subsecref \newrefsubsecname = \RSsectxt \RS@ifundefinedthmref \newrefthmname = theorem \RS@ifundefinedlemref \newreflemname = lemma \newrefpropname=Proposition \newrefcorname=Corollary \newrefremname=Remark \newrefdefname=Definition \newrefexaname=Example \newrefthmname=Theorem \newreflemname=Lemma \newrefconjname=Conjecture
Ambidexterity and Height
Abstract
We introduce and study the notion of semiadditive height for higher semiadditive -categories, which generalizes the chromatic height. We show that the higher semiadditive structure trivializes above the height and prove a form of the redshift principle, in which categorification increases the height by one. In the stable setting, we show that a higher semiadditive -category decomposes into a product according to height, and relate the notion of height to semisimplicity properties of local systems. We place the study of higher semiadditivity and stability in the general framework of smashing localizations of , which we call modes. Using this theory, we introduce and study the universal stable -semiadditive -category of semiadditive height , and give sufficient conditions for a stable -semiadditive -category to be -semiadditive.

1 Introduction
1.1 Background & Overview
Chromatic homotopy theory springs from the deep and surprising connection between the -category of spectra and the stack of formal groups. In particular, the height filtration on the latter is mirrored by the “chromatic height filtration” on the former. This connection begins with Quillen’s work on the complex cobordism spectrum MU, showing that the ring carries the universal formal group law. Formal group laws admit a notion of (-typical) height for every prime . This notion can be defined in terms of a certain sequence of classes as follows: If vanish, then the height is , and if is invertible, then the height is . This algebraic filtration has a spectrum level manifestation in the form of the Morava -theories , which are certain MU-algebras with the property . This suggests that the -s are concentrated at height exactly , and the corresponding Bousfield localizations can then be considered as the “monochromatic layers” of the chromatic height filtration. The process of -localization can be loosely thought of as completion with respect to followed by the inversion of .
By the work of Hopkins, Devinatz, and Smith (see [HS98, Theorems 9 and 4.12]), the -operations can be inductively lifted to finite spectra, without MU-module structure. More precisely, a finite -local spectrum is said to have type , if is the lowest integer for which the -localization of does not vanish. Given a type finite spectrum , there exists a self map , which induces a power of on -homology. The cofiber of this self map is then a type spectrum. This procedure allows us to construct a sequence of -s as iterated Moore spectra, which we may suggestively write as follows:
Just as localization with respect to (for any ) has the effect of -completion, one can think of the localization with respect to , as completion with respect to . Furthermore, localization with respect to the spectrum , can be thought of as completion with respect to , followed by the inversion of . It is known that the -localization factors through the -localization, and that they coincide for MU-modules (and also in general when ). Furthermore, the localizations turn out to be independent of all the choices and thus naturally constitute another, potentially larger, candidate for the “monochromatic layers” of the chromatic height filtration. While the question of whether the inclusion is strict for is open (known as the Telescope Conjecture), both candidates for the “monochromatic layers” play a pivotal role in homotopy theory.
The localizations and are known to possess several rather special and remarkable properties. Among them, the vanishing of the Tate construction for finite group actions ([Kuh04, GS96, HS96, CM17]). In [HL13], Hopkins and Lurie reinterpret this Tate vanishing property as -semiadditivity, and vastly generalize it by showing that the -categories are -semiadditive. In turn, this is exploited to obtain new structural results for . In [CSY18, Theorem B], the authors extended on [HL13] by classifying all the higher semiadditive localizations of with respect to homotopy rings. First, for all such localizations, -semiadditivity was shown to be equivalent to -semiadditivity. Second, the telescopic localizations , for various primes and heights , were shown to be precisely the maximal examples of such localizations (while the are the minimal). Concisely put, in the stable world, the higher semiadditive property singles out precisely the monochromatic localizations, which are parameterized by the chromatic height.
In this paper, we introduce a natural notion of semiadditive height for higher semiadditive -categories, which in the examples and reproduces the usual chromatic height , without appealing to the theory of formal groups. We then proceed to show that the semiadditive height is a fundamental invariant of a higher semiadditive -category, which controls many aspects of its higher semiadditive structure, and the behavior of local systems on -finite spaces valued in it. We also show that the semiadditive height exhibits a compelling form of the “redshift principle”, where categorification has the effect of increasing the height exactly by one. When restricting to stable -categories, we show that higher semiadditive -categories decompose completely according to the semiadditive height, which accounts for the monochromatic nature of the higher semiadditive localizations of . Finally, building on the work of Harpaz [Har17], we introduce and study universal constructions of stable -semiadditive -categories of height , and initiate their comparison with the chromatic examples.
The present work should be viewed as part of a more extensive program that aims to place chromatic phenomena in the categorical context of the interaction between higher semiadditivity and stability. Apart from providing new tools for the study of , we believe that this approach can elucidate the chromatic picture and unfold the rich and intricate structure hidden within.
1.2 Main Results
Height Theory
Recall that ambidexterity is a property of a space with respect to an -category , that allows “integrating” -families of morphisms between pairs of objects in in a canonical way [HL13, Construction 4.0.7]. In particular, integrating the constant -family on the identity morphism of each object, produces a natural endomorphism of the identity functor of . We call the -cardinality of , and think of it as multiplication by the “size of ” (the actual meaning of which depends on ).
An -category is called -semiadditive if every -finite space is -ambidextrous. Our notion of semiadditive height is defined in terms of cardinalities of such spaces. For starters, let us begin with a -semiadditive (i.e. semiadditive) -local -category . If is invertible in , then is rational and we consider it to be of “height ”. In contrast, if all objects of are -complete, we consider it to be of “height ”. To proceed, let us assume that is -semiadditive for some . In such a case, we can consider the -cardinalities of Eilenberg-MacLane spaces:
The definition of semiadditive height uses the maps in a manner which is analogous to how the -self maps are used in the definition of the chromatic height:
Definition (Semiadditive Height, 3.1.6, 3.1.11).
For every , we write
-
(1)
, if is invertible in .
-
(2)
, if is complete with respect to 111by 3.1.9, , if and only if is -complete..
-
(3)
, if is of height and .
To show that the semiadditive height of and is indeed , we need to get a handle on the -local and -local cardinalities of the Eilenberg-Maclane spaces . In [CSY18, Lemma 5.3.3], we have already shown that
for the -category of -local -modules. Thus, this -category is of height . Since tensoring with is conservative on -local spectra, this also readily implies that is of height . However, to show that is of height , one has to know that the map , induced by -localization, detects invertibility of elements. This result was established in [CSY18, Propostion 5.1.17] using the notion of “nil-conservativity”. Thus, we get that is of height as well.
The notion of semiadditive height allows us to contextualize various aspects of the -categories and pertaining to the chromatic height. At the bottom of the hierarchy, the -category can be shown to be -semiadditive by elementary arguments. This is strongly related to the fact that all connected -finite spaces are -acyclic and the cardinality of any (non-empty) -finite space is invertible. Thus, the higher semiadditive structure of is in a sense “trivial”. The higher semiadditivity of and for is more subtle precisely because not all connected -finite spaces are acyclic, and not all cardinalities are invertible. One might roughly say, that the complexity of the higher semiadditive structure grows with the height. Our first main result formalizes this as follows:
Theorem A (Bounded Height, 3.2.7).
Let be an -semiadditive -local -category, which admits all -finite limits and colimits. If is of height , then
-
(1)
is -semiadditive.
-
(2)
For every -connected nilpotent -finite space , the map is invertible.
-
(3)
For every -connected -finite space and , the fold map is invertible.
-
(4)
For every principal fiber sequence of -finite spaces
if is -connected and nilpotent, then .
Informally speaking, A states that the invertibility of has the effect of “trivializing” the higher semiadditive structure at levels . In particular, it shows that it exists, which is point (1). From point (2), we deduce that having height implies having height , so the conditions are of decreasing strength as the terminology suggests. Point (3) articulates a useful categorical consequence (and, in fact, a characterization) of having height , which does not refer directly to the higher semiadditive structure. This can be seen as a generalization of [CSY18, Theorem E], which is essentially the special case . Finally, point (4) can be used to reduce the computation of the -cardinalities of nilpotent -finite spaces to those of -finite ones, under the assumption that is of height . The case produces an explicit formula, which recovers Baez-Dolan’s classical homotopy cardinality (2.2.2). We note that the possible failure of point (4) for the principal fiber sequence , is precisely the obstruction for to have height .
In their work on algebraic -theory of ring spectra, Ausoni and Rognes have discovered a phenomena which they dubbed “chromatic redshift”. Roughly speaking, it is the tendency of , which is a spectrum constructed from the -category of perfect -modules, to be of chromatic complexity larger by one, than the ring spectrum (appropriately measured). While more precise conjectures regarding this phenomena were subsequently formulated and studied, a conceptual source for the chromatic redshift phenomena seems to remain unrevealed. Our next result concerns an analogue of the redshift phenomena for the semiadditive height. In this context the increase by one in height is a formal consequence of categorification. To state this formally, we first note that the definition of semiadditive height makes sense for individual objects. Namely, an object in an -semiadditive -category is of height for some , if acts invertibly on and of height , if it is complete with respect to . Second, we exploit the fact that the -category , of -semiadditive -categories and -finite colimit preserving functors, is itself -semiadditive. Thus, given an -semiadditive -category , we can consider the height of being lower equal (resp. greater than) , as an object of which we shall denote by (resp. ).
Theorem B (Semiadditive Redshift, 3.3.2).
Let be an -semiadditive -category. We have that (resp. ), if and only if (resp. ).
The higher semiadditive structure of is essentially given by taking colimits over -finite spaces. Hence, B is closely related to point (3) of A. As a concrete example, we can consider for a -local ring spectrum , the -category of -local left -modules. The space of objects of this -category is a commutative monoid for the direct sum operation. Moreover, the higher semiadditivity of the -category of modules endows this space with a higher commutative monoid structure in the sense of [Har17, Definition 5.10]. As a consequence of B, this higher commutative monoid is of height in the -category of higher commutative monoids. In a future work, we shall investigate the implications of this to the chromatic redshift in algebraic -theory in the sense of Ausani-Rognes.
Our main interest in the notion of higher semiadditivity is in its application to stable -categories. As it turns out, the two properties of higher semiadditivity and stability interact in a highly non-trivial way. First and foremost, in the presence of stability, the higher semiadditive structure turns out to decompose completely according to height:
Theorem C (Height Decomposition, 4.2.7).
Let be a stable idempotent complete -semiadditive -category for some . There is a canonical equivalence
were and are the full subcategories of objects of height and respectively.222We also treat the case , which is somewhat more subtle.
This result sheds light on the “monochromatic nature” of higher semiadditive phenomena in the stable world. Loosely speaking, the fact that the monochromatic layers, which have different heights, glue non-trivially (by means of the chromatic fracture square), obstructs the higher semiadditivity of non-monochromatic localizations of spectra.
In view of C, it makes sense to focus our attention on stable -categories of height exactly . In [HL13, Section 5.4] it is shown that the behavior of local systems of -local spectra on a -finite space , strongly depends on the level of connectedness of compared with . We show that some of these results hold for general stable -categories of height exactly . First of all, from A(3), it can be deduced that for an -connected -finite space , the inclusion functor of constant local-systems is fully faithful. The right orthogonal complement consists of local-systems whose global sections object (i.e. limit over ) vanishes. We prove the following:
Theorem D (Semisimplicity, 4.3.2).
Let be a stable -semiadditive -category such that , and let be an -connected -finite space. There is a canonical equivalence .
This result can be seen as a generalization of the “semisimplicity” of -valued local systems on -connected -finite spaces (compere [Lurb]). We also provide an explicit formula for the composition , as a “symmetrization” of the action of the -group . Intuitively, the “order” of , by which one has to divide, is precisely the -cardinality of , which is invertible by the assumption on the height of and the connectivity of (A).
Based on the classification of higher semiadditive localizations of with respect to homotopy rings in [CSY18, Theorem B], the authors proposed the conjecture that every stable -local presentable -semiadditive -category is in fact -semiadditive [CSY18, Conjecture 1.1.5]. In this paper, we prove a partial result in the direction of this conjecture. Given a stable -local presentable -category , we say that an object is of finite stable height if there exists a non-zero finite -local spectrum , such that . We also denote by the full subcategory of objects , for which for all of finite stable height.
Theorem E (Bounded Bootstrap, 5.5.17).
Let be a stable -local presentable -category. If is -semiadditive and , then it is -semiadditive. Moreover, in this case .
The condition is satisfied if for example for every , the mapping spectrum is -local for some integer . The proof of E, relies on the theory of modes, which we shall review next.
Mode Theory
In [Lura, Proposition 4.8.1.15], Lurie introduced a symmetric monoidal structure on the -category of presentable -categories and colimit preserving functors. Moreover, he showed that many familiar properties of presentable -categories can be characterized as having a (necessarily unique) module structure over certain idempotent algebras in [Lura, Section 4.8.2]. We call such idempotent presentable -categories modes. This notion was also considered in [GGN16] from the perspective of smashing localizations of . Given a mode , it is a property of a presentable -category to have a structure of a module over . The terminology is inspired by the idea that modes classify the possible “modes of existence” in which mathematical objects can occur, manifest, and behave. Most notably, the property of stability is equivalent to having a module structure over . Consequently, every stable presentable -category is canonically enriched in and colimit preserving functors between stable presentable -categories preserve this enrichment. This structure naturally plays a significant role in the study of stable -categories. In [Har17, Lemma 5.20], Harpaz showed that -semiadditivity is similarly characterized by having a module structure over the idempotent algebra of -commutative monoids. The case recovers the usual -category of commutative (i.e. ) monoids in spaces, which classifies ordinary semiadditivity. The mapping spaces of an -semiadditive -category obtain a canonical -commutative monoid structure, by analogy with the -enrichment of stable -categories.
In the final section of this paper, we develop the theory of modes further and apply it to the study of height in stable presentable higher semiadditive -categories. First, by the general theory of modes, is also a mode, which classifies the property of being at the same time stable and -semiadditive. Furthermore, using C, we show:
Theorem F (5.3.6).
For every , there exists a mode n333The letter (pronounced “tsadi”) is the first letter in the Hebrew word for “color”. The notation was chosen to indicate the close relationship with chromatic homotopy theory., which classifies the property of being stable, -local, -semiadditive and of height .
It is natural to compare n with , which is in a sense the universal -local height localization of spectra. Since is also -semiadditive and of semiadditive height , the theory of modes implies the existence of a unique colimit preserving symmetric monoidal functor . In the case , the functor is an equivalence and hence (5.3.7). In general, we show that exhibits as a smashing localization of n in the sense that admits a fully faithful right adjoint and there is a canonical isomorphism for all (5.5.14). For , the -category n also resembles in that the unique colimit preserving symmetric monoidal functor vanishes on all bounded above spectra (5.3.9), and that the right adjoint of the unique colimit preserving symmetric monoidal functor is conservative (5.3.10). We consider n to be a natural extension of , which is a universal home for phenomena of height .
In a previous draft of this paper, we proposed the conjecture that for every , the unique colimit preserving symmetric monoidal functor is an equivalence. However, this conjecture was soon disproved by Allen Yuan already in the case . More precisely, using the Segal Conjecture (now a theorem [Car84]), he has constructed a higher commutative monoid structure of height on the -complete sphere, as an object of the -category of -complete spectra. The details and some interesting applications of this example will appear in a separate paper by him.
Finally, the theory of modes allows us not only to analyze the implications of certain properties of presentable -categories, but also to enforce them in a universal way. For every mode and a presentable -category we can view as the universal approximation of by a presentable -category which satisfies the property classified by . For example, is the stabilization of [Lura, Example 4.8.1.23], and similarly, is the “-semiadditivization” of [Har17, Corollary 5.18]. As alluded to above, the non-trivial gluing in the chromatic fracture square, prevents from being higher semiadditive for . Employing the additive -derivation on the rings constructed in [CSY18, Section 4], we show that forcing even -semiadditivity on , has the effect of “dissolving the glue” in the chromatic fracture squares:
Theorem G (1-Semiadditive Splitting, 5.4.10).
