Invariant prime ideals in equivariant Lazard rings
Abstract
Let be an abelian compact Lie group. In this paper we compute the spectrum of invariant prime ideals of the -equivariant Lazard ring, or equivalently the spectrum of points of the moduli stack of -equivariant formal groups. We further show that this spectrum is homeomorphic to the Balmer spectrum of compact -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 with the Lazard ring [QuillenElementary]. Around the same time, tom Dieck introduced for every compact Lie group in [tomDieckBordism] an equivariant analog of , the homotopical -equivariant complex bordism . Letting be abelian, Cole, Greenlees and Kriz [CGK] found many years later the correct notion of an -equivariant formal group law. Recently, the first author generalized work of Hanke–Wiemeler [HankeWiemeler] and showed that is indeed the universal ring for -equivariant formal group laws, thus establishing an equivariant analog of Quillen’s theorem for the equivariant Lazard ring .
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 in terms of the height . Thus, the points of the moduli stack of formal groups correspond to pairs with if and only if . Hopkins and Smith [H-S98] showed that the same classification pertains to thick subcategories of finite spectra: Given a finite spectrum , its -homology defines a coherent sheaf over . Taking the support of in the Zariski spectrum 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 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 be an abelian compact Lie group. Then the spectrum of points of the moduli space of -equivariant formal groups is homeomorphic to the Balmer spectrum of finite -spectra, with the comparison map induced by a support theory based on complex bordism homology .
This establishes 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 -spectra has been computed completely in the papers [BalmerSanders] (the case ), [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 and decompose as the disjoint union of one copy of for every closed subgroup of . Thus, the correct notion of height of an -equivariant formal group over a field of characteristic consists of a pair: a height of a non-equivariant formal group and a closed subgroup such that is induced along the zig-zag .
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 of a finite -spectrum .
We will detail our results below in the language of invariant prime ideals. Crucially, we exhibit equivariant lifts of the classical and show that they provide in many cases a sequence of generators of invariant prime ideals. The non-equivariant 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 , as we now explain. Similarly to the non-equivariant case, is the ground ring of a flat Hopf algebroid , classifying -equivariant formal group laws and their strict isomorphisms. The associated stack is the moduli stack of -equivariant formal groups. Hence, the category of graded -comodules is equivalent to the category of quasi-coherent sheaves over .
Recall that an ideal of is called invariant (in the sense of Hopf algebroids) if it is a sub-comodule, i.e., if for the left and right unit . Every invariant prime ideal gives rise to a point of the moduli stack of prime ideals via the quotient field of . This defines a map from the set of invariant prime ideals to , 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 for a prime and (with being the -ideal for all ).
To describe the invariant prime ideals in the equivariant case we recall that contains universal Euler classes for all characters , 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 and a closed subgroup of we obtain an invariant prime ideal as the kernel of the composite
Here, is defined as the localization of away from all the Euler classes of non-trivial characters for . The ring is an algebraic version of geometric fixed points and indeed agrees with the coefficient ring of the -geometric fixed points of .
Theorem 1.2 (Theorem 4.7).
For every abelian compact Lie group the map
is a bijection.
Here, is the set of all closed subgroups of . Hence, as for the Balmer spectrum, the invariant prime ideals of decompose as a set as one copy of for every closed subgroup of . 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 if and only if
-
1.
is a subgroup of .
-
2.
or (in which case ), the components are a -group and .
Comparing with [BarthelGreenleesHausmann] we see that these correspond precisely to the inclusions in the Balmer spectrum, but with roles reversed: There is an inclusion if and only if there is an inclusion . Here, are the thick subcategories of with being Morava K-theory at the prime .
To show that indeed includes into when conditions and are satisfied, one can reduce to the case 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 of the elements which exhibit maximal height shifts (Definition 5.21, 5.24). Roughly speaking, is of height at the top group (i.e., it lies in the ideal ) while it is of height at the trivial group (i.e., it lies in the ideal but not in ). This is the algebraic analog of the existence of finite -spectra of underlying type whose -geometric fixed points are rationally non-trivial as in [BarthelHausmannNaumannNikolausNoelStapleton] and [KuhnLloydChromatic]. More precisely, is canonically defined only modulo a certain smaller ideal (analogously to only being defined uniquely up to the ideal ). More details are given in Section 5.3.
We further show that - at least over elementary abelian -groups - the elements give rise to generators of the invariant prime ideals:
Theorem 1.4 (6.1).
For all primes and the elements
generate the ideal . Here, denotes the projection to the first coordinates.
Suitable restrictions of the then form generators for the ideals at higher height , see Section 6. We emphasize that in contrast to the non-equivariant situation, the sequence of the is not a regular sequence. In fact, since 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 ). The torsion in the ring is closely linked to the torsion in the group of characters . Hence one might hope that is generated by a regular sequence whenever 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 is infinite, the set of closed subgroups 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 .
Theorem 1.5 (7.5).
The Zariski topology on has as basis the closed subsets which are
-
(i)
closed under upward inclusions, i.e., if and , then , and which
-
(ii)
are locally constant on in the sense that if there exists a neighborhood of such that for all .
Comparing with [BarthelGreenleesHausmann], we see that this description precisely matches the computation of the topology on the Balmer spectrum, with replaced by . Hence the assignment
yields a homeomorphism from (and hence ) to . In the last section we explain that this comparison map can be obtained less ad hoc via -homology:
Theorem 1.6 (Section 8.1).
Let be a finite -spectrum and an invariant prime ideal. Then the localization is non-trivial if and only if the -geometric fixed points are of type at , i.e, if and only if is in the Balmer support of .
This shows that
defines a universal support theory on finite -spectra and thus a homeomorphism . Here, is the Mackey functor recording for all closed subgroups of , and is defined as the set of invariant prime ideals at which the localization of this Mackey functor is non-trivial.
Our proof that is a support theory is independent of [BarthelHausmannNaumannNikolausNoelStapleton] and [BarthelGreenleesHausmann]; this already provides a continuous bijection , 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 as defining quasi-coherent sheaves on for every closed ; the union of the supports of these sheaves agrees with under the homeomorphism . To show this, we establish in 8.9 how the adjunction between and defined by geometric fixed points and pullback corresponds on the algebraic level to pullback and pushforward along the open immersion from 3.11.
1.2 Acknowledgements
We thank Robert Burklund, Jeremy Hahn and Allen Yuan for their interest and enlightening discussions about the elements 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 . 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 -algebra is a complete linearly topologized commutative -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 -algebras, indexed on a countable directed set with surjective transition maps. For us, the category of formal -schemes is the opposite of that of formal -algebras. It can be viewed as the category of ind-objects in affine -schemes, indexed by a countable directed set with closed immersions as transition maps. We will sometimes use the notation or for the formal -scheme associated to a formal -algebra , and for the formal -algebra associated to a formal -scheme . The product on affine -schemes induces one on formal -schemes, and this corresponds to the completed tensor product on formal -algebras.
We set . Given a countable set , we view as a formal scheme, namely as the colimit over all with finite. This corresponds to giving the product topology. If we just write , we will apply this construction to instead of .
For a compact Lie group , we will denote by its Pontryagin dual, which is always a discrete group. We use for the unit element in .
Definition 2.1.
Given a compact abelian Lie group , an -equivariant formal group over consists of a commutative group object in formal -schemes together with a group homomorphism of formal -schemes satisfying the following two conditions:
-
1.
For the composite , the augmentation ideal of the induced map is fpqc-locally on a free -module of rank , where .
-
2.
The topology on is generated by products of the ideals for and .
Remark 2.2.
In the above definition, one can easily replace by an arbitrary quasi-compact scheme , with formal -schemes being a suitable subcategory of the ind-category of , the category of schemes affine over .
Our definition differs in two aspects from that put forward in [StricklandMulti, Definition 2.15]. First, Strickland restricts to finite . Second, Strickland asks to be free of rank 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 a free -module of rank 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 -equivariant group. The category of -equivariant formal groups embeds into that of -equivariant groups and this inclusion has a right adjoint, called completion. Concretely, this replaces in the notation in Definition 2.1 by the formal -algebra . We will only use -equivariant groups to complete them to -equivariant formal groups.
Spelling out what we get in a more algebraic language if we fix a trivialization of the augmentation ideal gives us the notion of an equivariant formal group law.
Definition 2.4.
An -equivariant formal group law over is a quadruple
of a formal -algebra , a continuous comultiplication , a map of -algebras and an orientation , such that
-
(i)
the comultiplication is a map of -algebras which is cocommutative, coassociative and counital for the augmentation ,
-
(ii)
the map is compatible with the coproduct, and the topology on is generated by finite products of the kernels of the component functions for , and
-
(iii)
the element is regular and generates the kernel of .
