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

Invariant prime ideals in equivariant Lazard rings

Markus Hausmann and Lennart Meier
Abstract

Let AA be an abelian compact Lie group. In this paper we compute the spectrum of invariant prime ideals of the AA-equivariant Lazard ring, or equivalently the spectrum of points of the moduli stack of AA-equivariant formal groups. We further show that this spectrum is homeomorphic to the Balmer spectrum of compact AA-spectra, with the comparison map induced by equivariant complex bordism homology.

1 Introduction

Let us pose the question: what algebraic input do we need to develop equivariant versions of chromatic homotopy theory?

Chromatic homotopy theory studies stable homotopy theory through the lens of formal groups, building on Quillen’s identification of the complex bordism ring πMU\pi_{*}MU with the Lazard ring [QuillenElementary]. Around the same time, tom Dieck introduced for every compact Lie group AA in [tomDieckBordism] an equivariant analog of MUMU, the homotopical AA-equivariant complex bordism MUAMU_{A}. Letting AA be abelian, Cole, Greenlees and Kriz [CGK] found many years later the correct notion of an AA-equivariant formal group law. Recently, the first author generalized work of Hanke–Wiemeler [HankeWiemeler] and showed that πAMUA\pi_{*}^{A}MU_{A} is indeed the universal ring for AA-equivariant formal group laws, thus establishing an equivariant analog of Quillen’s theorem for the equivariant Lazard ring LAL_{A}.

Many structural features of stable homotopy theory can be explained through the chromatic perspective. The central notion of chromatic homotopy theory is that of height. Honda classified formal groups over a field of characteristic pp in terms of the height 0n0\leq n\leq\infty. Thus, the points of the moduli stack of formal groups FG\mathcal{M}_{FG} correspond to pairs (p,n)(p,n) with n=0n=0 if and only if p=0p=0. Hopkins and Smith [H-S98] showed that the same classification pertains to thick subcategories of finite spectra: Given a finite spectrum XX, its MUMU-homology MUXMU_{*}X defines a coherent sheaf over FG\mathcal{M}_{FG}. Taking the support of MUXMU_{*}X in the Zariski spectrum |FG||\mathcal{M}_{FG}| of points, we obtain a support theory on finite spectra. The thick subcategory theorem states that this support theory is the universal one. In other words, the induced map |FG|Spec(Spc)|\mathcal{M}_{FG}|\to\operatorname{Spec}(Sp^{c}) to the Balmer spectrum of finite spectra (cf. [BalmerSpectrum, BalmerSpectraSpectraSpectra]) is a homeomorphism.

We show the following equivariant analog (a more precise statement of which we give as 1.6):

Theorem 1.1.

Let AA be an abelian compact Lie group. Then the spectrum of points of the moduli space MAFGM^{A}_{FG} of AA-equivariant formal groups is homeomorphic to the Balmer spectrum of finite AA-spectra, with the comparison map induced by a support theory based on complex bordism homology (MUA)(MU_{A})_{*}.

This establishes MUAMU_{A} and the theory of equivariant formal groups as fundamental tools for building equivariant versions of chromatic homotopy theory.

For abelian groups as above, the Balmer spectrum of finite AA-spectra has been computed completely in the papers [BalmerSanders] (the case A=CpA=C_{p}), [BarthelHausmannNaumannNikolausNoelStapleton] (the finite abelian case) and [BarthelGreenleesHausmann] (the general abelian case). In a surprising turn of history, it had been the algebraic counterpart which had not been computed before. As a set, both |AFG||\mathcal{M}^{A}_{FG}| and Spec(SpAc)\operatorname{Spec}(\operatorname{Sp}_{A}^{c}) decompose as the disjoint union of one copy of |FG|Spec(Spc)|\mathcal{M}_{FG}|\cong\operatorname{Spec}(\operatorname{Sp}^{c}) for every closed subgroup of AA. Thus, the correct notion of height of an AA-equivariant formal group FF over a field of characteristic pp consists of a pair: a height nn of a non-equivariant formal group and a closed subgroup BAB\subset A such that FF is induced along the zig-zag AA/B{1}A\to A/B\leftarrow\{1\}.

The more subtle information lies in the topology of the spectrum, which encodes on the algebraic level how heights can deform and on the homotopical level the chromatic interdependence of the various geometric fixed points ΦBX\Phi^{B}X of a finite AA-spectrum XX.

We will detail our results below in the language of invariant prime ideals. Crucially, we exhibit equivariant lifts 𝐯n\mathbf{v}_{n} of the classical vnv_{n} and show that they provide in many cases a sequence of generators of invariant prime ideals. The non-equivariant vnv_{n} play an important role in many of the highlights of chromatic homotopy theory, like the greek-letter construction [Rav86], the construction of the Morava K-theories or the periodicity theorem [H-S98], and we hope that our equivariant lifts open the prospect to generalize these to the equivariant context.

1.1 Invariant prime ideals and statement of results

As indicated above, the main theorem can also be stated in terms of invariant prime ideals of the equivariant Lazard ring LAL_{A}, as we now explain. Similarly to the non-equivariant case, LAL_{A} is the ground ring of a flat Hopf algebroid (LA,SA)(L_{A},S_{A}), classifying AA-equivariant formal group laws and their strict isomorphisms. The associated stack is the moduli stack of AA-equivariant formal groups. Hence, the category of graded (LA,SA)(L_{A},S_{A})-comodules is equivalent to the category of quasi-coherent sheaves over MAFGM^{A}_{FG}.

Recall that an ideal II of LAL_{A} is called invariant (in the sense of Hopf algebroids) if it is a sub-comodule, i.e., if ηL(I)SA=ηR(I)SA\eta_{L}(I)S_{A}=\eta_{R}(I)S_{A} for the left and right unit ηL,ηR:LASA\eta_{L},\eta_{R}\colon L_{A}\to S_{A}. Every invariant prime ideal 𝔭\mathfrak{p} gives rise to a point of the moduli stack of prime ideals via the quotient field of LA/𝔭L_{A}/\mathfrak{p}. This defines a map from the set of invariant prime ideals Specinv(LA)\operatorname{Spec}^{\operatorname{inv}}(L_{A}) to |MAFG||M^{A}_{FG}|, which we show to be a homeomorphism in Theorem 4.7.

For the non-equivariant Lazard ring, Morava and Landweber [LandweberIdeals] showed that the invariant prime ideals are precisely the ideals Ip,n=(v0,,vn1)I_{p,n}=(v_{0},\ldots,v_{n-1}) for a prime pp and n{}n\in\mathbb{N}\cup\{\infty\} (with Ip,0I_{p,0} being the 0-ideal for all pp).

To describe the invariant prime ideals in the equivariant case we recall that LAL_{A} contains universal Euler classes eVe_{V} for all characters VAV\in A^{*}, and that equivariant Lazard rings are contravariantly functorial in continuous group homomorphisms. In particular, all equivariant Lazard rings are algebras over the non-equivariant Lazard ring.

Then, given a non-equivariant invariant prime ideal Ip,nI_{p,n} and a closed subgroup BB of AA we obtain an invariant prime ideal IAB,p,nLAI^{A}_{B,p,n}\subseteq L_{A} as the kernel of the composite

LAresABLBΦBLBΦBLBLL/Ip,n.L_{A}\xrightarrow{\operatorname{res}^{A}_{B}}L_{B}\to\Phi^{B}L_{B}\to\Phi^{B}L_{B}\otimes_{L}L/I_{p,n}.

Here, ΦBL\Phi^{B}L is defined as the localization of LAL_{A} away from all the Euler classes of non-trivial characters for BB. The ring ΦBLB\Phi^{B}L_{B} is an algebraic version of geometric fixed points and indeed agrees with the coefficient ring of the BB-geometric fixed points of MUBMU_{B}.

Theorem 1.2 (Theorem 4.7).

For every abelian compact Lie group AA the map

Sub(A)×Specinv(L)\displaystyle\operatorname{Sub}(A)\times\operatorname{Spec}^{\operatorname{inv}}(L)\to Specinv(LA)\displaystyle\operatorname{Spec}^{\operatorname{inv}}(L_{A})
(B,Ip,n)\displaystyle(B,I_{p,n})\mapsto IAB,p,n\displaystyle I^{A}_{B,p,n}

is a bijection.

Here, Sub(A)\operatorname{Sub}(A) is the set of all closed subgroups of AA. Hence, as for the Balmer spectrum, the invariant prime ideals of LAL_{A} decompose as a set as one copy of Specinv(L)\operatorname{Spec}^{\operatorname{inv}}(L) for every closed subgroup of AA. And similarly to the Balmer spectrum, the main work then lies in understanding the Zariski topology, in particular in determining the containments between invariant prime ideals associated to different subgroups.

We obtain the following:

Theorem 1.3 (Theorem 5.1).

There is an inclusion IAB,q,nIAB,p,nI^{A}_{B^{\prime},q,n^{\prime}}\subseteq I^{A}_{B,p,n} if and only if

  1. 1.

    BB is a subgroup of BB^{\prime}.

  2. 2.

    p=qp=q or n=0n=0 (in which case IAB,q,0=IAB,p,0I^{A}_{B^{\prime},q,0}=I^{A}_{B^{\prime},p,0}), the components π0(B/B)\pi_{0}(B^{\prime}/B) are a pp-group and nnrankp(π0(B/B))n\geq n^{\prime}-\operatorname{rank}_{p}(\pi_{0}(B^{\prime}/B)).

Comparing with [BarthelGreenleesHausmann] we see that these correspond precisely to the inclusions in the Balmer spectrum, but with roles reversed: There is an inclusion IAB,q,nIAB,p,nI^{A}_{B^{\prime},q,n^{\prime}}\subseteq I^{A}_{B,p,n} if and only if there is an inclusion PAB,p,nPAB,q,nP^{A}_{B,p,n}\subseteq P^{A}_{B^{\prime},q,n^{\prime}}. Here, P(B,p,n)={XSpAc|K(n)(ΦBX)=0}P(B,p,n)=\{X\in\operatorname{Sp}_{A}^{c}\ |\ K(n)_{*}(\Phi^{B}X)=0\} are the thick subcategories of SpAc\operatorname{Sp}_{A}^{c} with K(n)K(n) being Morava K-theory at the prime pp.

To show that IAB,q,nI^{A}_{B^{\prime},q,n^{\prime}} indeed includes into IAB,p,nI^{A}_{B,p,n} when conditions 11 and 22 are satisfied, one can reduce to the case A=𝐓A=\mathbf{T} the circle group where it is straightforward to describe explicit generators for the invariant prime ideals. The main step in ruling out further inclusions is the construction of equivariant refinements 𝐯¯n1LCpn\overline{\mathbf{v}}_{n-1}\in L_{C_{p}^{n}} of the elements vn1Lv_{n-1}\in L which exhibit maximal height shifts (Definition 5.21, 5.24). Roughly speaking, 𝐯¯n1\overline{\mathbf{v}}_{n-1} is of height 0 at the top group CpnC_{p}^{n} (i.e., it lies in the ideal ICpn,p,0I_{C_{p}^{n},p,0}) while it is of height nn at the trivial group (i.e., it lies in the ideal I{1},p,nI_{\{1\},p,n} but not in I{1},p,n1I_{\{1\},p,n-1}). This is the algebraic analog of the existence of finite CpnC_{p}^{n}-spectra of underlying type nn whose CpnC_{p}^{n}-geometric fixed points are rationally non-trivial as in [BarthelHausmannNaumannNikolausNoelStapleton] and [KuhnLloydChromatic]. More precisely, 𝐯¯n\overline{\mathbf{v}}_{n} is canonically defined only modulo a certain smaller ideal (analogously to vnv_{n} only being defined uniquely up to the ideal InI_{n}). More details are given in Section 5.3.

We further show that - at least over elementary abelian pp-groups - the elements 𝐯¯i\overline{\mathbf{v}}_{i} give rise to generators of the invariant prime ideals:

Theorem 1.4 (6.1).

For all primes pp and nn\in\mathbb{N} the elements

p1𝐯¯0,p2𝐯¯1,,pn1𝐯¯n2,𝐯¯n1p_{1}^{*}\overline{\mathbf{v}}_{0},p_{2}^{*}\overline{\mathbf{v}}_{1},\ldots,p_{n-1}^{*}\overline{\mathbf{v}}_{n-2},\overline{\mathbf{v}}_{n-1}

generate the ideal ICpnCpn,p,0I^{C_{p}^{n}}_{C_{p}^{n},p,0}. Here, pi:CpnCpip_{i}\colon C_{p}^{n}\to C_{p}^{i} denotes the projection to the first ii coordinates.

Suitable restrictions of the 𝐯¯n\overline{\mathbf{v}}_{n} then form generators for the ideals ICpnCpn,p,mI^{C_{p}^{n}}_{C_{p}^{n},p,m} at higher height mm, see Section 6. We emphasize that in contrast to the non-equivariant situation, the sequence of the pi𝐯¯i1p_{i}^{*}\overline{\mathbf{v}}_{i-1} is not a regular sequence. In fact, since ICpnCpn,p,0I^{C_{p}^{n}}_{C_{p}^{n},p,0} consists precisely of the Euler-class-power torsion, it does not contain a non-zero divisor and hence cannot be generated by a regular sequence (unless n=0n=0). The torsion in the ring LCpnL_{C_{p}^{n}} is closely linked to the torsion in the group of characters (Cpn)(C_{p}^{n})^{*}. Hence one might hope that IAB,nI^{A}_{B,n} is generated by a regular sequence whenever AA is a torus, and indeed that is the case in all the cases we understand (cf., Remark 6.7).

Finally, in order to describe the Zariski topology we need one additional ingredient. When AA is infinite, the set of closed subgroups Sub(A)\operatorname{Sub}(A) contains a non-trivial metric topology, turning it into a totally-disconnected compact Hausdorff space. Together with the inclusions between the invariant prime ideals, this topology determines the Zariski topology on Specinv(LA)\operatorname{Spec}^{\operatorname{inv}}(L_{A}).

Theorem 1.5 (7.5).

The Zariski topology on Specinv(LA)\operatorname{Spec}^{\operatorname{inv}}(L_{A}) has as basis the closed subsets CC which are

  1. (i)

    closed under upward inclusions, i.e., if IAB,q,nCI^{A}_{B^{\prime},q,n^{\prime}}\in C and IAB,q,nIAB,p,nI^{A}_{B^{\prime},q,n^{\prime}}\subseteq I^{A}_{B,p,n}, then IAB,p,nCI^{A}_{B,p,n}\in C, and which

  2. (ii)

    are locally constant on Sub(A)\operatorname{Sub}(A) in the sense that if IAB,p,nCI^{A}_{B,p,n}\in C there exists a neighborhood UU of BB such that IAB,p,nCI^{A}_{B^{\prime},p,n}\in C for all BUB^{\prime}\in U.

Comparing with [BarthelGreenleesHausmann], we see that this description precisely matches the computation of the topology on the Balmer spectrum, with IAB,p,nI^{A}_{B,p,n} replaced by PAB,p,nP^{A}_{B,p,n}. Hence the assignment

IAB,p,nPAB,p,nI^{A}_{B,p,n}\mapsto P^{A}_{B,p,n}

yields a homeomorphism from Specinv(LA)\operatorname{Spec}^{\operatorname{inv}}(L_{A}) (and hence |FGA||\mathcal{M}_{FG}^{A}|) to Spec(SpAc)\operatorname{Spec}(\operatorname{Sp}_{A}^{c}). In the last section we explain that this comparison map can be obtained less ad hoc via MUAMU_{A}-homology:

Theorem 1.6 (Section 8.1).

Let XX be a finite AA-spectrum and IAB,p,nI^{A}_{B,p,n} an invariant prime ideal. Then the localization (MUA)IAB,p,nX(MU_{A})_{I^{A}_{B,p,n}}\wedge X is non-trivial if and only if the BB-geometric fixed points ΦBX\Phi^{B}X are of type n\leq n at pp, i.e, if and only if PAB,p,nP^{A}_{B,p,n} is in the Balmer support of XX.

This shows that

Xsupp((MU¯A)X)Specinv(LA)|MAFG|X\mapsto\operatorname{supp}((\underline{MU}_{A})_{*}X)\subseteq\operatorname{Spec}^{\operatorname{inv}}(L_{A})\cong|M^{A}_{FG}|

defines a universal support theory on finite AA-spectra and thus a homeomorphism Specinv(LA)Spec(SpcA)\operatorname{Spec}^{\operatorname{inv}}(L_{A})\to\operatorname{Spec}(\operatorname{Sp}^{c}_{A}). Here, (MU¯A)X(\underline{MU}_{A})_{*}X is the Mackey functor recording (MUB)resABX(MU_{B})_{*}\operatorname{res}^{A}_{B}X for all closed subgroups BB of AA, and supp((MU¯A)X)\operatorname{supp}((\underline{MU}_{A})_{*}X) is defined as the set of invariant prime ideals at which the localization of this Mackey functor is non-trivial.

Our proof that Xsupp((MU¯A)X)X\mapsto\operatorname{supp}((\underline{MU}_{A})_{*}X) is a support theory is independent of [BarthelHausmannNaumannNikolausNoelStapleton] and [BarthelGreenleesHausmann]; this already provides a continuous bijection Specinv(LA)Spec(SpcA)\operatorname{Spec}^{\operatorname{inv}}(L_{A})\to\operatorname{Spec}(\operatorname{Sp}^{c}_{A}), providing a new proof of one half of their main theorems. This half has been dubbed the chromatic Smith chromatic fixed point theorem in [KuhnShort] and [BalderramaKuhnChromatic], where other proofs are given. To establish that our support theory is universal, we need to invoke [BarthelHausmannNaumannNikolausNoelStapleton] and [BarthelGreenleesHausmann] however.

We show in LABEL:prop:universalsupportMFG that the support theory from 1.6 can alternatively be built by viewing (MU¯A)X(\underline{MU}_{A})_{*}X as defining quasi-coherent sheaves on FGB\mathcal{M}_{FG}^{B} for every closed BAB\subseteq A; the union of the supports of these sheaves agrees with supp((MU¯A)X)\operatorname{supp}((\underline{MU}_{A})_{*}X) under the homeomorphism Specinv(LA)|MAFG|\operatorname{Spec}^{\operatorname{inv}}(L_{A})\cong|M^{A}_{FG}|. To show this, we establish in 8.9 how the adjunction between SpA\operatorname{Sp}^{A} and SpA/B\operatorname{Sp}^{A/B} defined by geometric fixed points and pullback corresponds on the algebraic level to pullback and pushforward along the open immersion FGA/BFGA\mathcal{M}_{FG}^{A/B}\subseteq\mathcal{M}_{FG}^{A} from 3.11.

1.2 Acknowledgements

We thank Robert Burklund, Jeremy Hahn and Allen Yuan for their interest and enlightening discussions about the elements 𝐯¯n\overline{\mathbf{v}}_{n} and transfers. We thank the Institut Mittag-Leffler for its hospitality during the research program Higher algebraic structures in algebra, topology and geometry and the Hausdorff Research Institute for Mathematics for its hospitality during the trimester program Spectral Methods in Algebra, Geometry, and Topology – both have provided a welcoming and stimulating work environment. Finally, the first author was supported by the Knut and Alice Wallenberg Foundation, and the second author was supported by the NWO grant VI.Vidi.193.111.

2 Equivariant formal groups

The aim of this section is to recall some basic definitions and properties about equivariant formal groups and equivariant formal group laws from [CGK, GreenleesFGL, StricklandMulti, Hau]. To make our paper more self-contained, we replicate also some of the proofs in our language, and we provide some small extensions of known results. Our treatment of equivariant formal groups is far from exhaustive and especially [StricklandMulti] contains a wealth of results we do not touch upon.

2.1 Basic definitions

In this subsection, we will recall the notions of an equivariant formal group and an equivariant formal group law over a commutative ring kk. The definition of an equivariant formal group law is due to Cole, Greenlees and Kriz [CGK] and our definition of an equivariant formal group will be a variant of that of Strickland [StricklandMulti].

For us, a formal kk-algebra is a complete linearly topologized commutative kk-algebra with a countable system of open ideals generating the topology. By [Yasuda, Section 5.2], this is equivalent to the full subcategory of pro-objects in commutative kk-algebras, indexed on a countable directed set with surjective transition maps. For us, the category of formal kk-schemes is the opposite of that of formal kk-algebras. It can be viewed as the category of ind-objects in affine kk-schemes, indexed by a countable directed set with closed immersions as transition maps. We will sometimes use the notation SpecR\operatorname{Spec}R or SpfR\mathrm{Spf}R for the formal kk-scheme associated to a formal kk-algebra RR, and 𝒪X\mathcal{O}_{X} for the formal kk-algebra associated to a formal kk-scheme XX. The product on affine kk-schemes induces one on formal kk-schemes, and this corresponds to the completed tensor product on formal kk-algebras.

We set S=SpeckS=\operatorname{Spec}k. Given a countable set MM, we view S×MS\times M as a formal scheme, namely as the colimit over all S×NS\times N with NMN\subseteq M finite. This corresponds to giving kMk^{M} the product topology. If we just write MM, we will apply this construction to Spec\operatorname{Spec}\mathbb{Z} instead of SS.

For a compact Lie group AA, we will denote by A=Hom(A,𝐓)A^{*}=\operatorname{Hom}(A,\mathbf{T}) its Pontryagin dual, which is always a discrete group. We use ϵ\epsilon for the unit element in AA^{*}.

Definition 2.1.

Given a compact abelian Lie group AA, an AA-equivariant formal group over kk consists of a commutative group object XX in formal kk-schemes together with a group homomorphism φ:S×AX\varphi\colon S\times A^{*}\to X of formal kk-schemes satisfying the following two conditions:

  1. 1.

    For the composite φϵ:SidS×ϵS×AφX\varphi_{\epsilon}\colon S\xrightarrow{\operatorname{id}_{S}\times\epsilon}S\times A^{*}\xrightarrow{\varphi}X, the augmentation ideal Iϵ=ker(Rk)I_{\epsilon}=\ker(R\to k) of the induced map is fpqc-locally on kk a free RR-module of rank 11, where R=𝒪XR=\mathcal{O}_{X}.

  2. 2.

    The topology on RR is generated by products of the ideals IV=ker(RφVk)I_{V}=\ker(R\xrightarrow{\varphi_{V}^{*}}k) for VAV\in A^{*} and φV:SidS×VS×AφX\varphi_{V}\colon S\xrightarrow{\operatorname{id}_{S}\times V}S\times A^{*}\xrightarrow{\varphi}X.

Remark 2.2.

In the above definition, one can easily replace Speck\operatorname{Spec}k by an arbitrary quasi-compact scheme SS, with formal SS-schemes being a suitable subcategory of the ind-category of AffS\operatorname{Aff}_{S}, the category of schemes affine over SS.

Our definition differs in two aspects from that put forward in [StricklandMulti, Definition 2.15]. First, Strickland restricts to finite AA^{*}. Second, Strickland asks ker(𝒪Xk)\ker(\mathcal{O}_{X}\to k) to be free of rank 11 instead of locally free (cf. [StricklandMulti, Proposition 2.10]). We changed it so that our definition satisfies descent. Note that we could have asked equivalently that the augmentation ideal is Zariski locally on kk a free RR-module of rank 11 because line bundles satisfy fpqc-descent.

Remark 2.3.

If we leave out the second condition in Definition 2.1, we get a notion we call an AA-equivariant group. The category of AA-equivariant formal groups embeds into that of AA-equivariant groups and this inclusion has a right adjoint, called completion. Concretely, this replaces RR in the notation in Definition 2.1 by the formal kk-algebra limV1,,VnAR/(IV1IVn)\lim_{V_{1},\dots,V_{n}\in A^{*}}R/(I_{V_{1}}\cdots I_{V_{n}}). We will only use AA-equivariant groups to complete them to AA-equivariant formal groups.

Spelling out what we get in a more algebraic language if we fix a trivialization of the augmentation ideal IϵI_{\epsilon} gives us the notion of an equivariant formal group law.

Definition 2.4.

An AA-equivariant formal group law over kk is a quadruple

(R,Δ,θ,y(ε))(R,\Delta,\theta,y(\varepsilon))

of a formal kk-algebra RR, a continuous comultiplication Δ:RR^R\Delta\colon R\to R\hat{\otimes}R, a map of kk-algebras θ:RkA\theta\colon R\to k^{A^{*}} and an orientation y(ϵ)Ry(\epsilon)\in R, such that

  1. (i)

    the comultiplication is a map of kk-algebras which is cocommutative, coassociative and counital for the augmentation θϵ:Rk\theta_{\epsilon}\colon R\to k,

  2. (ii)

    the map θ\theta is compatible with the coproduct, and the topology on RR is generated by finite products of the kernels of the component functions θV:Rk\theta_{V}\colon R\to k for VAV\in A^{*}, and

  3. (iii)

    the element y(ϵ)y(\epsilon) is regular and generates the kernel of θϵ\theta_{\epsilon}.

We refer to [CGK] and [Hau] for more information about equivariant formal group laws.

Remark 2.5.

If we want to remember the base of an equivariant formal group law, we sometimes also write it as a quintuple (k,R,Δ,θ,y(ε))(k,R,\Delta,\theta,y(\varepsilon)).

Lemma 2.6.

An AA-equivariant formal group (SpecR,φ)(\operatorname{Spec}R,\varphi) over kk together with an RR-linear isomorphism IϵRI_{\epsilon}\cong R of the augmentation ideal is equivalent datum to an AA-equivariant formal group law over kk.

Proof.

The maps Δ\Delta and θ\theta are induced by the multiplication on SpecR\operatorname{Spec}R and φ\varphi, respectively. The element y(ε)y(\varepsilon) corresponds to the trivialization of Iϵ=ker(Rk)I_{\epsilon}=\ker(R\to k). ∎

Given any equivariant formal group G=(φ:AX)G=(\varphi\colon A^{*}\to X) over S=SpeckS=\operatorname{Spec}k, we obtain for every VAV\in A^{*} a morphism φV:SidS×VS×AφX\varphi_{V}\colon S\xrightarrow{\operatorname{id}_{S}\times V}S\times A^{*}\xrightarrow{\varphi}X. If R=𝒪XR=\mathcal{O}_{X}, this corresponds to the morphism θV:Rk\theta_{V}\colon R\to k. Moreover, φ\varphi composed with left multiplication defines an AA^{*}-action on XX; for every VAV\in A^{*} this gives a map lV:RRl_{V}\colon R\to R. In terms of the data of an equivariant formal group law F=(k,R,Δ,θ,y(ε))F=(k,R,\Delta,\theta,y(\varepsilon)), this can explicitly be written as

lV:RΔR^RθV^idRR.l_{V}\colon R\xrightarrow{\Delta}R\hat{\otimes}R\xrightarrow{\theta_{V}\hat{\otimes}\operatorname{id}_{R}}R.

Given VAV\in A^{*}, we set

y(V)=lV(y(ϵ))R,y(V)=l_{V}(y(\epsilon))\in R,

which generates the kernel of θV1\theta_{V^{-1}}. If AA is trivial, we have RkyR\cong k\llbracket y\rrbracket. We want to describe an analog for general AA. A complete flag for AA is a sequence of characters f=V1,V2,(A)f=V_{1},V_{2},\ldots\in(A^{*})^{\mathbb{N}} such that every character appears infinitely often. Given such a flag and nn\in\mathbb{N}, we set

y(Wn)=y(Vn)y(Vn1)y(V1).y(W_{n})=y(V_{n})y(V_{n-1})\cdots y(V_{1}).

Then every element xx of RR can be written uniquely as

x=nanfy(Wn)x=\sum_{n\in\mathbb{N}}a_{n}^{f}y(W_{n}) (2.7)

for coefficients anfka_{n}^{f}\in k [Hau, Section 2.2]. Hence, as a kk-module, RR is isomorphic to a countable infinite product of copies of kk.

2.2 Lazard rings

Our aim in this subsection is to recall the definition of the universal ring for equivariant formal group laws and to clarify its universal property. Let us begin by considering a very strict form of morphisms of equivariant formal group laws.

Definition 2.8.

A morphism between AA-equivariant formal group laws (k1,R1,Δ1,θ1,y(ε)1)(k_{1},R_{1},\Delta_{1},\theta_{1},y(\varepsilon)_{1}) and (k2,R2,Δ2,θ2,y(ε)2)(k_{2},R_{2},\Delta_{2},\theta_{2},y(\varepsilon)_{2}) is a pair of maps f:k1k2f\colon k_{1}\to k_{2} and g:R1R2g\colon R_{1}\to R_{2} which are compatible with both the comultiplications Δ\Delta and the augmentations θ\theta and which send y(ε)1y(\varepsilon)_{1} to y(ε)2y(\varepsilon)_{2}.

This leads to a category AA-FGL of AA-equivariant formal group laws. In [CGK] it is shown that this category has an initial object FuniF^{\operatorname{uni}}, the ground ring of which is called the AA-equivariant Lazard ring and denoted LAL_{A}. In fact, the category of AA-equivariant formal group laws is equivalent to the category of commutative rings under LAL_{A}. To discuss this, note first that the forgetful functor

A-FGLCRing,(k,R,Δ,θ,y(ε))\displaystyle A\text{-FGL}\to\mathrm{CRing},\qquad(k,R,\Delta,\theta,y(\varepsilon)) k\displaystyle\mapsto k

into the category of commutative rings is cofibered in groupoids. Concretely this boils down to the following two observations:

  • Every morphism of AA-equivariant formal group laws whose first component f:k1k2f\colon k_{1}\to k_{2} is the identity map is an isomorphism. One observes indeed that the diagram

    R1\textstyle{R_{1}\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}\scriptstyle{\cong}R2\textstyle{R_{2}\ignorespaces\ignorespaces\ignorespaces\ignorespaces}\scriptstyle{\cong}k1=k2\textstyle{\prod_{\mathbb{N}}k_{1}=\prod_{\mathbb{N}}k_{2}}

    obtained from Eq. 2.7 commutes. We call such an isomorphism living over the identity a very strict isomorphism between AA-equivariant formal group laws, in order to distinguish from other kinds of isomorphisms.

  • Given a morphism f:k1k2f\colon k_{1}\to k_{2} and an AA-equivariant formal group law FF over k1k_{1}, one can define a pushforward fFf_{*}F over k2k_{2} with the usual universal property. Its underlying k2k_{2}-algebra is given by a completion of Rk1k2R\otimes_{k_{1}}k_{2}, where RR is the underlying k1k_{1}-algebra of FF (cf. [GreenleesFGL, Section 2.E], [Hau, Section 2.3]).

Note that the only automorphism of an AA-equivariant formal group law over kk which is also a very strict isomorphism is the identity map. Hence, given two AA-equivariant formal group laws over kk, there either exists a unique very strict isomorphism between them or none at all. For this reason it is usually harmless to identify two very strictly isomorphic AA-equivariant formal group laws, and we will often do so.

Now, given a map f:LAkf\colon L_{A}\to k there is an induced AA-equivariant formal group law fFunif_{*}F^{\operatorname{uni}} over kk obtained by pushing forward the universal AA-equivariant formal group law. Given an AA-equivariant FGL FF over kk, we can apply this to the first component f:LAkf\colon L_{A}\to k of the unique map of AA-equivariant formal group laws FuniFF^{\operatorname{uni}}\to F. The resulting morphism fFuniFf_{*}F^{\operatorname{uni}}\to F is necessarily a very strict isomorphism. So, as claimed above, we obtain:

Corollary 2.9.

The functor

CAlgLAA-FGL,(f:LAk)fFuni\displaystyle\mathrm{CAlg}_{L_{A}}\to A\text{-FGL},\qquad(f\colon L_{A}\to k)\mapsto f_{*}F^{\operatorname{uni}}

from commutative LAL_{A}-algebras is an equivalence of categories. An inverse is given by sending an AA-equivariant formal group law F=(k,R,Δ,θ,y(ε))F=(k,R,\Delta,\theta,y(\varepsilon)) to the first component f:LAkf\colon L_{A}\to k of the unique morphism FuniFF^{\operatorname{uni}}\to F.