For every , there is a unique equivalence of presentably symmetric monoidal -categories
In particular, we see that forcing -semiadditivity on makes it automatically -semiadditive. Noticing that both sides of G are modes, we can reinterpret it in terms of the properties classified by them. Namely, that every -semiadditive stable presentable -category whose mapping spectra are -local, is -semiadditive. With some additional effort, we deduce from it the stronger statement of E.
1.3 Organization
We shall now outline the content of each section of the paper.
In section 2, we recall and expand the theory of higher semiadditivity. We discuss the notion of cardinality for a -finite space in a higher semiadditive -category and the corresponding notion of amenability. We then give several examples of these notions in various higher semiadditive -categories, and relate the notion of amenability to the behavior of local systems, through the notion of acyclicity.
In section 3, we discuss the main notion of this paper, that of height in a higher semiadditive -category, defined in terms of the cardinalities of Eilenberg-MacLane spaces. We show that the higher semiadditive structure trivializes above the height (A) and exhibits a redshift principle of increasing by one under categorification (B).
In section 4, we study semiadditivity and height for stable -categories. After a general discussion on recollement, we show that a stable higher semiadditive -category splits as a product according to height (C). We then study local systems valued in a stable higher semiadditive -category of height and show how the notion of height is related to the phenomenon of semisimplicity of local systems (D). Finally, we use nil-conservative functors to show that semiadditive and chromatic height coincide for monochromatic localizations of spectra.
In section 5, we study the theory of modes, i.e. that of idempotent algebras in the category of presentable -categories. We show how algebraic operations on modes, such as tensor product and localization, translate into operations on the properties of presentable -categories classified by them. We then show that the main notions studied in this paper, higher semiadditivity and height, together with the more classical notion of chromatic height, are all encoded by modes (e.g. F). Using this theory, we study the interaction between the chromatic and the semiadditive heights through the interactions between the corresponding modes. In particular, we prove G and deduce from it E.
1.4 Conventions
Throughout the paper, we work in the framework of -categories (a.k.a. quasicategories), and in general follow the notation of [Lur09] and [Lura]. We shall also use the following terminology and notation most of which is consistent with [CSY18]:
-
(1)
We slightly diverge from [Lur09] and [Lura] in the following points:
-
(a)
We use the term isomorphism for an invertible morphism in an -category (i.e. equivalence).
-
(b)
We denote by the maximal -subgroupoid of an -category .
-
(c)
We write for the -category of presentable -categories and colimit preserving functors denoted in [Lur09] by .
-
(d)
We denote by the subcategory spanned by stable -categories and exact functors. Similarly, we denote by the full subcategory spanned by stable presentable -categories.
-
(a)
-
(2)
We say that a space is
-
(a)
-finite for , if and is contractible, or , the set is finite and all the fibers of the diagonal map are -finite444For , this is equivalent to having finitely many components, each of them -truncated with finite homotopy groups..
-
(b)
-finite or -finite, if it is -finite for some integer . For , we denote by the full subcategory spanned by -finite spaces.
-
(c)
-space, if all the homotopy groups of are -groups.
-
(a)
-
(3)
Given an -category ,
-
(a)
For every map of spaces , we write for the pullback functor and and for the left and right adjoints of whenever they exist.
-
(b)
Whenever convenient we suppress the canonical equivalence of -categories by identifying a space with the terminal map . In particular, for every -category , we write for and similarly and for and whenever they exist.
-
(c)
For every we write for and denote the fold (i.e. counit) map by . Similarly, we write for and denote the diagonal (i.e. unit) map by .
-
(a)
-
(4)
Given a map of spaces , we denote for every , the homotopy fiber of over by . We say that
-
(a)
an -category admits all -limits (resp. -colimits) if it admits all limits (resp. colimits) of shape for all .
-
(b)
a functor preserves -colimits (resp. -limits) if it preserves all colimits (resp. limits) of shape for all .
-
(a)
-
(5)
For every ,
-
(a)
by -finite (co)limits we mean (co)limits indexed by an -finite space.
-
(b)
We let (resp. ) be the subcategory spanned by -categories which admit -finite colimits (resp. limits) and functors preserving them.
-
(c)
For (resp. ) we write (resp. ) for the full subcategory of spanned by the -finite colimit (resp. limit) preserving functors.
-
(d)
We let be the subcategory spanned by the -semiadditive -categories and -semiadditive (i.e. -finite colimit preserving) functors.
-
(e)
Given an -operad , we say that is compatible with -indexed colimits for some collection of -categories , if admits -indexed colimits and every tensor operation of preserves -indexed colimits in each variable.
-
(f)
An -semiadditively -monoidal -category is an -monoidal -semiadditive -category which is compatible with -finite colimits.
-
(a)
-
(6)
If is a monoidal -category and is an -category enriched in , we write for the -mapping object of . We omit the subscript or superscript when they are understood from the context. In particular, when is closed, means . For every -category we have .
1.5 Acknowledgments
We would like to thank Tobias Barthel, Clark Barwick, Agnès Beaudry, Jeremy Hahn, Gijs Heuts, Mike Hopkins, and Tyler Lawson for useful discussions. We thank Allen Yuan for sharing with us his ideas regarding the topics of this paper. We also like to thank the entire Seminarak group, especially Shay Ben Moshe, for useful comments on the paper’s first draft.
The first author is supported by the Adams Fellowship Program of the Israel Academy of Sciences and Humanities. The second author is supported by ISF1588/18 and BSF 2018389.
2 Semiadditivity
In this section, we collect general facts regarding the notion of ambidexterity and its implications. We begin by reviewing some background material and most importantly (re)introduce the notion of cardinality for ambidextrous -finite spaces. We provide a variety of examples of cardinality in both the stable and the unstable settings, including those of relevance to chromatic homotopy theory. Of particular importance are the amenable spaces, whose cardinality is invertible. We continue the study of such spaces, which we began in [CSY18, Section 3], and in particular, establish its implications for the behavior of certain -finite limits and colimits. In the next section, the amenability of Eilenberg-MacLane spaces will play a central role in the definition of “semiadditive height” for higher semiadditive -categories, which is the main subject of this paper.
2.1 Preliminaries
In this subsection, we review some basic definitions and facts regarding ambidexterity from [HL13, Section 4], cardinality from [CSY18], and higher commutative monoids from [Har17, Section 5.2]. This subsection serves mainly to set up notation, terminology, and a convenient formulation of fundamental results.
Ambidexterity
Recall from [HL13, Section 4.1] the definition of ambidexterity:
Definition 2.1.1.
Let . A -finite map is called:
-
(1)
weakly -ambidextrous if it is an isomorphism, or is -ambidextrous.
-
(2)
-ambidextrous if it is weakly -ambidextrous, admits all -limits and -colimits and the the norm map is an isomorphism.
2.1.1 should be understood inductively on the level of truncatedness of . A -finite map, i.e. an isomorphism, is always -ambidextrous. If is -finite, then the diagonal map
is -finite and the ambidexterity of allows in turn the definition of by [HL13, Construction 4.1.8] (see also [CSY18, Definition 3.1.3]).
Remark 2.1.2.
A map is -ambidextrous if and only if all the fibers of are -ambidextrous spaces [HL13, Corollary 4.2.6(2), Corollary 4.3.6]. Moreover, the fibers of the diagonal are the path spaces of . In other words, is weakly -ambidextrous if and only if the path spaces of are -ambidextrous. Thus, -ambidexterity is ultimately a property of spaces.
By [HL13], the property of ambidexterity has the following useful characterization, which avoids the explicit inductive construction of the norm map:
Proposition 2.1.3.
Let be an -category and let be a -finite map. The map is -ambidextrous if and only if the following hold:
-
(1)
is weakly -ambidextrous.
-
(2)
admits all -limits and -colimits.
-
(3)
Either preserves all -colimits or preserves all -limits.
Proof.
By 2.1.2, we may assume that , in which case it is essentially [HL13, Proposition 4.3.9] and its dual [HL13, Remark 4.3.10]. We note that while the claim in [HL13] is stated under the stronger assumption that admits, and (resp. ) preserves, all small colimits (resp. limits), the proof uses only -colimits (resp. -limits). ∎
As a consequence, we can easily deduce that ambidexterity enjoys the following closure properties with respect to the -category:
Proposition 2.1.4.
Let be an -category and let be a -finite -ambidextrous space. The space is also -ambidextrous for:
-
(1)
the opposite -category of .
-
(2)
for an -category .
-
(3)
containing the final object and closed under -limits for all and .
-
(4)
containing the initial object and closed under -colimits for all and .
Proof.
First, (4) follows from (3) and (1), so it suffices to consider (1)-(3). In all cases, we proceed by induction on , so we may assume by induction that is weakly -ambidextrous. By 2.1.3, it suffices to verify that -limits and -colimit exist in and that the functor preserves -colimits or that preserves -limits. For (1), the claim follows from the fact that limits in are computed as colimits in and vice versa. For (2), we use the fact that limits and colimits in are computed pointwise. For (3), since -limits in coincide with -colimits in , it follows that -limits and -colimits are computed in in the same way as in . ∎
Cardinality
The main feature of ambidexterity is that it allows us to integrate families of morphisms in . That is, given a -ambidextrous map and we have a map (see [CSY18, Definition 2.1.11])
When , we can think of an element of as a map , and of as the sum of over the points of . In particular, we can integrate the identity morphism:
Definition 2.1.5.
Let and let be a -ambidextrous map. We have a natural transformation given by the composition
For a -ambidextrous space , we write and call the -cardinality of .
The name “-cardinality” can be explained as follows. For a given object , the map equals . Thus, we think of as the sum of the identity of with itself “ times”. Or in other words, as the result of multiplying by the “cardinality of ” on . The basic example which motivates the terminology and notation is the following:
Example 2.1.6.
Let be a semiadditive -category. For a finite set , viewed as a -finite space, the operation is simply the multiplication by the natural number, which is the cardinality of in the usual sense.
Remark 2.1.7.
For a general -ambidextrous map , the transformation can be understood as follows. Let be a -family of objects in . For each , let us denote the fiber of at by and the evaluation of at by . By [CSY18, Proposition 3.1.13], the map acts on each by . In other words, is just the -family of -cardinalities of the -family of spaces .
For a -ambidextrous space , the -limits and -colimits in are canonically isomorphic. This can be used to show the following:
Proposition 2.1.8.
Let and let be a - and -ambidextrous space. A functor preserves all -limits if and only if it preserves all -colimits. Moreover, if preserves all -(co)limits, then
Proof.
The -cardinality is additive in the following sense:
Proposition 2.1.9.
Let and be a map of spaces. If and are -ambidextrous then is -ambidextrous and for every ,
Informally, 2.1.9 says that the cardinality of the total space is the “sum over ” of the cardinalities of the fibers of .
Proof.
This follows from the Higher Fubini’s Theorem ([CSY18, Propostion 2.1.15]) applied to the identity morphism. ∎
Remark 2.1.10.
The various naturality properties enjoyed by the operations allow for useful abuses of notation:
-
(1)
Given an -colimit preserving functor , if is - and -ambidextrous, we get by 2.1.8 that . We therefore write , dropping the subscript , whenever convenient. As a special case, let be a full subcategory which is closed under -(co)limits. The cardinality coincides with the restriction of to .
-
(2)
When is monoidal and the tensor product preserves -colimits in each variable, the action of on any object can be identified with tensoring with (see [CSY18, Lemma 3.3.4]). We therefore sometimes identify with an element of .
-
(3)
Furthermore, for , the map can be also identified with multiplication by the image of under the map , which we also denote by .
All these abuses of notation are compatible with standard conventions when is a finite set (see 2.1.6).
Higher commutative monoids
Of particular interest are -semiadditive -categories, i.e. those for which all -finite spaces are ambidextrous. For , we recover the ordinary notion of a semiadditive -category. The central feature of semiadditive -categories is the existence of a canonical summation operation on their spaces of morphisms, endowing them with a commutative monoid structure. In [Har17, Section 5.2], an analogous theory of -commutative monoids is developed and applied to the study of -semiadditivity for all . In this section, we recall from [Har17] a part of this theory of higher commutative monoids and extend it to the case .
Definition 2.1.11.
[Har17, Definition 5.10, Proposition 5.14] Let , for , the -category of -commutative monoids in is given by
In the case , we simply write and refer to its objects as -commutative monoids555For , one indeed recovers the usual notion of a commutative (i.e. ) monoids in spaces by comparison with Segal objects [Lura, Section 2.4.2]..
In the case , evaluating at , the unique object of , gives an equivalence .
Remark 2.1.12.
We can understand the definition of as follows. An object consists of the “underlying space” , together with a collection of coherent operations for summation of -finite families of points in it. Indeed, for every , there is a canonical equivalence . Given in , the “right way” map is given simply by restriction, while the “wrong way” map encodes integration along the fibers. The functoriality with respect to composition of spans encodes the coherent associativity and commutativity of these integration operations.
Higher commutative monoids of different levels are related by “forgetful functors”.
Proposition 2.1.13.
Let and let . The restriction along the inclusion functor
induces a limit preserving functor
Proof.
It suffices to show that the preserves -finite colimits. By [Har17, Corollary 2.16], it suffices to show that the composite
preserves -finite colimits. Indeed, factors also as the composite
The first functor clearly preserves -finite colimits while the second one preserves -finite colimits by [Har17, Proposition 2.12]. ∎
We now extend the definition of to .
Definition 2.1.14.
For , we denote
As above we write .
When is presentable, is presentable for all , by [Har17, Lemma 5.17]. Moreover, can then be described as the colimit of in .
Lemma 2.1.15.
For , the forgetful functors
admit left adjoints and the colimit of the sequence
in is . In particular, is presentable.
Proof.
Since is presentable for all , by the adjoint functor theorem [Lur09, Corollary 5.5.2.9], the functor admits a left adjoint if and only if it is accessible and limit preserving. The functor is -accessible for any large enough such that -finite limits commute with -filtered colimits in , and by 2.1.13, is limit preserving. The second claim follows from the description of colimts in (see [Lur09, Theorem 5.5.3.18, Corollary 5.5.3.4]). ∎
In the theory of -semiadditivity, the -category plays an analogous role to that of the -category of commutative (i.e. ) monoids in spaces, in the the theory of ordinary (i.e. ) semiadditivity. In particular, the mapping space between every two objects in an -semiadditive -category has a canonical -commutative monoid structure. To see this, we begin by recalling the fundamental universal property of from [Har17]. For each , we have a forgetful functor , given by evaluation at .
Proposition 2.1.16.
Let . For every and , post-composition with the forgetful functor induces an equivalence of -categories
Proof.
As a corollary, for every -semiadditive -category we have a unique lift of the Yoneda embedding to a -enriched Yoneda embedding.
Corollary 2.1.17.
Let . For every , there is a unique fully faithful -semiadditive functor
whose composition with the forgetful functor is the Yoneda embedding.
Proof.
Taking in 2.1.16, shows that the ordinary Yoneda embedding
lifts uniquely to a fully faithful -finite limit preserving functor
∎
The -enriched Yoneda embedding corresponds to a functor
whose composition with the forgetful functor is the functor . Thus, we obtain a canonical structure of an -commutative monoid on each mapping space in . Informally, the “wrong way” maps for , in the higher commutative monoid structure on , are given by integration
Remark 2.1.18.
It is overwhelmingly likely that an -semiadditive -category can be canonically enriched in (for e.g. in the sense of [GH15] or [Hin16]), such that the -valued mapping objects coincide with our definition above. In case is further assumed to be presentable, this follows from the fact that is left tensored over (see [Har17, Lemma 5.20], 5.3.1 and [Lura, Proposition 4.2.1.33]).
2.2 Examples
We now review some examples of -semiadditive -categories and the behavior of cardinalities of -finite spaces in them.
Universal
It is proved in [Har17], that the following is the universal example of an -semiadditive -category. In particular, it shows that in general, the operations need not reduce to something “classical”:
Example 2.2.1 (Universal case).
Informally, the universality of 2.2.1 is reflected in its construction as follows. The collection of spaces is generated under -finite colimits from the point . The “right way” maps in encode the usual covariant functoriality of these colimits. The “wrong way” maps in encode the contravariant functoriality arising from “integration along the fibers”.