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 .
Lemma 2.6.
An -equivariant formal group over together with an -linear isomorphism of the augmentation ideal is equivalent datum to an -equivariant formal group law over .
Proof.
The maps and are induced by the multiplication on and , respectively. The element corresponds to the trivialization of . ∎
Given any equivariant formal group over , we obtain for every a morphism . If , this corresponds to the morphism . Moreover, composed with left multiplication defines an -action on ; for every this gives a map . In terms of the data of an equivariant formal group law , this can explicitly be written as
Given , we set
which generates the kernel of . If is trivial, we have . We want to describe an analog for general . A complete flag for is a sequence of characters such that every character appears infinitely often. Given such a flag and , we set
Then every element of can be written uniquely as
(2.7) |
for coefficients [Hau, Section 2.2]. Hence, as a -module, is isomorphic to a countable infinite product of copies of .
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 -equivariant formal group laws and is a pair of maps and which are compatible with both the comultiplications and the augmentations and which send to .
This leads to a category -FGL of -equivariant formal group laws. In [CGK] it is shown that this category has an initial object , the ground ring of which is called the -equivariant Lazard ring and denoted . In fact, the category of -equivariant formal group laws is equivalent to the category of commutative rings under . To discuss this, note first that the forgetful functor
into the category of commutative rings is cofibered in groupoids. Concretely this boils down to the following two observations:
-
•
Every morphism of -equivariant formal group laws whose first component is the identity map is an isomorphism. One observes indeed that the diagram
obtained from Eq. 2.7 commutes. We call such an isomorphism living over the identity a very strict isomorphism between -equivariant formal group laws, in order to distinguish from other kinds of isomorphisms.
-
•
Given a morphism and an -equivariant formal group law over , one can define a pushforward over with the usual universal property. Its underlying -algebra is given by a completion of , where is the underlying -algebra of (cf. [GreenleesFGL, Section 2.E], [Hau, Section 2.3]).
Note that the only automorphism of an -equivariant formal group law over which is also a very strict isomorphism is the identity map. Hence, given two -equivariant formal group laws over , 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 -equivariant formal group laws, and we will often do so.
Now, given a map there is an induced -equivariant formal group law over obtained by pushing forward the universal -equivariant formal group law. Given an -equivariant FGL over , we can apply this to the first component of the unique map of -equivariant formal group laws . The resulting morphism is necessarily a very strict isomorphism. So, as claimed above, we obtain:
Corollary 2.9.
The functor
from commutative -algebras is an equivalence of categories. An inverse is given by sending an -equivariant formal group law to the first component of the unique morphism .
Remark 2.10.
The above proof is an instance of a general characterization of initial objects in categories cofibered in groupoids , namely that pushforward defines an equivalence of with .
Remark 2.11.
Non-equivariantly the -algebra is often fixed to be the power series ring 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 . 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 ‘-equivariant formal group laws are represented by the -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 -equivariant formal groups in .
Definition 2.12.
Let be a group homomorphism and let be a -equivariant formal group. We define the corestriction to be the -equivariant formal group, which is the completion of the -equivariant group
Proposition 2.13.
For every injective group homomorphism , the functor from -equivariant formal groups to -equivariant formal groups is fully faithful. The essential image consists of those -equivariant formal groups where the homomorphism from factors through .
Proof.
For every -equivariant formal group , by construction is the same group object as in formal schemes with the structure morphism (since 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 -equivariant formal group law and a group homomorphism , there is an induced -equivariant formal group law over the same ring , given by completing at products of the ideals for those which are in the image of . This defines a functor from -equivariant formal group laws to -equivariant formal group laws, which induces a map on Lazard rings. Hence we obtain a functor
which we call the global Lazard ring. As shown in [Hau], is isomorphic to and our map 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 be a surjective group homomorphism and let be a -equivariant group. We define the coinduction to be the -equivariant group , where the target denotes the quotient of by the antidiagonal -action.
Lemma 2.16.
For a surjective group homomorphism and a -equivariant formal group, is an -equivariant formal group, i.e. needs no additional completion.
Proof.
If is a -equivariant formal group, then
The -algebra is isomorphic to . As products of complete rings are complete, the claim follows. ∎
Proposition 2.17.
Let be a surjective group homomorphism. As functors between -equivariant formal groups and -equivariant formal groups, is the left adjoint of .
Proof.
For an -equivariant group , define as . Then and are adjoints between -equivariant groups and -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 -equivariant formal group law over , we can define Euler classes in . Recall that for , we set . The corresponding Euler class is
In terms of the associated equivariant formal group , we have
with and being the composite . This implies:
Lemma 2.18.
For a given equivariant formal group law with notation as above:
-
1.
The Euler class is invertible iff .
-
2.
The Euler class is zero iff .
Thus, the vanishing or invertibility of Euler classes does not depend on chosen coordinates. This allows us to generalize these concepts to -equivariant formal groups in the following way:
Definition 2.19.
For an -equivariant formal group , we say that the Euler class is invertible if and that is zero if .
Informally, is invertible if and only if the images of and in are disjoint.
Example 2.20.
Let be the multiplicative group over . We choose the group homomorphism picking the units in . This defines the structure of a -equivariant group, and its completion is a -equivariant formal group we call . Let be the unique non-trivial character. One computes
Thus, , 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 -equivariant multiplicative formal group (given by a completion of ) along the map classifying .
We will use several times the following lemma, taken from [Hau, Corollary 2.8] and the explanation thereafter:
Lemma 2.21.
Let be a subgroup. Then the restriction map is surjective, with kernel generated by the Euler classes where is running over a generating set of .
Similarly, the following holds on the level of equivariant formal groups.
Proposition 2.22.
Let be an -equivariant formal group.
-
1.
Let be a surjective group homomorphism. Assume that for , the Euler class is invertible. Then is an isomorphism.
-
2.
Let be an injective group homomorphism. Assume that for , we have . Then is in the essential image of .
Proof.
Let . Fixing a coordinate Zariski-locally and choosing a complete flag, is defined by the directed system and each of these terms has underlying space .
In the first item, Lemma 2.18 implies that and intersect each other in only if . Thus, the underlying space of decomposes into closed subspaces , of which only finitely many are non-empty. This induces decompositions of the schemes and we obtain thus an -equivariant isomorphism on the level of formal schemes.
By construction, decomposes in the same way. On the unit copy, the map is an isomorphism since is one by 2.17. For the other copies, this follows by the -equivariance of the map .
For the second item: by definition, the structure morphism of factors as . 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 does not have a canonical coordinate; we will talk more about it in 2.25.
Corollary 2.23.
Let be a surjective group homomorphism. Then is a fully faithful embedding from the category of -equivariant formal groups to that of -equivariant formal groups. The image consists of those -equivariant formal groups such that is invertible for not in the image of .
Proof.
If is a -equivariant formal group, then has the property that is invertible for not in the image of by construction. By the preceding proposition, invertibility of these Euler classes characterizes the image of . Moreover, is fully faithful since 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 be a field and be an -equivariant formal group over . Denote by the Pontryagin dual of the subgroup Denote further by the obvious morphisms. Then , where is the non-equivariant formal group defined by .
Proof.
Assume , i.e. . By the -action, this implies for every . Setting , this implies . Moreover, if , this implies . Thus, is indeed a subgroup.
Since is a field, iff is not invertible. By Proposition 2.22, we thus have for some non-equivariant formal group over . One computes . ∎
For any , let be the localization of away from all Euler classes for . Our results above let us compute quite explicitely (cf. [GreenleesFGL, Corollary 6.4] and [Hau, Proposition 2.11]).
Proposition 2.25.
There is an isomorphism of the form
Proof.
The ring classifies -equivariant formal group laws such that is invertible for all . By Proposition 2.22, , where is the completion of at the augmentation ideal. The structure of determines the structure of a non-equivariant formal group law on , where is the projection; in particular, we obtain an isomorphism , where is the image of . Vice versa, and are determined by . In particular, is the composite . Thus we see that is determined by , plus a choice of mapping to under . Such are exactly those elements such that and with . This gives the result. ∎
Remark 2.26.
The elements in the previous proposition already come from elements , which are uniquely defined by the property that
for all . Here, denotes the tautological character for the circle group . In particular, equals the Euler class . The elements are natural in the sense that for every group homomorphism . 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.
For every and every non-torsion character , the sequence
is exact. In particular, all Euler classes for non-torsion characters are non-zero divisors.
-
2.
For every , the complete -Hopf algebra of the universal -equivariant formal group law is canonically isomorphic to the completion
where is the kernel of the restriction map . More generally, is a completion of . Under this identification
-
(a)
the comultiplication and the augmentations are induced by the maps and on completion, and
-
(b)
the elements are the image of under the completion map, where is the tautological character.
