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

On the CpC_{p}-equivariant dual Steenrod algebra

Krishanu Sankar and Dylan Wilson
Abstract

We compute the CpC_{p}-equivariant dual Steenrod algebras associated to the constant Mackey functors 𝔽¯p\underline{\mathbb{F}}_{p} and β„€Β―(p)\underline{\mathbb{Z}}_{(p)}, as β„€Β―(p)\underline{\mathbb{Z}}_{(p)}-modules. The CpC_{p}-spectrum 𝔽¯pβŠ—π”½Β―p\underline{\mathbb{F}}_{p}\otimes\underline{\mathbb{F}}_{p} is not a direct sum of R​O​(Cp)RO(C_{p})-graded suspensions of 𝔽¯p\underline{\mathbb{F}}_{p} when pp is odd, in contrast with the classical and C2C_{2}-equivariant dual Steenrod algebras.

Introduction

For over a decade, since the Hill-Hopkins-Ravenel solution of the Kervaire invariant one problem [HHR16], there has been great success in using exotic homotopy theories, like C2nC_{2^{n}}-equivariant homotopy theory and motivic homotopy theory, to study classical homotopy theory at the prime 2. A key foundational input to many of these applications is the computation of the appropriate version of the dual Steenrod algebra, 𝔽¯2βŠ—π”½Β―2\underline{\mathbb{F}}_{2}\otimes\underline{\mathbb{F}}_{2}, which was carried out by Hu-Kriz [HK01] in C2C_{2}-equivariant homotopy theory and by Voevodsky [Voe03] in motivic homotopy theory. One of the major obstacles to carrying out a similar program at odd primes is that we do not understand the structure of the dual Steenrod algebra in CpC_{p}-equivariant homotopy theory. The purpose of this paper is to make some progress towards this goal.

To motivate the statement of our main result, recall that we have the following description of the classical, pp-local dual Steenrod algebra as a β„€(p)\mathbb{Z}_{(p)}-algebra111We learned this fact from John Rognes. One proof is to base change the equivalence
BPβŠ—S0​[v1,…]S0≃℀(p)\mathrm{BP}\otimes_{S^{0}[v_{1},...]}S^{0}\simeq\mathbb{Z}_{(p)} to β„€(p)\mathbb{Z}_{(p)} and use that the Hurewicz image of the viv_{i}’s are p​tipt_{i}, mod decomposables.

β„€(p)βŠ—β„€(p)≃℀(p)βŠ—β¨‚icofib​(Ξ£|ti|​S0​[ti]βŸΆβ‹…p​tiS0​[ti]).\mathbb{Z}_{(p)}\otimes\mathbb{Z}_{(p)}\simeq\mathbb{Z}_{(p)}\otimes\bigotimes_{i}\mathrm{cofib}\left(\Sigma^{|t_{i}|}S^{0}[t_{i}]\stackrel{{\scriptstyle\cdot pt_{i}}}{{\longrightarrow}}S^{0}[t_{i}]\right).

Here the tensor product is taken over the sphere spectrum, S0​[x]S^{0}[x] denotes the free 𝔼1\mathbb{E}_{1}-algebra on a class xx, and the classes tit_{i} live in degree 2​piβˆ’22p^{i}-2. Modding out by pp causes each of the above cofibers to split into two classes related by a Bockstein; modding out by pp once more introduces the class Ο„0\tau_{0} and recovers Milnor’s computation of π’œβˆ—=Ο€βˆ—β€‹(𝔽pβŠ—π”½p)\mathcal{A}_{*}=\pi_{*}(\mathbb{F}_{p}\otimes\mathbb{F}_{p}), as an 𝔽p\mathbb{F}_{p}-algebra.

In the CpC_{p}-equivariant case our description involves a similar decomposition but is more complicated in two ways:

  • β€’

    Rather than extending the class tit_{i} to a map from S0​[ti]S^{0}[t_{i}] using the multiplication on β„€βŠ—β„€\mathbb{Z}\otimes\mathbb{Z}, we will want to choose as generators a mixture of ordinary powers of tit_{i} and of norms, N​(ti)N(t_{i}), of tit_{i}.

  • β€’

    Rather than modding out by the relation β€˜p​ti=0pt_{i}=0’ we will need to enforce the relation that β€˜ΞΈβ€‹ti=0\theta t_{i}=0’, where ΞΈ\theta is an equivariant lift of pp to an element in nontrivial R​O​(Cp)RO(C_{p})-degree. We will then also need to enforce the relation p​N​(ti)=0pN(t_{i})=0.

To make this precise, we will assume that the reader is comfortable with equivariant stable homotopy theory as used, for example, in [HHR16], and introduce the following conventions, in force throughout the paper:

  • β€’

    We will use ρCp\rho_{C_{p}} to denote the regular representation of CpC_{p}.

  • β€’

    We will use Ξ»\lambda to denote the representation of CpC_{p} on ℝ2=β„‚\mathbb{R}^{2}=\mathbb{C} where the generator acts by e2​π​i/pe^{2\pi i/p}.

  • β€’

    We denote by ΞΈ:SΞ»βˆ’2β†’S0\theta:S^{\lambda-2}\to S^{0} the map of CpC_{p}-spectra arising from the degree pp cover SΞ»β†’S2S^{\lambda}\to S^{2}. We’ll denote the cofiber of ΞΈ\theta by C​θC\theta. Note that the underlying nonequivariant spectrum of C​θC\theta is the Moore space M​(p)M(p).

  • β€’

    If XX is a spectrum, we will denote by N​(X)N(X) the Hill-Hopkins-Ravenel norm of XX, which is a CpC_{p}-equivariant refinement of the ordinary spectrum XβŠ—pX^{\otimes p}.

  • β€’

    We denote by β„€Β―\underline{\mathbb{Z}} and 𝔽¯p\underline{\mathbb{F}}_{p} the CpC_{p}-equivariant Eilenberg-MacLane spectra associated to the constant Mackey functors at β„€\mathbb{Z} and 𝔽p\mathbb{F}_{p}, respectively.

  • β€’

    We use βŠ—\otimes to denote the symmetric monoidal structure on genuine CpC_{p}-spectra, 𝖲𝗉Cp\mathsf{Sp}^{C_{p}} (often denoted by ∧\wedge, the smash product).

  • β€’

    The degree kk map

    SΞ»β†’SΞ»kS^{\lambda}\to S^{\lambda^{k}}

    is a pp-local equivalence when (k,p)=1(k,p)=1, so, when working pp-locally, we will often make this identification implicitly. For example, we may write

    S(p)ρCp=S(p)1+pβˆ’12​λ.S^{\rho_{C_{p}}}_{(p)}=S_{(p)}^{1+\frac{p-1}{2}\lambda}.
  • β€’

    We use π⋆​X\pi_{\star}X to denote the R​O​(Cp)RO(C_{p})-graded homotopy groups of a CpC_{p} spectrum, so that, when ⋆=Vβˆ’W\star=V-W is a virtual representation, Ο€Vβˆ’W​X=Ο€0​Map𝖲𝗉Cp​(SVβˆ’W,X)\pi_{V-W}X=\pi_{0}\mathrm{Map}_{\mathsf{Sp}^{C_{p}}}(S^{V-W},X).

Now we can give a somewhat ad-hoc description of the equivariant refinements of the building blocks in β„€βŠ—β„€\mathbb{Z}\otimes\mathbb{Z}.

Construction.

Let xx be a formal variable in an R​O​(Cp)RO(C_{p})-grading |x||x|. Define a CpC_{p}-spectrum as follows:

Tθ​(x):=Ξ£|x|​Cβ€‹ΞΈβŠ•Ξ£2​|x|​Cβ€‹ΞΈβŠ•β‹―βŠ•Ξ£(pβˆ’1)​|x|​Cβ€‹ΞΈβŠ•Ξ£|x|​ρCp​M​(p),T_{\theta}(x):=\Sigma^{|x|}C\theta\oplus\Sigma^{2|x|}C\theta\oplus\cdots\oplus\Sigma^{(p-1)|x|}C\theta\oplus\Sigma^{|x|\rho_{C_{p}}}M(p),

where M​(p)M(p) is the mod pp Moore spectrum. We denote the inclusion of the summand Ξ£i​|x|​C​θ\Sigma^{i|x|}C\theta by

xiβˆ’1​x^:Ξ£i​|x|​C​θ→Tθ​(x),x^{i-1}\hat{x}:\Sigma^{i|x|}C\theta\to T_{\theta}(x),

the restriction of x^\hat{x} to the bottom cell by xx, and the inclusion of the final summand by N​x^\widehat{Nx}. We denote by

N​x:S|x|​ρCpβ†’Tθ​(x)Nx:S^{|x|\rho_{C_{p}}}\to T_{\theta}(x)

the restriction of N​x^\widehat{Nx} to the bottom cell of the mod pp Moore spectrum.

