Odd primary analogs of Real orientations
Abstract.
We define, in -equivariant homotopy theory for , a notion of -orientation analogous to a -equivariant Real orientation. The definition hinges on a -space , which we prove to be homologically even in a sense generalizing recent -equivariant work on conjugation spaces.
We prove that the height Morava -theory is -oriented and that is -oriented. We explain how a single equivariant map completely generates the homotopy of and , expressing a height-shifting phenomenon pervasive in equivariant chromatic homotopy theory.
1. Introduction
The complex conjugation action on gives rise to a -equivariant space, , with fixed points . The subspace is invariant and equivalent as a -space to , the one-point compactification of the real regular representation of . A -equivariant ring spectrum is Real oriented if it is equipped with a map
such that the restriction
is the -suspension of the unit map . Such a Real orientation induces a homotopy ring map
with domain the spectrum of Real bordism [AM78, HK01]. These orientations have proved invaluable to the study of -local chromatic homotopy theory, leading to an explosion of progress surrounding the Hill–Hopkins–Ravenel solution of the Kervaire invariant one Problem [HHR16, GM17, HM17, KLW17, HLS18, HSWX19, BBHS19, LLQ20, LSWX19, HS20, BHSZ20, MSZ20].
The above papers solve problems, at the prime , that admit clear but often unapproachable analogs for odd primes. To give two examples, the primary Kervaire problem remains unresolved [HHR11], and substantially less precise information is known about odd primary Hopkins–Miller -theories [BC20, Conjecture 1.12].
To rectify affairs at , the starting point must be to find a -equivariant space playing the role of . This paper began as an attempt of the first two authors to understand a space proposed by the third.
Construction 1.1 (Wilson).
For any prime , let denote the fiber of the -equivariant multiplication map
where the codomain has trivial -action and the domain has -action cyclically permuting the terms. In other words, a map of spaces consists of the data of:
-
•
A -tuple of complex line bundles on .
-
•
A trivialization of the tensor product .
The action on is given by
Remark 1.2.
There is an equivalence of -spaces In general, the non-equivariant space underlying is equivalent to . The fixed points are equivalent to the classifying space , as can be seen by applying the fixed points functor to the defining fiber sequence for . The key point here is that the -fixed points of consist of the diagonal copy of , and is the fiber of the th tensor power map .
To formulate the notion of Real orientation, it is essential to understand the inclusion of the bottom cell
At an arbitrary prime, the analog of this bottom cell is described as follows:
Notation 1.3.
We let denote the cofiber of the unique non-trivial map of pointed -spaces from to . This is the spoke sphere, and it is a wedge of copies of with action on reduced homology given by the augmentation ideal in the group ring . We denote the suspension of the spoke sphere by either or , and 1.6 provides a natural inclusion
We will often also use to denote , and to denote its Spanier-Whitehead dual.
With this bottom cell in hand, we propose the following generalization of Real orientation theory:
Definition 1.4.
A -orientation of a -equivariant ring is a map of spectra
such that the composite
is the -suspension of the unit map .
Remark 1.5.
Applying the geometric fixed point functor to a -orientation we learn that the non-equivariant spectrum has in its homotopy groups.
Remark 1.6.
Let denote the -equivariant Eilenberg–MacLane spectrum associated to the constant Mackey functor. Then there is an equivalence of -equivariant spaces
Indeed, suspending and rotating the defining cofiber sequence gives rise to a cofiber sequence . Tensoring with and applying yields the defining fiber sequence for .
Under this identification, the natural inclusion is simply adjoint to the -suspension of the unit map . In particular, the identification gives a canonical -orientation of . In contrast, Bredon cohomology with coefficients in the Burnside Mackey functor cannot be -oriented, since is nonzero in the geometric fixed points.
In this paper we explore the interaction between -orientations and chromatic homotopy theory in the simplest possible case: chromatic height . Specifically, we study the following height -ring spectra:
Notation 1.7.
We let denote the height Lubin–Tate theory associated to the Honda formal group law over , with -action given by a choice of order element in the Morava stabilizer group. At , we let denote the -localized connective ring of topological modular forms with full level structure [Sto12]. The ring naturally admits an action by , and we restrict along an inclusion to view as a -equivariant ring spectrum.
The underlying homotopy groups of these spectra are given respectively by
We will review the -actions on the homotopy groups in Section 5.
Theorem 1.8.
For all primes , there exists a -orientation of the -equivariant Morava -theory .
Theorem 1.9.
The (-localized) -equivariant ring of topological modular forms with full level structure admits a -orientation.
Our second main result concerns the fact that, while
has distinct named generators, the conglomeration of them is generated under the -orientation by a single equivariant map .
Construction 1.10.
In Section 6, we will construct a map of -equivariant spectra
This map should be viewed as canonical only up to some indeterminacy, just as the classical class is only well-defined modulo . As was pointed out to the authors by Mike Hill, one choice of this map is given by norming a non-equivariant class in .
Construction 1.11.
Suppose a -equivariant ring is -oriented via a map
so that we may consider the composite
Using the dualizability of , this composite is equivalent to the data of a map
The non-equivariant spectrum underlying is (non-canonically) equivalent to a direct sum of copies of . In particular, by applying to the map one obtains a map from a rank free -module to .
Definition 1.12.
Given a -equivariant ring with a -orientation, the span of will refer to the subset of consisting of the image of the rank free -module constructed above.
Theorem 1.13.
For any -orientation of , the span of in is all of .
Theorem 1.14.
For any -orientation of the height Morava -theory , the span of inside maps surjectively onto .
Remark 1.15.
The map associated to a -oriented has an interpretation that may be more familiar to readers acquainted with the Hopkins–Miller computation of the fixed points of . Specifically, by definition there is a cofiber sequence
where is the transfer. It follows that the map determines a traceless element in , and the existence of such a traceless element was a key tool in the computations of [Nav10].
1.1. Homological and homotopical evenness
Non-equivariantly, complex orientation theory is intimately tied to the notion of evenness. A fundamental observation is that, since has a cell decomposition with only even-dimensional cells, any ring with must be complex orientable.
In -equivariant homotopy theory, a ring is called even if , and it is a basic fact that any even ring is Real orientable [HM17, §3.1].
In -equivariant homotopy theory, we propose the appropriate notion of evenness to be captured by the following definition, which we discuss in more detail in Section 3:
Definition 1.16.
We say that a -equivariant spectrum is homotopically even if the following conditions hold for all :
-
(1)
.