-
(a)
The special case of (2) for the trivial group is particularly important: Completing at the kernel of the augmentation yields a power series ring on generators , where is the Euler class of the -th projection . Moreover, the isomorphism
is natural in , where
-
•
the functoriality of is induced by the global functoriality of , and
-
•
the functoriality of is through the universal formal group law over .
For example, for any the square
commutes, where is the th power map, and sends to the -series with respect to the universal formal group law. This implies that the Euler class for the th power map on is sent to the -series under the completion map. Similar statements hold for the collection of for fixed and varying .
We further record, where again denotes the identity character:
Corollary 2.28.
Let be an element whose image in is of the form . Then is uniquely divisible by and the quotient restricts to at the trivial group.
Proof.
First we note that is a regular element by Proposition 2.27. Hence division by is always unique if possible. By induction on the corollary then follows from the following facts, for an element and its image :
-
1.
is divisible by if and only if .
-
2.
The leading coefficient of is equal to .
-
3.
If and hence is divisible by , then . ∎
Note that the global functoriality makes every equivariant Lazard ring an algebra over the non-equivariant Lazard ring , and that all restriction maps are -algebra maps. We have the following:
Proposition 2.29 ([Hau], [Com96]).
is free as a module over , for every abelian compact Lie group .
Corollary 2.30.
The exact sequences of -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 and be the ideal generated by . Then for every and every non-torsion character , the sequence
is exact. In particular, the Euler classes for non-torsion characters remain non-zero divisors in . In the terminology of [Hau], the assignment
(together with the image of under ) 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 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 .
Definition 3.1.
A strict isomorphism between two -equivariant formal group laws
over the same ground ring is a -linear isomorphism
of Hopf algebras over such that is sent to modulo , where is the augmentation ideal in . Explicitly, this means that , and for some unit which augments to .
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 -equivariant formal group laws.
Let be the category of strict isomorphisms of -equivariant formal group laws. More precisely, its objects are quadruples consisting of a commutative ring , two -equivariant formal group laws and over and a strict isomorphism between them. Morphisms between two such quadruples and are given by a pair of morphisms and with the same underlying map , such that
commutes.
Proposition 3.2.
The category has an initial object, whose underlying ring is a localization of an infinite polynomial ring over the Lazard ring .
By Remark 2.10, we can equivalently say that the functor
is an equivalence of categories. Here, we use that 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 for an -equivariant formal group law . Recall that after choosing a complete flag we can write every element uniquely as
Given , we have
where is defined as the product . This is a finite sum, since if there exists some where . We obtain:
Lemma 3.3.
The augmentation is a linear combination of whose coefficients are products of Euler classes.
Moreover, we have the following:
Lemma 3.4.
Let be an -equivariant formal group law. Then an element is a unit if and only if is a unit in for all .
Proof.
See [Hau, Lemma 2.3]. ∎
Proof of Proposition 3.2.
For every object we can define a new -equivariant formal group law for which the components and agree with those of , but is defined as , the preimage of the coordinate of under . Then we obtain a new object which is isomorphic in to via the commutative square
Note that the two vertical maps are in fact very strict isomorphisms of -equivariant formal group laws. In summary, every object of is isomorphic to one of the form , where and are given by the same -augmented -Hopf algebra and the strict isomorphism is the identity. This is the same data as a single -equivariant formal group law together with a second choice of coordinate for some unit which augments to . Up to very strict isomorphism we can further assume that is the push-forward along a map . We claim that the functor sending an -equivariant formal group law to the set of all units augmenting to is representable by an -algebra . Indeed, a presentation for is given by
(3.5) |
where is a complete flag starting with and is the linear combination expressing in terms of the coefficients with respect to ; see Lemma 3.3 and Lemma 3.4. Here we use that the units augmenting to are precisely the elements of the form , with as in the end of Section 2.1.
Thus, the functor
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 . In fact, the only difference in the proof is that the equation for the unit is now of the form so that the presentation for the analog of becomes
Since , this ring is indeed isomorphic to
There are functors
sending a strict isomorphism to its source and target, respectively, as well as an ‘identity’ functor
an ‘inverse’ functor
and a ‘composition’
which restrict to the full subcategory of those objects of where the isomorphism is given by the identity. By representability we obtain analogous source and target maps , an identity map , an inverse map and a composition map .
Corollary 3.7.
The pair together with the above structure defines a Hopf algebroid i.e. a cogroupoid object in commutative rings. The associated functor
is equivalent to the one sending a commutative ring to the groupoid of -equivariant formal group laws over and strict isomorphisms between them.
Remark 3.8.
The Hopf algebroid 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 -tuple of coordinates’ (cf., [Hau, Section 5.8]) also carries a unique grading for every such that the coordinates have degree . The source-target maps , the identity map , the inverse map and the composition map all preserve this grading, since they are defined through their effect on the respective coordinates. Evaluating these maps at a group yields the Hopf algebroids , which hence inherit a grading compatible with all restriction and inflation maps.
-
•
The groupoid-valued functor represented by admits a -action from multiplying the coordinate of the equivariant formal group law by a unit and acting on strict isomorphisms by multiplying by . This corresponds to an even grading on , putting e.g. in degree .
-
•
By its interpretation as representing the groupoid of -equivariant formal group laws and isomorphisms between them (see Remark 3.6), obtains the structure of a Hopf algebroid. An element in is of degree if applying the composition map to gives , where is the composition map of .
One can show that this is the same grading coming from the isomorphism
from [Hau, Theorem E].
Given a strict isomorphism of -equivariant formal group laws and a group homomorphism , we obtain an induced strict isomorphism by completion. This assignment is compatible with composition of strict isomorphisms. Therefore, the functor from Section 2.3 extends to a functor
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 represents the functor sending a commutative ring to the groupoid of -equivariant formal group laws and strict isomorphisms between them. As discussed in Remark 3.8, is naturally a graded Hopf algebroid. On the other hand, we have discussed in Remark 3.6 the ungraded Hopf algebroid classifying -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 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 is the same as the stack associated to the ungraded Hopf algebroid . Equivalently, it is the quotient of the stack associated to by the -action induced by the grading.
Proposition 3.9.
Sending an -equivariant formal group law to its underlying -equivariant formal group defines an equivalence from the stack associated to the graded Hopf algebroid to , the (pseudo-)functor sending a commutative ring to the groupoid of -equivariant formal groups over it.
Proof.
It suffices to show that is an fpqc-stack. Indeed: denote the stack associated to the graded Hopf algebroid by ; equivalently, this is the stack associated to the ungraded Hopf algebroid . Since the augmentation ideal of every -equivariant formal group is by definition fpqc-locally trivial, comes after fpqc-base change from an -equivariant formal group law (see Lemma 2.6). Thus, is essentially surjective as a functor of stacks. Moreover, it is fully faithful since isomorphisms between -equivariant formal group laws are precisely isomorphisms of the underlying -equivariant formal groups. Thus, it remains to show that satisfies fqpc-descent on morphisms and objects.
Given two formal -schemes and , the functor
is an fpqc-sheaf. Indeed, if we view and as ind-objects and , we can rewrite this as and equalizers commute with filtered colimits in sets. This easily implies that satisfies fpqc-descent on morphisms.
Let denote the directed set of finite multi subsets of (i.e. elements can occur more than once), ordered by inclusion. Sending an -equivariant formal group over to the system , where runs over all finite products of the for , defines a functor from -equivariant formal groups over to .
Given a descent datum for -equivariant formal groups for the fpqc-cover , we thus obtain a descent datum for a -diagram of affine schemes (with closed immersions as transition maps), which descends thus to a -diagram of closed immersion in . This diagram defines a formal -scheme . By descent for morphisms between formal -schemes, obtains a group structure and also a group homomorphism . Conditions (1) and (2) for an -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 to .
Proposition 3.11.
Let be a subgroup of an abelian group . Denote by the inclusion and by the projection.
-
(i)
The functor induces an open immersion whose image is the common non-vanishing locus of the Euler classes for all .
-
(ii)
The functor induces a closed immersion , inducing an equivalence of to the common vanishing locus of the for all .
Proof.
The first part is a reformulation of 2.23. The immersion is open since it is open after pullback to .
For the second, note that by Proposition 2.22 every -equivariant formal group such that for all is of the form for a -equivariant formal group. Moreover, by construction, this vanishing of Euler classes is true for all -equivariant formal groups of the form and thus characterizes the image of . The substack given by this image is closed since it is closed after pullback to . Moreover, is fully faithful by 2.13. ∎
Remark 3.12.