Remark 2.10.

The above proof is an instance of a general characterization of initial objects XX in categories cofibered in groupoids 𝒞F𝒟\mathcal{C}\xrightarrow{F}\mathcal{D}, namely that pushforward defines an equivalence of 𝒟F(X)/\mathcal{D}_{F(X)/-} with 𝒞\mathcal{C}.

Remark 2.11.

Non-equivariantly the kk-algebra RR is often fixed to be the power series ring ky(ε)k\llbracket y(\varepsilon)\rrbracket rather than a ring only isomorphic to it. With this convention, the category of formal group laws is isomorphic (not merely equivalent) to the category of commutative rings under LL. In other words, every very strict isomorphism of formal group laws is the identity. Equivariantly one needs to be a little more careful: The statement ‘AA-equivariant formal group laws are represented by the AA-equivariant Lazard ring’ is only true up to this notion of very strict isomorphism.

2.3 Global functorality

In this subsection, we will discuss both a covariant and a contravariant functoriality of the category of AA-equivariant formal groups in AA.

Definition 2.12.

Let α:BA\alpha\colon B\to A be a group homomorphism and let G=(BX)G=(B^{*}\to X) be a BB-equivariant formal group. We define the corestriction αG\alpha_{*}G to be the AA-equivariant formal group, which is the completion of the AA-equivariant group (AαBX)(A^{*}\xrightarrow{\alpha^{*}}B^{*}\to X)

Proposition 2.13.

For every injective group homomorphism α:BA\alpha\colon B\to A, the functor α\alpha_{*} from BB-equivariant formal groups to AA-equivariant formal groups is fully faithful. The essential image consists of those AA-equivariant formal groups where the homomorphism from AA^{*} factors through BB^{*}.

Proof.

For every BB-equivariant formal group GG, by construction αG\alpha_{*}G is the same group object as GG in formal schemes with the structure morphism ABGA^{*}\to B^{*}\to G (since ABA^{*}\to B^{*} is surjective, no completion is necessary). This implies fully faithfulness. ∎

Upon choosing coordinates, 2.12 corresponds to the construction of [Hau, Section 2.4]: given a BB-equivariant formal group law F=(k,R,Δ,θ,y(ε))F=(k,R,\Delta,\theta,y(\varepsilon)) and a group homomorphism α:BA\alpha\colon B\to A, there is an induced AA-equivariant formal group law αF\alpha_{*}F over the same ring kk, given by completing RR at products of the ideals IV=ker(RθVk)I_{V}=\ker(R\xrightarrow{\theta_{V}}k) for those VBV\in B^{*} which are in the image of α:AB\alpha^{*}\colon A^{*}\to B^{*}. This defines a functor α\alpha_{*} from BB-equivariant formal group laws to AA-equivariant formal group laws, which induces a map α:LALB\alpha^{*}\colon L_{A}\to L_{B} on Lazard rings. Hence we obtain a functor

𝐋:(abelian compact Lie groups)opcommutative rings\mathbf{L}\colon\text{(abelian compact Lie groups)}^{op}\to\text{commutative rings}

which we call the global Lazard ring. As shown in [Hau], LAL_{A} is isomorphic to πAMUA\pi_{*}^{A}MU_{A} and our map α\alpha^{*} corresponds to the restriction map on that level, explaining our terminology.

Remark 2.14.

By Pontryagin duality, the opposite category of abelian compact Lie groups is equivalent to the category of finitely generated abelian groups. Therefore, everything in this paper could alternatively be phrased in terms of finitely generated abelian groups rather than abelian compact Lie groups, and the more algebraically minded reader might prefer to do so.

In addition to this covariant functoriality, there is also a contravariant functoriality.

Definition 2.15.

Let α:AC\alpha\colon A\to C be a surjective group homomorphism and let G=(CX)G=(C^{*}\to X) be a CC-equivariant group. We define the coinduction αG\alpha^{*}G to be the AA-equivariant group (AX×CA)(A^{*}\to X\times_{C^{*}}A^{*}), where the target denotes the quotient of X×AX\times A^{*} by the antidiagonal CC^{*}-action.

Lemma 2.16.

For α:AC\alpha\colon A\to C a surjective group homomorphism and GG a CC-equivariant formal group, αG\alpha^{*}G is an AA-equivariant formal group, i.e. needs no additional completion.

Proof.

If G=(CSpecR)G=(C^{*}\to\operatorname{Spec}R) is a CC-equivariant formal group, then

SpecR×CASpecMapC(A,R).\operatorname{Spec}R\times_{C^{*}}A^{*}\cong\operatorname{Spec}\operatorname{Map}_{C^{*}}(A^{*},R).

The kk-algebra MapC(A,R)\operatorname{Map}_{C^{*}}(A^{*},R) is isomorphic to A/CR\prod_{A^{*}/C^{*}}R. As products of complete rings are complete, the claim follows. ∎

Proposition 2.17.

Let α:AC\alpha\colon A\to C be a surjective group homomorphism. As functors between CC-equivariant formal groups and AA-equivariant formal groups, α\alpha^{*} is the left adjoint of α\alpha_{*}.

Proof.

For an AA-equivariant group G=(AX)G=(A^{*}\to X), define α~G\widetilde{\alpha}_{*}G as CAXC^{*}\to A^{*}\to X. Then α\alpha^{*} and α~\widetilde{\alpha}_{*} are adjoints between CC-equivariant groups and AA-equivariant groups in the sense of Remark 2.3. Since completion is a right adjoint, the result follows from the previous lemma. ∎

2.4 Euler classes

Given an AA-equivariant formal group law F=(R,Δ,θ,y(ε))F=(R,\Delta,\theta,y(\varepsilon)) over kk, we can define Euler classes in kk. Recall that for VAV\in A^{*}, we set y(V)=lV(y(ϵ))Ry(V)=l_{V}(y(\epsilon))\in R. The corresponding Euler class is

eV=θϵ(y(V))=θV(y(ϵ))k.e_{V}=\theta_{\epsilon}(y(V))=\theta_{V}(y(\epsilon))\in k.

In terms of the associated equivariant formal group G=(φ:AX)G=(\varphi\colon A^{*}\to X), we have

S×φϵ,X,φVSSpeckθϵ,R,θVkSpeck/eVS\times_{\varphi_{\epsilon},X,\varphi_{V}}S\cong\operatorname{Spec}k\otimes_{\theta_{\epsilon},R,\theta_{V}}k\cong\operatorname{Spec}k/e_{V}

with S=SpeckS=\operatorname{Spec}k and φV\varphi_{V} being the composite SS×{V}S×AφXS\cong S\times\{V\}\subseteq S\times A^{*}\xrightarrow{\varphi}X. This implies:

Lemma 2.18.

For a given equivariant formal group law with notation as above:

  1. 1.

    The Euler class eVe_{V} is invertible iff S×φϵ,X,φVS=S\times_{\varphi_{\epsilon},X,\varphi_{V}}S=\varnothing.

  2. 2.

    The Euler class eVe_{V} is zero iff φV=φϵ\varphi_{V}=\varphi_{\epsilon}.

Thus, the vanishing or invertibility of Euler classes does not depend on chosen coordinates. This allows us to generalize these concepts to AA-equivariant formal groups in the following way:

Definition 2.19.

For an AA-equivariant formal group G=(φ:AX)G=(\varphi\colon A^{*}\to X), we say that the Euler class eVe_{V} is invertible if S×φϵ,X,φVS=S\times_{\varphi_{\epsilon},X,\varphi_{V}}S=\varnothing and that eVe_{V} is zero if φV=φϵ\varphi_{V}=\varphi_{\epsilon}.

Informally, eVe_{V} is invertible if and only if the images of S×{ϵ}S\times\{\epsilon\} and S×{V}S\times\{V\} in XX are disjoint.

Example 2.20.

Let 𝔾m=Spec[x±1]\mathbb{G}_{m}=\operatorname{Spec}\mathbb{Z}[x^{\pm 1}] be the multiplicative group over S=SpecS=\operatorname{Spec}\mathbb{Z}. We choose the group homomorphism φ:C2=(C2)𝔾m\varphi\colon C_{2}=(C_{2})^{*}\to\mathbb{G}_{m} picking the units {±1}\{\pm 1\} in \mathbb{Z}. This defines the structure of a C2C_{2}-equivariant group, and its completion is a C2C_{2}-equivariant formal group we call 𝔾^mC2\widehat{\mathbb{G}}_{m}^{C_{2}}. Let V(C2)V\in(C_{2})^{*} be the unique non-trivial character. One computes

Spec×φϵ,𝔾^mC2,φVSpecSpec×φϵ,𝔾m,φVSpecSpec/2.\operatorname{Spec}\mathbb{Z}\times_{\varphi_{\epsilon},\widehat{\mathbb{G}}_{m}^{C_{2}},\varphi_{V}}\operatorname{Spec}\mathbb{Z}\cong\operatorname{Spec}\mathbb{Z}\times_{\varphi_{\epsilon},\mathbb{G}_{m},\varphi_{V}}\operatorname{Spec}\mathbb{Z}\cong\operatorname{Spec}\mathbb{Z}/2.

Thus, eV=±2e_{V}=\pm 2, depending on the choice of coordinate. See also [StricklandMulti, Section 7] and [GreenleesFGL, Section 7] for more information on this and related examples. Note in particular that our example is the pushforward of the true C2C_{2}-equivariant multiplicative formal group (given by a completion of Spec[(C2×𝐓)]\operatorname{Spec}\mathbb{Z}[(C_{2}\times\mathbf{T})^{*}]) along the map [C2]\mathbb{Z}[C_{2}^{*}]\to\mathbb{Z} classifying ±1\pm 1.

Spf(t)\mathrm{Spf}(\mathbb{Z}\llbracket t\rrbracket)Spf(t)\mathrm{Spf}(\mathbb{Z}\llbracket t\rrbracket)Spec()\mathrm{Spec}(\mathbb{Z})(2)(2)
Figure 1: A schematic picture of 𝔾^mC2\widehat{\mathbb{G}}_{m}^{C_{2}} from Example 2.20

We will use several times the following lemma, taken from [Hau, Corollary 2.8] and the explanation thereafter:

Lemma 2.21.

Let BAB\subseteq A be a subgroup. Then the restriction map LALBL_{A}\to L_{B} is surjective, with kernel IABI_{A}^{B} generated by the Euler classes eVe_{V} where VV is running over a generating set of ker(AB)\ker(A^{*}\to B^{*}).

Similarly, the following holds on the level of equivariant formal groups.

Proposition 2.22.

Let GG be an AA-equivariant formal group.

  1. 1.

    Let α:AC\alpha\colon A\to C be a surjective group homomorphism. Assume that for Vim(CA)V\notin\operatorname{im}(C^{*}\to A^{*}), the Euler class eVe_{V} is invertible. Then ααGG\alpha^{*}\alpha_{*}G\to G is an isomorphism.

  2. 2.

    Let α:BA\alpha\colon B\to A be an injective group homomorphism. Assume that for Vker(AB)V\in\ker(A^{*}\to B^{*}), we have eV=0e_{V}=0. Then GG is in the essential image of α\alpha_{*}.

Proof.

Let G=(φ:AX)G=(\varphi\colon A^{*}\to X). Fixing a coordinate y(ϵ)y(\epsilon) Zariski-locally and choosing a complete flag, XX is defined by the directed system (SpecR/(y(V1)y(Vn)))n(\operatorname{Spec}R/(y(V_{1})\cdots y(V_{n})))_{n} and each of these terms has underlying space i=1nim(φVi)SpecR\bigcup_{i=1}^{n}\operatorname{im}(\varphi_{V_{i}})\subseteq\operatorname{Spec}R.

In the first item, Lemma 2.18 implies that im(φV)\operatorname{im}(\varphi_{V}) and im(φW)\operatorname{im}(\varphi_{W}) intersect each other in Spec𝒪X\operatorname{Spec}\mathcal{O}_{X} only if [V]=[W]A/C[V]=[W]\in A^{*}/C^{*}. Thus, the underlying space of (SpecR/(y(V1)y(Vn)))n(\operatorname{Spec}R/(y(V_{1})\cdots y(V_{n})))_{n} decomposes into closed subspaces νA/CViνim(φVi)\coprod_{\nu\in A^{*}/C^{*}}\bigcup_{V_{i}\in\nu}\operatorname{im}(\varphi_{V_{i}}), of which only finitely many are non-empty. This induces decompositions of the schemes SpecR/(y(V1)y(Vn)))\operatorname{Spec}R/(y(V_{1})\cdots y(V_{n}))) and we obtain thus an AA^{*}-equivariant isomorphism GA/CαGG\cong\coprod_{A^{*}/C^{*}}\alpha_{*}G on the level of formal schemes.

By construction, ααG\alpha^{*}\alpha_{*}G decomposes in the same way. On the unit copy, the map ααGG\alpha^{*}\alpha_{*}G\to G is an isomorphism since αααGαG\alpha_{*}\alpha^{*}\alpha_{*}G\to\alpha_{*}G is one by 2.17. For the other copies, this follows by the AA^{*}-equivariance of the map ααGG\alpha^{*}\alpha_{*}G\to G.

For the second item: by definition, the structure morphism φ:AX\varphi\colon A^{*}\to X of GG factors as ABφXA^{*}\to B^{*}\xrightarrow{\varphi^{\prime}}X. The result follows from 2.13. ∎

The second part is also true in the setting of equivariant formal group laws, as shown in [Hau, Lemma 2.7]. The analog of the first part becomes more complicated as ααG\alpha^{*}\alpha_{*}G does not have a canonical coordinate; we will talk more about it in 2.25.

Corollary 2.23.

Let α:AB\alpha\colon A\to B be a surjective group homomorphism. Then α\alpha^{*} is a fully faithful embedding from the category of BB-equivariant formal groups to that of AA-equivariant formal groups. The image consists of those AA-equivariant formal groups such that eVe_{V} is invertible for VV not in the image of BAB^{*}\to A^{*}.

Proof.

If GG is a BB-equivariant formal group, then αG\alpha^{*}G has the property that eVe_{V} is invertible for VV not in the image of BAB^{*}\to A^{*} by construction. By the preceding proposition, invertibility of these Euler classes characterizes the image of α\alpha^{*}. Moreover, α\alpha^{*} is fully faithful since GααGG\to\alpha_{*}\alpha^{*}G is an isomorphism by construction. ∎

The following proposition provides essentially a classification of equivariant formal groups over fields. The same statement already appears in [StricklandMulti, Corollary 8.3].

Proposition 2.24.

Let kk be a field and GG be an AA-equivariant formal group over kk. Denote by A/BA/B the Pontryagin dual of the subgroup {VA:eV=0}A\{V\in A^{*}:e_{V}=0\}\subseteq A^{*} Denote further by AqA/Bi{1}A\xrightarrow{q}A/B\xleftarrow{i}\{1\} the obvious morphisms. Then GqipGG\cong q^{*}i_{*}p_{*}G, where pGp_{*}G is the non-equivariant formal group defined by p:A{1}p\colon A\to\{1\}.

Proof.

Assume eV=0e_{V}=0, i.e. φϵ=φV\varphi_{\epsilon}=\varphi_{V}. By the AA^{*}-action, this implies φW=φVW\varphi_{W}=\varphi_{VW} for every WAW\in A^{*}. Setting W=V1W=V^{-1}, this implies eV1=0e_{V^{-1}}=0. Moreover, if φϵ=φW\varphi_{\epsilon}=\varphi_{W}, this implies φϵ=φVW\varphi_{\epsilon}=\varphi_{VW}. Thus, {VA:eV=0}A\{V\in A^{*}:e_{V}=0\}\subseteq A^{*} is indeed a subgroup.

Since kk is a field, eV=0e_{V}=0 iff eVe_{V} is not invertible. By Proposition 2.22, we thus have GqiΓG\cong q^{*}i_{*}\Gamma for some non-equivariant formal group Γ\Gamma over kk. One computes ΓpqiΓpG\Gamma\cong p_{*}q^{*}i_{*}\Gamma\cong p_{*}G. ∎

For any AA, let ΦAL=LA[eV1]\Phi^{A}L=L_{A}[e_{V}^{-1}] be the localization of LAL_{A} away from all Euler classes eVe_{V} for VϵV\neq\epsilon. Our results above let us compute ΦAL\Phi^{A}L quite explicitely (cf. [GreenleesFGL, Corollary 6.4] and [Hau, Proposition 2.11]).

Proposition 2.25.

There is an isomorphism of the form

ΦALL[(b0V)±1,biV|i>0,VA{ϵ}].\Phi^{A}L\cong L[(b_{0}^{V})^{\pm 1},b_{i}^{V}\ |\ i>0,V\in A^{*}-\{\epsilon\}].
Proof.

The ring ΦAL\Phi^{A}L classifies AA-equivariant formal group laws F=(R,Δ,θ,y(ε))F=(R,\Delta,\theta,y(\varepsilon)) such that eVe_{V} is invertible for all VϵV\neq\epsilon. By Proposition 2.22, Rmap(A,R^)R\cong\operatorname{map}(A^{*},\widehat{R}), where R^\widehat{R} is the completion of RR at the augmentation ideal. The structure of FF determines the structure of a non-equivariant formal group law pFp_{*}F on R^\widehat{R}, where p:A{1}p\colon A\to\{1\} is the projection; in particular, we obtain an isomorphism R^ky\widehat{R}\cong k\llbracket y\rrbracket, where yy is the image of y(ϵ)y(\epsilon). Vice versa, Δ\Delta and θ\theta are determined by pFp_{*}F. In particular, θϵ\theta_{\epsilon} is the composite map(A,R^)evϵR^k\operatorname{map}(A^{*},\widehat{R})\xrightarrow{\operatorname{ev}_{\epsilon}}\widehat{R}\to k. Thus we see that FF is determined by pFp_{*}F, plus a choice of y(ϵ)y(\epsilon) mapping to yy under evϵ\operatorname{ev}_{\epsilon}. Such y(ϵ)y(\epsilon) are exactly those elements (yV)map(A,ky)(y^{V})\in\operatorname{map}(A^{*},k\llbracket y\rrbracket) such that yϵ=yy^{\epsilon}=y and yV=b0V+b1Vy+y^{V}=b_{0}^{V}+b_{1}^{V}y+\cdots with b0Vk×b_{0}^{V}\in k^{\times}. This gives the result. ∎

Remark 2.26.

The elements biVΦALb_{i}^{V}\in\Phi^{A}L in the previous proposition already come from elements γiVLA\gamma_{i}^{V}\in L_{A}, which are uniquely defined by the property that

eϵτ=γ0V+γ1VeV1τ+γ2V(eV1τ)2++γnV(eV1τ)nLA×𝐓/(eV1τ)n+1,e_{\epsilon\otimes\tau}=\gamma_{0}^{V}+\gamma_{1}^{V}e_{V^{-1}\otimes\tau}+\gamma_{2}^{V}(e_{V^{-1}\otimes\tau})^{2}+\ldots+\gamma_{n}^{V}(e_{V^{-1}\otimes\tau})^{n}\in L_{A\times\mathbf{T}}/(e_{V^{-1}\otimes\tau})^{n+1},

for all nn\in\mathbb{N}. Here, τ𝐓\tau\in\mathbf{T}^{*} denotes the tautological character for the circle group 𝐓\mathbf{T}. In particular, γ0V\gamma_{0}^{V} equals the Euler class eVe_{V}. The elements γiV\gamma_{i}^{V} are natural in the sense that αγiV=γiαV\alpha^{*}\gamma_{i}^{V}=\gamma_{i}^{\alpha^{*}V} for every group homomorphism α:BA\alpha\colon B\to A. We refer to [Hau, Section 2.7] for more details on this construction.

2.5 The relationship between Lazard rings at different groups and their completions

While not needed for our classification of invariant prime ideals, it will be necessary for our study of containments between invariant prime ideals to have a deeper look upon how Lazard rings at different groups relate. These properties are all based on the identification of the global Lazard ring with equivariant complex bordism in [Hau].

Proposition 2.27 ([Hau], Proposition 5.50, Corollary 5.33, Lemma 5.28).
  1. 1.

    For every AA and every non-torsion character VAV\in A^{*}, the sequence

    0LAeVLAresAker(V)Lker(V)00\to L_{A}\xrightarrow{e_{V}\cdot}L_{A}\xrightarrow{\operatorname{res}^{A}_{\ker(V)}}L_{\ker(V)}\to 0

    is exact. In particular, all Euler classes eVLAe_{V}\in L_{A} for non-torsion characters VV are non-zero divisors.

  2. 2.

    For every AA, the complete LAL_{A}-Hopf algebra RR of the universal AA-equivariant formal group law is canonically isomorphic to the completion

    limn,V1,,VnA(LA×𝐓)/IV1IVn,\lim_{n\in\mathbb{N},V_{1},\ldots,V_{n}\in A^{*}}(L_{A\times\mathbf{T}})/I_{V_{1}}\cdots I_{V_{n}},

    where IVjI_{V_{j}} is the kernel of the restriction map (id,Vj):LA×𝐓LA(id,V_{j})^{*}\colon L_{A\times\mathbf{T}}\to L_{A}. More generally, R^nR^{\hat{\otimes}n} is a completion of LA×𝐓nL_{A\times\mathbf{T}^{n}}. Under this identification

    1. (a)

      the comultiplication RR^RR\to R\hat{\otimes}R and the augmentations θV:RLA\theta_{V}\colon R\to L_{A} are induced by the maps (idA,m):LA×𝐓LA×𝐓×𝐓(\operatorname{id}_{A},m)^{*}\colon L_{A\times\mathbf{T}}\to L_{A\times\mathbf{T}\times\mathbf{T}} and (idA,V):LA×𝐓LA(\operatorname{id}_{A},V)^{*}\colon L_{A\times\mathbf{T}}\to L_{A} on completion, and

    2. (b)

      the elements y(V)y(V) are the image of eVτLA×𝐓e_{V\otimes\tau}\in L_{A\times\mathbf{T}} under the completion map, where τ𝐓\tau\in\mathbf{T}^{*} is the tautological character.

The special case of (2) for AA the trivial group is particularly important: Completing L𝐓nL_{\mathbf{T}^{n}} at the kernel II of the augmentation L𝐓nLL_{\mathbf{T}^{n}}\to L yields a power series ring on nn generators Ly1,y2,,ynL\llbracket y_{1},y_{2},\ldots,y_{n}\rrbracket, where yiy_{i} is the Euler class of the ii-th projection 𝐓n𝐓\mathbf{T}^{n}\to\mathbf{T}. Moreover, the isomorphism

(L𝐓n)ILy1,y2,,yn(L_{\mathbf{T}^{n}})^{\wedge}_{I}\cong L\llbracket y_{1},y_{2},\ldots,y_{n}\rrbracket

is natural in 𝐓n\mathbf{T}^{n}, where

  • the functoriality of (L𝐓n)I(L_{\mathbf{T}^{n}})^{\wedge}_{I} is induced by the global functoriality of L𝐓nL_{\mathbf{T}^{n}}, and

  • the functoriality of Ly1,y2,,ynL\llbracket y_{1},y_{2},\ldots,y_{n}\rrbracket is through the universal formal group law over LL.

For example, for any nn\in\mathbb{N} the square

L𝐓\textstyle{L_{\mathbf{T}}\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}c\scriptstyle{c}[n]\scriptstyle{[n]^{*}}Ly\textstyle{L\llbracket y\rrbracket\ignorespaces\ignorespaces\ignorespaces\ignorespaces}[n]F\scriptstyle{[n]_{F}}L𝐓\textstyle{L_{\mathbf{T}}\ignorespaces\ignorespaces\ignorespaces\ignorespaces}c\scriptstyle{c}Ly\textstyle{L\llbracket y\rrbracket}

commutes, where [n]:𝐓𝐓[n]\colon\mathbf{T}\to\mathbf{T} is the nnth power map, and [n]F:LyLy[n]_{F}\colon L\llbracket y\rrbracket\to L\llbracket y\rrbracket sends yy to the nn-series [n]F(y)[n]_{F}(y) with respect to the universal formal group law. This implies that the Euler class eτnL𝐓e_{\tau^{n}}\in L_{\mathbf{T}} for the nnth power map on 𝐓\mathbf{T} is sent to the nn-series [n]F(y)[n]_{F}(y) under the completion map. Similar statements hold for the collection of LA×𝐓nL_{A\times\mathbf{T}^{n}} for fixed AA and varying 𝐓n\mathbf{T}^{n}.

We further record, where again τ:𝐓𝐓\tau\colon\mathbf{T}\to\mathbf{T} denotes the identity character:

Corollary 2.28.

Let xL𝐓x\in L_{\mathbf{T}} be an element whose image in LyL\llbracket y\rrbracket is of the form λyk+higher order terms\lambda y^{k}+\text{higher order terms}. Then xx is uniquely divisible by eτke_{\tau}^{k} and the quotient x/eτkx/e_{\tau}^{k} restricts to λ\lambda at the trivial group.

Proof.

First we note that eτL𝐓e_{\tau}\in L_{\mathbf{T}} is a regular element by Proposition 2.27. Hence division by eτe_{\tau} is always unique if possible. By induction on kk the corollary then follows from the following facts, for an element zL𝐓z\in L_{\mathbf{T}} and its image c(z)L𝐓c(z)\in L_{\mathbf{T}}:

  1. 1.

    zz is divisible by eτe_{\tau} if and only if res𝐓1(z)=0\operatorname{res}^{\mathbf{T}}_{1}(z)=0.

  2. 2.

    The leading coefficient of c(z)c(z) is equal to res𝐓1(z)\operatorname{res}^{\mathbf{T}}_{1}(z).

  3. 3.

    If resT1(z)=0\operatorname{res}^{T}_{1}(z)=0 and hence zz is divisible by eτe_{\tau}, then c(z/eτ)=c(z)/yc(z/e_{\tau})=c(z)/y. ∎

Note that the global functoriality makes every equivariant Lazard ring LAL_{A} an algebra over the non-equivariant Lazard ring LL, and that all restriction maps α:LALB\alpha^{*}\colon L_{A}\to L_{B} are LL-algebra maps. We have the following:

Proposition 2.29 ([Hau], [Com96]).

LAL_{A} is free as a module over LL, for every abelian compact Lie group AA.

Corollary 2.30.

The exact sequences of LL-modules in Part 1 of Proposition 2.27 are split exact. In particular, they remain exact after applying any additive functor.

The following special case is of particular importance to us:

Corollary 2.31.

Let n¯={}n\in\overline{\mathbb{N}}=\mathbb{N}\cup\{\infty\} and InLI_{n}\subseteq L be the ideal generated by (v0,,vn1)(v_{0},\ldots,v_{n-1}). Then for every AA and every non-torsion character VAV\in A^{*}, the sequence

0LA/IneVLA/InresAker(V)Lker(V)/In00\to L_{A}/I_{n}\xrightarrow{e_{V}\cdot}L_{A}/I_{n}\xrightarrow{\operatorname{res}^{A}_{\ker(V)}}L_{\ker(V)}/I_{n}\to 0

is exact. In particular, the Euler classes eVLAe_{V}\in L_{A} for non-torsion characters VAV\in A^{*} remain non-zero divisors in LA/InL_{A}/I_{n}. In the terminology of [Hau], the assignment

ALA/InA\mapsto L_{A}/I_{n}

(together with the image of eτe_{\tau} under L𝐓L𝐓/InL_{\mathbf{T}}\to L_{\mathbf{T}}/I_{n}) is a regular global group law.

3 The Lazard Hopf algebroid and its associated stack

3.1 Strict isomorphisms and the Lazard Hopf algebroid

In this subsection, we will introduce one of our main objects of study, the Hopf algebroid (LA,SA)(L_{A},S_{A}) for equivariant formal group laws. There is a hierarchy of notions of isomorphisms between (equivariant) formal group laws, namely

  • isomorphisms, which do not need to respect the coordinate and are thus really isomorphisms between the underlying (equivariant) formal groups;

  • strict isomorphisms, which respect the coordinate up to quadratic terms;

  • very strict isomorphisms as in Section 2.2, which respect the coordinate strictly.

Already classically, strict isomorphisms are especially relevant since the Hopf algebroid modeled on them gives (MU,MUMU)(MU_{*},MU_{*}MU).

Definition 3.1.

A strict isomorphism between two AA-equivariant formal group laws

(k,R1,Δ1,θ1,y(ε)1) and (k,R2,Δ2,θ2,y(ε)2)(k,R_{1},\Delta_{1},\theta_{1},y(\varepsilon)_{1})\quad\text{ and }\quad(k,R_{2},\Delta_{2},\theta_{2},y(\varepsilon)_{2})

over the same ground ring kk is a kk-linear isomorphism

φ:R1R2\varphi\colon R_{1}\xrightarrow{\cong}R_{2}

of Hopf algebras over kAk^{A^{*}} such that y(ε)1y(\varepsilon)_{1} is sent to y(ε)2y(\varepsilon)_{2} modulo Iε2I_{\varepsilon}^{2}, where IεI_{\varepsilon} is the augmentation ideal in R2R_{2}. Explicitly, this means that (φφ)Δ1=Δ2φ(\varphi\otimes\varphi)\circ\Delta_{1}=\Delta_{2}\circ\varphi, θ2φ=θ1\theta_{2}\circ\varphi=\theta_{1} and φ(y(ε)1)=xy(ε)2\varphi(y(\varepsilon)_{1})=x\cdot y(\varepsilon)_{2} for some unit xR2x\in R_{2} which augments to 1k1\in k.

By definition, strict isomorphisms need not preserve the coordinate, hence they are generally not morphisms of formal group laws in the sense of Section 2.2. On the other hand, every very strict isomorphism is both a strict isomorphism and an isomorphism in the category of AA-equivariant formal group laws.

Let SI\operatorname{SI} be the category of strict isomorphisms of AA-equivariant formal group laws. More precisely, its objects are quadruples (k,F1,F2,φ)(k,F^{1},F^{2},\varphi) consisting of a commutative ring kk, two AA-equivariant formal group laws F1F^{1} and F2F^{2} over kk and a strict isomorphism φ\varphi between them. Morphisms between two such quadruples (k1,F11,F12,φϵ)(k_{1},F_{1}^{1},F_{1}^{2},\varphi_{\epsilon}) and (k2,F21,F22,φ2)(k_{2},F_{2}^{1},F_{2}^{2},\varphi_{2}) are given by a pair of morphisms f1:F11F21f_{1}\colon F_{1}^{1}\to F_{2}^{1} and f2:F12F22f_{2}\colon F_{1}^{2}\to F_{2}^{2} with the same underlying map k1k2k_{1}\to k_{2}, such that

R11\textstyle{R_{1}^{1}\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}φϵ\scriptstyle{\varphi_{\epsilon}}f1\scriptstyle{f_{1}}R12\textstyle{R_{1}^{2}\ignorespaces\ignorespaces\ignorespaces\ignorespaces}f2\scriptstyle{f_{2}}R21\textstyle{R_{2}^{1}\ignorespaces\ignorespaces\ignorespaces\ignorespaces}φ2\scriptstyle{\varphi_{2}}R22\textstyle{R_{2}^{2}}

commutes.

Proposition 3.2.

The category SISI has an initial object, whose underlying ring SAS_{A} is a localization of an infinite polynomial ring LA[a1f,a2f,]L_{A}[a_{1}^{f},a_{2}^{f},\dots] over the Lazard ring LAL_{A}.