-
(2)
-
(3)
Remark 1.17.
In the presence of condition (1), condition (3) is equivalent to the statement that the transfer is surjective in degree . Conditions and constrain certain slices of , as we spell out in 3.14.
Remark 1.18.
A -spectrum is homotopically even, according to our definition above, if and only if it is even in the sense of [HM17, §3.1].
We prove the following theorem in Section 4.
Theorem 1.19.
If a -local -ring spectrum is homotopically even, then it is also -orientable.
The key point here, as we explain in Section 4, is that admits a slice cell decomposition with even slice cells. An even more fundamental fact, which turns out to be equivalent to the slice cell decomposition, is a splitting of the homology of :
Definition 1.20.
We say that a -spectrum is homologically even if there is a direct sum splitting
where each is equivalent, for some , to one of
Theorem 1.21.
The space is homologically even.
Remark 1.22.
The notion of homological evenness we propose in this paper restricts, when , to the notion studied by Hill in [Hil19, Definition 3.2]. Notably, our definition differs from Hill’s when .
Returning again to the group , work of Pitsch, Ricka, and Scherer relates a version of homological evenness to the study of conjugation spaces [PRS19]. An interesting example of a conjugation space, generalized in [HH18] and its in-progress sequel, is . It would be very interesting to develop a -equivariant version of conjugation space theory. Since is a form of (cf. 7), we wonder whether there is an interesting slice cell decomposition of .
Remark 1.23.
The slice cell structure on has many interesting attaching maps. The first non-trivial attaching map is a class , with fixed points the multiplication by map on . This class was previously studied by the third author [Wil17a, §3.2] and, independently, Mike Hill. The -equivariant is the familiar map .
1.2. A view to the future
The most natural next question, after those tackled in this paper, is the following:
Question 1.
Let , and fix a formal group of height over a perfect field of characteristic . When is the associated Lubin–Tate theory -orientable?
We have not fully answered this question even for , since we focus attention on the Honda formal group.
It seems likely that further progress on 1, at least for , must wait for work in progress of Hill–Hopkins–Ravenel, who have a program by which to understand the -action on Lubin–Tate theories. As the authors understand that work in progress, it is to be expected that the height Morava -theory has homotopy generated by copies of the reduced regular representation, . One expects to be able to construct Morava -theories, generated by a single , and we expect at least these Morava -theories to be homotopically even in the sense of this paper.
Question 2.
Can one construct homotopically even Morava -theories?
In light of the orientation theory of Section 2, it seems useful to know if Morava -theories admit norms. Indeed, at the Real Morava -theories all admit the structure of -algebras. Since the first Morava -theory should be , or perhaps , it seems pertinent to answer the following question first:
Question 3.
At the prime , what structure is carried by the -equivariant spectrum ? Is there an analog of the structure carried by ?
In another direction, one might ask about other finite subgroups of Morava stabilizer groups:
Question 4.
Is there an analog of the notion of -orientation related to the -actions on Lubin–Tate theories at the prime ?
One may also go beyond finite groups and ask for notions capturing other parts of the Morava stabilizer group, such as the central that acts on after -completion.
To make full use of all these ideas, one would like not only an analog of , but also an analog of at least one of or . Attempts to construct such analogs have consumed the authors for many years; we consider it one of the most intriguing problems in stable homotopy theory today.
Question 5.
(Hill–Hopkins–Ravenel [HHR11]) Does there exist a natural -ring spectrum, , with
-
•
Underlying, non-equivariant spectrum the smash product of copies of .
-
•
Geometric fixed points .
At , it should be the case that .
To the above we may add:
Question 6.
Does such a natural orient all -orientable -ring spectra, or at least all those that admit norms in the sense of Section 2?
Most of our attempts to build have proceeded via obstruction theory, while is naturally produced via geometry. It would be extremely interesting to see a geometric definition of an object . Alternatively, it would be very clarifying if one could prove that a reasonable does not exist. As some evidence in that direction, the authors doubt any variant of can be homotopically even.
Even if cannot be built, or cannot be built easily, it would be excellent to know whether it is possible to build -ring spectra .
Question 7.
Does there exist, for each prime , a -ring satisfying the following properties:
-
•
is the -localization of , and is the -localization of .
-
•
The homotopy groups are given by
with . The action on these generators should make into a copy of the reduced regular representation.
-
•
There is a -ring map .
-
•
is homotopically even, and in particular -orientable.
-
•
The underlying spectrum additively splits into a wedge of suspensions of .
-
•
We have for a generator of degree .
It is plausible that should come in many forms, in the sense of Morava’s forms of -theory [Mor89]. A natural form might be obtained by studying compactifications of the Gorbounov–Hopkins–Mahowald stack [GM00, Hil06] of curves of the form
Studying the uncompactified stack, it is possible to construct a -equivariant ring which is a analog of uncompleted Johnson-Wilson theory. The details of this construction will appear in forthcoming work of the second author.
Remark 1.24.
The -action on is naturally the restriction of an action by . In fact, most objects in this paper admit actions of , or at least of , but these are consistently ignored. The reader is encouraged to view this as an indication that the theory remains in flux, and welcomes further refinement.
Remark 1.25.
Since work of Quillen [Qui69], the notion of a complex orientation has been intimately tied to the notion of a formal group law. There are hints throughout this paper, particularly in Section 2 and Section 6, that the norm and diagonal maps on lead to equivariant refinements of the -series of a formal group. It may be interesting to develop the purely algebraic theory underlying these constructions, particularly if algebraically defined turn out to be of relevance to higher height Morava -theories.
1.3. Notation and Conventions
-
•
If is a -space, we use to denote the underlying non-equivariant space, and we use to denote the fixed point space. If is a -spectrum, we will use either or to denote the underlying spectrum, and we use to denote the geometric fixed points.
-
•
We fix a prime number , and throughout the paper all spectra and all (nilpotent) spaces are implicitly -localized. In the -equivariant setting, this means that we implicitly -localize both underlying and fixed point spaces and spectra.
-
•
If is a -space or spectrum, we use to denote the homotopy groups of , considered as a graded abelian group with -action. If is a -representation, we use to denote the set of homotopy classes of equivariant maps from to .
-
•
We let denote the cofiber of the -equivariant map , and we also use to refer to the suspension -spectrum of this -space. We let denote the Spanier-Whitehead dual of the -spectrum . Given a -representation and a -spectrum , we will use and to denote the set of homotopy classes of equivariant maps from and to .