With notation as in the preceding proposition, we have for every -equivariant formal group and for every -equivariant formal group . Thus, the substacks in the preceding proposition are retractive.
Example 3.13.
3.11 gives closed immersions of and into . 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 , 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 . The second, i.e. the closed immersion of , equals the vanishing locus of for . Its complement is an open substack equivalent to , the non-vanishing locus of all Euler classes of non-trivial characters.
For a non-cyclic group the situation is more complicated, and we cannot expect that the complement of the closed substack in can be expressed as a single open substack in general and vice versa. In general, the complement of in can be written as the union of the open substacks where runs over the minimal subgroups of not contained in . We have indicated the situation for in Fig. 3.
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 and the invariant prime ideals of . 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 , we can associate an ungraded Hopf algebroid to it (see e.g. [MeierOzornova, Section 4.1]).222Standard conventions force us to use as part of the notation of a general Hopf algebroid, while 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 . The stack associated to is by definition the stack associated to , i.e. the fpqc-stackification of the presheaf of groupoids represented by on the category of all schemes. We denote the resulting morphism by .
Definition 4.1.
Let be a Hopf algebroid with units and . An ideal is called invariant if . If is graded, we will assume that is also graded, i.e. generated by homogeneous elements.
It is easy to check that invariant ideals in a graded Hopf algebroid correspond exactly to graded subcomodules of and thus to ideal sheaves on . Here, we use that refines to an equivalence from quasi-coherent -modules to graded -comodules.
Following [LMB00, Section 5] and [STACKS, Tag 04XL] in the case of Artin stacks, we can associate a topological space to . To that purpose recall that is an open immersion if the pullback is an open immersion.
Definition 4.2.
For as above, define the underlying set of to consist of equivalence classes of morphisms for a field; the equivalence relation is generated by isomorphisms and , where is a field extension of . We call a subset of open if it is the image of for an open immersion .
Equivalently, we can characterize the opens as the images of those opens in that are invariant, i.e. have the same preimage along both the left and right unit . Indeed: by descent, an open immersion corresponds to an open immersion with an isomorphism over satisfying a cocycle condition. But the category of open immersions into some with isomorphisms over between them is equivalent to the discrete category of open subsets of , yielding the required equivalence. Since invariant opens in form a topology, we deduce that the opens in 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 is an affine scheme, where we get the usual topology.
Proposition 4.3.
Let be a graded Hopf algebroid with associated stack . Then:
-
1.
For every invariant prime ideal , the image of is closed in and for the generic point of (i.e. the point in corresponding to ) is generic (i.e. ).
-
2.
Mapping to defines an injection from the set of invariant prime ideals in to . Equipping with the subspace topology from , 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 , the set is invariant and hence the complement of defines an invariant open. Thus the image in is open. Moreover, . Thus is the complement of and thus closed.
Assume now that is an invariant prime ideal. Let be the image of the generic point of . Then and hence .
For the injectivity of , let be two invariant prime ideals with . By the first point, this implies that and hence . Thus, .
The continuity of follows from that of . An arbitrary closed set of is of the form for some ideal . Set , where the run over all invariant ideals containing . Then . If is a bijection, implies and this is closed by the first part. ∎
We warn the reader that in general, the preimage of an irreducible closed subset of won’t be irreducible in and thus does not correspond to an invariant prime ideal.
Proposition 4.4.
Let be a compact abelian Lie group. The underlying set of is in bijection with the product of the set of closed subgroups of and . For a given -equivariant formal group over a field, the point in is the pushforward of along , and the subgroup is Pontryagin dual to .
Proof.
Given a closed subgroup of and a non-equivariant formal group over a field , we obtain an -equivariant formal group via , where . By Proposition 2.24, every -equivariant formal group over a field is isomorphic to one of this form. Here, the projection is Pontryagin dual to the subgroup and thus is uniquely defined by . Moreover, necessarily . This implies that two -equivariant formal groups and defined over the same field are isomorphic if and only if they have the same associated closed subgroup and their completions and are isomorphic as non-equivariant formal groups, proving the claim. ∎
We recall that has been computed by Honda: its points are classified by a pair , where is the characteristic of the field and is the height of the formal group, with if . We will below always identify with such pairs.
Remark 4.5.
By 3.11, there is an open immersion for any closed subgroup and hence is homeomorphic to an open subset of . Given , the point corresponding to is for the projection. Indeed, for a -equivariant formal group over a field corresponding to , we have
The Pontryagin dual of is precisely since .
4.2 Invariant prime ideals of
As before, let be the localization of away from all Euler classes for and let . Denote by the ideal , i.e. the unique invariant prime ideal at height containing . We also include the case of . In this section, we will denote this ideal by to uniformize notation with respect to the residue characteristic.
Construction 4.6.
For every triple of a closed subgroup of , a prime and we define an invariant ideal as the preimage of along the composite map of Hopf algebroids
Note that the are indeed prime ideals since by 2.25 the quotient ring is of the form and hence an integral domain. Thus, the are prime as well.
To simplify notation, we will from now on often write instead of and likewise in similar situations.
Theorem 4.7.
The assignment
is a bijection. In other words, the ideals are pairwise different and constitute all the invariant prime ideals in .
The map from 4.3 is a homeomorphism.
Proof.
By 4.3, we know that injects into and the latter we computed to be as a set. Thus, it suffices to show that the element of associated to is precisely , where is a prime number if and if .
To spell this out concretely, let be the field of fractions of . We denote by the pushed forward -equivariant formal group law over and by the corresponding -equivariant formal group. Since hits the point corresponding to the prime ideal , the corresponding point in is represented by . By the classification in Proposition 4.4, we need to show three things:
-
1.
the set of such that in is precisely ,
-
2.
has characteristic (which is clear), and
-
3.
the pushforward along has height .
For the first, recall from Lemma 2.21 that is generated by the Euler classes for all . These Euler classes must vanish in since factors through . If is not in , then in and hence also in (as else ). Since injects into , the Euler class is actually nonzero in and hence invertible in . This shows the first point.
The pushforward is classified by the composite
The ideal maps to in and is hence contained in . Therefore, factors through and the height of is at least . It remains to show that is non-zero in and hence in . By definition of , the map factors over . We know by 2.25 that is an integral domain of the form . In particular, is non-trivial in and hence in . This shows that is of height as desired, which finishes the proof. ∎
Unraveling the definition, we obtain the following description of the ideal , where still denotes the kernel of the restriction map .
Lemma 4.8.
An element lies in if and only if there exists an -representation with , such that
Proof.
By definition, is the kernel of the composition
The composition is surjective with kernel , and the map inverts all Euler classes for -representations with . The product of Euler classes is an Euler class again. Hence, if is contained in , its image in must be annihilated by such an Euler class . We can extend to an -representation and find that is contained in , as desired.
For the opposite direction, if we assume that for some -representation with , then in . Since has trivial -fixed points, becomes invertible in and hence is taken to there. Therefore, is contained in . ∎
When is a torus, the ideal is easy to describe explicitly:
Corollary 4.9.
If is a torus, then .
Proof.
When is a torus, every non-trivial character is non-torsion and thus the map
is injective (Corollary 2.31). Hence, is equal to the kernel of
which is generated by and . ∎
We note the following useful corollary:
Corollary 4.10.
Let be a torus, and consider the augmentation ideal
i.e., the ideal generated by all the Euler classes. Then the intersection equals the -ideal.
For example, this shows that no element in the -equivariant Lazard ring is infinitely often divisible by the Euler class .
Proof.
Since is an invariant ideal of , so are all its powers and the intersection thereof. Moreover, equals the kernel of the completion map (cf. Section 2.5)
where is the rank of and the are the images of the Euler classes ranging through a basis of . Since is an integral domain, must be prime and hence an invariant prime ideal.
Therefore must be of the form (or rather its image under the projection ) for some subgroup and height . Note that is not contained in , hence in particular not in . This means that we must have . Moreover, given a character expressed in the chosen basis above, the image of under the completion map is given by
where is the universal formal group law pushed forward to . Since we assumed to be a finite height, the -series of is non-trivial whenever is non-zero. It follows that for any non-trivial the image of is non-trivial in . Hence contains no Euler class and we must have , i.e.,
as desired. ∎
5 Inclusions between invariant prime ideals
Non-equivariantly the ideals form ascending towers
essentially by definition. Except for the overlap at , there are no inclusions between the towers for different primes . For invariant prime ideals in the equivariant Lazard ring we saw that we have one tower
for every pair of a closed subgroup and prime . Again, there will be no interplay between the towers associated to different primes (except for the overlap at height ). However, there are additional inclusions connecting the towers for different subgroups at the same prime . 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 whenever . It is easy to see that maps non-trivially under whenever (since the target is free over by 2.25). Hence there cannot be an inclusion for and . Moreover, if we have . Hence we can reduce to studying containments between invariant prime ideals associated to the same prime .