A closely related example is the -category of -commutative monoids, which is shown in [Har17, Corollary 5.19, Corollary 5.21], to be the universal presentable -semiadditive -category. The Yoneda embedding induces a fully faithful -semiadditive (symmetric monoidal) functor
taking each -finite space to the “free -commutative monoid” on . From this we get that cardinalities in are computed essentially in the same way as in 777The relation between and is somewhat analogues to the relation between the -category of finite spectra and the -category of all spectra. . In section 5, we shall discuss more systematically the universality of (see 5.3.1).
Rational
There are however some situations in which the operations can be expressed in terms of classical invariants.
Example 2.2.2 (Homotopy cardinality).
For a -finite space , Baez and Dolan [Bae] define the homotopy cardinality of to be the following non-negative rational number
(1) |
This notion can be seen as a special case of the cardinality of a -finite space in a higher semiadditive -category as follows. We say that an -category is semirational if it is 0-semiadditive and for each , multiplication by is invertible in (e.g. or ). We shall see that a semirational -category which admits all 1-finite colimits is automatically -semiadditive and for every -finite space , we have
We note that formula (1) is completely determined by the additivity of the cardinality under coproducts and the following “multiplicativity property”: For every fiber sequence of -finite spaces
such that is connected, we have 888This follows from the long exact sequence in homotopy groups, and is reminiscent of the “additivity property” of the Euler characteristic. .
Remark 2.2.3.
In fact, we shall prove in 2.3.4 a somewhat sharper result. Let be a -semiadditive -category, which admits -finite colimits, and let be a -finite space. If satisfies the following condition:
-
The orders of the homotopy groups of are invertible on all objects of
Then is -ambidextrous and .
Chromatic
Examples of -semiadditive -categories of “higher height” appear naturally in chromatic homotopy theory. For a given prime and , let be the Morava -theory spectrum of height at the prime . One of the main results of [HL13] is that the localizations are -semiadditive. In particular, we can consider -local cardinalities of -finite spaces. For , we have and we recover the homotopy cardinality (2.2.2). Similarly, since is -local for all , for every -finite space whose homotopy groups have cardinality prime to , the -local cardinality of coincides with its homotopy cardinality for all (see 2.2.3). In particular, it is independent of . However, for the prime is not invertible in . Thus, there are -finite spaces (e.g. -finite -spaces), which are ambidextrous even though they do not satisfy condition of 2.2.3. For such spaces , the -local cardinality does depend on and in general does not (and can not) agree with the homotopy cardinality.999Note that the rationalization functor does not preserve colimits in general and so does not preserve cardinalities. It does however preserve colimits which are indexed on -finite spaces whose homotopy groups have order prime to .
To study the -local cardinalities of -finite spaces, it is useful to consider their image in Morava -theory. For every integer , we let be the Morava -theory associated with some formal group of height over , viewed as an object of . In particular, we have a (non-canonical) isomorphism
Example 2.2.4 (Chromatic cardinality).
The -category is -semiadditive [CSY18, Theorem 5.3.1] and hence one can consider cardinalities of -finite spaces in . We define the (-typical) height cardinality of to be
It makes sense to consider as , in which case we recover the homotopy cardinality (2.2.2). The technology of [Lur19] allows one to derive a rather explicit formula for , for heights as well. Let be the -adic free loop space of . One can show that the element belongs to the subring and satisfies . Applying this relation inductively we obtain the formula
(2) |
If happens to be a -space, then coincides with the ordinary free loop space . Thus, can be computed as the homotopy cardinality of the space of maps from the -dimensional torus to .
We shall not get here into the details of how formula (2) is deduced from the results of [Lur19], as we shall only need the following special case:
Proposition 2.2.5.
For all we have .101010For height this should be interpreted via the identity
This was proved independently in [CSY18, Lemma 5.3.3] by relating the cardinality to the symmetric monoidal dimension. However, we shall use the general formula (2) in some examples to illustrate interesting phenomena.
While the structure of the rings is not entirely understood in general, it follows from [BG18] and [BGH17] that:
Proposition 2.2.6.
For all and , the image of the unit map is and the kernel is precisely the nil-radical.
Proof.
Let be the Morava stabilizer group associated with . We have an action of on by commutative algebra maps and thus, the map factors through the fixed points . By [BG18, Lemma 1.33], we have
By [BGH17, Theorem 2.3.5], the -page of the descent spectral sequence
has a horizontal vanishing line. Since the spectral sequence is multiplicative, this implies that all elements in with positive filtration degree are nilpotent. Finally, since admits a ring map from the -complete sphere, the map is surjective. ∎
Thus, for every -finite space , the identity
holds up to nilpotents. We do not know, however, whether it holds in .
Categorical
Another family of examples of higher semiadditive -categories arises from category theory itself.
Proposition 2.2.7.
For every the -category is -semiadditive.
Proof.
The case is exactly [Har17, Proposition 5.26]. We now wish to show that is -semiadditive for every . By [Lura, Remark 4.8.1.6], both and admit closed symmetric monoidal structures, and by [Lura, Proposition 4.8.1.3], there exists a symmetric monoidal functor . By [Lura, Remark 4.8.1.8] and [Lur09, Proposition 5.3.6.2(2)], admits a right adjoint and thus preserves colimits. Hence, is -semiadditive by [CSY18, Corollary 3.3.2(2)].∎
Example 2.2.8 (Categorical cardinality).
Let and let . For every -finite space the -semiadditive structure of gives rise to a functor When it is shown in [Har17, Section 5.2] that is given by taking the constant colimit on . That is, it takes an object to the object . Since the forgetful functor preserves limits, and hence m-semiadditive, the same holds for . This is very suggestive of the idea that “multiplication by on ” is given by “summing each object with itself times”. A closely related example is discussed in [HL13, Example 4.3.11], where it is shown that is -semiadditive (in fact, every -truncated space, not necessarily -finite, is -ambidextrous).
Remark 2.2.9.
There is a different approach to the higher semiadditivity of , based on the notion of ambidextrous adjunctions of -categories. We sketch the argument to demonstrate the role of higher categorical structures as a useful perspective on ambidexterity phenomena. Given an -finite space , the adjunction
can be naturally promoted to an adjunction of -categories. Moreover, the unit and counit of , as -morphisms in the respective -categories of endofunctors, can be shown to have left adjoints and respectively. Thus, we are in a situation which is dual to the notion of an ambidextrous adjunction of [HSSS18, Definition 2.1]. It follows by an elementary argument that and exhibit as a left adjoint of (see [HSSS18, Remark 2.2]). Hence, by 2.1.3, the space is -ambidextrous and so is -semiadditive.
The functor that takes an -category to its opposite induces an equivalence . Hence, the -category is -semiadditive as well and the higher semiadditive structure is given by taking limits. For every -category with finite (co)products, the (co)product endows the space of objects with a (co)Cartesian commutative monoid structure. Using the -semiadditivity of and together with the -enriched Yoneda embedding provided by 2.1.17, this too can be generalized to all . Given , by the -semiadditivity of , the mapping space admits a canonical structure of an -commutative monoid. On the other hand, since is freely generated from a point under -finite colimits [Lura, Notation 4.8.5.2], we have
Definition 2.2.10.
For , we refer to the above -commutative monoid structure on the space of objects as the coCartesian structure. A completely analogous construction endows the space of objects of each with a Cartesian -commutative monoid structure.
As explained in [Har17, Section 5.2], the integration operations for the (co)Cartesian -commutative monoid structure on are given by taking -finite (co)limits. Finally, the -category of -semiadditive -categories and -semiadditive functors is a full subcategory of both and . Moreover, we have the following:
Proposition 2.2.11.
Let . The full subcategory (resp. ) is closed under colimits and in particular is -semiadditive.
2.3 Amenability
Let be a -ambidextrous map. Recall from [CSY18, Definition 3.1.7], that is called -amenable if is invertible. As with ambidexterity, amenability is a fiber-wise property ([CSY18, Corollary 3.1.16]). Thus, we shall be mainly interested in -amenable spaces. i.e. those whose -cardinality is invertible.
Remark 2.3.1.
We warn the reader not to confuse the condition that the natural transformation is invertible with the condition that the natural transformation is invertible, which is equivalent to being -ambidextrous (which is itself a prerequisite for defining ).
In this section, we extend some results from [CSY18] regarding amenability. The main point is that while -ambidexterity allows us to sum over -families of maps, -amenability allows us to average over -families of maps, which in turn facilitates “transfer arguments” along . We shall explore how this condition affects the higher semiadditive structure.
Closure properties
For a map of spaces, the condition of -amenability, as the condition of -ambidexterity, is fiber-wise. However, unlike -ambidextrous maps, -amenable maps are not closed under composition. To understand the situation, it suffices to consider the case . By the additivity of cardinality (2.1.9), we get . Assume for simplicity that is connected and that the fiber of is . The transformation equals at each point . Thus, it is tempting to presume that and hence if both and are invertible, then so is . However, for this reasoning to hold we need to know that is constant on with value . Alas, in general is not constant, even when is invertible, and need not equal (see 2.3.6). On the positive side, we show that must be constant if we require in addition to the invertibility of that is a principal fiber sequence.
Definition 2.3.2.
We call a map of spaces principal if it can be extended to a fiber sequence .
We note that for a principal map all the fibers are isomorphic even if the target is not connected.
Proposition 2.3.3.
Let and let be a principal -ambidextrous map of -ambidextrous spaces with fiber . For , if is invertible, then
and
Proof.
The base change of along itself is a map , which is a principal map with a section. Therefore is isomorphic to the projection Hence, , where is the projection. We get from [CSY18, Corollary 3.1.14] that
If is invertible, then is invertible, and thus by the above, we get ([CSY18, Proposition 3.1.13])
We can now integrate along and get
∎
As a simple application, we deduce that when the homotopy groups of a -finite space have invertible cardinality in , the notion of cardinality degenerates to the homotopy cardinality (2.2.3).
Proposition 2.3.4.
Let and let be a -finite space. Assume that admits -colimits and the order of each homotopy group of is invertible in . Then, is -ambidextrous and . In particular, if is connected, then it is -amenable.
Proof.
By 2.1.10, we may assume that is connected. We prove the claim for all connected -finite by induction on , where the case is trivial. We prove the claim for some , assuming it holds for . Choose a base point in and consider the connected component of the identity loop. The space satisfies the assumptions of the inductive hypothesis. Hence, is amenable and . Now, is just the coproduct of copies of . Hence, by 2.1.10 and the inductive hypothesis, we have
Moreover, is invertible as is invertible by assumption and is invertible by the inductive hypothesis. That is, is -amenable. Finally, consider the principal fiber sequence
Since is -amenable, by [CSY18, Proposition 3.1.17], the space is -ambidextrous and by 2.3.3, we have
∎
From 2.3.3, we also deduce that -amenable maps are partially closed under composition:
Corollary 2.3.5.
Let and let be a pair of composable maps of spaces. If and are -amenable and is principal, then is -amenable.
Proof.
We can check that is -amenable, by pulling back along for every point of . In other words, we can assume that . Taking to be the fiber of , we have a principal fiber sequence where and are -amenable; thus the result follows from 2.3.3. ∎
Counter examples
We conclude this subsection with a discussion of the necessity of the conditions in 2.3.3. For starters, if is not invertible, then the identity is (very much) false in general. The following examples show that the condition on the fiber sequence to be principal can also not be dropped. The first example shows that -cardinality need not be multiplicative even when the fiber and base space are -amenable. Moreover, in such case, the total space need not even be -amenable and so in particular, -amenable maps are not closed under composition.
Example 2.3.6.
Let be an odd prime and let . We consider the map classifying the cup-square operation on mod cohomology, and the associated fiber sequence
The only non-trivial homotopy groups of are , but the Postnikov invariant represented by in is non-zero. Using 2.2.4, we have and we can compute it using the fiber sequence
via the induced long exact sequence on homotopy groups. The only complication arises at the level of , where we need to compute the size of the kernel of the cup-square map
Namely, the number of -dimensional 2-forms over that square to zero. Since this is the number of 2-forms of rank lower or equal 1, one can write down an explicit combinatorial formula for it. This leads to the following explicit formula:
In particular, taking , we get , which is an invertible element in . It follows that is -amenable. Nevertheless,
which differs from
Moreover, and are both -amenable, but is not. Thus, the maps and are -amenable, but their composition is not.
The next example shows that -cardinality need not be multiplicative when the fiber and total space are -amenable. Moreover, the base space in this case need not be -amenable and so in particular the class of -amenable maps does not satisfy “left cancellation” (compare [CSY18, Theorem 2.4.5]).
Example 2.3.7.
Let and . Consider the (non-principal) fiber sequence
It can be shown using 2.2.4, that for we have
In particular,
Moreover, and are -amenable, but is not. Thus, the map and the composition are -amenable, but is not.
2.4 Acyclic Maps
In this subsection, we show that the amenability of a loop-space is equivalent to the “triviality” of limits and colimits over its classifying space. This characterization is interesting in that it does not directly involve the higher semiadditive structure.
Definitions & basic properties
We begin by introducing the notions of “acyclicity” and “triviality”, which are not immediately related to the theory of ambidexterity:
Definition 2.4.1.
Let be a map of spaces and let be an -category. We say that is -acyclic (resp. -trivial) if admits all -limits and -colimits and is fully faithful (resp. an equivalence).
It is a standard fact about adjoints, that is fully faithful, if and only if the counit is an isomorphism and if and only if the unit is an isomorphism. Similarly, is an equivalence, if furthermore is an isomorphism, or equivalently is an isomorphism. Like ambidexterity and amenability, acyclicity and triviality are fiber-wise conditions:
Proposition 2.4.2.
Let . A map of spaces is -acyclic (resp. -trivial) if and only if each fiber of is -acyclic (resp. -trivial).
Proof.
For each map we can form the following pullback square of spaces and the induced commutative square of -categories
By [CSY18, Lemma 2.2.3], we have
Thus, if is -acyclic (resp. -trivial), then is -acyclic (resp. -trivial). If (and hence also ) is surjective, then by the conservativity of (and ), the converse holds as well. Namely, if is -acyclic (resp. -trivial), then is -acyclic (resp. -trivial). Applying this to any section of yields the claim. ∎
The collections of -acyclic and -trivial spaces are also closed under extensions:
Corollary 2.4.3.
Let and let . If is -acyclic (resp. -trivial) and all the fibers of are -acyclic (resp. -trivial), then is -acyclic (resp. -trivial).
Proof.
By 2.4.2, it suffices to show that -acyclic (resp. -trivial) maps are closed under composition, which is clear from the definition. ∎
In presence of a compatible monoidal structure, one can check the acyclicity property on the unit:
Lemma 2.4.4.
Let be a space and let which is compatible with -colimits. The space is -acyclic if and only if the fold map is an isomorphism.
Proof.
Assume is an isomorphism. By assumption, for every , tensoring the isomorphism with gives the fold map . Hence, is an isomorphism for all . ∎
The following is the prototypical example of acyclicity:
Example 2.4.5 (Bousfield).
For , we can consider the -category of -local spectra. By 2.4.4, a space is -acyclic if and only if it is -acyclic in the sense of Bousfield, i.e. has the -homology of a point.
Remark 2.4.6.
For an -category , a space is -acyclic if the following equivalent conditions hold:
-
(1)
The fold map is an isomorphism for all .
-
(2)
The diagonal map is an isomorphism for all .
We warn the reader that for an individual object , it can happen that the map is an isomorphism, but is not (and vise versa). As a trivial example, consider with and any .
Relation to amenability
Under suitable ambidexterity assumptions, the notion of -acyclicity turns out to be closely related to that of -amenability. To begin with, recall that for a -ambidextrous space and an object , the map is given by the composition
Thus, if is -acyclic, then it is in particular -amenable. However, there is a deeper connection between acyclicity and amenability, which we first state on an object-wise level:
Proposition 2.4.7.
Let and let be a connected -ambidextrous space. For every , the following are equivalent:
-
(1)
The fold map is an isomorphism.
-
(2)
The diagonal is an isomorphism.