-
•
We let denote the fiber of the -equivariant multiplication map .
-
•
If is a classical commutative ring, we use to denote the -module given by the augmentation ideal . This is a rank -module with generators permuted by the reduced regular representation of . We similarly use to denote the -module that is isomorphic to with trivial action. We sometimes use to denote itself, and write free to denote a sum of copies of .
1.4. Acknowledgments
The authors thank Mike Hill, Mike Hopkins, and Doug Ravenel for inspiring numerous ideas in this document, as well as their consistent encouragement and interest in the work. We would especially like to thank Mike Hill for suggesting that our earlier definition of might be more conceptually viewed as a norm. We additionally thank Robert Burklund, Hood Chatham and Danny Shi for several useful conversations, and the anonymous referee for suggesting several improvements.
The first author was supported by NSF grant DMS-, and thanks Yuzhi Chen and Wenyun Liu for their hospitality during the writing of this paper. The second author was supported by an NSF GRFP fellowship under Grant No. 1745302. The third author was supported by NSF grant DMS-.
2. Orientation theory
Non-equivariantly, one may study complex orientations of any unital spectrum . However, if is further equipped with the a homotopy commutative multiplication, then the theory takes on extra significance: in this case, a complex orientation of provides an isomorphism .
In this section, we work out the analogous theory for -orientations. In particular, we find that the theory of -orientations takes on special significance for homotopy ring spectra that are equipped with a norm refining the underlying multiplication. Recall the following definition from the introduction:
Definition 2.1.
A -orientation of a unital -spectrum is a map
such that the composite
is equivalent to of the unit.
For any representation sphere , it is traditional to denote by the free -ring spectrum
Below, we extend this construction to take input not only representation spheres , but spoke spheres as well.
Definition 2.2.
For integers , let
where we consider as a -bimodule via the -map induced by the composite
In this composite, the first map is adjoint to the identity on and the second map is the canonical inclusion. Note that is a unital left module over .
Furthermore, given a -equivariant spectrum , we set
Construction 2.3.
Suppose that is a homotopy ring in -spectra, further equipped with a genuine norm map
which is unital and restricts on underlying spectra to the composite
where is the generator and is the -fold multiplication map.
If is -oriented by a map
then we may produce a map
as follows. First, the composite
where the map is the inclusion of the factor of corresponding to the identity in , extends to a map
since the target is a homotopy ring. Norming up, and combining the norm on with the diagonal map , we get a map
Finally, the extension of over provides a nullhomotopy of the composite
producing a map
We finish by extending scalars to , using the assumption that is a homotopy ring.
Construction 2.4.
If is -oriented then so too is the Postnikov truncation . The construction above is natural, and so we may form a map
Theorem 2.5.
Suppose is a -oriented homotopy ring, further equipped with a unital homotopy -module structure such that the unit
respects the underlying multiplication in the sense of 2.3. Then, with notation as above, the map
is an equivalence.
Proof.
By construction, it suffices to prove that the map
is an equivalence for each . This is clear on underlying spectra. On geometric fixed points we can factor this map as
being careful to interpret the source as a module (this is not a map of rings). Specifically, the above composite is one of unital -modules and, separately, one of -modules.
The second map is an equivalence by 2.6 below, so we need only prove the first map is an equivalence. Since in , the Atiyah-Hirzebruch spectral sequence computing has -page given by
The class is realized by applying geometric fixed point to the -orientation. The powers of are obtained from the unit of the unital -module structure. Using the -module structure, this implies that the spectral sequence degenerates and moreover that the first map is an equivalence. ∎
Lemma 2.6.
If is bounded above, and is a -space of finite type, then the map
is an equivalence.
Proof.
Write where the are skeleta for a -CW-structure on with each finite. Then the fiber of
becomes increasingly coconnective, and hence the map
is an equivalence. We are thus reduced to the case finite, where the result follows since is exact. ∎
Remark 2.7.
In practice, the conditions of 2.5 are often easy to check. For example, any -commutative ring in the homotopy category of -spectra will be equipped with a unital homotopy -module structure respecting the multiplication in the sense of 2.3 [HH16]. We choose to write 2.5 in generality because, even non-equivariantly, one occasionally studies orientations of rings that are not homotopy commutative (like Morava -theory at the prime ).
Corollary 2.8.
There is a natural equivalence
3. Evenness
In this section, we will introduce a notion of evenness in -equivariant homotopy theory. This is a generalization of the notion of evenness in non-equivariant homotopy theory. Evenness comes in two forms: homological evenness and homotopical evenness. Homological evenness is a -equivariant version of the condition that a spectrum have homology concentrated in even degrees, and homotopical evenness corresponds to the condition that a spectrum have homotopy concentrated in even degrees.
The main results in this section are 3.9, which shows that, under certain conditions, a bounded below homologically even spectrum admits a cell decomposition into even slice spheres (defined below), and 3.17, which shows that there are no obstructions to mapping in a bounded below homologically even spectrum to a homotopically even spectrum.
3.1. Homological Evenness
We begin our discussion of evenness with the definition of an even slice sphere.
Definition 3.1.
We say that a -equivariant spectrum is an even slice sphere if it is equivalent to one of the following for some :
A dual even slice sphere is the dual of an even slice sphere. The dimension of a (dual) even slice sphere is the dimension of its underlying spectrum.
Remark 3.2.
The phrase slice sphere is taken from [Wil17b, Definition 2.3], where a -equivariant slice sphere is defined to be a compact -equivariant spectrum, each of whose geometric fixed point spectra is a finite direct sum of spheres of a given dimension.
It is easy to check that the (dual) even slice spheres of 3.1 are slice spheres in this sense.
Remark 3.3.
In the case , the even slice spheres are precisely those of the form
for some .
Definition 3.4.
We say that a -equivariant spectrum is homologically even if there is an equivalence of -modules
where is a direct sum of even slice spheres of dimension .
Remark 3.5.
When , this recovers the notion of homological purity given in [Hil19, Definition 3.2]. However, when is odd, our definition of homological evenness differs from Hill’s definition of homological purity. The most important difference is that we allow the spoke spheres to appear in our definition. This is necessary for to be homologically even.
As in the non-equivariant case, homological evenness for a bounded below spectrum is equivalent to the existence of an even cell structure. To prove this, we need to recall the following definition from [Wil17b, HY18]:
Definition 3.6.
We say that a -equivariant spectrum is regular slice -connective if:
-
(1)
is -connective, and
-
(2)
is -connective.