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 . That is, we consider the -localized Lazard ring and denote its invariant prime ideals simply by , omitting the chosen prime . We further sometimes abbreviate to .
Hence our goal is to understand for which pairs of subgroups and natural numbers there is an inclusion
We will show the following:
Theorem 5.1.
There is an inclusion if and only if the following conditions are satisfied:
-
1.
is a subgroup of and is a -group.
-
2.
We have .
Hence, for example there are inclusions and , but is not contained in . 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 , the question whether is contained in does not depend on the ambient group , but only on and .
Theorem 5.1 can be interpreted as a statement about the heights of geometric fixed points of localizations of , in the following way. Recall from [HoveyStricklandComodules, Definition 4.1] that the height of an -algebra is the maximal such that ; equivalently, it is the minimal number such that . If there is no such , the height is understood to be infinite. If , the height is . Then we have the following:
Corollary 5.2.
Let be a closed subgroup, . If is not a -group, or if is a -group but , then the geometric fixed points are trivial. Otherwise, their height is given by
Proof.
We have if and only if there exists an element of not contained in which is mapped to under the geometric fixed point map. Since is defined precisely as the preimage of , this in turn is equivalent to not being contained in .
By Theorem 5.1 we know that if is not a -group or if is a -group but , then is not contained in . Since , this implies that the geometric fixed points are trivial.
If is a -group and , then the theorem tells us that and . Hence we have and , 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 if and only if in , the point corresponding to lies in the closure of the point corresponding to . Thus, Theorem 5.1 can be interpreted as a result about the topology of .
The proof of Theorem 5.1 takes up the remainder of this section.
5.1 Formal reduction to the -toral case
Our first step is the following:
Lemma 5.5.
If there is an inclusion , then is a subgroup of and .
Proof.
When is not a subgroup of we can choose a character which is trivial when restricted to but non-trivial when restricted to . Its Euler class then restricts to in , in particular it is contained in for all . On the other hand, its restriction to becomes an invertible element in the non-trivial ring and is hence non-trivial. Therefore is not an element of . It follows that cannot be contained in .
If , then is contained in but not in . This implies that (now thought of as an element of ) is contained in but not in , since is a non-trivial free module over by 2.25 Hence, again, cannot be contained in . ∎
Lemma 5.6.
Let be an inclusion of subgroups of and . Then there is an inclusion
if and only if there is an inclusion
if and only if there is an inclusion
Proof.
The first two statements are equivalent since the restriction map identifies with a quotient of , and the ideals and are the preimages of the ideals and under the quotient projection.
Phrased differently, the projection induces a closed embedding of the stack of -equivariant formal groups into the stack of -equivariant formal groups. On spectra, the image consists precisely of these with . This implies the desired equivalence by Remark 5.4.
For the second equivalence, recall the open embedding from Remark 4.5, sending the point to for every . 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 and and understand under what conditions there is an inclusion
or in other words whether the restriction map maps into the ideal .
Our next goal is to show that we can further reduce to the case where is a -group. For this we choose a prime and consider the Euler class , i.e., the pullback of along the th power map . The Euler class restricts to at the trivial group and is hence uniquely divisible by . We set to be the unique element satisfying (the reason for this choice of notation will become clear in Section 5.2). Under the completion map , the Euler class is sent to the -series of the universal formal group law. Hence, is sent to the quotient , whose leading coefficient equals . Since the restriction of any element in to the trivial group equals the leading coefficient of its image in , we see that We further set to be the restriction of , and find that it satisfies:
Here, denotes the restriction of the tautological character to . We are now ready to show:
Lemma 5.7.
If is not a -group, then there is no inclusion of the form .
Proof.
If is not a -group we can choose a surjection with a prime. Then the element satisfies the equation . Since is surjective, the character is non-trivial. Hence, is an element of and hence also of . Its restriction to the trivial group equals that of , which is and hence a unit in the (-localized) ring . In other words, is not an element of . Hence does not include into . ∎
Corollary 5.8.
If there is an inclusion , then is a subgroup of , and the quotient is -toral, i.e. a product of a -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 and stated there are satisfied we do have an inclusion . We start with the easiest case:
Lemma 5.9.
Let be an inclusion of subgroups of such that is a torus. Then there is an inclusion for all .
Proof.
We now turn to showing that there are inclusions . For this we show that is generated by plus one additional element which reduces to under the restriction map . We start with the case and recall again that the Euler class is sent to the -series under the completion map . Modulo , this -series is of the form + higher order terms. Hence 2.28 implies that there exists a unique element such that , and this element satisfies .
We then set for all , where is the multiplication-by- map on the circle. By functoriality, also restricts to at the trivial group. In fact, it already restricts to at . This is because is the kernel of and hence the restriction map factors through the trivial group. Applying to the defining equation for , we also obtain ; here and in the following, we will often abbreviate to .
Proposition 5.10.
For every and the element generates the kernel of
Hence, is generated by the regular sequence .
Corollary 5.11.
The ideal is generated by and the restriction of to .
Proof of Proposition 5.10.
Let be an element mapping to in , i.e. in the image of . By Lemma 4.8 we know that we have an equation of the form
(5.12) |
for some , and since by Lemma 2.21.
If is coprime to , then and become multiples of one another modulo : indeed, the corresponding characters in generate the same subgroup and thus Lemma 2.21 implies that and generate the same ideal in . It follows that Equation (5.12) gives rise to an equation
with only Euler classes for powers of appearing on the left hand side. We claim that must be divisible by the entire product , implying that is divisible by , as desired. To see this, recall that restricts to in for all . Hence we have that . Since is a regular element in by 2.29, this implies that is (uniquely) divisible by . This argument can be iterated by replacing by , and the statement follows. ∎
Remark 5.13.
One can show that more generally there exist elements for all uniquely determined by the equations
(5.14) |
where denotes the -adic valuation of a natural number. The element generates the kernel of
and its restriction to the trivial group is given by
We note also that every is prime, since the geometric fixed points are integral domains. The elements were previously considered in [Hau, Proposition 5.46], denoted there.
Remark 5.15.
The proof of Proposition 5.10 applies in a more general context. Let be a global group law in the sense of [Hau, Definition 5.1]. As the global Lazard ring is the initial global group law, there is a unique map . Assume that the map sends to , and that for every the Euler classes in are regular elements and is a regular element in (for example this is the case if is regular in and remains a regular element modulo ). Then the image of in generates the kernel of the composition
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 .
Proof.
By Corollary 5.11, is generated by and . Since is clearly contained in (even in ), we can reduce modulo and need to show that is taken to under the composition
But this is clear, since restricts to at the trivial group and lies inside . ∎
We now have all the ingredients to prove the ‘if’-direction in Theorem 5.1:
Corollary 5.17.
Let be an inclusion of subgroups of such that is -toral, and such that . Then there is an inclusion .
Proof.
By Lemma 5.6 we can assume that and that is the trivial subgroup. Hence, is a -toral group. Let denote the path component of the identity. We have by Lemma 5.9. Hence it suffices to show that is contained in . For this, making use of Lemma 5.6 once more, we can assume that and hence is a finite abelian -group. We write and choose a filtration of subgroups
such that every subquotient is a cyclic -group. By Corollary 5.16 and Lemma 5.6 we see that
The final inclusion follows from the assumption . 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 to construct an element which lies in the ideal and whose restriction to the trivial group lies outside of .
It turns out to be more natural to define such an element modulo a subideal of , namely the inflation of the ideal along the projection . That is, we construct an element
By Corollary 5.17 the restriction map takes into . Therefore, the restriction map takes into and we obtain an induced restriction map
We will see that, under this restriction, is sent to . Later in Section 6 we will show that in fact forms a generator of the quotient and that suitable inflations and restrictions of these elements generate all the invariant prime ideals at elementary abelian -groups.
We now turn to the construction of the element . We set and first define an element in the ring , whose restriction to then yields .
For every character , 2.27 yields a short exact sequence
where we use that identifies with the kernel of . The inflation map provides an -linear splitting of this exact sequence if we view as an -module via the same inflation map. Thus, we obtain a short exact sequence
(5.18) |
of -modules.
Remark 5.19.
We will see below in Section 6 that the ideal in generated by equals the invariant prime ideal . In particular, is again an integral domain. At this point it is only clear that is contained in .
We now consider the Euler class . We have , since is an elementary abelian -group and hence every character is -torsion. By exactness, it follows that is divisible by , for all . We want to define as the quotient of by the product over all the . For this we first need to check that the different are coprime. To understand this, we consider the following:
Lemma 5.20.
