar{s}_{j}\in\pi_{j\rho_{2}}^{C_{2}}MU^{(\!(G)\!)}\mid j\in J\}

be a collection of elements. Associated to S¯\bar{S}, we have a C2C_{2}-equivariant map

fS¯:jJΣjρ2jjJQjρ2(MU((G))).f_{\bar{S}}\colon\bigoplus_{j\in J}\Sigma^{j\rho_{2}}\mathbb{Z}_{-}^{\otimes j}\to\bigoplus_{j\in J}Q_{j\rho_{2}}\big{(}MU^{(\!(G)\!)}\big{)}.

We say that S¯\bar{S} generates a permutation summand if the adjoint GG-equivariant map

f~S¯:jJΣjρ2IndC2G(j)jJQjρ2(MU((G)))\tilde{f}_{\bar{S}}\colon\bigoplus_{j\in J}\Sigma^{j\rho_{2}}\operatorname{Ind}_{C_{2}}^{G}(\mathbb{Z}_{-}^{\otimes j})\to\bigoplus_{j\in J}Q_{j\rho_{2}}\big{(}MU^{(\!(G)\!)}\big{)}

is an isomorphism.

Given any element in the RO(C2)RO(C_{2})-graded homotopy of MU((G))MU^{(\!(G)\!)}, we can use the method of twisted monoid algebras from [22]. For each jJj\in J, we have a free associative algebra

𝕊0[s¯j]:=i=0Sijρ2,\mathbb{S}^{0}[\bar{s}_{j}]:=\bigvee_{i=0}^{\infty}S^{ij\rho_{2}},

and we have a canonical associative algebra map

𝕊0[s¯j]iC2MU((G))\mathbb{S}^{0}[\bar{s}_{j}]\to i_{C_{2}}^{\ast}MU^{(\!(G)\!)}

adjoint to the map defining s¯j\bar{s}_{j}. Using the norm maps on MU((G))MU^{(\!(G)\!)} and the multiplication, we get an associative algebra map

𝕊0[GS¯]=jJNC2G𝕊0[s¯j]MU((G)).\mathbb{S}^{0}[G\cdot\bar{S}]=\bigwedge_{j\in J}N_{C_{2}}^{G}\mathbb{S}^{0}[\bar{s}_{j}]\to MU^{(\!(G)\!)}.
Definition 2.2.

For JJ\subseteq\mathbb{N}, let

MU((G))/GS¯:=MU((G))/(Gs¯jjJ)=MU((G))𝕊0[GS¯]𝕊0.MU^{(\!(G)\!)}/G\cdot\bar{S}:=MU^{(\!(G)\!)}/\big{(}G\cdot\bar{s}_{j}\mid j\in J\big{)}=MU^{(\!(G)\!)}\underset{\mathbb{S}^{0}[G\cdot\bar{S}]}{\wedge}\mathbb{S}^{0}.

We call MU((G))/GS¯MU^{(\!(G)\!)}/G\cdot\bar{S} a quotient by permutation summands.

The following is a slight generalization of the Slice Theorem of Hill–Hopkins–Ravenel [22, Theorem 6.1]. Recall from [22] that

πρ2C2MU((G))(2)[Gr¯1,Gr¯2,]\pi_{*\rho_{2}}^{C_{2}}MU^{(\!(G)\!)}\cong{\mathbb{Z}}_{(2)}[G\cdot\bar{r}_{1},G\cdot\bar{r}_{2},\ldots]

for generators r¯i\bar{r}_{i} in πiρ2C2MU((G))\pi_{i\rho_{2}}^{C_{2}}MU^{(\!(G)\!)} introduced in (5.39) of [22]. In fact, the conditions on the classes s¯j\bar{s}_{j} guarantee that we can use them instead of the r¯j\bar{r}_{j} for jJj\in J. We then extend the set S¯\bar{S} to form a set of equivariant algebra generators, as in [22, Section 5]. Any set

S¯:={s¯jjJ}\bar{S}^{\prime}:=\{\bar{s}_{j}\mid j\notin J\}

of elements s¯jπjρ2C2MU((G))\bar{s}_{j}\in\pi_{j\rho_{2}}^{C_{2}}MU^{(\!(G)\!)} which generate a permutation summand will do to extend S¯\bar{S} to a set S¯S¯\bar{S}\cup\bar{S}^{\prime} of equivariant generators for πρ2C2MU((G))\pi_{*\rho_{2}}^{C_{2}}MU^{(\!(G)\!)}.

Definition 2.3.

Let JJ\subseteq\mathbb{N} and S¯\bar{S} and S¯\bar{S}^{\prime} be as above. Define

H¯[Gs¯1,Gs¯2,]:=H¯𝕊0𝕊0[Gs¯1,]H{\underline{{\mathbb{Z}}}}[G\cdot\bar{s}_{1},G\cdot\bar{s}_{2},\dots]:=H{\underline{{\mathbb{Z}}}}\underset{\mathbb{S}^{0}}{\wedge}\mathbb{S}^{0}[G\cdot\bar{s}_{1},\ldots]

and

H¯[Gs¯1,Gs¯2,]/GS¯:=H¯[Gs¯1,]𝕊0[GS¯]𝕊0.H{\underline{{\mathbb{Z}}}}[G\cdot\bar{s}_{1},G\cdot\bar{s}_{2},\dots]/G\cdot\bar{S}:=H{\underline{{\mathbb{Z}}}}[G\cdot\bar{s}_{1},\dots]\underset{\mathbb{S}^{0}[G\cdot\bar{S}]}{\wedge}\mathbb{S}^{0}\ .
Remark 2.4.

The spectrum H¯[Gs¯1,Gs¯2,]/GS¯H{\underline{{\mathbb{Z}}}}[G\cdot\bar{s}_{1},G\cdot\bar{s}_{2},\dots]/G\cdot\bar{S} is very simple. In fact, it is equivalent to H¯[GS¯]H{\underline{{\mathbb{Z}}}}[G\cdot\bar{S}^{\prime}], which itself is the smash product over jJj\not\in J of the norms, in the category of H¯H{\underline{{\mathbb{Z}}}}-modules, of H¯[s¯j]i=0H¯Sijρ2H{\underline{{\mathbb{Z}}}}[\bar{s}_{j}]\simeq\bigvee_{i=0}^{\infty}H{\underline{{\mathbb{Z}}}}\wedge S^{ij\rho_{2}} .

Theorem 2.5.

The slice associated graded of MU((G))/GS¯MU^{(\!(G)\!)}/G\cdot\bar{S} is the generalized Eilenberg–Mac Lane spectrum

H¯[Gs¯1,Gs¯2,]/(GS¯).H{\underline{{\mathbb{Z}}}}[G\cdot\bar{s}_{1},G\cdot\bar{s}_{2},\dots]/(G\cdot\bar{S}).
Proof.

We have a natural equivalence

MU((G))/GS¯MU((G))𝕊0[Gs¯1,]𝕊0[GS¯].MU^{(\!(G)\!)}/G\cdot\bar{S}\simeq MU^{(\!(G)\!)}\underset{\mathbb{S}^{0}[G\cdot\bar{s}_{1},\dots]}{\wedge}\mathbb{S}^{0}[G\cdot\bar{S}^{\prime}].

The result now follows exactly as [22, Slice Theorem 6.1], using the natural degree filtration on 𝕊0[GS¯]\mathbb{S}^{0}[G\cdot\bar{S}^{\prime}]. ∎

Letting S¯\bar{S} be the generators killed by the Quillen idempotent, this recovers the usual form of the slice associated graded for BP((G))BP^{(\!(G)\!)}. We could moreover always append this to any collection S¯\bar{S} we consider, which allows us to deduce all of the analogous results for BP((G))BP^{(\!(G)\!)}. We will do so without comment moving forward.

Remark 2.6.

The left action of MU((G))MU^{(\!(G)\!)} on itself always endows MU((G))/GS¯MU^{(\!(G)\!)}/G\cdot\bar{S} with a canonical MU((G))MU^{(\!(G)\!)}-module structure, and the same is true with BP((G))BP^{(\!(G)\!)} instead.

Notation 2.7.

In the homotopy of the spectrum BP((G))BP^{(\!(G)\!)}, let

v¯kG:=t¯kGπ(2k1)ρ2C2BP((G)),{\bar{v}}_{k}^{G}:=\bar{t}_{k}^{G}\in\pi_{(2^{k}-1)\rho_{2}}^{C_{2}}BP^{(\!(G)\!)},

as defined and considered in [7].

Definition 2.8.

For each m0m\geq 0, let Jm={kk>m}J_{m}=\{k\mid k>m\}. Let

S¯m={s¯2j1π(2j1)ρ2C2BP((G))jJm}\bar{S}_{m}=\big{\{}\bar{s}_{2^{j}-1}\in\pi_{(2^{j}-1)\rho_{2}}^{C_{2}}BP^{(\!(G)\!)}\mid j\in J_{m}\big{\}}

generate a permutation summand. When for each jJmj\in J_{m}, s¯2j1=v¯jG\bar{s}_{2^{j}-1}={\bar{v}}_{j}^{G}, we name the quotient

BP((G))/GS¯m=BP((G))m.BP^{(\!(G)\!)}/G\cdot\bar{S}_{m}=BP^{(\!(G)\!)}\langle m\rangle.

More generally, we say that the BP((G))BP^{(\!(G)\!)}-module

BP((G))/GS¯mBP^{(\!(G)\!)}/G\cdot\bar{S}_{m}

is a form of BP((G))mBP^{(\!(G)\!)}\langle m\rangle.

Notation 2.9.

Let vkGv_{k}^{G} be the restriction to the trivial group of v¯kG\bar{v}_{k}^{G}.

Remark 2.10.

Just as in [23], we note that since the underlying rings are all polynomial rings, the map

π{e}BP((G))π{e}BP((G))m=(2)[Gv1G,,GvmG]\pi_{\ast}^{\{e\}}BP^{(\!(G)\!)}\to\pi_{\ast}^{\{e\}}BP^{(\!(G)\!)}\langle m\rangle={\mathbb{Z}}_{(2)}[G\cdot{v}_{1}^{G},\dots,G\cdot{v}_{m}^{G}]

has a section.

A form of BP((G))mBP^{(\!(G)\!)}\langle m\rangle is a quotient module MM with the property that for any section, the composite

(2)[Gv1G,,GvmG]π{e}BP((G))π{e}M{\mathbb{Z}}_{(2)}[G\cdot{v}_{1}^{G},\dots,G\cdot{v}_{m}^{G}]\to\pi_{\ast}^{\{e\}}BP^{(\!(G)\!)}\to\pi_{\ast}^{\{e\}}M

is an isomorphism. The difference between the forms lies in the BP((G))BP^{(\!(G)\!)}-module structure, not in the underlying homotopy groups.

Corollary 2.11.

The slice associated graded for any form of BP((G))mBP^{(\!(G)\!)}\langle m\rangle is

H¯[Gv¯1G,,Gv¯mG].H{\underline{{\mathbb{Z}}}}[G\cdot{\bar{v}}_{1}^{G},\dots,G\cdot{\bar{v}}_{m}^{G}].
Definition 2.12.

Let kk and mm be natural numbers with 1km1\leq k\leq m. Let

S¯k,m={v¯jG0<j<k or j>m},\bar{S}_{k,m}=\big{\{}\bar{v}_{j}^{G}\mid 0<j<k\text{ or }j>m\big{\}},

and let

BP((G))k,m=BP((G))/GS¯k,m.BP^{(\!(G)\!)}\langle k,m\rangle=BP^{(\!(G)\!)}/G\cdot\bar{S}_{k,m}.
Remark 2.13.

As in Definition 2.8, we also define forms of BP((G))k,mBP^{(\!(G)\!)}\langle k,m\rangle as quotients by elements s¯2j1\bar{s}_{2^{j}-1}, for 0<j<k0<j<k or j>mj>m that generate permutation summands.

Corollary 2.14.

The slice associated graded for BP((G))k,mBP^{(\!(G)\!)}\langle k,m\rangle (or for any form) is

H¯[Gv¯kG,,Gv¯mG].H{\underline{{\mathbb{Z}}}}[G\cdot\bar{v}_{k}^{G},\dots,G\cdot\bar{v}_{m}^{G}].

One of the main examples we will analyze is

BP((G))m,m=BP((G))/(Gv¯1G,,Gv¯m1G,Gv¯m+1G,)BP^{(\!(G)\!)}\langle m,m\rangle=BP^{(\!(G)\!)}/\big{(}G\cdot{\bar{v}}_{1}^{G},\dots,G\cdot{\bar{v}}_{m-1}^{G},G\cdot{\bar{v}}_{m+1}^{G},\dots\big{)}

where m1m\geq 1.

The slice associated graded for BP((G))m,mBP^{(\!(G)\!)}\langle m,m\rangle is very simple, given by

H¯[Gv¯mG].H{\underline{{\mathbb{Z}}}}[G\cdot{\bar{v}}_{m}^{G}].

3. Chromatic Height of BP((G))m,mBP^{(\!(G)\!)}\langle m,m\rangle

In this section, we study the underlying chromatic height of the spectra BP((G))m,mBP^{(\!(G)\!)}\langle m,m\rangle.

Theorem 3.1.
  1. (1)

    For r=kmr=km where 0k2n10\leq k\leq 2^{n-1}, LK(r)ieBP((G))m,m≄L_{K(r)}i_{e}^{*}BP^{(\!(G)\!)}\langle m,m\rangle\not\simeq*;

  2. (2)

    For all other r0r\geq 0, LK(r)ieBP((G))m,mL_{K(r)}i_{e}^{*}BP^{(\!(G)\!)}\langle m,m\rangle\simeq*.

Proof.

Our proof will be similar to that of Proposition 7.4 and Theorem 7.5 in [7]. In this proof, let X=ieBP((G))m,mX=i_{e}^{*}BP^{(\!(G)\!)}\langle m,m\rangle. For any rr, there is a cofinal sequence J(i)=(j0,j1,,jr1)J(i)=(j_{0},j_{1},\ldots,j_{r-1}) of positive integers and generalized Moore spectra

MJ(i)=S0/(v0j0,,vr1jr1)M_{J(i)}=S^{0}/(v_{0}^{j_{0}},\ldots,v_{r-1}^{j_{r-1}})

with maps MJ(i+1)MJ(i)M_{J(i+1)}\to M_{J(i)} such that

LK(r)Xholimi(LrXMJ(i)).L_{K(r)}X\simeq\text{holim}_{i}\left(L_{r}X\wedge M_{J(i)}\right).

See [31, Prop. 7.10].

Since XX is a BPBP-module, it follows from [30, Cor. 1.10] that the natural map LrfXLrXL_{r}^{f}X\to L_{r}X is an equivalence (since it is a BPBP-equivalence between BPBP-local spectra). Therefore,

LrXMJ(i)LrfXMJ(i)XLrfMJ(i)Xvr1MJ(i)vr1XMJ(i).L_{r}X\wedge M_{J(i)}\simeq L_{r}^{f}X\wedge M_{J(i)}\simeq X\wedge L_{r}^{f}M_{J(i)}\simeq X\wedge v_{r}^{-1}M_{J(i)}\simeq v_{r}^{-1}X\wedge M_{J(i)}.

Here, we have used the fact that since MJ(i)M_{J(i)} is a type rr spectrum, its finite localization is the telescope [37, Prop. 3.2]. We have also used the fact that LrfL_{r}^{f} is smashing.

To prove (1), we assume rr is of the form kmkm with 0k2n10\leq k\leq 2^{n-1}. We will first show that under the map

BPvr1XMJ(i),BP\longrightarrow v_{r}^{-1}X\wedge M_{J(i)},

the image of vrπBPv_{r}\in\pi_{*}BP is nonzero. Note that

vr1XMJ(i)=vr1XMUMU/(v0j0,,vr1jr1)=vr1X/(v0j0,,vr1jr1).v_{r}^{-1}X\wedge M_{J(i)}=v_{r}^{-1}X\wedge_{MU}MU/(v_{0}^{j_{0}},\ldots,v_{r-1}^{j_{r-1}})=v_{r}^{-1}X/(v_{0}^{j_{0}},\ldots,v_{r-1}^{j_{r-1}}).

By an iterative application of the formula

vrC2n1vrC2n+γnvrC2n+j=1r1γnvjC2n(vrjC2n)2j(modIr)v_{r}^{C_{2^{n-1}}}\equiv v_{r}^{C_{2^{n}}}+\gamma_{n}v_{r}^{C_{2^{n}}}+\sum_{j=1}^{r-1}\gamma_{n}v_{j}^{C_{2^{n}}}(v_{r-j}^{C_{2^{n}}})^{2^{j}}\pmod{I_{r}}

(where Ir=(2,v1,,vr1)I_{r}=(2,v_{1},\ldots,v_{r-1})) in [7, Theorem 1.1], the images of vrj=(vrC2)jv_{r}^{j}=(v_{r}^{C_{2}})^{j} in (πX)/(v0,,vr1)(\pi_{*}X)/(v_{0},\ldots,v_{r-1}) are all nonzero for j1j\geq 1. This implies that their images are also nonzero in π(X/(v0,,vr1))\pi_{*}\left(X/(v_{0},\ldots,v_{r-1})\right). Therefore, the image of vrv_{r} in π(vr1X/(v0j0,,vr1jr1))\pi_{*}(v_{r}^{-1}X/(v_{0}^{j_{0}},\ldots,v_{r-1}^{j_{r-1}})) is nonzero. After taking the homotopy limit, the image of vrv_{r} under the map πBPπLK(r)X\pi_{*}BP\to\pi_{*}L_{K(r)}X will also be nonzero. It follows that πLK(r)X≄\pi_{*}L_{K(r)}X\not\simeq*.

To prove (2), we will consider two cases, based on the divisibility of rr by mm. If rr is not divisible by mm, then the degree of vrv_{r}, 2(2r1)2(2^{r}-1), is not divisible by 2(2m1)2(2^{m}-1). However, the homotopy groups of XX are concentrated in degrees that are divisible by 2(2m1)2(2^{m}-1). This implies that the multiplication by vrv_{r} map

Σ|vr|XX\Sigma^{|v_{r}|}X\longrightarrow X

induces the zero map on homotopy, and

πvr1Xvr1πX=0.\pi_{*}v_{r}^{-1}X\cong v_{r}^{-1}\pi_{*}X=0.

It follows that vr1Xv_{r}^{-1}X\simeq* and therefore LK(r)XL_{K(r)}X\simeq*.

Now, suppose mm divides rr. Let r=kmr=km for some k>2n1k>2^{n-1}. The result of [7, Proposition 7.3] implies that vr(2,v1,,vr1)v_{r}\in(2,v_{1},\ldots,v_{r-1}) so that vrq(v0j0,,vr1jr1)v_{r}^{q}\in(v_{0}^{j_{0}},\ldots,v_{r-1}^{j_{r-1}}) for some q>0q>0. Now,

XMJ(i)=XMUMU/(v0j0,,vr1jr1)X\wedge M_{J(i)}=X\wedge_{MU}MU/(v_{0}^{j_{0}},\ldots,v_{r-1}^{j_{r-1}})

and there is a Künneth spectral sequence [14, Theorem IV.4.1]

E2s,t=Tors,tMU(πX,πMU/(v0j0,,vr1jr1))πts(XMJ(i)).E_{2}^{s,t}=\operatorname{Tor}_{-s,t}^{MU_{*}}(\pi_{*}X,\pi_{*}MU/(v_{0}^{j_{0}},\ldots,v_{r-1}^{j_{r-1}}))\Longrightarrow\pi_{t-s}\left(X\wedge M_{J(i)}\right).

This is a cohomologically graded lower half-plane spectral sequence. As in the proof of Theorem 7.5(2) in [7], the fact that for some qq multiplication by vrqv_{r}^{q} raises filtration implies that every element in the homotopy groups of XMJ(i)X\wedge M_{J(i)} is killed by some finite power of vrv_{r}. It follows that LK(r)X=holimi(vr1XMJ(i))L_{K(r)}X=\text{holim}_{i}\left(v_{r}^{-1}X\wedge M_{J(i)}\right)\simeq*. ∎

4. Localized spectral sequences

In our computations below, we will make use of various localizations of the slice spectral sequence of quotients of MU((G))MU^{(\!(G)\!)}. In this section, we recall results from [24] and [39] that we will use here. As a reminder, we continue to let G=C2nG=C_{2^{n}}.

4.1. Some notation

Here, we introduce some notation. We refer the reader to [25] for more details.

Consider 22-local homotopy equivalence classes of representation spheres SVS^{V} where VV is a finite dimensional orthogonal representation. This is a semi-group with respect to the smash product. Let JO(G)JO(G) be the group completion.

Definition 4.1.

Define λj=λj(G)\lambda_{j}=\lambda_{j}(G) to be the 22-dimensional irreducible real representation of GG for which the generator γG\gamma\in G acts on 2{\mathbb{R}}^{2} by a rotation by 2π/2nj2\pi/2^{n-j}. We also have the one-dimensional sign representation σn=σ(G)\sigma_{n}=\sigma(G), for which the generator acts by multiplication by 1-1.

Note that λn1=2σn\lambda_{n-1}=2\sigma_{n} . There is an isomorphism of underlying abelian groups

JO(G){1,σn,λ0,,λn2}JO(G)\cong{\mathbb{Z}}\{1,\sigma_{n},\lambda_{0},\ldots,\lambda_{n-2}\}

where the equivalence sends SVS^{V} to VV.

Definition 4.2.

For each representation VV, there is a homotopy class

aV:S0SVa_{V}\colon S^{0}\to S^{V}

which corresponds to the inclusion of S0={0,}S^{0}=\{0,\infty\}. We call this the Euler class.

If VV is an orientable representation of dimension dd, we also get classes

uVπdVH¯.u_{V}\in\pi_{d-V}H\underline{{\mathbb{Z}}}.

We call these orientation classes.

We have commutative diagrams

S0\textstyle{S^{0}\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}aλi\scriptstyle{a_{\lambda_{i}}}aλi+1\scriptstyle{a_{\lambda_{i+1}}}Sλi\textstyle{S^{\lambda_{i}}\ignorespaces\ignorespaces\ignorespaces\ignorespaces}()2\scriptstyle{(-)^{2}}Sλi+1\textstyle{S^{\lambda_{i+1}}}

where the vertical arrow is a double cover. Therefore, aλia_{\lambda_{i}} divides aλi+1a_{\lambda_{i+1}} for each 0in20\leq i\leq n-2.

4.2. Localizations and isotropy separation

Definition 4.3.

For each 0in10\leq i\leq n-1, we have families

i=[C2i+1]={HHC2i+1}\mathcal{F}_{i}=\mathcal{F}[C_{2^{i+1}}]=\{H\mid H\subsetneq C_{2^{i+1}}\}

and n=𝒜ll\mathcal{F}_{n}=\mathcal{A}ll be the family of all subgroups of GG.

These families interpolate between 0={e}\mathcal{F}_{0}=\{e\} and n=𝒜ll\mathcal{F}_{n}=\mathcal{A}ll.

The universal and couniversal spaces for the family i\mathcal{F}_{i} can be written in very algebraic terms.

Proposition 4.4.

For 0in10\leq i\leq n-1, we have

EilimS(kλi)=S(λi)E\mathcal{F}_{i}\simeq\lim_{\to}S(k\lambda_{i})=S(\infty\lambda_{i})

and

E~iSλi=S0[aλi1].\widetilde{E}\mathcal{F}_{i}\simeq S^{\infty\lambda_{i}}=S^{0}[a_{\lambda_{i}}^{-1}].
Proof.

The representation λi\lambda_{i} has kernel exactly C2iGC_{2^{i}}\subset G, and the residual action of G/C2iG/C_{2^{i}} is faithful. The result follows. ∎

We now state two results of Meier–Shi–Zeng that we will use later.

Theorem 4.5 (Meier–Shi–Zeng [39]).

For XX a GG-spectrum with regular slice tower PXP^{\bullet}X, the spectral sequence associated to the tower E~iPX\widetilde{E}\mathcal{F}_{i}\wedge P^{\bullet}X, which corresponds to the aλia_{\lambda_{i}}-localized spectral sequence

aλi1E2s,t+α=aλi1πts+αGPttXπts+αG(aλi1X),αJO(G)a_{\lambda_{i}}^{-1}E_{2}^{s,t+\alpha}=a_{\lambda_{i}}^{-1}\pi_{t-s+\alpha}^{G}P_{t}^{t}X\Longrightarrow\pi_{t-s+\alpha}^{G}(a_{\lambda_{i}}^{-1}X),\quad\quad\alpha\in JO(G)

converges strongly.

Theorem 4.6 (Meier–Shi–Zeng [38]).

Let XX be a (1)(-1)-connected GG-spectrum. Let LiL_{i} be the line of slope (2i1)(2^{i}-1) through the origin. The following statements hold:

  1. (1)

    On the integer graded page, the map from the slice spectral sequence of XX to the aλia_{\lambda_{i}}-localized slice spectral sequence of XX induces an isomorphism on the E2E_{2}-page for the classes above LiL_{i}, and a surjection for the classes that are on LiL_{i}.

  2. (2)

    The map of spectral sequences above induces an isomorphism between differentials that originate from classes that are on or above LiL_{i}.

In what follows, we will compute heavily with localized slice spectral sequences. The following remark explains the advantages of this approach.

Remark 4.7.

It follows from Theorem 4.6 that all the differentials in the slice spectral sequence of XX that are on or above LiL_{i} can be immediately recovered from the aλia_{\lambda_{i}}-localized slice spectral sequence of XX by truncating off the latter spectral sequence below LiL_{i}. In particular, all the differentials in the slice spectral sequence of XX can be recovered by truncating off the aλ0a_{\lambda_{0}}-inverted slice spectral sequence below the horizontal line s=0s=0.

In addition, the aλia_{\lambda_{i}}-localized slice spectral sequence are individually easier to compute that the non-localized spectral sequences. Of course, by localizing, we loose the information below the line LiL_{i}, but the approach is to work inductively, starting with the aλn1a_{\lambda_{n-1}}-localization (which is the same as the aσna_{\sigma_{n}}-localization) and ending with the aλ0a_{\lambda_{0}}-localization. As we explained above, all differentials can be recovered from the aλ0a_{\lambda_{0}} localization so that at that stage, we have not actually lost any information at all.