Now suppose that RR is a CpC_{p}-ring spectrum equipped with a norm N​(R)β†’RN(R)\to R. If we have a class xβˆˆΟ€β‹†β€‹Rx\in\pi_{\star}R such that θ​x=0\theta x=0, it follows that pβ‹…N​(x)=0p\cdot N(x)=0 (see the proof of Lemma 4.4), so we may produce a map

S0βŠ•(S0​[N​x]βŠ—Tθ​(x))β†’R,S^{0}\oplus(S^{0}[Nx]\otimes T_{\theta}(x))\to R,

which only depends on the choice of the nullhomotopy witnessing θ​x=0\theta x=0.

We can now state our main theorem.

Theorem A.

There are equivariant refinements

ti:S2​piβˆ’1​ρCpβˆ’Ξ»β†’β„€Β―(p)βŠ—β„€Β―(p)t_{i}:S^{2p^{i-1}\rho_{C_{p}}-\lambda}\to\underline{\mathbb{Z}}_{(p)}\otimes\underline{\mathbb{Z}}_{(p)}

of the nonequivariant classes tiβˆˆΟ€βˆ—β€‹(β„€(p)βŠ—β„€(p))t_{i}\in\pi_{*}(\mathbb{Z}_{(p)}\otimes\mathbb{Z}_{(p)}) which satisfy the relation θ​ti=0\theta t_{i}=0. For any choice of witness for these relations, the resulting map

β„€Β―(p)βŠ—β¨‚iβ‰₯1(S0βŠ•(S0​[N​ti]βŠ—Tθ​(ti)))βŸΆβ„€Β―(p)βŠ—β„€Β―(p)\underline{\mathbb{Z}}_{(p)}\otimes\bigotimes_{i\geq 1}\left(S^{0}\oplus(S^{0}[Nt_{i}]\otimes T_{\theta}(t_{i}))\right)\longrightarrow\underline{\mathbb{Z}}_{(p)}\otimes\underline{\mathbb{Z}}_{(p)}

is an equivalence.

As an immediate corollary we have:

Corollary.

With notation as above, we have

𝔽¯pβŠ—π”½Β―p≃Λ​(Ο„0)βŠ—π”½Β―p𝔽¯pβŠ—β¨‚iβ‰₯1(S0βŠ•(S0​[N​ti]βŠ—Tθ​(ti))),\underline{\mathbb{F}}_{p}\otimes\underline{\mathbb{F}}_{p}\simeq\Lambda(\tau_{0})\otimes_{\underline{\mathbb{F}}_{p}}\underline{\mathbb{F}}_{p}\otimes\bigotimes_{i\geq 1}\left(S^{0}\oplus(S^{0}[Nt_{i}]\otimes T_{\theta}(t_{i}))\right),

where Ο„0\tau_{0} is dual to the Bockstein, in degree 11 and Λ​(Ο„0)=𝔽¯pβŠ•Ξ£β€‹π”½Β―p\Lambda(\tau_{0})=\underline{\mathbb{F}}_{p}\oplus\Sigma\underline{\mathbb{F}}_{p}. In particular, since 𝔽¯pβŠ—C​θ\underline{\mathbb{F}}_{p}\otimes C\theta is indecomposable at odd primes, the spectrum 𝔽¯pβŠ—π”½Β―p\underline{\mathbb{F}}_{p}\otimes\underline{\mathbb{F}}_{p} is not a direct sum of R​O​(Cp)RO(C_{p})-graded suspensions of 𝔽¯p\underline{\mathbb{F}}_{p} at odd primes.

Remark.

When p=2p=2 we have an accidental splitting 𝔽¯2βŠ—Cβ€‹ΞΈβ‰ƒΞ£Οƒβˆ’1​𝔽¯2βŠ•Ξ£Οƒβ€‹π”½Β―2\underline{\mathbb{F}}_{2}\otimes C\theta\simeq\Sigma^{\sigma-1}\underline{\mathbb{F}}_{2}\oplus\Sigma^{\sigma}\underline{\mathbb{F}}_{2}, where Οƒ\sigma is the sign representation.

Remark.

One can show that 𝔽¯pβŠ—Cβ€‹ΞΈβŠ—C​θ\underline{\mathbb{F}}_{p}\otimes C\theta\otimes C\theta splits as (𝔽¯pβŠ—C​θ)βŠ•(𝔽¯pβŠ—Ξ£Ξ»βˆ’1​C​θ)(\underline{\mathbb{F}}_{p}\otimes C\theta)\oplus(\underline{\mathbb{F}}_{p}\otimes\Sigma^{\lambda-1}C\theta). It follows that 𝔽¯pβŠ—π”½Β―p\underline{\mathbb{F}}_{p}\otimes\underline{\mathbb{F}}_{p} splits as a direct sum of cell complexes with at most 2 cells.

Our result raises a few natural questions which would be interesting to investigate.

Question 1.

When specialized to p=2p=2, how does our basis compare to the Hu-Kriz basis?

Question 2.

The geometric fixed points of β„€Β―(p)βŠ—β„€Β―(p)\underline{\mathbb{Z}}_{(p)}\otimes\underline{\mathbb{Z}}_{(p)} are given by (𝔽pβŠ—π”½p)​[b,bΒ―](\mathbb{F}_{p}\otimes\mathbb{F}_{p})[b,\overline{b}], where bΒ―\overline{b} is the conjugate of bb, a class in degree 22. It is possible to understand what happens to the generators tit_{i} and N​(ti)^\widehat{N(t_{i})} upon taking geometric fixed points. One is left with trying to understand the remaining class hit by t^i\hat{t}_{i} on geometric fixed points. We don’t know what this should be. One guess that seems consistent with computations is that this class is given, up to conjugating the Ο„i\tau_{i} and modding out by (b)(b), by:

Ο„iβˆ’1+bΒ―piβˆ’1βˆ’piβˆ’2​τiβˆ’2+β‹―+bΒ―piβˆ’1βˆ’1​τ0\tau_{i-1}+\overline{b}^{p^{i-1}-p^{i-2}}\tau_{i-2}+\cdots+\overline{b}^{p^{i-1}-1}\tau_{0}

It would be useful for computations to sort out what actually occurs.

Question 3.

Is it possible to profitably study the 𝔽¯p\underline{\mathbb{F}}_{p}-based Adams spectral sequence using this decomposition? Since 𝔽¯pβŠ—π”½Β―p\underline{\mathbb{F}}_{p}\otimes\underline{\mathbb{F}}_{p} is not flat over 𝔽¯p\underline{\mathbb{F}}_{p}, one would be forced to start with the E1E_{1}-term. But this is not an unprecedented situation (e.g. Mahowald had great success with the ko\mathrm{ko}-based Adams spectral sequence).

Question 4.

Can one describe the multiplication on π⋆​𝔽¯pβŠ—π”½Β―p\pi_{\star}\underline{\mathbb{F}}_{p}\otimes\underline{\mathbb{F}}_{p} in terms of our decomposition?

Relation to other work

As we mentioned before, we were very much motivated by the description of the C2C_{2}-equivariant dual Steenrod algebra given by Hu-Kriz [HK01]. That said, our generators are slightly different than the Hu-Kriz generators when we specialize to p=2p=2. For example, the generator t1t_{1} lives in degree 2​ρC2βˆ’Ξ»=22\rho_{C_{2}}-\lambda=2, whereas the Hu-Kriz generator ΞΎ1\xi_{1} lives in degree ρ=1+Οƒ\rho=1+\sigma.222In this low degree, it seems likely that, modulo decomposables, we have uσ​ξ1=t1u_{\sigma}\xi_{1}=t_{1} and that ΞΎ1\xi_{1} is recovered from t^1\hat{t}_{1} by restricting along 𝔽¯2βŠ—S1+σ→𝔽¯2βŠ—Ξ£2​C​θ\underline{\mathbb{F}}_{2}\otimes S^{1+\sigma}\to\underline{\mathbb{F}}_{2}\otimes\Sigma^{2}C\theta. Hill and Hopkins have also obtained a presentation of the C2nC_{2^{n}}-dual Steenrod algebra, using quotients of BP​𝐑\mathrm{BP}\mathbf{R} and its norms, which is similar in style to the one obtained here.