Let , and assume that restricts to a non-zero element under . Then is divisible by if and only if is divisible by .
Proof.
By exactness of (5.18), is divisible by if and only if in . Since is non-zero by assumption and is an integral domain, this is the case if and only if
which in turn is equivalent to being divisible by . ∎
Given a character , the restriction of along equals the non-trivial Euler class . Hence the lemma applies to being any product of Euler classes of the form with . We find that in the quotient , the Euler class is uniquely divisible by the product . To summarize:
Definition 5.21.
Let and . We define to be the unique element satisfying
and we set
Remark 5.22.
The element agrees with as defined in Section 5.1.
Remark 5.23.
Lemma 5.20 also applies in the ring itself (i.e., before quotienting by ) if we demand that is regular element, rather than just being non-zero. These two conditions are equivalent in since it is an integral domain. However, the Euler classes are not regular elements in , hence the lemma does not apply for a product of the . In fact is not divisible by before dividing out , even though it is divisible by each individual . One can see this by restricting from to : If was divisible by the product of all the , this would imply that its restriction was divisible by , since every restricts to . But is divisible by precisely once, since restricts to at the trivial group, cf. Section 5. It is for this reason that it is most natural to define and in this quotient. As we will see now, this matches nicely with the fact that is most naturally defined in the quotient .
Proposition 5.24.
-
1.
The element defines a class in the ideal , i.e., it is sent to under the map
-
2.
The restriction map
takes to .
Proof.
Part 1: The equation
reduces to the equation
in , where denotes the restriction of to . Note that each -character of the form is non-trivial. Hence forms Euler-power torsion and therefore maps to in the geometric fixed points.
Corollary 5.25.
If is a preimage of under the projection and is a subgroup of rank , then
Proof.
By the previous proposition, is an element of . As has rank , we know by Corollary 5.17 that must lie in .
If were an element of , then applying Corollary 5.17 to the inclusion of the trivial group into shows that is also an element of . This contradicts the fact that, modulo , we have . ∎
Corollary 5.26.
If is a -toral inclusion of subgroups of (i.e. is -toral) and , then does not include into .
Proof.
By Lemma 5.6 we can reduce to the case and . Let , and be a surjection. Let as in Corollary 5.25. Then, by the corollary, the restriction is an element of but not an element of . Therefore is an element of whose restriction to the trivial group is not contained in . In other words, is an element of but not an element of . Since by assumption we have and hence , this proves that does not include into . ∎
6 Generators for invariant prime ideals
In this section we show that over elementary abelian -groups the elements – 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.
For every torus and , the ideal is generated by the elements
where is the projection to the first factors.
-
2.
For every and every inclusion , the restriction map
maps surjectively onto .
Remark 6.2.
Implicit in the statement of the theorem is that each is well-defined modulo the ideal generated by . By definition, is an element of the quotient by the subideal generated by . Applying the theorem to rank and we see that this ideal is indeed generated by , 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 is generated by
where is any inclusion. The choice of inclusion will generally affect the resulting generators. For example, setting and the trivial group: If we choose to be the inclusion into the first factor, the composite becomes the identity. Hence we obtain that is generated by the elements and . If we alternatively use the inclusion into the second factor the composite becomes the constant map, yielding the generators and (i.e., the same ones as in Corollary 5.11, as equals ). Furthermore, it follows that generators for ideals of the form with a torus can be obtained as the union of Euler classes for a basis of together with a choice of generators for .
We prove Theorem 6.1 by induction on . Part 1 of the induction start is the statement that is the -ideal for any torus (4.9). For Part 2 we need to see that the restriction maps surjectively onto , which we know is generated by by 4.9. For we can consider the elements , which reduce to . It follows that the inflation of to via any choice of surjection gives an element of a quotient of which reduces to in . Since is generated by , the claim follows.
We now assume that Theorem 6.1 holds for an elementary abelian -group of rank and show it also holds for . For any we consider the surjection
with kernel generated by for the tautological -character pulled back to . We first claim that if is non-trivial, then the Euler class is a non-zero divisor in . To see this, we use that since is a torus we can apply the induction hypothesis to . In particular, we know by Part 1 that generates the ideal and hence is an integral domain. So we have to show that still acts regularly modulo . Since both Euler classes are regular, this is equivalent to showing that is regular modulo . We have an isomorphism
By Part 2 of the induction hypothesis, we know that restricts onto , hence the latter quotient identifies with . Again we know by the induction hypothesis that this quotient is an integral domain, and is clearly a non-trivial element. So the claim follows and we have shown that generates the Euler power torsion in (at height ) for characters inflated up from .
Hence to understand the full ideal it suffices to further divide by the Euler-power torsion for the remaining torsion characters in (there is no Euler-power torsion for non-torsion characters by 2.27). These torsion characters are of the form , where , is the restriction of to and . Furthermore we can assume that : Any has some power of the form and hence is a multiple of . Thus, is generated by Euler-power torsion for characters of the form .
Again it is beneficial to pass to the integral domain to understand the Euler-power torsion for these characters. We have the following:
Lemma 6.3.
Let be an elementary abelian -group, a torus and . Further let be an element satisfying
for some and collection of natural numbers . Then lies in the ideal generated by
Proof.
By applying to the defining property of (Definition 5.21) we obtain the equation
in . Restricting from to yields
(6.4) |
in the quotient . This uses that every character of extends to different characters of and that the restriction of lands in the ideal . For the rest of the proof we write for the element . With as in the statement of the lemma, we hence obtain an equation of the form
(6.5) |
and we need to show that is a multiple of . The Euler classes fit into short exact sequences of the form
analogously to Equation 5.18 above. As shown above, the induction hypothesis implies that the quotient is an integral domain. We know that restricts to at the trivial group. In particular it must restrict non-trivially under each . Hence the above short exact sequence together with the fact that is an integral domain implies: If an Euler class divides an element of the form , then divides . Applying this iteratively to Equation 6.5 (and using that is an integral domain by the induction hypothesis) we see that must divide the term . Dividing on both sides shows that is a multiple of , as desired. ∎
Corollary 6.6.
The quotient
is generated by the element
Proof.
We saw above that the quotient is generated by Euler-power torsion for characters of the form . An element of is such a torsion element if and only if it is the reduction of an element satisfying the conditions of the lemma. Since the reduction of equals , it follows that lies in the ideal generated by the latter.
As is Euler-power torsion itself, it hence forms a generator of . ∎
To finish the proof of Theorem 6.1: Setting in the corollary shows that generates the quotient . By the induction hypothesis we know that is generated by . Combined this proves Part 1 for the group .
For Part 2 and general , we first note that it suffices to show the statement for any choice of injection since any two only differ by postcomposition with an automorphism of . We can hence pick the canonical inclusion . By the induction hypothesis we know that surjects onto . From the diagram
we see that is contained in the image of the lower horizontal arrow. Furthermore, Corollary 6.6 implies that is generated by the restriction of an element of . This finishes the proof.
Remark 6.7.
Unlike the sequence , the sequence
isn’t regular. In fact, the ideal generated by these elements is precisely that of Euler-torsion.
This can be corrected by passing to a torus: The ideal is generated by the sequence . Here, each is the element of
(6.8) |
obtained as the inflation of along the projection to the first coordinate of . Since each successive quotient 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 is generated by a regular sequence of length .
7 The Zariski topology on the spectrum of invariant prime ideals
The goal of this section is to describe the Zariski topology on , or equivalently the topology on the space (Theorem 4.7). By definition, the closed subsets of are the subsets of the form
for some subset of . Hence we need to determine the collections of invariant prime ideals that arise as for some . We now fix a subset . Since a containment automatically implies for all , it suffices to understand – for every closed subgroup of – the smallest value of such that . In other words we need to determine the function
defined by
where we set . We note that with this definition the height function of the image of in (thought of as a one-element set) is constantly . This follows from the fact that is a free -module by 2.25; cf. also the proof of Theorem 4.7. Moreover, we have:
Example 7.1.
Let be a lift of . Then Corollary 5.25 implies that .
Our goal is to understand which functions arise as such height functions. In the previous section we showed that there are inclusions between invariant prime ideals associated to different subgroups of . These translate to conditions between the different values of : If is an inclusion of subgroups of such that is -toral, and is contained in , then is also contained in . In terms of the height function this translates to the inequality
This leads us to the following definition:
Definition 7.2.
A function is called admissible if it satisfies the inequality
for every -toral inclusion of closed subgroups of . Here, is -toral if is a product of a torus and a -group.