One final remark on the advantage of computing with the aλ0a_{\lambda_{0}}-localized spectral sequence is that it actually records information about the slice spectral sequences

E2s,t+=πts+GPttXπts+GXE_{2}^{s,t+\star}=\pi_{t-s+\star}^{G}P_{t}^{t}X\Rightarrow\pi_{t-s+\star}^{G}X

for any =kλ0\star=k\lambda_{0} where kk\in{\mathbb{Z}}. Indeed, we can recover all the differentials in the +kλ0*+k\lambda_{0}-graded spectral sequence by truncating the aλ0a_{\lambda_{0}}-localized spectral sequence below the horizontal line s=2ks=-2k. So, the localized spectral sequence contains much more information than simply the integer graded spectral sequence.

5. C2nC_{2^{n}}-geometric fixed points and quotients of BP((G))BP^{(\!(G)\!)}

As a proof-of-concept and for later computations, in this section, we will compute the homotopy of the geometric fixed points

ΦG(BP((G))/GS¯)(E~𝒫BP((G))/GS¯)G(aσ1BP((G))/GS¯)G\Phi^{G}(BP^{(\!(G)\!)}/G\cdot\bar{S})\simeq(\widetilde{E}\mathcal{P}\wedge BP^{(\!(G)\!)}/G\cdot\bar{S})^{G}\simeq(a_{\sigma}^{-1}BP^{(\!(G)\!)}/G\cdot\bar{S})^{G}

of quotients by permutation summands via the aσa_{\sigma}-localized slice spectral sequence.

On the one hand, we know the answer, since we know the homotopy type of the geometric fixed points.

Proposition 5.1.

We have a weak-equivalence of H𝔽2H{\mathbb{F}}_{2}-modules

ΦG(BP((G))/GS¯)H𝔽2Σ+(jJS2j).\Phi^{G}\big{(}BP^{(\!(G)\!)}/G\cdot\bar{S}\big{)}\simeq H{\mathbb{F}}_{2}\wedge\Sigma^{\infty}_{+}\left(\prod_{j\in J}S^{2^{j}}\right).
Proof.

The geometric fixed points functor is strong symmetric monoidal, and we have

ΦG(BP((G)))=H𝔽2.\Phi^{G}(BP^{(\!(G)\!)})=H{\mathbb{F}}_{2}.\qed

On the other hand, the aσa_{\sigma}-localization map of the slice spectral has a particular simple target, and this will tell us a great deal about any of the slice spectral sequences for these quotients.

5.1. General quotients of BP((G))BP^{(\!(G)\!)}

We now consider the aσa_{\sigma}-localized slice spectral sequence

aσ1E2s,t:=aσ1πtsGPtt(BP((G))/GS¯)πtsGaσ1(BP((G))/GS¯).a_{\sigma}^{-1}E_{2}^{s,t}:=a_{\sigma}^{-1}\pi_{t-s}^{G}P_{t}^{t}(BP^{(\!(G)\!)}/G\cdot\bar{S})\Longrightarrow\pi_{t-s}^{G}a_{\sigma}^{-1}(BP^{(\!(G)\!)}/G\cdot\bar{S}).

Inverting aσa_{\sigma} has the effect of killing the transfer from any proper subgroups. This means that the E2E_{2}-page of the aσa_{\sigma}-localizaed slice spectral sequence has a particular simple form:

aσ1E2,=𝔽2[NC2Gs¯iiJ][u2σ,aσ±1,aλ0±1,,aλn2±1].a_{\sigma}^{-1}E_{2}^{*,\star}=\mathbb{F}_{2}[N_{C_{2}}^{G}{\bar{s}}_{i}\mid i\not\in J][u_{2\sigma},a_{\sigma}^{\pm 1},a_{\lambda_{0}}^{\pm 1},\ldots,a_{\lambda_{n-2}}^{\pm 1}]\ .
Definition 5.2.

For each jj\in\mathbb{N}, let

f¯j=aρ¯2j1NC2Gs¯j.\bar{f}_{j}=a_{\bar{\rho}}^{2^{j}-1}N_{C_{2}}^{G}\bar{s}_{j}.

This definition a priori depends heavily on the choices of the s¯i\bar{s}_{i}. However, from the point of view of differentials, these choices will not matter, due to a small lemma.

Lemma 5.3.

Let s¯2m1\bar{s}_{2^{m}-1} be any element in degree (2m1)ρ2(2^{m}-1)\rho_{2} that generates a permutation summand. We have

NC2Gs¯2m1NC2Gv¯mGmod(NC2Gv¯1G,,NC2Gv¯m1G)+Im(tr),N_{C_{2}}^{G}\bar{s}_{2^{m}-1}\equiv N_{C_{2}}^{G}{\bar{v}}_{m}^{G}\mod(N_{C_{2}}^{G}{\bar{v}}_{1}^{G},\dots,N_{C_{2}}^{G}{\bar{v}}_{m-1}^{G})+Im(tr),

where Im(tr)Im(tr) denotes the image of the transfer.

In particular, f¯i\bar{f}_{i} is independent of the choice of s¯i\bar{s}_{i}, modulo the lower v¯j\bar{v}_{j}.

Proof.

We have

πρGG(BP((G)))/Im(tr)𝔽2[NC2Gv¯1G,NC2Gv¯2G,],\pi_{\ast\rho_{G}}^{G}\big{(}BP^{(\!(G)\!)}\big{)}/Im(tr)\cong{\mathbb{F}}_{2}\big{[}N_{C_{2}}^{G}\bar{v}_{1}^{G},N_{C_{2}}^{G}\bar{v}_{2}^{G},\dots\big{]},

and the map

xNC2Gxx\mapsto N_{C_{2}}^{G}x

gives a ring homomorphism

πρ2C2BP((G))πρGG(BP((G)))/Im(tr).\pi_{\ast\rho_{2}}^{C_{2}}BP^{(\!(G)\!)}\to\pi_{\ast\rho_{G}}^{G}\big{(}BP^{(\!(G)\!)}\big{)}/Im(tr).

Additionally, Weyl equivariance of the norm shows that for any γG\gamma\in G,

NC2G(γx)=γNC2G(x)NC2G(x)modIm(tr).N_{C_{2}}^{G}(\gamma x)=\gamma N_{C_{2}}^{G}(x)\equiv N_{C_{2}}^{G}(x)\mod Im(tr).

The lemma can be restated as saying that for any generator

s¯2m1Q(2m1)ρ2(πρ2C2BP((G))),\bar{s}_{2^{m}-1}\in Q_{(2^{m}-1)\rho_{2}}\big{(}\pi_{\ast\rho_{2}}^{C_{2}}BP^{(\!(G)\!)}\big{)},

we have

NC2Gs¯2m1=NC2Gv¯mGQ(2m1)ρG(𝔽2[NC2Gv¯1G,]).N_{C_{2}}^{G}\bar{s}_{2^{m}-1}=N_{C_{2}}^{G}\bar{v}_{m}^{G}\in Q_{(2^{m}-1)\rho_{G}}\big{(}{\mathbb{F}}_{2}[N_{C_{2}}^{G}\bar{v}_{1}^{G},\dots]\big{)}.

The above argument shows that the norm is a Weyl-equivariant ring homomorphism, and hence it induces a linear map

(Q(2m1)ρ2(πρ2C2BP((G))))GQ(2m1)ρG(𝔽2[NC2Gv¯1G,]).\Big{(}Q_{(2^{m}-1)\rho_{2}}\big{(}\pi_{\ast\rho_{2}}^{C_{2}}BP^{(\!(G)\!)}\big{)}\Big{)}_{G}\to Q_{(2^{m}-1)\rho_{G}}\big{(}{\mathbb{F}}_{2}\big{[}N_{C_{2}}^{G}\bar{v}_{1}^{G},\dots\big{]}\big{)}.

Both the source and target are isomorphic to 𝔽2{\mathbb{F}}_{2}, and choosing v¯mG\bar{v}_{m}^{G} as the generator of the source shows the map to be non-zero. It is therefore non-zero on any generator for the source. ∎

Corollary 5.4.

The E2E_{2}-term for the aσa_{\sigma}-localized slice spectral sequence for BP((G))BP^{(\!(G)\!)} is given by

𝔽2[aσ±1,aλn2±1,,aλ0±1][b,f¯1,],{\mathbb{F}}_{2}[a_{\sigma}^{\pm 1},a_{\lambda_{n-2}}^{\pm 1},\dots,a_{\lambda_{0}}^{\pm 1}][b,\bar{f}_{1},\dots],

where the bidegree of f¯i\bar{f}_{i} is (2i1,(2n1)(2i1))\big{(}2^{i}-1,(2^{n}-1)(2^{i}-1)\big{)} and where the bidegree of

b=u2σ/aσ2b=u_{2\sigma}/a_{\sigma}^{2}

is (2,2)(2,-2).

Corollary 5.5.

For any S¯\bar{S}, the aσa_{\sigma}-localized slice spectral sequence for BP((G))/GS¯BP^{(\!(G)\!)}/G\cdot\bar{S} is a module over that for BP((G))BP^{(\!(G)\!)}, and the E2E_{2}-term is the quotient

𝔽2[aσ±1,aλn2±1,,aλ0±1][b,f¯1,]/(f¯jjJ){\mathbb{F}}_{2}[a_{\sigma}^{\pm 1},a_{\lambda_{n-2}}^{\pm 1},\dots,a_{\lambda_{0}}^{\pm 1}][b,\bar{f}_{1},\dots]/\big{(}\bar{f}_{j}\mid j\in J\big{)}

We start with examining the aσa_{\sigma}-localized slice spectral sequence for BP((G))BP^{(\!(G)\!)}, since all other cases are modules over this.

Proposition 5.6.

Let s¯i\bar{s}_{i} for ii\in\mathbb{N} be any choice of permutation summand generators for πρ2C2BP((G))\pi_{*\rho_{2}}^{C_{2}}BP^{(\!(G)\!)}.

Then in the aσa_{\sigma}-localized slice spectral sequence for BP((G))BP^{(\!(G)\!)}, the differentials are determined by

d(k)(b2k)=f¯k+1,(k):=2n(2k+11)+1,k0.d_{\ell(k)}(b^{2^{k}})=\bar{f}_{k+1},\quad\ell(k):=2^{n}(2^{k+1}-1)+1,\quad k\geq 0. (5.1)
Proof.

Lemma 5.3 shows that

f¯jaρ¯(2j1)NC2Gv¯jG\bar{f}_{j}\equiv a_{\bar{\rho}}^{(2^{j}-1)}N_{C_{2}}^{G}{\bar{v}}_{j}^{G}

modulo the earlier generators f¯i\bar{f}_{i} with i<ji<j, so the Slice Differentials Theorem of [22] implies that we have the differentials (5.1). ∎

Remark 5.7.

The f¯i\bar{f}_{i} all lie on the line of slope 2n12^{n}-1 through the origin in the (ts,s)(t-s,s)-plane. This is a vanishing line for both the spectral sequence of aσ1BP((G))a_{\sigma}^{-1}BP^{(\!(G)\!)} and that of quotient by permutation summands, so the differential in (5.1) is the last possible on b2nb^{2^{n}}.

We next use this result to study the aσa_{\sigma}-localized slice spectral sequence of other quotients.

Definition 5.8.

Let A(J)A(J) be the set of non-negative integers rr such that the dyadic expansion r=ε0+ε12+ε24+r=\varepsilon_{0}+\varepsilon_{1}\cdot 2+\varepsilon_{2}\cdot 4+\ldots satisfies εi=0\varepsilon_{i}=0 if i+1Ji+1\not\in J.

Theorem 5.9.

The aσa_{\sigma}-localized slice spectral sequence of aσ1(BP((G))/GS¯)a_{\sigma}^{-1}(BP^{(\!(G)\!)}/G\cdot\bar{S}) can be completely described as follows:

  1. (1)

    the only non-trivial differentials are of lengths (k)=2n(2k+11)+1\ell(k)=2^{n}(2^{k+1}-1)+1 for some k0k\geq 0.

  2. (2)

    The E(k)E_{\ell(k)}-page is the module over

    𝔽2[f¯iik+1andiJ][b2k]{\mathbb{F}}_{2}[\bar{f}_{i}\mid i\geq k+1\ \text{and}\ i\not\in J][b^{2^{k}}]

    generated by the set of permanent cycles brb^{r} where 0r<2k0\leq r<2^{k} and rA(J)r\in A(J).

  3. (3)

    If k+1Jk+1\not\in J, then there are non-trivial differentials are multiples of

    d(k)(b2k)=f¯k+1d_{\ell(k)}(b^{2^{k}})=\bar{f}_{k+1}

    by the d(k)d_{\ell(k)}-cycles

    𝔽2[f¯iik+1andiJ][b2k+1]{brr<2kandrA(J)}.{\mathbb{F}}_{2}[\bar{f}_{i}\mid i\geq k+1\ \text{and}\ i\not\in J][b^{2^{k+1}}]\{b^{r}\mid r<2^{k}\ \text{and}\ r\in A(J)\}.

    There are no other differentials of that length.

  4. (4)

    If k+1Jk+1\in J, then E(k)=E(k+1)E_{\ell(k)}=E_{\ell(k+1)}.

  5. (5)

    Consequently, E𝔽2{brrA(J)}E_{\infty}\cong{\mathbb{F}}_{2}\{b^{r}\mid r\in A(J)\} .

Proof.

We will prove the statements by induction on kk. For the base case when k=0k=0, we have (0)=2n+1\ell(0)=2^{n}+1. The first possible non-trivial differential by sparseness is d(0)(b)=f¯1d_{\ell(0)}(b)=\bar{f}_{1}. Therefore, E(0)=E2E_{\ell(0)}=E_{2} and the class b0b^{0} is a permanent cycle. The claims hold.

Now, suppose that the E(k)E_{\ell(k)}-page is as claimed. If k+1Jk+1\in J, then b2kb^{2^{k}} is a d(k)d_{\ell(k)}-cycle, and hence a permanent cycle. Any element on the E(k)E_{\ell(k)}-page of the form brb^{r} with r<2k+1r<2^{k+1} is of the form r+εk2kr^{\prime}+\varepsilon_{k}\cdot 2^{k} for r<2kr^{\prime}<2^{k} with rA(J)r^{\prime}\in A(J) and εk{0,1}\varepsilon_{k}\in\{0,1\}. Since k+1Jk+1\in J, rA(J)r\in A(J). Using the module structure over the aσa_{\sigma}-localized spectral sequence of aσ1BP((G))a_{\sigma}^{-1}BP^{(\!(G)\!)}, the elements brb^{r} are also d(k)d_{\ell(k)}-cycles. By sparseness of the E(k)E_{\ell(k)}-page, it is impossible for brb^{r} to support a differential of length longer than (k)\ell(k) because there are no possible targets. Therefore, the elements brb^{r} are permanent cycles. This implies that E(k)=E(k+1)E_{\ell(k)}=E_{\ell(k+1)} which proves our claims when k+1Jk+1\in J.

On the other hand, if k+1Jk+1\not\in J, we have a non-trivial differential d(k)(b2k)=f¯k+1d_{\ell(k)}(b^{2^{k}})=\bar{f}_{k+1}. For α1\alpha\geq 1, consider the element f¯f¯k+1αbr+2kt\bar{f}\bar{f}_{k+1}^{\alpha}b^{r+2^{k}t} with 0r<2k0\leq r<2^{k}, t0t\geq 0 even, and f¯\bar{f} a monomial in the f¯i\bar{f}_{i}’s for i>k+1i>k+1 and iJi\not\in J. Such an element is the target of the d(k)d_{\ell(k)}-differential on f¯f¯k+1α1br+2k(t+1)\bar{f}\bar{f}_{k+1}^{\alpha-1}b^{r+2^{k}(t+1)}. It follows that the E(k)+1E_{\ell(k)+1}-page is a polynomial algebra over

𝔽2[f¯iik+2andiJ][b2k+1]{\mathbb{F}}_{2}[\bar{f}_{i}\mid i\geq k+2\ \text{and}\ i\not\in J][b^{2^{k+1}}]

generated by the already established permanent cycles brb^{r}, r<2kr<2^{k}. Since k+1Jk+1\not\in J, the set {rA(J)r<2k+1}\{r\in A(J)\mid r<2^{k+1}\} is equal to {rA(J)r<2k}\{r\in A(J)\mid r<2^{k}\}. This completes the induction step. ∎

As an immediate consequence, using Theorem 4.6, we have:

Theorem 5.10.

In the integer graded slice spectral sequence

E2s,t=πtsGPttBP((G))/GS¯πtsBP((G))/GS¯,E_{2}^{s,t}=\pi_{t-s}^{G}P^{t}_{t}BP^{(\!(G)\!)}/G\cdot\bar{S}\Rightarrow\pi_{t-s}BP^{(\!(G)\!)}/G\cdot\bar{S},

we have:

  1. (1)

    Above the line of slope 2n112^{n-1}-1, the E2E_{2}-page is isomorphic to the integer graded part of the E2E_{2}-page of the aσa_{\sigma}-localization of the spectral sequence, as described in Corollary 5.5.

  2. (2)

    The only non-trivial differentials whose sources lie on or above the line of slope 2n112^{n-1}-1 are in one-to-one correspondence with the non-trivial differentials of the integer graded aσa_{\sigma}-localized slice spectral sequence above this line. They are of lengths (k)\ell(k) for k+1Jk+1\not\in J, and are generated under the module structure of the spectral sequence for BP((G))BP^{(\!(G)\!)} by the differentials

    d(k)(u2σ2k)=aσ2k+1aρ¯2k+11NC2Gs¯2k+11d_{\ell(k)}(u_{2\sigma}^{2^{k}})=a_{\sigma}^{2^{k+1}}a_{\bar{\rho}}^{2^{k+1}-1}N_{C_{2}}^{G}\bar{s}_{2^{k+1}-1}

    for k+1Jk+1\not\in J.

We will now give some examples to illustrate the results above. We start with example for the group G=C2G=C_{2}. Since v¯jv¯k+1\bar{v}_{j}\equiv\bar{v}_{k+1} modulo the previous v¯j\bar{v}_{j}’s, we write v¯k+1=s¯2k+11\bar{v}_{k+1}=\bar{s}_{2^{k+1}-1}, but any choice of permutation summand generators gives the same results.

Example 5.11.

Consider BP/S¯BP_{{\mathbb{R}}}/\bar{S} for S¯={s¯2j1jJ}\bar{S}=\{\bar{s}_{2^{j}-1}\mid j\in J\}. The E2E_{2}-page are the aσa_{\sigma}-localized slice spectral sequence is

aσ1E2,=𝔽2[v¯iiJ][u2σ,aσ±1],a_{\sigma}^{-1}E_{2}^{*,\star}=\mathbb{F}_{2}[\bar{v}_{i}\mid i\not\in J][u_{2\sigma},a_{\sigma}^{\pm 1}],

the non-trivial differentials are generated by d2k+21(b2k)=f¯k+1d_{2^{k+2}-1}(b^{2^{k}})=\bar{f}_{k+1} and

E,𝔽2[aσ±1]{brrA(J)}.E_{\infty}^{*,\star}\cong{\mathbb{F}}_{2}[a_{\sigma}^{\pm 1}]\{b^{r}\mid r\in A(J)\}.

Figure 1 shows the example for BP2BP_{{\mathbb{R}}}\langle 2\rangle.

Refer to caption
Figure 1. The aσa_{\sigma}-localized slice spectral sequence of aσ1BP2a_{\sigma}^{-1}BP_{{\mathbb{R}}}\langle 2\rangle in integer degrees. The slice spectral sequence of BP2BP_{{\mathbb{R}}}\langle 2\rangle is obtained by removing the region below the horizontal line s=0s=0 and replacing =/2\bullet={\mathbb{Z}}/2 by copies of \mathbb{Z}, which reintroduces the transfers.

In this case, using Theorem 4.6 as in Theorem 5.10 together with the fact that the aσa_{\sigma}-localized spectral sequence records information about many JO(C2)JO(C_{2}) degrees of the slices spectral sequence (as noted in Remark 4.7), we can easily describe a large part of the (unlocalized) JO(C2)JO(C_{2})-graded slice spectral sequence of BP/S¯BP_{{\mathbb{R}}}/\bar{S}.

Corollary 5.12.

Let JO(C2)+JO(C2)JO(C_{2})^{+}\subseteq JO(C_{2}) be the elements of the form a+bσa+b\sigma with ab0a-b\geq 0. The JO(C2)+JO(C_{2})^{+} graded slice E2E_{2}-page of BP/S¯BP_{\mathbb{R}}/\bar{S} is

E2,(2)[v¯iiJ][u2σ,aσ]/(2aσ),JO(C2)+.\displaystyle E_{2}^{*,\star}\cong{\mathbb{Z}}_{(2)}[\bar{v}_{i}\mid i\not\in J][u_{2\sigma},a_{\sigma}]/(2a_{\sigma}),\quad\star\in JO(C_{2})^{+}\ . (5.2)
  1. (1)

    The only non-trivial differentials are of lengths (2k+21)(2^{k+2}-1) for some k0k\geq 0.

  2. (2)

    If k+1Jk+1\notin J, then the nontrivial-differentials are multiples of

    d2k+21(u2σ2k+r)=v¯k+1u2σraσ2k+21,rA(J)d_{2^{k+2}-1}(u_{2\sigma}^{2^{k}+r})=\bar{v}_{k+1}u_{2\sigma}^{r}a_{\sigma}^{2^{k+2}-1},\,\,\,r\in A(J)

    by the d2k+21d_{2^{k+2}-1}-cycles 𝔽2[v¯i,aσ|ik+1 and iJ]\mathbb{F}_{2}[\bar{v}_{i},a_{\sigma}\,|\,i\geq k+1\text{ and }i\notin J].

  3. (3)

    If k+1Jk+1\in J, then E2k+21=E2k+31.E_{2^{k+2}-1}=E_{2^{k+3}-1}.

  4. (4)

    Consequently, the EE_{\infty}-page is

    E,(2)[v¯iiJ][aσ]/(2aσ,v¯iaσ2i+11){u2σr,2u2σsrA(J),sA(J)}.E_{\infty}^{*,\star}\cong\mathbb{Z}_{(2)}[\bar{v}_{i}\mid i\not\in J][a_{\sigma}]/(2a_{\sigma},\bar{v}_{i}a_{\sigma}^{2^{i+1}-1})\{u_{2\sigma}^{r},2u_{2\sigma}^{s}\mid r\in A(J),s\notin A(J)\}.

As an explicit example, we show the computation of the slice spectral sequences of BP2,2BP_{{\mathbb{R}}}\langle 2,2\rangle, deduced from that of aσ1BP2,2a_{\sigma}^{-1}BP_{{\mathbb{R}}}\langle 2,2\rangle. The computation is illustrated in Figure 2.

Remark 5.13.

On the EE_{\infty}-page of BP2,2BP_{{\mathbb{R}}}\langle 2,2\rangle (the third picture of Figure 2), there is an exotic extension η=v¯1aσ\eta=\bar{v}_{1}a_{\sigma}, as shown by the dashed line. This extension follows from the work in [8]. More precisely, letting (n,k,b)=(2,1,0)(n,k,b)=(2,1,0) in Corollary 3.11 of [8] gives the exotic πMU\pi_{\star}MU_{{\mathbb{R}}}-multiplication v¯1u2σ=v¯2aσ4\bar{v}_{1}u_{2\sigma}=\bar{v}_{2}a_{\sigma}^{4}. It follows that

(v¯2aσu2σ)(v¯1aσ)=(v¯2a2σ)(v¯1u2σ)=(v¯2a2σ)(v¯2a4σ)=v¯22aσ6.(\bar{v}_{2}a_{\sigma}u_{2\sigma})\cdot(\bar{v}_{1}a_{\sigma})=(\bar{v}_{2}a_{2\sigma})\cdot(\bar{v}_{1}u_{2\sigma})=(\bar{v}_{2}a_{2\sigma})\cdot(\bar{v}_{2}a_{4\sigma})=\bar{v}_{2}^{2}a_{\sigma}^{6}.
Refer to caption
Refer to caption
Refer to caption
Figure 2. The aσa_{\sigma}-localized slice spectral sequences of aσ1BP2,2a_{\sigma}^{-1}BP_{{\mathbb{R}}}\langle 2,2\rangle (top). The middle figure is the slice spectral sequence of BP2,2BP_{{\mathbb{R}}}\langle 2,2\rangle and the bottom is its EE_{\infty}-page. A \square denotes (2){\mathbb{Z}}_{(2)}, a \bullet denotes /2{\mathbb{Z}}/2.

For the next two examples, the group G=C4G=C_{4} with generators s¯2i1=v¯iG\bar{s}_{2^{i}-1}=\bar{v}_{i}^{G}, but any choice of permutation summand generators gives the same results.

Example 5.14.

Consider BP((C4))2BP^{(\!(C_{4})\!)}\langle 2\rangle, so that J={jj>2}J=\{j\in{\mathbb{N}}\mid j>2\}. The E2E_{2}-page are the aσa_{\sigma}-localized slice spectral sequence is

aσ1E2,𝔽2[f¯1,f¯2,b].a_{\sigma}^{-1}E_{2}^{*,*}\cong{\mathbb{F}}_{2}[\bar{f}_{1},\bar{f}_{2},b]\ .

The spectral sequence has two types of differentials, namely

d5(b)=f¯1,andd13(b2)=f¯2.d_{5}(b)=\bar{f}_{1},\quad\text{and}\quad d_{13}(b^{2})=\bar{f}_{2}\ .

The class b4b^{4} is a permanent cycle, and we have

πC4aσ1BP((C4))2𝔽2[b4].\pi_{*}^{C_{4}}a_{\sigma}^{-1}BP^{(\!(C_{4})\!)}\langle 2\rangle\cong{\mathbb{F}}_{2}[b^{4}].

The computation is illustrated in Figure 3.