Furthermore, we say that is bounded below if it is regular slice -connective for some integer .
Lemma 3.7.
Let be a bounded below -spectrum with the property that is of finite type. Then is regular slice -connective if and only if is regular slice -connective.
Proof.
For the underlying spectrum, the follows from the fact that detects connectivity of bounded below -local spectra. For the geometric fixed points, we use the fact that , , detects connectivity of bounded below -local spectra which are of finite type, since a finitely generated -module is trivial if and only if it is trivial after tensoring with . ∎
Lemma 3.8.
Let denote an even slice sphere of dimension , and suppose that is regular slice -connective. Then we have .
Proof.
Proposition 3.9.
Suppose that is a bounded below, homologically even -equivariant spectrum with the property that is of finite type, so that there exists a splitting
where is a direct sum of -dimensional even slice spheres. Then admits a filtration such that for each .
Proof.
By assumption, we are given a splitting
where is a direct sum of -dimensional even slice spheres. By induction on , it will suffice to show that the dashed lifting exists in the diagram
since the cofiber of any such lift is a bounded below homologically even -spectrum with of finite type and whose -homology is .
3.2. Homotopical Evenness
We now introduce the homotopical version of evenness.
Definition 3.11.
We say that a -equivariant spectrum is homotopically even if the following conditions hold for all :
-
(1)
-
(2)
-
(3)
Remark 3.12.
All of the examples of homotopically even -spectra that we will encouter will satisfy the following condition for all :
-
(4)
We will say that a homotopically even -spectrum satisfies condition (4) if this holds.
In fact, the examples which we study satisfy even stronger evenness properties. We have chosen the weakest possible set of properties for which our theorems hold.
Remark 3.13.
If we assume condition (1), then we may rewrite conditions (3) and (4) as follows:
-
()
the transfer maps are surjective for all .
-
()
the restriction maps are injective for all .
This follows directly from the cofiber sequences defining and :
Remark 3.14.
Conditions - have some implications for the slice tower of any homotopically even , which can be read off from [HY18] (cf., [Wil17b, 3.5]). First, conditions and together imply that slices in degrees are trivial. Secondly, condition implies that the th slice is the zero-slice determined by the Mackey functor . However, when the implication of for slices is obscure, and many slices are unconstrained.
Remark 3.15.
Example 3.16.
The Eilenberg–Maclane spectra and are examples of homotopically even -spectra which satisfy condition (4). To verify this, we refer to the reader to the appendix of third author’s thesis [Wil17a, §A], where one may find a computation of the spoke graded homotopy groups of and .
At the prime , there are many examples of homotopically even -spectra in the literature, such as , and , where is equipped with the Goerss-Hopkins -action [HM17, HS20].
The main result of Section 5 is that the -spectra and the -spectrum are homotopically even and satisfy condition (4).
When trying to map a bounded below homologically even -spectrum into a homotopically even -spectrum, there are no obstructions:
Proposition 3.17.
Let be a homotopically even -spectrum, and suppose that is a -spectrum equipped with a bounded below filtration such that each is a direct sum of -dimensional even slice spheres.
Then, for any , every -equivariant map extends to an equivariant map .
Proof.
It suffices to prove by induction that any map extends to a map . Using the cofiber sequence
we just need to know that any map from the desuspension of an even slice sphere into is nullhomotopic. This follows precisely from the definition of homotopical evenness. ∎
If further satisfies condition (4), we have the stronger result:
Proposition 3.18.
Let be a homotopically even -ring spectrum which satisfies condition (4), and suppose that is a -spectrum equipped with a bounded below filtration such that each is a direct sum of -dimensional even slice spheres.
Then there is a splitting of the induced filtration on by -modules:
Proof.
We need to show that the filtration splits upon smashing with . Working by induction, we see that it suffices to show that all maps
where , are automatically null. Enumerating through all of the possible even slice spheres that can appear in and , and making use of the (non-canonical) equivalence
we find that this follows precisely from the hypothesis that is homotopically even and satisfies condition (4). ∎
4. The homological evenness of
The main goal of this section is to prove the following theorem:
Theorem 4.1.
The -spectrum is homologically even.
Noting that is of finite type, we may apply 3.9 and so deduce the following corollary:
Corollary 4.2.
There is a filtration of with subquotients as follows
Warning 4.3.
We believe that there is a filtration of the space that recovers upon applying , but we do not prove this here. As such, our name must be regarded as an abuse of notation: we do not prove that is of a -space . In light of the Dold-Thom theorem, it seems likely that the space could be defined as the th symmetric power of .
Remark 4.4.
As an application, we obtain the following analog of the fact that any ring spectrum with homotopy groups concentrated in even degrees admits a complex orientation:
Corollary 4.5.
Let be a homotopically even -ring spectrum. Then is -orientable.
Proof.
We devote the remainder of the section to the proof of 4.1. By 2.8, there is an equivalence
This is of finite type, so to prove 4.1 it will suffice to prove the following theorem and dualize:
Theorem 4.6.
As a -equivariant spectrum, is a direct sum of dual even slice spheres for all .
To prove this, we will construct a map in from a wedge of dual even slice spheres which is an equivalence on underlying spectra and geometric fixed points.
Construction 4.7.
The composition
is canonically null, and hence induces a map
On the other hand, letting
denote the canonical inclusion, there is the norm map
Since is a module over , this implies the existence of maps
for and .
We first show that the sum of these maps induces an equivalence on geometric fixed points:
Proposition 4.8.
Let
denote the direct sum of the maps . Then is an equivalence.
Proof.
We have an identification
Under this identification, the map
corresponds to the inclusion of into the left factor.
There are equivalences
and hence an isomorphism
Furthermore, the map
sends the fundamental class of to .
It follows that induces an isomorphism on homology, so is an equivalence. ∎
Our next task is to extend to a map that also induces an equivalence on underlying spectra. We will see that this can be accomplished by taking the direct sum with maps from induced even spheres, which are easy to produce. The main input is a computation of the homology of the underlying spectrum of as a -representation.
Lemma 4.9.
There is a -equivariant isomorphism
where lies in degree .
Proof.
There are equivariant isomorphisms
and
where and both lie in degree . Since is a unital -module, we obtain a map
of -modules. Since goes to zero in , it follows that this factors through a map
Examining the Künneth spectral sequence, we see that this map must be an isomorphism. ∎
The following theorem in pure algebra determines the structure of the mod reduction as a -representation:
Proposition 4.10 ([AF78, Propositions III.3.4-III.3.6]).