At odd primes, Caruso [Car99] studied the CpC_{p}-equivariant Steenrod algebra, π⋆​map​(𝔽p,𝔽p)\pi_{\star}\mathrm{map}(\mathbb{F}_{p},\mathbb{F}_{p}), essentially by comparing with the Borel equivariant Steenrod algebra and the geometric fixed point Steenrod algebra, and was able to compute the ranks of the integer-graded stems. There is also work of OruΓ§ [Oru89] computing the dual Steenrod algebra for the Eilenberg-MacLane spectra associated to Mackey fields (which does not include 𝔽¯p\underline{\mathbb{F}}_{p}).

In the Borel equivariant setting, the dual Steenrod algebra is given by the action Hopf algebroid for the coaction of the classical dual Steenrod algebra on Hβˆ—β€‹(B​Cp)H^{*}(\mathrm{B}C_{p}) (see [Gre88]).

There is also related work from the first and second authors. The first author produced a splitting of 𝔽¯pβŠ—π”½Β―p\underline{\mathbb{F}}_{p}\otimes\underline{\mathbb{F}}_{p} in [San19] using the symmetric power filtration. This summands in that splitting were roughly given by the homology of classifying spaces, and were much larger than the summands produced here. The second author and Jeremy Hahn showed [HW20] that 𝔽¯p\underline{\mathbb{F}}_{p} can be obtained as a Thom spectrum on Ωλ​SΞ»+1\Omega^{\lambda}S^{\lambda+1}. The Thom isomorphism then reduces the study of the dual Steenrod algebra to the computation of the homology of Ωλ​SΞ»+1\Omega^{\lambda}S^{\lambda+1}. Understanding the relationship between this picture and the one in this article is work in progress.

Acknowledgements

We thank Jeremy Hahn and Clover May for excellent discussions over the course of this project. We thank Mike Hill for generously sharing his time and his ideas, especially the idea to use the norm to build some of the generators of the dual Steenrod algebra. We apologize to those awaiting the publication of these results for the long delay.

1 Outline of the proof

To motivate our method of proof, let’s first revisit the classical story. We are interested in where the classes tiβˆˆΟ€βˆ—β€‹(β„€βŠ—β„€)t_{i}\in\pi_{*}(\mathbb{Z}\otimes\mathbb{Z}) come from, and why they are annihilated by pp.

Recall that the homology of ℂ​P∞\mathbb{C}P^{\infty} is a divided power algebra

Hβˆ—β€‹(ℂ​P∞)=Γ℀​{Ξ²1}H_{*}(\mathbb{C}P^{\infty})=\Gamma_{\mathbb{Z}}\{\beta_{1}\}

where Ξ²1\beta_{1} is dual to the first Chern class c1c_{1}. Write Ξ²(i):=Ξ³pi​(Ξ²1)\beta_{(i)}:=\gamma_{p^{i}}(\beta_{1}). Since ℂ​P∞=K​(β„€,2)\mathbb{C}P^{\infty}=K(\mathbb{Z},2), we have a map of spectra

ℂ​P+βˆžβ†’Ξ£2​℀\mathbb{C}P^{\infty}_{+}\to\Sigma^{2}\mathbb{Z}

and hence a homology suspension map

Οƒ:Hβˆ—β€‹(ℂ​P∞)β†’Ο€βˆ—βˆ’2​(β„€βŠ—β„€)\sigma:H_{*}(\mathbb{C}P^{\infty})\to\pi_{*-2}(\mathbb{Z}\otimes\mathbb{Z})

which annihilates elements decomposable with respect to the product structure on Hβˆ—β€‹(ℂ​P∞)H_{*}(\mathbb{C}P^{\infty}). We can take333Depending on ones preferences, this might be the conjugate of the generator you want; but we are only really concerned with these classes modulo decomposables. ti:=σ​(Ξ²(i))t_{i}:=\sigma(\beta_{(i)}). The relation p​ti=0pt_{i}=0 follows from the fact that p​β(i)p\beta_{(i)} is, up to a pp-local unit, decomposable as Ξ²(iβˆ’1)p\beta_{(i-1)}^{p} in Hβˆ—β€‹(ℂ​P∞)H_{*}(\mathbb{C}P^{\infty}).

In the equivariant case, we will proceed similarly.

  • Step 1.

    Compute the homology of K​(β„€Β―,Ξ»)K(\underline{\mathbb{Z}},\lambda) and use the homology suspension to define classes in π⋆​(β„€Β―βŠ—β„€Β―)\pi_{\star}(\underline{\mathbb{Z}}\otimes\underline{\mathbb{Z}}).

  • Step 2.

    Use information about the product structure on the homologies of K​(β„€Β―,Ξ»)K(\underline{\mathbb{Z}},\lambda) and K​(β„€Β―,2)K(\underline{\mathbb{Z}},2) to deduce relations for these classes, and hence produce the map described in Theorem A.

  • Step 3.

    Verify that the map in Theorem A is an equivalence by proving that it is an underlying equivalence and an equivalence on geometric fixed points.

The first step is carried out in Β§2 and Β§3 by identifying K​(β„€Β―,Ξ»)K(\underline{\mathbb{Z}},\lambda) with an equivariant version of ℂ​P∞\mathbb{C}P^{\infty} and then specializing a computation due to Lewis [Lew88], which we review in our context. The second step is carried out in Β§4. The third and final step is carried out in Β§6 using a lemma proven in Β§5 that allows us to check that the map on geometric fixed points is an equivalence by just verifying that the source and target have the same dimensions in each degree.

2 Homology of BCp​S1\mathrm{B}_{C_{p}}S^{1}

Recall that we have the CpC_{p}-space BCp​S1\mathrm{B}_{C_{p}}S^{1} classifying equivariant principal S1S^{1}-bundles. The following lemmas give two useful ways of thinking about this space.

Lemma 2.1.

The complex projective space ℙ​(ℂ​[z])\mathbb{P}(\mathbb{C}[z]) is a model for BCp​S1\mathrm{B}_{C_{p}}S^{1}, where the generator of CpC_{p} acts on ℂ​[z]\mathbb{C}[z] through ring maps by z↦e2​π​i/p​zz\mapsto e^{2\pi i/p}z. Here ℂ​[z]\mathbb{C}[z] is the ordinary polynomial ring over β„‚\mathbb{C}, and the projective space ℙ​(ℂ​[z])=(ℂ​[z]βˆ’{0})/(β„‚Γ—)\mathbb{P}(\mathbb{C}[z])=(\mathbb{C}[z]-\{0\})/(\mathbb{C}^{\times}) inherits an action in the evident way.

Lemma 2.2.

The space BCp​S1\mathrm{B}_{C_{p}}S^{1} is a model for K​(β„€Β―,Ξ»)K(\underline{\mathbb{Z}},\lambda).

Proof.

The map

ℙ​(ℂ​[z])β†’SPβˆžβ€‹(SΞ»)\mathbb{P}(\mathbb{C}[z])\to\mathrm{SP}^{\infty}(S^{\lambda})

to the infinite symmetric product, which sends a polynomial f​(z)f(z) to its set of roots (with multiplicity), is an equivariant homeomorphism. The group-completion of the latter is a model for K​(β„€Β―,Ξ»)K(\underline{\mathbb{Z}},\lambda) by the equivariant Dold-Thom theorem. But SPβˆžβ€‹(SΞ»)\mathrm{SP}^{\infty}(S^{\lambda}) is already group-complete: the monoid of connected components of the fixed points is β„•/p=β„€/p\mathbb{N}/p=\mathbb{Z}/p. ∎

Remark 2.3.

The reader may object that the definition of BCp​S1\mathrm{B}_{C_{p}}S^{1} makes no reference to Ξ»\lambda, so how does BCp​S1\mathrm{B}_{C_{p}}S^{1} know about this representation rather than Ξ»k\lambda^{k} for some kk coprime to pp? The answer is that, in fact, each of the Eilenberg-MacLane spaces K​(β„€Β―,Ξ»k)K(\underline{\mathbb{Z}},\lambda^{k}) coincide for such kk: we have an equivalence of β„€Β―\underline{\mathbb{Z}}-modules

Σλ​℀¯≃Σλk​℀¯\Sigma^{\lambda}\underline{\mathbb{Z}}\simeq\Sigma^{\lambda^{k}}\underline{\mathbb{Z}}

whenever (k,p)=1(k,p)=1. This follows from the computations in [FL04, Proposition 9.2], for example.

The filtration of ℂ​[z]\mathbb{C}[z] by the subspaces ℂ​[z]≀n\mathbb{C}[z]_{\leq n} of polynomials of degree at most nn gives a filtration of BCp​S1\mathrm{B}_{C_{p}}S^{1}.