By Remark 2.10, we can equivalently say that the functor

CAlgSA\displaystyle\mathrm{CAlg}_{S_{A}} SI\displaystyle\to\operatorname{SI}
(f:SAk)\displaystyle(f\colon S_{A}\to k) (k,(f1)Funi,(f2)Funi,id)\displaystyle\mapsto(k,(f_{1})_{*}F^{\operatorname{uni}},(f_{2})_{*}F^{\operatorname{uni}},\operatorname{id})

is an equivalence of categories. Here, we use that SISI is again cofibered in groupoids over commutative rings.

Before we prove the proposition, it will be good to review two general results about the maps θV:RkAk\theta_{V}\colon R\to k^{A^{*}}\to k for an AA-equivariant formal group law (k,R,Δ,θ,y(ε))(k,R,\Delta,\theta,y(\varepsilon)). Recall that after choosing a complete flag ff we can write every element xRx\in R uniquely as

x=nanfy(Wn)x=\sum_{n\in\mathbb{N}}a_{n}^{f}y(W_{n})

Given VAV\in A^{*}, we have

θV(x)=nanfθV(y(Wn))=nanfeVWn,\theta_{V}(x)=\sum_{n\in\mathbb{N}}a_{n}^{f}\theta_{V}(y(W_{n}))=\sum_{n\in\mathbb{N}}a_{n}^{f}e_{VW_{n}},

where eVWne_{VW_{n}} is defined as the product eVVneVVn1eVV1e_{V\cdot V_{n}}e_{V\cdot V_{n-1}}\cdots e_{V\cdot V_{1}}. This is a finite sum, since eVWn=0e_{VW_{n}}=0 if there exists some ini\leq n where Vi=V1V_{i}=V^{-1}. We obtain:

Lemma 3.3.

The augmentation θV\theta_{V} is a linear combination of anfa_{n}^{f} whose coefficients are products of Euler classes.

Moreover, we have the following:

Lemma 3.4.

Let F=(k,R,Δ,θ,y(ϵ))F=(k,R,\Delta,\theta,y(\epsilon)) be an AA-equivariant formal group law. Then an element xRx\in R is a unit if and only if θV(x)\theta_{V}(x) is a unit in kk for all VAV\in A^{*}.

Proof.

See [Hau, Lemma 2.3]. ∎

Proof of Proposition 3.2.

For every object (k,F1,F2,φ)SI(k,F^{1},F^{2},\varphi)\in\operatorname{SI} we can define a new AA-equivariant formal group law F~2\widetilde{F}^{2} for which the components k,R,Δk,R,\Delta and θ\theta agree with those of F1F^{1}, but y~(ε)2\widetilde{y}(\varepsilon)_{2} is defined as φ1(y(ε)2)\varphi^{-1}(y(\varepsilon)_{2}), the preimage of the coordinate of F2F^{2} under φ\varphi. Then we obtain a new object (k,F1,F~2,idR1)(k,F^{1},\widetilde{F}^{2},\operatorname{id}_{R_{1}}) which is isomorphic in SI\operatorname{SI} to (k,F1,F2,φ)(k,F^{1},F^{2},\varphi) via the commutative square

F1\textstyle{F^{1}\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}id\scriptstyle{id}id\scriptstyle{id}F~2\textstyle{\widetilde{F}^{2}\ignorespaces\ignorespaces\ignorespaces\ignorespaces}φ\scriptstyle{\varphi}F1\textstyle{F^{1}\ignorespaces\ignorespaces\ignorespaces\ignorespaces}φ\scriptstyle{\varphi}F2\textstyle{F^{2}}

Note that the two vertical maps are in fact very strict isomorphisms of AA-equivariant formal group laws. In summary, every object of SI\operatorname{SI} is isomorphic to one of the form (k,F1,F2,id)(k,F_{1},F_{2},\operatorname{id}), where F1F_{1} and F2F_{2} are given by the same kAk^{A^{*}}-augmented kk-Hopf algebra and the strict isomorphism is the identity. This is the same data as a single AA-equivariant formal group law FF together with a second choice of coordinate y(ε)2=xy(ε)1y(\varepsilon)_{2}=x\cdot y(\varepsilon)_{1} for some unit xRx\in R which augments to 11. Up to very strict isomorphism we can further assume that FF is the push-forward along a map LAkL_{A}\to k. We claim that the functor sending an AA-equivariant formal group law to the set of all units xRx\in R augmenting to 11 is representable by an LAL_{A}-algebra SAS_{A}. Indeed, a presentation for SAS_{A} is given by

SA=LA[a1f,a2f,][PV(1,a1f,a2f,)1|VA{ε}]S_{A}=L_{A}[a_{1}^{f},a_{2}^{f},\ldots][P_{V}(1,a_{1}^{f},a_{2}^{f},\ldots)^{-1}\ |\ V\in A^{*}-\{\varepsilon\}] (3.5)

where ff is a complete flag starting with ε\varepsilon and PVP_{V} is the linear combination expressing θV\theta_{V} in terms of the coefficients with respect to ff; see Lemma 3.3 and Lemma 3.4. Here we use that the units augmenting to 11 are precisely the elements of the form 1+n+anfy(Wn)1+\sum_{n\in\mathbb{N}^{+}}a_{n}^{f}y(W_{n}), with y(Wn)y(W_{n}) as in the end of Section 2.1.

Thus, the functor

CAlgSA\displaystyle\mathrm{CAlg}_{S_{A}} SI\displaystyle\to\operatorname{SI}
(f:SAk)\displaystyle(f\colon S_{A}\to k) (k,(f1)Funi,(f2)Funi,id)\displaystyle\mapsto(k,(f_{1})_{*}F^{\operatorname{uni}},(f_{2})_{*}F^{\operatorname{uni}},\operatorname{id})

is an equivalence of categories. ∎

Remark 3.6.

The same proof shows that the category of all (not necessarily strict) isomorphisms also has an initial object, whose underlying ring is SA[(a0f)±1]S_{A}[(a_{0}^{f})^{\pm 1}]. In fact, the only difference in the proof is that the equation for the unit xx is now of the form nanfy(Wn)\sum_{n\in\mathbb{N}}a_{n}^{f}y(W_{n}) so that the presentation for the analog of SAS_{A} becomes

LA[a0f,a1f,a2f,][(a0f)1,PV(a0f,a1f,a2f,)1|VA{ε}].L_{A}[a_{0}^{f},a_{1}^{f},a_{2}^{f},\ldots][(a_{0}^{f})^{-1},P_{V}(a_{0}^{f},a_{1}^{f},a_{2}^{f},\ldots)^{-1}\ |\ V\in A^{*}-\{\varepsilon\}].

Since PV(a0f,a1f,a2f,)=a0fPV(1,a1fa0f,a2fa0f,)P_{V}(a_{0}^{f},a_{1}^{f},a_{2}^{f},\ldots)=a_{0}^{f}P_{V}(1,\tfrac{a_{1}^{f}}{a_{0}^{f}},\tfrac{a_{2}^{f}}{a_{0}^{f}},\ldots), this ring is indeed isomorphic to

LA[(a0f)±1,a1fa0f,a2fa0f,][PV(1,a1fa0f,a2fa0f,)1|VA{ε}]SA[(a0f)±1].L_{A}[(a_{0}^{f})^{\pm 1},\tfrac{a_{1}^{f}}{a_{0}^{f}},\tfrac{a_{2}^{f}}{a_{0}^{f}},\ldots][P_{V}(1,\tfrac{a_{1}^{f}}{a_{0}^{f}},\tfrac{a_{2}^{f}}{a_{0}^{f}},\ldots)^{-1}\ |\ V\in A^{*}-\{\varepsilon\}]\cong S_{A}[(a_{0}^{f})^{\pm 1}].

There are functors

s,t:SIA-FGLs,t\colon\operatorname{SI}\to\text{A-FGL}

sending a strict isomorphism to its source and target, respectively, as well as an ‘identity’ functor

A-FGLSI,\text{A-FGL}\to\operatorname{SI},

an ‘inverse’ functor

i:SISIi\colon\operatorname{SI}\to\operatorname{SI}

and a ‘composition’

c:SI×SISI,c\colon\operatorname{SI}\times\operatorname{SI}\to\operatorname{SI},

which restrict to the full subcategory of those objects of SI\operatorname{SI} where the isomorphism φ\varphi is given by the identity. By representability we obtain analogous source and target maps s,t:LASAs,t\colon L_{A}\to S_{A}, an identity map SALAS_{A}\to L_{A}, an inverse map i:SASAi\colon S_{A}\to S_{A} and a composition map c:SASALASAc\colon S_{A}\to S_{A}\otimes_{L_{A}}S_{A}.

Corollary 3.7.

The pair (LA,SA)(L_{A},S_{A}) together with the above structure defines a Hopf algebroid i.e. a cogroupoid object in commutative rings. The associated functor

CRingGroupoids\mathrm{CRing}\to\mathrm{Groupoids}

is equivalent to the one sending a commutative ring kk to the groupoid of AA-equivariant formal group laws over kk and strict isomorphisms between them.

Remark 3.8.

The Hopf algebroid (LA,SA)(L_{A},S_{A}) has a natural grading. This can be constructed in three equivalent ways:

  • [Hau, Corollary 5.6] shows that the global Lazard ring admits a unique grading such that it defines a graded global group law. By the same arguments, the ‘universal global group law with a strict nn-tuple of coordinates’ 𝐋(n)\mathbf{L}^{(n)} (cf., [Hau, Section 5.8]) also carries a unique grading for every nn such that the coordinates have degree 2-2. The source-target maps s,t:𝐋𝐋(2)s,t\colon\mathbf{L}\to\mathbf{L}^{(2)}, the identity map 𝐋(2)𝐋\mathbf{L}^{(2)}\to\mathbf{L}, the inverse map i:𝐋(2)𝐋(2)i\colon\mathbf{L}^{(2)}\to\mathbf{L}^{(2)} and the composition map c:𝐋(2)𝐋(3)c\colon\mathbf{L}^{(2)}\to\mathbf{L}^{(3)} all preserve this grading, since they are defined through their effect on the respective coordinates. Evaluating these maps at a group AA yields the Hopf algebroids (LA,SA)(L_{A},S_{A}), which hence inherit a grading compatible with all restriction and inflation maps.

  • The groupoid-valued functor represented by (LA,SA)(L_{A},S_{A}) admits a 𝔾m\mathbb{G}_{m}-action from multiplying the coordinate of the equivariant formal group law by a unit uu and acting on strict isomorphisms by multiplying anfa_{n}^{f} by unu^{n}. This corresponds to an even grading on (LA,SA)(L_{A},S_{A}), putting e.g. anfa_{n}^{f} in degree 2n2n.

  • By its interpretation as representing the groupoid of AA-equivariant formal group laws and isomorphisms between them (see Remark 3.6), (LA,SA[(a0f)1])(L_{A},S_{A}[(a_{0}^{f})^{-1}]) obtains the structure of a Hopf algebroid. An element ss in SAS_{A} is of degree 2n2n if applying the composition map SA[(a0f)1]SA[(a0f)1]LASA[(a0f)1]S_{A}[(a_{0}^{f})^{-1}]\to S_{A}[(a_{0}^{f})^{-1}]\otimes_{L_{A}}S_{A}[(a_{0}^{f})^{-1}] to ss gives (a0f)nc(s)(a_{0}^{f})^{n}c(s), where c:SASALASAc\colon S_{A}\to S_{A}\otimes_{L_{A}}S_{A} is the composition map of (LA,SA)(L_{A},S_{A}).

One can show that this is the same grading coming from the isomorphism

(LA,SA)(πAMUA,πAMUAMUA)(L_{A},S_{A})\cong(\pi_{*}^{A}MU_{A},\pi_{*}^{A}MU_{A}\operatorname{\wedge}MU_{A})

from [Hau, Theorem E].

Given a strict isomorphism φ:F1F2\varphi\colon F_{1}\cong F_{2} of BB-equivariant formal group laws and a group homomorphism α:BA\alpha\colon B\to A, we obtain an induced strict isomorphism αφ:αF1αF2\alpha_{*}\varphi\colon\alpha_{*}F_{1}\cong\alpha_{*}F_{2} by completion. This assignment is compatible with composition of strict isomorphisms. Therefore, the functor 𝐋\mathbf{L} from Section 2.3 extends to a functor

𝐋:(abelian compact Lie groups)opHopf algebroids,\mathbf{L}\colon\text{(abelian compact Lie groups)}^{op}\to\text{Hopf algebroids},

which we call the global Lazard Hopf algebroid.

3.2 The moduli stack of equivariant formal groups

We have discussed above that the Hopf algebroid (LA,SA)(L_{A},S_{A}) represents the functor sending a commutative ring to the groupoid of AA-equivariant formal group laws and strict isomorphisms between them. As discussed in Remark 3.8, (LA,SA)(L_{A},S_{A}) is naturally a graded Hopf algebroid. On the other hand, we have discussed in Remark 3.6 the ungraded Hopf algebroid (LA,SA[a0±1])(L_{A},S_{A}[a_{0}^{\pm 1}]) classifying AA-equivariant formal group laws and all isomorphisms between them. It is easy to see that this is precisely the ungraded Hopf algebroid associated to the graded Hopf algebroid (LA,SA)(L_{A},S_{A}) in the sense of [MeierOzornova, Section 4.1].111[MeierOzornova] uses an algebraic grading convention, while we use a topological one; thus one has to double all degrees. We will follow [MeierOzornova, Definition 4.1] by defining the stack associated to a graded Hopf algebroid as the fpqc-stackification of the groupoid-valued functor corepresented by its associated ungraded Hopf algebroid. In particular, the stack associated to the graded Hopf algebroid (LA,SA)(L_{A},S_{A}) is the same as the stack associated to the ungraded Hopf algebroid (LA,SA[a0±1])(L_{A},S_{A}[a_{0}^{\pm 1}]). Equivalently, it is the quotient of the stack associated to (LA,SA)(L_{A},S_{A}) by the 𝔾m\mathbb{G}_{m}-action induced by the grading.

Proposition 3.9.

Sending an AA-equivariant formal group law to its underlying AA-equivariant formal group defines an equivalence from the stack associated to the graded Hopf algebroid (LA,SA)(L_{A},S_{A}) to FGA\mathcal{M}_{FG}^{A}, the (pseudo-)functor sending a commutative ring to the groupoid of AA-equivariant formal groups over it.

Proof.

It suffices to show that FGA\mathcal{M}_{FG}^{A} is an fpqc-stack. Indeed: denote the stack associated to the graded Hopf algebroid (LA,SA)(L_{A},S_{A}) by 𝒳A\mathcal{X}_{A}; equivalently, this is the stack associated to the ungraded Hopf algebroid (LA,SA[a0±1])(L_{A},S_{A}[a_{0}^{\pm 1}]). Since the augmentation ideal of every AA-equivariant formal group (X,φ)(X,\varphi) is by definition fpqc-locally trivial, (X,φ)(X,\varphi) comes after fpqc-base change from an AA-equivariant formal group law (see Lemma 2.6). Thus, 𝒳AFGA\mathcal{X}_{A}\to\mathcal{M}_{FG}^{A} is essentially surjective as a functor of stacks. Moreover, it is fully faithful since isomorphisms between AA-equivariant formal group laws are precisely isomorphisms of the underlying AA-equivariant formal groups. Thus, it remains to show that FGA\mathcal{M}_{FG}^{A} satisfies fqpc-descent on morphisms and objects.

Given two formal kk-schemes XX and YY, the functor

k-algebrasSet,KHomK(X×SpeckSpecK,Y×SpeckSpecK)k\text{-algebras}\to\operatorname{Set},\qquad K\mapsto\operatorname{Hom}_{K}(X\times_{\operatorname{Spec}k}\operatorname{Spec}K,Y\times_{\operatorname{Spec}k}\operatorname{Spec}K)

is an fpqc-sheaf. Indeed, if we view XX and YY as ind-objects (Xi)(X_{i}) and (Yj)(Y_{j}), we can rewrite this Hom\operatorname{Hom} as limicolimjHomK(Xi×SpeckSpecK,Yj×SpeckSpecK)\lim_{i}\operatorname{colim}_{j}\operatorname{Hom}_{K}(X_{i}\times_{\operatorname{Spec}k}\operatorname{Spec}K,Y_{j}\times_{\operatorname{Spec}k}\operatorname{Spec}K) and equalizers commute with filtered colimits in sets. This easily implies that FGA\mathcal{M}_{FG}^{A} satisfies fpqc-descent on morphisms.

Let Fin(A)\operatorname{Fin}(A^{*}) denote the directed set of finite multi subsets of AA^{*} (i.e. elements can occur more than once), ordered by inclusion. Sending an AA-equivariant formal group (X,φ)(X,\varphi) over kk to the system (Spec𝒪X/)(\operatorname{Spec}\mathcal{O}_{X}/\mathcal{I}), where \mathcal{I} runs over all finite products of the ker(𝒪Xϕ(V)k)\ker(\mathcal{O}_{X}\xrightarrow{\phi(V)^{*}}k) for VAV\in A^{*}, defines a functor from AA-equivariant formal groups over kk to Fun(Fin(A),Affk)\operatorname{Fun}(\operatorname{Fin}(A^{*}),\operatorname{Aff}_{k}).

Given a descent datum for AA-equivariant formal groups for the fpqc-cover TST\to S, we thus obtain a descent datum for a Fin(A)\operatorname{Fin}(A^{*})-diagram of affine schemes (with closed immersions as transition maps), which descends thus to a Fin(A)\operatorname{Fin}(A^{*})-diagram of closed immersion in Affk\operatorname{Aff}_{k}. This diagram defines a formal kk-scheme XX. By descent for morphisms between formal kk-schemes, XX obtains a group structure and also a group homomorphism φ:S×AX\varphi\colon S\times A^{*}\to X. Conditions (1) and (2) for an AA-equivariant formal group are fulfilled by construction. ∎

Remark 3.10.

As every equivariant formal group comes Zariski-locally from an equivariant formal group law, in our case only a Zariski-stackification was necessary to pass from (LA,SA[u±1])(L_{A},S_{A}[u^{\pm 1}]) to FGA\mathcal{M}_{FG}^{A}.

Proposition 3.11.

Let BAB\subseteq A be a subgroup of an abelian group AA. Denote by α:BA\alpha\colon B\to A the inclusion and by q:AA/Bq\colon A\to A/B the projection.

  1. (i)

    The functor qq^{*} induces an open immersion FGA/BFGA\mathcal{M}_{FG}^{A/B}\to\mathcal{M}_{FG}^{A} whose image is the common non-vanishing locus of the Euler classes eVe_{V} for all Vim((A/B)A)V\notin\operatorname{im}((A/B)^{*}\to A^{*}).

  2. (ii)

    The functor α\alpha_{*} induces a closed immersion FGBFGA\mathcal{M}_{FG}^{B}\to\mathcal{M}_{FG}^{A}, inducing an equivalence of FGB\mathcal{M}_{FG}^{B} to the common vanishing locus of the eVe_{V} for all Vker(AB)V\in\ker(A^{*}\to B^{*}).

Proof.

The first part is a reformulation of 2.23. The immersion is open since it is open after pullback to SpecLA\operatorname{Spec}L_{A}.

For the second, note that by Proposition 2.22 every AA-equivariant formal group such that eV=0e_{V}=0 for all Vker(AB)V\in\ker(A^{*}\to B^{*}) is of the form αG\alpha_{*}G for GG a BB-equivariant formal group. Moreover, by construction, this vanishing of Euler classes is true for all AA-equivariant formal groups of the form αG\alpha_{*}G and thus characterizes the image of α\alpha_{*}. The substack given by this image is closed since it is closed after pullback to SpecLA\operatorname{Spec}L_{A}. Moreover, α\alpha_{*} is fully faithful by 2.13. ∎

Remark 3.12.

With notation as in the preceding proposition, we have ααGG\alpha^{*}\alpha_{*}G\cong G for every BB-equivariant formal group GG and HqqHH\cong q_{*}q^{*}H for every A/BA/B-equivariant formal group HH. Thus, the substacks in the preceding proposition are retractive.

Example 3.13.

3.11 gives closed immersions of FG{1}\mathcal{M}_{FG}^{\{1\}} and FGC2\mathcal{M}_{FG}^{C_{2}} into FGC4\mathcal{M}_{FG}^{C_{4}}. The first is the common vanishing locus of all Euler classes (which equals the vanishing locus of the Euler class of one of the two generators of (C4)(C_{4})^{*}, cf. Proposition 2.24). Its complement is the open substack given by the non-vanishing locus of the Euler class of the generators and hence equivalent to FGC4/C2\mathcal{M}_{FG}^{C_{4}/C_{2}}. The second, i.e. the closed immersion of FGC2\mathcal{M}_{FG}^{C_{2}}, equals the vanishing locus of e[2]e_{[2]} for [2]:C4[2]C4𝐓[2]\colon C_{4}\xrightarrow{[2]}C_{4}\hookrightarrow\mathbf{T}. Its complement is an open substack equivalent to FGC4/C4\mathcal{M}_{FG}^{C_{4}/C_{4}}, the non-vanishing locus of all Euler classes of non-trivial characters.

11C2C_{2}C4C_{4}FG{1}\mathcal{M}_{FG}^{\{1\}}FGC4/C2\mathcal{M}_{FG}^{C_{4}/C_{2}}11C2C_{2}C4C_{4}FGC2\mathcal{M}_{FG}^{C_{2}}FGC4/C4\mathcal{M}_{FG}^{C_{4}/C_{4}}
Figure 2: Decompositions of FGC4\mathcal{M}_{FG}^{C_{4}} into open and closed substacks, using misty rose for open and lavender for closed

For a non-cyclic group AA the situation is more complicated, and we cannot expect that the complement of the closed substack FGB\mathcal{M}_{FG}^{B} in FGA\mathcal{M}_{FG}^{A} can be expressed as a single open substack FGA/C\mathcal{M}_{FG}^{A/C} in general and vice versa. In general, the complement of FGB\mathcal{M}_{FG}^{B} in FGA\mathcal{M}_{FG}^{A} can be written as the union of the open substacks FGA/C\mathcal{M}_{FG}^{A/C} where CC runs over the minimal subgroups of AA not contained in BB. We have indicated the situation for A=C2×C2A=C_{2}\times C_{2} in Fig. 3.

C2×C2C_{2}\times C_{2}{1}×C2\{1\}\times C_{2}C2×{1}C_{2}\times\{1\}Δ\Delta{1}×{1}\{1\}\times\{1\}FGΔ\mathcal{M}_{FG}^{\Delta}FGC2×(C2/C2)FG(C2/C2)×C2\,\mathcal{M}_{FG}^{C_{2}\times(C_{2}/C_{2})}\quad\!\!\cup\quad\!\mathcal{M}_{FG}^{(C_{2}/C_{2})\times C_{2}}C2×C2C_{2}\times C_{2}{1}×C2\{1\}\times C_{2}C2×{1}C_{2}\times\{1\}Δ\Delta{1}×{1}\{1\}\times\{1\}FGC2×C2/Δ\mathcal{M}_{FG}^{C_{2}\times C_{2}/\Delta}FG{1}×C2FGC2×{1}\,\mathcal{M}_{FG}^{\{1\}\times C_{2}}\quad\!\!\cup\quad\!\mathcal{M}_{FG}^{C_{2}\times\{1\}}
Figure 3: Decompositions of FGC2×C2\mathcal{M}_{FG}^{C_{2}\times C_{2}} into open and closed substacks, Δ\Delta being the diagonal subgroup

4 Points of the moduli stack of equivariant formal groups and invariant prime ideals

The goal of this section is to classify the points of FGA\mathcal{M}_{FG}^{A} and the invariant prime ideals of (LA,SA)(L_{A},S_{A}). Although the latter could be done without the former, we feel that both questions are of the same importance and the stack point of view makes some issues more transparent.

4.1 The space associated to a stack

As mentioned above, given a graded flat Hopf algebroid (A,Γ)(A,\Gamma), we can associate an ungraded Hopf algebroid (A,Γ[u±1])(A,\Gamma[u^{\pm 1}]) to it (see e.g. [MeierOzornova, Section 4.1]).222Standard conventions force us to use AA as part of the notation of a general Hopf algebroid, while AA stands in most of this article for a compact abelian Lie group. We trust that this does not cause confusion. The category of comodules over the latter is equivalent to that of graded comodules over (A,Γ)(A,\Gamma). The stack 𝒳\mathcal{X} associated to (A,Γ)(A,\Gamma) is by definition the stack associated to (A,Γ[u±1])(A,\Gamma[u^{\pm 1}]), i.e. the fpqc-stackification of the presheaf of groupoids represented by (A,Γ[u±1])(A,\Gamma[u^{\pm 1}]) on the category of all schemes. We denote the resulting morphism SpecA𝒳\operatorname{Spec}A\to\mathcal{X} by π\pi.

Definition 4.1.

Let (A,Γ)(A,\Gamma) be a Hopf algebroid with units ηL\eta_{L} and ηR\eta_{R}. An ideal IAI\subseteq A is called invariant if ηL(I)Γ=ηR(I)Γ\eta_{L}(I)\Gamma=\eta_{R}(I)\Gamma. If (A,Γ)(A,\Gamma) is graded, we will assume that II is also graded, i.e. generated by homogeneous elements.

It is easy to check that invariant ideals in a graded Hopf algebroid (A,Γ)(A,\Gamma) correspond exactly to graded subcomodules of AA and thus to ideal sheaves on 𝒳\mathcal{X}. Here, we use that π:QCoh(𝒳)QCoh(SpecA)ModA\pi^{*}\colon\operatorname{QCoh}(\mathcal{X})\to\operatorname{QCoh}(\operatorname{Spec}A)\simeq\operatorname{Mod}_{A} refines to an equivalence from quasi-coherent 𝒪𝒳\mathcal{O}_{\mathcal{X}}-modules to graded (A,Γ)(A,\Gamma)-comodules.

Following [LMB00, Section 5] and [STACKS, Tag 04XL] in the case of Artin stacks, we can associate a topological space to 𝒳\mathcal{X}. To that purpose recall that 𝒰𝒳\mathcal{U}\to\mathcal{X} is an open immersion if the pullback 𝒰×𝒳SpecASpecA\mathcal{U}\times_{\mathcal{X}}\operatorname{Spec}A\to\operatorname{Spec}A is an open immersion.

Definition 4.2.

For 𝒳\mathcal{X} as above, define the underlying set of |𝒳||\mathcal{X}| to consist of equivalence classes of morphisms x:SpecK𝒳x\colon\operatorname{Spec}K\to\mathcal{X} for KK a field; the equivalence relation is generated by isomorphisms and x(SpecLSpecKx𝒳)x\sim(\operatorname{Spec}L\to\operatorname{Spec}K\xrightarrow{x}\mathcal{X}), where LL is a field extension of KK. We call a subset of |𝒳||\mathcal{X}| open if it is the image of |𝒰||𝒳||\mathcal{U}|\to|\mathcal{X}| for an open immersion 𝒰𝒳\mathcal{U}\to\mathcal{X}.

Equivalently, we can characterize the opens as the images of those opens in SpecA\operatorname{Spec}A that are invariant, i.e. have the same preimage along both the left and right unit SpecΓ[u±1]SpecA\operatorname{Spec}\Gamma[u^{\pm 1}]\to\operatorname{Spec}A. Indeed: by descent, an open immersion 𝒰𝒳\mathcal{U}\to\mathcal{X} corresponds to an open immersion 𝒱SpecA\mathcal{V}\to\operatorname{Spec}A with an isomorphism 𝒱×SpecASpecΓ[u±1]SpecΓ[u±1]×SpecA𝒱\mathcal{V}\times_{\operatorname{Spec}A}\operatorname{Spec}\Gamma[u^{\pm 1}]\cong\operatorname{Spec}\Gamma[u^{\pm 1}]\times_{\operatorname{Spec}A}\mathcal{V} over SpecΓ[u±1]\operatorname{Spec}\Gamma[u^{\pm 1}] satisfying a cocycle condition. But the category of open immersions into some XX with isomorphisms over XX between them is equivalent to the discrete category of open subsets of XX, yielding the required equivalence. Since invariant opens in SpecA\operatorname{Spec}A form a topology, we deduce that the opens in |𝒳||\mathcal{X}| form a topology. One further checks that the map induced by any morphism of stacks is continuous.

Our definition coincides with that of [LMB00] and [STACKS] in the intersection of their domains, e.g. when 𝒳\mathcal{X} is an affine scheme, where we get the usual topology.

Proposition 4.3.

Let (A,Γ)(A,\Gamma) be a graded Hopf algebroid with associated stack 𝒳\mathcal{X}. Then:

  1. 1.

    For every invariant prime ideal IAI\subseteq A, the image of V(I)|SpecA|V(I)\subseteq|\operatorname{Spec}A| is closed in |𝒳||\mathcal{X}| and ηI=|π|(η)\eta_{I}=|\pi|(\eta) for η\eta the generic point of V(I)V(I) (i.e. the point in |SpecA||\operatorname{Spec}A| corresponding to II) is generic (i.e. {ηI}¯=|π|(V(I))\overline{\{\eta_{I}\}}=|\pi|(V(I))).

  2. 2.

    Mapping II to ηI\eta_{I} defines an injection from the set of invariant prime ideals Specinv(A)\operatorname{Spec}^{\operatorname{inv}}(A) in AA to |𝒳||\mathcal{X}|. Equipping Specinv(A)\operatorname{Spec}^{\operatorname{inv}}(A) with the subspace topology from SpecA\operatorname{Spec}A, this map is continuous; if it is a bijection, it is a homeomorphism.

Proof.

For the first point, observe first that for every invariant ideal II, the set V(I)V(I) is invariant and hence the complement of V(I)V(I) defines an invariant open. Thus the image in |𝒳||\mathcal{X}| is open. Moreover, |π|1(|π|(V(I)))=V(I)|\pi|^{-1}(|\pi|(V(I)))=V(I). Thus |π|(V(I))|\pi|(V(I)) is the complement of |π|(|SpecA|V(I))|\pi|(|\operatorname{Spec}A|\setminus V(I)) and thus closed.

Assume now that II is an invariant prime ideal. Let ηI\eta_{I} be the image of the generic point η\eta of V(I)V(I). Then {ηI}¯|π|(V(I))=|π|{η}¯{ηI}¯\overline{\{\eta_{I}\}}\subseteq|\pi|(V(I))=|\pi|\overline{\{\eta\}}\subseteq\overline{\{\eta_{I}\}} and hence {ηI}¯=|π|(V(I))\overline{\{\eta_{I}\}}=|\pi|(V(I)).

For the injectivity of SpecinvA|𝒳|\operatorname{Spec}^{\operatorname{inv}}A\to|\mathcal{X}|, let I,JAI,J\subseteq A be two invariant prime ideals with ηI=ηJ\eta_{I}=\eta_{J}. By the first point, this implies that |π|(V(I))=|π|(V(J))|\pi|(V(I))=|\pi|(V(J)) and hence V(I)=V(J)V(I)=V(J). Thus, I=JI=J.

The continuity of Specinv(A)|𝒳|\operatorname{Spec}^{\operatorname{inv}}(A)\to|\mathcal{X}| follows from that of SpecA|𝒳|\operatorname{Spec}A\to|\mathcal{X}|. An arbitrary closed set of SpecinvA\operatorname{Spec}^{\operatorname{inv}}A is of the form V(I)Specinv(A)V(I)\cap\operatorname{Spec}^{\operatorname{inv}}(A) for some ideal IAI\subseteq A. Set I=IJJI^{\prime}=\bigcap_{I\subseteq J}J, where the JAJ\subseteq A run over all invariant ideals containing II. Then V(I)SpecinvA=V(J)SpecinvAV(I)\cap\operatorname{Spec}^{\operatorname{inv}}A=V(J)\cap\operatorname{Spec}^{\operatorname{inv}}A. If |π|:SpecinvA|𝒳||\pi|\colon\operatorname{Spec}^{\operatorname{inv}}A\to|\mathcal{X}| is a bijection, |π|1(|π|(V(J)))=V(J)|\pi|^{-1}(|\pi|(V(J)))=V(J) implies |π|(V(J)SpecinvA))=|π|(V(J))|\pi|(V(J)\cap\operatorname{Spec}^{\operatorname{inv}}A))=|\pi|(V(J)) and this is closed by the first part. ∎

We warn the reader that in general, the preimage of an irreducible closed subset of |𝒳||\mathcal{X}| won’t be irreducible in |SpecA||\operatorname{Spec}A| and thus does not correspond to an invariant prime ideal.