Let denote the reduced regular representation of over , and let denote generators which are cyclically permuted by and satisfy . We set . Then the symmetric powers of decompose as follows:
Proof of 4.6.
Let be as in 4.8. It follows from 4.9 and 4.10 that the mod homology of splits as . Moreover, is an equivalence on geometric fixed points by 4.8.
It therefore suffices to show that, given any summand of isomorphic to , there is a map whose image is that summand. Taking the direct sum of with an appropriate collection of such maps, we obtain an -homology equivalence. Since both sides have finitely-generated free -homology, this must in fact be a -local equivalence, as desired.
To prove the remaining claim, it suffices to show that the mod Hurewicz map
is surjective in every degree. This follows from the following square
where the top horizontal arrow is a surjection because is a non-equivariant direct sum of spheres, and the right vertical arrow is a surjection by the proof of 4.9. ∎
5. Examples of homotopical evenness
In this section, we introduce our principal examples of homotopically even -ring spectra. By 4.5, they are also -orientable.
Our first examples are the the Morava -theories associated to the height Honda formal group. As we will recall in Section 5.1, admits an essentially unique -action by -automorphisms. We use this action to view as a Borel -equivariant -ring.
Our second example is the connective -ring of topological modular forms with full level structure. The group acts on via modification of the level structure, and we view as a -equivariant -ring via the inclusion . We will discuss this example in Section 5.2.
The main result of this section is the homotopical evenness of the above -ring spectra:
Theorem 5.1.
The Borel -equivariant height Morava -theories associated to the Honda formal group over are homotopically even and satisfy condition (4).
Theorem 5.2.
The -ring spectrum of connective topological modular forms with full level structure is homotopically even and satisfies condition (4).
Applying 4.5, we obtain the following corollary:
Corollary 5.3.
The -ring spectra and are -orientable.
5.1. Height Morava -theory
Given a pair , where is a perfect field of characterstic and is a formal group over of finite height , we may functorially associate an -ring , the Lubin-Tate spectrum or Morava -theory spectrum of [GH04, Lur18]. There is a non-canonical isomorphism
where and .
Given a prime and finite height , a formal group particularly well-studied in homotopy theory is the Honda formal group. The Honda formal group is defined over , so the Frobenius isogney may be viewed as a endomorphism
The Honda formal group is uniquely determined by the condition that in .
The endomorphism ring of the base change of to is the maximal order in the division algebra of Hasse invariant and center . By the functoriality of the Lubin-Tate theory construction, the automorphism group of over acts on . To keep our notation from becoming too burdensome, we set
There is a subgroup , which is unique up to conjugation. Indeed, such subgroups correspond to embeddings . Since is of degree over , it follows from a general fact about division algebras over local fields that such a subfield exists and is unique up to conjugation (cf. [Ser67, Application on pg. 138]). Using any such , we may view as a Borel -equivariant -ring spectrum.
Homotopical evenness of will follow from the computation of the homotopy fixed point spectral sequence for , which was first carried out by Hopkins and Miller and has been written down in [Nav10] and again reviewed in [HMS17]. We recall this computation below. The homotopy fixed point spectral sequence takes the form
so the first step is to compute the action of on .
This action may be determined as follows. Abusing notation, let denote a lift of the canonically defined element . The element is fixed modulo by the and in particular the -action on , so if we fix a generator we find that the element is divisible by . Set . Then the two key properties of are that:
-
(1)
.
-
(2)
is a unit in . As a consequence, is a unit in which is fixed by the -action [Nav10, pg. 498].
The existence of an element satisfying the above two conditions completely determines the action of on , as follows. First, let denote any unit, and set . Then continues to satisfy (1) and (2) above and determines a map of -representations
This determines a -equivariant map
which identifies with the graded completion of at the graded ideal generated by the kernel of the essentially unique nonzero map of -modules .
Remark 5.4.
In Section 7, we will see that the element is intimately related to the -orientability of . For later use, we note that it follows from the above analysis that the map induced by is an isomorphism.
Remark 5.5.
As pointed out by the referee, the element may also be described in terms of -theory. The class determines a function , and it follows from the formula in that . From this perspective, the crucial fact that is a unit in follows from the calculations in [Rav78, pgs. 438-439].
Using the above determination of the -action on , as well as 4.10, one may obtain with some work the following description of :
Proposition 5.6 (Hopkins–Miller, cf. [HMS17, Proposition 2.6]).
There is an exact sequence
(1) |
where and .
Finally, we must recall the differentials in the homotopy fixed point spectral sequence. We let denote equality up to multiplication by an element of . Then, as explained in [HMS17, §2.4], the spectral sequence is determined multiplicatively by the following differentials:
along with the fact that all differentials vanish on the image of the transfer map.
In particular, on the -page of the homotopy fixed point spectral sequence there are no elements in positive filtration in total degrees , or . Indeed, there are no elements at all in the -stem.
We now have enough information to establish the homotopical evenness of .
Proof of 5.1.
Let denote the periodicity element. Then is also invertible, so the -graded equivariant homotopy of is -periodic.
Therefore, using 3.13, we see that it suffices to show that:
-
(1)
.
-
(2)
.
-
(3)
The transfer map is a surjection.
-
(4)
The restriction map is an injection.
Condition (1) is immediate from the fact that is even periodic. Condition (2) is a direct consequence of the above computation of the homotopy fixed point spectral sequence. Condition (3) follows from the following two facts:
-
•
The short exact sequence (1) implies that is spanned by the image of the transfer.
-
•
On the -page of the homotopy fixed point spectral sequence, there are no positive filtration elements in stem .
Condition (4) follows from the fact that on the -page of the homotopy fixed point spectral sequence, there are no positive filtration elements in the zero stem. ∎
5.2. The spectrum as a form of
Recall from [Sto12] or [HL16] the spectrum of connective topological modular forms with full level structure.111The spectrum is obtained from the spectrum discussed in the references by taking the -equivariant connective cover. In this section we will consider as implictly -localized. It is a genuine -equivariant -ring spectrum with -fixed points , the (-localized) spectrum of connective topological modular forms. We view as a -spectrum via restriction along an inclusion .
This spectrum has been well-studied by Stojanoska [Sto12]. In particular, Stojanoska computes , where and a generator of acts by and . It follows that and span a copy of , so that . The corresponding family of elliptic curves is cut out by the explicit equation
For later use, we note down some facts about the associated formal group law.
Proposition 5.7.