Refer to caption
Refer to caption
Figure 3. The aσa_{\sigma}-localized slice spectral sequence of aσ1BP((C4))2a_{\sigma}^{-1}BP^{(\!(C_{4})\!)}\langle 2\rangle.
Example 5.15.

Consider BP((C4))2,2BP^{(\!(C_{4})\!)}\langle 2,2\rangle. In this case, the E2E_{2}-page are the aσa_{\sigma}-localized slice spectral sequence is

aσ1E2,=𝔽2[f¯2,b].a_{\sigma}^{-1}E_{2}^{*,*}={\mathbb{F}}_{2}[\bar{f}_{2},b]\ .

There is only one family of differentials, generated by

d13(b2)=f¯2d_{13}(b^{2})=\bar{f}_{2}

and the answer is

πC4aσ1BP((C4))2,2𝔽2[b4]{1,b}.\pi_{*}^{C_{4}}a_{\sigma}^{-1}BP^{(\!(C_{4})\!)}\langle 2,2\rangle\cong{\mathbb{F}}_{2}[b^{4}]\{1,b\}\ .

The computation is illustrated in Figure 4.

Refer to caption
Figure 4. The aσa_{\sigma}-localized slice spectral sequence of aσ1BP((C4))2,2a_{\sigma}^{-1}BP^{(\!(C_{4})\!)}\langle 2,2\rangle.

5.2. Application: Multiplicative structure

While the left action of MU((G))MU^{(\!(G)\!)} on itself always endows MU((G))/GS¯MU^{(\!(G)\!)}/G\cdot\bar{S} with a canonical MU((G))MU^{(\!(G)\!)}-module structure, and the same is true with BP((G))BP^{(\!(G)\!)} instead, much less is known for ring structures. We do have the following restrictive condition on quotients as a straightforward consequence of the techniques introduced above.

Theorem 5.16.

Let JJ\subseteq\mathbb{N} and S¯={s¯2j1jJ}\bar{S}=\{\bar{s}_{2^{j}-1}\mid j\in J\} be a set of generators for permutation summands. If there is a kJk\in J such that (k+1)J(k+1)\notin J, then BP((G))/GS¯BP^{(\!(G)\!)}/G\cdot\bar{S} does not have a ring structure in the homotopy category.

Proof.

If there is a ring structure on BP((G))/GS¯BP^{(\!(G)\!)}/G\cdot\bar{S}, then the map BP((G))/GS¯H¯BP^{(\!(G)\!)}/G\cdot\bar{S}\to H{\underline{{\mathbb{Z}}}} to the zero slice induces a map of multiplicative spectral sequences. This remains true after inverting aσa_{\sigma}. Since πG(aσ1H¯)𝔽2[b]\pi_{*}^{G}(a_{\sigma}^{-1}H{\underline{{\mathbb{Z}}}})\cong{\mathbb{F}}_{2}[b] and the map from πG(aσ1BP((G))/GS¯)\pi_{*}^{G}(a_{\sigma}^{-1}BP^{(\!(G)\!)}/G\cdot\bar{S}) to πG(aσ1H¯)\pi_{*}^{G}(a_{\sigma}^{-1}H{\underline{{\mathbb{Z}}}}) is the natural inclusion, the former is a subring of the latter. However, if kJk\in J and k+1Jk+1\not\in J, then b2k1b^{2^{k-1}} is nonzero in πG(aσ1BP((G))/GS¯)\pi_{*}^{G}(a_{\sigma}^{-1}BP^{(\!(G)\!)}/G\cdot\bar{S}), but its square b2kb^{2^{k}} is zero. This is a contradiction. ∎

Put another way, Theorem 5.16 says that the only possible BP((G))BP^{(\!(G)\!)}-module quotients BP((G))/GS¯BP^{(\!(G)\!)}/G\cdot\bar{S} by permutation summands which could be rings are the forms of BP((G))mBP^{(\!(G)\!)}\langle m\rangle. Even here, we know very little.

Example 5.17.

For G=C2G=C_{2}, BP1=kBP_{{\mathbb{R}}}\langle 1\rangle=k_{\mathbb{R}} and tmf1(3)tmf_{1}(3) is a form of BP2BP_{{\mathbb{R}}}\langle 2\rangle. Both admit commutative ring structures. For m>2m>2 we do not know if BPmBP_{{\mathbb{R}}}\langle m\rangle admits an associative ring structure.

For G=C4G=C_{4}, tmf1(5)tmf_{1}(5) is a form of BP((C4))1BP^{(\!(C_{4})\!)}\langle 1\rangle. For m>1m>1, we do not know if BP((C4))mBP^{(\!(C_{4})\!)}\langle m\rangle admits even an associative ring structure.

If we instead look only at the underlying spectrum, then work of Angeltveit and of Robinson shows that we have nice ring structures [1, 42]. This has been refined by Hahn–Wilson to show that this is still true in the category of MU((G))MU^{(\!(G)\!)} or BP((G))BP^{(\!(G)\!)}-modules [18].

Proposition 5.18.

For any JJ\subseteq\mathbb{N} and for any S¯\bar{S}, the spectrum

ieBP((G))/GS¯i_{e}^{\ast}BP^{(\!(G)\!)}/G\cdot\bar{S}

is an associative ieBP((G))i_{e}^{\ast}BP^{(\!(G)\!)}-algebra spectrum.

Proof.

The assumptions on S¯\bar{S} ensure that the sequence

(γis¯j0i2n11,jJ)\big{(}\gamma^{i}\bar{s}_{j}\mid 0\leq i\leq 2^{n-1}-1,j\in J\big{)}

forms a regular sequence in the homotopy groups of the even spectrum ieBP((G))i_{e}^{\ast}BP^{(\!(G)\!)}. The result follows from [18, Theorem A]. ∎

Remark 5.19.

The Hahn–Shi Real orientation shows the restriction to C2C_{2} of the spectrum BP((G))/GS¯BP^{(\!(G)\!)}/G\cdot\bar{S} always admits an EσE_{\sigma}-algebra structure [17].

6. The C2n1C_{2^{n-1}}-geometric fixed points

Let G=C2n1G^{\prime}=C_{2^{n-1}}, the subgroup of index two in G=C2nG=C_{2^{n}}. We extend the results of the previous section, considering the aλn2a_{\lambda_{n-2}}-localized slice spectral sequence for permutation quotients. This is again a spectral sequence of Mackey functors, now essentially for C4G/C2n2C_{4}\cong G/C_{2^{n-2}}. In this section, we study the C2G/C2n2C_{2}\cong G^{\prime}/C_{2^{n-2}}-equivariant level, since we can tell an increasingly complete story here. The C4C_{4}-fixed points are more subtle, as we will see in Section 7.

Note that since

iGλn2=λ(n1)1=2σ,i_{G^{\prime}}^{\ast}\lambda_{n-2}=\lambda_{(n-1)-1}=2\sigma,

the restriction to GG^{\prime} of the aλn2a_{\lambda_{n-2}}-localized slice spectral sequence for BP((G))/GS¯BP^{(\!(G)\!)}/G\cdot\bar{S} is the aσa_{\sigma}-localized slice spectral sequence for

iGBP((G))/GS¯.i_{G^{\prime}}^{\ast}BP^{(\!(G)\!)}/G\cdot\bar{S}.

Just as for the GG-geometric fixed points, we start by identifying the homotopy type. In this case, since

iGBP((G))BP((G))BP((G)),i_{G^{\prime}}^{\ast}BP^{(\!(G)\!)}\simeq BP^{((G^{\prime}))}\wedge BP^{((G^{\prime}))},

we have

ΦGBP((G))H𝔽2H𝔽2,\Phi^{G^{\prime}}BP^{(\!(G)\!)}\simeq H{\mathbb{F}}_{2}\wedge H{\mathbb{F}}_{2},

and all of the geometric fixed points we consider will take place in the category of modules over

A=H𝔽2H𝔽2.A=H{\mathbb{F}}_{2}\wedge H{\mathbb{F}}_{2}.

Composing with the localization map

iGBP((G))E~𝒫iGBP((G)),i_{G^{\prime}}^{\ast}BP^{(\!(G)\!)}\to\widetilde{E}\mathcal{P}\wedge i_{G^{\prime}}^{\ast}BP^{(\!(G)\!)},

the element NC2Gs¯iN_{C_{2}}^{G^{\prime}}\bar{s}_{i} gives us a polynomial in the dual Steenrod algebra.

Definition 6.1.

Let giπiAg_{i}\in\pi_{i}A be the image of NC2Gs¯iN_{C_{2}}^{G^{\prime}}\bar{s}_{i}.

Note that the residual C2G/C2n2C_{2}\cong G^{\prime}/C_{2^{n-2}}-action interchanges

NC2Gs¯i and γNC2Gs¯i,N_{C_{2}}^{G^{\prime}}\bar{s}_{i}\text{ and }\gamma N_{C_{2}}^{G^{\prime}}\bar{s}_{i},

while acting as the conjugation in the dual Steenrod algebra.

Lemma 6.2.

The GG^{\prime}-geometric fixed points of BP((G))/GS¯BP^{(\!(G)\!)}/G\cdot\bar{S} are the AA-module

A/(gi,g¯iiJ).A/(g_{i},\bar{g}_{i}\mid i\in J).

In general, the homotopy type of this module very heavily depends on the choices of generators. We have several cases where we can explicitly identify the images, however. Using [22, Proposition 2.57] and [39, Proposition 6.2], we see that for Hill–Hopkins–Ravenel generators v¯iG\overline{v}_{i}^{G} of BP((G))BP^{(\!(G)\!)}, the GG^{\prime}-geometric fixed points of their norms satisfy

ξi\displaystyle\xi_{i} =ΦGNC2Gv¯iG\displaystyle=\Phi^{G^{\prime}}N_{C_{2}}^{G}\overline{v}_{i}^{G}
ζi\displaystyle\zeta_{i} =ΦGNC2Gγv¯iG,\displaystyle=\Phi^{G^{\prime}}N_{C_{2}}^{G}\gamma\overline{v}_{i}^{G},

where ξi\xi_{i} are the Milnor generators of the mod 22 dual Steenrod algebra, and ζi\zeta_{i} are their dual.

6.1. Forms of BP((G))k,mBP^{(\!(G)\!)}\langle k,m\rangle

We can get much more explicit answers for the geometric fixed points with certain forms of BP((G))k,mBP^{(\!(G)\!)}\langle k,m\rangle, since here we can identify the geometric fixed points of the norms exactly.

Corollary 6.3.

The GG^{\prime}-geometric fixed points of BP((G))k,mBP^{(\!(G)\!)}\langle k,m\rangle are given by the AA-module

A/(ξi,ζii<k or i>m)A/(ξi,ζii<k)𝐴A/(ξj,ζjj>m).A/\big{(}\xi_{i},\zeta_{i}\mid i<k\text{ or }i>m\big{)}\simeq A/\big{(}\xi_{i},\zeta_{i}\mid i<k\big{)}\underset{A}{\wedge}A/\big{(}\xi_{j},\zeta_{j}\mid j>m\big{)}.

Writing this module in several ways makes working with this easier, as we can connect this with a series of modules and computations studied in [6].

Definition 6.4.

For any subset II of the natural numbers, let

MI=iIA/(ξi) and M¯I=iIA/(ζi).M_{I}=\bigwedge_{i\in I}A/(\xi_{i})\text{ and }\overline{M}_{I}=\bigwedge_{i\in I}A/(\zeta_{i}).

Let

Rk=EndA(M{1,,k1})R_{k}=\operatorname{End}_{A}\big{(}M_{\{1,\dots,k-1\}}\big{)}

and let

Am=M{m+1,m+2,}.A\langle m\rangle=M_{\{m+1,m+2,\dots\}}.

As the endomorphisms of a module, RkR_{k} is always an associative algebra. By [6], for any mm, AmA\langle m\rangle and A¯m\overline{A}\langle m\rangle are associative algebras as well. More surprisingly, by [6], we have

RkΣM{1,,k1}𝐴M¯{1,,k1},R_{k}\simeq\Sigma M_{\{1,\dots,k-1\}}\underset{A}{\wedge}\overline{M}_{\{1,\dots,k-1\}},

which allows us to rewrite ΦGBP((G))k,m\Phi^{G^{\prime}}BP^{(\!(G)\!)}\langle k,m\rangle.

Corollary 6.5.

For any kmk\leq m, we have

ΦGBP((G))k,mΣRk𝐴Am𝐴A¯m,\Phi^{G^{\prime}}BP^{(\!(G)\!)}\langle k,m\rangle\simeq\Sigma R_{k}\underset{A}{\wedge}A\langle m\rangle\underset{A}{\wedge}\overline{A}\langle m\rangle,

the suspension of an associative AA-algebra.

The extreme case of this is BP((G))m,mBP^{(\!(G)\!)}\langle m,m\rangle.

Corollary 6.6.

The GG^{\prime}-geometric fixed points of BP((G))m,mBP^{(\!(G)\!)}\langle m,m\rangle are given by the AA-module

ΣRm𝐴Am𝐴A¯m.\Sigma R_{m}\underset{A}{\wedge}A\langle m\rangle\underset{A}{\wedge}\overline{A}\langle m\rangle.

The homotopy of this AA-module is more complicated than one might have initially expected. These kinds of modules were studied by the authors [6], where we used a Baker–Lazarev style Adams spectral sequence based on H𝔽2H{\mathbb{F}}_{2}-homology, but in the category of AA-modules [3]. A remarkable feature of the case of BP((G))m,mBP^{(\!(G)\!)}\langle m,m\rangle is that this relative Adams spectral sequence completely determines the aσa_{\sigma}-localized slice spectral sequence.

6.2. A comparison of spectral sequences

Let P=PBP((G))m,mP_{\bullet}=P_{\bullet}BP^{(\!(G)\!)}\langle m,m\rangle be the slice covering tower of BP((G))m,mBP^{(\!(G)\!)}\langle m,m\rangle. That is, PtP_{t} is the homotopy fibre of the cannonical map

BP((G))m,mPt1BP((G))m,mBP^{(\!(G)\!)}\langle m,m\rangle\rightarrow P^{t-1}BP^{(\!(G)\!)}\langle m,m\rangle

where P=PBP((G))m,mP^{\bullet}=P^{\bullet}BP^{(\!(G)\!)}\langle m,m\rangle is the regular slice tower.

The slices PttBP((G))m,mP_{t}^{t}BP^{(\!(G)\!)}\langle m,m\rangle are non-trivial only in dimensions of the form t=2i(2m1)t=2i(2^{m}-1). Therefore we can “speed-up” the slice tower without losing any information. Define

P~t=P2t(2m1).\tilde{P}_{t}=P_{2t(2^{m}-1)}\ .

This re-indexes the slice tower, so that

P~tt=P2t(2m1)2t(2m1)BP((G))m,m.\tilde{P}^{t}_{t}=P^{2t(2^{m}-1)}_{2t(2^{m}-1)}BP^{(\!(G)\!)}\langle m,m\rangle.

Since there is an equivalence

ΦGBP((G))H𝔽2H𝔽2=A,\Phi^{G^{\prime}}BP^{(\!(G)\!)}\simeq H{\mathbb{F}}_{2}\wedge H{\mathbb{F}}_{2}=A,

ΦGP~\Phi^{G^{\prime}}\tilde{P}_{\bullet} is a covering tower converging to ΦGBP((G))m,m\Phi^{G^{\prime}}BP^{(\!(G)\!)}\langle m,m\rangle in the category of AA-modules.

Theorem 6.7.

The tower ΦGP~BP((G))m,m\Phi^{G^{\prime}}\tilde{P}_{\bullet}BP^{(\!(G)\!)}\langle m,m\rangle is an H𝔽2H{\mathbb{F}}_{2}-Adams resolution of ΦGBP((G))m,m\Phi^{G^{\prime}}BP^{(\!(G)\!)}\langle m,m\rangle in the category of AA-modules.

Proof.

Let Q=ΦGP~BP((G))m,mQ_{\bullet}=\Phi^{G^{\prime}}\tilde{P}_{\bullet}BP^{(\!(G)\!)}\langle m,m\rangle for convenience. Then QQ_{\bullet} is an H𝔽2H{\mathbb{F}}_{2}-Adams resolution of Q0=ΦGBP((G))m,mQ_{0}=\Phi^{G^{\prime}}BP^{(\!(G)\!)}\langle m,m\rangle in AA-modules if the following conditions are met for each i0i\geq 0 [40, Def. 2.2.1.3]:

  1. (1)

    QiiQ_{i}^{i} is a wedge of suspensions of H𝔽2H{\mathbb{F}}_{2}’s, and

  2. (2)

    the map QiQiiQ_{i}\rightarrow Q_{i}^{i} is monomorphic in H𝔽2H{\mathbb{F}}_{2}-homology.

We now verify the first condition. By definition,

Q00=ΦGP00BP((G))m,m=ΦGH¯=H𝔽2[b],Q_{0}^{0}=\Phi^{G^{\prime}}P^{0}_{0}BP^{(\!(G)\!)}\langle m,m\rangle=\Phi^{G^{\prime}}H\underline{\mathbb{Z}}=H{\mathbb{F}}_{2}[b],

with AA-module structure defined by the geometric fixed points of the reduction map BP((G))H¯BP^{(\!(G)\!)}\rightarrow H\underline{\mathbb{Z}}. By [22, Prop. 7.6], for each ii, v¯iG\overline{v}_{i}^{G} and its conjugate γv¯iG\gamma\overline{v}_{i}^{G} act trivially on H¯H\underline{\mathbb{Z}}, thus the geometric fixed points of NC2Gv¯iGN_{C_{2}}^{G^{\prime}}\overline{v}_{i}^{G} and NC2Gγv¯iGN_{C_{2}}^{G^{\prime}}\gamma\overline{v}_{i}^{G}, which are ξi\xi_{i} and ζi\zeta_{i}, act trivially on H𝔽2[b]H{\mathbb{F}}_{2}[b]. Therefore, as an AA-module, Q00j=0Σ2jH𝔽2Q_{0}^{0}\simeq\bigvee_{j=0}^{\infty}\Sigma^{2j}H{\mathbb{F}}_{2}. The Slice Theorem [22, Thm. 6.1] implies that for i>0i>0, QiiQ_{i}^{i} is a wedge of suspensions of Q00Q_{0}^{0}, thus the first condition is met.

We verify the second condition by an alternative construction of the slice covering tower of BP((G))m,mBP^{(\!(G)\!)}\langle m,m\rangle. As in [22, §6], let R=𝕊0[Gv¯mG]R=\mathbb{S}^{0}[G\cdot\overline{v}_{m}^{G}] be the homotopy refinement of BP((G))m,mBP^{(\!(G)\!)}\langle m,m\rangle, and MiM_{i} be the subcomplex of RR consisting of spheres of dimension 2i(2m1)\geq 2i(2^{m}-1). The arguments in [22, §6.1] tell us that

P~iBP((G))m,mRMi.\tilde{P}_{i}\simeq BP^{(\!(G)\!)}\langle m,m\rangle\wedge_{R}M_{i}.

Notice that GG^{\prime}-equivariantly, Mi+1MiM_{i+1}\subset M_{i} is the sub RR-module (v¯mG,γv¯mG)Mi(\overline{v}_{m}^{G},\gamma\overline{v}_{m}^{G})M_{i}, thus the quotient Mi/Mi+1M_{i}/M_{i+1} is equivalent to Mi/(v¯mG,γv¯mG)MiM_{i}/(\overline{v}^{G}_{m},\gamma\overline{v}_{m}^{G})M_{i}. Taking the GG^{\prime}-geometric fixed points on the cofibration

BP((G))m,mRMi+1BP((G))m,mRMiBP((G))m,mRMi/Mi+1,BP^{(\!(G)\!)}\langle m,m\rangle\wedge_{R}M_{i+1}\rightarrow BP^{(\!(G)\!)}\langle m,m\rangle\wedge_{R}M_{i}\rightarrow BP^{(\!(G)\!)}\langle m,m\rangle\wedge_{R}M_{i}/M_{i+1},

we obtain the cofibration

Qi+1QiQiiQi/(ξm,ζm)QiQ_{i+1}\rightarrow Q_{i}\rightarrow Q_{i}^{i}\simeq Q_{i}/(\xi_{m},\zeta_{m})Q_{i}

because ΦGNC2Gv¯mG=ξm\Phi^{G^{\prime}}N_{C_{2}}^{G^{\prime}}\overline{v}_{m}^{G}=\xi_{m} and ΦGNC2Gγv¯mG=ζm\Phi^{G^{\prime}}N_{C_{2}}^{G^{\prime}}\gamma\overline{v}_{m}^{G}=\zeta_{m}. Since ξm\xi_{m} and ζm\zeta_{m} have trivial image under AH𝔽2A\rightarrow H{\mathbb{F}}_{2}, the map QiQiiQ_{i}\rightarrow Q_{i}^{i} induces a monomorphism in H𝔽2H{\mathbb{F}}_{2}-homology. ∎

Corollary 6.8.

We have an isomorphism of spectral sequences between the relative Adams spectral sequence for A/(ξi,ζiim)A/(\xi_{i},\zeta_{i}\mid i\neq m) and the speeded-up aσa_{\sigma}-localized slice spectral sequence for iGBP((G))m,mi_{G^{\prime}}^{\ast}BP^{(\!(G)\!)}\langle m,m\rangle.

The dictionary here can be a little confusing, due to the scaling in the slice filtration. We record the un-scaled version here:

Remark 6.9.

A relative Adams drd_{r} corresponds to an ordinary aσa_{\sigma}-localized slice differential d2(2m1)r+1d_{2(2^{m}-1)r+1}.

Corollary 6.10.

The integer graded E2m+1E_{2^{m+1}}-page of the GG^{\prime}-equivariant aσa_{\sigma}-localized slice spectral sequence of iGBP((G))m,mi_{G^{\prime}}^{*}BP^{(\!(G)\!)}\langle m,m\rangle computing πGaσ1iGBP((G))m,m\pi_{*}^{G^{\prime}}a_{\sigma}^{-1}i_{G^{\prime}}^{*}BP^{(\!(G)\!)}\langle m,m\rangle is isomorphic to the E2E_{2}-page of the relative Adams spectral sequence of the spectrum ΦGiGBP((G))m,m\Phi^{G^{\prime}}i^{*}_{G^{\prime}}BP^{(\!(G)\!)}\langle m,m\rangle.

6.3. The Relative Adams spectral sequence for ΦGiGBP((G))m,m\Phi^{G^{\prime}}i^{*}_{G^{\prime}}BP^{(\!(G)\!)}\langle m,m\rangle

By Corollary 6.6, the GG^{\prime}-geometric fixed points of BP((G))m,mBP^{(\!(G)\!)}\langle m,m\rangle are a suspension of an associative ring spectrum. Because of this, we instead work with the associative algebra, since then the spectral sequence will be one of associative algebras.

One of the most surprising results from [6] was a decomposition of Am𝐴A¯m{A}\langle m\rangle\underset{A}{\wedge}\overline{A}\langle m\rangle, and hence of further quotients, into simpler, finite pieces. This makes our computations even easier.

Definition 6.11.

For each m1m\geq 1, let

A~m=Am𝐴M¯{m+1,,2m}.\tilde{A}\langle m\rangle=A\langle m\rangle\underset{A}{\wedge}\overline{M}_{\{m+1,\dots,2m\}}.
Proposition 6.12 ([6, Corollary 5.6]).

For each mm, we have a decomposition of AA-modules

Am𝐴A¯mk0Σ22m+1kA~m.A\langle m\rangle\underset{A}{\wedge}\overline{A}\langle m\rangle\simeq\bigvee_{k\geq 0}\Sigma^{2^{2m+1}k}\tilde{A}\langle m\rangle.
Proposition 6.13 ([6, Theorem 5.9]).

There is an associative algebra structure on A~m\tilde{A}\langle m\rangle such that the projection map

Am𝐴A¯mA~mA\langle m\rangle\underset{A}{\wedge}\overline{A}\langle m\rangle\to\tilde{A}\langle m\rangle

is a map of associative algebras.

This reduces the study of modules of the form

M𝐴Am𝐴A¯mM\underset{A}{\wedge}A\langle m\rangle\underset{A}{\wedge}\overline{A}\langle m\rangle

to the study of A~m\tilde{A}\langle m\rangle-modules of the form

M𝐴A~m.M\underset{A}{\wedge}\tilde{A}\langle m\rangle.

We apply this in the case of M=RmM=R_{m}, where we again have an associative algebra structure.

Definition 6.14.

Let

Rm=Rm𝐴A~m.R\langle m\rangle=R_{m}\underset{A}{\wedge}\tilde{A}\langle m\rangle.

By [6], the E2E_{2}-page of the relative Adams spectral sequence of RmR\langle m\rangle is given by

𝔽2[ξ1,,ξm,e]/e2mE(β2,β4,,β2m1)/(ξi+ξi+1β2i1im1),{\mathbb{F}}_{2}[\xi_{1},\ldots,\xi_{m},e]/e^{2^{m}}\otimes E(\beta_{{\mathchar 45\relax}2},\beta_{{\mathchar 45\relax}4},\ldots,\beta_{{\mathchar 45\relax}2^{m-1}})/\big{(}\xi_{i}+\xi_{i+1}\beta_{{\mathchar 45\relax}2^{i}}\mid 1\leq i\leq m-1\big{)},

where the bidegrees are given by:

|e|\displaystyle|e| =(2m+1,0)\displaystyle=(2^{m+1},0)
|ξm|\displaystyle|\xi_{m}| =(2m1,1)\displaystyle=(2^{m}-1,1)
|βk|\displaystyle|\beta_{{\mathchar 45\relax}k}| =(k,0).\displaystyle=(-k,0).

Since for 1i<m1\leq i<m, we have the relation

ξi=ξi+1β2i,\xi_{i}=\xi_{i+1}\beta_{{\mathchar 45\relax}2^{i}},

this simplifies as an algebra to

𝔽2[ξm]E(β2,β4,,β2m1)𝔽2[e2m+1]/(e2m+12m).{\mathbb{F}}_{2}[\xi_{m}]\otimes E(\beta_{{\mathchar 45\relax}2},\beta_{{\mathchar 45\relax}4},\ldots,\beta_{{\mathchar 45\relax}2^{m-1}})\otimes{\mathbb{F}}_{2}[e_{2^{m+1}}]/(e_{2^{m+1}}^{2^{m}}).
Notation 6.15.