-
(3)
is invertible.
Moreover, is then the inverse of .
Proof.
Let be a base point. First, we show that if the diagonal map is an isomorphism, then . We begin by reducing the claim to the fact that the map
equals the constant map on . Indeed, given that, by integrating along , we get ([CSY18, Propositions 2.1.15 and 3.1.13])
Now, recall that the diagonal is the unit of the adjunction Thus, if is an isomorphism at , then the map
is an isomorphism. Since , it follows by 2-out-of-3 that the map
is an isomorphism as well. Thus, it suffices to check that the maps and coincide after applying . The pullback square of spaces
gives by [CSY18, Proposition 3.1.13]
This concludes the proof of (2)(3) and that is the inverse of . A completely symmetric argument using the adjunction instead of shows that (1)(3) and that is the inverse of . The implication (3)(1) in the case that is invertible for all is given by [CSY18, Proposition 3.1.18]. One easily checks that all the arguments are, in fact, object-wise. Alternatively, one can run the argument on the full subcategory of objects on which is invertible. In particular, each of the conditions (1), (2), and (3) implies that is the inverse of and so, in particular, is invertible. Finally, the composition
is . Thus, by 2-out-of-3, we also have the implication (1)(2). ∎
From this we get:
Corollary 2.4.8.
Let and let be a connected space. The following conditions are equivalent:
-
(1)
is -acyclic and -ambidextrous.
-
(2)
is -amenable and admits -colimits.
In which case .
Proof.
We note that the assumption of -ambidexterity in (1) of 2.4.8 can not be relaxed to weak -ambidexterity. The following is a simple counter-example:
Example 2.4.9.
Let be the -category of -vector spaces and . It is clear that is weakly -ambidextrous (as is semiadditive) and that is -acyclic. However, is not -ambidextrous (and is not -amenable).
Remark 2.4.10.
The notion of acyclicity can be considered for a general (not necessarily -finite) space in any -category, without any assumptions on ambidexterity. However, in the presence of ambidexterity, 2.4.8 allows us to deduce the acyclicity of a given -finite space from the amenability of its loop space. This strategy was already employed in the proof of [CSY18, Theorem E] (exploiting the -semiadditivity of ).
We conclude this subsection by showing an analogue of the equivalence of (1) and (3) in 2.4.7, for the notions of -acyclicity and -triviality. As before, it is clear that a -trivial space is, in particular, -acyclic, but there is a better statement:
Proposition 2.4.11.
Let be a connected space and let be an -category which admits -colimits. Then, is -trivial if and only if is -acyclic.
Proof.
Let be a base point. The composition
is the identity. This implies that is an equivalence if and only if is. Moreover, it also implies that is essentially surjective and hence an equivalence if and only if it is fully faithful. Namely, if and only if is -acyclic. Since -acyclicity is a fiber-wise condition (2.4.2), is -acyclic if and only if is -acyclic. ∎
3 Height
In this section, we introduce the notion of “semiadditive height” for objects in higher semiadditive -categories, which is the central object of study in this paper. We establish here the most general properties of this notion, while those related to stability will be differed to the next section.
3.1 Semiadditive Height
The definition of height for a higher semiadditive -category depends on a choice of a prime . In fact, it suffices to have a certain “-typical” version of -semiadditivity, in which one requires ambidexterity only for -finite -spaces. To emphasize the relevant structure, and for some future applications, we shall develop the basic theory of height in this level of generality. However, for an -category which is -semiadditive and -local, the “-typical” version of higher semiadditivity will turn out to be equivalent to ordinary higher semiadditivity (3.2.6). Thus, for the applications considered in this paper, this point is of minor importance.
-Typical semiadditivity
We begin with a definition of the following “-typical” version of higher semiadditivity:
Definition 3.1.1.
Let be a prime and . We say that
-
(1)
An -category is -typically -semiadditive if all -finite -spaces are -ambidextrous.
-
(2)
A functor between such is -typically -semiadditive if it preserves all -finite -space colimits.
-
(3)
An -monoidal -category , for some -operad , is -typically -semiadditively -monoidal if it is -typically -semiadditive and is compatible with -finite -space colimits.
We denote by the subcategory of -typically -semiadditive -categories and -typically -semiadditive functors.
It is clear that an -semiadditive -category or functor are also -typically -semiadditive for every prime . It is useful to know that to verify -typical -semiadditivity, it suffices to consider only the “building blocks”:
Proposition 3.1.2.
Let .
-
(1)
An -category is -typically -semiadditive if and only if is -ambidextrous for all .
-
(2)
For , a -semiadditive functor is -typically -semiadditive if and only if it preserves -(co)limits for all .
Proof.
(1) The “only if” part is clear. Conversely, we need to show that if is -ambidextrous for all , then every -finite -space is -ambidextrous. Since is -semiadditive, we are reduced to the case that is connected. The Postnikov tower of can be refined to a tower of principal fibrations
such that the fiber of each is of the form for some . To show that is -ambidextrous, it suffices to show that each is -ambidextrous (as -ambidextrous maps are closed under composition). Finally, since -ambidexterity is a fiber-wise condition, this follows from the fact that is -ambidextrous.
(2) Follows by an analogous argument to (1). ∎
In a -typically -semiadditive -category , one can discuss cardinalities of -finite -spaces. As one might expect, the Eilenberg-MacLane spaces play a fundamental role, and so deserve a special notation:
Definition 3.1.3.
Let . For every integer , we define as a natural endomorphism of the identity functor of . We shall omit the superscript “”, whenever is clear from the context.
A fundamental example to keep in mind is the following:
Example 3.1.4.
Semiadditive height
In what follows it will be convenient to use the following terminology:
Definition 3.1.5.
Let and let be a natural endomorphism. An object is called
-
(1)
-divisible if is invertible.
-
(2)
-complete if for all -divisible .
We denote by and the full subcategories of spanned by the -divisible and -complete objects respectively.
Using the operations we can now define the semiadditive height:
Definition 3.1.6.
Let and let . For every we define and denote the (semiadditive) height of as follows:
-
(1)
, if is -divisible.
-
(2)
, if is -complete.
-
(3)
, if and .
We also extend the definition to as follows. For every , we write if and only if for all . Additionally, by convention for all , and or if and only if . We shall drop the subscript in , when the -category is clear from the context.
Remark 3.1.7.
We emphasize that the notation (and similarly etc.) asserts that satisfies a certain property, and does not mean that is a well-defined number, which can be compared with . We note that the only object in , which can simultaneously have height and is the zero object.
The motivating example for the definition of height is the following:
Example 3.1.8.
Let be a -semiadditive -category. An object is of height if and only if acts invertibly on , and of height if it is -complete.
The first thing to show is that the notion of height behaves as the terminology suggests:
Proposition 3.1.9.
Let and let be some integers. For every ,
-
(1)
If , then .
-
(2)
If , then .
Proof.
To prove (1), it suffices to show that if for some , then . We consider the principal fiber sequence
All maps and spaces in this sequence are -ambidextrous by assumption. Since , we have that is invertible. By 2.3.3, we get
Thus, is invertible as well, and hence . Claim (2) now follows from (1) by definition. ∎
Remark 3.1.10.
In any stable -category , a non-zero object can have height for at most one prime . In particular, if is -local, every object has height for every prime (in fact, this is ‘if an only if’). Nevertheless, in this case the notion of -height with respect to different primes allows one to treat “prime to phenomena” as “height phenomena” for primes .
It is also useful to consider the corresponding subcategories of objects having height in a certain range:
Definition 3.1.11.
Let and let . We define
the full subcategories of spanned by objects of height , and respectively. We also write , or , if , or respectively.
The above defined subcategories are themselves -typically -semiadditive:
Proposition 3.1.12.
Let . For every , the full subcategories and are closed under limits in . In particular, they are -typically -semiadditive, and are furthermore -semiadditive if is.
Proof.
An object belongs to if and only if is -divisible. Thus is closed under all limits which exist in . By definition, and are closed under limits in as well. Thus, for we have that is also closed under limits in . Finally, by 2.1.4(3), it follows that all these subcategories are -typically -semiadditive and are furthermore -semiadditive if is. ∎
Next, we consider the behavior of height with respect to higher semiadditive functors. It turns out that the height can only go down:
Proposition 3.1.13.
Let be a map in . For all and , if then . If is conservative, then the converse holds as well.
Proof.
This follows immediately from the fact that maps to . ∎
In contrast, the following example shows that a higher semiadditive functor need not preserve lower bounds on height:
Example 3.1.14.
The -semiadditive functor maps the -complete sphere , which is of height , to a non-zero object of height .
For an inclusion of a full subcategory, we can say a bit more:
Proposition 3.1.15.
Let and let be a full subcategory closed under -finite -space (co)limits. Given and , we have
-
(1)
if and only if .
-
(2)
implies .
Proof.
(1) follows from 3.1.13 applied to the inclusion . For (2), if , then for every , we have . We now observe that by (1), we have . Thus, for every , we have , which by definition means . ∎
In presence of a -typically -semiadditively monoidal structure, an upper bound on the height of the unit implies an upper bound on the height of the -category:
Corollary 3.1.16.
Let be -typically -semiadditively monoidal -category. For every , we have if and only if .
Proof.
Given , the functor is -typically -semiadditive. Thus, the claim follows from 3.1.13. ∎
3.2 Bounded Height
In this subsection, we study the implications for a higher semiadditive -category of having bounded height. These results generalize the previously discussed facts regarding the -semiadditive structure of semirational -categories (i.e. of height ), and fall under the slogan that “the higher semiadditive structure is trivial above the height”. As a by-product, we shall see that for a -semiadditive -local -category, there is no difference between -semiadditivity and -typical -semiadditivity.
Amenability & acyclicity
The results on amenability and acyclicity from 2.4 imply the following equivalent characterizations of height :
Proposition 3.2.1.
Let , which admits -(co)limits. The following properties are equivalent:
-
(1)
(i.e. is -amenable).
-
(2)
is -acyclic.
-
(3)
is -trivial.
We can therefore characterize the height of an -category in ways that do not make an explicit reference to the higher semiadditive structure.
Proof.
The following can be seen as a (-typical) generalization of the fact that a semirational -category is automatically -semiadditive:
Proposition 3.2.2.
Let such that , and assume admits -(co)limits. Then, is -typically -semiadditive.
Proof.
By 3.1.2(1), it suffice to prove that
-
is -amenable (and ,in particular, -ambidextrous) and admits all -colimits.
holds for all . We shall prove this by induction on . The base case is given by assumption. Assume holds for some and consider the fiber sequence
By the inductive hypothesis, is invertible and admits -(co)limits. Therefore by [CSY18, Propostion 3.1.17], the space is -ambidextrous. Moreover, by 2.3.3, is invertible as well. Finally, by 3.2.1(3), the diagonal functor is an equivalence and hence in particular admits -(co)limits. ∎
We now show that claims (1)-(3) of 3.2.1 extend to much wider classes of spaces:
Proposition 3.2.3.
Let such that . For every -finite -space , the following hold:
-
(1)
If is -connected, then is -amenable.
-
(2)
If is -connected, then is -acyclic.
-
(3)
If is -connected, then is -trivial.
Proof.
For (1), let be an -connected -finite -space. The Postnikov tower of can be refined to a tower of principal fibrations
such that the fiber of each is of the form for some . Since we assumed , we also have (3.1.9), and hence all the spaces are -amenable. By 2.3.5, the class of -amenable spaces is closed under principal extensions, and therefore is -amenable. Now, (2) follows from 2.4.8 and (1) applied to . Similarly, (3) follows from 2.4.11 and (2) applied to . ∎
Example 3.2.4.
Let be semirational, and so in particular of height (such as or ). By 2.2.2, for every -finite -space , the cardinality is a sum of positive rational numbers. Thus, is invertible if and only if is non-empty (i.e. -connected). The map is an equivalence for all , if and only if is connected, and is an equivalence if and only if is simply-connected.
Cardinality
Recall from 2.2.2, that the formula for the homotopy cardinality (1) could be deduced solely from the “multiplicativity property” with respect to fiber sequences. For higher heights, we have the following analogue:
Proposition 3.2.5.
Let such that . Given a principal fibration of -finite -spaces
if is -connected, then . In particular,
Proof.
The -local Case
In many situations of interest, the -category under consideration is -semiadditive, -local and admits all -finite colimits. In this case, is automatically -typically -semiadditive for all primes (3.2.2). Moreover, in this case there is essentially no difference between higher semiadditivity and -typical higher semiadditivity:
Proposition 3.2.6.
Let be a -semiadditive -local -category which admits all -finite limits and colimits. The -category is -typically -semiadditive if and only if it is -semiadditive.
Proof.
Assume that is -typically -semiadditive. By [HL13, Proposition 4.4.16] and the fact that the space is -ambidextrous, we get that is -semiadditive. To get higher semiadditivity, we first have by assumption that is -typically -semiadditive. Moreover, since is -local, it is -typically -semiadditive for all (3.2.2). Thus, is -ambidextrous for all primes and all integers . By inductive application of [HL13, Proposition 4.4.19], it follows that is -semiadditive. We note that in both [HL13, Proposition 4.4.16] and [HL13, Proposition 4.4.19], one assumes that admits all small limits and colimits. However, the proofs use only the limits and colimits which we assumed in the statement. ∎
When applied to -local -categories, the main results of this subsection can be summarized as follows:
Theorem 3.2.7.
Let be a -semiadditive -local -category, which admits all -finite limits and colimits. If is -typically -semiadditive such that , then is -semiadditive. Moreover, for every -finite space :
-
(1)
If is -connected and nilpotent, then is -amenable.
-
(2)
If is -connected, then is -acyclic.
-
(3)
If is -connected, then is -trivial.
Proof.
Since and admits -limits and colimits, it follows that is -typically -semiadditive (3.2.2). Since is -local, it follows that is in fact -semiadditive (3.2.6). For (1), we observe that if is -finite and nilpotent, then where ranges over primes and is a -finite -space which is contractible for almost all (e.g. [PS17, Theorem 5.7]). Since we have (2.1.10), the -amenability of follows from the -amenability of all the -s (see 3.2.3(1)). Finally, as in the proof of 3.2.3, (2) follows from (1) and 2.4.8 applied to and (3) follows from (2) and 2.4.11. ∎
Remark 3.2.8.
We note that the nilpotence condition in 3.2.7(1) is vacuous for simply connected spaces and hence relevant only for height . However, in this case it can not be dropped. Indeed, by 2.3.7, at the prime we have which is not-invertible. With a little more effort one can show that if , then is invertible for every connected -finite space , such that that the -Sylow subgroup of is normal. In other words, the -primary fusion in the fundamental group is the only obstruction for the invertibility of .
3.3 Semiadditive Redshift
By 2.2.8, the -category is -semiadditive and hence given , we can discuss for various primes , which is the height of as an object of . However, if itself is -typically higher semiadditive, then we have also defined , as the height of the objects of . These two notions of height do not coincide. Rather, we shall now show that exceeds exactly by one. As the semiadditive height generalizes the chromatic height, this can be viewed as a particular manifestation of the “redshift” principle. Roughly speaking, categorification tends to shift the height up by one. Before we begin, we need a general categorical lemma:
Lemma 3.3.1.
Let be an adjunction. If is an equivalence, then is fully faithful.
Proof.
The functor is fully faithful if and only if the unit is an isomorphism. The unit is part of a monad structure on making it a monoid in the homotopy category of , whose monoidal structure is given by composition. It therefore suffices to show that given a monoid in any monoidal (ordinary) category, if is invertible with respect to the monoidal structure, then the unit map is an isomorphism. We observe that the multiplication map is always a left inverse of . Thus, since is an equivalence, it follows that admits a left inverse as well. To conclude the proof, it remains to show that admits a right inverse. Since is invertible, it is in particular dualizable, and the dual of is its inverse. More precisely, the duality datum maps and are isomorphisms and exhibit as the inverse of . Consider the following commutative diagram
The clockwise composition is the identity and hence so is the counter-clockwise composition. It follows that the map
admits a right inverse. Since the functor is an equivalence, it follows that admits a right inverse. ∎
We can now state and prove the following “Semiadditive Redshift” result, which can be informally summarized as “”.