By the above considerations, any height function is admissible. When the group is finite, it turns out that the converse also holds: Any admissible function is realized by a height function . For positive dimensional there is an additional condition on top of admissibility. To state this condition, we recall that choosing an invariant Riemannian metric on also equips the set of closed subgroups with the Hausdorff metric, the underlying topology of which does not depend on the chosen metric on . This turns into a compact totally-disconnected metric space, in which a sequence of closed subgroups converges to another closed subgroup if and only if almost all are subgroups of and for every element the distance function converges to zero (see [tomDieckTransformationRepresentation, Section 5.6]). If , we have the following two implications about representations:
-
1.
Let be a representation of with . Thus, writing as a sum of characters , none of the is trivial. For sufficiently large , no contains (since is a closed proper subgroup) and thus .
-
2.
Let and be two characters of such that for all sufficiently large . If , then , in contradiction with the previous point. Thus .
We have the following:
Proposition 7.3.
For every finite subset , the height function is a locally constant function on .
Proof.
We start by noting that
The maximum of a finite number of locally constant functions is again locally constant. Hence it suffices to understand that is locally constant for any element of .
Now let denote a sequence of subgroups of converging to a subgroup . We need to show that for almost all . Without loss of generality we can assume that the are subgroups of . Replacing by if necessary we can further assume that .
We first assume that for some , and show that then also for almost all . If , there exists an -representation with , such that lies in the ideal . For all large enough we have , meaning that for these the restriction becomes invertible in . It follows that maps to the ideal generated by in , in other words is contained in .
For the other direction, we assume that is not contained in for some , and show that then also for almost all . For this we recall from Remark 2.26 the construction of elements for every character and satisfying the following three properties:
-
1.
for every group homomorphism .
-
2.
.
-
3.
for all abelian compact Lie groups .
Now, if is not contained in , it maps to a non-trivial element in
In other words, there exists an -representation with and pairwise different non-trivial characters such that is a polynomial over in the classes , , not all of whose coefficients are contained in . For all large enough we have that (i) and that (ii) all the characters restrict to pairwise different and non-trivial characters of . It then follows that for these the element equals the corresponding polynomial in the classes , implying that it maps to a non-trivial element in . In other words, is not contained in for large enough. This finishes the proof. ∎
Together with admissibility, this property characterizes the height functions of finite subsets of :
Proposition 7.4.
Given a function , the following are equivalent:
-
1.
There exists a finite subset such that .
-
2.
The function is admissible and locally constant.
Proof.
We have already shown the implication 1. 2.
It remains to show that given a locally constant admissible function , there exists a finite subset with . We start with the following claim: If is admissible, then given any pair of subgroups , there exists an element such that and . To see this, we distinguish between three cases:
-
(i)
If is not a subgroup of , we can choose a character which restricts to the trivial character over but to a non-trivial character over . Then has the desired properties, since and . Here and in the following, we set , and .
-
(ii)
If is a subgroup of with not a -group, we choose a prime dividing the order of and a surjection containing in the kernel. Then we set , where is the element introduced in Section 5.1. Then is an element of and its restriction to the trivial group is given by and hence a unit. It follows that , since . Moreover, , since . Hence, we can set to be any lift of to an element of .
-
(iii)
The remaining case is when is a subgroup of with a -group. Let be the minimum of the -rank of and the number , and choose a surjection . By 5.25 we know for that there exist an element such that and . (We set .) We can choose an embedding and restrict to an element . Then we have and (see 5.25). If , choose . It follows that and , since is admissible. Hence, has the desired properties.
Now given any such pair there exists an open neighbourhood of on which both and are constant. The for varying form an open cover of the compact space . Let be a finite subcover. We then set
to be the product of the corresponding elements. For any closed subgroup we have
For this gives , since for all . Any is contained in for some , yielding
In summary, the height function is less than or equal to everywhere and agrees with at itself. Once more we can apply that and are locally constant to find that and in fact agree on a neighborhood of . Letting vary, this yields an open cover of , for which we can choose a finite subcover . Hence, for every , there exists some such that . Finally, we define to be the set . Since is given by the maximum of the functions , it follows that has the desired property. ∎
Theorem 7.5.
The Zariski topology on has as a basis the closed sets
for all locally constant, admissible functions .
Proof.
A basis for the Zariski topology is given by the sets for all finite subsets of . As we saw above, is determined by its height function as
By Proposition 7.4, the functions that occur as height functions of finite subsets are precisely the locally constant admissible functions, which finishes the proof. ∎
8 Comparison with -spectra
In this final section, we discuss the relationship of the algebraic results of the previous sections with the theory of -spectra.
8.1 The universal support theory via -homology
We begin by comparing our classification of invariant prime ideals with the Balmer spectrum of compact -local -spectra. We recall from [BalmerSpectrum] that a prime ideal of a tensor-triangulated category is defined to be a thick tensor-ideal with the additional property that if is contained in , then or . The set of all prime ideals assembles to a topological space , the Balmer spectrum, with the topology generated by the closed sets for all objects .
Here, the support function , assigning a closed set of the Balmer spectrum to every object of , is the universal support theory in the sense of Balmer. This means that it is terminal among pairs of a topological space and a function
satisfying , , , and for all , and whenever there exists a distinguished triangle . See [BalmerSpectrum] for more details.
In the case of compact -local -spectra for an abelian compact Lie group , the Balmer spectrum was computed in [BalmerSanders, BarthelHausmannNaumannNikolausNoelStapleton, BarthelGreenleesHausmann]. Given a closed subgroup and , one defines a prime ideal
where denotes the nth Morava -theory, and is the -geometric fixed point spectrum of , a compact -local spectrum.
Theorem 8.1 ([BalmerSanders, BarthelHausmannNaumannNikolausNoelStapleton, BarthelGreenleesHausmann]).
The map
defines a bijection. Moreover, the topology on has a basis given by the closed sets defined by
where ranges through all admissible functions .
Here, ‘admissible’ is meant in the sense of Definition 7.2. Hence, comparing to Theorem 7.5, we see that the assignment defines a homeomorphism from to .333The reader may have noticed that in this section our admissible functions take values in , while in the last section they took values in . Likewise, we consider the condition here, and considered before. A shift by one shows that the topologies agree. The shift is caused by wanting to have having height in the last section. The computation of is also analogous to the one of 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 to take values in the invariant prime ideals of . The goal of the remainder of this section is to construct this universal support theory more intrinsically using -homology and the structure of equivariant formal group laws described in this paper. The idea is the following: Given a compact -local -spectrum , we can consider its equivariant complex bordism homology . Here, underlining indicates that we take the -Mackey functor valued homology of , i.e., we record the collection of for all closed subgroups of , together with restriction and transfer maps between them. We will always work at the fixed prime and -localize everything implicitly.
Since the coefficients are isomorphic to the Lazard ring and moreover the cooperations agree with , the groups form a graded -Mackey functor in -comodules. As such, we can take its support in the invariant prime ideals
Remark 8.2.
In general, is different from . Take for example , and , the circle with action given by reflection at a line. We have
The module is rational as well since . Rationally, splits into the coinvariants and the geometric fixed points , for the Euler class of the unique non-trivial character. The element from 5.21 becomes zero in the geometric fixed points, but restricts to and is thus not in ; thus becomes zero as well after localization at , and .
Remark 8.3 (Transfer maps).
The isomorphisms imply that on top of the contravariant restriction maps along group homomorphisms there also exist transfer maps for inclusions of finite index. In other words, the collection of all equivariant Lazard rings 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 on since the restriction is surjective. Inductively, we can further assume that . Furthermore, for since transfers are compatible with inflation maps (see, e.g., [SchGlobal, Theorem 4.2.6 ff.]). Hence it suffices to identify . We claim that it equals . Indeed, any transfer maps to zero in the geometric fixed points and is thus an element of . We know that generates . Hence, we can write for . Writing as for a non-trivial Euler class and , we obtain since by the definition of . We obtain since restricts to and thus and .
We will now see that our notion of support is another model for the universal support theory on compact -spectra.
We first note a major inconvenience: It is unclear whether is a finitely generated -module, even for compact . This is in contrast with the non-equivariant situation, where finite generation of for compact follows from the fact that is a polynomial ring and hence coherent. An analogous statement is unknown for equivariant Lazard rings. In particular, it is a priori unclear that is indeed a closed subset of and we have to prove this by hand, see Proposition 8.5 below.
The following proposition gives the relationship between the -homology support theory described above and geometric fixed points.
Proposition 8.4.
Let be a subgroup of , and a compact -spectrum. Then the following are equivalent:
-
(i)
The Mackey functor is trivial.
-
(ii)
The -module is trivial.
-
(iii)
The -geometric fixed points are of chromatic type .