Lemma 2.4.

There is a canonical equivalence

grk​BCp​S1β‰…SVk.\mathrm{gr}_{k}\mathrm{B}_{C_{p}}S^{1}\cong S^{V_{k}}.

where Vk=⨁0≀i≀kβˆ’1Ξ»iβˆ’kV_{k}=\bigoplus_{0\leq i\leq k-1}\lambda^{i-k}.

Proof.

This follows from a more general observation. If LL is a one-dimensional complex representation, and VV is an arbitrary complex representation, then the function assigning a linear map to its graph,

Homℂ​(L,V)βŸΆβ„™β€‹(VβŠ•L)βˆ’β„™β€‹(V),\mathrm{Hom}_{\mathbb{C}}(L,V)\longrightarrow\mathbb{P}(V\oplus L)-\mathbb{P}(V),

is an equivariant homeomorphism. So it induces an equivalence on one-point compactifications

SLβˆ¨βŠ—V≅ℙ​(VβŠ•L)/ℙ​(V).S^{L^{\vee}\otimes V}\cong\mathbb{P}(V\oplus L)/\mathbb{P}(V).

∎

The next proposition now follows from [Lew88, Proposition 3.1].

Proposition 2.5 (Lewis).

The above filtration on BCp​S1\mathrm{B}_{C_{p}}S^{1} splits after tensoring with β„€Β―\underline{\mathbb{Z}}, giving an equivalence

β„€Β―βŠ—BCp​S+1≃℀¯​{e0,e1,…}\underline{\mathbb{Z}}\otimes\mathrm{B}_{C_{p}}S^{1}_{+}\simeq\underline{\mathbb{Z}}\{e_{0},e_{1},...\}

where

|ek|=⨁0≀i≀kβˆ’1Ξ»iβˆ’k.|e_{k}|=\bigoplus_{0\leq i\leq k-1}\lambda^{i-k}.

In particular, for iβ‰₯1i\geq 1 we have |epi|=2​piβˆ’1​ρCp|e_{p^{i}}|=2p^{i-1}\rho_{C_{p}}.

We will also need some information about the multiplicative structure on homology.

Lemma 2.6.

Writing x≐yx\doteq y to mean that x=α​yx=\alpha y for some Ξ±βˆˆβ„€(p)Γ—\alpha\in\mathbb{Z}_{(p)}^{\times}, we have

e1p≐θ​ep,Β and ​epip≐p​epi+1​ for ​iβ‰₯1.e_{1}^{p}\doteq\theta e_{p},\text{ and }e_{p^{i}}^{p}\doteq pe_{p^{i+1}}\text{ for }i\geq 1.
Proof.

Using the model for BCp​S1\mathrm{B}_{C_{p}}S^{1} given by ℙ​(ℂ​[z])\mathbb{P}(\mathbb{C}[z]), we see that, in fact, ℙ​(ℂ​[z])\mathbb{P}(\mathbb{C}[z]) has the structure of a filtered monoid. It follows that the product in homology respects the filtration by the classes {ei}\{e_{i}\}. Thus, for iβ‰₯0i\geq 0, we have:

epip=βˆ‘j≀pi+1ci,j​eje_{p^{i}}^{p}=\sum_{j\leq p^{i+1}}c_{i,j}e_{j}

where the coefficients lie in π⋆​℀¯\pi_{\star}\underline{\mathbb{Z}}. When j<pi+1j<p^{i+1} we see that the virtual representations |ci,j||c_{i,j}| have positive virtual dimension and their fixed points also have positive virtual dimension. The homotopy of β„€Β―\underline{\mathbb{Z}} vanishes in these degrees (see, e.g., [FL04, Theorem 8.1(iv)]), so we must have

epip=ci,pi+1​epi+1e^{p}_{p^{i}}=c_{i,p^{i+1}}e_{p^{i+1}}

where |c0,p|=Ξ»βˆ’2|c_{0,p}|=\lambda-2 and |ci,pi+1|=0|c_{i,p^{i+1}}|=0 when iβ‰₯1i\geq 1. In both cases, the restriction map on π⋆​℀¯\pi_{\star}\underline{\mathbb{Z}} is injective in this degree, so the result follows from the nonequivariant calculation. ∎

3 Suspending classes

We begin with some generalities. If XX is any CpC_{p}-spectrum, we have the counit

Ξ£+βˆžβ€‹Ξ©βˆžβ€‹Xβ†’X\Sigma^{\infty}_{+}\Omega^{\infty}X\to X

which induces a map

Οƒ:β„€Β―βŠ—Ξ£+βˆžβ€‹Ξ©βˆžβ€‹Xβ†’β„€Β―βŠ—X,\sigma:\underline{\mathbb{Z}}\otimes\Sigma^{\infty}_{+}\Omega^{\infty}X\to\underline{\mathbb{Z}}\otimes X,

called the homology suspension. Just as in the classical case, Οƒ\sigma annihilates decomposable elements in Ο€βˆ—β€‹(β„€Β―βŠ—Ξ£+βˆžβ€‹Ξ©βˆžβ€‹X)\pi_{*}(\underline{\mathbb{Z}}\otimes\Sigma^{\infty}_{+}\Omega^{\infty}X).

Construction 3.1.

For iβ‰₯1i\geq 1, we define

ti:S2​piβˆ’1​ρCpβˆ’Ξ»β†’β„€Β―βŠ—β„€Β―t_{i}:S^{2p^{i-1}\rho_{C_{p}}-\lambda}\to\underline{\mathbb{Z}}\otimes\underline{\mathbb{Z}}

as the homology suspension of the element epiβˆˆΟ€2​piβˆ’1​ρCp​(β„€Β―βŠ—BCp​S1)e_{p^{i}}\in\pi_{2p^{i-1}\rho_{C_{p}}}(\underline{\mathbb{Z}}\otimes\mathrm{B}_{C_{p}}S^{1}). Here we use the identification

BCp​S1≃K​(β„€Β―,Ξ»)=Ξ©βˆžβ€‹Ξ£Ξ»β€‹β„€Β―.\mathrm{B}_{C_{p}}S^{1}\simeq K(\underline{\mathbb{Z}},\lambda)=\Omega^{\infty}\Sigma^{\lambda}\underline{\mathbb{Z}}.

4 Two relations in homology

We begin with a brief review of norms, transfers, and restrictions.

Remark 4.1 (Transfer and restriction).

Given a nonequivariant equivalence (SV)e≅Sn(S^{V})^{e}\cong S^{n}, we define

res:Ο€VXβ†’Ο€nXe,(x:SVβ†’X)↦(Snβ‰…(SV)eβ†’X)\mathrm{res}:\pi_{V}X\to\pi_{n}X^{e},\quad(x:S^{V}\to X)\mapsto(S^{n}\cong(S^{V})^{e}\to X)

and

trV:Ο€nXeβ†’Ο€VX,(y:Snβ†’Xe)↦(SVβ†’Cp+βŠ—SVβ‰…Cp+βŠ—Snβ†’Cp+βŠ—Xβ†’X).\mathrm{tr}_{V}:\pi_{n}X^{e}\to\pi_{V}X,\quad(y:S^{n}\to X^{e})\mapsto(S^{V}\to C_{p+}\otimes S^{V}\cong C_{p+}\otimes S^{n}\to C_{p+}\otimes X\to X).

For example, when V=Ξ»βˆ’2V=\lambda-2 and X=S0X=S^{0}, then trΞ»βˆ’2​(1)=ΞΈ\mathrm{tr}_{\lambda-2}(1)=\theta.

Changing the equivalence (SV)eβ‰…Sn(S^{V})^{e}\cong S^{n} has the effect of altering these classes by Β±1\pm 1; in our case the representations in question have canonical orientations so this will not be a concern. Given a map XβŠ—Yβ†’YX\otimes Y\to Y we have a relation:

tr​(xβŠ—res​(y))=tr​(x)βŠ—y.\mathrm{tr}(x\otimes\mathrm{res}(y))=\mathrm{tr}(x)\otimes y.
Remark 4.2 (Norms).

If a CpC_{p}-spectrum XX has a map N​(X)β†’XN(X)\to X, then, given an underlying class x:Snβ†’Xex:S^{n}\to X^{e}, we may define a norm by the composite

N​x:N​(Sn)=Sn​ρCpβ†’N​(X)β†’X.Nx:N(S^{n})=S^{n\rho_{C_{p}}}\to N(X)\to X.

The underlying nonequivariant class is given by res​(N​x)=∏g∈Cp(g​x)βˆˆΟ€p​n​Xe\mathrm{res}(Nx)=\prod_{g\in C_{p}}(gx)\in\pi_{pn}X^{e}.