Theorem 3.3.2 (Semiadditive Redshift).
Let be integers and let . If is -typically -semiadditive, then
-
(1)
if and only if .
-
(2)
if and only if .
In particular, an -semiadditive -category is of height if and only if, as an object of (and hence also ), it is of height .
Proof.
Denote . For (1), we observe that if and only if is -acyclic (3.2.1), namely that is fully faithful. On the other hand, the natural endomorphism of the identity functor of , acts on by the functor (2.2.8). Thus, , if and only if is invertible. Now, if is fully faithful, then and is in particular invertible. On the other hand, if is invertible, then is fully faithful by the dual of 3.3.1. Thus, if and only if .
For (2), we first assume that , and show that . Given such that , we need to show that every -finite colimit preserving functor must be zero. For every , we have . Since is -finite colimit preserving, we get also . Thus, by 2.4.7, we get . Since , the only object of height in is zero and hence . Thus, is zero, which proves that . Conversely, assume , to show that , consider the full subcategory spanned by objects of height in , which is also -typically -semiadditive (3.1.12). By definition, , and hence by (1), , so the inclusion functor must be zero. It follows that is zero and therefore .
Recall from 2.2.10, that for every , the space of objects is endowed with an -commutative coCartesian monoid structure making it an object of the -semiadditive -category . 3.3.2 has the following corollary:
Corollary 3.3.3.
Let , such that . The space of objects with the higher coCartesian structure satisfies , as an object of .
Proof.
Example 3.3.4.
Let be a -local ring spectrum. By 4.4.3 and 4.4.5, the -commutative monoid is of height in . In particular, this applies to Morava -theory . This suggests a relation between the “semiadditive redshift” of 3.3.2 and the “chromatic redshift” in algebraic -theory of Ausani-Rognes (see [AR08, AKQ19]). We shall explore this connection further in a future work.
The proof of 3.3.2 relies ultimately on the fact that is -amenable if and only if is -acyclic (2.4.8). In 2.4.11 we categorified this fact by showing that is -acyclic, if and only if is -trivial. Similarly, 3.3.2 can be categorified as follows. Let be the full subcategory spanned by the -semiadditive -categories such that .
Lemma 3.3.5.
The full subcategory is closed under -finite colimits.
Proof.
We have
Hence, in particular, admits -finite colimits and is therefore an object of the -semiadditive -category of large -categories, which admit -finite colimits and functors preserving them.
Proposition 3.3.6.
The -category is an object of height in .
4 Stability
So far, we have been considering general higher semiadditive -categories. In this section, we specialize to the stable world. First, using the general results on height from the previous section, we shall show that every stable higher semiadditive -category decomposes completely according to height. Second, inspired by [Lurb], we study semisimplicity properties of local systems valued in general stable -categories of semiadditive height . Finally, we show that and are indeed of semiadditive height .
4.1 Recollement
A central tool, which will be used several times in the following subsections, is that of a recollement of stable -categories following [BG16]. In this preliminary subsection, we collect several general facts regarding this notion. First, we provide criteria for a recollement to be split, in the sense that it has trivial “gluing data”. Second, we show that a decreasing intersection of a chain of recollements is again a recollement. Finally, we give special attention to recollements arising from “divisible” and “complete” objects with respect to a natural endomorphism of the identity functor in the sense of 3.1.5.
(Split) Recollement
We begin by recalling the notion of recollement in the context of stable -categories following the exposition in [Lura, Section A.8.1].
Definition 4.1.1.
Let be a stable -category and a full stable subcategory. We define the right orthogonal complement to be the full subcategory consisting of objects such that for all .
Recall from [Lura, Proposition A.8.20], that if the inclusion admits both a left adjoint and a right adjoint , then is a recollement of and in the sense of [Lura, Section A.8.1]. In particular, the -category can be identified with the -category of sections of the cartesian fibration over , classified by the functor
Definition 4.1.2.
We shall say that an inclusion of stable -categories , that admits both a left and a right adjoint, exhibits as a recollement of and .
Recollements in stable homotopy theory typically arise from smashing localizations, and amount to the existence of various fracture squares:
Example 4.1.3 (Arithmetic and chromatic squares).
For , the inclusion of the full subcategory admits both a left and a right adjoint, and we have
is the full subcategory spanned by -complete spectra. The recollement statement in this case recovers the classical -local arithmetic square for spectra. More generally, the full subcategory
exhibits as a recollement of and , where is a finite spectrum of type . In particular, the inclusion exhibits as a recollement of and . This recovers the classical telescopic fracture square at height . By the Smash Product Theorem,
is also a smashing localization, and similarly, the inclusion exhibits as a recollement of and . This recovers the classical chromatic fracture square at height .
Given a recollement , the functor encodes the “gluing data” in the construction of from the -categories and . A particularly simple instance of a recollement is when this gluing data is trivial:
Proposition 4.1.4.
Given a recollement , the following are equivalent:
-
(1)
The functor is zero.
-
(2)
The left adjoints of the inclusions induce an equivalence .
-
(3)
The left and right adjoints and of are isomorphic.
Proof.
We prove (1)(2)(3)(1). The implication (1)(2) follows from the identification of with the -category of sections of the cartesian fibration classified by . For the implication (2)(3), observe that the inclusion can be identified with the functor , for which the projection is both a left and a right adjoint. Finally, for every , we have . Thus, assuming (3) we have and so , which proves (1). ∎
Definition 4.1.5.
We say that a recollement is split if it satisfies the equivalent conditions of 4.1.4.
Recollement chains
We shall now study the behavior of chains of recollements.
Definition 4.1.6.
For , we say that a descending chain of full subcategories,
is a recollement chain, if each inclusion exhibits as a recollement of and . We also set
It turns out that under mild conditions, is itself a recollement.
Lemma 4.1.7.
Let , which admits sequential limits and colimits. For a recollement chain , the inclusion exhibits as a recollement of and .
Proof.
It suffices to show that the inclusion admits a left and a right adjoint. By symmetry, it suffices to consider only the left adjoint. By [Lur09, Proposition 5.2.7.8], for every we need to construct an object and a morphism , such that for all , the map
is an isomorphism. Let be the left adjoint of the inclusion and denote by the corresponding unit (i.e. localization) map, where we suppress the embedding functor . Since the -s are nested, we have canonical isomorphisms , and we abuse notation by denoting the composition also by . We now define
and take to be the cone map from the first object to the colimit in the diagram defining . For every , the map
is an isomorphism for each . Thus, by taking the limit over , we get an isomorphism
Precomposing with the isomorphism
we get an isomorphism
Unwinding the definitions, this isomorphism is given by precomposition with . ∎
To identify the right orthogonal complement of in , we need the following general categorical fact:
Lemma 4.1.8.
Let be a functor in , and denote by the full subcategory spanned by the objects , for which . If admits a fully faithful right adjoint , then .
Proof.
In one direction, for , we have for some . Hence, for every we have
Thus, . Conversely, let and be the unit and counit of the adjunction respectively. Since is fully faithful, is an isomorphism. By the zig-zag identities and 2-out-of-3, the map is also an isomorphism. Now, for every consider the fiber sequence
On the one hand, since is an isomorphism, and hence . On the other hand, if , then since , we also have and thus . This implies that is an isomorphism and so . ∎
Given a recollement chain as in 4.1.6, for each , we have a fully faithful embedding with left adjoint . We abuse notation by suppressing the inclusion and the canonical isomorphisms . We thus obtain a tower
of -categories under , which induces a functor
Proposition 4.1.9.
Let which admits sequential limits and colimits. Given a recollement chain , the functor admits a fully faithful right adjoint, whose essential image is . Thus, is a recollement of and .
Proof.
By 4.1.7, the inclusion exhibits as a recollement of and its right orthogonal complement and hence it suffices to identify . The objects of are precisely the for which . Thus, by 4.1.8, it suffices to show that admits a fully faithful right adjoint. Since is a left adjoint for all , by [HY17, Theorem B], the functor is a left adjoint and we denote its right adjoint by . We show that is fully faithful using the explicit description of the adjunction given in [HY17]. An object of consists of a sequence of objects together with structure isomorphisms . We shall write suppressing the structure isomorphisms. Composing the structure isomorphisms of with the corresponding unit (i.e. localization) maps , we get maps as follows:
By [HY17, Theorem B], the functor can be described explicitly on objects by the following formula:
To prove that is fully faithful, it suffices to show that the counit is an isomorphism. Since the collection of projection functors for all is jointly conservative, it suffices to show that
is an isomorphism for all and . By [HY17, Theorem 5.5], we can describe as the composition
where the first map is induced by the canonical projection . By cofinality, we can assume that the limit is taken over . By definition, for each , the fiber of lies in . Since is closed under sequential limits, it follows that the fiber of lies in and hence becomes an isomorphism after applying . This concludes the proof that is fully faithful and hence the proof of the claim. ∎
Corollary 4.1.10.
Let which admits sequential limits and colimits with a recollement chain . If then .
Proof.
By 4.1.9, is a recollement of and of , so that . ∎
Divisible and complete recollement
One way to get a recollement is by taking the divisible and complete objects with respect to a natural endomorphism of the identity functor. That is, given a stable -category and , we have the full subcategories and of (3.1.5). Assuming admits sequential limits and colimits, the inclusion admits both a left adjoint and a right adjoint , given respectively by “inverting ” on
and by taking the “-divisible part” of
Remark 4.1.11.
We warn the reader that although the above statements are well known and fairly intuitive, they are not as tautological as one might think. In particular, they might fail if is not assumed to be stable (or at least additive). We refer the reader to [BNT18, Appendix C], for a comprehensive treatment of a closely related situation.
Note that an object is -complete if and only if if and only if . In fact, the -completion functor
is the left adjoint to the inclusion .
Proposition 4.1.12.
Let which admits sequential limits and colimit and let . Then is a recollement of and .
Proof.
It follows from the discussion above that admits both adjoints. ∎
Our next goal is to give a characterization of when the said recollement is split in terms of the natural endomorphism .
Definition 4.1.13.
We say that a natural endomorphism is a semi-inverse of , if for every -divisible , the map is an inverse of .
The usefulness of the notion of semi-inverse is in that it allows us to characterize completeness in terms of divisibility:
Proposition 4.1.14.
Let which admits sequential limits and colimits. For every , if an object is -complete, then it is -divisible. If is a semi-inverse of , then the converse holds as well.
Proof.
Note that all natural endomorphisms of commute by the interchange law, so in particular . For an -complete object we have . For every , the map is zero on and hence is invertible on . By passing to the limit, is invertible on . Conversely, assume that acts invertibly on . If is a semi-inverse of , then for every -divisible , the map acts as zero on . Thus the pointed space must be contractible as acts both invertibly and as zero on it. This implies that is -complete. ∎
The above lemma leads us to the following characterization of split recollement:
Proposition 4.1.15.
Let be a stable -category which admits sequential limits and colimits. The recollement associated with a natural endomorphism is split if and only if admits a semi-inverse. In which case,
Proof.
Let be a semi-inverse of . To show that for all it suffices to show that is contractible for all . By definition, is zero on , so it suffices to observe that is invertible on by 4.1.14. Conversely, if , we have for every a natural decomposition with and . In this case, the map is a semi-inverse of . ∎
Remark 4.1.16.
In 4.1.3, the recollement corresponds to the endomorphism . However, not every recollement arises in such a way. For example, for the recollement is not induced by any endomorphism of the identity functor. We do note however, that every split recollement must arise from an endomorphism of , because we can take to be the idempotent projecting onto . In this case, itself is a semi-inverse of .
4.2 Height Decomposition
Let be now a stable -semiadditive -category. We shall use the general machinery of (split) recollement to show that splits into a product of -categories according to height. By definition, an object is of height if it is -divisible. Similarly, is of height if it is -complete, which by 3.1.9, is if and only if it is complete with respect to all of . Accordingly,
Proposition 4.2.1.
Let and let . If admits sequential limits and colimits, then is a recollement of and .
Proof.
The full subcategory consists of the -divisible objects and is the full subcategory of -complete objects. Thus, the result follows from 4.1.12. ∎
Our next goal is to show that under suitable assumptions, this recollement is in fact split. For this we need the following:
Proposition 4.2.2.
Let and assume it admits sequential limits and colimits. For all , the map is a semi-inverse of . In particular, for , we have if and only if is -divisible.
Proof.
Corollary 4.2.3.
Let be a functor in and assume and are stable. For every and , if is of height or , then so is . The converse holds if is conservative.
Proof.
Remark 4.2.4.
Similarly, we can improve on 3.1.16 as follows:
Corollary 4.2.5.
Let be -typically -semiadditively monoidal -category. For every , we have or if and only if or respectively.
Proof.
Given , the functor is -typically -semiadditive. Thus, the claim follows from 4.2.3. ∎
Remark 4.2.6.
Again, in 4.2.5 it is not true that if then . For example, is a presentably symmetric monoidal -semiadditive -category, whose unit has height , although the -category itself does not.
The following is our main structure theorem for stable higher semiadditive -categories:
Theorem 4.2.7 (Height Decomposition).
Let for some .
-
(1)
For , and which is idempotent complete, the inclusions determine an equivalence of -categories
Moreover, is -typically -semiadditive and are -semiadditive.
-
(2)
For , and which admits sequential limits and colimits, is a recollement of and . In particular, if , then
If in addition is -semiadditively -monoidal for some -operad then and for all , are compatible with the -monoidal structure and the equivalences in both (1) and (2) promote naturally to an equivalence of -monoidal -categories.
Proof.
(1) We first prove the claim under the additional assumption that admits all sequential limits and colimits (and hence in particular idempotent complete). First, by 4.2.2, the map is a semi-inverse of . Hence, by 4.1.15, we obtain a direct product decomposition . The category is itself -semiadditive (3.1.12) and admits all sequential limits and colimits. Thus, we can continue decomposing inductively and get for all . Finally, by 3.2.2, each is in fact -typically -semiadditive. By 3.1.12, is also -typically -semiadditive and since it is also -complete for all , it is in particular -local, and hence -semiadditive by 3.2.6.
For a general as in the claim, we use a semiadditive version of the Yoneda embedding to reduce to the presentable case. Namely, we shall show in 5.3.5, that there exists a presentable stable -semiadditive -category and an -semiadditive fully faithful embedding . By 4.2.3, for each , we have fully faithful embedding and , hence also
By the left cancellation property of fully faithful embeddings, we get a fully faithful embedding
For each object , the height and components of in , are retracts of in . Thus, if is idempotent complete, these components belong to . It follows that the above fully faithful embedding is also essentially surjective.
(2) For every , we have by (1), that . Hence, we can switch the roles of and , and consider the embedding as exhibiting as a recollement of and . We thus obtain a recollement chain
By definition, , and so by 4.1.9, is a recollement of and
Finally, assume that is -semiadditively -monoidal. The full subcategories and for all consist of objects which are -divisible and -divisible respectively. It follows that and are compatible with the -monoidal structure and hence so is for all . Thus, by [Lura, Proposition 2.2.1.9], the -s and inherit an -monoidal structure such that the projections and are -monoidal. ∎
We conclude with some remarks regarding the sharpness of 4.2.7. In the case , the fact that is -semiadditive also implies that is a recollement of and , but there is no guaranty that the “gluing data” is trivial. That is, that decomposes as a direct product of and . Indeed, consider the -semiadditive -category The case corresponds to the recollement of 4.1.3. In this case and the gluing data is not trivial, as the rationalization of the -completion does not vanish in general. Having said that, for we do not know whether there even exists a stable -local presentable -category that is -semiadditive, but not -semiadditive [CSY18, Conjecture 1.1.5].
In the case , we do not know whether there exists a stable -semiadditive -category for which . We hence propose the following:
Conjecture 4.2.8 (Height Finiteness).
For every , the full subcategory
of objects of height , is trivial.
4.3 Semisimplicity
Classical representation theory tells us that in characteristic , representations of a finite group are semisimple. The -category of -representations of in any -category is equivalent to the -category of -valued local systems on . From the point of view of higher semiadditivity, characteristic corresponds to semiadditive height , and so it is natural to consider the analogous situation for higher heights. We shall show that given a stable -semiadditive -category of height , certain analogous semisimplicity phenomena hold for for every -finite -connected space (e.g. ). For the case these ideas were discussed in [Lurb].