We will use the following convenient notation:

β(2ϵ1+4ϵ2++2m1ϵm1):=β2ϵ1β4ϵ2β2m1ϵm1.\beta(2\epsilon_{1}+4\epsilon_{2}+\ldots+2^{m-1}\epsilon_{m-1}):=\beta_{{\mathchar 45\relax}2}^{\epsilon_{1}}\beta_{{\mathchar 45\relax}4}^{\epsilon_{2}}\cdots\beta_{{\mathchar 45\relax}2^{m-1}}^{\epsilon_{m-1}}\ .

We get elements,

β(0),β(2),,β(2m2)\beta(0),\beta(2),\ldots,\beta(2^{m}-2)

where β()\beta(\ell) has degree -\ell. Note that, as element on the E2E_{2}-page for RmR\langle m\rangle, for 00\leq\ell\leq\ell^{\prime}

β()β()={β()=0(()1)β(+)>0.\displaystyle\beta(\ell)\beta(\ell^{\prime})=\begin{cases}\beta(\ell^{\prime})&\ell=0\\ \left(\binom{\ell^{\prime}}{\ell}-1\right)\beta(\ell+\ell^{\prime})&\ell>0\ .\end{cases} (6.1)

In [6], we determined a number of differentials in these kinds of relative Adams spectral sequences.

Proposition 6.16 ([6, Corollary 7.5]).

In the relative Adams spectral sequence for A~m\widetilde{A}\langle m\rangle, for each 1im1\leq i\leq m and n0n\geq 0, we have differentials

d1+2i+1(e2i+2i+1n)=ξm2i+1ξi+1e2i+1n.d_{1+2^{i+1}}(e^{2^{i}+2^{i+1}n})=\xi_{m}^{2^{i+1}}\xi_{i+1}\cdot e^{2^{i+1}n}.

The spectrum RmR\langle m\rangle is an A~m\widetilde{A}\langle m\rangle-algebra. Therefore, by naturality and the relations on the E2E_{2}-page, in the relative Adams spectral sequence of RmR\langle m\rangle there are differentials

d1+2i+1(e2i+2i+1n)\displaystyle d_{1+2^{i+1}}(e^{2^{i}+2^{i+1}n}) =ξm1+2i+1e2i+1nβ(2m2i+1)\displaystyle=\xi_{m}^{1+2^{i+1}}e^{2^{i+1}n}\beta(2^{m}-2^{i+1}) (6.2)

for 0im10\leq i\leq m-1 and 0n<2m(i+1)0\leq n<2^{m-(i+1)}, provided that the target survives to the E1+2i+1E_{1+2^{i+1}}-page. We will see that all other differentials will be determined by these and the multiplicative structure of the spectral sequence.

We start with two useful lemmas.

Lemma 6.17.

If d1+2r(β()ek)d_{1+2r}(\beta(\ell)e^{k}) is non-zero, then

d1+2r(β()ek)=ξm1+2rβ(+2m2r)ekrd_{1+2r}(\beta(\ell)e^{k})=\xi_{m}^{1+2r}\beta(\ell+2^{m}-2r)e^{k-r}

for some 1rk1\leq r\leq k.

Proof.

Let

d1+2r(β()ek)=ξm1+2rβ()ek.d_{1+2r}(\beta(\ell)e^{k})=\xi_{m}^{1+2r}\beta(\ell^{\prime})e^{k^{\prime}}.

Then kkk^{\prime}\leq k so we let ss be a number such that k=ksk^{\prime}=k-s. Note that 0s<2m0\leq s<2^{m} and 0<r<2m0<r<2^{m}. Computing degrees, we obtain the equation

2m+1k1=(2m1)(1+2r)+2m+1(ks)\displaystyle 2^{m+1}k-\ell-1=(2^{m}-1)(1+2r)+2^{m+1}(k-s)-\ell^{\prime} (6.3)

This simplifies to

2m+()+2r=2m+1(rs).-2^{m}+(\ell^{\prime}-\ell)+2r=2^{m+1}(r-s).

Since 0,2m20\leq\ell,\ell^{\prime}\leq 2^{m}-2, we have

22m2m2.2-2^{m}\leq\ell^{\prime}-\ell\leq 2^{m}-2.

Furthermore, since 0<r<2m0<r<2^{m}, 0<2r<2m+10<2r<2^{m+1}. Therefore the absolute value of 2m+()+2r-2^{m}+(\ell^{\prime}-\ell)+2r is less than 2m+12^{m+1}. The equation above implies that this quantity is divisible by 2m+12^{m+1}. This implies that both sides of the equation must be zero. It follows that r=sr=s and =+2m2r\ell^{\prime}=\ell+2^{m}-2r. ∎

The next lemma is a straightforward but annoying exercise analyzing inequalities and we do not include the proof here.

Lemma 6.18.

Consider pairs (,k)(\ell,k), where \ell is even and 0,k2m10\leq\ell,k\leq 2^{m}-1. Define subsets of such pairs by

S={(,k):k},andS={(,k):<k}S=\{(\ell,k):k\leq\ell\},\quad\text{and}\quad S^{\prime}=\{(\ell,k):\ell<k\}

as follows. Let kk\leq\ell. Set

j\displaystyle j =max{0rm1:(2r)0mod2},\displaystyle=\mathrm{max}\left\{0\leq r\leq m-1:\binom{\ell}{2^{r}}\equiv 0\mod 2\right\},
i\displaystyle i =min{jrm1:(k2r)0mod2}.\displaystyle=\mathrm{min}\left\{j\leq r\leq m-1:\binom{k}{2^{r}}\equiv 0\mod 2\right\}.

Then letting

ϕ(,k)=((2m2i+1),k+2i)\phi(\ell,k)=(\ell-(2^{m}-2^{i+1}),k+2^{i})

gives a bijection ϕ:SS\phi\colon S\to S^{\prime}.

Remark 6.19.

If <k\ell<k and we fix i0i\geq 0, then for 2iκ<2i+12^{i}\leq\kappa<2^{i+1} and κ2i<κ\kappa-2^{i}\leq\ell<\kappa, if kk can be written in the form k=κ+2i+1nk=\kappa+2^{i+1}n, then ϕ1(,k)=(+2m2i+1,k2i)\phi^{-1}(\ell,k)=(\ell+2^{m}-2^{i+1},k-2^{i}). This formulation of the above bijection will be useful for proving the next result.

Theorem 6.20.

In the relative Adams spectral sequence of RmR\langle m\rangle, for

0k,2m10\leq k,\ell\leq 2^{m}-1

with \ell even, we have the following:

  1. (1)

    the class β()ek\beta(\ell)e^{k} is a permanent cycle if and only if kk\leq\ell;

  2. (2)

    if <k\ell<k, the class β()ek\beta(\ell)e^{k} supports a non-trivial differential, determined as follows. Fix i0i\geq 0. For 2iκ<2i+12^{i}\leq\kappa<2^{i+1} and κ2i<κ\kappa-2^{i}\leq\ell<\kappa, if k=κ+2i+1nk=\kappa+2^{i+1}n, then there is a differential

    d1+2i+1(β()ek)=ξm1+2i+1β(+2m2i+1)ek2i.d_{1+2^{i+1}}(\beta(\ell)e^{k})=\xi_{m}^{1+2^{i+1}}\beta(\ell+2^{m}-2^{i+1})e^{k-2^{i}}\ .

    These are the only non-trivial differentials.

Proof.

This is a multiplicative spectral sequence. At E2E_{2}-page, there is a vanishing line of slope 1/(2m1)1/(2^{m}-1) with intercept on the (ts)(t-s)-axis at 2m22^{m}-2 (the vanishing line is formed by the ξm\xi_{m}-multiples on β(2m2)\beta(2^{m}-2)). Furthermore, looking at the map of spectral sequences from A~n\widetilde{A}\langle n\rangle, we see that the class e2ie^{2^{i}} survives to the E1+2i+1E_{1+2^{i+1}}-page for 1im11\leq i\leq m-1. Therefore, the differentials drd_{r} are e2ie^{2^{i}}-linear for r<2i+1+1r<2^{i+1}+1. The first non-zero class in positive filtration is ξmβ(2m2)\xi_{m}\beta(2^{m}-2) which has topological degree (2m1)(2m2)=1(2^{m}-1)-(2^{m}-2)=1. Therefore, every element of E(β)E(\beta) is a permanent cycle and the spectral sequence is one of modules over this exterior algebra.

We will prove the following statements inductively on 0jm10\leq j\leq m-1:

  1. (1)

    For 2jk<2j+12^{j}\leq k<2^{j+1}, if k2j<kk-2^{j}\leq\ell<k, then

    d1+2j+1(β()ek+2j+1n)=ξm1+2j+1β(+2m2j+1)ek2j+2j+1n.d_{1+2^{j+1}}(\beta(\ell)e^{k+2^{j+1}n})=\xi_{m}^{1+2^{j+1}}\beta(\ell+2^{m}-2^{j+1})e^{k-2^{j}+2^{j+1}n}\ .
  2. (2)

    For 2jk<2j+12^{j}\leq k<2^{j+1} and kk\leq\ell, the class β()ek\beta(\ell)e^{k} is a permanent cycle.

  3. (3)

    There are no other non-trivial differentials until the E1+2j+2E_{1+2^{j+2}}-page.

We note that (1) and (2) inductively imply that any class β()ek\beta(\ell)e^{k} with k<2j+1k<2^{j+1} either supports a differential drd_{r} for r1+2j+1r\leq 1+2^{j+1}, or is a permanent cycle.

To prove the inductive claim, we start with j=0j=0, so that k=1k=1 in (1). Using that \ell is even, in (1), the range forces =0\ell=0. The first possible non-trivial differential for degree reasons is on ee, and this differential is forced by the d3d_{3}-differential

d3(e)=ξm3β(2m2)d_{3}(e)=\xi_{m}^{3}\beta(2^{m}-2)

in A~n\widetilde{A}\langle n\rangle. All d3d_{3}-s are then determined by e2e^{2}-linearity and given by

d3(e1+2n)=ξm3β(2m2)e2n.d_{3}(e^{1+2n})=\xi_{m}^{3}\beta(2^{m}-2)e^{2n}.

Here, we have used the fact that the differentials are linear over 𝔽2[ξm]E(β){\mathbb{F}}_{2}[\xi_{m}]\otimes E(\beta). For degree reasons, the classes β()e\beta(\ell)e for 2\ell\geq 2 are permanent cycles, proving (2). The differentials are e2e^{2}-linear and all other classes that could support a d3d_{3} are the product of e2e^{2} with permanent cycles. So they survive to the E5E_{5}-page.

Let i>0i>0 and assume that (1), (2), (3) hold for smaller values of 0j<i{0\leq j<i}. As noted above, the differentials in the spectral sequence of A~m\widetilde{A}\langle m\rangle implies the differentials

d1+2i+1(e2i+2i+1n)=ξm1+2i+1β(2m2i+1)e2i+1n,d_{1+2^{i+1}}(e^{2^{i}+2^{i+1}n})=\xi_{m}^{1+2^{i+1}}\beta(2^{m}-2^{i+1})e^{2^{i+1}n},

provided that the targets survive to the E1+2i+1E_{1+2^{i+1}}-page. By the induction hypothesis and Lemma 6.18, this is the case. This proves the differential of (1) for k=2ik=2^{i} and =0\ell=0.

Now, choose kk and \ell so that 2ik<2i+12^{i}\leq k<2^{i+1} and k2i<kk-2^{i}\leq\ell<k as in (1). In particular, >0\ell>0. The class [β()ek2i][\beta(\ell)e^{k-2^{i}}] is a permanent cycle by the induction hypothesis. Therefore,

d1+2i+1(β()ek+2i+1n)\displaystyle d_{1+2^{i+1}}(\beta(\ell)e^{k+2^{i+1}n}) =d1+2i+1([β()ek2i]e2i+2i+1n)\displaystyle=d_{1+2^{i+1}}([\beta(\ell)e^{k-2^{i}}]e^{2^{i}+2^{i+1}n})
=[β()ek2i]d1+2i+1(e2i)e2i+1n\displaystyle=[\beta(\ell)e^{k-2^{i}}]d_{1+2^{i+1}}(e^{2^{i}})e^{2^{i+1}n}
=[β()ek2i]ξm1+2i+1β(2m2i+1)e2i+1n\displaystyle=[\beta(\ell)e^{k-2^{i}}]\xi_{m}^{1+2^{i+1}}\beta(2^{m}-2^{i+1})e^{2^{i+1}n}
=ξm1+2i+1β()β(2m2i+1)ek2i+2i+1n.\displaystyle=\xi_{m}^{1+2^{i+1}}\beta(\ell)\beta(2^{m}-2^{i+1})e^{k-2^{i}+2^{i+1}n}\ .

The binomial expansion of 2m2i+12^{m}-2^{i+1} is

2m1++2i+2+2i+1.2^{m-1}+\ldots+2^{i+2}+2^{i+1}.

The bounds on \ell give 0<<2i+10<\ell<2^{i+1}, which guarantees that (2m2i+1)=0\binom{2^{m}-2^{i+1}}{\ell}=0 since >0\ell>0. So, by (6.1)

β()β(2m2i+1)=β(+2m2i+1).\beta(\ell)\beta(2^{m}-2^{i+1})=\beta(\ell+2^{m}-2^{i+1}).

We get a non-trivial differential as long as the target is non-zero, which is the case by the induction hypothesis and Lemma 6.18. This proves (1).

We next show that the classes β()ek\beta(\ell)e^{k} for kk\leq\ell and 2ik<2i+12^{i}\leq k<2^{i+1} are permanent cycles. Suppose that for 2irk2^{i}\leq r\leq k,

d1+2r(β()ek)=ξm1+2rβ(+2m2r)ekr.d_{1+2r}(\beta(\ell)e^{k})=\xi_{m}^{1+2r}\beta(\ell+2^{m}-2r)e^{k-r}.

The form of the differential comes from Lemma 6.17. Note that

krrr+2mr=+2m2r.k-r\leq\ell-r\leq\ell-r+2^{m}-r=\ell+2^{m}-2r.

This shows that the target is a permanent cycle by the induction hypothesis. We will now show that this target is actually killed by a shorter differential.

Since

+2m2r(2m2i+1)=2i+1+2r2i+1+k2k=2i+1k>0.\ell+2^{m}-2r-(2^{m}-2^{i+1})=2^{i+1}+\ell-2r\geq 2^{i+1}+k-2k=2^{i+1}-k>0.

Therefore, +2m2r>2m2i+1\ell+2^{m}-2r>2^{m}-2^{i+1} and so we can write

+2m2r\displaystyle\ell+2^{m}-2r =2m2j+1+¯,\displaystyle=2^{m}-2^{j+1}+\bar{\ell}, ¯<2j, 0ji\displaystyle\bar{\ell}<2^{j},\ 0\leq j\leq i
kr\displaystyle k-r =k¯+2h2j+2h+1n\displaystyle=\bar{k}+2^{h}-2^{j}+2^{h+1}n jh,k¯<2j, 0n\displaystyle j\leq h,\ \bar{k}<2^{j},\ 0\leq n

Let

=+2m2r(2n2h+1)=¯+2h+12j+1\ell^{\prime}=\ell+2^{m}-2r-(2^{n}-2^{h+1})=\bar{\ell}+2^{h+1}-2^{j+1}

and

κ=kr2h+1n+2h=k¯+2h+12j.\kappa^{\prime}=k-r-2^{h+1}n+2^{h}=\bar{k}+2^{h+1}-2^{j}.

Then, κ2h<κ\kappa^{\prime}-2^{h}\leq\ell^{\prime}<\kappa^{\prime} and 2hκ<2h+12^{h}\leq\kappa^{\prime}<2^{h+1}. So by the induction hypothesis,

d1+2h+1(ξm2r2h+1β()eκ+2h+1n)\displaystyle d_{1+2^{h+1}}(\xi_{m}^{2r-2^{h+1}}\beta(\ell^{\prime})e^{\kappa^{\prime}+2^{h+1}n}) =ξm2r2h+1(ξm1+2h+1β(+2m2h+1)eκ2h+2h+1n)\displaystyle=\xi_{m}^{2r-2^{h+1}}\left(\xi_{m}^{1+2^{h+1}}\beta(\ell^{\prime}+2^{m}-2^{h+1})e^{\kappa^{\prime}-2^{h}+2^{h+1}n}\right)
=ξm1+2rβ(+2m2r)ekr.\displaystyle=\xi_{m}^{1+2r}\beta(\ell+2^{m}-2r)e^{k-r}\ .

Therefore, the target was killed by a shorter differential. This proves (2).

Finally, (3) holds by the linearity of the differentials with respect to the d2i+2d_{2^{i+2}}-cycle e2i+1e^{2^{i+1}}. ∎

Finally, the correspondence between the aσa_{\sigma}-localized slice spectral sequence of iGBP((G))m,mi_{*}^{G^{\prime}}BP^{(\!(G)\!)}\langle m,m\rangle and the relative Adams spectral sequence is as follows:

Summary 6.21.

The E2E_{2}-page of the relative Adams spectral sequence of the geometric fixed points ΦGBP((G))m,m\Phi^{G^{\prime}}BP^{(\!(G)\!)}\langle m,m\rangle is, additively,

E2,Σ2m2𝔽2[ξm]E(β)𝔽2[e2m+1]E^{*,*}_{2}\cong\Sigma^{2^{m}-2}{\mathbb{F}}_{2}[\xi_{m}]\otimes E(\beta)\otimes{\mathbb{F}}_{2}[e_{2^{m+1}}]

where the shift preserves the filtration and adds 2m22^{m}-2 to the topological degree.

The correspondence between the slice spectral sequence and the relative Adams spectral sequence is as follows:

  1. (1)

    The elements bkb^{k} for 0k2m110\leq k\leq 2^{m-1}-1 correspond to β(2m2(k+1))\beta(2^{m}-2(k+1)). Note that b2m11b^{2^{m-1}-1} corresponds to β(0)=1\beta(0)=1, the multiplicative unit in the relative Adams spectral sequence of RmR\langle m\rangle.

  2. (2)

    For =2n1(2m1)+1\ell=2^{n-1}(2^{m}-1)+1, 0k<2m10\leq k<2^{m-1} and r0r\geq 0, the element bk+2m1b^{k+2^{m-1}} in the localized sliced spectral sequence supports the differential d(bk+2m1)=bk(f¯m+γf¯m)d_{\ell}(b^{k+2^{m-1}})=b^{k}(\bar{f}_{m}+\gamma\bar{f}_{m}), forced by the slice differential

    d(u2σ2m1)=aσ2m+11(v¯mG+γv¯mG).d_{\ell}(u_{2\sigma}^{2^{m-1}})=a_{\sigma}^{2^{m+1}-1}(\overline{v}_{m}^{G}+\gamma\overline{v}_{m}^{G}).

    These differentials are b2mb^{2^{m}}-linear, leaving behind bkb^{k} where the dyadic expansion of k=ε0+ε12+k=\varepsilon_{0}+\varepsilon_{1}\cdot 2+\ldots satisfies εm1=0\varepsilon_{m-1}=0. This family of differentials identifies f¯m\overline{f}_{m} with γf¯m\gamma\overline{f}_{m} in the whole slice spectral sequence.

  3. (3)

    After the slice differentials above, multiplication by either f¯m\bar{f}_{m} or γf¯m\gamma\overline{f}_{m} corresponds to multiplication by ξm\xi_{m} in the relative Adams spectral sequence.

  4. (4)

    The remaining elements bkb^{k} are in one-to-one correspondence with the elements in filtration s=0s=0 in the relative Adams spectral sequence, shifted by a degree 2m22^{m}-2. In particular, the element b2m+2m11b^{2^{m}+2^{m-1}-1} corresponds to e2m+1e_{2^{m+1}}.

6.4. The C2C_{2}-slice spectral sequence of BP((C4))2,2BP^{(\!(C_{4})\!)}\langle 2,2\rangle

As an application, we illustrate the above correspondence of spectral sequences by computing the C2C_{2}-slice spectral sequence of BP((C4))2,2BP^{(\!(C_{4})\!)}\langle 2,2\rangle.

The C2C_{2}-slice spectral sequence of BP((C4))2,2BP^{(\!(C_{4})\!)}\langle 2,2\rangle is determined by the relative Adams spectral sequence for

R2=A~2/(ζ5,)AEndA(M1),R\langle 2\rangle=\widetilde{A}\langle 2\rangle/(\zeta_{5},\ldots)\wedge_{A}\operatorname{End}_{A}(M_{1}),

whose computation follows from Section 6.3 above. The essential features were also completely computed in Section 7.3 of [6]. The E2E_{2}-page is

𝔽2[ξ2]E(β2)𝔽2[e8].\mathbb{F}_{2}[\xi_{2}]\otimes E(\beta_{{\mathchar 45\relax}2})\otimes\mathbb{F}_{2}[e_{8}].

There are only d3d_{3}- and d5d_{5}-differentials. The d3d_{3}-differentials are generated by

d3(e81+2)\displaystyle d_{3}(e_{8}^{1+2*}) =e82ξ23β2,\displaystyle=e_{8}^{2*}\xi_{2}^{3}\beta_{-2},

and the d5d_{5}-differentials are generated by

d5(e82+4)\displaystyle d_{5}(e_{8}^{2+4*}) =e84ξ25,\displaystyle=e_{8}^{4*}\xi_{2}^{5},
d5(e83+4β2)\displaystyle d_{5}(e_{8}^{3+4*}\beta_{{\mathchar 45\relax}2}) =e81+4ξ25β2.\displaystyle=e_{8}^{1+4*}\xi_{2}^{5}\beta_{{\mathchar 45\relax}2}.

The E2E_{2}-term of the aσa_{\sigma}-localized slice spectral sequence for aσ1iC2BP((C4))2,2a_{\sigma}^{-1}i_{C_{2}}^{*}BP^{(\!(C_{4})\!)}\langle 2,2\rangle is

𝔽2[u2σ,aσ±1,t¯2,γt¯2].\mathbb{F}_{2}[u_{2\sigma},a_{\sigma}^{\pm 1},\bar{t}_{2},\gamma\bar{t}_{2}].

As before, let b=u2σ/a2σb=u_{2\sigma}/a_{2\sigma}. The shortest differentials in this spectral sequence are the d7d_{7}-differentials, whose effects are to identify t¯2\bar{t}_{2} with γt¯2\gamma\bar{t}_{2}. The d7d_{7}-differentials are generated by

d7(b2+4)\displaystyle d_{7}(b^{2+4*}) =b4(t¯2+γt¯2)aσ3,\displaystyle=b^{4*}(\bar{t}_{2}+\gamma\bar{t}_{2})a_{\sigma}^{3},
d7(b3+4)\displaystyle d_{7}(b^{3+4*}) =b1+4(t¯2+γt¯2)aσ3.\displaystyle=b^{1+4*}(\bar{t}_{2}+\gamma\bar{t}_{2})a_{\sigma}^{3}.

After the d7d_{7}-differentials, we can then import the differentials from the relative Adams spectral sequence as explained in Summary 6.21. As a result, the Adams d3d_{3}-differentials become the slice d19d_{19}-differentials, generated by

d19(b5+8)=b1+8t¯23aσ9.d_{19}(b^{5+8*})=b^{1+8*}\bar{t}_{2}^{3}a_{\sigma}^{9}.

The Adams d5d_{5}-differentials become the slice d31d_{31}-differentials, generated by

d31(b9+16)\displaystyle d_{31}(b^{9+16*}) =b1+16t¯25aσ15,\displaystyle=b^{1+16*}\bar{t}_{2}^{5}a_{\sigma}^{15},
d31(b12+16)\displaystyle d_{31}(b^{12+16*}) =b4+16t¯25aσ15.\displaystyle=b^{4+16*}\bar{t}_{2}^{5}a_{\sigma}^{15}.

Figure 5 shows the aσa_{\sigma}-localized slice spectral sequence for aσ1iC2BP((C4))2,2a_{\sigma}^{-1}i_{C_{2}}^{*}BP^{(\!(C_{4})\!)}\langle 2,2\rangle. At E2E_{2}, the classes \bullet denote families of monomials formed by the classes aσ3t¯2a_{\sigma}^{3}\bar{t}_{2} and aσ3γt¯2}a_{\sigma}^{3}\gamma\bar{t}_{2}\} on the various powers of bb. At E8E_{8}, a \bullet depicted as the target of a d7d_{7}-differential becomes a copy of /2{\mathbb{Z}}/2, represented by a class of the form aσ3t¯2bka_{\sigma}^{3\ast}\bar{t}_{2}^{\ast}b^{k} for k2,3mod4k\neq 2,3\mod 4. Each \bullet depicted as the source of a d7d_{7}-differential is completely gone as the d7d_{7}-differentials are injective.

As in Theorem 4.6, to obtain the differentials in the C2C_{2}-slice spectral sequence of BP((C4))2,2BP^{(\!(C_{4})\!)}\langle 2,2\rangle, we truncate at the horizontal line of filtration s=0s=0 and remove the region below this line.

Refer to caption
Figure 5. The aσa_{\sigma}-localized slice spectral sequence of aσ1iC2BP((C4))2,2a_{\sigma}^{-1}i_{C_{2}}^{*}BP^{(\!(C_{4})\!)}\langle 2,2\rangle.

7. The C4C_{4}-localized slice spectral sequence of aλ1BP((C4))2,2a_{\lambda}^{-1}BP^{(\!(C_{4})\!)}\langle 2,2\rangle

In this section, we compute the integer-graded C4C_{4}-localized slice spectral sequence of aλ1BP((C4))2,2a_{\lambda}^{-1}BP^{(\!(C_{4})\!)}\langle 2,2\rangle, using all the tools that we have developed in the previous sections. This computation serves to demonstrate the robustness of our techniques as well as providing insights to higher differentials phenomena when generalized to higher heights.