Our goal in this section is to prove the following two lemmas.

Lemma 4.3.

The classes tiβˆˆΟ€2​piβˆ’1​ρCpβˆ’Ξ»β€‹(β„€Β―(p)βŠ—β„€Β―(p))t_{i}\in\pi_{2p^{i-1}\rho_{C_{p}}-\lambda}(\underline{\mathbb{Z}}_{(p)}\otimes\underline{\mathbb{Z}}_{(p)}) satisfy θ​ti=0\theta t_{i}=0.

Lemma 4.4.

The classes N​(ti)βˆˆΟ€(2​piβˆ’2)​ρCp​(β„€Β―(p)βŠ—β„€Β―(p))N(t_{i})\in\pi_{(2p^{i}-2)\rho_{C_{p}}}(\underline{\mathbb{Z}}_{(p)}\otimes\underline{\mathbb{Z}}_{(p)}) satisfy p​N​(ti)=0pN(t_{i})=0.

In fact, the second relation follows from the first.

Proof of Lemma 4.4 assuming Lemma 4.3.

Since p=tr​(1)p=\mathrm{tr}(1), the class p​N​(ti)pN(t_{i}) is the transfer of the class res​(ti)p\mathrm{res}(t_{i})^{p} into degree (2​piβˆ’2)​ρCp(2p^{i}-2)\rho_{C_{p}}. Notice that (2​piβˆ’2)​ρCpβˆ’|tip|=Ξ»βˆ’2(2p^{i}-2)\rho_{C_{p}}-|t_{i}^{p}|=\lambda-2 (after identifying the Ξ»k\lambda^{k} suspensions with Ξ»\lambda for (k,p)=1(k,p)=1), and the transfer of 11 into this degree is ΞΈ\theta, so we have

p​N​(ti)=θ​tip=0.pN(t_{i})=\theta t_{i}^{p}=0.

∎

Proof of Lemma 4.3.

By Lemma 2.6, we have e1p≐θ​epe_{1}^{p}\doteq\theta e_{p} so that θ​t1=σ​(θ​ep)=0\theta t_{1}=\sigma(\theta e_{p})=0, since Οƒ\sigma annihilates decomposables. For the remaining classes, consider the commutative diagram

K​(β„€Β―,Ξ»)+\textstyle{K(\underline{\mathbb{Z}},\lambda)_{+}\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}[ΞΈ]\scriptstyle{[\theta]}Σλ​℀¯\textstyle{\Sigma^{\lambda}\underline{\mathbb{Z}}\ignorespaces\ignorespaces\ignorespaces\ignorespaces}ΞΈ\scriptstyle{\theta}K​(β„€Β―,2)+\textstyle{K(\underline{\mathbb{Z}},2)_{+}\ignorespaces\ignorespaces\ignorespaces\ignorespaces}Ξ£2​℀¯\textstyle{\Sigma^{2}\underline{\mathbb{Z}}}

where [ΞΈ]=Ξ©βˆžβ€‹(ΞΈ)[\theta]=\Omega^{\infty}(\theta). Thus, to show that θ​ti=0\theta t_{i}=0 for iβ‰₯2i\geq 2, it is enough to show that [ΞΈ]βˆ—β€‹epi[\theta]_{*}e_{p^{i}} is decomposable in π⋆​(β„€Β―(p)βŠ—K​(β„€Β―,2)+)\pi_{\star}(\underline{\mathbb{Z}}_{(p)}\otimes K(\underline{\mathbb{Z}},2)_{+}) for iβ‰₯2i\geq 2.

Write

β„€Β―(p)βŠ—K​(β„€Β―,2)+=β„€Β―(p)​{Ξ³i​(Ξ²1)}\underline{\mathbb{Z}}_{(p)}\otimes K(\underline{\mathbb{Z}},2)_{+}=\underline{\mathbb{Z}}_{(p)}\{\gamma_{i}(\beta_{1})\}

where the elements Ξ³i​(Ξ²1)\gamma_{i}(\beta_{1}) are the standard module generators of Hβˆ—β€‹(ℂ​P∞;β„€)H_{*}(\mathbb{C}P^{\infty};\mathbb{Z}), and write Ξ²(i)=Ξ³pi​β1\beta_{(i)}=\gamma_{p^{i}}\beta_{1}. To show that [ΞΈ]βˆ—β€‹(epi)[\theta]_{*}(e_{p^{i}}) is decomposable for iβ‰₯2i\geq 2, it is enough to establish the following two claims:

  1. (a)

    [ΞΈ]βˆ—β€‹(epi)≐piβˆ’1​θuΞ»pi​(pβˆ’1)βˆ’1​β(i)[\theta]_{*}(e_{p^{i}})\doteq\frac{p^{i-1}\theta}{u_{\lambda}^{p^{i}(p-1)-1}}\beta_{(i)}, and

  2. (b)

    Ξ²(iβˆ’1)p≐p​β(i)\beta_{(i-1)}^{p}\doteq p\beta_{(i)}.

Claim (b) is just the classical computation of the product in homology for Hβˆ—β€‹(ℂ​P∞,β„€)H_{*}(\mathbb{C}P^{\infty},\mathbb{Z}). For claim (a), let ΞΉΞ»\iota_{\lambda} denote the fundamental class in cohomology for K​(β„€Β―,Ξ»)K(\underline{\mathbb{Z}},\lambda) and ΞΉ2\iota_{2} the same for K​(β„€Β―,2)K(\underline{\mathbb{Z}},2). Then we have [ΞΈ]βˆ—β€‹(ΞΉ2)=θ​ιλ[\theta]^{*}(\iota_{2})=\theta\iota_{\lambda} by design, and hence

[ΞΈ]βˆ—β€‹(ΞΉ2j)=ΞΈj​ιλj.[\theta]^{*}(\iota_{2}^{j})=\theta^{j}\iota_{\lambda}^{j}.

The map on homology is now determined by the relation

⟨[ΞΈ]βˆ—β€‹epi,ΞΉ2j⟩=ΞΈjβ€‹βŸ¨epi,ΞΉΞ»jβŸ©βˆˆΟ€β‹†β€‹β„€Β―(p).\langle[\theta]_{*}e_{p^{i}},\iota_{2}^{j}\rangle=\theta^{j}\langle e_{p^{i}},\iota^{j}_{\lambda}\rangle\in\pi_{\star}\underline{\mathbb{Z}}_{(p)}.

Since ΞΈj\theta^{j} is a transferred class, the value above is also a transfer, and hence determined by its restriction to an underlying class. But res​([ΞΈ])=[p]\mathrm{res}([\theta])=[p] and we clearly have [p]βˆ—β€‹(res​(epi))=pi​β(i)[p]_{*}(\mathrm{res}(e_{p^{i}}))=p^{i}\beta_{(i)}, which agrees with the restriction of piβˆ’1​θuΞ»pi​(pβˆ’1)βˆ’1​β(i)\frac{p^{i-1}\theta}{u_{\lambda}^{p^{i}(p-1)-1}}\beta_{(i)}. This completes the proof. ∎

5 Digression: Detecting equivalences nonequivariantly

The goal of this section is to establish a criterion for detecting equivalences of β„€Β―\underline{\mathbb{Z}}-modules. We recall that

℀¯Φ​Cp≃𝔽p​[b]\underline{\mathbb{Z}}^{\Phi C_{p}}\simeq\mathbb{F}_{p}[b]

where the class bb in degree 2 arises from taking the geometric fixed points of the Thom class uΞ»:SΞ»β†’Ξ£2​℀¯u_{\lambda}:S^{\lambda}\to\Sigma^{2}\underline{\mathbb{Z}}.

Proposition 5.1.

Let f:M→Nf:M\to N be a map of ℀¯\underline{\mathbb{Z}}-modules which are bounded below. Assume the following conditions are satisfied:

  1. (i)

    ff is an underlying equivalence.

  2. (ii)

    Ο€j​MΦ​Cp\pi_{j}M^{\Phi C_{p}} and Ο€j​NΦ​Cp\pi_{j}N^{\Phi C_{p}} are finite dimensional of the same rank, for all jj.

  3. (iii)

    Ο€βˆ—β€‹MΦ​Cp\pi_{*}M^{\Phi C_{p}} and Ο€βˆ—β€‹NΦ​Cp\pi_{*}N^{\Phi C_{p}} are graded-free 𝔽p​[b]\mathbb{F}_{p}[b]-modules.

Then ff is an equivalence.

We will deduce this proposition from the following one, which relates geometric and Tate fixed points.

Proposition 5.2.