Splitting local systems
The main result of this subsection is the following relation between acyclic maps and split recollement in the stable setting:
Proposition 4.3.1.
Let and let be a -acyclic and weakly -ambidextrous map. The functor is fully faithful and exhibits as a recollement of and . The recollement is split if and only if is -ambidextrous, in which case there is a canonical equivalence:
Proof.
By definition of acyclicity, is fully faithful. Hence, the recollement is split if and only the left and right adjoints and respectively of are isomorphic (4.1.4). If is -ambidextrous then and are isomorphic and the recollement is split. Conversely, if we have , then preserves all -colimits and hence is -ambidextrous (2.1.3). ∎
As a special case we obtain:
Theorem 4.3.2.
Let be -local such that . For every map of spaces with -connected -finite fibers, we have a canonical equivalence
Proof.
In the case , we can interpret 4.3.2 from the perspective of “higher representation theory” (see [Lurb]). For every space , we call an object unipotent if it belongs to the full subcategory generated by colimits from the trivial representations (i.e. constant local systems). It can be shown that every object fits into an essentially unique cofiber sequence
with and . If is -connected, then 4.3.2 implies that is constant and canonically splits.
Transfer idempotents
Given a stable -category and an equivalence , every object has an essentially unique decomposition , such that and . This allows us to define a natural endomorphism by the formula
The natural endomorphism is idempotent and realizes internally to the projection onto the essential image of in . Our next goal is to provide an explicit description of the idempotent for the split recollement of 4.3.1. To help guide the intuition, we begin with a closely related elementary example:
Example 4.3.3.
Let and let be a finite group with a normal subgroup . The map induces a fully faithful embedding
which is the “inflation” functor that takes a vector space with a -action to itself with the -action induced via . The essential image of consists of -representations on which the subgroup acts trivially. The adjoints and can be identified with the vector spaces of -coinvairant and -invariants respectively, equipped with the residual -action. Thus, the full subcategory is spanned by the -representations without non-trivial -fixed vectors and we have an equivalence
This equivalence is realized explicitly as follows. Since is semi-simple, we can split each -representation uniquely as with the space of -fixed vectors and . Consider the natural endomorphism:
and its normalization . Unwinding the definitions, we get
which is an explicit formula for the -equivariant projection onto the subspace .
In a similar fashion, we have:
Proposition 4.3.4.
Let and let be a -acyclic and -ambidextrous map. Consider
The natural endomorphism of is idempotent, and it realizes the projection onto the essential image of .
Proof.
First, for every we have and hence . Next, consider the commutative diagram (see 2.1.5):
The composition along the top row computes the restriction of along , which is therefore invertible and coincides with . It follows that is the identity on the essential image of and is zero on . Thus, equals the idempotent which projects onto the essential image of . ∎
Remark 4.3.5.
Example 4.3.6.
Let be of height . For an abelian -group with a subgroup , we consider the fiber sequence
The map satisfies the conditions of 4.3.2, and hence splits as a product of and its right orthogonal complement. Moreover, by 4.3.4 and 4.3.5, the idempotent is given by . The case is a “derived version” of 4.3.3.
4.4 Chromatic Examples
The main source of stable higher semiadditive -categories is chromatic homotopy theory. In this subsection, we shall address the semiadditive height of such -categories, using the properties of nil-conservative functors and the Nilpotence Theorem (see [HS98]). We begin by recalling the following notion:
Definition 4.4.1 ([CSY18, Definition 4.4.1]).
We call a functor in nil-conservative, if for every ring , if then 111111This notion is closely related to the notion of “nil-faithfulness” defined in [Bal16]..
We also recall that nil-conservative functors are conservative on the full subcategory of dualizable objects. In particular, we have the following consequence regarding height:
Proposition 4.4.2.
Let be a map in . If is -semiadditive, then so is . Furthermore, for every , if is of height or then so is and the converse holds if is nil-conservative.
Proof.
Since is -semiadditive and is a monoidal -semiadditive functor, by [CSY18, Corollary 3.3.2(2)], is -semiadditive as well. By 4.2.5, the height of (resp. ) is determined by the height of (resp. ). Moreover, if and only if is invertible and , if and only if is invertible. Thus, the claim follows from [CSY18, Corollary 4.4.5]. ∎
As a special case, we get:
Corollary 4.4.3.
Let be -semiadditive and let . For every integer , if is of height or then so is . The converse holds if tensoring with is nil-conservative.
Proof.
The claim follows from 4.4.2 for the colimit preserving symmetric monoidal functor . ∎
We now apply the above to higher semiadditive -categories arising in chromatic homotopy theory. We begin with a special case in which cardinalities of Eilenberg-MacLane spaces can be computed explicitly.
Proposition 4.4.4.
The -category satisfies .
Proof.
For , the claim is clear, so assume . This follows from the explicit formula given in 2.2.5. Indeed, for , we have and hence invertible. For , the element is a non-zero power of , and hence every is complete with respect to it as the -category is -complete. ∎
In [CSY18, Theorem C] we have shown that for a homotopy ring spectrum , the -category is -semiadditive if and only if it is -semiadditive if and only if for some integer . We now show that this is in fact the semiadditive height of .
Theorem 4.4.5.
Let be a homotopy ring spectrum121212In fact, it suffices to assume that is a weak ring in the sense of [CSY18, Definition 5.1.4].. If for some integer , then . In particular, we have
Proof.
5 Modes
In this section, we use the theory of idempotent algebras in , which we call modes, to further study the interaction of stability and higher semiadditivity.
5.1 Idempotent Algebras
We begin with a general discussion about idempotent algebras in symmetric monoidal -categories, as a means to encode properties, which induce “canonical structure”.
Definitions & characterizations
Following [Lura, Definition 4.8.2.1], given a symmetric monoidal -category , we say that a morphism in exhibits as an idempotent object of , if
is an isomorphism. By [Lura, Proposition 4.8.2.9], an idempotent object admits a unique commutative algebra structure for which is the unit. Conversely, the unit of a commutative algebra , exhibits it as an idempotent object if and only if the multiplication map is an isomorphism. In this case we call an idempotent algebra. More precisely, the functor which forgets the algebra structure and remembers only the unit map, induces an equivalence of -categories from the full subcategory of idempotent algebras to the full subcategory of idempotent objects [Lura, Proposition 4.8.2.9]. The fundamental feature of an idempotent algebra , is that the forgetful functor is fully faithful, and its essential image consists of those objects for which the map is an isomorphism [Lura, Proposition 4.8.2.10]. Thus, it is a property of an object in to have the structure of an -module. We shall say that classifies the property of being an -module.
Example 5.1.1.
For , the idempotent algebras are classically known as solid rings [BK72, Definition 2.1]. These include for example and . We note that for , the ring is still idempotent, classifying the property of being rational, but is not idempotent. The idempotent rings in correspond precisely to the smashing localizations.
Given an idempotent ring the forgetful functor admits a left adjoint
This is a localization functor, which can be thought of as forcing the property classified by in a universal way. In line with the standard terminology for localizations of spectra, we set:
Definition 5.1.2.
Let be a map in . We say that is a smashing localization if there exists an idempotent algebra in and an isomorphism in , such that is the composition
We note that for a smashing localization , there is always a fully faithful right adjoint , which is lax symmetric monoidal [Lura, Corollary 7.3.2.7], and we can identify the idempotent algebra with . To characterize smashing localizations, we first introduce some terminology. Let in , which admits a (lax symmetric monoidal) right adjoint . For every and , we have a natural map
which is the mate of
where is the counit of the adjunction. We say that the adjunction satisfies the projection formula if the map is an isomorphism for all . The map is compatible with the unit of the adjunction in the following sense:
Lemma 5.1.3.
For all , the following diagram is commutative
Proof.
Passing to the mates under the adjunction and unwinding the definitions, this follows from the zigzag identities. ∎
We can now characterize the smashing localizations among all localizations as those that satisfy the projection formula (compare with [MNN17, Proposition 5.29]):
Proposition 5.1.4.
A map in is a smashing localization if and only if it admits a fully faithful right adjoint (i.e. it is a localization) and the adjunction satisfies the projection formula.
Proof.
In the “only if” direction, the right adjoint of is the forgetful functor , which is fully faithful since is idempotent. For and , the projection formula transformation is the composition
The map is an isomorphism since is fully faithful. The map is an isomorphism because admits a structure of an -module and hence in the essential image of .
For the “if” direction, consider the commutative ring
By the projection formula, we have a natural isomorphism
Since is fully faithful, the unit map exhibits as a localization (as in [Lur09, Proposition 5.2.7.4]). Moreover, by 5.1.3, the unit map is induced from tensoring with the unit of . Thus, by [Lura, Proposition 4.8.2.4], is an idempotent ring. Moreover, by [Lura, Proposition 4.8.2.10], the forgetful functor is a symmetric monoidal equivalence onto the essential image of with the localized symmetric monoidal structure, which is equivalent to . ∎
Remark 5.1.5.
Let be presentably symmetric monoidal and let be an idempotent algebra in . The fully faithful forgetful functor admits also a right adjoint. Hence, if is moreover stable, then exhibits as a recollement of and its right orthogonal complement.
Poset structure
Another nice characterization of idempotent algebras is as the -cotruncated objects of :
Proposition 5.1.6.
Let . A commutative algebra is idempotent, if and only if for all the space is either empty or contractible. Moreover, it is non-empty if and only if is an -module.
Proof.
By [Lura, Proposition 3.2.4.7], the tensor product of commutative algebras is the coproduct in . Moreover, the multiplication map
is the categorical fold map. Thus, it follows by the dual of [Lur09, Lemma .5.5.6.15], that is idempotent if and only if it is -cotruncated. We note that [Lur09, Lemma .5.5.6.15] requires the -category to admit finite limits; however, the only limit used in the proof is the one defining the diagonal map, which in our case corresponds to the coproduct defining the fold map. Now, if there exists a map in , then is an -algebra and in particular an -module. Conversely, for every , we have maps
in . If is an -module, then the map is an isomorphism, and thus
is a map in . ∎
We also have the following non-commutative analogue of 5.1.6:
Proposition 5.1.7.
Let and let . For every algebra , the space is either empty or contractible and it is non-empty if and only if is an -module.
Proof.
As in the proof of 5.1.6, is an -module if and only if there exists a map in . If is an -module, then and
since is the initial object of . ∎
As a consequence of 5.1.6, the -category of idempotent algebras is a poset. This poset admits binary meets:
Proposition 5.1.8.
Let and . Then is an idempotent algebra which classifies the conjunction of the properties classified by and .
Proof.
Let be the unit of and the unit of . We consider the composition
After tensoring with the first map becomes an isomorphism and after tensoring with the second map becomes an isomorphism. Thus, exhibits as an idempotent object. Given an -module , tensoring with is an isomorphism. Thus, and hence is both an -module and an -module. Conversely, tensoring with is given by the composition:
Hence, if is both an -module and an -module, then tensoring and with is an isomorphism, and so tensoring with is an isomorphism as well. Thus, is an -module. ∎
Idempotent objects are also closed under sifted colimits in .
Proposition 5.1.9.
Let , and let be a sifted -category such that is compatible with -indexed colimits. Then
-
(1)
The -category admits -indexed colimits.
-
(2)
The forgetful functor preserves -indexed colimits.
-
(3)
Given a functor , the idempotent algebra classifies the conjunction of the properties classified by for all .
Proof.
By [Lura, Corollary 3.2.3.2] the -category admits -indexed colimits and the forgetful functor preserves -indexed colimits. Thus, to prove (1) and (2) it suffices to show that is closed under -indexed colimits. Since is sifted, the diagonal map is cofinal. Therefore, given a functor , the multiplication map
can be identified with the colimit of the multiplication maps for . Hence, if is an idempotent algebra for every , so is . We shall now prove (3). First, for every there is a canonical algebra map . Thus, every ()-module admits an -module structure for every . It remains to prove that if admits an -module structure for every , then
is an isomorphism. Since is compatible with -indexed colimits, the map above is the colimit of the isomorphisms . ∎
Consequently, under mild conditions on , the poset admits arbitrary meets.
Corollary 5.1.10.
Let which is compatible with filtered colimits. Then the poset is cocomplete. Moreover, given a functor , the idempotent algebra
classifies the conjunction of the properties classified by for all .
Proof.
First, admits an initial object which is . Second, by 5.1.8, admits binary coproducts. Since every filtered -category is sifted by [Lur09, Example 5.5.8.3] we get by 5.1.9 that admits filtered colimits. Since is a poset, we deduce that it is cocomplete. Furthermore, 5.1.8 and 5.1.9 also imply that the colimit classifies the conjunction of the properties classified by the idempotent algebras in the diagram. ∎
Under some conditions, disjoint idempotent algebras have also binary joins.
Proposition 5.1.11.
Let which is compatible with all small colimits and is -semiadditive. Let be idempotent algebras in . If , then is an idempotent algebra, which classifies the property of an object to be of the form , where is an -module and is an -module.
Proof.
By assumption, the tensor product preserves binary coproducts in each variable. Since is -semiadditive, we get that the tensor product also preserves binary products in each variable. Thus,
coincides with , which is an isomorphism. Thus, is an isomorphism, and therefore, is an idempotent algebra. The projection maps and induce the extension of scalars functors
which in turn induce a functor
Since and are left adjoints, by [HY17, Theorem B] the functor admits a right adjoint , which is given object-wise by . To complete the proof of the claim, it would suffice to show that is an equivalence. We do this by showing that is conservative and is fully faithful. By 3.3.1, in order to show that is fully faithful, it suffices to show that is an equivalence. For every -module , we have
To show that is conservative, it suffices to observe that the underlying -object of is the direct sum and both and are retracts of . ∎
5.2 Theory of Modes
We now specialize the notion of idempotent algebras to the -category of presentable -categories and colimit preserving functors.
Tensor of presentable -categories
Recall from [Lura, Proposition 4.8.1.15], that the -category admits a closed symmetric monoidal structure. The unit is , and for every , the internal hom and tensor product are given respectively by
It is worth spelling out in what sense the above formula for the tensor product is functorial. Given a functor in with right adjoint , the induced functor is the left adjoint of
From this we get that tensoring with preserves reflective localizations:
Lemma 5.2.1.
Let and in . If admits a fully faithful (resp. conservative) right adjoint, then so does .
Proof.
By the above formula for , its right adjoint is given by
where is the right adjoint of . Thus, if is fully faithful (resp. conservative), then post composition with is fully faithful (resp. conservative) as well. ∎
Remark 5.2.2.
On the other hand, if is itself fully faithful or conservative, then need not be. For example, let be the full subcategory of connective spectra. One can show that , while (e.g. by 5.2.10). Thus, tensoring the fully faithful inclusion with the category , produces the zero functor .
Another general fact which we shall require is the preservation of recollements under base change. Recall that is the full subcategory of stable presentable -categories.
Proposition 5.2.3.
Let and assume it is a recollement of and . For every , the morphism exhibits as a recollement of and .
Proof.
Let be the inclusion functor. We denote by and the left and right adjoints of respectively. We observe that is presentable as an accessible localization of , and hence both functors and are morphisms in . The adjunction induces an adjunction
Since and are both right adjoints, this adjunction restricts to an adjunction on the full subcategories spanned by the right adjoints on both sides,
On the other hand, the left adjoint of is and the left adjoint of is . It follows that is the left adjoint of . To conclude, admits a right adjoint (as a morphism in and also a left adjoint, given by . Furthermore, by 5.2.1, is also fully faithful. The -categories and are stable by [Lura, Proposition 4.8.2.18] because and are. Hence, we deduce that exhibits as a recollement of and
It remains to identify with . Note that is the full subcategory of spanned by objects on which the right adjoint of is zero. This right adjoint is given by
On the other hand, the inclusion is right adjoint to its left adjoint , and hence the right adjoint of is given by
This is a fully faithful functor whose essential image consists precisely of objects on which is zero. Thus, we have canonically identified with . ∎
Definition & examples of modes
We are now ready to introduce the central notion of this section:
Definition 5.2.4.