Remark 7.1.

Just like the integer-graded spectral sequence, the full RO(C4)RO(C_{4})-graded spectral sequence can be computed by the exact same method. We have opted to only compute the integer-graded slice spectral sequence because it is more convenient to present its diagrams.

Remark 7.2.

By the discussion in Section 4, the slice spectral sequence of the unlocalized spectrum BP((C4))2,2BP^{(\!(C_{4})\!)}\langle 2,2\rangle is completely determined by the aλa_{\lambda}-localized slice spectral sequence by truncating away the region below the line of filtration s=0s=0.

Remark 7.3.

The following facts are good to keep in mind while doing the computation:

  1. (1)

    The differentials with source on or above the line of slope 1 in the C4C_{4}-localized slice spectral sequence of aλ1BP((C4))2,2a_{\lambda}^{-1}BP^{(\!(C_{4})\!)}\langle 2,2\rangle are determined by the C4C_{4}-localized slice spectral sequence of aσ1BP((C4))2,2a_{\sigma}^{-1}BP^{(\!(C_{4})\!)}\langle 2,2\rangle computed in Example 5.15. These are all d13d_{13}-differentials.

  2. (2)

    Many of the differentials d31d_{\leq 31} are determined using the C2C_{2}-slice differentials computed in Section 6.4 and the Mackey functor structure (i.e. restriction and transfer).

  3. (3)

    The C4C_{4}-localized slice spectral sequence of aλ1BP((C4))2,2a_{\lambda}^{-1}BP^{(\!(C_{4})\!)}\langle 2,2\rangle is a module over the spectral sequence of aλ1MU((C4))a_{\lambda}^{-1}MU^{(\!(C_{4})\!)}, but very little of that structure is needed for the computation (see Section 7.3). Multiplication with respect to two key classes gives rise to periodicity of differentials and a vanishing line (Theorem 7.20). These phenomena determine all higher differentials (d>31d_{>31}).

7.1. Organization of the slice associated graded

For the rest of this section, we let

t¯1=v¯1C4andt¯2=v¯2C4.\bar{t}_{1}=\bar{v}_{1}^{C_{4}}\quad\text{and}\quad\bar{t}_{2}=\bar{v}_{2}^{C_{4}}.

From Corollary 2.14, the slice associated graded for BP((C4))2,2BP^{(\!(C_{4})\!)}\langle 2,2\rangle is H¯[t¯2,γt¯2]H\underline{\mathbb{Z}}[\bar{t}_{2},\gamma\bar{t}_{2}], so the E2E_{2}-page of our spectral sequence is obtained by aλa_{\lambda}-localization of the homotopy of this slice associated graded.

We organize the slice cells by powers of

𝔡¯t¯2:=NC2C4(t¯2).\bar{\mathfrak{d}}_{\bar{t}_{2}}:=N_{C_{2}}^{C_{4}}(\bar{t}_{2}).
Remark 7.4.

This mirrors the approach taken by Hill–Hopkins–Ravenel in [26] to compute the slice spectral sequence of BP((C4))1=BP((C4))1,1BP^{(\!(C_{4})\!)}\langle 1\rangle=BP^{(\!(C_{4})\!)}\langle 1,1\rangle, where they organized the slice cells by powers of 𝔡¯t¯1:=NC2C4(t¯1)\bar{\mathfrak{d}}_{\bar{t}_{1}}:=N_{C_{2}}^{C_{4}}(\bar{t}_{1}).

The slice cells are organized according to the following matrix:

(𝔡¯t¯20𝔡¯t¯21𝔡¯t¯22𝔡¯t¯20(t¯2,γt¯2)𝔡¯t¯21(t¯2,γt¯2)𝔡¯t¯22(t¯2,γt¯2)𝔡¯t¯20(t¯22,γt¯22)𝔡¯t¯21(t¯22,γt¯22)𝔡¯t¯22(t¯22,γt¯22))\displaystyle\begin{pmatrix}\bar{\mathfrak{d}}_{\bar{t}_{2}}^{0}&\bar{\mathfrak{d}}_{\bar{t}_{2}}^{1}&\bar{\mathfrak{d}}_{\bar{t}_{2}}^{2}&\cdots\\ \bar{\mathfrak{d}}_{\bar{t}_{2}}^{0}(\bar{t}_{2},\gamma\bar{t}_{2})&\bar{\mathfrak{d}}_{\bar{t}_{2}}^{1}(\bar{t}_{2},\gamma\bar{t}_{2})&\bar{\mathfrak{d}}_{\bar{t}_{2}}^{2}(\bar{t}_{2},\gamma\bar{t}_{2})&\cdots\\ \bar{\mathfrak{d}}_{\bar{t}_{2}}^{0}(\bar{t}_{2}^{2},\gamma\bar{t}_{2}^{2})&\bar{\mathfrak{d}}_{\bar{t}_{2}}^{1}(\bar{t}_{2}^{2},\gamma\bar{t}_{2}^{2})&\bar{\mathfrak{d}}_{\bar{t}_{2}}^{2}(\bar{t}_{2}^{2},\gamma\bar{t}_{2}^{2})&\cdots\\ \vdots&\vdots&\vdots&\ddots\end{pmatrix} (7.1)

To read this, note that iC2𝔡¯t¯2=t¯2γt¯2i^{*}_{C_{2}}\bar{\mathfrak{d}}_{\bar{t}_{2}}=\bar{t}_{2}\gamma\bar{t}_{2} so that every monomial in t¯2,γt¯2\bar{t}_{2},\gamma\bar{t}_{2} appears in some entry of the matrix.

Entries 𝔡¯t¯2j\bar{\mathfrak{d}}_{\bar{t}_{2}}^{j} in the matrix represent slice cells of the form

H¯S3jρ4,H\underline{\mathbb{Z}}\wedge S^{3j\rho_{4}},

where ρ4\rho_{4} is the regular representation of C4C_{4}. These are the regular (or non-induced) cells. Entries of the form 𝔡¯t¯2j(t¯2i,γt¯2i)\bar{\mathfrak{d}}_{\bar{t}_{2}}^{j}(\bar{t}_{2}^{i},\gamma\bar{t}_{2}^{i}) represent slice cells of the form

H¯(C4+C2SiC2(3jρ4+3iρ2))H¯(C4+C2S(6j+3i)ρ2).H\underline{\mathbb{Z}}\wedge({C_{4}}_{+}\wedge_{C_{2}}S^{i^{*}_{C_{2}}(3j\rho_{4}+3i\rho_{2})})\simeq H\underline{\mathbb{Z}}\wedge({C_{4}}_{+}\wedge_{C_{2}}S^{(6j+3i)\rho_{2}}).

These are the induced cells.

For the non-induced cells, the homotopy groups of H¯Sρ4H\underline{\mathbb{Z}}\wedge S^{*\rho_{4}} are computed in [26] and depicted in Figure 3 of that reference. For the induced cells, we get the induced Mackey functors

H¯(C4+Sρ2)C2C4H¯C2(Sρ2),H\underline{\mathbb{Z}}_{*}({C_{4}}_{+}\wedge S^{*\rho_{2}})\cong\uparrow_{C_{2}}^{C_{4}}H\underline{\mathbb{Z}}_{*}^{C_{2}}(S^{*\rho_{2}}),

(see Definition 2.6 of [26]) whose values are also known. After inverting aλa_{\lambda}, we obtain the E2E_{2}-page depicted in Figure 6.

We use the same notation as in Table 1 of [26] for the Mackey functors. Blue Mackey functors are supported on induced cells and represent multiple copies of the Mackey functor ^\hat{\bullet} of [26], supported on the various monomials in the matrix (7.1) that are not in the first row.

We name the classes that do not come from induced cells. First, there are classes of order four (which have non-trivial restrictions):

  • 𝔡¯t¯22iu(6ij)λu6iσajλ\bar{\mathfrak{d}}_{\bar{t}_{2}}^{2i}u_{(6i-j)\lambda}u_{6i\sigma}a_{j\lambda} in degree (24i2j,2j)(24i-2j,2j) for 0i0\leq i and j6ij\leq 6i.

Next, the classes of order two which do not come from induced cells are:

  • 𝔡¯t¯2iu2kσa3iλa(3i2k)σ\bar{\mathfrak{d}}_{\bar{t}_{2}}^{i}u_{2k\sigma}a_{3i\lambda}a_{(3i-2k)\sigma} in degree (3i+2k,9i2k)(3i+2k,9i-2k) for 0i0\leq i and 02k<3i0\leq 2k<3i. These are above the line of slope one.

  • 𝔡¯t¯2iujλu(3i1)σa(3ij)λaσ\bar{\mathfrak{d}}_{\bar{t}_{2}}^{i}u_{j\lambda}u_{(3i-1)\sigma}a_{(3i-j)\lambda}a_{\sigma} in degree (6i+2j1,6i2j+1)(6i+2j-1,6i-2j+1) for ii odd, 0i0\leq i and 0j0\leq j.

The induced cells are named by treating them as images of the transfer map from the corresponding classes in the C2C_{2}-slice spectral sequence.

Refer to caption
Figure 6. The E2E_{2}-page of the Mackey functor valued aλa_{\lambda}-localized slice spectral sequence of BP((C4))2,2BP^{(\!(C_{4})\!)}\langle 2,2\rangle.

7.2. The d7d_{7}-differentials

The first differentials are the d7d_{7}-differentials. They occur between classes supported on slice cells that are in the same column of the matrix (7.1). The d7d_{7}-differentials are all proven by restricting to the aσa_{\sigma}-localized slice spectral sequence of aσ1iC2BP((C4))2,2a_{\sigma}^{-1}i_{C_{2}}^{*}BP^{(\!(C_{4})\!)}\langle 2,2\rangle. More specifically, the restriction of certain classes to the C2C_{2}-spectral sequence support the d7d_{7}-differentials discussed in Section 6.4, and therefore by naturality and degree reasons their preimages must also support d7d_{7}-differentials in the C4C_{4}-spectral sequence. By naturality and degree reasons, these are all the d7d_{7}-differentials that can occur, after which we obtain the E8E_{8}-page.

Let b=uλaλb=\frac{u_{\lambda}}{a_{\lambda}}. The restriction of this class is the class b=u2σa2σb=\frac{u_{2\sigma}}{a_{2\sigma}} in the C2C_{2}-spectral sequence. We have d7d_{7}-differentials

d7(b2+4)\displaystyle d_{7}\left(b^{2+4*}\right) =trC2C4(t¯2a3σ2)b4\displaystyle=tr_{C_{2}}^{C_{4}}(\bar{t}_{2}a_{3\sigma_{2}})b^{4*}
d7(b3+4)\displaystyle d_{7}\left(b^{3+4*}\right) =trC2C4(t¯2a3σ2)b1+4.\displaystyle=tr_{C_{2}}^{C_{4}}(\bar{t}_{2}a_{3\sigma_{2}})b^{1+4*}.

Figure 7 depicts the E7E_{7}-page. The \circ classes are /4{\mathbb{Z}}/4’s coming from non-induced cells. The black \bullet classes are /2{\mathbb{Z}}/2’s coming from the non-induced cells, while the blue {\color[rgb]{0,0,1}\definecolor[named]{pgfstrokecolor}{rgb}{0,0,1}\bullet} classes are direct sums of /2{\mathbb{Z}}/2’s coming from the induced cells.

Figure 8 depicts the E8E_{8}-page. In that figure, all dots represent a copy of /2{\mathbb{Z}}/2, with the exception of the white \circ classes which represent /4{\mathbb{Z}}/4’s. The pink denotes a degree where, on a previous page, there was a /4{\mathbb{Z}}/4 but the generator supported a non-zero differential and is no longer present. As before, blue classes come from induced cells.

Refer to caption
Figure 7. The E7E_{7}-page of the aλa_{\lambda}-localized slice spectral sequence of aλ1BP((C4))2,2a_{\lambda}^{-1}BP^{(\!(C_{4})\!)}\langle 2,2\rangle.
Refer to caption
Figure 8. The E8E_{8}-page of the aλa_{\lambda}-localized slice spectral sequence of aλ1BP((C4))2,2a_{\lambda}^{-1}BP^{(\!(C_{4})\!)}\langle 2,2\rangle.

7.3. Strategy for computing higher differentials

Before computing the higher differentials (those of lengths 13\geq 13), we describe our strategy. There are two classes that will be very important for our computation. They are the classes

α\displaystyle\alpha =𝔡¯t¯28u24σa24λat (48,48),\displaystyle=\bar{\mathfrak{d}}_{\bar{t}_{2}}^{8}u_{24\sigma}a_{24\lambda}\quad\text{at $(48,48)$},
b32\displaystyle b^{32} =u32λ/a32λat (64,64).\displaystyle=u_{32\lambda}/a_{32\lambda}\quad\text{at $(64,-64)$.}

We will use the following crucial facts about the spectral sequence of aλ1MU((C4))a_{\lambda}^{-1}MU^{(\!(C_{4})\!)}. These results are the only significant computational inputs from the localized slice spectral sequence for aλ1MU((C4))a_{\lambda}^{-1}MU^{(\!(C_{4})\!)} and the much harder computations of [28].

Proposition 7.5.

The class α=𝔡¯t¯28u24σa24λ\alpha=\bar{\mathfrak{d}}_{\bar{t}_{2}}^{8}u_{24\sigma}a_{24\lambda} is a permanent cycle in the localized slice spectral sequence for aλ1MU((C4))a_{\lambda}^{-1}MU^{(\!(C_{4})\!)}. Consequently, the differentials are linear with respect to multiplication by α\alpha.

Proof.

This is a direct consequence of the Slice Differential Theorem of Hill–Hopkins–Ravenel [22, Theorem 9.9]. More precisely, all the differentials in the region on or above the line of slope 1 in the C4C_{4} localized slice spectral sequence for aλ1MU((C4))a_{\lambda}^{-1}MU^{(\!(C_{4})\!)} can be completely computed. ∎

Proposition 7.6.

In the localized slice spectral sequence for aλ1MU((C4))a_{\lambda}^{-1}MU^{(\!(C_{4})\!)}, we have the following facts:

  1. (1)

    The class b8=u8λ/a8λb^{8}=u_{8\lambda}/a_{8\lambda} supports a d15d_{15}-differential.

  2. (2)

    The class b16=u16λ/a16λb^{16}=u_{16\lambda}/a_{16\lambda} supports a d31d_{31}-differential.

  3. (3)

    The class b32=u32λ/a32λb^{32}=u_{32\lambda}/a_{32\lambda} is a d61d_{61}-cycle.

Proof.

All the claims are direct consequences of the computations for the C4C_{4} slice spectral sequence of BP((C4))2BP^{(\!(C_{4})\!)}\langle 2\rangle (see Table 1 in [28]). We will elaborate on each of them below.

For (1), the restriction of the class u8λu_{8\lambda} is u16σ2u_{16\sigma_{2}}, which supports the d31d_{31}-differential

d31(u16σ2)=v¯4C2a31σ2d_{31}(u_{16\sigma_{2}})=\bar{v}_{4}^{C_{2}}a_{31\sigma_{2}}

in the C2C_{2}-spectral sequence of MU((C4))MU^{(\!(C_{4})\!)}. Therefore by naturality of the restriction map, the class u8λu_{8\lambda} must support a differential of length at most 31. By a stem-wise computation, it is impossible for the class u8λu_{8\lambda} to support a differential of length 13\leq 13, as there are no possible targets. We will not carry out the details of this computation here, as [28, Section 7.3] already contains the arguments and results to show that u8λu_{8\lambda} supports a d15d_{15}-differential in the slice spectral sequence of MU((C4))MU^{(\!(C_{4})\!)}. It follows that the class u8λa8λ\frac{u_{8\lambda}}{a_{8\lambda}} must also support a d15d_{15}-differential in the aλa_{\lambda}-localized slice spectral sequence of aλ1MU((C4))a_{\lambda}^{-1}MU^{(\!(C_{4})\!)}. This proves (1).

For (2), the restriction of the class u16λu_{16\lambda} is u32σ2u_{32\sigma_{2}}, which supports the d63d_{63}-differential

d63(u32σ2)=v¯5C2a63σ2d_{63}(u_{32\sigma_{2}})=\bar{v}_{5}^{C_{2}}a_{63\sigma_{2}}

in the C2C_{2}-spectral sequence of MU((C4))MU^{(\!(C_{4})\!)}. Therefore by naturality of the restriction map, the class u16λu_{16\lambda} must support a differential of length at most 63. By a stem-wise computation, it is impossible for the class u16λu_{16\lambda} to support a differential of length 19\leq 19, as there are no possible targets. Again, we do not write down the details here, as the computation in [28] already shows that u16λu_{16\lambda} supports a d31d_{31}-differential in the slice spectral sequence of MU((C4))MU^{(\!(C_{4})\!)} (see the discussion in Section 11.2 and the chart in Section 13 of [28]). It follows that the class u16λa16λ\frac{u_{16\lambda}}{a_{16\lambda}} must also support a d31d_{31}-differential in the aλa_{\lambda}-localized slice spectral sequence of aλ1MU((C4))a_{\lambda}^{-1}MU^{(\!(C_{4})\!)}. This proves (2).

For (3), it is a consequence of the computation of aλ1BP((C4))2,2a_{\lambda}^{-1}BP^{(\!(C_{4})\!)}\langle 2,2\rangle that u32λu_{32\lambda} is a d61d_{61}-cycle in the slice spectral sequence of MU((C4))MU^{(\!(C_{4})\!)} (in fact, it can be shown that it supports the d63d_{63}-differential d63(u32λ)=N(v¯5a63σ2)u16λd_{63}(u_{32\lambda})=N(\bar{v}_{5}a_{63\sigma_{2}})u_{16\lambda}, but we will not prove it here). ∎

Multiplication by the classes α\alpha and b32b^{32} give the spectral sequence a large amount of structure which we will exploit in our computation. We describe this below, focusing on each class at a time.

7.3.1. Multiplication by α\alpha

The class α\alpha is extremely important for this computation. A key consequence of the behavior we describe here is that this allows us to compute differentials out of order, flipping back and forth between different pages of the spectral sequence without loosing the thread of its story. We make this precise now, starting with the following straightforward lemma.

Lemma 7.7.

In the C4C_{4}-localized slice spectral sequence of aλ1BP((C4))2,2a_{\lambda}^{-1}BP^{(\!(C_{4})\!)}\langle 2,2\rangle, on the E13E_{13}-page, we have:

  1. (1)

    Multiplication by α\alpha is injective.

  2. (2)

    Let xx be a class in bidegree (ts,s)(t-s,s). If

    s48(ts48)ands483(ts48)s-48\geq-(t-s-48)\quad\text{and}\quad s-48\leq 3(t-s-48)

    then xx is α\alpha divisible.

See Figure 9.

The key result is then the following:

Proposition 7.8.

Let r13r\geq 13. Suppose that dr(x)=yd_{r}(x)=y is a non-trivial differential on the ErE_{r}-page of the localized slice spectral sequence of aλ1BP((C4))2,2a_{\lambda}^{-1}BP^{(\!(C_{4})\!)}\langle 2,2\rangle. Then yy is α\alpha-free in the sense that no multiple αiy\alpha^{i}y is zero at ErE_{r}. Consequently, xx is also α\alpha-free.

Proof.

First, note that if yy is α\alpha-free, the linearity of the differentials with respect to multiplication by α\alpha implies that xx must also be α\alpha-free.

We prove that yy is α\alpha-free by induction on rr. If r=13r=13, the claim follows immediately from the fact that all classes are α\alpha-free at E13E_{13} (Lemma 7.7). Suppose that the claim holds for all r<rr^{\prime}<r, that dr(x)=yd_{r}(x)=y is a non-trivial differential and that yy is α\alpha-torsion. Then there exists i>0i>0, r<rr^{\prime}<r and zz such that

dr(z)=αiy.d_{r^{\prime}}(z)=\alpha^{i}y.

Choose a minimal such ii, so that αi1y\alpha^{i-1}y is non-zero at ErE_{r}. A comparison of degrees then implies that the bidegree of zz satisfies the conditions of Lemma 7.7 (2), so that zz is α\alpha-divisible. It cannot be the case that dr(z/α)=αi1yd_{r^{\prime}}(z/\alpha)=\alpha^{i-1}y since this contradicts the minimality of ii. So we must have that dr′′(z/α)=v0d_{r^{\prime\prime}}(z/\alpha)=v\neq 0 for some r′′<r<rr^{\prime\prime}<r^{\prime}<r. But by the induction hypothesis, vv is then α\alpha-free which means that dr′′(z)=αv0d_{r^{\prime\prime}}(z)=\alpha v\neq 0, which is also a contradiction. ∎

Remark 7.9.

We will show in Section 7.4.3 that α\alpha is killed by a d61d_{61}-differential. This will imply that for any permanent cycle xx, the class αx\alpha x must be hit by a differential of length at most 6161.

We now explain the upshot of Proposition 7.8. Given any class yy at E13E_{13}, there is a unique class xx which is not α\alpha-divisible (so is in the complement of the region of Lemma 7.7 (2)) and such that y=αixy=\alpha^{i}x for some i0i\geq 0. We say that xx generates an α\alpha-free family, where the family is the collection of classes {αixi0}\{\alpha^{i}x\mid i\geq 0\}.

Now, Proposition 7.8 implies that α\alpha-free families come in pairs: one family in the pair, generated by xx say, supports differentials which truncate the second family in the pair, generated by yy say. (In fact, by Remark 7.9, the differentials must be of the form dr(αix)=αi+1yd_{r}(\alpha^{i}x)=\alpha^{i+1}y.) All classes in the α\alpha-free family generated by xx are then gone at the Er+1E_{r+1}-page, having supported a no-trivial differential. The class yy is now α\alpha-torsion, and by Proposition 7.8, it cannot support any further differential. This allows us to discard yy from the rest of the computation, making the spectral sequence effectively sparser. Furthermore, we may now run differentials out of order if we find a unique possibility for pairing α\alpha-free families, even if this is through very long differentials.

Remark 7.10.

This is in fact a common behavior for spectral sequences. For example, what we have here is very similar to the situation explained in a certain elliptic spectral sequence [5, Section 6], where there the role of α\alpha is played by the class κ¯\bar{\kappa}.

7.3.2. Multiplication by b32b^{32}

Multiplication by the permanent cycle b32b^{32} acts as a periodicity generator for most of the spectral sequence. More precisely, we have:

Lemma 7.11.

Let r13r\geq 13. Multiplication by b32b^{32} is injective on the ErE_{r}-page for classes on or below the line of slope 1 through the origin.

It follows that if a differential has both source and target on or below the line of slope 1 through the origin, then dr(x)=yd_{r}(x)=y occurs if and only if dr(b32x)=b32yd_{r}(b^{32}x)=b^{32}y occurs. Differentials whose source and target are above the line of slope 1 through the origin are determined by the aσa_{\sigma}-localized spectral sequence. Some differentials, fall in neither category in the sense that they cross the line of slope one. That is, the source is on or below the line of slope one and the target is above. For these differentials, the target may be b32b^{32}-torsion while the source is not.

As one does the computation however, one sees that the target of such differentials have bidegree (ts,s)(t-s,s) such that

s(ts)+14.s\leq(t-s)+14.

This can be seen from the d13d_{13}-differentials that are obtained from the aσa_{\sigma}-localized spectral sequence and the structure of the E14E_{14}-page. Since the longest differential is a d61d_{61} and classes are concentrated in degrees with tst-s even, classes strictly below the line

s=(ts)60s=(t-s)-60

cannot support differentials that cross the line of slope 1 through the origin. Therefore, to completely determine the differentials of the spectral sequence using b32b^{32}-linearity and α\alpha-linearity, it is sufficient to determine:

  • The d13d_{13}’s with source on or above the line of slope 1 through the origin, all obtained from the aσa_{\sigma}-localized spectral sequence.

  • The differentials on classes with source of bidegree (ts,s)(t-s,s) where is in the rectangular region:

    s\displaystyle s (ts)\displaystyle\leq(t-s) s\displaystyle s (ts)\displaystyle\geq-(t-s)
    s\displaystyle s (ts)188\displaystyle\geq(t-s)-188 s\displaystyle s <(ts)+96\displaystyle<-(t-s)+96

This region is larger than what is needed in practice, but the goal of this discussion is simply to illustrate the strategy and make a rough estimate on what differentials need to be determined. As we go through the computation, we learn that the region that determines all differentials is in fact smaller but, a priori, this is not clear.

7.3.3. Summary

To summarize, we just have to focus on the classes in the shaded rectangular region of Figure 9 which is the union of a cone and a rectangle. Once we have figured out the fate of all the classes in this region, we can propagate by the classes α\alpha and b32b^{32} to obtain the rest of the differentials. Furthermore, once an α\alpha-multiple of a class gets truncated by a differential, that class can no longer support differentials and can be disregarded from future arguments.

Refer to caption
Figure 9. Features of the aλa_{\lambda}-localized slice spectral sequence of aλ1BP((C4))2,2a_{\lambda}^{-1}BP^{(\!(C_{4})\!)}\langle 2,2\rangle

7.4. Differentials of length at least 13

7.4.1. d13d_{13}-differentials

By degree reasons, the next possible differentials are the d13d_{13}-differentials.

The differentials on or above the line of slope 1 are all obtained by computing the aσa_{\sigma}-localized spectral sequence, as explained in Section 5.1. This spectral sequence is depicted in Figure 4. The differentials are summarized in the following proposition.

Proposition 7.12.

The d13d_{13}-differentials that are on or above the line of slope 1 are generated by

  1. (1)

    d13((𝔡¯t¯22u6σa6λ)i+4k)=𝔡¯t¯23u2σa9λa7σ(𝔡¯t¯22u6σa6λ)i1+4kd_{13}\left((\bar{\mathfrak{d}}_{\bar{t}_{2}}^{2}u_{6\sigma}a_{6\lambda})^{i+4k}\right)=\bar{\mathfrak{d}}_{\bar{t}_{2}}^{3}u_{2\sigma}a_{9\lambda}a_{7\sigma}\cdot(\bar{\mathfrak{d}}_{\bar{t}_{2}}^{2}u_{6\sigma}a_{6\lambda})^{i-1+4k}
    i=1,2i=1,2, k0k\geq 0.