The -series of the formal group law associated to is given by the following formula:
It follows that we have the following formulas for and :
and
Proof.
This is an elementary computation using the method of [Sil09, §IV.1]. ∎
Remark 5.8.
Let , so that . Then we have
so that
Note that this element generates as a -algebra with -action. In Section 7, we will relate this element to the -orientation of .
In his thesis, the third author has computed the slices of (cf. [HHR16, §4]):
Proposition 5.9 ([Wil17a, Corollary 3.2.1.10]).
Given a -equivariant spectrum , let denote the th slice of . The slices of are of the form:
We now turn to the proof of 5.2. Given the computation of the slices of in 5.9, this will follow from 4.6 and the following proposition:
Proposition 5.10.
Let be a -spectrum whose slices are of the form , where is a direct sum of dual even slice -spheres. Then is homotopically even and satisfies condition (4).
Using the slice spectral sequence, the proof of 5.10 reduces to the following lemma:
Lemma 5.11.
Let denote a dual even slice sphere. Then is homotopically even and satisfies condition (4).
Proof.
If , then this follows from the fact that for all .
If , then this follows from the fact that is homotopically even, since the definition of homotopically even is invariant under -suspension.
6. and a formula for its span
In this section, given a -oriented -ring spectrum , we will define a class
When , our construction agrees with the class in the homotopy of a Real oriented -ring spectrum. Just as is well-defined modulo , we will see that is well-defined modulo the transfer. We will also give a formula for the image of in the the underlying homotopy of in terms of the classical element and the -action.
To define , we first construct a class , and then we take its image along the -orientation . To begin, we recall an analogous construction of the classical element .
6.1. The non-equivariant as a th power
We recall some classical, non-equivariant theory that we will generalize to the equivariant setting in the next section.
Notation 6.1.
We let denote a generator of the stable homotopy group .
Since is an infinite loop space, its suspension spectrum is a non-unital ring spectrum. This allows us to make sense of the following definition.
Definition 6.2.
We define the class to be , the th power of the degree generator.
There are at least two justifications for naming this class , which might more commonly be defined as the coefficient of in the -series of a complex-oriented ring. The relationship is expressed in the following proposition:
Proposition 6.3.
Let denote a (non-equivariant) homotopy ring spectrum, equipped with a complex orientation
which can be viewed as a class . Then the composite
records, up to addition of a multiple of , the coefficient of in the -series .
Proof.
Consider the -fold multiplication map of infinite loop spaces
Applying to the above, we obtain a map
By the definition of the formal group law associated to the complex orientation, the class is sent to the formal sum
The commutativity of the formal group law ensures that this power series is invariant under cyclic permutation of the .
The composite
that we must compute can be read off as the coefficient of the product in the power series . We may of course consider other degree monomials in the , such as . The coefficient in of any such degree monomial will be an element of . Summing these coefficients over all the possible degree monomials, we obtain the the coefficient of in the single variable power series .
Our claim is that this sum differs from the coefficient of by a multiple of . The reason is that is the unique monomial invariant under the cyclic permutation of the . For example, the coefficients of and will all be equal, so their sum is a multiple of . ∎
Remark 6.4.
The integral homology is a divided power ring on the Hurewicz image of . In particular, the Hurewicz image of is a multiple of times a generator of .
Consider the ring spectrum together with its canonical complex orientation
The integral homology is the symmetric algebra on the image, under this map, of . In particular, the Hurewicz image of in is sent to times an indecomposable generator of . By [Mil60], this provides another justification for the name .
Remark 6.5.
One might ask whether higher , with , can be defined in . A classical argument with topological -theory [Mos68] shows that the Hurewicz image of inside of is generated as a -module by powers of . For larger than , is not simply times a generator of , so it is impossible to lift the corresponding indecomposable generators of to . However, it may be possible to lift multiples of such generators.
Finally, we record the following proposition for later use:
Proposition 6.6.
Let denote a (non-equivariant) homotopy ring spectrum, equipped with a map
that induces the zero homomorphism on (in particular, is not a complex orientation). Then the image of in is a multiple of .
Proof.
Let denote the cofiber of .
We recall first that, -locally, the spectrum
admits a splitting as . Indeed, since is the lowest positive degree element in the -local stable stems, most of the attaching maps in the standard cell structure for are automatically -locally trivial. The only possibly non-trivial attaching map is between the th cell and the bottom cell, and this attaching map is detected by the action on .
By cellular approximation, must factor through , and again the lack of elements in the -local stable stems ensures a further factorization of through . Thus, to determine the image of in , it suffices to consider the composite
There is by definition a cofiber sequence . By the assumption that is trivial on , must factor as a composite
We now finish by noting that the composite must be a multiple of , because otherwise would split as . ∎
6.2. The equivariant as a norm
As we defined the non-equivariant to be the th power of a degree class, we similarly define an equivariant to be the norm of a degree class. We thank Mike Hill for suggesting this conceptual way of constructing . To see that is equipped with norms, we will make use of the following proposition:
Proposition 6.8.
There is an equivalence of -equivariant spaces
where denotes the -equivariant Eilenberg–Maclane spectrum associated to the constant Mackey functor.
Proof.
This is 1.6. ∎
Construction 6.9.
The above proposition equips the space with a natural norm, meaning a map
Indeed, any -equivariant infinite loop space , like , is equipped with a norm
This norm is applied to the -spectrum map
that is induced from the identity on .
Convention 6.10.
For the remainder of this section we fix a (non-canonical) equivalence
The natural map of -spaces
then induces an (again, non-canonical) equivalence
giving classes
Choosing our non-canonical equivalence appropriately, we may suppose that the -action on is given by the rules
-
(1)
, if
-
(2)
Definition 6.11.
We let
denote the norm of . Explicitly, norming the non-equivariant map yields a map
and we may compose this with the norm map of 6.9 to make the class
Remark 6.12.
Of course, the choice of the class above is not canonical. We view this as a mild indeterminancy in the definition of , related to the fact that the classical should only be well-defined modulo . As we will see later, many formulas we write for will similarly be well-defined only modulo transfers.
6.3. A formula for in terms of
Our next aim will be to give an explicit formula for the image of in the underlying homotopy of a -oriented cohomology theory. Our formula is stated as 6.21. To begin its derivation, our first order of business is to give a different formula for modulo transfers:
Proposition 6.13.
In , the class and the class differ by times a transferred class. In particular, is divisible by , and the class is the restriction of a class in .
Proof.