Let MM be a β„€Β―\underline{\mathbb{Z}}-module which is both bounded above and below. Then the natural map

MΦ​Cp​[bβˆ’1]β†’Mt​CpM^{\Phi C_{p}}[b^{-1}]\to M^{tC_{p}}

is an equivalence.

Proof of Proposition 5.1 assuming Proposition 5.2.

By assumption (i), it is enough to check that fΦ​Cpf^{\Phi C_{p}} is an equivalence; by assumption (ii), it is enough to check that Ο€βˆ—β€‹(fΦ​Cp)\pi_{*}(f^{\Phi C_{p}}) is an injection; and by assumption (iii) it is enough to check that Ο€βˆ—β€‹(fΦ​Cp)​[bβˆ’1]\pi_{*}(f^{\Phi C_{p}})[b^{-1}] is an injection.

Again by (i), the map ft​Cpf^{tC_{p}} is an equivalence. So, from the diagram

MΦ​Cp​[bβˆ’1]\textstyle{M^{\Phi C_{p}}[b^{-1}]\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}NΦ​Cp​[bβˆ’1]\textstyle{N^{\Phi C_{p}}[b^{-1}]\ignorespaces\ignorespaces\ignorespaces\ignorespaces}Mt​Cp\textstyle{M^{tC_{p}}\ignorespaces\ignorespaces\ignorespaces\ignorespaces}≃\scriptstyle{\simeq}Nt​Cp\textstyle{N^{tC_{p}}}

we see that it is enough to check that the vertical maps are injective on homotopy. More generally, we show that whenever XX is a bounded below β„€Β―\underline{\mathbb{Z}}-module, the map

Ο€βˆ—β€‹XΦ​Cp​[bβˆ’1]β†’Ο€βˆ—β€‹Xt​Cp\pi_{*}X^{\Phi C_{p}}[b^{-1}]\to\pi_{*}X^{tC_{p}}

is injective. Indeed, by Proposition 5.2 and the fact that the Tate construction commutes with limits of Postnikov towers (see, e.g., [NS18, I.2.6]), we have

limn((τ≀n​X)Φ​Cp​[bβˆ’1])→≃limn(τ≀n​X)t​Cp≃Xt​Cp.\lim_{n}\left((\tau_{\leq n}X)^{\Phi C_{p}}[b^{-1}]\right)\stackrel{{\scriptstyle\simeq}}{{\to}}\lim_{n}(\tau_{\leq n}X)^{tC_{p}}\simeq X^{tC_{p}}.

Therefore, we need only check that

Ο€βˆ—β€‹XΦ​Cp​[bβˆ’1]β†’Ο€βˆ—β€‹limn((τ≀n​X)Φ​Cp​[bβˆ’1])\pi_{*}X^{\Phi C_{p}}[b^{-1}]\to\pi_{*}\lim_{n}\left((\tau_{\leq n}X)^{\Phi C_{p}}[b^{-1}]\right)

is injective. Since the maps XΦ​Cpβ†’(τ≀n​X)Φ​CpX^{\Phi C_{p}}\to(\tau_{\leq n}X)^{\Phi C_{p}} have increasingly connective fibers, we can replace the left hand side by (limnΟ€βˆ—β€‹(τ≀n​X)Φ​Cp)​[bβˆ’1](\lim_{n}\pi_{*}(\tau_{\leq n}X)^{\Phi C_{p}})[b^{-1}] and reduce to showing that

(limnΟ€βˆ—β€‹(τ≀n​X)Φ​Cp)​[bβˆ’1]β†’limnΟ€βˆ—β€‹((τ≀n​X)Φ​Cp​[bβˆ’1])(\lim_{n}\pi_{*}(\tau_{\leq n}X)^{\Phi C_{p}})[b^{-1}]\to\lim_{n}\pi_{*}\left((\tau_{\leq n}X)^{\Phi C_{p}}[b^{-1}]\right)

is injective. Finally, this reduces to showing that the kernel of

limnΟ€βˆ—β€‹(τ≀n​X)Φ​Cpβ†’limnΟ€βˆ—β€‹((τ≀n​X)Φ​Cp​[bβˆ’1])\lim_{n}\pi_{*}(\tau_{\leq n}X)^{\Phi C_{p}}\to\lim_{n}\pi_{*}\left((\tau_{\leq n}X)^{\Phi C_{p}}[b^{-1}]\right)

consists of elements annihilated by a power of bb. This is clear because, for each jj, the system {Ο€j​(τ≀n​X)Φ​Cp}n\{\pi_{j}(\tau_{\leq n}X)^{\Phi C_{p}}\}_{n} is eventually constant. ∎

Proof of Proposition 5.2.

Let β„°\mathcal{E} denote the full subcategory of β„€Β―\underline{\mathbb{Z}}-modules MM for which

MΦ​Cp​[bβˆ’1]β†’Mt​CpM^{\Phi C_{p}}[b^{-1}]\to M^{tC_{p}}

is an equivalence. Then β„°\mathcal{E} is stable, closed under retracts, and closed under suspending by representation spheres.

The map MΦ​Cp​[bβˆ’1]β†’Mt​CpM^{\Phi C_{p}}[b^{-1}]\to M^{tC_{p}} is one of ℀¯Φ​Cp=𝔽p​[b]\underline{\mathbb{Z}}^{\Phi C_{p}}=\mathbb{F}_{p}[b]-modules, and hence one of 𝔽p\mathbb{F}_{p}-modules, so it must be a retract of

(M/p)Φ​Cp​[bβˆ’1]=MΦ​Cp​[bβˆ’1]/pβ†’Mt​Cp/p=(M/p)t​Cp.(M/p)^{\Phi C_{p}}[b^{-1}]=M^{\Phi C_{p}}[b^{-1}]/p\to M^{tC_{p}}/p=(M/p)^{tC_{p}}.

Thus M/pβˆˆβ„°M/p\in\mathcal{E} if and only if Mβˆˆβ„°M\in\mathcal{E}. So, by replacing MM with M/pM/p and considering the Postnikov tower, we are reduced to proving the proposition in the case where Mβˆˆπ–¬π—ˆπ–½β„€Β―β™‘M\in\mathsf{Mod}^{\heartsuit}_{\underline{\mathbb{Z}}} is a Mackey functor which is a module over 𝔽¯p\underline{\mathbb{F}}_{p}.

In particular, MeM^{e} is an 𝔽p​[Cp]\mathbb{F}_{p}[C_{p}]-module. Let Ξ³\gamma denote the generator of CpC_{p} so that 𝔽p​[Cp]=𝔽p​[Ξ³]/(1βˆ’Ξ³)p\mathbb{F}_{p}[C_{p}]=\mathbb{F}_{p}[\gamma]/(1-\gamma)^{p}. Let Fj​MβŠ†MF_{j}M\subseteq M be the sub-Mackey functor generated by (1βˆ’Ξ³)j​MeβŠ†Me(1-\gamma)^{j}M^{e}\subseteq M^{e}. This is a finite filtration with associated graded pieces given by Mackey functors with trivial underlying action. So, since β„°\mathcal{E} is a thick subcategory, we are reduced to the case when MM is a discrete 𝔽¯p\underline{\mathbb{F}}_{p}-module with trivial underlying action.

For the next reduction we recall some notation. If NN is any Mackey functor, denote by NCpN_{C_{p}} the Mackey functor NβŠ—Cp+N\otimes C_{p+} and, if AA is an abelian group, denote by AΒ―tr\underline{A}_{\mathrm{tr}} the Mackey functor whose transfer map is the identity on AA and whose restriction map is multiplication by pp. We also recall that the transfer extends to a map of Mackey functors tr:NCpβ†’N\mathrm{tr}:N_{C_{p}}\to N.