Proposition 4.4.

Let AA be a compact abelian Lie group. The underlying set of |FGA||\mathcal{M}_{FG}^{A}| is in bijection with the product of the set Sub(A)\operatorname{Sub}(A) of closed subgroups of AA and |FG||\mathcal{M}_{FG}|. For a given AA-equivariant formal group GG over a field, the point in |FG||\mathcal{M}_{FG}| is the pushforward of GG along p:A{e}p\colon A\to\{e\}, and the subgroup is Pontryagin dual to A/{VA:eV=0}A^{*}/\{V\in A^{*}:e_{V}=0\}.

Proof.

Given a closed subgroup BB of AA and a non-equivariant formal group Γ\Gamma over a field kk, we obtain an AA-equivariant formal group via qiΓq^{*}i_{*}\Gamma, where AqA/Bi{1}A\xrightarrow{q}A/B\xleftarrow{i}\{1\}. By Proposition 2.24, every AA-equivariant formal group GG over a field kk is isomorphic to one of this form. Here, the projection AA/BA\to A/B is Pontryagin dual to the subgroup {VA:eV=0}\{V\in A^{*}:e_{V}=0\} and thus BB is uniquely defined by GG. Moreover, necessarily ΓpG\Gamma\cong p_{*}G. This implies that two AA-equivariant formal groups G1G_{1} and G2G_{2} defined over the same field kk are isomorphic if and only if they have the same associated closed subgroup and their completions pG1p_{*}G_{1} and pG2p_{*}G_{2} are isomorphic as non-equivariant formal groups, proving the claim. ∎

We recall that |FG||\mathcal{M}_{FG}| has been computed by Honda: its points are classified by a pair (p,n)(p,n), where p0p\geq 0 is the characteristic of the field and n¯={}n\in\overline{\mathbb{N}}=\mathbb{N}\cup\{\infty\} is the height of the formal group, with n=0n=0 if p=0p=0. We will below always identify |FG||\mathcal{M}_{FG}| with such pairs.

Remark 4.5.

By 3.11, there is an open immersion FGA/BFGA\mathcal{M}_{FG}^{A/B}\to\mathcal{M}_{FG}^{A} for any closed subgroup BAB\subseteq A and hence |FGA/B||\mathcal{M}_{FG}^{A/B}| is homeomorphic to an open subset of |FGA||\mathcal{M}_{FG}^{A}|. Given CA/BC\subseteq A/B, the point corresponding to (C,p,n)(C,p,n) is (q1(C),p,n)(q^{-1}(C),p,n) for q:AA/Bq\colon A\to A/B the projection. Indeed, for a A/BA/B-equivariant formal group GG over a field corresponding to (C,p,n)(C,p,n), we have

{VA:eV=0}=q({V(A/B):eV=0})=q(((A/B)/C)).\{V\in A^{*}:e_{V}=0\}=q^{*}(\{V\in(A/B)^{*}:e_{V}=0\})=q^{*}(((A/B)/C)^{*}).

The Pontryagin dual of A/q(((A/B)/C))A^{*}/q^{*}(((A/B)/C)^{*}) is precisely q1(C)q^{-1}(C) since A/q1(C)(A/B)/CA/q^{-1}(C)\cong(A/B)/C.

4.2 Invariant prime ideals of (LA,SA)(L_{A},S_{A})

As before, let ΦBL=LB[eV1]\Phi^{B}L=L_{B}[e_{V}^{-1}] be the localization of LBL_{B} away from all Euler classes eVe_{V} for VϵV\neq\epsilon and let ΦBS=SBLBΦBL\Phi^{B}S=S_{B}\otimes_{L_{B}}\Phi^{B}L. Denote by Ip,nI_{p,n} the ideal (p,v1,v2,,vn1)L(p,v_{1},v_{2},\dots,v_{n-1})\subseteq L, i.e. the unique invariant prime ideal at height nn containing pp. We also include the case of Ip,0=0I_{p,0}=0. In this section, we will denote this ideal by I0,0I_{0,0} to uniformize notation with respect to the residue characteristic.

Construction 4.6.

For every triple (B,p,n)(B,p,n) of a closed subgroup BB of AA, a prime pp and n¯n\in\overline{\mathbb{N}} we define an invariant ideal IAB,p,nLAI^{A}_{B,p,n}\subseteq L_{A} as the preimage of ΦBLIp,nΦBL\Phi^{B}L\cdot I_{p,n}\subseteq\Phi^{B}L along the composite map of Hopf algebroids

(LA,SA)(LB,SB)(ΦBL,ΦBS).(L_{A},S_{A})\to(L_{B},S_{B})\to(\Phi^{B}L,\Phi^{B}S).

Note that the ΦBLIp,n\Phi^{B}L\cdot I_{p,n} are indeed prime ideals since by 2.25 the quotient ring ΦBL/ΦBLIp,n\Phi^{B}L/\Phi^{B}L\cdot I_{p,n} is of the form (L/Ip,n)[(b0V)±1,biV|i>0,VA{ϵ}](L/I_{p,n})[(b_{0}^{V})^{\pm 1},b_{i}^{V}\ |\ i>0,V\in A^{*}-\{\epsilon\}] and hence an integral domain. Thus, the IAB,p,nI^{A}_{B,p,n} are prime as well.

To simplify notation, we will from now on often write ΦBL/Ip,n\Phi^{B}L/I_{p,n} instead of ΦBL/ΦBLIp,n\Phi^{B}L/\Phi^{B}L\cdot I_{p,n} and likewise in similar situations.

Theorem 4.7.

The assignment

Sub(A)×|FG|\displaystyle\operatorname{Sub}(A)\times|\mathcal{M}_{FG}| \displaystyle\to Specinv(LA)\displaystyle\operatorname{Spec}^{\operatorname{inv}}(L_{A})
(B,p,n)\displaystyle(B,p,n) \displaystyle\mapsto IAB,p,n\displaystyle I^{A}_{B,p,n}

is a bijection. In other words, the ideals IAB,p,nI^{A}_{B,p,n} are pairwise different and constitute all the invariant prime ideals in LAL_{A}.

The map Specinv(LA)|FGA|\operatorname{Spec}^{\operatorname{inv}}(L_{A})\to|\mathcal{M}_{FG}^{A}| from 4.3 is a homeomorphism.

Proof.

By 4.3, we know that Specinv(LA)\operatorname{Spec}^{\operatorname{inv}}(L_{A}) injects into |FGA||\mathcal{M}_{FG}^{A}| and the latter we computed to be Sub(A)×|FG|\operatorname{Sub}(A)\times|\mathcal{M}_{FG}| as a set. Thus, it suffices to show that the element of Sub(A)×|FG|\operatorname{Sub}(A)\times|\mathcal{M}_{FG}| associated to IAB,p,nSpecinv(LA)I^{A}_{B,p,n}\in\operatorname{Spec}^{\operatorname{inv}}(L_{A}) is precisely (B,(p,n))(B,(p,n)), where pp is a prime number if n>0n>0 and 0 if n=0n=0.

To spell this out concretely, let kk be the field of fractions of LA/IAB,p,nL_{A}/I^{A}_{B,p,n}. We denote by FF the pushed forward AA-equivariant formal group law over kk and by GG the corresponding AA-equivariant formal group. Since SpeckSpecLA\operatorname{Spec}k\to\operatorname{Spec}L_{A} hits the point corresponding to the prime ideal IAB,p,,nI^{A}_{B,p,,n}, the corresponding point in |FGA||\mathcal{M}_{FG}^{A}| is represented by SpeckGFGA\operatorname{Spec}k\xrightarrow{G}\mathcal{M}_{FG}^{A}. By the classification in Proposition 4.4, we need to show three things:

  1. 1.

    the set of VAV\in A^{*} such that eV=0e_{V}=0 in kk is precisely ker(AB)\ker(A^{*}\to B^{*}),

  2. 2.

    kk has characteristic pp (which is clear), and

  3. 3.

    the pushforward pGp_{*}G along Ap{1}A\xrightarrow{p}\{1\} has height nn.

For the first, recall from Lemma 2.21 that IAB=ker(res:LALB)I^{A}_{B}=\ker(\operatorname{res}\colon L_{A}\to L_{B}) is generated by the Euler classes eVe_{V} for all Vker(AB)V\in\ker(A^{*}\to B^{*}). These Euler classes must vanish in kk since LAkL_{A}\to k factors through LA/IABL_{A}/I_{A}^{B}. If VV is not in ker(AB)\ker(A^{*}\to B^{*}), then eV0e_{V}\neq 0 in LBL_{B} and hence also in ΦBL/Ip,n\Phi^{B}L/I_{p,n} (as else ΦBL/Ip,n=0\Phi^{B}L/I_{p,n}=0). Since LA/IAB,p,,nL_{A}/I^{A}_{B,p,,n} injects into ΦBL/Ip,n\Phi^{B}L/I_{p,n}, the Euler class eVe_{V} is actually nonzero in LA/IAB,p,,nL_{A}/I^{A}_{B,p,,n} and hence invertible in kk. This shows the first point.

The pushforward pGp_{*}G is classified by the composite

g:LLALA/IAB,p,nk.g\colon L\to L_{A}\to L_{A}/I^{A}_{B,p,n}\to k.

The ideal Ip,nLAI_{p,n}\cdot L_{A} maps to 0 in ΦBL/Ip,n\Phi^{B}L/I_{p,n} and is hence contained in IAB,p,nI^{A}_{B,p,n}. Therefore, gg factors through L/Ip,nL/I_{p,n} and the height of pGp_{*}G is at least nn. It remains to show that vnv_{n} is non-zero in LA/IAB,p,nL_{A}/I^{A}_{B,p,n} and hence in kk. By definition of IAB,p,nI^{A}_{B,p,n}, the map LAΦBL/Ip,nL_{A}\to\Phi^{B}L/I_{p,n} factors over LA/IAB,p,nL_{A}/I^{A}_{B,p,n}. We know by 2.25 that ΦBL/Ip,n\Phi^{B}L/I_{p,n} is an integral domain of the form (L/Ip,n)[(b0V)±1,biV](L/I_{p,n})[(b_{0}^{V})^{\pm 1},b_{i}^{V}]. In particular, vnv_{n} is non-trivial in ΦBL/Ip,n\Phi^{B}L/I_{p,n} and hence in LA/IAB,p,nL_{A}/I^{A}_{B,p,n}. This shows that pGp_{*}G is of height nn as desired, which finishes the proof. ∎

Unraveling the definition, we obtain the following description of the ideal IAB,p,nI^{A}_{B,p,n}, where IABI^{A}_{B} still denotes the kernel of the restriction map LBLAL_{B}\to L_{A}.

Lemma 4.8.

An element xLAx\in L_{A} lies in IAB,p,nI^{A}_{B,p,n} if and only if there exists an AA-representation WW with WB=0W^{B}=0, such that

xeW(IAB,Ip,n).x\cdot e_{W}\in(I^{A}_{B},I_{p,n}).
Proof.

By definition, IAB,p,nI^{A}_{B,p,n} is the kernel of the composition

LAresABLBLB/LBIp,nΦBL/ΦBLIp,n.L_{A}\xrightarrow{\operatorname{res}^{A}_{B}}L_{B}\to L_{B}/L_{B}\cdot I_{p,n}\to\Phi^{B}L/\Phi^{B}L\cdot I_{p,n}.

The composition LALBLB/Ip,nL_{A}\to L_{B}\to L_{B}/I_{p,n} is surjective with kernel (IAB,Ip,n)(I^{A}_{B},I_{p,n}), and the map LB/Ip,nΦBL/Ip,nL_{B}/I_{p,n}\to\Phi^{B}L/I_{p,n} inverts all Euler classes eW¯e_{\overline{W}} for BB-representations W¯\overline{W} with W¯B=0\overline{W}^{B}=0. The product of Euler classes is an Euler class again. Hence, if xLAx\in L_{A} is contained in IAB,p,nI^{A}_{B,p,n}, its image x¯\overline{x} in LB/Ip,nL_{B}/I_{p,n} must be annihilated by such an Euler class eW¯e_{\overline{W}}. We can extend W¯\overline{W} to an AA-representation WW and find that xeWx\cdot e_{W} is contained in (IAB,Ip,n)(I^{A}_{B},I_{p,n}), as desired.

For the opposite direction, if we assume that xeW(IAB,Ip,n)x\cdot e_{W}\in(I^{A}_{B},I_{p,n}) for some AA-representation WW with WB=0W^{B}=0, then x¯eresABW=0\overline{x}\cdot e_{\operatorname{res}^{A}_{B}W}=0 in LB/Ip,nL_{B}/I_{p,n}. Since resABW\operatorname{res}^{A}_{B}W has trivial BB-fixed points, eresABWe_{\operatorname{res}^{A}_{B}W} becomes invertible in ΦBL/Ip,n\Phi^{B}L/I_{p,n} and hence x¯\overline{x} is taken to 0 there. Therefore, xx is contained in IAB,p,nI^{A}_{B,p,n}. ∎

When BB is a torus, the ideal IAB,p,nI^{A}_{B,p,n} is easy to describe explicitly:

Corollary 4.9.

If BB is a torus, then IAB,p,n=(IAB,Ip,n)I^{A}_{B,p,n}=(I^{A}_{B},I_{p,n}).

Proof.

When BB is a torus, every non-trivial character is non-torsion and thus the map

LB/Ip,nΦBL/Ip,nL_{B}/I_{p,n}\to\Phi^{B}L/I_{p,n}

is injective (Corollary 2.31). Hence, IB,p,nAI_{B,p,n}^{A} is equal to the kernel of

LALA/Ip,nLB/Ip,n,L_{A}\to L_{A}/I_{p,n}\to L_{B}/I_{p,n},

which is generated by IABI^{A}_{B} and Ip,nI_{p,n}. ∎

We note the following useful corollary:

Corollary 4.10.

Let AA be a torus, nn\in\mathbb{N} and consider the augmentation ideal

I=ker(LA/Ip,nL/Ip,n),I=\ker(L_{A}/I_{p,n}\to L/I_{p,n}),

i.e., the ideal generated by all the Euler classes. Then the intersection J=kIkJ=\cap_{k\in\mathbb{N}}I^{k} equals the 0-ideal.

For example, this shows that no element in the 𝐓\mathbf{T}-equivariant Lazard ring L𝐓L_{\mathbf{T}} is infinitely often divisible by the Euler class ee.

Proof.

Since II is an invariant ideal of LA/Ip,nL_{A}/I_{p,n}, so are all its powers and the intersection thereof. Moreover, JJ equals the kernel of the completion map (cf. Section 2.5)

LA/Ip,nL/Ip,ny1,,yr,L_{A}/I_{p,n}\to L/I_{p,n}\llbracket y_{1},\ldots,y_{r}\rrbracket,

where rr is the rank of AA and the yiy_{i} are the images of the Euler classes ranging through a basis of AA^{*}. Since L/Ip,ny1,,yrL/I_{p,n}\llbracket y_{1},\dots,y_{r}\rrbracket is an integral domain, JJ must be prime and hence an invariant prime ideal.

Therefore JJ must be of the form IAB,p,mI^{A}_{B,p,m} (or rather its image under the projection LALA/Ip,nL_{A}\to L_{A}/I_{p,n}) for some subgroup BB and height mnm\geq n. Note that vnv_{n} is not contained in II, hence in particular not in JJ. This means that we must have m=nm=n. Moreover, given a character V=V1k1VrkrAV=V_{1}^{\otimes k_{1}}\otimes\dots\otimes V_{r}^{\otimes k_{r}}\in A^{*} expressed in the chosen basis V1,,VrV_{1},\ldots,V_{r} above, the image of eVe_{V} under the completion map is given by

F([k1]F(y1),,[kr]F(yr))L/Ip,ny1,,yr,F([k_{1}]_{F}(y_{1}),\dots,[k_{r}]_{F}(y_{r}))\in L/I_{p,n}\llbracket y_{1},\ldots,y_{r}\rrbracket,

where FF is the universal formal group law pushed forward to L/Ip,nL/I_{p,n}. Since we assumed nn to be a finite height, the [k][k]-series of FF is non-trivial whenever kk is non-zero. It follows that for any non-trivial VV the image of eVe_{V} is non-trivial in L/Ip,ny1,,yrL/I_{p,n}\llbracket y_{1},\ldots,y_{r}\rrbracket. Hence JJ contains no Euler class eVe_{V} and we must have B=AB=A, i.e.,

J=IAA,p,n=(0)LA/Ip,n,J=I^{A}_{A,p,n}=(0)\subseteq L_{A}/I_{p,n},

as desired. ∎

5 Inclusions between invariant prime ideals

Non-equivariantly the ideals Ip,nLI_{p,n}\subseteq L form ascending towers

(0)=Ip,0Ip,1Ip,2,(0)=I_{p,0}\subseteq I_{p,1}\subseteq I_{p,2}\subseteq\ldots,

essentially by definition. Except for the overlap at (0)(0), there are no inclusions between the towers for different primes pp. For invariant prime ideals in the equivariant Lazard ring LAL_{A} we saw that we have one tower

IAB,p,0IAB,p,1IAB,p,2I^{A}_{B,p,0}\subseteq I^{A}_{B,p,1}\subseteq I^{A}_{B,p,2}\subseteq\ldots

for every pair of a closed subgroup BB and prime pp. Again, there will be no interplay between the towers associated to different primes (except for the overlap at height 0). However, there are additional inclusions connecting the towers for different subgroups B,BB,B^{\prime} at the same prime pp. This relationship between the heights at different subgroups is one of the essential properties of equivariant formal groups. It is closely related to the blue-shift phenomenon in stable homotopy theory. We say more about this in Section 8 below.

To see that there is no inclusion between towers associated with different primes we note that p=v0IAB,p,np=v_{0}\in I^{A}_{B,p,n} whenever n1n\geq 1. It is easy to see that pp maps non-trivially under LAΦBL/Iq,nL_{A}\to\Phi^{B^{\prime}}L/I_{q,n^{\prime}} whenever qpq\neq p (since the target is free over L/Iq,nL/I_{q,n} by 2.25). Hence there cannot be an inclusion IAB,p,nIAB,q,nI^{A}_{B,p,n}\subseteq I^{A}_{B^{\prime},q,n^{\prime}} for n1n\geq 1 and pqp\neq q. Moreover, if n=0n=0 we have IAB,p,0=IAB,q,0I^{A}_{B,p,0}=I^{A}_{B,q,0}. Hence we can reduce to studying containments between invariant prime ideals associated to the same prime pp.

New convention: For this reason and to simplify notation we from now on and for the rest of the paper implicitly localize at a fixed prime pp. That is, we consider the pp-localized Lazard ring LAL_{A} and denote its invariant prime ideals simply by IAB,nI^{A}_{B,n}, omitting the chosen prime pp. We further sometimes abbreviate IAA,nI^{A}_{A,n} to IA,nI_{A,n}.

Hence our goal is to understand for which pairs of subgroups B,BB,B^{\prime} and natural numbers n,nn,n^{\prime} there is an inclusion

IAB,nIAB,n.I^{A}_{B,n}\subseteq I^{A}_{B^{\prime},n^{\prime}}.

We will show the following:

Theorem 5.1.

There is an inclusion IAB,nIAB,nI^{A}_{B,n}\subseteq I^{A}_{B^{\prime},n^{\prime}} if and only if the following conditions are satisfied:

  1. 1.

    BB^{\prime} is a subgroup of BB and π0(B/B)\pi_{0}(B/B^{\prime}) is a pp-group.

  2. 2.

    We have nn+rankp(π0(B/B))n^{\prime}\geq n+\operatorname{rank}_{p}(\pi_{0}(B/B^{\prime})).

Hence, for example there are inclusions I𝐓𝕋,nI𝐓1,nI^{\mathbf{T}}_{\mathbb{T},n}\subseteq I^{\mathbf{T}}_{1,n} and ICpkCpk,nICpk1,n+1I^{C_{p^{k}}}_{C_{p^{k}},n}\subseteq I^{C_{p^{k}}}_{1,n+1}, but ICpkCpk,nI^{C_{p}^{k}}_{C_{p}^{k},n} is not contained in ICpk1,n+k1I^{C_{p}^{k}}_{1,n+k-1}. In fact the theorem can be formally reduced to checking those three special cases, as we will see below. Note also that the theorem in particular says that given a chain of inclusions BBAB^{\prime}\subseteq B\subseteq A, the question whether IAB,nI^{A}_{B,n} is contained in IAB,nI^{A}_{B^{\prime},n^{\prime}} does not depend on the ambient group AA, but only on B,B,nB,B^{\prime},n and nn^{\prime}.

Theorem 5.1 can be interpreted as a statement about the heights of geometric fixed points of localizations of LAL_{A}, in the following way. Recall from [HoveyStricklandComodules, Definition 4.1] that the height ht(R)ht(R) of an LL-algebra RR is the maximal nn such that R/InR0R/I_{n}\cdot R\neq 0; equivalently, it is the minimal number nn such that In+1R=RI_{n+1}\cdot R=R. If there is no such nn, the height is understood to be infinite. If R=0R=0, the height is 1-1. Then we have the following:

Corollary 5.2.

Let BAB\subseteq A be a closed subgroup, nn\in\mathbb{N}. If π0(A/B)\pi_{0}(A/B) is not a pp-group, or if π0(A/B)\pi_{0}(A/B) is a pp-group but rankp(π0(A/B))>n\operatorname{rank}_{p}(\pi_{0}(A/B))>n, then the geometric fixed points ΦA((LA)IAB,n)\Phi^{A}((L_{A})_{I^{A}_{B,n}}) are trivial. Otherwise, their height is given by

ht(ΦA((LA)IAB,n))=nrankp(π0(A/B)).ht(\Phi^{A}((L_{A})_{I^{A}_{B,n}}))=n-\operatorname{rank}_{p}(\pi_{0}(A/B)).
Proof.

We have ImΦA((LA)IAB,n)=ΦA((LA)IAB,n)I_{m}\cdot\Phi^{A}((L_{A})_{I^{A}_{B,n}})=\Phi^{A}((L_{A})_{I^{A}_{B,n}}) if and only if there exists an element of LAL_{A} not contained in IAB,nI^{A}_{B,n} which is mapped to ImΦALAI_{m}\cdot\Phi^{A}L_{A} under the geometric fixed point map. Since IA,mI_{A,m} is defined precisely as the preimage of ImΦALAI_{m}\cdot\Phi^{A}L_{A}, this in turn is equivalent to IAA,mI^{A}_{A,m} not being contained in IAB,nI^{A}_{B,n}.

By Theorem 5.1 we know that if π0(A/B)\pi_{0}(A/B) is not a pp-group or if π0(A/B)\pi_{0}(A/B) is a pp-group but rankp(π0(A/B))>n\operatorname{rank}_{p}(\pi_{0}(A/B))>n, then IAA,0I^{A}_{A,0} is not contained in IAB,nI^{A}_{B,n}. Since I0=(0)I_{0}=(0), this implies that the geometric fixed points ΦA((LA)IAB,n)=I0ΦA((LA)IAB,n)\Phi^{A}((L_{A})_{I^{A}_{B,n}})=I_{0}\cdot\Phi^{A}((L_{A})_{I^{A}_{B,n}}) are trivial.

If π0(A/B)\pi_{0}(A/B) is a pp-group and r=rankp(π0(A/B))nr=\operatorname{rank}_{p}(\pi_{0}(A/B))\leq n, then the theorem tells us that IAA,nrIAB,nI^{A}_{A,n-r}\subseteq I^{A}_{B,n} and IAA,nr+1IAB,nI^{A}_{A,n-r+1}\not\subseteq I^{A}_{B,n}. Hence we have InrΦA((LA)IAB,n)ΦA((LA)IAB,n)I_{n-r}\cdot\Phi^{A}((L_{A})_{I^{A}_{B,n}})\neq\Phi^{A}((L_{A})_{I^{A}_{B,n}}) and IAnr+1ΦA((LA)IAB,n)=ΦA((LA)IAB,n)I^{A}_{n-r+1}\cdot\Phi^{A}((L_{A})_{I^{A}_{B,n}})=\Phi^{A}((L_{A})_{I^{A}_{B,n}}), as claimed. ∎

Remark 5.3.

The techniques of this paper can be used to compute the height of geometric fixed points for many complex oriented theories. We give one example of this in Proposition 8.6 below.

Remark 5.4.

There is an inclusion IAB,nIAB,nI^{A}_{B,n}\subseteq I^{A}_{B^{\prime},n^{\prime}} if and only if in |FG,(p)A||\mathcal{M}_{FG,(p)}^{A}|, the point corresponding to IAB,nI^{A}_{B^{\prime},n^{\prime}} lies in the closure of the point corresponding to IAB,nI^{A}_{B,n}. Thus, Theorem 5.1 can be interpreted as a result about the topology of |FG,(p)A||\mathcal{M}_{FG,(p)}^{A}|.

The proof of Theorem 5.1 takes up the remainder of this section.

5.1 Formal reduction to the pp-toral case

Our first step is the following:

Lemma 5.5.

If there is an inclusion IAB,nIAB,nI^{A}_{B,n}\subseteq I^{A}_{B^{\prime},n^{\prime}}, then BB^{\prime} is a subgroup of BB and nnn^{\prime}\geq n.

Proof.

When BB^{\prime} is not a subgroup of BB we can choose a character VAV\in A^{*} which is trivial when restricted to BB but non-trivial when restricted to BB^{\prime}. Its Euler class eVe_{V} then restricts to 0 in LBL_{B}, in particular it is contained in IAB,nI^{A}_{B,n} for all nn. On the other hand, its restriction to BB^{\prime} becomes an invertible element in the non-trivial ring ΦBL/In\Phi^{B^{\prime}}L/I_{n^{\prime}} and is hence non-trivial. Therefore eVe_{V} is not an element of IB,nAI_{B^{\prime},n^{\prime}}^{A}. It follows that IAB,nI^{A}_{B,n} cannot be contained in IAB,nI^{A}_{B^{\prime},n^{\prime}}.

If n<nn^{\prime}<n, then vnv_{n^{\prime}} is contained in InI_{n} but not in InI_{n^{\prime}}. This implies that vnv_{n^{\prime}} (now thought of as an element of LAL_{A}) is contained in IAB,nI^{A}_{B,n} but not in IAB,nI^{A}_{B^{\prime},n^{\prime}}, since ΦBL/In\Phi^{B^{\prime}}L/I_{n^{\prime}} is a non-trivial free module over L/InL/I_{n^{\prime}} by 2.25 Hence, again, IAB,nI^{A}_{B,n} cannot be contained in IAB,nI^{A}_{B^{\prime},n^{\prime}}. ∎

Lemma 5.6.

Let BBB^{\prime}\subseteq B be an inclusion of subgroups of AA and n,nn,n^{\prime}\in\mathbb{N}. Then there is an inclusion

IB,nAIB,nAI_{B,n}^{A}\subseteq I_{B^{\prime},n^{\prime}}^{A}

if and only if there is an inclusion

IB,nBIB,nBI_{B,n}^{B}\subseteq I_{B^{\prime},n^{\prime}}^{B}

if and only if there is an inclusion

IB/B,nB/BIB/B,nB/B.I_{B/B^{\prime},n}^{B/B^{\prime}}\subseteq I_{B^{\prime}/B^{\prime},n^{\prime}}^{B/B^{\prime}}.
Proof.

The first two statements are equivalent since the restriction map resAB:LALB\operatorname{res}^{A}_{B}\colon L_{A}\to L_{B} identifies LBL_{B} with a quotient of LAL_{A}, and the ideals IAB,nI^{A}_{B,n} and IAB,nI^{A}_{B^{\prime},n^{\prime}} are the preimages of the ideals IBB,nI^{B}_{B,n} and IBB,nI^{B}_{B^{\prime},n^{\prime}} under the quotient projection.

Phrased differently, the projection LALBL_{A}\to L_{B} induces a closed embedding of the stack of BB-equivariant formal groups into the stack of AA-equivariant formal groups. On spectra, the image consists precisely of these IAB,nI^{A}_{B^{\prime\prime},n} with BBB^{\prime\prime}\subseteq B. This implies the desired equivalence by Remark 5.4.

For the second equivalence, recall the open embedding |FGB/B||FGB||\mathcal{M}_{FG}^{B/B^{\prime}}|\subseteq|\mathcal{M}_{FG}^{B}| from Remark 4.5, sending the point (B/B,n)(B^{\prime\prime}/B^{\prime},n) to (B,n)(B^{\prime\prime},n) for every BBBB^{\prime}\subseteq B^{\prime\prime}\subseteq B. Since the closure relation among points in a subspace can be detected in the subspace, Remark 5.4 gives the result. ∎

Taken together, the previous two lemmas allow us to reduce to the case B=AB=A and B=1B^{\prime}=1 and understand under what conditions there is an inclusion

IAA,nIA1,n,I^{A}_{A,n}\subseteq I^{A}_{1,n^{\prime}},

or in other words whether the restriction map LALL_{A}\to L maps IAA,nI^{A}_{A,n} into the ideal InI_{n^{\prime}}.

Our next goal is to show that we can further reduce to the case where π0A\pi_{0}A is a pp-group. For this we choose a prime qq and consider the Euler class eτqL𝐓e_{\tau^{q}}\in L_{\mathbf{T}}, i.e., the pullback of eτL𝐓e_{\tau}\in L_{\mathbf{T}} along the qqth power map [q]:𝐓𝐓[q]\colon\mathbf{T}\to\mathbf{T}. The Euler class eτqe_{\tau^{q}} restricts to 0 at the trivial group and is hence uniquely divisible by eτe_{\tau}. We set x~0(q)L𝐓\widetilde{x}_{0}^{(q)}\in L_{\mathbf{T}} to be the unique element satisfying eτq=x~0(q)eτe_{\tau^{q}}=\widetilde{x}_{0}^{(q)}\cdot e_{\tau} (the reason for this choice of notation will become clear in Section 5.2). Under the completion map L𝐓LeτL_{\mathbf{T}}\to L\llbracket e_{\tau}\rrbracket, the Euler class eτqe_{\tau^{q}} is sent to the qq-series [q]F(eτ)[q]_{F}(e_{\tau}) of the universal formal group law. Hence, x~0(q)\widetilde{x}_{0}^{(q)} is sent to the quotient [q]F(eτ)/eτ[q]_{F}(e_{\tau})/e_{\tau}, whose leading coefficient equals qq. Since the restriction of any element in L𝐓L_{\mathbf{T}} to the trivial group equals the leading coefficient of its image in LeτL\llbracket e_{\tau}\rrbracket, we see that res𝐓1x~0(q)=qL.\operatorname{res}^{\mathbf{T}}_{1}\widetilde{x}_{0}^{(q)}=q\in L. We further set x0(q)LCqx_{0}^{(q)}\in L_{C_{q}} to be the restriction of x~0(q)\widetilde{x}_{0}^{(q)}, and find that it satisfies:

x0(q)eτ¯\displaystyle x_{0}^{(q)}\cdot e_{\overline{\tau}} =0\displaystyle=0
resCq1x0(q)\displaystyle\operatorname{res}^{C_{q}}_{1}x_{0}^{(q)} =q\displaystyle=q

Here, τ¯\overline{\tau} denotes the restriction of the tautological character τ𝐓\tau\in\mathbf{T}^{*} to CpC_{p}. We are now ready to show:

Lemma 5.7.