Identifying with and using the nonunital -ring structure on , we obtain a map
under which the norm class maps to the image of . The conclusion of the proposition then follows from 6.14 below. ∎
Lemma 6.14.
Let denote the reduced regular representation of over , and let denote generators which are cyclically permuted by and satisfy . We set .
Then is divisible by , and and differ by a transferred class in .
Proof.
To see that is divisible by , we expand it out in terms of the basis of :
It is clear from linearity of the Frobenius modulo that is divisible by . Our next goal is to show that is a transferred class. It is clearly fixed by the -action, so we wish to show that its image in
is zero. Since times any fixed point of is the transfer of an element, there is an isomorphism
By 4.10, there is an isomorphism of -representations
so that any choice of -equivariant map which is nonzero on restricts to an isomorphism
A choice of such a map may be made as follows. First, let denote the equivariant map sending each to . This induces a map which sends to . We now need to show that the image of under is also equal to . Writing
we find that its image of is equal to
as desired. ∎
6.13 can be read as the statement that is a formula for , if one is only interested in modulo transfers. We often find this formula for to be more useful in computational contexts.
Convention 6.15.
For the remainder of this section, we fix a -ring together with a -orientation
Definition 6.16.
The -orientation of gives rise to a map
which under our fixed identification of is given by a map
By mapping in the first of the copies of , and then projecting to the first of the copies of , we obtain the underlying complex orientation of .
Warning 6.17.
Notation 6.18.
Using 6.2, the underlying complex orientation of gives rise to a class .
Notation 6.19.
Recall our fixed non-canonical identification . Let correspond to the th copy of , so that we have
-
(1)
if , and
-
(2)
.
Then a generic class
may be written as
where .
The key relationship between the equivariant and non-equivariant is expressed in the following lemma:
Lemma 6.20.
The class maps to plus a multiple of in
Proof.
The class maps to for some collection of elements .
At last, we are ready to state the main result of this section:
Theorem 6.21.
Suppose that the underlying homotopy groups are torsion-free. Then the class is given, modulo transfers, by the class
Proof.
Remark 6.22.
Consider the class
of 6.21. If in this formula we replace by , for an arbitrary class , the resulting expression differs from the original by
This is exactly the transfer, in , of . Thus, altering by a multiple of does not change the class modulo transfers.
7. The span of in height theories
In this section, we use the formula of 6.21 to compute the span of in the height theories and , which we verified were -orientable in Section 5. Our main result, stated in Theorems 7.3 and 7.4, proves that the span of generates the homotopy of these theories in a suitable sense. This demonstrates a height-shifting phenomenon in equivariant homotopy theory: though these theories are height classically, the fact that their homotopy is generated by indicates that they should be regarded as height objects in -equivariant homotopy theory.
Notation 7.1.
Let denote a -ring spectrum, equipped with a -orientation
Precomposition with then yields a map
which by the dualizability of is equivalent to a map of -spectra
Engaging in a slight abuse of notation, we will throughout this section denote this map by
Definition 7.2.
Given a -oriented -ring , applying gives a homomorphism of -modules
The main theorems of this section are as follows:
Theorem 7.3.
Suppose that
is any -orientation of . Then the map is an isomorphism of -modules, and thus also of -modules.
Theorem 7.4.
Suppose that
is any -orientation of . Then the image of in maps surjectively onto the degree component of .
Remark 7.5.
Note that the map of 7.3 is a map of rank free -modules. Thus, it is an isomorphism if and only if its mod reduction is, which is a map of rank vector spaces over .
Similarly, the degree component is a rank vector space over , generated by . The map of 7.4 factors through the mod reduction of its domain, after which it becomes a map of rank vector spaces over .
Both Theorems 7.3 and 7.4 thus reduce to a question of whether maps of rank vector spaces over are isomorphisms. These maps are furthermore equivariant, or maps of -modules, with the actions of given by reduced regular representations. We will therefore find 7.7 below particularly useful. First, we recall some basic facts from representation theory.
Recollection 7.6.
Given two -modules and , the space inherits the structure of a -module via conjugation, where sends to . Then there is an identification
so that the transfer determines a linear map
Lemma 7.7.
Let denote the -module corresponding to the reduced regular representation of . Then a homomorphism
is an isomorphism if and only if is for any transferred homomorphism . More precisely, is a local -algebra, with maximal ideal the ideal of transferred homomorphisms.
Proof.
Note that is a uniserial -module, i.e. its submodules are totally ordered by inclusion. Since the endomorphism ring of a uniserial module over a Noetherian ring is local [Lam01, Proposition 20.20], the ring is local.
There is an identification , so we obtain a ring homomorphism
Since this homomorphism is clearly surjective, we learn that its kernel must be equal to the maximal ideal of .
On the other hand, for any and , we have
where the last equality follows from the fact that the transfer is zero on . It follows that lies in the maximal ideal of .
Finally, the equivalence
shows that the maximal ideal is equal to the image of for dimension reasons. ∎
Proof of 7.3.
Recall that is a free -module with basis and . In light of 7.5, it suffices to analyze the image of in its mod reduction, which is a free -module generated by the reductions of and . By combining 7.7 with 6.21, it suffices to show that a basis for this rank -module is given by the mod reduction of classes
Here, refers to the class of 6.18, which depends on the chosen -orientation. By combining 6.22 and 5.7, we may as well set to be . Using the formulas of [Sto12, Lemma 7.3] (cf. 5.8), we calculate
These clearly generate all of modulo , as desired. ∎
Proof of 7.4.
By arguments analogous to those in the previous proof, it suffices to check that
reduce to generators of the degree component of . By 6.22, we may assume that in is the element defined in Section 5.1. Under this assumption, the classes of interest become and its translates under the action on . As noted in 5.4, these span . ∎
References
- [AF78] Gert Almkvist and Robert Fossum. Decomposition of exterior and symmetric powers of indecomposable -modules in characteristic and relations to invariants. In Séminaire d’Algèbre Paul Dubreil, 30ème année (Paris, 1976–1977), volume 641 of Lecture Notes in Math., pages 1–111. Springer, Berlin, 1978.
- [AM78] Shôrô Araki and Mitutaka Murayama. -cohomology theories. Japan. J. Math. (N.S.), 4(2):363–416, 1978.
- [BBHS19] Agnes Beaudry, Irina Bobkova, Michael Hill, and Vesna Stojanoska. Invertible -Local -Modules in -Spectra. 2019. arXiv:1901.02109.