A mode is an idempotent algebra in . We denote by
the full subcategory spanned by modes.
Applying the preceding results on idempotent algebras to , we get the following:
Proposition 5.2.5.
-
(1)
is a (large) poset.
-
(2)
is co-complete. Moreover, the colimit of a diagram of modes classifies the conjunction of the properties classified by the modes in the diagram.
-
(3)
is the initial mode and it classifies the empty property (which is always satisfied).
-
(4)
for all .
-
(5)
The forgetful functor preserves sifted colimits.
-
(6)
is the terminal mode and it classifies the property of being equivalent to .
-
(7)
For , if , then their join is given by , and it classifies the property of being a direct product of an -module and an -module.
Proof.
(1) follows from 5.1.6. (2) follows from 5.1.10 and the fact that is closed symmetric monoidal. (3) follows from the fact that is the unit of . (4) follows from 5.1.8. (5) follows from 5.1.9. (6) follows from the fact that is a zero object of . (7) follows from 5.1.11, since is -semiadditive ([HL13, Example 4.3.11]). ∎
In addition to the initial and terminal modes, we also have the following (far from exhaustive) list of modes:131313All these can be found in [Lura, Section 4.8.2] with the exception of (3), which can be deduced from 5.2.10.
Example 5.2.6.
-
(1)
is the boolean mode which classifies the property of being equivalent to a poset (i.e. the mode of propositional logic).
-
(2)
is the discrete mode, which classifies the property of being equivalent to an ordinary category (i.e. the mode of ordinary, as opposed to “higher”, mathematics).
-
(3)
is the discrete additive mode, which classifies the property of being equivalent to an ordinary additive category.
-
(4)
is the pointed mode, which classifies the property of having a zero object.
-
(5)
is the stable mode, which classifies the property of being stable.
Given a mode , the fully faithful forgetful functor admits a left adjoint (i.e. localization), which is given by
This procedure should be thought of as forcing to be in the mode in a universal way.
Example 5.2.7.
For the stable mode , the -category is the stabilization [Lura, Example 4.8.1.23]. Similarly, for the discrete mode , the -category is the -truncation [Lura, Remark 4.8.2.17].
The general results for idempotent algebras have the following implication:
Proposition 5.2.8.
Let be a mode and which is an -module. The fully faithful embedding induces an equivalence of -categories
In particular, every is an -module.
Proof.
Since is an -module, by 5.1.7, there is a map of algebras and the claim follows. ∎
The -categorical structure of allows further constructions of modes beyond those provided by 5.2.5. Primarily, modes can be localized.
Localization of modes
Given a mode , every -module is by definition left-tensored over , and hence in particular enriched over [Lura, Proposition 4.2.1.33]. For every we denote by the corresponding hom-object in . The -enrichment of an -module can be explicitly described via the -enriched Yoneda embedding:
Definition 5.2.9.
Let be a mode, let be a reflective full subcategory, and let be an -module. We say that an object is -local if for every , the enriched hom-object lies in Furthermore, we say that itself is -local, if every object of is -local.
Proposition 5.2.10.
Let be a mode and an accessible reflective full subcategory, which is compatible with the symmetric monoidal structure of . Let be the left adjoint of the inclusion . The composition
exhibits as a mode. Moreover, for every -module the -category can be canonically identified with the full subcategory of -local objects in . In particular, classifies the property of being an -local -module.
Proof.
By [Lura, Proposition 2.2.1.9], the -category admits a canonical symmetric monoidal structure, such that the left adjoint promotes to a symmetric monoidal functor. In particular, the unit of this symmetric monoidal structure is the composition
where is the unit of the symmetric monoidal structure of . We need to show that is an equivalence, or equivalently, that its right adjoint is an equivalence. Let be the tensor product functor. Since the composition
is an equivalence, so is the composition of the right adjoints. It follows that is essentially surjective. To complete the proof, we shall show that is also fully faithful. Write as the composition
Let be the right adjoint of . It follows that is the composition of the right adjoint of and . Since admits a fully faithful right adjoint, by 5.2.1, the functor has a fully faithful right adjoint as well. Thus, it suffices to show that is fully faithful. For this, consider the commutative diagram
Taking the right adjoints, we see that the composition of with the right adjoint of , which is fully faithful by 5.2.1, is fully faithful. It follows that must be fully faithful as well. This concludes the proof that exhibits as a mode.
Now, we want to analyze the property classified by . Since is symmetric monoidal, is a commutative algebra over . Thus, every -module is, in particular, an -module. Given an -module , it is an -module if and only if the composition
is an equivalence. The first functor is an equivalence since is an -module. Thus, by 2-out-of-3, the composition is an equivalence if and only if is an equivalence. The functor admits by 5.2.1 a fully faithful right adjoint. To describe its essential image, we consider the commutative diagram of Yoneda embeddings
We see that is identified with the full subcategory of -local objects in . ∎
Remark 5.2.11.
5.2.10 need not hold for a reflective subcategory , which is not assumed in advance to be compatible with the symmetric monoidal structure. Indeed, the inclusion of co-connective spectra is reflective with a left adjoint . However, we have
Indeed, is the -category of spectrum objects in , and thus is zero. But by 5.2.1, the functor
admits a fully faithful right adjoint, so must be zero as well.
Many further examples of modes can be constructed using 5.2.10:
Example 5.2.12.
Consider the subcategory of -truncated spaces. Every is an -module and an object is -local if and only if it is -truncated. Thus, is a mode which classifies the property that every object is -truncated. Namely, the property of being equivalent to a -category (compare with 5.2.11). The cases and , reproduce the terminal mode , the boolean mode , and the discrete mode respectively.
Another important family of examples is the Bousfield localizations:
Example 5.2.13.
For every the full subcategory of -local spectra is a mode and we have that in , if and only if and are Bousfield equivalent. For every , we write and for a prime .
Remark 5.2.14.
Given , if , then is a mode (5.2.5(7)). However, it is usually not a localization of , and in particular, it is not the same as . For example, for every we have
Note that the right-hand side is -semiadditive, while the left-hand side is not even -semiadditive. In 5.4.10, we shall show that in a sense, this is the difference between the left and right-hand sides.
As with Bousfield localization, a particularly nice kind of localizations of modes, is given by the ones which are smashing in the sense of 5.1.2. The smashing localizations of modes have a very simple characterization:
Proposition 5.2.15.
A localization of modes is smashing, if and only if the (fully faithful) right adjoint of admits a further right adjoint.
Proof.
In one direction, the forgetful functor admits a right adjoint for every by [Lura, Remark 4.2.3.8]. Conversely, by 5.1.4, it suffices to show that if the right adjoint of admits a further right adjoint, the adjunction satisfies the projection formula. Since is then colimit preserving, the natural transformation in the projection formula
is a natural transformation between two functors , which are colimit preserving in each variable. Equivalently, these are colimit preserving functors [Lura, Section 4.8.1]. Thus, it suffices to check that becomes an isomorphism after whiskering along the equivalence
This amounts to verifying the case , in which, by unwinding the definitions, is the identity and so in particular an isomorphism. ∎
Remark 5.2.16.
Every mode provides by definition a smashing localization . Furthermore, every map of modes induces a smashing localization of the -categories of modules . 5.2.15 characterizes those smashing localizations of , which arise from smashing localizations of
A particular instance of mode localizations arises from divisible and complete objects with respect to a natural endomorphism of the identity functor. More generally,
Proposition 5.2.17.
Let and . The -categories and are accessible localizations of and hence in particular presentable. If moreover , and is given by tensoring with , then the full subcategories are compatible with the symmetric monoidal structure of and are thus symmetric monoidal localizations of .
Proof.
To show that and are accessible localizations of , we use [Lur09, Propostion 5.5.4.15], by which it suffices to realize them as the full subcategories of -local objects with respect to a suitable (small) set of morphisms in . Since is presentable, it is -compactly generated for some cardinal . In particular, the subcategory of -compact objects is essentially small and generates under colimits. For , we take to be the collection of maps for in (a set of representatives of) . For , we can take to be the collection of maps for in (a set of representatives of) -compact objects for large enough so that is -compactly generated and the inclusion is -accessible.
For , the assumption on implies that for all we have and similarly for the adjoint . In particular, the class of -divisible objects is closed under tensoring and exponentiation by any object of . Now, the class of morphisms in , which are mapped to isomorphisms under the localization , is the set of maps in such that for every , the map
is an isomorphism. For any , we have that . Thus, by adjointness, . The argument for is similar but simpler. It again suffices to show that for and , the object is -complete. This follows from the fact that for any -divisible , the object is -divisible. ∎
5.3 Modes of Semiadditivity
In this subsection, we apply the general theory of modes to study the interaction of stability and higher semiadditivity. In particular, we introduce and study the mode which classifies the property of being stable, -semiadditive and of semiadditive height , and compare it with .
Semiadditivity & stability
It is a fundamental result of [Har17], that higher semiadditivity is classified by a mode. More precisely, by 2.1.15 the forgetful functor
admits a left adjoint . We consider as an object of via this left adjoint.
Proposition 5.3.1.
Let . is a mode, which classifies -semiadditivity. Moreover, for every , there is a canonical equivalence141414This can be compared with the fact that the stabilization can be identified with , the -category of spectrum objects in .
Proof.
We first treat the case . The fact that is a mode, which classifies -semiadditivity is exactly [Har17, Corollary 5.21]. Consider now the inclusion . To prove that , note that [Har17, Corollary 5.21] and [Har17, Corollary 5.18] identify the left-hand side and the right-hand side respectively as left adjoints to .
We now consider the case . For verious , the left adjoints of the forgetful functors can be considered as maps in . Thus, by 2.1.15 and 5.2.5(5), we have that
is a mode, classifying the property of being -semiadditive for every . In other words, is a mode classifying -semiadditivity. Finally, since is closed symmetric monoidal, for every we have
∎
Remark 5.3.2.
To study the interaction of higher semiadditivity and stability we introduce the stable -semiadditive mode:
Definition 5.3.3.
For every , we define the -category of -commutative monoids in spectra:
We denote [∞] simply by (and observe that ).
It is an immediate consequence of 5.3.1 and 5.2.5(4), that for all values , the -category [m] is a mode which classifies the property of being -semiadditive and stable.
The mode [m] plays an analogous role to for stable -semiadditive -categories (even for the non-presentable ones). In particular, the -enriched Yoneda embedding of an -semiadditive -category can be further lifted to [m], provided that the -category is stable. To see this, for a stable -category , we already have a limit preserving spectral Yoneda embedding (e.g. as can be deduced from [Lura, Corollary 1.4.2.23])
By 2.1.16, if is in addition -semiadditive, we have a canonical equivalence
Thus, we get a [m]-enriched Yoneda embedding functor
We note that this functor is fully faithful, exact and -semiadditive. Using the [m]-enriched Yoneda embedding, we can characterize the semiadditive height of an object.
Proposition 5.3.4.
Let and let . For every the object is of height (resp. ) if and only if is of height (resp. ), for every object .
Proof.
Using the [m]-enriched Yoneda embedding for , the functors
are -semiadditive and jointly conservative. Thus, by 4.2.3, is of height or if and only if is so for all . ∎
Another useful application of the [m]-enriched Yoneda embedding was already used in the proof of 4.2.7:
Proposition 5.3.5.
For every , there exists and a fully faithful, -semiadditive and exact embedding .
Proof.
We take
which satisfies all the required properties. ∎
Modes of semiadditive height
We can further concentrate on stable higher semiadditive -categories of particular semiadditive height. We shall now show that this property is also classified by a mode.
Theorem 5.3.6.
For every prime and , there exists a mode n which classifies the property of being stable -local -semiadditive of height 151515We keep implicit in the notation n, by analogy with and .. Moreover, n can be canonically identified with for every .
Proof.
We first consider the case . For every , we have that
is a symmetric monoidal localization of (5.2.17). By 5.3.4, for every -local , an object is -local, if and only if . Hence, we can apply 5.2.10 to deduce that is itself a mode classifying the property of being stable -local -semiadditive and of height . Finally, by 3.2.7, every such -category is -semiadditive and hence we can take n to be .
For , we have
Each is a symmetric monoidal localization of (5.2.17). Thus, (p),∞ is also an accessible reflective subcategory of (p) ([Lur09, Proposition 5.4.7.10]), which is compatible with the symmetric monoidal structure. It follows from 5.2.10, that (p),∞ is a mode. Moreover, (p),∞ classifies the property of being stable -local -semiadditive and of height for all , which is the same as being of height . ∎
Example 5.3.7.
In the case , we have
As with the -s, the modes n are disjoint for different -s.
Proposition 5.3.8.
For all , we have .
Proof.
The -category is -semiadditive in which every object is of height both and , and hence must be the zero object. ∎
Another aspect in which n resembles , is that it kills all bounded above spectra:
Proposition 5.3.9.
For , the map of modes vanishes on all bounded above spectra.
Proof.
Denote by the map of modes. The class of spectra on which vanishes is closed under colimits and desuspensions in . Hence, by a standard devissage argument, it suffices to show that vanishes on and for all primes . First, and for are -divisible. Since is -semiadditive, and are -divisible as well, but all objects of n are -complete, and so . Thus, it remains to show that . For every we denote by the fiber of the fold map Applying , we get a fiber sequence
Since n is by definition of height , it follows from 3.2.3(2), that the second map above is an isomorphism for and hence . We observe that can be written as a filtered colimit . Thus, we also get
∎
Corollary 5.3.10.
For , the right adjoint of the map of modes is conservative.
Proof.
The map of modes is given by the composition
The right adjoints and of and respectively, are fully faithful embeddings and hence in particular conservative. The right adjoint of can be identified with the functor
which evaluates at the point. Thus, is conservative by [Har17, Lemma 5.17]. Combining the above, the right adjoint of the functor
is given by , and is therefore conservative.
Now, the right adjoint of , is given by the composition of the right adjoints . Let be a map in n with fiber , such that is an isomorphism. It follows that and hence is co-connective. By 5.3.9, we get and hence . By the zig-zag identities, is a retract of and hence we also get . Finally, since is conservative, we get and hence is an isomorphism. This concludes the proof that , the right adjoint of , is conservative. ∎
Remark 5.3.11.
4.2.8 is equivalent to the statement that .
5.4 1-Semiadditive Decomposition
As we recalled in 4.1.3, while for the -category itself is not even -semiadditive, it is an iterated recollement of the -categories for , which are -semiadditive. The theory of modes allows us to enforce -semiadditivity on in a universal way, by tensoring it with the -semiadditive mode . We shall show that enforcing even -semiadditivity on , “dissolves the glue” which holds the monochromatic layers together, and decomposes it into a product of for .
Remark 5.4.1.
The fact that is a recollement of for , is closely related to the fact that for , every exact functor must be zero. The fact that the recollement becomes split after imposing -semiadditivity follows from the fact that every -semiadditive functor must be zero (5.4.9). Note that by 4.2.3, a -semiadditive functor must be zero because every object of is of height , and thus must be sent to the only object of height in , which is zero. The main “non-formal” ingredient in the proof of 5.4.10 (which makes essential use of the theory developed in [CSY18, Section 4]) is that it suffices to assume merely -semiadditivity.
Divisible and complete cardinalities
We begin with some general observations and constructions. Every is a module over [1] and thus, every induces a natural transformation . If in addition is presentably symmetric monoidal, then there is a unique functor in (see 5.1.6), which induces a map . The natural transformation can be identified with the one induced by the image of in . We now restrict to a particular subset of consisting of elements which have a simple description, as they are of an unstable origin.
Definition 5.4.2.
Let be the -linear span of elements of the form , for a finite group.
We observe that the action of on the objects of any stable -semiadditive -category is natural with respect to -semiadditive functors.
Proposition 5.4.3.
Let be a -semiadditive functor between stable presentable -semiadditive -categories. For every , we have .
Proof.
By -semiadditivity, it suffices to consider the case for a finite group, which follows from the fact that is -semiadditive. ∎
Remark 5.4.4.