If π0A\pi_{0}A is not a pp-group, then there is no inclusion of the form IAA,nIA1,nI^{A}_{A,n}\subseteq I^{A}_{1,n^{\prime}}.

Proof.

If π0A\pi_{0}A is not a pp-group we can choose a surjection f:ACqf\colon A\to C_{q} with qpq\neq p a prime. Then the element fx0(q)LAf^{*}x_{0}^{(q)}\in L_{A} satisfies the equation fx0(q)efτ¯=0f^{*}x_{0}^{(q)}\cdot e_{f^{*}\overline{\tau}}=0. Since ff is surjective, the character fτ¯Af^{*}\overline{\tau}\in A^{*} is non-trivial. Hence, fx0(q)f^{*}x_{0}^{(q)} is an element of IAA,0I^{A}_{A,0} and hence also of IAA,nI^{A}_{A,n}. Its restriction to the trivial group equals that of x0(q)x_{0}^{(q)}, which is qq and hence a unit in the (pp-localized) ring L/InL/I_{n^{\prime}}. In other words, fx0(q)f^{*}x_{0}^{(q)} is not an element of IA1,nI^{A}_{1,n^{\prime}}. Hence IAA,0I^{A}_{A,0} does not include into IA1,nI^{A}_{1,n^{\prime}}. ∎

Combined with Lemmas 5.5 and 5.6 we obtain:

Corollary 5.8.

If there is an inclusion IAB,nIAB,nI^{A}_{B,n}\subseteq I^{A}_{B^{\prime},n^{\prime}}, then BB^{\prime} is a subgroup of BB, nnn^{\prime}\geq n and the quotient B/BB/B^{\prime} is pp-toral, i.e. a product of a pp-group and a torus.

5.2 Proof of inclusions

The next goal is to prove the ‘if’ part of Theorem 5.1, i.e., to show that if the conditions on B,B,nB,B^{\prime},n and nn^{\prime} stated there are satisfied we do have an inclusion IAB,nIAB,nI^{A}_{B,n}\subseteq I^{A}_{B^{\prime},n^{\prime}}. We start with the easiest case:

Lemma 5.9.

Let BBB^{\prime}\subseteq B be an inclusion of subgroups of AA such that B/BB/B^{\prime} is a torus. Then there is an inclusion IAB,nIAB,nI^{A}_{B,n}\subseteq I^{A}_{B^{\prime},n} for all nn.

Proof.

By Lemma 5.6 we can reduce to B=AB=A and B=1B^{\prime}=1 the trivial group. Hence AA is a torus. Lemma 4.9 then implies that IA1,nI^{A}_{1,n} is equal to the ideal generated by InI_{n} and the augmentation ideal IA1I^{A}_{1}, whereas IAA,nI^{A}_{A,n} is generated by InI_{n} only. Clearly the latter is contained in the former. ∎

We now turn to showing that there are inclusions ICpkCpk,nICpk1,n+1I^{C_{p^{k}}}_{C_{p^{k}},n}\subseteq I^{C_{p^{k}}}_{1,n+1}. For this we show that ICpkCpk,nI^{C_{p^{k}}}_{C_{p^{k}},n} is generated by InI_{n} plus one additional element which reduces to vnv_{n} under the restriction map LCpk/InL/InL_{C_{p^{k}}}/I_{n}\to L/I_{n}. We start with the case k=1k=1 and recall again that the Euler class eτpe_{\tau^{p}} is sent to the pp-series [p]F(eτ)[p]_{F}(e_{\tau}) under the completion map L𝐓LeτL_{\mathbf{T}}\to L\llbracket e_{\tau}\rrbracket. Modulo InI_{n}, this pp-series is of the form vneτpnv_{n}e_{\tau}^{p^{n}} + higher order terms. Hence 2.28 implies that there exists a unique element ψp(n)L𝐓/In\psi_{p}^{(n)}\in L_{\mathbf{T}}/I_{n} such that eτp=ψp(n)eτpne_{\tau^{p}}=\psi_{p}^{(n)}e_{\tau}^{p^{n}}, and this element satisfies res𝐓1(ψp(n))=vn\operatorname{res}^{\mathbf{T}}_{1}(\psi_{p}^{(n)})=v_{n}.

We then set ψpk(n)=[pk1](ψp(n))L𝐓/In\psi_{p^{k}}^{(n)}=[p^{k-1}]^{*}(\psi_{p}^{(n)})\in L_{\mathbf{T}}/I_{n} for all k2k\geq 2, where [pk1][p^{k-1}] is the multiplication-by-pk1p^{k-1} map on the circle. By functoriality, ψpk(n)\psi_{p^{k}}^{(n)} also restricts to vnv_{n} at the trivial group. In fact, it already restricts to vnv_{n} at LCpk1/InL_{C_{p^{k-1}}}/I_{n}. This is because Cpk1C_{p^{k-1}} is the kernel of [pk1][p^{k-1}] and hence the restriction map factors through the trivial group. Applying [pk1][p^{k-1}]^{*} to the defining equation for ψp(n)\psi_{p}^{(n)}, we also obtain epk=ψpk(n)epk1pne_{p^{k}}=\psi_{p^{k}}^{(n)}e_{p^{k-1}}^{p^{n}}; here and in the following, we will often abbreviate eτme_{\tau^{m}} to eme_{m}.

Proposition 5.10.

For every k1k\geq 1 and nn\in\mathbb{N} the element ψpk(n)\psi_{p^{k}}^{(n)} generates the kernel of

ϕ𝐓Cpk:L𝐓/InLCpk/InΦCpkL/In.\phi^{\mathbf{T}}_{C_{p^{k}}}\colon L_{\mathbf{T}}/I_{n}\to L_{C_{p^{k}}}/I_{n}\to\Phi^{C_{p^{k}}}L/I_{n}.

Hence, I𝐓Cpk,nI^{\mathbf{T}}_{C_{p^{k}},n} is generated by the regular sequence v0,v1,,vn1,ψpk(n)v_{0},v_{1},\ldots,v_{n-1},\psi_{p^{k}}^{(n)}.

Corollary 5.11.

The ideal ICpkCpk,nI^{C_{p^{k}}}_{C_{p^{k}},n} is generated by InI_{n} and the restriction ψ¯(n)pk\overline{\psi}^{(n)}_{p^{k}} of ψpk(n)\psi_{p^{k}}^{(n)} to LCpk/InL_{C_{p^{k}}}/I_{n}.

Proof of Proposition 5.10.

Let xL𝐓/Inx\in L_{\mathbf{T}}/I_{n} be an element mapping to 0 in ΦCpkL/In\Phi^{C_{p^{k}}}L/I_{n}, i.e. in the image of ICpk,n𝐓I_{C_{p^{k}},n}^{\mathbf{T}}. By Lemma 4.8 we know that we have an equation of the form

xe1b1epk1bpk1=yepk=yψpk(n)x\cdot e_{1}^{b_{1}}\cdots e_{p^{k-1}}^{b_{p^{k}-1}}=y\cdot e_{p^{k}}=y^{\prime}\cdot\psi_{p^{k}}^{(n)} (5.12)

for some bib_{i}\in\mathbb{N}, yL𝐓/Iny\in L_{\mathbf{T}}/I_{n} and y=y(epk/ψpk(n))y^{\prime}=y\cdot(e_{p^{k}}/\psi_{p^{k}}^{(n)}) since ICpk𝐓=(epk)I_{C_{p^{k}}}^{\mathbf{T}}=(e_{p^{k}}) by Lemma 2.21.

If ll is coprime to pp, then epile_{p^{i}l} and epie_{p^{i}} become multiples of one another modulo epke_{p^{k}}: indeed, the corresponding characters in (Cpk)(C_{p^{k}})^{*} generate the same subgroup and thus Lemma 2.21 implies that epile_{p^{i}l} and epie_{p^{i}} generate the same ideal in LCpkL_{C_{p^{k}}}. It follows that Equation (5.12) gives rise to an equation

xe1a0epa1epk1ak1=yψpk(n),x\cdot e_{1}^{a_{0}}\cdot e_{p}^{a_{1}}\cdots e_{p^{k-1}}^{a_{k-1}}=y^{\prime\prime}\cdot\psi_{p^{k}}^{(n)},

with only Euler classes for powers of pp appearing on the left hand side. We claim that yy^{\prime\prime} must be divisible by the entire product e1a0epa1epk1ak1e_{1}^{a_{0}}\cdot e_{p}^{a_{1}}\cdots e_{p^{k-1}}^{a_{k-1}}, implying that xx is divisible by ψpk(n)\psi_{p^{k}}^{(n)}, as desired. To see this, recall that ψpk(n)\psi_{p^{k}}^{(n)} restricts to vnv_{n} in LCpl/InL_{C_{p^{l}}}/I_{n} for all l<kl<k. Hence we have that 0=res𝐓Cpl(yψm(n))=res𝐓Cpl(y)vn0=\operatorname{res}^{\mathbf{T}}_{C_{p^{l}}}(y^{\prime\prime}\cdot\psi_{m}^{(n)})=\operatorname{res}^{\mathbf{T}}_{C_{p^{l}}}(y^{\prime\prime})\cdot v_{n}. Since vnv_{n} is a regular element in LCpl/InL_{C_{p^{l}}}/I_{n} by 2.29, this implies that yy^{\prime\prime} is (uniquely) divisible by eple_{p^{l}}. This argument can be iterated by replacing yy^{\prime\prime} by y/eply^{\prime\prime}/e_{p^{l}}, and the statement follows. ∎

Remark 5.13.

One can show that more generally there exist elements ψm(n)L𝐓/In\psi_{m}^{(n)}\in L_{\mathbf{T}}/I_{n} for all mm\in\mathbb{N} uniquely determined by the equations

em=t|m(ψt(n))pνp(mt)n,e_{m}=\prod_{t|m}(\psi_{t}^{(n)})^{p^{\nu_{p}(\frac{m}{t})\cdot n}}, (5.14)

where νp()\nu_{p}(-) denotes the pp-adic valuation of a natural number. The element ψm(n)\psi_{m}^{(n)} generates the kernel of

ϕ𝐓Cm:L𝐓/InLCm/InΦCmL/In\phi^{\mathbf{T}}_{C_{m}}\colon L_{\mathbf{T}}/I_{n}\to L_{C_{m}}/I_{n}\to\Phi^{C_{m}}L/I_{n}

and its restriction to the trivial group is given by

res𝐓1(ψ(n)m)={0if m=1vnif m=pl and l>0qif m=ql with qp prime and l>01otherwise.\operatorname{res}^{\mathbf{T}}_{1}(\psi^{(n)}_{m})=\begin{cases}0&\text{if $m=1$}\\ v_{n}&\text{if $m=p^{l}$ and $l>0$}\\ q&\text{if $m=q^{l}$ with $q\neq p$ prime and $l>0$}\\ 1&\text{otherwise.}\end{cases}

We note also that every ψm(n)\psi_{m}^{(n)} is prime, since the geometric fixed points ΦCmL/In\Phi^{C_{m}}L/I_{n} are integral domains. The elements ψm(0)\psi_{m}^{(0)} were previously considered in [Hau, Proposition 5.46], denoted ψm\psi_{m} there.

Remark 5.15.

The proof of Proposition 5.10 applies in a more general context. Let XX be a global group law in the sense of [Hau, Definition 5.1]. As the global Lazard ring 𝐋\mathbf{L} is the initial global group law, there is a unique map 𝐋X\mathbf{L}\to X. Assume that the map L=𝐋(1)X(1)L=\mathbf{L}(1)\to X(1) sends InI_{n} to 0, and that for every l=0,,kl=0,\ldots,k the Euler classes eple_{p^{l}} in X(𝐓)X(\mathbf{T}) are regular elements and vnv_{n} is a regular element in X(Cpl)=X(𝐓)/eplX(C_{p^{l}})=X(\mathbf{T})/e_{p^{l}} (for example this is the case if vnv_{n} is regular in X(𝐓)X(\mathbf{T}) and eple_{p^{l}} remains a regular element modulo vnv_{n}). Then the image of ψpk(n)\psi_{p^{k}}^{(n)} in X(𝐓)X(\mathbf{T}) generates the kernel of the composition

Φ𝐓Cpk:X(𝐓)X(Cpk)ΦCpkX.\Phi^{\mathbf{T}}_{C_{p^{k}}}\colon X(\mathbf{T})\to X(C_{p^{k}})\to\Phi^{C_{p^{k}}}X.

For example, this applies to the coefficients of many Borel-equivariant complex oriented spectra, which can be used to compute their blue-shift numbers. We make use of this in Proposition 8.6.

Corollary 5.16.

We have an inclusion ICpkCpk,nICpk1,n+1I^{C_{p^{k}}}_{C_{p^{k}},n}\subseteq I^{C_{p^{k}}}_{1,n+1}.

Proof.

By Corollary 5.11, ICpkCpk,nI^{C_{p^{k}}}_{C_{p^{k}},n} is generated by InI_{n} and ψ¯(n)pk\overline{\psi}^{(n)}_{p^{k}}. Since InI_{n} is clearly contained in ICpk1,n+1I^{C_{p^{k}}}_{1,n+1} (even in ICpk1,nI^{C_{p^{k}}}_{1,n}), we can reduce modulo InI_{n} and need to show that ψ¯(n)pk\overline{\psi}^{(n)}_{p^{k}} is taken to 0 under the composition

LCpk/Inres𝐓1L/InL/In+1.L_{C_{p^{k}}}/I_{n}\xrightarrow{\operatorname{res}^{\mathbf{T}}_{1}}L/I_{n}\to L/I_{n+1}.

But this is clear, since ψ¯(n)pk\overline{\psi}^{(n)}_{p^{k}} restricts to vnv_{n} at the trivial group and vnv_{n} lies inside In+1I_{n+1}. ∎

We now have all the ingredients to prove the ‘if’-direction in Theorem 5.1:

Corollary 5.17.

Let BBB^{\prime}\subseteq B be an inclusion of subgroups of AA such that B/BB/B^{\prime} is pp-toral, and n,nn^{\prime},n\in\mathbb{N} such that nn+rankp(π0(B/B))n^{\prime}\geq n+\operatorname{rank}_{p}(\pi_{0}(B/B^{\prime})). Then there is an inclusion IAB,nIAB,nI^{A}_{B,n}\subseteq I^{A}_{B^{\prime},n^{\prime}}.

Proof.

By Lemma 5.6 we can assume that A=BA=B and that B=1B^{\prime}=1 is the trivial subgroup. Hence, AA is a pp-toral group. Let A0A^{0} denote the path component of the identity. We have IAA0,nIA1,nI^{A}_{A^{0},n^{\prime}}\subseteq I^{A}_{1,n^{\prime}} by Lemma 5.9. Hence it suffices to show that IAA,nI^{A}_{A,n} is contained in IAA0,nI^{A}_{A^{0},n^{\prime}}. For this, making use of Lemma 5.6 once more, we can assume that A0=1A^{0}=1 and hence AA is a finite abelian pp-group. We write m=rankp(A)m=\operatorname{rank}_{p}(A) and choose a filtration of subgroups

1A1A2Am=A1\subseteq A_{1}\subseteq A_{2}\subseteq\dots\subseteq A_{m}=A

such that every subquotient Ai/Ai1A_{i}/A_{i-1} is a cyclic pp-group. By Corollary 5.16 and Lemma 5.6 we see that

IAA,nIAAm1,n+1IAAm2,n+2IA1,n+mIA1,n.I^{A}_{A,n}\subseteq I^{A}_{A_{m-1},n+1}\subseteq I^{A}_{A_{m-2},n+2}\subseteq\dots\subseteq I^{A}_{1,n+m}\subseteq I^{A}_{1,n^{\prime}}.

The final inclusion follows from the assumption nn+rankp(π0(B/B))=n+mn^{\prime}\geq n+\operatorname{rank}_{p}(\pi_{0}(B/B^{\prime}))=n+m. This finishes the proof. ∎

5.3 Proof of non-inclusions

For the ‘only if’ direction we need to rule out further inclusions between prime ideals by constructing elements whose restrictions exhibit a large ‘height shift’. The goal is for every nn\in\mathbb{N} to construct an element xLCpn+1x\in L_{C_{p}^{n+1}} which lies in the ideal ICpn+1,0I_{C_{p}^{n+1},0} and whose restriction to the trivial group lies outside of InLI_{n}\subseteq L.

It turns out to be more natural to define such an element modulo a subideal of ICpn+1,0I_{C_{p}^{n+1},0}, namely the inflation of the ideal ICpn,0I_{C_{p}^{n},0} along the projection pCpn:Cpn+1Cpn×CpCpnp_{C_{p}^{n}}\colon C_{p}^{n+1}\cong C_{p}^{n}\times C_{p}\to C_{p}^{n}. That is, we construct an element

𝐯¯nICpn×Cp,0/pCpnICpn,0.\overline{\mathbf{v}}_{n}\in I_{C_{p}^{n}\times C_{p},0}/p^{*}_{C_{p}^{n}}I_{C_{p}^{n},0}.

By Corollary 5.17 the restriction map LCpnLL_{C_{p^{n}}}\to L takes ICpn,0I_{C_{p}^{n},0} into InI_{n}. Therefore, the restriction map LCpn+1LL_{C_{p^{n+1}}}\to L takes pICpn,0p^{*}I_{C_{p}^{n},0} into InI_{n} and we obtain an induced restriction map

LCpn×Cp,0/pCpnICpn,0L/In.L_{C_{p}^{n}\times C_{p},0}/p^{*}_{C_{p}^{n}}I_{C_{p}^{n},0}\to L/I_{n}.

We will see that, under this restriction, 𝐯¯n\overline{\mathbf{v}}_{n} is sent to vnv_{n}. Later in Section 6 we will show that 𝐯¯n\overline{\mathbf{v}}_{n} in fact forms a generator of the quotient ICpn×Cp,0/pCpnICpn,0I_{C_{p}^{n}\times C_{p},0}/p^{*}_{C_{p}^{n}}I_{C_{p}^{n},0} and that suitable inflations and restrictions of these elements generate all the invariant prime ideals at elementary abelian pp-groups.

We now turn to the construction of the element 𝐯¯n\overline{\mathbf{v}}_{n}. We set A=CpnA=C_{p}^{n} and first define an element 𝐯n\mathbf{v}_{n} in the ring LA×𝐓/pAIA,0L_{A\times\mathbf{T}}/p_{A}^{*}I_{A,0}, whose restriction to LA×Cp/pAIA,0L_{A\times C_{p}}/p^{*}_{A}I_{A,0} then yields 𝐯¯n\overline{\mathbf{v}}_{n}.

For every character VAV\in A^{*}, 2.27 yields a short exact sequence

0LA×𝐓eVτLA×𝐓(id,V1)LA0,0\to L_{A\times\mathbf{T}}\xrightarrow{e_{V\otimes\tau}\cdot}L_{A\times\mathbf{T}}\xrightarrow{(\operatorname{id},V^{-1})^{*}}L_{A}\to 0,

where we use that (id,V1):AA×𝐓(\operatorname{id},V^{-1})\colon A\to A\times\mathbf{T} identifies AA with the kernel of (Vid):A×𝐓𝐓(V\otimes\operatorname{id})\colon A\times\mathbf{T}\to\mathbf{T}. The inflation map LALA×𝐓L_{A}\to L_{A\times\mathbf{T}} provides an LAL_{A}-linear splitting of this exact sequence if we view LA×𝐓L_{A\times\mathbf{T}} as an LAL_{A}-module via the same inflation map. Thus, we obtain a short exact sequence

0LA×𝐓/pAIA,0eVτLA×𝐓/pAIA,0(id,V1)LA/IA,00.0\to L_{A\times\mathbf{T}}/p_{A}^{*}I_{A,0}\xrightarrow{e_{V\otimes\tau}}L_{A\times\mathbf{T}}/p_{A}^{*}I_{A,0}\xrightarrow{(\operatorname{id},V^{-1})^{*}}L_{A}/I_{A,0}\to 0. (5.18)

of LA/IA,0L_{A}/I_{A,0}-modules.

Remark 5.19.

We will see below in Section 6 that the ideal in LA×𝐓L_{A\times\mathbf{T}} generated by pAIAA,0p_{A}^{*}I^{A}_{A,0} equals the invariant prime ideal IA×𝐓A×𝐓,0I^{A\times\mathbf{T}}_{A\times\mathbf{T},0}. In particular, LA×𝐓/pAIA,0L_{A\times\mathbf{T}}/p_{A}^{*}I_{A,0} is again an integral domain. At this point it is only clear that pAIA,0Ap_{A}^{*}I_{A,0}^{A} is contained in IA×𝐓A×𝐓,0I^{A\times\mathbf{T}}_{A\times\mathbf{T},0}.

We now consider the Euler class eϵτpLA×𝐓/IA,0e_{\epsilon\otimes\tau^{p}}\in L_{A\times\mathbf{T}}/I_{A,0}. We have (id,V1)(eϵτp)=eVp=0(\operatorname{id},V^{-1})^{*}(e_{\epsilon\otimes\tau^{p}})=e_{V^{-p}}=0, since AA is an elementary abelian pp-group and hence every character is pp-torsion. By exactness, it follows that eϵτpe_{\epsilon\otimes\tau^{p}} is divisible by eVτe_{V\otimes\tau}, for all VAV\in A^{*}. We want to define 𝐯¯n\overline{\mathbf{v}}_{n} as the quotient of eϵτpe_{\epsilon\otimes\tau^{p}} by the product over all the eVτe_{V\otimes\tau}. For this we first need to check that the different eVτe_{V\otimes\tau} are coprime. To understand this, we consider the following:

Lemma 5.20.

Let x,yLA×𝐓/pAIA,0x,y\in L_{A\times\mathbf{T}}/p_{A}^{*}I_{A,0}, and assume that yy restricts to a non-zero element under LA×𝐓/pAIA,0(id,V1)LA/IA,0L_{A\times\mathbf{T}}/p_{A}^{*}I_{A,0}\xrightarrow{(\operatorname{id},V^{-1})^{*}}L_{A}/I_{A,0}. Then xx is divisible by eVτe_{V\otimes\tau} if and only if xyx\cdot y is divisible by eVτe_{V\otimes\tau}.

Proof.

By exactness of (5.18), xx is divisible by eVτe_{V\otimes\tau} if and only if (id,V1)x=0(\operatorname{id},V^{-1})^{*}x=0 in LA/IA,0L_{A}/I_{A,0}. Since (id,V1)(y)(\operatorname{id},V^{-1})^{*}(y) is non-zero by assumption and LA/IA,0L_{A}/I_{A,0} is an integral domain, this is the case if and only if

(id,V1)(xy)=(id,V1)x(id,V1)y=0,(\operatorname{id},V^{-1})^{*}(x\cdot y)=(\operatorname{id},V^{-1})^{*}x\cdot(\operatorname{id},V^{-1})^{*}y=0,

which in turn is equivalent to xyx\cdot y being divisible by eVτe_{V\otimes\tau}. ∎

Given a character V2VV_{2}\neq V, the restriction of eV2τe_{V_{2}\otimes\tau} along (id,V1)(\operatorname{id},V^{-1})^{*} equals the non-trivial Euler class eV2V1LA/IA,0e_{V_{2}V^{-1}}\in L_{A}/I_{A,0}. Hence the lemma applies to yy being any product of Euler classes of the form eVτe_{V^{\prime}\otimes\tau} with VVV^{\prime}\neq V. We find that in the quotient LA×𝐓/pAIA,0L_{A\times\mathbf{T}}/p_{A}^{*}I_{A,0}, the Euler class eϵτpe_{\epsilon\otimes\tau^{p}} is uniquely divisible by the product VAeVτ\prod_{V\in A^{*}}e_{V\otimes\tau}. To summarize:

Definition 5.21.

Let nn\in\mathbb{N} and A=CpnA=C_{p}^{n}. We define 𝐯nLA×𝐓/pAIA,0\mathbf{v}_{n}\in L_{A\times\mathbf{T}}/p^{*}_{A}I_{A,0} to be the unique element satisfying

eϵτp=𝐯nVAeVτLA×𝐓/pAIA,0,e_{\epsilon\otimes\tau^{p}}=\mathbf{v}_{n}\cdot\prod_{V\in A^{*}}e_{V\otimes\tau}\in L_{A\times\mathbf{T}}/p^{*}_{A}I_{A,0},

and we set

𝐯¯n=resA×𝐓A×Cp(𝐯n)LA×Cp/pAIA,0.\overline{\mathbf{v}}_{n}=\operatorname{res}^{A\times\mathbf{T}}_{A\times C_{p}}(\mathbf{v}_{n})\in L_{A\times C_{p}}/p_{A}^{*}I_{A,0}.
Remark 5.22.

The element 𝐯¯0\overline{\mathbf{v}}_{0} agrees with x0(p)LCpx_{0}^{(p)}\in L_{C_{p}} as defined in Section 5.1.

Remark 5.23.

Lemma 5.20 also applies in the ring LA×𝐓L_{A\times\mathbf{T}} itself (i.e., before quotienting by pAIA,0p_{A}^{*}I_{A,0}) if we demand that (id,V)y(\operatorname{id},V)^{*}y is regular element, rather than just being non-zero. These two conditions are equivalent in LA/IA,0L_{A}/I_{A,0} since it is an integral domain. However, the Euler classes are not regular elements in LAL_{A}, hence the lemma does not apply for yy a product of the eVτe_{V^{\prime}\otimes\tau}. In fact eϵτpe_{\epsilon\otimes\tau^{p}} is not divisible by VAeVτ\prod_{V\in A^{*}}e_{V\otimes\tau} before dividing out pAIA,0p_{A}^{*}I_{A,0}, even though it is divisible by each individual eVτe_{V\otimes\tau}. One can see this by restricting from A×𝐓A\times\mathbf{T} to 𝐓\mathbf{T}: If eϵτpe_{\epsilon\otimes\tau^{p}} was divisible by the product of all the eVτe_{V\otimes\tau}, this would imply that its restriction eτpL𝐓e_{\tau^{p}}\in L_{\mathbf{T}} was divisible by eτpne_{\tau}^{p^{n}}, since every eVτe_{V\otimes\tau} restricts to eτe_{\tau}. But eτpL𝐓e_{\tau^{p}}\in L_{\mathbf{T}} is divisible by eτe_{\tau} precisely once, since eτp/eτe_{\tau^{p}}/e_{\tau} restricts to pp at the trivial group, cf. Section 5. It is for this reason that it is most natural to define 𝐯n\mathbf{v}_{n} and 𝐯¯n\overline{\mathbf{v}}_{n} in this quotient. As we will see now, this matches nicely with the fact that vnv_{n} is most naturally defined in the quotient L/InL/I_{n}.

Proposition 5.24.
  1. 1.

    The element 𝐯¯n\overline{\mathbf{v}}_{n} defines a class in the ideal IA×Cp,0/pAIA,0I_{A\times C_{p},0}/p_{A}^{*}I_{A,0}, i.e., it is sent to 0 under the map LA×Cp,0/pAIA,0ΦA×CpL.L_{A\times C_{p},0}/p_{A}^{*}I_{A,0}\to\Phi^{A\times C_{p}}L.

  2. 2.

    The restriction map

    LA×Cp,0/pAIA,0L/InL_{A\times C_{p},0}/p_{A}^{*}I_{A,0}\to L/I_{n}

    takes 𝐯¯n\overline{\mathbf{v}}_{n} to vnv_{n}.

Proof.

Part 1: The equation

eϵτp=𝐯nVAeVτe_{\epsilon\otimes\tau^{p}}=\mathbf{v}_{n}\cdot\prod_{V\in A^{*}}e_{V\otimes\tau}

reduces to the equation

0=𝐯¯nVAeVτ¯0=\overline{\mathbf{v}}_{n}\cdot\prod_{V\in A^{*}}e_{V\otimes\overline{\tau}}

in LA×Cp/pAIA,0L_{A\times C_{p}}/p_{A}^{*}I_{A,0}, where τ¯\overline{\tau} denotes the restriction of τ𝐓\tau\in\mathbf{T}^{*} to CpC_{p}. Note that each (A×Cp)(A\times C_{p})-character of the form Vτ¯V\otimes\overline{\tau} is non-trivial. Hence 𝐯¯n\overline{\mathbf{v}}_{n} forms Euler-power torsion and therefore maps to 0 in the geometric fixed points.

For Part 2, we first restrict 𝐯n\mathbf{v}_{n} to L𝐓L_{\mathbf{T}}. This restriction takes pAIA,0p_{A}^{*}I_{A,0} into InI_{n} by Corollary 5.17, and sends each eVτe_{V\otimes\tau} to eτe_{\tau}. It follows that, modulo InI_{n}, we have an equation

res𝐓A×𝐓𝐯neτpn=eτp.\operatorname{res}_{\mathbf{T}}^{A\times\mathbf{T}}\mathbf{v}_{n}\cdot e_{\tau}^{p^{n}}=e_{\tau^{p}}.

Hence, modulo InI_{n}, res𝐓A×𝐓𝐯n\operatorname{res}_{\mathbf{T}}^{A\times\mathbf{T}}\mathbf{v}_{n} equals the element ψp(n)\psi_{p}^{(n)} from Proposition 5.10, whose restriction to the trivial group is vnv_{n}. This finishes the proof. ∎

Corollary 5.25.

If xnx_{n} is a preimage of 𝐯¯n\overline{\mathbf{v}}_{n} under the projection LCpn+1LCpn+1/pCpnICpn,0L_{C_{p}^{n+1}}\to L_{C_{p}^{n+1}}/p_{C_{p}^{n}}^{*}I_{C_{p}^{n},0} and BCpn+1B\subseteq C_{p}^{n+1} is a subgroup of rank 0mn+10\leq m\leq n+1, then

xnICpn+1B,n+1mICpn+1B,nm.x_{n}\in I^{C_{p}^{n+1}}_{B,n+1-m}-I^{C_{p}^{n+1}}_{B,n-m}.
Proof.

By the previous proposition, xnx_{n} is an element of ICpn+1Cpn+1,0I^{C_{p}^{n+1}}_{C_{p}^{n+1},0}. As A/BA/B has rank (n+1m)(n+1-m), we know by Corollary 5.17 that xnx_{n} must lie in ICpn+1B,n+1mI^{C_{p}^{n+1}}_{B,n+1-m}.

If xnx_{n} were an element of ICpn+1B,nmI^{C_{p}^{n+1}}_{B,n-m}, then applying Corollary 5.17 to the inclusion of the trivial group into BB shows that xnx_{n} is also an element of ICpn+11,nI^{C_{p}^{n+1}}_{1,n}. This contradicts the fact that, modulo InI_{n}, we have resCpn+11(xn)=resCpn+11(𝐯¯n)=vn\operatorname{res}^{C_{p}^{n+1}}_{1}(x_{n})=\operatorname{res}^{C_{p}^{n+1}}_{1}(\overline{\mathbf{v}}_{n})=v_{n}. ∎

Corollary 5.26.

If BBB^{\prime}\subseteq B is a pp-toral inclusion of subgroups of AA (i.e. B/BB/B^{\prime} is pp-toral) and n<n+rankp(π0(B/B))n^{\prime}<n+\operatorname{rank}_{p}(\pi_{0}(B/B^{\prime})), then IAB,nI^{A}_{B,n} does not include into IAB,nI^{A}_{B^{\prime},n^{\prime}}.

Proof.