- [BC20] Prasit Bhattacharya and Hood Chatham. On the -orientability of vector bundles. 2020. arXiv:2003.03795.
- [BHSZ20] Agnes Beaudry, Michael Hill, XiaoLin Danny Shi, and Mingcong Zeng. Models of Lubin-Tate spectra via Real bordism theory. 2020. arXiv:2001.08295.
- [GH04] P. G. Goerss and M. J. Hopkins. Moduli spaces of commutative ring spectra. In Structured ring spectra, volume 315 of London Math. Soc. Lecture Note Ser., pages 151–200. Cambridge Univ. Press, Cambridge, 2004.
- [GM00] V. Gorbounov and M. Mahowald. Formal completion of the Jacobians of plane curves and higher real -theories. J. Pure Appl. Algebra, 145(3):293–308, 2000.
- [GM17] J. P. C. Greenlees and Lennart Meier. Gorenstein duality for real spectra. Algebr. Geom. Topol., 17(6):3547–3619, 2017.
- [HH16] Michael Hill and Michael Hopkins. Equivariant symmetric monoidal structures. 2016. arXiv:1610.03114.
- [HH18] Michael Hill and Michael Hopkins. Real Wilson Spaces I. 2018. arXiv:1806.11033.
- [HHR11] M. A. Hill, M. J. Hopkins, and D. C. Ravenel. On the 3-primary Arf-Kervaire invariant problem. 2011. Unpublished note available at https://web.math.rochester.edu/people/faculty/doug/mypapers/odd.pdf.
- [HHR16] M. A. Hill, M. J. Hopkins, and D. C. Ravenel. On the nonexistence of elements of Kervaire invariant one. Ann. of Math. (2), 184(1):1–262, 2016.
- [Hil06] Michael Anthony Hill. Computational methods for higher real K-theory with applications to tmf. ProQuest LLC, Ann Arbor, MI, 2006. Thesis (Ph.D.)–Massachusetts Institute of Technology.
- [Hil19] Michael Hill. Freeness and equivariant stable homotopy. 2019. arXiv:1910.00664.
- [HK01] Po Hu and Igor Kriz. Real-oriented homotopy theory and an analogue of the Adams-Novikov spectral sequence. Topology, 40(2):317–399, 2001.
- [HL16] Michael Hill and Tyler Lawson. Topological modular forms with level structure. Invent. Math., 203(2):359–416, 2016.
- [HLS18] Drew Heard, Guchuan Li, and XiaoLin Danny Shi. Picard groups and duality for Real Morava -theories. 2018. arXiv:1810.05439.
- [HM17] Michael A. Hill and Lennart Meier. The -spectrum and its invertible modules. Algebr. Geom. Topol., 17(4):1953–2011, 2017.
- [HMS17] Drew Heard, Akhil Mathew, and Vesna Stojanoska. Picard groups of higher real -theory spectra at height . Compos. Math., 153(9):1820–1854, 2017.
- [HS20] Jeremy Hahn and XiaoLin Danny Shi. Real orientations of Lubin-Tate spectra. Invent. Math., 221(3):731–776, 2020.
- [HSWX19] Michael A. Hill, XiaoLin Danny Shi, Guozhen Wang, and Zhouli Xu. The slice spectral sequence of a -equivariant height- Lubin-Tate theory. 2019. arXiv:1811.07960.
- [HY18] Michael A. Hill and Carolyn Yarnall. A new formulation of the equivariant slice filtration with applications to -slices. Proc. Amer. Math. Soc., 146(8):3605–3614, 2018.
- [KLW17] Nitu Kitchloo, Vitaly Lorman, and W. Stephen Wilson. Landweber flat real pairs and -cohomology. Adv. Math., 322:60–82, 2017.
- [Lam01] T. Y. Lam. A first course in noncommutative rings, volume 131 of Graduate Texts in Mathematics. Springer-Verlag, New York, second edition, 2001.
- [LLQ20] Guchuan Li, Vitaly Lorman, and J.D. Quigley. Tate blueshift and vanishing for Real oriented cohomology. 2020. arXiv:1910.06191.
- [LSWX19] Guchuan Li, XiaoLin Danny Shi, Guozhen Wang, and Zhouli Xu. Hurewicz images of real bordism theory and real Johnson-Wilson theories. Adv. Math., 342:67–115, 2019.
-
[Lur18]
Jacob Lurie.
Elliptic Cohomology II: Orientations.
2018.
Available at http://www.math.ias.edu/ lurie/. - [Mil60] J. Milnor. On the cobordism ring and a complex analogue. I. Amer. J. Math., 82:505–521, 1960.
- [Mor89] Jack Morava. Forms of -theory. Math. Z., 201(3):401–428, 1989.
- [Mos68] Robert E. Mosher. Some stable homotopy of complex projective space. Topology, 7:179–193, 1968.
- [MSZ20] Lennart Meier, XiaoLin Danny Shi, and Mingcong Zeng. Norms of Eilenberg-MacLane spectra and Real Bordism. 2020. arXiv:2008.04963.
- [Nav10] Lee S. Nave. The Smith-Toda complex does not exist. Ann. of Math. (2), 171(1):491–509, 2010.
- [PRS19] Wolfgang Pitsch, Nicolas Ricka, and Jerome Scherer. Conjugation Spaces are Cohomologically Pure. 2019. arXiv:1908.03088.
- [Qui69] Daniel Quillen. On the formal group laws of unoriented and complex cobordism theory. Bull. Amer. Math. Soc., 75:1293–1298, 1969.
- [Rav78] Douglas C. Ravenel. The non-existence of odd primary Arf invariant elements in stable homotopy. Math. Proc. Cambridge Philos. Soc., 83(3):429–443, 1978.
- [Ser67] J.-P. Serre. Local class field theory. In Algebraic Number Theory (Proc. Instructional Conf., Brighton, 1965), pages 128–161. Thompson, Washington, D.C., 1967.
- [Sil09] Joseph H. Silverman. The arithmetic of elliptic curves, volume 106 of Graduate Texts in Mathematics. Springer, Dordrecht, second edition, 2009.
- [Sto12] Vesna Stojanoska. Duality for topological modular forms. Doc. Math., 17:271–311, 2012.
- [Wil17a] Dylan Wilson. Equivariant, parametrized, and chromatic homotopy theory. 2017. Thesis (Ph.D.)–Northwestern University.
- [Wil17b] Dylan Wilson. On categories of slices. 2017. arXiv:1711.03472.