Now consider the two exact sequences

0β†’im​(tr)β†’Mβ†’M/im​(t​r)β†’00\to\mathrm{im}(\mathrm{tr})\to M\to M/\mathrm{im}(tr)\to 0
0β†’ker​(tr)β†’MΒ―treβ†’im​(tr)β†’00\to\mathrm{ker}(\mathrm{tr})\to\underline{M}^{e}_{\mathrm{tr}}\to\mathrm{im}(\mathrm{tr})\to 0

If NN is any Mackey functor with Ne=0N^{e}=0, then Nβˆˆβ„°N\in\mathcal{E} since then N=NΦ​CpN=N^{\Phi C_{p}} is bounded above and hence NΦ​Cp​[bβˆ’1]=0N^{\Phi C_{p}}[b^{-1}]=0. Thus, from the exact sequences above, we are reduced to the case where MM is of the form VΒ―tr\underline{V}_{\mathrm{tr}} for an 𝔽p\mathbb{F}_{p}-vector space VV (with trivial action). Now recall that (𝔽¯p)tr=Ξ£2βˆ’Ξ»β€‹π”½Β―p(\underline{\mathbb{F}}_{p})_{\mathrm{tr}}=\Sigma^{2-\lambda}\underline{\mathbb{F}}_{p} and hence VΒ―tr=Ξ£2βˆ’Ξ»β€‹VΒ―\underline{V}_{\mathrm{tr}}=\Sigma^{2-\lambda}\underline{V}. So we are reduced to showing that the constant Mackey functor VΒ―\underline{V} lies in β„°\mathcal{E}, where VV is an 𝔽p\mathbb{F}_{p}-vector space with trivial action. This certainly holds for V=𝔽pV=\mathbb{F}_{p}, and in general we have

V¯Φ​Cp≃𝔽¯pΦ​CpβŠ—π”½pV,\underline{V}^{\Phi C_{p}}\simeq\underline{\mathbb{F}}^{\Phi C_{p}}_{p}\otimes_{\mathbb{F}_{p}}V,

since geometric fixed points commutes with colimits, and

Vt​Cp≃𝔽pt​CpβŠ—π”½pVV^{tC_{p}}\simeq\mathbb{F}_{p}^{tC_{p}}\otimes_{\mathbb{F}_{p}}V

by direct calculation. (Notice this holds even when VV is infinite-dimensional). This completes the proof. ∎

6 Proof of the main theorem

We are now ready to prove the main theorem. Recall that we have constructed classes

tiβˆˆΟ€2​piβˆ’1​ρCpβˆ’Ξ»β€‹(β„€Β―(p)βŠ—β„€Β―(p)),t_{i}\in\pi_{2p^{i-1}\rho_{C_{p}}-\lambda}(\underline{\mathbb{Z}}_{(p)}\otimes\underline{\mathbb{Z}}_{(p)}),

and shown that θ​ti=0\theta t_{i}=0 and p​N​(ti)=0pN(t_{i})=0. With notation as in the introduction, let

Xi=(S0βŠ•(S0​[N​ti]βŠ—Tθ​(ti)))X_{i}=\left(S^{0}\oplus(S^{0}[Nt_{i}]\otimes T_{\theta}(t_{i}))\right)

and

X=⨂iβ‰₯1(S0βŠ•(S0​[N​ti]βŠ—Tθ​(ti)))X=\bigotimes_{i\geq 1}\left(S^{0}\oplus(S^{0}[Nt_{i}]\otimes T_{\theta}(t_{i}))\right)

Then, choosing nullhomotopies which witness θ​ti=0\theta t_{i}=0, we get a map:

f:β„€Β―(p)βŠ—β¨‚iβ‰₯1(S0βŠ•(S0​[N​ti]βŠ—Tθ​(ti)))βŸΆβ„€Β―(p)βŠ—β„€Β―(p)f:\underline{\mathbb{Z}}_{(p)}\otimes\bigotimes_{i\geq 1}\left(S^{0}\oplus(S^{0}[Nt_{i}]\otimes T_{\theta}(t_{i}))\right)\longrightarrow\underline{\mathbb{Z}}_{(p)}\otimes\underline{\mathbb{Z}}_{(p)}

The main theorem is then the statement:

Theorem 6.1.

The map ff is an equivalence.

Proof.

Combine Proposition 5.1 with the two lemmas below. ∎

Lemma 6.2.

The map fef^{e} is an underlying equivalence.

Proof.

First observe that, by our construction in the proof of Lemma 4.4, the map N​(ti)^\widehat{N(t_{i})} restricts to the map tipβˆ’1​t^it_{i}^{p-1}\hat{t}_{i}, since the nullhomotopy witnessing p​N​(ti)=0pN(t_{i})=0 was chosen to restrict to the nullhomotopy chosen for p​tippt_{i}^{p} that came from the already chosen nullhomotopy of p​tipt_{i}. The upshot is that the map

S0βŠ•S0​[N​ti]βŠ—Tθ​(ti)β†’β„€Β―βŠ—β„€Β―S^{0}\oplus S^{0}[Nt_{i}]\otimes T_{\theta}(t_{i})\to\underline{\mathbb{Z}}\otimes\underline{\mathbb{Z}}

restricts on underlying spectra to the map

S0​[ti]/(p​ti)β†’β„€βŠ—β„€S^{0}[t_{i}]/(pt_{i})\to\mathbb{Z}\otimes\mathbb{Z}

obtained just from the relation p​ti=0pt_{i}=0 and extended via the multiplicative structure.

In particular, on mod pp homology fef^{e} induces a ring map

𝔽p​[ti]βŠ—Ξ›β€‹(xi)→𝔽p​[ΞΎi]βŠ—Ξ›β€‹(Ο„i).\mathbb{F}_{p}[t_{i}]\otimes\Lambda(x_{i})\to\mathbb{F}_{p}[\xi_{i}]\otimes\Lambda(\tau_{i}).

We know that tit_{i} maps to ΞΎi\xi_{i} and that β​xi=ti\beta x_{i}=t_{i}, so that β​(fβˆ—e​(xi))=ΞΎi\beta(f_{*}^{e}(x_{i}))=\xi_{i}. Modulo decomposables, Ο„i\tau_{i} is the only element whose Bockstein is ΞΎi\xi_{i}. So xix_{i} must map to Ο„i\tau_{i}, mod decomposables. It follows that fef^{e} is a mod pp equivalence, and hence an equivalence. ∎

Lemma 6.3.

(β„€Β―βŠ—X)Φ​Cp(\underline{\mathbb{Z}}\otimes X)^{\Phi C_{p}} and (β„€Β―βŠ—β„€Β―)Φ​Cp(\underline{\mathbb{Z}}\otimes\underline{\mathbb{Z}})^{\Phi C_{p}} are free 𝔽p​[b]\mathbb{F}_{p}[b]-modules, finite-dimensional in each degree, and isomorphic as graded vector spaces over 𝔽p\mathbb{F}_{p}.

Proof.

If YY is any CpC_{p}-spectrum, then

(β„€Β―(p)βŠ—Y)Φ​Cp=𝔽p​[b]βŠ—YΦ​Cp≃𝔽p​[b]βŠ—π”½p(𝔽pβŠ—YΦ​Cp)(\underline{\mathbb{Z}}_{(p)}\otimes Y)^{\Phi C_{p}}=\mathbb{F}_{p}[b]\otimes Y^{\Phi C_{p}}\simeq\mathbb{F}_{p}[b]\otimes_{\mathbb{F}_{p}}(\mathbb{F}_{p}\otimes Y^{\Phi C_{p}})

is a free 𝔽p​[b]\mathbb{F}_{p}[b]-module. Applying this in the cases Y=XY=X and Y=β„€Β―Y=\underline{\mathbb{Z}}, we see that each is a free 𝔽p​[b]\mathbb{F}_{p}[b], evidently finite-dimensional in each degree. So it suffices to prove that

𝔽pβŠ—XΦ​Cp≅𝔽pβŠ—(𝔽p​[b])\mathbb{F}_{p}\otimes X^{\Phi C_{p}}\cong\mathbb{F}_{p}\otimes(\mathbb{F}_{p}[b])

as graded vector spaces. Notice that we can write, as graded vector spaces,

𝔽pβŠ—XiΦ​Cp≅𝔽p​[d(iβˆ’1),ΞΎi]βŠ—π”½pΛ​(Οƒiβˆ’1,Ο„i)/(d(iβˆ’1)p,d(iβˆ’1)​τi,d(iβˆ’1)pβˆ’1​σiβˆ’1,Οƒiβˆ’1​τi),\mathbb{F}_{p}\otimes X_{i}^{\Phi C_{p}}\cong\mathbb{F}_{p}[d_{(i-1)},\xi_{i}]\otimes_{\mathbb{F}_{p}}\Lambda(\sigma_{i-1},\tau_{i})/(d_{(i-1)}^{p},d_{(i-1)}\tau_{i},d_{(i-1)}^{p-1}\sigma_{i-1},\sigma_{i-1}\tau_{i}),

where |Οƒiβˆ’1|=2​piβˆ’1βˆ’1|\sigma_{i-1}|=2p^{i-1}-1 and |d(iβˆ’1)|=2​piβˆ’1|d_{(i-1)}|=2p^{i-1}. Indeed, t^i\hat{t}_{i}, on geometric fixed points, gives rise to two classes; one we are calling d(iβˆ’1)d_{(i-1)} and the other we are calling Οƒiβˆ’1\sigma_{i-1}. Similarly, N​(ti)^\widehat{N(t_{i})}, on geometric fixed points, gives rise to two classes: one we are calling ΞΎi\xi_{i} and the other Ο„i\tau_{i}, in their usual degrees. The relations are the ones needed to ensure that the monomials not arising from geometric fixed points of elements in XiX_{i} are omitted.