By Lemma 5.6 we can reduce to the case A=BA=B and B=1B^{\prime}=1. Let r=rankp(π0(A))r=\operatorname{rank}_{p}(\pi_{0}(A)), and q:ACprq\colon A\to C_{p}^{r} be a surjection. Let xn+r1LCpn+rx_{n+r-1}\in L_{C_{p}^{n+r}} as in Corollary 5.25. Then, by the corollary, the restriction resCpn+rCpr(xn+r1)\operatorname{res}^{C_{p}^{n+r}}_{C_{p}^{r}}(x_{n+r-1}) is an element of ICprCpr,nI^{C_{p}^{r}}_{C_{p}^{r},n} but not an element of ICpr1,n+r1I^{C_{p}^{r}}_{1,n+r-1}. Therefore x=q(resCpn+rCpr(xn+r1))x=q^{*}(\operatorname{res}^{C_{p}^{n+r}}_{C_{p}^{r}}(x_{n+r-1})) is an element of IAA,nI^{A}_{A,n} whose restriction to the trivial group is not contained in In+r1I_{n+r-1}. In other words, xx is an element of IAA,nI^{A}_{A,n} but not an element of IA1,n+r1I^{A}_{1,n+r-1}. Since by assumption we have nn+r1n^{\prime}\leq n+r-1 and hence IA1,nIA1,n+r1I^{A}_{1,n^{\prime}}\subseteq I^{A}_{1,n+r-1}, this proves that IAA,nI^{A}_{A,n} does not include into IA1,nI^{A}_{1,n^{\prime}}. ∎

Combined with Corollaries 5.8 and 5.17, this proves Theorem 5.1.

6 Generators for invariant prime ideals

In this section we show that over elementary abelian pp-groups the elements 𝐯¯n\overline{\mathbf{v}}_{n} – together with the Euler classes – generate all invariant prime ideals under restriction and inflation maps. More precisely, we show the following theorem:

Theorem 6.1.
  1. 1.

    For every torus BB and nn\in\mathbb{N}, the ideal ICpn×B,0=ICpn×BCpn×B,0I_{C_{p}^{n}\times B,0}=I^{C_{p}^{n}\times B}_{C_{p}^{n}\times B,0} is generated by the elements

    p1(𝐯¯0),p2(𝐯¯1),,pn1(𝐯¯n2),pn(𝐯¯n1),p_{1}^{*}(\overline{\mathbf{v}}_{0}),p_{2}^{*}(\overline{\mathbf{v}}_{1}),\ldots,p_{n-1}^{*}(\overline{\mathbf{v}}_{n-2}),p_{n}^{*}(\overline{\mathbf{v}}_{n-1}),

    where pi:Cpn×BCpip_{i}\colon C_{p}^{n}\times B\to C_{p}^{i} is the projection to the first ii factors.

  2. 2.

    For every mm\in\mathbb{N} and every inclusion i:CpnCpn+mi\colon C_{p}^{n}\to C_{p}^{n+m}, the restriction map

    (i×B):LCpn+m×BLCpn×B(i\times B)^{*}\colon L_{C_{p}^{n+m}\times B}\to L_{C_{p}^{n}\times B}

    maps ICpn+m×B,0I_{C_{p}^{n+m}\times B,0} surjectively onto ICpn×B,mI_{C_{p}^{n}\times B,m}.

Remark 6.2.

Implicit in the statement of the theorem is that each pi(𝐯¯i1)p_{i}^{*}(\overline{\mathbf{v}}_{i-1}) is well-defined modulo the ideal generated by p1(𝐯¯0),,pi1(𝐯¯i2)p_{1}^{*}(\overline{\mathbf{v}}_{0}),\ldots,p_{i-1}^{*}(\overline{\mathbf{v}}_{i-2}). By definition, pi(𝐯¯i1)p_{i}^{*}(\overline{\mathbf{v}}_{i-1}) is an element of the quotient by the subideal generated by pi1ICpi1,0p_{i-1}^{*}I_{C_{p}^{i-1},0}. Applying the theorem to rank n1n-1 and B=0B=0 we see that this ideal is indeed generated by p1(𝐯¯0),,pi1(𝐯¯i2)p_{1}^{*}(\overline{\mathbf{v}}_{0}),\ldots,p_{i-1}^{*}(\overline{\mathbf{v}}_{i-2}), so the sequence of elements makes sense. Hence, the theorem and sequence should be interpreted in an inductive manner.

Combining both parts it follows that ICpn×B,mI_{C_{p}^{n}\times B,m} is generated by

(i×B)p1(𝐯¯0),(i×B)p2(𝐯¯1),,(i×B)pn+m1(𝐯¯n+m2),(i×B)𝐯¯n+m1,(i\times B)^{*}p_{1}^{*}(\overline{\mathbf{v}}_{0}),(i\times B)^{*}p_{2}^{*}(\overline{\mathbf{v}}_{1}),\ldots,(i\times B)^{*}p_{n+m-1}^{*}(\overline{\mathbf{v}}_{n+m-2}),(i\times B)^{*}\overline{\mathbf{v}}_{n+m-1},

where i:CpnCpn+mi\colon C_{p}^{n}\to C_{p}^{n+m} is any inclusion. The choice of inclusion will generally affect the resulting generators. For example, setting n=m=1n=m=1 and BB the trivial group: If we choose i1:CpCp2i_{1}\colon C_{p}\to C_{p}^{2} to be the inclusion into the first factor, the composite p1i1p_{1}\circ i_{1} becomes the identity. Hence we obtain that ICp,1I_{C_{p},1} is generated by the elements 𝐯¯0\overline{\mathbf{v}}_{0} and i1(𝐯¯1)i_{1}^{*}(\overline{\mathbf{v}}_{1}). If we alternatively use the inclusion i2:CpCp2i_{2}\colon C_{p}\to C_{p}^{2} into the second factor the composite p1i2p_{1}\circ i_{2} becomes the constant map, yielding the generators v0=pv_{0}=p and i2(𝐯¯1)i_{2}^{*}(\overline{\mathbf{v}}_{1}) (i.e., the same ones as in Corollary 5.11, as ψ¯p(1)\overline{\psi}_{p}^{(1)} equals i2(𝐯¯1)i_{2}^{*}(\overline{\mathbf{v}}_{1})). Furthermore, it follows that generators for ideals of the form IACpn×B,mI^{A}_{C_{p}^{n}\times B,m} with BB a torus can be obtained as the union of Euler classes (eV)V(e_{V})_{V\in\mathcal{B}} for a basis \mathcal{B} of ker(A(Cpn×B))\ker(A^{*}\to(C_{p}^{n}\times B)^{*}) together with a choice of generators for ICpn×BCpn×B,mI^{C_{p}^{n}\times B}_{C_{p}^{n}\times B,m}.

We prove Theorem 6.1 by induction on nn. Part 1 of the induction start n=0n=0 is the statement that IBB,0I^{B}_{B,0} is the 0-ideal for any torus BB (4.9). For Part 2 we need to see that the restriction LCpm×BLBL_{C_{p}^{m}\times B}\to L_{B} maps ICpm×B,0I_{C_{p}^{m}\times B,0} surjectively onto IB,mI_{B,m}, which we know is generated by ImI_{m} by 4.9. For i=0,,m1i=0,\ldots,m-1 we can consider the elements 𝐯¯iICpi+1,0/piICpi,0\overline{\mathbf{v}}_{i}\in I_{C_{p}^{i+1},0}/p_{i}^{*}I_{C_{p}^{i},0}, which reduce to viL/Ii1v_{i}\in L/I_{i-1}. It follows that the inflation of 𝐯¯i\overline{\mathbf{v}}_{i} to CpmC_{p}^{m} via any choice of surjection Cpm×BCpi+1C_{p}^{m}\times B\to C_{p}^{i+1} gives an element of a quotient of ICpm×B,0I_{C_{p}^{m}\times B,0} which reduces to viv_{i} in LB/Ii1LBL_{B}/I_{i-1}\cdot L_{B}. Since Ii/Ii1I_{i}/I_{i-1} is generated by viv_{i}, the claim follows.

We now assume that Theorem 6.1 holds for an elementary abelian pp-group AA of rank nn and show it also holds for A×CpA\times C_{p}. For any mm\in\mathbb{N} we consider the surjection

LA×𝐓×B/pAIA,mLA×Cp×B/pAIA,m,L_{A\times\mathbf{T}\times B}/p^{*}_{A}I_{A,m}\to L_{A\times C_{p}\times B}/p^{*}_{A}I_{A,m},

with kernel generated by eτpe_{\tau^{p}} for τ\tau the tautological 𝐓\mathbf{T}-character pulled back to A×𝐓×BA\times\mathbf{T}\times B. We first claim that if VAV\in A^{*} is non-trivial, then the Euler class pA(eV)p_{A}^{*}(e_{V}) is a non-zero divisor in LA×Cp×B/pAIA,mL_{A\times C_{p}\times B}/p^{*}_{A}I_{A,m}. To see this, we use that since 𝐓×B\mathbf{T}\times B is a torus we can apply the induction hypothesis to LA×𝐓×BL_{A\times\mathbf{T}\times B}. In particular, we know by Part 1 that pAIA,mp^{*}_{A}I_{A,m} generates the ideal IA×𝐓×B,mI_{A\times\mathbf{T}\times B,m} and hence LA×𝐓×B/pAIA,mL_{A\times\mathbf{T}\times B}/p_{A}^{*}I_{A,m} is an integral domain. So we have to show that pA(eV)p^{*}_{A}(e_{V}) still acts regularly modulo eτpe_{\tau^{p}}. Since both Euler classes are regular, this is equivalent to showing that eτpe_{\tau^{p}} is regular modulo pA(eV)p^{*}_{A}(e_{V}). We have an isomorphism

LA×𝐓×B/(pAIA,m,pA(eV))Lker(V)×𝐓×B/pker(V)(resAker(V)IA,m).L_{A\times\mathbf{T}\times B}/(p_{A}^{*}I_{A,m},p^{*}_{A}(e_{V}))\cong L_{\ker(V)\times\mathbf{T}\times B}/p^{*}_{\ker(V)}(\operatorname{res}^{A}_{\ker(V)}I_{A,m}).

By Part 2 of the induction hypothesis, we know that IA,mI_{A,m} restricts onto Iker(V),m+1I_{\ker(V),m+1}, hence the latter quotient identifies with Lker(V)×𝐓×B/pker(V)Iker(V),m+1L_{\ker(V)\times\mathbf{T}\times B}/p^{*}_{\ker(V)}I_{\ker(V),m+1}. Again we know by the induction hypothesis that this quotient is an integral domain, and eτpe_{\tau^{p}} is clearly a non-trivial element. So the claim follows and we have shown that pAIA,mp^{*}_{A}I_{A,m} generates the Euler power torsion in LA×Cp×BL_{A\times C_{p}\times B} (at height mm) for characters inflated up from AA.

Hence to understand the full ideal IA×Cp×B,mI_{A\times C_{p}\times B,m} it suffices to further divide by the Euler-power torsion for the remaining torsion characters in (A×Cp×B)(A\times C_{p}\times B)^{*} (there is no Euler-power torsion for non-torsion characters by 2.27). These torsion characters are of the form Vτ¯kV\otimes\overline{\tau}^{k}, where VAV\in A^{*}, τ¯\overline{\tau} is the restriction of τ𝐓\tau\in\mathbf{T}^{*} to CpC_{p} and k{1,,p1}k\in\{1,\ldots,p-1\}. Furthermore we can assume that k=1k=1: Any Vτ¯kV\otimes\overline{\tau}^{k} has some power of the form Vτ¯V^{\prime}\otimes\overline{\tau} and hence eVτ¯e_{V^{\prime}\otimes\overline{\tau}} is a multiple of eVτ¯ke_{V\otimes\overline{\tau}^{k}}. Thus, IA×Cp×B,m/pAIA,mI_{A\times C_{p}\times B,m}/p^{*}_{A}I_{A,m} is generated by Euler-power torsion for characters of the form Vτ¯V\otimes\overline{\tau}.

Again it is beneficial to pass to the integral domain LA×𝐓×B/pAIA,mL_{A\times\mathbf{T}\times B}/p_{A}^{*}I_{A,m} to understand the Euler-power torsion for these characters. We have the following:

Lemma 6.3.

Let A=CpnA=C_{p}^{n} be an elementary abelian pp-group, BB a torus and mm\in\mathbb{N}. Further let xLA×𝐓×B/pAIA,mx\in L_{A\times\mathbf{T}\times B}/p_{A}^{*}I_{A,m} be an element satisfying

xVAeVτnV=yeτpx\cdot\prod_{V\in A^{*}}e_{V\otimes\tau}^{n_{V}}=y\cdot e_{\tau^{p}}

for some yy and collection of natural numbers nVn_{V}. Then xx lies in the ideal generated by

pA×𝐓resA×Cpm×𝐓A×𝐓(𝐯n+m).p_{A\times\mathbf{T}}^{*}\operatorname{res}^{A\times C_{p}^{m}\times\mathbf{T}}_{A\times\mathbf{T}}(\mathbf{v}_{n+m}).
Proof.

By applying pA×Cpm×𝐓p_{A\times C_{p}^{m}\times\mathbf{T}}^{*} to the defining property of 𝐯n+m\mathbf{v}_{n+m} (Definition 5.21) we obtain the equation

eτp=pA×Cpm×𝐓(𝐯n+m)V(A×Cpm)eVτe_{\tau^{p}}=p_{A\times C_{p}^{m}\times\mathbf{T}}^{*}(\mathbf{v}_{n+m})\cdot\prod_{V\in(A\times C_{p}^{m})^{*}}e_{V\otimes\tau}

in LA×Cpm×𝐓×B/pA×CpmIA×Cpm,0L_{A\times C_{p}^{m}\times\mathbf{T}\times B}/p_{A\times C_{p}^{m}}^{*}I_{A\times C_{p}^{m},0}. Restricting from A×CpmA\times C_{p}^{m} to A=CpnA=C_{p^{n}} yields

eτp=pA×𝐓resA×Cpm×𝐓A×𝐓(𝐯n+m)VAeVτpme_{\tau^{p}}=p_{A\times\mathbf{T}}^{*}\operatorname{res}^{A\times C_{p}^{m}\times\mathbf{T}}_{A\times\mathbf{T}}(\mathbf{v}_{n+m})\cdot\prod_{V\in A^{*}}e_{V\otimes\tau}^{p^{m}} (6.4)

in the quotient LA×𝐓×B,m/pAIA,mL_{A\times\mathbf{T}\times B,m}/p_{A}^{*}I_{A,m}. This uses that every character of AA extends to pmp^{m} different characters of A×CpmA\times C_{p}^{m} and that the restriction of IA×Cpm,0LA×CpmI_{A\times C_{p}^{m},0}\subseteq L_{A\times C_{p}^{m}} lands in the ideal IA,mLAI_{A,m}\subseteq L_{A}. For the rest of the proof we write zz for the element pA×𝐓resA×Cpm×𝐓A×𝐓(𝐯n+m)p^{*}_{A\times\mathbf{T}}\operatorname{res}^{A\times C_{p}^{m}\times\mathbf{T}}_{A\times\mathbf{T}}(\mathbf{v}_{n+m}). With xx as in the statement of the lemma, we hence obtain an equation of the form

xVAeVτnV=zVAeVτpmyx\cdot\prod_{V\in A^{*}}e_{V\otimes\tau}^{n_{V}}=z\cdot\prod_{V\in A^{*}}e_{V\otimes\tau}^{p^{m}}\cdot y (6.5)

and we need to show that xx is a multiple of zz. The Euler classes eVτe_{V\otimes\tau} fit into short exact sequences of the form

0LA×𝐓×B/pAIA,meVτLA×𝐓×B/pAIA,m((id,V1)×B)LA×B/pAIA,m0,0\to L_{A\times\mathbf{T}\times B}/p_{A}^{*}I_{A,m}\xrightarrow{e_{V\otimes\tau}}L_{A\times\mathbf{T}\times B}/p_{A}^{*}I_{A,m}\xrightarrow{((\operatorname{id},V^{-1})\times B)^{*}}L_{A\times B}/p_{A}^{*}I_{A,m}\to 0,

analogously to Equation 5.18 above. As shown above, the induction hypothesis implies that the quotient LA×B/pAIA,mL_{A\times B}/p_{A}^{*}I_{A,m} is an integral domain. We know that zz restricts to vn+mL/In+mv_{n+m}\in L/I_{n+m} at the trivial group. In particular it must restrict non-trivially under each ((id,V1)×B)((\operatorname{id},V^{-1})\times B)^{*}. Hence the above short exact sequence together with the fact that LA×B/pAIA,mL_{A\times B}/p_{A}^{*}I_{A,m} is an integral domain implies: If an Euler class eVτe_{V\otimes\tau} divides an element of the form zαz\cdot\alpha, then eVτe_{V\otimes\tau} divides α\alpha. Applying this iteratively to Equation 6.5 (and using that LA×𝐓×B/pAIA,mL_{A\times\mathbf{T}\times B}/p_{A}^{*}I_{A,m} is an integral domain by the induction hypothesis) we see that VAeVτnV\prod_{V\in A^{*}}e_{V\otimes\tau}^{n_{V}} must divide the term VAeVτpmy\prod_{V\in A^{*}}e_{V\otimes\tau}^{p^{m}}\cdot y. Dividing on both sides shows that xx is a multiple of zz, as desired. ∎

Corollary 6.6.

The quotient

IA×Cp×B,m/pAIA,mI_{A\times C_{p}\times B,m}/p^{*}_{A}I_{A,m}

is generated by the element

resA×Cpm×Cp×BA×Cp×B(pA×Cpm×Cp𝐯¯n+m)=pA×Cp(resA×Cpm×CpA×Cp𝐯¯n+m).\operatorname{res}^{A\times C_{p}^{m}\times C_{p}\times B}_{A\times C_{p}\times B}(p_{A\times C_{p}^{m}\times C_{p}}^{*}\overline{\mathbf{v}}_{n+m})=p_{A\times C_{p}}^{*}(\operatorname{res}^{A\times C_{p}^{m}\times C_{p}}_{A\times C_{p}}\overline{\mathbf{v}}_{n+m}).
Proof.

We saw above that the quotient IA×Cp×B,m/pAIA,mI_{A\times C_{p}\times B,m}/p^{*}_{A}I_{A,m} is generated by Euler-power torsion for characters of the form Vτ¯V\otimes\overline{\tau}. An element x¯\overline{x} of LA×Cp×B/pAIA,mL_{A\times C_{p}\times B}/p_{A}^{*}I_{A,m} is such a torsion element if and only if it is the reduction of an element xLA×𝐓×B/pAIA,mx\in L_{A\times\mathbf{T}\times B}/p_{A}^{*}I_{A,m} satisfying the conditions of the lemma. Since the reduction of pA×𝐓resA×Cpm×𝐓A×𝐓(𝐯n+m)p_{A\times\mathbf{T}}^{*}\operatorname{res}^{A\times C_{p}^{m}\times\mathbf{T}}_{A\times\mathbf{T}}(\mathbf{v}_{n+m}) equals pA×CpresA×Cpm×CpA×Cp(𝐯¯n+m)p_{A\times C_{p}}^{*}\operatorname{res}^{A\times C_{p}^{m}\times C_{p}}_{A\times C_{p}}(\overline{\mathbf{v}}_{n+m}), it follows that x¯\overline{x} lies in the ideal generated by the latter.

As pA×CpresA×Cpm×CpA×Cp(𝐯¯n+m)p_{A\times C_{p}}^{*}\operatorname{res}^{A\times C_{p}^{m}\times C_{p}}_{A\times C_{p}}(\overline{\mathbf{v}}_{n+m}) is Euler-power torsion itself, it hence forms a generator of IA×Cp×B,m/pAIA,mI_{A\times C_{p}\times B,m}/p^{*}_{A}I_{A,m}. ∎

To finish the proof of Theorem 6.1: Setting m=0m=0 in the corollary shows that 𝐯¯n\overline{\mathbf{v}}_{n} generates the quotient IA×Cp×B,0/pAIA,0I_{A\times C_{p}\times B,0}/p_{A}^{*}I_{A,0}. By the induction hypothesis we know that IA,0I_{A,0} is generated by p1(𝐯¯0),,pn1(𝐯¯n2),𝐯¯n1p_{1}^{*}(\overline{\mathbf{v}}_{0}),\ldots,p_{n-1}^{*}(\overline{\mathbf{v}}_{n-2}),\overline{\mathbf{v}}_{n-1}. Combined this proves Part 1 for the group A×CpA\times C_{p}.

For Part 2 and general mm, we first note that it suffices to show the statement for any choice of injection i:A×CpCpn+m+1i\colon A\times C_{p}\to C_{p}^{n+m+1} since any two only differ by postcomposition with an automorphism of Cpn+m+1C_{p}^{n+m+1}. We can hence pick the canonical inclusion A×CpA×Cpm×CpA\times C_{p}\to A\times C_{p}^{m}\times C_{p}. By the induction hypothesis we know that IA×Cpm,0I_{A\times C_{p}^{m},0} surjects onto IA,mI_{A,m}. From the diagram

IA×Cpm,0\textstyle{I_{A\times C_{p}^{m},0}\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}pA×Cpm\scriptstyle{p_{A\times C_{p}^{m}}^{*}}res\scriptstyle{\operatorname{res}}IA,m\textstyle{I_{A,m}\ignorespaces\ignorespaces\ignorespaces\ignorespaces}pA\scriptstyle{p_{A}^{*}}IA×Cpm×Cp×B,0\textstyle{I_{A\times C_{p}^{m}\times C_{p}\times B,0}\ignorespaces\ignorespaces\ignorespaces\ignorespaces}res\scriptstyle{\operatorname{res}}IA×Cp×B,m\textstyle{I_{A\times C_{p}\times B,m}}

we see that pA(IA,m)p_{A}^{*}(I_{A,m}) is contained in the image of the lower horizontal arrow. Furthermore, Corollary 6.6 implies that IA×Cp×B,m/pAIA,mI_{A\times C_{p}\times B,m}/p^{*}_{A}I_{A,m} is generated by the restriction of an element of IA×Cpm×Cp×B,mI_{A\times C_{p}^{m}\times C_{p}\times B,m}. This finishes the proof.

Remark 6.7.

Unlike the sequence v0,,vn1v_{0},\ldots,v_{n-1}, the sequence

p1(𝐯¯0),p2(𝐯¯1),,pn1(𝐯¯n2),𝐯¯n1p_{1}^{*}(\overline{\mathbf{v}}_{0}),p_{2}^{*}(\overline{\mathbf{v}}_{1}),\ldots,p_{n-1}^{*}(\overline{\mathbf{v}}_{n-2}),\overline{\mathbf{v}}_{n-1}

isn’t regular. In fact, the ideal ICpn,0I_{C_{p}^{n},0} generated by these elements is precisely that of Euler-torsion.

This can be corrected by passing to a torus: The ideal I𝐓nCpn,0I^{\mathbf{T}^{n}}_{C_{p}^{n},0} is generated by the sequence p1(𝐯0),p2(𝐯1),,pn1(𝐯n2),𝐯¯n1p_{1}^{*}(\mathbf{v}_{0}),p_{2}^{*}(\mathbf{v}_{1}),\ldots,p_{n-1}^{*}(\mathbf{v}_{n-2}),\overline{\mathbf{v}}_{n-1}. Here, each pi+1(𝐯i)p_{i+1}^{*}(\mathbf{v}_{i}) is the element of

L𝐓n/(p1(𝐯0),p2(𝐯1),,pi(𝐯i1))LCpi×𝐓ni/ICpi×𝐓ni,0L_{\mathbf{T}^{n}}/(p_{1}^{*}(\mathbf{v}_{0}),p_{2}^{*}(\mathbf{v}_{1}),\ldots,p_{i}^{*}(\mathbf{v}_{i-1}))\cong L_{C_{p}^{i}\times\mathbf{T}^{n-i}}/I_{C_{p}^{i}\times\mathbf{T}^{n-i},0} (6.8)

obtained as the inflation of 𝐯iLCpi×𝐓/ICpi×𝐓,0\mathbf{v}_{i}\in L_{C_{p}^{i}\times\mathbf{T}}/I_{C_{p}^{i}\times\mathbf{T},0} along the projection to the first coordinate of 𝐓ni\mathbf{T}^{n-i}. Since each successive quotient LCpi×𝐓ni/ICpi×𝐓ni,0L_{C_{p}^{i}\times\mathbf{T}^{n-i}}/I_{C_{p}^{i}\times\mathbf{T}^{n-i},0} is an integral domain, the regularity of the sequence is clear once we have demonstrated the isomorphism claimed in (6.8). Similarly one shows that I𝐓nCpn,mI^{\mathbf{T}^{n}}_{C_{p}^{n},m} is generated by a regular sequence of length n+mn+m.

To establish the isomorphism, first note that (eτp)(pCpi1×𝐓(𝐯i1))(e_{\tau^{p}})\subseteq(p^{*}_{C_{p}^{i-1}\times\mathbf{T}}(\mathbf{v}_{i-1})) in the quotient ring LCpi1×𝐓ni+1/pCpi1ICpi1,0L_{C_{p}^{i-1}\times\mathbf{T}^{n-i+1}}/p^{*}_{C_{p}^{i-1}}I_{C_{p}^{i-1},0} by Lemma 6.3. Thus, 6.6 with m=0m=0 gives

L𝐓n/(p1(𝐯0),p2(𝐯1),,pi(𝐯i1))LCpi×𝐓ni/(p1(𝐯¯0),p2(𝐯¯1),,pi(𝐯¯i1)),L_{\mathbf{T}^{n}}/(p_{1}^{*}(\mathbf{v}_{0}),p_{2}^{*}(\mathbf{v}_{1}),\ldots,p_{i}^{*}(\mathbf{v}_{i-1}))\cong L_{C_{p}^{i}\times\mathbf{T}^{n-i}}/(p_{1}^{*}(\overline{\mathbf{v}}_{0}),p_{2}^{*}(\overline{\mathbf{v}}_{1}),\ldots,p_{i}^{*}(\overline{\mathbf{v}}_{i-1})),

and the claimed isomorphism becomes 6.1.

7 The Zariski topology on the spectrum of invariant prime ideals

(e)=I{1},0(e)=I_{\{1\},0}(e,v0)=I{1},1(e,v_{0})=I_{\{1\},1}(e,v0,v1)=I{1},2(e,v_{0},v_{1})=I_{\{1\},2}(e,v0,v1,v2)=I{1},3(e,v_{0},v_{1},v_{2})=I_{\{1\},3}(e,v0,v1,v2,)=I{1},(e,v_{0},v_{1},v_{2},\ldots)=I_{\{1\},\infty}ICp,0=(𝐯¯0)I_{C_{p},0}=(\overline{\mathbf{v}}_{0})ICp,1=(𝐯¯0,i1𝐯¯1)=(v0,i2𝐯¯1)I_{C_{p},1}=(\overline{\mathbf{v}}_{0},i_{1}^{*}\overline{\mathbf{v}}_{1})=(v_{0},i_{2}^{*}\overline{\mathbf{v}}_{1})ICp,2=(𝐯¯0,i1𝐯¯1,i1𝐯¯2)=(v0,i2𝐯¯1,i2𝐯¯2)=(v0,v1,i3𝐯¯2)I_{C_{p},2}=(\overline{\mathbf{v}}_{0},i_{1}^{*}\overline{\mathbf{v}}_{1},i_{1}^{*}\overline{\mathbf{v}}_{2})=(v_{0},i_{2}^{*}\overline{\mathbf{v}}_{1},i_{2}^{*}\overline{\mathbf{v}}_{2})=(v_{0},v_{1},i_{3}^{*}\overline{\mathbf{v}}_{2})ICp,3=(𝐯¯0,i1𝐯¯1,i1𝐯¯2,i1𝐯¯3)==(v0,v1,v2,i4𝐯¯3)I_{C_{p},3}=(\overline{\mathbf{v}}_{0},i_{1}^{*}\overline{\mathbf{v}}_{1},i_{1}^{*}\overline{\mathbf{v}}_{2},i_{1}^{*}\overline{\mathbf{v}}_{3})=\ldots=(v_{0},v_{1},v_{2},i_{4}^{*}\overline{\mathbf{v}}_{3})ICp,=(𝐯¯0,i1𝐯¯1,i1𝐯¯2,)I_{C_{p},\infty}=(\overline{\mathbf{v}}_{0},i_{1}^{*}\overline{\mathbf{v}}_{1},i_{1}^{*}\overline{\mathbf{v}}_{2},\ldots)
Figure 4: A picture of Specinv(LCp)\operatorname{Spec}^{\operatorname{inv}}(L_{C_{p}}), localized at pp, including different choices of generators, with ij:CpCpki_{j}\colon C_{p}\to C_{p}^{k} denoting the jjth canonical inclusion (generators arising from the further inclusions CpCpkC_{p}\to C_{p}^{k} are omitted). The yellow area depicts the closure of the point ICp,1I_{C_{p},1}.

The goal of this section is to describe the Zariski topology on Specinv(LA)\operatorname{Spec}^{\operatorname{inv}}(L_{A}), or equivalently the topology on the space |FGA||\mathcal{M}_{FG}^{A}| (Theorem 4.7). By definition, the closed subsets of Specinv(LA)\operatorname{Spec}^{\operatorname{inv}}(L_{A}) are the subsets of the form

V(X)={IB,nSpecinv(LA)|XIB,n}V(X)=\{I_{B,n}\in\operatorname{Spec}^{\operatorname{inv}}(L_{A})\ |\ X\subseteq I_{B,n}\}

for some subset XX of LAL_{A}. Hence we need to determine the collections of invariant prime ideals that arise as V(X)V(X) for some XX. We now fix a subset XX. Since a containment XIAB,nX\subseteq I^{A}_{B,n} automatically implies XIAB,nX\subseteq I^{A}_{B,n^{\prime}} for all nnn^{\prime}\geq n, it suffices to understand – for every closed subgroup BB of AA – the smallest value of n{}n\in\mathbb{N}\cup\{\infty\} such that XIAB,nX\subseteq I^{A}_{B,n}. In other words we need to determine the function

htX:Sub(A)¯={1}{}\operatorname{ht}_{X}\colon\operatorname{Sub}(A)\to\overline{\mathbb{N}}_{-}=\{-1\}\cup\mathbb{N}\cup\{\infty\}

defined by

htX(B)=sup{n|XIB,nA},\operatorname{ht}_{X}(B)=\sup\{n\ |\ X\not\subseteq I_{B,n}^{A}\},

where we set sup()=1\sup(\varnothing)=-1. We note that with this definition the height function of the image of vnLv_{n}\in L in LAL_{A} (thought of as a one-element set) is constantly nn. This follows from the fact that ΦBL\Phi^{B}L is a free LL-module by 2.25; cf. also the proof of Theorem 4.7. Moreover, we have:

Example 7.1.

Let xnLCpn+1x_{n}\in L_{C_{p}^{n+1}} be a lift of 𝐯¯nLCpn+1/pCpnICpn,0\overline{\mathbf{v}}_{n}\in L_{C_{p}^{n+1}}/p_{C_{p}^{n}}^{*}I_{C_{p}^{n},0}. Then Corollary 5.25 implies that htxn(B)=nrk(B)\operatorname{ht}_{x_{n}}(B)=n-\operatorname{rk}(B).

Our goal is to understand which functions Sub(A)¯\operatorname{Sub}(A)\to\overline{\mathbb{N}}_{-} arise as such height functions. In the previous section we showed that there are inclusions between invariant prime ideals associated to different subgroups of AA. These translate to conditions between the different values of htX\operatorname{ht}_{X}: If BBB^{\prime}\subseteq B is an inclusion of subgroups of AA such that B/BB/B^{\prime} is pp-toral, and XX is contained in IAB,nI^{A}_{B,n}, then XX is also contained in IAB,n+rankp(π0(B/B))I^{A}_{B^{\prime},n+\operatorname{rank}_{p}(\pi_{0}(B/B^{\prime}))}. In terms of the height function this translates to the inequality

htX(B)htX(B)+rankp(π0(B/B)).\operatorname{ht}_{X}(B^{\prime})\leq\operatorname{ht}_{X}(B)+\operatorname{rank}_{p}(\pi_{0}(B/B^{\prime})).