  2. (2)

    d13(𝔡¯t¯22u4σa6λa2σ(𝔡¯t¯22u6σa6λ)i+4k)=𝔡¯t¯23a9λa9σ(𝔡¯t¯22u6σa6λ)i+4kd_{13}\left(\bar{\mathfrak{d}}_{\bar{t}_{2}}^{2}u_{4\sigma}a_{6\lambda}a_{2\sigma}\cdot(\bar{\mathfrak{d}}_{\bar{t}_{2}}^{2}u_{6\sigma}a_{6\lambda})^{i+4k}\right)=\bar{\mathfrak{d}}_{\bar{t}_{2}}^{3}a_{9\lambda}a_{9\sigma}\cdot(\bar{\mathfrak{d}}_{\bar{t}_{2}}^{2}u_{6\sigma}a_{6\lambda})^{i+4k}
    i=0,3i=0,3, k0k\geq 0.

  3. (3)

    d13(𝔡¯t¯25u14σa15λaσ(𝔡¯t¯22u6σa6λ)i+4k)=𝔡¯t¯26u10σa18λa8σ(𝔡¯t¯22u6σa6λ)i+4kd_{13}\left(\bar{\mathfrak{d}}_{\bar{t}_{2}}^{5}u_{14\sigma}a_{15\lambda}a_{\sigma}\cdot(\bar{\mathfrak{d}}_{\bar{t}_{2}}^{2}u_{6\sigma}a_{6\lambda})^{i+4k}\right)=\bar{\mathfrak{d}}_{\bar{t}_{2}}^{6}u_{10\sigma}a_{18\lambda}a_{8\sigma}\cdot(\bar{\mathfrak{d}}_{\bar{t}_{2}}^{2}u_{6\sigma}a_{6\lambda})^{i+4k}
    i=0,1i=0,1, k0k\geq 0.

To prove the d13d_{13}-differentials that are under the line of slope 1, we would like to first point out that the class 𝔡¯t¯2uλu2σa2λaσ\bar{\mathfrak{d}}_{\bar{t}_{2}}u_{\lambda}u_{2\sigma}a_{2\lambda}a_{\sigma} in bidegree (7,5)(7,5) is a permanent-cycle by degree reasons.

The class 𝔡¯t¯24u12σa12λ\bar{\mathfrak{d}}_{\bar{t}_{2}}^{4}u_{12\sigma}a_{12\lambda} in bidegree (24,24)(24,24) will also be important. By the Hill–Hopkins–Ravenel Slice differential theorem [22, Theorem 9.9]. This class supports the d13d_{13}-differential

d13(𝔡¯t¯24u12σa12λ)=𝔡¯t¯25u8σa15λa7σd_{13}(\bar{\mathfrak{d}}_{\bar{t}_{2}}^{4}u_{12\sigma}a_{12\lambda})=\bar{\mathfrak{d}}_{\bar{t}_{2}}^{5}u_{8\sigma}a_{15\lambda}a_{7\sigma}

in the slice spectral sequence of BP((C4))BP^{(\!(C_{4})\!)}. By naturality, this differential also appears in the slice spectral sequence of BP((C4))2,2BP^{(\!(C_{4})\!)}\langle 2,2\rangle. When applying the Leibniz rule, the class 𝔡¯t¯24u12σa12λ\bar{\mathfrak{d}}_{\bar{t}_{2}}^{4}u_{12\sigma}a_{12\lambda} (24,24)(24,24) acts as if it is a d13d_{13}-cycle for differentials whose sources are below the line of slope 1. More specifically, the target of the d13d_{13}-differential on this class multiplied with the source of another d13d_{13}-differential below the line of slope 1 is always 0.

Proposition 7.13.

We have the following d13d_{13}-differentials:

  1. (1)

    d13(b4)=𝔡¯t¯2uλu2σa2λaσd_{13}\left(b^{4}\right)=\bar{\mathfrak{d}}_{\bar{t}_{2}}u_{\lambda}u_{2\sigma}a_{2\lambda}a_{\sigma}

  2. (2)

    d13(b5)=𝔡¯t¯2u2λu2σaλaσd_{13}\left(b^{5}\right)=\bar{\mathfrak{d}}_{\bar{t}_{2}}u_{2\lambda}u_{2\sigma}a_{\lambda}a_{\sigma}

  3. (3)

    d13(𝔡¯t¯2u4λu2σaλaσ)=2𝔡¯t¯22u6σa6λd_{13}\left(\bar{\mathfrak{d}}_{\bar{t}_{2}}u_{4\lambda}u_{2\sigma}a_{-\lambda}a_{\sigma}\right)=2\bar{\mathfrak{d}}_{\bar{t}_{2}}^{2}u_{6\sigma}a_{6\lambda}

  4. (4)

    d13(𝔡¯t¯2u4λu2σaλaσb)=2𝔡¯t¯22u6σa6λbd_{13}\left(\bar{\mathfrak{d}}_{\bar{t}_{2}}u_{4\lambda}u_{2\sigma}a_{-\lambda}a_{\sigma}\cdot b\right)=2\bar{\mathfrak{d}}_{\bar{t}_{2}}^{2}u_{6\sigma}a_{6\lambda}\cdot b

  5. (5)

    d13(𝔡¯t¯2u4λu2σaλaσb2)=2𝔡¯t¯22u6σa6λb2d_{13}\left(\bar{\mathfrak{d}}_{\bar{t}_{2}}u_{4\lambda}u_{2\sigma}a_{-\lambda}a_{\sigma}\cdot b^{2}\right)=2\bar{\mathfrak{d}}_{\bar{t}_{2}}^{2}u_{6\sigma}a_{6\lambda}\cdot b^{2}

  6. (6)

    d13(𝔡¯t¯2u4λu2σaλaσb3)=2𝔡¯t¯22u6σa6λb3d_{13}\left(\bar{\mathfrak{d}}_{\bar{t}_{2}}u_{4\lambda}u_{2\sigma}a_{-\lambda}a_{\sigma}\cdot b^{3}\right)=2\bar{\mathfrak{d}}_{\bar{t}_{2}}^{2}u_{6\sigma}a_{6\lambda}\cdot b^{3}

Proof.

To prove (1), we will first prove the differential

d13(b12)=𝔡¯t¯2u9λu2σa6λaσ(d13(24,24)=(23,11))d_{13}\left(b^{12}\right)=\bar{\mathfrak{d}}_{\bar{t}_{2}}u_{9\lambda}u_{2\sigma}a_{-6\lambda}a_{\sigma}\,\,\,(d_{13}(24,-24)=(23,-11))

The source of this differential restricts to a class in the C2C_{2}-spectral sequence that supports a d31d_{31}-differential. By naturality and degree reasons, we must have the d13d_{13}-differential claimed above. Applying Leibniz with the class b8b^{8} in degree (16,16)(16,-16) proves (1).

The source of (2) restricts to a class that supports a d19d_{19}-differential in the C2C_{2}-spectral sequence. Therefore the source class must support either a d13d_{13}- or a d19d_{19}-differential. By naturality, it cannot be a d19d_{19}-differential because the target does not restrict to the target of the d19d_{19}-differential in the C2C_{2}-spectral sequence.

The targets of (3) is in the image of the transfer. The preimage is killed by a d19d_{19}-differential. Therefore by naturality and degree reasons, we have the d13d_{13}-differential claimed in (3).

Differentials (4) and (5) are obtained by applying the Leibniz rule using the class 𝔡¯t¯2uλu2σa2λaσ\bar{\mathfrak{d}}_{\bar{t}_{2}}u_{\lambda}u_{2\sigma}a_{2\lambda}a_{\sigma} (7,5)(7,5) with differentials (1) and (2) (and also using the gold relation).

It remains to prove differential (6). Consider the class 𝔡¯t¯2u7λu2σa4λ\bar{\mathfrak{d}}_{\bar{t}_{2}}u_{7\lambda}u_{2\sigma}a_{-4\lambda}. The restriction of this class is t¯2γt¯2u14σ2a8σ2\bar{t}_{2}\gamma\bar{t}_{2}u_{14\sigma_{2}}a_{-8\sigma_{2}}, which supports a d7d_{7}-differential in the C2C_{2}-spectral sequence. Therefore, the class 𝔡¯t¯2u7λu2σa4λ\bar{\mathfrak{d}}_{\bar{t}_{2}}u_{7\lambda}u_{2\sigma}a_{-4\lambda} also supports a d7d_{7}-differential in the C4C_{4}-spectral sequence. The existence of this d7d_{7}-differential shows that there is an exotic restriction of filtration 6 for the class 𝔡¯t¯2u7λu2σa4λaσ\bar{\mathfrak{d}}_{\bar{t}_{2}}u_{7\lambda}u_{2\sigma}a_{-4\lambda}a_{\sigma} (19,7)(19,-7). It must have nonzero restriction, restricting to the class t¯23u10σ2aσ2\bar{t}_{2}^{3}u_{10\sigma_{2}}a_{-\sigma_{2}} (19,1)(19,-1) after the E7E_{7}-page.

Since the class t¯23u10σ2aσ2\bar{t}_{2}^{3}u_{10\sigma_{2}}a_{-\sigma_{2}} (19,1)(19,-1) supports a d19d_{19}-differential in the C2C_{2}-spectral sequence, the class 𝔡¯t¯2u7λu2σa4λaσ\bar{\mathfrak{d}}_{\bar{t}_{2}}u_{7\lambda}u_{2\sigma}a_{-4\lambda}a_{\sigma} (19,7)(19,-7) cannot survive past the E25E_{25}-page. The only possibility is for it to support the d13d_{13}-differential claimed by (6). ∎

The same proof for differentials (1)–(6) can be used to prove six more differentials that are obtained by multiplying both the source and the target of each differential by (12,12)(12,12): 𝔡¯t¯2u6σa6λ\bar{\mathfrak{d}}_{\bar{t}_{2}}u_{6\sigma}a_{6\lambda} (note that we can’t just directly propagate by this class using the Leibniz rule because it supports a d5d_{5}-differential in SliceSS(BP((C4)))\operatorname{SliceSS}(BP^{(\!(C_{4})\!)})).

Proposition 7.14.

The following classes are d13d_{13}-cycles:

  1. (1)

    2b62b^{6} (12,12)(12,-12);

  2. (2)

    2𝔡¯t¯22u6λu6σ2\bar{\mathfrak{d}}_{\bar{t}_{2}}^{2}u_{6\lambda}u_{6\sigma} (24,0)(24,0);

  3. (3)

    𝔡¯t¯23u8λu8σaλaσ\bar{\mathfrak{d}}_{\bar{t}_{2}}^{3}u_{8\lambda}u_{8\sigma}a_{\lambda}a_{\sigma} (33,3)(33,3).

Proof.

For (1), the class 2𝔡¯t¯24u6λu12σa6λ2\bar{\mathfrak{d}}_{\bar{t}_{2}}^{4}u_{6\lambda}u_{12\sigma}a_{6\lambda} (36,12)(36,12) is a d13d_{13}-cycle by using the class 𝔡¯t¯2uλu2σa2λaσ\bar{\mathfrak{d}}_{\bar{t}_{2}}u_{\lambda}u_{2\sigma}a_{2\lambda}a_{\sigma} (7,5)(7,5) to apply the Leibniz rule to the d13d_{13}-differential

d13(𝔡¯t¯23u6λu8σa3λaσ)=2𝔡¯t¯24u2λu12σa10λ(d13(29,7)=(28,20)).d_{13}(\bar{\mathfrak{d}}_{\bar{t}_{2}}^{3}u_{6\lambda}u_{8\sigma}a_{3\lambda}a_{\sigma})=2\bar{\mathfrak{d}}_{\bar{t}_{2}}^{4}u_{2\lambda}u_{12\sigma}a_{10\lambda}\quad(d_{13}(29,7)=(28,20)).

Therefore, by Leibniz with the class 𝔡¯t¯24u12σa12λ\bar{\mathfrak{d}}_{\bar{t}_{2}}^{4}u_{12\sigma}a_{12\lambda} (24,24)(24,24), the class 2b62b^{6} (12,12)(12,-12) is also a d13d_{13}-cycle.

(2) is proven by the exact same method, by using the class 𝔡¯t¯2uλu2σa2λaσ\bar{\mathfrak{d}}_{\bar{t}_{2}}u_{\lambda}u_{2\sigma}a_{2\lambda}a_{\sigma} (7,5)(7,5) to apply the Leibniz rule to the d13d_{13}-differential

d13(𝔡¯t¯2u6λu2σa3λaσ)=2𝔡¯t¯22u2λu6σa4λ(d13(17,5)=(16,8)).d_{13}(\bar{\mathfrak{d}}_{\bar{t}_{2}}u_{6\lambda}u_{2\sigma}a_{-3\lambda}a_{\sigma})=2\bar{\mathfrak{d}}_{\bar{t}_{2}}^{2}u_{2\lambda}u_{6\sigma}a_{4\lambda}\quad(d_{13}(17,-5)=(16,8)).

For (3), the class 𝔡¯t¯2u8λu8σaλaσ\bar{\mathfrak{d}}_{\bar{t}_{2}}u_{8\lambda}u_{8\sigma}a_{\lambda}a_{\sigma} is a d13d_{13}-cycle in SliceSS(BP((C4)))\operatorname{SliceSS}(BP^{(\!(C_{4})\!)}). Therefore by naturality it is a d13d_{13}-cycle in SliceSS(BP((C4))2,2)\operatorname{SliceSS}(BP^{(\!(C_{4})\!)}\langle 2,2\rangle). ∎

Now, we can propagate all the differentials by the classes 𝔡¯t¯22u12σa12λ\bar{\mathfrak{d}}_{\bar{t}_{2}}^{2}u_{12\sigma}a_{12\lambda} (24,24)(24,24) and b8b^{8} (16,16)(16,-16). The d13d_{13}-differentials under the line of slope 1 are summarized in the following proposition.

Proposition 7.15.

The d13d_{13}-differentials that are under the line of slope 1 are

  1. (1)

    d13(b4+i+8j(𝔡¯t¯22u6σa6λ)k)=𝔡¯t¯2uλu2σa2λaσbi+8j(𝔡¯t¯22u6σa6λ)kd_{13}\left(b^{4+i+8j}\cdot(\bar{\mathfrak{d}}_{\bar{t}_{2}}^{2}u_{6\sigma}a_{6\lambda})^{k}\right)=\bar{\mathfrak{d}}_{\bar{t}_{2}}u_{\lambda}u_{2\sigma}a_{2\lambda}a_{\sigma}\cdot b^{i+8j}(\bar{\mathfrak{d}}_{\bar{t}_{2}}^{2}u_{6\sigma}a_{6\lambda})^{k},
    0i10\leq i\leq 1, j,k0j,k\geq 0.

  2. (2)

    d13(𝔡¯t¯2u4λu2σaλaσbi+8j(𝔡¯t¯22u6σa6λ)k)=2𝔡¯t¯22u6σa6λbi+8j(𝔡¯t¯22u6σa6λ)kd_{13}\left(\bar{\mathfrak{d}}_{\bar{t}_{2}}u_{4\lambda}u_{2\sigma}a_{-\lambda}a_{\sigma}\cdot b^{i+8j}(\bar{\mathfrak{d}}_{\bar{t}_{2}}^{2}u_{6\sigma}a_{6\lambda})^{k}\right)=2\bar{\mathfrak{d}}_{\bar{t}_{2}}^{2}u_{6\sigma}a_{6\lambda}\cdot b^{i+8j}(\bar{\mathfrak{d}}_{\bar{t}_{2}}^{2}u_{6\sigma}a_{6\lambda})^{k},
    0i30\leq i\leq 3, j,k0j,k\geq 0.

They are shown in Figure 10.

Remark 7.16.

We will see that these are the last non-trivial d13d_{13} differentials. However, at this point in the computation, there are possibilities for other non-trivial d13d_{13} differentials. Later, (in Lemmas 7.21 and 7.24) we will show that these do not occur.

Refer to caption
Figure 10. The E13E_{13}-page of the aλa_{\lambda}-localized slice spectral sequence of aλ1BP((C4))2,2a_{\lambda}^{-1}BP^{(\!(C_{4})\!)}\langle 2,2\rangle.

7.4.2. d19d_{19}-differentials

Proposition 7.17.

The following d19d_{19}-differentials exist:

  1. (1)

    d19(2b5)=tr(t¯23a9σ2)d_{19}\left(2b^{5}\right)=tr(\bar{t}_{2}^{3}a_{9\sigma_{2}}) (d19(10,10)=(9,9)d_{19}(10,-10)=(9,9))

  2. (2)

    d19(2b6)=tr(t¯23u2σ2a7σ2)d_{19}\left(2b^{6}\right)=tr(\bar{t}_{2}^{3}u_{2\sigma_{2}}a_{7\sigma_{2}}) (d19(12,12)=(11,7)d_{19}(12,-12)=(11,7))

  3. (3)

    d19(b9)=tr(t¯23u8σ2aσ2)d_{19}\left(b^{9}\right)=tr(\bar{t}_{2}^{3}u_{8\sigma_{2}}a_{\sigma_{2}}) (d19(18,18)=(17,1)d_{19}(18,-18)=(17,1))

  4. (4)

    d19(2b13)=tr(t¯23u16σ2a7σ2)d_{19}\left(2b^{13}\right)=tr(\bar{t}_{2}^{3}u_{16\sigma_{2}}a_{-7\sigma_{2}}) (d19(26,26)=(25,7)d_{19}(26,-26)=(25,-7))

Proof.

Differential (1) is obtained by applying transfer to the d19d_{19}-differential

d19(u10σ2a10σ2)=t¯23a9σ2d_{19}\left(\frac{u_{10\sigma_{2}}}{a_{10\sigma_{2}}}\right)=\bar{t}_{2}^{3}a_{9\sigma_{2}}

in the C2C_{2}-spectral sequence.

For differentials (2) and (3), the classes t¯23u2σ2a7σ2\bar{t}_{2}^{3}u_{2\sigma_{2}}a_{7\sigma_{2}} and t¯23u8σ2aσ2\bar{t}_{2}^{3}u_{8\sigma_{2}}a_{\sigma_{2}} are killed by d31d_{31}-differentials in the C2C_{2}-spectral sequence. Therefore their images under the transfer map must also be killed by differentials of lengths at most 31. The only possibilities are the differentials claimed.

Differential (4) is obtained by applying the transfer to the d19d_{19}-differential

d19(u26σ2a26σ2)=t¯23u16σ2a7σ2d_{19}\left(\frac{u_{26\sigma_{2}}}{a_{26\sigma_{2}}}\right)=\bar{t}_{2}^{3}u_{16\sigma_{2}}a_{-7\sigma_{2}}

in the C2C_{2}-spectral sequence. ∎

The same arguments in the proof above can be used to prove twelve more d19d_{19}-differentials, obtained by multiplying the four d19d_{19}-differentials in Proposition 7.17 by 𝔡¯t¯22u6λa6λ\bar{\mathfrak{d}}_{\bar{t}_{2}}^{2}u_{6\lambda}a_{6\lambda} (12,12)(12,12), 𝔡¯t¯24u12λa12λ\bar{\mathfrak{d}}_{\bar{t}_{2}}^{4}u_{12\lambda}a_{12\lambda} (24,24)(24,24), and 𝔡¯t¯26u18λa18λ\bar{\mathfrak{d}}_{\bar{t}_{2}}^{6}u_{18\lambda}a_{18\lambda} (36,36)(36,36).

Proposition 7.18.

The following d19d_{19}-differentials exist:

  1. (1)

    d19(2b14)=tr(t¯23u18σ2a9σ2)d_{19}(2b^{14})=tr(\bar{t}_{2}^{3}u_{18\sigma_{2}}a_{-9\sigma_{2}}) (d19(28,28)=(27,9))(d_{19}(28,-28)=(27,-9))

  2. (2)

    d19(2b14(𝔡¯t¯22u6σa6λ))=tr(t¯27u18σ2a3σ2)d_{19}\left(2b^{14}\cdot(\bar{\mathfrak{d}}_{\bar{t}_{2}}^{2}u_{6\sigma}a_{6\lambda})\right)=tr(\bar{t}_{2}^{7}u_{18\sigma_{2}}a_{3\sigma_{2}}) (d19(40,16)=(39,3))(d_{19}(40,-16)=(39,3))

  3. (3)

    d19(2b14(𝔡¯t¯22u6σa6λ)2)=tr(t¯211u18σ2a15σ2)d_{19}\left(2b^{14}\cdot(\bar{\mathfrak{d}}_{\bar{t}_{2}}^{2}u_{6\sigma}a_{6\lambda})^{2}\right)=tr(\bar{t}_{2}^{11}u_{18\sigma_{2}}a_{15\sigma_{2}}) (d19(52,4)=(51,15))(d_{19}(52,-4)=(51,15))

  4. (4)

    d19(2b14(𝔡¯t¯22u6σa6λ)3)=tr(t¯215u18σ2a27σ2)d_{19}\left(2b^{14}\cdot(\bar{\mathfrak{d}}_{\bar{t}_{2}}^{2}u_{6\sigma}a_{6\lambda})^{3}\right)=tr(\bar{t}_{2}^{15}u_{18\sigma_{2}}a_{27\sigma_{2}}) (d19(64,8)=(63,27))(d_{19}(64,8)=(63,27))

Proof.

We will prove differential (1) first. Consider the class tr(t¯219u18σ2a39σ2)tr(\bar{t}_{2}^{19}u_{18\sigma_{2}}a_{39\sigma_{2}}) (75,39)(75,39). This class must die on or before the E61E_{61}-page. There are three possible ways for this class to die. It can support a d31d_{31}-differential hitting the class 𝔡¯t¯212uλu36σa35λ\bar{\mathfrak{d}}_{\bar{t}_{2}}^{12}u_{\lambda}u_{36\sigma}a_{35\lambda} (74,70)(74,70); it can be the target of a d19d_{19}-differential from the class 2𝔡¯t¯28u14λu24σa10λ2\bar{\mathfrak{d}}_{\bar{t}_{2}}^{8}u_{14\lambda}u_{24\sigma}a_{10\lambda} (76,20)(76,20); or it can be the target of a d43d_{43}-differential from the class 2𝔡¯t¯26u20λu18σa2λ2\bar{\mathfrak{d}}_{\bar{t}_{2}}^{6}u_{20\lambda}u_{18\sigma}a_{-2\lambda} (76,4)(76,-4).

It is impossible for this class to support a d31d_{31}-differential because it is the transfer of a class that supports a d31d_{31}-differential in the C2C_{2}-slice spectral sequence, and the target does not transfer to the class 𝔡¯t¯212uλu36σa35λ\bar{\mathfrak{d}}_{\bar{t}_{2}}^{12}u_{\lambda}u_{36\sigma}a_{35\lambda} (74,70)(74,70).

The d43d_{43}-differential also cannot happen because the class 2𝔡¯t¯26u20λu18σa2λ2\bar{\mathfrak{d}}_{\bar{t}_{2}}^{6}u_{20\lambda}u_{18\sigma}a_{-2\lambda} (76,4)(76,-4) is the transfer of t¯212u40σ2a4σ2\bar{t}_{2}^{12}u_{40\sigma_{2}}a_{-4\sigma_{2}}, which is the target of a d31d_{31}-differential in the C2C_{2}-spectral sequence. Therefore it must be killed by a differential of length at most 31.

It follows that the d19d_{19}-differential

d19(2𝔡¯t¯28u14λu24σa10λ)=tr(t¯219u18σ2a39σ2)(d19(76,20)=(75,39))d_{19}(2\bar{\mathfrak{d}}_{\bar{t}_{2}}^{8}u_{14\lambda}u_{24\sigma}a_{10\lambda})=tr(\bar{t}_{2}^{19}u_{18\sigma_{2}}a_{39\sigma_{2}})\,\,\,(d_{19}(76,20)=(75,39))

exists. Applying Leibniz with respect to the class 𝔡¯t¯24u24σa24λ\bar{\mathfrak{d}}_{\bar{t}_{2}}^{4}u_{24\sigma}a_{24\lambda} (48,48)(48,48) proves (1).

Differentials (2), (3), (4) are proven by the exact same method. ∎

Now, we can propagate the d19d_{19}-differentials that we have proven by the classes b16b^{16} (32,32)(32,-32) and 𝔡¯t¯24u24σa24λ\bar{\mathfrak{d}}_{\bar{t}_{2}}^{4}u_{24\sigma}a_{24\lambda} (48,48)(48,48) to obtain the rest of the d19d_{19}-differentials.

Proposition 7.19.

The d19d_{19}-differentials are

  1. (1)

    d19(2bi+8j(𝔡¯t¯22u6σa6λ)k)=tr(t¯23a9σ2)bi5+8j(𝔡¯t¯22u6σa6λ)kd_{19}\left(2b^{i+8j}\cdot(\bar{\mathfrak{d}}_{\bar{t}_{2}}^{2}u_{6\sigma}a_{6\lambda})^{k}\right)=tr(\bar{t}_{2}^{3}a_{9\sigma_{2}})\cdot b^{i-5+8j}(\bar{\mathfrak{d}}_{\bar{t}_{2}}^{2}u_{6\sigma}a_{6\lambda})^{k}
    i=5,6i=5,6, j,k0j,k\geq 0

  2. (2)

    d19(b9+16i(𝔡¯t¯22u6σa6λ)k)=tr(t¯23u8σ2aσ2)b16i(𝔡¯t¯22u6σa6λ)kd_{19}\left(b^{9+16i}\cdot(\bar{\mathfrak{d}}_{\bar{t}_{2}}^{2}u_{6\sigma}a_{6\lambda})^{k}\right)=tr(\bar{t}_{2}^{3}u_{8\sigma_{2}}a_{\sigma_{2}})\cdot b^{16i}(\bar{\mathfrak{d}}_{\bar{t}_{2}}^{2}u_{6\sigma}a_{6\lambda})^{k}
    i,k0i,k\geq 0.