It follows that we have an isomorphism of graded vector spaces

𝔽pβŠ—XΦ​Cp≅𝔽p[ΞΎn:nβ‰₯1]βŠ—π”½p𝔽p[d(i):iβ‰₯0]βŠ—π”½pΞ›(Οƒj,Ο„k:jβ‰₯0,kβ‰₯1)/(d(i)p,d(iβˆ’1)Ο„i,d(i)pβˆ’1Οƒi,Οƒiβˆ’1Ο„i).\mathbb{F}_{p}\otimes X^{\Phi C_{p}}\cong\mathbb{F}_{p}[\xi_{n}:n\geq 1]\otimes_{\mathbb{F}_{p}}\mathbb{F}_{p}[d_{(i)}:i\geq 0]\otimes_{\mathbb{F}_{p}}\Lambda(\sigma_{j},\tau_{k}:j\geq 0,k\geq 1)/(d_{(i)}^{p},d_{(i-1)}\tau_{i},d_{(i)}^{p-1}\sigma_{i},\sigma_{i-1}\tau_{i}).

We are trying to show that this is isomorphic, as a graded vector space to

𝔽pβŠ—π”½p[b]≅𝔽p[ΞΎn:nβ‰₯1]βŠ—π”½pΞ›(Ο„i:iβ‰₯0)βŠ—π”½p𝔽p[b].\mathbb{F}_{p}\otimes\mathbb{F}_{p}[b]\cong\mathbb{F}_{p}[\xi_{n}:n\geq 1]\otimes_{\mathbb{F}_{p}}\Lambda(\tau_{i}:i\geq 0)\otimes_{\mathbb{F}_{p}}\mathbb{F}_{p}[b].

We may regard each vector space as a module over 𝔽p[ΞΎn:nβ‰₯0]\mathbb{F}_{p}[\xi_{n}:n\geq 0] in the evident way, and hence reduce to showing that the two vector spaces

V=Ξ›(Ο„i:iβ‰₯0)βŠ—π”½p𝔽p[b]V=\Lambda(\tau_{i}:i\geq 0)\otimes_{\mathbb{F}_{p}}\mathbb{F}_{p}[b]

and

W=𝔽p[d(i):iβ‰₯0]βŠ—π”½pΞ›(Οƒj,Ο„k:jβ‰₯0,kβ‰₯1)/(d(i)p,d(iβˆ’1)Ο„i,d(i)pβˆ’1Οƒi,Οƒiβˆ’1Ο„i)W=\mathbb{F}_{p}[d_{(i)}:i\geq 0]\otimes_{\mathbb{F}_{p}}\Lambda(\sigma_{j},\tau_{k}:j\geq 0,k\geq 1)/(d_{(i)}^{p},d_{(i-1)}\tau_{i},d_{(i)}^{p-1}\sigma_{i},\sigma_{i-1}\tau_{i})

are isomorphic. (Here recall that |Οƒi|=|Ο„i|=2​piβˆ’1|\sigma_{i}|=|\tau_{i}|=2p^{i}-1, |b|=2|b|=2, and |d(i)|=2​pi|d_{(i)}|=2p^{i}).

Let II range over sequences (a0,a1,…)(a_{0},a_{1},...) with 0≀ai≀pβˆ’20\leq a_{i}\leq p-2, JJ range over sequences (Ξ΅0,Ξ΅1,…)(\varepsilon_{0},\varepsilon_{1},...) with Ξ΅i∈{0,1}\varepsilon_{i}\in\{0,1\}, KK range over sequences (ΞΊ0,ΞΊ1,…)(\kappa_{0},\kappa_{1},...) with ΞΊi∈{0,1}\kappa_{i}\in\{0,1\}, and let Kβ€²K^{\prime} range over sequences (ΞΊ0β€²,ΞΊ1β€²,…)(\kappa^{\prime}_{0},\kappa^{\prime}_{1},...) with ΞΊiβ€²βˆˆ{0,1}\kappa^{\prime}_{i}\in\{0,1\}. We impose the following requirements on these sequences:

  • β€’

    Each sequence has finite support.

  • β€’

    If ΞΊiβ€²=1\kappa^{\prime}_{i}=1, then ΞΊi=1\kappa_{i}=1. (So Kβ€²K^{\prime} is otained from KK by changing some subset of 11s to 0s).

  • β€’

    Jβ‹…K=Iβ‹…K=(0,0,…)J\cdot K=I\cdot K=(0,0,...). That is: II and KK have disjoint support and JJ and KK have disjoint support.

Then VV has a basis of monomials

MI,J,K=(∏iβ‰₯0bai​pi)​τJ​(∏iβ‰₯0bΞΊi​(pβˆ’1)​pi)​τKβ€²M_{I,J,K}=(\prod_{i\geq 0}b^{a_{i}p^{i}})\tau_{J}(\prod_{i\geq 0}b^{\kappa_{i}(p-1)p^{i}})\tau_{K^{\prime}}

and WW has a basis of monomials

NI,J,K=dI​σJ​(∏iβ‰₯0d(i)(ΞΊiβˆ’ΞΊiβ€²)​(pβˆ’1))​τK′​[1]N_{I,J,K}=d_{I}\sigma_{J}(\prod_{i\geq 0}d_{(i)}^{(\kappa_{i}-\kappa^{\prime}_{i})(p-1)})\tau_{K^{\prime}[1]}

where K′​[1]=(0,ΞΊ0β€²,ΞΊ1β€²,…)K^{\prime}[1]=(0,\kappa^{\prime}_{0},\kappa^{\prime}_{1},...). These have the same number of basis elements in each dimension, so Vβ‰…WV\cong W. ∎

References

  • [Car99] JeffreyΒ L. Caruso, Operations in equivariant 𝐙/p{\bf Z}/p-cohomology, Math. Proc. Cambridge Philos. Soc. 126 (1999), no.Β 3, 521–541. MR 1684248
  • [FL04] KevinΒ K. Ferland and L.Β Gaunce Lewis, Jr., The R​O​(G)R{\rm O}(G)-graded equivariant ordinary homology of GG-cell complexes with even-dimensional cells for G=β„€/pG={\mathbb{Z}}/p, Mem. Amer. Math. Soc. 167 (2004), no.Β 794, viii+129. MR 2025457
  • [Gre88] J.Β P.Β C. Greenlees, Stable maps into free GG-spaces, Trans. Amer. Math. Soc. 310 (1988), no.Β 1, 199–215. MR 938918
  • [HHR16] M.Β A. Hill, M.Β J. Hopkins, and D.Β C. Ravenel, On the nonexistence of elements of Kervaire invariant one, Ann. of Math. (2) 184 (2016), no.Β 1, 1–262. MR 3505179
  • [HK01] PoΒ Hu and Igor Kriz, Real-oriented homotopy theory and an analogue of the Adams-Novikov spectral sequence, Topology 40 (2001), no.Β 2, 317–399. MR 1808224
  • [HW20] Jeremy Hahn and Dylan Wilson, Eilenberg–Mac Lane spectra as equivariant Thom spectra, Geom. Topol. 24 (2020), no.Β 6, 2709–2748. MR 4194302
  • [Lew88] L.Β Gaunce Lewis, Jr., The R​O​(G)R{\rm O}(G)-graded equivariant ordinary cohomology of complex projective spaces with linear 𝐙/p{\bf Z}/p actions, Algebraic topology and transformation groups (GΓΆttingen, 1987), Lecture Notes in Math., vol. 1361, Springer, Berlin, 1988, pp.Β 53–122. MR 979507
  • [NS18] Thomas Nikolaus and Peter Scholze, On topological cyclic homology, Acta Math. 221 (2018), no.Β 2, 203–409. MR 3904731
  • [Oru89] MeldaΒ Yaman OruΓ§, The equivariant Steenrod algebra, Topology Appl. 32 (1989), no.Β 1, 77–108. MR 1003301
  • [San19] KrishanuΒ Roy Sankar, Steinberg summands in the free 𝔽p\mathbb{F}_{p}-module on the equivariant sphere spectrum, 2019.
  • [Voe03] Vladimir Voevodsky, Reduced power operations in motivic cohomology, Publ. Math. Inst. Hautes Γ‰tudes Sci. (2003), no.Β 98, 1–57. MR 2031198