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 is always finitely generated.
We have the following corollary:
Proposition 8.5.
Let be a compact -spectrum. Then its support is a closed subset of .
Moreover, the assignment
is a support theory on compact -spectra.
Proof.
We consider the type function
which is locally constant by [BarthelGreenleesHausmann, Proposition 4.3]. By the previous proposition, is an element of if and only if . Since the support is closed under inclusion, Theorem 5.1 implies that if for every -toral inclusion . Thus is admissible in the sense of Definition 7.2. Theorem 7.5 implies that is closed, as desired.
For the second part, all required properties of a support theory follow easily from exactness of localization 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
Proposition 8.4 makes it clear that this map sends to and is hence bijective. Therefore we can conclude from the results of this paper that the topology on is at least as coarse as the one on . In other words, our proof of the existence of inclusions gives another proof of the analogous inclusion on the topological side. The fact that there are no further topological inclusions requires additional arguments, namely the existence of compact -spectra with ‘maximal type shifting behaviour’. See [BarthelHausmannNaumannNikolausNoelStapleton, Section 4] or [KuhnLloydChromatic, Section 7]. Knowing this, we see that is a universal support theory on compact -spectra.
It remains to give the proof of Proposition 8.4, which we will do in three steps:
: Since the -action on factors through and , it follows that we have an isomorphism . In particular the vanishing of the entire Mackey-functor implies the vanishing at the subgroup as a special case.
: Note that contains none of the non-trivial Euler classes for , hence the non-trivial Euler classes act invertibly on . It follows that we have an isomorphism
Modulo , the ring embeds into the field of fractions of , which is non-trivial. Moreover, is not contained in and hence is invertible in the localization. It follows that the localization is an -algebra of height , i.e., its vanishing detects compact spectra of type .
: Let be a compact -spectrum such that is of type . By the previous paragraph we know that , and we have to show that for all other closed subgroups , too. We first assume that does not contain . Then there exists a character which restricts to the trivial character for but to a non-trivial character for . It follows that is not contained in and hence acts invertibly on . On the other hand restricts to in , so it follows that and in particular , as desired. Hence we can assume that contains as a subgroup. Since the statement then no longer depends on the ambient group, we can reduce to the case .
Hence we need to show that . By induction on the pair (dimension, rank of ) it follows that all the localizations at smaller intermediate groups vanish, and hence the homotopy groups of are concentrated at . In particular this implies that the map is an isomorphism and all Euler classes for non-trivial characters act invertibly on . Now, if restricts to the trivial character in (such a always exists since we can assume that is a proper subgroup of ), its Euler class lies in the maximal ideal of .
If we knew that is finitely generated over we could apply Nakayama’s lemma to see directly that , as desired. Since we do not know this, we have to argue differently: By Corollary 5.2, we know that is trivial if is not a -group, and is of height otherwise (where negative heights again mean that the theory is trivial). Hence, what we want to show is that if is of type , then is of type . Indeed, this implies that is trivial.
This precise statement about 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 and by replacing by the compact -spectrum it suffices to show two special cases:
-
1.
If is a compact -spectrum of underlying type , then the type of is at least .
-
2.
If is a compact -spectrum of underlying type , then the type of is also .
For (1) it suffices to find a complex oriented theory of height such that is of height , where denotes the Borel theory associated to (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 -theory has this property, building on results of Hopkins-Kuhn-Ravenel on the -disivible group associated to [HKR00] and extending earlier work of Greenlees, Hovey and Sadofsky. Using Remark 5.15 we obtain similar results more generally:
Proposition 8.6.
Let be any complex oriented ring spectrum of height which is Landweber exact over , i.e., acts trivially on , is a regular element and is a unit modulo . Then is of height .
Proof.
We apply Remark 5.15 and check that its assumptions are satisfied. We have , with Euler classes given by the -series for the formal group law associated to . If is a power of , the leading term of this Euler class is a power of , which we assumed to be regular. Modulo , the leading term becomes which is a unit since is of height . Hence by Remark 5.15 we find that the pushforward of the element to , i.e., the element for , generates the kernel of the composite
The leading term of equals , which is not a unit. Hence is non-trivial and since acts trivially it must be of height . Furthermore, reducing modulo the leading coefficient of equals a power of . Since is a unit modulo , it follows that so is . Therefore is trivial, and hence is of height exactly . ∎
Example 8.7.
Given any Landweber exact complex oriented ring spectrum of height , its quotient satisfies the assumptions of the previous proposition. It follows that also in this case the height of equals . For example this applies to Johnson-Wilson spectra, to or to Morava -theories.
Remark 8.8.
Another approach to blue-shift questions like the above is via the chromatic Nullstellensatz of [BurklundSchlankYuan]: Let be an -ring of height , i.e. is non-trivial and is trivial. If is complex-oriented, this is equivalent to our previous definition of height, i.e. to the vanishing of -homology detecting compact spectra of type . We claim that the Tate spectra are of height . Indeed, by the results of [BurklundSchlankYuan] any such theory maps to Morava -theory , and is non-trivial -locally by [BarthelHausmannNaumannNikolausNoelStapleton, Theorem 3.4] or the argument above. Furthermore, must be trivial -locally since the converse would contradict the existence of compact -spectra of underlying type and -geometric fixed points of type , by the same argument referenced above.
Similarly, for (2) one needs to find a complex oriented theory of height such that is also of height . This is more elementary and satisfied by any -local complex oriented theory of height . To see this, note again that the Euler classes are given by the -series . Writing with coprime to , we find that is a unit multiple of . Modulo , for the leading term of is a power of , which by assumption is a unit in . It follows that, modulo , the coefficients of are given by , which is always non-trivial when is.
8.2 Change of groups and the structure of
For any -spectrum , the groups come equipped with the structure of a graded -comodule. Since the stack is the stack associated to this graded Hopf algebroid, we obtain an associated quasi-coherent sheaf on (see [MeierOzornova, Proposition 4.3]). This is the -th graded piece of a -valued homology theory on -spectra, whose -th piece is given by . We end this section by a closer look upon how the structure of relates to this homology theory.
Recall from 3.11 that for a closed subgroup , there is an open immersion and a closed immersion . We obtain corresponding adjunctions
and |
Believing that the structure of dictates the structure of the -category of -spectra, we expect a relation to the adjunctions
and |
Here, denoting by the family of subgroups of not containing and by the projection, we define and . Note that the definition of is made so that its underlying spectrum is and more generally for every . 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 and induces maps
which can be checked to be equivalences on geometric fixed points. The inverses | ||||
form the counit and unit of the adjunction. In particular, is fully faithful and its image agrees with that of . Since smashing with is idempotent and hence symmetric monoidal, is symmetric monoidal as well and so is .
Proposition 8.9.
The diagram
commutes, i.e. there are natural isomorphisms for and for .
Moreover, there is a natural isomorphism for .
No such isomorphism can be expected for and in general. One reason is that is not flat, but even a spectral sequence relating and seems not to exist for 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 . Here is obtained from by inverting for all , and the ring is defined as . 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 .
Lemma 8.10.
For and the projection, there are natural isomorphisms
and
of graded -comodules, where the map is defined as the composite .
Proof.
A model for is given by for the sum of all characters , for the set of characters not restricting to in or, equivalently, . Indeed, for if and only if , as this is equivalent to none of the restricting to in .
In other words, smashing with is the same as inverting all the maps for . For an -module, this is equivalent to inverting for .
We have
Replacing by 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 -variable and we can plug into this variable the left and right unit .
To construct the second isomorphism, note that as part of the global structure of equivariant , there is a ring map (cf. [LinskensNardinPol]). Applying yields a ring map . This induces a morphism
It is enough to show that this is an isomorphism for finite and hence for for . In this case, the map becomes
(8.11) |
The natural map
is an isomorphism since is a surjection with kernel generated by the Euler classes for those restricting trivially to ; these are exactly the images of the Euler classes for those restricting trivially to .
Thus, (8.11) becomes
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 for -spectra . Consider the commutative diagram
where the down-right arrow comes from applying to an -equivariant formal group classified by a morphism to , and the right-pointing horizontal morphisms come from the composition . The square is a pullback square by 3.11. Thus is faithfully flat and hence induces an equivalence of to graded -comodules.
The comodule corresponding to is . As in (the proof of) the first isomorphism in Lemma 8.10, we observe that this is isomorphic to
By the second isomorphism in Lemma 8.10, this is isomorphic to the comodule corresponding to , i.e. to . This establishes the first claimed isomorphism of sheaves.
Since the counit is an equivalence, we can assume for the proof of for that for some and we obtain from the first isomorphism in this case a natural isomorphism