They are shown in Figure 11.

Refer to caption
Figure 11. The E19E_{19}-page of the aλa_{\lambda}-localized slice spectral sequence of aλ1BP((C4))2,2a_{\lambda}^{-1}BP^{(\!(C_{4})\!)}\langle 2,2\rangle.

7.4.3. The vanishing theorem

Theorem 7.20 (Vanishing Theorem).

In the aλa_{\lambda}-localized slice spectral sequence for aλ1BP((C4))2,2a_{\lambda}^{-1}BP^{(\!(C_{4})\!)}\langle 2,2\rangle, we have the d61d_{61}-differential

d61(𝔡¯t¯23u16λu8σa7λaσ)=𝔡¯t¯28u24σa24λ(d61(49,13)=(48,48)).d_{61}(\bar{\mathfrak{d}}_{\bar{t}_{2}}^{3}u_{16\lambda}u_{8\sigma}a_{-7\lambda}a_{\sigma})=\bar{\mathfrak{d}}_{\bar{t}_{2}}^{8}u_{24\sigma}a_{24\lambda}\,\,\,(d_{61}(49,-13)=(48,48)).

Furthermore, any class of the form (𝔡¯t¯28u24σa24λ)x(\bar{\mathfrak{d}}_{\bar{t}_{2}}^{8}u_{24\sigma}a_{24\lambda})\cdot x must die on or before the E61E_{61}-page.

Proof.

In the aλa_{\lambda}-localized slice spectral sequence of aλ1BP((C4))a_{\lambda}^{-1}BP^{(\!(C_{4})\!)}, the class N(v¯4)u16σa31λN(\bar{v}_{4})u_{16\sigma}a_{31\lambda} must die on or before the E61E_{61}-page because it is the target of the predicted d61d_{61}-differential

d61(u16λaσ)=N(v¯4)u16σa31λ,d_{61}\left(u_{16\lambda}a_{\sigma}\right)=N(\bar{v}_{4})u_{16\sigma}a_{31\lambda},

obtained by norming up the d31d_{31}-differential d31(u16σ2)=v¯4a31σ2d_{31}(u_{16\sigma_{2}})=\bar{v}_{4}a_{31\sigma_{2}} in the C2C_{2}-spectral sequence. Therefore, if we multiply the target by 𝔡¯t¯211u32σa32λ\bar{\mathfrak{d}}_{\bar{t}_{2}}^{11}u_{32\sigma}a_{32\lambda}, the class 𝔡¯t¯211N(v¯4)u48σu48λ\bar{\mathfrak{d}}_{\bar{t}_{2}}^{11}N(\bar{v}_{4})u_{48\sigma}u_{48\lambda} (96,96)(96,96) must die on or before the E61E_{61}-page.

Under the map

aλ1SliceSS(BP((C4)))aλ1SliceSS(BP((C4))2,2),a_{\lambda}^{-1}\operatorname{SliceSS}(BP^{(\!(C_{4})\!)})\longrightarrow a_{\lambda}^{-1}\operatorname{SliceSS}(BP^{(\!(C_{4})\!)}\langle 2,2\rangle),

the class 𝔡¯t¯211N(v¯4)u48σu48λ\bar{\mathfrak{d}}_{\bar{t}_{2}}^{11}N(\bar{v}_{4})u_{48\sigma}u_{48\lambda} is sent to 𝔡¯t¯216u48σu48λ\bar{\mathfrak{d}}_{\bar{t}_{2}}^{16}u_{48\sigma}u_{48\lambda} (96,96)(96,96). By naturality and degree reasons, the only possibility that this class can die on or before the E61E_{61}-page is for it to be killed by a d61d_{61}-differential. This implies that the original class must also be killed by a d61d_{61}-differential in the aλa_{\lambda}-localized slice spectral sequence of aλ1BP((C4))a_{\lambda}^{-1}BP^{(\!(C_{4})\!)}. Furthermore, by the module structure, any class in the aλa_{\lambda}-localized slice spectral sequence of aλ1BP((C4))2,2a_{\lambda}^{-1}BP^{(\!(C_{4})\!)}\langle 2,2\rangle of the form (𝔡¯t¯216u48σu48λ)x(\bar{\mathfrak{d}}_{\bar{t}_{2}}^{16}u_{48\sigma}u_{48\lambda})\cdot x must die on or before the E61E_{61}-page.

The class 𝔡¯t¯28u24σa24λ\bar{\mathfrak{d}}_{\bar{t}_{2}}^{8}u_{24\sigma}a_{24\lambda} (48,48)(48,48) is also the target of a d61d_{61}-differential because after multiplying it by 𝔡¯t¯216u48σa48λ\bar{\mathfrak{d}}_{\bar{t}_{2}}^{16}u_{48\sigma}a_{48\lambda} (96,96)(96,96), it must die on or before the E61E_{61}-page. By degree reasons, the only possibility is for it to be killed by a d61d_{61}-differential. Since multiplication by 𝔡¯t¯216u48σa48λ\bar{\mathfrak{d}}_{\bar{t}_{2}}^{16}u_{48\sigma}a_{48\lambda} (96,96)(96,96) induces an injection on the E2E_{2}-page, and all the classes above the line of slope (1)(-1) with this class as the origin are all divisible by it, the claimed d61d_{61}-differential must occur.

Similarly, for any class of the form (𝔡¯t¯28u24σa24λ)x(\bar{\mathfrak{d}}_{\bar{t}_{2}}^{8}u_{24\sigma}a_{24\lambda})\cdot x, we can multiply it by 𝔡¯t¯216u48σa48λ\bar{\mathfrak{d}}_{\bar{t}_{2}}^{16}u_{48\sigma}a_{48\lambda} (96,96)(96,96) to deduce that the product must die on or before the E61E_{61}-page. It follows from the same reasoning as the previous paragraph that the original class must also die on or before the E61E_{61}-page. ∎

7.4.4. d31d_{31}-differentials

To prove the d31d_{31}-differentials, we will first prove the nonexistence of certain d13d_{13}-differentials.

Lemma 7.21.

At the E13E_{13}-page, we have

  1. (1)

    d13(𝔡¯t¯2u8λu2σa5λaσ)2𝔡¯t¯22u4λu6σa2λd_{13}(\bar{\mathfrak{d}}_{\bar{t}_{2}}u_{8\lambda}u_{2\sigma}a_{-5\lambda}a_{\sigma})\neq 2\bar{\mathfrak{d}}_{\bar{t}_{2}}^{2}u_{4\lambda}u_{6\sigma}a_{2\lambda} (d13(21,9)(20,4))(d_{13}(21,-9)\neq(20,4)).

  2. (2)

    d13(𝔡¯t¯23u8λu8σaλaσ)2𝔡¯t¯24u4λu12σa8λd_{13}(\bar{\mathfrak{d}}_{\bar{t}_{2}}^{3}u_{8\lambda}u_{8\sigma}a_{\lambda}a_{\sigma})\neq 2\bar{\mathfrak{d}}_{\bar{t}_{2}}^{4}u_{4\lambda}u_{12\sigma}a_{8\lambda} (d13(33,3)(32,16))(d_{13}(33,3)\neq(32,16)).

Proof.

Suppose (1) exists. By applying the Leibniz rule with respect to the classes 𝔡¯t¯24u12σa12λ\bar{\mathfrak{d}}_{\bar{t}_{2}}^{4}u_{12\sigma}a_{12\lambda} (24,24)(24,24) and b8b^{8} (16,16)(16,-16), the d13d_{13}-differential

d13(𝔡¯t¯25u16λu14σaλaσ)=2𝔡¯t¯26u12λu18σa6λ(d13(61,1)=(60,12))d_{13}(\bar{\mathfrak{d}}_{\bar{t}_{2}}^{5}u_{16\lambda}u_{14\sigma}a_{-\lambda}a_{\sigma})=2\bar{\mathfrak{d}}_{\bar{t}_{2}}^{6}u_{12\lambda}u_{18\sigma}a_{6\lambda}\quad(d_{13}(61,-1)=(60,12))

must also exist. Consider the class 𝔡¯t¯29u3λu26σa24λaσ\bar{\mathfrak{d}}_{\bar{t}_{2}}^{9}u_{3\lambda}u_{26\sigma}a_{24\lambda}a_{\sigma} in (59,49)(59,49). By Theorem 7.20, this class must die on or before the E61E_{61}-page. However, with the class 2𝔡¯t¯26u12λu18σa6λ2\bar{\mathfrak{d}}_{\bar{t}_{2}}^{6}u_{12\lambda}u_{18\sigma}a_{6\lambda} (60,12)(60,12) gone, there are no classes that could kill it or be killed by this class on or before the E61E_{61}-page. Contradiction.

Now, suppose (2) exists. By applying the Leibniz rule with respect to the classes 𝔡¯t¯24u12σa12λ\bar{\mathfrak{d}}_{\bar{t}_{2}}^{4}u_{12\sigma}a_{12\lambda} (24,24)(24,24) and b8b^{8} (16,16)(16,-16), the d13d_{13}-differential

d13(𝔡¯t¯27u16λu20σa5λaσ)=2𝔡¯t¯28u12λu24σa12λ(d13(73,11)=(72,24))d_{13}(\bar{\mathfrak{d}}_{\bar{t}_{2}}^{7}u_{16\lambda}u_{20\sigma}a_{5\lambda}a_{\sigma})=2\bar{\mathfrak{d}}_{\bar{t}_{2}}^{8}u_{12\lambda}u_{24\sigma}a_{12\lambda}\,\,\,(d_{13}(73,11)=(72,24))

must also exist. Consider the class 𝔡¯t¯211u3λu32σa30λaσ\bar{\mathfrak{d}}_{\bar{t}_{2}}^{11}u_{3\lambda}u_{32\sigma}a_{30\lambda}a_{\sigma} (71,61)(71,61). By Theorem 7.20, this class must die on or before the E61E_{61}-page. Just like the previous case, there is no class that could kill it or be killed by it on or before the E61E_{61}-page. Contradiction. ∎

Proposition 7.22.

We have the following d31d_{31}-differentials:

  1. (1)

    d31(tr(t¯211u10σ2a23σ2))=𝔡¯t¯28u18σa24λa6σd_{31}\left(tr(\bar{t}_{2}^{11}u_{10\sigma_{2}}a_{23\sigma_{2}})\right)=\bar{\mathfrak{d}}_{\bar{t}_{2}}^{8}u_{18\sigma}a_{24\lambda}a_{6\sigma};

  2. (2)

    d31(tr(t¯211u10σ2a23σ2)(𝔡¯t¯22u6σa6λ))=𝔡¯t¯28u18σa24λa6σ(𝔡¯t¯22u6σa6λ)d_{31}\left(tr(\bar{t}_{2}^{11}u_{10\sigma_{2}}a_{23\sigma_{2}})\cdot(\bar{\mathfrak{d}}_{\bar{t}_{2}}^{2}u_{6\sigma}a_{6\lambda})\right)=\bar{\mathfrak{d}}_{\bar{t}_{2}}^{8}u_{18\sigma}a_{24\lambda}a_{6\sigma}\cdot(\bar{\mathfrak{d}}_{\bar{t}_{2}}^{2}u_{6\sigma}a_{6\lambda});

  3. (3)

    d31(tr(t¯211u24σ2a9σ2)(𝔡¯t¯22u6σa6λ)i)=2𝔡¯t¯28u4λu24σa20λ(𝔡¯t¯22u6σa6λ)id_{31}\left(tr(\bar{t}_{2}^{11}u_{24\sigma_{2}}a_{9\sigma_{2}})\cdot(\bar{\mathfrak{d}}_{\bar{t}_{2}}^{2}u_{6\sigma}a_{6\lambda})^{i}\right)=2\bar{\mathfrak{d}}_{\bar{t}_{2}}^{8}u_{4\lambda}u_{24\sigma}a_{20\lambda}\cdot(\bar{\mathfrak{d}}_{\bar{t}_{2}}^{2}u_{6\sigma}a_{6\lambda})^{i},
    0i30\leq i\leq 3;

  4. (4)

    d31(tr(t¯211u24σ2a9σ2)b16(𝔡¯t¯22u6σa6λ)i)=2𝔡¯t¯28u4λu24σa20λb16(𝔡¯t¯22u6σa6λ)id_{31}\left(tr(\bar{t}_{2}^{11}u_{24\sigma_{2}}a_{9\sigma_{2}})\cdot b^{16}(\bar{\mathfrak{d}}_{\bar{t}_{2}}^{2}u_{6\sigma}a_{6\lambda})^{i}\right)=2\bar{\mathfrak{d}}_{\bar{t}_{2}}^{8}u_{4\lambda}u_{24\sigma}a_{20\lambda}\cdot b^{16}(\bar{\mathfrak{d}}_{\bar{t}_{2}}^{2}u_{6\sigma}a_{6\lambda})^{i},
    0i30\leq i\leq 3.

Proof.

To prove (1), first multiply the predicted target, 𝔡¯t¯28u18σa24λa6σ\bar{\mathfrak{d}}_{\bar{t}_{2}}^{8}u_{18\sigma}a_{24\lambda}a_{6\sigma} (42,54)(42,54), by 𝔡¯t¯216u48σa48λ\bar{\mathfrak{d}}_{\bar{t}_{2}}^{16}u_{48\sigma}a_{48\lambda} (96,96)(96,96). By Theorem 7.20 and degree reasons, the product must be killed by a differential of length 61. It follows that (1) must hold.

By Theorem 7.20 and degree reasons, the target of (2) must be killed by a differential of length at most 61. The only possible differential is the ones claimed.

To prove (3), note that in the aσ2a_{\sigma_{2}}-localized slice spectral sequence of iC2BP((C4))2,2i_{C_{2}}^{*}BP^{(\!(C_{4})\!)}\langle 2,2\rangle, we have the differential

d31(t¯211u24σ2a9σ2)=t¯216u8σ2a40σ2(d31(57,9)=(56,40)).d_{31}(\bar{t}_{2}^{11}u_{24\sigma_{2}}a_{9\sigma_{2}})=\bar{t}_{2}^{16}u_{8\sigma_{2}}a_{40\sigma_{2}}\,\,\,(d_{31}(57,9)=(56,40)).

Applying transfer to the target shows that the image of the target under the transfer map must be killed on or before the E31E_{31}-page. There are only two possibilities. Either the claimed d31d_{31}-differential exists, or it is killed by a d13d_{13}-differential from 𝔡¯t¯27u8λu20σa13λaσ\bar{\mathfrak{d}}_{\bar{t}_{2}}^{7}u_{8\lambda}u_{20\sigma}a_{13\lambda}a_{\sigma} (57,27)(57,27). By Lemma 7.21, the d13d_{13}-differential does not exist. Therefore the claimed d31d_{31}-differential happens for i=0i=0. The rest of the differentials in (3) and all the differentials in (4) are proven by the same method. ∎

We can propagate the differentials in Proposition 7.22 with respect to the classes 𝔡¯t¯28u24σa24λ\bar{\mathfrak{d}}_{\bar{t}_{2}}^{8}u_{24\sigma}a_{24\lambda} (48,48)(48,48) and b32b^{32} (64,64)(64,-64) to obtain the rest of the d31d_{31}-differentials.

Proposition 7.23.

The d31d_{31}-differentials are

  1. (1)

    d31(tr(t¯211u10σ2a23σ2)(𝔡¯t¯22u6σa6λ)i+4j)=𝔡¯t¯28u18σa24λa6σ(𝔡¯t¯22u6σa6λ)i+4jd_{31}\left(tr(\bar{t}_{2}^{11}u_{10\sigma_{2}}a_{23\sigma_{2}})\cdot(\bar{\mathfrak{d}}_{\bar{t}_{2}}^{2}u_{6\sigma}a_{6\lambda})^{i+4j}\right)=\bar{\mathfrak{d}}_{\bar{t}_{2}}^{8}u_{18\sigma}a_{24\lambda}a_{6\sigma}\cdot(\bar{\mathfrak{d}}_{\bar{t}_{2}}^{2}u_{6\sigma}a_{6\lambda})^{i+4j},
    i=0,1i=0,1, j0j\geq 0;

  2. (2)

    d31(tr(t¯23u24σ2a15σ2)b16i(𝔡¯t¯22u6σa6λ)j)=2𝔡¯t¯24u4λu12σa8λb16i(𝔡¯t¯22u6σa6λ)jd_{31}\left(tr(\bar{t}_{2}^{3}u_{24\sigma_{2}}a_{-15\sigma_{2}})\cdot b^{16i}(\bar{\mathfrak{d}}_{\bar{t}_{2}}^{2}u_{6\sigma}a_{6\lambda})^{j}\right)=2\bar{\mathfrak{d}}_{\bar{t}_{2}}^{4}u_{4\lambda}u_{12\sigma}a_{8\lambda}\cdot b^{16i}(\bar{\mathfrak{d}}_{\bar{t}_{2}}^{2}u_{6\sigma}a_{6\lambda})^{j},
    i,j0i,j\geq 0.

They are shown in Figure 12.

Refer to caption
Figure 12. The E31E_{31}-page of the aλa_{\lambda}-localized slice spectral sequence of aλ1BP((C4))2,2a_{\lambda}^{-1}BP^{(\!(C_{4})\!)}\langle 2,2\rangle.

7.4.5. d37d_{37}-differentials

To prove the d37d_{37}-differentials, we will first prove the nonexistence of certain d13d_{13}-differentials.

Lemma 7.24.

At the E13E_{13}-page, we have

  1. (1)

    d13(𝔡¯t¯212uλu36σa35λ)𝔡¯t¯213u34σa39λa5σd_{13}\left(\bar{\mathfrak{d}}_{\bar{t}_{2}}^{12}u_{\lambda}u_{36\sigma}a_{35\lambda}\right)\neq\bar{\mathfrak{d}}_{\bar{t}_{2}}^{13}u_{34\sigma}a_{39\lambda}a_{5\sigma};

  2. (2)

    d13(𝔡¯t¯212uλu36σa35λ(𝔡¯t¯22u6σa6λ))𝔡¯t¯213u34σa39λa5σ(𝔡¯t¯22u6σa6λ)d_{13}\left(\bar{\mathfrak{d}}_{\bar{t}_{2}}^{12}u_{\lambda}u_{36\sigma}a_{35\lambda}\cdot(\bar{\mathfrak{d}}_{\bar{t}_{2}}^{2}u_{6\sigma}a_{6\lambda})\right)\neq\bar{\mathfrak{d}}_{\bar{t}_{2}}^{13}u_{34\sigma}a_{39\lambda}a_{5\sigma}\cdot(\bar{\mathfrak{d}}_{\bar{t}_{2}}^{2}u_{6\sigma}a_{6\lambda});

  3. (3)

    d13(𝔡¯t¯211u3λu32σa30λaσ)𝔡¯t¯212u34σa36λa2σd_{13}\left(\bar{\mathfrak{d}}_{\bar{t}_{2}}^{11}u_{3\lambda}u_{32\sigma}a_{30\lambda}a_{\sigma}\right)\neq\bar{\mathfrak{d}}_{\bar{t}_{2}}^{12}u_{34\sigma}a_{36\lambda}a_{2\sigma};

  4. (4)

    d13(𝔡¯t¯211u3λu32σa30λaσ(𝔡¯t¯22u6σa6λ))𝔡¯t¯212u34σa36λa2σ(𝔡¯t¯22u6σa6λ)d_{13}\left(\bar{\mathfrak{d}}_{\bar{t}_{2}}^{11}u_{3\lambda}u_{32\sigma}a_{30\lambda}a_{\sigma}\cdot(\bar{\mathfrak{d}}_{\bar{t}_{2}}^{2}u_{6\sigma}a_{6\lambda})\right)\neq\bar{\mathfrak{d}}_{\bar{t}_{2}}^{12}u_{34\sigma}a_{36\lambda}a_{2\sigma}\cdot(\bar{\mathfrak{d}}_{\bar{t}_{2}}^{2}u_{6\sigma}a_{6\lambda});

  5. (5)

    d13(𝔡¯t¯29u19λu26σa8λaσ(𝔡¯t¯22u6σa6λ)i)2𝔡¯t¯210u15λu30σa15λ(𝔡¯t¯22u6σa6λ)id_{13}\left(\bar{\mathfrak{d}}_{\bar{t}_{2}}^{9}u_{19\lambda}u_{26\sigma}a_{8\lambda}a_{\sigma}\cdot(\bar{\mathfrak{d}}_{\bar{t}_{2}}^{2}u_{6\sigma}a_{6\lambda})^{i}\right)\neq 2\bar{\mathfrak{d}}_{\bar{t}_{2}}^{10}u_{15\lambda}u_{30\sigma}a_{15\lambda}\cdot(\bar{\mathfrak{d}}_{\bar{t}_{2}}^{2}u_{6\sigma}a_{6\lambda})^{i},
    0i30\leq i\leq 3.

Proof.

To prove (1), note that if the class 𝔡¯t¯212uλu36σa35λ\bar{\mathfrak{d}}_{\bar{t}_{2}}^{12}u_{\lambda}u_{36\sigma}a_{35\lambda} (74,70)(74,70) supports a d13d_{13}-differential, then applying the Leibniz rule with respect to the class 𝔡¯t¯24u12σa12λ\bar{\mathfrak{d}}_{\bar{t}_{2}}^{4}u_{12\sigma}a_{12\lambda} (24,24)(24,24) would show that the class 𝔡¯t¯28uλu24σa23λ\bar{\mathfrak{d}}_{\bar{t}_{2}}^{8}u_{\lambda}u_{24\sigma}a_{23\lambda} (50,46)(50,46) must support a differential of length at most 13. This is a contradiction because there are no possible targets. The nonexistence of differentials (2), (3), and (4) can be proven by the same method. The differentials in (5) follows from (1)-(4) by applying the Leibniz rule with respect to 𝔡¯t¯24u12σa12λ\bar{\mathfrak{d}}_{\bar{t}_{2}}^{4}u_{12\sigma}a_{12\lambda} (24,24)(24,24) and b16b^{16} (32,32)(32,-32). ∎

Proposition 7.25.

We have the following d37d_{37}-differentials for i=0,1i=0,1:

d37(𝔡¯t¯25u27λu14σa12λaσ(𝔡¯t¯22u6σa6λ)i)=𝔡¯t¯28u17λu24σa7λ(𝔡¯t¯22u6σa6λ)i.d_{37}(\bar{\mathfrak{d}}_{\bar{t}_{2}}^{5}u_{27\lambda}u_{14\sigma}a_{-12\lambda}a_{\sigma}\cdot(\bar{\mathfrak{d}}_{\bar{t}_{2}}^{2}u_{6\sigma}a_{6\lambda})^{i})=\bar{\mathfrak{d}}_{\bar{t}_{2}}^{8}u_{17\lambda}u_{24\sigma}a_{7\lambda}\cdot(\bar{\mathfrak{d}}_{\bar{t}_{2}}^{2}u_{6\sigma}a_{6\lambda})^{i}.
Proof.

To prove the differential when i=0i=0, we will show that the d37d_{37}-differential

d37(𝔡¯t¯213u27λu38σa12λaσ)=𝔡¯t¯216u17λu48σa31λ(d37(131,25)=(130,62))d_{37}(\bar{\mathfrak{d}}_{\bar{t}_{2}}^{13}u_{27\lambda}u_{38\sigma}a_{12\lambda}a_{\sigma})=\bar{\mathfrak{d}}_{\bar{t}_{2}}^{16}u_{17\lambda}u_{48\sigma}a_{31\lambda}\,\,\,(d_{37}(131,25)=(130,62))

exists. Propagating with respect to the class 𝔡¯t¯28u24σa24λ\bar{\mathfrak{d}}_{\bar{t}_{2}}^{8}u_{24\sigma}a_{24\lambda} (48,48)(48,48) would then prove the desired differential. Note that by Theorem 7.20, the class 𝔡¯t¯216u17λu48σa31λ\bar{\mathfrak{d}}_{\bar{t}_{2}}^{16}u_{17\lambda}u_{48\sigma}a_{31\lambda} (130,62)(130,62) must die on or before the E61E_{61}-page. There are two possibilities: either it supports a d37d_{37}-differential hitting 𝔡¯t¯219u8λu56σa49λaσ\bar{\mathfrak{d}}_{\bar{t}_{2}}^{19}u_{8\lambda}u_{56\sigma}a_{49\lambda}a_{\sigma} (129,99)(129,99), or the claimed differential exists. Suppose the first case happens, then we claim there is no possibility for the class 𝔡¯t¯213u27λu38σa12λaσ\bar{\mathfrak{d}}_{\bar{t}_{2}}^{13}u_{27\lambda}u_{38\sigma}a_{12\lambda}a_{\sigma} (131,25)(131,25) to die on or before the E61E_{61}-page. This is because if the class does die, then the only possibility is for it to support a d13d_{13}-differential hitting 2𝔡¯t¯214u23λu42σa19λ2\bar{\mathfrak{d}}_{\bar{t}_{2}}^{14}u_{23\lambda}u_{42\sigma}a_{19\lambda} (130,38)(130,38). However, if this d13d_{13}-differential exists, then by applying the Leibniz rule with respect to the class 𝔡¯t¯24u12σa12λ\bar{\mathfrak{d}}_{\bar{t}_{2}}^{4}u_{12\sigma}a_{12\lambda} (24,24)(24,24), the class 𝔡¯t¯29u27λu26σaσ\bar{\mathfrak{d}}_{\bar{t}_{2}}^{9}u_{27\lambda}u_{26\sigma}a_{\sigma} (107,1)(107,1) must also support a differential of length at most 13. This is a contradiction because we must have the d37d_{37}-differential

d37(𝔡¯t¯29u27λu26σaσ)=𝔡¯t¯212u17λu36σa19λ(d37(107,1)=(106,38))d_{37}(\bar{\mathfrak{d}}_{\bar{t}_{2}}^{9}u_{27\lambda}u_{26\sigma}a_{\sigma})=\bar{\mathfrak{d}}_{\bar{t}_{2}}^{12}u_{17\lambda}u_{36\sigma}a_{19\lambda}\,\,\,(d_{37}(107,1)=(106,38))

by the Vanishing Theorem and degree reasons (Theorem 7.20). It follows that the class 𝔡¯t¯29u27λu26σaσ\bar{\mathfrak{d}}_{\bar{t}_{2}}^{9}u_{27\lambda}u_{26\sigma}a_{\sigma} (107,1)(107,1) supports a d37d_{37}-differential.