This leads us to the following definition:

Definition 7.2.

A function f:Sub(A)¯f\colon\operatorname{Sub}(A)\to\overline{\mathbb{N}}_{-} is called admissible if it satisfies the inequality

f(B)f(B)+rankp(π0(B/B))f(B^{\prime})\leq f(B)+\operatorname{rank}_{p}(\pi_{0}(B/B^{\prime}))

for every pp-toral inclusion BBB^{\prime}\subseteq B of closed subgroups of AA. Here, BBB^{\prime}\subseteq B is pp-toral if B/BB/B^{\prime} is a product of a torus and a pp-group.

By the above considerations, any height function htX\operatorname{ht}_{X} is admissible. When the group AA is finite, it turns out that the converse also holds: Any admissible function is realized by a height function htX\operatorname{ht}_{X}. For positive dimensional AA there is an additional condition on top of admissibility. To state this condition, we recall that choosing an invariant Riemannian metric dd on AA also equips the set of closed subgroups Sub(A)\operatorname{Sub}(A) with the Hausdorff metric, the underlying topology of which does not depend on the chosen metric on AA. This turns Sub(A)\operatorname{Sub}(A) into a compact totally-disconnected metric space, in which a sequence (Bi)i(B_{i})_{i\in\mathbb{N}} of closed subgroups converges to another closed subgroup BSub(A)B\in\operatorname{Sub}(A) if and only if almost all BiB_{i} are subgroups of BB and for every element bBb\in B the distance function d(b,Bi)d(b,B_{i}) converges to zero (see [tomDieckTransformationRepresentation, Section 5.6]). If BiBB_{i}\to B, we have the following two implications about representations:

  1. 1.

    Let WW be a representation of BB with WB=0W^{B}=0. Thus, writing WW as a sum of characters ViV_{i}, none of the ViV_{i} is trivial. For sufficiently large ii, no ker(Vi)\ker(V_{i}) contains BiB_{i} (since ker(Vi)B\ker(V_{i})\subseteq B is a closed proper subgroup) and thus WBi=0W^{B_{i}}=0.

  2. 2.

    Let VV and WW be two characters of BB such that resBBiV=resBBiW\operatorname{res}^{B}_{B_{i}}V=\operatorname{res}^{B}_{B_{i}}W for all sufficiently large ii. If VWV\neq W, then (VW1)B=0(V\cdot W^{-1})^{B}=0, in contradiction with the previous point. Thus V=WV=W.

We have the following:

Proposition 7.3.

For every finite subset XLAX\subseteq L_{A}, the height function htX\operatorname{ht}_{X} is a locally constant function on Sub(A)\operatorname{Sub}(A).

Proof.

We start by noting that

htX(B)=max(ht{x}(B)|xX).\operatorname{ht}_{X}(B)=\max(\operatorname{ht}_{\{x\}}(B)\ |\ x\in X).

The maximum of a finite number of locally constant functions is again locally constant. Hence it suffices to understand that ht{x}\operatorname{ht}_{\{x\}} is locally constant for any element xx of LAL_{A}.

Now let (Bi)i(B_{i})_{i\in\mathbb{N}} denote a sequence of subgroups of AA converging to a subgroup BB. We need to show that ht{x}(Bi)=ht{x}(B)\operatorname{ht}_{\{x\}}(B_{i})=\operatorname{ht}_{\{x\}}(B) for almost all ii. Without loss of generality we can assume that the BiB_{i} are subgroups of BB. Replacing xx by resAB(x)\operatorname{res}^{A}_{B}(x) if necessary we can further assume that A=BA=B.

We first assume that xIAA,nx\in I^{A}_{A,n} for some nn, and show that then also xIABi,nx\in I^{A}_{B_{i},n} for almost all BiB_{i}. If xIAA,nx\in I^{A}_{A,n}, there exists an AA-representation WW with WA=0W^{A}=0, such that eWxe_{W}\cdot x lies in the ideal LAInL_{A}\cdot I_{n}. For all ii large enough we have WBi=0W^{B_{i}}=0, meaning that for these ii the restriction eresABiWe_{\operatorname{res}^{A}_{B_{i}}W} becomes invertible in ΦBiL\Phi^{B_{i}}L. It follows that resABix\operatorname{res}^{A}_{B_{i}}x maps to the ideal generated by InI_{n} in ΦBiL\Phi^{B_{i}}L, in other words xx is contained in IABi,nI^{A}_{B_{i},n}.

For the other direction, we assume that xx is not contained in IAA,nI^{A}_{A,n} for some nn, and show that then also xIABi,nx\notin I^{A}_{B_{i},n} for almost all BiB_{i}. For this we recall from Remark 2.26 the construction of elements γjV\gamma_{j}^{V} for every character VGV\in G^{*} and jj\in\mathbb{N} satisfying the following three properties:

  1. 1.

    α(γjV)=γjα(V)\alpha^{*}(\gamma_{j}^{V})=\gamma_{j}^{\alpha^{*}(V)} for every group homomorphism α:GG\alpha\colon G^{\prime}\to G.

  2. 2.

    γ0V=eV\gamma_{0}^{V}=e_{V}.

  3. 3.

    ΦGLL[eV±1,γjV|VG{ϵ},j>0]\Phi^{G}L\cong L[e_{V}^{\pm 1},\gamma_{j}^{V}\ |\ V\in G^{*}-\{\epsilon\},j>0] for all abelian compact Lie groups GG.

Now, if xx is not contained in IAA,nI^{A}_{A,n}, it maps to a non-trivial element in

ΦAL/In=L/In[eV±1,γjV|VA{ϵ},j>0].\Phi^{A}L/I_{n}=L/I_{n}[e_{V}^{\pm 1},\gamma_{j}^{V}\ |\ V\in A^{*}-\{\epsilon\},j>0].

In other words, there exists an AA-representation WW with WA=0W^{A}=0 and pairwise different non-trivial characters V1,,VkV_{1},\ldots,V_{k} such that eWxe_{W}\cdot x is a polynomial over LL in the classes γjVl\gamma_{j}^{V_{l}}, l=1,,k,j0l=1,\ldots,k,j\geq 0, not all of whose coefficients are contained in InI_{n}. For all ii large enough we have that (i) WBi=0W^{B_{i}}=0 and that (ii) all the characters V1,,VkV_{1},\ldots,V_{k} restrict to pairwise different and non-trivial characters of BiB_{i}. It then follows that for these ii the element eresABiWresABixe_{\operatorname{res}^{A}_{B_{i}}W}\cdot\operatorname{res}^{A}_{B_{i}}x equals the corresponding polynomial in the classes γjresABiVl\gamma_{j}^{\operatorname{res}^{A}_{B_{i}}V_{l}}, implying that it maps to a non-trivial element in ΦBiL/In\Phi^{B_{i}}L/I_{n}. In other words, xx is not contained in IABi,nI^{A}_{B_{i},n} for ii large enough. This finishes the proof. ∎

Together with admissibility, this property characterizes the height functions of finite subsets of LAL_{A}:

Proposition 7.4.

Given a function f:Sub(A)¯f\colon\operatorname{Sub}(A)\to\overline{\mathbb{N}}_{-}, the following are equivalent:

  1. 1.

    There exists a finite subset XLAX\subseteq L_{A} such that f=htXf=\operatorname{ht}_{X}.

  2. 2.

    The function ff is admissible and locally constant.

Proof.

We have already shown the implication 1. \Rightarrow 2.

It remains to show that given a locally constant admissible function ff, there exists a finite subset XLAX\subseteq L_{A} with f=htXf=\operatorname{ht}_{X}. We start with the following claim: If ff is admissible, then given any pair of subgroups B,BAB,B^{\prime}\subseteq A, there exists an element xB,BLAx_{B,B^{\prime}}\in L_{A} such that htxB,B(B)=f(B)\operatorname{ht}_{x_{B,B^{\prime}}}(B)=f(B) and htxB,B(B)f(B)\operatorname{ht}_{x_{B,B^{\prime}}}(B^{\prime})\leq f(B^{\prime}). To see this, we distinguish between three cases:

  1. (i)

    If BB is not a subgroup of BB^{\prime}, we can choose a character VAV\in A^{*} which restricts to the trivial character over BB^{\prime} but to a non-trivial character over BB. Then xB,B=eVvf(B)x_{B,B^{\prime}}=e_{V}\cdot v_{f(B)} has the desired properties, since hteVvn(B)=n\operatorname{ht}_{e_{V}\cdot v_{n}}(B)=n and hteVvn(B)=1\operatorname{ht}_{e_{V}\cdot v_{n}}(B^{\prime})=-1. Here and in the following, we set v1=0v_{-1}=0, v0=pv_{0}=p and v=1v_{\infty}=1.

  2. (ii)

    If BB is a subgroup of BB^{\prime} with π0(B/B)\pi_{0}(B^{\prime}/B) not a pp-group, we choose a prime qpq\neq p dividing the order of π0(B/B)\pi_{0}(B^{\prime}/B) and a surjection g:BCqg\colon B^{\prime}\to C_{q} containing BB in the kernel. Then we set y=vf(B)g(x0(q))y=v_{f(B)}\cdot g^{*}(x_{0}^{(q)}), where x0(q)LCqx_{0}^{(q)}\in L_{C_{q}} is the element introduced in Section 5.1. Then x0(q)x_{0}^{(q)} is an element of ICqCq,0I^{C_{q}}_{C_{q},0} and its restriction to the trivial group is given by qq and hence a unit. It follows that hty(B)=1\operatorname{ht}_{y}(B^{\prime})=-1, since htg(x0(q))(B)=1\operatorname{ht}_{g^{*}(x_{0}^{(q)})}(B^{\prime})=-1. Moreover, hty(B)=f(B)\operatorname{ht}_{y}(B)=f(B), since resBB(y)=vf(B)q\operatorname{res}^{B^{\prime}}_{B}(y)=v_{f(B)}\cdot q. Hence, we can set xB,Bx_{B,B^{\prime}} to be any lift of yy to an element of LAL_{A}.

  3. (iii)

    The remaining case is when BB is a subgroup of BB^{\prime} with π0(B/B)\pi_{0}(B^{\prime}/B) a pp-group. Let rr be the minimum of the pp-rank of π0(B/B)\pi_{0}(B^{\prime}/B) and the number f(B)+1f(B)+1, and choose a surjection g:B/BCprg\colon B^{\prime}/B\to C_{p}^{r}. By 5.25 we know for f(B)<f(B)<\infty that there exist an element xf(B)LCpf(B)+1x_{f(B)}\in L_{C_{p}^{f(B)+1}} such that htxf(B)(1)=f(B)\operatorname{ht}_{x_{f(B)}}(1)=f(B) and htxf(B)(Cpf(B)+1)=1\operatorname{ht}_{x_{f(B)}}(C_{p}^{f(B)+1})=-1. (We set x1=0x_{-1}=0.) We can choose an embedding CprCpf(B)+1C_{p}^{r}\to C_{p}^{f(B)+1} and restrict xf(B)x_{f(B)} to an element yLCpry\in L_{C_{p}^{r}}. Then we have hty(1)=f(B)\operatorname{ht}_{y}(1)=f(B) and hty(Cpr)=f(B)r\operatorname{ht}_{y}(C_{p}^{r})=f(B)-r (see 5.25). If f(B)=f(B)=\infty, choose y=1y=1. It follows that htg(y)(B)=f(B)\operatorname{ht}_{g^{*}(y)}(B)=f(B) and htg(y)(B)f(B)rf(B)\operatorname{ht}_{g^{*}(y)}(B^{\prime})\leq f(B)-r\leq f(B^{\prime}), since ff is admissible. Hence, g(y)g^{*}(y) has the desired properties.

Now given any such pair (B,B)(B,B^{\prime}) there exists an open neighbourhood UBU_{B^{\prime}} of BB^{\prime} on which both htxB,B\operatorname{ht}_{x_{B,B^{\prime}}} and ff are constant. The UBU_{B^{\prime}} for varying BB^{\prime} form an open cover of the compact space Sub(A)\operatorname{Sub}(A). Let UB1,,UBkU_{B_{1}^{\prime}},\dots,U_{B_{k}^{\prime}} be a finite subcover. We then set

xB=xB,B1xB,B2xB,Bkx_{B}=x_{B,B_{1}^{\prime}}\cdot x_{B,B_{2}^{\prime}}\cdots x_{B,B_{k}^{\prime}}

to be the product of the corresponding elements. For any closed subgroup BB^{\prime} we have

htxB(B)=min(htxB,Bj(B)|j=1,,k).\operatorname{ht}_{x_{B}}(B^{\prime})=\min(\operatorname{ht}_{x_{B,B_{j}^{\prime}}}(B^{\prime})\ |\ j=1,\dots,k).

For B=BB^{\prime}=B this gives htx(B)=f(B)\operatorname{ht}_{x}(B)=f(B), since htxB,Bj(B)=f(B)\operatorname{ht}_{x_{B,B_{j}^{\prime}}}(B^{\prime})=f(B) for all jj. Any BB^{\prime} is contained in UBiU_{B_{i}^{\prime}} for some ii, yielding

htxB(B)htxB,Bi(B)=htxB,Bi(Bi)f(Bi).\operatorname{ht}_{x_{B}}(B^{\prime})\leq\operatorname{ht}_{x_{B,B_{i}^{\prime}}}(B^{\prime})=\operatorname{ht}_{x_{B,B_{i}^{\prime}}}(B_{i}^{\prime})\leq f(B_{i}^{\prime}).

In summary, the height function htxB\operatorname{ht}_{x_{B}} is less than or equal to ff everywhere and agrees with ff at BB itself. Once more we can apply that htxB\operatorname{ht}_{x_{B}} and ff are locally constant to find that htxB\operatorname{ht}_{x_{B}} and ff in fact agree on a neighborhood VBV_{B} of BB. Letting BB vary, this yields an open cover of Sub(A)\operatorname{Sub}(A), for which we can choose a finite subcover VB1,,VBlV_{B_{1}},\dots,V_{B_{l}}. Hence, for every BSub(A)B\in\operatorname{Sub}(A), there exists some i{1,,l}i\in\{1,\dots,l\} such that f(B)=htxBif(B)=\operatorname{ht}_{x_{B_{i}}}. Finally, we define XX to be the set {xB1,,xBl}\{x_{B_{1}},\dots,x_{B_{l}}\}. Since htX\operatorname{ht}_{X} is given by the maximum of the functions htBi\operatorname{ht}_{B_{i}}, it follows that XX has the desired property. ∎

Theorem 7.5.

The Zariski topology on Specinv(LA)\operatorname{Spec}^{\operatorname{inv}}(L_{A}) has as a basis the closed sets

Vf={IB,n|n>f(B)}V_{f}=\{I_{B,n}\ |\ n>f(B)\}

for all locally constant, admissible functions f:Sub(A)¯f\colon\operatorname{Sub}(A)\to\overline{\mathbb{N}}_{-}.

Proof.

A basis for the Zariski topology is given by the sets V(X)V(X) for all finite subsets XX of LAL_{A}. As we saw above, V(X)V(X) is determined by its height function htX\operatorname{ht}_{X} as

V(X)={IB,n|n>htX(B)}.V(X)=\{I_{B,n}\ |\ n>\operatorname{ht}_{X}(B)\}.

By Proposition 7.4, the functions Sub(A)¯\operatorname{Sub}(A)\to\overline{\mathbb{N}}_{-} that occur as height functions of finite subsets are precisely the locally constant admissible functions, which finishes the proof. ∎

8 Comparison with AA-spectra

In this final section, we discuss the relationship of the algebraic results of the previous sections with the theory of AA-spectra.

8.1 The universal support theory via MUAMU_{A}-homology

We begin by comparing our classification of invariant prime ideals with the Balmer spectrum of compact pp-local AA-spectra. We recall from [BalmerSpectrum] that a prime ideal 𝔭\mathfrak{p} of a tensor-triangulated category 𝒞\mathcal{C} is defined to be a thick tensor-ideal with the additional property that if XYX\otimes Y is contained in 𝔭\mathfrak{p}, then X𝔭X\in\mathfrak{p} or Y𝔭Y\in\mathfrak{p}. The set of all prime ideals assembles to a topological space Spec(𝒞)\operatorname{Spec}(\mathcal{C}), the Balmer spectrum, with the topology generated by the closed sets supp(X)={𝔭|X𝔭}\operatorname{supp}(X)=\{\mathfrak{p}\ |\ X\notin\mathfrak{p}\} for all objects X𝒞X\in\mathcal{C}.

Here, the support function supp()\operatorname{supp}(-), assigning a closed set of the Balmer spectrum to every object of 𝒞\mathcal{C}, is the universal support theory in the sense of Balmer. This means that it is terminal among pairs (T,σ)(T,\sigma) of a topological space TT and a function

σ:ob(𝒞){closed subsets of T}\sigma\colon\operatorname{ob}(\mathcal{C})\to\{\text{closed subsets of }T\}

satisfying σ(0)=\sigma(0)=\emptyset, σ(1)=T\sigma(1)=T, σ(XY)=σ(X)σ(Y)\sigma(X\oplus Y)=\sigma(X)\cup\sigma(Y), σ(ΣX)=σ(X)\sigma(\Sigma X)=\sigma(X) and σ(XY)=σ(X)σ(Y)\sigma(X\otimes Y)=\sigma(X)\cap\sigma(Y) for all X,Yob(𝒞)X,Y\in\operatorname{ob}(\mathcal{C}), and σ(X)σ(Y)σ(Z)\sigma(X)\subseteq\sigma(Y)\cup\sigma(Z) whenever there exists a distinguished triangle XYZΣXX\to Y\to Z\to\Sigma X. See [BalmerSpectrum] for more details.

In the case of compact pp-local AA-spectra SpAc\operatorname{Sp}_{A}^{c} for an abelian compact Lie group AA, the Balmer spectrum was computed in [BalmerSanders, BarthelHausmannNaumannNikolausNoelStapleton, BarthelGreenleesHausmann]. Given a closed subgroup BB and n{}n\in\mathbb{N}\cup\{\infty\}, one defines a prime ideal

P(B,n)={XSpAc|K(n)(ΦBX)=0},P(B,n)=\{X\in\operatorname{Sp}_{A}^{c}\ |\ K(n)_{*}(\Phi^{B}X)=0\},

where K(n)K(n) denotes the nth Morava KK-theory, and ΦBX\Phi^{B}X is the BB-geometric fixed point spectrum of XX, a compact pp-local spectrum.

Theorem 8.1 ([BalmerSanders, BarthelHausmannNaumannNikolausNoelStapleton, BarthelGreenleesHausmann]).

The map

Sub(A)ׯ\displaystyle\operatorname{Sub}(A)\times\overline{\mathbb{N}} Spec(SpAc)\displaystyle\to\operatorname{Spec}(\operatorname{Sp}_{A}^{c})
(B,n)\displaystyle(B,n) P(B,n)\displaystyle\mapsto P(B,n)

defines a bijection. Moreover, the topology on Spec(SpAc)\operatorname{Spec}(\operatorname{Sp}_{A}^{c}) has a basis given by the closed sets defined by

{P(B,n)|nf(B)}\{P(B,n)\ |\ n\geq f(B)\}

where ff ranges through all admissible functions Sub(A)¯={}\operatorname{Sub}(A)\to\overline{\mathbb{N}}=\mathbb{N}\cup\{\infty\}.

Here, ‘admissible’ is meant in the sense of Definition 7.2. Hence, comparing to Theorem 7.5, we see that the assignment IB,nP(B,n)I_{B,n}\mapsto P(B,n) defines a homeomorphism from Specinv(LA)\operatorname{Spec}^{\operatorname{inv}}(L_{A}) to Spec(SpAc)\operatorname{Spec}(\operatorname{Sp}_{A}^{c}).333The reader may have noticed that in this section our admissible functions ff take values in ¯\overline{\mathbb{N}}, while in the last section they took values in ¯\overline{\mathbb{N}}_{-}. Likewise, we consider the condition nf(B)n\geq f(B) here, and considered n>f(B)n>f(B) before. A shift by one shows that the topologies agree. The shift is caused by wanting to have vnv_{n} having height nn in the last section. The computation of Spec(SpAc)\operatorname{Spec}(\operatorname{Sp}_{A}^{c}) is also analogous to the one of Specinv(LA)\operatorname{Spec}^{\operatorname{inv}}(L_{A}) in the way that computing the underlying set is relatively straightforward (and is in fact known for all compact Lie groups), with most work going into understanding the topology.

Hence, we can view the universal support theory of SpAc\operatorname{Sp}_{A}^{c} to take values in the invariant prime ideals of LAL_{A}. The goal of the remainder of this section is to construct this universal support theory more intrinsically using MUAMU_{A}-homology and the structure of equivariant formal group laws described in this paper. The idea is the following: Given a compact pp-local AA-spectrum XX, we can consider its equivariant complex bordism homology (MU¯A)X(\underline{MU}_{A})_{*}X. Here, underlining MUAMU_{A} indicates that we take the AA-Mackey functor valued homology of XX, i.e., we record the collection of (MUB)(resABX)(MU_{B})_{*}(\operatorname{res}^{A}_{B}X) for all closed subgroups BB of AA, together with restriction and transfer maps between them. We will always work at the fixed prime pp and pp-localize everything implicitly.

Since the coefficients πAMUA\pi_{*}^{A}MU_{A} are isomorphic to the Lazard ring LAL_{A} and moreover the cooperations πAMUAMUA\pi_{*}^{A}MU_{A}\wedge MU_{A} agree with SAS_{A}, the groups (MUA)X(MU_{A})_{*}X form a graded AA-Mackey functor in (LA,SA)(L_{A},S_{A})-comodules. As such, we can take its support in the invariant prime ideals

supp((MU¯A)X)={IAB,n|((MU¯A)X)IAB,n=0}Specinv(LA).\operatorname{supp}((\underline{MU}_{A})_{*}X)=\{I^{A}_{B,n}\ |\ ((\underline{MU}_{A})_{*}X)_{I^{A}_{B,n}}=0\}\subseteq\operatorname{Spec}^{\operatorname{inv}}(L_{A}).
Remark 8.2.

In general, supp((MU¯A)X)\operatorname{supp}((\underline{MU}_{A})_{*}X) is different from supp((MUA)X)\operatorname{supp}((MU_{A})_{*}X). Take for example A=C2A=C_{2}, p=2p=2 and X=SσX=S^{\sigma}, the circle with action given by reflection at a line. We have

MU(res1C2Sσ)I1,0=MU(S1)0.MU_{*}(\operatorname{res}_{1}^{C_{2}}S^{\sigma})_{I_{1,0}}=MU_{*}(S^{1})_{\mathbb{Q}}\neq 0.

The module (MUC2)(Sσ)I1,0C2(MU_{C_{2}})_{*}(S^{\sigma})_{I_{1,0}^{C_{2}}} is rational as well since pI1,0C2p\notin I_{1,0}^{C_{2}}. Rationally, (MUC2)(Sσ)(MU_{C_{2}})_{*}(S^{\sigma}) splits into the coinvariants (MUC2)(Sσ)C2=0(MU_{C_{2}})_{*}(S^{\sigma})_{C_{2}}=0 and the geometric fixed points (MUC2)(Sσ)[e1](MU_{C_{2}})_{*}(S^{\sigma})[e^{-1}], for ee the Euler class of the unique non-trivial character. The element 𝐯¯0\overline{\mathbf{v}}_{0} from 5.21 becomes zero in the geometric fixed points, but restricts to 22 and is thus not in I1,0C2I_{1,0}^{C_{2}}; thus (MUC2)(Sσ)[e1](MU_{C_{2}})_{*}(S^{\sigma})[e^{-1}] becomes zero as well after localization at I1,0C2I_{1,0}^{C_{2}}, and (MUC2)(Sσ)I1,0C2=0(MU_{C_{2}})_{*}(S^{\sigma})_{I_{1,0}^{C_{2}}}=0.

Remark 8.3 (Transfer maps).

The isomorphisms LAπAMUAL_{A}\cong\pi^{A}_{*}MU_{A} imply that on top of the contravariant restriction maps along group homomorphisms there also exist transfer maps LBLAL_{B}\to L_{A} for inclusions BAB\subseteq A of finite index. In other words, the collection of all equivariant Lazard rings LAL_{A} forms a ‘global Green functor’ on the family of abelian compact Lie groups, in the sense of [SchGlobal, Definition 5.1.3]. While we have no general interpretation of transfers in terms of equivariant formal groups, we can compute them as follows: By Frobenius reciprocity, it suffices to compute trBA:LBLA\operatorname{tr}_{B}^{A}\colon L_{B}\to L_{A} on 1LB1\in L_{B} since the restriction is surjective. Inductively, we can further assume that A/BCpA/B\cong C_{p}. Furthermore, trBA(1)=qtr{1}Cp(1)\operatorname{tr}_{B}^{A}(1)=q^{*}\operatorname{tr}_{\{1\}}^{C_{p}}(1) for q:AA/BCpq\colon A\to A/B\cong C_{p} since transfers are compatible with inflation maps (see, e.g., [SchGlobal, Theorem 4.2.6 ff.]). Hence it suffices to identify tr{1}Cp(1)\operatorname{tr}_{\{1\}}^{C_{p}}(1). We claim that it equals 𝐯¯0LCp\overline{\mathbf{v}}_{0}\in L_{C_{p}}. Indeed, any transfer maps to zero in the geometric fixed points and is thus an element of ICp,0I_{C_{p},0}. We know that 𝐯¯0\overline{\mathbf{v}}_{0} generates ICp,0I_{C_{p},0}. Hence, we can write tr{1}Cp(1)=x𝐯¯0\operatorname{tr}_{\{1\}}^{C_{p}}(1)=x\cdot\overline{\mathbf{v}}_{0} for xLCpx\in L_{C_{p}}. Writing xx as x0+xex_{0}+x^{\prime}\cdot e for ee a non-trivial Euler class and x0Lx_{0}\in L, we obtain tr{1}Cp(1)=x0𝐯¯0\operatorname{tr}_{\{1\}}^{C_{p}}(1)=x_{0}\cdot\overline{\mathbf{v}}_{0} since e𝐯¯0=0e\cdot\overline{\mathbf{v}}_{0}=0 by the definition of 𝐯¯0\overline{\mathbf{v}}_{0}. We obtain x0p=res{1}Cptr{1}Cp(1)=px_{0}\cdot p=\operatorname{res}_{\{1\}}^{C_{p}}\operatorname{tr}_{\{1\}}^{C_{p}}(1)=p since 𝐯¯0\overline{\mathbf{v}}_{0} restricts to pp and thus x0=1x_{0}=1 and tr{1}Cp(1)=𝐯¯0\operatorname{tr}_{\{1\}}^{C_{p}}(1)=\overline{\mathbf{v}}_{0}.

We will now see that our notion of support is another model for the universal support theory on compact AA-spectra.

We first note a major inconvenience: It is unclear whether (MUA)X(MU_{A})_{*}X is a finitely generated LAL_{A}-module, even for compact XX. This is in contrast with the non-equivariant situation, where finite generation of MUXMU_{*}X for compact XX follows from the fact that LL is a polynomial ring and hence coherent. An analogous statement is unknown for equivariant Lazard rings. In particular, it is a priori unclear that supp((MU¯A)X)\operatorname{supp}((\underline{MU}_{A})_{*}X) is indeed a closed subset of Specinv(LA)\operatorname{Spec}^{\operatorname{inv}}(L_{A}) and we have to prove this by hand, see Proposition 8.5 below.

The following proposition gives the relationship between the MUAMU_{A}-homology support theory described above and geometric fixed points.

Proposition 8.4.

Let BB be a subgroup of AA, n¯n\in\overline{\mathbb{N}} and XX a compact AA-spectrum. Then the following are equivalent:

  1. (i)

    The Mackey functor ((MU¯A)(X))IAB,n((\underline{MU}_{A})_{*}(X))_{I^{A}_{B,n}} is trivial.

  2. (ii)

    The (MUB)(MU_{B})_{*}-module ((MUB)(resABX))IBB,n((MU_{B})_{*}(\operatorname{res}^{A}_{B}X))_{I^{B}_{B,n}} is trivial.

  3. (iii)

    The BB-geometric fixed points ΦB(X)\Phi^{B}(X) are of chromatic type >n>n.

We give two proofs of this proposition below, one using the results of [BalmerSanders, BarthelHausmannNaumannNikolausNoelStapleton, BarthelGreenleesHausmann] and one independent of these results. The latter one is complicated by the fact that we don’t know whether (MUA)X(MU_{A})_{*}X is always finitely generated.

We have the following corollary:

Proposition 8.5.

Let XX be a compact AA-spectrum. Then its support supp((MU¯A)X)\operatorname{supp}((\underline{MU}_{A})_{*}X) is a closed subset of Specinv(LA)\operatorname{Spec}^{\operatorname{inv}}(L_{A}).

Moreover, the assignment

ob(SpAc)Specinv(LA);Xsupp((MU¯A)X)\operatorname{ob}(\operatorname{Sp}_{A}^{c})\to\operatorname{Spec}^{\operatorname{inv}}(L_{A})\ ;\ X\mapsto\operatorname{supp}((\underline{MU}_{A})_{*}X)

is a support theory on compact AA-spectra.

Proof.

We consider the type function

typeX:Sub(A)¯;Xtype(ΦBX),\operatorname{type}_{X}\colon\operatorname{Sub}(A)\to\overline{\mathbb{N}}\ ;\ X\mapsto type(\Phi^{B}X),

which is locally constant by [BarthelGreenleesHausmann, Proposition 4.3]. By the previous proposition, IB,nI_{B,n} is an element of supp((MU¯A)X)\operatorname{supp}((\underline{MU}_{A})_{*}X) if and only if ntypeX(B)n\geq\operatorname{type}_{X}(B). Since the support supp((MU¯A)X)\operatorname{supp}((\underline{MU}_{A})_{*}X) is closed under inclusion, Theorem 5.1 implies that IB,nsupp((MU¯A)X)I_{B^{\prime},n}\in\operatorname{supp}((\underline{MU}_{A})_{*}X) if IB,nrkp(π0(B/B))supp((MU¯A)X)I_{B,n-\operatorname{rk}_{p}(\pi_{0}(B/B^{\prime}))}\in\operatorname{supp}((\underline{MU}_{A})_{*}X) for every pp-toral inclusion BBB^{\prime}\subseteq B. Thus typeX\operatorname{type}_{X} is admissible in the sense of Definition 7.2. Theorem 7.5 implies that supp((MU¯A)X)\operatorname{supp}((\underline{MU}_{A})_{*}X) is closed, as desired.

For the second part, all required properties of a support theory follow easily from exactness of localization ()IAB,n(-)_{I^{A}_{B,n}} except for the one on the interplay with smash products. This in turn follows from the third characterization in Proposition 8.4, since the type of a smash product of two compact spectra is the maximum of the two types. ∎

By the universal property of the Balmer spectrum, we obtain a continuous map

Specinv(LA)Spec(SpAc).\operatorname{Spec}^{\operatorname{inv}}(L_{A})\to\operatorname{Spec}(\operatorname{Sp}_{A}^{c}).