If is further assumed to be colimit preserving, then for all elements . However, we shall be interested in applying 5.4.3 to functors , which are not a priori colimit preserving (e.g. right adjoints).
Recall from [CSY18, Theorem 4.3.2], that for every -semiadditive stable presentably symmetric monoidal -category , the ring is equipped with a canonical additive -derivation
Proposition 5.4.5.
The subset is closed under multiplication and the additive -derivation inside . Consequently, is a semi--ring and the inclusion is a homomorphism of semi--rings.
-semiadditive decomposition
We would now like to apply the above to the -categories . By [CSY18, Proposition 4.3.4], the construction of the additive -derivation is functorial with respect to colimit preserving symmetric monoidal functors. In particular, assuming that , both maps
commute with . By 2.2.6, the image of is the canonical copy obtained from the map
Since is the image of a semi--ring map we obtain a surjective map of semi--rings
The semi--ring structure on can be explicitly described.
Lemma 5.4.6.
The unique semi--ring structure on is given by
In particular, if , then .
Proof.
For , we denote by the -adic valuation of .
Proposition 5.4.7.
For every and we have:
-
(1)
If , then is -divisible.
-
(2)
If , then is -complete.
Proof.
For (1), if , then is invertible. It follows that is invertible by nil-conservativity [CSY18, Corollary 5.1.17]. For (2), let such that . We need to show that for every , the element acts non-invertibly on . Since tensoring with is conservative, we may replace by . Without loss of generality, we may choose to be a ring spectrum. Thus, is a -module and the action of on it is by its image under the map , which we denote by . Thus, it would suffice to show that is nilpotent. By the Nilpotence Theorem, it suffices to show that the image of under the map
is nilpotent, as for all . Consider the commutative diagram
We see that the image of in is contained in the image of , which lies in (2.2.6). Since is -power torsion, the map
factors through a finite quotient and hence the image of in is nilpotent. ∎
Proposition 5.4.8.
For every , there exists an element , such that is -complete for and -divisible for .
Proof.
First, we observe that it suffices to construct elements such that is -complete and is -divisible for . Indeed, we can then define , which satisfies the required properties. To construct for a specific , we proceed as follows. For , we take . We get that for , the element acts as zero, while for , the -category is -complete, and hence by 4.1.14, it is -divisible. Now, we treat the case . For every , we shall denote by its image in under the map . By 5.4.7, it suffices to construct , such that and for all . Using 5.4.5, we define the element
By 2.2.5, we have for all . Thus, by 5.4.6, we have for all , and in particular . It follows that there exists an integer with , such that . We define On the one hand, we have by construction , and on the other, for by the ultrametric property of the -adic valuation. ∎
Corollary 5.4.9.
For all , every -semiadditive functor is zero.
Proof.
We are now ready to prove that -semiadditivity forces to decompose into its monochromatic pieces.
Theorem 5.4.10 (1-Semiadditive Decomposition).
For every , there is an equivalence of modes
Proof.
For , the claim is which is true because is -semiadditive. It thus suffices to prove by induction on , that the functor
given by the product of the respective symmetric monoidal localizations, becomes an equivalence after tensoring with . We know that is a recollement of and . It follows that is a recollement of and (5.2.3). By the inductive hypothesis, we have
Thus, is a recollement of and . It is therefore suffices to show that the gluing data given by the functor
is zero (4.1.4). We observe that is given as a composition of a right and a left adjoint between -semiadditive -categories and hence is -semiadditive. Thus, by 5.4.9, we must have and hence the corresponding recollement is split. To conclude, the localization functors for induce a functor , which becomes an equivalence after tensoring with . In particular, this is a symmetric monoidal equivalence and hence an equivalence of modes. ∎
Remark 5.4.11.
By tensoring 5.4.10 with , we also get
5.5 Semiadditive vs. Stable Height
As we recalled in the introduction, for every , the localization functors of spectra , and can be thought of as restriction to heights , and respectively, as measured by the -self maps. It is natural to compare this notion of height with the semiadditive one considered in this paper. In this subsection, we phrase the notion of height classified by , and (which for disambiguation we shall call stable height) in the language of modes and establish some comparison results with the notion of semiadditive height. Using that, we shall prove the bounded version of the “Bootstrap Conjecture” (E), regarding -semiadditivity vs. -semiadditivity for stable presentable -categories.
Stable Height
By 5.2.13, the -categories , and are themselves modes. Our first goal is to show that the properties classified by them can be profitably reinterpreted in terms of the following notion:
Definition 5.5.1.
Given a stable -category , for every , we define and denote the stable height of as follows:
-
(1)
, if for some (hence any) finite spectrum of type .
-
(2)
, if for every of stable height .
-
(3)
, if and .
By convention, for all , and if and only if . We also extend the definition to as follows: For every , we write if and only if for all .
Remark 5.5.2.
Since can be constructed as a cofiber of a self map of , it is clear that implies . Consequently, also implies .
As with the semiadditive height, it is useful to consider the corresponding subcategories of objects of stable height in a certain range:
Definition 5.5.3.
Let and let . We denote by , , and the full subcategories of spanned by objects of stable height , , and . In addition, we write , , or , if , , or respectively.
Example 5.5.4.
In the special case we have by definition
We note that the subcategory is rather large. First, for every , we have
Therefore, for every -module spectrum , we have
which implies that . Since the subcategory is closed under limits, it contains all bounded below -complete spectra. In contrast, the -category satisfies , even though it contains many objects of unbounded height.
We now show that the modes , and classify the properties of having the corresponding stable height.
Proposition 5.5.5.
For every , the modes , and classify the properties of being stable -local of stable height , and respectively.
Proof.
We begin with . It follows from 5.2.10, that is a mode, which classifies stable -local -categories. Thus, it suffice to show that an object in a stable -category is -local, if and only if . By definition, is -local, if for every , the mapping spectrum belongs to . Since the corepresentable functor is exact, we have a canonical isomorphism
It follows that if , then is -local. Since the collection of functors for all is also jointly conservative, the converse holds as well.
We now move to . We first show that if , then is -local for all . For an -acyclic spectrum, is -acyclic as well and hence . Since , it follows that
and hence is -local. Conversely, assume that is -local, and let such that . We have
where is the Spanier-Whithead dual of , which is itself of type . Since is -local and hence -local, we get . This implies that and so .
Finally, for we observe that a spectrum is -local if and only if it is both -local and -local. Thus,
Hence, it classifies the property of being stable -local of both stable height and , i.e exactly . ∎
Similarly, we also treat the case of stable height .
Corollary 5.5.6.
The -category is a mode, which classifies the property of being stable of stable height .
Proof.
We have
and thus is an accessible reflective subcategory of , which is compatible with the symmetric monoidal structure [Lur09, Proposition .5.4.7.10]. It follows from 5.2.10, that is a mode. Moreover, classifies the property that the -category is stable such that every object is -local for all . By 5.5.5, the said condition on is equivalent to for all , which is the same as . Thus, classifies the property of being stable of stable height . ∎
5.5.5 has the following immediate corollaries regarding stable height for stable presentable -categories:
Corollary 5.5.7.
For every -local , we have canonical equivalences
for every .
Corollary 5.5.8.
Every -local is a recollement of and for every .
Proof.
Let be a stable presentable -category. For each let be the right adjoint of the inclusion and the left adjoint of the inclusion . We define
Proposition 5.5.9.
For every -local and . If for all , then . In particular, if , the collection of functors is jointly conservative.
Proof.
Let , such that for all . We shall show by induction that for all , and hence . Assuming by induction that we have
Therefore and hence
It follows that
Hence, for all we get
and so . We can take the base of the induction to be , in which there is nothing to prove. ∎
When further assuming -semiadditivity, we get the following:
Proposition 5.5.10.
For every -local , we have a canonical equivalence:
Comparing heights
For a stable higher semiadditive -category, it is natural to compare the stable height with the semiadditive height. First,
Lemma 5.5.11.
Let be -local. For all , we have .
We next observe that for the -category , which is the mode of stable height , is -semiadditive of semiadditive height (4.4.5). Therefore, there is a map of modes , making an algebra over n.
Proposition 5.5.12.
Let be -local. For all and , we have
Proof.
On the one hand, for we have (5.3.8) and so . On the other, as is a n-module. Tensoring these with yields the claim. ∎
Given tensoring the map of modes with , yields a map .
Proposition 5.5.13.
Let be -local. For every , the map admits a fully faithful right adjoint , which exhibits as a recollement of and (). In particular, if , then .
Proof.
As a consequence, we get a tight connection between and n.
Corollary 5.5.14.
The map of modes is a smashing localization.
Bootstrap of Semiadditivity
Based on the classification of higher semiadditive localizations of with respect to homotopy rings, the authors proposed in [CSY18, Conjecture 1.1.5] the “Bootstrap Conjecture”, stating that if a presentable stable -local -category is -semiadditive, then it is automatically -semiadditive. Using the -semiadditive decomposition of 5.4.10 we now provide some partial results in the direction of proving this conjecture. First, given a -local and , the -category is a module over and hence over n. It follows that is -semiadditive and of height . More generally, if is -semiadditive, then by 5.5.10, we have . From this one can deduce that, if is -semiadditive and every object of is of bounded stable height, then is -semiadditive. However, we shall show that having such a bound on the stable height of the objects of is an unnecessarily strong restriction, and it in fact suffices to assume merely that .
Since every stable presentable -category of stable height exactly (for some ) is -semiadditive, it has an action of , and a fortiori of the subring (see 5.4.2). We begin with a generalization of 5.4.8.
Proposition 5.5.15.
For every , there exists , such that every of stable height is -complete if and -divisible if .
Proof.
By 5.4.8, there exists , such that is -complete if and -divisible if . Since is an -module, the action of on is via its image in . Thus, it suffices to show that for every , if is -complete (resp. -divisible), then is also -complete (resp. -divisible). We observe that is -divisible if and only if is invertible, in which case is -divisible as well. On the other hand, if is -complete, then tensoring with is conservative on . The object is dualizable with dual and so tensoring with is also conservative. Finally, to show that is -complete, it suffices to show that tensoring with is conservative on , via the left tensoring of over . For every , we have isomorphisms
Hence, if , then by the conservativity of , we get that for all and hence . ∎
From this we derive a strengthening of 5.4.9.
Proposition 5.5.16.
Let be -local. If , and , then every 1-semidditive functor is zero.
Proof.
We shall construct an element such that is -divisible and is -complete. Since is -semiadditive, we shall get that it takes every to an object of which is both -complete and -divisible (5.4.3). This will imply that for all and hence that is zero. By 5.5.15, there is an such that is -divisible for and is -complete for . By 5.5.10, we have
and hence itself is -complete. As for , by 5.5.9, we have a jointly conservative collection of functors for . Moreover, all the -s are -semiadditive, as a composition of a left and a right adjoint, and thus is -divisible as well. ∎
As a corollary, we get the following partial result in the direction of [CSY18, Conjecture 1.1.5]:
Theorem 5.5.17.
Let be -local, such that . If is -semiadditive, then is -semiadditive. Moreover, in this case for all and there is a canonical decomposition
Proof.
By 5.5.8, is a recollement of and . Since is -semiadditive and
every 1-semiadditive functor is zero by 5.5.16. It follows that is a split recollement of and , and hence a recollement of and (i.e. we may switch the roles). By 4.1.10, we get
For every , the -category is -semiadditive of semiadditive height , hence itself is -semiadditive and for every we have (5.5.12)
∎
References
- [AKQ19] Gabriel Angelini-Knoll and JD Quigley. Chromatic complexity of the algebraic K-theory of . arXiv preprint arXiv:1908.09164, 2019.
- [AR08] Christian Ausoni and John Rognes. The chromatic red-shift in algebraic K-theory. L’Enseignement Mathématique, 54(2):13–15, 2008.
- [Bae] John C. Baez. Euler characteristic versus homotopy cardinality. http://math.ucr.edu/home/baez/cardinality/cardinality.pdf.
- [Bal16] Paul Balmer. Separable extensions in tensor-triangular geometry and generalized Quillen stratification. Ann. Sci. Éc. Norm. Supér.(4), 49(4):907–925, 2016.
- [BG16] Clark Barwick and Saul Glasman. A note on stable recollements. arXiv preprint arXiv:1607.02064, 2016.
- [BG18] Irina Bobkova and Paul G. Goerss. Topological resolutions in -local homotopy theory at the prime 2. Journal of Topology, 11(4):917–956, 2018.
- [BGH17] Agnes Beaudry, Paul G Goerss, and Hans-Werner Henn. Chromatic splitting for the -local sphere at . arXiv preprint arXiv:1712.08182, 2017.
- [BK72] AK Bousfield and DM Kan. The core of a ring. Journal of Pure and Applied Algebra, 2(1):73–81, 1972.
- [BNT18] Ulrich Bunke, Thomas Nikolaus, and Georg Tamme. The Beilinson regulator is a map of ring spectra. Advances in Mathematics, 333:41–86, 2018.
- [Car84] Gunnar Carlsson. Equivariant stable homotopy and segal’s burnside ring conjecture. Annals of Mathematics, pages 189–224, 1984.
- [CM17] Dustin Clausen and Akhil Mathew. A short proof of telescopic Tate vanishing. Proceedings of the American Mathematical Society, 145(12):5413–5417, 2017.
- [CSY18] Shachar Carmeli, Tomer M Schlank, and Lior Yanovski. Ambidexterity in chromatic homotopy theory. arXiv preprint arXiv:1811.02057, 2018.
- [GGN16] David Gepner, Moritz Groth, and Thomas Nikolaus. Universality of multiplicative infinite loop space machines. Algebraic & Geometric Topology, 15(6):3107–3153, 2016.
- [GH15] David Gepner and Rune Haugseng. Enriched -categories via non-symmetric -operads. Advances in mathematics, 279:575–716, 2015.
- [GS96] John P.C. Greenlees and Hal Sadofsky. The Tate spectrum of -periodic complex oriented theories. Mathematische Zeitschrift, 222(3):391–405, 1996.
- [Har17] Yonatan Harpaz. Ambidexterity and the universality of finite spans. arXiv preprint arXiv:1703.09764, 2017.
- [Hin16] Vladimir Hinich. Dwyer-Kan localization revisited. Homology Homotopy Appl., 18(1):27–48, 2016.
- [HL13] Michael Hopkins and Jacob Lurie. Ambidexterity in -local stable homotopy theory. http://www.math.harvard.edu/ lurie/, 2013.
- [HS96] Mark Hovey and Hal Sadofsky. Tate cohomology lowers chromatic Bousfield classes. Proceedings of the American Mathematical Society, 124(11):3579–3585, 1996.
- [HS98] Michael J. Hopkins and Jeffrey H. Smith. Nilpotence and stable homotopy theory II. Ann. of Math. (2), 148(1):1–49, 1998.
- [HSSS18] Marc Hoyois, Pavel Safronov, Sarah Scherotzke, and Nicolò Sibilla. The categorified Grothendieck-Riemann-Roch theorem. arXiv preprint arXiv:1804.00879, 2018.
- [HY17] Asaf Horev and Lior Yanovski. On conjugates and adjoint descent. Topology and its Applications, 232:140–154, 2017.
- [Kuh04] Nicholas J. Kuhn. Tate cohomology and periodic localization of polynomial functors. Inventiones mathematicae, 157(2):345–370, 2004.
- [Lura] Jacob Lurie. Higher algebra. http://www.math.harvard.edu/ lurie/.
- [Lurb] Jacob Lurie. Representation theory in intermediate characteristic. http://www.claymath.org/downloads/utah/Notes/Lurie2.pdf.
- [Lur09] Jacob Lurie. Higher topos theory, volume 170 of Annals of Mathematics Studies. Princeton University Press, Princeton, NJ, 2009.
- [Lur19] Jacob Lurie. Elliptic cohomology III: Tempered cohomology. Available from the author’s webpage, 2019.
- [MNN17] Akhil Mathew, Niko Naumann, and Justin Noel. Nilpotence and descent in equivariant stable homotopy theory. Advances in Mathematics, 305:994–1084, 2017.
- [PS17] Matan Prasma and Tomer M Schlank. Sylow theorems for -groups. Topology and its Applications, 222:121–138, 2017.