The second differential, when i=1i=1, is proven by the same method. ∎

Proposition 7.26.

The d37d_{37}-differentials are

  1. (1)

    d37(𝔡¯t¯24u8λu12σa4λ(𝔡¯t¯22u6σa6λ)i+4j)=𝔡¯t¯27u18σa21λa3σ(𝔡¯t¯22u6σa6λ)i+4jd_{37}\left(\bar{\mathfrak{d}}_{\bar{t}_{2}}^{4}u_{8\lambda}u_{12\sigma}a_{4\lambda}\cdot(\bar{\mathfrak{d}}_{\bar{t}_{2}}^{2}u_{6\sigma}a_{6\lambda})^{i+4j}\right)=\bar{\mathfrak{d}}_{\bar{t}_{2}}^{7}u_{18\sigma}a_{21\lambda}a_{3\sigma}\cdot(\bar{\mathfrak{d}}_{\bar{t}_{2}}^{2}u_{6\sigma}a_{6\lambda})^{i+4j},
    i=0,1i=0,1, j0j\geq 0;

  2. (2)

    d37(𝔡¯t¯2u8λu2σa5λaσ(𝔡¯t¯22u6σa6λ)i+4j)=𝔡¯t¯24u8σa12λa4σ(𝔡¯t¯22u6σa6λ)i+4jd_{37}\left(\bar{\mathfrak{d}}_{\bar{t}_{2}}u_{8\lambda}u_{2\sigma}a_{-5\lambda}a_{\sigma}\cdot(\bar{\mathfrak{d}}_{\bar{t}_{2}}^{2}u_{6\sigma}a_{6\lambda})^{i+4j}\right)=\bar{\mathfrak{d}}_{\bar{t}_{2}}^{4}u_{8\sigma}a_{12\lambda}a_{4\sigma}\cdot(\bar{\mathfrak{d}}_{\bar{t}_{2}}^{2}u_{6\sigma}a_{6\lambda})^{i+4j},
    i=0,3i=0,3, j0j\geq 0;

  3. (3)

    d37(2𝔡¯t¯22u7λu6σaλ(𝔡¯t¯22u6σa6λ)i+4j)=𝔡¯t¯25u10σa15λa5σ(𝔡¯t¯22u6σa6λ)i+4jd_{37}\left(2\bar{\mathfrak{d}}_{\bar{t}_{2}}^{2}u_{7\lambda}u_{6\sigma}a_{-\lambda}\cdot(\bar{\mathfrak{d}}_{\bar{t}_{2}}^{2}u_{6\sigma}a_{6\lambda})^{i+4j}\right)=\bar{\mathfrak{d}}_{\bar{t}_{2}}^{5}u_{10\sigma}a_{15\lambda}a_{5\sigma}\cdot(\bar{\mathfrak{d}}_{\bar{t}_{2}}^{2}u_{6\sigma}a_{6\lambda})^{i+4j},
    i=0,1i=0,1, j0j\geq 0;

  4. (4)

    d37(2b12+16i(𝔡¯t¯22u6σa6λ)j)=𝔡¯t¯23u3λu8σa6λaσb16i(𝔡¯t¯22u6σa6λ)jd_{37}\left(2b^{12+16i}\cdot(\bar{\mathfrak{d}}_{\bar{t}_{2}}^{2}u_{6\sigma}a_{6\lambda})^{j}\right)=\bar{\mathfrak{d}}_{\bar{t}_{2}}^{3}u_{3\lambda}u_{8\sigma}a_{6\lambda}a_{\sigma}\cdot b^{16i}(\bar{\mathfrak{d}}_{\bar{t}_{2}}^{2}u_{6\sigma}a_{6\lambda})^{j},
    i,j0i,j\geq 0;

  5. (5)

    d37(𝔡¯t¯2u11λu2σa8λaσb16i(𝔡¯t¯22u6σa6λ)j)=𝔡¯t¯24uλu12σa11λb16i(𝔡¯t¯22u6σa6λ)jd_{37}\left(\bar{\mathfrak{d}}_{\bar{t}_{2}}u_{11\lambda}u_{2\sigma}a_{-8\lambda}a_{\sigma}\cdot b^{16i}(\bar{\mathfrak{d}}_{\bar{t}_{2}}^{2}u_{6\sigma}a_{6\lambda})^{j}\right)=\bar{\mathfrak{d}}_{\bar{t}_{2}}^{4}u_{\lambda}u_{12\sigma}a_{11\lambda}\cdot b^{16i}(\bar{\mathfrak{d}}_{\bar{t}_{2}}^{2}u_{6\sigma}a_{6\lambda})^{j},
    i,j0i,j\geq 0.

They are shown in Figure 13.

Proof.

All the differentials can be proven immediately from the Vanishing Theorem and degree reasons (Theorem 7.20 and Lemma 7.24), Proposition 7.25, and propagation with respect to the classes 𝔡¯t¯28u24σa24λ\bar{\mathfrak{d}}_{\bar{t}_{2}}^{8}u_{24\sigma}a_{24\lambda} (48,48)(48,48) and b32b^{32} (64,64)(64,-64). ∎

Refer to caption
Figure 13. The E37E_{37}-page of the aλa_{\lambda}-localized slice spectral sequence of aλ1BP((C4))2,2a_{\lambda}^{-1}BP^{(\!(C_{4})\!)}\langle 2,2\rangle.

7.4.6. d43d_{43}-differentials

Proposition 7.27.

The following d43d_{43}-differentials exist for i=0,1i=0,1:

d43(𝔡¯t¯212u40λu36σa4λ(𝔡¯t¯22u6σa6λ)i)=tr(t¯231u58σ2a35σ2)(𝔡¯t¯22u6σa6λ)i.d_{43}(\bar{\mathfrak{d}}_{\bar{t}_{2}}^{12}u_{40\lambda}u_{36\sigma}a_{-4\lambda}\cdot(\bar{\mathfrak{d}}_{\bar{t}_{2}}^{2}u_{6\sigma}a_{6\lambda})^{i})=tr(\bar{t}_{2}^{31}u_{58\sigma_{2}}a_{35\sigma_{2}})\cdot(\bar{\mathfrak{d}}_{\bar{t}_{2}}^{2}u_{6\sigma}a_{6\lambda})^{i}.
Proof.

When i=0i=0, note that by the Vanishing Theorem (Theorem 7.20), the class tr(t¯231u58σ2a35σ2)tr(\bar{t}_{2}^{31}u_{58\sigma_{2}}a_{35\sigma_{2}}) (151,35)(151,35) must die on or before the E61E_{61}-page. There are two possibilities. Either the claimed differential occurs, or it supports a d55d_{55}-differential hitting 2𝔡¯t¯220u15λu60σa45λ2\bar{\mathfrak{d}}_{\bar{t}_{2}}^{20}u_{15\lambda}u_{60\sigma}a_{45\lambda} (150,90)(150,90). The second case does not occur because the class 2𝔡¯t¯220u15λu60σa45λ2\bar{\mathfrak{d}}_{\bar{t}_{2}}^{20}u_{15\lambda}u_{60\sigma}a_{45\lambda} (150,90)(150,90) needs to support a d61d_{61}-differential killing the class 𝔡¯t¯225u74σa75λaσ\bar{\mathfrak{d}}_{\bar{t}_{2}}^{25}u_{74\sigma}a_{75\lambda}a_{\sigma} (149,151)(149,151), or else no class would be able to kill 𝔡¯t¯225u74σa75λaσ\bar{\mathfrak{d}}_{\bar{t}_{2}}^{25}u_{74\sigma}a_{75\lambda}a_{\sigma} (149,151)(149,151) on or before the E61E_{61}-page and we would reach a contradiction with Theorem 7.20.

The second differential is proven by the same method. ∎

Proposition 7.28.

The d43d_{43}-differentials are

  1. (1)

    d43(b16+32i(𝔡¯t¯22u6σa6λ)j+4k)=tr(t¯27u10σ2a11σ2)b32i(𝔡¯t¯22u6σa6λ)j+4kd_{43}\left(b^{16+32i}\cdot(\bar{\mathfrak{d}}_{\bar{t}_{2}}^{2}u_{6\sigma}a_{6\lambda})^{j+4k}\right)=tr(\bar{t}_{2}^{7}u_{10\sigma_{2}}a_{11\sigma_{2}})\cdot b^{32i}(\bar{\mathfrak{d}}_{\bar{t}_{2}}^{2}u_{6\sigma}a_{6\lambda})^{j+4k},
    i,k0i,k\geq 0, j=0,3j=0,3;

  2. (2)

    d43(b24+32i(𝔡¯t¯22u6σa6λ)j+4k(𝔡¯t¯22u8λu6σa2λ))=tr(t¯27u26σ2a5σ2)b32i(𝔡¯t¯22u6σa6λ)j+4k(𝔡¯t¯22u8λu6σa2λ)d_{43}\left(b^{24+32i}\cdot(\bar{\mathfrak{d}}_{\bar{t}_{2}}^{2}u_{6\sigma}a_{6\lambda})^{j+4k}(\bar{\mathfrak{d}}_{\bar{t}_{2}}^{2}u_{8\lambda}u_{6\sigma}a_{-2\lambda})^{\ell}\right)\\ =tr(\bar{t}_{2}^{7}u_{26\sigma_{2}}a_{-5\sigma_{2}})\cdot b^{32i}(\bar{\mathfrak{d}}_{\bar{t}_{2}}^{2}u_{6\sigma}a_{6\lambda})^{j+4k}(\bar{\mathfrak{d}}_{\bar{t}_{2}}^{2}u_{8\lambda}u_{6\sigma}a_{-2\lambda})^{\ell},
    i,k0i,k\geq 0, j=0,1j=0,1, =0,1,2\ell=0,1,2.

They are shown in Figure 14.

Proof.

All the differentials can be proven immediately from the Vanishing Theorem and degree reasons (Theorem 7.20), Proposition 7.27, and propagation with respect to the classes 𝔡¯t¯28u24σa24λ\bar{\mathfrak{d}}_{\bar{t}_{2}}^{8}u_{24\sigma}a_{24\lambda} (48,48)(48,48) and b32b^{32} (64,64)(64,-64). ∎

7.4.7. d55d_{55}-differentials

Proposition 7.29.

The d55d_{55}-differentials are

  1. (1)

    d55(tr(t¯23u26σ2a17σ2)b32i(𝔡¯t¯22u6σa6λ)j+4k)=𝔡¯t¯26u16σa18λa2σb32i(𝔡¯t¯22u6σa6λ)j+4kd_{55}\left(tr(\bar{t}_{2}^{3}u_{26\sigma_{2}}a_{-17\sigma_{2}})\cdot b^{32i}(\bar{\mathfrak{d}}_{\bar{t}_{2}}^{2}u_{6\sigma}a_{6\lambda})^{j+4k}\right)=\bar{\mathfrak{d}}_{\bar{t}_{2}}^{6}u_{16\sigma}a_{18\lambda}a_{2\sigma}\cdot b^{32i}(\bar{\mathfrak{d}}_{\bar{t}_{2}}^{2}u_{6\sigma}a_{6\lambda})^{j+4k},
    i,k0i,k\geq 0, j=0,3j=0,3;

  2. (2)

    d55(tr(t¯23u26σ2a17σ2)b8+32i(𝔡¯t¯22u6σa6λ)j+4k(𝔡¯t¯22u8λu6σa2λ))=𝔡¯t¯26u16σa18λa2σb8+32i(𝔡¯t¯22u6σa6λ)j+4k(𝔡¯t¯22u8λu6σa2λ)d_{55}\left(tr(\bar{t}_{2}^{3}u_{26\sigma_{2}}a_{-17\sigma_{2}})\cdot b^{8+32i}(\bar{\mathfrak{d}}_{\bar{t}_{2}}^{2}u_{6\sigma}a_{6\lambda})^{j+4k}(\bar{\mathfrak{d}}_{\bar{t}_{2}}^{2}u_{8\lambda}u_{6\sigma}a_{-2\lambda})^{\ell}\right)\\ =\bar{\mathfrak{d}}_{\bar{t}_{2}}^{6}u_{16\sigma}a_{18\lambda}a_{2\sigma}\cdot b^{8+32i}(\bar{\mathfrak{d}}_{\bar{t}_{2}}^{2}u_{6\sigma}a_{6\lambda})^{j+4k}(\bar{\mathfrak{d}}_{\bar{t}_{2}}^{2}u_{8\lambda}u_{6\sigma}a_{-2\lambda})^{\ell},
    i,k0i,k\geq 0, j=0,1j=0,1, =0,1,2\ell=0,1,2.

They are shown in Figure 14.

Proof.

All the differentials can be deduced from the Vanishing Theorem and degree reasons (Theorem 7.20), and propagation with respect to the classes 𝔡¯t¯28u24σa24λ\bar{\mathfrak{d}}_{\bar{t}_{2}}^{8}u_{24\sigma}a_{24\lambda} (48,48)(48,48) and b32b^{32} (64,64)(64,-64). ∎

7.4.8. d61d_{61}-differentials

Proposition 7.30.

We have the following d61d_{61}-differentials:

  1. (1)

    d61(𝔡¯t¯2u16λu2σa13λaσ(𝔡¯t¯22u8λu6σa2λ)i(𝔡¯t¯22u6σa6λ)j+4kb32)=𝔡¯t¯26u18σa18λ(𝔡¯t¯22u8λu6σa2λ)i(𝔡¯t¯22u6σa6λ)j+4kb32d_{61}\left(\bar{\mathfrak{d}}_{\bar{t}_{2}}u_{16\lambda}u_{2\sigma}a_{-13\lambda}a_{\sigma}\cdot(\bar{\mathfrak{d}}_{\bar{t}_{2}}^{2}u_{8\lambda}u_{6\sigma}a_{-2\lambda})^{i}(\bar{\mathfrak{d}}_{\bar{t}_{2}}^{2}u_{6\sigma}a_{6\lambda})^{j+4k}b^{32\ell}\right)\\ =\bar{\mathfrak{d}}_{\bar{t}_{2}}^{6}u_{18\sigma}a_{18\lambda}\cdot(\bar{\mathfrak{d}}_{\bar{t}_{2}}^{2}u_{8\lambda}u_{6\sigma}a_{-2\lambda})^{i}(\bar{\mathfrak{d}}_{\bar{t}_{2}}^{2}u_{6\sigma}a_{6\lambda})^{j+4k}b^{32\ell},
    (i,j)=(0,0),(0,1),(1,0),(1,1),(2,0),(2,1),(3,3),(3,0)(i,j)=(0,0),(0,1),(1,0),(1,1),(2,0),(2,1),(3,-3),(3,0), k,0k,\ell\geq 0;

  2. (2)

    d61(2𝔡¯t¯24u15λu12σa3λ(𝔡¯t¯22u8λu6σa2λ)i(𝔡¯t¯22u6σa6λ)j+4kb32)=𝔡¯t¯29u26σa27λaσ(𝔡¯t¯22u8λu6σa2λ)i(𝔡¯t¯22u6σa6λ)j+4kb32d_{61}\left(2\bar{\mathfrak{d}}_{\bar{t}_{2}}^{4}u_{15\lambda}u_{12\sigma}a_{-3\lambda}\cdot(\bar{\mathfrak{d}}_{\bar{t}_{2}}^{2}u_{8\lambda}u_{6\sigma}a_{-2\lambda})^{i}(\bar{\mathfrak{d}}_{\bar{t}_{2}}^{2}u_{6\sigma}a_{6\lambda})^{j+4k}b^{32\ell}\right)\\ =\bar{\mathfrak{d}}_{\bar{t}_{2}}^{9}u_{26\sigma}a_{27\lambda}a_{\sigma}\cdot(\bar{\mathfrak{d}}_{\bar{t}_{2}}^{2}u_{8\lambda}u_{6\sigma}a_{-2\lambda})^{i}(\bar{\mathfrak{d}}_{\bar{t}_{2}}^{2}u_{6\sigma}a_{6\lambda})^{j+4k}b^{32\ell},
    (i,j)=(0,0),(0,1),(1,3),(1,0),(2,4),(2,3),(3,4),(3,3)(i,j)=(0,0),(0,1),(1,-3),(1,0),(2,-4),(2,-3),(3,-4),(3,-3), k,0k,\ell\geq 0.

They are shown in Figure 14.

Proof.

All the differentials can be deduced from the Vanishing Theorem and degree reasons (Theorem 7.20), and propagation with respect to the classes 𝔡¯t¯28u24σa24λ\bar{\mathfrak{d}}_{\bar{t}_{2}}^{8}u_{24\sigma}a_{24\lambda} (48,48)(48,48) and b32b^{32} (64,64)(64,-64). ∎

Refer to caption
Figure 14. The d43d_{43} ( blue), d55d_{55} ( magenta), and d61d_{61}-differentials (black) in the aλa_{\lambda}-localized slice spectral sequence of aλ1BP((C4))2,2a_{\lambda}^{-1}BP^{(\!(C_{4})\!)}\langle 2,2\rangle.
Refer to caption
Figure 15. The EE_{\infty}-page of the aλa_{\lambda}-localized slice spectral sequence of aλ1BP((C4))2,2a_{\lambda}^{-1}BP^{(\!(C_{4})\!)}\langle 2,2\rangle.

References

  • [1] Vigleik Angeltveit. Topological Hochschild homology and cohomology of AA_{\infty} ring spectra. Geom. Topol., 12(2):987–1032, 2008.
  • [2] Shôrô Araki. Orientations in τ\tau-cohomology theories. Japan. J. Math. (N.S.), 5(2):403–430, 1979.
  • [3] Andrew Baker and Andrej Lazarev. On the Adams spectral sequence for RR-modules. Algebr. Geom. Topol., 1:173–199, 2001.
  • [4] Tilman Bauer. Computation of the homotopy of the spectrum tmf. In Groups, homotopy and configuration spaces, volume 13 of Geom. Topol. Monogr., pages 11–40. Geom. Topol. Publ., Coventry, 2008.
  • [5] Agnès Beaudry, Irina Bobkova, Viet-Cuong Pham, and Zhouli Xu. The topological modular forms of P2\mathbb{R}P^{2} and P2P2\mathbb{R}P^{2}\wedge\mathbb{C}P^{2}. J. Topol., 15(4):1864–1926, 2022.
  • [6] Agnès Beaudry, Michael A. Hill, Tyler Lawson, XiaoLin Danny Shi, and Mingcong Zeng. Quotient rings of H𝔽2H𝔽2{H}\mathbb{F}_{2}\wedge{H}\mathbb{F}_{2}. Trans. Amer. Math. Soc., 374(3):8949–8988, 2021.
  • [7] Agnès Beaudry, Michael A. Hill, XiaoLin Danny Shi, and Mingcong Zeng. Models of Lubin-Tate spectra via real bordism theory. Adv. Math., 392:Paper No. 108020, 58, 2021.
  • [8] Agnès Beaudry, Michael A. Hill, XiaoLin Danny Shi, and Mingcong Zeng. Transchromatic extensions in motivic bordism. Proc. Amer. Math. Soc. Ser. B, 10:76–90, 2023.
  • [9] Mark Behrens and Kyle Ormsby. On the homotopy of Q(3)Q(3) and Q(5)Q(5) at the prime 2. Algebr. Geom. Topol., 16(5):2459–2534, 2016.
  • [10] Irina Bobkova and Paul G. Goerss. Topological resolutions in K(2)K(2)-local homotopy theory at the prime 2. J. Topol., 11(4):918–957, 2018.
  • [11] E. S. Devinatz, M. J. Hopkins, and J. H. Smith. Nilpotence and stable homotopy theory. I. Ann. of Math. (2), 128(2):207–241, 1988.
  • [12] Christopher L. Douglas, John Francis, André G. Henriques, and Michael A. Hill, editors. Topological modular forms, volume 201 of Mathematical Surveys and Monographs. American Mathematical Society, Providence, RI, 2014.
  • [13] Zhipeng Duan, Guchuan Li, and XiaoLin Danny Shi. Vanishing lines in chromatic homotopy theory. arXiv e-prints, page arXiv:2204.08600, April 2022.
  • [14] A. D. Elmendorf, I. Kriz, M. A. Mandell, and J. P. May. Rings, modules, and algebras in stable homotopy theory, volume 47 of Mathematical Surveys and Monographs. American Mathematical Society, Providence, RI, 1997. With an appendix by M. Cole.
  • [15] Michikazu Fujii. Cobordism theory with reality. Math. J. Okayama Univ., 18(2):171–188, 1975.
  • [16] P. Goerss, H.-W. Henn, M. Mahowald, and C. Rezk. A resolution of the K(2)K(2)-local sphere at the prime 3. Ann. of Math. (2), 162(2):777–822, 2005.
  • [17] Jeremy Hahn and XiaoLin Danny Shi. Real orientations of Lubin-Tate spectra. Invent. Math., 221(3):731–776, 2020.
  • [18] Jeremy Hahn and Dylan Wilson. Quotients of even rings. arXiv e-prints, page arXiv:1809.04723, September 2018.
  • [19] Jeremy Hahn and Dylan Wilson. Real topological Hochschild homology and the Segal conjecture. Adv. Math., 387:Paper No. 107839, 17, 2021.
  • [20] Hans-Werner Henn. On finite resolutions of K(n)K(n)-local spheres. In Elliptic cohomology, volume 342 of London Math. Soc. Lecture Note Ser., pages 122–169. Cambridge Univ. Press, Cambridge, 2007.
  • [21] Thomas Hewett. Finite subgroups of division algebras over local fields. J. Algebra, 173(3):518–548, 1995.
  • [22] M. A. Hill, M. J. Hopkins, and D. C. Ravenel. On the nonexistence of elements of Kervaire invariant one. Ann. of Math. (2), 184(1):1–262, 2016.
  • [23] Michael Hill and Tyler Lawson. Automorphic forms and cohomology theories on Shimura curves of small discriminant. Adv. Math., 225(2):1013–1045, 2010.
  • [24] Michael A. Hill. The equivariant slice filtration: a primer. Homology Homotopy Appl., 14(2):143–166, 2012.
  • [25] Michael A. Hill, M. J. Hopkins, and D. C. Ravenel. The slice spectral sequence for certain RO(Cpn)RO(C_{p^{n}})-graded suspensions of H𝐙¯H\underline{\bf Z}. Bol. Soc. Mat. Mex. (3), 23(1):289–317, 2017.
  • [26] Michael A. Hill, Michael J. Hopkins, and Douglas C. Ravenel. The slice spectral sequence for the C4C_{4} analog of real KK-theory. Forum Math., 29(2):383–447, 2017.
  • [27] Michael A. Hill and Lennart Meier. The C2C_{2}-spectrum Tmf1(3){\rm Tmf}_{1}(3) and its invertible modules. Algebr. Geom. Topol., 17(4):1953–2011, 2017.
  • [28] Michael A Hill, XiaoLin Danny Shi, Guozhen Wang, and Zhouli Xu. The slice spectral sequence of a C4C_{4}-equivariant height-4 Lubin-Tate theory. Memoir of the American Mathematical Society, to appear, 2018.
  • [29] Michael J. Hopkins and Jeffrey H. Smith. Nilpotence and stable homotopy theory. II. Ann. of Math. (2), 148(1):1–49, 1998.
  • [30] Mark Hovey. Bousfield localization functors and Hopkins’ chromatic splitting conjecture. In The Čech centennial (Boston, MA, 1993), volume 181 of Contemp. Math., pages 225–250. Amer. Math. Soc., Providence, RI, 1995.
  • [31] Mark Hovey and Neil P. Strickland. Morava KK-theories and localisation. Mem. Amer. Math. Soc., 139(666):viii+100, 1999.
  • [32] Po Hu and Igor Kriz. Real-oriented homotopy theory and an analogue of the Adams-Novikov spectral sequence. Topology, 40(2):317–399, 2001.
  • [33] Nitu Kitchloo and W. Stephen Wilson. On the Hopf ring for ER(n)ER(n). Topology Appl., 154(8):1608–1640, 2007.
  • [34] Peter S. Landweber. Conjugations on complex manifolds and equivariant homotopy of MUMU. Bull. Amer. Math. Soc., 74:271–274, 1968.
  • [35] Guchuan Li, XiaoLin Danny Shi, Guozhen Wang, and Zhouli Xu. Hurewicz images of real bordism theory and real Johnson-Wilson theories. Adv. Math., 342:67–115, 2019.
  • [36] Mark Mahowald and Charles Rezk. Topological modular forms of level 3. Pure Appl. Math. Q., 5(2, Special Issue: In honor of Friedrich Hirzebruch. Part 1):853–872, 2009.
  • [37] Mark Mahowald and Hal Sadofsky. vnv_{n} telescopes and the Adams spectral sequence. Duke Math. J., 78(1):101–129, 1995.
  • [38] Lennart Meier, XiaoLin Danny Shi, and Mingcong Zeng. A stratification of the equivariant slice filtration. arXiv e-prints, page arXiv:2310.12105, October 2023.
  • [39] Lennart Meier, XiaoLin Danny Shi, and Mingcong Zeng. The localized slice spectral sequence, norms of Real bordism, and the Segal conjecture. Adv. Math., 412, 2023.
  • [40] D. C. Ravenel. Complex cobordism and stable homotopy groups of spheres, volume 121 of Pure and Applied Mathematics. Academic Press, Inc., Orlando, FL, 1986.
  • [41] Douglas C. Ravenel. The non-existence of odd primary Arf invariant elements in stable homotopy. Math. Proc. Cambridge Philos. Soc., 83(3):429–443, 1978.
  • [42] Alan Robinson. Obstruction theory and the strict associativity of Morava KK-theories. In Advances in homotopy theory (Cortona, 1988), volume 139 of London Math. Soc. Lecture Note Ser., pages 143–152. Cambridge Univ. Press, Cambridge, 1989.