Proposition 8.4 makes it clear that this map sends IAB,nI^{A}_{B,n} to P(B,n)P(B,n) and is hence bijective. Therefore we can conclude from the results of this paper that the topology on Spec(SpAc)\operatorname{Spec}(\operatorname{Sp}_{A}^{c}) is at least as coarse as the one on Specinv(LA)\operatorname{Spec}^{\operatorname{inv}}(L_{A}). In other words, our proof of the existence of inclusions IB,nIB,nI_{B,n}\subseteq I_{B^{\prime},n^{\prime}} gives another proof of the analogous inclusion P(B,n)P(B,n)P(B^{\prime},n^{\prime})\subseteq P(B,n) on the topological side. The fact that there are no further topological inclusions requires additional arguments, namely the existence of compact AA-spectra with ‘maximal type shifting behaviour’. See [BarthelHausmannNaumannNikolausNoelStapleton, Section 4] or [KuhnLloydChromatic, Section 7]. Knowing this, we see that Xsupp((MU¯A)X)X\mapsto\operatorname{supp}((\underline{MU}_{A})_{*}X) is a universal support theory on compact AA-spectra.

It remains to give the proof of Proposition 8.4, which we will do in three steps:

(i)(ii)(i)\Rightarrow(ii): Since the LAL_{A}-action on (MUB)(X)(MU_{B})_{*}(X) factors through resAB:LALB\operatorname{res}^{A}_{B}\colon L_{A}\to L_{B} and IAB,n=(resAB)1)(IBB,n)I^{A}_{B,n}=(\operatorname{res}^{A}_{B})^{-1})(I^{B}_{B,n}), it follows that we have an isomorphism ((MUB)(X))IAB,n((MUB)(X))IBB,n((MU_{B})_{*}(X))_{I^{A}_{B,n}}\cong((MU_{B})_{*}(X))_{I^{B}_{B,n}}. In particular the vanishing of the entire Mackey-functor (MU¯(X))IAB,n(\underline{MU}_{*}(X))_{I^{A}_{B,n}} implies the vanishing at the subgroup BB as a special case.

(ii)(iii)(ii)\Longleftrightarrow(iii): Note that IBB,nI^{B}_{B,n} contains none of the non-trivial Euler classes for BB, hence the non-trivial Euler classes act invertibly on ((MUB)(X))IBB,n((MU_{B})_{*}(X))_{I^{B}_{B,n}}. It follows that we have an isomorphism

((MUB)(X))IBB,n((ΦBMUB)ΦBX)IBB,nΦB((MUB)IBB,n)ΦBX.((MU_{B})_{*}(X))_{I^{B}_{B,n}}\cong((\Phi^{B}MU_{B})_{*}\Phi^{B}X)_{I^{B}_{B,n}}\cong\Phi^{B}((MU_{B})_{I^{B}_{B,n}})_{*}\Phi^{B}X.

Modulo InI_{n}, the ring ΦB((MUB)IBB,n)\Phi^{B}((MU_{B})_{I^{B}_{B,n}})_{*} embeds into the field of fractions of ΦBL/In\Phi^{B}L/I_{n}, which is non-trivial. Moreover, vnv_{n} is not contained in IBB,nI^{B}_{B,n} and hence is invertible in the localization. It follows that the localization (ΦBMUB)IBB,n(\Phi^{B}MU_{B})_{I^{B}_{B,n}} is an MUMU-algebra of height nn, i.e., its vanishing detects compact spectra of type n+1\geq n+1.

(iii)(i)(iii)\Rightarrow(i): Let XX be a compact AA-spectrum such that ΦB(X)\Phi^{B}(X) is of type n\geq n. By the previous paragraph we know that ((MUB)(X))IAB,n=0((MU_{B})_{*}(X))_{I^{A}_{B,n}}=0, and we have to show that ((MUB~)(X))IAB,n=0((MU_{\widetilde{B}})_{*}(X))_{I^{A}_{B,n}}=0 for all other closed subgroups B~\widetilde{B}, too. We first assume that B~\widetilde{B} does not contain BB. Then there exists a character VAV\in A^{*} which restricts to the trivial character for B~\widetilde{B} but to a non-trivial character for BB. It follows that eVe_{V} is not contained in IAB,nI^{A}_{B,n} and hence acts invertibly on (MUB~)IAB,n(MU_{\widetilde{B}})_{I^{A}_{B,n}}. On the other hand eVe_{V} restricts to 0 in MUB~MU_{\widetilde{B}}, so it follows that (MUB~)IAB,n=0(MU_{\widetilde{B}})_{I^{A}_{B,n}}=0 and in particular ((MUB~)(X))IAB,n=0((MU_{\widetilde{B}})_{*}(X))_{I^{A}_{B,n}}=0, as desired. Hence we can assume that B~\widetilde{B} contains BB as a subgroup. Since the statement then no longer depends on the ambient group, we can reduce to the case B~=A\widetilde{B}=A.

Hence we need to show that (MUA)(X))IAB,n=0(MU_{A})_{*}(X))_{I^{A}_{B,n}}=0. By induction on the pair (dimension, rank of π0(A/B)\pi_{0}(A/B)) it follows that all the localizations at smaller intermediate groups BB~AB\subseteq\widetilde{B}\subseteq A vanish, and hence the homotopy groups of (MUA)IAB,nX(MU_{A})_{I^{A}_{B,n}}\wedge X are concentrated at AA. In particular this implies that the map ((MUA)IAB,n)X((ΦAMUA)IAB,n)ΦAX((MU_{A})_{I^{A}_{B,n}})_{*}X\to((\Phi^{A}MU_{A})_{I^{A}_{B,n}})_{*}\Phi^{A}X is an isomorphism and all Euler classes eVe_{V} for non-trivial characters act invertibly on ((MUA)IAB,n)X((MU_{A})_{I^{A}_{B,n}})_{*}X. Now, if VV restricts to the trivial character in BB^{*} (such a VV always exists since we can assume that BB is a proper subgroup of AA), its Euler class eVe_{V} lies in the maximal ideal IAB,nI^{A}_{B,n} of ((MUA)IAB,n)((MU_{A})_{I^{A}_{B,n}})_{*}.

If we knew that ((MUA)IAB,n)X((MU_{A})_{I^{A}_{B,n}})_{*}X is finitely generated over (MUA)(MU_{A})_{*} we could apply Nakayama’s lemma to see directly that ((MUA)IAB,n)X=0((MU_{A})_{I^{A}_{B,n}})_{*}X=0, as desired. Since we do not know this, we have to argue differently: By Corollary 5.2, we know that ΦA((MUA)IAB,n)\Phi^{A}((MU_{A})_{I^{A}_{B,n}})_{*} is trivial if π0(A/B)\pi_{0}(A/B) is not a pp-group, and is of height nrankp(π0(A/B))n-\operatorname{rank}_{p}(\pi_{0}(A/B)) otherwise (where negative heights again mean that the theory is trivial). Hence, what we want to show is that if ΦBX\Phi^{B}X is of type n\geq n, then ΦAX\Phi^{A}X is of type nrankp(π0(A/B))\geq n-\operatorname{rank}_{p}(\pi_{0}(A/B)). Indeed, this implies that ((MUA)IAB,n)X((ΦAMUA)IAB,n)ΦAX((MU_{A})_{I^{A}_{B,n}})_{*}X\cong((\Phi^{A}MU_{A})_{I^{A}_{B,n}})_{*}\Phi^{A}X is trivial.

This precise statement about ΦAX\Phi^{A}X is one of the main results of [BarthelGreenleesHausmann], building on [BalmerSanders] and [BarthelHausmannNaumannNikolausNoelStapleton]. Hence, using these results, Proposition 8.4 follows. Alternatively, rather than importing we can reprove the above statement using the methods from this paper. Note that by induction on the rank of A/BA/B and by replacing XX by the compact A/BA/B-spectrum ΦB¯X\underline{\Phi^{B}}X it suffices to show two special cases:

  1. 1.

    If XX is a compact CpkC_{p^{k}}-spectrum of underlying type n\geq n, then the type of ΦCpkX\Phi^{C_{p^{k}}}X is at least n1n-1.

  2. 2.

    If XX is a compact 𝐓\mathbf{T}-spectrum of underlying type n\geq n, then the type of Φ𝐓X\Phi^{\mathbf{T}}X is also n\geq n.

For (1) it suffices to find a complex oriented theory EE of height nn such that ΦCpkE¯\Phi^{C_{p^{k}}}\underline{E} is of height n1\geq n-1, where E¯=F(EG+,E)\underline{E}=F(EG_{+},E) denotes the Borel theory associated to EE (see the proofs of [BarthelHausmannNaumannNikolausNoelStapleton, Corollary 3.12] or [BarthelGreenleesHausmann, Proposition 6.10] for details on this argument). In [BarthelHausmannNaumannNikolausNoelStapleton] it was shown that Morava EE-theory E=EnE=E_{n} has this property, building on results of Hopkins-Kuhn-Ravenel on the pp-disivible group associated to EnE_{n} [HKR00] and extending earlier work of Greenlees, Hovey and Sadofsky. Using Remark 5.15 we obtain similar results more generally:

Proposition 8.6.

Let EE be any complex oriented ring spectrum of height nn which is Landweber exact over MU/In1=MU/(v0,,vn2)MU/I_{n-1}=MU/(v_{0},\ldots,v_{n-2}), i.e., In1I_{n-1} acts trivially on πE\pi_{*}E, vn1πEv_{n-1}\in\pi_{*}E is a regular element and vnv_{n} is a unit modulo vn1v_{n-1}. Then ΦCpkE¯\Phi^{C_{p^{k}}}\underline{E} is of height n1n-1.

Proof.

We apply Remark 5.15 and check that its assumptions are satisfied. We have (E¯𝐓)=Ee(\underline{E}^{\mathbf{T}})_{*}=E_{*}\llbracket e\rrbracket, with Euler classes ene_{n} given by the nn-series [n]F(e)[n]_{F}(e) for the formal group law associated to EE. If nn is a power of pp, the leading term of this Euler class is a power of vn1v_{n-1}, which we assumed to be regular. Modulo vn1v_{n-1}, the leading term becomes vnv_{n} which is a unit since E/vn1E/v_{n-1} is of height nn. Hence by Remark 5.15 we find that the pushforward of the element ψpk(n1)\psi_{p^{k}}^{(n-1)} to EeE_{*}\llbracket e\rrbracket, i.e., the element i=0ai[pk1]F(e)\sum_{i=0}^{\infty}a_{i}[p^{k-1}]_{F}(e) for i=0aiei=[p]F(e)/epn1\sum_{i=0}^{\infty}a_{i}e^{i}=[p]_{F}(e)/e^{p^{n-1}}, generates the kernel of the composite

EeEe/[pk]F(e)=(E¯Cpk)(ΦCpkE¯).E_{*}\llbracket e\rrbracket\to E_{*}\llbracket e\rrbracket/[p^{k}]_{F}(e)=(\underline{E}^{C_{p^{k}}})_{*}\to(\Phi^{C{p^{k}}}\underline{E})_{*}.

The leading term of ψpk(n1)\psi_{p^{k}}^{(n-1)} equals vn1v_{n-1}, which is not a unit. Hence ΦCpkE¯\Phi^{C{p^{k}}}\underline{E} is non-trivial and since In1I_{n-1} acts trivially it must be of height n1\geq n-1. Furthermore, reducing modulo vn1v_{n-1} the leading coefficient of ψpk(n1)\psi_{p^{k}}^{(n-1)} equals a power of vnv_{n}. Since vnv_{n} is a unit modulo vn1v_{n-1}, it follows that so is ψpk(n1)\psi_{p^{k}}^{(n-1)}. Therefore (ΦCpkE¯)/vn1(\Phi^{C{p^{k}}}\underline{E})/v_{n-1} is trivial, and hence (ΦCpkE¯)(\Phi^{C{p^{k}}}\underline{E}) is of height exactly n1n-1. ∎

Example 8.7.

Given any Landweber exact complex oriented ring spectrum EE of height nn, its quotient E/In1E/I_{n-1} satisfies the assumptions of the previous proposition. It follows that also in this case the height of ΦCpkE\Phi^{C_{p^{k}}}E equals n1n-1. For example this applies to Johnson-Wilson spectra, to MU[vn1]MU[v_{n}^{-1}] or to Morava EE-theories.

Remark 8.8.

Another approach to blue-shift questions like the above is via the chromatic Nullstellensatz of [BurklundSchlankYuan]: Let EE be an EE_{\infty}-ring of height nn, i.e. K(n)EK(n)_{*}E is non-trivial and K(n+1)EK(n+1)_{*}E is trivial. If EE is complex-oriented, this is equivalent to our previous definition of height, i.e. to the vanishing of EE-homology detecting compact spectra of type n+1\geq n+1. We claim that the Tate spectra EtCp=ΦCpkE¯E^{tC_{p}}=\Phi^{C_{p^{k}}}\underline{E} are of height n1n-1. Indeed, by the results of [BurklundSchlankYuan] any such theory EE maps to Morava EE-theory EnE_{n}, and ΦCpkEn¯\Phi^{C_{p^{k}}}\underline{E_{n}} is non-trivial K(n1)K(n-1)-locally by [BarthelHausmannNaumannNikolausNoelStapleton, Theorem 3.4] or the argument above. Furthermore, ΦCpkE¯\Phi^{C_{p^{k}}}\underline{E} must be trivial K(n)K(n)-locally since the converse would contradict the existence of compact CpkC_{p^{k}}-spectra XX of underlying type nn and CpkC_{p^{k}}-geometric fixed points of type n1n-1, by the same argument referenced above.

Similarly, for (2) one needs to find a complex oriented theory EE of height nn such that Φ𝐓E¯\Phi^{\mathbf{T}}\underline{E} is also of height nn. This is more elementary and satisfied by any pp-local complex oriented theory EE of height nn. To see this, note again that the Euler classes ene_{n} are given by the nn-series [n]F(e)Ee[n]_{F}(e)\in E_{*}\llbracket e\rrbracket. Writing n=pkmn=p^{k}\cdot m with mm coprime to pp, we find that [n]F(e)[n]_{F}(e) is a unit multiple of [pk]F(e)[p^{k}]_{F}(e). Modulo InI_{n}, for k>0k>0 the leading term of [pk]F(e)[p^{k}]_{F}(e) is a power of vnv_{n}, which by assumption is a unit in E/InE_{*}/I_{n}. It follows that, modulo InI_{n}, the coefficients of Φ𝐓E¯\Phi^{\mathbf{T}}\underline{E} are given by E/In((e))E_{*}/I_{n}((e)), which is always non-trivial when E/InE_{*}/I_{n} is.

8.2 Change of groups and the structure of FGA\mathcal{M}_{FG}^{A}

For any AA-spectrum XX, the groups MU2A(X)MU_{2*}^{A}(X) come equipped with the structure of a graded (LA,SA)(L_{A},S_{A})-comodule. Since the stack FGA\mathcal{M}_{FG}^{A} is the stack associated to this graded Hopf algebroid, we obtain an associated quasi-coherent sheaf A(X)\mathcal{F}^{A}(X) on FGA\mathcal{M}_{FG}^{A} (see [MeierOzornova, Proposition 4.3]). This is the 0-th graded piece of a QCoh(FGA)\operatorname{QCoh}(\mathcal{M}_{FG}^{A})-valued homology theory on AA-spectra, whose ii-th piece iA(X)\mathcal{F}_{i}^{A}(X) is given by MU2+iA(X)MU_{2*+i}^{A}(X). We end this section by a closer look upon how the structure of FGA\mathcal{M}_{FG}^{A} relates to this homology theory.

Recall from 3.11 that for a closed subgroup BAB\subseteq A, there is an open immersion j:FGA/BFGAj\colon\mathcal{M}_{FG}^{A/B}\to\mathcal{M}_{FG}^{A} and a closed immersion i:FGBFGAi\colon\mathcal{M}_{FG}^{B}\to\mathcal{M}_{FG}^{A}. We obtain corresponding adjunctions

QCoh(FGA){\operatorname{QCoh}(\mathcal{M}_{FG}^{A})}QCoh(FGA/B){\operatorname{QCoh}(\mathcal{M}_{FG}^{A/B})}j\scriptstyle{j^{*}}j\scriptstyle{j_{*}} and QCoh(FGA){\operatorname{QCoh}(\mathcal{M}_{FG}^{A})}QCoh(FGB).{\operatorname{QCoh}(\mathcal{M}_{FG}^{B}).}i\scriptstyle{i^{*}}i\scriptstyle{i_{*}}

Believing that the structure of FGA\mathcal{M}_{FG}^{A} dictates the structure of the \infty-category SpA\operatorname{Sp}_{A} of AA-spectra, we expect a relation to the adjunctions

SpA{\operatorname{Sp}_{A}}SpA/B{\operatorname{Sp}_{A/B}}Φ¯B\scriptstyle{\underline{\Phi}^{B}}PA/B\scriptstyle{P^{*}_{A/B}}          and          SpA{\operatorname{Sp}_{A}}SpB.{\operatorname{Sp}_{B}.}resBA\scriptstyle{\operatorname{res}_{B}^{A}}coindBA\scriptstyle{\operatorname{coind}_{B}^{A}}

Here, denoting by [B]\mathcal{F}[B] the family of subgroups of AA not containing BB and by q:AA/Bq\colon A\to A/B the projection, we define Φ¯B(X)=(E~[B]X)B\underline{\Phi}^{B}(X)=(\widetilde{E}\mathcal{F}[B]\operatorname{\wedge}X)^{B} and PA/B(Y)=E~[B]qXP^{*}_{A/B}(Y)=\widetilde{E}\mathcal{F}[B]\operatorname{\wedge}q^{*}X. Note that the definition of Φ¯B\underline{\Phi}^{B} is made so that its underlying spectrum is ΦB\Phi^{B} and more generally ΦC/BΦ¯BΦC\Phi^{C/B}\underline{\Phi}^{B}\simeq\Phi^{C} for every BCAB\subseteq C\subseteq A. For more details on the first adjunction, see [LewisMaySteinberger, Section II.9], [HillPrimer, Section 4.1] and [MeierShiZeng, Section 2.2]. For the benefit of the reader, we give a brief sketch of its basic properties: The adjunction between qq^{*} and ()H(-)^{H} induces maps

idΦ¯BPA/B\displaystyle\operatorname{id}\to\underline{\Phi}^{B}P^{*}_{A/B}\qquad andPA/BΦ¯BE~[B](),\displaystyle\text{and}\qquad P^{*}_{A/B}\underline{\Phi}^{B}\to\widetilde{E}\mathcal{F}[B]\operatorname{\wedge}(-),
which can be checked to be equivalences on geometric fixed points. The inverses
Φ¯BPA/Bid\displaystyle\underline{\Phi}^{B}P^{*}_{A/B}\xrightarrow{\simeq}\operatorname{id}\qquad andidE~[B]()PA/BΦ¯B\displaystyle\text{and}\qquad\operatorname{id}\to\widetilde{E}\mathcal{F}[B]\operatorname{\wedge}(-)\xrightarrow{\simeq}P^{*}_{A/B}\underline{\Phi}^{B}

form the counit and unit of the adjunction. In particular, PA/BP^{*}_{A/B} is fully faithful and its image agrees with that of E~[B]\widetilde{E}\mathcal{F}[B]\operatorname{\wedge}. Since smashing with E~[H]\widetilde{E}\mathcal{F}[H] is idempotent and hence symmetric monoidal, PA/BP^{*}_{A/B} is symmetric monoidal as well and so is Φ¯B\underline{\Phi}^{B}.

Proposition 8.9.

The diagram

SpA{\operatorname{Sp}_{A}}SpA/B{\operatorname{Sp}_{A/B}}QCoh(FGA){\operatorname{QCoh}(\mathcal{M}_{FG}^{A})}QCoh(FGA/B){\operatorname{QCoh}(\mathcal{M}_{FG}^{A/B})}Φ¯B\scriptstyle{\underline{\Phi}^{B}}A\scriptstyle{\mathcal{F}^{A}}PA/B\scriptstyle{P^{*}_{A/B}}A/B\scriptstyle{\mathcal{F}^{A/B}}j\scriptstyle{j^{*}}j\scriptstyle{j_{*}}

commutes, i.e. there are natural isomorphisms jA(X)A/B(Φ¯BX)j^{*}\mathcal{F}^{A}(X)\cong\mathcal{F}^{A/B}(\underline{\Phi}^{B}X) for XSpAX\in\operatorname{Sp}_{A} and jA/B(Y)A(PA/BY)j_{*}\mathcal{F}^{A/B}(Y)\cong\mathcal{F}^{A}(P^{*}_{A/B}Y) for YSpA/BY\in\operatorname{Sp}_{A/B}.

Moreover, there is a natural isomorphism iB(Z)A(coindBAZ)i_{*}\mathcal{F}^{B}(Z)\cong\mathcal{F}^{A}(\operatorname{coind}_{B}^{A}Z) for ZSpBZ\in\operatorname{Sp}_{B}.

No such isomorphism can be expected for ii^{*} and resAB\operatorname{res}^{A}_{B} in general. One reason is that ii^{*} is not flat, but even a spectral sequence relating ii^{*} and resAB\operatorname{res}^{A}_{B} seems not to exist for AA not a torus for reasons related to Remark 8.2.

Before we prove the proposition, we need a lemma, in which we will use the Hopf algebroid (Φ¯BLA,Φ¯BSA)(\underline{\Phi}^{B}L_{A},\underline{\Phi}^{B}S_{A}). Here Φ¯BLA\underline{\Phi}^{B}L_{A} is obtained from LAL_{A} by inverting eVe_{V} for all Vim((A/B)A)V\notin\operatorname{im}((A/B)^{*}\to A^{*}), and the ring Φ¯BSA\underline{\Phi}^{B}S_{A} is defined as Φ¯BLALASALAΦ¯BLA\underline{\Phi}^{B}L_{A}\otimes_{L_{A}}S_{A}\otimes_{L_{A}}\underline{\Phi}^{B}L_{A}. This classifies strict isomorphism between equivariant formal group laws where the relevant Euler classes are invertible on source and target. Since the invertibility of Euler classes only depends on the underlying equivariant formal group and not on the choice of coordinate, this simplifies to SALAΦ¯BLAS_{A}\otimes_{L_{A}}\underline{\Phi}^{B}L_{A}.

Lemma 8.10.

For YSpBY\in\operatorname{Sp}_{B} and q:AA/Bq\colon A\to A/B the projection, there are natural isomorphisms

Φ¯BLALA(MUA)(qY)Φ¯BLALA(MUA)(PA/BY)(Φ¯BMUA)(Y)\underline{\Phi}^{B}L_{A}\otimes_{L_{A}}(MU_{A})_{*}(q^{*}Y)\xrightarrow{\cong}\underline{\Phi}^{B}L_{A}\otimes_{L_{A}}(MU_{A})_{*}(P^{*}_{A/B}Y)\xrightarrow{\cong}(\underline{\Phi}^{B}MU_{A})_{*}(Y)

and

Φ¯BLALA/B(MUA/B)(Y)(Φ¯BMUA)(Y)\underline{\Phi}^{B}L_{A}\otimes_{L_{A/B}}(MU_{A/B})_{*}(Y)\xrightarrow{\cong}(\underline{\Phi}^{B}MU_{A})_{*}(Y)

of graded (Φ¯BLA,Φ¯BSA)(\underline{\Phi}^{B}L_{A},\underline{\Phi}^{B}S_{A})-comodules, where the map LA/BΦ¯BLAL_{A/B}\to\underline{\Phi}^{B}L_{A} is defined as the composite LA/BqLAΦ¯BLAL_{A/B}\xrightarrow{q^{*}}L_{A}\to\underline{\Phi}^{B}L_{A}.

Proof.

A model for E~[B]\widetilde{E}\mathcal{F}[B] is given by SWS^{\infty W} for WW the sum of all characters V𝒱V\in\mathcal{V}, for 𝒱\mathcal{V} the set of characters VV not restricting to 11 in BB or, equivalently, Vim((A/B)A)V\notin\operatorname{im}((A/B)^{*}\to A^{*}). Indeed, WC=0W^{C}=0 for CAC\subseteq A if and only if BCB\subseteq C, as this is equivalent to none of the V𝒱V\in\mathcal{V} restricting to 11 in CC.

In other words, smashing with E~[B]\widetilde{E}\mathcal{F}[B] is the same as inverting all the maps S0SVS^{0}\to S^{V} for V𝒱V\in\mathcal{V}. For an MUAMU_{A}-module, this is equivalent to inverting eVe_{V} for V𝒱V\in\mathcal{V}.

We have

Φ¯BLALA(MUA)(qY)\displaystyle\underline{\Phi}^{B}L_{A}\otimes_{L_{A}}(MU_{A})_{*}(q^{*}Y) (MUA)(qY)[eV1:V𝒱]\displaystyle\cong(MU_{A})_{*}(q^{*}Y)[e_{V}^{-1}:V\in\mathcal{V}]
π(MUAqYE~[B])A\displaystyle\cong\pi_{*}(MU_{A}\wedge q^{*}Y\wedge\widetilde{E}\mathcal{F}[B])^{A}
π(Φ¯B(MUAPA/BY))A/B\displaystyle\cong\pi_{*}(\underline{\Phi}^{B}(MU_{A}\wedge P^{*}_{A/B}Y))^{A/B}
π(Φ¯BMUAY)A/B\displaystyle\cong\pi_{*}(\underline{\Phi}^{B}MU_{A}\wedge Y)^{A/B}
(Φ¯BMUA)(Y).\displaystyle\cong(\underline{\Phi}^{B}MU_{A})_{*}(Y).

Replacing qYq^{*}Y by PA/BYP^{*}_{A/B}Y in the chain of isomorphisms above yields a similar chain of isomorphisms. All the isomorphisms are isomorphisms of comodules since the isomorphisms are natural in the MUAMU_{A}-variable and we can plug into this variable the left and right unit MUAMUAMUAMU_{A}\to MU_{A}\wedge MU_{A}.

To construct the second isomorphism, note that as part of the global structure of equivariant MUMU, there is a ring map qMUA/BMUAq^{*}MU_{A/B}\to MU_{A} (cf. [LinskensNardinPol]). Applying Φ¯B=(E~[H])B\underline{\Phi}^{B}=(-\wedge\widetilde{E}\mathcal{F}[H])^{B} yields a ring map MUA/BΦ¯BMUAMU_{A/B}\to\underline{\Phi}^{B}MU_{A}. This induces a morphism

Φ¯BLALA/B(MUA/B)(Y)(Φ¯BMUA)(Y).\underline{\Phi}^{B}L_{A}\otimes_{L_{A/B}}(MU_{A/B})^{*}(Y)\to(\underline{\Phi}^{B}MU_{A})^{*}(Y).

It is enough to show that this is an isomorphism for finite YY and hence for Y=(A/B)/(B/B)+A/B+Y=(A/B)/(B^{\prime}/B)_{+}\cong A/B^{\prime}_{+} for BBAB\subseteq B^{\prime}\subseteq A. In this case, the map becomes

Φ¯BLALA/BLB/B(Φ¯BMUA)(A/B+).\underline{\Phi}^{B}L_{A}\otimes_{L_{A/B}}L_{B^{\prime}/B}\to(\underline{\Phi}^{B}MU_{A})^{*}(A/B^{\prime}_{+}). (8.11)

The natural map

LALA/BLB/BLBL_{A}\otimes_{L_{A/B}}L_{B^{\prime}/B}\to L_{B^{\prime}}

is an isomorphism since LALBL_{A}\to L_{B^{\prime}} is a surjection with kernel generated by the Euler classes eVe_{V} for those VAV\in A^{*} restricting trivially to BB^{\prime}; these are exactly the images of the Euler classes eWe_{W} for those W(A/B)W\in(A/B)^{*} restricting trivially to B/BB^{\prime}/B.

Thus, (8.11) becomes

Φ¯BLALAMUA(A/B+)Φ¯BLALALB(Φ¯BMUA)(A/B+),\underline{\Phi}^{B}L_{A}\otimes_{L_{A}}MU_{A}^{*}(A/B^{\prime}_{+})\cong\underline{\Phi}^{B}L_{A}\otimes_{L_{A}}L_{B^{\prime}}\to(\underline{\Phi}^{B}MU_{A})^{*}(A/B^{\prime}_{+}),

which is an isomorphism by the first part. Similar to the first part, all isomorphisms are isomorphisms of comodules again. ∎

Proof of 8.9: .

We establish first the isomorphism jA(X)A/B(Φ¯BX)j^{*}\mathcal{F}^{A}(X)\cong\mathcal{F}^{A/B}(\underline{\Phi}^{B}X) for AA-spectra XX. Consider the commutative diagram

SpecLA\textstyle{\operatorname{Spec}L_{A}\ignorespaces\ignorespaces\ignorespaces\ignorespaces}SpecΦ¯BLA\textstyle{\operatorname{Spec}\underline{\Phi}^{B}L_{A}\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}φ\scriptstyle{\varphi}SpecLA/B\textstyle{\operatorname{Spec}L_{A/B}\ignorespaces\ignorespaces\ignorespaces\ignorespaces}FGA\textstyle{\mathcal{M}_{FG}^{A}\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}\pigpenfontLFGA/B,\textstyle{\mathcal{M}_{FG}^{A/B}\ignorespaces\ignorespaces\ignorespaces\ignorespaces,}j\scriptstyle{j}

where the down-right arrow comes from applying qq_{*} to an AA-equivariant formal group classified by a morphism to SpecΦ¯BLA\operatorname{Spec}\underline{\Phi}^{B}L_{A}, and the right-pointing horizontal morphisms come from the composition LA/BqLAΦ¯BLAL_{A/B}\xrightarrow{q^{*}}L_{A}\to\underline{\Phi}^{B}L_{A}. The square is a pullback square by 3.11. Thus φ\varphi is faithfully flat and hence φ\varphi^{*} induces an equivalence of QCoh(FGA/B)\operatorname{QCoh}(\mathcal{M}_{FG}^{A/B}) to graded (Φ¯BLA,Φ¯BSA)(\underline{\Phi}^{B}L_{A},\underline{\Phi}^{B}S_{A})-comodules.

The comodule corresponding to jXAj^{*}\mathcal{F}_{X}^{A} is Φ¯BLALA(MUA)2X\underline{\Phi}^{B}L_{A}\otimes_{L_{A}}(MU_{A})_{2*}X. As in (the proof of) the first isomorphism in Lemma 8.10, we observe that this is isomorphic to

Φ¯BLALA(MUA)2(XE~[B])\displaystyle\underline{\Phi}^{B}L_{A}\otimes_{L_{A}}(MU_{A})_{2*}(X\wedge\widetilde{E}\mathcal{F}[B]) Φ¯BLALA(MUA)2(PA/BΦ¯BX)\displaystyle\cong\underline{\Phi}^{B}L_{A}\otimes_{L_{A}}(MU_{A})_{2*}(P^{*}_{A/B}\underline{\Phi}^{B}X)
(Φ¯BMUA)2(Φ¯BX).\displaystyle\cong(\underline{\Phi}^{B}MU_{A})_{2*}(\underline{\Phi}^{B}X).

By the second isomorphism in Lemma 8.10, this is isomorphic to the comodule corresponding to Φ¯BXA/B\mathcal{F}_{\underline{\Phi}^{B}X}^{A/B}, i.e. to Φ¯BLALA/B(MUA/B)2(Φ¯BX)\underline{\Phi}^{B}L_{A}\otimes_{L_{A/B}}(MU_{A/B})_{2*}(\underline{\Phi}^{B}X). This establishes the first claimed isomorphism of sheaves.

Since the counit Φ¯BPA/BidSpA/B\underline{\Phi}^{B}P^{*}_{A/B}\to\operatorname{id}_{\operatorname{Sp}_{A/B}} is an equivalence, we can assume for the proof of jA/B(Y)A(PA/BY)j_{*}\mathcal{F}^{A/B}(Y)\cong\mathcal{F}^{A}(P^{*}_{A/B}Y) for YSpA/BY\in\operatorname{Sp}_{A/B} that Y=Φ¯BXY=\underline{\Phi}^{B}X for some XSpAX\in\operatorname{Sp}_{A} and we obtain from the first isomorphism in this case a natural isomorphism

jjA/B(Y)jjjA(X)jA(X)A/